content
stringlengths
1
15.9M
\section{Introduction}\label{intro} \subsection{Background}\label{background} Consider a finite and positive measure $\mu$ of compact and infinite support $\textrm{supp}(\mu)\subseteq\mathbb{C}$. Given such a measure and any $q>0$, we can define the sequence of monic polynomials $\{P_n(z;\mu,q)\}_{n=0}^{\infty}$ by letting $P_n(z;\mu,q)$ be any monic polynomial satisfying \[ \|P_n(z;\mu,q)\|_{L^q(\mu)}=\inf\{\|Q_n\|_{L^q(\mu)}:Q_n=z^n+\mbox{ lower order terms}\}; \] a property called the \textit{extremal property} of the polynomials $P_n(z;\mu,q)$. For $q>1$, the strict convexity of the norm implies such a polynomial is unique, while it need not be when $0<q\leq1$ (see page 84 in \cite{StaTo}, see also Proposition \ref{nonunique} in the appendix for a treatment of the case $q=1$). When the meaning is clear, we will often omit the $z$, $\mu$, or $q$ dependence of $P_n(z;\mu,q)$ in our notation. By dividing each $P_n(\mu,q)$ by its $L^q(\mu)$ norm, we obtain the sequence of normalized polynomials $\{p_n(\mu,q)\}_{n=0}^{\infty}$ (we use the word ``norm" here loosely as it is not technically a norm when $q<1$). In case $q=2$, the polynomial $p_n(\mu,q)$ is just the orthonormal polynomial for the measure $\mu$. For an extensive introduction to the general theory of orthogonal polynomials - especially on the real line and the unit circle - we refer the reader to the references \cite{NevaiOP,OPUC1,Rice,StaTo,Suetin,Szego} and references therein. We will consider measures whose support is contained in some compact and simply connected set $\overline{G}$ along with finitely many points not in $\overline{G}$. We will also assume that $G$ is a region with analytic boundary (as defined on page 42 in \cite{GarnMar}) and that the logarithmic capacity (see section \ref{tools} below) of $G$ is equal to $1$. One of our main tools for studying these extremal polynomials is the conformal map $\psi$ mapping the exterior of the closed unit disk $\overline{\bbD}$ to $\overline{\mathbb{C}}\setminus\overline{G}$ and satisfying $\psi(\infty)=\infty$ and $\psi'(\infty)>0$. We will denote the inverse function to $\psi$ by $\phi$. Since $G$ has analytic boundary, the map $\psi$ can be extended to be univalent (that is, holomorphic and injective) on a slightly larger region, namely the exterior of the closed disk of radius $\tilde{\rho}$ for some $\tilde{\rho}<1$. From now on we will assume that $\rho$ is some fixed number in the interval $(\tilde{\rho},1)$. If we define $A_r:=\{z:r\leq|z|\leq1\}$ for every $r\in[\rho,1]$ and define $G_r:=\psi(A_r)$ then $\psi$ and $\phi$ provide a one-to-one correspondence between measures on $A_{\rho}$ and $G_{\rho}$. Given a measure $\lambda$ on $G_{\rho}$, we will denote the corresponding measure $\kappa$ on $A_{\rho}$ by $\phi_*\lambda$. By this we mean that for all $f\in C(\overline{G})$, we have \[ \int_{G_{\rho}}f(z)d\lambda(z)=\int_{A_{\rho}}f(\psi(w))d\kappa(w). \] Similarly, we can write $\lambda=\psi_*\kappa$. For example, the equilibrium measure for $\overline{G}$ can written as $\psi_*(\frac{d\theta}{2\pi})$ (see Theorem 3.1 in \cite{ToTrans}). Central to the theory of $L^q$ extremal polynomials on a smooth Jordan curve is the analog of \textit{Szeg\H{o}'s Theorem} on the unit circle. This can be stated as the following theorem, which follows from Theorem 7.1 in \cite{Geronimus}: \begin{theorem}\label{geronseven}\cite{Geronimus} If $\mu$ is a finite measure on an analytic Jordan curve $\Gamma$ having capacity $1$, then \begin{align}\label{szegstat} \lim_{n\rightarrow\infty}\|P_n(\mu,q)\|^q_{L^q(\mu)}=\exp\left(\int_0^{2\pi}\log((\phi_*\mu)'(\theta)))\frac{d\theta}{2\pi}\right). \end{align} \end{theorem} Any measure for which the right hand side of (\ref{szegstat}) is finite will be called a \textit{Szeg\H{o} measure} on $\partial\mathbb{D}$. Szeg\H{o}'s Theorem can also be stated for measures on the real line (see Theorem 1.1 in \cite{FinGap2} for a precise statement). We note here that Szeg\H{o}'s Theorem for analytic curves - as we have stated it - does not require $\mu$ to be a probability measure. There has also been considerable research on orthogonal polynomials for measures supported on regions. Substantial results were introduced by Carleman in \cite{Carleman} and major achievements in the field since then include the works of Ullman \cite{Ull1,Ull2}, Suetin \cite{Suetin}, Lubinsky \cite{LubBerg}, Mi\~{n}a-D\'{i}az \cite{MDPoly}, Saff \cite{SaffConj}, Stylianopoulos \cite{Corner}, Totik \cite{ToTrans}, and Widom \cite{Wid2} among others. Recently, Nazarov, Volberg, and Yuditskii showed in \cite{Koosis} that the appropriate analog of Szeg\H{o}'s Theorem holds when $\mu$ can be written as the sum of a measure carried by $\mathbb{D}=\{z:|z|<1\}$, a Szeg\H{o} measure on $\partial\mathbb{D}$ with no singular part, and a pure point measure carried by the compliment of $\overline{\bbD}$. Their result motivates our investigation in many ways. We provide leading order asymptotics for the monic orthogonal polynomial norms in a related setting. A special case of our main theorem (Theorem \ref{bigone} below) applies to measures similar to those considered in \cite{Koosis} although we allow for a singular component to the measure on $\partial\mathbb{D}$, but we only allow for finitely many pure points outside $\overline{\bbD}$. Our results can also be motivated by the conjecture in \cite{SaffConj}. If a measure on $\mathbb{D}$ is given by $w(z)d^2z$ where $w>0$ Lebesgue almost everywhere in some annulus with outer boundary $\partial\mathbb{D}$, the conjecture asserts that \[ \lim_{n\rightarrow\infty}\frac{p_{n+1}(z;\mu,2)}{zp_{n}(z;\mu,2)}=1 \] uniformly on compact subsets of $\overline{\bbC}\setminus\overline{\bbD}$. We will settle this conjecture in the affirmative if the weight factors as $w(re^{i\theta})=h(re^{i\theta})f(\theta)g(r)$ with $h$ continuous and non-vanishing on $\overline{\bbD}$, $f$ a Szeg\H{o} weight, and $1\in\textrm{supp}(g(r)dr)$. Indeed our main theorem considers measures that can be thought of as perturbations of such measures. We consider several different kinds of perturbations, including adding finitely many pure points outside $\mathbb{D}$, allowing a singular component to the angular measure with weight $f(\theta)$, and the addition of a measure whose density at the boundary is negligible compared to $w(z)d^2z$. Throughout this paper, for a measure $\gamma$ (on any set), we denote \[ c_t(\gamma)=\int_{\mathbb{C}}|z|^t\,d\gamma(z) \] where we do \textit{not} insist $t\in\mathbb{N}$. We will see that these ``moments" provide the appropriate rate of decay of the norms of the extremal polynomials. One of our main results is the following: \begin{theorem}\label{bigone} Consider the measure $\tilde{\mu}(re^{i\theta})=h(re^{i\theta})\left(\nu(\theta)\otimes\tau(r)\right)+\sigma_2(re^{i\theta})$ where \begin{enumerate} \item $h(z)$ is a continuous function on $\overline{\bbD}$ that is non-vanishing in a neighborhood of $\partial\mathbb{D}$, \item $\sigma_2$ is a measure carried by $A_{\rho}$ that satisfies $\lim_{t\rightarrow\infty}c_t(\sigma_2)c_{t}(\tau)^{-1}=0$, \item $\nu$ is a measure on the unit circle such that $\nu'(\theta)>0$ Lebesgue almost everywhere, \item $\tau$ is a measure on $[\rho,1]$ such that $1\in\textrm{supp}(\tau)$. \end{enumerate} Let $\mu$ be the measure on $\mathbb{C}$ be given by \[ \mu=\psi_*\tilde{\mu}+\sigma_1+\sum_{j=1}^m\alpha_j\delta_{z_j}+\sum_{j=1}^{\ell}\beta_j\delta_{\zeta_j} \] where $\textrm{supp}(\sigma_1)\subseteq G$, $\alpha_j,\beta_j>0$, $z_j\not\in\overline{G}$ for all $j\in\{1,\ldots,m\}$, and $\zeta_j\in\partial G$ for all $j\in\{1,\ldots,\ell\}$. Then \begin{align}\label{bigconcl} \lim_{n\rightarrow\infty}\frac{\|P_n(z;\mu,q)\|^q_{L^q(\mu)}}{c_{qn}(\tau)}= \exp\left(\int_0^{2\pi}\log\left(h(e^{i\theta})\nu'(\theta)\right)\,\frac{d\theta}{2\pi}\right)\prod_{j=1}^m|\phi(z_j)|^q. \end{align} \end{theorem} \noindent\textit{Remark.} We do not actually need $h$ to be continuous. The proof of Theorem \ref{bigone} will show that $h$ need only satisfy the following conditions: \begin{enumerate} \item $0<\delta_1<h(re^{i\theta})<\delta_2<\infty$ for all $r\in[\rho,1]$ and $\theta\in[0,2\pi]$, \item $\int_0^{2\pi}\log(h(re^{i\theta}))\frac{d\theta}{2\pi}$ is a contunous funtion of $r\in[\rho,1]$, \item for any polynomial $\Phi$, any $k\in\mathbb{N}$, and any $q>0$ it holds that \[ \lim_{r\rightarrow1^-}\int_0^{2\pi}e^{ik\theta}|\Phi(\psi(re^{i\theta}))|^qh(re^{i\theta})\frac{d\theta}{2\pi}= \int_0^{2\pi}e^{ik\theta}|\Phi(\psi(e^{i\theta}))|^qh(e^{i\theta})\frac{d\theta}{2\pi}. \] \end{enumerate} For example, if $D^+=\mathbb{D}\cap\{z=x+iy:x>0\}$ and $D^-=\mathbb{D}\cap\{z=x+iy:x\leq0\}$ then $h(z)=\chi_{D^+}(z)+2\chi_{D^-}(z)$ is a function $h$ to which Theorem \ref{bigone} applies. \vspace{2mm} During the preparation of this manuscript, we discovered the recent results of Baratchart and Saff, which are outlined in \cite{BaraSaff}. They consider measures on the unit disk that in may ways resemble the measures we consider in Theorem \ref{bigone}. They obtain similar results on the asymptotic behavior of the monic orthogonal polynomial norms, though Theorem \ref{bigone} seems to be more general. The factor of $\prod_{j=1}^m|\phi(z_j)|^q$ in (\ref{bigconcl}) is exactly what one would expect given the results of \cite{Kalicurve,Kaliarc,KaliKon,KiSi,Koosis,PanLi,PrYd}. We will call a measure $\mu$ as in the statement of Theorem \ref{bigone} a \textit{push-forward of a product measure}. Let us consider some examples of measures to which we can apply Theorem \ref{bigone}. \vspace{2mm} \noindent\textbf{Example.} If we set $q=2$, $d\nu=\frac{d\theta}{2\pi}$, $d\tau=2rdr$, $\sigma_1=\sigma_2=0$, and $\ell=m=0$ then we are dealing with measures of the form $h(z)d^2z$ for a function $h$ continuous and non-vanishing on $\overline{\bbD}$. Such measures with an added H\"{o}lder continuity assumption on $h$ were considered by Suetin in \cite{Suetin}. Theorem \ref{bigone} recovers the leading term in the conclusion of Theorem 3.1 in \cite{Suetin}. \vspace{2mm} \noindent\textbf{Example.} If we set $G=\mathbb{D}$, $\tau=\delta_1$, and $h=1$ then we recover a result similar to that of \cite{Koosis} (when $q=2$) that allows for a singular component of the measure on $\partial\mathbb{D}$, but only finitely many pure points outside $\overline{\bbD}$. If we further set $\sigma_1=\sigma_2=0$ then we can recover the result from Theorem 2.2 in \cite{Kalicurve} (for any $G$ with analytic boundary). \vspace{2mm} \noindent\textbf{Example.} Let us set $G=\mathbb{D}$ and $\tau=\frac{6}{\pi^2}\sum_{j=1}^{\infty}j^{-2}\delta_{1-2^{-j}}$, $d\nu=\frac{d\theta}{2\pi}$, $h=1$, $\ell=m=0$, and $\sigma_2=\sigma_1=0$. If $s$ is a sufficiently large power of $2$ then \[ c_s(\tau)=\frac{6}{\pi^2}\sum_{j=1}^{\infty}j^{-2}(1-2^{-j})^s\geq \frac{6}{\pi^2\log_2(s)^2}\left(1-\frac{1}{s}\right)^{s}\geq\frac{C}{\log_2(s)^2} \] for some constant $C>0$. Theorem \ref{bigone} implies that in this example, the extremal polynomial norms do \textit{not} decay like $O(n^{-1})$ as $n\rightarrow\infty$. \vspace{2mm} \noindent\textbf{Example.} Let us set $G=\mathbb{D}$, $\tau=(1-r)dr$, $d\nu=d\nu_{ac}$, $\ell=m=0$, and $\sigma_2=\sigma_1=0$. In this case, we have $d\mu(z)=w(z)d^2z$ where the weight $w$ vanishes on the boundary. Theorem \ref{bigone} still applies to this measure, and we will see below that we can still derive the asymptotics of the extremal polynomials outside $\overline{\bbD}$. \vspace{2mm} Theorem \ref{bigone} provides the asymptotic behavior of the norms of the $L^q(\mu)$-extremal polynomials for general $q\in(0,\infty)$. We can also deduce the behavior of the extremal polynomials outside the compact set $\overline{G}$, i.e. we can prove what is often referred to as \textit{strong asymptotics}. If $\mu$ is of the form considered in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$ then we can prove the following: \begin{enumerate} \item\label{szegasy} there are polynomials $\{y_n\}_{n\in\mathbb{N}}$ (depending on $P_n(\mu,q)$ and $q$) of degree $m$ and a function $S=S_q$ analytic and non-vanishing in $\overline{\mathbb{C}}\setminus\overline{\bbD}$ and positive at $\infty$ so that \begin{align* \lim_{n\rightarrow\infty}\frac{P_n(\psi(z);\mu,q)S(z)}{y_n(\psi(z))z^{n-m}S(\infty)}=1 \end{align*} uniformly on compact subsets of $\overline{\mathbb{C}}\setminus\overline{\bbD}$, \item\label{weakk} the probability measures $|p_n(z;\mu,q)|^qd\mu$ converge weakly to the equilibrium measure for $\overline{G}$ as $n\rightarrow\infty$, \item\label{sumo} for any $z\in G$, we have $\sum_{n=0}^{\infty}|p_n(z;\mu,2)|^2<\infty$. \end{enumerate} Item (\ref{sumo}) follows from an argument based on Christoffel functions and the associated minimization problem. We will discuss this in more detail in Section \ref{christo} and for all values of $q>0$. The function $S$ in item (\ref{szegasy}) will be of the form given in (\ref{szegfundef}) below. We will see that the polynomial $y_n$ in item (\ref{szegasy}) has a single zero near each $z_i$ for $i\in\{1,\ldots,m\}$ and shares all of its zeros with $P_n(z;\mu,q)$. \subsection{Tools and Methods}\label{tools} In an effort to fix notation and for the reader's convenience, we will now provide a brief summary of the main tools that we will use in our proofs. In some of our proofs we will make heavy use of the Szeg\H{o} function, which we now define. For a Szeg\H{o} measure $\gamma$ on $\partial\mathbb{D}$ with Radon-Nikodym derivative given by $\gamma'(\theta)$ we define a function $S(z;q)$, which is analytic on $\{z:|z|>1\}$ by \begin{align}\label{szegfundef} S(z;q)=\exp\left(-\frac{1}{2q\pi}\int_0^{2\pi}\log(\gamma'(\theta))\frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta\right) \qquad,\qquad |z|>1, \end{align} which we will often denote by $S(z)$ if the intended value of $q$ is clear. By the same argument as in the proof of Theorem 2.4.1(ii) in \cite{OPUC1}, we know $S(z;q)\in\mathbb{H}^q(\overline{\bbC}\setminus\overline{\bbD})$. We should mention that different authors have used different definitions of the Szeg\H{o} function. The one we use was also used in \cite{Geronimus,Suetin}, while \cite{OPUC1,OPUC2,PrYd} prefer to define the Szeg\H{o} function slightly differently and with domain $\mathbb{D}$. It follows from equation (6.5) in \cite{Geronimus} that \[ |S(e^{i\theta};q)|^q=\lim_{r\searrow1}|S(re^{i\theta};q)|^q=\gamma'(\theta)\qquad,\qquad a.e.\,\,\theta\in[0,2\pi]. \] We will also need potential theoretic objects such as the equilibrium measure, logarithmic potential, and Green function of a compact set. We refer the reader to the books \cite{GarnMar,Ransford,SaffTot} for additional background in potential theory and to \cite{SimPot,StaTo} for extensive applications of these ideas to orthogonal polynomials. Given a finite measure $\gamma$ of compact support, we can define its \textit{logarithmic potential} \[ U^{\gamma}(z):=\int_{\mathbb{C}}\log\frac{1}{|z-w|}d\gamma(w), \] though for some values of $z$, the integral may be $+\infty$. We define the \textit{equilibrium measure} of a compact set $K$ as the unique probability measure $\omega_K$ satisfying \[ \int_K\int_K\log\frac{1}{|z-w|}d\omega_K(z)d\omega_K(w)= \inf\left\{\int_K\int_K\log\frac{1}{|z-w|}d\gamma(z)d\gamma(w):\gamma(K)=1=\gamma(\mathbb{C})\right\} \] provided the right hand side is finite. In this case we call the left hand side the \textit{logarithmic energy} of $\omega_K$ and denote it by $E(\omega_K)$. It is always true that the support of the equilibrium measure $\omega_K$ is contained in the boundary of $K$ (see Theorem 3.7.6 in \cite{Ransford}). We define the \textit{logarithmic capacity} of the compact set $K$ as $e^{-E(\omega_K)}$ and denote it by $\mbox{cap}(K)$. In this paper, we will always assume that $\mbox{cap}(\overline{G})=1$ and consequently (with $\psi$ defined as in Section \ref{background}) $\psi'(\infty)=1$. In this case, we can write \begin{align* \psi(z)=z+\xi_0+\frac{\xi_1}{z}+\frac{\xi_2}{z^2}+\cdots\qquad,\qquad\xi_i\in\mathbb{C}. \end{align*} A measure $\gamma$ with compact support is called a \textit{regular measure} if \[ \lim_{n\rightarrow\infty}\|P_n(\gamma,2)\|^{1/n}_{L^2(\gamma)}=\mbox{cap}(\textrm{supp}(\gamma)). \] The equilibrium measure will play an important role in Section \ref{strong}. We will mention regularity again in Section \ref{product} and it is a major topic throughout \cite{StaTo}. Armed with the notions of equilibrium measure and capacity, we can define the Green function with pole at infinity of a compact set $K$ (of positive capacity) as \begin{align}\label{greedef} g_{\overline{\bbC}\setminus K}(z;\infty):=-U^{\omega_K}(z)-\log(\mbox{cap}(K)). \end{align} It follows from Theorem 4.4.4 in \cite{Ransford} that the Green's function is conformally invariant, i.e. if $K_1$ and $K_2$ are simply connected compact sets in the plane and $\mathcal{F}$ is the conformal map that sends the compliment of $K_1$ to the compliment of $K_2$ mapping $\infty$ to itself and having positive derivative there, then \[ g_{\overline{\bbC}\setminus K_1}(z;\infty)=g_{\overline{\bbC}\setminus K_2}(\mathcal{F}(z);\infty). \] \vspace{2mm} We also include here a brief discussion of Faber polynomials (see \cite{MDFaber} for extensive background and references). We will denote these polynomials by $F_n(z)$ and they are defined as the polynomial part of the Laurent expansion of $\phi^n(z)$ around $\infty$. Since we are assuming $\mbox{cap}(\overline{G})=1$, we recover from formula (1.4) in \cite{MDFaber} the following two facts: \begin{enumerate} \item\label{monicfaber} $F_n(z)$ is a monic polynomial of degree $n$, \item\label{faberdecay} for $\rho<|z|\leq1$ we have \[ F_n(\psi(z))=z^n+O(\tilde{\rho}^n) \] where the implied constant is uniformly bounded from above in the annulus considered. \end{enumerate} In case $G=\mathbb{D}$, we have $F_n(z)=z^n$. Many of the proofs in \cite{Suetin} and the proof of the main theorem in \cite{Koosis} rely heavily on generalized Faber polynomials. We will use Faber polynomials to obtain upper bounds on the $L^q$ norms of the extremal polynomials. \vspace{2mm} The remainder of the paper is organized as follows. In Section \ref{product}, we prove Theorem \ref{bigone}. One key step will be to use Faber polynomials and look at weak limits of the measures $\left\{\frac{|F_n(z)|^qd\mu}{c_{qn}(\tau)}\right\}_{n\in\mathbb{N}}$. In Section \ref{strong} we will discuss strong asymptotics of the extremal polynomials for measures of the form considered in Theorem \ref{bigone}. In Section \ref{christo} we will discuss Christoffel functions and their behavior on the set $\overline{G}$, especially inside the region $G$. A major theme throughout will be the many similarities with the theory of orthogonal polynomials on the unit circle (OPUC). Many of our results produce interesting corollaries and we will point these out as we go. Throughout this paper, we will let $\Gamma_r$ be the contour given by $\{\psi(z):|z|=r\}$ for $r>\tilde{\rho}$ and $\mathcal{G}_r$ will denote the region bounded by $\Gamma_r$. \vspace{2mm} \noindent\textbf{Acknowledgements.} It is a pleasure to thank Barry Simon for encouraging me to pursue this line of inquiry and for much useful discussion. I would also like to thank M. Lukic for his help with the reference \cite{Geronimus}. \section{Push-Forward of Product Measures on the Disk}\label{product} In this section, we will derive norm asymptotics for the extremal polynomials corresponding to measures of the form considered in Theorem \ref{bigone}. We will use the Faber polynomials in conjunction with the extremal property to eventually derive an upper bound in the proof of Theorem \ref{bigone} and we will use subharmonicity of appropriate functions to derive a lower bound. For the remainder of this section, we will let $q>0$ be fixed but arbitrary and we will denote $P_n(z;\mu,q)$ by $P_n(\mu)$ and $\|P_n(\mu)\|_{L^q(\mu)}$ by $\|P_n(\mu)\|_{\mu}$ when there is no possibility for confusion. We begin with the following crude estimate: \begin{proposition}\label{nuregreg} If $\mu$ is as in Theorem \ref{bigone} then $\mu$ is regular. \end{proposition} \begin{proof} We will in fact show that $\mu$ satisfies Widom's criterion (see Section 4.1 in \cite{StaTo}) from which regularity immediately follows by Theorem 4.1.6 in \cite{StaTo}. For each $r\in(\rho,1]$, the equilibrium measure of the curve $\Gamma_r$ is absolutely continuous with respect to arc-length measure with continuous derivative bounded above and below by positive constants (see Theorem II.4.7 in \cite{GarnMar}; the constants are allowed to depend on $r$). Let $C$ be a carrier of $\mu$ (i.e. $\mu(C)=\mu(\mathbb{C})$). Since $\nu'(\theta)>0$ Lebesgue almost everywhere, we conclude that \[ \lambda_r(C\cap\Gamma_r)=\ell(\Gamma_r) \] for $\tau$ almost every $r\in(\rho,1]$ where $\lambda_r$ is arc-length measure on $\Gamma_r$ and $\ell(\Gamma_r)$ is the length of the curve $\Gamma_r$. It follows that there is a sequence $r_n\rightarrow1$ such that $\omega_{\Gamma_{r_n}}(C)=1$ while clearly $\mbox{cap}(\Gamma_{r_n})\rightarrow1$. This shows $\mu$ satisfies Widom's criterion. \end{proof} We will now begin developing the ideas necessary to prove the more refined estimate of $\|P_n(\mu)\|^q_{\mu}$ given in Theorem \ref{bigone}. We begin with a lemma that immediately highlights the importance of Faber polynomials to our results. \begin{lemma}\label{momentous} Let $\mathcal{N}\subseteq\mathbb{N}$ be a subsequence such that \[ w\mbox{-}\lim_{{n\rightarrow\infty}\atop{n\in\mathcal{N}}}\frac{|F_n(z)|^qd\mu(z)}{a_n}=d\gamma \] where $\gamma$ is a measure on $\partial G$ and $\{a_n\}_{n\in\mathbb{N}}$ is a sequence of positive real numbers satisfying $\lim_{n\rightarrow\infty}a_na_{n+1}^{-1}=1$. Then for any fixed $k\in\mathbb{N}$, we have \[ w\mbox{-}\lim_{{n\rightarrow\infty}\atop{n\in\mathcal{N}}}\frac{|F_{n-k}(z)|^qd\mu}{a_n}=d\gamma. \] \end{lemma} \begin{proof} Recall our notation $G_{\rho}=\{\psi(z):\rho\leq|z|\leq1\}$. It is clear from our earlier discussion of Faber polynomials (specifically fact (\ref{faberdecay})) that all weak limits in question are measures on $\partial G$ and that $F_n$ has no zeros in $G_{\rho}$ for all sufficiently large $n$. Now, let $f$ be a continuous function on $G_{\rho}$. We have \begin{align*} \int_{G_{\rho}}f(z)\frac{|F_n(z)|^q}{a_n}d\mu(z)-&\int_{G_{\rho}}f(z)\frac{|F_{n-k}(z)|^q}{a_n}d\mu(z)=\\ &=\int_{G_{\rho}}f(z)\left(1-\frac{|F_{n-k}(z)|^q}{|F_n(z)|^q}\right)\frac{|F_n(z)|^q}{a_n}d\mu(z)\\ &=\int_{G_{\rho}}f(z)\left(1-\frac{|\phi(z)|^{q(n-k)}+O(\tilde{\rho}^n)}{|\phi(z)|^{qn}+O(\tilde{\rho}^n)}\right)\frac{|F_n(z)|^q}{a_n}d\mu(z)\\ &\rightarrow\int_{\partial G}f(z)\left(1-|\phi(z)|^{-qk}\right)d\gamma(z)\\ &=0 \end{align*} since $|\phi(z)|=1$ when $z\in\partial G$. \end{proof} Our next lemma will identify some ideal choices for the sequence $\{a_n\}_{n\in\mathbb{N}}$ of Lemma \ref{momentous}. \begin{lemma}\label{onemoment} Let $\gamma$ be a probability measure on the unit interval $[0,1]$, let $c_n$ denote the $n^{th}$ moment of $\gamma$. The following are equivalent: \begin{enumerate} \item\label{suppq} $1\in\textrm{supp}(\gamma)$, \item\label{nthrot} $\lim_{n\rightarrow\infty}c_n^{1/n}=1$, \item\label{ratoi} $\lim_{n\rightarrow\infty}c_{q(n+1)}c_{qn}^{-1}=1$. \end{enumerate} \end{lemma} \begin{proof} It is obvious that (\ref{suppq})$\Rightarrow$(\ref{nthrot}) and (\ref{ratoi})$\Rightarrow$(\ref{suppq}) so we need only prove that (\ref{nthrot})$\Rightarrow$(\ref{ratoi}). To this end, we have \[ \frac{c_{qn+q}}{c_{qn}}=1+\frac{\int_0^1r^{qn}(r^q-1)d\gamma(r)}{\int_0^1r^{qn}d\gamma(r)}. \] If $\lim_{n\rightarrow\infty}c_n^{1/n}=1$ then the measures $\frac{r^{qn}d\gamma(r)}{\int_0^1r^{qn}d\gamma(r)}$ converge weakly to the point mass at $1$ as $n\rightarrow\infty$, which implies the desired conclusion. \end{proof} Now we can prove the following lemma, which will be of critical importance in our proof of Theorem \ref{bigone}. \begin{lemma}\label{weakboundary} Let $\kappa$ be a measure on $\overline{G}$ and $\gamma$ a measure on $\partial\mathbb{D}$ and let $\mathcal{N}\subseteq\mathbb{N}$ be a subsequence such that \[ w\mbox{-}\lim_{{n\rightarrow\infty}\atop{n\in\mathcal{N}}}\frac{|F_n(z)|^q}{a_n}d\kappa=d(\psi_*\gamma) \] where $\{a_n\}_{n\in\mathbb{N}}$ is as in Lemma \ref{momentous}. Then \[ \limsup_{{n\rightarrow\infty}\atop{n\in\mathcal{N}}}\frac{\|P_n(\kappa)\|^q_{\kappa}}{a_n}\leq \exp\left(\int_0^{2\pi}\log(\gamma'(\theta))\frac{d\theta}{2\pi}\right). \] \end{lemma} \begin{proof} By the extremal property, we have $\|P_n(\kappa)\|^q_{\kappa}\leq \|F_{n-k}(z)P_k(\psi_*\gamma)\|^q_{\kappa}$. By Lemma \ref{momentous}, we can write \[ \int_{\overline{G}}\frac{|P_k(z;\psi_*\gamma)|^q|F_{n-k}(z)|^q}{a_n}d\kappa(z) \rightarrow\int_{\partial G}|P_k(z;\psi_*\gamma)|^qd(\psi_*\gamma) \] as $n\rightarrow\infty$ through $\mathcal{N}$. Therefore \[ \limsup_{{n\rightarrow\infty}\atop{n\in\mathcal{N}}}a_{n}^{-1}\|P_n(z;\kappa)\|^q_{\kappa} \leq \|P_k(\psi_*\gamma)\|^q_{\psi_*\gamma} \] for every $k>0$. Since $k$ here is arbitrary, we can take the infimum over all $k$, which is no larger than the limit as $k$ tends to infinity. The result now follows from Theorem \ref{geronseven}. \end{proof} As a second preparatory step for the proof of Theorem \ref{bigone}, we need to understand how to deal with pure points outside of $\overline{G}$. The following lemma is a consequence of the remark following the statement of Theorem 1 in \cite{What} (although Theorem 1 in \cite{What} is only stated for the orthonormal polynomials ($q=2$), the same proof works for the extremal polynomials in any $L^q(\mu)$ space). \begin{lemma}\label{closetozone}\cite{What} Let $\mu$ be a finite measure carried by $\overline{G}\bigcup\{z_1,\ldots,z_m\}$ where $\overline{G}$ is simply connected and each $z_i\not\in\overline{G}$. Then for any $\delta>0$, there is $N_\delta$ such that if $n>N_{\delta}$ then $\{u:|u-z_i|<\delta\}$ has at least one zero of $P_n(\mu)$ for each $i=1,2,\ldots,m$. \end{lemma} \noindent\textit{Remark.} We will refine Lemma \ref{closetozone} later in this section (see Corollary \ref{closertozone}). \vspace{2mm} \noindent It is also shown in \cite{What} that pure points of $\mu$ inside $G$ need not attract zeros of $P_n(\mu)$. The case of pure points on the boundary of $G$ is more subtle (see Section 10.13 in \cite{OPUC2} for more results on point perturbation). The following calculation will be useful also. \begin{proposition}\label{psiint} If $x\not\in G$ and $r\in[\rho,1]$ then \[ \int_0^{2\pi}\log|\psi(re^{i\theta})-x|^q\frac{d\theta}{2\pi}=\log|\phi(x)|^q. \] \end{proposition} \begin{proof} First, consider the case when $x\not\in\overline{G}$. It is clear that $\mbox{cap}(\overline{\mathcal{G}}_{r})=r$. Define $\psi_r(z)=\psi(rz)$ on $\{z:|z|>\tilde{\rho}r^{-1}\}$. Then we calculate \begin{align*} \log|\phi(x)|^q&=\int_0^{2\pi}\log|e^{i\theta}-\phi(x)r^{-1}|^q\frac{d\theta}{2\pi}+q\log(r)\\ &=-qU^{\omega_{\overline{\bbD}}}\left(\frac{\phi(x)}{r}\right)+q\log(r)\\ &=qg_{\overline{\bbC}\setminus\overline{\bbD}}(\phi(x)r^{-1},\infty)+q\log(r)\\ &=qg_{\overline{\bbC}\setminus\overline{\mathcal{G}}_{r}}(x,\infty)+q\log(\mbox{cap}(\overline{\mathcal{G}}_{r}))\\ &=\int_{\Gamma_r}\log|y-x|^qd\omega_{\overline{\mathcal{G}}_{r}}(y)\\ &=\int_0^{2\pi}\log|\psi_r(e^{i\theta})-x|^q\frac{d\theta}{2\pi}. \end{align*} The first line follows from Example 0.5.7 in \cite{SaffTot}. The second line is just the definition of the logarithmic potential. The third line then follows from (\ref{greedef}) above and the fact that $\overline{\bbD}$ has logarithmic capacity $1$. The fourth line then follows from the conformal invariance of the Green's function (see Section \ref{tools}). The fifth line follows as the third did from the first. Finally, the last line follows from the definition of equilibrium measure as given in Theorem 3.1 in \cite{ToTrans}. The case $x\in\partial G$ follows by dominated convergence as in Example 0.5.7 in \cite{SaffTot}. \end{proof} Before we proceed with the proof of Theorem \ref{bigone}, we need to make an observation. The upper bound will be obtained using the above lemmas, while for the lower bound we will invoke subharmonicity of a particular integrand. This is simple enough when $q\geq1$ because every $H^1$ function is the Poisson integral of its boundary values (see Theorem 17.11 in \cite{Rudin}). However, some care is required when $0<q<1$. We simply note here that Theorem 17.11(c) in \cite{Rudin} combined with a well-known $L^q$ inequality (see page 74 in \cite{Rudin}) imply that if $f\in H^q(\overline{\bbC}\setminus\overline{\bbD})$ then \begin{align}\label{subharm} \int_0^{2\pi}|f(e^{i\theta})|^q\frac{d\theta}{2\pi}\geq|f(\infty)|^q. \end{align} Now we are ready to prove Theorem \ref{bigone}. \vspace{2mm} \noindent\textit{Proof of Theorem \ref{bigone}.} For now, let us assume that $\ell=m=0$ in our definition of $\mu$. For any $k\in\mathbb{N}$, we have \begin{align*} \lim_{n\rightarrow\infty}\int_{\overline{G}}\frac{\phi(z)^k|F_n(z)|^q}{c_{qn}(\tau)}d\mu(z)&= \lim_{n\rightarrow\infty}\frac{\int_{\rho}^{1}\int_{0}^{2\pi}r^{k+qn}e^{ik\theta}h(re^{i\theta})d\nu(\theta)d\tau(r)}{c_{qn}(\tau)}+ \lim_{n\rightarrow\infty}\frac{\int_{\overline{\bbD}}z^{k}|z^{n}|^qd\sigma_2}{c_{qn}(\tau)}+o(1)\\ &=\int_0^{2\pi}e^{ik\theta}h(e^{i\theta})d\nu(\theta)=\int_{\partial G}\phi(z)^kd(\psi_*(h\nu)). \end{align*} It follows that the measures $\frac{|F_n(z)|^2}{c_{qn}(\tau)}d\mu$ converge weakly to $d(\psi_*(h\nu))$ as measures on $\overline{G}$. The upper bound in this case now follows from Lemma \ref{weakboundary}. If we add finitely many pure points outside $G$, we get the desired upper bound by placing a single zero at each $z_i$ and $\zeta_i$. More precisely, if we define the polynomials $y_{\infty}(z)$ and $\Upsilon_{\infty}(z)$ by \begin{align}\label{ydef} y_{\infty}(z)=\prod_{j=1}^m(z-z_j)\qquad,\qquad\Upsilon_{\infty}(z)=\prod_{j=1}^m(z-\zeta_j) \end{align} we have \[ \|P_n(\mu)\|^q_{\mu}\leq\|y_{\infty}\Upsilon_{\infty}P_{n-m-\ell}(|y_{\infty}(z)\Upsilon_{\infty}(z)|^q\mu)\|^q_{\mu}= \|P_{n-m-\ell}(|y_{\infty}(z)\Upsilon_{\infty}(z)|^q\mu)\|^q_{|y_{\infty}(z)\Upsilon_{\infty}(z)|^q\mu} \] and then proceed as in the case when $\ell=m=0$ and apply Proposition \ref{psiint}. For the lower bound, Lemma \ref{closetozone} implies that for each $z_i$, we can choose a sequence $\{w_{i,n}\}_{n\in\mathbb{N}}$ so that $P_n(w_{i,n};\mu)=0$ and $\lim_{n\rightarrow\infty}w_{i,n}=z_i$ (we will establish later that such a sequence has a unique tail, but we do not need this now). Define \begin{align}\label{whydefs} y_n(z)=\prod_{j=1}^m(z-w_{j,n}) \end{align} (so that $y_n(z)\rightarrow y_{\infty}(z)$ pointwise). We now can calculate \begin{align}\label{philow} \|P_n(z;\mu)\|^q_{\mu}\geq \int_{\rho}^1\int_0^{2\pi}\left|\frac{P_{n}(\psi(re^{i\theta}))}{y_n(\psi(re^{i\theta}))}\right|^q \prod_{j=1}^m|\psi(re^{i\theta})-w_{j,n}|^qh(re^{i\theta})d\nu_{ac}(\theta)d\tau(r) \end{align} For $|z|>1$ and $r\in[\rho,1]$, define the functions \[ S_{r,n}(z)=\exp\left(-\frac{1}{2q\pi}\int_0^{2\pi}\log\left(\prod_{j=1}^m|\psi(re^{i\theta})-w_{j,n}|^q h(re^{i\theta})\nu'(\theta)\right)\frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta\right). \] By our discussion in Section \ref{tools}, we can rewrite (\ref{philow}) as \[ \|P_n(z;\mu)\|^q_{\mu}\geq \int_{\rho}^1\int_0^{2\pi} \left|\frac{P_{n}(\psi(re^{i\theta}))}{e^{i(n-m)\theta}y_n(\psi(re^{i\theta}))}\right|^q|S_{r,n}(e^{i\theta})|^q\frac{d\theta}{2\pi} d\tau(r) \] (notice that we arbitrarily added a factor of $e^{-i(n-m)\theta}$ to the integrand, which is acceptable since it is inside the absolute value bars). For each fixed $r$, we invoke the subharmonicity of the integrand (or equation (\ref{subharm})) to obtain \begin{align}\label{lowermoment} \|P_n(z;\mu)\|^q_{\mu}\geq\int_{\rho}^1r^{qn-qm}|S_{r,n}(\infty)|^q d\tau(r). \end{align} Since $w_{j,n}$ converges to $z_j$ as $n\rightarrow\infty$ for each $j$ (by construction), we find that \[ \liminf_{n\rightarrow\infty}\frac{\|P_n(z;\mu)\|^q_{\mu}}{c_{qn}(\tau)}\geq \exp\left(\int_0^{2\pi}\log\left(h(e^{i\theta})\nu'(\theta))\right)\,\frac{d\theta}{2\pi}\right) \prod_{j=1}^m|\phi(z_j)|^q \] by Proposition \ref{psiint}. This is the desired lower bound. \begin{flushright} $\Box$ \end{flushright} \vspace{2mm} The proof of Theorem \ref{bigone} produces several interesting corollaries. The first of these shows that certain parts of the measure $\mu$ contribute only negligibly to the norm of the extremal polynomial. The following corollary is reminiscent of Theorem 2.4.1(vii) in \cite{OPUC1}. \begin{corollary}\label{singzero} If $\mu$ is as in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$ then \begin{align*} &\lim_{n\rightarrow\infty}\bigg(\int_{\overline{\bbD}}|p_n(\psi(re^{i\theta});\mu)|^q\,h(re^{i\theta})d\nu_{\textrm{sing}}(\theta)d\tau(r)+ \int_{\overline{G}}|p_n(z;\mu)|^qd\sigma_1(z)\,+\\ &\qquad\qquad\qquad+\int_{\overline{\bbD}}|p_n(\psi(re^{i\theta});\mu)|^q\,d\sigma_2(re^{i\theta})+ \sum_{j=1}^m\alpha_j|p_n(z_j;\mu)|^q+\sum_{j=1}^{\ell}\beta_j|p_n(\zeta_j;\mu)|^q\bigg)=0. \end{align*} \end{corollary} \begin{proof} Let us write $\mu=\mu^0+\mu^1$ where $\mu^0=\psi_*(h(\nu\otimes\tau))+\sum_{j=1}^m\alpha_j\delta_{z_j}$. Then \begin{align}\label{negl} \frac{\|P_n(\mu)\|^q_{\mu}}{c_{qn}(\tau)}=\frac{\|P_n(\mu)\|^q_{\mu^0}}{c_{qn}(\tau)}+ \frac{\|P_n(\mu)\|^q_{\mu^1}}{c_{qn}(\tau)}. \end{align} The proof of Theorem \ref{bigone} shows that the left hand side of (\ref{negl}) and the first term on the right hand side of (\ref{negl}) both converge to the right hand side of (\ref{bigconcl}). This shows that everything except $\mu^0$ contributes only negligibly to the norm of $p_n(z;\mu)$. To show that the pure points outside $\overline{G}$ contribute only negligibly to the norm, we keep our definition of $w_{1,n}$ from the proof of Theorem \ref{bigone} and we write $\mu^0=\mu^0_1+\alpha_1\delta_{z_1}$. We can now calculate \begin{align*} 1&\geq \frac{\int_{\mathbb{C}}\left|\frac{P_{n}(z;\mu)}{z-w_{1,n}}\right|^q|z-w_{1,n}|^qd\mu^0_1}{\|P_n(\mu)\|^q_{\mu}} +\alpha_1|p_n(z_1)|^q\\ &\geq\frac{\|P_{n-1}(|z-w_{1,n}|^q\mu^0_1)\|^q_{|z-w_{1,n}|^q\mu^0_1}}{\|P_n(\mu)\|^q_{\mu}} +\alpha_1|p_n(z_1)|^q\\ &=\frac{c_{qn}(\tau)\exp\left(\int_0^{2\pi}\log\left(h(e^{i\theta})\nu'(\theta)|\psi(e^{i\theta})-w_{1,n}|^q\right)\frac{d\theta}{2\pi} \right)\prod_{j=2}^m|\phi(z_j)|^q} {c_{qn}(\tau)\exp\left(\int_0^{2\pi}\log\left(h(e^{i\theta})\nu'(\theta)\right)\frac{d\theta}{2\pi}\right) \prod_{j=1}^m|\phi(z_j)|^q}+\\ &\qquad+\alpha_1|p_n(z_1)|^q+o(1)\\ &=1+o(1)+\alpha_1|p_n(z_1)|^q, \end{align*} which implies the desired conclusion for $z_1$. An identical proof works for each $z_j$ for $j=2,3,\ldots,m$. \end{proof} \noindent\textit{Remark.} As a consequence of Corollary \ref{singzero}, we see that if $K\subseteq G$ is compact, then \[ \int_K|p_n(z;\mu,q)|^qd\mu(z)\rightarrow0 \] as $n\rightarrow\infty$. \vspace{2mm} An additional consequence of Theorem \ref{bigone} is the following corollary, which is a refinement of Lemma \ref{closetozone}. \begin{corollary}\label{closertozone} Let $\mu$ be as in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$. There exists a $\delta>0$ and $N\in\mathbb{N}$ such that for all $n\geq N$, the polynomial $P_n(\mu)$ has a single zero in $\{u:|u-z_i|<\delta\}$ for each $i\leq m$. If we denote this zero by $w_{i,n}$, then there is an $a>0$ so that $|w_{i,n}-z_i|\leq e^{-an}$ for all large $n$. \end{corollary} \begin{proof} Lemma \ref{closetozone} establishes the existence of at least one zero of $P_n(\mu)$ in $\{u:|u-z_i|<\delta\}$ for all $i$ and all large $n$. Now, fix $\epsilon>0$ (but small) and let $\{w_1,\ldots,w_{t(n)}\}$ denote the collection of zeros of $P_n(\mu)$ outside $\Gamma_{1+\epsilon}$. Define for $|z|>1$ the functions \[ S_{r,n}(z)=\exp\left(-\frac{1}{2q\pi}\int_0^{2\pi}\log\left(\prod_{j=1}^{t(n)}|\psi(re^{i\theta})-w_{j}|^q h(re^{i\theta})\nu'(\theta)\right)\frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta\right). \] As in the proof of Theorem \ref{bigone}, we calculate \begin{align} \nonumber\frac{\|P_n(z;\mu)\|^q_{\mu}}{c_{qn}(\tau)}&\geq \frac{\int_{\rho}^1r^{qn-qt(n)}|S_{r,n}(\infty)|^q d\tau(r)}{c_{qn}(\tau)}\geq \frac{\int_{\rho}^1r^{qn-qt(n)}|S_{r,n}(\infty)|^q d\tau(r)}{c_{qn-qt(n)}(\tau)}\\ \label{attract}&=\frac{\int_{\rho}^1r^{qn-qt(n)}\exp\left(\frac{1}{2\pi}\int_0^{2\pi}\log\left( h(re^{i\theta})\nu'(\theta)\right)d\theta\right) d\tau(r)}{c_{qn-qt(n)}(\tau)}\prod_{j=1}^{t(n)}|\phi(w_j)|^q, \end{align} where we used Proposition \ref{psiint}. From this expression, it follows that $n-t(n)$ tends to infinity as $n\rightarrow\infty$, for if it did not, then since $|\phi(w_j)|>1+\epsilon$ for every $j\leq t(n)$, we would have $\|P_n(z;\mu)\|^{1/n}_{\mu}>1+\epsilon$ for all $n$ in some subsequence $\mathcal{N}\subseteq\mathbb{N}$, which violates the fact that $\mbox{cap}(\overline{G})=1$ and $\mu$ is regular (see Theorem III.3.1 in \cite{SaffTot}). Since $n-t(n)\rightarrow\infty$, the first factor in (\ref{attract}) converges to $\exp\left(\frac{1}{2\pi}\int_0^{2\pi}\log\left(h(e^{i\theta})\nu'(\theta)\right)d\theta\right)$ as $n\rightarrow\infty$ while the left hand side has limit given by the right hand side of (\ref{bigconcl}). If for each $i\in\{1,\ldots,m\}$ we pick a sequence $\{w_{i,n}\}_{n\in\mathbb{N}}$ as in the proof of Theorem \ref{bigone} then the corresponding factor in the product (\ref{attract}) converges to $|\phi(z_i)|^q$ as $n\rightarrow\infty$. Therefore, it must be that \[ \limsup_{n\rightarrow\infty}\prod_{j=1, w_j\neq w_{i,n}}^{t(n)}|\phi(w_j)|^q\leq1. \] However, each factor in this product is larger than $(1+\epsilon)^q$. We conclude that $t(n)=m$ for all sufficiently large $n$. This implies $P_n(\mu)$ has a single zero near each $z_j$ for $j=1,\ldots,m$ when $n$ is sufficiently large. The proof of the exponential attraction now proceeds exactly as in the last portion of the proof of Theorem 8.1.11 in \cite{OPUC1}. \end{proof} \noindent\textit{Remark.} Corollary \ref{closertozone} tells us that the polynomial $P_n(\mu,q)$ has a single zero extremely close to $z_i$ for each $i\in\{1,\ldots,m\}$ and the remaining $n-m$ zeros are placed so as to minimize the $L^q(\mu)$ norm with respect to a varying weight. It would be interesting to look at the measure $\mu$ on $\mathbb{D}\cup\{z_1,\ldots,z_m\}$ given by $d\mu=d^2z+\sum_{j=1}^m\delta_{z_j}$ (where $d^2z$ refers to area measure on the unit disk) and see if the results from \cite{MDPoly} continue to hold in this case, where the polynomial weight would be $y_{\infty}(z)$ (see (\ref{ydef}) above). \vspace{2mm} The upper bound in the proof of Theorem \ref{bigone} came from Lemma \ref{weakboundary}, which applies to arbitrary finite measures (not just product measures). We can also state the lower bound used in the proof of Theorem \ref{bigone} in a more general form. \begin{proposition}\label{genlow} Let $\tilde{\mu}$ be a measure on $\overline{G}$ so that $\tilde{\mu}\geq\mu$ and $\mu$ is the push-forward (via $\psi$) of the measure $w(re^{i\theta})\frac{d\theta}{2\pi} d\tau(r)$ where $1\in\textrm{supp}(\tau)$ and $w\in L^1(\frac{d\theta}{2\pi}\otimes d\tau(r))$. Then \[ \|P_n(\tilde{\mu})\|^q_{\tilde{\mu}}\geq\int_0^1r^{nq}\exp\left(\int_0^{2\pi}\log(w(re^{i\theta}))\frac{d\theta}{2\pi}\right)d\tau(r). \] \end{proposition} \noindent\textit{Remark.} The statement here is very general because we do not insist on any continuity of $w$. \begin{proof} By the inequality of the measures and the extremal property, we have \[ \|P_n(\tilde{\mu})\|^q_{\tilde{\mu}}\geq\|P_n(\tilde{\mu})\|^q_{\mu}\geq\|P_n(\mu)\|^q_{\mu}, \] so it suffices to put the desired bound on $\|P_n(\mu)\|^q_{\mu}$. Let $X\subseteq[0,1]$ be the collection of all $r$ so that $w(re^{i\theta})\frac{d\theta}{2\pi}$ is a Szeg\H{o} measure on $\partial\mathbb{D}$. The proposition is trivial unless $\tau(X)>0$. Therefore, we assume this is the case, and for $r\in X$ we define \[ S_r(z)= \exp\left(-\frac{1}{2q\pi}\int_0^{2\pi}\log\left(w(re^{i\theta})\right)\frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta\right) \quad,\quad|z|>1 \] and write \[ \|P_n(\mu)\|^q_{\mu}\geq\int_Xr^{nq}|S_r(\infty)|^qd\tau(r) \] as in (\ref{lowermoment}). This is the desired lower bound. \end{proof} We conclude this section with an example showing how one can apply Lemma \ref{weakboundary} to a region without analytic boundary. \vspace{2mm} \noindent\textbf{Example.} Let $G$ be the region $\left\{z:|z^3-1|<1\right\}$ and assume $q>1$. Notice that $G$ has capacity $1$ since $\phi(z)^3=z^3-1$ (see the example in Section 3 of \cite{MDFaber}). Define the polynomials $Q_n$ for $n$ a multiple of $3$ by $Q_{3m}(z)=(z^3-1)^m$. \begin{figure}[h!]\label{GDroq} \begin{center} \includegraphics[scale=.77]{GDrop.pdf} \caption{The region $G$ of the example.} \end{center} \end{figure} Let $\tau$ be a probability measure on $(0,1)$ with $1\in\textrm{supp}(\tau)$. The region $G$ can be decomposed into level sets $\Xi_r$ where \[ \Xi_r=\left\{z:|z^3-1|=r\right\} \] and $r$ runs from $0$ to $1$. Let $\nu_r$ be arc-length measure on each component of $\Xi_r$ and let $h(z)$ be a function that is continuous on $\overline{G}$ and is invariant under rotations by $\tfrac{2\pi}{3}$ so that $\phi_*(h\nu_1)$ has $\mathbb{Z}_3$ symmetry as a measure on $\partial\mathbb{D}$ as in Example 1.6.14 in \cite{OPUC1}. Let us define $\mu$ by \[ \int_{\overline{G}}f(z)d\mu(z)=\int_0^1\int_{\Xi_r}f(z)h(z)d\nu_r(z)d\tau(r). \] Consider the measure $h\nu_1$ on $\partial G$. If $m\in\mathbb{N}$ is fixed, then by the extremal property we have that for any choice of complex numbers $a_0,\ldots,a_{m-1}$ and $a_m=1$ \[ \|P_{3n}(h\nu_1,q)\|_{h\nu_1}^q\leq\left\|\sum_{j=0}^ma_jQ_{3(j+n-m)}(z)\right\|_{h\nu_1}^q=\left\|\sum_{j=0}^ma_j\phi(z)^{3(j+n-m)}\right\|_{h\nu_1}^q. \] Following the proof of the upper bound in Theorem 7.1 in \cite{Geronimus} we get \begin{align}\label{siv} \|P_{3n}(h\nu_1,q)\|_{h\nu_1}^q\leq \int_0^{2\pi}\left|1+\sum_{k=1}^m\gamma_ke^{3ki\theta}\right|^qd\phi_*(h\nu_1) \end{align} for any $m\leq n$ and any choice of constants $\gamma_1,\ldots,\gamma_m$. The assumed $\mathbb{Z}_3$ symmetry of the measure implies that $P_{3m}(z;\phi_*(h\nu_1),q)=R_m(z^3)$ for some monic polynomial $R_m$ of degree $m$ (this follows from the uniqueness of the extremal polynomial in the case $q>1$; see Example 1.6.14 in \cite{OPUC1}). Therefore, we can choose $\gamma_1,\ldots,\gamma_m$ appropriately so that the right hand side of (\ref{siv}) is equal to $\|P_{3m}(\phi_*(h\nu_1),q)\|_{\phi_*(h\nu_1)}^q$. The reasoning of Lemma \ref{weakboundary} then implies \[ \limsup_{n\rightarrow\infty}\|P_{3n}(h\nu_1,q)\|_{h\nu_1}^q\leq\exp\left(\int_0^{2\pi}\log\left(\phi_*(h\nu_1)'(\theta)\right)\frac{d\theta}{2\pi}\right). \] Now, as in Lemma \ref{weakboundary}, we calculate (for $f\in C(\overline{G})$) \begin{align*} c_{qm}(\tau)^{-1}\int_{\overline{G}}f(z)|Q_{3m}(z)|^qd\mu(z)&= c_{qm}(\tau)^{-1}\int_{0}^1\left(\int_{\Xi_r}f(z)h(z)d\nu_r(z)\right)r^{qm}d\tau(r)\\ &\rightarrow \int_{\Xi_1}f(z)h(z)d\nu_1(z) \end{align*} as $m\rightarrow\infty$. Therefore, the measures $\frac{|Q_{3m}|^q}{c_{qm}(\tau)}d\mu$ converge weakly to $hd\nu_1$ and the reasoning of Lemma \ref{weakboundary} implies \[ \limsup_{n\rightarrow\infty}\frac{\|P_{3n}(z;\mu,q)\|^q_{L^q(\mu)}}{c_{qn}(\tau)}\leq \exp\left(\int_0^{2\pi}\log(\phi_*(h\nu_1)'(\theta))\frac{d\theta}{2\pi}\right). \] \begin{flushright} $\Box$ \end{flushright} \vspace{2mm} In the next section, we explore more detailed asymptotic properties of the polynomials $P_n(z;\mu,q)$ and $p_n(z;\mu,q)$. \section{Strong Asymptotics for Extremal Polynomials}\label{strong} The main idea of Theorem \ref{bigone} is that the asymptotic behavior of the extremal polynomial norms is comparable to the behavior of the $L^q$ norms $\{\|\phi(z)^n\|_{L^q(\mu)}\}_{n\in\mathbb{N}}$. It should not be surprising then that in some cases we can make a stronger statement about the extremal polynomials' resemblance to $\phi(z)^n$ in certain regions of the plane, and this is what we call \textit{strong asymptotics}. Theorems \ref{szegconv} and \ref{boundarymeasure} will provide us with detailed information about the behavior of $P_n(z;\mu,q)$ outside of $\overline{G}$ and near the boundary of $G$. In Section \ref{christo} we will see how $P_n(z;\mu,2)$ behaves inside $G$ (see Corollary \ref{summable}). In the previous section we established that the polynomial $P_n(\mu,q)$ has a single zero near each pure point of $\mu$ outside of $\overline{G}$ (for large $n$) and asymptotically, all other zeros tend to $\overline{G}$. If we label the zero of $P_n(\mu,q)$ near $z_j$ as $w_{j,n,q}$, let us recall the definition \begin{align* y_n(z;q)=\prod_{j=1}^m(z-w_{j,n,q}), \end{align*} which can be uniquely defined for all sufficiently large $n$ by Corollary \ref{closertozone}. It will be convenient for us to define \begin{align}\label{lamden} \Lambda_n(z;\mu,q)=\frac{P_n(z;\mu,q)}{y_n(z;q)} \end{align} for all sufficiently large $n$. We also recall the definition \begin{align}\label{essen} S_{r,n}(z;q)=\exp \left(-\frac{1}{2q\pi}\int_0^{2\pi}\log\left(h(re^{i\theta})\nu'(\theta)|y_n(\psi(re^{i\theta});q)|^q\right) \frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta\right) \end{align} for $r\in[\rho,1]$ and $z\in\overline{\mathbb{C}}\setminus\overline{\bbD}$. We begin by considering the behavior of $P_n(z;\mu,q)$ when $z\not\in\overline{G}$ and $q>0$. We will prove a result reminiscent of the convergence result in Theorem 2.4.1(iv) in \cite{OPUC1} and the corollary in \cite{Koosis}. \begin{theorem}\label{szegconv} Let $S_{r,n}(z;q)$ be defined as in (\ref{essen}). If $\mu$ is a measure as in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$ and $q>0$ then \[ \frac{\Lambda_n(\psi(z);\mu,q)S_{1,\infty}(z;q)}{z^{n-m}S_{1,\infty}(\infty;q)}\rightarrow1 \] as $n\rightarrow\infty$ uniformly on compact subsets of $\overline{\mathbb{C}}\setminus\overline{\bbD}$. \end{theorem} \begin{proof} Let $q>0$ be fixed throughout this proof and denote $S_{r,n}(z;q)$ by $S_{r,n}(z)$ and $\Lambda_n(z;\mu,q)$ by $\Lambda_n(z;\mu)$. We showed in Section \ref{product} (see equation (\ref{subharm})) that if $r\in[\rho,1]$ then \begin{align}\label{schwint} r^{q(n-m)}S_{r,n}(\infty)^q\leq \int_0^{2\pi}|\Lambda_n(\psi(re^{i\theta});\mu)S_{r,n}(e^{i\theta})|^q\frac{d\theta}{2\pi} \end{align} for all $r\in[\rho,1]$. Let us fix some $t<1$. If we divide both sides of (\ref{schwint}) by $c_{qn}(\tau)$ and then integrate in the variable $r$ from $t$ to $1$ with respect to $\tau$, then both sides converge to $S_{1,\infty}(\infty)^q$ as $n\rightarrow\infty$ (by Theorem \ref{bigone}). Therefore, (\ref{schwint}) is optimal in that we cannot multiply the right hand side by a factor smaller than $1$ and have the inequality remain valid for all $r\in[t,1]$ when $n$ is sufficiently large. It follows that for any $\epsilon>0$, there exists a sequence $\{r_n\}_{n=1}^{\infty}$ converging to $1$ from below as $n\rightarrow\infty$ so that \begin{align}\label{rbound} r_n^{q(n-m)}S_{r_n,n}(\infty)^q\geq (1-\epsilon)\int_0^{2\pi}|\Lambda_n(\psi(r_ne^{i\theta});\mu)S_{r_n,n}(e^{i\theta})|^q\frac{d\theta}{2\pi}. \end{align} By a standard argument, we can choose our sequence $\{r_n\}_{n=1}^{\infty}$ converging to $1$ from below so that (\ref{rbound}) remains true for some sequence $\epsilon_n$ tending monotonically to $0$ from above. Let $a_n:=\|\Lambda_n(\psi(r_ne^{i\theta});\mu)S_{r_n,n}(e^{i\theta})\|_{L^q(\frac{d\theta}{2\pi})}$. By using (\ref{schwint}) and (\ref{rbound}) we see that \begin{align}\label{infvalu} 1-\epsilon_n\leq \left|\lim_{z\rightarrow\infty}\frac{\Lambda_n(\psi(r_nz);\mu)S_{r_n,n}(z)}{a_nz^{n-m}}\right|^q \leq1. \end{align} Let \[ f_n(z)=\frac{\Lambda_n(\psi(r_nz);\mu)S_{r_n,n}(z)} {a_nz^{n-m}}. \] Clearly $\|f_n\|_{H^q(\overline{\bbC}\setminus\overline{\bbD})}=1$ for all $n$ so we can find a subsequence $\mathcal{N}\subseteq\mathbb{N}$ so that $f_n$ converges uniformly on compact subsets of $\overline{\bbC}\setminus\overline{\bbD}$ to some (analytic) function $\tilde{f}$ (by a normal families argument and Lemma 1.1 in \cite{Kalicurve}). Equation (\ref{infvalu}) shows that $\tilde{f}(\infty)=1$ while $\|\tilde{f}\|_{H^q(\overline{\bbC}\setminus\overline{\bbD})}\leq\limsup_{n\in\mathcal{N}}\|f_n\|_{H^q(\overline{\bbC}\setminus\overline{\bbD})}=1$ (see Lemma 1.2 in \cite{Kalicurve}). However, clearly $\|\tilde{f}\|_{H^q(\overline{\bbC}\setminus\overline{\bbD})}\geq1$ since $\tilde{f}(\infty)=1$. This means $\int_0^{2\pi}|\tilde{f}(re^{i\theta})|^q\frac{d\theta}{2\pi}=1$ for all $r\geq1$. Furthermore, the proof of Corollary \ref{closertozone} and the Hurwitz Theorem imply $\tilde{f}$ is non-vanishing in $\overline{\bbC}\setminus\overline{\bbD}$ and so $f^q$ is an analytic function on this domain. This implies \[ \textrm{Re}\left[\int_0^{2\pi}|\tilde{f}(re^{i\theta})|^q-\tilde{f}(re^{i\theta})^q\frac{d\theta}{2\pi}\right]=0\qquad,\qquad r>1 \] and it follows easily that $\tilde{f}\equiv1$. The same argument applies to any subsequence of $\{f_n\}_{n\in\mathbb{N}}$ so $f_n$ converges to $1$ uniformly on compact subsets of $\overline{\bbC}\setminus\overline{\bbD}$. Equations (\ref{schwint}) and (\ref{rbound}) show that $a_n=(1+\delta_n)r_n^{(n-m)}S_{r_n,n}(\infty)$ with $\delta_n\rightarrow0$ as $n\rightarrow\infty$. Therefore, if $|z|>1$, we have \begin{align}\label{exterior} \frac{\Lambda_n(\psi(r_nz))S_{r_n,n}(z)}{(r_nz)^{(n-m)}S_{r_n,n}(\infty)}\rightarrow1 \end{align} and the convergence is uniform on compact subsets of $\overline{\mathbb{C}}\setminus\overline{\bbD}$. Now choose $w\not\in\overline{\bbD}$. Notice that the function $H(z)=(e^{i\theta}+z)(e^{i\theta}-z)^{-1}$ is a conformal map from $\overline{\bbC}\setminus\overline{\bbD}$ to the left half-plane. Therefore, $H(z/r_n)$ converges to $H(z)$ as $n\rightarrow\infty$ and the convergence is uniform on compact subsets of $\overline{\bbC}\setminus\overline{\bbD}$ and uniform in $\theta$. This observation and Dominated Convergence imply \begin{align}\label{sconv} \frac{S_{r_n,n}(w/r_n)S_{1,\infty}(\infty)}{S_{r_n,n}(\infty)S_{1,\infty}(w)}\rightarrow1 \end{align} as $n\rightarrow\infty$ and the convergence is uniform on compact subsets of $\overline{\bbC}\setminus\overline{\bbD}$. By plugging in $z=w/r_n$ into (\ref{exterior}) and using the uniformity of convergence on compact subsets we recover \[ \frac{\Lambda_n(\psi(w))S_{1,\infty}(w)}{w^{n-m}S_{1,\infty}(\infty)}\rightarrow1, \] which proves convergence. To prove the uniformity, let $K\subseteq\overline{\bbC}\setminus\overline{\bbD}$ be a compact set. There is a compact set $K_1\subseteq\overline{\bbC}\setminus\overline{\bbD}$ such that for all sufficiently large $n$, every $x\in K$ can be written as $r_nx_1^{(n)}$ for some $x_1^{(n)}\in K_1$. Then \begin{align}\label{unif} \frac{\Lambda_n(\psi(w))S_{1,\infty}(w)}{w^{n-m}S_{1,\infty}(\infty)}&= \frac{\Lambda_n(\psi(r_nw_1^{(n)}))S_{r_n,n}(w_1^{(n)})}{(r_nw_1^{(n)})^{n-m}S_{r_n,n}(\infty)} \cdot\frac{S_{1,\infty}(w)S_{r_n,n}(\infty)}{S_{r_n,n}(w_1^{(n)})S_{1,\infty}(\infty)}. \end{align} We have already shown in (\ref{exterior}) that the first factor in (\ref{unif}) tends to $1$ uniformly on compact subsets as $n\rightarrow\infty$. Similarly, we have shown in (\ref{sconv}) that the second factor in (\ref{unif}) tends to $1$ uniformly on compact subsets as $n\rightarrow\infty$. \end{proof} \noindent\textit{Remark.} One can in fact conclude that in the proof of Theorem \ref{szegconv}, the functions $f_n$ converge to $1$ in $H^q(\overline{\bbC}\setminus\overline{\bbD})$ (see Theorem 1 in \cite{Keldysh}). \vspace{2mm} Theorem \ref{szegconv} yields the following Corollary, which says that the strong asymptotic behavior of the polynomials $P_n(\mu,q)$ is in some sense independent of $\tau$. \begin{corollary}\label{stoutlim} Let $\mu$ be as in Theorem \ref{szegconv} and let $\kappa$ be the measure on $\partial G$ given by $|y_{\infty}|^2\psi_*(h\nu)$. Uniformly on compact subsets of $\overline{\mathbb{C}}\setminus\overline{G}$ we have \[ \lim_{n\rightarrow\infty}\frac{\Lambda_n(z;\mu,q)}{P_{n-m}(z;\kappa,q)}=1. \] \end{corollary} \begin{proof} By Theorem \ref{szegconv} above and Theorem 2.1 in \cite{Kalicurve}, both $\frac{\Lambda_n(z;\mu,q)}{\phi(z)^{n-m}}$ and $\frac{P_{n-m}(z;\kappa,p)}{\phi(z)^{n-m}}$ converge to $S_{1,\infty}(\infty)S_{1,\infty}(\phi(z))^{-1}$ uniformly on compact subsets of $\overline{\mathbb{C}}\setminus\overline{G}$ so the claim follows. \end{proof} Now that we have some information about the behavior of $P_n(z;\mu,q)$ outside $\overline{G}$, we will consider what happens close to the boundary of $G$. Our next result is motivated in part by Theorem 9.3.1 in \cite{OPUC2}. As in Theorem \ref{szegconv}, we will consider all $q>0$. \begin{theorem}\label{boundarymeasure} If $\mu$ is as in Theorem \ref{bigone}, $q>0$, and $\nu$ is a Szeg\H{o} measure on $\partial\mathbb{D}$ then \[ w\mbox{-}\lim_{n\rightarrow\infty}|p_n(z;\mu,q)|^qd\mu(z)=d\omega_{\overline{G}}(z) \] as measures on $\mathbb{C}$. \end{theorem} \begin{proof} Let $q>0$ be fixed and denote by $p_n$ the polynomial $p_n(z;\mu,q)$. Corollary \ref{singzero} and the remark following it imply that any weak limit of the measures $\{|p_n|^qd\mu\}_{n\in\mathbb{N}}$ must be a measure on $\partial G$ and that we may without loss of generality assume that $\sigma_1=\sigma_2=0$, $\ell=0$, and $\nu$ is purely absolutely continuous with respect to Lebesgue measure. Let us recall the definition of $S_{r,n}(z)=S_{r,n}(z;q)$ from (\ref{essen}) for $r\in[\rho,1]$ and $z\in\overline{\mathbb{C}}\setminus\overline{\bbD}$. We showed in Theorem \ref{bigone} that \begin{align}\label{ratione} \int_{\rho}^1\int_0^{2\pi}\frac{\left|e^{-i(n-m)\theta}\Lambda_n(\psi(re^{i\theta}))\right|^q |S_{r,n}(e^{i\theta})|^q}{c_{qn}(\tau)|S_{1,n}(\infty)|^q}\frac{d\theta}{2\pi} d\tau(r)\rightarrow1 \end{align} as $n\rightarrow\infty$. For fixed $n\in\mathbb{N}$ and $r\in[\rho,1]$, let $\{u_1,\ldots,u_{\eta_n(r)}\}$ be the zeros of $\Lambda_n(\psi(rz))$ ($\eta_n\in\mathbb{N}_0$) lying outside of $\overline{\bbD}$, each listed a number of times equal to its multiplicity as a zero. We may then define the Blaschke product \[ B_{r,n}(z)=\prod_{j=1}^{\eta_n(r)}\frac{z-u_j}{z\bar{u}_j-1}\cdot\frac{\bar{u}_j}{|u_j|}. \] With this notation, we may define $J_{r,n}(z)$ so that \begin{align}\label{atinf} z^{-(n-m)}\Lambda_n(\psi(rz))S_{r,n}(z)=B_{r,n}(z)J_{r,n}(z). \end{align} From (\ref{atinf}), we know that $J_{r,n}(z)$ is analytic and non-vanishing in $\overline{\bbC}\setminus\overline{\bbD}$ so we may write \begin{align}\label{conatinf} J_{r,n}(z)^{q/2}=J_{r,n}(\infty)^{q/2}+g_{r,n}(z) =\left(\frac{r^{n-m}S_{r,n}(\infty)}{B_{r,n}(\infty)}\right)^{q/2}+g_{r,n}(z), \end{align} where $g_{r,n}(z)$ is in $\mathbb{H}^2(\overline{\bbC}\setminus\overline{\bbD})$ and is orthogonal to the constant functions in $\mathbb{H}^2(\overline{\bbC}\setminus\overline{\bbD})$ (that is, $g_{r,n}(z)\in\mathbb{H}^2_0(\overline{\bbC}\setminus\overline{\bbD})$ in the notation of \cite{Duren}). Notice that \[ |e^{-i(n-m)\theta}\Lambda_n(\psi(re^{i\theta}))S_{r,n}(e^{i\theta})|^q=|J_{r,n}(e^{i\theta})^{q/2}|^2. \] If we plug (\ref{conatinf}) into (\ref{ratione}), we get \begin{align}\label{gneg} \int_{\rho}^1\frac{r^{qn-qm}|S_{r,n}(\infty)|^q}{c_{qn}(\tau)|S_{1,n}(\infty)|^qB_{r,n}(\infty)^q}+ \frac{\|g_{r,n}\|^2_{\mathbb{H}^2}}{c_{qn}(\tau)|S_{1,n}(\infty)|^q}d\tau(r) \rightarrow1 \end{align} as $n\rightarrow\infty$. However, $B_{r,n}(\infty)^{-q}>1$ and the second term is always non-negative, so we conclude that the first term in (\ref{gneg}) has integral tending to $1$ as $n\rightarrow\infty$ and hence \begin{align}\label{tozero} \int_{\rho}^1\frac{\|g_{r,n}\|^2_{\mathbb{H}^2}}{c_{qn}(\tau)|S_{1,n}(\infty)|^q}d\tau(r)\rightarrow0 \end{align} as $n\rightarrow\infty$. Now fix $k\in\mathbb{N}$. We have \begin{align} &\label{momint} \int_{G_{\rho}}\phi(z)^k|p_n(z)|^qd\mu(z)\\ \nonumber&\qquad=\int_{\rho}^1\int_0^{2\pi} \frac{r^ke^{ik\theta}|e^{-i(n-m)\theta}\Lambda_n(\psi(re^{i\theta}))|^q|S_{r,n}(e^{i\theta})|^q} {c_{qn}(\tau)|S_{1,n}(\infty)|^q}\frac{d\theta}{2\pi} d\tau(r)+o(1)\\ \nonumber&\qquad=\int_{\rho}^1\int_0^{2\pi}\frac{r^ke^{ik\theta}|r^{q(n-m)/2}S_{r,n}(\infty)^{q/2} B_{r,n}(\infty)^{-q/2}+g_{r,n}(e^{i\theta})|^2} {c_{qn}(\tau)|S_{1,n}(\infty)|^q} \frac{d\theta}{2\pi} d\tau(r)+o(1)\\ &\label{htwoproof}\qquad=\int_{\rho}^1\int_0^{2\pi}\frac{r^{k+q(n-m)}e^{ik\theta}|S_{r,n}(\infty)|^q} {c_{qn}(\tau)|S_{1,n}(\infty)|^qB_{r,n}(\infty)^{q}}\frac{d\theta}{2\pi} d\tau(r) +\int_{\rho}^1\int_0^{2\pi}\frac{r^{k}e^{ik\theta}|g_{r,n}(e^{i\theta})|^2} {c_{qn}(\tau)|S_{1,n}(\infty)|^q} \frac{d\theta}{2\pi} d\tau(r)\\ \nonumber&\qquad\qquad+\int_{\rho}^1\int_0^{2\pi}\frac{r^{k+q(n-m)/2}e^{ik\theta}S_{r,n}(\infty)^{q/2} \cdot2\textrm{Re}[g_{r,n}(e^{i\theta})]}{c_{qn}(\tau)|S_{1,n}(\infty)|^qB_{r,n}(\infty)^{q/2}} \frac{d\theta}{2\pi} d\tau(r)+o(1) \end{align} as $n\rightarrow\infty$. If we send $n$ to infinity, the first term in (\ref{htwoproof}) converges to $0$ since $k\in\mathbb{N}$. The second term in (\ref{htwoproof}) can be bounded from above in absolute value by \begin{align}\label{twothreebound} \int_{\rho}^1\frac{\|g_{r,n}\|^2_{\mathbb{H}^2}}{c_{qn}(\tau)|S_{1,n}(\infty)|^q}d\tau(r), \end{align} which tends to $0$ by (\ref{tozero}). By applying the Schwartz inequality, the third term in (\ref{htwoproof}) can be bounded from above in absolute value by \[ \left(\int_{\rho}^1\frac{r^{2k+q(n-m)}S_{r,n}(\infty)^q}{c_{qn}(\tau)S_{1,n}(\infty)^qB_{r,n}(\infty)^q}d\tau(r)\right)^{1/2} \left(\int_{\rho}^1\frac{4|\int_0^{2\pi}e^{ik\theta}\textrm{Re}[g_{r,n}(e^{i\theta})]\frac{d\theta}{2\pi}|^2} {c_{qn}(\tau)S_{1,n}(\infty)^q}d\tau(r)\right)^{1/2} \] The first factor tends to $1$ as $n\rightarrow\infty$ (as in (\ref{gneg})). After applying Jensen's inequality to the second factor, we can bound it from above by twice the square root of (\ref{twothreebound}). Therefore the integral (\ref{momint}) tends to $0$ as $n\rightarrow\infty$. We conclude that if $\gamma$ is a weak limit point of the measures $\{|p_n(\mu)|^qd\mu\}_{n\in\mathbb{N}}$, then for every $k\in\mathbb{N}$ we have \[ \int_{\partial G}\phi(z)^kd\gamma=0. \] This implies that $\gamma$ is induced (via $\psi$) by a measure $\kappa$ on $\partial\mathbb{D}$ with no non-trivial moments, i.e. $d\kappa=\frac{d\theta}{2\pi}$ and it follows that $\gamma$ is the equilibrium measure for $\overline{G}$ (see Theorem 3.1 in \cite{ToTrans}). \end{proof} Theorem \ref{boundarymeasure} yields the following corollary, which can be interpreted in terms of the Christoffel functions discussed in Section \ref{christo} (see (\ref{lambdn}) below). \begin{corollary}\label{kernelmeasure} Under the hypotheses of Theorem \ref{boundarymeasure}, we have \[ w\mbox{-}\lim_{n\rightarrow\infty}\frac{K_n(z,z)}{n+1}d\mu(z)=d\omega_{\overline{G}} \] as measures on $\mathbb{C}$ where $K_n(z,w)=\sum_{j=0}^np_j(z;\mu,2)\overline{p_j(w;\mu,2)}$ is the reproducing kernel for the measure $\mu$ and polynomials of degree at most $n$. \end{corollary} \noindent\textit{Remark.} Since $\mu$ is regular, one can use a polynomial approximation argument, Corollary \ref{singzero}, and the results in \cite{WeakCD} to arrive at a different proof of Corollary \ref{kernelmeasure}. Theorem \ref{boundarymeasure} is of course much stronger. \vspace{2mm} In the next section, we will consider the behavior of the Christoffel functions on $\overline{G}$. \section{Christoffel Functions}\label{christo} In this section we will turn our attention to an interesting minimization problem. For each $n\in\mathbb{N}$ and $q>0$, let us define the \textit{Christoffel function} $\lambda_n(z;\mu,q)$ by \begin{align* \lambda_n(z;\mu,q)=\inf\left\{\int_{\mathbb{C}}|Q(w)|^qd\mu(w):Q(z)=1\,,\,\textrm{deg}(Q)\leq n\right\}. \end{align*} For $z\in\mathbb{C}$ fixed, $\lambda_n(z;\mu,q)$ is obviously non-increasing (as $n\rightarrow\infty$) and positive, so we may define $\lambda_{\infty}(z;\mu,q)=\lim_{n\rightarrow\infty}\lambda_n(z;\mu,q)$. It is clear that \[ \lambda_{\infty}(z;\mu,q)=\inf\left\{\int_{\mathbb{C}}|Q(w)|^q\,d\mu(w): \, Q(z)=1\, ,\,Q \mbox{ polynomial}\right\}. \] The behavior of $\lambda_{\infty}(z;\mu,q)$ is particularly easy to describe when $z\in\partial G$. \begin{proposition}\label{christoboundary} If $\mu$ is any measure with support in $\overline{G}$ and $G$ has analytic boundary then $\lambda_{\infty}(x;\mu,q)=\mu(\{x\})$ for all $x\in\partial G$ and all $q>0$. \end{proposition} \noindent\textit{Remark.} For Proposition \ref{christoboundary}, we do not need to assume $\mbox{cap}(\overline{G})=1$. \begin{proof} Fix $x\in\partial G$. It is obvious that $\lambda_n(x;\mu,q)\geq\mu(\{x\})$ for every $n\in\mathbb{N}$, so it remains to show the reverse inequality holds in the limit. Since $\partial G$ is analytic, we can define a conformal map $\varphi:G\rightarrow\mathbb{D}$ satisfying $\varphi(x)=1$. By a well-known argument, this map $\varphi$ has an analytic continuation to some open set $U\supseteq\overline{G}$. Define \[ f_n(z):=3^{-n}(\varphi(z)+2)^n\qquad,\qquad z\in U \] so that $f_n(x)=1=\|f_n\|_{L^{\infty}(\overline{G})}$. By Theorem 2.5.7 in \cite{StSh} there exists a sequence of polynomials $\{W_n\}_{n\in\mathbb{N}}$ so that $\|W_n-f_n\|_{L^{\infty}(\overline{G})}<n^{-1}$ (we do not assume $W_n$ has degree $n$). It follows that for each $n\in\mathbb{N}$ there is a constant $a_n=1+o(1)$ (as $n\rightarrow\infty$) so that $a_nW_n(x)=1$. Then (with $E_n=W_n-f_n$) \[ \lambda_n(x;\mu,q)\leq\int_{\overline{G}}|a_nW_n(z)|^qd\mu(z)=(1+o(1))\int_{\overline{G}}|f_n(z)+E_n|^qd\mu(z) \rightarrow\mu(\{x\}) \] by Dominated Convergence. \end{proof} \noindent\textit{Remark.} For results producing more precise asymptotics of $\lambda_n(z;\mu,2)$ for $z\in\partial G$ under stronger hypotheses on $\mu$, see \cite{LubBerg,ToTrans}. \vspace{2mm} Now let us focus on $x\in G$. For measures supported on the unit circle, it is known (see Theorem 2.5.4 in \cite{OPUC1}) that if $\nu$ is a Szeg\H{o} measure then $\lambda_{\infty}(z;\nu,q)>0$ for all $z\in\mathbb{D}$ and $q\in(0,\infty)$. We will prove an analog for the kinds of measures considered in Theorem \ref{bigone}. Before we can do this, we need to define some auxiliary notation. For $x$ interior to $\Gamma_{1}$, define \[ \xi(x)=\frac{1}{2}\left(1+\inf\{r:x\in\mathcal{G}_r\,,\, r\geq\rho\}\right). \] For each $r\in[\xi(x),1]$, let $\chi_{r,x}$ be the conformal map from $\mathbb{D}$ to $\mathcal{G}_r$ that sends $0$ to $x$ and satisfies $\chi_{r,x}'(0)>0$. Denote the inverse to $\chi_{r,x}$ by $\varphi_{r,x}$. The following lemma will be useful: \begin{lemma}\label{varphideriv} With the above notation, it holds that $\varphi_{r,x}$ converges to $\varphi_{1,x}$ uniformly on some open set containing $\overline{G}$ as $r\rightarrow1$ and there is an $s\in(\xi(x),1)$ and positive constants $\lambda_1$ and $\lambda_2$ such that \[ \lambda_1<|\varphi_{r,x}'(z)|<\lambda_2 \] for all $r\in[s,1]$ and $z\in\overline{G}$. \end{lemma} \noindent\textit{Remark.} The proof of the lemma will actually show that when $r$ is sufficiently close to $1$, $\varphi_{r,x}$ is defined on all of $\overline{G}$ so the statement of the lemma makes sense. \begin{proof} By the Carath\'{e}odory Convergence Theorem (see Theorem 3.1 in \cite{Dur}), the maps $\varphi_{r,x}$ converge to $\varphi_{1,x}$ uniformly on compact subsets of $G$ as $r\rightarrow1^-$ (see also Theorem 3 in \cite{SnipWard}). Since $G$ has analytic boundary, a simple argument shows that each $\varphi_{r,x}$ can be univalently continued outside of $\overline{G}$ when $r$ is sufficiently close to $1$ and in fact all such $\varphi_{r,x}$ have a common domain of holomorphy containing $\overline{G}$. A normal families argument then implies $\varphi_{r,x}$ converges to $\varphi_{1,x}$ uniformly on some open set containing $\overline{G}$ as $r\rightarrow1$. We can then use the Cauchy integral formula to conclude that $\varphi_{r,x}'$ converges to $\varphi_{1,x}'$ on a smaller open set containing $\overline{G}$. This means that when $r$ is sufficiently close to $1$, we have $\|\varphi_{r,x}'\|_{L^{\infty}(\Gamma_1)}\leq2\|\varphi_{1,x}'\|_{L^{\infty}(\Gamma_1)}$. The same arguments can be applied to $\{\chi_{r,x}\}_{r\in[\xi(x),1]}$, which proves the claim. \end{proof} As a final preparatory step, we will need the following lemma, which is a slight refinement of Lemma 1.1 in \cite{Kalicurve}. \begin{lemma}\label{hpbound} If $q\in(0,\infty)$ and $w\in\mathcal{G}_r$ then there is a constant $\beta_w$ so that for every $f\in H^q(\mathcal{G}_r)$, \[ |f(w)|^q\leq\beta_w\int_{\Gamma_r}|f(z)|^qd|z|. \] Furthermore, the constant $\beta_w$ may be taken uniform for all $r$ sufficiently close to $1$ (but perhaps depending on $w$). \end{lemma} \begin{proof} The inequality follows from Lemma 1.1 in \cite{Kalicurve} and the equivalence of the spaces $E^q(\mathcal{G}_r)$ and $H^q(\mathcal{G}_r)$ (see Chapter 10 in \cite{Duren}), so we need only focus on the uniformity. If $q\geq1$, then this is a simple consequence of Jensen's inequality and the fact that $H^1$ functions are the Cauchy integral of their boundary values (see Theorem 10.4 in \cite{Duren}), so we need only focus on the case $0<q<1$. To this end, let $g$ be the function harmonic in $\mathcal{G}_r$ satisfying $g(\psi(re^{i\theta}))=|f(\psi(re^{i\theta}))|^q$ almost everywhere on $\Gamma_r$. Let $\omega_{r,w}$ by the harmonic measure for the region $\mathcal{G}_r$ and the point $w$. Then by the subharmonicity of $f$, we have \begin{align*} |f(w)|^q\leq g(w)=\int_{\Gamma_r}g(z)d\omega_{r,w}(z)&\leq \left\|\frac{d\omega_{r,w}}{d|z|}\right\|_{L^{\infty}(\Gamma_r)}\int_{\Gamma_r}g(z)d|z|\\ &=\left\|\varphi_{r,w}'\right\|_{L^{\infty}(\Gamma_r)}\int_{\Gamma_r}|f(z)|^qd|z|. \end{align*} We can now apply Lemma \ref{varphideriv} with $x=w$ to provide uniformity in the constant $\beta_w$. \end{proof} Now we are ready to prove the main theorem of this section. \begin{theorem}\label{pbigo} If $\mu$ and $G$ are as in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$, then $\lambda_{\infty}(z;\mu,q)>0$ for all $z\in G$ and $q\in(0,\infty)$. \end{theorem} \begin{proof} Since $h$ is bounded from below and $\lambda_n(z;\mu,q)$ increases as we increase $\mu$, we may assume that $\mu=\nu_{ac}\otimes\tau$. In the region $G_{\rho}$ we may write (for $f$ continuous) \begin{align}\label{moveto} \int_{G_{\rho}}f(z)d\mu(z)= \int_{\rho}^1\int_{\Gamma_t}f(z)\tilde{w}(z)d|z| d\tau(t) \end{align} where $\tilde{w}$ is a weight on $G_{\rho}$. In fact, we can write explicitly \begin{align}\label{proofofszego} \tilde{w}(z)=\frac{1}{2\pi}\cdot\nu'\left(\frac{\phi(z)}{|\phi(z)|}\right)\frac{|\phi'(z)|}{|\phi(z)|} \end{align} (we identify $\nu'(e^{i\theta})$ and $\nu'(\theta)$). As in \cite{MDCurve}, define $\Delta_{r,q}(z)$ by \begin{align}\label{szegrpdef} \Delta_{r,q}(z)=\exp\left(\frac{1}{2q\pi i}\oint_{\Gamma_r}\log\left(\tilde{w}(\zeta)\right)\frac{1+\overline{\varphi_r(\zeta)} \varphi_r(z)}{\varphi_r(\zeta)-\varphi_r(z)}\varphi_r'(\zeta)d\zeta\right) \end{align} for each $r\in[\rho,1]$ so that $|\Delta_{r,q}(\zeta))|^q=\tilde{w}(\zeta)$ for almost every $\zeta\in\Gamma_r$ ((\ref{proofofszego}) implies the integral in (\ref{szegrpdef}) converges). Now fix $y\in G$ and let $Q(z)$ be any polynomial so that $Q(y)=1$ (we make no assumptions on the degree of $Q$). Let $s\in(\rho,1)$ be so that $y$ is interior to $\Gamma_s$ and so the constant $\beta_y$ of Lemma \ref{hpbound} may be chosen independently of $t\in[s,1]$. We calculate \begin{align}\label{pdelta} \|Q\|^q_{L^q(\mu)}\geq\int_{s}^1\int_{\Gamma_t}|Q(z)\Delta_{t,q}(z)|^q\,d|z|d\tau(t) \geq \beta_y^{-1}\int_s^1|\Delta_{t,q}(y)|^qd\tau(t) \end{align} by Lemma \ref{hpbound}. The function $\Delta_{t,q}(y)$ is expressed as an exponential so the fact that $\nu$ is a Szeg\H{o} measure on $\partial\mathbb{D}$ implies $\Delta_{t,q}(y)$ is never equal to $0$ for any $t$. Therefore, $|\Delta_{t,q}(y)|^q$ is not the zero function and so the integral on the far right on (\ref{pdelta}) is not equal to zero. We have therefore obtained a lower bound for the far left hand side of (\ref{pdelta}) that is independent of the degree of $Q$. Taking the infimum over all such $Q$ proves the theorem. \end{proof} Recall the definition of $K_n(z,w)$ from Corollary \ref{kernelmeasure}. By equation (2.16.6) in \cite{Rice}, one has \begin{align}\label{lambdn} \lambda_n(z;\mu,2)=\frac{1}{K_n(z,z;\mu)}. \end{align} This and Theorem \ref{pbigo} for the case $q=2$ yields a proof of the following corollary: \begin{corollary}\label{summable} If $\tilde{\mu}\geq\mu$ and $\mu$ is as in Theorem \ref{bigone} with $\nu$ a Szeg\H{o} measure on $\partial\mathbb{D}$ then \[ \sum_{n=0}^{\infty}|p_n(z;\tilde{\mu},2)|^2<\infty \] for all $z\in G$. \end{corollary} Now that we have some understanding of $\lambda_{\infty}(x;\mu,q)$ for all $x\in G$ when $\mu$ is of the form considered in Theorem \ref{bigone}, we want to try to calculate it exactly. Our next result will show that one can reduce the problem to considering only measures on $G=\mathbb{D}$ and only the point $x=0$. Indeed, take any $x_0\in G$ and let $\varphi$ be the conformal map of $G$ to $\mathbb{D}$ sending $x_0$ to $0$ and satisfying $\varphi'(x_0)>0$. By the injectivity of $\varphi$ on $\overline{G}$ (we used Carath\'{e}odory's Theorem here; see Theorem I.3.1 in \cite{GarnMar}), we can push any measure $\mu$ on $\overline{G}$ forward via $\varphi$ to get a measure $\varphi_*\mu$ on $\overline{\bbD}$ as in Section \ref{intro}. With this notation, we can prove the following result: \begin{proposition}\label{into} With $x_0$, $\mu$ and $\varphi$ as above, we have $\lambda_{\infty}(x_0;\mu,q)=\lambda_{\infty}(0;\varphi_*\mu,q)$ for all $q\in(0,\infty)$. \end{proposition} \noindent\textit{Remark.} We do not exclude the possibility that $G=\mathbb{D}$ and $\varphi$ is an automorphism of the disk. \vspace{2mm} \noindent\textit{Remark.} If $\tau\neq\delta_1$, the resulting measure $\varphi_*\mu$ may not be of the form considered in Theorem \ref{bigone}. \vspace{2mm} \begin{proof} Fix $q\in(0,\infty)$. Given $\epsilon>0$, let $T$ be a polynomial so that $\|T\|^q_{L^q(\varphi_*\mu)}<\lambda_{\infty}(0;\varphi_*\mu,q)+\epsilon$ and $T(0)=1$. Then $\tilde{Q}:=T\circ\varphi$ is a function on $\overline{G}$ satisfying $\|\tilde{Q}\|^q_{L^q(\mu)}=\|T\|^q_{L^q(\varphi_*\mu)}$ and $\tilde{Q}(x_0)=1$. Now let $Q$ be a polynomial satisfying $\||Q|^q-|\tilde{Q}|^q\|_{L^{\infty}(\overline{G})}<\epsilon$ and $Q(x_0)=1$ (such a $Q$ exists by the same reasoning as in the proof of Proposition \ref{christoboundary}). It follows at once that $\lambda_{\infty}(x_0;\mu,q)\leq\lambda_{\infty}(0;\varphi_*\mu,q)+2\epsilon$ and one direction of the inequality follows by sending $\epsilon\rightarrow0$. The reverse inequality follows by an argument symmetric to the one just given. \end{proof} \noindent\textit{Remark.} If we set $\tau=\delta_1$, Proposition \ref{into} can be used to provide a new proof of Proposition 2.2.2 in \cite{OPUC1} and a new proof of Theorem 2.5.4 in \cite{OPUC1}. \vspace{2mm} Proposition \ref{into} allows us to calculate $\lambda_{\infty}(x;\mu,q)$ by considering only measures on $\overline{\bbD}$ and only the point $0$. If $\mu$ happens to be supported on $\partial G$, then $\varphi_*\mu$ is supported on $\partial\mathbb{D}$ so that $\lambda_{\infty}(0;\varphi_*\mu,q)$ is in fact independent of $q$ (see Theorem 2.5.4 in \cite{OPUC1}) so the same must be true of $\lambda_{\infty}(x;\mu,q)$. However, the following example shows that the value of $\lambda_{\infty}(0;\mu,q)$ is in general not as easily calculated when $\textrm{supp}(\mu)\nsubseteq\partial G$. \vspace{2mm} \noindent\textbf{Example.} Let us consider the special case of Corollary \ref{summable} where $G=\mathbb{D}$, $h=1$, and $z=0$. Let us further assume $\tau$ and $\nu$ are both probability measures. Fix any $N\in\mathbb{N}$ and let $Q_N(z)$ be a polynomial of degree at most $N$ satisfying $Q_N(0)=1$. Then for any $r<1$ we have \[ \int_0^{2\pi}|Q_N(re^{i\theta})|^2d\nu(\theta)\geq\lambda_N(0;\nu,2) \] because $Q_N(rz)$ is still a polynomial of degree $N$ in $z$ that is equal to $1$ at $0$. Integrating both sides in the variable $r$ with respect to $\tau$ from $0$ to $1$, we obtain $\lambda_N(0;\mu,2)\geq\lambda_N(0;\nu,2)$. Sending $N\rightarrow\infty$ we obtain $\lambda_{\infty}(0;\mu,2)\geq\lambda_{\infty}(0;\nu,2)>0$ (see equation (2.2.3) in \cite{OPUC1}). However, if $0\in\textrm{supp}(\tau)$ then the reverse inequality is false unless $d\nu=\frac{d\theta}{2\pi}$ (we still assume $\nu$ is a Szeg\H{o} measure on $\partial\mathbb{D}$), i.e. it is true that $\lambda_{\infty}(0;\mu,2)>\lambda_{\infty}(0;\nu,2)$. To see this, recall Proposition 2.16.2 in \cite{Rice}, which tells us that $Q_{n,z}(w):=K_n(z,w;\mu)K_n(z,z;\mu)^{-1}$ satisfies $Q_{n,z}(z)=1$ and $\|Q_{n,z}(z)\|^2_{\mu}=\lambda_n(z;\mu,2)$. Corollary \ref{summable} tells us that $\{Q_{n,0}(w)\}_{n\in\mathbb{N}}$ is uniformly bounded on $\{z:|z|\leq r_1\}$ for any $r_1<1$. By Montel's Theorem this is a normal family so we may take $n\rightarrow\infty$ through some subsequence $\mathcal{N}\subseteq\mathbb{N}$ so that $\{Q_{n,0}(w)\}_{n\in\mathcal{N}}$ converges uniformly to a function $Q_{\infty,0}(w)$, which is analytic in $\{z:|z|<r_1\}$ and $Q_{\infty,0}(0)=1$. By continuity and the fact that if $d\nu\neq\frac{d\theta}{2\pi}$ then $\lambda_{\infty}(0;\nu,2)<1$, it must be that \[ \int_0^{2\pi}|Q_{\infty,0}(re^{i\theta})|^2d\nu(\theta)>\frac{1+\lambda_{\infty}(0;\nu,2)}{2} \] for all $r$ sufficiently small (say $r<r_0$). By Dominated Convergence, the same must be true for all $Q_{n,0}(z)$ for $n$ sufficiently large and $n\in\mathcal{N}$. We conclude that for sufficiently large $n\in\mathcal{N}$, we have \begin{align*} \lambda_n(0;\mu,2)&=\|Q_{n,0}(z)\|^2_{\mu}=\int_0^{r_0}\int_0^{2\pi}|Q_{n,0}(re^{i\theta})|^2d\nu(\theta)d\tau(r)+ \int_{r_0}^{1}\int_0^{2\pi}|Q_{n,0}(re^{i\theta})|^2d\nu(\theta)d\tau(r)\\ &>\frac{1+\lambda_{\infty}(0;\nu,2)}{2}\tau([0,r_0])+\lambda_{\infty}(0;\nu,2)\tau((r_0,1])\\ &=\frac{1-\lambda_{\infty}(0;\nu,2)}{2}\tau([0,r_0])+\lambda_{\infty}(0;\nu,2). \end{align*} Since $\lambda_n(0;\mu,2)$ is decreasing in $n$, $\tau([0,r_0])>0$ and $\lambda_{\infty}(0;\nu,2)<1$, the desired conclusion follows. \begin{flushright} $\Box$ \end{flushright} \vspace{4mm} \section{Appendix}\label{appendix} \subsection{Non-uniqueness when $q=1$.}\label{pequal} On page 84 in \cite{StaTo}, it is stated that one does not have uniqueness of the $L^q$-extremal polynomial when $0<q<1$. This is a correct statement, but we show here that this can be extended to include the case $q=1$. \begin{proposition}\label{nonunique} If $\mu$ is a finite measure supported on $[-2,-1]\cup[1,2]$ and $\mu(A)=\mu(-A)$ for all measurable sets $A$, then one does not have uniqueness of the $L^1$-extremal polynomial $P_n(\mu,1)$ for every odd $n$. \end{proposition} \begin{proof} Suppose for contradiction that $P_{2n+1}(\mu,1)$ can be uniquely defined. By the symmetry of the measure, we must have that $P_{2n+1}(0;\mu,1)=0$. We may then write $P_{2n+1}(z;\mu,1)=zQ_n(z)$, for some polynomial $Q_n$ of degree $2n$ and satisfying $Q_n(x)=Q_n(-x)$ for all $x\in\mathbb{R}$. For $a\in(-1,1)$, define \[ P_{2n+1}^{(a)}(z)=(z-a)Q_n(z) \] so that $P_{2n+1}^{(0)}(z)=P_{2n+1}(z;\mu,1)$. We then have \begin{align*} \frac{\partial}{\partial a}\|P_{2n+1}^{(a)}\|_{L^1(\mu)}&=\frac{\partial}{\partial a}\left( \int_{-2}^{-1}(a-z)|Q_n(z)|d\mu(z)+\int_{1}^{2}(z-a)|Q_n(z)|d\mu(z)\right)\\ &=\int_{-2}^{-1}|Q_n(z)|d\mu(z)-\int_{1}^{2}|Q_n(z)|d\mu(z)=0, \end{align*} which contradicts our uniqueness assumption. \end{proof} If in Proposition \ref{nonunique} we also assume $\mu$ has no pure points then an alternative proof can be found by appealing to Theorem 2.1 in \cite{Pinkus}. \vspace{10mm}
\section*{1. Introduction } ~ In this paper, we will mainly discuss Jordan product determined points on matrix algebras. Before proceeding let us fix some symbols and notations in this paper. Let $M_n(R)$ be the algebra of all $n\times n$ matrices over a unital commutative ring $R$ with 6 invertible. Matrix units are denoted by $e_{ij}$ and the Jordan product ``$\circ$'' is defined as $x\circ y=xy+yx$. The identity matrix is denoted by $I$. By $M_n(R)^2$ we denote the $R$-linear span of all elements of the form $xy$ where $x,y\in M_n(R)$. The concept of zero product (resp. Jordan product, Lie product) determined algebras was introduced by Bre\v{s}ar et al. \cite{bresar}. According to \cite{bresar}, $M_n(B)$ ($n\geq 2$) is zero product determined where $B$ is a unital algebra. If $B$ is a unital algebra with 2 invertible, then $M_n(B)$ ($n\geq 3$) is zero Jordan product determined. From the results above, we can study the linear maps preserving zero product (resp. Jordan product) and derivable (resp. Jordan derivable) at zero point respectively. Wang et al. \cite{Wang1,Wang2} showed that (1) if a symmetric bilinear map $\{\cdot ,\cdot\}$ : $M_n(R)\times M_n(R)\to X$ satisfies the condition that $\{u,u\}=\{e,u\}$ whenever $u^2=u$, then there exists a linear map $f$ from $M_n(R)$ to $X$ such that $\{x,y\}=f(x\circ y)$ for all $x,y\in M_n(R)$; and (2) if an invertible linear map $\delta$ on $M_n(R)$ preserves identity-product, then it is a Jordan automorphism; and a linear map $\sigma$ on $M_n(R)$ is derivable at the identity matrix if and only if it is an inner derivation. Zhu et al. \cite{Zhu1} showed that for every $G\in M_n$, det$G=0$, is an all-multiplicative point in $M_n$. Gong and Zhu \cite{Gong} considered the case of Jordan all-multiplicative point in $M_n$. Zhu et al. \cite{Zhu2} showed that a matrix $G$ is an all-derivable point in $M_n$ if and only if $G\neq 0$. Zhao et al. \cite{Zhao} showed that every element of the algebra of all upper triangular matrices is a Jordan all-derivable point. Motivated by the concepts and results above, we will consider Jordan product determined points in matrix algebras. For $A\in M_n(R)$, we say that $A$ is a Jordan product determined point if for every $R$-module $X$ and every symmetric $R$-bilinear map $\{\cdot, \cdot\}$ : $M_n(R)\times M_n(R)\to X$ the following two conditions are equivalent: (i) there exists a fixed element $w\in X$ such that $\{x,y\}=w$ whenever $x\circ y=A$, $x,y\in M_n(R)$; (ii) there exists an $R$-linear map $T:M_n(R)^2\to X$ such that $\{x,y\}=T(x\circ y)$ for all $x,y\in M_n(R)$. We say that $G\in M_n(R)$ is a Jordan all-multiplicative point in $M_n(R)$ if for every $M_n(R)$-module $X$ and every Jordan multiplicative $R$-linear map $\varphi$ : $M_n(R)\to X$ at $G$ (i.e. $\varphi (S\circ T)=\varphi (S)\circ \varphi (T)$ for any $S,T\in M_n(R)$, $S\circ T=G$) with $\varphi (I)=I$ is a multiplicative mapping in $M_n(R)$. We say that $H\in M_n(R)$ is a Jordan all-derivable point in $M_n(R)$ if for every $M_n(R)$-module $X$ and every Jordan derivable $R$-linear map $\varphi$ : $M_n(R)\to X$ at $H$ (i.e. $\varphi (S\circ T)=\varphi (S)\circ T+S\circ \varphi (T)$ for any $S,T\in M_n(R)$, $S\circ T=H$) with $\varphi (I)=0$ is a Jordan derivation in $M_n(R)$. The above two definitions are somewhat different from \cite{Gong} and \cite{Zhao}. In this paper, we will prove that every matrix unit $e_{ij}$ is a Jordan product determined point in $M_n(R)$ when $n\geq 3$. As an application of the result above, we will show that every matrix unit $e_{ij}$ is a Jordan all-multiplicative point and a Jordan all-derivable point respectively. This paper is organized as follows. Section 2 concerns Jordan product determined points in $M_n(R)$, and we obtain the major results Theorem 2.2 and Theorem 2.3 in this paper. In Section 3, we get some corollaries by applying the main results in section 2. \section*{2. Jordan product determined points in $M_n(R)$} ~ According to \cite{bresar}, we give the lemma below.\\~\\ {\bf Lemma 2.1.} For $A\in M_n(R)$, $A$ is a Jordan product determined point if and only if for every $R$-module $X$ and every symmetric $R$-bilinear map $\{\cdot ,\cdot \}$ satisfy the condition (i), the following condition (iii) holds true. (iii) for every $x_t,y_t\in M_n(R)$ with $\sum_{t=1}^lx_t\circ y_t=0$, $t=1,2,\dots ,l$, $\sum_{t=1}^l\{x_t,y_t\}=0$ holds true.\\~\\ {\bf Proof.} Obviously, the ``only if'' part holds true.\\Conversely, if the condition (iii) holds true, we can define $R$-linear map $T$ : $M_n(R)^2\to X$ as $T(\sum_tx_t\circ y_t)=\sum_t\{x_t,y_t\}$ according to \cite{bresar}. Then $T$ satisfies condition (ii). We only need to prove that $T$ is well defined. Indeed, if condition (iii) is fulfilled, $T$ is well defined obviously. Hence $A$ is a Jordan product determined point. $\Box$\\~\\ {\bf Theorem 2.2.} $e_{ss}$, $s\in \{1,2,\dots ,n\}$, is a Jordan product determined point in $M_n(R)$ when $n\geq 3$.\\~\\ {\bf Proof.} Let $s$ be a fixed number, $X$ be an $R$-module, $\{\cdot,\cdot \}$ : $M_n(R)\times M_n(R)\to X$ be a symmetric (i.e. $\{x,y\}=\{y,x\}$) $R$-bilinear map. Now we assume that there exists a fixed element $w\in X$ such that $\{x,y\}=w$ whenever $x\circ y=e_{ss}$, $x,y\in M_n(R)$. Throughout the proof, $i,j,k,m$ will denote arbitrary indices. We begin by noticing that $(\frac {1}{2}e_{ss})\circ e_{ss}=e_{ss}$ and as $\{\cdot,\cdot\}$ is symmetric, then \begin{equation} \{\frac {1}{2}e_{ss},e_{ss}\}=w=\{e_{ss},\frac {1}{2}e_{ss}\}. \end{equation} Next we suppose $n\geq 3$ and divide the proof into three steps. Step 1. In this step, we assume $i\neq s,j\neq s,k\neq s \; \mathrm{and} \; m\neq s$. Case 1.1. $i\neq m \; \mathrm{and} \; j\neq k$.\\Since $(\frac {1}{2}e_{ss}+e_{ij})\circ e_{ss}=e_{ss}$ and as $\{\cdot, \cdot\}$ is symmetric, we have that \begin{equation} \{e_{ij},e_{ss}\}=0=\{e_{ss},e_{ij}\}. \end{equation} Noting $(\frac {1}{2}e_{ss}+e_{ij})\circ (e_{ss}+e_{km})=e_{ss}$, it follows $\{\frac {1}{2}e_{ss}+e_{ij},e_{ss}+e_{km}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, applying (1) and (2), this yields \begin{equation} \{e_{ij},e_{km}\}=0. \end{equation} Case 1.2. $i,j,k$ are distinct.\\From $(\frac {1}{2}e_{ss}+e_{ik})\circ (e_{ss}+e_{kk}-e_{ii})=e_{ss}$, we obtain that $\{\frac {1}{2}e_{ss}+e_{ik}, e_{ss}+e_{kk}-e_{ii}\}=w$. Because $\{\cdot, \cdot\}$ is symmetric, it follows from (1) and (2) that \begin{equation} \{e_{ik},e_{kk}\}=\{e_{ik},e_{ii}\}=\{e_{ii},e_{ik}\}=\{e_{kk},e_{ik}\}. \end{equation} Now we assume $n>3$. As $(\frac {1}{2}e_{ss}+e_{ij}+e_{ik})\circ (e_{ss}+e_{jk}-e_{kk})=e_{ss}$, we derive $\{\frac {1}{2}e_{ss}+e_{ij}+e_{ik}, e_{ss}+e_{jk}-e_{kk}\}=w$. Using (1), (2) and (3), this can be reduced to \begin{equation} \{e_{ij},e_{jk}\}=\{e_{ik},e_{kk}\}. \end{equation} (4) together with (5) yield \begin{equation} \{e_{ij}, e_{jk}\}=\{e_{ik}, e_{kk}\}=\{e_{ik}, e_{ii}\}=\{e_{ii}, e_{ik}\}. \end{equation} Since $\{\cdot, \cdot\}$ is symmetric, it follows that \begin{equation} \{e_{ii}, e_{ik}\}=\{e_{ik}, e_{ii}\}=\{e_{kk}, e_{ik}\}=\{e_{jk},e_{ij}\}. \end{equation} Case 1.3. $i\neq j$.\\By $(\frac{1}{2}e_{ss}+\frac{1}{2}e_{ii}+e_{ij}-\frac{1}{2}e_{jj})\circ (e_{ss}+e_{ji}-e_{ii}+e_{jj})=e_{ss}$, it is clear that $\{\frac {1}{2}e_{ss}+\frac {1}{2}e_{ii}+e_{ij}-\frac {1}{2}e_{jj}, e_{ss}+e_{ji}-e_{ii}+e_{jj}\}=w$. Then we can get from (1), (2), (3) and (4) that \begin{equation} \{e_{ij}, e_{ji}\}=\frac{1}{2}\{e_{ii}, e_{ii}\}+\frac{1}{2}\{e_{jj}, e_{jj}\}. \end{equation} Step 2. In this step, we consider some of the indices of the matrix units equal to $s$. Case 2.1. $i\neq s, j\neq s, k\neq s \; \mathrm{and} \; m\neq s$.\\For $(\frac{1}{2}e_{ss}-e_{si})\circ (e_{ss}-e_{ii})=e_{ss}$, we have that $\{\frac {1}{2}e_{ss}-e_{si}, e_{ss}-e_{ii}\}=w$. Applying (1) and (2), it can be reduced to $\{e_{si}, e_{ss}\}=\{e_{si}, e_{ii}\}$. As $\{\cdot, \cdot\}$ is symmetric, then it follows that \begin{equation} \{e_{si}, e_{ss}\}=\{e_{si}, e_{ii}\}=\{e_{ss}, e_{si}\}=\{e_{ii}, e_{si}\}. \end{equation} Since $(\frac{1}{2}e_{ss}-e_{is})\circ (e_{ss}-e_{ii})=e_{ss}$, a similar discussion as above shows that \begin{equation} \{e_{is}, e_{ss}\}=\{e_{is}, e_{ii}\}=\{e_{ss}, e_{is}\}=\{e_{ii}, e_{is}\}. \end{equation} Let $j\neq k$, then we have $(\frac {1}{2}e_{ss}+e_{sj})\circ (e_{ss}+e_{km}-e_{jj})=e_{ss}$ and it implies that $\{\frac {1}{2}e_{ss}+e_{sj}, e_{ss}+e_{km}-e_{jj}\}=w$. As $\{\cdot, \cdot\}$ is symmetric and by (1), (2) and (9), this yields \begin{equation} \{e_{sj}, e_{km}\}=0=\{e_{km}, e_{sj}\}\;\mathrm {if}\; j\neq k. \end{equation} From $(\frac{1}{2}e_{ss}+e_{si})\circ (e_{ss}-2e_{si})=e_{ss}$, we derive $\{\frac{1}{2}e_{ss}+e_{si}, e_{ss}-2e_{si}\}=w$. Then it follows from (1) and (9) that \begin{equation} \{e_{si}, e_{si}\}=0. \end{equation} Since $(\frac {1}{2}e_{ss}+e_{sj}-\frac{1}{2}e_{si})\circ (e_{ss}+e_{si}-e_{jj})=e_{ss}$ when $i\neq j$, then we have $\{\frac {1}{2}e_{ss}+e_{sj}-\frac{1}{2}e_{si}, e_{ss}+e_{si}-e_{jj}\}=w$. Using (1), (2), (9), (11) and (12), it is clear that \begin{equation} \{e_{sj}, e_{si}\}=0\;\mathrm {if}\; i\neq j. \end{equation} If $j\neq m$, from $(\frac {1}{2}e_{ss}+e_{js})\circ (e_{ss}+e_{km}-e_{jj})=e_{ss}$ we can get that $\{\frac {1}{2}e_{ss}+e_{js}, e_{ss}+e_{km}-e_{jj}\}=w$. As $\{\cdot, \cdot\}$ is symmetric and by (1), (2) and (10), it follows \begin{equation} \{e_{js}, e_{km}\}=0=\{e_{km}, e_{js}\}\;\mathrm {if}\; j\neq m. \end{equation} For $(\frac{1}{2}e_{ss}+e_{is})\circ (e_{ss}-2e_{is})=e_{ss}$, we have that $\{\frac{1}{2}e_{ss}+e_{is}, e_{ss}-2e_{is}\}=w$. Applying (1) and (10), it can be reduced to \begin{equation} \{e_{is}, e_{is}\}=0. \end{equation} Assume $i\neq j$, it follows from $(\frac {1}{2}e_{ss}+e_{js}-\frac{1}{2}e_{is})\circ (e_{ss}+e_{is}-e_{jj})=e_{ss}$ that $\{\frac {1}{2}e_{ss}+e_{js}-\frac{1}{2}e_{is}, e_{ss}+e_{is}-e_{jj}\}=w$. Using (1), (2), (10), (14) and (15), this yields \begin{equation} \{e_{js}, e_{is}\}=0 \;\mathrm {if}\; i\neq j. \end{equation} Case 2.2. $i\neq s, k\neq s \; \mathrm{and} \;j\neq s$.\\Since $(\frac {1}{2}e_{ss}+e_{ki}-e_{ks}-\frac {1}{2}e_{ii})\circ (e_{ss}+e_{is})=e_{ss}$ if $k\neq i$, then we have that $\{\frac {1}{2}e_{ss}+e_{ki}-e_{ks}-\frac {1}{2}e_{ii}, e_{ss}+e_{is}\}=w$. By (1), (2), (10) and (16), this yields $\{e_{ki}, e_{is}\}=\{e_{ks}, e_{ss}\}$ if $k\neq i$. For our purpose, it is more convenient to rewrite this equation as $\{e_{ik}, e_{ks}\}=\{e_{is}, e_{ss}\}$ if $i\neq k$. As $\{\cdot, \cdot\}$ is symmetric, then we can conclude from the above equation and (10) that \begin{equation} \{e_{ik}, e_{ks}\}=\{e_{is}, e_{ss}\}=\{e_{ii}, e_{is}\}=\{e_{is}, e_{ii}\}=\{e_{ss}, e_{is}\}=\{e_{ks}, e_{ik}\}\;\mathrm {if}\; i\neq k. \end{equation} If $i\neq k$, then we have $(\frac {1}{2}e_{ss}+e_{ik}-e_{sk}-\frac{1}{2}e_{ii})\circ (e_{ss}+e_{si})=e_{ss}$. By a similar discussion as above, this yields \begin{equation} \{e_{sk}, e_{ki}\}=\{e_{ss}, e_{si}\}=\{e_{si}, e_{ii}\}=\{e_{ii}, e_{si}\}=\{e_{si}, e_{ss}\}=\{e_{ki}, e_{sk}\}\;\mathrm {if}\; i\neq k. \end{equation} If $j\neq k$, from $(\frac {1}{2}e_{ss}+e_{sk}+e_{jk}-\frac {1}{2}e_{js})\circ (e_{ss}+e_{js}-e_{kk})=e_{ss}$ we have that $\{\frac {1}{2}e_{ss}+e_{sk}+e_{jk}-\frac {1}{2}e_{js}, e_{ss}+e_{js}-e_{kk}\}=w$. For $\{\cdot, \cdot\}$ is symmetric, applying (1), (2), (14), (15), (17) and (18), it follows \begin{equation} \{e_{sk}, e_{js}\}=\{e_{jk}, e_{kk}\}=\{e_{js}, e_{sk}\}\;\mathrm {if}\;j\neq k. \end{equation} Case 2.3. $i\neq s$.\\Since $(2e_{si}+\frac {1}{2}e_{is}-e_{ii})\circ (e_{si}+\frac {1}{4}e_{is}+\frac {1}{2}e_{ii})=e_{ss}$, we obtain $\{2e_{si}+\frac {1}{2}e_{is}-e_{ii}, e_{si}+\frac {1}{4}e_{is}+\frac {1}{2}e_{ii}\}=w$. Using (1), (12), (15), (17) and (18), this can be reduced to $\{e_{si}, e_{is}\}+\{e_{is}, e_{si}\}=\{e_{ss}, e_{ss}\}+\{e_{ii}, e_{ii}\}$. As $\{e_{si}, e_{is}\}=\{e_{is}, e_{si}\}$, it leads to \begin{equation} \{e_{si}, e_{is}\}=\{e_{is}, e_{si}\}=\frac {1}{2}\{e_{ss}, e_{ss}\}+\frac {1}{2}\{e_{ii}, e_{ii}\}. \end{equation} Step 3. Now concluding from case 1.1 and case 2.1, we can obtain \begin{equation} \{e_{ij}, e_{km}\}=0\;\mathrm{for \;every}\; i, j, k, m,\;\mathrm{if}\; i\neq m \; \mathrm{and} \; j\neq k. \end{equation} If $i,j,k$ are distinct and $n=3$, it follows from (4), (17), (18) and (19) that \begin{equation} \{e_{ij}, e_{jk}\}=\{e_{ik}, e_{kk}\}=\{e_{ii}, e_{ik}\}=\{e_{ik}, e_{ii}\}=\{e_{kk}, e_{ik}\}=\{e_{jk}, e_{ij}\}. \end{equation} If $i,j,k$ are distinct and $n>3$, from (6), (7), (17), (18) and (19) we have the equations above as well. So (22) holds true whenever $n \geq 3$.\\ From case 1.3 and case 2.3, we have \begin{equation} \{e_{ij}, e_{ji}\}=\frac {1}{2}\{e_{ii}, e_{ii}\}+\frac {1}{2}\{e_{jj}, e_{jj}\}\;\mathrm {if}\;i\neq j. \end{equation} Let $\sum_{t=1}^lx_t\circ y_t=0$ where $x_t, y_t\in M_n(R)$, $t=1,2,\ldots ,l$. We write $x_t$ and $y_t$ for \begin{equation}\nonumber x_t=\sum_{i=1}^n\sum_{j=1}^na_{ij}^te_{ij},\; y_t=\sum_{k=1}^n\sum_{m=1}^nb_{km}^te_{km}. \end{equation} Then for all $i$ and $m$ we have that \begin{equation} \sum_{t=1}^l\sum_{j=1}^n(a_{ij}^tb_{jm}^t+b_{ij}^ta_{jm}^t)=0. \end{equation} Now we will show $\sum_{t=1}^l\{x_t, y_t\}=0$. Note that \begin{eqnarray*} \sum_{t=1}^l\{x_t, y_t\}&=&\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{k=1}^n\sum_{m=1}^n\{a_{ij}^te_{ij}, b_{km}^te_{km}\}\\&=&\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{k=1}^n\sum_{m=1}^na_{ij}^tb_{km}^t\{e_{ij}, e_{km}\}. \end{eqnarray*} According to our assumptions, it follows from (21), (22), (23) and (24) that \begin{eqnarray*} \sum_{t=1}^l\{x_t, y_t\} &= &\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{k=1 \atop k\neq j}^n\sum_{m=1 \atop m\neq i}^na_{ij}^tb_{km}^t\{e_{ij}, e_{km}\}+\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{m=1 \atop m\neq i}^na_{ij}^tb_{jm}^t\{e_{ij}, e_{jm}\}\\ &+& \sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{k=1 \atop k\neq j}^na_{ij}^tb_{ki}^t\{e_{ij}, e_{ki}\}+\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^na_{ij}^tb_{ji}^t\{e_{ij}, e_{ji}\}\\ &= &\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^n\sum_{m=1 \atop m\neq i}^n(a_{ij}^tb_{jm}^t+b_{ij}^ta_{jm}^t)\{e_{ij}, e_{jm}\}\\ &+&\frac {1}{2}\sum_{t=1}^l\sum_{i=1}^n\sum_{j=1}^na_{ij}^tb_{ji}^t\left(\{e_{ii}, e_{ii}\}+\{e_{jj}, e_{jj}\}\right)\\ &=& \sum_{i=1}^n\sum_{m=1 \atop m\neq i}^n\left[\sum_{t=1}^l\sum_{j=1}^n(a_{ij}^tb_{jm}^t+b_{ij}^ta_{jm}^t)\{e_{im}, e_{mm}\}\right]\\ &+&\frac{1}{2}\sum_{i=1}^n\left[\sum_{t=1}^l\sum_{j=1}^n(a_{ij}^tb_{ji}^t+ b_{ij}^ta_{ji}^t)\{e_{ii}, e_{ii}\}\right]\\ &=& 0. \end{eqnarray*} By Lemma 2.1, $e_{ss}$ is a Jordan product determined point. $\Box$\\~\\ {\bf Theorem 2.3.} $e_{pq}$, $p\neq q$, is a Jordan product determined point in $M_n(R)$ when $n\geq 3$.\\~\\ {\bf Proof.} Let $p, q$ be the distinct fixed indices, $\{\cdot, \cdot\}$ : $M_n(R)\times M_n(R)\to X$ be a symmetric (i.e. $\{x, y\}=\{y, x\}$) $R$-bilinear map where $X$ is an $R$-module. And we assume that there exists a fixed element $w\in X$ such that $\{x,y\}=w$ whenever $x\circ y=e_{pq}$, $x,y\in M_n(R)$. Throughout the proof, $i,j,k,m$ will denote arbitrary indices. According to Lemma 2.1 and Step 3 in the proof of Theorem 2.2, we only need to verify (21), (22) and (23) hold true when $n\geq 3$. Now we suppose $n\geq 3$ and divide the proof into several steps. Step 1. In this step, we assume $i\neq q \; \mathrm {and} \; m\neq p$.\\Since $e_{ps}\circ e_{sq}=e_{pq}\:(s=1,2,\ldots ,n)$ and as $\{\cdot, \cdot\}$ is symmetric, hence we have \begin{equation} \{e_{ps}, e_{sq}\}=w=\{e_{sq}, e_{ps}\},\; s=1,2,\ldots ,n. \end{equation} Choosing $s\neq j\; \mathrm {and} \;s\neq k$, then we obtain $e_{ps}\circ (e_{sq}+e_{km})=e_{pq}$ and $(e_{ps}+e_{ij})\circ e_{sq}=e_{pq}$ respectively. For $\{\cdot, \cdot\}$ is symmetric, applying (25), it follows \begin{equation} \{e_{ps}, e_{km}\}=\{e_{km}, e_{ps}\}=\{e_{ij}, e_{sq}\}=\{e_{sq}, e_{ij}\}=0 \;\mathrm{if}\:s\neq j\; \mathrm {and} \; s\neq k. \end{equation} Given $s\neq j \; \mathrm {and} \; s\neq k$, then if $i\neq m\; \mathrm {and} \;j\neq k$ we have $(e_{ps}+e_{ij})\circ (e_{sq}+e_{km})=e_{pq}$. It is clear that $\{e_{ps}+e_{ij},e_{sq}+e_{km}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, using (25) and (26), this yields \begin{equation} \{e_{ij}, e_{km}\}=0=\{e_{km}, e_{ij}\}\:\mathrm{if}\:i\neq m\; \mathrm {and} \;j\neq k. \end{equation} Step 2. In this step, we suppose $i\neq q\; \mathrm {and} \;m\neq p$.\\Assuming $s\neq i\; \mathrm {and} \;s\neq m$, then if $i\neq m\; \mathrm {and} \;i\neq p$ we can verify $(e_{ps}+e_{im})\circ (e_{sq}+e_{mm}-e_{ii})=e_{pq}$. It follows that $\{e_{ps}+e_{im}, e_{sq}+e_{mm}-e_{ii}\}=w$. Since $\{\cdot, \cdot\}$ is symmetric, by (25) and (27), this yields \begin{equation} \{e_{im}, e_{mm}\}=\{e_{im}, e_{ii}\}=\{e_{ii}, e_{im}\}\:\mathrm {if}\: i\neq m\; \mathrm {and} \;i\neq p. \end{equation} From $e_{pm}\circ (e_{mq}+e_{mm}-e_{pp})=e_{pq}$, we have that $\{e_{pm}, e_{mq}+e_{mm}-e_{pp}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, it follows from (25) that \begin{equation} \{e_{pm}, e_{mm}\}=\{e_{pm}, e_{pp}\}=\{e_{pp}, e_{pm}\}. \end{equation} Choosing $s\neq j \; \mathrm {and} \; s\neq m$, if $i,j,k$ are distinct we obtain $(e_{ps}+e_{ij}+e_{im})\circ (e_{sq}+e_{jm}-e_{mm})=e_{pq}$. Then it leads to $\{e_{ps}+e_{ij}+e_{im}, e_{sq}+e_{jm}-e_{mm}\}=w$. Applying (25) and (27), this yields \begin{equation} \{e_{ij}, e_{jm}\}=\{e_{im}, e_{mm}\}\;\mathrm {if}\;i,j,k\;\mathrm{are distinct}. \end{equation} Since $\{\cdot, \cdot\}$ is symmetric, if $i,j,m$ are distinct, we can conclude from (28), (29) and (30) that \begin{equation} \{e_{ij}, e_{jm}\}=\{e_{im}, e_{mm}\}=\{e_{ii}, e_{im}\}=\{e_{im}, e_{ii}\}=\{e_{mm}, e_{im}\}=\{e_{jm}, e_{ij}\}. \end{equation} Step 3. In this step, we assume $i\neq q\;\mathrm {and}\;m\neq p$.\\Choosing $s\neq i\; \mathrm {and} \; s\neq m$, we suppose $n>3, \:i\neq p,\: i\neq m\; \mathrm {and} \;m\neq q$. As $(e_{ps}+\frac {1}{2}e_{ii}+e_{im}-\frac{1}{2}e_{mm})\circ (e_{sq}+e_{mi}-e_{ii}+e_{mm})=e_{pq}$, it follows that $\{e_{ps}+\frac {1}{2}e_{ii}+e_{im}-\frac{1}{2}e_{mm}, e_{sq}+e_{mi}-e_{ii}+e_{mm}\}=w$. Using (25), (27) and (31), if $n>3,\:i\neq p,\: i\neq m\; \mathrm {and} \;m\neq q$ we have that \begin{equation} \{e_{im},e_{mi}\}=\frac{1}{2}\{e_{ii}, e_{ii}\}+\frac{1}{2}\{e_{mm}, e_{mm}\}. \end{equation} Step 4. In this step, we assume $i=q\; \mathrm {and} \;m\neq p$. Case 4.1. $j\neq p\; \mathrm {and} \;k\neq q$.\\By (27), we have $\{e_{km}, e_{qj}\}=0$ if $m\neq q\; \mathrm {and} \;j\neq k$. As $\{\cdot, \cdot\}$ is symmetric, this yields \begin{equation} \{e_{qj},e_{km}\}=\{e_{km}, e_{qj}\}=0\;\mathrm {if}\; m\neq q\; \mathrm {and} \;j\neq k. \end{equation} Case 4.2. $j\neq p\; \mathrm {and} \; k=q$.\\Since $(e_{pq}+e_{qj})\circ (e_{qq}-e_{jj})=e_{pq}$ if $j\neq q$, then we have $\{e_{pq}+e_{qj}, e_{qq}-e_{jj}\}=w$. For $\{\cdot, \cdot\}$ is symmetric, it follows from (25) and (27) that \begin{equation} \{e_{qj}, e_{qq}\}=\{e_{qj}, e_{jj}\}=\{e_{jj}, e_{qj}\}=\{e_{qq}, e_{qj}\}\;\mathrm {if}\; j\neq q. \end{equation} Noting $(e_{pq}+e_{qj}+e_{pp})\circ (e_{qq}+e_{qm}-e_{jj}-e_{pm})=e_{pq}$ if $j\neq q\; \mathrm {and} \;m\neq q$, then we obtain $\{e_{pq}+e_{qj}+e_{pp}, e_{qq}+e_{qm}-e_{jj}-e_{pm}\}=w$. Applying (25), (27), (31), (33) and (34), it can be reduced to \begin{equation} \{e_{qj}, e_{qm}\}=0 \;\mathrm{if}\; j\neq q\; \mathrm {and} \;m\neq q. \end{equation} Case 4.3. $j=p\; \mathrm {and} \;k\neq q$.\\From $(\frac {1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp})\circ (e_{pq}-2e_{qp}+e_{qq})=e_{pq}$, we have that $\{\frac {1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}, e_{pq}-2e_{qp}+e_{qq}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, it follows from (25) and (27) that \begin{equation} \{e_{qp}, e_{qq}\}-\{e_{pp}, e_{qp}\}-2\{e_{qp}, e_{qp}\}=0. \end{equation} Since $(\frac{1}{2}e_{pp}+e_{pq}+2e_{qp})\circ (e_{pq}-2e_{qp}+\frac{1}{2}e_{qq})=e_{pq}$, a similar discussion shows that \begin{equation} \{e_{qp}, e_{qq}\}-\{e_{pp}, e_{qp}\}-4\{e_{qp}, e_{qp}\}=0. \end{equation} Because $\{\cdot, \cdot\}$ is symmetric, comparing (36) with (37), we get that \begin{equation} \{e_{qp}, e_{qp}\}=0, \end{equation} \begin{equation} \{e_{qp}, e_{qq}\}=\{e_{pp}, e_{qp}\}=\{e_{qp},e_{pp}\}=\{e_{qq}, e_{qp}\}. \end{equation} Noting $(\frac{1}{2} e_{pp}+\frac{1}{2}e_{pq}+e_{qp})\circ (e_{pq}-2e_{qp}+e_{qq}+e_{km})=e_{pq}$ if $k\neq p\; \mathrm {and} \;m\neq q$, we have that $\{\frac{1}{2} e_{pp}+\frac{1}{2}e_{pq}+e_{qp}, e_{pq}-2e_{qp}+e_{qq}+e_{km}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, applying (25), (27) and (36), if $k\neq p\; \mathrm {and} \; m\neq q$ it leads to \begin{equation} \{e_{qp}, e_{km}\}=0=\{e_{km}, e_{qp}\}. \end{equation} Case 4.4. $j=p\; \mathrm {and} \;k=q$.\\If $m\neq q$, then we can verify that \begin{equation}\nonumber (\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}-e_{mm})\circ (e_{pq}-2e_{qp}+e_{qq}+e_{qm}+e_{pm})=e_{pq}, \end{equation} \begin{equation}\nonumber (\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}+\frac {1}{2}e_{mm})\circ (e_{pq}-2e_{qp}+e_{qq}-2e_{qm}+e_{pm})=e_{pq}. \end{equation} Hence we have that \begin{equation}\nonumber \{\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}-e_{mm}, e_{pq}-2e_{qp}+e_{qq}+e_{qm}+e_{pm}\}=w, \end{equation} \begin{equation}\nonumber \{\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}+\frac {1}{2}e_{mm}, e_{pq}-2e_{qp}+e_{qq}-2e_{qm}+e_{pm}\}=w. \end{equation} For $\{\cdot, \cdot\}$ is symmetric, it follows from (25), (27), (31), (38), (39) and (40) that \begin{equation}\nonumber \{e_{qp}, e_{pm}\}+\{e_{qp},e_{qm}\}-\{e_{mm}, e_{qm}\}=0, \end{equation} \begin{equation}\nonumber \{e_{qp}, e_{pm}\}-2\{e_{qp},e_{qm}\}-\{e_{mm}, e_{qm}\}=0. \end{equation} Thus if $m\neq q$ we have that \begin{equation} \{e_{qp},e_{qm}\}=0=\{e_{qm},e_{qp}\}, \end{equation} \begin{equation} \{e_{qp}, e_{pm}\}=\{e_{mm}, e_{qm}\}=\{e_{qm}, e_{mm}\}=\{e_{pm}, e_{qp}\}. \end{equation} Step 5. In this step, we assume $i\neq q\; \mathrm {and} \;m=p$. Case 5.1. $j\neq p\; \mathrm {and} \;k\neq q$.\\From (27), if $i\neq p\; \mathrm {and} \;j\neq k$ it follows that \begin{equation} \{e_{ij}, e_{kp}\}=0=\{e_{kp}, e_{ij}\}. \end{equation} Case 5.2. $j=p\; \mathrm {and} \;k\neq q$.\\Since $(e_{pq}+e_{ip}-e_{kq})\circ (e_{qq}+e_{kp})=e_{pq}$ if $i\neq p\; \mathrm {and} \;k\neq p$, it is clear that $\{e_{pq}+e_{ip}-e_{kq}, e_{qq}+e_{kp}\}=w$. Applying (25), (27) and (31), it can be reduced to \begin{equation} \{e_{ip}, e_{kp}\}=0. \end{equation} Case 5.3. $j\neq p\; \mathrm {and} \;k=q$.\\By (40), if $i\neq p\; \mathrm {and} \;j\neq q$ we have that \begin{equation} \{e_{ij}, e_{qp}\}=0. \end{equation} Case 5.4. $j=p\; \mathrm {and} \;k=q$.\\If $i\neq p\; \mathrm {and} \;i\neq q$, then we can verify that \begin{eqnarray*} (\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}-e_{ii})\circ (e_{pq}-2e_{qp}+e_{qq}+2e_{ip}+e_{iq})=e_{pq},\\ (\frac{1}{2}e_{pp}+\frac{1}{2}e_{pq}+e_{qp}+\frac{1}{2}e_{ii})\circ (e_{pq}-2e_{qp}+e_{qq}-e_{ip}+e_{iq})=e_{pq}. \end{eqnarray*} According to our assumptions, using (25), (27), (31), (38), (39) and (40), we have \begin{equation} \{e_{pp},e_{ip}\}+\{e_{qp}, e_{iq}\}-2\{e_{ii}, e_{ip}\}+2\{e_{qp},e_{ip}\}=0, \end{equation} \begin{equation} \{e_{pp},e_{ip}\}-2\{e_{qp}, e_{iq}\}+\{e_{ii}, e_{ip}\}+2\{e_{qp},e_{ip}\}=0. \end{equation} Comparing the two equations above, as $\{\cdot, \cdot\}$ is symmetric, it follows that \begin{equation} \{e_{qp}, e_{iq}\}=\{e_{ii},e_{ip}\}=\{e_{iq},e_{qp}\}=\{e_{ip},e_{ii}\} \;\mathrm{if}\; i\neq p\; \mathrm {and} \;i\neq q. \end{equation} Substituting (48) into (46), we obtain \begin{equation} \{e_{pp},e_{ip}\}-\{e_{ii}, e_{ip}\}+2\{e_{qp},e_{ip}\}=0\;\mathrm {if} \;i\neq p\; \mathrm {and} \;i\neq q. \end{equation} Noting $(e_{pq}+e_{ip})\circ (e_{ii}-e_{pp}+2e_{qq})=e_{pq}$ if $i\neq p\; \mathrm{and}\;i\neq q$, then we have that $\{e_{pq}+e_{ip}, e_{ii}-e_{pp}+2e_{qq}\}=w$. Since $\{\cdot, \cdot\}$ is symmetric, it follows from (25) and (27) that \begin{equation} \{e_{ip}, e_{ii}\}=\{e_{ip}, e_{pp}\}=\{e_{pp}, e_{ip}\}=\{e_{ii}, e_{ip}\}\;\mathrm{if}\; i\neq p \; \mathrm {and} \;i\neq q. \end{equation} Substituting (50) into (49), we derive \begin{equation} \{e_{qp}, e_{ip}\}=0=\{e_{ip}, e_{qp}\}\;\mathrm{if}\;i\neq p\; \mathrm {and} \;i\neq q. \end{equation} Step 6. In this step, we assume $i=q\; \mathrm {and} \;m=p$. Case 6.1. $j\neq p\; \mathrm {and} \;k\neq q$.\\From (27), if $j\neq k$ it is clear that \begin{equation} \{e_{qj},e_{kp}\}=\{e_{kp}, e_{qj}\}=0. \end{equation} Case 6.2. $j\neq p\; \mathrm {and} \; k=q$.\\From (41), we have \begin{equation} \{e_{qj}, e_{qp}\}=\{e_{qp}, e_{qj}\}=0\;\mathrm{if}\;j\neq q. \end{equation} Case 6.3. $j=p\; \mathrm {and} \;k\neq q$.\\From (51), it follows \begin{equation} \{e_{qp}, e_{kp}\}=\{e_{kp}, e_{qp}\}=0 \;\mathrm{if}\;k\neq p. \end{equation} Case 6.4. $j=p\; \mathrm {and} \;k=q$.\\From (38), we obtain that \begin{equation} \{e_{qp}, e_{qp}\}=0. \end{equation} Step 7. In this step, we assume $i=q\; \mathrm {and} \;m\neq p$.\\If $n>3,\:j\neq p,\: j\neq q, \:j\neq m\; \mathrm {and} \;m\neq q$, choosing $s\neq q\; \mathrm {and} \;s\neq m$, then we can verify $(e_{ps}+e_{mm}-e_{jm})\circ (e_{sq}+e_{qm}+e_{qj})=e_{pq}$. It follows that $\{e_{ps}+e_{mm}-e_{jm}, e_{sq}+e_{qm}+e_{qj}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, by (25) and (27), if $n>3,\:j\neq p, \:j\neq q,\: j\neq m\; \mathrm {and} \;m\neq q$ we have that \begin{equation} \{e_{mm}, e_{qm}\}=\{e_{jm}, e_{qj}\}=\{e_{qj}, e_{jm}\}=\{e_{qm}, e_{mm}\}. \end{equation} In the case of $n=3$, if $q, j, m$ are distinct we can conclude from (34) and (42) that \begin{equation} \{e_{qj}, e_{jm}\}=\{e_{qm}, e_{mm}\}=\{e_{qq}, e_{qm}\}=\{e_{qm}, e_{qq}\}=\{e_{mm}, e_{qm}\}=\{e_{jm}, e_{qj}\}. \end{equation} In the case of $n>3$, if $q, j, m$ are distinct, from (34), (42) and (56) we have the same equations as well. So (57) holds true whenever $n\geq 3$. Step 8. In this step, we assume that $i\neq q\; \mathrm {and} \; m=p$.\\If $n>3,\:i\neq p,\: i\neq j,\: j\neq p\; \mathrm {and} \; j\neq q$, choosing $s\neq p\; \mathrm {and} \;s\neq i$, then we have that $(e_{ps}+e_{jp}-e_{ip})\circ (e_{sq}+e_{ij}+e_{ii})=e_{pq}$. From the assumptions this yields $\{e_{ps}+e_{jp}-e_{ip}, e_{sq}+e_{ij}+e_{ii}\}=w$. Because $\{\cdot, \cdot\}$ is symmetric, if $n>3,\:i\neq p,\: i\neq j,\: j\neq p\; \mathrm {and} \;j\neq q$ it follows from (25) and (27) that \begin{equation} \{e_{jp}, e_{ij}\}=\{e_{ip}, e_{ii}\}=\{e_{ij}, e_{jp}\}=\{e_{ii}, e_{ip}\}. \end{equation} If $i,j,p$ are distinct and $n=3$, then we can conclude from (48) and (50) that \begin{equation} \{e_{ij}, e_{jp}\}=\{e_{ip}, e_{pp}\}=\{e_{ii}, e_{ip}\}=\{e_{ip}, e_{ii}\}=\{e_{pp}, e_{ip}\}=\{e_{jp}, e_{ij}\}. \end{equation} If $i,j,p$ are distinct and $n>3$, from (48), (50) and (58) we obtain the equations above as well. Hence, (59) holds true whenever $n\geq 3$. Step 9. Selecting $s\neq p\; \mathrm {and} \;s\neq m$, then if $m\neq p\;\mathrm{and} \;m\neq q$ we can verify $(e_{ps}+\frac{1}{2}e_{pp}+e_{mp}-\frac{1}{2}e_{mm})\circ (e_{sq}+e_{ss}+e_{pm}-e_{pp}+e_{mm})=e_{pq}$. It implies that $\{e_{ps}+\frac{1}{2}e_{pp}+e_{mp}-\frac{1}{2}e_{mm}, e_{sq}+e_{ss}+e_{pm}-e_{pp}+e_{mm}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, by (25), (27), (31) and (59), if $m\neq p\;\mathrm{and} \;m\neq q$ this yields \begin{equation} \{e_{mp}, e_{pm}\}=\frac {1}{2}\{e_{pp}, e_{pp}\}+\frac {1}{2}\{e_{mm}, e_{mm}\}=\{e_{pm}, e_{mp}\}. \end{equation} Choosing $s\neq q\; \mathrm {and} \;s\neq m$, then if $m\neq p\;\mathrm{and} \;m\neq q$ we have $(e_{ps}+e_{mq}+\frac {1}{2}e_{qq}-\frac{1}{2}e_{ss}-\frac {1}{2}e_{mm})\circ (e_{sq}+e_{qm}-e_{qq}+e_{mm})=e_{pq}$. It is clear that $\{e_{ps}+e_{mq}+\frac {1}{2}e_{qq}-\frac{1}{2}e_{ss}-\frac {1}{2}e_{mm}, e_{sq}+e_{qm}-e_{qq}+e_{mm}\}=w$. Because $\{\cdot, \cdot\}$ is symmetric, if $m\neq p\;\mathrm{and} \;m\neq q$ it follows from (25), (27), (31) and (57) that \begin{equation} \{e_{mq}, e_{qm}\}=\frac {1}{2}\{e_{mm}, e_{mm}\}+\frac{1}{2}\{e_{qq},e_{qq}\}=\{e_{qm},e_{mq}\}. \end{equation} Since $(e_{pp}+\frac{5}{4}e_{pq}+e_{qp}+e_{qq})\circ (e_{pp}-\frac{3}{4}e_{pq}-e_{qp}+e_{qq})=e_{pq}$, then we have that $\{e_{pp}+\frac{5}{4}e_{pq}+e_{qp}+e_{qq}, e_{pp}-\frac{3}{4}e_{pq}-e_{qp}+e_{qq}\}=w$. As $\{\cdot, \cdot\}$ is symmetric, using (25), (27) and (38), it can be reduced to \begin{equation} \{e_{pq}, e_{qp}\}=\frac{1}{2}\{e_{pp}, e_{pp}\}+\frac{1}{2}\{e_{qq}, e_{qq}\}=\{e_{qp}, e_{pq}\}. \end{equation} Step 10. If $j\neq p\; \mathrm {and} \;j\neq q$, we can verify that $(e_{pp}+e_{pq}-4e_{qp}-3e_{qq}+e_{qj}+e_{jj})\circ (2e_{pp}-4e_{qp}-e_{qq}-4e_{jp}-2e_{jq}+e_{jj})=e_{pq}$. It follows that $\{e_{pp}+e_{pq}-4e_{qp}-3e_{qq}+e_{qj}+e_{jj}, 2e_{pp}-4e_{qp}-e_{qq}-4e_{jp}-2e_{jq}+e_{jj}\}=w$. Since $\{\cdot, \cdot\}$ is symmetric, applying (25), (27), (31), (38), (39), (40), (53), (54), (57), (59), (61) and (62), if $j\neq p\; \mathrm {and} \;j\neq q$ we have that \begin{equation} \{e_{qj}, e_{jp}\}=\{e_{qp}, e_{qq}\}=\{e_{qp}, e_{pp}\}=\{e_{qq}, e_{qp}\}=\{e_{pp}, e_{qp}\}=\{e_{jp}, e_{qj}\}. \end{equation} Step 11. Now, if $i\neq m\; \mathrm {and} \;j\neq k$ we can conclude from step 1, step 4, step 5 and step 6 that \begin{equation} \{e_{ij}, e_{km}\}=0. \end{equation} If $i,j,m$ are distinct, it follows from step 2, step 7, step 8 and step 10 that \begin{equation} \{e_{ij}, e_{jm}\}=\{e_{im}, e_{mm}\}=\{e_{ii}, e_{im}\}=\{e_{im}, e_{ii}\}=\{e_{mm}, e_{im}\}=\{e_{jm}, e_{ij}\}. \end{equation} If $i\neq m$ and $n=3$, from (60), (61) and (62) we have that \begin{equation} \{e_{im}, e_{mi}\}=\frac{1}{2}\{e_{ii}, e_{ii}\}+\frac{1}{2}\{e_{mm}, e_{mm}\}. \end{equation} If $i\neq m$ and $n>3$, from step 3 and step 9 we have the same equations as well. So we complete the proof. $\Box$ \section*{3. Several applications} ~ Now, we will give two applications of the theorems above.\\~\\{\bf Definition 3.1.} We say that $G\in M_n(R)$ is a Jordan all-multiplicative point in $M_n(R)$ if for every $M_n(R)$-module $X$ and every Jordan multiplicative $R$-linear map $\varphi$ : $M_n(R)\to X$ at $G$ (i.e. $\varphi (S\circ T)=\varphi (S)\circ \varphi (T)$ for any $S,T\in M_n(R),\: S\circ T=G$) with $\varphi (I)=I$ is a multiplicative mapping in $M_n(R)$.\\~\\{\bf Corollary 3.2.} Every matrix units $e_{ij}$ in $M_n(R)$, $n\geq 3$, is a Jordan all-multiplicative point.\\~\\{\bf Proof.} Let $X$ be an $M_n(R)$-module, $\varphi$ be a Jordan multiplicative $R$-linear map at $e_{ij}$. Then it follows \begin{equation}\nonumber \varphi (I)=I, \;\varphi (S\circ T)=\varphi (S)\circ \varphi (T)\ \mathrm{for\;all}\: S,T\in M_n(R),\:S\circ T=e_{ij}. \end{equation} Set $\{S,T\}=\varphi (S) \circ \varphi (T)$ for any $S,T\in M_n(R)$, thus $\{\cdot,\cdot\}$ is a symmetric $R$-bilinear map. Since $\varphi$ is a multiplicative map at $e_{ij}$, we have \begin{equation}\nonumber \{S,T\}=\varphi (S)\circ \varphi (T)=\varphi (e_{ij}) \ \mathrm{for\;all}\: S,T\in M_n(R)\;\mathrm{with}\;S\circ T=e_{ij}. \end{equation} As $e_{ij}$ is a Jordan product determined point in $M_n(R)$, then there exists an $R$-linear map $\phi$ : $M_n(R)^2\to X$ such that $\{S,T\}=\phi(S\circ T)$ for all $S,T\in M_n(R)$. So \begin{equation}\nonumber \varphi (S)\circ \varphi (T)=\{S,T\}=\phi (S\circ T),\; \forall \:S,T\in M_n(R). \end{equation} Set $S=I$ in the equation above, then we have $\varphi (T)=\phi (T)$ for every $T\in M_n(R)$. It follows that \begin{equation}\nonumber \varphi(S\circ T)=\varphi (S)\circ \varphi (T),\;\forall\; S,T\in M_n(R). \end{equation} Hence $\varphi$ is a multiplicative mapping in $M_n(R)$. The proof is completed. $\Box$ \\~\\{\bf Definition 3.3.} We say that $H\in M_n(R)$ is a Jordan all-derivable point in $M_n(R)$ if for every $M_n(R)$-module $X$ and every Jordan derivable $R$-linear map $\varphi$ : $M_n(R)\to X$ at $H$ (i.e. $\varphi (S\circ T)=\varphi (S)\circ T+S\circ \varphi (T)$ for any $S,T\in M_n(R),\: S\circ T=H$) with $\varphi (I)=0$ is a Jordan derivation in $M_n(R)$.\\~\\{\bf Corollary 3.4.} Every matrix units $e_{ij}$ in $M_n(R)$, $n\geq 3$, is a Jordan all-derivable point.\\~\\{\bf Proof.} Let $X$ be an $M_n(R)$-module, $\tau$ be a Jordan derivable $R$-linear map at $e_{ij}$. Then we have \begin{equation}\nonumber \tau (I)=0,\; \tau (S\circ T)=\tau (S)\circ T+S\circ \tau (T)\;\mathrm{ for\; all}\;S,T\in M_n(R),\;S\circ T=e_{ij}. \end{equation} Set $\{S,T\}=\tau (S)\circ T+S\circ \tau (T)$ for any $S,T\in M_n(R)$, so $\{\cdot,\cdot\}$ is a symmetric $R$-bilinear map. Since $\tau$ is a derivable map at $e_{ij}$, it follows \begin{equation}\nonumber \{S,T\}=\tau (S)\circ T+S\circ \tau (T)=\tau (e_{ij})\;\mathrm{ for\; all}\;S,T\in M_n(R)\;\mathrm{with}\;S\circ T=e_{ij}. \end{equation} As $e_{ij}$ is a Jordan product determined point in $M_n(R)$, then there exists an $R$-linear map $\psi$ : $M_n(R)^2\to X$ such that $\{S,T\}=\psi(S\circ T)$ for all $S,T\in M_n(R)$. Hence \begin{equation}\nonumber \tau (S)\circ T+S \circ \tau (T)=\{S,T\}=\psi (S\circ T),\; \forall \:S,T\in M_n(R). \end{equation} Set $S=I$ in the equation above, then we have $\tau (T)=\psi (T)$ for every $T\in M_n(R)$. It follows that \begin{equation}\nonumber \tau(S\circ T)=\tau (S)\circ T+S\circ \tau (T),\; \forall \:S,T\in M_n(R). \end{equation} Then $\tau$ is a Jordan derivation in $M_n(R)$. The proof is completed. $\Box$\\
\section{Introduction} The earliest phases of star formation have been the subject of intense observational and theoretical investigation over the past four decades. Nevertheless, together with the very late stages of star formation (formation of planetary systems), the very early processes of molecular cloud fragmentation into high-density ``cores'' and the subsequent early contraction of these cores before protostars form in their centers remain the two least constrained and most debated aspects of star formation theory (McKee and Ostriker 2007). The reason for this lack of strong constraints is that it is difficult to extract physical conditions in star forming regions from observable quantities, such as dust column or molecular line maps; each observable is typically affected by multiple factors which are nontrivial to deconvolve. As a result, theoretical models of cloud fragmentation and core evolution span a large range of possibilities. A diverse array of mutually interacting microphysical processes influence the evolution of molecular clouds to differing extents: gravity, turbulence, magnetic fields, UV and cosmic ray (CR) ionization, and a complex network of chemical reactions. Without strong observational constraints on the relative importance of these processes, competing star formation theories have arisen based on different assumptions of the physics driving the formation of prestellar cores and their collapse to form protostars. The path to the appearance of a protostar depends sensitively on the initial conditions and the relative importance of the different physical processes driving the formation and contraction of prestellar molecular cloud cores.The most important distinction among different calculations of prestellar collapse pertains to the nature of any forces opposing gravity during the prestellar collapse. In this context, we can identify two broad classes of models: non-magnetic and magnetic. Non-magnetic collapsing models include early numerical and semi-analytic solutions of the Euler equations of hydrodynamic collapse (e.g., Larson 1969; Penston 1969; Yorke \& Kr\"ugel 1977; Shu 1977) \footnote{Note however that Shu's singular isothermal sphere models refer to protostellar cores in which a central object exists, since the initial conditions for the inside-out collapse are singular at the center. Thus, this class of models cannot be used to study the early prestellar phase of the core evolution.}, as well as numerical studies of the gravo-turbulent formation and evolution of bound condensations formed by converging flows in turbulent model clouds (see Mac Low \& Klessen 2004 and Scalo \& Elmegreen 2004 for recent reviews). The evolution of cores formed by turbulence-driven fragmentation is compatible with models of pure hydrodynamical collapse (e.g., Gong \& Ostriker 2009). Turbulence-initiated core formation models can be generally separated into models of rapid global collapse, where the entire molecular cloud is viewed as an evanescent structure (e.g., Hartmann et al.\ 2001; Heitsch \& Hartmann 2008), and models with some initial decaying turbulent support that extends the lifetime of clouds (e.g., Krumholz et al.\ 2006; Tan et al.\ 2006). Magnetic models include magnetohydrodynamical calculations of the magnetically mediated contraction of prestellar cores, which is typically slower than the pure hydrodynamical collapse (with lower infall velocities and longer evolution timescales, e.g.\ Fiedler \& Mouschovias 1993; Li \& Shu 1996; Tassis \& Mouschovias 2007), and result in a characteristic hourglass-like configuration of the magnetic field threading the core (see Mouschovias \& Ciolek 1999 for a review). In magnetic models, the lifetime of the parent molecular cloud is controlled by the initial value of the mass-to-magnetic-flux ratio. If this ratio is smaller than a critical value (the cloud is {\em magnetically subcritical}), then the cloud as a whole is long-lived as the magnetic field can support it against gravity. In this case, {\em magnetically supercritical} cores form through the process of ambipolar diffusion and collapse to form protostars, while the cloud envelope remains supported by the field. If the mass to magnetic flux ratio of the parent cloud is larger than the critical value, then the cloud as a whole is unstable and short-lived, while the formation and collapse of cores also proceeds in shorter timescales. The coupling between chemistry and dynamics is an essential step in connecting the dynamics of prestellar cores to observations, and has long been recognized as critical in extracting information from molecular line observations. The problem has been pursued for the past decade in a series of ground-breaking studies. Bergin \& Langer (1997) identified, through studies of static media, ``early'' and ``late'' molecules (molecules which peak in abundance either in the initial or in the later stages of the evolution of a molecular cloud fragment). Aikawa et al. (2001) compared observations of L1544 with coupled chemical/dynamical models. For the dynamics, they used the analytical similarity solution of Larson (1969) and Penston (1969), as well as a ``slowed down'' version of it (using an overall slow-down factor) to approximate the slower magnetic collapse, and found that their fast models are better fits to L1544. Li et al. (2002) and Shematovich (2003) also modeled L1544, using a non-magnetic model and the Li (1999) magnetic model for the dynamics, which treats the magnetic field in an approximate way including only magnetic pressure forces and assuming a spherical geometry. They also included grain-surface reactions in their calculations, and they found the magnetic model to be a better fit to the observations. They showed that the magnetic collapse is accelerating, and therefore a uniform slow-down of the non-magnetic collapse is not a good approximation, because in magnetic collapse the amount of time spent at different densities changes with respect to the non-magnetic case. Aikawa et al. (2003) updated the Aikawa et al. (2001) calculation to include grain-surface reactions, and they found that moderate retardation factors (factor of 3) are not in as strong disagreement with the data as in Aikawa et al. (2001), however they still found that the faster models are preferred. Li et al. (2002) however noted that their preferred Larson-Penston solution features very large infall velocities that cannot be reconciled with observations of L1544. Lee et al. (2003) modeled the chemistry of various prestellar cores to fit Bonnor-Ebert and Plummer-like radial density profiles using dust continuum and line emission. Lee et al. (2004) looked at the chemical evolution of prestellar cores aiming to produce line profiles for prestellar and protostellar cores. For modeling the dynamical evolution of prestellar cores they used a sequence of Bonnor-Ebert profiles, however the assignment of times and velocities to each profile was not treated self-consistently. For the protostellar cores they used the inside-out collapse model of Shu (1977). Flower et al. (2005), using a homologous collapse treatment of prestellar cores, emphasized the effect of grain growth on the chemistry. Aikawa et al. (2005) revisited the prestellar core problem, but instead of using an analytic solution for the dynamics, they used post-processing of a numerical simulation of the collapse of Bonnor-Ebert spheres (one very close to equilibrium and one very unstable), and found significant differences in the system chemistry depending on the initial conditions. This result was confirmed by Keto \& Caselli (2008) who coupled one-dimensional dynamical simulations with a simple CO chemistry network and radiative transfer. In a follow-up study, Keto \& Caselli (2010), starting from a Bonnor-Ebert sphere of central density $2\times 10^{4} \rm {\, cm^{-3}}$ in unstable equilibrium, evolving quasi-statically, found that at a density of $2\times 10^{7} \rm {\, cm^{-3}}$ the abundance of CO is much lower than observations in L1544, and argued for a higher CO desorption rate from grains. On certain issues a consensus is forming among these studies (see, e.g., Bergin \& Tafalla 2007 for a recent review). One such issue is the sensitivity of the chemical state of the contracting core to the duration of the collapse. Characteristic examples are the carbon-bearing molecules, which tend to become depleted on the ice mantles of dust grains and are therefore most prominent in the early stages of evolution of a core, in contrast to nitrogen-bearing molecules which are not depleted and, as a result, their abundance increases in later stages. However, core age is not the only contributing factor in determining molecular abundances. Observations have established the existence of chemical differentiation in prestellar cores of otherwise similar properties, which however show strong variations in the abundances of a number of chemical species. Characteristic examples of chemical differentiation can be seen in observations of prestellar cores L1544, L1498, L1517B which show CO depletion (Tafalla et al. 2002, 2004), while L1521E (a core of similar central density with L1517B) shows no CO depletion (Tafalla \& Santiago 2004). This differentiation may be a result of the dynamics of the collapse as suggested by, e.g., Shematovich et al. (2003), and thus reflect the initial conditions in the parent star forming region. In this paper, we self-consistently follow the chemistry and dynamics of contracting prestellar cores. We do so by adding to the system of equations of Eulerian hydrodynamics, or MHD (as appropriate), one continuity equation with sources and sinks given by the chemical reactions network {\em for each species} followed. In this way, the dynamics and non-equilibrium chemistry evolve simultaneously. This is particularly important in the case of MHD models, as abundances of charged species have a feedback effect on the dynamics, since the magnetic field couples only to charged molecules and grains, which in turn determine the evolution of the magnetic field and thus the dynamics of the contraction. In a companion publication (Tassis et al.~2011, in preparation, hereafter Paper II), we discuss in more detail the interplay between chemistry and dynamics in magnetic models. Our paper is organized as follows. In \S \ref{dynamics} we discuss our dynamical models (hydrodynamic and MHD) of core evolution. The chemical network we implement is discussed in \S \ref{chemistry}. Details on the parameter study we have conducted are given in \S \ref{parstu}. We present our results in \S \ref{res} and we discuss our findings and conclusions in \S \ref{thedisc}. \section{Dynamical Models of Molecular Cloud Cores}\label{dynamics} \subsection{Hydrodynamic Core Models} We model the contraction of a non-magnetic, self-gravitating, isothermal prestellar core in spherical symmetry. The core is assumed to be embedded within a uniform-density isothermal cloud. We solve the equations of hydrodynamics in spherical symmetry: \begin{subequations} \beq \frac{\partial \rho}{\partial t} + \frac{1}{r^2} \frac{\partial (r^2 \rho v_r)}{\partial r}=0 ,\label{HDcont} \eeq \beq \frac{\partial \rho_j}{\partial t} + \frac{1}{r^2} \frac{\partial (r^2 \rho_j v_r)}{\partial r}=S_j ,\label{HDcontChem} \eeq \beq \frac{\partial (\rho v_r)}{\partial t} + \frac{1}{r^2} \frac{(\partial r^2 \rho v_r^2)}{\partial r} = -C_{\rm s}^2 \frac{\partial \rho}{\partial r} + \rho g_r , \label{HDmom} \eeq \beq g_r=-\frac{4 \pi G}{r^2} \int_0^r \! \rho r'^2 \, dr', \eeq \end{subequations} where $v_r$ is the radial velocity, $\rho$ is the total density of the gas, and $\rho_j$ is the density of each individual chemical species followed (listed in Table \ref{table1}). We solve a density continuity equation for each of the chemical species in the model (equation \ref{HDcontChem}). In these equations, $S_j$ are source/sink terms describing mass interchange between the species due to chemical reactions. The initial state is one of uniform density equal to $10^3 {\rm \, cm^{-3}}$ and zero velocity. The radius of the core is $0.4$ pc, making it significantly thermally supercritical. At the outer boundary, we require constant pressure and zero velocity. This way, the core experiences no influx of mass. \subsection{MagnetoHydroDynamic Core Models} We follow Basu \& Mouschovias (1994) in modeling the formation and contraction of prestellar cores in self-gravitating, magnetic, rotating, isothermal model clouds in the presence of ambipolar diffusion and magnetic braking. In the presence of an ordered magnetic field the cloud relaxes rapidly along field lines until thermal-pressure forces balance gravity, before any significant contraction perpendicular to the field lines takes place (Fiedler \& Mouschovias 1993; Desch \& Mouschovias 2001; Kunz \& Mouschovias 2010). Force balance along magnetic field lines is maintained even when the contraction becomes dynamic in the perpendicular direction. Consequently, the thin-disk approximation is adopted for the model cloud cores (Ciolek \& Mouschovias 1993). The model cloud is axisymmetric with the rotation axis coinciding with the axis of symmetry, which is taken to be the z-axis of a cylindrical polar coordinate system ($r,\phi,z$). The upper and lower surfaces of the disk are located at $z=\pm Z(r,t)$. The cloud is embedded in an ``external'' medium of constant thermal pressure $P_{\rm ext}$, and density $\rho_{\rm ext}$, and is threaded by a magnetic field which, far from the cloud, becomes uniform and equal to a ``reference'' value $B_{\rm ref}$. The thin disk evolution equations, obtained by integrating the two-fluid (plasma and neutrals) MHD equations over $z$, are \begin{subequations} \beq\label{thin1} \frac{\partial \sigma_{\rm n}}{\partial t} + \frac{1}{r}\frac{\partial (r \sigma_{\rm n} v_{{\rm n},r})}{\partial r} = 0, \eeq \beq\label{thin2} \begin{split} \frac{\partial (\sigma_{\rm n} v_{{\rm n},r})}{\partial t} + \frac{1}{r}\frac{\partial( r \sigma_{\rm n} v_{{\rm n},r}^2)}{\partial r} & = -C_{\rm eff}^2 \frac{\partial \sigma_{\rm n}}{\partial r} + \sigma_{\rm n} g_r \\ &+\frac{L_{\rm n}^2}{\sigma_{\rm n} r^3}+ F_{{\rm M},r} , \end{split} \eeq \beq\label{thin3} \frac{\partial L_{\rm n}}{\partial t} + \frac{1}{r}\frac{\partial( r L_{\rm n} v_{{\rm n},r})}{\partial r} = \frac{1}{2\pi} r B_{z,{\rm eq}} B_{\phi,Z}, \eeq \beq\label{thin4} \frac{\partial B_{z,{\rm eq}}}{\partial t} + \frac{1}{r}\frac{\partial (r B_{z,{\rm eq}} v_{{\rm i},r})}{\partial r} = 0, \eeq \beq\label{thin5} \rho_{\rm n}(r,t)C^2 = P_{\rm ext} + \frac{\pi}{2}G \sigma_{\rm n}^2(r,t), \eeq \beq\label{thin6} \sigma_{\rm n}(r,t) = \int_{-Z(r,t)}^{+Z(r,t)} \rho_{\rm n} dz = 2 \rho_{\rm n}(r,t)Z(r,t), \eeq \beq\label{thin7} C_{\rm eff}^2 = \frac{\pi}{2}G \sigma_{\rm n}^2 \frac{[3P_{\rm ext} + (\pi/2)G \sigma_{\rm n}^2]}{[P_{\rm ext} + (\pi/2)G\sigma_{\rm n}^2]^2} C^2, \eeq \beq\label{thin8} \begin{split} F_{{\rm M},r} &= \frac{1}{2\pi} \left\{ B_{z,{\rm eq}}\left( B_{r,Z}-Z\frac{\partial B_{z,{\rm eq}}}{\partial r}\right) + \frac{1}{2} \frac{\partial Z}{\partial r} \right .\\ &\times \left[ B_{r,Z}^2 + 2B_{z,{\rm eq}}\left(B_{r,Z}\frac{\partial Z}{\partial r}\right) \right .\\ &\left. \left. +B_{\phi,Z}^2 +\left(B_{r,Z}\frac{\partial Z}{\partial r}\right)^2 \right] \right\}, \end{split} \eeq \beq\label{thin9} B_{\phi,Z}(r) = - 2 \frac{(4\pi \rho_{\rm ext})^{1/2}}{B_{\rm ref}} \frac{\Phi}{r} (\Omega-\Omega_{\rm b}), \eeq \beq\label{thin10} B_{r,Z}(r) = - \int_{0}^{\infty}dr' r' [B_{z,{\rm eq}}(r')-B_{\rm ref}]\mathcal{M}(r,r'), \eeq \beq\label{thin11} g_{\rm r}(r) = 2\pi G \int_{0}^{\infty}dr' r' \sigma_{\rm n}(r')\mathcal{M}(r,r'), \eeq \beq\label{thin12} \begin{split} \mathcal{M}(r,r') =& \frac{d}{dr}\left[ \int_{0}^{\infty} dk J_0(kr)J_0(kr') \right]\\ =&\frac{2}{\pi}\frac{d}{dr}\left[ \frac{1}{r_>}K\left( \frac{r_<}{r_>}\right) \right], \end{split} \eeq \beq\label{thin13} \frac{\partial \sigma_{\rm j0}}{\partial t} + \frac{1}{r}\frac{\partial (r \sigma_{\rm j0} v_{{\rm n},r})}{\partial r} = S_{\rm j0}, \eeq \beq\label{thin14} \frac{\partial \sigma_{\rm j+}}{\partial t} + \frac{1}{r}\frac{\partial (r \sigma_{\rm j+} v_{{\rm i},r})}{\partial r} = S_{\rm j+}, \eeq \beq\label{thin15} v_{{\rm i},r} = v_{{\rm n},r} + \frac{<\tau_{\rm ni}>}{\sigma_n}F_{{\rm M},r}, \eeq \end{subequations} \noindent where $\sigma_{\rm n}$ and $\rho_{\rm n}$ are the column and volume densities of neutrals respectively, $v_{{\rm i},r}$ and $v_{{\rm n},r}$ are the radial ion and neutral velocities respectively, $L_{\rm n}=\sigma_{\rm n} v_{{\rm n},\phi} r$ is the angular momentum per unit area, and $\Phi=\int^r_0 dr' r'B_{z,{\rm eq}}$ is the magnetic flux. Equation (\ref{thin1}) is the mass continuity equation of the neutrals. Equation (\ref{thin3}) gives the $\phi$ component of the neutral force equation, while the $r$-component is given by Equation (\ref{thin2}). Equation (\ref{thin5}) states that the thermal pressure at the equatorial plane in each magnetic flux tube balances the external and gravitational pressures. The introduction of the effective sound speed, $C_{\rm eff}$, accounts for the contribution of the radial force exerted by the external pressure $P_{\rm ext}$ on the upper and lower surfaces of the flaring disk, since the latter are not horizontal. Furthermore, in the expression for the total radial magnetic force (per unit area), $F_{{\rm M},r}$, the first two terms are the magnetic tension force and the magnetic pressure force within the cloud, respectively. The remaining terms are corrections due to the variation of the half-thickness of the cloud, $Z$, with $r$. Equation (\ref{thin4}) is Faraday's law in the two-fluid approximation with the magnetic flux frozen in the ion fluid. The toroidal component of the magnetic field ($B_{\phi,Z}$) is determined assuming a steady state condition on the transport of angular momentum by torsional Alfv{\'e}n waves in the external medium and given by equation (\ref{thin9}). The radial component of the magnetic field at the top surface of the thin disk ($B_{r,Z}$) is given by equation (\ref{thin10}), obtained by assuming a force-free external medium, whereby the half-thickness of the disk ($Z$) is obtained from equation (\ref{thin6}) after $\rho_{\rm n}(r,t)$ is calculated from equation (\ref{thin5}). The $r$-component of the gravitational field (\ref{thin11}) is calculated from Poisson's equation for a thin disk. $J_0$ is the zeroth-order Bessel function of the first kind, $K$ is the complete elliptic integral of the first kind, $r_< = {\rm min}[r,r']$, and $r_> = {\rm max}[r,r']$. Equation (\ref{thin13}) is the mass continuity equation for neutral gas phase chemical species advected with the neutral bulk velocity $v_{{\rm n},r}$. Equation (\ref{thin14}) is the mass continuity equation for the gas phase ions and all the grain mantle species (since dust grains are overwhelmingly negatively charged in the densities we are studying) advected with the velocity of the ions $v_{{\rm i},r}$. $S_{\rm j0}$ and $S_{\rm j+}$ represent the source/sink terms due to chemical reactions. Note that the ions are advected with a different velocity than the neutrals, so two equations are required in the magnetic model to replace Eq.~(\ref{HDcontChem}) of the non-magnetic model. Equation (\ref{thin15}) gives the velocity of the ions and is obtained from the algebraic solution of the force equation for the ion fluid where the acceleration terms have been neglected because of the small inertia of these species. The mean (momentum exchange) collision time of a neutral particle with ions in equation (\ref{thin15}) is \beq <\tau_{\rm ni}>=1.56\frac{<\mu_{\rm i}>m_{\rm p}+2m_{\rm p}}{\rho_{\rm i} <\sigma w>_{\rm in}}, \eeq where $m_{\rm p}$ is the proton mass, $\rho_{\rm i}=\sum_{\rm j+} \rho_{\rm j+}$ is the total ion density, the mean molecular weight of the ions is \[<\mu_{\rm i}>=\sum_{\rm j+} m_{\rm j+}n_{\rm j+}/\sum_{\rm j+} n_{\rm j+}\,,\] $n_{\rm j+}$ is the number volume density for each one of the ions considered, and $<\sigma w>_{\rm in}$ is the average collisional rate between ions and hydrogen molecules. The value of $<\sigma w>_{\rm in}$ varies slightly with the mass of the ion (McDaniel \& Mason 1973); we use the canonical value $<\sigma w>_{\rm in}=1.69\times 10^{-9}$ ${\rm cm^3s^{-1}}$ (Ciolek \& Mouschovias 1993). Each magnetic model is initiated at a reference state as in Basu \& Mouschovias (1994). The reference state has a column density profile of the form \begin{equation} \sigma_{\rm n,ref}(r) = \frac{\sigma_{\rm c,ref}} {\left[ 1+(r/l_{\rm ref})^2\right]^{3/2}}\,, \end{equation} which implies an almost uniform column density for $r\ll l_{\rm ref}$, which is where the core forms. The high-$r$ tail has little effect on the calculations and is there for numerical convenience, allowing the calculation of the disk gravitational potential to converge. The profile of the reference state angular velocity is obtained by using a linear dependence of specific angular momentum on mass, corresponding to the angular momentum distribution of a uniformly rotating disk, yielding \begin{equation} \Omega_{\rm ref}(r) = 2\Omega_{\rm c,ref}\left(\frac{l_{\rm ref}}{r}\right)^2 \left[1-\frac{1}{\sqrt{1+(r/l_{\rm ref})^2}}\right]\,. \end{equation} The reference state is first allowed to relax to an equilibrium configuration, while ambipolar diffusion and chemistry are turned off. It is from this equilibrium state that the subsequent evolution is initiated. For this reason, the initial central density to which each model relaxes is not always the same but it depends on the initial mass to flux ratio (see Fig.~\ref{n_vs_t}). The free parameters in the magnetic models are the central density of the initial reference state (which we take to be $10^3 {\rm \, cm^{-3}}$); the background angular velocity $\Omega_b = 10^{-15} {\rm \, rad \, s^{-1}}$, approximately corresponding to the rate of galactic rotation in the solar neighborhood; the reference state angular velocity ($\Omega_{\rm ref} = 0.48 {\rm \, km \, s^{-1} \, pc^{-1}}$); the external pressure $P_{\rm ext} = 0.1 \pi G \sigma^2_{\rm c,ref}/2$ corresponding to $10\%$ of the central ``gravitational'' pressure; the characteristic size of the column density profile $l_{\rm ref} = 1 {\rm \, pc}$; the external density $\rho_{\rm ext} = 10 {\rm cm^{-3}}$; and the mass to magnetic flux ratio, which is varied (see discussion of parameter study below). With the exception of the magnetic mass to flux ratio, the dynamical evolution of the collapsing core is not sensitive to any of the free parameters (Basu \& Mouschovias 1995). At the outer boundary of the disk we require continuity of thermal as well as magnetic pressures (see Tassis \& Mouschovias 2005, \S 5). No influx of mass is allowed at the outer boundary, so the total mass of the cloud is conserved. \begin{table*} \begin{center} \caption{\label{table1} Chemical Species Considered} \resizebox{\textwidth}{!}{ \begin{tabular}{l l l l l l l l l l l l l l l l} \tableline\tableline \multicolumn{16}{c}{gas phase species}\\%\\ \hline \tableline ${\rm H^+}$ & H & $\rm H_2^+$ & ${\rm H_2}$ & ${\rm H_3^+}$ & He & ${\rm He^+}$ & ${\rm C^+}$ & C & CH & ${\rm CH^+}$ & ${\rm CH_2^+}$ & ${\rm CH_2}$ & N & ${\rm N^+}$ & ${\rm CH_3}$ \\ ${\rm NH^+}$ & ${\rm CH_3^+}$ & NH & ${\rm NH_2^+}$ & O & ${\rm CH_4}$ & ${\rm CH_4^+}$ & ${\rm O^+}$ & ${\rm NH_2}$ & ${\rm CH_5^+}$ & OH & ${\rm OH^+}$ & ${\rm NH_3^+}$ & ${\rm NH_3}$ & ${\rm H_2O}$ & ${\rm NH_4^+}$ \\ ${\rm H_2O^+}$ & ${\rm H_3O^+}$ & ${\rm C_2}$ & ${\rm C_2^+}$ & ${\rm C_2H^+}$ & ${\rm C_2H}$ & ${\rm C_2H_2^+}$ & ${\rm C_2H_2}$ & CN & ${\rm CN^+}$ & ${\rm HCN^+}$ & ${\rm C_2H_3^+}$ & HCN & HNC & ${\rm Si^+}$ & ${\rm C_2H_4^+}$ \\ ${\rm H_2NC^+}$ & Si & ${\rm N_2}$ & ${\rm CO^+}$ & ${\rm HCNH^+}$ & CO & ${\rm N_2^+}$ & HCO & ${\rm N_2H^+}$ & ${\rm HCO^+}$ & ${\rm H_2CO}$ & ${\rm H_2CO^+}$ & NO & ${\rm NO^+}$ & ${\rm H_3CO^+}$ & ${\rm CH_3OH}$ \\ ${\rm O_2}$ & ${\rm O_2^+}$ & ${\rm CH_3OH_2^+}$ & ${\rm C_3^+}$ & ${\rm C_3H^+}$ & ${\rm C_2N^+}$ & ${\rm CNC^+}$ & ${\rm C_3H_3^+}$ & ${\rm CH_3CN}$ & ${\rm C_3H_2}$ & ${\rm CO_2}$ & ${\rm CO_2^+}$ & ${\rm HCO_2^+}$ & ${\rm HC_3N}$ & \\ \tableline \multicolumn{16}{c}{grain mantle species}\\%\\ \hline \tableline H & C & CO & ${\rm H_2CO}$ & Si & ${\rm C_2}$ & ${\rm O_2}$ & CH & OH & NO & ${\rm CH_2}$ & ${\rm H_2O}$ & ${\rm CO_2}$ & ${\rm CH_3}$ & ${\rm CH_4}$ & HNC \\ ${\rm C_2H_2}$ & ${\rm HC_3N}$ & ${\rm N_2}$ & CN & NH & HCN & ${\rm C_2H}$ & ${\rm NH_3}$ & ${\rm CH_3CN}$ & ${\rm CH_3OH}$ & ${\rm NH_2}$ & N & O & ${\rm H_2}$ & HCO & ${\rm C_3H_2}$ \\ ${\rm CH_2OH}$ & & & & & & & & & & & & & & \\ \tableline \end{tabular}} \end{center} \end{table*} \begin{table} \begin{center} \caption{\label{table2} Initial Abundances} \begin{tabular}{l l} \tableline\tableline ${\rm H}$ & $1.00 \times 10^{-2}$ \\ ${\rm H_2}$ & $4.95 \times 10^{-1}$ \\ ${\rm He}$ & $1.40 \times 10^{-1}$ \\ ${\rm C^+}$ & $7.30 \times 10^{-5}$ \\ ${\rm N}$ & $2.14 \times 10^{-5}$ \\ ${\rm O}$ & $1.76 \times 10^{-4}$ \\ ${\rm Si}$ & $2.00 \times 10^{-8}$ \\ \tableline \tableline \end{tabular} \end{center} \end{table} \section{Chemical Network}\label{chemistry} We use a subset of the UMIST chemical reaction network (Woodall et al.\ 2007), which includes data on 420 species and 13 elements followed in gas-phase reactions. Each chemical species is treated as a separate fluid in our formalism. The source and sink terms on the right-hand side of Eq.~(\ref{HDcontChem}) and in Eqs.~(\ref{thin13}) and (\ref{thin14}) describe the interactions between chemical species. They are given by: \begin{eqnarray} S_j &=& \frac{\partial \rho_j}{\partial t} \nonumber \\ & = &\!\!\!\!\! \sum_{k\ne j, l \ne j} \!\!\!\!\! \alpha_{jkl} \rho_k \rho_l - \!\!\! \sum_{k\ne j, l} \!\!\!\! \alpha_{klj} \left( 1\!\!+\! \delta_{lj}\right)\rho_l \rho_j + \!\! \sum_{k\ne j} \!\! \left( \beta_{jk} \rho_k \!\! - \!\! \beta_{kj}\rho_j \right)\!, \nonumber \\ \end{eqnarray} where $\alpha_{klj}$ is the rate coefficient for the production of species $j$ due to the interaction of species $k$ and $l$, $\beta_{jk}$ is the rate of production of species $j$ from species $k$ by photoreactions, and $\delta_{lj}$ is the Kronecker delta function. In our models the source terms are implemented in an operator-split fashion, and the change of abundances due to the source terms is updated after the advection terms. The species we consider are listed in Table 1, and they include 78 gas-phase atomic and molecular species, and 33 grain ice mantle species, that are composed of the elements H, He, C, N, O, and Si as a representative low ionization potential metal. The chemical network follows 1553 reactions. All the models studied in this work use the isothermal approximation (small deviations from isothermality have been shown to not affect the dynamical evolution of cores significantly, e.g. Keto \& Caselli 2010). We adopt as a fiducial temperature $T=10$K, consistent with line intensity ratios of different molecular transitions showing that starless cores are almost isothermal, with $T\sim 10$ K, (Tafalla et al.\ 1998, 2002, 2004), although there may be small variations of the order of a few K (e.g., Evans et al.\ 2001). In our parameter studies, we test for the effect of small temperature variations on our chemical abundances. Since the prestellar cores we model are embedded in a molecular cloud we assume a visual extinction $A_{\rm V}=3$ mag at the outer boundary, so that the results are not much affected by photodissociation. For the inner layers of the core we obtain the visual extinction from the column density to the surface using the empirical conversion relation $N_{\rm H_2}/A_{\rm V}=9.4 \times 10^{20}$ ${\rm cm^{-2}mag^{-1}}$ (Bohlin et al.\ 1978). For our deeply embedded cores, no ${\rm H_2}$ self-shielding is required; CO self-shielding on the other hand is accounted for based on the model of Lee et al.\ (1996). We assume a standard MRN distribution of grains (Mathis, Rumpl \& Nordsieck 1977), assumed to be well-mixed with the gas. Since most of the grains are negatively charged at the densities of interest in prestellar cores, we also take into account the electrostatic attraction between ions and grains through an enhanced rate of accretion for the ions (Rawlings et al.\ 1992). We use a fixed grain abundance ($n_{\rm grain}/n_{\rm neutral} = 10^{-12}$) and charge (we take all grains to be singly negatively charged). Grain processes modeled include freeze-out, surface reactions, thermal desorption, and cosmic ray heating. We have used a sticking probability of 0.3. The discrete nature of grains means that chemistry on their surfaces should properly be modeled by stochastic methods (e.g.; Stantcheva, Shematovich, \& Herbst 2002; Chang et al.\ 2007). However, those methods are computationally very expensive, especially for a large chemical network and the large number of HD and MHD models we run. Instead, we use the rate equations method, with modifications as suggested by Vasyunin et al. (2009), which brings the results of this method into good agreement with Monte Carlo modeling for the range of temperatures we are concerned with. The grain reaction network is taken from Hasegawa \& Herbst (1993), as are the rates for desorption by cosmic ray heating of grains (updated for new binding energies where necessary). Where available, we use binding energies determined in the lab (see Willacy 2007). Other binding energies are taken from Hasegawa \& Herbst (1993). The ${\rm CO_2}$ chemistry on the grains is based on Garrod \& Pauly (2011)\footnote{ with the activation barrier for the ${\rm CO+O \rightarrow CO_2}$ reaction on the grain ice increased from 290 K to 340 K (within the experimental range).}. The gas phase reactions are better constrained as they have been studied for a longer time, and there appears to be convergence in the literature among different groups (e.g., Aikawa et al.\ 2003) with the exception of sulphur-bearing molecules (Wakelam, Herbst \& Selsis 2006), which, however, are not considered in our model. In our study we aim to assess the effect of the dynamics on chemical abundances, while the chemical network remains fixed. We use low-metal initial abundances\footnote{with the exception of Si, the abundance of which is higher than in the low-metal model}, given in Table 2. The calculations are initiated with all elements other than molecular hydrogen in the form of atoms. Of course in nature there is some chemical evolution taking place during the formation of the parent cloud. A first attempt to calculate initial abundances for molecular clouds for one particular cloud formation scenario was presented by Hassel et al. (2010). However, these results are model-dependent and will differ for different cloud formation models. We will return to the issue of initial abundances in a future publication. \section{Parameter Study}\label{parstu} For each class of models (non-magnetic or magnetic), we examine the dependence of the molecular abundances and their evolution on four parameters: the C/O ratio, the cosmic ray ionization rate, the temperature, and a parameter controlling the time available for chemical evolution. In non-magnetic models, the latter is a time interval prior to collapse during which chemistry evolves but the core is assumed to not evolve dynamically. In magnetic models, it is the initial mass-to-magnetic-flux ratio which controls the amount of magnetic support: a smaller initial mass to magnetic flux ratio implies a larger amount of magnetic support against gravity, and hence a slower dynamical evolution for the core. Our ``reference'' non-magnetic model has a C/O ratio of 0.4, a cosmic ray ionization rate of $\zeta = 1.3 \times 10^{-17} {\rm \, s^{-1}}$, a temperature $T=10 {\rm \, K}$ and an initial collapse delay time of 1 Myr. Our ``reference'' magnetic model similarly has a C/O ratio of 0.4, $\zeta = 1.3 \times 10^{-17} {\rm \, s^{-1}}$, $T=10 {\rm \, K}$, and a central mass-to-flux ratio equal to the critical value for collapse (Mouschovias \& Spitzer 1976). For non-magnetic models, we examine two additional values of delay: zero, and 10 Myr. For magnetic models, we examine two additional values of the initial mass to magnetic flux ratio: 1.3 times the critical value (a faster-evolving, magnetically supercritical model), and 0.7 of the critical value (a slower, magnetically subcritical model). The three values of the delay and the three values of the mass-to-flux ratio set our dynamical model variations. When these models are combined with reference values for the C/O ratio, $\zeta$, and $T$, we will refer to them as the six basic dynamical models. The carbon-to-oxygen ratio is varied from its reference value by keeping the abundance of C constant and changing that of O (as in Terzieva \& Herbst 1998). The two other values of C/O ratio examined are 1 and 1.2. We have varied each of the six basic dynamical models by changing its C/O ratio to one of the values above, and we have thus studied 12 models (6 magnetic and 6 non-magnetic) with the C/O ratio varied from its reference value. We have also considered different temperatures for each of the six basic dynamical models. We have varied the temperature by a factor of $\sim 1.5$ from its reference value of $10$ K and examined models with $T=7$ K and $T=15$ K, and we have thus studied 12 models (6 magnetic and 6 non-magnetic) with a temperature differing from its reference value. Finally, to test the effect of the cosmic ray ionization rate we have studied two additional magnetic and two non-magnetic models, which are, in each case, identical to the corresponding ``reference'' models, but with different cosmic-ray ionization rates: namely, a factor of four above ($\zeta = 5.2 \times 10^{-17}$ $s^{-1}$) and below ($\zeta = 3.3 \times 10^{-18}$ $s^{-1}$) its reference value (covering the range of observational estimates e.g.: MaCall et al.\ 2003; Hezareh et al.\ 2008). Thus, we have examined a total of 34 models (6 basic dynamical models, 12 with varied C/O ratios, 12 with varied $T$, and 4 with varied $\zeta$). These 34 models are depicted in our figures in \S \ref{res} with different line types, thicknesses, colors, and symbols, as follows: \begin{itemize} \item Line type (solid, dashed, dotted) and symbols (X, diamond, cross) depict the evolutionary timescale (``reference'', ``slow'', and ``fast'' respectively). \item Color denotes the type of model (magnetic or non-magnetic), with bluish hues corresponding to non-magnetic models, and reddish hues corresponding to magnetic models. \item The line thickness/symbol size (for red and blue lines and symbols) indicate the values of the cosmic ray ionization rates: thin red/blue solid lines and small red/blue symbols correspond to the low value of $\zeta$, while thick red/blue solid lines and large red/blue symbols to the high value of $\zeta$, with all other parameters fixed at their ``reference'' values. \item The effect of different C/O ratio values is shown as a shaded area (orange for magnetic runs, and cyan for non-magnetic runs). The edge of the shaded area in contact with the corresponding (``fast'', ``reference'', or ``slow'', magnetic or non-magnetic) model line corresponds to C/O=0.4, while the edge of the shaded area furthest away from the line corresponds to C/O=1.2 \item Finally, the effect of the temperature is shown with the colors brown for magnetic runs and purple for non-magnetic runs, either as a shaded area, or as line thickness/symbol size. In the latter case, the thick brown/purple lines (or the large brown/purple symbols) correspond to $T=15$ K and the thin brown/purple lines (small brown/purple symbols) to $T=7$ K. \end{itemize} \section{Results}\label{res} \subsection{Dynamics} \begin{figure} \plotone{plotnvst_center.eps} \caption{\label{n_vs_t} (a) Central number density, $n_{\rm c}$, as a function of time. Symbols indicate the times at which full radial snapshots are taken. Blue lines: non-magnetic models (dotted/crosses: no delay; solid/x: 1 Myr delay; dashed/diamonds: 10 Myr delay). Red lines: magnetic models (dotted/crosses: supercritical model; solid/x: critical model; dashed/diamonds: subcritical model). The thin and thick red solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value, and bracket the normal-thickness red solid line. The yellow shaded areas show the effect of a varying C/O ratio, and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. The brown shaded areas show the effect of varying the temperature from 7K to 15K, the reference value being 10K.} \end{figure} \begin{figure*} \plotone{plotdynamics_multi.eps} \caption{\label{comp_dynamics_rad} Radial profiles of: the number volume density, $n$ (panel a); velocities of the neutrals in units of the isothermal sound speed (panel b); and (edge-on for the magnetic runs) column density (panel c). Red lines: magnetic runs; Blue lines: non-magnetic runs. Each curve corresponds to a snapshot in time when the volume density is an order of magnitude higher than the previous one. Time elapsed per snapshot in magnetic runs: 3.53 Myr, 4.4 Myr, 4.6 Myr, 4.65 Myr; in non-magnetic runs: 1.88 Myr, 2.01 Myr, 2.05 Myr, 2.06 Myr.} \end{figure*} Figure \ref{n_vs_t} shows the evolution of the central density of contracting core models with time. Red lines correspond to magnetic runs, and blue lines to non-magnetic runs. The different magnetic runs correspond to different values of the mass-to-magnetic-flux ratio: critical for the solid lines/ x symbols; 0.7 times the critical value (magnetically subcritical, slower-evolving magnetic model) for the dashed lines/diamonds; and 1.3 times the critical value (magnetically supercritical, faster-evolving magnetic model) for the dotted lines/ + symbols. The different non-magnetic runs correspond to different assumed delay times, during which the chemistry operates, before the core is allowed to evolve dynamically (collapse): 1 Myr delay for the solid lines/ x symbols; 10 Myr delay (slowest-evolving non-magnetic model) for the dashed lines/diamonds; and no delay (fastest-evolving non-magnetic model) for the dotted lines/ x symbols. The symbols mark the times when full radial outputs are obtained and analyzed, and correspond to successive enhancements of the central density by an order of magnitude. Different dynamical models differ with respect to the two fundamental ingredients affecting the chemical evolution of a core: density, and time spent at each density. Non-magnetic models with substantial time delays (representing the time it takes for some level of turbulent support to dissipate) can take as much time as substantially magnetically subcritical models to reach very high densities. However, once the collapse sets in (support dissipates), the evolution occurs essentially within one free-fall timescale, so that the bulk of the time delay is spent at the initial, low density ($10^3 {\rm cm ^{-3}}$). In contrast, slow magnetic models allow the core to spend substantial time at {\em higher} densities, and as a result the chemical state of these models are expected to be different. The thin and thick red lines correspond to magnetically critical models but have $\zeta$ a factor of four above and below the ``reference'' value, and they bracket the normal-thickness red solid line. Because the ambipolar diffusion timescale depends on the degree of ionization, which in turn depends on the cosmic-ray ionization rate, $\zeta$ affects the evolutionary timescale of the magnetic models: a higher degree of ionization implies increased coupling between magnetic field and neutral fluid, and a resulting slower evolution, as magnetic forces are more effective in opposing gravity (see Paper II). In addition to the mass-to-flux ratio and the degree of ionization, the evolutionary timescale of the magnetic models is also affected by the C/O ratio. The effect of different values of the C/O ratio can be seen in Fig.~\ref{n_vs_t} as the orange shaded area, which denotes a range of C/O values between 0.4 (reference) and 1.2. The reason for this dependence is that the C/O value affects the chemical reaction network and consequently the degree of ionization at each density, resulting to a different evolution rate. In general, higher values of the C/O ratio result in slower evolution, and the effect is more pronounced for models with a low mass-to-flux ratio. Finally, the evolutionary timescale of the magnetic models also depends on the temperature. The effect of different $T$ values is shown in Fig.~\ref{n_vs_t} as the brown shaded area, which denotes a range of $T$ values between $7$ K and $15$ K. The dependence is stronger for slower models, and it enters through the dependence of the degree of ionization on temperature (see Paper II). The evolutionary timescale increases with increasing temperature, and the difference between $10$ K and $15$ K is significantly larger than the difference between $7$ K and $10$ K, although the fractional increase is comparable (about 50\% in each case). Radial profiles of the density, velocity, and (edge-on for the magnetic runs) column density for the magnetic and non-magnetic reference models and for the different time snapshots corresponding to the x symbols in Fig.~\ref{n_vs_t} are given in Fig.~\ref{comp_dynamics_rad}. Again the magnetic model is represented by red lines and the non-magnetic model by blue lines. Each snapshot corresponds to a time when the volume density is an order of magnitude higher than the previous one. The time elapsed between the initialization of each simulation and each snapshot for the magnetic model is 3.53 Myr, 4.4 Myr, 4.6 Myr, 4.65 Myr. The corresponding times for the non-magnetic model are 1.89 Myr, 2.02 Myr, 2.05 Myr, 2.06 Myr. The first snapshot in both cases corresponds to the initial configuration of each model. Radial profiles of the volume density are shown in the left panel. These are power-law profiles with a flat inner region, the boundary of which is indicated with a circle for the magnetic runs and a square for the non-magnetic runs. Profiles of the infall velocity, in units of the sound speed, are shown in the middle panel. Here the difference between the magnetic and the non-magnetic models is most striking, as for the magnetic model the infall is always subsonic, while for the non-magnetic model the infall velocities are substantially supersonic throughout the evolution of the core and for most of its radial extent. The right panel shows radial profiles of the column density, which have similar properties with the volume density profile. The masses of the different model cores are not exactly the same. This is because the mass of each core depends on the temperature in the case of non-magnetic cores, and the mass-to-flux ratio, temperature, C/O ratio, and cosmic-ray ionization rate in the case of magnetic cores. In spite of these dependencies, we have selected the initial conditions of our models so that the core masses remain roughly comparable: the non-magnetic ``reference'' model core mass is $15 {\rm \, M_\odot}$ and the magnetic ``reference'' model core mass is $20 {\rm \, M_\odot}$, while in general the core masses in each class of models do not differ from this value by more than a factor of $\sim 2$. A shared feature of the volume and column density radial profiles with important observational implications is the size of the flat-profile inner region. It is well-known that prestellar cores have uniform column density inner regions (Ward-Thompson et al.~1994; Ward-Thompson et al.~1999; Bacmann et al.~2000; Shirley et al.~2000; Schnee et al.~2010). The angular size of the {\em column density} profile flat inner region, which is observationally measurable in well resolved cores, combined with an estimate for the distance can yield the physical size of the column density profile flat inner region, which is of the same size as the {\em volume density profile flat inner region}, as can be directly seen by comparing the left and the right panels of Fig.~\ref{comp_dynamics_rad}. The latter can be used to estimate the central volume density of a collapsing core (Tassis \& Yorke 2011). \begin{figure*} \plotone{testNEQ.eps} \caption{\label{NEQ} Ratio of total (including all possible channels) production and destruction rates as a function of central density, for the magnetic (left panel) and the non-magnetic (right panel) reference models. The inset in the left panel shows a zoom-in of the density range between $10^4$ and $10^6$ $cm^{-3}$, with a linear y axis between $P/D=0.5$ and $1.5$. } \end{figure*} \subsection{Non-equilibrium Chemistry, Production and Destruction of Commonly Studied Molecules} \begin{figure*} \plotone{plotcenterreac_short.eps} \caption{\label{reacfig} Top panels: central abundances for two commonly observed molecules (CO and NH$_3$), for the non-magnetic (left column) and magnetic (right column) reference models. Solid lines: gas phase abundance; dashed lines: grain ice mantle abundance. Middle and bottom panels: percentage contribution of most important reactions to the total production (black lines) and destruction (red lines) rates for the two molecules.} \end{figure*} The need for the inclusion of non-equilibrium chemistry in our dynamical models is explicitly demonstrated in Fig. \ref{NEQ}. Here we plot the ratio $P/D$ between the total production and destruction rates of various molecules, as a function of density, at the center of our reference core models. The left panel corresponds to the magnetic reference model, and the right panel to the non-magnetic reference model. The total production and destruction rates are defined as including all chemical processes leading to the production or destruction of a specific molecule respectively. Each line corresponds to a different molecule. Note that the y-axis in the two plots is logarithmic and extends over ten orders of magnitude. A large number of molecules are very much out of equilibrium ($P/D$ is very far from 1), and they remain so throughout each simulation. In the magnetic run, where more time is available for the system to approach chemical equilibrium even at high densities, the majority of molecules eventually move closer to $P/D =1$; in the non-magnetic model the reverse trend is observed, and as the density increases and the evolution timescale decreases, most molecules deviate further away from $P/D = 1$. However, even for the magnetic model the spread about $P/D=1$ is substantial, as demonstrated by the inset on the left panel, which zooms in around central densities of $10^4 - 10^6$ particles per cm$^3$ and $P/D$ within 50\% of unity displayed on a linear scale. In the inset, the values of $P/D$ exhibit a spread of more than 20\% around unity. The effect of non-equilibrium chemistry becomes obvious when we examine the dependence of central molecular abundances on central density for different dynamical models. If the chemistry was in equilibrium, then the central volume density would uniquely determine the central molecular abundances. This, however, is not the case, as we can see in Fig. \ref{reacfig}, where we examine the central abundances, production, and destruction rates due to various reactions for one carbon-bearing (CO) and one nitrogen-bearing (NH$_3$) molecule. The left column corresponds to the non-magnetic reference model (1 Myr delay), and the right column corresponds to the magnetic reference model (critical mass-to-flux ratio). The top row shows the evolution with central density (bottom axis) and with time (top axis) of the central abundances of CO (blue lines) and NH$_3$ (purple lines). Solid lines correspond to gas-phase abundances and dashed lines correspond to abundances on grain ice mantles. The middle and bottom rows show the most important production (black lines) and destruction (red lines) reactions for the two molecules. The importance of each reaction is quantified by the percentage by which the specific reaction contributes to the total production/destruction rate of each molecule (y-axis). Comparing the top left and right panels, we can immediately see that the central abundances of the two molecules at the same central density are very different for different dynamical models, a result of the chemistry being out of equilibrium. Because each model spends different amounts of time at each density (see Fig. \ref{n_vs_t}), non-equilibrium chemistry results to significantly different central abundances. Note that since the magnetic model evolves more slowly, the gas phase abundances of both molecules in the magnetic core model decrease more steeply with density, as the molecules freeze out on the grain mantles; correspondingly, there is a faster increase in the ice mantle abundance of each molecule. It is evident from the middle and bottom panels that any specific reaction reaches its maximum importance at different densities for different dynamical models. Freeze out onto grain mantles (solid red line), which becomes increasingly important at higher densities, is a prominent example. Furthermore, for each molecule different reactions dominate the production and destruction channels for different dynamical models. \subsection{Evolution of central abundances}\label{evca} \begin{figure*} \plotone{plotcentabund_big_final.eps} \caption{\label{comp_cen_abund} Central abundances for representative molecules as a function of central density, for different dynamical models, different C/O ratios, and different cosmic ray ionization rates. Reddish hues correspond to magnetic models (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and bluish hues to non-magnetic models (dashed: 10 Myr delay; solid: 1 Myr delay; dotted: no delay). The thin and thick solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value. The orange/cyan shaded areas show the effect of a varying C/O ratio in magnetic/non-magnetic models and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. } \end{figure*} \begin{figure*} \plotone{plotcentabund_T_final.eps} \caption{\label{T_cen_abund} Effect of temperature on central abundances of representative molecules. Blue/red lines: reference non-magnetic/magnetic models. Thick/thin brown lines: $T=15$ K/ $T=7$ K magnetic models; all other parameters kept at their reference values. Thick/thin purple lines: $T=15$ K/ $T=7$ K non-magnetic models; all other parameters kept at their reference values. } \end{figure*} \begin{figure} \plotone{statT10T15.eps} \caption{\label{twostatic} Evolution of N$_2$H$^+$ abundance for $T=10$ K (black line) and $T=15$ K (red line) in a static medium of density $2\times 10^4 {\rm , cm^{-3}}$. } \end{figure} \begin{figure*} \plotone{plotcentabundG_big_multi.eps} \caption{\label{gmices} As in Fig.~\ref{comp_cen_abund}, for grain mantle ices. Reddish hues correspond to magnetic models (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and bluish hues to non-magnetic models (dashed: 10 Myr delay; solid: 1 Myr delay; dotted: no delay). The thin and thick solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value. The orange/cyan shaded areas show the effect of a varying C/O ratio for magnetic/non-magnetic models and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. The brown/purple shaded areas show the effect of varying temperature for magnetic/non-magnetic models and correspond to a range of $T$ values between $7$ K and $15$ K (reference value is $10$ K). } \end{figure*} Figures \ref{comp_cen_abund} and \ref{T_cen_abund} show the evolution with central density of the central abundances of selected molecules in the gas phase. Figure \ref{comp_cen_abund} shows the effect of the dynamics, C/O ratio, and $\zeta$ on the gas-phase abundances, while Fig.~ \ref{T_cen_abund} shows the effect of the temperature. In Fig.~\ref{comp_cen_abund}, reddish hues correspond to magnetic models and bluish hues correspond to non-magnetic models. Different line types correspond to different dynamical models, with dashed lines representing the ``slowest'' models (subcritical magnetic model and 10 Myr delay non-magnetic model); solid lines representing the ``intermediate'' models (critical magnetic model and 1 Myr delay non-magnetic model); and dotted lines representing the ``fast'' models (supercritical magnetic model and no-delay non-magnetic model). The shaded bands (cyan for non-magnetic models, orange for magnetic models) show the effect of the C/O ratio on the abundances (the edge of the band adjacent to a model line corresponds to C/O=0.4 and the edge furthest away from the line to C/O=1.2). The line thickness is representative of the value of the cosmic ray ionization rate, with normal thickness lines corresponding to $\zeta = 1.3 \times 10^{-17} {\rm \, s^{-1}}$, and thick (thin) lines corresponding to values of $\zeta$ a factor of four above (below) that value. A general feature of the central gas-phase abundance evolution curves of Fig.~\ref{comp_cen_abund} is their dependence on the dynamical evolution model. Magnetic models, which evolve more slowly between different central densities and spend more time at each value of the density, have an abundance peak at low densities, and generally exhibit a faster drop of the gas-phase abundance with increasing central density, due to efficient freeze-out onto grain ice mantles. As a result, the abundance evolution curves appear convex. In contrast, non-magnetic models evolve faster, chemistry has less time in which to operate at each density, the abundances peak and drop at higher densities, and the non-magnetic abundance evolution curves are generally concave. The quantitative explanation of this behavior can be given by comparing the dependence of the freeze-out timescale and the collapse timescale on density for different models. The steep drop in molecular abundances (especially in carbon-bearing molecules) occurs at the density where the freeze-out timescale becomes shorter than the collapse timescale (Walmsley 1991). In magnetic models this happens at much lower densities ($\sim 10^4 {\rm \, cm^{-3}}$) than in non-magnetic models ($\sim 10^6 {\rm \, cm^{-3}}$). The results of Fig.~\ref{comp_cen_abund} make it obvious that different collapse delay times at the initial phases of non-magnetic core evolution are not a good proxy for the effect of magnetic fields. Although the slowest non-magnetic model curves (dashed curves, 10 Myr delay) are generally closer to the magnetic model curves, the difference is still significant even in the qualitative behavior of the abundances (concave vs convex). The reason for this is that allowing more time for chemical evolution at $\sim 10^3 {\rm \, cm^{-3}}$ does not have the same effect, from a chemistry perspective, as spending a similar amount of time at higher densities, which is what happens in magnetic models. The behavior of certain molecules, such as H$_2$CO, H$_3^+$, NH, and CH$_3$, is quite different for the non-magnetic models. While the slowest non-magnetic and all the magnetic models have a qualitative similar behavior that resembles that of other molecules (the abundance builds up with increasing density, peaks at $\sim 10^4 {\rm \, cm^{-3}}$, and then decreases at even higher density), the faster non-magnetic models, instead of a peak, exhibit a local minimum at these densities. The molecules with the greatest sensitivity to the C/O ratio are, unsurprisingly, the oxygen-bearing molecules, the abundance of which is generally lower for higher C/O ratios. Molecules with more than one oxygen atoms are affected more than molecules with a single oxygen atom (see, e.g., CO versus CO$_2$). Other molecules sensitive to the C/O ratio include CH, CH$^+$, CH$_3$, and NH. As expected, H$_3^+$ responds directly to the change in the cosmic-ray ionization rate, as it is a direct ionization product (see Paper II for more details). Molecules that are primarily produced through reactions involving direct ionization products, such as H$_3^+$ and He$^+$ (e.g., CH$^+$) or free electrons, (e.g., NH$_2$, NH, and CH$_3$) are similarly affected, even in the non-magnetic models. The magnetic models have an additional response to $\zeta$ through the modification of their evolutionary timescale (see Fig.~\ref{n_vs_t}). For this reason, the change in molecular abundances, especially in the case of the largest-$\zeta$ magnetic model (which is also by far the slowest-evolving model) can be dramatic and feature very significant depletion in relatively low densities. The biggest difference is seen in molecules which, for moderate evolutionary timescales, show relatively modest depletion (e.g. N-bearing molecules). Figure \ref{T_cen_abund} shows the effect of the temperature on the evolution with density of the central abundances of a few representative molecules. The blue and red lines correspond to the reference non-magnetic and magnetic models (the same ones shown with the normal-thickness blue and red lines in Fig.~\ref{comp_cen_abund}). High-temperature ($T=15$ K) models are shown with the thick brown and purple lines, for magnetic and non-magnetic models respectively, while low-temperature ($T=7$ K) models are shown with the thin brown and purple lines; all other parameters are kept at their reference values. There are some significant differences between the molecular abundances in the 15K models compared to those in the 10K models, especially for nitrogen-bearing compounds. This is a result of two effects: (1) the reaction \begin{equation}\label{react:nhplus} \hbox{N}^+ + \hbox{H}_2 \longrightarrow \hbox{NH}^+ + \hbox{H} \end{equation} has a small activation barrier of 85 K. At 15 K the temperature is high enough for this barrier to be overcome and hence this reaction is much more efficient than at 10K (see also Wakelam et al. 2010). This reaction is the start of the formation of nitrogen-bearing molecules, and an increase in its efficiency results in higher abundances of these molecules. (2) At 15 K the grains are warm enough for the desorption of N$_2$ ice to proceed more rapidly than at 10K (the rate of desorption is approximately 10$^{10}$ times greater at 15 K compared to 10K). The way in which these processes act to produce the model results depends on the conditions in the models. As can be seen in Figure \ref{T_cen_abund}, increasing the temperature to 15K results in an increase in the abundance of nitrogen bearing molecules for the non-magnetic models (bold purple curves). The opposite effect is seen for the magnetic models (bold brown curves). The difference between the two types of models is due to the timescales over which they evolve. The non-magnetic models in Fig. \ref{T_cen_abund} evolve quickest and reach their final density within 2 Myr, whereas the magnetic models take between 5.3 and 5.5 Myr (Fig.~\ref{n_vs_t}). Taking the non-magnetic models first: there is an increase in gaseous N$_2$ caused by its faster desorption from the grains (see Fig.~\ref{gmices}). This means that the N$_2$ abundance in the gas is higher at earlier times. The increase in N$_2$ increases the destruction rate of H$_3^+$ as they react forming N$_2$H$^+$, and results in the sharp decrease in H$_3^+$ abundance seen in Fig.~\ref{T_cen_abund}. Most of the N$_2$H$^+$ will be recycled back into N$_2$ by recombination with electrons, but a small amount ($\sim$ 5\%; Molek et al.~2009) will form NH. This can go on to form other gaseous molecules, or it can freeze out onto the grains, where it is quickly hydrogenated to form ammonia ice. Ammonia has a much higher binding energy than N$_2$ and once formed remains on the grains at these temperatures. The effect of the enhanced rate for reaction~(\ref{react:nhplus}) can also be seen in the increased abundance of NH in the gas phase in the 15 K model. The magnetic models are indicated by red/brown lines in Fig.~\ref{T_cen_abund}, where the bold brown line is the 15 K model. The increased temperature here has exactly the opposite effect on nitrogen molecular abundances compared to the non-magnetic models. In this case the abundance of N$_2$H$^+$ falls at 15 K, as do the abundances of most other nitrogen-bearing species. The abundance of H$_3^+$ is roughly the same for all temperature models. The reason for this variance is the differing dynamical timescales of the models considered. The magnetic models remain at a relatively low density for $>$ 2 Myr, before the onset of rapid collapse. During this prolonged period of time the chemistry evolves quite differently depending on the temperature. This is illustrated in Fig.~\ref{twostatic} which shows the evolution of the N$_2$H$^+$ abundance with time for two static models with a density of 2 $\times$ 10$^4$ cm$^{-3}$ and with temperatures of 10 K (black) and 15 K (red). At 15 K the nitrogen chemistry evolves faster, with X(N$_2$H$^+$) reaching a peak value of $\gtrsim 10^{-13} {\rm \, cm^{-3}}$ at 1 Myr, compared to $9\times 10^{-14} {\rm \, cm^{-3}}$ at almost 2 Myr for the 10 K model. In the magnetic models, rapid collapse does not begin until after $\sim$4 Myr. From Fig.~\ref{twostatic} we can see that the N$_2$H$^+$ abundance is falling rapidly at this time and is more than an order of magnitude higher at 10 K than at 15 K. At 15 K much more nitrogen has been incorporated into NH$_3$ ices by this stage in magnetic models (see Fig.~\ref{gmices}) and less is available in the gas phase. Hence the early evolution of the collapse is quite different and results in the different abundances seen in Fig.~\ref{T_cen_abund}. Figure \ref{gmices} shows the behavior of central abundances of various molecules in grain ice mantles, as a function of central density. Colors and lines are as in Fig.~\ref{comp_cen_abund}, however here the effect of temperature is also overplotted with a brown shaded area for the magnetic models, and a purple shaded area for the non-magnetic models. In all models, the composition of ice mantles is dominated by H$_2$O, in accordance with observations of absorption features in the spectra of background stars viewed through dense clouds taken in IR (e.g., Whittet 2003 and references therein). In observed dark, quiescent regions (where cores are most likely to be prestellar and the effect of the protostellar UV field is minimal), CO and CO$_2$ are also observed to be important constituents of ice mantles: the contributions relative to the abundance of water ice is at the level of 10-40\% for CO (Chiar et al.\ 1995; Gibb et al.\ 2004; Chiar et al.\ 2011; \"{O}berg et al.\ 2011) and 20-30\% for CO$_2$ (Whittet et al.\ 2009; Gibb et al.\ 2004; \"{O}berg et al.\ 2011). Our models are in general agreement with these observations. There are, however, a few exceptions. The CO ice abundance in the 10 Myr delay time non-magnetic model shows a great sensitivity to temperature, with the 15 K model abundance being much lower (more than 3 orders of magnitude) than the rest of the models and observations. Similarly, the CO$_2$ abundance is decreased by much compared to observations for the 7 K versions of the 10 Myr delay and 1 Myr delay non-magnetic models, and all of the magnetic models. In observed clouds CH$_3$OH is also found at the few \% level relative to the water ice abundance (Boogert et al.\ 2011; \"{O}berg et al.\ 2011); in our models the CH$_3$OH abundance is at the order of $\lesssim 1\%$. Nitrogen-bearing ices are dominated in our models by NH$_3$ ($\lesssim 10\%$ of the water ice abundance), whereas CN ice is very underabundant, consistent with observations (Gibb et al.\ 2004, Chiar et al.\ 2011). NH$_3$ ice is shown to be a good chronometer, with its abundance almost monotonically increasing with the evolutionary timescale. HCN is systematically more abundant than HNC. N$_2$ ice is very sensitive to temperature as discussed above, and its abundance decreases with increasing temperature. Molecular oxygen ice abundance has a sensitive dependence on the elemental C/O ratio (cyan and yellow bands for magnetic and non-magnetic models respectively), and its abundance decreases as oxygen elemental abundance decreases, as expected. It is this dependence that dominates the range in possible values for the O$_2$ ice abundance. The evolutionary trends with central density followed in our models by various mantle ice abundances are generally similar: the abundance builds up fast as a result of efficient molecule freeze-out onto grains and high enough (A$_{\rm V} > 3$) extinction, and saturates at higher values of the central density. \subsection{Radial Abundance Profiles} \begin{figure*} \plotone{plotradabund_46.eps} \caption{\label{comp_rad_abund} Radial profiles of gas-phase abundances of three commonly observed molecules, taken when the core has reached a central density of $10^4 {\rm \, cm^{-3}}$ (upper row) and $10^6 {\rm \, cm^{-3}}$ (lower row). Reddish hues correspond to magnetic models (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and bluish hues to non-magnetic models (dashed: 10 Myr delay; solid: 1 Myr delay; dotted: no delay). The thin and thick solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value. The orange/cyan shaded areas show the effect of a varying C/O ratio for magnetic/non-magnetic runs, and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. The brown/purple shaded areas show the effect of varying temperature for magnetic/non-magnetic models and correspond to a range of $T$ values between $7$ K and $15$ K (reference value is $10$ K).} \end{figure*} Figure \ref{comp_rad_abund} shows radial profiles of gas-phase abundances of representative molecules, taken when the core has a central density of $10^4 {\rm \, cm^{-3}}$ (upper row) and $10^6 {\rm \, cm^{-3}}$ (lower row). Colors and lines are the same as in Fig.~\ref{gmices}. At high densities (lower row), both magnetic and non-magnetic models have experienced significant central depletion in all molecules (although generally to different extent). This is consistent with the qualitative picture of Fig.~\ref{comp_cen_abund}: at high densities the molecular abundances in all models have decreased significantly. At lower densities however (upper row), magnetic models exhibit a qualitatively different behavior from non-magnetic models, at least for certain molecules. The difference is most striking in the case of CO, which for non-magnetic models, is centrally-peaked at a central density of $10^4 {\rm \, cm^{-3}}$. In contrast, at the same density, magnetic models have already suffered significant depletion at the center. In the case of NH$_3$ and CN, however, the radial profiles are quite similar (centrally peaked or flat) in both magnetic and non-magnetic models, even at low densities. As we will see in \S \ref{DM}, this effect can be used to discriminate between non-magnetic and magnetic models. An important point to note is that the plots in Fig.~\ref{comp_rad_abund} are in a logarithmic scale, and, as a result, most of the core mass is located in the larger scales depicted in the plots. For this reason, even in the cases where some or all models have undergone significant central depletion, when a core is not resolved the observed abundances sample mostly the outer layers where little if any depletion has occurred. This results in degeneracies among models with otherwise significant differences in central depletion, when a core is unresolved. This point is prominently demonstrated in \S \ref{totab}. \begin{figure*} \epsscale{0.8} \plotone{plottotabund_big_final.eps} \caption{\label{comp_tot_abund} Abundances {\em mass-averaged over the entire core} for representative molecules at four different stages in the core's evolution (quantified by the core central density), for different dynamical models, different C/O ratios, and different cosmic ray ionization rates. Reddish hues correspond to magnetic models (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and bluish hues to non-magnetic models (dashed: 10 Myr delay; solid: 1 Myr delay; dotted: no delay). The thin and thick solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value. The orange/cyan shaded areas show the effect of a varying C/O ratio for magnetic/non-magnetic models, and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. The brown/purple {\em lines} show the effect of varying temperature for magnetic/non-magnetic models. Thin brown/purple lines: $T = 7$ K; thick brown/purple lines: $T = 15$ K (reference value is $10$ K). } \end{figure*} \begin{figure*} \plotone{plotncritabund.eps} \caption{\label{ncritt_abund} As in Fig.\ \ref{comp_tot_abund}, but mass-averaged only over regions of the core with densities above the critical density $n_{\rm crit}$ for excitation of commonly observed lines for each molecule: CO(J=1$\rightarrow$0), 115 GHz, $n\ge n_{\rm crit} =2.18\times 10^3 {\rm \, cm^{-3}}$; HCO$^+$(J=1$\rightarrow$0) 89.2 GHz, $n \ge n_{\rm crit} =1.63\times 10^5 {\rm \, cm^{-3}}$; NH$_3$(J,K=1,1), 23 GHz, $n \ge 10^5 {\rm \, cm^{-3}}$, see text; N$_2$H$^+$(J=1$\rightarrow$0), 93.17 GHz, $n \ge n_{\rm crit} =2.25\times 10^5 {\rm \, cm^{-3}}$; HCN(JF=12$\rightarrow$01), 88.631 GHz, $n \ge n_{\rm crit} =2.18\times 10^6 {\rm \, cm^{-3}}$; CN(N=1$\rightarrow$0), 113.17 GHz, $n \ge n_{\rm crit} =1.3\times 10^6 {\rm \, cm^{-3}}$. } \end{figure*} \subsection{Total Abundances -- Unresolved Cores}\label{totab} Figure \ref{comp_tot_abund} shows the total gas-phase abundances of selected molecules in the entire core (i.e. abundances mass-averaged over the entire extent of the ``unresolved'' core), as a function of evolutionary stage, quantified by the core central density (x-axis). More advanced evolutionary stages are to the right side of the horizontal axis. The effect of C/O ratio is shown with the shaded orange and cyan regions (for magnetic and non-magnetic models respectively), while the effect of temperature is shown with thin/thick brown and purple lines (thin for $T=7$ K, thick for $T=15$ K, brown for magnetic models, purple for non-magnetic models). The effect of $\zeta$ is only shown for the reference magnetic and non-magnetic models, and is represented by the thickness of the solid blue and red lines (thicker line corresponds to higher $\zeta$). A striking feature of the total gas-phase abundances is that there is little evolution with central density. The cause of this effect is the dominant contribution of the outer layers of the core (the lower-density regions) to the total core mass and their weight in determining the total core abundances as would be observed in unresolved cores. Since the abundances in the outer layers of the cores do not evolve much, the total core abundances also remain relatively stable with evolutionary stage. Because of the power-law abundance profiles, this feature is generally preserved even if one averages only over radii with densities above the critical density for excitation of a specific molecular line instead of over the entire core. Of course, the fact that a specific line is observed already places a lower limit on the central density of the core, equal to $n_{\rm crit}$ for the specific line. However, little evolution of the abundance with time is seen beyond that limit, unless $n_{\rm crit}$ is very high ($\ge 10^6 {\rm \, cm^{-3}}$). This can be seen in Fig.\ \ref{ncritt_abund} , where we have plotted abundances of selected molecules averaged over regions with densities higher than the critical density $n_{\rm crit}$ for the following molecular lines: CO(J=1$\rightarrow$0), 115 GHz, $n\ge n_{\rm crit} =2.18\times 10^3 {\rm \, cm^{-3}}$; HCO$^+$(J=1$\rightarrow$0) 89.2 GHz, $n \ge n_{\rm crit} =1.63\times 10^5 {\rm \, cm^{-3}}$; NH$_3$(JK=11), 23 GHz, $n \ge 10^5 {\rm \, cm^{-3}}$[Although the critical density for this transition is much lower ($\sim 2\times 10^3 {\rm \, cm^{-3}}$), this line requires much higher densities before the line is thermalized (Evans 1989)]; N$_2$H$^+$(J=1$\rightarrow$0), 93.17 GHz, $n \ge n_{\rm crit} =2.25\times 10^5 {\rm \, cm^{-3}}$; HCN(JF=12$\rightarrow$01), 88.631 GHz, $n \ge n_{\rm crit} =2.18\times 10^6 {\rm \, cm^{-3}}$; and CN(N=1$\rightarrow$0), 113.17 GHz, $n \ge n_{\rm crit} =1.3\times 10^6 {\rm \, cm^{-3}}$. These critical densities are for $T =10$ K. All molecular data are taken from the Leiden Atomic and Molecular Database (Sch\"{o}ier et al. 2005).\footnote{http://www.strw.leidenuniv.nl/$\sim$moldata/} The large range of possible abundance values for different models at a fixed evolutionary stage (central density) demonstrates the large potential of chemical differentiation in prestellar cores. Even cores that appear identical in, e.g., their total column density profiles, can exhibit large variations in their molecular abundances as a result of a diversity in other local physical parameters (temperature, ionization rate, initial elemental abundance ratios, and dynamics). However, the large variation in molecular abundances among different models is neither monotonic nor systematic, and the abundances resulting from different parameters are not well-separated. Instead, we often see a nonlinear dependence on certain parameters (a characteristic example is the ionization rate), as well as largely overlapping ranges, resulting in strong model degeneracies. We conclude that discrimination between models is not straight-forward, but instead requires the use of a variety of different observations and constraints, in combination with a statistical study accounting for as many different molecular abundances as possible. In the next sections, we will discuss certain quantities, such as depletion measures and abundance ratios, which also hold promise as tools for the discrimination among dynamical models and physical parameters. We caution the reader that the abundances in Fig.~\ref{comp_tot_abund} are not to be compared directly with abundances derived from observations beyond an uncertainty of a factor of at least several. There are various reasons for possible discrepancies. First of all, the mass-averaged abundances presented here are not appropriate when the core is at least partly resolved; in this case, a more detailed study is needed, starting with the radial profile at a specific evolutionary snapshot for each model, and properly accounting for the beam size and the physical scale it corresponds to at the distance of the core. Second, chemical reaction rates have appreciable uncertainties (Wakelam et al.~2010). We have not studied the effect of such uncertainties here, as our purpose was to investigate the effect of varying the core dynamics and other physical parameters on molecular abundances. Finally, obtaining a molecular abundance from observations involves several approximations as well as assuming a value for the H$_2$ column density (which is not always well-constrained). Nevertheless, we have verified, for the following cores, that molecular abundances obtained from observations are generally within our model ranges, or only deviate from it within uncertainties: L1498 (Tafalla et al.~ 2004, 2006); L1517B (Tafalla et al.~2004, 2006; Crapsi 2005); L1544 (Caselli et al.~1999; Jorgensen et al.~2004); L134N (Wakelam et al.~2006); TMC-1(CP) (Smith, Herbst \& Chang 2004); L1689B (Jorgensen et al.~2004). For an improved comparison with observations, the appropriate methodology would be to start from our dynamical/chemical models, perform radiative transfer calculations, and directly compare predicted line intensities with observed ones. We plan to return to such a comparison in a future study. \subsection{Depletion Measures}\label{DM} \begin{figure*} \epsscale{0.8} \plotone{plotX30thXtot_big_final.eps} \caption{\label{delta} Central deviation of molecular abundance with respect to the average abundance over the entire core ($\Delta$, see text), as a function of central density. The dotted line corresponds to $\Delta = 1$ (no central enhancement or depletion). Reddish hues correspond to magnetic models (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and bluish hues to non-magnetic models (dashed: 10 Myr delay; solid: 1 Myr delay; dotted: no delay). The thin and thick solid lines correspond to cosmic ray ionization rate $\zeta$ a factor of four above and below the ``reference'' value. The shaded areas show the effect of a varying C/O ratio, and correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. The brown/purple shaded areas show the effect of varying temperature for magnetic/non-magnetic models and correspond to a range of $T$ values between $7$ K and $15$ K (reference value is $10$ K).} \end{figure*} The qualitatively different behavior between the radial profiles of magnetic and non-magnetic models seen in Fig.~\ref{comp_rad_abund} can be used to discriminate between models. We define the {\em central deviation} $\Delta$ (central depletion or enhancement) of the molecular abundance as the ratio between the integrated abundance through the central $30\%$\footnote{We have confirmed that the results are similar if this fraction is changed to 10\% or 20\%.} of the radial extent of the core over the integrated abundance over the entire core. Figure \ref{delta} shows the dependence of $\Delta$ on evolutionary stage (here quantified by the value of the central density) for all models (colors and lines as in Fig.~\ref{comp_rad_abund}). The horizontal dotted line marks the value $\Delta =1$. Values of $\Delta$ above the line indicate that the particular molecule is centrally enhanced. Values of $\Delta$ close to the line indicate a flat abundance profile, while values of $\Delta$ below the line correspond to centrally depleted molecules. Despite the significant overlap between magnetic and non-magnetic models both in the radial profiles and the total abundances discussed in Figs.~\ref{comp_rad_abund} and \ref{comp_tot_abund} respectively, for many molecules the magnetic and non-magnetic models not only are well separated, but also show little sensitivity to model parameters such as the evolutionary timescale, the temperature, the C/O ratio, and the cosmic-ray ionization rate. Molecules which show such good separation between magnetic and non-magnetic models include CO, CO$_2$, HCO$^+$, NO, OH, H$_2$), and HC$_3$N. In contrast, H$_3^+$ and CN show a particularly degenerate behavior between different models. \begin{figure} \plotone{plottotabundratio_C2O_single.eps} \caption{\label{ratioC2O} Gas-phase abundance ratio between HCN and OH, mass-averaged over the entire core, which has good potential for discriminating between C/O ratio values. Red/blue symbols correspond to magnetic/non-magnetic models with $T=10$ K. Diamonds': ``slow'' models; X: ``reference'' models; crosses: ``fast'' models. The size of the red/blue symbols represents the value of $\zeta$ (larger symbols correspond to larger $\zeta$). Brown/purple symbols correspond to magnetic/non-magnetic models with $T=7$ K (small symbols) and $T=15$ K (large symbols), and a ``reference'' value for $\zeta$. The C/O ratio of each model is shown on the vertical axis, while the value of the abundance ratio is shown on the horizontal axis. Points correspond to a central $H_2$ density of $10^6 {\rm cm^{-3}}$, however the core-average abundance ratios evolve little with density.} \end{figure} \begin{figure*} \plotone{plottotabundratio_time.eps} \caption{\label{ratioTime} Gas-phase abundance ratios, mass-averaged over the entire core, with high potential for discrimination among evolutionary timescales for the core. Diamonds': ``slow'' models; X: ``reference'' models; crosses: ``fast'' models. Red/blue symbols correspond to magnetic/non-magnetic models with $T=10$ K and C/O =0.4. The size of the red/blue symbols represents the value of $\zeta$ (larger symbols correspond to larger $\zeta$). Orange/cyan points correspond to magnetic/non-magnetic models with $T=10$ K, C/O=1.2, and a ``reference'' value for $\zeta$. Brown/purple symbols correspond to magnetic/non-magnetic models with $T=7$ K (small symbols) and $T=15$ K (large symbols), C/O=0.4, and a ``reference'' value for $\zeta$. The time it takes the core to reach a central density of $10^6 {\rm cm^{-3}}$ is shown on the vertical axis, while the value of the abundance ratio is shown on the horizontal axis. Points correspond to a central $H_2$ density of $10^6 {\rm cm^{-3}}$, however the core-average abundance ratios evolve little with density.} \end{figure*} For non-magnetic models, CO-bearing molecules, and in particular CO, CO$_2$ and HCO$^+$, tend to show no depletion ($\Delta>1$) at early stages of the collapse, and only become depleted when the central density reaches $10^5$ - $10^6 {\rm \, cm^{-3}}$. In contrast, for magnetic models, the same molecules are already centrally depleted by central densities of $\sim 10^4 {\rm \, cm^{-3}}$. NH$_3$ on the other hand shows no sign of central depletion at central densities of $\sim 10^4 {\rm \, cm^{-3}}$ for both magnetic and non-magnetic models. It only becomes depleted at later stages of the collapse ($10^6$ - $10^7 {\rm \, cm^{-3}}$ for non-magnetic models, and $10^5$ - $10^6 {\rm \, cm^{-3}}$ for magnetic models). N$_2$H$^+$ shows a similar behavior, however its behavior is complicated by its sensitivity to temperature in the case of non-magnetic models (see also discussion in \S \ref{evca}). Note also that in all N-bearing molecules, the high cosmic ray ionization rate magnetic model is always depleted, as a result of the very long evolutionary timescale (see also discussion of Fig.\ref{comp_cen_abund}). This difference in behavior between CO and NH$_3$ is consistent with observations (e.g., Caselli et al.~1999; Bacmann et al.~2003; Crapsi et al.~2007). \subsection{Abundance ratios of molecules with discriminatory potential} Ratios between abundances of different molecules have been long considered as promising tools to overcome model uncertainties and degeneracies. However, most abundance ratios exhibit large variations and no consistent trend with any single model parameter when other parameters are varied. We have identified several gas-phase abundance ratios (averaged over the entire core) that are most affected by a single model parameter and show little variation as a result of the rest and thus show promising discriminatory potential. We plot these in Figs.~\ref{ratioC2O} and \ref{ratioTime}. Figure \ref{ratioC2O} shows the values of the abundance ratio between HCN and OH, mass-averaged over the entire core, for our various models. This ratio can serve as an indicator of the C/O ratio. Different models are plotted with different symbols as discussed in the caption. The C/O ratio, shown in the vertical axis, exhibits a trend (increases with increasing value of each abundance ratio). Similarly, Figure \ref{ratioTime} shows three abundance ratios that trace the evolutionary timescale of a core. These ratios are ${\rm X_{NH_3}/X_{CO}}$, ${\rm X_{NH_2}/X_{CO}}$, and ${\rm X_{NH_3}/X_{HCO^+}}$. All the ratio values are taken at the time when the central density in each model has reached $10^6 {\rm \, cm^{-3}}$. However, the averaged abundance over the entire core shows little time evolution, as shown in Fig.~\ref{comp_tot_abund}. Each abundance ratio examined in these two figures shows an appreciable scatter at a fixed value of the parameter that it can help constrain (C/O ratio or evolutionary timescale, respectively). However, a general trend is obvious, and there are significant parts of the C/O or timescale parameter space that can be excluded with measurements of each ratio. \section{Discussion and Conclusions}\label{thedisc} We have combined dynamical and non-equilibrium chemical modeling of both magnetic and non-magnetic contracting molecular cloud cores, in order to predict molecular abundances as a function of core evolutionary stage, and identify observables with high promise in discriminating among core models. To this end, we have undertaken an extensive parameter study, following a total of 34 different models, and varying: \\ (a) the core evolutionary timescale, controlled by the {\em collapse delay time}, in the case of non-magnetic models, and the initial {\em mass-to-magnetic-flux ratio}, i.e. the amount of magnetic support against gravity, in the case of magnetic models; \\ (b) the elemental C/O ratio in the core; \\ (c) the cosmic-ray ionization rate, which affects the chemistry in both magnetic and non-magnetic models, but also the evolutionary timescale in magnetic models, by affecting the extent of coupling between the magnetic field and the neutral fluid; and \\ (d) the temperature. Our main results can be summarized as follows. We have explicitly demonstrated that many reactions in our chemical network are out of equilibrium during the dynamical evolution of the core, necessitating the inclusion of non-equilibrium chemistry and the coupled modeling of chemistry and dynamics to properly follow the molecular abundances. For different dynamical models, the molecular abundances at the center of the core are not the same, even for identical values of the central number density; rather, the amount of time that a model core has spent {\em at each density} is what determines the actual molecular abundances at each evolutionary stage. Therefore it is not possible to properly account for magnetic effects in core chemistry by using a collapse retardation factor in non-magnetic models. Even if the overall final age of the core is the same, the fact that the non-magnetic core model has spent very different fractions of this age at various values of the central density than a magnetic core model has, results in significantly different molecular abundances. Abundances in magnetic models peak at much lower densities and decrease much faster at higher densities. The evolution of grain mantle ice abundances is qualitatively similar in magnetic and non-magnetic models, with a fast initial increase due to freeze-out, and an abundance saturation at higher densities and later times. The value of that saturation however is different in different models. We find that the abundance of NH$_3$ ice scales roughly with the evolutionary timescale of the model, varying by a factor of 4 for our model range, which corresponds to an evolutionary timescale range of about 10 Myr. NH$_3$ ice is thus a potentially important chronometer. Radial profiles of gas-phase abundances reveal that the most significant difference between magnetic and non-magnetic models is at the central parts of the core, while most of the core extent tends to evolve much less and is much more similar among different models. However, as only a small fraction of the mass is concentrated at the central layers of the core, the central differences are smeared out in unresolved cores. As a result, the evolution of the total, mass-averaged molecular abundances with central density in any particular model is much milder, if at all present. The situation is similar with the total abundances of grain mantle ices. Differences in the C/O ratio tend to affect mostly the abundance of the oxygen-bearing molecules, while the cosmic-ray ionization rate $\zeta$ has its largest effect on the nitrogen-bearing molecules. In the case of magnetic models, the direct effect of the cosmic-ray ionization rate on chemistry is compounded with its indirect effect on core evolution timescale, and as a result magnetic models are much more sensitive to changes in $\zeta$. The many, sometimes competing, sometimes compounded, effects of different model parameters on molecular abundances result in severe degeneracies among different models. For this reason, we have identified {\em molecular abundance ratios} which may help to constrain, even in unresolved cores, specific model parameters: the evolutionary timescale (ratios between NH$_3$ and CO; NH$_2$ and CO; NH$_3$ and HCO$^+$) and the C/O ratio (ratio between HCN and OH). In addition, we have demonstrated that the {\em central deviation} $\Delta$ (central depletion or enhancement) of molecules can be another useful tool in model discrimination. While for many molecules values of $\Delta$ are degenerate among different models, there are several molecules for which there is a clear separation between magnetic and non-magnetic models. For these molecules the value of $\Delta$ even shows very little sensitivity to other model parameters. Measurements of the central depletion of such molecules therefore have an excellent potential for constraining dynamical models of core evolution. For this reason, it is very important to place adequate emphasis into understanding differential depletion and grain chemistry effects as much as possible, and thus gain confidence in the chemistry part of these calculations, which still features significant uncertainties (eg., Wakelam et al.\ 2010). The dynamical models we have used in this study have low dimensionality: the non-magnetic models are spherically symmetric (one-dimensional), while the magnetic ones are axially symmetric, and integrated over the z-direction (1.5 dimensions). We note that as a result of the distinct geometries, the magnetic model does not converge to the non-magnetic model in the limit of zero magnetic field. A zero magnetic field disk model cannot be followed with the current numerical scheme, as the thin-disk approximation breaks down in the absence of a magnetic field. Although higher-dimensional calculations are desirable (e.g., van Weeren et al.~2009), it is currently not possible to perform a parameter study as extensive as the one undertaken here using higher-dimensional models: a single model in this study (including non-equilibrium chemistry) requires more than 800 CPU hours for the magnetic runs and more than 200 CPU hours for the non-magnetic runs. We are however in the process of implementing true 2D calculations for select cases presented here, to assess the effect of low dimensionality on our calculations. Because the model cores are not multi-dimensional, multiplicity or fragmentation is by definition neither accounted for nor allowed. This however is not a severe disadvantage in our case, as multiplicity at these early stages of collapse and for the low densities considered here (up to $10^7 { \rm \, cm^{-3}}$) is not observed (Schnee et al.~2010). Finally our work has not accounted for turbulent mixing; the flows in our model are laminar. Any turbulent mixing would result in making the radial profiles shallower (Xie et al. 1995). However, dense molecular cloud cores are generally not observed to have significant amounts of turbulence since their line widths are approximately sonic (Myers 1983; Barranco \& Goodman 1998; Goodman et al.\ 1998; Kirk, Johnstone, \& Tafalla 2007). \acknowledgements{We thank Paul Goldsmith, Talayeh Hezareh, and the anonymous referee for insightful comments that have improved this paper. This work was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. \copyright 2012. All rights reserved.}
\section{Introduction} The \emph{visibility polygon} of a simple polygon $\Poly$ from a viewpoint $q$ is the set of all points of $\Poly$ that can be seen from $q$, where two points $p$ and $q$ can see each other whenever the segment $pq$ is contained in $\Poly$. The visibility polygon is a fundamental concept in computational geometry and one of the first problems studied in planar visibility. The first correct and optimal algorithm for computing the visibility polygon from a point was found by Joe and Simpson\cite{js-clvpa-87}. It computes the visibility polygon from a point in linear time and space. We refer the reader to the survey of O'Rourke \cite{r-v-04} and the book of Gosh~\cite{g-vap-07} for an extensive discussion of such problems. In this paper we look for an algorithm that computes the visibility polygon of a given point and uses few variables. This kind of algorithm not only provides an interesting trade-off between running time and memory needed, but is also useful in portable devices where important hardware constraints are present (such as the ones found in digital cameras or mobile phones). In addition, this model has direct applications in concurrent environments where several devices with limited memory resources perform some computation on a large centralized input. Since several devices may access the input at the same time, allowing writing to the input memory can result in compromising its integrity. A significant amount of research has focused on the design of algorithms that use few variables, some of them even dating from the 80s \cite{mp-ssls-80}. Although many models exist, most of the research considers that the input is in some kind of read-only data structure. In addition to the input values, we are allowed to use few additional variables (typically a variable holds a logarithmic number of bits). One of the most studied problems in this setting is that of selection. For any constant $\epsilon \in (0,1)$, Munro and Raman \cite{mr-sromswmdm-96} gave an algorithm that runs in $O(n^{1+\epsilon})$ time and uses $O(1/\epsilon)$ variables. Frederickson~\cite{Frederickson87} extended this result to the case in which $s$ working variables are available (and $s\in \Omega (\log n)\cap O(2^{\log n/\log^* n})$). Raman and Ramnath~\cite{rr-iubtstsls-98} gave several exact and approximation algorithms for the case in which fewer variables are available. Among other results, they provide a $2/3$-approximation of the median that runs in $O(sn^{1+1/s})$ time using $O(s)$ variables (for $s\in o(\log n)$), or $O(n\log n)$ time, using $O(\log n)$ variables. More recently Chan \cite{Chan} provided several lower bounds for performing selection with few variables. In recent years there has been a growing interest in finding algorithms for geometric problems that use a constant number of variables. An early example is the well-known gift-wrapping algorithm (also known as Jarvis march~\cite{j-ich-73}), which can be used to report the points on the convex hull of a set of $n$ points in $O(n\Hout)$ time using a constant number of variables, where \Hout\ is the number of vertices on the convex hull. Recently, Asano and Rote~\cite{ar-cwagp-09} and afterwards Asano {\em et al.} \cite{amw-cwaspsp-10,abbkmrs-mcasp-11} gave efficient methods for computing well-known geometric structures, such as the Delaunay triangulation, the Voronoi diagram, a polygon triangulation, and a minimum spanning tree (MST) using a constant number of variables. These algorithms run in $O(n^2)$ time (except computing the MST, which needs $O(n^3)$ time). Observe that, since these structures have linear size, they are not stored but reported. Prior to this work, there was no algorithm for computing the visibility polygon in memory-constrained models. Indeed, this problem was explicitly posed as an open problem by Asano {\em et al.}~\cite{amrw-cwagp-10} for the case in which only a constant number of variables are allowed. \subsubsection*{Results} In this paper we present a novel approach for computing the visibility polygon of a given point inside a simple polygon. It is easy to see that reflex vertices have a much larger influence on the visibility polygon than convex vertices. Therefore, whenever possible we express the running time of our algorithms not only in terms of $n$, the complexity of $\Poly$, but also in terms of $\Rin$ and $\Rout$ (the number of reflex vertices of $\Poly$ that are present in the input and in the output, respectively). This approach continues a line of research relating the combinatorial and computational properties of polygons to the number of their reflex vertices. We refer the reader to \cite{bdhilm-ghsc-07,akpv-got-12,bchm-gwp-10} and references found therein for a deep review of existing similar results. In Section~\ref{sec:Preliminaries} we begin the paper with some preliminaries, followed by some observations and basic algorithms in Section~\ref{sec:SimpleAlgorithm}. In Section \ref{section:SequentialAlgorithms} we give an output-sensitive algorithm that reports the vertices of the visibility polygon in $O(n\Rout)$ time using $O(1)$ variables. Using this algorithm as a stepping stone, in Section \ref{Section:Divide and conquer} we present a divide-and-conquer algorithm. This algorithm runs in $O(\frac{nr}{2^{s}}+n\log^2 r)$ time (or $O(\frac{nr}{2^{s}}+n\log r)$ randomized expected time) using $O(s)$ variables (for any $s\in O(\log r)$), giving an exponential trade-off between running time and space. {\bf Remark:} prior to this research there was no known method for computing visibility polygons using few variables. Following the conference version of this paper~\cite{bkls-cvpufv-11}, De {\em et al.}~\cite{dmn-seavpsp-12} provided a linear-time algorithm that uses $O(\sqrt{n})$-variables. Parallel to this research, Barba {\em et al.}~\cite{bklss-sttosba-13} gave a general method for transforming stack-based algorithms into memory constrained workspaces. Since Joe and Simpson's algorithm for computing the visibility polygon~\cite{js-clvpa-87} is stack-based, their approach can be used for this problem as well. \iffalse \begin{table} \centering \begin{tabular}{|c|c||l|} \hline Space & Time & Notes \\ \hline $\Theta(1)$ & $O(n\Rout)$ & Thm.~\ref{theoImprov}. Output Sensitive. \\ \hline $O(\log r)$ & $O(\frac{nr\log s}{s2^{s}})$ & Thm.~\ref{theo_determ}. \\ \hline $\Omega(\log r)\cap O(r)$ & $O(n\log^2\Rin)$ & Thm.~\ref{theo_determ}. \\ \hline $\Omega(\log r)\cap O(r)$ & $O(n\log r)$ & Thm.~\ref{theo_randomi}. Randomized expected time. \\ \hline $\Omega(n)$ & $O(n)$ & Given in~\cite{js-clvpa-87} . \\ \hline \end{tabular} \caption{Running time of the known memory-constrained algorithms for computing the visibility polygon as a function of the working space $s$. Unless otherwise stated, running times are deterministic.} \label{tabresults} \end{table} \fi \section{Preliminaries} \label{sec:Preliminaries} \subsubsection*{Model definition and considerations on input/output precision} We use a slight variation of the constant workspace model, introduced by Asano and Rote~\cite{ar-cwagp-09}. In this model the input of the problem resides in a read-only data structure and we are allowed to perform random access to any of the values of the input in constant time. An algorithm can use a constant number of variables and we assume that each variable or pointer contains a data word of $O(\log n)$ bits. Implicit storage consumption required by recursive calls is also considered part of the workspace. This model is also referred as {\em log space}~\cite{AB09} in the complexity literature. Many other similar models have been studied. We note that in some of them (like the {\em streaming}~\cite{gk-seocqs-01} or the {\em multi-pass} model~\cite{cc-mpga-07}) the values of the input can only be read once or a fixed number of times. As in the constant workspace model of Asano and Rote~\cite{ar-cwagp-09}, our model allows scanning the input as many times as necessary. However, our model differs from theirs in two aspects: we are allowed to use a workspace of $O(s)$ variables (instead of $O(1)$), and we do not require random access to the vertices of the input. The input to our problem is a simple polygon $\Poly$ in a read-only data structure and a point $q$ in the plane, from where the visibility polygon needs to be computed. We do not make any assumptions on whether the input coordinates are rational or real numbers (in some implicit form). The only operations that we perform on the input are determining whether a given point is above, below or on a given line and determining the intersection point between two lines. In both cases, the line is defined as passing through two points of the input, hence both operations can be expressed as finding the root of linear equations whose coefficients are values of the input. We assume that these two operations can be done in constant time. Moreover, if the coordinates of the input are algebraic values, we can express the coordinates of the output as ``the intersection point of the line passing through points $p_i$ and $p_j$ and the line passing through $p_k$ and $p_l$" (where $p_i,p_j,p_k$ and $p_l$ are vertices of the input). \subsubsection*{Definitions and basic properties} We say that a point $p$ is \emph{visible} from $q$ (with respect to $\Poly$) if and only if the segment $pq$ is contained in $\Poly$ (note that we regard $\Poly$ as a closed subset of the plane). The set of points visible from $q$ is called the \emph{visibility polygon} of \Poly, and is denoted \Vis. Note that if $q$ is outside of \Poly\ then \Vis\ is by definition empty. Thus, when considering visibility with respect to polygons, we always assume that $q$ is inside the polygon. From now on, we assume that $q$ is fixed, hence we omit the ``with respect to $q$" term. We assume we are given $\Poly$ as a list of its vertices in counterclockwise order along its boundary (denoted by $\bd$). Let $p_0$ be a point on $\bd$ closest to $q$ on the horizontal line passing through $q$. It is easy to see that $p_0$ is visible and can be computed in linear time. In the following, we will treat $p_0$ as a vertex of \Poly\ (even though it does not need to be one). By implicitly reordering the vertices of the input, we can assume that we are given the vertices of $\Poly$ in counterclockwise direction starting from $p_0$ (i.e., $\Poly = (p_0, \ldots, p_n)$). For simplicity of exposition, we assume that the vertices of $\Poly$ are in a weak general position; that is, we assume that there do not exist two vertices $p,p'\in\Poly$ such that $p, p'$, and $q$ are aligned (but we note that the algorithms can be extended easily for the general case). Along this paper, we will often work with polygonal chains (instead of polygons). However, we will restrict our scope to polygonal chains contained in $\bd$. For any two points $a,b$ on $\bd$, there is a unique path from $a$ to $b$ that travels counterclockwise along $\bd$; let $\C = \chain(a,b)$ be the set of points traversed in this path, including both $a$ and $b$ (this set is called the {\em chain} between $a$ and $b$). We say that $a$ and $b$ are the \emph{endpoints} of $\mathcal C$, and we refer to the rest of the points on $\mathcal C$ as its \emph{interior points}. We now extend the concept of visibility to chains. Due to technical reasons, we define this concept for chains contained in $\bd$ whose endpoints are visible from $q$. We say that a chain $\C$ is \emph{independent} if and only if its endpoints are visible points of $\Poly$. Given a chain $\C=\chain(a,b)$ with endpoints $a$ and $b$, let \regionC\ denote the polygon enclosed by the union of \C\ and the segments $qa$ and $qb$ (equivalently, we use the notation $\region(a,b)$). Given a point $q\in \mathbb{R}^2$ and an independent chain \C\ with endpoints $a$ and $b$, we say that a point $p\in \mathbb{R}^2$ is \emph{visible} from $q$ with respect to \C\ if and only if $p$ is visible from $q$ with respect to \regionC. The set of points that are visible with respect to $\C$ is called the \emph{visibility polygon} of $\C$, and is denoted \VisC. We start by observing that both concepts of visibility are equivalent (hence we need not distinguish between them). \begin{observation} If $\mathcal C = \chain(a,b)$ is an independent subchain, then \regionC\ is a simple polygon. Moreover, a point $x\in \C$ is visible with respect to $\Poly$ if and only if it is visible with respect to $\C$. \end{observation} The above observation certifies that visibility within independent chains is well-defined. Since we will only consider independent chains, from now on we omit the ``with respect to $\Poly$" (or to $\C$), and simply say that a point $p$ is visible. True to its name, the visibility polygon of an independent chain $\C \subseteq \bd$ can be computed without having knowledge of the remainder of $\Poly$. \begin{observation}\label{DivideObs} Let $\mathcal C = \chain(a,b)$ be an independent chain such that $\C = (a, p_i, \ldots, p_j, b)$ for some $0<i<j<n$. Let $x$ be a visible interior point of $\C$ lying on the edge $p_k p_{k+1}$ for some $i\leq k < j$. If we let $\mathcal C_1 = \{a, p_i, \ldots, p_k, x\}$ and $\mathcal C_2 = \{x, p_{k+1}, \ldots, p_j,b\}$, then $$\VisC = \mathrm{Vis}_{\mathcal C_1}(q) \cup \mathrm{Vis}_{\mathcal C_2}(q),$$ $$\mathrm{Vis}_{\mathcal C_1}(q) \cap \mathrm{Vis}_{\mathcal C_2}(q) = qx.$$ \end{observation} Given a point $p$ on the plane, let $\rho_q(p)$ be the ray emanating from $q$ and passing through $p$. We define $\theta_q(p)$ as the angle that $\rho_q(p)$ makes with the positive $x$-axis, $0\leq \theta_q(p) < 2\pi$. We call $\theta_q(p)$ the \emph{CCW-angle} of $p$. \begin{figure}[tb!] \centering \includegraphics{polygon} \caption{Left: general setting, vertices that are reflex with respect to $q$ are shown with a white point (black otherwise). Right: the visibility polygon $\Vis$. } \label{fig:polygon} \end{figure} We also need to define what a \emph{reflex} vertex is in our context. Given any vertex $p_k$, the line $\ell_k$ passing through $p_k$ and $q$ splits $\Poly \setminus \ell_k$ into disjoint components. A vertex $p_k$ is \emph{reflex with respect to $q$} if the angle at the vertex interior to $\Poly$ is more than $\pi$ and the vertices $p_{k-1}$ and $p_{k+1}$ lie on the same connected component of $\mathbb{R}^2\setminus \ell_k$ (see Fig. \ref{fig:polygon}). Observe that any vertex that is reflex with respect to $q$ is a reflex vertex (in the usual sense), but the converse is not true. Since the point $q$ is fixed, from now on we omit the ``with respect to $q$" term and simply refer to these points as reflex. Note that being reflex is a local property that can be verified in $O(1)$ time. Intuitively speaking, reflex vertices with respect to $q$ are the vertices where important changes occur in the visibility polygon. That is, where the polygon boundary can change between visible or not-visible. Let $\Rin$ be the number of reflex of vertices of $\Poly$. We also define $\Rout$ as the number of reflex vertices of $\Poly$ that are present in $\Vis$. Naturally, we always have $\Rout\leq \Rin < n$. Given two points $p$ and $p'$ on a chain $\C=\chain(a,b)$, we say that $p$ lies \emph{before} $p'$ (resp. $p'$ lies \emph{after} $p$) if, when we walk from $a$ towards $b$ along $\C$, we first pass through $p$ and then through $p'$. We say that a chain is {\em visible} if all the points of the chain are visible. A visible chain $\C=\chain(a,b)$ is \emph{CCW-maximal} if no other visible chain starting at $a$ strictly contains ${\mathcal C}$. In this case, we say that $\C$ is the maximal chain {\em starting} at $a$ and {\em ending} at $b$. Given a visible reflex vertex $v$ on a chain $\C$, we say that a point $w\neq v$ on $\bd$ is the \emph{shadow} of $v$ if $w$ is collinear with $q$ and $v$, and $w$ is visible from $q$ (this point is denoted by $\textsc{Shadow}$($v,\C$)). Due to the general position assumption, $w$ is uniquely defined and must be an interior point of an edge. That is, each visible reflex vertex is uniquely associated to a shadow point (and vice versa). We say that a visible reflex vertex $v$ is of \emph{type R} (resp. \emph{type L}) if its shadow lies after (resp. before) $v$; see Fig.~\ref{fig:RightLeftReflex and ConeDefs} (left). Equivalently, a vertex $v$ is of type R if $q,v,x$ and $q,v,y$ make a \emph{right} turn (where $x$ and $y$ are the predecessor and the successor of $v$ on $\C$, respectively). Analogously, vertices of type L make a \emph{left} turns instead. A \emph{ray shooting query} is a basic operation that, given an independent chain \C\ and a point $x\in \C$, considers the ray $\rho_q(x)$ and reports the last visible point in $\rho_q(x)$ with respect to \C\ (i.e., the one furthest away from $q$). Observe that when $x$ is a visible reflex vertex we obtain its shadow. We denote the output of this operation by \textsc{RayShooting}$(x, \mathcal{C})$. It is easy to see that a ray shooting query can be performed in linear time, using only $O(1)$ extra variables, by scanning the edges of \C\ one by one and computing their intersections with $\rho_q(x)$. Finally, for $\C = \chain(a,b)$ we define $\coneC$ as the cone with apex $q$ that contains every point in the plane with CCW-angle between $\theta_q(a)$ and $\theta_q(b)$; see Fig.~\ref{fig:RightLeftReflex and ConeDefs} (right). \section{Understanding the visibility polygon} \label{sec:SimpleAlgorithm} The basic scheme of our algorithms is to partition the input polygon into independent subchains so that the visibility within one subchain is unaffected by the others. We will use $\bd$ as the starting chain, thus the first chain will be closed, but the following chains will be open. Notice that since \Poly\ is simple, any chain of it will be simple too. In this section we present some observations about the independence between chains. \begin{figure}[tb] \centering \includegraphics{RightLeftReflex} \caption{\small Left: $v$ is a reflex vertex of type R, while $v'$ is of type L. The chain between $v$ and $\textsc{Shadow}(v,\mathcal C)$, together with the segment joining them, bound a simple polygon containing no visible points other than $v$ and $\textsc{Shadow}(v,\mathcal C)$. Observe that the chain $\mathcal{C}(\textsc{Shadow}(v,\mathcal C),\textsc{Shadow}(v',\mathcal C))$ is CCW-maximal. Right: A polygonal chain $\mathcal C = \chain(a,b)$ and its associated polygon $\region(a,b) = \regionC$. Point $x$ is visible while points $y$ and $z$ are not. Note that every visible point of $\C$ lies inside $\coneC$.} \label{fig:RightLeftReflex and ConeDefs} \end{figure} \begin{observation}\label{obs:VisibleChains} Let $v$ be a visible reflex vertex of $\Poly$ of type R (resp. L) whose shadow is $w$. The $\chain(v, w)$ (resp. $\chain(w, v)$) contains no visible point other than its endpoints. In particular, a visible chain cannot contain an interior reflex vertex. \end{observation} The following important lemma characterizes the endpoints of CCW-maximal chains. \begin{lemma}\label{lem_key} Let $\C =\chain(a,b)$ be an independent chain, let $p\in \C$ be a visible point and let $v$ the first visible reflex vertex encountered when walking from $p$ towards $b$ (or $b$ if none exists). Let $\chain(p,p')$ be the CCW-maximal chain starting at $p$. The point $p'$ is either equal to $v$ (if $v=b$ or $v$ is of type R), or equal to the shadow of $v$ (if $v$ is of type L). \end{lemma} \begin{proof} Clearly, all the points lying after $p$ are visible if and only if $p'=b$. If $p' \neq b$, then, when walking on $\C$, $p'$ is the last visible point of the chain before a change in visibility occurs (i.e. points at distance $\varepsilon >0$ lying after $p'$ are not visible for sufficiently small values of $\varepsilon$). This can happen for only two reasons: either $p'$ is a reflex vertex of type R, or there is some reflex vertex $v'$ of type L such that $p'$ is the shadow of $v'$. In the former case, $p'$ is equal to $v$ since no reflex vertex lies in the interior of $\chain(p,p')$ by Observation~\ref{obs:VisibleChains}. In the latter case, $v'$ must be the first visible reflex vertex lying after $p$ since no point between $p'$ and $v'$ can be visible by Observation~\ref{obs:VisibleChains}. \end{proof} The following result is a direct consequence of Lemma~\ref{lem_key} that allows us to report $\VisC$ of an independent chain $\C$ with no interior reflex vertices. \begin{corollary}\label{corollary:CharacterizationVisibleReflex} Let $p \in \bd$ be a visible point such that $p$ is not a reflex vertex of type R. Let $v$ be the first visible reflex vertex lying after $p$ and let $w$ be its shadow. The following statements hold: \begin{itemize} \vspace{-1mm}\item[-] If $v$ is of type R, then $Chain(p, v)$ is CCW-maximal. \vspace{-1mm}\item[-] If $v$ is of type L, then $Chain(p, w)$ is CCW-maximal. \end{itemize} \end{corollary} \iffalse \begin{lemma}\label{DivideLemma} Let $\mathcal C = \{p_i, \ldots, p_j\}$ be an independent subchain of $\bd$ (i.e. $p_i, p_j$ are visible). Let $x$ be a visible interior point of $\mathcal C$ lying on the edge $p_k p_{k+1}$ for some $i\leq k< j$. Let $\mathcal C_1 = \{p_i, \ldots, p_k, x\}$ and $\mathcal C_2 = \{x, p_{k+1}, \ldots, p_j\}$, then $$\VisC = \mathrm{Vis}_{\mathcal C_1}(q) \cup \mathrm{Vis}_{\mathcal C_2}(q),$$ $$\mathrm{Vis}_{\mathcal C_1}(q) \cap \mathrm{Vis}_{\mathcal C_2}(q) = qx.$$ \end{lemma} \begin{proof} For every $p\in \VisC$, the segment $pq$ is contained in $\region(p_i, p_j)$ (since $p$ is $\mathcal C$-visible). That is, no point of the subchain $\mathcal C =\mathcal C_1\cup \mathcal C_2$ crosses $pq$, which implies that $p$ is $\mathcal C_1$-visible or $\mathcal C_2$-visible (depending on whether $p$ belongs to $\mathcal C_1$ or to $\mathcal C_2$), and it is both if and only if $p=x$. Thus, $\VisC \subseteq \mathrm{Vis}_{\mathcal C_1}(q) \cup \mathrm{Vis}_{\mathcal C_2}(q)$. The other inclusion is proven by contradiction. Assume that there exists a point $p\in \mathrm{Vis}_{\mathcal C_1}(q)$ such that $p\notin qx$, and the segment $pq$ intersects $\mathcal C_2$ at a point $w$ (since otherwise $p$ would be $\mathcal C$-visible). By definition of $\mathcal C_1$-visible, $pq$ is contained in the polygon $\regone = \{q, p_i, \ldots, p_{k}, x, q\}$, hence $w$ lies in the interior of $\regone$. Since $x$ is a visible point, $\regone$ and $\ensuremath{{{\mathcal R_{\mathcal C_{2}}}}}$ have disjoint interiors, therefore $p_j$ does not belong to $\regone$. Thus, by the Jordan curve Theorem, the curve $\gamma$ contained on $\mathcal C_2$ joining $w$ with $p_m$ either crosses $\mathcal C_1$, in which case $\mathcal P$ would not be a simple polygon; or $\gamma$ crosses either the edge $qp_i$ or $qx$, in which case we have a contradiction since both $p_i$ and $x$ are visible points of $\mathcal P$. Therefore, $pq$ does not intersect $\mathcal C_2$ and consequently $p$ is $\mathcal C$-visible. Note that this also implies that $p\notin \mathrm{Vis}_{\mathcal C_2}(q)$ and hence $\mathrm{Vis}_{\mathcal C_1}(q) \cap \mathrm{Vis}_{\mathcal C_2}(q) = qx$. The proof when $p\in \mathrm{Vis}_{\mathcal C_2}(q)$ is analogous. \end{proof} \rodrigo{What about making the previous lemma an observation, and removing the proof? The result looks rather obvious to me, but the proof is rather convoluted---given the simplicity of the result} Lemma~\ref{DivideLemma} allows us to divide the problem of finding $\Vis$ into a series of subproblems (computing the visibility polygon of a polygonal subchain) that can be solved independently. \fi \section{Output Sensitive Algorithm}\label{section:SequentialAlgorithms} In this section we present an algorithm that will be used as stepping stone for our divide and conquer algorithm presented in Section~\ref{Section:Divide and conquer}. This base algorithm computes the visibility polygon of an independent chain using $O(1)$ extra variables. For this purpose, we introduce an operation that we call $\textsc{NextVisReflex}$. This operation receives an independent chain $\C=\chain(a,b)$ and a visible point $p$ on $\C$. Its objective is to compute the next \emph{visible} reflex vertex lying after $p$ on \C, along with its shadow. That is, let $v$ be the first reflex vertex lying after $p$ on $\C$. If $v$ is visible, then $\textsc{NextVisReflex}(p,\C)$ should return $v$ and its shadow. Otherwise, we know that at some point when walking along the path from $p$ to $v$ we change from a visible to a non-visible region. By Lemma~\ref{lem_key}, this change occurs at the shadow of some visible reflex vertex. In this case, $\textsc{NextVisReflex}(p,\C)$ should return this reflex vertex (as well as its shadow). For well-definedness purposes, we say that $\textsc{NextVisReflex}(p,\C)$ should return $b$ if $\C$ contains no reflex vertex. The following observation (illustrated in Fig.~\ref{fig:ImprovedAlgorithm}) allows us to compute $\textsc{NextVisReflex}(p,\C)$ efficiently. \begin{figure}[tb] \centering \includegraphics{ImprovedAlgorithm} \caption {\small When the first reflex vertex $v$ lying after $p$ is not visible, there must exist a visible reflex vertex in $\chain(p,b)$ angularly located between $p$ and $v$. The one with smallest CCW-angle inside $\region(p,v)$ ($v'$ in the figure) will determine the change between visible and non-visible regions between $p$ and $v$.} \label{fig:ImprovedAlgorithm} \end{figure} \begin{observation}\label{obs:SmallestAngle} For any independent chain $\mathcal{C}=\chain(a,b)$, and visible point $p\in \C$ that is not an R-type reflex vertex, let $v$ be the first reflex vertex encountered when walking from $p$ towards $b$ on $\C$ (or $b$ if none exists). If $v$ is not visible, then the CCW-maximal chain starting at $p$ ends at the shadow of the L-type reflex vertex in $\region(p,v)$ with smallest CCW-angle. \end{observation} \begin{lemma}\label{lem_next} For any independent chain $\C$ of $n$ vertices and a visible point $p\in \C$, \linebreak$\textsc{NextVisReflex}(p,\C)$ can be computed in $O(n)$ time using $O(1)$ additional variables. \end{lemma} \begin{proof} Let $\mathcal{C}=\chain(a,b)$ and let $v$ be the first reflex vertex lying after $p$ on $\C$. In $O(n)$ time we can perform a ray shooting query: if $v$ is visible then we output it (and its shadow) as the result of $\textsc{NextVisReflex}(p,\C)$. Otherwise, we use Observation~\ref{obs:SmallestAngle} and compute the reflex vertex on $\chain(p,b)$ with smallest CCW-angle (among those that are in $\region(p,v)$). This vertex is found by walking along $\chain(p,b)$ and keeping track of every time we enter or leave $\region(p,v)$. Note that since $p$ is visible and $\C$ is simple, we can only enter or leave $\region(p,v)$ when we cross line segment $qv$, hence this can be checked in constant time per edge of \C. Since a constant number of operations is needed per vertex, at most $O(n)$ time will be needed for computing $\textsc{NextVisReflex}(p,\C)$. Finally note that Observation~\ref{obs:SmallestAngle} holds whenever $p$ is not an R-type reflex vertex. Thus, if $p$ is a reflex vertex of type R, it suffices to first compute its shadow, and return the same as $\textsc{NextVisReflex}(\textsc{Shadow}(p,\C),\C)$ would. \end{proof} Given an independent chain $\C = (c_1, \ldots, c_n )$, our base algorithm works as follows: start from $c_1$, use $\textsc{NextVisReflex}$ to obtain the next visible reflex vertex, and report the CCW-maximal chain starting from $c_1$. We then repeat the procedure starting from the last reported vertex until we reach $c_n$ (see the details in Algorithm \ref{alg:subroutine}). \begin{theorem}\label{theo:Sequential algorithm} Algorithm~\ref{alg:subroutine} reports the visibility polygon of an independent chain of $n$ vertices and $\overline{r}$ visible reflex vertices in counterclockwise order in $O(n\overline{r})$ time, using $O(1)$ additional variables. \end{theorem} \begin{proof} Correctness of the algorithm is given by Lemma \ref{lem_next} and Corollary~\ref{corollary:CharacterizationVisibleReflex}. It is easy to verify that Algorithm \ref{alg:subroutine} uses a constant number of variables, hence it remains to show a bound on the running time. Notice that at each iteration of the algorithm we report a visible reflex vertex. Hence, we can charge the cost of $\textsc{NextVisReflex}$ operation to the reported vertex. Since no vertex is reported twice and operation $\textsc{NextVisReflex}$ needs linear time, the total running time is bounded by $O(n\overline{r})$. \end{proof} \begin{algorithm} \begin{algorithmic}[1] \STATE $c_{\mathrm{start}} \leftarrow$ $c_1$ (or $c_{\mathrm{start}}\gets \textsc{Shadow}(c_1,\C)$ if $c_1$ is an R-type reflex vertex) \REPEAT \STATE $v\gets \textsc{NextVisReflex}(c_{\mathrm{start}},\C)$ \IF{$v=c_n$} \STATE (* The remainder of the chain is visible *) \STATE $c_{\mathrm{stop}} \gets c_n$ \STATE $c_{\mathrm{next}}\gets c_n$ \ELSE \STATE (* We found next visible reflex $v$. The reported chain will depend on the type of $v$ *) \IF{$v$ is of type R} \STATE $c_{\mathrm{stop}} \gets v$ \STATE $c_{\mathrm{next}}\gets \textsc{Shadow}(v,\C)$ \ELSE \STATE $c_{\mathrm{stop}} \gets \textsc{Shadow}(v,\C)$ \STATE $c_{\mathrm{next}}\gets v$ \ENDIF \ENDIF \STATE Report every vertex between $c_{\mathrm{start}}$ and $c_{\mathrm{stop}}$ \STATE $c_{\mathrm{start}}\gets c_{\mathrm{next}}$ \UNTIL{$c_{\mathrm{start}} = c_n$} \end{algorithmic} \caption{Computing the visibility polygon of an independent chain $\C=(c_1,\ldots c_n)$} \label{alg:subroutine} \end{algorithm} \section{A divide-and-conquer approach}\label{Section:Divide and conquer} We now consider the case in which we are allowed a slightly larger amount of variables. We parametrize the running time of our algorithms by the number of working variables allowed, which we denote by $s$. Our aim is to obtain an algorithm whose running time decreases as $s$ grows. Using the result of the previous section as base algorithm, we now present a divide-and-conquer approach to solve the problem. The general scheme of our algorithm is the natural one: choose a reflex vertex $z$ inside the cone $\coneC$, perform a ray shooting query to find the visible point in the direction of $z$, and split the polygonal chain into two smaller independent subchains $\mathcal C_1, \mathcal C_2$ (see Fig.~\ref{fig:OutlineAlgorithm}, left). We repeat the process recursively, until either $(1)$ a chain $\mathcal C$ has a constant number of reflex vertices (see Fig.~\ref{fig:OutlineAlgorithm}, right) or $(2)$ the depth of the recursion is such that we would exceed the number of allowed working variables. Whenever either of these two conditions is met, we compute the visibility polygon of the chain using Algorithm \ref{alg:subroutine}. See a scheme of this approach in Algorithm~\ref{alg:DivideAndConquer}. \begin{figure}[tb] \centering \includegraphics{OutlineAlgorithm} \includegraphics{Casek=0} \caption{Left: Split of $\mathcal C$ into two subchains $\mathcal C_1, \mathcal C_2$ using a visible point $x$ in the direction of a reflex vertex $z$. Right: A polygonal chain $\mathcal C$ with no reflex vertices inside the cone $\coneC$, only one subchain of $\mathcal C$ is visible. } \label{fig:OutlineAlgorithm} \end{figure} \begin{algorithm} \caption{Given a polygonal chain $\mathcal C = (c_1, \ldots, c_n)$ such that $c_1, c_n$ are both visible points of $\mathcal P$ and a positive integer depth $d$ (initially $1$), compute $\VisC$} \label{alg:DivideAndConquer} \begin{algorithmic}[1] \STATE $k\gets$ number of reflex vertices of $\mathcal C$ inside the cone $\coneC$ \IF{$k \leq 2$ or $d\geq h(s)$} \STATE\label{algoconst} Run Algorithm~\ref{alg:subroutine} on $\C$ \ELSE \STATE $v\leftarrow \textsc{findPartitionVertex}(\mathcal C)$ \STATE $x\gets \textsc{RayShooting}(v,\C)$ \STATE Algorithm~\ref{alg:DivideAndConquer}$ (\{c_1, \ldots, x\},d+1)$ \STATE Algorithm~\ref{alg:DivideAndConquer}$ (\{x, \ldots, c_n\},d+1)$ \ENDIF \end{algorithmic} \end{algorithm} In order to control the depth of the recursion we use a depth counter, hence the algorithm stops dividing once $d=h(s)$ (for some value $h(s)\in O(s)$ that will be determined later). Note that the split direction is decided by subroutine $\textsc{FindPartitionVertex}(\mathcal C)$. This procedure should give a direction so that the resulting subchains ${\mathcal C_1}$ and ${\mathcal C_2}$ have roughly the same complexity. Naturally, it must also run reasonably fast and use $O(s)$ variables. In this section we propose two different methods to choose the partition vertex. The first one uses the approximate median finding algorithm of Raman and Ramnath~\cite{rr-iubtstsls-98}. The second one is randomized, and simply chooses a random reflex vertex among those lying in the cone $\coneC$. We first show that, regardless of the partition method used, the visibility polygon will be correctly computed. \begin{lemma}\label{lem_correctdiv} Algorithm~\ref{alg:DivideAndConquer} correctly reports the visibility polygon of an independent chain in counterclockwise order, using $O(s)$ variables. \end{lemma} \begin{proof} The divide-and-conquer approach repeatedly partitions the input into independent chains. Each subchain will eventually be reported. By Observation~\ref{DivideObs}, the union of reported vertices is equal to the visibility polygon, hence correctness is derived from the correctness of Algorithm~\ref{alg:subroutine}. Regarding space, the subroutines called in this algorithm (\textsc{findPartitionVertex} and \linebreak\textsc{RayShooting}) use $O(s)$ and $O(1)$ variables, respectively. Once the procedure finishes, their working space can be reused for further calls, hence we never use more than $O(s)$ working space at the same time. It remains to consider the memory used implicitly for handling the recursion. Since each step of the algorithm needs a constant number of variables, the total memory needed will be proportional to the recursion depth. Since $h(s)\in O(s)$, the claim is shown. \end{proof} In what follows we present two implementations of $\textsc{FindPartitionVertex}$. \subsection{Deterministic variant}\label{secdeter} We start by giving a deterministic algorithm for $\textsc{FindPartitionVertex}(\mathcal C)$ using $O(s)$ extra variables. For this purpose, we will use the algorithms by Raman and Ramnath presented in~\cite{rr-iubtstsls-98}, which compute an approximation of the median value of a given set using reduced workspace. Given a chain $\C$, let $\Theta = \{\theta_q(v_i) : v_i\text{ is a reflex vertex of $\C$ lying inside }\coneC\}$. An element $\theta_q(v_i)\in \Theta$ is called a \emph{$2/3$-median of $\Theta$} if there are at most $2|\Theta|/3$ elements in $\Theta$ smaller that $\theta_q(v_i)$ and at most $2|\Theta|/3$ greater that $\theta_q(v_i)$. Given two elements $z,z'$ of $\Theta$ such that $z<z'$, we say that $z,z'$ is an \emph{approximate median pair} if at most $|\Theta|/2$ elements of $\Theta$ are smaller than $z$, at most $|\Theta|/2$ lie between $z$ and $z'$, and at most $|\Theta|/2$ elements are greater that $z'$. Notice that if $z,z'$ is an approximate median pair, then either $z$ or $z'$ is a $2/3$-median of $\Theta$. Moreover, we can determine which of the two is a $2/3$-median with one scan of $\C$. Thus, we say that every approximate median pair induces a $2/3$-median of $\Theta$. Raman and Ramnath~\cite{rr-iubtstsls-98} presented two algorithms to find an approximate median pair. The first one is used whenever $s\in o(\log \Rin)$, and allows us to find a $2/3$-median of $\Theta$ in $O(sn\Rin^{1/s})$ time using $O(s)$ variables (Lemma 5 of~\cite{rr-iubtstsls-98}, assigning their parameter $p$ to $\Rin^{1/s}$). For the case $s\in \Omega(\log \Rin)$, they propose another algorithm (stated in Lemma 3 of~\cite{rr-iubtstsls-98}) that computes an approximate median pair in $O(n\log \Rin)$ time, using $O(\log \Rin)$ variables. We note that Raman and Ramnath used these algorithms to afterwards obtain the exact median, but in here we only need an approximation. $\textsc{FindPartitionVertex}(\mathcal C)$ will execute the approximate median pair algorithm, and return the reflex vertex $v$ that induces a $2/3$-median of $\Theta$. By construction, each of the two cones obtained from $\coneC$, by shooting a ray through $v$, will contain at most $2/3$ of the reflex vertices in $\coneC$. Let $P(n,\Rin)$ be the running time of the approximate median finding algorithm on a chain $\C$ of length $n$ with $\Rin$ reflex vertices inside $\coneC$ (that is, $P(n,r)=O(sn\Rin^{1/s})$ if $s\in o(\log \Rin)$, $P(n,r)=O(n\log \Rin)$ otherwise). We set $h(s)=s \log_{3/2}2 \approx 1.71 s$ (if $s\in o(\log r)$) or $s=\log_{3/2}r$ (otherwise). Observe that in both cases we have $h(s)\in O(s)$. We now prove an upper bound on the running time of Algorithm~\ref{alg:DivideAndConquer} with this implementation of $\textsc{FindPartitionVertex}(\mathcal C)$. \begin{theorem}\label{theo_determ} For any $s\in O(\log r)$, $\VisC$ can be computed in $O(\frac{nr}{2^{s}}+n\log^2 r)$ time using $O(s)$ variables. \end{theorem} \begin{proof} Consider any node $u_i$ of the recursion tree of the algorithm and let $\mathcal C_i$ be the chain processed at this node. Let $n_i$ and $r_i$ be the size of $\mathcal C_i$ and the number of reflex vertices lying inside $\Delta_{\mathcal C_i}$, respectively. If $u_i$ is a non-terminal node of the recursion, then the running time at $u_i$ is bounded by $O(P(n_i,r_i))$. Note that a vertex of the input can only appear in at most two chains of the same depth. Hence, the total cost of all non-terminal nodes of a fixed level $j$ in the recursion tree is bounded by $\sum_{u_i} O(P(n_i, r_i)) \leq O(P(n,r))$. By definition, there are at most $h(s)$ levels of recursion, hence the time spent in all the non-terminal executions of the algorithm is bounded by $O(P(n,r)h(s))$. It remains to consider the time spent in the terminal nodes. Each terminal node $u_i$ will need $O(n_i\overline{r}_i)$ time, where $\overline{r}_i$ denotes the number of visible reflex vertices on $\mathcal C_i$. Recall that terminal nodes are only executed whenever either $\C_i$ has a constant number of reflex vertices or we have reached $h(s)$ levels of recursion. Further note that, at each level of recursion at least a third of the reflex vertices are discarded. In particular, we have that $\overline{r_i}\leq r(\frac{2}{3})^{h(s)}$. Similar to non-terminal nodes, vertices cannot be present in more than two terminal nodes. Thus, the total time spent in the terminal nodes is bounded by $$\sum_{u_i} O(n_i\overline{r_i}) \leq \sum_{u_i} O(n_ir(2/3)^{h(s)})) \leq O(nr(2/3)^{h(s)}).$$ Therefore, the total time spent by the algorithm becomes $O(P(n,r)h(s)+nr(2/3)^{h(s)})$. This expression can be simplified by distinguishing between different workspace sizes. \begin{description} \item[Case $s\in o(\log r)$.] In this case we have $P(n,r)h(s)=O(s^2nr^{1/s})$, and in particular $O(s^2nr^{1/s}) \in O(nr^{1/3})$ (since $s$ is a parameter that can be chosen to be at least 3). Further recall that $h(s)=s\log_{3/2}2$, and in particular $(2/3)^{h(s)}=2^{-s}$. Thus, the second term simplifies to $\frac{nr}{2^s}$. Since $s\leq \log_2(r)/3$ and the running time decreases as $s$ grows, we have $\frac{nr}{2^s} \geq \frac{nr}{r^{1/3}} \in \Omega(n\sqrt{r})$. That is, the running time of our algorithm is expressed as the sum of two terms that, when $s\in o(\log r)$, the first one is at most $O(nr^{1/3})$ whereas the second one is at least $\Omega(n\sqrt{r})$. Asymptotically speaking, the first one can be ignored, and the running time is dominated by $O(\frac{nr}{2^s})$ (i.e., applying Algorithm~\ref{alg:subroutine} to each terminal node). \item[Case $s\in\Theta(\log r)$.] In this case we can use the faster method of Raman and Ramnath (i.e. $P(n,r)=O(n\log r)$). Recall that in this case we set $h(s)=\log_{3/2}r$, hence $(2/3)^{h(s)}=1/r$. In particular, the running time of the second term simplifies to $O(nr(2/3)^{h(s)})= O(n)$. Since $s\in\Theta(\log r)$, the running time is dominated by the first term (i.e., finding the split direction), which is $O(P(n,r)h(s))=O(n\log^2 r)$. \end{description} Observe that in both cases the running time is bounded by $O(\frac{nr}{2^{s}}+n\log^2 r)$, hence the claim holds. \end{proof} {\bf Remark} Note that, although we only consider algorithms that use up to $O(\log r)$ variables, one could study what happens whenever more space is allowed. However, we note that increasing the size of our workspace will not reduce the running time, since the time bottleneck of this approach is determined by the approximate median method of Raman and Ramnath. \subsection{Randomized approach}\label{second} Whenever $s\in \Theta (\log n)$, the running time of the previous algorithm is dominated by the \textsc{FindPartitionVertex} procedure. Motivated by this, in this section we consider a faster (albeit randomized) partition method. The randomized method proceeds as follows: let $k$ be the number of reflex vertices of $\C$ lying inside $\coneC$ (note that $k$ can be computed in linear time by scanning $\C$). Select a random number $i$ between $1$ and $k$ (uniformly at random). The idea is to output the $i$-th reflex vertex inside $\coneC$ (computed by walking counterclockwise along $\C$). However, we must first check that this vertex will make a balanced partition. In order to check so, we make another scan of $\mathcal C$, and we count the angular rank of the $i$-th reflex vertex (among the reflex vertices in $\coneC$). If its rank is between $k/3$ and $2k/3$, we use it as partition vertex. Otherwise, we pick another random number and repeat the process until such a vertex is found. \begin{lemma}\label{lem_runningtime} The randomized version of \textsc{FindPartitionVertex} has expected running time $O(n)$. \end{lemma} \begin{proof} The probability that, when choosing a reflex vertex uniformly at random, we pick one whose rank is between $1/3$ and $2/3$ is exactly $1/3$. If each time we make the choices independently, we are performing Bernoulli trials whose probability of success is $1/3$. Hence, the expected number of times we have to choose a random index is a constant (three in this case). For each try we only need to check its rank (which can be done in $O(n)$ time by performing two scans of the input). Therefore we conclude that the expected running time of \textsc{FindPartitionVertex} is $O(n)$. \end{proof} \begin{theorem}\label{theo_randomi} For any $s\in O(\log r)$, $\VisC$ can be computed in $O(\frac{nr}{2^{s}}+n\log r)$ expected time using $O(s)$ variables. \end{theorem} \begin{proof} The proof is identical to the proof of Theorem \ref{theo_determ}, just taking into account that now $P(n,r)=O(n)$, hence the running time of the second term is decreased by a $\log r$ factor. \end{proof} Observe that this algorithm is faster than the previous one when $s\in \Theta(\log r)$. We conclude by summarizing the different algorithms presented in this paper. \begin{theorem} Given a polygon $\Poly$ of $n$ vertices, out of which $r$ are reflex, the visibility polygon of a point $q\in\Poly$ can be computed in $O(n\Rout)$ time using $O(1)$ variables (where $\Rout$ is the number of reflex vertices of $\Vis$) or $O(\frac{nr}{2^{s}})$ time using $O(s)$ variables (for any $s\in o(\log r)$). If $\Omega(\log r)$ variables are available, the running time decreases to $O(n\log^2 r)$ time or $O(n\log r)$ randomized expected time. \end{theorem} \bibliographystyle{abbrv}
\section{Introduction} The main symmetry of torus links, $T(p,q)=T(q,p)$, is a trivial geometric fact. However, it is hardly visible on the level of standard diagrams; the braids $(\sigma_1 \sigma_2 \ldots \sigma_{p-1})^q$ and $(\sigma_1 \sigma_2 \ldots \sigma_{q-1})^p$ do not even have the same number of crossings. We propose a diagram of the fibre surface of torus links that exhibits the symmetry of the parameters $p$ and $q$. The following description is motivated by A'Campo's new t\^ete-\`a-t\^ete vision of the monodromy of isolated plane curve singularities~\cite{AC}. It relies on the fact that the fibre surface of the torus link $T(p,q)$ retracts on a complete bipartite graph of type $\theta_{p,q}$~\cite{Ph}. This fact can be seen explicitly in Figure~1, where the fibre surface of the torus knot $T(3,4)$ is drawn as a union of $12$ ribbons along the edges of the graph $\theta_{3,4}$ embedded in ${\mathbb R}^3$. We will shortly explain this in detail. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{tete1.eps}}}$}} \caption{} \end{figure} Ribbon diagrams offer a lot of possibilities to perform cobordisms by cutting ribbons. For example, the three cuts shown in Figure~2 result in a disjoint union of two trefoil knots. It turns out that many ribbon cuts correspond to smoothings of certain crossing in the standard diagrams of torus links. This suggests to look at links associated with subgraphs of the complete bipartite graphs $\theta_{p,q} \subset {\mathbb R}^3$. Let us call these bipartite graph links. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{tete2.eps}}}$}} \caption{} \end{figure} \begin{theorem} The family of bipartite graph links coincides with the family of strongly quasipositive links. \end{theorem} Bipartite graph links contain various well-studied classes of links, for example positive braids links and Lorenz links. We will single out these classes in terms of bipartite graphs. As an application, we obtain a curious duality on standard diagrams of twisted torus links. This is the content of Section~3. Given two fixed natural numbers $p,q \geq 2$, we may ask which links can be obtained by cutting ribbons of the embedded complete bipartite graph $\theta_{p,q}$. This innocent question turns out to be a hard one. In fact, it may be related to the adjacency problem for plane curve singularities. We will offer perspectives on this in Section~4. The last section is devoted to a notion of density that comes naturally with bipartite graph links. As we will see, links with a high density share at least one property with torus links: their signature invariant has a high defect from maximality. \section*{Acknowledgements} This manuscript was greatly influenced by various people. Special thanks go to Norbert A'Campo, Peter Feller, Christian Graf and Masaharu Ishikawa for their inspiring inputs. \section{Ribbon diagrams for strongly quasipositive links} We need precise definitions for bipartite graph links, quasipositive surfaces and strongly quasipositive links before proving Theorem~1. Let $U,L \subset {\mathbb R}^3$ be two skew lines. We will fix $U=\{x=0,\, z=1\}$, $L=\{x=y,\, z=0\}$, for simplicity. Let $\Gamma \subset {\mathbb R}^3$ be a finite union of straight line segments, each one having one endpoint on $U$ and one on $L$. Thus $\Gamma$ is an embedded bipartite graph. The ribbon surface associated with $\Gamma$ is made up of ribbons, one for each edge of $\Gamma$, whose projections onto the $(y,z)$-plane are immersions (see Figures~1, 2 and~4 for an illustration). \begin{definition} A link in ${\mathbb R}^3$ is called a bipartite graph link, if it is the boundary of a ribbon surface in the above sense. \end{definition} We claim that the links associated with complete bipartite graphs are precisely torus links. In fact, given an embedded complete bipartite graph $\Gamma \subset {\mathbb R}^3$ with $p$ and $q$ vertices on $U$ and $L$, respectively, we may deform the corresponding link $L(\Gamma)$ into the standard braid diagram $(\sigma_1 \sigma_2 \ldots \sigma_{p-1})^q$ of $T(p,q)$. This deformation is performed in two steps. \medskip (1) Each vertex on the lower line $L$ is adjacent to precisely $p$ edges of $\Gamma$, the union of which we call a fork. Thus $\Gamma$ consists of $q$ forks that are piled in some sense. Stretching each of the vertices on $U$ into an interval allows us to separate all forks in the $(y,z)$-projection. This is shown on the top left of Figure~3, for the case $p=3$, $q=4$. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{movie1.eps}}}$}} \bigskip \bigskip \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{movie2.eps}}}$}} \caption{} \end{figure} (2) Each fork can be split into $p-1$ forks with two teeth connecting a pair of stretched vertices, by a suitable isotopy. This is illustrated at the bottom of Figure~3, for a single fork with $4$ teeth. The resulting surface diagram is shown on the top right of Figure~3 and can easily be identified as the canonical Seifert surface associated with the braid $(\sigma_1 \sigma_2 \ldots \sigma_{p-1})^q$. A first naive and false guess is that bipartite graph links are positive braid links. The bipartite graph knot depicted in Figure~4 is obtained from the ribbon diagram of Figure~1 by $4$ ribbon cuts. It is isotopic to the non-fibred positive twist knot $5_2$, which is not a positive braid knot, since these are all fibred~\cite{St}. For the same reason, it is not possible to obtain the knot $5_2$ from any of the standard diagrams of the torus knot $T(3,4)$ by smoothing any number of crossings. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{tete5.eps}}}$}} \caption{} \end{figure} There is still a weaker notion of positivity inherited by bipartite graph links: strong quasipositivity. As many notions of positivity, strong quasipositivity was introduced by Lee Rudolph~\cite{Ru1}. The definition of strong quasipositivity involves certain surfaces, altough it is an attribute for links. \begin{definition} An embedded compact surface $S \subset {\mathbb R}^3$ is called positive, if it is isotopic to an incompressible subsurface of the fibre surface of a positive torus link. \end{definition} Here incompressibility simply means that the inclusion of $S$ into the fibre surface induces an injective map on the level of fundamental groups. \begin{definition} A link in ${\mathbb R}^3$ is called strongly quasipositive, if it is the boundary of a quasipositive surface. \end{definition} Alternatively, quasipositive surfaces may be defined as Legendrian ribbons with respect to the standard contact strucure on ${\mathbb R}^3$~\cite{BI}. As the name suggests, strongly quasipositive links include positive links. However, this is a non-trivial fact, due to Rudolph~\cite{Ru2}. Using either of the definitions, we immediately see that bipartite graph links are strongly quasipositive. Indeed, all ribbon surfaces constructed above are incompressible subsurfaces of the ribbon surface associated with an embedded complete bipartite graph. The latter are fibre surfaces of torus links, since they are isotopic to the canonical Seifert surfaces of these. In order to show the converse, we need yet another desription of strongly quasipositive links, which is in fact the original one~\cite{Ru1}. \begin{definition} A link in ${\mathbb R}^3$ is strongly quasipositive, if it is the closure of a strongly quasipositive braid $\beta$ in some braid group $B_n$, i.e. a finite product of words of the form $$\sigma_{i,j}=(\sigma_i \sigma_{i+1} \ldots \sigma_{j-2}) \sigma_{j-1} (\sigma_i \sigma_{i+1} \ldots \sigma_{j-2})^{-1},$$ for $1 \leq i<j \leq n-1$. \end{definition} For obvious reasons, we call the words $\sigma_{i,j}$ generalised positive crossings (see Figure~5). Strongly quasipositive links bound canonical Seifert surfaces made of one disc for each braid strand and one band for each generalised crossing. \begin{figure}[ht] \scalebox{1.0}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{positivecrossing.eps}}}$}} \caption{} \end{figure} A canonical surface is shown at the bottom right of Figure~6, for the braid $\sigma_1 \sigma_{1,3} \sigma_3 \sigma_2 \sigma_3$. In that diagram all bands are represented by forks with two teeth. By looking at the whole sequence of diagrams of Figure~6, we realise that the ribbon surface of any embedded bipartite graph can be deformed into the canonical surface of a strongly quasipositive braid. Note that forks consisting of one edge give no contribution to the canonical surface diagram, since the correponding ribbons can be removed by an isotopy. Conversely, every canonical surface can be deformed into a bipartite graph surface by shrinking each braid strand to a point. In this way we obtain a bipartite graph where all points on the lower line $L$ have valency two. This proves Theorem~1. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{movie3.eps}}}$}} \caption{} \end{figure} \section{Twisted torus links} As we saw in the preceding section, there is a natural way of deforming the ribbon surface of an embedded bipartite graph $\Gamma$ into the canonical surface of a strongly quasipositive braid. The resulting braid is positive, if and only if all forks of $\Gamma$ are complete, meaning that their endpoints form sets of consecutive vertices of $\Gamma$ on the upper line (as in Figure~3). By turning an embedded bipartite graph $\Gamma$ upside down, we obtain another embedded bipartite graph $\widetilde{\Gamma}$ that gives rise to a different braid, in general. Note that the forks of $\widetilde{\Gamma}$ need not be complete, even if the ones of $\Gamma$ are. Therefore we are not able to extend the duality on torus link diagrams to positive braids. However, we may define a duality on the more restricted class of twisted torus link diagrams. By the work of Birman and Kofman~\cite{BK}, these represent Lorenz links. Twisted torus links admit a natural description in terms of bipartite graphs. Let $a_1 \geq a_2 \geq \ldots \geq a_n$ be a finite decreasing sequence of natural numbers. We define an embedded bipartite graph $\Gamma(a_1, \ldots, a_n)$ with $a_1$ vertices on the upper line $U$ as a union of $n$ forks, where the $k$-th fork has the first $a_k$ points on $U$ as endpoints (see Figure~7 for an illustration). Braid diagrams associated with graphs of type $\Gamma(a_1, \ldots, a_n)$ are precisely twisted torus link diagrams. Turning $\Gamma(a_1, \ldots, a_n)$ upside down gives rise to a dual graph $\Gamma(b_1, \ldots, b_n)$ with $b_1=n$, $m=a_1$. In fact, thinking of the numbers $a_1, a_2, \ldots, a_n$ as the coefficients of a Young diagram, the numbers $b_1, b_2, \ldots, b_n$ correspond to the dual Young diagram. For example, the dual graph of $\Gamma(4,4,3,2,2)$ is $\Gamma(5,5,3,2)$. The same involution has been described by Birman and Kofman (\cite{BK}, Corollary~4) and, in terms of Lorenz links, by Dehornoy (\cite{De}, Proposition~1.18.). Within the framework of bipartite graphs, it is not hard to figure out a duality on an even larger class of diagrams. \begin{figure}[ht] \scalebox{1.0}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{graphs.eps}}}$}} \caption{$\Gamma(4,4,3,2,2)$ and $\Gamma(5,5,3,2)$} \end{figure} \section{Combinatorial adjacency} Torus links are prototypes of links associated with isolated singularities of complex plane curves. Classically, the torus link $T(p,q)$ is defined as the intersection of the singular curve $\{z^p+w^q=0\}$ with the unit $3$-sphere $\{|z|^2+|w|^2=1\} \subset {\mathbb C}^2$. A generic deformation of that curve will transform it into a smooth one, e.g. $z^p+w^q+t$, $t \in [0,1]$. Carefully chosen deformations may give rise to simpler singularities, e.g. $z^p+w^q+t(z^a+w^b)$, $t \in [0,1]$, where $a \leq p$, $b \leq q$. After a suitable biholomorphic coordinate change around $0 \in {\mathbb C}^2$, the singularity $z^p+w^q+t(z^a+w^b)$ becomes $x^a+y^b$, provided $t>0$. Formally, a deformation of a singularity of $f \in {\mathbb C}[z,w]$ at $0 \in {\mathbb C}^2$ is a polynomial in three variables $H(t,z,w)$ with the following properties: \begin{enumerate} \item $H(0,z,w)=f(z,w)$; \item for all $t \in [0,1]$, the restriction $H_t(z,w)=H(t,z,w)$ has an isolated singularity at $0 \in {\mathbb C}^2$; \item for all $t \in (0,1]$, the singularity of $H_t$ at $0$ is equivalent to the singularity of $H_1$ at $0$ (via a local biholomorphic coordinate change). \end{enumerate} We say that the singularity of $H_1$ at $0$ is adjacent to the singularity of $H_0=f$, as well as their links. For more details on versal deformations and adjacency of singularities, we refer the reader to Siersma's dissertation~\cite{Si}. An explicit solution to the adjacency problem is not known, not even for singularities of type $z^p+w^q$. Let us note that the link $T(a,b)$ is adjacent to $T(p,q)$, provided $a \leq p$, $b \leq q$. However, this is not a necessary condition. For example, the links $T(2,n)$ are adjacent to $T(3,4)$, for all $n \leq 6$ (these are the only ones, apart from $T(3,3)$, for genus reasons). More generally, for $a,b,c \in {\mathbb N}$ with $c \leq a$, the function $H(t,z,w)=z^a+(w^b+tz)^c$ exhibits a deformation of the singularity $z^a+w^{bc}$ into $z^{ab}+w^c$. This surprisingly easy deformation was found by Peter Feller, based upon ideas of Ishikawa, Nguyen and Oka~\cite{ITO}. The verification requires basic knowledge about Newton polygons. Substituting $x=w^b+tz$ transforms $z^a+(w^b+tz)^c$ into $(\frac{x-w^b}{t})^a+x^c$; the latter is equivalent to $y^{ab}+x^c$, provided $c \leq a$. In this section we propose a combinatorial notion of adjacency for bipartite graph links, motivated by the above list of algebraic adjacencies. A complete bipartite graph of type $\theta_{p,q}$ consists of $p$ forks with $q$ teeth. Given a fork $f$ and two natural numbers $a_1, a_2 \geq 1$ with $a_1+a_2=q$, let us define a splitting of $f$ by assembling the first $a_1$ and the last $a_2$ endpoints of $f$ into two forks $f_1$, $f_2$, respectively (see Figure~8). On the level of braid diagrams, this is simply smoothing a crossing. We say that an embedded bipartite graph $\Gamma$ is adjacent to $\theta_{p,q}$, if it is obtained from $\theta_{p,q}$ by a finite number of fork splittings, where splittings may be applied to up- and down-pointing forks, possibly with iteration (see Figure~2). The corresponding link $L(\Gamma)$ is called combinatorially adjacent to $T(p,q)$ \footnote{This definition arose from discussions with Peter Feller}. As a first example, we note that the torus link $T(a,b)$ is combinatorially adjacent to $T(p,q)$, provided $a \leq p$, $b \leq q$ (this is most easily seen in two steps, via $T(a,q)$ resp. $T(p,b)$). A more interesting splitting is shown in Figure~8; it shows that $T(2,6)$ is combinatorially adjacent to $T(3,4)$. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{tete3.eps}}}$}} \caption{} \end{figure} This is a special case of the following fact: for $a,b,c \in {\mathbb N}$, $c \leq a$, the torus link $T(ab,c)$ is obtained from $T(a,bc)$ by smoothing an appropriate set of crossings, $(b-1)(a-c)$ in number (see~\cite{Ba}, Proposition~1, and Figure~9 for $(a,b,c)=(3,2,7)$ and $(2,3,7)$). This is the combinatorial counterpart to the above statement on algebraic adjacencies. \begin{figure}[ht] \scalebox{0.7}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{toruscut3.eps}}}$}} \caption{} \end{figure} The exciting thing about combinatorial adjacency is that it allows smoothings of crossings in the two standard diagrams of torus links simultaneously. As we just observed, there is a big overlap between algebraic and combinatorial adjacency, where both notions apply. We do not know whether this is more than a pure coincidence. The following question should nevertheless be justified: for which $a,b,p,q \in {\mathbb N}$ is $T(a,b)$ combinatorially adjacent to $T(p,q)$? \section{Density of bipartite graph links} Let $\Gamma \subset \theta_{p,q}$ be an embedded bipartite graph with $p+q$ vertices. We define the density of $\Gamma$ as $$d(\Gamma)=\frac{e(\Gamma)}{pq},$$ where $e(\Gamma)$ is the number of edges of $\Gamma$. This carries over to bipartite graph links $L$, by taking the supremum over all bipartite graph representatives for $L$: $$d(L)=\sup_{L=L(\Gamma)} d(\Gamma).$$ \begin{proposition} \quad \begin{enumerate} \item[(1)] The supremum is actually a maximum, i.e. there exists an embedded bipartite graph $\Gamma$ with $d(\Gamma)=d(L)$,\\ \item[(2)] $d(K) \leq \frac{2}{b}-\frac{\chi}{b^2}$, for all bipartite graph knots $K$, where $b$ and $\chi$ denote the braid index and the maximal Euler characteristic of the knot $K$, respectively. \\ \end{enumerate} \end{proposition} \begin{corollary} $d(L)=1 \Leftrightarrow L$ is a torus link. \end{corollary} \begin{proof}[Proof of Proposition~1] Let $\Gamma$ be an embedded bipartite graph representative for $L$. The corresponding ribbon surface $F=F(\Gamma)$ is a Seifert surface of maximal Euler characteristic for the link $L$, by Bennequin's inequality~\cite{Be}: $\chi(F)=\chi(L)$. In terms of $\Gamma$, we compute $$\chi(F)=p+q-e(\Gamma)=p+q-pqd(\Gamma),$$ thus $$d(\Gamma)=\frac{p+q-\chi(L)}{pq}.$$ This implies the second item of the Proposition, since $p,q \geq b$ and $\chi$ is non-positive for knots (except for the trivial knot, for which the statement is true anyway). Choose $N \in {\mathbb N}$ so that $N \geq |\chi(L)|$ and $N>\frac{3}{d(L)}$. If $p,q \geq N$ then $$d(\Gamma)=\frac{1}{q}+\frac{1}{p}+\frac{-\chi(L)}{pq} \leq \frac{3}{N}<d(L).$$ Therefore, in order to come close to $d(L)$, the graph $\Gamma$ must satisfy $p,q \leq N$. This implies the first item, since there are only finitely many embedded bipartite graphs with $p,q \leq N$. \end{proof} A link is called special alternating, if it has a diagram which is both positive and alternating~\cite{Mu}. Since positive links are quasipositive~\cite{Ru2}, it makes sense to talk about their density. \begin{proposition} Let $K$ be a special alternating knot of braid index at least $3$. Then either $K=T(3,4)$ or $d(K) \leq \frac{24}{25}$. \end{proposition} The restriction to knots of braid index at least $3$ is only for the sake of simplicity of the statement; special alternating knots of braid index $2$ are torus links of type $T(2,n)$, $n \geq 3$. \begin{proof}[Proof of Proposition~2] Let $K=L(\Gamma)$ be represented by an embedded bipartite graph $\gamma \subset \theta_{p,q}$. If $K \neq T(3,4)$ then $d(K) \neq 1$, since the only alternating torus knots are $T(2,n)$ , $n \in {\mathbb N}$, and $T(3,4)$. Suppose, for contradiction, that $\frac{24}{25}<d(K)<1$. Then the graph $\Gamma$ has more than $25$ edges (in particular, one of the numbers $p,q$ is at least $6$). Moreover, $\Gamma$ contains a complete bipartite graph of type $\theta_{3,6}$ as a subgraph. Indeed, if no $(3,6)$-bipartite subgraph of $\Gamma$ was complete, then $d(\Gamma) \leq \frac{17}{18}<\frac{24}{25}$. In terms of Seifert surfaces, this means that the ribbon surface $F(\Gamma)$ associated with $\Gamma$ contains the fibre surface of the torus link $T(3,6)$ as an incompressible subsurface. A direct computation, carried out in~\cite{GLM}, shows that the symmetrised Seifert matrix of the fibre surface of $T(3,6)$ has non-maximal signature ($\sigma(T(3,6))=8<10$). This fact descends to the surface $F(\Gamma)$, since maximality of the signature is preserved under taking minors of symmetric matrices. We conclude $$\sigma(K)<2g(K),$$ where $g(K)$ denotes the genus of the knot $K$. On the other hand, since $K$ is positive, its Rasmussen invariant $s(K)$ coincides with twice the genus: $s(K)=2g(K)$. But $K$ is also alternating, so $\sigma(K)=s(K)$, a contradiction (see~\cite{Ra} for the last two facts). \end{proof} A careful generalisation of the above proof shows that a large density implies $\frac{\sigma}{2g} \leq \frac{3}{4}$. The reader might suspect there is a kind of converse, i.e. a small density implies $\frac{\sigma}{2g} \approx 1$. However, this is not true, as shows the following example. \begin{example} Let $n \geq 5$ and let $\Gamma_n \subset \theta_{n,n+1}$ be the embedded bipartite subgraph obtained from the complete bipartite graph $\theta_{4,5}$ by the extension shown in Figure~10, for $n=9$. We observe that $K_n=L(\Gamma_n)$ is a knot of genus $g(K_n)=g(T(4,5))=6$. A direct computation yields $\sigma(K_n) \leq 10$, thus $\frac{\sigma_{K_n}}{2g(K_n)} \leq \frac{5}{6}$ (this is due to the fact that the torus link $T(4,4)$ has non-maximal signature). Moreover, the braid index of $K_n$ grows linearly in $n$, similarly as for twist knots with increasing twist number. This follows for example from the inequality of Morton and Franks-Williams~\cite{Mo},~\cite{FW}. The inequality of Proposition~1 implies $$\lim_{n \to \infty} d(K_n)=0.$$ Therefore a small density does not necessarily imply $\frac{\sigma}{2g} \approx 1$. \begin{figure}[ht] \scalebox{1.0}{\raisebox{-0pt}{$\vcenter{\hbox{\epsffile{graph.eps}}}$}} \caption{} \end{figure} \end{example} The proof of Proposition~2 suggests to study bipartite graph links via graph minor theory. For example, we may ask if there exist finitely many embedded bipartite graphs $\Gamma_1, \ldots, \Gamma_n \subset {\mathbb R}^3$ such that the following holds: the signature of a bipartite graph knot $K=L(\Gamma)$ is maximal, $\sigma(K)=2g(K)$, if and only if $\Gamma$ contains none of the graphs $\Gamma_i$ as embedded minors.
\section{Introduction} In recent years, the coupling property, the strong Feller property, and gradient estimates have been intensively investigated for linear stochastic differential equations driven by L\'evy processes on $\R^d$, see e.g. \cite{PZ, W10a, W10b, SW, BSW, SSW, SW2, KS2, KS} and references within. In these references the shift-invariance of the Lebesgue measure plays an essential role. When the state space is infinite-dimensional so that the Lebesgue measure is no longer available, we need a reference measure which is quasi-invariant under a reasonable class of shift transforms. Typical examples of the reference measure include the Wiener measure on the continuous path space and the Gaussian measure on a Hilbert space, see Section 5 for details. The purpose of this paper is to investigate regularity properties of linear SDEs driven by L\'evy processes on a Banach space equipped with such a nice reference measure. To ensure the quasi-invariance of the solution, a strong enough linear drift term will be needed. On the other hand, concerning (semi-)linear SDEs on Hilbert spaces, when the noise is a cylindrical $\aa$-stable process, many regularity results derived in finite dimensions can be extended to the infinite-dimensional setting (see \cite{X1,X2,WJ}); and when the noise has a non-trivial Gaussian part, the regularity properties can be derived by using the drift part and the Gaussian part (see e.g. \cite{Z, DZ, BN, RW03}). But there seems to be no results concerning the strong Feller and coupling properties for SDEs driven by purely jump non-cylindrical L\'evy processes. In this paper we intend to investigate these properties for linear SDEs driven by non-cylindrical L\'evy noise on Banach spaces. Let $(\mathbb B}\def\vp{\varphi, \|\cdot\|_\mathbb B}\def\vp{\varphi)$ be a Banach space and let $\mu$ be a probability measure on $\mathbb B}\def\vp{\varphi$ having full support. Let $\mathbb B}\def\vp{\varphi'$ be the dual space of $\mathbb B}\def\vp{\varphi$ with $\langle \cdot, \cdot\rangle$ the duality between $\mathbb B}\def\vp{\varphi$ and $\mathbb B}\def\vp{\varphi'$. Let $(\H,\|\cdot\|_\H)$ be another Banach space which is densely and continuously embedded into $ \mathbb B}\def\vp{\varphi$ such that for any $h\in\H$, $\mu$ is quasi-invariant under the shift $x\mapsto x+h$; that is, there exists a non-negative measurable function $\vp_h$ on $\mathbb B}\def\vp{\varphi$ such that \beq\label{1.1} \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z-h)= \vp_h(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z).\end{equation} Let $L_t$ be a L\'evy process on $\mathbb B}\def\vp{\varphi$ with L\'evy measure $\nu$. Recall that a $\sigma} \def\ess{\text{\rm{ess}}$-finite measure $\nu$ on $\mathbb B}\def\vp{\varphi$ is called a L\'evy measure if $ \nu (\{0\})=0$ and the mapping from $\mathbb B}\def\vp{\varphi'$ to $\R$ given by $$\mathbb B}\def\vp{\varphi'\ni a\mapsto \exp\bigg[\int_\mathbb B}\def\vp{\varphi\big(\cos\,\<x,a\>-1\big)\, \nu (\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D x)\bigg]$$ is the characteristic function of a random variable on $\mathbb B}\def\vp{\varphi$. Note that since $\cos$ is an even function, one may replace $\nu$ by the symmetric measure $\nu+\nu^*$ as in \cite{App1}, where $\nu^{*}(A)=\nu(-A)$ for any $A\in\B$. When $\mathbb B}\def\vp{\varphi$ is a Hilbert space, $\nu$ is a L\'evy measure if and only if $\nu(\{0\})=0$ and $\int_\mathbb B}\def\vp{\varphi (1\land\|x\|_\mathbb B}\def\vp{\varphi^2)\,\nu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D x)<\infty$; while in general, $\nu$ is a L\'evy measure provided $\nu(\{0\})=0$ and $\int_\mathbb B}\def\vp{\varphi (1\land\|x\|_\mathbb B}\def\vp{\varphi )\,\nu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D x)<\infty$ (see \cite{App0,App1}). Let $\sigma} \def\ess{\text{\rm{ess}}: \mathbb B}\def\vp{\varphi\to\mathbb B}\def\vp{\varphi$ be a bounded linear operator and let $(A,\D(A))$ be a linear operator on $\mathbb B}\def\vp{\varphi$ generating a $C_0$ semigroup $(T_s)_{s\ge 0}$. Consider the following linear SDE on $\mathbb B}\def\vp{\varphi$: \beq\label{E1} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= AX_t\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t + \sigma} \def\ess{\text{\rm{ess}} \,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_t.\end{equation} For any $x\in \mathbb B}\def\vp{\varphi$, the solution with initial data $x$ is \beq\label{S} X_t^x= T_t x +\int_0^t T_{t-s} \sigma} \def\ess{\text{\rm{ess}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s,\ \ \ t\ge 0.\end{equation} See \cite{CM, PZ1, App0, App1} for the detailed construction of this solution. Let $\B_b(\mathbb B}\def\vp{\varphi)$ be the class of all bounded measurable functions on $\mathbb B}\def\vp{\varphi$. We aim to investigate the coupling property and the strong Feller property for the associated Markov semigroup $$P_t f(x):= \E f(X_t^x),\ \ \ t\ge 0, x\in \mathbb B}\def\vp{\varphi, f\in \B_b(\mathbb B}\def\vp{\varphi).$$ Recall that the solution has successful coupling if and only if (cf. \cite{Lin,CG}) $$\lim_{t\to \infty} \|P_t(x,\cdot)- P_t(y,\cdot)\|_{var}=0,\ \ \ x,y\in\mathbb B}\def\vp{\varphi,$$ where $P_t(x,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D y)$ is the transition kernel of $P_t$ and $\|\cdot\|_{var}$ is the total variation norm. Let $\rr_0$ be a non-trivial non-negative measurable function on $\mathbb B}\def\vp{\varphi$ such that \beq\label{C} \nu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\ge \rr_0(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)=:\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\end{equation} holds. Thus, the L\'evy process considered here is essentially different from the cylindrical $\aa$-stable process used in \cite{X1,X2}. Indeed, for $\mathbb B}\def\vp{\varphi$ being a Hilbert space with ONB $\{e_i\}_{i\ge 1}$, the L\'evy measure (if exists) for a cylindrical L\'evy process is supported on $\cup_{i\ge 1} \R e_i$ and hence, is singular w.r.t. e.g. a non-trivial Gaussian probability measure $\mu$. Assume \paragraph{(A)} {\rm Ker}$(\sigma} \def\ess{\text{\rm{ess}})=\{0\}$ and $T_s \mathbb B}\def\vp{\varphi\subset \sigma} \def\ess{\text{\rm{ess}} \H$ holds for any $s>0.$ \ Obviously, {\bf(A)} implies that for any $s>0$, the operator $\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s: \mathbb B}\def\vp{\varphi\to\H$ is well defined. \begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}\label{T1.1} Assume {\bf (A)}. Suppose that $\nu_0$ in $(\ref{C})$ is infinite; i.e.\ $\nu_0(\mathbb B}\def\vp{\varphi)=\infty.$ \begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate} \item[$(1)$] If for any $h\in\H$ \beq\label{N}\sup_{\vv\in (0,1)} \vp_{\vv h}(\cdot+\vv h)<\infty,\ \ \mu\text{-a.e.},\end{equation} then for any $f\in \B_b(\mathbb B}\def\vp{\varphi)$ and $t>0$, $P_tf$ is directionally continuous; i.e. $\lim_{\vv\to 0} P_t f(x+\vv y)= P_t f(x)$ holds for any $x,y\in\mathbb B}\def\vp{\varphi.$ \item[$(2)$] If for any $s>0$ \beq\label{1.5'}\sup_{\|y\|_\mathbb B}\def\vp{\varphi\le 1} \vp_{\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(\cdot+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)<\infty,\ \ \mu\text{-a.e.},\end{equation} then $P_t$ is strong Feller for $t>0$; i.e. $P_t \B_b(\mathbb B}\def\vp{\varphi)\subset C_b(\mathbb B}\def\vp{\varphi).$ \end{enumerate} \end{thm} A simple example for $\nu_0(\mathbb B}\def\vp{\varphi)=\infty$ to hold is as follows. Let $z\to \|z\|_\mathbb B}\def\vp{\varphi$ have a strictly positive distribution density function $\rr$ under the probability measure $\mu$, for instance it is the case when $\mu$ is the Wiener measure (see Subsection 5.1 below). Let $r_0\in (0,\infty]$, and let $\aa\in (0,2)$ when $\mathbb B}\def\vp{\varphi$ is a Hilbert space and $\aa\in (0,1)$ otherwise. Then $$\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z):= \ff{1_{(0,r_0)}(\|z\|_\mathbb B}\def\vp{\varphi)}{\rr(\|z\|_\mathbb B}\def\vp{\varphi)\|z\|^{1+\aa}_\mathbb B}\def\vp{\varphi}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)$$ is a L\'evy measure on $\mathbb B}\def\vp{\varphi$ with $\nu_0(\mathbb B}\def\vp{\varphi)=\infty.$ This measure is an infinite-dimensional version of the $\aa$-stable jump measure. Modifying arguments from \cite[Theorem 3.1]{W10a} and \cite[Theorem 1.1]{SW2} where the coupling property has been investigated in the finite-dimension setting, we have the following two assertions on the coupling property with estimates on the convergence rate. For $r>0$ and $z\in\mathbb B}\def\vp{\varphi$, let $B(z,r)=\{y\in \mathbb B}\def\vp{\varphi: \|z-y\|_\mathbb B}\def\vp{\varphi<r\}$ be the open ball at $z$ with radius $r$. \begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}\label{T1.2} Assume {\bf (A)}. Suppose that $\nu_0$ in $(\ref{C})$ is finite; i.e.\ $\nu_0(\mathbb B}\def\vp{\varphi)<\infty$, $\sigma} \def\ess{\text{\rm{ess}}$ is invertible with $\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|_\mathbb B}\def\vp{\varphi<\infty$, and $\|T_s\|_{\mathbb B}\def\vp{\varphi}\le c$ holds for some constant $c>0$ and all $s>0.$ \begin{itemize} \item[{\rm (i)}] If there exist $z_0\in\mathbb B}\def\vp{\varphi$ and $r_0>0$ such that \beq\label{C2} \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv):=\sup_{s\ge \vv, \|x\|_\mathbb B}\def\vp{\varphi\le 1} \int_{B(z_0, r_0)} \ff{\vp_{\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s x}(z)^2\rr_0(z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sx)^2}{ \rr_0(z) }\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)<\infty,\ \ \vv>0,\end{equation} then there exists a constant $C>0$ such that \beq\label{Coupling} \|P_t(x,\cdot)-P_t(x+y,\cdot)\|_{var} \le C (1+\|y\|_\mathbb B}\def\vp{\varphi) \inf_{\vv\in (0,1)} \bigg(\vv +\sqrt{\ff{ {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}}{ t}}\bigg),\ \ t>0,\ x,y\in\mathbb B}\def\vp{\varphi \end{equation} holds. \item[{\rm (ii)}]If there exist $z_0\in \mathbb B}\def\vp{\varphi$ and $r_0>0$ such that \beq\label{C22}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_2(\vv):=\sup_{s\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le 1} \int_{B(z_0, r_0)} \frac{\varphi_{\sigma^{-1}T_s x}(z)^2\vee 1}{\rho_0(z)}\,\mu(dz)<\infty,\ \ \vv>0,\end{equation} then there exist two constants $C>0 $ such that for all $x,y\in\mathbb B}\def\vp{\varphi$ and $t>0$, \begin{equation}\label{Coupling1}\begin{aligned} \|P_t(x,\cdot)-P_t(y,\cdot)\|_{var}\le C (1+\|x-y\|_{\mathbb B}\def\vp{\varphi})\inf_{\varepsilon\in (0,1)}\!\!\bigg(\varepsilon+ \sqrt{\frac{{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_2(\vv)}}{{t}}} \bigg).\end{aligned} \end{equation} \end{itemize} \end{thm} Using $\rr_0\land 1$ in place of $\rr_0$, one may replace \eqref{C2} by $$\tilde} \def\Ric{\text{\rm{Ric}}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv):=\sup_{s\ge \vv, \|x\|_\mathbb B}\def\vp{\varphi\le 1} \int_{B(z_0, r_0)} \ff{\vp_{\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s x}(z)^2}{1\land \rr_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)<\infty,\ \ \vv>0.$$ If $\inf_{z\in B(x_0,r_0)}\rho_0(z)>0$, then this condition and \eqref{C22} are equivalent. But in general (\ref{C2}) and (\ref{C22}) are incomparable. Next, it is easy to see that the convergence rate implied by (\ref{Coupling}) or \eqref{Coupling1} is in general slower than $\ff 1 {\ss t}$. Our next result shows that if $\vp$ and $\rr_0$ are regular enough, the convergence could be exponentially fast. \begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}\label{T1.3} Assume {\bf (A)}. Suppose that $\nu_0$ in $(\ref{C})$ is finite with $\ll_0:=\nu_0(\mathbb B}\def\vp{\varphi)\in (0,\infty)$, $\|T_s\|_\mathbb B}\def\vp{\varphi\le c\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll s}$ and \beq\label{Z1} \int_{\mathbb B}\def\vp{\varphi} \Big(|\rr_0(z)-\rr_0(z+h)|+\rr_0(z)|\vp_h(z)-1|\Big)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\le c\|h\|_\H,\ \ \|h\|_\H\le 1\end{equation} holds for some constants $c,\ll>0$ and all $s\ge 0$. If \beq\label{Z2} \sup_{t\ge 1}\ff 1 {1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 t}} \int_0^t\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 r} \Big(\sup_{\|z\|_\mathbb B}\def\vp{\varphi\le 1}\sup_{s\ge r}\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sz\|_\H\Big)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r <\infty,\end{equation} then there exists a constant $C>0$ such that \beq\label{Z3} \|P_t(x,\cdot)-P_t(y,\cdot)\|_{var} \le C (1+\|x-y\|_\mathbb B}\def\vp{\varphi)\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ff{\ll_0\ll t}{\ll_0+\ll}},\ \ x,y\in \mathbb B}\def\vp{\varphi, t\ge 0.\end{equation} \end{thm} Following the line of \cite[Section 3]{W11}, one may also naturally investigate gradient estimates and derivative formula for $P_t$. It is not difficult to present a formal result under a condition similar to \cite[(3.1)]{W11}, for instance: \begin} \def\beq{\begin{equation}} \def\F{\scr F{prp}\label{PP0} Assume that $\{h\in\H: \sup_{s\in [0,1]}\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sh\|_\H<\infty\}$ is dense in $\mathbb B}\def\vp{\varphi$. If there exists a non-negative function $g$ on $\mathbb B}\def\vp{\varphi$ such that $ \nu_0(\{g>0\})=\infty$, $\rr_0g$ is bounded and Lipschitz continuous in $\|\cdot\|_\H$, and \beq\label{T1-2} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} q(t)&:=\sup_{\|h\|_\H\in (0,1]}\bigg\{\Big(1+ \ff{\mu(|\vp_h-1|)} {\|h\|_\H}\Big)\int_0^\infty \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-t\nu_0(1-\exp[-rg])}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \\ &\qquad\qquad\qquad \quad+ \ff{\mu\big(|g-g(\cdot-h)|\big)} {\|h\|_\H}\int_0^\infty r\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-t\nu_0(1-\exp[-rg])}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\bigg\} <\infty,\ \ t>0,\end{split}\end{equation} then there exists a constant $C_1>0$ such that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\nabla} \def\pp{\partial} \def\EE{\scr E_y P_t f(x)|:&=\,\limsup_{\vv\downarrow 0} \ff 1\vv |P_tf(x+\vv y)-P_tf(x)|\\ &\le \,C_1\|f\|_\infty q(t) \int_0^t \|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy\|_\H\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s,\ \ f\in\B_b(\mathbb B}\def\vp{\varphi), t>0, x,y\in \mathbb B}\def\vp{\varphi.\end{split}\end{equation*} Suppose moreover that $\|T_s\|_\mathbb B}\def\vp{\varphi\le c\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll s}$ for some constants $c,\ll>0$ and all $s\ge 0$. Then $$\|P_t(x,\cdot)-P_t(y,\cdot)\|_{var} \le C_2 (1+\|x-y\|_\mathbb B}\def\vp{\varphi)\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{- \ll t },\ \ x,y\in \mathbb B}\def\vp{\varphi, t\ge 0$$ holds for some constant $C_2>0.$ \end{prp} Unfortunately, in the moment we do not have any non-trivial example in infinite dimensions to illustrate condition (\ref{T1-2}). Indeed, it seems that in infinite dimensions the uniform norm of the gradient of $P_t$ $$\|\nabla} \def\pp{\partial} \def\EE{\scr E P_t\|_\infty:= \sup\{|\nabla} \def\pp{\partial} \def\EE{\scr E_y P_t f(x)|:\ \|y\|_\mathbb B}\def\vp{\varphi\le 1, x\in \mathbb B}\def\vp{\varphi, \|f\|_\infty\le 1\}$$ is most likely infinite for any $t>0.$ The intuition is that comparing with a cylindrical noise given in \cite[Assumption 2.2]{X2}, which is strong enough along single directions so that the noise might not take values in $\mathbb B}\def\vp{\varphi$, our non-cylindrical L\'evy process seems too weak to imply a bounded gradient estimate of $P_t$. Nevertheless, we are able to estimate the uniform gradient of a modified version of $P_t$ (cf. Proposition \ref{PP2} below), which implies the desired exponential convergence in (\ref{Z3}). We remark that the derivative formula and gradient estimate are investigated in \cite{T,Zhang} for SDEs on $\R^d$ driven by L\'evy noises, where in \cite{T} the process may contain a diffusion part but extensions of the main results to infinite dimensions are not yet available, while in \cite{Zhang} the main result was also extended to a class of semi-linear SPDEs driven by cylindrical $\aa$-stable processes. Both papers are quite different from the present one, where we aim to describe regularity properties of the semigroup merely using the L\'evy measure of the noise. \ We will prove Theorems \ref{T1.1} (also Proposition \ref{PP0}), \ref{T1.2} and \ref{T1.3} in the following three sections respectively. In Section 5 we present two specific examples, with $\mu$ the Wiener measure on a Brownian path space and the Gaussian measure on an Hilbert space respectively, to illustrated these results. \section{Proofs of Theorem \ref{T1.1} and Proposition \ref{PP0}} The key technique of the study is the coupling by change of measure. For readers' convenience, let us briefly recall the main idea of the argument. To investigate e.g. the continuity of $P_tf$ along $y\in \mathbb B}\def\vp{\varphi$, for any $x\in\mathbb B}\def\vp{\varphi$ we construct a family of processes $\{X_\cdot^\vv\}_{\vv\in [0,1)}$ and the associated probability densities $\{R_\vv\}_{\vv\in {[0,1)}}$ such that \begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate} \item[(1)] $X_0^\vv = x+\vv y,\ X_t^\vv= X_t^0,\ \vv\in [0,1),\ t>0;$ \item[$(2)$] Under the probability $R_\vv \P$, the process $X_\cdot^\vv$ is associated to the transition semigroup $(P_s)_{s\ge 0};$ \item[$(3)$] $\lim_{\vv\to 0} R_\vv =R_0=1$ holds in $L^1(\P).$ \end{enumerate} Then, for any bounded measurable function $f$ and $t>0$, $$\lim_{\vv\to 0} P_t f(x+\vv y)= \lim_{\vv\to 0} \E\big[R_\vv f(X_t^\vv)\big] =\lim_{\vv\to 0} \E\big[R_\vv f(X_t^0)\big] = \E \big[R_0 f(X_t^0)\big] = P_t f(x).$$ To realize this idea in the present setting, the following Lemma \ref{L2.1} will play a crucial role. For fixed $t>0$, let $\LL$ be the distribution of $L:=(L_s)_{s\in [0,t]}$ which is a probability measure on the paths pace \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} W_t=\big\{w: [0,t]\to \mathbb B}\def\vp{\varphi\ &\text{is\ right-continuous\ having left limits}\big\} \end{split}\end{equation*}equipped with the Skorokhod metric. For any $w\in W_t$, let $$w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s):= \sum_{s\in [0,t], \DD w_s \ne 0} \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_{(\DD w_s, s)},$$ which records jumps of the path $w$, where $\DD w_s= w_s-w_{s-}$. Let $$w(g)=\int_{\mathbb B}\def\vp{\varphi\times [0,t]}g(z,s)\,w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)=\sum_{s\in [0,t], \DD w_s\ne 0} g(\DD w_s, s),\ \ \ g\in L^1(w).$$ A function $g$ on $\mathbb B}\def\vp{\varphi$ will be also regarded as a function on $\mathbb B}\def\vp{\varphi\times [0,t]$ by letting $g(z,s)=g(z)$ for $(z,s)\in\mathbb B}\def\vp{\varphi\times [0,t].$ Moreover, write $L=L^1+L^0$, where $L^1$ and $L^0$ are two independent L\'evy processes with L\'evy measure $\nu-\nu_0$ and $\nu_0$ respectively, and $L^0$ does not have a Gaussian term. Let $\LL^1$ and $\LL^0$ be the distributions of $L^1$ and $L^0$ respectively. We have $\LL=\LL^1*\LL^0.$ Repeating the proof of \cite[Lemma 2.1]{W11} where $\mathbb B}\def\vp{\varphi=\R^d$, we have the following result. \begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L2.1} For any $h\in L^1(W_t\times\mathbb B}\def\vp{\varphi\times [0,t]; \LL^0\times \nu_0\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)$, \beq\label{F1}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} h(w,z,s)\, \LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &=\int_{W_t}\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} h(w-z1_{[s,t]},z, s)\,w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s).\end{split}\end{equation}\end{lem} To prove Theorem \ref{T1.1}, we also need the following two more lemmas. \begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L2.2} Let $y\in\mathbb B}\def\vp{\varphi$ such that $\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y\in\H$ for any $s>0$, and let $g$ be a non-negative measurable function on $\mathbb B}\def\vp{\varphi$ such that $\nu_0(g):=\int_\mathbb B}\def\vp{\varphi g\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\nu_0 <\infty$ and $w(g)>0$ for $\LL^0$-a.e. $w$. Let $$\Phi_\vv(w,z,s)= \ff{ \vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)(\rr_0 g)(z-\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)}{w(g) + g(z-\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)},\ \ \vv\ge 0.$$ If $(\ref{N})$ holds for any $h\in \H$, then $\{\Phi_\vv\}_{\vv\in [0,1)}$ is uniformly integrable w.r.t. $\LL^0\times\mu\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$ on $W_t\times \mathbb B}\def\vp{\varphi\times [0,t].$ \end{lem} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Since $\vp_0\equiv 1$, applying (\ref{F1}) to $h(w,z,s)= \ff{g(z)}{w(g)}$ we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{WW1}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{W_t \times\mathbb B}\def\vp{\varphi\times [0,t]} \Phi_0(w,z,s)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \int_{W_t \times\mathbb B}\def\vp{\varphi\times [0,t]} \ff{g(z)}{w(g)+g(z)}\, \LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \int_{W_t} \LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{g(z)}{w(g)}\,w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)\\ &= 1.\end{split}\end{equation} Next, by (\ref{1.1}) and the integral transform $z\mapsto z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy$, for any $F\in \B_b(W_t\times \mathbb B}\def\vp{\varphi\times [0,t])$ we have \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{WW}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} F(w, z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y, s)\Phi_0(w,z,s)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \\ &= \int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F(w, z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y, s)(\rr_0g)(z)}{w(g)+g(z)}\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \\ &=\int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} F(w, z, s)\Phi_\vv(w,z,s)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{split}\end{equation} Letting $F=1$ and combining this with (\ref{WW1}), we conclude that $\{\Phi_\vv\}_{\vv\in [0,1)}$ are probability densities w.r.t. $\LL^0\times \mu\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$. Moreover, applying (\ref{WW}) to $F(w,z,s)= 1_{\{\Phi_\vv>R\}}$ for $R>0$ and letting $$\eta(w,z,s)= \sup_{\vv\in (0,1)} \ff{(\rr_0g)(z)}{w(g)+g(z)} \vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)$$ which is finite $\LL^0\times \mu\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$-a.e., we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\sup_{\vv\in (0,1)}\int_{W_t\times \mathbb B}\def\vp{\varphi\times [0,t]} (\Phi_\vv 1_{\{\Phi_\vv>R\}})(w,z,s)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \\ &\le\int_{W_t\times \mathbb B}\def\vp{\varphi\times [0,t]} (\Phi_0 1_{\{\eta >R\}})(w,z,s)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \end{split}\end{equation*} which goes to zero as $R\to\infty$ by the dominated convergence theorem. \end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L2.3} Let $E$ be a topology space and $C_b(E)$ be the class of all bounded continuous functions on $\mathbb B}\def\vp{\varphi$. Let $\mu_0$ be a finite measure on the Borel $\sigma} \def\ess{\text{\rm{ess}}$-field $\B$ such that $C_b(E)$ is dense in $L^1(\mu_0)$. Let $\{f_n\}_{n\ge 1}$ be a sequence of uniformly integrable functions w.r.t. $\mu_0$ such that $$\lim_{n\to\infty} \int_E (Ff_n)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0= \int_E (Ff_0)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0 $$ holds for some $f_0\in L^1(\mu_0)$ and all $F\in C_b(E)$. Then it holds also for any $F\in \B_b(E).$ \end{lem} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Let $\vv(R)= \sup_{n\ge 1} \mu_0 (|f_n-f_0|1_{\{|f_n-f_0|>R\}})$ which goes to zero as $R\to\infty$. For any $F\in \B_b(E)$, let $\{F_m\}_{m\ge 1}\subset C_b(E)$ such that $\|F_m\|_\infty\le \|F\|_\infty$ and $\mu_0(|F_m-F|)\le \ff 1 m.$ Then \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \bigg|\int_E F(f_n-f_0)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0\bigg|& \le \bigg|\int_E F(f_n-f_0)1_{\{|f_n-f_0|\le R\}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0\bigg|+\|F\|_\infty\vv(R)\\ &\le \bigg|\int_EF_m(f_n-f_0)1_{\{|f_n-f_0|\le R\}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0\bigg|+\|F\|_\infty\vv(R)+\ff R m\\ &\le \bigg|\int_EF_m(f_n-f_0)\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\mu_0\bigg|+2\|F\|_\infty\vv(R)+\ff R m.\end{split}\end{equation*} By first letting $n\to\infty$ then $m\to\infty$ and finally $R\to\infty$, we complete the proof.\end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}[Proof of Theorem \ref{T1.1}] (1) Let $f\in \B_b(\mathbb B}\def\vp{\varphi)$ and $x,y\in\mathbb B}\def\vp{\varphi$ be fixed. For any $\vv>0$, let $$F_\vv(w)= f\bigg(T_t (x+\vv y) +\int_0^tT_{t-s}\sigma} \def\ess{\text{\rm{ess}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w_s\bigg),$$ where $\int_0^t T_{t-s}\sigma} \def\ess{\text{\rm{ess}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w_s$ is the It\^o stochastic integral which is $\LL$-a.e. well-defined. Let e.g. $g = \ff 1 {\rr_0\lor 1}.$ We have $\nu_0(g)<\infty$ and, since $\nu_0(\mathbb B}\def\vp{\varphi)=\infty$ and $g>0$, $w(g)>0$ for $\LL^0$-a.e. $w$. Then, by (\ref{S}) and Lemma \ref{L2.1} for $$h(w^0,z,s)= \ff{F_0(w^1+w^0+(z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)1_{[s,t]})g(z)}{w^0(g)+g(z)},$$ we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &P_t f(x+\vv y) \\ &= \E F_\vv(L^1+L^0)\\ &= \int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F_\vv(w^1+w^0) g(z)}{w^0(g)}\,w^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)\\ &=\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F_0(w^1+w^0+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y 1_{[s,t]})g(z)}{w^0(g)}\,w^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)\\ &= \int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F_0(w^1+w^0+(z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y) 1_{[s,t]})g(z)}{w^0(g)+ g(z) }\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \\ &= \int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F_0(w^1+w^0+(z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y) 1_{[s,t]})(\rr_0g)(z)}{w^0(g)+ g(z) }\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{split}\end{equation*} Since $\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y\in \H$ so that (\ref{1.1}) implies $$\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z- \vv \sigma} \def\ess{\text{\rm{ess}}^{-1} T_s y)= \vp_{\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z),$$ by using the integral transform $z\mapsto z -\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy$ and noting that $\LL=\LL^1 *\LL^0$, we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{F4'}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & P_t f(x+\vv y)\\ &= \int_{W_t} \LL(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\int_{\mathbb B}\def\vp{\varphi\times [0,t]} \ff{F_0(w +z 1_{[s,t]})(\rr_0 g)(z-\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)}{w^0(g) + g(z-\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)}\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &=\int_{W_t} \LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1) \int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} F_0(w^1+w^0+z1_{[s,t]}) \Phi_\vv(w^0,z,s) \,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\,\mu(d z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{split}\end{equation} Therefore, it suffices to show that $$ \lim_{\vv\to 0} \int_{W_t\times \mathbb B}\def\vp{\varphi\times [0,t]}(F \Phi_\vv)(w,z,s) \,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w) \, \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) \,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s =\int_{W_t\times \mathbb B}\def\vp{\varphi\times [0,t]}(F \Phi_0)(w,z,s) \,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w) \, \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s $$ holds for any $F\in \B_b(W_t\times \mathbb B}\def\vp{\varphi\times [0,t]).$ According to (\ref{WW}), this holds provided $F\in C_b(W_t\times\mathbb B}\def\vp{\varphi\times [0,t]).$ Since the Borel $\sigma} \def\ess{\text{\rm{ess}}$-field on the Polish space $W_t\times \mathbb B}\def\vp{\varphi\times [0,t]$ is induced by bounded continuous functions, $C_b(W_t\times \mathbb B}\def\vp{\varphi\times [0,t])$ is dense in $L^1(\LL^0\times \mu\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)$. Thus, the desired assertion follows from Lemmas \ref{L2.2} and \ref{L2.3}. (2) For any sequence $\{y_n\}\subset \mathbb B}\def\vp{\varphi$ converging to $0$ as $n\to \infty$, define $$\Psi_n(w,z,s)= \ff{\varphi_{\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y_n}(z) (\rr_0g)(z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy_n)}{w(g)+g(z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy_n)},\ \ n\ge 1.$$ Using $\sigma} \def\ess{\text{\rm{ess}}^{-1} T_s y_n$ to replace $\vv h$ in the proof of Lemma \ref{L2.2}, we see that (\ref{1.5'}) implies that $\{\Psi_n\}_{n\ge 1}$ is uniformly integrable w.r.t. $\LL^0\times \mu\times \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$ on $W_t\times\mathbb B}\def\vp{\varphi\times [0,t]$. Therefore, using $\sigma} \def\ess{\text{\rm{ess}}^{-1} T_s y_n$ to replace $\vv h$ in the proof of (1), we obtain $\lim_{n\to \infty} P_t f(x+y_n)= P_t f(x)$ for any $f\in \B_b(\mathbb B}\def\vp{\varphi), t>0$ and $x\in \mathbb B}\def\vp{\varphi.$ \end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}[Proof of Proposition \ref{PP0}] Since $\{h\in\H: \sup_{s\in [0,1]}\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sh\|_\H<\infty\}$ is dense in $\mathbb B}\def\vp{\varphi$, it suffices to prove for $y\in\H$ such that $\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy\|_\H\le 1$ for $s\in [0,1].$ Since the boundedness of $\rr_0g$ implies $\nu_0(g)<\infty$ and $\nu_0(\{g>0\})=\infty$ implies $w(g)>0, \LL^0$-a.e., (\ref{F4'}) holds true. By (\ref{F4'}) and (\ref{T1-2}) we have \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{WFY1}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\ff{|P_tf(x+\vv y)-P_tf(x)|}{\vv} \\ &\le \ff{\|f\|_\infty}\vv \int_{W_t\times\mathbb B}\def\vp{\varphi\times [0,t]} |\Phi_\vv(w,z,s)-\Phi_0(w,z,s)|\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s,\ \ \vv>0. \end{split}\end{equation} Since $\rr_0g$ is bounded and Lipschitz continuous in $\|\cdot\|_\H$, there exists a constant $c_1>0$ such that \beq\label{WFY2} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &|\Phi_\vv(w,z,s)-\Phi_0(w,z,s)|\\ &\le \ff{|\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)-1|}{w(g)} +\bigg|\ff{(\rr_0g)(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)}{w(g)+g (z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)}-\ff{(\rr_0g)(z)}{w(g)+g(z)}\bigg|\\ &\le \ff{|\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)-1|+c_1\|\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy\|_\H}{w(g)} +\ff{c_1|g(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)-g(z)| }{w(g)^2}.\end{split}\end{equation} Moreover, according to \cite[Lemma 2.2]{W11} with $\mathbb B}\def\vp{\varphi$ in place of $\R^d$, for any $\theta>0$, we have $$\int_{W_t} \ff{\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)}{w(g)^\theta}= \ff 1 {\Gamma} \def\gg{\gamma(\theta)} \int_0^\infty r^{\theta-1} \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-t\nu_0(1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-rg})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r.$$ Combining this with (\ref{WFY1}) and (\ref{WFY2}) and letting $\vv\to 0$, we obtain the desired gradient estimate. According to the proof of Theorem \ref{T1.3} in Section 4 with $P_t^1$ replaced by $P_t$, this along with the assumption on $T_t$ implies the second assertion.\end{proof} \section{Proof of Theorem \ref{T1.2}} By the triangle inequality for $\|\cdot\|_{var}$, it suffices to prove both assertions for small enough $\|y\|_\mathbb B}\def\vp{\varphi$. \subsection{Case (i)} Let $\|y\|_\mathbb B}\def\vp{\varphi\le \ff {1\land (\ff{r_0}2)} {1+c\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|_\mathbb B}\def\vp{\varphi}$, which implies that \beq\label{BB} \|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y\|_\mathbb B}\def\vp{\varphi+\|y\|_\mathbb B}\def\vp{\varphi \le 1\land \ff{r_0} 2,\ \ \ s\in [0,t].\end{equation} Moreover, since $\|P_t (x,\cdot)-P_t(x+y,\cdot)\|_{var}\le 2$ holds for all $x,y\in\mathbb B}\def\vp{\varphi$ and $t>0$, we only have to prove the desired inequality for large $t>0$. From now on, let us assume $t\ge 2$ and (\ref{BB}). Now, let $t\ge 2$ and $x,y\in \mathbb B}\def\vp{\varphi$ such that (\ref{BB}) holds. Since $T_s \sigma} \def\ess{\text{\rm{ess}}$ is bounded in $\mathbb B}\def\vp{\varphi$ uniformly in $s$, for any $z\in\mathbb B}\def\vp{\varphi$, $$J^z(w):= T_t z+ \int_0^t T_{t-s}\sigma} \def\ess{\text{\rm{ess}}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w_s$$ is $\LL$-a.e. (also $\LL^1$-a.e. and $\LL^0$-a.e.) defined. Moreover, due to (\ref{S}) and $L=L^1+L^0$, \beq\label{PP}X_t^z= J^z(L)= J^z(L^1+L^0),\ \ \ z\in \mathbb B}\def\vp{\varphi, t>0.\end{equation} Next, let $$ \tau_1(w)= \inf\{s>0: \DD w_s\ne 0\}, \ \ \tau_{i+1}(w)= \inf\{s>\tau_i(w):\ \DD w_s\ne 0\},\ \ \ i\ge 1.$$ Since $\ll_0=\nu_0(\mathbb B}\def\vp{\varphi)\in (0,\infty),$ we have $\P(\tau_1(L^0)\ge s)=\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 s}\in (0,1)$ for $s>0$, and $\tau_i(L^0)\uparrow \infty$ as $i\uparrow \infty$. Moreover, let $$N_s(w)=\#\{i\ge 1:\ \tau_i(w)\le s\},\ \ \ s\ge 0.$$ Then $\{N_s(L^0)\}_{s\in [0,t]}$ is a Poisson process with parameter $\ll_0$. Similarly, let $$ \tilde} \def\Ric{\text{\rm{Ric}} \tau_1(w)= \inf\{s> 1: \DD w_s\ne 0\}, \ \ \tilde} \def\Ric{\text{\rm{Ric}} \tau_{i+1}(w)= \inf\{s>\tilde} \def\Ric{\text{\rm{Ric}} \tau_i(w):\ \DD w_s\ne 0\},\ \ \ i\ge 1$$ and $$\tilde} \def\Ric{\text{\rm{Ric}} N_s(w)= N_{s+1}(w)-N_1(w)=\#\{i\ge 1:\ \tilde} \def\Ric{\text{\rm{Ric}}\tau_i\le s+1\}=\#\{i\ge 1: \ 1<\tau_i\le s+1\},\ \ s\in [0,t-1].$$ Then $\{\tilde} \def\Ric{\text{\rm{Ric}} N_s(L^0)\}_{s\in [0,t-1]}$ is a Poisson process with parameter $\ll_0$, which is independent of $\{\tau_1(L^0)> \vv\}=\{N_\vv(L^0)=0\}$ for $\vv\in (0,1).$ Finally, let \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\xi_i(w)= 1_{B(z_0,\ff{r_0}2)}(\DD w_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}),\\ &\tilde} \def\Ric{\text{\rm{Ric}}\xi_i(w) = \ff{\rr_0(\DD w_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}+ \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}y)}{\rr_0(\DD w_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)})} \big(1_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}y, \ff{r_0}2)}\vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}y}\big)(\DD w_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(w)}),\ \ \ i\ge 1.\end{split}\end{equation*} We have \beq\label{2*1} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{\mathbb B}\def\vp{\varphi\times [1,t]}1_{B(z_0,\ff{r_0}2)(z)}\, w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)= \sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}(w)} \xi_i(w),\\ &\int_{\mathbb B}\def\vp{\varphi\times [1,t]} \ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)}{\rr_0(z)} \big(1_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y, \ff{r_0}2)}\vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy}\big)(z)\,w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)= \sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}(w)} \tilde} \def\Ric{\text{\rm{Ric}} \xi_i(w), \end{split} \end{equation} where we set $\sum_{i=1}^0 =0$ by convention. From now on, we will simply denote $$\tau_i=\tau_i(L^0),\ \tilde} \def\Ric{\text{\rm{Ric}}\tau_i=\tilde} \def\Ric{\text{\rm{Ric}}\tau_i(L^0),\ \xi_i=\xi_i(L^0),\ N_s= N_s(L^0),\ \tilde} \def\Ric{\text{\rm{Ric}} N_s=\tilde} \def\Ric{\text{\rm{Ric}} N_s(L^0).$$ To characterize the coupling property of the solution, we first prove the following relation formula for $X_t^x$ and $X_t^{x+y}.$ \begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L3.1} For any $f\in \B_b(\mathbb B}\def\vp{\varphi)$ and $\vv\in (0,1)$, $$\E\bigg\{f(X_t^x)1_{\{\tau_1>\vv\}}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\xi_i\bigg\}= \E\bigg\{f(X_t^{x+y})1_{\{\tau_1>\vv\}}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\tilde} \def\Ric{\text{\rm{Ric}}\xi_i\bigg\}.$$\end{lem} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Since $\vv\in (0,1)$, $\{\tau_1(w)>\vv\}= \{\tau_1(w+z1_{[s,t]})>\vv\}$ holds for $s\in [1,t]$ and $z\in\mathbb B$. Moreover, by the definition of $J^x$ we have $$J^x(w^1+w^0)+ T_{t-s} \sigma} \def\ess{\text{\rm{ess}} z= J^x(w^1+w^0 +z1_{[s,t]}).$$ By Lemma \ref{L2.1} for $$h(w^0,z,s) = f(J^x(w^1,+w^0)+ T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z)1_{\{\tau_1\ge \vv\}\times B(z_0,\ff{r_0}2)\times [1,t]}(w^0,z,s) $$ with fixed $w^1$ and using (\ref{2*1}), we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split}&\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^x(w^1+w^0)+T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z) 1_{\{\tau_1>\vv\}}(w^0)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^x(w^1+w^0 + z1_{[s,t]})) 1_{\{\tau_1>\vv\}}(w^0+z1_{[s,t]})\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \int_{W_t^2}1_{\{\tau_1>\vv\}}(w^0)f(J^x(w^1+w^0 ) )\, \LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} \, w^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s).\end{split}\end{equation*} Combining this with (\ref{PP}) and the first equation in (\ref{2*1}) we arrive at \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation} \label{2*2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split}&\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^x(w^1+w^0)+T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z) 1_{\{\tau_1>\vv\}}(w^0)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \E\bigg\{f(X_t^x)1_{\{\tau_1>\vv\}}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\xi_i\bigg\}.\end{split}\end{equation} On the other hand, noting that $$J^x(w^1+w^0)+ T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z= J^{x+y} (w^1+w^0+(z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)1_{[s,t]}),$$ by Lemma \ref{L2.1} and the integral transform $z\mapsto z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y,$ we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^x(w^1+w^0)+T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z) 1_{\{\tau_1>\vv\}}(w^0)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &=\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^{x+y}(w^1+w^0 +\{z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy\}1_{[s,t]}))\\ &\qquad\qquad\qquad\qquad \qquad\qquad\qquad\qquad\times 1_{\{\tau_1>\vv\}}(w^0+ \{z-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy\}1_{[s,t]})\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &=\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{[1,t]}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \int_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy,\ff{r_0}2)} f(J^{x+y}(w^1+w^0 +z1_{[s,t]})) \\ &\qquad\qquad\qquad\qquad \qquad\qquad\times 1_{\{\tau_1>\vv\}}(w^0+ z1_{[s,t]})\ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y)}{\rr_0(z)} \vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y}(z)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &=\int_{W_t^2}1_{\{\tau_1>\vv\}}(w^0)f(J^{x+y}(w^1+w^0 ) ) \,\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0) \int_{\mathbb B}\def\vp{\varphi\times [1,t]}\ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy)}{\rr_0(z)}\\ &\qquad\qquad\qquad\qquad \qquad\qquad\times (1_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_s y,\ff{r_0}2)}\vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy})(z)\, w^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s).\end{split}\end{equation*} Combining this with (\ref{PP}) and the second equation in (\ref{2*1}), we conclude that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split}&\int_{W_t^2}\LL^1(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^1)\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w^0)\int_{B(z_0,\ff{r_0}2)\times [1,t]} f(J^x(w^1+w^0)+T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z) 1_{\{\tau_1>\vv\}}(w^0)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\\ &= \E\bigg\{f(X_t^{x+y})1_{\{\tau_1>\vv\}}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\tilde} \def\Ric{\text{\rm{Ric}}\xi_i\bigg\}.\end{split}\end{equation*} The desired formula follows from this and (\ref{2*2}). \end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L3.2} Given $\tilde} \def\Ric{\text{\rm{Ric}} N$, $\{\xi_i\}$ and $\{\tilde} \def\Ric{\text{\rm{Ric}} \xi_i\}$ are two conditionally i.i.d. sequences with $$\E(\xi_i|\tilde} \def\Ric{\text{\rm{Ric}} N)= \E(\xi_i^2|\tilde} \def\Ric{\text{\rm{Ric}} N)= \ff{\nu_0(B(z_0,\ff{r_0}2))}{\ll_0},$$ and $$\E(\tilde} \def\Ric{\text{\rm{Ric}}\xi_i|\tilde} \def\Ric{\text{\rm{Ric}} N)=\ff{\nu_0(B(z_0,\ff{r_0}2))}{\ll_0},\ \ \ \E(\tilde} \def\Ric{\text{\rm{Ric}}\xi_i^2|\tilde} \def\Ric{\text{\rm{Ric}} N)\le \ff{ \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\tilde} \def\Ric{\text{\rm{Ric}}\tau_1)}{\ll_0},\ \ i\ge1.$$\end{lem} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Since $\{\DD L^0_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}\}$ are i.i.d. and independent of $\tilde} \def\Ric{\text{\rm{Ric}} N$ with common distribution $\ff 1 {\ll_0} \nu_0$, and since $\tilde} \def\Ric{\text{\rm{Ric}} \tau_i$ is determined by $\tilde} \def\Ric{\text{\rm{Ric}} N$, it is clear that both $\{\xi_i\}$ and $\{\tilde} \def\Ric{\text{\rm{Ric}}\xi_i\}$ are conditionally i.i.d. sequences given $\tilde} \def\Ric{\text{\rm{Ric}} N$. Moreover, we have $$\E(\xi_i^2|\tilde} \def\Ric{\text{\rm{Ric}} N)=\E(\xi_i|\tilde} \def\Ric{\text{\rm{Ric}} N)=\E \xi_i = \ff{\nu_0(B(z_0,\ff{r_0}2))}{\ll_0}. $$ Noting that $\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)=\rr_0(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)$ and $\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z+h)=\vp_{-h}(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)$, we have \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \E(\tilde} \def\Ric{\text{\rm{Ric}} \xi_i|\tilde} \def\Ric{\text{\rm{Ric}} N) &= \ff 1 {\ll_0} \int_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i} y,\ff{r_0}2)} \ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y)}{\rr_0(z)}\vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y}(z)\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &=\ff 1 {\ll_0} \int_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y,\ff{r_0}2)} \rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y) \vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y}(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &=\ff 1 {\ll_0} \int_{B(z_0,\ff{r_0}2)} \rr_0(z)\, \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &=\ff{\nu_0(B(z_0,\ff{r_0}2))}{\ll_0}.\end{split}\end{equation*} Moreover, since $\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y\|_\mathbb B}\def\vp{\varphi\le 1\land \ff{r_0}2$ and $\tilde} \def\Ric{\text{\rm{Ric}}\tau_i\ge\tilde} \def\Ric{\text{\rm{Ric}}\tau_1$, we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \E({\tilde} \def\Ric{\text{\rm{Ric}} \xi_i}^2|\tilde} \def\Ric{\text{\rm{Ric}} N) &= \ff 1 {\ll_0} \int_{B(z_0-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y,\ff{r_0}2)} \ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y)^2}{\rr_0(z)^2}\vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y}(z)^2\,\nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\le \ff 1 {\ll_0} \int_{B(z_0,r_0)} \ff{\rr_0(z+\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i} y)^2 \vp_{-\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tilde} \def\Ric{\text{\rm{Ric}}\tau_i}y}(z)^2}{\rr_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\le \ff{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\tilde} \def\Ric{\text{\rm{Ric}}\tau_1)}{\ll_0}.\end{split}\end{equation*}This completes the proof. \end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}[Proof of Theorem \ref{T1.2} (i)] As explained in the beginning of this section, we assume that $t\ge 2$ and let $y$ satisfy (\ref{BB}). By Lemma \ref{L3.1} and $\tilde} \def\Ric{\text{\rm{Ric}}\tau_1\ge\tau_1$, for any $f\in \B_b(\mathbb B}\def\vp{\varphi)$ with $\|f\|_\infty\le 1$ we have \beq\label{3.3} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & \Big|\E\big(f(X_t^x)-f(X_t^{x+y})\big)1_{\{\tau_1>\vv\}}\Big|\\ &\le \E\bigg|1-\ff 1 {\nu_0(B(z_0,\ff{r_0}2))(t-1)}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\xi_i\bigg| + \E \bigg\{1_{\{\tilde} \def\Ric{\text{\rm{Ric}}\tau_1>\vv\}} \bigg|1-\ff 1 {\nu_0(B(z_0,\ff{r_0}2))(t-1)}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\tilde} \def\Ric{\text{\rm{Ric}}\xi_i\bigg| \bigg\}.\end{split}\end{equation} Noting that $\tilde} \def\Ric{\text{\rm{Ric}}\tau_1$ is determined by $\tilde} \def\Ric{\text{\rm{Ric}} N$, we obtain from Lemma \ref{L3.2} that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & \E \bigg\{1_{\{\tilde} \def\Ric{\text{\rm{Ric}}\tau_1>\vv\}} \bigg|1-\ff 1 {\nu_0(B(z_0,\ff{r_0}2))(t-1)}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\tilde} \def\Ric{\text{\rm{Ric}}\xi_i\bigg| \bigg\}^2\\ &=\E\bigg\{1_{\{\tilde} \def\Ric{\text{\rm{Ric}}\tau_1>\vv\}}\bigg(\ff {\sum_{i,j=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}} \E(\tilde} \def\Ric{\text{\rm{Ric}} \xi_i\tilde} \def\Ric{\text{\rm{Ric}} \xi_j|\tilde} \def\Ric{\text{\rm{Ric}} N)} {\nu_0(B(z_0,\ff{r_0}2))^2(t-1)^2}- \ff {2\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}} \E(\tilde} \def\Ric{\text{\rm{Ric}}\xi_i|\tilde} \def\Ric{\text{\rm{Ric}} N)} {\nu_0(B(z_0,\ff{r_0}2))(t-1)} +1\bigg)\bigg\}\\ &\le \E\bigg\{1_{\{\tilde} \def\Ric{\text{\rm{Ric}}\tau_1>\vv\}}\bigg(\ff {\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}^2-\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}{\ll_0^2(t-1)^2}+ \ff{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\tilde} \def\Ric{\text{\rm{Ric}}\tau_1)}{\ll_0\nu_0(B(z_0,\ff{r_0}2))^2(t-1)^2} -\ff {2\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}} {\ll_0(t-1)} +1\bigg)\bigg\}\\ &\le \ff{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}{\nu_0(B(z_0,\ff{r_0}2))^2(t-1)}.\end{split}\end{equation*} Similarly and even simpler, we have $$\E\bigg(1-\ff 1 {\nu_0(B(z_0,\ff{r_0}2))(t-1)}\sum_{i=1}^{\tilde} \def\Ric{\text{\rm{Ric}} N_{t-1}}\xi_i\bigg)^2\le \ff{1}{\nu_0(B(z_0,\ff{r_0}2))(t-1)}.$$ Combining these with (\ref{3.3}) and noting that $t-1\ge 1$, we arrive at $$\Big|\E\big(f(X_t^x)-f(X_t^{x+y})\big)1_{\{\tau_1>\vv\}}\Big|\le \ff{C_1\ss{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}}{\ss t}$$ for some constant $C_1>0$ independent of $t,x,y$ and $\vv\in (0,1)$. Therefore, there exists a constant $C>0$ independent of $t,x,y$ and $\vv\in (0,1)$ such that for $\|f\|_\infty\le 1$, \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |P_t f(x)-P_t f(x+y)|&\le \ff{C_1\ss{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho(\vv)}}{\ss t} +\E\Big|\big(f(X_t^x)-f(X_t^{x+y})\big)1_{\{\tau_1\le\vv\}}\Big|\\ &\le \ff{C_1\ss{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}}{\ss t} + 2\P(\tau_1\le\vv)\\ &= \ff{C_1\ss{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}}{\ss t}+2(1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0\vv})\\ &\le C\bigg(\vv+\ff{\ss{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_1(\vv)}}{\ss t}\bigg).\end{split}\end{equation*} This completes the proof. \end{proof} \subsection{Case (ii)} For every $\eta>0$, define ${\nu}_\eta$ on $\mathbb B}\def\vp{\varphi$ as follows: \begin{equation*} {\nu}_\eta(A) = \begin{cases} \nu(A), & \text{if\ \ } \nu(\mathbb B}\def\vp{\varphi)<\infty;\\ \nu(A\setminus \{z: \|z\|_{\mathbb B}\def\vp{\varphi}<\eta\}), & \text{if\ \ } \nu(\mathbb B}\def\vp{\varphi)=\infty, \end{cases} \end{equation*} where $A\in \B$. Then $\nu_\eta$ is a finite measure on $(\mathbb B}\def\vp{\varphi,\B)$. Recall that for any two finite measures $\pi_1$ and $\pi_2$ on $(\mathbb B}\def\vp{\varphi,\mathscr{B})$, $\pi_1\wedge\pi_2:=\pi_1-(\pi_1-\pi_2)^+$, where $(\pi_1-\pi_2)^{\pm}$ refers to the Jordan-Hahn decomposition of the signed measure $\pi_1-\pi_2$. In particular, $\pi_1\wedge\pi_2=\pi_2\wedge\pi_1$, and $$(\pi_1\wedge \pi_2)(\mathbb B}\def\vp{\varphi)=\frac{1}{2}\big(\pi_1(\mathbb B}\def\vp{\varphi)+\pi_2(\mathbb B}\def\vp{\varphi)-\|\pi_1-\pi_2\|_{var}\big).$$ The following is an extension of the main result in \cite{SW2} to the infinite-dimensional setting. \begin{thm}\label{th1} Let $X_t$ be the process determined by \eqref{E1}. Assume that $\sigma$ is invertible, and that there exist $\eta,\varrho>0$ such that \begin{equation}\label{th2233} \gamma(\eta, \varrho,\vv):=\inf_{t\ge\vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\big\{\nu _\eta\wedge (\delta_{\sigma^{-1}T_tx}*\nu_\eta)\big\}(\mathbb B}\def\vp{\varphi)>0 \end{equation} holds for any $\vv>0$. Then there exists a constant $C >0 $ such that for all $x,y\in\mathbb B}\def\vp{\varphi$ and $t>0$, \begin{equation}\label{th21} \|P_t(x,\cdot)-P_t(x+y,\cdot)\|_{var} \le C \big(1+\|y\|_{\mathbb B}\def\vp{\varphi}\big)\inf_{\vv\in (0,1)} \bigg(\vv+ \frac{1}{\sqrt{\gamma(\eta, \varrho,\vv)t}} \bigg). \end{equation} \end{thm} We postpone the proof to the end of this subsection and present the proof of Theorem \ref{T1.2} (ii). \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}[Proof of Theorem \ref{T1.2} (ii)] Without loss of generality, we assume that $0\notin B(z_0, r_0 )$. Otherwise, we may take $z_0'\in B(z_0,r_0)$ and $r_0'>0$ such that $0\notin B(z_0',r_0')\subset B(z_0,r_0)$, and use $B(z_0',r_0')$ to replace $B(z_0,r_0).$ Moreover, we take $\varrho\in(0,1)$ small enough such that $\|\sigma^{-1}T_t x\|\le 1\land \ff{r_0}4$ holds for all $\|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho$ and $t>0$. By \eqref{C}, \eqref{1.1} and the Cauchy-Schwarz inequality, for any $t\ge \vv$ and $\eta\in (0,\ff{r_0}4)$, \begin{equation*}\label{th211}\begin{aligned}&\inf_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho} \big\{\nu_\eta\wedge (\delta_{\sigma^{-1}T_t x }*\nu_\eta)\big\}(\mathbb B}\def\vp{\varphi)\\ &\ge \inf_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\int_{B(x_0,\ff{r_0}2)}\Big(\rho_0(z)\wedge \big(\rho_0(z-{\sigma^{-1}T_t x })\varphi_{\sigma^{-1}T_t x }(z)\big)\Big)\, \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\ge {\inf_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho} \Big(\int_{B(x_0,\ff{r_0}2)} \varphi_{\sigma^{-1}T_t x }(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\Big)^2}\\ &\qquad\qquad \times\bigg[{\sup_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)^2}{\rho_0(z)\wedge \big(\rho_0(z-{\sigma^{-1}T_t x })\varphi_{\sigma^{-1}T_t x }(z)\big)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)}\bigg]^{-1}. \end{aligned}\end{equation*} Since the measure $\mu$ has full support, \begin{equation*}\label{th211}\begin{aligned}&\inf_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\int_{B(x_0,\ff{r_0}2)} \varphi_{\sigma^{-1}T_t x }(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &=\inf_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\int_{B(x_0,\ff{r_0}2)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z- \sigma^{-1}T_t x)\\ &\ge \int_{B(x_0, \ff{r_0}4)} \,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)>0. \end{aligned}\end{equation*} On the other hand, by \eqref{C22}, for any $t\ge \varepsilon$, \begin{equation*}\label{th211}\begin{aligned}&{\sup_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)^2}{\rho_0(z)\wedge \big(\rho_0(z-{\sigma^{-1}T_t x })\varphi_{\sigma^{-1}T_t x }(z)\big)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)} \\ &\le \sup_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\bigg[\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)^2}{\rho_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)+\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)}{\rho_0(z-\sigma^{-1}T_t x)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\le \sup_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\bigg[\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)^2}{\rho_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)+\int_{B(x_0,\ff{r_0}2)}\frac{\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z-\sigma^{-1}T_t x )}{\rho_0(z-\sigma^{-1}T_t x )}\bigg]\\ &\le\sup_{t\ge \vv, \|x\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\bigg[\int_{B(x_0,\ff{r_0}2)}\frac{\varphi_{\sigma^{-1}T_t x }(z)^2}{\rho_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)+\int_{B(x_0,r_0)}\frac{1}{\rho_0(z)}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\bigg]\\ &<\infty.\end{aligned}\end{equation*} The required assertion \eqref{Coupling1} follows from the conclusions above and \eqref{th21}. \end{proof} \begin{proof}[Proof of Theorem \ref{th1}] As indicated in the proof of Theorem \ref{T1.2} (i), we only have to prove the result for $\|x-y\|_\mathbb B}\def\vp{\varphi\le\varrho$ and $t\ge 1$. To this end, we modify the argument from the proof of \cite[Theorem 1.1]{SW2}. For any $\eta>0$, let $L^\eta$ be a compound Poisson process on $\mathbb B}\def\vp{\varphi$ with L\'{e}vy measure $\nu_\eta$ such that $L^\eta$ and $L-L^\eta$ are independent L\'{e}vy processes. Then the random variables $$ X_t^{\eta,x}:=T_tx+\int_0^t T_{t-s}\sigma\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s^\eta $$ and $$ X_t^x-X_t^{\eta,x}:=\int_0^t T_{t-s}\sigma\,d(L_s-L_s^\eta) $$ are independent. Denote by $\mu_{\eta,t}$ the law of random variable $$ X_t^{\eta,0} := X_t^{\eta,x}-T_tx = \int_0^tT_{t-s}\sigma\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s^\eta. $$ Construct a sequence $\{\tau_i\}$ of i.i.d.\ random variables which are exponentially distributed with intensity $C_\eta=\nu_\eta(\mathbb B}\def\vp{\varphi)$, and introduce a further sequence $\{U_i\}$ of i.i.d.\ random variables on $\mathbb B}\def\vp{\varphi$ with law $\bar{\nu}_\eta=\nu_\eta/C_\eta$. We will assume that the random variables $\{U_i\}$ are independent of the sequence $\{\tau_i\}$. Then, according to \cite[Examples, Section 2]{App1}, $ L_t^\eta=\sum_{i=1}^{N_t}U_i$ for every $t\ge0$, where $N_t:=\sup\{k: \sum_{i=1}^k\tau_i\le t\}$, for $\sum_{i\in \varnothing}:=0$ by convention, is a Poisson process of intensity $C_\eta$. Therefore, the random variable \begin{equation}\label{proofs0} 1_{\{\tau_1\le t\}}\sum_{k=1}^\infty1_{\{N_t=k\}} \Big(T_{t-\tau_1}\sigma U_1+\cdots+T_{t-(\tau_1+\cdots+\tau_k)}\sigma U_k\Big) \end{equation} has the probability distribution $\mu_{\eta,t}$. Let $P_t(x,\cdot)$ and $P_t$ be the transition kernel and the transition semigroup of the Ornstein-Uhlenbeck process $X^x_t$. Similarly, we denote by $P^\eta_t(x,\cdot)$ and $P^\eta_t$ the transition kernel and the transition semigroup of $X_t^{\eta,x}$, and by $Q^\eta_t(x,\cdot)$ and $Q^\eta_t$ the transition kernel and the transition semigroup of $X_t^x-X_t^{\eta,x}$. By the independence of the processes $X_t^{\eta,x}$ and $X_t^x-X_t^{\eta,x}$, we get \begin{equation}\label{proofs2}\begin{aligned} \|P_t(x,\cdot)-P_t(y,\cdot)\|_{var} &= \sup_{\|f\|_\infty\le 1}\big|P_tf(x)-P_tf(y)\big|\\ &=\sup_{\|f\|_\infty\le 1} \big|P_t^\eta Q_t^\eta f(x)-P_t^\eta Q_t^\eta f(y)\big|\\ &\le \sup_{\|h\|_\infty\le 1} \big|P^\eta_th(x)-P^\eta_th(y)\big|\\ &=\sup_{\|h\|_\infty\le 1}\Big|\E(h(X_t^{\eta,x}))-\E(h(X_t^{\eta,y}))\Big|. \end{aligned}\end{equation} Following the argument leading to \cite[(2.11)]{SW2}, we may write $$ \E f\bigl(X_t^{\eta,x}\bigr) =\int_\mathbb B}\def\vp{\varphi f\bigl(T_tx+z\bigr)\,\mu_{\eta,t}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) =f\bigl(T_tx\bigr)\,\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t}+Hf(x),\ \ f\in\B_b(\mathbb B}\def\vp{\varphi) $$ for $$ Hf(x) =\sum_{k=1}^\infty \int_{I_{t,k}} C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\int_{\mathbb B}\def\vp{\varphi} f\bigl(T_tx+z\bigr)\,\mu_{t_1,\cdots,t_k}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z),$$ where \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &I_{t,k}:= \Big\{(t_1,\cdots, t_k,t_{k+1})\in (0,\infty)^{k+1}: \ \sum_{i=1}^k t_i\le t<\sum_{i=1}^{k+1}t_i\Big\},\\ & \mu_{t_1,\cdots,t_{k}} := (\bar \nu_\eta)^k\circ J_{t_1,\cdots, t_k}^{-1},\\ &J_{t_1,\ldots,t_k}(y_1,\ldots,y_k) :=T_{t-t_1}\sigma y_1+\cdots+T_{t-(t_1+\cdots+t_k)}\sigma y_k, \ \ y_1,\cdots, y_k\in\mathbb B}\def\vp{\varphi.\end{split}\end{equation*} Then, for any $t\ge1$ and $\vv\in (0,1)$, \begin{equation}\label{proofs3}\begin{aligned} &\sup_{\|h\|_\infty\le 1}\Big|\E(h(X_t^{\eta,x}))-\E(h(X_t^{\eta,y}))\Big|\\ &\le \sup_{\|h\|_\infty\le 1}\bigg|\E\Big(\!\big(h(X_t^{\eta,x})-h(X_t^{\eta,y})\big)1_{\{\tau_1\le \vv\}}\!\Big)\bigg| + \sup_{\|h\|_\infty\le 1}\bigg|\E\Big(\!\big(h(X_t^{\eta,x})-h(X_t^{\eta,y})\big)1_{\{\tau_1\ge \vv\}}\!\Big)\bigg|\\ &\le 2\P( \tau_1\le \vv)+2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t} +\sum_{k=1}^\infty \int_{I_{t,k}\cap\{(0,\infty)^{k+1}:\,t_1\ge\vv\}}C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\\ &\qquad\qquad\times\sup_{\|h\|_\infty\le 1} \bigg|\int_{\mathbb B}\def\vp{\varphi} h\big(T_tx+z\big)\,\mu_{t_1,\cdots,t_k}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) - \int_{\mathbb B}\def\vp{\varphi} h\big(T_ty+z\big)\,\mu_{t_1,\cdots,t_k}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\biggr|\\ &=2(1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta\vv})+2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t} +\sum_{k=1}^\infty \int_{I_{t,k}\cap\{(0,\infty)^{k+1}:\,t_1\ge\vv\}} C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\\ &\qquad\qquad\times\sup_{\|h\|_\infty\le 1} \bigg|\int_{\mathbb B}\def\vp{\varphi} h\big(T_t(x-y)+z\big)\,\mu_{t_1,\cdots,t_k}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) - \int_{\mathbb B}\def\vp{\varphi} h(z)\,\mu_{t_1,\cdots,t_k}(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\bigg|\\ &\le 2C_\eta\vv+ 2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t} \\ &\quad+\sum_{k=1}^\infty\int_{I_{t,k}\cap\{(0,\infty)^{k+1}:\,t_1\ge\vv\}} C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\\ &\qquad\qquad\qquad\qquad\qquad\qquad\quad\times\|\delta_{T_t(x-y)}*\mu_{t_1,\cdots,t_k}-\mu_{t_1,\cdots,t_k}\|_{var}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}. \end{aligned}\end{equation} To estimate $\|\delta_{T_t(x-y)}*\mu_{t_1,\cdots,t_k}-\mu_{t_1,\cdots,t_k}\|_{var}$ for any $t_1\ge \vv$ and $ t\ge t_1+\cdots+t_k$, we will use the Mineka and Lindvall-Rogers couplings for random walks as in \cite{SW, SW2}. The remainder of this part is based on steps 4 and 5 in the proof of \cite[Theorem 1.1]{SW2}. In order to ease notations, we set $\mathsf{n}:=\bar{\nu}_\eta$ and $\mathsf{n}^{a}:=\delta_a*\bar\nu_\eta$ for any $a\in\mathbb B}\def\vp{\varphi$. For any $i\ge 1$, let $(U_i,\Delta U_i)\in \mathbb B}\def\vp{\varphi \times \mathbb B}\def\vp{\varphi$ be a pair of random variables with the following distribution $$ \mathbb P\big((U_i,\Delta U_i)\in C\times D\big) = \begin{cases} \qquad \frac 12 (\mathsf{n}\wedge\mathsf{n}^{-a_i})(C), & \text{if\ \ } D=\{a_i\};\\ \qquad \frac 12 (\mathsf{n}\wedge\mathsf{n}^{a_i})(C), & \text{if\ \ } D= \{-a_i\};\\ \big(\mathsf{n}- \frac 12 (\mathsf{n}\wedge\mathsf{n}^{-a_i}+\mathsf{n}\wedge\mathsf{n}^{a_i})\big)(C), & \text{if\ \ } D=\{0\}; \end{cases} $$ where $C\in\B$, $a_i=\sigma^{-1}\, T_{t_1+\cdots+t_i}\,(x-y)$ and $D$ is any of the following three sets: $\{-a_i\}$, $\{0\}$ or $\{a_i\}$. It follows that, cf.\ see \cite[Lemma 3.2]{SW}, \begin{align*} \mathbb P\big(\Delta U_i=-a_i\big) =\frac{1}{2}\big(\mathsf{n}\wedge\big(\delta_{a_i}*\mathsf{n})\big)(\mathbb B}\def\vp{\varphi) =\frac{1}{2}\big(\mathsf{n}\wedge\big(\delta_{-a_i}*\mathsf{n})\big)(\mathbb B}\def\vp{\varphi) =\mathbb P(\Delta U_i=a_i). \end{align*} It is clear that the distribution of $U_i$ is $\mathsf{n}$. Let $U_i'=U_i+\Delta U_i$. We claim that the distribution of $U_i'$ is also $\mathsf{n}$. Indeed, for any $C\in\mathscr{B}$, \begin{align*} &\mathbb P(U_i'\in C)\\ &=\mathbb P(U_i-a_i\in C, \Delta U_i=-a_i) + \mathbb P(U_i+a_i\in C, \Delta U_i=a_i) +\mathbb P(U_i \in A, \Delta U_i=0)\\ &= \frac 12\left(\delta_{-a_i}*(\mathsf{n}\wedge\mathsf{n}^{a_i})\right)(C) +\! \frac12\left(\delta_{a_i}*(\mathsf{n}\wedge\mathsf{n}^{-a_i})\right)(C) \! +\! \left(\!\!\mathsf{n}-\!\! \frac 12\,\big(\mathsf{n}\wedge\mathsf{n}^{-a_i}+\mathsf{n}\wedge\mathsf{n}^{a_i}\big)\!\!\right)(C)\\ &=\mathsf{n}(C), \end{align*} where we have used that $$ \delta_{a_i}*(\mathsf{n}\wedge\mathsf{n}^{-a_i})=\mathsf{n}\wedge\mathsf{n}^{a_i}\quad\textrm{ and }\quad\delta_{-a_i}*(\mathsf{n}\wedge\mathsf{n}^{a_i})=\mathsf{n}\wedge\mathsf{n}^{- a_i}. $$ Without loss of generality, we can assume that the pairs $(U_i,U_i')$ are independent for all $i\ge1$. Now we construct the coupling $$ (S_k,S_k')_{k\ge1} =\left(\sum_{i=1}^k T_{t-(t_1+\cdots+t_i)}\sigma U_i\big), \sum_{i=1}^kT_{t-(t_1+\cdots+t_i)}\sigma U'_i\big)\right)_{k\ge1} $$ of $$ (S_k)_{k\ge1}:=\bigg(\sum_{i=1}^kT_{t-(t_1+\cdots+t_i)}\sigma U_i\bigg)_{k\ge1}. $$ Since $U'_i-U_i=\Delta U_i$ is either $\pm a_i$ or $0$, we know that \begin{align*} (S_k'- S_k)_{k\ge1}=\left(\sum_{i=1}^kT_{t-(t_1+\cdots+t_i)}\sigma(U'_i-U_i)\big)\right)_{k\ge1}=\left(\sum_{i=1}^kT_{t-(t_1+\cdots+t_i)} \sigma\Delta U_i\big)\right)_{k\ge1} \end{align*} is a random walk on $\mathbb B}\def\vp{\varphi$ whose steps are independent and attain the values $-T_t(x-y)$, $0$ and $T_t(x-y)$ with probabilities $\frac 12(1-p_i)$, $p_i$ and $\frac 12(1-p_i)$, respectively; the values of the $p_i$ are given by \begin{align* p_i &= \left(\mathsf{n}- \tfrac 12 (\mathsf{n}\wedge\mathsf{n}^{-a_i}+\mathsf{n}\wedge\mathsf{n}^{a_i})\right)(\mathbb B}\def\vp{\varphi) = 1-\mathsf{n}\wedge\mathsf{n}^{-a_i}(\mathbb B}\def\vp{\varphi). \end{align*} Note that $\mu_{t_1,\cdots,t_k}$ is the law of the random variable $ \sum_{i=1}^kT_{t-(t_1+\cdots+t_i)}\sigma U_i. $ We get \begin{equation}\label{proofs6} \|\delta_{T_t(x-y)}*\mu_{t_1,\cdots,t_k}-\mu_{t_1,\cdots,t_k}\|_{var} \le 2\,\mathbb P(T^S>k), \end{equation} where $$ T^S=\inf\{i\ge1\::\: S_{i}=S^{\prime}_{i}+T_t(x-y)\}. $$ From \eqref{th2233} we get that for all $i\ge1$, $t_1\ge \varepsilon$, $t\ge t_1+\cdots+t_k$ and $x, y\in\mathbb B}\def\vp{\varphi$ with $\|x-y\|_{\mathbb B}\def\vp{\varphi}\le \varrho$, \begin{equation}\label{proofcon}\begin{aligned} \frac 12(1-p_i) &= \frac{1}{2}\big(\mathsf{n}\wedge\big(\delta_{-a_i}*\mathsf{n})\big)(\mathbb B}\def\vp{\varphi)\\ &\ge \frac{1}{2}\inf_{s\ge \varepsilon, \|z\|_{\mathbb B}\def\vp{\varphi}\le \varrho}\mathsf{n}\wedge (\delta_{\sigma^{-1}T_sz}*\mathsf{n})(\mathbb B}\def\vp{\varphi)\\ &=\frac{1}{2C_\eta}\,\gamma(\eta,\varrho,\varepsilon)>0. \end{aligned}\end{equation} We will now estimate $\mathbb P(T^S>k)$. Let $V_i$, $i\geq 1$, be independent symmetric random variables on $\mathbb B}\def\vp{\varphi$, whose distributions are given by $$ \mathbb P(V_i=z) = \begin{cases} \frac 12(1-p_i), &\text{if\ \ } z=-T_t(x-y);\\ \frac 12(1-p_i), &\text{if\ \ } z=T_t(x-y);\\ \qquad p_i, &\text{if\ \ } z=0. \end{cases} $$ Set $Z_k:=\sum_{i=1}^k V_i$. We have seen earlier that $$ T^S=\inf\{k\ge 1\::\: Z_k=T_t(x-y)\}. $$ For any $k\ge1$, let $$ \kappa=\kappa(k):=\#\big\{i\::\: i\le k\textrm{ and }V_i\neq 0\big\} $$ and set $\tilde{Z}_k :=\sum_{i=1}^k\tilde{V}_i$, where $\tilde{V}_i$ denotes the $i$th $V_j$ such that $V_j\neq 0$. Then, $\tilde{Z}_k$ is a symmetric random walk on $\mathbb B}\def\vp{\varphi$ with iid steps which are either $-T_t(x-y)$ or $T_t(x-y)$ with probability $1/2$. Define $$ T^{\tilde{Z}}:=\inf\{k\ge 1\::\: \tilde{Z}_k=T_t(x-y)\}. $$ By \eqref{proofcon}, \begin{equation}\label{lll1}\begin{aligned} \mathbb P(T^S>k) &=\mathbb P\left(T^S>k,\; \kappa\ge \frac{1}{2C_\eta}\,\gamma(\eta,\varrho,\varepsilon)k\right) +\mathbb P\left(T^S>k,\, \kappa\le \frac{1}{2C_\eta}\, \gamma(\eta,\varrho,\varepsilon)k\right)\\ &\le\mathbb P\left(T^{\tilde{Z}}> \frac{1}{2C_\eta}\,\gamma(\eta,\varrho,\varepsilon)k\right) +\mathbb P\bigg(\kappa\le \frac{1}{2}\sum_{i=1}^k(1-p_i)\bigg)\\ &\le\mathbb P\left(T^{\tilde{Z}}>\frac{1}{2C_\eta}\,\gamma(\eta,\varrho,\varepsilon)k\right) +\mathbb P\bigg(\Big|\kappa-\sum_{i=1}^k(1-p_i)\Big|\ge\frac{1}{2}\sum_{i=1}^k(1-p_i)\bigg). \end{aligned}\end{equation} Note that $ \kappa=\kappa(k)=\sum_{i=1}^k\zeta_i, $ where $\zeta_i = 1_{\{V_i\neq 0\}}$, ${1\le i\le k}$, are independent random variables with $\mathbb P(\zeta_i = 0) = p_i$ and $\mathbb P (\zeta_i=1)=1-p_i$. Chebyshev's inequality shows that \begin{equation}\label{lll2}\begin{aligned} \mathbb P\bigg(\Big|\kappa-\sum_{i=1}^k(1-p_i)\Big|\ge\frac{1}{2}\sum_{i=1}^k(1-p_i)\bigg) &\le \frac{4 var(\kappa)}{\Big(\sum_{i=1}^k(1-p_i)\Big)^2}\\ &=\frac{4\sum_{i=1}^kp_i(1-p_i)}{\Big(\sum_{i=1}^k(1-p_i)\Big)^2}\\ &\le \frac{4(1-C_\eta^{-1}\gamma(\eta,\varrho,\varepsilon))\sum_{i=1}^k(1-p_i)}{\Big(\sum_{i=1}^k(1-p_i)\Big)^2}\\ &\le\frac{4(1-C_\eta^{-1}\gamma(\eta,\varrho,\varepsilon))}{C_\eta^{-1}\gamma(\eta,\varrho,\varepsilon)k}. \end{aligned}\end{equation} For the second and the last inequalities we have used \eqref{proofcon}. On the other hand, by \cite[Lemma 2.3]{SW2}, \begin{align*} \mathbb P\bigg(T^{\tilde{Z}}>\frac{1}{2C_\eta}\gamma(\eta,\varrho,\varepsilon)k\bigg) &=\mathbb P\bigg(\Big\langle\max_{i\le \big[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\big]}\tilde{Z}_i, \theta^*\Big\rangle< \| T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}\bigg)\\ &=\mathbb P\bigg(\max_{i\le \big[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\big]}\|\tilde{Z}_i\|_{\mathbb B}\def\vp{\varphi}< \|T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}\bigg)\\ &\leq 2\,\mathbb P\left(0\le \Big\|\tilde{Z}_{\big[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\big]}\Big\|_{\mathbb B}\def\vp{\varphi} \leq \|T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}\right), \end{align*}where in the first equality $\theta^*$ is an element in the dual space $E^*$ of the Banach space $E$ such that the duality $\langle T_t(x,y), \theta^*\rangle=\|T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}$, and in the second equality $\|\tilde{Z}_i\|_{\mathbb B}\def\vp{\varphi}=\langle \tilde{Z}_i, \theta^*\rangle$ for $i\ge1$. From the construction above, we know that $(\|\tilde{Z}_k\|_{\mathbb B}\def\vp{\varphi})_{k\ge1}$ is a symmetric random walk on $\R$ with iid steps with values $\pm \|T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}$. Using the central limit theorem we find for sufficiently large values of $k\geq k_0$ and some constant $C_0=C_0(k_0)\ge 1$ \begin{equation}\label{lll3}\begin{aligned} &\mathbb P\left(T^{\tilde{Z}}>\frac{1}{2C_\eta}\,\gamma(\eta,\varrho,\varepsilon)k\right)\\ &\le 2\,\mathbb P\left(0\leq \frac{\Big\|\tilde{Z}_{\big[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\big]}\Big\|_{\mathbb B}\def\vp{\varphi}}{\|T_t(x-y)\|_{\mathbb B}\def\vp{\varphi}\sqrt{\big[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\big]}} \leq {\left[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\right]}^{-1/2}\right)\\ &\leq \frac{C_0}{\sqrt{2\pi}} \int_{0}^{ {\left[\frac{\gamma(\eta,\varrho,\varepsilon)k}{2C_\eta}\right]}^{-1/2}} e^{-u^2/2}\,du\\ &\leq \frac{C_0\sqrt{C_\eta}}{\sqrt{\pi \gamma(\eta,\varrho,\varepsilon)k}}. \end{aligned}\end{equation} Combining \eqref{lll1}, \eqref{lll2} and \eqref{lll3} gives for all $x,y\in\mathbb B}\def\vp{\varphi$ with $\|x-y\|_{\mathbb B}\def\vp{\varphi} \le\varrho$, $t\ge (t_1+\cdots+t_k)\vee 1$, $t_1\ge \varepsilon$ and $k\geq k_0$ that $$ \mathbb P\big(T^S>k\big) \le \frac{C_0\sqrt{C_\eta}}{\sqrt{\pi\gamma(\eta,\varrho,\varepsilon) k}}+\frac{4(1-C_\eta^{-1}\gamma(\eta,\varrho,\varepsilon))}{C_\eta^{-1}\gamma(\eta,\varrho,\varepsilon)k}. $$ According to the estimate above and \eqref{proofs6}, we can find an integer $k_0$ and a constant $C_1>0$ such that $$ \|\delta_{T_t(x-y)}*\mu_{t_1,\cdots,t_k}-\mu_{t_1,\cdots,t_k}\|_{var} \le C_1\bigg(\frac{1}{\sqrt {\gamma(\eta, \varrho,\vv)k}}+\frac{1}{\gamma(\eta, \varrho,\vv)k}\bigg),\ \ k\ge k_0,\vv\in (0,1),t\ge 1 $$ holds for all $x,y\in\mathbb B}\def\vp{\varphi$ with $\|x-y\|_{\mathbb B}\def\vp{\varphi}\le \varrho$ and $(t_1,\cdots, t_{k+1})\in I_{t,k}\cap\{(0,\infty)^{k+1}:\,t_1\ge\vv\}.$ Combining this with \eqref{proofs2} and \eqref{proofs3}, we obtain that for all $x, y\in\mathbb B}\def\vp{\varphi$ with $\|x-y\|_{\mathbb B}\def\vp{\varphi}\le \varrho$, $t\ge1$ and $\vv>0$, \begin{equation}\begin{aligned}\label{proofs8} &\| P_t(x,\cdot)-P_t(y,\cdot)\|_{var}\\ &\le 2C_\eta\vv+2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t} +2\sum_{k=1}^{k_0} \int_{I_{t,k}} C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\\ &\quad +\frac{C_1}{\sqrt {\gamma(\eta, \varrho,\vv)}}\sum_{k=1}^\infty\frac{1}{\sqrt{k}} \int_{I_{t,k}}C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\\ &\quad +\frac{C_1}{ {\gamma(\eta, \varrho,\vv)}}\sum_{k=1}^\infty\frac{1}{{k}} \int_{I_{t,k}}C_\eta^{k+1}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta(t_1+\cdots+t_{k+1})}\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_1\cdots \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t_{k+1}\\ &\le 2C_\eta\vv+2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t}\bigg(1+ C_\eta\sum_{k=1}^{k_0} \frac{C_\eta^{k}t^k}{k!}\bigg)\\ &\quad +\frac{C_1C_\eta}{\sqrt {\gamma(\eta, \varrho,\vv)}} \sum_{k=1}^\infty\frac{C_\eta^k t^k}{\sqrt{k}\,k!}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t}+\frac{C_1C_\eta}{{\gamma(\eta, \varrho,\vv)}} \sum_{k=1}^\infty\frac{C_\eta^k t^k}{{k}\,k!}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-C_\eta t}\\ &\le C_2\bigg(\vv+\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\frac{1}{2}C_\eta t}+ \frac{1}{\sqrt{\gamma(\eta, \varrho,\vv)t}}+ \frac{1}{{\gamma(\eta, \varrho,\vv)t}}\bigg) \end{aligned}\end{equation}holds for some constant $C_2>0$ depending only on $C_\eta$ and $C_1.$ To finish the proof, let \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t:= \inf_{\vv>0} \Big(\vv+ \ff 1 {\ss{\gg(\eta,\varrho,\vv)t}}\Big),\\ &\vv_t= \sup\Big\{\vv>0:\ \vv^2\gg(\eta,\varrho, \vv)\le\ff 1 t\Big\},\ \ t\ge 1.\end{split}\end{equation*} Then it is easy to see that $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t,\vv_t\downarrow 0$ as $t\uparrow\infty$ and $\vv_t\ge \ff 1 {\ss{C_\eta t}}.$ Moreover, since $\gg(\eta,\rr,\vv)$ is increasing in $\vv$, for any $\vv\in (0,\vv_t)$, $$\vv +\ff 1 {\ss{\gg(\eta,\rr,\vv)t}} \ge\lim_{\vv'\uparrow \vv_t} \ff 1 {\ss{\gg(\eta,\rr,\vv')t}}\ge \lim_{\vv'\uparrow \vv_t} \vv'=\vv_t.$$ So, $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t\ge \vv_t$. Therefore, there exists a constant $C_3>0$ such that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \inf_{\vv>0}\bigg(\vv+\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\frac{1}{2}C_\eta t}+ \frac{1}{\sqrt{\gamma(\eta, \varrho,\vv)t}}+ \frac{1}{{\gamma(\eta, \varrho,\vv)t}}\bigg) &\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t +\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ff 1 2 C_\eta t}+\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t^2 \\ &\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t+ \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ff 1 {2\vv_t^2}} +\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t^2\\ & \le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t+ \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ff 1 {2\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t^2}} +\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t^2\\ &\le C_3 \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_t,\qquad \ t\ge 1.\end{split}\end{equation*} Combining this with (\ref{proofs8}) we complete the proof.\end{proof} \section{Proof of Theorem \ref{T1.3}} Let $L^1, L^0, \LL^1,\LL^0$ be in Section 3.1. In particular, $L^0$ is a compound Poisson process with jump measure $\nu_0$. Then $L^0$ can be formulated as $$L^0_t=\sum_{i=0}^{N_t} \xi_i, \quad t>0,$$ where $N_t:= \#\{s\in [0,t]:\ \DD L^0_s\ne 0\}$, $\xi_i=\DD L^0_{\tau_i}$ for $\tau_i$ the $i$-th jump time of $L^0$. It is well-known that $N$, $\{\xi_i\}$ are independent, $N$ is the Poisson process with parameter $\ll_0$, and $\{\xi_i\}$ have common distribution $\ff 1 {\ll_0}\nu_0.$ To derive exponential convergence of $P_t$ in the total variational norm, we make use of the decomposition \begin{equation}\label{D} \begin{split} P_t f(x)&= \E \big(1_{\{N_t=0\}}f(X_t^x)\big) +P_t^1f(x),\\ P_t^1f(x)&= \E\big(1_{\{N_t\ge 1\}} f(X_t^x)\big),\quad f\in \B_b(\mathbb B}\def\vp{\varphi), t\ge 0, x\in \mathbb B}\def\vp{\varphi.\end{split}\end{equation} Since when $t\to\infty, \E \big(1_{\{N_t=0\}}f(X_t^x)\big)$ decays exponentially fast, it suffices to prove the exponential convergence of $P_t^1.$ To this end, we first consider the gradient estimate of $P_t^1.$ \begin} \def\beq{\begin{equation}} \def\F{\scr F{prp}\label{PP2} Assume {\bf (A)} and suppose that $(\ref{Z1})$ and $(\ref{Z2})$ hold. Let $$\Gamma} \def\gg{\gamma_t:= \ff 1 {1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 t}} \int_0^t\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 r}\Big(\sup_{\|z\|_\mathbb B}\def\vp{\varphi\le 1}\sup_{s\ge r}\|\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sz\|_\H\Big)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r <\infty,\ \ t>0.$$ Then $$\|\nabla} \def\pp{\partial} \def\EE{\scr E P_t^1f\|_\infty\le c \Gamma} \def\gg{\gamma_t \|f\|_\infty,\ \ t>0, f\in \B_b(\mathbb B}\def\vp{\varphi).$$\end{prp} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} The proof is modified from that of \cite[Theorem 3.1]{W10b}. It suffices to prove \beq\label{GGG1} |\nabla} \def\pp{\partial} \def\EE{\scr E_{z_0}P_t^1f(x)|\le c \Gamma} \def\gg{\gamma_t\|f\|_\infty,\ \ z_0, x\in\mathbb B}\def\vp{\varphi, \|z_0\|_\B\le 1, f\in \B_b(\mathbb B}\def\vp{\varphi).\end{equation} To prove this inequality, we first establish a formula for $P_t^1$ as in \cite[(3.8)]{W10b} where $\sigma} \def\ess{\text{\rm{ess}}=I$ is considered. Recall that for a random variable $(\xi,\tau)$ on $\mathbb B}\def\vp{\varphi\times [0,t)$ such that the distribution of $(L^0,\xi,\tau)$ is $$g(w,z,s) \LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w) \nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s,$$ \cite[Corollary 2.3]{W10a} implies that \beq\label{ABC1} \E\big\{(F1_{\{U>0\}})(L^0)\big\} = \E \ff{F1_{\{U>0\}}}{U}(L^0+ \xi 1_{[\tau,t]})\end{equation} holds for positive measurable function $F$ on $W_t$, where \beq\label{ABC2} U(w):= \sum_{s\in [0,t): \DD w_s\ne 0} g(w-\DD w_s 1_{[s,t]}, \DD w_s, s).\end{equation} Now, let $(\xi,\tau)$ be independent of $(L^1,L^0)$ with distribution $\ff 1 {t\ll_0} 1_{[0,t]}(s) \nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.$ We have $g(w,z,s)=\ff 1 {t\ll_0} 1_{[0,t]}(s)$, so that $$U(L^0+\xi 1_{[\tau,t]}) =\ff{N_t +1}{\ll_0 t} >0.$$ Therefore, letting $Y_t= \int_0^{t}T_{t-s} \sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L^1_s$ which is independent of $(L^0, \xi,\tau)$, combining (\ref{S}) with (\ref{ABC1}) we obtain \beq\label{ABC3} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split}P_t^1f(x+\vv z_0) &= \E \bigg\{f\bigg(Y_t + T_t(x+\vv z_0) +\int_0^t T_{t-s} \sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s^0\bigg) 1_{\{N_t\ge 1\}}\bigg\}\\ & =\ll_0 t\, \E\bigg\{\ff{ f\big( Y_t + T_tx +\int_0^t T_{t-s} \sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \big\{L^0+(\xi+\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_\tau z_0)1_{[\tau,t]}\big\}_s\big)}{N_t+1}\bigg\}.\end{split}\end{equation} On the other hand, it is easy to see from (\ref{1.1}) that the distribution of $(L^0, \xi+\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_\tau z_0,\tau)$ is $$\ff{\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_sz_0}(z)\rr_0(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sz_0)1_{[0,t]}}{t\ll_0\rr_0(z)}\,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\, \nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s =: g(w,z,s) \,\LL^0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D w)\, \nu_0(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.$$ According to (\ref{ABC2}) we have $\{U(L^0)>0\}= \{N_t\ge 1\}$ and $$U(L^0) = \ff 1 {\ll_0 t} \sum_{i=1}^{N_t} \ff{\varphi_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i} z_0}(\xi_i) \rr_0(\xi_i-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i} z_0)}{\rr_0(\xi_i)}.$$ So, applying (\ref{ABC1}) to $FU$ in place of $F$, we obtain $$\E\big\{(FU)(L^0)1_{\{N_t\ge 1\}}\big\} = \E\big\{(F1_{\{U>0\}})(L^0 +(\xi+\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_\tau z_0)1_{[\tau,t]}\big\}.$$ Taking $n_t(w)= \sum_{s\le t}1_{\{\nabla} \def\pp{\partial} \def\EE{\scr E w_s\ne 0\}}$ such that $N_t= n_t(L^0)$, and letting $$F(w)= \ff{f\big(Y_t+T_t x+ \int_{\mathbb B}\def\vp{\varphi\times [0,t]}T_{t-s}\sigma} \def\ess{\text{\rm{ess}} z \, w(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s)\big)}{n_t(w)} 1_{\{n_t(w)\ge 1\}},$$ we arrive at \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\ff 1 {\ll_0 t} \E \bigg\{f\bigg(Y_t +T_t x +\int_0^t T_{t-s}\sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s^0\bigg) \ff{1_{\{N_t\ge 1\}}}{N_t} \sum_{i=1}^{N_t} \ff{\varphi_{\vv\sigma} \def\ess{\text{\rm{ess}}^{-1} T_{\tau_i} z_0}(\xi_i) \rr_0(\xi_i-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i} z_0)}{\rr_0(\xi_i)}\bigg\}\\ &= \E\bigg\{\ff{ f\big(( Y_t + T_tx +\int_0^t T_{t-s} \sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \big\{L^0+(\xi+\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_\tau z_0)1_{[\tau,t]}\big\}_s\big)}{N_t+1}\bigg\}.\end{split}\end{equation*} Combining this with (\ref{ABC3}) and noting that $X_t^x=Y_t +T_t x +\int_0^t T_{t-s} \sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_s^0$ due to (\ref{S}), we obtain $$P_t^1f(x+\vv z_0)= \E \bigg\{f (X_t^x) \ff{1_{\{N_t\ge 1\}}}{N_t} \sum_{i=1}^{N_t} \ff{\varphi_{\vv\sigma} \def\ess{\text{\rm{ess}}^{-1} T_{\tau_i} z_0}(\xi_i) \rr_0(\xi_i-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i} z_0)}{\rr_0(\xi_i)}\bigg\}.$$ Therefore, \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{GGG2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\ff{|P_t^1f(x+\vv z_0)-P_t^1 f(x)|}{\vv}\\ &=\E\bigg\{f(X_t^x)1_{\{N_t\ge 1\}} \ff 1 {N_t}\sum_{i=1}^{N_t} \ff{\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(\xi_i)\rr_0(\xi_i-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0)-\rr_0(\xi_i)}{\vv\rr_0(\xi_i)}\bigg\}\\ &\le \ff{\|f\|_\infty}{\varepsilon\ll_0} \E\bigg\{1_{\{N_t\ge 1\}} \ff 1 {N_t}\sum_{i=1}^{N_t}\int_{\mathbb B}\def\vp{\varphi} \big| \vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(z)\rr_0(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0)-\rho_0(z)\big|\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\bigg\}\end{split}\end{equation} holds for any $\vv>0.$ Moreover, it follows from (\ref{1.1}) and (\ref{Z1}) that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{\mathbb B}\def\vp{\varphi} \big| \vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(z)\rr_0(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0)-\rho_0(z)\big|\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\le \int_{\mathbb B}\def\vp{\varphi} \big| \rr_0(z-\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0)-\rr_0(z)\big|\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(z)\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)+ \int_{\mathbb B}\def\vp{\varphi}\rr_0(z)\big|\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(z)-1\big|\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &= \int_{\mathbb B}\def\vp{\varphi} \big|\rr_0(z)- \rr_0(z+\vv\sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0)\big|\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)+ \int_{\mathbb B}\def\vp{\varphi}\rr_0(z)\big|\vp_{\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0}(z)-1\big|\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &\le c\|\vv \sigma} \def\ess{\text{\rm{ess}}^{-1}T_{\tau_i}z_0\|_\H\le c\vv \sup_{\|z\|_\B\le 1}\sup_{s\ge \tau_1}\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\tau_sz\|_\H \end{split}\end{equation*} holds for small enough $\vv>0$ and some constant $c>0$. Combining this with (\ref{GGG2}) and using the fact that the conditional distribution of $\tau_1$ under $N_t\ge 1$ is $\ff {\ll_0\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0s}1_{[0,t]}}{1-\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0t}} \, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$, we obtain $$\ff{|P_t^1f(x+\vv z_0)-P_t^1 f(x)|}{\vv}\le c\Gamma} \def\gg{\gamma_t\|f\|_\infty $$ for small enough $\vv>0.$ Then (\ref{GGG1}) follows by letting $\vv\to 0$. \end{proof} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}[Proof of Theorem \ref{T1.3}] By (\ref{D}) and Proposition \ref{PP2} we have \beq\label{DD2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |P_t f(x)-P_t f(y)|&\le 2\|f\|_\infty \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0 t} + |P_t^1f(x)-P_t^1f(y)|\\ &\le 2\|f\|_\infty\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0t} + c \Gamma} \def\gg{\gamma_t \|f\|_\infty\|x-y\|_\mathbb B}\def\vp{\varphi.\end{split}\end{equation} Since $\|T_s\|_\mathbb B}\def\vp{\varphi\le c\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll s}$, it follows from (\ref{S}) that $$\|X_t^x-X_t^y\|_\mathbb B}\def\vp{\varphi\le c\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll t}\|x-y\|_\mathbb B}\def\vp{\varphi,\ \ x,y\in \mathbb B}\def\vp{\varphi, t\ge 0.$$ Combining this with (\ref{DD2}) and using the Markov property, we arrive at \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & |P_t f(x)-P_t f(y)|\\ &\le \E |P_s f(X_{t-s}^x)-P_sf(X_{t-s}^y)|\\ &\le 2\|f\|_\infty\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0s} + c \Gamma} \def\gg{\gamma_s \|f\|_\infty\|X_{t-s}^x-X_{t-s}^y\|_\mathbb B}\def\vp{\varphi \\ &\le c_1\|f\|_\infty(1+\|x-y\|_\mathbb B}\def\vp{\varphi) \big\{\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll_0s} \lor (\Gamma} \def\gg{\gamma_s \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\ll (t-s)})\big\},\ \ \ s\in (0,t) \end{split}\end{equation*} for some constant $c_1>0$. Taking $s= \ff{\ll t}{\ll_0+\ll}$ and using (\ref{Z2}), we prove the desired estimate for $t\ge \ff{\ll_0+\ll}\ll.$ The proof is then finished since the inequality trivially holds for some constant $C>0$ for $t\le \ff{\ll_0+\ll}\ll.$ \end{proof} \section{Two specific models} In the following two examples we take the reference measure $\mu$ to be the Wiener measure on the Brownian path space, and the Gaussian measure on a separable Hilbert space, respectively. \subsection{Wiener measure} Let $\mathbb B}\def\vp{\varphi=\{x\in C([0,1];\R^d):\ x_0=0\}$, and let $\mu$ be the Wiener measure on $\mathbb B}\def\vp{\varphi$, i.e. the distribution of the $d$-dimensional Brownian motion $(B_s)_{s\in [0,1]}$. Let $\H=\{h\in \mathbb B}\def\vp{\varphi: \int_0^1|\dot h_s|^2\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s<\infty\}$ be the Cameron-Martin space. Then $(\mathbb B}\def\vp{\varphi, \H,\mu)$ is known as the Wiener space (see \cite[Chapter 1]{Ma}). By the Cameron-Martin theorem (or the Girsanov theorem), (\ref{1.1}) holds for \beq\label{1.2}\vp_h(z)= \exp\bigg[\int_0^1\<\dot h_s, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_s\>-\ff 1 2 \int_0^1| \dot h_s|^2\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg],\end{equation} where $ \int_0^1\<\dot h_s, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_s\>$ is the It\^o stochastic integral w.r.t. $(z_s)_{s\in [0,1]}$, which is the Brownian motion under $\mu$. Let $(\mathbb B}\def\vp{\varphi,\H,\mu)$ be the Wiener space specified above, and let $\DD$ be the Laplace operator on $[0,1]$ with Dirichlet boundary condition at $0$, and with either Dirichlet or Neumann boundary condition at $1$. We call $\DD$ the Dirichlet or the Dirichlet-Neumann Laplacian on $[0,1].$ Let $P_t$ be the semigroup associated with the SDE $$\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= \DD X_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t+\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_t,$$ where $L_t$ is a L\'evy process on $\mathbb B}\def\vp{\varphi$ with L\'evy measure $\nu$, and let $\nu_0$ satisfy $(\ref{C})$. \begin} \def\beq{\begin{equation}} \def\F{\scr F{prp}\label{P4.1} $(1)$ If $\nu_0(\mathbb B}\def\vp{\varphi)=\infty$, then $P_t$ is strong Feller for any $t>0.$ $(2)$ If $\nu_0(\mathbb B}\def\vp{\varphi)<\infty$ and there exist $z_0\in\mathbb B}\def\vp{\varphi$ and $r_0>0$ such that $\inf_{B(z_0,r_0)}\rr_0>0$, then $$\|P_t(x,\cdot)-P_t(y,\cdot)\|_{var}\le \ff{C(1+\|x-y\|_\mathbb B}\def\vp{\varphi)}{\log (1+t)},\ \ \ t>0, x,y\in\mathbb B}\def\vp{\varphi$$ holds for some constant $C>0.$ $(3)$ If $\rr_0$ is Lipschitz continuous and $\ll_0:=\nu_0(\mathbb B}\def\vp{\varphi)\in (0,\infty)$, then $(\ref{Z3})$ holds for $\ll>0$ the first eigenvalue of $\DD$ on $[0,1]$ under the underlying boundary condition. \end{prp} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} By the gradient estimate for the (Dirichlet or Dirichlet-Neumann) heat semigroup $T_s$ on the interval $[0,1]$ (see e.g.\ \cite[Section 2.4]{WangBook} and the references therein), there exists a constant $c_1>0$ such that $$\Big|\ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D}{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r}(T_s y)(r) \Big|\le \ff{c_1\|y\|_\mathbb B}\def\vp{\varphi}{\ss s},\ \ \ s>0, r\in [0,1], y\in \mathbb B}\def\vp{\varphi.$$ Then \beq\label{GFY} \|T_s y\|_\H\le \ff{c_1\|y\|_\mathbb B}\def\vp{\varphi}{\ss s},\ \ \ s>0, y\in\mathbb B}\def\vp{\varphi.\end{equation} Therefore, {\bf (A)} holds for $\sigma} \def\ess{\text{\rm{ess}}= I$. By (\ref{1.2}) and (\ref{GFY}), for $\mu$-a.e. Bownian path $z$, we have $$\sup_{\|y\|_\mathbb B}\def\vp{\varphi\le 1} \vp_{T_s y}(z+T_s y)\le \sup_{\|h\|_\H\le c_1s^{-1/2}} \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ \int_0^1\<\dot h_u, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_u \> -\ff {1}{2}\int_0^1 |\dot h_u|^2\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D u } <\infty.$$ Thus, (\ref{1.5'}) holds and Theorem \ref{T1.1} implies the first assertion. Next, noting that $$\int_\mathbb B}\def\vp{\varphi \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ 2\int_0^1\<\dot h_r, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_r\>-2\int_0^1|\dot h_r|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r}\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)=1,\ \ \ h\in\H,$$ we obtain \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \int_\mathbb B}\def\vp{\varphi \vp_{T_s y}(z)^2\,\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) &= \int_\mathbb B}\def\vp{\varphi \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{2\<\ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D }{d r} (T_s y)_r, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_r\> -\int_0^1 |\ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D}{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r} (T_s y)_r|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r}\, \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)\\ &= \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ \int_0^1 |\ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D}{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r} (T_s y)_r|^2\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r}\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ c_1^2\|y\|_\mathbb B}\def\vp{\varphi^2/s},\ \ \ s>0, y\in \mathbb B}\def\vp{\varphi.\end{split}\end{equation*} This implies that $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_2(\vv)\le c_2\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ c_1^2/\vv}$ for some constant $c_2>0$ and all $\vv\in (0,1)$. Thus, the second assertion follows from Theorem \ref{T1.2}. Finally, to prove (3) it suffices to verify (\ref{Z1}) and (\ref{Z2}) in Theorem \ref{T1.3}. Since (\ref{Z2}) follows from (\ref{GFY}), we only have to prove (\ref{Z1}). By the Lipschitz continuity of $\rr_0$, there exist constants $c_3, c_4>0$ such that $$|\rr_0(z)-\rr_0(z+h)|\le c_3\|h\|_\mathbb B}\def\vp{\varphi\le c_3\|h\|_\H$$ and $$\mu(\rr_0^2)\le c_4\E\sup_{s\in [0,1]}\big(1+| B_s|^2\big)<\infty.$$ Moreover, $$\mu\big((\vp_h-1)^2\big)=\mu(\vp_h^2)-1 =\E\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{2\int_0^1\<\dot h_s,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\>-\|h\|_\H^2}-1=\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\|h\|_\H^2}-1\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega \|h\|_\H^2$$ holds for $\|h\|_\H\le 1.$ Then (\ref{Z1}) holds for some constant $c>0.$ \end{proof} \subsection{Gaussian measure} Let $\mathbb B}\def\vp{\varphi$ be a separable Hilbert space with ONB $\{e_k\}_{k\ge 1}$, and $\mu$ the Gaussian measure with trace class covariance operator $Q$ such that $Q e_k=q_k^{-1} e_k, q_k>0$ and $\sum_{k=1}^\infty q_k^{-1}<\infty$ (see \cite[Chapter 2]{DZ}). Coordinating $z\in \mathbb B}\def\vp{\varphi$ by $(z_k=\<z,e_k\>)_{k\ge 1}$, we have \beq\label{EE0} \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)=\prod_{k=1}^\infty \mu_k(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_k),\ \ \mu_k(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_k)=\ff{\ss{q_k}}{\ss{2\pi}}\exp\Big[-\ff{q_kz_k^2}{2}\Big]\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_k,\ k\ge 1.\end{equation} Next, let $A$ be the self-adjoint operator on $\mathbb B}\def\vp{\varphi$ with $A e_k=-\ll_k e_k$, $\ll_k\ge 0$ for $k\ge 1$ and \beq\label{EE} \bb(\vv):=\sup_{k\ge 1} \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-\vv \ll_k}q_k^2<\infty,\ \ \vv>0.\end{equation} Let $L_t$ be a L\'evy process on $\mathbb B}\def\vp{\varphi$ with L\'evy measure $\nu$ satisfying (\ref{C}). Let $P_t$ be the Markov semigroup associated to the linear SDE $$\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= AX_t \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t + \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D L_t.$$ \begin} \def\beq{\begin{equation}} \def\F{\scr F{prp} $(1)$ If $\nu_0(\mathbb B}\def\vp{\varphi)=\infty$, then $P_t$ is strong Feller for $t>0$. $(2)$ If $\nu_0(\mathbb B}\def\vp{\varphi)<\infty$ and there exist $z_0\in\mathbb B}\def\vp{\varphi$ and $r_0>0$ such that $c_0:=\inf_{B(z_0,r_0)}\rr_0>0$, then $(\ref{Coupling1})$ holds for $$ \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_2(\vv)= \ff 1 {c_0}\bigg[1+\exp\Big(\sup_{k\ge 1} q_k \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-2\vv \ll_k}\Big)\bigg]<\infty,\ \ \vv>0.$$ If, in particular, $q_k\approx k^{(1+\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)}$ and $\ll_k\approx k^{2/d}$ for some constants $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho, d>0$ and large $k$, then there exists a constant $C>0$ such that $$\|P_t(x,\cdot)-P_t(y,\cdot)\|_{var}\le \ff{C(1+\|x-y\|_\mathbb B}\def\vp{\varphi)}{t^{2/(4+d(1+\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho))}},\ \ \ t>0, x,y\in\mathbb B}\def\vp{\varphi.$$ $(3)$ Suppose $\ll:=\inf_{k\ge 1}\ll_k>0$. Then $(\ref{Z3})$ holds for any Lipschitz continuous $\rr_0$ with $\ll_0:=\nu_0(\mathbb B}\def\vp{\varphi)\in (0,\infty).$ \end{prp} \begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Let $\H=\{h\in \mathbb B}\def\vp{\varphi: \sum_{k=1}^\infty h_k^2q_k^2<\infty\}.$ By (\ref{EE}) it is easy to check that {\bf (A)} holds for $\sigma} \def\ess{\text{\rm{ess}}=I$. Moreover, by (\ref{EE0}), for any $h\in \H$ we have $\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z-h)=\varphi_h(z)\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z)$ for \beq\label{EE1} \varphi_h(z)= \exp\bigg[ \sum_{k=1}^\infty \Big(q_k h_k z_k-\ff 1 2 q_k h_k^2\Big)\bigg],\ \ h_k=\<h,e_k\>, k\ge 1.\end{equation} Then it is easy to see from (\ref{EE}) that there exists a constant $c_1>0$ such that $$\sup_{\|y\|\le 1} \varphi_{T_s y}(z+T_s y)\le \exp\big[\|z\|_\mathbb B}\def\vp{\varphi^2 + c_1\bb(2s)\big]<\infty,\ \ z\in \mathbb B}\def\vp{\varphi.$$ Therefore, the first assertion follows from Theorem \ref{T1.1}. Next, since $\sigma} \def\ess{\text{\rm{ess}}=I$, it follows from (\ref{EE1}) that \begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{WWY2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \int_\mathbb B}\def\vp{\varphi \varphi_{\sigma} \def\ess{\text{\rm{ess}}^{-1}T_sy}(z)^2\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z) &= \prod_{k=1}^\infty \ff{\ss{q_k}}{\ss{2\pi}}\int_\R \exp\Big[q_k(T_sy)_k^2- \ff 12 q_k(z_k-2(T_sy)_k)^2\Big]\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D z_k\\ &= \exp\bigg[\sum_{k=1}^\infty q_k (T_s y)_k^2\bigg]= \exp\bigg[\sum_{k=1}^\infty q_k \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-2\ll_k s} y_k^2\bigg]\\ &\le \exp\Big[\|y\|_\mathbb B}\def\vp{\varphi^2 \sup_{k\ge 1} q_k \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-2\ll_k s}\Big].\end{split}\end{equation} Thus, due to (\ref{EE}), Theorem \ref{T1.2} holds for the claimed $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_2(\vv)$. Finally, under \eqref{EE}, we have $\sup_{k\ge 1}q_k \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-s\ll_k}<\infty$ for $s>0$, which implies (\ref{Z2}). Moreover, replacing $T_s y$ by $h$ in (\ref{WWY2}) we obtain $$\mu(\vp_h^2)-1 =\exp\Big[\sum_{k\ge 1} q_k h_k^2\Big] -1\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{c_2\|h\|_\H^2}-1\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{c_2} \|h\|_\H^2,\ \ \|h\|_\H\le 1$$ for some constant $c_2>0.$ Then as in the proof of Proposition \ref{P4.1} we prove (\ref{Z1}). \end{proof} \paragraph{Acknowledgement.} The authors would like to thank the referee as well as an anonymous expert for very useful comments and corrections. \begin} \def\beq{\begin{equation}} \def\F{\scr F{thebibliography}{99} \bibitem{App0} D. Applebaum, \emph{Martingale-valued measures, Ornstein-Uhlenbeck processes with jumps and operator self-decomposability in Hilbert space,} S\'{e}minaire de Probabilit\'{e}s 39, 171--196. Lect. Notes in Math. 1874, Springer, Berlin, 2006. \bibitem{App1} D. Applebaum, \emph{L\'{e}vy processes and stochastic integrals in Banach spaces,} Probab. Math. Stat. 27(2007), 75--88. \bibitem{BSW} B. B\"ottcher, R. L. Schilling, J. Wang, \emph{Constructions of coupling processes for L\'evy processes,} Stoch. Proc. Appl. 121(2011), 1201--1216. \bibitem{CM} A. Chojnowska-Michalik, \emph{On processes of Ornstein-Uhlenbeck type in Hilbert space,} Stochastics 21(1987), 251--286. \bibitem{CG} M. Cranston, A. Greven, \emph{Coupling and harmonic functions in the case of continuous time Markov processes,} Stoch. Proc. Appl. 60(1995), 261--286. \bibitem{DZ} G. Da Prato, J. Zabczyk, \emph{Stochastic Equations in Infinite Dimensions,} Cambridge University Press, Cambridge, 1992. \bibitem{BN} B. Goldys, J. M. A. M. van Neerven, \emph{Transition semigroups of Banach space-valued Ornstein-Uhlenbeck processes,} Acta Appl. Math. 76(2003), 283--330. \bibitem{KS} V. Knopova, R.L. Schilling, \emph{A note on the existence of transition probability densities for L\'evy processes,} Forum Math. 25(2013), 125--149. \bibitem{KS2} V. Knopova, R.L. Schilling, \emph{Transition density estimates for a class of L\'evy and L\'evy-type processes,} J. Theor. Probab. 25(2012), 144--170. \bibitem{Lin} T. Lindvall, \emph{Lectures on the Coupling Methods,} Wiley, New York, 1992. \bibitem{Ma} P. Malliavin, \emph{Stochastic Analysis,} Springer, Berlin, 1997. \bibitem{PZ1} S. Peszat, J. Zabczyk, \emph{Stochastic Partial Equations with L\'{e}vy Noise,} Cambridge University Press, Cambridge, 2007. \bibitem{X2} E. Priola, A. Shirikyan, L. Xu, J. Zabczyk, \emph{Exponential ergodicity and regularity for equations with L\'evy noise,} Stoch. Proc. Appl. 122(2012), 106--133. \bibitem{PZ} E. Priola, J. Zabczyk, \emph{Densities for Ornstein-Uhlenbeck processes with jumps,} Bull. Lond. Math. Soc. 41(2009), 41--50. \bibitem{X1} E. Priola, J. Zabczyk, \emph{Structural properties of semilinear SPDEs driven by cylindrical stable processes,} Probab. Theory Relat. Fields 149(2011), 97--137. \bibitem{RW03} M. R\"ockner, F.-Y. Wang, \emph{Harnack and functional inequalities for generalized Mehler semigroups,} J. Funct. Anal. 203(2003), 237--261. \bibitem{SW} R. L. Schilling, J. Wang, \emph{On the coupling property of L\'evy processes,} Inst. Henri Poinc. Probab. Stat. 47(2011), 1147--1159. \bibitem{SW2} R. L. Schilling, J. Wang, \emph{On the coupling property and the Liouville theorem for Ornstein-Uhlenbeck processes,} J. Evol. Equat. 12(2012), 119--140. \bibitem{SSW} R. L. Schilling, P. Sztonyk, J. Wang, \emph{Coupling property and gradient estimates of L\'evy processes via the symbol,} Bernoulli. 18(2012), 1128--1149. \bibitem{T} A. Takeuchi, \emph{The Bismut-Elworthy-Li type formulae for stochastic differential equations with jumps, } J. Theor. Probab. 23(2010), 576--604. \bibitem{W10b} F.-Y. Wang, \emph{Gradient estimate for Ornstein-Uhlenbeck jump processes,} Stoch. Proc. Appl. 121(2011), 466--478. \bibitem{W10a} F.-Y. Wang, \emph{Coupling for Ornstein-Uhlenbeck processes with jumps,} Bernoulli 17(2011), 1136--1158. \bibitem{W11} F.-Y. Wang, \emph{Derivative formula and Harnack inequality for jump processes,} preprint, arXiv: 1104.5531v4. \bibitem{WangBook} F.-Y. Wang, \emph{Functional Inequalities, Markov Processes and Spectral Theory}, Science Press, Beijing 2005. \bibitem{WJ} J. Wang, \emph{Linear evolution equations with cylindrical L\'{e}vy noise: gradient estimates and exponential ergodicity,} preprint. \bibitem{Z} J. Zabczyk, \emph{Linear stochastic systems in Hilbert spaces; spectral properties and limit behaviour,} Banach Center Pub. 41(1985), 591--609. \bibitem{Zhang} X. Zhang, \emph{Derivative formula and gradient estimate for SDEs driven by $\aa$-stable processes,} preprint, arXiv:1204.2630. \end{thebibliography} \end{document}
\section{Introduction} Let $(\Omega,\mathcal{F},\mathbb{P})$ be a probability space. Let $G=(V,E,\sim)$ be a nonoriented connected locally finite graph without loops. Let $(a_{e})_{e\in E}$ be a sequence of positive initial weights associated to each edge $e\in E$. Let $(X_n)_{n\in{{\Bbb N}}}$ be a random process that takes values in $V$, and let $\mathcal{F}_n=\sigma(X_0,\ldots,X_n)$ be the filtration of its past. For any $e\in E$, $n\in{{\Bbb N}}\cup\{\infty\}$, let \begin{equation} Z_n(e)=a_e+ \sum_{k=1}^n\hbox{1 \hskip -7pt I}_{\{\{X_{k-1},X_k\}=e\}} \end{equation} be the number of crosses of $e$ up to time $n$ plus the initial weight $a_e$. Then $(X_n)_{n\in{{\Bbb N}}}$ is called Edge Reinforced Random Walk (ERRW) with starting point $i_0\in V$ and weights $(a_e)_{e\in E}$, if $X_0=i_0$ and, for all $n\in{{\Bbb N}}$, \begin{equation} \label{def-vsirw} \mathbb{P}(X_{n+1}=j~|~\mathcal{F}_n)={{\mathbbm 1}}_{\{j\sim X_n\}}\frac{Z_n(\{X_n,j\})} {\sum_{k\sim X_n} Z_n(\{X_n,k\})}. \end{equation} The Edge Reinforced Random Walk was introduced in 1986 by Diaconis \cite{coppersmith}; on finite graphs it is a mixture of reversible Markov chains, and the mixing measure can be determined explicitly (the so-called Coppersmith-Diaconis measure, or "magic formula" \cite{diaconis2}, see also \cite{keane-rolles1,rolles1}), which has applications in Bayesian statistics \cite{diaconis-rolles,bacallado1,bacallado2}. On infinite graphs, the research has focused so far on recurrence/transience criteria. On acyclic or directed graphs, the walk can be seen as a random walk in an \textit{independent } random environment \cite{pemantle4}, and a recurrence/transience phase transition was first observed by Pemantle on trees \cite{collevecchio1,keane-rolles2,pemantle4}. In the case of infinite graphs with cycles, recurrence criteria and asymptotic estimates were obtained by Merkl and Rolles on graphs of the form ${{\Bbb Z}}\times G$, $G$ finite graph, and on a certain two-dimensional graph \cite{merkl-rolles1,merkl-rolles4,merkl-rolles2,rolles2}, but recurrence on ${{\Bbb Z}}^2$ was still unresolved. Also, this original ERRW model \cite{coppersmith} has triggered a number of similar models of self-organization and learning behaviour; see for instance Davis \cite{davis}, Limic and Tarr\`es \cite{limic,limic-tarres}, Pemantle \cite{Pem07}, Sabot \cite{sabot1,sabot2}, Tarr\`es \cite{tarres2,tarres3} and T\'oth \cite{toth3}, with different perspectives on the topic. Our first result relates the ERRW to the Vertex-Reinforced Jump Process (VRJP), conceived by Werner and studied by Davis and Volkov \cite{dv1,dv2}, Collevechio \cite{collevecchio2,collevecchio3} and Basdevant and Singh \cite{bs}. We call VRJP with weights $(W_e)_{e\in E}$ a continuous-time process $(Y_t)_{t\ge0}$ on $V$, starting at time $0$ at some vertex $i_0\in V$ and such that, if $Y$ is at a vertex $i\in V$ at time $t$, then, conditionally on $(Y_s, s\le t)$, the process jumps to a neighbour $j$ of $i$ at rate $W_{\{i,j\}}L_j(t)$, where $$L_j(t):=1+\int_0^t \hbox{1 \hskip -7pt I}_{\{Y_s=j\}}\,ds.$$ The main results of the paper are the following. In Section 2, Theorem \ref{annealed}, we represent the ERRW in terms of a VRJP with independent gamma conductances. Section 3 is dedicated to showing, in Theorem \ref{meas}, that the VRJP is a mixture of time-changed Markov jump processes, with a computation of the mixing law. In Section 6, we interpret that mixing law with the supersymmetric hyperbolic sigma model introduced by Disertori, Spencer and Zirnbauer in \cite{dsz} and related to the Anderson model. We prove positive recurrence of VRJP and ERRW in any dimension for large reinforcement in Corollaries \ref{rec} and \ref{rec2}, using a localization result of Disertori and Spencer \cite{ds}, and transience of VRJP in dimension $d\ge3$ at small reinforcement in Corollary \ref{trans} using a delocalization result of Disertori, Spencer, Zirnbauer \cite{dsz}. Shortly after this paper appeared electronically, Angel, Crawford and Kozma \cite{ack} proposed another proof of recurrence of ERRW under similar assumptions, without making the link with statistical physics. Equivalent results were also obtained for the VRJP, using the existence of a limiting environment proved in this paper. \section{From ERRW to VRJP.} It is convenient here to consider a time changed version of $(Y_s)_{s\ge0}$: consider the positive continuous additive functional of $(Y_s)_{s\ge0}$ $$ A(s)= \int_0^s {1\over L_{Y_u}(u)} du = \sum_{x\in V} \log(L_x(s)) , $$ and the time changed process $$ X_t= Y_{A^{-1}(t)}. $$ Let $(T_i(t))_{i\in V}$ be the local time of the process $(X_t)_{t\ge0}$ $$ T_x(t)=\int_0^t {{\mathbbm 1}}_{\{X_u=x\}} du. $$ \begin{lemma} \label{invtime} The inverse functional $A^{-1}$ is given by $$ A^{-1}(t)= \int_0^t e^{T_{X_u}(u)} du=\sum_{i\in V}(e^{T_i(t)}-1). $$ The law of the process $X_t$ is described by the following: conditioned on the past at time $t$, if the process $X_t$ is at the position $i$, then it jumps to a neighbor $j$ of $i$ at rate $$ W_{i,j} e^{T_i(t)+T_j(t)}. $$ \end{lemma} \begin{proof} First note that \begin{equation} \label{tlx} T_x(A(s))= \log(L_x(s)), \end{equation} since $$ (T_x(A(s)))'= A'(s){{\mathbbm 1}}_{\{X_{A(s)}=x\}}={1\over L_{Y_s}(s)} {{\mathbbm 1}}_{\{Y_s=x\}}. $$ Hence, \begin{eqnarray*} (A^{-1}(t))'&=&{1\over A'(A^{-1}(t))} = L_{X_t} (A^{-1}(t)) = e^{T_{X_t}(t)}, \end{eqnarray*} which yields the expression for $A^{-1}$. It remains to prove the last assertion: \begin{eqnarray*} {{\Bbb P}}(X_{t+dt}=j | {{\mathcal F}}_t) &=& {{\Bbb P}}(Y_{A^{-1}(t+dt)}=j | {{\mathcal F}}_t) \\ &=& W_{X_t,j}(A^{-1})'(t) L_{j}(A^{-1}(t)) dt \\ &=& W_{i,j} e^{T_{X_t}(t)}e^{T_j(t)} dt \end{eqnarray*} \end{proof} In order to relate ERRW to VRJP, let us first define the following process $(\tilde{X}_t)_{t\in{{\Bbb R}}_+}$, initially introduced by Rubin, Davis and Sellke \cite{davis,sellke2}, which we call here continuous-time ERRW with weights $(a_e)_{e\in E}$ and starting at $\tilde{X}_0:=i_0$ at time $0$. \begin{itemize} \item Define on each edge $e\in E$ independent point processes (alarm times) as follows. Let $(\tau^e_k)_{e\in E,k\in{{\Bbb Z}}_+}$ be independent exponential random variables with parameter 1 and define $$ V^e _k =\sum_{l=0}^{k-1}{1\over a_e+l} \tau_l^e,\;\;\; \forall k\in{{\Bbb N}}.$$ \item Each edge $e\in E$ has its own clock, denoted by $\tilde T_e(t)$, which only runs when the process $(\tilde{X}_t)_{t\ge0}$ is adjacent to $e$. This means that if $e=\{i,j\}$, then $\tilde T_{\{i,j\}}(t)= \tilde T_i(t)+\tilde T_j(t)$, where $\tilde T_i(t)$ is the local time of the process $\tilde X$ at vertex $i$ and time $t$. \item When the clock of an edge $e\in E$ rings, i.e. when $\tilde T_e(t)= V^e_k$ for some $k>0$, then $\tilde{X}_t$ crosses it instantaneously (of course, this can happen only when $\tilde X$ is adjacent to $e$). \end{itemize} \psset{xunit= 1.05cm,yunit= 1cm} \pspicture(0,-2.5)(15,2) \psset{linewidth=.5pt} \psline{|->}(0,0)(8.5,0) \psline{|->}(0,1)(8.5,1) \psline{|->}(0,-1)(8.5,-1) \psline{|->}(0,-2)(8.5,-2) \pscircle*(1.1,0) {.05} \pscircle*(2.8,0) {.05} \pscircle*(4.2,0){.05} \pscircle*(5,0){.05} \pscircle*(5.2,0){.05} \pscircle*(5.31,0){.05} \pscircle*(0.9,1){.05} \pscircle*(2.3,1){.05} \pscircle*(3.1,1){.05} \pscircle*(4,1){.05} \pscircle*(4.15,1){.05} \pscircle*(1.3,-1) {.05} \pscircle*(2.5,-1) {.05} \pscircle*(4.6,-1){.05} \pscircle*(5.1,-1){.05} \pscircle*(5.9,-1){.05} \pscircle*(6.31,-1){.05} \pscircle*(0.5,-2){.05} \pscircle*(1.9,-2){.05} \pscircle*(3.4,-2){.05} \pscircle*(6,-2){.05} \pscircle*(7.15,-2){.05} \rput[u](1,0.6){$V_1^{e_0}$} \rput[u](2.3,0.6){$V_2^{e_0}$} \rput[u](1.3,-.4){$V_1^{e_1}$} \rput[u](3,-.4){$V_2^{e_1}$} \rput[u](1,-1.4){$V_1^{e_2}$} \rput[u](2.5,-1.4){$V_2^{e_2}$} \rput[u](0.7,-2.4){$V_1^{e_3}$} \rput[u](2,-2.4){$V_2^{e_3}$} \rput[b](4.5,1){$\cdots$} \rput[b](6,0){$\cdots$} \rput[u](9.5,1.5){time-line of} \rput[l](9.35,1){$e_0$} \rput[l](9.35,0){$e_1$} \rput[l](9.35,-1){$e_2$} \rput[l](9.35,-2){$e_3$} \endpspicture \medskip Let $\tau_n$ be the $n$-th jump time of $(\tilde{X}_t)_{t\ge0}$, with the convention that $\tau_0:=0$. \begin{lemma} \label{icd} (Davis \cite{davis}, Sellke \cite{sellke2}) Let $(X_n)_{n\in{{\Bbb N}}}$ (resp. $(\tilde{X}_{t})_{t\ge0}$) be an ERRW (resp. continuous-time ERRW) with weights $(a_e)_{e\in E}$, starting at some vertex $i_0\in V$. Then $(\tilde{X}_{\tau_n})_{n\ge0}$ and $(X_n)_{n\ge0}$ have the same distribution. \end{lemma} \begin{proof} The argument is based on the memoryless property of exponentials, and on the observation that, if $A$ and $B$ are two independent random variables of parameters $a$ and $b$, then $\mathbb{P}[A<B]=a/(a+b)$. \end{proof} \begin{theorem} \label{annealed} Let $(\tilde{X}_{t})_{t\ge0}$ be a continuous-time ERRW with weights $(a_e)_{e\in E}$. Then there exists a sequence of independent random variables $W_e\sim\it{Gamma}(a_e,1)$, $e\in E$, such that, conditionally on $(W_e)_{e\in E}$, $(\tilde X_{t})_{t\ge0}$ has the same law as the time modification $(X_t)_{t\ge 0}$ of the VRJP with weights $(W_e)_{e\in E}$. In particular, the ERRW $(X_n)_{n\ge 0}$ is equal in law to the discrete time process associated with a VRJP in random independent conductances $W_e\sim \it{Gamma}(a_e,1)$. \end{theorem} \begin{proof} For any $e\in E$, define the simple birth process $\{N_t^{e},t\ge0\}$ with initial population size $a_e$, by $$N_t^{e}:=a_e+\sup\left\{k\in{{\Bbb N}}\text{ s.t. }V_k^{e}\le t\right\}.$$ This process is sometimes called the Yule process: by a result of D. Kendall \cite{kendall}, there exists $W_e:=\lim N_t^{e}e^{-t}$, with distribution $Gamma(a_e,1)$, such that, conditionally on $W_e$, $\{N^{e}_{f_{W_e}(t)}, t\ge0\}$ is a Poisson point process with unit parameter, where $$f_W(t):=\log(1+t/W).$$ Let us now condition on $(W_e)_{e\in E}$: $N^{e}$ increases between times $t$ and $t+dt$ with probability $W_ee^t\,dt=(f^{-1}_{W_e})'(t)\,dt$. A similar characterization of the timelines is also used in \cite{tarres3}, Lemma 4.7. If $\tilde{X}$ is at vertex $x$ at time $t$, it jumps to a neighbour $y$ of $x$ at rate $W_{x,y}e^{T_x(t)+T_y(t)}$. \end{proof} \section{The mixing measure of VRJP.} Next we study VRJP. Given fixed weights $(W_e)_{e\in E}$, we denote by $(Y_t)_{t\ge 0}$ the VRJP and $(X_t)_{t\ge 0}$ its time modification defined in the previous Section, starting at site $X_0:=i_0$ at time $0$ and $(T_i(t))_{i\in V}$ its local time. It is clear from the definition that the joint process $\Theta_t=(X_t, (T_i(t))_{i\in V})$ is a time continuous Markov process on the state space $V\times {{\Bbb R}}_+^V$ with generator $\tilde L$ defined on $C^\infty$ bounded functions by $$ \tilde L (f)(i,T)= \left ({\partial \over \partial T_i}f\right)(i,T) + L(T)(f(.,T))(i), \;\;\; \forall (x,T)\in V\times {{\Bbb R}}_+^V, $$ where $L(T)$ is the generator of the jump process on $V$ at frozen $T$ defined for $g\in {{\Bbb R}}^V$: $$ L(T)(g)(i)=\sum_{j\in V} W_{i,j} e^{T_i+T_j} (g(j)-g(i)), \;\;\; \forall i\in V. $$ We denote by ${{\Bbb P}}_{i_0,T}$ the law of the Markov process with generator $\tilde L$ starting from the initial state $(i_0,T)$. Note that the law of $(X_t, T(t)-T)$ under ${{\Bbb P}}_{i_0,T}$ is equal to the law of the process starting from $(i_0, 0)$ with conductances $$ W_{i,j}^T=W_{i,j}e^{T_i+T_j}. $$ For simplicity, we let ${{\Bbb P}}_i:={{\Bbb P}}_{i,0}$. We show, in Proposition \ref{pconv}, that for finite graphs the centred occupation times converge a.s., and calculate the limiting measure in Theorem \ref{meas} i). In Theorem \ref{meas} ii) we show that the VRJP $(Y_s)_{s\ge 0}$ (as well as $(X_t)_{t\ge 0}$) is a mixture of time-changed Markov jump processes. This limiting measure can be interpreted as a supersymmetric hyperbolic sigma model. We are grateful to a few specialists of field theory for their advice: Denis Perrot who mentioned that the limit measure of VRJP could be related to the sigma model, and Krzysztof Gawedzki who pointed out reference \cite{dsz}, which actually mentions a possible link of their model with ERRW, suggested by Kozma, Heydenreich and Sznitman, cf \cite{dsz} Section 1.5. Note that when $G$ is a tree, if the edges are for instance oriented towards the root, letting $V_e= e^{U_{{\overline e}}-U_{{\underline e}}}$, the random variables $(V_e)$ are independent and are distributed according to an inverse gaussian law. This was understood in previous works on VRJP \cite{dv1,dv2,collevecchio2,collevecchio3,bs}. Theorems \ref{annealed} and \ref{meas} enable us to retrieve, in Section \ref{BCP} the limiting measure of ERRWs, computed by Coppersmith and Diaconis in \cite{coppersmith} (see also \cite{keane-rolles1}), by integration over the random gamma conductances $(W_e)_{e\in E}$. This explains its renormalization constant, which had remained mysterious so far. \begin{proposition} \label{pconv} Suppose that $G$ is finite and set $N=\vert V \vert$. For all $i\in V$, the following limits exist ${{\Bbb P}}_{i_0}$ a.s. $$ U_i =\lim_{t\to \infty} T_i(t) - {t\over N}. $$ \end{proposition} \begin{theorem} \label{meas} Suppose that $G$ is finite and set $N= \vert V\vert$. \noindent i) Under ${{\Bbb P}}_{i_0}$, $(U_i)_{i\in V}$ has the following density distribution on ${{\mathcal H}}_0=\{(u_i), \; \sum u_i=0\}$ \begin{eqnarray}\label{density} {1\over (2\pi)^{(N-1)/2}} e^{u_{i_0}} e^{-H(W,u)} \sqrt{D(W,u)}, \end{eqnarray} where $$ H(W,u)=2 \sum_{\{i,j\}\in E} W_{i,j} \sinh^2\left({{1\over 2}(u_i-u_j)}\right) $$ and $D(W,u)$ is any diagonal minor of the $N\times N$ matrix $M(W,u)$ with coefficients $$ m_{i,j}=\left\{ \begin{array}{ll} W_{i,j} e^{u_i+u_j} & \hbox{ if $i\neq j$} \\ -\sum_{k\in V} W_{i,k} e^{u_i+u_k} &\hbox{ if $i=j$} \end{array}\right. $$ \noindent ii) Let $C$, resp. $D$, be positive continuous additive functionals of $X$, resp. $Y$: $$C(t)=\sum_{i\in V}(e^{2T_i(t)}-1),\;\;\; D(s)=\sum_{i\in V} L_i^2(s)-1$$ and let $$Z_t=X_{C^{-1}(t)} \; (= Y_{ D^{-1}(t)}).$$ Then, conditionally on $(U_i)_{i\in V}$, $Z_t$ is a Markov jump process starting from $i_0$, with jump rate from $i$ to $j$ $$ {1\over 2} W_{i,j}e^{U_j-U_i}. $$ In particular, the discrete time process associated with $(Y_s)_{s\ge0}$ is a mixture of reversible Markov chains with conductances $W_{i,j} e^{U_i+U_j}$. \end{theorem} \noindent N.B.: 1) the density distribution in \eqref{density} is with respect to the Lebesgue measure on ${{\mathcal H}}_0$ which is $\prod_{i\in V\setminus\{j_0\}} du_i$ for any choice of $j_0$ in $V$. We simple write $du$ for any of the $\prod_{i\in V\setminus\{j_0\}} du_i$. \noindent 2) The diagonal minors of the matrix $M(W,u)$ are all equal since the sum on any line or column of the coefficients of the matrix are null. By the matrix-tree theorem, if we let $\mathcal{T}$ be the set of spanning trees of $(V,E,\sim)$, then $D(W,u)=\sum_{T\in\mathcal{T}}\prod_{\{i,j\}\in\mathcal{T}}W_{\{i,j\}}e^{u_i+u_j}$. \begin{remark} Remark that usually a result like ii) makes use of de Finetti's theorem: here, we provide a direct proof exploiting the explicit form of the density. In Section 5, we apply Theorem \ref{annealed} and Theorem \ref{meas} i) ii) to give a new proof of Diaconis-Coppersmith formula including its de Finetti part. \end{remark} \begin{remark} The fact that (\ref{density}) is a density is not at all obvious. Our argument is probabilistic: (\ref{density}) is the law of the random variables $(U_i)$. It can also be explained directly as a consequence of supersymmetry, see (5.1) in \cite{dsz}. The fact that the measure (\ref{density}) normalizes at 1 is a fundamental property, which plays a crucial role in the localization and delocalization results of Disertori and Spencer \cite{ds,dsz}. \end{remark} \begin{remark} \label{r3} ii) implies that the VRJP $(Y_s)$ is a mixture of Markov jump processes. More precisely, let $(U_i)_{i\in V}$ be a random variable distributed according to (\ref{density}) and, conditionally on $U$, $Z$ be the Markov jump process with jump rates from $i$ to $j$ given by ${1\over 2} W_{i,j} e^{U_j-U_i}$. Then the time changed process $(Z_{B^{-1}}(s))_{s\ge 0}$ with $$ B(t)= \sum_{i\in V} \sqrt{1+l_i^Z(t)}-1, $$ where $(l_i^Z(t))$ is the local time of $Z$ at time $t$, has the law of the VRJP $(Y_s)$ with conductances $W$. \end{remark} \section{Proof of the Proposition \ref{pconv} and Theorem \ref{meas}} \subsection{Proof of Proposition \ref{pconv}} \renewcommand{{\mathbb E}}{\varepsilon} \renewcommand{\tilde{E}}{\tilde{\varepsilon}} By a slight abuse of notation, we also use notation $L(T)$ for the $N\times N$ matrix $M(W^T,T)$ of that operator in the canonical basis. Let $\hbox{1 \hskip -7pt I}$ be the $N\times N$ matrix with coefficients equal to $1$, i.e. $\hbox{1 \hskip -7pt I}_{i,j}=1$ for all $i$, $j$ $\in V$, and let $I$ be the identity matrix. Let us define, for all $T\in {{\Bbb R}}^V$, \begin{eqnarray}\label{QT} Q(T):=-\int_0^\infty \left(e^{uL(T)}-\frac{\hbox{1 \hskip -7pt I}}{N}\right)\,du, \end{eqnarray} which exists since $e^{uL(T)}$ converges towards $\hbox{1 \hskip -7pt I}/N$ at exponential rate. Then $Q(T)$ is a solution of the Poisson equation for the Markov Chain $L(T)$, namely $$L(T)Q(T)=Q(T)L(T)=I-\frac{\hbox{1 \hskip -7pt I}}{N}.$$ Observe that $L(T)$ is symmetric, and thus $Q(T)$ as well. For all $T\in{{\Bbb R}}^V$ and $i$, $j$ $\in V$, let $E_i^T(\tau_j)$ denote the expectation of the first hitting time of site $j$ for the continuous-time process with generator $L(T)$. Then $$Q(T)_{i,j}=\frac{1}{N}E_i^T(\tau_j)+Q(T)_{j,j}$$ by the strong Markov property applied to (\ref{QT}). As a consequence, $Q(T)_{j,j}$ is nonpositive for all $j$, using $\sum_{i\in V} Q(T)_{i,j}=0$. Let us fix $l\in V$. We want to study the asymptotics of $T_l(t)-t/N$ as $t\to\infty$: \begin{align} \nonumber T_l(t)-\frac{t}{N}&=\int_0^t \left(\hbox{1 \hskip -7pt I}_{\{X_u=l\}}-\frac{1}{N}\right)\,du=\int_0^t (L(T(u))Q(T(u)))_{X_u,l}\,du\\ \nonumber &=\int_0^t {\tilde L} (Q(.)_{.,l})(X_u,T(u))\,du-\int_0^t \frac{\partial}{\partial T_{X_u}}Q(T(u))_{X_u,l}\,du\\ \label{poisson} &=Q(T(t))_{X_t,l}-Q(0)_{X_0,l}+M_l(t)-\int_0^t \frac{\partial}{\partial T_{X_u}}Q(T(u))_{X_u,l}\,du, \end{align} where $$M_l(t):=-Q(T(t))_{X_t,l}+Q(0)_{X_0,l}+\int_0^t {\tilde L} (Q(.)_{.,l})(X_u,T(u))\,du$$ is a martingale for all $l$. Recall that ${\tilde L}$ is the generator of $(X_t,T(t))$. The following lemma shows in particular the convergence of $Q(T(t))_{k,l}$ for all $k$, $l$, as $t$ goes to infinity. It is a purely determistic statement, which does not depend on the trajectory of the process $X_t$ (as long as it only performs finitely many jumps in a finite time interval), but only on the added local time in $W^T$. \begin{lemma} For all $k$, $l$ $\in V$, $Q(T(t))_{k,l}$ converges as $t$ goes to infinity, and $$\int_0^\infty\left|\frac{\partial}{\partial T_{X_u}}Q(T(u))_{X_u,l}\right|\,du<\infty.$$ \end{lemma} \begin{proof} For all $i$, $k$, $l$ $\in V$, let us compute $\frac{\partial}{\partial T_i}Q(T)_{k,l}$ : by differentiation of the Poisson equation, $$\frac{\partial}{\partial T_i}Q(T)_{k,l}=-\left(Q(T)\left(\frac{\partial}{\partial T_i}L\right)Q(T)\right)_{k,l}.$$ Now, for any real function $f$ on $V$, $$\frac{\partial}{\partial T_i}Lf(k) = \begin{cases} \sum_{j\sim i}W_{i,j}^T(f(j)-f(i))&\text{ if }k=i\\ W_{i,k}^T(f(i)-f(k))&\text{ if }k\sim i, k\not=i\\ 0&\text{ otherwise.} \end{cases} $$ Hence $$\frac{\partial}{\partial T_i}Lf(k) =\sum_{j\sim i}W_{i,j}^T(f(j)-f(i))(\hbox{1 \hskip -7pt I}_{\{i=k\}}-\hbox{1 \hskip -7pt I}_{\{j=k\}})$$ and, therefore, \begin{align} \nonumber \frac{\partial}{\partial T_i}Q(T)_{k,l}&=\sum_{j\sim i}W_{i,j}^T(Q(T)_{k,i}-Q(T)_{k,j})(Q(T)_{i,l}-Q(T)_{j,l})\\ &\label{derivq0}=\sum_{j\sim i}W_{i,j}^TQ(T)_{k,\nabla_{i,j}}Q(T)_{\nabla_{i,j},l}=\sum_{j\sim i}W_{i,j}^TQ(T)_{\nabla_{i,j},k}Q(T)_{\nabla_{i,j},l}, \end{align} where we use the notation $f(\nabla_{i,j}):=f(j)-f(i)$ in the second equality, and the fact that $Q(T)$ is symmetric in the third one. In particular, for all $l\in V$ and $t\ge0$, \begin{equation} \label{derivq} \frac{d}{dt}Q(T(t))_{l,l}=\frac{\partial}{\partial T_{X_t}}Q(T(t))_{l,l}=\sum_{j\sim X_t}W_{X_t,j}\left(Q(T(t))_{\nabla_{X_t,j},l}\right)^2. \end{equation} Now recall that $Q(T(t))_{l,l}$ is nonpositive for all $t\ge0$; therefore it must converge, and $$\int_0^\infty \sum_{j\sim X_t}W_{X_t,j}\left(Q(T(t))_{\nabla_{X_t,j},l}\right)^2\,dt=(Q(T(\infty))-Q(0))_{l,l}<\infty.$$ The convergence of $Q(T(t))_{k,l}$ now follows from Cauchy-Schwarz inequality, using \eqref{derivq0}: for all $t\ge s$, \begin{align*} \left|(Q(T(t))-Q(T(s)))_{k,l}\right| &=\int_s^t \sum_{j\sim X_u}W_{X_u,j}^TQ(T(u))_{\nabla_{X_u,j},k}Q(T(u))_{\nabla_{X_u,j},l}\,du\\ &\le\sqrt{(Q(T(t))-Q(T(s)))_{k,k}}\sqrt{(Q(T(t))-Q(T(s)))_{l,l}}; \end{align*} thus $Q(T(t))_{k,l}$ is Cauchy sequence, which converges as $t$ goes to infinity. Now, using again Cauchy-Schwarz inequality, \begin{align*} &\int_0^\infty\left|\frac{\partial}{\partial T_{X_u}}Q(T(u))_{X_u,l}\right|\,du \\ &=\int_0^\infty\left| \sum_{j\sim X_u}W_{X_u,j}^TQ(T(u))_{\nabla_{X_u,j},X_u}Q(T(u))_{\nabla_{X_u,j},l}\right| \,du\\ &\le\sqrt{\sum_{k\in V}(Q(T(\infty))-Q(T(0)))_{k,k}}\sqrt{(Q(T(\infty))-Q(T(0)))_{l,l}}, \end{align*} which enables us to conclude. \end{proof} Next, we show that $(M_l(t))_{t\ge0}$ converges, which will complete the proof: indeed, this implies that the size of the jumps in that martingale goes to $0$ a.s., and therefore, by (\ref{poisson}), that $Q(T(t))_{X_t,l}$ must converge as well, again by (\ref{poisson}). Let us compute the quadratic variation of the martingale $(M_l(t))_{t\ge0}$ at time $t$: \begin{align*} &\left(\frac{d}{d{\mathbb E}}\mathbb{E}\left((M_l(T(t+{\mathbb E}))-M_l(t))^2|\mathcal{F}_t\right)\right)_{{\mathbb E}=0}\\ &=\left(\frac{d}{d{\mathbb E}}\mathbb{E}\left((Q(T(t+{\mathbb E}))_{X_{t+{\mathbb E}},l}-Q(T(t))_{X_{t},l})^2|\mathcal{F}_t\right)\right)_{{\mathbb E}=0}\\ &=R(T(t))_{X_t,l} \end{align*} where, for all $(i,l,T)\in V\times V\times{{\Bbb R}}^V$, we let $$R(T)_{i,l}:={\tilde L}(Q^2(.)_{.,l})(i,T)-2Q(T)_{i,l}{\tilde L}(Q(.)_{.,l})(i,T);$$ here $Q^2(T)$ denotes the matrix with coefficients $(Q(T)_{i,j})^2$, rather than $Q(T)$ composed with itself. But \begin{align*} {\tilde L}(Q^2(.)_{.,l})(i,T)&=2\left(Q(T)\right)_{i,l}\left(\frac{\partial}{\partial T_{i}}Q(T)\right)_{i,l}+\left(L(T)Q^2(T)_{.,l}(i)\right)_{i,l}\\ Q(T)_{i,l}{\tilde L}(Q(.)_{.,l})(i,T)&=\left(Q(T)\right)_{i,l}\left(\frac{\partial}{\partial T_{i}}Q(T)\right)_{i,l}+Q(T)_{i,l}(L(T)Q(T)_{.,l}(i))_{i,l}, \end{align*} so that \begin{align*} R(T)_{i,l}&=L(T)(Q^2(T)_{.,l})_{i,l}-2Q(T)_{i,l}(L(T)Q(T)_{.,l})_{i,l}\\ &=\sum_{j\sim i}W_{i,j}^T\left((Q(T)_{j,l})^2-(Q(T)_{i,l})^2\right) -2Q(T)_{i,l}\sum_{j\sim i}W_{i,j}^T\left(Q(T)_{j,l}-Q(T)_{i,l}\right)\\ &=\sum_{j\sim i}W_{i,j}^T\left(Q(T)_{\nabla_{i,j},l}\right)^2=\frac{\partial}{\partial T_i}Q(T)_{l,l}, \end{align*} using \eqref{derivq0} in the last equality. Thus $$<M_l,M_l>_\infty=\int_0^\infty\frac{d}{du}Q(u)_{l,l}\,du=Q(T(\infty))_{l,l}-Q(0)_{l,l}\le-Q(0)_{l,l}<\infty.$$ Therefore $(M_l(t))_{t\ge0}$ is a martingale bounded in $L^2$, which converges a.s. \begin{remark} Once we know that $T_i(t)-t/N$ converges, then $T_i(\infty)=\infty$ for all $i\in V$, hence $Q(T(\infty))_{l,l}=0$, and the last inequality is in fact an equality, i.e. $<M_l,M_l>_\infty=-Q(0)_{l,l}$. \end{remark} \subsection{Proof of Theorem \ref{meas} i)} We consider, for $i_0\in V$, $T\in {{\Bbb R}}^V$, $\lambda \in {{\mathcal H}}_0$ \begin{eqnarray}\label{Psi} \Psi(i_0, T, \lambda)=\int_{{{\mathcal H}}_0} e^{u_{i_0}} e^{i<\lambda, u>} \phi(W^T,u) du, \end{eqnarray} where \begin{eqnarray} \label{notphi} \phi(W^T,u)= e^{-H(W^T,u)} \sqrt{ D(W^T,u)}, \end{eqnarray} and $W^T_{i,j}= W_{i,j} e^{T_i+T_j}$. We will prove that $$ {1\over \sqrt{2\pi}^{N-1}} \Psi(i_0, T, \lambda) ={{\Bbb E}}_{i_0,T}\left( e^{i<\lambda, U>}\right), $$ for all $i_0\in V$, $T\in {{\Bbb R}}^V$. \begin{lemma} The function $\Psi$ is solution of the Feynman-Kac equation $$ i\lambda_{i_0} \Psi(i_0, T , \lambda) +(\tilde L \Psi)(i_0,T,\lambda)=0. $$ \end{lemma} \begin{proof} Let $\overline T_i = T_i -{1\over N} \sum_{j\in V} T_j$. With the change of variables $ \tilde u_i= u_i+\overline T_i$, we obtain \begin{eqnarray}\label{Psitilde} \Psi(i_0, T, \lambda)=\int_{{{\mathcal H}}_0} e^{\tilde u_{i_0}-\overline T_{i_0}} e^{i<\lambda, \tilde u-\overline T>} \phi(W^T,\tilde u- \overline T) d \tilde u \end{eqnarray} Remark now that $H(W^T, \tilde u -\overline T)= H(W^T, \tilde u -T)$ since $H(W^T,u)$ only depends on the differences $u_i-u_j$. We observe that the coefficients of the matrix $M(W^T,u)$ only contain terms of the form $W_{i,j}e^{u_i+T_i+u_j+T_j}$, hence $$ \sqrt{ D(W^T, \tilde u -\overline T)}= e^{{N-1\over N} \sum_j T_j} \sqrt{D(W,\tilde u)}. $$ Finally, $<\lambda, \overline T>=<\lambda, T>$ since $\lambda\in {{\mathcal H}}_0$. This implies that \begin{eqnarray}\label{Psitilde2} \Psi(i_0, T, \lambda)=\int_{{{\mathcal H}}_0} e^{\sum_j T_j} e^{\tilde u_{i_0}- T_{i_0}} e^{i<\lambda, \tilde u- T>} e^{-H(W^T,\tilde u -T)} \sqrt{D(W,\tilde u)} d\tilde u. \end{eqnarray} We have \begin{eqnarray*} &&{\partial \over \partial T_{i_0}} H(W^T, \tilde u -T) \\ &=& {\partial \over \partial T_{i_0}}\left( 2 \sum_{\{i,j\} \in E } W_{i,j} e^{T_i+T_j}\sinh^2\left( {1\over 2}(\tilde u_i-\tilde u_j -T_i+T_j)\right)\right) \\ &=& 2\sum_{j\sim i_0} W_{i_0,j} e^{T_{i_0}+T_j}\left( \sinh^2\left({1\over 2}( \tilde u_{i_0}-\tilde u_j -T_{i_0}+T_j)\right) -{1\over 2}\sinh\left(\tilde u_{i_0}-\tilde u_j -T_{i_0}+T_j\right)\right) \\ &=& \sum_{j\sim i_0} W_{i_0,j} e^{T_{i_0}+T_j}\left( e^{-\tilde u_{i_0}+\tilde u_j +T_{i_0}-T_j} -1\right) \\ &=& e^{-(\tilde u_{i_0}-T_{i_0})}L(T)(e^{\tilde u-T})(i_0). \end{eqnarray*} Hence, \begin{eqnarray*} &&-{\partial \over \partial T_{i_0}}\Psi(i_0, T, \lambda) \\ &=& \int_{{{\mathcal H}}_0}\left(i\lambda_{i_0}e^{\tilde u_{i_0}-T_{i_0}}+L(T)(e^{\tilde u-T})(i_0)\right) e^{\sum_j T_j} e^{i<\lambda, \tilde u-T>} e^{-H(W^T,\tilde u- T)} \sqrt{ D(W,\tilde u)} d\tilde u \\ &=& i\lambda_{i_0} \Psi(i_0,T,\lambda) +(L(T)\Psi)(i_0,T,\lambda). \end{eqnarray*} This gives \begin{eqnarray*} (\tilde L \Psi)(i_0, T, \lambda) = -i\lambda_{i_0} \Psi(i_0,T,\lambda). \end{eqnarray*} \end{proof} Since $\Psi$ is a solution of the Feynman-Kac equation we deduce that for all $t>0$, $i_0\in V$, $\lambda\in {{\mathcal H}}_0$, $T\in {{\Bbb R}}^V$, $$ \Psi(i_0,T,\lambda)={{\Bbb E}}_{i_0,T}\left( e^{i<\lambda, \overline T (t)>} \Psi(X_t, T(t), \lambda)\right), $$ where we recall that $\overline T_i (t)= T_i(t)-t/N $. Let us now prove that $\Psi(X_t, T(t), \lambda)$ is dominated and that ${{\Bbb P}}_{i_0}$ a.s. \begin{eqnarray}\label{cv} \lim_{t\to \infty} \Psi(X_t, T(t), \lambda) =\sqrt{2\pi}^{N-1}. \end{eqnarray} By the matrix-tree theorem, we have, denoting by ${\mathcal T} $ the set of spanning trees of $G$, and using again notation $\phi$ in \eqref{notphi}, \begin{eqnarray}\label{domination} \nonumber e^{u_{i_0}} \phi(W^T,u) &=& e^{u_{i_0}} e^{-H(W^T,u)} \sqrt{ \sum_{\Lambda \in {\mathcal T}} \prod_{\{i,j\}\in \Lambda} W^T_{i,j} e^{u_i+u_j}} \\ \nonumber &\le& e^{N\max_{i\in V} \vert u_i\vert }e^{-{1\over 2}\sum_{\{i,j\} \in V} W^T_{i,j} (u_i-u_j)^2}\sqrt{ D(W^T,0)} \\ &\le & \left( \sum_{i\in V} e^{N u_i}+e^{-Nu_i}\right) e^{-{1\over 2}\sum_{\{i,j\} \in V} W^T_{i,j} (u_i-u_j)^2}\sqrt{ D(W^T,0)} \end{eqnarray} This is a gaussian integrand: for any real $a$ and $i_0\in V$, \begin{eqnarray*} &&\int_{{{\mathcal H}}_0} e^{a u_{i_0}} e^{-{1\over 2}\sum_{\{i,j\} \in V} W^T_{i,j} (u_i-u_j)^2}\sqrt{ D(W^T,0)} du \\ &=& e^{-{{1\over 2} a^2} Q(T)_{i_0,i_0}}\int e^{-{1\over 2}<U-aQ(T)_{i_0,\cdot}, L(T)(U-aQ(T)_{i_0,\cdot})>}\sqrt{ D(W^T,0)} du \\ &=& e^{-{{1\over 2} a^2} Q(T)_{i_0,i_0}} (2\pi)^{(N-1)/2}. \end{eqnarray*} where $Q(T)$ is defined at the beginning of Section 4.1. Therefore for all $i_0\in V$, $(T_i)\in {{\Bbb R}}^V$ $$ \vert \Psi(i_0,T,\lambda)\vert \le 2 \sum_{i\in V} (2\pi)^{(N-1)/2} e^{-{1\over 2} N^2 Q(T)_{i,i}}, $$ Using \eqref{derivq}, $Q(T(t))_{i,i}$ increases in $t$, hence $$\vert \Psi(X_t, T(t), \lambda)\vert \le 2 \sum_{i\in V} (2\pi)^{(N-1)/2} e^{-{1\over 2} N^2 Q(0)_{i,i}},$$ for all $t\ge 0$. Let us prove now (\ref{cv}). We have \begin{eqnarray*} &&\Psi(X_t, T(t), \lambda) \\ &=& \int e^{i<\lambda,u>} e^{u_{X_t}} e^{-2\sum_{\{i,j\} \in E} W^{T(t)}_{i,j} \sinh^2({1\over 2}(u_i-u_j))}\sqrt{D(W^{T(t)},u)} du \\ &=& \int e^{i<\lambda,u>} e^{u_{X_t}} e^{-2\sum_{\{i,j\} \in E} e^{2t/N} W^{\overline T(t)}_{i,j} \sinh^2({1\over 2}(u_i-u_j))}\sqrt{D(W^{\overline T(t)},u)} e^{(N-1)t/N}du. \end{eqnarray*} Changing to variables $\tilde u_i=e^{t/N} u_i$, we deduce that $\Psi(X_t, T(t), \lambda)$ equals \begin{eqnarray*} \int e^{i<\lambda, e^{-t/N}\tilde u>} e^{e^{-t/N} \tilde u_{X_t}} e^{-2\sum_{\{i,j\} \in E} W^{\overline T(t)}_{i,j} e^{2t/N} \sinh^2({1\over 2} e^{-t/N} (\tilde u_i-\tilde u_j))} \sqrt{D(W^{\overline T(t)},e^{-t/N} \tilde u)} d\tilde u. \end{eqnarray*} Since $\lim_{t\to \infty} \overline T_i(t)=U_i$, the integrand converges pointwise to the Gaussian integrand $$ e^{-{1\over 2} \sum_{\{i,j\} \in V} W^{U}_{i,j} (\tilde u_i-\tilde u_j)^2} \sqrt{D(W^{U},0 )}, $$ whose integral is $\sqrt{2\pi}^{N-1}$. Consider $\overline U_i=\sup_{t\ge 0} \overline T_i(t)$ and $\underline U_i= \inf_{t\ge 0} \overline T_i(t)$. Proceeding as in (\ref{domination}) the integrand is dominated for all $t$ by \begin{eqnarray*} &&e^{Ne^{-t/N} \max_{i\in V} \vert \tilde u_i\vert }e^{-{1\over 2} \sum_{\{i,j\} \in V} W^{\overline T(t) }_{i,j} (\tilde u_i-\tilde u_j)^2} \sqrt{D(W^{\overline T(t)},0 )} \\ &\le & \left( \sum_{i\in V} e^{N \tilde u_i}+e^{-N\tilde u_i}\right)e^{-{1\over 2} \sum_{\{i,j\} \in V} W^{\underline U}_{i,j} (\tilde u_i-\tilde u_j)^2} \sqrt{D(W^{\overline U},0 )}. \end{eqnarray*} which is integrable, which yields (\ref{cv}) by dominated convergence. \subsection{Proof of Theorem \ref{meas} ii)} The same change of variables as in (\ref{Psitilde2}), applied to $T_i=\log \lambda_i$, implies that, for any $j_0\in V$ and $(\lambda_i)_{i\in V}$ positive reals, $$ \frac{\prod_{i \in V} \lambda_i}{\sqrt{2\pi}^{N-1}}e^{u_{j_0}-\log(\lambda_{j_0})}e^{-{1\over 2}\sum_{\{i,j\}\in E} W_{i,j} \lambda_i\lambda_j\left(e^{{1\over 2}(u_j-u_i)}\sqrt{{\lambda_i\over\lambda_j}} -e^{{1\over 2}(u_j-u_i)}\sqrt{{\lambda_j\over\lambda_i}}\right)^2} \sqrt{D(W,u)} $$ is the density of a probability measure, which we call $\nu^{\lambda,j_0}$ (using that \eqref{density} defines a probability measure). Remark that this density can be rewritten as $$ {\prod_{i\in V} \lambda_i \over \sqrt{2\pi}^{N-1}}e^{u_{j_0}-\log(\lambda_{j_0})}e^{-{1\over 2}\sum_i\sum_{j\sim i} W_{i,j}(\lambda_i^2e^{u_j-u_i}- \lambda_i\lambda_j) } \sqrt{D(W,u)} $$ Let $(U_i)$ be a random variable distributed according to (\ref{density}), and, conditionally on $U$, let $(Z_t)$ be the Markov jump process starting at $i_0$, and with jump rates from $i$ to $j$ $$ {1\over 2} W_{i,j}e^{U_j-U_i}. $$ Let $({{\mathcal F}}^Z_t)_{t\ge 0}$ be the filtration generated by $Z$, and let $E^U_i$ be the law of the process $Z$ starting at $i$, conditionally on $U$. We denote by $(l_i(t))_{i\in V}$ the vector of local times of the process $Z$ at time $t$, and consider the positive continuous additive functional of $Z$ $$ B(t)=\int_0^t {1\over 2}{1\over \sqrt{ 1+l_{Z_u}(u)}}du=\sum_{i\in V} \left(\sqrt{ 1+l_i(t)}-1\right), $$ and the time changed process $$ \tilde Y_s =Z_{B^{-1}(s)}. $$ Let us first prove that the law of $U$ conditioned on ${{\mathcal F}}^Z_t$ is \begin{eqnarray}\label{Ucond} {{\mathcal L}}(U | {{\mathcal F}}_t^Z) = \nu^{\lambda(t), Z_t}, \end{eqnarray} where $\lambda_i(t)= \sqrt{1+ l_i(t)}$. Indeed, let $t>0$: if $\tau_1, \ldots, \tau_{K(t)}$ denote the jumping times of the Markov process $Z_t$ up to time $t$, then for any positive test function, \begin{align*} &E^{U}_{i_0}\left( \psi(\tau_1, \ldots, \tau_{K(t)}, Z_{\tau_1}, \ldots, Z_{\tau_{K(t)}})\right)= \\ &\sum_{k=0}^\infty \sum_{i_1, \ldots, i_k} (\prod_{l=0}^{k-1} {1\over 2} W_{i_l, i_{l+1}}) \int_{[0,t]^k} \psi((t_j), (i_j))e^{U_{i_k}-U_{i_0}}e^{-{1\over 2} \sum_{l=0}^{k-1} \left(\sum_{j\sim i_l} W_{i_l,j} e^{U_j-U_{i_l}}\right) (t_{l+1}-t_l)} dt_1\cdots dt_k \end{align*} with the convention $t_{k+1}=t$. Hence, for any test function $G$, \begin{eqnarray*} {{\Bbb E}} \left( G(U)| {{\mathcal F}}^Z_t\right) = {\int_{{{\mathcal H}}_0} G(u) e^{u_{Z_t}}e^{-H(W,u)-{1\over 2}\sum_{i\in V} \left(\sum_{j\sim i} W_{i,j} e^{u_j-u_i}\right)l_i(t)}\sqrt{D(W,u)} du\over \int_{{{\mathcal H}}_0} e^{u_{Z_t}}e^{-H(W,u)-{1\over 2}\sum_{i\in V} \left(\sum_{j\sim i} W_{i,j} e^{u_j-u_i}\right)l_i(t)}\sqrt{D(W,u)} du}. \end{eqnarray*} Using that we can write $H(W,u)= {1\over 2} \sum_{i\in V}\sum_{j\sim i} (e^{u_j-u_i}-1)$, and introducing adequate constants in the numerator and denominator we have \begin{eqnarray*} &&{{\Bbb E}} \left( G(U)| {{\mathcal F}}^Z_t\right) \\ &=& {\sqrt{2\pi}^{-(N-1)} \int_{{{\mathcal H}}_0} G(u) (\prod \lambda_i) e^{u_{Z_t}-\log\lambda_{Z_t}}e^{-{1\over 2}\sum_{i}\sum_{j\sim i} W_{i,j}(\lambda_i(t)^2e^{ u_j-u_i}- \lambda_i(t)\lambda_j(t))}\sqrt{D(W,u)} du\over \sqrt{2\pi}^{-(N-1)}\int_{{{\mathcal H}}_0} (\prod \lambda_i) e^{u_{Z_t}-\log\lambda_{Z_t}}e^{-{1\over 2}\sum_{i}\sum_{j\sim i} W_{i,j}(\lambda_i(t)^2e^{ u_j-u_i}- \lambda_i(t)\lambda_j(t))}\sqrt{D(W,u)} du} \end{eqnarray*} (recall that $\lambda_i(t)=\sqrt{1+l_i(t)}$). The denominator is 1 since it is the integral of the density of $\nu^{\lambda(t), Z_t}$. This proves (\ref{Ucond}). Subsequently, by (\ref{Ucond}), conditioned on $({{\mathcal F}}_t^Z)$, if the process $Z$ is at $i$ at time $t$, then it jumps to a neighbour $j$ of $i$ with rate $$ {1\over 2} W_{i,j} {{\Bbb E}}^{\nu^{\lambda(t),i}}\left( e^{U_j-U_i}\right) = {1\over 2} W_{i,j} {\lambda_j(t)\over \lambda_i(t)}. $$ In order to conclude, we now compute the corresponding rate for $\tilde Y$: by definition, $$ B'(t)={1\over 2}{1\over \sqrt{ 1+l_{Z_t}(t)}}. $$ Therefore, similarly as in the proof of Lemma \ref{invtime}, \begin{eqnarray*} {{\Bbb P}}\left( \tilde Y_{s+ds}=j | {{\mathcal F}}_t^Z\right)&=& {{\Bbb P}}\left(Z_{B^{-1}(s+ds)}=j | {{\mathcal F}}_t^Z\right) \\ &=& {1\over 2} W_{Y_s,j} {1\over B'(B^{-1}(s))} {\lambda_j(B^{-1}(s)) \over \lambda_{Y_s}(B^{-1}(s))}\,ds \\ &=& W_{Y_s,j} \lambda_j(B^{-1}(s))\,ds. \end{eqnarray*} Let $(\tilde{l}_i(s))$ be the local time of the process $\tilde Y$. Then \begin{eqnarray*} (\tilde{l}_i(B(t))'= B'(t) {{\mathbbm 1}}_{\{\tilde Y_{B(s)=i}\}}={1\over 2} (1+l_i(t))^{-{1\over 2}}{{\mathbbm 1}}_{\{Z_t=i\}}. \end{eqnarray*} This implies \begin{equation} \label{ltl} \tilde{l}_i(B(t))= \sqrt{1+l_i(t)} -1 \end{equation} and $$ {{\Bbb P}}\left( \tilde Y_{s+ds}=j | {{\mathcal F}}_t^Z\right)= W_{\tilde Y_s,j} (1+\tilde{l}_j(s))\,ds$$ This means that the annealed law of $\tilde Y$ is the law of a VRJP with conductances $(W_{i,j})$ (this is the content of remark \ref{r3}). Therefore, the process defined, for all $t\ge0$, by $\tilde Y_{A^{-1}(t)}=Z_{(A\circ B)^{-1}(t)}$, is equal in law to $(X_t)_{t\ge0}$; let us denote by $T$ its local time, and show that $T_i(t)-t/N$ converges to $U_i$ as $t\to\infty$, which will complete the proof. First note, using \eqref{tlx} and \eqref{ltl}, that, for all $i\in V$, $$T_i((A\circ B)(t))=\log(\tilde{l}_i(B(t))+1)=\log(1+l_i(t))/2.$$ On the other hand, conditionally on $U$, the Markov Chain $Z$ has invariant measure $(Ce^{2U_i})_{i\in V}$, $C:=(\sum_{i\in V}e^{2U_i})^{-1}$, so that $l_i(t)/(Ce^{2U_i}t)$ converges to $1$ as $t\to\infty$, for all $i\in V$. Therefore, for all $i\in V$, $$T_i(t)-T_{i_0}(t)=\frac{1}{2}\log\left(\frac{1+l_i((A\circ B)^{-1}(t))}{1+l_i((A\circ B)^{-1}(t))}\right),$$ which converges towards $U_i-U_{i_0}$ as $t\to\infty$, which enables us to conclude. \section{Back to Diaconis-Coppersmith formula}\label{BCP} It follows from de Finetti's theorem for Markov chains \cite{diaconis-freedman,rolles1} that the law of the ERRW is a mixture of reversible Markov chains; its mixing measure was explicitly described by Coppersmith and Diaconis (\cite{coppersmith}, see also \cite{keane-rolles1}). Theorems 1 and 2 enable us to retrieve this so-called Coppersmith-Diaconis formula, including its de Finetti part: they imply that the ERRW $(X_n)_{n\in{{\Bbb N}}}$ follows the annealed law of a reversible Markov chain in a random conductance network $x_{i,j}=W_{i,j} e^{U_{i}+U_j}$ where $W_e\sim Gamma(a_e,1)$, $e\in E$, are independent random variables and, conditioned on $W$, the random variables $(U_i)$ are distributed according to the law (\ref{density}). Let us compute the law it induces on the random variables $(x_e)$. The random variable $(x_e)$ is only significant up to a scaling factor, hence we consider a 0-homogeneous bounded measurable test function $\phi$; by Theorem 2, \begin{eqnarray*} &&{{\Bbb E}}\left( \phi((x_e))\right) \\ &=& {1\over \sqrt{ 2\pi}^{N-1}}\int_{{{\Bbb R}}_+^E\times {{\mathcal H}}_0} \phi(x) \left(\prod_{e\in E} {1\over \Gamma(a_e)}W_e^{a_e}e^{-W_e} \right) e^{u_{i_0}} \sqrt{D(W,u)}e^{-H(W,u)}{dW\over W} du \end{eqnarray*} where we write ${dW\over W}=\prod_{e\in E}{dW_e\over W_e}$. Changing to coordinates $\overline u_i=u_i-u_{i_0}$ yields \begin{eqnarray*} C(a)\int_{{{\Bbb R}}_+^E\times {{\Bbb R}}^{V\setminus\{i_0\}}} \phi(x) \left(\prod_{e\in E} W_e^{a_e}e^{-W_e} \right) e^{-\sum_{i\neq i_0} \overline u_i} \sqrt{D(W, \overline u)}e^{-H(W,\overline u)} {dW\over W} d\overline u \end{eqnarray*} with $d\overline u=\prod_{i\neq i_0} d\overline u_i$ and $C(a)= {1\over \sqrt{ 2\pi}^{N-1}}\left(\prod_{e\in E} {1\over \Gamma(a_e)}\right)$. But $$ -\sum_{e\in E} W_e -H(W,\overline u)= -{1\over 2} \sum_{\{i,j\}\in E} W_{i,j}e^{\overline u_i+\overline u_j}\left(e^{-2\overline u_j}+e^{-2\overline u_i}\right). $$ The change of variables $$\left((x_{i,j}=W_{i,j}e^{\overline u_i+\overline u_j})_{\{i,j\}\in E}, (v_i=e^{-2\overline u_i})_{i\in V\setminus\{i_0\}}\right), $$ with $v_{i_0}=1$ implies $$ -\sum_{e\in E}W_e -H(W,\overline u)= -{1\over 2} \sum_{i\in V } v_i x_i, $$ where $x_i=\sum_{j\sim i} x_{i,j}$, and ${{\Bbb E}}\left( \phi((x_e))\right)$ is equal to the integral \begin{eqnarray*} C'(a)\int \phi(x) \left(\prod_{e\in E} x_e^{a_e} \right) \left( \prod_{i\in V} v_i^{(a_i+1)/2}\right) v_{i_0}^{-{1\over 2}} \sqrt{D(x)} e^{-{1\over 2} \sum_{i\in V } v_i x_i} \left(\prod_{e\in E} {dx_e\over x_e}\right) \left( \prod_{i\neq i_0} {dv_i\over v_i}\right), \end{eqnarray*} with $a_i=\sum_{j\sim i} a_{i,j}$, $D(x)$ determinant of any diagonal minor of the $N\times N$ matrix $$ m_{i,j}=\left\{ \begin{array}{ll} x_{i,j} & \hbox{ if $i\neq j$} \\ -\sum_{k\sim i} x_{i,k} &\hbox{ if $i=j$} \end{array}\right. $$ and $$ C'(a)= {2^{-N+1}\over \sqrt{ 2\pi}^{N-1}}\left(\prod_{e\in E} {1\over \Gamma(a_e)}\right). $$ Let $e_0$ be a fixed edge: we normalize the conductance to be 1 at $e_0$ by changing to variables $$\left(\left(y_e={x_e\over x_{e_0}}\right)_{e\neq e_0}, \left( z_i=x_{e_0} {v_i}\right)_{i\in V}\right), $$ with $y_{e_0}=1$. Now, observe that $$ \left(\prod_{e\in E} {dx_e\over x_e}\right) \left( \prod_{i\neq i_0} {dv_i\over v_i}\right)= \left(\prod_{e\in E, e\neq e_0} {dy_e\over y_e}\right) \left( \prod_{i \in V} {dz_i\over z_i}\right). $$ We deduce that ${{\Bbb E}}\left( \phi((x_e))\right)$ equals the integral \begin{eqnarray*} C(a)\int_{{{\Bbb R}}_+^V\times {{\Bbb R}}_+^{E\setminus\{e_0\}}} \phi(y) \left(\prod_{e\in E} y_e^{a_e} \right) \left( \prod_{i\in V} z_i^{a_i/2}\right) z_{i_0}^{-{1\over 2}} \sqrt{D(y)} e^{-{1\over 2} \sum_{i\in V } z_i y_i} \left({dy\over y}\right)\left({dz\over z}\right), \end{eqnarray*} with ${dy\over y}= \prod_{e\neq e_0} {dy_e\over y_e}$ and ${dz\over z}=\prod_{i\in V} {dz_i\over z_i}$. Therefore, integrating over the variables $z_i$ \begin{eqnarray*} {{\Bbb E}}\left( \phi((x_e))\right) = C''(a)\int_{{{\Bbb R}}_+^{E\setminus\{e_0\}}} \phi(y) y_{i_0}^{{1\over 2}} {\left(\prod_{e\in E} y_e^{a_e} \over \prod_{i\in V} y_i^{(a_i+1)/2}\right)} \sqrt{D(y)} \left({dy\over y}\right), \end{eqnarray*} where $$ C''(a)={2^{1-N-\sum_{e\in E} a_e}\over \pi^{(N-1)/2}}{\Gamma(a_{i_0}/2) \prod_{i\neq i_0} \Gamma((a_i+1)/2)\over \prod_{e\in E} \Gamma(a_e)} $$ which is Diaconis-Coppersmith formula: the extra term $(\vert E\vert -1)!$ in \cite{keane-rolles1,diaconis-rolles} arises from the normalization of $(x_e)_{e\in E}$ on the simplex $\Delta=\{\sum x_e=1\}$ (see Section 2.2 \cite{diaconis-rolles}). \section{The supersymmetric hyperbolic sigma model} We first relate VRJP to the supersymmetric hyperbolic sigma model studied in Disertori, Spencer and Zirnbauer \cite{dsz,ds}. For notational purposes, we restrict our attention to the $d$-dimensional lattice, that is, our graph is ${{\Bbb Z}}^d$ with $x\sim y$ if $\vert x-y\vert_1 =1$. We denote by $E$ the set of edges $E=\{\{i,j\}, \; i\sim j\}$. For a subset $\Lambda \subset {{\Bbb Z}}^d$ we denote by $E_{\Lambda}$ the set of edges with both extremities in $\Lambda$. We start by a description of the measures defined in \cite{dsz,ds}. Let $V\subset {{\Bbb Z}}^d$ be a connected finite subset containing $0$. Let $\beta_{i,j}$, $i,j\in V$, $i\sim j$ be some positive weights on the edges, and ${\mathbb E}=(\epsilon_i)_{i \in V}$ be a vector of non-negative reals, $\epsilon \neq 0$. Let $\mu_V^{{\mathbb E},\beta}$ be a generalization of the measure studied in \cite{ds} (see (1.1)-(1.7) in that paper), namely \begin{align*} d\mu_V^{{\mathbb E},\beta}(t):&=\left(\prod_{j\in V}\frac{dt_j}{\sqrt{2\pi}}\right)e^{-\sum_{k\in V}t_k}e^{-F_V^\beta(\nabla t)}e^{-M_V^{{\mathbb E}}(t)}\sqrt{\mathsf{det}\,\,A_V^{{\mathbb E},\beta}}\\ &=\left(\prod_{j\in V}\frac{dt_j} {\sqrt{2\pi}}\right)e^{-F_V^\beta(\nabla t)}e^{-M_V^{\mathbb E}(t)}\sqrt{\mathsf{det}\,\,D_V^{{\mathbb E},\beta}} \end{align*} where $A_V^{{\mathbb E},\beta}=A^{{\mathbb E},\beta}$ and $D_V^{{\mathbb E},\beta}=D^{{\mathbb E},\beta}$ are defined by, for all $i$, $j$ $\in V$, by $$ A_{ij}^{{\mathbb E},\beta}=e^{t_i}D_{ij}^{{\mathbb E},\beta} e^{t_j}= \begin{cases} 0&|i-j|>1\\ -\beta_{ij}e^{t_i+t_j}&|i-j|=1\\ \sum_{l\sim i, l\in V}\beta_{il}e^{t_i+t_l}+{\mathbb E}_i e^{t_i}&i=j\\ \end{cases} $$ \begin{align*} F_V^\beta(\nabla t)&:=\sum_{\{i,j\}\in E_V}\beta_{ij}(\cosh(t_i-t_j)-1)\\ M_V^{\mathbb E}(t)&:=\sum_{i\in V}{\mathbb E}_i(\cosh t_i-1). \end{align*} The fact that $\mu_V^{{\mathbb E},\beta}$ is a probability measure can be seen as a consequence of supersymmetry (see (5.1) in \cite{dsz}). This is also a consequence of Theorem \ref{meas} i), cf later. The measure $\mu_V^{{\mathbb E},\beta}$ is directly related to the measure (\ref{density}) defined in theorem \ref{meas} as follows. Let add an extra point $\delta$ to $V$, $\tilde V=V\cup\{\delta\}$, and extra edges $\{i,\delta\}$ connecting any site $i\in V$ to $\delta$, i.e. $\tilde E_V=E_V\cup\cup_{i\in V}\{i,\delta\}$. Consider the VRJP on $\tilde V$ starting at $\delta$ and with conductances $W_{i,j}=\beta_{i,j}$, if $i\sim j$ in $V$, and $W_{i,\delta}=\epsilon_i$. Let us again use notation $(U_i)_{i\in\tilde{V}}$ for the limiting centred occupation times of VRJP on $\tilde V$ starting at $\delta$, and consider the change of variables, from ${{\mathcal H}}_0$ into ${{\Bbb R}}^V$, which maps $u_i$ to $t_i:=u_i-u_\delta$. Then, by Theorem \ref{meas}, for any test function $\phi$, letting $\iota$ be the canonical injection ${{\Bbb R}}^V\longrightarrow{{\Bbb R}}^{\tilde{V}}$, \begin{align*} \mathbb{E}^{W}_\delta(\phi(U-U_\delta ))&=\frac{1}{(2\pi)^{\vert V\vert/2}}\int_{{{\mathcal H}}_0}\phi(u-u_\delta)e^{u_\delta}e^{-H(W,u)} \sqrt{D(W,u)}\,du\\ &=\frac{1}{(2\pi)^{\vert V\vert/2}}\int_{{{\Bbb R}}^V}\phi(t)e^{-\sum_{i\in V}t_i}e^{-H(W,\iota(t))} \sqrt{D(W,\iota(t))}\left(\prod_{i\not=\delta}\,dt_i\right)\\ &=\mu_V^{{\mathbb E},\beta}\left(\phi(t)\right), \end{align*} which means that $U-U_\delta$ is distributed according to $\mu_V^{{\mathbb E},\beta}$. Indeed $A_V^{\mathbb E}$ is the restriction to $V\times V$ of the matrix $M(W,\iota(t))$ (which is on $\tilde{V}\times \tilde{V}$) (so that $\mathsf{det}\,\,A_V^{\mathbb E}=D(W,\iota(t))$), and $F_V(\nabla t)+M_V^{\mathbb E}(t)=H(W,\iota(t))$. We will be interested in the VRJP on finite subsets of ${{\Bbb Z}}^d$ starting at 0. In order to apply directly results of \cite{ds} we consider the VRJP on ${{\Bbb Z}}^d$ with an extra point $\delta$ uniquely connected to 0 and with $W_{0,\delta}=\epsilon_0=1$, $W_{i,j}=\beta_{i,j}$, $i\sim j$ in ${{\Bbb Z}}^d$. Clearly, the trace on ${{\Bbb Z}}^d$ of the VRJP starting from $\delta$ has the law of the VRJP on ${{\Bbb Z}}^d$ starting from 0. When $V$ contains $0$ the limiting occupation time $U_i -U_\delta$ of the VRJP on $\tilde V=V\cup\{\delta\}$ starting at $\delta$ is then distributed according to $d\mu_V^{\delta_0,\beta}$, where $\delta_0$ is the Dirac at 0. Set, for all $\beta>0$, $$I_\beta:=\sqrt{\beta}\int_{-\infty}^\infty\frac{dt}{\sqrt{2\pi}}e^{-\beta(\cosh t-1)}, $$ which is strictly increasing in $\beta$. Let $\beta^d_c$ be defined as the unique solution to the equation $$ I_{\beta^d_c}e^{\beta^d_c(2d-2)} (2d-1)=1 $$ for all $d>1$, $\beta^d_c:=\infty$ if $d=1$. If the parameters $\beta_e$ are constant equal to $\beta$ over all edges $e$, then Theorem 2 in \cite{ds} readily implies that VRJP over any graph ${{\Bbb Z}}^d$ is recurrent for $\beta<\beta^d_c$ (i.e. for large reinforcement). \begin{theorem}[Disertori and Spencer \cite{ds}, Theorem 2]\label{exp} Suppose that $\beta_e=\beta$ for all $e$. Then there exists a constant $C_0:=2d/(2d-1)>0$ such that, for all $\Lambda\subset {{\Bbb Z}}^d$ finite connected containing 0, $x\in\Lambda$, $0<\beta <\beta^d_c$, $$ \mu_\Lambda^{\eta\delta_0,\beta}\left( e^{t_x/2} \right)\le C_0 I_{\eta}\left[ I_\beta e^{\beta(2d-2)}(2d-1)\right]^{\vert x\vert}. $$ \end{theorem} \begin{corollary}\label{rec} For $0<\beta <\beta^d_c$, let $(Y_n)$ be the discrete time process associated with the VRJP on ${{\Bbb Z}}^d$ starting from 0 with constant conductance $\beta$. Then $(Y_n)$ is a mixture of reversible positive recurrent Markov chains. \end{corollary} \begin{proof} We prove this for the VRJP on ${{\Bbb Z}}^d$ with an extra point $\delta$ connected to 0 only, and conductances $W_{x,y}=\beta$ and $W_{0,\delta}=1$, which is clearly stronger. On a finite connected subset $V\subset {{\Bbb Z}}^d$ containing 0, we know from Theorem \ref{meas} that $(Y_n)_{n\in{{\Bbb N}}}$, the discrete-time process associated with $(Y_s)_{s\ge 0}$, is a mixture of reversible Markov chains with conductances $c_{x,y}=\beta e^{t_x+t_y}$, where $(t_x)_{x\in V}$ has law $\mu_V^{\delta_0,\beta}$. Now Theorem \ref{exp} implies that $\mu_V^{\delta_0,\beta}( (c_{e}/c_{\delta,0})^{1/4})$ decreases exponentially with the distance from $e$ to 0: indeed, by Cauchy-Schwarz inequality, \begin{eqnarray*} \mu_V^{\delta_0,\beta}( (c_{x,y}/c_{\delta,0})^{1/4})&\le& \left[\mu_V^{\delta_0,\beta}(e^{t_x/2})\mu_V^{\delta_0,\beta}(e^{(t_y-t_0)/2})\right]^{1/2} \\ &\le& C\left[\mu_V^{\delta_0,\beta}(e^{t_x/2}) \mu_V^{\delta_0,\beta}(e^{{1\over 2} (\cosh(t_0)-1)}e^{t_y/2})\right]^{1/2} \\ &\le& 2 C\left[\mu_V^{\delta_0,\beta}(e^{t_x/2})\mu_V^{\delta_0/2,\beta}(e^{t_y/2})\right]^{1/2} \end{eqnarray*} for some $C>0$ such that $\vert z \vert\le 4\log C + \cosh (z)-1$. This implies that there exists constants $c_1>0$, $c_2>0$, such that $\mu_V^{\delta_0,\beta}((c_{x,y}/c_{\delta,0})>e^{-c_1 \vert x\vert}) \le e^{-c_2 \vert x\vert}$. Following the proof of lemma 5.1 of \cite{merkl-rolles4} it implies that $(Y_n)$ is a mixture of positive recurrent Markov chains. \end{proof} By Theorems \ref{annealed} and \ref{meas}, the ERRW with constant initial weights $a>0$ corresponds to the case where $(\beta_{e})_{e\in E}$ are independent random variables with $\hbox{Gamma} (a,1)$ distribution for some parameter $a>0$: we can infer a similar localization and recurrence result for $a$ small enough. This requires an extension of Theorem \ref{exp} for random gamma weights $(\beta_e)_{e\in E}$: we propose one in the following Proposition \ref{expvrjp}, in the same line of proof as in \cite{ds}. \begin{proposition} \label{expvrjp} Let $a$ be a positive real, and assume that the conductances $(\beta_{e})_{e\in E}$ are i.i.d. with law Gamma$(a,1)$. Denote by ${{\Bbb E}}$ the expectation with respect to the random variables $(\beta_e)_{e\in E}$. For all $d\ge 1$, there exists $a_c^d>0$, $\delta\in(0,1)$, such that if $a<a_c^d$, there exists $C_0>0$ depending only on $a$ and $d$ such that for all $x\sim y$, $$ {{\Bbb E}}\left(\mu_\Lambda^{\eta \delta_0,\beta}(e^{t_x/2})\right)\le C_0 I_\eta\delta^{\vert x\vert}, \;\;\; {{\Bbb E}}\left((\beta_{x,y})^{{1\over 4}} \mu_\Lambda^{\eta \delta_0,\beta}(e^{t_x/2})\right)\le C_0I_\eta \delta^{\vert x\vert} $$ independently on the finite connected subset $\Lambda$ containing 0 and $x$, $y$. \end{proposition} \begin{corollary}\label{rec2} The ERRW on ${{\Bbb Z}}^d$ starting at 0 with constant initial weight $a>0$ is a mixture of positive recurrent Markov chains for $a<a_c^d$ (where $a_c^d$ is defined in Proposition \ref{expvrjp}). \end{corollary} \begin{remark} Corollary \ref{rec2} also holds on any graph of bounded degree, and for possibly non-constant weights $(a_e)_{e\in E}$ with $a_e<a_c$ for some $a_c>0$. Indeed, the proof of Proposition \ref{expvrjp} also holds for independent (not necessarily i.i.d.) conductances $(\beta_e)$ with $\mathbb{E}(\sqrt{\beta_e}(\log(1+\beta_e^{-1}))$ sufficiently small, when the graph is of bounded degree. \end{remark} \begin{corollary} \label{trans} For any $d\ge3$, there exists $\beta_c(d)$ such that, for all $\beta>\beta_c(d)$, VRJP on ${{\Bbb Z}}^d$ with constant conductance $\beta$ is transient. \end{corollary} \begin{proof} (Proposition \ref{expvrjp}) The strategy is to follow the proof of \cite{ds}, Theorem 2, and to truncate the random variables $\beta_e$ at adequate positions. For convenience we provide a self-contained proof but the only new input compared to \cite{ds}, Theorem 2, lies in the threshold argument (\ref{thres1}--\ref{thres3}). Let us define, for any $\Lambda\subset {{\Bbb Z}}^d$, $(\epsilon_i)_{i\in \Lambda}\in{{\Bbb R}}_+^\Lambda$ $$d\nu_\Lambda^{{\mathbb E} ,\beta}(t):=\left(\prod_{i\in \Lambda}\frac{dt_j} {\sqrt{2\pi}}\right)e^{-F^\beta_\Lambda(\nabla t)}e^{-M_\Lambda^{{\mathbb E}}(t)},$$ which is not a probability measure in general. We fix now a finite connected subset $\Lambda\subset {{\Bbb Z}}^d$ containing 0, and $x$. Let $\Gamma_x$ be the set of non-intersecting paths in $\Lambda$ from $0$ to $x$. For notational purposes, any element $\gamma$ in $\Gamma_x$ is defined here as the set of non-oriented edges in the path. We let $\Lambda_\gamma$ and $\Lambda_\gamma^c$ be respectively the set of vertices in the path and its complement. We say that an edge $e$ is adjacent to the path $\gamma$ if $e$ is not in $\gamma$ and has one adjacent vertex in $\gamma$, i.e. if $e=\{i,j\}$ with $i\in\Lambda_\gamma, j\not\in\Lambda_\gamma$; we write $e\sim \gamma$. We first proceed similarly to (3.1)-(3.4) in \cite{ds}, Lemma 2. Let ${\mathcal T}_\Lambda$ be the set of spanning trees of $\Lambda$. By the matrix-tree theorem $$ \det(A_\Lambda^{\eta\delta_0,\beta})= \eta e^{t_0} \sum_{T\in {\mathcal T}_\Lambda} \prod_{\{i,j\}\in T} \beta_{\{i,j\}} e^{t_{i}+t_{j}}. $$ In a spanning tree $T$ there is a unique path between 0 and $x\in \Lambda$. Decomposing this sum depending on this path we deduce $$ \det(A_\Lambda^{\eta\delta_0,\beta})= \eta e^{t_0} \sum_{\gamma\in\Gamma_x} \left( \prod_{\{i,j\}\in \gamma} \beta_{\{i,j\}} e^{t_i+t_j}\right) \sum_{T'\in {\mathcal T}_\Lambda^\gamma} \prod_{{\{i,j\}}\in T'} \beta_{\{i,j\}} e^{t_{i}+t_{j}} $$ where ${\mathcal T}_\Lambda^\gamma$ is the set of subsets $T'\subset E_\Lambda\setminus \gamma$ such that $\gamma\cup T'$ is a spanning tree. By the matrix-tree theorem, we have \begin{eqnarray} \label{MTA} \sum_{T'\in {\mathcal T}_\Lambda^\gamma} \prod_{{\{i,j\}}\in T'} \beta_{\{i,j\}} e^{t_{i}+t_{j}}= \det ( A_{\Lambda_\gamma^c}^{\epsilon, \beta}) \end{eqnarray} where $(\epsilon_i)_{i\in \Lambda_\gamma^c}$ is the vector defined by $${\mathbb E}_i:=\sum_{k\in\Lambda_\gamma, k\sim i}\beta_{\{i,k\}}e^{t_k}, \;\;\; \forall i\in \Lambda_\gamma^c$$ It follows that \begin{eqnarray} \label{DDD} \mathsf{det}\, D_\Lambda^{\eta\delta_0,\beta}=\eta e^{-t_x}\sum_{\gamma\in\Gamma_x}\left(\prod_{e\in\gamma}\beta_{e}\right) \,\,\mathsf{det}\, D_{\Lambda_\gamma^c}^{{\mathbb E},\beta}. \end{eqnarray} Let us define, similarly as in (2.12) and (2.14) in \cite{ds}, for $t_\gamma=t_{|\Lambda_\gamma}$ the restriction of $t$ to the vertices on the path $\gamma$, \begin{align} \label{zzz} Z_{\Lambda_\gamma^c}^{\gamma,\beta}(t_{\gamma})&:=\nu_{\Lambda_\gamma^c}^{\eta\delta_0,\beta} \left(\sqrt{\mathsf{det}\, D_{\Lambda_\gamma^c}^{{\mathbb E},\beta}}e^{-F_{\partial\gamma}^\beta(\nabla t)}\right)\\ \nonumber F_{\partial\gamma}^\beta(\nabla t)&:=\sum_{k\in\Lambda_\gamma, j\in\Lambda_\gamma^c, k\sim j}\beta_{kj}(\cosh(t_j-t_k)-1). \end{align} Now \begin{align} \nonumber \mu_\Lambda^{\eta\delta_0,\beta}\left(e^{t_x/2}\right)&=\nu_\Lambda^{\eta\delta_0,\beta}\left(\sqrt{\mathsf{det}\, D_\Lambda^{\eta\delta_0,\beta} e^{t_x}}\right) =\sqrt{\eta}\nu_\Lambda^{\eta\delta_0,\beta}\left(\sqrt{\sum_{\gamma\in\Gamma_x}\prod_{e\in\gamma}\beta_e \,\,\mathsf{det}\, D_{\Lambda_\gamma^c}^{{\mathbb E},\beta}}\right)\\ \label{estmu} &\le\sqrt{\eta} \sum_{\gamma\in\Gamma_x}\left(\prod_{e\in\gamma}\sqrt{\beta_e}\right)\nu_{\Lambda_\gamma}^{\eta\delta_0,\beta}\left(Z_{\Lambda_\gamma^c}^{\gamma,\beta}(t_\gamma)\right), \end{align} using (\ref{DDD}) in the second equality and, in the inequality, that for all $\gamma\in\Gamma_x$, $$d\nu_\Lambda^{\eta\delta_0,\beta}(t)=d\nu_{\Lambda_\gamma}^{\eta\delta_0,\beta}(t)d\nu_{\Lambda_\gamma^c}^{\eta\delta_0,\beta}(t)e^{-F_{\partial\gamma}(\nabla t)}.$$ The new argument compared to theorem \ref{exp} which allows to handle the case of random parameters $\beta$ is the following truncation. Given $\gamma\in\Gamma_x$, let $(\tilde\beta_e)$ be the set of random variables defined by \begin{eqnarray} \label{thres1} \tilde\beta_e=\left\{\begin{array}{ll} \min(\beta_e,1),&\hbox{ if $e\sim \gamma$,} \\ \beta_{e}, &\hbox{otherwise.} \end{array} \right. \end{eqnarray} First note that, trivially, \begin{eqnarray}\label{thres2} e^{-F^\beta_{\partial \gamma}(\nabla t)}\le e^{-F^{\tilde \beta}_{\partial \gamma}(\nabla t)}. \end{eqnarray} On the other hand, identity (\ref{MTA}) implies that \begin{eqnarray}\label{thres3} \det(D_{\Lambda_\gamma^c}^{{\mathbb E},\beta})\le \det(D_{\Lambda_\gamma^c}^{\tilde{E},\tilde \beta})\left( \prod_{e\sim \gamma} \max(\beta_{e},1)\right), \end{eqnarray} where $(\tilde\epsilon_i)_{i\in \Lambda_\gamma^c}$ is the vector defined by $$\tilde{E}_i:=\sum_{k\in\Lambda_\gamma, i\sim k}\tilde\beta_{\{i,k\}}e^{t_k}, \;\;\; \forall i\in \Lambda_\gamma^c$$ (In the last argument we used that, for any $\{i,j\}$ adjacent to $\gamma$, $\beta_{i,j}=\tilde \beta_{i,j}\max(1,\beta_{i,j})$). Therefore \begin{eqnarray}\label{thres4} Z_{\Lambda_\gamma^c}^{\gamma,\beta}(t_{\gamma})\le Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_{\gamma}) \prod_{e\sim \gamma} \sqrt{\max(\beta_{e},1)} \end{eqnarray} with $Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_\gamma)$ defined as in (\ref{zzz}) with $\epsilon, \beta$ replaced by $\tilde{E},\tilde\beta$. Hence we can replace $\beta$ by $\tilde \beta$ at the cost of the term $\prod_{e\sim \gamma} \sqrt{\max(\beta_{e},1)}$. \noindent The following lemma, which adapts Lemma 3 \cite{ds}, provides an upper bound of $Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_\gamma)$. \begin{lemma} \label{estz} For any configuration of $t_\gamma=t_{| \Lambda_\gamma}$, $Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_\gamma)\le e^{\sum_{e\sim \gamma}\tilde\beta_{e}}.$ \end{lemma} \begin{proof} We have $$ Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_\gamma) =\int \left({\prod_{j\in \Lambda_\gamma^c} dt_j\over \sqrt{2\pi}}\right) e^{-F_{\Lambda_\gamma^c}^{\tilde\beta}(\nabla t)-F^{\tilde\beta}_{\partial \gamma}(\nabla t)} \sqrt{ \det( D_{\Lambda_\gamma^c}^{\tilde{\mathbb E}, \tilde\beta})} $$ Let $t^*=\max\{t_k, \; k\in \Lambda_\gamma\}$. We perform the following translation in the variables $t_j\rightarrow t_j+t^*$ for $j\in \Lambda_\gamma^c$: in the previous integral the term $F_{\Lambda_\gamma^c}^{\tilde\beta}(\nabla t)$ does not change, the term $F^{\tilde\beta}_{\partial \gamma}(\nabla t)$ becomes $$ \sum_{k\in\Lambda_\gamma, j \in\Lambda_\gamma^c, k\sim j}\tilde\beta_{kj}(\cosh(t_j+t^*-t_k)-1), $$ and the term $\det(D_{\Lambda_\gamma^c}^{\tilde{\mathbb E}, \tilde\beta})$ is replaced by $\det(D_{\Lambda_\gamma^c}^{e^{-t^*}\tilde{\mathbb E}, \tilde\beta})$. Since $t^*-t_k\ge0$ we have $$ \cosh (t_j+t^*-t_k)-1\ge e^{t_k-t^*}(\cosh(t_j)-1)+(e^{t_k-t^*}-1). $$ This implies that $$ \sum_{k\in\Lambda_\gamma, j \in\Lambda_\gamma^c, k\sim j}\tilde\beta_{kj}(\cosh(t_j+t^*-t_k)-1)\ge M_{\Lambda_\gamma^c}^{e^{-t^*} \tilde {\mathbb E}}(t) + \sum_{k\in\Lambda_\gamma, j \in\Lambda_\gamma^c, k\sim j} \tilde\beta_{k,j} (e^{t_k-t^*}-1), $$ and $$ Z_{\Lambda_\gamma^c}^{\gamma,\tilde\beta}(t_\gamma)\le e^{\sum_{k\in\Lambda_\gamma, j \in\Lambda_\gamma^c, k\sim j} \tilde\beta_{k,j} (1-e^{t_k-t^*})} \mu_{\Lambda_\gamma^c}^{e^{-t^*}\tilde{\mathbb E},\tilde\beta}(1) \le e^{\sum_{e\sim \gamma}\tilde\beta_{e}}, $$ using that $\mu_{\Lambda_{\gamma}^c}^{e^{-t^*}\tilde{\mathbb E},\tilde\beta}$ is a probability measure. \end{proof} Combining \eqref{estmu}, (\ref{thres4}), Lemma \ref{estz}, and integration on the variables $(\nabla t_e)_{e\in \gamma}$, we obtain that $$ \mu_{\Lambda}^{\eta\delta_0, \beta}\left( e^{t_x/2}\right) \le I_{\eta}\sum_{\gamma\in\Gamma_x}\left( \prod_{e\sim \gamma} \sqrt{\max(\beta_{e},1)}e^{\min(\beta_e,1)}\right) \left(\prod_{e\in \gamma} I_{\beta_e}\right) $$ We set $$ \hat I_a ={{\Bbb E}}(I_\beta)={\Gamma(a+{1\over 2})\over \Gamma(a)} \int_{-\infty}^\infty \cosh(t)^{-a-{1\over 2}} dt, $$ and $$ \hat J_a = {{\Bbb E}}(\max (\beta,1)e^{\min(\beta,1)}). $$ where $\beta$ is a gamma random variable with parameter $a$. Clearly, $\hat I_a$ and $\hat J_a$ tend respectively to 0 and 1 when $a$ tends to 0. Integrating on the random variables $\beta_e$ we deduce that there exists a constant $C_0$ such that $$ {{\Bbb E}}\left(\mu_\Lambda^{\eta\delta_0,\beta}(e^{t_x/2})\right)\le C_0 \left( \hat I_a (\hat J_a)^{2d-2}(2d-1)\right)^{\vert x\vert}, $$ $$ {{\Bbb E}}\left((\beta_{x,y})^{{1\over 4}}\mu_\Lambda^{\eta\delta_0,\beta}(e^{t_x/2})\right)\le C_0 \left( \hat I_a (\hat J_a)^{2d-2}(2d-1)\right)^{\vert x\vert}, $$ for any $y\sim x$. This provides the exponential decrease for all $a>0$ such that $$ \hat I_a (\hat J_a)^{2d-2}(2d-1)<1. $$ \end{proof} \begin{proof}(Corollary \ref{rec2}) For any connected finite set $\Lambda$ containing 0, by Theorems \ref{annealed} and \ref{meas}, the ERRW on $\Lambda$ starting 0 and with constant initial parameter $a$ is a mixture of reversible Markov chains in conductance $c_{x,y}=\beta_{x,y} e^{t_x+t_y}$, where $\beta_{x,y}$ are Gamma$(a,1)$ independent random variables. As in Corollary \ref{rec}, there exists a constant $C>0$ such that \begin{eqnarray*} {{\Bbb E}}\left(\mu_\Lambda^{\delta_0,\beta}( (c_{x,y}/c_{\delta,0})^{1/4})\right) &\le& C{{\Bbb E}}\left( (\beta_{x,y})^{{1\over 4}}\left[\mu_\Lambda^{\delta_0,\beta}(e^{t_x/2})\mu_\Lambda^{\delta_0/2,\beta}(e^{t_y/2})\right]^{1/2} \right) \\ &\le& C{{\Bbb E}}\left( (\beta_{x,y})^{{1\over 4}}\mu_\Lambda^{\delta_0,\beta}(e^{t_x/2})\right)^{1/2} {{\Bbb E}}\left((\beta_{x,y})^{{1\over 4}}\mu_\Lambda^{\delta_0/2,\beta}(e^{t_y/2})\right)^{1/2} \end{eqnarray*} The rest of the proof is similar to the proof of Corollary \ref{rec}. \end{proof} \begin{proof} (Corollary \ref{trans}) Fix $d\ge3$. Let $\Lambda_n=\{i\in {{\Bbb Z}}^d, \; \| i\|_\infty \le n\}$ be the ball centred at $0$ with radius $n$ and $\partial \Lambda_n=\{i\in {{\Bbb Z}}^d, \; \| i\|_\infty = n\}$ its boundary. Let $E_n$ be the set of edges contained in $\Lambda_n$. Disertori, Spencer and Zirnbauer prove in Theorem 1 \cite{dsz} (see also their remark above) that, for any $m>0$, there exists $\tilde{\beta}_c(m)$ such that, for any $\beta>\tilde \beta_c(m)$, for any $n\in{{\Bbb N}}$, $x$, $y$ $\in\Lambda_n$, \begin{equation} \label{dsz} \mu_{\Lambda_n}^{\delta_0,\beta}\left(\cosh^m(t_x-t_y)\right)\le2; \end{equation} the result is stated for constant pinning, but its proof does not require that assumption, as we checked through careful reading. We consider the VRJP on ${{\Bbb Z}}^d$ with constant conductances $W_{i,j}=\beta$ and denote by ${{\Bbb P}}_0^\beta(.)$ its law starting from $0$. We denote by $P_0^c$ the law of the Markov chain in conductances $c_{i,j}= \beta e^{t_i+t_j}$ starting from 0, where $(t_i)$ is distributed according to $\mu_{\Lambda_n}^{\delta_0, \beta}$. Let $H_{\partial\Lambda_n}$ be the first hitting time of the boundary $\partial \Lambda_n$ and $\tilde H_0$ be the first return time to the point $\delta$. Let $R(0,\partial\Lambda_n)$ (resp. $R(0,\partial\Lambda_n,c)$) be the effective resistance between $0$ and $\partial\Lambda_n$ for conductances $1$ (resp. $c_{i,j}$). Classically $$ c_{0} R(0, \partial \Lambda_n,c) = {1\over P^c_0 (H_{\partial \Lambda_n}<\tilde H_0)} $$ with $c_0=\sum_{j\sim 0} c_{0,j}$. By Theorem \ref{meas} and Jensen's inequality, \begin{eqnarray} {1\over {{\Bbb E}}^\beta_0( H_{\partial\Lambda_n} <\tilde H_0 )}\le \mu_{\Lambda_n}^{\delta_0, \beta} ({1\over P^c_0 (H_{\partial \Lambda_n}<\tilde H_0)}) &\le& \mu_{\Lambda_n}^{\delta_0, \beta} ( c_0 R(0, \partial \Lambda_n,c)) \label{inbd} \end{eqnarray} Let us now show that for all $\beta>\tilde{\beta}_c(2)$, \begin{eqnarray}\label{Resistance} \mu_{\Lambda_n}^{\delta_0,\beta}\left[c_0 R(0,\partial\Lambda_n,c)\right]\le 16d R(0,\partial\Lambda_n) \end{eqnarray} This will enable us to conclude: since $\limsup R(0,\partial \Lambda_n) <\infty$, \eqref{inbd} and \eqref{Resistance} imply that $ \mathbb{P}^\beta_0( \tilde H_0 =\infty)>0$. Indeed, let $\theta$ be the unit flow from 0 to $\partial \Lambda_n$ which minimizes the $L^2$ norm. Then $$ R(0,\partial \Lambda_n, c) \le \sum_{\{i,j\}\in E_n} {1\over c_{i,j}} \theta^2(i,j), $$ and $$ R(0,\partial \Lambda_n) = \sum_{\{i,j\}\in E_n} \theta^2(i,j). $$ Now, for all $\beta>\tilde \beta_c(2)$, using identity \eqref{dsz}, $$ \mu_{\Lambda_n}^{\delta_0, \beta} ( {c_0\over c_{i,j}})\le \sum_{l\sim0} \mu_{\Lambda_n}^{\delta_0, \beta}(e^{2(t_0-t_i)})^{1/2} \mu_{\Lambda_n}^{\delta_0, \beta}(e^{2(t_l-t_j)})^{1/2} \le 16d $$ \end{proof} \medskip \noindent {\bf Acknowledgment.} We are particularly grateful to Krzysztof Gawedzki for a helpful discussion on the hyperbolic sigma model, and for pointing out reference \cite{dsz}. We would also like to thank Denis Perrot and Thomas Strobl for suggesting a possible link between the limit measure of VRJP and sigma models. We are also grateful to Margherita Disertori for a useful discussion on localization results on the hyperbolic sigma model. \footnotesize \bibliographystyle{plain}
\section{Introduction} \label{sec: intro} Large Hadron Collider experiment (LHC) is now operating and reports many important results on new physics beyond the standard model (SM). Although positive signals have not been reported so far, those are expected to be found in near future, because the hierarchy problem of the SM strongly suggests the existence of new physics at the TeV scale or below. On the other hand, many new physics models have been theoretically proposed. Among those, the supersymmetric model is very attractive because it guarantees the stability of the Higgs mass to its radiative corrections and gives a clue to solve the hierarchy problem. In addition, the supersymmetry (SUSY) plays a crucial role to realize the grand unification of known gauge interactions of the SM at a certain high energy scale. Details of supersymmetric model, such as the mass spectrum of sparticles, depend highly on how SUSY is broken. So far, a variety of SUSY breaking mechanisms has been proposed~\cite{BookDrees}. Among those, the gauge mediation scenario~\cite{Dine:1993yw} attracts an attention, because it gives a solution to dangerous SUSY flavor problems. In this scenario, the breaking occurs at lower energy scale than those of other SUSY breaking scenarios, so that the superpartner of graviton, the gravitino, is likely to be the lightest supersymmetric particle (LSP). The gravitino mass is predicted to be in the range between ${\cal O}(10)$eV and ${\cal O}(1)$GeV. In this article, we focus on the low-scale gauge mediation model providing a gravitino with ${\cal O}(10)$eV mass. Such a scenario is well motivated because it is completely free from severe cosmological constraints~\cite{Feng:2010ij} such as Big-Bang Nucleosynthesis~\cite{Kawasaki:2008qe} and large scale structure formation of our universe~\cite{Viel:2005qj}. Collider signals of the low-scale gauge mediation scenario depend on what the next lightest superparticle (NLSP) is, which decays only into gravitino and its superpartner. Though there are many candidates for NLSP, we focus on the stau NLSP in this article, which is predicted in wide parameter region of the scenario. When the gravitino mass is of ${\cal O}(10)$eV, the stau NLSP decays into a $\tau$-lepton and a gravitino with the lifetime of $10^{-15}$--$10^{-11}$sec. The decay length (the lifetime times the speed of light) of the stau NLSP is therefore much shorter than the typical size of collider detectors, and the traditional supersymmetric signal, namely, multi-jets associated with missing energy and $\tau$-leptons, is expected at the LHC experiment. Such a signal is, however, generally predicted in various SUSY breaking scenarios. We show in this article that, even if the NLSP decays before reaching inner trackers of collider detectors, we can use the impact-parameter information about the decay products of the NLSP to study various properties of superparticles. In particular, the impact parameter is available for charged tracks caused by decay products of $\tau$-lepton at the stau NLSP decay. If the decay product of stau is found to have large impact parameter, it strongly suggests that the underlying SUSY breaking scenario is low-energy gauge mediation. Furthermore, the impact parameter is also utilized to precisely measure the spectrum of sparticles such as squark, neutralino, and stau masses. This is because two tau leptons produced by the cascade decay of a squark can be distinguished with each other by using the impact parameter. In addition, we may be able to determine the lifetime of the NLSP (i.e., stau in the present study) using the impact parameter distribution. When mass and lifetime of the stau NLSP are measured, it is possible to determine the gravitino mass assuming that the stau decays into gravitino and tau. The scale of SUSY breaking in the low-energy gauge mediation scenario is, therefore, obtained. It has been already shown that such studies can be easily performed once the $e^+e^-$ linear collider becomes available \cite{Matsumoto:2011fk}. Here, we consider the case of the LHC. We will see that the measurement of the mass spectrum as well as the determination of the lifetime of the NLSP can be performed at the LHC with the help of impact parameter information. This article is organized as follows. In the next section, we consider some properties of the stau NLSP in the low-energy gauge mediation scenario and discuss how the impact parameter from the NLSP decay is utilized in determinations of sparticle masses and NLSP lifetime. Our simulation framework is summarized in section \ref{sec: framework}, in which a representative point and several strategies to reduce backgrounds are shown. In section \ref{sec: results}, simulation results for the measurements of sparticle masses and lifetime of the stau NLSP are discussed. Section \ref{sec: summary} is devoted to summary of our studies. \section{Utilizing impact parameter} \label{sec: impact parameter} In this section, we discuss how the impact parameter is utilized in order to determine the mass spectrum of sparticles and the lifetime of stau NLSP. We first briefly review some properties of the stau NLSP and define the impact parameter. Then, we discuss basic strategies for the measurement of the mass spectrum and the lifetime of NLSP with the use of the impact parameter. \subsection{Stau NLSP and impact parameter} Since the LHC is a hadron collider, colored sparticles such as squarks and gluino are expected to be produced at first, which decay into the stau NLSP, the super partner of $\tau$-lepton, through several cascade channels. The stau NLSP then decays into a $\tau$-lepton and a gravitino with the following lifetime, \begin{eqnarray} \tau_{\tilde \tau} = 48 \pi M_{\rm pl}^2 \left(\frac{m_{3/2}^2}{m_{\tilde \tau}^5}\right) \simeq 5.9 \times 10^{-12}~[{\rm sec}] \times \left( \frac{m_{3/2}}{10{\rm eV}} \right)^2 \left( \frac{100{\rm GeV}}{m_{\tilde \tau}} \right)^5, \label{eq: lifetime} \end{eqnarray} where $M_{\rm pl} \simeq 2.4 \times 10^{18}$GeV, $m_{\tilde \tau}$, and $m_{3/2}$ are reduced Planck mass, stau mass, and gravitino mass, respectively. It turns out from above formula that the decay length (the lifetime $\times$ the speed of light) of the stau NLSP is estimated to be $\sim {\cal O}(100)\mu$m when gravitino and stau masses are $\sim 10$eV and $\sim 100$GeV, respectively. On the other hand, the decay length of $\tau$-lepton which is one of main backgrounds against the stau signal, is 87$\mu$m, so that the decay of stau NLSP into very light gravitino can be, in principal, detected if we can reduce SM backgrounds efficiently. Since the decay length of the stau NLSP is, at most, ${\cal O}(1)$mm in the parameter region of our interest, the stau NLSP decays before reaching the first pixel detector, which is located at 5cm (4cm) away from the beam line in the ATLAS detector~\cite{Aad:2009wy} (CMS detector~\cite{Bayatian:2006zz}). The lifetime of the stau NLSP is, as a result, difficult to be measured using methods usually applied to detect long-lived particles, such as methods by observing charged tracks \cite{Ishiwata:2008tp, Kaneko:2008re, Asai:2008sk, Asai:2011wy}. On the other hand, the lifetime of the stau NLSP may still be possible to be determined using the distribution of the impact parameter, which is obtained by $\tau$-jets from the stau NLSP decay. The impact parameter is defined as the shortest distance to the track from the interaction point. The positional resolution of the ATLAS detector along the longitudinal direction is $\Delta_L \sim 100 \mu$m, which is not good compared to that along the transverse direction, $\Delta_T \sim 10 \mu$m.\footnote{ Details of those performances are found in the section "Tracking" in Ref.~\cite{Aad:2009wy}. } Thus, we use the transverse impact parameter which is defined by \begin{eqnarray} d_I \equiv \left| {\bf x}_T^I - \frac{{\bf x}_T^I \cdot {\bf P}_T^I}{|{\bf P}_T^I|^2}{\bf P}_T^I \right|, \label{eq: impact parameter} \end{eqnarray} where ${\bf x}_T^I$ and ${\bf P}_T^I$ are transverse decay point of the $I$-th $\tau$-lepton and transverse momentum of the tau-jets from the $I$-th $\tau$-lepton decay, respectively. Note that the summation over the index $I$ should not be taken here. The above formula is used in our simulation studies, which will be presented in following sections. We expect that, at the LHC experiment, the distribution of the impact parameter $d_I$ is obtained by measuring the shortest distance (projected onto the transverse-plane) to the $\tau$-jet track from the interaction point. \subsection{Impact parameter for mass measurements} \label{sec: IP for masses} We next consider how the impact parameter is utilized in mass measurements of sparticles. At the LHC, colored sparticles such as gluino and squarks are expected to be produced copiously, and non-colored sparticles are then produced through cascade decays of the colored ones. The chain of the cascade decay is, for example, composed of following processes; First colored sparticle decays into a neutralino/chargino by emitting a quark which is observed as a jet. Next a neutralino/chargino decays into a slepton by emitting a lepton. Finally, a slepton decays into a LSP by again emitting a lepton. It is needless to say that the LSP passes through the detector without giving any signatures, which is, instead, observed as a missing energy. This chain (called "golden mode") is frequently used to measure the mass spectrum of sparticles in various supersymmetric scenarios by using several kinematical endpoints~\cite{Hinchliffe:1996iu, Hinchliffe:1998ys}. \begin{figure}[t] \begin{center} \includegraphics[origin=b, angle=0, width=8cm]{dchain.eps} \caption{\small Typical decay chain in the low-scale gauge mediation scenario.} \label{fig: cascade chain} \end{center} \end{figure} In the case of the low-scale gauge mediation scenario, we have a similar decay chain. One of the examples is shown in Fig.~\ref{fig: cascade chain}, where $\tilde{q}_R$, $\tilde{\chi}^0_1$, $\tilde{\tau}_1$, and $\tilde{G}$ are right-handed squark, lightest neutralino, lightest stau, and gravitino, respectively. The character $j$ denotes a jet which originates in a quark from the $\tilde{q}_R$ decay.\footnote {In the gluino production event, the gluino decays into a squark by also emitting a jet. Because the mass difference between gluino and squark is much smaller than that between squark and neutralino in the parameter region of our interest, we can discriminate between a (soft) jet from the gluino decay and a (hard) jet from the squark decay.} Following the terminology used in studies of the golden mode, we call $\tau$ from the $\tilde{\chi}^0_1$-decay $\tau^{\rm (near)}$ and that from the $\tilde{\tau}_1$-decay $\tau^{\rm (far)}$. Importantly, the impact parameter of $\tau^{\rm (far)}$ is expected to be larger than that of $\tau^{\rm (near)}$, which is of great help for the event reconstruction. In our analysis, we focus on the signal from the decay of a squark shown in Fig.~\ref{fig: cascade chain}. Kinematics of its decay chain is, as a result, governed by following four sparticle masses; the masses of squark ($m_{\tilde q}$), lightest neutralino ($m_{\tilde{\chi}_1^0}$), lightest stau ($m_{\tilde{\tau}_1}$), and gravitino ($m_{3/2}$). Since the gravitino mass is of the order of 10eV, only the upper bound on the mass is expected to be obtained. On the other hand, the existence of large impact parameters in signal events strongly suggests the (low-scale) gauge mediation scenario. We therefore perform our analysis with simply postulating that the gravitino mass is much smaller than those of other sparticles, namely, with treating the gravitino as a massless particle. Three independent kinematical endpoints are then enough to determine the mass spectrum of sparticles. With the help of the impact parameter, many kinematical variables are now available. Among those, we use the invariant mass between two $\tau$-leptons ($M_{\tau^{\rm (near)} \tau^{\rm (far)}}$), that between jet and near $\tau$-lepton ($M_{j \tau^{\rm (near)}}$), and the $M_{T2}$ variable from leading two jets ($M_{T2,jj}$). The upper limit on the invariant mass $M_{\tau^{\rm (near)} \tau^{\rm (far)}}$ is given by \begin{eqnarray} M_{\tau^{\rm (near)} \tau^{\rm (far)}}^{\rm max} = m_{\tilde{\chi}_1^0} \sqrt{1 - m^2_{\tilde{\tau}_1}/m^2_{\tilde{\chi}_1^0}}, \label{eq: mass_tautau} \end{eqnarray} where the mass of $\tau$-lepton is set to be zero in above formula. In addition, the upper limit on the distribution of the invariant mass $M_{j \tau^{\rm (near)}}$ is given by the following formula, \begin{eqnarray} M_{j \tau^{\rm (near)}} ^{\rm max} = m_{\tilde{q}} \sqrt{ \left(1 - m^2_{\tilde{\chi}_1^0}/m^2_{\tilde{q}}\right) \left(1 - m^2_{\tilde{\tau}_1}/m^2_{\tilde{\chi}_1^0}\right)}. \label{eq: mass_jtau} \end{eqnarray} The last kinematical variable used in our analysis is the $M_{T2}$ variable~\cite{Lester:1999tx} constructed from highest two jets; denoting the momenta of highest two jets as ${\bf p}$ and ${\bf p}'$, we define \begin{eqnarray} M_{T2,jj}(m_{\rm miss}) = \min_{{\bf k}_T + {\bf k}_T^{\prime} = {\psll}^{\rm eff}_T} \left[ \max \left\{ M_T({\bf p}_T, {\bf k}_T), M_T({\bf p}'_T, {\bf k}'_T) \right\} \right], \end{eqnarray} where $M_T$ is the transverse mass and $m_{\rm miss}$ is the ``test mass.'' Here, because we construct the $M_{T2}$ variable only from highest two jets, ${\psll}^{\rm eff}_T$ should be understood as the vector sum of transverse momenta of all the activities other than highest two jets and missing momentum $\psl_T$: ${\psll}^{\rm eff}_T = \psl_T + \sum_i {\bf p}_{\tau-jet~i} + \sum_i {\bf p}^{\prime}_{\tau-jet~i}$. The upper limit of this variable is then given by \begin{eqnarray} M_{T2,jj}^{\rm max}(m_{\rm miss}) = \frac{m^2_{\tilde{q}} - m^2_{\tilde{\chi}_1^0}}{2m_{\tilde{q}}} + \sqrt{ \left( \frac{m^2_{\tilde{q}} - m^2_{\tilde{\chi}_1^0}}{2m_{\tilde{q}}} \right)^2 + m_{\rm miss}^2 }. \label{eq: MT2_jj} \end{eqnarray} Using three kinematical endpoints given in eqs.(\ref{eq: mass_tautau}), (\ref{eq: mass_jtau}) and (\ref{eq: MT2_jj}), we fit the sparticle masses $m_{\tilde q}$, $m_{\tilde{\chi}_1^0}$, and $m_{\tilde{\tau}_1}$. We will see that, though the cascade chain always involves $\tau$s as lepton emissions, the spectrum can be determined accurately because of information about the impact parameter. \subsection{Impact parameter for stau lifetime measurement} \label{sec: IP for lifetime} Information about the lifetime of the stau NLSP is imprinted in the distribution of the impact parameter. The impact parameter, however, depends not only on the lifetime of the stau NLSP but also on its mass and velocity. With the use of the strategy discussed in previous subsection, the stau mass is measured precisely. On the other hand, since the gravitino produced from the stau decay cannot be detected, the velocity cannot be determined on event-by-event basis, which makes it difficult to determine the lifetime using the impact parameter distribution. Once the mass spectrum of the superparticles are known, however, we expect to acquire information about the velocity distribution of the $\tilde{\tau}_1$ in supersymmetric events. In the present case (where the mass spectrum of a simple gauge mediation model is assumed), we may understand that the underlying scenario is indeed the low-scale gauge mediation from the experimentally measured mass spectrum as well as the confirmation of the existence of long-lived stau. Even if we cannot specify the complete structure of the underlying model, we may still be able to measure the masses of superparticles which are most important for the determination of the velocity distribution of $\tilde{\tau}_1$ (i.e., the masses of $\tilde{q}$, $\tilde{\chi}^0_1$, and $\tilde{\tau}_1$), as we have discussed in the previous section. Then, once those information becomes available, one will be able to obtain the velocity distribution with, for example, Monte Carlo simulation. In our analysis, we assume that the velocity distribution of $\tilde{\tau}_1$ can be understood once the superparticles are discovered. The detailed study of the methods of determining the velocity distribution is beyond the scope of this paper, so we simply assume that the averaged velocity of the produced $\tilde{\tau}_1$ can be obtained with some accuracy and determine the lifetime using the $\tilde{\tau}_1$. Although a better determination of the lifetime of $\tilde{\tau}_1$ may be possible if we can somehow obtain and use the information about the velocity distribution of $\tilde{\tau}_1$, we can still have a relatively good determination of the lifetime using the averaged velocity as we will describe. The procedure to measure the lifetime of the stau NLSP is therefore the following. \begin{enumerate} \item[(i)] We first assume that the averaged velocity of the stau NLSP, denoted as $\bar{\beta}_{\tilde{\tau}_1}$, is somehow understood. Then, we generate $\tilde{\tau}_1$ with the fixed velocity $\bar{\beta}_{\tilde{\tau}_1}$ and make signal templates of the distribution of the transverse impact parameter $d_I$. The template is prepared for wide range of stau lifetime. The production angle of the stau NLSP is assumed to be isotropic in generating the events for the template. \item[(ii)] Distribution of the impact parameter expected at the LHC experiment is obtained by using Monte Carlo simulation. \item[(iii)] Comparing the templates obtained in the (i) with the actual distribution obtained in (ii), we study how well we can constrain the lifetime of $\tilde{\tau}_1$ by $\chi^2$--analysis. By varying the value of $\bar{\beta}_{\tilde{\tau}_1}$ used in making the templates, we also discuss the uncertainty related to the determination of the velocity distribution. \end{enumerate} Here, we have a few comments on the above method. First comment is on the effect of gluino production. The mass difference between gluino and squark is much smaller than that between squark and neutralino in the parameter region of our interest. In addition, it will be possible to select signal events with the desirable squark decay chain ($\tilde{q} \to \tilde{\chi}^0_1 \to \tilde{\tau}_1 \to \tilde{G}$) by applying appropriate kinematical cuts. Thus, the averaged value of the boost factor is expected to depend weakly on the gluino mass. We have checked this statement quantitatively by simulating signal events with several choices of the gluino mass. Second comment is on how the transverse impact parameter $d_I$ depends on the production angle of the stau NLSP. One might worry if we may compare the actual impact parameter distribution with theoretical templates obtained by postulating isotropic distribution of the production angle. We have checked that this potential problem can be solved by only using $\tau$-jets with small pseudo-rapidity. \section{Simulation framework} \label{sec: framework} Before showing our results, we summarize the framework of our simulation study. We first mention a representative point and simulation tools used in the study. Next we discuss the strategy to suppress combinatorial backgrounds of signal events caused by the existence of two decay chains. We finally consider the SM backgrounds and discuss kinematical cuts used to reduce those backgrounds. \subsection{Representative point \& simulation tools} The representative point used in our simulation study has been chosen by adopting the minimal model of the gauge mediation symmetry breaking \cite{Dine:1993yw} with the following underlying parameters; the SUSY breaking scale ($\varLambda = 30$TeV), the messenger mass scale ($M_{\rm mess} = 300$TeV), the number of SU(5) messenger fields ($N_{\bf 5} = 5$), and the ratio of vacuum expectation values of two Higgs fields ($\tan\beta = 15$). The ISAJET package~\cite{ISAJET} is used in order to calculate the spectrum and branching fractions of sparticles. Resultant masses and branching fractions of sparticles relevant to the study are summarized in Table~\ref{tab: point} for the case of the gravitino mass of 9.7eV which corresponds to the decay length of the stau of 500$\mu$m. It can be seen that the model is consistent with current LHC data~\cite{Kats:2011qh}. In our analysis, we consider the LHC experiment with the center of mass energy of $\sqrt{s}=14{\rm TeV}$. Then, the signal cross section, which is the sum of the cross sections for gluino and squark productions, is estimated to be \begin{eqnarray} \sigma_{\tilde{g} \tilde{g}} = 0.129{\rm pb}, \quad \sigma_{\tilde{q} \tilde{g}} = 0.922{\rm pb}, \quad \sigma_{\tilde{q} \tilde{q}} = 0.879{\rm pb}. \end{eqnarray} We focus on the decay chain involving a right-handed squark, as shown in Fig.~\ref{fig: cascade chain}. As a result, a typical signal event consists of two energetic jets, four $\tau$-leptons ($\tau$-jets or leptons), and a missing energy in the transverse direction. \begin{table*}[t] \begin{center} \begin{tabular}{c|clll} & Mass (GeV) & \multicolumn{2}{c} {Branching fractions} \\ \hline $\tilde{g}$ & 1096.6 & Br($\tilde{g} \to \tilde{q} q$) = 0.89. \\ $\tilde{u}_L$ & 951.1 & Br($\tilde{u}_L \to \tilde{\chi}^\pm_2 d$) = 0.34, & Br($\tilde{u}_L \to \tilde{\chi}^\pm_1 d$) = 0.32, & \\ & & Br($\tilde{u}_L \to \tilde{\chi}^0_4 u$) = 0.18, & Br($\tilde{u}_L \to \tilde{\chi}^0_2 u$) = 0.15. \\ $\tilde{u}_R$ & 922.0 & Br($\tilde{u}_R \to \tilde{\chi}^0_1 u$) = 0.96. \\ $\tilde{\chi}^0_1$ & 197.3 & Br($\tilde{\chi}^0_1 \to \tilde{\tau}^\pm_1 \tau^\mp$) = 0.35, & Br($\tilde{\chi}^0_1 \to \tilde{e}^\pm_R e^\mp$) = 0.32, \\ & & Br($\tilde{\chi}^0_1 \to \tilde{\mu}^\pm_R \mu^\mp$) = 0.32. \\ $\tilde{e}_R$ & 130.0 & \multicolumn{2}{l} {Br($\tilde{e}_R \to \tilde{\tau}^\pm_1 \tau^\mp e$) $\simeq$ 1.00.} \\ $\tilde{\tau}_1$ & 126.2 & Br($\tilde{\tau}_1 \to \tau \tilde{G}$) = 1.00. \\ \hline \end{tabular} \caption{\small Masses and branching fractions of sparticle in our representative point.} \label{tab: point} \end{center} \end{table*} For parton-level event generation and hadronization, we employ the HERWIG code~\cite{HERWIG, HERWIGSUSY}. Generated events are passed through the PGS code~\cite{PGS4} for simulating detector effects. Fake $\tau$-jets from QCD processes and heavy meson decays are involved in the study. For tau-jets, we smear the transverse vertex position of the parton using Gaussian distribution with the error $\Delta d_I = 10\mu$m. \subsection{Charge subtraction method} Background reduction is the most important task in our analysis to determine the mass spectrum of sparticles, because all kinematical endpoints do not have sharp edge structures due to the energy leakage by $\nu_{\tau}$ emissions from $\tau$-decays. Expected backgrounds in those measurements are as follows: (i) A number of fake $\tau$-jets are expected at the hadron collider. (ii) A signal event results in multiple $\tau$-leptons, and hence there are combinatorial backgrounds in the analysis involving $\tau$-jets. In order to reduce these backgrounds, we adopt the method of the charge subtraction. We expect four $\tau$-leptons (two $\tau^+$ and two $\tau^-$) in one event. We therefore have three ways to pair the $\tau$ leptons, $(\tau^\pm, \tau^\mp)_1$, $(\tau^\pm, \tau^\mp)_2$, and $(\tau^\pm, \tau^\pm)$. In each event (involving four $\tau$s), we take the data using the method $(\tau^\pm, \tau^\mp)_1 + (\tau^\pm, \tau^\mp)_2 - (\tau^\pm, \tau^\pm)$, then the wrong opposite-sign pairing is expected be canceled by the subtraction of the same-sign pairing. This method works very well when $\tau$-leptons are produced in the process $\tilde{\chi}^0 \to \tilde{\tau} \to \tilde{G}$; this is due to the fact that $\tilde{\chi}^0$ decays into $\tilde{\tau}^+\tau^-$ and $\tilde{\tau}^-\tau^+$ with equal probability. In addition, this method can be applied to the determination of the $M_{j \tau^{\rm (near)}}$-endpoint. We simply collect $(j, \tau, \tau)$ events using the charge subtraction. In each paring, $\tau^{\rm (near)}$ is identified as the $\tau$ lepton which has a shorter impact parameter. It should be also noted that the charge subtraction method can reduce backgrounds from fake $\tau$-jets from QCD processes, because the QCD fake events are charge-blind in a good approximation at high energy processes. \subsection{Kinematical cuts to reduce $t\bar{t}$ backgrounds} \label{sec: selection cuts} The most serious SM background for our study is the $t\bar{t}$ production. Thus, we concentrate on this background. In order to reduce this background, we impose following kinematical cuts: \begin{itemize} \item Large missing transverse energy, $\ptsl >$ 150GeV. \item At least, four leptons, $e$, $\mu$ with $p_T >$ 20GeV or $\tau$-jet with $p_T >$ 25GeV. \item Two hard jets, $j_1$($j_2$) with $p_T >$ 200(100)GeV (and no b-jets). \end{itemize} Here, we use the label, $i = 1$ or 2, for a jet ($j_i$) in decreasing order of $p_T$. We also take account of both electrons and muons, because $\tau$-lepton often decays leptonically and selectron and smuon decay into gravitino by emitting electron and muon directly. The requirement for jets to have $p_T >$ 200(100)GeV is very important to reduce the $t\bar{t}$ background, because jets with $p_T > m_W$ ($m_W$ is the mass of weak gauge boson) from top quark decays are rather rare. In Table~\ref{table: cut flaw}, we summarize the cut flow in our simulation study with assuming that the integrated luminosity is 100fb$^{-1}$, where we take the gravitino mass of 9.7eV again. It is possible to apply tighter kinematical cuts for further reductions of the backgrounds. For instance, the requirement $\ptsl >$ 200GeV, $p_T(j_1) >$ 250GeV and $p_T(j_2) >$ 150GeV in addition to the basic kinematical cuts shown in Table~\ref{table: cut flaw} will reduce 80\% of the $t\bar{t}$ background (we have, as a result, $\sim$400 $t\bar{t}$ events), while this also reduces 30\% of SUSY signals (we have, as a result, $\sim$28000 SUSY events). \begin{table} \begin{center} \begin{tabular}{l|rr} Selection cut & SUSY & $t\bar{t}$ \\ \hline (0) Generated events & 273,600 & 49,610,000 \\ (1) $\ptsl >$ 150GeV & 190,504 & 1,644,160 \\ (2) \# of leptons $\geq$ 4 & 70,641 & 55,537 \\ (3) \# of b-jets = 0 & 56,228 & 41,673 \\ (4) $j_1$ with $p_T >$ 200GeV & 49,819 & 5,915 \\ (5) $j_2$ with $p_T >$ 100GeV & 41,007 & 1,984 \\ \hline \end{tabular} \caption{\small Numbers of signal (SUSY) and background ($t\bar{t}$) events after applying kinematical cuts with ${\cal L} =$ 100fb$^{-1}$. All SUSY processes are included in our event generation.} \label{table: cut flaw} \end{center} \end{table} \section{Simulation results} \label{sec: results} We are now in position to present several results of our simulation study, which are obtained based on arguments in previous sections. We first show the results for measurements of sparticle masses, and discuss how accurately these masses can be determined at the LHC. We next show that the lifetime of the stau may be determined by using the distribution of the transverse impact parameter of $\tau$-jets from the stau NLSP decay. We estimate how accurately the lifetime can be determined. \subsection{Sparticle masses} \label{sec: results mass} The strategy to determine the mass spectrum is the use of kinematical endpoints of several variables. We study how the endpoints behaves using generated events which pass through the basic cuts discussed in section \ref{sec: selection cuts}. For the simulation study of sparticle mass measurement, the decay length of the stau NLSP is set to be 500$\mu$m (corresponding to the gravitino mass of 9.7eV) \subsubsection{Endpoint on $M_{\tau\tau}$} The first kinematical variable used in the analysis for the mass spectrum is the invariant mass of two $\tau$-leptons in the decay chain of a squark. In Fig.~\ref{fig: endpoints} (upper panel), the distribution of the invariant mass ($M_{\tau\tau}$) after applying the charge subtraction method is shown. We can see a clear edge at $M_{\tau\tau} \simeq$ 150 GeV. In order to extract the location of the endpoint, we use the following fitting function, \begin{equation} f(M_{\tau\tau}) = \left\{ \begin{array}{ll} A (M_{\tau\tau} - M_{\tau\tau}^{\rm fit}) + C & : M_{\tau\tau} < M_{\tau\tau}^{\rm fit} \\ B (M_{\tau\tau} - M_{\tau\tau}^{\rm fit}) + C & : M_{\tau\tau} > M_{\tau\tau}^{\rm fit} \end{array} \right., \label{eq: fitting function} \end{equation} where $A$, $B$, $M_{\tau\tau}^{\rm fit}$, and $C$ are parameters to fit the shape of the distribution around the endpoint. With the use of this bilinear function for the fitting, the location of the endpoint is determined to be $M_{\tau \tau}^{\rm fit} = 151.2 \pm 14.5$GeV. Notice that the underlying value (the input value on the simulation) is 151.6 GeV. \begin{figure}[p] \begin{center} \includegraphics[origin=b, angle=0, width=8.4cm]{Mtautau.eps} \\ \vspace{0.5cm} \includegraphics[origin=b, angle=0, width=8.4cm]{Mjtau.eps} \\ \vspace{0.5cm} \includegraphics[origin=b, angle=0, width=8.4cm]{MT2jj.eps} \caption{\small (Upper panel) Distribution of the invariant mass between two tau-jets, $M_{\tau\tau}$. (Middle panel) Distribution of the invariant mass between hard jet and near $\tau$-jet, $M_{j\tau^{\rm (near)}}$. (Lower panel) Distribution of the $M_{T2,jj}$ variable defined in Eq.~(\ref{eq: MT2_jj}) with $m_{\rm miss}$ being zero.} \label{fig: endpoints} \end{center} \end{figure} \subsubsection{Endpoint on $M_{q \tau^{\rm (near)}}$} Second kinematical variable we use is the invariant mass between $\tau$-lepton and jet emitted by the decay of a squark. Using information about the impact parameter, it is possible to distinguish near and far $\tau$-leptons with high efficiency. For each ($\tau$,~$\tau$)-pair, we identify the $\tau$-jet whose track has a larger impact parameter than the other as the far tau-jet, while the $\tau$-jet with a smaller impact parameter is regarded as the near $\tau$-jet. After this identification, both combinations of $(j_1, \tau^{\rm (near)})$ and $(j_2,~\tau^{\rm (near)})$ are used to calculate $M_{q \tau^{({\rm near})}}$. In the analysis, we also require that the pair of two tau-jets should satisfy $M_{\tau\tau} < M_{\tau \tau}^{\rm fit} =$ 155.72GeV in order to reduce fake-QCD and combinatorial backgrounds. The distribution of the invariant mass ($M_{q \tau^{\rm (near)}}$) after applying the charge subtraction method is shown in Fig.~\ref{fig: endpoints} (middle panel). The endpoint is, again, fitted by using the bilinear function given in Eq.~(\ref{eq: fitting function}). It then turns out that the location of the endpoint on $M_{q \tau^{\rm (near)}}$ is $M_{q \tau^{\rm (near)}}^{\rm fit} = 700.0 \pm 0.1$GeV. Notice that the underlying value is now 692.3GeV. \subsubsection{Endpoint on $M_{T2,jj}$} The last kinematical variable is $M_{T2,jj}$ defined by two hard jets as visible particles, as mentioned in section \ref{sec: IP for masses}. In our analysis, we take the test mass ($m_{\rm miss}$) in Eq.~(\ref{eq: MT2_jj}) to be zero, so that the endpoint of this kinematical variable gives $M_{T2,jj}^{\rm max}(0) = (m_{\tilde{q}}^2 - m_{\tilde{\chi}_1^0}^2) / m_{\tilde{q}}$. The distribution of $M_{T2,jj}(0)$ is shown in Fig.~\ref{fig: endpoints} (lower panel). It can be seen that a very clear endpoint exists at $M_{T2,jj}(0) \simeq 870$GeV. As in the cases of previous kinematical variables, we fit the shape of the distribution around the endpoint by the bilinear function. The endpoint of the distribution is then obtained as $M_{T2,jj}^{\rm fit}(0) = 875.1 \pm 6.7$GeV (the underlying value is 879.8GeV). \subsubsection{Mass determination} When the gravitino mass is neglected, the masses of sparticles, $m_{\tilde{q}}$, $m_{\tilde{\chi}_1^0}$, and $m_{\tilde{\tau}_1}$, are determined by three kinematical endpoints of the variables, $M_{\tau \tau}^{\rm max}$, $M_{q \tau^{\rm (near)}}^{\rm max}$, and $M_{T2,jj}^{\rm max}(0)$. Using analytic expressions for the these endpoints shown in eqs.(\ref{eq: mass_tautau}), (\ref{eq: mass_jtau}), and (\ref{eq: MT2_jj}), the masses of sparticles are determined by minimizing the following $\chi^2$ function, \begin{equation} \chi_{\rm m}^2 = \left[ \frac{M_{\tau \tau}^{\rm max} - M_{\tau \tau}^{\rm fit}} {\Delta M_{\tau \tau}^{\rm fit}} \right]^2 + \left[ \frac{M_{q \tau^{\rm (near)}}^{\rm max} - M_{q \tau^{\rm (near)}}^{\rm fit}} {\Delta M_{q \tau^{\rm (near)}}^{\rm fit}} \right]^2 + \left[ \frac{M_{T2,jj}^{\rm max}(0) - M_{T2,jj}^{\rm fit}(0)} {\Delta M_{T2,jj}^{{\rm fit}}(0)} \right]^2, \label{eq:chi2_mass} \end{equation} where $M^{\rm fit}$ denotes the center value of the measured endpoint, and $\Delta M^{\rm fit}$ is its (statistical) error. After minimizing $\chi^2_{\rm m}$ by varying the input values, $m_{\tilde{q}}$, $m_{\tilde{\chi}_1^0}$, and $m_{\tilde{\tau}_1}$, we obtain following results; the right-handed squark mass is $m_{\tilde{q}} = 915.9 \pm 6.4$GeV (the true value is 922.0GeV), the lightest neutralino mass is $m_{\tilde{\chi}^0_1} = 193.4 \pm 19.5$GeV (the true value is 197.3GeV), and the lightest stau mass is $m_{\tilde{\tau}_1} = 120.5 \pm 18.1$GeV (the true value is 126.2GeV). Here, the gravitino mass is taken to be zero (i.e., negligibly small). \subsection{Lifetime of the stau NLSP} \label{sec: results lifetime} The strategy to determine the lifetime of the stau NLSP is the use of the distribution of the transverse impact parameter. After showing the distribution for several input values of $c\tau_{\tilde \tau}$, we discuss how accurately the lifetime can be determined at the LHC. \subsubsection{Distribution of the impact parameter} Distribution of the transverse impact parameter ($d_I$) obtained from hadronically decays of $\tau$-leptons is shown in Fig.~\ref{fig: di} with the use of generated events which are passed through the kinematical cuts discussed in previous section. Four distributions are shown in this figure with choices of the decay length of the stau NLSP to be $c\tau_{\tilde \tau} =$ 1, 100, 500, and 900$\mu$m, respectively. It is clearly seen that the distribution depends on the decay length of the stau NLSP as expected. \begin{figure}[t] \begin{center} \includegraphics[origin=b, angle=0, width=8.4cm]{di.eps} \caption{\small Distribution of the transverse impact parameter of $\tau$-jets after applying the basic kinematical cuts in previous section. We show distributions in the SUSY model with $c\tau_{\tilde \tau} =$ 1, 100, 500 and 900$\mu$m, respectively, and also show the distribution in $t\bar{t}$ production.} \label{fig: di} \end{center} \end{figure} It is also seen that a number of events are found in inner bins with $d_I \ll c\tau_{\tilde \tau}$ and that a broad tail-structure exists in the region $d_I \gg c\tau_{\tilde \tau}$. In fact, in both regions, backgrounds are expected to contribute to the distribution. In small $d_I$ region, the distribution is dominated by background $\tau$-jets such as QCD-originated fake ones. Though most of those backgrounds do not have finite $d_I$ at the parton-level, the backgrounds acquire finite values of $d_I$ at the detector-level because of the limited resolution for the vertexing. On the other hand, in the region of $d_I \gg c\tau_{\tilde \tau}$, backgrounds come from decays of heavy hadrons such as $D$ or $B$ mesons. Those hadrons sometimes produce fake $\tau$-jets after flying a sizable distance. In order to eliminated these backgrounds efficiently, we vary and optimize the upper and lower endpoints of the bins which are used for the $\chi^2$ analysis as we change $c\tau_{\tilde \tau}^{\rm (test)}$ (where $c\tau_{\tilde \tau}^{\rm (test)}$ is the test value of the decay length used to generate a template of $d_I$ distribution). \subsubsection{Lifetime estimation} According to the strategy to estimate the lifetime discussed in section \ref{sec: IP for lifetime}, we now study how well we can constrain the lifetime of $\tilde{\tau}_1$. First, we use the template generated with the true value of the averaged velocity of the stau NLSP, which is $\bar{\beta}_{\tilde{\tau}} = 0.88 c$ in our representative point. In order to see how the result depends on the underlying value of the lifetime of $\tilde{\tau}_1$, here we use several values of $c\tau_{\tilde \tau}$ in generating the impact-parameter distribution. In addition, in preparing the templates for $d_I$ distributions, TAUOLA library~\cite{TAUOLA} is used to simulate the $\tau$ decay event, which enable us to deal with chirality and finite lifetime of $\tau$-leptons. The range of the test value is taken to be 10$\mu$m--1100$\mu$m every 10$\mu$m. With the use of the templates, we perform $\chi^2$-analysis to determine the lifetime. In our analysis, only $\tau$-jets satisfying $0.5 \times c\tau_{\tilde \tau}^{\rm (test)} < d_I < 2.0 \times c\tau_{\tilde \tau}^{\rm (test)}$ with small pseudo-rapidity, $|\eta| < 1.0$, are adopted. The size of bin used in each distribution is set to be $0.15 \times c\tau_{\tilde \tau}^{\rm (test)}$ and, as a result, we have 10 bins in total. The $\chi^2$ variable to estimate the lifetime of the stau NLSP is therefore given by \begin{equation} \chi^2({c\tau_{\tilde \tau}^{\rm (test)}}) \equiv \sum_{i = 1}^{10} \left[ \frac{N^{\rm (th)}_i(c\tau_{\tilde \tau}^{\rm (test)}) - N^{\rm (exp)}_i}{\Delta N_i} \right]^2, \end{equation} where $N^{\rm (th)}_i(c\tau_{\tilde \tau}^{\rm (test)})$ denotes the number of signals in the $i$-th bin obtained by the template for a given $ c\tau_{\tilde \tau}^{\rm (test)}$, while $N^{\rm (exp)}_i$ is the one obtained by using generated events. We only involves the statistical error as $\Delta N_i \equiv \sqrt{N^{\rm (exp)}_i}$ in the analysis. The degrees of the freedom in this $\chi^2$-test is therefore $(10 - 1) = 9$, and the hypothesis is excluded at 95\% C.L. when $\chi^2 > 16.92$. \begin{figure}[t] \begin{center} \includegraphics[origin=b, angle=0,width=8.4cm]{chi2_088.eps} \caption{\small $\chi^2$-values as a function of the test lifetime (the test decay length) of the stau NLSP, $c\tau_{\tilde \tau}^{\rm (test)}$, for the underlying values of $c\tau_{\tilde \tau} =$ 100 (red, solid), 300 (green, dashed), 500 (blue, dotted), 700 (violet, dot-dashed), and 900 (cyan, dot-dot-dashed) $\mu$m. Here, the averaged velocity is taken to be $\bar{\beta}_{\tilde{\tau}}=0.88$.} \label{fig: chi2} \end{center} \end{figure} Resultant $\chi^2$-values as a function of the test lifetime (the test decay length) of the stau NLSP, $c\tau_{\tilde \tau}^{\rm (test)}$, for underlying values of $c\tau_{\tilde \tau} =$ 100 (red, solid), 300 (green, dashed), 500 (blue, dotted), 700 (violet, dot-dashed), and 900 (cyan, dot-dot-dashed) $\mu$m are shown in Fig.~\ref{fig: chi2}. From these results, the lifetime (the decay length) of the stau NLSP in each case is determined at 95\% C.L. to be \begin{eqnarray} \begin{array}{rcll} 50\mu{\rm m} & \lesssim~c\tau_{\tilde \tau}~\lesssim & 110\mu{\rm m} & {\rm (underlying}~c\tau_{\tilde \tau} = 100\mu{\rm m}), \\ 240\mu{\rm m} & \lesssim~c\tau_{\tilde \tau}~\lesssim & 330\mu{\rm m} & {\rm (underlying}~c\tau_{\tilde \tau} = 300\mu{\rm m}), \\ 410\mu{\rm m} & \lesssim~c\tau_{\tilde \tau}~\lesssim & 540\mu{\rm m} & {\rm (underlying}~c\tau_{\tilde \tau} = 500\mu{\rm m}), \\ 570\mu{\rm m} & \lesssim~c\tau_{\tilde \tau}~\lesssim & 800\mu{\rm m} & {\rm (underlying}~c\tau_{\tilde \tau} = 700\mu{\rm m}), \\ 810\mu{\rm m} & \lesssim~c\tau_{\tilde \tau}~\lesssim & 1060\mu{\rm m} & {\rm (underlying}~c\tau_{\tilde \tau} = 900\mu{\rm m}). \\ \end{array} \end{eqnarray} We can see that, if the correct value of $\bar{\beta}_{\tilde{\tau}}$ is used, the analysis based on the impact parameter distribution gives a good estimate of the lifetime (the decay length) of the stau NLSP with accuracy of about 30\% when $c\tau_{\tilde \tau} > 100\mu$m. If more precise information about the velocity distribution of $\tilde{\tau}_1$ is available, better estimate of the lifetime may be obtained. So far, we have neglected the uncertainty arising from the determination of the velocity distribution of $\tilde{\tau}_1$. As we have mentioned, the detailed study of the uncertainty in the velocity distribution is beyond the scope of this paper. In our analysis, however, we estimated the uncertainty of the $\bar{\beta}_{\tilde{\tau}}$ determination related to the errors in mass measurements and also to the production process in order to demonstrate that the $\bar{\beta}_{\tilde{\tau}}$ can be obtained with some accuracy. First, in order to study the effects of the errors in the mass measurements, we generated the events using the different sparticle mass spectrum from our representative point. Here, we used the mass spectrum predicted from the simple gauge mediation model except for $\tilde{\tau}_1$ because we found that the error of $m_{\tilde{\tau}_1}$ is the largest among the reconstructed masses in the previous subsection. Then we generated the full SUSY events by varying $m_{\tilde{\tau}_1}$ by $\pm 20$GeV, and found that the value of $\bar{\beta}_{\tilde{\tau}}$ changes by $\sim 0.01$. In addition, the dominant SUSY process may not be well understood in the actual situation. If so, it may be reasonable to estimate $\bar{\beta}_{\tilde{\tau}}$ by assuming the process we use for our analysis, which is the process shown in Fig.\ \ref{fig: cascade chain}. We generated events corresponding to such a process (using the correct mass relation). Then, $\bar{\beta}_{\tilde{\tau}}$ is found to be $\sim 0.93$. Thus, a relatively larger uncertainty of $\Delta\bar{\beta}_{\tilde{\tau}}\sim 0.05$ is expected if the dominant SUSY process cannot be understood. To see how this affects the determination of the lifetime, we calculate the $\chi^2$ variable using the template with $\bar{\beta}_{\tilde{\tau}}=0.83$ and $0.93$. The results are shown in Figs.~\ref{fig:chi2_0.83} and \ref{fig:chi2_0.93}. We can see that error related to the uncertainty to the averaged velocity is $\sim 50-100\mu{\rm m}$ if $\Delta\bar{\beta}_{\tilde{\tau}}\sim 0.05$. Even with such an uncertainty, we can still have a relatively good determination of $\tau_{\tilde \tau}$. Thus, the impact parameter will be a powerful tool to measure the lifetime at the LHC. \begin{figure}[t] \begin{center} \includegraphics[origin=b, angle=0,width=8.4cm]{chi2_083.eps} \caption{\small Same as Fig.~\ref{fig: chi2}, except for $\bar{\beta}_{\tilde{\tau}}=0.83$.} \label{fig:chi2_0.83} \end{center} \begin{center} \includegraphics[origin=b, angle=0,width=8.4cm]{chi2_093.eps} \caption{\small Same as Fig.~\ref{fig: chi2}, except for $\bar{\beta}_{\tilde{\tau}}=0.93$.} \label{fig:chi2_0.93} \end{center} \end{figure} Finally, we consider how well we can estimate the gravitino mass if we assume that $\tilde{\tau}_1$ decays into the gravitino and $\tau$-lepton. In the case with the underlying gravitino mass of 9.7eV, we have shown that the lifetime of the stau is estimated between 410$\mu$m and 540$\mu$m for $\bar{\beta}_{\tilde{\tau}}=$~0.88. In addition, the uncertainty related to $\bar{\beta}_{\tilde{\tau}}$ is estimated to be about 40$\mu$m, as one can see from Figs.~\ref{fig:chi2_0.83} and~\ref{fig:chi2_0.93}. Then, by using the measured stau mass $m_{\tilde{\tau}} = 120.5 \pm 18.1$GeV and the center value of the estimated NLSP lifetime $c\tau_{\tilde{\tau}} = 475\mu$m, we obtain $m_{3/2} = 8.3\pm 1.8\pm 0.8\pm 1.0$eV, where the errors originate in the uncertainties of $m_{\tilde{\tau}}$, $c\tau_{\tilde{\tau}}$ and $\bar{\beta}_{\tilde{\tau}}$, respectively. \section{Summary} \label{sec: summary} We have proposed a method to determine the mass spectrum of sparticles and the lifetime of the stau NLSP at the LHC when the mass of the gravitino LSP is of the order of 10eV. Though the decay length of the stau NLSP is very short, which is of the order of 100--1000$\mu$m, it is still possible to deeply study the model by utilizing the transverse impact parameter of tau-jets from the decay of the stau NLSP. We have first discussed the mass measurement of sparticles using a typical cascade decay chain of a squark shown in Fig.~\ref{fig: cascade chain}. This SUSY event involves, at least, four $\tau$-leptons, which makes it difficult to analyze the signal because of combinatorial backgrounds. Information about the impact parameter of the tau-jet, however, resolves the problem, and we have shown that the mass spectrum of sparticles can be determined accurately through the kinematical endpoints of $M_{\tau \tau}$, $M_{j \tau^{ \rm (near)} }$, and $M_{T2,jj}$. We have also discussed the determination of the lifetime of the stau NLSP using the distribution of the transverse impact parameter. The impact parameter depends not only on the lifetime but also on the velocity of the stau NLSP. We have therefore developed a strategy to estimate the velocity by utilizing a simulation with information about the mass spectrum obtained in the previous stage. We have shown that, if the velocity distribution of $\tilde{\tau}_1$ is somehow understood, the lifetime of the stau NLSP is determined with the accuracy of about 30\% as far as its decay length is larger than $\sim 100 \mu$m. Thus, if the underlying model of the SUSY breaking is low-scale gauge mediation with the gravitino mass of ${\cal O}(10)$eV, the LHC may have a chance to acquire some information about the gravitino mass. \section*{Acknowledgments} This work is supported by Grant-in-Aid for Scientific research from the Ministry of Education, Science, Sports, and Culture (MEXT), Japan, Nos.\ 21740174 \& 22244031 (S.M.), No.\ 22540263 (T.M.), and No.\ 22244021 (M.A., S.M. and T.M.), by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan, and by JSPS Research Fellowships for Young Scientists, MEXT, Japan (T.I.). M.A. also acknowledges support from the German Research Foundation (DFG) through grant BR 3954/1-1
\section{Introduction} The possibility that we live on a brane embedded in a higher-dimensional space has been widely explored since one of its earliest string theory manifestations, the Horava-Witten model \cite{HW}. Brane-worlds afford intriguing opportunities to go beyond the standard model of particle physics and cosmology, but they also present challenges for reproducing known physics. The existence of moduli in these models is a generic feature with potentially important cosmological consequences. If unstabilized, they can conflict with big bang nucleosynthesis or fifth force constraints. This was the case in the original proposal of Randall and Sundrum (RS1) \cite{RS1}, which elegantly addressed the weak-scale hierarchy problem. These issues were resolved by introducing a bulk scalar field to stabilize the radion, the modulus determining the size of the extra dimension \cite{GW2}. Moduli are natural candidates to serve as the inflaton, and the radion is foremost among them. However it has been difficult to make radion inflation work in practice. Ref.\ \cite{rapid} proposed the modulus of large extra dimensions \cite{ADD} as the inflaton, but this required an inflaton potential with a rather peculiar shape to have inflation at the TeV scale \cite{jcextra}. In the framework of warped extra dimensions, most of the effort has been either on inflation driven by motion of branes within the bulk, or else conventional inflation on one of the branes modified by an unconventional Friedmann equation due to the extra dimensions. We are not aware of previous proposals for getting inflation from the radion in the RS-I (orbifold bounded by two branes) scenario.\footnote{a recent reference using different moduli fields is \cite{sundrum}.} It would be difficult to make this work, first because the Goldberger-Wise mechanism for stabilizing the radion does not give it a flat potential, and secondly because at Planck-scale values of the canonically normalized radion field, which could in principle have given chaotic inflation, the radion has already passed the barrier in the potential separating the compactified and decompactified regimes of the theory. In fact the light KK modes of the extra dimension become excited, invalidating the 4D effective description, long before the radion can reach the Planck scale. In this work we present a model of a warped 5D radion as inflaton that overcomes these difficulties. We incorporate a warped background solution that has negative curvature but is not AdS; instead it is conformally flat, and relies on a linearly varying bulk scalar field for the source of its negative bulk energy density. The weak scale hierarchy problem is naturally solved, and inflation occurs at a high scale, with weak couplings of the inflaton to matter on the standard model brane. We are able to relate the potential of the radion to the bulk scalar field potentials on the branes in a rather simple way, and to find radion potentials (starting from brane potentials of exponential form) that are flat enough for inflation with a certain amount of tuning of parameters of the brane potentials. Generically, one always expects decompactified flat space to be a solution to the Einstein equations, and so the radion potential naturally has degenerate minima representing compact and noncompact fifth dimension, separated by a barrier. In our model the barrier is at Planckian field values, and we can find examples where superPlanckian excursions during inflation lead to an observable tensor component in the perturbations. Other examples have a small tensor component with inflation ending close to the hilltop. Because the minima are degenerate, we have the situation of topological inflation \cite{Vilenkin,Linde:1994wt}: the universe divides into 4D and 5D domains separated by eternally inflating domain walls where the radion is near the top of the potential. This guarantees the existence of regions where the inflaton is arbitrarily close to the potential maximum (up to the limitation imposed by quantum fluctuations of the field) so that one does not have to rely upon fine-tuning of the initial position of the inflaton to get enough inflation. In section \ref{sect2} we introduce the static solution with stabilized radion and contrast it to the RS solution, especially with regard to the hierarchy problem. In section \ref{sect3} we perturb away from the static solution to find time-dependent ones, in the approximation that the back-reaction is small. We derive the effective 4D theory in section \ref{sect4}, computing the mass of the radion and the spectrum of Kaluza-Klein excitations for the radion and bulk scalar. We then construct some inflationary models using exponential brane potentials and work out their predictions for the CMB power spectrum, using both analytic and numerical techniques. Section \ref{sect5} derives the couplings of the radion to the standard model, enabling us to compute the reheat temperature at the end of inflation and to estimate the radiative corrections to the inflaton potential. We give conclusions in section \ref{sect6} and the bulk Einstein and scalar field equations in the appendix. \section{Unperturbed static solution} \label{sect2} We begin by introducing the framework for our 5D brane-world model. The action for 5D gravity coupled to the stabilizing scalar field, $\Phi =\Phi \left( y,t\right) $ with a bulk potential $V\left( \Phi \right) $ and potentials on the branes $V_{0,1}\left( \Phi \right) $ is \begin{eqnarray} S &= 2 \int_0^{1} d^{\,5}x\, \sqrt{g} \left(- \frac{1}{2\kappa_5^2}R + \frac{1}{2}\partial^{\mu}\Phi\partial_{\mu}\Phi -V(\Phi) \right) -\int d^{\,4}x\, \sqrt{g_4} \left. V_0\right|_{y=0} -\int d^{\,4}x\,\sqrt{g_4} \left. V_1\right|_{y=1} \nonumber\\ &\quad -\kappa_5^2\int d^{\,4}x\,\sqrt{g_4} \left. \left[K\right]\right|_{y=0} -\kappa_5^2\int d^{\,4}x\,\sqrt{g_4} \left. \left[K\right]\right|_{y=1} \label{action} \end{eqnarray} The extra dimension is an $S^{1}/Z_{2}$ orbifold (hence the factor of $2$) and all functions are symmetric under $y\rightarrow -y$. Thus we consider the interval $y\in \left[ 0,1\right] $ for the compact extra dimension with the Planck $\left( y_{0}=0\right) $ and TeV $\left( y_{1}=1\right) $ branes located at the orbifold fixed points. We also include extrinsic curvature terms $[K]$ that are needed for properly defining the variations of the action. The most general metric ansatz that respects 3D homogeneity and isotropy can be written as \begin{equation} ds^{2}=e^{2N\left( t,y\right) }dt^{2}-e^{2A\left( t,y\right) }d\mathbf{x} ^{2}-e^{2B\left( t,y\right) }dy^{2} \label{metric_general} \end{equation} The field equations for this ansatz are well known and we reproduce them in appendix \ref{appA}. \subsection{Static solutions} It will be useful for what follows to construct an exact solution to the coupled bulk scalar field and Einstein equations, which is conformally flat: \begin{equation} ds^{2} =e^{2n\left( y\right) }\left( dt^{2}-d\mathbf{x}^{2}-b^{2}dy^{2}\right) \label{metric_stat} \end{equation} Notice that this differs from AdS$_5$ in Randall-Sundrum coordinates where the $e^{2n\left( y\right) }$ factor does not multiply $dy^2$. The negative vacuum energy density provided by the bulk scalar potential is not constant, so the geometry is not AdS. This solution is well-known (see \textit{e.g.,} \cite{myers}-\cite{polchinski1}) in the context of strings propagating in a spacetime with subcritical dimensionality, compensated by a spatially linearly varying dilaton similar to our bulk scalar. With this ansatz the Einstein equations in the bulk reduce to \begin{align} -3\left( n^{\prime \prime }+n^{^{\prime }2}\right) & =\kappa_5^{2}\left( \frac{1}{2}\Phi ^{\prime 2}+b^{2}e^{2n}\,V\left( \Phi \right) \right) \label{EE_static1} \\ -6\left( n^{^{\prime }}\right) ^{2}& =\kappa_5^{2}\left( -\frac{1}{2}\Phi ^{\prime 2}+b^{2}e^{2n}\,V\left( \Phi \right) \right) \label{EE_static2} \end{align} As usual, the scalar field equation of motion \begin{equation} \Phi ^{\prime \prime }+3n^{\prime }\Phi ^{\prime }-b^{2}e^{2n}\frac{dV}{% d\Phi }=0 \label{scalar_eq_stat} \end{equation} is not independent of the Einstein equations, but can be obtained by taking linear combinations of (\ref{EE_static1}, \ref{EE_static2}) and derivatives thereof. By choosing the bulk potential to take a special form, $V= -\bar V \exp(c_{{\scriptscriptstyle} V}\Phi)$, one finds that the solutions for $n$ and $\Phi$ are simple linear functions of $y$: $n(y)=n_0-kby$, $\Phi(y) = \Phi_0 + c_{{\scriptscriptstyle} \Phi}y$. Here $k$ plays the role of the AdS curvature scale, as in the usual RS solution, and by analogy we will so refer to it, even though the solution is not AdS. The integration constants $n_0$ and $\Phi_0$ can be absorbed into the normalizations of $b$ and $\bar V$ respectively without loss of generality, so we set them to zero. The equations of motion then imply that $\bar V = -{9 k^2\over 2\kappa_5^2}$, $c_{{\scriptscriptstyle} V} = 2{\kappa_5\over\sqrt{3}}$, $c_{{\scriptscriptstyle}\Phi} = {\sqrt{3}kb\over\kappa_5}$. With these choices, the bulk scalar potential is given by \begin{equation} V(\Phi) = -{3\over 2} \mu_5^2 k^2 \exp\left(2\mu_5^{-1}\Phi\right) \label{bulk_pot_2} \end{equation} where we define \begin{equation} \mu_5 = {\sqrt{3}\over \kappa_5} \end{equation} and the solution is \begin{equation} n(y) = -kby;\quad \Phi = \mu_5 k b y \end{equation} In contrast to the original RS construction, we do not include a bulk cosmological constant; the negative bulk potential $V(\Phi)$ is responsible for the curvature of the 5D geometry. Similar negative potentials have been studied in the context of Golberger-Wise stabilization of bulk fields in AdS$_5$ \cite{dewolfe}, where it was observed that the unboundedness of the potential does not lead to instabilities in the AdS background \cite{BF}. The present case is similar; in the analogous subcritical string theory situation where the bulk scalar is the dilaton, in the string frame the negative coefficient of the potential is interpreted as a negative bulk cosmological constant, which would give rise to an AdS background if the dilaton were held fixed. In the present situation we have in addition the couplings of the bulk field to the branes, which prevent the field from running away. \subsection{Junction conditions} The Israel junction conditions (see {\it e.g.,} ref.\ \cite{KO}) and the boundary conditions for the scalar field are given by \begin{equation} \left.b^{-1}e^{-n} n^{\prime }\right|_{y_{i}-\epsilon}^{y_{i}+\epsilon} =\left.\pm \frac{\kappa_5^{2}V_{i}}{% 3}\right|_{y_{i}},\quad b^{-1}e^{-n}\left. \Phi ^{\prime }\right|_{y_{i}-\epsilon}^{y_{i}+\epsilon} = \left. \pm \frac{\partial V_{i} }{\partial \Phi }\right|_{y_{i}} \label{junction2} \end{equation} where the $\pm$ apply respectively at $y_0 = 0$ and $y_1=1$. Then explicitly, \begin{align} 2 \mu_5^2 k & = V_{0}(0),\qquad\quad 2 \mu_5^2 k = -\,e^{n_{1}}\,V_{1}(\Phi_1) \label{junction1_stat} \\ 2 \mu_5 k& = \frac{\partial V_{0}}{\partial \Phi }(0),\quad\quad\ 2 \mu_5 k = -\,e^{n_{1}}\,\frac{\partial V_{1}}{\partial \Phi }(\Phi_1) \label{junction2_stat} \end{align} where $e^{n_{1}}=e^{-kb}$ and $\Phi_1 = \mu_5 kb$ are respectively the warp factor and the scalar field value on the TeV brane. Let us recall the physical significance of the junction conditions. In the pure RS model with no scalar field, the two junction conditions require fine-tuning of the brane tensions (here represented by the values of the potentials $V_0$ and $V_1$ on the branes). For generic values of the tensions, one would obtain a bulk solution that is not static, and which contains a black hole in the extra dimension \cite{Kaloper:1999sm}. Thus the two tunings can be interpreted as (1) the usual setting of the 4D cosmological constant to zero, to obtain a static solution, and (2) the tuning of the bulk black hole mass to zero. The radion $b$ is exactly massless in this solution, and can thus take any value. Next consider the addition of the bulk scalar field. In the usual implementations of the Goldberger-Wise mechanism, the back-reaction of the scalar field on the metric is taken to be small, and one simply solves the bulk scalar equation in the AdS$_5$ background. This is a second order equation, and the two constants of integration in the solution are determined by the two additional boundary conditions involving $dV_i/d\Phi$. This generically induces a potential for the radion that is minimized at some value $b_0$. In our solution, the situation is somewhat different; we have singled out a particularly simple form of the scalar field solution, which is only compatible with certain choices of $dV_i/d\Phi$ at the boundaries. It is still true that the radion is stabilized as in the generic GW mechanism, but there is an additional tuning of potential parameters needed to maintain the linear solution for $\Phi$. This extra tuning is on the same footing as that of the bulk black hole mass in the pure RS solution. We suspect that our results can be generalized to more complicated solutions in which this tuning is relaxed, but for ease of computation, we will adhere to this special situation. The upshot is that the junction conditions amount to three tunings of brane potential parameters, plus one condition that fixes the value of the radion $b$, and hence the warp factor on the TeV brane. Let us illustrate with a simple example of brane potentials which is similar to that made by GW: \begin{equation} V_0 = m^2_0(\Phi + u_0)^2,\quad V_1 = -m^2_1(\Phi+ u_1)^2 \end{equation} Solving the junction conditions, we obtain three constraints that can be regarded as fine-tunings on the parameters $m_0$, $u_0$ and $u_1$, \begin{equation} m_0^2 = {k\over 2},\quad u_0 = 2\mu_5,\quad u_1 = {\mu_5}\left(2-kb\right) \end{equation} while $m_1$ can be regarded as adjustable, and its value determines the warp factor through \begin{equation} e^{-kb_0} = {m_0^2\over m_1^2} \end{equation} One notices a disadvantage of our approach relative to the usual one; we have had to build the hierarchy of scales into our original choice of brane couplings to get an exponentially small warp factor, as opposed to having generic Planck-scale values for the brane potential parameters. The next example shows that this problem can be ameliorated if the potentials take an exponential form. \subsection{Exponential brane potentials} In our subsequent construction of an inflationary solution, we will make use of a different choice of brane potentials, which are exponentials in $\Phi$: \begin{align} V_{0}& =\phantom{-}\Lambda _{0}\,e^{\alpha_{0}\Phi/\mu_5 } -\Delta_{0}\,e^{\beta_{0}\Phi/\mu_5 } \label{brane_pot1} \\ V_{1}& =-\Lambda _{1}\,e^{\alpha_{1}\Phi/\mu_5 } +\Delta_{1}\,e^{\beta_{1}\Phi /\mu_5} \label{brane_pot2} \end{align} The junction conditions lead to \begin{eqnarray} \Lambda_{0}&=&2\mu_5^2k\left( \frac{1-\beta_{0}}{\alpha _{0}-\beta _{0}}\right),\quad\qquad\qquad \Delta _{0} = 2\mu_5^2k\left( \frac{1-\alpha _{0}}{\alpha _{0}-\beta _{0}}\right) \label{0brane_pot_param0}\\ \Lambda_{1}&=&2\mu_5^2k\left( \frac{1-\beta_{1}}{\alpha _{1}-\beta _{1}}\right)e^{(1-\alpha_1)kb_0},\quad \Delta _{1} = 2\mu_5^2k\left( \frac{1-\alpha _{1}}{\alpha _{1}-\beta _{1}}\right)e^{(1-\beta_1)kb_0} \label{0brane_pot_param1} \end{eqnarray} For a given desired value of $b_0$ determining the hierarchy of scales between the two branes, there are thus four free parameters $\alpha_i$, $\beta_i$ that can eventually be used to tune the potential of the radion to a form suitable for inflation, while the dimensionful couplings $\Delta_i$ and $\Lambda_i$ are fixed in terms of these. Notice that if $\alpha_1$ and $\beta_1$ happen to be moderately close to 1 ({\it e.g.,} $\sim 1.1$), the explicit hierarchy can be ameliorated between the dimensionful parameters on the Planck brane versus those on the TeV, while yielding a sufficiently large value of $kb_0\sim 37$ to explain the TeV scale, since then $\Lambda_1/\Lambda_0\sim e^{(1-\alpha_1)kb_0}$ and $\Delta_1/\Delta_0\sim e^{(1-\beta_1)kb_0}$. \section{Time-dependent solutions} \label{sect3} Our goal now is to extend the static solution of the previous section to dynamical ones in which the radion is displaced from its stable equilibrium value $b_0$. In refs.\ \cite{KO},\cite{JCHF}-\cite{KT2}, 5D solutions that also include the dynamics of a bulk scalar field were found for the case of a single brane. In these papers the search for solutions was simplified by assuming that the metric functions and the bulk scalar were separable functions of $t$ and $y$, which we will justify. The key to our approach will be to allow for a large excursion of the radion field (which plays the role of the inflaton) in response to linear perturbations of the metric and bulk scalar. Although such an ansatz is not guaranteed to be consistent, we will show that for suitable choices of the parameters for the exponential brane potentials (\ref{brane_pot1},\ref{brane_pot2}) it is justified, and will allow us to find physically interesting inflationary solutions. \subsection{Small perturbations with large radion fluctuations} We make an ansatz for general scalar perturbations similar to that in ref.\ \cite{CGK}, \begin{eqnarray} ds^{2}&=& e^{2n(y)}\left[e^{2F(x,y)}\eta _{\mu \nu }dx^{\mu}dx^{\nu } -b_{0}^{2}\,e^{2\varphi(t) +2G(x,y) }dy^{2}\right] \label{metricKK_aug}\\ \Phi(x,y) &=& -\mu_5\left(n(y) +\varphi(t)\right)+ \delta\Phi(x,y) \label{scalarKK_aug} \end{eqnarray} It reduces to the static solution of the previous section when $F = G = \delta\Phi = \varphi(t) = 0$. Normally then, one would linearize in all of these quantities. However we will find a more general solution for the perturbations by treating only $F,\,G,\,\delta\Phi$ as being small, while working to all orders in $\varphi(t)$ (and $\dot\varphi$). This corresponds to allowing for large excursions of the radion during inflation. We will justify it {\it a postieriori} by choosing special values of the brane potentials that lead to a sufficiently flat inflaton potential. The perturbations $F,\,G,\,\delta\Phi$ are constrained by the nondynamical Einstein equations, namely those containing only first derivatives. The $G_{05}$ equation gives \begin{equation} (\dot G + \mu_5^{-1}\dot{\delta\Phi})n' + (F' + \mu_5^{-1}\delta\Phi')\dot\varphi -\dot F' =0 \label{05_derived} \end{equation} Similarly the off-diagonal $G_{ij}$ equation gives \begin{equation} 2\partial_i\partial_j F + \partial_i\partial_j G =0 \label{ij_derived} \end{equation} It can be integrated to find \begin{equation} G(x,y) = -2F(x,y) + G_0(t,y) \label{G_pert} \end{equation} where an untenable term linear in $\mathbf{x}$ has been set to zero. The $G_{0j}$ equation \begin{equation} (\partial_j F + \mu_5^{-1}\partial_j \delta\Phi)\dot\varphi =0 \label{0j_derived} \end{equation} can be solved for the bulk scalar perturbations, using eq. (\ref{G_pert}) : \begin{equation} \delta\Phi(x,y) = -\mu_5 F(x,y) + \delta\Phi_0(t,y) \label{dP_pert} \end{equation} Lastly the $G_{i5}$ equation, after substituting eqs.\ (\ref{G_pert}) and (\ref{dP_pert}), reduces to \begin{equation} \partial_j (F' + 3n'F) =0 \label{i5_derived} \end{equation} The general solution to (\ref{i5_derived}) is \begin{equation} F(x,y) = \hat{F}(x)e^{-3n(y)} + F_0(t,y) \label{F_pert} \end{equation} We find that the $\hat{F}(x)$ terms in the solutions to the metric and scalar field perturbations, eqs. (\ref{G_pert},\,\ref{dP_pert},\,\ref{F_pert}), vanish at linear order from the off-diagonal constraint equations and so $\hat{F}(x)$ is undetermined. Therefore it can be consistently set to zero. Now let us specialize the general ansatz to a form that is adapted for the cosmological solutions we seek: \begin{eqnarray} ds^{2}&=& e^{2n(y)}\left[e^{2F_1( \varphi(t),y) + 2F_2(\varphi(t),y)}\left( dt^{2}-a^{2}(t)\, d\mathbf{x}^{2}\right) -b_{0}^{2}\,e^{2\varphi \left( t\right) -4F_2(\varphi(t),y) }dy^{2}\right] \label{metric_pert}\\ F_1(\varphi,y) &=& n(y)f(\varphi(t))\,,\quad F_2(\varphi,y)=f_2(\varphi(t))e^{(3+\varepsilon)(n_1-n(y))} \label{f2metric_pert} \\ \Phi(t,y) &=& \mu_5\left( kb_0y - \varphi(t)\right) +\delta\Phi(\varphi(t),y) \label{scalar_pert} \end{eqnarray} where $n_1=n(1)=-kb_0$ and $\varepsilon$ is an adjustable parameter.\footnote{We will see that the choice $\epsilon = -1$ leads to small back-reaction of $f$, $f_2$ and $\delta\Phi$ while allowing for large $\varphi$} It reduces to the previous static solution when the radion fluctuation $\varphi(t)$ vanishes, if $f$, $f_2$ and $\delta\Phi$ also vanish. One can interpret $f,f_2$ and $\delta\Phi$ as small back-reactions induced on the metric and scalar field, respectively, by the (possibly large) radion fluctuation $\varphi$. This ansatz does not provide an exact solution to the equations of motion, but it is a good approximation at leading order in $f$, $f_2$ and $\delta\Phi$ as long as these quantities are small, in a sense to be specified. We will find situations where it is possible to reliably consider large excursions of the radion during inflation, even superPlanckian ones. It will be shown that the functions $f$, $f_2$ and $\delta\Phi$ can be algebraically derived from the brane potentials $V_0$ and $V_1$. In turn they determine the potential for the radion in the 4D effective theory. We will find that the radion potential derived from this ansatz is dominated by contributions from $f$ and its bulk scalar counterpart (the part that is flat in the extra dimension). The other parts of the perturbations, those with exponential $y$-dependence, turn out to be relevant solely for the stabilization of the static solution. \subsection{The $G_{05}$ equation and junction conditions} To arrive at the functional form of the scalar perturbations we examine the off-diagonal Einstein equation. Linearizing the left hand side we find \begin{equation} -\dot{\varphi}\left( 1+f-\frac{df}{d\varphi } - 3F_2 + \frac{dF_2}{d\varphi}\right) =\mu_5^{-2} \,\dot{\Phi} \Phi' \label{05_pert} \end{equation} This motivates us to parametrize $\delta\Phi(\varphi,y) = \delta\Phi_1+\delta\Phi_2$ where \begin{equation} \mu_5^{-1}\delta\Phi_1(\varphi,y) = g(\varphi)\, , \qquad\qquad \mu_5^{-1}\delta\Phi_2(\varphi,y) = g_2(\varphi)e^{(3+\varepsilon)(n_1-n(y))} \label{scalar1_time} \end{equation} This separation is useful, because it will be necessary to require that $f,f_2,g,g_2\ll 1$ but not necessarily that $\varphi\ll 1$. With this parametrization, the dominant ``1'' terms which are zeroth order in the perturbations cancel as in our previous analysis and the linearized $G_{05}$ equation becomes \begin{equation} -f +\frac{df}{d\varphi }-\frac{dg}{d\varphi}= e^{(3+\varepsilon)k(y-b_0)}\left((1+\varepsilon)\frac{df_2}{d\varphi} +\frac{dg_2}{d\varphi} -(3+\varepsilon)(f_2 + g_2)\right) \label{05_reduced_new} \end{equation} It is satisfied everywhere in the bulk if \begin{eqnarray} \frac{df}{d\varphi}- \frac{dg}{d\varphi} &=& f\label{05_reduced}\\ (1+\varepsilon)\frac{df_2}{d\varphi} +\frac{dg_2}{d\varphi} &=& (3+\varepsilon)(f_2+ g_2) \label{05_f2g2} \end{eqnarray} and has negligible $y$-dependence when \begin{equation} f_2,g_2 \ll \varphi \qquad \text{and}\qquad f_2,g_2 \ll 1 \label{small_f2g2} \end{equation} In our numerical analysis we find eqs. (\ref{05_reduced},\,\ref{05_f2g2}) hold to a high degree of accuracy when the back-reaction is small. Moreover, for the cases we study, eq. (\ref{05_f2g2}) will reduce to eq. (\ref{05_reduced}). Now let us reconsider the junction conditions in the presence of the perturbations: \begin{eqnarray} 2\mu_5^2k\,\left.\left(1+f-(1+\varepsilon)f_2e^{-(3+ \varepsilon)(n_i-n_1)}\right) \,e^{-n_i-\varphi }\right|_{y=y_i}&=& \pm \left.V_{i}(\Phi_i +\delta\Phi)\right|_{y=y_i} \quad \label{junction1_pert} \\ 2\mu_5 k\,\left.\left( 1 + \left((3+\varepsilon)g_2+2f_2\right)e^{-(3+ \varepsilon)(n_i-n_1)}\right) \,e^{-n_i-\varphi }\right|_{y=y_i} &=& \pm \frac{\partial V_{i}}{\partial \Phi }\left.(\Phi_i + \delta\Phi)\right|_{y=y_i} \quad \label{junction2_pert} \end{eqnarray} where now $\Phi_0 = -\mu_5\varphi $ and $\Phi_1= \mu_5(kb_0-\varphi)$ take the place of the unperturbed scalar field at $y=0$ and $y=1$. These are algebraic equations that determine the back-reaction functions $f,g$ and $f_2,g_2$. At low energies the functions all have a linear dependence on $\varphi$ and are related as $g = f$ and $g_2 = (1+\varepsilon)f_2$. However, when $\varepsilon=-1$, $g_2$ vanishes at linear order in $\varphi$ so that the bulk scalar perturbations are effectively independent of $y$. Even when $\varphi$ is large, this choice simplifies the subsequent analysis without affecting the radion dynamics, so we will adopt $\varepsilon=-1$ from now on. The linearized jump conditions can be solved to first order in $g$ by Taylor-expanding $V_0'$ about $\Phi = - \mu_5\varphi$ at $y=0$: \begin{equation} g \cong - g_2e^{2n_1} + {2k e^{-\varphi}\left(1+ (2f_2 + 2g_2)e^{2n_1}\right) - \mu_5^{-1}\,V'_0\over V''_0} \label{geq_new} \end{equation} where the prime on potentials denotes $d \over d\Phi$. Expanding next for $V_0$ in eq. (\ref{junction1_pert}) we find \begin{equation} f = -1 + {e^{\varphi}\over 2 \mu_5^2 k}\left( V_0 + \mu_5 (g+g_2e^{2n_1}) V'_0 \right) \label{feq_new} \end{equation} Similarly, at $y=1$, we expand $V'_1$ about $\Phi= \Phi_1 - \mu_5\varphi$ and find for arbitrary stabilizing potentials using (\ref{geq_new}) and (\ref{feq_new}) \begin{equation} f_2 = -g_2\left[\frac{1}{2} +{V_0''V_1''e^{n_1}e^{\varphi}(1-e^{2n_1})\over 4k( V_0''+V_1''e^{3n_1})}\right] -\left[{\frac{1}{2}(V_0''+V_1''e^{n_1}) +\frac{e^{\varphi}}{4\mu_5k}(V_0''V_1'e^{n_1} -V_0'V_1''e^{n_1}) \over V_0''+V_1''e^{3n_1}} \right] \label{f2eq} \end{equation} Then expanding for $V_1$ in eq. (\ref{junction1_pert}) \begin{eqnarray} g_2 = G_2^{-1}& \left[ 2ke^{-\varphi}\mu_5(V_0'+V_1'e^{n_1}) + \left[e^{2n_1}\left((V_0 +V_1e^{n_1})V_1''e^{n_1} -(V_0' +V_1'e^{n_1})V_1'e^{n_1}\right) \right.\right. \nonumber\\ & + \left.\left. (V_0 +V_1e^{n_1})V_0'' -(V_0' +V_1'e^{n_1})V_0' \right] \left(1-e^{2n_1}\right)^{-1}\right] \label{g2eq} \end{eqnarray} where \begin{equation} G_2 = V_0'V_1''e^{3n_1} -V_1'e^{n_1}V_0'' \end{equation} A fine-tuned choice of brane potentials such that $V_0 = -V_1e^{n1}$ is tempting for simplifying the solutions since then $g_2 =f_2 =0$. However, our study of the KK excitations (section \ref{KK_sect}) shows that this choice is precluded by the presence of a tachyon instability. The expressions for the back-reaction functions may nevertheless be simplified by neglecting the subdominant contributions involving extra powers of the warp factor: \begin{eqnarray} g &\cong& {2k e^{-\varphi} - \mu_5^{-1}\,V'_0\over V''_0} \label{geq} \\ f &\cong& -1 + {e^{\varphi}\over 2 \mu_5^2 k}\left( V_0 + \mu_5g V'_0 \right) \label{feq} \\ g_2 &\cong& -g -\frac{2ke^{-\varphi}\mu_5\left(1+f \right) + \mu_5^{-1}V_1e^{n_1}}{V_1'e^{n_1}} \label{g2eq} \\ f_2 &\cong& -g_2\left( 1 +\frac{e^{\varphi}}{4k}V_1''e^{n_1} \right) - \frac12 -\frac{e^{\varphi}}{4k}\left(\mu_5^{-1}V_1'e^{n_1} + V_1''e^{n_1}g \right) \label{f2eq} \end{eqnarray} Our results for $f,g$ and $f_2,g_2$ are derived from the jump conditions and not from the Einstein equations, so it is not obvious that they satisfy the relation (\ref{05_reduced_new}) from the $G_{05}$ equation. However, it is simple to show that this indeed is the case, at linear order. Differentiating the first jump condition (at $y=0$ or $y=1$) in (\ref{junction1_pert}) with respect to $\varphi$ using (\ref{scalar1_time}) gives \begin{eqnarray} \left.\frac{dV_i}{d\Phi}\right|_{y=y_i} &=& \left.\frac{d\varphi}{d\Phi}\frac{dV_i}{d\varphi}\right|_{y=y_i} = \mu_5\left.\left(-1 + \frac{dg}{d\varphi} + \frac{dg_2}{d\varphi}e^{-2(n-n_1)}\right)^{-1}\frac{dV_i}{d\varphi}\right|_{y=y_i} \end{eqnarray} Eliminating ${dV_i\over d\Phi}$ using the second jump condition (at $y=0$ or $y=1$) in (\ref{junction2_pert}) and linearizing, the dominant ``1'' terms again balance and we arrive directly at (\ref{05_reduced_new}) evaluated on the branes (which also guarantees that (\ref{05_reduced_new}) is satisfied in the bulk since both sides of the equation vanish). This is similar to what was found in ref.\ \cite{CGK} where the linearized off-diagonal Einstein equation, $G_{\mu 5}$ evaluated on the branes was shown to be equivalent to one of the junction conditions of the metric so that no new constraints arise. For small $\varphi \ll 1$ we can Taylor-expand about the background and arrive at simple expressions for the back-reaction functions \begin{eqnarray} g &=& f , \quad g = \frac{-\sigma_{0} \left(\varphi-g_2e^{-2kb_0}\right)}{ 1 -\sigma_{0}} \simeq -\frac{\sigma_{0}\varphi}{ 1 -\sigma_{0}} \\ g_2 &=& 0, \quad f_2 = \frac12\frac{\varphi(\sigma_{1} - \sigma_{0})}{\left(1 -\sigma_{0} - e^{-2kb_0}(1 -\sigma_{1})\right) } \simeq \frac12\frac{\varphi(\sigma_{1} - \sigma_{0})}{(1 -\sigma_{0})} \end{eqnarray} where the jump conditions \begin{equation} \pm \mu_5^{-1} e^{n_i} V_{i}'(\Phi)\vert _{\Phi_i} = \pm e^{n_i} \mu_5^{-2} V_{i}(\Phi)\vert _{\Phi_i} = 2k \end{equation} have motivated us to parametrize the brane potentials such that \begin{equation} \pm \, e^{n_i} V_{i}''(\Phi)\vert _{\Phi_i} = 2k(1-\sigma_{i}) \label{BPS_shift} \end{equation} This has a convenient physical interpretation. When $\sigma_i = 0$ the radion is massless and the background solution will be BPS in the sense of the solution generating technique utilized in ref.\ \cite{dewolfe}. In general, the $\sigma_i$ depend on parameters of the brane potentials, and they determine the zero-mode masses for the radion and the bulk scalar field in terms of the curvature scale $k$. The effects of various choices of $\sigma_i$ will be explored in section \ref{sect4}. \subsection{The remaining Einstein equations} So far we have succeeded in finding the back-reaction as a function of the radion fluctuation $\varphi$, but we have not yet determined the dynamics of $\varphi$. This comes from solving the remaining Einstein equations. They are difficult to solve exactly due to their explicit dependence on $y$ through the warp factor $n(y)$ which also appears in the metric and scalar perturbations. However, when we supplement the restrictions (\ref{small_f2g2}) on $f_2$ and $g_2$ by analogous ones for $f,\, g$, \begin{equation} f,g\ll \varphi \qquad \text{and}\qquad f,g\ll 1 \label{small_fg} \end{equation} then it is possible to linearize to obtain some equations that have negligible $y$-dependence. The radion will induce only a small back-reaction on the 4D slices of the 5D spacetime relative to the static solution, \begin{equation} ds_{4}^{2}\simeq e^{2n\left( y\right) }\left( dt^{2}-a^{2}d\mathbf{x}% ^{2}\right) \left( 1+2nf\left( \varphi \right) + 2f_2(\varphi)e^{-2(n-n_1)} +\cdots \right) \end{equation} Similar to the off-diagonal equation (\ref{05_reduced_new}), zeroth order terms vanish since these involve factoring out $\varphi$ dependence with the background solutions. The remaining equations can then be averaged over the extra dimension to find the radion dynamics: \begin{equation} \int_0^1 dy \sqrt{g}\,G_N^M = \kappa_5^2 \int_0^1 dy \sqrt{g}\,T_N^M \label{EE_averaging} \end{equation} It was shown in reference \cite{CGRT} that the cosmology of the radion in the RS model can be obtained equivalently from the effective action or from averaging the Einstein equations as in (\ref{EE_averaging}). In Section 4.1 we will construct the 4D effective action that leads to the bulk-averaged equations of motion. The bulk equations to be averaged are given in the Appendix, eqs.\ (\ref{00_gen})-(\ref{55_gen}). Collecting $t$ derivatives to the right hand side, Using eq.\ (\ref{bulk_pot_2}) for $V$ and (\ref{scalar_pert}, \ref{scalar1_time}) for $\Phi$, linearizing eq. (\ref{EE_averaging}) with $\kappa_5^{2}=3\mu_5^{-2}$ and integrating over $y$ we find to first order in the back-reaction functions \begin{eqnarray} \left(\frac{\dot{a}}{a}\right)^2 + \frac{\dot{a}}{a}\dot{\varphi}\left(1+\Upsilon_1\right) -\dot{\varphi}^2 \left(\frac{1}{2} + \Upsilon_2 \right) = 3k^2e^{-2\varphi}\left(f - g -\Upsilon_5 \right) \label{00_linear_up} \\ 2\frac{\ddot{a}}{a} + \left(\frac{\dot{a}}{a}\right)^2 + \left(2\frac{\dot{a}}{a}\dot{\varphi} + \ddot{\varphi} \right)\left(1 + \Upsilon_1 \right) + \dot{\varphi}^2\left(\frac{5}{2} + \Upsilon_3\right) = 9k^2e^{-2\varphi}\left(f - g -\Upsilon_5 \right) \label{ii_linear_up}\\ \frac{\ddot{a}}{a} + \left(\frac{\dot{a}}{a}\right)^2 + \ddot{\varphi}\left(\frac{\Upsilon_1}{2} + \Upsilon_4'\right) + 3\frac{\dot{a}}{a}\dot{\varphi} \left(\frac{\Upsilon_1}{2} + \Upsilon_4\right) + \frac{\dot{\varphi}^2}{2}\left(1+ {\Upsilon_2 + \Upsilon_3 \over 2} +2\Upsilon_4\right) \nonumber \\ = k^2e^{-2 \varphi}\left(4f - 3g - \Upsilon_6\right) \label{55_linear_up} \end{eqnarray} where primes denote $\frac{d}{d\varphi}$ and \begin{eqnarray} \Upsilon_1 \equiv -\frac{2}{3}f'\left(1 - 3kb_0e^{-3kb_0}\Omega \right), \quad \Upsilon_2 \equiv -\frac{1}{2}\Upsilon_1 - g'- 3e^{-3kb_0}\left(f_2'+ g_2'\right)(1-e^{kb_0})\Omega\quad \\ \Upsilon_3 \equiv -\Upsilon_2 - 4g' - 12e^{-3kb_0}\left(f_2' + g_2' \right)(1-e^{kb_0})\Omega, \quad \Upsilon_4 \equiv 3e^{-3kb_0}f_2'(1-e^{kb_0})\Omega\quad\\ \Upsilon_5 \equiv g_2e^{-3kb_0}(1-e^{kb_0})\Omega , \quad \Upsilon_6 \equiv 3e^{-3kb_0}(2f_2+5g_2)(1-e^{kb_0})\Omega \quad \end{eqnarray} with \begin{equation} \Omega = (1 - e^{-3kb_0})^{-1} \label{omega_int} \end{equation} All of the $\Upsilon_i$ terms are subdominant to other similar terms appearing in the equations, so it is consistent to drop them and keep only the explicit $f$ and $g$ source terms (notice that $\Upsilon_{5,6}$ are exponentially suppressed by powers of the warp factor). This demonstrates our earlier assertion that the radion dynamics are dominated by the Planck brane potential, which determines $f$ and $g$, while $f_2$ and $g_2$ are needed for stabilization only. The remaining Einstein equations then take the simple FRW-like form \begin{eqnarray} \left( \frac{\dot{a}}{a}\right) ^{2}+\frac{\dot{a}}{a}\dot{\varphi}-\frac{% \dot{\varphi}^{2}}{2}&=&3k^{2}e^{-2\varphi }\left( f-g\right) \label{00_linear} \\ 2\frac{\ddot{a}}{a}+\left( \frac{\dot{a}}{a}\right) ^{2}+\ddot{\varphi}+2% \frac{\dot{a}}{a}\dot{\varphi}+\frac{5}{2}\dot{\varphi}^{2} &=& 9k^{2}e^{-2\varphi }\left( f-g\right) \label{ii_linear} \\ \frac{\ddot{a}}{a}+\left( \frac{\dot{a}}{a}\right) ^{2}+\frac{\dot{\varphi}% ^{2}}{2}& =&k^{2}e^{-2\varphi }\left( 4f-3g\right) \label{55_linear} \end{eqnarray} These equations are augmented by the linearized Klein-Gordon equation \begin{equation} \ddot{\varphi}+3\frac{\dot{a}}{a}\dot{\varphi}+\dot{\varphi}^{2} =4k^{2}e^{-2\varphi }\left( f-\frac{3}{2}g\right) \label{scalar_eq_linear} \end{equation} found at this level of approximation by integrating (\ref{scalar_eq_general}) or equivalently by starting from the 4D effective action (\ref{eff_act_Jordan}) in Jordan frame which we derive in the next section. Of course, not all four of the equations (\ref{00_linear}-\ref{scalar_eq_linear}) are independent. As was observed in ref.\ \cite{JCHF1}, the linear combination of equations (\ref{scalar_eq_linear})$-$(\ref{ii_linear}) + 2$\times$(\ref{55_linear}) is equivalent to eq.\ (\ref{00_linear}). Furthermore, $d\over dt$(\ref{00_linear})$+H(3\times(\ref{00_linear})-(\ref{ii_linear}))/ \dot\varphi + (\ref{scalar_eq_linear})+(\ref{00_linear})$ is equivalent to eq.\ (\ref{55_linear}). The two constraints arise as a consequence of the gauge symmetries, namely reparametrizations of the $t$ and $y$ coordinates \cite{JCHF1},\cite{CGRT}. \section{Inflation in the effective 4D theory} \label{sect4} It is not difficult to derive the form of the effective action that gives rise to the 4D equations of motion (\ref{00_linear})-(\ref{scalar_eq_linear}). Starting from the 5D action (\ref{action}) and the metric (\ref{metric_general}) one arrives at \begin{eqnarray} \begin{split} S = 2 \int_0^{1}d^{\,5}x\, \sqrt{g} \left[\mu_5^2\left(- e^{-2N}(\dot{A}^2 + \dot{A}\dot{B}) +2e^{-2B}N'^2 \right) + \frac{1}{2}\partial^{\mu}\Phi\partial_{\mu}\Phi -V(\Phi) \right]\\ -\int d^{\,4}x \sqrt{g_4}\left. V_0\right|_{y=0} -\int d^{\,4}x\sqrt{g_4} \left. V_1\right|_{y=1} \end{split} \label{5Daction2} \end{eqnarray} The boundary contributions from the compact extra dimension exactly cancel the extrinsic curvature terms and so do not appear in this expression. Substituting for the metric functions $A,B$ and $N$ in (\ref{metric_pert}), for $V$ and $\Phi$ with (\ref{bulk_pot_2}) and (\ref{scalar_pert}), and using the jump condition (\ref{junction1_pert}) to rewrite $V_i$ in terms of the perturbations one can linearize about $\Phi = \mu_5(kb_0y - \varphi)$. After integrating over $y$ we arrive at the effective action \begin{eqnarray} S = \frac{2\mu_5^2}{3k\Omega} \int a^3e^{\varphi} dt\left[ - \left(\frac{\dot{a}}{a}\right)^2 - \frac{\dot{a}}{a}\dot{\varphi}\left(1 + \Upsilon_1 \right) + \dot{\varphi}^2\left(\frac{1}{2} + \Upsilon_2\right) -3k^2e^{-2\varphi}\left(f -g - \Upsilon_5 \right)\right] \label{4Daction1} \end{eqnarray} Neglecting the small $\Upsilon_i$ terms as before, we can write the simplified effective action in Jordan frame as \begin{equation} S = {1\over 2\kappa_4^2}\int dt\, a^3\, e^{\varphi}\left( -6\left({\dot a\over a} +\frac12\dot\varphi\right)^2 + \frac{9}{2}\dot\varphi^2 - 2\kappa_4^2 V_{r,J}(\varphi)\right) \label{eff_act_Jordan} \end{equation} where $V_{r,J }= 9k^2\kappa_4^{-2} e^{-2\varphi}(f-g)$ and the 4D Newton's constant $\kappa_4^2 = 8\pi G$ is found from \begin{equation} \kappa_4^{-2} = 2\kappa_5^{-2}\int^{1}_0 b_0e^{-3kb_0y}dy = \frac{2\mu_5^2}{9k\Omega} \label{4D_Newton} \end{equation} It is straightforward to show that this action implies the equations of motion (\ref{00_linear})-(\ref{scalar_eq_linear}) to linear order in $f,g$ when we make use of eq.\ (\ref{05_reduced}). One can go to the Einstein frame by performing the Weyl transformation \begin{equation} a\left( t\right) =\gamma \bar{a}\left(\tau\right), \qquad dt =\gamma\, d\tau,\qquad \gamma =e^{-\frac{\varphi }{2}} \label{conf_trans} \end{equation} after which the action takes the form \begin{equation} S = {1\over 2\kappa_4^2}\int d\tau\, \bar a^3\, \left( -6\left({\dot{\bar a}\over \bar a}\right)^2 + \frac92\dot\varphi^2 - 2\kappa_4^2 V_r(\varphi)\right) \label{eff_act_Einstein} \end{equation} where dots now denote ${d\over d\tau}$, and the radion no longer mixes with the scale factor. The Einstein frame potential is \begin{equation} V_r = {9k^2\over \kappa_4^2} e^{-3\varphi}(f-g) \label{radpot} \end{equation} and $\varphi$ is related to the canonically normalized radion field $\phi$ by \begin{equation} \phi = {3\over \sqrt{2}}\,\kappa_4^{-1}\,\varphi = {3\over \sqrt{2}}\,m_p\, \varphi \equiv \mu_4\,\varphi \label{cannorm} \end{equation} where $m_p$ is the reduced Planck mass. The Friedmann equation then takes the usual form \begin{equation} \bar H^2 = {\kappa_4^2\over 3}\left(\frac12\dot\phi^2 + V_r\right) \end{equation} as does the Klein-Gordon equation \begin{equation} \ddot{\phi} + 3\bar{H}\dot{\phi} = -\frac{dV_r}{d\phi} \end{equation} We will continue to do some of the subsequent analysis using the dimensionless field $\varphi$ however, to avoid having to repeatedly write $\mu_4$. \subsection{Radion stability} To summarize the results to this point, we have shown that in the class of 5D warped models we are considering, the radion potential is given by \begin{equation} V_r(\phi) = 9 k^2 m_p^2\, e^{-3\phi/\mu_4}(f-g) \label{radpot} \end{equation} where the functions $f$ and $g$ are determined by the Planck brane potential $V_0$ to leading order for small $f,g$, by eqs.\ (\ref{geq}, \ref{feq}). The TeV brane potential $V_1$ has virtually no effect because of the suppression of $f_2$ and $g_2$ by powers of the warp factor in eqs.\ (\ref{geq_new}-\ref{feq_new}). We are assuming that $\phi=0$ is a stable equilibrium point. Since $f=g= f'-g' =0$ at $\phi=0$ (see eq.\ (\ref{05_reduced})), it is clear that $\phi=0$ is a critical point of the potential, and $V_r$ also vanishes at this point, leading to a Minkowski solution. To compute the radion mass, notice that $f''-g'' = f'$. Therefore \begin{equation} m_{r}^{2}=\left. \frac{d^{2}\bar{V}_{r}}{d\phi ^{2}}\right\vert _{\phi=0} ={9 k^{2}m_p^2\over \mu_4^2}f'(0) = 2 k^2 f'(0) = 2 k^2 g'(0) \label{radion_mass1} \end{equation} Using eq.\ (\ref{geq_new}), we find that $g'(0) = 1 - 2k/V''_0(0)$. (Recall that $g' = {dg\over d\varphi}$ and $V_0'' = {d^2V_0\over d\Phi^2}$.) Consider for example the exponential potentials (\ref{brane_pot1}-\ref{brane_pot2}). The radion mass is determined by $V''_0(0) = 2k[\alpha_0^2(1-\beta_0) - \beta_0^2(1-\alpha_0)]/(\alpha_0-\beta_0)$. Recalling our discussion of the hierarchy problem, where it was pointed out that values of $\alpha_1$ and $\beta_1$ close to unity alleviate the need for a strong explicit hierarchy between the scales of the potentials on the two branes, we are motivated to define \begin{equation} \alpha_1 = 1 + \hat\alpha,\qquad \beta_1 = 1 + \hat\beta \label{a1b1defs} \end{equation} and for consistency when comparing parameters \begin{equation} \alpha_0 = 1 + \alpha,\qquad \beta_0 = 1 + \beta \label{abdefs} \end{equation} Then it is straightforward to show that $g'(0) = \alpha\beta(1+\alpha\beta)/(1-\alpha\beta)$. The radion mass squared is thus given by \begin{equation} m_r^2 = -2 k^2 \alpha\beta\, {1+\alpha\beta\over 1-\alpha\beta} \label{radmass} \end{equation} and there is a second zero mode whose mass has a similar expression in terms of the $\hat\alpha$ and $\hat\beta$ parameters. Hence $m_r^2$ is generically nonzero, and it is positive as long as $-1 < \alpha \beta < 0$ or $\alpha\beta > 1$. As for the remaining parameters we will show in the next subsection that $0 < \hat\alpha\hat\beta \ll 1$ is required in order to avoid a tachyon instability in the second light state, and to address the hierarchy problem. We emphasize an important difference between our model and the RS model supplemented by Goldberger-Wise stabilization of the radion. In the latter, the radion mass is suppressed by the warp factor and so is naturally at the TeV scale, whereas in ours, whose background geometry is different, it is natually at the scale $k$; we must do one tuning of parameters, $\alpha\ll 1$, to make it small enough for inflation. This will be advantageous when we consider the robustness of the radion-as-inflaton potential against quantum corrections from standard model physics later on. \subsection{Kaluza-Klein excitations}\label{KK_sect} One might wonder if it is sufficient to stabilize the radion without also considering the zero mode of the bulk scalar field. As ref.\ \cite{CGK} has shown, there is only a single KK tower for the two fields. However, in our model there is a second zero mode in the tower, the sign of whose mass squared depends on the choice of brane potentials. We will show that this state can be made much heavier than the radion so that it is consistent to neglect it along with the higher KK modes. Our main point is to show that the nonzero KK modes are not excited during inflation, which if it happened would invalidate our effective 4D description. We will find that these modes have masses of order $k$, while the second zero mode hass mass $\sqrt{\sigma_1} k$, with $\sigma_i$ defined in (\ref{BPS_shift}). The details of the calculations are left to Appendix \ref{appB}. Following Csaki, Graesser and Kribs (CGK) \cite{CGK} we consider the spectrum of perturbations around the background solution. The metric ansatz used to describe the scalar fluctuations differs slightly from theirs since we are working with a $y$-coordinate in which the background is conformally flat% \begin{eqnarray} ds^{2}&=&e^{2n(y) }\left[ e^{2F(x,y)}\eta _{\mu \nu }dx^{\mu}dx^{\nu }-e^{2G(x,y)}dy^{2}\right] \label{metricKK} \\ \Phi (x,y)&=& \Phi_b(y) +\delta\Phi(x,y) =\mu_5 kby +\delta\Phi(x,y) \label{scalarKK} \end{eqnarray} A similar analysis was subsequently done by Kofman, Martin and Peloso (KMP) \cite{Kofman:2004tk}. They generalized the CGK results by constructing a gauge invariant combination of scalar perturbations which ensures the hermiticity and orthogonality of modes in the mass spectrum, defined here as \begin{equation} v \equiv z \left( F + \frac{\delta \Phi}{\mu_{5}}\right) \label{defv} \end{equation} with \begin{equation} z \equiv \sqrt{2} \frac{\Phi_{b}'}{n'} e^{\frac{3n}{2}} = -\sqrt{2} \mu_5 e^{-\frac{3ky}{2}} \label{defz} \end{equation} $v$ is analogous to the Mukhanov-Sasaki variable \cite{mukhanov,sasaki} in 4D inflationary cosmology. Expanding the action (\ref{action}) to second order and diagonalizing in terms of the different scalars, one finds separable solutions $v(x,y) = \sum_{j}Q_j(x)\tilde v_j(y)$ such that the quadratic action of the KK modes can be reduced to \cite{Kofman:2004tk} \begin{equation} S = \sum_{j} C_j \int d^{\,4} x \; Q_j \left[ -\Box - m_j^2 \, \right] Q_j \label{KKactions} \end{equation} with the normalization coefficients \begin{eqnarray} C_j &\equiv& \frac{1}{2}\int_0^{1} dy\,b_0 \,\tilde{v}_j^2 + \frac{\mu_5^2}{k} \left.e^{3n}\tilde F_j^2\right|_{y_0=0}^{y_1=1} \notag \\ &=& \frac{1}{2}\int_0^{1}dy\, b_0e^{3n} \, \left(\frac{1}{kb_0}\tilde F'_j- 3\tilde F_j\right)^2 + \frac{\mu_5^2}{k} \left.e^{3n}\tilde F_j^2\right|_{y_0=0}^{y_1=1} \label{KKnorm} \end{eqnarray} where $\tilde F_j$ is related to $\tilde v_j$ as $F$ is to $v$ in (\ref{defv}). The solutions are determined by the jump conditions (\ref{junction1_KK},\,\ref{junction2_KK}) and the two nondynamical constraints (\ref{KKgauge},\,\ref{mu5_KK}) originating from the linearized off-diagonal Einstein equations to determine $G$ and $\delta\Phi$. To calculate the mass spectrum in our model we incorporate the results of KMP and CGK except in the application of the boundary conditions. Unlike the present work, refs.\ \cite{CGK},\cite{Kofman:2004tk} assumed a convenient stiff potential limit for the stabilizing potentials that forces $\delta \Phi$ to vanish on the branes with the result that only the radion zero mode appears with the infinite tower of KK excitations. Relaxing this assumption gives rise to the bulk scalar zero mode. One can also understand the two light modes as being the moduli of the positions of the two branes in the bulk \cite{brax2003cosmological}. General solutions obtained from the action expanded at second order for $v_j$ are equivalent to those obtained for $F_j$ from the linearized Einstein equations \cite{Kofman:2004tk}. We prefer to solve for $F_j$ similarly to CGK since the boundary conditions and normalization look simpler in this language. The heavier KK excitations and the zero modes take the form% \begin{eqnarray} \tilde F_j &=& e^{-\frac{3}{2}kb_0y}\left(A_j \sin (\sqrt{\lambda}_j b_0y) + B_j \cos (\sqrt{\lambda}_j b_0y)\right) \label{KKsol} \\ \tilde F_{z} &=& e^{-\frac{3}{2}kb_0y}\left(A_{z} e^{\sqrt{\lambda}_{z}b_0y} + B_{z} e^{-\sqrt{\lambda}_{z} b_0y}\right) \label{radsol} \end{eqnarray} where \begin{equation} \lambda_j = m_j^2 -\frac{9k^2}{4} \quad, \quad \lambda_{z} = \frac{9k^2}{4} - m_{z}^2 \label{KKradlambda} \end{equation} for $4m_j^2 \geq 9k^2$ and $m_{z}^2 \ll k^2$ respectively. The index $j\ge0$ is an integer which labels the heavier KK modes, while $z=r,s$ stands for the two light (radion or bulk scalar) zero-modes. The complete solutions are obtained by applying the linearized boundary conditions to determine the ratio of integration constants and the mass. The lightest KK excited state has $\lambda_0 = 0$ with $m_0^2 = \frac{9}{4}k^2$. Heavier modes must be found numerically except for some special cases which we display in Table \ref{tab:table1}. \begin{table}[h] \begin{center} \begin{tabular}{l|l|l} $\qquad\quad\sigma_i$ & \, $m_j^2$ (KK modes) & \, $m_{z}^2$ (zero modes) \\[5pt]\hline & & \\[-12pt] $\sigma_0 <0 \,,\sigma_1>0 \,,$ \, & \, $m_0^2 = \frac{9}{4}k^2$ & \, $m_{r,s}^2 \simeq -2\sigma_0 k^2 , \, \sigma_1 k^2 $ \\ $|\sigma_0|\ll|\sigma_1| \ll 1$ &\, $m_{j+1}^2 \simeq \frac{9}{4}k^2+ \left[\left(j + \frac{1}{2}\right)\frac{\pi}{b_0}\right]^2$ \, & \\[5pt] \hline & & \\[-12pt] $\sigma_i =0$ & \, $\frac{9}{4}k^2 + \left(\frac{j\pi}{b_0}\right)^2 $ &\, 0 \\[5pt] $|\sigma_i| \gg 1$ & \, $\frac{9}{4}k^2+ \left(\frac{j \pi}{b_0}\right)^2$ & \, $2k^2$ \\[5pt] $\sigma_0 = \sigma_1 ,\, |\sigma_i| \ll 1$ \, & \, $\frac{9}{4}k^2+ \left(\frac{j\pi}{b_0}\right)^2$& \, $m_{r,s}^2 \simeq -2\sigma_0 k^2 , \, \sigma_0 k^2 $ \end{tabular} \caption{\label{tab:table1} \textbf{A summary of the mass spectrum for various brane potentials parametrized by $\sigma_i$}. The entries correspond to the general case where tachyons are absent, followed by the massless radion, the stiff potential limit and fine-tuned brane potentials.} \end{center} \end{table} Recall from the discussion in the previous subsection that $\sigma_1=\hat\alpha\hat\beta \ll 1$ is needed to explain the hierarchy between the Planck and TeV branes. If $|\sigma_0| \ll |\sigma_1| \ll 1$ so as to achieve a large mass gap between the radion and the other modes, then the KK masses are well-approximated by \begin{equation} m_{j+1}^2 \cong \frac{9}{4}k^2+ k^2\left[\left(j + \frac{1}{2}\right)\frac{\pi}{kb_0}\right]^2 \label{KKmass_heavy} \end{equation} for $j\ge 0$. Thus all KK mode masses are at least order $k$, independent of the precise details of the stabilizing potentials. We previously showed that the radion mass depends on $V_0$, the potential on the Planck brane; see eq.\ (\ref{radion_mass1})). Similarly the bulk scalar zero mode mass depends on the TeV brane potential $V_1$. In the case where the brane potentials are tuned such that $|\sigma_i| \ll 1$ (eq.\ (\ref{BPS_shift})) the zero mode masses are approximately \begin{equation} m_{r}^2 \cong -2\sigma_0k^2 \, , \, \sigma_1 k^2 \label{radmass2} \end{equation} To avoid tachyonic instabilities we must impose $\sigma_0 < 0$ and $\sigma_1 > 0$. This result agrees with the analysis of criteria regarding the presence of tachyons for brane world models utilizing the Goldberger-Wise mechanism carried out in reference \cite{lesgourgues2004}. Comparing (\ref{radmass2}) with (\ref{radmass}), for the exponential brane potential $V_0$ (\ref{brane_pot1}) which implies $\sigma_0 = \alpha\beta$, we corroborate at first order in $\sigma_0$ the more exact expression for the radion mass given in the previous subsection. We will henceforth assume that $\sigma_1=\hat\alpha\hat\beta$ from $V_1$ (\ref{brane_pot2}) is large enough so that the bulk scalar zero mode is much heavier than the radion: \begin{equation} |\sigma_0| \ll {\sigma_1} \ll 1 \label{radmass_restrict} \end{equation} We will verify that $m_s \gg H$ (the Hubble scale during inflation) so that it is consistent to ignore the bulk scalar zero mode during inflation. The KK modes are much heavier and thus can also be considered as frozen. \subsection{Radion inflation: an explicit model} \label{two-exp} Let us now consider the full radion potential (valid to leading order in $f,g$) corresponding to the brane potentials (\ref{brane_pot1}-\ref{brane_pot2}), with the restrictions (\ref{0brane_pot_param0}-\ref{0brane_pot_param1}) and the definitions (\ref{abdefs}). Ignoring the small exponential terms involving $f_2$ and $g_2$ we find that \begin{equation} g = {(\alpha-\beta)e^{\beta\varphi} + \beta(1+\alpha)e^{(\beta-\alpha)\varphi} -\alpha(1+\beta)\over \alpha(1+\beta)^2 - \beta(1+\alpha)^2 e^{(\beta-\alpha)\varphi}} \label{geq_inf} \end{equation} and the combination $(f-g)$ that enters the radion potential (\ref{radpot}) is \begin{eqnarray} f-g &=& {1+g\over\alpha-\beta}\left[\beta-\alpha +\alpha(1+\beta)e^{-\beta\varphi} - \beta(1+\alpha)e^{-\alpha\varphi}\right]\nonumber\\ &+& g {\alpha\beta\over \alpha-\beta} \left( e^{-\beta\varphi}-e^{-\alpha\varphi} \right) \end{eqnarray} In the case we are interested in where $|\hat\alpha|,|\hat\beta|\ll 1$ in (\ref{a1b1defs}), the expressions for $f_2$ and $g_2$ (\ref{g2eq},\,\ref{f2eq}) take the simple form then \begin{eqnarray} f_2 &\simeq& -\frac12 f +g + \frac12\hat\alpha\hat\beta(\varphi-\varphi^2) \label{f2eq2} \\ g_2 &\simeq& f -g + \frac12\hat\alpha\hat\beta\varphi^2 \label{g2eq2} \end{eqnarray} Since $\varphi\sim 1$ during inflation while $\hat\alpha\hat\beta\ll 1$, it follows that if $f$ and $g$ are small then so are $f_2$ and $g_2$ and the consistency conditions (\ref{small_f2g2}) are satisfied. To design a suitable potential for inflation, we have the two dimensionless parameters $\alpha,\beta$ and one dimensionful scale $k$ at our disposal, subject to the restriction (\ref{radmass_restrict}) on $\sigma_0=\alpha\beta$. In addition $\sigma_1 = \hat\alpha\hat\beta$ must satisfy (\ref{radmass_restrict}) and we must ensure that $f,g\ll 1$ for consistency of the effective theory. Notice from (\ref{geq_inf}) that if $|\alpha|\sim|\beta|$, then generically $g\sim 1$ even if $|\alpha|,|\beta|\ll 1$. However if \begin{equation} |\alpha| \ll |\beta| \label{abineq} \end{equation} (or $\beta\ll\alpha$, giving qualitatively similar results) then $f$ and $g$ can be made parametrically small: both vanish in the limit $\alpha\to 0$. This motivates our choice of parameter values. We have also remarked that the radion potential has a second degenerate (Minkowski) minimum because of the decompactification limit as $\varphi\to\infty$. Thus the generic form of the potential is that of a hilltop model with a barrier separating the minima at $\varphi=0$ and $\varphi=\infty$. We find that this barrier can be made very flat when $\beta$ is close to $-3$, for $0< \alpha\ll 1$, making it possible to get inflation from near the top of the barrier. The shape of the potential is illustrated for $\alpha=10^{-7}$ and a few values of $\beta$ near $-3$ in figure \ref{potfig}. Figure \ref{corrfig} shows that the condition $f,g\ll 1$ is satisfied in the region $0 < \varphi < 3$ where inflation starting from the top of the barrier might take place. Choosing $\hat\alpha = 10^{-2}$ and $\hat\beta = 6 \times 10^{-3}$ for the fiducial values of $\alpha=10^{-7}$ and $\beta=-3$ is sufficient to ensure eqs. (\ref{f2eq2},\,\ref{g2eq2}) for $f_2$ and $g_2$ are good approximations and that the mass squared of the second light zero mode is much larger than that of the radion. We will presently show that it also exceeds the Hubble scale during inflation. \DOUBLEFIGURE[t]{pot.eps, width=\hsize}{fg.eps, width=\hsize}{Radion potential for $\alpha=10^{-7}$ and $\beta=-2.9,\, -3\, -3.1$. \label{potfig}}{Logarithm of the functions $g$ and $f-g$ that must be $\ll 1$ in the region of validity of the potential, for the case $\alpha=10^{-7}$ and $\beta=-3$. \label{corrfig}} \DOUBLEFIGURE[h]{traj-fig.eps, width=\hsize}{ns.eps, width=\hsize} {$\varphi$ as a function of number of $e$-foldings $N$ for $\alpha=10^{-7}$, $\beta=-3$ potential, starting $\Delta\varphi = 0.01$ from the top of the barrier. Inset shows end of inflation.\label{infsol}}{Spectral index as a function of number of $e$-foldings until the end of inflation, corresponding to fig.\ 3.\label{nsfig}} We have numerically integrated the inflaton equations of motion starting close to the top of the barrier. Taking the number of $e$-foldings $N$ as the time variable, these can be written in the first-order form, \begin{equation} {d\varphi\over dN} = {\pi\over \mu_4 H};\qquad {d\pi\over dN} = -3\pi - {1\over \mu_4 H}{dV_r\over d\varphi} \label{eoms} \end{equation} where $\pi = \dot\varphi$ is the canonical momentum and the Hubble parameter is given by $H^2 = \frac16 \pi^2 + \frac13 V_r$ in Planck units. (Recall that $\mu_4$ arises in the relation between the canonically normalized radion $\phi$ and the dimensionless one $\varphi$.) For $\alpha=10^{-7}$, $\beta=-3$, the maximum of the potential is at $\varphi_m = 3.01$. Starting at $\varphi=\varphi_m-\delta\varphi$ with $\delta\varphi=0.05$, for example, gives $N_e= 685$ $e$-foldings of inflation. $\varphi$ as a function of $N$ is shown in figure \ref{infsol}. Although the total number of $e$-foldings depends upon how small $\delta\varphi$ is taken to be, the spectral index at the time of horizon crossing (nominally 60 $e$-foldings before the end of inflation) is insensitive to the initial condition, as long as at least 60 $e$-foldings of inflation occurred. This behavior has been observed in a number of previous studies of hilltop inflation models \cite{racetrack}-\cite{d3d7}. To compute the power spectrum and spectral index we have used the prescription (see eq.\ (43) of ref.\ \cite{lyth}) \begin{equation} {\cal P} = {1\over 150\pi^2}\,{V_r\over m_p^4\epsilon},\qquad n_s = 1 + {d\ln{\cal P}\over dN} \label{norm} \end{equation} in terms of the slow roll parameter $\epsilon = \frac12 m_p^2 (V_r'/V_r)^2 = \frac19 m_p^2 ({d V_r \over d\varphi} /V_r)^2$. These are to be evaluated at horizon crossing, approximately 60 $e$-foldings before the end of inflation. More precisely, we take the crossing of the scale relevant for the COBE normalization of the power spectrum to be $N_{\rm{\scriptscriptstyle} COBE} = 56.5$ \cite{better}, a number that depends only logarithmically on the scale of inflation; since the spectrum changes slowly with $N$, this is an adequate approximation for our purposes. Fig.\ \ref{nsfig} shows the spectral index $n_s$ as a function of $N_e-N$, the number of $e$-foldings until the end of inflation. We see that $n_s=0.96$ at the scale of horizon crossing, in agreement with the WMAP5 central value. The normalization of the power spectrum at this point, $\sqrt{{\cal P}} = 2\times 10^{-5}$, implies that the curvature scale is given by $k = 0.015\, m_p$. Eq.\ (\ref{radmass}) shows that the radion mass is \begin{equation} m_r\cong \sqrt{6\alpha}k = 1.2\times 10^{-5} \,m_p = 3\times 10^{13} {\rm\ GeV}. \label{radmass13} \end{equation} {}From fig.\ \ref{potfig} the scale of inflation $m_I = V^{1/4}$ is given by $m^4_I \cong 3\times 10^{-8}\cdot 9 k^2 m_p^2$ in the flat region of the potential, hence \begin{equation} m_I = 2.7\times 10^{-3} m_p. \end{equation} The first slow roll parameter is $\epsilon = 1.1\times 10^{-4}$, giving a small tensor-to-scalar ratio of $r= 1.8\times 10^{-3}$. This is related to the fact that the canonically normalized inflaton field does not change by an amount much greater than the Planck scale \cite{lyth-bound}; in our model $\Delta\phi = \frac{3}{\sqrt{2}} m_p \Delta\varphi \cong 6 m_p$. This is consistent with a refined version of the Lyth bound \cite{mack}, $\Delta\phi \cong 6M_p\, r^{1/4}$ (where $M_p = \sqrt{8\pi}\, m_p$). We investigated the effect of varying the parameters $\alpha$ and $\beta$ away from the fiducial values $\alpha = 10^{-7}$ and $\beta=-3$. Holding $\beta$ fixed and varying $\alpha$, we find that for small $\alpha$ the spectral index asymptotes to its maximum value of $n_s = 0.963$ (figure \ref{ns-alpha}). Figure \ref{k-alpha} shows that the curvature scale $k$, determined by the normalization of the power spectrum, follows a power-law relation $k\sim \alpha^{-1/2}$ for very small values of $\alpha$. Recall that the prefactor of the inflaton potential is proportional to $k^2 m_p^2$; however the actual scale of inflation does not rise despite the increase in $k$; the $\epsilon$ slow roll parameter asymptotes to $1.2\times 10^{-4}$ at the smallest values of $\alpha$, which is nearly the same as its value at $\alpha=10^{-7}$. \DOUBLEFIGURE[t]{ns-alpha-old.eps, width=\hsize}{k-alpha-old.eps, width=\hsize} {Spectral index versus $\log_{10}\alpha$ for $\beta=-3$ model. \label{ns-alpha}}{Log of curvature scale $k$ versus $\log_{10}\alpha$ for $\beta=-3$ model.\label{k-alpha}} \DOUBLEFIGURE[h]{ns-beta-new.eps, width=\hsize}{k-beta-new.eps, width=\hsize} {Spectral index versus $\beta$ for $\alpha=10^{-7}$ model. \label{ns-beta}}{Log of curvature scale $k$ versus $\beta$ for $\alpha=10^{-7}$ model.\label{k-beta}} The corresponding results of varying $\beta$ at fixed $\alpha=10^{-7}$ are shown in figures \ref{ns-beta} and \ref{k-beta}. We see that it is possible to achieve a higher value of the spectral index than by varying only $\alpha$; the maximum value is $n_s=0.99$ at $\beta=-3.1$. \subsection{Model with larger tensors} In the previous model, constructed from brane potentials consisting of just two exponentials, there was not enough freedom to find examples with a tensor-to-scalar ratio $r$ exceeding $10^{-3}$. However with the addition of just one more exponential term, we can overcome this limitation. Consider the Planck brane potential \begin{equation} V_{0} =\Lambda _{0}e^{\alpha _{0}\Phi /\mu _{5}}-\Delta _{0}e^{\beta _{0}\Phi /\mu _{5}}-\Omega _{0}e^{\delta _{0}\Phi /\mu _{5}} \label{extended} \end{equation} The back-reaction functions $f$ and $g$ are given by \begin{eqnarray} f &=&-1+{e^{\varphi}} \left( { \Lambda_0}\,{e^{-{ \alpha_0}\,\varphi}}-{ \delta_0}\,{e^{-{ \beta_0}\,\varphi}}+{ \Omega_0}\,{e^{-{ \delta_0}\,\varphi}}+{ g}\, \left( { \Lambda_0}\,{ \alpha_0}\,{e^{-{ \alpha_0}\,\varphi}}-{ \delta_0}\,{ \beta_0}\,{e^{-{ \beta_0}\,\varphi}}+{ \Omega_0}\,{ \delta_0}\,{e^{-{ \delta_0}\,\varphi}} \right) \right) \nonumber\\ g &=& -{\frac {{e^{-\varphi}}-{ \Lambda_0}\,{ \alpha_0}\,{e^{-{ \alpha_0}\,\varphi}}+{ \Delta_0 }\,{ \beta_0}\,{e^{-{ \beta_0}\,\varphi}}-{ \Omega_0}\,{ \Delta_0}\,{e^{-{ \Delta_0}\, \varphi}}}{-{ \Lambda_0}\,{{ \alpha_0}}^{2}{e^{-{ \alpha_0}\,\varphi}}+{ \Delta_0}\,{{ \beta_0}}^{2}{e^{-{ \beta_0}\,\varphi}}-{ \Omega_0}\,{{ \Delta_0}}^{2}{e^{-{ \Delta_0 }\,\varphi}}}} \end{eqnarray} which lead to the radion potential (\ref{radpot}), proportional to $e^{-3\varphi}(f-g)$. Solving the unperturbed junction conditions similarly to eq.\ (% \ref{0brane_pot_param0}-\ref{0brane_pot_param1}), we obtain \begin{eqnarray} {\Lambda_{0}\over 2\mu_5^2 k} &=&{1\over {\beta-\alpha }} \left( {\beta +\frac{\Omega _{0}}{2\mu _{5}^{2}k}\,\left(\beta-\delta\right) }\right)\nonumber\\ {\Delta_{0} \over 2\mu_5^2 k}&=&{1\over {\beta-\alpha }} \left( {\alpha +\frac{\Omega _{0}}{2\mu _{5}^{2}k}% }\left({\alpha -\delta }\right)\right) \label{0brane_pot_param0_2} \end{eqnarray} where we have reparametrized $\delta_0=\delta +1$ similarly to (\ref{abdefs}). This model therefore has two parameters in addition to the previous one, $\delta$ and $\Omega_0$, the latter of which we find convenient to exchange for ${\Omega _{0}}/{2\mu _{5}^{2}k}\equiv h\alpha$. It will turn out that interesting values of ${\Omega _{0}}/{2\mu _{5}^{2}k}$ are of order $\alpha$, so that $h$ is of order unity. \EPSFIGURE[b]{hpot.eps, width=0.5\hsize} {Potential of extended model for $h=0,1,2$, and $\alpha=1.5 \times 10^{-9}$, $\beta=-3$, $\delta=-2.9$. \label{hpot}} \subsubsection{Analytic treatment} We have seen that taking $\alpha\ll 1$ is needed to get acceptable values of the spectral index. The potential can be approximated by a simpler expression in this regime, that allows us to make analytic predictions for the tensor-to-scalar ratio $r$. By further tunings of parameters, we can achieve a potential that is nearly linear during inflation,\footnote{See ref.\ \cite{eva} for another recent example leading to a linear potential, or ref.\ \cite{cliff} for other string-derived potentials supposed to be valid at Planckian field values.} leading to a large $\epsilon$ parameter and $\eta\cong 0$. This is the best case for getting a large tensor signal from the model. Linear behavior in $\varphi$ occurs if some of the exponents in $V_r$ are quite small. Taking $\delta$ is close to $-3$ realizes this possibility. First consider the small-$\alpha$ approximation to the potential, which can be written as% \begin{equation} V_{r}\cong9k^{2}m_{P}^{2}\alpha \left[ \frac{1-\delta h-3he^{-( \delta +3) \varphi }-e^{-3\varphi }( 1+3\varphi -h( \delta +3)) }{3+\alpha e^{3\varphi }( 4(1-\delta h)-3h( 1+\delta) ^{2}e^{-( \delta +3) \varphi }) }\right] \label{extend_rad_pot} \end{equation} This is a good approximation in the inflationary region, between $\varphi =0$ and the maximum of the potential. To see the turnover of the potential at large $\varphi$, it is important to keep the term of $O(\alpha)$ in the denominator; neglecting it leads to a potential that levels out at large $\varphi$ instead of having a maximum. Next consider the case where $0 < \delta+3\ll 1$ so that the terms of order $e^{-3\varphi}$ in the numerator are subdominant. This leads to \begin{equation} V_{r}\cong9k^{2}m_{P}^{2}\alpha \left[ \frac{1-\delta h-3he^{-( \delta +3) \varphi }}{3+ \alpha V_{\delta h}(\varphi)}\right] \label{extend_rad_pot_approx} \end{equation} The term $\alpha V_{\delta h}(\varphi)$ is $O(1)$ near the maximum of the potential, but it becomes subdominant away from the maximum, and the $\varphi$-dependence is dominated by the numerator. In this region, since $|\delta+3|\ll 1$, the dependence is in fact linear. We can write $V_r\sim (1 -\xi + 3\xi\varphi)$, where $\xi= h(3+\delta)$. This potential has slow-roll parameters $\eta\cong 0$ and \begin{equation} \epsilon = \frac12 \left({m_p^2\over\mu_4^2}\right) \left( V_r'\over V\right)^2 = {\xi^2\over (1-\xi+3\xi\varphi)^2} \label{epsilon} \end{equation} Inflation ends when $\epsilon\sim 1$, {\it i.e.,} \begin{equation} \varphi_e = {2\xi-1\over 3\xi} \label{phi_end} \end{equation} It is straightforward to solve the slow-roll equation of motion for the linear potential. Setting $d\pi/dN=0$ in eq.\ (\ref{eoms}), we have $d\varphi/dN = -(m_p/\mu_4)^{2} V'_r/V_r$. This gives a simple quadratic function for $N(\varphi)$. Using (\ref{phi_end}), one can find $\varphi$ as a function the number of $e$-foldings before the end of inflation, \begin{equation} \varphi(N) = \frac13\left(\sqrt{1+4N}+1 - \xi^{-1}\right) \label{phiN} \end{equation} Notice that with $N\cong 60$ and $\xi\cong 1$, $\varphi(60)\cong 5$, in agreement with the position of the maximum of the potential in fig.\ \ref{hpot}. Substituting this into (\ref{epsilon}) we get a simple result in which the dependence on the slope parameter $\xi$ drops out completely: \begin{equation} \epsilon = {1\over 4N+1} \cong 0.0044 \label{epsN} \end{equation} where we have taken $N=56.5$ to correspond to the COBE scale. This gives a tensor-to-scalar ratio of \begin{equation} r = 16\epsilon = 0.07 \label{rlin} \end{equation} which is somewhat higher than the projected sensitivity of the Planck experiment, $r_{\rm min} \cong 0.05$ \cite{planck}. Moreover the spectral index is \begin{equation} n_s = 1 - 6\epsilon = 0.974 \label{nslin} \end{equation} The prediction that $r$ is independent of the parameters of the potential in this regime depends on the assumption that the linear behavior in the numerator of (\ref{extend_rad_pot_approx}) dominates over the $\varphi$-dependence from $V_{\delta h}(\varphi)$ in the denominator. In the next subsection we will quantify this by solving the equations of motion numerically using the full potential. Using eqs.\ (\ref{phiN}-\ref{epsN}), the power spectrum (\ref{norm}) in this model is $P=3(150\pi^2)^{-1}\alpha \xi k^2 m_p^2$ $(4N+1)^{3/2}$, allowing us to determine $\sqrt{\alpha} k$ through the normalization: \begin{equation} \sqrt{\alpha}k =7.6\, \xi^{-1/2} \times 10^{-6}\, m_p \label{ak} \end{equation} The radion mass differs from that in (\ref{radmass13}) only by an extra factor of $\sqrt{1+\xi}\sim 1$. \subsubsection{Numerical analysis} Numerical analysis shows that the linear behavior of the potential is fully achieved for somewhat smaller values of $\alpha\mbox{\raisebox{-.6ex}{~$\stackrel{<}{\sim}$~}} 10^{-11}$ than we considered previously. For larger values of $\alpha$, the linear shape is not fully realized and the tensor ratio does not reach its maximum value of $r=0.07$. This is illustrated in figures \ref{tensor-h} and \ref{tensor-d}, which show the dependence of $r$ and $n_s$ on $h$ and $\delta$ respectively, when $\alpha = 1.5\times 10^{-9}$. On the other hand, by decreasing $\alpha$ to $1.5\times 10^{-12}$ and tuning $\delta$ closer to $-3$, the predictions (\ref{rlin}) and (\ref{nslin}) based on the linear potential are borne out, both for $r$ and for $n_s$. In all cases, values of $h(3+\delta)$ of order unity are needed to get the maximum tensor signal. \DOUBLEFIGURE[h]{tensor-h.eps, width=\hsize}{tensor-d-new.eps, width=\hsize} {Tensor ratio and $1-n_s$ versus $h$ for $\alpha=1.5 \times 10^{-9}$, $\beta=-3$, $\delta=-2.9$. \label{tensor-h}}{Tensor ratio and $1-n_s$ versus $\delta$ for $\alpha=1.5 \times 10^{-9}$, $\beta=-3$, $h=3$ \label{tensor-d}} \DOUBLEFIGURE[h]{h-smalla.eps, width=\hsize}{d-smalla.eps, width=\hsize} {Same as fig.\ 10, but for $\alpha = 1.5\times 10^{-12}$ and $\delta = -2.99$. \label{h-smalls}}{Same as fig.\ 11, but for $\alpha = 1.5\times 10^{-12}$ and $h = 100$. \label{d-smalla}} It is encouraging that the model is able to produce values of $r$ that exceed the minimum value of $r\cong 0.05$, which is estimated to be at the threshold for detection by the Planck satellite \cite{planck}. It is novel that we find such a regime in a hilltop potential \cite{hilltop}, for which the tensor ratio is typically unobservably small, $r< 0.002$. We have checked that for the examples shown here, the maximum value of $f,g$ in the inflationary region is $f=0.02$, so the condition $f,g\ll 1$ is still satisfied and the $O(f^2,g^2)$ corrections to our approximation for the potential are under control. Consistent with the analytic prediction (\ref{ak}), the curvature scale $k$ starts to exceed $m_p$ in the large-tensor models that have the smallest values of $\alpha$. For example with $1.5\times 10^{-11}$, $\delta = -2.99$ and $h=200$, we obtain $r=0.066$ and $k=1.4 m_p$. Examples with nearly as large tensors exist with $k < m_p$, such as $\alpha=5\times 10^{-11}$, $\delta = -2.985$, $h=100$, yielding $r=0.06$ and $k=0.8\, m_p$. However it is the combination $\sqrt{\alpha}k$ that appears in the potential, and this remains subPlanckian, so the fact that $k$ itself may exceed the Planck scale can be regarded as merely an artifact of our parametrization. \section{Coupling of radion to standard model} \label{sect5} An interesting feature of this model that distinguishes it from most other models of inflation is that the couplings of the inflaton to the standard model are exactly specified, since the radion is a component of the higher-dimensional metric. This enables us to address the details of reheating in a way that is usually not possible without making additional assumptions. Moreover, we can compute the quantum corrections to the radion potential to check whether its exotic form is radiatively stable. The most natural situation is that the SM fields are localized on the TeV brane. As an example, consider a Higgs field $H$. Using the metric (\ref{metric_pert}) and the conformal transformation (\ref{conf_trans}), we find that in the Einstein frame, its coupling to $\varphi$ is \begin{equation} {\cal L}_{\varphi,H} = e^{-\varphi-2kb_0f + 2f_2}(\partial H)^2 - e^{-2\varphi-4kb_0f + 4f_2} m_H^2 H^2 \label{smcoupling} \end{equation} where we have renormalized the field and the mass to absorb warp factors so that $m_H$ is the already-warped mass, $m_H\mbox{\raisebox{-.6ex}{~$\stackrel{<}{\sim}$~}}$ TeV. For small $\varphi$, $f(\varphi)\sim \alpha \varphi$ and $f_2(\varphi)\sim \hat\alpha\hat\beta \varphi$ which in the inflationary models we have considered is a negligible correction to the leading $\varphi$ dependence in the exponents, giving the usual result that the radion couples to the trace of the stress-energy tensor. \subsection{Reheating} Using the conventional theory of perturbative reheating, one estimates the reheat temperature as $T_r \sim g_*^{-1/4}(\Gamma M_p)^{1/2}$ where $\Gamma$ is the decay rate of the inflaton and $g_*$ the number of relativistic degrees of freedom. Using the interaction (\ref{smcoupling}) and recalling the relation (\ref{cannorm}) between $\varphi$ and the canonically normalized radion $\phi$ gives a rate of order $\Gamma \sim g_*m_H^4/(16\pi m_p^2 m_r)$, since there are $\sim g_*$ additional degrees of freedom besides the Higgs into which the radion can decay. For $m_H\sim $ 100 GeV, this leads to a reheat temperature that is too low even for nucleosynthesis, $T_r\sim 10^{-3}$ eV. However, there exists a much more efficient channel for reheating, by decay of the radion into gauge bosons. Although the radion does not couple to gauge bosons at tree level due to the tracelessness of their stress energy tensor, the conformal anomaly induces a coupling at one loop, of the form \cite{CGK} \begin{equation} {\alpha\over 8\pi}\varphi F_{\mu\nu} F^{\mu\nu} \end{equation} where $\alpha$ is the fine structure constant. An analogous term is present for gluons and the electroweak vector bosons. The decay rate due to this operator is much faster than that due to (\ref{smcoupling}) since the derivatives are of order $m_r/2$ as opposed to $m_H$.\footnote{we thank Neil Barnaby for pointing out the possibility of perturbative decay into gauge bosons}\ \ The decay rate is approximately \begin{equation} \Gamma \sim {\alpha^2 m_r^3\over 3^2\, 2^{13}\,\pi^3\, m_p^2} \end{equation} Using the radion mass (\ref{radmass13}) and $\alpha=0.1$ for QCD, this gives a reheat temperature of order \begin{equation} T_r \sim 10^7 {\rm\ GeV} \end{equation} which is high enough for electroweak baryogenesis, and low enough to avoid the gravitino problem. It is possible that reheating could be more efficient than indicated by this perturbative estimate, due to parametric resonance by the same coupling. Production of massless particles can be a particularly efficient form of preheating (see for example ref.\ \cite{Neil}). \subsection{Radiative corrections to inflaton potential} The radion potentials we have derived have exotic shapes compared to simple renormalizable potentials, and this has enabled us to produce distinctive signatures, like a larger tensor ratio than would be expected. One should always be concerned whether such peculiar potential shapes are radiatively stable. A second advantage of knowing the radion's couplings to matter is that we can address this question quantitatively. Consider the coupling (\ref{smcoupling}) to a Higgs field. To compute the contribution to the effective potential for $\varphi$, we can take $\varphi$ to be constant and renormalize $H$ to have a canonical kinetic term. Then the contribution to the Coleman-Weinberg potential is \begin{equation} \Delta V \sim e^{-2\varphi} {m_H^4\over 64\pi^2} \end{equation} Since $m_H^4 \ll k^2 m_p^2$ (the prefactor of the radion potential), this is a negligible correction. Of course the more relevant contributions come from integrating out particles at some intermediate scale $m_{\rm int}$, if there is new physics above the TeV scale. Even then however we are safe as long as $m_{\rm int}^4 \ll k^2 m_p^2$. One correction that does not appear in the Coleman-Weinberg potential is the cosmological constant on the branes, a constant shift to the brane potentials (\ref{brane_pot1}-\ref{brane_pot2}). Such terms can be studied in the extended model with brane potential (\ref{extended}) in the limit where $\delta_0=0$ and by adding a similar term to the TeV brane. We have checked that our conclusions based of the potential with two exponentials in \ref{two-exp} are not very sensitive to the addition of the constant term. Fig.\ \ref{ccpot} shows the effect of the constant term on the potential: for $h>0$, the flat region of the hilltop is shortened, leading to a shorter period of inflation, while for $h<-1$ it is destabilized. The spectral index as a function of $h$ is shown for the same potentials in fig.\ \ref{ccns}; values up to $h\cong 20$ are compatible with the $1\sigma$ WMAP5 allowed region, $n_s=0.963\pm 0.015$. \DOUBLEFIGURE[h]{cc_effect.eps, width=\hsize}{ns-cc-effect.eps, width=\hsize} {Effect of constant addition to brane potential on inflaton potential, parametrized as eq.\ (4.17) with ${\Omega _{0}}/{2\mu _{5}^{2}k}\equiv h\alpha$, for $\alpha=1.5\times 10^{-9}$, $\beta=-3$, and a range of values of $h=-10,\, 10,\dots,90$.\label{ccpot}}{Spectral index as a function of $h$ for the potentials of fig.\ 14. \label{ccns}} \section{Discussion} \label{sect6} We have explored a new framework for inflation, where the radion in a warped 5D compactification is the inflaton. Our approach differs from the usual Randall-Sundrum setup in that we work in the linear dilaton background (a bulk scalar that varies linearly with the extra dimension) rather than AdS$_5$. The bulk scalar stabilizes the radion by the Goldberger-Wise mechanism. Knowing its unperturbed solution exactly gives us greater computational control over the back-reaction when we perturb the radion away from the minimum of its potential. We derived an approximate analytic formula for the radion potential that depends mainly on the form of bulk scalar potential term on the Planck brane. Similar to the RS model, the hierarchy between the weak and Planck scales arises in our model without having to build it into the 5D Lagrangian. One can check that for the dimensionless parameter values we needed for inflation, the dimensionful brane potential parameters (\ref{0brane_pot_param0},\, \ref{0brane_pot_param1}) are dominated by the terms $\Lambda_0$ and $\Delta_1$, both of order $\mu_5^2 k$ for a warp factor consistent with $kb_0 \sim 37$ needed to solve the hierarchy problem. On the other hand, the mass of the radion does not get warped down to the weak scale; $m_r$ is fixed to be around $10^{13}$ GeV by the normalization of the CMB power spectrum. Moreover the couplings of the radion are not enhanced by the inverse warp factor; they remain Planck-suppressed. This tends to lead to a low reheat temperature, but also protects the inflaton potential from large radiative corrections. Choosing a sum of two exponentials as the brane potentials, we found that the radion potential suitable for inflation was generically of the hilltop form, but (with some tuning of dimensionless couplings) having a shape that could be much flatter than the generic negative quadratic form. Expanding the ansatz on the Planck brane to a sum of three exponential enabled us to find potentials where the descent from the hilltop is linear and consistent with a large tensor contribution $r=0.07$, detectable by Planck. An advantage of this model is that the couplings of the radion to matter on the branes can be computed explicitly. The radion couples to the trace of the stress energy tensor of a given particle, and to the one-loop conformal anomaly for states that are classically conformally invariant. We showed that this leads to reheating into gauge bosons by perturbative decay, with a reheat temperature of order $T_r\sim 10^7$ GeV. We did not explore the possibility of parametric resonance, which might significantly increase $T_r$. A direction for further development is to see if our scenario can be realized in a string theory compactification. It is well-known that warped-throat-like linear dilaton solutions arise in the near horizon region of a stack of NS 5-branes \cite{CHS} (see also section 14.1 of \cite{polchinski2}). Moreover, in the same way as the Klebanov-Strassler throats \cite{KS} are smoothly capped versions of anti-de Sitter throats near a stack of D3-branes, there exist known smoothly capped throat-like linear dilaton solutions \cite{MN}, \cite{Chamseddine}. So far these throats and their compactifications have not been as widely studied as the analogous Klebanov-Strassler throat. In ref.\ \cite{Greene} it was proposed that dynamically weakening gravity at early times could provide the low-entropy initial conditions that are needed for inflation to get started. In the Jordan frame our effective 4D action (\ref{eff_act_Jordan}) has a time-dependent Newton's constant $\hat{G}(\phi) = Ge^{-\phi /\mu _{4}}$ when the radion is displaced from its equilibrium value, which could potentially realize this scenario. In this frame, 4D gravity is effectively weak when the radion is far from its stable minimum, which is also the condition needed for inflation. During inflation $\hat{G}(\phi)$ grows to its normal strength and then oscillates about this value as the radion decays. It would be interesting to further investigate the extent to which our model is compatible with this proposal. Another feature of our model is that the mass spectrum of general scalar perturbations, including the zero modes of the radion and bulk scalar field and the infinite tower of KK excitations, is completely known in terms of $k$ and the parameters of the brane potentials. All modes can be normalized and their exact 4D effective actions determined. This presents an opportunity for constructing a model of assisted inflation \cite{Liddle:1998jc} which we will explore elsewhere \cite{JCJT2}. \paragraph*{\textbf{Acknowledgements:}} We thank Andrew Frey, Alex Maloney and Johannes Walcher for information about the linear dilaton background, Neil Barnaby for insights about reheating and preheating, and Cliff Burgess for discussions about radiative corrections to the brane potentials.
\section{Introduction} The standard model (SM) describes the electroweak interactions through a non-abelian gauge group $SU(2)_L \otimes U(1)_Y$, which includes self-interactions of gauge bosons. Because the $Z$ boson carries no electric charge, a coupling between a $Z$ boson and a photon is not permitted. The $Z\gamma$ production in the SM is dominated by the lowest-order Feynman diagrams shown in Fig.~\ref{ISRFSR}. \begin{figure}[htbp] \subfigure[]{\label{fig:SM1}\includegraphics[width=0.2\textwidth]{SM1}} \subfigure[]{\label{fig:SM2}\includegraphics[width=0.2\textwidth]{SM2}} \subfigure[]{\label{fig:SM1}\includegraphics[width=0.2\textwidth]{SM3}} \subfigure[]{\label{fig:SM2}\includegraphics[width=0.2\textwidth]{SM4}} \caption{Feynman diagrams for leading-order $Z\gamma$ production in the SM: (a) and (b) initial-state radiation from one of the initial-state partons, (c) and (d) final-state radiation from one of the final-state leptons. } \label{ISRFSR} \end{figure} \begin{figure}[htbp] \subfigure[]{\label{fig:ISR1}\includegraphics[width=0.22\textwidth]{AC2}} \subfigure[]{\label{fig:ISR2}\includegraphics[width=0.22\textwidth]{AC1}} \caption{\small Feynman diagrams illustrating anomalous $Z\gamma$ production with a $ZZ\gamma$ vertex (a) and a $Z\gamma\gamma$ vertex (b) . } \label{ACs} \end{figure} An excess in the number of high-energy photons can be a sign of new physics, e.g.,~supersymmetry, as described in Ref.~\cite{zgammaSUSY} or new heavy fermions with nonstandard couplings to the gauge bosons, as discussed in Ref.~\cite{zgammaSUSY2}. Such an excess of high-energy photons can be described by assuming only Lorentz and local $U(1)_{em}$ gauge invariant $ZZ\gamma$ and $Z\gamma\gamma$ trilinear gauge boson vertices of the form shown in Fig.~\ref{ACs}, using an effective theory with eight complex coupling parameters, $h^V_i$, where $i$ = $1,~2,~3,~4$ and $V=Z$ or $\gamma$ \cite{hzp}. Here, the couplings parameters $h^V_1$ and $h^V_3$ ($h^V_2$ and $h^V_4$) are associated with dimension-six (dimension-eight) operators which allow for an interaction between a $Z$ boson and a photon. To conserve tree-level unitarity at asymptotically high energies, one can introduce form factors dependent on the square of the partonic center-of-mass energy, $\hat{s}$, given by $h^V_i = h^V_{0i}/(1+\hat{s}/\Lambda^2)^n$, where $\Lambda$ is the mass scale at which the new physics responsible for anomalous couplings is introduced \cite{baur}. These anomalous gauge boson couplings would give rise to an excess of photons at high transverse momentum, $p_T^\gamma$, which can be searched for by measuring the total production cross section and the differential cross section $d\sigma/dp_T^\gamma$ for $Z\gamma\rightarrow \ell^+\ell^-\gamma$ ($\ell\ell\gamma$ henceforth) production. If no evidence of new physics is seen, we can place limits on the real components of the $CP$-even coupling parameters, $h^V_{03}$ and $h^V_{04}$, for $\Lambda = 1.2$ and $1.5$ TeV. Following Ref.~\cite{baur}, we choose form-factor powers for the unitarity scaling dimensions of $n=3$ for $h^V_3$ and $n=4$ for $h^V_4$. $Z\gamma$ production has been previously studied at collider experiments \cite{zgammaprev8, d0prev, zgammaprev5, zgammaprev2, zgammaprev7, zgammaprev6, zgammaprev4, zgammaprev3, zgammaprev1}, and because the value of $\Lambda$ greatly affects the scale of anomalous $Z\gamma$ production, we choose to perform this analysis for the values of $\Lambda$ that were used by the recent D0 \cite{zgammaprev8, d0prev} and CDF \cite{zgammaprev5} analyses. This choice of $\Lambda$ differs from the value used by the ALEPH \cite{zgammaprev2}, CMS \cite{zgammaprev6}, DELPHI \cite{zgammaprev4}, L3 \cite{zgammaprev3}, and OPAL \cite{zgammaprev1} collaborations. We present measurements of the inclusive cross section and differential cross section for $Z\gamma$ production in the electron and muon channels using a data sample corresponding to an integrated luminosity of 6.2 $\pm$ 0.4 fb$^{-1}$ collected at $\sqrt{s} = 1.96$ TeV by the D0 detector at the Fermilab Tevatron Collider between June 2006 and July 2010. These results provide a significant improvement in the sensitivity to anomalous $ZZ\gamma$ and $Z\gamma\gamma$ production compared to a previous D0 publication, utilizing the same channels and an integrated luminosity of 1 fb$^{-1}$ \cite{zgammaprev8}. In addition to increasing the size of the data set, we also combine with a previous result in the same channels \cite{zgammaprev8}, along with another D0 result \cite{d0prev} that used 3.6 fb$^{-1}$ of $Z\gamma\rightarrow\nu\nu\gamma$ production to place stringent limits on $Z\gamma$ anomalous couplings. \section{ The D0 Detector} The D0 detector \cite{dzero1,dzero2, dzero3, dzero4, dzero5} consists of a central tracking system contained within a 2~T superconducting solenoidal magnet, surrounded by a central preshower (CPS) detector, a liquid-argon sampling calorimeter, and an outer muon system. The tracking system, consisting of a silicon microstrip tracker (SMT) and a scintillating fiber tracker (CFT), provides coverage for charged particles in the pseudorapidity range $|\eta| \lesssim 3$~\cite{d0_coordinate}. The CPS is located immediately before the inner layer of the calorimeter and has about one radiation length of absorber followed by several layers of scintillating strips. The calorimeter consists of a central cryostat sector (CC) with coverage $|\eta| \lesssim 1.1$ and two end calorimeters (EC) which extend coverage to $|\eta| \approx 4.2$. The electromagnetic (EM) section of the calorimeter is segmented into four longitudinal layers (EM$i$, $i=1,~4$) with transverse segmentation of $\Delta\eta\times\Delta\phi = 0.1\times 0.1$, except in EM3, where it is $0.05\times 0.05$. The muon system resides beyond the calorimeter and consists of a layer of tracking detectors and scintillation trigger counters before a 1.8 T iron toroidal magnet, followed by two similar layers after the toroid. The coverage of the muon system corresponds to a pseudorapidity range $|\eta| < 2$. \section{Event selection} Candidate $Z\gamma$ events are selected in the $e^+e^-\gamma$ and $\mu^+\mu^- \gamma$ ($ee\gamma$ and $\mu\mu\gamma$ henceforth) final states. The $p\bar{p}$ interaction vertex must be reconstructed within $\pm 60$~cm of the center of the D0 detector along the beam ($z$) axis. For the electron channel, a sample of candidate $Z$-boson events is collected with a suite of single-electron triggers. The electrons are selected by requiring an EM cluster in either the CC ($|\eta| < 1.1$) or EC ($1.5 < |\eta| < 2.5$) regions of the EM calorimeter with transverse momentum $p_T > 25$ (15)~GeV/$c$ for the electron candidate with the highest (next-to-highest) transverse energy contained within a cone of radius $\Delta\mathcal{R}=\sqrt{(\Delta \eta)^2 + (\Delta \phi)^2}=0.2$, centered on the axis of the EM shower. At least 90\% of the cluster energy must be deposited within the EM section of the calorimeter. Electron candidates, with a shower shape consistent with that of an electron, are required to be spatially matched to a track and to be isolated in both the calorimeter and tracking detectors. To suppress jets and photons misidentified as electrons, a likelihood discriminant is built using a set of variables sensitive to differences in tracker activity and energy deposits in the calorimeter: the number of tracks and the scalar sum of the transverse momentum of all tracks within $\Delta\mathcal{R} < 0.4$ of the EM cluster, the fraction of energy deposited in the EM section of the calorimeter, the longitudinal and transverse shower profile in the calorimeter, and the ratio between the transverse energy in the calorimeter and the transverse momentum of the electron associated track. To further suppress jets misidentified as electrons, in particular, for high instantaneous luminosity conditions, a neural network algorithm is trained on Drell-Yan $Z/\gamma^* \to e^+e^-$ and jet data, using information from the calorimeter and CPS: the numbers of cells above a threshold in EM1 within $\Delta\mathcal{R} < $ 0.2 and 0.2 $< \Delta\mathcal{R} <$ 0.4 of the EM cluster, the number of CPS clusters within $\Delta\mathcal{R} < $ 0.1 of the EM cluster, and the squared-energy weighted width of the energy deposit in the CPS. Events where both electrons are contained within the EC are excluded because of the small signal acceptance. Candidate events where the $Z$ boson decays into two muons are collected using a suite of single-muon triggers. Within the muon channel, muon candidates are required to be within $|\eta| < 2$ and matched to a well-isolated track in both the tracker and the calorimeter with transverse momentum $p_T > 15$ GeV/$c$. The highest $p_{T}$ muon must have $p_T$ $>$ 20 GeV/$c$. Both muon candidates are required to originate from within 2 cm of the interaction point in the $z$ direction. Photon candidates in both the electron and muon channels are required to have transverse momentum $p_T^{\gamma} > 10$ GeV/$c$ within a cone of radius $\Delta\mathcal{R}=0.2$ centered around the EM shower in the CC. The rapidity of the photon, $\eta^\gamma$, is required to be $|\eta^\gamma|<1.1$. Additionally, the photon candidate must satisfy the following requirements: (i) at least $90\%$ of the cluster energy is deposited in the EM calorimeter; (ii) the calorimeter isolation variable $I = [E_{\text{tot}}(0.4)-E_{\text{EM}}(0.2)]/E_{\text{EM}}(0.2) <$~0.15, where $E_{\text{tot}}(0.4)$ is the total energy in a cone of radius $\Delta\mathcal{R}=0.4$ and $E_{\text{EM}}(0.2)$ is the EM energy in a cone of radius $\Delta\mathcal{R}=0.2$; (iii) the energy-weighted cluster width in the EM3 layer is consistent with that for an EM shower; (iv) the scalar sum of the $p_T$ of all tracks, $p_{T_{\text{trk}}}^{\text{sum}}$, originating from the interaction point in an annulus of $0.05<\Delta\mathcal{R}<0.4$ around the cluster is less than 2.0 GeV/$c$; (v) the EM cluster must not be spatially matched to either a reconstructed track or to energy depositions in the SMT or CFT detectors that are compatible with a trajectory of an electron \cite{HOR}; and (vi) an output larger than 0.1 of an artificial neural network ($O_{NN}$)~\cite{ANN5} that combines information from a set of variables sensitive to differences between photons and jets in the tracking detector, the calorimeter, and the CPS detector. The dilepton invariant mass, $M_{\ell\ell}$, is required to be greater than 60 GeV/$c^2$, and the photon must be separated from each lepton by $\Delta\mathcal{R}_{\ell\gamma} > 0.7$. Additionally, each lepton must be separated from a jet by $\Delta\mathcal{R}_{\ell j} > 0.5$. In the electron and muon channels, we select 1002 and 1000 data events, respectively. In order to reduce the contribution of final-state radiation (FSR), subset data samples are defined with the requirement that the reconstructed three-body invariant mass, $M_{\ell\ell\gamma}$, exceeds 110 GeV/$c^2$. With this additional requirement, 304 and 308 data events are selected in the electron and muon channels, respectively. \subsection{Background subtraction} The selected sample is contaminated by a small admixture of $Z+$jet events in which a jet is misidentified as a photon. To estimate this background in the electron channel, the fraction of jets that pass the photon selection criteria but fail either the $p_{T_{\text{trk}}}^{\text{sum}}$ or the shower width requirement, as determined by using a dijet data sample, is parametrized as a function of $p_T^{\gamma}$ and $\eta_{\gamma}$ (ratio method). The background from $Z$+jet production is then estimated starting from a data sample obtained by reversing the requirements either on $p_{T_{\text{trk}}}^{\text{sum}}$ or on the shower width requirement, and applying the same parametrization. A systematic uncertainty associated with the estimation of the number of real photons in the data samples is due to the finite size of the dijet background sample. After subtracting the estimated background from the data in the electron channel, we estimate 926 $\pm$ 53 (stat.)~$\pm$ 19 (syst.)~signal events when no $M_{\ell\ell\gamma}$ requirement is applied, and 255 $\pm$ 15 (stat.)~$\pm$ 5 (syst.)~signal events with $M_{\ell\ell\gamma}>110$ GeV/$c^2$. To estimate the background in the muon channel, we use a matrix method to estimate the $Z$+jet background contribution. After applying all of the selection criteria described above, a tighter requirement on $O_{NN}$ is used to classify the data events into two categories, depending on whether the photon candidate passes ($p$) or fails ($f$) this requirement. The corresponding number of events compose a 2-component vector ($N_p$, $N_f$). Thus, the sample composition is obtained by resolving a linear system of equations $(N_{p}, N_{f})^T = {\cal E}\times (N_{Z\gamma}, N_{Zj})^T$, where $N_{Z\gamma}$ ($N_{Zj}$) is the true number of $Z+\gamma$ ($Z+$jet) events in the fiducial region. The $2\times2$ efficiency matrix ${\cal E}$ contains the photon $\varepsilon_{\gamma}$ and jet $\varepsilon_{\rm jet}$ efficiencies that are estimated using photon and jet Monte Carlo (MC) samples and validated in data. Based on these studies, the efficiencies are parametrized as a function of the photon candidates' $\eta^\gamma$ with 1.5\% and 10\% relative systematic uncertainties for $\varepsilon_{\gamma}$ and $\varepsilon_{\rm jet}$ respectively. Having subtracted the estimated background from data in the muon channel, we estimate 947 $\pm$ 40 (stat.)~$\pm$ 16 (syst.)~signal events when no $M_{\ell\ell\gamma}$ requirement is applied, and 285 $\pm$ 24 (stat.)~$\pm$ 2 (syst.)~signal events requiring $M_{\ell\ell\gamma}>110$ GeV/$c^2$. As a cross-check, the $Z+$jet background is also estimated through a fit to the shape of the $O_{NN}$ distribution in data for both electron and muon channels, using MC templates constructed from simulated photon and jet events. The results are in good agreement with those obtained from the ratio and matrix methods. \section{Results} \subsection{Total cross section} The total cross section for $\ell\ell\gamma$ production is obtained from the ratio of the acceptance-corrected $\ell\ell\gamma$ rate for $M_{\ell\ell} > 60$ GeV/$c^2$, $\Delta \mathcal{R}_{\ell\gamma} > 0.7$, $p_T^\gamma > 10$ GeV/$c$, and $|\eta^\gamma| < 1$, to the total acceptance-corrected dilepton rate for $M_{\ell\ell} > 60$ GeV/$c^2$. Henceforth, these acceptance requirements are referred to as the generator-level requirements. We utilize this method because uncertainties associated with the trigger efficiencies, reconstruction efficiencies, and integrated luminosity are larger than the theoretical uncertainties and cancel in the ratio. This ratio is multiplied by a theoretical estimate for the total cross section for inclusive $Z/\gamma^*\rightarrow \ell\ell$ production for $M_{\ell\ell}>60$ GeV/$c^2$: \begin{equation} \label{xsect} \sigma_{Z\gamma} \times \mathcal{B}= \frac{ \kappa N_{\ell\ell\gamma}^{\text{data}}\left(A\times\epsilon_{ID}\right)_{\ell\ell\gamma}^{-1}}{N_{\ell\ell}^{\text{data}}\left(A\times\epsilon_{ID}\right)_{\ell\ell}^{-1}} \times (\sigma_Z\times \mathcal{B})_{\text{FEWZ}}^{\text{NNLO}}. \end{equation} Here, $N_{\ell\ell}^{\text{data}}$ and $N_{\ell\ell\gamma}^{\text{data}}$ are the number of measured $Z$ and background-subtracted $Z\gamma$ events in the data sample, respectively. The factor $(\sigma_Z\times \mathcal{B})_{\text{FEWZ}}^{\text{NNLO}}$ is calculated with the {\sc fewz} next-to-next-to-leading-order (NNLO) generator \cite{fewz1}-\cite{fewz2}, with the CTEQ6.6 parton distribution functions (PDF) \cite{CTEQ66}. The {\sc fewz} theoretical prediction is 262.9 $\pm$ 8.0 pb, where the dominant uncertainty is from the choice of PDF. The term $\mathcal{B}$ is the branching fraction for $Z/\gamma^*\rightarrow \ell\ell$, which in the SM is 3.4\% for either electrons or muons. The factor $\kappa$ corrects for the resolution effects that would cause events not to pass the selections on the generator-level quantities, e.g.~a generator-level photon with $p_T^\gamma < 10$ GeV/$c$, but to pass the reconstruction requirements, e.g.~a reconstructed photon with $p_T^\gamma > 10$ GeV/$c$. This factor is only used for $Z\gamma\rightarrow \ell\ell\gamma$ events, and corrects for the photon energy smearing that dominates in the first $p_T^\gamma$ bin. The muon $p_T$ resolution affects both $Z/\gamma^*\rightarrow\ell\ell$ and $Z\gamma\rightarrow \ell\ell\gamma$ and the corresponding correction cancel in the ratio of cross sections. For the events that pass the generator-level requirements, the factors $(A\times\epsilon_{ID})_{\ell\ell\gamma}$ and $(A\times\epsilon_{ID})_{\ell\ell}$ provide the fraction of events that pass the analysis requirements, with all acceptances measured relative to the kinematic requirements at the generator level for the $\ell\ell$ and $\ell\ell\gamma$ final states, respectively. Events migrate between bins in $p_T^\gamma$ because of finite detector resolution, and these effects are taken into account in calculating $(A\times\epsilon_{ID})_{\ell\ell\gamma}$ as a function of $p_T^\gamma$, while $(A\times\epsilon_{ID})_{\ell\ell}$ is calculated for the entire $\ell\ell$ sample. To estimate $\kappa$ and $A\times\epsilon_{ID}$, we use inclusive $Z/\gamma^{*}\rightarrow \ell\ell$ events generated with the {\sc pythia}~\cite{pythia} generator with final-state radiation simulated using {\sc photos}~\cite{photos} and the CTEQ6.1L~\cite{cteq61} PDF set. Because {\sc pythia} is a leading-order (LO) generator and does not reproduce the observed $p_T^Z$ spectrum in data, generated events are weighted to reflect the $p_T^Z$ distribution observed in Ref.~\cite{zpt}. Events are then traced through the D0 detector using a simulation based on {\sc geant} \cite{geant}. Data events from random beam crossings are overlaid on the simulated interactions to reproduce the effects of multiple $p\bar{p}$ interactions and detector noise. Simulated interactions that take into account the observed differences between data and simulation are reweighted, e.g., the $z$ coordinate of the vertex, instantaneous luminosity, trigger efficiency, lepton identification (ID) efficiency, photon ID efficiency, and resolution effects. Here, the factor $(A\times\epsilon_{ID})_{\ell\ell}$ has values of 0.15 (0.17) in the electron channel (muon channel). When no constraints on $M_{\ell\ell\gamma}$ are applied, the factor $\kappa$ has average values of $0.83 \pm 0.01$ (stat.)~and $0.85 \pm 0.01$ (stat.)~for the electron and muon channels, respectively, and the average value of $(A\times\epsilon_{ID})_{\ell\ell\gamma}$ is 0.12 for both the electron and muon channels. Values for $(A\times\epsilon_{ID})_{\ell\ell\gamma}$ and $\kappa$ are similar for the subsample requiring $M_{\ell\ell\gamma}>110$ GeV/$c^2$. To account for systematic uncertainty on the migration into the sample from generated events with $p_T^\gamma <$ 10 GeV/$c$, we conservatively vary the number of events produced outside the generator-level requirements in the {\sc pythia} simulation by $\pm$20\%, found as an upper estimate in studies of photon energy resolution in this kinematic regime, to measure the effect on the final cross section measurement. We find that the effect introduces a 1.5\% systematic uncertainty on the total cross section. The dominant uncertainty corresponding to the calculation of $A\times\epsilon_{ID}$ is due to choice of the PDF set. There are 20 free parameters in the CTEQ6.1L parametrization of the PDF that reflect fits to data from previous experiments. The uncertainties on acceptance and efficiencies due to the PDF parametrization are estimated using the CTEQ6.1M PDF uncertainties, following Ref.~\cite{PDFerrs}. We find a total PDF uncertainty of 3.5\%, dominated by the uncertainty on the acceptance-correction to the full geometrical lepton acceptance. The photon ID efficiency is determined from a simulated sample of photons and is estimated to have an uncertainty of 10\% for $p_T^\gamma<$ 15 GeV/$c$ and 3\% for $p_T^\gamma>$ 15 GeV/$c$. To reduce the contribution of FSR in the data samples, we calculate the cross section with and without the $M_{\ell\ell\gamma}>110$ GeV/$c^2$ requirement. To combine the electron and muon channels, we utilize the method in Ref.~\cite{BLUE}, which averages the results of measurements with correlated systematic uncertainties. We assume the PDF and photon ID efficiency uncertainties to be 100\% correlated between the two channels. The total cross section results can be found in Tables \ref{xsect} and \ref{xsect_llg110}. The measurements are consistent with the NLO {\sc mcfm} \cite{MCFM} prediction using CTEQ6.6 PDF set \cite{CTEQ66} and the renormalization and factorization scales evaluated at the mass of the $W$ boson, $M_W=80$ GeV/$c^2$. The PDF uncertainties associated with the SM prediction are evaluated following Ref.~\cite{PDFerrs}. We reevaluate the values for the $p_T^\gamma$ spectrum calculated by NLO {\sc mcfm} with the renormalization and factorization scales set to 160 GeV/$c^2$ and again at 40 GeV/$c^2$ and use these as estimates of the theoretical uncertainty of 1 standard deviation relative to the central NLO {\sc mcfm} value. \begin{table}[h!] \begin{ruledtabular} \begin{center} \caption{\label{xsect} Summary of the total cross-section measurements, when no $M_{\ell\ell\gamma}$ requirement is applied, for individual channels, combined channels, and the NLO {\sc mcfm} calculation with associated PDF and scale uncertainties. } \begin{tabular}{ l l l } & & \multicolumn{1}{c}{$\sigma_{Z\gamma} \times \mathcal{B}$ [fb]} \\ \hline \multicolumn{2}{l}{$ee\gamma$ data} & 1026 $\pm$ 62 (stat.) $\pm$ 60 (syst.) \\ \multicolumn{2}{l}{$\mu\mu\gamma$ data} & 1158 $\pm$ 53 (stat.) $\pm$ 70 (syst.) \\ \hline \multicolumn{2}{l}{$\ell\ell\gamma$ combined data} & 1089 $\pm$ 40 (stat.) $\pm$ 65 (syst.) \\ \hline \multicolumn{2}{l}{NLO {\sc mcfm}} & 1096 $\pm$ 34 (PDF) $^{+2}_{-4}$ (scale) \end{tabular} \end{center} \end{ruledtabular} \end{table} \begin{table}[h!] \begin{ruledtabular} \begin{center} \caption{\label{xsect_llg110} Summary of the total cross-section measurements, with the $M_{\ell\ell\gamma}>110$ GeV/$c^2$ requirement, for individual channels, combined channels, and the NLO {\sc mcfm} calculation with associated PDF and scale uncertainties. } \begin{tabular}{ l l l } & & \multicolumn{1}{c}{$\sigma_{Z\gamma} \times \mathcal{B}$ [fb]} \\ \hline \multicolumn{2}{l}{$ee\gamma$ data} & 281 $\pm$ 17 (stat.) $\pm$ 11 (syst.) \\ \multicolumn{2}{l}{$\mu\mu\gamma$ data} & 306 $\pm$ 28 (stat.) $\pm$ 11 (syst.) \\ \hline \multicolumn{2}{l}{$\ell\ell\gamma$ combined data} & 288 $\pm$ 15 (stat.) $\pm$ 11 (syst.) \\ \hline \multicolumn{2}{l}{NLO {\sc mcfm}} & 294 $\pm$ 10 (PDF) $^{+1}_{-2}$ (scale) \\ \end{tabular} \end{center} \end{ruledtabular} \end{table} \subsection{Differential cross section $d\sigma/dp_T^\gamma$} We use the matrix inversion technique \cite{unfolding} to unfold the experimental resolution and extract $d(\sigma_{Z\gamma}\times\mathcal{B})/dp_T^\gamma$ ($d\sigma/dp_T^\gamma$ henceforth), the differential cross section for $Z\gamma\rightarrow \ell\ell\gamma$, as a function of the true $p_T^\gamma$. The elements of the smearing matrix between true and reconstructed $p_T^\gamma$ bins are estimated using the full simulation of the detector response on a sample of $Z\gamma$ events generated using {\sc pythia}. Then, the matrix is inverted to obtain the unsmeared spectrum. We confirm that the unfolding procedure introduces a negligible bias. Following Ref.~\cite{bincentering}, the position of the data points are plotted in Figs.~\ref{unfolded} and \ref{unfolded_llg110} at the value of $p_T^\gamma$ where the cross section equals the average value for that bin. The theoretical uncertainties associated with the choice of PDF and the renormalization and factorization scales are determined analogously to the theoretical prediction for the total production cross section. The combined differential cross sections $d\sigma/dp_T^\gamma$ are shown in Figs.~\ref{unfolded} and \ref{unfolded_llg110} for no $M_{\ell\ell\gamma}$ requirement and $M_{\ell\ell\gamma}>110$ GeV/$c^2$, respectively. The values associated with Figs.~\ref{unfolded} and \ref{unfolded_llg110} are given in Tables \ref{unfoldedvals} and \ref{unfoldedvals_llg110}. \begin{figure}[htbp] \includegraphics[width=0.45\textwidth]{Unfolded_merged} \caption{Unfolded $d\sigma/dp_T^\gamma$ distribution with no $M_{\ell\ell\gamma}$ requirement for combined electron and muon data compared to the NLO {\sc mcfm} prediction. } \label{unfolded} \end{figure} \begin{figure}[htbp] \includegraphics[width=0.45\textwidth]{Unfolded_merged_llg110} \caption{Unfolded $d\sigma/dp_T^\gamma$ distribution with $M_{\ell\ell\gamma}>110$ GeV/$c^2$ for combined electron and muon data compared with the NLO {\sc mcfm} prediction. } \label{unfolded_llg110} \end{figure} \begin{table*} \begin{ruledtabular} \caption{\label{unfoldedvals} Summary of the unfolded differential cross section $d\sigma/dp_T^\gamma$, when no $M_{\ell\ell\gamma}$ requirement is applied, and NLO {\sc mcfm} predictions with PDF and scale uncertainties} \begin{tabular}{ D{,}{-}{-1} c l l } \multicolumn{2}{c}{ } & \multicolumn{1}{c}{$\ell\ell\gamma$ combined data} & \multicolumn{1}{c}{NLO {\sc mcfm}} \\ \hline \multicolumn{1}{c}{$p_T^\gamma$ bin } & $p_T^\gamma$ center & \multicolumn{2}{c}{$d\sigma/d p_T^\gamma $ } \\ \multicolumn{1}{c}{\mbox{[GeV/$c$]}} & [GeV/$c$] & \multicolumn{2}{c}{ [fb/(GeV/$c$)]} \\ \hline 10,15 & 12.4 & 111.14 $\pm$ 4.40 (stat.)~$\pm$ 11.99 (syst.) & 104.02 $\pm$ 4.10 (PDF) $^{+1.4}_{-1.2}$ (scale) \\ 15 , 20 & 17.2 & 51.41 $\pm$ 3.83 (stat.)~$\pm$ 2.65 (syst.) & 57.13 $\pm$ 2.23 (PDF) $^{+1.3}_{-1.8}$ (scale) \\ 20 , 25 & 22.5 & 25.34 $\pm$ 2.74 (stat.)~$\pm$ 1.13 (syst.) & 28.77 $\pm$ 0.43 (PDF) $^{+1.1}_{-0.7}$ (scale) \\ 25 , 30 & 27.5 & 8.08 $\pm$ 1.45 (stat.)~$\pm$ 0.40 (syst.) & 10.16 $\pm$ 0.26 (PDF) $^{+0.7}_{-0.5}$ (scale) \\ 30 , 40 & 34.4 & 3.23 $\pm$ 0.60 (stat.)~$\pm$ 0.17 (syst.) & 4.15 $\pm$ 0.16 (PDF) $^{+0.34}_{-0.19}$ (scale) \\ 40 , 60 & 48.5 & 1.70 $\pm$ 0.26 (stat.)~$\pm$ 0.088 (syst.) & 1.60 $\pm$ 0.061 (PDF) $^{+0.008}_{-0.010}$ (scale) \\ 60 , 100 & 76.5 & 0.34 $\pm$ 0.079 (stat.)~$\pm$ 0.018 (syst.) & 0.42 $\pm$ 0.017 (PDF) $^{+0.028}_{-0.028}$ (scale) \\ 100 , 200 & 124.5 & 0.038 $\pm$ 0.014 (stat.)~$\pm$ 0.002 (syst.) & 0.052 $\pm$ 0.001 (PDF) $^{+0.003}_{-0.001}$ (scale) \\ \end{tabular} \end{ruledtabular} \end{table*} \begin{table*} \begin{ruledtabular} \caption{\label{unfoldedvals_llg110} Summary of the unfolded differential cross section $d\sigma/dp_T^\gamma$, with the $M_{\ell\ell\gamma}>110$ GeV/$c^2$ requirement, and NLO {\sc mcfm} predictions with PDF and scale uncertainties .} \begin{tabular}{ D{,}{-}{-1} c l l } \multicolumn{2}{c}{ } & \multicolumn{1}{c}{$\ell\ell\gamma$ combined data} & \multicolumn{1}{c}{NLO {\sc mcfm}}\\ \hline \multicolumn{1}{c}{$p_T^\gamma$ bin } & $p_T^\gamma$ center & \multicolumn{2}{c}{$d\sigma/d p_T^\gamma $ } \\ \multicolumn{1}{c}{\mbox{[GeV/$c$]}} & [GeV/$c$] & \multicolumn{2}{c}{ [fb/(GeV/$c$)]} \\ \hline 10 , 15 & 13.7 & 13.57 $\pm$ 1.87 (stat.)~$\pm$ 2.43 (syst.) & 13.48 $\pm$ 0.48 (PDF) $^{+0.25}_{-0.51}$ (scale) \\ 15 , 20 & 17.2 & 14.87 $\pm$ 2.17 (stat.)~$\pm$ 2.30 (syst.) & 12.25 $\pm$ 0.47 (PDF) $^{+0.29}_{-0.36}$ (scale) \\ 20 , 25 & 22.0 & 7.91 $\pm$ 1.76 (stat.)~$\pm$ 0.81 (syst.) & 8.94 $\pm$ 0.25 (PDF) $^{+0.13}_{-0.35}$ (scale) \\ 25 , 30 & 27.4 & 5.30 $\pm$ 1.15 (stat.)~$\pm$ 0.44 (syst.) & 6.13 $\pm$ 0.21(PDF) $^{+0.016}_{-0.25}$ (scale) \\ 30 , 40 & 34.5 & 3.08 $\pm$ 0.57 (stat.)~$\pm$ 0.33 (syst.) & 3.71 $\pm$ 0.15 (PDF) $^{+0.012}_{-0.14}$ (scale) \\ 40 , 60 & 48.6 & 1.73 $\pm$ 0.26 (stat.)~$\pm$ 0.17 (syst.) & 1.57 $\pm$ 0.061 (PDF) $^{+0.004}_{-0.094}$ (scale) \\ 60 , 100 & 76.5 & 0.34 $\pm$ 0.079 (stat.)~$\pm$ 0.019 (syst.) & 0.42 $\pm$ 0.017 (PDF) $^{+0.028}_{-0.028}$ (scale) \\ 100 , 200 & 124.5 & 0.038 $\pm$ 0.014 (stat.)~$\pm$ 0.002 (syst.) & 0.052 $\pm$ 0.001 (PDF) $^{+0.003}_{-0.001}$ (scale) \\ \end{tabular} \end{ruledtabular} \end{table*} \begin{table}[h!] \begin{ruledtabular} \begin{center} \caption{\label{1Dlimits} Summary of the 1D limits on the $ZZ\gamma$ and $Z\gamma\gamma$ coupling parameters at the 95\% C.L. } \begin{tabular}{ l c c c } & & & $\ell\ell\gamma$ 7.2 fb$^{-1}$ \\ & \multicolumn{2}{c}{$\ell\ell\gamma$ 6.2 fb$^{-1}$} & $\nu\nu\gamma$ 3.6 fb$^{-1}$ \\ & $\Lambda=1.2$ TeV & $\Lambda=1.5$ TeV & $\Lambda = 1.5$ TeV \\ \hline $|h^Z_{03}|<$ & 0.050 & 0.041 & 0.026\\ $|h^Z_{04}|<$ & 0.0033 & 0.0023 & 0.0013\\ \hline $|h^\gamma_{03}|<$ & 0.052 & 0.044 & 0.027\\ $|h^\gamma_{04}|<$ & 0.0034 & 0.0023 & 0.0014\\ \end{tabular} \end{center} \end{ruledtabular} \end{table} \section{Limits on Anomalous Couplings} To set limits on anomalous trilinear gauge boson couplings, we generate $Z\gamma$ events for different values of the anomalous couplings using the NLO Monte Carlo generator of Ref.~\cite{baur}. SM Drell-Yan production is included by reweighting the $p_T^\gamma$ spectrum to {\sc mcfm} for vanishing anomalous couplings. As shown in Fig.~\ref{ACplot}, anomalous $Z\gamma$ couplings would contribute to an excess of high-energy photons as compared to the SM prediction. We apply the following generator-level requirements: $M_{\ell\ell}>60$ GeV/$c^2$, $\Delta \mathcal{R}_{\ell\gamma} > 0.7$, $p_T^\gamma>10$ GeV/$c$, $|\eta^\gamma|<1$, and $M_{\ell\ell\gamma}>$110 GeV/$c^2$, generate $p_T^\gamma$ templates as a function of the anomalous couplings, and use the known acceptance and resolution functions to fold the predicted generator-level distribution into a reconstruction-level distribution for $p_T^\gamma$. Using Poisson statistics for $p_T^\gamma > 30$ GeV/$c$, we define a likelihood function to compare the combined electron and muon channels with a predicted distribution for given values of anomalous couplings. In the absence of any significant deviation from the SM prediction, we set one-dimensional (1D) and two-dimensional (2D) limits on the anomalous coupling parameter values at the 95\% C.L. A combined log-likelihood function using all data is defined by the sum of the individual log-likelihood functions of the electron and muon channels. We include the effect of systematic uncertainties associated with transforming a Monte Carlo $p_T^\gamma$ template from the generator-level into a reconstructed distribution and find that these uncertainties contribute to the value of the calculated limits on the order of 1\%. We generate a $10 \times 10$ grid of templates for the $p_T^\gamma$ distribution as a function of $h^V_{03}$ and $h^V_{04}$ for $|h^V_{03}|<0.1$ and $|h^V_{04}|<0.01$, while setting all other coupling parameters to zero, and the limits are derived by varying about the maxima of the log-likelihood functions for the 95\% C.L \cite{PDGstats}. Results for the 1D limits for $\Lambda = 1.2$ TeV and 1.5 TeV are shown in Table \ref{1Dlimits}. The 1D and 2D limits on the anomalous coupling parameters are shown in Figs.~\ref{2D_1200} and \ref{2D_1500}, utilizing the electron and muon channels. In these figures, the dotted lines represent the theoretical limits on the anomalous coupling values, beyond which $S$-matrix unitarity is violated. Because the $h_{04}^V$ parameters come from dimension-eight operators, the limits are more constrained than those of $h_{03}^V$ couplings, which are dimension-six. \begin{figure}[h!] \begin{minipage}[b]{0.9\linewidth} \centering \includegraphics[scale=0.45]{Unfolded_merged_llg110_AC} \caption{\label{ACplot} The SM prediction and anomalous $Z\gamma$ coupling production compared with the unfolded $d\sigma/dp_T^\gamma$ for combined muon and electron channels for $p_T^\gamma > 30$ GeV/$c$ and $M_{\ell\ell\gamma}>110$ GeV/$c^2$.} \end{minipage} \end{figure} We combine these results with those of a previous D0 $Z\gamma$ analysis \cite{d0prev} . In that analysis, the 1D and 2D limits on the anomalous couplings parameter were calculated using a data sample corresponding to an integrated luminosity of 1 fb$^{-1}$ of data collected between October~2002 and Feburary~2006 (3.6 fb$^{-1}$ of data collected between October~2002 and September~2008) in the $ee\gamma$ and $\mu\mu\gamma$ channels ($\nu\nu\gamma$ channel), for $\Lambda = 1.5$ TeV. Results can be found in Fig.~\ref{2D_1500_wprev} and Table \ref{1Dlimits}. \begin{figure} \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hZ_lambda1200}} \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hg_lambda1200}} \caption{The 2D (contour) and 1D (cross) limits on the anomalous coupling parameters for (a) $ZZ\gamma$ and (b) $Z\gamma\gamma$ vertices at the 95\% C.L. for $\Lambda=1.2$ TeV. Limits on $S$-matrix unitarity are represented by the dotted lines. } \label{2D_1200} \end{figure} \begin{figure} \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hZ_lambda1500}} \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hg_lambda1500}} \caption{The 2D (contour) and 1D (cross) limits on the anomalous parameters for (a) $ZZ\gamma$ and (b) $Z\gamma\gamma$ vertices at the 95\% C.L. for $\Lambda=1.5$ TeV. Limits on $S$-matrix unitarity are represented by the dotted lines. } \label{2D_1500} \end{figure} \begin{figure}[h!] \begin{minipage}[b]{1.0\linewidth} \centering \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hZ_lambda1500_wprev}} \subfigure[]{\label{fig:ISR1}\includegraphics[scale=.45]{2D_Limits_hg_lambda1500_wprev}} \caption{\label{2D_1500_wprev} The 2D (contour) and 1D (cross) limits on coupling parameters for (a) $ZZ\gamma$ and (b) $Z\gamma\gamma$ vertices at the 95\% C.L. for $\Lambda=1.5$ TeV. Limits on $S$-matrix unitarity are represented by the dotted lines. } \end{minipage} \end{figure} \section{Conclusions} We have measured the differential and total cross sections for $Z\gamma\rightarrow \ell\ell\gamma$ production in $p\bar{p}$ collisions using the D0 detector at the Tevatron Collider with and without a $M_{\ell\ell\gamma}>110$ GeV/$c^2$ requirement. Both the total production cross sections and differential cross sections $d\sigma/dp_T^\gamma$ are consistent with the SM at NLO predicted by {\sc mcfm} \cite{MCFM}. We observe no deviation from SM predictions and place 1D and 2D limits on the $CP$-even anomalous $Z\gamma$ couplings for $\Lambda=1.2$ and 1.5 TeV. When combining with the previous D0 analyses, the limits are comparable to those found in the most recent CDF result \cite{zgammaprev5}, which uses $\approx 5$ fb$^{-1}$ in the $\ell\ell\gamma$ and $\nu\bar{\nu}\gamma$ channels and $\Lambda=1.5$ TeV. Our results include the first unfolded photon differential cross section $d\sigma/dp^\gamma_T$, as well as the most precise measurement of the total production cross section of $Z\gamma\rightarrow \ell\ell\gamma$. \begin{acknowledgements} \input acknowledgement.tex \end{acknowledgements} \input{Bib.tex} \end{document}
\section{Introduction} The six vertex model is one of the most studied solvable lattice model in statistical mechanics \cite{Baxter82}. Using the Bethe ansatz for eigenvectors and eigenvalues of the transfer matrix, Lieb \cite{Lieb67_Oct,Lieb67_April} and Sutherland \cite{Sutherland67} solved the regular square lattice zero-field six-vertex model with periodic boundary conditions. By computing the largest eigenvalue they obtained explicit solutions for the free-energy per site in the thermodynamic limit. The Bethe ansatz solution for a periodic square lattice with $L^2$ sites requires that for each eigenvalue one has to solve $n$ non-linear equations in $n$ unknowns (the `wave numbers' $k_1,\ldots,k_n$). This can only be done explicitly in the large-$L$ limit under the `string hypothesis'. Usually one can solve those equations explicitly only for the largest and near-largest eigenvalues, which limits one to considering the full infinite square lattice. On the infinite square lattice vertex operator methods are available (see e.g. \cite{Jimbo_Miwa94}), which have resulted in multiple integral expression for correlation functions of the model. The situation is different when one considers fixed boundary conditions. Izergin and Korepin constructed an explicit determinant solution for the case of `domain wall boundary conditions' \cite{Izergin87,Korepin82}. Alternative representations for this same partition function have also been proposed by other authors. See for instance \cite{Stroganov04,Stroganov06,Khoro_Paku05,Lascoux07,Galleas10,Galleas11} and references therein. In \cite{Baxter87}, Baxter constructs the partition function for a general $Z$-invariant six-vertex model, which is given explicitly (apart from a simple factor) by a Bethe ansatz type expression. However, in this case the `wave numbers' do not need to be evaluated from a complicated set of simultaneous equations: they are known explicitly. The disadvantage of Baxter's expression however is that it is given in terms of a large summation over the symmetric group. The main results of this paper is to rewrite Baxter's expression as a multiple contour integral over a factorised polynomial kernel. As an application of this result we specialise it to the regular square lattice with two specific boundary conditions. This allows us to write a new representation for the partition function of the six vertex model with domain wall boundary conditions as a multiple contour integral. Likewise we derive an off-shell multiple integral expression for the scalar product of two Bethe vectors of the six-vertex model transfer matrix. Such formul\ae are important as recent developments in the computation of three-point functions in $\mathcal{N}=4$ Super Yang-Mills (SYM) seem to require manageable expressions for the norms of Bethe vectors for general values of the parameters. With this in mind, in Section~\ref{sec:3pt} we illustrate how the integral formul\ae presented here can be embedded in the framework of \cite{Escobedo_2011}. \section{The six vertex model} \label{sec:vertexmodel} The exact solvability of a vertex model in the sense of Baxter \cite{Baxter_book} is intimately related to the solutions of the Yang-Baxter equation. This equation reads \begin{equation} \label{YB} \mathcal{L}_{12} (\lambda - \mu) \mathcal{L}_{13} (\lambda) \mathcal{L}_{23} (\mu) = \mathcal{L}_{23} (\mu) \mathcal{L}_{13} (\lambda) \mathcal{L}_{12} (\lambda - \mu) \end{equation} where $\mathcal{L}_{ij} \in \mbox{End}( \mathbb{V}_i \otimes \mathbb{V}_j )$. The complex variables $\lambda$ and $\mu$ correspond to spectral parameters while $\mathbb{V}_i$ denotes a complex vector space. We refer to the solutions of (\ref{YB}) as $\mathcal{L}$-matrices and for the six vertex model the corresponding $\mathcal{L}$-matrix is invariant under the $U_{q}[\widehat{\alg{sl}}(2)]$ algebra in the fundamental representation. In that case $\mathbb{V}_i \simeq \mathbb{C}^2$ and the associated $\mathcal{L}$-matrix explicitly reads \begin{equation} \label{su2} \mathcal{L} = \left( \begin{array}{cccc} a & 0 & 0 & 0 \\ 0 & b & c & 0 \\ 0 & c & b & 0 \\ 0 & 0 & 0 & a \end{array} \right) \end{equation} where \begin{equation} \label{weightsdef} a (\lambda) = \sinh{(\lambda+\gamma)},\quad b (\lambda) = \sinh{(\lambda)}\quad \text{and}\quad c(\lambda) = \sinh{(\gamma)}. \end{equation} To characterise the vertex model associated with a solution of the Yang-Baxter equation we shall employ the following notation, \begin{equation} \label{decomp} \mathcal{L} = \sum_{\alpha,\beta,\gamma,\delta} \mathcal{L}^{\beta \delta}_{\alpha \gamma} \; \hat{e}_{\alpha \beta} \otimes \hat{e}_{\gamma \delta} \; , \end{equation} where $\hat{e}_{\alpha \beta} = \ket{\alpha} \bra{\beta}$ where the vectors $\ket{\alpha}$, as well as their duals, form a basis of $\mathbb{V}_i$. In this way we associate the Boltzmann weight $\mathcal{L}^{\beta \delta}_{\alpha \gamma}$ to the vertex configuration $\{\alpha, \beta, \gamma, \delta \}$ as shown in the \figref{fig:bw}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.6,line width=0.4mm] \draw (-2.3,0) node[left] {{\large $\mathcal{L}^{\beta \delta}_{\alpha \gamma} \;\;\; \simeq \;\;\;$}}; \draw (-1,0) node [left] {$\alpha$} -- (1,0) node[right]{$\beta$}; \draw (0,-1) node[below]{$\gamma$} -- (0,1) node[above]{$\delta$}; \end{tikzpicture} \caption{Vertex $\{\alpha, \beta, \gamma, \delta \}$ and its Boltzmann weight.} \label{fig:bw} \end{figure} For the six vertex model the indices $\alpha$ and $\beta$ in (\ref{decomp}) run through the set $\{\uparrow , \downarrow \}$ while $\gamma$ and $\delta$ run through $\{ \rightarrow , \leftarrow \}$ defined as \begin{equation} \state{\uparrow} = \state{\rightarrow} = \left( \begin{matrix} 1 \cr 0 \end{matrix} \right) \in \mathbb{V}_i \qquad \mbox{and} \qquad \state{\downarrow} = \state{\leftarrow} = \left( \begin{matrix} 0 \cr 1 \end{matrix} \right) \in \mathbb{V}_i \; . \end{equation} From (\ref{su2}) we thus have six possible vertex configurations which are depicted in the \figref{fig:vertex6}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] \draw[>->] (-1,0) -- (1,0) ; \draw[>->] (0,-1) node[below]{$a$} -- (0,1); \draw[<-<] (4,0) -- (6,0); \draw[<-<] (5,-1) node[below]{$a$} -- (5,1); \draw[>->] (9,0) -- (11,0) ; \draw[<-<] (10,-1) node[below]{$b$} -- (10,1); \draw[<-<] (14,0) -- (16,0) ; \draw[>->] (15,-1) node[below]{$b$} -- (15,1); \draw[>-<] (19,0) -- (21,0); \draw[<->] (20,-1) node[below]{$c$} -- (20,1); \draw[<->] (25,0) -- (27,0); \draw[>-<] (26,-1) node[below]{$c$} -- (26,1); \end{tikzpicture} \caption{The 6 vertex configurations of the model.} \label{fig:vertex6} \end{figure} The vertices described in \figref{fig:vertex6} can be juxtaposed in a two-dimensional lattice with $N$ rows and $L$ columns as shown in the \figref{fig:lattice}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] {\scriptsize \draw (-2,0) node[left]{$\alpha_{N,1}$} -- (10,0) node[right]{$\alpha_{N,L+1}$}; \draw (0,-2) -- (0,2) node[above]{$\beta_{N+1,1}$}; \draw (2,-2) -- (2,2); \draw (3,2) node[above]{\dots}; \draw (4,-2) -- (4,2); \draw (5,2) node[above]{\dots}; \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2) node[above]{$\;\;\;\;\; \beta_{N+1,L}$}; \begin{scope}[yshift=-2.2cm] \draw (-2,0) node[left]{\vdots} -- (10,0) node[right]{\vdots}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \end{scope} \begin{scope}[yshift=-4.4cm] \draw (-2,0) node[left]{\vdots} -- (10,0) node[right]{\vdots}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \end{scope} \begin{scope}[yshift=-6.6cm] \draw (-2,0) node[left]{$\alpha_{1,1}$} -- (10,0) node[right]{$\alpha_{1,L+1}$}; \draw (0,-2) node[below]{$\beta_{1,1}$} -- (0,2); \draw (2,-2) -- (2,2); \draw (3,-2) node[below]{\dots}; \draw (4,-2) -- (4,2); \draw (5,-2) node[below]{\dots}; \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \draw (8,-2) node[below]{$\beta_{1,L}$} -- (8,2); \end{scope}} \end{tikzpicture} \caption{Two dimensional square lattice with boundary conditions index by $\alpha_{i,j}$ and $\beta_{i,j}$ who take values in $\{\leftarrow,\rightarrow\}$ and $\{\uparrow,\downarrow\}$ respectively.} \label{fig:lattice} \end{figure} The probability of having the vertex $\{ \alpha_{i,j}, \alpha_{i,j+1}, \beta_{i,j} , \beta_{i+1,j} \}$ at the Cartesian coordinates $(i,j)$ is weighted by the factor $\mathcal{L}_{\alpha_{i,j} \beta_{i,j}}^{\alpha_{i,j+1} \beta_{i+1,j} }$ and the partition function of the system is formed by the summation over all possible configurations of the product of vertex weights. More precisely, the partition function $Z$ is defined by \begin{equation} \label{partition} Z = \sum_{\{ \alpha_{i,j} , \beta_{i,j} \}} \prod_{i=1}^{N} \prod_{j=1}^{L} \mathcal{L}_{\alpha_{i,j} \beta_{i,j}}^{\alpha_{i,j+1} \beta_{i+1,j} } \; . \end{equation} It is possible to generalise this construction in the following way to include inhomogeneities while maintaining integrability. To each horizontal line $i$ of the square lattice we associate a variable $\lambda_i$, and to each vertical $j$ a variable $\mu_j$. The local Boltzmann weight for a vertex $\{ \alpha_{i,j}, \alpha_{i,j+1}, \beta_{i,j} , \beta_{i+1,j} \}$ is then given by $\mathcal{L}_{\alpha_{i,j} \beta_{i,j}}^{\alpha_{i,j+1} \beta_{i+1,j} }(\lambda_i-\mu_j)$ and the inhomogeneous partition function is defined as \begin{equation} \label{partition2} Z(\{\lambda_i\},\{\mu_j\}) = \sum_{\{ \alpha_{i,j} , \beta_{i,j} \}} \prod_{i=1}^{N} \prod_{j=1}^{L} \mathcal{L}_{\alpha_{i,j} \beta_{i,j}}^{\alpha_{i,j+1} \beta_{i+1,j} } (\lambda_i-\mu_j) \; . \end{equation} Boundary conditions play a fundamental role in the evaluation of the partition function \eqref{partition}. In this paper we will consider periodic boundary conditions in the horizontal and vertical directions ($\alpha_{i,L+1}=\alpha_{i,1}$ and $\beta_{N+1,j}=\beta_{j,1}$) as well as fixed boundary conditions with specified values of $\alpha_{i,j}$ and $\beta_{i,j}$ at the borders. \subsection{Monodromy and transfer matrix.} Each row of the two-dimensional lattice formed by the juxtaposition of vertices can be conveniently characterised by a matrix $\mathcal{T} (\lambda) = \mathcal{T} (\lambda , \{ \mu_j \})$ usually referred to as monodromy matrix. This matrix has components \begin{equation} \label{eq:mono1} \mathcal{T}^{\alpha', \beta'_1 \dots \beta'_L}_{\alpha, \beta_1 \dots \beta_L} (\lambda , \{ \mu_j \}) = \mathcal{L}^{\alpha_1 \beta'_1}_{\alpha \; \beta_1} (\lambda - \mu_1) \mathcal{L}^{\alpha_2 \beta'_2}_{\alpha_1 \beta_2} (\lambda - \mu_2) \dots \mathcal{L}^{\alpha_L \beta'_{L-1}}_{\alpha_{L-1} \beta_{L-1}} (\lambda - \mu_{L-1}) \mathcal{L}^{\alpha' \; \beta'_L}_{\alpha_L \beta_L} (\lambda - \mu_L) \end{equation} which is diagrammatically represented in the \figref{fig:mono}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.4,line width=0.4mm] \draw (-2.3,0) node[left]{{\large $\mathcal{T}^{\alpha', \beta'_1 \dots \beta'_L}_{\alpha, \beta_1 \dots \beta_L} \;\;\; \simeq \;\;\;$}}; \draw (-2,0) node[left]{$\alpha$} -- (14,0) node[right]{$\alpha'$}; \draw (0,-2) node[below]{$\beta_1$} -- (0,2) node[above]{$\beta'_1$}; \draw (2,-2) node[below]{$\beta_2$} -- (2,2) node[above]{$\beta'_2$}; \draw (4,-2) -- (4,2); \draw (5,-2) node{\dots}; \draw (5,2) node{\dots}; \draw (6,-2) -- (6,2); \draw (7,-2) node{\dots}; \draw (7,2) node{\dots}; \draw (8,-2) -- (8,2); \draw (10,-2) node[below]{$\beta_{L-1}$} -- (10,2) node[above]{$\beta'_{L-1}$}; \draw (12,-2) node[below]{$\beta_L$} -- (12,2) node[above]{$\beta'_L$}; \end{tikzpicture} \caption{Diagrammatic representation of the monodromy matrix.} \label{fig:mono} \end{figure} In this way the two-dimensional lattice described in the \figref{fig:lattice} can be completely built in terms of the operators $\mathcal{T} (\lambda , \{ \mu_j \})$. In a more compact form we have \begin{equation} \label{mono} \mathcal{T}(\lambda, \{ \mu_j \}) = \mathcal{L}_{\mathcal{A} 1} (\lambda - \mu_1) \dots \mathcal{L}_{\mathcal{A} L} (\lambda - \mu_L). \end{equation} which is an operator on the tensor product space $\mathbb{V}_{\mathcal{A}} \otimes \mathbb{V}_{1} \otimes \dots \otimes \mathbb{V}_{L}$. The space $\mathbb{V}_{\mathcal{A}}$ is usually called the auxiliary space while we shall refer to the tensor product $\mathbb{V}_{1} \otimes \dots \otimes \mathbb{V}_{L}$ as the quantum space. Thus the monodromy matrix $\mathcal{T}(\lambda)$ is a matrix in the auxiliary space whose elements $\mathcal{T}^{\alpha'}_{\alpha}$ act on the quantum space. In the diagrammatic representation of the monodromy matrix given in \figref{fig:mono}, the contraction of indices labelling the horizontal edges represent the product in the auxiliary space while the product in the quantum space is given by the contraction of indices labelling the vertical edges. For the six vertex model the monodromy matrix can be conveniently denoted as \begin{equation} \label{monorep} \mathcal{T} (\lambda) = \mathcal{T}(\lambda, \{ \mu_j \}) = \left( \begin{matrix} A(\lambda) & B (\lambda) \cr C (\lambda) & D (\lambda) \end{matrix} \right) \; . \end{equation} Here and in the following we suppress where possible the dependence on $\{\mu_j\}$ for clarity of notation. \section{Periodic boundary conditions} The evaluation of the partition function (\ref{partition}) depends drastically on the boundary conditions chosen. In this section we shall consider periodic, or toroidal boundary conditions ($\alpha_{i,L+1}=\alpha_{i,1}$ and $\beta_{N+1,j}=\beta_{j,1}$), and review the basics of the quantum inverse scattering method (QISM) for computing the six-vertex partition function in this case. From \eqref{partition} and \eqref{eq:mono1} it is a standard construction that the partition function for toroidal boundary conditions can be written as \begin{equation} Z(\{\lambda_i\},\{\mu_j\}) = \mathop{\mathrm{Tr}}_{\mathbb{C}^{2\otimes L}} \left(\tau(\lambda_1)\cdots \tau(\lambda_{N}) \right), \end{equation} where $\tau(\lambda)$ is the transfer matrix defined by \begin{equation} \label{transfmatrix} \tau(\lambda) = \tau(\lambda,\{\mu_j\}) = \mathop{\mathrm{Tr}}_{\mathcal{A}} \mathcal{T}(\lambda) = A(\lambda)+D(\lambda). \end{equation} Due to the Yang-Baxter equation the transfer matrices $\tau(\lambda_i)$ all commute and are simultaneously diagonalisable. The partition function is therefore given by \begin{equation} Z(\{\lambda_i\},\{\mu_j\}) = \sum_{\ell=1}^{2^L} \prod_{i=1}^N \Lambda_{\ell}(\lambda_i), \end{equation} where $\Lambda_{\ell}$ is the $\ell$th eigenvalue of the transfer matrix $\tau$. \subsection{Bethe ansatz diagonalisation} Let us first define the `pseudo-vacuum' states $\ket{\Psi_{\uparrow}}$ and $\ket{\Psi_{\downarrow}}$ by \begin{equation} \label{states} \ket{\Psi_\uparrow} = \bigotimes_{i=1}^{L} \ket{\uparrow} \qquad \qquad \mbox{and} \qquad \qquad \ket{\Psi_\downarrow} = \bigotimes_{i=1}^{L} \ket{\downarrow} \; . \end{equation} An important ingredient in the QISM setup for diagonalising the transfer matrix $\tau(\lambda)$ is the action of the monodromy matrix elements on the states $\ket{\Psi_{\uparrow}}$ and $\ket{\Psi_{\downarrow}}$. For $\ket{\Psi_{\uparrow}}$ these are given by \begin{align} \label{action} A(\lambda) \ket{\Psi_{\uparrow}} &= \prod_{j=1}^L a(\lambda - \mu_j) \ket{\Psi_{\uparrow}} & B(\lambda) \ket{\Psi_{\uparrow}} & = * \nonumber \\ C(\lambda) \ket{\Psi_{\uparrow}} &= 0 & D(\lambda) \ket{\Psi_{\uparrow}} &= \prod_{j=1}^L b(\lambda - \mu_j) \ket{\Psi_{\uparrow}}\; . \end{align} The relations (\ref{action}) follow from the definitions (\ref{mono}) and (\ref{monorep}) together with the triangularity of $\mathcal{L}_{\mathcal{A} j}$ on local states. Due to the Yang-Baxter equation (\ref{YB}) one can readily show that the monodromy matrix (\ref{mono}) satisfies the following quadratic relation \begin{equation} \label{yba} \mathcal{R}(\lambda - \nu) \left[\mathcal{T}(\lambda) \otimes \mathcal{T}(\nu)\right] = \left[\mathcal{T}(\nu) \otimes \mathcal{T}(\lambda)\right] \mathcal{R}(\lambda - \nu) \end{equation} which is known commonly as the Yang-Baxter algebra. Here $\mathcal{R} = P \mathcal{L}$ where $P$ stands for the standard permutation matrix. Thus (\ref{yba}) yields commutation relations for the elements of the monodromy matrix (\ref{monorep}). Among the relations encoded in (\ref{yba}) we shall make use of the following ones, \begin{align} A(\lambda) B(\nu) &= \frac{a(\nu - \lambda)}{b(\nu - \lambda )} B(\nu) A(\lambda) - \frac{c(\nu - \lambda)}{b(\nu - \lambda )} B(\lambda) A(\nu), \nonumber \\ D(\lambda ) B(\nu) &= \frac{a(\lambda - \nu)}{b(\lambda - \nu)} B(\nu ) D(\lambda ) - \frac{c(\lambda - \nu)}{b(\lambda - \nu )} B(\lambda) D(\nu ), \label{alg}\\ B(\lambda) B(\nu ) &= B(\nu ) B(\lambda ) \; .\nonumber \end{align} We are now in a position to state the main theorem of the algebraic Bethe ansatz, or quantum inverse scattering method. \begin{theorem} Let the numbers $\lambda_1,\ldots,\lambda_n$ be solutions of the equations \begin{equation} \label{eq:bae} \prod_{j=1}^L \frac{a(\lambda_i - \mu_j)}{b(\lambda_i - \mu_j) } = (-1)^{n-1}\prod_{k=1}^n \frac{a(\lambda_i - \lambda_k)}{a(\lambda_k - \lambda_i) }. \end{equation} Then, \begin{equation} \label{eq:BAeigvec} \ket{\lambda_1,\ldots,\lambda_n} = B(\lambda_1)\cdots B(\lambda_n) \ket{\Psi_{\uparrow}} \end{equation} are eigenvectors of the transfer matrix $\tau(\lambda,\{\mu_j\})$ with eigenvalues $\Lambda(\lambda,\{\mu_j\})$ given by \begin{align} \label{eq:BAeigval} \Lambda(\lambda,\{\mu_j\}) &= \prod_{j=1}^L a(\lambda - \mu_j) \prod_{k=1}^n \frac{a(\lambda_k-\lambda)}{b(\lambda_k-\lambda)} + \prod_{j=1}^L b(\lambda - \mu_j) \prod_{k=1}^n \frac{a(\lambda-\lambda_k)}{b(\lambda-\lambda_k)} \; . \end{align} \end{theorem} \begin{proof} The theorem follows from the definition \eqref{transfmatrix} of the transfer matrix, the action \eqref{action} on the pseudo-vacuum, and relations \eqref{alg}. In particular, the expression for the eigenvalue originates from \eqref{action} and the first terms on the right hand side of \eqref{alg}. Equations \eqref{eq:bae} imply that unwanted terms arising from the other terms on the right hand side of \eqref{alg} cancel. \end{proof} An important problem in the theory of solvable lattice models is to provide manageable expressions for correlation functions. These are of the form \begin{equation} \bra{\nu_1,\ldots,\nu_m} \mathcal{O} \ket{\lambda_1,\ldots,\lambda_n}, \end{equation} where $\mathcal{O}$ is some operator. The simplest case concerns an expression for the norm of Bethe states, \begin{equation} \label{eq:scprod} N(\{\nu_i\},\{\lambda_j \}) = \bra{\nu_1,\ldots,\nu_n} \lambda_1,\ldots,\lambda_n\rangle. \end{equation} Below we shall see that \eqref{eq:scprod} is equal to a special case of the partition function \eqref{partition2} with fixed boundary conditions. In the next section we show that such partition functions can be realised as multiple contour integrals over a factorised kernel. \section{Fixed boundary conditions} In this section we review some known results for the partition function \eqref{partition2} in the case of fixed boundary conditions. In particular we recall Baxter's formula for such partition functions for general graphs and boundary conditions, expressed in terms of a sum over the symmetric group \cite{Baxter87}. \\ Let us first look at the well known case of domain wall boundary conditions. \subsection{Domain wall boundary conditions} \label{sec:altern} For the case of domain wall boundaries with $N=L$, the edges on the vertical and horizontal borders in Figure \ref{fig:lattice} assume a particular configuration. Naturally one needs to be careful when choosing the boundaries since an inappropriate choice can render a trivial (null) partition function or force the system to freeze in a particular configuration. Domain wall boundaries for the six vertex model were first introduced by Korepin \cite{Korepin82}, and in our notation correspond to the choice $\alpha_{i,1} = \rightarrow$, $\beta_{L+1,j} = \uparrow$ and $\alpha_{i,L+1} = \leftarrow$, $\beta_{1,j} = \downarrow$. This special boundary condition is illustrated in \figref{fig:latticedw}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] \draw[>-<] (-2,0) -- (10,0); \draw[->] (0,-2) -- (0,2); \draw[->] (2,-2) -- (2,2); \draw[->] (4,-2) -- (4,2); \draw[->] (6,-2) -- (6,2); \draw[->] (8,-2) -- (8,2); \begin{scope}[yshift=-2.2cm] \draw[>-<] (-2,0) -- (10,0); \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \end{scope} \begin{scope}[yshift=-4.4cm] \draw[>-<] (-2,0) -- (10,0); \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \end{scope} \begin{scope}[yshift=-6.6cm] \draw[>-<] (-2,0) -- (10,0); \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \end{scope} \begin{scope}[yshift=-8.8cm] \draw[>-<] (-2,0) -- (10,0) ; \draw[<-] (0,-2) -- (0,2); \draw[<-] (2,-2) -- (2,2); \draw[<-] (4,-2) -- (4,2); \draw[<-] (6,-2) -- (6,2); \draw[<-] (8,-2) -- (8,2); \end{scope} \end{tikzpicture} \caption{Domain wall boundary condition.} \label{fig:latticedw} \end{figure} Using the prescriptions given in \secref{sec:vertexmodel}, we can see that the top and bottom boundary represent the states $\ket{\Psi_\uparrow}$ and $\ket{\Psi_\downarrow}$ defined in \eqref{states}. The left and right boundaries in Figure \ref{fig:latticedw} select the elements of the monodromy matrix forming the operator $\mathcal{O} = \prod_{i=1}^{L} B (\lambda_i , \{ \mu_j \})$. Thus our partition function can be written as \begin{equation} \label{pf1} Z(\{\lambda_i\}, \{\mu_j\}) = \bra{\Psi_{\downarrow}} \prod_{i=1}^{L} B (\lambda_i , \{ \mu_j \}) \ket{\Psi_{\uparrow}} \; , \end{equation} where we emphasise that the rapidities $\{\lambda_i\}$ may be chosen arbitrarily and do not have to satisfy the Bethe ansatz equations. It is well known that \eqref{pf1} can be written as the Izergin-Korepin determinant \cite{Izergin87, Korepin82} as well as sums over the permutation group \cite{Khoro_Paku05, Galleas11}. \subsection{Scalar product} The scalar product \eqref{eq:scprod} can be written in a similar fashion as the domain boundary partition function. From the definition of the Bethe eigenstates in \eqref{eq:BAeigvec} it follows immediately that the scalar product \eqref{eq:scprod} can be written as \begin{equation} \label{scprod1} N(\{\nu_i\},\{\lambda_j\})= \bra{\Psi_{\uparrow}} \prod_{i=1}^{n} C (\nu_i , \{ \mu_k \}) \prod_{i=1}^{n} B (\lambda_j , \{ \mu_k \}) \ket{\Psi_{\uparrow}} \; , \end{equation} where the sets of numbers $\{\nu_i\}$ and $\{\lambda_j\}$ each satisfy the Bethe equations \eqref{eq:bae}. Graphically \eqref{scprod1} is depicted in \figref{fig:latticescprod}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] \draw[>-<] (-2,0) -- (12,0) node[right]{$\lambda_n$}; \draw[->] (0,-2) -- (0,2) node[above]{$\mu_1$}; \draw[->] (2,-2) -- (2,2); \draw (4,-2) -- (4,2) ; \draw (5,2) node[above]{\dots}; \draw (6,-2) -- (6,2) ; \draw[->] (8,-2) -- (8,2); \draw[->] (10,-2) -- (10,2) node[above]{$\mu_L$}; \begin{scope}[yshift=-2.2cm] \draw (-2,0) node[left]{$\vdots$} -- (12,0) node[right]{$\vdots$}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \draw (10,-2) -- (10,2); \end{scope} \begin{scope}[yshift=-4.4cm] \draw[>-<] (-2,0) -- (12,0) node[right]{$\lambda_1$}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \draw (10,-2) -- (10,2); \end{scope} \begin{scope}[yshift=-6.6cm] \draw[<->] (-2,0) -- (12,0) node[right]{$\nu_1$}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \draw (10,-2) -- (10,2); \end{scope} \begin{scope}[yshift=-8.8cm] \draw (-2,0) node[left]{$\vdots$} -- (12,0) node[right]{$\vdots$}; \draw (0,-2) -- (0,2); \draw (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (6,-2) -- (6,2); \draw (8,-2) -- (8,2); \draw (10,-2) -- (10,2); \end{scope} \begin{scope}[yshift=-11cm] \draw[<->] (-2,0) -- (12,0) node[right]{$\nu_n$} ; \draw[>-] (0,-2) -- (0,2); \draw[>-] (2,-2) -- (2,2); \draw (4,-2) -- (4,2); \draw (5,-2) node[below]{\dots}; \draw (6,-2) -- (6,2); \draw[>-] (8,-2) -- (8,2); \draw[>-] (10,-2) -- (10,2); \end{scope} \end{tikzpicture} \caption{Scalar product boundary condition.} \label{fig:latticescprod} \end{figure} \subsection{Perimeter Bethe ansatz} \label{sec:perimeter} \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] \draw[dashed] (0,0) circle(6); \draw (160:7) node[left]{9} -- (60:7) node[right]{12}; \draw (15:-7) node[left]{8} -- (35:7) node[right]{1}; \draw (40:-7) node[left]{7} -- (-10:7) node[right]{3}; \draw (60:-7) node[below]{6} -- (20:7) node[right]{2}; \draw (-30:7) node[right]{4} -- (125:7) node[left]{10}; \draw (90:7) node[above]{11} -- (-110:7) node[below]{5}; \end{tikzpicture} \caption{Example of a general planar graph inside a domain $\mathcal{D}$ without multiple crossings at one point for $m=6$.} \label{fig:Zinv} \end{figure} In this section, we review the result of Baxter \cite{Baxter87} for a general $Z$-invariant six-vertex model, adjusting some of his notation to serve our purposes \footnote{in particular, note that Baxter labels the endpoints in an anti-clockwise direction, and writes his formula with respect to the {\it left} rapidities}. The general model is considered on a simply connected convex planar domain $\mathcal{D}$, with $m$ straight lines within it, starting and ending at the boundary \cite{Baxter78,Baxter87}. No three lines are allowed to intersect at a common point, (see Figure~\ref{fig:Zinv} for an example). We fix some point on the boundary, and label the ends of the lines as $1,2,\dots,2m$, in a clockwise direction. The line with endpoints $i$ and $j$ is referred to as `the line $(i,j)$', where $1\leq i<j\leq2m$. We associate a `rapidity' $v_i$ with each endpoint $i$, and for each line $(i,j)$ impose the constraint \begin{equation} v_j=v_i-\gamma. \end{equation} Baxter uses the terminology `right' rapidity and `left' rapidity of a line $(i,j)$ for $v_i$ and $v_j$ respectively. We also define the set of all rapidities $V=\{v_1,\dots,v_{2m} \}$, as well as the set $U$ of all right rapidities. A six-vertex model is constructed by placing arrows on the edges of the lattice so that each vertex (intersection) has two in-pointing arrows in and two out-pointing arrows (the ice-rule). A vertex at the crossing of lines $(i,j)$ and $(k,l)$ with $i<k<j<l$, as in Figure~\ref{fig:vertexinZ}, is given a weight $w(i,j|k,l)$ defined below. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.2,line width=0.4mm] \draw[dashed] (0,0) circle(6); \draw[color=lightgray] (160:7) -- (60:7); \draw[color=lightgray] (40:-7) -- (-10:7); \draw[color=lightgray] (60:-7) -- (20:7); \draw[color=lightgray] (-30:7) -- (125:7); \draw (15:-7) node[left]{$j$} -- (35:7) node[right]{$i$}; \draw (90:7) node[above]{$l$} -- (-110:7) node[below]{$k$} ; \end{tikzpicture} \caption{A vertex at the crossing of lines $(i,j)$ and $(k,l)$ with $i<k<j<l$ has weight $w(i,j|k,l)$.} \label{fig:vertexinZ} \end{figure} The six possible arrangements as viewed from the boundary points $k$ and $j$ are shown in Figure~\ref{fig:6vertexZinv}. \begin{figure}[h]\centering \begin{tikzpicture}[scale=0.5,line width=0.4mm] \draw[>->] (15:-1) -- (35:1); \draw[<-<] (90:1) -- (-110:1) node[below]{$w_1$}; \draw[xshift=4cm,<-<] (15:-1) -- (35:1); \draw[xshift=4cm,>->] (90:1) -- (-110:1) node[below]{$w_2$}; \draw[xshift=8cm,>->] (15:-1) -- (35:1); \draw[xshift=8cm,>->] (90:1) -- (-110:1) node[below]{$w_3$}; \draw[xshift=12cm,<-<] (15:-1) -- (35:1); \draw[xshift=12cm,<-<] (90:1) -- (-110:1) node[below]{$w_4$}; \draw[xshift=16cm,>-<] (15:-1) -- (35:1); \draw[xshift=16cm,<->] (90:1) -- (-110:1) node[below]{$w_5$}; \draw[xshift=20cm,<->] (15:-1) -- (35:1); \draw[xshift=20cm,>-<] (90:1) -- (-110:1) node[below]{$w_6$}; \end{tikzpicture} \caption{Six vertex configurations on vertices inside $\mathcal{D}$.} \label{fig:6vertexZinv} \end{figure} Boltzmann weights are assigned to the six configurations as follows (in the notation of this paper): \begin{align} w_1=w_2 &=1,\nonumber \\ w_3=w_4 &=\frac{b(v_l-v_k)}{a(v_l-v_k)} = \frac{\sinh(v_l-v_k)}{\sinh(v_l-v_k+\gamma)},\\ w_5=w_6 &=\frac{c(v_l-v_k)}{a(v_l-v_k)}=\frac{\sinh(\gamma)}{\sinh(v_l-v_k+\gamma)}.\nonumber \end{align} We also fix boundary conditions, with $m$ arrows pointing into $\mathcal{D}$ and $m$ arrows pointing out. We specify these boundary arrows with the set of endpoint locations where an out arrow occurs, $X=\{ x_1,\dots,x_m \}$. The partition function of the model is then a function of $X$ and $V$. We now come to the main result (for our purposes) of \cite{Baxter87}: \begin{theorem}[Baxter] \label{th:baxter} Let $V=\{v_1,\ldots,v_{2m}\}$ be the set of all rapidities, and let $U$ be the set of right rapidities. The partition function $Z(X|V)$ for the Z invariant six-vertex model is given by \begin{align} Z(X|V)&:= \sum_{\text{cfgs.}} \prod_{(i,j|k,l)} w(i,j|k,l) \nonumber \\ &= C^{-1} \sum_{P} \prod_{1\leq i<j\leq m} \frac{a(u_j - u_i)}{b(u_j - u_i)} \prod_{i=1}^m \phi_{x_i}(u_i) \end{align} where \begin{align} \label{eq:phidef} \phi_x(u)&= \prod_{j=1}^{x-1} a(v_j-u) \prod_{j=x+1}^{2m} b(v_j-u), \end{align} and the pre-factor is defined by \begin{equation} C=(-1)^{n_\uparrow}\sinh(\gamma)^{-m}\prod_{1\leq i,j\leq m} a(u_i-u_j)^2, \end{equation} where $n_\uparrow$ is the number of out-point arrows with coordinate $x\leq m$ and the variables $u_i$ refer to the elements of the set of right rapidities $U$. \end{theorem} \section{Integral expression} In this section we will derive a multiple integral expression for the partition function \eqref{partition2} in the case of fixed boundary conditions. Theorem~\ref{th:baxter} provides an explicit expression derived by Baxter \cite{Baxter87} for such partition functions for general graphs and boundary conditions. This expression is in terms of a large sum over the symmetric group and in this form not immediately useful. We hope our integral expressions are more manageable and may lead to further progress for correlation functions as well as higher rank cases. We wish to write the sum over permutations as an integral, summing over residues. We introduce auxiliary integration variables $w_i$, $i=1,\dots,m$. This leads to the following result. \begin{theorem} Let $V=\{v_1,\ldots,v_{2m}\}$ be the set of all rapidities, and let $U$ be the set of left rapidities. \begin{multline} Z(X|V) = \frac{C^{-1} }{(2\pi\mathrm{i})^m}\oint \dots \oint \frac{\prod_{i=1}^m\prod_{j=i+1}^{m} a(w_j-w_i)b(w_i-w_j)} {\prod_{i,j=1}^{m} b(w_i-u_j) } \prod_{i=1}^{m} \phi_{x_i}(w_i)\ \mathrm{d} w_1 \dots \mathrm{d} w_{m}. \label{eq:Baxter_int} \end{multline} The contours of integration are around the poles at $w_i=u_j$, $i,j=1,\ldots,m$ and the prefactor is again given by \begin{equation} C=(-1)^{n_\uparrow}\sinh(\gamma)^{-m}\prod_{1\leq i,j\leq m} a(u_i-u_j)^2, \end{equation} and $\phi_x(w)$ and $n_\uparrow$ are as in Theorem~\ref{th:baxter}. \end{theorem} \begin{proof} The factor $\prod_{j>i} b(w_i-w_j)$ in the denominator implies that the only nonzero contributions arise from residues at different poles, i.e. if $\{w_1,\ldots,w_m\}$ is a permutation of $\{u_1,\ldots,u_m\}$. \end{proof} \subsection{Domain wall boundary conditions} \label{sec:integralformDWBC} We now consider the result of Baxter \cite{Baxter87}, specialised to a $L$ by $L$ square lattice with domain wall boundary conditions. Using the notation $\bar{\mu}_j=\mu_{L-j+1}$, we have $4L$ rapidities though only $2L$ are independent. The full set of rapidities are given by \begin{equation} V=\{\mu_1,\dots,\mu_L, \bar{\lambda}_1,\dots,\bar{\lambda}_L,\bar{\mu}_1-\gamma,\dots,\bar{\mu}_L-\gamma,\lambda_1-\gamma,\dots,\lambda_L-\gamma \} \end{equation} while the set of left rapidities is \bea U=\{ u_1,\dots,u_{2L} \}=\{\mu_1,\dots,\mu_L ,\bar{\lambda}_1,\dots,\bar{\lambda}_L\}. \eea We take boundary conditions of out arrows being on the vertical lines, and in arrows on the horizontal lines according to \figref{fig:latticedw}. Thus the positions of the out arrows are \bea X=\{ 1,\dots, L, 2L+1, \dots, 3L \}. \eea With these definitions we have that the domain wall boundary partition function \eqref{pf1} is equal to \begin{equation} Z(\{\lambda_i\},\{\mu_j\}) = \prod_{i,j=1}^L a(\lambda_i-\mu_j) Z(X|V). \end{equation} The domain wall boundaries specialisation of \eqref{eq:Baxter_int} implies that the only nonzero contributions arise from residues at $w_i=\mu_i$ for $i=1,\ldots,L$, and if $\{w_{L+1},\ldots,w_{2L}\}$ is a permutation of $\{\lambda_{1},\ldots,\lambda_{L}\}$. With this specialisation, and the re-labelling of the integration variables $w_{L+i} \rightarrow w_i$, the formula becomes \begin{multline}\label{eq:int} Z(\{\lambda_i\},\{\mu_j\}) = \left(\frac{\sinh\gamma}{2\pi \mathrm{i}}\right)^L \oint \dots \oint \frac{\prod_{i=1}^L\prod_{j=i+1}^{L} a({w}_j-{w}_i)b({w}_j-{w}_i)} {\prod_{i,j=1}^L b({w}_i-\lambda_j)}\\ \times \prod_{i=1}^L \prod_{j=1}^{i-1} b(\bar{\mu}_j-{w}_i)\prod_{j=i+1}^{L} a({w}_i-\bar{\mu}_j)\ \mathrm{d}{w}_1 \dots \mathrm{d} {w}_L \end{multline} where the integrals are around the poles at ${w}_i=\lambda_j$. \subsubsection{Homogeneous limit} \label{sec:homogeneous} The integral formula \eqref{eq:int}, enables us to readily calculate the homogeneous limit of the partition function. With the changes of variables \bea\label{eq:cofv} w=\frac{1}{2}\log \bigg( \frac{1-t x}{t- x} \bigg),\quad \lambda = \frac{1}{2}\log \bigg(\frac{1-t z}{t- z}\bigg),\quad \mu = \frac{1}{2}\log \bigg(\frac{1-t y}{t- y}\bigg),\quad t=\e^{\gamma}, \eea we obtain \begin{multline} Z= \frac{c^{L^2}}{(2\pi \mathrm{i})^L} \oint \dots \oint \frac{\prod_{i=1}^L\prod_{j=i+1}^{L} (x_i-x_j)(1+\kappa x_j+x_ix_j)} {\prod_{i,j=1}^L (z_j-x_i)} \\ \times \prod_{i=1}^L\prod_{j=1}^L \frac{(y_j-z_i)(1+\kappa z_i+ y_jz_i)}{(1+\kappa y_j +y_j^2)^{1/2}(1+\kappa z_i +z_i^2)^{1/2}} \\ \times \prod_{i=1}^L \bigg( \prod_{j=1}^i \frac{1}{(1+\kappa x_i +x_i y_j)} \prod_{j=i}^L \frac{1}{(y_j-x_i)} \bigg)\ \mathrm{d} x_1 \dots \mathrm{d} x_L, \end{multline} where $\kappa=-(t+1/t)$. We now set $z_i=z$ and $y_i=y$ for all $i$ ({\it i.e. } $\gl_i=\gl$, $\mu_i=\mu$), \begin{multline} Z= \frac{c^{L^2}}{(2\pi \mathrm{i})^L} \oint \dots \oint \frac{\prod_{i=1}^L \prod_{j=i+1}^{L} (x_i-x_j)(1+\kappa x_j+x_ix_j)} {\prod_{i=1}^L (z-x_i)^L} \\ \times \frac{(y-z)^{L^2}(1+\kappa z+ y z)^{L^2}}{(1+\kappa y +y^2)^{L/2}(1+\kappa z +z^2)^{L/2}}\prod_{i=1}^L \frac{1}{(z_j-x_i)} \\ \times \prod_{j=1}^i \frac{1}{(1+\kappa x_i +x_i y)^L} \prod_{j=i}^L \frac{1}{(y-x_i)^{L+1-i}}\ \mathrm{d} x_1 \dots \mathrm{d} x_L. \end{multline} Also the limit $a=b=c=1$, where the partition function counts the number of alternating sign matrices \cite{Kuperberg96}, is obtained as follows. We first set $a(\lambda,\mu)=b(\lambda,\mu)$, {\it i.e.} $\lambda-\mu=-\gamma/2+\mathrm{i} \pi/2$. In the transformed variables we achieve this by setting $z=0$ and $y=1$, corresponding to $\lambda=-\gamma/2$ and $\mu=-\mathrm{i}\pi/2$. We also normalise each of these weights to $1$ by dividing through by $a(\lambda,\mu)$ and thus obtain \begin{multline}\label{eq:hom} Z_{\rm ASM}= \frac{(c/a)^{L^2}}{(2+\kappa)^{L^2/2}(2\pi \mathrm{i})^L} \oint \dots \oint \frac{\prod_{i=1}^L\prod_{j=i+1}^{L} (x_i-x_j)(1+\kappa x_j+x_ix_j)} {\prod_{i=1}^L (-x_i)^{L}} \\ \times \prod_{i=1}^L (1+(\kappa +1)x_i )^i (1-x_i)^{-L+i-1} \ \mathrm{d} x_1 \dots \mathrm{d} x_L. \end{multline} (Note that we would finally get $(a=b=c=1)$ by setting $\gamma=\mathrm{i} \pi/3$, {\it i.e.} $\kappa=-1$). Alternatively we set $\kappa=\tau-2$, and write equation \eqref{eq:hom} as the following constant term expression: \begin{multline} A_L(\tau)= {\rm CT}\left[ \prod_{i=1}^L\prod_{j=i+1}^{L} (x_j-x_i)(1+(\tau-2) x_j+x_ix_j)\right.\\ \left.\prod_{i=1}^L \frac{1}{x_i^{L-1}(1-x_i)^{L-i+1}(1+(\tau-1) x_i)^{i}} \right], \end{multline} which yields polynomials in $\tau$ corresponding to the enumeration of alternating sign matrices, by number of $-1$s. This latter formula is very similar to expressions obtained in \cite{DiFZJ2007,Zeilb07,FonsecaZJ09,deGier_Lascoux_Sorrell10,FonsecaNadeau11} as polynomial solutions of the $q$-deformed Knizhnik-Zamolodchikov equation and their relation to the combinatorics of alternating sign matrices and symmetric plane partitions. \subsection{Scalar Product} \label{sec:scalar_product} Next we consider the result of Baxter \cite{Baxter87} specialised to the $2n \times L$ rectangular lattice depicted in \figref{fig:latticescprod}. This specialisation renders the norm of Bethe vectors (\ref{scprod1}), and according to the conventions of \Secref{sec:perimeter} we set $m=L+2n$. We have rapidities \bea \label{vv} V=\{ \mu_1,\dots, \mu_{L}, \bar{\lambda}_1 , \dots , \bar{\lambda}_n , \nu_1 , \dots , \nu_n, \bar{\mu}_1 - \gamma , \dots , \bar{\mu}_L - \gamma, \bar{\nu}_1 -\gamma, \dots , \bar{\nu}_n -\gamma , \lambda_1 -\gamma , \dots , \lambda_n -\gamma \} \nonumber \\ \eea and the set of left rapidities is then given by \bea \label{uu} U=\{ \mu_1,\dots, \mu_{L}, \bar{\lambda}_1 , \dots , \bar{\lambda}_n , \nu_1 , \dots , \nu_n \} \; . \eea Here we use the definitions $\bar{\mu}_j = \mu_{L+1-j}$, $\bar{\lambda}_j = \lambda_{n+1-j}$ and $\bar{\nu}_j = \nu_{n+1-j}$. In order to characterise the scalar product (\ref{scprod1}) we also need to specify the location of the out arrows $X$. We have these out arrows appearing in three distinct blocks with positions: \bea\label{xx} X=\{ 1,\dots,L, L+n+1, \dots, L+2n,2L+2n+1,\dots,2L+3n \} \eea Similarly to the case of domain wall boundaries, the norm (\ref{scprod1}) consists of the partition function $Z(X|V)$ up to a normalisation factor arising from the normalisation of the statistical weights. Thus we have \begin{equation} N(\{\nu_i\},\{\lambda_j\}) = \prod_{i=1}^n \prod_{j=1}^L a(\lambda_i-\mu_j) a(\nu_i-\mu_j) Z(X|V). \end{equation} with $Z(X|V)$ computed using (\ref{vv}), (\ref{uu}) and (\ref{xx}). With this specialisation the integrand of (\ref{eq:Baxter_int}) has poles only at $w_i = \mu_i$ for $i=1, \dots , L$, (see Appendix \ref{spc}) and thus the integration over this particular subset of variables is trivial. After performing the trivial integrations and some re-arrangments detailed in Appendix~\ref{spc} we are left with \begin{align} \label{eq:intsp} N(\{\nu_i\},\{\lambda_j\}) = & \frac{(-1)^{(L-1)n} (\sinh{\gamma})^{2n} }{(2\pi \mathrm{i})^{2n}} \prod_{i,j=1}^{n} \frac{1}{a(\lambda_i - \nu_j) a(\nu_j - \lambda_i) a(\nu_i - \nu_j)^2} \nonumber \\ & \times \oint \dots \oint \mathrm{d}{w}_{1} \dots \mathrm{d} w_{2n} \frac{\prod_{i=1}^{2n}\prod_{j=i+1}^{2n} a({w}_{j}-{w}_{i}) b({w}_{j}-{w}_{i})} {\prod_{i=1}^{2n} \prod_{j=1}^{n} b({w}_{i}-\lambda_j) b({w}_{i}-\nu_j) } \nonumber\\ & \quad \times \prod_{i=1}^{n} \prod_{j=1}^{L} a({w}_{i}-\mu_j) b(\mu_j - {w}_{n+i}) \prod_{i=1}^{n} \prod_{j=1}^{n} a({w}_{i}-\nu_j)a(\nu_j - {w}_{n+i})\nonumber\\ &\quad\times \prod_{i=1}^{n} \prod_{j=i+1}^{n} b(\nu_j - {w}_{i}) b({w}_{2n+1-i} - \nu_j ) a(\nu_i - {w}_{j}) a({w}_{2n+1-j} - \nu_i ) \; . \end{align} \section{Three-point functions in $\mathcal{N}=4$ Super Yang-Mills} \label{sec:3pt} The formul\ae (\ref{eq:intsp}) for the scalar product of Bethe vectors is valid for arbitrary values of the complex parameters $\{\nu_i\}$ and $\{\lambda_j\}$. On the other hand, the expression (\ref{eq:intsp}) will yield the norm of the transfer matrix eigenstate (\ref{eq:BAeigvec}) only when the set $\{\nu_i\}$ is related to the set $\{\lambda_j\}$, i.e. $\nu_j = \lambda^{*}_j$, and with the set $\{\lambda_j\}$ subjected to the Bethe ansatz equations (\ref{eq:bae}). In that case Slavnov's formula \cite{Slavnov_89} provides a rather simple expression for the scalar product in terms of a determinant of a $n \times n$ matrix. For generic parameters $\{\nu_i\}$ and $\{\lambda_j\}$, one would have to consider the formul\ae of \cite{Korepin82,Korepin_book}, given in terms of a large sum over partitions. However, the recent developments in the computation of three-point functions in $\mathcal{N}=4$ Super Yang-Mills (SYM) seems to require manageable expressions for the norms of Bethe vectors for general values of the parameters and here we shall illustrate how the integral formula (\ref{eq:intsp}) can be embedded in the framework of \cite{Escobedo_2011}. Three-point correlation functions of primary operators in $\mathcal{N}=4$ SYM theory are constrained by conformal invariance to have the form \begin{equation} \langle \mathcal{O}_i (x_i) \mathcal{O}_j (x_j) \mathcal{O}_k (x_k) \rangle \; = \frac{\sqrt{\mathcal{N}_i \mathcal{N}_j \mathcal{N}_k} \; c_{ijk}}{| x_i - x_j |^{\Delta_i + \Delta_j - \Delta_k } | x_j - x_k |^{\Delta_j + \Delta_k - \Delta_i } | x_k - x_i |^{\Delta_k + \Delta_i - \Delta_j }}, \end{equation} where $\mathcal{N}_j$ are normalization factors and $\Delta_i$ are the respective conformal dimensions. The dimensions $\Delta_i$ are already fixed in the computation of two-point functions and a large literature is devoted to that problem in the planar limit of the $\mathcal{N}=4$ SYM, exploring integrable structures. For a detailed revision on this subject see for instance \cite{Beisert} and references therein. Here the quantity we wish to evaluate is the structure constant $c_{ijk}$, which has been worked out to leading order in \cite{Escobedo_2011} employing the spin chain picture of single trace operators introduced in \cite{Minahan_2002}. See also \cite{Okuyama_2004} and \cite{Roiban_2004} for previous works on $\mathcal{N}=4$ SYM three-point functions within the spin chain picture. The structure constant $c_{ijk}$ in the $\mathcal{N}=4$ SYM admits the perturbative expansion $c_{ijk} = \frac{1}{N_c} c^{(0)}_{ijk} + \frac{\theta}{N_c} c^{(1)}_{ijk} + \frac{\theta^2}{N_c} c^{(2)}_{ijk} + \ldots$, where $N_c$ and $\theta$ are respectively the number of colours and the 't Hooft coupling constant. We shall consider the leading order term $c^{(0)}_{123}$ as described in \cite{Escobedo_2011} and in order to compute it from our results we first need to consider a special limit usually refereed as rational limit. We start by performing the change of variables \begin{equation} \label{rat1} \lambda = \bar{\gamma} (u - \tfrac{\mathrm{i}}{2}) \quad \quad \mbox{and} \quad \quad \gamma = \mathrm{i} \bar{\gamma} \; , \end{equation} and we also set the inhomogeneities $\mu_j = 0$. From (\ref{su2})-(\ref{monorep}) we find that \begin{equation} \label{rat2} \lim_{\genfrac{}{}{0pt}{}{\bar{\gamma} \rightarrow 0}{\mu_j \rightarrow 0}} \frac{1}{\sinh{\bar{\gamma}}^{L}} \mathcal{T}(\lambda) = \overline{\mathcal{T}}(u) \; , \end{equation} where $\overline{\mathcal{T}}(u)$ corresponds to the monodromy matrix considered in \cite{Escobedo_2011}. In practice the above limit corresponds to considering our results with statistical weights $a$, $b$ and $c$ given by \begin{eqnarray} \label{rat} a &=& u + \tfrac{\mathrm{i}}{2}, \nonumber \\ b &=& u - \tfrac{\mathrm{i}}{2}, \nonumber \\ c &=& \mathrm{i} \; . \end{eqnarray} Now we consider three independent $XXX$-type spin chains with periodic boundary conditions which we shall refer as $\mathcal{A}_1$, $\mathcal{A}_2$ and $\mathcal{A}_3$. Those spin chains can be obtained as the logarithmic derivative of the transfer matrix (\ref{transfmatrix}) up to normalisation factors after considering the rational limit above described. This is a standard construction and a detailed description can be found in \cite{Korepin_book}. Within the spin chain picture introduced in \cite{Minahan_2002}, each single trace operator $\mathcal{O}_i$ in the $SU(2)$ sector of the $\mathcal{N}=4$ SYM at one-loop order is then represented by the spin chain $\mathcal{A}_i$ under the zero momentum condition. The state of each spin chain $\mathcal{A}_j$ with $L_j$ sites is characterised by rapidities satisfying Bethe ansatz equations. For instance we use the set of $m_1$ rapidities $\{ u_1 , \dots , u_{m_1} \}$ to characterise the spin chain $\mathcal{A}_1$, the $m_2$ rapidities $\{ v_1 , \dots , v_{m_2} \}$ for $\mathcal{A}_2$ and $m_3$ rapidities $\{ w_1 , \dots , w_{m_3} \}$ for $\mathcal{A}_3$. They are subjected to the following Bethe ansatz equations: \begin{align} \label{eq:BAE} \left[ \frac{a(u_i)}{b(u_i)} \right]^{L_1} &= \prod_{\stackrel{k=1}{k \neq i}}^{m_1} - \frac{a(u_i - u_k + \frac{\mathrm{i}}{2})}{a(u_k - u_i + \frac{\mathrm{i}}{2}) }, \qquad \qquad i = 1, \dots , m_1\; , \nonumber \\ \left[ \frac{a(v_i)}{b(v_i)} \right]^{L_2} &= \prod_{\stackrel{k=1}{k \neq i}}^{m_2} - \frac{a(v_i - v_k + \frac{\mathrm{i}}{2})}{a(v_k - v_i + \frac{\mathrm{i}}{2}) }, \qquad \qquad i = 1, \dots , m_2\; , \nonumber \\ \left[ \frac{a(w_i)}{b(w_i)} \right]^{L_3} &= \prod_{\stackrel{k=1}{k \neq i}}^{m_3} - \frac{a(w_i - w_k + \frac{\mathrm{i}}{2})}{a(w_k - w_i + \frac{\mathrm{i}}{2}) }, \qquad \qquad i = 1, \dots , m_3 \; , \end{align} and the zero momentum condition translates into \begin{equation} \prod_{i=1}^{m_1} \frac{a(u_i)}{b(u_i)} = \prod_{i=1}^{m_2} \frac{a(v_i)}{b(v_i)} = \prod_{i=1}^{m_3} \frac{a(w_i)}{b(w_i)} = 1 \; . \end{equation} We remark that now we are considering the statistical weights $a$, $b$ and $c$ given in (\ref{rat}). Next we define the functions \begin{align} E_2 &= \prod_{j=1}^{m_2} \left[ \frac{a(v_j)}{b(v_j)} \right]^{L_2}, & F_2 &= \prod_{i=1}^{m_2} \prod_{j=i+1}^{m_2} \frac{a(v_i - v_j + \frac{\mathrm{i}}{2})}{b(v_i - v_j + \frac{\mathrm{i}}{2})}, \nonumber \\ F_1 &= \prod_{i=1}^{m_1} \prod_{j=i+1}^{m_1} \frac{a(u_i - u_j + \frac{\mathrm{i}}{2})}{b(u_i - u_j + \frac{\mathrm{i}}{2})}, & G_2 &= \prod_{j=1}^{m_2} \prod_{i=j+1}^{m_2} \frac{a(v_i - v_j + \frac{\mathrm{i}}{2})}{b(v_i - v_j + \frac{\mathrm{i}}{2})}, \end{align} and separate the set of rapidities $\{ u \} = \{ u_1 , \dots , u_{m_1} \}$ into two subsets $\{ \alpha \} $ and $\{ \bar{\alpha} \} $ such that $\{ \alpha \} \cup \{ \bar{\alpha} \} = \{ u \}$. There are $2^{m_1}$ ways to partitionate $\{ u \}$ into $\{ \alpha \} $ and $\{ \bar{\alpha} \}$, and each possibility is simply recasted as \begin{equation} \{ \alpha \} = \{ u_1 , \dots , u_{n_1} \} \qquad \mbox{and} \qquad \{ \bar{\alpha} \} = \{ \bar{u}_1 , \dots , \bar{u}_{\bar{n}_1} \} , \end{equation} with $m_1 = n_1 + \bar{n}_1$. For each partitioning we also define the functions \begin{equation} E(\bar{\alpha}) = \prod_{j=1}^{\bar{n}_1} \left[ \frac{a(\bar{u}_j)}{b(\bar{u}_j)} \right]^{L_1 +1}, \qquad \qquad F(\alpha ,\bar{\alpha}) = \prod_{i=1}^{n_1} \prod_{j=1}^{\bar{n}_1} \frac{a(u_i - \bar{u}_j + \frac{\mathrm{i}}{2})}{b(u_i - \bar{u}_j + \frac{\mathrm{i}}{2})} \; \; , \end{equation} and following \cite{Escobedo_2011} the structure constant for three-point functions at leading order in planar $\mathcal{N}=4$ SYM is given by \begin{equation} \label{eq:c0} c^{(0)}_{123} = \sqrt{\frac{L_1 L_2 L_3}{\mathcal{N}_1 \mathcal{N}_2 \mathcal{N}_3}} \frac{E_2 G_2}{F_2 F_1} (-1)^{m_1} \sum_{\alpha , \bar{\alpha}} E(\bar{\alpha}) F(\alpha , \bar{\alpha}) S_2 (\alpha) S_3 (\bar{\alpha}) , \end{equation} under the constraint $m_1 =m_2 + m_3$. The normalisation factors are \begin{align} \label{eq:norms} \mathcal{N}_1 = & (-1)^{m_1} \prod_{j=1}^{m_1} b(u_j)^{-L_1} a(u^{*}_j)^{-L_1} a(u_j) b(u^{*}_j) \prod_{i=1}^{m_1} \prod_{j=i+1}^{m_1} \frac{b(u_i - u_j + \frac{\mathrm{i}}{2})}{a(u_i - u_j + \frac{\mathrm{i}}{2})} \nonumber \\ & \times \prod_{j=1}^{m_1} \prod_{i=j+1}^{m_1} \frac{b(u^{*}_i - u^{*}_j + \frac{\mathrm{i}}{2})}{a(u^{*}_i - u^{*}_j + \frac{\mathrm{i}}{2})} N_{L_1}(\{ u^{*} \} , \{ u \} ), \nonumber \\ \mathcal{N}_2 = & (-1)^{m_2} \prod_{j=1}^{m_2} b(v_j)^{-L_2} a(v^{*}_j)^{-L_2} a(v_j) b(v^{*}_j) \prod_{i=1}^{m_2} \prod_{j=i+1}^{m_2} \frac{b(v_i - v_j + \frac{\mathrm{i}}{2})}{a(v_i - v_j + \frac{\mathrm{i}}{2})} \nonumber \\ & \times \prod_{j=1}^{m_2} \prod_{i=j+1}^{m_2} \frac{b(v^{*}_i - v^{*}_j + \frac{\mathrm{i}}{2})}{a(v^{*}_i - v^{*}_j + \frac{\mathrm{i}}{2})} N_{L_2}(\{ v^{*} \} , \{ v \} ), \nonumber \\ \mathcal{N}_3 = & (-1)^{m_3} \prod_{j=1}^{m_3} b(w_j)^{-L_3} a(w^{*}_j)^{-L_3} a(w_j) b(w^{*}_j) \prod_{i=1}^{m_3} \prod_{j=i+1}^{m_3} \frac{b(w_i - w_j + \frac{\mathrm{i}}{2})}{a(w_i - w_j + \frac{\mathrm{i}}{2})} \nonumber \\ & \times \prod_{j=1}^{m_3} \prod_{i=j+1}^{m_3} \frac{b(w^{*}_i - w^{*}_j + \frac{\mathrm{i}}{2})}{a(w^{*}_i - w^{*}_j + \frac{\mathrm{i}}{2})} N_{L_3}(\{ w^{*} \} , \{ w \} ) \; \; , \end{align} where $N_{L}$ denotes the scalar product of Bethe vectors (\ref{eq:intsp}) for a lattice of length $L$ after the limit (\ref{rat1}). The function $N_{L}$ turns out to be explicitly given by \begin{align} \label{eq:splim} N_L (\{\nu \},\{\lambda\})& = \frac{1}{(2\pi \mathrm{i})^{2n}} \prod_{i,j=1}^{n} \frac{1}{a(\lambda_i - \nu_j+ \frac{\mathrm{i}}{2}) a(\nu_j - \lambda_i+ \frac{\mathrm{i}}{2}) a(\nu_i - \nu_j+ \frac{\mathrm{i}}{2})^2} \nonumber \\ & \times \oint \dots \oint \mathrm{d}{w}_{1} \dots \mathrm{d} w_{2n} \frac{\prod_{i=1}^{2n}\prod_{j=i+1}^{2n} a({w}_{j}-{w}_{i}+ \frac{\mathrm{i}}{2}) b({w}_{j}-{w}_{i}+ \frac{\mathrm{i}}{2})} {\prod_{i=1}^{2n} \prod_{j=1}^{n} b({w}_{i}-\lambda_j+ \frac{\mathrm{i}}{2}) b({w}_{i}-\nu_j+ \frac{\mathrm{i}}{2}) } \nonumber\\ & \; \; \times \prod_{i=1}^{n} [ a({w}_{i}) b({w}_{n+i}) ]^L \prod_{i=1}^{n} \prod_{j=1}^{n} a({w}_{i}-\nu_j+ \tfrac{\mathrm{i}}{2})a(\nu_j - {w}_{n+i}+ \tfrac{\mathrm{i}}{2})\nonumber\\ & \; \; \times \prod_{i=1}^{n} \prod_{j=i+1}^{n} b(\nu_j - {w}_{i}+ \tfrac{\mathrm{i}}{2}) b({w}_{2n+1-i} - \nu_j+ \tfrac{\mathrm{i}}{2} ) a(\nu_i - {w}_{j}+ \tfrac{\mathrm{i}}{2}) a({w}_{2n+1-j} - \nu_i + \tfrac{\mathrm{i}}{2}), \end{align} with the functions $a$ and $b$ as described in (\ref{rat}). The number of excitations $n$ is understood by the number of rapidities in the argument. In its turn the function $S_2 (\alpha)$ is given by \begin{align} \label{eq:s2} S_2 (\alpha) = & \prod_{j=1}^{n_1} b(u_j)^{m_3 - L_1} a(u_j) \prod_{j=1}^{m_2} a(v_j)^{m_3 - L_1} b(v_j) \prod_{j=1}^{m_2} \prod_{i=j+1}^{m_2} \frac{b(v_i - v_j + \frac{\mathrm{i}}{2})}{a(v_i - v_j + \frac{\mathrm{i}}{2})} \nonumber \\ & \times N_{L_1 - m_3} (\{ v \} , \{ u \}) \; \delta_{n_1 , m_2}\; , \end{align} while \begin{align} \label{eq:s3} S_3 (\bar{\alpha}) = & \prod_{j=1}^{m_3} b(w_j)^{-m_3} a(w_j) \prod_{j=1}^{\bar{n}_1} a(\bar{u}_j)^{-m_3} b(\bar{u}_j) \prod_{i=1}^{m_3} \prod_{j=i+1}^{m_3} \frac{b(w_i - w_j + \frac{\mathrm{i}}{2})}{a(w_i - w_j + \frac{\mathrm{i}}{2})} \nonumber \\ & \times Z_{m_3} (\{ \bar{u} \} ) Z_{m_3} (\{ w \} ) \; \delta_{\bar{n}_1 , m_3} \; . \end{align} In the expressions (\ref{eq:s2}) and (\ref{eq:s3}), $\delta_{i,j}$ stands for the standard Kronecker delta and $Z_{m_3}$ consists of the six-vertex model partition function with domain wall boundaries (\ref{eq:int}) for a lattice with dimensions $m_3 \times m_3$ in the limit (\ref{rat1})-(\ref{rat}) previously discussed. For completeness it is given by \begin{multline} \label{eq:int1} Z_{m_3} (\{\lambda\}) = \frac{(-1)^{m_3 (m_3 -1)/2}}{(2\pi)^{m_3}} \oint \dots \oint \frac{\prod_{i=1}^{m_3}\prod_{j=i+1}^{m_3} a({w}_j-{w}_i + \tfrac{\mathrm{i}}{2})b({w}_j-{w}_i+ \tfrac{\mathrm{i}}{2})} {\prod_{i,j=1}^{m_3} b({w}_i-\lambda_j+ \frac{\mathrm{i}}{2})} \\ \times\prod_{i=1}^{m_3} b({w}_i)^{i-1} a({w}_i)^{m_3 -i}\ \mathrm{d}{w}_1 \dots \mathrm{d}{w}_{m_3} \; , \end{multline} recalling that $a$ and $b$ correspond to the functions in (\ref{rat}). In this way, to compute the structure constant leading term $c^{(0)}_{123}$ according to the prescriptions of \cite{Escobedo_2011}, one needs to substitute (\ref{eq:splim}) in (\ref{eq:norms}) and (\ref{eq:s2}), and then insert the results into (\ref{eq:c0}). Similarly we also need to substitute (\ref{eq:int1}) in (\ref{eq:s3}) and replace the results in (\ref{eq:c0}). In summary, the evaluation of $c^{(0)}_{123}$ is thus reduced to the evaluation of the countour integrals (\ref{eq:splim}) and (\ref{eq:int1}). \section{Conclusion} The main result of this paper is to write the partition function for a general $Z$-invariant six-vertex model as a multiple contour integral over a factorised polynomial kernel. This allows us to write a new representation for the partition function of the six vertex model with domain wall boundary conditions as a multiple contour integral. Likewise we derive an off-shell multiple integral expression for the scalar product of two Bethe vectors of the six-vertex model transfer matrix. The study of two-point functions in $\mathcal{N}=4$ Super Yang-Mills (SYM) has advanced dramatically due to the presence of integrable structures, and the computation of three-point functions at weak coupling seems to be benefited as well \cite{Escobedo_2011}. In this sense we have also illustrated in Section~\ref{sec:3pt} how the integral formul\ae presented in Sections~\ref{sec:integralformDWBC} and \ref{sec:scalar_product} can be used in that context. Although in a different fashion, recently the connection was realised between the partition function of a system of statistical mechanics and the computation of norms of Bethe vectors in a particular limit \cite{Gromov_2011}. We hope the results presented here to help understanding this relation in more general cases. The integral formul\ae we obtain are very similar to those in multi-variate polynomial solutions of the $q$-deformed Knizhnik-Zamolodchikov equation, a connection which deserves future exploration. In \cite{DiFZJ2007,Zeilb07,FonsecaZJ09,deGier_Lascoux_Sorrell10,FonsecaNadeau11} methods have been developed to compute with such multi-variate expressions which we hope will be applicable in the current context as well. We further hope that our representation for the scalar product can be extended to compute correlation functions, providing an alternative to the methods developed in \cite{KitanineMT99,IzerginKMT99,KitanineMT00}, and also constitutes an approach which may be generalised to higher rank solvable lattice models. \section*{Acknowledgments} Our warm thanks goes to Michael Wheeler for instructive discussions, and we gratefully acknowledge financial support from the Australian Research Council (ARC).
\section{Introduction} A two dimensional electron gas (2DEG) in strong perpendicular magnetic field exhibits the regime of quantum Hall effect\cite{qhe} (QHE). One of the peculiar phenomena specific to this regime is the appearance of one dimensional (1D) {\em chiral} edge states, which are quantum analogs of skipping orbits. Recent extensive experimental studies\cite{firstMZ, Heiblum, Glattli, Litvin, Basel} of these sates have led to the emergence of a new field in condensed matter physics dubbed the electron optics. On the theoretical side, there are two main points of view on the physics of quantum Hall (QH) edge states. One group of theories \cite{group1} suggests that at integer values of the Landau levels filling factor the edge excitations are free chiral \textit{fermions}. The second group of theories is based on the concept of the edge magneto-plasmon picture.\cite{group2} The fundamental edge excitations in these theories are the charged and neutral collective \textit{boson} modes. The domain where these two approaches meet each other is the \textit{low-energy} effective theory.\cite{eff-theory} In the framework of this theory, both fermion and boson excitations are two forms of the same entity. Namely, they can be equivalently rewritten in terms of each other: $$ \psi(x,t) \sim \exp[i\phi(x,t)] $$ where $\psi(x,t)$ is the fermion field, and $\phi(x,t)$ is the boson field. However, this transformation is highly nonlinear, and in the presence of strong Coulomb interaction fermions are not stable and decay into the boson modes which are the eigenstates of the edge Hamiltonian. Results of tunneling spectroscopy experiments \cite{Chang} reasonably agree with the free-electron description of edge states. However, the first experiment on Aharonov-Bohm (AB) oscillations of a current through the electronic Mach-Zehnder (MZ) interferometer \cite{firstMZ} has shown that the phase coherence of edge states is strongly suppressed at energies, which are inversely proportional to the interferometer's size. Moreover, several subsequent experiments on MZ interferometers at filling factor $\nu = 2$ have shown puzzling results on finite bias dephasing\cite{Heiblum, Basel, Glattli, Litvin} theoretically studied in \cite{Neder, Sim, Chalker,our,Sukh-Che}. Namely, the visibility of AB oscillations in these experiments is found to have a lobe-type pattern as a function of the applied voltage bias. Such results are difficult to explain in terms of the fermion picture, while they all follow naturally from the plasmon physics,\cite{our} where the Coulomb interaction plays a crucial role. Thus the boson picture of edge excitations might be more appropriate. In contrast to mentioned above non-local experiments, some local measurements seem to be not able to differentiate between two physical pictures of edge states. For example, both theories predict Ohmic behavior of the tunneling current, unless it is renormalized by a non-linear dispersion of plasmons. Moreover, the equilibrium distribution of the bosons is equivalent to the one of fermions (see the demonstration of this fact in Section \ref{sec-distr-long}). Therefore, it might be interesting to investigate non-equilibrium local properties of edge states. Non-equilibrium behavior of 1D systems has been a subject of intensive theoretical \cite{1d} and experimental \cite{1d-exp} studies for a long time. However, only recently it has become possible to measure an electron distribution at quantum Hall edge $f(\epsilon)$ as a function of energy $\epsilon$ with high precision. \cite{exper} The main idea of the experimental technique is to restore the function $f(\epsilon)$ by measuring the differential conductance $\mathcal{G}$ of tunneling between two edges through a single level in a quantum dot: \begin{equation} \mathcal{G}(\epsilon)\propto \partial f(\epsilon)/\partial\epsilon, \end{equation} where $\epsilon$ is the energy of the quantum dot level, controlled by the gate voltage $V_g$. This technique has been used in experiments [\onlinecite{exper2}] in order to investigate the energy relaxation at QH edge states at filling factor $\nu = 2$. The schematics of these experiments is shown in Fig.\ \ref{nu2}. The main result is that the electron distribution relaxes toward local equilibrium Fermi distribution, and the energy splits equally between the two edge channels. \begin{figure}[h] \epsfxsize=7cm \epsfbox{equilib.eps} \caption{Schematics of the experiments [\onlinecite{exper}] and [\onlinecite{exper2}]. The shaded region is filled by the 2D electron gas in the regime of the quantum Hall effect. At filling factor $\nu = 2$ there are two chiral edge states shown by the blue (the outer channel) and the black (the inner channel) lines. The QPC of the transparency $T$ and biased with the voltage difference $\Delta\mu$, injects electrons into the outer channel, and thus creates a non-equilibrium electron distribution. After the propagation along the QH edge, the distribution is detected at distances $L$ from the source with the help of a quantum dot with a single level controlled by the gate voltage $V_g$. } \vspace{-3mm} \label{nu2} \end{figure} The first theoretical models, based on the fermion picture [\onlinecite{nigg}] and on the plasmon approach [\onlinecite{giovan}], have come qualitatively to \textit{identical} conclusions. Namely, both works predict equal distribution of the energy between the edge channels, in agreement with the experimental findings. In other words, based on the results of Refs.\ [\onlinecite{nigg}] and [\onlinecite{giovan}] alone, the experimentalists are not able\cite{exper2} to discriminate between two alternative descriptions of the physics of QH edge. Thus, it seems to be important and timely to reanalyze the problem of the energy relaxation at the QH edge in order to make new, model specific and distinct predictions that can be verified experimentally. This is exactly the purpose of the present work. Here we show that the Coulomb interaction strongly affects the spectrum of collective edge excitations and leads to the formation of charged and dipole plasmons modes, which propagate with different velocities.\cite{citeus} They carry away the energy of electrons injected through the QPC and equally distribute it between edge channels at distances $L_{\rm ex}$ from the QPC. In addition to this observation, which agrees with findings of previous works, \cite{nigg,giovan} we stress that the same process splits the wave packets of injected electrons, and leads to strong coupling of electrons to the noise of the QPC current. The regime of weak injection, i.e., when the transparency of the QPC is small, $T\ll 1$, deserves a special consideration. In this regime the current noise at relevant time scales becomes markovian, and as a result, the function $-\partial f(\epsilon)/\partial\epsilon$ acquires a Lorentzian shape. (This effect resembles a well known phenomenon of the homogeneous level broadening.) Interestingly, the width of the Lorentzian scales as $T\Delta\mu$ at small $T$, where $\Delta\mu$ is the voltage bias applied to a QPC. In contrast, the width of the eventual equilibrium Fermi distribution of thermalized electrons scales as $\sqrt{T}\Delta\mu$. If thermalization takes place at longer distances, $L_{\rm eq}\gg L_{\rm ex}$, then the intermediate regime described here may be observed in experiment with a weak injection. This would indicate that interactions strongly affects the physics at the edge and that the fermion picture becomes {\em inappropriate}. In order to theoretically describe the experiments [\onlinecite{exper2}] and to quantitatively elaborate the physical picture, we use the \textit{non-equilibrium} bosonization technique, which has been introduced in our previous work [\onlinecite{our-phas}]. The main idea of this approach is based on the fact that in a 1D chiral system one can find a non-equilibrium density matrix by solving equations of motion for plasmons with non-trivial boundary conditions. Then, one can rewrite an average over the non-equilibrium state of an interacting system in terms of the full counting statistics (FCS) generators\cite{Levitov} of the current at the boundary. In the situation considered in this paper, because of chirality of QH edge states, interactions do not affect the transport through the QPC alone. This leads to a great simplification, because in the markovian limit the FCS generator for free electrons is known.\cite{Levitov} The structure of the paper is following: In Sec.\ \ref{sec-noneq} we describe the non-equilibrium bosonization technique in some details. Next, we use this technique in Sec.\ \ref{sec-corr} in order to find the electron correlation function for different distances from the QPC. Finally, we use these results to find the electron distribution function in Sec.\ \ref{sec-distr}, and present our conclusions in Sec.\ \ref{sec-conclus}. Several important technical steps and the phenomena resulting from the non-linearity of the spectrum of plasmons are described in Sec.\ \ref{ap-energy} and Appendices. \section{Non-equilibrium bosonization} \label{sec-noneq} We note, that the relevant energy scales in recent mesoscopic experiments with the QH edge state \cite{exper,Heiblum,Basel,Glattli,Litvin} are very small. Therefore, it is appropriate to use the low-energy effective theory \cite{eff-theory} of the QH edge. One of the advantages of this theory is that it allows to take into account strong Coulomb interactions in a straightforward way.\cite{our} However, an additional complication arises from the fact that in experiments \cite{exper,exper2} the injection into one of the two edge channels creates a strongly non-equilibrium state. We, therefore, start by recalling in this section the method of non-equilibrium bosonization, proposed earlier in Ref.\ [\onlinecite{our-phas}], which is suitable for solving the type of a problem that we face. Throughout the paper, we set $e=\hbar=1$. \subsection{Fields and Hamiltonian} According to the effective theory of QH edge,\cite{eff-theory} the collective fluctuations of the charge densities $\rho_{\alpha}(x)$ of the two edge channels, $\alpha= 1,2$, at filling factor $\nu=2$ are the only relevant degrees of freedom at low energies. These charge densities may be expressed in terms of the \textit{chiral} boson fields, $\phi_{\alpha}(x)$, \begin{equation} \rho_{\alpha}(x)=(1/2\pi)\partial_x\phi_{\alpha}(x), \label{rho} \end{equation} which satisfy the following commutation relations: \begin{equation} [\phi_{\alpha}(x),\phi_{\beta}(y)]=i\pi\delta_{\alpha\beta}{\rm sgn}(x-y), \label{fields} \end{equation} The vertex operator \begin{equation} \psi_\alpha(x) = \frac{1}{\sqrt{a}}\,e^{i\phi_\alpha(x)} \label{psi} \end{equation} annihilates an electron at point $x$ in the edge channel $\alpha$. The constant $a$ in the prefactor is the ultraviolet cutoff, which is not universal and will be omitted and replaced by other normalizations. One can easily check, with the help of the commutation relations (\ref{fields}), that the operators (\ref{psi}) indeed create a local charge of the value 1 at point $x$, and satisfy fermionic commutation relations. Close to the Fermi level the spectrum of electrons may be linearized, therefore the free-fermion part $\mathcal{H}_0$ of the total QH edge Hamiltonian $\mathcal{H} = \mathcal{H}_0 + \mathcal{H}_{\rm int}$ takes the following form: \begin{equation} \mathcal{H}_0 =-iv_F\sum\limits_{\alpha} \int dx \,\psi_\alpha^\dagger\partial_x\psi_\alpha, \end{equation} where the bare Fermi velocity $v_F$ is assumed to be the same for electrons at both edge channels. The second contribution to the edge Hamiltonian describes the density-density Coulomb interaction, \begin{equation} \mathcal{H}_{\rm int} = (1/2)\sum_{\alpha,\beta} \iint dx dy\, U_{\alpha\beta}(x-y) \rho_\alpha(x)\rho_\beta(y), \label{Hamiltonian} \end{equation} which is assumed to be screened at distances $d$ smaller than the characteristic length scale $L$ in experiments [\onlinecite{exper,exper2,Heiblum,Basel,Glattli,Litvin}], i.e., $L\gg d$. Therefore, we may write: \begin{equation} U_{\alpha\beta}(x-y) = U_{\alpha\beta}\delta (x-y). \label{short-r} \end{equation} Screening may occur due to the presence of either a back gate, or several top gates. We show below that the assumption (\ref{short-r}) results in the linear spectrum of charge excitations. This approximation seems to be reasonable, agrees well with some experimental observations such as an Ohmic behavior of the QPC conductance at low voltage bias, and eventually does not strongly affect our main results. Nevertheless, below we relax this assumption and investigate the effects of weak and strong non-linearities in the spectrum of charge excitations. After taking into account the relations (\ref{rho}) and (\ref{psi}) and applying the point splitting procedure, we arrive at the edge Hamiltonian of the quadratic form in boson fields: \begin{equation} \hspace*{-1pt}\mathcal{H} = \frac{1}{8\pi^2}\!\sum_{\alpha,\beta} V_{\alpha\beta}\!\!\int dx \partial_x\phi_\alpha(x)\partial_x\phi_\beta(x), \label{hamilt} \end{equation} which nevertheless contains free fermion contribution as well as the Coulomb interaction potential: \begin{equation} V_{\alpha\beta} = 2\pi v_F \delta_{\alpha\beta} + U_{\alpha\beta}. \label{kernel} \end{equation} Equations (\ref{fields}), (\ref{psi}), (\ref{hamilt}) and (\ref{kernel}) complete the description of the QH edge at low energies. The experimentally found \cite{exper,exper2} electron distribution function at the outmost QH edge channel is given by the expression: \begin{equation} f(\epsilon) = \int dt e^{-i\epsilon t}\langle\psi_1^\dag(L,t)\psi_1(L,0)\rangle. \label{distr-ferm} \end{equation} Rewriting this expression via the boson fields we finally obtain \begin{subequations} \begin{eqnarray} f(\epsilon)& =&\int dt e^{-i\epsilon t}K(t), \label{distr-phi-1}\\ K(t)& = &\langle e^{-i\phi_1(L,t)} e^{i\phi_1(L,0)}\rangle . \label{distr-phi-2} \end{eqnarray}\label{distr-phi} \end{subequations} where we have introduced the electron correlation function $K$, evaluated at coincident points at distance $L$ from the QPC. The proportionality coefficient in (\ref{distr-phi-2}) may be corrected later from the condition that $f(\epsilon)$ takes a value $1$ for energies well below the Fermi level (see, however, the discussion in Sec.\ \ref{ap-energy} for further details). In equilibrium, in order to evaluate the correlation function on the right hand side of this equation one may now follow a standard procedure \cite{Giamarchi} of imposing periodic boundary conditions on the boson fields and diagonalizing the Hamiltonian (\ref{hamilt}). In our case, however, the average in (\ref{distr-phi}) has to be taken over a \textit{non-equilibrium} state created by a QPC. Attempting to express such a state entirely in terms of bosonic degrees of freedom is a complicated and not a best way to proceed. We circumvent this difficulty by applying a non-equilibrium bosonization technique proposed in our earlier work [\onlinecite{our-phas}]. This technique is outlined below in some detail. \subsection{Equations of motion, boundary conditions,\\ and FCS} \label{NB} The Hamiltonian (\ref{hamilt}), together with the commutation relations (\ref{fields}), generates equations of motion for the fields $\phi_{\alpha}$, which have to be complemented with boundary conditions:\cite{Ines} \begin{subequations} \label{eom} \begin{eqnarray} \partial_t\phi_{\alpha}(x,t) = -\frac{1}{2\pi}\sum_\beta V_{\alpha\beta}\partial_x\phi_{\beta}(x,t), \label{eoma}\\ \partial_t\phi_{\alpha}(0,t) = -2\pi j_{\alpha}(t). \label{eomb} \end{eqnarray} \end{subequations} The last equation follows from the charge continuity condition $\partial_t \rho_{\alpha}+\partial_x j_{\alpha}=0$ and the definition (\ref{rho}). Thus the operator $j_{\alpha}(t)$ describes a current through the boundary $x=0$ in the channel $\alpha$. For the convenience, we place a QPC in the outer channel $\alpha=1$ right before the boundary, so that the operator $j_{1}(t)$ describes an outgoing QPC's current. The key idea of the non-equilibrium bosonization approach is to replace the average in Eq.\ (\ref{distr-phi}) by the average over temporal fluctuations of currents $j_{\alpha}$, the statistics of which is assumed to be known. Indeed, although in general the fields $\phi_{\alpha}$ influence fluctuations of the currents $j_{\alpha}$, leading to such effects as the dynamical Coulomb blockade \cite{dynamicalCB} and cascade corrections to noise, \cite{nagaev} in the case of chiral fields describing QH edge states no back-action effects arise. \cite{Sukh-Che,our} As a consequence, at integer filling factors the electron transport through a single QPC is not affected by interactions, which seem to be an experimental fact. \cite{exper,Basel} Therefore, by solving equations (\ref{eom}) one may express the correlation functions of the fields $\phi_{\alpha}$ in terms of the generator of full counting statistics (FCS):\cite{Levitov} \begin{equation} \chi_{\alpha}(\lambda,t)=\langle e^{i\lambda Q_{\alpha}(t)}e^{-i\lambda Q_{\alpha}(0)}\rangle. \label{fcs} \end{equation} Here averaging is taken over \textit{free} electrons, and the operators \begin{equation} Q_{\alpha}(t)=\int_{-\infty}^t dt' j_{\alpha}(t') \label{Q} \end{equation} may be viewed as a total charge in the channel $\alpha$ to the right of the boundary at $x=0$. To prove the connection of the electron correlations in (\ref{distr-phi}) to the generating functions (\ref{fcs}), we come back to the discussion of the interaction effects, which are in fact encoded in a solution of the equations of motion (\ref{eoma}). The long-range character of the Coulomb interaction leads to the logarithmic dispersion in the spectrum of collective charge excitations, the physical consequences of which are discussed in Secs.\ \ref{sec-distr} and \ref{ap-energy}. For a moment, to simplify equations (\ref{eoma}), we have assumed screening of the Coulomb potential at distances $d$ shorter than the characteristic length scale $v_F/\Delta\mu$, which is of the order of few microns in recent experiments. Nevertheless, it is very natural to assume that the screening length $d$ is much larger than the distance $a$ between edge channels, $d \gg a$, which does not exceed few hundreds nanometers. Therefore, one can write \begin{equation} U_{\alpha\beta} =\pi u,\quad u/v_F\sim\log(d/a)\gg 1, \label{approx} \end{equation} i.e., the in-channel interaction strength is approximately equal to the intra-channel. As a result, the spectrum of collective charge excitations splits into two modes: a fast charged mode $\tilde\phi_1$ with the speed $u$, and a slow dipole mode $\tilde\phi_2$ with the speed $v \simeq v_F$. It is important to stress that the condition $d\gg a$, leading to (\ref{approx}), results in a sort of universality: the solution of equations of motion (\ref{eoma}) in terms of the charged and dipole mode, \begin{subequations} \label{transform-phi} \begin{eqnarray} \phi_1(x,t) = \frac{1}{\sqrt{2}}[\tilde{\phi}_1(x-ut)+\tilde{\phi}_2(x-vt)]\\ \phi_2(x,t) = \frac{1}{\sqrt{2}}[\tilde{\phi}_1(x-ut)-\tilde{\phi}_2(x-vt)] \end{eqnarray} \end{subequations} is only weakly sensitive to perturbations of our model, in particular to those that account for different bare Fermi velocities of edge channels and slightly different interaction strengths. Applying now boundary conditions (\ref{eomb}) to the result (\ref{transform-phi}), we finally solve equations of motion in terms of the boundary currents: \begin{subequations} \label{sol-full} \begin{multline} \phi_1(x,t) = -\pi\!\int_{-\infty}^{t_u}\!\!\!dt'[j_{1}(t')+j_{2}(t')]\\ -\pi\!\int_{-\infty}^{t_v}\!\!\!dt'[j_{1}(t')-j_{2}(t')], \end{multline}\vspace{-25pt} \begin{multline} \phi_2(x,t) = -\pi\!\int_{-\infty}^{t_u}\!\!\!dt'[j_{1}(t')+j_{2}(t')]\\ +\pi\!\int_{-\infty}^{t_v}\!\!\!dt'[j_{1}(t')-j_{2}(t')], \end{multline} \end{subequations} where we have introduced notations \begin{equation} t_u = t- x/u,\quad t_v = t- x/v. \label{notations} \end{equation} Finally, using the definition (\ref{Q}), we arrive at the solution in the compact form \begin{equation} \phi_{1}(x,t)=-\pi[Q_{1}(t_u)+Q_{2}(t_u)+Q_{1}(t_v)-Q_{2}(t_v)], \label{sol-fast} \end{equation} and to a similar expression for the inner channel. The physical meaning of this result is rather simple: when charges are injected into the channel $\alpha=1$ and $2$, they excite charged and dipole mode (note the minus sign in the fourth term on the right hand side) which have different propagation speeds $u$ and $v$. As a result, these charges arrive at the observation point $x$ with different time delays $x/u$ and $x/v$, and make a contribution to the field $\phi_1$ at different times (\ref{notations}). \begin{figure}[h] \epsfxsize=8cm \epsfbox{split.eps} \caption{Schematic illustration of the Coulomb interaction effect at the HQ edge at filling factor $\nu=2$. The electron wave-packet of the charge $e$ created in the outer edge channel (black line) decays into two eigenmodes of the Hamiltonian (\ref{hamilt}), the charged and dipole mode, which propagate with different speeds and carry the charge $e/2$ in the outer channel. As a result, the wave packets do not overlap at distances larger than their width, and contribute independently to the electron correlation function with the coupling constant $\lambda = \pi$. \cite{footnote1} Similar situation arises when an electron is injected in the inner channel (blue line), however in this case the charged and dipole states carry opposite charges at the outer channel. Thus, there are four independent contributions to the correlation function in the outer edge channel. } \vspace{-3mm} \label{fig-plasmon} \end{figure} When substituting this result into the correlation function in Eq.\ (\ref{distr-phi-2}) one may use the statistical independence of the current fluctuations at different channels and split the exponential functions accordingly: \begin{multline} \label{corr-long} K(t)=\langle e^{i\pi [Q_1(t_u)+Q_1(t_v)]}e^{-i\pi [Q_1(t_u-t)+Q_1(t_v-t)]}\rangle \\ \times \langle e^{i\pi [Q_2(t_u)-Q_2(t_v)]}e^{-i\pi [Q_2(t_u-t)-Q_2(t_v-t)]}\rangle . \end{multline} In the rest of the paper we will be interested in the correlation function at relatively long distances $L\gg v\tau_c$, where $\tau_c \simeq 1/\Delta\mu$ is the correlation time of fluctuations of the current through a QPC. (We show below that at this length scale the energy exchange between two channels takes place.) In this case, the partitioned charges $Q_\alpha$, taken at different times $t_u$ and $t_v$, are approximately not correlated, as illustrated in Fig.\ \ref{fig-plasmon}. This assumption is quite intuitive and may be easily checked using Gaussian approximation. We, finally, arrive at the following important result: \begin{equation} \label{corr-short} K(t) =\chi_1^2(\pi, t)\chi_2(-\pi,t)\chi_2(\pi,t), \end{equation} i.e., the electronic correlation function (\ref{corr-long}) may indeed be expressed in terms of the FCS generator (\ref{fcs}). \section{Electron correlation function} \label{sec-corr} The expression (\ref{corr-short}) presents formally a full solution of the problem of evaluation of an electron correlation function. Generators of the FCS for free electrons in this expression, defined as (\ref{fcs}), may be represented as a determinant of a single particle operator,\cite{Levitov} and eventually, evaluated, e.g., numerically. However, a further analytical progress is possible in a number of situations, which are important for understanding physics of the energy relaxation processes. In particular, we show in this section that for the case of equilibrium fluctuations of the boundary currents, the correlation function (\ref{corr-short}) as well as the electron distribution function (\ref{distr-phi}) acquire an equilibrium free-fermionic form. The electron correlation function may also be found analytically away from equilibrium for the case of a Gaussian noise. Interestingly, in the short-time limit, $t\ll 1/\Delta\mu$, the main contribution to the correlation function comes from zero-point fluctuations of boundary currents, and it behaves as a free-fermion correlator, i.e., it scales as $1/t$. In the long-time limit, $t\gg 1/\Delta\mu$, the non-equilibrium zero-frequency noise dominates, and the electron correlation function decays exponentially with time. This is exactly the limit where a non-Gaussian markovian noise should also be taken into account. \subsection{Gaussian noise} In the context of the noise detection physics \cite{Levitov,Edwards} the dimensionless counting variable $\lambda$ in the expression (\ref{corr-short}) for the FCS generator plays the role of a coupling constant. Typically, it is small, $\lambda\ll 1$, so that the contributions of high-order cumulants of current noise to the detector signal are negligible. \cite{Edwards} In contrast, in the physical situation that we consider in the present paper $\lambda=\pm\pi$, implying that the shape of the distribution function may be strongly affected by high-order current cumulants. Nevertheless, it is instructive to first consider Gaussian fluctuations, simply truncating the cumulant expansion at second order in $\lambda$. In this case the correlation function (\ref{corr-long}) may be evaluated exactly. The are many reasons for starting the analysis from considering an example of a Gaussian noise: First of all, in equilibrium the current fluctuations in a chiral 1D system are always Gaussian. Second, as we show in the Appendix \ref{ap-supr}, the dispersion of the charged and dipole modes leads to a suppression of higher order cumulants at large distances $L$. Finally, on the Gaussian level it is more easy to investigate and compare contributions of zero-point fluctuations and of non-equilibrium noise to the electron correlation function. Thus, expanding the logarithm of the right hand side of the Eq.\ (\ref{fcs}) to second order in $\lambda$ and accounting for the Eq.\ (\ref{Q}), we obtain \begin{equation} \log[\chi_{\alpha}(\lambda,t)]=i\lambda \langle j_{\alpha}\rangle t -\lambda^2J_{\alpha}(t). \label{logchi} \end{equation} Here the Gaussian contribution of current fluctuations $\delta j_\alpha(t) \equiv j_\alpha(t) - \langle j_\alpha\rangle$ is given by the following integral \begin{equation} J_{\alpha}(t) = \frac{1}{2\pi}\int\frac{d\omega S_{\alpha}(\omega)}{\omega^2+\eta^2}(1-e^{-i\omega t}), \quad \eta\to 0, \label{gaussian} \end{equation} where the non-symmetrized noise power spectrum is defined as \begin{equation} \label{S} S_\alpha(\omega) = \int dt e^{i\omega t}\langle\delta j_\alpha(t)\delta j_\alpha(0)\rangle. \end{equation} In what follows, we apply this result for the evaluation of the electron correlation function in the case of equilibrium boundary conditions and in the case of a Gaussian noise far away from equilibrium. \subsubsection{Equilibrium boundary conditions} \label{b-f-eq} One may propose the following simple test of the non-equilibrium bosonization method: Let us consider an infinite QH edge. In equilibrium, the charge densities and edge currents exhibit thermal fluctuations. This is the case, in particular, for the currents $j_\alpha$ through the cross-section $x=0$, which are considered to be boundary conditions for the field $\phi_\alpha$ in our theory. Therefore, one may evaluate the electron correlation function using these boundary conditions and compare it with the result of the standard equilibrium bosonization technique,\cite{Giamarchi} applied to a chiral 1D system.\cite{our} In equilibrium, $\langle j_{\alpha}\rangle=0$. The current noise power spectrum is given by the fluctuation-dissipation relation\cite{FDT} \begin{equation} S_\alpha(\omega)\equiv\int dt e^{i\omega t}\langle j_\alpha(t)j_\alpha(0)\rangle = \frac{1}{2\pi}\, \frac{\omega}{1-e^{-\beta\omega}}. \label{seq} \end{equation} Substituting this expression into the equation (\ref{gaussian}), one obtains \begin{equation} \label{int-corr} \log[\chi_\alpha(\lambda,t)] = -\frac{\lambda^2}{4\pi^2} \int \frac{d\omega}{\omega}\frac{1-e^{-i\omega t}}{1-e^{-\beta\omega}}\,. \end{equation} This integral may be evaluated by expanding the integrand in Boltzmann factors $e^{\pm\beta\omega}$ and integrating each term. Substituting the result (for $\lambda = \pi$) into Eq.\ (\ref{corr-short}), we arrive at the following expression for the electron correlation function in the case of equilibrium boundary conditions: \begin{equation} \label{corr-thermal} K(t)\propto \frac{\beta^{-1}}{\sinh(\pi t /\beta)}, \end{equation} which is, in fact, the equilibrium fermionic correlation function. The straightforward calculations of the integral (\ref{distr-phi-1}) gives, naturally, the equilibrium distribution function, $f_1(\epsilon)=1/(1+e^{\beta\epsilon})\equiv f_F(\epsilon)$, where we have fixed the normalization constant, as explained above. Thus for chiral, interacting quasi-1D systems with a linear spectrum equilibrium bosons also implies equilibrium distribution of fermions. It is instructive to compare this result with the known expression for the electron correlation function at $\nu=2$, found earlier in Ref.\ [\onlinecite{our}] with the help of the standard bosonization technique: \begin{equation} \label{corr-known} K(t)=\beta^{-1}\Big[\sinh\Big(\frac{x-y-vt}{v\beta/\pi}\Big)\sinh\Big(\frac{x-y-ut}{u\beta/\pi}\Big)\Big]^{-1/2} \end{equation} For $x = y$, details of the interaction leading to wave-packet splitting (see Fig.\ \ref{fig-plasmon}) vanish, and one obtains the expression (\ref{corr-thermal}), thus validating our approach. Moreover, the free-fermionic character of the correlation function at coincident points (\ref{corr-thermal}) justifies the assumption underlying the non-equilibrium bosonization procedure that the FCS generators (\ref{fcs}) may be taken as for free electrons. \subsubsection{Gaussian noise away from equilibrium} For a QPC far away from equilibrium, $\beta\Delta\mu\gg 1$, one may simply set the temperature to zero. Straightforward calculations based on the scattering theory \cite{Buttiker} give the following result for the spectral density of noise (\ref{S}) of a QPC: \begin{equation} S_{\alpha}(\omega)= S_{\rm q}(\omega) + R_{\alpha} T_{\alpha} S_{\rm n}(\omega), \label{spectr} \end{equation} where $T_\alpha=1-R_\alpha$ is the transparency of a QPC (i.e., the average occupation in the channel $\alpha$), $S_{\rm q}(\omega) =(1/2\pi)\omega \theta(\omega)$ is the quantum, ground-state spectral function, and $S_{\rm n}(\omega) =\sum_\pm S_{\rm q}(\omega\pm\Delta\mu) -2S_{\rm q}(\omega),$ is the non-equilibrium contribution (see Fig.\ \ref{noise}). Note, that the noise power (\ref{spectr}) differs from the one for a non-chiral case. \cite{Edwards} \begin{figure}[h] \epsfxsize=8cm \epsfbox{noise.eps} \caption{Two contributions to the spectral density of noise (\ref{spectr}). {\em Left panel}: Quantum contribution $S_{\rm q}(\omega)$ generated by the incoming Fermi sea. This contribution vanishes at low frequencies $S_{\rm q}(0)= 0$, but dominates the behavior of the correlator (\ref{gaussian}) at short times, $t\Delta\mu\ll 1$. {\em Right panel}: Non-equilibrium contribution $S_{\rm n}(\omega)$ dominates at long times $t\Delta\mu\gg 1$, i.e., in markovian limit.} \vspace{-3mm} \label{noise} \end{figure} Evaluating the integral (\ref{gaussian}), we arrive at the following expression \begin{equation} J_{\alpha}(t) = (1/4\pi^2)[\log t + 2R_\alpha T_\alpha f(\Delta\mu t)], \end{equation} where the logarithm of time originates from the quantum contribution $S_{\rm q}$, and the dimensionless function $f(\Delta\mu t)$, describing non-equilibrium noise, is given by the integral \begin{equation} \label{f} f(\Delta\mu t) = \int_0^{1}\!\!\!ds\,\frac{1 - s}{s^2}\,[1-\cos(\Delta\mu s)]. \end{equation} This function has a quadratic behavior $f(\Delta\mu t) \simeq (\Delta\mu t)^2/4$ at short times $\Delta\mu t\ll 1$, while in the long-time (markovian) limit, $\Delta\mu t\gg 1$, the dominant contribution to this function is linear in time: $f(\Delta\mu t)\simeq (\pi/2)|\Delta\mu t|$. For the purpose of further analysis we need the electron correlation function in the long-time limit. Taking into account that $\langle j_\alpha\rangle = \Delta\mu T_\alpha /2\pi$, we find the cumulant generating function \begin{eqnarray} \log [\chi_\alpha ] &=& \frac{i\lambda}{2\pi} \Delta\mu T_\alpha t \nonumber \\ &-&\Big(\frac{\lambda}{2\pi}\Big)^2(\log t - \pi R_\alpha T_\alpha |\Delta\mu t|),\;\, \Delta\mu t\gg 1.\quad \end{eqnarray} Finally, substituting this result into the equation (\ref{corr-short}), and setting $T_1=T$ and $T_2=1$ according to the situation shown in Fig.\ \ref{nu2}, we obtain the electron correlation function in the long-time limit: \begin{equation} K(t) \propto t^{-1}e^{i\Delta\mu T t- \pi RT\Delta\mu|t|/2}, \quad\Delta\mu t\gg 1. \label{cf-gauss} \end{equation} Note, that the expression (\ref{cf-gauss}) contains the quantum contribution in the form of a single pole, as for free fermions, as well as the non-equilibrium contribution in the form of an exponential envelop, those width is determined by the noise power at zero frequency, $S_1(0) = RT\Delta\mu/2\pi$. The phase shift of the correlator is determined by the ``average'' voltage bias $\langle\Delta\mu\rangle=\Delta\mu T$ of the incoming stream of electrons, diluted by the QPC. In the next section we show that this mean-field like effect of the dilution is strongly modified by a non-Gaussian component of noise. \subsection{Non-Gaussian markovian noise} Here we consider non-Gaussian noise and show that the contribution of high-order cumulants of current to the correlation function is not small. Note, that the quantum ground state part of the current noise, $S_{\rm q}$, that dominates at short times, is pure Gaussian. Therefore, the denominator in the expression (\ref{cf-gauss}) remains unchanged. In the long time, markovian limit, the dominant contribution to the FCS generator comes from the non-equilibrium part of noise, which, e.g., is described by $S_{\rm n}$ in Gaussian case. For a QPC, the markovian FCS generator is given by the well known expression \cite{Levitov} for a Binomial process: \begin{equation} \chi_{1}(\lambda,t)=(R+Te^{i\lambda})^N, \end{equation} where $N=\Delta\mu t/2\pi$ is the total number of electrons that contribute to noise. Applying the analytical continuation $\lambda\to \pi$, we obtain \begin{equation} \log[\chi_{1}(\pi,t)]=\frac{\Delta\mu t}{2\pi}\big[\log|T-R|+i\pi\theta(T-R)\big]. \label{logchi-qpc} \end{equation} Substituting this expression to the correlation function (\ref{corr-short}), we arrive at the result \begin{equation} \label{cf-nongauss} K(t)\propto t^{-1}e^{i\theta(T-R)\Delta\mu t +\log|T-R|\Delta\mu|t|/\pi}, \end{equation} where the imaginary part of the exponent determines the effective voltage bias, while the real part is responsible for dephasing. Interestingly, at the point $T = 1/2$ the dephasing rate is divergent, and the effective voltage bias drops to zero abruptly for $T<1/2$. It has been predicted in Ref.\ [\onlinecite{our-phas}] that this behavior may lead to a phase transition in the visibility of Aharonov-Bohm oscillations in electronic Mach-Zehnder interferometers. We will argue below that no sharp transition arises in the electron distribution function. However, it leads to its strong asymmetry with respect to the average voltage bias $\langle\Delta\mu\rangle=T\Delta\mu$ of the outer channel. \section{Electron distribution function} \label{sec-distr} In this section we use the results (\ref{corr-thermal}), (\ref{cf-gauss}) and (\ref{cf-nongauss}) for the correlation function of electrons to evaluate and analyze the electronic distribution function. We start by noting that the experiments [\onlinecite{exper,exper2}] are done in a particular regime of strong partitioning $T\approx 0.5$ at the QPC injecting current to the channel $\alpha=1$. This detail, which seem to be irrelevant from the first glance, is in fact of crucial importance. Indeed, as it follows from the expressions (\ref{cf-gauss}) and (\ref{cf-nongauss}), the main contribution to the integral (\ref{distr-phi-1}) for the correlation function comes from times $t$ of the order of the correlation time $\tau_c\simeq 1/\Delta\mu$, where our results based on the markovian noise approximation are, strictly speaking, not valid. However, the numerical calculations show that the non-equilibrium distribution in this regime is very close to the equilibrium one. Therefore, the actual equilibration of electrons, which is reported in the experiment [\onlinecite{exper}] to occur at distances $v/\Delta\mu$, may in fact take place at even longer distances $L_{\rm eq}\gg v/\Delta\mu$ due to an unknown mechanism (not considered here). Indeed, if the chiral LL model considered in our paper is valid, then neither the strong interaction between electrons of two edges taken alone, nor the weak dispersion of plasmons resulting from a long-range character of Coulomb interaction may lead to the equilibration, because the systems remain integrable. Thus it seems to be reasonable to assume that the equilibration length $L_{\rm eq}$ may indeed be quite long. Therefore, in order to explore the physics of collective charge excitations at intermediate distances we propose to consider a regime of weak injection at the QPC: $T\ll 1$. Firstly, we note that in this case our results (\ref{cf-gauss}) and (\ref{cf-nongauss}) may indeed be used to evaluate the electron distribution function, because the main contribution to the integral (\ref{distr-phi-1}) arises from markovian time scales. Secondly, and more importantly, in this regime the electron distribution function acquires a strongly non-equilibrium form and the width of the order of $T\Delta\mu$, which plays a role of the new energy scale. Moreover, the advantage of the weak injection regime is that it allows to investigate a non-trivial evolution of the distribution function, which arises due to bosonic, collective character of excitations and goes via several well distinguishable steps. \subsection{Short distances} At distances of the order of the energy exchange length scale \begin{equation} \label{len-nongauss} L_{\rm ex}\equiv v/\Delta\mu \end{equation} the initial double step distribution function is strongly perturbed by the interaction between channels. As we argued in Sec.\ \ref{NB}, at distances $L\gg L_{\rm ex}$ the charged and dipole modes split and make independent contributions to the electron correlation function. Therefore, we may rely on the result (\ref{cf-nongauss}). Applying the limit $T\ll 1$ to this expression and evaluating the Fourier transform, we find: \begin{equation} \label{distr-nongauss} -\frac{\partial f(\epsilon)}{\partial\epsilon} = \frac{\Gamma_{\rm ng}/\pi}{\epsilon^2 + \Gamma_{\rm ng}^2}, \quad \Gamma_{\rm ng}=2T\Delta\mu/\pi. \end{equation} Here, the missing prefactor in the correlation function has been fixed by the requirement that $f(\epsilon)=1$ at $\epsilon\to-\infty$. Thus, we conclude that energy derivative of the distribution function acquires a narrow Lorentzian peak, which is shifted with respect to the average bias $\langle\Delta\mu\rangle=T\Delta\mu$ and centered at $\epsilon=0$. The last effect is a unique signature of the non-Gaussian character of noise. Because of the electron-hole symmetry of the Binomial process, in the limit $R\ll 1$ the Lorentzian peak obviously has a width $\Gamma_{\rm ng}=2R\Delta\mu/\pi$ and centered at $\epsilon=\Delta\mu$. We stress again, that the result (\ref{distr-nongauss}) holds only for small enough energies close to the Fermi level, namely, for $|\epsilon| < \Delta\mu$, where the main contribution arises from the noise in markovian limit. In fact, the result (\ref{distr-nongauss}) fails at large energies in somewhat non-trivial way. Namely, it is easy to see that any electron distribution function has to satisfy the sum rule \begin{equation} \langle\Delta\mu\rangle\equiv\epsilon_0+\int\limits_{\epsilon_0}^\infty d\epsilon\,f(\epsilon) =-\int\limits_{-\infty}^\infty d\epsilon\, \epsilon\, \partial f(\epsilon)/\partial\epsilon, \label{sum-rule1} \end{equation} where $\epsilon_0$ is the cutoff well below the Fermi level, and the ``average'' bias $\langle\Delta\mu\rangle=T\Delta\mu$ in the case of linear dispersion of plasmons. This sum rule simply expresses the requirement of the conservation of the charge current and implies certain amount of asymmetry in the distribution function. In the present case, such an asymmetry arises in the power-law tails of the function $-\partial f(\epsilon)/\partial\epsilon$ and originates from quantum non-equilibrium noise. It can be seen in Fig.\ \ref{fig-distrib}, where the results of numerical calculations are shown. Moreover, at energies of the order of $\Delta\mu$ the power-law behavior of the function (\ref{distr-nongauss}) has to have a cut-off, because the QPC does not provide energy much larger than the voltage bias. Quantitatively, this follows from the conservation of the energy. We demonstrate below that for the system with linear dispersion of plasmons, the heat flux in the outer channel can be written entirely in terms of the single-electron distribution function (in unites $e=\hbar=1$), \begin{equation} I_{\rm m} = (1/2\pi)\int d\epsilon\,\epsilon\,[f(\epsilon)-\theta(\langle\Delta\mu\rangle-\epsilon)], \label{flux-def} \end{equation} as in the case of free electrons. We use the subscript ``${\rm m}$'' in order to emphasize the fact that it is this quantity that has been measured in the experiment [\onlinecite{exper2}]. In Sec.\ \ref{ap-energy} we show that at distances $L\gg L_{\rm ex}$ the total heat flux injected at a QPC splits equally between two edge channels, therefore integrating Eq.\ (\ref{flux-def}) by parts and substituting the heat flux for a double-step distribution, we obtain \begin{equation} I_{\rm m} = -\frac{(T\Delta\mu)^2}{4\pi}-\frac{1}{4\pi}\int d\epsilon\, \epsilon^2 \frac{\partial f(\epsilon)}{\partial\epsilon}=\frac{TR(\Delta\mu)^2}{8\pi} \label{sum-rule2} \end{equation} for $L\gg L_{\rm ex}$. One can see from Eq.\ (\ref{distr-nongauss}) that indeed the power-law behavior has to have a cut-off at $|\epsilon|\sim\Delta\mu$. We stress, however, that this summation rule is less universal than the one given by Eq.\ (\ref{sum-rule1}), because it accounts only a single-particle energy of electrons and fails in the case of a non-linear spectrum of plasmons, considered in Sec. \ref{ap-energy} in detail. \subsection{Intermediate distances} \label{sec-distr-interm} So far we have considered the case of a linear spectrum of plasmons. This is a reasonable assumption, taking into account the fact that non-linear corrections in the spectrum of plasmons lead to a nonlinear corrections in the Ohmic conductance of a QPC. However, the experiments [\onlinecite{exper,exper2}] seem to be done in the Ohmic regime. Nevertheless, even in the case of a weak non-linearity in the spectrum of the the both modes of the sort\cite{footnote2} \begin{equation} k_j(\omega) = \omega/v_j + \gamma_j \omega^2{\rm sign}(\omega),\quad v_1=u,\; v_2=v, \label{spectrum} \end{equation} barely seen in the conductance of a QPC, an intermediate length scale $L_{\rm g}$ may arise at which high-order cumulants of current are suppressed, and the noise becomes effectively Gaussian. This situation occurs when the wave packets of the original width $v/(T\Delta\mu)$ overlap. A simple estimate using the nonlinear correction (\ref{spectrum}) gives the length scale \begin{equation} L_{\rm g}=1/\gamma(T\Delta\mu)^2, \quad \gamma\equiv \min(\gamma_j). \label{len-gauss} \end{equation} We support this conclusion by rigorous calculations in Appendix \ref{ap-supr}. The non-linearity in the spectrum is weak, if $\gamma vT\Delta\mu\ll 1$. This implies that $L_{\rm g}\gg L_{\rm ex}$, and leads to the possibility to observe non-Gaussian effects at distances $L_{\rm ex}\ll L\ll L_{\rm g}$, discussed in the previous section. Obviously, the same requirement also guarantees that dispersion corrections to the Ohmic conductance of a QPC are small. This allows us to neglect corrections to the quantum part of the electron correlation function and to use the result (\ref{cf-gauss}) for a Gaussian noise. Substituting this result to the equation (\ref{distr-phi-1}), we obtain \begin{equation} \label{distr-gauss} -\frac{\partial f(\epsilon)}{\partial\epsilon} = \frac{\Gamma_{\rm ng}/\pi}{(\epsilon-\langle\Delta\mu\rangle)^2 + \Gamma_{\rm g}^2}, \quad \Gamma_{\rm g}=\pi T\Delta\mu/2, \end{equation} in the case $L_{\rm g}\ll L\ll L_{\rm eq}$. One can see, that the width of the function (\ref{distr-gauss}) is almost twice as large compared to the one the function (\ref{distr-nongauss}). Moreover, the function (\ref{distr-gauss}) satisfies the sum rule (\ref{sum-rule1}). Therefore, we do not expect any asymmetry in the high-energy tails of this function, in contrast to the situation with the non-Gaussian noise. The comparison of distribution functions in these two regimes is shown in Fig.\ \ref{fig-distrib}. \begin{figure}[h] \epsfxsize=8cm \epsfbox{distribs.eps} \caption{Energy derivative of the electron distribution function, $-\partial f/\partial\epsilon$, is shown for different distances $L$ from the QPC injecting current. The transparency of the QPC is set to $T = 0.05$ and voltage bias is $\Delta\mu = 40\, \mu V$. {\em Black line}: $-\partial f(\epsilon)/\partial\epsilon$ for short distances $L_{\rm ex}\ll L\ll L_{\rm g}$, so that the noise is Gaussian (\ref{distr-nongauss}). {\em Red line}: $-\partial f(\epsilon)/\partial\epsilon$ for intermediate distances $L_{\rm g}\ll L\ll L_{\rm eq}$, where the noise is Gaussian (\ref{distr-gauss}). {\em Blue line}: The derivative of the Fermi distribution function at the temperature given by Eq.\ (\ref{fermi-width}). The dashed line is a guide for eyes at the energy equal to the effective voltage bias $\langle\Delta\mu\rangle = T\Delta\mu = 2\, \mu V$. {\em Inset}: The same distribution functions are shown in the integrated form. They are shifted vertically by $0.2$ intervals for clarity.} \vspace{-3mm} \label{fig-distrib} \end{figure} So far we have considered a situation where both charged and dipole mode are dispersive. If for some reason the dispersion of one of the modes is negligible, then higher order cumulants are suppressed only by a factor of two. The derivative of the electron distribution function in this situation is given by the Lorentzian \begin{equation} \frac{\partial f(\epsilon)}{\partial\epsilon} = \frac{(\Gamma_{\rm ng}+\Gamma_{\rm g})/ 2\pi}{(\epsilon - \langle\Delta\mu\rangle/2)^2 + (\Gamma_{\rm ng}+\Gamma_{\rm g})^2/4} \end{equation} centered at $\langle\Delta\mu\rangle/2 = \Delta\mu T /2$ with the width $(\Gamma_{\rm ng}+\Gamma_{\rm g})/2 = (1/\pi +\pi/4)\Delta\mu T$. This is because one mode brings only Gaussian component of the markovian noise, while the other one brings full non-Gaussian noise. \subsection{Long distances} \label{sec-distr-long} Next, we consider the distribution function at long distances, $L\gg L_{\rm eq}$, after the equilibration takes place. The temperature of the eventual equilibrium distribution may be found from the conservation of energy. In the next section we show that the heat flux produced at QPC splits equally between two edge states. In the situation of linear dispersion the distribution function acquires the form \begin{equation} f(\epsilon)=\frac{1}{1+e^{(\epsilon-\langle\Delta\mu\rangle)/\Gamma_{\rm eq}}}. \label{eq-distr} \end{equation} The possibility of such equilibration process is suggested by the fact that the equilibrium distribution of bosons implies the equilibrium distribution for electrons, as has been shown in Sec.\ \ref{b-f-eq}. Obviously, the distribution (\ref{eq-distr}) satisfies the sum rule (\ref{sum-rule1}), while the energy conservation condition (\ref{sum-rule2}) may now be used in order to find the effective temperature: \begin{equation} \label{fermi-width} \Gamma_{\rm eq} = \sqrt{3T/2\pi^2}\Delta\mu, \end{equation} where we have used $T\ll 1$. We conclude that the width of the equilibrium distribution scales as $\sqrt{T}$, in contrast to the case of a non-equilibrium distribution at shorter distances from the current source, where it scales linear in $T$. Therefore, if $T$ is small, an equilibrium and non-equilibrium distributions may easily be distinguished, as illustrated in Fig.\ \ref{fig-distrib}. In the situation where the dispersion can not be neglected, the equilibrium distribution of fermions is not given by the Fermi function (\ref{eq-distr}). This situation deserves a separate consideration, which is provided in the next section. \section{Measured and total heat fluxes} \label{ap-energy} We have seen that in the case of weakly dispersive plasmons, $\gamma v T\Delta\mu\ll 1$, the non-linearity in the spectrum leads to the suppression of high-order cumulants of current noise at relatively long distances, which strongly affects the distribution function. On the other hand, the direct contribution of the non-linear correction in the spectrum to local physical quantities, such as the QPC conductance and the heat flux, is small and has been so far neglected. Nevertheless, it may manifest itself experimentally in a quite remarkable way. In this section we show that the non-linearity in the the plasmon spectrum contributes differently to the measured heat flux (\ref{flux-def}) and to the actual heat flux expected from the simple evaluation of the Joule heat. As we demonstrate below, this may, under certain circumstances, explain the missing energy paradox in the experiment [\onlinecite{exper2}]. We start by noting that the measured flux (\ref{flux-def}) at the distance $L$ form the QPC may be expressed entirely in terms of the excess noise spectrum ${\mathbb S}_\alpha(\omega)\equiv S_\alpha(\omega)-(1/2\pi)\omega\theta(\omega)$ of edge channels right after the QPC, where $S_\alpha(\omega)$ is defined in (\ref{S}). Namely, in Appendix \ref{ap-energy-ap} we derive the following result: \begin{multline} \label{tot-energy} I_{\rm m}(L) =\frac{1}{4}\int_{-\infty}^\infty d\omega \{{\mathbb S}_1(\omega)[1+\cos(\Delta k L)] \\ + {\mathbb S}_2(\omega)[1-\cos(\Delta k L)]\}, \end{multline} where $\Delta k\equiv k_1(\omega)-k_2(\omega)$, and $k_j(-\omega)=-k_j(\omega)$. Importantly, this result holds for an arbitrary non-linear spectrum $k_j(\omega)$ of the charged and dipole modes, and for a non-Gaussian noise in general, i.e., high-order cumulants simply do not contribute. One can easily find two important limits of Eq.\ (\ref{tot-energy}): for $L = 0$ we immediately obtain an expected result \begin{equation} I_{\rm m}(0) = \frac{1}{2}\int_{-\infty}^{\infty} d\omega \,{\mathbb S}_1(\omega), \end{equation} while at $L\to\infty$ the cosine in (\ref{tot-energy}) acquires fast oscillations, and we get \begin{equation} I_{\rm m}(\infty) = \frac{1}{2}\int_{-\infty}^{\infty} d\omega [{\mathbb S}_1(\omega)+{\mathbb S}_2(\omega)]. \label{I-at-infty} \end{equation} To be more precise, this happens at $L\gg L_{\rm ex}=v/\Delta\mu$. At zero temperature ${\mathbb S}_2$ vanishes, and the single-electron heat flux $I_{\rm m}$, created at the QPC, splits equally between edge channels: $I_{\rm m}(\infty)=I_{\rm m}(0)/2$. Note also, that in the case of linear dispersion ${\mathbb S}_\alpha=T_\alpha R_\alpha S_{\rm n}$, where $S_{\rm n}$ is shown in Fig.\ \ref{noise}. In the next step, we rewrite the same measured flux in terms of the plasmon distributions $n_j(k)=\langle \tilde a^\dagger_j(k)\tilde a_j(k)\rangle$, see Appendix \ref{ap-energy-ap}: \begin{eqnarray} I_{\rm m}(\infty) &=& \frac{1}{4\pi}\sum_j\int\limits \frac{dk}{k} \, \omega_j^2(k)n_j(k) + I_{\rm q}, \label{measured-f}\\ I_{\rm q} &=& \frac{1}{8\pi}\sum_j\int \frac{dk}{k} [\omega_j^2(k)-(v_jk)^2], \label{quantum-f} \end{eqnarray} where $v_j=\partial\omega_j/\partial k$ are the plasmon speeds at $k=0$. The term $I_{\rm q}$ is the contribution to the measured flux from quantum smearing of the zero-temperature electron distribution function $f_0(\epsilon)$ close to the Fermi level, which originates from a non-linear dispersion of plasmons. Here an important remark is in order. The integral (\ref{quantum-f}) may diverge at large $k$ and has to be cut off at the upper limit. This is because there is no guarantee of the free-fermionic behavior of the correlator $K(t)$ at short times and of the zero-temperature electron distribution function $f_0(\epsilon)$ at large energies. Thus, the integral (\ref{flux-def}) has to be also cut off, which is what in fact is done in experiment. In contrast, the spectrum of plasmons is linear at small $k$, and thus the distribution function $f_0(\epsilon)$ has a free-fermionic behavior close to the Fermi level. Our definition of $I_{\rm q}$ corresponds to the normalization of $f_0(\epsilon)$ to have a discontinuity of the value $-1$ at $\epsilon=\langle\Delta\mu\rangle$. The experimentally measured $I_{\rm q}$ may differ from the one defined in (\ref{quantum-f}) by a constant, which is, on the other hand, independent on the voltage bias $\Delta\mu$. Next, we note that the actual total heat flux in the case of a non-linear dispersion of plasmons acquires the completely different form \cite{footnote3} \begin{equation} I_{\rm h}= \frac{1}{2\pi}\sum_j\int dk \frac{\partial\omega_j}{\partial k}\omega_j(k)\,n_j(k), \label{actual-f} \end{equation} and thus in general $I_{\rm m}\neq I_{\rm h}/2$, contrary to what has been assumed in the experiment [\onlinecite{exper2}]. This may explain the missing energy paradox. Indeed, assuming the low $\omega$ spectrum of the general form \begin{equation} k_j = \omega/v_j + \gamma_j \omega^{\ell_j},\quad j=1,2, \end{equation} where $\gamma_j$ are small, and equilibration of plasmons at $L\to \infty$, i.e., $n_j(k)=n_B(\omega_j/\Gamma_{\rm eq})=1/[\exp(\omega_j/\Gamma_{\rm eq})-1]$, we obtain the missing heat flux as \begin{equation} I_{\rm m}-I_{\rm q}-I_{\rm h}/2=\sum_{j=1,2}c_j\gamma_jv_j\Gamma_{\rm eq}^{\ell_j+1}, \label{paradox} \end{equation} where the constants $c_j=(1/4\pi)\int dz z^{\ell_j+1}n_B(z)$ are of the order of $1$. This result implies that experimentally, the missing heat flux may be found investigating its bias dependence and the spectrum of plasmons. \begin{figure}[h]\begin{center} \epsfxsize=5cm \vspace{3mm} \epsfbox{logar.eps} \caption{Typical spectrum of charged plasmon in the case of the Coulomb interaction screened at distances $d\ll 1/k$. This spectrum is concave, i.e., $\partial\omega/\partial k < \omega(k)/k$.} \vspace{-4mm} \label{fig-coulomb}\end{center} \end{figure} Let us consider an example where the dispersion of charged plasmon at small $k$ arises from the screened Coulomb interaction: \cite{group2} \begin{multline} \label{coulomb-exact} \omega/k = 2[\mathcal{K}_0(ka)-\mathcal{K}_0(kd)]\\ =2\ln(d/a)-(1/2)(kd)^2\ln(2/kd), \end{multline} where $a$ is the high-energy cutoff, $d$ is the distance to the gate, such as $ka\ll kd\ll 1$, and $\mathcal{K}_0$ is the MacDonald function. The low-$k$ asymptotics of this spectrum is illustrated in Fig.\ \ref{fig-coulomb}. One can see that the spectrum is concave, so that in this case the measured heat flux (\ref{measured-f}) is larger than the half of the actual heat flux (\ref{actual-f}). In addition, the effect is weak, because $kd\sim 0.1$ in the experiment [\onlinecite{exper2}]. Thus, the dispersion of the Coulomb interaction potential alone is not able to explain the missing flux paradox. Various mechanisms of convex dispersion are still possible and will be investigated elsewhere. \section{Conclusions} \label{sec-conclus} Earlier theoretical works on quantum Hall edge states at integer filling factors may be divided into two groups: fermion based and boson based theories. Recent interference experiments suggest that the boson approach might be more appropriate for the description of the edge physics. However, both groups of theories give the same predictions for the local physical quantities at equilibrium. Moreover, the first theoretical works based on fermion \cite{nigg} and boson \cite{giovan} approaches and addressing the non-equilibrium local measurements, have not been able to make qualitatively distinct predictions. In this paper we show that it is nevertheless \textit{possible} to test and differentiate between two approaches with the local non-equilibrium measurements. We address recent experiments [\onlinecite{exper,exper2}] with quantum Hall edge states at filling factor 2, where an energy relaxation process has been investigated by creating a non-equilibrium state at the edge with the help of a QPC and reading out the electron distribution downstream using a quantum dot. We use the non-equilibrium bosonization approach \cite{our-phas} in order to describe the gradual relaxation of initially non-equilibrium, double-step electron distribution to its equilibrium form. In the framework of this approach the non-equilibrium initial state is encoded in the boundary conditions for the equations of motions that depend on interactions. We show that the electrons excite two plasmons: fast charged and slow dipole mode. Thus the electron correlation function (\ref{corr-short}) is expressed in terms of the four contributions, each having form of the generator of FCS of free electrons with the coupling constant $\lambda = \pi$. Evaluating the Fourier transform of this function, we find the electron distribution function. Before reaching eventual equilibrium form, the distribution function evolves via several steps, where its energy derivative acquires a Lorentzian shape: \begin{equation} \frac{\partial f(\epsilon)}{\partial\epsilon} = \frac{\Gamma/ \pi}{(\epsilon - \epsilon_0)^2 + \Gamma^2}. \quad |\epsilon|\lesssim\Delta\mu, \label{final-r} \end{equation} Here the width $\Gamma$ and centering $\epsilon_0$ take different values in different regimes. Each of the regimes, summarized below and illustrated in Fig.\ \ref{fig-scales}, has its own dominant process: (i) First, after tunneling through the QPC, electrons excite plasmons, which then split in two eigenmodes: one is charged fast mode with the speed $u$, and the other is slow dipole mode with the speed $v$. This process takes place at distances $L_{\rm ex}=v/\Delta\mu$, where $\Delta\mu$ is the voltage bias across the QPC. In this regime Eq.\ (\ref{corr-short}) applies, which eventually leads to the the distribution (\ref{final-r}) with the width $\Gamma=\Gamma_{\rm ng} = 2\Delta\mu T/\pi$, centered at $\epsilon_0 =0$. (ii) Next, a weak dispersion of plasmons, e.g., of the form $k=\omega/v+\gamma\omega^2{\rm sign}(\omega)$, leads to broadening of wave-packets of the energy width $\epsilon$ and to their overlap. This process takes place at distances $L \gg 1/\gamma \epsilon^2$. As a result, high-order cumulants of the current injected at the QPC are suppressed at distances $L \gg L_{\rm g}= 1/\gamma (T\Delta\mu)^2$, the noise becomes Gaussian, and the derivative of the electron distribution function acquires the shape (\ref{final-r}) with the width $\Gamma= \Gamma_{\rm g} = \pi \Delta\mu T/2$, centered at $\epsilon_0 = \Delta\mu T$. (iii) A situation is possible, where the dispersion of one mode, most likely of the charged mode, is much stronger than the dispersion of the second mode, i.e., $\gamma_1 \gg \gamma_2$. In this case, the previously described regime splits in two separate regimes. First, at distances $L= 1/\gamma_1(T\Delta\mu)^2$ the contribution of the charged mode to high-order cumulants of noise become suppressed, which leads to the distribution (\ref{final-r}) with the parameters $\Gamma = (\Gamma_{\rm ng} + \Gamma_{\rm g})/2$ and $\epsilon_0 = \Delta\mu T/2$. Then, at longer distances $L= 1/\gamma_2(T\Delta\mu)^2$ the noise becomes fully Gaussian. (iv) The interaction may lead to broadening of the wave-packets, but they do not decay, which implies that the interaction alone does not lead to the equilibration. This means that a different, weaker process may lead to the equilibration at distances $L_{\rm eq}$ much longer than the discussed above length scales. In the tunneling regime, $T\ll 1$ the width of the eventual equilibrium distribution scales as $\sqrt{T}$, in contrast to the above regimes, where it scales as $T$. Thus, to observe the described variety of regimes, we propose to perform measurements at large voltage biases and low transparencies of the QPC utilized to inject electrons. Finally, we suggest a possible explanation of the paradox of missing heat flux in the experiment [\onlinecite{exper2}]. So far we have summarized the effects of weak dispersion, which lead to appearance of intermediate length scales. We have found that in the case of a strongly nonlinear dispersion of plasmons, the measured heat flux $I_{\rm m}$ in the outmost edge channel, experimentally determined with the procedure described by Eq.\ (\ref{flux-def}), is different from the actual heat flux per channel $I_{\rm h}/2$, defined by Eq.\ (\ref{actual-f}). The screened Coulomb interaction leads to a rather weak dispersion of the charged plasmon, and the effect is of the opposite sign, because the spectrum in this case is concave. Nevertheless, other mechanisms of the convex dispersion are possible. They will be considered elsewhere. \begin{figure}[h]\begin{center} \epsfxsize=9cm \epsfbox{scales.eps} \caption{Different length scales for energy relaxation processes and corresponding distribution functions in each regime are schematically shown. {\em Red curve}: The initial double-step distribution function. {\em Black curve}: At distances $L\gg \Delta\mu/v$ the distribution function is strongly asymmetric with respect to the ``average'' bias $\langle\Delta\mu\rangle = T\Delta\mu$. {\em Green curve}: At distances $L\gg 1/\gamma (T\Delta\mu)^2$ the distribution function is a Lorentzian with the width that scales as $T\Delta\mu$. {\em Blue curve}: The final equilibrium Fermi function at large distances. For small transparencies its width scales as $\sqrt{T}\Delta\mu$.} \vspace{-3mm} \label{fig-scales}\end{center} \end{figure} \begin{acknowledgments} We would like to thank Fr\'ed\'eric Pierre for the clarification of the experimental details and Pascal Degiovanni for fruitful discussions. This work has been supported by the Swiss National Science Foundation. \end{acknowledgments}
\section{Introduction} The control and manipulation of single electrons in mesoscopic systems constitutes one of the key ingredients in nanoelectronics. The study of single-electron sources\cite{ebbecke04,feve07,blumenthal07,moskalets08,keeling08,kaestner08b,mahe10,battista11} in the high-frequency regime has attracted a great interest due to their potential application in quantum electron optics experiments, in metrology, and in quantum information processing based on fermionic systems.\cite{beenakker05,samuelsson04,samuelsson05,feve08,splettstoesser09,mcneil11,hermelin11} In this work we study the time evolution of a quantum dot (QD) tunnel coupled to a single electronic reservoir, as depicted schematically in Fig.~\ref{fig_scheme}(a). In the presence of some time-dependent voltage modulations, this system defines the building block of the typical single-electron source, namely the mesoscopic capacitor.\cite{buttiker93} In the linear-response regime, the relaxation behavior of such a mesoscopic capacitor has been extensively studied theoretically\cite{nigg06,nigg08,ringel08,nigg09,mora10,hamamoto10,lopez11,filippone11,kashuba11} and experimentally,\cite{gabelli06} revealing the quantization of the charge relaxation resistance.\cite{buttiker93,nigg06,gabelli06,ringel08,mora10,hamamoto10,kashuba11} On the other hand, the application of \textit{nonlinear} periodic potentials to the mesoscopic capacitor yields the controlled emission and absorption of electrons at giga-hertz frequencies.\cite{feve07,mahe10} From these experiments the average charge as well as current correlations\cite{mahe10,albert10,partmentier11} after each cycle of the potential applied have been extracted. These results demonstrate the importance of investigating the dynamics of this kind of single-electron sources. In some of the recent realizations\cite{gabelli06,feve07,mahe10} the Coulomb interaction is weak; however, in small-sized QDs the Coulomb blockade is, in general, strong and it is, therefore, desirable to include it in the theoretical analysis \cite{splettstoesser10a,hamamoto10,mora10,lopez11,filippone11,kashuba11} since it may even dominate time-dependent phenomena, see e.g. Ref.~\onlinecite{reckermann10}. The time-evolution of interacting quantum dots after the coupling to the leads has been switched on, has, e.g., been studied in Refs.~\onlinecite{schoeller10,andergassen10,karrasch10,andergassen11} and references therein. \begin{figure}[b] \includegraphics[width=0.9\linewidth,clip]{fig1.eps} \caption{(Color online) Schematics of the models: a) Single level QD with Coulomb interaction, $U$, coupled to a normal lead with a tunneling strength $\Gamma$. Dot occupations can be measured via the current passing through a nearby quantum point contact (QPC) capacitively coupled to the dot. b) QD attached to an additional superconducting contact. c) QD coupled to a ferromagnetic lead.} \label{fig_scheme} \end{figure} Here we investigate the exponential relaxation of a QD towards its equilibrium state after its has been brought out of equilibrium by applying, e.g., a voltage step pulse. We consider a voltage pulse that affects the occupation of only a single orbital energy level. The level can be spin split due to Coulomb interaction. In an earlier work,\cite{splettstoesser10a} some of the present authors investigated the decay of charge and spin of such a single level QD. It was found that the relaxation of charge and spin are given by rates which differ from each other due to Coulomb repulsion. Since the reduced density matrix of a QD with a single orbital level with spin is four dimensional, there are thus three rates which govern the relaxation of the diagonal elements of the density matrix towards equilibrium (plus one which is always zero and corresponds to the stable stationary state). In addition to the rates that govern charge and spin there is a third rate that appears in the relaxation of a single level QD with spin and with interaction. This additional rate is the subject of this paper. Interestingly, this additional time scale is independent of the interaction and of the dot's level position. It is shown to be related to two-particle effects and appears, e.g., in the time-evolution of the mean squared deviations of the charge from its equilibrium value. We study in detail the perturbations leading to a relaxation of the system with the additional decay rate only, and find that it is indeed related to two-particle correlations. We also propose a procedure to separately read out the different relaxation rates occurring in the dynamics of the QD exploiting the sensitivity of a nearby quantum point contact to the occupation of the QD, see Fig.~\ref{fig_scheme} (a). In order to further clarify the properties of the additional time scale, we extend our study to two other setups: a QD proximized by an extra, superconducting electrode and tunnel coupled to a normal lead; and a QD tunnel coupled to a ferromagnetic lead, see Fig.~\ref{fig_scheme} (b) and (c). \section{Model} We consider a quantum dot coupled to an electronic reservoir. We assume that the single-particle level spacing in the dot is larger than all other energy scales, so that only one, spin-degenerate level of the QD spectrum is accessible. At a certain time $t_0$ the system is brought out of equilibrium, e.g. by applying a gate potential, and afterwards relaxes to an equilibrium dictated by the Hamiltonian $H=H_\mathrm{D}+H_\mathrm{T}+H_\mathrm{res}$. The Hamiltonian $H_\mathrm{D}$ of the decoupled dot \begin{equation} H_\mathrm{D}=\sum_{\sigma}\epsilon d_{\sigma}^{\dagger}d^{}_{\sigma}+U\hat{n}_{\uparrow}\hat{n}_{\downarrow}\ , \label{eq_hdot} \end{equation} contains the spin-degenerate level $\epsilon$ and the on-site Coulomb energy $U$ for double occupation of the dot. The creation (annihilation) operator of an electron with spin $\sigma=\uparrow,\downarrow$ on the dot is denoted by $d_{\sigma}^{\dagger}\left(d^{}_{\sigma}\right)$ and $\hat{n}_{\sigma}$ is the corresponding number operator. The reservoir is modeled by the Hamiltonian $H_\mathrm{res}=\sum_{k,\sigma}\epsilon_{k}c_{k\sigma}^{\dagger}c^{}_{k\sigma}$, in which $c_{k\sigma}^{\dagger}\left(c^{}_{k\sigma}\right)$ creates (annihilates) an electron with spin $\sigma$ and momentum $k$ in the lead. The coupling between the dot and the reservoir is described by the tunneling Hamiltonian $H_\mathrm{T}=\sum_{k,\sigma}(Vc_{k\sigma}^{\dagger}d^{}_{\sigma}+\text{H.c.})$, where $V$ is a tunneling amplitude, which we assume to be independent of momentum and spin. By considering a constant density of states $\nu$ in the reservoir, the tunnel coupling strength $\Gamma$ is defined as $\Gamma=2\pi\nu |V|^2$. In the remainder of this paper, we focus on the relaxation behavior of the quantum dot to its equilibrium state and in particular on how this relaxation manifests itself in measurable quantities. We are not interested in the dynamics of the reservoir, thus the trace over its degrees of freedom is performed to obtain the reduced density matrix of the QD. The Hilbert space is spanned by the four eigenstates of the decoupled dot Hamiltonian, $ \{|\chi\rangle\}$, where $|0\rangle$ represents the unoccupied dot, the dot is in the state $|\sigma\rangle$ when being singly occupied with spin $\sigma=\uparrow,\downarrow$, and $|d\rangle$ is the state of double occupation. The energies related to these states are $E_0=0,E_\sigma=\epsilon$ and $E_\mathrm{d}=2\epsilon+U$, where we set the electrochemical potential of the reservoir to zero. As we consider spin-conserving tunneling events, the off-diagonal elements of the reduced density matrix evolve independently of the diagonal ones (which are the occupation probabilities). We can, therefore, consider these probabilities alone, which arranged in a vector are given by $\mathbf{P}=(p_0,p_{\uparrow},p_{\downarrow},p_d)^\mathrm{T}$ and fulfill the condition $\sum_jp_j(t)=1$. The time evolution of the occupation probabilities is governed by the generalized master equation \begin{equation} \frac{d\mathbf{P}(t)}{dt}=\int_{t_0}^{t}\mathbf{W}(t,t')\mathbf{P}(t')dt'\ , \label{eq_master} \end{equation} where the matrix elements $W_{\chi,\chi'}(t,t')$ of the kernel $\mathbf{W}(t,t')$ describe transitions from the state $|\chi'\rangle$ at time $t'$ to a state $|\chi\rangle$ at time $t$. We consider now the dynamics of the system after being brought out of equilibrium at time $t_0$. Since for $t>t_0$ the total Hamiltonian is time independent, the transition matrix elements depend only on the time difference $t-t'$, i.e. $\mathbf{W}(t,t')\rightarrow\mathbf{W}(t-t')$. Furthermore, we are interested in the exponential decay towards equilibrium. To be more specific, we will therefore consider only the leading, time-independent, prefactor of the exponential functions. Time-dependent corrections to the pre-exponential functions, that generally may appear,\cite{schoeller10,buttiker00} are disregarded. Furthermore, when focussing on times $t$ distant from the switching time $t_0$, such that the difference $t-t_0$ is hence much larger than the decay time of the kernel $\mathbf{W}(t-t')$, we can replace the lower limit of the integral in Eq.~(\ref{eq_master}) by $-\infty$. Expanding the probability vector $\mathbf{P}(t')$ in Eq.~(\ref{eq_master}) around the measuring time $t$ we find\cite{splettstoesser10a} \begin{equation} \frac{d\mathbf{P}(t)}{dt}=\sum_{n=0}^\infty\frac{1}{n!} \partial^n{\mathbf{W}}\cdot\frac{d^n\mathbf{P}(t)}{dt^n}\ . \label{eq_masterexpand} \end{equation} Here we introduced the Laplace transform of the kernel $\mathbf{W}(z)=\int_{-\infty}^{t}\mathbf{W}(t-t')e^{-z(t-t')}dt'$, with $\mathbf{W}=\left.\mathbf{W}(z)\right|_{z=0}$ and the $n$-th derivative of the kernel with respect to the Laplace variable $\partial^n\mathbf{W}=\left[\partial^n\mathbf{W}(z)/\partial z^n\right]_{z=0}$. The formal solution of Eq.~(\ref{eq_masterexpand}) is given by \begin{equation} \mathbf{P}(t) = \exp({\mathbf{A}t})\mathbf{P}^\mathrm{in}\ , \label{eq_exp} \end{equation} which depends on the initial probability vector $\mathbf{P}^\mathrm{in}$ at $t=t_0$, where the initial values for the system parameters are given by the ones just after the switching time $t_0$. The matrix $\mathbf{A}$ includes Markovian and non-Markovian processes.\cite{braggio06} In the following, we consider the limit of weak coupling between quantum dot and reservoir and limit ourselves to a perturbation expansion up to second order in $\Gamma$, which is valid for the regime where the tunnel coupling $\Gamma$ is much smaller than the energy scale set by the temperature $k_\mathrm{B}T$. The perturbative expansion of $\mathbf{A}$ is $\mathbf{A}=\mathbf{A}^{(1)}+\mathbf{A}^{(2)}$ with $\mathbf{A}^{(1)}=\mathbf{W}^{(1)}$ and $\mathbf{A}^{(2)}=\mathbf{W}^{(2)}+\partial\mathbf{W}^{(1)}\cdot\mathbf{W}^{(1)}$, where the number in the superscript represents the power of $\Gamma$ included in the transition matrix $\mathbf{W}$. Notice that the first non-Markovian correction, i.e. the term $\partial\mathbf{W}^{(1)}\cdot\mathbf{W}^{(1)}$ is present in second-order in the tunnel coupling. The evaluation of the kernel within a perturbative expansion can be performed using a real-time diagrammatic technique,\cite{konig96a,konig96b} which has been used in Ref.~\onlinecite{splettstoesser10a} in order to extract the exponential decay of spin and charge in the system studied here. Considering Eq.~(\ref{eq_exp}), we see that the rates defining the decay of the state into equilibrium are found from the eigenvalues of the matrix $\mathbf{A}$, which turn out to be real and non-positive. The matrix $\mathbf{A}$ is not Hermitian, as expected since we deal with a dissipative system, and hence has different left and right eigenvectors, $\mathbf{l}_i$ and $\mathbf{r}_i$. \section{Results} \subsection{Relaxation of the reduced density matrix} The time-dependent probability vector, $\mathbf{P}(t)$, can be expressed in terms of the right eigenvectors of $\mathbf{A}$, each being related to a decay with a different rate. The left eigenvectors determine the observable that decay with a single time scale only, see also the appendix. \\ In the following we discuss the exponential relaxation towards equilibrium of the vector of occupation probabilities, in first order in the tunneling strength $\Gamma$. \subsubsection{Noninteracting limit} We start by briefly discussing the simplest case of a single spinless particle. This limit is obtained, when a magnetic field much larger than the temperature is applied, $B\gg k_\mathrm{B}T$. The Hilbert space of the system is two dimensional and spanned by the states $|0\rangle$ and $|1\rangle$ for the empty and singly-occupied dot respectively, whose occupation probabilities are arranged in the vector $\mathbf{P}=(p_0,p_1)^T$. The decay to the stationary state is governed by matrix $\tilde{\mathbf{A}}^{(1)}$ (defined equivalently to $\mathbf{A}^{(1)}$ but for the two-dimensional Hilbert space for the problem at hand) which contains a single relaxation rate, namely the tunnel coupling $\Gamma$, as intuitively expected. We now include the spin degree of freedom but disregard interactions. The system is described by two independent Hilbert spaces spanned by the states $|0_{\sigma}\rangle$ and $|1_{\sigma}\rangle$ with $\sigma=\uparrow,\downarrow$. The probability vector for each spin $\sigma$ can be written in terms of the eigenvalues and eigenvectors of the matrix $\tilde{\mathbf{A}}^{(1)}$ (for the two-dimensional Hilbert space) as \begin{equation} \mathbf{P}_{\sigma}(t)=\mathbf{P}^\mathrm{eq}_{\sigma}+e^{-\gamma_{\sigma} t}\left(\begin{array}{c} 1\\ -1 \end{array}\right)\left[\langle \hat{n}_{\sigma}\rangle^{\mathrm{eq}}-\langle \hat{n}_{\sigma}\rangle^{\mathrm{in}}\right] \label{eq_psigma} \end{equation} where the right eigenvector corresponding to the eigenvalue zero of $\tilde{\mathbf{A}}^{(1)}$ defines the occupation probabilities for the equilibrium state, $\mathbf{P}^{\text{eq}}_{\sigma}=(p_{0,\sigma}^{\mathrm{eq}},p_{1,\sigma}^{\mathrm{eq}})^\mathrm{T}=(1-f(\epsilon),f(\epsilon))^\mathrm{T}$, with the Fermi function $f(\epsilon)=[1+\exp(\beta\epsilon)]^{-1}$ and the inverse temperature $\beta=1/k_\mathrm{B} T$. Furthermore, $\hat{n}_{\sigma}=(0,1)$ is the vector representation of the number operator for dot electrons with spin $\sigma$, whose initial/equilibrium expectation value is obtained by multiplying it from the left into the initial/equilibrium probability vector, $\langle\hat{n}_{\sigma}\rangle^{\mathrm{in/eq}}=\hat{n}_{\sigma}\cdot\mathbf{P}^\mathrm{in/eq}_{\sigma}$. The rate $\gamma_{\sigma}=\Gamma$ is obtained as the negative of the non-zero eigenvalue of $\tilde{\mathbf{A}}^{(1)}$, with the corresponding left eigenvector being $\mathbf{l}_{\sigma}=(0,1)-\langle\hat{n}_{\sigma}\rangle^{\text{eq}}(1,1)$. The time evolution of the occupation of each spin state is governed by a single decay rate $\Gamma$, \begin{equation} \langle \hat{n}_{\sigma}\rangle(t)=\langle \hat{n}_{\sigma}\rangle^{\mathrm{eq}}\left(1-e^{-\Gamma t}\right)+\langle \hat{n}_{\sigma}\rangle^{\mathrm{in}}e^{-\Gamma t}.\label{nsigma} \end{equation} This equation can be obtained making use of the fact that the time evolution of the expectation value of any operator, which describes an observable of the QD, is given by projecting its vector representation from the left onto Eq. (\ref{eq_psigma}). The time evolution of the total charge of the dot, $\langle \hat{n}\rangle(t)=\langle \hat{n}_{\uparrow}\rangle(t)+\langle \hat{n}_{\downarrow}\rangle(t)$, is also determined by a single relaxation rate $\gamma_{\sigma}=\Gamma$. This means that both charge and spin, which are quantities related with single-particle processes, do not evolve independently from each other and the corresponding decay is given by the same rate. A similar non-interacting problem has been studied \textit{non-pertubatively} in Refs. \onlinecite{moskalets08} and \onlinecite{battista11}. As a next step we consider the squared deviation of the charge from its equilibrium value, $[\hat{n}-\langle \hat{n}\rangle^{\text{eq}}]^2$. Its time evolution is obtained from Eq.~(\ref{eq_psigma}) as \begin{eqnarray} &&\langle[\hat{n}-\langle \hat{n}\rangle^{\text{eq}}]^2\rangle(t)-[\langle\hat{n}\rangle^{\text{eq}}]^2\\ &&=\sum_{\sigma=\uparrow,\downarrow}[1+\langle\hat{n}_{\sigma}\rangle^{\text{eq}}]\langle\hat{n}_{\sigma}\rangle(t)+2\langle\hat{n}_{\uparrow}\hat{n}_{\downarrow}\rangle(t)\nonumber \end{eqnarray} The last, two-particle term of this expression exhibits a decay rate given by $\exp(-2\Gamma t)$. This is in contrast to the spinless case, where such a term does not appear since double occupation is not possible. Such an additional exponential decay with the rate $2\Gamma$ appears directly in the time evolution of the probability vector, when considering the full two-particle Hilbert space spanned by the basis $\{|0\rangle, |\uparrow\rangle, |\downarrow\rangle, |d\rangle\}$. In this basis, Eq.~(\ref{eq_exp}) for the non-interacting regime can be written as: \begin{eqnarray} \label{ptot} \mathbf{P}(t) = \mathbf{P}^\mathrm{eq}+\left(\begin{array}{c} -\left[1-f(\epsilon)\right]\\ \frac{1}{2}\left[1-2f(\epsilon)\right]\\ \frac{1}{2}\left[1-2f(\epsilon)\right]\\ f(\epsilon) \end{array}\right) e^{-\Gamma t}\left(\langle\hat{n}\rangle^\mathrm{in}-\langle\hat{n}\rangle^\mathrm{eq}\right)\nonumber \\ +\left(\begin{array}{c} 0\\ \frac{1}{2}\\ -\frac{1}{2}\\ 0 \end{array}\right) e^{-\Gamma t}\langle\hat{s}\rangle^\mathrm{in} +\left( \begin{array}{c} -1\\1\\1\\-1 \end{array} \right) e^{-2\Gamma t}\left( \langle\hat{m}\rangle^\mathrm{in}-\langle\hat{m}\rangle^\mathrm{eq} \right) \nonumber\\ \end{eqnarray} where as before, $\mathbf{P}^{\mathrm{eq}}$ defines the state at equilibrium. The decaying part of the probability vector can be divided into three contributions which appear depending on how the initial state at $t_0$ differs from the equilibrium state. Deviations of charge and spin from their equilibrium value relax with the same rate $\Gamma$. The corresponding expectation values are calculated by multiplying the probability vector Eq.~(\ref{ptot}) from the left with the vector representation of the operators $\hat{n}=(0,1,1,2)$ and $\hat{s}=(0,1,-1,0)$ which represent the charge and spin, respectively, in this two-particle basis. The two left eigenvectors of the matrix $\mathbf{A}^{\mathrm{(1)}}$ with the same eigenvalue $-\Gamma$, are given by $\mathbf{l}_n=\hat{n}-\langle\hat{n}\rangle^{\mathrm{eq}}(1,1,1,1)$ and $\mathbf{l}_s=\hat{s}$. The third contribution to the decay of the system into the equilibrium comes from the relaxation rate $2\Gamma$, which enters the probability vector in connection with a quantity $\hat{m}$, defined by the operator in vector notation \begin{equation} \hat{m}=\left( 0,f(\epsilon),f(\epsilon),-1+2f(\epsilon) \right). \label{m} \end{equation} The left eigenvector of $\mathbf{A}^{(1)}$ with the eigenvalue $-2\Gamma$ is given by $\hat{m}-\langle\hat{m}\rangle^\mathrm{eq}(1,1,1,1)$. In contrast to charge and spin, the quantity represented by $\hat{m}$ does not have a straightforward intuitive interpretation, since it depends on the quantum dot parameters at $t>t_0$ and on the temperature and chemical potential of the reservoir via the Fermi functions. \subsubsection{Finite Coulomb interaction $U$} From now on we assume a finite on-site Coulomb repulsion $U$ on the dot. Analogously to the noninteracting case discussed before, from Eq.~(\ref{eq_exp}) we can write the time-dependent probability vector in terms of contributions exhibiting different decay times \begin{widetext} \begin{eqnarray} \label{eq_solution} \mathbf{P}(t) & = & \mathbf{P}^\mathrm{eq} +\frac{1}{1-f(\epsilon)+f(\epsilon+U)}\left(\begin{array}{c} -[1-f(\epsilon)]\\ \frac{1}{2}\left[1-f(\epsilon)-f(\epsilon+U)\right]\\ \frac{1}{2}\left[1-f(\epsilon)-f(\epsilon+U)\right]\\ f(\epsilon+U) \end{array}\right)e^{-\gamma_n t}\left(\langle\hat{n}\rangle^\mathrm{in}-\langle\hat{n}\rangle^\mathrm{eq}\right) \nonumber \\ && +\left(\begin{array}{c} 0\\ \frac{1}{2}\\ -\frac{1}{2}\\ 0 \end{array}\right) e^{-\gamma_s t}\langle\hat{s}\rangle^\mathrm{in}+\left( \begin{array}{c} -1\\1\\1\\-1 \end{array} \right) e^{-\gamma_m t}\left( \langle\hat{m}\rangle^\mathrm{in}-\langle\hat{m}\rangle^\mathrm{eq} \right)\ . \end{eqnarray} \end{widetext} Again, $\mathbf{P}^\mathrm{eq}$ is the eigenvector of $\mathbf{A}^{(1)}=\mathbf{W}^{(1)}$ with the zero eigenvalue and represents the equilibrium state in lowest order in the tunnel coupling (the explicit form of the four-dimensional matrix $\mathbf{A}^{(1)}$, together with its entire set of eigenvalues and eigenvectors, is given in the Appendix). In the two-particle basis \mbox{$ \{|\chi\rangle\}=\{|0\rangle, |\uparrow\rangle, |\downarrow\rangle, |d\rangle\}$}, again $\hat{n}=\left(0,1,1,2\right)$ represents the charge operator, and $\hat{s}=\left(0,1,-1,0\right)$ represents the spin operator. The form of the operator $\hat{m}$ is modified by the presence of finite Coulomb interaction; the explicit form will be discussed later in this sub-section (see Eq. (\ref{eq_def_m}) below). The initial and equilibrium expectation values for these operators, entering in the above Eq.~(\ref{eq_solution}), are obtained as $\langle\hat{o}\rangle^{\mathrm{in/eq}}=\hat{o}\cdot\mathbf{P}^\mathrm{in/eq}$, with $\hat{o}=\hat{s},\hat{n},\hat{m}$. Explicit expressions for $\langle\hat{n}\rangle^{\mathrm{eq}}$ and $\langle\hat{m}\rangle^{\mathrm{eq}}$ are shown below. The negative of the other three eigenvalues of $\mathbf{A}^{(1)}$ directly determine the decay of charge, spin,\cite{splettstoesser10a} and the quantity denoted by $\hat{m}$. These decay rates read \begin{subequations} \begin{eqnarray} \gamma_n&=&\Gamma\left[1+f(\epsilon)-f(\epsilon+U)\right]\label{eq_lcharge1}\\ \gamma_s&=&\Gamma\left[1-f(\epsilon)+f(\epsilon+U)\right]\label{eq_lspin1}\\ \gamma_m&=&2\Gamma. \label{eq_lmal1} \end{eqnarray} \end{subequations} Notice that due to interaction, the relaxation rates for charge and spin ($\gamma_n$ and $\gamma_s$ respectively) differ from each other and depend on the level position $\epsilon$, in contrast to the non-interacting case. Their dependence on the level position is shown in Fig.~\ref{fig_decay}. In the region for $-U<\epsilon<0$, $\gamma_n$ is enhanced as the charge decays into the twofold degenerate state of single-occupation, whereas the spin relaxation in first order in $\Gamma$ is suppressed, since spin-flip processes are not possible. However, the third decay rate, $\gamma_m$, remains fully energy independent as in the case with $U=0$. \begin{figure}[h] \includegraphics[width=0.91\linewidth,clip]{fig2.eps} \caption{(Color online) Decay rates $\gamma_m$ (blue, dashed line), $\gamma_n$ (red, dash-dotted line) and $\gamma_s$ (green, solid line) in units of $\Gamma$ as a function of the dot level position $\epsilon$. The temperature is $k_\mathrm{B}T=1.5\Gamma$ and the interaction energy is $U=10\Gamma$. } \label{fig_decay} \end{figure} The right eigenvectors occurring in Eq.~(\ref{eq_solution}) each represent a change to the steady state density matrix that decays exponentially with rate $\gamma_i$ ($i=n,s,m$). Therefore, a system being brought out of equilibrium by a symmetric deviation between $p_\uparrow$ and $p_\downarrow$ only, is decaying with a rate $\gamma_s$. A deviation from equilibrium in which the occupation of the even sector, $p_0+p_\mathrm{d}$ is symmetrically shifted from the odd sector, $p_\uparrow+p_\downarrow$, is governed solely by the relaxation rate $\gamma_m$. This right eigenvector is found to play an important role also in the low-temperature renormalization of this model.~\cite{saptsov} An energy-dependent change in the occupation probabilities as prescribed by the second vector in Eq.~(\ref{eq_solution}) yields a decay of the total charge of the system with the rate $\gamma_n$. The conditions under which specific deviations from the equilibrium state should be performed in order to obtain a specific decay rate, are discussed in the following Section. The attribution of these relaxation rates to the charge, spin, and $\hat{m}$ arises from the independent decay of these quantities, due to the explicit form of the \textit{left} eigenvectors of $\mathbf{A}^{(1)}$. The spin operator coincides with the left eigenvector associated to the eigenvalue $-\gamma_s$ and since it has a vanishing equilibrium value, the time evolution of its expectation value is given by \begin{equation}\label{eq_srelax} \begin{array}{cccc} (0, & 1, & -1, & 0) \end{array}\cdot \left(\begin{array}{c} p_0(t)\\ p_\uparrow(t)\\ p_\downarrow(t)\\ p_\mathrm{d}(t) \end{array}\right) = \langle\hat{s}\rangle(t)=e^{-\gamma_s t}\langle\hat{s}\rangle^\mathrm{in}. \end{equation} Equivalently, the left eigenvector corresponding to the eigenvalue $-\gamma_n$, is $\hat{n}-\langle\hat{n}\rangle^\mathrm{eq}(1,1,1,1)$. It contains the charge operator $\hat{n}$ and its equilibrium value $\langle\hat{n}\rangle^\mathrm{eq}=2f(\epsilon)/\left[1+f(\epsilon)-f(\epsilon+U)\right]$. Hence, for the time evolution of the charge we find \begin{eqnarray}\label{eq_nrelax} \nonumber \begin{array}{cccc} (0, & 1, & 1, & 2) \end{array} &\cdot& \left(\begin{array}{c} p_0(t)\\ p_\uparrow(t)\\ p_\downarrow(t)\\ p_\mathrm{d}(t) \end{array}\right) -\langle\hat{n}\rangle^\mathrm{eq}\\ \nonumber &=& \langle\hat{n}\rangle(t) -\langle\hat{n}\rangle^\mathrm{eq}\\ &=&e^{-\gamma_n t}\left(\langle\hat{n}\rangle^\mathrm{in}-\langle\hat{n}\rangle^\mathrm{eq}\right) . \end{eqnarray} \begin{figure}[h] \includegraphics[width=0.9\linewidth,clip]{fig3.eps} \caption{Equilibrium value of the quantity $\hat{m}$ as a function of the dot level position $\epsilon$. The other parameters are: $k_\mathrm{B}T=1.5\Gamma$ and $U=10\Gamma$. } \label{fig_meh} \end{figure} The quantity decaying with the rate $\gamma_m$ alone is related to the left eigenvector $\hat{m}-\langle\hat{m}\rangle^\mathrm{eq}(1,1,1,1)$, where the operator $\hat{m}$ is given by \begin{eqnarray} \hat{m}=\frac{1}{1-f(\epsilon)+f(\epsilon+U)} \left( \begin{array}{c} 0\\ f(\epsilon+U)\\ f(\epsilon+U)\\ -1+f(\epsilon)+f(\epsilon+U) \end{array}\right)^T\ . \label{eq_def_m} \end{eqnarray} Its expectation value follows a time evolution equivalent to the one for the charge in Eq.~(\ref{eq_nrelax}): $\langle\hat{m}\rangle(t) -\langle\hat{m}\rangle^\mathrm{eq}=e^{-\gamma_m t}\left(\langle\hat{m}\rangle^\mathrm{in}-\langle\hat{m}\rangle^\mathrm{eq}\right)$. Its equilibrium value $\langle\hat{m}\rangle^\mathrm{eq}=f(\epsilon)f(\epsilon+U)/\left[1-f(\epsilon)+f(\epsilon+U)\right]$, plotted in Fig.~\ref{fig_meh}, is - in contrast to spin and charge - not sensitive to the regime of single occupation on the quantum dot. Instead, it exhibits a feature close to the electron-hole symmetric point of the Anderson model, indicating that $\hat{m}$ represents a quantity which is affected by two-particle effects and it decays with a rate that is not modified by the Coulomb interaction $U$. Already for the noninteracting case, we found that the rate $2\Gamma$ appears as a consequence of introducing two particles in the system, and we considered the deviations from equilibrium charge as a quantity involving two-particle processes leading to such a decay rate. Also in the case for finite Coulomb interaction, the time-dependent mean squared deviations $\langle[\hat{n}-\langle \hat{n}\rangle^\mathrm{eq}]^2\rangle(t)$ are suitable to reveal the relaxation rate $\gamma_m=2\Gamma$. Their time evolution is obtained by means of Eq.~(\ref{eq_solution}) and reads \begin{eqnarray} \langle[\hat{n}-\langle \hat{n}\rangle^\mathrm{eq}]^2\rangle(t) - [\langle \hat{n}\rangle^\mathrm{eq}]^2 & = & C\cdot\langle \hat{n}\rangle(t)-2\cdot\langle \hat{m}\rangle(t)\nonumber\\ \label{eq_variance} \end{eqnarray} \begin{figure}[h] \includegraphics[width=0.9\linewidth,clip]{fig4.eps} \caption{(Color online) Equilibrium electron-hole occupation $\langle \hat{n}_{eh}\rangle^\mathrm{eq}$ (red, solid line) and the coefficient $S$ (blue, dashed line) as a function of the dot level position $\epsilon$. The other parameters are: $k_\mathrm{B}T=1.5\Gamma$ and $U=10\Gamma$.} \label{fig_coeff} \end{figure} where in front of the time-dependent charge $\langle \hat{n}\rangle(t)$ the following coefficient appears: \begin{equation} C=-2\langle p_\mathrm{d}-p_0\rangle^\mathrm{eq}+S \\ \label{ceh} \end{equation} with \begin{equation} S=-\frac{1-f(\epsilon)-f(\epsilon+U)}{1-f(\epsilon)+f(\epsilon+U)}.\ \label{eq_coeff} \end{equation} The quantity $\langle p_\mathrm{d}-p_{0}\rangle^\mathrm{eq}=-\left[1-f(\epsilon)-f(\epsilon+U)\right]/\left[1+f(\epsilon)-f(\epsilon+U)\right]$ is the difference between the probability of doubly occupied and empty dot in equilibrium, which can also be related with the occupation of electrons and holes, $\langle \hat{n}_{eh}\rangle=2\langle p_{\mathrm{d}}-p_{0}\rangle$. The behavior of $\langle \hat{n}_{eh}\rangle^\mathrm{eq}$ is shown in Fig.~\ref{fig_coeff}. For $\epsilon<-U$, when the dot is doubly occupied, $\langle \hat{n}_{eh}\rangle^\mathrm{eq}=2$; for $-U<\epsilon<0$, when one electron and one hole are present in the system (singly occupied dot), $\langle \hat{n}_{eh}\rangle^\mathrm{eq}=0$; and for $\epsilon>0$, when the system is completely ``filled with holes'' (empty dot), $\langle \hat{n}_{eh}\rangle^\mathrm{eq}=-2$. The quantity $S$ is also shown in Fig.~\ref{fig_coeff} (blue dashed line), exhibiting a sign change around $\epsilon=-U/2$, the point at which the Anderson model is electron-hole symmetric. By replacing $\epsilon\rightarrow-\epsilon-U$, we go from the electron-like to the hole-like behavior, finding an inversion in the sign of $S$, $S\rightarrow-S$. The function $S$ therefore indicates whether the spectrum of the quantum dot is electron-like or hole-like.\par The mean squared deviations of the charge from its value at equilibrium is an example for a physical quantities showing a decay with $\gamma_m$; it also includes the charge relaxation rate $\gamma_n$, which is found independently from the time evolution of the charge. Equivalently also the time-resolved charge variance, $\langle[\hat{n}-\langle \hat{n}\rangle(t)]^2\rangle(t)$, or the time-resolved spin variance,\cite{notacrooker} $\langle(\hat{s})^2\rangle(t)$, contain a contribution decaying with $\gamma_m$. \subsection{Response to an external perturbation} We now consider in detail which external perturbations are necessary in order to induce a decay of the \textit{full} occupation probability vector with one certain relaxation rate only, in a controlled way. Furthermore, we address the conditions under which a single decay rate can be extracted more easily from the occupation of a single state by a measurement with a nearby quantum point contact (QPC). We first address the case of an infinitesimal perturbation (linear response). A small variation of the gate potential leads to a decay of the charge governed by the charge relaxation rate $\gamma_n$. Similarly, the infinitesimal variation of the Zeeman splitting in the dot yields a decay with the spin relaxation rate $\gamma_s$. In order to obtain a decay of the state with the rate $\gamma_m$ only, it is not sufficient to modulate the gate voltage, also the two-particle term in the Hamiltonian, $U n_\uparrow n_\downarrow$, needs to be varied. The on-site repulsion $U$ could be changed, for example, by tuning the carrier density in a nearby two-dimensional electron gas, thereby controlling the screening of the electron-electron interaction in the dot. From Eq.~(\ref{eq_solution}) we know that a dynamics given only by $\gamma_m$ is obtained if the occupation of the even states are changed in the same direction, opposite to that of the single occupied states; this condition is fulfilled if infinitesimal variations of the gate, $\epsilon\rightarrow\epsilon+d\epsilon$, and of the interaction, $U\rightarrow U+dU$, obey the relation: \begin{equation} dU = -\frac{1+2 \exp(\beta\epsilon)+\exp(\beta[2\epsilon+U])}{1+\exp(\beta\epsilon)}d\epsilon. \label{dvarm} \end{equation} This expression is represented in terms of field lines in Fig.~\ref{fig_field}. An infinitesimal change tangential to the field line passing through the point corresponding to the initial values of $\epsilon$ and $U$ leads to a pure decay with $\gamma_m$. For parameter variations that are not infinitesimal (beyond linear response), a change only of the gate voltage results in a decay of the state with both rates $\gamma_n$ and $\gamma_m$. From Eq.~(\ref{eq_solution}) we find that a finite variation of the energy level and the interaction from an initial condition $(\epsilon_0,U_0)$ to $(\epsilon=\epsilon_0+\Delta\epsilon,U=U_0+\Delta U)$ resulting in a relaxation containing solely $\gamma_n$, satisfies the equation \begin{equation} \beta\Delta U=-\beta\Delta\epsilon+\ln[2-e^{-\beta\Delta\epsilon}]. \label{varn} \end{equation} A relaxation given \textit{only} by the rate $\gamma_m$ is found when the relation: \begin{equation} U=U_0+\frac{1}{\beta}\ln\left[\frac{e^{\beta(\epsilon_0-\epsilon)}\left(1+e^{\beta\epsilon_0}\right)}{1+e^{\beta\epsilon}+e^{\beta(\epsilon+\epsilon_0+U_0)}-e^{\beta(2\epsilon_0+U_0)}}\right] \label{varm} \end{equation} is fulfilled. For different values of $\epsilon_0$ and $U_0$, Eq.~(\ref{varm}) produces again the field lines shown in Fig.~\ref{fig_field}. Therefore, finite variations of the parameters between two points lying on \textit{the same} field line yield a dynamics for the entire occupation probabilities vector $\mathbf{P}$ governed only by $\gamma_m$. Obviously, a generic variation in both $\epsilon$ and $U$ which does not fulfill the conditions specified by Eqs.~(\ref{varn}) or (\ref{varm}) exhibits a dynamics of the probabilities with two time scales: $\gamma_n$ and $\gamma_m$. \begin{figure}[h] \epsfig{file=fig5.eps,width=3.0in} \caption{Field lines describing variations of $\epsilon$ and $U$ that lead to a response of the system with only the rate $\gamma_m$.} \label{fig_field} \end{figure} In Fig.~\ref{fig_field} it is observed that in the region $\epsilon>-U/2$ the field lines are approximately horizontal, i.e, only the interaction $U$ needs to be varied while keeping the level position constant in order to see a dynamics of the probability governed by $\gamma_m$ only. In fact, in this regime the QD is predominantly empty and variations of the interaction strength $U$ do not affect the occupation of the dot. This is the reason why this variation yields a dynamics in which the rate $\gamma_n$ does not contribute. On the other hand, in the region for $\epsilon<-U/2$ in order to avoid that the number of particles on the dot changes, which would lead to a relaxation with rate $\gamma_n$, a variation in $U$ needs to be accompanied by an opposite variation in $\epsilon$, that is $\Delta\epsilon=-\Delta U$. The crossover between the two regimes appears around the symmetry point of the Anderson model, $\epsilon=-U/2$. Importantly, it is also possible to read out either the rate $\gamma_n$ or the rate $\gamma_m$ by varying the gate voltage only (and, thus, not fulfilling Eqs.~(\ref{dvarm}) and (\ref{varm})), which is easier to realize in an experiment. This can be done by measuring an observable that is sensitive to only one occupation probability, for instance the probability of the quantum dot being empty. Such a time-resolved read-out of the probability can be achieved by considering a QPC located nearby the system and tuned such that it conducts only if the QD is empty.~\cite{field93, vander04,mueller10} In the simplest model of the QPC, which assumes a very fast response, the operator corresponding to the current in the QPC is given by \begin{equation} \label{iqp} \hat{i}_{QPC}=i_0\left(1,0,0,0\right), \end{equation} where $i_0$ is a constant current, given by the characteristics of the QPC potential. The expectation value of the QPC current is simply $\langle\hat{i}_{QPC}\rangle(t)=i_0\langle p_0\rangle(t)$. In this way, the QPC effectively measures the dynamics of the occupation probability $p_0$. According to Eq.~(\ref{eq_solution}), a modulation of the gate in which the initial value $\langle\hat{m}\rangle^{\mathrm{in}}$ equals the equilibrium value $\langle\hat{m}\rangle^{\mathrm{eq}}$ leads to a pure decay with $\gamma_n$. Instead, for a decay given by $\gamma_m$ either the factor $\langle\hat{n}\rangle^{\mathrm{in}}-\langle\hat{n}\rangle^{\mathrm{eq}}$ or the factor $\left[1-f(\epsilon)\right]/\left[1-f(\epsilon)+f(\epsilon+U)\right]$ in Eq.~(\ref{eq_solution}) has to vanish. \begin{figure}[h] \includegraphics[width=0.9\linewidth,clip]{fig6.eps} \caption{(Color online) Logplot of the current in the QPC as a function of the time after a finite variation of $\epsilon$. The on-site Coulomb repulsion $U$ is constant and takes the value $U=5 k_BT$. Dashed blue line: $\epsilon$ changes from $\epsilon_0=10 k_BT$ to $\epsilon=-10 k_BT$, its slope yields the relaxation rate $\gamma_m$. Red dot-dashed line: in this case $\epsilon_0 =10 k_BT$ to $\epsilon=2 k_BT$, and the slope leads to $\gamma_n$. The black line is obtained if $\epsilon$ changes from $\epsilon_0=-13 k_BT$ to $\epsilon=2 k_BT$, in which both rates $\gamma_m$ and $\gamma_n$ are present. In all cases we have subtracted the corresponding value for the current in the long-time limit.} \label{lines} \end{figure} Results for the QPC current for different variations of the level position $\epsilon$ while $U$ is kept constant, are shown in the logarithmic plot in Fig.~\ref{lines}. For clarity, we also subtracted the corresponding current in the long time limit, $\langle\hat{i}\rangle^{\mathrm{eq}}$. In particular, for a fixed value of $U$ equal to $5k_BT$, we find that if the level position is changed from $\epsilon_0=10 k_BT$ to $\epsilon=-10 k_BT$, the time evolution of $p_0$ is governed entirely by the rate $\gamma_m$, giving rise to the straight, blue-dashed line in Fig.~\ref{lines}. Its slope is given by $\gamma_m$, making it possible to extract this relaxation rate from measurements of the current in the QPC. However we can obtain a dynamics of $p_0$ given mainly by the rate $\gamma_n$ by performing a variation in $\epsilon$ from $\epsilon_0=10 k_BT$ to $\epsilon=2 k_BT$ which results in the red dot-dashed straight line in Fig.~\ref{lines}; again, the slope yields the corresponding relaxation rate which takes the value $\gamma_n=1.12~\Gamma$. Finally, we show an example in which variations from $\epsilon_0=-13 k_BT$ to $\epsilon=2 k_BT$ (solid black line) produce a dynamics of $p_0$ which includes two exponential decays with rates $\gamma_m$ and $\gamma_n$. As a result, the curve exhibits a change in the slope, showing that a single rate will not be obtained by arbitrary variations of the parameters. \subsection{Second-order corrections in the tunnel coupling} In the previous sections we investigated the relaxation rates in first order in the tunnel coupling strength $\Gamma$. However, corrections due to higher order tunneling processes appear when the tunnel coupling gets stronger. Besides quantitative corrections, this reveals an interesting new aspect. In second order in the tunnel coupling, the matrix $\mathbf{A}^{(2)}$ included in the exponential decay takes the form $\mathbf{A}^{(2)}=\mathbf{W}^{(2)}+\partial \mathbf{W}^{(1)}\cdot \mathbf{W}^{(1)}$. The second-order corrections to the relaxation rates for charge and spin are given by:\cite{splettstoesser10a} \begin{eqnarray} \gamma_\mathrm{n}^{(2)} & = & \sigma(\epsilon,\Gamma,U)\frac{\partial}{\partial\epsilon}\gamma_n+ \sigma_\Gamma(\epsilon,\Gamma,U)\gamma_\mathrm{n}\nonumber\\ & &+2 \frac{ f(\epsilon+U)W_{0\mathrm{d}} + \left[1-f(\epsilon)\right]W_{\mathrm{d}0} } {1-f(\epsilon)+f(\epsilon+U)} \label{eq_charge_corr}\\ \gamma_\mathrm{s}^{(2)} & = & \sigma(\epsilon,\Gamma,U) \frac{\partial}{\partial\epsilon}\gamma_\mathrm{s}+\sigma_\Gamma(\epsilon,\Gamma,U)\gamma_\mathrm{s}+2W_\mathrm{sf}\ . \label{eq_spin_corr} \end{eqnarray} These corrections contain renormalization terms as well as real cotunneling contributions. On one hand, the renormalization terms contain an effect due to the level renormalization $\epsilon\rightarrow\epsilon+\sigma(\epsilon,\Gamma,U)$, with $\sigma(\epsilon,\Gamma,U)=\Gamma[\phi(\epsilon+U)-\phi(\epsilon)]$, $\phi(\epsilon)=\frac{1}{2\pi}\mathrm{Re}\Psi\left(\frac{1}{2}+i\frac{\beta\epsilon}{2\pi}\right)$ and $\Psi(x)$ is the digamma function. On the other hand, the renormalization of the tunnel coupling appears, $\Gamma\rightarrow\Gamma[1+\sigma_\Gamma(\epsilon,\Gamma,U)]$, with $\sigma_\Gamma(\epsilon,\Gamma,U)= -S\left[\Gamma\phi'(\epsilon)+\Gamma\phi'(\epsilon+U)-\frac{2}{U}\sigma(\epsilon,\Gamma,U)\right]$ and where $S$ was defined in Eq.~(\ref{eq_coeff}). Real cotunneling contributions are manifest in terms of spin flips, $W_\mathrm{sf}$, and coherent transitions changing the particle number on the dot by $2$, $W_{0d}$ and $W_{d0}$. These cotunneling terms read \begin{eqnarray} W_\mathrm{sf} & = & -\frac{\Gamma}{\beta}\left[\Gamma\phi''(\epsilon)+\Gamma\phi''(\epsilon+U)-\frac{2}{U}\sigma'(\epsilon,U)\right]\\ W_{\mathrm{d}0} & = & - \frac{2 \Gamma} {e^{\beta(2\epsilon+U)}-1} \left[\Gamma\phi'(\epsilon)+\Gamma\phi'(\epsilon+U)-\frac{2}{U}\sigma(\epsilon,U)\right],\nonumber\\ \end{eqnarray} and $W_{0\mathrm{d}} = \exp[\beta(2\epsilon+U)] W_{\mathrm{d}0}$. The way in which the cotunneling contributions enter in the respective charge and spin relaxation rates is related to the deviation of the state of the QD from equilibrium, given by Eq.~(\ref{eq_solution}) in first order in $\Gamma$. As an example we discuss the correction to the charge decay rate, second line of Eq.~(\ref{eq_charge_corr}). There the factor $2$ appears due to the change in the charge by $\pm2$ in a process bringing the dot from zero to double occupation and vice versa.~\cite{leijnse09} The fraction with which the transition from zero to double occupation, $W_{d0}$, enters the correction to the charge relaxation rate, $\gamma_n^{(2)}$, is given by the deviation from equilibrium of $\mathbf{P}(t)$ in the direction of $p_0$, of the contribution which actually decays with $\gamma_n$ only. This is the first component of the second vector in Eq.~(\ref{eq_solution}). Equivalently, the transition from double to zero occupation, $W_{0d}$, enters with the fraction given by the fourth component of the same vector, namely by the deviation from equilibrium of $\mathbf{P}(t)$ in the direction of $p_d$. Strikingly, in contrast to the charge and spin relaxation rates, $\gamma_m$ does not get renormalized at all by second order tunneling processes: \begin{eqnarray} \gamma_m^{(2)} & = & \gamma_m^{(1)}\sigma_\Gamma(\epsilon,\Gamma,U)+\frac{1-f(\epsilon)-f(\epsilon+U)}{1-f(\epsilon)+f(\epsilon+U)}\left(W_{0d}-W_{d0}\right) \nonumber\\ \label{m2ndorder} & = & 0\ , \end{eqnarray} The reason for this is that the contribution due to $\Gamma$ renormalization and those due to coherent processes between empty and doubly occupied dot, cancel each other. The lack of second order corrections, confirms that this relaxation rate is related to a quantity which is not sensitive to the Coulomb interaction. The fact that corrections are missing, is also found using a renormalization-group approach.~\cite{saptsov}. Another important aspect of this missing second-order correction is that it is due to an exact cancelation of the contribution due to virtual second order processes, namely the $\Gamma$-renormalization, with real cotunneling contributions. This is in contrast to, e.g. the conductance, where only the real cotunneling processes contribute far from resonances, while renormalization terms are limited to the resonant regions. \subsection{Hybrid systems} Until now, we considered the quantum dot to be coupled to a normal conducting lead. However, the vicinity of a superconducting or a ferromagnetic reservoir induces correlations between electrons and holes or between charge and spin, respectively. In the following we study, in first order in the tunnel coupling strength $\Gamma$, the influence of induced correlations on the relaxation rates of the dot. The charge response of a noninteracting mesoscopic scattering region coupled to both normal and superconducting leads has been studied in Refs.~\onlinecite{Pilgram02,xing07}. \subsubsection{Proximity to a superconducting lead} In the previous sections we have seen that the rate $\gamma_m$, which together with the time decay of charge and spin determines the relaxation of the QD to the equilibrium state, is independent of the level position and the Coulomb interaction and that it enters in the time evolution of quantities sensitive to two-particle effects. It is therefore expected that the rate $\gamma_m$ will directly influence the relaxation of the charge towards the equilibrium in a setup that naturally mixes the empty and doubly occupied states of the dot. This situation is obtained if the QD is not only coupled to a normal lead (with tunnel coupling strength $\Gamma$) but also to an \textit{additional} superconducting contact (with tunnel coupling strength $\Gamma_S$), as shown in Fig.~\ref{fig_scheme} (b). We consider only the case when the superconductor is kept at the same chemical potential as the normal lead and we set both chemical potentials to zero. The only purpose of the extra lead is here to induce superconducting correlations on the dot via the proximity effect. To the original Hamiltonian, $H_D+H_\mathrm{res}+H_T$, we now add the Hamiltonian for the superconducting contact and its tunnel coupling to the QD, \begin{eqnarray} H_\mathrm{S} & = & \sum_{k\sigma}\epsilon_{\mathrm{S} k}c^\dagger_{\mathrm{S} k\sigma}c^{}_{\mathrm{S} k\sigma}-\sum_k\left(\Delta c^{}_{\mathrm{S}-k\downarrow}c^{}_{\mathrm{S} k\uparrow}+\mathrm{H.c.}\right)\nonumber\\ &&+\sum_{k,\sigma}(V_\mathrm{S} c_{\mathrm{S}k\sigma}^{\dagger} d_{\sigma}+\mathrm{H.c})\, . \end{eqnarray} where $c^{(\dagger)}_{\mathrm{S} k\sigma}$ is the the annihilation (creation) operator of electrons in the lead. In the limit of a large superconducting gap $\Delta$ the effect of the additional contact can be cast in an effective Hamiltonian of the dot which includes a coupling between electrons and holes in the QD, $H_{D}^\mathrm{(eff)}=H_D -\Gamma_s/2(d_{\downarrow}^{\dagger}d_\uparrow^{\dagger}+ \mathrm{H.c})$. The eigenstates of the proximized dot are the states of single occupation $|\sigma\rangle$ and other two states which are superpositions of the empty and double occupied states of the dot (due to Andreev reflection): \begin{equation} |\pm\rangle=\frac{1}{\sqrt{2}}\sqrt{1\mp\frac{\delta}{2\epsilon_A}}|0\rangle\mp\frac{1}{\sqrt{2}}\sqrt{1\pm\frac{\delta}{2\epsilon_A}}|d\rangle \label{eq_pmvectors} \end{equation} with energies given by $E_{\pm}=\delta/2\pm\epsilon_A$, where the level detuning between $|0\rangle$ and $|d\rangle$ is $\delta=2\epsilon+U$ and $2\epsilon_A=\sqrt{\delta^2+\Gamma_{s}^{2}}$ is the energy splitting between the $|+\rangle$ and $|-\rangle$ states.\cite{braggio11,eldridge10} In the new basis $\left\{|+\rangle,|\uparrow\rangle,|\downarrow\rangle,|-\rangle\right\}$, the vector representing the charge operator is expressed as $\hat{n}=\left(-\sqrt{2+\frac{\delta}{\epsilon_A}},1,1,\sqrt{2-\frac{\delta}{\epsilon_A}}\right)$ and we expect that the effect of the mixing of electrons and holes will be visible in its time evolution. In first order in the tunnel-coupling strength to the normal reservoir $\Gamma$ and assuming $\Gamma\ll\Gamma_{S}$, we find the relaxation rates \begin{eqnarray} \gamma_{S,1}&=&\Gamma\left[1+f\left(\epsilon-E_-\right)-f\left(E_+-\epsilon\right)\right]\\ \gamma_{S,s}&=& \Gamma\left[1-f(\epsilon-E_-)+f(E_+-\epsilon)\right]\\ \gamma_{S,2}&=&2\Gamma. \end{eqnarray} Remarkably the eigenvalue $-2\Gamma=-\gamma_{S,2}$ remains unaffected, i.e. $\gamma_m=\gamma_{S,2}$ is not modified by the presence of the additional superconducting lead. The spin on the dot, which is determined by the occupation probabilities of singly occupied states, still decays with a single relaxation rate given by $\gamma_{S,s}$, i.e. $\mathbf{l}_{s}=\left(0,1,-1,0\right)$ is an eigenvector of the kernel $\mathbf{A}^{(1)}$ (in the proximized basis). In contrast, the decay of the charge to its equilibrium value is given by \begin{eqnarray} \langle\hat{n}\rangle_{SC}(t) & = & \frac{1}{2} \left[ \langle \hat{n}\rangle^{\mathrm{in}}-\langle \hat{n}\rangle^{\mathrm{eq}} \right] \left(e^{-\gamma_{S,2} t}+e^{-\gamma_{S,1}t} \right) +\langle \hat{n}\rangle^{\mathrm{eq}}\nonumber\\ & & +a_{SC} \frac{1}{2}\left[\langle x\rangle^{\mathrm{in}}-\langle x\rangle^{\mathrm{eq}}\right] \left(e^{-\gamma_{S,2} t}-e^{-\gamma_{S,1}t}\right) \nonumber \\ & & +\frac{1}{2}\left[\langle y\rangle^{\mathrm{in}}-\langle y\rangle^{\mathrm{eq}}\right] \left(e^{-\gamma_{S,2} t}-e^{-\gamma_{S,1}t}\right) \label{eq_nsct} \end{eqnarray} with \begin{eqnarray} a_{SC} & = & (2-k_--k_+)\frac{f\left(E_+-\epsilon\right)-f\left(E_--\epsilon\right)}{f\left(E_+-\epsilon\right)+f\left(E_--\epsilon\right)} \nonumber \label{factors_nsct} \end{eqnarray} and where we defined the difference in the occupation of the $|\pm\rangle$ states, $x=p_+-p_-$ and the quantity $y=(k_--1)p_++(k_+-1)p_-$, with $k_{\pm}=\mp\sqrt{2\pm\frac{\delta}{\epsilon_A}}$. The charge evolves with two different time scales, $\gamma_{S,1}$ and $\gamma_{S,2}=\gamma_m$, instead of only one as in the normal case. This is a direct consequence of the mixing of the states $|0\rangle$ and $|d\rangle$ induced by the superconducting contact. This effect opens the possibility to extract this rate by measuring the time evolution of the charge in the proximized dot. \subsubsection{Ferromagnetic lead} Even though the presence of a superconducting lead couples electrons and holes, the relaxation rate $\gamma_m$ has not been modified. Since we associate this rate with processes involving two particles each with spin $\sigma$, it is expected that if the spin symmetry is broken by introducing a ferromagnetic contact, the rate $\gamma_m$ will now be the sum of the tunneling rates for spin up and spin down electrons. In order to verify this, we consider the Hamiltonian used for the normal case and assume a spin-dependent density of states in the only reservoir attached to the quantum dot, see Fig.~\ref{fig_scheme} (c). This leads to spin-dependent tunnel couplings, $\Gamma_\uparrow$ and $\Gamma_\downarrow$, which are included in the corresponding transition matrix $\mathbf{A}^{(1)}$. Diagonalization of $\mathbf{A}^{(1)}$ yields the three relaxation rates: \begin{eqnarray} \gamma_{F,1} & = & \Gamma +\frac{1}{2}\sqrt{\left(\Delta\Gamma\right)^2+4\Gamma_\uparrow\Gamma_\downarrow\left[f(\epsilon)-f(\epsilon+U)\right]^2} \\ \gamma_{F,2} & = & \Gamma-\frac{1}{2}\sqrt{\left(\Delta\Gamma\right)^2+4\Gamma_\uparrow\Gamma_\downarrow\left[f(\epsilon)-f(\epsilon+U)\right]^2} \\ \gamma_{F,m}&=& 2 \Gamma \end{eqnarray} with $\Gamma=\frac{1}{2} \left(\Gamma_\downarrow+\Gamma_\uparrow\right) $ and $\Delta\Gamma=\Gamma_\uparrow-\Gamma_\downarrow$. As in the normal case, there is an eigenvalue which does not depend on the level position nor on the interaction but on the sum of the different tunneling rates: $-2\Gamma\rightarrow-\left(\Gamma^{\uparrow}+\Gamma^{\downarrow}\right)$. The appearance of such a combination of the spin-dependent tunneling strengths in the relaxation rate, confirms the statement that two-particle processes involving electrons with both spin polarizations are at the basis of the decay rate $\gamma_m$. Due to the ferromagnetic lead, the dynamics of spin and charge are now mixed. The corresponding time evolution in first order in the tunnel coupling takes the form: \begin{eqnarray} \langle \hat{s}\rangle_F(t) & = & \frac{1}{2}\langle \hat{s}\rangle^{\mathrm{in}}(e^{-\gamma_{F,1}t}+e^{-\gamma_{F,2}t})\nonumber\\ & & +a_s\langle \hat{s}\rangle^{\mathrm{in}}(e^{-\gamma_{F,1}t}-e^{-\gamma_{F,2}t}) \label{spin_ferro}\\ & & +b_s\left[\langle \hat{n}\rangle^{\mathrm{in}}-\langle \hat{n}\rangle^{\mathrm{eq}}\right](e^{-\gamma_{F,1}t}-e^{-\gamma_{F,2}t}) \nonumber \\ \langle\hat{n}\rangle_F(t) & = & \frac{1}{2}\left[\langle \hat{n}\rangle^{\mathrm{in}}-\langle \hat{n}\rangle^{\mathrm{eq}}\right](e^{-\gamma_{F,1}t}+e^{-\gamma_{F,2}t}) +\langle \hat{n}\rangle^{\mathrm{eq}}\nonumber \\ & & +a_c\left[\langle \hat{n}\rangle^{\mathrm{in}}-\langle \hat{n}\rangle^{\mathrm{eq}}\right](e^{-\gamma_{F,1}t}-e^{-\gamma_{F,2}t})\nonumber \\ & & +b_c\langle \hat{s}\rangle^{\mathrm{in}}(e^{-\gamma_{F,1}t}-e^{-\gamma_{F,2}t})\label{eq_decay_ferro} \end{eqnarray} where we introduced the abbreviations: \begin{eqnarray} a_s & = & \frac{\Gamma[f(\epsilon)-f(\epsilon+U)]}{2\sqrt{\Delta\Gamma^2+4\Gamma_\uparrow\Gamma_\downarrow\left[f(\epsilon)-f(\epsilon+U)\right]^2}} \nonumber \\ b_s & = &\frac{\Delta\Gamma[1+f(\epsilon)-f(\epsilon+U)]}{2\sqrt{\Delta\Gamma^2+4\Gamma_\uparrow\Gamma_\downarrow\left[f(\epsilon)-f(\epsilon+U)\right]^2}} \nonumber \\ a_c & = & a_s\nonumber \\ b_c & = & \frac{\Delta\Gamma[1-f(\epsilon)+f(\epsilon+U)]}{2\sqrt{\Delta\Gamma^2+4\Gamma_\uparrow\Gamma_\downarrow\left[f(\epsilon)-f(\epsilon+U)\right]^2}}. \nonumber \end{eqnarray} The last term in Eq.~(\ref{spin_ferro}) shows that at finite time $t$ the initial charge influences the time evolution of the spin; similarly, the initial spin enters explicitly in the dynamics of the charge, Eq.~(\ref{eq_decay_ferro}). These terms persist in the non-interacting limit, revealing that the coupled evolution of charge and spin including two relaxation rates (which for the non-interacting case take the form $\gamma_{F,1}=\Gamma_\uparrow$ and $\gamma_{F,2}=\Gamma_\downarrow$) is a direct consequence of the presence of the ferromagnetic contact. In contrast, the factor $a_s=a_c$ vanishes for $U=0$ implying that it stems from the combined effect of the Coulomb interaction and the breaking of the spin symmetry. As expected the independent evolution of charge and spin is recovered in the limit $\Gamma_\uparrow=\Gamma_\downarrow$. The mixing of the dynamics of both, charge and spin, induced here by a ferromagnetic lead was found in Ref. \onlinecite{splettstoesser10a} for the case of lifted spin-degeneracy in the dot due to a finite Zeeman splitting. Note that for the hybrid as well as for the normal system, the sum of the energy-dependent relaxation rates equals $2\Gamma$, as long as the tunnel coupling $\Gamma$ is treated in first order, only. \section{Conclusion} We have studied the different time scales present in the evolution of the reduced density matrix of a single-level QD with Coulomb interaction and tunnel coupled to a single reservoir, after being brought out of equilibrium. Besides the relaxation rates for charge and spin, we find an additional rate $\gamma_m=2\Gamma$, which is independent of the energy level of the dot as well as of the interaction strength. This relaxation is related to the presence of two particles in the dot and is found to be not sensitive to the Coulomb interaction. The time evolution of the square deviations of the charge from its equilibrium value is proposed as a physical quantity related with processes involving two-particles leading to the rate $2\Gamma$. In order to further elucidate the properties of this decay, we analyzed the response of the system to specific variations of both, the interaction strength $U$ and the level position $\epsilon$, finding that $\gamma_m$ can be extracted from time-resolved measurements of the current passing through a nearby quantum point contact. Additionally, we analyzed two other setups: a dot proximized by a superconductor and coupled to a normal reservoir, and a dot coupled to a ferromagnetic lead. In the hybrid normal-superconducting systems, we found that the time-resolved read-out of the charge represents another possibility to get access to the rate $\gamma_m$. \begin{acknowledgments} We thank Michael Moskalets, Roman Riwar and Maarten Wegewijs for fruitful discussion. Financial support by the Ministry of Innovation, NRW, the DFG via SPP 1285 and KO 1987/5, the European Community's Seventh Framework Programme under Grant Agreement No. 238345 (GEOMDISS), as well as the Swiss National Science Foundation, the Swiss centers of excellence MaNEP and QSIT and the European Marie Curie ITN, NanoCTM is acknowledged. \end{acknowledgments}
\section{\label{sec:intro}Introduction} In the field of high temperature superconductor (HTS) research and development, a significant hurdle has proved to be the anisotropic properties of the available materials. For a typical YBa$_2$Cu$_3$O$_7$ (YBCO) sample, the variation in the critical current density with the angle of an applied magnetic field, $J_c(\theta)$, can exceed a factor ten at high fields \cite{Coulter1999}. Successful methods of tackling this issue have been developed, most notable of which has been the incorporation of artificial pinning centers in the form of self-assembled BaZrO$_3$ nanocolumns \cite{Driscoll2004}. These have introduced a so-called $c$-axis peak to the pinning profile, enhancing $J_c$ for fields applied in the $c$-direction, perpendicular to the film surface. Rather than solving the problem, however, this has merely shifted the region of concern to intermediate angles, between the two peaks, where $J_c$ remains low. Adding BaZrO$_3$ nanoparticles creates a more isotropic profile but the effect varies substantially with field and temperature \cite{Miura2011}. In order to fully overcome this issue by engineering appropriate defect structures, we need first to understand the origin of the $J_c(\theta)$ dependence. Remarkably, no consensus model of $J_c(\theta)$ exists, highlighting the general weakness in established models of $J_c$ to reproduce the phenomenology of critical currents. One proposed origin of at least some of the features of $J_c(\theta)$ gaining momentum has been the electronic mass anisotropy of the materials \cite{Xu1994,Kumar1994,CivaleJLTP2004,*CivaleAPL2004}. However, the physical link between this intrinsic material anisotropy and (extrinsic) properties such as $J_c$ has not been made clear. While application of this approach to understand the physics of pristine samples of clean superconductors \cite{Kidszun2011} may have some merit, its direct transfer to explaining the pinning in technological materials incorporating artificial pinning centers, resulting in ``effective electronic mass anisotropy ratios'' that bear no relation to the fundamental material parameters \cite{Gutierrez2007} is questionable. We therefore begin by reviewing the physics of the electron mass anisotropy scaling approach and discuss whether this approach is applicable to high-$J_c$ pinning-engineered samples. Additional to this, many recent pinning-optimized samples \cite{Maiorov2007} exhibit features such as pronounced shoulders \cite{Chudy2010} and flat angular regions \cite{Mikheenko2011} in $J_c(\theta)$ that cannot be explained by the naive summation of an $ab$-plane intrinsic pinning peak, a $c$-axis correlated pinning peak and a random pinning contribution characterized by a geometrical anisotropy of whatever origin. The presence of similar features throughout the literature from the earliest samples produced \cite{Roas1990,BraithwaiteJLTP1993a,*BraithwaiteJLTP1993b} to present-day `champion' samples \cite{Maiorov2009} means that these can not be dismissed as due to poor sample quality or inaccurate measurement, but must be acknowledged as genuine effects of an optimized pinning landscape. Combine this with the complex temperature and magnetic field dependence of $J_c$ now starting to be revealed \cite{Xu2010} and it is apparent that a deeper understanding of the effects of a complex flux pinning landscape is required. Here we demonstrate that an analysis of the statistical effects of combining $c$-axis correlated pinning and $ab$-plane correlated pinning (whether intrinsic, extrinsic, or surface in origin) can explain all the features observed to date in pinning-optimized samples. Isotropic pinning resulting from point defects or incorporated second-phase nanoparticles contributes to this picture in the form of `bridges' between these defining pinning populations. The aim of this paper is to show that the model that results from this analysis, termed the vortex path model, provides a consistent and detailed explanation of the experimental features that have risen to prominence in recent studies. Using this model, $J_c(\theta)$ is decomposed into statistically distinguishable components that can be related back to distinct aspects of the sample microstructure that determines $J_c$ --- a link that is missing in other approaches. \section{\label{sec:models}Models of $J_c(\theta)$} \subsection{\label{sec:anisotropy}Electron mass anisotropy scaling} The electron mass anisotropy scaling approach has recently become a common method of analyzing $J_c(\theta)$ data obtained from diverse species of superconducting materials \cite{Choi2004,Paturi2008,Iida2010,Pardo2011}. In some cases misinterpretation of the process has lead to simple fits of the Ginzburg-Landau mass anisotropy function to $J_c(\theta)$ being employed \cite{Crisan2010,*Dang2010}. It is therefore worth reviewing the reasoning behind this approach to understand the limits of its applicability. If the scaling approach is valid then this affects how we should interpret the various features of $J_c(\theta)$. The scaling approach was originally proposed by Blatter \emph{et al.}\ \citetext{\citealp{[(a) ]Blatter1992,*[(b) ]Blatter1994,*[(c) ]Blatter2008}(a)}, with a similar formulation being independently proposed by Hao and Clem \cite{Hao1992}. The idea is to apply results derived for an isotropic superconductor to an anisotropic superconductor through a rescaling of the magnetic field, temperature, characteristic lengths, etc. The scaling rules are obtained by finding a transformation which recasts the Gibbs free energy of an anisotropic superconductor into the form for an isotropic material. The primary result is then a scaling rule such as \begin{equation} Q(\theta,H,T,\xi,\lambda_L,\varepsilon,\delta) = s_Q\tilde{Q}(\varepsilon_\theta H,T/\varepsilon,\xi,\lambda_L,\delta/\varepsilon). \label{eq:scalingrule} \end{equation} Blatter \emph{et al.}\ then apply this rule to calculate critical currents for the case of single-vortex weak collective pinning, that is pinning by randomly distributed point pins. The results are $J_c^\parallel = J_c^c$, independent of $\theta$, for the in-plane critical current density (as usually measured in thin film experiments), where $J_c^c$ is the in-plane $J_c$ with the field applied in the $c$-direction, and $J_c^\perp = \varepsilon_\theta J_c^c$ for the out-of-plane critical current density, with $\varepsilon_\theta^2 = \varepsilon(\theta) = \cos^2\theta + \varepsilon^2\sin^2\theta$, where $\varepsilon^2 = m_{ab}/m_c$ is the electronic mass anisotropy. For consistency throughout our paper, and in contrast to Blatter, we have taken $\theta = 0$ to be perpendicular to the $ab$-planes. The result $J_c^\parallel = \mathrm{constant}$ for low fields is consistently ignored by those who use the scaling approach to analyse $J_c(\theta)$ data. We are not aware of any report claiming to have confirmed this result experimentally, thereby validating the scaling hypothesis. Many samples reported in this paper and elsewhere have wide regions of $J_c^\parallel = \mathrm{constant}$ at low field. However, these samples also have correlated $c$-axis defects such as grain boundaries, threading dislocations and twin plane intersections and it would be puzzling to assume these made no contribution to pinning in these circumstances. The scaling procedure that is widely used in the literature is based on an expression postulated in a subsequent review by Blatter \emph{et al.}\ \citetext{\citealp{Blatter1994}(b)} for the high field case, defined by $H(\theta) > \beta_{sb}(J_c^c/J_d)H_{c2}(\theta)$ where $\beta_{sb} \approx 5$ and $J_d$ is the depairing current density. In this case, no physical model of $J_c$ to which the scaling rules can directly be related is proposed. Rather, the authors simply assert that at high field, $J_c$ will scale in accordance with $H \rightarrow \varepsilon_\theta H$, i.e. \begin{equation} J_c^\parallel(\theta,H) = J_c^c(\varepsilon_\theta H). \label{eq:massanisotropyscaling} \end{equation} To evaluate the status of this assertion we recall that the scaling rule, Eq.~(\ref{eq:scalingrule}), is a non-unique mathematical transformation that allows us to write the equilibrium free energy for an anisotropic superconductor in the same form as for an isotropic superconductor. It is not a physical theory where a \emph{measurement} of a physical quantity may actually give a different answer depending on the value of another physical variable. In particular, a \emph{measurement} of the field in the sample does not scale with these rules. Any change in the critical current must therefore have an indirect mechanism, which this approach does not address. The critical current density is a non-equilibrium property of the sample; formally, it is a parameter in a constitutive equation such as $E = E_0(J/J_c)^n$. There is no formal derivation possible from Eq.~(\ref{eq:scalingrule}) to Eq.~(\ref{eq:massanisotropyscaling}); rather, Eq.~(\ref{eq:massanisotropyscaling}) is a new hypothesis. The justification for the adoption of this hypothesis, noted in both \citetext{\citealp{Blatter1994}(b)} and \cite{Hao1992}, is a perceived similarity between the high field behavior of $J_c(\theta)$ and the field-scaling rule for the free energy. The justification given is empirical, not theoretical. This justification, however, is not well founded. If a broad peak in $J_c(\theta)$ is determined by a variation in pinning density with angle, for example arising from the interaction of correlated defects, rather than from the mass anisotropy of the material, then we would still expect $J_c(\theta)$ data to be approximately scaleable in this way. Typically, for HTS samples, critical currents in large out-of-plane fields behave as $J_c(H{\parallel}c) \propto H^{-\alpha}$. If, for an in-plane field, we have a higher density of correlated pinning defects but otherwise similar physics we would expect $J_c(H{\parallel}ab) \propto (H/k)^{-\alpha}$, where $k$ is a measure of the change in pinning density between the field directions. We can scale $H \rightarrow H\varepsilon(k,\theta)$ so the data coincide at these two angles using an equation of the same form as the mass anisotropy, $\varepsilon^2(k,\theta) = \cos^2\theta + k^2\sin^2\theta$. Further, if $J_c(\theta)$ varies smoothly and broadly then $\varepsilon(k,\theta)$ will also give a reasonable fit to all the intermediate angle data. The electron mass anisotropy expression can therefore be a reasonable fitting function whether the origin of the behavior is in fact electronic mass anisotropy or simply pinning density anisotropy arising from correlated defects. Experimentalists who employ this approach assume that correlated pinning can only be effective and therefore influence their data over narrow angular ranges. The evidence from samples implanted at different angles by heavy ions \cite{Holzapfel1993, Strickland2009} or the addition of self-assembled nanocolumns \cite{Goyal2005} suggests this confidence is misplaced. Similar to the low field result, those who apply this scaling also consistently ignore the lower bound to the field range over which it is applicable and do not attempt to verify any crossover from the low field to the high field regime. The lower bound quantified above can be significant for HTS since $H_{c2}$ is large; at low temperature it may easily be tens of teslas, beyond most common experimental capability. Aside from these criticisms the scaling approach suffers from one further fatal flaw when applied to pinning-optimized samples: it ignores the pinning summation problem. If we add random isotropic defects to a sample that already possesses a high density of $ab$-plane pinning but much lower $c$-axis pinning then we expect the effect on $J_c(\theta)$ about the $c$-axis direction to be much stronger than the effect about the $ab$-plane. This is exactly what is seen experimentally \cite{Miura2011,Chudy2010,Long2005}. As individual vortices interact with multiple defects, correlated and uncorrelated, the effect on $J_c$ of additions to the pinning landscape depends on the pinning already present and the scaling approach has no way of taking this into account. The use of the high field scaling approach to analyze $J_c(\theta)$ has led to a situation where a correspondence between the experimental data and the known defect structures of HTS films is missing. We think this approach is wrong and should be abandoned. In the next section we recapitulate a model that can distinguish how different defect populations within the sample contribute to $J_c$, and that can enumerate how these distinct contributions combine to yield the overall $J_c(\theta)$. \subsection{\label{sec:VPM}The vortex path model} A statistical mechanical approach to modeling the response of a superconductor to a particular population of pinning defects was first proposed in \cite{Long2007} and formally presented in \cite{Long2008}. It was subsequently termed the vortex path model in \cite{Paturi2010} since at its core, it considers the multiplicity of possible pinned vortex paths through a sample, as determined by its defect structure, to quantify the relative pinning strength for fields applied at different angles. It has recently been used to describe experimental data where the electronic mass anisotropy approach has been found inadequate \cite{Mikheenko2010}. Here we do not repeat the full exposition of the model but instead give a general theoretical context as an introduction. The model is constructed in the spirit of a maximum entropy approach \cite{JaynesPR1957a,*JaynesPR1957b}. If we consider any macroscopic property of a system, where the macroscopic state can be related to a multiplicity of microstates, then to predict the behavior of the property we maximize the Shannon entropy of the distribution of microstates of the system subject to the known constraints. For elementary constraints this leads to the ubiquitous functions of physics; for example, the Gaussian function is obtained upon maximizing the Shannon entropy subject to the constraints that the mean and the variance of the distribution of microstates are fixed, while the (truncated) Lorentzian function is obtained when only the variance is fixed \cite{Carazza1976}. The critical current density of a superconductor is just such a macroscopic property arising from the multiplicity of microstates in which vortices are pinned at a particular angle, and following \cite{Long2008} we can write \begin{equation} J_c(\theta) = J_0 g(\theta), \label{eq:Jc} \end{equation} where $g(\theta)$ is the density distribution of these pinning microstates normalized such that ${\int_0^\pi}g(\theta)d\theta = 1$ and the proportionality factor $J_0 = {\int_0^\pi}J_c(\theta)d\theta$. In \cite{Long2008}, a Lorentz force definition of the critical current was used to define $J_0$ in terms of the pinning force. However, this definition of $J_c$ is unnecessary, and instead Eq.~\ref{eq:Jc} itself serves to define the critical current. We can then maximize the Shannon entropy \begin{equation} H = -{\int_0^\pi}g(\theta)\ln(g(\theta))d\theta \label{eq:entropy} \end{equation} with appropriate constraints in order to predict $J_c(\theta)$. Without any constraints the result is a uniform distribution and we obtain a constant $J_c$, \begin{subequations} \label{eq:VPMequations} \begin{equation} J_c(\theta) = \frac{J_0}{\pi}. \label{eq:uniform} \end{equation} For a sample containing correlated defects within the $ab$-plane and along the $c$-axis we know that $J_c(\theta)$ may exhibit peaks centered on the directions of the defects. These peaks are broadened by the vortices' interactions with the orthogonal defect populations. The vortex path model introduces constraints on $g(\theta)$ by requiring the vortex microstates to follow a path which interacts with these defects. In the usual geometry for $J_c$ measurements we define $\theta = \tan^{-1}(y/z)$, where $\hat{y}$ lies within the plane of the sample and $\hat{z}$ is perpendicular to it. For a peak centered on $\hat{y}$, we constrain the pinned vortex path to comprise $m$ steps of average $y$-distance $\lambda$, so $y = m\lambda$ is the vortex length in the $y$-direction, and $m$ steps of length $z_i$ in the $z$-direction, where the distribution of lengths $z_i$ is $p(z_i)$. The total $z$ displacement is then given by $z = \sum z_i$, with a distribution $p(z)$. The principle of maximum entropy can be applied and we hypothesize entropy maximizing distributions of $p(z)$ and hence $g(\theta)$, leading to different results for $J_c(\theta)$ through a transformation of random variables from $(y, z)$ to $\theta$. If the distribution $p(z)$ is Gaussian, we obtain \begin{equation} J_c(\theta) = \frac{J_0}{\sqrt{2\pi}\Gamma\sin^2\theta}\exp\left(-\frac{1}{2\Gamma^2\tan^2\theta}\right), \label{eq:gaussian} \end{equation} while if $p(z)$ is Lorentzian, we obtain \begin{equation} J_c(\theta) = \frac{1}{\pi}\frac{J_0\Gamma}{\cos^2\theta + \Gamma^2\sin^2\theta}. \label{eq:lorentzian} \end{equation} \end{subequations} We term Eqs.~(\ref{eq:gaussian}) and (\ref{eq:lorentzian}) angular Gaussian and angular Lorentzian functions, respectively. In Eq.~(\ref{eq:gaussian}), $\Gamma = \sigma/\sqrt{m}\lambda$, where $m\sigma^2$ is the variance of the Gaussian, while in Eq.~(\ref{eq:lorentzian}), $\Gamma = \gamma/\lambda$, where $m\gamma$ is the scale factor of the Lorentzian. Note that Eq.~(\ref{eq:lorentzian}) incorporates the special case $\Gamma = 1$ for which it reduces to the uniform function of Eq.~(\ref{eq:uniform}). To explain this further, the choice of a Gaussian distribution for $p(z)$ indicates that there exists a population of orthogonal defects that broaden the peak with the only added constraints that $p(z)$ has a mean and a variance, i.e.\ that the lengths of pinned vortex segments in the $z$-direction form a normal distribution. If multiple defect populations exist spanning different length scales, we may achieve a `heavy tailed' statistics; that is, no convergence to a Gaussian and the distribution becomes Lorentzian. In reality this will always be a truncated Lorentzian due to the finite sample size, thus sidestepping any mathematical qualms regarding the use of the Lorentzian distribution. The Lorentzian is a scale independent distribution; consequently the number of steps, $m$, drops out of Eq.\ (\ref{eq:lorentzian}). We note in passing that the vortex paths have a fractal character in this case. We have plotted the entropy for the uniform and angular functions in Appendix A. Some further details on the extrema of Eq.~(\ref{eq:gaussian}) are listed in Appendix B. This model naturally allows vortex interactions with multiple defect populations to be taken into account. For interactions with multiple Gaussian populations we have $\Gamma^2 = \Gamma_1^2 + \Gamma_2^2 + \ldots$, while for multiple Lorentzians, $\Gamma = \Gamma_1 + \Gamma_2 + \ldots$. These results are derived in Appendix C. Whether a particular peak is due to a unique population of defects or a compound population is a question of interpretation that needs to be addressed through the experimental evidence. As a method of inference, the principle of maximum entropy further allows us to make the following observations. If we can describe $J_c(\theta)$ using Eqs.~(\ref{eq:VPMequations}) then we have found the correct constraints for the problem and no better description of $J_c(\theta)$ is possible unless we can find weaker constraints and higher entropy functions to fit the data, which is highly unlikely. If $J_c(\theta)$ cannot be described by these equations then further constraints on the microstates exist which have not been taken into account. The model thus provides a direct test of the high-field electron mass anisotropy scaling hypothesis, Eq.~(\ref{eq:massanisotropyscaling}), since the model as presented takes no account of electron mass anisotropy. \section{\label{sec:results}Results and discussion} In this section we fit $J_c(\theta)$ data for a wide variety of defect-engineered samples, both LTS and HTS, to combinations of Eqs.~(\ref{eq:VPMequations}). In the vortex path model each equation corresponds to one set of constraints for one set of microstates. There is no expectation that a single equation enumerates all the possible microstates and so fitting involves choosing what, in the field of probability, is termed a mixture distribution. For example, with both $ab$-plane and $c$-axis correlated defects present, both a so-called ``$ab$-plane'' peak and a ``$c$-axis'' peak in $J_c$ are possible. Thus, the combination of two orthogonal defect populations can lead to a mixture of two (or possibly more) distributions. The equations above have been written for a peak centered on the $ab$-plane; for a $c$-axis peak the $\theta$ coordinate will be modified accordingly. Likewise, although not explicitly considered here, it is perfectly possible to account for inclined defect structures, for example due to off-axis irradiation, vicinal substrates, inclined substrate deposition or ion-beam assisted deposition, within the framework of the vortex path model The fitting has been done ``by eye'' with the assistance of a non-linear least squares fitting algorithm. In some data sets there exist asymmetries or drift which we have not attempted to correct. A formal statistical best fit procedure is possible but such a development is not warranted for the interpretation we are attempting here. More importantly we are attempting to interpret these distributions in terms of the underlying defect structure which contributes to each set of microstates or pinned vortex paths. This interpretation is of necessity tentative due to our limited knowledge of both the microstructure of the samples and the effectiveness of different elements of that microstructure in pinning. \subsection{\label{sec:Nb}Nanostructured Nb films} We begin by presenting data obtained on Nb thin films nanostructured to incorporate finely-tuned arrays of vertical pores. The sample preparation details and $J_c(B)$ data for these samples were published elsewhere \cite{Dinner2011}. The samples have a structure of nano-engineered pores etched vertically through the 36 nm thickness of the film with an average pore diameter of 75 nm and an inter-pore spacing of about 140 nm. The $J_c$ values for different samples are variable, but for the best samples approach the depairing current density $J_d$ at self-field. The incorporation of the nano-pores increases $J_c$ by up to a factor of 50 over the virgin films, so clearly the pores contribute significant pinning. \begin{figure} \includegraphics{Nb} \captionof{figure}{\label{fig:Nb}$J_c(\theta)$ at 5 K for a Nb thin film incorporating an array of vertical columnar pores: experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{purple}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{brown}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:Nb}.} \end{figure} \begin{figure} \captionof{table}{\label{tab:Nb}Parameters of fit components in Fig.~\ref{fig:Nb}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}cdd} & Angular function & & \multicolumn{1}{c}{\textrm{0.1 T}} & \multicolumn{1}{c}{\textrm{0.2 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 1.140 & -\\ \multirow{2}{*}{OOP} & \multirow{2}{*}{Gaussian (\textcolor{purple}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & 0.351\\ & & $\Gamma$ & - & 1.000\\ \multirow{2}{*}{IP} & \multirow{2}{*}{Gaussian (\textcolor{brown}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.015 & 0.013\\ & & $\Gamma$ & 0.043 & 0.030\\ \multirow{2}{*}{IP} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.083 & 0.426\\ & & $\Gamma$ & 0.249 & 0.400\\ \end{tabular} \end{ruledtabular} \end{figure} Measurements of $J_c(\theta)$ on one of these samples are shown in Fig.~\ref{fig:Nb}. We note that this data is qualitatively similar to much of the literature data obtained on YBCO and observe that, following the approach of \cite{CivaleJLTP2004,*CivaleAPL2004}, such data would be interpreted as comprising three distinct angular regimes: (i) a sharp in-plane peak in the region of $\pm90^\circ$, attributed to staircase vortices predominantly following the intrinsic pinning $ab$-planes; (ii) a broader underlying in-plane peak attributed to mass anisotropy effects; and (iii) an enhanced $J_c$ near the out-of-plane direction at $0^\circ$ (in this case manifesting itself as a flat angular dependence over an extended angular range), attributed to correlated out-of-plane pinning. Evidently, in the case of this data on Nb, interpretations (i) and (ii) are manifestly incorrect, and yet the features remain. Applying the vortex path model to this data, we also identify three components. However, our attribution of the origin of these components is somewhat different. In the vortex path model, each component is identified as resulting from the statistical combination of a dominant pinning population, giving correlated pinning aligned with the centre of the peak, and a pinning species providing a pinning force orthogonal to this direction, that broadens the peak. Thus, the two components (brown and green) centered on $\pm90^\circ$ represent two distinct combinations of dominant in-plane pinning with a pinning population orthogonal to this (i.e.\ out-of-plane). Interestingly the two components have similar peak height. We are not certain as to the origin of this effect but it may be due to the frequency of interaction of the vortex path with the orthogonal defects being similar. The two defect populations operate over different length scales (reflected in the widths of the peaks). Since this particular system is extremely simple, clean and well-defined, the only likely sources of correlated pinning are the surface interfaces (the film thickness being comparable to the coherence length) and the patterned vertical pores. We therefore identify the broader peak as arising from the interaction of the surface pinning with large-scale vertical defects comprising the pore structures themselves, while the narrow peak is the result of the interaction of the surface pinning with the (small-scale) surface roughness caused by the presence of the pores. It is then clear why these two interactions should occur with the same frequency. The third component in the low-field data appears in the form of a constant background. This is the uniform function. A better understanding of the origin of this component comes from considering the behavior under a larger applied field, where it is seen that this component narrows into a rather broad angular Gaussian. The angular peak width of this component is similar to that of the broad in-plane peak, and so we attribute this third component to the dominant out-of-plane pinning facilitated by the pores, broadened by the interaction with the surface pinning. The scale factor is somewhat larger because instead of being limited by the length of the pore (i.e.\ the film thickness), it is now determined by the distance the vortex is able to travel in the in-plane direction while maintaining the macroscopic field orientation. Since this pinning mechanism is effective out to angles of about $\pm60^\circ$, this distance is approximately $\tan 60^\circ = \sqrt{3}$ times the film thickness. At the lower field the weaker vortex-vortex interactions remove the statistical constraints on the majority of vortex paths formed between the pores and the surface pinning, thus resulting in the large uniform contribution. It is to be emphasized from these results that the effect of the introduction of perfectly strong, perfectly correlated out-of-plane pinning defects is not necessarily the creation of a strong out-of-plane peak in the critical current. The absence of such a peak, as here, cannot therefore be taken as evidence of the absence of correlated pinning. Here, the existence of such pinning centers has a determining effect on the critical current, but it does not manifest itself as a simple out-of-plane peak. Instead it contributes at all angles, through both the broadening of the in-plane peak and the addition of a broad and almost flat component centered on the out-of-plane direction. This is of critical importance for the interpretation of other data, for example in similar experiments to these recently conducted on YBCO \cite{Bagarinao2012} which also lack an out-of-plane peak in $J_c(\theta)$. Likewise, `shoulders' in the angular $J_c$ dependence at an intermediate angle between the in-plane and out-of-plane directions are a natural result of the interplay of in-plane and out-of-plane correlated pinning, and not a signature of some unorthodox type of pinning at an arbitrary angle. As we shall see, such shoulders are an increasingly prominent feature of pinning-optimized samples, and one which the electron mass anisotropy scaling approach is entirely unable to address. \subsection{\label{sec:PLD}PLD YBCO films} Turning now to HTS materials, we begin with the simplest data available for an epitaxial 300 nm thick YBCO film, without nanostructural modifications, prepared by pulsed laser deposition (PLD) on a single crystal SrTiO$_3$ substrate \cite{Ercolano2011}, and we see that all the features just described are again present (Fig.~\ref{fig:PLD}). There is no need to invoke anything beyond the known microstructure of the sample to explain all of the observed features. In this thicker film, rather than surface pinning, effective in-plane pinning is provided by the intrinsic pinning due to the planar crystal structure of the sample, with non-superconducting planes separated by roughly the coherence length. Significant out-of-plane pinning is provided by the array of well-known structural defects resulting from the columnar growth mode of PLD samples that are the underlying cause of the high $J_c$ in thin films \cite{Wimbush2009}. \begin{figure} \includegraphics{PLD} \captionof{figure}{\label{fig:PLD}$J_c(\theta)$ at 77 K for a PLD YBCO film (data from \cite{Ercolano2011}): experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{brown}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:PLD}.} \end{figure} \begin{figure} \captionof{table}{\label{tab:PLD}Parameters of fit components in Fig.~\ref{fig:PLD}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}cdd} & Angular function & & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{3 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 0.600 & 0.132\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{brown}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.020 & 0.004\\ & & $\Gamma$ & 0.075 & 0.035\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.103 & 0.113\\ & & $\Gamma$ & 0.473 & 0.396\\ \end{tabular} \end{ruledtabular} \end{figure} At 1 T, we observe three contributions to the pinning profile: two $ab$-centered angular Gaussians and a uniform component. This indicates two distinct $c$-axis defect populations, each broadening the $ab$-peak by a different amount. The peak heights of the two Gaussians again coincide, suggesting that the populations of orthogonal defects are correlated with each other as was the case in the Nb films. We hypothesize that the explanation is the same: the population of longer $c$-axis defects (e.g.\ threading dislocations which penetrate the full film thickness) also creates small scale disorder (e.g.\ surface roughness) which vortices lying close to the $ab$-plane orientation see as a defect population available to provide the small amount of $c$-axis pinning required to broaden their peak. If we label the scale factors for these individual populations as $\Gamma_1$ and $\Gamma_2$ then the broad peak is most probably of scale factor $\Gamma = (\Gamma_1^2 + \Gamma_2^2)^{1/2}$, i.e.\ we observe the contribution of pinning paths that interact with both species. Where the two distributions are strongly disparate, as here, the influence of the small scale disorder on the broader peak becomes negligible. Note that the $c$-axis defects present must of necessity also be contributing to the uniform component as this is the only component with a non-zero magnitude at $\theta = 0$. All other defects in the sample also contribute to the uniform component. At 3 T, the peak magnitudes of the narrow $ab$-peak and the uniform component are the same. The broad $ab$-peak is now of greater magnitude than these components. This is the same trend as was observed for Nb. This suggests that the only effective $c$-axis pinning remaining at this field is now provided by the through-thickness dislocations that match the frequency of the defects broadening the narrow $ab$-peak. These defects remain effective in broadening the $ab$-peak as at 1 T. However, the matching field of these defects is such that they are no longer effective pins in their own right at 3 T, and therefore the uniform component is diminished in magnitude. \subsection{\label{sec:MOD}MOD YBCO tapes} The real power of the vortex path model, however, is revealed in its application to technologically-relevant HTS samples. In materials presently being developed for application, pinning effects dominate at all angles, necessitating consideration of the effects of combining multiple different defect populations. It is in these samples that the vortex path model comes into its own, explaining every aspect of the observed data in a coherent, self-consistent manner. \begin{figure} \includegraphics{MOD} \captionof{figure}{\label{fig:MOD}$I_c(\theta)$ at 77 K for MOD YBCO tape: experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{purple}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:MOD}. Where there exists significant asymmetry in the data set, the first half of the data has been fitted.} \end{figure} \begin{figure} \captionof{table}{\label{tab:MOD}Parameters of fit components in Fig.~\ref{fig:MOD}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}cddd} & Angular function & & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{2 T}} & \multicolumn{1}{c}{\textrm{3 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 117.8 & 55.45 & 28.15\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{purple}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.331 & 0.461 & 1.755\\ & & $\Gamma$ & 0.056 & 0.065 & 0.133\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 10.83 & 4.21 & 5.19\\ & & $\Gamma$ & 0.664 & 0.962 & 1.69\\ \end{tabular} \end{ruledtabular} \end{figure} Angular $I_c$ data for a production sample of American Superconductor Amperium\texttrademark tape with a superconducting YBCO layer fabricated by metal-organic deposition (MOD) are shown in Fig.~\ref{fig:MOD}. The major contributions to $I_c(\theta)$ can be fitted with three angular functions at all fields. A well-defined $ab$-plane peak increases in magnitude as the applied field increases. This is paired with an extremely broad $c$-axis peak that broadens further under increasing field, resulting in the characteristic shoulder shape. Note that these shoulders are in approximately the same position as for the LTS film. The gradual broadening of the $c$-axis peak is consistent with the observation that the $ab$-plane pinning is increasing in strength, and will therefore contribute more to the broadening of the $c$-axis peak. Because the fit comprises just two angular Gaussian functions centered on the orthogonal directions, this indicates that there exists a single population of $ab$-plane defects and a single population of $c$-axis defects which are the dominant contributors to the angle-dependent pinning. The two peaks arise from the combination of these same two pinning populations as outlined previously. The sample additionally has a significant uniform component at all fields. We have seen in the Nb data how this can arise from correlated defects. In these films it may also include a substantial contribution from point defects. An assembly of point defects may pin a single vortex or they may act in combination with correlated defects. To emphasize the importance of this interpretation, which differs from the standard view, we note that in the data measured at 1 T and 2 T there is a small angular region around $\pm75^\circ$ where neither of the angular Gaussian functions has any magnitude. It is not credible to suppose that in precisely this region the correlated defects present in the sample provide no contribution to $I_c$, while contributing significantly at all other angles. As evidence to support this view, we have seen in the PLD sample that the $ab$-peak has a width extending over this region, while in this sample at 3 T we can again see directly the contribution of correlated defects in this region. However, we also know that in these pinning-engineered samples, the incorporation of nanoparticles --- non-correlated defects --- has increased the overall $I_c$. Therefore both nanoparticles and correlated defects must be contributing to the uniform component. These samples differ from previously described MOD samples \cite{Long2005} in having much stronger $c$-axis pinning. Presumably this is due to twin planes or twin plane intersections, which are the only known high density $c$-axis defects to occur naturally in MOD samples. Again, the vortex path model enables us to relate the pinning profile observed in $I_c(\theta)$ to microstructural features of the sample that can be investigated experimentally. The fitting obtained for samples on technical substrates using angular Gaussians is seldom as good as for angular Lorentzians, as seen throughout this paper and in \cite{Long2008}. This is because the samples are also disordered in their grain orientation, with an out-of-plane mosaic spread around $5^\circ$. With the Lorentzian components, this actually helps to produce better fits as it generates the same result as having a collection of slightly differing length scales for the same defect population across different grains. As explained in \cite{Long2008}, a Lorentzian is equivalent to a mixture of Gaussians and the disorder helps to create a spread of Gaussians. Conversely, when we have a Gaussian, the grain misorientation is working against a good fit. A Gaussian requires a single scale parameter but we don't have one because defects in differently oriented grains project onto our measurement axes differently. On this basis, it would be appropriate to smooth the fits over an angular range approximating the mosaic spread in the sample. Further counting against a good Gaussian fitting, we will never truly have a single defect population giving all the correlated $J_c(\theta)$ --- there are bound to be minor defect populations also contributing. This is most likely the case in this sample where in the 2 T and 3 T data a second $c$-axis peak can just about be resolved. Nonetheless, even without accounting for these finer details, we obtain a high level of accuracy with very few components. \subsection{\label{sec:NbO}YBCO with Ba$_2$YNbO$_6$ additions} We now turn to defect-engineered samples in the true meaning of the term; that is to say samples with incorporated defects designed to influence the pinning profile in particular ways. The first of these we will consider is the introduction of self-assembled $c$-axis aligned nanocolumns of second-phase material intended to provide strong, effective flux pinning in the $c$ direction where $J_c$ is usually lowest. In the present case, this has been achieved through the addition of the double-perovskite Ba$_2$YNbO$_6$ to PLD YBCO films \cite{Ercolano2010}. This material is incorporated into the YBCO matrix as $\sim$100 nm long, 10-15 nm diameter rods aligned with the $c$-axis of the YBCO and has enabled a two to three times enhancement in $J_c(B{\parallel}c)$ compared with pure YBCO films. The data shown in Fig.~\ref{fig:NbO} are for a 300 nm thick film comprising 5 mol\% Ba$_2$YNbO$_6$, resulting in nanocolumns having an average lateral spacing of around 40 nm. \begin{figure} \includegraphics{NbO} \captionof{figure}{\label{fig:NbO}$J_c(\theta)$ at 77 K for a PLD YBCO film with Ba$_2$YNbO$_6$ additions (data from \cite{Ercolano2011}): experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{purple}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{brown}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:NbO}. Where there exists significant asymmetry in the data set, the first half of the data has been fitted.} \end{figure} \begin{figure} \captionof{table}{\label{tab:NbO}Parameters of fit components in Fig.~\ref{fig:NbO}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}cdd} & Angular function & & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{3 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 1.040 & 0.273\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Lorentzian / Gaussian (\textcolor{purple}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & \textrm{L }0.223 & \textrm{G }0.075\\ & & $\Gamma$ & \textrm{L }1.012 & \textrm{G }0.524\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Lorentzian / Gaussian (\textcolor{brown}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & \textrm{L }0.050 & \textrm{G }0.138\\ & & $\Gamma$ & \textrm{L }0.302 & \textrm{G }0.050\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & 0.125\\ & & $\Gamma$ & - & 0.570\\ \end{tabular} \end{ruledtabular} \end{figure} At low field (1 T), we once again require three components to fit the data: a uniform component and two Lorentzian peaks. The presence of Lorentzian rather than Gaussian peaks may be caused by additional lattice disorder introduced into the YBCO by the nanocolumns due to their large lattice mismatch and consequent straining of the matrix. Significant distortion of the YBCO lattice has been observed in samples of this composition. Additionally, splaying of the nanorods about the $c$-axis means the direction of their pinning is not well-defined, but instead encompasses a distribution of its own. Consequently, instead of a single defect population yielding a single scale parameter there are multiple Gaussians present which have merged to produce a Lorentzian lineshape. In contrast to the pure YBCO films, here the two peaks are not both $ab$-plane peaks, but rather one of them is a strong $c$-axis peak as a result of the highly effective nanocolumnar pins. In fact, at this field, these pins are so effective that the magnitude of the $c$-axis peak is greater than that of the $ab$-plane peak. This in turn explains the absence of the second $ab$-plane peak in this data: the strength of the $c$-axis pinning is such that it is completely dominant in its interaction with the available $ab$-pinning. This results in the strong uniform component ranging across the angles usually occupied by the broad $ab$-plane peak. At higher field (3 T), the nanocolumnar pinning begins to lose its strength, a common phenomenon in samples utilizing present flux pinning strategies. This is because the overall density and therefore matching field is relatively low for this defect population. Clearly the pinning centers are still present, but they lose in dominance to the $ab$-plane pinning, and the broad peak associated with their interaction with this pinning source is recovered. The response at this field is extremely similar to the pure YBCO PLD films, with the addition of a small $c$-axis peak due to the residual effect of the additional engineered $c$-axis pinning. The higher field thus again helps to distinguish the different pinning contributions. The change in peak shape from Lorentzian to Gaussian likely results from vortex-vortex interactions at high field acting to constrain the vortices, diminishing both the effectiveness of the pinning and reducing the effect of splay. We also see the previously noted phenomenon of the two $ab$-plane peaks coinciding in magnitude, and additionally here the uniform component also coincides. The uniform component is about twice the magnitude seen in the pure YBCO, reflecting the continuing contribution of the engineered $c$-axis pinning to the uniform component, not just to the $c$-axis peak. \subsection{\label{sec:TaO}YBCO with Gd$_3$TaO$_7$ additions} In the attempt to progress beyond BaZrO$_3$ as an effective second-phase pinning inclusion for YBCO, and to increase the effectiveness of this type of artificial pinning center, the pyrochlore rare earth tantalate \emph{R}$_3$TaO$_7$ material was successfully introduced, resulting in significant $J_c$ enhancement compared to BaZrO$_3$-containing samples across the angular range \cite{Harrington2009}. Due to its close lattice match to YBCO, in contrast to BaZrO$_3$ and Ba$_2$YNbO$_6$, \emph{R}$_3$TaO$_7$ addition readily produces extremely fine and highly linear through-thickness nanorods without splaying, and with no poisoning effect reducing the critical temperature. By a simple variation in deposition rate, the nanostructure of the \emph{R}$_3$TaO$_7$ additions can be controlled, either freezing the additions into a fine dispersion of nanoparticles of $\sim$5 nm diameter (at high deposition rate), or allowing time for diffusion-mediated growth into extended nanorods of 5-10 nm diameter and 10-15 nm spacing (at low deposition rate). This provides a unique `lever' offering unprecedented control over the nature of the flux pinning in samples incorporating this additive \cite{Harrington2010}. \begin{figure*} \includegraphics{TaO} \captionof{figure}{\label{fig:TaO}$J_c(\theta)$ at 77 K for PLD YBCO films with Gd$_3$TaO$_7$ additions, deposited at different rates to induce nanoparticle or nanorod formation (data from \cite{Harrington2010}): experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{purple}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{pink}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{cyan}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{brown}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:TaO}.} \end{figure*} \begin{figure} \captionof{table}{\label{tab:TaO}Parameters of fit components in Fig.~\ref{fig:TaO}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}ccdddd} & \multirow{2}{*}{Angular function} & & \multicolumn{2}{c}{\textrm{High rate}} & \multicolumn{2}{c}{\textrm{Low rate}}\\ & & & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{3 T}} & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{3 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 2.293 & 0.534 & 1.084 & 0.669\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{purple}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & - & 0.117 & 0.043\\ & & $\Gamma$ & - & - & 0.263 & 0.244\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{pink}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & - & 0.207 & 0.055\\ & & $\Gamma$ & - & - & 0.630 & 1.100\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{cyan}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & - & 0.099 & 0.027\\ & & $\Gamma$ & - & - & 1.444 & 2.816\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{brown}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & 0.050 & 0.011 & 0.003\\ & & $\Gamma$ & - & 0.154 & 0.045 & 0.030\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.253 & 0.128 & - & -\\ & & $\Gamma$ & 0.384 & 0.465 & - & -\\ \end{tabular} \end{ruledtabular} \end{figure} The data shown in Fig.~\ref{fig:TaO} for 500 nm thick YBCO films prepared by PLD with 1.5 mol\% Gd$_3$TaO$_7$ additions exemplify the pinning control that can be achieved with this additive, where the anisotropy is seen to vary from broad $ab$-peak in the case of a high-rate deposition through to strong $c$-axis peak for low-rate deposition. Supported by TEM imaging, the simple additive model of pinning used in \cite{Harrington2010} to describe this data provides an adequate explanation of the primary features, that is to say the existence of a broad, flat profile with residual $ab$-peaks in the case of the isotropic nanoparticle pinning produced at high growth rate giving way to a strong $c$-axis peak in the presence of the nanocolumns formed at low growth rate. However, application of the vortex path model to these data allows us to decompose the behavior further. In the case of the sample deposited at high rate, we see the simplest behavior observed to date --- just two components: one uniform and one broad $ab$-peak. Similar to the pure PLD YBCO film, we attribute the broad $ab$-peak to dominant intrinsic $ab$-plane pinning broadened by the type of through-thickness defects characteristic of PLD. The effect of the nanoparticle inclusions is to strongly enhance the uniform component with the result that the sharp $ab$-peak previously observed in films of this type is obscured. This is an example of similar-strength pinning centers adding non-linearly to generate the total pinning force, and evidences the failure of a simple additive pinning model. As the applied field is increased, the effectiveness of the nanoparticulate pinning centers diminishes, leading to a reduction in the uniform component, and the sharp $ab$-plane peak is again distinguished. Conversely, in the case of the sample deposited at low rate, the most prominent feature is the $c$-axis peak resulting from the nanorods. The vortex path model represents this as three angular Gaussians of increasing width, all centered on the $c$-axis. Here, due to the extreme linearity and crystalline perfection of the nanorods, no tendency towards a Lorentzian lineshape is observed, and distinct broadening defect populations can be resolved. Three components of differing width imply three distinct populations of broadening defects available to interact with the nanorods. Based on the limited microstructural data available on these samples, we speculate that, in common with the MOD samples, strong $ab$-plane pinning is one of the broadening populations, while another broadening defect may be specific $ab$-plane defects such as stacking faults. These tentative assignments can be verified or refuted through experiments designed to control the density of specific defects within the sample and correlate this with $J_c(\theta)$. As the field is increased, these three components become better resolved, broadening and developing the characteristic shoulders also seen in the MOD samples as the intrinsic $ab$-plane pinning becomes relatively more effective than the artificial $c$-axis pinning. The advantage of the vortex path model in analyzing this data is that we can see how the basic components of correlated and uncorrelated defects are combining to produce the overall $J_c(\theta)$ behavior. That particular peaks are narrow or broad does not mean that whole pinning populations are suddenly appearing or disappearing but rather that the relative strengths and the statistics of how these populations combine is being altered. \subsection{\label{sec:NbOTaO}YBCO with Ba$_2$YNbO$_6$ + Gd$_3$TaO$_7$ additions} Recently, the importance to pinning optimization of achieving segmented nanocolumnar pins rather than continuous nanocolumns was suggested \cite{Chen2009} to mitigate the effects of thermally activated depinning. One method by which this has been achieved is through the simultaneous incorporation of the two types of pinning additive already considered, Ba$_2$YNbO$_6$ and Gd$_3$TaO$_7$. What forms in this case is a complex composite of YBCO, Ba$_2$Y(Nb,Ta)O$_6$ segmented nanorods, in-plane Y$_2$O$_3$ platelets and nanoparticles of the YBa$_2$Cu$_4$O$_8$ superconducting phase, where the Gd partially substitutes for Y in all four phases and the Ta and Nb content goes entirely into forming nanorods \cite{Ercolano2011}. The $J_c(\theta)$ dependence as a function of the applied field for a 300 nm thick sample of this composition prepared by PLD is shown in Fig.~\ref{fig:NbOTaO}. \begin{figure*} \includegraphics{NbOTaO} \captionof{figure}{\label{fig:NbOTaO}$J_c(\theta)$ at 77 K for a PLD YBCO film with Ba$_2$YNbO$_6$ + Gd$_3$TaO$_7$ additions (data from \cite{Ercolano2011}; 2 T data unpublished data from the same sample): experiment ($\bullet$), full fit (\textcolor{red}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}), fit components (\textcolor{blue}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{purple}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{pink}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{cyan}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{orange}{\protect\rule[0.5ex]{0.4cm}{1.5pt}},\textcolor{green}{\protect\rule[0.5ex]{0.4cm}{1.5pt}}). The fitting parameters for these data are summarized in Table \ref{tab:NbOTaO}. The small $ab$-peak has been ignored in the fitting.} \end{figure*} \begin{figure*} \captionof{table}{\label{tab:NbOTaO}Parameters of fit components in Fig.~\ref{fig:NbOTaO}.} \begin{ruledtabular} \begin{tabular}{r@{\extracolsep{1ex}}l@{\extracolsep{\fill}}cdddddd} & Angular function & & \multicolumn{1}{c}{\textrm{1 T}} & \multicolumn{1}{c}{\textrm{2 T}} & \multicolumn{1}{c}{\textrm{3 T}} & \multicolumn{1}{c}{\textrm{4 T}} & \multicolumn{1}{c}{\textrm{5 T}} & \multicolumn{1}{c}{\textrm{6 T}}\\ \colrule & Uniform (\textcolor{blue}{\rule[0.5ex]{0.4cm}{1.5pt}}) & $J_0$ & 1.166 & 0.829 & 0.801 & 0.531 & 0.311 & 0.148\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{purple}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.125 & 0.105 & 0.070 & 0.033 & 0.035 & 0.025\\ & & $\Gamma$ & 0.091 & 0.113 & 0.176 & 0.153 & 0.151 & 0.152\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{pink}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.348 & 0.384 & 0.356 & 0.256 & 0.185 & 0.118\\ & & $\Gamma$ & 0.257 & 0.371 & 0.722 & 1.21 & 1.43 & 1.55\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{cyan}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.404 & 0.368 & 0.163 & 0.045 & 0.023 & 0.015\\ & & $\Gamma$ & 0.672 & 0.831 & 1.49 & 3.56 & 4.00 & 3.64\\ \multirow{2}{*}{$c$} & \multirow{2}{*}{Gaussian (\textcolor{orange}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & 0.218 & 0.140 & 0.055 & - & - & -\\ & & $\Gamma$ & 1.63 & 2.38 & 3.50 & - & - & -\\ \multirow{2}{*}{$ab$} & \multirow{2}{*}{Gaussian (\textcolor{green}{\rule[0.5ex]{0.4cm}{1.5pt}})} & $J_0$ & - & - & - & 0.070 & 0.118 & 0.145\\ & & $\Gamma$ & - & - & - & 0.491 & 0.422 & 0.400\\ \end{tabular} \end{ruledtabular} \end{figure*} From the low field data, the most striking feature is the large sharp $c$-axis peak, far stronger and better defined than we have seen in any other data. Analysis in terms of the vortex path model begins to give us some idea of why this is --- the coincidence of no less than four components centered on the $c$ direction combine to give this peak. Its sharpness is caused by the combination of two narrow components of similar magnitude, while its baseline is boosted by the contribution of two additional broad components, as well as a large uniform component. The effect of this large uniform component is to make the $ab$-peak look small, but once again this is not due to any particular lack of $ab$-pinning in these samples, but rather the strong relative dominance of the $c$-axis pinning. In the same way that strong $ab$-pinning can obscure a $c$-axis peak despite the existence of significant $c$-axis pinning, uncommonly strong $c$-axis pinning as has been achieved here can also obscure the $ab$-peak. The uniform component is similar in magnitude to the nanorod-containing samples of the previous section. By analogy with the earlier niobate addition data, we identify the broader of the two sharp $c$-axis peaks as resulting from the segmented nanorods, in interaction with the intrinsic $ab$-plane pinning, observing that it varies with field in the same way, diminishing and broadening out to a relatively flat response by 3 T. This leaves the sharper $c$-axis peak to be identified, and we note that it tracks the broader peak in magnitude. It also remains relatively constant throughout the entire field range, changing little in width or relative magnitude. This suggests two things: firstly that it has a high matching field, i.e.\ it results from a high-density field-independent population of defects, and secondly that its width probably arises from a property such as the angular spread of the nanorods. Hence this is an unusual peak in that it may arise from a single defect species only, interacting with itself through its own disorder. The broader $c$-axis components then likely arise due to interactions between the segmented nanorods and the in-plane Y$_2$O$_3$ platelets and YBa$_2$Cu$_4$O$_8$ nanoparticles. These unique features are a direct result of the unique microstructure of these samples, with the Y$_2$O$_3$ platelets being of similar dimension to the nanorod segments. It is therefore unsurprising that they greatly extend the angular range over which the nanorods can effectively pin, facilitating pinning of vortices between adjacent nanorod segments at high field angles. These four $c$-axis components can be traced through to 3 T, by which point the $c$-axis pinning is significantly reduced in strength, as is commonly observed for all artificial pinning centers known to date. The two broad $c$-axis components continue broadening strongly with increasing applied field. To higher fields, the dominance of the $c$-axis pinning is overcome, and familiar features relating to the $ab$-pinning begin to emerge. The resultant combination of these similar strength features yields extremely complex angular variations in $J_c$ that are nonetheless resolvable into a few primitive components that describe the subtlest features of the data extremely well. At 4 T, an interesting crossover occurs. It is at this field that the strongest nanorod component is broadened to such an extent that it has developed significant shoulders that result in a quite notable response in the $J_c(\theta)$. It is appropriate that such significant microstructural features should continue to give a marked response in the $J_c(\theta)$ even though their contribution to the overall $J_c$ is now diminished. At the same time, the broadest $c$-axis peak vanishes, to be replaced with the usual $ab$-pinning peak that we attribute to intrinsic pinning being broadened by through-thickness defects. The disappearance is assumed to be a consequence of the applied field becoming greater than the matching field of the relevant $c$-axis defect species. We note a couple of interesting points about the fitting. Again we observe coincidences in the peak heights of multiple components. These are seen to occur not only at the center of the distribution but also at other points where the distributions reach a local extremum. At an extremum $\partial g/\partial \theta = 0$ and the distribution is locally unconstrained. This coincidence may therefore have an interpretation in terms of the maximization of entropy for a mixture distribution which we do not presently understand. The other point we make is to remind the reader that each component has only a single scale parameter. This means that both the shape and the position of the shoulders are highly constrained by the form of the angular Gaussian distribution we have derived. It is difficult to conceive of other primitive functions which could so well reproduce these features of this data. \section{\label{sec:conclusion}Conclusion} We have addressed the question of whether the application of the electron mass anisotropy scaling approach to field angular $J_c(\theta)$ data of pinning-optimized samples is theoretically justified, and have concluded that it is not. In experimental support of this determination, we have presented data measured on thin films of the isotropic superconductor Nb that exhibit features commonly attributed to electron mass anisotropy, reinforcing the fact that $J_c$ is a wholly extrinsic material parameter and demanding an alternative explanation that can be related to the microstructure or other extrinsic properties of the sample. We provide this explanation in terms of the vortex path model, and have applied the same model to a wide range of contemporary pinning-engineered samples, from which a number of general results emerge. Firstly, that the introduction of perfectly strong, perfectly correlated out-of-plane pinning defects does not necessarily result in a strong out-of-plane peak in the $J_c(\theta)$, and therefore that the absence of such a peak cannot be taken as evidence of the absence of correlated pinning. Secondly, that pinning from different sources is not a simple additive process, but rather depends on what pinning is already present. In this context, common manifestations of strong out-of-plane correlated pinning are a broadening of the in-plane peak or an almost flat angular variation in $J_c(\theta)$ about the out-of-plane direction, while commonly-observed shoulders in the angular $J_c$ dependence are a result of the interplay between in-plane and out-of-plane correlated pinning of similar strength, and not a signature of exotic pinning at an arbitrary angle. In particular we emphasize that a broad in-plane peak is evidence of strong out-of-plane correlated pinning, and not a reflection of electron mass anisotropy. Our assignment of defect populations to particular peaks is based on the available knowledge of the sample microstructure. Correct identification of the individual pinning populations requires a correlation between $J_c(\theta)$ and microstructural analysis over a family of samples in which the density of these defect populations is systematically varied and the effect on the different components of $J_c$ quantified. These hypotheses can then be tested by measuring samples in those temperature and field ranges which best distinguish between the different defect populations. It appears from our analysis of available data that measuring the field dependence of the angular profile is very useful for determining the origin of pinning effects. The evolution of the magnitude of a component allows us to determine the matching field for a contributing defect population. The evolution of the shape also correlates with matching field effects, e.g.\ out-of-plane components broaden towards the in-plane direction as the field increases due to the high matching field of the intrinsic pinning. In the context of the principle of maximum entropy, the success of the vortex path model in describing the data shows that no further constraints beyond those we have imposed on the vortex path statistics are necessary to explain $J_c(\theta)$. Thus we can conclude with confidence that the electronic mass anisotropy plays no role in determining the form of $J_c(\theta)$ for any of the various samples studied. If we find samples which do not immediately fit the vortex path model equations then different constraints are operating in those systems. An obvious example is the need to extend the model to cases where $J_c(\theta)$ has asymmetries due to oblique correlated defect populations. The success of the vortex path model in this instance also suggests that other angular data influenced by the pinning landscape could be better described by Eqs.~(\ref{eq:VPMequations}), or at least other expressions derived from maximum entropy considerations. The Ginzburg-Landau mass anisotropy expression is often applied simply because no alternative angular expressions have been proposed. For example, we can see no reason why the angular dependence of the irreversibility field or the flux flow resistivity should follow an expression involving electronic mass anisotropy. \begin{acknowledgments} The authors are grateful to R.\ B.\ Dinner for supplying the Nb thin film samples, and to G.\ Ercolano and S.\ A.\ Harrington for providing digital data sets of their measurements, including some previously unpublished data. We acknowledge helpful discussions with J.\ H\"{a}nisch on the application of the electron mass anisotropy scaling procedure. This work was funded by the New Zealand Ministry of Science and Innovation. \end{acknowledgments}
\section{Introduction} Polarizabilities represent fundamental properties of hadrons, encoding their linear response to externally applied fields. Experimentally, they manifest themselves, e.g., in the non-Born part of the low-energy Compton scattering amplitude. While the leading low-energy response is controlled by the static polarizabilities found in the presence of constant external fields, at subsequent orders of a derivative expansion, the effective hadron Hamiltonian becomes sensitive to temporal and spatial structures in the applied fields. The present study focuses, specifically, on weak, spatially constant external electric fields $\vec{E} $ applied to a neutron. When the fields are sufficiently weak, nonlinear effects, i.e., terms in the effective neutron Hamiltonian of higher order than quadratic in the external field, can be neglected, and, to zeroth order in spatial derivatives as well as second order in temporal derivatives of $\vec{E} $, the neutron mass shift can be expanded as \begin{equation} \Delta m = m(\vec{E} )-m(0) = -\frac{1}{2} \alpha_{E} \vec{E}^2 \ -\frac{1}{2} \gamma_{E1} \vec{\sigma } \cdot (\vec{E} \times \dot{\vec{E} } ) \ -\frac{1}{2} \alpha_{E\nu } \dot{\vec{E} }^2 \ + \ldots \label{heff} \end{equation} The static electric polarizability $\alpha_{E} $ has been investigated in a number of lattice studies, cf., e.g., \cite{wilepap,shintani,elpol,polprog,alexan,detlat}. Here, the goal is to perform a first evaluation of the electric spin polarizability $\gamma_{E1} $ within lattice QCD. An effort in this direction is particularly timely in view of an ongoing experimental program at HI$\gamma $S to measure this polarizability in isolation for the first time, albeit in a proton; previously, only the forward and backward spin polarizabilities $\gamma_{0} $, $\gamma_{\pi } $ of the proton \cite{spexp1,spexp2}, which contain $\gamma_{E1} $ in combination with other polarizabilities, had been accessed experimentally. \section{External field and four-point function scheme} Consider a spatially constant external electromagnetic gauge field $A_i $ generating an electric field $E_i = \partial_0 A_i $. It enters the lattice link variables $U_i $ as an additional phase, $U_i \rightarrow \exp (iaq A_i ) \cdot U_i $, where $a$ denotes the lattice spacing and $q$ the quark electric charge matrix. To evaluate only contributions to the neutron mass shift quadratic in $\vec{E}$, it is sufficient to expand the relevant neutron two-point function in powers of $\vec{E}$ from the outset. Thus, one can perform a fully dynamical calculation using existing unperturbed gauge ensembles: Expanding $\exp (iaq A_i ) = 1+iaq A_i -(a^2 q^2 /2) A_i^2 +\ldots $ and inserting into the lattice action decomposes the latter into an unperturbed part and an external field-dependent part, $S=S_0 + S_{ext} $, where $S_{ext} $ is essentially a standard $j\cdot A$ coupling of the external field to the vector current\footnote{Additional contact terms proportional to $A_i^2 $, cf.~\cite{elpol,polprog}, can be excluded in the present context, as seen below.}. As a consequence, the neutron two-point function expands as \begin{equation} \left\langle N_{\beta } (y) \bar{N}_{\alpha } (x) \right\rangle = \begin{array}{c} \int [DU] [D\psi ] [D\bar{\psi } ] \ \exp (-S_0 ) \ \left( 1 - S_{ext} + S_{ext}^{2} /2 + \ldots \right) N_{\beta } (y) \bar{N}_{\alpha } (x) \\ \hline \int [DU] [D\psi ] [D\bar{\psi } ] \ \exp (-S_0 ) \ \left( 1 - S_{ext} + S_{ext}^{2} /2 + \ldots \right) \end{array} \end{equation} Performing the quark integrations yields the diagrammatic representation for the contributions to the neutron two-point function quadratic in $\vec{A} $ depicted in Fig.~\ref{diagrams}. \begin{figure}[h] \vspace{-0.5cm} \hspace{1.1cm} \epsfig{file=diagrams2.ps,width=13cm} \vspace{-0.9cm} \caption{Diagrams contributing to the neutron mass shift; crosses denote $j\cdot A$ insertions.} \label{diagrams} \vspace{-0.15cm} \end{figure} In the present study, only the connected contributions in Fig.~\ref{diagrams} were evaluated. The four-point function scheme at which one has thus arrived is particularly suited to isolate the spin polarizability $\gamma_{E1} $ from the other contributions appearing in (\ref{heff}): As will be specified in more detail below, each vertex in Fig.~\ref{diagrams} is associated with an external field insertion $\vec{A} = A_1 \vec{e}_{1} + A_2 \vec{e}_{2} $; however, in view of (\ref{heff}), only contributions $\sim A_1 A_2 $ \linebreak are of \pagebreak interest, which can be isolated by populating one vertex with only $A_1 $ and the other vertex with only $A_2 $ in each diagram (and, thus, any contact terms proportional to $A_i^2 $ can also be disregarded, as noted above). This not only discards the signals from the contributions controlled by $\alpha_{E} $ and $\alpha_{E\nu } $ in (\ref{heff}), but also their statistical noise\footnote{Note that evaluating the difference between the mass shifts obtained with neutrons polarized in two opposing directions does not achieve this; that eliminates the signal, but not the noise associated with the $\alpha_{E} $ and $\alpha_{E\nu } $ contributions.}. The very specific control over the different contributions afforded by the present scheme thus aids in reducing the numerical uncertainties. Choosing the neutron state to be polarized in the 3-direction, a suitable external gauge field is \begin{equation} \vec{A} = ( a_1^0 + a_1^1 t ,\ a_2^0 + a_2^1 t + a_2^2 t^2 ,\ 0) \ , \label{gena} \end{equation} which generates a constant $E_1 $ and an $E_2 $ varying linearly with time, as called for by (\ref{heff}). An ambiguity remains in the choice of the constants $a_1^0 $, $a_2^0 $ and $a_2^1 $; fixing $E_1 $ and $\dot{E}_2 $ only determines $a_1^1 $ and $a_2^2 $. In particular, the finite spatial lattice torus is distinct from an infinite spatial domain in that the constant potentials $a_i^0 $ cannot be wholly eliminated using gauge transformations. Due to the torus boundary conditions, only the residual discrete invariance $A_i \longrightarrow A_i +2\pi /q_{min} L$ remains, where $L$ is the spatial extent of the torus and $q_{min}$ the smallest unit of electrical charge in the theory. Fractionally different $a_i^0 $ correspond to different physics; namely, they represent flavor-dependent Bloch momenta in view of the minimal substitution $p_i \rightarrow (p_i -q a_i^0 )$. Thus, despite the dependences on the constants $a_1^0 $, $a_2^0 $ and $a_2^1 $ being partially related to one another via space-time symmetries, the neutron mass shift is influenced by additional physics distinct from the spin polarizability effect, \begin{equation} \Delta m = c_1 \vec{\sigma } \cdot (\vec{A} \times \vec{E} ) \ -\frac{1}{2} \gamma_{E1} \vec{\sigma } \cdot (\vec{E} \times \dot{\vec{E} } ) \label{haeff} \end{equation} even after having restricted the calculation to contributions $\sim A_1 A_2 $. The additional Bloch momentum dependence\footnote{\label{foot3}Note that a separate dependence of the form $\vec{A} \times \dot{\vec{E} } $ can be excluded as follows: Working in Euclidean space, each time derivative acquires a factor $i$ compared to Minkowski space. The terms in (\ref{haeff}) both contain an odd number of time derivatives, and $\Delta m $ will correspondingly be extracted from the imaginary part of the neutron two-point function. By contrast, any putative physical $\vec{A} \times \dot{\vec{E} } $ dependence cannot (and will explicitly be seen not to) arise in the imaginary part.} must be disentangled from the polarizability effects in order to extract $\gamma_{E1} $. To this end, several choices of external fields of the type (\ref{gena}) will be considered and contrasted below. \section{Extracting the neutron mass shift in the adiabatic approximation} In the presence of external fields of the form (\ref{gena}), the Hamiltonian of the system under consideration is time-dependent. One facet of this which deserves to be noted is the connection to the Bloch momenta discussed above. One can, e.g., write $A_1 $ in (\ref{gena}) in two equivalent ways: $A_1 = a_1^0 + a_1^1 t \equiv a_1^1 (t-t_0 )$, i.e., a translation of the system by the time $t_0 $ corresponds to the introduction of a Bloch momentum $a_1^0 = -a_1^1 t_0 $. This, of course, simply expresses the fact that the electric field $E_1 $ accelerates or decelerates the Bloch currents. Correspondingly, the Bloch momentum dependence of the neutron mass shift \pagebreak will manifest itself in a characteristic time dependence of the neutron two-point function, as will be seen below. In general, the time evolution operator for a time-dependent Hamiltonian acquires an intricate time dependence. Here, an adiabatic approximation will be adopted to interpret the measured behavior of the neutron two-point function: Since the external field can be taken to be arbitrarily weak, it seems plausible to consider the case that the strong dynamics equilibrate the system rapidly on the scale of the temporal variation of the external field. In this situation, the behavior of the two-point function (now projected onto zero momentum states polarized in the 3-direction) is, at times $t$ sufficiently large for excited states to have decayed, \begin{equation} G_{\uparrow} (p=0,t) = \sum_{\vec{y} } \ \mbox{Tr} \left( \frac{1+\gamma_{4} }{2} (1-i\gamma_{3} \gamma_{5} ) \left\langle N (y) \bar{N} (x) \right\rangle \right) \ \ = \ \ W \ \exp \left( -\int dt^{\prime } \, m(t^{\prime } ) \right) \end{equation} Inserting expansions of the normalization and the mass\footnote{Note that the omission of a linear term $m^{(1)} $ in the ansatz (\ref{massexp}) implies that the absence of a permanent net electric current or electric dipole moment in an unperturbed neutron has already been put in.} in the external fields (where the superscript specifies the order in the external field), \begin{eqnarray} W &=& W_0 + W^{(1)} [\vec{A} (t)] + W^{(2)} [\vec{A} (t)] +\ldots \\ m &=& m_0 + m^{(2)} [\vec{A} (t)] +\ldots \label{massexp} \end{eqnarray} one obtains specifically for the contribution quadratic in the external field \begin{equation} G^{(2)}_{\uparrow} (p=0,t) \ \ = \ \ W_0 \ \exp (-m_0 \, t) \left( \frac{W^{(2)} [\vec{A} (t)]}{W_0 } -\int dt^{\prime } \, m^{(2)} [\vec{A} (t^{\prime } )] \right) \label{corr2} \end{equation} In view of this, the mass shift $m^{(2)} [\vec{A} (t)]$ can be obtained, after dividing out the unperturbed correlator $G^{(0)} (p=0,t) = W_0 \ \exp (-m_0 \, t)$, from the temporal slope of the ratio {\em only at a stationary point} of the time evolution, where the time dependences of $W^{(2)} [\vec{A} (t)]$ and $m^{(2)} [\vec{A} (t)]$ are relegated $\mbox{\hspace{1.5cm} } $ \vspace{-0.2cm} \begin{table}[h] \hspace{0.05cm} \begin{tabular}{|c|c|c|c|} \hline $am_l $ & $am_s $ & $ m_{\pi } $ & \# configs \\ \hline \hline 0.01 & 0.05 & 357 MeV & 448 \\ \hline 0.05 & 0.05 & 759 MeV & 425 \\ \hline \end{tabular} \caption{$N_f =2+1$, $20^3 \times 64 $ MILC asqtad \hspace{8.6cm} $\mbox{\ \ } $ ensembles with $a=0.124\, \mbox{fm} $ used in the \hspace{8.43cm} $\mbox{\ \ } $ present investigation.} \label{tabmilc} \end{table} \vspace{-4cm} \hspace{6.3cm} \parbox{8.02cm}{to quadratic order in $t$. Physically, this occurs when the quark Bloch currents cease to flow and turn around into the opposite direction due to the forcing by the external electric field. Only at such a stationary point do the strong dynamics have the opportunity to form a bona fide neutron, the polarizability of which can thus be extracted only in the vicinity \linebreak of that particular point in time.} \section{Numerical results} The neutron two-point function was evaluated using the MILC ensembles listed in Table~\ref{tabmilc}. Each configuration was HYP-smeared and chopped into two $20^3 \times 32 $ sublattices with temporal Dirichlet boundary conditions. Domain wall valence quarks were employed. To improve statistics, averages over neutron spin in the positive and negative 3-directions were taken, as well as over the external field (\ref{gena}) and its rotation by $\pi /4$ around the 3-axis. Figs.~\ref{fig759ab}-\ref{fig357c} display results for the ratio $R_2 (t) = G_{\uparrow }^{(2)} (p=0,t) / G^{(0)} (p=0,t)$ for several cases of external field. Consider first Figs.~\ref{fig759ab} and \ref{fig759c}, which pertain to the case of pion mass $m_{\pi } = 759$ MeV. To exhibit the different effects at play, Fig.~\ref{fig759ab}~(left) shows the result of merely using an \pagebreak external field of the form $\vec{A} = (a_1 ,\ e_2 (t-t_0 ) ,\ 0)$, which is expected to yield a time-independent effective dynamics isolating the $\vec{\sigma } \cdot (\vec{A} \times \vec{E} )$ dependence in (\ref{haeff}). Indeed, the correlator ratio $R_2 (t)$ exhibits linear behavior, corresponding to a constant neutron mass shift determined by the slope of $R_2 (t)$, cf.~(\ref{corr2}). \begin{figure}[t] \vspace{-0.1cm} \epsfig{file=data05.01.ps,width=4.9cm,angle=-90} \hspace{0.1cm} \epsfig{file=data05.02.ps,width=4.9cm,angle=-90} \caption{Neutron two-point function ratio $R_2 $ obtained in the $m_{\pi } =759\, \mbox{MeV} $ ensemble, in the presence of the external fields $\vec{A} = (a_1 ,\ e_2 (t-t_0 ) ,\ 0)$ (left) and $\vec{A} = (a_1 ,\ \dot{e}_2 (t-t_0 )^2 /2 ,\ 0)$ (right), where $t_0 =6a$. The neutron source is located at $t=0$, electromagnetic fields are in Gaussian units. Linear and quadratic fits, respectively, were performed using the fit range $t/a \in [4,8]$.} \label{fig759ab} \end{figure} Continuing with the more complex external field $\vec{A} = (a_1 ,\ \dot{e}_2 (t-t_0 )^2 /2 ,\ 0)$, the expected behavior of the correlator ratio $R_2 (t)$ is parabolic in the vicinity of the stationary point $t=t_0 $, since now the electric field entering the $\vec{\sigma } \cdot (\vec{A} \times \vec{E} )$ dependence of (\ref{haeff}) is linear in time, crossing zero at $t=t_0 $. On the other hand, an additional mass shift proportional to $\vec{\sigma } \cdot (\vec{A} \times \dot{\vec{E} } )$ would manifest itself as an additional linear term in $R_2 (t)$, with the effect of shifting the parabola away from $t=t_0 $. Fig.~\ref{fig759ab}~(right), depicting the corresponding numerical result, corroborates the expected behavior; there is no shift of the parabolic time dependence away from $t=t_0 $, and therefore no additional dependence on the combination $\vec{\sigma } \cdot (\vec{A} \times \dot{\vec{E} } )$, as expected on general grounds, cf.~footnote~\ref{foot3} further above. One can moreover verify that the quadratic coefficient of the parabola in Fig.~\ref{fig759ab}~(right) is compatible with the slope in Fig.~\ref{fig759ab}~(left) within numerical error, consistent with the above interpretation. \begin{figure}[h] \vspace{0.15cm} \hspace{0.05cm} \epsfig{file=data05.12.ps,width=4.9cm,angle=-90} \caption{Neutron two-point function ratio $R_2 $ ob- \hspace{7.53cm} $\mbox{\ \ } $ tained in the $m_{\pi } =759\, \mbox{MeV} $ ensemble, given the ex- \hspace{7.62cm} $\mbox{\ \ } $ ternal field $\vec{A} = (e_1 (t-t_0 ) ,\ \dot{e}_2 (t-t_0 )^2 /2 ,\ 0)$, where \hspace{7.58cm} $\mbox{\ \ } $ $t_0 =6a$. The neutron source is located at $t=0$, and \hspace{7.58cm} $\mbox{\ \ } $ electromagnetic fields are in Gaussian units. Cubic \hspace{7.54cm} $\mbox{\ \ } $ fit was performed using the fit range $t/a \in [4,8]$.} \label{fig759c} \vspace{-8.67cm} \end{figure} \hspace{7.3cm} \parbox{7.02cm}{Finally, employing the full external field $\vec{A} = (e_1 (t-t_0 ) ,\ \dot{e}_2 (t-t_0 )^2 /2 ,\ 0)$, one would expect to see cubic behavior of $R_2 (t)$ around the stationary point $t=t_0 $, since now the time dependences of $A$ and $E$ in $\vec{\sigma } \cdot (\vec{A} \times \vec{E} )$ combine to render the latter proportional to $(t-t_0 )^2 $. If this were the only contribution to the neutron mass shift, the slope of $R_2 (t)$ at the inflection point $t=t_0 $ would vanish; on the other hand, an additional contribution to the mass shift proportional to $\vec{\sigma } \cdot (\vec{E} \times \dot{\vec{E} } )$ would manifest itself as an additional linear term in $R_2 (t)$, the slope of which could therefore be read off at the inflection point of the overall cubic behavior. This slope, of course, determines the electric \hspace{-0.039cm} spin \hspace{-0.039cm} polarizability \hspace{-0.039cm} $\gamma_{E1} $ \hspace{-0.039cm} of \hspace{-0.039cm} the \hspace{-0.039cm} neutron} which is the ultimate objective of the analysis. Indeed, the corresponding numerical result for $R_2 (t)$ shown in Fig.~\ref{fig759c} exhibits the expected behavior, including a nonvanishing slope at the stationary point $t=t_0 $. From this slope, one extracts the value $\gamma_{E1} = 0.0057(8) \cdot 10^{-4} \, \mbox{fm}^{4} $, from connected contributions only, at the pion mass $m_{\pi } = 759$ MeV. \begin{figure}[t] \vspace{-0.2cm} \epsfig{file=data01.01.ps,width=4.9cm,angle=-90} \hspace{0.1cm} \epsfig{file=data01.02.ps,width=4.9cm,angle=-90} \caption{Correlator ratios $R_2 $ analogous to Fig.~2, except in the $m_{\pi } =357\, \mbox{MeV} $ ensemble, with fits performed using the fit range $t/a \in [3,5]$.} \label{fig357ab} \vspace{-0.1cm} \end{figure} Turning to Figs.~\ref{fig357ab} and \ref{fig357c}, which present analogous results for the lighter pion mass $m_{\pi } = 357$ MeV, the principal difference lies in the substantially larger statistical uncertainties, which preclude a clear identification of the functional form of the neutron correlator beyond the time $t=5a$ in the plots. As a result, fits in a symmetric time range around the stationary point $t=6a$, as employed in the $m_{\pi } = 759$ MeV case, are not viable at the present level of statistics. Instead, pending improvement of the numerical accuracy, a fit in the time range $t/a\in [3,5]$ was used to identify the preliminary value $\gamma_{E1} = 0.031(21) \cdot 10^{-4} \, \mbox{fm}^{4} $, from connected contributions only, at the pion mass $m_{\pi } = 357$ MeV. Nevertheless, the data depicted in Figs.~\ref{fig357ab} and \ref{fig357c} display behavior entirely analogous to the one seen in the heavier pion mass case for small times, where the numerical fluctuations are under control. Finally, Fig.~\ref{figeft} relates the two data points for the spin polarizability $\gamma_{E1} $ extracted above to partially quenched chiral perturbation theory \cite{detcsb}, adjusted to match the present connected diagram calculation. The corresponding expression depends on a number of low energy constants, all but one of which are comparatively well determined from either experiment or previous lattice calcula- $\mbox{\hspace{0.5cm}}$ \begin{figure}[h] \vspace{-0.54cm} \hspace{0.05cm} \epsfig{file=data01.12.ps,width=4.9cm,angle=-90} \caption{Correlator ratio $R_2 $ analogous to Fig.~3, \hspace{7.78cm} $\mbox{\ \ } $ except\, in\, the\, $m_{\pi } =357\, \mbox{MeV} $\, ensemble,\, with\, fit \hspace{7.8cm} $\mbox{\ \ } $ performed using the fit range $t/a \in [3,5]$.} \label{fig357c} \vspace{-6.94cm} \end{figure} \hspace{7.3cm} \parbox{7.02cm}{tions on the same ensembles. The lone poorly determined parameter, $g_1 $, has been adjusted such as to generate the two curves in Fig.~\ref{figeft} which pass through the upper and lower ends of the error bar of the $\gamma_{E1} $ measurement at $m_{\pi } = 357$ MeV. This yields a rather stringent estimate of $g_1 = -0.15\,\,^{+0.01+0.04}_{-0.01-0.10}\,$, where the first uncertainty quantifies the spread depicted in Fig.~\ref{figeft}, and the second one was obtained by varying the other low energy constants within reasonable bounds. Note that errors due to the truncation of the chiral expansion were not estimated.} \newpage \section{Summary} The present investigation provides first lattice QCD results for the electric spin polarizability $\gamma_{E1} $ of the neutron, albeit including connected diagrams only. The extraction of this quantity was facilitated by the adopted four-point function approach, which allows for the exact elimination of other contributions to the neutron mass shift in the external electric field considered, namely, ones associated with the static polarizability $\alpha_{E} $ and the dispersion polarizability $\alpha_{E\nu} $. In addition, the effects of constant external gauge fields, corresponding to quark Bloch momenta on a finite spatial torus, had to be carefully disentangled from the spin polarizability effect itself. A clear signal for $\mbox{\hspace{0.5cm}}$ \begin{figure}[h] \vspace{-0.18cm} \hspace{0.05cm} \epsfig{file=new_eftdata_short.ps,width=4.9cm,angle=-90} \caption{Measured lattice data in relation to par- \hspace{7.7cm} $\mbox{\ \ } $ tially quenched chiral perturbation theory, cf.~text.} \label{figeft} \vspace{-6.67cm} \end{figure} \hspace{7.3cm} \parbox{7.02cm}{ the connected contribution to $\gamma_{E1} $ was obtained at the pion mass $m_{\pi } = 759\, \mbox{MeV} $; also a preliminary extraction at $m_{\pi } = 357\, \mbox{MeV} $ proved possible, although a refinement of the statistical accuracy is desirable. The most striking feature of the connected contribution to $\gamma_{E1} $ obtained here is its smallness compared to the result expected for full QCD from chiral perturbation theory \cite{detcsb}, namely, a negative $\gamma_{E1} $ larger in modulus by roughly two orders of magnitude. This suggests that the electric spin polarizability $\gamma_{E1} $ of the neutron is dominated by the disconnected contributions.} \vspace{0.099cm} \section*{Acknowledgments} Discussions with F.~X.~Lee, presenting related work at this conference, and W.~Detmold are gratefully acknowledged. Computations were performed using Chroma \cite{chroma} on U.S.~DOE/USQCD resources at Fermilab. This work was supported by U.S.~DOE grant DE-FG02-96ER40965.
\section{Introduction} The permanent of an $m \times n$ matrix $A$ (height $m$ and width $n$) satisfying $m \leq n$ is defined as \[ \perm A = \sum_{\pi \in S_{m,n}} \prod_{i=1}^n a_{i,\pi(i)}, \] where $S_{m,n}$ is the set of one-to-one functions from $[m]=\{1,\dots,m\}$ to $[n]=\{1,\dots,n\}$. When $m=n$, that is when $A$ is a square matrix, $S_{m,n} = S_n$ the set of permutations on $[n]$. In this paper we will study asymptotics of permanents of large matrices with positive, independent and identically distributed elements. Permanents of random matrices of similar type have been studied in a number of papers. In \cite{Girko71} and \cite{Girko72} Girko proved that \[ \lim_{n \to \infty}\frac{\log |\perm A| - \mathbb{E}\log |\perm A|}{n} \to 0, \] (without estimating $\mathbb{E}\log |\perm A|$) for $n \times n$ square matrices $A$ with independent elements, when either characteristic functions or Laplace transforms of elements is of the form $\exp(-c|t|^{\alpha})$ (and some other finite man cases). Working in the context of perfect matchings on random bipartite graphs, Janson \cite{Janson94} proved central limit theorems for permanents of matrices with $0$-$1$ iid elements. In a series of papers Rempa{\l}a and Weso{\l}owski studied the permanents of large rectangular matrices with identically distributed elements of non-zero mean and finite variance (allowing some correlation among elements in each column). In the case of iid elements, relying on earlier results of van Es and Helmers \cite{EsHelmers88} and Borovskikh and Korolyuk \cite{KB92}, they proved central limit theorems \cite{RempalaWesolowski99} for $(\perm A) / \mathbb{E}\perm A$ and later certain strong laws of large numbers \cite{RW02}. See also Chapter 3 in \cite{RempalaWesolowski08} for a self-contained discussion of these results. Recently Tao and Vu \cite{TaoVu08} obtained significantly different behavior for $n \times n$ matrices $A$ with independent mean zero Bernoulli $\pm 1$ elements. They showed that with high probability $|\perm(A)| = n^{(\frac{1}{2}+o(1))n}$. The above results demonstrate the contrast between the non-zero mean, finite variance case and the Bernoulli case, which can be summarized as \begin{equation}\label{eq:old_limits} \lim_{m,n \to \infty}\frac{\log |\perm A|}{m \log n} = \left\{\begin{array}{r l} 1, & \text{ in the case of the finite variance and non zero mean,} \\ \frac{1}{2}, & \text{ in the Bernoulli case with zero mean for } m=n. \end{array}\right. \end{equation} In the non-zero finite mean case ($\mu$ being mean of the elements) the value of the limit, and especially the upper bounds, can be inferred by calculating the first moment $\mathbb{E}(\perm A)=\binom{n}{m}m!\mu^n$, and in the Bernoulli case from the second moment $\mathbb{E}(|\perm A|^2)=n!$. In this paper we will calculate the value of this limit under the assumption that elements are positive and have power law decaying tails $\mathbb{P}(\xi \geq t) = t^{-1/\beta +o(1)}$. In the case $\beta > 1$ elements have infinite mean which prevents us from guessing the value of the limit. Actually in Theorem \ref{thm:main} we will observe a first order phase transition in the limit at $\beta =1$, when the mean becomes infinite. \section{Setup and the Result} In the text we will assume that for $m \leq n$, $A_{m,n}$ is an $m \times n$ matrix ($A_n$ when $m=n$) with independent positive elements distributed as $\xi$. Note that we will drop the subscripts when there is no confusion. Assuming that the matrices are constructed on a common probability space, theorems below give strong laws of large numbers for $\frac{\log \perm A_n}{n\log n}$ (in particular they imply a weak law of large numbers without the assumption that $A_n$ are given on a common probability space). Extensions to rectangular matrices are given in Section \ref{sec:non-square}. \begin{theorem}\label{thm:main} Let $\xi$ be a positive random variable satisfying \begin{equation}\label{eq:heavy_tail_assumption} \lim_{t \to \infty}\frac{\log \mathbb{P}(\xi \geq t)}{\log t} = - \frac{1}{\beta}, \end{equation} for some $\beta > 0$. If $(A_n)_n$ is a sequence $n \times n$ matrices on a common probability space with elements which are independent and identically distributed as $\xi$, then almost surely \begin{equation}\label{eq:main_result} \lim_{n \to \infty} \frac{\log \perm A_n}{n \log n} = \max(1,\beta). \end{equation} \end{theorem} Random variable $\xi$ in \eqref{eq:heavy_tail_assumption} has finite variance for $\beta < 1/2$, finite mean for $\beta < 1$, and infinite mean for $\beta > 1$ when we observe a limit different from the values in \eqref{eq:old_limits}. The following result generalizes the case $\beta < 1$. It does not require the finite variance assumption and gives the general lower bounds and the upper bounds in the case of finite mean uniformly over all submatrices of linear size. Note that for an $m \times n$ matrix $A=(a_{ij})$ any matrix $B=(a_{ij})_{i\in I,j\in J}$, where $I \subset [m]$, $J \subset [n]$ is called a submatrix of of $A_n$. \begin{theorem}\label{thm:general_bounds} \begin{itemize} Assume that $(A_n)_n$ is a sequence of $n \times n$ matrices on a common probability space with elements which are independent and identically distributed as $\xi$, and let $0 < \alpha < 1$. \item[i)] We have \begin{equation}\label{eq:general_lower_bounds} \liminf_{n \to \infty} \min_{(k,B)}\frac{\log \perm B}{k \log k} \geq 1, \end{equation} where the minimum is taken over all integers $\alpha n \leq k \leq n$ and all $k \times k$ submatrices $B$ of $A_n$. \item[ii)] If $\xi$ has a finite mean then \begin{equation}\label{eq:general_upper_bounds} \limsup_{n \to \infty} \max_{(k,B)}\frac{\log \perm B}{k \log k} = 1, \end{equation} where the maximum is again taken over all integers $\alpha n \leq k \leq n$ and all $k \times k$ submatrices $B$ of $A_n$. \end{itemize} \end{theorem} Condition \eqref{eq:heavy_tail_assumption} in Theorem \ref{thm:main} is satisfied with $\beta > 1$ for many common heavy tail distributions including Pareto distribution, L\'{e}vy distribution, Inverse-Gamma distribution, Beta-prime distribution and many more. We are particularly interested in the case when $\xi$ has Pareto distribution with parameter $\beta$, that is $\mathbb{P}(\xi \geq t) =t^{-1/\beta}$ for $t \geq 1$. Actually in Section \ref{sec:upper_bounds} the upper bounds in Theorem \ref{thm:main} will be proven in the Pareto case and then extended to the general case via simple stochastic domination. Note that when the convergence \eqref{eq:heavy_tail_assumption} fails to hold, one cannot guarantee the existence of the limit in \eqref{eq:main_result} (see Example \ref{ex:no_convergence} in Section \ref{sec:non-square}). However, the upper bound on the $\limsup$ in \eqref{eq:heavy_tail_assumption} will imply the upper bound in \eqref{eq:main_result} and similarly the lower bound for $\liminf$. \begin{remark}\label{rem:matchings} From a more combinatorial point of view permanents can be interpreted in the context of saturated matchings (or perfect matchings for $m=n$) of bipartite graphs. For a bipartite graph $G=(V,E)$ let $V=V_1 \cup V_2$, $|V_1| \leq |V_2|$ be a decomposition of the vertex sets into subsets so that no two vertices in $V_i$ are connected by an edge. Saturated matchings of $G$ can be defined as subsets $\mathcal{M} \subset E$ of the edge set with the property that every vertex is adjacent to at most one edge in $\mathcal{M}$ and that every vertex in the smaller component $V_1$ is adjacent to at least one edge in $\mathcal{M}$. For $m \leq n$ and an $m\times n$ matrix $A=(a_{ij})$ containing only elements $0$ and $1$, construct a bipartite graph $G$ with $m+n$ vertices $\{v_1, \dots v_m, w_1, \dots w_n\}$ so that $v_i$ and $w_j$ are connected by an edge if and only if $a_{ij}=1$. Clearly every one-to-one function $\pi\colon [m] \to [n]$ for which $\prod_{i=1}^n a_{i\pi(i)}$ is non-vanishing corresponds to a saturated matching on $G$ in $1-1$ manner. Therefore $\perm A$ is equal to the number of saturated matchings on $G$. For general matrices $A$ one can construct the graph by drawing an edge between $v_i$ and $w_j$ whenever $a_{ij} \neq 0$ and putting the weight $a_{ij}$ on this edge. Then $\perm A$ can be interpreted as the total weight of all the saturated matchings on $G$ (a weight of a matching being the product of the weight on its edges). All the results in this paper can be interpreted in this way. \end{remark} In the following section we prove the upper bounds in Theorem \ref{thm:main} and in Section \ref{sec:lower_bounds} we provide the lower bounds and prove Theorem \ref{thm:general_bounds}. In the last section we will extend the results to a large class of rectangular matrices and show an example demonstrating that, in general without \eqref{eq:heavy_tail_assumption} Theorem \ref{thm:main} fails to hold. \section{Proof of the upper bounds in Theorem \ref{thm:main}}\label{sec:upper_bounds} The exact calculations needed for the proof of the upper bounds in Theorem \ref{thm:main} are easier to perform when we are given a concrete distribution of $\xi$ to work with. The proof will be provided for the Pareto case, but first we will see how this yields the upper bounds in Theorem \ref{thm:main} for the general case. \begin{remark}\label{rem:stochastic_domination} Throughout the paper we will use the following two simple observations. \emph{i)} For any $m \times n$ matrix $A$ and $\lambda \in \mathbb{R}$ we have that $\perm (\lambda A) = \lambda^m \perm A$. Thus the value of the limit of $\log \perm A_n /(n \log n )$ in Theorem \ref{thm:main} (as well as $\liminf$ and $\limsup$) is unchanged if we replace the generic random variable $\xi$ by random variables $\lambda \xi$, for any $\lambda >0$. \emph{ii)} Assume that we are given two random variables $\xi_1$ and $\xi_2$ with right continuous cumulative distribution functions $F_1(t)$ and $F_2(t)$ (and denote $F_i^-(t) = \lim_{s \uparrow t}F_i(s)$). If $\xi_2$ stochastically dominates $\xi_1$, that is $F_2(t) \leq F_1(t)$ for any $t> 0$ (equivalently $\mathbb{P}(\xi_1 \geq t) \leq \mathbb{P}(\xi_2 \geq t)$), then one can enlarge the probability space $\Omega$ that supports $\xi_1$ and construct a version of $\xi_2$ on the larger probability space which dominates $\xi_1$ pointwise. For example, if $\xi_1$ is defined on $(\Omega,\mathbf{P})$ then on $(\Omega \times [0,1], \mathbf{P}\times d\lambda)$, where $d\lambda$ is the Lebesgue measure on $[0,1]$, we define $u = \lambda F_1(\xi_1) + (1-\lambda)F_1^-(\xi_1)$. It is easy to check that $u$ is uniformly distributed on $[0,1]$. Knowing this it is also easy to check that $\overline{\xi}_2 = \inf\{t:F_2(t) \geq u\}$ has the same distribution as $\xi_2$ and that $F_1^-(\xi_1) \leq u \leq F_2(\overline{\xi_2})$. The last inequalities imply that $\xi_1 \leq \overline{\xi}_2$ everywhere. Thus if $(A_{m,n}^1)$ is a sequence of $m \times n$ matrices with independent elements distributed as $\xi_1$ defined on a common probability space, one can enlarge the probability space and construct on it a sequence of $m \times n$ matrices $(A_{m,n}^2)$ whose elements are independent and distributed as $\xi_2$ such that for every $n$, the $ij$th element of $A_{m,n}^1$ is not larger than the corresponding element of $A_{m,n}^2$. In particular, if $\xi_1$ and $\xi_2$ are almost surely positive then $\perm A_{m,n}^1 \leq \perm A_{m,n}^2$. The analogous claim holds when $\xi_1$ dominates $\xi_2$. \end{remark} \begin{proof}[Proof of the upper bounds in Theorem \ref{thm:main} assuming it holds for the Pareto case] Fix $\epsilon>0$ and take $M>1$ so that $\mathbb{P}(\xi \geq t) \leq t^{-1/(\beta+\epsilon)}$ holds for all $t \geq M$. Denote by $\overline{\xi}_{\beta+\epsilon}$ a Pareto distributed random variable with parameter $\beta+\epsilon$ and observe that $\mathbb{P}(\xi \geq t) \leq \mathbb{P}(M\overline{\xi}_{\beta+\epsilon} \geq t)$ holds for all $t$. Assuming the statement holds for the Pareto case, Remark \ref{rem:stochastic_domination} implies that almost surely \[ \limsup_n \frac{\log \perm A_n}{n \log n} \leq \beta + \epsilon. \] Since $\epsilon > 0$ was arbitrary the claim follows. \end{proof} The rest of this section is devoted to the proof of the upper bounds in the Pareto case in which we show explicit calculations. A useful observation which we will use extensively is the fact that if $\xi$ is a Pareto distributed random variable with parameter $\beta$, then $Y = (\log \xi)/\beta$ has exponential distribution with rate $1$, that is $\mathbb{P}(Y \geq t) = e^{-t}$ for $t\geq 0$. We start by proving some basic estimates for maxima of independent exponential random variables. \begin{lemma} \label{lemma: Max} Let $n \geq 2$ and $Y_i$, $1 \leq i \leq n$ be independent exponential random variables with rate 1 and $ R=\frac{\max_{1 \leq i \leq n} Y_i}{\log n}$. \begin{itemize} \item[i)] For any positive $t$ \begin{equation} \label{eq: Max1} \mathbb{P}(R \leq t) \leq \exp\bigl(-n^{1-t}\bigr), \textrm{ and } \ \mathbb{P}(R \geq t) \leq n^{1-t}. \end{equation} \item[ii)] The expectation of $e^{R}$ can be bounded as \begin{equation} \label{eq: Max2} \mathbb{E}\left(e^R\right) \leq \exp\biggl(1+\frac{1}{\log n -1}\biggr). \end{equation} \end{itemize} \end{lemma} \begin{proof} i) Both inequalities are straightforward. First we calculate \begin{equation} \label{eq: RDown1} \mathbb{P} (R \leq t) = \mathbb{P} (\max_{1 \leq i \leq n}Y_i \leq t\log n) = \left(1-e^{-t\log n}\right)^n = \left(1-n^{-t}\right)^n. \end{equation} Now applying the inequality $1-x \leq e^{-x}$ to the right hand side of (\ref{eq: RDown1}) we obtain the first inequality in (\ref{eq: Max1}). Using (\ref{eq: RDown1}) and the inequality $(1-x)^n \geq 1-nx$ which holds for all positive integers $n$ and all $0 < x < 1$ we prove the second inequality in (\ref{eq: Max1}): \begin{equation*} \mathbb{P}(R \geq t) = 1- (1-n^{-t})^n \leq n^{1-t}. \end{equation*} ii) Using the second inequality in (\ref{eq: Max1}) we obtain for $t \geq 1$ $$ \mathbb{P} \left(e^{R} \geq t\right) = \mathbb{P}(R \geq \log t) \leq \frac{n}{n^{\log t}} = \frac{n}{t^{\log n}}, $$ from where we get \begin{multline*} \mathbb{E}\left(e^R\right) = \int_0^{\infty}\mathbb{P}\left(e^R \geq t\right) dt \leq e+\int_e^{\infty}\frac{n}{t^{\log n}} dt = e+\frac{n}{\log n -1}\frac{1}{e^{\log n-1}} \\ = e \left(1 + \frac{1}{\log n -1}\right) \leq \exp\biggl(1+\frac{1}{\log n- 1}\biggr) \end{multline*} \end{proof} The idea of the proof of the upper bounds in the Pareto case is to estimate (by evaluating the expectation) the number of permutations $\pi$ for which the product $\prod_{i=1}^n \xi_{i\pi(i)}$ will lie in some given interval. The key estimate is provided in Lemma \ref{prop: BoundZMax}. We will only consider the intervals not exceeding $(n\sqrt{\log n})^{\beta n}$, since as the following lemma shows, the largest product $\prod_{i}\xi_{i\pi(i)}$ typically does not exceed this value. \begin{lemma} \label{lemma: MaxPerm} If $(Y_{i,j})_{i,j}$ are independent exponential random variables with rate 1 then for any $\lambda>0$ \begin{equation} \label{eq: MaxPermStrong} \sum_{n = 2}^{\infty}\mathbb{P}\left(\max_{\pi \in S_n}\sum_{i=1}^n Y_{i,\pi(i)} \geq n\log n + n \frac{\log\log n}{\lambda}\right) < \infty. \end{equation} \end{lemma} \begin{proof} Denote $R_i := \frac{\max_{1 \leq j \leq n}Y_{i,j}}{\log n}$. From the definition of $R_i$ it is obvious that $$ \max_{\pi \in S_n} \sum_{i=1}^n Y_{i,\pi(i)} \leq \log n \left(\sum_{i=1}^n R_i\right).$$ Using the inequality (\ref{eq: Max2}) we have \begin{multline*} \mathbb{P} \left(\max_{\pi \in S_n}\sum_{i=1}^n Y_{i,\pi(i)} \geq n\log n + n \frac{\log\log n}{\lambda}\right) \leq \mathbb{P}\left(\sum_{i=1}^n R_i \geq n + n \frac{\log\log n}{\lambda\log n}\right) \\ \leq \mathbb{E}\left(e^{\sum_{i=1}^nR_{i} - n - n \frac{\log\log n}{\lambda\log n}}\right) = \left(\frac{\mathbb{E}\left(e^{R_{1}}\right)}{e(\log n)^{\frac{1}{\lambda \log n}}}\right)^n \leq \exp\biggl(\Big(\frac{1}{\log n -1} - \frac{\log \log n}{\lambda \log n}\Big)n\biggr). \end{multline*} The right hand side above is summable in $n$ which proves the lemma. \end{proof} \begin{lemma} \label{prop: BoundZMax} Let $(Y_{i,j})_{i,j}$ be independent exponential random variables with rate 1 and \begin{equation}\label{eq:number_in_interval} Z_{n,k}=\bigg|\biggl\{\pi \in S_n: (k-1)n \leq \sum_{i=1}^nY_{i,\pi(i)} < kn \biggr\}\biggr|. \end{equation} Then for any $\gamma>1$ and $\lambda > 1$ we have \begin{equation} \label{eq: BoundZMax} \sum_{n = 2}^{\infty} \mathbb{P}\biggl(\bigcup_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}}\left\{Z_{n,k} > \mathbb{E}(Z_{n,k})^{\gamma}\right\}\biggr) < \infty. \end{equation} \end{lemma} \begin{proof} First by Markov's inequality \begin{align*} \label{eq: BoundZMarkov} \mathbb{P}\biggl(\bigcup_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}}\left\{Z_{n,k} > \mathbb{E}(Z_{n,k})^{\gamma}\right\}\biggr) & \leq \sum_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \mathbb{P}\left(Z_{n,k} > \mathbb{E}(Z_{n,k})^{\gamma}\right) \\ & \leq \sum_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{1}{ \mathbb{E}(Z_{n,k})^{\gamma-1}}. \end{align*} Let's calculate the expectation of $Z_{n,k}$. Clearly for any fixed $\pi \in S_n$ we have $$\mathbb{E}(Z_{n,k})=n!\mathbb{P}\biggl((k-1)n \leq \sum_{i=1}^nY_{i,\pi(i)} < kn \biggr).$$ Since for a fixed $\pi$, $\sum_{i=1}^nY_{i,\pi(i)}$ is the sum of $n$ independent exponential random variables with mean $1$, it has Gamma density $\frac{x^{n-1}e^{-x}}{(n-1)!}$. Therefore the expectation of $Z_{n,k}$ is given by \begin{equation} \label{eq: Z mean} \mathbb{E}(Z_{n,k})=n! \int_{(k-1)n}^{kn} \frac{x^{n-1}e^{-x}}{(n-1)!}\ dx = n \int_{(k-1)n}^{kn} x^{n-1}e^{-x}\ dx. \end{equation} In particular we have \begin{equation} \label{eq: upper bound for Z mean} \mathbb{E}(Z_{n,k}) \leq e^{-(k-1)n}\int_{(k-1)n}^{kn}nx^{n-1}dx \leq \left(kn\right)^{n}e^{-(k-1)n}. \end{equation} Furthermore by (\ref{eq: Z mean}) \begin{equation} \label{eq: Z1} \mathbb{E}(Z_{n,k})=n \int_{(k-1)n}^{kn} x^{n-1}e^{-x}\ dx \geq e^{-kn}\int_{(k-1)n}^{kn}nx^{n-1}dx = e^{-kn}n^n(k^n-(k-1)^n). \end{equation} In particular $\mathbb{E}(Z_{n,1}) \geq \left( \frac{n}{e}\right)^n$, which implies that the series $\sum_{n=1}^{\infty}\mathbb{E}(Z_{n,1})^{1-\gamma}$ converges to a finite limit. Therefore we are left to prove \begin{equation} \label{eq: Z2} \sum_{n =2}^{\infty} \sum_{2 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{1}{ \mathbb{E}(Z_{n,k})^{\gamma-1}} < \infty. \end{equation} Again using (\ref{eq: Z1}) and the inequality $k^n \geq (k-1)^n + n(k-1)^{n-1}$ we obtain $$ \mathbb{E}(Z_{n,k}) \geq e^{-kn}n^{n+1}(k-1)^{n-1} \geq e^{-kn}n^n(k-1)^n, $$ from where \begin{equation}\label{eq: BoundUp} \sum_{2 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{1}{ \mathbb{E}(Z_{n,k})^{\gamma-1}} \leq \frac{e^{n (\gamma-1)}}{n^{n(\gamma-1)}} \sum_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{e^{kn(\gamma-1)}}{k^{n (\gamma-1)}}. \end{equation} The function $g(t)=e^{tn(\gamma-1)}t^{-n (\gamma-1)}$ is convex since $$g''(t)=n(\gamma-1) \frac{e^{tn(\gamma-1)}}{t^{n (\gamma-1) +2}}(n (\gamma-1) (t-1)^2+1) \geq 0.$$ Therefore for any $1 \leq k \leq \log n + \frac{\log \log n}{\lambda}$ we have \begin{equation} \label{eq: gConvex1} \frac{e^{kn(\gamma-1)}}{k^{n (\gamma-1)}} = g(k) \leq \max\left\{g(1),g\left(\log n + \frac{\log \log n}{\lambda}\right)\right\}. \end{equation} For any $n$ large enough we have $$ g\left(\log n + \frac{\log \log n}{\lambda}\right) = \frac{n^{n(\gamma-1)}(\log n)^{\frac{n (\gamma-1)}{\lambda}}}{(\log n + \frac{\log \log n}{\lambda})^{n (\gamma-1)}} \geq \left(\frac{n}{2(\log n)^{1-1/\lambda}}\right)^{n (\gamma-1)} \geq e^{n (\gamma-1)} = g(1), $$ and, for such $n$, using (\ref{eq: gConvex1}) we also get $$ \frac{e^{kn(\gamma-1)}}{k^{n (\gamma-1)}} \leq g\left(\log n + \frac{\log \log n}{\lambda}\right) = \frac{n^{n(\gamma-1)}(\log n)^{\frac{n (\gamma-1)}{\lambda}}}{(\log n + \frac{\log \log n}{\lambda})^{n (\gamma-1)}} \leq \frac{n^{n(\gamma-1)}}{(\log n)^{n(\gamma-1)(1-1/\lambda)}}. $$ Thus, for $n$ large enough (\ref{eq: BoundUp}) yields \begin{align*} \sum_{2 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{1}{ \mathbb{E}(Z_{n,k})^{\gamma-1}} & \leq \frac{e^{n (\gamma-1)}}{n^{n(\gamma-1)}} \frac{n^{n(\gamma-1)}}{(\log n)^{n(\gamma-1)(1-1/\lambda)}}\left(\log n + \frac{\log \log n}{\lambda}\right) \\ & = \left(\frac{e}{(\log n)^{1-1/\lambda}}\right)^{n(\gamma-1)}\biggl(\log n + \frac{\log \log n}{\lambda}\biggr). \end{align*} The expression on the right hand side is summable in $n$ which proves (\ref{eq: Z2}) and thus also (\ref{eq: BoundZMax}). \end{proof} Now we are ready to finish the proof of upper bounds. \begin{proof}[Proof of the upper bounds in Theorem \ref{thm:main} for the Pareto case] Define $Z_{n,k}$ as in \eqref{eq:number_in_interval} and fix an arbitrary $\gamma > 1$. Lemmas \ref{lemma: MaxPerm} and \ref{prop: BoundZMax} and Borel-Cantelli lemma imply that almost surely there exists a positive integer $n_0$ such that for all $n \geq n_0$ we have \begin{align*} & \max_{\pi \in S_n}\sum_{i=1}^nY_{i,\pi(i)} \leq n \log n + \frac{n \log\log n}{2} \ \ {\rm and} \\ & Z_{n,k} \leq \mathbb{E}(Z_{n,k})^{\gamma}, \text{ for each } 1 \leq k \leq \log n + \frac{\log \log n}{2}. \end{align*} Using (\ref{eq: upper bound for Z mean}) the following inequalities are almost surely satisfied for $n$ large enough \begin{multline} \label{eq: PermEstimate} \perm A = \sum_{\pi \in S_n} e^{\beta \sum_{i=1}^{n}Y_{i,\pi(i)}} \leq \sum_{1 \leq k \leq \log n + \frac{\log \log n}{2}}e^{\beta kn}Z_{n,k} \leq \sum_{1 \leq k \leq \log n + \frac{\log \log n}{2}}e^{\beta kn}\mathbb{E}(Z_{n,k})^{\gamma} \\ \leq e^{\gamma n} n^{\gamma n} \sum_{1 \leq k \leq \log n + \frac{\log \log n}{2}}k^{\gamma n}e^{(\beta - \gamma) kn} \\ \leq e^{\gamma n} n^{\gamma n} \left(\log n + \frac{\log \log n}{2}\right) \max_{1 \leq \tau \leq \log n+\frac{\log \log n}{2 }}\Big(\tau^{\gamma n}e^{ (\beta - \gamma)\tau n}\Big). \end{multline} If $\beta > 1$ and $\gamma$ is such that $\beta > \gamma > 1$, $\tau^{\gamma n}e^{ (\beta - \gamma)\tau n}$ is an increasing function in $\tau$ and thus for $n$ large enough $$ \perm A \leq e^{\gamma n}n^{\beta n} (\log n)^{\frac{(\beta - \gamma)n}{2}} \left(\log n + \frac{\log \log n}{2}\right )^{\gamma n +1}. $$ This yields $$ \frac{\log \perm A}{n \log n} \leq \frac{\gamma}{\log n} + \beta + \frac{\beta - \gamma}{2}\frac{\log \log n}{\log n} + \left (\gamma + \frac{1}{n}\right)\frac{\log \left(\log n + \frac{\log \log n}{2}\right)}{\log n}, $$ from where clearly $$ \limsup_{n \to \infty} \frac{\log \perm A}{n \log n} \leq \beta. $$ In the case $\beta \leq 1$ we want to maximize the function $ \tau^{\gamma n}e^{(\beta - \gamma)\tau n}$. Write $e^{h(\tau)}:= \tau^{\gamma n}e^{(\beta - \gamma)\tau n}$. We get $$h(\tau)=\gamma n \log \tau + (\beta - \gamma) \tau n, \ \ h'(\tau) = \frac{\gamma n}{\tau} + (\beta - \gamma)n, \ \ h''(\tau)=- \frac{\gamma n}{\tau^2} < 0.$$ Therefore function $h$ is concave and the maximum occurs when $h'(\tau)=0$, that is when $\tau = \frac{\gamma}{\gamma-\beta}$ at which the value of the function $e^{h(\tau)}$ is equal to $\left(\frac{\gamma}{(\gamma - \beta)e}\right)^{\gamma n}$. From (\ref{eq: PermEstimate}) we get $$ \perm A \leq n^{\gamma n}\left(\log n + \frac{\log \log n}{2}\right)\left(\frac{\gamma}{\gamma - \beta}\right)^{\gamma n} $$ and so $$ \frac{\log \perm A}{n \log n} \leq \gamma + \frac{\log \left(\log n + \frac{\log \log n}{2}\right)}{n \log n} +\gamma\frac{\log \gamma - \log (\gamma - \beta)}{\log n} \to \gamma, $$ as $n\to \infty$. Since $\gamma >1$ was arbitrary the claim follows. \end{proof} \section{Lower bounds and the proof of Theorem \ref{thm:general_bounds}}\label{sec:lower_bounds} In this section we prove Theorem \ref{thm:general_bounds} as well as the lower bounds in Theorem \ref{thm:main}. An important ingredient is the use of stochastic domination to reduce certain technical issues to iid $0$, $1$ matrices. The following result proven by Hall \cite{Hall48} and Mann and Ryser in \cite{Mann_Ryser53} provides lower bounds for permanents of such matrices (see also Theorem 1.2 in Chapter 4 of \cite{Minc78}). \begin{proposition}\label{thm:lower_bounds_for_01} Let $A$ be an $m \times n$ matrix, $m\leq n$ whose all elements are equal to $0$ or $1$. Assume that each row of $A$ contains at least $k$ elements equal to $1$. If $k \geq m$, then \begin{equation}\label{eq:lower_bounds_on_01_1} \perm A \geq \frac{k!}{(k-m)!}. \end{equation} If $k < m$ and $\perm A > 0$ then \begin{equation}\label{eq:lower_bounds_on_01_2} \perm A \geq k!. \end{equation} \end{proposition} As discussed in the introduction (see Remark \ref{rem:matchings}) permanents of matrices with $0$, $1$ elements can be viewed as the number of saturated matchings on corresponding bipartite graphs. To ensure the positivity of the permanent, when applying \eqref{eq:lower_bounds_on_01_2}, we will exploit this connection through the classical Hall's marriage theorem, which can be easily stated in this setting (see \cite{Hall35}). \begin{theorem}\label{thm:positivity_of_01} Let $G=(V,E)$ be a bipartite graph and let $V = V_1 \cup V_2$ be a decomposition of the vertex set so that no two vertices in $V_i$ are connected by an edge, $i=1,2$. Assuming $|V_1| \leq |V_2|$, there exists a saturated matching on $G$ if and only if for any subset $W \subset V_1$ we have $|W| \leq |\{v: v\sim w, w \in W\}|$. \end{theorem} Restating the above theorem in terms of permanents of $0$, $1$ matrices yields the following lemma. \begin{lemma}\label{lemma:permanents_and_matchings} Let $B$ be an $m \times n$ matrix whose all elements are either $0$ or $1$. If for any $1 \leq k \leq m$ any $k \times (n-k+1)$ submatrix of $B$ has at least one element equal to $1$, then $\perm B \geq 1$. \end{lemma} All the necessary applications of Proposition \ref{thm:lower_bounds_for_01} and Lemma \ref{lemma:permanents_and_matchings} are summarized in Lemma \ref{lemma:lower_bound_for_submatrices} which, in particular, proves the lower bounds in Theorem \ref{thm:general_bounds}. \begin{remark}\label{ex:stirling} Recall that Stirling's formula says that \[ \lim_{n \to \infty} n!e^nn^{-(n+1/2)} = \sqrt{2\pi}. \] In particular there are constants $c_1< c_2$ so that for any $n$ and $1 \leq k \leq n-1$ \begin{equation}\label{eq:stirling_for_binom} c_1 \frac{n^{n+1/2}}{k^{k+1/2}(n-k)^{n-k+1/2}} \leq \binom{n}{k} \leq c_2 \frac{n^{n+1/2}}{k^{k+1/2}(n-k)^{n-k+1/2}}. \end{equation} \end{remark} \begin{lemma}\label{lemma:lower_bound_for_submatrices} Let $\xi$ be a positive random variable and let $A_n$ be a sequence of $n \times n$ matrices whose elements are independent and identically distributed as $\xi$. For any $0 < \alpha < 1$ and any $\delta > 0$ there exists $r>0$ with the following property: Almost surely there exists $n_0$ such that for any $n \geq n_0$ and any $\alpha n \leq k \leq n$, any $k \times k$ submatrix $B$ of $A_n$ satisfies $\perm B \geq r^k k^{(1-\delta)k}$. \end{lemma} \begin{proof} Let $q>0$ be such that $\mathbb{P}(\xi \leq q) < \eta$, where $\eta$ is to be chosen later. Define the random variable $\tilde{\xi} = \mathbf{1}_{(\xi \geq q)}$. and define the matrix $\tilde{A}_n = (\tilde{\xi}_{ij})$. Let $\mathfrak{B}_n$ denote the event that some row of $\tilde{A}_n$ contains more than $\alpha \delta n $ zeros and let $\mathfrak{C}_n$ denote the event that for some $k_1$ and $k_2$ satisfying $\alpha n \leq k_1 + k_2$ there exists a $k_1 \times k_2$ submatrix of $\tilde{A}_n$ containing only zeros. By Lemma \ref{lemma:permanents_and_matchings} on the event $\frak{C}_n^c$ any $k\times k$ submatrix of $\tilde{A}_n$ has a positive permanent, for $\alpha n \leq k \leq n$. Furthermore on the event $\frak{B}_n^c$ every $k \times k$ submatrix of $\tilde{A}_n$ for $k \geq \alpha n$ contains at least $(1-\delta)k$ ones. Thus on the event $\frak{B}_n^c \cap \frak{C}_n^c$ by \eqref{eq:lower_bounds_on_01_2} we have for any $k \geq \alpha n$ and any $k \times k$ submatrix $B$ of $A_n$ \[ \perm B \geq q^k \perm \tilde{B} \geq q^k \lfloor (1-\delta)k\rfloor! \geq \Big(\frac{q(1-\delta)}{e}\Big)^kk^{(1-\delta)k}, \] where $\tilde{B}$ is the submatrix of $\tilde{A}_n$ having the same rows and columns as $B$ in $A_n$. Note that the last inequality above holds for $n$ large enough by Stirling's approximation. Thus we only need to prove that the probabilities of the events $\frak{B}_n \cup \frak{C}_n$ are summable (since then they happen only finitely many times almost surely). To end this observe that the average number of $1$s in every row and column of $\tilde{A}_n$ is greater than $n(1-\eta)$, so for $\eta < \alpha \delta$ by standard large deviation arguments there exists a constant $C$ such that $\mathbb{P}(\frak{B}_n) \leq Cne^{-n(\alpha\delta-\eta)/C}$, which is clearly summable. For $\frak{C}_n$ use union bound to obtain \begin{multline}\label{eq:large_deviations_for _rectangles} \mathbb{P}(\frak{C}_n) \leq 2\sum_{\alpha n \leq k_1+k_2 \leq n \atop k_1 \geq k_2}\binom{n}{k_1}\binom{n}{k_2}\eta^{k_1k_2} \leq 2\sum_{\alpha n/2 \leq k_1 \leq n} \binom{n}{k_1} \sum_{1\leq k_2 \leq n}\binom{n}{k_2}\eta^{k_1k_2} \\ \leq 2\sum_{\alpha n/2 \leq k_1 \leq n} \binom{n}{k_1} \Big(\Big(1+\eta^{k_1}\Big)^n-1\Big) \leq 2\sum_{\alpha n/2 \leq k_1 \leq n} \binom{n}{k_1} (2\eta)^{k_1}, \end{multline} for $\eta$ small enough. It is easy to check the last inequality by writing $\big(1+\eta^{k_1}\big)^n-1 = \eta^{k_1} \sum_{\ell=0}^{n-1}\big(1+\eta^{k_1}\big)^\ell$ and bounding each of the terms on the right hand side. To prove that the right hand side above is summable, observe that for $\eta=\eta(\alpha)$ sufficiently small the following inequalities hold for $\alpha n/2 \leq k_1 \leq n$ \[ (2\eta)^{k_1/2} \leq (2\eta)^{\alpha n/4} \leq (1-\sqrt{2\eta})^{(1-\alpha/2)n} \leq (1-\sqrt{2\eta})^{n-k_1}. \] Plugging this back into \eqref{eq:large_deviations_for _rectangles} we get \[ \mathbb{P}(\frak{C}_n) \leq 2 \sum_{\alpha n/2 \leq k_1 \leq n} \binom{n}{k_1} (2\eta)^{k_1/2}(1-\sqrt{2\eta})^{n-k_1}. \] The right hand side is just twice the probability that the sum of $n$ independent random variables having value $1$ with probability $\sqrt{2\eta}$ and value $0$ otherwise, is greater than $\alpha n/2$. Choosing $\sqrt{2\eta} < \alpha /2$, large deviation principle implies that this probability is exponentially small, and thus summable in $n$. This finishes the proof. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:general_bounds}] i) Taking an arbitrary $\delta > 0$ by Lemma \ref{lemma:lower_bound_for_submatrices} we can find $r > 0$ small enough so that almost surely for $n$ large enough \[ \frac{\log \perm B}{k \log k} \geq \frac{\log r}{\log k} + (1-\delta), \] for any $\alpha n \leq k \leq n$ and any $k \times k$ submatrix $B$ of $A_n$. Thus almost surely \[ \liminf_n \min_{B,k}\frac{\log \perm B}{k \log k} \geq 1-\delta. \] Since $\delta > 0$ was arbitrary the claim follows. ii) Let $\tilde{\xi}$ be parameter 1 Pareto distributed random variable. By Markov inequality, for all $t \geq \mathbb{E}(\xi)$ we have \[ \mathbb{P}(\xi \geq t) \leq \frac{\mathbb{E}(\xi)}{t} = \mathbb{P}(\mathbb{E}(\xi)\tilde{\xi} \geq t), \] Thus $\mathbb{E}(\xi) \tilde{\xi}$ stochastically dominates $\xi$ from above and by Remark \ref{rem:stochastic_domination} we can construct a sequence $(\tilde{A}_n)$ of random $n \times n$ matrices whose elements are independent and identically distributed as $\tilde{\xi}$, such that for any $\alpha n \leq k \leq n$ and any $k \times k$ submatrix $B$ of $A_n$ for the corresponding submatrix $\tilde{B}$ of $\tilde{A}_n$ we have $\perm B \leq \mathbb{E}(\xi)^k \perm \tilde{B}$. Thus it is enough to prove the claim in the case when we replace $\xi$ with $\tilde{\xi}$. If $\tilde{B}$ is an arbitrary $k \times k$ submatrix of $\tilde{A}_n$ then after permuting the rows and columns we can assume that $\tilde{B}$ is at the intersection of the first $k$ rows and columns. Denote by $\tilde{B}^c$ the matrix at the intersection of the other $n-k$ rows and columns and observe that since $\tilde{A}_n$ has positive elements $\perm \tilde{A}_n \geq \perm \tilde{B} \perm \tilde{B}^c$. Furthermore since all the elements are larger than $1$ we have $\perm \tilde{B}^c \geq (n-k)!$ and thus \[ \frac{\perm \tilde{B}}{k \log k} \leq \frac{\perm \tilde{A}_n}{k \log k} - \frac{\log(n-k)!}{k \log k} \leq \frac{\perm \tilde{A}_n}{k \log k} - \frac{(n-k)(\log(n-k) -1) - c}{k \log k}, \] for some $c> 0$, where the second inequality inequality follows from Stirling's formula (note that we can assume that $k<n$, since for $k=n$ the upper bounds have been proven in the previous section). By the upper bounds in Theorem \ref{thm:main}, for any $\epsilon >0$ almost surely there is $n_0$ such that for all $n \geq n_0$ we have $\perm \tilde{A}_n \leq (1+\epsilon) n\log n$. For such $n$ we have \begin{multline*} \frac{\perm \tilde{B}}{k \log k} \leq (1+\epsilon)\frac{n \log n}{k \log k} - \frac{(n-k)\log(n-k)}{k \log k} + \frac{n-k}{k \log k} + \frac{c}{k \log k}\\ \leq \frac{1}{\alpha}\Big(\frac{(1+\epsilon)\log n}{\log n + \log \alpha} -1\Big) + \frac{n-k}{k \log k} + \frac{c}{k \log k} + \frac{n}{k} -\Big(\frac{n}{k}-1\Big)\frac{\log (n-k)}{\log k}. \end{multline*} For a fixed $\alpha$ the first three terms on the right hand side vanish in the limit and it suffices to show that \begin{equation}\label{eq:useful_lowerbound} \limsup_{n \to \infty}\max_{2 \leq k < n}\Big(\frac{n}{k} -\Big(\frac{n}{k}-1\Big)\frac{\log (n-k)}{\log k}\Big) \leq 1. \end{equation} Denoting $n=tk$, where $t \geq 1$ we have \[ \frac{n}{k} -\Big(\frac{n}{k}-1\Big)\frac{\log (n-k)}{\log k} = 1-\frac{(t-1)\log(t-1)}{\log k}. \] When taking supremum one can assume that $1 < t < 2$, that is $k > n/2$. In that case the claim follows from the fact that $s \mapsto s \log s$ is bounded on $(0,1)$. \end{proof} \begin{proof}[Proof of the lower bounds in Theorem \ref{thm:main}] The lower bounds for $\beta \leq 1$ follow from Theorem \ref{thm:general_bounds} i), so in the rest of the proof we will assume that $\beta > 1$. First define the random variable $Y =(\log \xi)/\beta$ and observe that \[ \lim_{t \to \infty}\frac{\log \mathbb{P}(Y \geq t)}{t} = -1, \] and thus for any $\epsilon >0$ we have $\mathbb{P}(Y \geq t) \geq \exp(-t(1+\epsilon))$, for $t$ large enough. Let $(Y_{ij})$ be an array of independent random variables distributed as $Y$, and for a fixed $n$ define \[ Q_k = \max_{1 \leq i \leq n-k+1} Y_{ki}, \text{ for } 1 \leq k \leq n. \] Following the same arguments as in the proof of the first inequality in \eqref{eq: Max1} one can prove that for any $t'>0$ \[ \mathbb{P}(Q_1 \leq t'\log n) \leq \exp(-n^{1-t'(1+\epsilon)}), \] holds for $n$ large enough. In particular if we fix $0<t'<1$, $\epsilon>0$ such that $t'(1+\epsilon) < 1$ and $0<\rho < 1$ then for $n$ large enough \begin{align}\label{eq:lower_bounds_estimates_on_almost_exponential} \sum_{1 \leq k \leq \rho n}\mathbb{P}\left(\frac{Q_k}{\log (n-k+1)} \leq t'\right) &\leq \sum_{1 \leq k \leq \rho n} \exp\big(-(n-k+1)^{1-t'(1+\epsilon)}\big) \nonumber \\ & \leq \rho n\exp\bigl(-(1-\rho)^{1-t'(1+\epsilon)}n^{1-t'(1+\epsilon)}\bigr). \end{align} We will need this estimate later in the proof. Now we will run a greedy algorithm to extract a large submatrix of $A_n$ with a large permanent. Starting from the first row of $A_n$ we pick a largest possible admissible element in each row. First define $A^{(1)} = A_n$ and $m_1$ as the smallest index between $1$ and $n$, such that $\zeta_1 := \xi_{1,m_1} = \max_{1 \leq i \leq n} \xi_{1,i}$. Next we inductively construct the matrix $A^{(k+1)}$ from $A^{(k)}$ by deleting the first row and the $m_{k}$-th column of the matrix $A^{(k)}$. At each step $\zeta_k$ is defined as the largest element in the first row of the matrix $A^{(k)}$ and $m_k$ as the first column which contains such an element. Note that conditioned on the elements of the first $k-1$ rows of the matrix $A_n$ the matrix $A^{(k)}$ is just an $(n-k+1) \times (n-k+1)$ matrix with independent elements distributed as $\xi$. This immediately implies that $\zeta_1, \dots , \zeta_n$ are independent and that the first row in $A^{(k)}$ is distributed as $(\xi_{k,1}, \dots \xi_{k,n-k+1})$. Therefore \begin{equation}\label{eq:distribution_of_max_in_rows} \Big(\frac{ \log \zeta_1}{\beta}, \dots , \frac{\log \zeta_n}{\beta}\Big) \stackrel{d}{=} (Q_1, \dots Q_n). \end{equation} Now fix some $0 < t < t' < 1$ and $t/t' < \rho < 1$. Then for $n$ large enough \begin{equation}\label{eq: HelpBound} t \frac{n \log n}{\log n + \log (1-\rho)} \leq t'(n\rho-1). \end{equation} Let $k = \lfloor \rho n\rfloor$ and consider the $(n-k) \times (n-k)$ submatrix $B=A^{(k+1)}$ which is at the intersection of the last $(n-k)$ rows and columns $[n]\backslash\{m_i:1 \leq i \leq k\}$. Consider the complement submatrix $B^c$ which lies at the intersection of the first $k$ rows and columns $\{m_i:1 \leq i \leq k\}$. Since $\zeta_i$ is the largest element in the $i$-th row of $B^c$ we have $\perm B^c \geq \prod_{i=1}^k\zeta_i$, and by \eqref{eq:distribution_of_max_in_rows} we have that \begin{equation}\label{eq:lower_bound_on_the_extreacted_part} \mathbb{P}\Big(\frac{\log \perm B^c}{\beta n \log n} \leq t\Big) \leq \mathbb{P}\Big(\sum_{i=1}^kQ_i \leq tn\log n\Big)\leq \mathbb{P}\Big(\sum_{i=1}^k\frac{Q_i}{\log(n-i+1)} \leq \frac{t n \log n}{\log ((1-\rho)n)}\Big). \end{equation} Now using \eqref{eq: HelpBound} we see that if the event under the probability on the right hand side happens, then $Q_i \leq t'\log(n-i+1)$ happens for some $1 \leq i \leq k$. Therefore \eqref{eq:lower_bounds_estimates_on_almost_exponential} and the sub-additivity imply that for any $\epsilon > 0$ such that $t'(1+\epsilon < 1$ we have \begin{equation}\label{eq: ThirdBound} \mathbb{P}\Big(\frac{\log \perm B^c}{\beta n \log n} \leq t\Big)\leq \sum_{i=1}^k\mathbb{P}\Big(\frac{Q_i}{\log(n-i+1)} \leq t'\Big) \leq \rho n\exp\bigl(-(1-\rho)^{1-t'(1+\epsilon)}n^{1-t'(1+\epsilon)}\bigr), \end{equation} for $n$ large enough. Since the right hand side is summable in $n$ we see that almost surely $\frac{\log \perm B^c}{ n \log n} \leq \beta t$ happens for only finitely many integers $n$. On the other hand, Lemma \ref{lemma:lower_bound_for_submatrices} implies that for any $\delta > 0$ there is $r>0$ such that almost surely \[ \perm B \geq r^{n-k}(n-k)^{(1-\delta)(n-k)} \geq \Big(r(1-\rho)^{1-\delta}\Big)^{(1-\rho)n}n^{(1-\delta)(1-\rho)n}. \] Since $A_n$ has positive elements, the inequality $\perm A_n \geq \perm B \perm B^c$ holds and thus almost surely for $n$ large enough \[ \frac{\log \perm A_n}{n \log n} \geq \beta t + (1-\delta)(1-\rho) + \frac{(1-\rho)(\log r + (1-\delta)\log(1-\rho)}{\log n}. \] Taking the limit as $n \to \infty$ and then $t \uparrow 1$ (which forces $\rho \uparrow 1$) yields the claim. \begin{comment} This is quite obvious since we can surely find some constant $C$ such that $\displaystyle \rho ne^{-(1-\rho)^{1-t'}n^{1-t'}} \leq \frac{C}{n^2}$. From here Borel-Cantelli lemma implies that almost surely $\displaystyle \liminf_{n \to \infty}\frac{\max_{\pi \in S_n} \sum_{i = 1}^n Y_{i,\pi(i)}}{n \log n} \geq t$ for any $t<1$. Since $t<1$ was arbitrary we obtain the lower bound in (\ref{eq: MaxPerm}). \end{comment} \begin{comment} \begin{remark} Notice that the expressions on the right hand side of ... are summable in $n$. This in fact gives us stronger statement that $\lim_{n \to \infty} \frac{\sum_{i=1}^n Y_{i, \pi(i)}}{n \log n} = 1$ almost surely. \end{remark} \end{comment} \end{proof} \begin{comment} \begin{lemma} \label{prop: BoundMaxY} Let $(Y_{i,j})_{i,j}$ be independent exponential random variables with rate 1. Then \begin{equation} \label{eq: BoundMaxY} \lim_{n \to \infty}\mathbb{P}\left(\max_{\pi \in S_n}\sum_{i=1}^n Y_{i,\pi(i)} > n\log n + \frac{n \log \log n}{2}\right) = 0. \end{equation} \end{lemma} \begin{proof} Denote $R_{n,i} := \frac{\max_{1 \leq j \leq n}Y_{i,j}}{\log n}$ so $\max_{\pi \in S_n} \sum_{i=1}^n Y_{i,\pi(i)} \leq \log n \sum_{i=1}^n R_{n,i}$ and \begin{align*} \mathbb{P} \left(\max_{\pi \in S_n}\sum_{i=1}^n Y_{i,\pi(i)} > n\log n + \frac{n \log \log n}{2}\right) & \leq \mathbb{P}\left(\sum_{i=1}^n R_{n,i} > n + \frac{n \log \log n}{2\log n}\right) \\ & \leq \mathbb{E}\left(e^{\left(\sum_{i=1}^nR_{n,i} - n - \frac{n \log \log n}{2\log n}\right)}\right) \\ & = \left(\frac{\mathbb{E}\left(e^{R_{n,1}}\right)}{e(\log n)^{\frac{1}{2 \log n}}}\right)^n. \end{align*} To estimate $\mathbb{E}\left(e^{R_{n,1}}\right)$ we calculate for $t>1$ and $n \geq 2$ \begin{multline} \label{eq: ProbR} \mathbb{P} \left(e^{R_{n,1}} \geq t\right) = \mathbb{P}(R_{n,1} \geq \log t) = 1- \mathbb{P} (\max_{1 \leq j \leq n}Y_{1,j} < \log n \log t) \\ = 1-(1-e^{-\log n \log t})^n = 1- (1-\frac{1}{t^{\log n}})^n \leq \frac{n}{t^{\log n}}. \end{multline} Here we used the inequality $(1-x)^n \geq 1-nx$ for all positive integers $n$ and all $0 < x < 1$. Now, assuming $n \geq 3$, the expectation is estimated as $$ \mathbb{E}\left(e^{R_{n,1}}\right) = \int_0^{\infty}\mathbb{P}\left(e^{R_{n,1}} \geq t\right) dt \leq e+\int_e^{\infty}\frac{n}{t^{\log n}} dt = e+\frac{n}{\log n -1}\frac{1}{e^{\log n-1}} = e \left(1 + \frac{1}{\log n -1}\right) \leq e^{1+\frac{1}{\log n- 1}}. $$ Plugging this back into we get $$ \mathbb{P} \left(\max_{\pi \in S_n}\sum_{i=1}^n Y_{i,\pi(i)} > n\log n + \frac{n \log \log n}{2}\right) \leq \left(\frac{\mathbb{E}\left(e^{R_{n,1}}\right)}{e(\log n)^{\frac{1}{2 \log n}}}\right)^n \leq \left(\frac{e^{\frac{1}{\log n -1}}}{(\log n)^{\frac{1}{2 \log n}}}\right)^n = e^{\left(\frac{1}{\log n -1} - \frac{\log \log n}{2 \log n}\right)n} $$ Since $\frac{1}{\log n -1} - \frac{\log \log n}{2 \log n} \to -\infty$ as $n \to \infty$ the right hand side of (\ref{}) converges to $0$ which proves the proposition. \end{proof} \end{comment} \section{Non-square matrices and the necessity of \eqref{eq:heavy_tail_assumption}}\label{sec:non-square} In this section we sketch how the above arguments extend to a large class of rectangular matrices. We will still assume that elements are sampled independently from a distribution supported on $\mathbb{R}^+$, but will now allow the width of the matrix to be significantly larger than the height, in particular it will suffice for the height to grow polynomially in the logarithm of the width. The precise condition under the method extends is that matrix $A_n$ is $m_n \times n$, that is has height $m_n$ and width $n$, and the height satisfies the condition \begin{equation}\label{eq:condition_on_the_height} \liminf_{n}\frac{m_n \log \log n}{(\log n)^2} > 1. \end{equation} Observe that for an $m \times n$ matrix with iid elements of mean $\mu$ we have $\mathbb{E}(\perm A_n) = \binom{n}{m} m! \mu^m$ which demonstrates that the scaling function $n \log n$ will have to be replaced by $m_n \log n$. \begin{theorem}\label{thm:main-rectangular} Let $\xi$ be a positive random variable satisfying \eqref{eq:heavy_tail_assumption} for some $\beta > 0$. If $(A_n)_n$ is a sequence of $m_n \times n$ matrices on a common probability space with elements which are independent and identically distributed as $\xi$ and satisfying \eqref{eq:condition_on_the_height}, then almost surely \begin{equation}\label{eq:main_result_rectangular} \lim_{n \to \infty} \frac{\log \perm A_n}{m_n \log n} = \max(1,\beta). \end{equation} \end{theorem} The uniformity over all submatrices of linear size holds as well. \begin{theorem}\label{thm:general_bounds-rectangular} Assume that $(A_n)$ is a sequence of $m_n \times n$ matrices on a common probability space with elements which are independent and identically distributed as $\xi$ and satisfying \eqref{eq:condition_on_the_height}, and let $0 < \alpha < 1$. \begin{itemize} \item[i)] We have \begin{equation}\label{eq:general_lower_bounds_rectangular} \liminf_{n \to \infty} \min_{(k_1, k_2,B)}\frac{\log \perm B}{k_1 \log k_2} \geq 1, \end{equation} where the minimum is taken over all pairs of integers $(k_1, k_2)$ satisfying $\alpha m_n \leq k_1 \leq m_n$, $\alpha n \leq k_2 \leq n$ and $k_1 \leq k_2$ and all $k_1 \times k_2$ submatrices $B$ of $A_n$. \item[ii)] If $\xi$ has a finite mean then \begin{equation}\label{eq:general_upper_bounds_rectangular} \limsup_{n \to \infty} \max_{(k_1, k_2, B)}\frac{\log \perm B}{k_1 \log k_2} = 1, \end{equation} where the minimum is taken over all pairs of integers $(k_1,k_2)$ satisfying $\alpha m_n \leq k_1 \leq m_n$, $\alpha n \leq k_2 \leq n$ and $k_1 \leq k_2$ and all $k_1 \times k_2$ submatrices $B$ of $A_n$. \end{itemize} \end{theorem} The proofs of these theorems are modifications of the arguments in the previous two sections. We will briefly sketch how the modifications go and at which points one needs to do a more careful analysis. Note that, to simplify notation, we will drop the ceiling and the floor notation throughout the section. \begin{proof}[Sketch of the proof of the upper bounds in Theorem \ref{thm:main-rectangular}] As before, by stochastic domination, it suffices to prove the claim when elements are Pareto distributed. To end this one needs to prove a version of Lemma \ref{lemma: MaxPerm} which states that when $(Y_{i,j})$ are independent exponentially distributed with rate one and $\lambda > 1$, we have \[ \sum_{n = 2}^{\infty}\mathbb{P}\left(\max_{\pi \in S_{m_n,n}}\sum_{i=1}^{m_n} Y_{i,\pi(i)} \geq m_n\log n + m_n \frac{\log\log n}{\lambda}\right) < \infty. \] Proceeding as in the proof of Lemma \ref{lemma: MaxPerm} one is left to show that \[ \sum_{n=2}^\infty\exp\biggl(\Big(\frac{1}{\log n -1} - \frac{\log \log n}{\lambda \log n}\Big)m_n\biggr) < \infty, \] for some $\lambda > 1$. By \eqref{eq:condition_on_the_height} for $\lambda > 1$ small enough the expression in the exponent is not larger than $-\alpha \log n$ for some $\alpha > 1$ which yields the claim. Next one defines the analog of \eqref{eq:number_in_interval} as \begin{equation}\label{eq:Z_for_rectangular} Z_{n,k}=\bigg|\biggl\{\pi \in S_{m_n,n} : (k-1)m_n \leq \sum_{i=1}^{m_n}Y_{i,\pi(i)} < km_n \biggr\}\biggr|, \end{equation} and needs to prove \eqref{eq: BoundZMax}. Calculating expectation of $Z_{n,k}$ as in \eqref{eq: Z mean} yields \begin{equation}\label{eq:expectation_of_Z_rectangular} \binom{n}{m_n}e^{-km_n}m_n^{m_n}(k^{m_n} - (k-1)^{m_n}) \leq \mathbb{E}(Z_{n,k}) \leq \binom{n}{m_n}e^{-(k-1)m_n}(km_n)^{m_n}. \end{equation} For $k=1$ this yields $\mathbb{E}(Z_{n,1})\geq \binom{n}{m_n}m_n!e^{-m_n}$ which handles the sum $\sum_{n \geq 2}\mathbb{E}(Z_{n,1})^{1-\gamma}$. We are left to prove the analog of \eqref{eq: BoundUp}, that \[ \sum_{2 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{1}{ \mathbb{E}(Z_{n,k})^{\gamma-1}} \leq \frac{e^{m_n (\gamma-1)}}{\binom{n}{m_n}^{\gamma-1}m_n^{m_n(\gamma-1)}} \sum_{1 \leq k \leq \log n + \frac{\log \log n}{\lambda}} \frac{e^{km_n(\gamma-1)}}{k^{m_n (\gamma-1)}} \] is summable in $n$. Again by the convexity of $g(t)=e^{tm_n(\gamma-1)}t^{-m_n (\gamma-1)}$ and the fact that $g(1) \leq g(\log n + \frac{1}{\lambda} \log \log n)$, proceeding as before one is left to prove that \[ \left(\frac{ne}{m_n (\log n)^{1-1/\lambda}}\right)^{m_n(\gamma-1)}\frac{\log n + \frac{1}{\lambda}\log \log n}{\binom{n}{m_n}^{\gamma-1}} \] is summable in $n$. Since $(\log n)^{-\kappa m_n}$ is summable, for any $\kappa > 0$, it is enough to show that $\binom{n}{m_n} \geq (cn/m_n)^{m_n}$, for some $c> 0$. Taking $\log$s on both sides and using \eqref{eq:stirling_for_binom} (we can assume that $m_n < n$) it is enough to show that \[ (n-m_n +1/2) (\log n - \log (n-m_n)) \geq m_n \log c + \frac{1}{2}\log m_n, \] which holds for $c< 1$ and $n$ large enough (since the left hand side is positive and the right hand side negative). This proves \eqref{eq: BoundZMax} for \eqref{eq:Z_for_rectangular}. To finish the proof assume that both \begin{align*} & \max_{\pi \in S_{m_n,n}}\sum_{i=1}^{m_n}Y_{i,\pi(i)} \leq m_n \log n + \frac{m_n \log\log n}{\lambda} \ \ {\rm and} \\ & Z_{n,k} \leq \mathbb{E}(Z_{n,k})^{\gamma}, \text{ for each } 1 \leq k \leq \log n + \frac{\log \log n}{\lambda}, \end{align*} hold for some $\lambda > 1$, which is true for $n$ large enough almost surely. The same calculations as in the proof of of the upper bounds in Theorem \ref{thm:main} and \eqref{eq:expectation_of_Z_rectangular} yield \[ \perm A \leq \binom{n}{m_n}^{\gamma}n^{(\beta-\gamma)m_n}m_n^{m_n\gamma}(\log n)^{\frac{\beta - \gamma}{\lambda}m_n}e^{m_n\gamma}\Big(\log n + \frac{1}{\lambda}\log \log n\Big)^{\gamma m_n} \] and \[ \frac{\log \perm A}{m_n \log n} \leq \gamma\frac{\log \binom{n}{m_n}}{m_n \log n} + \beta -\gamma + \gamma \frac{\log m_n}{\log n}+ \frac{\beta - \gamma}{\lambda}\frac{\log \log n}{\log n} + \frac{\gamma}{\log n} + \gamma \frac{\log (\log +\frac{1}{\lambda} \log \log n)}{\log n}. \] The last three terms vanish in the limit and so it remains to prove \[ \limsup_{n \to \infty}\Big(\frac{\log \binom{n}{m_n}}{m_n \log n} + \frac{\log m_n}{\log n}\Big) \leq 1. \] After applying \eqref{eq:stirling_for_binom} we are left with \[ \limsup_{n} \Big(\frac{n}{m_n}-\Big(\frac{n}{m_n} -1\Big)\frac{\log (n-m_n)}{\log n}\Big) \leq 1. \] which follows from \eqref{eq:useful_lowerbound}. For $\beta \leq 1$ one can repeat the calculations, or simply refer to stochastic domination. \end{proof} The proof of Theorem \ref{thm:general_bounds-rectangular} is based on the following equivalent of Lemma \ref{lemma:lower_bound_for_submatrices}. \begin{lemma}\label{lemma:lower_bound_for_submatrices_rectangular} Let $\xi$ be a positive random variable and let $A_n$ be a sequence of $m_n \times n$ matrices whose elements are independent and identically distributed as $\xi$ and which height $m_n$ satisfies \eqref{eq:condition_on_the_height}. For any $0 < \alpha < 1$ and any $\delta > 0$ there exists $r>0$ with the following property: Almost surely there exists $n_0$ such that for any $n \geq n_0$ and any pair of integers $(k_1,k_2)$ satisfying $\alpha m_n \leq k_1 \leq m_n$, $\alpha n \leq k_2 \leq n$, and $k \leq k_2$, any $k_1 \times k_2$ submatrix $B$ of $A_n$ satisfies $\perm B \geq r^{k_1} k_2^{(1-\delta)k_1}$. \end{lemma} \begin{proof}[Sketch of the proof of Lemma \ref{lemma:lower_bound_for_submatrices_rectangular}] One follows the proof of Lemma \ref{lemma:lower_bound_for_submatrices}. In the definitions one needs to write $n$ for the width of the matrix and $m_n$ for the height, for example $\mathcal{C}_n$ is defined as the event that for some pair of integers $(k_1,k_2)$ satisfying $1 \leq k_1 \leq m_n$, $1 \leq k_2 \leq n$ and $k_1 + k_2 \geq \alpha n$ some $k_1 \times k_2$ submatrix of $A_n$ contains only zeros, and $\mathcal{B}_n$ is defined as before. On $\mathcal{B}_n^c \cap \mathcal{C}_n^c$ one has \[ \perm B \geq \left\{\begin{array}{ll} q^{k_1} \lfloor(1-\delta)k_2 \rfloor!, & \text{ for } k_1 \geq (1-\delta)k_2\\ q^{k_1} \frac{\lfloor(1-\delta)k_2 \rfloor!}{(\lfloor(1-\delta)k_2 \rfloor - k_1)!}, & \text{ for } k_1 <(1-\delta)k_2. \end{array}\right. \] In the first case the logarithm of the right hand side will be bounded from below by \[ k_1 \log q + (1-\delta)k_2 \log ((1-\delta)k_2) - (1-\delta)k_2 \geq k_1 (\log q -1 + \log(1-\delta)) + (1-\delta)k_1\log k_2, \] and in the second \begin{multline*} k_1 \log q + ((1-\delta)k_2+1/2)\log ((1-\delta)k_2) - ((1-\delta)k_2-k_1+1/2)\log ((1-\delta)k_2 - k_1) + k_1 \\ \geq k_1\log ((1-\delta)k_2) + k_1(1+ \log q), \end{multline*} which is sufficient in both cases. Probability of the event $\mathcal{B}_n$ is estimated as before. For $\mathcal{C}_n$ one can follow the arguments in \eqref{eq:large_deviations_for _rectangles} starting with \[ \mathbb{P}(\frak{C}_n) \leq 2\sum_{\alpha n/2 \leq k_2 \leq n} \binom{n}{k_2} \sum_{1\leq k_1 \leq m_n}\binom{m_n}{k_1}\eta^{k_1k_2}. \] This inequality follows from the simple fact that $m_n \leq n$ and $k_1 \leq k_2$ imply that $\binom{m_n}{k_2}\binom{n}{k_1} \leq \binom{m_n}{k_1}\binom{n}{k_2}$. \end{proof} \begin{proof}[Sketch of the proof of Theorem \ref{thm:general_bounds-rectangular}] Lower bounds follow from Lemma \ref{lemma:lower_bound_for_submatrices_rectangular}. For the upper bounds again use Markov's inequality and reduce to the case when elements of $A_n$ are parameter 1 Pareto distributed. Similarly as before observe that for any $k_1 \times k_2$ submatrix $B$, any term in the sum defining $\perm B$ can be expanded in $\binom{n-k_1}{m_n - k_1}(m_n-k_1)!$ ways to a term in the sum defining $\perm A_n$. Since all elements of $A_n$ are greater or equal than $1$ we have \begin{multline*} \log \perm B \leq \log \perm A_n - \log \left(\binom{n-k_1}{m_n - k_1}(m_n-k_1)!\right) \\ \leq \log \perm A_n - (n-k_1+1/2)\log(n-k_1) + (n-m_n+1/2)\log (n-m_n) + (m_n-k_1). \end{multline*} Terms $m_n - k_1$, $\log (n-m_n)$ and $\log (n-k_1)$ are $o(k_1 \log k_2)$. After disregarding them, by the proven upper bounds in Theorem \ref{thm:main-rectangular}, almost surely for any $\epsilon > 0$ one has for $n$ large enough \[ \frac{\log \perm B}{k_1 \log k_2} \leq \frac{\log n}{\log k_2}\left(\epsilon\frac{m_n}{k_1} + \frac{m_n \log n + (n-m_n)\log(n-m_n) - (n-k_1)\log(n-k_1)}{k_1 \log n}\right). \] We are left to prove that \begin{equation}\label{eq:left_to_prove} - (n-k_1) \log(n-k_1) + (n-m_n)\log(n-m_n) + (m_n-k_1)\log n \leq o(m_n \log n), \end{equation} uniformly in $\alpha m_n \leq k_1 \leq m_n$. Since the left hand side (as a function of $k_1$) is increasing on $(0, n(1-1/e))$ and decreasing on $(n(1-1/e),1)$ it suffices to prove \eqref{eq:left_to_prove} for $k_1 =m_n$ when $m_n \leq n(1-1/e)$ and for $k_1 = n(1-1/e)$ when $n(1-1/e) < m_n \leq n$ (assuming $\alpha < 1-1/e$ which is of no loss of generality). For $k_1 = m_n$ the left hand side is $0$ so there is nothing to prove, and for $k_1 = n(1-1/e)$ the left hand side is equal to \[ \frac{n}{e} - (n-m_n)\big(\log n - \log (n-m_n)\big), \] which is not larger than an $o(n \log n)$ term, and this suffices in the case $n(1-1/e) < m_n \leq n$. \end{proof} \begin{proof}[Sketch of the proof of the lower bounds in Theorem \ref{thm:main-rectangular}] The lower bounds for $\beta \leq 1$ case follow directly from Theorem \ref{thm:general_bounds-rectangular} i). For the case $\beta > 1$ one can follow the arguments almost verbatim. In \eqref{eq:lower_bounds_estimates_on_almost_exponential} one needs to replace $\rho n$ by $\rho m_n$ in the upper limit in the sum and in the factor in the end. Moreover the greedy algorithm is run for $k=\rho m_n$ steps, where $\rho$ is defined as in \eqref{eq: HelpBound} with $n$ replaced by $m_n$ (except under $\log$s). One defines the submatrices $B$ and $B^c$ as before and arrives at the analog of \eqref{eq: ThirdBound} \[ \mathbb{P}\Big(\frac{\log \perm B^c}{\beta m_n \log n} \leq t\Big)\leq \sum_{i=1}^k\mathbb{P}\Big(\frac{Q_i}{\log(n-i+1)} \leq t'\Big) \leq \rho m_n\exp\bigl(-(1-\rho)^{1-t'(1+\epsilon)}n^{1-t'(1+\epsilon)}\bigr). \] To finish the proof apply Lemma \ref{lemma:lower_bound_for_submatrices_rectangular} on $B$ identically as Lemma \ref{lemma:lower_bound_for_submatrices} in Section \ref{sec:lower_bounds}. \end{proof} This example shows that Theorem \ref{thm:main} in general fails when the limit in \eqref{eq:heavy_tail_assumption} does not exist. Actually this is possible at arbitrary small oscillations of the sequence in \eqref{eq:heavy_tail_assumption}. We present the argument for square matrices. \begin{example}\label{ex:no_convergence} Let $S=\{k_i\}$ be a set of positive integers labeled so that $k_{i+1} > 2k_i$. Fix $C_2 > C_1 > \lambda > 1$ and for every $k \geq 1$ define the following sequences of positive real numbers \[ t_k = \exp(\lambda^k), \ \tilde{p}_k' = \exp(- \lambda^k/C_1), \text{ and } p_k' = \left\{\begin{array}{ll}\exp(- \lambda^k/C_1), & \text{ for }k \notin S \\ \exp(- \lambda^k/C_2), & \text{ for } k \in S. \end{array} \right. \] Clearly both series $\sum_k p_k'$ and $\sum_k \tilde{p}_k'$ converge, so we can normalize the sequences with the its sums $Z$ and $\tilde{Z}$ respectively and obtain sequences $p_k = p_k'/Z$ and $\tilde{p}_k = \tilde{p}_k'/\tilde{Z}$. Let $\xi$ and $\tilde{\xi}$ be random variables supported on the set $\{t_k\}$ with distributions $\mathbb{P}(\xi = t_k) = p_k$ and $\mathbb{P}(\tilde{\xi} = t_k) = \tilde{p}_k$. Observing that the mappings $t \mapsto \mathbb{P}(\xi \geq t)$ and $t \mapsto \mathbb{P}(\tilde{\xi} \geq t)$ are constant on $(t_k , t_{k+1}]$ and that $\mathbb{P}(\xi \geq t_k) \leq 2p_k$, for all $k \in S$ large enough and for infinitely many $k \notin S$ as well, and $\mathbb{P}(\tilde{\xi} \geq t_k) \leq 2\tilde{p}_k$, for all $k$ large enough, it is easy to see that \begin{align} &\liminf_{t \to \infty} \frac{\log \mathbb{P}(\xi \geq t)}{\log t} = -\frac{1}{C_1/\lambda} < -\frac{1}{C_2} = \limsup_{t \to \infty} \frac{\log \mathbb{P}(\xi \geq t)}{\log t}, \label{eq:example_bounds_1} \\ &\liminf_{t \to \infty} \frac{\log \mathbb{P}(\tilde{\xi} \geq t)}{\log t} = -\frac{1}{C_1/\lambda} < -\frac{1}{C_1} = \limsup_{t \to \infty} \frac{\log \mathbb{P}(\tilde{\xi} \geq t)}{\log t} \label{eq:example_bounds_2}. \end{align} As usual let $(A_n)$ denote a sequence of $n \times n$ matrices on a common probability space with independent elements distributes as $\xi$. We will prove that \begin{equation}\label{eq:example_statement} \liminf_n \frac{\log \perm A_n}{n \log n} \leq C_1, \text{ and } \limsup_n \frac{\log \perm A_n}{n \log n} = C_2 \end{equation} To get the upper bound on $\liminf$ take a sequence $(\ell_i)$ of positive integers such that $k_i < \ell_i < k_{i+1}$ and that sequences $(\ell_i - k_i)$ and $(k_{i+1} -\ell_i)$ are strictly increasing. Define integers $n_i = \exp(\lambda^{\ell_i}/C_1)$. By a simple union bound the probability that $A_n$ contains an element $t_\ell$, for some $\ell \geq k_{i+1}$ is bounded from above by \[ 2n_i^2p_{k_{i+1}} =\frac{1}{Z}\exp(2\lambda^{\ell_i}/C_1) 2\exp(-\lambda^{k_{i+1}} /C_2), \] for $i$ large enough. This expression is summable in $i$, so almost surely $A_{n_i}$ does not contain elements greater than $t_{k_{i+1}}$ for $i$ large enough. Thus to prove the first inequality in \eqref{eq:example_statement} one can assume that elements in $A_{n_i}$ are distributed as $\xi\mathbf{1}_{(\xi < t_{k_{i+1}})}$. Next observe that $\xi\mathbf{1}_{(\xi < t_{k_{i+1}})}$ is stochastically dominated by $t_{k_i}\tilde{\xi}$, that is \[ \mathbb{P}(t \leq \xi < t_{k_{i+1}}) \leq \mathbb{P}(t_{k_i}\tilde{\xi} \geq t). \] While the inequality is trivial for $t\geq t_{k_{i+1}}$ and for $t \leq t_{k_i}$, for $t_{k_i} < t < t_{k_{i+1}}$ it follows from the fact that for $J \subset (t_{k_i},t_{k_{i+1}})$ \[ \mathbb{P}(\xi \in J) =\frac{\sum_{k : t_k\in J}p_k'}{Z} \leq \frac{\sum_{k : t_k\in J}\tilde{p}_k'}{\tilde{Z}} = \mathbb{P}(\tilde{\xi} \in J), \] since $p_k' = \tilde{p}_k'$, for $t_k \in J$ and $\tilde{Z} \leq Z$. Thus if $\tilde{A}_n$ is the sequence of $n \times n$ matrices whose elements are identical and distributed as $\tilde{\xi}$ then \[ \liminf_{n} \frac{\log \perm A_n}{n \log n} \leq \lim_i \frac{n_i\log t_{k_i} + \log \perm \tilde{A}_{n_i}}{n_i \log n_i} \leq \lim_i \frac{\lambda^{k_i}}{\lambda^{\ell_i}/C_1} + C_1 = C_1. \] Here the second inequality follows form the upper bounds in Theorem \ref{thm:main} and \eqref{eq:example_bounds_2}. To prove the second relation in \eqref{eq:example_statement} fix $k\in S$, $\epsilon > 0$ and define the integer $n = n_k = \exp((1+\epsilon)\lambda^k/C_2)$. We proceed with a greedy algorithm analogous to the one in the proof of the lower bound in Theorem \ref{thm:main}. With the probability $1- (1-p_k)^n$ there is an element in the first row of $A^{(0)} = A_n$ equal to $t_k$. On this event take the first such element, remove the corresponding column and the first row from $A_n$ and obtain the $(n-1) \times (n-1)$ matrix $A^{(1)}$ which is independent of the first row and is distributed as $A_{n-1}$. Now repeat the step with $A^{(1)}$ instead of $A^{(0)}$ and proceed recursively as long as one is successful at each step. For $0 < \rho < 1$ one will not be able to proceed till step $\rho n $ with probability at most \[ \sum_{(1-\rho) n \leq i \leq n}(1-p_k)^i \leq \frac{(1-p_k)^{(1-\rho)n}}{p_k} \leq 3Z\exp\Big(-\frac{1-\rho}{Z}\exp(\epsilon\lambda^k/C_2)+\lambda^k/C_2\Big), \] if $k$ is chosen large enough. The right hand side is clearly summable in $k$ and thus almost surely for $k$ large enough and $n=n_k$ constructed as above, the above algorithm will be successful for $\rho n$ steps. In that case one can get lower bound on $\perm A_n$ as in the proof of the lower bounds in Theorem \ref{thm:main}: The matrix at the intersection of the first $\rho n$ rows and the removed columns is bounded from below by the product of extracted elements, that is $t_k^{\rho n}$ and the matrix at the intersection of the last $(1-\rho)n$ rows and non-removed columns is bounded from below by $((1-\rho) n)!$ (since all of it's elements are greater or equal than $1$). Therefore for $k$ large enough and $n$ constructed as above $\perm A_n \geq t_k^{\rho n} ((1-\rho) n)!$. Since $\log ((1-\rho) n)!/(n \log n) \to 1-\rho$ and \[ \frac{\log t_k^{\rho n}}{n \log n} \geq \frac{\rho\log t_k}{\log n} \geq \frac{\rho \lambda^k }{(1+\epsilon)\lambda^k /C_2} \to \frac{C_2 \rho}{1+\epsilon}, \] we obtain that almost surely \[ \limsup_{n \to \infty} \frac{\log \perm A_n}{n \log n} \geq \frac{C_2 \rho}{1+\epsilon} + 1- \rho. \] By sending $\rho \to 1$ and $\epsilon \to 0$ we get $C_2$ as the lower bound on the $\limsup$, and by Theorem \ref{thm:main} and \eqref{eq:example_bounds_1} it is equal to $C_2$. \end{example} \textbf{Acknowledgments} The author would like to thank Professor Sourav Chatterjee for suggesting this problem and for generous help and guidance with this paper.
\section{Introduction} This paper deals with a century-old issue, the so called ``radiation reaction" force, in classical electrodynamics and in general relativity. On one hand, it is clear that energy and momentum carried away by radiation from an accelerated charge should be compensated by a force which is going to reduce the energy and momentum of the particle accordingly. On the other hand, the particle in 3+1 flat space-time does not interact with its own field because it is fully concentrated on the light cone. The relation to the energy loss in the nonrelativistic dipole radiation and its relativistic extension leads to the well known Abraham-Lorentz-Dirac force \begin{eqnarray} f^\mu_{4D} = {2e^2 \over 3}\left(\dddot{x}^\mu - \ddot{x}^\nu \ddot{x}_\nu \dot{x}^\mu\right) \quad,\label{4d} \end{eqnarray} where the dot is a derivative over the proper time ${d/d\tau}$. Its derivation comes from ``the large-distance" discussion, based on the amount of the energy/momentum fluxes through some distant surface (large sphere, etc). Although in principle such an approach can/was applied for scalar/electromagnetic/gravitational radiations, in some cases it is technically difficult. In particular, the radiation and its corresponding ultrarelativistic sources move through nearby paths in curved spaces, making the radiation calculation highly non-trivial (see e.g. \cite{Khriplovich:1973qf}) . It would be more satisfactory logically and much easier practically to use some local derivation. In this paper we provide examples in support of a local derivation. We stumbled on this issue while trying to assess the braking force for a gravitational radiation of an ultrarelativistic particle moving in a particular curved space (the so called thermal or black-hole AdS$_5$ in 4+1 dimensions). This problem is related to the practically important problem of jet quenching in the quark-gluon plasma. Since the calculations are very different in AdS$_5$ we relegate the analysis of jet quenching by braking radiation to the companion paper~\cite{Shuryak:2011ge}. The issue we study in this paper is whether one can define and calculate {\it the self-interaction force}, using only the particle's own field related to its past trajectory. This question was first addressed by Dirac for electrodynamics in flat 3+1 dimensions \cite{DIRAC} and extended to curved 3+1 dimensions by Dewitt~\cite{DEWITT}. In this work we show that this is also possible for electrodynamics in odd dimensional space-times, namely 2+1 and 4+1 dimensions. The idea that it is possible was inspired by the so called MSTQW approach in gravitational setting \cite{Mino:1996nk,Quinn:1996am}. In it the remarkable ``self-force" expression was suggested \begin{eqnarray} ma^a &=& m^2 u^b u^c \int_{-\infty}^{\tau^-}d\tau' u^{a'} u^{b'} ( {1\over 2} \nabla^a G_{bca'b'} \\ \nonumber && - \nabla_b G_{c\,\,a'b'}^{\,\,a} -{1 \over 2} u^a u^d \nabla_d G_{bca'b'}, ) \end{eqnarray} in which the integral is done over proper time and the past world line of the particle till the regulated present time $\tau^-$. $G$ is the retarded Green function for the Einstein equation with the particle as the source. Note that the bracket is just the Chrystoffel force for a gravity perturbation, induced by the past gravity field of the particle itself. Does this expression (or its simpler analogs) actually work? The first obvious try, for electromagnetic radiation in 3+1 flat space-time, produces zero because in flat 4d space-time the retarded propagator is totally localized on the light cone $(x^\alpha-r^\alpha)(x_\alpha-r_\alpha)=0$. There are simply no points on the particle path that can intersect the past light cone sustained by its present location! (It produces the well-known Lienard-Wiechert expression for the retarded fields we all learned/taught in classical E/M courses). This upset can be remedied by realizing that such form of the retarded propagator is not the generic case, and in fact it is nonzero inside the light cone in other space-times. Therefore, we decided to calculate it for the nearby space-time dimensions and see if the results make sense. We also calculate the radiative losses by standard large-distance method for comparison. We assume that the motion of the charge is prefixed by some external non-electromagnetic forces and do not discuss its origin. We will define and compute the local back-reaction force on a particle originating from its own electromagnetic fields. (Of course, this force should be a part of the total force that determines the given charge trajectory, but we do not specify its effect on the path ). Although the self-force is in general expected to be nonlocal in character, and given by the integral over the past trajectory, the situation is simplified in the highly relativistic limit, $\gamma\to\infty$. Indeed we will see that in this case it is defined by a small range in the proper time interval $\tau\sim {1\over\gamma^2}$, so that the leading $\gamma$-contribution depends only on the local data of the motion. The resulting leading-$\gamma$ behavior of the self-force takes a similar form as in (\ref{4d}) in terms of local derivatives of the motion, which would be the odd-dimensional analogue of the Abraham-Lorentz-Dirac force in 3+1 dimensions. Another feature is that the integrals with the retarded propagator diverge and need to be correctly regulated: we will see this feature explicitly in our analysis shortly. \section{The self-force in 2+1 dimensions \label{sec21}} The case of 2+1 dimensions is the simplest one to consider. (Below we will see that it contains all the basic features of the radiation reaction force in any odd space-time dimensions because of some recursive relations between the propagators involved.). As already mentioned in the introduction, the main new feature in odd spacetimes compared to the conventional flat 3+1 dimensional space-time is that the massless retarded propagator from a given source has support $inside$ the light-cone. This implies that at a given moment/position of the charge, the electromagnetic field acting on it obtain some contributions from the past trajectory of the charge. A moving charge $e$ with a given trajectory $x^\mu(\tau)$ gives a relativistic current \begin{eqnarray} j^\mu(x)=e\int d\tau\, \delta^{(3)}\left(x-x(\tau)\right) \dot{x}^\mu(\tau)\quad,\label{current} \end{eqnarray} where $\tau$ is the proper time normalized as $\dot{x}^\mu \dot{x}_\mu =+1$. In the covariant gauge $\partial_\mu A^\mu=0$, the Maxwell equation becomes \begin{eqnarray} \Box A^\mu= j^\mu\quad,\quad \Box\equiv\partial_\mu\partial^\mu\quad, \end{eqnarray} which has a formal retarded solution as \begin{eqnarray} A^\mu(x)&=&\int d^3x'\, \Delta_R(x-x')j^\mu(x')\nonumber\\ &=&e\int d\tau'\,\Delta_R\left(x-x(\tau')\right)\dot{x}^\mu(\tau')\quad,\label{ret} \end{eqnarray} using (\ref{current}). $\Delta_R(x)$ is the massless retarded propagator in 2+1 dimensions which is given by \begin{eqnarray} \Delta_R(x)={\theta(x^0)\over 2\pi}{\theta(x^2)\over\sqrt{x^2}}\quad,\quad x^2\equiv x^\mu x_\mu\quad,\label{prop} \end{eqnarray} which has support in the entire forward light-cone. It is clear that only the past trajectory contributes to the electromagnetic field at a given moment, and we are interested in the self-force acting on the moving charge itself. This amounts to computing the covariant Lorentz force \begin{eqnarray} f^\mu_{3D}=e F^{\mu\nu}\left(x(\tau)\right) \dot{x}_\nu(\tau) \quad,\label{3d} \end{eqnarray} where $F^{\mu\nu}\left(x(\tau)\right)$ are the field strengths of the retarded electromagnetic field (\ref{ret}) at proper time $\tau$, induced by the past trajectory of the charge, $x^\mu(\tau')$ with $\tau'<\tau$. As usual, one encounters local divergences in computing (\ref{3d}) coming from the region near $\tau'=\tau$, and one has to regularize and absorb divergences by renormalizing physical parameters of the moving charge such as its mass. In this section, we regularize divergences covariantly by cutting off proper time integral \begin{eqnarray} \int^\tau_{-\infty} d\tau' \to \int^{\tau-\epsilon}_{-\infty}d\tau'\quad,\label{cutoff} \end{eqnarray} with a small $\epsilon>0$, and let $\epsilon\to 0$ after renormalizing the particle mass. In the appendix, we will present another regularization by averaging (\ref{3d}) over a small sphere around $x(\tau)$ of radius $r=\epsilon$ and taking $\epsilon\to 0$ after renormalization. The results are identical, which is at least a useful consistency check of our results. From (\ref{ret}), (\ref{3d}), and using the fact that \begin{eqnarray} \partial^\mu \Delta_R(x)=2 x^\mu \Delta_R'(x)+\delta^{\mu 0}{\delta(x^0)\over 2\pi}{\theta(x^2)\over\sqrt{x^2}}\quad, \end{eqnarray} where \begin{eqnarray} \Delta_R'(x)=\frac{\theta(x^0)}{2\pi}\left(\frac{\delta(x^2)}{\sqrt{x^2}}-\frac 12\,\frac{\theta(x^2)}{x^2\sqrt{x^2}}\right)\quad, \label{T6} \end{eqnarray} the unregularized bare Lorentz force is \begin{eqnarray} &&f^\mu_{3D}=2e^2\int d\tau'\,\Delta_R'\left({\bf X}\right) {\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau)\nonumber\\ &&+ e^2\delta^{[\mu 0}\int d\tau'\,{\delta\left({\bf X}^0\right)\over 2\pi}{\theta\left({\bf X}^2\right)\over \sqrt{{\bf X}^2}} \dot{x}^{\nu]}(\tau')\dot{x}_\nu (\tau)\quad,\label{unreg} \end{eqnarray} where ${\bf X}^\mu(\tau,\tau')\equiv x^\mu(\tau)-x^\mu(\tau')$ and $[\mu,\nu]=\mu\nu-\nu\mu$. To identify the local divergences near $\tau'=\tau$, we expand the quantities in terms of $(\tau-\tau')\equiv\epsilon$ as \begin{eqnarray} &&{\bf X}^\mu(\tau,\tau')=\epsilon\dot{x}^\mu(\tau) -{\epsilon^2\over 2}\ddot{x}^\mu(\tau)+{\epsilon^3\over 6}\dddot{x}^\mu(\tau)+{\cal O}(\epsilon^4)\,,\nonumber\\ &&\dot{x}^\nu(\tau')=\dot{x}^\nu(\tau)-\epsilon\ddot{x}^\nu(\tau)+{\epsilon^2\over 2}\dddot{x}^\nu(\tau)+{\cal O}(\epsilon^3)\,, \nonumber\\ && {\bf X}^2=\epsilon^2-{\epsilon^4\over 12}\ddot{x}^\mu\ddot{x}_\mu +{\cal O}(\epsilon^5)\,,\label{epsexp} \end{eqnarray} using the identities such as $\dot{x}^\mu\ddot{x}_\mu=0$ and $\dot{x}^\mu\dddot{x}_\mu=-\ddot{x}^\mu\ddot{x}_\mu$. From these, one obtains after some algebra \begin{eqnarray} &&{\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau)\\\label{expand} &&={\epsilon^2\over 2}\ddot{x}^\mu(\tau)-{\epsilon^3\over 3}\left(\dddot{x}^\mu+\ddot{x}^\nu\ddot{x}_\nu\dot{x}^\mu\right)(\tau) +{\cal O}(\epsilon^4)\quad,\nonumber \end{eqnarray} and \begin{eqnarray} \Delta_R'({\bf X})=-{\theta(\epsilon)\over 4\pi}\left({1\over\epsilon^3}-{\delta(\epsilon)\over \epsilon^2}\right) +{\cal O}\left(1\over\epsilon\right)\quad, \end{eqnarray} so that the proper time integral of the first term in (\ref{unreg}) near the $\epsilon=0$ region becomes \begin{eqnarray} -{e^2\over 4\pi}\int d\epsilon\,\theta(\epsilon)\left({1\over\epsilon}-\delta(\epsilon)\right)\ddot{x}^\mu(\tau) +{\cal O}(\epsilon^1)\quad,\label{loginf} \end{eqnarray} which is logarithmically divergent. This divergence is proportional to the covariant acceleration $\ddot{x}^\mu$ and is readily absorbed by renormalizing the mass \begin{eqnarray} m_{ren}=m_{bare}-{e^2\over 4\pi}\log(\epsilon)\quad,\label{mren} \end{eqnarray} where $\epsilon$ here means a cutoff with a slight abuse of notation. This is equivalent to having a counterterm added to (\ref{unreg}) so that the renormalized self-force now takes a form \begin{eqnarray} &&f^\mu_{3D}=-{e^2\over 2\pi}\lim_{\epsilon\to 0^+}\int d\tau'\,\theta(\tau-\tau'-\epsilon)\label{renormalized}\\&&\times\left({{\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau) \over\left({\bf X}^2\right)^{3\over 2}} -{1\over 2}{\ddot{x}^\mu(\tau)\over (\tau-\tau')}\right)\quad,\nonumber \end{eqnarray} considering only the first term in (\ref{unreg}). The finite contribution coming from the piece proportional to $\delta(\epsilon)$ in (\ref{loginf}) is ambiguous up to the value of $\theta(0)$. In fact, the second term in (\ref{unreg}) gives rise to a divergence which is also proportional to $\theta(0)$. One can choose to have $\theta(0)=0$ removing these ambiguities as a choice of regularization. Alternatively, once we introduce a cutoff $\tau-\tau'>\epsilon$ they simply don't appear for any finite $\epsilon>0$, and hence the limit $\epsilon\to 0^+$ after taking care of the logarithmic divergence via (\ref{mren}) is insensitive to them. Unlike in 3+1 dimensions where the self-force is the local Abraham-Lorentz-Dirac force (\ref{4d}), in 2+1 dimensions the self-force (\ref{renormalized}) is non-local and receives contributions from the entire past tail of the charged particle. This non-locality is due to the peculiar fact that the retarded propagator (\ref{prop}) has support on the entire forward part of the light cone as we noted earlier. However, for ultra-relativistic motion we now show that the leading $\gamma$ behavior of (\ref{renormalized}) is completely local. Let's denote the $n$-th derivative of $x^\mu$ with respect to proper time $\tau$ as $x_{(n)}^\mu\equiv {d^n x^\mu\over d\tau^n}$. In the large $\gamma$-limit, $x^\mu_{(n)}$ is of order ${\cal O}(\gamma^n)$. In general, for any combinations of $x^\mu_{(n)}$ the powers of $\gamma$ simply add up \begin{eqnarray} \Pi_i x_{(n_i)} \le {\cal O}\left(\gamma^{\sum_i n_i}\right) \quad,\label{powers} \end{eqnarray} where we allow the inequality due to exceptions of some lowest contractions such as $\dot{x}^\mu\dot{x}_\mu=1$ and $\dot{x}^\mu \ddot{x}_\mu=0$. These exceptions will not affect our derivation and conclusion, because what will be important is that the maximum powers of $\gamma$ is $\sum_i n_i$. We will prove that the leading $\gamma$ behavior of (\ref{renormalized}) is coming from the proper time interval $\tau-\tau'\equiv \epsilon\sim {1\over\gamma^2}$: we will first assume this and expand (\ref{renormalized}) in small $\epsilon$, then show the consistency of the assumption in our final results. Using (\ref{epsexp}) and (\ref{powers}), one can show the expansion \begin{eqnarray} &&{\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau)\\\label{expand2} &&={\epsilon^2\over 2}\ddot{x}^\mu(\tau)-{\epsilon^3\over 3}\left(\dddot{x}^\mu+\ddot{x}^\nu\ddot{x}_\nu\dot{x}^\mu\right)(\tau) +{\cal O}(\epsilon^4\gamma^6)\quad,\nonumber\\ && {\bf X}^2=\epsilon^2-{\epsilon^4\over 12}\ddot{x}^\nu\ddot{x}_\nu +{\cal O}(\epsilon^5\gamma^5)\quad, \end{eqnarray} Inserting the expansion into (\ref{renormalized}), yields \begin{eqnarray} &&{1\over 2\epsilon}\left({1\over\left(1-{\epsilon^2\over 12}\ddot{x}^\nu\ddot{x}_\nu\right)^{3\over 2}} -1\right)\ddot{x}^\mu\\&& -{1\over 3\left(1-{\epsilon^2\over 12}\ddot{x}^\nu\ddot{x}_\nu\right)^{3\over 2}}\left(\dddot{x}^\mu+\ddot{x}^\nu\ddot{x}_\nu\dot{x}^\mu\right)\nonumber +{\cal O}\left(\gamma^4\right)\quad, \end{eqnarray} In estimating ${\cal O}\left(\gamma^4\right)$ we already used the assumption $\epsilon\sim {1\over\gamma^2}$. This is justified because of the expression in the denominator which effectively cutoffs the $\epsilon$-integral and confines it to $\epsilon\sim (\ddot{x})^{-1}\sim {1\over\gamma^2}$. The dominant contribution in the large $\gamma$ limit then comes from the second line which is ${\cal O}\left(\gamma^5\right)$, while others are all ${\cal O}\left(\gamma^4\right)$, so that the leading self-force becomes \begin{eqnarray} f^{\mu}_{3D}&\approx&{e^2\over 6\pi}\left(\int_0^\infty{d\epsilon\over \left(1-{\epsilon^2\over 12}\ddot{x}^\nu\ddot{x}_\nu\right)^{3\over 2}}\right) \ddot{x}^\nu\ddot{x}_\nu \dot{x}^\mu+{\cal O}(\gamma^2)\nonumber\\ &=& -{e^2\over\sqrt{3}\pi}\sqrt{-\ddot{x}^\nu\ddot{x}_\nu} \,\,\dot{x}^\mu+{\cal O}(\gamma^2)\quad, \end{eqnarray} which is ${\cal O}\left(\gamma^3\right)$. This is our 2+1 dimensional version of the Abraham-Lorentz-Dirac force. We apply our result to the ultra-relativistic circular motion \begin{eqnarray} x^\mu(\tau)=\left(\gamma\tau, \rho\,{\rm cos}(\gamma\omega\tau), \rho\,{\rm sin}(\gamma\omega\tau)\right)\quad, \label{T10} \end{eqnarray} where ${1\over\gamma^2}=(1-v^2)\ll 1$ with $v=\rho\,\omega\approx 1$. Because $\ddot{x}^\nu \ddot{x}_\nu=-\rho^2 \omega^4\gamma^4 \approx -\omega^2 \gamma^4$, the proper time integral is confined to \begin{eqnarray} |\tau'-\tau|\approx\frac 1{\omega\gamma^2}\ll 1\quad, \label{T11} \end{eqnarray} as discussed before, and the leading covariant self-force is \begin{eqnarray} f^\mu_{3D}=-{e^2\over\sqrt{3}\pi} \omega\gamma^2 \dot{x}^\mu\label{ressec2} \quad, \end{eqnarray} which is longitudinal. The common (non-covariant) longitudinal drag force is \begin{eqnarray} \vec{\bf f}_L\approx -\frac{e^2}{\sqrt{3}\pi}\omega\gamma^2\vec{v}\quad. \label{T14} \end{eqnarray} Therefore, the external force that is needed to maintain the circular motion is \begin{eqnarray} \vec{\bf f}_{ext}= m_{ren}{d^2\vec{x}\over dt^2} -\vec{\bf f}_L\quad, \end{eqnarray} and the rate of work done to the particle is given by \begin{eqnarray} P=\vec{\bf f}_{ext}\cdot\vec{v}\approx \frac{e^2}{\sqrt{3}\pi}\omega\gamma^2\quad, \label{work} \end{eqnarray} where we used ${d^2\vec{x}\over dt^2}\cdot\vec{v}=0$ for the circular motion. In section (\ref{farsec}), we will compute the far-field radiation of this synchrotron motion, and find an agreement between the total power radiated at large distance and the work done locally to the charge (\ref{work}). \section{Local self-force in 4+1 dimensions} In this section, we will perform similar computations as in the previous section but in 4+1 dimensional spacetime. We will therefore be brief in explaining the details of the derivation while presenting our results. The Maxwell's equations sourced by a charge $e$ in the radiative gauge are solved as before, \begin{eqnarray} A^\mu(x) &=&e\int d\tau'\,\Delta_R\left(x-x(\tau')\right)\dot{x}^\mu(\tau'),\label{ret5d} \end{eqnarray} where the retarded propagator in 4+1 dimensions is \begin{eqnarray} \Delta_R(x)= -\frac{\theta(x^0)}{2\pi^2}\left(\frac{\delta(x^2)}{\sqrt{x^2}}-\frac 12\,\frac{\theta(x^2)}{x^2\sqrt{x^2}}\right). \label{F3} \end{eqnarray} Note the analogy with (\ref{prop}), which in fact stems from the generic relationship \begin{eqnarray} \Delta_{4+1}(x)=-\frac 1\pi\,\frac{d}{dx^2}\,\Delta_{2+1}(x), \label{F4} \end{eqnarray} which is readily shown from the momentum space representation of the retarded propagators and the recursive property of the integer Bessel functions. Modulo normalizations, (\ref{F4}) extends to all odd space-time dimensions. From (\ref{ret5d}) and (\ref{F3}), one easily writes down the unregularized self-force simialr to (\ref{unreg}) before as \begin{eqnarray} f^\mu_{5D}= 2e^2\int d\tau'\,\Delta_R'\left({\bf X}\right) {\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau),\label{unreg5d} \end{eqnarray} where in 4+1 dimensions we have \begin{eqnarray} \Delta'_R(x)=-\frac{\theta(x^0)}{2\pi^2}\left(\frac{\delta'(x^2)}{\sqrt{x^2}}-\frac{\delta(x^2)}{(x^2)^{3\over 2}} +\frac 34 \frac{\theta(x^2)}{(x^2)^{5\over 2}}\right). \label{F6} \end{eqnarray} In (\ref{unreg5d}), we have dropped terms that are proportional to $\delta\left({\bf X}^0\right)$ due to our regularization scheme that we explain in the following. However, we should caution the readers that we haven't checked the regularization scheme independency of our results, contrary to the previous 2+1 dimensional case. We will regularize the divergences appearing in (\ref{unreg5d}) by replacing $\Delta'_R(x)$ with \begin{eqnarray} \Delta^{'\epsilon}_R(x)=-\frac{\theta(x^0)}{2\pi^2}\left(\frac{\delta'(x^2-\epsilon^2)}{\sqrt{x^2}}-\frac{\delta(x^2-\epsilon^2)}{(x^2)^{3\over 2}} +\frac 34 \frac{\theta(x^2-\epsilon^2)}{(x^2)^{5\over2}}\right),\nonumber\\ \label{regprop} \end{eqnarray} and taking the $\epsilon\to 0$ limit after removing divergences by renormalization. To identify the divergences in (\ref{unreg5d}), the necessary small $\epsilon$-expansion reads as \begin{eqnarray} &&{\bf X}^{[\mu}(\tau,\tau')\dot{x}^{\nu]}(\tau')\dot{x}_\nu(\tau)\label{expand5d}\\ &&={\epsilon^2\over 2}\ddot{x}^\mu(\tau)-{\epsilon^3\over 3}\left(\dddot{x}^\mu+\ddot{x}^\nu\ddot{x}_\nu\dot{x}^\mu\right)(\tau) \nonumber\\ && +{\epsilon^4\over 8}\left(\ddddot{x}^\mu-\ddddot{x}^\nu \dot{x}_\nu \dot{x}^\mu +{2\over 3} \ddot{x}^\nu\ddot{x}_\nu \ddot{x}^\mu\right)\nonumber\\ &&-{\epsilon^5\over 30}\left({x}^{(5)\mu}-{x}^{(5)\nu} \dot{x}_\nu \dot{x}^\mu -{5\over 4} \ddddot{x}^\nu\dot{x}_\nu \ddot{x}^\mu\right) +{\cal O}(\epsilon^6\gamma^8),\nonumber \end{eqnarray} where $x^{(5)}\equiv {d^5 x\over d\tau^5}$. Upon inserting (\ref{expand5d}) into (\ref{unreg5d}), the divergences of the self-force are found to be \begin{eqnarray} f^\mu_{5D}&\sim& {e^2\over 16\pi^2}{1\over\epsilon^2}\ddot{x}^\mu\label{leading}\\ &+&{3e^2\over 32\pi^2}\log(\epsilon)\left(\ddddot{x}^\mu-\ddddot{x}^\nu \dot{x}_\nu \dot{x}^\mu +{2\over 3} \ddot{x}^\nu\ddot{x}_\nu \ddot{x}^\mu\right).\nonumber \end{eqnarray} The first leading divergence can readily be absorbed into the renormalized mass \begin{eqnarray} m_{ren}=m_{bare}-{e^2\over 16\pi^2}{1\over\epsilon^2}, \end{eqnarray} whereas the nature of the second term of logarithmic divergence is unclear to us. We will simply choose our regularization scheme to remove it minimally. We expect this minimal subtraction to be consistent with the far-field radiation formulae, although we will only check this for the 2+1 dimensional case below. After removing the divergences, the finite contribution can be computed in the leading $\gamma$ approximation. One finds that the $\delta({\bf X^2})$ and $\delta'({\bf X^2})$ terms in $\Delta'_R({\bf X})$ in (\ref{unreg5d}) give us the leading $\gamma^6$ contributions to the self-force, \begin{eqnarray} f^\mu_{5D}\sim{e^2\over 8\pi^2}\left(\ddddot{x}^\mu+\ddot{x}^\nu\ddot{x}_\nu \ddot{x}^\mu- \ddddot{x}^\nu\dot{x}_\nu \dot{x}^\mu\right)+{\cal O}(\gamma^5), \end{eqnarray} which is completely local. In deriving the above, one has to expand \begin{eqnarray} {\bf X}^2=\epsilon^2-{\epsilon^4\over 12}\ddot{x}^\mu\ddot{x}_\mu -{\epsilon^5\over 12}\left(\ddddot{x}^\mu\dot{x}_\mu +2 \ddot{x}^\mu\dddot{x}_\mu\right)+{\cal O}(\epsilon^6\gamma^6)\nonumber. \end{eqnarray} However, for circular motion that we are interested in, the longitudinal component of the above force that is related to the rate of work done simply vanishes. Therefore, we are led to seek the next leading term in ${\cal O}(\gamma^5)$. The next leading term is quasi-local, that is, it comes from the region $\epsilon\sim{1\over\gamma^2}$ as was the case in 2+1 dimensions before. It is given by \begin{eqnarray} &&-{e^2\over 40\pi^2}\left(\int_0^\infty d\epsilon\,{1\over\left(1-{\epsilon^2\over12}\ddot{x}^\nu\ddot{x}_\nu\right)^{5\over 2}}\right)\\&&\times\big(x^{(5)\nu}\dot{x}_\nu \dot{x}^\mu +{35\over 8}\ddddot{x}^\nu\dot{x}_\nu \ddot{x}^\mu +{25\over 4}\ddot{x}^\nu\dddot{x}_\nu\ddot{x}^\mu\big)\nonumber\\ &&=-{e^2\over 10\sqrt{3}\pi^2}{\big(x^{(5)\nu}\dot{x}_\nu \dot{x}^\mu +{35\over 8}\ddddot{x}^\nu\dot{x}_\nu \ddot{x}^\mu +{25\over 4}\ddot{x}^\nu\dddot{x}_\nu\ddot{x}^\mu\big)\over\sqrt{-\ddot{x}^\nu\ddot{x}_\nu}}.\nonumber \end{eqnarray} Applying the above to the ultra-relativistic synchrotron motion, we see that only the first term contributes, and we obtain the leading longitudinal force as \begin{eqnarray} f^\mu_{5D}\sim -{e^2\over 10\sqrt{3}\pi^2}\gamma^4\omega^3 \dot{x}^\mu, \end{eqnarray} which leads to the rate of work done to the system, \begin{eqnarray} P={e^2\over 10\sqrt{3}\pi^2}\gamma^4\omega^3 . \end{eqnarray} \section{Far-field synchrotron radiation in 2+1 dimensions\label{farsec}} In this section, we compute the far-field radiation of the ultra-relativistic circular motion and check that the resulting total rate of radiation matches the rate of work done locally to the charge (\ref{work}), which is our first nontrivial consistency check for the ``self-force" approach and the subtractions we have applied. The retarded gauge field from a point-like charge motion is \begin{eqnarray} A^\mu(t,\vec x)={e\over 2\pi}\int^{t_*(t,\vec x)}_{-\infty}dt'\,{1\over \sqrt{(t-t')^2- (\vec x-\vec x'(t'))^2}} {dx'^\mu\over dt'},\nonumber \end{eqnarray} where $t_*(t,\vec x)<0$ is determined by the intersection between the past light-cone from $(t,\vec x)$ and the particle trajectory \begin{eqnarray} (t-t_*)^2-(\vec x-\vec x'(t_*))^2=0\quad.\label{deft} \end{eqnarray} For the synchrotron motion we are interested in \begin{eqnarray} x^\mu&=&(t,\rho\cos(\omega t),\rho\sin(\omega t))\nonumber\\ &=&\left(\gamma\tau, \rho\,{\rm cos}(\gamma\omega\tau), \rho\,{\rm sin}(\gamma\omega\tau)\right)\quad, \end{eqnarray} the field strengths can be written explicitly as in the appendix. We will start our discussion with the formulae (\ref{app10}), (\ref{app11}), and (\ref{app12}). We are interested in the far-field asymptotics at $r\to\infty$, where we introduce a polar coordinate system $(r,\phi)$ on the spatial two dimensional plane. The expression for the total far-field radiation is \begin{eqnarray} {d E\over dt}&=& \lim_{r\to\infty}r\int_0^{2\pi}d\phi\, T_{0i}\hat n^i\label{4s1}\\ &=&\lim_{r\to\infty}r\int_0^{2\pi}d\phi\, F_{12}\left(F_{01}\sin\phi-F_{02}\cos\phi\right),\nonumber \end{eqnarray} where $\hat n=(\cos\phi,\sin\phi)$ is the unit radial vector. To compute the leading large $r$ asymptotics of the field strengths, it is useful to note \begin{eqnarray} t'_*&=&-r+{\cal O}(1)\quad,\quad {\partial t'_*\over\partial t}={1\over 1+v\sin(\omega t'_*-\phi)},\label{4s2}\\ {\partial t'_*\over\partial x^1}&=& {-\cos\phi\over 1+v\sin(\omega t'_*-\phi)}\quad,\quad {\partial t'_*\over\partial x^2}= {-\sin\phi\over 1+v\sin(\omega t'_*-\phi)},\nonumber \end{eqnarray} in the $r\to\infty$ limit. These quantities appear in the numerators of the field strength expressions, and one can see from the structure of the denominators in the above that in the limit $v\to 1$, these quantities are highly peaked around the angle $\phi_c$ determined by the condition \begin{eqnarray} \phi_c=\omega t'_*(\phi_c)+{\pi\over 2}.\label{4s3} \end{eqnarray} We note that $\omega t'_*(\phi)$ is the azimuthal angle of the particle at the intersection of the past light-cone from $(r,\phi)$. This condition has a simple geometrical meaning: the light pulse emitted from one moment of the trajectory is highly collimated in the direction of the instant velocity, and travels with the speed of light. Below, we will see that the leading $\gamma$ contribution is indeed confined in the small angular range $\delta\phi\sim {1\over\gamma^3}$ around $\phi_c$. For that purpose, it will be useful to have the formula \begin{eqnarray} {\partial t'_*\over\partial \phi}={\rho\sin(\omega t'_*-\phi)\over 1+v\sin(\omega t'_*-\phi)},\label{4s4} \end{eqnarray} so that defining a new convenient angular variable $a$ instead of $\phi$ \begin{eqnarray} a\equiv \omega t'_*(\phi)-\phi+{\pi\over 2}\quad,\label{4s5} \end{eqnarray} we have a relation \begin{eqnarray} {da\over d\phi} = -{1\over 1+v\sin(\omega t'_*-\phi)}=-{1\over 1-v\cos a}\,\label{4s6} \end{eqnarray} which can be integrated as \begin{eqnarray} a-v\sin a = -(\phi-\phi_c)\quad,\label{4s7} \end{eqnarray} using the boundary condition (\ref{4s3}). As $\phi$ runs in the range $(0,2\pi)$, $a$ also runs one cycle $(0,2\pi)$ monotonically with a very steep gradient of order ${\cal O}(\gamma^2)$ around a small region near $a=0$ ($\phi=\phi_c$) that can be seen in (\ref{4s6}). Therefore, the width $\delta a$ of the collimated light pulse near $a=0$ will translate to the width in $\phi$ \begin{eqnarray} \delta \phi \sim {\delta a\over\gamma^2},\label{4ss1} \end{eqnarray} near the center $\phi=\phi_c$. We will see shortly that $\delta a \sim {1\over\gamma}$ leading to $\delta\phi\sim {1\over\gamma^3}$. Another important large $r$ expansion is the one for the proper distance, \begin{eqnarray} (t')^2-(\vec x-\vec x')^2=(t')^2-r^2+2r\rho\cos(\omega t'-\phi)-\rho^2\quad,\nonumber \end{eqnarray} that appears in the denominators in the field strength expressions. Since the $t'$ integration starts from $t'_*$ to $-\infty$, it is more convenient to shift $t'$ integration by \begin{eqnarray} t'\to -t+t'_*\quad, \end{eqnarray} upon which $t\in(0,\infty)$, and the proper distance becomes \begin{eqnarray} t^2-2 t'_* t +2r\rho\sin(a-\omega t)-2r\rho\sin a,\label{4s9} \end{eqnarray} where we used (\ref{4s5}) and the definition of $t'_*$ (\ref{deft}). Note that $t_*=-r+{\cal O}(1)$ in $r\to\infty$, and one can consider two parametric regions of the $t$-integral: (1) $t\ll |t_*'|\approx r$ and (2) $t\ge |t_*'|\approx r$. For (1), one can clearly neglect the $t^2$ term in (\ref{4s9}) and the proper distance is ${\cal O}(r)$. For (2) the proper distance becomes of order ${\cal O}(r^2)$, and considering that this enters the denominators of the field strength expressions, we see that the region (2) of $t$ integral gives us a sub-dominant large $r$ behavior of field strengths. One concludes that the leading large $r$ value of field strengths arises only from the range (1), and therefore we can neglect the $t^2$ term in (\ref{4s9}) when computing leading large $r$ asymptotics, and the proper distance can simply be replaced by \begin{eqnarray} 2r\left(t+\rho\sin(a-\omega t)-\rho\sin a\right),\label{4s10} \end{eqnarray} for this purpose. It is straightforward to use (\ref{app10}), (\ref{app11}), and (\ref{app12}) to compute the large $r$ asymptotics of the field strengths. After some algebra, one finds that \begin{eqnarray} F_{12}=\left(F_{01}\sin\phi-F_{02}\cos\phi\right)\quad, \end{eqnarray} at leading order in $r\to\infty$, and \begin{eqnarray} F_{12}&\to& {e\over 2\pi}{v\omega^{1\over 2}\over \sqrt{2r}}{1\over 1-v\cos a}\int_0^\infty dt\,\Bigg\{\nonumber\\ &&{-\cos(a-t)\over\sqrt{t+v\sin(a-t)-v\sin a}}\label{4s11}\\&&+{1\over 2}{v\sin(a-t)\left(\cos(a-t)-\cos a\right) \over \left(t+v\sin(a-t)-v\sin a\right)^{3\over 2}}\Bigg\},\nonumber \end{eqnarray} where we have rescaled $t$ by $\omega t \to t$. The above expression is a general result for arbitrary velocity $v=\rho\omega$ and the angle $a$ (or equivalently $\phi$). Now, we need to find the leading $\gamma$ behavior of the total radiated power (\ref{4s1}), \begin{eqnarray} {dE\over dt} = \lim_{r\to\infty} r\int_0^{2\pi}da\,\left|{da\over d\phi}\right|^{-1}|F_{12}|^2,\label{4s12} \end{eqnarray} where we changed the angle integration from $\phi$ to $a$ for convenience. By careful inspection of the above integral, it can be shown that the leading $\gamma$ behavior of ${\cal O}(\gamma^2)$ arises from the narrow range of $(\delta t,\delta a)\sim {1 \over\gamma}$ around $(t,a)=0$, and one can for example expand the denominator as \begin{eqnarray} t+v\sin(a-t)-v\sin a \approx {1\over 2}\left({1\over\gamma^2}+a^2\right)t -{1\over 2} a t^2 +{1\over 6}t^3,\nonumber \end{eqnarray} up to the relevant order. One also expands \begin{eqnarray} \left|{da\over d\phi}\right|={1\over 1-v\cos a}\approx {2\over {1\over\gamma^2}+a^2},\label{4s14} \end{eqnarray} as well as the numerator, and by rescaling the integration variables $(t,a)\to {1\over\gamma}(t,a)$, the leading $\gamma$-piece can easily be shown to reduce to \begin{eqnarray} {d E\over dt} = {e^2\over (2\pi)^2}\gamma^2 \omega \int_{-\infty}^{\infty} da\, {1\over 1+ a^2} |f(a)|^2, \end{eqnarray} where \begin{eqnarray} f(a)=\partial_a\left(\int_0^\infty dt\,{(a-t)\over\sqrt{{1\over 2}(1+a^2)t-{1\over 2} a t^2 +{1\over 6} t^3} }\right). \end{eqnarray} The integral \begin{eqnarray} \int_{-\infty}^{\infty} da\, {1\over 1+ a^2} |f(a)|^2 \end{eqnarray} is $(2\pi) {2\over \sqrt{3}}$ up to 4 digits numerically. Assuming this numerical result to be exact, we obtain for the total power radiated \begin{eqnarray} {dE\over dt} = {e^2 \over \sqrt{3}\pi}\omega\gamma^2,\label{4s15} \end{eqnarray} which agrees precisely with the rate of work done locally by the ``self-force" (\ref{work}) derived in section \ref{sec21}. For completeness, in Fig.\ref{fig1}, we plot the angular distribution of the radiation power, ${dP\over da}={1\over 1+ a^2} |f(a)|^2$, as a function of $a$ (recall that the true $a$ is given by $a\over\gamma$). The plot can be translated to the angular distribution in $\phi$ using the relation (\ref{4s7}). \begin{figure}[t] \centering \includegraphics[width=8cm]{fig1.eps} \caption{ The angular distribution of radiated power in the leading $\gamma$ approximation. See (\ref{4s7}) for the relation between $a$ (more precisely ${a\over \gamma}$) and $\phi$. \label{fig1}} \end{figure} \section{Summary } We have shown that, unlike in the 3+1 dimensional space-time, the self-force can be defined to be nonzero in odd-dimensional space-times, as the retarded propagators in these cases contain a theta-function part or inside-the-light-cone contributions. Furthermore, in the ultrarelativistic case it can be put in local form, depending only on the instantaneous derivatives of the particle motion. We have explicitly derived such expressions for the self-force, in 2+1 and 4+1 dimensions. In the former case we have also calculated the radiation intensity at large distances and checked that it matches the work done by the self-force numerically. We expect the same in 4+1 dimensions with minimal subtraction. We think it is perhaps the first instance of an entirely local and consistent derivation of the radiation braking force. Although the MSTQW approach \cite{Mino:1996nk,Quinn:1996am} is 15 years old and it had inspired our work, we are not aware of its explicit tests in the gravitational setting either. Needless to say, much more work is needed in order to find the exact applicability domain of the ``self-force" approach, in flat and curved space-times. In~\cite{Shuryak:2011ge} we have suggested that the MSTQW equation when adapted to thermal AdS may be of relevance to jet quenching in ultrarelativistic collisions such as RHIC and LHC. \vskip .25cm {\bf Acknowledgments.} \vskip .2cm This work was supported in parts by the US-DOE grant DE-FG-88ER40388. \section{Appendix: Alternative regularization of the self-force in 2+1 dimensions} In this appendix, we present another way of regularizing the self-force in 2+1 dimensions, by averaging the Lorentz force \begin{eqnarray} f^\mu_{3D}=e F^{\mu\nu}\left(x(\tau)\right) \dot{x}_\nu(\tau) \quad,\label{app1} \end{eqnarray} around a small circle of radius $r$, and taking $r\to 0$ limit after renormalizing the diverging mass. We will restrict ourselves to the case of ultra-relativistic circular motion for simplicity, \begin{eqnarray} x^\mu(\tau)=\left(\gamma\tau, \rho\,{\rm cos}(\gamma\omega\tau), \rho\,{\rm sin}(\gamma\omega\tau)\right)\quad, \label{app5} \end{eqnarray} where $v=\rho\,\omega\approx 1$, and will find that the leading $\gamma$ result of the finite self-force agrees with the one in section \ref{sec21}, which is a confirming check for our results. Starting from the expression of retarded gauge potential (\ref{ret}), \begin{eqnarray} A^\mu(x) =e\int d\tau'\,\Delta_R\left(x-x(\tau')\right)\dot{x}^\mu(\tau')\quad,\label{app2} \end{eqnarray} with \begin{eqnarray} \Delta_R(x)={\theta(x^0)\over 2\pi}{\theta(x^2)\over\sqrt{x^2}}\quad,\quad x^2\equiv x^\mu x_\mu\quad,\label{app3} \end{eqnarray} we need to compute field strengths at radius $r$ from the position of the charge at $\tau=0$ in the two-dimensional spatial plane $\vec x=(x^1,x^2)$. Let us show the computation $F_{01}=\partial_0 A_1-\partial_1 A_0$ in some detail to set-up notations and procedures, and we will present other components of $F_{\mu\nu}$ at the end without much details. The expression (\ref{app2}) gives us \begin{eqnarray} A_0(t,\vec x)={e\over 2\pi}\int^{t'_*(t,\vec x)}_{-\Lambda}\, {dt'\over\sqrt{(t-t')^2-(\vec x-\vec x')^2}}\quad,\label{app6} \end{eqnarray} where $\vec x'\equiv \vec x(t')$ is the past trajectory (\ref{app5}) parameterized here in terms of regular time $t'=\gamma \tau'$, and the integration starts from $t'_*(t,\vec x)<0$ which is determined by the intersection between the past light-cone from the position $(t,\vec x)$ and the particle trajectory, \begin{eqnarray} (t-t'_*)^2=\left(\vec x-\vec x'(t'_*)\right)^2\quad.\label{app7} \end{eqnarray} The expression itself is infrared divergent and we introduced a cutoff $\Lambda$, but the field strengths which are physical are completely IR finite as one takes $\Lambda\to\infty$ at the end, as will be clear in the following. To compute $\partial_1 A_0$, one needs to evaluate the variation of (\ref{app6}) with respect to $x^1\to x^1+\delta x^1$. The variation will shift both the integration range through $t'_*$ and the integrand. The former variation naively gives us the contribution which is proportional to the value of the integrand at $t'=t'_*$ which happens to be divergent due to (\ref{app7}). The variation of the integrand also gives us an integral which is divergent near $t'=t'_*$. However, since the original (\ref{app6}) is completely finite near $t'=t'_*$ these divergences are mere artifacts of improper manipulations, and in fact the two divergences cancel with each other. A better way of handling them is the following. From the expression of $A_0$ at $\vec x+\delta\vec x$, \begin{eqnarray} A_0(\vec x+\delta\vec x)={e\over 2\pi}\int^{t'_*+{\partial t'_*\over\partial\vec x}\cdot\delta \vec x}_{-\Lambda}\, {dt'\over\sqrt{(t-t')^2-(\vec x+\delta\vec x-\vec x')^2}},\nonumber\\ \label{app8} \end{eqnarray} one shifts the $t'$ integral by $t'\to t'+{\partial t'_*\over\partial\vec x}\cdot\delta\vec x$ so that the new $t'$ integral starts at the same $t'_*$ while the integrand gets additional contributions \begin{eqnarray} && A_0(\vec x+\delta\vec x)=\label{app9}\\&& {e\over 2\pi}\int^{t'_*}_{-\Lambda}\, {dt'\over\sqrt{(t-t'-{\partial t'_*\over\partial\vec x}\cdot\delta\vec x)^2-(\vec x+\delta\vec x-\vec x'-{\partial\vec x'\over\partial t'} {\partial t_*'\over\partial\vec x}\cdot\delta\vec x)^2}},\nonumber \end{eqnarray} which is free from the divergence near $t'=t'_*$. There is also a shift in the IR cutoff \begin{eqnarray} \Lambda\to\Lambda-{\partial t'_*\over\partial\vec x}\cdot\delta\vec x\quad,\end{eqnarray} but it can be easily shown that it becomes irrelevant in the final results, and we omit it in (\ref{app9}). Taking difference between (\ref{app9}) and (\ref{app6}) to first order in $\delta\vec x$, one readily computes $\vec\partial A_0$ as (after putting $t=0$) \begin{eqnarray} && \vec\partial A_0 =\label{app10}\\ && {-e\over 2\pi}\int^{t'_*}_{-\infty}dt'\,{{\partial t'_*\over\partial\vec x} t'-(\vec x-\vec x') +(\vec x-\vec x')\cdot{\partial \vec x'\over\partial t'} {\partial t'_*\over\partial\vec x} \over \left((t')^2 -(\vec x-\vec x')^2\right)^{3\over 2}},\nonumber \end{eqnarray} where we removed $\Lambda$ to infinity as the final integral is finite. By similar steps, one obtains \begin{eqnarray} &&\partial_0 \vec A={-e\over 2\pi} \int^{t'_*}_{-\infty}dt'\,\Bigg({{\partial^2\vec x'\over\partial t'^2}{\partial t'_*\over\partial t}\over \sqrt{(t')^2-(\vec x-\vec x')^2}}\label{app11}\\ &&+{{\partial\vec x'\over\partial t'}\left((1-{\partial t'_*\over\partial t})t' -(\vec x-\vec x')\cdot{\partial\vec x'\over\partial t'}{\partial t'_*\over\partial t}\right)\over \left((t')^2 -(\vec x-\vec x')^2\right)^{3\over 2}}\Bigg).\nonumber \end{eqnarray} The above (\ref{app10}) and (\ref{app11}) give us the field strength $F_{0i}=\partial_0 A_i-\partial_i A_0$, $i=1,2$. The expressions for ${\partial t'_*\over \partial t}$ and $\partial t'_*\over\partial \vec x$ that appear in the above can be easily obtained from the relation (\ref{app7}). Finally, $F_{12}=\partial_1 A_2-\partial_2 A_1$ is written as \begin{eqnarray} &&F_{12}={e\over 2\pi} \int^{t'_*}_{-\infty}dt'\,\Bigg({\epsilon_{ij}{\partial^2\vec x'^i\over\partial t'^2}{\partial t'_*\over\partial x^j}\over \sqrt{(t')^2-(\vec x-\vec x')^2}}\label{app12}\\ &&-{\epsilon_{ij}{\partial\vec x'^i\over\partial t'}\left({\partial t'_*\over\partial x^j}t' -(x^j-x'^j) +(\vec x-\vec x')\cdot{\partial\vec x'\over\partial t'}{\partial t'_*\over\partial x^j}\right)\over \left((t')^2 -(\vec x-\vec x')^2\right)^{3\over 2}}\Bigg),\nonumber \end{eqnarray} where $\epsilon_{12}=-\epsilon_{21}=+1$. We will apply the above formulae to our case of circular motion (\ref{app5}). We then consider a small circle of radius $r$ from the position of the charge at $t=0$, that is, $(\rho,0)$. We let the azimuthal angle of the circle be $\phi$, so that a point on the circle has the coordinate $\vec x=(\rho+r\cos\phi,r\sin\phi)$. We will compute the Lorentz force on the points in the circle and take an average over $\phi$ before taking the limit $r\to 0$. Noting the 3-velocity ${dx^\mu\over d\tau}\equiv u^\mu =\gamma (1,0,v)$ at $t=0$ (recall $v=\rho\omega$), let's first look at the transverse component of the self-force, \begin{eqnarray} f^1=e\gamma \left(F_{01}-v F_{12}\right)\quad. \end{eqnarray} Near $r\to 0$, the useful expansions of some quantities are \begin{eqnarray} t'_*&=& \gamma^2 xr + {\cal O}(r^2)\,\,,\,\,x\equiv -({v\sin\phi+\sqrt{1-v^2\cos^2\phi}})\nonumber,\\ {\partial t'_*\over \partial x^1}&=&{\cos\phi\over {x}+v\sin\phi}\quad,\quad {\partial t'_*\over \partial x^2}={\sin\phi-\gamma^2 v x \over {x}+v\sin\phi},\nonumber\\ {\partial t'_*\over \partial t}&=&{\gamma^2 x\over {x}+v\sin\phi}\quad,\label{app13} \end{eqnarray} and from these, one can derive that the $\phi$-averaged transverse force has a divergence near $r\to 0$ from the term \begin{eqnarray} f^1&\sim& {e^2\gamma^2 v\omega\over 2\pi}\int^{\gamma^2 x r} dt'{1\over\sqrt{(t'-\gamma^2 rx)(t'-\gamma^2 r x')}}\nonumber\\ &\sim&{e^2\gamma^2 v\omega \over 2 \pi} \log\left(1\over r\right)=-{e^2\over 2\pi}\log\left(1\over r\right)\ddot{x}^1, \label{app14} \end{eqnarray} where $x'=-({v\sin\phi-\sqrt{1-v^2\cos^2\phi}})>0$. In deriving this, we have used an important expansion \begin{eqnarray} (t')^2-(\vec x-\vec x')^2 \approx {1\over\gamma^2}(t'-\gamma^2 rx)(t'-\gamma^2 r x'),\label{app15} \end{eqnarray} near $r\to 0$ , ${t'\over r}\sim {\cal O}(1)$ limit, which gives one factor of $\gamma$ in (\ref{app14}). This divergence is precisely of the same character of the mass renormalization we encountered in section \ref{sec21}, and one can absorb it by \begin{eqnarray} m_{ren}=m_{bare}+{e^2\over 2\pi}\log\left(1\over r\right)\quad. \end{eqnarray} We are more interested in the leading $\gamma$ behavior of longitudinal self-force, \begin{eqnarray} f^2=e\gamma F_{02}\quad,\label{app15} \end{eqnarray} and using (\ref{app13}), averaging over $\phi$ and taking $r\to0$ limit, one arrives at a finite expression, \begin{eqnarray} f^2&=&-{e^2\gamma^3\over 2 \pi}\int^\infty_0 dt'\, \Bigg({v\omega\sin(\omega t')\over\sqrt{ (t')^2-4\rho^2\sin^2\left(\omega t'\over 2\right)}}\nonumber\\ &+& {\rho\left(1-v^2\cos(\omega t')\right)\left(\sin(\omega t')-\omega t'\right)\over \left((t')^2-4\rho^2\sin^2\left(\omega t'\over 2\right)\right)^{3\over 2}}\Bigg),\label{app16} \end{eqnarray} where we changed the variable $t'\to-t'$ in the integration. The last step is to find a leading $\gamma$ behavior of (\ref{app16}). Rescaling $\omega t'\to t$, we have \begin{eqnarray} f^2&=&-{e^2\gamma^3 v\omega\over 2\pi} \int^\infty_0 dt\, \Bigg({\sin t\over\sqrt{ t^2-4v^2\sin^2\left(t\over 2\right)}}\nonumber\\ &+& {\left(1-v^2\cos t\right)\left(\sin t - t\right)\over \left(t^2-4v^2\sin^2\left(t\over 2\right)\right)^{3\over 2}}\Bigg).\label{app17} \end{eqnarray} To study the large $\gamma$ behavior of the integral in (\ref{app17}) replacing $v^2=1-{1\over\gamma^2}$, one writes the integral after some manipulations \begin{eqnarray} &&\int_0^\infty dt\,\partial_t\left({(1-\cos t)\over \sqrt{t^2-4\sin^2\left(t\over 2\right)}}\right)\\ && +{1\over\gamma^2}\int_0^\infty dt\,{4\sin t\sin^2\left(t\over 2\right)+(\sin t-t)\cos t\over \left(t^2-4\sin^2\left(t\over 2\right)+{4\over\gamma^2}\sin^2\left(t\over 2\right)\right)^{3\over 2}},\nonumber \end{eqnarray} neglecting additional ${\cal O}(\gamma^{-2})$ contributions. The first integral is completely localized at $t=0$ giving a value of $-\sqrt{3}$. In the second integral, one can show that a leading ${\cal O}(1)$ result comes from a range of small $t\sim {\cal O}\left(1\over\gamma\right)$, and one has an approximation in the denominator, \begin{eqnarray} t^2-4\sin^2\left(t\over 2\right)+{4\over\gamma^2}\sin^2\left(t\over 2\right)\approx {t^2\over\gamma^2}+{1\over 12}t^4 ,\label{app18} \end{eqnarray} where higher order terms can be shown to be irrelevant in the leading $\gamma$ contributions. The $t^4$ term in (\ref{app18}) effectively cutoffs the $t$ integral to be $t\le {1\over \gamma}$, which we have seen before in section \ref{sec21} (recall the proper time is $\tau={t\over\gamma}$). The small $t$ expansion for the numerator up to relevant order is \begin{eqnarray} 4\sin t\sin^2\left(t\over 2\right)+(\sin t-t)\cos t\approx{5\over 6}t^3 ,\label{app20} \end{eqnarray} giving us the second integral \begin{eqnarray} {1\over\gamma^2}\int_0^\infty dt\,{{5\over 6}t^3\over\left({t^2\over\gamma^2}+{1\over 12}t^4\right)^{3\over 2}} ={5\over 3}\sqrt{3}.\label{app21} \end{eqnarray} In total, the integral in (\ref{app17}) gives us $(-1+{5\over 3})\sqrt{3}={2\over\sqrt{3}}+{\cal O}(\gamma^{-2})$, so that the leading $\gamma$ result of $f^2$ is \begin{eqnarray} f^2\sim -{e^2\omega\gamma^3 \over \sqrt{3}\pi}\sim -{e^2\omega \gamma^2\over\sqrt{3}\pi} \dot{x}^2,\label{app19} \end{eqnarray} which agrees precisely with (\ref{ressec2}) in section \ref{sec21}.
\section{Introduction} How can a broadcast station communicate separate messages to two receivers using a single antenna? Two well known strategies \cite{el2010lecture} for transmitting information over broadcast channels are % % superposition coding \cite{C72,B73} and Marton over-binning using correlated auxiliary random variables \cite{M79}. In this paper, we prove that these strategies can be adapted to the quantum setting by constructing random codebooks and matching decoding measurements that have asymptotically vanishing error in the limit of many uses of the channel. Sending classical data over a quantum channel is one of the fundamental problems of quantum information theory~\cite{wilde2011book}. Single-letter formulas are known for classical-quantum point-to-point channels \cite{H98,SW97} and multiple access channels \cite{winter2001capacity}. Classical-quantum channels are a useful abstraction for studying general quantum channels and correspond to the transmitters being restricted to classical encodings. Codes for classical-quantum channels (c-q channels), when augmented with an extra optimization % over the possible input states, directly generalize to codes for quantum channels. Furthermore, it is known that classical encoding (coherent-state encoding using classical Gaussian codebooks) is sufficient to achieve the capacity of the lossy bosonic channel, which is a realistic model for optical communication links \cite{PhysRevLett.92.027902}. Previous work on quantum broadcast channels includes \cite{YHD2006,guha2007classical,DHL10}. Ref.~\cite{YHD2006} considers both classical and quantum communication over quantum broadcast channels and proves a superposition coding inner bound similar to our Theorem~\ref{thm:sup-coding-inner-bound}. Ref.~\cite{guha2007classical} discusses classical communication over a bosonic broadcast channel, and % % % % % % % % Ref.~\cite{DHL10} considers a Marton rate region for quantum communication. In this paper, we derive % % two achievable rate regions for classical-quantum broadcast channels % % % by exploiting error analysis techniques developed in the context of quantum interference channels \cite{FHSSW11,S11a}. % % % In Section~\ref{sec:superposition-coding}, we prove achievability of the superposition coding inner bound (Theorem~\ref{thm:sup-coding-inner-bound}), by using a quantum simultaneous nonunique decoder at one of the receivers. Yard {\it et al}. independently proved the quantum superposition coding inner bound \cite{YHD2006}, but our proof is arguably simpler and more in the spirit of its classical analogue~\cite{el2010lecture}. In Section~\ref{sec:marton} we prove that the quantum Marton rate region with no common message is achievable (Theorem~\ref{thm:marton-no-common}). In the Marton coding scheme, the sub-channels to each receiver are essentially point-to-point, but it turns out that a technique that we call the ``projector trick'' seems to be necessary in our proof. % % % % % % We discuss open problems and give an outlook for the future % in Section~\ref{sec:conclusion}. \section{Preliminaries} \subsubsection{Notation} We denote classical random variables as $X,U,W$, whose realizations are elements of the respective finite alphabets $\mathcal{X}, \mathcal{U}, \mathcal{W}$. Let $p_X$, $p_U$, $p_W$ denote their corresponding probability distributions. We denote quantum systems as $A$, $B$, and $C$ and their corresponding Hilbert spaces as $\mathcal{H}^{A}$, $\mathcal{H}^{B}$, and $\mathcal{H}^{C}$. We represent quantum states of a system~$A$ with a density operator $\rho^{A}$, which is a positive semi-definite operator with unit trace. Let $H(A)_{\rho}\equiv-\text{Tr}\left\{ \rho^{A}\log\rho^{A}\right\}$ denote the von Neumann entropy of the state $\rho^{A}$. % % A classical-quantum channel, $\mathcal{N}^{X\to B}$, is represented by the set of $|\mathcal{X}|$ possible output states $\{ \rho^B_x \equiv \mathcal{N}^{X\to B}\!(x) \}$, meaning that a classical input of $x$ leads to a quantum output $\rho^B_x$. % In a communication scenario, the decoding operations performed by the receivers correspond to quantum measurements on the outputs of the channel. A quantum measurement % is a positive operator-valued measure (POVM) $\left\{ \Lambda_{m}\right\}_{m\in\left\{ 1,\ldots,|\mathcal{M}|\right\} }$ on the system $B^n$, the output of which we denote $M^{\prime}$. To be a valid POVM, the set of $|\mathcal{M}|$ operators $\Lambda_{m}$ must all be positive semi-definite and sum to the identity: $\Lambda_{m} \geq 0, \,\,\, \sum_{m}\Lambda_{m}=I$. % % % % % \subsubsection{Definitions} % % % % % We define a classical-quantum-quantum broadcast channel as the following map:% \begin{equation} x\rightarrow\rho_{x}^{B_{1}B_{2}},\label{eq:cqq-broadcast}% \end{equation} where $x$ is a classical letter in an alphabet $\mathcal{X}$ and $\rho_{x}^{B_{1}B_{2}}$ is a density operator on the tensor product Hilbert space for systems $B_{1}$ and $B_{2}$. The model is such that when the sender inputs a classical letter $x$, Receiver~1 obtains system $B_{1}$, and Receiver~2 obtains system $B_{2}$. Since Receiver~1 does not have access to the $B_2$ part of the state $\rho_{x}^{B_{1}B_{2}}$, we model his state as $\rho_{x}^{B_{1}} = \text{Tr}_{B_2}\!\!\left[ \rho_{x}^{B_{1}B_{2}} \right]$, where $\text{Tr}_{B_2}$ denotes the partial trace over Receiver~2's system. \subsubsection{Information processing task} The task of communication over a broadcast channel is to use $n$ independent instances of the channel in order to communicate with Receiver~1 at a rate $R_1$ and to Receiver~2 at a rate $R_2$. % More specifically, the sender chooses a pair of messages $(m_1,m_2)$ from message sets $\mathcal{M}_i\equiv \left\{ 1,2,\ldots,|\mathcal{M}_i|\right\} $, where $|\mathcal{M}_i|=2^{nR_{i}}$, and encodes these messages into an $n$-symbol codeword $x^{n}\!\left( m_1,m_2\right)\in \mathcal{X}^n$ suitable as input for the $n$ channel uses. The output of the channel is a quantum state % of the form: \begin{align} \mathcal{N}^{\otimes n}\!\left( x^{n}(m_1,m_2) \right) & \equiv \rho_{x^{n}\left( m_1,m_2\right) }^{B_{1}^{n}B_{2}^{n}} \ \ \in \mathcal{H}^{B_1^n B_2^n}. % \end{align} where $ \rho_{x^{n}}^{B_{1}^{n}B_{2}^{n}} \equiv \rho_{x_1}^{B_{11}B_{21}} \otimes \cdots \otimes \rho_{x_n}^{B_{1n}B_{2n}}. $ To decode the message $m_1$ intended for him, Receiver~1 performs a POVM $\left\{ \Lambda_{m_1}\right\} _{m_1\in\left\{ 1,\ldots,|\mathcal{M}_1|\right\} }$ on the system $B_{1}^n$, the output of which we denote $M^{\prime}_1$. % % Receiver~2 similarly performs a POVM $\left\{ \Gamma_{m_2}\right\} _{m_2\in\left\{ 1,\ldots ,|\mathcal{M}_2|\right\} }$ on the system $B_{2}^n$, and the random variable associated with the outcome is denoted $M^{\prime}_2$. An error occurs whenever either of the receivers decodes the message incorrectly. % % % % % The probability of error for a particular message pair $(m_1,m_2)$ is \begin{align*} p_{e}\!\left( m_1,m_2\right) \equiv \text{Tr}\! \left\{ \left( I-\Lambda_{m_1}\otimes\Gamma_{m_2}\right) \rho_{x^{n}\left( m_1,m_2\right) }^{B_{1}^{n}B_{2}^{n}} \right\}, \end{align*} where the measurement operator $\left( I-\Lambda_{m_1}\otimes\Gamma_{m_2}\right)$ represents the complement of the correct decoding outcome. \begin{definition}% An $(n,R_1,R_2,\epsilon)$ broadcast channel code consists of a codebook $\{x^n(m_1,m_2)\}_{m_1\in \mathcal{M}_1, m_2\in \mathcal{M}_2}$ % and two decoding POVMs $\left\{ \Lambda_{m_1}\right\}_{m_1\in \mathcal{M}_1}$ and $\left\{ \Gamma_{m_2}\right\}_{m_2\in \mathcal{M}_2}$ such that the average probability of error $\overline{p}_{e}$ is bounded from above as \begin{align} \overline{p}_{e} & \equiv \frac{1}{|\mathcal{M}_1||\mathcal{M}_2|}\sum_{m_1,m_2}p_{e}\!\left( m_1,m_2\right) % \leq \epsilon. \end{align} \end{definition} A rate pair $\left( R_{1},R_{2}\right) $ is \textit{achievable} if there exists an $\left( n,R_{1}-\delta,R_{2}-\delta,\epsilon\right) $ quantum broadcast channel code for all $\epsilon,\delta>0$ and sufficiently large $n$. When devising coding strategies for c-q channels, the main obstacle to overcome is the construction of a decoding POVM that correctly decodes the messages. Given a set of positive operators $\{ \Pi^\prime_m \}$ which are suitable for detecting each message, we can construct a POVM by \emph{normalizing} them using the square-root measurement \cite{H98,SW97}: \begin{align} \Lambda_{m} & \equiv \left( \sum_{k}\Pi_{k}^{\prime} \right)^{\!\!\!-\frac{1}{2}} \!\! \Pi_{m}^{\prime} \left( \sum_{k}\Pi_{k}^{\prime} \right)^{\!\!-\frac{1}{2}}\!\!. \label{eq:square-root-POVM-generic} \end{align} Thus, the search for a decoding POVM is reduced to the problem of finding positive operators $\Pi^\prime_m$ apt at detecting and distinguishing the output states produced by each of the possible input messages ($\text{Tr}\!\left[\Pi^\prime_m\ \rho_m\right] \geq 1- \epsilon$ and $\text{Tr}\!\left[\Pi^\prime_{m}\ \rho_{m^\prime\neq m}\right] \leq \epsilon$). \section{Superposition coding inner bound} \label{sec:superposition-coding} % % % One possible strategy for the broadcast channel is to send a message at a rate that is low enough so that both receivers are able to decode. % Furthermore, if we assume that Receiver~1 has a better reception signal, then the sender can encode a further message \emph{superimposed} on top of the common message that Receiver 1 will be able to decode \emph{given} the common message. % The sender encodes the common message at rate $R_2$ using a codebook generated from a probability distribution $p_W(w)$, and the additional message for Receiver~1 at rate $R_1$ using a conditional codebook with distribution $p_{X|W}(x|w)$. % % % % % \begin{theorem}[Superposition coding inner bound] \label{thm:sup-coding-inner-bound} A rate pair $\left( R_{1},R_{2}\right) $ is achievable for the quantum broadcast channel in (\ref{eq:cqq-broadcast}) if it satisfies the following inequalities:% % % % % % % % % % % % % \begin{align} R_1 & \leq I(X; B_1 | W)_\theta, \\ R_2 &\leq I(W; B_2)_\theta, \\ R_1 + R_2 &\leq I(X;B_1)_\theta, \end{align} where the above information quantities are with respect to a state $\theta^{WXB_1B_2} $ of the form \be \sum_{w,x} p_W(w)p_{X|W}(x|w) \ \ketbra{w}{w}^{W} \otimes \ketbra{x}{x}^{X} \otimes \rho_x^{B_1B_2}.\label{eq:code-state} \ee \end{theorem} \begin{IEEEproof} The new idea in the proof is to exploit superposition encoding and a quantum simultaneous nonunique decoder for the decoding of the first receiver \cite{C72,B73} instead of the quantum successive decoding used in \cite{YHD2006}. We use a standard HSW\ decoder for the second receiver \cite{H98,SW97}. % % % \textbf{Codebook generation.} The sender randomly and independently generates $M_{2}$ sequences $w^{n}\!\left( m_{2}\right) $ according to the product distribution $ p_{W^{n}}\!\left( w^{n}\right) \equiv\prod\limits_{i=1}^{n}p_{W}\!\left(w_{i}\right)$. % For each sequence $w^{n}\!\left( m_{2}\right) $, the sender then randomly and conditionally independently generates $M_{1}$ sequences $x^{n}\!\left( m_{1},m_{2}\right) $ according to the product distribution: $p_{X^{n}|W^{n}}\!\left( x^{n}|w^{n}\!\left( m_{2}\right) \right) \equiv \prod\limits_{i=1}^{n}p_{X|W}\!\left( x_{i}|w_{i}\!\left( m_{2}\right) \right)$. % The sender then transmits the codeword $x^{n}\!\left( m_{1},m_{2}\right) $ if she wishes to send $\left( m_{1},m_{2}\right) $. \textbf{POVM\ Construction}. We now describe the POVMs that the receivers employ in order to decode the transmitted messages. % First consider the state we obtain from (\ref{eq:code-state}) by tracing over the $B_{2}$ system:% \[ \rho^{WXB_{1}}=\sum_{w,x}p_{W}\!\left( w\right) \ p_{X|W}\!\left( x|w\right) \ \left\vert w\right\rangle\!\left\langle w\right\vert ^{W}\otimes\left\vert x\right\rangle\!\left\langle x\right\vert ^{X}\otimes\rho_{x}^{B_{1}}. \] Further tracing over the $X$ system gives% \begin{align*} \rho^{WB_{1}} & =\sum_{w}p_{W}\!\left( w\right) \ \left\vert w\right\rangle\! \left\langle w\right\vert ^{W}\otimes\sigma_{w}^{B_{1}}, \end{align*} where $\sigma_{w}^{B_{1}}\equiv \sum_{x}p_{X|W}\!\left( x|w\right) \rho_{x}^{B_{1}}$. For the first receiver, we exploit a square-root decoding POVM\ as in \eqref{eq:square-root-POVM-generic} based on the following positive operators: % % % % % % % % % \be \Pi_{m_{1},m_{2}}^{\prime}\equiv\Pi\ \Pi_{\!W^{\!n}\!(m_2)} % \ \Pi_{X^{n}\left( m_{1},m_{2}\right) }\ \Pi_{\!W^{\!n}\!(m_2)} \ \Pi, \label{eq:rec-1-POVM} \ee where we have made the abbreviations% \begin{align*} \Pi \equiv\Pi_{\rho,\delta}^{B_{1}^{n}} & , \,\,\,\,\,\,\,\,\, \Pi_{\!W^{\!n}\!(m_2)} \equiv\Pi_{\sigma_{W^{n}\left( m_{2}\right) },\delta}^{B_{1}^{n}},\\ \Pi_{X^{n}\left( m_{1},m_{2}\right) } & \equiv\Pi_{\rho_{X^{n}\left( m_{1},m_{2}\right) },\delta}^{B_{1}^{n}}. \end{align*} The above projectors are weakly typical projectors \cite[Section 14.2.1]{wilde2011book} defined with respect to the states $\rho^{\otimes n}$, $\sigma^{B_{1}^{n}}_{W^{n}\left( m_{2}\right)}$, and $\rho^{B_{1}^{n}}_{X^{n}\left( m_{1},m_{2}\right)}$. Consider now the state in (\ref{eq:code-state}) as it looks from the point of view of Receiver~2. If we trace over the $X$ and $B_{1}$ systems, we obtain the following state:% \begin{align*} \rho^{WB_{2}} =\sum_{w}p_{W}\left( w\right) \ \left\vert w\right\rangle \left\langle w\right\vert ^{W}\otimes\sigma_{w}^{B_{2}}, \end{align*} where $\sigma_{w}^{B_{2}}\equiv \sum_{x}p_{X|W}\left( x|w\right) \rho_{x}^{B_{2}}$. For the second receiver, we exploit a standard HSW\ decoding POVM\ that is with respect to the above state---it is a square-root measurement as in \eqref{eq:square-root-POVM-generic}, based on the following positive operators: \be \Pi_{m_2}^{\prime B_{2}^{n} } =\Pi_{\rho,\delta}^{B_{2}^{n} }\ \Pi_{\sigma_{W^{\!n}\!(m_2)},\delta}^{B_{2}}\ \Pi_{\rho,\delta}^{B_{2}^{n}}, \label{eq:rec-2-POVM} \ee % % % % % % % % % % % where the above projectors are weakly typical projectors defined with respect to $\rho^{\otimes n}$ % and $\sigma^{B_{2}^{n}}_{W^{n}\left( m_{2}\right)}$. \textbf{Error analysis}. % We now analyze the expectation of the average error probability for the first receiver with the POVM defined by (\ref{eq:square-root-POVM-generic}) and (\ref{eq:rec-1-POVM}): \begin{multline*} \!\!\!\!\!\!\!\!\!\!\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \frac{1}{M_{1}M_{2}}\sum_{m_{1},m_{2}% }\text{Tr}\left\{ \left( I-\Gamma_{m_{1},m_{2}}^{B_{1}^{n}}\right) \rho_{X^{n}\left( m_{1},m_{2}\right) }^{B_{1}}\right\} \right\} \\ =\frac{1}{M_{1}M_{2}}\sum_{m_{1},m_{2}}\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \left( I-\Gamma_{m_{1},m_{2}}^{B_{1}^{n}}\right) \rho_{X^{n}\left( m_{1},m_{2}\right) }^{B_{1}}\right\} \right\}. \end{multline*} Due to the above exchange between the expectation and the average and the symmetry of the code construction (each codeword is selected randomly and independently), it suffices to analyze the expectation of the average error probability for the first message pair $\left( m_{1}=1,m_{2}=1\right) $, i.e., the last line above is equal to $ \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \left( I-\Gamma _{1,1}^{B_{1}^{n}}\right) \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} . $ Using the Hayashi-Nagaoka operator inequality (Lemma~\ref{lem:HN-inequality} in the appendix), % % % % % % % % % % % % % we obtain the following upper bound on this term: \begin{multline} % 2 \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \left( I-\Pi_{1,1}^{\prime}\right) \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \label{eq:after-HN}\\ \qquad \qquad +4 \!\!\!\!\!\!\! \sum_{\left( m_{1},m_{2}\right) \neq\left( 1,1\right) }\!\!\!\!\!\!\!\!\!\! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{m_{1},m_{2}}^{\prime} \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\}. \end{multline} We begin by bounding the term in the first line above. Consider the following chain of inequalities:% \begin{align*} & \!\!\!\!\!\!\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{1,1}^{\prime} \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \\ & = \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi\ \Pi_{\!W^{\!n}\!(1)} \Pi_{X^{n}\left( 1,1\right) }\ \Pi_{\!W^{\!n}\!(1)} \ \Pi \ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \\ & \geq\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{X^{n}\left( 1,1\right) }\ \ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \\ & \ \ \ \ \ -\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \left\Vert \rho_{X^{n}\left( 1,1\right) }^{B_{1}}-\Pi\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}% \ \Pi\right\Vert _{1}\right\} \\ & \ \ \ \ \ -\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \left\Vert \rho_{X^{n}\left(1,1\right) }^{B_{1}} -\Pi_{\!W^{\!n}\!(1)} \ \rho_{X^{n}\left(1,1\right) }^{B_{1}}\ \Pi_{\!W^{\!n}\!(1)} \right\Vert _{1}\right\} \\ & \geq1-\epsilon-4\sqrt{\epsilon}, \end{align*} where the first inequality follows from the inequality% \begin{equation} \text{Tr}\left\{ \Lambda\rho\right\} \leq\text{Tr}\left\{ \Lambda \sigma\right\} +\left\Vert \rho-\sigma\right\Vert _{1}, \label{eq:trace-inequality} \end{equation} which holds for all $\rho$, $\sigma$, and $\Lambda$ such that $0\leq \rho,\sigma,\Lambda\leq I$. The second inequality follows from the Gentle Operator Lemma for ensembles (see Lemma~\ref{lem:gentle-operator} in the appendix) and the properties of typical projectors for sufficiently large $n$. % % % % % % % % % % % % % % % % % % We now focus on bounding the term in the second line of (\ref{eq:after-HN}). We can expand this term as follows:% \be \begin{split} \sum_{m_{1}\neq1}\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{m_{1},1} ^{\prime}\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \qquad \qquad\\[-2mm] \qquad \qquad +\sum_{\substack{m_{1}, \\ m_{2} \neq1}}\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{m_{1},m_{2}}^{\prime}\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\}. \end{split} \label{eq:error-terms} \ee Consider the term in the first line of (\ref{eq:error-terms}):% \begin{align*} & \sum_{m_{1}\neq1} \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{m_{1},1}^{\prime}\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \\ & =\sum_{m_{1}\neq1}\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \text{Tr}\left\{ \Pi\ \Pi_{\!W^{\!n}\!(1)} \ \Pi_{X^{n}\left( m_{1},1\right) } \ \Pi_{\!W^{\!n}\!(1)} \ \Pi\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \\ & \leq K \sum_{\!m_{1}\neq1}\! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \! \left\{ \text{Tr}\!\left[ \Pi\ \Pi_{\!W^{\!n}\!(1)} \rho_{X^{n}\left( m_{1},1\right) } \Pi_{\!W^{\!n}\!(1)} \Pi\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right] \! \right\} \\ & = K \sum_{\!m_{1}\neq1} \mathop{\mathbb{E}}_{\scriptscriptstyle W^{\!n}} \bigg\{ \text{Tr}[ \Pi_{\!W^{\!n}\!(1)} \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!|\!W^{\!n}} \left\{ \rho_{X^{n}\left(m_{1},1\right) }\right\} \ \Pi_{\!W^{\!n}\!(1)} \ \\[-2mm] & \hspace{48mm} \Pi\ \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!|\!W^{\!n}} \left\{ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \Pi\ ] \bigg\} \\ & = K \sum_{m_{1}\neq 1} \!\! \mathop{\mathbb{E}}_{\scriptscriptstyle W^{\!n}} \!\! \left\{ \text{Tr}\left\{ \Pi\ \Pi_{\!W^{\!n}\!(1)} \sigma_{W^{n}\left( 1\right) }\ \Pi_{\!W^{\!n}\!(1)} \ \Pi\ \sigma_{W^{n}\left( 1\right) } \right\} \right\} \\ & \leq K \,2^{-n\left[ H\left( B_{1}|W\right) -\delta\right] } % \sum_{m_{1}\neq1} \!\! \mathop{\mathbb{E}}_{\scriptscriptstyle W^{\!n}} \!\! \left\{ \text{Tr}\left\{ \Pi\ \Pi_{\!W^{\!n}\!(1)} \ \Pi\ \sigma_{W^{n}\left( 1\right) }\right\} \right\} \\ & \leq K\, 2^{-n\left[H\left( B_{1}|W\right) -\delta\right] }\ \sum_{m_{1}\neq1} \mathop{\mathbb{E}}_{\scriptscriptstyle W^{\!n}} \!\! \left\{ \text{Tr}\left\{ \sigma_{W^{n}\left( 1\right) }\right\} \right\} \\ & \leq 2^{-n\left[ I\left( X;B_{1}|W\right) -2\delta\right] }\ M_{1}, \end{align*} where we define $K \equiv 2^{n\left[ H\left( B_{1}|WX\right) +\delta\right] }$. The first inequality is due to the \emph{projector trick} inequality \cite{GLM10,S11a,FHSSW11} which states that \begin{align} \Pi_{X^{n}\left( m_{1},1\right) } & \leq 2^{n\left[ H\left( B_{1}|WX\right) +\delta\right] }\,\rho_{X^{n}\left( m_1,1\right) }^{B_{1}}. \label{eq:projector-trick} % % % % % % % \end{align} The second inequality follows from the properties of typical projectors: $\Pi_{\!W^{\!n}\!(1)} \sigma_{W^{n}\left( 1\right) }\ \Pi_{\!W^{\!n}\!(1)} \leq 2^{-n\left[H\left( B_{1}|W\right) -\delta\right] }\Pi_{\!W^{\!n}\!(1)} $. Now consider the term in the second line of (\ref{eq:error-terms}):% \begin{align*} & \!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}}\!\!\!\! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \!\! \left\{ \text{Tr}\left\{ \Pi_{m_{1},m_{2}}^{\prime}\ \rho_{X^{n}\left( 1,1\right) }^{B_{1}}\right\} \right\} \\ & = \!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}} \!\!\!\! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \!\!\! \left\{ \text{Tr}\!\! \left[ \Pi \Pi_{W^{\!n}(m_2) } \Pi_{X^{n}\left( m_{1},m_{2}\right) } \Pi_{\!W^{\!n}\!(m_2)} \Pi\ \rho_{X^{n}\left(1,1\right) }^{B_{1}}\right] \right\} \\ & = \!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}} \text{Tr}\left[ \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \bigg\{ \Pi_{\!W^{\!n}\!(m_2)} \ \Pi_{X^{n}\left( m_{1},m_{2}\right) }\ \Pi_{\!W^{\!n}\!(m_2)} \right\} \\[-5mm] & \hspace{47mm} \Pi \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \rho_{X^{n}\left( 1,1\right) }^{B_{1} } \right\} \Pi\ \bigg] \\ & = \!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}} \!\! \text{Tr}\left\{ \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \!\!\left\{ \Pi_{\!W^{\!n}\!(m_2)} \Pi_{X^{n}\left( m_{1},m_{2}\right) } \Pi_{\!W^{\!n}\!(m_2)} \right\} \ \Pi \rho^{\otimes n} \Pi \right\} \\ & \leq 2^{-n\left[ H\left( B_{1}\right) -\delta\right] } \!\!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}} \!\!\! \text{Tr}\!\!\left[ \! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \!\!\! \left\{ \Pi_{\!W^{\!n}\!(m_2)} \Pi_{X^{n}\left( m_{1},m_{2}\right) } \Pi_{\!W^{\!n}\!(m_2)} \right\} \! \Pi \right] \\ & = 2^{-n\left[ H\left( B_{1}\right) -\delta\right] } \!\!\! \sum_{\substack{m_{1}, \\ m_{2} \neq1}} \!\! \mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \! \text{Tr}\left[ \Pi_{X^{n}\left( m_{1},m_{2}\right) } \Pi_{\!W^{\!n}\!(m_2)} \Pi \Pi_{\!W^{\!n}\!(m_2)} \right] \end{align*} \begin{align*} & \leq 2^{-n\left[ H\left( B_{1}\right) -\delta\right] } \sum_{m_{2} \neq1,\ m_{1}}\mathop{\mathbb{E}}_{\scriptscriptstyle X^{\!n}\!\!,W^{\!n}} \left\{ \text{Tr}\left\{ \Pi_{X^{n}\left( m_{1},m_{2}\right) }\right\} \right\} \\ & \leq 2^{-n\left[ H\left( B_{1}\right) -\delta\right] }\ 2^{n\left[ H\left( B_{1}|WX\right) +\delta\right] }\ M_{1}\ M_{2}\\ & = 2^{-n\left[ I\left( WX;B_{1}\right) -2\delta\right] }\ M_{1}\ M_{2}\\ & = 2^{-n\left[ I\left( X;B_{1}\right) -2\delta\right] }\ M_{1}\ M_{2}. \end{align*} % The equality $I(WX;B_1)=I(X;B_1)$ follows from the way the codebook is constructed (i.e., the Markov chain $W-X-B$), as discussed also in \cite{S11a}. This completes the error analysis for the first receiver. The proof for the second receiver is analogous to the point-to-point HSW theorem. The appendix gives the details for this and ties the coding theorem together so that the sender and two receivers can agree on a strategy that has asymptotically vanishing error probability in the large $n$ limit. \end{IEEEproof} \section{Marton coding scheme} \label{sec:marton} We now prove that the Marton inner bound is achievable for quantum broadcast channels. % % % % % The Marton scheme depends on auxiliary random variables $U_1$ and $U_2$, \emph{binning}, and the properties of strongly typical sequences and projectors. % \begin{theorem}[Marton inner bound] \label{thm:marton-no-common} Let % $\{\rho^{B_1B_2}_x \}$ % be a classical-quantum broadcast channel and $x=f(u_1,u_2)$ be a deterministic function. The following rate region is achievable: \bea R_1 &\leq& I(U_1; B_1)_\theta, \nonumber \\ R_2 &\leq& I(U_2; B_2)_\theta, \\ R_1+R_2 &\leq& I(U_1; B_1)_\theta + I(U_2; B_2)_\theta - I(U_1; U_2)_\theta, \nonumber \eea where the information quantities are with respect to the state: $$ \theta^{U_1U_2B_1B_2} =\! \sum_{u_1,u_2} \! p(u_1,u_2) \ketbra{u_1}{u_1}^{U_1} \otimes \ketbra{u_2}{u_2}^{U_2} \otimes \rho_{f(u_1,u_2)}^{B_1B_2}. $$ \end{theorem} \begin{IEEEproof} Consider the classical-quantum broadcast channel % $\{\mathcal{N}(x) \equiv \rho^{B_1B_2}_x \}$, % and a deterministic mixing function: $f: \mathcal{U}_1 \times \mathcal{U}_2 \to \mathcal{X}$. % Using the mixing function as a pre-coder to the broadcast channel $\mathcal{N}$, we obtain a channel $\mathcal{N'}$ defined as: \be % \mathcal{N}'(u_1,u_2) \equiv \rho^{B_1B_2}_{f(u_1,u_2)} \equiv \rho^{B_1B_2}_{u_1,u_2}. % \ee % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % \textbf{Codebook construction}. Define two auxiliary indices $\ell_1 \in [1, \ldots, L_1]$, $L_1= 2^{n[I(U_1;B_1)-\delta]}$ and $\ell_2 \in [1, \ldots, L_2]$, $L_2 = 2^{n[I(U_2;B_2)-\delta]}$. For each $\ell_1$ generate an i.i.d.~random sequence $u_1^n(\ell_1)$ according to $p_{U_1^n}(u_1^n)$. Similarly we choose $L_2$ random i.i.d.~sequences $u_2^n(\ell_2)$ according to $p_{U_2^n}(u_2^n)$. % Partition the sequences $u_1^n(\ell_1)$ into $2^{nR_1}$ different bins $B_{m_1}$. Similarly, partition the sequences $u_2^n(\ell_2)$ into $2^{nR_2}$ bins $C_{m_2}$. % For each message pair $(m_1,m_2)$, % the sender selects a sequence $\big(u_1^n(\ell_1),u_2^n(\ell_2) \big) \in \left( B_{m_1} \times C_{m_2} \right) \ \cap \ \mathcal{A}^n_{p_{U_1U_2},\delta}$, such that each sequence is taken from the appropriate bin and the sender demands that they are strongly jointly-typical (otherwise declaring failure). The codebook $x^n(m_1,m_2)$ is deterministically constructed from $\big(u_1^n(\ell_1),u_2^n(\ell_2) \big)$ by applying the function $x_i=f(u_{1i},u_{2i})$. % % % % % % % % % % % % % % % % \textbf{Transmission}. Let $(\ell_1,\ell_2)$ denote the pair of indices of the joint sequence $(u_1^n(\ell_1), u_2^n(\ell_2))$ which was chosen as the codeword for message $(m_1,m_2)$. % % Expressed in terms of these indices the output of the channel is \be \rho_{u_1^n(\ell_1),u_2^n(\ell_2)}^{B_1^nB_2^n} = \bigotimes_{i \in [n]} \rho_{f(u_{1i}(\ell_1),u_{2i}(\ell_2))}^{B_1B_2} % % % % % % \equiv \rho_{\ell_1,\ell_2}. \ee Define the following states: \be \omega^{B_1}_{u_1} \equiv \sum_{u_2} p_{U_2|U_1}(u_2|u_1) \rho^{B_1}_{u_1,u_2}, \,\,\,\,\, \label{avg-both-single-copy} \bar{\rho} \equiv \sum_{u_1} p(u_1) \omega^{B_1}_{u_1}. \ee \textbf{Decoding}. % % % The detection POVM for Receiver 1, \ $\left\{ \Lambda_{\ell_1}\right\}_{\ell_1 \in [1,\ldots,L_1]}$, is constructed by using the square-root measurement as in \eqref{eq:square-root-POVM-generic} based on the following combination of strongly typical projectors: % % % % % % % % % % % % % \begin{align} % % % % % % \Pi_{\ell_1}^{\prime} \equiv \Pi^n_{\bar{\rho},\delta} \ \Pi_{u^n_1(\ell_1)} \ \Pi^n_{\bar{\rho},\delta}. % % % % \end{align} % % % % The projectors $\Pi_{u^n_1(\ell_1)}$ and $\Pi^n_{\bar{\rho},\delta}$ are defined with respect to the states $\omega_{u_1^n(\ell_1) }$ and $\bar{\rho}^{\otimes n}$ given in (\ref{avg-both-single-copy}). % % % % % Note that we use {\it strongly} typical projectors in this case \cite[Section 14.2.3]{wilde2011book}). % % Knowing $\ell_1$ and the binning scheme, Receiver~1 can deduce the message $m_1$ from the bin index. Receiver~2 uses a similar decoding strategy to obtain $\ell_2$ and infer $m_2$. % % % % % % \textbf{Error analysis}. An error occurs if one (or more) of the following error events occurs. \begin{description} \item[$E_0$:] An encoding error occurs whenever there is no jointly typical sequence in $B_{m_1} \times C_{m_2}$ for some message pair $(m_1,m_2)$. \item[$E_1$:] A decoding error occurs at Receiver 1 if $L_1^\prime \neq \ell_1$. \item[$E_2$:] A decoding error occurs at Receiver 2 if $L_2^\prime \neq \ell_2$. \end{description} The probability of an encoding error $E_0$ is bounded like in the classical Marton scheme \cite{M79,el2010lecture,cover1998comments}. % To see this, we use Cover's counting argument \cite{cover1998comments}. The probability that two random sequences $u_1^n$, $u_2^n$ chosen according to the marginals are jointly typical is $2^{-nI(U_1;U_2)}$ and since there are on average $2^{n[I(U_1;B_1)-R_1]}$ and $2^{n[I(U_2;B_2)-R_2]}$ sequences in each bin, the expected number of jointly-typical sequences that can be constructed from each combination of bins is \be 2^{n[I(U_1;B_1)-R_1]} 2^{n[I(U_2;B_2)-R_2]}2^{-nI(U_1;U_2)}. \ee Thus, if we choose $R_1 + R_2 +\delta \leq I(U_1; B_1) + I(U_2; B_2) - I(U_1; U_2)$, then the expected number of % % strongly jointly-typical sequences in $B_{m_1} \times C_{m_2}$ is much larger than one. To bound the probability of error event $E_1$, we use the Hayashi-Nagaoka operator inequality (Lemma~\ref{lem:HN-inequality}): \begin{align} & \Pr(E_1) = \frac{1}{L_1} \sum_{\ell_1} {\rm Tr}\;\left[ (I - \Lambda_{\ell_1}) \rhoNulul \right] \nonumber \\ & \leq \frac{1}{L_1 L_2} \sum_{\ell_1} \bigg( 2\underbrace{ {\rm Tr}\;\!\!\!\left[ (I - \Pi^n_{\bar{\rho},\delta} \Pi_{u^n_1(\ell_1)} \Pi^n_{\bar{\rho},\delta} ) \rhoNulul \right]}_{(T1) } % \nonumber \\[-2mm] & \qquad\qquad\qquad \ \ \ \ + 4 \underbrace{ \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \Pi_{u^n_1(\ell_1^{\prime})} \Pi^n_{\bar{\rho},\delta} \rhoNulul \right] }_{(T2)} \bigg). \nonumber \\[-7mm] \nonumber % \end{align}% Consider the following lemma \cite[Property 14.2.7]{wilde2011book}. % % % % \begin{lemma} \label{lemma-avg-typ-proj-works} The state $\rhoNulul$ is well supported by both the averaged state projector: $ {\rm Tr}\;\!\!\big[ \Pi^n_{\bar{\rho},\delta} \ \rhoNulul \big] \geq 1 - \epsilon, \ \forall \ell_1, \ell_2, $ and % % % the $\omega^{B_1}_{u_1}$ conditionally typical projector: $ {\rm Tr}\;\!\!\!\left[ \Pi_{u^n_1(\ell_1)} \ \rhoNulul \right] \geq 1 - \epsilon, \ \forall \ell_2, $ when $u_1^n(\ell_1)$ and $u_2^n(\ell_2)$ are strongly jointly typical. % \end{lemma} To bound the first term (T1), we use the following argument: % \begin{align} 1 - (T1) & = {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \Pi_{u^n_1(\ell_1)} \Pi^n_{\bar{\rho},\delta} \ \rhoNulul \right] \nonumber \\ & = {\rm Tr}\;\!\!\!\left[ \Pi_{u^n_1(\ell_1)} \ \Pi^n_{\bar{\rho},\delta} \rhoNulul \Pi^n_{\bar{\rho},\delta} \right] \nonumber \\ & \geq % {\rm Tr}\;\!\!\!\left[ \Pi_{u^n_1(\ell_1)} \ \rhoNulul \right] % % % - \| \Pi^n_{\bar{\rho},\delta} \rhoNulul \Pi^n_{\bar{\rho},\delta} - \rhoNulul \|_1 \nonumber \\ & \geq (1 - \epsilon) - 2\sqrt{\epsilon}, \end{align} where the inequalities follow from \eqref{eq:trace-inequality} and Lemma~\ref{lemma-avg-typ-proj-works}. This use of Lemma~\ref{lemma-avg-typ-proj-works} demonstrates why the Marton coding scheme selects the sequences $u_1^n(\ell_1)$ and $u_2^n(\ell_2)$ such that they are strongly jointly typical. To bound the second term, we begin by applying a variant of the projector trick from \eqref{eq:projector-trick}. In the below, note that the expectation $ \mathop{\mathbb{E}}_{U_1,U_2} $ over the random code is with respect to the product distribution $p_{U_1^n}(u_1^n) p_{U_2^n}(u_2^n)$: \begin{align*} & \!\!\! \mathop{\mathbb{E}}_{U_1,U_2} \!\!\left\{ (T2) \right\} = \mathop{\mathbb{E}}_{U_1,U_2} \left\{ \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \Pi_{U^n_1(\ell_1^\prime)} \Pi^n_{\bar{\rho},\delta} \ \rhoNulul \right] \right\} \\ % % & \leq 2^{n[H(B_1|U_1)+\delta]} \mathop{\mathbb{E}}_{U_1,U_2} \left\{ \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \ \omgNulpr \ \Pi^n_{\bar{\rho},\delta} \ \rhoNulul \right] \right\}.\\ \intertext{We continue the proof using averaging over the choice of codebook and the properties of typical projectors:} & = 2^{n[H(B_1|U_1)+\delta]} \mathop{\mathbb{E}}_{U_2} \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \ \mathop{\mathbb{E}}_{U_1} \!\left\{ \omgNulpr \right\} \ \Pi^n_{\bar{\rho},\delta} \mathop{\mathbb{E}}_{U_1} \!\left\{\rhoNulul\right\} \right] \\ & = 2^{n[H(B_1|U_1)+\delta]} \mathop{\mathbb{E}}_{U_2} \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \ \bar{\rho} \ \Pi^n_{\bar{\rho},\delta} \ \mathop{\mathbb{E}}_{U_1} \!\left\{\rhoNulul\right\} \right] \\ & \leq 2^{n[H(B_1|U_1)+\delta]} 2^{-n[H(B_1)- \delta]} \mathop{\mathbb{E}}_{U_1,U_2} \sum_{\ell'_1\neq \ell_1} {\rm Tr}\;\!\!\!\left[ \Pi^n_{\bar{\rho},\delta} \ \rhoNulul \right] \\ & \leq 2^{n[H(B_1|U_1)+\delta]} 2^{-n[H(B_1)- \delta]} \mathop{\mathbb{E}}_{U_1,U_2} \sum_{\ell'_1\neq \ell_1} 1 \\ & \leq L_1 2^{-n[I(U_1;B_1) - 2\delta]}. \end{align*} Therefore, if we choose $2^{nR_1} = L_1 \leq 2^{n[I(U_1;B_1) - 3\delta]}$, the probability of error will go to zero in the asymptotic limit of many channel uses. The analysis of the event $E_2$ is similar. % \end{IEEEproof} \section{Conclusion} \label{sec:conclusion} We have proved quantum generalizations of the superposition coding inner bound \cite{C72,B73} and the Marton rate region with no common message \cite{M79}. The key ingredient in both proofs was the use of the projector trick. A natural followup question would be to combine the two strategies to obtain the Marton coding scheme with a common message. A much broader goal would be to extend all of network information theory to the study of quantum channels. To accomplish this goal, it would be helpful to have a powerful tool that generalizes El Gamal and Kim's classical packing lemma \cite{el2010lecture} to the quantum domain. The packing lemma is sufficient to prove all of the known coding theorems in network information theory. % At the moment, it is not clear to us whether such a tool exists for the quantum case, but evidence in favor of its existence is that \ 1) one can prove the HSW coding theorem by using conditionally typical projectors only \cite[Exercise 19.3.5]{wilde2011book}, \ 2) we have solved the quantum simultaneous decoding conjecture for the case of two senders \cite{FHSSW11,S11a}, and \ 3) we have generalized two important coding theorems in the current paper (with proofs somewhat similar to the classical proofs). Ideally, such a tool would allow quantum information theorists to prove quantum network coding theorems by appealing to it, rather than having to analyze each coding scheme in detail on a case by case basis. We acknowledge discussions with Patrick Hayden, Omar Fawzi, Pranab Sen, and Saikat Guha during the development of \cite{FHSSW11,S11a}. % I.~Savov acknowledges support from FQRNT and NSERC. M.~M.~Wilde acknowledges support from the Centre de Recherches Math\'ematiques. \bibliographystyle{IEEEtran}
\section*{Introduction} The invariant subspace problem asks whether every bounded linear operator on an infinite dimensional separable Banach space admits a non-trivial closed invariant subspace. A classical result of M. Aronszajn and K.T. Smith \cite{AS} asserts that the problem has a positive answer for compact operators. This result was extended by V. Lomonosov \cite{L} for operators on complex Banach spaces that commute with a non-trivial compact operator. Recently G. Sirotkin \cite{Sir} has presented a version of Lomonosov's theorem for real spaces. It is also known that the problem, in its full generality, has a negative answer. Indeed P. Enflo \cite{E} and subsequently C. J. Read \cite{R1},\cite{R2} have provided several examples of operators on non-reflexive Banach spaces that do not admit a non-trivial invariant subspace. In particular, in a profound study concerning spaces admitting operators without non trivial invariant subspaces, C. J. Read has proven that every separable Banach space that contains either $c_0$ or a complemented subspace isomorphic to $\ell_1$ or $J_\infty$, admits an operator without non trivial closed invariant subspaces \cite{R4}. A comprehensive study of Read's methods of constructing operators with no non trivial invariant subspaces can be found in \cite{GR1}, \cite{GR2}. Also recently a non-reflexive hereditarily indecomposable (HI) Banach space $\mathfrak{X}_K$ with the ``scalar plus compact'' property has been constructed \cite{AH}. This is a $\mathcal{L}_\infty$ space with separable dual, resulting from a combination of HI techniques with the fundamental J. Bourgain and F. Delbaen construction \cite{BD}. As consequence, the space $\mathfrak{X}_K$ satisfies the Invariant Subspace Property (ISP). Moreover, recently a $\mathcal{L}_\infty$ space, containing $\ell_1$ isomorphically, with the scalar plus compact property, has been constructed \cite{AHR}. Let us point out, that the latter shows that Read's result concerning separable spaces containing $\ell_1$ as a complemented subspace, is not extended to separable spaces containing $\ell_1$. All the above results provide no information in either direction within the class of reflexive Banach spaces. The importance of a result in this class, concerning the Invariant Subspace Problem, is reflected in the concluding phrase of C. J. Read in \cite{R4} where the following is stated. ``It is clear that we cannot go much further until and unless we solve the invariant subspace problem on a reflexive Banach space.'' The aim of the present work is to construct the first example of a reflexive Banach space $\mathfrak{X}_{_{^\text{ISP}}}$ with the Invariant Subspace Property, which is also the example of a Banach space with the hereditary Invariant Subspace Property. This property is not proved for the aforementioned space $\mathfrak{X}_K$. It is notable that no subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ has the ``scalar plus compact'' property. More precisely, the strictly singular operators\footnote[1]{A bounded linear operator is called strictly singular, if its restriction on any infinite dimensional subspace is not an isomorphism.} on every subspace $Y$ of $\mathfrak{X}_{_{^\text{ISP}}}$ form a non separable ideal (in particular, the strictly singular non-compact are non-separable). The space $\mathfrak{X}_{_{^\text{ISP}}}$ is a hereditarily indecomposable space and every operator $T\in\mathcal{L}(\mathfrak{X}_{_{^\text{ISP}}})$ is of the form $T = \lambda I + S$ with $S$ strictly singular. We recall that there are strictly singular operators in Banach spaces without non-trivial invariant subspaces \cite{R3}. On the other hand, there are spaces where the ideal of strictly singular operators does not coincide with the corresponding one of compact operators and every strictly singular operator admits a non-trivial invariant subspace. The most classical spaces with this property are $L^p[0,1], 1\leqslant p <\infty$ and $C[0,1]$. This is a combination of Lomonosov-Sirotkin theorem and the classical result, due to V. Milman \cite{M}, that the composition $TS$ is a compact operator, for any $T,S$ strictly singular operators, on any of the above spaces. In \cite{ABM}, Tsirelson like spaces satisfying similar properties are presented. The possibility of constructing a reflexive space with ISP without the ``scalar plus compact'' property emerged from an earlier version of \cite{ABM}. The following describes the main properties of the space $\mathfrak{X}_{_{^\text{ISP}}}$. \begin{thm*} There exists a reflexive space $\mathfrak{X}_{_{^\text{ISP}}}$ with a Schauder basis $\{e_n\}_{n\in\mathbb{N}}$ satisfying the following properties. \begin{enumerate} \item[(i)] The space $\mathfrak{X}_{_{^\text{ISP}}}$ is hereditarily indecomposable. \item[(ii)] Every seminormalized weakly null sequence $\{x_n\}_{n\in\mathbb{N}}$ has a subsequence generating either $\ell_1$ or $c_0$ as a spreading model. Moreover every infinite dimensional subspace $Y$ of $\mathfrak{X}_{_{^\text{ISP}}}$ admits both $\ell_1$ and $c_0$ as spreading models. \item[(iii)] For every $Y$ infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and every $T\in\mathcal{L}(Y,\mathfrak{X}_{_{^\text{ISP}}}),\; T = \lambda I_{_{Y,\mathfrak{X}_{_{^\text{ISP}}}}} + S$ with $S$ strictly singular. \item[(iv)] For every $Y$ infinite dimensional subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ the ideal $\mathcal{S}(Y)$ of the strictly singular operators is non separable. \item[(v)] For every $Y$ subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and every $Q,S,T$ in $\mathcal{S}(Y)$ the operator $QST$ is compact. Hence for every $T\in\mathcal{S}(Y)$ either $T^3 = 0$ or $T$ commutes with a non zero compact operator. \item[(vi)] For every $Y$ infinite dimensional closed subspace of $X$ and every $T\in\mathcal{L}(Y)$, $T$ admits a non-trivial closed invariant subspace. In particular every $T\neq\lambda I_Y$, for $\lambda\in\mathbb{R}$ admits a non-trivial hyperinvariant subspace. \end{enumerate} \end{thm*} It is not clear to us if the number of operators in property (v) can be reduced. For defining the space $\mathfrak{X}_{_{^\text{ISP}}}$ we use classical ingredients like the coding function $\sigma$, the interaction between conditional and unconditional structure, but also some new ones which we are about to describe. In all previous HI constructions, one had to use a mixed Tsirelson space as the unconditional frame on which the HI norm is built. Mixed Tsirelson spaces appeared with Th. Schlumprecht space \cite{Sch}, twenty years after Tsirelson construction \cite{T}. They became an inevitable ingredient for any HI construction, starting with the W.T. Gowers and B. Maurey celebrated example \cite{GM}, and followed by myriads of others \cite{AD2},\cite{AT} etc. The most significant difference in the construction of $\mathfrak{X}_{_{^\text{ISP}}}$ from the classical ones, is that it uses as an unconditional frame the Tsirelson space itself. As it is clear to the experts, HI constructions based on Tsirelson space, are not possible if we deal with a complete saturation of the norm. Thus the second ingredient involves saturation under constraints. This method was introduced by E. Odell and Th. Schlumprecht \cite{OS1},\cite{OS2} for defining heterogeneous local structure in HI spaces, a method also used in \cite{ABM}. By saturation under constraints we mean that the operations $(\frac{1}{2^n},\mathcal{S}_n)$ (see Remark \ref{remxx1.5}) are applied on very fast growing families of averages, which are either $\alpha$-averages or $\beta$-averages. The $\alpha$-averages have been also used in \cite{OS1},\cite{OS2}, while $\beta$-averages are introduced to control the behaviour of special functionals. It is notable that although the $\alpha,\beta$-averages do not contribute to the norm of the vectors in $\mathfrak{X}_{_{^\text{ISP}}}$, they are able to neutralize the action of the operations $(\frac{1}{2^n},\mathcal{S}_n)$ on certain sequences and thus $c_0$ spreading models become abundant. This significant property yields the structure of $\mathfrak{X}_{_{^\text{ISP}}}$ described in the above theorem. Let us briefly describe some further structural properties of the space $\mathfrak{X}_{_{^\text{ISP}}}$. The first and most crucial one is that for a $(n,\varepsilon)$ special convex combination (see Definition \ref{defscc}) $\sum_{i\in F}c_ix_i$, with $\{x_i\}_{i\in F}$ a finite normalized block sequence, we have that \begin{equation*} \|\sum_{i\in F}c_ix_i\|\leqslant \frac{6}{2^n} + 12\varepsilon \end{equation*} This evaluation is due to the fact that the space is built on Tsirelson space and differs from the classical asymptotic $\ell_1$ HI spaces (i.e. \cite{AD2},\cite{AT}) where seminormalized $(n,\varepsilon)$ special convex combinations exist in every block subspace. A consequence of the above, is that the frequency of the appearance of RIS sequences is significantly increased, which among others yields the following. Every strictly singular operator maps sequences generating $c_0$ spreading models to norm null ones. Furthermore, we classify weakly null sequences into sequences of rank 0, namely norm null ones, sequences of rank 1, namely sequences generating $c_0$ as a spreading model and sequences of rank 2 or 3, namely sequences generating $\ell_1$ as a spreading model. The main result concerning these ranks is the following. If $Y$ is an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T$ is a strictly singular operator on $Y$, then it maps sequences of non zero rank, to sequences of strictly smaller rank. Combining the above properties we conclude property (v) of the above theorem.\vskip5pt We thank G. Costakis for bringing to our attention G. Sirotkin's paper \cite{Sir}. \section{The norming set of the space $\mathfrak{X}_{_{^\text{ISP}}}$} In this section we define the norming set $W$ of the space $\mathfrak{X}_{_{^\text{ISP}}}$. This set is defined with the use of the sequence $\{\mathcal{S}_n\}_n$ which we remind below and also families of $\mathcal{S}_n$-admissible functionals. As we have mentioned in the introduction, the set $W$ will be a subset of the norming set $W_T$ of the Tsirelson space. \subsection*{The Schreier families} The Schreier families is an increasing sequence of families of finite subsets of the naturals, first appeared in \cite{AA}, inductively defined in the following manner. Set $\mathcal{S}_0 = \big\{\{n\}: n\in\mathbb{N}\big\}$ and $\mathcal{S}_1 = \{F\subset\mathbb{N}: \#F\leqslant\min F\}$. Suppose that $\mathcal{S}_n$ has been defined and set $\mathcal{S}_{n+1} = \{F\subset\mathbb{N}: F = \cup_{j = 1}^k F_j$, where $F_1 <\cdots< F_k\in\mathcal{S}_n$ and $k\leqslant\min F_1\}$ If for $n,m\in\mathbb{N}$ we set $\mathcal{S}_n*\mathcal{S}_m = \{F\subset\mathbb{N}: F = \cup_{j = 1}^k F_j$, where $F_1 <\cdots< F_k\in\mathcal{S}_m$ and $\{\min F_j: j=1,\ldots,k\}\in\mathcal{S}_n\}$, then it is well known that $\mathcal{S}_n*\mathcal{S}_m = \mathcal{S}_{n+m}$. \begin{ntt} A sequence of vectors $x_1 <\cdots<x_k$ in $c_{00}$ is said to be $\mathcal{S}_n$-admissible if $\{\min\supp x_i: i=1,\ldots,k\}\in\mathcal{S}_n$. Let $G\subset c_{00}$. A vector $f\in G$ is said to be an average of size $s(f) = n$, if there exist $f_1,\ldots,f_d\in G, d\leqslant n$, such that $f = \frac{1}{n}(f_1+\cdots+f_d)$. A sequence $\{f_j\}_j$ of averages in $G$ is said to be very fast growing, if $f_1<f_2<\ldots$, $s(f_j)>2^{\max\supp f_{j-1}}$ and $s(f_j) > s(f_{j-1})$ for $j>1$. \end{ntt} \subsection*{The coding function} Choose $L = \{\ell_k: k\in\mathbb{N}\}, \ell_1 >2$ an infinite subset of the naturals such that: \begin{enumerate} \item[(i)] For any $k\in\mathbb{N}$ we have that $\ell_{k+1} > 2^{2\ell_k}$ and \item[(ii)] $\sum_{k=1}^\infty \frac{1}{2^{\ell_k}}<\frac{1}{1000}$. \end{enumerate} Decompose $L$ into further infinite subsets $L_1, L_2$. Set \begin{eqnarray*} \mathcal{Q} &=& \big\{\big((f_1,n_1),\ldots,(f_m, n_m)\big): m\in\mathbb{N}, \{n_k\}_{k=1}^m\subset\mathbb{N}, f_1 <\ldots <f_m\in c_{00}\\ &&\text{with}\;f_k(i)\in\mathbb{Q},\;\text{for}\;i\in\mathbb{N}, k=1,\ldots,m\} \end{eqnarray*} Choose a one to one function $\sigma:\mathcal{Q}\rightarrow L_2$, called the coding function, such that for any $\big((f_1,n_1),\ldots,(f_m, n_m)\big)\in\mathcal{Q}$, we have that \begin{equation*} \sigma\big((f_1, n_1),\ldots,(f_m,n_m)\big) > 2^{n_m}\cdot\max\supp f_m \end{equation*} \begin{rmk} For any $n\in\mathbb{N}$ we have that $\#L\cap\{n,\ldots,2^{2n}\}\leqslant 1$. \label{remark1.1} \end{rmk} \subsection*{The norming set} The norming set $W$ is defined to be the smallest subset of $c_{00}$ satisfying the following properties:\vskip3pt \noindent {\bf 1.} The set $\{\substack{+\\[-2pt]-} e_n\}_{n\in\mathbb{N}}$ is a subset of $W$, for any $f\in W$ we have that $-f\in W$, for any $f\in W$ and any $I$ interval of the naturals we have that $If\in W$ and $W$ is closed under rational convex combinations. Any $f = \substack{+\\[-2pt]-} e_n$ will be called a functional of type 0.\vskip3pt \noindent {\bf 2.} The set $W$ contains any functional $f$ which is of the form $f = \frac{1}{2^n}\sum_{j=1}^d\alpha_j$, where $\{\alpha_j\}_{j=1}^d$ is an $\mathcal{S}_n$-admissible and very fast growing sequence of $\alpha$-averages in $W$. If $I$ is an interval of the naturals, then $g = \substack{+\\[-2pt]-}If$ is called a functional of type I$_\alpha$, of weight $w(g) = n$.\vskip3pt \noindent {\bf 3.} The set $W$ contains any functional $f$ which is of the form $f = \frac{1}{2^n}\sum_{j=1}^d\beta_j$, where $\{\beta_j\}_{j=1}^d$ is an $\mathcal{S}_n$-admissible and very fast growing sequence of $\beta$-averages in $W$. If $I$ is an interval of the naturals, then $g = \substack{+\\[-2pt]-}If$ is called a functional of type I$_\beta$, of weight $w(g) = n$.\vskip3pt \noindent {\bf 4.} The set $W$ contains any functional $f$ which is of the form $f = \frac{1}{2}\sum_{j=1}^df_j$, where $\{f_j\}_{j=1}^d$ is an $\mathcal{S}_1$-admissible special sequence of type I$_\alpha$ functionals. This means that $w(f_1)\in L_1$ and $w(f_j) = \sigma\big(\big(f_1,w(f_1)\big),\ldots,\big(f_{j-1},w(f_{j-1})\big)\big)$, for $j>1$. If $I$ is an interval of the naturals, then $g = \substack{+\\[-2pt]-}If$ is called a functional of type II with weights $\widehat{w}(g) = \{w(f_j) : \ran f_j\cap I\neq\varnothing\}$.\vskip3pt We call an $\alpha$-average any average $\alpha\in W$ of the form $\alpha = \frac{1}{n}\sum_{j=1}^df_j, d\leqslant n$, where $f_1<\cdots<f_d\in W$. We call a $\beta$-average any average $\beta\in W$ of the form $\beta = \frac{1}{n}\sum_{j=1}^df_j, d\leqslant n$, where $f_1,\ldots,f_d\in W$ are functionals of type II, with disjoint weights $\widehat{w}(f_j)$. In general, we call a convex combination any $f\in W$ that is not of type 0, I$_\alpha$, I$_\beta$ or II.\vskip3pt For $x\in c_{00}$ define $\|x\| = \sup\{f(x): f\in W\}$ and $\mathfrak{X}_{_{^\text{ISP}}} = \overline{(c_{00}(\mathbb{N}),\|\cdot\|)}$. Evidently $\mathfrak{X}_{_{^\text{ISP}}}$ has a bimonotone basis. One may also describe the norm on $\mathfrak{X}_{_{^\text{ISP}}}$ with an implicit formula. Indeed, for some $x\in\mathfrak{X}_{_{^\text{ISP}}}$, we have that \begin{equation*} \|x\| = \max\big\{\|x\|_0,\;\|x\|_{II},\;\sup\{\frac{1}{2^n}\sum_{j=1}^d\|E_jx\|_{k_j}^\alpha\},\;\sup\{\frac{1}{2^n}\sum_{j=1}^d\|E_jx\|_{k_j}^\beta\}\big\} \end{equation*} where the inner suprema are taken over all $n\in\mathbb{N}$, all $\mathcal{S}_n$-admissible intervals $\{E_j\}_{j=1}^d$ of the naturals and $k_1<\cdots<k_d$ such that $k_j > 2^{\max E_{j-1}}$ for $j>1$. By $\|x\|_{II}$ we denote\\ $\|x\|_{II} = \sup\{f(x): f\in W$ is a functional of type II$\}$\\ whereas for $j\in\mathbb{N}$, by $\|x\|_j^\alpha$ we denote\\ $\|x\|_j^\alpha = \sup\{\alpha(x): \alpha\in W$ is an $\alpha$-average of size $s(\alpha) = j\}$ Similarly, by $\|x\|_j^\beta$ we denote\\ $\|x\|_j^\beta = \sup\{\beta(x): \beta\in W$ is a $\beta$-average of size $s(\beta) = j\}$. \begin{rmk} Very fast growing sequences of $\alpha$-averages have been considered by E. Odell and Th. Schlumprecht in \cite{OS1}, \cite{OS2} and were also used in \cite{ABM}. However, $\beta$-averages are a new ingredient, introduced to control the behaviour of type II functionals on block sequences. The $\beta$-averages can also be used to provide an alternative and simpler approach of the main result in \cite{OS2}. As we have mentioned in the introduction, the $\|z\|_j^\alpha, \|z\|_j^\beta$, which are averages, do not contribute to the norm of the vector $z$. On the other hand, the $\{\|\cdot\|_j^\alpha\}_j, \{\|\cdot\|_j^\beta\}_j$ have a significant role for the structure of the space $\mathfrak{X}_{_{^\text{ISP}}}$. \end{rmk} \begin{rmk} The norming set $W$ can be inductively constructed to be the union of an increasing sequence of subsets $\{W_m\}_{m=0}^\infty$ of $c_{00}$, where $W_0 = \{\substack{+\\[-2pt]-} e_n\}_{n\in\mathbb{N}}$ and if $W_m$ has been constructed, then set $W_{m+1}^\alpha$ to be the closure of $W_m$ under taking $\alpha$-averages, $W_{m+1}^{\text{I}_\alpha}$ to be the closure of $W_{m+1}^\alpha$ under taking type I$_\alpha$ functionals, $W_{m+1}^{\text{I}_\beta}$ to be the closure of $W_{m+1}^{\text{I}_\alpha}$ under taking type I$_\beta$ functionals, $W_{m+1}^{\text{II}}$ to be the closure of $W_{m+1}^{\text{I}_\beta}$ under taking type II functionals, $W_{m+1}^\beta$ to be the closure of $W_{m+1}^{\text{II}}$ under taking $\beta$-averages and finally $W_{m+1}$ to be the closure of $W_{m+1}^\beta$ under taking rational convex combinations. \label{remark1.2} \end{rmk} \subsection*{Tsirelson space} Tsirelson's initial definition \cite{T} of the first Banach space not containing any $\ell_p, 1\leqslant p<\infty$ or $c_0$, concerned the dual of the so called Tsirelson norm which was introduced by T. Figiel and W. B. Johnson \cite{FJ} and satisfies the following implicit formula. \begin{equation*} \|x\|_T = \max\big\{\|x\|_0,\;\sup\{\frac{1}{2}\sum_{j=1}^d\|E_jx\|_T\}\big\} \end{equation*} where $x\in c_{00}$ and the inner supremum is taken over all successive subsets of the naturals $d \leqslant E_1 <\cdots <E_d$. Tsirelson space $T$ is defined to be the completion of $(c_{00},\|\cdot\|_T)$. In the sequel by Tsirelson norm and Tsirelson space we will mean the norm and the corresponding space from \cite{FJ}. As is well known, a norming set $W_T$ of Tsirelson space is the smallest subset of $c_{00}$ satisfying the following properties.\vskip3pt \noindent {\bf 1.} The set $\{\substack{+\\[-2pt]-} e_n\}_{n\in\mathbb{N}}$ is a subset of $W_T$, for any $f\in W_T$ we have that $-f\in W_T$, for any $f\in W_T$ and any $E$ subset of the naturals we have that $Ef\in W_T$ and $W_T$ is closed under rational convex combinations.\vskip3pt \noindent {\bf 2.} The set $W_T$ contains any functional $f$ which is of the form $f = \frac{1}{2}\sum_{j=1}^df_j$, where $\{f_j\}_{j=1}^d$ is a $\mathcal{S}_1$ admissible sequence in $W_T$. \begin{rmk} The following are well known facts about Tsirelson space.\label{remark1.3} \begin{enumerate} \item[(i)] The norming set $W_T$ can be inductively constructed to be the union of an increasing sequence of subsets $\{W_T^m\}_{m=0}^\infty$ of $c_{00}$, in a similar manner as above. \item[(ii)] The set $W_T^\prime$, which is the smallest subset of $c_{00}$ satisfying the following properties, also is a norming set for Tsirelson space.\vskip3pt \noindent {\bf 1.} The set $\{\substack{+\\[-2pt]-} e_n\}_{n\in\mathbb{N}}$ is a subset of $W_T^\prime$, for any $f\in W_T^\prime$ we have that $-f\in W_T^\prime$ and for any $f\in W_T^\prime$ and any $E$ subset of the naturals we have that $Ef\in W_T^\prime$.\vskip3pt \noindent {\bf 2.} The set $W_T^\prime$ contains any functional $f$ which is of the form $f = \frac{1}{2}\sum_{j=1}^df_j$, where $\{f_j\}_{j=1}^d$ is a $\mathcal{S}_1$ admissible sequence in $W_T^\prime$. \end{enumerate} \end{rmk} \begin{rmk}It is easy to check that the norming set $W_T$ of Tsirelson space is closed under $(\frac{1}{2^n},\mathcal{S}_n)$ operations, namely for any $f_1<\cdots<f_d$ in $W_T$\; $\mathcal{S}_n$-admissible, the functional $\frac{1}{2^n}\sum_{j=1}^df_j\in W_T$. This explains that the norming set $W$ of the space $\mathfrak{X}_{_{^\text{ISP}}}$ is a subset of $W_T$. Therefore Tsirelson space is the unconditional frame on which the norm of $\mathfrak{X}_{_{^\text{ISP}}}$ is built. As we mentioned in the introduction, $\mathfrak{X}_{_{^\text{ISP}}}$ is the first HI construction which uses Tsirelson space instead of a mixed Tsirelson one.\label{remxx1.5} \end{rmk} As it is shown in \cite{CJT} (see also \cite{CS}), an equivalent norm on Tsirelson space is described by the following implicit formula. For $x\in c_{00}$ set \begin{equation*} |\!|\!|x|\!|\!| = \max\big\{\|x\|_0,\;\sup\{\frac{1}{2}\sum_{j=1}^{2d}|\!|\!|E_jx|\!|\!|\}\big\} \end{equation*} where the inner supremum is taken over all successive subsets of the naturals $d \leqslant E_1 <\cdots <E_{2d}$. Then, for any $\{c_k\}_{k=1}^n\subset\mathbb{R}$, the following holds. \begin{equation} \|\sum_{k=1}^nc_ke_k\|_T \leqslant |\!|\!|\sum_{k=1}^nc_ke_k|\!|\!| \leqslant 3\|\sum_{k=1}^nc_ke_k\|_T \label{tsirelsonequivalence} \end{equation} \begin{rmk} A norming set $W_{(T,|\!|\!|\cdot|\!|\!|)}$ for $(T,|\!|\!|\cdot|\!|\!|)$ is also defined in a similar manner as $W_T$.\label{remarknormingthreebars} \end{rmk} \subsection*{Special convex combinations} Next, we remind the notion of the $(n,\varepsilon)$ special convex combinations, (see \cite{AD2},\cite{AGR},\cite{AT}) which is one of the main tools, used in the sequel. \begin{dfn} Let $x = \sum_{k\in F}c_ke_k$ be a vector in $c_{00}$. Then $x$ is said to be a $(n,\varepsilon)$ basic special convex combination (or a $(n,\varepsilon)$ basic s.c.c.) if: \begin{enumerate} \item[(i)] $F\in\mathcal{S}_n, c_k\geqslant 0$, for $k\in F$ and $\sum_{k\in F}c_k = 1$. \item[(ii)] For any $G\subset F, G\in\mathcal{S}_{n-1}$, we have that $\sum_{k\in G}c_k < \varepsilon$. \end{enumerate} \end{dfn} The next result is from \cite{AMT}. For a proof see \cite{AT}, Chapter 2, Proposition 2.3. \begin{prp} For any $M$ infinite subset of the naturals, any $n\in\mathbb{N}$ and $\varepsilon>0$, there exists $F\subset M, \{c_k\}_{k\in F}$, such that $x = \sum_{k\in F}c_ke_k$ is a $(n,\varepsilon)$ basic s.c.c. \label{prop1.5} \end{prp} \begin{dfn} Let $x_1 <\cdots<x_m$ be vectors in $c_{00}$ and $\psi(k) = \min\supp x_k$, for $k=1,\ldots,m$. Then $x = \sum_{k=1}^mc_kx_k$ is said to be a $(n,\varepsilon)$ special convex combination (or $(n,\varepsilon)$ s.c.c.), if $\sum_{k=1}^mc_ke_{\psi(k)}$ is a $(n,\varepsilon)$ basic s.c.c.\label{defscc} \end{dfn} \section{Basic evaluations for special convex combinations} In this section we prove the basic inequality for block sequences in $\mathfrak{X}_{_{^\text{ISP}}}$, with the auxiliary space actually being Tsirelson space. This will allow us to evaluate the norm of $(n,\varepsilon)$ special convex combinations and it is critical throughout the rest of the paper. \begin{dfn} Let $f\in W$ be a functional of type I$_\alpha$ or I$_\beta$, of weight $w(f) = n$, $f = \frac{1}{2^n}\sum_{j=1}^df_j$. Then, by definition, there exist $F_1<\cdots<F_p$ successive intervals of the naturals such that: \begin{enumerate} \item[(i)] $\cup_{i=1}^pF_i = \{1,\ldots,d\}$ \item[(ii)] $\{\min\supp f_j: j\in F_i\}\in\mathcal{S}_{n-1}$, for $i=1,\ldots,p$ \item[(iii)] $\{\min\supp f_{\min F_i}: i=1,\ldots,p\}\in\mathcal{S}_1$ \end{enumerate} Set $g_i = \frac{1}{2^{n-1}}\sum_{j\in F_i}f_j$, for $i=1,\ldots,p$. We call $\{g_i\}_{i=1}^p$ a Tsirelson analysis of $f$. \end{dfn} \begin{rmk} If $f\in W$ is a functional of type I$_\alpha$ or I$_\beta$ and $\{f_i\}_{i=1}^p$ is a Tsirelson analysis of $f$, then $f_i\in W$, $\{f_i\}_{i=1}^p$ is $\mathcal{S}_1$-admissible and $f = \frac{1}{2}\sum_{i=1}^pf_i$, although $\{f_i\}_{i=1}^p$ may not be a very fast growing sequence of $\alpha$-averages or $\beta$-averages. Moreover, if $w(f)>1$, then $f_i$ is of the same type as $f$ and $w(f_i) = w(f) - 1$ for $i=1,\ldots,p$. \label{remark2.3} \end{rmk} \subsection*{The tree analysis of a functional $\mathbf{f\in W}$} A key ingredient for evaluating the norm of vectors in $\mathfrak{X}_{_{^\text{ISP}}}$ is the analysis of the elements $f$ of the norming set $W$. This is similar to the corresponding concept that has occurred in almost all previous HI and related constructions (i.e. \cite{AD1}, \cite{AD2}, \cite{AH}, \cite{AT}). Next we briefly describe the tree analysis in our context. For any functional $f\in W$ we associate a family $\{f_\lambda\}_{\lambda\in\Lambda}$, where $\Lambda$ is a finite tree which is inductively defined as follows. Set $f_\varnothing =f$, where $\varnothing$ denotes the root of the tree to be constructed. If $f$ is of type 0, then the tree analysis of $f$ is $\{f_\varnothing\}$. Otherwise, suppose that the nodes of the tree and the corresponding functionals have been chosen up to a height $p$ and let $\lambda$ be a node of height $|\lambda| = p$. If $f_\lambda$ is of type 0, then don't extend any further and $\lambda$ is a maximal node of the tree. If $f_\lambda$ is of type I$_\alpha$ or I$_\beta$, set the immediate successors of $\lambda$ to be the elements of the Tsirelson analysis of $f_\lambda$. If $f_\lambda$ is of type II, $f = \frac{1}{2}\sum_{j=1}^df_j$, set the immediate successors of $\lambda$ to be the $\{f_j\}_{j=1}^d$. If $f_\lambda$ is a convex combination, which includes $\alpha$-averages and $\beta$-averages, $f_\lambda = \sum_{j=1}^dc_jf_j$, set the immediate successors of $\lambda$ to be the $\{f_j\}_{j=1}^d$. By Remark \ref{remark1.2} it follows that the inductive construction ends in finitely many steps and that the tree $\Lambda$ is finite. \begin{rmk} Let $f\in W$ and $\{f_\lambda\}_{\lambda\in\Lambda}$ be a tree analysis of $f$. Then for any $\lambda\in\Lambda$ not a maximal node, such that $f_\lambda$ is not a convex combination, we have that $f_\lambda = \frac{1}{2}\sum_{\mu\in\scc(\lambda)}f_\mu$, where $\{f_\mu\}_{\mu\in\scc(\lambda)}$ are $\mathcal{S}_1$-admissible and by $\scc(\lambda)$ we denote the immediate successors of $\lambda$ in $\Lambda$. \label{remark2.4} \end{rmk} \begin{rmk} In a similar manner, for any $f\in W_T^\prime$ (see Remark \ref{remark1.3}\;(ii)), the tree analysis of $f$ is defined. \end{rmk} \begin{prp} Let $x = \sum_{k\in F}c_ke_k$ be a $(n,\varepsilon)$ basic s.c.c. and $G\subset F$. Then the following holds. \begin{equation*} \|\sum_{k\in G}c_ke_k\|_T \leqslant\frac{1}{2^n}\sum_{k\in G}c_k + \varepsilon \end{equation*}\label{prop2.1} \end{prp} \begin{proof} Let $f\in W_T^\prime$. We may assume that $\supp f\subset G$. Set $G_1 = \{k\in \supp f: |f(e_k)|\leqslant\frac{1}{2^n}\}, G_2 = \supp f\setminus G_1$. Then clearly $|G_1f(\sum_{k\in G}c_ke_k)|\leqslant \frac{1}{2^n}\sum_{k\in G}c_k$. We will show by induction that $G_2\in\mathcal{S}_{n-1}$. Let $\{f_\lambda\}_{\lambda\in\Lambda}$ be a tree analysis of $G_2f$. Then it is easy to see that $h(\Lambda)\leqslant n-1$. For $\lambda$ a maximal node in $\Lambda$, we have that $\supp f_\lambda\in\mathcal{S}_0$. Assume that for any $\lambda\in\Lambda, |\lambda| = k>0$ we have that $\supp f_\lambda\in\mathcal{S}_{n-1-k}$ and let $\lambda\in\Lambda$, such that $|\lambda| = k-1$. Then $f_\lambda = \frac{1}{2}\sum_{j=1}^df_{\lambda_j}$, where $|\lambda_j| = k$, $\supp f_{\lambda_j}\in\mathcal{S}_{n-k-1}$ for $j=1,\ldots,d$ and $\{\min\supp f_{\lambda_j}:j=1,\ldots,d\}\in\mathcal{S}_1$. Then $\supp f_\lambda = \cup_{j=1}^d\supp f_{\lambda_j}\in \mathcal{S}_{n-1-(k-1)}$. The induction is complete and it follows that $G_2 = \supp G_2f\in \mathcal{S}_{n-1}$ and therefore $G_2f(\sum_{k\in G}c_ke_k)\leqslant \sum_{k\in G_2}c_k <\varepsilon$. Hence, $|f(\sum_{k\in G}c_ke_k)| < \frac{1}{2^n}\sum_{k\in G}c_k + \varepsilon$. \end{proof} \begin{prp}[Basic Inequality] Let $\{x_k\}_k$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ such that $\|x_k\|\leqslant 1$, for all $k$ and let $f\in W$. Set $\phi(k) = \max\supp x_k$, for all $k$. Then there exists $g\in W_{(T,|\!|\!|\cdot|\!|\!|)}$ (see Remark \ref{remarknormingthreebars}) such that $2g(e_{\phi(k)}) \geqslant f(x_k)$, for all $k$. \label{basic} \end{prp} \begin{proof} Let $\{f_\lambda\}_{\lambda\in\Lambda}$ be a tree analysis of $f$. We will inductively construct $\{g_\lambda\}_{\lambda\in\Lambda}$ such that for any $\lambda\in\Lambda$ the following are satisfied. \begin{enumerate} \item[(i)] $g_\lambda\in W_{(T,|\!|\!|\cdot|\!|\!|)}$ and $2g_\lambda(e_{\phi(k)}) \geqslant f_\lambda(x_k)$, for any $k$. \item[(ii)] $\supp g_\lambda \subset \{\phi(k): \ran f_\lambda\cap\ran x_k\neq\varnothing\}$ \end{enumerate} For $\lambda\in\Lambda$ a maximal node, if there exists $k$ such that $\ran f_\lambda\cap\ran x_k\neq\varnothing$, set $g_\lambda = e^*_{\phi(k)}$. Otherwise set $g_\lambda = 0$. Let $\lambda\in\Lambda$ be a non-maximal node, and suppose that $\{g_\mu\}_{\mu>\lambda}$ have been chosen. We distinguish two cases.\vskip3pt \noindent {\em Case 1:} $f_\lambda$ is a convex combination (i.e. $f_\lambda$ is not of type 0, I$_\alpha$, I$_\beta$, or II). If $f_\lambda = \sum_{\mu\in\scc(\lambda)}c_\mu f_\mu$, set $g_\lambda = \sum_{\mu\in\scc(\lambda)}c_\mu g_\mu$.\vskip3pt \noindent {\em Case 2:} $f_\lambda$ is not a convex combination. If $f_\lambda = \frac{1}{2}\sum_{j=1}^df_{\mu_j}$, where $\scc(\lambda) = \{\mu_j\}_{j=1}^d$ such that $f_{\mu_1}<\cdots<f_{\mu_d}$, set \begin{eqnarray*} G_\lambda &=&\{k: \ran f_\lambda\cap\ran x_k\neq\varnothing\}\\ G_1 &=& \{k\in G_\lambda:\;\text{there exists at most one}\;j\;\text{with}\;\ran f_{\mu_j}\cap\ran x_k\neq\varnothing\}\\ G_2 &=& \{k\in G_\lambda:\;\text{there exist at least two}\;j\;\text{with}\;\ran f_{\mu_j}\cap\ran x_k\neq\varnothing\}\\ I_j &=& \{k\in G_1: \ran x_k\cap\ran f_{\mu_j}\neq\varnothing\}\quad\text{for}\;j = 1,\ldots,d. \end{eqnarray*} Observe that $\#G_2\leqslant d-1$. For $j = 1,\ldots,d$ set $g_j^\prime = g_{\mu_j}|_{\phi(I_j)}$ and for $k\in G_2$ set $g_k = e^*_{\phi(k)}$. It is easy to check that if we set $g_\lambda = \frac{1}{2}\big(\sum_{j=1}^dg_j^\prime + \sum_{k\in G_2}g_k\big)$, then $g_\lambda$ is the desired functional. The induction is complete. Set $g = g_\varnothing$ \end{proof} \begin{rmk} In the previous constructions (see \cite{AD1}, \cite{AD2}, \cite{AH}, \cite{AT}), the basic inequality is used for estimating the norm of linear combinations of block vectors which are RIS. In the present paper the basic inequality is stronger, as it is able to provide upper estimations for any block vectors. Moreover, RIS sequences are defined in a different manner as in previous constructions and they also play a different role, which will be discussed in the sequel. \end{rmk} \begin{cor} Let $\{x_k\}_{k}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ such that $\|x_k\|\leqslant 1, \{c_k\}_k\subset\mathbb{R}$ and $\phi(k) = \max\supp x_k$ for all $k$. Then: \begin{equation*} \|\sum_kc_kx_k\| \leqslant 6\|\sum_kc_ke_{\phi(k)}\|_T \end{equation*}\label{cor2.20} \end{cor} \begin{proof} Let $f\in W$. Apply the basic inequality and take $g\in W_{(T,|\!|\!|\cdot|\!|\!|)}$, such that if $\phi(k) = \max\supp x_k$ and $y_k = \sgn(c_k)x_k$ for all $k$, we have that $2g(e_{\phi(k)}) \geqslant f(y_k),\;\text{for any}\;k$. It follows that \begin{equation*} 2g(\sum_k|c_k|e_{\phi(k)})\geqslant f(\sum_kc_kx_k). \end{equation*} Therefore, applying \eqref{tsirelsonequivalence}, we get \begin{equation*} \|\sum_kc_kx_k\| \leqslant 2|\!|\!|\sum_k|c_k|e_{\phi(k)}|\!|\!| = 2|\!|\!|\sum_kc_ke_{\phi(k)}|\!|\!|\leqslant 2\cdot 3\|\sum_kc_ke_{\phi(k)}\|_T \end{equation*} \end{proof} \begin{cor} Let $x = \sum_{k=1}^mc_kx_k$ be a $(n,\varepsilon)$ s.c.c. in $\mathfrak{X}_{_{^\text{ISP}}}$, such that $\|x_k\|\leqslant 1$, for $k=1,\ldots,m$. If $F\subset\{1,\ldots,m\}$, then \begin{equation*} \|\sum_{k\in F}c_kx_k\| \leqslant \frac{6}{2^n}\sum_{k\in F}c_k + 12\varepsilon. \end{equation*} In particular, we have that $\|x\|\leqslant \frac{6}{2^n} + 12\varepsilon$.\label{cor2.21} \end{cor} \begin{proof} Set $\phi(k) = \max\supp x_k, \psi(k) = \min\supp x_k$. Corollary \ref{cor2.20} yields that $\|\sum_{k\in F}c_kx_k\| \leqslant 6\|\sum_{k\in F}c_ke_{\phi(k)}\|_T$. Since, according to the assumption, $\sum_{k\in F}c_ke_{\psi(k)}$ is a $(n,\varepsilon)$ basic s.c.c., it easily follows that $\sum_{k\in F}c_ke_{\phi(k)}$ is a $(n,2\varepsilon)$ basic s.c.c. By Proposition \ref{prop2.1} the result follows. \end{proof} \begin{cor} The basis of $\mathfrak{X}_{_{^\text{ISP}}}$ is shrinking.\label{cor2.22} \end{cor} \begin{proof} Suppose that it is not. Then there exist $x^*\in\mathfrak{X}_{_{^\text{ISP}}}^*, \|x^*\| = 1$, a normalized block sequence $\{x_k\}_{k\in\mathbb{N}}$ in $\mathfrak{X}_{_{^\text{ISP}}}$ and $\delta>0$, such that $x^*(x_k)>\delta$, for all $k\in\mathbb{N}$. Choose $n\in\mathbb{N}$, such that $\frac{1}{2^n}<\frac{\delta}{12}$ and $\varepsilon>0$, such that $\varepsilon<\frac{\delta}{24}$. By Proposition \ref{prop1.5} there exists $F$ a subset of $\mathbb{N}$, such that $x = \sum_{k\in F}c_kx_k$ is a $(n,\varepsilon)$ s.c.c. By Corollary \ref{cor2.21} we have that $\delta > \|x\| \geqslant x^*(x) >\delta$. A contradiction, which completes the proof. \end{proof} \begin{prp} The basis of $\mathfrak{X}_{_{^\text{ISP}}}$ is boundedly complete.\label{prop2.23} \end{prp} \begin{proof} Assume that it is not. Then there exist $\varepsilon>0$ and $\{x_k\}_{k\in\mathbb{N}}$ a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that $\|x_k\| > \varepsilon$ and $\|\sum_{k=\ell}^{\ell+m}x_k\|\leqslant 1$, for all $\ell,m\in\mathbb{N}$. Choose $k_0$ such that $d = \min\supp x_{k_0} > \frac{2}{\varepsilon}$. Set $F_1 = \{k_0\}$ and inductively choose $F_1,\ldots,F_d$, intervals of the naturals such that \begin{enumerate} \item[(i)] $\max F_j + 1 = \min F_{j+1}$, for $j<d$ and \item[(ii)] $\#F_j > \max\{\#F_{j-1}, 2^{\max\supp x_{\max F_{j-1}}}\}$, for $1<j\leqslant d$. \end{enumerate} Then, if we set $y_j = \sum_{k\in F_j}x_k$, we have that $\|\sum_{j=1}^dy_j\|\leqslant 1$. On the other hand, notice that for $j=1,\ldots,d$, there exists $\alpha_j$ an $\alpha$-average in $W$, such that \begin{enumerate} \item[(i)] $\ran \alpha_j\subset \ran y_j$, therefore $\{\alpha_j\}_{j=1}^d$ is $\mathcal{S}_1$-admissible. \item[(ii)] $s(\alpha_j) = \#F_j$, therefore $\{\alpha_j\}_{j=1}^d$ is very fast growing. \item[(iii)] $\alpha_j(y_j) > \varepsilon$ \end{enumerate} From the above it follows $f = \frac{1}{2}\sum_{j=1}^d\alpha_j$ is a functional of type I$_\alpha$ in $W$ and $f(\sum_{j=1}^dy_j) > \frac{\varepsilon\cdot d}{2} > 1$. Since this cannot be the case, the proof is complete. \end{proof} These last two results and a well known result due to R. C. James \cite{J}, allow us to conclude the following. \begin{cor} The space $\mathfrak{X}_{_{^\text{ISP}}}$ is reflexive.\label{cor2.24} \end{cor} \begin{dfn} Let $F$ either be $\mathbb{N}$ or an initial segment of the natural numbers. A block sequence $\{x_k\}_{k\in F}$ is said to be a $(C,\{n_k\}_{k\in F})$\;$\alpha$-rapidly increasing sequence (or $(C,\{n_k\}_{k\in F})$\;$\alpha$-RIS), for a positive constant $C\geqslant 1$ and a strictly increasing sequence of naturals $\{n_k\}_{k\in F}$, if $\|x_k\| \leqslant C$ for all $k\in F$ and the following conditions are satisfied. \begin{enumerate} \item[(i)] For any $k\in F$, for any functional $f$ of type I$_\alpha$ of weight $w(f) = j < n_k$ we have that $|f(x_k)| < \frac{C}{2^j}$ \item[(ii)] For any $k\in F$ we have that $\frac{1}{2^{n_{k+1}}}\max\supp x_{k} < \frac{1}{2^{n_k}}$ \end{enumerate}\label{def3.10} \end{dfn} \begin{rmk} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that there exist a positive constant $C$ and $\{n_k\}_{k\in\mathbb{N}}$ strictly increasing naturals, such that $\|x_k\|\leqslant C$ for all $k$ and condition (i) from Definition \ref{def3.10} is satisfied. Then passing, if necessary, to a subsequence, $\{x_k\}_{k\in\mathbb{N}}$ is $(C,\{n_k\}_{k\in\mathbb{N}})$\;$\alpha$-RIS.\label{rem3.11} \end{rmk} \begin{dfn} Let $n\in\mathbb{N}, C\geqslant 1, \theta>0$. A vector $x\in\mathfrak{X}_{_{^\text{ISP}}}$ is called a $(C,\theta,n)$ vector if the following hold. There exist $0<\varepsilon<\frac{1}{36C2^{3n}}$ and $\{x_k\}_{k=1}^m$ a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ with $\|x_k\|\leqslant C$ for $k=1,\ldots,m$ such that \begin{itemize} \item[(i)] $\min\supp x_1 \geqslant 8C2^{2n}$ \item[(ii)] There exist $\{c_k\}_{k=1}^m\subset [0,1]$ such that $\sum_{k=1}^mc_kx_k$ is a $(n,\varepsilon)$ s.c.c. \item[(iii)] $x = 2^n\sum_{k=1}^mc_kx_k$ and $\|x\| \geqslant \theta$. \end{itemize} If moreover there exist $\{n_k\}_{k=1}^m$ strictly increasing natural numbers with $n_1 > 2^{2n}$ such that $\{x_k\}_{k=1}^m$ is $(C,\{n_k\}_{k=1}^m)$ $\alpha$-RIS, then $x$ is called a $(C,\theta,n)$ exact vector.\label{defvector} \end{dfn} \begin{rmk} Let $x$ be a $(C,\theta,n)$ vector in $\mathfrak{X}_{_{^\text{ISP}}}$. Then, using Corollary \ref{cor2.21} we conclude that $\|x\| < 7C$. \label{remexactest} \end{rmk} \section{The $\alpha,\beta$ indices} To each block sequence we will associate two indices related to $\alpha$ and $\beta$ averages. In this section we will show that every normalized block sequence $\{x_n\}_n$ has a further normalized block sequence $\{y_n\}_n$ such that on it both indices $\alpha$ and $\beta$ are equal to zero. As we will show in the next section, this is sufficient, for a sequence to have a subsequence generating a $c_0$ spreading model. \begin{dfn} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ that satisfies the following. For any $n\in\mathbb{N}$, for any very fast growing sequence $\{\alpha_q\}_{q\in\mathbb{N}}$ of $\alpha$-averages in $W$ and for any $\{F_k\}_{k\in\mathbb{N}}$ increasing sequence of subsets of the naturals, such that $\{\alpha_q\}_{q\in F_k}$ is $\mathcal{S}_n$-admissible, the following holds. For any $\{x_{n_k}\}_{k\in\mathbb{N}}$ subsequence of $\{x_k\}_{k\in\mathbb{N}}$, we have that $\lim_k\sum_{q\in F_k}|\alpha_q(x_{n_k})| = 0$. Then we say that the $\alpha$-index of $\{x_k\}_{k\in\mathbb{N}}$ is zero and write $\al\big(\{x_k\}_k\big) = 0$. Otherwise we write $\al\big(\{x_k\}_k\big) > 0$. \end{dfn} \begin{dfn} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ that satisfies the following. For any $n\in\mathbb{N}$, for any very fast growing sequence $\{\beta_q\}_{q\in\mathbb{N}}$ of $\beta$-averages in $W$ and for any $\{F_k\}_{k\in\mathbb{N}}$ increasing sequence of subsets of the naturals, such that $\{\beta_q\}_{q\in F_k}$ is $\mathcal{S}_n$-admissible, the following holds. For any $\{x_{n_k}\}_{k\in\mathbb{N}}$ subsequence of $\{x_k\}_{k\in\mathbb{N}}$, we have that $\lim_k\sum_{q\in F_k}|\beta_q(x_{n_k})| = 0$. Then we say that the $\beta$-index of $\{x_k\}_{k\in\mathbb{N}}$ is zero and write $\be\big(\{x_k\}_k\big) = 0$. Otherwise we write $\be\big(\{x_k\}_k\big) > 0$. \end{dfn} \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. Then the following assertions are equivalent. \begin{enumerate} \item[(i)] $\al\big(\{x_k\}_k\big) = 0$ \item[(ii)] For any $\varepsilon>0$ there exists $j_0\in\mathbb{N}$ such that for any $j\geqslant j_0$ there exists $k_j\in\mathbb{N}$ such that for any $k\geqslant k_j$, and for any $\{\alpha_q\}_{q=1}^d$ $\mathcal{S}_j$-admissible and very fast growing sequence of $\alpha$-averages such that $s(\alpha_q) > j_0$, for $q=1,\ldots,d$, we have that $\sum_{q=1}^d|\alpha_q(x_k)| < \varepsilon$. \end{enumerate}\label{prop3.3} \end{prp} \begin{proof} It is easy to prove that (i) follows from (ii), therefore we shall only prove the inverse. Suppose that (i) is true and (ii) is not. Then there exists $\varepsilon>0$ such that for any $j_0\in\mathbb{N}$ there exists $j\geqslant j_0$, such that for any $k_0\in\mathbb{N}$, there exists $k\geqslant k_0$ and $\{\alpha_q\}_{q=1}^d$ a $\mathcal{S}_j$-admissible and very fast growing sequence of $\alpha$-averages with $s(\alpha_q) > j_0$, for $q=1,\ldots,d$, such that $\sum_{q=1}^d|\alpha_q(x_k)| \geqslant \varepsilon$. We will inductively choose a subsequence $\{x_{n_i}\}_{i\in\mathbb{N}}$ and $\{\alpha^i\}_{i\in\mathbb{N}}$ a very fast growing sequence of $\alpha$-averages, such that $|\alpha^i(x_{n_i})| > \frac{\varepsilon}{2}$, for any $i$. This evidently yields a contradiction. For $j_0 = 1$, there exists $j_1\geqslant 1$, such that there exists a subsequence $\{x_{k_j}\}_{j\in\mathbb{N}}$ of $\{x_k\}_{k\in\mathbb{N}}$, a sequence $\{\alpha_q\}_{q\in\mathbb{N}}$ of $\alpha$-averages with $s(\alpha_q) > 1$ for all $q\in\mathbb{N}$ and $\{F_j\}_{j\in\mathbb{N}}$ a sequence of increasing intervals of the naturals, such that: \begin{enumerate} \item[(i)] $\{\alpha_q\}_{q\in F_j}$ is very fast growing and $\mathcal{S}_{j_1}$-admissible. \item[(ii)] $\sum_{q\in F_j}|\alpha_q(x_{k_j})|\geqslant \varepsilon$. \item[(iii)] If $F_j^\prime = F_j\setminus\{\min F_j\}$, then $\{\alpha_q\}_{q\in\cup_j F_j^\prime}$ is very fast growing. \end{enumerate} Since $\al\big(\{x_k\}_k\big) = 0$, we have that $\lim_j\sum_{q\in F_j^\prime}|\alpha_q(x_{k_j})| = 0$. Choose $j$ such that $|\alpha_{\min F_j}(x_{k_j})| >\frac{\varepsilon}{2}$ and set $n_1 = k_j, \alpha^1 = \alpha_{\min F_j}$. Suppose that we have chosen $n_1 <\cdots <n_p$ and $\{a^i\}_{i=1}^p$ a very fast growing sequence of $\alpha$-averages, such that $|\alpha^i(x_{n_i})| > \frac{\varepsilon}{2}$, for $i=1,\ldots,p$. Set $j_0 = \max\{s(\alpha^p),\;2^{\max\supp\alpha^p}\}$ and repeat the first inductive step to find an $\alpha$-average $\alpha$ with $s(\alpha)>j_0$ and $x_k>x_{n_p}$, $x_k > \alpha^p$, such that $|\alpha(x_k)|\geqslant\frac{\varepsilon}{2}$. Set $x_{n_{p+1}} = x_k$ and $\alpha^{p+1} = \alpha|_{\ran x_k}$. The inductive construction is complete and so is the proof. \end{proof} The proof of the next proposition is identical to the proof of the previous one. \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. Then the following assertions are equivalent. \begin{enumerate} \item[(i)] $\be\big(\{x_k\}_k\big) = 0$ \item[(ii)] For any $\varepsilon>0$ there exists $j_0\in\mathbb{N}$ such that for any $j\geqslant j_0$ there exists $k_j\in\mathbb{N}$ such that for any $k\geqslant k_j$, and for any $\{\beta_q\}_{q=1}^d$ $\mathcal{S}_j$-admissible and very fast growing sequence of $\beta$-averages such that $s(\beta_q) > j_0$, for $q=1,\ldots,d$, we have that $\sum_{q=1}^d|\beta_q(x_k)| < \varepsilon$. \end{enumerate}\label{prop3.4} \end{prp} \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a seminormalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that either $\al\big(\{x_k\}_k\big) > 0$, or $\be\big(\{x_k\}_k\big) > 0$. Then there exists $\theta>0$ and a subsequence $\{x_{n_k}\}_{k\in\mathbb{N}}$ of $\{x_k\}_{k\in\mathbb{N}}$, that generates an $\ell_1^n$ spreading model, with a lower constant $\frac{\theta}{2^n}$, for all $n\in\mathbb{N}$. More precisely, for every $n\in\mathbb{N}$ and $F\subset\mathbb{N}$ with $\{\min\supp x_{n_k}: k\in F\}\in\mathcal{S}_n$ and $\{c_k\}_k\subset\mathbb{R}$, we have that $\|\sum_{k\in F}c_kx_{n_k}\| \geqslant \frac{\theta}{2^n}\sum_{k\in F}|c_k|$. In particular, for any $k_0,n\in\mathbb{N}$, there exists $F$ a finite subset of $\mathbb{N}$ with $\min F\geqslant k_0$ and $\{c_k\}_{k\in F}$, such that $x = 2^n\sum_{k\in F}c_kx_{n_k}$ is a $(C,\theta,n)$ vector, where $C = \sup\{\|x_k\|: k\in\mathbb{N}\}$. If moreover $\{x_k\}_k$ is $(C^\prime,\{n_k\}_k)$ $\alpha$-RIS, then $x$ can be chosen to be a $(C^{\prime\prime},\theta,n)$ exact vector, where $C^{\prime\prime} = \max\{C,C^\prime\}$. \label{prop3.5} \end{prp} \begin{proof} Assume that $\al\big(\{x_k\}_k\big) > 0$. Then there exist $\ell\in\mathbb{N}, \varepsilon>0, \{\alpha_q\}_{q\in\mathbb{N}}$ a very fast growing sequence of $\alpha$-averages, $\{F_k\}_{k\in\mathbb{N}}$ increasing subsets of the naturals such that $\{\alpha_q\}_{q\in F_k}$ is $\mathcal{S}_\ell$-admissible for all $k\in\mathbb{N}$ and $\{x_{n_k}\}_{k\in\mathbb{N}}$ a subsequence of $\{x_k\}_{k\in\mathbb{N}}$, such that $\sum_{q\in F_k}|\alpha_q(x_{n_k})| > \varepsilon$, for all $k\in\mathbb{N}$. Pass, if necessary ,to a subsequence, again denoted by $\{x_{n_k}\}_{k\in\mathbb{N}}$, generating some spreading model. By changing the signs and restricting the ranges of the $\alpha_q$, we may assume that $\sum_{q\in F_k}\alpha_q(x_{n_k}) > \varepsilon$, for all $k\in\mathbb{N}$ and $\ran \alpha_q\subset \ran x_{n_k}$ for any $q\in F_k$ and $k\in\mathbb{N}$. Set $\theta = \frac{\varepsilon}{2^\ell}$. Let $k_0,n\in\mathbb{N}$ and choose $0<\eta<\frac{1}{36C2^{3n}}$. By Proposition \ref{prop1.5} there exists $F$ a finite subset of $\{n_k: k\geqslant \max\{k_0, 8C2^{2n}\}\}$ and $\{c_k\}_{k\in F}$, such that $x^\prime = \sum_{k\in F}c_kx_{n_k}$ is a $(n,\eta)$ s.c.c. Set $f = \frac{1}{2^{\ell + n}}\sum_{k\in F}\sum_{q\in F_{n_k}}\alpha_q$. Then $f$ is a functional of type I$_\alpha$ in $W$ and $f(x^\prime) > \frac{\varepsilon}{2^{\ell+n}} = \frac{\theta}{2^n}$. Therefore $x = 2^nx^\prime$ is the desired $(C,\theta,n)$ vector. If moreover $\{x_k\}_k$ is $(C^\prime,\{n_k\}_k)$ $\alpha$-RIS, obvious modifications yield that $x$ can be chosen to be a $(C^{\prime\prime},\theta,n)$ exact vector. Arguing in the same way, for any $n\in\mathbb{N}$, for any $F\subset\mathbb{N}$ with $\{\min\supp x_{n_k}: k\in F\}\in\mathcal{S}_n$ and $\{c_k\}_k\subset\mathbb{R}$, we have that $\|\sum_{k\in F}c_kx_{n_k}\| > \frac{\theta}{2^n}\sum_{k\in F}|c_k|$. The proof is exactly the same if $\be\big(\{x_k\}_k\big) > 0$. \end{proof} \subsection*{Block sequences with $\alpha$-index zero} In this subsection we show that sequences $\{x_k\}_{k\in\mathbb{N}}$ with $x_k$ a $(C,\theta,n_k)$ vector, with $\{n_k\}_k$ strictly increasing have $\alpha$-index zero. Also also prove that sequences with $\alpha$-index zero have $\alpha$-RIS subsequences. \begin{prp} Let $\{x_k\}_k$ be a bounded block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ with $\al\big(\{x_k\}_k\big) = 0$. Then it has a subsequence that is $(2C,\{n_k\}_k)$ $\alpha$-RIS, where $C = \sup\{\|x_k\|: k\in\mathbb{N}\}$.\label{prpalzeroalris} \end{prp} \begin{proof} Applying Proposition \ref{prop3.3} we have the following. There exists $j_0\in\mathbb{N}$ such that for every $j\geqslant j_0$ there exists $k_j\in\mathbb{N}$ such that for every $k\geqslant k_j$ and $\{\alpha_q\}_{q=1}^d$ very fast growing and $\mathcal{S}_j$ admissible sequence of $\alpha$-averages with $s(\alpha_q)>j_0$ for $q=1,\ldots,d$, we have that $\sum_{q=1}^d|\alpha_q(x_k)| < C$. We shall show that for every $j\geqslant j_0$ and $k_0\in\mathbb{N}$, there exists $k\geqslant k_0$ such that for every $f\in W$ of type I$_\alpha$ and $w(f) = n < j$, we have that $|f(x_k)| < \frac{2C}{2^n}$. If this is shown to be true, then by Remark \ref{rem3.11} we are done. Fix $j\geqslant j_0$ and $k_0\in\mathbb{N}$. Set $k = \max\{j_0, k_0, k_{j_0}\}$ and let $f\in W$ with $w(f) = n < j$. Then $f$ is of the form $f = \frac{1}{2^n}\sum_{q=1}^d\alpha_q$, where $\{\alpha_q\}_{q=1}^d$ is a very fast growing and $\mathcal{S}_n$ admissible sequence of $\alpha$-averages. We may clearly assume that $\ran \alpha_1\cap \ran x_k\neq\varnothing$. Then $\{\alpha_q\}_{q=2}^d$ is very fast growing with $s(\alpha_q) > \min\supp x_k \geqslant j_0$ and it is $S_j$ admissible, as it is $\mathcal{S}_n$ admissible and $n<j$. We conclude the following. \begin{equation*} |f(x_k)| \leqslant \frac{1}{2^n}(|\alpha_1(x_k)| + \sum_{q=2}^d|\alpha_q(x_k)|) <\frac{1}{2^n}(C + C) \end{equation*} \end{proof} \begin{lem} Let $x = 2^n\sum_{k=1}^mc_kx_k$ be a $(C,\theta,n)$ vector in $\mathfrak{X}_{_{^\text{ISP}}}$. Let also $\alpha$ be an $\alpha$-average in $W$ and set $G_\alpha = \{k: \ran\alpha\cap\ran x_k\neq\varnothing\}$. Then the following holds. \begin{equation*} |\alpha(x)| < \min\big\{\frac{C2^n}{s(\alpha)}\sum_{k\in G_\alpha}c_k,\; \frac{6C}{s(\alpha)}\sum_{k\in G_\alpha}c_k + \frac{1}{3\cdot2^{2n}}\big\} + 2C2^n\max\{c_k: k\in G_\alpha\} \end{equation*}\label{lem3.6} \end{lem} \begin{proof} If $\alpha = \frac{1}{p}\sum_{j=1}^df_j$. Set \begin{eqnarray*} E_1 &=& \{k\in G_\alpha:\;\text{there exists at most one}\;j\;\text{with}\;\ran f_j\cap\ran x_k\neq\varnothing\}\\ E_2 &=& \{1,\ldots,m\}\setminus E_1\\ J_k &=& \{j: \ran f_j\cap\ran x_k\neq\varnothing\}\quad\text{for}\;k\in E_2. \end{eqnarray*} Then it is easy to see that \begin{equation} |\alpha(\sum_{k\in E_1}c_kx_k)| \leqslant \frac{C}{p}\sum_{k\in G_\alpha}c_k \label{lem3.6eq1} \end{equation} Moreover \begin{equation} |\alpha(\sum_{k\in E_2}c_kx_k)| < 2C\max\{c_k: k\in G_\alpha\}\label{lem3.6eq2} \end{equation} To see this, notice that \begin{equation*} |\alpha(\sum_{k\in E_2}c_kx_k)| \leqslant \frac{1}{p}\sum_{k\in E_2}c_k\big(\sum_{j\in J_k}|f_j(x_k)|\big) < \max\{c_k: k\in G_\alpha\}\frac{2Cp}{p} \end{equation*} Set $J = \{j:$ there exists $k\in E_1$ such that $\ran f_j\cap\ran x_k\neq\varnothing\}$ and for $j\in J$ set $G_j = \{k\in E_1: \ran f_j\cap\ran x_k\neq\varnothing\}$. Then the $G_j$ are pairwise disjoint and $\cup_{j\in J}G_j = E_1$. For $j\in J$, Corollary \ref{cor2.21} yields that \begin{equation*} |f_j(\sum_{k\in G_j}c_kx_k)| \leqslant \frac{6C}{2^n}\sum_{k\in G_j}c_k + \frac{1}{3\cdot2^{3n}} \end{equation*} Therefore \begin{equation} |\alpha(\sum_{k\in E_1}c_kx_k)| \leqslant \frac{1}{p}\sum_{j\in J}|f_j(\sum_{k\in G_j}c_kx_k)| \leqslant \frac{6C}{2^np}\sum_{k\in G_\alpha}c_k + \frac{1}{3\cdot2^{3n}}\label{lem3.6eq3} \end{equation} Then \eqref{lem3.6eq1} and \eqref{lem3.6eq3} yield the following. \begin{equation} |\alpha(\sum_{k\in E_1}c_kx_k)| \leqslant \min\big\{\frac{C}{s(\alpha)}\sum_{k\in G_\alpha}c_k,\; \frac{6C}{2^ns(\alpha)}\sum_{k\in G_\alpha}c_k + \frac{1}{3\cdot2^{3n}}\big\}\label{lem3.6eq4} \end{equation} By summing up \eqref{lem3.6eq2} and \eqref{lem3.6eq4} the result follows. \end{proof} \begin{lem} Let $x$ be a $(C,\theta,n)$ vector in $\mathfrak{X}_{_{^\text{ISP}}}$. Let also $\{a_q\}_{q=1}^d$ be a very fast growing and $\mathcal{S}_j$-admissible sequence of $\alpha$-averages, with $j<n$. Then the following holds. \begin{equation*} \sum_{q=1}^d|\alpha_q(x)| < \frac{6C}{s(\alpha_1)} + \frac{1}{2^n} \end{equation*}\label{lem3.7} \end{lem} \begin{proof} Assume that $x = 2^n\sum_{k=1}^mc_kx_k$ such that the assumptions of Definition \ref{defvector} are satisfied. Set $q_1 = \min\{q: \ran \alpha_q\cap\ran x\neq\varnothing\}$. For convenience assume that $q_1 = 1$. Then by Lemma \ref{lem3.6} we have that \begin{equation} |\alpha_1(x)| < \frac{6C}{s(\alpha_1)} + \frac{7}{18\cdot 2^{2n}} \label{lem3.7eq1} \end{equation} Set \begin{eqnarray*} J_1 &=& \{q>1:\;\text{there exists at most one}\;k\;\text{such that}\;\ran\alpha_q\cap\ran x_k\neq\varnothing\}\\ J_2 &=& \{q>1: q\notin J_1\}\\ G^q &=& \{k: \ran \alpha_q\cap\ran x_k\neq\varnothing\}\quad\text{for}\;q>1.\\ G_1 &=& \{k:\;\text{there exists}\;q\in J_1\;\text{with}\;\ran\alpha_q\cap\ran x_k\neq\varnothing\} \end{eqnarray*} Then $\{\min\supp x_k: k\in G_1\setminus\{\min G_1\}\}\in\mathcal{S}_j$, hence $\sum_{k\in G_1}c_k< \frac{1}{18C2^{3n}}$. It is easy to check that \begin{equation} \sum_{q\in J_1}|\alpha_q(x)| \leqslant 2^jC2^n\|\sum_{k\in G_1}c_kx_k\| < 2^{n-1}C2^{n}\frac{1}{18C2^{3n}} = \frac{1}{36\cdot 2^n} \label{lem3.7eq2} \end{equation} For $q\in J_2$, Lemma \ref{lem3.6} yields that \begin{eqnarray*} |\alpha_q(x)| &<& \frac{C2^n}{s(\alpha_q)}\sum_{k\in G^q}c_k + 2C2^n\max\{c_k: k\in G^q\}\\ &<& \frac{C2^n}{\min\supp x}\sum_{k\in G^q}c_k + 2C2^nc_{k_q} \end{eqnarray*} where $k_q\in G^q$, such that $c_{k_q} = \max\{c_k: k\in G^q\}$. Then $\{\min\supp x_{k_q}: q\in J_2\setminus\{\min J_2\}\}\in\mathcal{S}_j$. By the above we conclude that \begin{equation} \sum_{q\in J_2}|\alpha_q(x)| < \frac{2C2^n}{\min\supp x} + \frac{8}{36\cdot2^{2n}} < \frac{1}{4\cdot2^n} + \frac{8}{36\cdot2^{2n}} \label{lem3.7eq3} \end{equation} Summing up \eqref{lem3.7eq1}, \eqref{lem3.7eq2} and \eqref{lem3.7eq3}, the desired result follows. \end{proof} \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence of $(C,\theta,n_k)$ vectors in $\mathfrak{X}_{_{^\text{ISP}}}$ with $\{n_k\}_k$ strictly increasing. Then $\al\big(\{x_k\}_k\big) = 0$.\label{prop3.8} \end{prp} \begin{proof} We shall make use of Proposition \ref{prop3.3}. Let $\varepsilon>0$ and choose $j_0\in\mathbb{N}$ such that $\frac{6C}{j_0} < \frac{\varepsilon}{2}$. For $j\geqslant j_0$, choose $k_j$, such that $\frac{1}{2^{n_{k_j}}}<\frac{\varepsilon}{2}$. For $k\geqslant k_j$, Lemma \ref{lem3.7} yields that if $\{\alpha_q\}_{q=1}^d$ is a very fast growing and $\mathcal{S}_j$-admissible sequence of $\alpha$-averages and $s(\alpha_q)>j_0$, for $q=1,\ldots,d$, we have that \begin{equation*} \sum_{q=1}^d|\alpha_q(x_k)| < \frac{6C}{j_0} + \frac{1}{2^{n_k}} < \frac{\varepsilon}{2} + \frac{\varepsilon}{2} = \varepsilon \end{equation*} \end{proof} \begin{prp} Let $x$ be a $(C,\theta,n)$ vector in $\mathfrak{X}_{_{^\text{ISP}}}$. Then for any $f\in W$ functional of type I$_\alpha$, such that $w(f) = j < n$, we have that $|f(x)| < \frac{7C}{2^j}$.\label{cor3.9} \end{prp} \begin{proof} Let $f = \frac{1}{2^j}\sum_{q=1}^d\alpha_j$ be a functional of type I$_\alpha$ with weight $w(f) = j < n$. Then Lemma \ref{lem3.7} yields that \begin{equation*} |f(x)| \leqslant \frac{1}{2^j}\big( \sum_{q=1}^d|\alpha_q(x)|\big) < \frac{1}{2^j}\big(\frac{6C}{s(\alpha_1)} + \frac{1}{2^n}\big)\leqslant \frac{7C}{2^j} \end{equation*} \end{proof} The following proposition follows immediately from Proposition \ref{cor3.9} and Remark \ref{rem3.11}. \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a block sequence of $(C,\theta,n_k)$ vectors in $\mathfrak{X}_{_{^\text{ISP}}}$ with $\{n_k\}_k$ strictly increasing. Then passing, if necessary, to a subsequence, $\{x_k\}_{k\in\mathbb{N}}$ is $(7C,\{n_k\}_k)$\;$\alpha$-RIS. \label{rem3.12} \end{prp} \subsection*{Block sequences with $\beta$-index zero.} In this subsection we first prove that every block sequence of $(C,\theta,n_k)$ exact vectors with $\{n_k\}_k$ strictly increasing, has $\beta$-index zero. This yields that every block sequence has a further block sequence with both $\alpha, \beta$ indices equal to zero. We start with the following technical lemma. Its meaning becomes more transparent in the following Corollary \ref{corx3.14} and Lemmas \ref{lem3.14}, \ref{lem3.15}. \begin{ntt} Let $x = 2^n\sum_{k=1}^mc_kx_k$ be a $(C,\theta,n)$ exact vector, with $\{x_k\}_{k=1}^m$ $(C,\{n_k\}_{k=1}^m)$ $\alpha$-RIS. Let also $f = \frac{1}{2}\sum_{j=1}^df_j$ be a type II functional. Set \begin{eqnarray*} I_0 &=& \{j: n\leqslant w(f_j) < 2^{2n}\}\\ I_1 &=& \{j: w(f_j) < n\}\\ I_2 &=& \{j: 2^{2n} \leqslant w(f_j) < n_1\}\\ J_k &=& \{j: n_k\leqslant w(f_j) < n_{k+1}\},\;\text{for}\;k<m\;\text{and}\quad J_m = \{j: n_m\leqslant w(f_j)\} \end{eqnarray*} \end{ntt} Under the above notation the following lemma holds. \begin{lem} Let $x = 2^n\sum_{k=1}^mc_kx_k$ be a $(C,\theta,n)$ exact vector in $\mathfrak{X}_{_{^\text{ISP}}}, n\geqslant 2$, with $\{x_k\}_{k=1}^m$ $(C,\{n_k\}_{k=1}^m)$ $\alpha$-RIS. Let also $f = \frac{1}{2}\sum_{j=1}^df_j$ be a functional of type II. Then there exists $F_f\subset \{k: \ran f\cap \ran x_k\neq\varnothing\}$ with $\{\min\supp x_k: k\in F_f\}\in\mathcal{S}_2$ such that \begin{eqnarray*} |f(x)| &<& 7C\#I_0 + \frac{C}{2}\bigg(\sum_{k=2}^m\sum_{j\in J_k}\frac{2^{n_k}}{2^{w(f_j) + n_{k-1}}} + \sum_{k=1}^{m-1}\sum_{j\in J_k}\frac{2^n}{2^{w(f_j)}}\\ &&+ \sum_{j\in I_1}\frac{7}{2^{w(f_j)}} + \sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}}\bigg) + C2^n\sum_{k\in F_f}c_k \end{eqnarray*}\label{lem3.13} \end{lem} \begin{proof} Notice that $\{J_k\}_{k=1}^m$ are disjoint intervals of $\{1,\ldots,d\}$ and that $g_k = \frac{1}{2}\sum_{j\in J_k}f_j\in W$, for $k=1,\ldots,m$. Set $F_f = \{k: \ran g_k\cap\ran x_k\neq\varnothing\}$. It easily follows that $\{\min\supp x_k: k\in F_f\}\in\mathcal{S}_2$ and that \begin{equation} \frac{2^n}{2}\sum_{k=1}^m|\sum_{j\in J_k}f_j(c_kx_k)| \leqslant C2^n\sum_{k\in F_f}c_k \label{lem3.13eq1} \end{equation} Let $k_0\leqslant m, j\in J_{k_0}$. Then \begin{equation} 2^n|f_j(\sum_{k<k_0}c_kx_k)| < C\frac{2^{n_{k_0}}}{2^{w(f_j) + n_{k_0 - 1}}}\;\;\text{and}\;\; 2^n|f_j(\sum_{k>k_0}c_kx_k)| < C\frac{2^n}{2^{w(f_j)}} \label{lem3.13eq2} \end{equation} Proposition \ref{cor3.9} yields that for $j\in I_1$ we have that $|f_j(x)| < \frac{7C}{2^{w(f_j)}}$ and hence \begin{equation} \frac{1}{2}\sum_{j\in I_1}|f_j(x)| < \frac{C}{2}\sum_{j\in I_1}\frac{7}{2^{w(f_j)}} \label{lem3.13eq3} \end{equation} For $j\in I_2$ we have that $|f_j(x)| < \frac{C2^n}{2^{w(f_j)}}$ and therefore \begin{equation} \frac{1}{2}\sum_{j\in I_2}|f_j(x)| < \frac{C}{2}\sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}} \label{lem3.13eq4} \end{equation} By Remark \ref{remexactest} yields that $\|x\| < 7C$, and since $I_0$ is an interval, it follows that $\frac{1}{2}\sum_{j\in I_0}f_j\in W$. Therefore \begin{equation} \frac{1}{2}|\sum_{j\in I_0}f_j(x)| < 7C \label{lem3.13eq5} \end{equation} Summing up \eqref{lem3.13eq1} to \eqref{lem3.13eq5} the desired result follows. \end{proof} The next corollary will be useful in the next sections, when we define the notion of dependent sequences. \begin{cor} Let $x$ be a $(C,\theta,n)$ exact vector in $\mathfrak{X}_{_{^\text{ISP}}}, n\geqslant 3$. Let also $f = \frac{1}{2}\sum_{j=1}^df_j$ be a functional of type II. with $\widehat{w}(f)\cap\{n,\ldots,2^{2n}\} = \varnothing$. Then \begin{equation*} |f(x)| < \frac{C}{2^n} + \frac{C}{2^{2n}} + \sum_{\{j:\;w(f_j)<n\}}\frac{4C}{2^{w(f_j)}} \end{equation*}\label{corx3.14} \end{cor} \begin{proof} Let $x = 2^n\sum_{k=1}^mc_kx_k$ with $\{x_k\}_{k=1}^m$ $(C,\{n_k\}_{k=1}^m)$ $\alpha$-RIS. Apply Lemma \ref{lem3.13}. Then the following holds. \begin{eqnarray} |f(x)| &<& \frac{C}{2}\bigg(\sum_{k=2}^m\sum_{j\in J_k}\frac{2^{n_k}}{2^{w(f_j) + n_{k-1}}} + \sum_{k=1}^{m-1}\sum_{j\in J_k}\frac{2^n}{2^{w(f_j)}}\label{corx3.14eq1}\\ &&+ \sum_{j\in I_1}\frac{7}{2^{w(f_j)}} + \sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}}\bigg) + \frac{1}{36\cdot2^{2n}}\nonumber \end{eqnarray} Notice the following. \begin{equation} \sum_{k=2}^m\sum_{j\in J_k}\frac{2^{n_k}}{2^{w(f_j) + n_{k-1}}} \leqslant \frac{1}{2^{n_1}} < \frac{1}{2^{2n}} \label{corx3.14eq2} \end{equation} \begin{equation} \sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}} + \sum_{k=1}^{m-1}\sum_{j\in J_k}\frac{2^n}{2^{w(f_j)}} = 2^n\big(\sum_{\{j: w(f_j)\geqslant 2^{2n}\}}\frac{1}{2^{w(f_j)}}\big) \leqslant \frac{2}{2^n}\label{corx3.14eq3} \end{equation} Applying \eqref{corx3.14eq2} and \eqref{corx3.14eq3} to \eqref{corx3.14eq1} the result follows. \end{proof} \begin{lem} Let $x = 2^n\sum_{k=1}^mc_kx_k$ be a $(C,\theta,n)$ exact vector in $\mathfrak{X}_{_{^\text{ISP}}}, n\geqslant 2$, with $\{x_k\}_{k=1}^m$ $(C,\{n_k\}_{k=1}^m)$ $\alpha$-RIS. Let also $\beta$ be a $\beta$-average. Then there exists $F_\beta\subset \{k: \ran \beta\cap \ran x_k\neq\varnothing\}$ with $\{\min\supp x_k: k\in F_\beta\}\in\mathcal{S}_2$ such that \begin{equation*} |\beta(x)| < \frac{8C}{s(\beta)} + C2^n\sum_{k\in F_\beta}c_k \end{equation*}\label{lem3.14} \end{lem} \begin{proof} If $\beta = \frac{1}{p}\sum_{q=1}^dg_q$, then by definition the $g_q$ are functionals of type II with disjoint weights $\widehat{w}(g_q)$. For convenience, we may write $g_q = \frac{1}{2}\sum_{j\in G_q}f_j$, where the index sets $G_q, q=1,\ldots,d$ are pairwise disjoint. Notice that for $j_1,j_2\in G, j_1\neq j_2$ we have that $w(f_{j_1})\neq w(f_{j_2})$. By slightly modifying the previously used notation, set $G = \cup_{q=1}^dG_q$ and \begin{eqnarray*} I_0 &=& \{j\in G: n\leqslant w(f_j) < 2^{2n}\}\\ I_1 &=& \{j\in G: w(f_j) < n\}\\ I_2 &=& \{j\in G: 2^{2n} \leqslant w(f_j) < n_1\}\\ J_k &=& \{j\in G: n_k\leqslant w(f_j) < n_{k+1}\},\;\text{for}\;k<m\quad\text{and}\\ J_m &=& \{j\in G: n_m\leqslant w(f_j)\} \end{eqnarray*} By Remark \ref{remark1.1} there exists at most one $q_0\leqslant d$, with $\widehat{w}(f_{q_0})\cap\{n,\ldots,2^{2n}\}\neq\varnothing$ and if such a $q_0$ exists, then $\#\widehat{w}(f_{q_0})\cap\{n,\ldots,2^{2n}\}\leqslant 1$. Apply Lemma \ref{lem3.13}. Then for $q=1,\ldots,d$ there exists $F_q \subset \{x_k: \ran \beta\cap \ran x_k\neq\varnothing\}$ with $\{\min\supp x_k: k\in F_q\}\in\mathcal{S}_2$ such that \begin{eqnarray} 2^n|\beta(x)| &<& \frac{7C}{p} + \frac{C}{2p}\bigg(\sum_{k=2}^m\sum_{j\in J_k}\frac{2^{n_k}}{2^{w(f_j) + n_{k-1}}} + \sum_{k=1}^{m-1}\sum_{j\in J_k}\frac{2^n}{2^{w(f_j)}}\label{lem3.14eq1}\\ &&+ \sum_{j\in I_1}\frac{7}{2^{w(f_j)}} + \sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}}\bigg) + \frac{1}{p}C2^n\sum_{q=1}^d\sum_{k\in F_q}c_k\nonumber \end{eqnarray} Just as in the proof of Corollary \ref{corx3.14}, notice the following. \begin{equation} \sum_{k=2}^m\sum_{j\in J_k}\frac{2^{n_k}}{2^{w(f_j) + n_{k-1}}} < \frac{1}{2^{2n}} \label{lem3.14eq2} \end{equation} \begin{equation} \sum_{j\in I_2}\frac{2^n}{2^{w(f_j)}} + \sum_{k=1}^{m-1}\sum_{j\in J_k}\frac{2^n}{2^{w(f_j)}} \leqslant \frac{2}{2^n}\label{lem3.14eq3} \end{equation} By the definition of the coding function $\sigma$ we get \begin{equation} \sum_{j\in I_1}\frac{7}{2^{w(f_j)}} < \frac{7}{1000} \label{lem3.14eq4} \end{equation} \begin{equation} \frac{1}{p}C2^n\sum_{q=1}^d\sum_{k\in F_q}c_k\leqslant C2^n\max\big\{\sum_{k\in F_q}c_k: q=1,\ldots,p\big\} = C2^n\sum_{k\in F_{q_0}}c_k \label{lem3.14eq5} \end{equation} for some $1\leqslant q_0\leqslant d$. Set $F_\beta = F_{q_0}$ and apply \eqref{lem3.14eq2} to \eqref{lem3.14eq5} to \eqref{lem3.14eq1} to derive the desired result. \end{proof} \begin{lem} Let $x$ be a $(C,\theta,n)$ exact vector in $\mathfrak{X}_{_{^\text{ISP}}}, n\geqslant 4$. Let also $\{\beta_q\}_{q=1}^d$ be a very fast growing and $\mathcal{S}_j$-admissible sequence of $\beta$-averages with $j\leqslant n-3$. Then we have that \begin{equation*} \sum_{q=1}^d|\beta_q(x)| < \sum_{q=1}^d\frac{8C}{s(\beta_q)} + \frac{1}{2^n} \end{equation*}\label{lem3.15} \end{lem} \begin{proof} Set \begin{eqnarray*} J_1 &=& \{q:\;\text{there exists at most one}\; k\;\text{such that}\;\ran\beta_q\cap\ran x_k\neq\varnothing\}\\ J_2 &=& \{1,\ldots,d\}\setminus J_2\\ G_1 &=& \{k:\;\text{there exists}\;q\in J_1\;\text{with}\;\ran\beta_q\cap\ran x_k\neq\varnothing\} \end{eqnarray*} Then $\{\min\supp x_k: k\in G_1\setminus\{\min G_1\}\}\in\mathcal{S}_{j+1}$ and it is easy to check that \begin{equation} \sum_{q\in J_1}^d|\beta_q(x)| \leqslant 2^j2^n\|\sum_{k\in G_1}c_kx_k\| < \frac{2}{9\cdot2^n} \label{lem3.15eq1} \end{equation} For $q\in J_2$, choose $F_q\subset\{1,\ldots,m\}$ as in Lemma \ref{lem3.14} and set $F = \cup_{q\in J_2}F_q$. Then $\{\min\supp x_k: k\in F\setminus\{\min F\}\}\in\mathcal{S}_{n-1}$, therefore $\sum_{q\in J_2}\sum_{k\in F_q}c_k < \frac{1}{9C2^{3n}}$. Lemma \ref{lem3.14} yields that \begin{equation} \sum_{q\in J_2}|\beta_q(x)| < \sum_{q\in J_2}\frac{8C}{s(\beta_q)} + \frac{1}{9\cdot2^{2n}} \label{lem3.15eq2} \end{equation} Combining \eqref{lem3.15eq1} and \eqref{lem3.15eq2}, the result follows. \end{proof} \begin{prp} Let $\{x_k\}_k$ be a block sequence of $(C,\theta,n_k)$ exact vectors in $\mathfrak{X}_{_{^\text{ISP}}}$ with $\{n_k\}_k$ strictly increasing. Then $\al\big(\{x_k\}_k\big) = 0$ as well as $\be\big(\{x_k\}_k\big) = 0$.\label{cor3.16} \end{prp} \begin{proof} Proposition \ref{prop3.8} yields that $\al\big(\{x_k\}_k\big) = 0$. To prove that $\be\big(\{x_k\}_k\big) = 0$, we shall make use of Proposition \ref{prop3.4}. Let $\varepsilon>0$ and choose $j_0\in\mathbb{N}$ such that \begin{equation*} \frac{8C}{j_0} < \frac{\varepsilon}{4} \end{equation*} For $j\geqslant j_0$ choose $k_j$, such that $n_{k_j} \geqslant j + 3$ and $\frac{1}{2^{n_{k_j}}}<\frac{\varepsilon}{4}$. For $k\geqslant k_j$, Lemma \ref{lem3.15} yields that if $\{\beta_q\}_{q=1}^d$ is a very fast growing and $\mathcal{S}_j$-admissible sequence of $\beta$-averages and $s(\beta_q)>j_0$, for $q=1,\ldots,d$, we have that \begin{equation*} \sum_{q=1}^d|\beta_q(x_k)| < \sum_{q=1}\frac{8C}{s(\beta_q)} + \frac{1}{2^{n_k}} < \frac{\varepsilon}{4} + \sum_{j>\min\supp x_k}\frac{8C}{2^j} + \frac{\varepsilon}{4} < \varepsilon \end{equation*} \end{proof} \begin{cor} Let $\{x_k\}_{k\in\mathbb{N}}$ be a normalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. Then there exists a further normalized block sequence $\{y_k\}_{k\in\mathbb{N}}$ of $\{x_k\}_{k\in\mathbb{N}}$, such that $\al\big(\{y_k\}_k\big) = 0$ as well as $\be\big(\{y_k\}_k\big) = 0$.\label{cor3.17} \end{cor} \begin{proof} If $\al\big(\{x_k\}_k\big) = 0$ and $\be\big(\{x_k\}_k\big) = 0$, then there is nothing to prove. Otherwise, if $\al\big(\{x_k\}_k\big) > 0$ or $\be\big(\{x_k\}_k\big) > 0$, apply Proposition \ref{prop3.5} to construct a block sequence $\{z_k\}_{k\in\mathbb{N}}$ of $(1,\theta,n_k)$ vectors, with $\{n_k\}_k$ strictly increasing. Then by Proposition \ref{prop3.8} $\al\big(\{z_k\}_k\big) = 0$ and Proposition \ref{rem3.12} yields, that passing, if necessary, to a subsequence, we have that $\{z_k\}_{k\in\mathbb{N}}$ is $(7,\{n_k\}_k)$\;$\alpha$-RIS. If $\be\big(\{z_k\}_k\big) = 0$, set $y_k = \frac{1}{\|z_k\|}z_k$ and $\{y_k\}_{k\in\mathbb{N}}$ is the desired sequence. Otherwise, if $\be\big(\{z_k\}_k\big) > 0$, apply once more Proposition \ref{prop3.5} to construct a block sequence $\{w_k\}_{k\in\mathbb{N}}$, of $(7,\theta^\prime,m_k)$ exact vectors, with $\{m_k\}_k$ strictly increasing. Proposition \ref{cor3.16} yields that $\al\big(\{w_k\}_k\big) = 0$, as well as $\be\big(\{w_k\}_k\big) = 0$. Set $y_k = \frac{1}{\|w_k\|}w_k$ and $\{y_k\}_{k\in\mathbb{N}}$ is the desired sequence. \end{proof} \section{$c_0$ spreading models} This section is devoted to necessary conditions for a sequence $\{x_k\}_k$ to generate a $c_0$ spreading model. At the beginning a Ramsey type result is proved concerning type II functionals acting on a block sequence $\{x_k\}_k$ with $\be\big(\{x_k\}_k\big) = 0$. Then conditions are provided for a finite sequence to be equivalent to the basis of $\ell_\infty^n$. This is critical for establishing the HI property and the properties of the operators in the space. Moreover it is shown that any block sequence $\{x_k\}_k$ with $\al\big(\{x_k\}_k\big) = 0$ and $\be\big(\{x_k\}_k\big) = 0$ contains a subsequence generating a $c_0$ spreading model. Another critical property related to sequences generating $c_0$ spreading models is that increasing Schreier sums of them define $\alpha$-RIS sequences. \subsection*{Evaluation of type II functionals on $\{x_k\}_k$ with $\be\big(\{x_k\}_k\big) = 0$} \begin{dfn} Let $x_1<x_2<x_3$ be vectors in $\mathfrak{X}_{_{^\text{ISP}}}$, $f = \frac{1}{2}\sum_{j=1}^df_j$ be a functional of type II, such that $\supp f\cap\ran x_i\neq\varnothing$, for $i=1,2,3$ and $j_0 = \min\{j: \ran f_j\cap \ran x_2\neq\varnothing\}$. If $\ran f_{j_0}\cap\ran x_3 = \varnothing$, then we say that $f$ separates $x_1,x_2,x_3$. \end{dfn} \begin{dfn} Let $i,j\in\mathbb{N}$. If there exists $f\in W$ a functional of type II, such that $i,j\in\widehat{w}(f)$, then we say that $i$ is compatible to $j$. \end{dfn} \begin{lem} Let $x_1 < x_2 <\cdots<x_m$ be vectors in $\mathfrak{X}_{_{^\text{ISP}}}$, such that there exist $\varepsilon>0$ and $\{f_k\}_{k=2}^{m-1}$ functionals of type II satisfying the following. \begin{enumerate} \item[(i)] $f_k$ separates $x_1,x_k,x_m$, for $k=2,\ldots,m-1$ \item[(ii)] If $f_k = \frac{1}{2}\sum_{j=1}^{d_k}f_j^k$ and $j_k = \min\{j: \ran f_j^k\cap\ran x_k\neq\varnothing\}$, then $w(f_{j_k}^k)$ is not compatible to $w(f_{j_\ell}^\ell)$ for $k\neq\ell$. \item[(iii)] $|f_k(x_m)| > \varepsilon$ for $k=2,\ldots,m-1$ \end{enumerate} Then there exists a $\beta$-average $\beta$ in $W$ of size $s(\beta) = m-2$ such that $\beta(x_m) > \varepsilon$.\label{lem4.3} \end{lem} \begin{proof} Set $g_k = \sgn(f_k(x_m))f_k|_{\ran x_m}$, for $k=2,\ldots,m-1$. Then $g_k$ is a functional of type II in $W$. We will show that the $g_k$ have disjoint weights $\widehat{w}(g_k)$. Towards a contradiction, suppose that there exist $2\leqslant k < \ell\leqslant m-1$ and $i\in\widehat{w}(g_k)\cap\widehat{w}(g_\ell)$. By (i) and the way type II functionals are constructed, it follows that $f_k|_{[\min\supp x_2,\ldots,\max\supp f^\ell_{j_\ell}]} = \substack{+\\[-2pt]-}f_\ell|_{[\min\supp x_2,\ldots,\max\supp f^\ell_{j_\ell}]}$. This contradicts (ii). By the above, it follows that if we set $\beta = \frac{1}{m-2}\sum_{k=2}^{m-1}g_k$, then $\beta$ is the desired $\beta$-average. \end{proof} \begin{lem} Let $x_1 < x_2 <\cdots<x_m$ be vectors in $\mathfrak{X}_{_{^\text{ISP}}}$, such that there exist $\varepsilon>0$ and $\{f_k\}_{k=2}^{m-1}$ functionals of type II satisfying the following. \begin{enumerate} \item[(i)] $f_k$ separates $x_1,x_k,x_m$, for $k=2,\ldots,m-1$ \item[(ii)] If $f_k = \frac{1}{2}\sum_{j=1}^{d_k}f_j^k$ and $j_k = \min\{j: \ran f_j^k\cap\ran x_k\neq\varnothing\}$, then $w(f_{j_k}^k) = w(f_{j_\ell}^\ell)$ for $k\neq\ell$. \item[(iii)] If $j_k^\prime = \min\{j: \ran f_j^k\cap\ran x_m\neq\varnothing\}$, then $w(f_{j_k^\prime}^k) \neq w(f_{j_\ell^\prime}^\ell)$ for $k\neq\ell$. \item[(iv)] $|f_k(x_m)| > \varepsilon$ for $k=2,\ldots,m-1$ \end{enumerate} Then there exists a $\beta$-average $\beta$ in $W$ of size $s(\beta) = m-2$ such that $\beta(x_m) > \varepsilon$.\label{lem4.4} \end{lem} \begin{proof} As before, set $g_k = \sgn(f_k(x_m))f_k|_{\ran x_m}$, for $k=2,\ldots,m-1$. Then $g_k$ is a functional of type II in $W$. We will show that the $g_k$ have disjoint weights $\widehat{w}(g_k)$. Suppose that there exist $2\leqslant k < \ell\leqslant m-1$ and $i\in\widehat{w}(g_k)\cap\widehat{w}(g_\ell)$. By (i), (ii) and the way type II functionals are constructed, it follows that \begin{equation*} f_k|_{[\min\supp x_2,\ldots,\min\supp x_{m}]} = \substack{+\\[-2pt]-}f_\ell|_{[\min\supp x_2,\ldots,\min\supp x_{m}]} \end{equation*} This leaves us no choice, but to conclude that $w(f_{j_k^\prime}^k) = w(f_{j_\ell^\prime}^\ell)$, a contradiction. It follows that if we set $\beta = \frac{1}{m-2}\sum_{k=2}^{m-1}g_k$, then $\beta$ is the desired $\beta$-average. \end{proof} \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a bounded block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that $\be\big(\{x_k\}_k\big) = 0$. Then for any $\varepsilon>0$, there exists $M$ an infinite subset of the naturals, such that for any $k_1 < k_2 < k_3\in M$, for any functional $f\in W$ of type II that separates $x_{k_1},x_{k_2},x_{k_3}$, we have that $|f(x_{k_i})| < \varepsilon$, for some $i\in\{1,2,3\}$.\label{prop4.5} \end{prp} \begin{proof} Suppose that this is not the case. Then by using Ramsey theorem \cite{Ra}, we may assume that there exists $\varepsilon>0$ such that for any $k < \ell < m\in\mathbb{N}$, we have that there exists $f_{k,\ell,m}$ a functional of type II, that separates $x_k, x_\ell, x_m$ and $|f_{k,\ell,m}(x_k)| > \varepsilon, |f_{k,\ell,m}(x_\ell)| > \varepsilon$ and $|f_{k,\ell,m}(x_m)| > \varepsilon$. For $1<k<m$, if $f_{1,k,m} = \frac{1}{2}\sum_{j=1}^{d_{k,m}}f_j^{k,m}$, set \begin{eqnarray*} i_{k,m} &=& \min\{j: \ran f_j^{k,m}\cap\ran x_1\neq\varnothing\}\\ j_{k,m} &=& \min\{j: \ran f_j^{k,m}\cap\ran x_k\neq\varnothing\}\\ j_{k,m}^\prime &=& \min\{j: \ran f_j^{k,m}\cap\ran x_m\neq\varnothing\} \end{eqnarray*} Notice, that for $1<k<m$, since $|f_{1,k,m}(x_1)| > \varepsilon$, it follows that \begin{equation*} \frac{1}{2^{w(f^{k,m}_{i_{k,m}})}} > \frac{\varepsilon}{\|x_1\|\max\supp x_1} \end{equation*} By applying Ramsey theorem once more, we may assume that there exists $n_1\in\mathbb{N}$, such that for any $1<k<m$, we have that $w(f^{k,m}_{i_{k,m}}) = n_1$ Arguing in the same way and diagonalizing, we may assume that for any $k>1$, there exists $n_k\in\mathbb{N}$ such that for any $m>k$, we have that $w(f^{k,m}_{j_{k,m}}) = n_k$. Set \begin{eqnarray*} A_1 &=& \big\{\{k,\ell\}\in[\mathbb{N}\setminus\{1\}]^2: n_k\neq n_\ell\;\text{and}\;n_k\;\text{is compatible to}\;n_\ell\big\}\\ A_2 &=& \big\{\{k,\ell\}\in[\mathbb{N}\setminus\{1\}]^2: n_k\neq n_\ell\;\text{and}\;n_k\;\text{is not compatible to}\;n_\ell\big\}\\ A_3 &=& \big\{\{k,\ell\}\in[\mathbb{N}\setminus\{1\}]^2: n_k = n_\ell\big\} \end{eqnarray*} Once more, Ramsey theorem yields that there exists $M$ an infinite subset of the naturals, such that $[M]^2\subset A_1, [M]^2\subset A_2$, or $[M]^2\subset A_3$. Assume that $[M]^2\subset A_1$ and for convenience assume that $M = \mathbb{N}\setminus\{1\}$. Choose $k_0>1$ such that $k_0 > \max\supp x_1$. Since $n_1$ is compatible to $n_2$ and in general $n_{k-1}$ is compatible to $n_{k}$, for $k>1$, it follows that there exists a functional $f = \frac{1}{2}\sum_{j=1}^df_j$ of type II in $W$, such that $\ran f\cap \ran x_1\neq\varnothing$ and for $k=1,\ldots,k_0$ there exists $j_k$, with $w(f_{j_k}) = n_k$, for $k=1,\ldots,k_0$. Since $\min\supp f_1\leqslant\max\supp x_1$ it follows that $\{f_j\}_{j=1}^d$ can not be $\mathcal{S}_1$-admissible, a contradiction. Assume next that $[M]^2\subset A_2$. Lemma \ref{lem4.3} yields that $\be\big(\{x_k\}_k\big) > 0$ and since this cannot be, we conclude that $[M]^2\subset A_3$, therefore there exists $n_0\in\mathbb{N}$, such that $n_k = n_0$, for any $k\in M$. Assume once more that $M = \mathbb{N}\setminus\{1\}$ and set \begin{equation*} B = \big\{\{k,\ell,m\}\in[\mathbb{N}\setminus\{1\}]^3: w(f^{1,k,m}_{j_{k,m}^\prime}) = w(f^{1,\ell,m}_{j_{\ell,m}^\prime})\big\} \end{equation*} If there exists $M$ an infinite subset of the naturals, such that $[M]^3\subset B^c$, Lemma \ref{lem4.4} yields that $\be\big(\{x_k\}_k\big) > 0$, therefore by one last Ramsey argument, there exists $M$ an infinite subset of the naturals, such that $[M]^3\subset B$. By the above, we conclude that for $m\geqslant4$, $\ran x_k\subset \ran f^{2,m}_{j_{2,m}}$ and $|f^{2,m}_{j_{2,m}}(x_k)| > 2\varepsilon$, for $k= 2,\ldots,m-2$. Set $f_m = f^{2,m}_{j_{2,m}}$ and let $f$ be a $w^*$ limit of some subsequence of $\{f_m\}_{m\in\mathbb{N}}$. Then $|f(x_k)|\geqslant 2\varepsilon$, for any $k\geqslant 2$. Corollary \ref{cor2.22} yields a contradiction and this completes the proof. \end{proof} \begin{rmk} The proof of Proposition \ref{prop4.5} is the only place where the condition $\be\big(\{x_k\}_k\big) = 0$ is needed. This makes necessary to introduce the $\beta$-averages and their use in the definition of the norm. \end{rmk} \subsection*{Finite sequences equivalent to $ \ell_\infty^n$ basis.} \begin{prp} Let $x_1<\cdots<x_n$ be a seminormalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that $\|x_k\|\leqslant 1$ for $k=1,\ldots,n$ and there exist $n+3\leqslant j_1 <\cdots< j_n$ strictly increasing naturals such that the following are satisfied. \begin{enumerate} \item[(i)] For any $k_0\in\{1,\ldots,n\}$, for any $k\geqslant k_0, k\in\{1,\ldots,n\}$, for any $\{\alpha_q\}_{q=1}^d$ very fast growing and $\mathcal{S}_j$-admissible sequence of $\alpha$-averages, with $j<j_{k_0}$ and $s(\alpha_1) > \min\supp x_{k_0}$, we have that $\sum_{q=1}^d|\alpha_q(x_k)| < \frac{1}{n\cdot2^n}$. \item[(ii)] For any $k_0\in\{1,\ldots,n\}$, for any $k\geqslant k_0, k\in\{1,\ldots,n\}$, for any $\{\beta_q\}_{q=1}^d$ very fast growing and $\mathcal{S}_j$-admissible sequence of $\beta$-averages, with $j<j_{k_0}$ and $s(\beta_1) > \min\supp x_{k_0}$, we have that $\sum_{q=1}^d|\beta_q(x_k)| < \frac{1}{n\cdot2^n}$. \item[(iii)] For $k = 1,\ldots,n-1$, the following holds: $\frac{1}{2^{j_{k+1}}}\max\supp x_k < \frac{1}{2^n}$. \item[(iv)] For any $1\leqslant k_1 < k_2 < k_3\leqslant n$, for any functional $f\in W$ of type II that separates $x_{k_1},x_{k_2},x_{k_3}$, we have that $|f(x_{k_i})| < \frac{1}{n\cdot2^n}$, for some $i\in\{1,2,3\}$. \end{enumerate} Then $\{x_k\}_{k=1}^n$ is equivalent to $\ell_\infty^n$ basis, with an upper constant $3 + \frac{3}{2^n}$. Moreover, for any functional $f\in W$ of type I$_\alpha$ with weight $w(f) = j < j_1$, we have that $|f(\sum_{k=1}^nx_k)| < \frac{3 + \frac{4}{2^n}}{2^j}$.\label{prop4.6} \end{prp} \begin{proof} By using Remark \ref{remark1.2}, we will inductively prove, that for any $\{c_k\}_{k=1}^n\subset[-1,1]$ the following hold. \begin{enumerate} \item[(i)] For any $f\in W$, we have that $|f(\sum_{k=1}^nc_kx_k)| < (3 + \frac{3}{2^n})\max\{|c_k|:k=1,\ldots,n\}$. \item[(ii)] If $f$ is of type I$_\alpha$ and $w(f)\geqslant 2$, then $|f(\sum_{k=1}^nc_kx_k)| < (1 + \frac{2}{2^n})\max\{|c_k|:k=1,\ldots,n\}$. \item[(iii)] If $f$ is of type I$_\alpha$ and $w(f) = j < j_1$, then $|f(\sum_{k=1}^nc_kx_k)| < \frac{3 + \frac{4}{2^n}}{2^j}\max\{|c_k|:k=1,\ldots,n\}$. \end{enumerate} For any functional $f\in W_0$ the inductive assumption holds. Assume that it holds for any $f\in W_m$ and let $f\in W_{m+1}$. If $f$ is a convex combination, then there is nothing to prove. Assume that $f$ is of type I$_\alpha, f = \frac{1}{2^j}\sum_{q=1}^d\alpha_q$, where $\{\alpha_q\}_{q=1}^d$ is a very fast growing and $\mathcal{S}_j$-admissible sequence of $\alpha$-averages in $W_m$. Set $k_1 = \min\{k: \ran f\cap\ran x_k\neq\varnothing\}$ and $q_1 = \min\{q: \ran \alpha_q\cap\ran x_{k_1}\neq\varnothing\}$. We distinguish 3 cases.\vskip3pt \noindent {\em Case 1:} $j<j_1$. For $q>q_1$, we have that $s(\alpha_q) > \min\supp x_{k_1}$, therefore we conclude that \begin{equation} \sum_{q>q_1}|\alpha_q(\sum_{k=1}^nc_kx_k)| < \frac{1}{2^n}\max\{|c_k|:k=1,\ldots,n\} \label{prop4.6eq1} \end{equation} while the inductive assumption yields that \begin{equation} |\alpha_{q_1}(\sum_{k=1}^nc_kx_k)| < (3 + \frac{3}{2^n})\max\{|c_k|:k=1,\ldots,n\} \label{prop4.6eq2} \end{equation} Then \eqref{prop4.6eq1} and \eqref{prop4.6eq2} allow us to conclude that \begin{equation} |f(\sum_{k=1}^nc_kx_k)| < \frac{3 + \frac{4}{2^n}}{2^j}\max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eqx1} \end{equation} Hence, (iii) from the inductive assumption is satisfied.\vskip3pt \noindent {\em Case 2:} There exists $k_0 < n$, such that $j_{k_0}\leqslant j < j_{k_0 + 1}$. Arguing as previously we get that \begin{equation} |f(\sum_{k>k_0}c_kx_k)| < \frac{3 + \frac{4}{2^n}}{2^{j_{k_0}}}\max\{|c_k|:k=1,\ldots,n\} \label{prop4.6eq3} \end{equation} and \begin{equation} |f(\sum_{k<k_0}c_kx_k)| < \frac{1}{2^n}\max\{|c_k|:k=1,\ldots,n\} \label{prop4.6eq4} \end{equation} Using \eqref{prop4.6eq3}, \eqref{prop4.6eq4}, the fact that $|f(x_{k_0})| \leqslant 1$ and $j_{k_0}\geqslant n+3$, we conclude that \begin{equation} |f(\sum_{k=1}^nc_kx_k)| < (1 + \frac{2}{2^n})\max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eqx2} \end{equation} \noindent {\em Case 3:} $j\geqslant j_n$ By using the same arguments, we conclude that \begin{equation} |f(\sum_{k=1}^nc_kx_k)| < (1 + \frac{1}{2^n})\max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eqx3} \end{equation} Then \eqref{prop4.6eqx1}, \eqref{prop4.6eqx2} and \eqref{prop4.6eqx3} yield that (ii) from the inductive assumption is satisfied. If $f$ is of type I$_\beta$, then the proof is exactly the same, therefore assume that $f$ is of type II, $f = \frac{1}{2}\sum_{j=1}^df_j$, where $\{f_j\}_{j=1}^d$ is an $\mathcal{S}_1$-admissible sequence of functionals of type I$_\alpha$ in $W_m$. Set \begin{eqnarray*} E &=& \{k: |f(x_k)| \geqslant \frac{1}{n\cdot2^n}\}\\ E_1 &=& \{k\in E:\;\text{there exist at least two}\;j\;\text{such that}\;\ran f_j\cap\ran x_k\neq\varnothing\} \end{eqnarray*} Then $\#E_1\leqslant 2$. Indeed, if $k_1<k_2<k_3\in E_1$, then $f$ separates $x_{k_1},x_{k_2}$ and $x_{k_3}$ which contradicts our initial assumptions. If moreover we set $J = \{j:$ there exists $k\in E\setminus E_1$ such that $\ran f_j\cap\ran x_k\neq\varnothing\}$, then for the same reasons we get that $\#J\leqslant 2$. Since for any $j$, we have that $w(f_j)\in L$, we get that $w(f_j)>2$, therefore: \begin{eqnarray} |f(\sum_{k \in E\setminus E_1}^nc_kx_k)| &<& (1 + \frac{2}{2^n})\max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eq5}\\ |f(\sum_{k \in E_1}^nc_kx_k)| &\leqslant& 2 \max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eq6}\\ |f(\sum_{k \notin E}^nc_kx_k)| &\leqslant& n\cdot\frac{1}{n\cdot2^n}\max\{|c_k|:k=1,\ldots,n\}\label{prop4.6eq7} \end{eqnarray} Finally, \eqref{prop4.6eq5} to \eqref{prop4.6eq7} yield the following. \begin{equation*} |f(\sum_{k=1}^nc_kx_k)| < (3 + \frac{3}{2^n})\max\{|c_k|:k=1,\ldots,n\} \end{equation*} This means that (i) from the inductive assumption is satisfied an this completes the proof. \end{proof} \subsection*{The spreading models of $\mathfrak{X}_{_{^\text{ISP}}}$.} In this subsection we show that every seminormalized block sequence has a subsequence which generates either $\ell_1$ or $c_0$ as a spreading model. \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a seminormalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, such that $\|x_k\|\leqslant 1$ for all $k\in\mathbb{N}$ and $\al\big(\{x_k\}_k\big) = 0$ as well as $\be\big(\{x_k\}_k\big) = 0$. Then it has a subsequence, again denoted by $\{x_k\}_{k\in\mathbb{N}}$ satisfying the following. \begin{enumerate} \item[(i)] $\{x_k\}_{k\in\mathbb{N}}$ generates a $c_0$ spreading model. More precisely, for any $n\leqslant k_1<\cdots<k_n$, we have that $\|\sum_{i=1}^nx_{k_i}\| \leqslant 4$. \item[(i)] There exists a strictly increasing sequence of naturals $\{j_n\}_{n\in\mathbb{N}}$, such that for any $n\leqslant k_1 < \cdots <k_n$, for any functional $f$ of type I$_\alpha$ with $w(f) = j<j_n$, we have that \begin{equation*} |f(\sum_{i=1}^nx_{k_i})| < \frac{4}{2^j} \end{equation*} \end{enumerate}\label{cor4.7} \end{prp} \begin{proof} By repeatedly applying Proposition \ref{prop4.5} and diagonalizing, we may assume that for any $n\leqslant k_1 < k_2 < k_3$, for any functional $f$ of type II that separates $x_{k_1},x_{k_2}$ and $x_{k_3}$, we have that $|f(x_{k_i})| < \frac{1}{n\cdot2^n}$, for some $i\in\{1,2,3\}$. Use Propositions \ref{prop3.3} and \ref{prop3.4} to inductively choose a subsequence of $\{x_k\}_{k\in\mathbb{N}}$, again denoted by $\{x_k\}_{k\in\mathbb{N}}$ and $\{j_k\}_{k\in\mathbb{N}}$ a strictly increasing sequence of naturals with $j_k\geqslant k+3$ for all $k\in\mathbb{N}$, such that the following are satisfied. \begin{enumerate} \item[(i)] For any $k_0\in\mathbb{N}$, for any $k\geqslant k_0$, for any $\{\alpha_q\}_{q=1}^d$ very fast growing and $\mathcal{S}_j$-admissible sequence of $\alpha$-averages, with $j<j_{k_0}$ and $s(\alpha_1) > \min\supp x_{k_0}$, we have that $\sum_{q=1}^d|\alpha_q(x_k)| < \frac{1}{k_0\cdot2^{k_0}}$. \item[(ii)] For any $k_0\in\mathbb{N}$, for any $k\geqslant k_0$, for any $\{\beta_q\}_{q=1}^d$ very fast growing and $\mathcal{S}_j$-admissible sequence of $\beta$-averages, with $j<j_{k_0}$ and $s(\beta_1) > \min\supp x_{k_0}$, we have that $\sum_{q=1}^d|\beta_q(x_k)| < \frac{1}{k_0\cdot2^{k_0}}$. \item[(iii)] For $k\in\mathbb{N}$, the following holds: $\frac{1}{2^{j_{k+1}}}\max\supp x_k < \frac{1}{2^k}$. \end{enumerate} It is easy to check that for $n\leqslant k_1<\cdots<k_n$, the assumptions of Proposition \ref{prop4.6} are satisfied. \end{proof} Propositions \ref{prop3.5} and \ref{cor4.7} yield the following. \begin{cor} Let $\{x_k\}_{k\in\mathbb{N}}$ be a normalized weakly null sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. Then it has a subsequence that generates a spreading model which is either equivalent to $c_0$, or to $\ell_1$.\label{cor4.8} \end{cor} \begin{dfn} A pair $\{x,f\}$, where $x\in\mathfrak{X}_{_{^\text{ISP}}}$ and $f\in W$ is called an $(n,1)$-exact pair, if the following hold. \begin{itemize} \item[(i)] $f$ is a functional of type I$_\alpha$ with $w(f) = n, \min\supp x\leqslant\min\supp f$ and $\max\supp x\leqslant \max\supp f$. \item[(ii)] There exists $x^\prime\in\mathfrak{X}_{_{^\text{ISP}}}$ a $(4,1,n)$ exact vector, such that $\frac{28}{29}<f(x^\prime)\leqslant 1$ and $x = \frac{x^\prime}{f(x^\prime)}$. \end{itemize} A pair $\{x,f\}$, where $x\in\mathfrak{X}_{_{^\text{ISP}}}$ and $f\in W$ is called an $(n,0)$-exact pair, if the following hold. \begin{itemize} \item[(i)] $f$ is a functional of type I$_\alpha$ with $w(f) = n, \min\supp x\leqslant\min\supp f$ and $\max\supp x\leqslant \max\supp f$. \item[(ii)] $x$ is a $(4,1,n)$ exact vector and $f(x) = 0$. \end{itemize} \label{defexactpair} \end{dfn} \begin{rmk} If $\{x,f\}$ is an $(n,1)$-exact pair, then $f(x) = 1$ and by Remark \ref{remexactest} we have that $1\leqslant\|x\|\leqslant 29$.\label{remestexactpair} \end{rmk} \begin{prp} Let $\{x_k\}_{k\in\mathbb{N}}$ be a normalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$, that generates a $c_0$ spreading model. Then there exists $\{F_k\}_{k\in\mathbb{N}}$ an increasing sequence of subsets of the naturals such that $\#F_k\leqslant\min F_k$ for all $k\in\mathbb{N}$ and $\lim_k\#F_k = \infty$ such that by setting $y_k = \sum_{i\in F_k}x_k$, there exists a subsequence of $\{y_k\}_{k\in\mathbb{N}}$, which generates an $\ell_1^n$ spreading model, for all $n\in\mathbb{N}$. Furthermore, for any $k_0,n\in\mathbb{N}$, there exists $F$ a finite subset of $\mathbb{N}$ with $\min F\geqslant k_0$ and $\{c_k\}_{k\in F}$, such that \begin{enumerate} \item[(i)]$x^\prime = 2^n\sum_{k\in F}c_ky_k$ is a $(4,1,n)$ exact vector. \item[(ii)] For any $\eta > 0$ there exists a functional $f_\eta$ of type I$_\alpha$ of weight $w(f_\eta) = n$ such that $f_\eta(x^\prime) > 1-\eta, \min\supp x^\prime\leqslant\min\supp f_\eta$ and $\max\supp f_\eta > \max\supp x^\prime$. \end{enumerate} In particular, if $f = f_{1/29}, x = \frac{x^\prime}{f(x^\prime)}$, then $\{x,f\}$ is an $(n,1)$-exact pair. \label{prop4.9} \end{prp} \begin{proof} Since $\{x_k\}_{k\in\mathbb{N}}$ generates a $c_0$ spreading model, Proposition \ref{prop3.5} yields that $\al\big(\{x_k\}_k\big) = 0$ as well as $\be\big(\{x_k\}_k\big) = 0$, therefore passing, if necessary, to a subsequence $\{x_k\}_{k\in\mathbb{N}}$, satisfies the conclusion of Proposition \ref{cor4.7}. Choose $\{F_k\}_{k\in\mathbb{N}}$ an increasing sequence of subsets of the naturals, such that the following are satisfied. \begin{enumerate} \item[(i)] $\#F_k \leqslant \min F_k$ for all $k\in\mathbb{N}$. \item[(ii)] $\#F_{k+1} > \max\big\{\#F_k,\;2^{\max\supp x_{\max F_k}}\big\}$, for all $k\in\mathbb{N}$. \end{enumerate} By Proposition \ref{prop4.5} and Remark \ref{rem3.11}, we have that $1 \leqslant \|y_k\| \leqslant 4$, for all $k\in\mathbb{N}$ and passing, if necessary, to a subsequence, $\{y_k\}_{k\in\mathbb{N}}$ is $(4,\{n_k\}_{k\in\mathbb{N}})$\;$\alpha$-RIS. Moreover it is easy to see, that for any $k\in\mathbb{N}, \eta > 0$, there exists an $\alpha$-average $\alpha$ of size $s(\alpha) = \#F_k$, such that $\alpha(y_k) > 1 - \eta$ and $\ran\alpha\subset y_k$. This yields that $\al\big(\{y_k\}_k\big) > 0$, therefore we may apply Proposition \ref{prop3.5} to conclude that $\{y_k\}_{k\in\mathbb{N}}$ has a subsequence generating an $\ell_1^n$ spreading model, for all $n\in\mathbb{N}$. We now prove the second assertion. Let $k_0,n\in\mathbb{N}$ and fix $0< \varepsilon < (36\cdot4\cdot2^{3n})^{-1}$. By taking a larger $k_0$, we may assume that $n_{k_0} > 2^{2n}$. Set $\varepsilon^\prime = \varepsilon(1-\varepsilon)$ Proposition \ref{prop1.5} yields that there exists $\{d_1,\ldots,d_m\}$ a finite subset of $\{k: k\geqslant k_0\}$ and $\{c_k\}_{k=1}^m$ such that $x^{\prime\prime} = \sum_{k=1}^mc_ky_{d_k}$ is a $(n,\varepsilon^\prime)$ s.c.c. It is straightforward to check that $x^\prime = 2^n\sum_{k=1}^{m-1}\frac{c_k}{1-c_m}y_{d_k}$ is a $(4,\theta,n)$ exact vector. If (ii) also holds, it will follow that $\theta \geqslant 1$. For some $\eta>0$, $k = 1,\ldots,m$, choose an $\alpha$-average $\alpha_k$ of size $s(\alpha_k) = \#F_{d_k}$, such that $\alpha_k(y_{d_k})> 1 - \eta$ and $\ran\alpha\subset y_k$. Set $f = \frac{1}{2^n}(\sum_{k = 1}^m\alpha_k)$, which is a functional of type I$_\alpha$ of weight $w(f) =n$ such that $f(x^\prime) > 1 - \eta$ and $\max\supp f > \max\supp x$. \end{proof} \begin{cor} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$. Then $Y$ admits a spreading model equivalent to $c_0$ as well as a spreading model equivalent to $\ell_1$.\label{cor4.10} \end{cor} \begin{proof} Assume first that $Y$ is generated by some normalized block sequence $\{x_k\}_{k\in\mathbb{N}}$. Corollary \ref{cor3.17} and Proposition \ref{cor4.7} yield that it has a further normalized block sequence $\{y_k\}_{k\in\mathbb{N}}$, generating a spreading model equivalent to $c_0$. Proposition \ref{prop4.9} yields that $\{y_k\}_{k\in\mathbb{N}}$ has a further block sequence generating an $\ell_1$ spreading model. Since any subspace contains a sequence arbitrarily close to a block sequence, the result follows. \end{proof} Corollary \ref{cor4.10} and Proposition \ref{prop4.9} yield the following. \begin{cor} Let $X$ be a block subspace of $\mathfrak{X}_{_{^\text{ISP}}}$. Then for every $n\in\mathbb{N}$ there exists $x\in X$ and $f\in W$, such that $\{x,f\}$ is an $(n,1)$-exact pair.\label{corexactpairexist} \end{cor} We remind that, as Propositions \ref{prop3.5} and \ref{cor4.7} state, if a sequence generates an $\ell_1$ spreading model, then passing, if necessary, to a subsequence, it generates an $\ell_1^k$ spreading model for any $k\in\mathbb{N}$. However, as the next proposition states, the space $\mathfrak{X}_{_{^\text{ISP}}}$ does not admit higher order $c_0$ spreading models. \begin{prp} The space $\mathfrak{X}_{_{^\text{ISP}}}$ does not admit $c_0^2$ spreading models.\label{rem4.11} \end{prp} \begin{proof} Towards a contradiction, assume that there is a sequence $\{x_k\}_{k\in\mathbb{N}}$ in $\mathfrak{X}_{_{^\text{ISP}}}$, generating a $c_0^2$ spreading model. Then it must be weakly null and we may assume that it is a normalized block sequence. By Proposition \ref{prop4.9}, it follows that there exist $\{F_k\}_{k\in\mathbb{N}}$ increasing, Schreier admissible subsets of the naturals and $c>0$ such that $\|\sum_{j=1}^n\sum_{i\in F_{k_j}}x_i\| \geqslant n\cdot c$ for any $n\leqslant k_1 <\ldots< k_n$. Since for any such $F_{k_1}<\cdots< F_{k_n}$ we have that $\cup_{j=1}^n F_{k_j}\in\mathcal{S}_2$, it follows that $\{x_k\}_{k\in\mathbb{N}}$ does not generate a $c_0^2$ spreading model. \end{proof} \begin{cor}Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$. Then $Y^*$ admits a spreading model equivalent to $\ell_1$ as well as a spreading model equivalent to $c_0^n$, for any $n\in\mathbb{N}$.\label{cor4.12} \end{cor} \begin{proof} Since $Y$ contains a sequence $\{x_k\}_{k\in\mathbb{N}}$ generating a spreading model equivalent to $c_0$, which we may assume is Schauder basic, then for any normalized $\{x_k^*\}_{k\in\mathbb{N}}\subset Y^*$, such that $x^*_k(x_m) = \delta_{n,m}$ for $n,m\in\mathbb{N}$, we have that passing, if necessary, to a subsequence, $\{x_k^*\}_{k\in\mathbb{N}}$ generates a spreading model equivalent to $\ell_1$. To see that $Y^*$ admits a spreading model equivalent to $c_0^n$ for any $n\in\mathbb{N}$, take the previously used sequence $\{x_k\}_{k\in\mathbb{N}}$. Working just like in the proof of Proposition \ref{prop4.9} find $\{F_k\}_{k\in\mathbb{N}}$ successive subsets of the natural such that $\min F_k\geqslant \#F_k$, for all $k\in\mathbb{N}$, if $y_k = \sum_{i\in F_k}x_i$ for all $k\in\mathbb{N}$, then $\{y_k\}_{k\in\mathbb{N}}$ is seminormalized and there exists a very fast growing sequence of $\alpha$-averages $\{\alpha_k\}_{k\in\mathbb{N}}\subset W$ such that $\lim\inf\alpha_k(\sum_{i\in F_k}x_i)\geqslant 1$. Then, if $c = \lim\sup_k\|y_k\|$,we evidently have that $\lim\inf_k\|\alpha_k\|\geqslant 1/c$ and since for any $n\in\mathbb{N}, F\in\mathcal{S}_n$, we have that $\frac{1}{2^n}\sum_{q\in F}\alpha_{q}$ is a functional of type I$_\alpha$ in $W$, it follows that $\|\sum_{q\in F}^d\alpha_{q}\|\leqslant 2^n$. This means that, $\{\alpha_k\}_{k\in\mathbb{N}}$ generates a spreading model equivalent to $c_0^n$, with an upper constant $2^n$. Let $I^*:\mathfrak{X}_{_{^\text{ISP}}}^*\rightarrow Y^*$ be the dual operator of $I:Y\rightarrow\mathfrak{X}_{_{^\text{ISP}}}$. Then $\{I^*(\alpha_k)\}_{k\in\mathbb{N}}$ generates a spreading model equivalent to $c_0^n$, for any $n\in\mathbb{N}$. Since $\|I^*\| = 1$, all that needs to be shown is that $\lim\inf_k\|I^*(\alpha_k)\| > 0$. Indeed, \begin{equation*} \liminf_k\|I^*(\alpha_k)\|\geqslant \liminf_k(I^*\alpha_k)(\frac{\sum_{i\in F_k}x_i}{c}) = \liminf_k\alpha_k(\frac{\sum_{i\in F_k}x_i}{c})\geqslant 1/c \end{equation*} \end{proof} \section{Properties of $\mathfrak{X}_{_{^\text{ISP}}}$ and $\mathcal{L}(\mathfrak{X}_{_{^\text{ISP}}})$} In this final section it is proved that $\mathfrak{X}_{_{^\text{ISP}}}$ is hereditarily indecomposable and the properties of the operators acting on infinite dimensional closed subspaces of $\mathfrak{X}_{_{^\text{ISP}}}$ are presented. \subsection*{Dependent sequences and the HI property of $\mathfrak{X}_{_{^\text{ISP}}}$.} In the first part of this subsection we introduce the dependent sequences, which are the main tool for proving the HI property of $\mathfrak{X}_{_{^\text{ISP}}}$ and studying the structure of the operators. \begin{dfn} A sequence of pairs $\{x_k, f_k\}_{k=1}^n$, is said to be a 1-dependent sequence (respectively a 0-dependent sequence) if the following are satisfied. \begin{enumerate} \item[(i)] $\{x_k, f_k\}$ is an $(m_k,1)$-exact pair (respectively an $(m_k,0)$-exact pair) for $k=1,\ldots,n$, with $m_1 > 4n2^{2n}$ \item[(ii)] $\max\supp f_k < \min\supp x_{k+1}$ for $k=1,\ldots,n-1$ \item[(iii)] $\{f_k\}_{k=1}^n$ is an $\mathcal{S}_1$-admissible special sequence of type I$_\alpha$ functionals in $W$, i.e. $f = \frac{1}{2}\sum_{k=1}^nf_k$ is a functional of type II in $W$. \end{enumerate} \label{defdependent} \end{dfn} \begin{prp} Let $X$ be a block subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $n\in\mathbb{N}$. Then there exist $x_1,\ldots,x_n$ in $X$ and $f_1,\ldots,f_n$ in $W$, such that $\{x_k, f_k\}_{k=1}^n$ is a 1-dependent sequence. \label{dependentsequenceexist} \end{prp} \begin{proof} Choose $m_1\in L_1$ with $m_1 > 4n2^{2n}$. By Corollary \ref{corexactpairexist} there exists $\{x_1, f_1\}$ an $(m_1,1)$-exact pair in $X$. Then $\min\supp f_1 \geqslant \min\supp x_1 > n$. Let $d<n$ and suppose that we have chosen $\{x_k, f_k\}$ $(m_k,1)$-exact pairs for $k=1,\ldots,d$ such that $\{f_k\}_{k=1}^m$ is a special sequence and $\max\supp f_k < \min\supp x_{k+1}$ for $k=1,\ldots,d$. Set $m_{d+1} = \sigma\big((f_1,m_1),\ldots,(f_d,m_d)\big)$. Then applying Corollary \ref{corexactpairexist} once more, there exists $\{x_{d+1},f_{d+1}\}$ an $m_{d+1}$-exact pair in $X$, such that $\max\supp f_d < \min\supp x_{d+1}$. The inductive construction is complete and $\{x_k,f_k,\}_{k=1}^n$ is a 1-dependent sequence. \end{proof} An easy modification of the above proof yields the following. \begin{cor} If $X, Y$ are block subspaces of $\mathfrak{X}_{_{^\text{ISP}}}$ and $n\in\mathbb{N}$, then a 1-dependent sequence $\{x_k,f_k,\}_{k=1}^{2n}$ can be chosen, such that $x_{2k-1}\in X$ and $x_{2k}\in Y$ for $k=1,\ldots,n$.\label{remarkforhi} \end{cor} \begin{prp} Let $\{(x_k,f_k)\}_{k=1}^{2n}$ be a 1-dependent sequence in $\mathfrak{X}_{_{^\text{ISP}}}$ and set $y_k = x_{2k-1} - x_{2k}$, for $k=1,\ldots,n$. Then we have that: \begin{enumerate} \item[(i)] $\frac{1}{n}\|\sum_{k=1}^{2n}x_k\| \geqslant 1$ \item[(ii)] $\frac{1}{n}\|\sum_{k=1}^ny_k\| \leqslant \frac{232}{n}$ \end{enumerate}\label{prop5.2} \end{prp} \begin{proof} Since $\frac{1}{2}\sum_{k=1}^{2n}f_k$ is a type II functional in $W$, it immediately follows that $\frac{1}{n}\|\sum_{k=1}^{2n}x_k\| \geqslant \frac{1}{2n}\sum_{k=1}^{2n}f_k(x_k) = 1$. By Remark \ref{remestexactpair} it follows that $1 \leqslant \|y_k\| \leqslant 58$, for $k=1,\ldots,n$. Set $y_k^\prime = \frac{1}{58}y_k$ and $j_k = m_{2k-1}-2$ for $k=1,\ldots,n$. We will show that the assumptions of Proposition \ref{prop4.6} are satisfied. From this, it will follow that $\frac{1}{n}\|\sum_{k=1}^ny_k\| \leqslant 58\frac{4}{n}$, which is the desired result. The first and second assumptions, follow from Lemmas \ref{lem3.7} and \ref{lem3.15} respectively and the definition of the 1-dependent sequence. The third assumption follows from the fact that, by the definition of the 1-dependent sequence, $\max\supp f_k > \max\supp x_k$, for $k=1,\ldots,2n$ and the definition of the coding function $\sigma$. It remains to be proven that the fourth assumption is also satisfied. Let $1\leqslant k_1<k_2<k_3\leqslant n$ and $g = \frac{1}{2}\sum_{j=1}^dg_j$ be a functional of type II that separates $y_{k_1}^\prime,y_{k_2}^\prime$ and $y_{k_3}^\prime$. Set $j_0 = \min\{j: \ran g_j\cap\ran y_{k_3}^\prime\neq\varnothing\}$ and assume first that $w(f_{j_0}) = m_{2k_3-1}$ Since $\supp g\cap\supp y_{k_1}\neq\varnothing$, it follows that $g_{j_0-1} = f_{2k_3-2}$ and there exists $I$ an interval of the naturals, $\ran y_{k_2}^\prime\subset I$, such that $g = I(\frac{1}{2}\sum_{k=1}^{j_0 - 1}f_k)$. This yields that $g(y_{k_2}^\prime) = 0$. Otherwise, if $w(f_{j_0}) \neq m_{2k_3-1}$, set $g^\prime = g|_{\ran y_{k_3}^\prime}$ and Corollary \ref{corx3.14} yields the following. \begin{equation*} |g^\prime(y_{k_3}^\prime)| < \frac{2\cdot4}{58}\frac{29}{28}\big(\frac{1}{2^{m_{2k_3-1}}} + \frac{1}{2^{2m_{2k_3-1}}} + \sum_{\substack{j\in\widehat{w}(g^\prime):\\w(g_j)<n}}\frac{4}{2^{w(g_j)}}\big) \end{equation*} Since $g$ separates $y_{k_1}, y_{k_2}$ and $y_{k_3}$, we have that $\min\widehat{w}(g^\prime) > p_0 = \min\supp x_1$, therefore \begin{equation*} \sum_{\substack{j\in\widehat{w}(g^\prime):\\w(g_j)<n}}\frac{1}{2^{w(g_j)}} < \sum_{p > p_0}\frac{1}{2^p} = \frac{1}{2^{p_0}} \leqslant \frac{1}{2^{32\cdot2^{2m_1}}} \end{equation*} By the choice of $m_1$, we conclude that $|g(y_{k_3}^\prime)| < \frac{1}{n2^{n}}$, which means that the fourth assumption is satisfied. \end{proof} The next proposition is proved by using similar arguments. \begin{prp} Let $\{(x_k,f_k)\}_{k=1}^n$ be a 0-dependent sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. Then we have that: \begin{equation*} \frac{1}{n}\|\sum_{k=1}^nx_k\| \leqslant \frac{112}{n} \end{equation*}\label{prop5.3} \end{prp} We pass to the main structural property of $\mathfrak{X}_{_{^\text{ISP}}}$. \begin{thm} The space $\mathfrak{X}_{_{^\text{ISP}}}$ is hereditarily indecomposable.\label{cor5.4} \end{thm} \begin{proof} It is enough to show that for $X, Y$ block subspaces of $\mathfrak{X}_{_{^\text{ISP}}}$, for any $\varepsilon>0$, there exist $x\in X$ and $y\in Y$, such that $\|x+y\|\geqslant 1$ and $\|x-y\| < \varepsilon$. Let $n\in\mathbb{N},$ such that $\frac{232}{n}<\varepsilon$. By Corollary \ref{remarkforhi} there exists a 1-dependent sequence $\{x_k,f_k,\}_{k=1}^{2n}$, such that $x_{2k-1}\in X$ and $x_{2k}\in Y$ for $k=1,\ldots,n$. Set $x = \frac{1}{n}\sum_{k=1}^nx_{2k-1}$ and $y = \frac{1}{n}\sum_{k=1}^nx_{2k}$. By applying Proposition \ref{prop5.2}, the result follows. \end{proof} \subsection*{The structure of $\mathcal{L}(Y,\mathfrak{X}_{_{^\text{ISP}}})$} For $Y$ a closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow\mathfrak{X}_{_{^\text{ISP}}}$ we show that $T = \lambda I_{Y,\mathfrak{X}_{_{^\text{ISP}}}} + S$ with $S:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ strictly singular. \begin{prp} Let $Y$ be a subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a linear operator, such that there exists $\{x_k\}_{k\in\mathbb{N}}$ a sequence in $Y$ generating a $c_0$ spreading model and $\lim\sup\dist(Tx_k,\mathbb{R}x_k) > 0$. Then $T$ is unbounded.\label{prop5.5} \end{prp} \begin{proof} Passing, if necessary, to a subsequence, there exists $1 > \delta > 0$, such that $\dist(Tx_k,\mathbb{R}x_k) > \delta$, for any $k\in\mathbb{N}$. Since $\{x_k\}_{k\in\mathbb{N}}$ generates a $c_0$ spreading model, it is weakly null. Set $y_k = Tx_k$ and assume that $T$ is bounded. It follows that passing, if necessary, to a subsequence of $\{x_k\}_{k\in\mathbb{N}}$, then $\{y_k\}_{k\in\mathbb{N}}$ also generates a $c_0$ spreading model. We may assume that $\{x_k\}_{k\in\mathbb{N}}$, as well as $\{y_k\}_{k\in\mathbb{N}}$ are block sequences with rational coefficients. And $\lim_k\|x_k\| = 1$, as well as $\lim_k\|y_k\| = 1$. If this is not the case pass, if necessary, to a further subsequence of $\{x_k\}_{k\in\mathbb{N}}$, such that both $\{x_k\}_{k\in\mathbb{N}}$ and $\{y_k\}_{k\in\mathbb{N}}$ are equivalent to some block sequences with rational coefficients $\{x_k^\prime\}_{k\in\mathbb{N}}, \{y_k^\prime\}_{k\in\mathbb{N}}$ respectively, and moreover $\lim_k\|x_k^\prime\| = 1$, as well as $\lim_k\|y_k^\prime\| = 1$. Set $Y^\prime = [\{x_k^\prime\}_{k\in\mathbb{N}}]$ and $T^\prime:Y^\prime\rightarrow\mathfrak{X}_{_{^\text{ISP}}}$, such that $T^\prime x_k^\prime = y_k^\prime$. It is easy to check that $T^\prime$ is also bounded and $\dist(T^\prime x_k^\prime,\mathbb{R}x_k^\prime) > \delta^\prime$, for some $\delta^\prime > 0$. Set $I_k = \ran(\ran x_k\cup\ran y_k)$ and passing, if necessary, to a subsequence, we have that $\{I_k\}_{k\in\mathbb{N}}$ is an increasing sequence of intervals of the naturals. We will choose $\{f_k\}_{k\in\mathbb{N}}\subset W$, such that $f_k(y_k) > \frac{\delta}{5}$, $f_k(x_k) = 0$ and $\ran f_k \subset I_k$, for all $k\in\mathbb{N}$. The Hahn-Banach Theorem, yields that for all $k\in\mathbb{N}$, there exists $f^\prime_k\in B_{\mathfrak{X}_{_{^\text{ISP}}}^*}$, such that $f^\prime_k(y_k) > \delta$, $f_k^\prime(x_k) = 0$ and $\ran f_k^\prime \subset I_k$, for all $k\in\mathbb{N}$. By the fact that $\mathfrak{X}_{_{^\text{ISP}}}$ is reflexive, it follows that $W$ is norm dense in $B_{\mathfrak{X}_{_{^\text{ISP}}}^*}$, therefore there exists $f_k^{\prime\prime}\in W$ with $\|f_k^{\prime} - f_k^{\prime\prime}\| < \frac{\delta}{4}$ and $\ran f_k^{\prime\prime}\subset I_k$, for all $k\in\mathbb{N}$. It follows that $f_k^{\prime\prime}(y_k) > \frac{3\delta}{4}, |f_k^{\prime\prime}(x_k)| < \frac{\delta}{4}$ and $f_k^{\prime\prime}(x_k)$ is rational, for all $k\in\mathbb{N}$. Furthermore, there exists $g_k\in W$, such that $g_k(x_k) > 1 - \frac{\delta}{4}$, $g_k(x_k)$ is rational and $\ran g_k\subset I_k$, for all $k\in\mathbb{N}$. Set $f_k = \frac{1}{2}(f_k^{\prime\prime} - \frac{f_k^{\prime\prime}(x_k)}{g_k(x_k)}g_k)$. By doing some easy calculations, it follows that the $f_k$ are the desired functionals. For the rest of the proof we may assume that the $\{x_k\}_k$ are normalized. By copying the proof of Proposition \ref{prop4.9}, for any $k_0\in\mathbb{N}, n\in\mathbb{N}$, there exists $F$ a finite subset of the naturals with $\min F \geqslant k_0$ and $\{c_k\}_{k\in F}$ such that \begin{enumerate} \item[(i)] $z = 2^n\sum_{k\in F}c_kx_k$ is a $(4,1,n)$ exact vector \item[(ii)] There exists a functional $f$ of type I$_\alpha$ with weight $w(f) = n$ such that $f(z) = 0, \max\supp f > \max\supp z$ and if $w = 2^n\sum_{k\in F}c_ky_k$, then $f(w) > \frac{\delta}{5}$. \end{enumerate} Using the above fact and arguing in the same way as in the proof of Proposition \ref{dependentsequenceexist}, for some $n\in\mathbb{N}$, we construct a sequence $\{z_k\}_{k=1}^n$ and $\{g_k\}_{k=1}^n$ such that $\{(z_k, g_k)\}_{k=1}^n$ is 0-dependent and if $w_k = Tz_k$, then $g_k(w_k) > \frac{\delta}{5}$ and $\ran g_k\cap\ran w_m = \varnothing$, for $k\neq m$. Then $f = \frac{1}{2}\sum_{k=1}^ng_k$ is a functional of type II and $\frac{1}{n}\|\sum_{k=1}^nw_k\| \geqslant \frac{1}{2n}\sum_{k=1}^ng_k(w_k) > \frac{\delta}{10}$. Moreover, Proposition \ref{prop5.3} yields that $\frac{1}{n}\|\sum_{k=1}^nz_k\| \leqslant\frac{112}{n}$. It follows that $\|T\| > \frac{n\cdot\delta}{1120}$. Since $n$ was randomly chosen, $T$ cannot be bounded, a contradiction which completes the proof. \end{proof} In \cite{F}, it is proven that if $X$ is a hereditarily indecomposable complex Banach space, $Y$ is a subspace of $X$ and $T:Y\rightarrow X$ is a bounded linear operator, then there exists $\lambda\in\mathbb{C}$, such that $T - \lambda I_{_{Y,X}}: Y \rightarrow X$ is strictly singular. Here we prove a similar result for $\mathfrak{X}_{_{^\text{ISP}}}$. \begin{thm} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a bounded linear operator. Then there exists $\lambda\in\mathbb{R}$, such that $T - \lambda I_{_{Y,\mathfrak{X}_{_{^\text{ISP}}}}}: Y \rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ is strictly singular.\label{prop5.6} \end{thm} \begin{proof} If $T$ is strictly singular, then evidently $\lambda = 0$ is the desired scalar. Otherwise, choose $Z$ an infinite dimensional closed subspace of $Y$, such that $T:Z\rightarrow\mathfrak{X}_{_{^\text{ISP}}}$ is an into isomorphism. Choose $\{x_k\}_{k\in\mathbb{N}}$ a normalized sequence in $Z$ generating a $c_0$ spreading model. Proposition \ref{prop5.5} yields that $\lim_k\dist(Tx_k,\mathbb{R}x_k) = 0$. Choose $\{\lambda_k\}_{k\in\mathbb{N}}$ scalars, such that $\lim_k\|Tx_k - \lambda_k x_k\| = 0$ and $\lambda$ a limit point of $\{\lambda_k\}_{k\in\mathbb{N}}$. We will prove that $S = T - \lambda I_{_{Y,\mathfrak{X}_{_{^\text{ISP}}}}}$ is strictly singular. Towards a contradiction, suppose that this is not the case. Then there exists $\{y_k\}_{k\in\mathbb{N}}$ a normalized sequence in $Y$ generating a $c_0$ spreading model and $\delta>0$, such that $\|Sy_k\| = \|(T - \lambda I_{_{Y,\mathfrak{X}_{_{^\text{ISP}}}}})y_k\| > \delta$, for all $k\in\mathbb{N}$. As previously, we may assume that $\{x_k\}_{k\in\mathbb{N}}, \{y_k\}_{k\in\mathbb{N}}$ as well as $\{Sy_k\}_{k\in\mathbb{N}}$ are all normalized block sequences generating $c_0$ spreading models. By Proposition \ref{prop5.5} and passing, if necessary, to a subsequence, there exists $\mu\in\mathbb{R}$, such that $\lim_k\|Sy_k - \mu y_k\| = 0$. Evidently $\mu\neq 0$, otherwise we would have that $\lim_k\|Sy_k\| = 0$. Pass, if necessary, to a further subsequence of $\{y_k\}_{k\in\mathbb{N}}$, such that $\sum_{k=1}^\infty\|Sy_k - \mu y_k\| < \frac{|\mu|}{232}$. Observe that $\lim_k\|Sx_k\| = 0$ and therefore we may pass, if necessary, to a subsequence of $\{x_k\}_{k\in\mathbb{N}}$, such that $\sum_{k=1}^\infty\|Sx_k\| < \frac{|\mu|}{232}$. Arguing in the same manner as in the proof of Proposition \ref{dependentsequenceexist}, for some $n\in\mathbb{N}$ construct $\{z_k\}_{k=1}^{2n}$ and $\{f_k\}_{k=1}^{2n}$ such that $z_{2k-1}$ is a linear combination of $\{y_k\}_{k\in\mathbb{N}}$, $z_{2k}$ is a linear combination of $\{x_k\}_{k\in\mathbb{N}}$ and $\{(z_k, f_k)\}_{k=1}^{2n}$ is a 1-dependent sequence. Set $f = \frac{1}{2}\sum_{k=1}^{2n}f_k$, which is a functional of type II in $W$. If $w_k = z_{2k-1} - z_{2k}$, Proposition \ref{prop5.2} yields that $\frac{1}{n}\|\sum_{k=1}^nw_k\| \leqslant \frac{232}{n}$. On the other hand, we have that \begin{eqnarray*} \frac{1}{n}\|\sum_{k=1}^nSw_k\| &\geqslant& \frac{1}{n}\big(\|\sum_{k=1}^nSz_{2k-1}\| - \|\sum_{k=1}^nSz_{2k}\|\big) \end{eqnarray*} \begin{eqnarray*} \phantom{a}&\geqslant& \frac{1}{n}\big(\|\sum_{k=1}^n\mu z_{2k-1}\| - \|\sum_{k=1}^n(Sz_{2k-1} - \mu z_{2k-1})\| - \frac{29|\mu|}{232}\big)\\ &\geqslant& \frac{1}{n}\big(\frac{|\mu|}{2}\sum_{k=1}^nf_{2k-1}(z_{2k-1}) - \frac{29|\mu|}{232} - \frac{29|\mu|}{232}\big)\\ &=& \frac{|\mu|}{2} - \frac{|\mu|}{4n} \geqslant \frac{|\mu|}{4} \end{eqnarray*} It follows that $\|S\| \geqslant \frac{n|\mu|}{928}$, where $n$ was randomly chosen. This means that $S$ is unbounded, a contradiction completing the proof. \end{proof} \subsection*{Strictly Singular Operators} In this subsection we study the action of strictly singular operators on Schauder basic sequences in subspaces of $\mathfrak{X}_{_{^\text{ISP}}}$. \begin{prp} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a linear bounded operator. Then the following assertions are equivalent. \begin{enumerate} \item[(i)] $T$ is not strictly singular. \item[(ii)] There exists a sequence $\{x_k\}_{k\in\mathbb{N}}$ in $Y$ generating a $c_0$ spreading model, such that $\{Tx_k\}_{k\in\mathbb{N}}$ is not norm convergent to 0. \end{enumerate}\label{prop5.7} \end{prp} \begin{proof} Assume first that $T$ is not strictly singular and let $Z$ be an infinite dimensional closed subspace of $Y$, such that $T|_Z$ is an isomorphism. Since any subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ contains a sequence generating a $c_0$ spreading model, then so does $Z$. Since $T|_Z$ is an isomorphism, the second assertion is true. Assume now that there exists $\{x_k\}_{k\in\mathbb{N}}$ a sequence in $Y$ generating a $c_0$ spreading model, such that $\{Tx_k\}_{k\in\mathbb{N}}$ does not norm converge to 0. By Proposition \ref{prop5.5} and passing, if necessary to a subsequence, there exists $\lambda\neq 0$, such that $\lim_k\|Tx_k - \lambda x_k\| = 0$. Passing, if necessary, to a further subsequence, we have that $\sum_{k=1}^\infty\|Tx_k - \lambda x_k\| < \infty$. But this means that $\{x_k\}_{k\in\mathbb{N}}$ is equivalent to $\{Tx_k\}_{k\in\mathbb{N}}$, therefore $T$ is not strictly singular. \end{proof} \begin{dfn} Let $\{x_k\}_k$ be a normalized block sequence $\mathfrak{X}_{_{^\text{ISP}}}$. We say that $\{x_k\}_k$ is of rank 1, if $\al\big(\{x_k\}_k\big) = 0$ and $\be\big(\{x_k\}_k\big) = 0$.\label{defrank1} \end{dfn} \begin{dfn} Let $\{x_k\}_k$ be a normalized block sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. We say that $\{x_k\}_k$ is of rank 2, if it satisfies one of the following. \begin{itemize} \item[(i)] $\alpha\big(\{x_k\}_k\big) = 0$ and for every $L$ infinite subset of the natural numbers, $\beta\big(\{x_k\}_{k\in L}\big) > 0$ \item[(ii)] For every $L$ infinite subset of the natural numbers $\alpha\big(\{x_k\}_{k\in L}\big) > 0$ and for every $C\geqslant 1, \theta > 0, \{n_j\}_j$ strictly increasing sequence of natural numbers, $F_j \subset L$ and $\{c_k^j\}_{k\in F_j}, j\in\mathbb{N}$ such that $w_j = 2^{n_j}\sum_{k\in F_j}c_k^jx_k$ are $(C,\theta,n_j)$ vectors for every $j\in\mathbb{N}$, we have that $\beta\big(\{w_j\}_j\big) = 0$ \end{itemize} \label{defrank2} \end{dfn} \begin{dfn} Let $\{x_k\}_k$ be a normalized block sequence. We say that $\{x_k\}_k$ is of rank 3, if for every $L$ infinite subset of the natural numbers, $\alpha\big(\{x_k\}_{k\in L}\big) > 0$ and there exist $C\geqslant 1, \theta > 0, \{n_j\}_j$ strictly increasing sequence of natural numbers, $F_j \subset L$ and $\{c_k^j\}_{k\in F_j}, j\in\mathbb{N}$ such that $w_j = 2^{n_j}\sum_{k\in F_j}c_k^jx_k$ are $(C,\theta,n_j)$ are vectors for every $j\in\mathbb{N}$, and $\beta\big(\{w_j\}_j\big) > 0$ \label{defrank3} \end{dfn} \begin{rmk} It follows easily from the definitions above, that every normalized block sequence has a subsequence that is of some rank. Moreover, if a normalized block sequence is of some rank, then any of its subsequences is of the same rank. We would also like to point out that we can neither prove nor disprove the existence of sequences of rank 3. The failure of the existence of such sequences, would yield that the composition of any two strictly singular operators defined on a subspace of $\mathfrak{X}_{_{^\text{ISP}}}$, is a compact one.\label{rmkrankhereditary} \end{rmk} \begin{dfn} Let $\{x_k\}_k$ be a weakly null sequence in $\mathfrak{X}_{_{^\text{ISP}}}$. If it norm null, then we say that it is of rank 0. If it is seminormalized, we say that $\{x_k\}_k$ is of rank $i$, if there exists a normalized block sequence $\{x_k^\prime\}_k$ which is of rank $i$, such that $\sum_k\|\frac{x_k}{\|x_k\|} - x_k^\prime\| < \infty$. \label{defrankweaklynull} \end{dfn} \begin{rmk} Every weakly null sequence in $\{x_k\}_k$ has a subsequence which is of some rank. Moreover Propositions \ref{prop3.5} and \ref{cor4.7} yield that $\{x_k\}_k$ has a subsequence which is of rank 1 if and only it admits a $c_0$ spreading model and it has a subsequence that is of rank 2 or 3 if and only if it admits $\ell_1$ as a spreading model.\label{rmksmrankrelation} \end{rmk} Proposition \ref{prop5.7} yields the following. \begin{prp} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a strictly singular operator. Then for every $\{x_k\}_k$ weakly null sequence in $Y$ which is of rank 1, we have that $\{Tx_k\}_k$ is of rank 0.\label{prp1to0} \end{prp} \begin{prp} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a strictly singular operator. Then for every $\{x_k\}_k$ weakly null sequence in $Y$ which is of rank 2, we have that $\{Tx_k\}_k$ has no subsequence which is of rank 2 or of rank 3.\label{rank2to1} \end{prp} \begin{proof} Towards a contradiction pass to a subsequence of $\{x_k\}_k$ and assume that there exist $\{x_k^\prime\}_k$, $\{y_k\}_k$ normalized block sequences, with $\sum_k\|\frac{x_k}{\|x_k\|} - x_k^\prime\| < \infty, \sum_k\|\frac{Tx_k}{\|Tx_k\|} - y_k\| < \infty$, $\{x_k^\prime\}_k$ satisfies either (i) or (ii) from Definition \ref{defrank2} and $\{y_k\}_k$ is of either rank 2 or 3. By Remark \ref{rmksmrankrelation}, we may assume that both $\{x_k^\prime\}_k$, $\{y_k\}_k$ generate $\ell_1$ as a spreading model. Setting $T^\prime:[\{x_k^\prime\}_k]\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ with $T^\prime x_k^\prime = y_k$ for all $k\in\mathbb{N}$, we have that $T^\prime$ is bounded and strictly singular. Arguing in a similar manner as in the proof of Proposition \ref{prop3.5}, we may choose $\{n_j\}_j$ a strictly increasing sequence of natural numbers, $\{F_j\}_j$ a strictly increasing sequence of natural numbers and $\{c_k^j\}_{k\in F_j}, j\in\mathbb{N}$ such that $z_j = 2^{n_j}\sum_{k\in F_j}c_k^jx^\prime_k$ and $w_j = T^\prime z_j = 2^{n_j}\sum_{k\in F_j}c_k^jy_k$ are $(1,\theta,n_j)$ vectors for every $j\in\mathbb{N}$. Proposition \ref{prop3.8} yields that $\alpha\big(\{z_j\}_j\big) = 0$ as well as $\alpha\big(\{w_j\}_j\big) = 0$. If $\{x_k^\prime\}_k$ satisfies (i) from Definition \ref{defrank2}, then by Proposition \ref{prpalzeroalris} we may assume that is is $(2,\{m_k\}_k)$ $\alpha$-RIS, therefore the $z_j$ can in fact have been chosen to be $(2,\theta,n_j)$ exact vectors. Proposition \ref{cor4.7} yields that $\beta\big(\{z_j\}_j\big) = 0$. We have concluded that $\{z_j\}_j$ is of rank 1 and by Proposition \ref{prp1to0} we have that $\{w_j\}_j$ is norm null, which contradicts the fact that $\|w_j\| \geqslant \theta$. If on the other hand, if $\{x_k^\prime\}_k$ satisfies (ii) from Definition \ref{defrank2}, then we have that $\beta\big(\{z_j\}_j\big) = 0$. Again, Proposition \ref{prp1to0} yields that $\{w_j\}_j$ is norm null, which cannot be the case. \end{proof} \begin{prp} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ be a strictly singular operator. Then every $\{x_k\}_k$ weakly null sequence in $Y$, has a subsequence $\{x_{k_n}\}_n$ such that $\{Tx_{k_n}\}_n$ is of rank 0, of rank 1, or of rank 2.\label{rank3to2} \end{prp} \begin{proof} If $\{Tx_k\}_k$ has a norm null subsequence then we are ok. Otherwise, pass to a subsequence of $\{x_k\}_k$, again denoted by $\{x_k\}_k$ and choose $\{x_k^\prime\}_k$, $\{y_k\}_k$ normalized block sequences, with $\sum_k\|\frac{x_k}{\|x_k\|} - x_k^\prime\| < \infty, \sum_k\|\frac{Tx_k}{\|Tx_k\|} - y_k\| < \infty$. By passing to a further subsequence and slightly perturbing the $x_k^\prime, y_k$, we may assume that $\min\supp x_k^\prime = \min\supp y_k$ for all $k\in\mathbb{N}$ and that $\{x_k^\prime\}_k$, $\{y_k\}_k$ are of some rank. Setting $T^\prime:[\{x_k^\prime\}_k]\rightarrow \mathfrak{X}_{_{^\text{ISP}}}$ with $T^\prime x_k^\prime = y_k$ for all $k\in\mathbb{N}$, we have that $T^\prime$ is bounded and strictly singular. Towards a contradiction, assume that $\{y_k\}_k$ satisfies the assumption of Definition \ref{defrank3}. By Remark \ref{rmksmrankrelation}, passing to a further subsequence, we have that $\{y_k\}_k$ generates an $\ell_1$ spreading model and by the boundedness of $T^\prime$, we may assume that so does $\{x_k^\prime\}_k$. Passing to an even further subsequence, we have that both $\{x_k^\prime\}_k$ and $\{y_k\}_k$ satisfy the conclusion of Proposition \ref{prop3.5}. Choose $C\geqslant 1, \theta > 0, \{n_j\}_j$ strictly increasing sequence of natural numbers, $F_j \subset L$ and $\{c_k^j\}_{k\in F_j}, j\in\mathbb{N}$ such that $w_j = 2^{n_j}\sum_{k\in F_j}c_k^jy_k$ are $(C,\theta,n_j)$ vectors for every $j\in\mathbb{N}$, and $\beta\big(\{w_j\}_j\big) > 0$. Since $\min\supp x_k^\prime = \min\supp y_k$ for all $k\in\mathbb{N}$, we have that $z_j = 2^{n_j}\sum_{k\in F_j}c_k^jx_k^\prime$ are $(C,\theta,n_j)$ vectors for every $j\in\mathbb{N}$. Proposition \ref{prop3.8} yields that $\alpha\big(\{z_j\}_j\big) = 0$ as well as $\alpha\big(\{w_j\}_j\big) = 0$. Since $\beta\big(\{w_j\}_j\big) > 0$, we may pass to a subsequence of $\{w_j\}_j$ that generates an $\ell_1$ spreading model and if $w_j^\prime = \frac{w_j}{\|w_j\|}$, then $\{w_j^\prime\}_j$ satisfies (i) from Definition \ref{defrank2}, it is therefore of rank 2. Once more, the boundedness of $T^\prime$ yields that if $z_j^\prime = \frac{z_j}{\|w_j\|}$, then $\{z_j^\prime\}_j$ generates an $\ell_1$ spreading model. Since $\alpha\big(\{z_j^\prime\}_j\big) = 0$, we conclude that $\beta\big(\{z_j^\prime\}_j\big) > 0$. We may therefore pass to a final subsequence of $\{z_j^\prime\}_j$ which is of rank 2. Since $T^\prime z_j^\prime = w_j^\prime$, Proposition \ref{rank2to1} yields a contradiction. \end{proof} \subsection*{The Invariant Subspace Property} \begin{thm} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $Q,S,T: Y\rightarrow Y$ be strictly singular operators. Then $QST$ is compact.\label{cor5.9} \end{thm} \begin{proof} Since $\mathfrak{X}_{_{^\text{ISP}}}$ is reflexive, it is enough to show that for any weakly null sequence $\{x_k\}_{k\in\mathbb{N}}$, we have that $\{QSTx_k\}_{k\in\mathbb{N}}$ norm converges to zero. Proposition \ref{rank3to2}, yields that passing, if necessary to a subsequence, $\{Tx_k\}_k$ is of rank 0, rank 1, or rank 2. If it is of rank 0, then it is norm null and we are done. If it is or rank 1, Proposition \ref{prp1to0} yields that $\{STx_k\}_k$ is of rank 0 and as previously we are done. Otherwise, $\{Tx_k\}_k$ is of rank 2. By Proposition \ref{rank2to1}, we may pass to a further subsequence, such that $\{STx_k\}_k$ is either of rank 0, or rank 1. If it is not of rank 0, then applying Proposition \ref{prp1to0} we have that $\{QSTx_k\}_{k\in\mathbb{N}}$ norm converges to zero and the proof is complete. \end{proof} \begin{cor}Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $S:Y\rightarrow Y$ be a non zero strictly singular operator. Then $S$ admits a non-trivial closed hyperinvariant subspace.\label{cor5.10} \end{cor} \begin{proof} Assume first that $S^3 = 0$. Then it is straightforward to check that $\ker S$ is a non-trivial closed hyperinvariant subspace of $S$. Otherwise, if $S^3 \neq 0$, then Theorem \ref{cor5.9} yields that $S^3$ is compact and non zero. Since $S$ commutes with its cube, by Theorem 2.1 from \cite{Sir}, it is enough to check that for any $\alpha,\beta\in\mathbb{R}$ such that $\beta\neq 0$, we have that $(\alpha I - S)^2 + \beta^2I \neq 0$. Since $S$ is strictly singular, it is easy to see that this condition is satisfied. \end{proof} \begin{cor}Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$ and $T:Y\rightarrow Y$ be a non scalar operator. Then $T$ admits a non-trivial closed hyperinvariant subspace.\label{cor5.11} \end{cor} \begin{proof} Theorem \ref{prop5.6} yields that there exist $\lambda\in\mathbb{R}$, such that $S = T - \lambda I$ is strictly singular, and since $T$ is not a scalar operator, we evidently have that $S$ is not zero. By Corollary \ref{cor5.10}, it follows that $S$ admits a non-trivial closed hyperinvariant subspace $Z$. It is straightforward to check that $Z$ also is a hyperinvariant subspace for $T$. \end{proof} In the final result, which is related to Proposition 3.1 from \cite{ADT}, we show that the ``scalar plus compact'' property fails in every subspace of $\mathfrak{X}_{_{^\text{ISP}}}$. \begin{prp} Let $Y$ be an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$. Then there exists a strictly singular, non compact operator $S:Y\rightarrow Y$. In fact, if $\mathcal{S}(Y)$ is the space of strictly singular operators on $Y$, then $\mathcal{S}(Y)$ is non-separable.\label{prop5.12} \end{prp} \begin{proof} By Corollary \ref{cor4.10}, there exists a sequence $\{x_k\}_{k\in\mathbb{N}}$ in $Y$ that generates a spreading model equivalent to $c_0$, say with an upper constant $c_1$ and by Corollary \ref{cor4.12}, there exists a sequence $\{x_k^*\}_{k\in\mathbb{N}}$ in $Y^*$ that also generates a spreading model equivalent to $c_0$, say with an upper constant $c_2$. Therefore $\{x_k\}_{k\in\mathbb{N}}$ and $\{x_k^*\}_{k\in\mathbb{N}}$ are weakly null and we may assume that they are Schauder basic and that $\dim(Y/[x_k]_k) = \infty$. We may also assume that there exist $\{z_k\}_{k\in\mathbb{N}}$ in $Y$ such that $\{x_k^*\}_{k\in\mathbb{N}}$ is almost biorthogonal to $\{z_k\}_{k\in\mathbb{N}}$. For $\varepsilon>0$, set $M_\varepsilon = \frac{4c_1}{\varepsilon}$. Choose a strictly increasing sequence of naturals $\{q_j\}_{j\in\mathbb{N}}$, such that $q_j\geqslant M_{1/2^{j+1}}$. Set $S:Y\rightarrow Y$, such that $Sx = \sum_{k=1}^\infty x_{q_k}^*(x)x_k$. Then: \begin{enumerate} \item[(i)] $S$ is bounded and non compact. \item[(ii)] $S$ is strictly singular. \end{enumerate} We first prove that it is bounded. Let $x\in Y, \|x\| = 1$, $x^*\in Y^*, \|x^*\| = 1$. For $j\geqslant 0$, set $B_j = \{k\in\mathbb{N}: 1/2^{j+1}<|x^*(x_k)|\leqslant 1/2^j\}$. Since $\{x_k\}_{k\in\mathbb{N}}$ generates $c_0$ as a spreading model, it follows that $B_j\leqslant M_{1/2^{j+1}}\leqslant q_j$. Set $C_j = \{k\in B_j: k\geqslant j\}, D_j = B_j\setminus C_j$. Evidently $\#D_j\leqslant j$ and it is easy to see that $\#\{q_k: k\in C_j\}\leqslant\min\{q_k: k\in C_j\}$, therefore, since $\{x_k^*\}_{k\in\mathbb{N}}$ generates a spreading model equivalent to $c_0$, it follows that \begin{equation*} |\sum_{k\in C_j}x^*(x_k)x^*_{q_k}(x)| < c_2\max\{|x^*(x_k)|: k\in C_j\} \end{equation*} Therefore $|\sum_{k\in B_j}x^*_{q_k}(x)x^*(x_k)|\leqslant c_2\max\{|x^*(x_k)|: k\in C_j\} + j/2^j\leqslant c_2/2^j + j/2^j$. From this it follows that \begin{equation*} \|Sx\|\leqslant \sum_{j=0}^\infty \frac{j + c_2}{2^j}\|x\| \end{equation*} The fact that $S$ is non compact follows easily if you consider the almost biorthogonals $\{z_k\}_{k\in\mathbb{N}}$ of $\{x^*_{q_k}\}_{k\in\mathbb{N}}$. Then $\{z_k\}_{k\in\mathbb{N}}$ is a seminormalized sequence in $Y$ and $\{Sz_k\}_{k\in\mathbb{N}}$ does not have a norm convergent subsequence. We now prove that $S$ is strictly singular. Suppose that it is not, then there exists $\lambda\neq 0$ such that $T = S - \lambda I$ is strictly singular. Since $\lambda I$ is a Fredholm operator and $T$ is strictly singular, it follows that $S = T + \lambda I$ is also a Fredholm operator, therefore $\dim(Y/S[Y]) < \infty$. The fact that $S[Y] \subset [x_k]_k$ and $\dim(Y/[x_k]_k) = \infty$ yields a contradiction. Moreover, for any further subsequence $\{x_k^*\}_{k\in L}$ of $\{x_{q_k}^*\}_{k\in\mathbb{N}}$, if we set $S_Lx = \sum_{k\in L} x_{k}^*(x)x_k$, then $S_L$ satisfies the same conditions. This yields that $\mathcal{S}(Y)$ contains an uncountable $\varepsilon$-separated set and is therefore non-separable. \end{proof} The last proof actually yields that if $Y$ is an infinite dimensional closed subspace of $\mathfrak{X}_{_{^\text{ISP}}}$, then the space of strictly singular, non-compact operators of $Y$ is non-separable. \subsection*{Some final remarks} We would like to mention that the structure of the dual of $\mathfrak{X}_{_{^\text{ISP}}}$ is unclear to us. In particular we cannot determine whether $\mathfrak{X}_{_{^\text{ISP}}}^*$ shares similar properties with $\mathfrak{X}_{_{^\text{ISP}}}$. For example, we do not know whether $\mathfrak{X}_{_{^\text{ISP}}}^*$ admits only $c_0$ and $\ell_1$ as a spreading model. However, the following holds. \begin{prp} Let $X$ be a reflexive Banach space. Then the following are equivalent. \begin{itemize} \item[(i)] The space $X$ satisfies the hereditary ISP. \item[(ii)] Every infinite dimensional quotient of $X^*$ satisfies ISP \end{itemize} If even more, every non scalar operator defined on an infinite dimensional closed subspace of $X$ admits a non trivial closed hyperinvariant subspace, then every non scalar operator defined on an infinite dimensional quotient of $X^*$ admits a non trivial closed hyperinvariant subspace.\label{dualityisp} \end{prp} \begin{proof} In the general setting, if $X$ is a Banach space, $T$ is a bounded linear operator on $X$ admitting a non trivial closed invariant subspace $Y$, then it is straightforward to check that $Y^\perp$ is a non trivial closed invariant subspace of $T^*$. If moreover $Y$ is $T$-hyperinvariant, then $Y^\perp$ is $S^*$-invariant, for every operator $S$ on $X$, which commutes with $T$. In the setting of reflexive spaces, all operators on $X^*$ are dual operators, hence we conclude the following. Let $X$ be a reflexive Banach space and $T$ be a bounded linear operator on $X$. \begin{itemize} \item[(a)] Then $T$ admits a non trivial closed invariant subspace if and only if $T^*$ admits a non trivial closed invariant subspace. \item[(b)] Moreover, $T$ admits a non trivial closed hyperinvariant subspace if and only if $T^*$ admits a non trivial closed hyperinvariant subspace. \end{itemize} We now proceed to prove the equivalence of assertions (i) and (ii). Assume that (i) holds, let $X^*/Y$ be an infinite dimensional quotient of $X^*$ and $T$ be a bounded linear operator on $X^*/Y$. Then $X^*/Y = Y_\perp^*$ and there is $S$ a bounded linear operator on $Y_\perp$ with $T = S^*$. Since $S$ admits a non trivial closed invariant subspace, by (a) so does $T$. If we moreover assume that $S$ admits a non trivial closed hyperinvariant subspace, then by (b), so does $T$. Conversely, if (ii) holds, assume that $Y$ is an infinite dimensional subspace of $X$ and $T$ is a bounded linear operator on $Y$. Then $Y^* = X^*/Y^\perp$. By the assumption, $T^*$ admits a non trivial invariant subspace and therefore, by (a) so does $T$. \end{proof} \begin{cor} Every infinite dimensional quotient of $\mathfrak{X}_{_{^\text{ISP}}}^*$ satisfies ISP. More precisely, every non scalar operator defined on an infinite dimensional quotient of $\mathfrak{X}_{_{^\text{ISP}}}^*$ admits a non trivial closed hyperinvariant subspace. \end{cor} We would also like to mention, that the method of constructing hereditarily indecomposable Banach spaces with saturation under constraints, using Tsirelson space as an unconditional frame, can be used to yield further results. For example, in \cite{AM} a reflexive hereditarily indecomposable Banach space $\mathfrak{X}_{_{^\text{usm}}}$ is constructed having the following property. In every subspace $Y$ of $\mathfrak{X}_{_{^\text{usm}}}$ there exists a weakly null normalized sequence $\{y_n\}_n$, such that every subsymmetric sequence $\{z_n\}_n$ is isomorphically generated as a spreading model of a subsequence of $\{y_n\}_n$.
\section{Introduction} The flavour problem can be viewed as the clash between the theoretical expectation of New Physics (NP) at the TeV scale and the experimental observations in Flavour Changing Neutral Current (FCNC) processes which severely constrain the scale $\Lambda_{\rm{NP}}$ of the NP beyond the $10^4$ TeV domain (for a review see e.g.~Ref.~\cite{Isidori:2010kg}). If we insist in keeping $\Lambda_{\textrm{NP}} \approx$ TeV for naturalness, then we have to conclude that the flavour structure of the NP is highly non-generic. The Minimal Flavour Violation (MFV) hypothesis~\cite{Chivukula:1987py,Hall:1990ac,Buras:2000dm,D'Ambrosio:2002ex} is a powerful organizing principle which states that the sources of flavour symmetry breaking of the NP are aligned to the Standard Model (SM) Yukawas. This ansatz provides an automatic suppression of the NP contribution to the flavour violating observables and thus a solution of the aforementioned flavour problem (see for instance Ref.~\cite{Bona:2007vi,Hurth:2008jc}). In particular the MFV hypothesis can be formulated as a symmetry principle: in absence of the Yukawa couplings the global symmetry of the SM is that of the gauge invariant kinetic terms. This flavour symmetry or a subgroup can be formally be restored by promoting the Yukawa couplings to spurions with definite transformation properties under the flavour group and the same holds in any extension of the SM. While in the quark sector the MFV ansatz unambiguously relates the sources of flavor breaking beyond the SM to the quark masses and the CKM matrix, the situation in the lepton sector is less straightforward. This is namely due to the introduction of new sources of flavour breaking related to neutrino masses and, since the mechanism itself generating neutrino masses is unknown, several scenarios can be envisaged. Starting from Ref.~\cite{Cirigliano:2005ck} many formulations of Minimal Lepton Flavour Violation (MLFV) have been proposed and analyzed~\cite{Cirigliano:2006su,Cirigliano:2006nu,Davidson:2006bd,Branco:2006hz,Gavela:2009cd,Filipuzzi:2009xr,Alonso:2011jd}. In this work we focus on a particular realization of MFV in the in the context of the Minimal Supersymmetric SM (MSSM) without R-parity (for a review see e.g.~Ref.~\cite{Barbier:2004ez}). Our analysis is moved by two simple observations about the MSSM: \begin{enumerate} \item The largest group of unitary transformations commuting with the gauge group is enhanced with respect to the SM case. This is namely $U(3)_{\hat{q}} \otimes U(3)_{\hat{u}^c} \otimes U(3)_{\hat{d}^c} \otimes U(3)_{\hat{e}^c} \otimes U(4)_{\hat{L}} \otimes U(1)_{\hat{h}_u}$, where the presence of the $U(4)_{\hat{L}}$ factor is due to the fact that the superfields $\hat{\ell}$ and $\hat{h}_d$ have the same quantum numbers \cite{Hall:1983id,Hempfling:1995wj,Nilles:1996ij,Nardi:1996iy,Chun:1999bq,Davidson:2000ne}. \item The MSSM (without right handed neutrinos) has already all the degrees of freedom sufficient to generate neutrino masses and mixings through R-parity Violating (RPV) interactions~\cite{Hall:1983id,Lee:1984kr,Lee:1984tn,Dawson:1985vr,Banks:1995by,Nardi:1996iy,Hirsch:2000ef,Hirsch:2004he,Bajc:2010qj}. \end{enumerate} Thus the aim of our work is to generalize the MFV expansion of the soft terms in the presence of the enlarged flavor symmetry and to include the RPV couplings as the original sources of flavor breaking. Such a MFV expansion is in principle expressed in terms of many free parameters. However, motivated by our second observation, we connect the RPV spurions with the neutrino sector observables, requiring that the flavour symmetry is broken in the minimal way compatible with the pattern of neutrino masses. In particular we will study the impact of such a framework on LFV processes, thus providing an alternative scenario of MLFV. Notice that the MSSM without R-parity contains a large set of lepton number violating parameters, making the connection between neutrino masses and the soft terms still ambiguous. Thus extra assumptions on the orientation of the RPV spurions in the flavour space are needed. Our approach towards the R-parity differs from that of Refs.~\cite{Nikolidakis:2007fc,Csaki:2011ge} in the fact that we do not aim at an explanation of the smallness of the RPV couplings. We simply treat them, in a more democratic way, on the same ground of all the other couplings of the superpotential. Remarkably, the values of the RPV couplings needed in order to fit neutrino masses are of the same order of magnitude of the SM Yukawas of the first and second families~\cite{Allanach:2003eb,Ellis:1998rj}. In the following we analyze two symmetry patterns based on the groups $SU(3)^4 \otimes SU(4)$ and $SU(3)^5 \otimes U(1)_L \otimes U(1)_B$. With the former flavour symmetry the breaking scale of lepton number is related to that of lepton flavour violation (LFV), thus implying small effects in LFV physics. On the other hand the latter flavour symmetry allows to break the lepton number independently by means of an abelian spurion, so that visible effects are in principle achievable. We finally study the correlations among the flavour changing charged lepton decays $\ell_i \rightarrow \ell_j \gamma$. \section{Minimal Flavour Violation without R-parity} \label{flavgroupMSSM} The starting point of the MFV idea is based on the observation that the largest group of unitary transformations commuting with the SM gauge group is \begin{equation} \label{kineticSM} G_{\rm{kin}}^{\rm{SM}} = U(3)_{q} \otimes U(3)_{u^c} \otimes U(3)_{d^c} \otimes U(3)_{e^c} \otimes U(3)_{\ell} \otimes U(1)_{h} \, . \end{equation} This corresponds to the global symmetry of the gauge invariant kinetic term of the SM fields \begin{equation} \Phi = \left( q_i, u^c_i, d^c_i, e^c_i, \ell_i, h \right) \, , \end{equation} with $i$ spanning over the three families. Notice that $\ell_i$ and $h$ have the same gauge quantum numbers and only the Lorentz structure prevents the global symmetry of the kinetic term from being larger. On the other hand the situation in the MSSM is qualitatively different since the supersymmetrization of the SM spectrum restore the symmetry between scalars and fermions, thus enhancing the global symmetry of the kinetic term. In order to make apparent this enhancement it is useful to define a generalized lepton multiplet $\hat{L}_{\alpha} = (\hat{\ell}_i, \hat{h}_d)$ and rewrite the set of chiral superfields of the MSSM in the following way \begin{equation} \hat{\Phi}= \left( \hat{q}_i, \hat{u}^c_i, \hat{d}^c_i,\hat{e}^c_i, \hat{L}_{\alpha}, \hat{h}_u \right) \, , \end{equation} where a second Higgs doublet is introduced in order to ensure anomaly cancellation. Then the global symmetry of the kinetic term \begin{equation} \int d^4 \theta \ \hat{\Phi}^{\dagger} e^{2 g \hat{V}} \hat{\Phi} \end{equation} turns out to be \begin{equation} \label{kineticMSSM} G_{\rm{kin}}^{\rm{MSSM}} = U(3)_{\hat{q}} \otimes U(3)_{\hat{u}^c} \otimes U(3)_{\hat{d}^c} \otimes U(3)_{\hat{e}^c} \otimes U(4)_{\hat{L}} \otimes U(1)_{\hat{h}_u} \, . \end{equation} Notice that this holds irrespectively of the fact that R-parity is or not an exact symmetry of the full MSSM lagrangian. We can decompose $G_{\rm{kin}}^{\rm{MSSM}}$ in abelian and non-abelian factors and identify a linear combination of the six $U(1)$ generators with the SM hypercharge. Then we define the generalized flavour group of the MSSM as \begin{equation} G_{F}= SU(3)_{\hat{q}} \otimes SU(3)_{\hat{u}^c} \otimes SU(3)_{\hat{d}^c} \otimes SU(3)_{\hat{e}^c} \otimes SU(4)_{\hat{L}} \, , \end{equation} while the abelian factors can be rearranged in the following way \begin{equation} G_{A} = U(1)_{\hat{u}^c} \otimes U(1)_{\hat{d}^c} \otimes U(1)_{\hat{e}^c} \otimes U(1)_{\hat{L}} \otimes U(1)_{B} \, , \end{equation} where $B$ is the baryon number. $G_F$ and $G_A$ are explicitly broken by the most general MSSM superpotential and soft lagrangian. Since the MSSM has many sources of flavour violation it is useful to have a rationale in order to select the origin of this breaking. Let us imagine that the flavour symmetry is broken at the scale $\Lambda_F$ by some unknown mechanism. Then, if the breaking of SUSY is due to a flavour universal mechanism (like in gauge mediation \cite{Giudice:1998bp}) and the scale of mediation $M$ is smaller than $\Lambda_F$, the flavor structure of the soft terms will be related to the breaking of the flavor symmetry in the supersymmetric sector. Having in mind such a MFV framework we \emph{assume} that the original source of flavour violation is given by the the couplings of the most general MSSM superpotential \begin{equation} \label{susyW} W= Y_U^{ij} \hat{q}_i \hat{u}^c_j \hat{h}_u + Y_D^{\alpha ij} \hat{L}_{\alpha} \hat{q}_i \hat{d}^c_j + \tfrac{1}{2} Y_E^{\alpha \beta i} \hat{L}_{\alpha} \hat{L}_{\beta} \hat{e}^c_i + \mu^{\alpha} \hat{h}_u \hat{L}_{\alpha} + \tfrac{1}{2} (\lambda^{''} )^{ijk} \hat{u}^c_i \hat{d}^c_j \hat{d}^c_k \, , \end{equation} where the gauge structure has been omitted for simplicity. Notice also the antisymmetry of the couplings $Y_E^{\alpha \beta i} = - Y_E^{\beta \alpha i}$ and $(\lambda^{''} )^{ijk} = - (\lambda^{''} )^{ikj}$. In order to formally restore the invariance with respect to the flavor group we treat the couplings in~\eq{susyW} as spurions, with quantum numbers under $G_F$: \begin{eqnarray} \label{YUspu} Y_U &\sim& (\bar{3},\bar{3},1,1,1) \\ \label{YDspu} Y_D &\sim& (\bar{3},1,\bar{3},1,\bar{4}) \\ \label{YEspu} Y_E &\sim& (1,1,1,\bar{3},6) \\ \label{muspu} \mu &\sim& (1,1,1,1,\bar{4}) \\ \label{lambdasspu} \lambda^{''} &\sim& (1,\bar{3},3,1,1) \, , \end{eqnarray} where our conventions are such that each chiral superfield in $\hat{\Phi}$ transforms according to the fundamental representation of the corresponding group factor in $G_F$. Following the MFV principle we can expand the soft terms (cf.~\app{appnotation} for the notation) by means of the spurions in~\eqs{YUspu}{lambdasspu} \begin{equation} \label{MFVexpSU4} \begin{array}{rcl} \left( \tilde{m}^2_q \right)^{i}_j & = & \tilde{m}^2 \left( c_q \delta^{i}_j + d^{1}_{q} \ Y_U^{ik} \left(Y_U^* \right)_{jk} + d^{2}_{q} \ Y_D^{\alpha ik} \left(Y_D^* \right)_{\alpha jk} \right) \\ \left( \tilde{m}^2_{u^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{u^c} \delta^{i}_j + d^{1}_{u^c} \ Y_U^{ki} \left(Y_U^* \right)_{kj} + d^{2}_{u^c} (\lambda^{''} )^{ikl} (\lambda^{''\ast} )_{jkl} \right) \\ \left( \tilde{m}^2_{d^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{d^c} \delta^{i}_j + d^{2}_{d^c} \ Y_D^{\alpha k i} \left(Y_D^* \right)_{\alpha k j} + d^{2}_{d^c} (\lambda^{''} )^{kil} (\lambda^{''\ast} )_{kjl} \right) \\ \left( \tilde{m}^2_{e^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{e^c} \delta^{i}_j + d^{1}_{e^c} \ Y_E^{\alpha \beta i} \left(Y_E^* \right)_{\alpha \beta j} \right) \\ \left( \tilde{m}^2_{L} \right)^{\alpha}_{\beta} & = & \tilde{m}^2 \left( c_{L} \delta^{\alpha}_{\beta} + d^{1}_{L} \ Y_E^{\alpha \gamma k} \left(Y_E^* \right)_{\beta \gamma k} + d^{2}_{L} \ Y_D^{\alpha k l} \left(Y_D^* \right)_{\beta k l} + d^3_L \, \mu^{\alpha} \mu^*_{\beta} / |\mu|^2 \right)\\ B^{\alpha}&= & \tilde{m}^2 \left( c_{B} \, \mu^{\alpha} / |\mu| + d^1_B \ Y_D^{\alpha k l} (Y_D^*)_{\beta k l} \, \mu^{\beta} / |\mu| + d^2_B \ Y_E^{\alpha \beta k} (Y_E^*)_{\gamma \beta k} \, \mu^{\gamma} / |\mu| \right) \\ A^{ij}_U &=& A \left( c_{A_U} Y^{ij}_U + d^1_{A_U} Y^{kj}_U \left( Y_D^* \right)_{\alpha k l} Y^{\alpha i l}_D + d^2_{A_U} Y^{ik}_U ( \lambda^{''*} )_{k l m } ( \lambda^{''} )^{j l m} \right. \\ & & \left. + d^3_{A_U} Y^{ik}_U \left( Y_U^* \right)_{l k} Y^{l j}_U \right) \\ A^{\alpha ij}_D &=& A \left( c_{A_D} Y^{\alpha ij}_D + d^1_{A_D} Y^{\alpha kj}_D \left(Y_U^* \right)_{kl} Y_U^{il} + d^2_{A_D} Y^{\beta ij}_D \left(Y_E^* \right)_{\beta \gamma k} Y_E^{\alpha \gamma k} \right. \\ & & + d^3_{A_D} Y^{\alpha ik}_D ( \lambda^{''*} )_{l k m } ( \lambda^{''} )^{l j m} + d^4_{A_D} Y^{\alpha il}_D (Y_D^*)_{\gamma k l} (Y_D)^{\gamma k j} \\ & & + d^5_{A_D} Y^{\alpha k j}_D (Y_D^*)_{\gamma k l} (Y_D)^{\gamma i l} + d^6_{A_D} Y^{\alpha kl}_D (Y_D^*)_{\gamma k l} (Y_D)^{\gamma i j} \\ & & \left. + d^7_{A_D} Y^{\beta ij}_D \mu^*_{\beta} \mu^{\alpha} / |\mu|^2 \right)\\ A^{\alpha \beta i}_E &=& A \left( c_{A_E} Y^{\alpha \beta i}_E + d^1_{A_E} Y^{[\alpha \gamma i} _E \left(Y^*_D \right)_{\gamma k l} Y^{\beta] k l} _D +d^2_{A_E} Y^{\alpha \beta k} _E \left(Y^*_E \right)_{\gamma \delta k} Y^{\gamma \delta i} _E \right. \\ & & \left. + d^3_{A_E} Y^{[\alpha \gamma k} _E \left(Y^*_E \right)_{\gamma \delta k} Y^{\beta] \delta i} _E + d^4_{A_E} Y^{ [\alpha \gamma i}_E \mu^{*}_{\gamma} \mu^{\beta]} / |\mu|^2 \right) \\ A_{\lambda^{''}}^{ijk} &=& A \left( c_{A_{\lambda^{''}}} (\lambda^{''})^{ijk} + d^1_{A_{\lambda^{''}}} (\lambda^{''})^{ljk} (Y_U^*)_{ml} (Y_U)^{mi} \right. \\ & & + d^2_{A_{\lambda^{''}}} (\lambda^{''})^{i[jl} (Y_D^*)_{\alpha m l} (Y_D)^{\alpha m k]} \\ & & \left. + d^3_{A_{\lambda^{''}}} (\lambda^{''})^{i[jm} (\lambda^{''*})_{lnm} (\lambda^{''})^{lnk]} + d^4_{A_{\lambda^{''}}} (\lambda^{''})^{imn} (\lambda^{''*})_{lmn} (\lambda^{''})^{ljk} \right) \, , \\ \end{array} \end{equation} where the squared brackets stand for anti-symmetrization and we also defined $\abs{\mu}^2 \equiv \sum_{\alpha=1, \dots , 4} \abs{\mu^{\alpha}}^2$. Notice that expansion is truncated at the third order in the spurions. Actually the higher order terms in the spurions can be phenomenologically neglected under the assumption that the dimensionless couplings of the MFV expansion are of $\mathcal{O}(1)$~\cite{D'Ambrosio:2002ex,Colangelo:2008qp}. In absence of R-parity all the neutral scalar components of $\hat{L}_{\alpha}$ and $\hat{h}_u$ develop a VEV which triggers the electroweak symmetry breaking $SU(2)_L \otimes U(1)_Y \to U(1)_Q$ \cite{Hall:1983id,Hempfling:1995wj}. Given the $SU(4)_{\hat{L}}$ symmetry it is always possible, without loss of generality, to redefine the $\hat{L}_{\alpha}$ superfield in such a way that only the fourth component acquires a VEV. From now on we work in such a basis and we define operatively the superfield $\hat{h}_d$ so that it corresponds to the component which develops a VEV, $\hat{h}_d \equiv \hat{L}_{4}$, while the leptons do not, $\hat{\ell}_i \equiv \hat{L}_{i}$. Despite our notation makes explicit the underlying non-abelian flavour symmetry $SU(3)^4 \otimes SU(4)$, it is also useful to translate it into the more common $SU(3)^5$ language. This connection is provided in~\app{appnotation}. Then we can formally split the superpotential in~\eq{susyW} in an RPC and an RPV term \begin{align} & W_{RPC}= y_U^{ij} \hat{q}_i \hat{u}^c_j \hat{h}_u + y_D^{ij} \hat{h}_d \hat{q}_i \hat{d}^c_j + y_E^{ij} \hat{h}_d \hat{\ell}_i \hat{e}^c_j + \mu \, \hat{h}_u \hat{h}_d \, , \\ & W_{RPV}= \mu^{i} \hat{h}_u \hat{\ell}_i + \tfrac{1}{2} \lambda^{ijk} \hat{\ell}_i \hat{\ell}_j \hat{e}^c_k + (\lambda^{'} )^{ijk} \hat{\ell}_i \hat{q}_j \hat{d}^c_k + \tfrac{1}{2} (\lambda^{''} )^{ijk} \hat{u}^c_i \hat{d}^c_j \hat{d}^c_k \, , \end{align} Analogously the MFV expansion of the soft terms in~\eq{MFVexpSU4} can be easily translated in the $SU(3)^5$ language by means of the dictionary given in~\eq{dictionary} of~\app{appnotation} \begin{equation} \label{MFVexpSU3lang} \begin{array}{rcl} \left( \tilde{m}^2_q \right)^{i}_j & = & \tilde{m}^2 \left( c_q \delta^i_j + d^{1}_{q} (y_U y_U^{\dagger})^i_j + d^{2}_{q} \left[ (y_D y_D^{\dagger})^i_j + (\lambda')^{lik} \lambda'^*_{ljk} \right] \right) \\ \left( \tilde{m}^2_{u^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{u^c} \delta^i_j + d^{1}_{u^c} (y_U^{\dagger} y_U )^i_j + d^{2}_{u^c} (\lambda^{''} )^{ikl} (\lambda^{''\ast} )_{jkl} \right) \\ \left( \tilde{m}^2_{d^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{d^c} \delta^i_j + d^{1}_{d^c} \left[ (y_D^{\dagger} y_D)^i_j + (\lambda')^{lki} \lambda'^*_{lkj} \right] + d^{2}_{d^c} (\lambda^{''} )^{kil} (\lambda^{''\ast} )_{kjl} \right) \\ \left( \tilde{m}^2_{e^c} \right)^{i}_j & = & \tilde{m}^2 \left( c_{e^c} \delta^i_j + d^{1}_{e^c} \left[ 2 (y_E^{\dagger} y_E)^i_j + \lambda^{lki} \lambda^*_{lkj} \right] \right) \\ \left( \tilde{m}^2_{\ell} \right)^{i}_{j} & = & \tilde{m}^2 \left( c_{L} \delta^{i}_{j} + d^{1}_{L} \left[ ( y_{E} y_E^{\dagger})^i_j + \lambda^{ilk} \lambda^*_{jlk} \right] + d^{2}_{L} (\lambda')^{ilk} \lambda'^*_{jlk} + d^3_L \, \mu^{i} \mu^*_j / |\mu|^2 \right) \\ \left( \tilde{m}^2_{d} \right)^{i} & = & \tilde{m}^2 \left( d^{1}_{L} \lambda^{ilk} (y_E^*)_{lk} + d^{2}_{L} (\lambda')^{ilk} (y_D^{*})_{lk} + d^3_L \, \mu^{i} \mu^* / |\mu|^2 \right) \\ \tilde{m}^2_{h_d}& = & \tilde{m}^2 \left(c_L +d^1_L \ \textrm{Tr}(y_E y_E^{\dagger}) + d^2_L \ \textrm{Tr}(y_D y_D^{\dagger}) + d^3_L \, \mu \, \mu^* / |\mu|^2 \right) \\ b &= & \tilde{m}^2 \left( c_{B} \mu / |\mu| + d^1_B \left[ \textrm{Tr}(y_D y_D^{\dagger}) \, \mu / |\mu| + (y_D)^{lk} \lambda'^*_{plk} \, \mu^{p} / |\mu| \right] \right. \\ & & \left. + d^2_B \left[ \textrm{Tr}(y_E y_E^{\dagger}) \, \mu / |\mu| + (y_E)^{lk} \lambda^*_{plk} \, \mu^{p} / |\mu| \right] \right) \\ b^i &=& \tilde{m}^2 \left( c_{B} \mu^i / |\mu|+ d^1_B \left[ (\lambda')^{ilk} (y^*_D)_{lk} \, \mu / |\mu| + (\lambda')^{ilk} \lambda'^*_{plk} \, \mu^p / |\mu| \right] \right. \\ & & \left. + d^2_B \left[ (\lambda)^{ilk} (y^*_E)_{lk} \, \mu / |\mu| + (\lambda)^{ilk} \lambda^*_{plk} \, \mu^p / |\mu| + (y_E y_E^{\dagger})^i_p \, \mu^p / |\mu| \right] \right)\\ a^{ij}_U &=& A \left( c_{A_U} y^{ij}_U + \dots \right) \\ a^{ij}_D &=& A \left( c_{A_D} y^{ij}_D + \dots \right) \\ a^{ij}_E &=& A \left( c_{A_E} y^{ij}_E + \ldots \right) \\ (a_{\lambda})^{ijk} &=& A \left( c_{A_E} \lambda^{ijk} + \dots \right) \\ (a_{\lambda^{'} })^{ijk}&=& A \left( c_{A_D} (\lambda')^{ ijk} + \dots \right) \\ a_{\lambda^{''}}^{ijk} &=& A \left( c_{A_{\lambda^{''}}} (\lambda^{''})^{ijk} + \dots \right) \, , \end{array} \end{equation} where we have omitted to expand the $a$ terms up at the third order in the spurions. Notice that the flavor violation in the lepton sector can be linked to the amount of R-parity violation. For instance the RPV couplings in the expansion of $\tilde{m}^2_{\ell}$ in~\eq{MFVexpSU3lang} are responsible for flavour violating mass insertions leading to processes like $\ell_i \rightarrow \ell_j \gamma$. Assuming that the RPV couplings are responsible for neutrino masses it is possible then to provide a link between LFV in the charged lepton sector and the neutrino observables. However, in absence of R-parity the MSSM superpotential is enriched by a large set of lepton number violating parameters, thus making the connection between neutrino masses and the soft terms somehow ambiguous. In order to retain some level of predictivity for the soft terms extra assumptions must be made on the orientation of the RPV spurions in the flavour space\footnote{Notice that this problem is present in many formulations of MLFV. See for instance Refs.~\cite{Cirigliano:2005ck,Davidson:2006bd,Alonso:2011jd}}. In the next section we will exemplify this point by considering a particular model of neutrino masses in which the RPV spurions are oriented along the $\mu^i$ and $\lambda'^{i33}$ directions. \section{Neutrino Masses without R-parity} In the previous section we have formally restored the invariance under the flavour group $G_F$ by promoting all the supersymmetric couplings in~\eq{susyW} to spurions. Now we want to provide the link between these spurions and the physical observables. Our guideline is to break the flavour group in a \emph{minimal} way, namely we consider the minimal amount of flavour breaking which is able to reproduce the correct pattern of fermion masses and mixings. In the limit of massless neutrinos, the connection of the spurions with the flavour structure of the charged fermions is straightforward. From the superpotential \begin{equation} W \supset Y_U^{ij} \, \hat{q}_i \hat{u}^c_j \hat{h}_u + Y_D^{4ij} \, \hat{h}_d \hat{q}_i \hat{d}^c_j + Y_E^{4ij} \, \hat{h}_d \hat{\ell}_{i} \hat{e}^c_j \, , \end{equation} we can identify the relevant spurions in terms of known physical observables, up to the parameter $\tan\beta \equiv v_u / v_d$. Indeed it is always possible to choose a basis such that \begin{equation} \label{RPCmassmatrices} Y_U^{ij} = (V^\dag \hat{m}_U)^{ij} / v_u \, , \qquad Y_D^{4ij} = \hat{m}_D^{ij} / v_d \, , \qquad Y_E^{4ij} = \hat{m}_E^{ij} / v_d \, , \end{equation} where $V$ is the CKM matrix and $\hat{m}_U$, $\hat{m}_D$, $\hat{m}_E$ are the diagonal charged-fermion masses. On the other hand the experimental evidence of neutrino masses and mixings makes clear that the flavour group must be further broken. The standard way to introduce neutrino masses in the context of supersymmetric MFV is to extend the field content of the MSSM by introducing three SM-singlet chiral superfields~\cite{Nikolidakis:2007fc,Csaki:2011ge} and thus applying the seesaw mechanism~\cite{Minkowski:1977sc,GellMann:1980vs,Yanagida:1979as,Glashow:1979nm,Mohapatra:1979ia,Schechter:1980gr,Schechter:1981cv}. Remarkably the MSSM without R-parity gives the possibility of generating neutrino masses and mixings without the need of additional ingredients. This is the approach we pursue in this work. As we are going to show soon, neutrino masses are fitted by moderate small values of the R-parity violating couplings $\mu^i/\mu$, $\lambda$ and $\lambda'$, of $\mathcal{O}(10^{-5})$ or even larger. From this point of view the issue of the smallness of neutrino masses could be brought back at the same conceptual level of understanding the flavour structure of the charged fermions, featuring Yukawa couplings also of $\mathcal{O}(10^{-6})$ as in the case of the electron. The formulae for the neutrino mass matrix in terms of the RPV couplings are collected for completeness in~\app{RPVandNuMasses}. The leading contributions can be schematically written as \begin{equation} \label{numassschem} m_\nu \quad \sim \quad \left(\frac{M_Z}{\tilde{m}}\right)^2 \frac{\mu_i \mu_j}{\tilde{m}} \, , \qquad \frac{3 \, \lambda'^2}{8 \pi^2} \frac{\hat{m}_D^2}{\tilde{m}} \, , \qquad \frac{\lambda^2}{8 \pi^2} \frac{\hat{m}_E^2}{\tilde{m}} \, , \end{equation} where for simplicity we set $M_1 \approx M_2 \approx A \approx \mu \equiv \tilde{m}$ and we neglected the flavour structure of $\lambda'$ and $\lambda$. Finally we comment about the baryon number violating coupling $\lambda''$. According to our guideline at the beginning of this section, this coupling does not give any contribution to the construction of fermion masses and mixings and thus should be absent as an irreducible source of flavor breaking. However $\lambda''$ can still be induced by the other spurions. If the $U(1)$ factors are part of the symmetry that we want to formally restore, then $\lambda''$ cannot be generated by the baryon number conserving couplings $Y_U$, $Y_E$, $Y_D$ and $\mu$. On the other hand if we consider only the non-abelian symmetry $SU(3)^4 \otimes SU(4)$, the coupling $\lambda''$ can be induced in a holomorphic way~\cite{Csaki:2011ge}: \begin{equation} \lambda'' \sim Y_U (Y_D)^2 (Y_E)^3 \, , \end{equation} where the proper contractions with the $SU(3)_{\hat{q}}$, $SU(3)_{\hat{e}^c}$ and $SU(4)_{\hat{L}}$ epsilon tensors are understood. Actually, it turns out that the tensor structure forces the invariant to span over RPV couplings and light generation Yukawas, thus providing an automatic suppression of $\lambda''$. Remarkably we are able to satisfy the bounds from proton decay without invoking any \emph{ad hoc} conservation or small breaking of the $U(1)_B$ symmetry, but just requiring our minimality condition on the identification of the flavor spurions. \subsection{A toy model} \label{toy} When all the RPV spurions are switched on there is an overabundance of free parameters, which cannot all be fixed by the constraints from the neutrino sector. Hence we are going to consider scenarios in which only a minimal number of spurions are switched on in order to reproduce neutrino masses and mixings. Our goal is to show how to link the RPV spurions with the neutrino observables, by means of a simplified model of neutrino masses featuring only the RPV couplings $\mu^i$ and $\lambda'^{i33}$. Notice that we work in the basis $\vev{\tilde{\nu}_i} = 0$. We have checked explicitly that, within these set of RPV couplings, this condition can be consistently achieved with the MFV expansion in~\eq{MFVexpSU3lang}. The neutrino mass matrix can be split in a tree level and a one-loop term \begin{equation} m_\nu = m_\nu^{\rm{(tree)}} + m_\nu^{\rm{(loop)}} \, , \end{equation} whose diagonalization through the PMNS matrix $\hat{U}$ yields \begin{equation} m_\nu = \hat{U} \hat{m}_{\nu} \hat{U}^T \, , \end{equation} where $\hat{m}_\nu$ is the diagonal neutrino mass matrix. Then, assuming the dominance of the tree level contribution and requiring an orthogonality condition between the vectors $\mu^i$ and $\lambda'^{i33}$ (cf.~\app{RPVandNuMasses} for more details), the neutrino sector observables are fitted in the following analytical way: \begin{itemize} \item Tree level contribution \begin{equation} \label{eq:neutreet} (m_{\nu}^{(\rm{tree})})^{ij} = m_3 \, \hat{U}^{i3} \hat{U}^{j3} \, , \end{equation} where $m_3 \approx \sqrt{\Delta m^2_{\textrm{atm}}} = 4.9 \cdot 10^{-2} \ \textrm{eV}$. Taking $M_1 = M_2 \equiv \tilde{m} \gg M_Z$ in~\eq{eq:mefftree} of~\app{RPVandNuMasses}, we get \begin{equation} \label{eq:neutree} \frac{\mu^{i}}{\mu} = 2.4 \cdot 10^{-5} \left( \frac{ \tilde{m}}{1 \textrm{ TeV}} \right)^{1/2} \left( \frac{ \tan \beta}{10} \right) \hat{U}^{i3} \, . \end{equation} \item One-loop contribution \begin{equation} \label{eq:neuloopt} (m_{\nu}^{(\rm{loop})})^{ij} = m_2 \, \hat{U}^{i2} \hat{U}^{j2} \, , \end{equation} where $m_2 \approx \sqrt{\Delta m^2_{\textrm{sol}}} = 8.7 \cdot 10^{-3} \ \textrm{eV}$. Taking $\mu = \tilde{m}$ and at the leading order in the MFV expansion with $c_{A_D} = c_{d^c} = c_q = 1$ (cf.~\eq{mlpMFV} in~\app{RPVandNuMasses}), we get \begin{equation} \label{eq:neuloop} (\lambda')^{i33} = 3.6 \cdot 10^{-5} \left( \frac{ \tilde{m}}{1 \ \textrm{TeV}} \right)^{1/2} \left( \frac{ \tan \beta}{10} \right)^{-1/2} \hat{U}^{i2} \, . \end{equation} \end{itemize} \section{Analysis of LFV processes} \subsection{Case $SU(3)^4 \otimes SU(4)$} Once the relevant spurions are fixed in terms of the neutrino masses and mixings one can relate the MFV expansion to LFV processes. For our purposes it is enough to make an order of magnitude estimate of the processes induced in our MLFV setup, focusing just on the effects due to the non-diagonal entries in the sfermion mass matrices. In this case the normalized branching fractions for the processes $\ell_i \rightarrow \ell_j \gamma$ can be naively estimated in the following way~\cite{Masina:2002mv,Paradisi:2005fk}: \begin{equation} \label{eq:sketchBR} B_{\ell_i \rightarrow \ell_j\gamma}\equiv\frac{BR(\ell_i \rightarrow \ell_j \, \gamma)}{BR(\ell_i \rightarrow \ell_j \nu_i \bar{\nu}_j)} \approx \frac{\alpha^3} {G_F^2}\frac{\delta_{ij}^2}{\tilde{m}^4} \tan^2\beta \, , \end{equation} where, according to the MFV expansion, the flavour violating mass insertions $\delta_{ij}$ are expressed as combinations of neutrino masses and elements of the PMNS matrix. For instance in our toy model where only the couplings $\mu^i$ and $\lambda'^{i33}$ are switched on, $\delta_{ij}^{LL}$ reads (cf.~\eq{MFVexpSU3lang}) \begin{equation} \label{expansdll} \delta_{ij}^{LL}=\frac{\Delta_{ij}^{LL}}{\tilde{m}_\ell^2} \approx \frac{1}{c_L}\left( d^{2}_{L} (\lambda')^{i33} \lambda'^*_{j33} + d^3_L \, \frac{\mu^{i} \mu^*_j}{|\mu|^2} \right) \, , \end{equation} where $\Delta_{ij}^{LL}$ is the flavour violating part of $\tilde{m}_\ell^2$. As it is evident from~\eq{expansdll}, the mass insertions scale like the square of the RPV parameters. Given the following estimation of the branching fractions in~\eq{eq:sketchBR} \begin{equation} B_{\ell_i \rightarrow \ell_j\gamma} \approx 10^{-27} \ {\left(\frac{\tilde{m}}{1 \mbox{TeV}}\right)}^{-4} {\left(\frac{\tan\beta}{10}\right)}^{2} \left( \frac{\lambda' }{10^{-5}} \right)^4 \, , \end{equation} one concludes that it is not possible to accomplish observable rates, in view of the current experimental bounds showed in~\Table{tabsummarylfv}. \begin{table}[htbp] \begin{center} \begin{tabular}{|l|l|} \hline LFV process & Bound \\ \hline $BR(\mu \rightarrow e\,\gamma)$ & $2.4 \times 10^{-12}$ \cite{Uchiyama:2011zz}\\ \hline $BR(\tau \rightarrow e\,\gamma)$ & 3.3 $\times 10^{-8}$ \cite{Aubert:2009tk} \\ \hline $BR(\tau \rightarrow \mu\,\gamma)$ & 4.5 $\times 10^{-8}$ \cite{Hayasaka:2007vc}\\ \hline \end{tabular} \end{center} \caption{ Current experimental bounds on $\ell_i \rightarrow \ell_j \gamma$ processes. } \label{tabsummarylfv} \end{table} It should be also mentioned that besides the contributions related to the MFV mass insertions there are other ones due to the mixing between neutrinos and neutralinos (or charginos and charged leptons) and RPV vertices (see e.g.~\cite{Carvalho:2002bq,deCarlos:1996du}) which are relevant for the calculation of the LFV branching fractions. These latter contributions, however, remain of the same order of magnitude of those due to the mass insertions, thus leading to non-observable rates for the values of the RPV couplings fitting neutrino masses. Let us mention that rates of $\mu \rightarrow e \, \gamma$ closer to the experimental sensitivity can be obtained when neutrino masses are fitted by trilinears featuring first families indices, like for instance $\lambda'^{i11}$. In such a case the suppression due to the down-quark mass in the expression of the neutrino mass matrix (cf.~\eq{mlpMFV} in~\app{RPVandNuMasses}) allows for larger values of $\lambda'^{i11}$ even of $\mathcal{O}(10^{-2})$. However such a large coupling may be in conflict with other flavour violating observables \cite{Barbier:2004ez}. A complete analysis of such scenarios and a more realistic model for neutrino masses is postponed to future works. In the next subsection we are going to present a simple solution in order to achieve observable rates within the neutrino mass model considered here. \subsection{Case $SU(3)^5 \otimes U(1)_{L} \otimes U(1)_{B}$} \label{SU35LB} We have previously seen that the contribution to LFV processes are generically well below the present experimental bounds. This is due to fact that the spurions responsible for neutrino masses break simultaneously both the total lepton number and the non-abelian part of the flavor group. As it has been shown in~\cite{Cirigliano:2005ck,Gavela:2009cd}, in order to have measurable rates for the flavor changing radiative charged lepton decays, one has to separate the source of breaking of lepton number from that of LFV. This leads us to consider a different scenario based on another subgroup of the original kinetic symmetry $G^{\rm{MSSM}}_{\rm{kin}}$ (cf.~\eq{kineticMSSM}). We assume that the symmetry that we want to formally restore is given by $SU(3)^5 \otimes U(1)_{L} \otimes U(1)_B$, where $L$ and $B$ are the total lepton and baryon number. The $U(1)_B$ factor is needed in order to properly suppress dangerous contributions to the proton decay rate (for the relevant bounds see for instance Ref.~\cite{Smirnov:1996bg,Bhattacharyya:1998dt}). In this setup the R-parity violating couplings $\mu_i$, $\lambda$, $\lambda'$ and $\lambda''$ can be split in two parts, one responsible for the breaking of lepton and baryon number and the other for the breaking of the flavor group \begin{equation} \mu^i=\varepsilon_{L} \tilde{\mu}^i \, , \qquad \lambda=\varepsilon_{L} \tilde{\lambda} \, , \qquad \lambda'=\varepsilon_{L} \tilde{\lambda}^{'} \, , \qquad \lambda''=\varepsilon_{B} \tilde{\lambda}^{''} \, . \end{equation} The quantum numbers of the flavor spurions under $SU(3)^5 \otimes U(1)_{L} \otimes U(1)_{B}$ are given by \begin{eqnarray} y_U & \sim & (\bar{3},\bar{3},1,1,1)_{(0,0)} \\ y_D & \sim & (\bar{3},1,\bar{3},1,1)_{(0,0)} \\ y_E & \sim & (1,1,1,\bar{3},\bar{3})_{(0,0)} \\ \tilde{\mu}^{i} & \sim & (1,1,1,1,\bar{3})_{(0,0)} \\ \tilde{\lambda} & \sim & (1,1,1,\bar{3},3)_{(0,0)} \\ \tilde{\lambda}^{'} & \sim & (\bar{3},1,\bar{3},1,\bar{3})_{(0,0)} \\ \tilde{\lambda}^{''} & \sim & (1,\bar{3},3,1,1)_{(0,0)} \\ \varepsilon_L & \sim & (1,1,1,1,1)_{(-1,0)} \\ \varepsilon_B & \sim & (1,1,1,1,1)_{(0,+1)} \end{eqnarray} where the subscripts label the abelian quantum numbers. The corresponding MFV expansion of the soft terms is reported in~\eq{MFVexpSU3BL} of~\app{gtSU35}. In this case the rates of the LFV processes are dominated by the lepton and baryon number preserving (but flavor changing) slepton mass insertions, which depend only on the parameters $\tilde{\mu}$, $\tilde{\lambda}$, $\tilde{\lambda}^{'}$. Other RPV vertex contributions depend on quantities which violate total lepton number and hence are suppressed by the $\varepsilon_{L}$ factor. As we are going to show, peculiar correlations among physical observables will emerge due the MFV expansion. In the case that only the spurions $\mu^i$ and $\lambda'^{i33}$ are switched on, the relevant off-diagonal terms $i \neq j$ induced by these two spurions in $\left( \tilde{m}^2_{\ell} \right)^{i}_{j}$ and $a^{ij}_E$ are given by (cf.~\eq{MFVexpSU3BL} in~\app{groupth}) \begin{equation} \left( \tilde{m}^2_{\ell} \right)^{i}_{j} = \tilde{m}^2 \left( d^{2}_{\ell} (\tilde{\lambda}')^{i33} \tilde{\lambda}'^*_{j33} + d^3_{\ell} \, \frac{\tilde{\mu}^{i} \tilde{\mu}^*_j}{\abs{\mu}^2} \right) \, , \quad a^{ij}_E = A \, y_E^{jj} \left( d^4_{a_E} \frac{\tilde{\mu}^*_j \tilde{\mu}^i}{\abs{\mu}^2} + d^5_{a_E} \tilde{\lambda}'^*_{j33} (\tilde{\lambda}')^{i33} \right) \, . \end{equation} In our framework the LL mass insertions, and thus $\left( \tilde{m}^2_{\ell} \right)^{i}_{j}$, give the dominant contribution to the LFV processes\footnote{The term $a_E$ is responsible for the LR mass insertions. However, according to the analysis of Refs.~\cite{Masina:2002mv,Paradisi:2005fk}, $\delta^{LR}$ is negligible provided that $\delta_{ij}^{LR} \ll \left(m_{\ell_i} / \tilde{m} \right) \tan\beta \ \delta_{ij}^{LL}$. In our case, assuming all the coefficients of the MFV expansion to be of order one, this condition translates into $\abs{ \left( \tilde{m} / A \right) \tan\beta } \gg 1$. }. Focusing on $\delta^{LL}$, using (\ref{eq:neutreet}) and (\ref{eq:neuloopt}), we get: \begin{multline} \label{eq:susyinsertion} \left( \delta^{LL} \right)^{i}_{j} = \frac{1}{ c_{\ell}} \left[ d^3_\ell \frac{\tilde{\mu}^i \tilde{\mu}^{*}_j}{\abs{\mu}^2}+d^2_\ell (\tilde{\lambda}')^{i33} \tilde{\lambda}'^*_{j33} \right] \\ = \frac{1}{\varepsilon_{L}^2 c_{\ell}} \left[ d^3_\ell {\left(\frac{ \tan\beta}{M_Z}\right)}^2 \left(\frac{M_1 M_2}{M_1 c_W^2 +M_2 s_W^2} - \frac{M_Z^2}{\mu} \sin 2 \beta \right) \sqrt{\Delta m^2_{\textrm{atm}}} \, \hat{U}^{i3} (\hat{U}^{j3})^* \right. \\ \left. +d^2_\ell \frac{8 \pi^2 \tilde{m}^2}{3 \mu \tan\beta \, m_b^2} \, \sqrt{\Delta m^2_{\textrm{sol}}} \, \hat{U}^{i2} (\hat{U}^{j2})^* \right] \,. \end{multline} Notice that the factor $1/\varepsilon_L^2$ in the mass insertions implies an enhancement of $1/\varepsilon_L^4$ in the rates. Indeed it is possible to estimate the normalized branching ratios in the following way \begin{equation} B_{\ell_i \rightarrow \ell_j\gamma} \approx 10^{-27} \left( \frac{1}{\varepsilon_L} \right)^4 {\left(\frac{\tilde{m}}{1 \ \mbox{TeV}}\right)}^{-4} {\left(\frac{\tan\beta}{10}\right)}^{2} {\left(\frac{\lambda'}{10^{-5}}\right)}^4 \, , \end{equation} for values of $\varepsilon_L \sim 10^{- (3 \div 4)}$ the rates of the three relevant processes can get close to the experimental sensitivities, depending on the values of SUSY parameters. Notice that the coupling $\varepsilon_L$ cannot be arbitrarily small. Indeed, by imposing the relations in~\eqs{eq:neutree}{eq:neuloop} and by requiring that the flavor violating parameters $\tilde{\mu}$ and $\tilde{\lambda}'$ are at most of order one, we can estimate the lower bound $\varepsilon_L \gtrsim 10^{-5}$. Given the potential detectability of these processes it is now interesting to compute the ratios among the normalized branching fractions of the LFV channels. Considering~\eq{eq:sketchBR} and~\eq{eq:susyinsertion} we can parametrize the ratios among the normalized LFV branching fractions as \begin{equation} \label{ratiobf} \frac{B_{\ell_j \rightarrow \ell_i\gamma}}{B_{\ell_k \rightarrow \ell_m \gamma}}=\frac{|\hat{U}^{i2} (\hat{U}^{j2})^* +c \, \hat{U}^{i3} (\hat{U}^{j3})^*|^2}{|\hat{U}^{m2} (\hat{U}^{k2})^* +c \, \hat{U}^{m3} (\hat{U}^{k3})^*|^2} \, , \end{equation} where the constant $c$ is given by \begin{equation} \label{eq:cest} c \approx 1.4 \times 10^{-1} \left(\frac{d_\ell^3}{d_\ell^2}\right) {\left(\frac{\tan\beta}{10}\right)}^{3} {\left(\frac{\mu}{1 \ \mbox{TeV}}\right)}{\left(\frac{\tilde{m}}{1 \ \mbox{TeV}}\right)}^{-2}{\left(\frac{M_G}{300 \ \mbox{GeV}}\right)} \, , \end{equation} with $M_1 = M_2 \equiv M_G$. From~\eq{eq:cest} it is evident that, depending on the SUSY parameters, the mass insertions are dominated either by the trilinear ($c \ll 1$) or the bilinear ($c \gg 1$) couplings. It is possible then to identify two asymptotic regimes in which these ratios have a simple analytical expression: \begin{itemize} \item $\abs{c} \ll 1$ In this case the mass insertions are dominated by the contribution from the trilinear couplings $(\tilde{\lambda}')^{i33}$. Since the ratios $B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow \mu \gamma}$ and $B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow e \gamma}$ show only a slight dependence from the Maiorana phase $\delta$, we take $\delta=0,\pi$ and obtain \begin{equation} \frac{B_{\mu \rightarrow e\gamma}}{B_{\tau \rightarrow \mu \gamma}}= \frac{|\hat{U}^{12}|^2}{|\hat{U}^{32}|^2} = \frac{s_{12}^2 c_{13}^2}{(\mp c_{23} s_{12} s_{13} - c_{12}s_{23})^2} \approx 0.53 \div 1.75 \, , \end{equation} \begin{equation} \frac{B_{\mu \rightarrow e\gamma}}{B_{\tau \rightarrow e \gamma}} = \frac{|\hat{U}^{22}|^2}{|\hat{U}^{32}|^2} = \frac{(c_{12}c_{23} \mp s_{12} s_{13} s_{23})^2}{(\mp c_{23} s_{12} s_{13} - c_{12}s_{23})^2} \approx 0.37 \div 2.4 \, , \end{equation} where the extrema of the range are obtained by scanning over the 2-$\sigma$ values of the mixing angles (cf.~\Table{tab:summaryneu}). In this case, the three normalized branching ratios are of the same order of magnitude. Notice that the LFV effects depend on the PMNS matrix $\hat{U}^{i2}$, differently with respect to other MLFV setups (cf.~for instance~\Table{tab:summary}). \item $\abs{c} \gg 1$ In this case the mass insertions are dominated by the bilinear couplings $\mu^i$. Then we can derive the following functional behaviors for the two relevant ratios \begin{equation} \frac{B_{\mu \rightarrow e\gamma}}{B_{\tau \rightarrow \mu \gamma}}= \frac{|\hat{U}^{13}|^2}{|\hat{U}^{33}|^2}= \frac{s_{13}^2}{c_{13}^2c_{23}^2} \approx 0.007 \div 0.07 \, , \end{equation} \begin{equation} \frac{B_{\mu \rightarrow e\gamma}}{B_{\tau \rightarrow e \gamma}} = \frac{|\hat{U}^{23}|^2}{|\hat{U}^{33}|^2} =\frac{s^2_{23}}{c_{23}^2} \approx 0.7 \div 1.6 \, . \end{equation} Compared to the previous case we observe an enhancement of $B_{\tau \rightarrow \mu\gamma}$ with respect to $B_{\tau \rightarrow e\gamma}$ and $B_{\mu \rightarrow e\gamma}$ . This results coincides with the one found in Ref.~\cite{Alonso:2011jd} in the case of inverted hierarchy of neutrino masses. \end{itemize} \begin{table}[htbp] \begin{center} \begin{tabular}{|c|c|c|} \hline Observable & Best fit & 2-$\sigma$ \\ \hline $\Delta m_{\rm atm}^2$ & $2.50 \times 10^{-3} \ {\mbox{eV}}^2$ & $(2.25-2.68) \times 10^{-3} \ {\mbox{eV}}^2$ \\ \hline $\Delta m_{\rm sol}^2$ & $7.59 \times 10^{-5} \ {\mbox{eV}}^2$ & $(7.24-7.99) \times 10^{-5} \ {\mbox{eV}}^2$ \\ \hline $\sin^2 \theta_{12}$ & $0.312$ & $0.28-0.35$ \\ \hline $\sin^2 \theta_{23}$ & $0.52$ & $0.41-0.61$ \\ \hline $\sin^2 \theta_{13}$ & $0.013$ & $0.004-0.028$ \\ \hline \end{tabular} \end{center} \caption{Experimental values of the neutrino sector observables as reported in Ref.~\cite{Schwetz:2011zk}. For the PMNS matrix we have considered the PDG parametrization \cite{Nakamura:2010zzi}. The Dirac phase $\delta$ varies in the range $[0,2\pi]$. } \label{tab:summaryneu} \end{table} Furthermore, in order to study the general case we vary the parameter $c$ in the range $[-100,100]$ and the parameters of the neutrino sector according to~\Table{tab:summaryneu} as before. The results are plotted in~\fig{fig:ben2s13}. In~\fig{fig:ben2delta} we report the correlation between the two ratios $B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow \mu\gamma}$ and $B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow e\gamma}$. \begin{figure}[htbp] \centering \subfloat{\includegraphics[width=8 cm, height= 6 cm, angle=360]{ratio1cc.eps}} \subfloat{\includegraphics[width=8 cm, height= 6 cm, angle=360]{ratio2cc.eps}} \caption{Ratios between normalized branching fractions of as function of $|c|$.} \label{fig:ben2s13} \end{figure} \begin{figure}[htbp] \centering \includegraphics[width=10cm,height=8cm,angle=360]{brcheck.eps} \caption{$B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow \mu\gamma}$ versus $B_{\mu \rightarrow e\gamma}/B_{\tau \rightarrow e\gamma}$. The parameter $c$ is varied in the range $[-100, 100]$. The blue region is characterized by $\abs{c} = 0$ while the green one by $\abs{c} \gg 1$.} \label{fig:ben2delta} \end{figure} From~\fig{fig:ben2s13} we see that, away from the two asymptotic regimes, there are regions of strong enhancement/suppression of the ratios. Indeed the values of the $c$ parameter for which $B_{\ell_i \rightarrow \ell_j\gamma} \rightarrow 0$ are: \begin{align} & B_{\mu \rightarrow e\gamma} \to 0 \quad \longrightarrow \quad c =\mp \frac{c_{12}s_{23}s_{12}}{s_{23}s_{13}} \approx \mp (2.2 \div 9.4) \qquad (\delta=0,\pi) \, , \\ & B_{\tau \rightarrow e\gamma} \rightarrow 0 \quad \longrightarrow \quad c = \pm \frac{c_{12}s_{23}s_{12}}{s_{23}s_{13}} \approx \pm (2.2 \div 9.4) \qquad (\delta=0,\pi) \, , \\ & B_{\tau \rightarrow \mu\gamma} \rightarrow 0 \quad \longrightarrow \quad c = \frac{c_{12}^2}{c_{13}^2} \approx (0.66 \div 0.72) \, . \end{align} Finally, for the purpose of comparison, we show in~\Table{tab:summary} the LFV parameters predicted by various MLFV models. \begin{table}[htbp] \begin{center} \begin{tabular}{|c|c|} \hline Model & Flavor violating parameter \\ \hline Minimal field content \cite{Cirigliano:2005ck} & $\propto (\hat{U} \hat{m}_{\nu}^2\hat{U}^{\dagger})^{ij}$ \\ \hline Extended field content + CP limit \cite{Cirigliano:2005ck}& $\propto (\hat{U}\hat{m}_{\nu}\hat{U}^{\dagger})^{ij}$ \\ \hline Extended field content + leptogenesis \cite{Cirigliano:2006nu,Branco:2006hz} & $\propto (\hat{U}\hat{m}_{\nu}^{1/2}\,H^2\,\hat{m}_{\nu}^{1/2}\hat{U}^{\dagger})^{ij}$ \\ \hline $SU(3)_{\ell} \otimes SU(3)_N \to SU(3)_{\ell+N} $ \cite{Alonso:2011jd} & $\propto (\hat{U}\frac{1}{\hat{m}^2_{\nu}}\hat{U}^{\dagger})^{ij}$ \\ \hline MSSM wihtout R-parity (toy model) & $\propto( \hat{U}^{i2}\hat{U}^{*j2}+c\,\hat{U}^{i3}\hat{U}^{*j3}) $\\ \hline \end{tabular} \end{center} \caption{Comparative summary of MLFV models.} \label{tab:summary} \end{table} We conclude commenting on other LFV processes like $\mu \rightarrow e$ conversions and $\ell_i \rightarrow \ell_j \ell_k \ell_k$ decays, not considered until now. In our case these processes are determined by $\gamma$-penguin type diagrams~\cite{Arganda:2005ji,Brignole:2004ah} and turn out to have the same flavor structure of $\ell_i \rightarrow \ell_j \gamma$. This implies in particular that the decays in three leptons have similar patterns of enhancement/suppression of those discussed above. Notice however that the radiative decays are the processes most severely constrained by the experiments. \section{Conclusions} In this work we have presented the general supersymmetric version of MFV including also the RPV terms as the irreducible sources of the flavour symmetry breaking. If the RPV couplings are responsible for neutrino masses, the framework can be also viewed as an extension of MFV to the lepton sector. An important aspect stressed throughout the paper is that the global symmetry of the kinetic term of the MSSM lagrangian is enhanced with respect to that of the SM. Indeed the superfields $\hat{\ell}_i$ and $\hat{h}_d$ can be rearranged in a 4-dimensional flavour multiplet $\hat{L}$, whose kinetic term is invariant under $U(4)_{\hat{L}}$ unitary transformations. This gives us the possibility to consider as the most general flavour symmetry the non-abelian group $SU(3)^4 \otimes SU(4)_{\hat{L}}$. In such a case the breaking of the total lepton number and that of lepton flavour number are linked together, thus generically implying small effects in LFV physics. On the other hand the separation between the breaking of lepton number and lepton flavour number leads to an interesting phenomenology. This is the motivation to consider our second scenario based on the $SU(3)^5 \otimes U(1)_L \otimes U(1)_B$ flavour symmetry. This last option yields peculiar correlations among the branching ratios of the $\ell_i \rightarrow \ell_j \gamma$ processes. Several interesting possibilities could be considered for future investigations both from a theoretical and a phenomenological point of view. \section*{Acknoledgments} We thank Federica Bazzocchi, Ben Grinstein, Jernej Kamenik and Federico Mescia for useful discussions and Gino Isidori for encouragement. We also thank Andrea Romanino and Marco Serone for further useful comments. This work was supported in part by the National Science Foundation under Grant No.~1066293 and the hospitality of the Aspen Center for Physics. The work of LDL was supported by the DFG through the SFB/TR 9 ``Computational Particle Physics''; he would also like to thank the CP3-Origins group at the University of Southern Denmark for the warm hospitality and the partial support during the final stages of this work.
\section{Introduction} \label{sec:1} The thermal freeze-out of weakly interacting particles (so-called 'WIMPs') with masses at the electroweak symmetry breaking scale $E_{\rm ew}= 1/\sqrt{2^{1/2}G_{\rm F}}\approx 246$~GeV leads to a relic density agreeing with the observed one $\Omega_{\rm dm}h^2=0.1123\pm0.0035$ \cite{Lee1977,Jarosik2011}. This so-called WIMP miracle provides a natural solution to the dark matter problem. In the present-day Universe, WIMPs can still annihilate in regions with large mass concentrations. The Galactic Center \cite{Aharonian2006,Albert2006}, dwarf galaxies in the halo of the Milky Way \cite{Aliu2009,Albert2008,Aleksi'c2011,Aharonian2008,Acciari2010}, and galaxy clusters \cite{Aleksi'c2010,Ackermann2010} have been studied to obtain limits on dark matter annihilation emission. Annihilation gives rise to pairs of heavy quarks, leptons or vector bosons, which decay leading to the emission of gamma rays, electrons, positrons, and neutrinos. The secondary radiation shows a peak at $E_{\rm ew}/20\approx 10$~GeV reflecting the high multiplicity of secondary particles in a typical annihilation event. The canonical peak energy lies at the high end of the bandwidth of the Fermi-LAT detector, and at the low end of the accessible energy range for ground-based Cherenkov telescopes. A very well motivated WIMP candidate is the Lightest Supersymmetric Particle (LSP): the neutralino \cite{Bertone2005}. Annihilating neutralinos produce vector bosons, leptons or quarks. The differential gamma-ray flux \cite{Evans2004} produced through their subsequent hadronization and decay from a source at distance $D$ and volume $V$ is given by: \begin{equation} \left(\frac{\mathrm{d}\Phi}{\mathrm{d}E}\right)_{\pi^{0}}=\frac{1}{4\pi}\frac{f\left\langle\sigma_{A}\,\nu\right\rangle}{2m_{\chi}^{2}}\frac{\mathrm{d}N_{\gamma}}{\mathrm{d}E}\frac{1}{{D}^{2}}\int\limits_{\rm{M87}} \! \mathrm{d}V\,\rho_{\rm NFW}^2 \end{equation} where $f$ is the boost factor that accounts for enhancement due to sub-halo clumping of dark matter in the M87 halo, $\left\langle \sigma_{A}\nu\right\rangle$ is the thermally averaged annihilation cross section, $m_{\chi}$ is the mass of the dark matter particle, and $\mathrm{d}N_{\gamma}/\mathrm{d}E$ is the gamma photon spectrum coming from the decay of neutral pions from hadronization in the annihilation process (prompt pion emission) \cite{Saxena2011}. $\rho$ is taken to be $\rho_{\rm NFW}=\rho_{\rm NFW}(r)$ the Navarro-Frenk-White dark matter density profile \cite{Navarro1996} of the dark matter halo obtained from numerical simulations. A minimum mass scale of $10^{-5}M_\odot$ for the sub-halo clum\-ping has been inferred from the transfer function of density perturbations in the early Universe for weakly interacting dark matter particles \cite{Green2004}. However, tidal interactions with the baryon-dominated cores of dark matter halos and supernova feedback could destroy these structures to a large extent. Therefore, the boost factor $f=\int \rho^2 \rm{d}V/\int\rho_{\rm NFW}^2\rm{d}V$ is introduced as a free parameter to account for the unknown enhancement due to sub-halo clumping. The inverse-Compton interaction is the up-scattering of photons by high-energy charged particles. The differential gamma-ray flux given by Eq. (2) describes inverse-Compton scattering off the cosmic microwave background by relativistic electrons and positrons. $b\left(E'\right)$ is the total rate of electron/positron energy loss due to inverse-Compton scattering as in \cite{Cirelli2009}, $P\left(E,E'\right)$ is the differential power emitted into photons of energy $E$ by an electron/positron with energy $E'$, and $\mathrm{d}N_{e}/\mathrm{d}\widetilde{E}$ is the spectrum of secondary electrons and positrons \cite{Saxena2011}: \begin{eqnarray} \left(\frac{\mathrm{d}\Phi}{\mathrm{d}E}\right)_{\rm{IC}}&=& \frac{1}{E}\frac{f\left\langle\sigma_{A}\,\nu\right\rangle}{4\pi m_{\chi}^{2}}\frac{1}{{D}^{2}}\int\limits_{\rm{M87}} \! \mathrm{d}V\,\rho_{\rm NFW}^2\left(r\right)\nonumber \\ &&\times\int\limits_{m_{e}}^{m_{\chi}}\mathrm{d}E'\,\frac{P\left(E,E'\right)}{b\left(E'\right)}\int\limits_{E'}^{m_{\chi}}\mathrm{d}\widetilde{E}\,\frac{\mathrm{d}N_{e}}{\mathrm{d}\widetilde{E}} \end{eqnarray} The main challenge in constraining the putative dark matter annihilation component is to discriminate it against gam\-ma rays from astrophysical sources and cosmic ray interactions. Here, we show as an exemplary case study how recent multi-frequency data of the center of the Virgo cluster, harboring the giant cD galaxy M87 with its gamma-ray emitting radio jet, can be used to constrain dark matter particles. Note that we do not refer to the inverse-Compton scattering in so-called leptophilic dark matter annihilation scenarios, but include the inverse-Compton emission component due to the electrons and positrons from the decay of isospin-symmetric annihilation products. We adopt a distance of $(16.5\pm1)$~Mpc \cite{NED} to M87 and therefore omit redshift corrections to the energy throughout the paper. Section 2 describes the data sets chosen for the study. The radiation code employed to model the spectral energy distribution of M87 is explained in Section 3, and physical parameters inferred from the fit of the data are briefly discussed to show that the model is a viable interpretation of the nonthermal particle content in the jet of M87. The generic model for the weakly interacting massive particles and their radiative signatures is described in Section 4, and a fit of a component from dark matter annihilation to the excess above the astrophysical model is presented. Finally, we discuss the results and draw conclusions regarding future search strategies. \section{Data} \label{sec:2} The dataset used here is comprised of observations of M87 by the Chandra X-ray Observatory, Fermi-LAT, MAGIC, and the H.E.S.S. system of Cherenkov telescopes \cite{Harris2009,Abdo2009,Berger2011,Aharonian2006a}. As presented in \cite{Abdo2009}, additional observational data exist for the radio-to-optical regime. However, we do not use these to constrain the fit, but rather only require the projected emission to not exceed these observations, since in this wavelength regime the unavoidable contamination due to starlight and dust from the central region of M87 is very difficult to assess. There may also be hidden nonthermal components in this energy region unrelated to the emission region where the high-energy emission originates from. The data sample was not taken contemporaneously, but we carefully checked to avoid inclusion of an observation containing a significant flare. All the data used here can thus be considered a representative long-term average, reasonably well describing the steady and low-state spectral energy distribution of the high-energy emission component in M87. \section{Synchrotron-self-Compton emission} \label{sec:3} The nonthermal emission from the relativistic jet emerging from the nucleus of M87 follows a spectral energy distribution across almost 20 orders of magnitude which can be described by the synchrotron and self-Compton radiation produced by electrons (and positrons) accelerated in the jet, presumably by shock waves. Whereas the high amplitude flares track single shock waves or magnetic reconnection events in the expanding flow, guided by the helical structure of the magnetic field, the steady-state (long-term average) emission which dominates the energy output from the jet of M87 corresponds to the superposition of a large number of shock waves \cite{Blandford1979}, a strong stationary shock such as the reconfinement shock \cite{Marscher2008} or the emission from a sheath surrounding the spine of the jet \cite{Tavecchio2004}. To obtain a fit for the broad-band spectral energy distribution of M87, we used an implementation of the SSC model \cite{Ruger2010} in which the cooling break in the electron spectrum is self-consistently determined, and the cross section for Compton scattering in the Klein-Nishina regime is accurately treated. We also took into account the inhomogeneous nature of the emission zone, by considering only the emission from the inner jet for the fit, i.e. we considered the radio to optical emission, related to the emission from larger scales in the jet, as an upper bound for the model. Adopting for the Doppler factor $\delta=3.9$ (bulk Lorentz factor $\Gamma=2.3$) as in the model fitted to the data by \cite{Abdo2009} we obtain a fit with $\chi^2/\nu=2.5$. Physical parameters of the emission zone are the source radius $R_b=3.5\times 10^{13}$ cm and the magnetic field strength $B=3$ G. For the injected electron power law distribution with exponential cut-off, we obtain a differential slope of $s=2.2$, maximum Lorentz factor $\Gamma_{max}=10^8$ and normalization factor $K=10^{6}\,\rm{cm^{-3}s^{-1}}$ (cf. \cite{Ruger2010}). This results in an injection luminosity of $L_{inj}=3\times10^{41}$~erg~s$^{-1}$ consistent with the energetics of the jet inferred from its large scale radio structure \cite{Hardee1982}. Model fits including the low-energy continuum \cite{Abdo2009} yield different results, but fail to produce an acceptable fit including the very high energy observations. For the black hole mass of M87 $M_{\rm BH}= 6.4\times10^9M_{\odot}$, the Eddington luminosity is given by $L_{\rm E}= 8.32\times 10^{47} $ erg~s$^{-1}$. Thus, the nonthermal power release amounts to only $\sim 10^{-6}L_{\rm E}$. Although a sub-Eddington state of the black hole is generally expected for high-peaked blazars, the extremely low power is peculiar. In fact, it is not possible to model the data by simply adjusting the Doppler factor of an SSC fit obtained for high-peaked blazars, i.e. by changing the inclination of the jet axis with respect to the observer. The idea of treating M87 as a misaligned blazar \cite{Mucke2001} does not seem to be sufficient. Furthermore the large difference between break energy and cut-off is not typical for blazar spectra. This might be related to the extremely low accretion rate, indicating that the AGN is fading out due to a lack of accretable matter, or to the fact that the observed emission is dominated by the emission from a low-bulk-Lorentz-factor sheath surrounding the jet. The SSC fit of the data obtained in this way provides an accurate and practically unique model for the SED, and the small size of the emission region is in line with observations of short-time variability. \begin{figure} \includegraphics[width=0.45\textwidth]{Fig1.pdf} \caption{Result of the combined fit including the spectral energy distributions due to the SSC mechanism in the jet of M87 and the annihilation of dark matter particles. The boost factor $f$ accounts for sub-halo clumping, and $m$ denotes the generic WIMP mass.} \label{fig:1} \end{figure} \section{Fitting the excess emission component with dark matter annihilation} \label{sec:4} \begin{figure*} \includegraphics[width=0.95\textwidth]{Fig2.pdf} \caption{Spectral energy distribution of the best-fit combined SSC model for M87 and dark matter model, including all data points used in this analysis \cite{Harris2009,Abdo2009,Berger2011,Aharonian2006a}. The inset legend provides the mass and boost factor for the generic WIMP with cross section $\left\langle \sigma_{A}\nu\right\rangle=3\times10^{-24}\,\rm{cm}^3\rm{s}^{-1}$.} \label{fig:2} \end{figure*} To now fit an additional component due to dark matter annihilation, and study whether the statistical agreement between model and observations can thereby be further improved, we assume a generic species of annihilating WIMPs with rest mass in the GeV-TeV range and a thermally averaged annihilation cross section of $\left\langle \sigma_{A}\nu\right\rangle=3\times10^{-24}\,\rm{cm}^3\rm{s}^{-1}$. It has been shown that many models with multi-TeV masses provide large pair annihilation cross sections which are still in agreement with the thermal freeze-out of these particles in the early Universe \cite{Profumo2005}. The choice of the annihilation cross section also reflects the possible existence of boost factors in the particle physics sector and should be considered as moderate upper limit (cf. \cite{Pinzke2011}). For the dark matter distribution of M87, we use the analysis of \cite{McLaughlin1999}, resulting in a description of the halo according to the Navarro-Frenk-White \cite{Navarro1996} model. The normalization of the intensity of the resulting emission is furthermore fixed by choosing the boost factor from unresolved substructure in the halo of M87, which from recent numerical experiments \cite{Springel2008} is expected to be of order $f=10^{2}-10^{3}$. The spectra for the emission due to the decay of charged and neutral pions from hadronization in the annihilation process are generated using the DarkSUSY code \cite{Gondolo2004}. A total of $10^6$ realizations of neutralinos are produced to obtain the average secondary spectra for generic WIMPs motivated by supersymmetric theories. Here we also include a treatment of the inevitable contribution~of inverse-Compton emission from energetic electrons/positrons from the decay of the charged pions up-scat\-ter\-ing cosmic microwave background photons. By using the known number density of these 2.7 K background photons ($413$~cm$^{-3}$), the respective contribution from inverse-Comp\-ton emission can be of the same order as the prompt pion decay emission. This results in a telltale ``double hump structure'' of the dark matter related emission. Employing a $\chi^{2}$-test, we search for the best-fit model of the active galactic nucleus (AGN) as discussed in the previous section, and the complete dark matter related model. The results of the fit are shown in Fig. 1. A minimal $\chi^{2}$ value of 1.6 is achieved for a particle mass of 4.7 TeV and a boost factor of 812, both well within the range of values accessible in state-of-the-art particle physics and numerical models, as discussed in the previous section. Assuming that the main multi-wavelength features of such a generic dark matter model can be described by two additional parameters (neutralino mass and boost factor) compared to the SSC model alone, it can be shown by nonlinear regression analysis that the deviation of the data points with respect to the SSC model amounts to $2\sigma$. We note that this result is not significant enough to claim the realization of a certain dark matter model, but even our simplified assumptions show the capability of multi-wavelength methods for unvealing the characteristics of the constituents of dark matter. In Fig. 2, we show the resulting spectral energy distribution from the combined SSC-DMA model. \section{Discussion and conclusions} \label{sec:5} We show that while synchrotron-self-Compton emission can fit the observations reasonably well, the introduction of a component due to annihilating dark matter particles margi\-nal\-ly improves the agreement between theoretical modeling and multi-frequency observations of the long-term average SED of the radio galaxy M87. It is essential to include the inverse-Compton emission component in the dark matter annihilation model. The method thus shows great sensitivity in ruling out dark matter models at the TeV scale which is innate to some supersymmetric extensions of the Standard Model, while annihilation cross section and additional boost from substructures are concordant with the paradigm of thermal freeze-out of such a particle at the onset of hierarchical structure formation. Although the significance of this result is not sufficient to claim evidence of a specific particle, it is encouraging further studies. Since the SSC component generally shows variability while the DMA component remains steady, we emphasize that the significance of the marginal excess reported here would increase during low states of the SSC emission. Therefore we encourage to extend the temporal coverage of multi-wavelength observations of M87. Measuring the putative \linebreak dark matter ``double hump structure`` in the spectrum with a high significance would open up the possibility to extract the dark matter properties (WIMP mass, annihilation cross section and boosting due to substructure) more accurately. The high mass scale of the generic WIMP favored in our model here is in-line with recent findings at the LHC \cite{Baer2009}. A high mass scale may also help in explaining the measured extragalactic gamma-ray background (EGB) \cite{Elsasser2005}, after a careful reassessment of the Fermi-measured EGB \cite{Abdo2010}. Detection of high-mass WIMPs by elastic scattering experiments \cite{CDMSIICollaboration2010} will be difficult due to the suppression of the event rate by the low number density of WIMPs in the Galactic halo. \bibliographystyle{springer} \balance
\section*{Introduction} If $A$ is a Banach algebra one can view it as an abstract ring and consider its algebraic K-theory or one can take the topology into account and then consider its topological K-theory. There is a natural map from the former to the latter and so one can form the homotopy fibre of this map giving the \emph{relative} K-theory. Karoubi's relative Chern character \cite{KarCR, Kar87, CK} is a homomorphism \[ K_{i}^{\mathrm{rel}}(A) \to HC_{i-1}(A) \] mapping relative K-theory to continuous cyclic homology. It was an idea of Karoubi \cite{KaroubiConnexions, KarCR} that the relative Chern character could be used for the construction of regulators. In accordance with this idea Hamida \cite{HamidaBorel} established, for $A=\mathbf C$ the field of complex numbers, a precise relation between the relative Chern character and the Borel regulator \cite{Borel1} \[ K_{2n-1}(\mathbf C) \to \mathbf C. \] Karoubi's construction works equally well in the case of ultrametric Banach algebras and building on previous work by Hamida \cite{HamidaCR} we proved the $p$-adic analogue of the above result in \cite{TammeBorel}, giving the precise relation of the relative Chern character with the $p$-adic Borel regulator introduced by Huber and Kings \cite{HK}. Changing perspective, let $X$ be a smooth variety over $\mathbf C$. Again we have the algebraic K-theory of $X$ but we can also consider the topological K-theory of the complex manifold associated to $X$. It is natural to ask for a generalization of the previous results to this situation. Here the analogue of the cyclic homology groups are the quotients of the de Rham cohomology by the Hodge filtration and Borel's regulator is replaced by Beilinson's regulator mapping algebraic K-theory to Deligne-Beilinson cohomology. In this setup we proved a comparison between the relative Chern character and Beilinson's regulator in \cite{TammeBeil}. It is the goal of the present paper to prove the $p$-adic analogue of this result. Let $R$ be a complete discrete valuation ring with field of fractions $K$ of characteristic $0$ with perfect residue field $k$ of characteristic $p$ and consider a smooth $R$-scheme $X$. The $p$-adic analogue of Beilinson's regulator is the rigid syntomic regulator, i.e.~the Chern character with values in rigid syntomic cohomology, introduced by Gros \cite{Gros} and developed systematically by Besser \cite{Besser}. We introduce topological and hence relative K-theory of $X$, and relative cohomology groups $H_{\mathrm{rel}}^{*}(X,n)$ mapping naturally to the rigid syntomic cohomology groups. These are the target of the relative Chern character in the $p$-adic situation. Our main result is \begin{thm* Let $X$ be a smooth $R$-scheme and $i >0$. The diagram \[ \xymatrix{ K_{i}^{\mathrm{rel}}(X) \ar[d]_{\ch_{n,i}^{\mathrm{rel}}} \ar[r] & K_{i}(X) \ar[d]^{\ch_{n,i}^{\mathrm{syn}}}\\ H^{2n-i}_{\mathrm{rel}}(X, n) \ar[r] & H^{2n-i}_{\mathrm{syn}}(X,n) } \] commutes. \end{thm*} If $X$ is proper the lower horizontal map is in fact an isomorphism and both groups are given by $H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n}H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)$ which, in turn, is naturally isomorphic to the weight $n-1$ part $HC_{i-1}^{(n-1)}(X_K)$ in the $\lambda$-decomposition of the cyclic homology of $X_K/K$ \cite{WeibelHodgeCyclic}. Moreover, for projective $X$ and finite $k$, Parshin's conjecture would imply that also the upper horizontal map is rationally an isomorphism. One of the possible advantages of this approach to the syntomic regulator is that Karoubi's constructions give quite explicit formulas. For instance, in the case $X=\Spec(R)$, these have been used in the comparison of Karoubi's regulator with the $p$-adic Borel regulator \cite{TammeBorel} and in computer calculations by Choo and Snaith \cite{ChooSnaith}. Another motivation to study the relative Chern character and its relation to the syntomic regulator goes back to an idea of Besser. In contrast to the Beilinson regulator or Soul\'e's \'etale $p$-adic regulator, the rigid syntomic regulator explicitly depends on the choice of the local model $X/R$ of the generic fibre $X_{K}/K$. In computations (e.g.~\cite{BesserK1surface,Besser-deJeu}) this leads to integrality assumptions one would like to remove. Besser proposed the use of Karoubi's relative Chern character in order to obtain a model independent replacement for the syntomic regulator. In fact, one can define topological and hence relative K-theory of $K$-schemes using the associated rigid space, and the techniques of this paper give a relative Chern character $K_{i}^{\mathrm{rel}}(X_{K}) \to H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n}$. If $X$ is a smooth, proper $R$-scheme there is a natural map $K_{*}^{\mathrm{rel}}(X) \to K_{*}^{\mathrm{rel}}(X_{K})$ and the relative Chern character for $X$ factors through the relative Chern character of $X_{K}$. In particular, if one assumes Parshin's conjecture, the relative Chern character would give a description of the syntomic regulator for $X$ smooth, projective over $R$ with finite residue field solely in terms of the generic fibre $X_{K}$. In general, a good understanding of topological K-theory is still missing. From the Theorem we also get the following corollary (see \ref{apr1201}): \begin{cor*} Assume that $k$ is finite and $X/R$ is smooth and projective. Then \[ \xymatrix@C+0.5cm{ K_{i}^{\mathrm{rel}}(X) \ar[d]^{\ch_{n,i}^{\mathrm{rel}}} \ar[r] & K_{i}(X) \ar[d]^{r_{p}}\\ H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n} \ar[r]^-{\exp} & H^{1}(K, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))) } \] commutes. \end{cor*} Here $r_{p}$ is the \'etale $p$-adic regulator and $\exp$ is the Bloch-Kato exponential map of the $p$-adic $\mathrm{Gal}(\ol K/K)$-representation $H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))$. This corollary may be seen as a generalization of the main result of Huber-Kings \cite[Theorem 1.3.2]{HK} which is the case $X=\Spec(R)$ and amounts to the commutativity of \[ \xymatrix{ K_{2n-1}(R) \ar[r]^-{r_{p}} \ar[dr]_-{p\text{-adic Borel}} & H^{1}(K, \mathbf{Q}_{p}(n))\\ & K=D_{\mathrm{dR}}(\mathbf{Q}_{p}(n)) \ar[u]_{\exp} } \] to all smooth, projective $R$-schemes (see our Corollary \ref{cor:HK}). \smallskip A result related to ours is proven by Chiarellotto, Ciccioni, and Mazzari in \cite{Mazzari}. They provide an alternative construction of the rigid syntomic regulator in terms of higher Chow groups and syntomic cycle classes. A key step in their construction is the compatibility of the de Rham and rigid cycle classes under the specialization map from de Rham cohomology of the generic fibre to rigid cohomology of the special fibre, whereas our result in some sense rests on the compatibility of topological and rigid Chern classes. For an interpretation of the relative cohomology groups, introduced here, in stable homotopy theory of schemes we refer the reader to \cite{DegliseMazzari}. \smallskip Let us describe the contents of this paper in more detail. Karoubi's original construction of the relative Chern character uses integration of certain $p$-adic differential forms over standard simplices. A reformulation of this construction is given in Section \ref{sec:Karoubi}. The key ingredient that enables us to compare the relative Chern character and the rigid syntomic regulator is a new description of the former in Section \ref{sec:Construction} that is similar to the construction of Chern class maps on higher K-groups by Beilinson \cite{Beilinson}, Huber \cite{Huber}, and Besser \cite{Besser}. This is made possible by the functorial complexes of Besser \cite{Besser} and Chiarellotto-Ciccioni-Mazzari \cite{Mazzari}. Their construction is recalled in Section \ref{sec:Functorial} with some simplifications coming from the systematic use of Gro{\ss}e-Kl\"onne's dagger spaces \cite{GK, GKRigid}. The necessary background from rigid resp. ``dagger'' geometry is collected in Section \ref{sec:Rigid}. In Section \ref{sec:K} we recall Karoubi and Villamayor's definition of topological K-theory for ultrametric Banach rings \cite{KV, Calvo} and extend it to smooth, separated $R$-schemes. The necessary comparison between the two constructions of the relative Chern character is proven in Theorem \ref{thm:KaroubisRelChern}. The main comparison theorem (Theorem \ref{thm:Comparison}) then follows rather formally. Applications are given in Subsection \ref{ssec:applications}. \subsection*{Acknowledgments} The results presented here emerged from my thesis \cite{TammeDiss}. It is a pleasure to thank my advisor Guido Kings for his interest and constant support. I would like to thank the California Institute of Technology, where most of this work was done, and especially Matthias Flach for their hospitality and the Deutsche Forschungsgemeinschaft for financial support. Furthermore, I would like to thank Nicola Mazzari for interesting discussions about this work. \subsection{Notation}\label{sec:notation} For $p \in \mathbf{N}_{0}=\{0,1,2,\dots\}$ we denote by $[p]$ the finite set $\{0,\dots, p\}$ with its natural order. The category of finite ordered sets with monotone maps is the \emph{simplicial category} $\Delta$. The unique injective map $[p-1] \to [p]$ that does not hit $i$ is denoted by $\partial^{i}$. Similarly, $s^{i}\colon [p+1] \to [p]$ is the unique surjective map such that $s^{i}(i)=s^{i}(i+1)$. These induce morphisms $\partial_{i}, s_{i}$ (resp. $\partial^{i}, s^{i}$) on every (co)simplicial object called (co)face and (co)degeneracy morphisms, respectively. For a group object $G$ we define the simplicial objects $E_{\dot}G$ and $B_{\dot}G$ by $E_{p}G=G^{\times (p+1)}$ and $B_{p}G = G^{\times p}$ with the usual faces and degeneracies (see e.g.~\cite[0.2]{HK}). \section{Preliminaries on rigid geometry} \label{sec:Rigid} In the definition of rigid cohomology one usually works with rigid analytic spaces and their de Rham cohomology. However, as de Rham cohomology is not well behaved for rigid spaces one has to introduce some overconvergence condition. An elegant approach to do this is to replace rigid spaces by Gro{\ss}e-Kl\"onne's dagger spaces \cite{GK, GKRigid}. We recall some basic definitions and facts which will be needed in the rest of the paper. Let $R$ be a complete discrete valuation ring with field of fractions $K$ of characteristic 0 and residue field $k$ of characteristic $p>0$. Fix an absolute value $|\,.\,|$ on $K$. \begin{Nr}[Rings]\label{Nr:DaggerRings} The $K$-algebra of \emph{overconvergent} power series in $n$ variables $\ul x = (x_{1}, \dots, x_{n})$ is \( K\<\ul x\>^{\dag} := \left\{\sum a_{\nu}\ul x^{\nu} \big|a_{\nu}\in K, \exists\rho>1:|a_{\nu}|\rho^{|\nu|}\xrightarrow{|\nu|\to\infty} 0\right\}. \) We denote by $R\<\ul x\>^{\dag}$ the sub-$R$-algebra of overconvergent power series with $R$-coefficients. A $K$- (resp. $R$-)\emph{dagger algebra} is a quotient of some $K\<\ul x\>^{\dag}$ (resp. $R\<\ul x\>^{\dag}$). The algebra of overconvergent power series carries the Gau{\ss} norm $\left|\sum_{\nu} a_{\nu}\ul x^{\nu}\right| =\sup_{\nu}|a_{\nu}|$. Its completion with respect to this norm is the Tate algebra of convergent power series $K\<\ul x\> =\{\sum_{\nu}a_{\nu}x^{\nu} \big| |a_{\nu}| \xrightarrow{|\nu|\to\infty} 0\}$. Similarly, the completion of $R\<x\>^{\dag}$ is $R\<x\>$. Quotients of these are called $K$- (resp. $R$-)\emph{affinoid}. These are Banach algebras. Up to equivalence the quotient norm on a dagger or affinoid algebra does not depend on the chosen representation as a quotient of an algebra of overconvergent, respectively, convergent power series. If $A=R[\ul x]/I$ is an $R$-algebra of finite type its \emph{weak completion} is the $R$-dagger algebra $A^{\dag}:= R\<\ul x\>^{\dag}/IR\<\ul x\>^{\dag}$. The categories of $R$- resp. $K$-dagger and affinoid algebras admit tensor products. E.g., if $A=R\<\ul x\>^{\dag}/I$ and $B=R\<\ul y\>^{\dag}/J$ are $R$-dagger algebras their tensor product is given by $A \otimes_{R}^{\dag} B = R\<\ul x, \ul y\>^{\dag}/(I+J)$. \end{Nr} \begin{Nr}[Spaces]\label{Nr:DaggerSpaces} We only sketch the main points here, referring the reader to \cite{GK, GKRigid} for details. Similarly as one defines rigid analytic spaces that are locally isomorphic to max-spectra of $K$-affinoid algebras with a certain Grothendieck topology, \emph{dagger spaces} are defined by taking the max-spectra $\Sp(A)$ of $K$-dagger algebras $A$ as building blocks \cite[2.12]{GKRigid}. For any dagger space $X$ one has an associated rigid space $X^{\mathrm{rig}}$ (``completion of the structure sheaf'') and a natural map of G-ringed spaces $X^{\mathrm{rig}} \xrightarrow{u} X$ which is an isomorphism on the underlying G-topological spaces \cite[2.19]{GKRigid}. There exists a \emph{dagger analytification functor} $X\mapsto X^{\dag}$ from the category of $K$-schemes of finite type to the category of $K$-dagger spaces. There is a natural morphism of locally G-ringed spaces $X^{\dag} \xrightarrow{\iota} X$ which is final for morphisms from dagger spaces to $X$ \cite[3.3]{GKRigid}. We also need the notion of \emph{weak formal schemes} (\cite[Ch. 3]{GK} and originally \cite{Mer}). Let $A$ be an $R$-dagger algebra and $\ol A = A\otimes_{R} k$. Then $D(\ol f) \mapsto A\<\frac{1}{f}\>^{\dag}$ where $f\in A$ lifts $\ol f\in \ol A$ defines a sheaf of local rings on the topological space $\Spec(\ol A)$. The corresponding locally ringed space is the \emph{weak formal $R$-scheme} $\Spwf(A)$. A general weak formal $R$-scheme is a locally ringed space that is locally isomorphic to some $\Spwf(A)$. Sending $\Spwf(A)$ for an $R$-dagger algebra $A$ to $\Sp(A\otimes_{R} K)$ induces the \emph{generic fibre} functor $\mathscr X \mapsto \mathscr X_{K}$ from weak formal $R$-schemes to $K$-dagger spaces and there is a natural \emph{specialization map} $\spec\colon \mathscr X_{K} \to \mathscr X$. Taking the weak completion of finitely generated $R$-algebras induces the functor $X \mapsto \widehat X$ from $R$-schemes of finite type to weak formal $R$-schemes. There is a natural morphism of dagger spaces from the generic fibre of the weak completion of $X$, $\widehat X_{K}$, to the dagger analytification $X_{K}^{\dag}$ of the generic fibre $X_{K}$ of $X$. This is an open immersion if $X$ is separated and an isomorphism if $X$ is proper over $R$ (cf. \cite[Proposition 0.3.5]{BerCohomRig}). For example, if $X=\mathbb A^{1}_{R}$ then $\widehat X_{K}=\Sp(K\<x\>^{\dag})$ is the closed ball of radius 1 in $(\mathbb A^{1}_{K})^{\dag}$. \end{Nr} \section{Preliminaries on K-theory} \label{sec:K} \begin{Nr} Let $A$ be an ultrametric Banach ring with norm $|\,.\,|$, e.g. an affinoid algebra with a fixed norm. In \cite{KV} Karoubi and Villamayor introduce K-groups of $A$ that we will denote by $K_{\top}^{-i}(A), i\geq 1$, which were further studied by Calvo \cite{Calvo}. A convenient way to define them is the following: Set \[ A\<x_{0}, \dots, x_{n}\> := \left\{\sum a_{\nu}\ul x^{\nu} \big|a_{\nu}\in A, |a_{\nu}|\xrightarrow{|\nu|\to\infty} 0\right\} \] and $A\<\Delta^n\> := A\<x_{0}, \dots, x_{n}\>/(\sum_{i} x_{i} -1)$. Then $[n]\mapsto A\<\Delta^n\>$ becomes a simplicial ring in a natural way and hence $B_{\dot}\GL(A\<\Delta^{\dot}\>)$ is a bisimplicial set (cf. \ref{sec:notation}). For any bisimplicial set $S_{\dot\dot}$, we denote by $\pi_*(S_{\dot\dot})$ the homotopy groups of the underlying diagonal simplicial set. We define \[ K^{-i}_{\top}(A) := \pi_{i}(B_{\dot}\GL(A\<\Delta^{\dot}\>)), \quad i\geq 1. \] That this definition coincides with the original one in \cite{KV, Calvo} follows from an argument of Anderson \cite[Prop. 7.3]{TammeDiss}. For our purposes it is important to know that one can compute the K-theory of affinoid algebras using dagger algebras. More precisely, let $R$ be a complete discrete valuation ring with field of fractions $K$ of characteristic 0 and residue field $k$ of characteristic $p>0$ as before. We define the simplicial ring $R\<\Delta^{\dot}\>^{\dag}$ by $R\<\Delta^{n}\>^{\dag} = R\<x_{0}, \dots, x_{n}\>^{\dag}/(\sum_{i} x_{i} -1)$ with the obvious structure maps. For any $R$-dagger algebra $A$ we set $A\<\Delta^{\dot}\>^{\dag}=A\otimes_{R}^{\dag} R\<\Delta^{\dot}\>^{\dag}$ (cf. \ref{Nr:DaggerRings}) and define topological K-groups by \[ K^{-i}_{\top}(A) := \pi_{i}(B_{\dot}\GL(A\<\Delta^{\dot}\>^{\dag})), \quad i\geq 1. \] Using Calvo's techniques we have shown in \cite[Prop. 7.5]{TammeDiss} that these agree with the Karoubi-Villamayor K-groups of the completion $\widehat A$ of $A$: \[ K^{-i}_{\top}(A) \cong K^{-i}_{\top}(\widehat A), \quad i\geq 1. \] \end{Nr} Now let $X = \Spec(A)$ be an affine scheme of finite type over $R$. \begin{dfn} We define the topological K-groups of $X$ to be the K-groups of the (weak) completion of $A$: \[ K^{-i}_{\top}(X) := K^{-i}_{\top}(A^{\dag}) = K^{-i}_{\top}(\widehat A), \quad i\geq 1. \] \end{dfn} \begin{rems}\label{nov0201} (i) Note the similarity with topological complex K-theory: If $X$ is a smooth separated scheme of finite type over $\mathbf{C}$, and $A_{\dot}$ denotes the simplicial ring of smooth functions $X(\mathbf{C}) \times \Delta^{\dot} \to \mathbf{C}$, then $\pi_{i}(B_{\dot}\GL(A_{\dot})) \cong K^{-i}_{\top}(X(\mathbf{C}))$ is the connective complex K-theory of the manifold $X(\mathbf{C})$. \noindent (ii) Let $\pi\in R$ be a uniformizer of $R$ and $B$ any $R$-dagger or affinoid algebra. Then $(\pi) \subseteq B$ is topologically nilpotent. Calvo \cite{Calvo} proved that the reduction $B \to B/(\pi)$ induces an isomorphism $K^{-i}_{\top}(B) \xrightarrow{\cong} K^{-i}_{\top}(B/(\pi))$. This last group is the Karoubi-Villamayor K-theory of $B/(\pi)$. In particular, if $B/(\pi)$ is regular, this coincides with the Quillen K-theory: $K^{-i}_{\top}(B) \xrightarrow{\cong} K_{i}(B/(\pi))$ \cite[3.14]{Gersten}. \noindent (iii) If, in the situation of the definition, $\pi$ is invertible on $X = \Spec(A)$, i.e., the special fibre $X_{k}$ is empty, then the completion $\widehat A$ is the zero ring and the topological K-theory of $X$ vanishes. \end{rems} \begin{Nr}[Connection with algebraic K-theory] Recall that for any ring $A$, the Karoubi-Villamayor K-groups \cite{KV} can be defined as \[ KV_i(A) = \pi_i\left(B_\dot\GL(A[\Delta^\dot])\right), \quad i\geq 1, \] where $A[\Delta^\dot]$ is the simplicial ring with $A[\Delta^n] = A[x_0, \dots, x_n]/(\sum_i x_i -1)$. There is a natural map from the Quillen K-group $K_i(A)$ to $KV_i(A)$ which is an isomorphism when $A$ is regular. Since we are only interested in the case of regular rings, we will in the following identify $K_i(A) = KV_i(A)$. Consider a smooth affine $R$-scheme $X = \Spec(A)$ as above. There is a natural map $A[\Delta^\dot] \to A^\dag\<\Delta^\dot\>^\dag$. We define the bisimplicial set \begin{equation}\label{nov0904} F(X) := F(A) := B_\dot\GL\left(A[\Delta^\dot]\right) \times_{B_\dot\GL\left(A^\dag\<\Delta^\dot\>^\dag\right)} E_\dot\GL\left(A^\dag\<\Delta^\dot\>^\dag\right) \end{equation} and the \emph{relative K-groups of $X$} \[ K_{i}^{\mathrm{rel}}(X) := \pi_{i} F(A),\quad i>0. \] We will also need the following finite level variant of $F(A)$: \begin{equation}\label{dez1202} F_{r}(A) := B_\dot\GL_{r}\left(A[\Delta^\dot]\right) \times_{B_\dot\GL_{r}\left(A^\dag\<\Delta^\dot\>^\dag\right)} E_\dot\GL_{r}\left(A^\dag\<\Delta^\dot\>^\dag\right) \end{equation} so that $F(A) = \varinjlim_{r} F_{r}(A)$ and $K_{i}^{\mathrm{rel}}(X) = \varinjlim_{r} \pi_{i}(F_{r}(A))$. Since the projection $E_\dot\GL(A^\dag\<\Delta^\dot\>^\dag) \to B_\dot\GL(A^\dag\<\Delta^\dot\>^\dag)$ is a Kan fibration on the diagonal simplicial sets and since $E_\dot\GL(A^\dag\<\Delta^\dot\>^\dag)$ is contractible we get: \end{Nr} \begin{lemma}\label{lem:KRel} There are long exact sequences \[ \dots \to K_{i}^{\mathrm{rel}}(X) \to K_{i}(X) \to K^{-i}_{\top}(X) \to K_{i-1}^{\mathrm{rel}}(X) \to \dots \] \end{lemma} We extend the definition of topological and relative K-theory to smooth, separated $R$-schemes $X$ of finite type as follows. Write $\Delta^\dot_R := \Spec(R[\Delta^\dot])$. Since K-theory for regular schemes satisfies Zariski descent we have isomorphisms \begin{equation}\label{eq:AlgKDescent} K_{i}(X) \cong \varinjlim_{U_{\dot}\to X} \pi_{i}\holim_{[q]\in\Delta}(B_{\dot}\GL(U_{q}\times_R\Delta^\dot_R)), \quad i\geq 1 \end{equation} where $U_{\dot}\to X$ runs through all finite affine open coverings of $X$ viewed as simplicial schemes (cf. \cite[Prop. 18.1.5]{Huber}). In analogy to \eqref{eq:AlgKDescent} we define \begin{gather*} K^{-i}_{\top}(X) := \varinjlim_{U_{\dot}\to X} \pi_{i} \holim_{[q]\in\Delta} B_{\dot}\GL(\widehat U_{q} \times_R \widehat\Delta^{\dot}_R), \quad i\geq 1, \text{ and}\\ K_{i}^{\mathrm{rel}}(X):= \varinjlim_{U_{\dot}\to X} \pi_{i} \holim_{[q]\in\Delta} F(U_{q}). \end{gather*} Here $\widehat \Delta^p_R$ is the weak completion of the algebraic standard simplex $\Delta^{p}_R$ so that if $U=\Spec(A)$ then $\widehat U \times_R \widehat\Delta^{p}_R = \Spwf(A^{\dag} \otimes_{R}^{\dag} R\<\Delta^p\>^{\dag})$ (see \ref{Nr:DaggerRings}). These groups are contravariantly functorial in $X$. \begin{lemma}\label{lemma:RelKSmoothSeparated} \begin{enumerate} \item If $X=\Spec(A)$ is affine and smooth over $R$ these definitions coincide with the earlier ones. \item There are long exact sequences \[ \dots \to K_{i}^{\mathrm{rel}}(X) \to K_{i}(X) \to K^{-i}_{\top}(X) \to K_{i-1}^{\mathrm{rel}}(X) \to \dots \] as before. \end{enumerate}• \end{lemma} \begin{proof} (i) Calvo's theorem \ref{nov0201}(ii) implies that for a smooth affine $R$-scheme $U$ we have weak equivalences \[ B_{\dot}\GL(\widehat U \times_{R} \widehat\Delta^{\dot}_R) \xrightarrow{\sim} B_{\dot}\GL(U_{k} \times_{k} \Delta^{\dot}_{k}). \] Since $\holim$ preserves weak equivalences between fibrant simplicial sets, for any open affine covering $U_{\dot} \to X$, viewed as simplicial scheme, we get weak equivalences \[ \holim_{[q]\in\Delta} B_{\dot}\GL(\widehat U_{q} \times_{R} \widehat \Delta^{\dot}_R) \xrightarrow{\sim} \holim_{[q]\in\Delta} B_{\dot}\GL((U_{q})_{k}\times_{k} \Delta^{\dot}_{k}). \] Taking $\pi_{i}$ and the limit over all finite affine coverings yields isomorphisms \begin{multline*} \varinjlim_{U_{\dot}\to X} \pi_{i} \holim_{[q]\in\Delta} B_{\dot}\GL(\widehat U_{q} \times_{R} \widehat \Delta^{\dot}_R) \xrightarrow{\cong} \varinjlim_{U_{\dot}\to X} \pi_{i} \holim_{[q]\in\Delta} B_{\dot}\GL((U_{q})_{k}\times_{k} \Delta^{\dot}_{k}) \\ \overset{\eqref{eq:AlgKDescent}}{\cong} K_{i}(X_{k}) \overset{\text{Calvo}}{\cong} K^{-i}_{\top}(A^{\dag}) \end{multline*} which is (i) for $K^{-*}_{\top}$. Using the five lemma the result for $K_{*}^{\mathrm{rel}}$ follows from this, \eqref{eq:AlgKDescent} again, and (ii). (ii) follows from Lemma \ref{lem:KRel} and the fact that $\holim$ preserves homotopy fibrations of fibrant simplicial sets \cite[Lemma 5.12]{ThomasonAlgK}. \end{proof} \begin{rem} Using $K$-dagger or $K$-affinoid algebras in the above constructions we get a notion of topological K-theory for rigid $K$-spaces. It is likely that this coincides with the one defined by Ayoub \cite{Ayoub} using the stable homotopy category of schemes and rigid varieties, respectively. \end{rem} \section{Preliminaries on functorial complexes} \label{sec:Functorial} For our construction of the relative Chern character we need functorial complexes computing the different cohomology theories involved. The main work has been done before by Huber \cite{Huber}, Besser \cite{Besser}, and Chiarellotto-Ciccioni-Mazzari \cite{Mazzari}. The only difference in our approach is the systematic use of dagger spaces also initiated by Huber and Kings \cite{HK} which simplifies the construction of the rigid and syntomic complexes. \begin{Nr}[Godement resolutions]\label{Nr:Godement} We recall some facts on Godement resolutions (see \cite[Sections 3 and 4]{Mazzari} for more details and references). To a morphism of sites $u\colon P \to X$ and an abelian sheaf $\mathcal F$ on $X$ one associates a cosimplicial sheaf $[p]\mapsto (u_{*}u^{*})^{p+1}\mathcal F$ on $X$ where the structure maps are induced by the unit and the counit of the adjoint pair $(u^{*}, u_{*})$ between the categories of abelian sheaves on $P$ and $X$. The associated complex of sheaves on $X$ will be denoted by $\Gd_{P}\mathcal F$. There is a canonical augmentation $\mathcal F \to \Gd_{P}\mathcal F$ which is a quasi-isomorphism if $u^{*}$ is exact and conservative (e.g. \cite[Lemma 3.4.1]{Ivorra}). We want to use this in the situation where $P$ is a certain set of points of (the topos associated with) $X$ with the discrete topology. The first case is that of a \emph{scheme} $X$. Here we take $P=Pt(X)$ to be the set of all points of the underlying topological space of $X$. Then $u^{*}$ is given by $\mathcal F \mapsto \coprod_{x\in X}\mathcal F_{x}$ which is exact and conservative and $\mathcal F \xrightarrow{\simeq} \Gd_{Pt(X)}\mathcal F$ is the usual Godement resolution. The second case is that of a dagger space $X$. Here it is not enough to take just the usual points of $X$. Instead, one has to use the set of prime filters on $X$ (introduced in \cite{vdPS}) as alluded to in \cite{Besser} and carried out in \cite[Section 3]{Mazzari}. We take $P=Pt(X)$ to be the set of prime filters (\cite[Def. 3.6]{Mazzari}) on the rigid space $X^{\mathrm{rig}}$ associated with $X$ with the discrete topology. Then there are morphisms of sites $Pt(X) \xrightarrow[\text{\cite[3.6]{Mazzari}}]{\xi} X^{\mathrm{rig}} \xrightarrow[\text{cf. \ref{Nr:DaggerSpaces}}]{u} X$. Since $u$ is the identity on underlying G-topological spaces we get from \cite[Lemma 3.8]{Mazzari} that $\xi^{-1}u^{-1}$ is exact and conservative. Hence for any abelian sheaf $\mathcal F$ on $X$ the augmentation $\mathcal F \to \Gd_{Pt(X)}X$ is a quasi-isomorphism. It is important to note that in both cases the complex $\Gd_{P}\mathcal F$ consists of flabby sheaves. This follows automatically since on a discrete site every sheaf is flabby and direct images of flabby sheaves are flabby. More generally, if $\mathcal F^{*}$ is a bounded below complex of abelian sheaves on $X$ we can apply $\Gd_{P}$ to each component to get a double complex. We then denote by $\Gd_{P} \mathcal F^{*}$ its associated total complex. It follows from a simple spectral sequence argument that the induced morphism $\mathcal F^{*} \to \Gd_{P}\mathcal F^{*}$ is a quasi-isomorphism. An important feature of the Godement resolution is their functorial behavior: If \[ \xymatrix{ Q\ar[r] \ar[d] & P \ar[d] \\ Y \ar[r]^{f} & X} \] is a commutative diagram of sites, and $\mathcal F$ (resp. $\mathcal G$) is a sheaf on $X$ (resp. $Y$), then a morphism $\mathcal F \to f_{*}\mathcal G$ induces a morphism $\Gd_{P}\mathcal F \to f_{*}\Gd_{Q}\mathcal G$ compatible with the augmentations \cite[Lemma 3.2]{Mazzari}. \end{Nr} \begin{Nr}[{Analytic de Rham cohomology}]\label{mar2005} Let $X$ be a smooth $K$-dagger space. There is a notion of differential forms on $X$ (cf. \cite[\S 4]{GKRigid}): For an open affinoid $U=\Sp(A)$ the differential $d\colon A=\Gamma(U,\O_{X}) \to \Gamma(U,\Omega^{1}_{X/K})$ is universal for $K$-derivations of $A$ in finite $A$-modules. As usual one constructs the de Rham complex $\Omega^{*}_{X/K}$ and defines \[ H^{*}_{\mathrm{dR}}(X/K) := \mathbb{H}^{*}(X, \Omega^{*}_{X/K}). \] We define a complex of $K$-vector spaces, functorial in the $K$-dagger space $X$ \[ R\Gamma_{\mathrm{dR}}(X/K):=\Gamma(X, \Gd_{Pt(X)}\Omega^{*}_{X/K}). \] Since the complex $\Gd_{Pt(X)}\Omega^{*}_{X/K}$ consists of flabby sheaves which are acyclic for $\Gamma(X,\,.\,)$ we have natural isomorphisms \[ H^{*}(R\Gamma_{\mathrm{dR}}(X/K)) \cong H^{*}_{\mathrm{dR}}(X/K). \] \end{Nr} \begin{Nr}[{Algebraic de Rham cohomology}] Here $X$ is a smooth, separated scheme of finite type over $K$. Its de Rham cohomology is by definition the hypercohomology of the complex of K\"ahler differential forms: \[ H^{*}_{\mathrm{dR}}(X/K) := \mathbb{H}^{*}(X, \Omega^{*}_{X/K}). \] It is equipped with the \emph{Hodge filtration} constructed as follows: By Nagata's compactification theorem and Hironaka's resolution of singularities there exists a \emph{good compactification of $X$}, i.e. an open immersion $X\overset{j}{\hookrightarrow} \ol X$ of $X$ in a smooth, proper $K$-scheme $\ol X$ such that the complement $D = \ol X - X$ is a divisor with normal crossings. On $\ol X$ one has the complex $\Omega^{*}_{\ol X/K}(\log D)$ of differential forms with logarithmic poles along $D$. There are isomorphisms \[ H^{*}_{\mathrm{dR}}(X/K) \cong \mathbb{H}^{*}(\ol X,\Omega^{*}_{\ol X/K}(\log D)), \] the Hodge--de Rham spectral sequence $E_{1}^{p,q} = H^{q}(\ol X, \Omega^{p}_{\ol X/K}(\log D)) \Rightarrow H^{*}_{\mathrm{dR}}(X/K)$ degenerates at $E_{1}$, and the induced filtration $F^\dot H^{*}_{\mathrm{dR}}(X/K)$ is independent of the choice of $\ol X$. It is given by \[ F^{n} H^{*}_{\mathrm{dR}}(X/K) = \mathbb{H}^{*}(\ol X, \Omega^{\geq n}_{\ol X/K}(\log D)), \] $\geq n$ denoting the naive truncation. Since the good compactifications of $X$ form a directed set with respect to maps under $X$ and taking the colimit along a directed set is exact, to get functorial complexes computing algebraic de Rham cohomology together with its Hodge filtration we could take the colimit of the $\Gamma(\ol X, \Gd_{Pt(\ol X)}\Omega^{*}_{\ol X/K}(\log D))$ along the system of good compactifications $\ol X$ of $X$. However, for the comparison with analytic de Rham cohomology it is technically easier to use the following variant (see \cite[Prop. 4.9]{Mazzari}): Let $X^{\dag}$ be the dagger analytification of $X$ (cf. \ref{Nr:DaggerSpaces}) and $Pt(X^{\dag})$ as in \ref{Nr:Godement}, $Pt(\ol X)$ the usual set of points of the good compactification $\ol X$ of $X$. We can form the disjoint sum $Pt(X^{\dag}) \sqcup Pt(\ol X)$ viewed as site with the discrete topology to get a commutative diagram of sites \begin{equation}\label{eq:DiagSites} \begin{split} \xymatrix{ Pt(X^{\dag}) \ar[d] \ar[rr] && Pt(X^{\dag}) \sqcup Pt(\ol X) \ar[d] \\ X^{\dag} \ar[r]^{\iota}_{\text{cf. \ref{Nr:DaggerSpaces}}} & X \ar[r]^{j} & \ol X. } \end{split} \end{equation} There are natural morphisms $\Omega^{*}_{\ol X/K}(\log D) \to j_{*}\Omega^{*}_{X/K} \to j_{*}\iota_{*}\Omega^{*}_{X^{\dag}/K}$ which together with \eqref{eq:DiagSites} induce a natural map \begin{equation}\label{eq:GdCompMap} \Gd_{Pt(X^{\dag}) \sqcup Pt(\ol X)}\Omega^{*}_{\ol X/K}(\log D) \to j_{*}\iota_{*}\Gd_{Pt(X^{\dag})}\Omega^{*}_{X^{\dag}/K}. \end{equation} Thus we are led to define \begin{equation}\label{eq:FilDRCompl} F^{n}R\Gamma_{\mathrm{dR}}(X/K) := \varinjlim_{X\hookrightarrow \ol X}\Gamma(\ol X, \Gd_{Pt(X^{\dag}) \sqcup Pt(\ol X)}\Omega^{\geq n}_{\ol X/K}(\log D)) \end{equation}• where the limit runs over the directed set of good compactifications of $X$. It follows from the discussion above that there are natural isomorphisms \[ H^{*}(F^{n}R\Gamma_{\mathrm{dR}}(X/K)) \cong F^{n}H^{*}_{\mathrm{dR}}(X/K), \] and \eqref{eq:GdCompMap} induces natural comparison maps \begin{equation}\label{eq:ComplCompMap} F^{n}R\Gamma_{\mathrm{dR}}(X/K) \to R\Gamma_{\mathrm{dR}}(X^{\dag}/K). \end{equation} \end{Nr} \section{The relative Chern character} \label{sec:Construction} As before $R$ denotes a complete discrete valuation ring with field of fractions $K$ of characteristic $0$ and residue field $k$ of characteristic $p>0$. Let $\mathsf{Sm}_R$ be the category of smooth, separated $R$-schemes of finite type. For $X \in \mathsf{Sm}_{R}$ we have its generic fibre $X_{K}$ with dagger analytification $X_{K}^{\dag}$, and its weak completion $\widehat X$ with generic fibre $\widehat X_{K}$, related by the following morphisms of locally G-ringed spaces \[ \widehat X_{K} \subseteq X_{K}^{\dag} \xrightarrow{\iota} X_{K}. \] In particular, we have morphisms of complexes \begin{equation}\label{eq:CompDeRhamRigid} F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \xrightarrow{\eqref{eq:ComplCompMap}} R\Gamma_{\mathrm{dR}}(X_{K}^{\dag}/K) \xrightarrow{\text{by functoriality}} R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K). \end{equation} \begin{Nr}\label{dez1201} We denote by $\mathsf{Ch}$ the category of complexes of abelian groups. For a morphism $A\xrightarrow{f} B$ in $\mathsf{Ch}$ we denote by $\MF(A\to B) := \Cone(A \to B)[-1]$ the mapping fibre. It has the following property: If $C$ is a complex, the morphisms $C \to \MF(A\to B)$ are in one-to-one correspondence with pairs $(g,h)$ where $g\colon C \to A$ is a morphism of complexes and $h\colon C \to B[-1]$ is a homotopy such that $dh+hd=f\circ g$. \end{Nr} \begin{dfn} \label{nov0905} For every integer $n$ we define a functor $R\Gamma_{\mathrm{rel}}(\,.\,,n)\colon \mathsf{Sm}_{R}^{\mathrm{op}} \to \mathsf{Ch}$ by \[ R\Gamma_{\mathrm{rel}}(X,n) := \MF(F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K)\xrightarrow{\eqref{eq:CompDeRhamRigid}} R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K)) \] and \emph{relative cohomology groups} $H^{*}_{\mathrm{rel}}(X,n) := H^{*}(R\Gamma_{\mathrm{rel}}(X,n))$. \end{dfn} \begin{rems}\label{nov0203} (i) These are closely related to rigid syntomic cohomology, see Lemma \ref{nov0202} below. The complex $R\Gamma_{\mathrm{rel}}(X,n)$ can also be interpreted as the syntomic $P$-complex $\mathbb R\Gamma_{\mathrm f,1}(X,n)$ of \cite[2.2]{BesserColeman} for the polynomial $P=1$. \noindent (ii) Since $X/R$ is smooth, the de Rham cohomology of $\widehat X_{K}$ is just the rigid cohomology of the special fibre $X_{k}$ (see Section \ref{sec:rigidcohom} below). Hence the relative cohomology groups sit in exact sequences \[ \dots \to H^{i}_{\mathrm{rel}}(X,n) \to F^{n}H^{i}_{\mathrm{dR}}(X_{K}/K) \to H^{i}_{\mathrm{rig}}(X_{k}/K) \to \dots \] \noindent (iii) If $X/R$ is proper then $\widehat X_{K} = X_{K}^{\dag}$ (cf. \ref{Nr:DaggerSpaces}) and by GAGA \cite[Kor. 4.5]{GK} $R\Gamma_{\mathrm{dR}}(X_{K}^{\dag}/K) \simeq R\Gamma_{\mathrm{dR}}(X_{K}/K)$, where $\simeq$ denotes a quasi-isomorphism. Hence $R\Gamma_{\mathrm{rel}}(X, n) \simeq \MF(F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to R\Gamma_{\mathrm{dR}}(X_{K}/K))$ in this case, and the degeneration of the Hodge--de Rham spectral sequence yields isomorphisms \[ H^i_\mathrm{rel}(X,n) \cong H^{i-1}_\mathrm{dR}(X_K/K)/F^nH^{i-1}_\mathrm{dR}(X_K/K). \] \noindent (iv) For an interpretation of the relative cohomology in terms of stable $\mathbf A^{1}$-homotopy theory we refer the reader to \cite{DegliseMazzari}. \end{rems} The goal of this section is to construct relative Chern character maps which will be homomorphisms \[ \ch_{n,i}^{\mathrm{rel}}\colon K_{i}^{\mathrm{rel}}(X) \to H^{2n-i}_{\mathrm{rel}}(X,n). \] We first describe an abstract formalism to obtain homomorphisms from the homotopy groups of certain simplicial sets to the cohomology of suitable functorial complexes and then specialize this to the construction of the relative Chern character and, in the next section, of the syntomic regulator. This formalization of the constructions makes it easier to compare them afterwards. \begin{Nr} \label{dez0701} We view complexes in $\mathsf{Ch}$ either homologically $\dots\to C_i \xrightarrow{d} C_{i-1} \to \dots$ or cohomologically $\dots\to C^{-i} \xrightarrow{d} C^{-i+1}\to \dots$ using the convention $C_i = C^{-i}$. Given $A, B \in \mathsf{Ch}$, we denote by $\ul\Hom(A,B)$ the mapping complex. In degree $i$ it is given by $\prod_{p} \Hom(A^{p},B^{p+i})$ with differential $f\mapsto f\circ d^{A} - (-1)^{i}d^{B}\circ f$. In particular, cycles in degree $i$ are given by $Z^i\ul\Hom(A,B)=\Hom_{\mathsf{Ch}}(A,B[i])$. If $C^{\dot,\dot}$ is a double complex, the differential of the total complex is given on $C^{p,q}$ by $d^{\mathrm{horiz}} + (-1)^{p}d^{\mathrm{vert}}$ ($p$ is the horizontal coordinate). \end{Nr} \begin{Nr} \label{nov1205} We consider the following setup: $\mathsf{S}$ is a category, $a\colon \mathsf{Sm}_R \to \mathsf{S}$ a functor, $\Gamma_0\colon \mathsf{Sm}_R^{\mathrm{op}} \to \mathsf{Ch}$ and $\Gamma_1\colon \mathsf{S}^\mathrm{op} \to \mathsf{Ch}$ are functorial complexes, and we have a natural transformation $\Gamma_0 \to \Gamma_1\circ a$. We fix a morphism $E \to B$ in $\mathsf{Sm}_R$. Consider a map $X \xrightarrow{f} B$ in $\mathsf{Sm}_R$ together with a map $a(X) \xrightarrow{g} a(E)$ in $\mathsf{S}$ such that \[ \xymatrix@[email protected]{ & a(E)\ar[d]\\ a(X) \ar[ur]^g \ar[r]_{a(f)} & a(B) } \] commutes, in other words, an element $(f,g) \in B(X) \times_{a(B)(a(X))} a(E)(a(X))$, where $B(X):= \Hom_{\mathsf{Sm}_R}(X,B)$, etc.. By abuse of notation we write $(B\times_{a(B)} a(E))(X)$ for this set. Then the pair $(f,g)$ gives a commutative diagram \[ \xymatrix@[email protected]{ \Gamma_0(B) \ar[r] \ar[d]_{f^*} & \Gamma_1(a(B)) \ar[d]_{a(f)^*} \ar[r] & \Gamma_1(a(E)) \ar[dl]^-{g^*}\\ \Gamma_0(X) \ar[r] & \Gamma_1(a(X)) } \] in $\mathsf{Ch}$ and hence a morphism of complexes, i.e. a zero cycle in the $\ul\Hom$-complex \[ \MF\left(\Gamma_0(B) \to \Gamma_1(a(E))\right) \to \MF\left(\Gamma_0(X) \to \Gamma_1(a(X))\right). \] This construction induces a morphism of complexes \begin{multline*} \mathbf{Z}\left[(B\times_{a(B)} a(E))(X)\right] \to \\ \ul\Hom\left(\MF\left(\Gamma_0(B) \to \Gamma_1(a(E))\right), \MF\left(\Gamma_0(X) \to \Gamma_1(a(X))\right) \right) \end{multline*} where $\mathbf{Z}[\,.\,]$ is the free abelian group considered as a complex in degree $0$. If $E_\dot \to B_\dot$ is a morphism of simplicial objects in $\mathsf{Sm}_R$, and $X^\dot$ is a cosimplicial object in $\mathsf{Sm}_R$, then $([p],[q])\mapsto\left(B_p\times_{a(B_{p})} a(E_p)\right)(X^q)$ is a bisimplicial set, and we get a natural map of complexes \begin{multline}\label{dez0702} \Tot\mathbf{Z}\left[\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right] \to \\ \ul\Hom\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right), \MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right) \right). \end{multline} Here $\Gamma_0(B_\dot), \Gamma_0(X^\dot)$, etc. are defined as the direct sum total complexes and we view the simplicial, respectively cosimplicial direction as the horizontal one. On the left hand side, the vertical direction is that coming from $X^{\dot}$. Using the sign conventions from \ref{dez0701}, we have to introduce a sign $(-1)^{q(q-1)/2}$ in bidegree $(p,q)$ in order that \eqref{dez0702} is a morphism of complexes. On homology \eqref{dez0702} induces for every integer $*$ a map \begin{multline}\label{nov0801} H_i\left(\Tot\mathbf{Z}\left[\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right]\right) \to \\ \Hom\left(H^*\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right)\right), H^{*-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right) \right) \right). \end{multline} In particular, any class $c\in H^{2n}\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right)\right)$ gives by composing \eqref{nov0801} with the evaluation at $c$ a map \begin{multline}\label{nov0901} \phantom{.}^{*}c\colon H_i\left(\Tot\mathbf{Z}\left[\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right]\right) \to \\ H^{2n-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right)\right). \end{multline} \end{Nr}% \begin{dfn} A \emph{regulator datum} is a tuple $\omega$ consisting of (1) a category $\mathsf{S}$ together with a functor $a\colon \mathsf{Sm}_{R} \to \mathsf{S}$, (2) functors $\Gamma_{0}\colon \mathsf{Sm}_{R}^{\mathrm{op}} \to \mathsf{Ch}, \Gamma_{1}\colon \mathsf{S}^{\mathrm{op}} \to \mathsf{Ch}$ together with a natural transformation $\Gamma_{0} \to \Gamma_{1}\circ a$, (3) a morphism of simplicial objects $E_{\dot}\to B_{\dot}$ in $\mathsf{Sm}_{R}$, and (4) a class $c \in H^{2n}\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right)\right)$. To simplify notation, we denote such a regulator datum by $\omega = (\mathsf{S}, E_{\dot} \to B_{\dot}, \Gamma_{0} \to \Gamma_{1}\circ a, c)$. \end{dfn} \begin{lemma}\label{nov1602} A regulator datum $\omega$ induces for every cosimplicial object $X^{\dot}$ in $\mathsf{Sm}_{R}$ and $i\geq 0$ a homomorphism \[ \reg_{i}(\omega)\colon \pi_i\left(\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right) \to H^{2n-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right)\right) \] \end{lemma} \begin{proof} The desired homomorphism $\reg_{i}(\omega)$ is the composition \begin{align* & \pi_i\left(\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right) \\ \longrightarrow& H_i\left(\mathbf{Z}\left[\diag \left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right]\right) \qquad \text{ Hurewicz}\\ \overset{\cong}{\longrightarrow}& H_i\left(\Tot \mathbf{Z}\left[\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right]\right) \qquad \text{ by Eilenberg-Zilber}\\ \overset{\!\phantom{.}^{*}c}{\longrightarrow}& H^{2n-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right)\right) \qquad \text{ by \eqref{nov0901}} \qedhere \end{align*} \end{proof} We record the following naturality properties which are easily established. They will be used in the comparison of the relative with the syntomic Chern character in the next section. Consider two regulator data $\omega = (\mathsf{S}, E_{\dot}\to B_{\dot}, \Gamma_{0}\to\Gamma_{1}\circ a, c)$ and $\omega' = (\mathsf{S}', E_{\dot}\to B_{\dot}, \Gamma_{0}'\to\Gamma_{1}'\circ a', c')$ with the same $E_{\dot}\to B_{\dot}$. Assume moreover, that we have a functor $b\colon \mathsf{S} \to \mathsf{S}'$ such that $b\circ a \cong a'$, natural transformations $\Gamma_0 \to \Gamma_0'$, $\Gamma_1 \to \Gamma_1'\circ b$ and a natural homotopy $h$ between the compositions $\Gamma_0 \to \Gamma_0' \to \Gamma_1'\circ a'$ and $\Gamma_0 \to \Gamma_1\circ a \to \Gamma_1'\circ b\circ a\cong \Gamma_{1}'\circ a'$. For every map $Z\to Y$ in $\mathsf{Sm}_R$ these induce a map $\MF(\Gamma_0(Y) \to \Gamma_1(a(Z))) \to \MF(\Gamma_0'(Y) \to \Gamma_1'(a'(Z)))$ (cf. \ref{dez1201}). \begin{lemma}\label{nov1301} If $c$ maps to $c'$ by the natural map $H^{2n}\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right)\right)\to H^{2n}\left(\MF\left(\Gamma_0'(B_\dot) \to \Gamma_1'(a'(E_\dot))\right)\right)$ then for every $X^{\dot}$ the diagram \[ \xymatrix@[email protected]{ \pi_i\left(\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right) \ar[r]^-{\reg_{i}(\omega)} \ar[d] & H^{2n-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right)\right) \ar[d] \\ \pi_i\left(\left(B_\dot\times_{a'(B_{\dot})} a'(E_\dot)\right)(X^\dot)\right) \ar[r]^-{\reg_{i}(\omega')} & H^{2n-i}\left(\MF\left(\Gamma_0'(X^\dot) \to \Gamma_1'(a'(X^\dot))\right)\right) } \] commutes. \end{lemma} We now consider regulator data $\omega = (\mathsf{S}, E_{\dot}\to B_{\dot}, \Gamma_{0}\to\Gamma_{1}\circ a, c)$ and $\omega' = (\mathsf{S}, E'_{\dot}\to B'_{\dot}, \Gamma_{0}\to\Gamma_{1}\circ a, c')$ with the same category $\mathsf{S}$ and complexes $\Gamma_{0}, \Gamma_{1}$, and assume that we have a commutative diagram of simplicial objects in $\mathsf{Sm}_R$ \[ \[email protected]@C-0.3cm{ E_\dot \ar[r] \ar[d] & E'_\dot\ar[d]\\ B_\dot \ar[r] & B'_\dot. } \] \begin{lemma}\label{nov1302} If $c'\in H^{2n}\left(\MF\left(\Gamma_0(B'_\dot) \to \Gamma_1(a(E'_\dot))\right)\right)$ maps to the class $c \in H^{2n}\left(\MF\left(\Gamma_0(B_\dot) \to \Gamma_1(a(E_\dot))\right)\right)$ by the induced map, then for every $X^{\dot}$ the diagram \[ \xymatrix{ \pi_i\left(\left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot)\right) \ar[r] \ar[dr]_{\reg_{i}(\omega)}& \pi_i\left(\left(B'_\dot\times_{a(B'_{\dot})} a(E'_\dot)\right)(X^\dot)\right) \ar[d]^{\reg_{i}(\omega')} \\ & H^{2n-i}\left(\MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right)\right) } \] commutes. \end{lemma} We now construct a regulator datum $\omega^{\mathrm{rel}}_{n,(r)}$ that produces the relative Chern character. Let $\mathsf{Smf}_{R}$ be the category of smooth weak formal $R$-schemes, $a\colon \mathsf{Sm}_R \to \mathsf{Smf}_{R}$ the weak completion functor $X \mapsto \widehat X$, $\Gamma_0$ given by $X \mapsto F^nR\Gamma_\mathrm{dR}(X_K/K)$, $\Gamma_1$ given by $\mathscr X \mapsto R\Gamma_\mathrm{dR}(\mathscr X_K/K)$ and the natural transformation $\Gamma_0 \to \Gamma_1\circ a$ given by \eqref{eq:CompDeRhamRigid}. For $E_\dot \to B_\dot$ we take $E_\dot\GL_{r} \to B_\dot\GL_{r}$. As cosimplicial object in $\mathsf{Sm}_R$ we will always take $X^\dot=X\times_R\Delta_{R}^\dot = \Spec(A[\Delta^\dot])$ for some affine $X= \Spec(A)$ in $\mathsf{Sm}_R$. With these choices we have \begin{align} & \left(B_\dot\times_{a(B_{\dot})} a(E_\dot)\right)(X^\dot) = F_{r}(A), && \text{ cf. \eqref{dez1202},} \label{nov1201}\\ & \MF\left(\Gamma_0(X^\dot) \to \Gamma_1(a(X^\dot))\right) = R\Gamma_\mathrm{rel}(X\times_{R}\Delta_{R}^\dot, n) && \text{ by Definition \ref{nov0905}.}\label{nov1202} \end{align} For the equality \eqref{nov1201} we use that $\widehat\GL_r(\widehat X) = \Hom_{\mathsf{Smf}_{R}}(\widehat X, \widehat\GL_r) \cong \GL_r(A^\dag)$. To get a regulator datum we need to specify the class $c$. This is accomplished by \begin{lemma}\label{nov1204} For each $r\geq 1$ the natural map \begin{multline*} H^{2n}\left(\MF(F^{n}R\Gamma_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \to R\Gamma_{\mathrm{dR}}(E_{\dot}\widehat\GL_{r,K}/K))\right) \to \\ H^{2n}(F^{n}R\Gamma_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K)) = F^{n}H^{2n}_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \end{multline*} is an isomorphism. \end{lemma} \begin{proof} This follows from the long exact sequence for the cohomology of a cone together with the fact that $E_{\dot}\widehat\GL_{r,K} $ is a contractible simplicial dagger space, hence has no cohomology in positive degrees (cf. \cite[Lemma 2.11]{TammeDiss}). \end{proof} In particular, there is a unique class \begin{equation}\label{nov1401} \ch_{n,(r)}^\mathrm{rel} \in H^{2n}\left(\MF(F^{n}R\Gamma_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \to R\Gamma_{\mathrm{dR}}(E_{\dot}\widehat\GL_{r,K}/K))\right) \end{equation} which is mapped to the degree $2n$ component $\ch_{n,(r)}^\mathrm{dR} \in F^{n}H^{2n}_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K)$ of the universal Chern character class in de Rham cohomology. Since the $\ch_{n,(r)}^{\mathrm{dR}}$ are compatible for varying $r\geq 1$, so are the $\ch_{n,(r)}^{\mathrm{syn}}$. We also need the following lemma. At this point it is crucial to work with dagger spaces. \begin{lemma}\label{nov1206} The natural map $R\Gamma_{\mathrm{rel}}(X,n) \to R\Gamma_{\mathrm{rel}}(X \times \widehat\Delta^{\dot}_{R},n)$ is a quasi-isomorphism for any $X\in \mathsf{Sm}_{R}$. \end{lemma} \begin{proof} It suffices to check this for both components of the cone separately. We show that $R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K) \to R\Gamma_{\mathrm{dR}}(\widehat X_{K} \times \widehat\Delta_{K}^{\dot}/K)$ is a quasi-isomorphism. Using in addition the degeneration of the Hodge--de Rham spectral sequence, the claim for $F^nR\Gamma_\mathrm{dR}(X_K/K) \to F^nR\Gamma_\mathrm{dR}(X_K\times \Delta^\dot_K/K)$ follows similarly. By construction $R\Gamma_{\mathrm{dR}}(\widehat X_{K} \times \widehat\Delta_{K}^{\dot}/K)$ is the direct sum total complex of a double complex in the second quadrant. The filtration by columns gives a convergent spectral sequence in the second quadrant (cf. \cite[5.6.1]{Weibel} for the dual homological case) \[ E_{1}^{p,q} = H^{q}_{\mathrm{dR}}(\widehat X_{K}\times\widehat\Delta_{K}^{-p}/K) \Rightarrow H^{p+q}(R\Gamma_{\mathrm{dR}}(\widehat X_{K} \times \widehat\Delta_{K}^{\dot}/K)). \] The differential $d_{1}$ is induced from the cosimplicial structure of $\widehat\Delta_{K}^{\dot}$. The homotopy invariance of de Rham cohomology of dagger spaces \cite[Prop. 5.8]{GK} implies that $d_{1}^{p,q}$ is the identity if $p$ is even and zero if $p$ is odd. Hence $E_{2}^{0,q} = H^{q}_{\mathrm{dR}}(\widehat X_{K}/K)$, $E_{2}^{p,q}=0$ if $p<0$. It follows that the edge morphism $E_{2}^{0,q} = H^{q}_{\mathrm{dR}}(\widehat X_{K}/K) \to H^{q}(R\Gamma_{\mathrm{dR}}(\widehat X_{K} \times \widehat\Delta_{K}^{\dot}/K))$ is an isomorphism. \end{proof} \begin{dfn}\label{nov1601} Let $X = \Spec(A)$ be a smooth affine $R$-scheme of finite type. Let $\omega^{\mathrm{rel}}_{n,(r)}$ be the regulator datum $(\mathsf{Smf}_{R}, E_{\dot}\GL_{r} \to B_{\dot}\GL_{r}, F^{n}R\Gamma_{\mathrm{dR}}((.)_{K}/K)\to R\Gamma_{\mathrm{dR}}((\widehat.)_{K}/K), \ch_{n,(r)}^{\mathrm{rel}})$. By \ref{nov1602} this gives homomorphisms \[ \reg_{i}(\omega^{\mathrm{rel}}_{n,(r)})\colon \pi_i(F_{r}(A)) \to H^{2n-i}\left(R\Gamma_\mathrm{rel}(X\times \Delta^\dot,n)\right) \overset{\text{\ref{nov1204}}}{\cong} H^{2n-i}_\mathrm{rel}(X,n), \] which are compatible for varying $r\geq 1$. The \emph{relative Chern character} is defined to be the colimit \[ \ch_{n,i}^\mathrm{rel}\colon K_i^\mathrm{rel}(X) = \varinjlim_{r}\pi_i(F_{r}(A)) \xrightarrow{\varinjlim_{r}\reg_{i}(\omega^{\mathrm{rel}}_{n,(r)})} H^{2n-i}_\mathrm{rel}(X,n). \] \end{dfn} We use Jouanolou's trick to extend this definition to all schemes in $\mathsf{Sm}_{R}$. According to Jouanolou and Thomason \cite[4.4]{WeibelKH} such a scheme $X$ admits a Jouanolou torsor $W\xrightarrow{p} X$, i.e. $W$ is affine and $p$ is a torsor for some vector bundle on $X$. \begin{lemma} In the above situation the map $p^*\colon R\Gamma_\mathrm{rel}(X,n) \to R\Gamma_\mathrm{rel}(W,n)$ is a quasi-isomorphism. \end{lemma} \begin{proof} It is enough to show that $p$ induces a quasi-isomorphism on both components of the cone. We show that $R\Gamma_\mathrm{dR}(\widehat X_K/K) \xrightarrow{p^*} R\Gamma_\mathrm{dR}(\widehat W_K/K)$ is a quasi-isomorphism. Using moreover the degeneration of the Hodge--de Rham spectral sequence, the proof for $F^nR\Gamma_\mathrm{dR}$ is similar. Choose a finite open covering $X=\bigcup_{\alpha\in A} U_{\alpha}$ such that $p^{-1}(U_{\alpha}) \to U_{\alpha}$ is isomorphic to a trivial vector bundle $\mathbb A^{r}_{U_{\alpha}} \to U_{\alpha}$. Let $U_{\dot} \to X$ be the \v{C}ech nerve of this covering and denote by $p^{-1}(U_{\dot}) \to W$ its base change to $W$. Since $\{(\widehat U_\alpha)_K\}_{\alpha\in A}$ is an admissible covering of $\widehat X_K$ it follows that $R\Gamma_\mathrm{dR}(\widehat X_K/K) \to R\Gamma_\mathrm{dR}((\widehat U_\dot)_K/K)$ and similarly $R\Gamma_\mathrm{dR}(\widehat W_K/K) \to R\Gamma_\mathrm{dR}(\widehat{p^{-1}(U_\dot)}_K/K)$ are quasi-isomorphisms. Hence we are reduced to the case that $W\to X$ is of the form $\mathbb A^r_X \to X$. Then the claim follows from homotopy invariance for the de Rham cohomology of dagger spaces \cite[Prop. 5.8]{GK} \end{proof} \begin{dfn}\label{nov0601} Let $X$ be in $\mathsf{Sm}_{R}$ and choose a Jouanolou torsor $W \xrightarrow{p} X$ . We define the \emph{relative Chern character} to be the composition \[ K_i^\mathrm{rel}(X) \xrightarrow{p^*} K_i^\mathrm{rel}(W) \xrightarrow{\ch_{n,i}^\mathrm{rel}} H^{2n-i}_{\mathrm{rel}}(W, n) \xrightarrow[\cong]{(p^*)^{-1}} H^{2n-i}_{\mathrm{rel}}(X, n). \] \end{dfn} One checks that this does not depend on the choice of $W \to X$ using the fact that for two Jouanolou torsors $W\to X, W'\to X$, the fibre product $W\times_X W' \to X$ is again a Jouanolou torsor. \section{Comparison with the rigid syntomic regulator} \label{sec:Comparison} The main technical problem in the construction of the rigid syntomic regulator is the construction of functorial complexes computing rigid and rigid syntomic cohomology. This was solved by Besser \cite{Besser}. An alternative construction of the regulator using cycle classes and higher Chow groups instead of K-theory is given in \cite{Mazzari}. We recall Besser's construction with some improvements from \cite{Mazzari}. As in \cite{HK} the systematic use of dagger spaces simplifies the construction a little bit. Let $R$ be as before and assume moreover that the residue field $k$ of $R$ is perfect. Let $K_{0} \subseteq K$ be the field of fractions of the ring of Witt vectors of $k$. \subsection{Rigid cohomology} \label{sec:rigidcohom} We consider the category $\mathsf{Sch}_{k}$ of separated schemes of finite type over $k$ which admit a closed immersion in a flat weakly formal $R$-scheme $\mathscr Y$ with smooth special fibre $\mathscr Y_k$. For $X\in\mathsf{Sm}_{R}$ the special fibre $X_{k}$ is in $\mathsf{Sch}_{k}$, as we can take the closed immersion of $X_k$ in the weak completion $\widehat X$ of $X$. Let $X$ be in $\mathsf{Sch}_{k}$ and choose an embedding $X \hookrightarrow \mathscr Y$ as above. The \emph{rigid cohomology} of $X$ with coefficients in $K$ is by definition the de Rham cohomology of the tube $]X[_{\mathscr Y} := \spec^{-1}(X) \subset \mathscr Y_{K}$ (cf. \ref{Nr:DaggerSpaces}) of $X$ in $\mathscr Y$: \[ H^{*}_{\mathrm{rig}}(X/K) = H^{*}_{\mathrm{dR}}(]X[_{\mathscr Y}/K) \] \cite{Berthelot}, \cite[Prop. 8.1]{GK}. Up to isomorphism this is independent of the choice of $\mathscr Y$. Following Besser we define \[ R\Gamma_{\mathrm{rig}}(X/K)_{\mathscr Y} := R\Gamma_{\mathrm{dR}}(]X[_{\mathscr Y}/K). \] This complex is functorial only in the \emph{pair} $(X,\mathscr Y)$. To obtain complexes functorial in the $k$-scheme $X$ we proceed as in \cite[\S 4]{Besser}. Define the category of \emph{rigid pairs} $\mathcal{RP}$: Objects are pairs $(X, j\colon X \hookrightarrow \mathscr Y)$ where $X$ and $j$ are as above. We will often abbreviate such a pair as $(X, \mathscr Y)$. Morphisms $(X', \mathscr Y') \to (X, \mathscr Y)$ are pairs of morphisms $(f\colon X'\to X, F\colon ]X'[_{\mathscr Y'} \to ]X[_{\mathscr Y})$ such that the diagram \[\xymatrix{ ]X'[_{\mathscr Y'} \ar[d]_{\spec} \ar[r]^F & ]X[_{\mathscr Y} \ar[d]^{\spec}\\ X' \ar[r]^f & X } \] commutes. The category $\mathcal{RP}$ substitutes Besser's category $\mathcal{RT}$ of rigid triples. Note that there is a natural functor $\mathsf{Sm}_{R}\to \mathcal{RP}$ taking $X$ to the pair $(X_{k}, \widehat X)$. With this replacement Besser's construction goes through word by word and yields the following \begin{prop}[{\cite[Prop. 4.9, Cor. 4.22]{Besser}}]\label{prop:BesserRigComplexes} \begin{enumerate} \item There exists a functor $R\Gamma_{\mathrm{rig}}(\,.\,,/k)\colon \mathsf{Sch}_{k}^{\mathrm{op}} \to \mathsf{Ch}_{K}$ to the category of complexes of $K$-vector spaces such that $H^{*}(R\Gamma_{\mathrm{rig}}(X/K)) = H^{*}_{\mathrm{rig}}(X/K)$ functorially. If $K$ is absolutely unramified, i.e. $K = K_{0}$, and $\sigma$ is the Frobenius on $K_{0}$, there exists a natural $\sigma$-semilinear Frobenius-endomorphism $\phi$ on $R\Gamma_{\mathrm{rig}}(X/K_{0})$. \item There exists a functor $\mathcal{RP}^{\mathrm{op}} \to \mathsf{Ch}_{K}$, $(X, \mathscr Y) \mapsto \widetilde{R\Gamma}_{\mathrm{rig}}(X/K)_{\mathscr Y}$ together with $\mathcal{RP}$-functorial quasi-isomorphisms \[ R\Gamma_{\mathrm{dR}}(]X[_{\mathscr Y}/K) = R\Gamma_{\mathrm{rig}}(X/K)_{\mathscr Y} \xleftarrow{\simeq} \widetilde{R\Gamma}_{\mathrm{rig}}(X/K)_{\mathscr Y} \xrightarrow{\simeq} R\Gamma_{\mathrm{rig}}(X/K). \] \end{enumerate} \end{prop} \subsection{Rigid syntomic cohomology} \begin{Nr}\label{Nr:QuasiPullback} Recall that the \emph{homotopy pullback} of a diagram of complexes $A \xrightarrow{f} C \xleftarrow{g} B$ is by definition the complex $A\tilde\times_{C}B := \Cone(A\oplus B \xrightarrow{f-g} C)[-1]$. It fits in a diagram \[ \xymatrix{ A \tilde\times_{C} B \ar[r]^-{\tilde f} \ar[d]_{\tilde g} & B \ar[d]^{g}\\ A \ar[r]^{f} & C } \] which is commutative up to \emph{canonical} homotopy, given by the projection to the $C$-component of the cone. If $f$ is a quasi-isomorphism, so is $\tilde f$. \end{Nr} \begin{Nr}\label{Nr:TildeComplexes} Let $X$ be in $\mathsf{Sm}_{R}$. Then $]X_{k}[_{\widehat X} = \widehat X_{K} \subseteq X_{K}^{\dag}$ and we have natural maps of complexes, functorial in $X$, \[ F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \xrightarrow{\eqref{eq:CompDeRhamRigid}} R\Gamma_{\mathrm{dR}}(]X_{k}[_{\widehat X}/K) \xleftarrow{\simeq} \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X} \xrightarrow{\simeq} R\Gamma_{\mathrm{rig}}(X_{k}/K). \] Define $F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K)$ to be the homotopy pullback of the left two arrows above, a complex quasi-isomorphic to $F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K)$ which admits natural maps \[ F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \to \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X} \xrightarrow{\simeq} R\Gamma_{\mathrm{rig}}(X_{k}/K). \] On $R\Gamma_{\mathrm{rig}}(X_{k}/K_{0})$ we have the Frobenius $\phi$ and the natural map to $R\Gamma_{\mathrm{rig}}(X_{k}/K)$. We define the complex \[ \Phi(n)(X_{k}) :=\Cone\left(R\Gamma_{\mathrm{rig}}(X_{k}/K_{0}) \to R\Gamma_{\mathrm{rig}}(X_{k}/K_{0}) \oplus R\Gamma_{\mathrm{rig}}(X_{k}/K) \right) \] where the map is given by $\omega\mapsto \left((1-\frac{\phi}{p^{n}})\omega,\omega\right)$. This complex is functorial in the $k$-scheme $X_{k}$ and there are natural maps \begin{equation}\label{eq:MapFilPhi} \xymatrix@1{ F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \ar[r]^{(\ref{eq:MapFilPhi}.1)} & \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X} \ar[r]\ar@/^1pc/[rr]^{(\ref{eq:MapFilPhi}.2)} & R\Gamma_{\mathrm{rig}}(X_{k}/K) \ar[r] & \Phi(n)(X_{k}) } \end{equation} \end{Nr} \begin{dfn} We define the \emph{syntomic complex of $X$ twisted by $n$} \[ R\Gamma_{\mathrm{syn}}(X,n):=\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \xrightarrow{-\eqref{eq:MapFilPhi}} \Phi(n)(X_{k})). \] Its cohomology groups will be denoted by $H^{*}_{\mathrm{syn}}(X,n)$. \end{dfn} \begin{rem} Writing down the iterated cone construction explicitly one sees that $R\Gamma_{\mathrm{syn}}(X,n)$ is isomorphic to the complex \[ \Cone\left(R\Gamma_{\mathrm{rig}}(X_{k}/K_{0}) \oplus F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \to R\Gamma_{\mathrm{rig}}(X_{k}/K_{0}) \oplus R\Gamma_{\mathrm{rig}}(X_{k}/K)\right)[-1] \] where the map is given by $(x,y) \mapsto ((1-\frac{\phi}{p^{n}})x, x-y)$. From this one sees that the fundamental Proposition 6.3 of \cite{Besser} also holds for our definition of rigid syntomic cohomology. Hence all further constructions of \cite{Besser} work equally well in our setting. In particular, there are natural maps \begin{equation}\label{eq:MapSynDeRham} H^{*}_{\mathrm{syn}}(X,n) \to F^{n}H^{*}_{\mathrm{dR}}(X_{K}/K). \end{equation} In fact, it is possible to construct a natural chain of quasi-isomorphisms connecting our version of the rigid syntomic complex with Besser's. But since we do not need this we omit the lengthy and technical details. \end{rem} We now give Besser's construction of the rigid syntomic regulator \cite[Thm. 7.5]{Besser} in the setup of the present paper. By \cite[Prop. 7.4]{Besser} and the discussion following it, for every positive integers $n, r$ there exists a class \begin{equation}\label{nov1501} \ch_{n,(r)}^{\mathrm{syn}} \in H^{2n}_{\mathrm{syn}}(B_{\dot}\GL_{r,R}, n), \end{equation} the universal $n$-th syntomic Chern character class, uniquely determined by the fact that it is mapped to the degree $2n$ component $\ch_{n,(r)}^\mathrm{dR}$ of the universal de Rham Chern character class in $F^{n}H^{2n}_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K)$ under \eqref{eq:MapSynDeRham}. These are compatible for varying $r$. Let $a=(\,.\,)_{k}\colon \mathsf{Sm}_{R} \to \mathsf{Sch}_{k}$ be the special fibre functor. We define the syntomic regulator datum $\omega^{\mathrm{syn}}_{n,(r)} := (\mathsf{Sch}_{k}, B_{\dot}\GL_{r,R} \xrightarrow{\id} B_{\dot}\GL_{r,R}, F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(\,.\,/K) \xrightarrow{-\eqref{eq:MapFilPhi}} \Phi(n)((\,.\,)_{k}), \ch_{n,(r)}^{\mathrm{syn}})$. Now let $X=\Spec(A)$ be an affine scheme in $\mathsf{Sm}_{R}$ and as always $X^{\dot} = X \times_{R} \Delta_{R}^{\dot}$. With these choices we have \begin{align*} & \left(B_\dot\GL_{r,R}\times_{a(B_{\dot}\GL_{r,R})} a(B_\dot\GL_{r,R})\right)(X^\dot) = B_\dot\GL_{r,R}(A[\Delta^\dot]), \\%&& \text{ cf. \eqref{nov0904},} \label{nov1201}\\ & \MF\left(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X^\dot) \to \Phi(n)(X_{k}^\dot)\right) = R\Gamma_\mathrm{syn}(X\times_{R}\Delta_{R}^\dot, n), \\ & \MF\left(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_\dot\GL_{r,R}) \to \Phi(n)(B_\dot\GL_{r,k})\right) = R\Gamma_\mathrm{syn}(B_\dot\GL_{r,R},n). \end{align*} Similarly as in Lemma \ref{nov1206} one shows that $R\Gamma_\mathrm{syn}(X,n) \to R\Gamma_\mathrm{syn}(X\times_{R}\Delta_{R}^\dot, n)$ is a quasi-isomorphism. \begin{dfn} For $i\geq 1$ the \emph{syntomic Chern character} or \emph{regulator} is given by \begin{multline*} \ch_{n,i}^{\mathrm{syn}}\colon K_i(X) = \varinjlim_{r} \pi_{i}\left(B_{\dot}GL_{r}(A[\Delta^{\dot}])\right) \xrightarrow{\varinjlim_{r}\reg_{i}(\omega^{\mathrm{syn}}_{n,(r)})} \\ H^{2n-i}(R\Gamma_\mathrm{syn}(X\times_{R}\Delta_{R}^\dot, n)) \cong H^{2n-i}_\mathrm{syn}(X,n). \end{multline*} \end{dfn} Using that the natural map $B_\dot\GL(A) \to \diag B_\dot\GL(A[\Delta^\dot])$ induces an isomorphism in homology with $\mathbf{Z}$-coefficients (cf.~the proof of Lemma \ref{mar2003}), it is easy to check that this construction is equivalent to Besser's in the affine case. Again, this is extended to all schemes in $\mathsf{Sm}_{R}$ using Jouanolou's trick (cf.~\ref{nov0601}). \subsection{The comparison} \begin{lemma}\label{nov0202} There exist complexes $\widetilde{R\Gamma}_\mathrm{rel}(X,n)$, functorial in $X\in \mathsf{Sm}_{R}$, together with maps \[ R\Gamma_\mathrm{rel}(X,n) \xleftarrow{\simeq} \widetilde{R\Gamma}_\mathrm{rel}(X,n) \to R\Gamma_\mathrm{syn}(X,n), \] the left pointing arrow being a quasi-isomorphism. These induce natural maps \begin{equation}\label{mar0701} H^{*}_{\mathrm{rel}}(X,n) \to H^{*}_{\mathrm{syn}}(X, n) \end{equation} which are isomorphisms if $X$ is proper and $*\not\in \{2n, 2n+1, 2n+2\}$. \end{lemma} \begin{proof} Consider the following diagram of complexes \begin{equation}\label{eq:MapRelSyn} \begin{split} \xymatrix@[email protected]{ F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \ar[d]^{\eqref{eq:CompDeRhamRigid}} & F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \ar[d]^{(\ref{eq:MapFilPhi}.1)} \ar[l]_-\simeq \ar@{=}[r] & F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \ar[d]^{-\eqref{eq:MapFilPhi}}\\ R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K) & \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X} \ar[l]_-{\simeq} \ar[r]^{-(\ref{eq:MapFilPhi}.2)} & \Phi(n)(X_{k}) } \end{split} \end{equation} The left square commutes up to canonical homotopy (cf. \ref{Nr:QuasiPullback}), the right square strictly commutes. We set $\widetilde{R\Gamma}_\mathrm{rel}(X,n) = \MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \to \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X})$. The desired maps are induced by the maps in the diagram together with the homotopy which makes the left hand square commute (cf. \ref{dez1201}). The second statement follows from weight considerations as in the proof of \cite[Prop. 8.6]{Besser}. \end{proof} We can now formulate the main result of this paper: \begin{thm}\label{thm:Comparison} For every $X$ in $\mathsf{Sm}_{R}$ and $i \geq 1$ the diagram \[ \xymatrix@C+0.5cm{ K_{i}^{\mathrm{rel}}(X) \ar[r] \ar[d]_{\ch_{n,i}^{\mathrm{rel}}} & K_{i}(X) \ar[d]^{\ch_{n,i}^{\mathrm{syn}}} \\ H^{2n-i}_{\mathrm{rel}}(X,n) \ar[r]^-{\eqref{mar0701}} & H^{2n-i}_{\mathrm{syn}}(X,n) } \] commutes. \end{thm} \begin{proof} By construction of the maps in the diagram, we may suppose that $X = \Spec(A)$ is affine. Write $A_{k} = A\otimes_{R} k$, so that $X_{k} = \Spec(A_{k})$. We split the diagram up into the following smaller diagrams, and show that every single one of them commutes. \begin{equation}\begin{split}\label{nov1502} \xymatrix@C+0.3cm{ K_i^\mathrm{rel}(X)\ar[d]_{\ch_{n,i}^{\mathrm{rel}}} \ar[dr]^{\widetilde\ch_{n,i}^\mathrm{rel}} \ar[r] & K_i(X,X_k) \ar[r] \ar[dr]^{\widetilde{\ch}_{n,i}^\mathrm{syn}} & K_i(X) \ar[d]^{\ch_{n,i}^\mathrm{syn}} \\ H^{2n-i}_\mathrm{rel}(X,n) & H^{2n-i}(\widetilde{R\Gamma}_\mathrm{rel}(X,n)) \ar[l]^-{\text{cf. \ref{nov0202}}}_-{\cong} \ar[r]_-{\text{cf. \ref{nov0202}}}& H^{2n-i}_\mathrm{syn}(X,n) } \end{split} \end{equation} Here $K_{i}(X, X_{k}) := \pi_{i}\left(B_{\dot}\GL(A[\Delta^{\dot}]) \times_{B_{\dot}\GL(A_{k}[\Delta^{\dot}])} E_{\dot}\GL(A_{k}[\Delta^{\dot}])\right)$. Since $A_{k}[\Delta^{\dot}] \cong A^{\dag}\<\Delta^{\dot}\>\otimes_{R} k$ the map $K_{i}^{\mathrm{rel}}(X) \to K_{i}(X)$ factors through $K_{i}(X,X_{k})$. All vertical resp. diagonal maps are induced by a compatible family of maps for each finite level $r$. We can thus restrict to a fixed finite level $r$. To ease notation we write $E_{\dot} := E_{\dot}\GL_{r,R}, B_{\dot}:=B_{\dot}\GL_{r,R}$. The diagonal maps arise as follows: Since the left hand square in \eqref{eq:MapRelSyn} commutes up to canonical homotopy we are in the situation of Lemma \ref{nov1301}. In particular, we have a natural isomorphism (cf. \eqref{eq:MapRelSyn}) \begin{multline*} H^{2n}(\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_{\dot}/K) \to \widetilde{R\Gamma}_{\mathrm{rig}}(E_{\dot,k}/K)_{\widehat E_{\dot}})) \xrightarrow{\cong} \\ H^{2n}\left(\MF(F^{n}R\Gamma_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \to R\Gamma_{\mathrm{dR}}(E_{\dot}\widehat\GL_{r,K}/K))\right) \end{multline*} and we define $\widetilde\ch^{\mathrm{rel}}_{n,(r)}$ to be the class mapping to $\ch^{\mathrm{rel}}_{n,(r)}$ (cf. \eqref{nov1401}) under this isomorphism. We let $\widetilde\ch_{n,i}^{\mathrm{rel}}$ be induced by the regulator data $\widetilde\omega^{\mathrm{rel}}_{n,(r)} = (\mathsf{Smf}_{R}, E_{\dot} \to B_{\dot}, F^n\widetilde{R\Gamma}_\mathrm{dR}(./K) \to \widetilde{R\Gamma}_\mathrm{rig}((.)_{k}/K)_{\widehat{(.)}}, \widetilde\ch^{\mathrm{rel}}_{n,(r)})$. Then it is clear from Lemma \ref{nov1301} and the constructions that the left triangle commutes. The map $\widetilde\ch^\mathrm{syn}_{n,i}$ is induced by the data $\widetilde\omega^\mathrm{syn}_{n,(r)} = (\mathsf{Sch}_k, E_\dot \to B_\dot, F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(./K) \xrightarrow{-\eqref{eq:MapFilPhi}} \Phi(n)((.)_{k}), \widetilde\ch_{n,(r)}^\mathrm{syn})$ where $\widetilde\ch_{n,(r)}^{\mathrm{syn}}$ is defined as follows: We have a commutative diagram \[ \[email protected]@C-0.3cm{ E_{\dot} \ar[r] \ar[d] & B_{\dot}\ar[d]^{\id}\\ B_{\dot} \ar[r]^-{\id} & B_{\dot} } \] which induces a map \begin{multline*} H^{2n}_{\mathrm{syn}}(B_{\dot}, n) \cong H^{2n}\left(\MF\left(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_\dot/K) \to \Phi(n)(B_{\dot,k})\right)\right) \to \\ H^{2n}\left(\MF\left(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_\dot/K) \to \Phi(n)(E_{\dot,k})\right)\right) \end{multline*} and $\widetilde\ch_{n,(r)}^{\mathrm{syn}}$ is by definition the image of $\ch_{n,(r)}^{\mathrm{syn}}$ (see \eqref{nov1501}) by this map. It is then clear from Lemma \ref{nov1302} that the right triangle in \eqref{nov1502} commutes. It remains to show that the middle parallelogram in \eqref{nov1502} commutes. For this we apply Lemma \ref{nov1301} to the regulator data $\widetilde\omega^{\mathrm{rel}}_{n}$ and $\widetilde\omega^{\mathrm{syn}}_{n}$: The special fibre functor $\mathsf{Sm}_{R} \to \mathsf{Sch}_{k}$ factors naturally as $\mathsf{Sm}_{R}\xrightarrow{X\mapsto \widehat X} \mathsf{Smf}_{R} \xrightarrow{\mathscr X\mapsto \mathscr X_{k}} \mathsf{Sch}_{k}$. Moreover, we have a natural transformation between functors $\mathsf{Smf}_{R}^{\mathrm{op}}\to \mathsf{Ch}, \widetilde{R\Gamma}_{\mathrm{rig}}((.)_{k}/K)_{(.)} \to \Phi(n)((.)_{k})$ given by $-\eqref{eq:MapFilPhi}$. Since \[ \xymatrix@[email protected]{ F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \ar[d]^{(\ref{eq:MapFilPhi}.1)} \ar@{=}[r] & F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(X/K) \ar[d]^{-\eqref{eq:MapFilPhi}}\\ \widetilde{R\Gamma}_{\mathrm{rig}}(X_{k}/K)_{\widehat X} \ar[r]^{-(\ref{eq:MapFilPhi}.2)} & \Phi(n)(X_{k}) } \] commutes for every $X\in \mathsf{Sm}_{R}$, we have a natural map \begin{multline}\label{nov1503} H^{2n}\left(\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_{\dot}/K) \to \widetilde{R\Gamma}_{\mathrm{rig}}(E_{\dot,k}/K)_{\widehat E_{\dot}})\right) \to \\ H^{2n}\left(\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_\dot/K) \to \Phi(n)(E_{\dot,k}))\right). \end{multline} If we show that under this map $\widetilde\ch_{n,(r)}^{\mathrm{rel}}$ maps to $\widetilde\ch_{n,(r)}^{\mathrm{syn}}$ then Lemma \ref{nov1301} implies the desired commutativity. But indeed, \eqref{nov1503} fits in a commutative diagram \[ \xymatrix@[email protected]{ H^{2n}\left(\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_{\dot}/K) \to \widetilde{R\Gamma}_{\mathrm{rig}}(E_{\dot,k}/K)_{\widehat E_{\dot}})\right) \ar[r]^-{\cong} \ar[d]_{\eqref{nov1503}} & H^{2n}( F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_{\dot}/K))\\ H^{2n}\left(\MF(F^{n}\widetilde{R\Gamma}_{\mathrm{dR}}(B_\dot/K) \to \Phi(n)(E_{\dot,k}))\right) \ar[ur]_-{\cong} & F^{n}H^{2n}_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \ar@{=}[u] } \] where the two isomorphisms are established as in Lemma \ref{nov1204} and, by the constructions, both, $\widetilde\ch_{n,(r)}^{\mathrm{rel}}$ and $\widetilde\ch_{n,(r)}^{\mathrm{syn}}$, map to $\ch_{n,(r)}^{\mathrm{dR}}$ on the right hand side. \end{proof} \subsection{Applications}\label{ssec:applications} In this section we assume that the residue field $k$ of $R$ is finite. Let $X$ be a smooth and proper $R$-scheme. The \'etale Chern character class induces a map $\ch_{n,i}^{\mathrm{\acute{e}t}}\colon K_{i}(X) \to H^{2n-i}_{\mathrm{\acute{e}t}}(X_{K}, \mathbf{Q}_{p}(n))$. It follows from the crystalline Weil conjectures \cite{ChiarellottoLeStum} and Faltings' crystalline comparison theorem \cite{FaltingsCrystalline} that $H^{2n-i}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))^{G_{K}}=0$ for $i>0$, where $\ol K$ denotes an algebraic closure of $K$ and $G_{K}=\Gal(\ol K/K)$. Hence the Hochschild-Serre spectral sequence for $X_{\ol K} \to X_{K}$ induces an edge morphism $H^{2n-i}_{\mathrm{\acute{e}t}}(X_{K}, \mathbf{Q}_{p}(n)) = F^{1}H^{2n-i}_{\mathrm{\acute{e}t}}(X_{K}, \mathbf{Q}_{p}(n)) \to H^{1}(G_{K}, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n)))$ and the composition \begin{equation} r_{p}\colon K_{i}(X) \xrightarrow{\ch_{n,i}^{\mathrm{\acute{e}t}}} H^{2n-i}_{\mathrm{\acute{e}t}}(X_{K}, \mathbf{Q}_{p}(n)) \xrightarrow{\text{edge}} H^{1}(G_{K}, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))) \end{equation}• is the \emph{\'etale $p$-adic regulator}. According to the de Rham comparison theorem \cite{FaltingsCrystalline} we have an isomorphism of filtered vector spaces \[ D_{\mathrm{dR}}(H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))) \cong H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)(n) \] where the twist by $n$ on the right hand side only shifts the filtration. Hence \[ D_{\mathrm{dR}}(H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n)))/F^{0} \cong H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n} \cong H^{2n-i}_{\mathrm{rel}}(X,n) \] (see Remark \ref{nov0203}(iii)). In particular, the Bloch-Kato exponential for the $G_{K}$-re\-pre\-sen\-ta\-tion $H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))$ is a map \[ \exp\colon H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n} \to H^{1}(G_{K}, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))) \] and from Nizio\l's work we get the \begin{cor}\label{apr1201} For each smooth, projective $R$-scheme $X$ the diagram \[ \xymatrix@C+0.5cm{ K_{i}^{\mathrm{rel}}(X) \ar[d]^{\ch_{n,i}^{\mathrm{rel}}} \ar[r] & K_{i}(X) \ar[d]^{r_{p}}\\ H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n} \ar[r]^-{\exp} & H^{1}(G_{K}, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))) } \] commutes. \end{cor} \begin{proof} It follows from Nizio\l's work \cite{NiziolImg, NiziolCrys} and the comparison with Besser's syntomic cohomology \cite[Prop. 9.9]{Besser} that there is a natural map $H^{*}_{\mathrm{syn}}(X,n) \to H_{\mathrm{\acute{e}t}}^{\ast}(X_{K}, \mathbf{Q}_{p}(n))$ which is compatible with Chern classes \cite[Cor. 9.10]{Besser}. By \cite[Prop. 9.11]{Besser} the composition $H^{2n-i-1}_{\mathrm{dR}}(X_{K}/K)/F^{n} \to H^{2n-i}_{\mathrm{syn}}(X,n) \to H^{2n-i}_{\mathrm{\acute{e}t}}(X_{K}, \mathbf{Q}_{p}(n)) \to H^{1}(G_{K}, H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n)))$ is the Bloch-Kato exponential for $H^{2n-i-1}_{\mathrm{\acute{e}t}}(X_{\ol K}, \mathbf{Q}_{p}(n))$. Hence the claim follows from the comparison of the relative Chern character with the syntomic regulator in Theorem \ref{thm:Comparison}. \end{proof} \begin{rem}\label{nov1504} As in the proof of Lemma \ref{lemma:RelKSmoothSeparated} it follows from Calvo's Theorem \cite{Calvo} that $K_{\top}^{-i}(X) \cong K_{i}(X_{k})$. For $X_{k}$ smooth and projective these groups are conjectured to be torsion (Parshin's conjecture). This would imply that $K_{i}^{\mathrm{rel}}(X) \to K_{i}(X)$ is rationally an isomorphism. It follows from \cite{Harder} and \cite{QuillenK} that this conjecture is true for $\dim X_{k} \leq 1$. \end{rem} From the previous corollary together with our earlier work \cite{TammeBorel, TammeDiss}, we get a new proof of the main result of \cite{HK}: Huber and Kings introduce the \emph{$p$-adic Borel regulator} $r_{\mathrm{Bo},p}\colon K_{2n-1}(R) \to K$ by imitating the construction of the classical Borel regulator for the field of complex numbers, replacing the van Est isomorphism by the Lazard isomorphism. \begin{cor}[Huber-Kings \cite{HK}] \label{cor:HK} Let $K$ be a finite extension of $\mathbf{Q}_{p}$. The diagram \[ \xymatrix{ K_{2n-1}(R) \ar[d]_{\frac{(-1)^{n}}{(n-1)!}r_{\mathrm{Bo},p}} \ar[dr]^{r_{p}}\\ K=D_{\mathrm{dR}}(\mathbf{Q}_{p}(n)) \ar[r]^-{\exp} & H^{1}(G_{K}, \mathbf{Q}_{p}(n)) } \] commutes. \end{cor} \begin{rem*} The factor $\frac{(-1)^{n}}{(n-1)!}$ appears since Huber and Kings use Chern classes in the normalization of both, the \'etale and the $p$-adic Borel regulator, whereas we used Chern character classes in the definition of the \'etale regulator. \end{rem*} \begin{proof} Apply the previous corollary with $X=\Spec(R)$, $i=2n-1$. As mentioned in Remark \ref{nov1504}, $K_{2n-1}^{\mathrm{rel}}(R) \to K_{2n-1}(R)$ is rationally an isomorphism. In Theorem \ref{thm:KaroubisRelChern} we will show that the present version of the relative Chern character coincides with Karoubi's original construction. For this we showed in \cite[Corollary 7.23]{TammeDiss} that the diagram \[ \[email protected]{ K^{\mathrm{rel}}_{2n-1}(R) \ar[r] \ar[d]_{\ch_{n,2n-1}^{\mathrm{rel}}} & K_{2n-1}(R) \ar[dl]^{\frac{(-1)^{n}}{(n-1)!}r_{\mathrm{Bo},p}} \\ K } \] commutes. \end{proof} \section{Comparison with Karoubi's original construction} \label{sec:Karoubi} As before $R$ denotes a complete discrete valuation ring with field of fractions $K$ of characteristic $0$ and residue field $k$ of characteristic $p>0$. In this section we compare the relative Chern character of Section \ref{sec:Construction} with Karoubi's original construction \cite{KarCR, Kar87} in the form of \cite{TammeDiss}. For any smooth, affine $R$-scheme $X=\Spec(A)$ this is a homomorphism \[ \ch_{n,i}^{\mathrm{Kar}}\colon K_{i}^{\mathrm{rel}}(X) \to \mathbb{H}^{2n-i-1}(\widehat X_{K}, \Omega^{<n}_{\widehat X_{K}/K}) = H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})). \] We recall the main steps in its construction. \begin{Nr} We begin with some preliminaries concerning integration (see also \cite[Appendix]{TammeBorel}). Consider the polynomial ring $\mathbf{Q}[\Delta^{p}] = \mathbf{Q}[x_{0},\dots, x_{n}]/(\sum_{i}x_{i}-1)$. There is a well defined integration map $\int_{\Delta^{p}}\colon \Omega^{p}_{\mathbf{Q}[\Delta^{p}]/\mathbf{Q}} \to \mathbf{Q}$ sending an algebraic $p$-form $\omega$ to the integral of $\omega$, considered as a smooth $p$-form, over the real standard simplex $\mathbf \Delta^{p}:=\{(x_{0}, \dots, x_{p}) \in \mathbf{R}^{p+1}\,|\,\sum_{i} x_{i} = 1, \forall i: 0\leq x_{i}\leq 1\} \subseteq \mathbf{R}^{p+1}$ with orientation given by the form $dx_{1}\cdots dx_{p}$. This integral is in fact a rational number. It satisfies Stokes's formula \begin{equation}\label{mar2006} \int_{\Delta^{p}} d\omega = \sum_{i=0}^{p} (-1)^{i} \int_{\Delta^{p-1}} (\partial^{i})^{*}\omega. \end{equation} Similarly one can define an integration map $\int_{\Delta^{p}\times\Delta^{q}}\colon \Omega^{p+q}_{\mathbf{Q}[\Delta^{p}]\otimes_{\mathbf{Q}}\mathbf{Q}[\Delta^{q}]/\mathbf{Q}} \to \mathbf{Q}$, again given by the usual integration over the product of real simplices $\mathbf \Delta^{p}\times \mathbf \Delta^{q}$. There is a canonical decomposition of $\mathbf \Delta^{p}\times \mathbf \Delta^{q}$ into copies of the standard $p+q$-simplex $\mathbf \Delta^{p+q}$, indexed by all $(p,q)$-shuffles $(\mu,\nu)$ (cf.~the proof of Theorem \ref{thm:KaroubisRelChern} below). It follows from this and the analogous formula for smooth differential forms that \begin{equation}\label{mar2007} \int_{\Delta^{p}\times\Delta^{q}} \omega = \sum_{(\mu,\nu)} \sgn(\mu,\nu) \int_{\Delta^{p+q}} (\mu,\nu)^{*}\omega, \end{equation} where the sum runs over all $(p,q)$-shuffles and $(\mu,\nu)^{*}\omega\in \Omega^{p+q}_{\mathbf{Q}[\Delta^{p+q}]/\mathbf{Q}}$ is the pullback of $\omega \in \Omega^{p+q}_{\mathbf{Q}[\Delta^{p}]\otimes_{\mathbf{Q}}\mathbf{Q}[\Delta^{q}]/\mathbf{Q}}$ to the simplex corresponding to the shuffle $(\mu,\nu)$. Tensoring the $\mathbf{Q}$-linear integration map with $K$, we get $\int_{\Delta^{p}}\colon \Omega^{p}_{K[\Delta^{p}]/K} \to K$. More generally, for a $K$-algebra $A$ we can define an integration map $\int_{\Delta^{p}}\colon \Omega^{n}_{A[\Delta^{p}]/K} \to \Omega^{n-p}_{A/K}$ using the decomposition $\Omega^{n}_{A[\Delta^{p}]/K} \cong \bigoplus_{k+l=n} \Omega^{k}_{A/K} \otimes_{K} \Omega^{l}_{K[\Delta^{p}]/K}$. If $A$ is a $K$-dagger algebra one can show similarly as in \cite[Appendix]{TammeBorel} that by continuity this extends uniquely to a map \[ \int_{\Delta^{p}}\colon \Omega^{n}_{A\<\Delta^{p}\>^{\dag}/K,\mathrm f} \to \Omega^{n-p}_{A/K, \mathrm f}, \] where the subscript $\mathrm f$ indicates that we consider differential forms for dagger algebras (cf. \ref{mar2005}). The analogues of \eqref{mar2006} and \eqref{mar2007} remain valid. \end{Nr} \begin{Nr} Let $Y_{\dot}$ be a simplicial dagger space. A \emph{simplicial} $n$-form $\omega$ on $Y_{\dot}$ is a collection of $n$-forms $\omega_{p}$ on $Y_{p} \times \widehat\Delta^{p}_{K}$, $p\geq 0$, satisfying $(1\times\phi_{\Delta})^{*}\omega_{p} = (\phi_{X}\times 1)^{*}\omega_{q}$ on $Y_{p}\times\widehat\Delta^{q}_{K}$ for all monotone maps $\phi\colon [q]\to [p]$, $\phi_{\Delta},\phi_{X}$ denoting the induced (co)simplicial structure maps. We will often denote $\omega_{p}$ by $\omega|_{Y_{p}\times\widehat\Delta^{p}_{K}}$. The space of all simplicial $n$-forms is denoted by $D^{n}(Y_{\dot})$. Applying the wedge product and exterior differential component-wise makes $D^{*}(Y_{\dot})$ into a commutative differential graded algebra. For a $K$-dagger space $X$ we write $\Omega^{*}(X)$ for the complex of global sections $\Gamma(X, \Omega^{*}_{X/K})$. Define $\Omega^{*}(Y_{\dot})$ as the total complex of the cosimplicial complex $[p]\mapsto \Omega^{*}(Y_{p})$. By Dupont's Theorem \cite[Thm. 5.6]{TammeDiss} \begin{equation}\label{eq:DefI} I\colon D^{*}(Y_{\dot}) \to \Omega^{*}(Y_{\dot}), \quad \omega \mapsto \sum_{k} \int_{\Delta^{k}} \omega|_{Y_{k}\times\widehat\Delta^{k}_{K}} \end{equation} is a quasi-isomorphism. We have a decomposition $\Omega^n_{X\times\widehat\Delta^k_{K}/K} \cong \bigoplus_{p+q=n} \Omega^{p,q}_{X\times\widehat\Delta^k_{K}/K}$ where $\Omega^{p,q}_{X\times\widehat\Delta^k_{K}/K} := \operatorname{pr}_X^*\Omega^p_{X/K} \otimes \operatorname{pr}_{\widehat\Delta^k_{K}}^*\Omega^q_{\widehat\Delta^k_{K}/K}$ and hence the filtration $F_X^\dot\Omega^*(X\times\widehat\Delta^k_{K})$ with respect to the first degree. This induces a filtration $F^\dot$ on $D^{*}(Y_{\dot})$ and $I$ is a filtered quasi-isomorphism when $\Omega^{*}(Y_{\dot})$ carries the filtration given by $F^{n}\Omega^{*}(Y_{\dot}) = \Omega^{\geq n}(Y_{\dot})$. \end{Nr} \begin{Nr} Let $X$ be an affine scheme in $\mathsf{Sm}_{R}$ with generic fibre $X_{K}$. Recall the complexes $F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K)$ from \eqref{eq:FilDRCompl}. Similarly as in \eqref{eq:GdCompMap}, \eqref{eq:ComplCompMap} there is a natural map \begin{equation}\label{eq:KarCompMap1} F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to \Gamma(X_{K}, \Gd_{Pt(X_{K}^{\dag})\sqcup Pt(X_{K})} \Omega^{\geq n}_{X_{K}/K}) \end{equation} where the complex on the right computes the hypercohomology of $\Omega^{\geq n}_{X_{K}/K}$ on $X_{K}$. Since $X_{K}$ is affine the sheaves $\Omega^{i}_{X_{K}/K}$ are acyclic and hence the coaugmentation \begin{equation}\label{eq:KarCompMap2} \Gamma(X_{K}, \Gd_{Pt(X_{K}^{\dag})\sqcup Pt(X_{K})} \Omega^{\geq n}_{X_{K}/K}) \leftarrow \Gamma(X_{K}, \Omega^{\geq n}_{X_{K}/K}) \end{equation} is a quasi-isomorphism. The maps \eqref{eq:KarCompMap1}, \eqref{eq:KarCompMap2}, and \[ \Gamma(X_{K}, \Omega^{\geq n}_{X_{K}/K}) \to \Gamma(X_{K}^{\dag}, \Omega^{\geq n}_{X_{K}^{\dag}/K}) \to \Gamma(\widehat X_{K}, \Omega^{\geq n}_{\widehat X_{K}/K}) \] give a chain of morphisms connecting $F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K)$ with $\Gamma(\widehat X_{K}, \Omega^{\geq n}_{\widehat X_{K}/K})$ where the arrow \eqref{eq:KarCompMap2} pointing in the wrong direction is a quasi-isomorphism. All these morphisms are functorial in $X\in \mathsf{Sm}_{R}$. We denote this by \begin{equation}\label{eq:KarCompMap3} F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to\xleftarrow{\simeq} \Gamma(\widehat X_{K}, \Omega^{\geq n}_{\widehat X_{K}/K}). \end{equation} If $Y_{\dot}$ is a simplicial smooth, affine $R$-scheme this and the quasi-isomorphism $I$ from \eqref{eq:DefI} give a natural chain of morphisms \begin{equation}\label{eq:KarCompMap4} F^{n}R\Gamma_{\mathrm{dR}}(Y_{K,\dot}/K) \to\xleftarrow{\simeq} F^{n}D^{*}(\widehat Y_{K,\dot}) \end{equation} \end{Nr} We apply \eqref{eq:KarCompMap4} to $B_{\dot}\GL_{r,R}$ to get a map \begin{equation}\label{eq:MapDeRhamDupont} F^{n}H^{2n}_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \to H^{2n}(D^{*}(B_{\dot}\widehat\GL_{r,K})). \end{equation} Similarly as in Lemma \ref{nov1204} there is an isomorphism \begin{equation}\label{mar0801} H^{2n}\left(\MF(F^{n}D^{*}(B_{\dot}\widehat\GL_{r,K}) \to D^{*}(E_{\dot}\widehat\GL_{r,K}))\right) \xrightarrow{\cong} H^{2n}(D^{*}(B_{\dot}\widehat\GL_{r,K})) \end{equation} and we define \begin{equation}\label{eq:DefChnKar} \ch_{n,(r)}^{\mathrm{Kar}} \in H^{2n}\left(\MF(F^{n}D^{*}(B_{\dot}\widehat\GL_{r,K}) \to D^{*}(E_{\dot}\widehat\GL_{r,K}))\right)\end{equation} to be the unique element whose image under \eqref{mar0801} coincides with the image of $\ch_{n,(r)}^{\mathrm{dR}}$ under \eqref{eq:MapDeRhamDupont}. \begin{rem*} Originally, Karoubi used Chern-Weil theory to construct the relevant characteristic classes. This is the reason for the use of the dga $D^{*}(B_{\dot}\widehat\GL_{K})$. One advantage is that it gives more explicit formulas for the relative Chern character. For instance, these are used for the comparison theorem in \cite{TammeBorel} and in the work of Choo and Snaith \cite{ChooSnaith}. It was checked in \cite[Prop. 5.13]{TammeDiss} that the present approach yields the same classes as the Chern-Weil theoretic one. \end{rem*} To describe Karoubi's version of the relative Chern character we first need a Lemma. Let $X=\Spec(A)$ be a regular, affine $R$-scheme. Define the bisimplicial set $F^\flat(A) := B_\dot\GL(A) \times_{B_\dot\GL(A^\dag\<\Delta^\dot\>)} E_\dot\GL(A^\dag\<\Delta^\dot\>)$ and similarly $F^{\flat}_{r}(A)$ for every finite level $r$. These are bisimplicial subsets of $F(A)$, respectively $F_{r}(A)$, from \eqref{nov0904} and \eqref{dez1202}. \begin{lemma}\label{mar2003} The induced map on complexes $\mathbf{Z} F^\flat(A) \to \mathbf{Z} F(A)$ is a quasi-isomorphism. \end{lemma} \begin{proof} By the Eilenberg-Zilber theorem it suffices to show that $\mathbf{Z}\diag F^{\flat}(A) \to \mathbf{Z}\diag F(A)$ is a quasi-isomorphism. For this it is enough to show that the map $\diag F^{\flat}(A) \to \diag F(A)$ is acyclic. By the definitions we have a pullback square \begin{equation}\label{mar0802}\begin{split} \[email protected]@R-0.2cm{ \diag F^{\flat}(A) \ar[r]\ar@{->>}[d] & \diag F(A) \ar@{->>}[d] \\ B_{\dot}\GL(A) \ar[r] & \diag B_{\dot}\GL(A[\Delta^{\dot}]). } \end{split} \end{equation} Since $\diag E_{\dot}\GL(A^{\dag}\<\Delta^\dot\>) \to \diag B_{\dot}\GL(A^{\dag}\<\Delta^\dot\>)$ is a Kan fibration, so are the vertical maps in \eqref{mar0802}. It thus suffices to show that the lower horizontal map is acyclic (cf.~\cite[(4.1)]{Berrick}). Let $(\,.\,)^{+}$ denote Quillen's plus construction. Then $\pi_{i}(B_{\dot}\GL(A)^{+})=K_{i}(A), i\geq 1$, are the Quillen K-groups of $A$. The map $B_{\dot}\GL(A) \to B_{\dot}\GL(A)^{+}$ is acyclic. We have a commutative diagram \[ \[email protected]{ B_{\dot}\GL(A) \ar[r] \ar[d]_{\text{acyclic}} & \diag B_{\dot}\GL(A[\Delta^{\dot}]) \ar[d]^{\simeq}\\ B_{\dot}\GL(A)^{+} \ar[r]^-{\simeq} & \diag B_{\dot}\GL(A[\Delta^{\dot}])^{+} } \] where the lower horizontal map is a weak equivalence by the homotopy invariance for K-theory of regular rings, and the right vertical map is a weak equivalence since $B_{\dot}\GL(A[\Delta^{\dot}])$ has the homotopy type of an H-space. Hence the upper horizontal map is acyclic, as desired. \end{proof} Karoubi's relative Chern character $\ch_{n,i}^{\mathrm{Kar}}$ is the composition of the Hurewicz map \begin{equation}\label{mar1801} K_{i}^{\mathrm{rel}}(X) \to H_{i}(\diag F(A), \mathbf{Z}) \cong H_{i}(\diag F^\flat(A), \mathbf{Z}) \cong \varinjlim_{r} H_{i}(\diag F^{\flat}_{r}(A),\mathbf{Z}) \end{equation} with a map induced by a system of maps $H_{i}(\diag F^\flat_{r}(A), \mathbf{Z}) \to H^{2n-i-1}(\Omega^{<n}(\widehat X_{K}))$ for varying $r$ constructed as follows (cf. \cite[Remark 3.6(ii)]{TammeDiss}): An $i$-simplex $\sigma$ in $\diag F^\flat_{r}(A)$ defines a pair of morphisms $\widehat X_{K}\xrightarrow{\sigma_0} B_{i}\widehat\GL_{r,K}$ and $\widehat X_{K} \times \widehat\Delta^{i}_{K} \xrightarrow{\sigma_1} E_{i}\widehat\GL_{r,K}$ such that \[ \xymatrix{ \widehat X_{K} \times \widehat \Delta^{i}_{K} \ar[r]^-{\sigma_1} \ar[d]_{\mathrm{proj.}} & E_{i}\widehat\GL_{r,K}\ar[d]^{p} \\ \widehat X_{K} \ar[r]^-{\sigma_0} & B_{i}\widehat\GL_{r,K} } \quad\text{and hence} \xymatrix@C+0.5cm{ & E_{i}\widehat\GL_{r,K}\times\widehat\Delta^{i}_{K} \ar[d]^{p\times\id_{\Delta^{i}}} \\ \widehat X_{K} \times \widehat\Delta^{i}_{K} \ar[ur]^{(\sigma_1,\pr_{\Delta^{i}})} \ar[r]^-{\sigma_0\times\id_{\Delta^{i}}} & B_{i}\widehat\GL_{r,K} \times\widehat\Delta^{i}_{K} } \] commute. We can write \begin{equation}\label{mar2004} \ch_{n,(r)}^{\mathrm{Kar}} = (\omega_0, \omega_1) \end{equation} with $\omega_0 \in F^nD^{2n}(B_\dot\widehat\GL_{r,K}), \omega_1\in D^{2n}(E_\dot\GL_{r,K})$ and $d\omega_1=p^*\omega_0$. Then \begin{multline*} ((\sigma_0\times\id_{\Delta^i})^*\omega_0, (\sigma_1,\pr_{\Delta^i})^*\omega_1) :=\\ ((\sigma_0\times\id_{\Delta^i})^*\omega_0|_{E_{i}\widehat\GL_{r,K}\times\widehat\Delta^{i}_{K}}, (\sigma_1,\pr_{\Delta^i})^*\omega_1|_{E_{i}\widehat\GL_{r,K}\times\widehat\Delta^{i}_{K}}) \end{multline*} is a cycle of degree $2n$ in $\MF\left(F^n_{\widehat X_K}\Omega^{*}(\widehat X_{K}\times\widehat\Delta^{i}_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{i}_{K})\right)$. We now integrate along $\Delta^{i}$ and use the quasi-isomorphism $\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K})) \xrightarrow{\sim} \Omega^{<n}(\widehat X_{K})[-1]$ induced by the projection to the second component to get the map \[ F^\flat_{r}(A) \ni \sigma \mapsto \int_{\Delta^i} (\sigma_{1},\pr_{\Delta^{i}})^*\omega_{1} \in \Omega^{<n}(\widehat X_{K})^{2n-i-1}. \] This induces a well defined homomorphism \begin{equation}\label{eq:FastKaroubisRelChern} H_{i}(\diag F^\flat_{r}(A), \mathbf{Z}) \to H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})) \end{equation} compatible with varying $r$. By definition Karoubi's relative Chern character is the composition \begin{equation}\label{eq:KaroubisRelChern} \ch_{n,i}^{\mathrm{Kar}}\colon K_{i}^{\mathrm{rel}}(X) \xrightarrow{\eqref{mar1801}} \varinjlim_{r} H_{i}(\diag F^{\flat}_{r}(A), \mathbf{Z}) \xrightarrow{\eqref{eq:FastKaroubisRelChern}} H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})). \end{equation} To compare this with our version of the relative Chern character we need the following \begin{lemma}\label{lemma:ComparisonMap} Let $X$ be a smooth, affine $R$-scheme. There is a natural map \begin{equation}\label{eq:ComparisonMap} H^{*}_{\mathrm{rel}}(X,n) \to H^{*}\left(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}))\right) \cong H^{*-1}(\Omega^{<n}(\widehat X_{K})). \end{equation} \end{lemma} \begin{proof} We have $H^{*}_{\mathrm{rel}}(X,n) = \MF(F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K))$. Since $X$ is affine we have the natural chain of morphisms $F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to\xleftarrow{\simeq} \Omega^{\geq n}(\widehat X_{K})$ from \eqref{eq:KarCompMap3} and similarly $R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K) \xleftarrow{\simeq} \Omega^{*}(\widehat X_{K})$. These together induce the desired chain of maps \begin{equation*} \MF\left(F^{n}R\Gamma_{\mathrm{dR}}(X_{K}/K) \to R\Gamma_{\mathrm{dR}}(\widehat X_{K}/K)\right) \to \xleftarrow{\simeq} \MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K})). \qedhere \end{equation*} \end{proof} \begin{thm}\label{thm:KaroubisRelChern} Let $X$ be a smooth, affine $R$-scheme. The composition of the relative Chern character $\ch_{n,i}^{\mathrm{rel}}$ of Definition \ref{nov1601} with the comparison map \eqref{eq:ComparisonMap} coincides with Karoubi's relative Chern character $\ch_{n,i}^{\mathrm{Kar}}$ \eqref{eq:KaroubisRelChern}. \end{thm} \begin{proof} Let $X=\Spec(A)$. Using the quasi-isomorphism $\mathbf{Z} F^\flat(A) \xrightarrow{\simeq} \mathbf{Z} F(A)$ we can describe the composition \begin{equation}\label{eq:RelChernNaive0} K_{i}^{\mathrm{rel}}(X) \xrightarrow{\ch_{n,i}^{\mathrm{rel}}} H^{2n-i}_{\mathrm{rel}}(X, n) \xrightarrow{\eqref{eq:ComparisonMap}} H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})) \end{equation} explicitly as follows: We have a natural map (cf. the construction in Lemma \ref{lemma:ComparisonMap}) \begin{multline*} H^{2n}\left(\MF(F^{n}R\Gamma_{\mathrm{dR}}(B_{\dot}\GL_{r,K}/K) \to R\Gamma_{\mathrm{dR}}(E_{\dot}\widehat\GL_{r,K}))\right) \to \\ H^{2n}\left(\MF(\Omega^{\geq n}(B_{\dot}\widehat\GL_{r,K}) \to \Omega^{*}(E_{\dot}\widehat\GL_{r,K}))\right), \end{multline*} and we denote the image of $\ch_{n,(r)}^{\mathrm{rel}}$ (see \eqref{nov1401}) under this map by \[ \widetilde{\ch}_{n,(r)}^{\mathrm{rel}} \in H^{2n}(\MF(\Omega^{\geq n}(B_\dot\widehat\GL_{r,K}) \to \Omega^*(E_\dot\widehat\GL_{r,K}))). \] As in \eqref{nov0801} we have a map \begin{equation}\label{nov1603} H_i(\Tot\mathbf{Z} F^\flat_{r}(A)) \to \Hom\binom{H^{2n}(\MF(\Omega^{\geq n}(B_{\dot}\widehat\GL_{r,K}) \to \Omega^{*}(E_{\dot}\widehat\GL_{r,K}))),\qquad}{\qquad H^{2n-i}(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K})))}, \end{equation} that we can compose with the evaluation $\mathrm{eval}_{\widetilde{\ch}_{n,(r)}^{\mathrm{rel}}}$ to get \begin{equation}\label{mar2001} H_i(\Tot\mathbf{Z} F^\flat_{r}(A)) \to H^{2n-i}(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K}))). \end{equation} Since $\widehat X_{K} \times \widehat\Delta^{p}_{K}$ is an affinoid dagger space, hence the higher cohomology of coherent sheaves on $\widehat X_{K} \times \widehat\Delta^{p}_{K}$ vanishes, the argumentation of Lemma \ref{nov1206} shows that the coaugmentation $\Omega^{*}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K})$ is a quasi-isomorphism. It induces an isomorphism \begin{equation}\label{mar2002} H^{2n-i}\left(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K}))\right) \cong H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})). \end{equation} Now \eqref{eq:RelChernNaive0} equals the composition of \begin{multline} K_{i}^{\mathrm{rel}}(X)=\pi_i(F(A)) \xrightarrow{\text{Hurewicz}} H_i(\mathbf{Z}\diag F(A)) \overset{\text{\ref{mar2003}}}{\cong} \\ H_i(\mathbf{Z}\diag F^\flat(A)) \overset{\text{Eilenberg-Zilber}}{\cong} H_i(\Tot \mathbf{Z} F^\flat(A)) \cong \varinjlim_{r} H_{i}(\Tot \mathbf{Z} F^{\flat}_{r}(A)) \end{multline} with the map induced on the direct limit by \eqref{mar2001} and \eqref{mar2002}. From the definitions of $\ch_{n,(r)}^\mathrm{rel}$ in \eqref{nov1401} and $\ch_{n,(r)}^{\mathrm{Kar}}$ in \eqref{eq:DefChnKar} it is clear that $\widetilde{\ch}_{n,(r)}^{\mathrm{rel}} = I(\ch_{n,(r)}^{\mathrm{Kar}})$ where we still denote by $I$ the isomorphism induced by $I$ (see \eqref{eq:DefI}) on the cohomology of the respective mapping fibres. Thus we have to show that \begin{equation}\label{eq:diagKar \xymatrix@C+0.5cm{ H_{i}(\Tot \mathbf{Z} F^\flat_{r}(A)) \ar[d]_-{\eqref{nov1603}} & H_{i}(\mathbf{Z}\diag F^\flat_{r}(A)) \ar[l]_{\text{Eilenberg-Zilber}}^{\cong} \ar[dd]^{\eqref{eq:FastKaroubisRelChern}}\\ \Hom\binom{H^{2n}(\MF(\Omega^{\geq n}(B_{\dot}\widehat\GL_{r,K}) \to \Omega^{*}(E_{\dot}\widehat\GL_{r,K}))),}{H^{2n-i}(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K})))} \ar[d]_{\mathrm{eval}_{I(\ch_{n}^{\mathrm{Kar}})}} \\ H^{2n-i}(\MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K}\times\widehat\Delta^{\dot}_{K}))) \ar[r]_-{\cong}^-{\eqref{mar2002}} & H^{2n-i-1}(\Omega^{<n}(\widehat X_{K})) \end{equation} commutes. The following lemma shows that, on the level of complexes, \eqref{mar2002} is given by sending an element of the mapping fibre $(\eta_{0},\eta_{1})$ to the integral of $\eta_{1}$ along the simplex $\Delta^\dot$. \begin{lemma}\label{lemma:ExplQuInv} A quasi-inverse of $\Omega^{*}(\widehat X_{K}) \xrightarrow{\simeq} \Omega^{*}(\widehat X_{K}\times \widehat\Delta^{\dot}_{K})$ is given by integration over the standard simplices: \[ \Omega^{k+p}(\widehat X_{K}\times \widehat\Delta^{p}_{K}) \to \Omega^{k}(\widehat X_{K}), \quad \omega \mapsto (-1)^{p(p-1)/2}\int_{\Delta^{p}} \omega \] \end{lemma} \begin{proof} The integration map is obviously left inverse to the inclusion of $\Omega^{*}(\widehat X_{K})$ as the zeroth column in the double complex $\Omega^{*}(\widehat X_{K} \times \widehat\Delta^{\dot}_{K})$. So we only have to check that the integration indeed defines a morphism of complexes. This is straight forward using the following relative version of Stokes' formula \[ \int_{\Delta^{p}} d\omega = (-1)^{p}d\int_{\Delta^{p}} \omega + \sum_{i=0}^{p} (-1)^{i} \int_{\Delta^{p-1}} (\partial^{i})^{*}\omega \] and keeping in mind that the total differential on $\Omega^{q}(\widehat X_{K}\times\widehat\Delta^{p}_{K})$ is given by $\omega \mapsto (-1)^{p}d\omega + \sum_{i=0}^{p}(-1)^{i}(\partial^{i})^{*}\omega$. \end{proof} To show the commutativity of diagram \eqref{eq:diagKar}, we start with a bisimplex $\sigma \in F^\flat_{r}(A)_{p,i-p}$ so that the degree of $\sigma$ in $\Tot\mathbf{Z} F^\flat_{r}(A)$ is $i$. First we compute its image in $\Omega^{<n}(\widehat X_{K})$ going counterclockwise. As before, $\sigma$ gives a commutative diagram \[ \xymatrix{ \widehat X_{K} \times \widehat\Delta_{K}^{i-p} \ar[d]_{\mathrm{proj.}} \ar[r]^-{\sigma_1} & E_{p}\widehat\GL_{r,K} \ar[d] \\ \widehat X_{K}\ar[r]^-{\sigma_0} & B_{p}\widehat\GL_{r,K}. } \] The component of $I(\ch_{n,(r)}^{\mathrm{Kar}})$ in simplicial degree $p$ is given by $(\int_{\Delta^{p}} \omega_{0}, \int_{\Delta^{p}}\omega_{1})$ (cf. \eqref{mar2004}). Hence the image of $\sigma$ in the lower left corner of \eqref{eq:diagKar} is given by $(-1)^{(i-p)(i-p-1)/2}(\sigma_{0}^{*}\int_{\Delta^{p}}\omega_{0}, \sigma_{1}^{*}\int_{\Delta^{p}}\omega_{1}) \in \MF(\Omega^{\geq n}(\widehat X_{K}) \to \Omega^{*}(\widehat X_{K} \times \widehat\Delta^{i-p}_{K}))^{2n-p}$ and its image in $\Omega^{<n}(\widehat X_{K})$ by \[ \int_{\Delta^{i-p}} \sigma_{1}^{*}\int_{\Delta^{p}} \omega_{1} = \int_{\Delta^{i-p}\times\Delta^{p}} (\sigma_{1}\times\id_{\Delta^{p}})^{*}\omega_{1}. \] Note that there is no sign because the signs introduced via \eqref{nov1603} (cf. \eqref{dez0702}) and the integration map of Lemma \ref{lemma:ExplQuInv} cancel out. Next we compute the image of $\sigma$ going through diagram \eqref{eq:diagKar} clockwise. To do this, we use the shuffle map which is an inverse of the Eilenberg-Zilber isomorphism (cf. \cite[8.5.4]{Weibel}): Recall that a $(p,i-p)$ shuffle $(\mu,\nu)$ is a permutation $(\mu_{1}, \mu_{2}, \dots, \mu_{p}, \nu_{1}, \dots, \nu_{i-p})$ of $\{1, 2,\dots, i\}$ such that $\mu_{1} < \mu_{2} < \dots < \mu_{p}$ and $\nu_{1} < \dots < \nu_{i-p}$. It determines a map $\mu\colon [i] \to [p]$ in the category $\Delta$ given by $\mu=s^{\mu_{1}-1}\circ s^{\mu_{2}-1}\circ\cdots\circ s^{\mu_{p}-1}$ and similarly $\nu\colon [i] \to [i-p]$ such that $(\mu,\nu)\colon [i] \to [p]\times [i-p]$ is a non degenerate $i$-simplex of the simplicial set $\Delta^{p}\times\Delta^{i-p}$. All such simplices arise in this way. On any simplicial resp. cosimplicial object we have induced maps $\mu^{*}, \nu^{*}$, respectively $\mu_{*}, \nu_{*}$. In particular we have $\mu^{*}\colon E_{p}\widehat\GL_{r,K}\to E_{i}\widehat\GL_{r,K}$ resp. $B_{p}\widehat\GL_{r,K}\to B_{i}\widehat\GL_{r,K}$ and $\nu_{*}\colon \widehat\Delta^{i}_{K}\to \widehat\Delta^{i-p}_{K}$. Recall that we started with a $(p,i-p)$-simplex $\sigma$ in $F_{r}^\flat(A)$. Then $(\mu,\nu)^{*}\sigma$ is an $i$-simplex in $\diag F_{r}^{\flat}(A)$. The shuffle map sends $\sigma$ to \begin{equation*}\label{nov1901} \sum_{(\mu,\nu)} \sgn(\mu,\nu) (\mu,\nu)^{*}\sigma, \end{equation*} where the sum runs over all $(p,i-p)$-shuffles $(\mu,\nu)$. Its image under the Karoubi construction \eqref{eq:FastKaroubisRelChern} is then given by \[ \sum_{(\mu,\nu)} \sgn(\mu,\nu) \int_{\Delta^{i}} (((\mu,\nu)^{*}\sigma)_{1},\pr_{\Delta^{i}})^{*} \omega_{1}|_{E_{i}\widehat\GL_{r,K}\times\widehat\Delta^{i}_{K}}. \] Thus, to complete the proof we have to show the equality \begin{multline}\label{nov1903} \int_{\Delta^{i-p}\times\Delta^{p}} (\sigma_{1}\times\id_{\Delta^{p}})^{*} \omega_{1}|_{E_{p}\widehat \GL_{K}\times\widehat\Delta^{p}_{K}}= \\ \sum_{(\mu,\nu)} \sgn(\mu,\nu) \int_{\Delta^{i}} (((\mu,\nu)^{*}\sigma)_{1},\pr_{\Delta^{i}})^{*} \omega_{1}|_{E_{i}\widehat\GL_{K}\times\widehat\Delta^{i}_{K}}. \end{multline} Describing the map that appears in the right hand side of the formula more explicitly we have that for a $(p,i-p)$-shuffle $(\mu,\nu)$ the induced map $\widehat X_{K} \times \widehat\Delta^{i}_{K} \xrightarrow{((\mu,\nu)^{*}\sigma)_{1}} E_{i}\widehat\GL_{K}$ is given by $\mu^{*}\circ \sigma_{1} \circ (\id_{\widehat X_{K}} \times \nu_{*})$. Hence we have a commutative diagram \[ \xymatrix@C+2cm{ \widehat X_{K} \times \widehat\Delta^{i-p}_{K} \times \widehat\Delta^{p}_{K} \ar[r]^{\sigma_{1}\times\id_{\Delta^{p}}} & E_{p}\widehat\GL_{r,K}\times\widehat\Delta^{p}_{K}\\ \widehat X_{K} \times \widehat\Delta^{i-p}_{K} \times \widehat\Delta^{i}_{K} \ar[r]^{\sigma_{1}\times\id_{\Delta^{i}}} \ar[u]^{\id_{X} \times \id_{\Delta^{i-p}}\times \mu_{*}} & E_{p}\widehat\GL_{r,K}\times\widehat\Delta^{i}_{K} \ar[u]_{\id_{E_{p}\GL}\times \mu_{*}} \ar[d]^{\mu^{*}\times\id_{\Delta^{i}}} \\ \widehat X_{K}\times\widehat\Delta^{i}_{K} \ar[r]^{(((\mu,\nu)^{*}\sigma)_{1},\pr_{\Delta^{i}})} \ar[u]^{(\id_{X}\times \nu_{*},\pr_{\Delta^{i}})} & E_{i}\widehat\GL_{r,K}\times \widehat\Delta^{i}_{K}. } \] The composition of the left two vertical arrows is $\id_{\widehat X_{K}}\times(\nu_{*}, \mu_{*})$. The map $(\nu_{*}, \mu_{*})$ is exactly the map corresponding to the shuffle $(\mu,\nu)$ in the standard decomposition of $\Delta^{i-p} \times \Delta^{p}$ into $i$-simplices. It follows (cf.~\eqref{mar2007}) that \begin{equation}\label{nov1902} \int_{\Delta^{i-p}\times \Delta^{p}} (?)=\sum_{(\mu,\nu)}\sgn(\mu,\nu)\int_{\Delta^{i}} (\id_{X}\times(\nu_{*},\mu_{*}))^{*}(?). \end{equation} Since $\omega_{1}\in D^{2n-1}(E_{\dot}\widehat\GL_{r,K})$ is a simplicial differential form we have $(\mu^{*}\times\id_{\Delta^{i}})^{*}\omega_{1}|_{E_{i}\widehat\GL_{r,K}\times\widehat\Delta_{K}^{i}} = (\id_{E_{p}\widehat\GL_{r,K}} \times \mu_{*})^{*}\omega_{1}|_{E_{p}\widehat\GL_{r,K}\times\widehat\Delta_{K}^{p}}$. Hence we get \[ (\id_{X}\times(\nu_{*},\mu_{*}))^{*}(\sigma_{1}\times\id_{\Delta^{p}})^{*}\omega_{1}\big|_{E_{p}\widehat\GL_{r,K}\times\Delta^{p}} = (((\mu,\nu)^{*}\sigma)_{1}, \pr_{\Delta^{i}})^{*}\omega_{1}\big|_{E_{i}\widehat\GL_{r,K}\times\Delta^{i}}. \] Setting $(?)= (\sigma_{1}\times\id_{\Delta^{p}})^{*}\omega_{1}|_{E_{p}\widehat\GL_{r,K}\times\Delta^{p}}$ in \eqref{nov1902} we get the desired equality \eqref{nov1903}. \end{proof}
\section{Introduction} Snyder showed long ago that the continuous symmetries of space-time can be made consistent with a lattice through the construction of a covariant noncommutative algebra.\cite{Snyder:1946qz} The algebra that Snyder proposed is a nontrivial unification of space-time with the Poincar\'e algebra. The lattice is a result of the discrete representations of Snyder's algebra. It is not a classical lattice because only one spatial coordinate can be determined in a measurement. Here we generalize the Snyder algebra to get a nontrivial unification of space-time with the super-Poincar\'e algebra. Our supersymmetric extension is minimal and also has discrete representations corresponding to a spatial lattice. The lattice is compatible with supersymmetry transformations, and the lattice spacing is half that of the bosonic case. The supersymmetry generators act in a nonstandard way on the space-time. As discussed in \cite{us}, the translation group generated by the momenta is not associated with discrete translations on the spatial lattice. The system presented here might be useful in formulating an alternative discretization of supersymmetric field theories, where supersymmetry transformations are consistently implemented on the lattice.\cite{Joseph:2011xy} Supersymmetry has been shown useful in improving renormalizability properties of noncommutative field theories,\cite{Ruiz:2000hu} and this may turn out to be the case for the Snyder model as well. We shall not examine field theory here, but note that the lattice appears to be the appropriate setting for studying field theory in this case. Supersymmetric extensions of the Snyder algebra in arbitrary space-time dimensions were previously constructed by Hatsuda and Siegel.\cite{Hatsuda:2003wt} Their general approach is based on supersymmetric de Sitter algebras.\cite{Gates:1983nr} We shall give an alternative construction which attaches a pair of Grassmann odd spinors to de Sitter space, but does not require the spinors to generate a supersymmetric de Sitter algebra. Nevertheless, the super-Poincar\'e algebra is recovered upon projecting to Minkowski space-time. The algebra combines $N=1$ supersymmetry with the Snyder algebra, and in contrast to \cite{Hatsuda:2003wt}, the fermionic coordinates are all anticommuting. The Snyder algebra and its supersymmetric extensions are characterized by a deformation parameter $\Lambda$, which is proportional to one over the lattice spacing. One can define the action of the supersymmetry generators on superspace in the $\Lambda\rightarrow\infty$ limit. Superspace cannot be defined for finite $\Lambda$ (except in one-dimension), since the operators associated with the space-time coordinates are not simultaneously diagonalizable. On the other hand, the momentum operators commute, and so we can define a `super-momentum space' for finite $\Lambda$. We shall introduce fields on this space and write down differential representations for the supersymmetry generators. Discrete representations of the Snyder algebra were examined in \cite{us}, where it was argued that there are two distinct Hilbert spaces. (This is a result of the projection from de Sitter space.) The two are distinguished by $SU(2)$ quantum numbers, which are integer in one case and half-integer in the other. In our supersymmetric extension of the model we can construct fermionic raising and lowering operators that combine the integer and half-integer states in a single graded space. The raising and lowering operators simultaneously change the spin by $\pm 1/2$, and the location on the lattice to the nearest neighbor. Care must be taken in defining an involution for the fermionic operators. Negative norm states result if complex conjugation connects Lorentz spinors in the standard way. On the other hand, negative norm states are absent upon adopting an alternative involution which maps one de Sitter spinor to the other. The outline of this article is as follows: In section 2 we review the derivation of the Snyder algebra starting from de Sitter space. De Sitter spinors are introduced in section 3. Their projection to Minkowski space yields the minimal supersymmetric Snyder algebra. Discrete representations are examined in section 4 and an involution is introduced in section 5 which eliminates negative norm states. A differential representation for the supersymmetry generators on super-momentum space is given in section 6. Concluding remarks are made in section 7, including the construction of the extended supersymmetric Snyder algebra. \section{Snyder algebra} \setcounter{equation}{0} The original derivation of the Snyder algebra starts with a de Sitter manifold. Say that the latter is coordinatized by $ P_M$, $\;M=0,1,2,3,4$, which are all commuting and subject to the constraint \be P_M P_N\;\eta_{\mbox{\tiny dS}}^{MN} =1\;,\label{dstrspc}\ee where $\eta_{\mbox{\tiny dS}}$ is the de Sitter metric $\eta_{\mbox{\tiny dS}}=$diag$(-1,1,1,1,1)$. We take $ P_M$ to be dimensionless. Denote by $\ell^{MN}=-\ell^{NM}$ the generators of the de Sitter group with commutation relations \be [\ell^{MN},\ell^{RS}]= i(\eta_{\mbox{\tiny dS}}^{RM} \ell^{NS}-\eta_{\mbox{\tiny dS}}^{RN} \ell^{MS}+\eta^{SM}_{\mbox{\tiny dS}}\ell^{RN}-\eta_{\mbox{\tiny dS}}^{SN}\ell^{RM})\;,\label{lrntzalgbr} \ee Assuming $ P_M$ transforms as a five-vector with respect to the de Sitter group, one has \beqa [\ell^{MN}, P^R]&=& i(\eta_{\mbox{\tiny dS}}^{RM} P^N -\eta_{\mbox{\tiny dS}}^{RN} P^M)\label{capis5vctr} \eeqa The projection of this system to four dimensions was obtained by defining the four-momentum vector $p^m$, $\;m=0,1,2,3$, according to \be p^m=\Lambda\frac {P^m}{P^4} \;,\label{rdctnofmom}\ee and identifying $\ell^{4m}$ with the space-time four vector $x^m$, up to the dimensionful parameter $\Lambda$. For this Snyder could consistently set \be \ell^{mn}=x^mp^n-x^np^m \qquad\qquad \ell^{4m} =\Lambda x^m \label{orbtldsgn} \ee and then derive the algebra for the two Lorentz vectors from (\ref{lrntzalgbr}) and (\ref{capis5vctr}), \be [x^m,x^n]=\frac i{\Lambda^2}\ell^{mn} \qquad\quad [x^m,p^n]=i\Bigl(\eta^{mn} +\frac{p^mp^n}{\Lambda^2}\Bigr)\qquad\quad [p^m,p^n]=0 \label{xmxn}\;,\ee where $\eta=$diag$(-1,1,1,1)$ is the Minkowski metric. $\Lambda$ is a deformation parameter, and the Heisenberg algebra is recovered in the limit $\Lambda\rightarrow\infty$. From (\ref{dstrspc}) and (\ref{rdctnofmom}) one gets a mass upper bound, $-p^mp_m\le\Lambda^2$. The spatial coordinates $x_{\tt i}$, along with the orbital angular momentum $\ell^{\tt ij}$, ${\tt i,j,...}=1,2,3$, generate the $SO(4)$ subgroup of the de Sitter group. The discrete spectra of the position operators follows from the discrete representations of $SO(4)$. \section{Minimal supersymmetric extension} \setcounter{equation}{0} To get the supersymmetric version of the algebra we attach two spinor degrees of freedom to de Sitter space. Define $\Theta_{\tt A}$ and $\bar\Theta_{\tt A}$, ${\tt A}=1,2,3,4, $ to be two conjugate four-dimensional spinors which commute with $P^M$ and $\ell^{MN}$, and satisfy the anti-commutation relations \be \{\Theta_{\tt A},\bar\Theta_{\tt B}\}=-i\delta_{\tt AB} \qquad\qquad \{\Theta_{\tt A},\Theta_{\tt B}\}= \{\bar\Theta_{\tt A},\bar\Theta_{\tt B}\}=0 \label{Capthtbrkts} \ee Assume they give a spin contribution to the generators of the de Sitter group. Denote by $j^{MN}=-j^{NM}$ the sum of the orbital and spin contributions, \be j^{MN}=\ell^{MN} - \bar\Theta\Sigma^{MN}\Theta \;,\label{dstrgnrtrs}\ee where $\Sigma^{MN}=-\Sigma^{NM}$ define the four-by-four spinor representation of the algebra \be [\Sigma^{MN},\Sigma^{RS}]= \eta_{\mbox{\tiny dS}}^{RM} \Sigma^{NS}-\eta_{\mbox{\tiny dS}}^{RN} \Sigma^{MS}+\eta^{SM}_{\mbox{\tiny dS}} \Sigma^{RN}-\eta_{\mbox{\tiny dS}}^{SN} \Sigma^{RM} \ee Then $\Theta$ and $\bar\Theta $ transform as de Sitter spinors, while $P^M$ remains a five-vector under de Sitter transformations, now generated by $j^{MN}$, \beqa [j^{MN},\bar\Theta]&=&i\bar\Theta \Sigma^{MN} \cr &&\cr [j^{MN}, \Theta]&=&-i\Sigma^{MN}\Theta\cr &&\cr [j^{MN}, P^R]&=& i(\eta_{\mbox{\tiny dS}}^{RM} P^N -\eta_{\mbox{\tiny dS}}^{RN} P^M)\cr &&\cr [j^{MN},j^{RS}]&=& i(\eta_{\mbox{\tiny dS}}^{RM} j^{NS}-\eta_{\mbox{\tiny dS}}^{RN} j^{MS}+\eta^{SM}_{\mbox{\tiny dS}} j^{RN}-\eta_{\mbox{\tiny dS}}^{SN} j^{RM})\;\label{ttllrntzalgbr} \eeqa To project to four space-time dimensions we again assume (\ref{rdctnofmom}) and (\ref{orbtldsgn}), leading to the Snyder algebra (\ref{xmxn}) for the bosonic operators $x^m$ and $p^m$. For the spinor representations, we choose \be \Sigma^{mn}=\pmatrix{\sigma^{mn} &\cr & \bar\sigma^{mn}\cr}\qquad\qquad \Sigma^{4m}=\frac 12 \pmatrix{&\sigma^m\cr\bar\sigma^m &\cr}\;, \ee where we follow conventions in Wess and Bagger:\cite{WB} $\sigma^0=\bar\sigma^0=-\BI_{2\times 2}$ and $\bar\sigma^{\tt i}=-\sigma^{\tt i}$, where $\sigma^{\tt i},\;{\tt i}=1,2,3$, are the three Pauli matrices. The $2\times 2$ Lorentz matrices $\sigma^{mn}$ and $\bar\sigma^{mn}$ are defined by \beqa \sigma^{mn} =\frac 14 (\sigma^m\bar \sigma^n - \sigma^n\bar \sigma^m) &\qquad\quad &\bar \sigma^{mn} =\frac 14 (\bar\sigma^m \sigma^n -\bar \sigma^n \sigma^m)\label{lrntzspin} \eeqa They satisfy the identities \beqa 2\sigma^{mn}\sigma^s &=& \eta^{ms}\sigma^n -\eta^{ns}\sigma^m +i\epsilon^{mnsr}\sigma_r \cr & &\cr 2\sigma^s\bar \sigma^{mn} &=& -\eta^{ms}\sigma^n +\eta^{ns}\sigma^m +i\epsilon^{mnsr}\sigma_r \cr & &\cr 2\bar\sigma^{mn}\bar\sigma^s &=& \eta^{ms}\bar\sigma^n -\eta^{ns}\bar\sigma^m -i\epsilon^{mnsr}\bar\sigma_r \cr & &\cr 2\bar\sigma^s\sigma^{mn} &=& -\eta^{ms}\bar\sigma^n +\eta^{ns}\bar\sigma^m -i\epsilon^{mnsr}\bar\sigma_r \cr &&\cr [\sigma^{mn},\sigma^{rs}]&=&\eta^{rm} \sigma^{ns}-\eta^{rn} \sigma^{ms}+\eta^{sm} \sigma^{rn}-\eta^{sn} \sigma^{rm}\cr &&\cr [\bar\sigma^{mn},\bar\sigma^{rs}]&=&\eta^{rm}\bar \sigma^{ns}-\eta^{rn} \bar\sigma^{ms}+\eta^{sm}\bar \sigma^{rn}-\eta^{sn} \bar\sigma^{nm} \eeqa Next we express the two four-dimensional spinors $\Theta$ and $\bar\Theta$ in terms of four two-dimensional Lorentz spinors $Q_\alpha $, $\bar Q_{\dot \alpha}$, $\theta^\alpha $ and $\bar \theta^{\dot \alpha}$, $\; \alpha,\dot\alpha=1,2$. We write $\bar\Theta$ as a row matrix and $\Theta$ as a column matrix according to \be \bar \Theta=\frac 1{\sqrt{\Lambda}}\pmatrix{\Lambda\theta \;&\;\bar Q +i\theta\sigma^n p_n }\qquad\qquad \Theta=\frac 1{\sqrt{\Lambda}}\pmatrix{-Q+ip_n\sigma^n\bar\theta\cr\Lambda\bar\theta }\label{redcaptht}\;,\ee $\Lambda$ being the same dimensionful parameter appearing in (\ref{rdctnofmom}). Since $\Theta$ and $\bar\Theta$ commute with $p^m$, the four two-dimensional spinors must also commute with $p^m$. From the anti-commutation relations (\ref{Capthtbrkts}) for $\Theta$ and $\bar\Theta$, it follows: \noindent $i)$ that $\theta^\alpha $ and $\bar \theta^{\dot \alpha}$ are all anti-commuting Grassmann coordinates\footnote{This is in contrast to \cite{Hatsuda:2003wt}, where the anticommutators of fermionic coordinates are in general a linear combination of space-time coordinates, Lorentz generators and additional $SO(N)$ generators.}, i.e., \be \{\theta^\alpha,\theta^\beta\}=\{\bar \theta^{\dot \alpha},\bar \theta^{\dot \beta}\} =\{\theta^\alpha,\bar \theta^{\dot \beta}\}=0\;,\label{grsmncrdnts} \ee $ii)$ that $Q_\alpha $ and $\bar Q_{\dot \alpha}$ are canonically conjugate to $\theta^\alpha $ and $\bar \theta^{\dot \alpha}$, respectively, \be \{Q_\alpha,\theta^\beta\}=i\delta_{\alpha}^{\beta}\qquad \quad \{\bar Q_{\dot \alpha},\bar\theta^{\dot\beta}\}=-i\delta_{\dot \alpha}^{\dot \beta} \qquad\quad \{Q_\alpha,\bar\theta^{\dot\beta}\}= \{\bar Q_{\dot \alpha},\theta^\beta\}=0 \label{susyhisnbrg}\;,\ee and $iii)$ that $Q_\alpha$ and $\bar Q_{\dot \alpha}$ are super-translation generators, \be\{Q_\alpha,\bar Q_{\dot \alpha}\}= 2{\sigma_{\alpha\dot\alpha}}^m p_m\qquad\quad\{Q_\alpha,Q_\beta\}=\{\bar Q_{\dot \alpha},\bar Q_{\dot \beta}\}=0 \label{susyalgbr}\ee Finally, since $\Theta$ and $\bar\Theta$ commute with coordinates $x^m$, it follows: \noindent $iv)$ that $\theta$ and $\bar\theta$ also commute with $x^m$, while \beqa [Q_\alpha,x^m]=\Bigl( \sigma^m + \frac {\sigma^np_np^m}{\Lambda^2}\Bigr)_{\alpha\dot\alpha}\bar\theta^{\dot\alpha} &\qquad\quad & [\bar Q_{\dot \alpha},x^m]= -\theta^\alpha\Bigl( \sigma^m + \frac {\sigma^np_np^m}{\Lambda^2}\Bigr)_{\alpha\dot\alpha}\quad \label{4dsnyderQQdx}\eeqa $Q_\alpha$, $\bar Q_{\dot \alpha}$ and $p^m$ generate the $N=1$ super-translation group. Upon including the Lorentz generators $j^{mn}$, which using (\ref{redcaptht}) can be expressed as \be j^{mn}=x^mp^n-x^np^m +\theta\sigma^{mn}Q-\bar Q \bar\sigma^{mn}\bar\theta -\epsilon^{mnrs}(\theta\sigma_r\bar\theta) p_s \;, \label{Lrntzgnrtrs}\ee we get the super-Poincar\'e group. From the commutation relations with $j^{mn}$, \beqa [Q_\alpha,j^{mn}]= i(\sigma^{mn}Q)_\alpha &\qquad\quad & [\bar Q_{\dot \alpha},j^{mn}]= -i(\bar Q \bar \sigma^{mn})_{\dot \alpha}\cr &&\cr [\theta^\alpha,j^{mn}]= -i(\theta\sigma^{mn})^\alpha &\qquad\quad & [\bar \theta^{\dot \alpha},j^{mn}]=i(\bar\sigma^{mn}\bar\theta)^{\dot\alpha}\cr &&\cr [x^r,j^{mn}]= i(x^m\eta^{rn}- x^n\eta^{rm}) &\qquad\quad & [p^r,j^{mn}]= i(p^m\eta^{rn}- p^n\eta^{rm}) \label{spnrvctrtrns}\;,\eeqa it follows that $Q_\alpha$, $\bar Q_{\dot \alpha}$, $\theta^\alpha $ and $\bar \theta^{\dot \alpha}$ transform as Lorentz spinors, while $x^m$ and $p^m$ as Lorentz vectors. Eqs. (\ref{grsmncrdnts})-(\ref{4dsnyderQQdx}), along with (\ref{xmxn}), defines a minimal supersymmetric extension of the Snyder algebra. It is expressed above in terms of $\theta^\alpha$, $\bar\theta^{\dot\alpha}$ and $x^m$, in addition to the $N=1$ super-translation generators $Q_\alpha$, $\bar Q_{\dot \alpha}$ and $p^m$. Alternatively, instead of $x^m$, we can define the following space-time coordinates in the supersymmetric theory: \beqa X^m=\frac 1{\Lambda}j^{4m} &=&x^m+\frac 1{2\Lambda^2}\Bigl(\bar Q\bar\sigma^m Q+i\theta\sigma^np_n\bar\sigma^mQ-i\bar Q\bar\sigma^m\sigma^np_n\bar \theta\Bigr)\cr&&\cr&&\qquad\;\;-\frac 1{\Lambda^2} p^m\theta\sigma^np_n\bar\theta-\frac 12 \Bigl(1-\frac{p^np_n}{\Lambda^2} \Bigr)\theta\sigma^m\bar\theta\;\label{capx}\eeqa It forms a closed algebra with the Poincar\'e generators. In contrast to (\ref{xmxn}), one gets \be [X^m,X^n]=\frac i{\Lambda^2}j^{mn} \qquad\quad [X^m,p^n]=i\Bigl(\eta^{mn} +\frac{p^mp^n}{\Lambda^2}\Bigr)\qquad\quad [p^m,p^n]=0 \label{capxmxn}\;,\ee where $j^{mn}$ are the Lorentz generators (\ref{Lrntzgnrtrs}). The latter can be re-expressed in terms of $X^n$ using (\ref{capx}). $X^n$'s commutators with the spinors are given by \beqa [X^m,\theta]&=&\frac i{2\Lambda^2}\;(\bar Q+i \theta \sigma^n p_n)\bar \sigma^m \cr &&\cr [X^m,\bar\theta]&=&\frac i{2\Lambda^2}\;\bar\sigma^m (Q-ip_n\sigma^n\bar\theta) \cr &&\cr [X^m,\bar Q+i \theta \sigma^n p_n]&=&\frac i{2}\theta \sigma^m \cr &&\cr [X^m, Q-i \sigma^n p_n\bar \theta]&=&\frac i{2} \sigma^m\bar \theta\label{cmtrsfcX} \eeqa Eqs. (\ref{capxmxn}) and (\ref{cmtrsfcX}), along with (\ref{grsmncrdnts})-(\ref{susyalgbr}), give an alternative definition of the supersymmetric Snyder algebra. Due to the nonstandard commutation relations of $Q_\alpha$, $\bar Q_{\dot\alpha}$ and $p_m$ with the coordinate operators $X^m$ (or $x^m$), the super-translation group acts in a nonstandard fashion on space-time. \section{Discrete representations} \setcounter{equation}{0} We regard $X_{\tt i}$, ${\tt i}=1,2,3$, as the position operators for the supersymmetric Snyder algebra. They are the spatial components of $X^n$. From (\ref{capxmxn}), $X_{\tt i}$, along with $j^{\tt ij}$, generate another $SO(4)$ group. It follows that $X_{\tt i}$, like $x_{\tt i}$, have discrete spectra. Here we can define \be A_{\tt i} =\frac 12\Bigl( J_{\tt i} +\Lambda X_{\tt i}\Bigr)\qquad\qquad B_{\tt i} =\frac 12\Bigl( J_{\tt i} -\Lambda X_{\tt i}\Bigr)\;,\label{su2chargs}\ee where $ J_{\tt i}=\frac12\epsilon_{\tt ijk}\,j^{\tt jk} $ is the rotation generator. They satisfy two $su(2)$ algebras \beqa [ A_{\tt i}, A_{\tt j}] &=&i\epsilon_{\tt ijk} A_{\tt k}\cr & &\cr [ B_{\tt i}, B_{\tt j}] &=&i\epsilon_{\tt ijk} B_{\tt k}\cr & &\cr [ A_{\tt i}, B_{\tt j}] &=&0 \; \label{ofralgbr} \eeqa $ A_{\tt i} A_{\tt i}$, $ B_{\tt i} B_{\tt i}$, $ A_3$ and $ B_3$ form a complete set of commuting operators.\footnote{The Casimir operators $ A_{\tt i} A_{\tt i}$ and $ B_{\tt i} B_{\tt i}$ are independent, unlike their counterparts in the non supersymmetric theory.\cite{us}} We denote their eigenvectors by $|j_A,j_B,m_A,m_B>$, \beqa A_{\tt i} A_{\tt i}\;|j_A,j_B,m_A,m_B> &=& j_A(j_A+1)\;|j_A,j_B,m_A,m_B>\cr & &\cr B_{\tt i} B_{\tt i}\;|j_A,j_B,m_A,m_B> &=& j_B(j_B+1)\;|j_A,j_B,m_A,m_B>\cr & &\cr A_3\;|j_A,j_B,m_A,m_B> &=& m_A|j_A,j_B,m_A,m_B>\cr & &\cr B_3\;|j_A,j_B,m_A,m_B> &=& m_B\;|j_A,j_B,m_A,m_B> \;,\label{eigneqAB} \eeqa where $ m_A=-j_A,1-j_A,...,j_A\;,$ and $ m_B=-j_B,1-j_B,...,j_B\;$. $j_A$ and $j_B$ take values $0,\frac 12,1,\frac 32,...\;,$ and label the $SO(4)$ representations. $|j_A,j_B,m_A,m_B>$ is also an eigenvector of $ J_3$ and $ X_3$, whose corresponding eigenvalues are $m_A+m_B$ and $(m_A-m_B)/{\Lambda}$, respectively. Applying $ A_\pm$ or $ B_\pm$ changes $|J_3|$ by $1$ and $|X_3|$ by $1/\Lambda$. $SO(4)$ group representations are also present in the non-supersymmetric version of the theory, and there $m_A$ and $m_B$ are either both integer or both half-integer.\cite{us} It followed that the associated (orbital) angular momentum operators had integer eigenvalues, and also that the eigenvalues of the position operators $x_{\tt i}$ were evenly spaced at intervals of $\Lambda^{-1}$. We show below that for the supersymmetric Snyder algebra, eigenvalues of the position operators $X_{\tt i}$ are evenly spaced at intervals of $\frac 12\Lambda^{-1}$, and that both integer and half-integer values of $m_A$ and $m_B$ occur in the representation of the graded algebra. This means that both integer and half-integer values for $j_A$ and $j_B$ occur in a representation of the supersymmetry algebra. From (\ref{spnrvctrtrns}) and (\ref{cmtrsfcX}) one can construct various linear combinations of the spinors which act as raising and lowering operators with respect to the eigenvalues of the operators $X_{\tt i}$ and $J_{\tt i}$, or equivalently, $A_{\tt i}$ and $B_{\tt i}$. For the case of ${\tt i}=3$, we can define the raising and lowering operators $a^{\tt A}_\pm$, ${\tt A}=1,2,3,4$, according to \beqa \sqrt{2}\; a^1_\pm&=&\bar\Theta_1\mp i\bar\Theta_3\;\;=\;\;\sqrt{\Lambda}\theta^1 \mp \frac i{\sqrt{\Lambda}}(\bar Q+i\theta\sigma^np_n)_1\cr &&\cr \sqrt{2}\; a^2_\pm&=&\bar\Theta_2\pm i\bar\Theta_4\;\;=\;\;\sqrt{\Lambda}\theta^2 \pm \frac i{\sqrt{\Lambda}}(\bar Q+i\theta\sigma^np_n)_2\cr &&\cr \sqrt{2}\; a^3_\pm&=&\Theta_3\pm i\Theta_1\;\;=\;\;\sqrt{\Lambda}\bar\theta^1 \mp \frac i{\sqrt{\Lambda}}( Q-i\sigma^np_n\bar\theta)_1\cr &&\cr \sqrt{2}\; a^4_\pm&=&\Theta_4\mp i\Theta_2\;\;=\;\;\sqrt{\Lambda}\bar\theta^2 \pm \frac i{\sqrt{\Lambda}}( Q-i\sigma^np_n\bar\theta)_2\eeqa Their nonvanishing anticommutators are \be\{a^1_\mp, a^3_{\pm}\}=\{a^2_{\pm},a^4_{\mp}\}=\pm 1\label{aanticmtrs}\ee From (\ref{spnrvctrtrns}) and (\ref{cmtrsfcX}) one has \be [X_3, a^{\tt A}_\pm]=\pm \frac 1{2\Lambda}\;a^{\tt A}_\pm\; \label{ralx3} \ee \be [J_3, a^{1,4}_\pm]= \frac 12 a^{1,4}_\pm\quad\qquad [J_3, a^{2,3}_\pm]= -\frac 12 a^{2,3}_\pm\label{J3cmtrwapm}\ee Say that $|m_A,m_B>$ is an eigenvector of $ X_3$ and $J_3$ with eigenvalue is $(m_A-m_B)/{\Lambda}$ and $m_A+m_B$, respectively, where for convenience we ignore the dependence on the indices $j_A$ and $j_B$. Then $a^{\tt A}_\pm|m_A,m_B>$ are also eigenvectors of $ X_3$ and $J_3$. From (\ref{ralx3}), the $X_3$ eigenvalue of $a^{\tt A}_\pm|m_A,m_B>$ is $(m_A-m_B\pm\frac 1{2})/{\Lambda}$. From (\ref{J3cmtrwapm}), the $ J_3$ eigenvalue of $a^{1,4}_\pm|m_A,m_B>$ is $m_A+m_B+\frac 1{2}$, while the $ J_3$ eigenvalue of $a^{2,3}_\pm|m_A,m_B>$ is $m_A+m_B-\frac 1{2}$. So application of $ a^{\tt A}_\pm$ simultaneously changes $|J_3|$ by $\frac12$ and $|X_3|$ by $\frac 1{2\Lambda}$. Up to degenerate states, \beqa a^{1,4}_+|m_A,m_B>&\sim & \Big|m_A+\frac 12,m_B\Big> \cr&&\cr a^{1,4}_-|m_A,m_B>&\sim &\Big |m_A,m_B+\frac 12\Big>\cr&&\cr a^{2,3}_+|m_A,m_B>&\sim & \Big|m_A,m_B-\frac 12\Big> \cr&&\cr a^{2,3}_-|m_A,m_B>&\sim & \Big|m_A-\frac 12,m_B\Big> \label{actnfcaops}\eeqa It follows that, unlike what happens in the non-supersymmetric theory, both integer and half-integer values of $m_A$ and $m_B$ occur in the representations of the supersymmetric algebra. Consequently, both integer and half-integer values of $j_A$ and $j_B$ occur in the representations. In the non-supersymmetric theory, the eigenvalues of the position operator are regularly spaced at intervals of $\Lambda^{-1}$, whereas here they are spaced at intervals of $\frac 12\Lambda^{-1}$. $a^{\tt A}_\pm$ were defined to raise and lower eigenvalues associated with the $3-$direction. Raising and lowering operators can also be constructed for the $1-$ and $2-$directions, and they yield the same spectra for the position and angular momentum operators. \section{Involutions} \setcounter{equation}{0} There are two approaches to introducing an involution, or complex conjugation, of the algebra. The first which we discuss below connects Lorentz spinor $\theta$ with $\bar\theta$ and $Q$ with $\bar Q$ in the usual way. It leads to negative norm states. No negative norm states result from an alternative involution. The latter maps one de Sitter spinor to the other. Complex conjugation (which we denote by $*$) in four space-time dimensions standardly relates spinor representations to conjugate representations according to \be (\theta^\alpha)^*=\bar \theta^{\dot\alpha}\qquad\quad (Q_\alpha)^*=\bar Q_{\dot\alpha}\, \label{cclrntspnr}\ee Its action on the de Sitter spinors $\Theta$ and $\bar\Theta$ is then \be (\bar\Theta^*)=\pmatrix{ &\BI\cr-\BI &\cr}\Theta \qquad\qquad(\Theta^*)=\bar\Theta\pmatrix{ &\BI\cr-\BI &\cr}\label{cconcpth} \ee It is then easy to show that \be \{\Theta_{\tt A}^*,\bar\Theta_{\tt B}^*\}=i\delta_{\tt AB} \qquad\qquad \{\Theta_{\tt A}^*,\Theta_{\tt B}^*\}= \{\bar\Theta_{\tt A}^*,\bar\Theta_{\tt B}^*\}=0 \,\label{stranticmtr}\ee This means that complex is consistent with the supersymmetric Snyder algebra generated by $x^m$, $p_m$, $\theta^\alpha$ $,\bar\theta^{\dot\alpha}$, $ Q_\alpha$ and $\bar Q_{\dot\alpha}$, since all anticommutators between Lorentz spinors followed from (\ref{Capthtbrkts}). We assume that $p_m$ and $x^m$ are real. It follows that all of the de Sitter group generators $X^m$ and $j^{mn}$ are also real. For the raising and lowering operators we get \be (a_\pm^1)^*=a_\mp^3 \qquad\quad (a_\pm^2)^*=a_\mp^4\,\label{ccfapm} \ee From the anticommutation relations (\ref{aanticmtrs}), the norm-squared of eigenvectors of $A_3$ and $B_3$ are related by \beqa \Big|a^1_+|m_A,m_B>\Big|^2&=&-\;\Big|a^3_-|m_A,m_B>\Big|^2-\; \Big||m_A,m_B>\Big|^2\cr &&\cr \Big|a^2_-|m_A,m_B>\Big|^2&=&-\;\Big|a^4_+|m_A,m_B>\Big|^2-\; \Big||m_A,m_B>\Big|^2\;,\eeqa implying the existence of negative norm states. Alternatively, one can define \be b^{\tt A}_-=\bar\Theta_{\tt A}\qquad \qquad b^{\tt A}_+=i\Theta_{\tt A}\;,\ee which satisfy the usual algebra of fermionic creation and annihilation operators, \be\{b_+^{\tt A},b_-^{\tt B}\}=\delta^{\tt AB} \qquad\quad\{b_+^{\tt A}, b_+^{\tt B}\}=\{b_-^{\tt A}, b_-^{\tt B}\}=0\ee However from (\ref{cconcpth}), $b_+^{\tt A}$ is not the complex conjugate of $b_-^{\tt A}$, again implying the existence of negative norm states, e.g. $\Big|b^1_-|m_A,m_B>\Big|^2=-\Big|b^3_+|m_A,m_B>\Big|^2$. We note from (\ref{actnfcaops}), that while $b_\pm^{\tt A}$ acting on $|m_A,m_B>$ is an eigenvector of $ J_3$, it is not an eigenvector of $ X_3$, $ A_3$ or $B_3$. On the other hand, negative norm states are absent if we replace the $*$ by another involution-operation, which we denote by $\star$, satisfying $(x^\star)^\star =x$, $(xy)^\star=y^\star x^\star$ and $i^\star=-i$. We define it in terms of the de Sitter spinors according to \be \bar\Theta_{\tt A}^\star =i\Theta_{\tt A} \label{stroncpth}\;, \ee in contrast to (\ref{cconcpth}). (\ref{stranticmtr}) is again satisfied (with $\star$ now replacing $*$), and so the $\star$ involution is consistent with the supersymmetric Snyder algebra. We assume that $p_m$ and $x^m$ are real with respect to $\star$. In order that the de Sitter generators $X^m$ and $j^{mn}$ are real under the $\star$ involution we need that \be \sigma^{m\star}=-\bar\sigma^m \qquad\quad \bar\sigma^{m\star}=-\sigma^m \label{strsgma} \ee The action of the $\star$ involution on Lorentz spinors is more involved than the previous complex conjugation (\ref{cclrntspnr}). Using (\ref{redcaptht}) and (\ref{strsgma}), we get \beqa (\theta^\alpha)^\star &=&-\frac i\Lambda (Q-i \sigma^np_n\bar\theta )_{\alpha}\cr &&\cr(\bar \theta^{\dot\alpha})^\star &=&\frac i\Lambda (\bar Q+i\theta \sigma^np_n )_{\dot\alpha}\cr &&\cr Q_{\alpha}^\star &=& -\frac 1\Lambda ( \bar Q\bar\sigma^n)^\alpha p_n-i\Lambda \theta^\alpha \Bigl(1-\frac{p^np_n}{\Lambda^2}\Bigr)\cr &&\cr \bar Q_{\dot\alpha}^\star &=&-\frac 1\Lambda ( \bar\sigma^nQ)^{\dot\alpha} p_n+ i\Lambda\bar\theta^{\dot\alpha}\Bigl(1-\frac{p^np_n}{\Lambda^2}\Bigr)\eeqa From (\ref{stroncpth}), the $\star$ involution of the raising and lowering operators is given by \be (a_\pm^{1})^\star=\mp a_\mp^{3}\qquad\quad (a^2_\pm)^\star=\pm a_\mp^{4}\;,\label{cvlnapm}\ee in contrast to (\ref{ccfapm}), or simply, \be (b_\pm^{\tt A})^\star=b_\mp^{\tt A}\ee We can then introduce a set of states $\{|cv> \}$, corresponding to Clifford vacuum states , i.e., $ b_-^{\tt A}|cv>=0\;.$ They should form a representation of the (non-supersymmetric) Snyder algebra, i.e., the algebra generated by the bosonic operators $x_{\tt i}$ and $p_{\tt i}$, since these operators commute with $b_\pm^{\tt A}$. Two such infinite-dimensional representations were found in \cite{us}. To the Clifford vacuum we add all states obtained by acting with $b_+^{\tt A}$ to obtain a representation of the supersymmetric theory. There are no negative norm states in this case, because hermitian conjugation is with respect to $\star$, i.e., hermiticity for any vectors $|\psi>$ and $|\phi >$ here means $<\psi|\phi>^\star=<\phi|\psi>$. \section{Momentum-dependent superfields} \setcounter{equation}{0} Superspace is standardly coordinatized by space-time coordinates and Grassmann odd variables. This is not possible for finite $\Lambda$, since the space-time coordinates (either $x^m$ or $X^m$) do not commute amongst themselves in this case, and moreover, they have discrete spectra. On the other hand, since $[p_m,p_n]=0$, we can define a momentum superspace spanned by $p^m$, $\theta^\alpha $ and $\bar \theta^{\dot \alpha}$, and then write down fields on this space. Using $\frac\partial{\partial\theta^\alpha}\theta^\beta=\delta^\beta_\alpha$, $\frac\partial{\partial\bar\theta^{\dot\alpha}}\bar\theta^{\dot\beta}=\delta^{\dot\beta}_{\dot \alpha}$ and $\frac\partial{\partial\theta^{\alpha}}\bar\theta^{\dot\beta}=\frac\partial{\partial\bar\theta^{\dot\alpha}}\theta^{\beta}=0$, one can represent the four spinors in (\ref{redcaptht}) by \be \bar \Theta=\pmatrix{\sqrt{\Lambda}\theta \;&\;-\frac i{\sqrt{\Lambda}}\frac\partial{\partial\bar\theta} }\qquad\qquad \Theta=\pmatrix{-\frac i{\sqrt{\Lambda}}\frac\partial{\partial\theta}\cr\sqrt{\Lambda}\bar\theta }\ee Then $Q_\alpha$, $\bar Q_{\dot \alpha}$ , $x^m$ and $j^{mn}$ are given by differential operators \beqa Q_\alpha &=&i\Bigl(\frac\partial{\partial\theta^\alpha} +{\sigma_{\alpha\dot\alpha}}^mp_m\bar\theta^{\dot\alpha}\Bigr)\cr & &\cr \bar Q_{\dot \alpha} &=&-i\Bigl(\frac\partial{\partial\bar\theta^{\dot \alpha}}+\theta^\alpha {\sigma_{\alpha\dot\alpha}}^mp_m\Bigr) \cr & &\cr X^m&=&i \Bigl(\frac\partial{\partial p_m} +\frac{p^mp^n}{\Lambda^2}\frac\partial{\partial p^n}\Bigr)-\frac 12\theta\sigma^m\bar\theta+\frac1{2\Lambda^2} \frac\partial{\partial\bar\theta}\bar\sigma^m \frac\partial{\partial\theta} \cr & &\cr j^{mn}&=&i\Bigl(-p_m\frac\partial{\partial p_n}+p_n\frac\partial{\partial p_m}+\theta\sigma^{mn}\frac\partial{\partial\theta} +\frac\partial{\partial\bar\theta}\bar\sigma^{mn} \bar\theta\Bigr)\label{sprsndrdfrntnlrep}\eeqa As is usual, one can construct fermionic operators $D_\alpha$ and $\bar D_{\dot \alpha}$ which anticommute with the super-translation generators $Q_\alpha$ and $\bar Q_{\dot\alpha}$, and use them to reduce supersymmetric representations. They are \beqa D_\alpha &=&i\Bigl(\frac\partial{\partial\theta^\alpha} -{\sigma_{\alpha\dot\alpha}}^mp_m\bar\theta^{\dot\alpha}\Bigr)\cr & &\cr \bar D_{\dot \alpha} &=&-i\Bigl(\frac\partial{\partial\bar\theta^{\dot \alpha}}-\theta^\alpha {\sigma_{\alpha\dot\alpha}}^mp_m\Bigr) \;,\eeqa satisfying \be\{D_\alpha,\bar D_{\dot \alpha}\}= -2{\sigma_{\alpha\dot\alpha}}^m p_m\qquad\quad\{D_\alpha,D_\beta\}=\{\bar D_{\dot \alpha},\bar D_{\dot \beta}\}=0\label{DDdgralgbr}\ee A chiral field $ \Phi$ on super-momentum space satisfies \be \bar D_{\dot\alpha} \Phi=0\; \ee This is solved by \be \Phi= {\cal F}(p,\theta)\;e^{-\theta p\cdot \sigma\bar\theta} \ee Assuming ${\cal F}(p,\theta)$ to be a bosonic field, it can be expanded in terms of two bosonic component fields and a spinor femionic field. We define the action of an operator ${\cal O}$ on the function ${\cal F}(p,\theta)$ according to $ {\cal O} \Phi= [{\cal O}{\cal F}]\;e^{-\theta p\cdot \sigma\bar\theta}$. Applying the super-translation generators, one gets \beqa Q_\alpha {\cal F} &=& i \frac {\partial{\cal F}}{\partial \theta^\alpha} \cr &&\cr\bar Q_{\dot\alpha} {\cal F} &=& -2i(\theta \sigma\cdot p)_{\dot\alpha}{\cal F} \label{qqbroncrl}\eeqa Similar considerations can be made for anti-chiral fields $\bar{\Phi}$, which satisfy \be D_\alpha \bar\Phi=0\; \ee It is solved by \be \bar{ \Phi}= \bar {\cal F}(p,\bar\theta)\;e^{\theta p\cdot \sigma\bar\theta} \ee If we define the action of an operator ${\cal O}$ on the function $\bar {\cal F}(p,\bar\theta)$ by $ {\cal O}\bar \Phi= [{\cal O}\bar {\cal F}]\;e^{\theta p\cdot\sigma\bar\theta}$, then in contrast (\ref{qqbroncrl}), \beqa Q_\alpha \bar{\cal F} &=& 2i(p\cdot \sigma\bar\theta)_\alpha \bar{\cal F} \cr &&\cr \bar Q_{\dot\alpha}\bar {\cal F} &=&- i \frac{\partial \bar {\cal F}}{\partial\bar \theta^{\dot\alpha} } \eeqa \section{Concluding remarks} \setcounter{equation}{0} Our derivation of the supersymmetric Snyder algebra closely follows Snyder's work.\cite{Snyder:1946qz} In particular, it uses his projection to Minkowski space-time. Other projections from de Sitter space have been considered.\cite{Hatsuda:2003wt} Snyder's algebra also has been derived starting from relativistic particle dynamics.\cite{Jaroszkiewicz}, \cite{Romero:2004er},\cite{Romero:2006pe},\cite{Banerjee:2006wf},\cite{Chatterjee:2008bp},\cite{Stern:2010ri} Particle dynamics on Snyder space has also been studied in \cite{Mignemi:2011gr},\cite{Lu:2011fh}. Ref. \cite{Stern:2010ri}, in particular, begins from the reparametrization invariant action of a relativistic particle. One arrives at the Snyder algebra (or more precisely, its three-dimensional Euclidean subalgebra) from a particular gauge condition which fixed the reparametrization freedom. It should also be possible to obtain the supersymmetric Snyder algebra starting from an action principle for supersymmetric particles, for example \cite{Galvao:1980cu}. As the system has two first class constraints it will require to gauge constraints to eliminate all gauge degrees of freedom. There may exist a choice of conditions whereby the supersymmetric Snyder algebra is realized by the Dirac brackets. Finally, a number of generalizations of our construction are possible. One can consider supersymmetric Snyder algebras in different space-time dimensions and also extended supersymmetry. For the three-dimensional Euclidean version of the supersymmetry Snyder algebra, see \cite{Beppefest}. Concerning extended supersymmetry, its construction is straightforward. For this we can introduce $2N$ de Sitter spinors $\Theta^{\tt a}_{\tt A}$ and $\bar\Theta^{\tt a}_{\tt A}$, ${\tt a}=1,...N$, satisfying \beqa \{\Theta^{\tt a}_{\tt A},\bar\Theta^{\tt b}_{\tt B}\}&=&-i\delta_{\tt AB} \delta^{\tt ab}\cr &&\cr \{\bar\Theta^{\tt a}_{\tt A},\bar\Theta^{\tt b}_{\tt B}\}&=&\frac 2\Lambda E^+_{\tt AB} Z_+^{\tt ab}\cr &&\cr \{\Theta^{\tt a}_{\tt A},\Theta^{\tt b}_{\tt B}\}&=&\frac 2\Lambda E^-_{\tt AB} Z_-^{\tt ab}\;, \label{xtnddCapthtbrkts} \eeqa where $Z_+^{\tt ab}=-Z_+^{\tt ba}$ and $Z_-^{\tt ab}=-Z_-^{\tt ba}$ are central charges and we define \be E^+=\pmatrix{0&0\cr 0&-\epsilon_{\dot\alpha\dot\beta}} \qquad\qquad E^-= \pmatrix{\epsilon_{\alpha\beta}&0\cr0&0}\ee The de Sitter generators (\ref{dstrgnrtrs}) now contain a sum over ${\tt a}=1,...N$. Upon writing \be \bar \Theta^{\tt a}=\frac 1{\sqrt{\Lambda}}\pmatrix{\Lambda\theta^{\tt a} \;&\;\bar Q^{\tt a} +i\theta^{\tt a}\sigma^n p_n }\qquad\qquad \Theta^{\tt a}=\frac 1{\sqrt{\Lambda}}\pmatrix{-Q^{\tt a}+ip_n\sigma^n\bar\theta^{\tt a}\cr\Lambda\bar\theta^{\tt a} }\;,\ee and substituting into (\ref{xtnddCapthtbrkts}), we recover the extended supersymmetry algebra \beqa \{Q^{\tt a}_\alpha,\bar Q^{\tt b}_{\dot \alpha}\}&=& 2{\sigma_{\alpha\dot\alpha}}^m p_m\delta^{\tt ab}\cr &&\cr\{Q^{\tt a}_\alpha,Q^{\tt b}_\beta\}&=&2\epsilon_{\alpha\beta}Z_-^{\tt ab} \cr &&\cr\{\bar Q^{\tt a}_{\dot \alpha},\bar Q^{\tt b}_{\dot \beta}\}&=&-2\epsilon_{\dot\alpha\dot\beta}Z_+^{\tt ab}\;, \eeqa upon projecting to Minkowski space-time. From (\ref{xtnddCapthtbrkts}), it also follows that $\theta^{{\tt a}\alpha}$ and $\bar\theta^{{\tt a}\dot\alpha}$ are Grassmann odd variables, which are canonically conjugate, respectively, to $Q^{\tt a}_\alpha$ and $\bar Q^{\tt a}_{\dot\alpha}$, i.e., \be \{Q^{\tt a}_\alpha,\theta^{\beta^{\tt b}}\}=i\delta_{\alpha}^{\beta}\delta^{\tt ab}\qquad \quad \{\bar Q^{\tt a}_{\dot \alpha},\bar\theta^{\dot\beta{\tt b}}\}=-i\delta_{\dot \alpha}^{\dot \beta}\delta^{\tt ab}\qquad\quad \{Q^{\tt a}_\alpha,\bar\theta^{\dot\beta{\tt b}}\}= \{\bar Q^{\tt a}_{\dot \alpha},\theta^{\beta{\tt b}}\}=0 \label{susyhisnbrg}\;\ee \bigskip {\Large {\bf Acknowledgments} } \noindent L. G. was supported by the High Energy Section of ICTP. A.S. was supported in part by the DOE, Grant No. DE-FG02-10ER41714. \bigskip
\section{Introduction} We consider the Cauchy problem for the following type of semilinear wave equations with small data: \begin{align} & \dal u= F(\pa u), & & (t,x) \in (0, \infty) \times \R^2, \label{WaveEq}\\ & u(0,x)=\eps f(x), \ \pa_t u(0,x)=\eps g(x), && x\in \R^2, \label{InitData} \end{align} where $u$ is a real-valued unknown function of $(t,x)\in [0, \infty)\times \R^2$, $\dal=\pa_t^2-\Delta=\pa_t^2-(\pa_1^2+\pa_2^2)$, and $\pa u=(\pa_0 u, \pa_1 u, \pa_2 u)$ with the notation $\pa_0=\pa_t=\pa/\pa t$, and $\pa_{j}=\pa/ \pa x_j$ ($j=1,2$). We suppose that $f,g \in C_0^{\infty}(\R^2;\R)$. $\eps$ is a small positive parameter. We assume that the nonlinear term $F=F(q)$ is a $C^\infty$ function of $q=(q_0, q_1, q_2)\in \R^3$. The local existence of classical solutions to \eqref{WaveEq}--\eqref{InitData} is well known, and we are interested in the sufficient condition for the global existence of the solutions, and also in the asymptotic behavior of global solutions. We say that the {\it small data global existence} (or SDGE) holds if for any $f,g\in C^\infty_0(\R^2)$ there is a positive constant $\eps_0$ such that \eqref{WaveEq}--\eqref{InitData} possesses a global solution $u\in C^\infty([0,\infty)\times \R^2)$ for $0<\eps\le \eps_0$. The case of cubic nonlinearity, that is to say $F(q)=O(|q|^3)$ near $q=0$, is critical in two space dimensions, because SDGE holds for some nonlinearity and fails for others. For example, SDGE does not hold for $F=(\pa_t u)^3$, however SDGE holds when $F$ is cubic and the {\it null condition for cubic terms} (we refer to it as the {\it cubic null condition}) is satisfied (see Godin~\cite{God93}). To be more specific, we say that $F$ satisfies the cubic null condition if $F^{(c)}(-1, \omega_1, \omega_2)=0$ for any $\omega=(\omega_1, \omega_2)\in \Sph^1$, where $F^{(c)}$ denotes the cubic homogeneous part of $F$, that is, $$ F^{(c)}(q)=\lim_{\lambda \to +0} \lambda^{-3} F(\lambda q). $$ The typical example satisfying the cubic null condition is $$ F(\pa u)=\sum_{a=0}^2 c_a(\pa_a u)\left(|\pa_t u|^2-|\nabla_x u|^2\right) $$ with arbitrary constants $c_a$. The null condition was first introduced for systems of nonlinear wave equations with quadratic nonlinearity in three space dimensions as a sufficient condition to ensure SDGE (see Klainerman~\cite{Kla86} and Christodoulou~\cite{Chr86}; note that the case of quadratic nonlinearity is critical in three space dimensions). The terms satisfying the null condition form an important class of nonlinearity. We do not go into details, but the global existence under the null condition for the quasi-linear systems, even with quadratic nonlinearity, in two space dimensions is also studied by many authors (see \cite{Ali01:01}, \cite {Ali01:02}, \cite{Hos95}, \cite{Kat93}, and \cite{Kat95} for example). Another important class of nonlinearity is the {\it nonlinear dissipation}. To explain the situation clearly, we consider the following equation in general space dimensions: \begin{align} & \dal u= -|\pa_t u|^{p-1} (\pa_t u), & & (t,x) \in (0, \infty) \times \R^n, \label{DisWaveEq}\\ & u(0,x)=\varphi(x), \ \pa_t u(0,x)=\psi(x), && x\in \R^n, \label{DisInitData} \end{align} where $p>1$. It is known that there is a global solution $u\in C\bigl([0,\infty); H^1(\R^n)\bigr)\cap C^1\bigl([0,\infty); L^2(\R^n)\bigr)$ for $(\varphi, \psi)\in X_p:=H^2(\R^n)\times(H^1(\R^n)\cap L^{2p}(\R^n))$ (see Lions--Strauss~\cite{LioStr65} for instance). Here no smallness of the data is required. We define the energy norm by $$ \|u(t)\|_E^2:=\frac{1}{2}\int_{\R^n} \left(|\pa_t u(t,x)|^2+|\nabla_x u(t,x)|^2\right)dx. $$ Mochizuki-Motai~\cite{MocMot95} proved that if $n\ge 1$ and $1<p\le 1+2/n$, then the energy decays to zero, namely $\lim_{t\to\infty} \|u(t)\|_E=0$ for initial data belonging to a dense subset of $X_p$ (see also Todorova-Yordanov~\cite{TodYor07}). On the other hand, in the case of $n\ge 2$ and $p>1+2/(n-1)$, it is also proved in \cite{MocMot95} that the energy does not decay for a class of small initial data in $X_p$. To sum up, the result in \cite{MocMot95} for $n=2$ can be read as follows: The energy decays for initial data in a dense subset of $X_p$ if $1<p\le 2$, while the energy does not decay for small initial data if $p>3$. There is a gap in the conditions for $p$, and it is quite interesting to investigate what happens for \eqref{DisWaveEq} with $n=2$ and $p=3$ (or equivalently \eqref{WaveEq} with $F(\pa u)=-|\pa_t u|^2\pa_t u$). The result above suggests that SDGE holds for the nonlinearity $F(\pa u)=-|\pa_t u|^2\pa_t u$, though it does not satisfy the cubic null condition. Hence it is natural to expect that there is a sufficient condition for SDGE which is weaker than the cubic null condition and includes also the nonlinear dissipative terms. Agemi conjectured that the condition \begin{equation} \label{AHK} F^{(c)}(-1, \omega_1,\omega_2)\ge 0, \quad \omega=(\omega_1,\omega_2)\in \Sph^1 \end{equation} implies SDGE. This conjecture was proved to be true by Hoshiga~\cite{Hos08} and Kubo~\cite{Kub07} independently. Moreover some asymptotic pointwise behavior of global solutions under \eqref{AHK} is obtained in \cite{Kub07} (see also Hayashi-Naumkin-Sunagawa~\cite{HayNauSun08} and Sunagawa~\cite{Sun06} for related results on nonlinear Schr\"odinger equations and nonlinear Klein-Gordon equations, respectively). Let us restrict our attention to the case of the nonlinear dissipative term $F(\pa u)=-|\pa_t u|^2\pa_t u$. Then, it follows from \cite{Kub07} that $$ |\pa u(t,x)|\le Cr^{-1/2} \left(\log \frac{r}{|r-t-1|+1}\right)^{-1/2}, \quad r\ge \frac{t}{2}+1 $$ with some positive constant $C$, where $r=|x|$. This estimate is improved in \cite{Mur11} as follows: $$ |\pa u(t,x)|\le Ct^{-1/2} \min\left\{\left(\log t \right)^{-1/2}, \eps (1+|r-t|)^{-\vartheta}\right\} $$ for $(t,x)\in [2,\infty)\times \R^2$ with $0<\vartheta<1/2$. However this is still insufficient in order to say something about the decay of the energy. In this paper we will further improve the pointwise estimate of $\pa u$ under the condition \eqref{AHK}, and show that the energy decays to zero if we consider \eqref{WaveEq} with $F(\pa u)=-|\pa_t u|^2\pa_t u$ (or \eqref{DisWaveEq} with $n=2$ and $p=3$ in other words) for small $C^\infty_0$-data (in fact, complex-valued solutions will be considered). \section{The main result and its applications} \subsection{Global existence and asymptotic pointwise behavior} In what follows, we consider the initial value problem \eqref{WaveEq}--\eqref{InitData} for complex-valued data with the nonlinearity \begin{equation} \label{CExp} F(\pa u)= \sum_{a,b,c=0}^2 p_{abc} (\pa_a u)(\pa_b u)(\overline{\pa_c u}) \end{equation} with complex constants $p_{abc}$ in order that we can catch up two kinds of interesting nonlinearities $F(\pa u)=-|\pa_t u|^2(\pa_t u)$ and $F(\pa u)=i |\pa_t u|^2 (\pa_t u)$ (see Subsections~\ref{App02} and \ref{App03} below). Here and hereafter, $\overline{z}$ denotes the complex conjugate of $z \in \C$ and the symbol $i$ always stands for $\sqrt{-1}$. We also use the notation $\jb{z}=\sqrt{1+|z|^2}$. The following theorem is our main result. \begin{thm}\label{Global} Assume that \eqref{CExp} is satisfied. We also assume \begin{equation} \label{ComplexAHK} \realpart F(\hat{\omega}) \ge 0, \quad \hat{\omega} (\omega_0,\omega)\in \{-1\}\times \Sph^1. \end{equation} Let $\mu$ be a sufficiently small positive constant satisfying $0<\mu<1/10$, say. Then, for any $f,g\in C^\infty_0(\R^2;\C)$, there exists a positive constant $\eps_0$ such that \eqref{WaveEq}--\eqref{InitData} admits a unique global classical solution $u\in C^\infty\bigl([0,\infty)\times\R^2; \C\bigr)$ for any $\eps \in (0,\eps_0]$. Moreover, there is a function $P_0$ of $(\sigma, \omega)\in \R\times \Sph^1$, which satisfies \begin{align} \label{Concl02} |P_0(\sigma, \omega)|\le C\eps \jb{\sigma}^{\mu-1},\quad (\sigma, \omega)\in \R\times \Sph^1 \end{align} with a positive constant $C$, such that \begin{align} \label{Concl01} \pa u(t,x)=\hat{\omega}(x) t^{-1/2} P(\log t, |x|-t, |x|^{-1}x) +O\bigl(\eps t^{4\mu-(3/2)}\jb{t-|x|}^{-3\mu}\bigr) \end{align} for any $(t,x)\in [1,\infty)\times \R^2$, where $\hat{\omega}(x)=(-1, |x|^{-1}x_1, |x|^{-1} x_2)$ for $x=(x_1, x_2)\in \R^2$, and $P=P(\tau,\sigma, \omega)$ is defined as a solution to \begin{align} \begin{cases} \pa_{\tau} P(\tau,\sigma, \omega) =-\left({F(\hat{\omega})}/2\right)|P(\tau,\sigma,\omega)|^2 P(\tau,\sigma, \omega), & \tau>0, \\ P(0,\sigma,\omega)=P_0(\sigma,\omega)& \end{cases} \label{ode} \end{align} with $\hat\omega=(-1,\omega)$. Here the constant $C$ is independent of $\eps$. \end{thm} \begin{rmk} (1) It is well known that if $f=g=0$ for $|x|\ge R$, then $u(t,x)=0$ for $|x|\ge t+R$ (see H\"ormander~\cite{Hoe97} for instance). Hence $t$ is equivalent to $\jb{t+|x|}$ in $\supp u(t,\cdot)$ when $t\ge 1$. \\ (2) $P(\tau,\sigma,\omega)$ in the above theorem can be explicitly solved as \begin{equation} \label{ExpS} P(\tau,\sigma,\omega)= P_0(\sigma,\omega) \frac{ \exp\left(-i\Theta(\tau,\sigma, \omega) \right) } {\sqrt{1+\left(\realpart{F}(\hat{\omega})\right) |P_0(\sigma,\omega)|^2\tau}} \end{equation} with \begin{equation} \label{Phase} \Theta(\tau,\sigma, \omega)=\frac{1}{2}\left(\imagpart F(\hat{\omega})\right) \int_{0}^{\tau} \frac{|P_0(\sigma, \omega)|^2}{{1+\left(\realpart F(\hat{\omega})\right)|P_0(\sigma,\omega)|^2\tau'}}d\tau'. \end{equation} From \eqref{ExpS}, we have \begin{equation} \label{ExpS1} |P(\tau,\sigma, \omega)|\le \frac{|P_0(\sigma,\omega)|}{\sqrt{1+\left(\realpart F(\hat{\omega})\right) |P_0(\sigma,\omega)|^2\tau}}. \end{equation} (3) We can add higher order nonlinear terms to \eqref{CExp}, but the result becomes slightly weaker from the viewpoint of the estimate for the remainder term: Theorem~\ref{Global} remains valid if we replace \eqref{CExp} by $$ F(\pa u) =\sum_{a,b,c=0}^2 p_{abc}(\pa_a u)(\pa_b u)\left(\overline{\pa_c u}\right) +O(|\pa u|^4)\quad \text{ near $\pa u=0$}, $$ $F$ in \eqref{ComplexAHK} as well as in \eqref{ode} by $F^{(c)}$, and $O(\eps t^{4\mu-(3/2)} \jb{t-|x|}^{-3\mu})$ in \eqref{Concl01} by $O(\eps t^{\mu-1}\jb{t-|x|}^{-1/2})$. \qed \end{rmk} We will prove Theorem~\ref{Global} in Section~\ref{Proof0101} after some preparation given in Sections~\ref{Preliminary}, \ref{Reduction} and \ref{LemmaODE}. Compared to the method by Kubo~\cite{Kub07}, the most different point in our proof is the choice of the equation \eqref{ode} for the asymptotics. Careful treatment of the factor $\jb{t-|x|}$ is also quite important in our improvement. In the following subsections, we discuss what we can see from Theorem~\ref{Global} particularly in the cases of $F(\pa u)= -|\pa_t u|^2\pa_t u$ (Subsection~\ref{App02}) and $F(\pa u)= i|\pa_t u|^2\pa_t u$ (Subsection~\ref{App03}). \subsection{The case of nonlinear dissipation: The decay of the energy} \label{App02} We focus on the case where the inequality in \eqref{ComplexAHK} is strict, i.e., \begin{equation} \label{DampingAss} C_0:=\min_{\omega\in \Sph^1} \realpart F(\hat{\omega})>0. \end{equation} The typical example satisfying \eqref{DampingAss} is $F(\pa u)=-|\pa_t u|^2(\pa_t u)$. In this case, it follows from \eqref{Concl02} and \eqref{ExpS1} that $$ |P(\tau,\sigma, \omega)|\le \min\left\{\frac{1} {\sqrt{C_0 \tau}}, C\eps \jb{\sigma}^{\mu-1}\right\}. $$ Hence, by \eqref{Concl01} we can find a positive constant $C$ such that \begin{equation} \label{DecayDamping} |\pa u(t,x)|\le C t^{-1/2}\min\left\{(\log t)^{-1/2}, \eps \jb{t-|x|}^{\mu-1}\right\} \end{equation} for $(t,x)\in [2,\infty)\times \R^2$. This estimate says that $\pa u$ decays like $(t \log t)^{-1/2}$ along the line $l_\sigma:=\{(t,x);\, |x|-t=\sigma\}$ for each $\sigma\in \R$. On the other hand, for the solution $u_0$ to the free wave equation $\dal u_0=0$ with $C^\infty_0$-data, it is known that $\pa u_0$ decays at the rate of $t^{-1/2}$ along $l_\sigma$. Hence \eqref{DecayDamping} tells us that $\pa u$ has a gain of $(\log t)^{-1/2}$ in the pointwise decay compared to $\pa u_0$. Moreover, from \eqref{DecayDamping} we obtain the following decay of the energy: \begin{cor}\label{EnergyDecay} Suppose that \eqref{CExp} and \eqref{DampingAss} are fulfilled, and let $0<\mu<1/10$. Then for any $f, g\in C^\infty_0(\R^2; \C)$, the global solution $u$ to \eqref{WaveEq}--\eqref{InitData} satisfies $$ \|u(t)\|_E^2\le C\eps^{\frac{1}{1-\mu}}(\log t)^{-\frac{1-2\mu}{2-2\mu}},\quad t\ge 2 $$ with some positive constant $C$, provided that $\eps$ is sufficiently small. \end{cor} The proof for this result will be given in Section~\ref{Omake}. This corollary says that the energy non-decay result of Mochizuki-Motai~\cite{MocMot95} fails for $n=2$ and $p=3$. It also suggests that the energy decay result in \cite{MocMot95} holds also for $n=2$ and $2< p\le 3$, but this is still an open problem (note that the energy decay for $n=2$ and $p=3$ established here is only for small $C^\infty_0$-data). \subsection{The case without dissipation: Logarithmic correction of the phase.} \label{App03} We say that the global solution $u$ to \eqref{WaveEq}--\eqref{InitData} is asymptotically free (in the energy sense) if there is $(\varphi_+, \psi_+) \in \dot{H}^1(\R^2)\times L^2(\R^2)$ such that $$ \lim_{t\to \infty} \|u(t)-u_+(t)\|_E=0, $$ where $u_+$ is the solution to the free wave equation $\dal u_+=0$ in the energy class with initial data $(u_+,\pa_t u_+)=(\varphi_+,\psi_+)$ at $t=0$. Here $\dot{H}^1(\R^2)$ is the completion of $C^\infty_0(\R^2)$ with respect to the norm $\|\phi\|_{\dot{H}^1(\R^2)}=\|\nabla\phi\|_{L^2(\R^2)}$. It is proved in Katayama~\cite{Kat11b} (see also \cite{Kat11a}) that $u$ is asymptotically free if and only if there is $U=U(\sigma,\omega)\in L^2(\R\times \Sph^1)$ such that $$ \lim_{t\to\infty} \bigl\|\pa u(t,\cdot)-\hat{\omega}(\cdot)\widetilde{U}(t,\cdot) \bigr\|_{L^2(\R^2)}=0, $$ where $\hat{\omega}(x)=(-1, x_1/|x|, x_2/|x|)$ and $\widetilde{U}(t,x)=|x|^{-1/2} U(|x|-t, x/|x|)$. Now we consider the case of \begin{equation} \realpart F(\hat{\omega})=0,\quad \omega\in \Sph^1, \end{equation} which is stronger than \eqref{ComplexAHK} but weaker than the cubic null condition. In this case, $P$ can be written as \begin{equation}\label{ExpSS} P(\tau, \sigma, \omega)=P_0(\sigma, \omega) \exp\left( -i\widetilde\Theta(\sigma, \omega) \tau\right) \end{equation} with $\widetilde\Theta(\sigma, \omega)= \left(\imagpart F(\hat\omega)\right)|P_0(\sigma, \omega)|^2/2(\in \R)$. Hence we find from \eqref{Concl01} that $\pa u$ decays at the same rate of $t^{-1/2}$ as the derivatives of the free solution $u_0$ along the line $l_\sigma$ for each $\sigma$. By \eqref{Concl02}, \eqref{Concl01}, and \eqref{ExpSS}, we get \begin{align} \bigl\| \pa u(t,\cdot)-\hat{\omega}(\cdot)\widetilde{P}(t, \cdot) \bigr\|_{L^2(\R^2)} \le & C\eps t^{2\mu-(1/2)} \bigl\||\cdot|^{-1/2}\jb{t-|\cdot|}^{-\mu-(1/2)}\bigr\|_{L^2(\R^2)} \nonumber\\ \le & C\eps t^{2\mu-(1/2)}\to 0\quad (t\to \infty), \end{align} where $\widetilde{P}(t, x)=|x|^{-1/2}P(\log t, |x|-t, x/|x|)$. Moreover \eqref{Concl02} and \eqref{ExpSS} lead to $$ \lim_{t\to\infty}\|P(\log t,\cdot,\cdot)\|_{L^2((-\infty, -t)\times \Sph^1)}=0. $$ Therefore we find that $u$ is asymptotically free if and only if there is $U=U(\sigma,\omega)\in L^2(\R\times\Sph^1)$ such that $$ \lim_{\tau\to\infty} \|P(\tau,\cdot,\cdot)-U(\cdot,\cdot)\|_{L^2(\R\times\Sph^1)}=0. $$ If we assume $\imagpart F(\hat{\omega})=0$ for any $\omega\in \Sph^1$ in addition, then $P(\tau,\sigma,\omega)=P_0(\sigma, \omega)$. Furthermore we see from \eqref{Concl02} that $P_0\in L^2(\R\times \Sph^1)$. Hence we conclude that the solution $u$ is asymptotically free. For this case, we have $F(\hat{\omega})\equiv0$ and the cubic null condition (for the complex case) is satisfied. Typical examples are $F=(\pa_a u)\bigl(|\pa_t u|^2-|\pa_1u|^2-|\pa_2u|^2\bigr)$ or $(\overline{\pa_a u})\bigl((\pa_t u)^2-(\pa_1u)^2-(\pa_2u)^2\bigr)$, as well as $F=(\pa_a u)\bigl((\pa_b u)(\overline{\pa_c u})-(\pa_c u)(\overline{\pa_b u})\bigr)$ for $a,b,c=0,1,2$. The situation is different if $\imagpart F(\hat\omega)\not\equiv 0$. For example, take $F(\pa u)=i|\pa_t u|^2(\pa_t u)$ so that $F(\hat{\omega})\equiv -i$. Then it is easy to show $\|u(t)\|_E=\|u(0)\|_E$. By \eqref{ExpSS} we get $$ P(\tau,\sigma,\omega)=P_0(\sigma,\omega)\exp\left(\frac{i}{2}|P_0(\sigma,\omega)|^2\tau \right), $$ and we can easily see that $P(\tau)$ does not converge to any function in $L^2(\R\times \Sph^1)$ as $\tau\to \infty$ unless $P_0\equiv 0$. Because of the conservation of the energy, we can show that if $P_0\equiv 0$ then $(f,g)\equiv (0,0)$. Hence we see that the global solution $u$ for small $\eps$ is not asymptotically free unless $(f, g)\equiv (0,0)$, though the energy is preserved. Such a phenomenon never occurs in the real-valued case. \section{Preliminaries}\label{Preliminary} We introduce \begin{align*} S:=& t\pa_t+\sum_{j=1}^2 x_j \pa_j \\ L=&(L_1,L_2):=(t\pa_1+x_1\pa_t, t\pa_2+x_2\pa_t), \ \Omega:= x_1\pa_2-x_2\pa_1, \end{align*} and we set $$ \Gamma=(\Gamma_0,\Gamma_1, \ldots, \Gamma_6)=(S, L_1, L_2, \Omega, \pa_0,\pa_1,\pa_2). $$ With a multi-index $\alpha=(\alpha_0, \alpha_1, \ldots, \alpha_6)\in \ZP^7$, we write $\Gamma^\alpha=\Gamma_0^{\alpha_0}\Gamma_1^{\alpha_1}\cdots \Gamma_6^{\alpha_6}$, where $\ZP$ denotes the set of nonnegative integers. For a smooth function $\psi=\psi(t,x)$ and a nonnegative integer $s$, we define $$ |\psi(t,x)|_s=\sum_{|\alpha|\le s} |\Gamma^\alpha \psi(t,x)|, \quad \|\psi(t)\|_s=\sum_{|\alpha|\le s} \|\Gamma^\alpha \psi(t,\cdot)\|_{L^2(\R^2)}. $$ It is easy to see that $[\dal, \Omega]=[\dal, L_j]=[\dal, \pa_a]=0$ for $j=1,2$ and $a=0,1,2$, where $[A,B]=AB-BA$ for the operators $A$ and $B$. We also have $[\dal, S]=2\dal$. Therefore for any $\alpha=(\alpha_0,\alpha_1,\ldots,\alpha_6)\in \ZP^7$ and a smooth function $\psi$, we have \begin{equation} \label{Comm01} \dal \Gamma^\alpha \psi=(\Gamma_0+2)^{\alpha_0}\Gamma_1^{\alpha_1}\cdots\Gamma_6^{\alpha_6} \dal \psi=: \widetilde\Gamma^\alpha \dal \psi. \end{equation} We can check that we have $[\Gamma_a, \Gamma_b]=\sum_{0\le c\le 7} A^{ab}_c\Gamma_c$ and $[\Gamma_a, \pa_b]=\sum_{0\le c\le 2} B^{ab}_c \pa_c$ with appropriate constants $A^{ab}_c$ and $B^{ab}_c$. Hence for any $\alpha, \beta\in \ZP^7$, and any nonnegative integer $s$, there exist positive constants $C_{\alpha,\beta}$ and $C_s$ such that \begin{align} \label{Comm02} & |\Gamma^\alpha\Gamma^\beta\psi(t,x)|\le C_{\alpha,\beta} |\psi(t,x)|_{|\alpha|+|\beta|},\\ \label{Comm03} & C_s^{-1} |\pa \psi(t,x)|_s\le \sum_{0\le a\le 2}\sum_{|\gamma|\le s} |\pa_a\Gamma^\gamma\psi(t,x)|\le C_s|\pa \psi(t,x)|_s \end{align} for any smooth function $\psi$. Using these vector fields, we obtain a good decay estimate for the solution to the inhomogeneous wave equation. Let $0<T\le \infty$. \begin{lem}[H\"ormander's $L^1$--$L^\infty$ estimate]\label{Ho} Let $v$ be a smooth solution to $$ \dal v(t,x)=\Psi(t,x), \quad (t,x)\in (0, T)\times \R^2 $$ with initial data $v=\pa_t v=0$ at $t=0$. Then there exists a universal positive constant $C$ such that $$ \jb{t+|x|}^{1/2}|v(t,x)|\le C\int_0^t \left(\int_{\R^2} \frac{|\Psi(\tau, y)|_1}{\jb{\tau+|y|}^{1/2}}dy\right) d\tau, \quad 0\le t<T. $$ \end{lem} See H\"ormander \cite{Hoe88} for the proof (similar estimates for arbitrary space dimensions are also available there). The vector fields in $\Gamma$ also play an important role in the reduction of the analysis of semilinear wave equations to that of the corresponding ordinary differential equations (or ODEs in short). We use the polar coordinates $(r,\theta)$ to write $x=(x_1,x_2)=(r\cos \theta, r\sin \theta)$ with $r=|x|$ and $\theta\in \R$. We put $\omega=(\omega_1,\omega_2)=x/|x|=(\cos\theta, \sin\theta)$. Then we have $\pa_r=\sum_{j=1}^2\omega_j\pa_j$, and $\pa_\theta=x_1\pa_2-x_2\pa_1=\Omega$. We also have \begin{equation} \pa_1=\omega_1\pa_r-\frac{\omega_2}{r}\pa_\theta,\quad \pa_2=\omega_2\pa_r+\frac{\omega_1}{r}\pa_\theta. \label{f001} \end{equation} We put $$ \pa_\pm=\pa_t\pm \pa_r. $$ Then, for a smooth function $\psi$, we get \begin{equation} \label{pcd} r^{1/2}\dal \psi =\pa_+\pa_-(r^{1/2}\psi)-r^{-3/2}\left(\pa_\theta^2\psi+\frac{1}{4}\psi\right). \end{equation} Since $$ \frac{1}{r}\pa_\theta=\frac{1}{r}(x_1\pa_2-x_2\pa_1)=\frac{1}{t}(\omega_1 L_2-\omega_2 L_1), $$ it follows from \eqref{f001} that \begin{equation} \label{L202} \sum_{j=1}^2|\pa_j\psi(t,x)-\omega_j\pa_r\psi(t,x)|\le \left|\frac{1}{r}\pa_\theta\psi(t,x)\right| \le C\jb{t+r}^{-1}|\psi(t,x)|_1 \end{equation} for any $\psi\in C^1([0, T)\times \R^2)$. We put $\omega_0=-1$ and $\hat\omega=(\omega_0,\omega_1,\omega_2)$. For simplicity of exposition, we introduce $$ D_\pm=\pm \frac{1}{2} \pa_{\pm}=\frac{1}{2} (\pa_r\pm \pa_t). $$ \begin{lem}\label{Rewrite01} There exists a positive constant $C$ such that \begin{equation} \label{f203} |\pa \psi(t,x)-\hat \omega D_-\psi(t,x)| \le C \jb{t+r}^{-1}|\psi(t,x)|_1 \end{equation} for $\psi\in C^1([0, T)\times\R^2)$ and $(t,x)\in [0, T)\times (\R^2\setminus\{0\})$. \end{lem} \begin{proof} Because $S=t\pa_t+r\pa_r$, if we put $L_r=t\pa_r+r\pa_t=\sum_{j=1}^2 \omega_j L_j$, then we get $$ D_+=\frac{1}{2}(\pa_r+\pa_t)=\frac{1}{2(t+r)} (S+L_r). $$ Since we have $\pa_t=-D_-+D_+$ and $\pa_r=D_-+D_+$, we find that $|\pa_t \psi(t,x)-\omega_0 D_-\psi(t,x)|$ and $|\pa_r \psi(t,x)-D_-\psi(t,x)|$ are bounded by $C \jb{t+r}^{-1} |\psi(t,x)|_1$. Hence we obtain \eqref{f203} with the help of \eqref{L202}. \end{proof} We define $$ \Lambda_T:=\{(t,x)\in [0, T)\times \R^2; r\ge t/2\ge 1\}. $$ Observe that we have $r^{-1}\le 4(1+t+r)^{-1}$ for $(t,x)\in \Lambda_T$. We compute $D_-(r^{1/2}\psi)=r^{1/2}D_-\psi+r^{-1/2}\psi/4$. Thus from \eqref{f203} we obtain the following. \begin{cor}\label{Rewrite02} There is a positive constant $C$ such that $$ |r^{1/2}\pa \psi(t,x)-\hat\omega D_-\bigl(r^{1/2}\psi(t,x)\bigr)| \le C \jb{t+r}^{-1/2}|\psi(t,x)|_1 $$ for $(t,x)\in \Lambda_T$ and $\psi\in C^1([0, T)\times \R^2)$. \end{cor} The following lemma is due to Lindblad \cite{Lin90}. \begin{lem}\label{Rewrite03} For any nonnegative integer $s$, there exists a positive constant $C_s$ such that \begin{equation} \label{L201} |\pa \psi(t,x)|_s\le C_s \jb{t-r}^{-1}|\psi(t,x)|_{s+1},\quad (t,x)\in [0, T)\times \R^2 \end{equation} for any $\psi\in C^{s+1}([0, T)\times \R^2)$. \end{lem} \begin{proof} In view of \eqref{Comm02} and \eqref{Comm03}, it suffices to prove \eqref{L201} for $s=0$. As in the proof of Lemma~\ref{Rewrite01}, we put $L_r=t\pa_r+r\pa_t$. Then we have $$ (t-r)\pa_t\psi=\frac{1}{t+r}\left(tS-rL_r\right)\psi,\quad (t-r)\pa_r\psi=\frac{1}{t+r} \left(tL_r-rS\right)\psi. $$ Hence we get $\jb{t-r}(|\pa_t\psi|+|\pa_r\psi|)\le C|\psi|_1$, which leads to \eqref{L201} for $s=0$ with the help of \eqref{L202}. \end{proof} \section{Reduction to simplified equations} \label{Reduction} Let $0<T\le \infty$, and let $u$ be the solution to \eqref{WaveEq} on $[0, T)\times \R^2$. We suppose that \begin{equation} \label{Supp01} \supp f\cup \supp g \subset B_R \end{equation} for some $R>0$, where $B_M=\{x\in \R^2;|x|\le M\}$ for $M>0$. Then, from the finite propagation property, we have \begin{equation} \label{Supp02} \supp u(t,\cdot)\subset B_{t+R},\quad 0\le t<T. \end{equation} In what follows, we put $r=|x|$, $\omega=(\omega_1, \omega_2)=|x|^{-1}x$, and $\omega_0=-1$. We write $\hat{\omega}=(\omega_0,\omega)=(-1,\omega)$. We define $$ U(t, x):=D_-\bigl(r^{1/2} u(t, x) \bigr),\quad (t, x)\in [0,T)\times (\R^2\setminus\{0\}). $$ From \eqref{WaveEq} and \eqref{pcd}, we get \begin{equation} \label{Red01} \pa_+U(t, x)= -\frac{1}{2t}F(\hat{\omega}) |U(t, x)|^2U(t, x)+H(t, x), \end{equation} where $H=H(t,x)$ is given by \begin{align*} H=& -\frac{1}{2} \left(r^{1/2}F(\pa u)-t^{-1}F(\hat{\omega})|U|^2U\right) {}-\frac{1}{2r^{3/2}}\left(\pa_\theta^2 u+\frac{1}{4}u\right). \end{align*} \eqref{Red01} plays an important role in our analysis. We also need the following to estimate the generalized derivative $\Gamma^\alpha u$ for a multi-index $\alpha=(\alpha_0, \alpha_1, \ldots, \alpha_6)\in \ZP^7$: \eqref{Comm01} and \eqref{WaveEq} yield $\dal(\Gamma^\alpha u)=\widetilde{\Gamma}^\alpha\left(F(\pa u)\right)$, which leads to \begin{equation} \label{Red02} \pa_+U^{(\alpha)}= -\frac{1}{2t}F(\hat{\omega}) \left( 2|U|^2U^{(\alpha)}+U^2\overline{U^{(\alpha)}}\right) +H_\alpha, \end{equation} where we have put $$ U^{(\alpha)}(t,x):= D_-\bigl(r^{1/2}\Gamma^\alpha u(t,x)\bigr),\\ $$ and $H_\alpha=H_\alpha(t,x)$ is given by \begin{align*} H_\alpha=& -\frac{1}{2} \left(r^{1/2}\widetilde{\Gamma}^\alpha F(\pa u)- t^{-1}F(\hat{\omega})\bigl(2|U|^2U^{(\alpha)}+U^2 \overline{U^{(\alpha)}}\bigr)\right) \\ & {}-\frac{1}{2r^{3/2}}\left(\pa_\theta^2 \Gamma^\alpha u+\frac{1}{4} \Gamma^\alpha u\right). \end{align*} Note that $U=U^{(\alpha)}$ if $|\alpha|=0$. We put \begin{equation} \label{DefLamTR} \Lambda_{T,R}:=\{(t,x)\in [0, T)\times \R^2;\, 1\le t/2\le r\le t+R\}. \end{equation} Note that we have $$ (1+t+r)^{-1}\le r^{-1}\le 2t^{-1}\le 3(1+t)^{-1}\le 3(R+2)(1+t+r)^{-1} $$ for $(t,x)\in \Lambda_{T,R}$. In other words, the weights $\jb{t+r}^{-1}$, $(1+t)^{-1}$, $r^{-1}$, and $t^{-1}$ are equivalent to each other in $\Lambda_{T,R}$. For nonnegative integer $s$, we define \begin{equation} \label{Basis300} |u(t,x)|_{\sharp,s}:=|\pa u(t,x)|_s+\jb{t+r}^{-1}|u(t,x)|_{s+1}. \end{equation} For $|\alpha|\le s$, Corollary~\ref{Rewrite02} and \eqref{Comm03} immediately imply that \begin{equation} |U^{(\alpha)}(t,x)|\le C_s r^{1/2}|u(t,x)|_{\sharp,s} \label{Basic301} \end{equation} for $(t,x)\in \Lambda_{T, R}$ with some positive constant $C_s$. Our final goal in this section is to prove the following. \begin{prp}\label{Pro301} Suppose that \eqref{CExp} is fulfilled. Let $s$ be a positive integer, and suppose that $1\le |\alpha|\le s$. Then there are positive constants $C$ and $C_s$ such that \begin{align} \label{Basic321} |H(t,x)| \le & C \left(t^{-1/2}|u|_{\sharp,0}^2|u|_1+t^{-3/2}|u|_2 \right),\\ |H_\alpha(t,x)|\le & C_s (t^{-1/2}|u|_{\sharp,s}^2|u|_{s+1} +r^{1/2}|\pa u|_{s-1}^3+t^{-3/2}|u|_{s+2}) \label{Basic322} \end{align} for $(t,x)\in \Lambda_{T,R}$. \end{prp} Before we proceed to the proof of Proposition~\ref{Pro301}, we show one lemma. For $a,b,c=0,1,2$, we define $$ Q_{abc}:=r^{1/2}(\pa_a u)(\pa_b u)(\overline{\pa_c u})-r^{-1} \omega_a\omega_b\omega_c |U|^2U. $$ For a multi-index $\alpha\in \ZP^7$, we also define $$ \widetilde{Q}_{abc}^{(\alpha)} :=r^{1/2}\Gamma^\alpha\left((\pa_a u)(\pa_b u)(\overline{\pa_c u})\right) -r^{-1}\omega_a\omega_b\omega_c\left(2|U|^2U^{(\alpha)}+U^2\overline{U^{(\alpha)}}\right). $$ \begin{lem}\label{L302} Let $|\alpha|\le s$. Then there are positive constants $C$ and $C_s$ such that \begin{align} \label{L3021} |Q_{abc}|\le & C t^{-1/2}|u|_{\sharp,0}^2|u|_1,\\ \label{L3022} |\widetilde{Q}_{abc}^{(\alpha)}|\le & C_s (t^{-1/2} |u|_{\sharp, s}^2|u|_{s+1} {}+r^{1/2}|\pa u|_{s-1}^3) \end{align} for $(t,x)\in \Lambda_{T,R}$. \end{lem} \begin{proof} Suppose that $(t,x)\in \Lambda_{T,R}$. By \eqref{Basic301} and Corollary~\ref{Rewrite02}, we get \begin{align} |Q_{abc}|=& r^{-1}\bigl( (r^{1/2}\pa_a u)(r^{1/2}\pa_b u)(\overline{r^{1/2}\pa_c u}) {}-\omega_a\omega_b\omega_c|U|^2U\bigr) \nonumber\\ \le & Cr^{-1}(|r^{1/2}\pa u|+|U|)^2|r^{1/2}\pa u-\hat \omega U| \nonumber\\ \le & Ct^{-1/2}|u|_{\sharp,0}^2|u|_1. \label{Basic311} \end{align} Let $|\alpha|\le s$. In view of \eqref{Comm02} and \eqref{Comm03} we obtain \begin{equation} \label{Basic312} \left|\Gamma^\alpha\bigl( (\pa_a u)(\pa_b u)(\overline{\pa_c u})\bigr)-R_{abc}^{(\alpha)}\right| \le C_s|\pa u|_{[s/2]}^2|\pa u|_{s-1}\le C_s|\pa u|_{s-1}^3 \end{equation} with some positive integer $C_s$, where $$ R_{abc}^{(\alpha)}=(\pa_a\Gamma^\alpha u)(\pa_b u)(\overline{\pa_c u}) {}+(\pa_a u)(\pa_b \Gamma^\alpha u)(\overline{\pa_c u})+(\pa_a u)(\pa_b u)(\overline{\pa_c \Gamma^\alpha u}). $$ Similarly to \eqref{Basic311}, we get \begin{align} & |r^{1/2}R_{abc}^{(\alpha)}-r^{-1}\omega_a\omega_b\omega_c(2|U|^2U^{(\alpha)}+ U^2\overline{U^{(\alpha)}})| \le C_st^{-1/2} |u|_{\sharp,s}^2|u|_{s+1}. \label{Basic313} \end{align} The desired estimate \eqref{L3022} follows from \eqref{Basic312} and \eqref{Basic313}. \end{proof} \begin{proof}[Proof of Proposition~$\ref{Pro301}$] Suppose that $(t,x)\in \Lambda_{T,R}$. Recalling \eqref{CExp}, we get $r^{1/2}F(\pa u)-r^{-1}F(\hat{\omega})|U|^2U=\sum_{a,b,c} p_{abc} Q_{abc}$, and Lemma~\ref{L302} yields \begin{equation} \label{PL301} \left|r^{1/2}F(\pa u)-r^{-1}F(\hat{\omega})|U|^2U\right|\le Ct^{-1/2}|u|_{\sharp,0}^2|u|_1. \end{equation} Using \eqref{Basic301} and Lemma~\ref{Rewrite03} we get \begin{align} \left|(r^{-1}-t^{-1})F(\hat{\omega})|U|^2U\right|\le & Ct^{-1}r^{-1}\jb{t-r}r^{3/2}|u|_{\sharp,0}^3\nonumber\\ \le & Ct^{-1/2}|u|_{\sharp,0}^2 (\jb{t-r}|\pa u|+|u|_1) \nonumber\\ \le & Ct^{-1/2}|u|_{\sharp,0}^2 |u|_1 . \label{PL302} \end{align} It is clear that we have \begin{equation} \label{PL304} r^{-3/2}|\pa_\theta^2 u+u/4|\le Ct^{-3/2}|u|_2. \end{equation} Now \eqref{Basic321} follows from \eqref{PL301}, \eqref{PL302}, and \eqref{PL304}. Using \eqref{L3022} instead of \eqref{L3021}, we can show \eqref{Basic322} in a similar manner. \end{proof} \section{A key lemma for ordinary differential equations} \label{LemmaODE} Let $t_0\ge 1$. Motivated by \eqref{Red01}, we consider the initial value problem for the following kind of ODE: \begin{align} \label{ODE} & z'(t)=-\frac{K}{2t} |z(t)|^2z(t)+{J}(t), \quad t_0<t<\infty,\\ \label{ODEinit} & z(t_0)=z_0, \end{align} where $K, z_0 \in \C$, and ${J} \in C\bigl([t_0, \infty);\C\bigr)$. The following lemma is a refinement of the ones obtained in Hayashi-Naumkin-Sunagawa~\cite{HayNauSun08} and Sunagawa~\cite{Sun06}. \begin{lem}\label{ODELemma02} Let $\eps>0$, $\sigma\in \R$, $\rho>1$, $0<\mu<\rho-1$, and $\kappa\ge 0$. Suppose that $\realpart K\ge 0$. We assume that there are positive constants $K_0$ and $E_0$ such that \begin{align*} |K|\le & K_0,\\ |z_0|\le & E_0 \eps\jb{\sigma}^{-\kappa-\rho+1}, \\ |{J}(t)|\le & E_0 \eps \jb{\sigma}^{-\kappa} t^{-\rho},\quad t\ge t_0. \end{align*} We also assume that there is a positive constant $c_0$ such that \begin{equation} c_0^{-1}\jb{\sigma}< t_0<c_0\jb{\sigma}. \label{CondT0} \end{equation} Then \eqref{ODE}--\eqref{ODEinit} admits a unique global solution $z\in C^1([t_0,\infty);\C)$, and there is a positive constant $C_0=C_0(E_0, c_0, \rho)$ such that we have \begin{equation} \label{ODEBound} |z(t)|\le C_0\eps \jb{\sigma}^{-\kappa-\rho+1} ,\quad t_0\le t<\infty. \end{equation} Moreover there is a positive constant $\eps_1=\eps_1(K_0, E_0, c_0, \rho, \mu)$ such that if $0<\eps\le \eps_1$, then we can find $p_{0}\in \C$ satisfying \begin{align} \label{ODEaSymp} |z(t)-p(\log t)|\le & C_1 \eps \jb{\sigma}^{-\kappa-\mu}t^{-\rho+\mu+1},\quad t\ge t_0, \\ \label{ODEprof} |p_0|\le & C_1 \eps \jb{\sigma}^{-\kappa-\rho+1}, \end{align} where $p(t)$ is a solution to \begin{align} \left\{\begin{array}{ll} p'(\tau)=-\frac{K}{2}|p(\tau)|^2p(\tau),& \tau>0,\\ p(0)=p_{0}, \end{array}\right. \label{ReducedODE} \end{align} and $C_1=C_1(K_0, E_0, c_0, \rho, \mu)$ is a positive constant. \end{lem} \begin{proof} We put $\nu=\kappa+\rho-1(>0)$. Let $z$ be the solution to \eqref{ODE}--\eqref{ODEinit} for $t_0<t<T_0$ with some $T_0(>t_0)$. Since $\realpart K\ge 0$, it follows from \eqref{ODE} that $$ \frac{d}{dt}(|z(t)|^2)=-\frac{\realpart K}{t}|z(t)|^4 +2\realpart({J}(t)\overline{z(t)}) \le 2|J(t)|\,|z(t)|, $$ which yields \begin{align*} |z(t)|\le & |z_0|+ \int_{t_0}^t |{J}(\tau)|d\tau\le E_0\eps\left(\jb{\sigma}^{-\nu} + \jb{\sigma}^{-\kappa}\int_{t_0}^\infty \tau^{-\rho} d\tau\right) \\ =& E_0\eps\left(\jb{\sigma}^{-\nu}+\jb{\sigma}^{-\kappa}\frac{t_0^{-\rho+1}}{\rho-1}\right) \le E_0\left(1+\frac{c_0^{\rho-1}}{\rho-1}\right)\eps\jb{\sigma}^{-\nu} \end{align*} for $t_0\le t<T_0$. With this {\it a priori} bound and the local existence theorem for ODEs, we can easily show the global existence of the unique solution $z$ to \eqref{ODE}--\eqref{ODEinit}, and we obtain \eqref{ODEBound}. Now we turn our attention to the asymptotic behavior of $z$. In what follows, $C$ stands for various positive constants that can be determined only by $K_0$, $E_0$, $c_0$, $\rho$, and $\mu$. The actual value of $C$ may change line by line. Let $\xi$ and $\eta$ be the solution to the system of ODEs \begin{align} & \xi'(t)= -i \frac{\imagpart K}{2t\eta(t)} |\xi(t)|^2 \xi(t) +{J}(t)\sqrt{\eta(t)}, && t>t_0, \label{EqXi} \\ & \eta'(t)=\frac{\realpart K}{t}|\xi(t)|^2, && t>t_0, \label{EqEta}\\ & \xi(t_0)=z_0,\quad \eta(t_0)=1. \label{SysData} \end{align} We can easily check that $z(t)={\xi(t)}/{\sqrt{\eta(t)}}$ as long as $(\xi, \eta)$ exists. By the local existence theorem for ODEs, there is a local solution $(\xi, \eta)$ on $[t_0, T_1)$ with some $T_1(>t_0)$. For $t_0<\tau_0<T_1$, we put $$ M_{\tau_0}:=\sup_{t_0\le t\le \tau_0} |\xi(t)|. $$ Then we get $0\le \eta'(t)\le K_0 M_{\tau_0}^2t^{-1}$ for $t_0\le t\le {\tau_0}$, which leads to \begin{equation} \label{BoundEta} 1\le \eta(t) \le 1+K_0 M_{\tau_0}^2 \log(t/t_0), \quad t_0\le t\le \tau_0. \end{equation} By \eqref{EqXi} we obtain $$ \frac{d}{dt}|\xi(t)|^2= 2\realpart\left(\overline{\xi(t)} \, J(t)\sqrt{\eta(t)}\right) \le 2|\xi(t)||{J}(t)|\sqrt{\eta(t)} , $$ which, in combination with \eqref{BoundEta}, yields \begin{align} |\xi(t)| \le & |z_0|+\int_{t_0}^t |J(\tau)| \sqrt{\eta(\tau)} d\tau \nonumber\\ \le & C\eps\left(\jb{\sigma}^{-\nu} +\jb{\sigma}^{-\kappa}\int_{t_0}^t \tau^{-\rho} \left(1+M_{\tau_0} \sqrt{\log \frac{\tau}{t_0}}\right)d\tau\right) \label{B401} \end{align} for $t_0\le t\le \tau_0$. Since we have $\log t\le C_\mu t^{2\mu}$ for $t\ge 1$ with a positive constant $C_\mu$, and since $-\rho+\mu+1<0$, we get $$ \int_{t_0}^t \tau^{-\rho}\sqrt{\log \frac{\tau}{t_0}}d\tau\le Ct_0^{-\mu}\int_{t_0}^\infty \tau^{-\rho+\mu}d\tau\le Ct_0^{-\rho+1} \le C\jb{\sigma}^{-\rho+1},\quad t\ge t_0. $$ Hence we obtain from \eqref{B401} that \begin{equation} \label{B402} M_{\tau_0}\le C\eps\jb{\sigma}^{-\nu}+C\eps \jb{\sigma}^{-\nu} M_{\tau_0}\le C\eps\jb{\sigma}^{-\nu}+\frac{1}{2}M_{\tau_0} \end{equation} for $0<\eps\le \eps_2:=1/(2C)$ (with the constant $C$ from \eqref{B402}), which yields \begin{equation} \label{BoundXi} \left(\sup_{t_0\le t\le \tau_0}|\xi(t)|=\right)M_{\tau_0} \le C\eps\jb{\sigma}^{-\nu}, \end{equation} provided that $0<\eps\le \eps_2$. With the {\it a priori} bound \eqref{BoundXi} as well as \eqref{BoundEta}, we see that the solution $(\xi, \eta)$ exists globally in time, and we also have \begin{equation} \label{BoundXiEta} |\xi(t)|\le C\eps\jb{\sigma}^{-\nu},\ 1\le\eta(t)\le 1+C \eps^2\log \frac{t}{t_0} \le 1+C\eps^2 \left(\frac{t}{t_0}\right)^{2\mu} \end{equation} for $t\ge t_0$, provided that $0<\eps\le \eps_2$. We assume $0<\eps\le \eps_2$ from now on. We put $$ \Theta(t)=\frac{\imagpart K}{2}\int_{t_0}^t \frac{|\xi(\tau)|^2}{\tau\eta(\tau)}d\tau, \quad t\ge t_0. $$ Then \eqref{EqXi} implies $$ \xi(t)=e^{-i\Theta(t)}\left(z_0+ \int_{t_0}^t e^{i\Theta(\tau)}J(\tau)\sqrt{\eta(\tau)}d\tau \right), \quad t\ge t_0. $$ We define $$ z_+:= z_0+\int_{t_0}^\infty e^{i\Theta(\tau)} J(\tau)\sqrt{\eta(\tau)}d\tau, \quad \xi_+(t):=e^{-i\Theta(t)}z_+. $$ Then we obtain \begin{align} |\xi_+(t)-\xi(t)|\le & \int_{t}^\infty| J(\tau)| \sqrt{\eta(\tau)}d\tau\nonumber\\ \le & C\eps\jb{\sigma}^{-\kappa}\int_t^\infty \tau^{-\rho} \left(1+\eps_2 \left(\frac{\tau}{t_0}\right)^\mu\right)d\tau\nonumber\\ \le & C \eps (\jb{\sigma}^{-\kappa}t^{-\rho+1}+\jb{\sigma}^{-\kappa-\mu} t^{-\rho+\mu+1})\nonumber\\ \le & C\eps \jb{\sigma}^{-\kappa-\mu}t^{-\rho+\mu+1}, \quad t\ge t_0. \label{B454} \end{align} Especially we have $$ |z_+-z_0|\le C\eps \jb{\sigma}^{-\kappa-\mu}t_0^{-\rho+\mu+1} \le C\eps \jb{\sigma}^{-\nu}, $$ which shows \begin{equation} \label{B456} |z_+|\le C\eps\jb{\sigma}^{-\nu}. \end{equation} Observing that $|z_+|=|\xi_+(t)|$, we obtain from \eqref{B454} and \eqref{B456} that \begin{align} \label{B457} \left||\xi(t)|^2-|z_+|^2\right|\le & (|\xi(t)|+|z_+|)|\xi(t)-\xi_+(t)| \le C\eps^2\jb{\sigma}^{-\nu-\kappa-\mu}t^{-\rho+\mu+1} \end{align} for $t\ge t_0$. We define $$ \eta_\infty(t):=1+(\realpart K)\left(|z_+|^2 \log \frac{t}{t_0} +\int_{t_0}^\infty \frac{|\xi(\tau)|^2-|z_+|^2}{\tau}d\tau\right), \quad t\ge 1. $$ Since \eqref{EqEta} implies that $$ \eta(t)=1+(\realpart K)\left(|z_+|^2 \log \frac{t}{t_0} +\int_{t_0}^t \frac{|\xi(\tau)|^2-|z_+|^2}{\tau}d\tau\right), \quad t\ge t_0 $$ we obtain from \eqref{B457} that \begin{align} \label{B458} |\eta(t)-\eta_\infty(t)|\le K_0\int_t^\infty\frac{\bigl||\xi(\tau)|^2-|z_+|^2\bigr|}{\tau}d\tau\le C\eps^2 \jb{\sigma}^{-\nu-\kappa-\mu}t^{-\rho+\mu+1} \end{align} for $t\ge t_0$. Especially we have $\eta_\infty(t_0)\ge 1-C\eps^2$, which leads to \begin{equation} \label{B459} \eta_\infty(t)\ge 1-C\eps^2-K_0|z_+|^2\log t_0\ge 1-C\eps^2-C \eps^2 \jb{\sigma}^{-2\nu}\log t_0, \quad t\ge 1 \end{equation} with the help of \eqref{B456}. By \eqref{CondT0}, there is a positive constant $c_1=c_1(c_0, \rho)$ such that we have $$ \jb{\sigma}^{-2\nu}\log t_0\le \jb{\sigma}^{-2(\rho-1)} \log (c_0\jb{\sigma})\le c_1, \quad \sigma\in \R. $$ If we put $$ \eps_1:=\min\left\{\sqrt{\frac{1}{4C}}, \sqrt{\frac{1}{4c_1C}}, \eps_2\right\} $$ with the constant $C$ coming from \eqref{B459}, then we get \begin{equation} \label{LBXi1} \eta_\infty(t)\ge \frac{1}{2},\quad t\ge 1 \end{equation} for $0<\eps\le \eps_1$. From now on, we assume that $0<\eps \le \eps_1$. We set \begin{align*} \Theta_\infty(t):=& \frac{\imagpart K}{2} \int_{t_0}^t \frac{|z_+|^2}{\tau \eta_\infty(\tau)}d\tau, \quad t\ge 1\\ \Theta_0:=& \frac{\imagpart K}{2} \int_{t_0}^\infty \left(\frac{|\xi(\tau)|^2}{\tau\eta(\tau)}-\frac{|z_+|^2}{\tau\eta_\infty(\tau)}\right)d\tau. \end{align*} From \eqref{BoundXiEta}, \eqref{B456}, \eqref{B457}, \eqref{B458}, and \eqref{LBXi1}, we obtain \begin{align*} \left|\frac{|\xi(\tau)|^2}{\eta(\tau)}-\frac{|z_+|^2}{\eta_\infty(\tau)}\right| \le & \left|\frac{|\xi(\tau)|^2-|z_+|^2}{\eta(\tau)}\right|+|z_+|^2\left| \frac{1}{\eta(\tau)}-\frac{1}{\eta_\infty(\tau)}\right| \\ \le & C\eps^2\jb{\sigma}^{-\nu-\kappa-\mu}\tau^{-\rho+\mu+1} \end{align*} for $\tau\ge t_0$, which yields \begin{align} |(\Theta_\infty(t)+\Theta_0)-\Theta(t)|\le & \frac{K_0}{2} \int_{t}^\infty \left|\frac{|\xi(\tau)|^2}{\tau\eta(\tau)}-\frac{|z_+|^2}{\tau\eta_\infty(\tau)}\right|d\tau \nonumber\\ \le & C\eps^2 \jb{\sigma}^{-\nu-\kappa-\mu} t^{-\rho+\mu+1}, \quad t\ge t_0. \label{Ma} \end{align} We define $$ \xi_\infty(t):=e^{-i\Theta_\infty(t)}(e^{-i\Theta_0}z_+),\quad t\ge 1. $$ Then \eqref{Ma} leads to \begin{align} |\xi_\infty(t)-\xi_+(t)|\le & |e^{-i(\Theta_\infty(t)+\Theta_0)}-e^{-i\Theta(t)}|\,|z_+| \nonumber\\ \le & |(\Theta_\infty(t)+\Theta_0)-\Theta(t)|\,|z_+| \le C\eps^3 \jb{\sigma}^{-2\nu-\kappa-\mu} t^{-\rho+\mu+1}, \end{align} which, together with \eqref{B454}, yields \begin{equation} \label{B480} |\xi_\infty(t)-\xi(t)|\le C\eps \jb{\sigma}^{-\kappa-\mu}t^{-\rho+\mu+1},\quad t\ge t_0. \end{equation} Observing that $|\xi_\infty(t)|=|z_+|$ by definition, we find that \begin{align*} \xi_\infty'(t)=& -i\Theta_\infty'(t)\xi_\infty(t)= -i\frac{\imagpart K}{2t\eta_\infty(t)} |\xi_\infty(t)|^2\xi_\infty(t), \\ \eta_\infty'(t)=& \frac{\realpart K}{t}|\xi_\infty(t)|^2 \end{align*} for $t>1$. Hence if we put $z_\infty(t)=\xi_\infty(t)/\sqrt{\eta_\infty(t)}$, then we get $$ z_\infty'(t)= -\frac{K}{2t}|z_\infty(t)|^2z_\infty(t),\quad t>1. $$ It follows from \eqref{BoundXiEta}, \eqref{B456}, \eqref{B458}, \eqref{LBXi1}, and \eqref{B480} that \begin{align} |z(t)-z_\infty(t)| \le & \frac{|\xi(t)-\xi_\infty(t)|}{\sqrt{\eta(t)}}+|z_+| \left|\frac{1}{\sqrt{\eta(t)}}-\frac{1}{\sqrt{\eta_\infty(t)}}\right| \nonumber\\ \le & C\eps \jb{\sigma}^{-\kappa-\mu} t^{-\rho+\mu+1}, \quad t\ge t_0. \label{B481} \end{align} Finally we put $p_0:=z_\infty(1)$, and let $p$ be the solution to \eqref{ReducedODE}. It is clear that $z_\infty(t)=p(\log t)$ for $t\ge 1$. By \eqref{B456} and \eqref{LBXi1}, we get \begin{equation} \label{B482} |p_0|=\left|\frac{\xi_\infty(1)}{\sqrt{\eta_\infty(1)}}\right|\le C\eps\jb{\sigma}^{-\nu}. \end{equation} We obtain \eqref{ODEaSymp} and \eqref{ODEprof} from \eqref{B481} and \eqref{B482}. This completes the proof. \end{proof} \section{Proof of Theorem~\ref{Global}} \label{Proof0101} In this section we prove Theorem~\ref{Global}. Let $u$ be a smooth solution to \eqref{WaveEq}--\eqref{InitData} on $[0,T)\times \R^2$ for some $T>0$. For a positive integer $k$, and positive constants $\lambda$ and $\mu$, we define \begin{align*} e_{k,\lambda,\mu}[u](T):=& \sup_{(t,x)\in [0,T)\times \R^2}\bigl\{ (1+t)^{1/2} \jb{t-r}^{1-\mu}|\pa u(t,x)|\\ & \qquad\qquad\qquad {}+(1+t)^{(1-\lambda)/2}\jb{t-r}^{1-\mu}|\pa u(t,x)|_k\bigr\}, \end{align*} where $r=|x|$. Our first aim here is to prove the following, from which the global existence part of Theorem~\ref{Global} follows: \begin{prp}\label{AP01} Suppose that the assumptions in Theorem~$\ref{Global}$ are fulfilled. Let $k\ge 2$ and $0<\mu<1/10$. If $0< 2k\lambda \le \mu/6$, then we can find a positive constant $m_0=m_0(k, \lambda, \mu)$ having the following property: For any $m\ge m_0$ there is a positive constant $\eps_0=\eps_0(m, k, \lambda, \mu)$ such that \begin{equation} \label{Boot01} e_{k,\lambda,\mu}[u](T)\le m\eps \end{equation} implies \begin{equation} \label{Boot-F} e_{k,\lambda,\mu}[u](T)\le \frac{m}{2}\eps, \end{equation} provided that $0<\eps\le \eps_0$. \end{prp} {\it Proof.} Assume that \eqref{Boot01} is satisfied. In the following we always suppose that $0\le t<T$. We also suppose that $m\ge 1$, and that $\eps$ is small enough to satisfy $m\eps \le 1$. The letter $C$ in this proof stands for a positive constant which may depend on $k$, $\lambda$, and $\mu$, but are independent of $m$, $\eps$, and $T$. The proof is divided into several steps. \medskip {\bf Step 1: The energy estimates.} For $l\le 2k$, it follows from \eqref{Boot01} that \begin{align} |F(\pa u)|_{l}\le & C\left(|\pa u|^2|\pa u|_l+|\pa u|_k^2|\pa u|_{l-1}\right) \nonumber\\ \le & \frac{C^*}{2}m^2\eps^2(1+t)^{-1}|\pa u|_l+Cm^2\eps^2(1+t)^{\lambda-1}|\pa u|_{l-1} \label{Boot02} \end{align} with a positive constant $C^*=C^*(k,\lambda,\mu)$, where terms including $|\pa u|_{l-1}$ should be neglected if $l=0$. From the energy inequality and \eqref{Boot02} with $l=0$, we get $$ \|\pa u(t)\|_{0}\le C\eps+C^* m^2\eps^2\int_0^t (1+\tau)^{-1}\|\pa u(\tau)\|_{0} d\tau. $$ Gronwall's lemma implies \begin{equation} \|\pa u(t)\|_{0}\le C\eps (1+t)^{C^*m^2\eps^2}. \label{Ind00} \end{equation} We are going to prove that there are positive constants $B_l=B_l(k, \lambda, \mu)$ for $0\le l\le 2k$ such that \begin{equation} \label{Ind} \|\pa u(t)\|_l \le B_l \eps (1+t)^{C^*m^2\eps^2+\lambda l} \end{equation} for $0\le l\le 2k$. Indeed \eqref{Ind} for $l=0$ follows from \eqref{Ind00}. Suppose that \eqref{Ind} is true for some $0\le l\le 2k-1$. Then, by \eqref{Boot02} we get \begin{align*} \|\pa u(t)\|_{l+1}\le & C\eps+C^*m^2\eps^2\int_0^t (1+\tau)^{-1}\|\pa u(\tau)\|_{l+1} d\tau\\ & {}+CB_l m^2\eps^3\int_0^t (1+\tau)^{C^*m^2\eps^2+\lambda(l+1)-1}d\tau. \end{align*} Then Gronwall's lemma yields \begin{align*} \|\pa u(t)\|_{l+1}\le C\eps (1+t)^{C^*m^2\eps^2}+C\frac{B_l}{\lambda(l+1)}m^2\eps^3(1+t)^{C^*m^2\eps^2+\lambda(l+1)}, \end{align*} which inductively implies the desired result because $m\eps\le 1$. \medskip {\bf Step 2: Decay estimates of generalized derivatives of higher order.} We suppose that $\eps$ is so small to satisfy $C^*m^2\eps^2 \le \mu/6$, where the constant $C^*$ is from \eqref{Ind}. Then \eqref{Ind} for $l=2k$ implies \begin{equation} \|\pa u(t)\|_{2k}\le C\eps (1+t)^{\mu/3}, \label{Hoe00} \end{equation} because we have assumed $2k\lambda\le \mu/6$. \eqref{Hoe00} yields \begin{align*} \int_{\R^2} |F(\pa u)(\tau,y)|_{2k}dy \le & Cm\eps (1+\tau)^{(\lambda-1)/2}\|\pa u(\tau)\|_{2k}^2 \\ \le & Cm\eps^3(1+\tau)^{\mu-(1/2)} \le C\eps (1+\tau)^{\mu-(1/2)}, \end{align*} because $\lambda/2<\mu/3$. It is well known that we have $|u_0(t,x)|_{2k-1}\le C\eps(1+t)^{-1/2}$ for the solution $u_0$ to $\dal u_0=0$ with initial data $(u_0,\pa_t u_0) =(\eps f, \eps g)$ at $t=0$ (see \cite{Hoe97} for instance). Hence, by Lemma~\ref{Ho} we get \begin{align} (1+t)^{1/2}|u(t,x)|_{2k-1}\le & C\eps+C\int_0^t \int_{\R^2} \frac{|F(\pa u)(\tau, y)|_{2k}}{(1+\tau)^{1/2}}dy d\tau\nonumber\\ \le & C\eps(1+t)^{\mu} \label{Hoe01} \end{align} for $(t,x)\in [0,T)\times \R^2$. From Lemma~\ref{Rewrite03} we obtain \begin{equation} \label{Kla01} |\pa u(t,x)|_{2k-2}\le C\jb{t-r}^{-1}|u(t,x)|_{2k-1}\le C\eps (1+t)^{\mu-(1/2)}\jb{t-r}^{-1} \end{equation} for $(t, x)\in [0,T)\times \R^2$. Suppose that $R$ is a positive number to satisfy \eqref{Supp01}, so that we have \eqref{Supp02}. Recall the definition \eqref{DefLamTR} of $\Lambda_{T,R}$. We put $\Lambda_{T,R}^{\rm c}:=\bigl([0,T)\times \R^2\bigr)\setminus \Lambda_{T,R}$. If we have either $t/2<1$ or $r< t/2$, then we get $$ \jb{t-r}^{-1}\le C\jb{t+r}^{-1}. $$ If $r>t+R$, then we have $u(t,x)=0$ by \eqref{Supp02}. Hence \eqref{Kla01} implies \begin{equation} \label{BF01} \sup_{(t,x)\in \Lambda_{T,R}^c}(1+t)^{1/2}\jb{t-r}^{1-\mu}|\pa u(t,x)|_{2k-2}\le C\eps. \end{equation} \medskip {\bf Step 3: Decay estimates for generalized derivatives of lower order.} We suppose that $(t,x)=(t,r\omega)\in \Lambda_{T,R}$ throughout this step. Recall that $t^{-1}$, $r^{-1}$, $(1+t)^{-1}$, and $\jb{t+r}^{-1}$ are equivalent to each other. We define $U$, $U^{(\alpha)}$, $H$, and $H_\alpha$ as in Section~\ref{Reduction}. By \eqref{Hoe01} and \eqref{Kla01}, we get \begin{equation} \label{E501} |u(t,x)|_{\sharp,2k-2}\le C\eps t^{\mu-(1/2)}\jb{t-r}^{-1}, \end{equation} where $|u(t,x)|_{\sharp, 2k-2}$ is defined by \eqref{Basis300}. Hence we obtain from \eqref{Basic301} that $$ \sum_{|\alpha|\le 2k-2}|U^{(\alpha)}(t,x)|\le C\eps t^{\mu}\jb{t-r}^{-1}. $$ If $t/2=r$ or $t/2=1$, then we have $t^{\mu}\le C\jb{t-r}^{\mu}$, and we obtain \begin{equation} \label{BF03} \sum_{|\alpha|\le 2k-2}|U^{(\alpha)}(t,x)|\le C\eps \jb{t-r}^{\mu-1} \text{ when either $t/2=r$ or $t/2=1$}. \end{equation} By Corollary~\ref{Rewrite02} and \eqref{Hoe01}, we get \begin{align} \label{f401} t^{1/2}|\pa u|_l\le & C \sum_{|\alpha|\le l} |r^{1/2}\pa \Gamma^\alpha u| \le C\sum_{|\alpha|\le l} |U^{(\alpha)}|+C \eps t^{\mu-1},\quad l\le 2k-2. \end{align} Recall that $0<\mu<1/10$. By \eqref{Hoe01}, \eqref{E501}, and Proposition~\ref{Pro301} we get \begin{equation} |H(t,x)|\le C\left(\eps^3 t^{3\mu-2}\jb{t-r}^{-2} {}+\eps t^{\mu-2}\right) \le C\eps t^{3\mu-2}\jb{t-r}^{-2\mu}. \label{E502} \end{equation} We define $t_{0, \sigma}=\max\{2, -2\sigma\}$ for $\sigma\le R$. Note that the line segment $\{(t, (t+\sigma)\omega);\, 0\le t< T\}$, with $\sigma\le R$ and $\omega\in \Sph^1$ being fixed, meets the boundary of $\Lambda_{T,R}$ at the point $\left(t_{0, \sigma}, (t_{0,\sigma}+\sigma)\omega\right)$. We have \begin{equation} \label{517} (1+R)^{-1}\jb{\sigma} \le t_{0,\sigma}\le 2(1+|\sigma|)\le 2\sqrt{2} \jb{\sigma},\quad \sigma\le R. \end{equation} We define $$ V_{\sigma,\omega}(t)=U\bigl(t, (t+\sigma)\omega\bigr) $$ for $0\le t\le T$, $\sigma\le R$, and $\omega\in \Sph^1$. Then \eqref{Red01} leads to \begin{equation} \label{Red01a} V_{\sigma, \omega}'(t)= -\frac{1}{2t}F(\hat{\omega})|V_{\sigma,\omega}(t)|^2V_{\sigma, \omega}(t)+H(t, (t+\sigma)\omega) \end{equation} for $t_{0,\sigma}\le t<T$. Note that by \eqref{BF03} we have \begin{equation} \label{BF04} |V_{\sigma, \omega}(t_{0,\sigma})| \le C\eps \jb{\sigma}^{\mu-1}. \end{equation} for $\sigma\le R$ and $\omega\in \Sph^1$. By \eqref{E502}, \eqref{517}, and \eqref{BF04}, we can apply Lemma~\ref{ODELemma02} to \eqref{Red01a} (with $\rho=2-3\mu$, and $\kappa=2\mu$): \eqref{ODEBound} implies that \begin{equation} \label{E409} |V_{\sigma, \omega}(t)|\le C\eps \jb{\sigma}^{\mu-1},\quad t\ge t_{0,\sigma}, \end{equation} where $C$ is a constant independent of $\sigma$ and $\omega$ (note that $0<\mu<1/10<\rho-1$ and $\kappa+\rho-1=1-\mu$). Now we get $|U(t,r\omega)|=|V_{r-t,\omega}(t)|\le C\eps\jb{t-r}^{\mu-1}$, and with the help of \eqref{f401} for $l=0$ we obtain \begin{equation} \label{f410} \sup_{(t,x)\in \Lambda_{T,R}} (1+t)^{1/2}\jb{t-r}^{1-\mu} |u(t,x)| \le C\eps. \end{equation} Let $|\alpha|\le k$. For a nonnegative integer $s$, we set $$ {\mathcal U}^{(s)}(t,x):=\sum_{|\alpha|\le s}|U^{(\alpha)}(t,x)|. $$ By \eqref{f401} we get \begin{equation} |\pa u(t,x)|_{|\alpha|-1}\le C\left(t^{-1/2}{\mathcal U}^{(|\alpha|-1)}(t,x)+\eps t^{\mu-(3/2)}\right). \label{B410} \end{equation} We obtain from \eqref{E501}, \eqref{B410}, and Proposition~\ref{Pro301} that \begin{align} |H_\alpha(t,x)|\le & C\left( \eps^3 t^{3\mu-2}\jb{t-r}^{-2} {}+\eps t^{\mu-2}+\eps^3t^{3\mu-4}+t^{-1} \bigl({\mathcal U}^{(|\alpha|-1)}(t,x)\bigr)^3\right), \nonumber\\ \le & C\eps t^{3\mu-2}\jb{t-r}^{-2\mu}+Ct^{-1}\bigl({\mathcal U}^{(|\alpha|-1)}(t,x)\bigr)^3. \label{f421} \end{align} We put $$ V^{(\alpha)}_{\sigma, \omega}(t)=U^{(\alpha)}\bigl(t, (t+\sigma)\omega\bigr) $$ for $0\le t< T$, $\sigma\le R$, and $\omega\in \Sph^1$. From \eqref{Red02} we get \begin{align*} \bigl(V^{(\alpha)}_{\sigma,\omega}\bigr)'(t)=& -\frac{F(\hat{\omega})}{2t} \left(2|V_{\sigma,\omega}(t)|^2 V^{(\alpha)}_{\sigma,\omega}(t) +\bigl(V_{\sigma,\omega}(t)\bigr)^2\overline{V^{(\alpha)}_{\sigma,\omega}(t)}\right)\\ &+H_\alpha(t, (t+\sigma) \omega) \end{align*} for $t_{0,\sigma}\le t<T$. Hence by \eqref{BF04} and \eqref{f421} we obtain \begin{align*} \frac{d}{dt}|V^{(\alpha)}_{\sigma,\omega}(t)|^2=& -\realpart\left(\frac{F(\hat{\omega})}{t} \left(2|V_{\sigma,\omega}(t)|^2 |V^{(\alpha)}_{\sigma,\omega}(t)|^2 +\bigl(V_{\sigma,\omega}(t)\bigr)^2\bigl(\overline{V^{(\alpha)}_{\sigma,\omega}(t)}\bigr)^2\right) \right)\\ &+2\realpart\left(H_\alpha(t, (t+\sigma)\omega) \overline{V^{(\alpha)}_{\sigma,\omega}}(t)\right)\\ \le & \frac{2C_*\eps^2}{t} |V^{(\alpha)}_{\sigma,\omega}(t)|^2 {}+C\left(\eps t^{3\mu-2}\jb{\sigma}^{-2\mu}+t^{-1}\bigl({\mathcal V}_{\sigma, \omega}^{(|\alpha|-1)}(t)\bigr)^3 \right)\,|V^{(\alpha)}_{\sigma,\omega}(t)| \end{align*} for $t_{0,\sigma}\le t<T$, where ${\mathcal V}_{\sigma,\omega}^{(s)}(t):=\sum_{|\alpha|\le s}|V_{\sigma, \omega}^{(\alpha)}(t)|$, and $C_*=C_*(k,\lambda, \mu)$ is a positive constant. Therefore it follows from \eqref{BF03} that \begin{align*} t^{-C_*\eps^2}|V^{(\alpha)}_{\sigma,\omega}(t)| \le & t_{0,\sigma}^{-C_*\eps^2} |V^{(\alpha)}_{\sigma,\omega}(t_{0,\sigma})| {}+C\eps \jb{\sigma}^{-2\mu} \int_{t_{0,\sigma}}^t \tau^{-C_*\eps^2+3\mu-2}d\tau \\ &{}+C\int_{t_{0,\sigma}}^t \tau^{-C_*\eps^2-1} \bigl({\mathcal V}_{\sigma, \omega}^{(|\alpha|-1)}(\tau)\bigr)^3 d\tau\\ \le & C\eps\jb{\sigma}^{\mu-1-C_*\eps^2}+C\int_{t_{0,\sigma}}^t \tau^{-C_*\eps^2-1} \bigl({\mathcal V}_{\sigma, \omega}^{(|\alpha|-1)}(\tau)\bigr)^3 d\tau \end{align*} for $t_{0,\sigma}\le t <T$, which leads to $$ t^{-C_*\eps^2}{\mathcal V}^{(l)}_{\sigma,\omega}(t) \le C\eps\jb{\sigma}^{\mu-1}+C\int_{t_{0,\sigma}}^t \tau^{-C_*\eps^2-1} \bigl({\mathcal V}_{\sigma, \omega}^{(l-1)}(\tau)\bigr)^3 d\tau $$ for $t_{0,\sigma}\le t<T$ and $1\le l \le k$. Using this inequality, we are going to prove that \begin{align} \label{BF} {\mathcal V}_{\sigma, \omega}^{(l)}(t) \le \widetilde{B}_l \eps t^{3^{l-1} C_* \eps^2} \jb{\sigma}^{\mu-1} \end{align} for $t_{0,\sigma}\le t < T$ and $1\le l\le k$ with some positive constant $\widetilde{B}_l=\widetilde{B}_l(k,\lambda, \mu)$. By \eqref{E409} we have ${\mathcal V}_{\sigma, \omega}^{(0)}(t)\le \widetilde{B}_0\eps\jb{\sigma}^{\mu-1}$ with a positive constant $\widetilde{B}_0=\widetilde{B}_0(k,\lambda,\mu)$. Hence we get \begin{align*} t^{-C_*\eps^2}{\mathcal V}^{(1)}_{\sigma,\omega}(t) \le & C\eps\jb{\sigma}^{\mu-1}+C\widetilde{B}_0^3\eps^3\jb{\sigma}^{3\mu-3}\int_{2}^\infty \tau^{-C_*\eps^2-1} d\tau\le \widetilde{B}_1\eps\jb{\sigma}^{\mu-1} \end{align*} with $\widetilde{B}_1=C+CC_*^{-1}\widetilde{B}_0^3$, which leads to $$ {\mathcal V}_{\sigma, \omega}^{(1)}(t)\le \widetilde{B}_1 \eps \jb{\sigma}^{\mu-1} t^{C_*\eps^2}, $$ and \eqref{BF} for $l=1$ is shown. Next, suppose that \eqref{BF} is true for some $l$ with $1\le l \le k-1$. Then we get \begin{align*} t^{-C_*\eps^2}{\mathcal V}^{(l+1)}_{\sigma,\omega}(t) \le & C\eps\jb{\sigma}^{\mu-1}+C\widetilde{B}_l^3 \eps^3\jb{\sigma}^{3\mu-3} \int_{2}^t \tau^{(3^{l}-1)C_*\eps^2-1} d\tau\\ \le & \widetilde{B}_{l+1}\eps \jb{\sigma}^{\mu-1} t^{(3^{l}-1)C_*\eps^2} \end{align*} with $\widetilde{B}_{l+1}=C+CC_*^{-1} (3^{l}-1)^{-1}\widetilde{B}_l^{3}$, and we obtain $$ {\mathcal V}_{\sigma, \omega}^{(l+1)}(t) \le \widetilde{B}_{l+1}\eps \jb{\sigma}^{\mu-1} t^{3^{l}C_*\eps^2}, $$ which is \eqref{BF} with $l$ replaced by $l+1$. Now \eqref{BF} for $1\le l \le k$ is established. By \eqref{BF}, we obtain $$ \sum_{|\alpha|\le k}|U^{(\alpha)}(t,x)|\le C\eps (1+t)^{3^{k-1}C_*\eps^2}\jb{t-r}^{\mu-1},\quad (t,x)\in \Lambda_{T,R}, $$ which yields \begin{equation} \label{BF404} |\pa u(t,x)|_k\le C\eps (1+t)^{3^{k-1}C_*\eps^2-(1/2)}\jb{t-r}^{\mu-1},\quad (t,x)\in \Lambda_{T,R} \end{equation} with the help of \eqref{f401}. If we choose sufficiently small $\eps$ to satisfy $$ 3^{k-1}C_*\eps^2\le \lambda/2, $$ it follows from \eqref{BF404} that \begin{equation} \label{BFF} \sup_{(t,x)\in \Lambda_{T, R}} (1+t)^{(1-\lambda)/2}\jb{t-r}^{1-\mu}|\pa u(t,x)|_k\le C\eps. \end{equation} \medskip {\bf The final step.} By \eqref{BF01}, \eqref{f410}, and \eqref{BFF}, we see that there exist two positive constants $\eps_0=\eps_0(m, k, \lambda, \mu)$ and $C_0=C_0(k, \lambda, \mu)$ such that \begin{equation} \label{QQ} e_{k,\lambda, \mu}[u](T)\le C_0\eps \end{equation} for $0<\eps\le \eps_0$ (note that we have $2k-2\ge k$ for $k\ge 2$). If $m\ge \max\{2C_0, 1\}$, then \eqref{QQ} implies \eqref{Boot-F} immediately. \qed \medskip We are in a position to prove Theorem~\ref{Global}. \begin{proof}[Proof of Theorem~$\ref{Global}$] Let the assumptions in Theorem~\ref{Global} be fulfilled. Suppose that $u$ is a local solution to \eqref{WaveEq}--\eqref{InitData} on $[0,T)\times \R^2$ for some $T>0$. We fix $\lambda$ and $k$ as in Proposition~\ref{AP01}. We also fix $m$ satisfying $m\ge m_0$ and $$ \sup_{x\in \R^2}\jb{r}^{1-\mu}(|\pa u(t,x)|+|\pa u(t,x)|_k)\bigr|_{t=0} \le \frac{m}{2}{\eps}, \quad \eps>0, $$ where $m_0=m_0(k, \lambda, \mu)$ is from Proposition~\ref{AP01}. Let $\eps_0=\eps_0(m, k, \lambda, \mu)$ also be from Proposition~\ref{AP01}. We put $$ T_*=\sup\left\{t\in [0,T)\,; e_{k, \lambda, \mu}(t)\le m\eps \right\}. $$ By the choice of $m$, we have $T_*>0$. Moreover we get $T_*=T$ for $0<\eps\le \eps_0$, because if $T_*<T$ then Proposition~\ref{AP01} implies $e_{k,\lambda,\mu}(T_*)<m\eps/2$ for $0<\eps\le \eps_0$, and we obtain from the continuity of $e_{k,\lambda,\mu}$ that $e_{k,\lambda,\mu}(t_*)\le m\eps$ for some $t_*>T_*$, which contradicts the definition of $T_*$. Therefore we see that $e_{k,\lambda, \mu}(t)$ cannot exceed $m\eps$ as long as the solution $u$ exists. This {\it a priori} estimate and the local existence theorem implies the global existence of the solution $u$. We also see that \eqref{Boot01} holds for some large $m$. Now we turn our attention to the asymptotic behavior. Because we have \eqref{Boot01}, the estimates in the proof of Proposition~\ref{AP01} are valid. We go back to \eqref{Red01a}, and apply Lemma~\ref{ODELemma02} with $z(t)=V_{\sigma,\omega}(t)$, $K=F(\hat\omega)$, $J(t)=H(t,(t+\sigma)\omega)$, $t_0=t_{0,\sigma}$ for each fixed $(\sigma, \omega)\in (-\infty, R]\times \Sph^1$ (note that we can take $\rho=2-3\mu$ and $\kappa=2\mu$ because of \eqref{E502}, \eqref{517}, and \eqref{BF04}). Then we see that there exists $P_0=P_0(\sigma, \omega)$ such that \begin{align} \label{B601} & |V_{\sigma, \omega}(t)-P(\log t, \sigma, \omega)|\le C\eps t^{4\mu-1} \jb{\sigma}^{-3\mu}, \quad t\ge t_{0,\sigma},\\ \label{B602} & |P_0(\sigma, \omega)|\le C\eps \jb{\sigma}^{\mu-1}, \end{align} where $P=P(\tau, \sigma, \omega)$ is the solution to \eqref{ode}, and $C$ is a constant independent of $\sigma$ and $\omega$. Recalling that $D_-\bigl(r^{1/2} u(t,x)\bigr)=U(t,x)=V_{r-t,\omega}(t)$ with $r=|x|$ and $\omega=|x|^{-1}x$, we obtain from \eqref{B601} that \begin{equation} \label{B603} |D_-\bigl(r^{1/2}u(t,x)\bigr) {}-P(\log t, r-t, \omega)|\le C\eps t^{4\mu-1}\jb{t-r}^{-3\mu} \end{equation} for $(t,x)\in \Lambda_{T,R}$. By Corollary~\ref{Rewrite02} and \eqref{Hoe01} we get \begin{equation} \label{B604} \bigl|r^{1/2} \pa u(t,x)-\hat{\omega}(x) D_-\bigl(r^{1/2}u(t,x)\bigr)\bigr|\le C\eps (1+t)^{\mu-1}, \end{equation} where $\hat{\omega}(x)=(-1, x/|x|)$. Since \eqref{B602} and \eqref{ExpS1} yield $|P(\tau,\sigma, \omega)|\le C\eps \jb{\sigma}^{\mu-1}$, we have \begin{align} |(r^{-1/2}-t^{-1/2})P(\log t, r-t, \omega)|\le & C\eps \frac{|t-r|}{\sqrt{t}\sqrt{r}(\sqrt{t}+\sqrt{r})}\jb{t-r}^{\mu-1} \nonumber\\ \le C\eps t^{-3/2}\jb{t-r}^{\mu} \label{B605} \end{align} for $(t,x)\in \Lambda_{T,R}$. To sum up the estimates \eqref{B603}, \eqref{B604} and \eqref{B605}, we arrive at \begin{equation} \label{B606} |\pa u(t,x) {}-\hat{\omega}(x) t^{-1/2}P(\log t, r-t, \omega)| \le C\eps t^{4\mu-(3/2)}\jb{t-r}^{-3\mu} \end{equation} for $(t,x)\in \Lambda_{T,R}$. In order to extend \eqref{B606} outside of $\Lambda_{T,R}$, we just have to extend the definition of $P_0$ by setting $P_0(\sigma, \omega)=0$ for $\sigma>R$. Indeed, both $u(t,x)$ and $P(\log t, r-t, \omega)$ vanish for $|x|\ge t+R$. On the other hand, if $r<t/2$ or $1< t<2$, we have \begin{equation} |\pa u(t,x)|\le C\eps t^{-1/2}\jb{t-r}^{\mu-1}\le C\eps t^{\mu-(3/2)}, \end{equation} and \begin{align} t^{-1/2}|P(\log t, r-t, \omega)| \le & C\eps t^{-1/2}\jb{t-r}^{\mu-1} \le C\eps t^{\mu-(3/2)}. \end{align} Therefore \eqref{B606} is valid for all $(t,x)\in [1, \infty)\times \R^2$, which shows \eqref{Concl01}. This completes the proof. \end{proof} \section{Proof of Corollary~\ref{EnergyDecay}}\label{Omake} Let the assumptions of Corollary~\ref{EnergyDecay} be fulfilled. Suppose that we have \eqref{Supp01} for some $R>0$. Then we get \eqref{Supp02}. We always assume $t\ge 2$ throughout this proof. Since we have $R+|t-r|\le \sqrt{2}(1+R)\jb{t-r}$, \eqref{DecayDamping} leads to \begin{equation} \label{DecayDamping02} |\pa u(t,x)|\le C t^{-1/2} \min\{(\log t)^{-1/2}, \eps(R+|t-r|)^{\mu-1}\}, \end{equation} where $r=|x|$. We put $m_\eps(t):=\eps^{1/(1-\mu)}(\log t)^{1/(2-2\mu)}$. Here $m_\eps(t)$ is chosen so that we have $\eps (R+t-r)^{\mu-1}=\eps m_\eps(t)^{\mu-1}=(\log t)^{-1/2}$ for $r=t+R-m_\eps(t)$. For small $\eps>0$ we have $0<m_\eps(t)<t$, and we get $0<t+R-m_\eps (t)\le t+R$. Then it follows from \eqref{Supp02} that $$ \|u(t)\|_E^2=\frac{1}{2}\int_{|x|\le t+R} |\pa u(t,x)|^2 dx=I_1+I_2, $$ where $$ I_1=\frac{1}{2}\int_{|x|\le t+R-m_\eps(t)} |\pa u(t,x)|^2 dx,\quad I_2=\frac{1}{2}\int_{t+R-m_\eps(t)\le |x|\le t+R} |\pa u(t,x)|^2 dx. $$ Note that we have $t^{-1}r\le t^{-1}(t+R)\le 1+R/2$ for $0\le r\le t+R$. Hence, switching to the polar coordinate, we get from \eqref{DecayDamping02} that \begin{align*} I_1 \le & C\eps^2 \int_0^{t+R-m_\eps(t)} t^{-1} (R+|t-r|)^{2\mu-2} r dr\\ \le & C\eps^2 \int_0^{t+R-m_\eps(t)} (R+t-r)^{2\mu-2}dr\le C\eps^2m_\eps(t)^{2\mu-1} =C\eps^{\frac{1}{1-\mu}} (\log t)^{-\frac{1-2\mu}{2-2\mu}}. \end{align*} Similarly it follows from \eqref{DecayDamping02} that \begin{align*} I_2 \le & C \int_{t+R-m_\eps(t)}^{t+R} t^{-1}(\log t)^{-1} r dr \\ \le & C(\log t)^{-1} \int_{t+R-m_\eps(t)}^{t+R} dr = C\eps^{\frac{1}{1-\mu}} (\log t)^{-\frac{1-2\mu}{2-2\mu}}. \end{align*} This completes the proof. \qed \medskip \section*{Acknowledgments} The first author (S.~K.) is supported by Grant-in-Aid for Scientific Research (C) (No.~23540241), JSPS. The third author (H.~S.) is supported by Grant-in-Aid for Young Scientists~(B) (No.~22740089), MEXT.
\section{Introduction} Fibonacci numbers, Pell numbers and their generalizations have been studying for a long time. One of these generalizations was given by Miles in 1960. Miles [8] defined generalized order-$k$ Fibonacci numbers(GO$k$F) as \begin{equation} f_{k,n}=\sum\limits_{j=1}^{k}f_{k,n-j}\ \end{equation for $n>k\geq 2$, with boundary conditions: $f_{k,1}=f_{k,2}=f_{k,3}=\cdots =f_{k,k-2}=0$ and $f_{k,k-1}=f_{k,k}=1.$\newline Er [2] defined $k$ sequences of generalized order-$k$ Fibonacci numbers ($k SO$k$F) as; for $n>0,$ $1\leq i\leq k \begin{equation} f_{k,n}^{\text{ }i}=\sum\limits_{j=1}^{k}c_{j}f_{k,n-j}^{\text{ }i}\ \ \end{equation with boundary conditions for $1-k\leq n\leq 0,$ \begin{equation*} f_{k,n}^{\text{ }i}=\left\{ \begin{array}{l} 1\text{ \ \ \ \ \ if \ }i=1-n, \\ 0\text{ \ \ \ \ \ otherwise, \end{array \right. \end{equation* where $c_{j}$ $(1\leq j\leq k)$ are constant coefficients, $f_{k,n}^{\text{ i}$ is the $n$-th term of $i$-th sequence of order-$k$ generalization. For c_{j}=1$, $k$-th sequence of this generalization involves the Miles generalization(1) for $i=k,$ i.e \begin{equation*} f_{k,n}^{k}=f_{k,k+n-2}. \end{equation*} Kili\c{c} and Ta\c{c}c\i \lbrack 4] defined $k$ sequences of generalized order-$k$ Pell numbers ($k$SO$k$P) as; for $n>0,$ $1\leq i\leq k \begin{equation} p_{k,n}^{\text{ }i}=2p_{k,n-1}^{\text{ }i}+p_{k,n-2}^{\text{ }i}+\cdots +\ p_{k,n-k}^{\text{ }i}\ \end{equation with initial conditions for $1-k\leq n\leq 0,$ \begin{equation*} p_{k,n}^{\text{ }i}=\left\{ \begin{array}{l} 1\text{ \ \ \ \ \ if \ }i=1-n, \\ 0\text{ \ \ \ \ \ otherwise, \end{array \right. \end{equation* where $p_{k,n}^{\text{ }i}$ is the $n$-th term of $i$-th sequence of order k $ generalization. Kaygisiz and \c{S}ahin [3] defined $k$ sequences of the generalized order-$k$ Van der Laan numbers($k$SO$k$V) as; for $n>0,$ $1\leq i\leq k \begin{equation} v_{k,n}^{i}=\sum\limits_{j=2}^{k}v_{k,n-j}^{i}\ \end{equation with initial conditions for $1-k\leq n\leq 0,$ \begin{equation*} v_{k,n}^{i}=\left\{ \begin{array}{lc} 1\ \ \ \ \ \ \ \ \text{ \ if }i-n=k, & \\ 0\ \text{ \ \ \ \ \ \ \ \ otherwise} & \end{array \right. \end{equation* for $1-k\leq n\leq 0,$ where $v_{k,n}^{i}$ is the $n$-th term of $i$-th sequence. \bigskip MacHenry [5] defined generalized Fibonacci polynomials $(F_{k,n}(t))$ where t_{i}$ $(1\leq i\leq k)$ are constant coefficients of the core polynomia \begin{equation*} P(x;t_{1},t_{2},\ldots ,t_{k})=x^{k}-t_{1}x^{k-1}-\cdots -t_{k}, \end{equation* which is denoted by the vector \begin{equation*} t=(t_{1},t_{2},\ldots ,t_{k}). \end{equation* $F_{k,n}(t)$ is defined inductively by \begin{eqnarray} F_{k,n}(t) &=&0,\text{ }n<1 \\ F_{k,1}(t) &=&1 \notag \\ F_{k,2}(t) &=&t_{1} \notag \\ F_{k,n+1}(t) &=&t_{1}F_{k,n}(t)+\cdots +t_{k}F_{k,n-k+1}(t). \notag \end{eqnarray For example the generalized Fibonacci polynomials for $k=4$ and $k=5$ are \begin{equation*} 1,t_{1},t_{2}+t_{1}^{2},t_{3}+2t_{1}t_{2}+t_{1}^{3},t_{4}+2t_{1}t_{3}+t_{2}^{2}+t_{1}^{4}+3t_{1}^{2}t_{2},... \end{equation*} \bigskip\ an \begin{eqnarray*} &&1,\text{ }t_{1},\text{ }t_{2}+t_{1}^{2},\text{ t_{3}+2t_{1}t_{2}+t_{1}^{3},\text{ t_{4}+2t_{1}t_{3}+t_{2}^{2}+t_{1}^{4}+3t_{1}^{2}t_{2},\text{ } \\ &&t_{1}^{5}+4t_{1}^{3}t_{2}+3t_{1}^{2}t_{3}+3t_{1}t_{2}^{2}+2t_{1}t_{4}+2t_{2}t_{3}+t_{5}, \\ &&2t_{1}t_{5}+2t_{2}t_{4}+6t_{1}t_{2}t_{3}+t_{2}^{3}+t_{3}^{2}+t_{1}^{6} \allowbreak 3t_{1}^{2}t_{4}+4t_{1}^{3}t_{3}+5t_{1}^{4}t_{2}+6t_{1}^{2}t_{2}^{2},... \end{eqnarray* respectively. \bigskip MacHenry studied on these polinomials and obtain very useful properties of these polynomials in [6,7]. \begin{remark} $f_{k,n},$ $f_{k,n}^{\text{ }i},$ $p_{k,n}^{\text{ }i}$, $v_{k,n}^{i}$ and F_{k,n}(t)$ are GO$k$F (1), $k$SO$k$F (2), $k$SO$k$P (3), $k$SO$k$V (4) and generalized Fibonacci polynomials (5) respectively, then \newline $i)$ substituting $c_{j}=t_{i}$ in (2) and generalized Fibonacci polynomials, for $1\leq i, j\leq k$ we obtai \begin{equation*} F_{k,n}(t)=f_{k,n-1}^{\text{ }1}, \end{equation* $ii)$ substituting $t_{1}=2$ and $t_{i}=1$ for $2\leq i\leq k$ in generalized Fibonacci polynomials, we obtai \begin{equation*} F_{k,n}(t)=p_{k,n}^{\text{ }k}, \end{equation* $iii)$ substituting $t_{1}=0$ and $t_{i}=1$ for $2\leq i\leq k$ in generalized Fibonacci polynomials, we obtain \begin{equation*} F_{k,n}(t)=v_{k,n}^{k}, \end{equation* $iv)$ substituting $t_{i}=1$ in generalized Fibonacci polynomials, we obtain \begin{equation*} F_{k,n}(t)=f_{k,k+n-2}. \end{equation*} \end{remark} Minc [9] found Hessenberg matrix whose permanent are subscripted generalized order-$k$ Fibonacci numbers. Ocal [10] gave various Hessenberg matrices whose determinants and permanents are subscripted generalized order-$k$ Fibonacci numbers. In this paper we derive determinantal and permanental representation of generalized Fibonacci polynomials using various Hessenberg matrices. \section{The determinantal representations} \bigskip An $n\times n$ matrix $A_{n}=(a_{ij})$ is called lower Hessenberg matrix if a_{ij}=0$ when $j-i>1$ i.e. \begin{equation*} A_{n}=\left[ \begin{array}{ccccc} a_{11} & a_{12} & 0 & \cdots & 0 \\ a_{21} & a_{22} & a_{23} & \cdots & 0 \\ a_{31} & a_{32} & a_{33} & \cdots & 0 \\ \vdots & \vdots & \vdots & & \vdots \\ a_{n-1,1} & a_{n-1,2} & a_{n-1,3} & \cdots & a_{n-1,n} \\ a_{n,1} & a_{n,2} & a_{n,3} & \cdots & a_{n,n \end{array \right] \end{equation*} \begin{thrm} $\bigskip \lbrack 1]$ $A_{n}$ be the $n\times n$ lower Hessenberg matrix for all $n\geq 1$ and define $\det (A_{0})=1,$ then \begin{equation*} \det (A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation} \det (A_{n})=a_{n,n}\det (A_{n-1})+\sum\limits_{r=1}^{n-1}((-1)^{n-r}a_{n,r}\pro \limits_{j=r}^{n-1}a_{j,j+1}\det (A_{r-1})). \end{equation} \end{thrm} \bigskip \begin{thrm} Let $k\geq 2$ be an integer$,$ $F_{k,n}(t)$ be the generalized Fibonacci Polynomial (5) and $Q_{k,n}=(q_{rs})$ $n\times n$ Hessenberg matrix$,$ wher \begin{equation*} q_{rs}=\left\{ \begin{array}{c} i^{\left\vert r-s\right\vert }.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ }-1\leq r-s<k\text{ },\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation} Q_{k,n}=\left[ \begin{array}{cccccc} t_{1} & it_{2} & 0 & 0 & \cdots & 0 \\ i & t_{1} & it_{2} & 0 & \cdots & 0 \\ i^{2}\frac{t_{3}}{t_{2}^{2}} & i & t_{1} & it_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & i^{k-4}\frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & \cdots & 0 \\ & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] \label{kuka} \end{equation the \begin{equation*} \det (Q_{k,n})=F_{k,n+1}(t) \end{equation* where $t_{0}=1$ and $i=\sqrt{-1}.$ \end{thrm} \bigskip \begin{proof} \bigskip To prove $\det (Q_{k,m})=F_{k,m+1}(t)$, we use the mathematical induction on $m$. The result is true for m=1$ by hypothesis. Using Theorem (2.1) we have \begin{eqnarray*} \det (Q_{k,m+1}) &=&q_{m+1,m+1}\det (Q_{k,m})+\sum\limits_{r=1}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=1}^{m-k+1}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (Q_{k,r-1})\right) \\ &&+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r \frac{t_{m-r+2}}{t_{2}^{(m-r+1)}}\prod\limits_{j=r}^{m}it_{2}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m}) \\ &&+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r}\frac{t_{m-r+2}} t_{2}^{(m-r+1)}}.i^{m+1-r}.t_{2}^{(m-r+1)}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r}t_{m-r+2}.i^{m+1-r}.\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}t_{m-r+2}\det (Q_{k,r-1}) \\ &=&t_{1}\det (Q_{k,m})+t_{2}\det (Q_{k,m-1})+\cdots +t_{k}\det (Q_{k,m-(k-1)}) \end{eqnarray* From the hypothesis and the definition of generalized Fibonacci polynomials we obtai \begin{equation*} \det (Q_{k,m+1})=t_{1}F_{k,m+1}+t_{2}F_{k,m}+\cdots +t_{k}F_{k,m-(k-2)}=F_{k,m+2}. \end{equation* Therefore, the result is true for all positive integers. \end{proof} \bigskip \begin{exam} We obtain $5$-th generalized Fibonacci polynomial for $k=6$, using Theorem 2. \begin{equation*} F_{6,5}(t)=\det \left[ \begin{array}{cccc} t_{1} & it_{2} & 0 & 0 \\ i & t_{1} & it_{2} & 0 \\ \frac{-t_{3}}{t_{2}^{2}} & i & t_{1} & it_{2} \\ \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}}{t_{2}^{2}} & i & t_{1 \end{array \right] =t_{4}+2t_{1}t_{3}+t_{2}^{2}+t_{1}^{4}+3t_{1}^{2}t_{2}. \end{equation*} \end{exam} \bigskip \begin{cor} \bigskip $\left[ 10\right] $ Let $k\geq 2$ be an integer$,$ $f_{k,n}$ be the generalized order-$k$ Fibonacci numbers (1) and $C_{k,n}=(c_{rs})$ $n\times n $ Hessenberg matrix$,$ wher \begin{equation*} c_{rs}=\left\{ \begin{array}{l} i^{\left\vert r-s\right\vert }\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ }-1\leq r-s<k\text{ },\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* the \begin{equation*} \det (C_{k,n})=f_{k,k+n-1} \end{equation* where $i=\sqrt{-1}.$ \end{cor} \bigskip \begin{proof} \bigskip It is direct from Theorem 2.2 for $t_{i}=1.$ \end{proof} \begin{thrm} \bigskip Let $k\geq 2$ be an integer$,$ $F_{k,n}(t)$ be the generalized Fionacci Polynomial (5) and $B_{k,n}=(b_{ij})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} b_{ij}=\left\{ \begin{array}{l} -t_{2}\text{ \ \ \ \ \ \ \ \ \ \ \ if \ \ \ }j=i+1, \\ \frac{t_{i-j+1}}{t_{2}^{(i-j)}}\text{\ \ \ \ \ \ \ \ \ \ \ \ \ if\ \ \ \ \ 0\leq i-j<k\text{,} \\ 0\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise \end{array \right. \end{equation* i.e. \begin{equation} B_{k,n}=\left[ \begin{array}{cccccc} t_{1} & -t_{2} & 0 & 0 & \cdots & 0 \\ 1 & t_{1} & -t_{2} & 0 & \cdots & 0 \\ \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \cdots & 0 \\ & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] \label{beka} \end{equation the \begin{equation*} \det (B_{k,n})=F_{k,n+1}(t). \end{equation* where $t_{0}=1.$ \end{thrm} \bigskip \begin{proof} \bigskip To prove $\det (B_{k,m})=F_{k,m+1}(t)$, we use the mathematical induction on $m$. The result is true for m=1$ by hypothesis. Using Theorem (2.1) we have \begin{eqnarray*} \det (B_{m+1,k}) &=&b_{m+1,m+1}\det (B_{k,m})+\sum\limits_{r=1}^{m}((-1)^{m+1-r}b_{m+1,r}\pro \limits_{j=r}^{m}b_{j,j+1}\det (B_{r-1,k})) \\ &=&t_{1}\det (B_{k,m})+\sum\limits_{r=1}^{m-k+1}((-1)^{m+1-r}b_{m+1,r}\pro \limits_{j=r}^{m}b_{j,j+1}\det (B_{r-1,k})) \\ &&+\sum\limits_{r=m-k+2}^{m}((-1)^{m+1-r}b_{m+1,r}\pro \limits_{j=r}^{m}b_{j,j+1}\det (B_{r-1,k})) \\ &=&t_{1}\det (B_{k,m})+\sum\limits_{r=m-k+2}^{m}((-1)^{m+1-r}.\frac{t_{m-r+2 }{t_{2}^{(m-r+1)}}\prod\limits_{j=r}^{m}(-t_{2})\det (B_{r-1,k})) \\ &=&t_{1}\det (B_{k,m}) \\ &&+\sum\limits_{r=m-k+2}^{m}((-1)^{m+1-r}.\frac{t_{m-r+2}}{t_{2}^{(m-r+1)} .(-1)^{m+1-r}t_{2}^{(m-r+1)}\det (B_{r-1,k})) \\ &=&t_{1}\det (B_{k,m})+\sum\limits_{r=m-k+2}^{m}(t_{m-r+2}.\det (B_{r-1,k})) \\ &=&t_{1}\det (B_{k,m})+t_{2}\det (B_{k,m-1})+\cdots +t_{k}\det (B_{k,m-(k-1)}). \end{eqnarray* From the hypothesis and the definition of generalized Fibonacci polynomials we obtai \begin{equation*} \det (Q_{m+1,k})=t_{1}F_{k,m+1}+t_{2}F_{k,m}+\cdots +t_{k}F_{k,m-(k-2)}=F_{k,m+2}. \end{equation* Therefore, the result is true for all positive integers. \end{proof} \bigskip \begin{exam} \bigskip We obtain $6$-th generalized Fibonacci polynomial for $k=4$, using Theorem 2. \begin{equation*} F_{4,6}(t)=\det \left[ \begin{array}{ccccc} t_{1} & -t_{2} & 0 & 0 & 0 \\ 1 & t_{1} & -t_{2} & 0 & 0 \\ \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} & 0 \\ \frac{t_{4}}{t_{2}^{3}} & \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} \\ 0 & \frac{t_{4}}{t_{2}^{3}} & \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1 \end{array \right] =2t_{1}t_{4}+2t_{2}t_{3}+t_{1}^{5}+3t_{1}t_{2}^{2}+3t_{1}^{2}t_{3} \allowbreak 4t_{1}^{3}t_{2}. \end{equation*} \end{exam} \bigskip \begin{cor} \bigskip $\left[ 10\right] $Let $k\geq 2$ be an integer$,$ $f_{k,n}$ be the generalized order-$k$ Fibonacci numbers (1) and $M_{k,n}=(m_{ij})$ be an n\times n$ lower Hessenberg matrix such tha \begin{equation*} m_{ij}=\left\{ \begin{array}{l} -1\text{ \ \ \ \ \ \ \ \ \ \ \ if \ \ \ }j=i+1, \\ 1\text{\ \ \ \ \ \ \ \ \ \ \ \ \ if\ \ \ \ \ }0\leq i-j<k\text{,} \\ 0\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise \end{array \right. \end{equation* the \begin{equation*} \det (M_{k,n})=f_{k,k+n-1} \end{equation*} \end{cor} \bigskip \begin{proof} It is direct from Theorem 2.5 for $t_{i}=1.$ \end{proof} \begin{cor} \bigskip \bigskip\ If we rewrite Theorem (2.2) and Theorem (2.5) \newline i) for $t_{i}=c_{j}$ $(1\leq i,j\leq k),$ we obtai \begin{equation*} \det (Q_{k,n})=f_{k,n}^{1\text{ }} \end{equation* an \begin{equation*} \det (B_{k,n})=f_{k,n}^{1\text{ }} \end{equation* respectively, \newline ii) for $t_{1}=2$ and $t_{i}=1$ for $2\leq i\leq k,$ we obtai \begin{equation*} \det (Q_{k,n})=p_{k,n+1}^{k\text{ }} \end{equation* an \begin{equation*} \det (B_{k,n})=p_{k,n+1}^{k\text{ }} \end{equation* respectively, \newline iii) for $t_{1}=0$ and $t_{i}=1$ for $2\leq i\leq k,$ we obtai \begin{equation*} \det (Q_{k,n})=v_{k,n+1}^{k\text{ }} \end{equation* an \begin{equation*} \det (B_{k,n})=v_{k,n+1}^{k\text{ }} \end{equation* respectively. Where $f_{k,n}^{1\text{ }}$, $p_{k,n}^{k\text{ }}$, $v_{k,n}^{ \text{ }}$ be the $k$ sequences of generalized order-$k$ Fibonacci, Pell and Van der Laan numbers. Matrices $Q_{k,n}$ and $B_{k,n}$ are as in (7) and (8), respectively. \end{cor} \bigskip \section{The permanent representations} \bigskip Let $A=(a_{i,j})$ be a square matrix of order $n$ over a ring R. The permanent of $A$ is defined b \begin{equation*} \text{per}(A)=\sum\limits_{\sigma \in S_{n}}\prod\limits_{i=1}^{n}a_{i,\sigma (i)} \end{equation* where $S_{n}$ denotes the symmetric group on $n$ letters. Let $A_{i,j}$ be the $(i,j)$-th minor of matrix $A.$ The \begin{equation*} \text{per}(A)=\sum\limits_{k=1}^{n}a_{i,k}\text{per}(A_{i,k})=\su \limits_{k=1}^{n}a_{k,j}\text{per}(A_{k,j}) \end{equation* for any $i,j$. \bigskip \begin{thrm} $\left[ 10\right] $Let $A_{n}$ be $n\times n$ lower Hessenberg matrix for all $n\geq 1$ and define per$(A_{0})=1.$ Then \begin{equation*} \text{per}(A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation} \text{per}(A_{n})=a_{n,n}\text{per}(A_{n-1})+\sum\limits_{r=1}^{n-1}(a_{n,r \prod\limits_{j=r}^{n-1}a_{j,j+1}\text{per}(A_{r-1})). \end{equation} \end{thrm} \bigskip \begin{thrm} \bigskip \bigskip Let $k\geq 2$ be an integer, $F_{k,n}(t)$ be the generalized Fibonacci Polynomial and $H_{k,n}=(h_{rs})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} h_{rs}=\left\{ \begin{array}{l} i^{r-s}.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}\text{ \ \ \ \ \ \ \ \ if \ }-1\leq r-s<k\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation} H_{k,n}=\left[ \begin{array}{cccccc} t_{1} & -it_{2} & 0 & 0 & \cdots & 0 \\ i & t_{1} & -it_{2} & 0 & \cdots & 0 \\ i^{2}\frac{t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & i^{k-4}\frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & \cdots & 0 \\ & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] \end{equation the \begin{equation*} \text{per}(H_{k,n})=F_{k,n+1}(t) \end{equation* where $t_{0}=1$ and $i=\sqrt{-1}.$ \end{thrm} \bigskip \begin{proof} \bigskip Since the proof is similar to the proof of Theorem (2.2) by using Theorem (3.1) we omit the detail. \end{proof} \bigskip \begin{exam} \bigskip We obtain $7$-th generalized Fibonacci Polynomials for $k=5$, using Theorem 3. \begin{eqnarray*} F_{5,7} &=&\text{per}\left[ \begin{array}{cccccc} t_{1} & -it_{2} & 0 & 0 & 0 & 0 \\ i & t_{1} & -it_{2} & 0 & 0 & 0 \\ \frac{-t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & 0 & 0 \\ \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & 0 \\ \frac{t_{5}}{t_{2}^{4}} & \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}}{t_{2}^{2 } & i & t_{1} & -it_{2} \\ 0 & \frac{t_{5}}{t_{2}^{4}} & \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}} t_{2}^{2}} & i & t_{1 \end{array \right] \\ &=&2t_{1}t_{5}+2t_{2}t_{4}+6t_{1}t_{2}t_{3}+t_{2}^{3}+t_{3}^{2}+t_{1}^{6} \allowbreak 3t_{1}^{2}t_{4}+4t_{1}^{3}t_{3}+5t_{1}^{4}t_{2}+6t_{1}^{2}t_{2}^{2}. \end{eqnarray*} \end{exam} \begin{cor} \bigskip \bigskip \bigskip $\left[ 10\right] $Let $k\geq 2$ be an integer$,$ $f_{k,n}$ be the generalized order-$k$ Fibonacci numbers and H_{k,n}=(h_{rs})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} h_{rs}=\left\{ \begin{array}{l} i^{r-s}\text{\ \ \ \ \ \ \ if \ }-1\leq r-s<k\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* the \begin{equation*} \text{per}(H_{k,n})=f_{k,k+n-1} \end{equation*} \end{cor} \bigskip \begin{proof} \bigskip It is direct from Theorem 3.2 for $t_{i}=1.$ \end{proof} \begin{thrm} Let $k\geq 2$ be an integer$,$ $F_{k,n}(t)$ be the generalized Fibonacci Polynomial and $L_{k,n}=(l_{ij})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} l_{ij}=\left\{ \begin{array}{l} \frac{t_{i-j+1}}{t_{2}^{(i-j)}}\text{\ \ \ \ \ \ \ \ \ \ \ if \ }-1\leq i-j< \text{,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation} L_{k,n}=\left[ \begin{array}{cccccc} t_{1} & t_{2} & 0 & 0 & \cdots & 0 \\ 1 & t_{1} & t_{2} & 0 & \cdots & 0 \\ \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & t_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \cdots & 0 \\ & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] . \end{equation where $t_{0}=1,$ the \begin{equation*} \text{per}(L_{k,n})=F_{k,n+1}. \end{equation*} \end{thrm} \bigskip \begin{proof} \bigskip This is similar to the proof of \ Theorem 2.5 using Theorem 3.1. \end{proof} \bigskip \begin{cor} $\left[ 9\right] $\bigskip Let $k\geq 2$ be an integer$,$ $f_{k,n}$ be the generalized order-$k$ Fibonacci numbers and $D_{k,n}=(d_{ij})$ be an n\times n$ lower Hessenberg matrix such tha \begin{equation*} d_{ij}=\left\{ \begin{array}{l} 1\text{\ \ \ \ \ \ \ \ \ \ \ if \ }-1\leq i-j<k\text{,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* the \begin{equation*} \text{per}(D_{k,n})=f_{k,k+n-1}. \end{equation*} \end{cor} \begin{proof} \bigskip It is direct from Theorem 3.5 for $t_{i}=1.$ \end{proof} \begin{cor} If we rewrite Theorem (3.2) and Theorem (3.5) \newline i) for $t_{i}=c_{j}$ $(1\leq i, j\leq k), $ we obtai \begin{equation*} \text{per}(H_{k,n})=f_{k,n}^{\text{ }1} \end{equation* an \begin{equation*} \text{per}(L_{k,n})=f_{k,n}^{\text{ }1} \end{equation* respectively, \newline ii) for $t_{1}=2$ and $t_{i}=1$ $(2\leq i\leq k),$ we obtain \begin{equation*} \text{per}(H_{k,n})=p_{k,n+1}^{\text{ }k} \end{equation* an \begin{equation*} \text{per}(L_{k,n})=p_{k,n+1}^{\text{ }k} \end{equation* respectively, \newline iii) for $t_{1}=0$ and $t_{i}=1$ $(2\leq i\leq k),$ we obtain \begin{equation*} \text{per}(H_{k,n})=v_{k,n+1}^{\text{ }k} \end{equation* an \begin{equation*} \text{per}(L_{k,n})=v_{k,n+1}^{\text{ }k} \end{equation* respectively. Where $f_{k,n}^{k\text{ }}$, $p_{k,n}^{k\text{ }}$, $v_{k,n}^{ \text{ }}$ be the $k$ sequences of generalized order-$k$ Fibonacci, Pell and Van der Laan numbers. Matrices $H_{k,n}$ and $L_{k,n}$ are as in (10) and (11), respectively. \end{cor} \bigskip
\section{Introduction} The top quark, discovered at Fermilab in 1995, completed the three generation structure of the Standard Model, SM. The top quark is distinguished by its large mass, ($m \sim 172~$GeV). Precision measurements in the top quark sector will shed light on the electroweak symmetry breaking mechanism and indirectly on the Higgs mechanism of elementary particle mass generation. In particular, the measurement of the top quark pairs ($t\bar{t}$) production cross section is an important test of QCD perturbative calculations as well as an estimation of one of the major background sources for several new physics signatures. The full hadronic $t\bar{t}$ is important background for several analyses beyond the Standard Model: producton of new particle(s) decaying to many hadronic jets in association with missing trasverse momentum, predicted by SUSY; search for the Higgs boson in the all hadronic ($bb+$jets) final state, such as associated production with vector boson or top pair. In proton-proton collisions, top-antitop pairs are created when a parton from each colliding proton interact through the strong force. The production mechanisms at the LHC are the gluon-gluon fusion (85\%) and $q\bar{q}$ annihilation (15\%) Within the SM, the top quark decays into a $W$ boson and a $b$-quark almost ${100\%}$ of the time. The W boson subsequently decays into either a pair of quarks or a lepton-neutrino pair. In the $t\bar{t}$ production in the fully hadronic final state both $W$s decay hadronically. The experimental signature of fully hadronic $t\bar{t}$ is characterized by a nominal six-jet topology with $b$-jets. This channel has the advantage of a large BR, top pair decay in full hadronic signature in (${44\%}$) of the case, although it suffers from a large QCD multijet background.\\ The ATLAS \cite{ATLAS} detector at the LHC covers nearly the entire solid angle around the collision point. It consists of an inner tracking detector divided in three independent subsystems immersed in a 2T magnetic field generated by the central solenoid that reconstructs charged particle trajectories and measures their momentum, an electromagnetic calorimeter that identifies and measures the electrons and photons and a hadronic calorimeter that identifies jets formed by the hadronization of quarks and muon spectrometer that identifies muons and measures their deflections in the magnetic field produced by a toroid magnet system (4T). \vspace{-0.37cm} \section{Analysis method} \vspace{-0.01cm} During the 2010, ATLAS recorded ${\sim 36\ \mathrm{pb^{-1}}}$ of p-p collisions at center of mass energy of 7 TeV.\\ The data sample has been collected with unprescaled multijet triggers, requesting at least four jets with pseudorapidity\footnote[1]{In the right-handed ATLAS coordinate systems, the pseudorapidity $\eta$ is defined as $\eta = - \ln[\tan(\theta/2)]$, where the polar angle $\theta$ is measured with respect to the LHC beamline. Transverse energy is defined $E_T = E \sin\theta$.}, $|\eta|$, less than 3.2 and transverse energy, $E_T$, great than 30 GeV; the event selection requirement is of at least four jets with $E_T>60~$~ GeV \cite{koei}. The trigger efficiency for a jet with its $E_T = 60$ GeV is 90\%. The identification of jets originating from $b$-quarks ($b$-tagging) is performed using a secondary vertex-based tagger algorithm which reconstructs the inclusive vertex formed by the decay products of bottom hadron.\\ \section{Background modeling} The QCD multijet background shape is modelled by a data-driven method applying the tag-rate function. The signal is defined as events with 6 or more jets with 2 $b$-tagged jets. The baseline tag rate function (TRF) to be used for the backgroud modelling in the signal region are derived from the 5 jets control bin with 2 $b$-tags, as seen in Fig. \ref{TRF}. The tag rates are evaluted separately for 1 $b$- and 2 $b$-tag events, because the event topology in the 2 $b$-tag case is different due to gluon splitting process ($g\rightarrow b\bar{b}$). The background tag rate for a jet (TR) is defined as: \begin{equation} TR_{nb_{bin},nj_{bin}} = \frac{njet_{tag}}{njet_{all}} \times \frac{nj_{bin}}{nb_{bin} \times {_nj_{bin}}C_{nb_{bin}}} \end{equation} The $TR_{nb_{bin},nj_{bin}}$ function is calculated in each jet and b-tag multiplicity bin separately. The variables $njet_{all}$ and $njet_{tag}$ are respectively the total number of jets and the number of $b$-tagged jets, the symbol C stands for the binomial coefficient $(_{n}C_{r} = \frac{n!}{(n-r)!r!})$ for jets to be selected as $b$-tagged jets in the event. The variables ${njet_{tag}}/{nb_{bin}}$ and ${njet_{all}}/{nj_{bin}}$ correspond to the number of tagged events in $nb_{bin}$ and $nj_{bin}$ and the number of pretag events in ${nj_{bin}}$, respectively. The coefficient $_{n}C_{r}$ is a factor to make this probability on a per-jet basis. $TR_{nb_{bin},nj_{bin}}$ is to be applied to a jet to obtain an event weight. The TRF is parametrized as a function of the jet $p_T$ and $\eta$ to take into account a possible dependence on those quantities. The weight for an event with 6 jet (2 $b$-tag) is obtained applying as follows the TRF: \begin{equation} w(6j,2b)= \sum_{k=1}^{nj_{bin}} TR^k_{njb_{bin},5j} \times \frac{{_{nj_{bin}}}C_{nb_{bin}}}{nj_{bin}} \end{equation} After applying TRF kinematic distribution are compared in control region, 4 jet 2 $b$-tag events, see fig. \ref{kin}.\\ \begin{figure} \includegraphics[scale=0.23]{fig_01a.pdf} \caption{Background tag rate function as a function of jet $p_T$ from data \cite{koei}.}\label{TRF} \end{figure} \begin{figure} \includegraphics[scale=0.23]{fig_02a.pdf} \caption{Kinematic distributions for 4-jet 2 $b$-tag events (Control Region) \cite{koei}. }\label{kin} \end{figure} \section{Fitting procedure} The background normalization, for the cross section measurement, is extracted from a fit to the mass $\chi^2$ variable \begin{equation} \chi^2 = \sum_{i=1}^2 (\frac{m_{jjb}^{(i)}-m_{top}}{\sigma_{top}})^2 +(\frac{m_{jj}^{(i)}-m_{W}}{\sigma_{W}})^2 \end{equation} Only the leading 6 jets are considered in $\chi^2$ computation for $N_{jet}\ge~6$ events. In the case of 2 $b$-tag events (signal) there are 6 combination, and we select the combination with the minimum $\chi^2$ to build the final $\chi^2$ distribution. \subsection{$t\bar{t}$ cross section measurement analysis with 2010 data} The final mass $\chi^2$ distribution is fitted with signal and background template. Figure \ref{fig:fit1} shows the fit result of the minimum mass $\chi^2$ distribution. The cross section $\sigma_{t\bar{t}}$ is obtained b \begin{equation} \sigma_{t\bar{t}} = \frac{N_{obs} \times f_S}{\epsilon \times \int L dt} \end{equation} where $f_s$ is the signal fraction of 6.4\% from the fit; the factor $\epsilon$ of 1.8\% includes signal acceptance and branching franction. \begin{figure}[h!] \includegraphics[scale=0.24]{fig_09.pdf}\caption{Fit result of minimum mass $\chi^2$ distribution in 6-jet with 2 $b$-tag events (Signal Region) \cite{koei}.}\label{fig:fit1} \end{figure} \begin{table}[] \begin{center} \caption{Fit results for events with at least 6-jets and 2 $b$-tags} \begin{tabular}{|c|c|} \hline \textbf{Source} & \textbf{Number of events} \\ \hline Background & 1097.0 \\ $t\bar{t}$ & 75.0$\pm $46.5 (stat.)\\ \hline Data ($N_{obs}$) & 1172\\ \hline \end{tabular} \end{center} \end{table} The fitted cross section with 36 $\mathrm{pb^{-1}}$ is $\sigma_{pp \rightarrow t\bar{t}} = 118 \ \pm \ 73 (\mathrm{ stat.} ) \ \pm\ 48 (\mathrm{ syst.} ) \ \pm \ 4 (\mathrm{ lumi.} ) \mathrm{pb}$. The significance of the fitted value is $1.6 \ \sigma$ whereas the expected sensitivity was $2.2 \ \sigma$. The observed one-sided upper limit at 95\% confidence level is $\sigma_{t\bar{t}}<~261\mathrm{ pb}$. The total systematic uncertainty on the $t\bar{t}$ cross-section is 41\%. The most important systematic sources considered are Jet energy scale uncertainty(JES), Trigger efficiency, b-tagging, Background modeling, Initial \& Final State Radiation (ISR \& FSR). Table \ref{syst} shows a summary of the most important individual contributions. \begin{table}[h!] \begin{center} \caption{Summary of the most important individual contributions on the systematic uncertainties on the fitted cross-section value} \begin{tabular}{|c|c|} \hline \textbf{Source} & $\Delta \sigma/ \sigma$ \\ \hline JES & 17\%\\ Trigger & 10\%\\ b-tagging & 29\%\\ Background modeling & 7\%\\ ISR/FSR & 16 \%\\ \hline \end{tabular}\label{syst} \end{center} \end{table} \vspace{-0.8cm} \section{ Outlook: $b$-jet trigger} The ATLAS trigger and data acquisition system is based on three levels. Trigger levels must provide a rejection to reduce the 40 MHz bunch-crossing rate to an output of about few hundred Hz. The level 1 is hardware based, uses the calorimeter and muon spectrometer with coarse granularity; the level 2 is software-based, exploits regions of interest identified by the level 1 and accesses data from all sub detectors with full granularity; the Event Filter (EF) runs offline-quality software-based algorithms. \subsection{Online $b$-tagging} The identification of jets stemming from the hadronization of $b$-quarks is made possible by the\\relatively long lifetime of hadrons containing $b$-quarks (lifetime of the order of $1.5\ ps$ corresponding to $c\tau\approx 450\mu m$). This allows the identification of $b$-jets from the one containing only lighter quarks. Given the high instantaneous luminosity the LHC will deliver in the next year, $b$-tagging at HLT is a possibility for collecting $t\bar{t}$ in the full hadronic final state with an acceptable data taking rate. \vspace{+0.2cm} \subsection{Future analysis} At the high instantaneous luminosity foreseen for LHC data taking, most multi-jet trigger will be prescaled unless their threshold and jet multiplicity are constantly increased to keep the trigger rate under control. For this reason ATLAS has put in place a combination of multijet and b-jet trigger to efficiency select events with final states containing several $b$-jets.\\ The $b$-jet trigger for hadronic top requires four EF-jets with $E_T>~30~$GeV at electromagnetic scale and 1 $b$-jet with $E_T>~10~$GeV at EM scale and tight instance for the $b$-tagging criteria, as seen in Fig. \ref{b1}. The signal efficiency is $\sim40\%$. \begin{figure}[h!] \includegraphics[scale=0.27]{1b-4j.pdf}\caption{Trigger rate for 1b/4j topology. LVL1, LVL2 and EF rate of a b-jet trigger requiring at least four jets in the event and at least one $b$-tagged jet \cite{btrig}.} \label{b1} \end{figure} \vspace{-0.8cm} \section{Conclusions} The analysis is performed using 36 $\mathrm {pb^{-1}}$ of pp collisions produced at the LHC with a center-of-mass energy of $\sqrt{s} = 7$ TeV and recorded with the ATLAS detector. A 95\% confidence level limit is set at 261 pb, compatible with the expected Standard Model cross-section of $165^{+11}_{-16}$ pb. \vspace{-0.5cm} \bigskip
\section{Introduction} Structure formation and evolution are thought to proceed via hierarchical merger events and accretion of smaller units to form larger systems driven by gravity with dark matter (DM) dominating the gravitational field. Thus, merging plays a key role in driving the build-up of structures on both small and large scales. Clusters are located at the crossing point of filamentary structures which drive the accretion towards their DM potential wells. Accretion of smaller systems happens preferentially in the filaments and the products are, in turn, accreted by larger clusters (e.g. \citealt{2005Natur.435..629S}). Investigating massive merging clusters implies having a snapshot of the regions of the Universe where the highest-mass structures are created, which offers constraints and insights to both cosmology and astrophysics. Indeed, the study of galaxy clusters allows not only to highlight the structure of the Universe and to quantify several cosmological parameters but also to study the dynamics and evolution of the baryonic fraction in a deep potential well \citep{2009arXiv0906.4370B}. In most cases the dynamical state of clusters has been derived from X-ray observations (e.g. \citealt{1992csg..conf...49J} and \citealt{2002ASSL..272...79B}). X-ray studies of merging clusters are useful to provide information on the evolution of the intra-cluster medium (ICM, e.g. \citealt{2002ASSL..272....1S}), for example on how the ICM is heated to the high observed temperatures and how entropy structure is generated (\citealt{2001ApJ...546...63T}, \citealt{2010A&A...511A..85P}), in particular through the effect of shocks (e.g. \citealt{2007PhR...443....1M} for a review). Merging events are detected by substructures in the surface brightness distribution of X-ray images (e.g. \citealt{2001A&A...378..408S}). By their characterisation we are able to study the cluster dynamics and mass assembly history. For example, combining X-ray and spectroscopic information \cite{2011ApJ...728...27O} analysed the merging cluster Abell~2744. The identification of substructures was crucial to determine the merging direction and approximately establish the status of the core passage. However, accretion events can be also identified by considering all infalling groups in the proximity of a cluster \citep{2009MNRAS.400..937M} and deriving their properties. These groups can be detected through the X-ray emission of their gas (e.g. \citealt{2000ARA&A..38..289M}) and through the distribution of galaxies and optical light. Direct detection of X-ray emission from groups is not always feasible: indeed, small associations of a few galaxies are normally undetectable in X-rays \citep{2009ApJ...704..564F} due to their weak emission, and even richer groups, once captured by a massive cluster, are likely to lose almost all their gas. Nevertheless, it is likely that the galaxies of accreted groups will give rise to significant galaxy over-densities within the parent cluster \citep{2006PASP..118..517B}. Combining these two complementary approaches makes it possible to explore the history of formation and evolution of clusters, where several snapshots of the accretion pattern are highlighted just by observing one cluster out to several times its radius. In order to understand the role of merging in the evolution of clusters with similar mass and the connection with their large scale structure (LSS) environment we need a sample of galaxy clusters with small mass range and different dynamical states. With this purpose we have selected a statistically complete cluster sample drawn from the ROSAT ESO Flux Limited X-ray (REFLEX) survey \citep{2001A&A...369..826B}. The 13 distant X-ray luminous (DXL, see e.g. \citealt{2004astro.ph..2533Z} and \citealt{2005AdSpR..36..667Z} for details) clusters have luminosity $\mathrm{L_X^{bol}=0.5 - 4\times 10^{45} \ erg \ s^{-1}}$, masses $\mathrm{M_{500}= 0.5- 1.1\times 10^{15}M_\odot}$\footnote{$\mathrm{M_\Delta}$ (where $\mathrm{\Delta=500,200}$) is defined as $\mathrm{M_\Delta=(4 \pi/3) \Delta \rho_c R_{\Delta}^3}$ where $\mathrm{R_\Delta}$ is the radius at which the density of a cluster is equal to $\Delta$ times the critical density of the Universe ($\mathrm{\rho_c}$). Throughout our analysis we adopt the X-ray estimate of $\mathrm{R_\Delta}$ (unless it is otherwise specified).} and are located within a narrow redshift interval ($\mathrm{z=0.27-0.31}$). The DXL sample is a powerful instrument to investigate the mass assembly of the clusters and the evolution of galaxies therein, together with the exploration of the link between large-scale structure, substructure and galaxy population. Moreover, this snapshot of the Universe is comparable to N-body simulations including hydrodynamics (e.g. \citealt{2009arXiv0906.4370B}, \citealt{2011ApJ...728...54Z}), allowing us to better investigate the physics that regulates cluster evolution across the cosmic web. For instance, one can understand the variation of the sub-halo mass function, traced by galaxies and the amount of substructure in clusters; estimate time scales from the dynamical state of the gas; and understand the physical processes that drive its behaviour. In fact, DXL can be represented as a sequence of cluster dynamical states, starting from early stages of merging events (including several components of different mass) towards strong cool core clusters. Most of the detailed X-ray analysis of DXL clusters has been already performed ( \citealt{2004astro.ph..2533Z}, \citealt{2004A&A...413...49Z}, \citealt{2005AdSpR..36..667Z}, \citealt{2006A&A...456...55Z}, hereafter referred to as Z06, \citealt{2005A&A...442..827F}). These studies, focused on the ICM, have provided valuable information on the dynamical state, AGN feedback and chemical enrichment of these clusters. Optical analysis of the DXL sample has started in parallel. In particular, the attention was focused on two clusters with different dynamical states: an ongoing major merger \citep{2007A&A...470..425B} and a relaxed cluster (\citealt{2009A&A...500..947B}, hereafter referred to as B09), in order to compare their kinematics and galaxy distributions with their X-ray properties. \citet{2007A&A...470..425B} found, associated with the merging cluster A\,2744, two large-scale filaments along which blue galaxies exhibited enhanced star formation activity. This study was followed up in B09 who explored the existing link between the fraction of passively evolving galaxies and the assembly state of the cluster. \citet{2008A&A...483..727P} then suggested, from observations of three DXL clusters, that the intra cluster light has multiple origins, possibly linked to the dynamical state of the cluster. In this paper we present the results from the study of RXCJ1131.9\textendash 1955 (alias Abell 1300), a post-merging cluster at $\mathrm{z \sim 0.3075}$ with a ``dumbbell'' cD galaxy \citep{1997A&AS..124..283P} at its centre and prominent filaments visible in the galaxy density distribution. The definition of post-merging cluster dates back to \citet{1997A&A...326...34L} who state that this cluster has undergone a merger but the merging phase may be nearly over. We support this statement after the inspection of the shape of the X-ray emission (Fig.~\ref{BCG_fig}), not clearly separated from the dark matter halo, and the displacement observed between the X-ray peak and the Brightest Cluster Galaxy (BCG), as better explained later on. This paper is organised as follows: in Section 2 we describe our dataset and our preliminary analysis; in Section 3 we give a morphological overview of the cluster and its large scale structure using optical (photometric and spectroscopic) and X-ray analyses; in Section 4 we discuss our results comparing them with simulations and previous works; we draw our conclusions in Section 5. Throughout our analysis we adopt the AB magnitude system (unless otherwise specified) and the following cosmological values: $\mathrm{H_0=70} \mathrm{\ km\ s^{-1}\ Mpc^{-1}}$, $\Omega_\mathrm{M}=0.3$ and $\Omega_\mathrm{\Lambda}=0.7$. At the cluster distance 1$\arcmin$ corresponds to 270 kpc. In all figures hereafter North is up and East to the left, unless otherwise specified. \section{Observations, Data Reduction and Analysis} \subsection{Wide-field Imaging} Optical photometry was carried out using the Wide Field Imager (WFI, \citealt{1999Msngr..95...15B}) mounted on the Cassegrain focus of the ESO/MPG-2.2~m telescope at La Silla, Chile. The data presented here were obtained as part of a heterogeneous programme during MPG observing time in visitor mode (P.I.: B\"{o}hringer). The observations of A1300 in the B, V and R passbands were performed in 2001, between January 27th and February 1st, in photometric conditions. They were divided into sequences of 8 dithered sub-exposures for a total exposure time of $\mathrm{3150 \ s}$ for the V band and $\mathrm{3600 \ s}$ for the R and B bands. Filter curves can be found in \citet{2001A&A...379..740A} and on the web-page of the La Silla Science Operation Team\footnote{http://www.eso.org/sci/facilities/lasilla/instruments/wfi/inst/ filters}. Standard stars were observed in all the four nights: three Landolt fields \citep{1992AJ....104..340L} were targeted for a total of 35-50 standard stars per field (SA98, SA101 and SA104). The WFI data were reduced using the data reduction system developed for the ESO Imaging Survey (EIS, \citealt{1997Msngr..87...23R}) and its associated EIS/MVM image processing library version 1.0.1 ($Alambic$\footnote{http://www.eso.org/sci/activities/projects/eis/survey\_$ $release. html}). For more details on the transformation of raw images into reduced ones see \citet{2008A&A...483..727P}. \begin{table} \begin{center} \begin{tabular}[t]{cccc} \hline Passband & ZP & $k$ & CT \\ \hline B & 24.55$\pm$0.01 & 0.20$\pm$0.01 & 0.24 \\ V & 24.15$\pm$0.01 & 0.14$\pm$0.01 & -0.12 \\ R & 24.43$\pm$0.01 & 0.09$\pm$0.01 & 0.01 \\ \hline \end{tabular} \end{center} \caption[zp work]{Photometric solutions determined in this work. Column 1 gives the passband, column 2 the zero-points in the Vega magnitude system, column 3 the extinction coefficient and column 4 the colour term. Two of the three quantities are reported with respective errors in the Vega magnitude. These best-fit parameters were obtained from a two-parameter fit from about 850 to 2000 measurements across the WFI field for each passband, the colour term being fixed. } \label{zp_work} \end{table} Source detection and photometry were based on SExtractor \citep{1996A&AS..117..393B} both for standard and science images. Magnitudes were calibrated to the Johnson-Cousins system using \citet{1992AJ....104..340L} standard stars whose magnitudes were obtained using a 10 arcsec-wide circular aperture, which were adequate as judged from determining the average growth curve of all the measured stars. Photometric standards were observed over a rather broad range of airmasses, but science frames were taken at the best airmass; in this way we were able to obtain photometric solutions (e.g. zero-points) for the calibration of reduced scientific images by merging the measurements of standard stars for each passband. The number of non-saturated Landolt stars per field did not allow independent solutions to be determined for each of the eight chips of WFI. Hence calibration had to rely upon solutions based on measurements taken across all chips. Although the EIS data reduction system includes a photometric pipeline for the automatic determination of photometric solutions, these were determined interactively using the IRAF\footnote{IRAF is the Image Reduction and Analysis Facility, a general purpose software system for the reduction and analysis of astronomical data. IRAF is written and supported by the IRAF programming group at the National Optical Astronomy Observatories (NOAO) in Tucson, Arizona. NOAO is operated by the Association of Universities for Research in Astronomy (AURA), Inc. under cooperative agreement with the National Science Foundation.} task $fitparams$. This choice allows the interactive rejection of individual measurements, stars, and chips. Photometric solutions with minimum scatter were obtained by a two-parameter linear fit with about 850 photometric points for the B-band, about 900 for the V-band and more than 2000 for the R-band, the atmospheric extinction coefficient in each band being set equal to that listed as the median value obtained by the EIS team\footnote{http://www.eso.org/sci/activities/projects/eis/surveys/readme/ 70000027}. In general, zero-points and colour terms are consistent with those obtained by the EIS team or by the 2p2 Telescope Team\footnote{http://www.eso.org/sci/facilities/lasilla/instruments/wfi/zero- points}, as can be seen by comparing Tables \ref{zp_work}\textendash \ref{zp_2p2}. \begin{table} \begin{center} \begin{tabular}[t]{cccc} \hline Passband & ZP & $k$ & CT \\ \hline B & 24.65 & 0.23 & 0.24 \\ V & 24.19 & 0.15 & -0.12 \\ R & 24.47 & 0.12 & 0.01 \\ \hline \end{tabular} \end{center} \caption{Median values for all the photometric solutions in the Vega magnitude system based on three parameter fits obtained by the ESO Deep Public Survey (DPS) team.} \label{zp_eis} \end{table} \begin{table} \begin{center} \begin{tabular}[t]{cccc} \hline Passband & ZP & $k$ & CT \\ \hline B & 24.81$\pm$0.05 & 0.22$\pm$0.015 & 0.25$\pm$0.01 \\ V & 24.15$\pm$0.04 & 0.11$\pm$0.01 & -0.13$\pm$0.01 \\ R & 24.47$\pm$0.04 & 0.07$\pm$0.01 & 0.00$\pm$0.00 \\ \hline \end{tabular} \end{center} \caption{Definitive photometric solutions obtained by the 2p2 Telescope Team from observations of standard stars in perfectly photometric nights, where a bunch of standard star fields were moved around each chip of WFI. All parameters were fitted simultaneously as free parameters, with good airmass and colour range, and around 300 stars per fit. The Table below gives the average solutions in the Vega magnitude system over all chips.} \label{zp_2p2} \end{table} As for science images, source extraction and photometry were obtained after matching the BVR images of each target to the worst seeing (0.92$\arcsec$ for the V band), using the IRAF task $psfmatch$, and taking into account the weight-maps associated with the individual images, produced by $Alambic$. A common configuration file was used to produce three catalogues per target, after evaluating the seeing and the zero-point for individual images. The R-band image having the deepest exposure was used as the detection image, where sources are defined by an area with a minimum number of 5 pixels above a threshold of $1\sigma$ of the background counts. Source photometry in individual passbands was extracted in fixed circular apertures (between $\mathrm{1.2\arcsec}$ and $\mathrm{10\arcsec}$ in diameter) or in flexible elliptical apertures (Kron-like, \citealt{1980ApJS...43..305K}) with a Kron-factor of 2.5 and a minimum radius of 3.5 pixels. For our analysis we used the total magnitudes (Kron-like). Object magnitudes were corrected for Galactic extinction according to the \citet{1998ApJ...500..525S} galactic reddening maps (from NASA/IPAC Extragalactic Database, NED) and converted to the AB system according to the response function of the optical system (see \citealt{2004A&A...428..339A}). The output catalogues were successively culled of fake sources by hand and by using masks before photometric redshifts were determined. In addition, stars and galaxies could be safely identified on the basis of their surface brightness profile and optical colours down to $\mathrm{R = 21.5}$. Fainter than this limit, number counts are dominated by galaxies, so that all detected objects with $\mathrm{R > 21.5}$ are assumed to be galaxies. This assumption is supported by several tests that we ran, comparing colours and magnitudes derived with different methods. Information from SExtractor flags was also taken into account in these tests. \begin{figure} \includegraphics[width=\hsize]{deep_errors.png} \caption{Comparison of number counts for the entire region of A1300 imaged with WFI in this work and three deep fields/wide area surveys: VVDS (red circles), COSMOS (green stars) and SDSS (orange diamonds). All errors, but those of COSMOS (for which errors were available, \citealt{2004AJ....127..180C}), were obtained using the modified Poisson statistics in \citet{1986ApJ...303..336G}. } \label{deep_survey} \end{figure} Depth and quality of the final catalogues were also determined. Galaxy number counts in the observed field were compared with deep number counts from several surveys (i.e. VVDS, VIMOS VLT Deep Survey, \citealt{2003A&A...410...17M}; COSMOS, Cosmic Evolution Survey, \citealt{2004AJ....127..180C}; and SDSS, Sloan Digital Sky Survey, \citealt{2001AJ....122.1104Y}). All errors, but the ones from COSMOS (for which errors were available, \citealt{2004AJ....127..180C}), were obtained using Poissonian statistics from \citet{1986ApJ...303..336G}. Our galaxy number counts exceed those obtained from observations of deep fields such as VVDS \citep{2003A&A...410...17M} and COSMOS \citep{2004AJ....127..180C} for $\mathrm{18.5 \leq R \leq 23.5 \ mag}$, as shown in Fig.~\ref{deep_survey}. On the other hand, they begin to drop below the galaxy number counts in deep fields/wide area surveys at $\mathrm{R > 23.5}$, where the number of background galaxies dominates the number of likely cluster members. The bright end is comparable with the number counts of SDSS \citep{2001AJ....122.1104Y}. Assuming as a completeness limit the magnitude at which the observed counts are equal to 50\% of those in the deep fields/wide area surveys, we thus conclude that our R-band selected catalogues are complete down to $\mathrm{\sim24\ mag_{AB}}$ with respect to the VVDS. We identify and fit the cluster red-sequence from the colour-magnitude diagram (CMD, Fig.~\ref{CMD}) of all the galaxies in the imaged region of A1300 within a cluster-centric distance of $\mathrm{5.66 {\arcmin}}$, corresponding to $\mathrm{R_{200}} = 1.53$ Mpc at the cluster redshift (cf. Z06). The best fit (dashed line in Fig.~\ref{CMD}), obtained through recursive $\mathrm{3\sigma}$ clipping, is described by the linear relation: \begin{equation} (B-R)=(2.823 \pm 0.090) -(0.048 \pm 0.005)\times R. \label{CMD_equation} \end{equation} \begin{figure} \includegraphics[angle=90,width=\hsize]{col_mag_diagram.png} \caption{Colour-magnitude diagram (B-R versus R) for all galaxies found in the field of A1300 comprised within $\mathrm{R_{200}}$ of the cluster. Red stars represent galaxies defining the red sequence within $\mathrm{3\sigma}$ from the best-fit within the photometric errors, blue triangles show the galaxies bluer then the red sequence of $\mathrm{3\sigma}$ within the photometric errors, green diamonds all galaxies redder than the red sequence galaxies of $\mathrm{3\sigma}$ within the photometric errors. The dashed line represent the red sequence best fit and the dotted line the $\pm 1\sigma$ scatter around the red sequence. The catalogue is cut for the magnitude limit of $\mathrm{B=24.9}$ and $\mathrm{R=24.5}$.} \label{CMD} \end{figure} The rms scatter around the red sequence fit is 0.09 mag (represented by the dotted lines in Fig.~\ref{CMD}). We also compiled a CMD for the spectroscopic and photometric members separately as in B09: since these fits did not show a remarkable difference w.r.t. the one in Fig.~\ref{CMD}, we decided to use equation \ref{CMD_equation}, being based on larger statistics. The identification of the red sequence allowed us to distinguish between red (objects within $\mathrm{3\sigma}$ from the red sequence, likely to be old passively evolving objects) and blue (below the red sequence, likely to be star forming objects) galaxies and to trace their distribution in the cluster field. \subsection{Spectroscopic Data} \label{spectro_data_descr} Multi-Object Spectroscopy (MOS) was performed between the periods of May 21-23, 2004 and January 12-19, 2005 as part of the ESO GO large programme 169.A-0595 (P.I. B\"ohringer), carried out in visitor and service modes. The main aim of this program was to observe the largest number of galaxies lying in the same area of the sky for seven out of the 13 DXL clusters. Low resolution ($\mathrm{R=200}$, LR-Blue grism) spectroscopy was carried out with VIMOS (VIsible Multi-Object Spectrograph, \citealt{2003SPIE.4841.1670L}) mounted on VLT-UT3 at Paranal Observatory (ESO), Chile. VIMOS is a wide-field imager and multi-object spectrograph operating in the visible (from 3600 to 10000 \AA{}), with an array of 4 identical CCDs with a field of view (FOV) of $\mathrm{7\arcmin \times 8\arcmin}$ each and $0.205\ \arcsec$ pixel scale, separated by a gap between each quadrant of $\mathrm{\sim2\arcmin}$. To provide a good coverage of the cluster central region and to extend the analysis to the cluster outskirts we used three pointings that partially overlap in the centre and reach beyond a distance of 4 Mpc in the E-W direction from the cluster X-ray centre (Fig.~\ref{vimos_pointings}). \begin{figure} \includegraphics[width=\hsize]{rxcj1131_dens_map_vimos_masks_real1.jpg} \caption{R-band image of A1300 on which are overlaid density contours (in green) and the three VIMOS pointings. Each set of four boxes with the same colour represents one VIMOS pointing.} \label{vimos_pointings} \end{figure} Objects with $\mathrm{I\leq 22.5}$ were selected from the pre-imaging with VLT-VIMOS, corresponding to a limiting magnitude of approximately $\mathrm{I^\star + 3}$ galaxy at the redshift of the cluster (0.3075, see \citealt{1998ApJ...497..188C}). The pre-imaging was done in the I-band in order to select targets based only on stellar mass (e.g. \citealt{1994ApJS...95..107W}) and avoiding any colour bias. The catalogue on the pre-imaging was produced running SExtractor \citep{1996A&AS..117..393B} which allowed to classify galaxies with $\mathrm{I \leq 20}$ as bright and objects with $\mathrm{20 < I \leq 22.5}$ as faint, in order to observe them with two different masks (prepared using the VMMPS\footnote{http://www.eso.org/sci/observing/phase2/VIMOS/VMMPS. html} tool from ESO) for the same Observation Block (OB). Slits of $1 \arcsec$ width were used for an expected uncertainty on the observed velocities of $\mathrm{250-300\ km\ s^{-1}}$. Previous works (\citealt{2007A&A...470..425B} and B09) have confirmed these expectations and shown that with these uncertainties it is possible to establish the membership of a galaxy and the global cluster dynamics for the massive clusters of REFLEX\textendash DXL. At the average redshift of the DXL ($\mathrm{z\sim0.3}$) the LR-Blue grism samples several important spectral features: [OII], [OIII], CaII$_{\mathrm{H+K}}$, $\mathrm{H_\beta}$, $\mathrm{H_\delta}$ lines and the 4000\,\AA \,break. The combination of these features allows us to characterise the spectral type of galaxies (e.g. by the 4000\,\AA \,break), the present star formation rate (e.g. by the [OII] line) and nuclear activity (by the [OII]/[OIII] line ratio). The spectrum in this wavelength range does not suffer from fringing and spectroscopic redshifts up to $\mathrm{z\sim0.8}$ can be determined. The spectroscopic observations provided about 900 spectra which were reduced using the dedicated software VIPGI\footnote{VIPGI (VIMOS Interactive Pipeline and Graphical Interface, \citealt{2005PASP..117.1284S}). This software was developed by the VIRMOS Consortium to handle the reduction of the VIMOS data for the VVDS (\citealt{2003A&A...410...17M})}. VIPGI allows to calibrate all spectra through a user-friendly interface, applying flat fields corrections, sky subtraction, spectrum extraction and wavelength calibration. Data reduction followed the standard approach, as described in the VIPGI manual and in \citet{2005PASP..117.1284S}: lines were fitted and matched with available line catalogues. Furthermore, we used the template fitting procedure EZ (Easy redshift, \citealt{2010PASP..122..827G}), that allows to fit spectral templates to the continuum when no evident features (e.g. emission lines) were present. Although EZ mainly relies on a $\mathrm{\chi^2}$ template-fitting procedure it allows also to choose the best template after an analysis by eye. We assigned different flags to the redshifts according to their reliability: $flag=0$ was given when it was not possible to assign a redshift, $flag=1$ meant a confidence in the redshift within 25\%, $flag=2$ a reliability of the solution comprised between 25\% and 50\%, $flag=3$ between 50\% and 75\% and $flag=4$ between 75\% and 100\%. In order to ensure reliable cluster memberships we only use galaxies with spectral flags larger than 2 in the following. We expanded our sample of spectroscopic redshifts with 96 publicly available ones from NED. \begin{figure*} \includegraphics[width=0.48\hsize]{rxcj1131_OK_caustics.png} \includegraphics[width=0.48\hsize]{RXCJ1131_veldisp.png} \caption{Left panel: phase-space diagram of the confirmed 230 spectroscopic cluster members for A1300. Crosses mark the candidate cluster members prior to the interloper rejection, while red diamonds identify the confirmed cluster members. The dashed lines track the caustics defined by \citet{2004ApJ...600..657K} for comparison. Right panel: velocity distribution of the spectroscopic cluster member galaxies. The black solid histogram shows the distribution of all cluster members; the distribution of velocities for blue and red members (as defined from their position in the colour-magnitude diagram, cf. Section \ref{photometric_redshifts_section}) is shown by the dashed red and dotted blue histograms, respectively. } \label{members} \end{figure*} \subsubsection{Cluster membership} \label{photometric_redshifts_section} Spectroscopic cluster members were identified using the same combination of techniques summarised by \citet{2006A&A...456...23B}. First, we removed the obvious interlopers by excluding all galaxies with peculiar velocities in excess of $\mathrm{\pm 4000\ km\ s^{-1}}$ from the robust cluster redshift of $0.3048 \pm 0.0044$ (calculated using the \citet{1990AJ....100...32B} biweight estimators for robust mean and scale). Peculiar velocities were corrected for cosmological redshift and velocity errors and set to rest-frame using the standard recipe of \citet{1980A&A....82..322D}. To the remaining galaxies we applied first the weighted gap selection method described by \citet{1993ApJ...404...38G} and then the phase-space rejection criterion of \citet[cf. also \citealt{2004ApJ...600..657K}]{1996MNRAS.279..349D} to identify less evident interlopers. In all our calculations we assumed as cluster centre the peak of the X-ray surface brightness map, which lies within $8\arcsec$ from the BCG. Our analysis yields a total of 230 dynamically bound galaxies within a projected cluster-centric distance of $\mathrm{4.5\ h_{70}^{-1}\ Mpc}$, with a robust rest-frame velocity dispersion of $\mathrm{987 \pm 101\ km\ s^{-1}}$ (the errors on velocity dispersion and later on mass are derived via jackknife from the catalogue of confirmed cluster members). Fig.~\ref{members} shows the distribution of cluster members in phase-space and their velocity distribution. The spectroscopic redshifts enabled us to train the photometric redshift solutions obtained from Le PHARE (PHotometric Analysis for Redshift Estimations, S. Arnouts \& O. Ilbert), a publicly available\footnote{http://www.cfht.hawaii.edu/~arnouts/LEPHARE/cfht\_lephare/ lephare.html} software based on the $\mathrm{\chi^2}$ template-fitting procedure. The photometric data in three optical bands (B, V, R) allowed us to trace the Balmer break of galaxies as a function of redshift up to $\mathrm{z \sim 1.5}$. We adopt the included Virgo cluster template set \citep{2003A&A...402...37B}, as this yields the highest quality photometric redshifts. This was defined as the smallest achievable fraction of catastrophic failures $\mathrm{\eta=|z_p -z_s|/(1+z_s)>0.15}$ (where $\mathrm{z_p}$ and $\mathrm{z_s}$ are the photometric and spectroscopic redshifts, respectively) combined with the best possible accuracy $\mathrm{\sigma_{\Delta z/(1+z_s)}}$ measured with the normalised median absolute deviation $\mathrm{1.48 \times median(|\Delta z|/(1+z))}$ (as done in \citealt{2006A&A...457..841I}), where $\mathrm{\Delta z = z_p-z_s}$. The value of these parameters for all galaxies (without any selection in magnitude) and for bright galaxies ($\mathrm{R<R^\star +1}$) were respectively: $\mathrm{\eta = 0.21}$, $\mathrm{\sigma_{\Delta z/(1+z_s)} = 0.07}$, $\mathrm{\eta_{Bright} = 0.07}$ and $\mathrm{\sigma_{Bright,\Delta z/(1+z_s)} = 0.03}$. We removed the Blue Compact Dwarf template as this was found to increase the scatter and the number of catastrophic failures. We use the option AUTO\_ADAPT of Le PHARE to correct our zero-points based on a sub-sample of bright galaxies and then we apply the result to the whole catalogue. We chose the template set and checked whether to apply extinction correction on several templates using modified \citet{2000ApJ...533..682C} attenuation law and several others with the colour excess E(B-V) values ranging between 0 and 0.3 and with a step of 0.05. Eventually, we did not use any extinction correction as we obtained our best result using the provided templates. \citet{2009ApJ...690.1236I} implemented an improved method to compute photometric redshifts taking into account the emission line contribution using relations between the UV continuum and the emission line fluxes associated with star-formation activity (like $\mathrm{[OII],\ [OIII],\ H_\beta, \ H_\alpha}$). The authors compared the template curves with and without emission lines and found that the expected line fluxes can change up to 0.4 mag in the colour. Therefore, we decided to add the emission line contributions to the SED templates after verifying an improvement in the comparison between the spectroscopic and photometric redshifts, also according to the parameters $\eta$ and $\mathrm{\sigma_{\Delta z/(1+z_s)}}$. \begin{figure} \includegraphics[width=\hsize]{zphot_zspec_comp.png} \caption{Comparison of spectroscopic and photometric redshifts. Red dots represent all galaxies brighter than $\mathrm{R^\star +1}$. The continuous line is for $\mathrm{z_p=z_s}$, dashed and dotted lines are for $\mathrm{z_p=z_s\pm0.05(1+z_s)}$ and $\mathrm{z_p=z_s\pm0.15(1+z_s)}$, respectively.} \label{z_comparison} \end{figure} In Fig.~\ref{z_comparison} we show the comparison between $\mathrm{z_s}$ and $\mathrm{z_p}$. The continuous line represents $\mathrm{z_p=z_s}$, the two sets of dashed and dotted lines are for $z_p=z_s\pm0.05(1+z_s)$ and $z_p=z_s\pm0.15(1+z_s)$, respectively. The discrepancies in the comparison between $\mathrm{z_s}$ and $\mathrm{z_p}$ is due partly to the small number (i.e. 3) of bands used for the fit and partly to the fraction of catastrophic failures provoked by the misinterpretation of some features. \cite{2007A&A...470..425B} applied their photometric redshift analysis to a simulated catalogue in order to check the robustness of their $\mathrm{z_p}$ against the contamination of high-z outliers. They found that wrong identifications mainly lay outside the cluster photometric range ( $\mathrm{ z\sim 0.3}$). Thus, after a visual inspection of the outliers in our catalogue at $\mathrm{z \sim 0.3}$ we noticed that the best fit SED templates were typical of late type galaxies, for which the identification of the 4000~$\mathrm{ \AA}$ break is more problematic. Consistently, all these galaxies were classified as faint objects in the VIMOS program (section \ref{spectro_data_descr}) and their spectra revealed some line emissions. Photometric cluster membership was defined from the distribution of photometric redshifts of spectroscopically confirmed members: we used the biweight mean \citep{1990AJ....100...32B} of their photo-z in order to define the centre of the cluster and a scatter of $\mathrm{\pm 1 \sigma=0.06}$. The redshift interval of the selected cluster members is [0.23,0.35] and the mean $\mathrm{z_{phot}}$ is at $\mathrm{z=0.29}$. Choosing this interval we had a contamination (considering the whole field of the cluster) of 42\% (of which 26\% in foreground and 16\% in background) and we lost 35\% of the spectroscopic sources (of which 18\% had lower photometric redshifts and 17\% higher ones with respect to the spectroscopic values). For the bright sources (marked in Fig.~\ref{z_comparison} as red dots) we had a contamination of 35\% (of which 25\% in foreground and 10\% in background) and we measured losses of 18\% of the spectroscopic sources (of which 14\% had lower photometric redshifts and 4\% higher ones with respect to the spectroscopic values). Spectroscopic redshifts were used instead of photometric ones, when available. Based on these values, we select a total of 4462 photometric cluster members, 2108 of which are classified as red galaxies and the remaining 2354 as blue, based on their position in the CMD with respect to the red sequence fit given by equation \ref{CMD_equation}. \subsection{X-ray imaging} \begin{figure*} \includegraphics[width=0.50\hsize]{rxcj1131_vertical_central_xray_new.jpg} \includegraphics[width=0.495\hsize]{x_ray_sb_plus_contours_new1.jpg} \caption{Left panel: BVR image of the central region of A1300 with an overlaid composite X-ray emission contour map. Right panel: composite X-ray emission map with contours. The upper left inset shows the Brightest Cluster Galaxy (BCG) for which different cuts and different scales were used for the three bands, in order to highlight the extended halo and the two nuclei separated by about $1.2\arcsec$. On the right: XMM-Newton image of A1300 in the 0.5-2 keV and smoothed with a Gaussian of $\mathrm{\sigma=4\arcsec}$. Z06 report a suspicious X-ray point source at the position (11:31:54.6, -19:55:43), which corresponds to the position of the BCG}. \label{BCG_fig} \end{figure*} The ROSAT X-ray emission from A1300 was first investigated by \citet{1997A&A...326...34L} who noticed a displacement between the soft (i.e. 0.1-0.4 keV) and hard (i.e. 0.4-2.4 keV) X-ray emission maxima. A1300 was observed also by XMM-Newton (Fig.~\ref{BCG_fig}, right panel) in AO-1 as part of the REFLEX-DXL cluster sample in July 2001 (Z06). The total exposure time was $\mathrm{8.8\ ks}$ for EPN and about 14 ks for each of the EMOS detectors. The data were subjected at this stage to a solar flare cleaning process and the observation was found to be quite clean from contamination so that almost all exposure time survived this process. In fact, after data cleaning from solar flare events we had 8.8 ks for the EPN and 12 ks for the EMOS detectors. The EPN data were corrected for the out-of-time events in the usual way. For a more detailed description of the XMM-Newton data reduction see Z06. The most important global cluster properties resulting from the XMM-Newton data analysis are the bolometric X-ray luminosity $L=\mathrm{1.80 (\pm 0.15) \times 10^{45}\ erg\ s^{-1} }$, the ICM temperature $T=\mathrm{9.2\pm 0.4\ keV}$ and the mass $\mathrm{M_{500}=5.2 (\pm 3.0) \times 10^{14} M_{\sun}}$ (Z06). The right panel of Fig.~\ref{BCG_fig} shows an XMM-Newton image of A1300 in the 0.5-2 keV band, smoothed with a Gaussian of $\mathrm{\sigma=4\ \arcsec}$. Z06 report a suspicious X-ray point source at the position (11:31:54.6, -19:55:43), which corresponds to the position of the BCG. The cluster exhibits an elliptical morphology according to the classification of the dynamical state based on X-ray imaging \citep{1992csg..conf...49J}. The X-ray peak is displaced with respect to the BCG by 36 kpc. The left panel of Fig.~\ref{BCG_fig} shows the same X-ray contours of the right panel superimposed on the BVR image of the central part of A1300 whose BCG presents 2 nuclei at its centre (upper left corner of Fig.~\ref{BCG_fig}). In order to obtain the maximal information from our X-ray data we performed the analysis using the method of the PSF reconstruction as explained in \citet{2009ApJ...704..564F}. In particular, after the point source removal, we computed the background estimate. This procedure allowed us to highlight not only the dynamical shape of the cluster, but also to determine the significance of the structures already seen in an appropriately smoothed X-ray image, as better explained in Section \ref{substructure_id_paragraph} and Appendix \ref{substructures_optical}. This enables us to reconstruct the dynamical history of the cluster through the analysis of the groups and their link with the large scale structure of A1300. \section{Results and Morphological Overview} \subsection{Cluster morphology and large-scale galaxy distribution} \label{density_maps_comp_section} The red sequence fit (Fig.~\ref{CMD} and Equation~\ref{CMD_equation}) allows us to separate red and blue galaxies, for which we derived the galaxy density distribution using an adaptive kernel smoothing. This method is a refinement of the basic procedure discussed in B09 and provides a more accurate estimate of the background counts and noise, while retaining all significant information about substructure and large-scale structure, also in the outskirts of the cluster (i.e. beyond $\mathrm{R}_{200}$). \begin{figure} \includegraphics[width=\hsize]{density_map_blue_red_bkg_r200.png} \caption{Density contours for red and blue photometric members of A1300 smoothed with an adaptive kernel. Red galaxies (in red) are more concentrated in the central regions, while blue galaxies (in blue) are preferentially located in filamentary structures. The green star shows the peak of the X-ray emission, the region of the cluster within $\mathrm{R_{200}}$ is marked by the green circle, while the three green dashed-lined boxes highlight the regions selected for estimating of the background. The contours have a significance of at least $5 \sigma$ w.r.t. the background and follow a square root scale.} \label{density_maps} \end{figure} Red galaxies (red contours in Fig.~\ref{density_maps}) are concentrated towards the innermost region of the cluster with a mean galaxy density of 1.57 galaxies $\mathrm{arcmin^{-2}}$), which is highly significant w.r.t. the background. Conversely blue galaxies (in blue) are more scattered and mostly located in extended structures beyond $\mathrm{R_{200}}$. The over-density information from the galaxy distribution can be used to estimate the amount of galaxy mass belonging to the cluster and compare it to that in the filamentary regions in the NE and SE. Using the X-ray peak as a reference centre and $\mathrm{R_{200}}$ as the fiducial radius of the cluster, we find that the cumulative luminosity of all red galaxies in the NE (SE) filament is about 30\% (33\%) of the total. Such a high fraction of red (i.e. likely passively-evolving) galaxies beyond $\mathrm{R_{200}}$ suggests that the infalling galaxies may have evolved already along the large-scale structure before falling into A1300. This might have happened in the lower-mass groups through the same filaments. \begin{figure} \includegraphics[width=\hsize]{density_map_contour_plot_multi_2_new.png} \caption{Density contours for bright ($\mathrm{R< R^{\star} +1}$, top panel) and faint ($\mathrm{R\geq R^{\star} +1}$,bottom panel) photometric members (both red and blue galaxies) of A1300 using an adaptive kernel method. The green star represents the peak of the X-ray emission while the region of the cluster within $\mathrm{R_{200}}$ is marked by the green circle. The contours have a significance of at least $5 \sigma$ w.r.t. the background (corresponding to 0.41 galaxies ${\rm arcmin^{-2}}$ for the bright cluster members and 3.32 galaxies ${\rm arcmin^{-2}}$ for the faint ones) and follow a square root scale.} \label{density_map_bright} \end{figure} Further comparing the density maps for all bright ($\mathrm{R< R^{\star} +1}$) and for faint ($\mathrm{R\geq R^{\star} +1}$) galaxies (both red and blue ones, Fig.~\ref{density_map_bright}), two different behaviours can be identified. On one hand, bright galaxies are found at larger ratio close to the X-ray peak (as confirmed by the density profiles in Fig.~\ref{density_profiles}), with some clumps following the overall direction of the filaments, while on the other hand, faint galaxies (present in larger numbers) show more over-dense clumps all across the cluster and its outskirts. Hence it seems that, while massive galaxies trace better the inner region of the cluster and the surrounding large-scale structure, faint galaxies dominate in the outskirts (at $\rm R \gtrsim 1.3\ R_{200}$) and can provide insight to the substructure at the smaller scales. \begin{figure} \includegraphics[width=\hsize]{density_profiles.png} \caption{ Density profile for bright (red dashed line) and faint (blue solid line) photometric members (both red and blue galaxies) of A1300. The density profile of the bright galaxies is normalised to the same total number of faint galaxies in order to compare it with that of the faint galaxies. Bright objects are mostly located in the central region of the cluster while faint ones dominate in the outskirts.} \label{density_profiles} \end{figure} \subsection{Temperature, Pressure and Entropy} \label{temp_entr_section} The X-ray properties for the DXL sample were first investigated in \citet{2004A&A...413...49Z}. A1300 exhibits a temperature gradient and cool intra-cluster gas in the centre, suggesting that cooling cores are not only found in clusters with symmetric and regular X-ray images, but can also be found in elongated, very disturbed clusters (such as A1300). An estimate of the cooling time was given in Z06: $\mathrm{t_{cool}\approx10\ Gyr}$, assuming a gas temperature $\mathrm{T=10^8\ K}$ and a density $\mathrm{n_p= 5.8 \times 10^{-3}\ cm^{-3}}$. The region where the cooling time is smaller than the age of the Universe at the cluster redshift was found to be within a cluster-centric distance of $27\arcsec$ ($\sim$120 kpc) at maximum. \begin{figure*} \includegraphics[width=0.48\hsize]{entropy_red_gals_new_new.jpg} \includegraphics[width=0.48\hsize]{temperature_map_red_gals.jpg} \caption{ On the left: Entropy pseudo-map of A1300. The colour bar is in arbitrary units. On the right: Temperature map of A1300. The colour bar is in keV (For details on both pseudo-maps see \citealt{2005A&A...442..827F}). In both panels red contours outline the over-density of red galaxies while the yellow circle represents the region of the cluster within $\mathrm{R_{200}}$. } \label{entropy_map} \end{figure*} A more qualitative analysis of the temperature, pressure and entropy\footnote{The entropy is an important diagnostic parameter because it determines the structure of the ICM recording its thermodynamic history. We adopt the definition of \citet{2005RvMP...77..207V} for the entropy: $\mathrm{K= k_b T n_e^{-2/3}}$, where $\mathrm{k_b}$ is the Boltzmann constant, T is the temperature in keV units and $\mathrm{n_e}$ is the electron density.} maps was performed by \citet[we address the reader to their Fig.~11]{2005A&A...442..827F} who named A1300 the ``Whirlpool'' cluster of galaxies because of the features in its temperature map. \citet{2005A&A...442..827F} found a central East-West ridge of high temperature that may reflect the compression of the central region between the two main merging components. A distorted cool core, which partially preserves the characteristic low entropy, could be responsible for the complex temperature structure of the cluster. Moreover, the same authors found a large scatter in the entropy profile arguing that it reflects a high degree of substructure in the cluster. Therefore, we compare the spatial distribution of galaxies and hot gas in order to investigate consistent behaviours of both components. In particular, the entropy map ( left panel of Fig.~\ref{entropy_map}) provides a reliable record of the gas history \citep{2005RvMP...77..207V}. The distribution of the member galaxies (red in particular) follows closely the entropy features, suggesting that galaxies track the information provided by the gas in the central regions out to larger cluster-centric radii. A similar behaviour can be seen in the temperature map (right panel of Fig.~\ref{entropy_map}) where the hot features are probably caused by shocks resulted from the collisions of different merging components. This comparison reveals a direct correspondence between the substructures traced by the gas and the galaxies. After the impact of two massive clusters it is likely that the coupled gas and dark matter components of each cluster started to swing around a common centre yielding the characteristic whirlpool shape seen in Fig.~\ref{entropy_map}. The projection effects play an important role as, to our knowledge, there is no other cluster with a spiral like shape of the galaxy density distribution which resembles the gas temperature or entropy map. The merging process could also have affected the star formation activity of the galaxies. The interaction with the ICM could have provoked the quenching of the star formation, possibly, after a starburst (e.g. \citealt{2004ApJ...601..197P}). Taking this topic is far from the aim of this paper as further investigation will rely on the ongoing spectral index analysis (Ziparo et al. in preparation). \subsection{X-ray Surface Brightness and Identification of Substructures} \label{substructure_id_paragraph} \begin{figure} \includegraphics[width=\hsize]{dens_map_new_groups_psf_rec.png} \caption{Surface brightness in the $\mathrm{0.5-2\ keV}$ band of A1300 using the technique described in \citet{2009ApJ...704..564F}. Several extended sources are detected around the cluster, with a significance larger than 4$\sigma$ w.r.t. the background. Superimposed are the density contours of red and blue cluster members. The yellow circles represent $\mathrm{R_{200}}$ of each group identified with the same ID of Table \ref{groups_table} (number 5 is A1300). The continuous white circle shows how the X-ray emission of the cluster would look like if it were spherically symmetric and the dashed white circle a possible group that disturbs the original symmetry but is now dissolved inside the cluster. The units of the colour bar are in counts/sec/pixel. } \label{psf_rec_dens_maps} \end{figure} In order to minimize the impact of point sources and isolate the X-ray emission due to the diffuse hot ICM, we apply a novel technique to enhance the significance of extended sources and filter out the point sources from the X-ray surface brightness map \citep{2009ApJ...704..564F}. Although this method was originally designed for group identification in wide-field/survey areas, it proved to be also quite efficient in identifying and characterising smaller subsystems within or close to clusters. In particular, where point sources were detected in the XMM-{\it Newton} maps (and confirmed by archival {\it Chandra} data), this technique enabled us to reduce their X-ray appearance and better remove their contribution from the extended emission (for further details and explanations see \citealt{2009ApJ...704..564F}). Figure \ref{psf_rec_dens_maps} shows the X-ray surface brightness in the $\mathrm{0.5-2\ keV}$ band obtained using the wavelet+PSF reconstruction. It is possible to identify several extended sources in the proximity of the cluster (with a significance larger than 4$\sigma$ w.r.t. the background in the optical). However the X-ray emission itself is not enough to establish their membership to the A1300 system. \begin{figure} \includegraphics[width=\hsize]{groups_appr_rec_new_col_fil.jpg} \caption{WFI R band image of A1300. The ellipses represent the groups observed in this field: the shape of all groups is related to their X-ray emission (Fig.~\ref{psf_rec_dens_maps}). Black ellipse identify groups with a high probability to belong to the cluster i.e. at a significance larger than 5$\sigma$ w.r.t. the background, the green ones have a lower probability. Group number 5 is the main cluster of which region inside $\mathrm{R_{200}}$ is represented by the black dashed circle. The small symbols are the spectroscopic member galaxies of the cluster: the blue squares are all the galaxies approaching with respect to the observer, the red diamonds are the receding ones. Group 4 and 10 correspond to the substructures found in the DS-test (red circles in Fig.~\ref{dstest}).} \label{groups_appr_rec} \end{figure} Comparison with the distribution of cluster members shows that a few extended X-ray sources match the position of over-densities in the galaxy distribution maps. The results of this combined analysis and identification of the groups in X-rays and optical are described extensively in Appendix \ref{substructures_optical}. All groups detected and possibly related to the cluster are marked in Fig.~\ref{groups_appr_rec}: in black are those belonging to the LSS of A1300 (detected at more than 5$\sigma$ w.r.t. the background), whereas those in green have a lower probability to be at the same redshift of the cluster. The groups selected as related to the cluster (ID numbers 4, 6 and 10) are also highlighted by yellow circles in Fig.~\ref{psf_rec_dens_maps}. They all lie in the outskirts of the cluster, thus providing a tool to probe the accretion regions and to investigate the large-scale dynamics of A1300. The central region of the cluster (marked by the solid white circle in Fig.~\ref{psf_rec_dens_maps}) is spherically symmetric. This region has a radius of 3.09 arcmin (corresponding to about $\mathrm{835\ kpc}$ at the cluster redshift) and emphasizes a possible forward shock towards SW, followed closely by the galaxy density distribution. This shock is possibly related to the candidate radio relic found in previous studies (\citealt{1999MNRAS.302..571R} and \citealt{2011arXiv1102.1901G}). \begin{figure} \includegraphics[width=\hsize]{shock.png} \caption{0.5-2 keV surface brightness profile of A1300 extracted from the centre of the curvature of the shock towards the SW quadrant, where the candidate radio relic was found (\citealt{1999MNRAS.302..571R}, \citealt{2011arXiv1102.1901G}) at the position (11:31:46.8, -19:56:44). The errors are 1$\sigma$, the red continuous line shows the best fit model while the dashed line shows the same model in the region of the shock without the enhancement. The green dashed lines constrain the position of the relic which corresponds to the enhancement we find in the profile. } \label{shock} \end{figure} Fig.~\ref{shock} shows the 0.5-2 keV surface density radial profile extracted in the SW quadrant in the direction of the radio relic (the position of which is highlighted by the green dashed lines). We used the position of the relic to tentatively identify the curvature of the shock (as already done, for example, by \citealt{2011ApJ...728...82M}). We find an enhancement in the surface brightness coincident with the position of the relic. We use only pn data in which this enhancement is more evident after masking all point sources and extended emissions outside the cluster. Our best fit model is obtained with a projected emissivity profile in which the density jump in the position of the relic is ${\rm \rho_1/\rho_0 = 1.30 \pm 0.15}$, where $\rho_0$ is the density of the unperturbed gas and $\rho_1$ is the density of the gas after the shock. The density jump allows us to derive an upper limit to the strength of this possible shock. Assuming a monoatomic gas with ${\rm \gamma=5/3}$ and using equation 1 of \cite{2010ApJ...715.1143F} we obtain a Mach number $\mathcal{M}=1.20 \pm 0.10$. \cite{2005A&A...442..827F} define the cluster merging direction with the North-South on the basis of their X-ray qualitative analysis. Using X-ray and optical information (thus through a more detailed study) we are able to better define this direction. However the study of the dynamics of this cluster is not straightforward. In fact, the X-ray surface brightness map shows a more complex morphology in the outer regions of the cluster, with multiple extended structures around and beyond $\mathrm{R_{200}}$ and a globally asymmetric shape. This matches the distribution of photometric cluster members (cf. Fig.~\ref{psf_rec_dens_maps}), suggesting that A1300 is accreting matter along filamentary structures. In particular, the extended structure seen to the SE of the main cluster in the distribution of galaxies matches the position of Group 10, suggesting that this system may be infalling along a filament which extends from SE to NW. In addition, we highlight with a dashed white circle in Fig.~\ref{psf_rec_dens_maps} a possible group which is now part of the cluster: its identity is lost in X-rays but it is visible in the optical image as a concentration of galaxies and in Fig.~\ref{psf_rec_dens_maps} as a density peak. To the NE of the core, the symmetry is disturbed by group 4 (shown in Fig.~\ref{group_snapshots_and_CMD} with overlaid X-ray contours) which lies mostly inside the $\mathrm{R_{200}}$ of the cluster. Even if this group is very close to the cluster, it preserved its identity both in the X-ray and in the optical: its X-ray emission stands out even if it is already embedded with that of the cluster (Fig.~\ref{psf_rec_dens_maps}). It is also possible to trace a red sequence (Fig.~\ref{group_snapshots_and_CMD}) and to identify a BCG in the corresponding region in the optical image. The X-ray mass of this group is $M_{200}=\mathrm{1.35\times10^{14}\ M_{\sun}}$ (obtained adopting the same scaling relations of \citealt{2010ApJ...709...97L}, i.e. after assuming a beta profile and removing embedded point sources, see Appendix~\ref{substructures_optical} for details), i.e. 10\% of that of the cluster. We are thus able to reconstruct the accretion pattern of A1300, which looks to be entering a phase of dynamical relaxation in its inner region, while still accreting mass in its outskirts. \subsubsection{Substructures from kinematics} We apply the kinematical DS-test \citep{1988AJ.....95..985D} to the dataset of available spectroscopic cluster members, to identify substructures based on the combination of their position and velocity. This provides a complementary approach to detecting over-densities in the galaxy distribution and has the potential to identify substructures bound to the cluster but still retaining their kinematical identity. \begin{figure} \includegraphics[width=\hsize]{RXCJ1131_NEW_DStest.png} \caption{Substructure skyplot showing the results of the DS-test for the spectroscopic members of A1300. Circles are centred on the positions of member galaxies; their radius is proportional to $e^{P_{DS}}$, where $P_{DS}$ is the DS deviation parameter for each galaxy. Red circles mark groups of galaxies which exhibit significant deviations ($>3\sigma$) from the local velocity distribution.} \label{dstest} \end{figure} The test was iterated over $10^4$ re-samplings of the velocities and yields a DS statistics $P_{DS} = 4\times10^{-4}$, implying that the cluster has a significant degree of sub-structure (the whole test being significant at $5.5\sigma$). The $\kappa$-test of \citet{1996ApJ...458..435C} was also run for comparison on the same dataset with consistent results. The results of the DS-test are shown in Fig.~\ref{dstest}. We identify kinematical substructures as groups of galaxies with values of the DS parameter beyond a critical value of 2.58 (calculated from the re-sampling statistics). Two groups, both with a significance above the $3\sigma$ level, were detected. A first group to the NE of the cluster centre consists of 12 cluster members with a mean peculiar velocity of +1494$\,\rm{km~s^{-1}}$ w.r.t the mean cluster redshift and a velocity dispersion of 582$\,\rm{km~s^{-1}}$. This group appears quite compact on the plane of the sky, and its mean position (11:32:09.4, -19:50:43.8) is consistent with that found in X-rays for group 4. All 12 galaxies lie within the $\rm{R_{200}}$ of this group. After correcting for velocity errors, we derive a tentative mass estimate from the velocity dispersion (using the relation given by \citealt{2006A&A...456...23B}), finding a mass of $\mathrm{(1.17 \pm 0.24)\times10^{14}\ M_{\sun}}$, in good agreement with the X-ray estimate (cf. Appendix \ref{substructures_optical}). Another group of 7 galaxies to the SE of the cluster centre exhibits large systematic deviations. These galaxies appear less concentrated than those identifying the previous group, however they all lie within $R_{200}$ of group 10, their mean position (11:33:02.8, -20:12:09.33) being only 53$\arcsec$ away from the X-ray peak (Table \ref{groups_table} and ellipse number 10 in Fig.~\ref{groups_appr_rec}). We find a mean velocity of -774$\,\rm{km~s^{-1}}$ w.r.t the cluster mean velocity and a dispersion of 452$\,\rm{km~s^{-1}}$, from which we infer a mass of $\mathrm{(4.37 \pm 2.99)\times10^{13}\ M_{\sun}}$. Although with larger uncertainty, also for this group we find an agreement with the X-ray value. \subsection{Linking merging configuration and substructures} The combination of the X-ray, photometric and spectroscopic data enables us to investigate the dynamics of the cluster and to link them to its main substructures. Figure \ref{groups_appr_rec} shows all spectroscopic cluster members grouped by their peculiar velocity w.r.t the cluster mean velocity: receding from the observer in red, approaching in blue. Ellipses mark the X-ray groups detected with the PSF reconstruction technique, groups encircled in black being those with a high probability to belong to the cluster. The cluster $\mathrm{R_{200}}$, marked with the black dashed circle, highlights the presence of three groups (ID 4, 6, 10) entering the virialised region of the cluster. The dynamical configuration of A1300 is quite complex, as already anticipated in the previous sections. X-ray and entropy maps reveal signs of a major merger and the presence of three X-ray groups. Two of these (ID 4 and 10) find confirmation from both the distribution of galaxies and the kinematical substructures and they seem to be accreted onto the main cluster, while a third group (ID 6) may be following a third accretion direction in a filament pointing towards the observer (unfortunately we do not have spectroscopic information to confirm it). Northwards of the main cluster, Group 4 shows a significantly higher recession velocity by +1494$\,\rm{km~s^{-1}}$. Our findings suggest that this group may be part of the northern filament; from its peculiar velocity we can infer that the filament lies between the observer and the cluster. This is mirrored by the global velocity distribution of galaxies to the North of the cluster, which shows a systematic velocity of +456$\,\rm{km~s^{-1}}$ for galaxies outside the core ($\sim$1 Mpc away from the cluster centre). Towards the South, a filamentary structure is traced by photometric spectroscopic cluster members. Group 10 is detected in the same region. It is particularly significant in X-rays and in the optical as well, showing a rich red sequence (see Fig.~\ref{group_snapshots_and_CMD}) and a well-defined over-density of galaxies (see Tab.~\ref{groups_table}). It is also detected from kinematics, with a peculiar velocity of -774$\,\rm{km~s^{-1}}$. Consistently, the overall velocity field in the southern outskirts is negative, with a mean of -354$\,\rm{km~s^{-1}}$. This suggests that the filament in which Group 10 is embedded reaches the cluster from behind. \begin{figure} \includegraphics[width=\hsize]{histo_vel_fil.png} \caption{Velocity distribution for galaxies further than 1 Mpc from the cluster centre. Red histogram: northern outskirts; blue histogram: southern outskirts.} \label{histo_vel_north_south} \end{figure} The velocity distributions in the two regions (1 Mpc northwards of the cluster core and 1 Mpc southwards of it) are shown in Fig.~\ref{histo_vel_north_south}. We compare the two distributions by means of a two-sided Kolmogorov-Smirnov test: this shows that the two distributions are significantly different, yielding a KS parameter of 0.018 (i.e. a significance of $3.3\sigma$). We can estimate the expected infall velocity around $\mathrm{R_{200}}$ for a group infalling from a large distance. Assuming a mass of $10^{15} \mathrm{M_{\sun}}$ for the main cluster as calculated for system 5 (cf. Appendix A), we obtain a velocity of about $\mathrm{2000\ km/s}$ at a distance of ${\rm 2\ Mpc}$ from the cluster. This allows us to estimate for each filament the angle w.r.t the plane of the sky: we find that the northern filament has an angle of about 40$^{\circ}$, while the southern has a lower angle of about 20$^{\circ}$. In addition, given the estimates of total luminosity for the two filaments (cf. Section \ref{density_maps_comp_section}), we estimate that A1300 will accrete about 60\% of its current total mass the next Gyr from infalling material. \section{Discussion} A1300 is a post-merging cluster at $\mathrm{z\sim 0.3}$ with a ``dumbbell'' cD galaxy at its centre and prominent filaments visible in the galaxy density distribution. The BCG position is not coincident with the position of the X-ray emission peak, on the other hand, this cluster exhibits an elongated X-ray emission and a disturbed galaxy distribution in the optical. As already found by B09 in other two DXL clusters (and in the past in other clusters e.g. by \citealt{1980ApJ...236..351D}), also A1300 hosts a centrally located population of galaxies dominated by an old, passively evolving stellar populations which define the red sequence. The study of the CMD enables us to measure the magnitude gap between the BCG and the second brightest galaxy. We can tentatively use it as a proxy on the dynamical state of the clusters (as already done in greater detail for a sample of galaxy clusters by \citealt{2010MNRAS.409..169S}). Table \ref{mag_gap_table} provides a comparison of the gap statistics for the REFLEX-DXL clusters studied so far. RXCJ0014.3-3022, which is in an early phase of merging, displays the smallest magnitude gap. A1300 (RXCJ1131.9-1955), classified as a post-merging system \citep{1997A&A...326...34L}, has an intermediate value of $\mathrm{\Delta m}$, while RXCJ2308.3-0211, which is a very regular cluster with a cool core, has the largest gap. This supports the conclusions by \citet{2010MNRAS.409..169S} i.e. the time passed since the last major merger is directly reflected in the growth of the BCG; in this way the magnitude gap increases with the ageing of the cluster. \begin{table} \centering \resizebox{\hsize}{!}{ \begin{tabular}[!t]{ccccc} \hline RXCJ & a & b & $\mathrm{\Delta mag}$ & Dynamical State \\ \hline 0014.3 \textendash 3022 & $-0.037 \pm 0.003 $ & $2.935 \pm 0.238$ & $\sim 0.1$ & Merging \\ 1131.9 \textendash 1955 & $-0.048 \pm 0.005 $ & $2.823 \pm 0.090$ & $\sim 0.8$ & Post-merging \\ 2308.3 \textendash 0211 & $-0.038 \pm 0.003 $ & $2.933 \pm 0.323$ & $\sim 1.1$ & Cool core \\ \hline \end{tabular} } \caption{Optical properties for 3 DXL clusters: col.1 gives the name of the cluster as in the REFLEX catalogue, col.2 and col.3 the fit parameters of the best red sequence fit (where the fit is represented by the equation $\mathrm{y=ax+b}$), col.4 the gap between the two brightest galaxies and col.5 the dynamical state.} \label{mag_gap_table} \end{table} Table \ref{mag_gap_table} also provides information on the parameters of the best fitting line to the red sequence for the 3 DXL clusters. The red sequence fit in the CMD (Fig.~\ref{CMD}) allows us also to separate red and blue galaxies, for which we derive the galaxy density distribution (Fig.~\ref{density_maps}). The density contours of red galaxies reveal a vortex-like shape reminiscent of the temperature and the entropy features (Fig.~\ref{entropy_map}). The filamentary structures are highlighted mostly by the red galaxies departing from the central region of the cluster and extending beyond $\mathrm{R_{200}}$. This suggests that infalling galaxies have already evolved before starting their infall towards A1300, possibly in former groups embedded in the surrounding feeding filaments. In fact, the gravitational potential well of the groups could affect the evolution of their galaxies \citep{2004ogci.conf..426L}. Furthermore, the extent to which galaxies are pre-processed in groups before falling into clusters depends on the mass of galaxies \citep{2009MNRAS.400..937M}, suggesting that massive galaxies are more easily segregated in the group halos. A possible shock front (consistent with a Mach number $\mathcal{M}=1.20 \pm 0.10$) was also identified in the southern part of the cluster coincident with a candidate relic (\citealt{1999MNRAS.302..571R} and \citealt{2011arXiv1102.1901G}) found in the radio bands. Radio halos are diffuse radio emission located at the centre of clusters, while relics are usually elongated or arc-shaped and found in the peripheral regions of clusters (e.g. \citealt{2011arXiv1102.1572V} for a recent review). Most of the merging clusters host both radio halo and relics \citep{2002ASSL..272..197G}, likely originated by acceleration of electrons through shocks and turbulence \citep{2011arXiv1106.0591B}. The Mach Number that we find from the surface brightness profile is consistent with a shock originating from a merging of two progenitors with comparable mass \citep{2007PhR...443....1M}. The gas motion after the merging could have played an important role though. Surprisingly also the galaxy distribution exhibits a sharp edge in the SW. This coupled conduct of gas and galaxies is also visible in the analysis of the temperature and entropy maps. In particular, we overlaid the density contours of the red galaxies of the cluster (mostly populating the inner region) which nicely follow the entropy (and temperature) features, suggesting that the galaxies track the information provided by the gas in the central region out to larger cluster-centric radii. The projection effects combined with the status of the merging play an important role as, to our knowledge, there is no other cluster with a spiral like shape and the galaxy density distribution which reminds the gas temperature or entropy map. \citet{2011ApJ...728...54Z} performed simulations of cluster merging starting with two cool core clusters with different masses and total mass distributions (gas and dark matter) represented by a Navarro, Frenk and White (NFW, \citealt{1997ApJ...490..493N}) profile. The status of A1300 seems to be represented by one of the simulations where two parent clusters of approximately the same mass ($\mathrm{M_{200}\sim 6 \times 10^{14}\ M_{\odot}}$) are approaching one another with an initial impact parameter $\mathrm{b\approx 464\ kpc}$ (fig.~4 for simulation S2 in \citealt{2011ApJ...728...54Z}). However, we can not exclude the possibility represented by the same simulation with an impact parameter $\mathrm{b\approx 932\ kpc}$ (fig.~5 for simulation S3 in \citealt{2011ApJ...728...54Z}). In these simulation the gas cores do not collide at the first core passage but sideswipe, creating a bridge of stripped, low entropy gas that stretches between the two DM cores. In fact, as high entropy gas floats and low-entropy gas sinks, the low entropy region marks the positions of the core of the cluster progenitors and the high entropy is due to the presence of shocks. The entropy map of the simulation finds a pretty good correspondence to the projected entropy map of A1300: here one plume of cold gas is located in correspondence of the BCG, showing up as a low entropy channel in the central region of the cluster, surrounded by higher entropy gas (with a characteristic whirlpool shape). The absence of a symmetric plume could be due to the fact that just one of the two cluster progenitors was a cool core (whose central part still survive at the centre of A1300) similarly massive. The other entropy plume is located to the North and is not compact as the previous one. This could be an indication that A1300 is a complex system in which minor mergers played an important role in the mass assembly and the major merger happened between clusters of different dynamical states. The comparison with these aforementioned simulations of \citet{2011ApJ...728...54Z}, even if the matching is not perfect, allows us to approximately date the merging of the cluster progenitors: about 3 $(\pm 1)$ Gyr could have passed since when their regions within $\mathrm{R_{200}}$ touched for the first time. Tentatively tracing the previous cores in the entropy map, we can estimate the current projected impact angle which is likely to be $\mathrm{\sim 130\ deg}$ in the plane of the sky. This suggests that we are witnessing an almost face-on merging of clusters which allow us to better understand the status of the system related to the projection effects: as in the simulations, the two clusters collide with a relatively low initial impact parameter and after the first core passage the low entropy regions trace the two progenitors. Moreover, the presence of the shock found in the same position of the radio relic suggests that more than 2 Gyr passed since the merging; its weakness ($\mathcal{M}=1.20 \pm 0.10$) could constrain the age of the merging to around 3 Gyr. Furthermore, the degree of gas mixing as a function of radius (fig.~13 for simulation S2 or S3 in \citealt{2011ApJ...728...54Z}) is very efficient in the off-axis merging, because the cores feel their mutual interaction and are stripped before the final merging. The disturbed shape of the X-ray entropy, temperature and surface brightness 2-D distributions reflect a high degree of substructure in the cluster, fundamental for the interpretation of different moments of the assembly history. The surface brightness, in fact, is asymmetric (Fig.~\ref{psf_rec_dens_maps}) and exhibits some excess of emission due to the influence of the filamentary structure through which groups fall. There is a mutual perturbation among the cluster and the infalling groups which keep their identities while they experience the attraction of the cluster, as appears evident from the analysis of the red and blue galaxy fractions as a function of position and luminosity (Appendix A). The substructures found in A1300 show several episodes of accretion: a filament shows up in the DS-test (Fig.~\ref{dstest}) and in the galaxy distribution (Fig.~\ref{density_maps}) we detect groups entering into the virial region of the cluster (Fig.~\ref{psf_rec_dens_maps}) and over-densities in the inner part of the cluster without an X-ray identification. The last case refers to some peaks that we observe inside $\mathrm{R_{200}}$ by the density map. In fact, when a group enters in a cluster, it is stripped almost immediately of its hot gas \citep{2006PASP..118..517B}. Conversely, group galaxies can keep orbiting around the cluster centre for some time even after being captured in the deep potential well. Thus, they may retain part of their common DM halo remaining effectively bound together for longer times and appearing as galaxy over-densities. This happens mainly because of the difference in the behaviour of collisional and non-collisional components of clusters or groups (galaxies and ICM respectively) as explained for example by \citet{2002ASSL..272....1S}. Moreover, we compared the velocity distributions of the northern and southern components (both at a distance larger than 1 Mpc from the core) of the cluster. The NE filament (in which the Group 4 is embedded, Fig.~\ref{group_snapshots_and_CMD}, corresponding to the northern red circles of the DS test, Fig.~\ref{dstest}) is receding w.r.t. the observer while the southern one (in which is embedded the Group 10, Fig.~\ref{group_snapshots_and_CMD}, corresponding to the southern red circles of the DS test, Fig.~\ref{dstest}) appears approaching to us. In this way we are able to give a three dimensional picture of the mass assembly history using all data available. The substructure analysis revealed also that the total mass of our X-ray selected groups is around 20\% of that of the cluster. One of them (Group 4, $\mathrm{M_{200}\approx10^{14}\ M_{\sun}}$) is already embedded in the virial region of the cluster (also in terms of gas, as it is possible to see from the X-ray emission in Fig.~\ref{psf_rec_dens_maps}) and contributes alone to half of this value. This confirms the relation suggested by a recent simulation \citep{2011arXiv1105.1397C}, i.e. that richness of the largest subgroup is typically 20\% of that of its host cluster (even if this relation exhibits a wide scatter). Fig.~\ref{BCG_fig} shows the elongated shape of the galaxy distribution in the central region of the cluster, compared with the X-ray emission: the brightest galaxies, together with their satellites, trace the disturbed X-ray emission. A bit further from the central emission of the gas we detect groups both in the X-rays and in the optical bands (as galaxy concentration). The X-ray surface brightness (better investigated in the PSF reconstructed image, Fig.~\ref{psf_rec_dens_maps}) appears elliptical at the centre but a projected filament to the South suggests the direction of a past accretion event. The X-ray surface brightness appears elongated towards the direction of the filaments in which the infalling groups are embedded, thus the ongoing mergers will modify again the shape of the cluster as it happens in the North-East. Combining all the available information, A1300 could have been the result of the merging of two clusters of similar mass. The collision happened most likely with a small impact parameter (as inferred from the entropy and the temperature maps) and this shock heated the gas, yielding all the features visible in X-rays. A1300 still accretes via filaments and groups, which suggests that this cluster is still in the formation process (cf. \citealt{1997MNRAS.291..353B}). Indeed, it could be experiencing a relaxation at the observed time, just after the merging, in which the gas starts settling down and the different components mix. According to our estimations the cluster will increase its total mass about 60\% through filaments (including groups) in the next Gyr with a medium velocity of $\mathrm{\sim2000\ km\ s^{-1}}$. \section{Conclusions} We investigated the recent assembly history of the REFLEX-DXL post-merging cluster A1300 through the X-ray and optical (spectroscopy and photometry) data. Our main results are summarized as follows: \begin{itemize} \item[$\bullet$] The galaxy distribution of red sequence galaxies reveals a filamentary structure departing from the inner area of the cluster which extends beyond $\mathrm{R_{200}}$. This suggests that galaxies could be pre-processed in the groups that we identified embedded in the filaments surrounding A1300. \item[$\bullet$] The distribution of the red galaxies of the cluster (mostly concentrated at the centre) exhibits a marked correspondence with the entropy (and temperature) features, suggesting that the galaxies track the information provided by the gas in the central region out to larger cluster-centric radii. The projection effects combined with the status of the merging play an important role in this configuration. \item[$\bullet$] The X-ray surface brightness distribution is clearly affected by the presence of filamentary structures through which groups enter the virial radius of the cluster. \item[$\bullet$] A possible forward shock (consistent with a Mach number $\mathcal{M}=1.20 \pm 0.10$) was identified in the southern part of the cluster coincident with a candidate relic (\citealt{1999MNRAS.302..571R} and \citealt{2011arXiv1102.1901G}) found in the radio bands. Surprisingly also the galaxy distribution presents a sharp edge in the SW. \item[$\bullet$] A comparison with simulations suggests that 3~($\pm 1$)~Gyr elapsed since the two cluster progenitors (still visible in the entropy map) started merging with a likely projected impact parameter between 462 and 932 kpc. \end{itemize} A1300 could have been disturbed by another cluster of similar mass. The collision happened most likely with a small impact parameter and this shock heated the gas, yielding all the features visible in X-rays. The cluster is now likely to experience a relaxation, just after the merging, in which the gas starts to settle down and the different components mix. The brightest galaxies of A1300 are coincident with the X-ray peaks in the central region (Fig.~\ref{BCG_fig}) or X-ray signatures of substructure (Fig.~\ref{psf_rec_dens_maps}): these massive galaxies could be reminiscent of former Brightest Group Galaxy (BGG) that have merged to form the main cluster. It is thus possible to study different stages of a merging even by analysing a single cluster: A1300 is a good example for this purpose as it shows a filament through which it accretes not only individual galaxies but also groups; disrupted ones visible in the optical but not any more in X-rays and other groups still in the process of crossing $\mathrm{R_{200}}$. \section*{Acknowledgments} We would like to thank the anonymous referee for the constructive comments, which led to a significant improvement of the quality of this paper. FZ acknowledges the support from and participation in the International Max-Planck Research School on Astrophysics at the Ludwig-Maximilians University. DP acknowledges the kind hospitality at the Max-Planck-Institut f\"{u}r extraterrestrische Physik (MPE). FZ thanks Olivier Ilbert, Mara Salvato and Francesco Pace for useful discussions.
\section*{Appendix \thesection\protect\indent\quad #1} } \renewcommand{\theequation}{\thesection.\arabic{equation}} \def\hybrid{\topmargin 0pt \oddsidemargin 0pt \headheight 0pt \headsep 0pt \textwidth 6.5in \textheight 9.0in \marginparwidth 0.0in \parskip 5pt plus 1pt \jot = 1.5ex} \catcode`\@=11 \def\marginnote#1{} \newcount\hour \newcount\minute \newtoks\amorpm \hour=\time\divide\hour by60 \minute=\time{\multiply\hour by60 \global\advance\minute by-\hour} \edef\standardtime{{\ifnum\hour<12 \global\amorpm={am}% \else\global\amorpm={pm}\advance\hour by-12 \fi \ifnum\hour=0 \hour=12 \fi \number\hour:\ifnum\minute<10 0\fi\number\minute\the\amorpm}} \edef\militarytime{\number\hour:\ifnum\minute<100\fi\number\minute} \newcommand{\textcolor{yellow}}{\textcolor{yellow}} \newcommand{\textcolor{red}}{\textcolor{red}} \newcommand{\textcolor{blue}}{\textcolor{blue}} \newcommand{\textcolor{green}}{\textcolor{green}} \newcommand{\textcolor{white}}{\textcolor{white}} \newcommand{\textcolor{magenta}}{\textcolor{magenta}} \let\wht=\widehat \let\wtd=\widetilde \def\draftlabel#1{{\@bsphack\if@filesw {\let\thepage\relax \xdef\@gtempa{\write\@auxout{\string \newlabel{#1}{{\@currentlabel}{\thepage}}}}}\@gtempa \if@nobreak \ifvmode\nobreak\fi\fi\fi\@esphack} \gdef\@eqnlabel{#1}} \def\@eqnlabel{} \def\@vacuum{} \def\draftmarginnote#1{\marginpar{\raggedright\scriptsize\tt#1}} \def\draft{ \oddsidemargin -.5truein % \def\@oddfoot{\footnotesize \sl preliminary draft \hfil \rm\thepage\hfil\sl\today\quad\militarytime} % \let\@evenfoot\@oddfoot \overfullrule 3pt \let\label=\draftlabel \let\marginnote=\draftmarginnote \def\@eqnnum{(\theequation)\rlap{\kern\marginparsep\tt\@eqnlabel}% \global\let\@eqnlabel\@vacuum} } \newcommand{\,{\rm Tr}\,}{\,{\rm Tr}\,} \newcommand{1\!\!1}{1\!\!1} \def\partial{\partial} \def\varepsilon{\varepsilon} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\<{\langle} \def\>{\rangle} \def\frac{\rm d}{{\rm d}z}{\frac{\rm d}{{\rm d}z}} \def\frac{\rm d}{{\rm d}\lambda}{\frac{\rm d}{{\rm d}\lambda}} \def\nonumber{\nonumber} \def{\mathbb H}{{\mathbb H}} \def{\mathbb R}{{\mathbb R}} \def{\mathbb P}{{\mathbb P}} \def{\mathbb Z}{{\mathbb Z}} \def{\mathcal F}{{\mathcal F}} \let\wtd=\widetilde \def{\rm Tr}{{\rm Tr}} \def\one#1{#1^{\raise5pt\hbox{$\scriptstyle\!\!\!\!1$}}\,{}} \def\two#1{#1^{\raise5pt\hbox{$\scriptstyle\!\!\!\!2$}}\,{}} \def\onetwo#1{#1^{\raise5pt\hbox{$\scriptstyle\!\!\!\!\!{12}$}}\,{}} \def\mathop{\otimes}{\mathop{\otimes}} \def\vartheta{\vartheta} \defe{e} \def{\mathbf x}{{\mathbf x}} \def{\mathbf p}{{\mathbf p}} \def{\mathcal A}{{\mathcal A}} \def{\mathbb Q}{{\mathbb Q}} \newtheorem{theorem}{Theorem}[section] \newtheorem{lm}[theorem]{Lemma} \newtheorem{prop}[theorem]{Proposition} \newtheorem{corollary}[theorem]{Corollary} \theoremstyle{definition} \newtheorem{df}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{remark}[theorem]{Remark} \newtheorem{claim}[theorem]{Claim} \newtheorem{xca}[theorem]{Exercise} \newtheorem{cor}[theorem]{Corollary} \theoremstyle{remark} \newtheorem{hypothesis}[theorem]{Hypothesis} \newtheorem{conjecture}[theorem]{Conjecture} \allowdisplaybreaks \begin{document} \title[Teichm\"uller spaces of Riemann surfaces with orbifold points] {Teichm\"uller spaces of Riemann surfaces with orbifold points of arbitrary order and cluster variables} \author{Leonid Chekhov$^{\ast,\dag}$}\thanks{$^{\ast}$Steklov Mathematical Institute and Laboratoire Poncelet, Moscow, Russia}\thanks{$^{\dag}$Concordia University, Montr\'eal, Canada. Email: [email protected].} \author{Michael Shapiro$^\diamondsuit$}\thanks{$^\diamondsuit$Mathematical Department, Michigan State University, East Lansing, USA} \maketitle \begin{abstract} We generalize a new class of cluster type mutations for which exchange transformations are given by reciprocal polynomials. In the case of second-order polynomials of the form $x+2\cos{\pi/n_o}+x^{-1}$ these transformations are related to triangulations of Riemann surfaces of arbitrary genus with at least one hole/puncture and with an arbitrary number of orbifold points of arbitrary integer orders $n_o$. We propose the dual graph description of the corresponding Teichm\"uller spaces, construct the Poisson algebra of the Teichm\"uller space coordinates, propose the combinatorial description of the corresponding geodesic functions and find the mapping class group transformations. \end{abstract} \section{Introduction} Since their appearance, cluster variables~\cite{FZ} find applications in geometry. An important example of the cluster variables in the case where exchange polynomial is of the second-order is provided by $\lambda$-lengths~\cite{ThSh},~\cite{Penn1} of curves that make ideal triangle partitions of fundamental domains of Riemann surfaces with punctures. These systems were generalized in~\cite{Fock1},~\cite{Fock2} to the case of Riemann surfaces with holes. At the same time, a combinatorial description of geodesic functions in terms of the dual variables, the shear coordinates, as well as their quantization, was developed in~\cite{ChF}. Amazingly enough, transition from punctures to holes does not effectively change the corresponding cluster algebra. Generalizations of Teichm\"uller spaces of Riemann surfaces to the case of bordered Riemann surfaces~\cite{KaufPen} or ciliated Riemann surfaces~\cite{FG} were constructed. The corresponding cluster algebras were developed in~\cite{GSV1,FG1,FST}, whereas the geometrical pattern underlying the bordered Riemann surfaces was identified with that of Riemann surfaces with ${\mathbb Z}_2$-orbifold points in~\cite{Ch1},~\cite{Ch1a}, where the corresponding mutations (flips) in terms of the shear coordinates were constructed. These flips preserve the sets of geodesic functions; on the level of cluster variables the corresponding transformations were considered in~\cite{FST-Orb}; the mutations of cluster variables were there again given by the standard two-term relations. In \cite{ChM}, the description of Teichm\"uller spaces of Riemann surfaces with holes and with orbifold points of order two and three was given. It was, however, an ambiguity in identifying a boundary component with an orbifold point; in the present paper, we demonstrate that this ambiguity is due to choice of the type of an orbifold point corresponding to the given boundary component; we therefore provide the combinatorial description of Riemann surfaces with holes and with orbifold points of arbitrary orders. We show that mutations for orbifold points of order greater than two are given by three-, not two-term transformations determined by a second-order reciprocal polynomial. We prove the Laurent phenomenon and positivity property for these transformations. The positive coefficients of Laurent polynomials however become not necessarily integers in the presence of orbifold points of order greater than three. On the shear-coordinate side, we define the complete set of real-valued coordinates, construct all the geodesic functions for such surfaces, all the mapping-class-group transformations, and prove the regularity condition, that is, that all elements of the corresponding Fuchsian group are hyperbolic or parabolic ones except precisely elements conjugate to passing around orbifold points. We therefore have a regular (up to exactly the indicated orbifold points) Riemann surface with holes for any choice of the introduced real coordinates, and vice versa; these coordinates parameterize therefore the corresponding Teichm\"uller spaces of Riemann surfaces with holes and with orbifold points. The mutations of orbifold triangulations are particular cases of more general construction. In \cite{GSV} some generalization of cluster transformations were described that preserve Poisson bracket and have some universal properties. Until recently no applications of these transformations were found. We were happy to see manifestation of our formulas in mutations of orbifold triangulations. Motivated by that we propose a new algebraic construction of generalized cluster algebras with mutations given by reciprocal polynomials. Using the tools of the standard cluster algebra \cite{FZ1}, \cite{FZ2}, we prove that the Laurent phenomenon holds true in this case as well. For algebras of order greater than two, we do not know whether the positivity property holds in general; it however holds in all tested examples, so we formulate it as a conjecture. We also prove that generalized cluster algebras of finite type satisfy the same Cartan--Killing classification as the standard cluster algebras. \vskip 2mm \noindent{\bf Acknowledgements.} The authors are grateful to Anna Felikson, Pavel Tumarkin, Sergey Fomin, and Dylan Thurston for many enlighting conversations and, specially, to Alek Vainshtein for valuable comments improving our paper. The work of L.Ch. was supported in part by the Ministry of Education and Science of the Russian Federation (contract 02.740.11.0608), by the Russian Foundation for Basic Research (Grant Nos. 10-01-92104-JF$\_$a, 11-01-00440-a, 09-02-93105-CNRS-a, and 11-01-12037-ofi-m-2011), by the Grant of Supporting Leading Scientific Schools of the Russian Federation NSh-8265.2010.1, and by the Program Mathematical Methods for Nonlinear Dynamics. Michael Shapiro was supported in part by grants DMS-0800671 and DMS-1101369. \section{Generalized cluster algebra}\label{s-algebra} \noindent We briefly remind the definition of cluster algebra. An integer $n\times n$ matrix $B$ is called \emph{skew-symmetrizable} if there exists an integer diagonal $n\times n$ matrix $D=diag(d_1,\dots,d_n)$, such that the product $BD$ is a skew-symmetric matrix, i.e., $b_{ij}d_j=-b_{ji}d_i$. { \color{black} Let ${\mathbb P}$ be \emph{a semifield } equipped with commutative multiplication $\cdot$ and addition $\oplus$. We assume that the multiplicative group of ${\mathbb P}$ is a free abelian group. ${\mathbb P}$ is \emph{a coefficient group} of cluster algebra. ${\mathbb Z}{\mathbb P}$ is the integer group ring, ${\mathcal F}$ is a field of rational functions in $n$ independent variables with coefficients in the field of fractions of ${\mathbb Z}{\mathbb P}$. ${\mathcal F}$ is called an ambient field. \begin{df} \emph{A seed} is a triple $({\mathbf x},{\mathbf p},B)$, where \begin{itemize} \item ${\mathbf p}=(p_{x}^\pm)_{x\in{\mathbf x}}$, a $2n$-tuple of elements of ${\mathbb P}$ is a \emph{coefficient tuple} of cluster ${\mathbf x}$; \item ${\mathbf x}=\{x_1,\dots,x_n\}$ is a collection of algebraically independent rational functions of $n$ variables which generates ${\mathcal F}$ over the field of fractions of ${\mathbb Z}{\mathbb P}$; \item $B$ is a skew-symmetrizable \emph{exchange matrix}. \end{itemize} The part ${\mathbf x}$ of seed $({\mathbf x},{\mathbf p},B)$ is called \emph{cluster}, elements $x_i\in{\mathbf x}$ are called \emph{cluster variables}, and $B$ is called \emph{exchange matrix}. \end{df} \begin{df}[seed mutation] For any $k$, $1\le k\le n$ we define \emph{the mutation} of seed $({\mathbf x},{\mathbf p},B)$ in direction $k$ as a new seed $({\mathbf x}',{\mathbf p}',B')$ in the following way: \begin{equation}\label{eq:MatrixMutation} b'_{ij}=\left\{ \begin{array}{ll} -b_{ij}, & \hbox{ if } i=k \hbox{ or } j=k; \\ b_{ij}+\frac{|b_{ik}|b_{kj}+b_{ik}|b_{kj}|}{2}, & \hbox{ otherwise.} \end{array} \right. \end{equation} \begin{equation}\label{eq:ClusterMutation} x'_i=\left\{ \begin{array}{ll} x_i, & \hbox{ if } i\ne k; \\ \frac{p^+_k\prod_{b_{jk}>0} x_j^{b_{jk}}+p^-_k\prod_{b_{ji}<0} x_j^{-b_{ji}}}{x_k}, & \hbox{ otherwise.} \end{array} \right. \end{equation} \begin{eqnarray}\label{eq:CoeffMutation} p'^\pm_k &=& p^\mp_k \\ \hbox{ for } i\ne k\qquad p'^+_i/p'^-_i &=& \left\{ \begin{array}{ll} (p^+_k)^{b_{ki}}p^+_i/p^-_i, & \hbox{ if } b_{ki}\ge 0; \\ (p^-_k)^{b_{ki}}p^+_i/p^-_i, & \hbox{ if } b_{ki}\le 0; \\ \end{array} \right. \end{eqnarray} \end{df} \noindent We write $({\mathbf x}',{\mathbf p}',B')=\mu_k\left(({\mathbf x},{\mathbf p},B)\right)$. Notice that $\mu_k(\mu_k(({\mathbf x},{\mathbf p},B)))=({\mathbf x},{\mathbf p},B)$. We say that two seeds are \emph{mutation-equivalent} if one is obtained from the other by a sequence of seed mutations. Similarly we say that two clusters or two exchange matrices are \emph{mutation-equivalent}. For any skew-symmetrizable matrix $B$ we define \emph{initial seed} $$({\mathbf x},{\mathbf p},\!B)=(\!\{x_1,\dots,x_n\}\!,\!\{p_1^\pm,\ldots,p_n^\pm\}\!,\!B),$$ where $B$ is the \emph{initial exchange matrix}, ${\mathbf x}=\{x_1,\dots,x_n\}$ is the \emph{initial cluster}, ${\mathbf p}=\{p_1^\pm,\ldots,p_n^\pm\}$ is the \emph{initial coefficient tuple}. } {\it Cluster algebra} ${\mathcal A}(B)$ associated with the skew-sym\-met\-ri\-zab\-le $n\times n$ matrix $B$ is a subalgebra of ${\mathbb Q}(x_1,\dots,x_n)$ generated by all cluster variables of the clusters mutation-equivalent to the initial seed $(x,B)$. Cluster algebra ${\mathcal A}(B)$ is called \emph{of finite type} if it contains only finitely many cluster variables. In other words, all clusters mutation-equivalent to initial cluster contain totally only finitely many distinct cluster variables. Two most important properties of cluster algebra are Laurent phenomenon~\cite{FZ1} and finite type classification~\cite{FZ2}. More exactly, Lurent phenomenon states that any cluster variable is expressed as a Laurent polynomial in terms of the initial cluster. The remarkable finite type classification claims that cluster algebras of finite type are in one-to-one correspondence with the Dynkin diagrams of finite type. \subsection{Generalized cluster transformations} Now we introduce more general cluster transformations. Assume that $B$ is a skew-symmetrizable integer matrix such that all elements in its $k$th row are divisible by $d_k$. Define $\beta_{kj}=b_{kj}/d_k$. \begin{lm} Let $B'=\mu_l(B)$ is obtained from $B$ by mutation in direction $l$. Then, all entries $b'_{kj}$ of $k$th row of $B'$ are divisible by $d_k$. \end{lm} \begin{proof} The statement follows immediately from matrix mutation~\ref{eq:MatrixMutation}. \end{proof} Let us fix $d_k$ for all $k$ from $1$ to $n$ and we assume that all elements $b_{kj}$ of $k$th row of integer skew-symmetrizable matrix $B$ are divisible by $d_k$. For a collection ${\mathbf p}_i=(p_{i;0},\ldots,i_{i;d_i})$ we define the polynomial $\theta_i[{\mathbf p}_i](u,v)=\sum_{\ell=0}^{d_i} p_{i;\ell} u^\ell v^{d_i-\ell}$ be a polynomial of degree $d_i$. The corresponding inhomogeneous polynomial we denote by $\rho_i[{\mathbf p}_i]=\rho_i[{\mathbf p}_i](t)=\sum_{\ell=0}^{d_i} p_{i;\ell} t^\ell$. Note that, $\theta_i[{\mathbf p}_i](u,v)=u^{d_i}\rho_i[{\mathbf p}_i](v/u)$. We define a generalized seed $q$ of a generalized cluster algebra as a triple\newline $q=({\mathbf x}(q),\overline{{\mathbf p}(q)},B(q))$, where ${\mathbf x}=(x_1(q),\ldots,x_n(q))$ is a $n$-tuple of cluster variables in seed $q$, $\overline{{\mathbf p}(q)}=({\mathbf p}_1(q),\ldots,{\mathbf p}_n(q))$ is $n$-tuple of coefficient collections ${\mathbf p}_i(q)$, ${\mathbf p}_i(q)=(p_{i;0}(q),\ldots, p_{i;d_i}(q))$ is a $d_i+1$-tuple of coefficients of $\theta_i[q]$, and, finally, $B(q)$ is an exchange $n\times n$ matrix. Generalized cluster mutations are described by the following formulas: Exchange matrix is mutated in direction $k$ by the rule ~\ref{eq:MatrixMutation}. Introduce $u_{j;>0}=\prod_{\beta_{j,\ell}>0} x_\ell^{\beta_\ell}$, $u_{j;<0}=\prod_{\beta_{j,\ell}<0} x_\ell^{-\beta_\ell}$, Mutation of cluster variables is given by the rule $(\{x_i'\},\{{\mathbf p}_i'\},B')=\mu_k(\{x_i\},\{{\mathbf p}_i\},B)$: \begin{equation}\label{eq:GenClusterMutation} x'_i=\left\{ \begin{array}{ll} x_i, & \hbox{ if } i\ne k; \\ \frac{\theta_k(u_{k;>0},u_{k;<0})}{x_k}, & \hbox{ otherwise.} \end{array} \right. \end{equation} Coefficients mutate by the following generalized rule: \begin{eqnarray}\label{eq:GenCoeffMutation} p'_{k;\ell} &=& p_{k;d_k-\ell} \\ \hbox{ for } i\ne k\qquad p'_{i;j}/p'_{i;r} &=& \left\{ \begin{array}{ll} (p_{k;0})^{b_{ki}}p_{i;r}/p_{i;j}, & \hbox{ if } b_{ki}\ge 0; \\ (p_{k;d_k})^{b_{ki}}p_{i;r}/p_{i;j}, & \hbox{ if } b_{ki}\le 0; \\ \end{array} \right. \end{eqnarray} \begin{remark} Note that if we assume that coefficients of all $\theta_i$ do not change under mutation then the corresponding inhomogeneous polynomial $\rho_i$ is reciprocal of degree $d_i$, i.e., $t^{d_i}\rho(1/t)=\rho(t)$. \end{remark} \begin{theorem}\label{thm:generalizedLaurent} (Laurent property) Any generalized cluster variable is a Laurent polynomial in initial cluster variables $x_i$. \end{theorem} \begin{proof} The proof uses the "caterpillar lemma"~\cite{FZ1}. \begin{lm}\label{lm:caterpillar} Assume that a generalized exchange pattern on $T_{n,m}$ satisfies the following conditions: \begin{enumerate} \item For any edge the polynomial $P$ does not depend on $x_k$ and is not divisible by any $x_i$, $i\in[n]$. \item Each exchange polynomial has nonnegative coefficients \item For any three edges labeled by $i,j,i$\newline \smallskip {\psset{unit=0.7} \begin{pspicture}(-3,0.8)(3,-0.8) \pcline[linewidth=2pt](-3,0)(3,0) \pscircle[linewidth=1pt,fillstyle=solid,fillcolor=blue](-3,0){0.2} \pscircle[linewidth=1pt,fillstyle=solid,fillcolor=blue](-1,0){0.2} \pscircle[linewidth=1pt,fillstyle=solid,fillcolor=blue](1,0){0.2} \pscircle[linewidth=1pt,fillstyle=solid,fillcolor=blue](3,0){0.2} \rput(-2,0.2){\makebox(0,0)[cb]{$i$}} \rput(-2,-0.2){\makebox(0,0)[ct]{$P$}} \rput(0,0.2){\makebox(0,0)[cb]{$j$}} \rput(0,-0.2){\makebox(0,0)[ct]{$Q$}} \rput(2,0.2){\makebox(0,0)[cb]{$i$}} \rput(2,-0.2){\makebox(0,0)[ct]{$R$}} \end{pspicture} } \smallskip\newline we have $L\cdot Q_0^b\cdot P=R|_{x_j\leftarrow\frac{Q_0}{x_j}},$ where $b$ is a negative integer, $Q_0=Q|_{x_i=0}$, and $L$ is a Laurent monomial whose coefficient lies in $\mathbb A$ and is coprime with $P$. \end{enumerate} Then each element $x_i(t)$ for $i\in[n]$, $t\in T_{n,m}$ is a Laurent polynomial in $x_1(t_0),\ldots,x_n(t_0)$ with coefficients in $\mathbb A$. \end{lm} By definition of generalized cluster transformation $P=\theta_i[t],\ Q=\theta_j[t'],\ R=\theta_i[t'']$. Note that parts (1) and (2) are evidently satisfied by generalized cluster mutations. It remains to proof part (3). note that if $x_i$ is not included into any monomial of $Q$ then generalized mutation with labels $i$ and $j$ commute and the latter mutation is inverse to the former, namely, $x_i(t''')=x_i(t)$. Consider the case when $x_i$ enters a monomial of $Q$. For simplicity denote $b_{ji}(t')$ by $a$. By our assumptions $a\ne 0$. Moreover, without loss of generality we can assume that $a>0$, otherwise we replace $B$ by $-B$. Since $Q$ is determined by the homogeneous polynomial of two variables where only one variable contains a positive power of $x_i$ then $x_i$ enters all monomial of $Q$ but one. Hence, $Q_0=Q|_{x_i=0}$ is a monomial. Moreover, $Q_0=p_{j;0}\prod_{b_{jk}(t')>0} x_k^{b_{jk}(t')}$. Note that $P=\theta_i(u_{>0}(t'), u_{<0}(t'))$. By the mutation rule~\ref{eq:ClusterMutation}, \begin{equation*} b_{il}(t'')=\left\{ \begin{array}{ll} b_{il}(t'), & \hbox{ if } b_{jl}(t')\le 0;\\ b_{il}(t')+a b_{jl}(t'), & \hbox{ otherwise.} \end{array} \right. \end{equation*} By the definition of generalized cluster transformation $R=\theta_i(u_{>0}(t''), u_{<0}(t''))$. For $q=t'$ or $t''$ introduce $\tau_q=\prod_\ell x_\ell(q)^{\beta_{i\ell}(q)}=\frac{u_{>0}(q)}{u_{<0}(q)}$. Finally, \begin{multline*} \tau_{t''}|_{x_j\leftarrow \frac{Q_0}{x_j}}=\prod_{l\ne j} x_l^{\beta_{il}(t')}\cdot \left(\frac{Q_0}{x_j}^{-a/d_i}\right)\prod_{\beta_{jl}(t')>0} x_l^{a\beta_{jl}(t')}=\\ =\prod_l x_l^{\beta_{il}(t)}\left(p_{j;0}\prod_{b_{jl}(t')>0} x_l^{b_{jl}(t')}\right)^{-a/d} \left(\prod_{b_{jl}(t')>0} x_l^{a \beta_{jl}(t')}\right)=p_{j;0}^{-a/d_i}\tau_{t'}=\frac{1}{p_{j;0}^{a/d_i}\tau_{t}} \end{multline*} It is enough to notice that \ref{eq:GenClusterMutation} and \ref{eq:GenCoeffMutation} imply that $\rho_{t;i}(\tau_{t})=\rho_{t'';i}\left(\tau_{t''}|_{x_j\leftarrow \frac{Q_0}{x_j}}\right)$. Therefore, $R|_{x_j\leftarrow \frac{Q_0}{x_j}}=P\cdot\hat L$, where $\hat L$ is a Laurent monomial. \end{proof} \begin{theorem}\label{thm:generalizedCartanKilling} Generalized cluster algebras of finite type satisfy the same Cartan-Killing classification as the standard cluster algebras. \end{theorem} \begin{proof} The proof repeats the one of~\cite{FZ2}. The only differences make the proofs of the fact that the only finite type generalized cluster algebras of rank two correspond to $A_2, B_2, G_2$ types. It is checked by direct computation similar to one in \cite{FZ2}. Note first that in $A_2$-case formulas for generalized cluster transformation coincide with formulas for the standard cluster transformation. In $B_2$-case the polynomial degrees of theta-polynomials are two and one. Set the theta polynomials in the initial cluster $\theta_1(u,v)=a u^2+b u v+c v^2$, $\theta_2(u,v)=p u+q v$. Then, we immediately obtain $(x,y)\overset{\mu_1}\longleftrightarrow(\mu_1)(x_1,y)\overset{\mu_2}\longleftrightarrow (x_1,y_1)\overset{\mu_1}\longleftrightarrow (x_2,y_1)\overset{\mu_2}\longleftrightarrow (x_2,y_2)\overset{\mu_1}\longleftrightarrow (x,y_2)\overset{\mu_2}\longleftrightarrow (x,y)$, where $x_1=(a+by+cy^2)/x$, $y_1=(px+qa+bqy+cqy^2)/xy$, $x_2=(a^2q^2+2ap q x+acq^2 y^2+abq^2 y+bpq x y+p^2 x^2)/x y^2$, $y_2=(q a+p x)/y$. Similar computations lead to the 8-cycle in $G_2$-case. Note that degrees of polynomials $\deg(\theta_1)=3$, $\deg(\theta_2)=1$. We set $\theta_1(u,v)=a u^3+b u^2 v+c u v^2+ d v^3$, $\theta_2(u,v)=p u+q v$. $(x,y)\overset{\mu_1}\longleftrightarrow(\mu_1)(x_1,y)\overset{\mu_2}\longleftrightarrow (x_1,y_1)\overset{\mu_1}\longleftrightarrow (x_2,y_1)\overset{\mu_2}\longleftrightarrow (x_2,y_2)\overset{\mu_1}\longleftrightarrow (x_3,y_2)\overset{\mu_2}\longleftrightarrow (x_3,y_3)\overset{\mu_1}\longleftrightarrow (x,y_3)\overset{\mu_2}\longleftrightarrow (x,y)$, where $x_1=(a+b y+c y^2+d y^3)/x$, $y_1=(p x+aq +bq y+cq y^2+dq y^3)/x y$, $x_2=(a^3q^3 +2 a^2 cq^3 y^2+2a^2 d q^3 y^3+3 a^2 p q^2 x+2a^2 b q^3 y+2a b d q^3 y^4 +3 a p^2 q x^2 +4 a b p q^2 x y+a c^2 q^3 y^4+2acd q^3 y^5 +ad^2q^3 y^6+3ac p q^2 x y^2+3ad p q^2 x y^3+ a b^2 q^3 y^2 +2 a b c q^3 y^3 +bc p q^2 x y^3 +bd p q^2 x y^4 + p^3 x^3+b^2 p q^2 x y^2+2 b p^2 q x^2 y+p^2 cq x^2 y^2)/x^2 y^3$, $y_2=(q^2 a^2+abq^2 y+acq^2 y^2+adq^2 y^3+2 a p q x+bp q x y+p^2 x^2)/xy^2 $, $x_3=( a^2 d q^3 y^3+ a^2 c q^3 y^2+acp q^2 x y^2+ a^2 b q^3 y+2 ab p q^2 x y+bp^2 q x^2 y+ a^3 q^3 +3 a^2 p q^2 x+3 a p^2 q x^2 +p^3 x^3)/x y^3$, $y_3=(p x+a q )/y$. \end{proof} We note that according to~\cite{GSV} the generalized cluster transformation preserves presymplectic structure compatible with cluster algebra structure. Similarly, the secondary generalized cluster transformation preserves the compatible Poisson bracket. \begin{theorem}\label{thm:generalizedCompatible}(\cite{GSV}) Poisson structure compatible with a cluster algebra is compatible with the corresponding generalized cluster transformations. \end{theorem} In the next section we describe generalized cluster structure associated with triangulated surfaces with orbifold points. \section{Teichm\"uller space of surfaces with holes and orbifold points of arbitrary order}\label{s:geometry} We now demonstrate how the above mutations with reciprocal polynomials of the second order appear in the description of Teichm\"uller spaces of Riemann surfaces of arbitrary genus with nonzero number of holes (punctures) and with an arbitrary number $r$ of orbifold points of arbitrary orders. \subsection{The ideal triangle decompositions of orbifold Riemann surfaces and cluster variables}\label{ss:clusters} We now present the geometric pattern underlying the algebraic construction of cluster variables corresponding to orbifold Riemann surfaces. Particular cases of Riemann surfaces with orbifold points of order 2 and 3 are discussed in~\cite{Ch1,Ch2}, \and\cite{ChM}. For relation between skew-symmetrizable cluster algebras of finite mutation type and surfaces with orbifold points see \cite{FST-Orb}. We consider a regular Riemann surface $\Sigma_{g,s,r}$ of genus $g$ with $s>0$ holes and with a number $r\ge 0$ of orbifold points; orders of these points $p_i$, $i=1,\dots,r$ are positive integers greater than one. We introduce the marking on the set of orbifold points splitting this set into nonintersecting subsets $\delta_k$ (maybe empty), $k=1,\dots,s$, $\sum_{k=1}^s|\delta_k|=r$. We then assign the subset $\delta_k$ to the $k$th hole for every $k$ and introduce a cyclic ordering inside each subset $\delta_k$. To construct the generalization of the ideal triangle decomposition \cite{Fock1,Penn1}, we first remove from the surface all the hyperbolic domains of holes bounded by the corresponding perimeter geodesic lines (with their closures, which are these perimeter lines). Second, we choose for each orbifold point from the set $\delta_k$ a domain containing this point and bounded by a geodesic curve whose both ends spiral to the $k$th hole as shown in Fig.~\ref{fi:saucer-pan}.\footnote{In the case of a $\mathbb Z_2$-orbifold point, this domain has zero area because the corresponding geodesics goes directly to the orbifold point, reflects at it, and repeats its path in the opposite direction.} We remove from the remaining part of the surface all such domains. The remaining part of $\Sigma_{g,s,r}$ admits splitting into ideal triangles; a copy of this splitting can be drawn as a connected ideal polygon in the Poincar\'e disc; the sides of this polygon are of two sorts: those that are not pre-images of geodesics going around orbifold points must be pairwise identified; to a side that is a pre-image of the geodesics going around ${\mathbb Z}_p$ orbifold point we attach (from outside) an equilateral ideal $p$-gone with the orbifold point situated at its geodesic center; this $p$-gone is the $p$-fold covering of the removed domain enclosing the orbifold point. An example of a fundamental domain of the Riemann surface $\Sigma_{1,1,2}$ of genus one with one hole and with two ${\mathbb Z}_3$ orbifold points is in Fig.~\ref{fi:disc}. \begin{figure}[tb] {\psset{unit=0.5} \begin{pspicture}(-9,1)(9,-7) \rput(-5,-3){ \pscircle[linewidth=0.5pt](0,0){4} \rput(0,0){\psarc[linecolor=green, linewidth=0.5pt](5.657,0){4}{135}{225}} \rput(0,0){\psarc[linecolor=blue, linewidth=0.5pt](4,-1.657){1.657}{90}{225}} \rput{90}(0,0){\psarc[linecolor=green, linewidth=0.5pt](5.657,0){4}{135}{225}} \rput{90}(0,0){\psarc[linecolor=blue, linewidth=0.5pt](4,-1.657){1.657}{90}{225}} \rput{180}(0,0){\psarc[linecolor=green, linewidth=0.5pt](5.657,0){4}{135}{225}} \rput{180}(0,0){\psarc[linecolor=blue, linewidth=0.5pt](4,-1.657){1.657}{90}{225}} \rput{270}(0,0){\psarc[linecolor=green, linewidth=0.5pt](5.657,0){4}{135}{225}} \rput{270}(0,0){\psarc[linecolor=blue, linewidth=0.5pt](4,-1.657){1.657}{90}{225}} \rput(0,0){\makebox(0,0){\large $\star$}} } \rput(0,-3){\pcline[linewidth=1pt]{->}(-0.5,0)(0.5,0)} \rput(5,0){ \psellipse[linecolor=blue, linewidth=0.5pt](0,0)(2,0.3) \psbezier[linewidth=0.5pt](-4,-4)(-2.5,-2.5)(-2,-1.5)(-2,0) \psbezier[linewidth=0.5pt](4,-4)(2.5,-2.5)(2,-1.5)(2,0) \psbezier[linewidth=0.5pt](-4,-4)(-1.5,-5.5)(2.5,-5)(4,-4) \psframe[linecolor=white, fillstyle=solid, fillcolor=white](-1.8,-5.3)(0.3,-4) \psbezier[linestyle=dashed, linewidth=0.5pt](-4,-4)(-1.5,-5.5)(2.5,-5)(4,-4) \psbezier[linewidth=0.5pt](-3,-4.2)(-2.5,-4.3)(-1.5,-4.5)(-1.5,-6.5) \psbezier[linewidth=0.5pt](1.5,-4.6)(1,-4.7)(0.5,-4.7)(-1.5,-6.5) \pscircle[linecolor=white, fillstyle=solid, fillcolor=white](-1.5,-6.5){0.1} \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-0.35,-5.5)(-0.8,-6)(-1.46,-6.2)(-1.63,-5.3) \psbezier[linecolor=green, linewidth=0.5pt](-0.35,-5.5)(0.1,-5)(2.45,-3.2)(2.55,-2.2) \psbezier[linecolor=green, linewidth=0.5pt](-1.63,-5.3)(-1.8,-4.4)(2.35,-2.7)(2.41,-1.9) \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.3,-1.6)(-2,-1.5)(2.4,-1.8)(2.55,-2.2) \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.2,-1.35)(-1.8,-1.25)(2.3,-1.5)(2.41,-1.9) \psbezier[linecolor=green, linewidth=0.5pt](-2.3,-1.6)(-2.2,-1.9)(2,-1.45)(2.1,-1.15) \psbezier[linecolor=green, linewidth=0.5pt](-2.2,-1.35)(-2.1,-1.65)(2,-1.25)(2.08,-.95) \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.03,-.65)(-1.4,-0.5)(1.9,-.85)(2.1,-1.15) \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.03,-.5)(-1.4,-.37)(1.86,-.65)(2.08,-.95) \psbezier[linecolor=green, linewidth=0.5pt](-2.03,-.65)(-1.95,-0.87)(1.3,-.75)(1.6,-.55) \psbezier[linecolor=green, linewidth=0.5pt](-2.01,-.5)(-1.92,-0.78)(1.2,-.67)(1.5,-.45) \rput(-4,-4){\makebox(0,0){ $\bullet$}} \rput(4,-4){\makebox(0,0){ $\bullet$}} \rput(-1.5,-6.5){\makebox(0,0){\large $\star$}} } \end{pspicture} } \caption{\small The right part represents an example of a regular genus zero Riemann surface with a ${\mathbb Z}_4$-orbifold point marked by a star; the bounding geodesic line is the image of a side of an ideal equilateral square in the Poincar\'e disc depicted in the left part. Both ends of the geodesic line spiral asymptotically to the closed geodesic that is the boundary of a hole.} \label{fi:saucer-pan} \end{figure} \begin{figure}[tb] {\psset{unit=0.5} \begin{pspicture}(-10,-5)(8,5) \pscircle[linewidth=0.5pt](0,0){4} \psarc[linecolor=green, linewidth=0.5pt](-4,1.64){1.64}{-90}{45} \psarc[linecolor=green, linewidth=0.5pt](0,5.66){4}{-135}{-45} \psarc[linecolor=green, linewidth=0.5pt](4,1.64){1.64}{135}{270} \psarc[linestyle=dashed,linecolor=green, linewidth=0.5pt](4,-2.31){2.31}{90}{210} \rput{142}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.314,0){1.616}{-68}{68}} \rput{172}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.04,0){.56}{-82}{82}} \rput{146}(0,0){\psarc[linecolor=green, linewidth=0.5pt](-4.45,0){1.95}{-32}{64}} \rput{165}(0,0){\psarc[linecolor=green, linewidth=0.5pt](-4.14,0){1.07}{-75}{-14}} \psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-.67,-4.99){3.06}{30}{135} \rput{57.5}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.09,0){.85}{-78}{78}} \rput{95}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.40,0){1.86}{-65}{65}} \rput{66}(0,0){\psarc[linecolor=green, linewidth=0.5pt](-4.285,0){1.535}{6}{69}} \rput{89}(0,0){\psarc[linecolor=green, linewidth=0.5pt](-4.666,0){2.4}{-59}{25}} \psarc[linecolor=green, linewidth=0.5pt](-4,-1.64){1.64}{-45}{90} \rput(3,-0.65){\makebox(0,0){$\star$}} \pscircle[linestyle=dashed, linecolor=blue, linewidth=0.5pt](3,-.65){0.4} \rput(-1.15,-2.5){\makebox(0,0){$\star$}} \pscircle[linestyle=dashed, linecolor=blue, linewidth=0.5pt](-1.15,-2.5){0.5} \rput(-2.65,0.5){\makebox(0,0)[lt]{$b$}} \rput(-2.7,-0.45){\makebox(0,0)[lb]{$a$}} \rput(2.2,1.3){\makebox(0,0)[rt]{$b$}} \rput(-0.55,1.5){\makebox(0,0)[ct]{$a$}} \end{pspicture} } \caption{\small The Poincar\'e disc with depicted fundamental domain for the genus-one surface $\Sigma_{1,1,2}$ with one hole and two ${\mathbb Z}_3$ orbifold points marked by $\star$). Solid lines constitute the boundary of a fundamental domain and dashed lines are sides of the related ideal triangles. The sides with labels $a$ and $b$ are pairwise identified. Inscribed circles with centers at the ${\mathbb Z}_3$ orbifold points all have the radius $\frac12\log3$.} \label{fi:disc} \end{figure} \begin{figure}[tb] {\psset{unit=0.5} \begin{pspicture}(6,-5)(0,5) \pscircle[linewidth=0.5pt](0,0){4} \newcommand{\PATTERN}{% \psarc[linecolor=green, linewidth=1pt](4.62,0){2.3}{120}{240} \rput{10}(0,0){\rput(2.3,0){\makebox(0,0)[rc]{$c$}}} \rput{-20}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.06,0){.7}{100}{260}} \rput{10}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.26,0){1.45}{110}{250}} \rput{-20}(0,0){\rput(3.9,0){\makebox(0,0)[rc]{$a$}}} \rput{10}(0,0){\rput(3.8,0){\makebox(0,0)[rc]{$b$}}} } \rput{30}(0,0){\PATTERN} \rput{90}(0,0){\PATTERN} \rput{150}(0,0){\PATTERN} \rput{210}(0,0){\PATTERN} \rput{270}(0,0){\PATTERN} \rput{330}(0,0){\PATTERN} \rput{60}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \rput{-60}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \pcline[linestyle=dashed,linecolor=green, linewidth=1pt](-4,0)(4,0) \pcline[linewidth=2pt]{->}(6,0)(7,0) \rput(0,0){\makebox(0,0){$\star$}} \rput(-0.6,0.1){\makebox(0,0)[cb]{$c_3$}} \rput(1,-0.9){\makebox(0,0)[ct]{$c_2$}} \rput(.4,1.2){\makebox(0,0)[cb]{$c_2$}} \end{pspicture} \begin{pspicture}(-6,-5)(0,5) \pscircle[linewidth=0.5pt](0,0){4} \newcommand{\PATTERN}{% \rput{20}(0,0){\psarc[linecolor=green, linewidth=1pt](4.62,0){2.3}{120}{240}} \rput{30}(0,0){\rput(2.2,-0.3){\makebox(0,0)[rc]{$c'$}}} \rput{-20}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.06,0){.7}{100}{260}} \rput{10}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.26,0){1.45}{110}{250}} \rput{-20}(0,0){\rput(3.9,0){\makebox(0,0)[rc]{$a$}}} \rput{10}(0,0){\rput(3.8,0){\makebox(0,0)[rc]{$b$}}} } \rput{30}(0,0){\PATTERN} \rput{90}(0,0){\PATTERN} \rput{150}(0,0){\PATTERN} \rput{210}(0,0){\PATTERN} \rput{270}(0,0){\PATTERN} \rput{330}(0,0){\PATTERN} \rput{80}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \rput{-40}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \rput{20}(0,0){ \pcline[linestyle=dashed,linecolor=green, linewidth=1pt](-4,0)(4,0) } \rput(0,0){\makebox(0,0){$\star$}} \rput(-0.6,0.1){\makebox(0,0)[cb]{$c'_3$}} \rput(0.3,-0.2){\makebox(0,0)[ct]{$c'_2$}} \rput(.4,1.2){\makebox(0,0)[cb]{$c'_2$}} \end{pspicture} } \caption{\small The mutation in the ideal $p$-gone (here $p=6$), which is the $p$-fold covering of the domain around a $\mathbb Z_p$-orbifold point (marked by $\star$). We perform standard $2$-term mutations (\ref{diagonal}) on the set of cluster variables to come from the pattern in the left-hand side to the one in the right-hand side.} \label{fi:p-gone} \end{figure} We use the standard geometric correspondence between cluster variables and $\lambda$-lengths: at each point at the absolute that is a vertex of an ideal triangle we set an horocycle; the $\lambda$-length $\ell_a$ is then the geodesic length of the part of side $a$ enclosed between two horocycles based at its endpoints. The correspondence reads: \begin{equation} \label{lambda-l} a=e^{\ell_a/2}, \end{equation} where $\ell_a$ is understood as the signed distance between horocycles: $\ell_a$ is negative when the corresponding horocycles overlap. Here $a$ is the cluster variable associated with the edge. For an ideal quadrangle with the (cyclically enumerated) sides $a_i$, $i=1,\dots,4$ and with diagonals $d$ and $d'$ we have the standard two-term cluster relation \begin{equation} \label{diagonal} dd'=a_1a_3+a_2a_4. \end{equation} We now consider mutations for cluster variables of the ideal $p$-gone corresponding to a $\mathbb Z_p$-orbifold point. We consider the pattern in the left-hand side of Fig.~\ref{fi:p-gone} and perform a sequence of mutations (\ref{diagonal}) to come to the pattern in the right-hand side. For a equilateral $p$-gone, the cluster variables $c_k$ for $k$-diagonal are \begin{equation} \label{ck} c_k=c\frac{\sin(\pi k/p)}{\sin(\pi/p)},\qquad (c_1= c). \end{equation} The easy combinatorics then yields \begin{lm}\label{cc-prime} $\lambda$-lengths satisfy the following relation $$ cc'=a^2+2\cos(\pi/p)ab+b^2. $$ \end{lm} \begin{remark} Relation in Lemma~\ref{cc-prime} is a generalized cluster mutation described in Sec.~\ref{s-algebra}. Note that since we use only two-term transformations \ref{diagonal} to prove Lemma~\ref{cc-prime}, the positivity property for generalized mutations of such form follows from the one for the standard mutations. \end{remark} To simplify the description, it is convenient to introduce the notion of {\em petal surface}. Petals are the domains containing orbifold points that were removed on the second step of constructing the ideal triangulation. To the $k$th hole, we associate the bouquet of $|\delta_k|$ petals (with no petals if $\delta_k$ is empty), each petal carries, besides its cluster variable, the number $\omega_p=2\cos(\pi/p)$. The mutation then occurs inside the corresponding ideal triangle (painted by a light color), and we have three cases depending on whether the adjacent sides are petals themselves: \begin{center} {\psset{unit=1.2} \begin{pspicture}(-1.5,-2.5)(1.5,2.5) \psarc[linewidth=1pt,fillstyle=solid,fillcolor=yellow](0,.923){1.155}{30}{150} \psarc[linewidth=1pt,fillstyle=solid,fillcolor=yellow](0,2.077){1.155}{210}{330} \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](-1,1.5)(0,2.5)(0,0.5)(-1,1.5) \pscircle*(-1,1.5){0.05} \pscircle*(1,1.5){0.05} \pcline[linewidth=1pt]{->}(0,0.3)(0,-0.3) \psarc[linewidth=1pt,fillstyle=solid,fillcolor=yellow](0,-.923){1.155}{210}{330} \psarc[linewidth=1pt,fillstyle=solid,fillcolor=yellow](0,-2.077){1.155}{30}{150} \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](1,-1.5)(0,-2.5)(0,-0.5)(1,-1.5) \pscircle*(-1,-1.5){0.05} \pscircle*(1,-1.5){0.05} \rput(-0.5,1.5){\makebox(0,0)[cc]{$\omega_p$}} \rput(-0.1,1.5){\makebox(0,0)[cc]{$c$}} \rput(0,2.2){\makebox(0,0)[cb]{$a$}} \rput(0,0.8){\makebox(0,0)[ct]{$b$}} \rput(-1.1,1.5){\makebox(0,0)[rc]{$v_1$}} \rput(1.1,1.5){\makebox(0,0)[lc]{$v_2$}} \rput(0.5,-1.5){\makebox(0,0)[cc]{$\omega_p$}} \rput(0.05,-1.5){\makebox(0,0)[cc]{$c'$}} \rput(0,-2.2){\makebox(0,0)[ct]{$b$}} \rput(0,-0.8){\makebox(0,0)[cb]{$a$}} \rput(-1.1,-1.5){\makebox(0,0)[rc]{$v_1$}} \rput(1.1,-1.5){\makebox(0,0)[lc]{$v_2$}} \end{pspicture} } {\psset{unit=1.2} \begin{pspicture}(-1.5,-2.5)(1,2.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=yellow](-1,1.5)(1,4.5)(1,-1.5)(-1,1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](-1,1.5)(0,2.5)(.5,1.5)(-1,1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](-1,1.5)(0,0.5)(.5,1.5)(-1,1.5) \pscircle*(-1,1.5){0.05} \pcline[linewidth=1pt]{->}(0,0.3)(0,-0.3) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=yellow](-1,-1.5)(1,-4.5)(1,1.5)(-1,-1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](-1,-1.5)(0,-2.5)(.5,-1.5)(-1,-1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](-1,-1.5)(0,-0.5)(.5,-1.5)(-1,-1.5) \pscircle*(-1,-1.5){0.05} \rput(-0.4,1.75){\makebox(0,0)[cc]{$\omega_p$}} \rput(-0.4,1.25){\makebox(0,0)[cc]{$\omega_q$}} \rput(0,1.8){\makebox(0,0)[lc]{$c$}} \rput(0.6,1.5){\makebox(0,0)[lc]{$a$}} \rput(0,1.2){\makebox(0,0)[lc]{$b$}} \rput(-0.4,-1.75){\makebox(0,0)[cc]{$\omega_p$}} \rput(-0.4,-1.25){\makebox(0,0)[cc]{$\omega_q$}} \rput(0,-1.8){\makebox(0,0)[lc]{$c'$}} \rput(0,-1.2){\makebox(0,0)[lc]{$b$}} \rput(0.6,-1.5){\makebox(0,0)[lc]{$a$}} \end{pspicture} } {\psset{unit=1.2} \begin{pspicture}(-1.5,-2.5)(1.5,2.5) \pscircle[linewidth=1pt,linecolor=white,fillstyle=solid,fillcolor=yellow](0,1.5){1.1} \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,1.5)(-0.3,2.8)(1.3,1.7)(0,1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,1.5)(-0.3,.2)(1.3,1.3)(0,1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,1.5)(-1,2.5)(-1,.5)(0,1.5) \pscircle*(0,1.5){0.05} \pcline[linewidth=1pt]{->}(0,0.3)(0,-0.3) \pscircle[linewidth=1pt,linecolor=white,fillstyle=solid,fillcolor=yellow](0,-1.5){1.1} \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,-1.5)(0.3,-2.8)(-1.3,-1.7)(0,-1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,-1.5)(0.3,-.2)(-1.3,-1.3)(0,-1.5) \psbezier[linewidth=1pt,fillstyle=solid,fillcolor=green](0,-1.5)(1,-2.5)(1,-.5)(0,-1.5) \pscircle*(0,-1.5){0.05} \rput(-0.4,1.5){\makebox(0,0)[cc]{$\omega_p$}} \rput(0.25,1.85){\makebox(0,0)[cc]{$\omega_q$}} \rput(0.25,1.15){\makebox(0,0)[cc]{$\omega_r$}} \rput(-0.85,1.5){\makebox(0,0)[rc]{$c$}} \rput(0.6,1){\makebox(0,0)[lc]{$a$}} \rput(0.6,2){\makebox(0,0)[lc]{$b$}} \rput(0.4,-1.5){\makebox(0,0)[cc]{$\omega_p$}} \rput(-0.25,-1.85){\makebox(0,0)[cc]{$\omega_r$}} \rput(-0.25,-1.15){\makebox(0,0)[cc]{$\omega_q$}} \rput(0.82,-1.5){\makebox(0,0)[lc]{$c'$}} \rput(-0.6,-1){\makebox(0,0)[rc]{$b$}} \rput(-0.6,-2){\makebox(0,0)[rc]{$a$}} \end{pspicture} } \end{center} In all the three cases above, the mutation law is given by Lemma~\ref{cc-prime}. In the first case, we transfer the cluster variable from one set $\delta_k$ to another set (if the vertices $v_1$ and $v_2$ are distinct); these transformations may also change the cyclic ordering of orbifold points inside a set $\delta_k$. Note that the label $\omega_p$ remains assigned to the transformed edge. \subsection{Fat graph description for Riemann surfaces with holes and with ${\mathbb Z}_p$ orbifold points} In this subsection and in the rest of the paper, we use the graphs dual to the above ideal triangle decompositions of Riemann surfaces. These graphs are especially useful when describing the Fuchsian groups $\Delta_{g,s,r}$ of Riemann surfaces $\Sigma_{g,s,r}$ and the corresponding geodesic functions. \begin{df}\label{def-pend} {\rm We call a fat graph (a graph with the prescribed cyclic ordering of edges entering each vertex) $\Gamma_{g,s,r}$ a {\em spine of the Riemann surface} $\Sigma_{g,s,r}$ with $g$ handles, $s>0$ holes, and $r$ orbifold points of the corresponding orders $p_i$, $i=1,\dots,r$, if \begin{itemize} \item[(a)] this graph can be embedded without self-intersections in $\Sigma_{g,s,r}$; \item[(b)] all vertices of $\Gamma_{g,s,r}$ are three-valent except exactly $r$ one-valent vertices (endpoints of ``pending'' edges), which are placed at the corresponding orbifold points; \item[(c)] for an orbifold point from the set $\delta_k$, the corresponding pending edge protrudes towards the interior of the face of the graph containing the $k$th hole; the cyclic ordering of pending edges pointing towards the interior of this face coincide with that of orbifold points in the set $\delta_k$; \item[(d)] upon cutting along all edges of $\Gamma_{g,s,r}$ the Riemann surface $\Sigma_{g,s,r}$ splits into $s$ polygons each containing exactly one hole and being simply connected upon gluing this hole. \end{itemize} Edges of this graph are labeled by distinct integers $\alpha=1,2,\dots,6g-6+3s+2r$, and to each edge we set into the correspondence the real number $Z_\alpha$. } \end{df} The first homotopy groups $\pi_1(\Sigma_{g,s,r})$ and $\pi_1(\Gamma_{g,s,r})$ coincide because each closed path in $\Sigma_{g,s,r}$ can be homotopically transformed to a closed path in $\Gamma_{g,s,r}$ (taking into account paths that go around orbifold points) in a unique way. The standard statement in hyperbolic geometry is that conjugacy classes of elements of a Fuchsian group $\Delta_{g,s,r}$ are in the 1-1 correspondence with homotopy classes of closed paths in the Riemann surface $\Sigma_{g,s.r}={\mathbb H}^2_+/\Delta_{g,s,r}$ and that the ``actual'' length $\ell_\gamma$ of a hyperbolic element $\gamma\in\Delta_{g,s,r}$ coincides with the minimum length of curves from the corresponding homotopy class; it is then the length of a unique closed geodesic line belonging to this class. When orbifold points are present, the Fuchsian group contains besides hyperbolic elements also elliptic elements corresponding to rotations about these orbifold points. The corresponding generators ${\wtd F}_i$, $i=1,\dots,r$, of the rotations through $2\pi/p_i$ are conjugates of the matrices \begin{equation} \label{F-p} {\wtd F}_i=U_iF_{p_i}U_i^{-1},\qquad F_p= \left(\begin{array}{cc} 0 & 1\\ -1 & -w\end{array}\right),\quad w=2\cos{\pi/p}, \quad p\ge 2. \end{equation} The real numbers $Z_\alpha$ in Definition~\ref{def-pend} are the $h$-lengths \cite{Penn1}: they are called the {\em (Thurston) shear coordinates} \cite{ThSh},\cite{Bon2} in the case of punctured Riemann surface (without boundary components). We preserve this notation and this term also in the case of orbifold surfaces. These coordinates are related to the cross-ratio relation for two adjacent ideal triangles constituting the ideal quadrangle with the respective vertices (in the cyclic order) $a,b,c,d$ and diagonal $bd$. At the same time, they are related to the cluster variables $a_i$ corresponding to the sides of the corresponding quadrangle. We have \begin{equation} \label{four-term} e^{Z}=-\frac{(b-c)(d-a)}{(b-a)(d-c)}=\frac{a_1a_3}{a_2a_4}, \end{equation} and we obtain the parameter $Z_\alpha$ choosing $\{a,b,c,d\}=\{-1,0,e^{Z_\alpha},\infty\}$. For example, the fat graph corresponding to the pattern in Fig.~\ref{fi:disc} is depicted in Fig.~\ref{fi:treegraph}. \begin{figure}[tb] {\psset{unit=0.6} \begin{pspicture}(-10,-5)(8,5) \pscircle[linewidth=0.5pt](0,0){4} \psarc[linecolor=green, linewidth=0.5pt](-4,1.64){1.64}{-90}{45} \psarc[linecolor=green, linewidth=0.5pt](0,5.66){4}{-135}{-45} \psarc[linecolor=green, linewidth=0.5pt](4,1.64){1.64}{135}{270} \psarc[linestyle=dashed,linecolor=green, linewidth=0.5pt](4,-2.31){2.31}{90}{210} \rput{142}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.314,0){1.616}{-68}{68}} \rput{172}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.04,0){.56}{-82}{82}} \psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-.67,-4.99){3.06}{30}{135} \rput{57.5}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.09,0){.85}{-78}{78}} \rput{95}(0,0){\psarc[linestyle=dashed, linecolor=green, linewidth=0.5pt](-4.40,0){1.86}{-65}{65}} \psarc[linecolor=green, linewidth=0.5pt](-4,-1.64){1.64}{-45}{90} \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.83,2.83)(-1.4,1.4)(2,0)(4,0) \psbezier[linecolor=green, linestyle=dashed, linewidth=0.5pt](-2.83,2.83)(-1.4,1.4)(1,-1.732)(2,-3.464) \rput{90}(0,0){\psarc[linecolor=green, linestyle=dashed, linewidth=0.5pt](0,5.66){4}{-135}{-45}} \pcline[linewidth=1pt](-3.1,0.5)(-2.6,0) \pcline[linewidth=1pt](-2.6,1)(-1.9,0.3) \pcline[linewidth=1pt](-3.2,-0.4)(-2.6,0) \pcline[linewidth=1pt](-2.8,-0.9)(-2.2,-0.4) \pcline[linewidth=1pt](-1.53,-2.5)(-1.4,-0.8) \pcline[linewidth=1pt](-.73,-2.5)(-.6,-0.9) \pcline[linewidth=1pt](1.7,.3)(3.4,-.485) \pcline[linewidth=1pt](1.4,-.2)(3.1,-.985) \pcline[linewidth=1pt](1.4,1.4)(2.7,1.8) \pcline[linewidth=1pt](1.5,.8)(2.9,1.2) \pcline[linewidth=1pt](-0.1,1.8)(.9,.8) \pcline[linewidth=1pt](.4,2.4)(1.4,1.4) \pcline[linecolor=blue, linewidth=1pt](-2.2,-0.4)(-1.4,-0.8) \pcline[linecolor=blue, linewidth=1pt](-1.9,0.3)(-.9,-0.2) \pcline[linecolor=blue, linewidth=1pt](-.6,-0.9)(1.4,-.2) \pcline[linecolor=blue, linewidth=1pt](1.05,.4)(-.9,-0.2) \pcline[linecolor=blue, linewidth=1pt](1.5,.8)(1.7,.3) \pcline[linecolor=blue, linewidth=1pt](.9,.8)(1.05,.4) \psbezier[linecolor=white, linewidth=14pt](.25,2.1)(-2.15,4.4)(-2.7,4.7)(-4.2,3.2) \psbezier[linecolor=white, linewidth=14pt](-4.2,3.2)(-5.8,1.6)(-5.4,-3.05)(-3,-0.65) \psbezier[linestyle=dashed, linewidth=1pt](-.1,1.8)(-2.1,3.8)(-2.7,4.3)(-4,3) \psbezier[linestyle=dashed, linewidth=1pt](-4,3)(-5.3,1.7)(-5.2,-2.4)(-3.2,-0.4) \psbezier[linestyle=dashed, linewidth=1pt](.4,2.4)(-2.2,5)(-2.7,5.1)(-4.4,3.4) \psbezier[linestyle=dashed, linewidth=1pt](-4.4,3.4)(-6.3,1.5)(-5.6,-3.7)(-2.8,-0.9) \psbezier[linecolor=white, linewidth=14pt](2.8,1.5)(5.7,2.4)(2.5,4.75)(0,4.75) \psbezier[linecolor=white, linewidth=14pt](0,4.75)(-2.5,4.75)(-5,2.9)(-2.85,.75) \psbezier[linestyle=dashed, linewidth=1pt](2.7,1.8)(5.3,2.6)(2,4.5)(0,4.5) \psbezier[linestyle=dashed, linewidth=1pt](0,4.5)(-2,4.5)(-4.6,3)(-2.6,1) \psbezier[linestyle=dashed, linewidth=1pt](2.9,1.2)(6.15,2.2)(3,5)(0,5) \psbezier[linestyle=dashed, linewidth=1pt](0,5)(-3,5)(-5.4,2.8)(-3.1,0.5) \rput(3,-0.65){\makebox(0,0){$\star$}} \rput(-1.15,-2.5){\makebox(0,0){$\star$}} \rput(-1,0.1){\makebox(0,0)[cb]{$Y_2$}} \rput(1,-0.6){\makebox(0,0)[ct]{$Y_3$}} \rput(2.3,0.4){\makebox(0,0)[cb]{$Y_4$}} \rput(-4,-2.3){\makebox(0,0)[rt]{$A$}} \rput(4.7,2.6){\makebox(0,0)[rt]{$B$}} \rput(2.8,-1.3){\makebox(0,0)[lt]{$Z_2$}} \rput(-0.4,-2.3){\makebox(0,0)[lt]{$Z_1$}} \end{pspicture} } \caption{\small The fat graph corresponding to the ideal triangle partition of the fundamental domain in Fig.~\ref{fi:disc}. The real numbers $Z_1$ and $Z_2$ are associated to the pending edges and $A$, $B$, and $Y_2$, $Y_3$, and $Y_4$ to the inner edges.} \label{fi:treegraph} \end{figure} \subsection{The Fuchsian group $\Delta_{g,s,r}$ and geodesic functions}\label{ss:geodesic} We now describe combinatorially the conjugacy classes of the Fuchsian group $\Delta_{g,s,r}$. Every time the path homeomorphic to a (closed) geodesic $\gamma$ goes through the edge with the label $\alpha$ we insert~\cite{Fock1} the matrix of the M\"obius transformation \begin{equation} \label{XZ} X_{Z_\alpha}=\left( \begin{array}{cc} 0 & -e^{Z_\alpha/2}\\ e^{-Z_\alpha/2} & 0\end{array}\right). \end{equation} We also have the ``right'' and ``left'' turn matrices to be set in proper places when a path makes the corresponding turns at three-valent vertices, \begin{equation} \label{R} R=\left(\begin{array}{cc} 1 & 1\\ -1 & 0\end{array}\right), \qquad L= R^2=\left(\begin{array}{cc} 0 & 1\\ -1 & -1\end{array}\right). \end{equation} New elements of the Fuchsian group correspond to rotations of geodesic lines when going around orbifold points indicated by star-vertices; for a ${\mathbb Z}_p$ orbifold point we then insert the matrix $F_p$ (\ref{F-p}) into the corresponding string of $2\times2$-matrices (when we go around the orbifold point counterclockwise as in Fig.~\ref{fi:corner}(a)). When the order of the orbifold point is larger than two, we can go around it $k$ times; {\em we then have to insert the matrix $(-1)^{k+1}F_p^k$ into the product of $2\times2$-matrices}. For example, parts of geodesic functions in the three cases in Fig.~\ref{fi:corner} read \begin{equation} \label{XZFXZ} \begin{array}{ll} \hbox{(a)}\quad & \dots X_XLX_ZF_pX_ZLX_Y\dots, \\ \hbox{(b)}\quad & \dots X_XLX_Z(-F^2_p)X_ZRX_X\dots, \\ \hbox{(c)}\quad & \dots X_YRX_Z(F^3_p)X_ZLX_Y\dots. \\ \end{array} \end{equation} Note that $F_p^p=(-1)^{p-1}{\mathbb E}$, so going around the ${\mathbb Z}_p$ orbifold point $p$ times merely corresponds to avoiding this orbifold point due to the simple equality (note that $X_S^2=-{\mathbb E}$ and $L^2=-R$) $$ X_XLX_Z(-1)^{p-1}F_p^pX_ZLX_Y=X_XLX_Z^2LX_Y=-X_XL^2X_Y=X_XRX_Y. $$ (For the ${\mathbb Z}_2$ orbifold points this pattern was first proposed by Fock and Goncharov \cite{FG}; the graph morphisms were described in \cite{Ch2}.) \begin{figure}[tb] {\psset{unit=0.7} \begin{pspicture}(-2,-3)(2,2) \pcline[linewidth=1pt](-1.5,1)(1.5,1) \pcline[linewidth=1pt](-1.5,0)(-0.5,0) \pcline[linewidth=1pt](1.5,0)(0.5,0) \pcline[linewidth=1pt](-.5,0)(-.5,-1) \pcline[linewidth=1pt](.5,0)(.5,-1) \rput(0,-1){\makebox(0,0){$\star$}} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{<-}(-1.5,0.2)(-.5,0.2) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](-.5,0){.2}{0}{90} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](-.3,0)(-.3,-1) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](0,-1){.3}{-180}{0} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](.3,0)(.3,-1) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](.5,0){.2}{90}{180} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{<-}(.5,0.2)(1.5,0.2) \rput(-1.3,1.2){\makebox(0,0)[cb]{$X$}} \rput(1.3,1.2){\makebox(0,0)[cb]{$Y$}} \rput(0.8,-.7){\makebox(0,0){$Z$}} \rput(0,-1.8){\makebox(0,0){$F_p$}} \rput(0,-2.4){\makebox(0,0){(a)}} \end{pspicture} \begin{pspicture}(-2,-3)(2,2) \pcline[linewidth=1pt](-1.5,1)(1.5,1) \pcline[linewidth=1pt](-1.5,0)(-0.5,0) \pcline[linewidth=1pt](1.5,0)(0.5,0) \pcline[linewidth=1pt](-.5,0)(-.5,-1) \pcline[linewidth=1pt](.5,0)(.5,-1) \rput(0,-1){\makebox(0,0){$\star$}} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{<-}(-1.5,0.2)(-.5,0.2) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](-.5,0.05){.15}{0}{90} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](-.35,0.05)(-.35,-1) \psarc[linecolor=blue, linewidth=0.7pt](0.075,-1){0.275}{-180}{0} \psarc[linecolor=blue, linewidth=0.7pt](-0.075,-1){0.275}{-180}{0} \psarc[linecolor=blue, linewidth=0.7pt](0,-1){0.2}{0}{180} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](.35,0.05)(.35,-1) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](-.4,0.05){.75}{0}{90} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{->}(-1.5,0.8)(-.4,0.8) \rput(-1.3,1.2){\makebox(0,0)[cb]{$X$}} \rput(1.3,1.2){\makebox(0,0)[cb]{$Y$}} \rput(0.8,-.7){\makebox(0,0){$Z$}} \rput(0,-1.8){\makebox(0,0){$-F^2_p$}} \rput(0,-2.4){\makebox(0,0){(b)}} \end{pspicture} \begin{pspicture}(-2,-3)(2,2) \pcline[linewidth=1pt](-1.5,1)(1.5,1) \pcline[linewidth=1pt](-1.5,0)(-0.5,0) \pcline[linewidth=1pt](1.5,0)(0.5,0) \pcline[linewidth=1pt](-.5,0)(-.5,-1) \pcline[linewidth=1pt](.5,0)(.5,-1) \rput(0,-1){\makebox(0,0){$\star$}} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{->}(1.5,0.2)(.5,0.2) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](.5,0.1){.1}{90}{180} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](-.4,0.1)(-.4,-1) \psarc[linecolor=blue, linewidth=0.7pt](0.05,-1){.35}{-180}{0} \psarc[linecolor=blue, linewidth=0.7pt](-0.05,-1){.35}{-180}{0} \psarc[linecolor=blue, linewidth=0.7pt](0,-1){.2}{-180}{0} \psarc[linecolor=blue, linewidth=0.7pt](0.05,-1){.25}{0}{180} \psarc[linecolor=blue, linewidth=0.7pt](-0.05,-1){.25}{0}{180} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt](.4,0.1)(.4,-1) \psarc[linecolor=blue, linestyle=dashed, linewidth=1pt](.3,0.1){.7}{90}{180} \pcline[linecolor=blue, linestyle=dashed, linewidth=1pt]{<-}(1.5,0.8)(.3,0.8) \rput(-1.3,1.2){\makebox(0,0)[cb]{$X$}} \rput(1.3,1.2){\makebox(0,0)[cb]{$Y$}} \rput(0.8,-.7){\makebox(0,0){$Z$}} \rput(0,-1.8){\makebox(0,0){$F^3_p$}} \rput(0,-2.4){\makebox(0,0){(c)}} \end{pspicture} } \caption{\small Part of a graph with a pending edge. Its endpoint with the orbifold point is directed toward the interior of the boundary component this point is associated with. The variable $Z$ corresponds to the respective pending edge. We present four typical examples of geodesics undergoing single (a), double (b), and triple (c) rotations at the ${\mathbb Z}_p$ orbifold point.} \label{fi:corner} \end{figure} An element of a Fuchsian group has then the typical structure \begin{equation} \label{Pgamma} P_{\gamma}=LX_{Y_n}RX_{Y_{n-1}}\cdots RX_{Y_2}LX_{Z_1}(-1)^{k+1}F^k_p X_{Z_1}RX_{Y_1}, \end{equation} where $Y_i$ are variables of ``internal'' edges and $Z_j$ are those of pending edges. The corresponding {\em geodesic function} \begin{equation} \label{G} G_{\gamma}\equiv \,{\rm Tr}\, P_\gamma=2\cosh(\ell_\gamma/2) \end{equation} is expressed via the actual length $\ell_\gamma$ of the closed geodesic on the Riemann surface. \begin{remark}\label{rm-positivity} Note that the combinations $$ RX_y=\left(\begin{array}{cc} e^{-Y/2} & -e^{Y/2}\\ 0 & e^{Y/2}\end{array}\right)\quad \hbox{and}\quad LX_y=\left(\begin{array}{cc} e^{-Y/2} & 0 \\ -e^{-Y/2} & e^{Y/2}\end{array}\right) $$ as well as products of any number of these matrices have the sign structure $\left(\begin{array}{cc} + & -\\ - & +\end{array}\right)$, so the trace of any of $P_\gamma$ in the absence of orbifold points is a sum of exponentials with positive integer coefficients; this sum always include the terms $e^{Y_1/2+\cdots+Y_n/2}$ and $e^{-Y_1/2-\cdots-Y_n/2}$ being therefore always greater or equal two thus describing a hyperbolic or parabolic element; the latter is possible only if $Y_1+\cdots+Y_n=0$ and only if the turn matrices in (\ref{Pgamma}) are all either $R$ or $L$, which corresponds to a path going along the boundary of a face; all such passes are homeomorphic to the hole boundaries, and the condition that the sum of $Y_i$ equals zero indicates the degeneration of a hole into a puncture. \end{remark} The group generated by elements (\ref{F-p}) together with translations along $A$- and $B$-cycles and around holes not necessarily produces a regular (metrizable) surface because its action is not necessarily discrete. We formulate the necessary and sufficient conditions for producing a {\em regular} surface in terms of graphs (see \cite{Ch1a} for the ${\mathbb Z}_2$ orbifold point case).\footnote{In what follows, we call a Riemann surface regular if it is locally a smooth constant-curvature surface everywhere except exactly $r$ orbifold points.} To formulate the regularity condition, we interpret passages around orbifold points as paths in the $p$-fold covering of the geodesic neighborhood in Fig.~\ref{fi:saucer-pan}. For this, we take the subgraph in Fig.~\ref{fi:p-gone-dual} dual to the ideal triangle decomposition in Fig.~\ref{fi:p-gone}. When splitting the equilateral $p$-gone into ideal triangles we break the $p$-fold symmetry, so now the shear coordinates $Z_i$, $i=1,\dots,p$, on the fat graph edges dual to the corresponding $p$-gone sides and $Y_j$, $j=2,\dots,p-2$, on the edges dual to the diagonals of the $p$-gone are different. \begin{figure}[tb] {\psset{unit=0.8} \begin{pspicture}(5,-5)(-5,5) \pscircle[linewidth=0.5pt](0,0){4} \newcommand{\PATTERN}{% \psarc[linecolor=green, linewidth=1pt](4.62,0){2.3}{120}{240} \rput{-20}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.06,0){.7}{100}{260}} \rput{10}(0,0){\psarc[linestyle=dashed, linewidth=1pt](4.26,0){1.45}{110}{250}} } \rput{30}(0,0){\PATTERN} \rput{90}(0,0){\PATTERN} \rput{150}(0,0){\PATTERN} \rput{210}(0,0){\PATTERN} \rput{270}(0,0){\PATTERN} \rput{330}(0,0){\PATTERN} \rput{60}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \rput{-60}(0,0){ \psarc[linestyle=dashed,linecolor=green, linewidth=1pt](8,0){6.93}{150}{210} } \pcline[linestyle=dashed,linecolor=green, linewidth=1pt](-4,0)(4,0) \rput(0,0){\makebox(0,0){$\star$}} \pcline[linecolor=red, linewidth=10pt](-0.8,0.8)(-0.8,-0.8) \pcline[linecolor=red, linewidth=10pt](-0.8,-0.8)(-2.5,-1.1) \pcline[linecolor=red, linewidth=10pt](-0.8,0.8)(-2.1,2.1) \pcline[linecolor=red, linewidth=10pt](-0.8,0.8)(1,1.5) \pcline[linecolor=red, linewidth=10pt](-0.8,-0.8)(0.4,-1.9) \pcline[linecolor=red, linewidth=10pt](1,1.5)(0.6,3) \pcline[linecolor=red, linewidth=10pt](1,1.5)(2.8,0.9) \pcline[linecolor=red, linewidth=10pt](0.4,-1.9)(-0.6,-2.8) \pcline[linecolor=red, linewidth=10pt](0.4,-1.9)(2.2,-2.1) \pcline[linecolor=white, linewidth=8pt](-0.8,0.8)(-0.8,-0.8) \pcline[linecolor=white, linewidth=8pt](-0.8,-0.8)(-2.5,-1.1) \pcline[linecolor=white, linewidth=8pt](-0.8,0.8)(-2.1,2.1) \pcline[linecolor=white, linewidth=8pt](-0.8,0.8)(1,1.5) \pcline[linecolor=white, linewidth=8pt](-0.8,-0.8)(0.4,-1.9) \pcline[linecolor=white, linewidth=8pt](1,1.5)(0.6,3) \pcline[linecolor=white, linewidth=8pt](1,1.5)(2.8,0.9) \pcline[linecolor=white, linewidth=8pt](0.4,-1.9)(-0.6,-2.8) \pcline[linecolor=white, linewidth=8pt](0.4,-1.9)(2.2,-2.1) \rput(4.1,0){\makebox(0,0)[lc]{$1$}} \rput(3.8,1.5){\makebox(0,0)[lc]{$e^{i\phi}$}} \rput(2.2,3.4){\makebox(0,0)[lb]{$e^{2\pi i/p}$}} \rput(0.5,4.1){\makebox(0,0)[lb]{$e^{2\pi i/p+i\phi}$}} \rput(-2.2,3.3){\makebox(0,0)[rb]{$e^{4\pi i/p}$}} \rput(3.1,-2.5){\makebox(0,0)[lt]{$e^{-2\pi i/p+i\phi}$}} \rput(2,-3.3){\makebox(0,0)[lt]{$e^{-2\pi i/p}$}} \rput(-0.2,-4.1){\makebox(0,0)[rt]{$e^{-4\pi i/p+i\phi}$}} \rput(-2.2,-3.3){\makebox(0,0)[rt]{$e^{-4\pi i/p}$}} \rput(0,1.5){\makebox(0,0)[cb]{$Y_2$}} \rput(-1.1,0){\makebox(0,0)[rc]{$Y_3$}} \rput(-0.3,-1){\makebox(0,0)[lb]{$Y_4$}} \rput(2.2,.8){\makebox(0,0)[rt]{$Z_1$}} \rput(-0.1,2.6){\makebox(0,0)[lc]{$Z_2$}} \rput(-1.7,1.2){\makebox(0,0)[ct]{$Z_3$}} \rput(-1.9,-0.3){\makebox(0,0)[rt]{$Z_4$}} \rput(-0.3,-2.1){\makebox(0,0)[rc]{$Z_5$}} \rput(1.2,-1.3){\makebox(0,0)[lt]{$Z_6$}} \end{pspicture} } \caption{\small The (tree-like) subgraph dual to the ideal triangle decomposition of the ideal equilateral $p$-gone in Fig.~\ref{fi:p-gone}. All the variables $Z_\alpha$ and $Y_\beta$ are determined by the cross-ratio relations in the corresponding ideal quadrangles based on the points from the sets $\{e^{2\pi i k/p}\}$ and $\{e^{2\pi i k/p+i\phi}\}$, $k=0,\dots,p-1$.} \label{fi:p-gone-dual} \end{figure} We first identify the parameter $Z$ in (\ref{XZFXZ}) to be \begin{equation} \label{eZ} e^Z=\frac{\sin\left(\pi/p-\phi/2\right)}{\sin(\phi/2)}, \end{equation} We can then derive the explicit relations between the parameter $Z$ in (\ref{eZ}) and the variables $Z_i$ and $Y_j$ determined by the standard cross-ratio relations (\ref{four-term}). The vertices of the ideal $p$-gone are situated at the points $e^{i2\pi k/p}$, $k=0,\dots,p-1$, and $p$ copies of the vertex of an additional ideal triangle adjacent to the $p$-gone side are $e^{i\phi+i2\pi ik/p}$, $k=0,\dots,p-1$. Using the cross-ratio relations to calculate $Z_i$ and $Y_j$ we find the exact relations between these variables and the variable $Z$ given by (\ref{eZ}): \begin{eqnarray} e^{Z_1}&=&e^Z\frac{\sin(2\pi/p)}{\sin(\pi/p)},\nonumber\\ e^{Z_p}&=&e^Z\frac{\sin(\pi/p)}{\sin(2\pi/p)},\nonumber\\ e^{Z_k}&=&e^Z\frac{\sin\bigl((k-1)\pi/p\bigr)}{\sin(k\pi/p)},\quad k=2,\dots,p-1,\label{ZK}\\ e^{Y_k}&=&\frac{\sin\bigl((k+1)\pi/p\bigr)}{\sin\bigl((k-1)\pi/p\bigr)}, \quad k=2,\dots,p-2.\label{YK} \end{eqnarray} The following $2\times 2$-matrix equalities can be verified directly: \begin{equation} \label{equivalence} X_ZF_pX_Z=X_{Z_1}LX_{Z_2}=X_{Z_{k}}LX_{Y_{k}}LX_{Z_{k+1}}=X_{Z_{p-1}}LX_{Z_p}, \ k=2,\dots,p-2. \end{equation} We then have the following lemma. \begin{lm}\label{lem-length} We have the following explicit $2\times 2$-matrix relations for the shear variables in the equilateral $p$-gone in Fig.~\ref{fi:p-gone-dual} given by (\ref{ZK}) and (\ref{YK}): \begin{eqnarray} X_ZF_pX_Z&=&X_{Z_1}LX_{Z_2}\nonumber\\ X_Z(-F_p^2)X_Z&=&X_{Z_1}RX_{Y_2}LX_{Z_3}\nonumber\\ \vdots&{}&\label{P-gone->pending}\\ X_Z(-1)^{k-1}F_p^{k}X_Z&=&X_{Z_1}RX_{Y_2}R\cdots RX_{Y_k}LX_{Z_{k+1}}, \ k=2,\dots,p-2\nonumber\\ X_Z(-1)^{p}F_p^{p-1}X_Z&=&X_{Z_1}R X_{Y_{2}}R\cdots RX_{Y_{p-2}}R X_{Z_p}. \nonumber \end{eqnarray} \end{lm} The {\it proof} uses equalities from (\ref{equivalence}) for constructing the longer chain using that $X_SX_S=-{\mathbb E}$ for any variable $S$ and that $L^2=-R$. For example, we obtain the r.h.s. of the second equality in (\ref{P-gone->pending}) multiplying $X_{Z_1}LX_{Z_2}\cdot X_{Z_{2}}LX_{Y_{2}}LX_{Z_{3}}$ whereas the l.h.s. merely becomes $X_ZF_pX_Z\cdot X_ZF_pX_Z=X_Z(-1)F_p^2 X_Z$. All other equalities are obtained if we continue this chain of multiplications. Due to Lemma~\ref{lem-length}, all ``rotations'' about orbifold points $X_Z(-1)^{k+1}F_p^k X_Z$ are now presented as the standard products of matrices $X_S$ (with real $S$) alternated with the matrices of left and right turns (\ref{R}), which means that the conclusion of Remark~\ref{rm-positivity} is valid in this case as well: as soon as in the original spine $\Gamma_{g,s,r}$ all the parameters $Z_\alpha$ of pending edges are real, all the geodesic functions constructed on a Riemann surface with orbifold points are Laurent polynomials with {\em positive integer coefficients} of the ``new'' real variables $Z^{(\alpha,p)}_i$, $Y^{(p)}_j$ (where the superscripts $\alpha,p$ indicate that these variables are completely determined by the original variable $Z_\alpha$ and the order $p$ of the orbifold point) and ``old'' variables of ``internal'' edges of the spine $\Gamma_{g,s,r}$. corresponding to usual partitions into ideal triangles. So, again, in the trace of every product of form (\ref{Pgamma}) we necessarily have the term $2\cosh (\sum_{\beta=1}^n X_\beta)$, where the sum ranges all edges (new and internal ones) the corresponding path goes over, and we let $X$ denote the variables on all these edges disregarding their origins. Every such trace is therefore a positive number greater or equal two, and the corresponding element of the group will be either hyperbolic or parabolic (the latter is possible only for geodesics around holes and only if a hole reduces to a puncture). The only elliptic elements are precisely conjugates of $F_p^k$. We therefore come to the theorem \begin{theorem} \label{lem-metric} We have a metrizable Riemann surface for {\em any} choice of real numbers $Z_\alpha$ associated to the edges of an original spine $\Gamma_{g,s,r}$. The converse statement is also true: for {\em any} metrizable Riemann surface $\Sigma_{g,s,r}$ we have a spine $\Gamma_{g,s,r}$ with real numbers associated to its edges such that the lengths of geodesics on $\Sigma_{g,s,r}$ are given by traces of products (\ref{Pgamma}) corresponding to paths in the spine. \end{theorem} The {\em proof} of the second statement was performed in \cite{Ch2} for ${\mathbb Z}_2$-orbifold points. It is based on the (obvious) existence of the ideal triangle decomposition described in Sec.~\ref{ss:clusters} for any metrizable Riemann surface and can be straightforwardly generalized to the case of orbifold points of any type. We have therefore parameterized all possible regular surfaces in terms of the $(6g-6+3s+2r)$-tuple of real coordinates $\{Z_\alpha\}$. \begin{corollary} The decorated Teichm\"uller space ${\mathfrak T}^{H}_{g,s,r}$ of Riemann surfaces with holes and orbifold points is the space ${\mathbb R}^{6g-6+3s+2r}$ of real parameters on the edges of a spine $\Gamma_{g,s,r}$. \end{corollary} \section{Mapping class group transformations}\label{s:mcg} \subsection{Poisson structure}\label{ss:Poisson} One of the most attractive properties of the graph description is a very simple Poisson algebra on the set of parameters $Z_\alpha$. The following result is the straightforward generalization of the theorem formulated for surfaces without marked points in~\cite{Fock1} and for surfaces with order-2 orbifold points in~\cite{FG} (see also \cite{Ch1}). \begin{theorem}\label{th-WP} In the coordinates $Z_\alpha $ on any fixed spine corresponding to a surface with orbifold points, the Weil--Petersson bracket $B_{{\mbox{\tiny WP}}}$ is given by \begin{equation} \label{WP-PB} \bigl\{f({\mathbf Z}),g({\mathbf Z})\bigr\}=\sum_{{\hbox{\small 3-valent} \atop \hbox{\small vertices $\alpha=1$} }}^{4g+2s+|\delta|-4} \,\sum_{i=1}^{3\!\!\mod 3} \left(\frac{\partial f}{\partial Z_{\alpha_i}} \frac{\partial g}{\partial Z_{\alpha_{i+1}}} - \frac{\partial g}{\partial Z_{\alpha_i}} \frac{\partial f}{\partial Z_{\alpha_{i+1}}}\right), \end{equation} where the sum ranges all the three-valent vertices of a graph and $\alpha_i$ are the labels of the cyclically (counterclockwise) ordered ($\alpha_4\equiv \alpha_1 $) edges incident to the vertex with the label $\alpha$ irrespectively on whether these edges are internal or pending edges of the graph. This bracket gives rise to the {\em Goldman bracket} on the space of geodesic length functions \cite{Gold}. \end{theorem} We identify the exchange matrix $B$ with the matrix of the Poisson relations for the variables $Z_\alpha$. The center of this Poisson algebra is provided by the proposition. \begin{prop}\label{prop12} The center of the Poisson algebra {\rm(\ref{WP-PB})} is generated by elements of the form $\sum Z_\alpha$, where the sum ranges all edges of $\Gamma_{g,s,r} $ belonging to the same boundary component taken with multiplicities. This means, in particular, that each pending edge, irrespectively on the type of orbifold point it corresponds to, contributes twice to such sums. The dimension of this center is obviously $s$. \end{prop} for the proof in the general case see Appendix~B of~\cite{ChP}. \subsection{Flip morphisms of fat graphs}\label{ss:flip} In this section, we present the complete list of mapping class group transformations that enable us to change numbers $|\delta_k|$ of orbifold points associated with the $k$th hole, change the cyclic ordering inside any of the sets $\delta_k$, flip any inner edge of the graph and, eventually, change the orientation of the geodesic spiraling to the hole perimeter (in the case where we have more than one hole).\footnote{These transformations are dual to mutations of cluster variables from Sec.~\ref{s-algebra}.} We can therefore establish a morphism between any two of the graphs belonging to the same class $\Gamma_{g,s,r}$ with the same (unordered) sets of orbifold point orders $\{p_i\}_{i=1}^r$. \subsubsection{Whitehead moves on inner edges}\label{sss:mcg} Given a spine $\Gamma$ of $\Sigma$ and assuming that the edge $\alpha$ has distinct endpoints, we may produce another spine $\Gamma _\alpha$ of $\Sigma$ by contracting and expanding edge $\alpha$ of $\Gamma $, the edge labeled $Z$ in Figure~\ref{fi:flip}. This transformation is dual to the mutation (\ref{diagonal}). We say that $\Gamma _\alpha$ arises from $\Gamma$ by a {\it Whitehead move} (or flip) along the edge $\alpha$. A labeling of edges of the spine $\Gamma$ implies a natural labeling of edges of the spine $\Gamma_\alpha$; we then obtain a morphism between the spines $\Gamma$ and $\Gamma_\alpha$. \begin{figure}[tb] \setlength{\unitlength}{1.5mm}% \begin{picture}(-60,27)(60,48) \thicklines \put(32,62){\line( -1,2){ 4}} \put(36,64){\line( -1,2){ 4}} \put(32,62){\line(-1,-2){ 4}} \put(36,60){\line(-1,-2){ 4}} \put(36,60){\line( 1, 0){20}} \put(36,64){\line( 1, 0){20}} \put(60,62){\line( 1, 2){ 4}} \put(60,62){\line( 1,-2){ 4}} \put(56,64){\line( 1, 2){ 4}} \put(56,60){\line( 1,-2){ 4}} \thinlines \put(21,62){\vector(-1, 0){ 0}} \put(21,62){\vector( 1, 0){ 5}} \thicklines \put(10,54){\line( -2,-1){ 8}} \put(8,58){\line( -2,-1){ 8}} \put(8,58){\line( 0,1){8}} \put(12,58){\line( 0,1){8}} \put(10,54){\line(2,-1){ 8}} \put(12,58){\line(2,-1){ 8}} \put(10,70){\line( -2,1){ 8}} \put(8,66){\line( -2,1){ 8}} \put(10,70){\line(2,1){ 8}} \put(12,66){\line(2,1){ 8}} \put( 4,74){\makebox(0,0)[lb]{$A$}} \put(16,74){\makebox(0,0)[rb]{$B$}} \put(14,62){\makebox(0,0)[lc]{$Z$}} \put(16,50){\makebox(0,0)[rt]{$C$}} \put( 4,50){\makebox(0,0)[lt]{$D$}} \put(32,52){\makebox(0,0)[lt]{$D - \phi(-Z)$}} \put(60,52){\makebox(0,0)[rt]{$C+\phi(Z)$}} \put(60,73){\makebox(0,0)[rb]{$B-\phi(-Z)$}} \put(32,73){\makebox(0,0)[lb]{$A+\phi(Z)$}} \put(47,66){\makebox(0,0)[cb]{$-Z$}} \color[rgb]{1,0,0} \thinlines \put(0.5,53){\line( 2,1){ 6}} \put(1,52){\line( 2,1){ 6}} \put(1.5,51){\line( 2,1){ 6}} \put(0.5,71){\line( 2,-1){ 6}} \put(19,72){\line( -2,-1){ 6}} \put(18.5,51){\line( -2,1){ 6}} \qbezier(6.5,56)(9,57.75)(9,60) \qbezier(6.5,68)(9,66.25)(9,64) \put(9,60){\line( 0,1){ 4}} \qbezier(7,55)(10,56.5)(10,62) \qbezier(13,69)(10,67.5)(10,62) \qbezier(7.5,54)(10,55.75)(12.5,54) \put(-1.5,72){\makebox(0,0)[cc]{\hbox{\small$1$}}} \put(-1.5,72){\circle{3}} \put(21,73){\makebox(0,0)[cc]{\hbox{\small$2$}}} \put(21,73){\circle{3}} \put(20.5,50){\makebox(0,0)[cc]{\hbox{\small$3$}}} \put(20.5,50){\circle{3}} \put(29,53.5){\line( 1,2){ 3}} \put(30,53){\line( 1,2){ 3}} \put(31,52.5){\line( 1,2){ 3}} \put(29,70.5){\line( 1,-2){ 3}} \put(62,71){\line( -1,-2){ 3}} \put(61,52.5){\line( -1,2){ 3}} \qbezier(34,58.5)(35.25,61)(38,61) \qbezier(58,58.5)(56.75,61)(54,61) \put(38,61){\line( 1,0){ 16}} \qbezier(33,59)(34.5,62)(38,62) \qbezier(59,65)(57.5,62)(54,62) \put(38,62){\line( 1,0){ 16}} \qbezier(32,59.5)(33.25,62)(32,64.5) \put(28,72.5){\makebox(0,0)[cc]{\hbox{\small$1$}}} \put(28,72.5){\circle{3}} \put(63,73){\makebox(0,0)[cc]{\hbox{\small$2$}}} \put(63,73){\circle{3}} \put(62,50.5){\makebox(0,0)[cc]{\hbox{\small$3$}}} \put(62,50.5){\circle{3}} \end{picture} \caption{\small Flip, or Whitehead move on the shear coordinates $Z_\alpha$. The outer edges can be pending, but the edge with respect to which the morphism is performed must be an internal edge. We also indicate the correspondences between geodesic paths under the flip.} \label{fi:flip} \end{figure} \begin{prop} {\rm \cite{ChF}}\label{propcase} Setting $\phi (Z)={\rm log}(1+e^Z)$ and adopting the notation of Fig.~\ref{fi:flip} for shear coordinates of nearby edges, the effect of a Whitehead move is \begin{equation} W_Z\,:\ (A,B,C,D,Z)\to (A+\phi(Z), B-\phi(-Z), C+\phi(Z), D-\phi(-Z), -Z) \label{abc} \end{equation} In the various cases where the edges are not distinct and identifying an edge with its shear coordinate in the obvious notation we have: if $A=C$, then $A'=A+2\phi(Z)$; if $B=D$, then $B'=B-2\phi(-Z)$; if $A=B$ (or $C=D$), then $A'=A+Z$ (or $C'=C+Z$); if $A=D$ (or $B=C$), then $A'=A+Z$ (or $B'=B+Z$). Any subset of edges $A$, $B$, $C$, and $D$ can be pending edges of the graph. \end{prop} We have the lemma establishing the properties of invariance w.r.t. the flip morphisms~\cite{ChF}. \begin{lm} \label{lem-abc} Transformation~{\rm(\ref{abc})} preserves the traces of products over paths {\rm(\ref{G})} (the geodesic functions) and transformation~{\rm(\ref{abc})} simultaneously preserves Poisson structure {\rm(\ref{WP-PB})} on the shear coordinates. \end{lm} \subsubsection{Whitehead moves on pending edges}\label{sss:pending} Choosing other representatives of the orbifold points in the Poincar\'e disc, we obtain different fundamental domains with different cyclic ordering of the (preimages) of the orbifold points $s_i$ $(i=1,\dots, |\delta_k|)$ possibly with transferring orbifold points from one set $\delta_k$ to another set $\delta_{k'}$. Analogously to the mutation in Fig.~\ref{fi:p-gone}, flipping the pending edge corresponds to choosing another fundamental domain, as shown in Fig.~\ref{fi:interchange-p-dual}. We take there $e^Z$ given by formula (\ref{eZ}) and $e^{Y_{1,2}}$ and $e^{\widetilde{Y_{1,2}}}$ given by the standard cross-ratio relations, for example, $$ e^{Y_2}=\frac{\left(1-e^{i\phi}\right)\left(e^{ib}-e^{2\pi i/p}\right)} {\left(e^{i\phi}-e^{ib}\right)\left(1-e^{2\pi i/p}\right)},\qquad e^{\widetilde{Y_2}}=\frac{\left(e^{-2\pi i/p+ib}-1\right)\left(e^{-2\pi i/p+i\phi}-e^{i\phi}\right)} {\left(e^{-2\pi i/p+i\phi}-e^{-2\pi i/p+ib}\right)\left(1-e^{i\phi}\right)}. $$ \begin{figure}[tb] {\psset{unit=0.8} \begin{pspicture}(-4,-5)(4,1) \psarc[linewidth=0.5pt](0,0){4}{210}{320} \psarc[linestyle=dashed, linecolor=blue, linewidth=1pt](0,0){1.65}{205}{325} \rput(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-5.65){4}{45}{135}} \rput{25}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.25){1.45}{20}{160}} \rput{-20}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.41){1.86}{25}{155}} \rput{-30}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.14){1.07}{15}{165}} \rput{35}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput{15}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput{-5}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput(0,0){ \psline[linewidth=0.5pt](0.2,-1.2)(0.4,-2.2) \psline[linewidth=0.5pt](-0.3,-1.25)(-0.1,-2.25) \psline[linewidth=0.5pt](1.6,-2.8)(0.4,-2.2) \psline[linewidth=0.5pt](-0.8,-2.95)(-0.1,-2.25) \psline[linewidth=0.5pt](0.2,-2.5)(1.2,-3) \psline[linewidth=0.5pt](0.2,-2.5)(-0.35,-3.05) } \rput(-0.9,-2.5){\makebox(0,0)[cc]{$Y_1$}} \rput(1.1,-2.1){\makebox(0,0)[cc]{$Y_2$}} \rput(0,-.9){\makebox(0,0)[cc]{$Z$}} \rput(1.6,-3.9){\makebox(0,0)[lt]{$e^{ib}$}} \rput(3,-3){\makebox(0,0)[lt]{$e^{2\pi i/p}$}} \rput(-3,-3){\makebox(0,0)[rt]{$1$}} \rput(0.2,-4.2){\makebox(0,0)[ct]{$e^{i\phi}$}} \rput(-1.2,-4.1){\makebox(0,0)[ct]{$e^{ia}$}} \end{pspicture} } {\psset{unit=0.8} \begin{pspicture}(2,-5)(-5,1) \rput(-6.5,-1){\psline[linewidth=2pt]{->}(0,0)(0.5,0)} \psarc[linewidth=0.5pt](0,0){4}{165}{280} \psarc[linestyle=dashed, linecolor=blue, linewidth=1pt](0,0){1.65}{160}{285} \rput{-40}(0,0){\psarc[linecolor=red, linewidth=0.5pt](0,-5.65){4}{45}{135}} \rput{-65}(0,0){\psarc[linecolor=red, linewidth=0.5pt](0,-4.25){1.45}{20}{160}} \rput{-20}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.41){1.86}{25}{155}} \rput{-30}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.14){1.07}{15}{165}} \rput{-55}(0,0){\psarc[linecolor=red, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput{-75}(0,0){\psarc[linecolor=red, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput{-5}(0,0){\psarc[linecolor=green, linewidth=0.5pt](0,-4.06){.7}{10}{170}} \rput(-3,-3){\makebox(0,0)[rt]{$1$}} \rput(0,-4.2){\makebox(0,0)[lt]{$e^{i\phi}$}} \rput(-1.2,-4.1){\makebox(0,0)[ct]{$e^{ia}$}} \rput(-3.9,-1.6){\makebox(0,0)[rt]{$e^{-\frac{2\pi i}{p}+ib}$}} \rput(-4.2,0){\makebox(0,0)[rt]{$e^{-\frac{2\pi i}{p}+i\phi}$}} \rput{-35}(0,0){ \psline[linewidth=0.5pt](-0.2,-1.2)(-0.4,-2.2) \psline[linewidth=0.5pt](0.3,-1.25)(0.1,-2.25) \psline[linewidth=0.5pt](-1.6,-2.8)(-0.4,-2.2) \psline[linewidth=0.5pt](0.6,-2.75)(0.1,-2.25) \psline[linewidth=0.5pt](-0.2,-2.5)(-1.2,-3) \psline[linewidth=0.5pt](-0.2,-2.5)(0.2,-2.9) } \rput(-0.8,-2.1){\makebox(0,0)[cc]{${\widetilde{Y_1}}$}} \rput(-2.2,-1){\makebox(0,0)[cc]{${\widetilde{Y_2}}$}} \rput(-.5,-0.6){\makebox(0,0)[cc]{${\widetilde{Z}}$}} \end{pspicture} } \caption{\small The transformation of dual variables ($h$-lengths) $\{Y_1, Y_2, Z\}\to \{{\widetilde{Y_1}},{\widetilde{Y_2}},{\widetilde{Z}}\}$ described by (\ref{morphism-pending}) with $w=2\cos(\pi/p)$. } \label{fi:interchange-p-dual} \end{figure} \begin{lm} \label{lem-pending1} The transformation~in Fig.~\ref{fi:interchange-p-dual} with $e^Z$ given by (\ref{eZ}) has the form \begin{equation} \label{morphism-pending} \{\tilde Y_1,\tilde Y_2,\tilde Z\}= \{Y_1-\log(1+we^{-Z}+e^{-2Z}),Y_2+\log(1+we^Z+e^{2Z}),-Z\} \end{equation} and is the morphism of the space ${\mathcal T}_{g,s,r}^H$. These morphisms preserve both Poisson structures {\rm(\ref{WP-PB})} and the geodesic functions. In Fig.~\ref{fi:interchange-p-dual} any (or both) of $Y$-variables can be variables of pending edges (the transformation formula is insensitive to it). \end{lm} {\bf Proof.} Verifying the preservation of Poisson relations (\ref{WP-PB}) is simple, whereas for traces over paths we have four cases, and in each of these cases we have the following $2\times2$-{\em matrix} equalities to be verified directly: \begin{eqnarray} X_{Y_2}LX_ZF_p^k(-1)^{k-1}X_ZLX_{Y_1}&=& X_{{\tilde Y}_2}RX_{\tilde Z}F_p^{p-k+1}(-1)^{p-k}X_{\tilde Z}RX_{{\tilde Y}_1},\nonumber\\ X_{Y_1}RX_ZF_p^k(-1)^{k-1}X_ZLX_{Y_1}&=& X_{{\tilde Y}_1}LX_{\tilde Z}F_p^{p-k+1}(-1)^{p-k}X_{\tilde Z}RX_{{\tilde Y}_1},\nonumber\\ X_{Y_2}LX_ZF_p^k(-1)^{k-1}X_ZRX_{Y_2}&=& X_{{\tilde Y}_2}RX_{\tilde Z}F_p^{p-k+1}(-1)^{p-k}X_{\tilde Z}LX_{{\tilde Y}_2}.\nonumber \end{eqnarray} Using flip morphisms in Fig.~\ref{fi:interchange-p-dual} and in formula (\ref{abc}), we establish a morphism between any two algebras corresponding to surfaces of the same genus, same number of boundary components, and same numbers of ${\mathbb Z}_p$-orbifold points of each sort $p$; the distribution of latter into the boundary components as well as the cyclic ordering inside each of the boundary component can be arbitrary. It is a standard tool that if, after a series of morphisms, we come to a graph of the same combinatorial type as the initial one (disregarding labeling of edges but distinguishing between different orbifold types of pending vertices), we associate a {\em mapping class group} operation to this morphism therefore passing from the groupoid of morphisms to the group of modular transformations. \begin{remark}\label{rem-hole} Another way of interpreting transformations (\ref{morphism-pending}) is as follows. We can imitate the above flips/mutations by introducing a new ``hole'' with possibly imaginary perimeter $P$ and considering the following chain of {\em standard} flips: \begin{center} {\psset{unit=0.7} \begin{pspicture}(-2.5,-3)(2.5,3) \pscircle(-1,0){0.5} \psarc[linewidth=1pt](-1,0){1}{15}{345} \pcline[linewidth=1pt](-0.1,0.25)(1,0.25) \pcline[linewidth=1pt](-0.1,-0.25)(1,-0.25) \pcline[linewidth=1pt](1,1.5)(1,0.25) \pcline[linewidth=1pt](1,-1.5)(1,-0.25) \pcline[linewidth=1pt](1.5,-1.5)(1.5,1.5) \rput(-1,1.4){\makebox(0,0){$P$}} \rput(.5,0.7){\makebox(0,0){$Z$}} \rput(1.9,1.2){\makebox(0,0){$X$}} \rput(1.9,-1.2){\makebox(0,0){$Y$}} \end{pspicture} } {\psset{unit=0.7} \begin{pspicture}(-3,-3)(3,3) \rput(-3,0){\makebox(0,0){$\to$}} \psbezier[linewidth=1pt](0,0.5)(-0.5,0)(-0.5,0)(0,-0.5) \psbezier[linewidth=1pt](0,0.5)(0.5,0)(0.5,0)(0,-0.5) \psbezier[linewidth=1pt](-0.25,0.75)(-1,0)(-1,0)(-0.25,-0.75) \psbezier[linewidth=1pt](0.25,0.75)(1,0)(1,0)(0.25,-0.75) \pcline[linewidth=1pt](-0.25,.75)(-0.25,1.5) \pcline[linewidth=1pt](0.25,.75)(0.25,1.5) \pcline[linewidth=1pt](-0.25,-.75)(-0.25,-1.5) \pcline[linewidth=1pt](0.25,-.75)(0.25,-1.5) \rput(-1.7,0){\makebox(0,0){$P+Z$}} \rput(1.5,0){\makebox(0,0){$-Z$}} \rput(2.2,1.2){\makebox(0,0){$X{+}\log(1{+}e^Z)$}} \rput(2.2,-1.2){\makebox(0,0){$Y{-}\log(1{-}e^{-Z})$}} \end{pspicture} } {\psset{unit=0.7} \begin{pspicture}(-3,-3)(3,3) \rput(-2,0){\makebox(0,0){$\to$}} \pscircle(1,0){0.5} \psarc[linewidth=1pt](1,0){1}{-165}{165} \pcline[linewidth=1pt](0.1,0.25)(-1,0.25) \pcline[linewidth=1pt](0.1,-0.25)(-1,-0.25) \pcline[linewidth=1pt](-1,1.5)(-1,0.25) \pcline[linewidth=1pt](-1,-1.5)(-1,-0.25) \pcline[linewidth=1pt](-1.5,-1.5)(-1.5,1.5) \rput(2.3,0){\makebox(0,0){$P$}} \rput(-.3,1.1){\makebox(0,0){$-Z{-}P$}} \rput(1,2){\makebox(0,0){$X{+}\log(1{+}e^Z){+}\log(1{+}e^{P{+}Z})$}} \rput(1,-2){\makebox(0,0){$Y{-}\log(1{-}e^{-Z}){-}\log(1{+}e^{-P{-}Z})$}} \end{pspicture} } \end{center} In this pattern, the variable $Z$ is not {\em completely} defined by the ideal triangle decomposition: there is an arbitrariness in choosing the ``mirror'' triangle (we must just choose it consistently). So, in the above sequence of flips, it is useful to shift the variable and consider $$ {\overline Z}=Z+\frac{P}2. $$ The transformation for the variables $X$ and $Y$ then just becomes (\ref{morphism-pending}), \begin{eqnarray} \left[\begin{array}{l} X \\ Y\\ \overline Z \end{array}\right]&\to& \left[\begin{array}{l} X+\log\bigl[(1+e^{\overline Z-P/2})(1+e^{\overline Z+P/2})\bigr] \\ Y-\log\bigl[(1+e^{-\overline Z+P/2})(1+e^{-\overline Z-P/2})\bigr]\\ -\overline Z \end{array}\right]=\nonumber\\ &=& \left[\begin{array}{l} X+\log\bigl[1+\omega_p e^{\overline Z}+e^{2\overline Z}\bigr] \\ Y-\log\bigl[1+\omega_p e^{-\overline Z}+e^{-2\overline Z}\bigr]\\ -\overline Z \end{array}\right],\nonumber \end{eqnarray} where $\omega_p=e^{P/2}+e^{-P/2}$. Since $P$ is not affected by the above sequence of flips, we can merely erase the corresponding loop and present it as flipping the pending edge, which carries besides the cluster variable $\overline Z$ also the ``index'' variable $\omega_p$, which is not affected by mutations and which is equal to $2\cos(\pi/p)$ in the geometric case. \end{remark} \subsubsection{Changing the spiraling direction} The last mapping class group transformation changes the sign of the hole perimeter: \begin{equation} \label{loopinvert} {\psset{unit=0.7} \begin{pspicture}(-5,-3)(7,1) \pcline[linewidth=1pt](-6,-0.5)(-4,-0.5) \pcline[linewidth=1pt](-6,-1.5)(-4,-1.5) \psbezier[linewidth=1pt](-4,-0.5)(-3,1)(-1,1)(-1,-1) \psbezier[linewidth=1pt](-4,-1.5)(-3,-3)(-1,-3)(-1,-1) \psbezier[linewidth=1pt](-3.2,-1)(-2.4,-0.2)(-2,-0.3)(-2,-1) \psbezier[linewidth=1pt](-3.2,-1)(-2.4,-1.8)(-2,-1.7)(-2,-1) \rput(-5.5,0.2){\makebox(0,0){$Y$}} \rput(-1.5,1){\makebox(0,0){$P$}} \pcline[linewidth=1pt]{<->}(0,-1)(2,-1) \pcline[linewidth=1pt](3,-0.5)(5,-0.5) \pcline[linewidth=1pt](3,-1.5)(5,-1.5) \psbezier[linewidth=1pt](5,-0.5)(6,1)(8,1)(8,-1) \psbezier[linewidth=1pt](5,-1.5)(6,-3)(8,-3)(8,-1) \psbezier[linewidth=1pt](5.8,-1)(6.6,-0.2)(7,-0.3)(7,-1) \psbezier[linewidth=1pt](5.8,-1)(6.6,-1.8)(7,-1.7)(7,-1) \rput(3.5,0.2){\makebox(0,0){$Y+P$}} \rput(7.5,1){\makebox(0,0){$-P$}} \rput(10,0){\makebox(0,0){.}} \pcline[linecolor=red, linewidth=0.5pt]{->}(-6,-.7)(-3.9,-.7) \pcline[linecolor=red, linewidth=0.5pt](-6,-1.3)(-3.9,-1.3) \psbezier[linecolor=red, linewidth=0.5pt](-3.9,-.7)(-3,.7)(-1.3,.8)(-1.3,-1) \psbezier[linecolor=red, linewidth=0.5pt](-3.9,-1.3)(-3,-2.7)(-1.3,-2.8)(-1.3,-1) \pcline[linecolor=red, linewidth=0.5pt]{->}(3,-.7)(5.1,-.7) \pcline[linecolor=red, linewidth=0.5pt](3,-1.3)(5.1,-1.3) \psbezier[linecolor=red, linewidth=0.5pt](5.1,-.7)(6,.7)(7.7,.8)(7.7,-1) \psbezier[linecolor=red, linewidth=0.5pt](5.1,-1.3)(6,-2.7)(7.7,-2.8)(7.7,-1) \end{pspicture} } \end{equation} \begin{lm}\label{lem-spiral} Transformation (\ref{loopinvert}) preserves the Poisson brackets and the set of geodesic functions. \end{lm} {\bf Proof}. The preservation of the Poisson bracket is obvious because the variable $P$ Poisson commutes with all other variables, whereas the preservation of geodesic functions follows from two matrix equalities: \begin{eqnarray} &&X_YLX_PLX_Y=X_{Y+P}LX_{-P}LX_{Y+P}, \nonumber \\ &&X_YRX_PRX_Y=X_{Y+P}RX_{-P}RX_{Y+P}. \nonumber \end{eqnarray} We can therefore enlarge the mapping class group of ${\mathcal T}^H_{g,s,r}$ by adding symmetries between sheets of the $2^s$-ramified covering of the ``genuine'' (nondecorated) Teichm\"uller space ${\mathcal T}_{g,s,r}$. The geometrical meaning of this transformation is clear: we change the direction of spiraling to the hole perimeter line for all lines of the ideal triangle decomposition that spiral to a given hole like in Fig.~\ref{fi:saucer-pan}. We can summarize as follows. \begin{theorem} The whole mapping class group of $\Sigma_{g,s,r}$ is generated by morphisms described by Lemmas~\ref{lem-abc},~\ref{lem-pending1}, and~\ref{lem-spiral}. \end{theorem} \section{Conjectures} Lemma~\ref{cc-prime} shows that generalized transformation $a^2+2\cos(\pi/p)ab+b^2$ appears as a flip in the presence of an orbifold point of order $p$. The generalized cluster algebra constructed in this way is a subalgebra of a bigger standard cluster algebra (maybe of infinite rank) associated with triangulated surface while generalized exchange relation are sequences of standard mutations. Note that the positivity of Laurent polynomials for cluster algebras associated with bordered surfaces is known by ~\cite{MSchW}. This implies the positivity of Laurent polynomials in generalized cluster algebra associated with triangulations of the surface with arbitrary orbifold points. We formulate the following conjecture. \begin{conjecture}\label{generalizedPositivity} If $\rho$ is a reciprocal polynomial with positive coefficients then any cluster variable of a generalized cluster algebra is expressed as a positive Laurent polynomial in the initial cluster. \end{conjecture} We checked by direct inspection that the statement holds for finite type rank 2 cluster algebras. The example above leads to a natural question: {\bf Question: } \emph{Is any generalized cluster algebra a subalgebra of some standard cluster algebra?}
\section{Introduction} A scale-invariant spectrum of gravitational radiation is a key prediction of inflation \cite{Starobinsky:1979ty,Rubakov:1982df}. Measuring the ratio of the amplitude of gravitational radiation to the amplitude of density perturbations, $r$, would be a direct probe of the inflationary energy scale. Such a measurement has eluded observation in the cosmic microwave background (CMB) thus far \cite{Komatsu:2010fb} and represents one of the key goals of future CMB observational missions \cite{:2006uk}. On the other hand, some authors have argued that phase transitions can mimic the scale-invariant inflationary signal \cite{Krauss:1991qu,JonesSmith:2007ne,Fenu:2009qf}. The mechanism through which this is accomplished is not trivial. The phase transition itself is not the source. Rather, energy is deposited into gravitational radiation via the self-ordering of fields as regions of spacetime become causally connected. The process begins with $\mathcal{N}$ scalar fields, $\phi_i$, subject to a temperature dependent potential, $V$. At high temperatures the potential looks quadratic with a single minimum at the origin. This state has $O(\mathcal{N})$ symmetry since the state is symmetric any any rotation in the $\mathcal{N}$-dimensional field space. Once the potential drops below a critical temperature, the field acquires a vacuum expectation value (VEV). This process spontaneously breaks the $O(\mathcal{N)}$ to a $O(\mathcal{N}-1)$ symmetry. The transition can be very fast, with a duration, $\tau$, that is much smaller than the Hubble time, $\tau \ll H^{-1}$. The field configuration of each Hubble volume will lie somewhere in the vacuum state, however each causally disconnected volume will be independent of the others. This process can be quite smooth, and there is no {\sl a priori} reason to believe such a process radiates. However, if the Universe is radiation (and later matter) dominated after the transition, the growth of the comoving Hubble scale continuously brings previously causally disconnected regions into contact. Regions thus acquire field gradients on the scale of the Hubble horizon. These gradients generate anisotropic stresses that source gravitational waves. The growth of the horizon acts as a high-pass filter on the spectrum of gravitational waves, {\sl freezing out} large-wavelength modes until they enter the horizon and become dynamical. As the horizon grows power is distributed at all scales between the physical Horizon size at the time of the transition, $H_c^{-1}$, and the Horizon size at the time of observation. In all previous studies, authors have relied on large-$\mathcal{N}$ approximations in order to calculate the gravitational wave signal. Here we make no such approximation and approach the problem numerically. We place $\mathcal{N}$ scalar fields on a lattice evolving in a Friedmann-Lema\^itre-Robertson-Walker background. Each lattice point is initialized with a field value derived from thermal initial conditions. If we allow the lattice spacing to be $H_0^{-1}$ at the beginning of the simulation each lattice point will settle into an independent position in the vacuum manifold. As the simulation progresses, the fields will evolve and align themselves. This self-ordering produces anisotropic stresses that source gravitational waves. This paper is organized as follows. In Section~\ref{GSB} we will introduce a toy model of spontaneous symmetry breaking. We outline our computational methods in Section~\ref{GWP} and present the spectra we calculate in our simulations. In Section~\ref{discussion} we will comment on the differences between the spectra we predict from self-ordering and the spectrum produced during inflation. \section{Global Symmetry Breaking} \label{GSB} The GUT scale naturally arises from particle physics as the scale at which the electroweak and strong couplings are of the same order of magnitude. It is likely that these forces are unified under some larger symmetry above the GUT scale and that this symmetry is broken as the Universe cools. Nevertheless, the nature of the GUT symmetry and how it is broken remains unknown. The transition could be first-order, in which case bubbles of the broken phase nucleate and coalesce. In this model the phase transition happens very rapidly---the entirety of the universe can end up in a unique state in less than a Hubble time. This process is likely to produce gravitational radiation \cite{Witten:1984rs,Kosowsky:1991ua,Kamionkowski:1993fg,Child:2010} as bubbles collide and coalesce. Conversely, the phase transition could be second-order. In this case the field smoothly transitions to the broken phase as the temperature of the universe drops. If the broken phase is not unique, that is if the vacuum state has some symmetry with respect to the field configuration, the effects of the {\sl existence} of this phase transition can lead to observational effects for many Hubble times. We begin with two assumptions: (1) that the Universe is radiation dominated at the time when the phase transition occurs, and (2) that the energy associated with the fields undergoing the phase transition is some small fraction, $\alpha$, of the total energy density at the time of the transition, $\rho_c$. The total density, at any time, is \begin{equation} \rho = \rho_{\rm rad} + \rho_{\phi} \end{equation} where \begin{equation} \label{tdeppot} \rho_\phi = \sum_i \frac{1}{2}\left(\dot{\phi}^2 + \frac{\left(\nabla \phi_i\right)^2}{a^2}\right) + V(\phi_i,T) \end{equation} and \begin{equation} \rho_{\rm rad} = (1-\alpha)a^{-4}\rho. \end{equation} Since the Universe is necessarily dominated by the radiation energy-density, we will only consider cases where $\alpha \ll 1$ so that the Universe is well described by assuming $H\propto a^{-4}$. The potential in \Deqn{tdeppot} is temperature dependent. To leading order in temperature, \begin{equation} \label{tpot} V(\phi_i,T) = m^2_{\rm eff}(T)\phi^2 + \frac{\lambda}{8}\left(\phi^4+\frac{v^4}{4}\right), \end{equation} where $\phi^2 = \sum_i \phi_i^2$. The temperature dependent effective mass can be parameterized by \begin{equation} m_{\rm eff}^2 = \frac{\lambda v^2}{8} \left(\frac{T}{T_c}-1\right). \end{equation} At temperatures higher than the critical temperature, $T_c$, the effective mass is positive, the potential has a unique minimum at the origin, and this minimum has full $O(\mathcal{N})$ symmetry, and at the origin \begin{equation} \left.m_{\rm eff}^2\right|_{\phi=0}= - \frac{\lambda v^2}{8}. \end{equation} After the phase transition the potential has an $O(\mathcal{N}-1)$ symmetric VEV \begin{equation} \label{potential} V(\phi_i) = V(\phi_i,0) = \frac{\lambda}{8} \left(\phi^2-\frac{v^2}{2}\right)^2, \end{equation} The phase transition occurs at the critical temperature, $T_c$, when the effective mass of the field vanishes. Although the field has a mean value, $\phi=0$, there is a variance associated with this value, \begin{equation} \sigma^2 = \left\langle\phi^2\right\rangle -\left\langle\phi \right\rangle^2 = \left\langle\phi^2\right\rangle, \end{equation} that sets the distribution of field values at the time of the transition. We can assume that at this time, each Hubble volume (sphere of radius $H_c^{-1}$) has a homogeneous field value, drawn from a Gaussian distribution (see Appendix~\ref{thermal}) \begin{equation} P(\phi) = \sqrt{\frac{1}{2\pi \sigma^2}}e^{-\frac{\phi^2}{2\sigma^2}}, \end{equation} and \begin{eqnarray} \nonumber \sigma^2 &=& \frac{T_c}{2\pi^2 }\int_0^\infty \frac{k^2e^{- k^2/H_c^2}}{k^2+m_{\rm eff}^2} dk \\ &=& \frac{H_cT_c}{4\pi^{3/2}}.\\ \nonumber \end{eqnarray} where the second equality comes from setting $m_{\rm eff}=0$. The temperature of the Universe at the beginning of the simulation, $T_c$, is related to the energy density of the Universe at that time, \begin{equation} \rho_c = \frac{\pi^2}{30}g_c T_c^4, \end{equation} where $g_c$ is the number of ultra-relativistic degrees of freedom at the time of the phase transition. We take $g_c = 1000$. The average energy density in the field is well approximated by setting $\phi=0$ in \Deqn{potential}, \begin{equation} \left<\rho_\phi \right> \approx \frac{\lambda v^4}{32}, \end{equation} which is some fraction, $\alpha$, of the total energy density \begin{equation} \frac{\lambda v^4}{32} = \alpha \rho_c = \alpha \frac{3m_{\rm pl}^2}{8\pi} H_c^2. \end{equation} This constrains the value of the VEV, \begin{equation} \label{defofvev} \frac{v^2}{2} = \sqrt{\frac{3\alpha}{\lambda \pi}}H_cm_{\rm pl}. \end{equation} This provides us with a good self-consistency check, namely, that the variance of the fluctuations of the filed are small compared to the VEV, $v/\sqrt{2}$. This ratio is given by \begin{equation} \frac{\sigma^2}{v^2/2} = \frac{T_0}{m_{\rm pl}}\frac{1}{4\pi}\sqrt{\frac{\lambda}{3\alpha}}, \end{equation} which is, in general, less that one if $T_0$ is significantly below $10^{19}\,{\rm GeV}$ and $\alpha$ and $\lambda$ are of similar order. \section{Gravitational Waves} \label{GWP} Evolving scalar fields on a discrete lattice is now a mature field of study. Scalar fields were first introduced to the lattice, in a Cosmological context, by {\sc LatticeEasy}~\cite{Felder:2000hq} and later by {\sc DeFROST} \cite{Frolov:2008hy} to study the non-linear dynamics of preheating after inflation. More recently the authors of \cite{Easther:2010qz} re-framed the question by moving the fundamental description of the fields from configuration space to momentum space. Even more recently the author of \cite{Huang:2011gf} introduced a versatile code that allows the user more control over the integrating scheme. Here we chose to use {\sc LatticeEasy} since we are interested in sub- and super-horizon scales, large lattices with efficient storage and a specific associated potential, \Deqn{potential}. This software natively evolves scalar fields according to the Klein-Gordon equation in an expanding background, \begin{equation} \ddot{\phi}_i + 3H \dot{\phi}_i - \frac{\nabla^2\phi_i}{a^2} + \frac{\partial V(\phi_i)}{\partial \phi} = 0, \end{equation} where we work in units where $c=\hbar=1$. The homogeneous background evolution is determined by \begin{equation} \label{firstfried} H^2 = \frac{8\pi}{3m_{pl}^2}\rho, \end{equation} where $\rho=\rho(t)$ is the homogeneous, average, energy density at time $t$. We couple {\sc LatticeEasy} to a code that evolves the metric perturbation using the methods of~\cite{Easther:2006vd,Easther:2007vj}. Since the lattice realizes the fields at discrete values of time, it is most convenient to perturb the metric, $h_{ij}$, in a synchronous gauge \begin{equation} \label{sync} ds^2 = dt^2 - a^2(t)\left[\delta_{ij} + h_{ij}\right]dx^idx^j. \end{equation} Additionally, the radiative part of $h_{ij}$ obeys the transverse-traceless conditions \begin{equation} h_i^i = 0\,\,\,{\rm and} \,\,\,\,h_{ij,j} =0. \end{equation} The radiative perturbations obey sourced Klein-Gordon equations \begin{equation} \ddot{h}_{ij} + 3H \dot{h}_{ij} - \frac{1}{a^2} \nabla^2 h_{ij} = \frac{16 \pi}{m^2_{\rm pl}} S^{TT}_{ij}, \label{equationofmotion} \end{equation} where the source term, $S^{TT}_{ij}$ is the transverse-traceless projection of the anisotropic stress tensor, \begin{equation} \label{aniso} S_{ij} = T_{ij} - \frac{\eta_{ij}}{3}T. \end{equation} We specify our model \Deqn{potential} and allow {\sc LatticeEasy} to evolve the fields and the scale factor. We can then calculate the source term of \Deqn{equationofmotion}, and evolve the six metric perturbations, $h_{ij}^{TT}$. We can always check our numerical stability by checking to see if the metric perturbations are still transverse-traceless; transverse-traceless metric perturbations require {\sl both} a transverse-traceless source and accurate evolution. At any point during the simulation, we can calculate the power spectrum of gravitational radiation. The stress-energy associated with metric perturbations is \cite{Misner:1974qy}, \begin{equation} T^{\rm gw}_{\mu\nu} = \frac{1}{32\pi} \left\langle h_{ij,\mu}h^{ij}_{\,\,\,,\nu}\right\rangle, \end{equation} where the brackets denote a spatial average over at least a few wavelengths. The $00$ component is the energy density, \begin{equation} \rho_{\rm{gw}} = \frac{t^{\mu}t^{\nu}}{32\pi} \left\langle h_{ij,\mu}h^{ij}_{,\nu}\right\rangle = \frac{1}{32 \pi} \sum_{i,j} \left\langle\dot{h}^2_{ij}\right\rangle, \label{gwdensity} \end{equation} where $t^{\mu} = (1,0,0,0)$. Finally, we can invoke Parseval's theorem (see \cite{Easther:2007vj}) to rewrite \Deqn{gwdensity} as \begin{equation} \rho_{\rm gw} =\frac{1}{32\pi} \frac{1}{V}\sum_{i,j}\int d^3\mathbf{k}\,\, \Bigl|\dot{h}_{ij}(t,\mathbf{k})\Bigr|^2, \label{omega0} \end{equation} where $V$ is the comoving volume over which the spatial average is being performed. We can then write \begin{equation} \frac{d\rho_{\rm gw}}{d\ln k} = \frac{k^3}{32\pi} \frac{1}{V} \sum_{i,j} \int d\Omega\, \Bigl| \dot{h}_{ij}^{\rm TT}(\eta,\mathbf{k}) \Bigr|^2, \label{omega} \end{equation} which can be transferred to present-day amplitude and frequency by ~\cite{Easther:2007vj,Price:2008hq}, \begin{equation} \Omega_{\rm gw,0}h^2 = \Omega_{\rm rad,0}h^2 \Biggl(\frac{g_0}{g_e}\Biggr)^{1/3} \frac{1}{\rho_{\rm tot, e}}\frac{d\rho_{\rm gw,e}}{d\ln k}, \label{omega1} \end{equation} where the $0$ and $e$ subscripts denote quantities defined today and the end of our simulations, respectively. We also keep the convention that $h$ absorbs the uncertainty in the present value of the Hubble parameter, $\Omega_{\rm rad,0}$ is the current fraction of the energy density in the form of radiation, and $\rho_{\rm tot, e}$ is the total energy density at the end of our simulations. The ratio, $g_0/g_{\rm e}$, is the number of degrees of freedom today to the number of degrees of freedom at matter/radiation equality. We approximate $g_0/g_{\rm e}=1/100$. \section{Results} \label{discussion} The first major difference between the {\sl structure} of the gravitational-wave spectrum from self-ordering and that predicted by inflation is the lack of power at high-frequencies. This cut-off feature exists because we only considering larger then Hubble length fluctuations as sources of gravitational waves as there will be no/short lived gradient terms to source the gravitational waves inside the Hubble volumes by second order phase transitions. \begin{figure}[htbp] \centering \includegraphics[width=3in]{ncomparison.eps} \caption{The present-day gravitational wave spectrum from self-ordering. From top to bottom, $\mathcal{N}=2,3,4,5,8,16$.} \label{fig:ncomparison} \end{figure} \begin{figure}[htbp] \centering \includegraphics[width=3in]{SOUcomparison.eps} \caption{The present-day gravitational wave spectrum from self-ordering. From top (rightmost) to bottom (leftmost), $\rho_c^{1/4}=10^{-3}\,m_{\rm pl}$, $\rho_c^{1/4}=10^{-4}\,m_{\rm pl}$, $\rho_c^{1/4}=10^{-5}\,m_{\rm pl}$.} \label{fig:soucomparison} \end{figure} We see in \cite{Easther:2006gt} that the cut-off frequency is related to the Hubble length at the time when the gravitational wave is generated \begin{equation} f_{\rm peak} = 6\times 10^{-10} \frac{k}{\sqrt{m_{\rm pl} H}}\,\rm{Hz}. \end{equation} This, along with and the first Friedmann equation, \begin{equation} H_c = \sqrt{\frac{8\pi}{3}}\frac{\sqrt{\rho_c}}{m_{\rm pl}}, \end{equation} allows us to determine where the cut-off should appear \begin{equation} \label{peakeq} f_{\rm peak} \sim 10^{11} \frac{\rho_c^{1/4}}{m_{\rm pl}}\,{\rm Hz}. \end{equation} For example, with $\rho_c^{1/4} = 10^{-4} m_{\rm pl}$ we expect the cutoff to be at $f = 10^{7}\, {\rm Hz}$ which agrees with the results of our simulation, shown in \Dfig{fig:ncomparison}. The characteristic amplitude and scale invariant nature of the spectrum was expected from the analytical approach of \cite{JonesSmith:2007ne,Krauss:2010df} and \cite{Fenu:2009qf}. It is interesting to note that the spectrum is scale invariant for all the cases in \Dfig{fig:ncomparison}, including those for $\mathcal{N}=2,3$; the analytical methods employed by \cite{Fenu:2009qf} assumed that $\mathcal{N}$ is large. In Fig.~\ref{fig:soucomparison} we observe two important scaling effects. First, we recover the fact that the high-frequency cutoff predicted by Eq.~\ref{peakeq} scales with $\rho^{1/4}$. We also see that the amplitude of the signal is proportional to the energy density of the Universe at the time of the transition, \[\Omega_{\rm gw}h^2 \propto \rho_c.\] This scaling is also suggested predicted in \cite{JonesSmith:2007ne,Krauss:2010df} and \cite{Fenu:2009qf}. A final step is to compare the numerical results here with the analytic arguments of \cite{JonesSmith:2007ne,Krauss:2010df} and \cite{Fenu:2009qf}. In both cases the authors use the model presented here to estimate the gravitational wave signal from reordering. The two sets of authors use slightly different parameterizations of the model; however, all authors arrive at the conclusion that there should be a scale-invariant gravitational wave spectrum from this transition In \cite{JonesSmith:2007ne,Krauss:2010df}, the authors estimate the power in gravitational waves from field reordering to be (Eq.~(10) in \cite{Krauss:2010df}) \begin{equation} \label{krasseq} \Omega^{\rm JKM}_{\rm gw}= \frac{1}{12\pi^3H_0^2}k^2 P(k,\tau_0), \end{equation} where $P(k,\tau)$ is calculated numerically, the Hubble constant today is $H$, $\tau_0$ is conformal time today and we've modified the normalization constant $1/12\pi^3$ to reflect a convention choice in \cite{JonesSmith:2007ne} and a typographical error in \cite{Krauss:2010df}. The strain power, $P$, ends up being a function only of $k\tau$. It peaks around $k\tau \approx 3.7$, although this corresponds to modes that entered the horizon during matter domination. To ensure that we're identifying modes that entered during radiation domination, we choose a mode, $k\approx 10/\tau_{\rm eq}$, \begin{equation} \label{maxpowerkrauss} P\left(\frac{k}{\tau_{\rm eq}},\tau_0\right) = \Omega_{\rm rad} P\left(\frac{k}{\tau_{\rm eq}},\tau_{\rm eq}\right), \end{equation} where we note that gravitational waves scale as a constant fraction of $\Omega_{\rm rad}$ after matter-radiation equality, the subscript ${\rm eq}$ indicates evaluating quantities at the time of radiation-matter equality. We can read off $P(k/\tau_{\rm eq},\tau_{\rm eq})$ from Fig.~1 in \cite{JonesSmith:2007ne}, \begin{equation} P\left(\frac{k}{\tau_{\rm eq}},\tau_0\right) \approx 1000 \Omega_{\rm rad}. \end{equation} So we can estimate the total gravitational wave energy per octave, \begin{equation} \label{estimate2} \Omega^{\rm JKM}_{\rm gw}= \frac{1}{12\pi^3H_0^2}\frac{10^2}{\tau_{\rm eq}^2}a_{\rm eq}^2 P\left(\frac{3.7}{\tau_{\rm eq}},\tau_{\rm eq}\right), \end{equation} where the physical wavevector now is $1/a_{eq}$ larger than the physical wavevector at the time of radiation-matter equality. We can make a crude estimate of the value of conformal time at radiation-matter equality, \begin{equation} \label{conftimenow}\tau_{\rm eq} = \frac{1}{H_{\rm eq}}\int_0^{a_{\rm eq}} da^\prime \left(\Omega_R+\Omega_Ma+\Omega_\Lambda a^4\right)^{-1/2}\sim\frac{1}{50H_0}, \end{equation} where the fractional energy densities come from \cite{Komatsu:2010fb}. Putting \Deqn{maxpowerkrauss} together with Eqs.~(\ref{estimate2},\ref{conftimenow}), we get an estimate, \begin{equation} \label{Jestimate} \Omega^{\rm JKM}_{\rm gw}h^2= \frac{99}{\mathcal{N}}\Omega_{\rm rad}h^2 \left(\frac{v^4}{4\mathcal{N}m_{\rm pl}^4}\right), \\ \end{equation} or, imposing our parameterization of the current-day Hubble constant, setting $\Omega_{\rm rad}h^2\approx 2\times10^{-5}$ and using Eqs.~(\ref{defofvev},\ref{firstfried}) \begin{equation} \Omega^{\rm JKM}_{\rm gw}h^2 = \frac{0.016}{\mathcal{N}} \frac{\alpha}{\lambda}\frac{\rho_c}{m_{\rm pl}^4}. \end{equation} In \cite{Fenu:2009qf}, the authors predict a scale-invariant power spectrum (Eq.~(5.2) of \cite{Fenu:2009qf}) \begin{equation} \label{fenuresult} \Omega^{\rm FFDG}_{\rm gw}h^2 \simeq \frac{511}{\mathcal{N}}\Omega_{\rm rad}h^2 \left(\frac{v}{\sqrt{2}m_{\rm pl}}\right)^4, \end{equation} where we use our definition of $v$ and our parameterization $\Omega_{gw}h^2$. Using Eqs.~(\ref{defofvev},\ref{firstfried}) along with $\Omega_{\rm rad}h^2 \approx 2\times10^{-5}$, the expression in \Deqn{fenuresult} reduces to \begin{equation} \label{Festimate} \Omega^{\rm FFDG}_{\rm gw}h^2 \simeq \frac{0.082}{\mathcal{N}}\frac{\alpha}{\lambda} \frac{\rho_c}{m_{\rm pl}^4}. \end{equation} These two estimates should vary from our simulations by one important factor. In both cases, the Universe is comprised only of the scalar fields. To preserve a radiation-dominated phase during and after the phase transition, we have, inherently, diluted the source by a factor of $\alpha$ which dilutes the analytic estimates Eqs.~(\ref{Jestimate},\ref{Festimate}) by a factor of $\alpha^2$. It is worth pointing out that some of the phase transitions we have simulated result in the production of global topological defects. Specifically, global strings for $\mathcal{N}=2$, global monopoles $\mathcal{N}=3$, and global textures for $\mathcal{N}>3$. Surprisingly, we find that the gravitational radiation produced is consistent with the large $\mathcal{N}$ approximation even for low values of $\mathcal{N}$, where the approximation is not valid (see our analytic estimates above). The gravitational wave backgrounds produced by global strings and monopoles are larger than those produced by textures. We will investigate these cases in more detail in a future publication. \section{Discussion} Phase transitions at high energies are a generic consequence of (almost) all models of high energy physics. Since there is no unique model of physics at this scale, we are forced to look for generic observational consequences at this scale. Second order phase transitions will not produce gravitational radiation over one Hubble time, yet the reordering of the fields that mediate the transition could produce characteristic gravitational radiation over a wide range of scales. Although this signature could be misinterpreted as the gravitational radiation from primordial quantum fluctuations, it might be possible to distinguish the two at very high frequencies. Such a high-frequency detection would also carry information about the energy scale at which the phase transition occurred. \begin{table} \caption{\label{tab:specN} Spectral amplitudes as a function of number of fields for simulations with $\left(\rho_c^{1/4}=10^{-4}m_{pl}, \alpha = \lambda = 0.1 \right)$. The numerical values, $\Omega_{\rm gw}h^2$, are an average value taken from the simulations, while the values in the second two columns are obtained from \Deqn{Jestimate} or \Deqn{Festimate}.} \begin{tabular}{cccc} \hline \hline $\mathcal{N}$ & $\Omega_{\rm gw} h^2$ & $\alpha^2 \Omega^{\rm JKM}_{\rm gw} h^2$ & $\alpha^2 \Omega^{\rm FFDG}_{\rm gw}h^2$ \\ \hline 2 & $1.0\times10^{-18}$ &$9.0\times10^{-21}$ & $4.1\times10^{-20}$ \\ 4 & $3.8\times10^{-20}$ & $4.0\times10^{-21}$& $2.1\times10^{-20}$ \\ 8 & $8.3\times10^{-21}$ & $2.0\times10^{-21}$& $1.0\times10^{-20}$ \\ 16 & $3.1\times10^{-21}$ & $1.0\times10^{-21}$& $5.1\times10^{-21}$ \\ \end{tabular} \end{table} \begin{table} \caption{\label{tab:specr} Spectral amplitudes as a function of $\rho_c$ simulations with $\left(\mathcal{N}=4, \alpha = \lambda = 0.1 \right)$. The numerical values, $\Omega_{\rm gw}h^2$, are an average value taken from the simulations, while the values in the second two columns are obtained from \Deqn{Jestimate} or \Deqn{Festimate}.} \begin{tabular}{cccc} \hline \hline $\rho_c^{1/4}(m_{pl})$ & $\Omega_{\rm gw} h^2$& $\alpha^2 \Omega^{\rm JKM}_{\rm gw} h^2$ & $\alpha^2\Omega^{\rm FFDG}_{\rm gw} h^2$ \\ \hline $10^{-3}$ & $4.7\times10^{-16}$ & $4.0\times10^{-17}$ & $2.1\times10^{-16}$ \\ $10^{-4}$ & $3.8\times10^{-20}$ & $4.0\times10^{-21}$& $2.1\times10^{-20}$ \\ $10^{-5}$ & $4.0\times10^{-24}$ & $4.0\times10^{-25}$& $2.1\times10^{-24}$ \\ \end{tabular} \end{table} We considered the phenomenological model of \cite{JonesSmith:2007ne,Krauss:2010df} and \cite{Fenu:2009qf} in which an $\mathcal{O(N)}$ symmetric false vacuum is dynamically broken into a $\mathcal{O}\left(\mathcal{N} - 1 \right)$ true vacuum, we find this produces a scale invariant gravitational wave spectrum whose amplitude depends inversely on number of fields. These results are summarized in Table~\ref{tab:specN}. Our numerical results suggest that the large $\mathcal{N}$ is not needed to make a scale invariant spectrum. Since the results of \cite{JonesSmith:2007ne,Krauss:2010df,Fenu:2009qf} are derived using a large $\mathcal{N}$ approximation for the amplitude, one does not expect these estimates to be a perfect estimator of the amplitude of the gravitational waves in the low-$\mathcal{N}$ limit. Additionally, the results for varying values of $\rho_c$ are given in Table~\ref{tab:specr}. We find that our simulations differ from analytic estimates by only a small amount and are more consistent at large $\mathcal{N}$, consistent with the fact that analytic methods are derived from a large $ \mathcal{N}$ expansion. \section{Acknowledgments} We thank Latham Boyle, Andrew Tolley, Harsh Mathur and Katherine Jones-Smith for useful discussions. JTG is supported by the National Science Foundation, PHY-1068080, and a Cottrell College Science Award from the Research Corporation. The work of XS and BV is supported by National Science Foundation grants PHY-0970074, PHY-0955929, and PHY-0758155, and the University of Wisconsin--Milwaukee Research Growth Initiative. Research at the Perimeter Institute for Theoretical Physics is supported by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Research \& Innovation. JTG would also like to thank the University of Wisconsin-Milwaukee for its generous hospitality while some of this research was completed. LRP would like to thank the Perimeter Institute for Theoretical Physics for their hospitality while some of this work was being completed there.
\section{Introduction} \subsection{The classical story}\label{subsec-classical-story} Schur functions play a central role in the beautiful circle of ideas around symmetric functions, symmetric and general linear group representations, and linear enumerative geometry. On the one hand, the Schur functions $s_\lambda$ corresponding to partitions $\lambda$ of $k$ give an integral orthonormal basis of the algebra of symmetric functions $\mathrm{\Lambda}$ in degree $k$. Under the Frobenius correspondence (see, for instance, \cite[Section 7.18]{Stanley2}) \begin{equation*} \bigoplus_{n\geq0}K_0({\mathbbm C}[S_n])\cong\mathrm{\Lambda}, \end{equation*} the Schur function $s_\lambda$ corresponds to the irreducible representation $L_\lambda$ associated to the partition $\lambda$. Recall that the product on the symmetric group side is given as follows. If $V$ is a representation of $S_k$ and $W$ is a representation of $S_\ell$, then \begin{equation} [V]\cdot[W]=[{\mathrm{Ind}}_{S_k\times S_\ell}^{S_{k+\ell}}(V\otimes W)]. \end{equation} On the other hand, the ring $\mathrm{\Lambda}_n$ obtained by setting $s_{(1^{n+1})}=s_{(1^{n+2})}=\ldots=0$ can be viewed as the character ring for polynomial representations of $GL_n({\mathbbm C})$. Under this correspondence, the image of $s_\lambda$ in $\mathrm{\Lambda}_n$ equals the character of the irreducible representation $V_\lambda$ associated to $\lambda$. This time the product is the usual product of characters, that is, $s_\mu s_\nu$ is the character of $V_\mu\otimes V_\nu$. Either of these two descriptions immediately implies that the structure coefficients of a product \begin{equation}\label{eqn-intro-lrc} s_\mu s_\nu=\sum_\lambda c_{\mu\nu}^\lambda s_\lambda \end{equation} are all non-negative integers; they are called \textit{Littlewood-Richardson coefficients}. The determination of these coefficients is a classical and important problem. By the above, $c_{\mu\nu}^\lambda$ has interpretations as \begin{enumerate} \item the multiplicity of $L_\lambda$ in ${\mathrm{Ind}}_{S_k\times S_\ell}^{S_{k\times\ell}}(L_\mu\otimes L_\nu)$, \item the multiplicity of $V_\lambda$ in $V_\mu\otimes V_\nu$, and \item the dimension of the space of $GL_n({\mathbbm C})$-invariant vectors in $V_\mu\otimes V_\nu\otimes V_\lambda^*$. \end{enumerate} \noindent The combinatorics of Young tableaux become an important tool at this point: partitions correspond to Young diagrams, and the coefficients $c_{\mu\nu}^\lambda$ can be computed by counting certain skew tableaux. We review this Littlewood-Richardson rule in Section \ref{subsec-even-lr}; more complete expositions can be found in \cite[Chapter 5]{Fulton} and \cite[Appendix A1]{Stanley2}. Schur functions also arise in linear enumerative geometry. The cohomology ring of the Grassmannian $\mathrm{Gr}(k,n)$ of complex $k$-planes in ${\mathbbm C}^n$ is isomorphic to the quotient of $\mathrm{\Lambda}_k$ by the ideal generated by all complete symmetric functions $h_m$ for $m>n-k$. Geometrically, elementary functions $e_m$ correspond to the Chern classes of the tautological $k$-plane bundle and complete functions $h_m$ correspond (up to sign) to the Chern classes of a complementary $(n-k)$-plane bundle. Under this identification, the image of the Schur function $s_\lambda$ in $H^*(\mathrm{Gr}(k,n))$ equals zero unless $\lambda_1\leq n-k$ and $\ell(\lambda)\leq k$. The nonzero Schur functions give a basis of the cohomology ring, and $s_\lambda$ is Poincar\'{e} dual to the class of a corresponding \textit{Schubert variety}. Relative to a fixed full flag $0=V_0\subset V_1\cdots\subset V_n={\mathbbm C}^n$, a point $W\in\mathrm{Gr}(k,n)$ is in the Schubert variety $\overline{C_\lambda}$ for the partition $\lambda$ if and only if \begin{equation*} \dim(W\cap V_i)\geq i+\lambda_i-(n-k) \end{equation*} for $1\leq i\leq n$. By \eqref{eqn-intro-lrc}, then, $c_{\mu\nu}^\lambda$ is the coefficient of $[\overline{C_\lambda}]$ in the cup product $[\overline{C_\mu}]\cdot[\overline{C_\nu}]$. Therefore we have a geometric interpretation of $c_{\mu\nu}^\lambda$ as \begin{enumerate} \item[(4)] the multiplicity of $[\overline{C_\lambda}]$ in the cup product $[\overline{C_\mu}]\cdot[\overline{C_\nu}]$. \end{enumerate} When $|\mu|+|\nu|=\dim(\mathrm{Gr}(k,n))=k(n-k)$, the coefficient $c_{\mu\nu}^{(n-k)^k}$ is the (finite!) number of $k$-planes satisfying the dimension criteria of both $\overline{C_\mu}$ and $\overline{C_\nu}$. For example, one can use this analysis to compute the number of lines which intersect five general 3-planes in $\mathbb{P}^6=\mathbb{P}({\mathbbm C}^7)$. All of the above is to say that Schur functions provide important insight into areas of geometry, algebra, and representation theory. The goal of this paper is to provide a small first step in this direction in the \textit{odd} setting, in a sense we will now describe. \subsection{Outline of this paper} An ongoing project initiated in joint work with Khovanov and Lauda is an attempt to give a construction of the Ozsv\'{a}th-Rasmussen-Szab\'{o} odd Khovanov homology \cite{ORS} via a 2-representation-theoretic approach analogous to Webster's construction of (even) Khovanov homology \cite{Web}, \cite{Web2}. As a byproduct of this investigation, an ``odd'' analogue $\mathrm{O}\Lambda$ of the algebra $\mathrm{\Lambda}$ was defined and found to admit signed analogues of many of the combinatorial properties of $\mathrm{\Lambda}$ \cite{EK}, \cite{EKL}. We gave two different candidates for elements playing the role analogous to that of Schur functions. The present work gives a third candidate, proves all three definitions are equivalent, and proves an odd analogue of the Littlewood-Richardson rule. Section \ref{subsec-odd-symmetric} reviews the definitions and basic properties of odd symmetric functions and odd Schur functions; the short Section \ref{subsec-boxterpretations} gives a convenient notation for the many signs which arise in the odd setting. Section \ref{subsec-odd-plactic} contains the new definition of odd Schur functions, and Section \ref{subsec-comparison} proves the following. \begin{unthm} The three definitions of odd Schur functions all coincide.\end{unthm} \noindent In Section \ref{subsec-even-lr} we review the even Littlewood-Richardson rule, and in Section \ref{subsec-odd-lr} we formulate and prove an odd analogue. \begin{unthm} The odd Littlewood-Richardson coefficient $c_{\mu\nu}^\lambda$ is a signed count of semistandard skew tableaux $S$ of shape $\lambda/\mu$ and content $\nu$ such that the row word of $S$ is Yamanouchi.\end{unthm} \noindent Finally, in Section \ref{subsec-hives}, we re-cast the odd Littlewood-Richardson rule in the language of Knutson-Tao hives. \subsection{Acknowledgments} The author thanks Mikhail Khovanov for many helpful conversations and suggestions. For the duration of this work, the author was supported by an NSF Graduate Research Fellowship. \section{Review of odd symmetric functions} \subsection{Odd symmetric polynomials}\label{subsec-odd-symmetric} Now we will briefly review the constructions of \cite{EK}, \cite{EKL}. Let $n\geq1$ and let \begin{equation*} \mathrm{OPol}_n={\mathbbm Z}\langle x_1,\ldots,x_n\rangle/(x_ix_j+x_jx_i\text{ if }i\neq j) \end{equation*} be the ring of skew polynomials in $n$ variables. The \textit{ascend-sorting} of a monomial $x_{i_1}\cdots x_{i_r}$ in $\mathrm{OPol}_n$ is defined to be the monomial obtained by sorting the subscripts $i_1,\ldots,i_r$ into non-decreasing order without introducing any sign. For example, the ascend-sorting of $x_1x_3x_2$ is $x_1x_2x_3$. We will sometimes write $:X:$ for the ascend-sorting of a monomial $X$. Although the generators $x_1,\ldots,x_n$ pairwise supercommute, the algebra $\mathrm{OPol}_n$ is not supercommutative: \begin{equation*} (x_1+x_2)x_1=x_1(x_1-x_2). \end{equation*} The algebra $\mathrm{OPol}_n$ is finite dimensional in each degree and has no zero divisors, but it does not have unique factorization: \begin{equation*} (x_1-x_2)^2=x_1^2+x_2^2=(x_1+x_2)^2. \end{equation*} Have the symmetric group $S_n$ act on the free algebra ${\mathbbm Z}\langle x_1,\ldots,x_n\rangle$ as \begin{equation}\label{eqn-symm-action} s_i(x_j)=\begin{cases}-x_{i+1}&\text{if }j=i,\\-x_i&\text{if }j=i+1,\\-x_j&\text{if }j\neq i,i+1,\end{cases} \end{equation} where $s_i$ is the transposition of $i$ and $i+1$ and the action of $s_i$ is a ring endomorphism. For $i=1,\ldots,n-1$, the $i$-th \textit{odd divided difference operator} is the ${\mathbbm Z}$-linear map $\partial_i:{\mathbbm Z}\langle x_1,\ldots,x_n\rangle\to{\mathbbm Z}\langle x_1,\ldots,x_n\rangle$ defined by \begin{equation}\begin{split} &\partial_i(x_j)=\begin{cases}1&\text{if }j=i,i+1,\\0&\text{otherwise,}\end{cases}\\ &\partial_i(fg)=\partial_i(f)g+s_i(f)\partial_i(g). \end{split}\end{equation} The second line is called the Leibniz rule. It is easy to check that \begin{equation*} \partial_i(x_jx_k+x_kx_j)=0 \end{equation*} for all $j,k$, so we can consider $\partial_i$ as an operator on $\mathrm{OPol}_n$. On $\mathrm{OPol}_n$, the kernel and image of $\partial_i$ coincide; by contrast with the even case, however, these do not equal the space of invariants or anti-invariants of the action of $s_i$. \begin{defn}The ring of \textit{odd symmetric polynomials} is the subring \begin{equation} \mathrm{O}\Lambda_n=\bigcap_{i=1}^{n-1}\ker(\partial_i)\subset\mathrm{OPol}_n. \end{equation}\end{defn} We endow $\mathrm{O}\Lambda_n$ with a ${\mathbbm Z}$- and a ${\mathbbm Z}/2$-grading, whose degree functions we denote by ${\rm deg}_{\mathbbm Z}$ and ${\rm deg}_s$ respectively. Write ${\rm deg}(f)=({\rm deg}_{\mathbbm Z}(f),{\rm deg}_s(f))$ for short. The degree of each $x_i$ is defined to be \begin{equation*} {\rm deg}(x_i)=(2,1). \end{equation*} We will consider $\mathrm{O}\Lambda_n$ as a superalgebra via the ${\mathbbm Z}/2$-grading. In particular, the product structure on $\mathrm{O}\Lambda_n\otimes\mathrm{O}\Lambda_n$ is, on homogeneous elements, \begin{equation*} (f\otimes g)(f'\otimes g')=(-1)^{{\rm deg}_s(g){\rm deg}_s(f')}(ff')\otimes(gg'). \end{equation*} \vspace{0.07in} For $i=1,\ldots,n$, let $\widetilde{x}_i=(-1)^{i-1}x_i$. We define elements of $\mathrm{OPol}_n$ for each $k\geq1$, \begin{equation*}\begin{split} &e_k=\sum_{1\leq i_1<\ldots<i_k\leq n}\widetilde{x}_{i_1}\cdots\widetilde{x}_{i_k},\\ &h_k=\sum_{1\leq i_1\leq\ldots\leq i_k\leq n}\widetilde{x}_{i_1}\cdots\widetilde{x}_{i_k}. \end{split}\end{equation*} Set $e_0=h_0=1$ and $e_k=h_k=0$ for $k<0$. If $k>n$, then $e_k=0$. Both these families of skew polynomials are odd symmetric, and they satisfy the relation \begin{equation*} \sum_{k=0}^\ell(-1)^{\frac{1}{2}k(k+1)}e_kh_{\ell-k}=0\qquad\text{if }\ell\geq1. \end{equation*} They also satisfy the \textit{odd defining relations}, \begin{equation}\label{eqn-ODRs}\begin{split} &e_ae_b=e_be_a\quad\text{if }a+b\text{ is even},\\ &e_ae_b+(-1)^ae_be_a=(-1)^ae_{a+1}e_{b-1}+e_{b-1}e_{a+1}\quad\text{if }a+b\text{ if odd}. \end{split}\end{equation} The relations \eqref{eqn-ODRs} also hold if all $e$'s are replaced by $h$'s. This next proposition follows from Section 2.1 of \cite{EKL}. \begin{prop}\label{prop-odd-properties} The algebra $\mathrm{O}\Lambda$ has a presentation by generators $e_0=1,e_1,e_2,\ldots$ and relations \eqref{eqn-ODRs}. A basis for $\mathrm{O}\Lambda$ in ${\mathbbm Z}$-degree $2k$ is given by all products $e_{\lambda_1}\cdots e_{\lambda_r}$ with $\lambda_1\geq\ldots\lambda_r\geq1$, $|\lambda_1|+\ldots+|\lambda_r|=k$. The same is true if all $e$'s are replaced by $h$'s.\end{prop} The products $e_\lambda=e_{\lambda_1}\cdots e_{\lambda_r}$ and $h_\lambda=h_{\lambda_1}\cdots h_{\lambda_r}$ with $\lambda_1\geq\ldots\geq\lambda_r$ are called \textit{odd elementary symmetric functions} and \textit{odd complete symmetric functions}, respectively. The sets of elementary and complete functions in ${\mathbbm Z}$-degree $2k$ are naturally indexed by all partitions of $k$. \vspace{0.07in} There are maps \begin{equation*}\begin{split} &\mathrm{OPol}_{n+1}\to\mathrm{OPol}_n\\ &x_i\mapsto x_i\text{ if }1\leq i\leq n,\\ &x_{n+1}\mapsto0. \end{split}\end{equation*} These induce maps $\mathrm{O}\Lambda_{n+1}\to\mathrm{O}\Lambda_n$ which send $h_k\mapsto h_k$ and $e_k\mapsto e_k$ for all $k$. The inverse limit in the category of graded rings of the resulting system is called the ring of \textit{odd symmetric functions}, \begin{equation*} \mathrm{O}\Lambda=\underleftarrow{\lim}\mathrm{O}\Lambda_n. \end{equation*} The ring $\mathrm{O}\Lambda$ can be given the structure of a Hopf superalgebra \cite{EK}. The coproduct is \begin{equation*} \Delta(h_k)=\sum_{i+j=k}h_i\otimes h_j,\qquad\Delta(e_k)=\sum_{i+j=k}e_i\otimes e_j. \end{equation*} As a Hopf superalgebra, $\mathrm{O}\Lambda$ is neither commutative nor cocommutative. \vspace{0.07in} As a result of Proposition \ref{prop-odd-properties}, we can define certain symmetries of $\mathrm{O}\Lambda$, \begin{equation*}\begin{split} &\psi_1(h_k)=e_k\\ &\qquad\text{bialgebra automorphism (not an involution),}\\ &\psi_2(h_k)=(-1)^{\frac{1}{2}k(k+1)}h_k\\ &\qquad\text{algebra involution (not a coalgebra homomorphism),}\\ &\psi_3(h_k)=h_k\\ &\qquad\text{algebra anti-involution (not a coalgebra homomorphism).} \end{split}\end{equation*} The map $\psi_1\psi_2$ is an involution, and both $\psi_1,\psi_2$ commute with $\psi_3$. The antipode of the Hopf superalgebra structure on $\mathrm{O}\Lambda$ is $S=\psi_1\psi_2\psi_3$. For a partition $\lambda=(\lambda_1,\ldots,\lambda_r)$, define \begin{equation*} \epsilon_\lambda=(-1)^{\sum_j\frac{1}{2}\lambda^T_j(\lambda^T_j-1)}, \end{equation*} where $\lambda^T$ is the transpose (or dual, or conjugate) partition to $\lambda$. That is, $\lambda^T_j$ is the height of the $j$-th column of $\lambda$. The involution $\psi_1\psi_2$ exchanges the elementary and complete bases up to sign, \begin{equation}\label{eqn-psi12-eh} \psi_1\psi_2(e_\lambda)=(-1)^{|\lambda|}\epsilon_{\lambda^T}h_\lambda,\qquad\psi_1\psi_2(h_\lambda)=(-1)^{|\lambda|}\epsilon_{\lambda^T}e_\lambda. \end{equation} \vspace{0.07in} In \cite{EK}, an odd analogue of the Schur functions was given as follows. From here on, we will identify partitions and Young diagrams without comment. Recall that a \textit{Young tableau} $T$ on a partition $\lambda$ with entries in an ordered alphabet $A$ is an assignment of an element of $A$ to each box of $\lambda$. When considering $\mathrm{O}\Lambda$ we will take $A={\mathbbm Z}_{>0}$ and when considering $\mathrm{O}\Lambda_n$ we will take $A=\lbrace1,\ldots,n\rbrace$. A tableau $T$ is called \textit{semistandard} if its entries are non-decreasing in rows (left to right) and strictly increasing in columns (top to bottom). The \textit{row word} $w_r(T)$ of a tableau $T$ is the word in the alphabet $A$ obtained by reading the entries of $T$ from left to right, bottom to top. For example, \begin{equation*} T=\young(112,23)\text{ is a semistandard Young tableau of shape }(3,2),\text{ and }w_r(T)=23112. \end{equation*} The \textit{content} of a tableau $T$ is the tuple $(a_1,\ldots,a_r)$, where $a_i$ is the number of entries of $T$ equal to $i$. For each partition $\lambda$, there is a unique Young tableau $T_\lambda$ of shape and content both equal to $\lambda$. For example, \begin{equation*} T_{(21)}=\young(11,2),\qquad T_{(311)}=\young(111,2,3). \end{equation*} If $T$ is a Young tableau, write $\mathrm{sh}(T)$ for its shape and $\mathrm{ct}(T)$ for its content. Let $SSYT(\lambda)$ be the set of semistandard Young tableaux of shape $\lambda$ and let $SSYT(\lambda,\mu)$ be the set of semistandard Young tableaux of shape $\lambda$ and content $\mu$. In order to use tableaux in the odd setting, define the sign of a tableau $T$, denoted $\text{sign}(T)$, to be the sign of the minimal length permutation which sorts $w_r(T)$ into non-decreasing order. For instance, with $T$ as above, \begin{equation*} \text{sign}(T)=(-1)^5=-1,\qquad\text{sign}(T_{(21)})=(-1)^2=1,\qquad\text{sign}(T_{(311)})=(-1)^7=-1. \end{equation*} For partitions $\lambda,\mu$, the \textit{odd Kostka number} $K_{\lambda\mu}$ is defined to be a signed count of semistandard tableaux of shape $\lambda$ and content $\mu$, \begin{equation} K_{\lambda\mu}=\text{sign}(T_\lambda)\sum_{T\inSSYT(\lambda,\mu)}\text{sign}(T). \end{equation} It is readily verified that as a matrix, $(K_{\lambda\mu})_{|\lambda|=|\mu|=k}$ is lower triangular and unimodular when the partitions basis is ordered lexicographically, so we can define the family of \textit{combinatorial (or Kostka) odd Schur functions} by the change of basis relation \begin{equation} h_\mu=\sum_{\lambda\vdash k}K_{\lambda\mu}s^K_\lambda, \end{equation} where $\mu\vdash k$. The following proposition combines several of the statements from Section 3.3 of \cite{EK}. \begin{prop} The family $\lbrace s_\lambda^K\rbrace_{\lambda\vdash k}$ is an integral unimodular basis for $\mathrm{O}\Lambda$ in ${\mathbbm Z}$-degree $2k$. This family is signed-orthonormal, \begin{equation} (s_\lambda^K,s_\mu^K)=\epsilon_\lambda\delta_{\lambda\mu}, \end{equation} with respect to the bilinear form of \cite{EK}. The involution $\psi_1\psi_2$ and the anti-involution $\psi_3$ act on $s_\lambda$ as \begin{equation}\label{eqn-psi-schur} \psi_1\psi_2(s_\lambda^K)=(-1)^{\ell(w_\lambda)+|\lambda|}s_{\lambda^T}^K,\qquad\psi_3(s_\lambda^K)=\epsilon_\lambda\text{sign}(T_\lambda)s_\lambda^K. \end{equation} Here, $w_\lambda$ is the element of $S_k$ which is combinatorially defined in Proposition 2.14 of \cite{EK} (it is the minimal representative of the unique double coset in $S_{\lambda^T}\backslash S_k/S_\lambda$ which gives a nonzero homomorphism from ${\mathrm{Ind}}_{S_\lambda}^{S_k}(V_\text{triv})$ to ${\mathrm{Ind}}_{S_{\lambda^T}}^{S_k}(V_\text{sign})$) and $\ell$ is the Coxeter length function.\end{prop} In \cite{EKL}, another family of Schur functions was introduced. Let \begin{equation*} \partial_{w_0}=\partial_1(\partial_2\partial_1)\cdots(\partial_{n-1}\cdots\partial_1), \end{equation*} a particular choice of longest odd divided difference operator (odd divided difference operators corresponding to other choices of reduced expression could differ by a factor of $-1$). For a skew polynomial $f$, let $f^{w_0}$ be the result of acting by $w_0$ on $f$ via the action \eqref{eqn-symm-action}. The \textit{odd-symmetrized Schur functions} are defined by \begin{equation} s_\lambda^s=(-1)^{\binom{n}{3}}\left(\partial_{w_0}(x_1^{\lambda_1}\cdots x_n^{\lambda_n}x_1^{n-1}x_2^{n-2}\cdots x_{n-1})\right)^{w_0}. \end{equation} For the motivation behind this name, see Section 2.2 of \cite{EKL}. The goal of Section \ref{subsec-comparison} is to prove that for all $\lambda$, $s_\lambda^s=s_\lambda^K$. \subsection{Boxterpretations}\label{subsec-boxterpretations} It will be convenient to have a systematic way to handle the signs which arise in the odd setting. Most combinatorial quantities arising in these signs can be described by counting certain boxes in some Young diagram or tableau; we call these descriptions ``boxterpretations.'' Here are some which have already arisen: \begin{itemize} \item For a Young diagram $\lambda$, the sum $\sum_j\frac{1}{2}\lambda^T_j(\lambda^T_j-1)$ can be described as: for each box $B$, add the number of boxes directly above $B$. This is the exponent of $-1$ used in defining the sign $\epsilon_\lambda$. \item For a Young diagram $\lambda$, the quantity $\ell(w_\lambda)$ can be described as: for each box $B$, add the number of boxes above and to the right of $B$. \item For a Young tableau $T$: for each box $B$, add the number of boxes above and with smaller entry. Then $\text{sign}(T)$ equals $-1$ raised to this count. \item As a special case of the previous, to obtain $\text{sign}(T_\lambda)$, for each box $B$, count the number of boxes above $B$. \end{itemize} We will use notations adapted from \cite{Fulton} to help condense these descriptions. If a box $B$ of a Young diagram is in a row above that of a box $B'$, we say $B$ is North of $B'$; if $B$ is North of or in the same row as $B'$, we say $B$ is north of $B'$. Likewise for East/east, West/west, and South/south. Write $N(B)$ for the number of boxes North of $B$, $sW(B)$ for the number of boxes southWest (both south and West) of $B$, and so forth. If we are considering a Young tableau and decorate a direction with one of $\lbrace<,\leq,>,\geq\rbrace$, this means to only count boxes with entry lower than (lower than or equal to, greater than, greater than or equal to) $B$. So $E^>(B)$ means the number of boxes East of $B$ with strictly greater entry than $B$. If we evaluate one of these counting functions on a diagram or a tableau, we mean to sum over all boxes, evaluating the function at each. For instance, \begin{equation*} SW^\geq(T)=\sum_{B\in T}\#\lbrace\text{boxes of }T\text{ SouthWest of }B\text{ with entry }\geq\text{ that of }B\rbrace. \end{equation*} One last decoration: $dN$ means directly North, that is, North and neither East nor West (and likewise for the other counts). \begin{example} Let \begin{equation*} T=\young(1122,2334,344,56). \end{equation*} Then $dN(T)=16$, $E^>(T)=28$, and $sW(T)=47$.\end{example} In this notation, we have the following boxterpretations: \begin{equation*} \epsilon_\lambda=(-1)^{dN(\lambda)},\qquad(-1)^{\ell(w_\lambda)}=(-1)^{NE(\lambda)},\qquad\text{sign}(T)=(-1)^{N^<(T)},\qquad\text{sign}(T_\lambda)=(-1)^{N(\lambda)}. \end{equation*} \section{Odd plactic Schur functions} \subsection{Definition and basic properties}\label{subsec-odd-plactic} Let $A=\lbrace a_1,a_2,\ldots\rbrace$ be an ordered alphabet. In practice, we will take $A={\mathbbm Z}_{>0}$ when working with $\mathrm{O}\Lambda$ and $A=\lbrace1,2,\ldots,n\rbrace$ when working with $\mathrm{O}\Lambda_n$. In order to add, multiply, and assign signs to tableaux, we will use the \textit{odd plactic ring} $\mathbb{Z}Pl$, which is the unital ring defined by \begin{equation}\begin{split} \text{generators:}\qquad&A\\ \text{relations:}\qquad&yzx=-yxz\qquad\text{if }x<y\leq z\qquad(K'),\\ &xzy=-zxy\qquad\text{if }x\leq y<z\qquad(K''). \end{split}\end{equation} When we want to emphasize that the alphabet in question is $\lbrace1,2,\ldots,n\rbrace$, we will sometimes write $\mathbb{Z}Pl_n$ instead of $\mathbb{Z}Pl$. The relations $(K'),(K'')$ are called \text{elementary Knuth transformations}. We define a map from the set of semistandard Young tableaux with entries in the alphabet $A$ to the odd plactic ring by \begin{equation}\begin{split} \lbrace\text{SSYTs}\rbrace&\to\mathbb{Z}Pl,\\ T&\mapsto w_r(T). \end{split}\end{equation} Since both the relations $(K'),(K'')$ are transpositions of letters with a minus sign, $\mathbb{Z}Pl_n$ sits as an intermediate quotient between a free algebra and the skew polynomial ring, \begin{equation*} {\mathbbm Z}\langle x_1,\ldots,x_n\rangle\twoheadrightarrow\mathbb{Z}Pl_n\twoheadrightarrow\mathrm{OPol}_n, \end{equation*} where the first map sends $\widetilde{x}_i$ to $i$ and the second map sends $i$ to $\widetilde{x}_i$. If $w$ is a word in $\mathbb{Z}Pl_n$, we will write $\widetilde{x}^w$ for the image of $w$ in $\mathrm{OPol}_n$. In particular, a semistandard Young tableau $T$ is sent to $\widetilde{x}^{w_r(T)}$. The utility of the plactic ring is in large part due to the following remarkable theorem. \begin{thm}[\cite{Fulton}, Section 2.1]\label{thm-word-tableau} Every word is equivalent, via relations $(K')$ and $(K'')$, to the row word $w_r(T)$ of a unique tableau $T$.\end{thm} \noindent Thus the set of all Young tableaux with entries in $A$ forms a basis of $\mathbb{Z}Pl$. We will informally refer to the multiplication of tableaux in the following; what we mean is the multiplication of their row words in $\mathbb{Z}Pl$. In terms of tableaux, the relations $(K')$ and $(K'')$ can be interpreted as ``bumping transformations'': \begin{equation*}\begin{split} (K')\qquad&\young(yz)\cdot\young(x)=-\young(xz,y)\qquad\text{if }x<y\leq z,\\ (K'')\qquad&\young(xz)\cdot\young(y)=-\young(xy,z)\qquad\text{if }x\leq y<z. \end{split}\end{equation*} For a detailed exposition of bumping, see Section 1.1 of \cite{Fulton}. \vspace{0.07in} If a word $w$ is known to be the row word of some tableau, then it is easy to reconstruct the tableau from the word. Since the row entries of a tableau never decrease from left to right and the column entries must always increase from top to bottom, reading the word $w$ from left to right until the first adjacent decreasing pair simply gives the bottom row of the tableau. Then continuing to read until the next adjacent decreasing pair gives the second to bottom row, and so forth. \begin{example} Using ${\mathbbm Z}_{>0}$ as the ordered alphabet, \begin{equation*} w=53422331112\qquad\text{corresponds to}\qquad\young(1112,2233,34,5). \end{equation*} \end{example} \begin{defn} Let $\lambda$ be a partition. Define an element of $\mathbb{Z}Pl_n$ by \begin{equation} \widehat{s}_\lambda=(-1)^{dN(\lambda)+N(\lambda)}\sum_{T\inSSYT(\lambda)}T. \end{equation} Its image in $\mathrm{OPol}_n$, \begin{equation}\label{eqn-defn-s-plactic} s_\lambda^p=(-1)^{dN(\lambda)+N(\lambda)}\sum_{T\inSSYT(\lambda)}\widetilde{x}^{w_r(T)}, \end{equation} is called the \textit{plactic odd Schur function} corresponding to $\lambda$.\end{defn} The label $p$ in the notation for the plactic odd Schur function is to distinguish it from the combinatorial odd Schur functions $s_\lambda^K$ of \cite{EK} and the odd-symmetrized Schur functions $s_\lambda^s$ of \cite{EKL}; once we prove these three objects are equal later in this section, we will drop the extra labels. \begin{rem} For the rest of the paper, we will not always specify whether we are working in $\mathrm{O}\Lambda$ or in $\mathrm{O}\Lambda_n$. Generally speaking, results will hold in $\mathrm{O}\Lambda$. The only change required in passing to $\mathrm{O}\Lambda_n$ is the understanding that certain elements become zero. It is easy to see that $\mathrm{O}\Lambda_n=\mathrm{O}\Lambda/(e_m:m>n)$, and it will follow from Theorem \ref{thm-schur-coincidence} that $s_\lambda=0$ in $\mathrm{O}\Lambda_n$ if and only if $\lambda$ has height greater than $n$. With this understood, the proofs and results in the rest of this paper work in either context.\end{rem} \begin{lem}\label{lem-schur-straight} If $\lambda=(1^k)$, then up to sign, all three Schur functions coincide with the corresponding elementary polynomial: \begin{equation} s_{(1^k)}^K=s_{(1^k)}^p=s_{(1^k)}^s=(-1)^{\frac{1}{2}k(k-1)}e_k. \end{equation} If $\lambda=(k)$, the combinatorial and plactic Schur functions coincide with the corresponding complete polynomial: \begin{equation} s_{(k)}^K=s_{(k)}^p=h_k. \end{equation}\end{lem} The last equation equals $s_{(k)}^s$ too, but we will prove this later. \begin{proof} The equality $s_{(1^k)}^K=(-1)^{\frac{1}{2}k(k-1)}e_k$ follows from Proposition 3.10 of \cite{EK} and the equality $s_{(1^k)}^s=(-1)^{\frac{1}{2}k(k-1)}e_k$ is Lemma 2.25 of \cite{EKL}. Since the row word of a semistandard Young tableau on the shape $(1^k)$ is $i_k\cdots i_1$ for positive integers $i_1<\ldots<i_k$ and it takes $\frac{1}{2}k(k-1)$ transpositions to ascend-sort the monomial $\widetilde{x}_{i_k}\cdots\widetilde{x}_{i_1}$, the equality with $s_{(1^k)}^p$ holds. It is similar but easier to show $s_{(k)}^p=h_k$, and $s_{(k)}^K=h_k$ is obvious because the only semistandard Young tableau of content $(k)$ is a row of 1's.\end{proof} \subsection{Comparison with previous definitions}\label{subsec-comparison} For a partition $\lambda$, let $\frac{i}{\lambda}$ be the Young diagram obtained by removing rows 1 through $i$ from the diagram corresponding to $\lambda$. Similarly, let $\frac{\lambda}{i}$, $i|\lambda$, and $\lambda|i$ be obtained by removing rows $i$ through the bottom, columns $1$ through $i$, and columns $i$ through the rightmost respectively. We say that a skew shape is a \textit{vertical strip} (respectively \textit{horizontal strip}) if no two of its boxes are in the same row (respectively column). We say that a diagram $\mu$ is obtained from $\lambda$ by adding a vertical strip if $\lambda\subset\mu$ and $\mu/\lambda$ is a vertical strip; likewise for horizontal strips. The following proposition was proved in \cite{EKL}. \begin{prop}[Odd Pieri rule, $e$-right odd-symmetrized version]\label{prop-pieri-s} Let $\lambda$ be a partition. Then \begin{equation}\label{eqn-pieri-s} s_\lambda^ss_{(1^k)}^s=\sum_\mu(-1)^{\left|\frac{i_1}{\lambda}\right|+\ldots+\left|\frac{i_k}{\lambda}\right|}s_\mu^s. \end{equation} The sum is over all $\mu$ obtained from $\lambda$ by adding a vertical strip of size $k$, and $i_1,\ldots,i_k$ are the rows of $\lambda$ to which a box was added.\end{prop} The plactic odd Schur functions satisfy the same relation. We prove the horizontal strip variant instead, for simplicity. \begin{prop}[Odd Pieri rule, $h$-right plactic version]\label{prop-pieri-p} Let $\lambda$ be a partition. Then \begin{equation}\label{eqn-pieri-p} (-1)^{NE(\lambda)}s_\lambda^ps_{(k)}^p=\sum_\mu(-1)^{NE(\mu)}(-1)^{\left|i_1|\lambda\right|+\ldots+\left|i_k|\lambda\right|}s_\mu^p. \end{equation} The sum is over all $\mu$ obtained from $\lambda$ by adding a horizontal strip of size $k$, and $i_1,\ldots,i_k$ are the columns of $\lambda$ to which a box was added.\end{prop} \begin{proof} Expanding all Schur functions as odd plactic sums, we want to prove \begin{equation*} (-1)^{dN(\lambda)+N(\lambda)+NE(\lambda)}\widetilde{x}^{w_r(T)}\widetilde{x}^{w_r(V)}=(-1)^{dN(\mu)+N(\mu)+NE(\mu)}(-1)^{\left|i_1|\lambda\right|+\ldots+\left|i_k|\lambda\right|}\widetilde{x}^{w_r(U)} \end{equation*} whenever $T$ is a Young tableau of shape $\lambda$, $V$ is a Young tableau of shape $(k)$, $TV=U$ in the even plactic ring with $\mathrm{sh}(U)=\mu$, and the boxes of $\mu/\lambda$ are in columns $i_1,\ldots,i_k$. This is because mod 2 the even and odd plactic rings are isomorphic (by the obvious map, $w_r(T)\mapsto w_r(T)$), so the set of products $\widetilde{x}^{w_r(T)}\widetilde{x}^{w_r(V)}$ and the set of terms $\widetilde{x}^{w_r(U)}$ which occur on the right-hand side of \eqref{eqn-pieri-p} are in bijection, with $(T,V)$ corresponding to $U$ if and only if $TV=U$ in the even plactic ring. Suppose the leftmost box of $V$ has entry $j$. If that box ends up in column $i$, we claim \begin{equation*} (-1)^{dN(\lambda)+N(\lambda)+NE(\lambda)}\widetilde{x}^{w_r(T)}\widetilde{x}_j=(-1)^{dN(\mu)+N(\mu)+NE(\mu)+\left|i|\lambda\right|}\widetilde{x}^{w_r(U)}. \end{equation*} For any partition $\nu$, $(-1)^{dN(\nu)+N(\nu)+NE(\nu)}=(-1)^{NW(\nu)}$. The new box of $\mu$ is not NorthWest of any other box (since it must be a southeast corner), so the sign discrepancy only counts those boxes NorthWest of the new box. The sign $(-1)^{\left|i|\lambda\right|}$ counts boxes NorthEast of the new box, so the overall sign is $(-1)^{\sum_j(\lambda_j-1)}$. And this is precisely the sign between $\widetilde{x}^{w_r(T)}\widetilde{x}_j$ and $\widetilde{x}^{w_r(U)}$, since bumping a box past a row of length $r$ incurs a sign of $(-1)^{r-1}$. Finally, note that as the boxes of $V$ are added one at a time, the signs cancel telescopically so as to yield the sign of equation \eqref{eqn-pieri-p}.\end{proof} We now have the tools necessary to prove the main result of this section. \begin{thm}\label{thm-schur-coincidence} The three notions of odd Schur function all coincide: for any partition $\lambda$, \begin{equation} s_\lambda^K=s_\lambda^p=s_\lambda^s. \end{equation}\end{thm} \begin{proof} The proof that $s_\lambda^K=s_\lambda^p$ is similar in spirit to the proof of the odd Pieri rule, $h$-right plactic version. By the same sort of analysis as in that proof, one shows that for a partition $\mu=(\mu_1,\ldots,\mu_r)$, \begin{equation*}\begin{split} h_\mu=h_{\mu_1}\cdots h_{\mu_r}&=\sum_{\mathrm{ct}(T)=\mu}(-1)^{NE(T)+NE^<(T)+dN(T)+N(T)}\widetilde{x}^{w_r(T)}\\ &=\sum_\lambda\sum_{T\inSSYT(\lambda,\mu)}(-1)^{NE(\lambda)+NE^<(T)}s_\lambda^p. \end{split}\end{equation*} Since $h_\mu=\sum_\lambda K_{\lambda\mu}s_\lambda^K$ and both $\lbrace h_\mu\rbrace,\lbrace s_\lambda^K\rbrace$ are integral bases of the ring of odd symmetric functions, it suffices to check \begin{equation*} (-1)^{NE(\lambda)+NE^<(T)}=(-1)^{N(\lambda)+N^<(T)} \end{equation*} whenever $T$ is a Young tableau of shape $\lambda$ (the right-hand side is the sign with which $T$ is counted in defining $K_{\lambda\mu}$; see Section \ref{subsec-odd-symmetric}). The sum computing the sign from a particular box $B$ of $T$ involves two types of other box: those Northwest and those NorthEast of $B$. For those NorthEast, the sign is identical. Those Northwest are ignored by the left-hand sign, but the entry of such a box is necessarily less than that of $B$ (since $T$ is semistandard), so the right-hand sign is $+1$. Hence $s_\lambda^K=s_\lambda^p$. We now use the Pieri rules (Propositions \ref{prop-pieri-s}, \ref{prop-pieri-p}) to prove $s_\lambda^p=s_\lambda^s$. For $\lambda=(1^k)$, this is true by Lemma \ref{lem-schur-straight}. Using this as a base case, we now induct on the width of $\lambda$, and within each particular width we induct with respect to the lexicographic order of $\lambda^T$. Since we have shown $s_\lambda^K=s_\lambda^p$, we can apply the involution $\psi_1\psi_2$ to equation \eqref{eqn-pieri-p} to show the plactic Schur functions obey an $e$-right Pieri rule of the same form and signs as the one obeyed by the Schur functions $s_\lambda^K$, \begin{equation}\label{eqn-pieri-p-2} s_\lambda^ps_{(1^k)}^p=\sum_\mu(-1)^{\left|\frac{i_1}{\lambda}\right|+\ldots+\left|\frac{i_k}{\lambda}\right|}s_\mu^p. \end{equation} Let $\lambda$ be a partition of width $\lambda_1=r$. Using the $e$-right Pieri rule, both types of Schur function satisfy \begin{equation*} s_{(\lambda^T_1)}\cdots s_{(\lambda^T_r)}=\sum_\mu\pm s_\mu, \end{equation*} where each $\mu$ has width at most $r$ and is lexicographically greater than or equal to $\lambda$. The coefficient of $s_\lambda$ on the right-hand side is $\pm1$. All the signs $\pm$ are the same for the two types of Schur function, so this allows us to solve for both $s_\lambda^p,s_\lambda^s$ in terms of elementary functions by the same expressions; hence $s_\lambda^p=s_\lambda^s$.\end{proof} \begin{cor} The span of the $\widehat{s}_\lambda$ in $\mathbb{Z}Pl_n$ is a subalgebra isomorphic to $\mathrm{O}\Lambda_n$, and this subalgebra is taken isomorphically onto $\mathrm{O}\Lambda_n\subset\mathrm{OPol}_n$ by the map $w\mapsto\widetilde{x}^w$.\end{cor} For the rest of this paper, we will drop the superscript labels on odd Schur functions. \section{Littlewood-Richardson rules} \subsection{The even Littlewood-Richardson rule}\label{subsec-even-lr} \quad\\ \begin{center}\textit{For this section only, we work in the even setting.}\end{center} \quad\\ Let $\mu,\nu,\lambda$ be partitions. The \textit{Littlewood-Richardson coefficient} $c_{\mu\nu}^\lambda$ is the coefficient of $s_\lambda$ in $s_\mu s_\nu$, \begin{equation*} s_\mu s_\nu=\sum_\lambda c_{\mu\nu}^\lambda s_\lambda. \end{equation*} If $|\mu|+|\nu|\neq|\lambda|$, then $c_{\mu\nu}^\lambda=0$. Schur functions are generating functions for semistandard Young tableaux of a given shape, \begin{equation*} s_\lambda=\sum_{T\inSSYT(\lambda)}x^{w_r(T)}. \end{equation*} It follows that, for any fixed semistandard Young tableau $T_0$ of shape $\lambda$, \begin{equation}\label{eqn-easy-even-lr} c_{\mu\nu}^\lambda=\#\lbrace U\inSSYT(\mu),V\inSSYT(\nu):UV=T_0\rbrace. \end{equation} Here, the product of tableaux is taken in the (even) plactic ring. If $T_0=T_\lambda$, then it is not hard to see that $UV=T_\lambda$ forces $V$ to be $T_\nu$; equation \eqref{eqn-easy-even-lr} becomes \begin{equation}\label{eqn-easy-even-lr-2} c_{\mu\nu}^\lambda=\#\lbrace U\inSSYT(\mu):UT_\nu=T_\lambda\rbrace. \end{equation} Since computing the set on the right-hand side of \eqref{eqn-easy-even-lr-2} can be tricky in practice, we would like to have a simpler combinatorial description of the $c_{\mu\nu}^\lambda$. The Littlewood-Richardson rule provides one of many such simpler descriptions. Good accounts of the rule and its proof are given in \cite[Chapter 5]{Fulton} and \cite[Appendix A1]{Stanley2}. We will review the terminology and statement here. \vspace{0.07in} Recall that a \textit{(Young) skew shape} $\lambda/\mu$ is the complement of a subdiagram $\mu\subseteq\lambda$. A \textit{semistandard skew tableau} is a skew shape which has been filled with entries from some ordered alphabet, subject to the same rules as for a semistandard tableau: entries must strictly increase in columns (top to bottom) and must not decrease in rows (left to right). We write $SSYT(\lambda/\mu)$ for the set of semistandard skew tableaux of shape $\lambda/\mu$ and $SSYT(\lambda/\mu,\nu)$ for the set of semistandard skew tableaux of shape $\lambda/\mu$ and content $\nu$. A word $w=w_1\cdots w_r$ in some ordered alphabet is called \textit{Yamanouchi} (or a \textit{reverse lattice word}) if, when read backwards, each initial subword has at least as many $a$'s as $b$'s whenever $a<b$. For example, $312211$ is Yamanouchi but $1221$ and $112$ are not. A skew tableau $S$ is called a \textit{Littlewood-Richardson tableau} if $w_r(S)$ is a Yamanouchi word. The following is Proposition 3, Chapter 5 of \cite{Fulton} and Theorem A1.3.3 of \cite{Stanley2}. \begin{thm}[(Even) Littlewood-Richardson Rule]\label{thm-even-lr} The coefficient $c_{\mu\nu}^\lambda$ equals the number of Littlewood-Richardson tableaux $S$ of shape $\lambda/\mu$ and content $\nu$.\end{thm} One specific bijection between the set described in the theorem and the set in equation \eqref{eqn-easy-even-lr-2} is described in the following section, where we use it to deduce an odd analogue of Theorem \ref{thm-even-lr}. \begin{example} Let $\mu\subseteq\lambda$ be a subdiagram and let $k=|\lambda|-|\mu|\geq1$. If $S$ is a Littlewood-Richardson tableau of shape $\lambda/\mu$ and content $(k)$, then no two boxes of $S$ can be in the same column (since all entries equal 1). And on any such skew shape $\lambda/\mu$, there is exactly one tableau of content $(k)$. Thus $c_{\mu(k)}^\lambda=1$ if $\lambda/\mu$ is a horizontal strip and equals 0 otherwise. We have deduced the (even) Pieri rule, \begin{equation} s_\mu s_{(k)}=\sum_{\substack{\lambda\\\lambda/\mu\text{ is a}\\\text{horizontal strip}}}s_\lambda. \end{equation} Using the standard involution $\omega$ on $\mathrm{\Lambda}$, or just arguing in analogy with the above, the same is true if $(k)$ is replaced by $(1^k)$ and ``horizontal'' is replaced by ``vertical.''\end{example} \begin{example} The lowest degree product which is not described by the Pieri rule is \begin{equation*} s_{21}s_{21}=s_{2211}+s_{222}+s_{3111}+2s_{321}+s_{33}+s_{411}+s_{42}. \end{equation*} \end{example} \subsection{The odd Littlewood-Richardson rule}\label{subsec-odd-lr} The sign of a Young tableau $T$ is the sign between its row word monomial $\widetilde{x}^{w_r(T)}$ and the ascend-sorting of that monomial. As explained in Section \ref{subsec-odd-symmetric}, this sign equals $(-1)^{N^<(T)}$. If $S$ is a skew tableau of shape $\lambda/\mu$, let $j$ be an element of the alphabet less than every entry of $S$ and let $\widehat{S}$ be the Young tableau of shape $\lambda$ formed by placing $j$ in each box of $\mu$ and filling the rest so as to match $S$. We then define the sign of $S$ to be \begin{equation*} \text{sign}(S)=\text{sign}(\widehat{S})=N^<(\widehat{S}). \end{equation*} When $\mu=(0)$, this reduces to the sign of a tableau as defined earlier. More generally, whenever we write a count $E, nW^\geq, dS,\ldots$ evaluated on $S$, read $\widehat{S}$ for $S$. \begin{example} To either alphabet ${\mathbbm Z}_{>0}$ or $\lbrace1,2,\ldots,n\rbrace$, we can always adjoin 0 and take $j=0$. If $\lambda=(3,3,2,1)$, $\mu=(2,1,1)$, and \begin{equation*} S=\young(::1,:12,:2,3),\quad\text{then}\quad\widehat{S}=\young(001,012,02,3)\quad\text{and}\quad\text{sign}(S)=(-1)^{18}=1. \end{equation*}\end{example} \begin{rem} We consider the partition $\mu$ to be part of the data of $S$. For example, both $(1,1)/(1)$ and $(2)/(1)$ consist of a single box, but if we fill these two boxes with equal entries, the resulting skew tableaux have opposite signs.\end{rem} \begin{defn} Let $\mu,\nu,\lambda$ be partitions. The \textit{odd Littlewood-Richardson coefficient} $c_{\mu\nu}^\lambda$ is the coefficient of $s_\lambda$ when $s_\mu s_\nu$ is expanded in the basis of odd Schur functions, \begin{equation*} s_\mu s_\nu=\sum_\lambda c_{\mu\nu}^\lambda s_\lambda. \end{equation*} If $|\mu|+|\nu|\neq|\lambda|$, then $c_{\mu\nu}^\lambda=0$.\end{defn} Note that the odd Pieri rules compute certain odd Littlewood-Richardson coefficients. Using the involution $\psi_1\psi_2$ and the anti-involution $\psi_3$, we know the odd Littlewood-Richardson coefficient $c_{\mu\nu}^\lambda$ whenever $\mu$ or $\nu$ has either height or width 1. If $Y,Z$ are two nonzero monomials in $\mathrm{OPol}_n$ such that $Y=\pm Z$, then let $\text{sign}(Y,Z)$ denote the sign between them; for example $\text{sign}(x_1x_2x_3,x_1x_3x_2)=-1$. \begin{lem}\label{lem-easy-lr} The odd Littlewood-Richardson coefficient $c_{\mu\nu}^\lambda$ is \begin{equation}\label{eqn-easy-lr} c_{\mu\nu}^\lambda=(-1)^{dN(\mu)+dN(\nu)+dN(\lambda)+N(\mu)+\sum_i(\lambda/\nu)_i\left|\frac{\nu}{i}\right|}\sum_{\substack{U\inSSYT(\mu)\\UT_\nu=T_\lambda}}(-1)^{N^<(U)}. \end{equation} In the summation condition, the product $UT_\nu$ is taken in the even plactic ring (so it is an equality up to sign in $\mathbb{Z}Pl_n$).\end{lem} \begin{proof} First note that \begin{equation}\begin{split} c_{\mu\nu}^\lambda&=(-1)^{dN(\mu)+dN(\nu)+dN(\lambda)+N(\mu)+N(\nu)+N(\lambda)}\sum_{\substack{U\inSSYT(\mu)\\V\inSSYT(\nu)\\UV=T_\lambda}}\text{sign}(\widetilde{x}^{w_r(U)}\widetilde{x}^{w_r(V)},\widetilde{x}^{w_r(T_\lambda)})\\ &=(-1)^{dN(\mu)+dN(\nu)+dN(\lambda)+N(\mu)+N(\nu)+N(\lambda)}\sum_{\substack{U\inSSYT(\mu)\\UT_\nu=T_\lambda}}\text{sign}(\widetilde{x}^{w_r(U)}\widetilde{x}^{w_r(T_\nu)},\widetilde{x}^{w_r(T_\lambda)}). \end{split}\end{equation} The first equality is immediate from the definition of $c_{\mu\nu}^\lambda$ and equation \eqref{eqn-defn-s-plactic}. The second follows from the following fact: if $U,V$ are semistandard tableaux of shapes $\mu,\nu$ and $UV=T_\lambda$ in the even plactic ring, then $V=T_\nu$ \cite[Section 5.2]{Fulton}. To turn $\widetilde{x}^{w_r(U)}\widetilde{x}^{w_r(T_\nu)}$ into $\widetilde{x}^{w_r(T_\lambda)}$, we proceed in two steps: \begin{enumerate} \item Ascend-sort each of $\widetilde{x}^{w_r(U)}$, $\widetilde{x}^{w_r(T_\nu)}$, and $\widetilde{x}^{w_r(T_\lambda)}$ separately. This incurs the sign $(-1)^{N^<(U)+N(\nu)+N(\lambda)}$. The monomials are now $:\widetilde{x}^{w_r(T_{\lambda/\nu})}::\widetilde{x}^{w_r(T_\nu)}:$ and $:\widetilde{x}^{w_r(T_\lambda)}:$, where the colons denote ascend-sorting. \item To sort these monomials together requires $\sum_i(\lambda/\nu)_i\left|\frac{\nu}{i}\right|$ transpositions. \end{enumerate} The lemma follows.\end{proof} \begin{thm}[Odd Littlewood-Richardson rule]\label{thm-lr} The odd Littlewood-Richardson coefficient $c_{\mu\nu}^\lambda$ is \begin{equation}\label{eqn-lr} c_{\mu\nu}^\lambda=(-1)^{N(\mu)+N(\lambda)}\sum_{\substack{S\inSSYT(\lambda/\mu,\nu)\\w_r(S)\text{ is Yamanouchi}}}(-1)^{N^<(S)}. \end{equation}\end{thm} \begin{proof} The Littlewood-Richardson bijection of \cite[Section 5.2]{Fulton}, \begin{equation}\label{eqn-lr-bij} \lbrace S\inSSYT(\lambda/\mu,\nu):w_r(S)\text{ is Yamanouchi}\rbrace\to\lbrace U\inSSYT(\mu,\lambda/\nu):\\UT_\nu=T_\lambda\rbrace, \end{equation} is described as follows. Let $U_0$ be a semistandard Young tableau of shape $\mu$, all of whose entries are less than all of those of $S$. Then form $T_S\inSSYT(\lambda)$ by giving it the entries of $U_0$ on $\mu\subseteq\lambda$ and the entries of $S$ on $\lambda/\mu$. Under the RSK correspondence, \begin{equation*} (T_\lambda,T_S)\leftrightarrow\left(\substack{\underline{t}\hspace{0.02in}\underline{u}\\\underline{x}\hspace{0.02in}\underline{v}}\right), \end{equation*} where $\underline{x},\underline{t}$ have length $|\mu|$ and $\underline{u},\underline{v}$ have length $|\nu|$. We can describe what these sub-words correspond to under the RSK correspondence as well: \begin{equation*} (U,U_0)\leftrightarrow\left(\substack{\underline{t}\\\underline{x}}\right),\qquad(T_\nu,T_\nu)\leftrightarrow\left(\substack{\underline{u}\\\underline{v}}\right) \end{equation*} for some $U\inSSYT(\mu,\lambda/\nu)$. The Littlewood-Richardson bijection \eqref{eqn-lr-bij} assigns $U$ to $S$. In order to prove the theorem, we have to relate the quantities $N^<(U)$ and $N^<(S)$. The odd RSK correspondence of \cite{EK} allows us to do this; it implies \begin{equation}\label{eqn-pf-lr-1}\begin{split} \text{sign}(\underline{x}\hspace{0.02in}\underline{v})&=(-1)^{dN(\lambda)+N(\lambda)+N^<(T_S)}\\ &=(-1)^{dN(\lambda)+N(\lambda)+N^<(U_0)+N^<(S)},\\ \text{sign}(\underline{x})&=(-1)^{dN(\mu)+N^<(U)+N^<(U_0)},\\ \text{sign}(\underline{v})&=(-1)^{dN(\nu)}. \end{split}\end{equation} Since we know the contents of $\underline{x}$, $\underline{v}$, and $\underline{x}\hspace{0.02in}\underline{v}$, the sign of the latest can be expressed in terms of the former two, \begin{equation}\label{eqn-pf-lr-2} \text{sign}(\underline{x}\hspace{0.02in}\underline{v})=\text{sign}(\underline{x})\text{sign}(\underline{v})(-1)^{\sum_i(\lambda/\nu)_i\left|\frac{\nu}{i}\right|}. \end{equation} Comparing the signs in equations \eqref{eqn-pf-lr-1} and \eqref{eqn-pf-lr-2}, \begin{equation*} (-1)^{dN(\mu)+dN(\nu)+dN(\lambda)+N(\lambda)+N^<(U)+N^<(S)+\sum_i(\lambda/\nu)_i\left|\frac{\nu}{i}\right|}=1. \end{equation*} Applying this to Lemma \ref{lem-easy-lr}, the theorem follows.\end{proof} \begin{example} The first interesting cancellation is $c_{(2,1)(2,1)}^{(3,2,1)}=0$ (in the even case, it equals 2). The Littlewood-Richardson skew tableaux of shape $(3,2,1)/(2,1)$ and content $(2,1)$ are \begin{equation*} \young(::1,:1,2),\quad\young(::1,:2,1). \end{equation*} They have $N^<(S)$ equal to $7$ and $6$, respectively.\end{example} \begin{rem} It follows from equation \eqref{eqn-psi-schur} that \begin{equation}\begin{split} &c_{\mu\nu}^\lambda=(-1)^{dN(\mu)+dN(\nu)+dN(\lambda)+N(\mu)+N(\nu)+N(\lambda)}c_{\nu\mu}^\lambda,\\ &c_{\mu\nu}^\lambda=(-1)^{NE(\mu)+NE(\nu)+NE(\lambda)}c_{\mu^T\nu^T}^{\lambda^T}. \end{split}\end{equation} These symmetries constrain some signs associated to Young diagrams. For instance, if $c_{\mu\mu}^\lambda\neq0$, then $(-1)^{dN(\lambda)+N(\lambda)}=1$. \end{rem} \subsection{Translation to Knutson-Tao hives}\label{subsec-hives} There are many combinatorial expressions for the even Littlewood-Richardson coefficients. Among them, the several which are expressible in terms of integer points of rational convex polytopes are especially interesting; one reason is that as result of Knutson and Tao's proof of the Saturation Conjecture \cite{KnutsonTao}, Klyachko's system of inequalities \cite{Klyachko} gives a necessary and sufficient criterion for a Littlewood-Richardson coefficient to be nonzero. These expressions in terms of polytopes include Gelfand-Zeitlin (GZ) patterns \cite{GelfandZelevinsky}, Berenstein-Zelevinsky triangles \cite{BerensteinZelevinsky}, the Littlewood-Richardson triangles of Pak and Vallejo \cite{PakVallejo}, and the honeycombs and hives of Knutson and Tao \cite{KnutsonTao}. As explained in the exposition of \cite{Buch}, the hive model is particularly convenient and flexible. In this section, we will write down the bijection of \cite{PakVallejo} from Littlewood-Richardson skew tableaux to hives, by way of Littlewood-Richardson triangles. In both the triangle and hive settings, the sign associated to a diagram will come from a quadratic form on the ambient space of the polytope in question. \vspace{0.07in} The following exposition follows \cite{PakVallejo} rather closely; our only contribution is to track the signs which arise in the odd case. For fixed $n\geq1$, we will work in $V_{\mathbbm Z}={\mathbbm Z}^{\binom{n+2}{2}}$ and $V=V_{\mathbbm Z}\otimes_{\mathbbm Z}{\mathbbm R}$, with coordinates $\lbrace a_{i,j}:0\leq i\leq j\leq n\rbrace$. We call elements of $V$ \textit{triangles} and write them pictorially as (for $n=3$) \begin{equation*} \begin{matrix}&&&a_{0,0}&&&\\\\&&a_{0,1}&&a_{1,1}&&&\\\\&a_{0,2}&&a_{1,2}&&a_{2,2}&&\\\\a_{0,3}&&a_{1,3}&&a_{2,3}&&a_{3,3}.\end{matrix} \end{equation*} \begin{defn} A \textit{Littlewood-Richardson triangle} is an element $A=(a_{i,j})\in V$ such that \begin{enumerate} \item $a_{i,j}\geq0$ for $1\leq i<j\leq n$, \item \begin{equation*} \sum_{p=0}^{i-1}a_{p,j}\geq\sum_{p=0}^ia_{p,j+1}\qquad\text{for }1\leq i\leq j<n,\text{ and} \end{equation*} \item \begin{equation*} \sum_{q=i}^ja_{i,q}\geq\sum_{q=i+1}^{j+1}a_{i+1,q}\qquad\text{for }1\leq i<j<n. \end{equation*} \end{enumerate} Note that $a_{0,j}<0$ and $a_{j,j}<0$ are both possible.\end{defn} Let $\bigtriangleup_{LR}$ denote the set of all Littlewood-Richardson triangles, a cone in $V$. Given a Littlewood-Richardson triangle $A=(a_{i,j})\in\bigtriangleup_{LR}$, define partitions $\lambda,\mu,\nu$ by \begin{equation*} \lambda_j=\sum_{p=0}^ja_{p,j},\qquad\mu_j=a_{0,j},\qquad\nu_i=\sum_{q=i}^ka_{i,q}. \end{equation*} Let $\bigtriangleup_{LR}(\lambda,\mu,\nu)$ be the set of Littlewood-Richardson triangles with fixed $\lambda,\mu,\nu$. Each set $\bigtriangleup_{LR}(\lambda,\mu,\nu)$ is a convex polytope in $V$. To a Littlewood-Richardson skew tableau $S\inSSYT(\lambda/\mu,\nu)$, associate an element $A_S$ of $\bigtriangleup_{LR}(\lambda,\mu,\nu)$ by setting \begin{equation}\begin{split} &a_{0,0}=0,\qquad a_{0,j}=\mu_j,\\ &a_{i,j}=\#\lbrace\text{entries equal to }i\text{ in row }j\text{ of }S\rbrace\quad(0<i\leq j). \end{split}\end{equation} \begin{lem}[\cite{PakVallejo}, Lemma 3.1] Suppose $|\mu|+|\nu|=|\lambda|$. Then the assignment $S\mapsto A_S$ is a bijection between the set of Littlewood-Richardson tableaux in $SSYT(\lambda/\mu,\nu)$ and the set of integer points of $\bigtriangleup_{LR}(\lambda,\mu,\nu)$.\end{lem} It is easy to read off $N^<(S)$ from the triangle $A_S=(a_{i,j})$. For each entry $a_{i,j}$, write $Y_{i,j}$ for the sum of all the $a_{p,q}$ in the shaded region below, where the $(i,j)$ place is the dot drawn with a hollow center: \vspace{0.15in} \begin{center}\includegraphics[width=2.5in]{oddlrc-quadratic-form.eps}\end{center} \vspace{0.15in} More formally, \begin{equation} Y_{i,j}=\sum_{p=0}^{i-1}\sum_{q=p}^{j-1}a_{p,q}. \end{equation} Consider the quadratic form \begin{equation} Q_\bigtriangleup(A)=\sum_{i,j}a_{i,j}Y_{i,j}=\sum_{i=0}^n\sum_{j=i}^n\left(a_{i,j}\sum_{p=0}^{i-1}\sum_{q=p}^{j-1}a_{p,q}\right). \end{equation} Then it is immediate from the description of the bijection above that \begin{equation*} N^<(S)=Q_\bigtriangleup(A_S), \end{equation*} so \begin{equation} c_{\mu\nu}^\lambda=(-1)^{N(\mu)+N(\lambda)}\sum_{A\in\bigtriangleup_{LR}(\lambda,\mu,\nu)\cap V_{\mathbbm Z}}(-1)^{Q_\bigtriangleup(A)}, \end{equation} as long as $\lambda,\mu,\nu$ all have at most $n$ parts. \begin{example}\label{ex-triangle} If \begin{equation*} S=\young(::1,:2,1), \end{equation*} then \begin{equation*} A_S=\begin{matrix}&&&0&&&\\\\&&2&&1&&\\\\&1&&0&&1&&\\\\0&&1&&0&&0&\end{matrix} \end{equation*} and $Q_\bigtriangleup(A_S)=6$. \end{example} We now translate this result into the language of hives. \begin{defn} A \textit{hive} is an element $H=(h_{i,j})\in V$ with $h_{0,0}=0$ which satisfies the inequalities \begin{equation}\label{eqn-hive-inequalities}\begin{split} &(R)\qquad h_{i,j}-h_{i,j-1}\geq h_{i-1,j}-h_{i-1,j-1}\text{ for }1\leq i<j\leq n,\\ &(V)\qquad h_{i-1,j}-h_{i-1,j-1}\geq h_{i,j+1}-h_{i,j}\text{ for }1\leq i\leq j<n,\\ &(L)\qquad h_{i,j}-h_{i-1,j}\geq h_{i+1,j+1}-h_{i,j+1}\text{ for }1\leq i\leq j<n. \end{split}\end{equation}\end{defn} Let $\mathfrak{H}$ be the set of all hives; this is a cone in $V$. The inequalities \eqref{eqn-hive-inequalities} have a geometric interpretation when we express $H$ as a triangle. Inside a triangle, there are three types of rhombi which can be made out of two adjacent smallest-size triangles: \vspace{0.15in} \begin{center}\includegraphics[width=2.5in]{oddlrc-rhombi.eps}\end{center} \vspace{0.15in} We call these right-slanted (R), vertical (V), and left-slanted (L) rhombi. The inequalities \eqref{eqn-hive-inequalities} say that the sum of the entries at the obtuse angles of any such rhombus is greater than or equal to the sum at the acute angles; the three inequalities are the right-slanted, vertical, and left-slanted cases, respectively. As with Littlewood-Richardson triangles, we associate three partitions to a hive, \begin{equation*} \lambda_j=h_{j,j}-h_{j-1,j-1},\qquad\mu_j=h_{0,j}-h_{0,j-1},\qquad\nu_i=h_{i,n}-h_{i-1,n}. \end{equation*} Let $\mathfrak{H}(\lambda,\mu,\nu)$ be the set of all hives with corresponding partitions $\lambda,\mu,\nu$. Each $\mathfrak{H}(\lambda,\mu,\nu)$ is a convex polytope in $V$. \begin{thm}[\cite{PakVallejo}, Theorem 4.1] Let $\Phi:V\to V$ be the linear map which takes $A=(a_{i,j})$ to $H=(h_{i,j})$, where \begin{equation}\label{eqn-a-to-h} h_{i,j}=\sum_{p=0}^i\sum_{q=p}^ja_{p,q}. \end{equation} Then $\Phi$ is a volume-preserving isomorphism and induces bijections $\bigtriangleup_{LR}(\lambda,\mu,\nu)\to\mathfrak{H}(\lambda,\mu,\nu)$ for all $\lambda,\mu,\nu$.\end{thm} As a matter of convention, let $h_{i,j}=0$ if either $i>j$, $i<0$, or $j>n$. It follows from equation \eqref{eqn-a-to-h} and an inclusion-exclusion argument that \begin{equation}\label{eqn-hij}\begin{split} &h_{i,j}=a_{i,j}+h_{i-1,j}+h_{i,j-1}-h_{i-1,j-1}\text{ if }0\leq i<j\leq n,\\ &h_{i,i}=a_{i,i}+h_{i-1,i}\text{ if }0\leq i=j\leq n. \end{split}\end{equation} Let $Q_\mathfrak{H}=Q_\bigtriangleup\circ\Phi^{-1}$. Then for a hive $H=(h_{i,j})$, equations \eqref{eqn-hij} imply \begin{equation} Q_\mathfrak{H}(H)=\sum_{i=1}^n\sum_{j=i}^nh_{i-1,j-1}\left(h_{i,j}-h_{i-1,j}-h_{i,j-1}+h_{i-1,j-1}\right)-\sum_{i=1}^{n-1}h_{i,i}^2. \end{equation} Note that the parenthesized term is non-negative by the right-slanted rhombus inequality $(R)$. It follows that if $\mu,\nu,\lambda$ are partitions with at most $n$ parts, then \begin{equation} N^<(S)=Q_\bigtriangleup(A_S)=Q_\mathfrak{H}(\Phi(A_S)), \end{equation} so \begin{equation} c_{\mu\nu}^\lambda=(-1)^{N(\mu)+N(\lambda)}\sum_{H\in\mathfrak{H}(\lambda,\mu,\nu)\cap V_{\mathbbm Z}}(-1)^{Q_\mathfrak{H}(H)}. \end{equation} The quantity $N(\mu)+N(\lambda)$ can also be expressed as a quadratic form in either the variables $(a_{i,j})$ or the variables $(h_{i,j})$. \begin{example} With $S$ and $A_S$ as in Example \ref{ex-triangle}, \begin{equation} \Phi(A_S)=\begin{matrix}&&&0&&&\\\\&&2&&3&&&\\\\&3&&4&&5&&\\\\3&&5&&6&&6&\end{matrix} \end{equation} and $Q_\mathfrak{H}(\Phi(A_S))=6$. \end{example} \bibliographystyle{alpha}
\section{The Question} \label{sec-intro} The MiniBooNE experiment, in the neutrino-mode, has reported an excess of $129 \pm 43$ (Stat $\oplus$Syst) electron-like events in the 200--475 MeV energy range~\cite{MB-EX}, a range largely unaffected by the \nmne\ oscillations at $\Delta m^2 > 1$ eV$^2$, which is expected to dominate the 475-1250 MeV range~\cite{MB-NMNE}. Assuming that the efficiency of these events is 0.25, similar to that of the \ne-induced electrons, the rate of the excess with respect to \nm-CC is $\simeq 3 \times 10^{-3}$. In the antineutrino mode, initially the MiniBOONE data were consistent with background in the 200--475 MeV range~\cite{MB-NMbNEb}. Recently, with additional antineutrino data, they report a 2$\sigma$ excess in this region~\cite{MB-NUFACT11}. MiniBOONE lacks the resolution to determine if the excess is due to $e^-$, $e^+$, or a photon. In neutrino interactions single photons do not occur: they occur in pairs from the $\pi^0 \rightarrow \gamma \gamma$ decay. Evidence of events with a single photon and nothing else in neutrino induced neutral current (NC) interaction will signal either new or unconventional physics. This anomaly needs to be checked experimentally. The low-energy MiniBOONE excess has motivated novel hypotheses invoking conventional, to unconventional, to entirely new physics processes. An explanation~\cite{BODEK-BREM} of the excess attributed to the internal bremsstrahlung associated with muons in \nm-induced charged current interaction (CC) has been refuted by MiniBooNE~\cite{MB-BREM}. Harvey, Hill, and Hill~\cite{3H-PRL, 3H-PRD} postulate a new anomaly-mediated interaction between the $\nu$-induced $Z^0$-boson, the photon, and a vector meson, such as $\omega$, coupling strongly to the target nucleus. A similar, single photon process has been envisioned by Jenkins \& Goldman~\cite{NCRAD-JG}. In the anomaly-mediated neutrino-photon (ANP) interaction, $\nu {\cal N} \rightarrow \nu {\cal N} \gamma$, where ${\cal N}$ is the target nucleus or nucleon, the observable is a single \boldmath { ${ { \gamma }} $} , approximately collinear with the incident neutrino direction in the lab frame with a recoiling ${\cal N}$ which remains intact and largely undetected; the form factor of ${\cal N}$ is expected to induce the single $\gamma$ to be pulled forward. The ANP cross-section is expected to increase with neutrino energy ($E_\nu$). The ANP interaction is akin to the exclusive $\pi^0$ production when the neutrino coherently scatters off a target nucleus (\boldmath {Coh$\pi^0$} ), {\boldmath $\nu + {\cal N} \rightarrow \nu + {\cal N} + \pi^0$}, where the only observable is $\pi^0 \rightarrow \gamma \gamma$. The phenomenology of ANP developed by Hill in~\cite{HILL09}, and largely focussed on the low energy neutrino interaction ($E_\nu \simeq {\cal O}(1)$~GeV), predicts a cross-section sufficiently large to explain the MiniBOONE anomaly. If true, the ANP interaction will also impact astrophysical processes such as neutron star cooling. Hypotheses involving new physics include: decay of a heavy neutrino into a light neutrino and a photon, $\nu_h \rightarrow \nu + \gamma$, by Gninenko~\cite{SG-NDKY, SG-NDKY10}; CP-violation in 3$\oplus$ 2 model by Maltoni \& Schwetz~\cite{CP32-MS}, and by Goldman {\it et al.}~\cite{CP32-GSM}; extra-dimensions in 3$\oplus$1 model by Pas {\it et al.}~\cite{ExDim-PPW}; and models by Giunti and Laveder~\cite{3p1-GL}; Lorentz-violation by Katori {\it et al.}~\cite{LV-KKT}; CPT-violation in 3$\oplus$1 model by Barger {\it et al.}~\cite{CPT31-BMW}; new gauge bosons with sterile neutrinos by Nelson \& Walsh~\cite{NGBSN-NW}; and soft-decoherence by Farzan {\it et al.}~\cite{SD-FSS}; etc. The NOMAD collaboration has reported a search of heavy-$\nu$ which mixes with $\nu_\tau$'s produced in the SPS proton-target, and then decays into $e^-e^+$ pair in the detector~\cite{NOMAD-HNU}. The search focussed on highly collinear (${\cal {C}} = \left [ 1 - cos \Theta_{ee} \right ] \leq 2 \times10^{-5}$) and high energy ($E_{ee} \geq 4$~GeV) $e^-e^+$ events, where $\Theta_{ee}$ and $E_{ee}$ are the polar angle and the energy of the converted photon, and resulted in a single event consistent with the estimated background. The present work extends this search with no cut on ${\cal C}$. The high resolution NOMAD data allow a sensitive search for neutrino induced single photon events (\boldmath {1${ { \gamma }} $} ). The detector measures the energy and emission angle of the photon. Additionally, the detector affords the redundancy to measure {\it in situ} all relevant backgrounds relieving the reliance on Monte Carlo simulation (MC) of the conventional processes. As regards the MiniBooNE low-energy anomaly, we note that the average energy of the SPS neutrino flux at NOMAD, $E_{\nu} \simeq 25$~GeV, is much higher than that of MiniBOONE, $E_{\nu} \simeq 1$~GeV. However, the NOMAD data provide a stringent check in a different energy range. If a mechanism explaining the anomaly were applicable to higher energies, such as ANP or $\nu$-decay hypotheses, then this search, even if negative, will furnish meaningful limits. \section{Beam and Detector} \label{sec-nomad} The Neutrino Oscillation MAgnetic Detector (NOMAD) experiment at CERN used a neutrino beam produced by the 450~GeV protons from the Super Proton Synchrotron (SPS) incident on a beryllium target and producing secondary $\pi^{\pm}$, $K^{\pm}$, and $K^0$ mesons. The positively charged mesons were focussed by two magnetic horns into a 290~m long evacuated decay pipe. Decays of $\pi^{\pm}$, $K^{\pm}$, and $K^0_L$ produced the SPS neutrino beam. The average neutrino flight path to NOMAD was 628~m, the detector being 836~m downstream of the Be-target. The SPS beamline and the neutrino flux incident at NOMAD are described in~\cite{NOMAD-FLUX}. The $\nu$-flux in NOMAD is constrained by the $\pi^{\pm}$ and $K^{\pm}$ production measurements in proton-Be collision by the SPY experiment ~\cite{SPY1, SPY2, SPY3} and by an earlier measurement conducted by Atherton {\it et al.}~\cite{ATHERTON}. The $E_\nu$-integrated relative composition of \nm:\nmb:\nel:\neb\ CC events, constrained {\it in situ} by the measurement of CC-interactions of each of the neutrino species, is $1.00: 0.025: 0.015:0.0015$. Thus, 97.5\% of the events are induced by neutrinos with a small anti-neutrino contamination, similar to that reported by MiniBOONE~\cite{MB-NMNE}. The NOMAD apparatus, described in~\cite{NOMAD-NIM}, was composed of several sub-detectors. The active target comprised 132 planes of $3 \times 3$~m$^2$ drift chambers (DC) with an average density similar to that of liquid hydrogen (0.1~gm/cm$^3$). On average, the equivalent material in the DC encountered by particles produced in a $\nu$-interaction was about half of a radiation length and a quarter of a hadronic interaction length ($\lambda$). The fiducial mass of the NOMAD DC-target, 2.7 tons, was composed primarily of carbon (64\%), oxygen (22\%), nitrogen (6\%), and hydrogen (5\%) yielding an effective atomic number, \boldmath {${\cal A}$} =12.8, similar to carbon. Downstream of the DC, there were nine modules of transition radiation detectors (TRD), followed by a preshower (PRS) and a lead-glass electromagnetic calorimeter (ECAL). The ensemble of DC, TRD, and PRS/ECAL was placed within a dipole magnet providing a 0.4~T magnetic field orthogonal to the neutrino beam line. Two planes of scintillation counters, $T_1$ and $T_2$, positioned upstream and downstream of the TRD, provided the trigger in combination with an anti-coincidence signal, ${\overline V}$, from the veto counter upstream and outside the magnet. Downstream of the magnet was a hadron calorimeter, followed by two muon-stations each comprising large area drift chambers and separated by an iron filter. The two stations, placed at 8- and 13-$\lambda$'s downstream of the ECAL, provided a clean identification of the muons. The schematic of the DC-tracker with a $\gamma$-conversion in the Y-Z view is shown in Figure~\ref{fig-evt1v01}. The charged tracks in the DC were measured with an approximate momentum ($p$) resolution of ${\rm {\sigma_p/p = 0.05/\sqrt{L} \oplus 0.008p/\sqrt{L^5} }}$ ($p$ in GeV/$c$ and $L$ in meters) with unambiguous charge separation in the energy range of interest. The experiment recorded over 1.7 million neutrino interactions in the active drift-chamber (DC) target in the range ${\cal {O}}(1) \leq E_\nu \leq 300$~GeV. \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.6\textwidth,angle=90] {12276-43379.pdf} \caption{Schematic of the DC tracker with a single $\gamma$-conversion event in NOMAD in Y (vertical) vs Z (horizontal) view. The energy and angle with respect to the neutrino direction of the \boldmath { ${ { \gamma }} $} \ are 3.1~GeV and 0.095~Radians. } \label{fig-evt1v01} \end{center} \end{figure} \section{The Exclusive Photon Signature} \label{sec-sig} In NOMAD the cleanest signature of a \boldmath { ${ { \gamma }} $} \ is a \boldmath {${\rm V}^0 $} \ arising from its conversion in the DC into $e^-$ and $e^+$ tracks as shown in Figure~\ref{fig-evt1v01}. The \boldmath { ${ { \gamma }} $} -momentum is reconstructed by measuring the 3-momenta of the associated $e^-$/$e^+$ tracks in the DC. The ECAL measures energy associated with the $e^-$/$e^+$ tracks plus any unconverted photons. Because the trigger counters are situated at the downstream end of the detector, only a fraction of events containing single photon will trigger the apparatus, namely events where the photon undergoes conversion in DC and the resulting $e^+$ or $e^-$ traverses the trigger counters. Events with unconverted photons reaching the ECAL will not trigger the apparatus. For example, about 29\% of the \boldmath {Coh$\pi^0$} \ events, containing a $\pi^0$ and `nothing else', trigger the apparatus; the loss arises from the unconverted photons and, among the converted photons, from the $e^-/e^+$ tracks that do not reach the downstream trigger counters. Single photon events will manifest as a \boldmath {${\rm V}^0 $} , composed of $e^-$ and $e^+$, with no additional energy in the ECAL. A converted photon is defined as a \boldmath {${\rm V}^0 $} \ such that the invariant mass of the $e^-$ and $e^+$ ($M_{ee}$) is less than 100~MeV which selects the converted photons with 95\% purity and 97\% efficiency. Furthermore, if the single photon is approximately collinear with the incident neutrino then the collinearity of the photon-track offers an additional kinematic handle. Because we do not have a definitive simulation for the single \boldmath { ${ { \gamma }} $} \ process, we use one of the photons from the \boldmath {Coh$\pi^0$} , while ignoring the other photon, as a {\it guide} for the signal. We define two signal boxes. ${\rm {Box1}}$ comprises single \boldmath { ${ { \gamma }} $} \ events with negligible energy in the ECAL associated with neutral particles (\boldmath { ${ { \gamma }} $} , neutron, etc.) quantified by the quantity, \boldmath {${\rm {PAN }}$ } = $\left [ E_\gamma - E_{Neut} \right ] / \left [ E_\gamma + E_{Neut} \right ]$, $ \geq 0.9$, where $E_\gamma$ is the energy of the photon derived from the $e^-$/$e^+$ tracks and $E_{Neut}$ is the ECAL energy associated with other neutral particles, i.e. not associated with $e^+$/$e^-$ tracks. (For calculation of $E_{Neut}$ see, for example, ~\cite{NOMAD-NMNT}.) The \boldmath {${\rm {PAN }}$ }\ $\geq 0.9$ cut is model independent. ${\rm {Box2}}$, a subset of ${\rm {Box1}}$, comprises single \boldmath { ${ { \gamma }} $} \ events provided that they are approximately collinear with the incident neutrino, characterized by the variable, {\boldmath $\zeta=E_{\gamma} \left [ 1-\cos(\theta_{\gamma}) \right ] \leq 0.05$}, where $E_{\gamma}$ and $\theta_{\gamma}$ are the photon's energy and polar angle with respect to the neutrino beam. (The variable $\zeta$ has the property that its distribution depends weakly on the incident neutrino energy.) Approximately 90\% of the photons from \boldmath {Coh$\pi^0$} \ have $\zeta \leq 0.05$. Although the $\zeta \leq 0.05$ cut will be somewhat model-dependent, at the SPS energies ($E_\nu \simeq 25$~GeV) most exotic models suggest that the single photon will be approximately collinear with the incident neutrino. The backgrounds to the single \boldmath { ${ { \gamma }} $} \ are all interactions where all daughter particles, except one photon, evade detection. The background includes the \boldmath {Coh$\pi^0$} \ interaction, the deep inelastic $\nu$-NC scattering (NC-DIS), and $\nu$-interactions occurring outside the fiducial volume (OBG). Note, that with a single converted \boldmath { ${ { \gamma }} $} \ in DC the coordinates of the $\nu$-interaction cannot be ascertained. The sample with \boldmath {${\rm {PAN }}$ }\ $< 0.9$ is composed of events with a \boldmath {${\rm V}^0 $} \ and neutral energy in the ECAL as expected from a \boldmath {$\pi^0 $} \ decay and constitutes an important control sample. All three backgrounds are determined {\it in situ} as described below. The analysis proceeds as follows. First, the sample with a single photon-conversion in DC is selected. Second, the three backgrounds to the single photon sample are calibrated using data outside ${\rm {Box1}}$. The analysis leads to an experimentally constrained prediction of the background events in the signal region, ${\rm {Box1}}$ and ${\rm {Box2}}$. Finally, boxes are opened and the predictions are compared with the observed data. \section{Selection of Single Photon Events} \label{sec-sel} We select $\nu$-induced events where a single photon converts (\boldmath {${\rm V}^0 $} ) within the fiducial volume of the DC target. The analysis uses the entire NOMAD data and the associated Monte Carlo (MC) samples described in ~\cite{NOMAD-XSEC}. The NC-DIS sample, defined by requiring that the generated invariant hadronic mass squared (${\rm {W^2}}$) be $\geq 1.96$~GeV$^2$, is normalized to $0.53 \times 10^6$ events which corresponds to 37\% of the \nm-CC. The normalization of the NC-Resonance (${\rm {W^2 < 1.96}}$) sample is set at 3.5\% of the NC-DIS. The \boldmath {Coh$\pi^0$} \ interaction is simulated using the Rein-Sehgal (RS) model~\cite{RS} which agrees well with our measurement~\cite{NOMAD-CohPi0}. The MC and data samples are subjected to a preselection requiring: (a) the presence of a converted photon whose reconstructed conversion point (${\rm {X}}$, ${\rm {Y}}$, ${\rm {Z}}$) be within the fiducial volume, ${\rm {|X,(Y-5)|\leq 130}}$~cm and \\ ${\rm {Z_{Min} \leq Z \leq 405}}$~cm where ${\rm {Z_{Min}}}$ is 5 and 35~cm for the two configurations of the detector composing more than 95\% of the NOMAD data; (b) no reconstructed muon ($\mu$ID-veto) and a single photon-induced \boldmath {${\rm V}^0 $} ; (c) the ${\rm {X}}$ and ${\rm {Y}}$ coordinate of the \boldmath { ${ { \gamma }} $} -vector when extrapolated back to the upstream most DC at ${\rm {Z_{Min}}}$ be within the transverse fiducial volume, ${\rm { |X,Y-5|_{PROJ} \leq130cm}}$, largely eliminating $\gamma$'s that enter from the sides; and (d) the energy of the $e^-e^+$ pair be $\geq 1.5$~GeV which reduces the NC-DIS and OBG background while having a small effect on the \boldmath {Coh$\pi^0$} \ sample which also serves as a guide for the single $\gamma$ signal. The preselection reduces the NC-DIS and data samples by more than a factor of one hundred. It is noteworthy that only a small fraction of the NC-Res pass the selection since most of the photons, coming from $\pi^0$ emitted at large angles, either fail to trigger the apparatus or have energies below $1.5$~GeV. Given the paucity of the NC-Resonance events, it has been added to the NC-DIS sample. Finally a negligible fraction of \nm-CC ($<10^{-5}$) pass the selection; consequently the CC sample has been ignored in this analysis. The preselection is presented in Table~\ref{tab-presel}. \begin{table}\centering \begin{tabular}{||c||c|c|c||c||} \hline Cut & \boldmath {Coh$\pi^0$} \ & NC-Res & NC-DIS MC & Data \\ \hline Fiducial Cut & 4,900 & 20,000 & 530,000 & 4,018,980 \\ $\mu$ID-veto \& Single $\gamma$-Induced \boldmath {${\rm V}^0 $} \ & 819 & 306 & 4,330 & 34,062 \\ ($M_{ee} \leq 100$~MeV) & & & & \\ ${\rm {|X,(Y-5)|_{PROJ}\leq130}}$~cm & 719 & 138 & 3,076 & 10,547 \\ $E_{ee} \geq 1.5$ GeV & 516 & 69 & 1,846 & 4,543 \\ & & & & \\ \hline \hline \end{tabular \caption{Preselection of events with a single, converted \boldmath { ${ { \gamma }} $} \ in \boldmath {Coh$\pi^0$} , NC-Resonance, and NC-DIS MC samples, and data. The MC samples, having a much larger statistics than data, are normalized to the expectation as described in the text. The ${\rm { |(X,Y-5)|_{PROJ} }}$ refer to X- and Y-position of the \boldmath { ${ { \gamma }} $} -vector when projected to the most upstream DC (${\rm { Z_{Min} }}$). } \label{tab-presel} \end{table} The final event selection follows the preselection with more stringent requirements. The vertex coordinates of \boldmath {${\rm V}^0 $} \ are required to be within \\ ${\rm {|X,(Y-5)|\leq 120}}$~cm and ${\rm {Z_{Min} \leq Z \leq 405}}$~cm. Two additional cuts (Clean-\boldmath {${\rm V}^0 $} ) are imposed to reduce outside background by requiring that there be no tracks upstream of the photon conversion and that there be no hits associated with the tracks composing the converted-\boldmath { ${ { \gamma }} $} \ in the most upstream DC. Finally, the $M_{ee}$ cut is tightened to $\leq 50$~MeV which increases the photon conversion purity to $\geq 98\%$ while reducing the efficiency to 93\%. Table~\ref{tab-sel} summarizes the selection of events in the MC samples. The preselected data are subjected to identical cuts. Distributions of the X, Y, and Z coordinates of the \boldmath {1${ { \gamma }} $} \ vertex in Data and the corresponding MC prediction, composed of \boldmath {Coh$\pi^0$} \, OBG, NC-DIS, agree. Only the control sample with \boldmath {${\rm {PAN }}$ } $< 0.9$ is examined to check and constrain the background prediction. The calibration of the three backgrounds is presented next. \begin{table}\centering {\small{ \begin{tabular}{||c||c|c||} \hline Cut & \boldmath {Coh$\pi^0$} -RS & NC-DIS $\oplus$ Resonance \\ \hline \hline Start & 516 & 1915 \\ Tighter Fid-Cuts & 483 & 1775 \\ $M_{ee}\leq50$~MeV \& Clean-\boldmath {${\rm V}^0 $} \ & 386 & 400 \\ \hline \hline \end{tabular \caption{Final selection of single $\gamma$ Events in the MC Samples: The MC samples have been normalized as presented in Section~\ref{sec-sel}.} \label{tab-sel} }} \end{table} \section{Constraining the Backgrounds and the \boldmath {${\rm {PAN }}$ }\ $< 0.9$ Sample } \label{sec-back} The prediction of the backgrounds to the single \boldmath { ${ { \gamma }} $} \ sample is data driven. The estimation of the \boldmath {Coh$\pi^0$} \ induced \boldmath {1${ { \gamma }} $} \ needs to be based on the observed \boldmath {2${\rm V}^0 $} \ sample where both photons convert in the DC. Furthermore, Monte Carlo simulations can neither reliably provide the OBG nor the NC-DIS contribution to the \boldmath {1${ { \gamma }} $} \ samle. These backgrounds need to be determined using data themselves. First, we present the calibration of the \boldmath {Coh$\pi^0$} \ contribution to the \boldmath {1${ { \gamma }} $} \ sample because it is the simplest. The analysis of the \boldmath {2${\rm V}^0 $} \ sample, where two photons and nothing else is observed, provides a measurement of the \boldmath {Coh$\pi^0$} \ process. It yields a rate with respect to the inclusive \nm-CC, $\left [ 3.21 \pm 0.46(stat \oplus syst) \right ] \times 10^{-3}$~\cite{NOMAD-CohPi0}. The measured kinematic distributions of the \boldmath {2${ { \gamma }} $} \ data are consistent with the RS-\boldmath {Coh$\pi^0$} \ model. Consequently we use the simulated RS-\boldmath {Coh$\pi^0$} \ events, normalize it to the measured cross-section, subject these to the \boldmath {1${ { \gamma }} $} \ analysis, and obtain the \boldmath {Coh$\pi^0$} \ contribution to the \boldmath {1${ { \gamma }} $} \ sample. The analysis yields a normalized prediction of \boldmath {Coh$\pi^0$} -induced \boldmath {1${ { \gamma }} $} \ with a 14.5\% precision, shown in Table~\ref{tab-FinalSel}. Next, we present the calibration of the OBG contribution to the \boldmath {1${ { \gamma }} $} \ sample. This, too, is accomplished using the \boldmath {2${\rm V}^0 $} \ sample as in~\cite{NOMAD-CohPi0}. The two reconstructed photon momentum vectors enable one to determine the $\nu$-interaction vertex by extrapolating the vectors upstream and finding the coordinates of their distance of closest approach (DCA). The procedure defines the DCA-vertex with coordinates denoted as DCA-X, DCA-Y, and DCA-Z. The DCA-vertex resolution is well understood. Among the \boldmath {2${ { \gamma }} $} \ sample 169 events, out of 550 events (see Table~\ref{tab-DCA-Z}), have a DCA vertex of origin upstream of the ${\rm {Z_{Min}}}$ providing a calibration for the OBG-\boldmath {1${ { \gamma }} $} \ as follows. A different data sample, OBG-Data, which includes charged tracks, is selected in which the vertex reconstructed using these charged tracks is upstream of the fiducial limit (${\rm {Z \leq Z_{Min}}}$). In this sample, the charged tracks are ignored and the events subjected to the \boldmath {2${ { \gamma }} $} \ analysis (see ~\cite{NOMAD-CohPi0}). This OBG-Data sample with \boldmath {2${ { \gamma }} $} \ yields 927 events with {\bf DCA-Z}${\rm {\leq Z_{Min}}}$ and 451 events with {\bf DCA-Z}${\rm {> Z_{Min}}}$ as shown in Table~\ref{tab-DCA-Z}. The number of \boldmath {2${ { \gamma }} $} \ events with {\bf DCA-Z}${\rm {> Z_{Min}}}$, originating from events occurring at ${\rm {Z\leq Z_{Min}}}$ but producing no visible tracks, is then calculated as $(451/927)\times 169 = 82.2 \pm 6.1$. This constitutes the OBG background to the \boldmath {2${ { \gamma }} $} \ sample. The so normalized OBG sample is then subjected to the \boldmath {1${ { \gamma }} $} \ analysis yielding a calibrated OBG prediction, with a 7.7\% error (driven by the statistics of 169 events), shown in Table~\ref{tab-FinalSel}. \begin{table} \begin{tabular}{|||c||c|c|||} \hline \hline & DCA-Z$\ge$Zmin & DCA-Z$<$Zmin \\ \hline \boldmath {2${ { \gamma }} $} \ Data & 381 & 169 \\ \boldmath {2${ { \gamma }} $} \ OBG-Data & 451 & 927 \\ \hline \hline \end{tabular} \caption{Events passing and failing the Z-cut in \boldmath {2${ { \gamma }} $} \ samples in Data and OBG-Data} \label{tab-DCA-Z} \end{table} Finally, we present the calibration of the NC-DIS contribution to the \boldmath {1${ { \gamma }} $} \ sample using the control region \boldmath {${\rm {PAN }}$ } $< 0.9$ dominated by events containing a photon-conversion accompanied by a neutral energy cluster in the ECAL. Figure~\ref{fig-Full-PAN} shows the observed \boldmath {${\rm {PAN }}$ }\ distribution for the full \boldmath {1${ { \gamma }} $} \ sample although the signal region \boldmath {${\rm {PAN }}$ } $\geq 0.9$ is not looked at till all backgrounds are finalized. The figure evinces agreement between data and the prediction for \boldmath {${\rm {PAN }}$ } $< 0.9$. Furthermore, the observed kinematic variables associated with the neutral cluster in ECAL are found to be consistent with the prediction. The observed momentum distribution of the photon is consistent with the prediction as evidenced by Figure~\ref{fig-BkgPan-Pg}. Lastly, Figure~\ref{fig-BkgPan-Zeta} shows that the measured \boldmath {$\zeta_{\gamma }$}\ distribution agrees with the MC prediction. Because the \boldmath {$\pi^0 $} -induced photons in NC-DIS will have a broader \boldmath {$\zeta_{\gamma }$}\ than those in \boldmath {Coh$\pi^0$} , we use \boldmath {$\zeta_{\gamma }$} $> 0.05$ to normalize the NC-DIS. The NC-DIS normalization factor is $1.13 \pm 0.14$. (This normalization factor has been applied to the NC-DIS in the figures.) The logic behind using this control sample to constrain the background in the signal region is that \boldmath {1${ { \gamma }} $} \ events come from $\pi^0$-dominated neutral current events where one of the photons evades detection. One concern, however, is that of a possible systematic bias when using the \boldmath {${\rm {PAN }}$ } $< 0.9$, a region where both photons are reconstructed in DC and ECAL, i.e. the two photons are emitted in the forward direction, to predict the NC-DIS in the \boldmath {${\rm {PAN }}$ } $\geq 0.9$ where the second photon misses the ECAL. To check this concern, ordinary NC events, where charged tracks define the event, and, thereby, relieving the forward bias in the \boldmath {1${ { \gamma }} $} \ reconstruction due to the trigger requirement, are selected in data and MC. A single \boldmath {1${ { \gamma }} $} \ is selected downstream of the event. The charged tracks are ignored, the \boldmath {1${ { \gamma }} $} \ is subjected to the current analysis. The Data/MC ratio is examined in $E_\gamma$ and $\zeta_\gamma$ variables. The check reveals that the shapes of these variables in MC are consistent with those of data: the ratio is unity within $\pm 5\%$ in the kinematic range, well covered by the 12\% error ascribed to the NC-DIS. Table~\ref{tab-FinalSel} presents the NC-DIS contribution to the \boldmath {1${ { \gamma }} $} \ sample. To sum up, we have a calibrated prediction of \boldmath {Coh$\pi^0$} , OBG, and NC-DIS contributions to the \boldmath {1${ { \gamma }} $} \ sample. Table~\ref{tab-FinalSel} lists the corresponding errors, all statistical in nature, as determined by the respective data control samples. Additionally, we estimate the error in the \boldmath {1${ { \gamma }} $} \ reconstruction to be $\leq 2.5\%$, which has a negligible effect on this analysis. As a final check, we reproduced the results presented in the ~\cite{NOMAD-HNU}. This exercise illustrates that the current analysis is two orders of magnitude less stringent than that in ~\cite{NOMAD-HNU} which was focussed on the search for a heavy neutrino coupling to $\nu_\tau$. \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth] {1v0_p16_pasymneut.pdf} \caption{Comparison of \boldmath {${\rm {PAN }}$ }, the energy asymmetry between the \boldmath {1${ { \gamma }} $} -momentum and the neutral energy in the ECAL, between Data (Symbol) and prediction: \boldmath {Coh$\pi^0$} \ (Blue-hatched), OBG (Green-dash-dot), NC-DIS (Red-dash), and total (Black-histogram). The signal region, \boldmath {${\rm {PAN }}$ } $\geq 0.9$, is not looked at till the analysis is complete.} \label{fig-Full-PAN} \end{center} \end{figure} \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth] {1v0_bkgpan_p03_pg_single.pdf} \caption{Comparison of $P_\gamma$ between data and MC in \boldmath {${\rm {PAN }}$ } $< 0.9$ region.} \label{fig-BkgPan-Pg} \end{center} \end{figure} \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth] {1v0_bkgpan_p08_zeta.pdf} \caption{Comparison of $\zeta_\gamma$ distribution of the \boldmath {1${ { \gamma }} $} \ sample with \boldmath {${\rm {PAN }}$ } $< 0.9$ in data and MC. The $\zeta_\gamma > 0.05$ region is used to normalize the NC-DIS.} \label{fig-BkgPan-Zeta} \end{center} \end{figure} \section{Results} \label{sec-sig} Figure~\ref{fig-evt1v01} shows a representative event in the signal region, \boldmath {${\rm {PAN }}$ } $\geq 0.9$. The first signal-box, ${\rm {Box1}}$, with \boldmath {${\rm {PAN }}$ } $\geq 0.9$, yields 155 observed events with a predicted background $129.2 \pm 8.5 \pm 3.3$ events yielding an excess of $25.8 \pm 15.5$ events. Figure~\ref{fig-SigPan-Pg} and Figure~\ref{fig-SigPan-Zeta} present the $P_\gamma$ and $\zeta_\gamma$ comparison between data and MC. Absence of significant excess leads us to interpret this as null evidence for excess of single photon events. Assuming the error on the background-prediction to be Gaussian, we derive an upper limit of $<51$ events at 90\% CL in ${\rm {Box1}}$. The second signal-box, ${\rm {Box2}}$, with \boldmath {${\rm {PAN }}$ } $\geq 0.9$ and \boldmath {$\zeta_{\gamma }$} $\leq 0.05$, yields 78 observed events with a predicted background $76.6 \pm 4.9 \pm 1.9$ events yielding an excess of $1.4 \pm 10.3$ events. Events in the ${\rm {Box2}}$ exhibit kinematic characteristics consistent with the background prediction. The vertex-distributions of the \boldmath {1${ { \gamma }} $} \ events in the signal-region are in agreement with the MC prediction; as are, within errors, the $P_\gamma$ and $\zeta_\gamma$ distributions, shown in Figure~\ref{fig-Sig-Pg} and Figure~\ref{fig-SigPan-Zeta}. In addition, the observed collinearity (${\cal C}$) of the photon matches that of the prediction, see Figure~\ref{fig-Sig-costh}. Assuming the error to be Gaussian, we derive an upper limit of $<18$ events at 90\% CL. Table~\ref{tab-FinalSel} presents the final enumeration. \begin{table}\centering {\small{ \begin{tabular}{|||c||c|c|c|c||c|||} \hline Cut & \boldmath {Coh$\pi^0$} -RS & NC-DIS$\oplus$Res & OBG & Total & Data \\ \hline \hline \boldmath {${\rm V}^0 $} \ sample & 385.9 & 400.1 & 341.3 & 1127.3 & 1149 \\ \hline {\bf {MC Error}} & & & & & \\ & 14.5\% & 12.0\% & 7.7\% & & \\ \hline {\bf {Background}} & & & & & \\ \boldmath {${\rm {PAN }}$ }\ $< 0.9$ & 353.9 & 347.7 & 296.5 & 998.1 $\pm$69.9$\pm$25.0 & 994 \\ \hline {\bf {Signal}} & & & & & \\ \boldmath {${\rm {PAN }}$ }\ $\geq 0.9$ & 32.0 & 52.4 & 44.8 & 129.2$\pm$8.5$\pm$3.3 & 155 \\ \boldmath {${\rm {PAN }}$ }\ $\geq 0.9$ $\oplus$ \boldmath {$\zeta_{\gamma }$} $\leq 0.05$ & 22.8 & 22.6 & 31.2 & 76.6$\pm$4.9$\pm$1.9 & 78 \\ \hline \hline \end{tabular \caption{The \boldmath {1${ { \gamma }} $} \ Sample: Presented are the normalized \boldmath {Coh$\pi^0$} , NC-DIS, and OBG predictions for the \boldmath {1${ { \gamma }} $} \ sample. The MC errors of the three components are all statistical in nature, as determined by the respective control samples. The systematic error due the \boldmath {${\rm V}^0 $} -reconstruction is shown under the `Total' column. Data are shown in the last column.} \label{tab-FinalSel} }} \end{table} The only remaining task is to set an upper limit on the rate of single $\gamma$ events. To determine the signal efficiency, we assume that the `signal' photon has kinematics similar to that of one of the photons from the \boldmath {Coh$\pi^0$} \ interaction (the other photon is removed from the simulation). Using the RS \boldmath {Coh$\pi^0$} -model, the signal efficiency for ${\rm {Box1}}$, \boldmath {${\rm {PAN }}$ }$\geq 0.9$, is 8.8\%. The efficiency for ${\rm {Box2}}$, \boldmath {${\rm {PAN }}$ }$\geq 0.9$ and $\zeta \leq 0.05$, is 8.0\%. The number of fully corrected \nm-CC in the same fiducial volume is measured to be $1.44 \times 10^{6}$. We obtain the following upper limits on the rate of single photon events in $\nu$-interactions: \begin{equation} \frac { \sigma {\rm { (Single-\gamma) }} } {\sigma (\nu_\mu {\cal A} \rightarrow \mu^- X)} < 4.0 \times 10^{-4} {\rm (90\% CL)} \label{eq-ccrat1} \end{equation} \begin{equation} \frac { \sigma {\rm {(Single, Forward-\gamma) }} } {\sigma (\nu_\mu {\cal A} \rightarrow \mu^- X)} < 1.6 \times 10^{-4} {\rm (90\% CL)} \label{eq-ccrat2} \end{equation} \vskip 0.75cm \noindent In summary, we have presented a search for single photon events in interactions of neutrinos with average energy $E_{\nu} \simeq 25$ GeV. All relevant backgrounds are constrained using data control samples. No significant excess is seen. Assuming that the hypothetical signal has kinematics similar to those of a photon from the \boldmath {Coh$\pi^0$} \ interaction, the analysis imposes an upper limit on the rate of excess of single photon events of $< 4.0 \times 10^{-4}$ per \nm-CC at 90\% CL; with an additional soft collinearity cut (permitting 90\% of $\gamma$ from \boldmath {Coh$\pi^0$} ) the limit is $< 1.6 \times 10^{-4}$ per \nm-CC at 90\% CL. Following the report on superluminal neutrinos by the OPERA collaboration ~\cite{OPERA}, we are conducting a specialized search for very forward $e^-e^+$ pairs. \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth] {1v0_sigpan_p03_pg_single.pdf} \caption{Comparison of $P_\gamma$ between data and MC in ${\rm {Box1}}$, \boldmath {${\rm {PAN }}$ } $\geq 0.9$ region.} \label{fig-SigPan-Pg} \end{center} \end{figure} \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth] {1v0_sigpan_p07_zeta.pdf} \caption{Comparison of $\zeta_\gamma$ distribution between data and MC in ${\rm {Box1}}$, \boldmath {${\rm {PAN }}$ } $\geq 0.9$ region.} \label{fig-SigPan-Zeta} \end{center} \end{figure} \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth] {1v0_sig_p03_pg_single.pdf} \caption{The Momentum of \boldmath {1${ { \gamma }} $} \ in ${\rm {Box2}}$: The observed distribution is consistent with the MC-prediction.} \label{fig-Sig-Pg} \end{center} \end{figure} \clearpage \newpage \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth] {1v0_sig_p11_1-costh__rebin1.pdf} \caption{Consistent collinearity of \boldmath {1${ { \gamma }} $} \ in data and MC in the ${\rm {Box2}}$. } \label{fig-Sig-costh} \end{center} \end{figure} \section*{Acknowledgments} We extend our grateful appreciations to the CERN SPS staff for the magnificent performance of the neutrino beam. We (CTK and SRM) warmly thank Bill Louis, Richard Hill, Chris Hill, James Jenkins and Terry Goldman for many stimulating discussions and insights. The experiment was supported by the following agencies: ARC and DIISR of Australia; IN2P3 and CEA of France, BMBF of Germany, INFN of Italy, JINR and INR of Russia, FNSRS of Switzerland, DOE, NSF, Sloan, and Cottrell Foundations of USA, and VP Research Office of the University of South Carolina.
\section{About the trend in azimuthal velocities}\label{sec:Vtrend} Among the various links between chemistry and kinematics in the solar neighbourhood is the trend in mean azimuthal velocity of stars with increasing metallicity for objects belonging chemically to the thin disc. The observational evidence for this trend can be traced back to \cite{Haywood08}, who found that metal-poor stars at low $[\alpha/\hbox{Fe}]$ in local samples have large angular momentum. As laid out in \cite{SBII} the low $[\alpha/\hbox{Fe}]$ stars constitute the younger part of the Galactic disc \citep[cf.][]{Matteucci90}. The \cite{SBI} model cannot make a firm prediction about such trends for the high $[\alpha/\hbox{Fe}]$ stars that formed at early epochs. It is still too uncertain what the radial abundance gradient and the radial behaviour of star formation approximately looked like. In particular either an inverse radial metallicity gradient at early times or different setups of inside-out formation can give rise to an inverse (i.e. positive) gradient of mean azimuthal velocity with $\hbox{[Fe/H]}$ for the older populations. We reserve the discussion of such possibilities to a paper in progress. Yet for the younger, low $[\alpha/\hbox{Fe}]$ stars very firm predictions are available \citep[for a discussion of these effects in an N-Body model see][]{Loebman11}. Observations imply a very flat age-metallicity relation during the past couple of Gyrs. Hence the links between chemistry and kinematics are dominated by the radial abundance gradient in the interstellar medium: With increasing age stellar populations acquire random energy, which brings them on more eccentric orbits letting them oscillate further away from the circular orbits defined by their angular momentum. We call this "blurring". As angular momentum is conserved in an axisymmetric potential, we see in the solar neighbourhood stars from inner radii at low azimuthal velocities and vice versa, so that the radial abundance gradient gives rise to a significant downtrend of mean azimuthal velocity with metallicity. Radial migration shuffles stars in angular momentum, a process called "churning", but the populations should still contain some residual information on where they were born: as stars from the inner Galaxy get scattered to larger angular momenta from where they can visit the Solar neighbourhood, the metal-rich population is still preferentially concentrated towards the inner Galactocentric radii, while the metal-poor outer disc populations are preferentially at large angular momentum. The explanation in \cite{SBII} focused on the fact that radial migration is largely responsible for us being able to find those migrated objects in large numbers at all and on that we can observe this gradient in contrast to classical expectation where the increasing asymmetric drift with age should have given an inverse gradient from an sloping age-metallicity relation. Unfortunately it evoked the widespread misconception that radial migration is responsible for the downtrend of azimuthal velocities with metallicity. This is not true, because in fact radial migration tends to weaken the observed trend. We can illustrate this by two extreme cases: Assume a radial abundance gradient of $-0.06 \,{\rm dex}/\,{\rm kpc}$ \citep[see][]{Luck11} in the young disc. In the first case we assume that stars do not change their initial angular momentum and that there is insignificant scatter of stellar abundances at birth at a certain Galactocentric radius $R_g$. We can hence write for the young population: \begin{equation} \hbox{[Fe/H]}(R_g) = \hbox{[Fe/H]}_0 + \frac{d\hbox{[Fe/H]}}{dR_g} (R_g - R_0) \end{equation} where $R_0 = 8.2$ and $\hbox{[Fe/H]}_0$ denote the Solar galactocentric radius and the local metallicity. Assuming a disc with constant circular velocity of $V_c = 233 \,{\rm km}\,{\rm s}^{-1}$ we can then replace the galactocentric radius $R_g$ using angular momentum conservation and resolve the linear relationship for the local azimuthal velocity $V$: \begin{equation} V_\phi(\hbox{[Fe/H]}) = V_c\frac{R_g}{R_0} = V_c \left(1 + \frac{R_g-R_0}{R_0} \right) = V_c \left(1 + \frac{\hbox{[Fe/H]} - \hbox{[Fe/H]}_0}{R_0}\frac{dR_g}{d\hbox{[Fe/H]}}\right) \end{equation} and for the chosen parameters we obtain the slope in $V_\phi$ (on the quite meagre baseline in metallicity that would be observable): \begin{equation} \frac{dV_\phi}{d\hbox{[Fe/H]}} = \frac{V_c dR_g}{R_0 d\hbox{[Fe/H]}} = -474 \,{\rm km}\,{\rm s}^{-1} / \,{\rm dex} . \end{equation} The result would be about $dV_\phi / d\hbox{[Fe/H]} \sim -300 \,{\rm km}\,{\rm s}^{-1} / \,{\rm dex}$ for the values used in \cite{SBI}, largely due to the steeper metallicity gradient assumed there. Measurement errors in metallicity would blur out the baseline and reduce the measured gradient, but this will not remove the stark contrast to what is observed in the data, as well as contributions from age-dependent evolution of the Galactic disc can hardly remove all difference on the super-solar metallicities. Radial migration with its redistribution in angular momentum, however, does a huge change to this relation: If we assume the other extreme case of complete radial mixing, the gradient will vanish, as the distribution of populations does not depend any more on where the stars have been born. So we see that in general radial migration reduces the gradient and is both detectable by the expansion of the baseline in metallicity and by the reduced gradient. The presence of a weak velocity-metallicity correlation in data confirms that there is considerable radial migration, but the mixing is not complete. \begin{figure} \begin{center} \epsfig{file=graphs/schoenrich_r_fig1.eps, angle=-90, width=\hsize} \caption[The metallicity plane of the Borkova et al. sample]{The metallicity plane in $[\hbox{Mg}/\hbox{Fe}]$ versus $\hbox{[Fe/H]}$ of the Borkova et al. sample. The depopulated region in the data advises a cut somewhere between $[\hbox{Mg}/\hbox{Fe}] = 0.14$ and $[\hbox{Mg}/\hbox{Fe}] = 0.18$ where we set the cut.} \label{fig:Borkoverview} \end{center} \end{figure} Let us examine those trends on a real data set. From the model of \cite{SBI} we expect that if we do not discard alpha rich stars the thick disc velocities will moderate any trend in the mean azimuthal velocities with mean metallicities. In the model they balance the young disc trend at intermediate metallicities, because the downtrend for the younger low $[\alpha/\hbox{Fe}]$ stars is compensated for by the increasing number of high asymmetric drift, high $[\alpha/\hbox{Fe}]$ stars of the chemical thick disc towards lower metallicities. In the lower metallicity regime ($\hbox{[Fe/H]} < -1.0$) we even encounter a mildly positive slope in the model, because of some peculiarities in early disc formation that will be discussed in an upcoming paper. In this light it is essential to get a clean cut in the metallicity plane, removing the high $[\alpha/\hbox{Fe}]$ population. The following exercise was first presented at the IAU assembly, but concerning the need for re-interpretation it seems appropriate to show it again. \figref{fig:Borkoverview} shows the metallicity plane as defined by $[\hbox{Mg}/\hbox{Fe}]$ as alpha element against $\hbox{[Fe/H]}$ in the \cite{Borkova05} sample. This sample is a homogenised compilation of high-resolution spectroscopic data from several studies of local stars \citep[e.g.][]{Bensby05, Fuhrmann04, Reddy03, Reddy06}. One should keep in mind a major caveat against over-interpretation of any data we use here: Almost all local high-resolution data have strong kinematic selection biases for increasing or decreasing the number of selected thin/thick disc stars. It should also be kept in mind that because of the strong correlations between single velocity components and also between metallicities and kinematics, any such sample is uncontrollably biased both in metallicities and in velocity space. Yet, it is still interesting to look at these trends, which are confirmed on GCS data in these proceedings by L. Casagrande. As advocated by the depopulated region at intermediate $[\hbox{Mg}/\hbox{Fe}]$ we cut the sample to keep only objects with $[\hbox{Mg}/\hbox{Fe}] < 0.18$, to get a chemical thin disc selection. In light of the observational errors there might be some residual contamination, which seems to be confirmed when we measure the gradient of metallicity against angular momentum: Since the local azimuthal velocities can be translated directly into the angular momentum of those stars and let us this way estimate the stellar guiding centre radii, a local sample can be used as a good indicator of the radial abundance gradient. For our cut we estimate $d\hbox{[Fe/H]}/dR = -0.04 \,{\rm dex}/\,{\rm kpc}$ a value that rises up to $-0.1 \,{\rm dex}/ \,{\rm kpc}$ depending on how much of the low angular momentum regime we remove from the sample and how far down in $[\hbox{Mg}/\hbox{Fe}]$ we move the cut (the radial gradient increases mildly for stricter selections which indicates that we are tossing out some residual old stars). Interestingly the gradient for the cleaner selections is a bit higher than in the \cite{Luck11} data, but this should not be taken too seriously in the light of the diverse biases in our sample. \begin{figure} \begin{center} \epsfig{file=graphs/schoenrich_r_fig2.eps, angle=-90, width=\hsize} \caption[Kinematics versus metallicities in the Borkova et al. thin disc]{Kinematics versus metallicities in the Borkova dataset in the subsample of stars with $[\hbox{Mg}/\hbox{Fe}] < 0.18$ (red data points). The blue line gives the linear fit of the sample compared to the binned heliocentric mean velocities shown in green. The purple line shows the velocity dispersions.}\label{fig:BorkV} \end{center} \end{figure} The trend of mean heliocentric $V$ velocity against metallicity in the low $[\hbox{Mg}/\hbox{Fe}]$ subsample is examined in \figref{fig:BorkV}. A linear fit (green line) yields a highly significant slope of $dV_\phi/d\hbox{[Fe/H]} = (-31.5 \pm 4.2) \,{\rm km}\,{\rm s}^{-1} / \,{\rm dex}$ and is plotted onto the data. The observed trend is a bit higher than what was found by \cite{Lee11} -- this can most likely be traced to the larger $[\alpha/\hbox{Fe}]$ errors of the SEGUE \citep[][]{Yanny09} sample with its low resolution spectra that likely results in larger contamination of the low $[\alpha/\hbox{Fe}]$ part with high $[\alpha/\hbox{Fe}]$ stars. Surprisingly our finding is not consistent with the result of \cite{Navarro11} on essentially the same data. The same result as with the linear fit is seen from the binned means, where we divided the sample into $0.1 \,{\rm dex}$ wide subsets. At the same time we plot the dispersions for each bin (purple line). We can compare this with the qualitative predictions of the different chemical evolution models: In classical chemical evolution models \citep[e.g.][]{Chiappini01} the thin disc density ridge is created by the local population running along this ridge from low to high metallicities. This implies a significant age trend with metallicity and hence via the age-dispersion relation the most metal poor stars should show the highest velocity dispersions with a clear downtrend towards the metal rich objects. This is not observed, while the data are consistent with the behaviour in the radial migration models that favour dispersions being rather flat with metallicity at low $[\alpha/\hbox{Fe}]$. Similarly the mean V velocity trend is hard to explain in the classical framework, but straight forward in the framework of radial migration models. Yet there is a little problem: The trend of V velocities with metallicity is considerably higher in the model (predicting up to $50 \,{\rm km}\,{\rm s}^{-1} / \,{\rm dex}$) than it is in the data. When we look at the above implications of the velocity gradient on metallicities, it becomes clear what went wrong in \cite{SBI}. When we set the radial abundance gradient in the model, we had to rely on older abundance gradient data that now seem to have indicated a too steep abundance gradient. At least it is significantly steeper than the value derived by \cite{Luck11}. The overestimated gradient places one of the main findings of the first paper on firm grounds, namely that radial migration is necessary for explaining the local metallicity distribution. Yet, it appears that with the too steep gradient we underestimated the need for stellar radial migration: The lower the gradient, the farther stars need to migrate in order to yield the same width of the local metallicity distribution. Strengthened migration should then result in a shallower slope of velocities with metallicity. As a side effect the stronger migration would result in a locally stronger and larger scale-height thick disc. The standard model from \cite{SBI} was already rather on the upper edge on how strong the thick disc may be. Again things fit together as on the decision between vertical energy and vertical action as conserved quantity \cite{SBI} decided for vertical energy. Solway et al. (in prep) and also the recent paper of \cite{Bird11} showed that vertical action is rather conserved than vertical energy. This will deliver some reduction in thick disc scale heights that can balance the stronger thick disc arising from increased radial migration. An impediment to modelling has so far been that despite its advantages the classic adiabatic approximation and also adiabatic modelling of stellar populations as put forward in \cite{B10} violate total energy conservation. The main effect of this short-coming is a significant underestimate of the Galactocentric radii populated by an orbit or in other words an under-prediction of inner disc stars in the intermediate and outer disc regions. Two solutions to this problem have been proposed: angular momentum correction by \cite{BM11} and the adiabatic potential of \cite{SBIII}, which directly restores total energy conservation by correcting the horizontal potential for the energy removed from the vertical motion. This solution pushes the high vertical energy orbits back to the outer disc, compensating locally for some of the difference between isothermal approximations and the adiabatic approximation. We are currently updating the \cite{SBI} model to include the adiabatic potential. A definitive treatment of the $([Fe/H],V_\phi)$ relation must await completion of this project. \begin{figure} \begin{center} \epsfig{file=graphs/schoenrich_r_fig3.eps, angle=-90, width=\hsize} \caption[Eccentricity distributions from the standard model at different altitudes]{The eccentricity distribution from \cite{SBI} at different altitudes above the plane at Solar Galactocentric radius. We used simple population masses, i.e. no selection function was applied.} \label{fig:eccdist} \end{center} \end{figure} \section{About eccentricity distributions} In recent studies starting with \cite{Sales09} and e.g. later \cite{Lee11} the comparison of eccentricity distributions has become a frequently used tool to assess possible scenarios for the history of the Galactic disc. \figref{fig:eccdist} shows the eccentricity distribution from the \cite{SBI} model without change of any parameter and applying no selection function, i.e. using the population masses as weight. The result looks quite similar to \cite{Lee11}, yet we do not want to book this as a win, as we would like to advocate against the use of eccentricity distributions for several reasons: First it is counter-intuitive to fold the wealth of information we have into a single quantity that hides essential information like the angular momentum distribution of the disc. As another point one needs to perform a real propagation of distance, proper motion and line-of-sight velocity errors that to date does not appear to have been fully carried out. This would be computationally expensive, because in contrast to simple velocity space errors have to be calculated on the orbital model from which the eccentricities must be derived. More important the derived eccentricity values depend on the assumed potential, which governs the orbit extension of a star derived from its estimated position and velocity. Another Achilles' heel of the method -- that comparisons of eccentricities utilise a different potential for the calculation of orbits than the theoretical models they compare to -- has in most cases been neglected. One may of course argue that eccentricities express the general circularity of orbits in the different approaches. To some part this is true, but there we come to a far more serious problem: The heating in models and even principal heating mechanisms are very weakly constrained \citep[see e.g. the discussion in][]{AB09}. Locally the eccentricities are quite directly related to azimuthal velocities in that high asymmetric drifts or respectively low rotational velocities imply large eccentricities. So we see the large uncertainty in the expected eccentricities by looking at the major changes in Fig. 2 of \cite{SBIII} that are induced by moderate variation in the assumed parameters of the disc. For example heating up the inner disc of the Galaxy a bit more, which is covered either by heating from a bar or respectively simply in the uncertainties intrinsic to molecular clouds as source of random energy, the number of high eccentricity visitors from the inner disc increases significantly. This would in the one-dimensional comparison be nearly indistinguishable from high eccentricity stars contributed e.g. by a minor merger in the outer regions of the disc. Summed up there may be physical entities (e.g. domination of the Solar neighbourhood by debris from some satellite) that can be ruled out by analysis of eccentricity distributions (though it is questionable why this should not be possible on full velocity space as well). However, we see that a large range of eccentricity distributions can be explained just by uncertainty in heating and other Galactic parameters, and moreover the use of eccentricity distributions in isolation hides a lot of valuable information. \section{Conclusions} Analysis of local spectroscopic samples reveals for the chemically young disc a significant downtrend of mean azimuthal velocities $V_\phi$ with metallicity, while there appears to be no notable change in velocity dispersion. Those samples have to be used with caution as they carry strong kinematic biases, but the observations are confirmed by kinematically unbiased data from \cite{CSA11}. While the qualitative picture is nicely consistent with predictions of radial migration models and has so far not been explained in the context of classical approaches, we stress that in particular the azimuthal velocity trend is not a consequence of radial migration of stars by itself. On the contrary the velocity trend is a natural consequence of the radial abundance gradient in the galactic disc. Radial mixing expands the baseline in metallicity on which the trend is observable, but moderates its slope. The fact that the observed slope is a bit smaller than predictions from the \cite{SBI} model can be reasoned by an underestimate of the amount of mixing that was caused by a too steep radial abundance gradient adopted in that approach. It is a pleasure to thank James Binney for helpful comments.
\section{Introduction} \label{sec:intro} Weak lensing has emerged as a powerful tool in cosmology and astrophysics (e.g., \cite{Refregier,MunshiEtal,HoekstraJain,MasseyKitching,SchrabbackEtal}). It probes the underlying total matter distribution, and can be used to tie the observed distribution of galaxies to that of the dark matter, the dominant component of the cosmic matter budget. So far, weak lensing has primarily been measured using the shear, through the statistical correlation of galaxy ellipticities (see \cite{BartelmannSchneider} for a review). In addition, weak lensing magnification has been detected through its effect on the number density of a flux-limited sample (magnification bias; \cite{ScrantonEtal05,HildebrandtEtal09}). Shear $\gamma$ and convergence $\k$, which are related to the magnification $\mu$ through $\mu = [(1-\kappa^2) + |\gamma|^2]^{-1}$, so that $\mu = 1+2\kappa$ to linear order, measure different properties of the density field. For an azimuthally symmetric lens, \ba \gamma =\:& \frac{\Delta\Sigma}{\Sigma_{\rm cr}};\quad \k = \frac{\Sigma}{\Sigma_{\rm cr}}\\ \Sigma_{\rm cr} =\:& \frac{c^2}{4\pi G}\frac{D_s}{D_L\,D_{Ls}} \label{eq:Scr}\\ \Delta\Sigma =\:& \bar\S(< r) - \S(r), \ea where $\Sigma(r)$ is the projected surface mass density of the lens, $r$ is the physical transverse distance on the lens plane, and $D_s,\:D_L,\:D_{Ls}$ denote the angular diameter distance out to the source, lens, and between lens and source, respectively. Note that the shear is proportional to the differential surface mass density $\Delta\S$, while the convergence probes $\S(r)$ itself. The different radial dependences of $\k$ and $\gamma$ can be used to break degeneracies between the lens mass and density profile shape \citep[][]{RozoSchmidt}. The convergence $\k$ has previously been measured using the observed number density of background sources. In a flux-limited survey, the observed number density is given by \begin{equation} n_{\rm obs} = \bar n [ 1 + \d_g + (5s-2) \k], \end{equation} where $\bar n$ is the mean number density, $\d_g$ denotes the intrinsic fluctuation in galaxy density, and $s = \partial\ln\bar n/\partial m_{\rm cut}$ for a sharp magnitude cut $m_{\rm cut}$. This can be generalized to surveys that are also limited by a size cut \citep[][]{schmidtetal09}. By cross-correlating foreground lenses with background sources widely separated in redshift, one can isolate this effect to measure $\k$. Note that the intrinsic fluctuations $\d_g$ act as noise, and any residual physical correlation with the lenses (e.g., due to photo-z uncertainties) will contaminate the signal. In this paper, we report on a detection of magnification around X-ray selected galaxy groups in COSMOS using galaxy sizes and fluxes. Related approaches have been proposed in \cite{jain02,BartelmannNarayan,BertinLombardi}; see also \cite{MenardEtal10} who used individual quasar magnitudes, and \cite{HuffGraves}. We rely on the observed properties of individual galaxies and do not use the number density of galaxies as done in magnification bias measurements. One key advantage of this technique is that the contamination due to residual physical correlation with the lenses is likely much smaller than in magnification bias measurements, due to the weaker correlation of galaxy sizes and luminosities with environment (\cite{MaltbyEtal,NairEtal,CrotonEtal}). Throughout, we adopt a flat $\Lambda$CDM cosmology with $h=0.72,\; \Omega_m = 0.258$ to calculate distances. Halo masses are defined through a mean interior density of $200\rho_{\rm crit}(z=0.38)$, where $z=0.38$ is the mean redshift of the lens sample. \section{Magnification Estimator} \label{sec:est} The data set we use consists of a set of values $\{d_i,m_i,z_i\}$ for each galaxy, denoting the size, magnitude, and photometric redshift. Let us consider how these observables change under lensing magnification. Throughout, we will work in the weak lensing regime. We can then write \ba d =\:& (1 + \eta\:\k) d_0\label{eq:dlens}\\ m =\:& m_0 + q \k\label{eq:mlens}\\ z =\:& z_0. \ea Here, a subscript $0$ denotes unlensed quantities, and in the last line we have assumed that photometric redshifts are not affected by lensing. In absence of noise and instrumental effects, the efficiency factors are $\eta = 1$ and $q = -5/\ln 10 \approx -2.17$, respectively (e.g., \cite{schmidtetal09}). In practice, these values can be different due to the instrument PSF, and other effects and in general depend on the properties of a given galaxy. In order to estimate $\k$ using \refeqs{dlens}{mlens}, we need to know the unlensed quantities $d_0, m_0$. These have some intrinsic distribution, and given the known distribution we can infer $\k$ statistically. Our starting point for the estimator is thus the joint distribution $p_0(d_0,m_0|z)$ of unlensed galaxy sizes and magnitudes at fixed redshift; the estimator is designed to \emph{not} use the observed number density of galaxies. We approximate $p_0$ as Gaussian in $m$ and $x \equiv \ln d$. Furthermore, we approximate the error distributions in $m$ and $x$ as Gaussian. This greatly simplifies the evaluation of the estimator, at the price of making it sub-optimal. The log-likelihood is then given by \ba -2 \ln P\left(\S\right) =\:& \sum_i \frac{1}{1-\rho^2(z_i)} \label{eq:LL}\\ \times& \Bigg\{ \frac{1}{\sigma_{x,i}^2} \left(x_i - \overline{x}(z_i) - \tilde\eta_i F_i\S\right)^2 + \frac{1}{\sigma_{m,i}^2}\left(m_i - \overline{m}(z_i) - \tilde q_i F_i\S\right)^2\vs - & 2 \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}} \left(x_i-\overline{x}(z_i)-\tilde\eta_iF_i\S\right)\ \left(m_i - \overline{m}(z_i) - \tilde q_i F_i\S\right) \Bigg\} \vs \sigma_{x,i}^2 =\:& \sigma_{x,{\rm meas},i}^2 + \sigma_{x,\rm int}^2(z_i) + \sigma_{xz,i}^2\\ \sigma_{m,i}^2 =\:& \sigma_{m,{\rm meas},i}^2 + \sigma_{m,\rm int}^2(z_i) + \sigma_{mz,i}^2\\ \tilde\eta_i =\:& \eta(d_i) + \varepsilon_x(z_i)\\ \tilde q_i =\:& q(m_i) + \varepsilon_m(z_i)\\ F_i =\:& F(z_i) = \Sigma^{-1}_{\rm cr}(z_i). \ea The total variance in $x$ and $m$ is given by the sum in quadrature of the measurement uncertainty, intrinsic dispersion, and propagated photometric redshift uncertainty. $\sigma_{x,\rm meas},\;\sigma_{m,\rm meas}$ are measured as a function of $d$ and $m$, respectively, through multiple images of the same galaxy from overlap regions of the HST ACS data \citep[][]{cosmoswl}. The propagated redshift uncertainty is approximated as ($a = x,\;m$) \begin{equation} \sigma_{az,i} = \frac{1}{2} |\bar a(z_{+,i})-\bar a(z_{-,i})|, \end{equation} where $z_{\pm,i}$ are the upper and lower 68\% confidence level values for the redshift of galaxy $i$. The corrections $\varepsilon_x,\;\varepsilon_m$ to $\tilde\eta$ and $\tilde q$ take into account the fact that a positive magnification pushes faint, small galaxies over the flux and size thresholds (\emph{lensing bias}). A similar effect exists for shear, but only at second order \citep[][]{schmidtetal09b}, while the magnification estimator is affected at leading order. The correlation between intrinsic galaxy sizes and magnitudes is quantified by the correlation coefficient $\rho$. The maximum-likelihood estimate $\hat\Sigma$ and its estimated variance are then given by \begin{equation} \hat\Sigma = \frac{\sum_i A_i\,F_i}{\sum_i B_i\,F_i^2}\;;\quad \sigma^2(\hat\Sigma) = \left(\sum_i B_i\,F_i^2\right)^{-1}, \end{equation} where \ba A_i \equiv\:& \frac{1}{1-\rho^2(z_i)}\Bigg\{ \frac{\tilde\eta_i}{\sigma_{x,i}^2} (x_i - \overline{x}(z_i)) + \frac{\tilde q_i}{\sigma_{m,i}^2} (m_i - \overline{m}(z_i))\vs &\hspace*{1.6cm} - \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}} \left[\tilde\eta_i (m_i - \overline{m}(z_i)) + \tilde q_i (x_i - \overline{x}(z_i)) \right] \Bigg \}\\ B_i \equiv\:& \frac{1}{1-\rho^2(z_i)}\left\{ \frac{\tilde\eta_i^2}{\sigma_{x,i}^2} + \frac{\tilde q_i^2}{\sigma_{m,i}^2} - 2 \frac{\rho(z_i)}{\sigma_{x,i} \sigma_{m,i}}\tilde\eta_i \tilde q_i \right\}. \ea The intrinsic dispersions $\sigma_{x,\rm int},\, \sigma_{m,\rm int}$ as well as $\rho$ are determined by fitting a bivariate Gaussian distribution to all galaxies before cuts on $x,\,m$ in a redshift slice. We subtract the mean measurement error in $x,\,m$ in quadrature from the total measured dispersion to obtain $\sigma_{x,\rm int},\,\sigma_{m,\rm int}$. The quantities $\overline{x},\,\overline{m}$ are determined by solving the equation $\hat\Sigma=0$ applied to all galaxies in a given redshift slice. Since $F\approx\rm const$ within a narrow redshift slice, this yields \ba &\sum_i \frac{1}{1-\rho^2(z_i)}\left\{ \frac{\tilde\eta_i}{\sigma_{x,i}^2} (x_i - \overline{x}(z_i)) - \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}} \tilde q_i (x_i - \overline{x}(z_i)) \right\} = 0\\ & \sum_i \frac{1}{1-\rho^2(z_i)}\left\{ \frac{\tilde q_i}{\sigma_{m,i}^2} (m_i - \overline{m}(z_i)) - \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}} \tilde\eta_i (m_i - \overline{m}(z_i)) \right\} = 0. \ea $\overline{x},\,\overline{m}$ do not correspond to the true mean intrinsic quantities, since only galaxies passing cuts are used to measure them. We can then write \ba \bar a(z) =\:& \left(\sum_i w_{a,i}\right)^{-1} \sum_i w_{a,i}\: a_i\label{eq:xbar},\\ w_{x,i} =\:& \frac{1}{1-\rho^2(z_i)}\left\{ \frac{\tilde\eta_i}{\sigma_{x,i}^2} - \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}}\tilde q_i \right\}\\ w_{m,i} =\:& \frac{1}{1-\rho^2(z_i)}\left\{ \frac{\tilde q_i}{\sigma_{m,i}^2} - \frac{\rho(z_i)}{\sigma_{x,i}\sigma_{m,i}}\tilde\eta_i \right\}, \ea where the sum runs over all galaxies passing cuts in the redshift slice. In order to determine $\varepsilon_x,\,\varepsilon_m$, we repeat the measurement of $\overline{x},\,\overline{m}$ with magnitude and size cuts varied by $\pm\Delta m,\;\pm\Delta x$, respectively, where $\Delta m = q(m_{\rm cut}) \k_0$ and $\Delta x = \eta(d_{\rm cut}) \k_0$, and we take $\k_0=0.02$ (the value of $\varepsilon_i$ does not depend significantly on $\k_0$). Then, \ba \varepsilon_x(z) =\:& \frac{\overline{x}_+(z) - \overline{x}_-(z)}{2\k_0}\\ \varepsilon_m(z) =\:& \frac{\overline{m}_+(z) - \overline{m}_-(z)}{2\k_0}, \ea where $\bar a_{\pm}$ indicate quantities measured when using cuts $m_{\rm cut}\pm \Delta m$ and $x_{\rm cut}\pm\Delta x$. In order to take into account interdependencies, we iterate the measurement of $\overline{x}(z),\,\overline{m}(z)$ until convergence. \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth,clip=true,trim=0cm 0.3cm 0cm 0cm] {xbar.eps} \includegraphics[width=0.49\textwidth,clip=true,trim=0cm 0cm 0cm 0cm] {mbar.eps} \caption{\textit{Upper panel:} Average source galaxy log size $\overline{x}(z)$ as a function of redshift, measured using \refeq{xbar}. The shaded band shows $\sigma_{x,\rm int}$. \textit{Lower panel:} Same, for source galaxy magnitudes ($\overline{m}(z)$, $\sigma_{m,\rm int}$). \label{fig:xbar}} \end{figure} \section{Data} \label{sec:data} Our data set consists of the HST ACS imaging of the COSMOS field \citep[][]{Koekemoeretal}, from which we take the galaxy sizes $d$ (in pixels of 0.03'') and magnitudes $m$ in the F814W band. For the size $d$ we use the variable-width Gaussian-filtered second moment of \cite{RRG}, while magnitudes $m$ are \texttt{MAG\_AUTO} from SExtractor \citep[][]{Sextractor}. We leave more sophisticated, optimized size and flux estimators for future work. The details of the reduction pipeline are described in \cite{cosmoswl}. The size estimator is PSF-deconvolved, and we thus expect systematic effects of PSF variations to be negligible. In order to avoid any contamination by a background-dependence of the estimated sizes, we only use sky regions within the lower peak (mean pixel background $< 0.0062$) of the essentially bimodal background distribution in the COSMOS ACS data, comprising $\sim 75$\% of the field (see Fig.~8 in \cite{cosmoswl}). For our source galaxy sample, we exclude galaxies in masked regions \citep[][]{IlbertEtal} and apply the cuts on size, magnitude and redshift summarized in \reftab{cuts}. The size cut approximately corresponds to a measured size (before PSF deconvolution) of 1.4 times the PSF size. Note that in the shear analysis, a significantly stricter cut ($d>3.6$) is employed. The magnitude cut is mostly determined by the reliability of the photometric redshifts. Finally, the redshift cut excludes a small portion of the sample which does not contribute significantly to the lensing signal-to-noise. We obtain a sample of $N_{\rm src}=250,500$ galaxies with a median redshift of 1.22. In the lensing analysis, we also impose the conditions $z_{i,-} > z_L$, $z_i > z_L+0.2$ for each source-lens pair. We apply the fitting procedure described in the last section, using overlapping redshift slices of half-width $\Delta z = 0.25$ separated by $\Delta z$. \reffig{xbar} illustrates the measured $\overline{x}(z),\; \overline{m}(z)$ as well as $\sigma_{x,\rm int},\; \sigma_{m,\rm int}$. We fix $\rho=-0.6$ to improve the stability of the fit. Allowing $\rho$ to vary does not significantly improve the fit. \reffig{eff} shows $\varepsilon_x$ and $\varepsilon_m$. Note that $\varepsilon_x < 0$: the number of faint, small galaxies passing the cuts for $\k > 0$ reduces the average size, counteracting the lensing effect on the sizes themselves. The analogous holds for magnitudes, where $\varepsilon_m > 0$. This effect can in principle be avoided by cutting on a quantity not affected by lensing (e.g., surface brightness) instead of magnitude and size. We leave the exploration of this for future work. \begin{table}[b] \centering \caption{Summary of cuts applied to the source galaxy sample.} \label{tab:cuts} \begin{tabular}{l|c} \hline $m$ & $< 25.8$~mag \\ $d$ & $ > 2$~pixels \\ $z$ & $\in [0.5,\;4]$ \\ \hline \end{tabular} \end{table} \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth,clip=true,trim=0cm 0.15cm 0cm 0cm]{q.eps} \includegraphics[width=0.49\textwidth,clip=true,trim=0cm 0.3cm 0cm 0.1cm] {xslope.eps} \includegraphics[width=0.49\textwidth,clip=true,trim=0cm 0cm 0cm 0cm] {mslope.eps} \caption{\textit{From top to bottom:~a)} Flux lensing efficiency $q$ as a function of magnitude $m$. \textit{b)} Lensing bias correction $\varepsilon_x$ for galaxy (log) sizes. The effective size lensing efficiency is given by $\tilde\eta = \eta + \varepsilon_x$. \textit{c)} Lensing bias effect $\varepsilon_m$ for magnitudes. Correspondingly, $\tilde q = q + \varepsilon_m$. \label{fig:eff}} \end{figure} While the estimators in \refsec{est} were designed to rely as much as possible on data alone, we do need simulations to determine the lensing efficiencies $\eta,\;q$. To simulate ACS data, we use the {\tt simage} software package \citep[][]{simage1,simage2,simage3}, as previously adopted for the Shear TEsting Programme \citep[STEP;][]{step2} and to calibrate the COSMOS shear analysis \citep[][]{cosmoswl}. The parameters were tuned to mimic the galaxy source counts and correlated background noise properties of the reduced and stacked COSMOS images. We model the intrinsic morphologies of source galaxies via shapelets \citep[][]{shapelets1,shapelets2}. We have adapted {\tt simage} to enable the superposition of an input convergence $\k_{\mathrm{sim}}$ and produced images at 15 values of $\k_{\mathrm{sim}}\in[-0.2,0.3]$. To determine how each simulated galaxy responds to magnification, we match galaxies in the lensed and unlensed ($\k_{\rm sim}=0$) simulations. We measure the average $x$ and $m$ as a function of the input $\k_{\rm sim} = -0.1\dots 0.1$. Then, we estimate $\eta$ through \begin{equation} \eta = \frac{\<x\>_{\k_{\rm sim}=\k} - \<x\>_{\k_{\rm sim}=0}}{\k}, \end{equation} and correspondingly for $q$. We have found that for galaxies passing cuts, $\eta$ is within $[0.9,1.1]$ and consistent with 1 for the whole sample (recall that we use a PSF-deconvolved size estimator). For the lens sample, we use X-ray selected groups since these have been well studied using shear \citep[][]{groups}. The entire COSMOS region has been mapped by {\sl XMM-Newton}, while the central region is covered by {\sl Chandra} observations \citep[][]{HasingerEtal07,CappellutiEtal09,ElvisEtal09}. These X-ray data have been used to construct a group catalog containing 211 extended X-ray sources, spanning the range $0<z_{\rm gr}<1$ \citep[][]{Finoguenov:2007,groups}. Group membership is assigned to galaxies in the COSMOS catalog based on the photometric redshifts and the red sequence. We include only groups with reliable optical associations which have more than 3 members, and we exclude close neighboring systems and groups near the edges of the field or masked regions (\textsc{flag\_include}=1 in \cite{GeorgeEtal11a}). We use the location of the most massive group galaxy (MMGG) located within the NFW scale radius of the X-ray centroid as the group center (MMGG$_{\rm scale}$ as defined in \cite{GeorgeEtal11b}). \cite{GeorgeEtal11b} show that this is likely the most reliable estimate for the group centers. We restrict the redshift range of groups to 0.2-0.6, close to the peak of the lensing sensitivity of the source sample. This selection also reduces the potential impact of catastrophic redshift errors which become more prevalent at $z_s > 1.5$, and yields $N_{\rm gr} = 61$ groups. \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth]{kappa-MC.eps} \caption{Mean reconstructed convergence $\k$ in simulations with magnification, as a function of the input $\k_{\rm sim}$. The thin solid line shows the identity while the thick solid line shows a quadratic fit. \label{fig:MC}} \end{figure} \section{Results} \label{sec:results} Before applying our estimators to data, we performed the calibration described in \refsec{est} on the unlensed simulations, and applied the estimators to the lensed simulations as a consistency check. Since the simulations do not contain redshift information, we set $F(z_i) = 1$. \reffig{MC} shows the mean reconstructed $\k$ as a function of the true $\k_{\rm sim}$, along with a quadratic fit, \begin{equation} \<\k_{\rm rec}\> = p_0 + p_1 \k_{\rm sim} + p_2 \k_{\rm sim}^2. \end{equation} Note that we expect some quadratic correction in this relation, since our estimators neglect all second- and higher-order terms. We found $p_1=0.95\pm0.05$, indicating that our estimator is close to unbiased, and $p_2\approx 0.6$. Turning to data, we collect source galaxies in bins of physical transverse radius $r$ around the $N_{\rm gr}$ X-ray groups, and apply the estimator for $\Sigma$. The second order corrections amount to at most $\sim 10$\% at the smallest radii, as shown in \cite{groups}, and can be neglected given the errors of this measurement. Since $\overline{x}(z),\:\overline{m}(z)$ are measured in finite redshift bins without any redshift weighting while we apply redshift weighting in the $\Sigma$ measurement, we need to subtract a constant offset $\S_{\rm off}$ from the measurement. $\Sigma_{\rm off}$ is measured by pairing all galaxies passing cuts in the COSMOS field with each group, yielding $\Sigma_{\rm off} = 151.9\:M_\odot/\rm pc^2$. This is equivalent to estimating $\<\S(r)\>_{\rm groups} - \<\S\>_{\rm COSMOS}$, where both averages employ the same redshift weighting. Note that a circle of $r\sim 1$~Mpc drawn around each group covers a significant fraction of the COSMOS field; thus, we necessarily subtract some of the signal in this way. If we instead restrict the source sample used for the $\S_{\rm off}$ estimate to galaxies separated by at least 1~Mpc in physical transverse distance from each group in the sample, $\S_{\rm off}$ decreases by $4.4\:M_\odot/\rm pc^2$, thus increasing $\S(r)$ uniformly by this amount. The measured $\S(r)$ is shown in \reffig{Sigma}. The error bars are those returned by the Gaussian estimators. Measurements of the variance of $\hat\Sigma$ applied to random sets of source galaxies indicate that these errors are correct to within $\sim$15\%. For $r \lesssim 1$~Mpc, the errors for different radial bins are independent. \begin{figure}[t] \centering \includegraphics[width=0.49\textwidth]{SigmaGroup-m25.8-BiGauss-joint.eps} \caption{Projected surface mass density $\Sigma$ (filled circles) measured around groups using the estimator described in \refsec{est}. The open circles show $\Delta\S$ measured using shear for the same group sample. The solid and dotted lines show $\Sigma$ and $\Delta\S$, respectively, for the joint best-fitting NFW model ($\lg (M/M_{\odot}) = 13.43$; $r < 1$~Mpc). Note that associated large-scale structure contributes significantly to $\S(r)$ for $r \gtrsim 1$~Mpc. \label{fig:Sigma}} \end{figure} We fit an NFW profile \citep[][]{NFW} with fixed concentration $c=4$ to the measurement, calculating $\S(r)$ as described e.g. in \cite{SchmidtRozo}. Furthermore, we restrict the fitting to $r < 1$~Mpc, beyond which the contribution from associated large-scale structure becomes important (see below). This yields a best-fit mass of $\lg (M/M_\odot) = 13.25\pm 0.16$ (statistical error), corresponding to a detection significance of $\sim 3.8\sigma$. This corresponds to $\sim 38$\% of the signal-to-noise of shear. Note that the precise value also depends on the definition of the offset $\S_{\rm off}$. The best-fit mass is consistent with the estimate from shear, $\lg (M/M_\odot) = 13.49\pm 0.07$ (\cite{groups}; again for $r < 1$~Mpc). In principle, a measurement of $\Sigma$ in addition to $\Delta\S$ can break the mass-concentration degeneracy, since the two quantities have different radial dependencies \citep[][]{RozoSchmidt}. Due to the limited signal-to-noise of this first measurement, a joint fit to the $\Sigma$ measurement together with $\Delta\S$ from shear only yields an improvement of $\sim 10$\% in the area of the error ellipse in the mass-concentration plane. The $\Sigma$ measurements at $r > 1$~Mpc are consistent with an estimate of the associated large-scale structure contribution (two-halo term) using the halo model, which yields $\S_{2h} \sim 10\:M_\odot/\rm pc^2$ with weak $r$-dependence for $r < 5$~Mpc. The corresponding contribution to the shear is much smaller, $\Delta\S_{2h} \lesssim 1\:M_\odot/\rm pc^2$, further exemplifying the complementarity of the two measurements. A detailed modeling of these data will be the subject of forthcoming work. The main systematic uncertainty of the $\S$ measurement is due to the lensing bias effect. We have quantified the impact of the measurement errors on $\varepsilon_x,\:\varepsilon_m$ by varying $m_{\rm cut},\:x_{\rm cut}$, resulting in an overall multiplicative systematic error on $\Sigma$ of 20\%. Residual effects from PSF and background variations have been found to be negligible. Increasing the minimum size cut from 2 to 3 pixels yields compatible results, as do estimators using sizes and magnitudes separately. The systematic uncertainty due to photometric redshift errors is similar to that of the shear (\cite{groups}; note that we have employed the same source-lens separation criteria as in that paper). Given the lower signal-to-noise of the magnification measurement, this systematic is negligible. \section{Summary and Discussion} \label{sec:concl} We have presented the first measurement of weak lensing magnification using galaxy sizes and fluxes. Measurements of galaxy sizes and magnitudes come for free with weak lensing surveys. We find a signal-to-noise only a factor of $\sim 2-3$ less than shear; thus it is worth pursuing this kind of independent weak lensing measurement. Magnification constitutes a different observable than shear, and the non-local relationship between the two can be used to break degeneracies e.g. when measuring halo profiles. Furthermore, the systematics affecting this measurement are largely independent from those of the shear: to first order, sizes are only sensitive to the size of the PSF, while magnitudes are PSF-independent. On the other hand, a knowledge of the shape of the PSF is necessary for shear measurements. This allows us to use less restrictive cuts. Conversely, a careful calibration of the lensing bias correction is necessary for the magnification measurement, while this effect is not present in the shear at linear order \citep[][]{schmidtetal09b}. Nevertheless, given sufficiently detailed simulations, this is a solvable problem. While the measurement presented here is limited by signal-to-noise and statistics, it is clear that for future, much larger surveys, robust size and flux estimators are essential to a successful application of this method. Given the already promising results, we believe a dedicated effort to develop such estimators for weak lensing will be worthwhile. Finally, our results show that there is significant lensing information in photometric galaxy samples in addition to the shear alone; this should be taken into consideration in the design of future weak lensing experiments. \acknowledgements We thank Jessica~Ford, Hendrik~Hildebrandt, Eric~Huff, Bhuvnesh~Jain, Donghui~Jeong, Eric~Jullo, James~Taylor, and Ludovic~van~Waerbeke for helpful discussions. FS is supported by the Gordon and Betty Moore Foundation. RM is supported by an STFC Advanced Fellowship. JR was supported by JPL, run by Caltech for NASA.
\section{Introduction} Inelastic neutron scattering has come to be recognized as indispensable in modern materials science, because a material's spin and lattice dynamics provides unique information about a system's Hamiltonian. A complete description of these excitations in momentum ($\vec{Q}$) and energy ($E$) space is needed to fully reconstruct the interactions that govern a material's behavior on the atomic scale. However, the technique normally requires a large volume of sample, often on the order of several cubic centimeters~\cite{Tomiyasu_2009}. This is a very significant limitation in research to develop new materials with novel functions. In neutron diffraction experiments, however, the development of time-of-flight (TOF) technique allowed the use of a white beam for increase in measurement efficiency compared with the conventional method using a monochromatic beam. Each wavelength (energy) component can be resolved by TOF, as shown in Fig.~\ref{fig:dgm_old}(a), and finally merged into a single diffraction $|\vec{Q}|$ pattern for a powder sample (time focusing) or a $\vec{Q}$ map for a single-crystal sample. Unfortunately, the same cannot be applied to inelastic scattering, because the different $E_{\rm i}$ (incident energy) and $E_{\rm f}$ (final energy) components are entangled at the same TOF for the same pixel on the detector, as shown in Fig.~\ref{fig:dgm_old}(b). Thus, it has been considered that either $E_{\rm i}$ or $E_{\rm f}$ must be monochromatized or must be analyzed, either of which incurs a large loss in neutron intensity. \begin{figure}[htbp] \begin{center} \includegraphics[width=3.2in, keepaspectratio]{fig_1} \end{center} \caption{\label{fig:dgm_old} (Color online) Plot of TOF against position without sequence chopper. Solid arrows indicate the most probable neutrons. Dotted lines sectionalize each $E_{\rm i}$ channel. (a) Diffraction. (b) Inelastic scattering. } \end{figure} Another remarkable method, called cross-correlation, was developed over four decades ago as an extension to the white-beam diffraction~\cite{Skold_1968}. The method basically involves extracting the elastic components and removing the inelastic components~\cite{Pellionisz_1971}. As shown in Fig.~\ref{fig:dgm_XC}, a special mechanical chopper modulates a white incident pulsed beam with a pseudorandom open/close sequence, and $N$-times cyclic phase shifts of the modulation generate a set of $N$ data with different $E_{\rm i}$ contrast. Then, on the basis of the contrast, the data for each $E_{\rm i}$ can be mathematically resolved. \begin{figure}[htbp] \begin{center} \includegraphics[width=3.2in, keepaspectratio]{fig_2} \end{center} \caption{\label{fig:dgm_XC} (Color online) Example of TOF--position diagrams with sequence chopper ($N=5$) at sequence phases $p=1$ (a) and $p=2$ (b). By cyclically shifting the phase, a raw data set of $I_{\rm obs}(1)$, $I_{\rm obs}(2)$, $I_{\rm obs}(3)$, $I_{\rm obs}(4)$, and $I_{\rm obs}(5)$ is obtained at a specific TOF and at a specific pixel on the detector. After the measurements, each $E_{\rm i}$ component of $I(1)$, $I(2)$, $I(3)$, $I(4)$, and $I(5)$ can be resolved mathematically. } \end{figure} The mathematical formalization is given below. Here, for convenience, parameters and functions are renamed and redefined from those in the original papers~\cite{Skold_1968,Pellionisz_1971}. The intensity detected at a specific TOF at a specific pixel of the detector, $I_{\rm obs}(p)$ $(p=1,\dots,N)$, is described by \begin{equation} \label{eq:corr_1} I_{\rm obs}(p) = \sum_{j=1}^{N}F(j+p)I(j) + B \end{equation} where $p$ is the phase shift of the sequence (phase of sequence chopper), $F(k)$ is the $k$-th element in the sequence $F$ consisting of only 0 (close) and 1 (open), $F(k+N)$ is defined to be equal to $F(k)$ for $k=1, \dots, N$, $j$ is an index for $E_{\rm i}$, $I(j)$ is the intensity coming from the $j$-th $E_{\rm i}$ component in a white incident pulsed beam for $j=1, \dots, N$, and $B$ is the background. The pseudorandom sequence $F$ is restricted by \begin{subequations} \begin{eqnarray} \label{eq:corr_2a} N = 2^{n} -1 \phantom{4} (n: \rm{integer}), \\ \label{eq:corr_2b} F^{\prime}(k) = 2F(k)-1, \\ \label{eq:corr_2c} \sum_{k=1}^{N}F^{\prime}(k) = 1, \\ \label{eq:corr_2d} \sum_{k=1}^{N}F^{\prime}(k)F^{\prime}(k+k^{\prime}) = (N+1)\delta_{0,k^{\prime}} -1. \end{eqnarray} \end{subequations} This type of sequence $F$ is currently called a maximum length sequence, which is generated by a simple recurrence formula and is widely applied in the field of digital communications~\cite{Golomb_1967}. Combining the above equations, one can resolve each $E_{\rm i}$ component, \begin{equation} I(j) = \frac{2}{N+1}\sum_{p=1}^{N}F^{\prime}(j+p)I_{\rm obs}(p) - \frac{2}{N+1}B. \label{eq:corr_3} \end{equation} It should be noted, however, that the method cannot be directly applied to inelastic scattering because the elastic components and their large statistical errors obfuscate the very weak inelastic components. This is probably why the method has not been realized thus far in an actual instrument dedicated to inelastic scattering. In fact, for the new CORELLI instrument under construction at the Spallation Neutron Source (SNS) in Oak Ridge, Tennessee, the method will be used mainly to study diffusive elastic scattering such as in frustrated systems and ionic conductors~\cite{Rosenkranz_2008}. This paper presents a modification to this method for a more practical white-beam inelastic neutron scattering setup. The proposed method has two novel aspects: the introduction of an inverse matrix representation and a proposed method for selective extraction. The former affords a different solution to Eq.~(\ref{eq:corr_1}) to resolve each $E_{\rm i}$ component in the white-beam data. The latter eliminates contamination by elastic components, which otherwise produce strong backgrounds. Finally, we present estimates of some instrumental specifications for the TOF polarized neutron spectrometer, POLANO, being constructed at J-PARC. \section{Inverse matrix representation} We formalize an alternative solution to Eq.~(\ref{eq:corr_1}) without the conditions Eqs.~(\ref{eq:corr_2a})--(\ref{eq:corr_2d}). Here, the measurement of $I_{\rm obs}(p)$ $(p=1, \dots, N)$ is the same as in the original method except for the kind of sequence. Ignoring $B$ for simplicity, Eq.~(\ref{eq:corr_1}) can be represented by \begin{equation} \vec{I}_{\rm obs} = \hat{F} \vec{I}, \label{eq:tomi_1} \end{equation} where $\vec{I}_{\rm obs}$ is the vector $\left[{I}_{\rm obs}(p)\right]$ $(p=1, \dots, N)$; $\hat{F}$ is the matrix $\left[\vec{F}_{p=1}, \vec{F}_{p=2}, \dots, \vec{F}_{p=N}\right]$; $\vec{F}_p$ is the sequence vector $\left[ F(j+p) \right]$ $(j=1, \dots, N)$; and $\vec{I}$ is the vector $\left[I(j)\right]$. Hence, one can resolve $\vec{I}$ by \begin{equation} \vec{I} = \hat{F}^{-1} \vec{I}_{\rm obs}. \label{eq:tomi_2} \end{equation} Consider, for example, the sequence (0, 1, 1, 0, 1) at $p=1$: \begin{eqnarray} \left[ \begin{array}{c} I_{\rm obs}(1) \\ I_{\rm obs}(2) \\ I_{\rm obs}(3) \\ I_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right] = \left[ \begin{array}{ccccc} 0 & 1 & 1 & 0 & 1 \\ 1 & 0 & 1 & 1 & 0 \\ 0 & 1 & 0 & 1 & 1 \\ 1 & 0 & 1 & 0 & 1 \\ 1 & 1 & 0 & 1 & 0 \\ \end{array} \right] \left[ \begin{array}{c} I(1) \\ I(2) \\ I(3) \\ I(4) \\ I(5) \\ \end{array} \right]. \label{eq:tomi_3} \end{eqnarray} Hence, one can obtain \begin{eqnarray} \left[ \begin{array}{c} I(1) \\ I(2) \\ I(3) \\ I(4) \\ I(5) \\ \end{array} \right] = \frac{1}{3} \left[ \begin{array}{ccccc} -1 & -1 & -1 & 2 & 2 \\ 2 & -1 & -1 & -1 & 2 \\ 2 & 2 & -1 & -1 & -1 \\ -1 & 2 & 2 & -1 & -1 \\ -1 & -1 & 2 & 2 & -1 \\ \end{array} \right] \left[ \begin{array}{c} I_{\rm obs}(1) \\ I_{\rm obs}(2) \\ I_{\rm obs}(3) \\ I_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right]. \label{eq:tomi_4} \end{eqnarray} Thus, almost all types of sequences can be used as long as $F^{-1}$ exists. Taking into account $B$ again, one can also identify a sequence to minimize $|F^{-1}(B,B,B,B,B)|$, for example, by trial and error with many numerical trials. It should be noted that this general matrix formalization is not considered superior to the maximum length sequence. However, the general matrix formalization does afford an advantage when the conditions of Eqs.~(\ref{eq:corr_2a})$-$(\ref{eq:corr_2d}) are not satisfied on actual instrumentation, for example, because of insufficient switching speed between 0 and 1 for high $E_{\rm i}$ range or high resolution. In this paper, we use the general matrix formalization only because selective extraction, proposed in the next section, also does not fulfill the conditions. \section{Selective extraction} We explain the proposed selective extraction method using the above example in Eq.~(\ref{eq:tomi_3}). First, one needs to prepare another chopper with the inverted sequence--from open/close to close/open, that is, from 1/0 to 0/1. The inverted chopper gives another data set, \begin{eqnarray} \left[ \begin{array}{c} J_{\rm obs}(1) \\ J_{\rm obs}(2) \\ J_{\rm obs}(3) \\ J_{\rm obs}(4) \\ J_{\rm obs}(5) \\ \end{array} \right] = \left[ \begin{array}{ccccc} 1 & 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 & 1 \\ 1 & 0 & 1 & 0 & 0 \\ 0 & 1 & 0 & 1 & 0 \\ 0 & 0 & 1 & 0 & 1 \\ \end{array} \right] \left[ \begin{array}{c} I(1) \\ I(2) \\ I(3) \\ I(4) \\ I(5) \\ \end{array} \right], \label{eq:tomi_5} \end{eqnarray} where $\left[ J_{\rm obs}(p) \right]$ is the raw data obtained for the phase $p$. Then, we consider the case where $I(3)$ is the elastic component for the targeted TOF and pixel of the detector. Our purpose is to remove $I(3)$. Thus, by selectively extracting only the arrays in which the third column is 0 (sequence chopper closed) from Eqs.~(\ref{eq:tomi_3}) and (\ref{eq:tomi_5}), one can reconstruct a good quality data set: \begin{eqnarray} \left[ \begin{array}{c} J_{\rm obs}(1) \\ J_{\rm obs}(2) \\ I_{\rm obs}(3) \\ J_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right] = \left[ \begin{array}{ccccc} 1 & 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 & 1 \\ 0 & 1 & 0 & 1 & 1 \\ 0 & 1 & 0 & 1 & 0 \\ 1 & 1 & 0 & 1 & 0 \\ \end{array} \right] \left[ \begin{array}{c} I(1) \\ I(2) \\ I(3) \\ I(4) \\ I(5) \\ \end{array} \right]. \label{eq:tomi_6} \end{eqnarray} Hence, \begin{eqnarray} \left[ \begin{array}{c} J_{\rm obs}(1) \\ J_{\rm obs}(2) \\ I_{\rm obs}(3) \\ J_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right] = \left[ \begin{array}{cccc} 1 & 0 & 1 & 0 \\ 0 & 1 & 0 & 1 \\ 0 & 1 & 1 & 1 \\ 0 & 1 & 1 & 0 \\ 1 & 1 & 1 & 0 \\ \end{array} \right] \left[ \begin{array}{c} I(1) \\ I(2) \\ I(4) \\ I(5) \\ \end{array} \right]. \label{eq:tomi_7} \end{eqnarray} This equation can be solved by dropping one array and using the inverse matrix, or by the least squares method. Further, we can reincorporate the background term $B$: \begin{eqnarray} \left[ \begin{array}{c} J_{\rm obs}(1) \\ J_{\rm obs}(2) \\ I_{\rm obs}(3) \\ J_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right] = \left[ \begin{array}{ccccc} 1 & 0 & 1 & 0 & 1 \\ 0 & 1 & 0 & 1 & 1 \\ 0 & 1 & 1 & 1 & 1 \\ 0 & 1 & 1 & 0 & 1 \\ 1 & 1 & 1 & 0 & 1 \\ \end{array} \right] \left[ \begin{array}{c} I(1) \\ I(2) \\ I(4) \\ I(5) \\ B \\ \end{array} \right]. \label{eq:tomi_8} \end{eqnarray} Hence, \begin{eqnarray} \left[ \begin{array}{c} I(1) \\ I(2) \\ I(4) \\ I(5) \\ B \\ \end{array} \right] = \left[ \begin{array}{ccccc} 0 & 0 & 0 & -1 & 1 \\ -1 & 0 & 0 & 0 & 1 \\ 0 & -1 & 1 & 0 & 0 \\ 0 & 0 & 1 & -1 & 0 \\ 1 & 1 & -1 & 1 & -1 \\ \end{array} \right] \left[ \begin{array}{c} J_{\rm obs}(1) \\ J_{\rm obs}(2) \\ I_{\rm obs}(3) \\ J_{\rm obs}(4) \\ I_{\rm obs}(5) \\ \end{array} \right]. \label{eq:tomi_9} \end{eqnarray} We emphasize that the selective extraction method is applicable not only when the third channel is elastic but also when an arbitrary channel is elastic. One can remove the elastic components at all TOFs and pixels with only the two data sets. \section{Statistical efficiency} Price and Sk$\ddot{\rm o}$ld reported that the cross-correlation method is not always better than the conventional monochromatic method in terms of statistical efficiency~\cite{Price_1970}. The reason is that white-beam data counts $\left[ I_{\rm obs}(p) \right]$ and $\left[ J_{\rm obs}(p) \right]$ are inevitably large and are accompanied by large statistical errors $\left[ \Delta I_{\rm obs}(p) \right]=\left[ \sqrt{I_{\rm obs}(p)} \right]$ and $\left[ \Delta J_{\rm obs}(p) \right]=\left[ \sqrt{J_{\rm obs}(p)} \right]$, which are propagated to $\left[ I(j) \right]$ with further enhancement by adding and subtracting $\left[ I_{\rm obs}(p) \right]$ and $\left[ J_{\rm obs}(p) \right]$. For example, $\{\Delta I(1)\}^{2} = (-1/3)^{2} \cdot I_{\rm obs}(1) + (-1/3)^{2} \cdot I_{\rm obs}(2) + (-1/3)^{2} \cdot I_{\rm obs}(3) + (2/3)^{2} \cdot I_{\rm obs}(4) + (2/3)^{2} \cdot I_{\rm obs}(5)$ in Eq.~(\ref{eq:tomi_4}), and $\{\Delta I(1)\}^{2} = (-1)^{2} \cdot J_{\rm obs}(4) + 1^{2} \cdot I_{\rm obs}(5)$ in Eq.~(\ref{eq:tomi_9}), where $\Delta I(1)$ denotes a statistical error of $I(1)$. In addition, since $N$ measurements with cyclic rotation of sequence are needed to resolve $\left[ I_{\rm obs}(p) \right]$ to $\left[ I(j) \right]$ with only one-measurement statistics, the use of white beam does not overall multiply the measurement efficiency so much. In the use of maximum length sequence, a statistical advantage can be obtained only for special $j$(s) ($E_{\rm i}$ channel(s)), in which the signal of interest is more than twice the average counts per channel~\cite{Price_1970}. Hence, the cross-correlation methods would be suitable only in the cases of phonon resonance, magnon resonance, or elastic scattering. The above situation is essentially the same as that in our modified cross-correlation method with selective extraction. Here, we roughly estimate the statistical efficiency based on the assumption that $\left[ I(j) \right]$ consists of only inelastic scattering counts with similar magnitude and without huge elastic components. First, in the cross-correlation methods without selective extraction including the original one and the aforementioned general-matrix one, the error of $I^{(\rm nsl)}(j)$ is described by \begin{equation} \begin{split} \{\Delta I^{(\rm nsl)}(j)\}^{2} = \sum_{p=1}^{N} (\hat{A}^{-1}_{jp})^{2} I_{\rm obs}^{(\rm nsl)}(p) \\ \sim a_{j}^{2} \langle I_{\rm obs}^{(\rm nsl)} \rangle \sim N_{\rm open} a_{j}^{2} \langle I^{(\rm nsl)} \rangle, \end{split} \label{eq:err_1a} \end{equation} where the superscript (\rm nsl) denotes the non-use of selective extraction, $\hat{A}^{-1}$ corresponds to the inverse matrix, $a_{j}^{2} = \sum_{p=1}^{N}(\hat{A}^{-1}_{jp})^{2}$, $\langle I_{\rm obs}^{(\rm nsl)} \rangle = \sum_{p=1}^{N}I_{\rm obs}^{(\rm nsl)}(p)/N$, $N_{\rm open}$ is the number of opening channels ($\sim N/2$), and $\langle I^{(\rm nsl)} \rangle = \sum_{j=1}^{N}I^{(\rm nsl)}(j)/N$. Hence, the statistical efficiency $\eta_{j}^{(\rm nsl)}$ is estimated by \begin{equation} \begin{split} \eta_{j}^{(\rm nsl)} \equiv \frac{I^{(\rm nsl)}(j)}{\Delta I^{(\rm nsl)}(j)} \\ \sim \left( \frac{c_{j}}{\sqrt{N_{\rm open}}a_{j}} \right) \cdot \sqrt{\langle I^{(\rm nsl)} \rangle} \\ \sim \left( \frac{\sqrt{2}c_{j}}{\sqrt{N}a_{j}} \right) \cdot \sqrt{\langle I^{(\rm nsl)} \rangle}, \end{split} \label{eq:err_2a} \end{equation} where $I^{(\rm nsl)}(j) \equiv c_{j}\langle I^{(\rm nsl)} \rangle$. Next, in the modified cross-correlation method, the error of $I^{(\rm sl)}(j)$ is described by \begin{equation} \begin{split} \{\Delta I^{(\rm sl)}(j)\}^{2} = \sum_{p=1}^{N} (\hat{B}^{-1}_{j^{\prime}p})^{2} \{K_{\rm obs}^{(\rm sl)}(p)\} \\ \sim \sum_{p=1}^{N} (\hat{B}^{-1}_{j^{\prime}p})^{2} \{I_{\rm obs}^{(\rm nsl)}(p)/2\} \\ \sim \frac{1}{2} b_{j^{\prime}}^{2} \langle I_{\rm obs}^{(\rm nsl)} \rangle \sim \frac{1}{2} N_{\rm open} b_{j^{\prime}}^{2} \langle I^{(\rm nsl)} \rangle, \end{split} \label{eq:err_1b} \end{equation} where the superscript (sl) denotes the use of selective extraction, $K_{\rm obs}^{(\rm sl)}(p)=I_{\rm obs}^{(\rm sl)}(p)$ or $J_{\rm obs}^{(\rm sl)}(p)$ like in Eq.~(\ref{eq:tomi_9}), $I_{\rm obs}^{(\rm sl)}(p) \sim J_{\rm obs}^{(\rm sl)}(p) \sim I_{\rm obs}^{(\rm nsl)}(p)/2$ since selective extraction involves the use of two data sets: one from the original chopper and the other from the inverted chopper, $\hat{B}^{-1}$ corresponds to the final inverse matrix like in Eq.~(\ref{eq:tomi_9}), $b_{j^{\prime}}^{2} = \sum_{p=1}^{N}(\hat{B}^{-1}_{j^{\prime}p})^{2}$, and $j^{\prime}$ denotes the row number in $\hat{B}^{-1}$ with which $I^{(\rm sl)}(j)$ can be obtained (e.g., $j^{\prime}=1$ for $j=1$ and $j^{\prime}=3$ for $j=4$ in Eq.~(\ref{eq:tomi_9})). Hence, the statistical efficiency $\eta_{j}^{(\rm sl)}$ is estimated by \begin{equation} \begin{split} \eta_{j}^{(\rm sl)} \equiv \frac{I^{(\rm sl)}(j)}{\Delta I^{(\rm sl)}(j)} = \frac{I^{(\rm nsl)}(j)/2}{\Delta I^{(\rm sl)}(j)} \\ \sim \left( \frac{c_{j}}{\sqrt{2}\sqrt{N_{\rm open}}b_{j^{\prime}}} \right) \cdot \sqrt{\langle I^{(\rm nsl)} \rangle} \\ \sim \left( \frac{c_{j}}{\sqrt{N}b_{j^{\prime}}} \right) \cdot \sqrt{\langle I^{(\rm nsl)} \rangle}, \end{split} \label{eq:err_2b} \end{equation} where $I^{(\rm sl)}(j)=I^{(\rm nsl)}(j)/2$ since selective extraction involves the use of two data sets again. Then, in the conventional monochromatic experiments, the efficiency $\eta_{j}^{\rm (mono)}$ required to obtain the same data set of all $E_{\rm i}$ ($j=1, \dots, N$) over the same total measurement time using monochromatic beam is estimated by \begin{equation} \eta_{j}^{\rm (mono)} = \frac{c_{j}\langle I^{(\rm nsl)} \rangle}{\sqrt{c_{j}\langle I^{(\rm nsl)} \rangle}} = \sqrt{c_{j}} \cdot \sqrt{\langle I^{(\rm nsl)} \rangle}, \label{eq:err_3} \end{equation} which corresponds to the measurements described by $\hat{A}=\hat{E}$ (identity matrix) in the general matrix formalization. Thus, the ratios of the statistical efficiencies are estimated by \begin{subequations} \begin{eqnarray} \frac{ \eta_{j}^{(\rm nsl)} }{ \eta_{j}^{\rm (mono)} } \sim \sqrt{ \frac { 2c_{j} } { Na_{j}^2 } }, \label{eq:err_4a} \\ \frac{ \eta_{j}^{(\rm sl)} }{ \eta_{j}^{\rm (mono)} } \sim \sqrt{ \frac { c_{j} } { Nb_{j^{\prime}}^2 } }. \label{eq:err_4b} \end{eqnarray} \end{subequations} Using example sequences summarized in the Appendix, we numerically calculated the values of $a_{j}$ and $b_{j^{\prime}}$ and the criteria of $c_{j}$ to retrieve the $I(j)$ of interest more efficiently than the use of conventional monochromatic beam. For the original cross-correlation method with maximum length sequence without selective extraction, the results are $a_{j}^{2}=0.44$, 0.23, 0.12, 0.062, 0.031, and 0.016 for all $j$ at $N=7$, 15, 31, 63, 127, and 255, respectively. Hence, $(\eta_{j}^{(\rm nsl)} / \eta_{j}^{\rm (mono)})>1$ gives the criteria of $c_{j}>1.5$, 1.8, 1.9, 2.0, 2.0, and 2.0, which are consistent with Price's criterion of $c_{j}>2$ for all the $N$; the $I(j)$ of interest should be more than twice the average counts per channel~\cite{Price_1970}. For the modified cross-correlation method, $b_{j^{\prime}}^{2}=0.88$, 0.39, 0.22, 0.15, 0.096, and 0.059 ($j^{\prime} \neq N$: $I(j)$) and $b_{N}^{2}=0.28$, 0.31, 0.63, 0.78, 0.20, and 0.30 ($j^{\prime} = N$: constant background). Hence, $(\eta_{j}^{(\rm sl)} / \eta_{j}^{\rm (mono)})>1$ ($j \neq N$: $I(j)$) gives $c_{j}>6.2$, 5.8, 6.7, 9.4, 12, and 15. The latter criteria are harder than the former ones in the absence of huge elastic components, as expected. In addition, since $c_{j}$ increases as $N$ increases in the latter, it would be better to set $N$ less than about 60. In this way, both the original and modified cross-correlation methods can efficiently give only the $I(j)$ components with relatively large $c_{j}$ among $\left[ I(j) \right]$. Therefore, it is important to tune the experimental conditions, such as the ranges of $E_{\rm i}$, TOF, and pixel used, such that the components of interest become the strongest in intensity among $\left[ I(j) \right]$. In this sense, we would like to emphasize that the modified method with selective extraction can remove not only an elastic channel but also an arbitrary one channel of no interest with relatively strong intensity among $\left[ I(j) \right]$; for example, spurious neutrons scattered on unexpected paths and spurious neutrons coming from the previous flame. In practice, (1) for elastic scattering, the original method without selective extraction could be safely applied, as has been reported in the past~\cite{Skold_1968, Pellionisz_1971, Rosenkranz_2008, Price_1970}, since elastic scattering is normally the strongest among $\left[ I(j) \right]$. However, the modified method might improve the data as an insurance to remove a spurious channel. (2) For quasielastic scattering and low-energy inelastic scattering, which will have the next strongest intensity and will necessarily overlap with the strongest elastic scattering in TOF, the modified method will be effective to remove the elastic scattering. (3) For higher energy modes with relatively weak intensity, one must remove not only an elastic channel but also quasielastic and low-energy inelastic channels, which contaminate the higher energy data as a source of statistical errors. Therefore, it would be better to narrow the $E_{i}$ range from white to quasi-monochromatic, which will allow us to avoid all the elastic, quasielastic, and low energy inelastic scatterings by TOF, in conjunction with the modified method to remove an arbitrary channel of no interest again. In any case, to efficiently utilize the cross-correlation methods, users should recognize that the optimal experimental condition will highly depend on the specific aim of the individual experiments. \section{Specifications for practical implementation} We are constructing a TOF polarized neutron spectrometer called POLANO at a decoupled moderator at J-PARC. Because polarization devices such as Heusler crystals and spin filters lose a large proportion of neutrons, we need a method to gain an increase of over a factor of 10 in measurement efficiency. Our method of white-beam inelastic neutron scattering is one candidate, though it is applicable to both unpolarized and polarized inelastic neutron scattering. In this section, taking this spectrometer as an example, we present some specifications for the sequence chopper. For this system, we set the distance between moderator and sample $L_{1}$ as 17.0 m, the distance between sample and detector $L_{2}$ as 2.0 m, and the distance between sequence chopper and sample $L_{3}$ as 2.0 m. The time width values at the decoupled moderator at J-PARC, $\Delta t_{\rm m}$, were used for each $E_{\rm i}$ (Table~\ref{tab:dt_sc}). For each $E_{\rm i}$, the energy resolution $\Delta E/E_{i}$ was kept $\leq0.04$. Then, using the analytical formula for energy resolution~\cite{Tomiyasu_2006}, we evaluated the opening time per $E_{\rm i}$ channel required at the sequence chopper, $\Delta t_{\rm sc1}$. Also, on the assumption that a sequence chopper is alternately closed without generating a sequence as an example, as shown in Fig.~\ref{fig:dgm_new}, we estimated the condition of $\Delta t_{\rm sc2}$ necessary to avoid contamination by the elastic tails from neighboring $E_{\rm i}$ channels. The condition can be estimated by $\Delta t_{\rm d} \leq \Delta t_{\rm ch}$, where $\Delta t_{\rm d}$ is the time width of the elastic component at the pixel on the detector with $E_{\rm i}$, and $\Delta t_{\rm ch}$ is the time difference between two TOFs of neighboring $E_{\rm i}$ channels at the pixel, as defined in Fig.~\ref{fig:dgm_new}. The two parameters are described by other parameters, \begin{subequations} \begin{eqnarray} \Delta t_{\rm d} &=& \frac{ L_{1}+L_{2} - \frac{\Delta t_{\rm m}}{\Delta t_{\rm m}+\Delta t_{\rm sc2}}(L_{1}-L_{3}) }{\frac{\Delta t_{\rm m}}{\Delta t_{\rm m}+\Delta t_{\rm sc2}}(L_{1}-L_{3})} \cdot \Delta t_{\rm m}, \\ \Delta t_{\rm ch} &=& \frac{L_{1}+L_{2}}{L_{1}-L_{3}} \cdot (2\Delta t_{\rm sc2}) \phantom{2}. \end{eqnarray} \end{subequations} \begin{figure}[htbp] \begin{center} \includegraphics[width=2.1in, keepaspectratio]{fig_3} \end{center} \caption{\label{fig:dgm_new} (Color online) Diagram showing TOF and position with sequence chopper in the estimated specifications for POLANO. Solid arrows indicate the most probable neutrons, and dotted arrows indicate most inaccurate neutrons. The sequence chopper is alternately closed using a set of counter-rotating disk choppers. The alternating closing parts do not generate the sequence. } \end{figure} As summarized in Table~\ref{tab:dt_sc}, there exists a solution for $\Delta t_{\rm sc}$. For example, over a wide $E_{\rm i}$ range of 10--80 meV, $\Delta t_{\rm sc} = 9$ $\mu$sec simultaneously satisfies the constraint of $\Delta E/E_{i} \leq 0.04$ and the avoidance of the tails from neighboring $E_{\rm i}$ channels. This opening time can be realized by a set of counter-rotating disk choppers with the following parameters: 700 mm-$\phi$, 20 mm/channel, and 350 Hz~\cite{comm_1}. In addition, for the sequence chopper, note that an arbitrary sequence can be generated by printing a sequence clockwise and then counterclockwise on counter-rotating disk choppers. The sequence chopper generates $0.5\cdot(\pi\cdot700 \phantom{1} \rm{(mm)})/20 \phantom{1} \rm{(mm)}=55$ $E_{\rm i}$ channels. \begin{table}[htbp] \begin{center} \caption{\label{tab:dt_sc} Results of numerical estimation of system specifications for POLANO. The values $\Delta t_{\rm sc1}$ and $\Delta t_{\rm sc2}$ are obtained so as to satisfy $\Delta E/E_{i} \leq 0.04$ and avoid contamination by the elastic tails from neighboring $E_{\rm i}$ channels, respectively. All the time widths are defined as full widths at half maximum. } \bigskip \begin{tabular}{cccc} \hline $E_{\rm i}$ (meV) & $\Delta t_{\rm m}$ ($\mu$sec) & $\Delta t_{\rm sc1}$ ($\mu$sec)& $\Delta t_{\rm sc2}$ ($\mu$sec)\\ \hline 5.0 & 50 & $\leq31$ & $\geq11$ \\ 10 & 32 & $\leq22$ & $\geq7$ \\ 20 & 20 & $\leq16$ & $\geq4$ \\ 40 & 13 & $\leq11$ & $\geq3$ \\ 80 & 9 & $\leq9$ & $\geq2$ \\ \hline \end{tabular} \end{center} \end{table} \section{Summary} We developed a modified cross-correlation method for an increase in measurement efficiency of inelastic neutron scattering. First, a different solution using an inverse matrix representation was formalized to resolve each $E_{\rm i}$ component in the white-beam data. Second, a method of selective extraction was proposed to avoid contamination by elastic components. Third, taking spectrometer POLANO at J-PARC as an example, practical specifications for the sequence chopper were estimated. Experimental situations giving high efficiency, however, would be quite limited and complex.
\section{Introduction} It is very likely that our Sun, like most stars, was born as part of a star cluster. The birth cluster environment is a self-gravitating system of gas and dust that collapses at a multitude of points to create its membership of stars. This process is inefficient, converting only 10\% to 30\% of the mass to stars while leaving 70\% to 90\% as gas \citep{LadaLada-2003}. As long as the gas remains, the stars are gravitationally bound and orbit chaotically throughout the cluster. This motion continues until the winds and ionizing radiation of the newborn stars blow away the remaining gas. At that point the cluster is no longer self-gravitating, it disperses, and the stars are free to move about the Galaxy. \par By examining the state of our Solar System as it exists today, we can place limits on the birth cluster's properties (i.e. mass, surface density, radius, etc.) and thus have a better understanding of the environment in which our Solar System was born. \par Meteorites provide one line of evidence. Samples taken from meteorites that solidified not long after the Solar System formed show daughter products of multiple radioactive isotopes. These radioactive isotopes most likely originated from stellar nucleosynthesis, and their presence suggests that there was a nearby supernova that enriched the Solar Nebula \citep{ThraneEtAl-2006, Wadhwa-2007, AllenEtAl-2007, ConnellyEtAl-2008, DupratTatischeff-2008, Adams-2010}. Such early exposure is likely only if the birth cluster includes a star $\geq 25 M_{\odot}$ in mass. This consideration suggests a birth cluster with at least 1000 members \citep{Adams-2010}. A priori such a large cluster is not unlikely. The young cluster mass function is $\frac{dN}{dM}\propto M^{-2}$ over the range from $\sim 10^{1}-10^{6}M_{\odot}$ \citep{LadaLada-2003,Chandar-2009,FallEtAl-2010}, which implies that $\sim\frac{1}{2}-\frac{2}{3}$ of stars are born in clusters larger than this.\footnote{We do caution here that the observations do not completely establish the existence of a mass function $dN/dM\propto M^{-2}$ over this full mass range. This powerlaw is seen in both the Galactic sample of \citet{LadaLada-2003} and the extragalactic samples of \citet{Chandar-2009} and \citet{FallEtAl-2010}. However, the \citet{LadaLada-2003} sample only includes clusters within 2 kpc of the Sun, the most massive of which is $\sim 10^3$ $M_\odot$. The \citet{Chandar-2009} and \citet{FallEtAl-2010} samples only go down to a mass of $\sim 10^4$ $M_\odot$. Thus the data do not rule out the possibility that there is a break in the mass function in the range $10^3 - 10^4$ $M_\odot$. However, it would be quite a coincidence for this discontinuity to coincide exactly with the mass range where the observations are incomplete. A simpler hypothesis is that $dN/dM \propto M^{-2}$ over the full mass range.} \par On the other hand, several authors have attempted to obtain upper limits on the birth cluster size by considering the effects of close encounters with other stars on the young Solar System. Close encounters would perturb the Solar Nebula during its formation. Any such perturbation must still permit the formation of a Solar System like the one we see today, with eight planets that all lie close to the same plane (inclination angles $\leq$ 3.5$^{\circ}$) and have small eccentricities (less than 0.2, or less than 0.09 is we exclude Mercury, whose eccentricity is pumped by perturbations from other planets). \par \citet{AdamsLaughlin-2001} determined an upper bound on the birth cluster size by performing Monte Carlo scattering simulations of close encounters between the Solar System and passing stars. They found that the maximum number of stars the birth cluster could have had without causing a Jovian planet to be excited to an eccentricity greater than 0.1 is $2200\pm1100$. However, observations of the properties of young, embedded star clusters were at the time quite limited, and they were forced to make a variety of assumptions about cluster properties. First, the relative velocities of the stars in their simulations were chosen randomly from a Maxwellian distribution with a standard deviation of 1 km/s; thus they had very few simulations done with higher velocities, and the velocity distribution was implicitly assumed to be independent of the cluster mass. Second, they assumed that the cluster's lifetime, and thus the time the Sun was exposed to close encounters, was on the order of 100 cluster crossing times. Lastly, they assumed that the cluster's radius was fixed and independent of its mass. This last assumption would lead a high mass cluster to have very high surface densities. \par More recent observations have shown that most of these assumptions are not satisfied in typical embedded clusters. Compiling observations of embedded clusters from \citet{ShirleyEtAl-2003}, \citet{FaundezEtAl-2003}, and \cite{FontaniEtAl-2005}, \citet{FallEtAl-2010} find that embedded clusters form a sequence of roughly constant surface density, not constant radius. Similarly, current observations favor cluster lifetimes closer to one to four crossing times \citep{Elmegreen-2000,HartmannEtAl-2001,LadaLada-2003,TanEtAl-2006, JeffriesEtAl-2011,ReggianiEtAl-2011}, not 100. After $\sim 10$ Myr, 90\% of the stars born in clusters have dispersed into the field, and by $\sim 100$ Myr roughly 99\% have done so \citep{AllenEtAl-2007,FallEtAl-2009, ChandarEtAl-2010}. We do not know the lifetime of the Sun's birth cluster, but in obtaining limits on its size it is important to consider the possibility both that the lifetime was very short, and that it was tens of Myr, rather than assuming the latter as was done in most previous work. \par In this paper we show that retaining the same theoretical framework as \citet{AdamsLaughlin-2001}, but using these more recent observations to set the properties of the embedded clusters leads to a dramatic relaxation of the \citet{AdamsLaughlin-2001} limit on the size of the Sun's birth cluster. In turn, this greatly eases the tension between the meteoritic and dynamical constraints. In particular, the lifetime of a $25 \ M_{\odot}$ star before it goes supernova is $\sim7.54$ Myr, longer than most clusters survive. A 75 $M_{\odot}$ star, however, has a lifetime of $\sim3.64$ Myr. \citet{Adams-2010} finds that to have a reasonable probability of producing such a star, a birth cluster of membership size $>10^{4}$ is needed. Again, such a large cluster is not a priori unlikely, given the observed young cluster mass function. \par Close encounters within the Sun's birth cluster have also been theorized to be responsible for other aspects of the Solar System's present-day architecture, particularly the drastic drop in object density in the Kuiper Belt at $\sim$45AU, and the Kuiper Belt object Sedna. Both of these phenomena are difficult to understand in our current picture of Solar System formation. The existence of Neptune, with its size and orbit, suggests that the surface density of the Solar Nebula should have been high enough to produce Kuiper Belt Objects (KBOs) out to distances greater than 50AU with a gradually declining object density. An unperturbed Solar Nebula, which tends to produce objects with low eccentricities, is hard to reconcile with Sedna's eccentricity of 0.84 and semi-major axis of 542 AU. A close encounter with another star could produce a disk truncation \citep{IdaEtAl-2000,KobayashiIda-2001} and scatter objects into high eccentricity orbits to produce Sedna \citep{BrasserEtAl-2006}. \par Close encounters could also affect other putative objects orbiting the Sun: the proposed companion brown dwarf Nemesis \citep{Raup-1984, Whitmire-1984} or the more recently proposed Tyche \citep{Matese-1999,MateseEtAl-2005,MateseWhitmire-2011}. Tyche has been proposed to have an orbit just inside the Inner Oort Clound (a semi-major axis of $\sim$5,000 AU) with a mass of 1-4 Juipter Masses. Given the low probability of capturing such a companion \citep{Tohline-2002}, if either Nemesis or Tyche exists, they must have formed with the Sun. To date, there has been no research on whether Nemesis nor Tyche could survive the dynamics of the birth cluster. \par In this paper we address all of these issues. We have constructed simulations similar those of \citet{AdamsLaughlin-2001}. We use these together with updated observational data on the properties of embedded clusters to determine an upper limit of our birth cluster's membership size and the effects of a close encounter on both the truncation on the Kuiper Belt and production of a Sedna-like object. In addition, we will use this data to determine the likelihood of Tyche staying bound to the Sun during its time within the birth cluster. \citet{AdamsEtAl-2006} have undertaken similar calculations, but survey a much more limited part of parameter space. In particular, they do not consider clusters of more than 10$^{3}$ stars. In Section 2 we will look at the simulation and method used in our experiment. In Section 3 we will then analyze the results of the simulation and finally in Section 4 we will make our conclusions. \section{Methods} To determine an upper limit on the birth cluster membership size, we want to determine the probability of a disruptive encounter as a function of cluster surface density and mass. By disruptive event, we mean a close encounter in which any of the Jovian planets are excited to an eccentricity greater than 0.1. In the process, we will also determine whether encounters with passing stars could be responsible for truncating the Kuiper Belt at $\sim$50AU, for producing a Sedna-like object, or for stripping any distant brown dwarf companion. \subsection{Determining Probability} To obtain the probability of an event occurring, we first perform numerical scattering simulations to determine the velocity-dependent cross-section for that event, $\langle\sigma\rangle$. We then calculate the average cross section times velocity, $\langle\sigma v\rangle$. We defer a description of our simulations to section 2.2. Simulations of star clusters suggest that the distribution of encounter velocities within a embedded stellar cluster is Maxwellian \citep{ProszkowAdams-2009}, so \begin{equation} \label{eq:sigma} \langle\sigma v\rangle=\frac{1}{\sigma^{3}_{v}}\sqrt{\frac{2}{\pi}}\int_{-\infty}^{\infty}\langle\sigma\rangle v^{3}{\it e}^\frac{-v^{2}}{2\sigma^{2}_{v}}dv \end{equation} where $\sigma_{v}$ is the standard deviation of the velocity distribution. If we then multiply this by the stellar number density of the cluster, $n$, we obtain the encounter rate, \begin{equation} \label{eq:Gamma} \Gamma=n\langle\sigma v\rangle. \end{equation} If we multiply $\Gamma$ by the total time the Sun was exposed in the cluster, we will have the expected number of events, $\Lambda$. This time will be roughly the time for which the cluster survives before mass evaporation leads it to disperse, $t_{life}$. Thus, we have \begin{equation} \label{eq:Lambda} \Lambda=\Gamma t_{life}. \end{equation} To compute the overall probability that an event will occur, we will assume that events follow a Poisson distribution, in which case the probability that an event will occur at least once is \begin{equation} \label{eq:PoissonOne} P_{event}=1-{\it e}^{-\Lambda}. \end{equation} The calculation is slightly more complicated for the Kuiper Belt, where there are many KBOs, and varying numbers can be affected by a single encounter. In our simulations we represent particular radial bins of the Kuiper Belt with a number $N_t$ of test particles. We use our simulation to determine $\langle\sigma\rangle_{k}$, the velocity-dependent cross section for an event (e.g.\ excitation or ejection from the Solar System) to affect $k$ of these $N_t$ test KBOs. Using equation \ref{eq:sigma}, we can then calculate $\langle\sigma v\rangle_{k}$ for each value of $k$. The rate at which the KBO population undergoes events is then given by \begin{equation} \label{eq:kGamma} \Gamma=n\sum_{k=1}^{N_t}\frac{k}{N_t}\langle\sigma v\rangle_{k}, \end{equation} where $n$ is the number density determined in equation \ref{eq:NumberDensity}. Once $\Gamma$ has been determined, we use equation (\ref{eq:Lambda}) to compute $\Lambda$. The expected fraction of KBOs that undergo a certain even is then \begin{equation} \label{eq:expectedfraction} \langle f\rangle=1-e^{-\Lambda}. \end{equation} The initial conditions of the birth cluster are unknown to us, so we cannot directly compute all of these quantities. However, we can estimate them in terms of the cluster surface density, $\Sigma_{c}$, and cluster mass, $M_{c}$. For a spherical cluster, the number density, velocity dispersion, and crossing time are \begin{equation} \label{eq:NumberDensity} n=\frac{3\pi^{1/2}\Sigma^{3/2}_{c}}{4\overline{m}M_{c}^{1/2}} = 700 M_4^{-1/2}\Sigma_{-1}^{3/2}\mbox{ pc}^{-3} \end{equation} \begin{equation} \begin{split} \label{eq:VelocityDispersion} \sigma_{v}=\left(\frac{3}{5}\alpha_{vir}\right)^{1/2}G^{1/2}(\pi \Sigma_{c}M_{c})^{1/4} \\ = 3.2 \alpha_{\rm vir}^{1/2} M_4^{1/4} \Sigma_{-1}^{1/4}\mbox{ km s}^{-1} \end{split} \end{equation} \begin{equation} \label{eq:CrossingTime} t_{cross} = \frac{M_{c}^{1/4}}{G^{1/2}(\pi\Sigma_{c})^{3/4}} = 0.62 M_{4}^{1/4} \Sigma_{-1}^{-3/4}\mbox{ Myr}, \end{equation} \begin{figure*} \plotone{Jovian01CT} \caption{The log of the probability of a close encounter exciting each Jovian planet's eccentricity to greater than 0.1, in one crossing time, as a function of cluster mass M$_{c}$ and surface density $\Sigma_{c}$. \label{fig:Jovian}} \end{figure*} \begin{figure*} \plotone{JovImpact} \caption{The effective impact parameter, $b=\sqrt{\langle\sigma \rangle_{v} / \pi}$ for exciting a Jovian to eccentricity greater than 0.1 vs.\ velocity of incoming star. They gray band indicates, given then number of trials at a given velocity, the smallest value of $b$ we can measure with 99\% confidence. The increase in the minimum $b$ with velocity is a result of ours having somewhat more simulations at low than at high velocity. The right panel shows a zoomed in portion of the full graph shown on the left. \label{fig:JovImpact}} \end{figure*} \begin{figure*} \plotone{JovianVir01CT} \caption{The log of the probability of a close encounter exciting each Jovian planet's eccentricity to greater than 0.1, in one crossing time, as a function of cluster mass M$_{c}$ and surface density $\Sigma_{c}$ in a sub-viral cluster ($\alpha_{vir}=$1/3). \label{fig:JovianVir}} \end{figure*} where $\alpha_{vir}$ is the virial parameter, $\overline{m}$ is the mean stellar mass, which we take to be 0.2$M_{\odot}$ \citep{Chabrier-2005} for our calculations, $M_{4}=M_{c}/10^{4}$ $M_{\odot}$, and $\Sigma_{-1}=\Sigma_{c}/0.1$ g cm$^{-2}$. Equations (\ref{eq:NumberDensity}) and (\ref{eq:VelocityDispersion}) provide estimates of the number density and velocity dispersion that we can use in our estimates. The relationship between the cluster lifetime $t_{\rm life}$ and the crossing time, however, is somewhat uncertain. As discussed in the Introduction, most clusters disperse rapidly, although exactly how rapidly is debated. For example, the best studied cluster containing O stars is the Orion Nebula Cluster ($M_4 = 0.46$, $\Sigma_{-1} = 1.2$ -- \citealt{Hillenbrand98a}), for which $t_{\rm cross} = 0.44$ Myr. Estimates of its age spread range from $\sim 1-3$ Myr \citep[e.g.][]{ReggianiEtAl-2011, JeffriesEtAl-2011}, corresponding to $\sim 1 - 6 t_{\rm cross}$, depending on the technique used to estimate protostellar ages and similar systematic uncertainties. On the other hand, $\sim 10\%$ of clusters survive for more than 10 Myr. We therefore consider two possible scenarios for the lifetime. The more likely one is that the Sun was born in one of the rapidly-dispersing clusters, and in this case we take $t_{\rm life} = t_{\rm cross}$; the true lifetime may be a factor of a few larger, but, as we will see below, this factor of a few is relatively insignificant. The other possibility is that the Sun was born in the minority of clusters with a longer lifetime. To represent this case, we take $t_{\rm life} = 30$ Myr, typical of this class. Finally, note that, in the rapidly-dispersing cluster scenario, for massive and high surface density clusters the cluster lifetime may be significantly less that the main sequence lifetime even of a very massive star. This is certainly the case in the ONC, for example, where star formation is complete and the cluster is likely in the process of dispersing, but $\Theta^1$ Ori C is still alive. This is not necessarily a problem for enrichment, however. Even for a dispersing cluster, the relatively low velocity dispersion of the stars implies that stars will not spread that far before the most massive stars go supernova. \subsection{Simulations} The simulations we use to measure $\langle\sigma\rangle$ consist of the Solar System and the scattering star system. The Solar System includes the Sun, the four Jovian planets, Tyche and 320 KBOs all within the same plane. We only place Tyche in the simulation because the semi-major axis of Nemesis is so large that we cannot reasonably evaluate it behavior with scattering simulations. Instead, a full (and much more expensive) N-body simulation of the entire birth cluster would be required. We treat the Jovian planets and KBOs as massless to conserve computing power. This means that we are unable to model secular processes produced by planet-planet interactions. However, since such processes generally take much longer than that $\sim 0.01$ Myr covered by our simulations, we could not model these in any event. We consider Tyche masses of zero, 0.02 $M_{\odot}$ and 0.08 $M_{\odot}$. We set the orbital properties of the Jovian planets to their present-day values and we randomize the initial orbital phases. We give Tyche a semi-major axis of 1,000 AU\footnote{While this is closer than the 5000 AU proposed by \citet{MateseEtAl-2005} and \citet{MateseWhitmire-2011}, the use of a similar but smaller distance greatly reduces the computational cost by allowing us to consider a smaller range of incoming star impact parameters.}, an eccentricity randomly between 0.0 and 0.4, and a random argument of periastron. We set up the Kuiper Belt up such that all objects are placed in 32 concentric circles ranging from 35 AU to 500 AU, in 15 AU intervals. Each circle has ten objects equally spaced in angle, and all objects have a zero eccentricity. For our statistical analysis, we then divide the KBOs into eight distance bins (35-80AU, 95-140AU, 155-200AU, 215-260AU, 275-320AU, 335-380AU, 395-440AU and 455-500AU), each bin containing four concentric circles. \par The incoming star systems consist of either a solo star or a binary star system. We first select the mass of the primary incoming star by drawing from a \citet{Chabrier-2005} IMF. We then determine if the primary has a companion. We set the probability of there being a companion to \begin{equation} \label{eq:binary} f_{binary} = \left\{ \begin{array}{lr} 0.2 & : m\;<\;0.5\;M_{\odot} \\ 0.2+0.8\left( \frac{m-0.5}{1.5} \right) & : 0.5\;M_{\odot}\;\leq\;m\;\leq 2.0\;M_{\odot} \\ 1.0 & : m\;>\;2.0\;M_{\odot} \end{array} \right. \end{equation} based on a rough fit to the mass-dependent companion fraction reported by \citet{Lada-2006}. If a companion exists, we also select its mass from the IMF, rejecting and redrawing if the mass exceeds the primary's mass. Finally, we select the orbital period of the binary by drawing from the distribution observed for field G star binaries \citep{DuquennoyMayor-1991} \begin{equation} \label{eq:tau} p(\log\tau )=\frac{1}{\sqrt{2\pi \sigma_{\log\tau}^{2}}}e^{\frac{-(\log\tau -\overline{\log\tau})^{2}}{2\sigma_{\log\tau}^{2}}} \end{equation} where $\tau$ is the period in days, $\sigma_{\log\tau}$=2.3, and $\overline{\log\tau}=4.8$. \par We determine the impact parameter of the encounter, $b$, randomly with probability proportional to $b$, with a maximum, $b_{max}$, of 2000 AU. We randomize the relative orientations of the incoming star, the Solar ecliptic, and the binary orbital plane. Finally, for each incoming system, we run the simulation for a series of relative velocities between the system and the Sun. We use velocities of $v$=0.1-2.5 km/s at intervals of 0.1 km/s and $v$=3.0-20.0 km/s at intervals of 0.5km/s. Over 2.1 million runs were preformed in total. \\ \\ \section{Results} \begin{figure} \plotone{KBOLost01CT} \caption{The log of the expected fraction of KBOs that will become unbound in one cluster crossing time in a given distance bin. In all cases, the inner bin and the outer bin represent the extreme values.} \label{fig:KBOLost01CT} \end{figure} \begin{figure} \plotone{KBOLostVir01CT} \caption{The log of the expected fraction of KBOs that will become unbound in a sub-viral cluster ($\alpha_{vir}=$1/3) in one cluster crossing time in a given distance bin. In all cases, the inner bin and the outer bin where the extreme values.} \label{fig:KBOLostVir01CT} \end{figure} \begin{figure} \plotone{KBOExcited01CT} \caption{The log of the expected fraction of KBOs that will be excited to an eccentricity greater than 0.5 but less that 1.0 in one cluster crossing time in a given distance bin. In all cases, the inner bin and the outer bin represent the extreme values.} \label{fig:KBOExcited01CT} \end{figure} \begin{figure} \plotone{KBOExcitedVir01CT} \caption{The log of the expected fraction of KBOs that will be excited to an eccentricity greater than 0.5 but less that 1.0 in a sub-viral cluster ($\alpha_{vir}=$1/3) in one cluster crossing time in a given distance bin. In all cases, the inner bin and the outer bin where the extreme values.} \label{fig:KBOExcitedVir01CT} \end{figure} \begin{figure} \epsscale{1.0} \plotone{Nemesis01CT} \caption{The log of the probability that, in one crossing time, Tyche becomes unbound.} \label{fig:Tyche} \end{figure} \begin{figure} \plotone{NemesisVir01CT} \caption{The log of the probability that, in one crossing time, Tyche becomes unbound in a sub-viral cluster ($\alpha_{vir}=$1/3). \label{fig:TycheVir} } \end{figure} In Sections \ref{sec:jovian} -- \ref{sec:tyche}, we first examine the effects of encounters on the Jovian planets, on the Kuiper Belt, and on Tyche in the more likely scenario where the Sun was born in a cluster with a lifetime $t_{\rm life} = t_{\rm cross}$. In Section \ref{sec:longlife} we examine how these results change in the scenario where the Sun's parent cluster was one of the $\sim 10\%$ that reach ages of $t_{\rm life} = 30$ Myr. \subsection{Effects of Close Encounters on the Jovian Planets} \label{sec:jovian} \begin{figure*} \epsscale{0.8} \plotone{JovianMyr30} \caption{The log of the probability of a close encounter exciting each Jovian planet's eccentricity to greater than 0.1 as a function of cluster mass M$_{c}$ and surface density $\Sigma_{c}$ with cluster lifetime of 30 Myr.} \label{fig:JovianMyr30} \end{figure*} \begin{figure} \plotone{KBOLostMyr30} \caption{The log of the expected fraction of KBOs that will become unbound with cluster lifetime of 30 Myr in a given distance bin. In all cases, the inner bin and the outer bin represent the extreme values.} \label{fig:KBOLostMyr30} \end{figure} \begin{figure} \plotone{KBOExcitedMyr30} \caption{The log of the expected fraction of KBOs that will be excited to an eccentricity greater than 0.5 but less that 1.0 with cluster lifetime of 30 Myr in a given distance bin. In all cases, the inner bin and the outer bin represent the extreme values.} \label{fig:KBOExcitedMyr30} \end{figure} Figure \ref{fig:Jovian} show the probabilities of each Jovian planet being excited to an eccentricity greater than 0.1 as a function of $M_{c}$ and $\Sigma_{c}$ in a fully relaxed embedded cluster (i.e. $\alpha_{vir}=$5/3 corresponding to $\sigma_{v}=\sqrt{(GM_{c})/R_{c}}$). We use a range of 10$^{-1}$-10$^{0.5}$ g cm$^{-2}$ for $\Sigma_{c}$ because this covers the range of embedded cluster properties complied by \citet{FallEtAl-2010}. Similarly, we use masses from 10$^{2}$-10$^{6}$M$_{\odot}$ because this covers the range of cluster masses seen in nearby galaxies and in the Milky Way \citep{LadaLada-2003,FallEtAl-2009,ChandarEtAl-2010}. Looking at the four graphs, we see that the probability is very low for all four planets in almost all combinations of $\Sigma_{c}$ and $M_{c}$ within the plausible range of mass and surface density. The probability is less than 1\% for all four Jovian planets when the mass of the cluster is greater than 10$^{4}M_{\odot}$. Neptune, the furthest planet from the Sun and thus the most likely to be excited, has excitation probabilities $<$1\% over most of parameter space, and reaches a high of 8.7\% at $M_{c}=10^2M_{\odot}$, $\Sigma_{c}=10^{0.5}$ g cm$^{-2}$. \par A curious feature of Fig.\ \ref{fig:Jovian} is that as the mass of the cluster increases, the probability of an event decreases. To understand this, we first must realize that the number of encounters is independent of $M_{c}$ at fixed $\Sigma_{c}$. This is because $\Lambda \propto \frac{n_{c}t_{cross}} {\sigma_{v}^{3}}M_{c} \propto M^{0}_{c}$. Next, looking at equation (\ref{eq:VelocityDispersion}) we see that as $M_{c}$ increases, so does the velocity dispersion. Figure \ref{fig:JovImpact} shows the velocity dependence of the cross section obtained though our simulation. We see that, at low velocities around 1 km/s, the cross section is high, comparable to the values obtained by \citet{AdamsLaughlin-2001}. However, at slightly higher velocities the cross section dramatically decreases, dropping under (200 AU)$^{2}$ at around 3 km/s. Thus, at higher $M_{c}$ the velocity dispersion increases yet the number of encounters does not. This explains the result in Fig.\ \ref{fig:Jovian}: higher mass clusters are less likely to increase a Jovian planet's eccentricity because they produce no more encounters (at fixed $\Sigma_{c}$), but reduce the effective cross section per encounter. \par Recent observational and theoretical evidence indicates that young embedded clusters can be sub-virial \citep{FureszEtAl2008,TobinEtAl-2009,OffnerEtAl-2009} and thus we make the same calculations for a cluster with a virial parameter that is one fifth the value of a fully relaxed cluster, $\alpha_{vir}=$1/3. Figure \ref{fig:JovianVir} shows probability of exciting a Jovian planet to an eccentricity greater than 0.1 in a sub-virial cluster. We see that the probability is higher than that of a relaxed cluster, which makes sense due to the lower velocity dispersions, but the overall probability is still small. Again, Neptune is the likeliest to be excited with $\sim$15\% probability at the high $\Sigma_{c}$, low $M_{c}$ extreme, but as in a relaxed cluster, a majority of parameter space produces probability below 1\% for all of the Jovian planets. This agrees with \citet{ProszkowAdams-2009} who show that the interaction rate in a sub-virial cluster is greater than that in a virialized one, but not by much. \subsection{Effects of Close Encounters on the Kuiper Belt} We next check if close encounters could truncate the Kuiper Belt or excite an object to a Sedna-like orbit. To determine if the a close encounter could truncate the Kuiper Belt, we compute the expected fraction of the KBOs in each of our eight initial distance bins (see Section 2.1 for derivation). The resulting expected fractions in the 35-80AU and 455-500AU bins, which represent the extremes of our initial Kuiper Belt, are show in Fig.\ \ref{fig:KBOLost01CT} for relaxed clusters and Fig.\ \ref{fig:KBOLostVir01CT} for sub-virial clusters. We find that the expected fraction of KBOs stripped in the 35-80AU bin is less than 1\% for most of parameter space, only going above 1\% in the high $\Sigma_{c}$, low $M_{c}$ extreme, where the maximum is $\sim$4\%. The expected fraction of KBOs stripped in the 455-500 AU bin reaches 88\% in the high $\Sigma_{c}$, low $M_{c}$ extreme. As we move away from this extreme, the probability decreases drastically. This appears to suggest that a birth cluster with $\Sigma_{c}$ and low $M_{c}$ could explain the truncation of the Kuiper Belt. However, comparing Fig.\ \ref{fig:KBOLost01CT} with Fig.\ \ref{fig:Jovian} we note that any cluster capable of destroying the outer Kuiper Belt while leaving the inner Kuiper Belt undisturbed would also disrupt Neptune's orbit. There is no part of parameter space where the Kuiper Belt is truncated but Neptune is not also perturbed. This is most likely because no single encounter can truncate the Kuiper Belt. Instead, multiple close stellar passes would be required, and the number required is sufficiently large as to make it extremely likely that one of those encounters would be close enough to perturb Neptune. \par To see if a close encounter could produce a Sedna-like object, we compute the expected fraction of the KBOs excited to an eccentricity between 0.5 and 1.0 in a given distance bin. We show the result in Fig. \ref{fig:KBOExcited01CT} for relaxed clusters and Fig.\ \ref{fig:KBOExcitedVir01CT} for sub-virial clusters. Again, we see that the greatest probability is found in the high $\Sigma_{c}$, low $M_{c}$ extreme, with a significant drop-off as we move away from this part of the parameter space. Since Sedna has a semi-major axis of 542 AU and there is only one such object known, we focus on the 455-500 AU distance bin, and we only need a relatively small probability. We see reasonable probabilities exist over much of our parameter space. We conclude that it is plausible that a close encounter could produce a Sedna-like object. \subsection{Effects of Close Encounters on Tyche} \label{sec:tyche} To determine if Tyche could survive the birth cluster, we compute the cross section for Tyche becoming unbound. We first check to see if there was any difference in the cross-section depending upon which Tyche mass was used (see section 2.2). Using a chi-square test, we find that our measurements of the number of Tyche losses as a function of incoming star velocity at our three different Tyche masses are consistent with each other to $\sim$98\% confidence, and thus there is no reason to reject the hypothesis that the probability of Tyche becoming unbound is independent of it mass. Thus, we use a combination of data from all three masses. Looking at Fig. \ref{fig:Tyche}, as before, the highest probability, $\sim$44\%, is in the high $\Sigma_{c}$, low $M_{c}$ extreme, with the probability decreasing as we move away from this extreme. The probability is $\sim$10\% for moderate $\Sigma_{c}$ and M$_{c}$. Given this, there is a small likelihood that Tyche would be stripped away. However, recall that we place Tyche at 1,000AU, and while it is theorized to have a semi-major axis of $\sim$5,000 AU. Thus the probability of stripping likely exceeds our estimate by a considerable factor. Even with this increased, though, stripping seems unlikely in a cluster with mass $>10^{4}M_{\odot}$. \subsection{Encounter Effects in Long-Lived Clusters} \label{sec:longlife} Finally, we consider the possibility that the natal cluster did not disperse at the point where its gas was expelled, and instead lived on as an open cluster. As discussed above, about 10\% of embedded clusters experience this fate. In this scenario, the Sun would be exposed to many more close encounters. To study the effects of this scenario, we use $t_{life}=30$ Myr in equation \ref{eq:Lambda} instead of using equation \ref{eq:CrossingTime} and repeat our analysis from the previous sections. \par Figure \ref{fig:JovianMyr30} shows the probability of the Jovian planets being excited to an eccentricity greater than 0.1 in such a cluster. As we found for the more typical short-lived cluster population, the probability of disruption is smallest for clusters of high mass and low surface density. Not surprisingly, excitation of the Jovians is much more likely in a long-lived cluster, and the only part of parameter space that has low probabilities is the high mass, low surface density extreme. Outside this regime all the Jovian planets are likely excited to an eccentricity greater than 0.1. Thus we conclude that the low eccentricities of the planets are consistent with the long-lived cluster scenario, but only if the cluster in question was both quite massive and of low surface density. \par Figures \ref{fig:KBOLostMyr30} and \ref{fig:KBOExcitedMyr30} show the expected fraction of KBOs lost and excited, respectively, in a long-lived cluster. We see that the expected fraction lost is very high for KBOs in all distance bins except in clusters of very high mass and low surface density. There is a fairly narrow strip of parameter space that allows the outer Kuiper Belt to be destroyed but leaves the inner Kuiper Belt relatively untouched. However, comparing to Figure \ref{fig:JovianMyr30}, we see that in this parameter regime it is also very likely that Neptune would be driven to high eccentricity. The underlying reason is that same as for short-lived clusters: it takes many encounters to truncate the outer Kuiper Belt, and any cluster capable of producing that many encounters is also likely to supply one close enough to disturb Neptune. The behavior of Sedna is also similar to that seen in the case of short-lived clusters: there is a reasonably large part of parameter space where the Jovian planets would not be disturbed, but there is a reasonable change of exciting a distant KBO to Sedna-like eccentricities. \section{Discussion} Our conclusion that encounters in the Sun's natal cluster are likely to have very little effect on the architecture of the Solar System, even for very massive clusters, is at odds with much previous work, which \citet{Adams-2010} summarizes to conclude that dynamical arguments imply that the Sun's birth cluster could not have contained more than several thousand stars. This conclusion, however, is based on work that assumed the cluster lifetime was much greater than current observational evidence shows, that the surface density of a cluster varies linearly with the cluster's mass, and that only simulated encounters at fairly low velocities. Each of these assumptions increases the probability of disruption encounters. For example \citet{LaughlinAdams-1998} estimated $\mean{\sigma}$=(230 AU)$^2$ for disruption of Jovian planets due to interactions with binary stars in the birth cluster through scattering simulations similar to ours. To obtain this value, they randomly chose the velocity of the incoming star from a normal distribution with $\sigma=1$ km s$^{-1}$. From this, they determined the rate of disruptive encounters is $\sim$0.13 per 100 Myr. \citet{AdamsLaughlin-2001} used the same velocity distribution and obtained similar results. We can immediately identify the effects of changing observational data which lead us to different conclusions. First, consulting equation (\ref{eq:VelocityDispersion}), we note that a cluster of mass $10^{4}M_{\odot}$ and surface density 1 g cm$^{-2}$ will have a velocity dispersion of 7 km s$^{-1}$ (if it is virialized) or 3 km s$^{-1}$ (for $\alpha_{vir}=1/5$). As a result most encounters will be at significantly higher velocities than Laughlin \& Adams assumed; consulting Fig.\ \ref{fig:JovImpact}, we see that this will reduce the cross section by roughly an order of magnitude. The earlier work was also based on a lifetime of 100 Myr, while modern observations tells us that only $\sim$1\% of stars remain in a cluster for that long, and only $\sim 10\%$ survive to even 10 Myr \citep{LadaLada-2003,FallEtAl-2009,ChandarEtAl-2010}. Again, switching to modern values leads to a 1-2 order of magnitude drop in the disruption probability. Together the decrease in cross section and in exposure time are sufficient to explain why we find no dynamical limit on the size of the Sun's parent cluster. \par Changing the assumed cluster velocity dispersion and lifetime also explains why our conclusions on the Kuiper Belt differs from that of earlier research. \citet{LestradeEtAl-2011} concluded that it was possible to truncate a debris disk around a main sequence star in an open cluster by a close encounter. To reach this conclusion, however, they assumed parabolic, coplanar encounters over a 100 Myr cluster lifetime, compared to our observationally-favored scenarios of mostly hyperbolic, non-coplanar encounters over a $\sim$1 Myr cluster dissolution timescale. \citet{Jimenez-Torres-2011} created a simulation that was able to produce Sedna-like objects but assumed incoming star velocities of 1, 2 and 3 km s$^{-1}$, closest approach distances of 200 AU and cluster lifetimes of 100 Myr. We concur with these authors that it is possible to truncate the Kuiper Belt in a cluster that lives 30 Myr, and that it may be possible to produce Sedna even in a shorter lived cluster. That the Sun might have been born in a long-lived cluster is by no means impossible, since $\sim 10\%$ of stars are. We simply point out that such a scenario is not typical. Moreover, we find that, regardless of the cluster lifetime, it is very difficult to truncate the outer Kuiper Belt without simultaneously disrupting the orbit of Neptune, a problem not considered in earlier work. Some other authors who have used the current observational data about cluster lifetime and velocities have partially anticipated the results presented here. \citet{BonnellEtAl-2001} argue that planet formation is unaffected in open clusters, and that it was possible for a the Sun to be born within a cluster whose density is greater than 10$^3$ stars pc$^{-3}$ if the cluster dissolves within $\sim$10 Myr. Using a cluster membership size of 1000 and a cluster lifetime of 10 Myr, \citet{AdamsEtAl-2006} showed that typical stellar passes are not close enough to appreciably enhance the eccentricity of Neptune. Their main conclusion was that planet-forming disks and newly formed solar systems generally survive their aggregates with little disruption. \citet{SpurzemEtAl-2009} constructed a N-body simulation of the Orion Nebula Cluster, an environment containing massive stars similar to the hypothetical progenitors of the supernova(e) that produce the isotopes we see in meteorites. Their conclusion was that the critical threshold for the survival of wider orbit planets (those similar to our Jovian planets) is a cluster lieftime less than 10$^8$ yr. \par We note that \citet{BrasserEtAl-2006} have shown through N-body simulations that both close encounters and cluster tidal effects have a reasonable probability of creating a Sedna-like object if the object begins at a semi-major axis $\ge$300AU, consistent with our conclusions (see Fig.\ \ref{fig:KBOExcited01CT} and Fig.\ \ref{fig:KBOExcitedVir01CT}). They hypothesize a two-step formation process for Sedna in which the outward migration of Neptune scatters a KBO into an orbit with large semi-major axis, and the object is scattered again by passing stars. Our work is not inconsistent with this scenario. \section{Conclusion} In this paper we revisit the tension between meteoritic and dynamical constraints on the size of the Solar System's parent star cluster. The meteorite evidence suggests that a supernova deposited short-lived radioactive isotopes near the forming Solar System. However, based on dynamical arguments about the likelihood of a close encounter disrupting the outer Solar System, \citet{AdamsLaughlin-2001} and \citet{Adams-2010} argue that the birth cluster could not have had no more than several thousand members. The probability of having a supernova within the range needed in a cluster of that size is only about 2\%. Given that we have only a single example (the Sun) of a star system that simultaneously has supernova enrichment and outer planets in circular orbits, this isn't necessarily impossible. It could simply be that our Solar System is quite unusual. However, we find that no such conclusion is warranted. We show that repeating the calculations of \citet{AdamsLaughlin-2001} but with embedded cluster properties drawn from recent observations, we can update the upper limit on birth cluster size to values where supernova pollution is much more probable. \par Indeed, we show that as the mass of the birth cluster increases, the relative velocities of the constituents increases and the effective impact parameters for Solar System-disrupting events decreases, yet the number of expected encounters is constant. Our result differs from previous work because we are using current observational evidence that suggests that the cluster's surface density is independent of its mass, and its lifetime is much shorter than previously believed. The removal of the upper bound from dynamics allows the cluster size to be large enough to be able to produce a massive star which can go supernova to seed the Solar System with the observed short lived radioactive isotopes. Indeed, at cluster sizes of 10$^4$-10$^5$ stars, multiple supernovae are likely, allowing each supernova to be further away and exposing the Solar System to less mass loss. \par A close encounter within the birth cluster has also been proposed to explain the existence of Sedna. We find that this is within reason, showing that it is possible to produce a small fraction of KBOs at $\sim$500 AU with eccentricity between 0.5 and 1. However, we find that it is highly unlikely that encounters in the birth cluster could be responsible for truncation of the Kuiper Belt. There is only a small region of parameter space that allows for the destruction of the outer Kuiper Belt while leaving the inner Kuiper Belt intact, and any cluster capable of disrupting the outer Kuiper Belt would also be very likely to excite Neptune to a higher eccentricity than we observe. Finally, we caution that disruption of planetary orbits by the gravity of passing stars may not be the only limiting factor when it comes to cluster size. \citet{Adams-2010} argues that in a cluster of more than $\sim 10^4$ stars, the ultraviolet radiation produced by massive stars would photoevaporate the Sun's protoplanetary disk, preventing formation of the outer planets. However, this conclusion is highly sensitive to the rate of mass loss due to photoevaporation, which is highly uncertain. The $10^4$ star limit is based on a loss rate taken from \citet{AdamsEtAl-2004}, but \citet{ErcolanoEtAl-2009} show that this rate is uncertain at the order of magnitude level. Moreover, it is not entirely clear that photoevaporation inhibits planet formation; instead, by raising the dust-to-gas ratio, it may trigger gravitational instability \citep{Throop-2005}. Due to these uncertainties, the question of whether photoevaporation might provide a limit on the cluster size remains an open one for future research. \acknowledgements We thank G.~Laughlin for extensive discussions and comments on the manuscript. MRK is supported by the Alfred P.~Sloan Foundation, the National Science Foundation through grants AST-0807739 and CAREER-0955300, and NASA through Astrophysics Theory and Fundamental Physics Grant NNX09AK31G and a Chandra Space Telescope Grant.
\section{\label{sec:Intro}Introduction} Complex materials often display rich phase diagrams where different types of instabilities lead to unusual properties related to various ordering phenomena. Understanding the interplay between lattice, charge, spin and orbital degrees of freedom is one of most challenging aspects of correlated electron systems. Barium vanadium sulfide, BaVS$_3$, is a particularly intriguing material because its structural, electrical and magnetic properties point to a combination of quasi-1D and 3D behavior. Its room temperature structure consists of hexagonal packing of quasi 1D-chains of face-sharing VS$_6$ octahedra directed along the \textbf{c} axis \cite{Gardner}. The ratio of inter-chain to intra-chain V-V distances being greater than 2, a high conductivity anisotropy ratio ($\sigma_c$/$\sigma_a$) can be expected. Surprisingly, $\sigma_c$/$\sigma_a$ is only about three \cite{Mihaly}. Lowering the temperature induces a zig-zag deformation of the V chains at $T_S$ = 240~K and reduces the crystal symmetry from hexagonal to orthorhombic \cite{Gardner}, but does not significantly change its conductive and magnetic properties. In both phases, it is a paramagnetic metal. Below $T_{MI}$ = 69~K it becomes a paramagnetic insulator (PI). The metal-insulator (MI) transition is driven by a Peierls instability \cite{Fagot1}, accompanied by a change to a monoclinic structure \cite{Inami}. At $T_N$ = 31 K, it undergoes a third phase transition to the lowest temperature (ground) state which has an incommensurate antiferromagnetic (AF) order in the (\textbf{a},\textbf{b}) plane \cite{H_NakamuraJPSJ}. Recent resonant magnetic x-ray scattering reveal additional incommensurability along the \textbf{c} axis in the AFI ground state, while the supporting time-dependent density functional theory simulations indicate that the spins lie within the (\textbf{a},\textbf{b}) plane and are polarized along the monoclinic \textbf{a} axis \cite{Leininger}. Diffuse X-ray scattering experiments \cite{Fagot1} have demonstrated that the MI transition is driven by a 1D instability of the electron gas. It leads to the formation of charge density waves (CDW) at the critical wave vector \boldmath$q$$_{c}$ = 2\boldmath$k$$_F$, preceded by pretransitional fluctuations. But the physical properties of BaVS$_3$ are not those of a standard quasi-1D system. First, the wave vector $q_{c}$\unboldmath = 0.5~\textbf{c*} (\textbf{c*} stands for the reciprocal lattice vector related to \textbf{c}) indicates that only one of the two $d$ electrons per unit cell ($uc$) participates directly in the CDW. On the other hand, instead of a typical Pauli behavior of a 1D electron gas in the metallic phase, it has a Curie-Weiss behavior with an effective magnetic moment of $\mu$ = 1.2 $\mu_{B}$ corresponding to approximately one localized spin per two vanadium sites \cite{Mihaly}. All these properties point to an unusual situation with $one$ electron per $uc$ delocalized along the \textbf{c*} direction driving the CDW formation. The $other$ is localized and responsible for the magnetic behavior. Local density approximation (LDA) calculations indicate an interplay between two different types of electron states at the Fermi level: two narrow $E_g$ bands, and one dispersive band with mainly $A_{1g}$ character extending along the \textbf{c*} direction \cite{Mattheiss,Whangbo}. But the LDA filling of the $A_{1g}$ band is almost complete ($n_{A_{1g}}$ $\approx$ 1.90 per $uc$), which overestimates the itinerant character, and is incompatible with the above-mentioned experimental findings. Dynamical mean-field theory (DMFT) calculations show that strong enough charge correlation and exchange effects can bring $n_{A_{1g}}$ down to $one$ per $uc$ \cite{Lechermann}. \begin{table} \begin{center} \begin{tabular}{lllll} \hline phase & AFI & mPI & oPM & hPM \\ $T_c$ & $T_N$= 31~K & $T_{MI}$= 69~K & $T_S$= 240~K & \\ structure & monoclinic & monoclinic & orthorhombic & hexagonal \\ space group & $Im$ & $Im$ & $Cmc2_1$ & $P6_3mmc$ \\ point group & $m$ (C$_{1h}$) & $m$ (C$_{1h}$) & $mm2$ (C$_{2v}$) & $\overline{3}m$~(D$_{3d}$)\\ N$_V$ (N$_{ineq.}$) & 4 (4) & 4 (4) & 2 (1) & 2 (1) \\ magnetism & AF & para & para & para \\ transport & insulator & insulator & metal & metal \\ \hline \end{tabular} \end{center} \caption{\label{tab_phases} Physical properties of BaVS$_3$ in its four phases: Anti-Ferromagnetic Insulator (AFI), monoclinic Paramagnetic Insulator (mPI), orthorhombic (oPM), and hexagonal (hPM) Paramagnetic Metal. $T_c$ stands for critical temperature, and N$_V$ (N$_{ineq.}$) the number of (inequivalent) V atoms per unit cell.} \end{table} Both types of electron states, delocalized (charge degrees of freedom), and localized (magnetic properties), are modified at the MI transition. A charge gap of about 50~meV \cite{Mihaly} and a spin gap of 10-20~meV \cite{NakamuraPRL} are opened at $T_{MI}$, even if the spin degrees of freedom quench only in the AFI phase. But the mechanism of their interplay is not yet completely understood. We have explored four BaVS$_3$ phases in single crystals by polarization dependent x-ray absorption spectroscopy (XAS) at the V L edges. After describing the experimental means for obtaining very high resolution contamination-free data we present the results and then discuss how the polarization dependence of the XAS data provides new insight into the subtle electronic structure changes that take place across a temperature range that covers all the phases of BaVS$_3$. \section{\label{sec:Exp}Experiment} Single crystals of BaVS$_3$ were grown by the tellurium flux method \cite{Berger}. Two samples from different batches were cleaved $in$-$situ$ along a plane parallel to the \textbf{c}-axis. After the cleaving at a pressure of 10$^{-9}$~mbar, they were transferred via a fast insertion lock to the analysis chamber at the base pressure of 4 x 10$^{-10}$~mbar. The experiments were performed using the helical undulator Dragon beam line ID08 at the European Synchrotron Radiation Facility (ESRF). At the V L$_3$ edge ($\approx$ 514~eV) the resolution was $\approx$ 100~meV. The incident light was normal to the sample surface and linearly polarized either parallel to the sample \textbf{c}-axis, or perpendicular to it. We periodically checked for energy shifts coming from the monochromator. The stability of the photon energy is estimated to be better than 40~meV. The V L edge XAS signal was measured in the total electron yield mode using the drain current from the sample holder. \begin{figure} \begin{center} \includegraphics [width=8.5cm,angle=0]{fFIG_1.eps} \caption{\label{fig:XAS_pol_dep} (Color online) Paramagnetic Metallic phase (T = 300~K) XAS spectra of the $in$-$situ$ cleaved sample for the two polarizations of the incoming light: \textbf{E}$\perp$\textbf{c} (red dots) and \textbf{E}$\parallel$\textbf{c} (black dots). The linear dichroism (LD = I$_{\perp}$-I$_{||}$) is also shown in the same graph (green line). The inset shows the spectra of a sample cleaved in air where the oxygen contamination is clearly visible above 530 eV. Extra features are observed at the V L$_3$ edge and the L$_2$ peak is broadened.} \end{center} \end{figure} \section{\label{sec:Res}Results} V L$_{2,3}$ XAS results for BaVS$_3$ at room temperature for the two polarizations of the incoming light, \textbf{E}$\perp$\textbf{c} and \textbf{E}$\parallel$\textbf{c} are presented in Fig.~\ref{fig:XAS_pol_dep}. The spectra are normalized so as to obtain the same difference in amplitude below the L$_3$ edge and at 534~eV. They are separated by the 2$p$ spin orbit coupling, into L$_3$ ($h\nu$ = 512-519~eV) and L$_2$ ($h\nu$ = 519-528~eV) regions. The V L$_3$ and L$_2$ maxima are at 516.2~eV and 523.2~eV, respectively, i.e. 7~eV apart. Their overall shape is closer to the spectra of vanadium metal \cite{Ilakovac} than to the spectra of vanadium oxides that tend to be more structured (see for example VO$_2$ \cite{Abbate} or V$_2$O$_3$ \cite{Park2000}). The O K XAS signal lies in the 530-545~eV region. By comparing with the inset in Fig.~\ref{fig:XAS_pol_dep} it can be seen that the surface contamination is negligibly small. In particular the hump at 540~eV is very weak while the feature at 531 eV may even be intrinsic to the BaVS$_3$ spectrum. A similar feature is present in elemental vanadium, where it has been attributed to a van Hove singularity at the boundary of the V metal Brillouin zone \cite{Scherz}. The inset shows clearly that the oxygen contamination affects the V L edge. Additional structures show up at the high energy side of the V L$_3$ peak, and the V L$_2$ structure is broadened. In this sense, at least the surface of the slightly Ti-doped sample previously studied \cite{Learmonth} was contaminated by oxygen. The V~L$_{2,3}$ maxima for the \textbf{E}$\parallel$\textbf{c} polarization are at slightly higher photon energy than for the \textbf{E}$\perp$\textbf{c} polarization. The shift is about 80~meV in the hPM phase. The peak position for V~L$_3$ with \textbf{E}$\perp$\textbf{c} excitation is 516.15~eV. In order to show more clearly the polarization dependence of the XAS intensity, we subtracted $I_{||}$ (\textbf{E}$\parallel$\textbf{c} spectra) from $I_{\perp}$ (\textbf{E}$\perp$\textbf{c} spectra). This linear dichroism (LD) curve is presented in Fig.~\ref{fig:XAS_pol_dep} for the hPM phase. The change from positive to negative LD simply reflects the shift in the main XAS peak according to the polarization of the incident photons. This derivative-like shape hardly changes except for differences in amplitude as a function of the temperature or phase (Fig.~\ref{fig:BaVS3_LD}). When cooling to below $T_S$ = 240~K (hPM to oPM), the shift increases to 110~meV. As a simplification we show the average spectrum taken at 300~K, 275~K and 250~K to represent the hPM phase, what we term the high temperature orthorhombic phase (oPM-HT) is taken as the average of spectra recorded at 225~K, 200~K, 175~K, and 150~K, and the low temperature orthorhombic phase (oPM-LT) spectrum is the average of data taken at 125~K, 100~K and 75~K. The mPI spectrum was recorded at 50~K and the AFI spectrum at 25~K. The shape of the main V L$_3$ XAS peak remains practically unchanged but the amplitude of the LD signal at 515.8~eV changes reflecting very small variations in the relative intensity of the $I_{||}$ and $I_{\perp}$ maxima. Another peak in the LD is observed at 514.2~eV, coinciding with the shoulder on the XAS data. Its amplitude is positive and remains practically unchanged across the range of measurements. An essentially equivalent situation is observed for the L$_2$ edge. Here structures are not well resolved. This can be explained by the difference in lifetime broadening of the two edges, which is 0.78~eV (0.28~eV) for the V L$_2$ (L$_3$) \cite{Campbell2001}.\\ \begin{figure} \begin{center} \includegraphics [width=8.5cm,angle=0]{fFIG_2.eps} \caption{\label{fig:BaVS3_LD} (Color online) LD as a function of temperature. Red/light-green/dark-green/blue/black curves correspond respectively to the hPM/oPM-HT/oPM-LT/mPI/AFI temperature ranges.} \end{center} \end{figure} A detailed examination of the low energy part of the spectra, brings to light structure at 512.8~eV. The LD data are shown in Fig.~\ref{fig:BaVS3_LD_PreEdge}. As the temperature is lowered to the oPM-LT temperature range the intensity is increased at 513.2~eV, where a relatively strong extra peak is clearly present in both the mPI and AFI phases. It should be underlined that we have observed the same low temperature behavior for $two$ samples from different preparations. Fig.~\ref{fig:BaVS3_LD_sec} shows the LD for the sample from the second batch demonstrating that the data are highly reproducible. \begin{figure} \begin{center} \includegraphics [width=8.5cm,angle=0]{fFIG_3.eps} \caption{\label{fig:BaVS3_LD_PreEdge} (Color online) LD changes with temperature in the pre-edge region. The spectra are shifted vertically for clarity. Inset shows details of the temperature changes in the oPM-LT range.} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics [width=8cm,angle=0]{fFIG_4.eps} \caption{\label{fig:BaVS3_LD_sec} (Color online) LD changes with temperature in the pre-edge region for the sample from another preparation is essentially the same.} \end{center} \end{figure} \section{Discussion} \label{sec:Disc} An elaborate theoretical interpretation of the BaVS$_3$ XAS spectrum is not presently available because of its structural complexity. We will therefore discuss our data in terms of a simplified analysis of orbital symmetries to help assess what information can be extracted from the XAS polarization dependence, along with our present knowledge of the physical properties of the material and the insight that can be gained from recent LDA-DMFT calculations \cite{Lechermann2007}. First we need to calculate the polarization dependence of the $2p$ to $3d$ dipole transition in the material to assess the polarization effects. Due to the inclination of the V-S octahedra in BaVS$_3$, it is convenient to present its V~$3d$ states as a linear combination of cubic $d$ orbitals (Eq.~\ref{d-orbitals} and Fig.~\ref{fig:Orbitale}): \begin{eqnarray} \label{d-orbitals} \lefteqn{A_{1g}=d_{z^2}}\nonumber\\ E_{g_1} &=& - \sqrt{\dfrac{2}{3}}d_{x^2-y^2}+\sqrt{\dfrac{1}{3}}d_{yz}\nonumber\\ E_{g_2} &=& \sqrt{\dfrac{2}{3}}d_{xy}-\sqrt{\dfrac{1}{3}}d_{xz}\\ e_{g_1} &=& \sqrt{\dfrac{1}{3}}d_{x^2-y^2}+\sqrt{\dfrac{2}{3}}d_{yz} \nonumber\\ e_{g_2} &=& \sqrt{\dfrac{1}{3}}d_{xy}+\sqrt{\dfrac{2}{3}}d_{xz},\nonumber \end{eqnarray} The 2$p$ $\rightarrow$ 3$d$ transition probabilities, according to dipolar selection rules, are given by \begin{equation} \label{transitions} P_{d} = \sum_{i=1}^3 \left| \langle p_i |\textbf{E} \cdot \textbf{r}| d \rangle \right|^2 \end{equation} where $p_i$ represents three V 2$p$ orbitals and $d$ can be any of the five V $3d$ orbitals: $A_{1g}$, $E_{g1}$, $E_{g2}$, $e_{g1}$, $e_{g2}$. $\textbf{E} \cdot \textbf{r}$ is the dipolar transition operator. \begin{figure}[h] \begin{center} \includegraphics [width=8.5cm,angle=0]{fFIG_5.ps} \caption{\label{fig:Orbitale} (Color online) Five BaVS$_3$ V 3d orbitals corresponding to Eq.~\ref{d-orbitals}: two e$_g$ (upper panels) and three t$_{2g}$ (lower panels). They are inserted in the orthorhombic phase deformed octahedra, with two inequivalent sulfur atoms, S$_1$ (orange) and S$_2$ (yellow). Axes (x,y,z) correspond to (a,b,c) crystallographic axes in the orthorhombic phase.} \end{center} \end{figure} The polarization dependent transition probabilities, $P^{\perp}$ for \textbf{E}$\perp$\textbf{c} and $P^{\parallel}$ for \textbf{E}$\parallel$\textbf{c}, are the following:\\ - Transitions to the higher energy $e_g$ states are not polarization dependent. For both polarizations their probability is $P_{e_g}^{\perp}$ = $P_{e_g}^{\parallel}$ = 0.4.\\ - Transitions to the $A_{1g}$ states are strongly polarization dependent, $P_{A_{1g}}^{\perp}$ = 0.1 and $P_{A_{1g}}^{\parallel}$ = 0.4.\\ - Transitions to $E_{g}$ states are also polarization dependent, $P_{E_{g}}^{\perp}$ = 0.5 and $P_{E_{g}}^{\parallel}$ = 0.2. Thus the total polarization dependent absorption intensity as a function of energy can be written as: \begin{eqnarray} \label{intensities} I_{\perp, \parallel}(\epsilon) \propto n_{A_{1g}}(\epsilon) P_{A_{1g}}^{\perp, \parallel} + [n_{E_{g1}}(\epsilon) + n_{E_{g2}}(\epsilon)] P_{E_{g}}^{\perp, \parallel} \nonumber \\ + [n_{e_{g1}}(\epsilon) + n_{e_{g2}}(\epsilon)] P_{e_{g}}^{\perp, \parallel}, \end{eqnarray} where $n_d(\epsilon)$ are unoccupied spectral function weights at a particular final (core-excited) state energy $\epsilon$. As the transitions to the $e_g$ states are equiprobable for both polarizations, the linear dichroism should be zero for excitations to states with this symmetry. The difference in the transition probabilities to $A_{1g}$ is equal to that of each individual $E_{g}$ orbital but of opposite sign. Thus an increase in the population of $A_{1g}$ states will tend to reduce the LD or give a negative signal while an increase in the $E_{g}$ population will tend to increase the LD when defined as $(I_{\perp}~-~I_{\parallel})(\epsilon)$. \begin{figure} \begin{center} \includegraphics [width=9cm,angle=0]{fFIG_6.eps} \caption{\label{fig:Frank_LDA_DMFT} Local spectral functions from LDA+DMFT calculations of the $t_{2g}$ states in the oPM phase where all V sites are equivalent (upper panel) and mean value of local spectral functions for four inequivalent V-sites for the $t_{2g}$ states in the mPI phase (lower panel). The LDA density of $e_{g}$ states shown in both panels is the one calculated for the oPM phase, as the calculation for the mPI phase is not available. Fermi level energy is set to 0. LDA and DMFT data have been adapted from the Figs. 2, 7 and 12 of Ref.\onlinecite{Lechermann2007}.} \end{center} \end{figure} Summarizing our observations, little difference in the polarization dependence can be detected across the T$_S$~=~240~K hPM to oPM phase transition despite the symmetry break which lifts the degeneracy of the two $E_{g}$ and two $e_{g}$ states as shown in Fig.~\ref{fig:Frank_LDA_DMFT} (upper panel). Lowering the temperature below 150~K (oPM-LT) induces an increase in the LD amplitude at 515.8~eV, while the intensity starts to build up at 513.2~eV, where a distinct peak appears below T$_{MI}$ = 69~K. When the temperature is further reduced below the mPI to AFI ($T_\chi$ = 30~K) phase transition the intensity at 515.8~eV again drops and the 513.2~eV peak intensity strengthens. This could mean that spin ordering has the effect of transfering one part of unoccipied $E_{g}$ states to lower energy. The increase in LD amplitude at 513.2 eV and 515.8~eV across the oPM-HT to oPM-LT temperature range might be interpreted as an increase of the $E_g$ weight in the corresponding small energy range. As the magnetic properties are related to the $E_g$ states, this interpretation is in line with the energy dependence of the ground state magnetic reflection at the (0.226 0.226 0.033) wave vector which shows two V L$_3$ peaks at similar positions \cite{Leininger}. On the other hand, in the oPM-LT temperature range, there is a formation of CDWs in chains, as shown by pretransitional diffuse lines in x-ray diagrams \cite{Fagot1}. It is a dynamical effect, with a typical lifetime of 10$^{-12}$~s \cite{PougetKMO}, which can induce local structure modifications via electron-phonon interaction. As the time scale of XAS is determined by the V L$_3$ edge core-hole-lifetime of 10$^{-15}$~s, it can probe short-time modifications of unoccupied states related to CDW fluctuations. Similar dynamical local structural change is observed at the Ni K-edge NEXAFS of PrNiO$_3$ \cite{Acosta} system. Over the oPM-HT temperature range all V sites are equivalent by symmetry, while in the mPI phase there are four inequivalent V-sites in the monoclinic unit cell \cite{FagotS}. Resonant diffraction measurements at the V K edge pointed to the possibility of an orbital ordering of $A0E0$ type ($A_{1g}$, equal, $E_{g1}$, equal) accompanying the CDW formation \cite{Fagot2}. Collective intersite orbital excitations measured in the dielectric signal support this hypothesis \cite{Ivek}. DMFT calculations confirm that the four inequivalent V-sites have different spectral functions but propose rather an $EE00$ occupation pattern \cite{Lechermann2007}. In terms of $unoccupied$ states, two sites have their $E_{g2}$ local spectral function maximum closer to E$_F$ and the other pair have the maximum pushed to higher energy, giving rise to a broad double peak when the mean value for four inequivalent V-sites is presented as in the Fig.~\ref{fig:Frank_LDA_DMFT} (lower panel). From these considerations we suppose that, below 150~K, the increased LD intensity at 513.2~eV and 515.8~eV is related to the strong high energy transfer of a part of unoccupied $E_{g}$ states. This shift is itself related to the formation of four inequivalent V-sites which accompanyies CDW formation in the pretransitional fluctuation regime. Possible core-hole effects are difficult to estimate without proper theoretical modeling. It is however reasonable to suppose that difference in core hole interaction across the phases is negligibly small. \section{Conclusion} \label{sec:Concl} BaVS$_3$ single crystal V L edge x-ray linear dichroism (LD) spectra show changes with temperature. Below 150~K there is a clear LD intensity increase at 513.2~eV and 515.8~eV. Simple symmetry analysis suggests the effect is related to rearrangements in $E_{g}$ and $A_{1g}$ states. Two V L$_3$ peaks at similar energies are observed in the energy scan of the ground state (0.226 0.226 0.033) reflection measured by the resonant magnetic x-ray scattering \cite{Leininger}. This confirms that the observed increase of the LD intensity is related to magnetically active $E_g$ states. DMFT calculations \cite{Lechermann2007} predict rearrangements of $E_{g}$ states in the mPI phase, together with CDW-related formation of four inequivalent vanadium sites with different $E_{g}$ and $A_{1g}$ spectral functions. Our data show that the $E_{g}$ rearrangements start at higher temperatures with the formation of CDWs in chains, affecting $A_{1g}$ states. We expect these experimental results to help ongoing efforts to model this complex material as they highlight the need to consider $A_{1g}$ and $E_{g}$ states as a whole to understand the physical properties of this unusual compound. \acknowledgments Fruitful discussions with P. Foury, J.-P. Pouget, S. Tomi\'{c}, F. Lechermann, S. Bari\v{s}i\'{c}, I. Kup\v{c}i\'{c}, M. Grioni, P. Ben\v{c}ok, M.-A. Arrio, P. Sainctavit and Y. Joly are gratefully acknowledged. The work in Lausanne was supported by the Swiss National Science Foundation and its NCCR ``MaNEP''.
\section{Introduction} The modern life always demands new technologies that should be fast, mobile and at low cost. It is then necessary to miniaturise current devices, and a question arises whether we need to adapt the existing technologies or build one ``ex novo''. While the limits of present-day technologies have been exposed, as miniaturisation of current computer architectures towards nanoscale will be hindered by quantum mechanical effects, on the other hand quantum mechanics itself can also help us by offering currently unused electrons properties to be efficiently exploited. The conventional classical devices rely on the charge of the electrons to produce energy, encode and manipulate information, but recently new and useful devices are being developed that utilise the spin of the electrons as a carrier and processor of information, bearing the promise to create a new generation of devices which will be smaller, more versatile and more robust than those currently made by semiconductors. In this context, the family of (III,Mn)V ferromagnetic semiconductors represent prominent candidates for the spintronics industry. In particular the (Ga,Mn)As compound has attracted much attention for its potential applications in non-volatile memories, spin-based optoelectronics and quantum computation.\cite{Ohno,Loss}. Ferromagnetic semiconductors are interesting also from a fundamental physics side, in particular to understand the nature of the magnetic interactions that underlie the ferromagnetism and to investigate the microscopic origin of the magnetic anisotropy in these compounds \cite{Tomas, Tomas_2}. It has been known that ferromagnetic (Ga,Mn)As films are characterised by a substantial magnetic anisotropy \cite{Anisotropy, Anisotropy_2} (consisting of a cubic and a uniaxial components), which has been in general related to strain, hole concentration and temperature.\cite{Dietl, Abolfath,Welp,Sawicki}. Several attempts have been performed to tune the magnetic anisotropy by designing 1D nano-objects using lithographic methods but the possibility of growing self organised ordered 1D structures has remained unexplored so far and triggered the work presented in this paper. Very recently, we applied atomic force microscopy and grazing incidence X-ray diffraction measurements to reveal the presence of ripples on the surface of (Ga,Mn)As layers grown on GaAs(001) substrates and buffer layers \cite{APL}. A connection between the surface anisotropy that characterises the distribution of the ripples and the uniaxial magnetic anisotropy has been suggested, deserving further investigation. In this paper we report a detailed study of the structural surface properties of (Ga,Mn)As by means of atomic force microscopy (AFM). This technique is regarded as a very useful tool to investigate the surface of non-conductive samples with a very high resolution at {\rm nm} scale \cite{Binning, review}. We have analysed a variety of (Ga,Mn)As samples with thickness ranging from $5$ to $25 {\rm nm}$, observing self-organised periodic ripples aligned along the $[1\overline{1}0]$ crystallographic direction on the surface of all the measured samples. By using a Fourier power spectral density (PSD) analysis \cite{Fang}, we have obtained a quantitative model of the periodicity of the surface ripples for each sample, calculating their amplitude and effective period. The amplitude is related to the root mean square roughness, which is around $(0.38 \pm 0.10) {\rm nm}$ for all the observed samples, while the effective period, which provides an estimate of the width of each ripple, stays around $(53 \pm 12) {\rm nm}$. We have further analyzed a handful of (Ga,Mn)(As,P) samples finding no significant change in their structural properties compared with the case of P-free compounds. We argue that the combination of the AFM data and the PSD analysis provides a complete and quantitative description of the surface of the (Ga,Mn)As compound. \section{Experimental data and analysis} (Ga,Mn)As films with different thickness, $5 {\rm nm}$, $7 {\rm nm}$ and $25 {\rm nm}$ and Mn concentrations, 6\% and 12\%, and (Ga,Mn)(As,P) films with different P concentrations, 3\%, 6\% and 9\%, have been deposited on GaAs(100) substrates by low temperature molecular beam epitaxy, the details are described elsewhere. \cite{preparation,prep2} AFM images were obtained by using an Asylum Research MFP-3D atomic force microscope and Asylum research silicon probes have been used as tips. Before performing measurements the samples have been cleaned with a 1:3 $HCl$:$H_2O$ to remove the superficial oxide layer. Measurements on uncleaned samples have shown the same surface structures but with additional surface contamination indicating that the cleaning was not modifying the surface morphology. The majority of the images were taken on $1\times 1 \mu{\rm m}^2$ areas along two different crystallographic direction, $[1\overline{1}0]$ and $[110]$. In Figs. \ref{image1}(a) and \ref{image2}(b,c) we show some typical images for two different samples. We can see, clearly, that some ripples grow aligned along the $[1\overline{1}0]$ direction. Similar surface structures have been observed in all the measured samples, including the phosphorate ones. \begin{figure}[t!] \centering \includegraphics [width=14cm]{Fig1.pdf} \caption{(a) AFM image of (Ga,Mn)As sample Mn351, $5{\rm nm}$, 2\% Mn; self-organised ripples grow along the $[1\overline{1}0]$ direction. (b) Histogram of the RMS roughness distribution across the horizontal scan lines for the same sample; its global RMS is $(0.38 \pm 0.04) {\rm nm}$.} \label{image1} \end{figure} To have an estimation of the periodicity of the features emerging from the AFM images we have used the Fourier analysis. The inset of Fig. \ref{image2}(a) shows a typical 2D PSD spectrum. One can clearly see the surface anisotropy from the shape of the graph, showing that periodicity across the horizontal direction is strongly enhanced compared to the vertical one (for an isotropic image, the 2D PSD spectrum should accumulate along the 45$^\circ$ diagonal). In order to properly take into account the observed anisotropy, and to achieve a more accurate quantitative information about the structural periodic features, we decompose our images into horizontal scan lines and apply Fourier analysis to each line, combining the data at the end. \begin{figure*}[tb] \includegraphics [width=16cm]{Fig2.pdf} \caption{AFM data and Fourier PSD analysis on the (Ga,Mn)As sample Mn381. (a) AFM image of a $5 \times 5 \mu{\rm m}^2$ area of the sample, with neat evidence of the ripples along the $[1\overline{1}0]$ direction; the inset shows the 2D PSD spectrum, clearly signaling the anisotropy of the surface. (b) 3D AFM image of an $1 \times 1 \mu{\rm m}^2$ area of the same sample. (c) Line-by-line PSD showing the Fourier pattern along each horizontal scan line, for the image in (b). Reconstructed image data from a filtered Fourier transform where only a narrow window of harmonics around the effective mean period $\tau \pm 2\delta\tau$ is retained, for the sample in (b).} \label{image2} \end{figure*} The heights along a horizontal scan line can be collected in a vector $z(y)=\{z_r\}$, where the index $r$ runs from $1$ to the number $n$ of points (e.g. $n=1024$). The fast Fourier transform can be computed as \begin{equation} \phi_s=\frac{1}{\sqrt{n}} \sum_{r=1} ^n z_r e^{2 \pi i (r-1)(s-1)/n} \end{equation} here $\phi_s$ is the Fourier amplitude corresponding to a (dimensionless) frequency $f_s=s-1$. Each harmonics, corresponding to a periodic function with period $T_s=L/f_s$, is given by: $z_r=\frac{1}{\sqrt{n}} \sum_{s=1} ^n \phi_s e^{-2 \pi i (r-1)(s-1)/n}$, where $L$ is the window length in ${\rm nm}$ (e.g. $L=1000$). The PSD is defined as $p_s=|\phi_s|^2$ [see Fig. \ref{image2}(c)]. For each horizontal scan line, we have calculated the RMS roughness as standard deviation of the $z(y)$ height data. The RMS can be equivalently computed from the total PSD and it is equal (up to a factor $\sqrt{2}$) to the amplitude ($A$) of the oscillations of the periodic structures: \begin{equation} A=\sqrt{2} {\rm RMS}=\frac{\sqrt{2 \sum_{s=1}^n p_s}}{n-1} \end{equation} An histogram showing an exemplary distribution of the RMS roughness across the horizontal scan lines is depicted in Fig. \ref{image1}(b). For each image we have calculated the ``effective'' global RMS (and its error) as the mean (and standard deviation) of the roughnesses per each horizontal scan line. The values of the global RMS vary between $0.28 {\rm nm}$ for a sample $25 {\rm nm}$ thick with $6\%$ Mn and $0.48 {\rm nm}$ for a sample $7 {\rm nm}$ thick with $6\%$ Mn. Concerning the periodicity of the structures, we define the mean frequency per line, $\overline{f}(y)$, as an average of the frequencies of all Fourier harmonics (along a given horizontal scan line) weighted with the corresponding PSDs: \begin{equation} \overline{f}(y)=\frac{\sum_{s=1}^{n/2}(s-1)p_s}{\sum_{s=1}^{n/2}p_s} \end{equation} Then, the global ``effective'' mean frequency ($\overline{f}$), characterising the oscillating ripple pattern in the whole image, is given by the mean of the $\overline{f}(y)$'s of each line $y$ along all horizontal scan lines, and the associated error $\delta\overline{f}$ is the standard deviation of the $\overline{f}(y)$'s. From the mean frequency $\overline{f}$ and the window length $L$ we obtain the ``effective'' period $\tau$ of the surface structures, \begin{equation} \tau=\frac{L}{\overline{f}}. \end{equation} Such an effective period $\tau$ is expected to give a quantitative estimate of the mean width of a single ripple observed in the AFM images. The uncertainty $\delta\tau$ on the effective period is calculated accordingly, from $\delta\overline{f}$. The regularity of the patterns and the {\it bona fide} characterisation of the morphograms by means of an effective Fourier component is further confirmed by observing that the relative error $\delta\tau/\tau$ on the effective period stays below 20\% for each image. Furthermore, by filtering only those $\lesssim 10$ harmonics (out of a total of $256$ or $512$) that lie within two standard deviations around the effective mean period (or, equivalently, in a frequency window $\overline{f} \pm 2 \delta\overline{f}$) one can reconstruct, via Fourier transform of the PSD, a theoretical periodic image that displays an overlap with the original experimental image of at least 50\%, for all the reported samples [see Fig. \ref{image2}(d)]. The effective period $\tau$ for all the $20$ different samples is about $(53 \pm 12) {\rm nm}$. \begin{table}[t!] \begin{tabular}{llllll} \hline \hline & Sample & Thickness& RMS (nm) & $\tau$ (nm)\\ \hline \hline 1 & Mn 351 & 5 nm, 2\% & 0.38 $\pm$ 0.04 & \quad 42.3 $\pm$ 4.2\\ 2 & Mn 352 & 5 nm, 6\% & 0.35 $\pm$ 0.07 & \quad49.7 $\pm$ 12.7 \\ 3 & Mn 556 & 5 nm, 6\% & 0.38 $\pm$ 0.05 & \quad 45.14 $\pm$ 11.5 \\ 4 & Mn 536 & 5 nm, 6\% & 0.31 $\pm$ 0.04 & \quad45.5 $\pm$ 11.3\\ 5 & Mn 396 & 7 nm, 6\% & 0.48 $\pm$ 0.06 & \quad 47.7 $\pm$ 8.2\\ 6 & Mn 381 & 7 nm, 6\% & 0.38 $\pm$ 0.04 & \quad 41.9 $\pm$ 6.4 \\ 7 & Mn 555 & 7 nm, 6\% & 0.3 $\pm$ 0.1 & \quad 47.7 $\pm$ 6.3\\ 8 & Mn 535 & 7 nm, 6\% & 0.34 $\pm$ 0.04 & \quad 45.4 $\pm$ 10.3\\ 9 & Mn 394 & 7 nm, 6\% & 0.35 $\pm$ 0.05 & \quad 59.8 $\pm$ 13.5\\ \hline 10 & Mn 437 & 25 nm, 12\% & 0.36 $\pm$ 0.06 & \quad 51.5 $\pm$ 8.6\\ 11 & Mn 438 & 25 nm, 12\% & 0.37 $\pm$ 0.07 & \quad 61.7 $\pm$ 11.4\\ 12 & Mn 439 & 25 nm, 12\% & 0.41 $\pm$ 0.05 & \quad 44.4 $\pm$ 5.1\\ \hline 13 & Mn 467 & 25 nm 6\% & 0.29 $\pm$ 0.07 & \quad 57.4 $\pm$ 8.6\\ 14 & Mn 554 & 25 nm, 6\% & 0.35 $\pm$ 0.04 & \quad 56.0 $\pm$ 16.9\\ 15 & Mn 499 & 25 nm, 6\% & 0.31 $\pm$ 0.05 & \quad 47.8 $\pm$ 11.8\\ 16 & Mn 330 & 25 nm, 6 \% & 0.31 $\pm$ 0.09 & \quad 65.9 $\pm$ 11.3\\ 17 & Mn 490 & 25 nm, 6\% & 0.28 $\pm$ 0.03 & \quad 52.5 $\pm$ 0.03\\ \hline 18 & Mn 491 & 25nm, 6\% P 6\% Mn & 0.28 $\pm$ 0.04 & \quad 51.5 $\pm$ 8.2\\ 19 & Mn 492 & 25nm, 3\% P 6\% Mn & 0.38 $\pm$ 0.08 & \quad 54.2 $\pm$ 12.7\\ 20 & Mn 498 & 25nm, 9\% P 6\% Mn & 0.37 $\pm$ 0.08 & \quad 52.61 $\pm$ 10.4\\ \hline \hline \end{tabular} \caption{Summary of the properties of (Ga,Mn)As samples analysed with the AFM. Notation: RMS is the global roughness; $\tau$ is the ``effective'' period of the ripples.}\label{table1} \end{table} In table \ref{table1} we summarise the main properties of the measured samples: thickness, \% Mn, effective period, and global RMS roughness. \begin{table}[t!] \begin{tabular}{lll} \hline \hline Sample Mn 381 & Scan area ($\mu{\rm m}^2$)& $\tau$ ({\rm nm})\\ \hline\hline 1 & 1x1 & 37.44 \\ 2 & 5x5 & 45.16 \\ 3 & 1x1 & 42.77 \\ 4 & 1x1 & 36.6 \\ 5 & 1x1 & 39.2 \\ 6 & 5x5 & 35 \\ 7 & 10x10 & 44.86 \\ 8 & 2x2 & 40.5\\ \hline \hline \end{tabular} \caption{Effective period $\tau$ calculated from the Fourier analysis on different areas of the sample Mn381, showing consistent values of the periodicity of the ripples.}\label{table2} \end{table} To confirm that the observed structures extend to the whole surface area of the grown (Ga,Mn)As compounds, we have performed additional measurements on one chosen sample, Mn381, selected for the particularly neat quality and visibility of its periodic surface ripples (see Fig.~\ref{image2}). We have acquired many AFM images on different areas of the sample with different size. The effective period $\tau$ for several $1 \times 1 \mu{\rm m}^2$, $2\times 2 \mu{\rm m}^2$, $5 \times 5 \mu{\rm m}^2$ and $10 \times 10 \mu{\rm m}^2$ AFM images is found to be stable at about $(40 \pm 5) {\rm nm}$. See Table \ref{table2} for a summary of the thickness and the Fourier periodicity figures of the ripples for different AFM images on this sample. \section{Conclusions} In conclusion we have studied the surface properties of (Ga,Mn)As compound with the AFM. From the analysis of different samples we can estimate a roughness of about $(0.38 \pm 0.10) {\rm nm}$. All of the measured samples show the formation of self-organised ripples along $[1\overline{1}0]$ crystallographic direction with an effective period of about $(53 \pm 12) {\rm nm}$, as estimated from a Fourier analysis. Completely analogous features (qualitatively and quantitatively) have been revealed for (Ga,Mn)(As,P) samples with varying P concentration. While it is known that epitaxially grown (Ga,Mn)(As,P) samples show magnetic properties comparable with P-free (Ga,Mn)As, the P layer does in general alter the strain state from compressive to tensile, inducing a modification of the magnetic anisotropy.\cite{prep2,muu} Nonetheless, we have noticed no significant difference in the structural surface properties, compared with the samples without P. A more extended case study with more finely tuned P concentration is certainly needed to support this conclusion, and will be the subject of further investigation. The presence of anisotropic surface structures in (Ga,Mn)As has been recently confirmed by grazing incidence X-ray diffraction measurements\cite{APL}, resulting in an estimate of the periodicity which is consistent with the findings reported in this work. An interesting future direction will be to investigate the mechanism at the heart of the formation of the observed ripples, in particular the role of lattice strain, and the correlation with the uniaxial magnetic anisotropy of (Ga,Mn)As. This is expected to shed new light on the magnetic properties of this material and in general of the family of (III,Mn)V ferromagnetic semiconductors. Developing a deeper understanding of how the ferromagnetic interactions between the local Mn moments, mediated by the itinerant holes, give rise to the observed magnetic properties, is essential in order to unleash the potential of (Ga,Mn)As to realise manyfold applications for the spintronics industry. \acknowledgments{Funding from the EU (Grant Nos. NAMASTE-214499 and SemiSpinNano-237375) is acknowledged.}
\section{Introduction} Many important physical phenomena are governed by hyperbolic systems of conservation laws. In one space dimension the standard conservation law has the form \begin{equation} \label{conslaw} q_t + f(q)_x = 0, \end{equation} where the components of $q \in \mathbb{R}^m$ are conserved quantities and the components of $f:\mathbb{R}^m \times \mathbb{R}^m \rightarrow \mathbb{R}^m$ are the corresponding fluxes. Very many numerical methods have been developed for the solution of \eqref{conslaw}; some of the most successful are the high resolution Godunov-type methods based on the use of Riemann solvers and nonlinear limiters. These and other methods are generally based on {\em flux-differencing} and make explicit use of the flux function $f$. Herein we also consider systems with spatially varying flux: \begin{equation} \label{varflux} \kappa(x)q_t + f(q,x)_x = \psi(q,x), \end{equation} and spatially varying linear systems not in conservation form: \begin{equation} \label{varlin} \kappa(x)q_t + A(x)q_x = \psi(q,x), \end{equation} as well as their two-dimensional extensions. Wave-propagation methods of up to second-order accuracy have been developed for such systems in, e.g. \cite{leveque1997,leveque2002}. These methods are based on wave-propagation Riemann solvers, which compute {\em fluctuations}, (i.e. traveling discontinuities) rather than fluxes and thus can be applied to \eqref{varflux} or \eqref{varlin} just as easily as to the conservation law \eqref{conslaw}. Second-order methods may often be the best choice in terms of a balance between computational cost and desired resolution, especially for problems with solutions dominated by shocks or other discontinuities with relatively simple structures between these discontinuities. For problems containing complicated smooth solution structures, where the accurate resolution of small scales is required (e.g. simulation of compressible turbulence, computational aeroacoustics, computational electromagnetism, turbulent combustion etc.), schemes with higher order accuracy are desirable. The purpose of this work is to present a numerical method that combines the advantages of wave-propagation solvers with high order of accuracy. The basic discretization approach was presented already in \cite{ketcheson2006}; here, we give a more detailed presentation and demonstrate the wide applicability of the method. The new method combines the notions of wave propagation (\cite{leveque1997,levequefvbook}) and the method of lines, and can in principle be extended to arbitrarily high order accuracy by the use of high order accurate spatial reconstructions and high order accurate ordinary differential equation (ODE) solvers. The implementation presented here is based on the fifth-order accurate weighted essentially non-oscillatory (WENO) reconstruction and a fourth-order accurate strong-stability-preserving Runge-Kutta (RK) scheme. We restrict our attention to problems in one or two dimensions, although the method may be extended in a straightforward manner to higher dimensions. Although the method described here can be applied to classical hyperbolic systems \eqref{conslaw}, in that case it is equivalent to a standard finite volume WENO flux-differencing scheme, as long as component-wise reconstruction and a conservative wave propagation Riemann solver (such as a Roe or HLL solver) are used. For this reason, we focus on problems in non-conservative form or with explicit spatial dependence in the flux or source terms, which can be challenging for traditional discretizations. Another approach to high order discretization of hyperbolic PDEs, referred to as the ADER method, has been developed by Titarev \& Toro \cite{titarev2002} and subsequent authors. That approach uses the Cauchy-Kovalewski procedure and has the advantage of leading to one-step time discretization. The method of lines approach used in the present work seems more straightforward and allows manipulation of the method's properties by the use of different time integrators, but requires the evaluation of multiple stages per time step. A similar class of conservative, well-balanced, and high-order accurate methods has been developed by Castro, Gallardo, Pares, and their coauthors; see, e.g. \cite{castro2008}. Those methods also use WENO reconstruction and Runge-Kutta time stepping in conjunction with Riemann solvers, and lead to a discretization with a form similar to that presented here and in \cite{ketcheson2006}. Those methods have recently been combined with the ADER approach; see \cite{dumbser2009}. The present method differs in the approach to reconstruction and the kind of Riemann solvers used. These differences result in some useful features: our method also handles systems \eqref{varflux}-\eqref{varlin} with capacity function $\kappa$ and it can make use of $f$-wave Riemann solvers \cite{bale2002} as well as wave-slope reconstruction and achieve high order convergence even for some problems with discontinuous coefficients. Finally, the method can immediately be applied to very many interesting problems because the implementation is based on Clawpack Riemann solvers, which are available for a great variety of hyperbolic systems. A potential drawback of the current implementation of our method is that, for well-balancing, it relies on discretizing the source term such that its effect is collocated at cell interfaces only. The methods described in this paper are implemented in the software package SharpClaw, which is freely available on the web at \url{http://www.clawpack.org}. SharpClaw employs the same interface that is used in Clawpack \cite{levequefvbook} for problem specification and setup, as well as for the necessary Riemann solvers. This makes it simple to apply SharpClaw to a problem that has been set up in Clawpack. These methods have also been incorporated into PyClaw\cite{PyClaw-2011-SISC}, which allows them to be run in parallel on large supercomputers. The paper is organized as follows. In \Sec{god-waveprop}, we present Godunov's method for linear hyperbolic PDEs in wave propagation form \cite{levequefvbook}. This method is extended to high order in \Sec{ho-waveprop} by introducing a high-order reconstruction based on cell averages. Generalization to spatially varying nonconservative linear systems is presented in \ref{sect:svnonconshypsys}. Further extensions and details of the method are presented in the remainder of Section 2. Numerical examples, including application to acoustics, elasticity, and shallow water waves,are presented in Section \ref{sect:applications}. \section{Semi-discrete wave-propagation} The wave-propagation algorithm was first introduced by LeVeque \cite{leveque1997} in 1997 in the framework of high resolution finite volume methods for solving hyperbolic systems of equations. The scheme is conservative, second-order accurate in smooth regions, and captures shocks without introducing spurious oscillations. In this section, we extend the wave-propagation algorithm to high order accuracy through use of high order reconstruction and time marching. For simplicity, we focus on the one dimensional scheme and then briefly describe the extension to two dimensions. \subsection{Riemann Problems and Notation} The notation for Riemann solutions used in this paper comes primarily from \cite{levequefvbook}, and is motivated by consideration of the linear hyperbolic system \begin{equation} \label{eq:linhypsys} q_t + A q_x = 0. \end{equation} Here $q\in\mathbb{R}^m$ and $A\in\mathbb{R}^{m\times m}$. System \eqref{eq:linhypsys} is said to be hyperbolic if $A$ is diagonalizable with real eigenvalues; we will henceforth assume this to be the case. Let $s^p$ and $r^p$ for $1\le p\le m$ denote the eigenvalues and the corresponding right eigenvectors of $A$ with the eigenvalues ordered so that $s^1 \leq s^2 \leq \ldots \leq s^m$. Consider the {\em Riemann problem} consisting of \eqref{eq:linhypsys} together with initial data \begin{align} \label{eq:riemann-data} q(x,0)=\left\{ \begin{array}{cc} q_l & x<0 \\ q_r & x>0 \end{array}\right. . \end{align} The solution for $t>0$ is piecewise constant with $m$ discontinuities, the $p$th one proportional to $r^p$ and moving at speed $s^p$. They can be obtained by decomposing the difference $q_r-q_l$ in terms of the eigenvectors $r^p$: \begin{align} \label{eq:wavedef} q_r-q_l=\sum_p \alpha^p r^p = \sum_p {\cal W}^p. \end{align} We refer to the vectors ${\cal W}^p$ as waves. Each wave is a jump discontinuity along the ray $x=s^p t$. An example solution is pictured in Figure \ref{fig:waveprop} for $m=3$. For brevity, we will sometimes refer to the Riemann problem with initial left state $q_l$ and initial right state $q_r$ as {\em the Riemann problem with initial states $(q_l,q_r)$.} \begin{figure} \centering \includegraphics[width=4in]{figures/riemann.eps} \caption{The wave propagation solution of the Riemann problem. The horizontal axis corresponds to space and the vertical axis to time.\label{fig:waveprop}} \end{figure} In a finite volume method, it is useful to define notation for the net effect of all left- or right-going waves: \begin{subequations}\label{eq:fluctdef} \begin{eqnarray} \Aop^-\Dq & \equiv & \sum_{p=1}^m \left(s^p\right)^- {\cal W}^p\\ \Aop^+\Dq & \equiv & \sum_{p=1}^m \left(s^p\right)^+ {\cal W}^p. \end{eqnarray} \end{subequations} Here and throughout, $(x)^\pm$ denotes the positive or negative part of $x$: $$(x)^-=\min(x,0) \quad \quad (x)^+=\max(x,0).$$ The symbols $\Aop^\pm\Delta q$, referred to as {\em fluctuations}, should be interpreted as single entities that represent the net effect of all waves travelling to the right or left. The notation is motivated by the constant coefficient linear system \eqref{eq:linhypsys}, in which case $\Aop^\pm\Delta q = A^{\pm}(q_r-q_l)$, where $A^-$ (respectively $A^+$) is the matrix obtained by setting all postive (respectively negative) eigenvalues of $A$ to zero. See \cite{leveque1997} or \cite{levequefvbook} for more details. The notation for waves and fluctuations defined in \eqref{eq:wavedef} and \eqref{eq:fluctdef} can also be used to describe numerical solutions of Riemann problems for nonlinear systems if the numerical solver approximates the solution by a series of propagating jump discontinuities, which is very often the case. Because the approximate Riemann solution for a nonlinear system depends not only on the difference $q_r-q_l$ but on the values of the states, we will sometimes employ for clarity the notation ${\cal W}^p(q_l,q_r)$ to denote the $p$th wave in the solution of the Riemann problem with initial states $(q_l,q_r)$. \begin{changebar} In this case the vectors $r^p$ used for the decomposition \eqn{eq:wavedef} are typically eigenvectors of an averaged Jacobian matrix $\bar A$, and the $s^p$ are the corresponding wave speeds. Then $\Aop^\pm\Delta q = \bar A^{\pm}(q_r-q_l)$. But there are other possible ways to define the $r^p$, $s^p$ and hence the fluctuations. For example, using the HLL approximate Riemann solver \cite{ha-la-vl} would always use only two waves regardless of the size of the system. Or, for the case of spatially varying flux functions, the left-going waves may be defined by eigenvectors of the Jacobian $f'(q_l)$ while the right-going waves are defined by eigenvectors of the Jacobian $f'(q_r)$. One could combine these to create a matrix $\bar A$ having this set of eigenvectors and the corresponding eigenvalues, but this is not required for implementation of the method. For these reasons we use the more general notation $\Aop^\pm\Delta q$ for the fluctuations defined by \eqref{eq:fluctdef}. \end{changebar} \subsection{First-order Godunov's method\label{sect:god-waveprop}} Consider the constant-coefficient linear system in one dimension \eqref{eq:linhypsys}. Taking a finite volume approach, we define the cell averages (i.e. the solution variables) \begin{align*} Q_i(t) = \frac{1}{\Delta x}\int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} q(x,t) dx, \end{align*} where the index $i$ and the quantity $\Delta x$ denote the cell index and the cell size, respectively. To solve the linear system \eqref{eq:linhypsys}, we initially approximate the solution $q(x,t)$ by these cell averages; that is, at $t=t_0$ we define the piecewise-constant function \begin{equation} \label{eq:pwc} \tilde{q}(x,t_0) = Q_i \mbox{ for } x\in (x_{i-\frac{1}{2}},x_{i+\frac{1}{2}}). \end{equation} The linear system \eqref{eq:linhypsys} with initial data $\tilde{q}$ consists locally of a series of Riemann problems, one at each interface $x_{i-\frac{1}{2}}$. The Riemann problem at $x_{i-\frac{1}{2}}$ consists of (\ref{eq:linhypsys}) with the piecewise constant initial data \begin{align*} q(x,0)=\left\{\begin{array}{ll} Q_{i-1} & x<x_{i-\frac{1}{2}} \\ Q_i & x>x_{i-\frac{1}{2}} \end{array}\right. . \end{align*} As discussed above, the solution of the Riemann problem is expressed as a set of waves obtained by decomposing the jump in $Q$ in terms of the eigenvectors of $A$: \begin{equation} \label{eq:waveDecompFV} Q_i-Q_{i-1}=\sum_p \alpha^p_{i-\frac{1}{2}} r^p_{i-\frac{1}{2}} = \sum_p {\cal W}^p(Q_{i-1},Q_i). \end{equation} Let $\tilde{q}(x,t_0+\Delta t)$ denote the exact evolution of $\tilde{q}$ after a time increment $\Delta t$. If we take $\Delta t$ small enough that the waves from adjacent interfaces do not pass through more than one cell, then we can integrate \eqref{eq:linhypsys} over $[x_{i-\frac{1}{2}},x_{{i+\frac{1}{2}}}] \times [0,\Delta t]$ and divide by $\Delta x$ to obtain \begin{align} \label{eq:1int} Q_i(t_0+\Delta t) - Q_i(t_0) & = -\frac{1}{\Delta x} \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A \, \tilde{q}_x(x,t_0+\Delta t) dx. \end{align} Here $\tilde{q}_x$ should be understood in the sense of distributions. \begin{figure} \centering \includegraphics[width=4in]{figures/fluctuations.eps} \caption{Time evolution of the reconstructed solution $\tilde{q}$ in cell $i$.\label{fig:qt-evol}} \end{figure} We can split the integral above into three parts, representing the Riemann fans from the two interfaces, and the remaining piece: \begin{align} \label{eq:3ints} \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A \tilde{q}_x dx & = \int_{x_{i-\frac{1}{2}}}^{x_{i-\frac{1}{2}}+s^R \Delta t} A \tilde{q}_x dx + \int_{x_{i+\frac{1}{2}}+s^L \Delta t}^{x_{i+\frac{1}{2}}} A \tilde{q}_x dx + \int_{x_{i-\frac{1}{2}}+s^R \Delta t}^{x_{i+\frac{1}{2}}+s^L \Delta t} A \tilde{q}_x dx. \end{align} The relevant regions are depicted in Figure \ref{fig:qt-evol}. Here we have defined $s^L=\min(s^1_{i+\frac{1}{2}},0)$ and $s^R=\max(s^m_{i-\frac{1}{2}},0)$. The third integral in \eqref{eq:3ints} vanishes because $\tilde{q}(x,\Delta t)$ is constant outside the Riemann fans, by the definition \eqref{eq:pwc}. Hence \eqref{eq:3ints} reduces to \begin{subequations} \label{eq:3intsb} \begin{align} \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A \tilde{q}_x dx & = \Delta t \sum_{p=1}^m \left(s^p_{i-\frac{1}{2}}\right)^+ {\cal W}^p_{i-\frac{1}{2}} \label{eq:3ints-waveDec} +\Delta t \sum_{p=1}^m \left(s^p_{i+\frac{1}{2}}\right)^- {\cal W}^p_{i+\frac{1}{2}} \\ & = \Delta t\left({\cal A}^+\Delta Q_{i-\frac{1}{2}}+{\cal A}^-\Delta Q_{i+\frac{1}{2}}\right),\label{eq:3ints-fluct} \end{align} \end{subequations} where the {\em fluctuations} ${\cal A}^-\Delta Q_{i+\frac{1}{2}}$ and ${\cal A}^+\Delta Q_{i-\frac{1}{2}}$ are defined by \begin{subequations}\label{pfluct} \begin{eqnarray} {\cal A}^-\Delta Q_{i+\frac{1}{2}} & \equiv & \sum_{p=1}^m \left(s^p_{i+\frac{1}{2}}\right)^- {\cal W}^p_{i+\frac{1}{2}}, \qquad {\cal W}^p_{i+\frac{1}{2}} \equiv {\cal W}^p(Q_{i},Q_{i+1})\\ {\cal A}^+\Delta Q_{i-\frac{1}{2}} & \equiv & \sum_{p=1}^m \left(s^p_{i-\frac{1}{2}}\right)^+ {\cal W}^p_{i-\frac{1}{2}}, \qquad {\cal W}^p_{i-\frac{1}{2}} \equiv {\cal W}^p(Q_{i-1},Q_i). \label{mfluct} \end{eqnarray} \end{subequations} Note again that the fluctuations ${\cal A}^+\Delta Q_{i-\frac{1}{2}}$ and ${\cal A}^-\Delta Q_{i+\frac{1}{2}}$ are motivated by the idea of a matrix-vector product but should be interpreted as single entities that represent the net effect of all waves travelling to the right or left. The upper-case $Q$ in the fluctuations is meant to emphasize that they are based on differences of cell averages. For instance, the fluctuation ${\cal A}^+\Delta Q_{i-\frac{1}{2}}$ corresponds to the effect of right-going waves from the Riemann problem with initial states $(Q_{i-1},Q_i)$. Substituting \eqref{eq:3ints-fluct} into \eqref{eq:1int}, we obtain the scheme $$ Q_i^{n+1}-Q_i^n = -\frac{\Delta t}{\Delta x} \left({\cal A}^+\Delta Q_{i-\frac{1}{2}} + {\cal A}^-\Delta Q_{i+\frac{1}{2}} \right). $$ Dividing by $\Delta t$ and taking the limit as $\Delta t$ approaches zero, we obtain the semi-discrete wave-propagation form of the (first-order) Godunov's scheme \begin{equation} \label{sdform_1} \frac{\partial Q_i}{\partial t} = -\frac{1}{\Delta x} \left({\cal A}^+\Delta Q_{i-\frac{1}{2}} + {\cal A}^-\Delta Q_{i+\frac{1}{2}} \right). \end{equation} Equation \eqref{sdform_1} constitutes a linear system of ordinary differential equations (ODEs) that may be integrated, for instance, with a Runge-Kutta method. It is clear from the derivation that this scheme reduces to the corresponding flux-differencing scheme when applied to systems written in conservation form, e.g. system (\ref{varflux}). \subsection{Extension to higher order\label{sect:ho-waveprop}} The method of the previous section is only first-order accurate in space. In order to improve the spatial accuracy, we replace the piecewise-constant approximation \eqref{eq:pwc} by a piecewise-polynomial approximation that is accurate to order $p$ in regions where the solution is smooth: \begin{align} \label{eq:pwp} \tilde{q}(x,t_0) & = \tilde{q}_i(x) \mbox{ for } x\in (x_{i-\frac{1}{2}},x_{i+\frac{1}{2}}), \end{align} where $$ \tilde{q}_i(x)=q(x,t_0)+{\cal O}(\Delta x^{p+1}). $$ Integration of $A\tilde{q}_x$ over $[x_{i-\frac{1}{2}},x_{{i+\frac{1}{2}}}]$ again yields \eqref{eq:3ints}, but now the third integral is non-zero in general, since $\tilde{q}$ is not constant outside the Riemann fans. Define \begin{changebar} \begin{align} q^R_{i-\frac{1}{2}} \equiv \lim_{x \to x_{i-\frac{1}{2}}^+} \tilde{q}_i(x) \qquad \qquad q^L_{i+\frac{1}{2}} \equiv \lim_{x \to x_{i+\frac{1}{2}}^-} \tilde{q}_i(x), \end{align} \end{changebar} where superscripts $L$ and $R$ refer respectively to the left and the right state of the interface considered. Then in place of \eqref{eq:3intsb}, we now obtain (neglecting terms of order ${\cal O}(\Delta t^2)$) \begin{subequations} \label{eq:3intsc} \begin{align} \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A \tilde{q}_x dx & \approx \Delta t \sum_{p=1}^m \left(s^p_{i-\frac{1}{2}}\right)^+ {\cal W}^p_{i-\frac{1}{2}} +\Delta t \sum_{p=1}^m \left(s^p_{i+\frac{1}{2}}\right)^- {\cal W}^p_{i+\frac{1}{2}}+ \int_{x_{i-\frac{1}{2}}+s^R\Delta t}^{x_{i+\frac{1}{2}}+s^L\Delta t} A \tilde{q}_x dx \label{eq:3intsc-waveDec}\\ & = \Delta t\left(\Aop^+\Delta {q}_{i-\frac{1}{2}}+\Aop^-\Delta {q}_{i+\frac{1}{2}}\right) + A(q^L_{i+\frac{1}{2}}-q^R_{i-\frac{1}{2}}).\label{eq:3intsc-fluct} \end{align} \end{subequations} The resulting fully-discrete scheme is thus $$ Q_i^{n+1}-Q_i^n = -\frac{\Delta t}{\Delta x} \left(\Aop^+\Delta {q}_{i-\frac{1}{2}}+ \Aop^-\Delta {q}_{i+\frac{1}{2}} + A(q^L_{i+\frac{1}{2}}-q^R_{i-\frac{1}{2}})\right). $$ We use the notation $\Aop^\pm\Delta {q}$ instead of ${\cal A}^\pm\Delta Q$ because the states in the Riemann problems are not the cell averages, but rather the reconstructed interface values. In other words, the fluctuations at $x_{i-\frac{1}{2}}$ are defined by \begin{align*} \Aop^\pm\Delta {q}_{i-\frac{1}{2}} & = \sum_{p=1}^m \left(s^p(q^L_{i-\frac{1}{2}},q^R_{i-\frac{1}{2}})\right)^\pm {\cal W}^p(q^L_{i-\frac{1}{2}},q^R_{i-\frac{1}{2}}). \end{align*} For instance, the fluctuation $\Aop^+\Delta {q}_{i-\frac{1}{2}}$ corresponds to the effect of right-going waves from the Riemann problem with initial states $(q^L_{i-\frac{1}{2}},q^R_{i-\frac{1}{2}})$. Moreover, we can view the term $A(q^L_{i+\frac{1}{2}}-q^R_{i-\frac{1}{2}})$ as the sum of both the left- and right-going fluctuations resulting from a Riemann problem with initial states $(q^R_{i-\frac{1}{2}},q^L_{i+\frac{1}{2}})$. It is natural to denote this term, which we refer to as a {\em total fluctuation}, by $\Aop\Delta q_i$: \begin{align*} \Aop\Delta q_i & = \sum_{p=1}^m \left(s^p(q^R_{i-\frac{1}{2}},q^L_{i+\frac{1}{2}})\right)^\pm {\cal W}^p(q^R_{i-\frac{1}{2}},q^L_{i+\frac{1}{2}}). \end{align*} Dividing by $\Delta t$ and taking the limit as $\Delta t$ approaches zero, we obtain the semi-discrete scheme \begin{equation} \label{sdform_1a} \frac{\partial Q_i}{\partial t} = -\frac{1}{\Delta x} \left(\Aop^+\Delta {q}_{i-\frac{1}{2}} + \Aop^-\Delta {q}_{i+\frac{1}{2}} + \Aop\Delta q_i\right). \end{equation} \subsection{Spatially varying linear systems\label{sect:svnonconshypsys}} Next we generalize the method to solve one-dimensional spatially varying nonconservative linear systems: \begin{equation} \label{eq:nlhypsys} q_t + A(x) q_x = 0. \end{equation} We again assume that $A$ is a constant function of $x$ within each cell, so we can write $A(x)=A_i(q)$. In the special case that $A$ is the Jacobian matrix of some function $f$, \eqref{eq:nlhypsys} corresponds to the conservation law \eqref{conslaw}. Our method can be applied to the general system \eqref{eq:nlhypsys} as long as the physically meaningful solution to the Riemann problem can be approximated. \begin{changebar} This may be nontrivial for a nonconservative nonlinear problem, as discussed in many papers such as \cite{ca-lef:sw,ca-di-ch-etal:wb,co-le-no-pe:multdist,dm-lef-mu,lef-tza:weak}. We do not go into this here, but assume that our numerical approach is to be applied to a problem for which the user has a means of solving the Riemann problem in terms of discontinuities (waves) ${\cal W}^p_{i-\frac{1}{2}}$ propagating at speeds $s^p_{i-\frac{1}{2}}$, and hence can define fluctuations. This is the case for any linearized Riemann solver, and often for other approximate solvers. \end{changebar} Then the scheme is given by \begin{equation} \label{sdform_1b} \frac{\partial Q_i}{\partial t} = -\frac{1}{\Delta x} \left(\Aop^+\Delta {q}_{i-\frac{1}{2}} + \Aop^-\Delta {q}_{i+\frac{1}{2}} + \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A_i \, \tilde{q}_x dx\right). \end{equation} In general, the integral in \eqref{sdform_1b} must be evaluated by quadrature; however, for the conservative system \eqref{conslaw}, the integral can be evaluated exactly, and is given by \begin{equation} \label{fluxint} \int_{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} A_i \, \tilde{q}_x dx = f(q^L_{i+\frac{1}{2}})-f(q^R_{i-\frac{1}{2}}). \end{equation} If the fluctuations are computed using a Roe solver or some other conservative wave-propagation Riemann solver, then the flux difference appearing in \eqref{fluxint} is equal to the sum of fluctuations from a fictitious ``internal'' Riemann problem for the current cell $i$, just as in the linear case above: \begin{equation} \label{fluxint-1} f(q^L_{i+\frac{1}{2}})-f(q^R_{i-\frac{1}{2}}) = \Aop^+\Delta {q}_i + \Aop^-\Delta {q}_i = \Aop\Delta q_i. \end{equation} Specifically, the fluctuations $\Aop^\pm\Delta {q}_i$ are those resulting from the Riemann problem with initial states $(q^R_{i-\frac{1}{2}},q^L_{i+\frac{1}{2}})$. Then we can write \eqref{sdform_1b} also as \begin{equation} \label{sdform_1c} \frac{\partial Q_i}{\partial t} = -\frac{1}{\Delta x} \left(\Aop^-\Delta {q}_{i+\frac{1}{2}} + \Aop^+\Delta {q}_{i-\frac{1}{2}} + \Aop\Delta q_i\right). \end{equation} Note that, for the conservative system \eqref{conslaw}, if a Roe solver or an $f$-wave solver (see Section \ref{subsec:F-wave}) is used, then the fluctuations are equal to the flux differences \begin{eqnarray} \label{lfluct} \Aop^-\Delta {q}_{i-\frac{1}{2}} & = & \hat{f}_{i-\frac{1}{2}} - f(q^L_{i-\frac{1}{2}}) \\ \Aop^+\Delta {q}_{i-\frac{1}{2}} & = & f(q^R_{i-\frac{1}{2}}) - \hat{f}_{i-\frac{1}{2}}, \label{rfluct} \end{eqnarray} where $\hat{f}_{i-\frac{1}{2}}$ is the numerical flux at $x_{i-\frac{1}{2}}$. Thus \eqref{sdform_1c} is in that case equivalent to the traditional flux-differencing method \begin{equation} \dd{Q_i}{t} = -\frac{1}{\Delta x}\left(\hat{f}_{i+\frac{1}{2}}-\hat{f}_{i-\frac{1}{2}}\right). \end{equation} In particular, the scheme is conservative in this case. \subsection{Capacity-form differencing} In many applications the system of conservation laws takes the form \begin{equation} \label{eq:capacitysys1D} \kappa(x)q_t + f(q)_x = 0, \end{equation} in one space dimension, or \begin{equation} \label{eq:capacitysys2D} \kappa(x,y)q_t + f(q)_x + g(q)_y = 0, \end{equation} in two dimensions, where $\kappa$ is a given function of space and is usually indicated as \emph{capacity function} (see \cite{leveque1997}). Systems like (\ref{eq:capacitysys1D}) and (\ref{eq:capacitysys2D}) arise naturally in the derivation of a conservation law, where the flux of a quantity is naturally defined in terms of one variable $q$, whereas it is a different quantity $\kappa q$ that is conserved. For instance, for the flow in a porous media, $\kappa$ would be the porosity. Note that a capacity function can also appear in systems that are not in conservation form, e.g. the quasilinear system \begin{changebar} (\ref{varlin}). \end{changebar} Several approaches can be used to reduce such a system to a more familiar conservation law. One natural approach is the capacity-form differencing \cite{leveque1997}, \begin{equation} \label{capacity-form-diff} \frac{\partial Q_i}{\partial t} = -\frac{1}{\kappa_i \Delta x} \left(\Aop^+\Delta {q}_{i-\frac{1}{2}} + \Aop^-\Delta {q}_{i+\frac{1}{2}} + \Aop\Delta q_i\right), \end{equation} where $\kappa_i$ is the capacity of the $i$th cell. This is a simple extension of (\ref{sdform_1a}) or (\ref{sdform_1c}) which ensures that $\sum \kappa_i Q_i$ is conserved (except possibly at the boundaries) and yet allows the Riemann solution to be computed based on $q$ as in the case $\kappa = 1$. \subsection{$f$-wave Riemann solvers and well-balancing\label{subsec:F-wave}} For application to conservation laws, it is desirable to ensure that the wave-propagation discretization is conservative. This can easily be accomplished by using an $f$-wave Riemann solver \cite{bale2002}. Use of $f$-wave solvers is also useful for problems with spatially varying flux function, as well as problems involving balance laws near steady state. Further uses of the $f$-wave approach can be found in \cite{ahmad:aiaa08, ahmad-lindeman:fwave, khouider-majda:balanced, NormanNairSemazzi2010, sniv-pasko:fwave}. The idea of the $f$-wave splitting for (\ref{conslaw}) is to decompose the flux difference $f(q_r) - f(q_l)$ into waves rather than the $q$-difference used in (\ref{eq:waveDecompFV}), i.e. we decompose the flux difference as a linear combination of the right eigenvectors $r^p$ of some Jacobian: \begin{equation} \label{eqn:f-waveDef} f(q_r) - f(q_l) = \sum_p \beta^p r^p = \sum_p \mathcal{Z}^p(q_l,q_r). \end{equation} The fluctuations are then defined as \begin{align*} \Aop^-\Dq & \equiv \sum_{p:s^p <0} \mathcal{Z}^p(q_l,q_r) & \Aop^+\Dq & \equiv \sum_{p:s^p >0} \mathcal{Z}^p(q_l,q_r) & \end{align*} Note that the total fluctuation in cell $i$ is given simply by $$ \Aop\Delta q_i = f\left(q^L_{i+\frac{1}{2}}\right) - f\left(q^R_{i-\frac{1}{2}}\right). $$ An advantage of particular interest is the possibility to include source terms directly into the $f$-wave decomposition. In fact, for balance laws that include non-hyperbolic terms, $$ q_t + f(q)_x = \psi(q,x), $$ one can easily extend this algorithm by first discretizing the source term to obtain values $\Psi_{i-\frac{1}{2}}$ at the cell interfaces and then compute the waves $\mathcal{Z}^p_{i-\frac{1}{2}}$ by splitting \begin{equation} \label{f-wave-balance} f(q_r) - f(q_l) - \Delta x \Psi(q_l,q_r,x) = \sum_p \beta^p r^p = \sum_p \mathcal{Z}^p(q_l,q_r). \end{equation} Here $\Psi(q_l,q_r,x)$ is some suitable average of $\psi(q,x),$ between the neighboring states. In Bale et al. \cite{bale2002}, it has been shown that for the second-order FV wave-propagation scheme implemented in Clawpack, the $f$-wave approach is very useful for handling source terms, especially in cases where the solution is close to a steady state because it leads to a well-balanced scheme. However, for our high order wave-propagation scheme, application of the $f$-wave algorithm with component-wise or characteristic-wise reconstruction \cite{Qiu2002187} (which take no account of the source term) does not lead to a method that is well-balanced, even though the source term is accounted for in the Riemann solves which compute $\Aop^\pm\Delta {q}_{i-\frac{1}{2}}$. This is because with the aforementined WENO approaches the reconstructed solution within each cell is not constant (i.e. $q^L_{i+\frac{1}{2}} \neq q^R_{i-\frac{1}{2}}$) and $\Aop\Delta q_i \neq 0$. In this work, we consider an extension of the $f$-wave well-balancing technique that is compatible with our higher order scheme. The extension is useful for problems in which the source term vanishes over the interior of each cell (i.e., its effect is concentrated at cell interfaces). This is the case, for instance, when considering the shallow water equations with piecewise-constant bathymetry (see Section \ref{subsubsec:2Dperturb-SWEs}). This well-balancing technique uses the $f$-wave Riemann solver in combination with a $f$-wave-slope reconstruction (see Section \ref{subsect:f-wave-slope-recon}). With this approach, the contribution of the source term is directly included in the Riemann solve at the cell's interface, i.e. (\ref{f-wave-balance}). The reconstruction methods considered in this work are presented in the next section. \subsection{Reconstruction\label{subsec:waveSlopeReconstruction}} The reconstruction \eqref{eq:pwp} should be performed in a manner that yields high order accuracy but avoids spurious oscillations near discontinuities. For this purpose, we use weighted essentially non-oscillatory (WENO) reconstruction \cite{shu2009}. The spatial accuracy of the method will in general be equal to that of the reconstruction. In the present work we employ fifth-order WENO reconstruction. \subsubsection{Scalar WENO Reconstruction} WENO reconstruction formulas are typically written in terms of the divided differences $\Delta Q_{{i \pm \frac{1}{2}}}/\Delta x$. It is possible to rewrite them in terms of the difference ratios \begin{equation} \label{theta-scalar} \theta_{{i-\frac{1}{2}},j} = \frac{\Delta Q_{{i-\frac{1}{2}}+j}}{\Delta Q_{i-\frac{1}{2}}}. \end{equation} as long as $\Delta Q_{i-\frac{1}{2}}\ne0$. Then the reconstructed interface values in cell $i$ are given by \begin{subequations} \label{recon-scalar} \begin{align} q^R_{i-\frac{1}{2}} & = Q_i - \phi(\theta_{{i-\frac{1}{2}},2-k},\dots,\theta_{{i-\frac{1}{2}},k-1})\Delta Q_{i-\frac{1}{2}} \\ q^L_{i+\frac{1}{2}} & = Q_i + \phi(\theta_{{i+\frac{1}{2}},1-k},\dots,\theta_{{i+\frac{1}{2}},k-2})\Delta Q_{i-\frac{1}{2}}, \end{align} \end{subequations} where $\phi$ is a particular nonlinear function that we will not write out here. The usefulness of \eqref{recon-scalar} is that it allows WENO reconstruction to be applied to waves directly by redefining $\theta$, as we do below. In the case that $\Delta Q_{i-\frac{1}{2}}\approx0$ (to near machine precision), the value of $\phi$ is set to zero. For systems of equations, the simplest approach to reconstruction is component-wise reconstruction, which consists of applying the scalar reconstruction approach \eqref{theta-scalar}-\eqref{recon-scalar} to each element of $q$. A more sophisticated approach is characteristic-wise reconstruction, in which an eigendecomposition of $q$ is performed, followed by reconstruction of each eigencomponent. For problems with spatially-varying coefficients, even the characteristic-wise reconstruction may not be satisfying, since it involves comparing coefficients of eigenvectors whose direction in state space varies from one cell to the next. In Clawpack, an alternative kind of TVD limiting known as {\em wave limiting} has been implemented and shown to be effective for such problems. \subsubsection{Wave-slope reconstruction} In order to implement a wave-type WENO limiter, we first solve a Riemann problem at each interface $x_{i-\frac{1}{2}}$, using the adjacent cell average values $Q_{i-1},Q_i$ as left and right states. This results in a set of waves ${\cal W}^p_{i-\frac{1}{2}}$, which are used solely for the purpose of reconstruction. This reconstruction is performed by replacing \eqref{theta-scalar} by \begin{align} \label{theta-wave} \theta^p_{{i-\frac{1}{2}},j} & = \frac{{\cal W}^p_{i-\frac{1}{2}+j}\cdot{\cal W}^p_{i-\frac{1}{2}}}{{\cal W}^p_{i-\frac{1}{2}}\cdot{\cal W}^p_{i-\frac{1}{2}}} \end{align} and replacing \eqref{recon-scalar} by \begin{subequations} \label{recon-wave} \begin{align} q^L_{i-\frac{1}{2}} & = Q_{i-1} + \sum_p \phi(\theta^p_{{i+\frac{1}{2}},1-k},\dots,\theta^p_{{i+\frac{1}{2}},k-2}) {\cal W}^p_{i-\frac{1}{2}} \\ q^R_{i-\frac{1}{2}} & = Q_i \ \ \ - \sum_p \phi(\theta^p_{{i-\frac{1}{2}},2-k},\dots,\theta^p_{{i-\frac{1}{2}},k-1}) {\cal W}^p_{i-\frac{1}{2}}, \end{align} \end{subequations} This approach takes into account the smoothness of the $p$th characteristic component of the solution by using the information arising from the $k$-cell stencils. It is intended to be similar to that used in Clawpack \cite{levequefvbook} and can be conveniently implemented using the same Riemann solvers supplied with Clawpack. \subsubsection{$f$-wave-slope reconstruction}\label{subsect:f-wave-slope-recon} Wave-slope reconstruction can also be performed using an $f$-wave Riemann solver. This is useful for computing near-equilibrium solutions of balance laws. The procedure is identical to that above, except that the Riemann problem is solved with the $f$-wave Riemann solver at each interface $x_{i-\frac{1}{2}}$, using the adjacent cell average values $Q_{i-1},Q_i$ as left and right states. Since the resulting $f$-waves have the form of a $q$ increment multiplied by the wave speed \cite{bale2002}, the waves ${\cal Z}$ are normalized by the corresponding wave speed before being used for reconstruction: \begin{align} \label{wavediv} {\cal W}^p_{i-\frac{1}{2}} & = \mathcal{Z}^p_{i-\frac{1}{2}} / s^p_{i-\frac{1}{2}}. \end{align} In this paper we assumed that the original hyperbolic system has no zero eigenvalues ($s^p \neq 0$) or eigenvalues that change sign between grid cells (i.e. the resonant case). The reconstruction procedure \eqref{theta-wave}-\eqref{recon-wave} is then applied to the waves computed by \eqref{wavediv}. For systems with source terms that are at a steady state, assuming that \eqref{f-wave-balance} holds, results in ${\cal Z}^p=0$ for all waves. Therefore, the $f$-wave-slope reconstruction will yield to a constant reconstructed function in regions where source terms and hyperbolic terms are balanced. Then all fluctuations computed in the update step will vanish, so the steady state will be preserved exactly. \subsection{Time integration} The semi-discrete scheme can be integrated in time using any initial-value ODE solver. Herein we use the ten-stage fourth-order strong-stability-preserving Runge-Kutta scheme of \cite{ketcheson2008}. This method has a large stability region and a large SSP coefficient, allowing use of a relatively large CFL number in practical computations. In all numerical examples of the next section, a CFL number of 2.45 is used. To summarize, the full semi-discrete algorithm used in each Runge-Kutta stage is as follows. \begin{enumerate} \setcounter{enumi}{-1} \item {\em (only if using wave-slope reconstruction)} Solve the Riemann problem at each interface $x_{i-\frac{1}{2}}$ using the adjacent cell average values $Q_{i-1},Q_i$ as left and right states. \item Compute the reconstructed piecewise function $\tilde{q}$, and in particular the states $q^R_{i-\frac{1}{2}},q^L_{i+\frac{1}{2}}$ in each cell, using component-wise, characteristic-wise, or wave-slope reconstruction. \item At each interface $x_{i-\frac{1}{2}}$, compute the fluctuations $\Aop^+\Delta {q}_{i-\frac{1}{2}}$ and $\Aop^-\Delta {q}_{i-\frac{1}{2}}$ by solving the Riemann problem with initial states $(q^L_{i-\frac{1}{2}},q^R_{i-\frac{1}{2}})$. \item Over each cell, compute the integral $\int A\,\tilde{q}_x dx$. For conservative systems this is just the total fluctuation $\Aop\Delta q_i$. \item Compute $\partial Q/\partial t$ using the semi-discrete scheme \eqref{sdform_1b}. \end{enumerate} Note again that, for conservative systems, the quadrature in step 3 requires nothing more than evaluating and differencing the fluxes. \subsection{Extension to Two Dimensions} In this section, we extend the numerical wave propagation method to two dimensions using a simple dimension-by-dimension approach. The method is applicable to systems of the form \begin{equation} q_t + A(x,y)q_x + B(x,y)q_y = 0 \end{equation} on uniform Cartesian grids. The 2D analog of the semi-discrete scheme \eqref{sdform_1c} is \begin{align} \label{sdform_2D} \begin{split} \frac{\partial Q_{ij}}{\partial t} = -\frac{1}{\Delta x\Delta y} \left({\cal A}^-\Delta {q}_{{i+\frac{1}{2}},j} + {\cal A}^+\Delta {q}_{{i-\frac{1}{2}},j} + {\cal A}\Delta {q}_{i,j} \right. \\ \left. +{\cal B}^-\Delta {q}_{i,{j+\frac{1}{2}}} + {\cal B}^+\Delta {q}_{i,{j-\frac{1}{2}}} + {\cal B}\Delta {q}_{i,j} \right). \end{split} \end{align} For the method to be high order accurate, the fluctuation terms like ${\cal A}^-\Delta {q}_{{i+\frac{1}{2}},j}$ should involve integrals over cell edges, while the total fluctuation terms like ${\cal A}\Delta {q}_{i,j}$ should involve integrals over cell areas. This can be achieved by forming a genuinely multidimensional reconstruction of $q$ and using, e.g., Gauss quadrature. An implementation following this approach exists in the SharpClaw software. For nonlinear problems containing shocks, the genuinely multidimensional reconstruction has been found to be inefficient (at least for some simple test problems), as it typically yields only a small improvement in accuracy over the dimension-by-dimension scheme given below (which formally only second-order accurate), but has a much greater computational cost on the same mesh. Similar observations have been reported in Zhang et al. \cite{Zhang-2011-WENO}, where both approaches have been throughly tested and compared for linear and nonlinear systems. For problems with shocks, at least for the simple test problems presented, the two schemes give comparable resolution on the same meshes, despite their difference in formal order of accuracy. We now describe the dimension-by-dimension scheme for a single Runge-Kutta stage. We first reconstruct piecewise-polynomial functions $\tilde{q}_j(x)$ along each row of the grid and $\tilde{q}_i(y)$ along each column, by applying a 1D reconstruction procedure to each slice. We thus obtain reconstructed values \begin{subequations} \begin{align} \tilde{q}_j^R(x_{i-\frac{1}{2}}) & \approx q(x_{i-\frac{1}{2}},y_j) \\ \tilde{q}_j^L(x_{i+\frac{1}{2}}) & \approx q(x_{i+\frac{1}{2}},y_j) \\ \tilde{q}_i^R(y_{i-\frac{1}{2}}) & \approx q(x_i,y_{i-\frac{1}{2}}) \\ \tilde{q}_i^L(y_{i+\frac{1}{2}}) & \approx q(x_i,y_{i+\frac{1}{2}}) \end{align} \end{subequations} for each cell $i,j$. The fluctuation terms in \eqref{sdform_2D} are determined by solving Riemann problems between the appropriate reconstructed values; for instance ${\cal B}^-\Delta {q}_{i,{j+\frac{1}{2}}}$ is determined by solving a Riemann problem in the $y$-direction with initial states $(q^L_{i,{j+\frac{1}{2}}},q^R_{i,{j+\frac{1}{2}}}).$ In the case of conservative systems or piecewise-constant coefficients, the total fluctuation terms ${\cal A}\Delta {q}_{i,j}$ and ${\cal B}\Delta {q}_{i,j}$ can be similarly determined by summing the left- and right-going fluctuations of an appropriate Riemann problem. Thus, for instance, ${\cal B}\Delta {q}_{i,j}$ is determined by solving Riemann problem in the $y$-direction with initial states $(q^R_{i,{j-\frac{1}{2}}},q^L_{i,{j+\frac{1}{2}}}).$ \section{Numerical applications} \label{sect:applications} In this section we present results of numerical tests using the wave propagation methods just described. The examples included are chosen to emphasize the advantages of the wave propagation approach. We make some comparisons with the well-known second-order wave propagation code Clawpack \cite{clawpack45,levequefvbook}. \subsection{Acoustics} \label{sect:acoustics} In this section, the high-order wave propagation algorithm is applied to the 1D equations of linear acoustics in piecewise homogeneous materials: \begin{subequations} \label{eq:acoustics} \begin{eqnarray} p_t + K(x,y) (u_x + v_y) & = & 0 \\ u_t + \frac{1}{\rho(x,y)} p_x & = & 0 \\ v_t + \frac{1}{\rho(x,y)} p_y & = & 0. \end{eqnarray} \end{subequations} Here $p$ is the pressure and $u,v$ are the x- and y-velocities, respectively. The coefficients $\rho$ and $K$, which vary in space, are the density and bulk modulus of the medium. We will also refer to the sound speed $c=\sqrt{K/\rho}$. Notice that in general since $\rho$ varies in space, the last two equations above are not in conservation form. This test case demonstrates that the proposed approach is able to solve hyperbolic system of equations written in nonconservative form. Of course, this system can be written in conservative form as follows: \begin{subequations} \label{eq:acoustics_cons} \begin{eqnarray} \epsilon_t - (u_x + v_y) & = & 0 \\ \rho(x,y) u_t - (K(x,y)\epsilon)_x & = & 0 \\ \rho(x,y) v_t - (K(x,y)\epsilon)_y & = & 0, \end{eqnarray} \end{subequations} Where $\epsilon=-p/K$ is the strain. As we will see, the latter form may be advantageous in terms of the accuracy that can be obtained. The grid is chosen so that the material is homogeneous in each computational cell, and an exact Riemann solver is used at each interface; for details of this solver see e.g. \cite{fogarty1999}. \subsubsection{One-dimensional acoustics} We first consider one-dimensional acoustic waves in a piecewise-constant medium with a single interface. Namely, we solve \eqref{eq:acoustics} on the interval $x\in[-10,10]$ with $$ (\rho,c) = \left\{\begin{array}{cr} (\rho_l,c_l) & x<0 \\ (\rho_r,c_r) & x>0 \end{array}\right. $$ We measure the convergence rate of the solution in order to verify the order of accuracy for smooth solutions. The initial condition is a compact, six-times differentiable purely right-moving pulse: \begin{align*} p(x,0) & = \frac{((x-x_0)-a)^6((x-x_0)+a)^6}{a^{12}} \xi(x-x_0) \\ u(x,0) & = p(x,0)/Z(x) \end{align*} where $$ \xi(x-x_0)=\left\{ \begin{array}{cc} 0 & \mbox{for } |x-x_0|> a) \\ 1 & \mbox{for } |x-x_0| \le a. \end{array} \right. $$ with $x_0=-4$ and $a=1$, and $Z(x) = \sqrt{K(x)\rho(x)}$. This condition was chosen to be sufficiently smooth to demonstrate the design order of the scheme and to give a solution that is identically zero at the material interface at the initial and final times. Table \ref{tbl:homogeneous} shows $L_1$ errors and convergence rates for propagation in a homogeneous medium with $\rho_l=c_l=\rho_r=c_r=1$. Here we use componentwise reconstruction. Specifically, we compute \begin{equation} \label{eq:defL1Err} E_{L_1} = \Delta x\sum_i |Q_i - \bar{Q}_i| \end{equation} where $\bar{Q}_i$ is a highly accurate solution cell average computed by characteristics or by using a very fine grid. For the acoustics problems in this section, $\bar{Q}$ is computed using characteristics and adaptive Gauss quadrature. Table \ref{tbl:homogeneous} indicates that in each case, the order of convergence is approximately equal to the design order of the discretization. \begin{table}[th] \caption{Errors for homogeneous acoustics test \label{tbl:homogeneous}} \begin{center} \begin{tabular}{c|cc|cc} & \multicolumn{2}{c|}{SharpClaw}& \multicolumn{2}{c}{Clawpack}\\ \hline mx & Error & Order& Error & Order\\ \hline 200 & 3.60e-02 & & 4.10e-02 & \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 400 & 3.65e-03 & 3.30& 1.30e-02 & 1.66\\ 800 & 1.85e-04 & 4.31& 3.61e-03 & 1.85\\ \shadeRow1600 & 7.35e-06 & 4.65& 8.94e-04 & 2.01\\ \hline \end{tabular} \end{center} \end{table} To test the accuracy in the presence of discontinuous coefficients we take $$\rho_l=c_l=1 \quad \rho_r=4 \quad c_r=1/2,$$ with an impedance ratio of $Z_r/Z_l=2$. As was noted in \cite{bale2002}, this system can also be solved in the conservative form \eqref{eq:acoustics_cons} using the $f$-wave approach. We include results of this approach, where we have also performed characteristic-wise rather than component-wise reconstruction. Results are shown in Table \ref{tbl:interface 1}. In this case all schemes exhibit a convergence rate below the formal order, even though the initial and final solutions are smooth. To investigate this further, we repeat the same test with a wider pulse by taking $a=4$. Results are shown in Table \ref{tbl:interface 1 wide}. For the latter test, we observe a convergence rate of approximately two for SharpClaw, one for Clawpack, and five for SharpClaw using the $f$-wave approach and characteristic-wise reconstruction. The last convergence rate is remarkable, considering that the solution is not differentiable when it passes through the material interface. Further investigation of the accuracy of this approach for more complicated problems with discontinuous coefficients is ongoing. In tests not shown here, Clawpack achieves approximately second-order accuracy when used with an $f$-wave Riemann solver for this problem. \begin{table}[th] \caption{Errors for acoustics interface with narrow pulse \label{tbl:interface 1}} \begin{center} \begin{tabular}{c|cc|cc|cc} & \multicolumn{2}{c|}{SharpClaw}& \multicolumn{2}{c|}{SharpClaw $f$-wave}& \multicolumn{2}{c}{Clawpack}\\ \hline mx & Error & Order& Error & Order& Error & Order\\ \hline 200 & 2.10e-01 & & 9.50e-02 & & 1.98e-01 & \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 400 & 5.98e-02 & 1.81& 1.42e-02 & 2.74 & 7.26e-02 & 1.45\\ 800 & 1.25e-02 & 2.26& 1.42 e-03 & 3.32 & 2.21e-02 & 1.71\\ \shadeRow1600 & 1.17e-03 & 3.42& 1.20e-04 & 3.56 & 7.86e-03 & 1.49\\ \hline \end{tabular} \end{center} \end{table} \begin{table}[th] \caption{$L^1$ Errors for acoustics interface problem with wide pulse (a=4)\label{tbl:interface 1 wide}} \begin{center} \begin{tabular}{c|cc|cc|cc} & \multicolumn{2}{c|}{SharpClaw}& \multicolumn{2}{c|}{SharpClaw $f$-wave}& \multicolumn{2}{c}{Clawpack}\\ \hline mx & Error & Order& Error & Order & Error & Order\\ \hline 200 & 9.67e-03 & & 5.01e-03 & & 5.23e-02 & \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 400 & 2.01e-03 & 2.27& 4.63e-04 & 3.44 & 2.32e-02 & 1.17\\ 800 & 4.89e-04 & 2.04& 2.51e-05 & 4.36 & 1.09e-02 & 1.09\\ \shadeRow1600 & 1.22e-04 & 2.00& 6.49e-07 & 5.12 & 5.26e-03 & 1.05\\ \hline \end{tabular} \end{center} \end{table} \subsubsection{A Two-dimensional sonic crystal} In this section we model sound propagation in a sonic crystal. A sonic crystal is a periodic structure composed of materials with different sounds speeds and impedances. The periodic inhomogeneity can give rise to {\em bandgaps} -- frequency bands that are completely reflected by the crystal. This phenomenon is widely utilized in photonics, but its significance for acoustics has only recently been considered. Photonic crystals can be analyzed quite accurately using analytic techniques, since they are essentially infinite size structures relative to the wavelength of the waves of interest. In contrast, sonic crystals are typically only a few wavelengths in size, so that the effects of their finite size cannot be neglected. For more information on sonic crystals, see for instance the review paper \cite{miyashita2005}. We consider a square array of square rods in air with a plane wave disturbance incident parallel to one of the axes of symmetry. The array is infinitely wide but only eight periods deep. The lattice spacing is 10 cm and the rods have a cross-sectional side length of 4 cm, so that the filling fraction is $0.16$. This crystal is similar to one studied in \cite{sanchis2001}, and it is expected that sound waves in the 1200-1800 Hz range will experience severe attenuation in passing through it, while longer wavelengths will not be significantly attenuated. A numerical instability very similar to that observed in 1D simulations in \cite{fogarty1998,fogarty1999} was observed when the standard Clawpack method was applied to this problem. The fifth-order WENO method with characteristic-wise limiting showed no such instability. Figure \ref{sclongwave} shows a snapshot of the RMS pressure distribution in space for a plane wave with $k=15$ incident from the left. The RMS (root mean square) pressure is computed as follows: \begin{align} p_\textup{RMS}(x,y) = \sqrt{\frac{1}{T}\int_t^{t+T}p^2(x,y,t)dt}. \end{align} This wave has a frequency of about 800 Hz, well below the partial band gap. As expected, the wave passes through the crystal without significant attenuation. In Figure \ref{sclongwave_slice}, the pressure is plotted along a slice in the $x$-direction approximately midway between rows of rods. \begin{figure} \centering \includegraphics[width=4in]{figures/scxyl.eps} \caption{Pressure in the sonic crystal for a long wavelength plane wave incident from the left.\label{sclongwave}} \end{figure} \begin{figure} \centering \includegraphics[width=3in]{figures/sclongwave_slice.eps} \caption{Pressure in the sonic crystal for a long wavelength plane wave incident from the left.\label{sclongwave_slice}} \end{figure} Figure \ref{sc1} shows a snapshot of the RMS pressure distribution in space for an incident plane wave with with frequency 1600 Hz, inside the partial bandgap. Notice that the wave is almost entirely reflected, resulting in a standing wave in front of the crystal. Figure \ref{sc1_slice} shows the RMS pressure along a slice in the $x$-direction. \begin{figure} \centering \includegraphics[width=4in]{figures/scxy.eps} \caption{Snapshot of RMS pressure distribution in space in the sonic crystal for a plane wave incident from the left.\label{sc1}} \end{figure} \begin{figure} \centering \includegraphics[width=3in]{figures/sc1_slice.eps} \caption{Snapshot of RMS pressure distribution in space in the sonic crystal along a slice.\label{sc1_slice}} \end{figure} \subsection{Nonlinear Elasticity in a Spatially Varying Medium} \label{subsec:stego} In this section we consider a more difficult test involving nonlinear wave propagation and many material interfaces. This problem was considered previously in \cite{leveque2002} and studied extensively in \cite{leveque2003}. Solitary waves were observed to arise from the interaction of nonlinearity and an effective dispersion due to material interfaces in layered media. Elastic compression waves in one dimension are governed by the equations \begin{subequations} \label{nel_pde} \begin{align} \epsilon_t(x,t)-u_x(x,t) & = 0 \\ (\rho(x)u(x,t))_t - \sigma(\epsilon(x,t),x)_x & = 0. \end{align} \end{subequations} where $\epsilon$ is the strain, $u$ the velocity, $\rho$ the density, and $\sigma$ the stress. This is a conservation law of the form \eqref{conslaw}, with \begin{align*} q(x,t) & = \left(\begin{array}{c} \epsilon \\ \rho(x) u \end{array}\right) & f(q,x) & = \left(\begin{array}{c} -u \\ -\sigma(\epsilon,x) \end{array}\right). \end{align*} Note that the density and the stress-strain relationship vary in $x$. The Jacobian of the flux function is \begin{align*} f'(q) = \left(\begin{array}{cc} 0 & -1/\rho(x) \\ -\sigma_\epsilon(\epsilon,x) & 0 \end{array}\right). \end{align*} In the case of the linear stress-strain relation $\sigma(x)=K(x)\epsilon(x)$, \eqref{nel_pde} is equivalent to the one-dimensional form of the acoustics equations considered in the previous section. We consider the piecewise constant medium studied in \cite{leveque2002,leveque2003}: \begin{align} \label{eq:layered} (\rho(x),K(x)) & = \left\{\begin{array}{cl} (1,1) & \mbox{if } j<x<(j+\frac{1}{2}) \mbox{ for some integer } j \\ (4,4) & \mbox{otherwise}, \end{array}\right. \end{align} with exponential stress-strain relation \begin{equation} \label{eq:exp-stress} \sigma(\epsilon,x)=\exp(K(x)\epsilon)-1. \end{equation} The initial condition is uniformly zero, and the boundary condition at the left generates a half-cosine pulse. We solve this problem using the $f$-wave solver developed in \cite{leveque2002}. \Fig{stego_comp} shows a comparison of results using Clawpack and SharpClaw on this problem, with 24 cells per layer. The SharpClaw results are significantly more accurate. \begin{figure} \centerline{ \subfigure[Strain.\label{fig:stego_comp_strain}]{\includegraphics[width=3in]{figures/stego_comp_strain.eps}} \subfigure[Stress.\label{fig:stego_comp_stress}]{\includegraphics[width=3in]{figures/stego_comp_stress.eps}}} \caption[Comparison of Clawpack and SharpClaw solutions of Stegoton problem.] {Comparison of Clawpack (red circles) and SharpClaw (blue squares) solution of the stegoton problem using 24 cells per layer. For clarity, only every third solution point is plotted. The black line represents a very highly resolved solution.\label{fig:stego_comp}} \end{figure} Solutions of \eqref{nel_pde} are time-reversible in the absence of shocks. As discussed in \cite{PhDDKetcheson2009,ketcheson2010}, the effective dispersion induced by material inhomogeneities suppresses the formation of shocks, for small amplitude initial and boundary conditions, rendering the solution time-reversible for very long times. This provides a useful numerical test. We solve the stegoton problem numerically up to time $T$, then negate the velocity and continue solving to time $2T$. The solution at any time $2T-t_0$, with $t_0\le T$, should be exactly equal to the solution at $t_0$. We take $T=600$ and $t_0=60$. \Fig{stego_tr_sc} shows the solution obtained using SharpClaw on a grid with 24 cells per layer. The $t=1140$ solution (squares) is in excellent agreement with the $t=60$ solution (solid line). In fact, the maximum point-wise difference has magnitude less than $2\times10^{-2}$. Using a grid twice as fine, with 48 cells per layer, reduces the point-wise difference to $1\times10^{-3}$. The Clawpack solution, computed on the same grid (24 cells per layer), is shown in \Fig{stego_tr_cp}. Again, the SharpClaw solution is noticeably more accurate. For a more detailed study of this time-reversibility test, we refer to \cite{ketcheson2010}. \begin{figure} \centerline{ \subfigure[SharpClaw.\label{fig:stego_tr_sc}]{\includegraphics[width=3in]{figures/stego_tr_sc.eps}} \subfigure[Clawpack.\label{fig:stego_tr_cp}]{\includegraphics[width=3in]{figures/stego_tr_cp.eps}}} \caption[Comparison of forward solution and time-reversed solution stegotons.]{Comparison of forward solution (black line) and time-reversed solution (symbols).\label{fig:stego_tr}} \end{figure} \subsection{Shallow Water Flow} The next example involves solution of the two-dimensional shallow water equations \begin{subequations} \label{eqn:shallowWater2DCons} \begin{align} \label{eqn:shallowWater2DConsMass} h_t + (hu)_x + (hv)_y & = 0 \\ \label{eqn:shallowWater2DConsMomx} (hu)_t + \left(\frac{1}{2}hu^2 + \frac{1}{2}gh^2\right)_x + (huv)_y & = -g h b_x \\ \label{eqn:shallowWater2DConsMomy} (hv)_t + (huv)_x + \left(\frac{1}{2}hv^2 + \frac{1}{2}gh^2\right)_y & = -g h b_y, \end{align} \end{subequations} where $h$, $u$ and $v$ are the depth of the fluid and the velocity components in $x$ and $y$ directions, respectively. The function $b(x,y)$ is the bottom elevation and the constant $g$ represents gravitational acceleration. Two test cases are presented: a radially symmetric dam-break problem over a flat bottom and a small perturbation of a steady state over a hump. A Roe solver with entropy fix is used in both cases. \subsubsection{Radial dam-break problem} \label{subsubsec:1Dradial-dam-break-problem} This problem consists in computing the flow induced by the instantaneous collapse of an idealized circular dam. It is widely used to benchmark numerical methods designed to simulate interfacial flows and impact problems. The domain considered is $[-1.25,1.25]\times[-1.25,1.25]$. The initial depth is \begin{align} h(x,y,t=0) & = \begin{cases} 2 & \mbox{ for } \sqrt{x^2+y^2} \le 1/2 \\ 1 & \mbox{ for } \sqrt{x^2+y^2} > 1/2, \end{cases} \end{align} and the initial velocity is zero everywhere. This tests the ability of the method to compute the 2D propagation of nonlinear waves and the extent to which symmetry is preserved in the numerical solution. In the presence of radial symmetry, system (\ref{eqn:shallowWater2DCons}) can be recast in the following form: \begin{subequations} \begin{align} \label{eqn:shallowWater1DRadSymMass} h_t + (hU)_r & = -\frac{hU}{r} \\ \label{eqn:shallowWater1DRadSymMomr} (hU)_t + \left(\frac{1}{2}hU^2 + \frac{1}{2}gh^2\right)_r & = -\frac{hU^2}{r}, \end{align} \label{eqn:shallowWater1DRadSym} \end{subequations} where $h$ is still the depth of the fluid, whereas $U$ and $r$ are the radial velocity and the radial position. An important feature of these equations is the presence of a source term, which physically arises from the fact that the fluid is spreading out and it is impossible to have constant depth and constant non-zero radial velocity. A first comparison between SharpClaw and Clawpack is performed by solving the 1D system (\ref{eqn:shallowWater1DRadSym}) on the interval $0\le r \le 2.5$. A wall boundary condition and non-reflecting boundary condition are imposed at the left and the right boundaries, respectively. The final time for the analysis is taken to be $t=1$. The classical $q$-wave Riemann solver based on the Roe linearization is used to solve the Riemann problem at each interface (see for instance \cite{levequefvbook} for details), where the left and the right states are computed by using the characteristic-wise WENO reconstruction. The gravitational acceleration is set to $g=1$. A highly resolved solution obtained with Clawpack on a grids with $25,600$ cells is used as a reference solution. It is well-known that high order convergence is not observed in the presence of shock waves \cite{Majda-1977} and typically the convergence rate is no greater than first order. However, if we plot the difference between the computed solution available at the cell's center and the reference solution conservatively averaged on the same grid, i.e. $E_i = |Q_i - \bar{Q}_i|$, then we can visualize where the errors are largest as well as their spatial structure. Figure \ref{fig:sw1DRad} shows this difference for the water height ($h$) on a grid with $800$ cells. The reference solution at $t=1$ is shown by the solid line in Figure \ref{fig:sw2DRadScatter125}. The largest errors in both solutions are near the shocks. In the smooth regions, the SharpClaw solution is more accurate than that of the Clawpack code. \begin{figure} \centering \includegraphics[trim=1cm 1cm 1cm 0cm, clip=true, width=5in] {figures/sw-L1Error-h-t1-800Cells.eps} \caption{Pointwise absolute error for the water height on a grid with $800$ cells. Solid line: SharpClaw solution; dashed line: Clawpack solution. \label{fig:sw1DRad}} \end{figure} Next we consider the same problem using the full 2D equations \eqref{eqn:shallowWater2DCons}. The SharpClaw and Clawpack codes are tested on two grids with $125 \times 125$ and $500 \times 500$ cells. The final time for the analysis is again taken to be $t=1$. Figure \ref{fig:sw2DRadScatter125} shows the water height $h$ computed with SharpClaw at each cell's center and $t=1.0$ as a function of the radial position. The radius is measured respect to the center of the initial condition. The 1D reference solution used before is also plotted for comparison. It can be seen that the scheme preserves a good radial symmetry, though it cannot resolve the shock near the origin. The grid is in fact too coarse. Clawpack results are not shown in this figure but indicate similar accuracy and similarly good symmetry. \begin{figure} \centering \includegraphics[trim=1cm 1cm 1cm 0cm, clip=true, width=5in]{figures/sw2DhScatterSharp125-Cut.eps} \caption{Solution for the 2D radial dam-break problem on a grid with $125 \times 125$ cells, plotted as a function of radius. \label{fig:sw2DRadScatter125}} \end{figure} The solutions obtained on the finer grid ($500 \times 500$ cells) are shown in Figure \ref{fig:sw2DRadScatter500}. The effect of the grid refinement is clearly visible. In fact, the solutions gets close to the reference solution. However, the density of the grid near the origin is still too coarse to resolve the shock near the origin to high accuracy. \begin{figure} \centering \includegraphics[trim=1.0cm 1cm 1cm 0cm, clip=true, width=5in]{figures/sw2DhScatterSharp500-Cut.eps} \caption{Solution for the 2D radial dam-break problem on a grid with $500 \times 500$ cells, plotted as a function of radius. \label{fig:sw2DRadScatter500}} \end{figure} \subsubsection{Perturbation of a steady state solution} \label{subsubsec:2Dperturb-SWEs} Conservation laws with source terms often have steady states in which the flux gradient are non-zero but exactly balanced by source terms. A good numerical scheme should be able to preserve such steady states and accurately model small perturbations around these conditions. A classical benchmark test case to investigate these properties is the small perturbation of a 2D steady state water given by LeVeque \cite{LeVeque-1998-QuasiSteady}. System (\ref{eqn:shallowWater2DCons}) is solved in a rectangular domain $[0,2] \times [0,1]$, with a bottom topography characterized by an ellipsoidal gaussian hump: $$ b(x,y) = 0.8 \exp(-5(x-0.9)^2 - 50(y-0.5)^2). $$ The surface is initially flat with $h(x,y,0)=1-b(x,y)$ except for $0.05<x<0.15$, where $h$ is perturbed upward by $\epsilon = 0.01$. The initial discharge in both direction is zero, i.e. $hu(x,y,0)=hv(x,y,0) = 0$. Non-reflecting (i.e., zero-extrapolation) conditions are imposed at all boundaries. The gravitational acceleration is set to $g=9.81$. An effort was made to achieve a well-balanced scheme using the $f$-wave approach combined with component-wise or characteristic-wise WENO reconstruction, but this was unsuccessful. This is not surprising, since the algorithm begins by reconstructing a non-constant function. Figure \ref{fig:noBalanced} shows the contour levels of the solution at $t=0.12$ on a fine grid with $600 \times 300$ cells, obtained with the $f$-wave Riemann solver and the component-wise reconstruction approach as a building block for the WENO scheme. The scheme is not well-balanced and spurious waves are generated around the hump. Similar results are obtained using characteristic-wise reconstruction. \begin{figure} \centering \subfigure[Component-wise reconstruction.\label{fig:noBalanced}]{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=3.2in]{figures/Std-fwave-noBalanced-t012.eps}} \subfigure[$f$-wave-slope reconstruction.\label{fig:balanced}]{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=3.2in]{figures/Pert2DSW-600x300-t012.eps}} \caption{Contour of the surface level $h + b$ at time $t=0.12$ computed with component-wise reconstruction and $f$-wave-slope reconstruction. Contour levels: $0.99942 : 0.000238: 1.00656$.\label{fig:comp-noBalanced-balanced}} \end{figure} In order to balance the scheme, the $f$-wave-slope reconstruction introduced in Section \ref{subsec:waveSlopeReconstruction} is used instead. In this approach, the WENO reconstruction is applied to waves computed by solving the Riemann problem at the cell's interface with the $f$-wave solver. The bathymetry is approximated by a piecewise-constant function so that its effect is concentrated at the cell interfaces. When the source term is included in these Riemann problems, the resulting waves vanish as shown in Figure \ref{fig:balanced}. In Figure \ref{fig:sw2DSlice-y0p5} the surface level a cross section along $y=0.5$ at time $t=0.06$ computed with both reconstruction approaches (and the $f$-wave Riemann solver) on a uniform mesh with $600 \times 300$ is plotted. This comparison illustrates the different nature of the two approaches. The $f$-wave-slope reconstruction method keeps the surface flat, whereas the component-wise reconstruction introduces spurious waves which have an amplitude of the order of the disturbance that we want to resolve. \begin{figure} \centering \includegraphics[trim=1.0cm 1cm 1cm 0cm, clip=true, width=5in]{figures/crossSection-y0p5-t006.eps} \caption{Surface level $h + b$ along a cross section at $y=0.5$ and time $t=0.06$. Solid line: $f$-wave Riemann solver and component-wise reconstruction; dashed line: $f$-wave Riemann solver and $f$-wave-slope reconstruction.\label{fig:sw2DSlice-y0p5}} \end{figure} Figure \ref{fig:Pert2DSWE-coraseFineGrids} shows the solution on two uniform meshes with $200 \times 100$ cells and $600 \times 300$ cells, computed using the $f$-wave-slope reconstruction approach. The results clearly indicate that the detailed structure of the evolution of such a small perturbation is resolved well even with the relatively coarse mesh. These results agree with those reported in \cite{LeVeque-1998-QuasiSteady}. \begin{figure} \centering \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-200x100-t012.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-600x300-t012.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-200x100-t024.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-600x300-t024.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-200x100-t036.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-600x300-t036.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-200x100-t048.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-600x300-t048.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-200x100-t060.eps}} \subfigure{\includegraphics[trim=1.5cm 3cm 1cm 3cm, clip=true, width=2.3in]{figures/Pert2DSW-600x300-t060.eps}} \caption{Contour of the surface level $h + b$. $f$-wave-slope reconstruction. 30 uniformly spaced contour lines. $t = 0.12$ from $0.99942$ to $1.00656$; $t = 0.24$ from $0.99318$ to $1.01659$; $t = 0.36$ from $0.98814$ to $1.01161$; $t = 0.48$ from $0.99023$ to $1.00508$; $t = 0.6$ from $0.995144$ to $1.00629$. Left: results with a $200\times100$ cells. Right: results with a $600\times300$ cells. \label{fig:Pert2DSWE-coraseFineGrids}} \end{figure} In order to investigate the accuracy of our scheme for smooth solutions we also have performed a convergence study at a fixed CFL of 0.3 for the same 2D problem. The smooth initial perturbation is given by \begin{align} \label{smooth} h(x,y,t=0) = \exp(-50(x-0.1)^2)/100 \end{align} The results of the convergence study are shown in Table \ref{tab:convergence-pert-shallow}. A highly resolved numerical simulation computed with Clawpack on a $40,000\times 20,000$ grid has been used as a reference solution. Since the bathymetry $b(x,y)$ is approximated by a piecewise-constant function, both discretizations are formally only second-order accurate, and the results are roughly consistent with this. Nevertheless, the SharpClaw discretization yields significantly more accurate results. \begin{table}[th] \caption{Convergence results for the smooth initial condition \eqref{smooth}. $L_2$-norm of the error as a function of the grid spacing. \label{tab:convergence-pert-shallow}} \begin{center} \begin{tabular}{c|cc|cc} & \multicolumn{2}{c|}{SharpClaw}& \multicolumn{2}{c}{Clawpack}\\ \hline $\Delta x=\Delta y$ & Error & Order & Error & Order\\ \hline 1/10 & 1.14e-2 & & 2.78e-2 & \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 1/20 & 7.00e-3 & 0.70 & 2.09e-2 & 0.41 \\ 1/100 & 8.11e-4 & 1.34 & 6.33e-3 & 0.74 \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 1/200 & 3.24e-4 & 1.32 & 2.40e-3 & 1.40 \\ 1/1000 & 2.93e-5 & 1.49 & 1.43e-4 & 1.75 \\ \rowcolor[rgb]{0.9, 0.9, 0.9} 1/2000 & 5.94e-6 & 2.30 & 3.81e-5 & 1.91 \\ \hline \end{tabular} \end{center} \end{table} \section{Conclusions} We have presented a general approach to extending the finite volume wave propagation algorithm to high order accuracy in one and two dimensions. The algorithm is based on a method-of-lines approach, wherein the semi-discrete scheme relies on high order reconstruction and computation of fluctuations, including a {\em total fluctuation} term arising inside each cell. By using WENO reconstruction and strong stability preserving time integration, high order accurate non-oscillatory results are obtained, as demonstrated through a variety of test problems. This algorithm has several desirable features. Like the second-order wave propagation algorithms in Clawpack \cite{leveque1997}, it is applicable to hyperbolic PDEs including linear nonconservative systems and nonlinear systems with spatially varying flux function. It has been shown to achieve high order accuracy even for problems with discontinuous coefficients. Finally, the algorithm can be adapted to give a well-balanced scheme for balance laws by use of the $f$-wave approach and a new {\em wave-slope reconstruction} technique. Hyperbolic systems of equations with both smooth and non-smooth solution have been used to test the properties and the capabilities of the proposed method. The schemes have been compared for linear acoustics and nonlinear elasticity problems in heterogeneous media and for the shallow water equations with and without bottom topography. Two types of Riemann solver have been used, i.e the classical ($q$-) wave algorithm and the $f$-wave approach. The new scheme performed well for all the test cases. It gives significantly better accuracy than Clawpack (on the same grid) for smooth problems. A drawback of our implementation of well-balancing is that it requires the effect of the source term to be approximated entirely at the cell interfaces. For shallow water equations with smooth bathymetry, this will reduce the formal accuracy to second order. Future work might explore the implementation of higher order accurate well-balancing for polynomial source terms using high order quadrature. In any case, in two dimensions, the presented dimension-by-dimension reconstruction approach is formally only second-order accurate (see \cite{Zhang-2011-WENO}); however, it gives improved accuracy over the second-order scheme implemented in Clawpack for the test problems considered. Further investigation of different approaches to multidimensional reconstruction for problems containing both shocks and rich smooth flow structures is a topic of future research. \subsection*{Acknowledgments} The authors are grateful to anonymous referees whose suggestions improved this paper. This work was supported in part by NSF grants DMS-0609661 and DMS-0914942.
\section{Introduction} The aim of this paper is to study the hitting time of some curved boundaries for the Bessel process. Our main motivations come from mathematical finance, optimal control and neurosciences. In finance Cox-Ingersoll-Ross processes are widely used to model interest rates. As an application, in this article we will consider the simulation of the first hitting time of a given level for the CIR by using its relation with the Bessel process. In neurosciences the firing time of a neuron is usually modelled as the hitting time of a stochastic process associated with the membrane potential behaviour (for introduction of noise in neuron systems see Part I Chapter 5 in \cite{gerstner}). The literature proposes different continuous stochastic models like for instance the family of integrate-and-fire models (see Chapter 10 in \cite{ermentrout}). Most of them are related to the Ornstein-Uhlenbeck process which appears in a natural way as extension of Stein's model, a classical discrete model. In Feller's model, generalized Bessel processes appear as a more realistic alternative to the Ornstein-Uhlenbeck process, see for instance \cite{feller-mod} for a comparison of these models. Therefore the interspike interval, which is interpreted as the first passage time of the membrane potential through a given threshold is closely related to the first hitting time of a curved boundary for some Bessel processes. Our main results and the main algorithm are obtained for the case of Bessel processes. We use in our numerical method the particular formula that we obtain for the hitting time of some curved boundaries for the Bessel process and the connection that exists between a Bessel process and the Euclidean norm of a Brownian motion when calculating the hitting position. As an application we consider the hitting time of a given level for the Cox-Ingersoll-Ross process. In order to obtain this, we use first of all the connection that exists between CIR processes and Bessel processes and secondly the method of images for this particular situation. The study of Bessel processes and their hitting times occupies a huge mathematical literature. Let us only mention few of them: A. G{\"o}ing-Jaeschke and M. Yor \cite{jaeschke_yor_2003} consider a particular case of CIR processes which are connected with radial Ornstein-Uhlenbeck processes and their hitting times; L. Alili and P. Patie \cite{alili_patie_2010} investigate as a special situation the Bessel processes via some boundary crossing identities for diffusions having the time inversion property; recently P.~Salminen and M.~Yor consider the hitting time of affine boundaries for the 3-dimensional Bessel process \cite{salminen_yor_2011}. In a recent paper Y. Hamana and H. Matsumoto \cite{hamana_matsumoto_2011} give explicit expressions for the distribution functions and the densities of the first hitting time of a given level for the Bessel process. Their results cover all the cases. Let us also mention a recent work of T. Byczkowski, J. Malecki and M. Ryznar \cite{byczkowski_malecki_ryznar_2011}. By using an integral formula for the density of the first hitting time of a given level of the Bessel process, they are able to obtain uniform estimates of the hitting time density function. In all these papers the formulas are explicit and are hard to use for a numerical purpose as they exhibit Bessel functions. The main idea of this paper is to get rid of this difficulty by using two important tools: first of all the method of images that allows to obtain, for some particular boundaries, an explicit form for the density of the hitting time and secondly the connection between $\delta$-dimensional Bessel processes and the Euclidean norm of a $\delta$-dimensional Brownian motion in order to get the simulated exit position. By coupling these ingredients we are able to construct a numerical algorithm easy to implement and very efficient which approaches the desired hitting time. We will use here a modified version of the {\sl random walk on spheres} method which was first introduced by Muller \cite{muller_56} in 1956. This procedure allows to solve a Dirichlet boundary value problem. The idea is to simulate iteratively, for the Brownian motion, the exit position from the largest sphere included in the domain and centered in the starting point. This exit position becomes the new starting point and the procedure goes on until the exit point is close enough to the boundary. Let us notice that the simulation of the exit time from a sphere is numerically costly. The method of images was introduced in 1969 by H.E. Daniels \cite{daniels_1969} as a tool to construct nonlinear boundaries for which explicit calculations for the exit distribution for the Brownian motion are possible. The method was developed also in H. R. Lerche \cite{lerche_1986}. We adapt this method for the Bessel process by using the explicit form of its density. For some particular curved boundaries we can explicitly evaluate the density of the Bessel hitting time. The paper is organized as follows. First we present some new results on hitting times for Bessel processes. Secondly we construct the new algorithm for approaching the hitting time, the so called Walk on Moving Spheres algorithm. Finally we present some numerical results and as a particular application the evaluation of the hitting time of a given level for the Cox-Ingersoll-Ross process. \section{Hitting time for Bessel processes} \noindent \par Bessel processes play an important role both in the study of stochastic processes like Brownian motion and in various theoretical and practical applications as for example in finance. Let us consider the $\delta$-dimensional Bessel process starting from $y$, solution of the following stochastic differential equation: \begin{equation} \label{bessel-delta} \left\{ \begin{array}{ll} Z^{\delta,y}_t& = Z^{\delta,y}_0+\displaystyle\frac{\delta-1}{2}\int_0^t (Z^{\delta,y}_s)^{-1} \mathrm{d} s + B_t,\\ Z^{\delta,y}_0&=y, \qquad y\geq0,\\ \end{array} \right. \end{equation} where $(B_t)_{t\geq 0}$ is a one dimensional Brownian motion. We denote \begin{equation} \nu = \frac{\delta}{2}-1 \end{equation} the \emph {index} of this process. We call $\delta$ the \emph {dimension} of the process. This terminology is coming from the fact that, in the case of positive integer $\delta \in \mathbb{N}$, a $\delta$-dimensional Bessel process can be represented as the Euclidean norm of a $\delta$-dimensional Brownian motion. This will be a key point in our numerical method. The density of this process starting from $y$ is given by: \begin{equation} \label{density} p_y(t,x)=\displaystyle\frac{x}{t}\displaystyle\left(\frac{x}{y}\right)^\nu \exp\left(-\displaystyle\frac{x^2+y^2}{2t}\right) I_\nu \left(\displaystyle\frac{xy}{t}\right),\quad \mbox { for } t>0, \, y>0, x\geq 0, \end{equation} where $I_\nu(z) $ is the Bessel function whose expression writes: \begin{equation} \label{bessel-function} I_\nu(z) = \displaystyle\sum_{n=0}^\infty \left( \displaystyle\frac{z}{2}\right) ^{\nu+2n} \displaystyle\frac{1}{n ! \Gamma (\nu+n+1)}. \end{equation} When starting from $y=0$, the density of $Z^{\delta,0}_t$ is: \begin{equation} \label{density-bessel-0} p_0(t,x) = \displaystyle\frac{1}{2^{\nu}}\displaystyle\frac{1}{t^{\nu+1}}\displaystyle\frac{1}{\Gamma (\nu+1)} x^{\delta-1}\exp\left(-\displaystyle\frac{x^2}{2t}\right), \mbox{ for } t>0, \, x\geq 0 . \end{equation} \subsection{The method of images for Bessel processes} In this section, we investigate the first hitting time of a curved boundary for the Bessel process starting from the origin. Let $\psi(t)$ denote the boundary, and introduce the following stopping time: \begin{equation} \label{taupsi} \tau_\psi=\inf\{ t\geq 0;\ Z^{\delta,0}_t\geq \psi(t)\}. \end{equation} For some suitable choice of the boundary, the distribution of $\tau_\psi$ can be explicitly computed. The idea is given by the following remark on the method of images (see for instance \cite{daniels_1969} for the origin of this method and \cite{lerche_1986} for a complete presentation): {\bf {Fundamental idea.}} Suppose that $F$ is a positive $\sigma$-finite measure satisfying some integrability conditions (to be specified later on) and define \begin{equation} u(t,x)=p_0(t,x)-\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+}p_y(t,x) F(\mathrm{d} y), \end{equation} for some real constant $a>0$. Let \begin{equation*} \psi (t) =\inf \{ x\in \mathbb{R}; u(t,x) < 0\}, \mbox { for all } t>0. \end{equation*} Then $u(t,x)$ is solution of the partial differential equation: \begin{equation} \label{PDE_bessel} \left\{ \begin{array}{ll} \displaystyle\frac{\partial u}{\partial t}(t,x) = \displaystyle\frac{1}{2} \displaystyle\frac{\partial ^2 u}{\partial x^2}(t,x) -\displaystyle\frac{\delta-1}{2} \displaystyle\frac{\partial}{\partial x}\left(\displaystyle\frac{1}{x} u(t,x)\right), & \mbox { on }\mathbb{R}_+\times \mathbb{R},\\ u(t,\psi (t)) = 0,& \mbox { for all } t>0,\\ u(0,.)=\delta_0(.) & \mbox { on } (-\infty, \psi (0+)].\\ \end{array} \right. \end{equation} From this remark we deduce an interesting expression for the hitting time. We can prove that: \begin{equation*} \tau_\psi =\inf\{t>0;\, u(t, Z^{\delta,0}_t)=0\}. \end{equation*} This means simply that in order to obtain informations on the hitting time it sufficies to look for $u(t, Z^{\delta, 0}_t)=0$. Let us express this in a general result \begin{thm} \label{thm_general_setting} Let $F(\mathrm{d} y)$ be a positive $\sigma$-finite measure such that $\int_0^\infty p_0(t,\sqrt {\varepsilon} y) F(\mathrm{d} y) <\infty $ for all $\varepsilon >0$. Let $a>0$ and define the function: \begin{equation} \label{general_u} u(t,x)=p_0(t,x)-\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+}p_y(t,x) F(\mathrm{d} y). \end{equation} Consider $\psi (t)$ such that $u(t,\psi (t))=0$. Then the probability density function of $\tau_\psi $ is given by \begin{equation} \label{distribution_tau} \mathbb{P}_0(\tau_\psi\in \mathrm{d} t) = \left[-\left.\displaystyle\frac{1}{2} \frac {\partial u}{\partial x}(t,x)\right|_{x=\psi(t)}+\left.\displaystyle\frac{1}{2} \frac {\partial u}{\partial x}(t,x)\right|_{x=0} - \left.\displaystyle\frac{\delta -1}{2x} u(t,x)\right|_{x=0}\right]\mathrm{d} t. \end{equation} \end{thm} \begin{proof} We will only point out the main ideas for the proof in this case as it follows mainly the ideas introduced in \cite{lerche_1986}. A complete description of the method and this result for the Brownian motion case can be found in \cite{lerche_1986}. Let us consider \begin{equation} \label{function-u-1} u(t,x)= p_0(t,x)-\displaystyle\frac{1}{a} \displaystyle\int_{\mathbb{R_+}} p_y(t,x)F(\mathrm{d} y), \end{equation} where $F(\mathrm{d} y)$ is a measure on $\mathbb{R}_+$. We consider $\psi(t)$ the solution of $u(t,\psi (t))=0$. Let us define as before $\tau_\psi = \inf\{ t\geq 0; Z^{\delta,0}_t\geq \psi (t)\}$. Then $u(t,x)\mathrm{d} x=\mathbb{P}(\tau_\psi>t,\ Z^{\delta,0}_t\in\mathrm{d} x)$ and \begin{equation} \label{proba_hitting_time} \mathbb{P}_0(\tau_\psi > t) =\displaystyle\int_0^{\psi(t)} u(t,x)\mathrm{d} x. \end{equation} In order to get the distribution of $\tau_\psi$ we have to evaluate the derivative of $\mathbb{P}_0(\tau_\psi > t )$. By using the equality (\ref{proba_hitting_time}) we obtain: \begin{equation} \begin{array}{ll} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) &= \left(-\psi '(t) u(t, \psi(t)) -\displaystyle\int_0^{\psi (t)} \displaystyle\frac{\partial u}{\partial t} (t,x) \mathrm{d} x\right)\mathrm{d} t\\ &= \left(-\displaystyle\frac{1}{2}\displaystyle\int_0^{\psi (t)} \displaystyle\frac{\partial^2 u}{\partial x^2}(t,x)\mathrm{d} x +\displaystyle\frac{\delta-1}{2}\displaystyle\int_0^{\psi(t)} \displaystyle\frac{\partial }{\partial x} \left(\displaystyle\frac{1}{x} u(t,x)\right)\mathrm{d} x\right)\mathrm{d} t, \end{array} \end{equation} as $u(t,x)$ is solution of the partial differential equation (\ref{PDE_bessel}). We obtain thus: \begin{equation} \label{proba_tau} \begin{array}{ll} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) & =\left( -\left.\displaystyle\frac{1}{2} \frac {\partial u}{\partial x}(t,x)\right|_{x=\psi(t)}+\displaystyle\frac{\delta -1}{2\psi(t)} u(t,\psi(t))+ \left.\left(\displaystyle\frac{1}{2} \frac {\partial u}{\partial x}(t,x) - \displaystyle\frac{\delta -1}{2x} u(t,x)\right)\right|_{x=0}\right)\mathrm{d} t. \end{array} \end{equation} As $\frac{\delta -1}{2\psi(t)} u(t,\psi(t))=0$ and this ends the proof of the theorem. \end{proof} The idea behind the method of images is that for some particular forms of $F(\mathrm{d} y)$ we can derive explicit formulas of the hitting time distribution. More precisely: \begin{proposition} \label{prop:first_boundary} Let, for $\delta=2\nu +2 > 0$ and $a > 0$ \begin{equation}\label{eq:ajout} \psi_a(t)=\sqrt{2t\log\displaystyle\frac{a}{\Gamma (\nu+1) t^{\nu+1}2^\nu}},\quad \mbox{ for }t\leq\left[\frac{a}{\Gamma(\nu+1)2^\nu}\right]^{\frac{1}{\nu+1}}. \end{equation} Then the probability density of $\tau_\psi$ is given by \begin{equation} \label{densite_tau} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) = \frac{1}{2at}\left( 2t \log \displaystyle\frac{a}{\Gamma (\nu +1)t^{\nu+1}2^\nu }\right)^ {\nu+1}\mathrm{d} t. \end{equation} \end{proposition} \begin{proof} By using the expression in $(\ref{density})$ we remark first that \begin{equation} \label{change-x-y} y^{2\nu+1} p_y(t,x)= x^{2\nu+1} p_x(t,y). \end{equation} Let us consider, as in Theorem \ref{thm_general_setting} \begin{equation} \label{function-u-2} u(t,x)= p_0(t,x)-\displaystyle\frac{1}{a} \displaystyle\int_{\mathbb{R_+}} p_y(t,x)F(\mathrm{d} y), \end{equation} with $F(\mathrm{d} y) = y^{2\nu+1} \ind{\{y >0\}}\mathrm{d} y$. In this situation the function $u$ defined in \eqref{function-u-2} writes \begin{equation} \label{specific-F} \begin{array}{ll} u(t,x) &= p_0(t,x) -\displaystyle\frac{1}{a} x^{2\nu +1}\\ & = \left( \displaystyle\frac{1}{2^{\nu}}\displaystyle\frac{1}{t^{\nu+1}}\displaystyle\frac{1}{\Gamma(\nu+1)}\exp\left(-\displaystyle\frac{x^2}{2t}\right)-\displaystyle\frac{1}{a}\right)x^{2\nu+1}. \end{array} \end{equation} For simplicity we will write $\psi $ instead of $\psi_a$. Following the result in the Theorem \ref{thm_general_setting}, we are looking for $\psi (t)$ such that $u(t,\psi (t))=0$. This yields~: \begin{equation} \label{curve} x=\psi(t)=\displaystyle\sqrt{2t\log\displaystyle\frac{a}{\Gamma (\nu+1) t^{\nu+1}2^\nu}} \end{equation} under the obvious condition $t^{\nu+1} \leq \frac{a}{\Gamma(\nu+1) 2^\nu}$.\\ We can now notice that: \begin{equation*} p_0(t,\psi (t)) = \displaystyle\frac{1}{a}(\psi(t))^{2\nu +1}, \end{equation*} and we can prove easily that: \begin{equation*} \displaystyle\frac{\partial u}{\partial x} (t,x)=(\delta -1)\displaystyle\frac{u(t,x)}{x} -\displaystyle\frac{x}{t}p_0(t,x). \end{equation*} We obtain, after replacing in (\ref{distribution_tau}) and after applying the Theorem \ref{thm_general_setting}, for this particular case: \begin{equation*} \begin{array}{ll} \mathbb{P}_0(\tau_\psi\in \mathrm{d} t)&=\displaystyle\frac{1}{2t}\,\,\psi (t) p_0(t,\psi(t))\mathrm{d} t\\ & = \displaystyle\frac{1}{2at}\ \psi ^{2\nu +2} (t)\mathrm{d} t\\ &= \displaystyle\frac{1}{2at}\left( 2t \log \displaystyle\frac{a}{\Gamma(\nu+1)t^{\nu+1}2^\nu}\right)^{\nu+1} \mathrm{d} t, \end{array} \end{equation*} which gives the desired result. \end{proof} The second boundary which allows to express explicit results is obtained by using the Markov property for the Bessel process. \begin{proposition} \label{prop:second_boundary} Let, for $\delta=2\nu+2 > 0$, $s>0$ and $a >0$ fixed, \begin{equation} \label{second_boundary} \psi_a(t)=\sqrt{\displaystyle\frac{2t(t+s)}{s}\left[ (\nu +1)\log \left(1+\frac{s}{t}\right) +\log a \right]}, \end{equation} for all $t\geq 0$ if $a\geq 1$ and for $t\leq \frac{s}{(\frac{1}{a})^{\frac{1}{\nu+1}}-1}$ if $0< a< 1$. Then the probability density function of the hitting time $\tau_\psi $ is given by: \begin{equation} \label{proba_tau_second_boundary} \mathbb{P}_0(\tau_\psi\in \mathrm{d} t)=\displaystyle\frac{1}{\Gamma (\nu +1)}\displaystyle\frac{1}{t}\left(\displaystyle\frac{t+s}{s}\right)^{\nu}\left[\log \left( a\left(\frac{t+s}{t}\right)^{\nu +1}\right)\right]^{\nu+1} \exp\left[-\displaystyle\frac{t+s}{s}\log \left(a \left(\frac{t+s}{t}\right)^{\nu+1}\right)\right] \mathrm{d} t. \end{equation} \end{proposition} \begin{proof} We will only sketch the proof as it follows the same ideas as the one of the Theorem \ref{thm_general_setting}. Let us consider the measure $F(\mathrm{d} y)= p_0(s,y)\mathrm{d} y$ for $s>0$ fixed. Then, when evaluating the corresponding $u(t,x)$ we have: \begin{equation*} \begin{array}{ll} u(t,x) & = p_0(t,x)-\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+} p_0(s,y) p_y(t,x)\mathrm{d} y \\ & = p_0(t,x)-\displaystyle\frac{1}{a} p_0(t+s,x)\\ & = \displaystyle\frac{1}{2^\nu}\displaystyle\frac{1}{\Gamma (\nu+1)}x^{2\nu +1}\left[\displaystyle\frac{1}{t^{\nu+1}} \exp\left( -\displaystyle\frac{x^2}{2t}\right) -\displaystyle\frac{1}{a}\displaystyle\frac{1}{(t+s)^{\nu +1}} \exp \left(-\displaystyle\frac{x^2}{2(t+s)}\right)\right],\\ \end{array} \end{equation*} by using the Markov property. We obtain the form of $\psi (t)$ by the condition $u(t,\psi (t))=0$ which gives: \begin{equation} \psi (t) = \sqrt{\displaystyle\frac{2t(t+s)}{s}\left[ (\nu +1)\log \left(1+\frac{s}{t}\right) +\log a \right]}, \quad \left\{ \begin{array}{l} \mbox{for } t\geq 0 \mbox{ if } a\geq 1 \\ \mbox{for } t\leq \frac{s}{(\frac{1}{a})^{\frac{1}{\nu+1}}-1} \mbox{ if } a < 1.\\ \end{array} \right. \end{equation} In order to obtain the distribution of $\tau_\psi$ one has only to evaluate: \begin{equation} \displaystyle\frac{\partial u}{\partial x} (t,x) = (\delta -1)\displaystyle\frac{u(t,x)}{x} -\displaystyle\frac{s}{t(t+s)}\, x\, p_0(t,x), \end{equation} and $\frac{u(t,x)}{x}$ for $x=0$ and $x=\psi (t)$ and replace the values in the general form (\ref{proba_tau}). The expression (\ref{proba_tau_second_boundary}) follows. \end{proof} \begin{rem} We can notice that the function $\psi_a(t)$ defined by (\ref{second_boundary}) satisfies for large times \begin{equation*} \left\{ \begin{array}{llcl} \psi_a(t)&\simeq & \sqrt{t} &{\mbox { for } }a =1, \\ \psi_a(t)&\simeq & t & {\mbox { for all } }a >1. \\ \end{array} \right. \end{equation*} In particular, we can approach large times by considering this kind of boundary. \end{rem} A new boundary can be obtained by using the Laplace transform of the square of the $\delta~$-dimensional Bessel process starting from $x$. More precisely: \begin{proposition} \label{prop:third__boundary} Let, for $\delta=2\nu +2 > 0$, $\lambda>0$ and $a>0$ fixed, \begin{equation} \label{psi_laplace} \psi_a(t)=\displaystyle\frac{\lambda t}{1+2\lambda t}+\ t\sqrt{ \left(\displaystyle\frac{\lambda }{1+2\lambda t}\right)^2 +\displaystyle\frac{2}{t} \log \displaystyle\frac{a(1+2\lambda t)^{\nu +1}}{2^\nu t^{\nu +1}\Gamma(\nu +1)}} \end{equation} for \begin{equation} \left\{ \begin{array}{ll} t\leq \frac{1}{(2^\nu\Gamma(\nu +1)/a)^{1/(\nu+1)}-2\lambda} & \mbox { if } \lambda <\displaystyle\frac{1}{2}\left(\displaystyle\frac{2^\nu \Gamma (\nu +1)}{a}\right)^{\frac{1}{\nu +1}}, \\ t\geq 0&\mbox { if }\lambda \geq \displaystyle\frac{1}{2}\left(\displaystyle\frac{2^\nu \Gamma (\nu +1)}{a}\right)^{\frac{1}{\nu +1}}. \end{array} \right. \end{equation} Then the probability density function of the hitting time is given by: \begin{equation} \label{proba_tau_third_boundary} \mathbb{P}_0(\tau_\psi\in \mathrm{d} t)=\sqrt{ \left(\displaystyle\frac{\lambda }{1+2\lambda t}\right)^2 +\displaystyle\frac{2}{t} \log \displaystyle\frac{a(1+2\lambda t)^{\nu +1}}{2^\nu t^{\nu +1}\Gamma(\nu +1)}}p_0(t,\psi(t))\mathrm{d} t. \end{equation} \end{proposition} \begin{proof} We present only the main ideas as the result follows as above from the general method in Theorem \ref{thm_general_setting} applied to the measure $F(\mathrm{d} y) = y^{2\nu +1} e ^{-\lambda y^2}\ind{\{y\geq 0\}}\mathrm{d} y$. For this measure $u(t,x)$ takes the form: \begin{equation*} \begin{array}{lcl} u(t,x) & = & p_0(t,x)-\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+} p_y(t,x) F(\mathrm{d} y)\\ &=& p_0(t,x) -\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+} p_y(t,x)y^{2\nu +1} e ^{-\lambda y^2}\mathrm{d} y\\ &=& p_0(t,x) -\displaystyle\frac{1}{a}\displaystyle\int_{\mathbb{R}_+}x^{2\nu+1}p_x(t,y) e ^{-\lambda y^2}\mathrm{d} y\\ &=& p_0(t,x) -\displaystyle\frac{x^{2\nu +1}}{a} \mathbb{E} \left(e^{-\lambda Z_t^{\delta,x}}\right).\\ \end{array} \end{equation*} By using the expression of the Laplace transform for $Z_t^{\delta,x}$ we obtain \begin{equation} \label{TL_boundary} u(t,x) = p_0(t,x) -\displaystyle\frac{x^{2\nu +1} }{a}\displaystyle\frac{1}{(1+2\lambda t)^{\delta/2}}\exp\left(-\displaystyle\frac{\lambda x}{1+2\lambda t}\right). \end{equation} We consider first the equality $u(t,\psi (t))=0$ in (\ref{TL_boundary}) and this gives the form of $\psi (t)$ in (\ref{psi_laplace}). Afterwards, we can evaluate once again in this particular situation \begin{equation*} \displaystyle\frac{\partial u}{\partial x}(t,x) = (\delta -1)\displaystyle\frac{u(t,x)}{x} -\left(\displaystyle\frac{x}{t} -\displaystyle\frac{\lambda t}{1+2\lambda t}\right) p_0(t,x). \end{equation*} For this particular case, there is only one no vanishing term in the expression (\ref{distribution_tau}) of $\mathbb{P}_0(\tau_\psi\!\in \!\mathrm{d} t)$, that is the term $-\left(\frac{x}{t} -\frac{\lambda t}{1+2\lambda t}\right) p_0(t,x)$ of $\frac{\partial u}{\partial x}(t,x)$ for $x=\psi (t)$ and this is exactly given by the right side of the formula (\ref{proba_tau_third_boundary}). \end{proof} \begin{corollary} The previous results writes for $\delta =2$ \begin{itemize} \item[$(1)$] For $a>0$, $t\leq a$ and $\psi(t)=\sqrt{2t\log \displaystyle\frac{a}{t}}$, the density of the hitting time $\tau_\psi$ is \begin{equation*} \label{densite_tau1} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) = \displaystyle\frac{1}{2a} \log \displaystyle\frac{a}{t}\mathrm{d} t. \end{equation*} \item[$(2)$] For $s>0$, $a>0$, $t\leq\frac{sa}{1-a}$ and $\psi (t) = \sqrt{\displaystyle\frac{2t(t+s)}{s}\log\left(a\displaystyle\frac{t+s}{t}\right)}$, the probability density function of $\tau_\psi$ is given by: \begin{equation*} \label{densite_tau2} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) = \displaystyle\frac{t+s}{t} \log\left ( a\displaystyle\frac{t+s}{t}\right ) \exp \left[-\displaystyle\frac{t+s}{t}\log \left( a\displaystyle\frac{t+s}{t}\right)\right]\mathrm{d} t. \end{equation*} \item[$(3)$] For $a>0$ and $\psi (t) =\displaystyle\frac{\lambda t}{1+2\lambda t}+t\sqrt{\left(\displaystyle\frac{\lambda}{1+2\lambda t}\right )^2 +\displaystyle\frac{2}{t}\log \displaystyle\frac{a(1+2\lambda t)}{t}}$, for \begin{equation} \left\{ \begin{array}{ll} t\geq 0, & \mbox { if } \lambda \geq {\displaystyle\frac{1}{2a}},\\ \\ t\leq\displaystyle \frac{a}{1-2\lambda a} &\mbox { if } \lambda <\displaystyle\frac{1}{2a}, \end{array} \right. \end{equation} the probability density function of $\tau_\psi$ is: \begin{equation*} \label{densite_tau3} \mathbb{P}_0(\tau_\psi \in \mathrm{d} t) = \sqrt{\left(\displaystyle\frac{\lambda}{1+2\lambda t}\right )^2 +\displaystyle\frac{2}{t}\log \displaystyle\frac{a(1+2\lambda t)}{t}} p_0(t, \psi(t))\mathrm{d} t. \end{equation*} \end{itemize} \end{corollary} \subsection{Approximation of the first hitting time for Bessel processes starting from the origin} In this section we will construct a stepwise procedure, the so-called \emph{random Walk on Moving Spheres (WoMS)} algorithm, which allows to approach the first time the standard Bessel process hits a given level $l>0$. Of course, this stopping time $\tau_l=\inf\{t> 0;\, Z^{\delta,x}_t=l\}$ can be characterized by its well-known Laplace transform computed by solving an eigenvalue problem. Indeed if $(Z^{\delta,x}_t,\,t\ge 0)$ is the Bessel process starting from $x$, of index $\nu=\frac{\delta}{2}-1$ then, for $\nu>0$ and $x\le l$, we get \[ \mathbb{E}_x[e^{-\lambda\tau_l}]=\frac{x^{-\nu}I_\nu(x\sqrt{2\lambda})}{l^{-\nu}I_\nu(l\sqrt{2\lambda})},\ x>0,\quad \mbox{and}\ \mathbb{E}_0\Big[ e^{-\lambda \tau_l} \Big]=\frac{(l\sqrt{2\lambda})^\nu}{2^\nu \Gamma(\nu+1)}\frac{1}{I_\nu(l\sqrt{2\lambda})}. \] Here $I_\nu$ denotes the modified Bessel function. This Laplace transform can be used to describe the tail distribution: Ciesielski and Taylor \cite{Ciesielski-Taylor} proved that, for $\delta=2\nu+2\in\mathbb{N}^*$, \[ \mathbb{P}_0(\tau_l>t)=\frac{1}{2^{\nu-1}\Gamma(\nu+1)}\,\sum_{k=1}^{\infty}\frac{j_{\nu,k}^{\nu-1}}{\mathcal{J}_{\nu+1}(j_{\nu,k})}\,e^{-\frac{j_{\nu,k}^2}{2l^2}t}. \] where $\mathcal{J}_\nu$ is the Bessel function of the first kind and $(j_{\nu,k})_{\nu,k}$ is the associated sequence of its positive zeros.\\ We are looking for a numerical approach for the hitting time and these formulas are not easy to handle and approach, in particular we can not compute the inverse cumulative function ! The aim of this section is to construct an easy and efficient algorithm without need of inverting the Laplace transform and without using discretization schemes since the hitting times are unbounded. In the next step we will extend this procedure to the hitting time of time-dependent boundaries like straight lines, useful in the description of the hitting time of a given level for the CIR process (the Laplace transform is then unknown). \subsubsection{Hitting time of a given level for the Bessel process with positive integer dimension $\delta$} \label{sect:0-horiz} Let us consider $\delta$ independent one-dimensional Brownian motions $(B^{(k)}_t,t\ge 0)$, $1\le k\le \delta$. Then the Bessel process of index $\delta$ starting from 0, satisfies the following property: \[ (Z^{\delta,0}_t,\,t\ge 0)\quad\mbox{has the same distribution as}\quad \left(\sqrt{\left(B^{(1)}_t\right)^2+\ldots+\left(B^{(\delta)}_t\right)^2},\, t\ge 0\right). \] Let \begin{equation} \label{tau-l} \tau_l = \inf \{ t\geq 0;\, Z^{\delta,0}_t \geq l\}. \end{equation} In particular, we can express $\tau_l$ by using the first time when the $\delta$-dimensional Brownian motion ${\bf B}=(B^{(1)},\ldots,B^{(\delta)})$ exits from the Euclidean ball $D$ centered in the origin with radius $l$. Approximating the exit time and the exit position for the 2-dimensional Brownian motion of a convex domain was already studied by Milstein \cite{Milstein-97}. He used the so-called random walk on spheres algorithm which allows to approach the exit location and the exit time through an efficient algorithm. The exit position is really simple to obtain (as it is uniformly distributed on the circle) while the exit time is much more difficult to approach. That is why we will construct an adaptation of this initial procedure in order to obtain nice and efficient results concerning the Bessel process exit time.\\[5pt] Let us introduce now our \emph{Walk on Moving Spheres} ($WoMS$) algorithm. We first define a continuous function $\rho:\mathbb{R}^2\to\mathbb{R}_+$ which represents the distance to the boundary of $D$: \begin{equation}\label{defderho} \rho(x)=\inf\{\Vert x-y\Vert;\ y\in D^{c}\}=l-\Vert x\Vert. \end{equation} For any small enough parameter $\varepsilon>0$, we will denote by $D^\varepsilon$ the sphere centered at the origin with radius $l-\varepsilon$ \begin{equation} \label{def_Depsilon } D^{\varepsilon} = \{ x\in D;\, \Vert x\Vert \leq l-\varepsilon \} = \{ x\in D;\, \rho (x) \geq \varepsilon \} . \end{equation} \centerline{\line(1,0){490}}\\ \emph{{\bf Algorithm (A1) for $\delta=2$.} Let us fix a parameter $0<\gamma<1$.\\ {\bf Initialization:} Set $X(0)=(X_1(0),X_2(0))=0$, $\theta_0=0$, $\Theta_0=0$, $A_0=\gamma^2l^2 e/2$.\\ {\bf First step:} Let $(U_1,V_1,W_1)$ be a vector of three independent random variables uniformly distributed on $[0,1]$. We set \begin{align*} \left\{\begin{array}{l} \theta_1=A_0 U_1V_1,\quad \Theta_1=\Theta_0+\theta_1,\\ [5pt] X(1)^\intercal=(X_1(1),X_2(1))^\intercal=X(0)^\intercal+\psi_{A_{0}}(\theta_1)\left(\begin{array}{c}\cos(2\pi W_1)\\ \sin(2\pi W_1)\end{array}\right), \end{array}\right. \end{align*} where \begin{equation} \label{defdepsi} \psi_a(t)=\sqrt{2t\log\frac{a}{t}},\quad t\le a, \ a>0. \end{equation} At the end of this step we set $A_1=\gamma^2\rho(X(1))^2 e/2$.\\ \mathversion{bold} {\bf The $n$-th step:}\mathversion{normal} While $X (n-1)\in D^\varepsilon$, simulate $(U_n,V_n,W_n)$ a vector of three independent random variables uniformly distributed on $[0,1]$ and define \begin{align}\label{eq:algo-n} \left\{\begin{array}{l} \theta_n=A_{n-1} U_nV_n,\quad \Theta_n=\Theta_{n-1}+\theta_n,\\ [5pt] X(n)^\intercal=(X_1(n),X_2(n))^\intercal=X(n-1)^\intercal+\psi_{A_{n-1}}(\theta_n)\left(\begin{array}{c}\cos(2\pi W_n)\\ \sin(2\pi W_n)\end{array}\right). \end{array}\right. \end{align} At the end of this step we set $A_n=\gamma^2\rho(X(n))^2 e/2$.\\ When $X(n-1)\notin D^\varepsilon$ the algorithm is stopped: we set $\theta_n=0$, $\Theta_n=\Theta_{n-1}$ and $X(n)=X(n-1)$.\\ {\bf Outcome:}\mathversion{normal} The hitting time $\Theta_n$ and the exit position $X(n)$.}\\ \centerline{\line(1,0){490}} \begin{rem} The WoMS algorithm describes a $D$-valued Markov chain $(X(n),\, n\ge 0)$. Each step corresponds to an exit problem for the $2$-dimensional Brownian motion. If $X(n)=x$ then we focus our attention to the exit problem of the ball centered in $x$ and of radius $(\psi_a(t),\, t\ge 0)$: the exit location corresponds to $X(n+1)$ and the exit time to $\theta_{n+1}$. Of course the choice of the parameter $a$ plays a crucial role since the moving sphere has to belong to the domain $D$ as time elapses... When the Markov chain $X$ is close to the boundary $\partial D$, we stop the algorithm and obtain therefore a good approximation of the exit problem of $D$. Comparison with the classical ($WoS$) algorithm: at each step, the $n$-th step of the classical walk on spheres ($WoS$) is based on the exit location and exit time, which are mutually independent, for the Brownian paths exiting from a ball centered in $X(n-1)$ and with radius $\gamma \rho(X(n-1))$. The exit location is uniformly distributed on the sphere while the exit time is characterized by its Laplace transform. Therefore, if one knows $X(n-1)$ then the diameter of the sphere is deterministic. For the $WoMS$ the center of the ball should also be $X(n-1)$ but the radius is random, smaller than $\gamma\rho(X(n-1))$. The exit location will also be uniformly distributed on the sphere but the exit time will be much easier to simulate: in particular, you don't need to evaluate the Bessel functions. \end{rem} The stochastic process $(X(n),\,n\ge 0)$ is an homogeneous Markov chain stopped at the first time it exits from $D^\varepsilon$. In the following, we shall denote $N^\varepsilon$ this stopping time which represents in fact the number of steps in the algorithm: \[ N^\varepsilon=\inf\{n\ge 0;\ X(n)\notin D^\varepsilon\}. \] We just notice that $X(N^\varepsilon)\in D^\varepsilon$ by its definition.\\[5pt] The algorithm (A1) is presented in the $2$-dimensional case. Of course we can construct a generalization of this procedure for the $\delta$-dimensional Bessel process when $\delta\in\mathbb{N}^*$. For notational simplicity, we use a slightly different method: instead of dealing with a Markov chain $(X(n),\, n\in\mathbb{N})$ living in $\mathbb{R}^\delta$ we shall consider its squared norm, which is also (surprisingly) a Markov chain. At each step, we shall construct a couple of random variables $(\xi_n,\,\chi(n))$ associated to an exit problem, the first coordinate corresponds to an exit time and the second one to the norm of the exit location.\\ We introduce some notations: $\mathcal{S}^\delta$ represents the unit ball in $\mathbb{R}^\delta$, and $\pi_1:\mathbb{R}^\delta\to \mathbb{R}$ the projection on the first coordinate. \\ \centerline{\line(1,0){490}}\\ \emph{{\bf Algorithm (A2).} Let us fix a parameter $0<\gamma<1$.\\ {\bf Initialization:} Set $\chi(0)=0$, $\xi_0=0$, $\Xi_0=0$, $A_0=(\gamma^2l^2 e/(\nu+1))^{\nu+1}\frac{\Gamma(\nu+1)}{2}$.\\ {\bf The $n$-th step:} While $\sqrt{\chi(n-1)}<l-\varepsilon$, we choose $U_n$ a uniform distributed random vector on $[0,1]^{\lfloor\nu\rfloor+2}$, $G_n$ a standard gaussian random variable and $V_n$ an uniformly distributed random vector on $\mathcal{S}^\delta$. Consider $U_n$, $G_n$ and $V_n$ independent. We set \begin{align}\label{eq:defalgo} \left\{\begin{array}{l} \xi_n=\Big(\frac{A_{n-1}}{\Gamma(\nu+1)2^\nu}\, U_n(1)\ldots U_n(\lfloor\nu\rfloor+2)\Big)^{\frac{1}{\nu +1}}\exp\Big\{-\frac{\nu-\lfloor\nu\rfloor}{\nu +1}G_n^2\Big\},\quad \Xi_n=\Xi_{n-1}+\xi_n,\\ [5pt] \chi(n)=\chi(n-1)+2\pi_1(V_n)\sqrt{\chi(n-1)}\psi_{A_{n-1}}(\xi_n)+\psi^2_{A_{n-1}}(\xi_n), \end{array}\right. \end{align} where \begin{equation} \label{defdepsi-new} \psi_a(t)=\sqrt{2t\log\frac{a}{\Gamma(\nu+1)t^{\nu+1}2^\nu}},\quad t\le t_{\rm max}(a):=\left[\frac{a}{\Gamma(\nu+1)2^\nu}\right]^{\frac{1}{\nu+1}}, \ a>0. \end{equation} At the end of this step we set \[ A_n=\Big(\gamma^2(l-\sqrt{\chi(n)})^2 e/(\nu+1)\Big)^{\nu+1}\frac{\Gamma(\nu+1)}{2}. \] When $\sqrt{\chi(n)}\ge l-\varepsilon$ the algorithm is stopped: we then set $\xi_n=0$, $\Xi_n=\Xi_{n-1}$ and $\chi(n)=\chi(n-1)$.\\ {\bf Outcome:} The hitting time $\Xi_n$ and the value of the Markov chain $\chi(n)$.}\\ \centerline{\line(1,0){490}} \\ It is obvious that for the particular dimension $\delta=2$, that is $\nu=0$, the stopping times obtained by the algorithm (A1) and (A2) have the same distribution. Moreover, for each $n$, $\chi(n)$ has the same distribution as $\Vert X(n) \Vert^2$. In other words, the number of steps of (A1) and (A2) are identical in law, the number of steps will be denoted in both cases $N^\varepsilon$. \begin{thm} \label{thm:algo} Set $\delta \in\mathbb{N}^*$. The number of steps $N^\varepsilon$ of the algorithm $WoMS$ (A2) is almost surely finite. Moreover, there exist a constant $C_\delta>0$ and $\varepsilon_0(\delta)>0$, such that \[ \mathbb{E}[N^\varepsilon]\le C_\delta|\log\varepsilon|, \mbox{ for all }\,\, \varepsilon\le\varepsilon_0(\delta). \] \end{thm} \begin{thm} \label{thm:algo-conv} Set $\delta\in\mathbb{N}^*$. As $\varepsilon$ goes to zero, $\Xi_{N^\varepsilon}$ converges in distribution towards $\tau_l$ the hitting time of the $\delta$-dimensional Bessel process (with cumulative distribution function $F$), which is almost surely finite. Moreover, for any $\alpha>0$ small enough, \begin{equation}\label{eq:thm:encadr} \Big(1-\frac{\varepsilon}{\sqrt{2\alpha\pi}}\Big)F^\varepsilon(t-\alpha)\le F(t)\le F^\varepsilon(t),\mbox { for all } \, t>0, \end{equation} where $F^\varepsilon(t):=\mathbb{P}(\Xi_{N^\varepsilon}\le t)$. \end{thm} These results and the key ideas of the proofs are adapted from the classical random walk on spheres ($WoS$), see \cite{Milstein-97}.\\ \noindent \emph{Proof of Theorem \ref{thm:algo}}.\\ \emph{Step 1.} Let us estimate the number of steps. Since $(\chi(n),\, n\ge 0)$ is a homogeneous Markov chain, we introduce the operator $P_xf$ defined, for any non-negative function $f:\mathbb{R}_+\to\mathbb{R}_+$, by \[ P_xf:=\int_{\mathbb{R}_+}f(y)\mathbb{P}(x, \mathrm{d} y), \] where $\mathbb{P}(x,\mathrm{d} y)$ is the transition probability of the Markov chain. By definition, $\chi(n+1)$ depends only on $\chi(n)$, $V_n$ and $\xi_n$. Let us note that, by construction, $V_n$ and $\xi_n$ are independent. Moreover using the result developed in the Appendix, the density of $\xi_n\Big(\frac{2^\nu\Gamma(\nu+1)}{A_{n-1}}\Big)^{\frac{1}{\nu+1}}$ is given by: \begin{equation}\label{eq:dens} \mu(r)=\frac{(\nu +1)^{\nu+2}}{\Gamma(\nu+2)}\ r^\nu\Big(-\log r\Big)^{\nu+1}\ind{[0,1]}(r). \end{equation} If we denote $\sigma^d$ the uniform surface measure on the unit sphere in $\mathbb{R}^d$, we get \begin{align}\label{eq:pxf} P_xf&=\int_{0}^1\int_{\mathcal{S^\delta}}f\Big(x+2\pi_1(u)\sqrt{x}K(x,r)+K^2(x,r)\Big)\mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u), \end{align} with $K(x,r)$ defined by \begin{equation}\label{eq:K} K(x,r)=\psi_{A}\Big(\Big[\frac{A}{2^\nu\Gamma(\nu+1)}\Big]^{\frac{1}{\nu+1}}r\Big), \end{equation} and $A$ depending on $x$ in the following way: \[ A=\Big(\frac{\gamma^2(l-\sqrt{x})^2 e}{\nu+1}\Big)^{\nu+1}\frac{\Gamma(\nu+1)}{2}. \] We can observe the following scaling property: $\psi_A(A^{\frac{1}{\nu+1}}t)=A^{\frac{1}{2\nu+2}}\psi_1(t)$. Therefore the definition of $\psi_1$ leads to \begin{equation} \label{eq:defK} K(x,r)=\gamma(l-\sqrt{x})\sqrt{\frac{er}{\nu+1}\log\frac{1}{r^{\nu+1}}}=\gamma(l-\sqrt{x})\sqrt{er(-\log r)}. \end{equation} \par\noindent\emph{ Step 2.} Using classical potential theory for discrete time Markov chains (see, for instance, Theorem 4.2.3 in \cite{Norris-97}), we know that \[ \phi(x)=\mathbb{E}_x\left(\sum_{n=0}^{N^\varepsilon-1}g(\chi(n))\right) \] satisfies, for any non-negative function $g$, \begin{equation} \label{eq:equat} \left\{ \begin{array}{l} \phi(x)=P_x\phi+g(x),\quad 0\le x < (l-\varepsilon)^2,\\ \phi((l- \varepsilon)^2)=0.\end{array}\right. \end{equation} In particular, for $g=1$, we obtain that $\phi(x)=\mathbb{E}_x[N^\varepsilon]$. In order to get an upper-bound for the averaged number of steps, it suffices to apply a comparison result. We choose the function \begin{equation} U^\varepsilon(x)=\{\log((l-\sqrt{x})/\varepsilon)-\log(1-\gamma)\}/(C_\delta\gamma^2) \end{equation} which satisfies $U^\varepsilon(x)\ge P_xU^\varepsilon+1$, for all $0<x<(l-\varepsilon)^2$ (see Lemma \ref{lem:ineg-log} for the definition of the constant and for the inequality) and $U^\varepsilon(x)\ge0$ for all $0<x<(l-\varepsilon)^2$. A classical comparison result related to the potential theory (see, for instance, Theorem 4.2.3 in \cite{Norris-97}) implies that $\mathbb{E}_x[N^\varepsilon]\le U^\varepsilon(x)$ for all $x\in [0,(l-\varepsilon)^2]$ and consequently leads to the announced statement.\hfill{$\Box$} \begin{lemma} \label{lem:ineg-log} Let us define, for small $\varepsilon >0$, $U^\varepsilon(x)=\{\log((l-\sqrt{x})/\varepsilon)-\log(1-\gamma)\}/(C_\delta\gamma ^2)$ for $x\in [0,l^2[$, where $\gamma$ is related to the definition of the $WoMS$. There exists a constant $C_\delta>0$, such that, for any $x\in]0,(l-\varepsilon)^2[$, the following inequality yields \[ P_xU^\varepsilon-U^\varepsilon(x)\le -1. \] We recall that $P_xU^\varepsilon$ is defined by \eqref{eq:pxf} and \eqref{eq:defK}. \end{lemma} \begin{proof} We will split the proof in several steps. \emph{Step 1.} First of all, we observe that $U^\varepsilon\ge -\log(1-\gamma)/(C_\delta\gamma^2)$ in the domain $[0,(l-\varepsilon)^2]$. Let us consider now $\chi(0)=x\in [0,(l-\varepsilon)^2]$ and $y$ in the support of the law of $\chi(1)$ and let us prove that $U^\varepsilon(y)\ge 0$. By the definition of $\chi(1)$ we obtain \[ \chi(1)\le \sup_{y\in[-1,1], t\in[0,t_{\rm max}(A)]}\ (x+2y\sqrt{x}\psi_{A}(t)+\psi^2_A(t)), \] where $A=\Big(\gamma^2(l-\sqrt{x})^2 e/(\nu+1)\Big)^{\nu+1}\frac{\Gamma(\nu+1)}{2}$ and both $\psi_A$ and $t_{\rm max}$ are defined by \eqref{defdepsi-new}. The right hand side of the preceding inequality is increasing with respect to $y$ so that: \[ \chi(1)\le \Big(\sqrt{x}+\sup_{t\in[0,t_{\rm max}(A)]}\ \psi_A(t)\Big)^2. \] Furthermore, for $a>0$ the maximum of the function $\psi_a$ is reached for $t_{\rm max}(a)=\frac{1}{e}\, \Big( \frac{a}{\Gamma(\nu+1)2^\nu}\Big)^{\frac{1}{\nu+1}}$ and is equal to \begin{equation} \label{eq:maxcalcul} \sup_{t\in[0,t_{\rm max}(a)]}\ \psi_a(t)=\left\{\frac{2(\nu+1)}{e}\left(\frac{a}{\Gamma(\nu+1)2^\nu}\right)^{\frac{1}{\nu+1}}\right\}^{1/2}. \end{equation} Finally using the definition of $A$ and the inequality $x\le(l-\varepsilon)^2$, we find the following lowerbound: \[ l-\sqrt{\chi(1)}\ge (l-\sqrt{x})(1-\gamma)\ge \varepsilon (1-\gamma). \] We can therefore conclude that, for any $y$ in the support of the law of $\chi(1)$ (even for $y\ge (l-\varepsilon)^2$), $U^\varepsilon(y)\ge 0$ which ensures that $U^\varepsilon$ is well defined and non-negative in the domain of the operator $P_x$.\\ \emph{Step 2.} Furthermore the Taylor expansion yields \begin{equation}\label{eq:taylor} U^\varepsilon(y)\le U^\varepsilon(x)+\frac{\sqrt{x}-\sqrt{y}}{C_\delta\gamma^2 (l-\sqrt{x})}-\frac{(\sqrt{x}-\sqrt{y})^2}{2C_\delta\gamma^2(l-\sqrt{x})^2}+\frac{(\sqrt{x}-\sqrt{y})^3}{3C_\delta\gamma^2(l-\sqrt{x})^3},\quad x,\ y\in [0,l^2[. \end{equation} If $\chi(0)=x$ and $y$ is in the support of the random variable $\chi(1)$, then \begin{align*} \sqrt{y}-\sqrt{x}&=\sqrt{x+2\pi_1(u)\sqrt{x}K(x,r)+K^2(x,r)}-\sqrt{x}\\ &\ge \pi_1(u)K(x,r). \end{align*} By the expansion \eqref{eq:taylor} and the definition of the operator $P_x$ given by \eqref{eq:pxf}, the following upper-bound for the operator $P_x$ holds \begin{align*} P_xU^\varepsilon&=\int_{0}^1\int_{\mathcal{S}^\delta}U^\varepsilon\Big(x+2\pi_1(u)\sqrt{x}K(x,r)+K^2(x,r)\Big)\mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u),\\ &\le U^\varepsilon(x)-\int_{0}^1\int_{\mathcal{S}^\delta}\frac{\pi_1(u)K(x,r)}{C_\delta\gamma^2(l-\sqrt{x})}\ \mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u)\\ &\qquad \qquad - \int_{0}^1\int_{S^\delta_+}\frac{\pi_1^2(u)K^2(x,r)}{2C_\delta\gamma^2(l-\sqrt{x})^2}\ \mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u)\\ &\qquad \qquad -\int_{0}^1\int_{\mathcal{S^\delta}}\frac{\pi_1^3(u)K^3(x,r)}{3C_\delta\gamma^2(l-\sqrt{x})^3}\ \mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u), \end{align*} where \begin{equation} \label{S_delta_+} S^\delta_+ : ={\{u\in \mathcal{S}^\delta: \ \pi_1 (u) > 0\}}. \end{equation} Due to symmetry properties, the first and the third integral terms vanish. Then \eqref{eq:defK} leads to \begin{align*} P_xU^\varepsilon&\le U^\varepsilon(x)- \frac{I}{C_\delta}\ \int_{S^\delta_+} \pi_1^2(u)\ \sigma_1^\delta(\mathrm{d} u) \end{align*} with \[ I=\frac{(\nu +1)^{\nu+2}e}{2\Gamma(\nu+2)}\int_0^1r^{\nu+1}(-\log r)^{\nu+2}\,\mathrm{d} r. \] The description of the probability density function in the Appendix leads to the following explicit value \[ I=\left( \frac{\nu+1}{\nu+2} \right)^{\nu+2}\frac{e}{\Gamma(\nu+2)}. \] In order to conclude the proof, it suffices to choose the particular constant \[ C_\delta=I\ \int_{S^\delta_+} \pi_1^2(u)\ \sigma_1^\delta(\mathrm{d} u). \] \end{proof} \noindent\emph{Proof of Theorem \ref{thm:algo-conv}}\\ The proof is splited in two parts. First the steps of the algorithm and the hitting time of the Bessel process of index $\nu$ shall be related to stopping times of a $\delta$-dimensional Brownian motion ($\nu=\frac{\delta}{2}-1$). Secondly we point out that the corresponding stopping times are close together by evaluating deviations of the Brownian paths.\\ {\emph { Step 1.}} Let ${\bf B}=(B^{(1)},B^{(2)},\ldots,B^{(\delta)})$ be a $\delta$-dimensional Brownian motion. Then the norm of ${\bf B}$ has the same distribution as a Bessel process of index $\nu$ (see, for instance \cite{Revuz-Yor-99}). Hence the first hitting time $\tau_l$ is identical in law to the stopping time \[ \mathbb{T}_l=\inf\{t\ge 0;\ {\bf B}_t\notin D \}, \] where $D$ is the Euclidean ball centered at the origin and of radius $l$. We introduce then a procedure in order to come close to $\mathbb{T}_l$. For the first step we shall focus our attention to the first exit time of a moving sphere centered at the origin and of radius $\psi_a(t)$ defined by \eqref{defdepsi-new}, we denote $\hat{\xi}_1$ this stopping time. Of course this moving sphere should always stay in $D$, so we choose $a$ such that the maximum of $\psi_a$ stays smaller than $l$. By \eqref{eq:maxcalcul}, we get \[ \sup_{t\le a}\psi_a(t)< l\Longleftrightarrow a< \frac{\Gamma(\nu+1)}{2}\left(\frac{el^2}{\nu+1}\right)^{\nu+1}. \] For $a=A_0=\frac{\Gamma(\nu+1)}{2}\left(\frac{e\gamma^2l^2}{\nu+1}\right)^{\nu+1}$ with a parameter $\gamma<1$, the condition is satisfied: $\sup_{t\le a}\psi_a(t)=\gamma^{2\nu+2} l<l$. Let us describe the law of $(\hat{\xi}_1,{\bf B}_{\hat{\xi_1}})$. The norm of the Brownian motion is identical in law with the Bessel process; therefore Proposition \ref{prop:first_boundary} implies that the density function of $\hat{\xi}_1$ is given by \eqref{densite_tau} with $a$ replaced by $A_0$. Using the law described in Proposition \ref{prop:append}, we can prove that $\hat{\xi}_1$ has the same distribution as \[ \left(\frac{ A_0}{\Gamma(\nu+1)2^\nu} \right)^{\frac{1}{\nu+1}} \ e^{-Z}, \] where $Z$ is Gamma distributed with parameters $\alpha=\nu+2$ and $\beta=\frac{1}{\nu+1}$. By construction we deduce that $\hat{\xi}_1 \stackrel{(d)}{=} \xi_1$ where $\xi_1$ is defined in the algorithm $WoMS$ (A2). Knowing the stopping time, we can easily describe the exit location since the Brownian motion is rotationnaly invariant: ${\bf B}_{\hat{\xi}_1}$ is then uniformly distributed on the sphere of radius $\psi_{A_0}(\hat{\xi}_1)$. Hence \[ (\hat{\xi}_1,\Vert{\bf B}_{\hat{\xi_1}}\Vert)\stackrel{(d)}{=}(\xi_1,\chi(1)),\quad\mbox{and}\ \hat{\xi}_1<\mathbb{T}_l. \] By this procedure we can construct a sequence of stopping times $(\hat{\xi}_n,\, n\ge 1)$ and define $\hat{\Xi}_n=\hat{\xi}_1+\ldots+\hat{\xi}_n$; $\hat{\Xi}_n$ is the first time after $\hat{\Xi}_{n-1}$ such that the Brownian motion exits from a sphere centered in ${\bf B}_{\hat{\Xi}_{n-1}}$ of radius $\psi_{a_n}$ initialized at time $\hat{\Xi}_{n-1}$. The moving sphere should stay in the domain $D$, so we choose \[ a_n=\Big(\gamma^2\Big(l-\sqrt{{\bf B}_{\hat{\Xi}_{n-1}}}\Big)^2 e/(\nu+1)\Big)^{\nu+1}\frac{\Gamma(\nu+1)}{2}. \] \begin{figure}[!h] \psfrag{D}{$\partial D$} \psfrag{n}{${\bf B}_{\hat{\Theta}_n}$} \psfrag{n+1}{${\bf B}_{\hat{\Theta}_{n+1}}$} \psfrag{n+2}{${\bf B}_{\hat{\Theta}_{n+2}}$} \psfrag{n+3}{${\bf B}_{\hat{\Theta}_{n+3}}$} \centerline{\includegraphics[width=8cm]{WoMS.eps}} \caption{Walk on Moving Spheres} \end{figure} Using the same arguments as before and by the Markov property for the Brownian motion, we obtain the identities in law \[ (a_n,\, n\ge 1)\stackrel{(d)}{=} (A_n,\, n\ge 1),\quad \Big(\hat{\Xi}_n, \Vert{\bf B}_{\hat{\Xi}_n}\Vert\Big)_{n\ge 1}\stackrel{(d)}{=}\Big(\Xi_n, \chi(n)\Big)_{n\ge 1} \] with $\hat{\Xi}_n<\mathbb{T}_l$ and $\Xi_n$, $A_n$, $\chi(n)$ defined in the algorithm $WoMS$ (A2). Consequently defining $\hat{N}^\varepsilon=\inf\{ n\ge 0; {\bf B}_{\hat{\Xi}_n}\notin D^\varepsilon\}$, the following identity yields \begin{equation}\label{eq:idenlaw} \Big( \hat{\Xi}_{\hat{N}^\varepsilon}, \Vert{\bf B}_{\hat{\Xi}_{\hat{N}^\varepsilon}} \Vert\Big)\stackrel{(d)}{=}\Big( \Xi_{N^\varepsilon}, \chi(N^\varepsilon) \Big)\quad \mbox{and}\quad \hat{\Xi}_{\hat{N}^\varepsilon}<\mathbb{T}_l. \end{equation} \emph{ Step 2.} Let us now estimate the difference between $\hat{\Xi}_{\hat{N}^\varepsilon}$ and $\mathbb{T}_l$. By \eqref{eq:idenlaw} we first deduce \begin{equation}\label{eq:upper-bound} F(t):=\mathbb{P}(\tau_l\le t)=\mathbb{P}(\mathbb{T}_l\le t)\le F^\varepsilon(t):=\mathbb{P}(\Theta_{N^\varepsilon}\le t),\quad t>0. \end{equation} Furthermore, for any small $\alpha>0$, \begin{align}\label{eq:decomp-under} 1-F(t)&=\mathbb{P}( \mathbb{T}_l> t , \,\hat{\Xi}_{\hat{N}^\varepsilon}\le t-\alpha )+\mathbb{P}( \mathbb{T}_l> t , \,\hat{\Xi}_{\hat{N}^\varepsilon}> t-\alpha )\nonumber\\ &\le\mathbb{P}( \mathbb{T}_l> t , \,\hat{\Xi}_{\hat{N}^\varepsilon}\le t-\alpha )+\mathbb{P}(\hat{\Xi}_{\hat{N}^\varepsilon}> t-\alpha )\nonumber\\ &\le\mathbb{P}( \mathbb{T}_l> t , \,\hat{\Xi}_{\hat{N}^\varepsilon}\le t-\alpha )+1-F^\varepsilon(t-\alpha). \end{align} At time $\hat{\Xi}_{\hat{N}^\varepsilon}$ the Brownian motion is in the $\varepsilon$-neighborhood of the boundary $\partial D$, hence \\ $l~-~\Vert{\bf B}_{\hat{\Xi}_{\hat{N}^\varepsilon}}\Vert ~\le~\varepsilon$. Using the strong Markov property, we obtain \begin{equation}\label{eq:interm} \mathbb{P}( \mathbb{T}_l> t , \,\hat{\Xi}_{\hat{N}^\varepsilon}\le t-\alpha )\le F^\varepsilon(t-\alpha)\sup_{y\in\partial D^\varepsilon}\mathbb{P}_y(\mathbb{T}_l>\alpha). \end{equation} Since the Brownian motion is rotationnaly invariant, it suffices to choose $y=(l-\varepsilon,0,\ldots,0)$. Due to the convexity of $\partial D$, the following upper-bound holds \begin{equation}\label{eq:lowboununif} \mathbb{P}_y(\mathbb{T}_l>\alpha)\le \mathbb{P}_0(\sup_{0\le t\le \alpha}\overline{B}_t^{(1)}<\varepsilon)=\mathbb{P}_0(2\vert \overline{B}_\alpha^{(1)} \vert <\varepsilon)\le \frac{\varepsilon}{\sqrt{2\alpha\pi}}. \end{equation} Combining \eqref{eq:upper-bound} for the upper-bound and \eqref{eq:decomp-under}, \eqref{eq:interm} and \eqref{eq:lowboununif} for the lower-bound yields the announced estimation \eqref{eq:thm:encadr}.\hfill{$\Box$} \mathversion{bold} \subsubsection{The first time the Bessel process of index $\nu$ hits a decreasing curved boundary} \mathversion{normal} The algorithm developed in the previous paragraph can be adapted to the problem of hitting a deacreasing curved boundary. Let us define : \begin{equation} \label{def:tau-line} \tau=\inf\{t\ge 0: \, Z_t^{\delta,0}=l(t)\},\quad \mbox{ where } l \ \mbox{is decreasing and }\ l(0)>0. \end{equation} \begin{assu}\label{assu} There exists a constant $\Delta_{{\rm min}}>0$ which bounds the derivative of $l$ \[ l'(t)\ge -\Delta_{\rm min},\quad \forall t\ge 0. \] \end{assu} The procedure then also consists in building a $WoMS$ which reaches a neighborhood of the boundary. But instead of dealing with a fixed boundary as in Section \ref{sect:0-horiz}, that is a ball of radius $l$, we shall in this section introduce the following moving boundary: the ball centered in the origin and of radius $l(t)$. The arguments developed in order to prove Theorem \ref{thm:algo} and Theorem \ref{thm:algo-conv} will be adapted to this new context. \\ \centerline{\line(1,0){490}}\\ \emph{{\bf Algorithm (A3):}\\ Let us define the following positive constants \begin{equation} \label{constante} L=\max\Big(l(0),\Delta_{\rm min}, \sqrt{\nu+1}\Big),\quad \kappa=\frac{2^\nu}{5^{\nu+1}L^{2\nu+2}}\,\Gamma(\nu+1). \end{equation} {\bf Initialization:} Set $\chi(0)=0$, $\xi_0=0$, $\Xi_0=0$, $A_0=\kappa\Big( l(0)- \sqrt{\chi(0)}\Big)^{2(\nu+1)}$.\\ \mathversion{bold} {\bf The $n$-th step:}\mathversion{normal} While the following condition holds \begin{equation*} l(\Xi_{n-1})-\sqrt{\chi(n-1)}>\varepsilon \end{equation*} (denoted by $\mathbb{C}(n-1)$), we simulate $U_n$ an uniform distributed random vector on $[0,1]^{\lfloor\nu\rfloor+2}$, $G_n$ a standard gaussian random variable and $V_n$ a uniformly distributed random vector on $\mathcal{S}^\delta$. $U_n$, $G_n$ and $V_n$ have to be independent. We then construct $(\xi_n,\Xi_n,\chi(n))$ using \eqref{eq:defalgo}. At the end of this step we set $A_n=\kappa \Big( l(\Xi_n)-\sqrt{\chi(n)}\Big)^{2(\nu+1)}$.\\ The algorithm stops when $\mathbb{C}(n-1)$ is not longer satisfied: we set $\xi_n=0$ and so $\Xi_n=\Xi_{n-1}$ and $\chi(n)=\chi(n-1)$.\\ {\bf Outcome} The exit position $\chi(n)$ and the exit time.}\\ \centerline{\line(1,0){490}}\\[5pt] Let us note that the stochastic process $(\chi(n),\,n\ge 0)$ is not a Markov chain since the sequence $(A_n)_{n\ge 0}$ depends on both $(\Xi_n, \chi(n))$. That is why we define the following Markov chain \begin{equation*} R_n:=(\Xi_n,\chi(n))\in \mathbb{R}_+^2 \end{equation*} stopped at the first time the condition $\mathbb{C}(n)$ is not satisfied. In the following, we shall denote $N^\varepsilon$ this stopping time (number of steps of the algorithm): \[ N^\varepsilon=\inf\{n\ge 0;\ l(\Xi_n)-\sqrt{\chi(n)}\le \varepsilon\}. \] \begin{thm} \label{thm:algo:line} The number of steps $N^\varepsilon$ of the algorithm $WoMS$ $(A3)$ is almost surely finite. Moreover, there exist a constant $C_\delta>0$ and $\varepsilon_0(\delta)>0$, such that \[ \mathbb{E}[N^\varepsilon]\le C_\delta |\log\varepsilon|, \mbox{ for all }\,\, \varepsilon\le\varepsilon_0(\delta). \] \end{thm} \begin{thm} \label{thm:algo-conv:line} As $\varepsilon$ goes to zero, $\Xi_{N^\varepsilon}$ converges in distribution towards $\tau$ defined by \eqref{def:tau-line} (with cumulative distribution function $F$), which is almost surely finite. Moreover, for any $\alpha>0$ small enough, \begin{equation}\label{eq:thm:encadr:line} \Big(1-\frac{\varepsilon}{\sqrt{2\alpha\pi}}\Big)F^\varepsilon(t-\alpha)\le F(t)\le F^\varepsilon(t),\mbox { for all } \, t>0, \end{equation} where $F^\varepsilon(t):=\mathbb{P}(\Xi_{N^\varepsilon}\le t)$. \end{thm} \noindent \emph{Proof of Theorem \ref{thm:algo:line}}.\\ The proof is based mainly on arguments already presented in Theorem \ref{thm:algo}. So we let the details of the proof to the reader and focus our attention to the main ideas.\\ {\bf (1)} The process $(\Xi_n, \chi(n))$ is a homogeneous Markov chain and the associated operator is given by \begin{equation}\label{eq:new-operator} P_{t,x}f:=\int_{(s,y)\in \mathbb{R}_+^2}f(s,y)\mathbb{P}\Big((t,x),(\mathrm{d} s,\mathrm{d} y)\Big), \end{equation} where $f$ is a non-negative function and $\mathbb{P}\Big((t,x),(\mathrm{d} s,\mathrm{d} y)\Big)$ is the transition probability of the chain. The chain starts with $(\Xi_0,\chi(0))=(0,0)$ and is stopped the first time when $l(\Xi_n)-\sqrt{\chi(n)}\le \varepsilon$. Classical potential theory ensures that \[ \phi(t,x)=\mathbb{E}_{t,x}\left( \sum_{n=0}^{N^\varepsilon-1}g(\Xi_n,\chi(n)) \right) \] is solution of the following equation \begin{align}\label{eq:sys:line} \left\{\begin{array}{l} \phi(t,x)=P_{t,x}\phi+g(t,x),\quad (t,x)\in D^\varepsilon,\\[5pt] \phi(t,x)=0, \quad \forall (t,x)\in\partial D^\varepsilon, \end{array}\right. \end{align} with $D^\varepsilon=\{(t,x)\in\mathbb{R}_+^2:\ l(t)-\sqrt{x}\le \varepsilon\}$. For the particular choice $g=1$, we obtain $\phi(t,x)=\mathbb{E}_{t,x}[N^\varepsilon]$, therefore the averaged number of step is given by $\phi(0,0)$. \\ {\bf (2)} In order to point out an upper-bound for the averaged number of steps, we use a comparison result: we are looking for a function $U(t,x)$ such that \begin{align}\label{sysavec1} \left\{\begin{array}{l} U(t,x)\ge P_{t,x}U+1,\quad \forall (t,x)\in D^\varepsilon\\[5pt] U(t,x)\ge 0, \quad \forall (t,x)\in\partial D^\varepsilon. \end{array}\right. \end{align} For such a particular function, we can deduce $\phi(t,x)\le U(t,x)$. Let us define \[ U(t,x)=c\log\left( \frac{l(t)-\sqrt{ x}}{\varepsilon} \right)1_{\{ l(t)-\sqrt{x} \ge 0 \}}, \] with some constant $c>0$ which shall be specified later on. The positivity assumption on the boundary $\partial D^ \varepsilon$ is trivial. Moreover since $l$ is a decreasing function, \eqref{eq:new-operator} implies \begin{align}\label{eq:ineg-inte} P_{t,x}U&=\int_{(s,y)\in\mathbb{R}_+^2}U(s,y)\mathbb{P}\Big((t,x),(\mathrm{d} s,\mathrm{d} y)\Big)\le \int_{(s,y)\in \mathbb{R}_+^2}U(t,y)\mathbb{P}\Big((t,x),(\mathrm{d} s,\mathrm{d} y)\Big). \end{align} By using the Taylor expansion, we get \begin{equation}\label{eq:taylor-} U(t,y)\le U(t,x)-c\ \frac{\sqrt{y}-\sqrt{x}}{l(t)-\sqrt{x}}-\frac{c}{2}\frac{(\sqrt{y}-\sqrt{x})^2}{(l(t)-\sqrt{x})^2}-\frac{c}{3}\frac{(\sqrt{y}-\sqrt{x})^3}{(l(t)-\sqrt{x})^3}, \quad (x,y)\in \mathbb{R}_+^2. \end{equation} Using similar arguments and similar bounds as those presented in Lemma \ref{lem:ineg-log}, the odd powers in the Taylor expansion don't play any role in the integral \eqref{eq:ineg-inte}. Therefore we obtain \begin{align*} P_{t,x}U&\le U(t,x)-\frac{c}{2} \int_{(s,y)\in\mathbb{R}_+^2}\frac{(\sqrt{y}-\sqrt{x})^2}{(l(t)-\sqrt{x})^2}\mathbb{P}\Big((t,x),(\mathrm{d} s,\mathrm{d} y)\Big)\\ &\le U(t,x)- \frac{c}{2}\int_{0}^1\int_{S^\delta_+}\frac{\pi_1^2(u)K^2(x,r)}{(l(t)-\sqrt{x})^2}\ \mu(r)\mathrm{d} r\,\sigma_{1}^\delta(\mathrm{d} u) \end{align*} where $S^\delta_+$ is given in \eqref{S_delta_+} and $K$ is defined by \eqref{eq:K} with $A=\kappa (l(s)-\sqrt{x})^{2(\nu+1)}$. We have now \[ P_{t,x}U\le U(t,x)-\frac{c(\nu+1)}{2}\Big( \frac{2K}{\Gamma(\nu+1)} \Big)^{\frac{1}{\nu+1}}\left( \int_{\mathcal{S}^\delta_+} \pi_1^2(u)\,\sigma_{1}^\delta(\mathrm{d} u)\right)\left(\int_0^1 r(-\log r)\mu(r)\mathrm{d} r\right). \] An appropriate choice of the constant $c$ leads to \eqref{sysavec1}. Finally we get \[ \mathbb{E}[N^\varepsilon]\le U(0,0)= c\,\log(l(0)/\varepsilon). \] \hfill{$\Box$}\\[5pt] \noindent \emph{Proof of Theorem \ref{thm:algo-conv:line}}.\\ The arguments are similar to those developped for Theorem \ref{thm:algo-conv}, the extension of the convergence result to curved boundaries is straightforward. That is why we shall not repeat the proof but just focus our attention on the only point which is quite different. We need to prove that the Markov chain $R_n=(\Xi_n,\chi(n))$ stays in the domain $D^0=\{(t,x):\ 0\le x\le l^2(t) \}$ so that the hitting time $\tau$ defined by \eqref{def:tau-line} satisfies $\tau>\Xi_{N^\varepsilon}$. In other words, if the Markov chain $R_n=(\Xi_n,\chi(n))$ for the $n$-th step is equal to $(s,x)$, then $R_{n+1}$ should belong to $\{(t,x):\ t\ge s, \ x\le l^2(t)\}$. In the $WoMS$ setting, for $t\ge s$, this means that the ball centered in $x$ and of time-dependent radius $\psi_A(t-s)$ always belongs as time elapses to the ball centered in $0$ of radius $l(t)$. We recall that \[ A=\kappa(l(s)-\sqrt{x})^{2(\nu+1)}. \] Therefore we shall prove that \begin{equation} \label{eq:amontrer} \forall t\ge s,\quad \psi_A(t-s)+\sqrt{x}\le l(t). \end{equation} In fact, due to Assumption \ref{assu} and the definition of $\psi_A$, it suffices to obtain \begin{equation}\label{eprou} \psi_A(t-s)\le l(s)-\sqrt{x}-\Delta_{{\rm min}}(t-s),\quad \forall s\le t\le s+W^2, \end{equation} where \[ W=\left(\frac{A}{\Gamma(\nu+1)2^\nu}\right)^{\frac{1}{2\nu+2}}=\left(\frac{\kappa}{\Gamma(\nu+1)2^\nu}\right)^{\frac{1}{2\nu+2}}(l(s)-\sqrt{x})=\frac{1}{L\sqrt{5}}\ (l(s)-\sqrt{x}). \] Due to the definition of the constant $L$, we have \[ 0\le W\le\frac{1}{2\Delta_{{\rm min}}} \frac{2(l(s)-\sqrt{x})\Delta_{\rm min}}{\sqrt{\frac{2\nu+2}{e}+4(l(s)-\sqrt{x})\Delta_{\rm min}}}\le \frac{1}{2\Delta_{{\rm min}}}\left\{\sqrt{ \frac{2\nu+2}{e}+4(l(s)-\sqrt{x})\Delta_{\rm min}} -\sqrt{ \frac{2\nu+2}{e} }\right\}. \] The right hand side of the preceding inequality is the positive root of the polynomial function $P(X)=\Delta_{\rm min}\,X^2+\sqrt{2(\nu+1)/e} X-(l(s)-\sqrt{x})$. We deduce that $P(W)\le 0$. By \eqref{eq:maxcalcul} and $P(W)\le 0$, we obtain \begin{align*} \sup_{t\ge s}\psi_A(t-s) &=\left( \frac{2(\nu+1)}{e} \right)^{1/2}\ W\\ &\le l(s)-\sqrt{x} -\Delta_{\rm min}\,W^2\\ &\le l(s)-\sqrt{x}-\Delta_{{\rm min}}(t-s),\quad \forall s\le t\le s+W^2. \end{align*} Finally we have proved \eqref{eprou} and so \eqref{eq:amontrer}. \hfill{$\Box$}\\ If Assumption \ref{assu} is not satisfied then it is difficult to have a general description of an iterated procedure in order to simulate hitting times. However the particular form of the function $\psi_a$ defined by \eqref{defdepsi-new} permits to describe a $WoMS$ algorithm for the square root boundaries. Let us therefore consider the following functions: \begin{equation}\label{eq:defde2} \psi_a(t)=\sqrt{2t\log\frac{a}{\Gamma(\nu+1)t^{\nu+1}2^\nu}}\quad \mbox{and}\quad f(t)=\sqrt{r-ut}, \end{equation} well defined for $t\le t_0:=\min\Big( \alpha^{\frac{1}{\nu+1}},\frac{r}{u} \Big)$ where $\alpha=a(\Gamma(\nu+1)2^\nu)^{-1}$.\\ The algorithm is essentially based on the following result (the constants $r$ and $u$ associated with the hitting problem of a square root boundary for the Bessel process shall be specified in the proof of Proposition \ref{prop:curved}) \begin{lemma} \label{lem:comparai} Let us define \begin{equation} \label{eq:defdeF} F_\nu(r,u)=\frac{1}{2}\left(\frac{er}{\nu+1}\right)^{\nu+1}\Gamma(\nu+1)\ e^{-u/2},\quad r>0,\ u>0. \end{equation} If $a=F_\nu(r,u)$ then \begin{equation}\label{comparai} \psi_a(t)\le f(t),\quad \mbox{for all}\ \ 0\le t\le \alpha^{\frac{1}{\nu+1}}. \end{equation} \end{lemma} \begin{proof} We are looking for a particular value $a$ depending on both $r$ and $u$ such that the following bound holds: $\psi_a(t)\le f(t)$, for all $0\le t\le t_0$. Since $t\le t_0$, it suffices to prove that \[ 2t\log\frac{\alpha}{t^{\nu+1}}\le r-ut \Longleftrightarrow g(t):=t\Big( 2\log \frac{\alpha}{t^{\nu+1}}+u\Big)\le r. \] Let us compute the maximum of the function $g$ on the interval $[0,t_0]$, with $t_0$ fixed, \[ g'(t)=2\log \frac{\alpha}{t^{\nu+1}}+u-2(\nu+1). \] We have \[ g'(t)=0 \Longleftrightarrow \log \frac{\alpha}{t^{\nu+1}}=\nu+1-\frac{u}{2}\Longleftrightarrow t^{\nu+1}=\alpha\exp\Big\{ \frac{u}{2}-\nu-1 \Big\}. \] In other words the maximum of the function $g$ is reached for $$ t_{max}=\alpha^{\frac{1}{\nu+1}}\exp\Big\{ \frac{u}{2(\nu+1)}-1 \Big\} $$ and is equal to $$ g(t_{max})=g_{\max}=2(\nu+1)\alpha^{\frac{1}{\nu+1}}e^{\frac{u}{2(\nu+1)}-1}. $$ Choosing $g_{\max}\le r$ we obtain in particular \eqref{comparai} that means $$ \alpha\le\left(\frac{er}{2(\nu+1)}\right)^{\nu+1}e^{-u/2}\Longleftrightarrow a\le \frac{1}{2}\left(\frac{er}{\nu+1}\right)^{\nu+1}\Gamma(\nu+1)e^{-u/2}. $$ For $a_0= \frac{1}{2}\left(\frac{er}{\nu+1}\right)^{\nu+1}\Gamma(\nu+1)e^{-u/2}$, we get \eqref{comparai} since $t_0=\alpha^{\frac{1}{\nu+1}}$. \end{proof} The aim is now to construct an algorithm which permits to approximate the hitting time of the square root boundary. Therefore we consider a Bessel process of dimension $\delta$ which hits the decreasing curved boundary $f(t)$ given by \eqref{eq:defde2}.\\[5pt] \centerline{\line(1,0){490}}\\ \emph{ \mathversion{bold} {\bf Algorithm (A4) -- the square root boundary: $l(t)=\sqrt{\beta_0-\beta_1 t}$ with $\beta_0>0$, $\beta_1>0$.}\\ \mathversion{normal} Let $\kappa\in]0,1[$.\\ {\bf Initialization:} Set $\chi(0)=0$, $\xi_0=0$, $\Xi_0=0$, $A_0=\kappa F_\nu(\beta_0,\beta_1)$.\\ \mathversion{bold} {\bf The $(n+1)$-th step :} \mathversion{normal} While the following condition holds \begin{equation*} l(\Xi_{n})-\sqrt{\chi(n)}>\varepsilon \mbox {(denoted by } \mathbb{C}(n) ) , \end{equation*} we define \begin{equation}\label{eq:defAn} A_n=\kappa F_\nu\Big((l(\Xi_n)-\sqrt{\chi(n)})^2, \beta_1\Big(1-\frac{\sqrt{\chi(n)}}{l(\Xi_n)}\Big) \Big) \end{equation} where $F_\nu$ is defined by \eqref{eq:defdeF}, and we simulate $U_{n+1}$ an uniform distributed random vector on $[0,1]^{\lfloor\nu\rfloor+2}$, $G_{n+1}$ a standard gaussian random variable and $V_{n+1}$ a uniformly distributed random vector on $\mathcal{S}^\delta$. $U_{n+1}$, $G_{n+1}$ and $V_{n+1}$ have to be independent. We then construct $(\xi_{n+1},\Xi_{n+1},\chi(n+1))$ using \eqref{eq:defalgo}. \\ The algorithm stops when $\mathbb{C}(n)$ is not longer satisfied : we set $\xi_{n+1}=0$ and so $\Xi_{n+1}=\Xi_{n}$ and $\chi(n+1)=\chi(n)$.}\\ \centerline{\line(1,0){490}} \begin{proposition} \label{prop:curved} The statements of Theorem \ref{thm:algo:line} and Theorem \ref{thm:algo-conv:line} are true for the Algorithm (A4) associated with the square root boundary. \end{proposition} \begin{proof} All the arguments developped for decreasing boundaries with lower-bounded derivatives are easily adapted to the square root boundary. We let the details to the reader and focus our attention to the following fact: the stochastic process $(\Xi_n,\chi(n),\, n\ge 0)$ stays in the domain $D^0$ defined by \[ D^0=\{ (t,x)\in\mathbb{R}_+^2:\ l(t)-\sqrt{x}>0\}. \] In the $WoMS$ setting, for $t\ge s$, this means that for $(\Xi_n,\chi(n))=(s,x)\in D^0 $ the following step leads to $\sqrt{\chi(n+1)}<l(\Xi_{n+1})$. By \eqref{eq:defalgo}, it suffices to prove that \begin{equation}\label{eq:a-dem} \sqrt{x}+\psi_{A}(t)<l(s+t),\quad\mbox{for all}\ t\in\{u\ge 0:\min(l(s+u),\psi_A(u))\ge 0\}, \end{equation} with $ A=\kappa F_\nu\Big( (l(s)-\sqrt{x})^2,\beta_1\Big(1-\frac{\sqrt{x}}{l(s)}\Big) \Big)$, since $\chi(n+1)\le (\sqrt{\chi(n)}+\psi_{A_n}(\xi_{n+1}))^2$. By Lemma \ref{lem:comparai} and due to the coefficient $\kappa$, we have \[ \psi_A(t)<\sqrt{(l(s)-\sqrt{x})^2-\beta_1\Big(1-\frac{\sqrt{x}}{l(s)}\Big) t }. \] Hence \begin{align*} (l(s+t)-\sqrt{x})^2-\psi_A(t)^2&>\Big(\sqrt{l(s)^2-\beta_1 t}-\sqrt{x}\Big)^2-(l(s)-\sqrt{x})^2+\beta_1\Big(1-\frac{\sqrt{x}}{l(s)}\Big)t \\ &> 2\sqrt{x}\Big( l(s)-\sqrt{l^2(s)-\beta_1 t} \Big)-\frac{\beta_1\sqrt{x}}{l(s)}\, t\ge 0. \end{align*} This leads directly to \eqref{eq:a-dem}. \end{proof} \begin{rem} The whole study points out a new efficient algorithm in order to simulate Bessel hitting times for given levels or curved boundaries. We can use this algorithm in two generalized situation: \begin{itemize} \item[(1)] We have assumed that the Bessel process starts from the origin. Of course the procedure presented here can also be applied to Bessel processes starting from $x>0$. It suffices to change the initialization step ! \item[(2)] We focused our attention to the Bessel process but we linked also the initial problem to the exit time of a $\delta$-dimensional Brownian motion from a ball of radius $l$. The Algorithm (A1) extended to higher dimensions can also be used in order to evaluate exit times of general compact domains whose boundary is regular. \end{itemize} \end{rem} \noindent \begin{minipage}{11cm} \section{Numerical results} In this part we will illustrate the previous results on some numerical examples. Let us figure first an outcome of our algorithm, the exit position from a sphere with radius depending on time. The figure opposite is giving this result for an radius $l=1$ and a precision $\varepsilon = 10^{-3}$. \\ Let us compare our algorithm with existing results. Consider the classical Euler scheme for a Brownian motion and evaluate the first hitting time and hitting position from a disk with given radius. \end{minipage} \begin{minipage}{6cm} \centerline{\includegraphics[height=5cm, width=5.5cm]{cercle.eps}} \end{minipage}\\[4pt] \noindent First of all we can verify that the distribution of the hitting time for the $WoMS$ algorithm matches the distribution of the hitting time of a given level for the $2-$dimensional Bessel process. Figure \ref{fig:histo_pos_sortie} gives this result for a starting disk with radius $1$, a precision $\varepsilon=10^{-3}$ and a number of simulations $N=20000$. In the Euler scheme the time step is $\Delta t = 10^{-4}$. \begin{figure}[H] \begin{center} \centerline{\includegraphics[height=5cm]{histogramme_position_sortie.eps}\includegraphics[height=5cm]{teta_uniforme.eps}} \caption{Distribution of the hitting time (Euler scheme and $WoMS$ algorithm (A1)) - Histogram of the angle for the exit position} \label{fig:histo_pos_sortie} \end{center} \end{figure} We can also test the fact that the exit position is uniformly distributed on the circle. In order to do this we can evaluate the angle of the exit position in our $WoMS$ procedure and show that it is an uniform distributed random variable with values in $[-\pi, \pi]$. Figure \ref{fig:histo_pos_sortie} also shows the histogram of the result for a disk of radius $1$ an $\varepsilon=10^{-3}$ and 20000 simulations. Let us now present a simulation with Algorithm (A2). We consider the hitting time of the level $l=2$ for the Bessel process of index $\nu=2$ and we illustrate Theorem \ref{thm:algo} by Figure \ref{graph1}. The curve represents the averaged number of steps versus the precision $\varepsilon=10^{-k}$, $k=1,\ldots, 7$. We can observe that the number of steps is better than suspected since the curve is sublinear. We obtain the following values (for $\gamma=0.9$ and $100\, 000$ simulations in order to evaluate the mean). \renewcommand{\arraystretch}{1.5} \begin{figure}[!h] \begin{minipage}{5cm} \includegraphics[height=5cm]{graph1.eps} \end{minipage}\hspace*{2cm} \begin{minipage}{5cm} \begin{tabular}{|c||c|c|c|c|} \hline $\varepsilon$ & $10^{-1}$ & $10^{-2}$ & $10^{-3}$ & $10^{-4}$ \\ \hline $\mathbb{E}[N^\varepsilon]$ & 4.0807 & 7.53902 & 9.50845 & 10.83133 \\ \hline\hline $\varepsilon$ & $10^{-5}$ & $10^{-6}$ & $10^{-7}$ & \\ \hline $\mathbb{E}[N^\varepsilon]$ & 10.94468 & 11.30869 & 11.62303 & \\ \hline \end{tabular} \end{minipage} \caption{Averaged number of step of Algorithm (A2) versus $\varepsilon$} \label{graph1} \end{figure} Finally we present the dependence of the averaged number of steps of Algorithm (A2) with respect to the dimension of the Bessel process. For that purpose, we simulate hitting time of the level $l=2$ with $\varepsilon=10^{-3}$, $\gamma=0.9$, $50\, 000$ simulations for each estimation of the averaged value, and the dimension of the Bessel process takes value in the set $\{2,3,\ldots,18\}$. \begin{figure}[!h] \renewcommand{\arraystretch}{1.5} \begin{minipage}{7cm} \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline $\nu$ & 0 & 0.5 & 1 & 1.5 & 2 \\ \hline $\mathbb{E}[N^\varepsilon]$ & 6.819 & 7.405 & 8.270 & 8.887 & 9.594 \\ \hline\hline $\nu$ & 2.5 & 3 & 3.5 & 4 & \\ \hline $\mathbb{E}[N^\varepsilon]$ & 10.256 & 10.542 & 10.995 & 11.096 & \\ \hline \end{tabular} \end{minipage}\hspace*{2cm} \begin{minipage}{5cm} \includegraphics[height=5cm]{graphnu.eps} \end{minipage} \caption{Averaged number of step of Algorithm (A2) versus $\delta=2\nu+2$} \end{figure} \section{Application to the Cox-Ingersoll-Ross process} We now aim to estimate the hitting time of a level $l>0$ for $(X^\delta_t,\, t\geq 0)$, a Cox-Ingersoll-Ross process. The CIR process is the solution of the following stochastic differential equation: \begin{equation} \label{CIR:def} \left\{ \begin{array}{l} \mathrm{d} X^\delta_t=(a+bX^\delta_t)\mathrm{d} t+c\sqrt{|X^\delta_t|}\mathrm{d} B_t,\\ X^\delta_0=x_0, \\ \end{array} \right. \end{equation} where $x_0\geq 0$, $a\geq 0$, $b\in\mathbb{R}$, $c>0$ and $(B_t,\, t\geq 0)$ is a standard Brownian motion. We denote here $\delta = 4a/c^2$. \\ We will first recall a connection between this stochastic process and $(Y^\delta(t),\, t\geq 0)$, the square of the Bessel process BESQ($\delta$), solution of the equation \begin{equation} \label{def:bessel} Y^\delta(t)=y_0+\delta t+2\int_0^t \sqrt{|Y^\delta (s)|}\mathrm{d} B_s,\quad t\geq 0. \end{equation} \begin{lemma} \label{lem:identite} The CIR process has the same distribution as $(\overline{X}_t,\, t\geq 0)$ which is defined by \begin{equation} \label{eq:cir-relation} \left\{ \begin{array}{l} \overline{X}_t=e^{bt}Y^\delta\Big( \frac{c^2}{4b}(1-e^{-bt}) \Big),\\ \overline{X}_0=Y^\delta(0), \end{array} \right. \end{equation} where $Y$ is the square of a Bessel process in dimension $\delta=4a/c^2$ (see \cite{Revuz-Yor-99}). \end{lemma} \begin{proof} Let us only sketch some ideas of the proof. Let $Y^\delta(t)$ be the square of the $\delta$-dimensional Bessel process. By applying It{\^o} formula, we get the stochastic differential equation satisfied by the process $\overline{X}_t$ \begin{align} \mathrm{d}\overline{X}_t&=b\overline{X}_t\,\mathrm{d} t+e^{bt} \mathrm{d}\Big(Y\Big( \frac{c^2}{4b}(1-e^{-bt}) \Big)\Big)\nonumber\\ &=b\overline{X}_t\,\mathrm{d} t+b\delta\frac{c^2}{4b}\,\mathrm{d} t+2e^{bt}\sqrt{|e^{-bt}\overline{X}_t|}\,\mathrm{d} B_{ \frac{c^2}{4b}(1-e^{-bt})}\nonumber\\ &=(a+b\overline{X}_t)\,\mathrm{d} t+2e^{\frac{bt}{2}}\sqrt{|\overline{X}_t|}\,\mathrm{d} B_{ \frac{c^2}{4b}(1-e^{-bt})},\label{eq:eds1} \end{align} where $\delta =4a/ c^2.$ Let us remark that \[ \frac{c^2}{4b}(1-e^{-bt})=\int_0^t \rho^2(s)\mathrm{d} s,\quad\mbox{with}\ \rho(t)=\frac{c}{2}\, e^{-\frac{bt}{2}}. \] We can deduce that there exists a Brownian motion $(\beta_t,\, t\geq 0)$ such that \begin{equation*} B_{ \frac{c^2}{4b}(1-e^{-bt})}=\int_0^t \rho(s)\mathrm{d} \beta_s, \end{equation*} for all $t\geq 0$. With this notation, the equation \eqref{eq:eds1} writes \begin{align*} \mathrm{d} \overline{X}_t&= (a+b\overline{X}_t)\,\mathrm{d} t+2e^{\frac{bt}{2}}\sqrt{|\overline{X}_t|}\rho(t)\, \mathrm{d} \beta_t\\ &=(a+b\overline{X}_t)\,\mathrm{d} t+c\sqrt{|\overline{X}_t|}\mathrm{d} \beta_t, \end{align*} and $\overline{X}_0=Y(0)$. This proves that the process $(\overline{X}_t,\, t\geq0)$ has the same distribution as the CIR process given by (\ref{CIR:def}). \end{proof} Let us consider the hitting time of a given level $l$ for the CIR process and denote it by $T_l$. This time is defined by: \begin{equation*} T_l=\inf\{ s\geq 0; X^\delta_s=l\}. \end{equation*} The previous Lemma \ref{lem:identite} gives also an equivalence (in distribution) connecting the hitting time of the CIR process and the hitting time of the square of a $\delta$-dimensional Bessel process. \begin{proposition}\label{prop:enfin} The hitting time $T_l$ of a level $l>0$ for a CIR process has the same distribution as $-\frac{1}{b}\log \Big(1-\frac{4b}{c^2}\tau_\psi\Big)$ where \[ \tau_\psi=\inf\Big\{ t\geq 0; Y^\delta(t)=l\Big(1-\frac{4b}{c^2}t \Big)\Big\}, \] and $Y^\delta$ is the square of a Bessel process of dimension $\delta=4a/c^2$. \end{proposition} \begin{proof} By using Lemma \ref{lem:identite}, $\tau_\psi$ has the same distribution as $\overline{T}_l$ given by \begin{equation} \label{def:tau_2} \overline{T}_l=\inf\Big\{s\geq 0;\ Y^\delta\Big( \frac{c^2}{4b}(1-e^{-bs}) \Big)=le^{-bs}\Big\}. \end{equation} Define $t=\frac{c^2}{4b}(1-e^{-bs})$, we have two situations:\\ {\bf First case:} If $b<0$ let $s=\eta(t)$ where \[ \eta(t)=-\displaystyle\frac{1}{b}\log\Big(1-\frac{4b}{c^2}t\Big),\quad\mbox{for}\ t\geq 0. \] The map $\eta$ is a strictly non-decreasing function, and we get thus \[ \overline{T}_l=\inf\Big\{\eta(t);\ t\geq 0,\ Y^\delta(t)=l\Big(1-\frac{4b}{c^2}t \Big)\Big\}=\eta\Big( \inf\Big\{t\geq 0;Y^\delta(t)=l\Big(1-\frac{4b}{c^2}t \Big)\Big\} \Big). \] {\bf Second case:} If $b\geq 0$, let also $s=\eta(t)$. In this case the variable $t$ takes its values only on the interval $\Big[0,\frac{c^2}{4b}\Big)$. So \[ \overline{T}_l=\inf\Big\{\eta(t);\ 0\le t\le \frac{c^2}{4b},\ Y^\delta(t)=l\Big(1-\frac{4b}{c^2} t \Big)\Big\}. \] The condition $0\le t\le \frac{c^2}{4b}$ can be omitted in the estimation of the infimum as the boundary to hit~: $1-\frac{4bt}{c^2}$ is negative outside this interval and the Bessel process is always positive. Furthermore the function $\eta$ is also non-decreasing for $b\geq 0$, and the result is thus obtained. \end{proof} \noindent {\bf Application of Algorithm (A4):}\\ An immediate consequence of Proposition \ref{prop:enfin} is that the hitting time $T_l$ is related to the first time the Bessel process of dimension $\delta=4a/c^2$ reaches the curved boundary: $f(t)=\sqrt{l\Big(1-\frac{4b}{c^2}t \Big)}$. We are able to apply Algorithm (A4) if $4a/c^2\in\mathbb{N}^*$ and $b>0$ (the boundary is then decreasing). Let us denote by $N^\varepsilon$ the number of steps of (A4) and $\Xi_{N^\varepsilon}$ the approximated hitting time of the Bessel process associated with the particular curved boundary $f$. Combining Proposition \ref{prop:curved} and Proposition \ref{prop:enfin} leads to $$ \left(1-\frac{\varepsilon}{\sqrt{2\alpha\pi}}\right)\mathbb{P}\Big(\Xi_{N^\varepsilon}\le \frac{c^2}{4b}\,(1-e^{-bt})-\alpha\Big)\le \mathbb{P}(T_l\le t)\le \mathbb{P}\Big(\Xi_{N^\varepsilon}\le \frac{c^2}{4b}\,(1-e^{-bt})\Big), $$ for $\alpha$ small enough and $t>0$.
\section{Introduction} The world-wide effort to directly detect \emph{gravitational waves} (GWs) for the first time is an ambitious project that unites the expertise from various fields in experimental and theoretical physics. A network of instruments, containing the Laser Interferometer Gravitational-wave Observatory (LIGO) \cite{Abbott:2007kv,Sigg:2008zz,Smith:2009bx}, VIRGO \cite{Acernese:2008zzf,Accadia:2011zz} and GEO600 \cite{Grote:2008zz,Luck:2010rs}, will soon reach a sensitivity where the signatures of coalescing compact binaries are expected to be seen above the noise level of the detectors a few times to hundreds of times per year~\cite{Abadie:2010cf}. In the case of binaries that consist of black holes (BHs) and/or neutron stars, the correct interpretation of the GW signals crucially depends on the quality of theoretically predicted \emph{template waveforms} that have to be used to identify the physical properties of the source. This paper focuses on waveform families of \emph{binary BHs} as they constitute one of the most promising sources of a first direct detection of GWs. Their modelling typically combines two very different approximation procedures. One describes the early inspiral of both objects through an asymptotic expansion in terms of the relative velocity $v/c$, where $c$ is the speed of light. As long as this quotient is small, the resulting post-Newtonian (PN) equations are an adequate representation of the dynamical evolution of the binary \cite{lrr-2006-4}. Because of the simple form of PN approximants that provide the GW signal in terms of differential equations or, in some cases, even in a closed form, they have long been the favourite tool for data-analysis applications. However, as the two BHs orbit around each other, they lose energy through the emission of GWs, and their distance shrinks along with an increase in velocity. Consequently, PN predictions become more and more inaccurate the closer the binary gets to merger. Different analytical modifications are known that try to enhance the convergence of the PN series, even close to merger, and one of the most successful methods is the \emph{effective-one-body} (EOB) approach \cite{Buonanno:1998gg,Buonanno:2000ef,Damour:1997ub,Damour:2000we}. Without further information, however, all these analytical schemes break down at some point prior to the merger of both BHs, and a second approach has to be used to model the dynamics from the late inspiral through the merger: \emph{numerical relativity} (NR). In NR, the full Einstein equations are usually solved discretely on a finite grid that is adapted to the movement of the two bodies, and the resolution in space and time is chosen fine enough to obtain a converging result. The GW content is extracted at finite radii and then extrapolated to infinity, or it is directly extracted at null infinity via Cauchy-characteristic extraction~\cite{Reisswig:2009rx,Babiuc:2010ze}. For current overviews of the field see for example \cite{Hannam:2009rd,Hinder:2010vn,Centrella:2010zf,McWilliams:2010iq, Sperhake:2011xk}. \begin{figure} \centering \includegraphics[width=0.9\textwidth]{figure1} \caption{The dominant spherical harmonic mode of the gravitational wave signal of two coalescing (nonspinning) BHs as a function of time. The different approximation schemes and their range of validity are indicated. Wavy lines illustrate the regime close to merger where analytical methods have to be bridged by NR.} \label{fig:longwave} \end{figure} Both numerical and analytical approaches have their limitations. The PN-based formulations are, by construction, not valid throughout the entire coalescence process; NR relies on computationally very expensive simulations that become increasingly challenging (and time-consuming) with larger initial separations, higher spin magnitudes of the BHs and higher mass-ratios $q = m_1/m_2$ ($m_i$ are the masses of the individual BHs and we use the convention $m_1 \geq m_2$). Thus, to build models of the complete inspiral, merger and ringdown signal, one has to combine information from both analytical and numerical approximations. See Fig.~\ref{fig:longwave} for an illustration of the dominant harmonic mode of a nonspinning binary. These `complete' waveforms are indispensable to perfect current search strategies. They constitute our best and most complete approximation of the real signals that we are trying to detect, which makes them ideal target waveforms in a simulated search to test existing analysis algorithms. The Numerical INJection Analysis (NINJA) project \cite{Aylott:2009ya,Ajith:2012tt} is dedicated to that question. The other important application of complete waveforms is to derive an analytical model from them which leads to an improved template bank in the search. The improvement manifests itself, e.g., in a wider detection range and a more accurate extraction of the physical information encoded in the signals. Ongoing searches with such templates in LIGO data are summarized for instance in \cite{Abadie:2011kd}. This paper briefly describes the efforts to build complete waveform models by combining analytical approximants and NR into individual signals and eventually entire waveform families. Our focus then turns to the question of how reliable and accurate such final models are. After all, one expects (and finds) a smooth connection between the two parts of a supposedly common GW signal, but the use in actual analysis algorithms of GW interferometers requires a much deeper error analysis with a quantitative understanding of the uncertainty introduced in the modelling process. \section{Concepts for constructing full waveform models} \label{sec:concepts} \subsection{EOBNR} The EOB formalism has been refined several times to incorporate additional information from NR. Depending on the number of available NR waveforms as well as the modifications introduced to the EOB description, various versions of such \emph{EOBNR} models have been developed~\cite{Buonanno:2007pf,Buonanno:2009qa,Damour:2007vq, Damour:2008te,Damour:2009kr,Pan:2009wj,Yunes:2009ef,Pan:2011gk}. It is beyond the scope of this paper to repeat the technical details of the EOB formalism and its extensions. For the sake of comparison to other approaches, however, we shall summarize the general strategy towards complete inspiral-merger-ringdown EOBNR models below. The main additions that allow for the description of the entire GW signal are (a) a generalization of the EOB formalism which introduces free parameters to be calibrated by NR simulations and (b) attaching a series of damped sinusoidal oscillations (quasinormal modes) representing the final stage of the BH ringdown (see, e.g., \cite{Berti:2009kk}). The proposed variants of EOBNR mainly differ in the way the original EOB description is modified and which free parameters are introduced. The most recent versions by Damour and Nagar \cite{Damour:2009kr} and Pan \emph{et al.}~\cite{Pan:2011gk} extend the standard EOB form through the following steps: \begin{itemize} \item Two unknown parameters representing the 4PN- and 5PN-order contributions are added to the radial potential [commonly referred to as $A(u)$] that enters the Hamiltonian. As for many quantities in the EOB framework, using \emph{Pad\'e} resummation \cite{Damour:1997ub} proves to be superior to the Taylor-expanded form (which is, however, not always true, see the discussion about a generalization to spinning BHs \cite{Pan:2009wj} and also \cite{Mroue:2008fu}). \item The radiation-reaction force and the waveform modes are written in a resummed, factorized manner \cite{Damour:2008gu}. Additional coefficients are introduced in the waveform, accounting for further, undetermined PN contributions and next-to-quasi-circular corrections. \item A sum of quasi-normal modes is attached to the inspiral-plunge EOBNR waveform over a certain time interval around the peak of the waveform mode. \end{itemize} The impact of NR on the above strategy is manifold. Some parameters (like the EOB-dynamical parameters introduced into the radial potential) are directly determined through minimizing the phase difference between the analytical and numerical GW. Other parameters are derived from independent (i.e., not EOB-related) fits of the numerical data, such as predictions of the final spin of the remnant BH or the maximum of the modulus of the GW. Note, however, that for a direct comparison (and thereby calibration), analytical and numerical waveforms have to be \emph{aligned}, i.e., a relative shift in time and phase has to be fixed by some minimization procedure. We shall find the same need in all construction algorithms for complete GW signals. In short, the characteristics of EOBNR constructions are that a well-adapted analytical description is extended and \emph{informed} by NR data, so that finally a time-domain description based on a set of differential equations provides the entire inspiral to plunge signal that is completed by attaching the ringdown waveform. \subsection{Phenomenological models} Although there is a common strategy in all modelling procedures described here, let us highlight a few distinct features of phenomenological waveform families as introduced by Ajith \emph{et al.}~\cite{Ajith:2007qp,Ajith:2007kx,Ajith:2009bn} and Santamar\'ia \emph{et al.}~\cite{Santamaria:2010yb}. These families are built by first constructing a finite set of complete \emph{hybrid} waveforms \cite{Ajith:2007kx,Boyle:2008ge,Boyle:2009dg,Santamaria:2010yb, Hannam:2010ky, MacDonald:2011ne,Boyle:2011dy} that are direct combinations of the available NR data with the appropriate waveforms obtained with some PN approximant (usually based on Taylor-expanded quantities). The construction of these hybrid waveforms may differ, as they can be based in the time or frequency domain, they can overlap both waveform parts at a single point or over an interval and they can impose various requirements on the smoothness of the transition. However, all hybridization procedures are based on finding the, in some sense, optimal alignment between two parts of the same waveform by exploiting the free relative time and phase shift. Different from the EOBNR approach, this combination of analytical and numerical waveforms does not immediately lead to a model allowing arbitrary physical parameters. Hybrid waveforms merely constitute the set of discrete \emph{target signals} that are represented in a next step as accurately as possible by a simple and convenient multi-parameter fit. This fit is separated from the analytical approach used to describe the inspiral of the hybrid waveform. For instance, the latest model of Ajith \emph{et al.}~\cite{Ajith:2009bn} employs a time-domain PN approximant commonly denoted by `TaylorT1' in the hybrid construction, but the final multi-parameter model is instead inspired by the form of a Fourier-domain PN approximant (see for instance \cite{Buonanno:2009zt} for an overview of the different PN approximants). The final fit that turns a set of hybrid waveforms distributed in the parameter space into an analytical model is a delicate procedure. Introducing an arbitrary (yet as small as possible) number of parameters to fit a relatively small number of hybrids is not difficult. These auxiliary parameters, however, have to be a smooth function of the \emph{physical parameters} (notably symmetric mass ratio and spins) themselves in order to allow for an interpolation of the parameter space. Only if the latter can be achieved, again with guidance from PN descriptions and the knowledge of quasinormal ringdown modes, the model becomes potentially useful for data-analysis purposes without increasing the rate of false alarms in a search process. In the end, the phenomenological descriptions mentioned above \cite{Ajith:2007qp,Ajith:2007kx,Ajith:2009bn,Santamaria:2010yb} are provided in terms of closed-form equations representing the GW signal in Fourier space. It should be noted that, although the procedure of combining PN and NR data in a first step and analytically modelling it in a second step is conceptually useful to analyse different error sources (see Sec.~\ref{sec:errors}), it is not entirely different from the EOBNR approach. If the inspiral model used in the hybrid would be EOB and an extended EOB description is chosen as the `phenomenological model', then we would recover the EOBNR construction. Likewise, if the EOBNR construction would calibrate its model against a complete hybrid signal instead of pure NR data, it would be conceptually no different from phenomenological constructions (which does not imply that one construction cannot be superior to the other). The important question ultimately is how \emph{flexible} and \emph{accurate} each individual strategy (with all its detailed distinctions) can predict the unknown real GW signal. We shall touch this question in Sec.~\ref{sec:errors}. For completeness, let us mention another phenomenological family that was constructed by Sturani \emph{et al.} \cite{Sturani:2010yv,Sturani:2010ju} as a first step to model waveforms of precessing binaries. In this approach, a Taylor-expanded time-domain approximant (`TaylorT4') is extended and finally fitted to NR data. Just like EOBNR (although less sophisticated), the resulting model is given in form of time-domain differential equations with quasinormal ringdown modes attached. \section{Physical range of waveform models} Understanding the concepts underlying the construction of complete waveform models is mainly interesting when we want to to compare various approaches, deduce why they lead to slightly different waveforms and, most importantly, assess the quality of individual families. In this section, however, we will first summarize the facts that are interesting for the actual usage of the waveforms in data-analysis applications. In particular, before applying the model to a set of physical parameters, one should have a clear perception of \emph{where in the parameter space} these models have been constructed. Although this range does not necessarily coincide with the range of parameters the model can be used with, it nevertheless is a good indication where it can be trusted most. The waveform models that have been introduced in Section~\ref{sec:concepts} are tailored to model binary BHs with comparable masses inspiralling on quasi-circular orbits. There are successful efforts to exploit the synergy of analytical methods and NR also for other scenarios, like the extreme mass-ratio regime \cite{Yunes:2009ef,Yunes:2010zj} or binary neutron star coalescences \cite{Damour:2007vq,Baiotti:2011am}. In this paper, we focus on binary BHs in the comparable-mass regime only, as they are the most promising sources for the upcoming generation of ground-based GW detectors whose detection and interpretation may require information both from PN and NR. \fulltable{A selection of recent complete waveform models for BH binaries with comparable masses on quasi-spherical orbits. We summarize the reference where the model was described, the approximate inspiral waveforms and NR codes that were employed, the parameter range in which each model was calibrated ($q$ is the mass ratio) and the number of parameters and NR simulations used to build the model. \label{tab:models}} \br Alias & Ref. & Inspiral & NR code & Calibration range & Calibrated parameters \\ \mr EOBNR & \cite{Damour:2009kr} & EOB & \texttt{SpEC}, \texttt{BAM} & $q \leq 4$ \& $q \to 0$ & 2 dynamical from $q=1$ \\ &&&& no spins & + fits from $q \in \{ 1,2,4 \}$ \\ EOBNR & \cite{Pan:2011gk} & EOB & \texttt{SpEC} & $q \leq 6$ \& $q \to 0$ & 2 dyn. + 4 waveform par./mode \\ &&&& no spins & 5 leading modes from 5 NR runs\\ Phenom\emph{B} & \cite{Ajith:2009bn} & T1 & \texttt{BAM} & $q \leq 4$ \& $q \to 0$ & 6 phase, 4 amplitude \\ &&&& aligned spins & from 24 NR simulations \\ Phenom\emph{C} & \cite{Santamaria:2010yb} & F2 & \texttt{BAM} & $q \leq 4$ & 6 phase, 3 amplitude \\ &&&& aligned spins & from 24 NR simulations \\ PhenSpin & \cite{Sturani:2010ju} & T4 & \texttt{MayaKranc} & $q = 1, \chi_i = 0.6$ & 2 phase param. \\ &&&& precession & from 24 NR sim.+4 Phenom\emph{B} \\ \br \endfulltable In Table~\ref{tab:models} we provide an overview of selected, recent models for this regime. Apart from an alias (partially adopted from the LIGO-Virgo collaboration \cite{LSC,Virgo}) we indicate the inspiral model which is either based on the EOB approach or derived from Taylor-expanded PN quantities. In the latter case, various different PN approximants are known depending on the details of the re-expansion and integration. The time-domain approximants are commonly referred to as \emph{TaylorTn} (where $n$ ranges from 1 to 4, and a fifth version has recently been suggested \cite{Ajith:2011ec}); a frequency-domain representation obtained via the stationary-phase approximation is denoted by \emph{TaylorF2}. For details, see \cite{Boyle:2007ft,Buonanno:2009zt,Brown:2007jx} and references therein. The NR codes that contributed to the construction of the given models are the Spectral Einstein Code (\texttt{SpEC}~\cite{Scheel:2006gg,Spec}), \texttt{BAM}~\cite{Bruegmann:2006at} and \texttt{MayaKranc}~\cite{Vaishnav:2007nm}, where the few \texttt{SpEC} waveforms are notably long and accurate, the \texttt{BAM} simulations provide the largest diversity in parameter space with moderately long waveforms, and \texttt{MayaKranc} waveforms are the only precessing simulations used to calibrate analytical models to date. Other distinctive features of the models listed in Table~\ref{tab:models} are for example as follows: \begin{itemize} \item Phenom\emph{B}/\emph{C} are closed-form frequency-domain representations of the GW; EOBNR and PhenSpin provide the signal in terms of time-domain differential equations. \item EOBNR models can readily be extended beyond the dominant spherical harmonic of the GW, whereas the phenomenological models and PhenSpin solely provide the signal in terms of the $\ell =2, m=\pm 2$ (spin-weighted) spherical harmonic modes. \item The PhenSpin model is a first attempt to model generic precessing spin configurations, but it is so far only calibrated to equal-mass systems and dimensionless spin magnitudes of 0.6. All other models in Table~\ref{tab:models} are only applicable to nonspinning systems or systems where the spin of each BH is aligned (or antialigned) with the total orbital angular momentum. \end{itemize} In the aligned-spin case, both phenomenological waveform families reduce the two spin parameters to one ``total'' spin \begin{equation} \chi = \frac{m_1 \chi_1 + m_2 \chi_2}{m_1 + m_2}, \end{equation} where $m_i$ are the individual masses and the dimensionless spin magnitudes are $\chi_i = \pm \vert \bi S _i \vert /m_i^2$ (the sign distinguishing aligned and antialigned configurations). As recently shown by Ajith~\cite{Ajith:2011ec}, this degeneracy in the spin parameters can be further optimized, and it will be an important goal for future models to describe as many physical effects as possible with the smallest possible number of parameters. In the nonspinning case, all waveforms presented here are parametrized in terms of the \emph{physical} parameters total mass and symmetric mass ratio (plus initial time and phase) but it may be useful both from the modelling and the search point of view to refrain from this parametrization strategy once all additional spin dynamics are included. Note that there is also an EOBNR model proposed that includes aligned-spin configurations~\cite{Pan:2009wj}, but this first exploratory study only employed two equal-mass simulations (performed with \texttt{SpEC}) with equal spins $\chi_1 = \chi_2 \approx \pm 0.44$. Apart from the listed facts, there are many more procedures involved in checking the validity of proposed models. Most importantly, it has been shown to some extent that the models mentioned here agree to reasonable accuracy with the waveforms they were derived from, but also with waveforms that were \emph{not} in the construction set. Thus, with an increasing number of available numerical simulations, all these models can not only be extended and refined, they can also be cross-checked extensively until, ideally, one can confidently interpolate over the entire parameter space independent of the set of waveforms actually used to calibrate the model. \section{Uncertainties in the modelling process} \label{sec:errors} Having briefly sketched some successful approaches to combine analytical and numerical methods in the GW-modelling process, let us recapitulate the choices that had to be made along the way. \begin{itemize} \item Which PN/EOB formulation should be employed? \item What \emph{physical parameters} in PN and NR are consistent with the other framework? \item Which NR resolution, extraction formalism etc. is sufficient? \item How long do the NR waveforms have to be? \item What is the appropriate way to match analytical and numerical data? \item How do the fitting parameters depend on physical quantities? \end{itemize} When constructing a complete waveform model, each of these questions has to be answered and different choices lead to the different results presented above. The important conclusion we shall draw from this is that none of the suggested models is based on an unambiguous construction, and the spread of possible results that different reasonable choices yield is a measure of the uncertainty within the modelling process. We shall first outline the basic concepts of evaluating these uncertainties in a way meaningful for GW searches and then summarize some results that have been obtained in the recent past. \subsection{Accuracy requirements for detection and parameter estimation} Let us recall the basic strategy of a matched-filter search (see, e.g., \cite{Finn:1992wt,Cutler:1994ys,lrr-2005-3}) that employs a theoretically predicted template bank of waveforms $h_2$ which in turn are characterized by the parameters $\theta$. Assume that a real GW signal $h_1$ is contained in the data stream. The search mainly relies on finding the maximum agreement between $h_1$ and $h_2 (\theta)$, for any $\theta$, as quantified by the inner product \begin{equation} \big \langle h_1 , h_2 \big \rangle = 4 \, {\rm Re} \int_{f_1}^{f_2} \frac{\tilde h_1(f) \, \tilde h_2^\ast(f)}{S_n(f)} \, df. \label{eq:innerprod} \end{equation} Here, $\tilde h_i$ are the Fourier transforms of $h_i$, $^\ast$ denotes the complex conjugation, and $S_n$ is the noise spectral density of the assumed GW detector (we assume stationary Gaussian noise with zero mean). The \emph{detection efficiency} can be expressed in terms of the minimal mismatch \begin{equation} \mathcal M = 1 - \max_{\theta} \frac{\big \langle h_1, h_2(\theta) \big \rangle}{\| h_1 \| \|h_2 (\theta) \|}, \qquad (\| h_i \| = \sqrt{% \big \langle h_i, h_i \big \rangle }), \label{eq:MM} \end{equation} which quantifies the loss in \emph{signal-to-noise ratio} (SNR) due to an inexact model, or equivalently yields the fraction of missed signals in a matched-filter search. If, for example, up to $x=10\%$ of the detectable signals may be missed due to a mismatch of real and modelled waveform, we can allow this mismatch to be at most $\mathcal M = 1 - \sqrt[3]{1-x} \approx 3.5\%$. Typically, one does not have access to an ideal target waveform and an approximate search family, so one commonly uses the mismatch between two supposedly equivalent, approximate waveforms to quantify the error of the modelling process itself. The other simplification that often comes with the restriction to discrete points in the parameter space is that instead of maximizing (\ref{eq:MM}) with respect to all parameters, one only exploits a free relative time and phase shift between the waveforms and varies along $\theta = (t_0, \phi_0)$. This can only be an upper bound on $\mathcal M$, which is of course sufficient if the values found are small enough. The uncertainty of estimating parameters in a search has recently \cite{Lindblom:2008cm,Damour:2010zb,Boyle:2011dy,MacDonald:2011ne} been based on the requirement that the error of the waveform model is \emph{indistinguishable} by the detector. With $h_1$ and $h_2$ denoting supposedly equivalent waveforms, this requirement reads \begin{equation} \| \delta h \| = \| h_1 - h_2 \| < \epsilon, \label{eq:indist} \end{equation} with $\epsilon$ being the fraction of the noise level we allow for the model uncertainty ($\epsilon \lesssim 1$). The parameters of $h_1$ and $h_2$ are deliberately kept the same in (\ref{eq:indist}), except for a time and phase shift that is used to minimize the distance. The other parameters (masses, spins) should be determined in the search with no bias introduced by the model; thus, we do not optimize over them in (\ref{eq:indist}). Note that (\ref{eq:indist}) can also be expressed in terms of the mismatch (also minimized over a time and phase shift only) by assuming the equal norms $\| h_1 \| = \| h_2 \| = \|h \|$ \cite{Flanagan:1997kp,McWilliams:2010eq}, \begin{equation} \mathcal M = \frac12 \, \frac{\| \delta h \|^2}{\|h\|^2} \qquad \Rightarrow \quad (\ref{eq:indist}) ~\Leftrightarrow~ \mathcal M < \frac{\epsilon^2}{2 \| h \|^2}. \label{eq:misdistance} \end{equation} Fulfilling (\ref{eq:indist}) or (\ref{eq:misdistance}) is certainly the ultimate goal for an accurate waveform model, and we can easily understand that if the waveform uncertainty $\delta h$ is not even detectable in the presence of instrument noise, it has no effect on the measurement. The converse, however, cannot be interpreted in this straightforward manner. If $\| \delta h \| > 1$, we would expect that the model uncertainty has \emph{some} effect on the parameter estimation, but \emph{which} effect it actually has must be quantified through the \emph{systematic} (i.e., model-induced) parameter bias which is defined as the difference between the parameter values of the best fitting template and the true parameters \cite{Damour:1997ub}. These errors should then be compared to statistical (i.e., noise-induced) errors, and there are explicit expressions available for both types of biases in the high SNR regime \cite{Flanagan:1997kp,Cutler:2007mi}. Here we just remind the reader that (\ref{eq:indist}) was derived as a \emph{sufficient} criterion to ensure that the systematic errors do not exceed the statistical parameter variance \cite{Flanagan:1997kp}. It is, however, not a necessary criterion, and we shall illustrate this explicitly in Sec.~\ref{sec:PNerr}. Next, we summarize some important results that have been obtained recently, all casting the question of accuracy of waveforms in the form just outlined. All the results quoted below use an Advanced LIGO noise curve \cite{Ajith:2007kx,advLIGO} with appropriate integration limits. \subsection{Errors in the NR regime} Quantifying errors is an important and very natural process for numerical integrations and the uncertainty of NR waveforms is usually given in terms of phase and amplitude error. Although these can be related to the quantities we have introduced above (see \cite{Lindblom:2008cm,Lindblom:2009ux,Lindblom:2010mh}) we will focus on publications here that directly analyse NR errors in terms of waveform mismatches (all mismatches quoted below are optimized over time and phase shifts of the waveforms). The `Samurai' project \cite{Hannam:2009hh} was a joint effort proving the consistency of NR waveforms by comparing numerical simulations of equal-mass, nonspinning binaries from five different NR codes. No completion of the waveforms with PN or EOB inspiral signals was considered as the focus laid primarily on the NR data and their errors. Therefore, the mismatches reported in \cite{Hannam:2009hh} are restricted to high frequencies; thus, high masses of the system (total mass $> 180 M_\odot$), and values of $\mathcal M < 0.1\%$ are found. Similarly, Santamar\'ia \emph{et al.}~\cite{Santamaria:2010yb} compare hybrid waveforms of nonspinning binaries with mass ratios 1 and 2 where the PN part was fixed, respectively, and NR data was produced either from the \texttt{BAM} or \texttt{Llama}~\cite{Pollney:2009yz} code. The maximal mismatch they find satisfies $\mathcal M < 0.2\%$. The effect of different resolutions used to calculate the NR part of a complete waveform was also analysed in \cite{Santamaria:2010yb} and by MacDonald \emph{et al.}~\cite{MacDonald:2011ne} who report that even the low-resolution run causes a difference to their best simulation (nonspinning, equal mass, \texttt{SpEC} code) with $\mathcal M$ not greater than $0.1\%$. Finally, various comparisons of analytical waveform models with NR data can be found in the literature, e.g., the recent EOBNR model by Pan \emph{et al.}~\cite{Pan:2011gk} exhibits a mismatch $\mathcal M \sim 0.5\%$ with an NR waveform (multiple harmonics) of a binary with mass ratio 6. Smaller mass ratios and fewer harmonics lead to smaller mismatches. Note that all these mismatches can be considered as small, at least in terms of detection, as only $\approx 1.5\% ~(0.6\%)$ of signals would be lost due to a mismatch of $\mathcal M = 0.5\%~ (0.2\%)$ (not even including additional optimizations over the physical parameters of the template bank). \subsection{Hybridization errors} There are different strategies to combine two parts of supposedly the same waveform into one signal. As one is forced to overlap analytical and numerical results in the late merger regime (as early as the NR simulation permits) one cannot expect that they agree perfectly, and there is some ambiguity about the way an `optimal alignment' of both waveform parts is defined. Most common both in EOBNR an phenomenological constructions is to decompose $h$ into phase $\phi$ and amplitude $A$ by $h = A \, e^{i \phi}$. The alignment of the PN/EOB and NR parts is then carried out by minimizing a phase difference, either over an entire interval or at discrete points, utilizing the frequency $\omega = d \phi/dt$ as well. A detailed analysis of various aspects involved in such procedures can be found in \cite{Santamaria:2010yb,MacDonald:2011ne}. The authors of \cite{Santamaria:2010yb} calculate in their example of a frequency-domain matching that the relative time shift between both waveform parts can be determined, in the best case, up to an uncertainty of $\delta t_0/M \approx 0.15$. Ref.~\cite{MacDonald:2011ne} complements this statement by estimating that $\delta t_0/M \lesssim 1$ is required for an accurate matching with $\mathcal M < 0.02\%$. In addition, the recommended matching interval is formulated in terms of the frequency evolution $\omega_1 \to \omega_2$ within this interval, and \cite{MacDonald:2011ne} suggests $(\omega_2 - \omega_1) / \omega_m \gtrsim 0.1$ (where $\omega_m$ is the transition frequency from PN to NR). Hannam \emph{et al.}~\cite{Hannam:2010ky} compared different hybridization schemes, including time and frequency-domain variants, and they show for an equal-mass, nonspinning binary that the resulting different hybrids have a mismatch of at most $\mathcal M = 0.03\%$. Again, we can summarize these results by stating that in the cases considered (mostly nonspinning, some aligned-spin configurations, mass-ratio close to unity) the hybridization involves some careful fine-tuning but the uncertainties introduced are acceptable for data-analysis purposes. \subsection{Uncertainty of the inspiral waveform -- NR length requirements} \label{sec:PNerr} Recently, several publications quantified the effect of different analytical waveform models that are completed with common late-inspiral, merger and ringdown data~\cite{Santamaria:2010yb,Hannam:2010ky,Damour:2010zb, MacDonald:2011ne,Boyle:2011dy,Ohme:2011zm}. The general approach in each of these articles is similar: different PN/EOB approximants are stitched to some given high-frequency data (Boyle~\cite{Boyle:2011dy} and Ohme \emph{et al.}~\cite{Ohme:2011zm} point out that only very limited information from NR is actually needed) and the slightly different signals are analysed in terms of their distance $\| \delta h \|$ or mismatch $\mathcal M$, at first optimized over time and phase shifts only. The results are sobering. Even approximants with nonspinning/spinning terms up to 3.5PN/2.5PN order (amplitude at 3PN/2PN order) induce hybrid-waveform disagreements that can reach mismatches of the order of a few to more than ten percent. That is at least an order of magnitude more than what has been found for NR and hybridization errors! Of course, there are again many details entering these calculations, the most crucial choices being (a) the two analytical models compared to each other, (b) the total mass of the system, (c) the other physical parameters of the system (mass ratio, spins), and (d) where the inspiral waveforms are connected to common NR data. In a conservative approach, (a) and (b) are dealt with by repeating the same analysis for various different models and calculating the mismatches for a range of masses (the waveforms themselves scale trivially with the total mass); the maximum of all of these numbers then represents the total uncertainty due to an ambiguous inspiral. However, astrophysical expectations of, e.g., the minimal constituent mass may considerably restrict the plausible range of total masses, which leads to more relaxed accuracy requirements particularly for higher mass ratios \cite{Boyle:2011dy,Ohme:2011zm}. One way to reduce the total modelling error is to shorten the PN contribution to the waveform and employ accordingly longer NR simulations. As stated before, NR data are extremely expensive to compute and therefore, estimating the required minimal simulation length is a major input to be provided by the waveform modelling community. The first efforts in this direction repeated the error analysis based on PN uncertainties simply for different matching points, thereby estimating how early one would have to match PN and NR to meet a given accuracy requirement. The results found in \cite{MacDonald:2011ne,Boyle:2011dy} suggest that, for given mass ratio and spin, numerical waveforms have to be \emph{much longer} than currently practical simulations. In particular, demanding an uncertainty indistinguishable by the detector $[$Eq.~(\ref{eq:indist})$]$, even for a moderate SNR ($\sim 10$), requires \emph{hundreds of NR orbits}. At first sight, these results are discouraging, but they are based on sufficient (but not necessary) criteria for individual waveforms. If instead the accuracy analysis is carried out for waveform families that allow for a continuous variation of all physical parameters, the corresponding error estimates are much smaller. Hannam \emph{et al.}~\cite{Hannam:2010ky} first presented mismatches that are not only optimized over phase and time shifts but also with respect to the total mass. Ohme \emph{et al.}~\cite{Ohme:2011zm} calculated mismatches that are fully optimized (also with respect to mass ratio and spin). This now allows us to quantify the errors not only in terms of waveform differences but also as modelling-based uncertainties in the determination of the parameters. In practice, \cite{Ohme:2011zm} reports fully optimized mismatches of the order of $1 \%$ and less, achieved with parameter uncertainties of $1\%$ for total mass and symmetric mass ratio and a total bias of $0.1$ for the spin parameter $\chi$. It was concluded that approximately ten orbits before merger of numerical data should be good enough for the modelling of many astrophysical systems. The reason behind such inconsistent conclusions about the required NR waveform length is that the authors take different approaches to define an accuracy goal for hybrid waveforms. From a purely theoretical point of view, every single waveform should be as accurate as possible, and any effect on the measurement should be excluded from the outset. This consequently leads to (\ref{eq:indist}) and (\ref{eq:misdistance}) and the request for hundreds of NR orbits. The other point of view is inspired by the immediate goal to detect and interpret GWs out of existing data, and although current waveform uncertainties significantly violate (\ref{eq:indist}), it is concluded that one can still potentially use waveforms incorporating $\approx$10 NR orbits for the science intended. \begin{figure} \centering \includegraphics[width=0.43\textwidth]{figure2} \caption{The parameter uncertainties for a binary system with $M=20M_\odot$, mass ratio 4 and $\chi = 0.5$. The ellipses illustrate the statistical 1-$\sigma$ uncertainty for SNR 10 and 20 (inner ellipse). They are obtained through a calculation of the Fisher information matrix (see \cite{Vallisneri:2007ev}) using the phenomenological model \cite{Santamaria:2010yb} with all its parameters. The systematic modelling bias is adopted from \cite{Ohme:2011zm}, and we use the same detector configuration here, i.e., the analytical fit of the design sensitivity for Advanced LIGO \cite{Ajith:2007kx} with a lower cutoff frequency at 20 Hz.} \label{fig:Biases} \end{figure} This much more optimistic conclusion relies on fully optimized mismatches, but also on the actual calculation of the systematic parameter bias. As we have noted before, not satisfying (\ref{eq:indist}) does not imply that the systematic bias exceeds the statistical one, and an illustration of that is given in Fig.~\ref{fig:Biases}. For the example of a binary with total mass $M = 20M_\odot$, mass ratio 4 and equal aligned spins with $\chi = 0.5$, we see that the modelling bias calculated in \cite{Ohme:2011zm} is outside of the 1-$\sigma$ ellipse (by which we mean the area that satisfies (\ref{eq:misdistance}) with $\epsilon = 1$) for all reasonable SNRs. The uncertainty of measuring the total mass $M$ and the symmetric mass-ratio $\eta$ is nevertheless dominated by the statistical error for moderate SNRs, as we can infer from the extension of the ellipses. (If we would measure the chirp mass $M_c = M \eta^{3/5}$ instead of the total mass, we would find the opposite relation, which highlights again that there is some non-negligible effect on measuring parameters, but it has to be interpreted carefully.) In the end, both views are important to assess the current state of the art in building GW models. Neither are 10 NR orbits good enough for every application nor are currently feasible waveform models entirely useless for detection or parameter estimation purposes. \subsection{Interpolation error} The final source of error we shall discuss here is the step from a discrete set of combined PN+NR waveforms to an analytical model. This involves a choice of the interpolation model that eventually represents the entire (phenomenological) waveform as a function of the physical parameters, or, less obvious in the EOBNR construction, the concrete dependence of certain parameters has to be fixed by hand (see, e.g., Fig.~5 in \cite{Pan:2011gk}). Investigations of how different choices affect the entire model have mostly been restricted to comparisons with NR data; mismatches of the entire signals in different points of the parameter space, however, have not been published to the extent other error sources have been analysed. An indication of how relevant these `interpolation errors' are is provided by the study of Damour, Nagar and Trias \cite{Damour:2010zb} who compared an EOBNR model \cite{Damour:2009kr} with phenomenological models \cite{Ajith:2007kx,Ajith:2009bn,Santamaria:2010yb}, showing that even the mismatches optimized over physical parameters (excluding the spin) exceed 3\% in some regions of the parameter space. At first sight, this might be surprising as the hybrids used to construct the models should be accurate enough for detection purposes (satisfying the 3\% mismatch criterion). The difference between the final model and hybrids is also reported to be $\mathcal M \lesssim 2\%$ ($\lesssim 5\%$ for the PhenSpin model). It should be noted, however, that the triangle inequality reads \begin{equation} \| h_{\rm model} - h_{\rm exact} \| \leq \| h_{\rm model} - h_{\rm hybrid} \| + \| h_{\rm hybrid} - h_{\rm exact} \|, \end{equation} which yields through the relation (\ref{eq:misdistance}) and its assumptions \begin{equation} \mathcal M (h_{\rm model}, h_{\rm exact} ) \leq \left( \sqrt{\mathcal M (h_{\rm model} , h_{\rm hybrid})} + \sqrt{\mathcal M (h_{\rm hybrid}, h_{\rm exact})} \right)^2. \label{eq:mm_triagle} \end{equation} Consequently, if the hybrids are accurate within, say, 2\% mismatch and the model does not deviate by more than 2\% from the set of hybrids, the resulting total uncertainty can nevertheless only be bounded to 8\%, which is far above the acceptable mismatch. It is clear from this rough estimation and the results from \cite{Damour:2010zb} that the interpolation of the final model has to be improved in the future, which can be done most easily by increasing the number of (NR/hybrid) waveforms it is constructed from. \section{Summary} Because of the rapid advance in numerical relativity and analytical waveform modelling, there are already a number of waveform models proposed that describe the entire inspiral, merger and ringdown signal of a BH binary. These models are already used to analyse data from GW laser interferometers \cite{Abadie:2011kd} and their impact will successively grow the more physical effects are understood and included in the model. One of the upcoming challenges is for instance the improved construction of full waveform models with precessing spins, and some fundamental questions concerning the choice of coordinate system have already been addressed~\cite{Schmidt:2010it,O'Shaughnessy:2011fx,Boyle:2011gg, O'Shaughnessy:2012vm}. Some attention of the modelling community has most recently been on investigating different error sources and ambiguities in the modelling process, and it was shown that different waveform models do not yet agree accurately enough so that their uncertainty can be neglected. Of course, current and future studies (such as the NINJA project \cite{Aylott:2009ya,Ajith:2012tt} and the NRAR collaboration) aim at testing various search algorithms and waveform models, and the `best model' will subsequently be refined by incorporating the results obtained in such analyses. It should be pointed out, however, that the variety of models we have today is very useful. In fact, most of the quantitative error analyses rely on the diversity different approaches generate, and if they eventually converge to an (almost) unambiguous description of the waveform, this will put GW astronomy on very solid ground. The prospects for that are rather good. Techniques to accurately simulate compact binaries and extract GWs are advancing, and the more efficient NR codes become the more they can provide large sets of waveforms spanning the parameter space. A few very long, very accurate simulations will greatly aid the analysis of fundamental questions related to the combination of analytical and numerical data, but there are already various established and well-tested strategies to match both waveform parts. In addition, most recent studies show that hybrid waveforms employing moderately long NR signals are already potentially useful for data-analysis applications, so refining existing models or introducing new complete waveform models on the basis of many of those NR waveforms should be a realistic goal to be accomplished before the advent of the advanced-detector era. The common framework to quantify the uncertainty in various waveform models has been introduced and used by several authors, which not only allows for a meaningful comparison of different approaches, it also sets the stage for the time when actually measured signals have to be interpreted. The question of how confident one can claim to identify the source of the signal will be of fundamental importance, and any signal that exceeds the uncertainty limits of all theoretically modelled waveforms will be equally exciting. \ack It is a pleasure to thank the organizers at Cardiff University for a very pleasant and enlightening conference NRDA2011/Amaldi 9. Many ideas and insights presented here are the result of numerous discussions with colleagues, with special thanks to Mark Hannam, Sascha Husa, Badri Krishnan, Stas Babak and Parameswaran Ajith. This work was supported by the IMPRS for Gravitational Wave Astronomy. \footnotesize \setlength{\bibsep}{0pt} \bibliographystyle{iopart-num}
\section{Introduction}\label{sec:intro} \begin{figure*}[t] \centering \includegraphics[scale=0.65]{FD_and_HD_DM_CRC.pdf} \caption{The full- and half-duplex modes for the cognitive relay channel.} \label{fig:FD_and_HD_DM_CRC} \end{figure*} \IEEEPARstart{T}{he} causal Cognitive Relay Channel (CRC) is a four-node channel with two senders and two receivers, in which the second sender obtains information from the first sender causally, then uses that to assist the transmissions of the first sender and its own message. Different from the assumption in the traditional cognitive channel that the secondary user knows the primary user's message non-causally, we propose several coding schemes in which the secondary user first decodes the primary user's message causally, then transmits the decoded message and its own message cognitively. In this paper, we study the cognitive relay channel in both full- and half-duplex modes. Analysis for the full-duplex mode gives us insights into the optimal coding schemes, while application to the half-duplex mode is more practical. In the full-duplex mode, there is no time division into sub-phases; both senders transmit all messages during the whole transmission. In the half-duplex mode, however, the transmission is divided into two phases with different message parts sent during each phase. In the first phase, the second user obtains a message from the first sender causally. In the second phase, these two senders transmit the messages concurrently. This cognitive relay channel has not been studied much in the literature. But it has tight relationships with the relay channel (RC), the interference channel (IC) and the cognitive interference channel (CIC). On the one hand, the second sender serves as a relay and helps forward the message from the first sender. On the other hand, these two senders interfere with each other during the transmission, and they can also cooperate cognitively. The closest channel to the CRC is the interference channel with source cooperation (IC with SC), in which both senders can exchange messages causally. Next, we review existing work related to the cognitive relay channel in both full- and half-duplex modes, then briefly summarize our main results. \subsection{Related work} \subsubsection{Full-duplex case} \emph{i) Relay channel:} Van der Meulen first proposes the concept of relay channel in \cite{Meulen71}. Cover and El Gamal further design several important techniques for relay channels, including decode-forward, compress-forward, and mixed decode-forward and compress-forward in \cite{Cover79}. A variation of the decode-forward scheme is partial decode-forward, in which the relay only decodes a part of the message from the source and forwards it to the destination instead of decoding the whole message. Kramer, Gastpar and Gupta \cite{Kramer2005TIT} extend these schemes to the multiple-node relay networks and propose several rate regions based on decode-forward, compress-forward and mixed strategies. Lim, Kim, El Gamal and Chung \cite{Lim11TIT} propose a new scheme called noisy network coding (NNC) based on compress-forward relaying. These relay coding techniques have been widely applied in other channels. For example, in \cite{Liang07a}, Liang and Kramer study the relay broadcast channel using the idea of rate splitting, block Markov encoding and partial decode-forward relaying. \emph{ii) Interference channel:} Carleial first introduces the interference channel and proposes inner and outer bounds as well as capacity results for several special cases in \cite{Carleial78}. Sato studies the capacity for the Gaussian interference channel with strong interference in \cite{Sato81TIT}. Han and Kobayashi propose the well-known Han-Kobayashi coding technique in \cite{Han} using rate splitting at the transmitters and joint decoding at the receivers, which to date achieves the largest rate region for the interference channel. Chong, Motani, Garg and El Gamal \cite{Chong08TIT} propose a variant scheme based on superposition coding, which achieves the same rate region as the original Han-Kobayashi scheme but has fewer auxiliary random variables and hence reduces the encoding and decoding complexities. \emph{iii) Cognitive interference channel:} The cognitive interference channel is another closely related channel, which plays a significant role in improving spectrum efficiency. Devroye, Mitran and Tarokh first propose the concept in \cite{Devroye2006TIT} and provide an achievable rate region based on combining Gelfand-Pinsker coding \cite{GP1980} with Han-Kobayashi scheme. They study both the genie-aided (non-causal) and the non genie-aided (causal) cases. Maric, Yates and Kramer determine the capacity region for the channel with very strong interference in \cite{Maric2007TIT}. Wu, Vishwanath and Arapostathis determine the capacity region for the weak interference case in \cite{Wu2007TIT}. Other coding schemes for the cognitive interference channel can be seen in \cite{JiangTIT2008, CaoACSSC2008,MaricEuroTele2008}. Jovicic and Viswanath \cite{Jovicic2009TIT} analyze the Gaussian cognitive channel and give the largest rate for the cognitive user under the constraint that the primary user experiences no rate degradation and uses single-user decoder. Rini, Tuninetti and Devroye \cite{Rini2011TIT} further propose several new inner, outer bounds and capacity results based on rate spitting, superposition coding, a broadcast channel-like binning scheme and Gelfand-Pinsker coding. An important technique used in all cognitive coding is the binning technique proposed by Gelfand and Pinsker in \cite{GP1980}. In Gelfand-Pinsker binning, the state of the channel is known at the input, but unknown at the output. Marton \cite{Marton79} proposes the double binning scheme and applies it to the broadcast channel. Kim, Sutivong and Cover \cite{Kim2008TIT} further analyze Gelfand-Pinsker binning to allow the decoding of a part of state information at the destination at a reduced information rate. Costa \cite{Costa83TIT} applies Gelfand-Pinsker binning to the Gaussian channel and proposes the well-known dirty paper coding (DPC) scheme, which achieves the same rate as if the channel is interference free. A surprising feature of DPC binning is that the transmit signal is independent of the state. \emph{iv) Interference channel with source cooperation:} Host-Madsen \cite{HostTIT2006} studies outer and inner bounds for the interference channel with either destination or source cooperation. The achievable rate for source cooperation is based on block Markov encoding and dirty paper coding, which includes the rate for decode-forward relaying but not the Han-Kobayashi region. Prabhakaran and Viswanath \cite{PrabhakaranTIT2011} investigate the Gaussian interference channel with source cooperation and propose an achievable rate region built on block Markov encoding, superposition coding and Han-Kobayashi scheme, but without binning, as well as several upper bounds on the sum rate. Wang and Tse \cite{WangTIT2011} study the Gaussian interference channel with conferencing transmitters and propose an achievable rate region within 6.5 bits/s/Hz of the capacity for all channel parameters. The channel is based on conferencing model, in which the common message parts are known though noiseless conference links between the two transmitters before each block transmission begins, hence there is no need for block Markovity. The scheme utilizes Marton's double binning for the cooperative private messages and superposition coding but not dirty paper coding for the non-cooperative private message parts. Cao and Chen \cite{CaoISIT2007} propose an achievable rate region for the interference channel with transmitter cooperation using block Markov encoding, rate splitting and superposition coding, dirty paper coding and random binning. This scheme includes the Han-Kobayashi region but not the decode-forward relaying rate. Yang and Tuninetti \cite{YangTIT2011} study the interference channel with generalized feedback (also known as source cooperation) and propose two schemes. The first scheme uses rate splitting and block Markov superposition coding only, in which the two users send common messages cooperatively. The second scheme extends the first one by using both block Markov superposition coding and binning, in which parts of both common and private messages are sent cooperatively. This scheme also achieves the Han-Kobayashi region but not the decode-forward relaying rate. We will discuss the schemes in \cite{CaoISIT2007, YangTIT2011} in more details in Section \ref{sec:FD_Tx_coop_compare}. Tandon and Ulukus \cite{TandonTIT2011} study an outer bound for the MAC with generalized feedback based on dependence balance \cite{HekstraTIT1989} and extend this idea to the interference channel with user cooperation. We will apply this outer bound in Section \ref{sec:FD_Outer_compare}. \subsubsection{Half-duplex case} For half-duplex communications, results also exist for the above channels, albeit fewer than in the full-duplex case. Host-Madsen and Zhang study capacity bounds for the half-duplex relay channel based on time-division in \cite{HostVTC2002, HostTIT2005} and give achievable rates for the Gaussian relay channel using partial decode-forward and compress-forward. Zhang, Jiang, Goldsmith and Cui \cite{Zhang2011TIT} study the half-duplex Gaussian relay channel with arbitrary correlated noises at the relay and destination. They also evaluate the achievable rates using decode-forward, compress-forward and amplify-forward, showing none of these schemes strictly outperforms the others. Peng and Rajan \cite{Peng2009ISIT} study the half-duplex Gaussian interference channel and compute several inner and outer bounds for transmitter or receiver cooperation. Transmitter cooperation uses decode-forward and divides the transmission into $3$ phases: $2$ broadcast phases and $1$ MIMO cooperative phase. Wu, Prabhakaran and Viswanath \cite{Wu2010ISIT} further study source cooperation for the half-duplex interference channel in the symmetric linear deterministic case and compute its sum rate. For the half-duplex cognitive interference channel, Devroye, Mitran and Tarokh \cite{Devroye2006TIT} propose four protocols in which the secondary user obtains the message from the primary user causally. Time-sharing these 4 protocols can achieve the Han-Kobayashi rate region but not the decode-forward relaying rate. Chatterjee, Tong and Oyman \cite{ChatterjeeISIT2010} further propose a new achievable rate region by a 2-phase scheme based on rate splitting, block Markov encoding, Gelfand-Pinsker binning and backward decoding. This scheme can only achieve the rate of decode-forward relaying, which is less than the partial decode-forward rate in the half-duplex mode. We will discuss these two schemes in more details in Section \ref{sec:numerical_sections}. \subsection{Summary of Main Results} In this paper, we fully define the cognitive relay channel in both the full- and half-duplex modes and propose several coding schemes based on partial decode-forward relaying, Gelfand-Pinsker binning and Han-Kobayashi coding. \subsubsection{Full-duplex case} The full-duplex cognitive relay channel is a four-node channel with two sender-receiver pairs $S_1$-$T_1$ and $S_2$-$T_2$, as in Figure \ref{fig:FD_and_HD_DM_CRC}(a). $S_1$ and $S_2$ want to transmit messages to $T_1$ and $T_2$, respectively. $S_2$ also serves as a relay by forwarding $S_1$'s message to $T_1$ while transmitting its own message to $T_2$. Since $S_2$ can both relay and apply cognitive coding at the same time, this gives rise to the name Cognitive Relay Channel (CRC). We propose two new coding schemes, in which the second scheme is built successively on top of the first one to illustrate the effect of each technique used. \begin{itemize} \item The first scheme is called \emph{partial decode-forward binning (PDF-binning)}, which utilizes rate splitting, block Markov encoding, partial decode-forward relaying, Gelfand-Pinsker binning and forward joint decoding across two blocks. $S_1$ divides its message into two parts: one as a private message sent to $T_1$ directly, the other as a forwarding message, which is sent to $T_1$ with the help of $S_2$. $S_2$ first causally decodes the forwarding message part from $S_1$, then uses the decoded codeword as the binning state. In this case, however, the binning also allows $S_2$ to forward a part of the state to $T_1$, who then uses joint decoding across two blocks to decode its messages from both $S_1$ and $S_2$. Different from state amplification in \cite{Kim2008TIT}, here we want to decode the state at a different receiver ($T_1$) from decoding the message ($T_2$). This scheme achieves the partial decode-forward relaying rate for user 1 and Gelfand-Pinsker rate for user 2. \item The second scheme is called \emph{Han-Kobayashi PDF-binning (HK-PDF-binning)}, which combines PDF-binning with Han-Kobayashi coding by having both users further split their messages. $S_1$ divides its message into three parts: one as the Han-Kobayashi (HK) private message decoded only at $T_1$, another as the HK public message decoded at both $T_1$ and $T_2$, and the final part as the forwarding message. $S_2$ divides its message into two parts: one as the HK private message and the other as the HK public message. There are three ideas additional to PDF-binning. First, in performing partial decode-forward, $S_2$ uses conditional binning instead of traditional binning to bin only its private message part. Second, although $T_1$ uses joint decoding in both schemes, the decoding rule here is relaxed as $T_1$ also decodes the public message from $S_2$ without requiring it to be correct. Third, instead of simple Gelfand-Pinsker decoding, $T_2$ uses joint decoding of the binning auxiliary random variable and the HK public messages from the two senders which are encoded independently of the state. HK-PDF-binning achieves both the Han-Kobayashi and the PDF-binning rate regions. \end{itemize} \subsubsection{Half-duplex case} For the half-duplex CRC, the transmission is divided into two phases as in Figure \ref{fig:FD_and_HD_DM_CRC} (b). In the first phase, $S_1$ transmits to $S_2$, $T_1$ and $T_2$. In the second phase, the two senders transmit messages simultaneously, during which $S_2$ can both relay and apply cognitive encoding. We adapt the above two coding schemes to the half-duplex case. The main challenges in adapting full-duplex schemes to the half-duplex mode include deciding which message parts should be sent in which phase and changing the destination decoding rule to joint decoding across both phases. Specifically, we propose two half-duplex (HD) schemes: HD-PDF-binning and HD-HK-PDF-binning. At the end of the first phase in both schemes, $S_2$ decodes a message part from $S_1$ then applies PDF-binning, but neither $T_1$ nor $T_2$ decode here. Both $T_1$ and $T_2$ only decode at the end of the second phase. There are several differences from full-duplex coding. First, not all message parts are sent in each phase. Second, there is no need for block Markovity, instead, we superposition codewords in the two phases of the same block. Third, we use joint decoding at the destinations over two phases of the same block instead of over two consecutive blocks. \subsubsection{Applications to Gaussian channels} When applied to the Gaussian channel, a major difference between PDF-binning and the traditional binning in dirty paper coding (DPC) \cite{Costa83TIT} is that we introduce a correlation between the transmit signal and the state. This correlation allows both binning and forwarding at the same time, thus helps improve the transmission rate for the first user and still allows the second user to achieve the interference-free rate. We derive the closed-form optimal binning parameter for each coding scheme. This PDF-binning parameter contains the DPC-binning parameter as a special case. Results show that the HK-PDF-binning scheme outperforms all existing schemes in both the full- and half-duplex modes for the cognitive relay channel. Our analysis also shows clearly the impact on rate region of each of the techniques used. Furthermore, the maximum rate for the primary sender is the rate of partial decode-forward relaying and the maximum rate for the secondary sender is the interference-free rate as in dirty paper coding. \section{CRC Channel Models} \subsection{Full-duplex DM-CRC model} The full-duplex cognitive relay channel consists of two input alphabet $\mathcal{X}_1, \mathcal{X}_2$, and three output alphabets $\mathcal{Y}_1, \mathcal{Y}_2, \mathcal{Y}$. The channel is characterized by a channel transition probability $p(y_1,y_2, y|x_1,x_2)$, where $x_1$ and $x_2$ are the transmit signals of $S_1$ and $S_2$, $y_1$, $y_2$ and $y$ are the received signals of $T_1$, $T_2$ and $S_2$. Figure \ref{fig:FD_and_HD_DM_CRC}(a) illustrates the channel model, where $W_1$ and $W_2$ are the messages of $S_1$ and $S_2$. For notation, we use upper case letters to indicates random variables and lower case letters to indicate their realizations. We use $x^n$ and $x_k^n$ to represent the vectors ($x_1, \ldots, x_n$) and ($x_k, \ldots, x_n$) respectively. The cognitive relay channel has tight relationships with the interference and the relay channels. For example, this channel model can be converted to the interference channel\cite{Carleial78} if $S_2$ does not forward any information to $T_1$. Similarly, this channel reduces to the relay channel \cite{Meulen71}, \cite{Cover79} if $S_2$ does not have any message for $T_2$. A ($2^{nR_1}$, $2^{nR_2}$, $n$) code, or a communication strategy for $n$ channel uses with rate pair ($R_{1}, R_2$), consists of the following: \begin{itemize} \item Two message sets $\mathcal{W}_{1}\times\mathcal{W}_{2}=[1, 2^{nR_1}]\times[1, 2^{nR_2}]$ and independent messages $W_{1}, W_2 $ uniformly distributed over $\mathcal{W}_{1}$ and $\mathcal{W}_{2}$, respectively. \item Two encoders: one maps message $w_{1}$ into codeword $x_1^n(w_1) \in \mathcal{X}_1^n$, and one maps $w_2$ and each received sequence $y^{k-1}$ into a symbol $x_{2k}(w_2,y^{k-1}) \in \mathcal{X}_2$. \item Two decoders: one maps $y_1^n$ into $\hat{w}_{1} \in \mathcal{W}_{1}$; one maps $y_2^n$ into $\hat{w}_2 \in \mathcal{W}_2$. \end{itemize} The probability of error when the message pair ($W_{1}, W_2$) is sent is defined as $P_e(W_{1}, W_2) = P\{(\widehat{W}_{1}, \widehat{W}_2) \neq (W_{1}, W_2)\}$. A rate pair ($R_{1}, R_2$) is said to be \emph{achievable} if, for any $\epsilon > 0 $, there exists a code such that the average error probability $P_e\leq \epsilon$ as $n\rightarrow\infty$. The capacity region is the convex closure of the set of all achievable rate pairs. \subsection{Half-duplex DM-CRC model} The half-duplex cognitive relay channel also consists of four nodes: two senders $S_1$, $S_2$ and two receivers $T_1$, $T_2$. $S_1$ wants to send a message to $T_1$. $S_2$ serves as a causal relay node and helps forward messages from $S_1$ to $T_1$, while also sending its own message to $T_2$. The transmission in the half-duplex mode is divided into two phases. In the first phase, $S_1$ transmits its message and $S_2$, $T_1$ and $T_2$ listen. In the second phase, both $S_1$ and $S_2$ transmit and $T_1$ and $T_2$ listen. This 2-phase transmission allows, for example, $S_2$ to decode a part of the message from $S_1$ in the first phase and then forwards this part with its own message in the second phase. Formally, the half-duplex cognitive relay channel consists of three input alphabet $\mathcal{X}_{11}$, $\mathcal{X}_{12}$, $\mathcal{X}_{22}$, and five output alphabets $\mathcal{Y}_{11}$, $\mathcal{Y}_{21}$, $\mathcal{Y}$, $\mathcal{Y}_{12}$, $\mathcal{Y}_{22}$. The channel is characterized by a channel transition probability \\$p_c(y_{11},y_{21}, y, y_{12},y_{22},|x_{11},x_{12},x_{22})$ defined as \begin{align}\label{eq:p_c} p_c(y_{11},y_{21}, y, y_{12},y_{22},|x_{11},x_{12},x_{22}) \begin{cases} p(y_{11},y_{21},y|x_{11}) &\quad \text{if $0 \leq t \leq \tau$}, \\ p(y_{12},y_{22}|x_{12},x_{22}) &\quad \text{if $\tau \leq t \leq 1$}, \end{cases} \end{align} where $t$ is the normalized transmission time within $1$ block, $x_{11}$ and $x_{21}$ refer to the transmit signals of $S_1$ in the first and second phases, respectively; $x_{22}$ refers to the transmit signal of $S_2$ in the second phase ($S_2$ does not send any signal in the first phase); $y_{11}$ and $y_{12}$ are the received signals of $T_1$ in the first and second phases; $y_{21}$ and $y_{22}$ are the received signals of $T_2$ in the two phases; and $y$ is the received signal of $S_2$ in the first phase. We assume the channel is memoryless. Figure \ref{fig:FD_and_HD_DM_CRC}(b) illustrates the channel model, where $W_1$ and $W_2$ are the messages of $S_1$ and $S_2$. We use the notation $x^{\tau n} = (x_1, x_2, \cdots, x_{\tau n})$ and $x^{\bar{\tau}n} = (x_{\tau n+1}, \cdots, x_n )$, which correspond to the codewords sent during the first and second phases. A ($2^{nR_1}$, $2^{nR_2}$, $n$) code, or a communication strategy for $n$ channel uses with rate pair ($R_{1}, R_2$), consists of the following: \begin{itemize} \item Two message sets $\mathcal{W}_{1}\times\mathcal{W}_{2}=[1, 2^{nR_1}]\times[1, 2^{nR_2}]$ and independent messages $W_{1}, W_2 $ that are uniformly distributed over $\mathcal{W}_{1}$ and $\mathcal{W}_{2}$. \item Three encoders: two that map message $w_{1}$ into codewords $x_{11}^n(w_1) \in \mathcal{X}_{11}^n$ and $x_{12}^n(w_1) \in \mathcal{X}_{12}^n$, and one that maps $w_2$ and $y^{\tau n}$ into a codeword $x_{22}^n(w_2,y^{\tau n}) \in \mathcal{X}_{22}^n$. \item Two decoders: One maps $y_1^n$ into $\hat{w}_{1} \in \mathcal{W}_{1}$; and one maps $y_2^n$ into $\hat{w}_2 \in \mathcal{W}_2$. \end{itemize} The probability of error, achievable rate and capacity region are defined in a similar way to the full-duplex case. \section{Full-duplex partial decode-forward binning schemes} \subsection{PDF-binning scheme}\label{sec:FD_DMC_Bin} The first scheme uses block Markov superposition encoding at $S_1$ and partial decode-forward relaying and Gelfand-Pinsker binning at $S_2$. $T_1$ uses joint decoding across two blocks while $T_2$ uses normal Gelfand-Pinsker decoding. The first sender $S_1$ splits its message $w_1$ into two parts ($w_{10}, w_{11}$), which correspond to the common (forwarding) and private parts. We use block Markov encoding at $S_1$, such that the current-block common message $w_{10}$ is superimposed on the previous-block common message $w_{10}'$. Then, message $w_{11}$ is superimposed on both $w_{10}'$ and $w_{10}$. The second sender $S_2$ decodes the previous common message $w_{10}'$ from the first sender $S_1$ then uses binning to bin against the codeword for this message part. Depending on the joint distribution between the binning auxiliary random variable and the state that $S_2$ can also forward a part of the state (i.e. message $w_{10}'$) to $T_1$. The encoding and decoding structure can be seen in Figure \ref{fig:FD_DM_PDF_Binning}, in which $w_{10}'$ corresponds to $w_{10[i-1]}$. \begin{figure}[t] \centering \includegraphics[scale=0.55]{FD_DM_PDF_Binning.pdf} \caption{Coding structure for the full-duplex PDF-binning scheme at block i. (SP stands for superposition) } \label{fig:FD_DM_PDF_Binning} \end{figure} \begin{thm} \label{thm:FD_DMC_Bin} The convex hull of the following rate region is achievable for the full-duplex cognitive relay channel using PDF-binning: \begin{align}\label{eq:FD_DMC_Bin} \bigcup_{\substack{P_{1}}} \left\{\begin{array}{ll} R_{1} &\leq I(U_{10}; Y|T_{10})+ I(X_1;Y_1|U_{10},T_{10})\\ R_{1} &\leq I(T_{10}, U_{10} , X_1; Y_1)\\ R_2 &\leq I(U_2;Y_2) - I( U_2; T_{10}) \end{array}\right. \end{align} where \begin{align*} P_{1} = p(t_{10})p(u_{10}|t_{10})p(x_1|t_{10}, u_{10})p(u_2|t_{10}) p(x_2|t_{10},u_2)p(y_1,y_2,y|x_1,x_2) \end{align*} \end{thm} \begin{rem}The maximum rate for each user. \begin{itemize} \item The first user $S_1$ achieves the maximum rate of partial decode-forward relaying if we set $U_2=\emptyset$, $X_2=T_{10}$. \begin{align}\label{eq:FD_DMC_Bin_maxR1} &R_{1}^{\max} =\max_{\substack{p(u_{10},x_2)p(x_1|u_{10},x_2)}} \min \{I(U_{10};Y|X_2)+I(X_1;Y_1|U_{10},X_2), I(X_1,X_2;Y_1)\} \end{align} In this case, there is no binning but only forwarding at $S_2$. \item The second user $S_2$ achieves the maximum rate of Gelfand-Pinsker's binning if we set $T_{10}=U_{10}=X_1$. \begin{align}\label{eq:FD_DMC_Bin_maxR2} R_{2}^{\max} = \max_{\substack{p(x_1,u_2)p(x_2|x_1,u_2)}} \{I(U_2;Y_2) - I(U_2;X_1)\} \end{align} In this case, there is no forwarding of the state at $S_2$. \end{itemize} \end{rem} \begin{proof} The transmission is done in $B$ blocks, each consists of $n$ channel uses. $S_1$ splits each message $w_1$ into two independent parts $(w_{10}, w_{11})$. During the first $B-1$ blocks, $S_1$ encodes and sends a message tuple ($w_{10[i-1]}, w_{10i},w_{11i}) \in [1,2^{nR_{10}}]\times [1,2^{nR_{10}}]\times [1,2^{nR_{11}}]$; $S_2$ encodes and sends message ($w_{10[i-1]},w_{2i}) \in [1,2^{nR_{10}}]\times[1,2^{nR_{2}}]$, where $i=1,2,\ldots,B-1$ denotes the block index. When $B\rightarrow \infty$, the average rate triple $\left(R_{10}\frac{B-1}{B},R_{11}\frac{B-1}{B},R_2\frac{B-1}{B}\right)$ approaches to ($R_{10},R_{11},R_2$). We use random codes and fix a joint probability distribution \begin{align*} p(t_{10})p(u_{10}|t_{10})p(x_1|t_{10}, u_{10})p(u_2|t_{10})p(x_2|t_{10},u_2). \end{align*} \subsubsection{Codebook generation} For each block i (We can also just generate two independent codebooks for the odd and even blocks to make the error events of two consecutive blocks independent \cite{Liang07a}.): \begin{itemize} \item Independently generate $2^{nR_{10}}$ sequences $t_{10}^n \sim \prod_{k=1}^{n}p(t_{10k})$. Index these codewords as $t_{10}^n(w_{10}')$, $w_{10}'\in [1,2^{nR_{10}}]$. \item For each $ t_{10}^n( w _{10} ' )$, independently generate $2^{nR_{10}}$ sequences $u_{10}^n \sim \prod_{k=1}^{n}p(u_{10k}|t_{10k})$. Index these codewords as $ u_{10}^n( w_{10}|w _{10} ')$, $w_{10} \in [1,2^{nR_{10}}]$. $w_{10}$ contains the common message of the current block, while $w_{10}'$ contains the common message of the previous block. \item For each $ t_{10}^n( w _{10} ' )$ and $ u_{10}^n( w_{10}|w _{10} ')$, independently generate $2^{nR_{11}}$ sequences $x_1^n \sim \prod_{k=1}^{n}p(x_{1k}|t_{10k}, u_{10k}$). Index these codewords as $x_1^n(w_{11},w_{10}|w_{10}')$, $w_{11} \in [1,2^{nR_{11}}]$, $w_{10} \in [1,2^{nR_{10}}]$. \item Independently generate $2^{n(R_{2}+ R_{2}')}$ sequences $u_2^n \sim \prod_{k=1}^{n}p(u_{2k})$. Index these codewords as $ u_2^n( w_{2}, v_{2})$, $w_{2} \in [1,2^{nR_{2}}]$ and $v_{2} \in [1,2^{nR_{2}'}]$. \item For each $t_{10}(w _{10} ')$ and $ u_2^n( w _{2}, v_{2})$, generate one $x_2^n \sim \prod_{k=1}^{n}p(x_{2k}|t_{10k}, u_{2k})$. Denote $x_2^n$ by $x_2^n(w_{10} ', w_{2}, v_2)$. \end{itemize} \subsubsection{Encoding} At the beginning of block $i$, let ($w_{10i},w_{11i},w_{2i}$) be the new messages to be sent in block $i$, and ($w_{10[i-1]},w_{11[i-1]},w_{2[i-1]}$) be the messages sent in block $i-1$. \begin{itemize} \item $S_1$ knows $w_{10[i-1]}$, in order to send ($w_{10i},w_{11i}$), $S_1$ transmits $x_1^n(w_{11i},w_{10i}|w_{10[i-1]})$. \item $S_2$ searches for a $v_{2i}$ such that \begin{align} (t_{10}^n{(w_{10[i-1]})}, u_2^n( w_{2i}, v_{2i}) ) \nonumber \in A_\epsilon^{(n)}(P_{T_{10} U_2}). \end{align} Such a $v_{2i}$ exists with high probability if \begin{align}\label{eq:FD_DMC_Bin_begin} R_{2}' \geq I( U_2; T_{10}). \end{align} $S_2$ then transmits $x_2^n(w_{10[i-1]}, w_{2i}, v_{2i})$ . \end{itemize} \subsubsection{Decoding} At the end of block $i$: \begin{itemize} \item $S_2$ knows $w_{10[i-1]}$ and declares message $\hat{w}_{10i}$ was sent if it is the unique message such that \begin{align*} (t_{10}^n{(w_{10[i-1]})}, u_{10}^n( \hat{w}_{10i} | w_{10[i-1]}), y^n(i)) \in A_\epsilon^{(n)}(P_{T_{10} U_{10} Y}), \end{align*} where $y^n(i)$ indicates the received signal at $S_2$ in block $i$. We can show that the decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{10} &\leq I(U_{10}; Y|T_{10}). \end{align} \item $T_1$ knows $w_{10[i-2]}$ and decodes ($w_{10[i-1]},w_{11[i-1]}$) based on the signals received at block $i-1$ and block $i$. It declares that message pair ($\hat{w}_{10[i-1]}, \hat{w}_{11[i-1]}$) was sent if it is the unique pair such that \begin{align*} ( t_{10}^n{(w_{10[i-2]})}, u_{10}^n(\hat{w}_{10[i-1]} | w_{10[i-2]}),x_1^n(\hat{w}_{11[i-1]},\hat{w}_{10[i-1]}|w_{10[i-2]}), y_1^n(i-1))&\in A_\epsilon^{(n)}(P_{T_{10} U_{10} X_1 Y_1})\\ \text{and} \quad (t_{10}^n{(\hat{w}_{10[i-1]})}, y_1^n(i)) &\in A_\epsilon^{(n)}(P_{T_{10} Y_1}). \end{align*} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{11} &\leq I(X_1;Y_1|U_{10},T_{10}) \nonumber \\ R_{10} + R_{11} &\leq I(T_{10}, U_{10} , X_1; Y_1). \end{align} \item $T_2$ treats $T_{10}$, a part of the signal from $S_1$, as the state and decodes $w_{2i}$ based on the signal received at block $i$. Specifically, $T_2$ decodes $w_{2i}$ directly using joint typicality between $u_2$ and $y_2$. It declares that message $\hat{w}_{2i}$ was sent if it is unique such that \begin{align*} (u_2^n(\hat{w}_{2i} ,\hat{v}_{2i} ), y_2^n(i)) \in A_\epsilon^{(n)}(P_{U_2 Y_2}) \end{align*} for some $\hat{v}_{2i}$. The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} \label{eq:FD_DMC_Bin_end} R_2 + R_2' &\leq I(U_2;Y_2). \end{align} \end{itemize} Let $R_1 = R_{10} + R_{11}$, apply Fourier-Motzkin Elimination \cite{Ziegler95} on constraints \eqref{eq:FD_DMC_Bin_begin}-\eqref{eq:FD_DMC_Bin_end}, we get the rate region in \eqref{eq:FD_DMC_Bin}. \end{proof} \begin{rem} While the idea of the basic PDF-binning scheme is straightforward, this scheme allows the understanding of binning to achieve the maximum rates of partial decode-forward relaying at user 1 as in \eqref{eq:FD_DMC_Bin_maxR1} and Gelfand-Pinsker coding at user 2 as in \eqref{eq:FD_DMC_Bin_maxR2}. The importance magnifies in the Gaussian application in Section \ref{sec:FD_Gaussian}. This scheme helps build the base for more complicated schemes later. \end{rem} \subsection{Han-Kobayashi PDF-binning scheme}\label{sec:FD_DMC_Combined} \begin{figure}[t] \centering \includegraphics[scale=0.55]{FD_DM_Combined.pdf} \caption{Coding structure for the full-duplex Han-Kobayashi PDF-binning scheme at block i. } \label{fig:FD_DM_Combined} \end{figure} Figure \ref{fig:FD_DM_Combined} illustrates the idea of the full-duplex Han-Kobayashi PDF-binning scheme. Built upon PDF-binning, each user further splits its message to incorporate Han-Kobayashi coding. Message $w_1$ is split into three parts: $w_{10}, w_{11}, w_{12}$, corresponding to the common (forwarding), public and private parts, and message $w_2$ is split into two parts: $w_{21}, w_{22}$, corresponding to the public and private parts. Take the transmission in block $i$ as an example. At $S_1$, the current common message $w_{10i}$ is superimposed on the previous commons message $w_{10[i-1]}$; message $w_{11i}$ is encoded independently of both $w_{10[i-1]}$ and $w_{10i}$; message $w_{12i}$ is then superimposed on all three messages $w_{10[i-1]}$, $w_{10i}$ and $w_{10i}$. $S_2$ decodes $\tilde{w}_{10[i-1]}$ of the previous block and uses conditional binning to bin its private part $w_{22[i]}$ against $\tilde{w}_{10[i-1]}$, conditionally on knowing the public part $w_{21[i]}$. At the end of block $i$, $T_1$ uses joint decoding over two blocks to decode a unique tuple ($\hat{w}_{10[i-1]},\hat{w}_{11[i-1]},\hat{w}_{12[i-1]})$ for some $\hat{w}_{21[i-1]}$ without requiring this message part to be correct. $T_2$ treats the codeword for $w_{10[i-1]}$ as the state and searches for a unique pair $(w_{21i}, w_{22i})$ for some $w_{11i}$. The detailed coding and decoding procedures are shown in the proof of Theorem \ref{thm:FD_DMC_combined} below. \begin{thm} \label{thm:FD_DMC_combined} The convex hull of the following rate region is achievable for the cognitive relay channel using HK-PDF-binning: \begin{align}\label{eq:FD_DMC_combined} \bigcup_{\substack{P_{2}}} \left\{\begin{array}{ll} R_{1} &\leq \min\{I_2 + I_5, I_6\}\\ R_{2} &\leq I_{12}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_7,I_8 \} + I_{13}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_3,I_4\} + I_{14}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_9, I_{10} \} + I_{11} - I_1 \\ 2R_{1} + R_{2} &\leq \min\{ I_2 + I_3,I_4\} +\min\{I_2 + I_9, I_{10} \} + I_{13}- I_1\\ R_{1} + 2R_{2} &\leq \min\{I_2 + I_7, I_8 \}+ I_{11} - I_1 + I_{14}- I_1 \end{array}\right. \end{align} where \begin{align} P_{2} = &p(t_{10})p(u_{10}|t_{10})p(u_{11})p(x_1|t_{10}, u_{10}, u_{11})p(u_{21} p(u_{22}|u_{21}, t_{10})p(x_2|t_{10},u_{21},u_{22})p(y_1,y_2,y|x_1,x_2), \end{align} and $I_1$ --- $I_{14}$ are defined as \begin{align} I_1 &= I( U_{22}; T_{10}|U_{21}) \nonumber \\ I_2 &= I(U_{10}; Y|T_{10}) \nonumber \\ I_3 &= I(X_1;Y_1|T_{10},U_{10}, U_{11}, U_{21}) \nonumber\\ I_4 &= I(U_{10}, X_1; Y_1|T_{10}, U_{11}, U_{21}) + I(T_{10};Y_1) \nonumber \\ I_5 &= I(U_{11}, X_1; Y_1|T_{10}, U_{10}, U_{21}) \nonumber \\ I_6 &= I(U_{10}, U_{11}, X_1; Y_1|T_{10}, U_{21})+ I(T_{10};Y_1) \nonumber \\ I_7 &= I(X_1, U_{21};Y_1|T_{10},U_{10}, U_{11}) \nonumber \\ I_8 &= I(U_{10}, X_1, U_{21}; Y_1|T_{10}, U_{11}) + I(T_{10};Y_1) \nonumber \\ I_9 &= I(U_{11}, X_1, U_{21}; Y_1|T_{10}, U_{10}) \nonumber \\ I_{10} &= I(T_{10}, U_{10}, U_{11}, X_1, U_{21}; Y_1) \nonumber\\ I_{11} &= I(U_{22};Y_2|U_{21},U_{11}) \nonumber \\ I_{12} &= I(U_{21},U_{22};Y_2|U_{11})\nonumber \\ I_{13} &= I(U_{11},U_{22};Y_2|U_{21})\nonumber \\ I_{14} &= I(U_{11},U_{21},U_{22};Y_2). \end{align} \end{thm} \begin{rem}\label{rem:FD_DMC_Combined_Inclusions} Inclusion of PDF-binning and Han-Kobayashi schemes. \begin{itemize} \item The HK-PDF-binning scheme becomes PDF-binning if $U_{11}=U_{21}=\emptyset$. \item The HK-PDF-binning scheme becomes the Han-Kobayashi scheme if $T_{10}=U_{10}=\emptyset$ and $X_{2} = U_{22}$. \item The maximum rates for $S_1$ and $S_2$ are the same as in the PDF-binning scheme in \eqref{eq:FD_DMC_Bin_maxR1} and \eqref{eq:FD_DMC_Bin_maxR2}. \end{itemize} \end{rem} \begin{proof} We use random codes and fix a joint probability distribution \begin{align*} p(t_{10})p(u_{10}|t_{10})p(u_{11})p(x_1|t_{10}, u_{10}, u_{11} p(u_{21})p(u_{22}|u_{21}, t_{10})p(x_2|t_{10},u_{21},u_{22}). \end{align*} \subsubsection{Codebook generation} For each block i (or for odd and even blocks): \begin{itemize} \item Independently generate $2^{nR_{10}}$ sequences $t_{10}^n \sim \prod_{k=1}^{n}p(t_{10k})$. Index these codewords as $t_{10}^n(w_{10}')$, $w_{10}'\in [1,2^{nR_{10}}]$. \item For each $t_{10}^n(w_{10}')$, independently generate $2^{nR_{10}}$ sequences $u_{10}^n \sim \prod_{k=1}^{n}p(u_{10k}|t_{10k})$. Index these codewords as $ u_{10}^n( w_{10}|w_{10} ')$, $w_{10} \in [1,2^{nR_{10}}]$. $w_{10}$ is the common message of the current block, while $w_{10}'$ is the common message of the previous block. \item Independently generate $2^{nR_{11}}$ sequences $u_{11}^n \sim \prod_{k=1}^{n}p(u_{11k})$. Index these codewords as $u_{11}^n(w_{11})$, $w_{11}\in [1,2^{nR_{11}}]$. \item For each $t_{10}^n(w_{10}')$, $u_{10}^n(w_{10}|w_{10}')$ and $u_{11}^n(w_{11})$, independently generate $2^{nR_{12}}$ sequences\\ $x_1^n \sim \prod_{k=1}^{n}p(x_{1k}|t_{10k}, u_{10k}, u_{11k}$). Index these codewords as $x_1^n(w_{12}|w_{11},w_{10},w_{10}')$, $w_{12} \in [1,2^{nR_{12}}]$. \item Independently generate $2^{nR_{21}}$ sequences $u_{21}^n \sim \prod_{k=1}^{n}p(u_{21k})$. Index these codewords as $u_{21}^n(w_{21})$, $w_{21}\in [1,2^{nR_{21}}]$. \item For each $u_{21}^n(w_{21})$, independently generate $2^{n(R_{22}+ R_{22}')}$ sequences $u_{22}^n \sim \prod_{k=1}^{n}p(u_{22k}|u_{21k})$. Index these codewords as $u_{22}^n(w_{22}, v_{22}|w_{21})$, $w_{22} \in [1,2^{nR_{22}}]$ and $v_{22} \in [1,2^{nR_{22}'}]$. \item For each $t_{10}(w_{10}')$, $u_{21}^n(w_{21})$ and $ u_{22}^n(w_{22}, v_{22}|u_{21})$, generate one $x_2^n \sim \prod_{k=1}^{n}p(x_{2k}|t_{10k}, u_{21i}, u_{22i})$. Denote $x_2^n$ by $x_2^n(w_{10}', w_{21}, w_{22}, v_{22})$. \end{itemize} \subsubsection{Encoding} At the beginning of block $i$, let ($w_{10i},w_{11i},w_{12i},w_{21i},w_{22i}$) be the new messages to be sent in block $i$, and ($w_{10[i-1]},w_{11[i-1]},w_{12[i-1]},w_{21[i-1]},w_{22[i-1]}$) be the messages sent in block $i-1$. \begin{itemize} \item $S_1$ knows $w_{10[i-1]}$, in order to send ($w_{10i},w_{11i},w_{12i}$), it transmits $x_1^n(w_{12}|w_{11i},w_{10i},w_{10[i-1]})$. \item $S_2$ searches for a $v_{22i}$ such that \begin{align}\label{eq:FD_DMC_Bin_binstep} (t_{10}^n{(w_{10[i-1]})}, u_{21}^n(w_{21i}), u_{22}^n(w_{22i}, v_{22i}|w_{21i}) ) \in A_\epsilon^{(n)}(P_{T_{10} U_{22}|U_{21}}). \end{align} Such a $v_{22i}$ exists with high probability if \begin{align}\label{eq:FD_DMC_combined_begin} R_{22}' \geq I( U_{22}; T_{10}|U_{21}). \end{align} $S_2$ then transmits $x_2^n(w_{10[i-1]},w_{21i}, w_{22i}, v_{22i})$. \end{itemize} \subsubsection{Decoding} At the end of block $i$: \begin{itemize} \item $S_2$ knows $w_{10[i-1]}$ and declares message $\hat{w}_{10i}$ was sent if it is the unique message such that \begin{align*} (t_{10}^n{(w_{10[i-1]})}, u_{10}^n(\hat{w}_{10i}|w_{10[i-1]}),y^n(i)) \in A_\epsilon^{(n)}(P_{T_{10} U_{10} Y}), \end{align*} where $y^n(i)$ indicates the received signal at $S_2$ in block $i$. We can show that the decoding error probability goes to 0 when $n\rightarrow \infty$ if \begin{align} R_{10} &\leq I(U_{10}; Y|T_{10}). \end{align} \item $T_1$ knows $w_{10[i-2]}$ and searches for a unique tuple ($\hat{w}_{10[i-1]},\hat{w}_{11[i-1]},\hat{w}_{12[i-1]})$ for some $\hat{w}_{21[i-1]}$ such that \begin{align}\label{eq:FD_DMC_Combined_Dec_T1} ( t_{10}^n{(w_{10[i-2]})}, u_{10}^n(\hat{w}_{10[i-1]}|w_{10[i-2]}),u_{11}^n{(w_{11[i-1]})}, &x_1^n(\hat{w}_{12[i-1]}|\hat{w}_{11[i-1]},\hat{w}_{10[i-1]},w_{10[i-2]}), \nonumber\\ u_{21}^n(\hat{w}_{21[i-1]}), y_1^n(i-1))&\in A_\epsilon^{(n)}(P_{T_{10} U_{10} U_{11} X_1 U_{21} Y_1})\nonumber\\ \text{and} \quad (t_{10}^n{(\hat{w}_{10[i-1]})}, y_1^n(i)) &\in A_\epsilon^{(n)}(P_{T_{10} Y_1}). \end{align} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{12} &\leq I(X_1;Y_1|T_{10},U_{10}, U_{11}, U_{21}) \nonumber \\ R_{10} + R_{12} &\leq I(U_{10}, X_1; Y_1|T_{10}, U_{11}, U_{21}) + I(T_{10};Y_1)\nonumber \\ R_{11} + R_{12} &\leq I(U_{11}, X_1; Y_1|T_{10}, U_{10}, U_{21})\nonumber \\ R_{10} + R_{11} + R_{12} &\leq I(U_{10}, U_{11}, X_1; Y_1|T_{10}, U_{21} + I(T_{10};Y_1)\nonumber \\ R_{12} + R_{21} &\leq I(X_1, U_{21};Y_1|T_{10},U_{10}, U_{11}) \nonumber \\ R_{10} + R_{12} + R_{21} &\leq I(U_{10}, X_1, U_{21}; Y_1|T_{10}, U_{11}) + I(T_{10};Y_1)\nonumber \\ R_{11} + R_{12} + R_{21} &\leq I(U_{11}, X_1, U_{21}; Y_1|T_{10}, U_{10})\nonumber \\ R_{10} + R_{11} + R_{12} + R_{21} &\leq I(T_{10}, U_{10}, U_{11}, X_1, U_{21}; Y_1). \end{align} \item $T_2$ treats $T_{10}^n{(w_{10[i-1]}')}$ as the state and decodes $(w_{21i}, w_{22i}, v_{22i})$ based on the signal received in block $i$. Specifically, $T_2$ searches for a unique $(\hat{w}_{21i}, \hat{w}_{22i})$ for some $(\hat{w}_{11i}, \hat{v}_{22i})$ such that \begin{align}\label{eq:FD_DMC_Combined_Dec_T2} (u_{11}^n(\hat{w}_{11i}), u_{21}^n(\hat{w}_{21i}), &u_{22}^n(\hat{w}_{22i},\hat{v}_{22i}|\hat{w}_{21i}), y_2^n(i)) \in A_\epsilon^{(n)}(P_{U_{11} U_{21} U_{22} Y_2}). \end{align} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align}\label{eq:FD_DMC_combined_end} R_{22} + R_{22}' &\leq I(U_{22};Y_2|U_{21},U_{11}) \nonumber \\ R_{21} + R_{22} + R_{22}' &\leq I(U_{21},U_{22};Y_2|U_{11}) \nonumber \\ R_{11} + R_{22} + R_{22}' &\leq I(U_{11},U_{22};Y_2|U_{21}) \nonumber \\ R_{11} + R_{21} + R_{22} + R_{22}' &\leq I(U_{11},U_{21},U_{22};Y_2). \end{align} \end{itemize} Applying Fourier-Motzkin Elimination to \eqref{eq:FD_DMC_combined_begin}-\eqref{eq:FD_DMC_combined_end}, we get rate region \eqref{eq:FD_DMC_combined}. \end{proof} \begin{rem}\label{rem:FD_DMC_Combined_features} Several features of the HK-PDF-binning scheme are worth noting: \begin{itemize} \item In encoding, $w_{10}$ and $w_{11}$ are encoded independently, then $w_{12}$ is superpositioned on both. This independent coding between the forwarding part ($w_{10}$) and Han-Kobayashi public part ($w_{11}$), rather than superposition, is important to ensure the rate region includes both PDF-binning and Han-Kobayashi regions. \item In the binning step \eqref{eq:FD_DMC_Bin_binstep} at $S_2$, we use conditional binning instead of the usual (unconditional) binning. The binning is only between the Han-Kobayashi private message part ($w_{22}$) and the state ($w_{10}'$), conditionally on knowing the Han-Kobayashi public messsage part $w_{21}$. This conditional binning is possible since $w_{21}$ is decoded at both destinations. \item In the decoding step \eqref{eq:FD_DMC_Combined_Dec_T2} at $T_2$, we use joint decoding of both the Gelfand-Pinsker auxiliary random variable ($u_{22}$) and the Han-Kobayashi public message parts ($w_{11}$ and $w_{21}$), instead of decoding Gelfand-Pinsker and Han-Kobayashi codewords separately. This joint decoding is possible since the codewords for $w_{11}$ and $w_{21}$ (i.e. $u_{11}^n$ and $u_{21}^n$) are independent of the state in Gelfand-Pinsker coding (i.e. $t_{10}^n$). Joint decoding at both $T_1$ \eqref{eq:FD_DMC_Combined_Dec_T1} and $T_2$ \eqref{eq:FD_DMC_Combined_Dec_T2} help achieve the largest rate region for this coding structure. \end{itemize} \end{rem} \subsection{Comparison with existing schemes for the interference channel with source cooperation}\label{sec:FD_Tx_coop_compare} In this section, we analyze in detail two existing schemes \cite{CaoISIT2007, YangTIT2011} for the interference channel with source cooperation, which are most closely related to the proposed schemes. The interference channel with source cooperation is a 4-node channel in which both $S_1$ and $S_2$ can receive signal from each other and use that cooperatively in sending messages to $T_1$ and $T_2$. This channel therefore includes the CRC as a special case (when $S_2$ sends no information to $S_1$). \subsubsection{IC with conferencing} Cao and Chen \cite{CaoISIT2007} propose an achievable rate region for the interference channel with source cooperation based on rate splitting, block Markov encoding, superposition encoding, dirty paper coding and random binning. Each user splits its message into three parts: common, private and cooperative messages and divides the cooperative message into cells. The second user generates independent codewords for the current common message ($u_2^n$), previous cooperative cell index ($s_2^n$) and current cooperative message ($w_2^n$). The codewords for the current private message are then superimposed on the current common message and previous cooperative cell index ($v_2^n|u_2^n, s_2^n$). Then, the first user treats the previous cooperative-cell-index codeword ($s_2^n$) as the state and jointly bins its codewords for the current common message ($n_1^n$), previous cooperative cell index ($h_1^n$) and current cooperative message ($g_1^n$). Finally, the codewords for the first user's private message ($m_1^n$) is conditionally binned with $s_2^n$ given $n_1^n$ and $h_1^n$. A two-step decoding with list decoding is then used at each destination. The common, private and cooperative message parts in \cite{CaoISIT2007} correspond roughly to our HK public, HK private and forwarding (common) part, respectively. As such, when applied to the CRC, their scheme differs from the proposed HK-PDF-binning scheme in the following aspects: \begin{itemize} \item Block Markovity is applied only on the HK private part, whereas in our scheme, block Markovity is applied on all message parts. \item Block Markovity is based on cell division of the previous cooperative message, while in our scheme, block Markovity is on the whole previous common message. This, however, is a minor difference since if each cell contains only one message, then cell index reduces to message index. \item The first user bins both its HK public and private parts (the user labels are switched in \cite{CaoISIT2007}), whereas we only bin the HK private part (see Remark \ref{rem:FD_DMC_Combined_features}). \item The scheme in \cite{CaoISIT2007} cannot achieve the decode-forward relaying rate because of no block Markovity between the current cooperative-message codeword ($w_2^n$) and the previous cooperative-cell codeword ($s_2^n$). In other words, there is no coherent transmission between the source and relay, which can be readily verified from the code distribution. Consider setting $V_1 = V_2 = U_1 = U_2 = 0$, $M_1 = M_2 = N_1 = N_2 = 0$ and $W_2 = S_2 = G_2 = H_2 = 0$ in equation (8) of \cite{CaoISIT2007}, then the code distribution reduces to \begin{align*} p(q)p(g_{1}|q)p(h_{1}|q)p(x_{1}|g_{1},q)p(x_{2}|h_{1},q) = p(q,g_{1},x_{1})p(q,h_{1},x_{2}) \neq p(q, x_{1}, x_{2} ), \end{align*} where $q$ is the time sharing variable. This distribution implies that the first user splits its message into two parts and independently encodes each of them (by $g_1$ and $h_1$). The second user then decodes one part in $g_1$ and forwards this part to the destination. But because of the independence between $g_1$ and $h_1$, the achievable rate is less than in coherent decode-forward relaying. Thus, the claim in Remark 2 of \cite{CaoISIT2007} that this scheme achieves the capacity region of the degraded relay channel is in fact unfounded. \end{itemize} \subsubsection{IC with generalized feedback} Yang and Tuninetti \cite{YangTIT2011} propose two schemes for the interference channel with generalized feedback based on block Markov superposition coding, binning and backward decoding. Since the first scheme is a special case of the second, we only analyze their second scheme. Each user splits its message into four parts: cooperative common ($w_{10c}$), cooperative private ($w_{11c}$), non-cooperative common ($w_{10n}$) and non-cooperative private ($w_{11n}$). Consider the transmission in block $b$. First, generate independent codewords for the previous cooperative-common messages of both users ($Q^n(w_{10c,b-1}, w_{20c,b-1})$). Then the cooperative-common ($w_{10c,b}$), non-cooperative common ($w_{10n,b}$) and non-cooperative private ($w_{11n,b}$) messages are superimposed on each other successively as $V_1$, $T_1$, $U_1$, respectively (according to $p(v_1, t_1, u_1|q)$). There are three binning steps after the above codebook generation. First, the codewords $S_1$, $S_2$ for the previous cooperative-private messages of both users are binned with each other given $Q$. Second, $V_1$, $U_1$ and $T_1$ are binned with $S_1$ and $S_2$ given $Q$. Third, the codeword $Z_1$ for the cooperative-private message ($w_{11c,b}$) is conditionally binned with $S_2$, $U_1$ and $T_1$ given $V_1$, $S_1$ and $Q$. Backward decoding is used, in which each destination applies relaxed joint decoding of all interested messages. The non-cooperative messages in \cite{YangTIT2011} correspond to our HK public and private parts. Their scheme has two cooperative message parts (the common is decoded at both destinations while the private is not), whereas the proposed HK-PDF-binning has only one common part. To compare these two schemes, we consider the following two special settings to make the message parts equivalent: i) Set the cooperative-common message ($w_{10c}$) to $\emptyset$: Their cooperative private message then corresponds to our forwarding (common) message. Their scheme differs markedly from HK-PDF-binning as follows. \begin{itemize} \item User 1 uses binning among the three message parts instead of superposition coding as in HK-PDF-binning. Block Markov superposition is also replaced by binning with the codeword for the previous cooperative message. \item User 2 applies joint binning of both the non-cooperative common and private parts instead of conditional binning of only the non-cooperative private part, given the non-cooperative common part (see Remark \ref{rem:FD_DMC_Combined_features}). \end{itemize} ii) Set the cooperative-private message ($w_{11c}$) to $\emptyset$: Their cooperative common message then corresponds to our forwarding (common) message. Their scheme is more similar to HK-PDF-binning, but there are several important differences as follows. \begin{itemize} \item User 1 now uses superposition coding, but superimposes all three message parts successively, whereas we generate codewords for the forwarding part and the HK public part independently (see Remark \ref{rem:FD_DMC_Combined_features}). \item User 2 also applies joint binning of both non-cooperative message parts instead of conditional binning, similar to case i). \item Destination 2 decodes the cooperative-common part of user 1, thus limits the rate of user 1 to below the decode-forward relaying rate because of the extra rate constraint at destination 2 (this applies even with relaxed decoding). In our proposed scheme, the forwarding part of user 1 is not decoded at destination 2. \end{itemize} As a result, both schemes in \cite{CaoISIT2007} and \cite{YangTIT2011}, when applied to the CRC, achieve the Han-Kobayashi region but not the decode-forward relaying rate for the first user. Thus, the maximum rates for user 1 in both schemes are smaller than in \eqref{eq:FD_DMC_Bin_maxR1}. Another point is that, in both \cite{CaoISIT2007} and \cite{YangTIT2011}, joint decoding of both the state and the binning auxiliary random variables is used at the destinations, but this joint decoding is invalid and results in a rate region larger than is possible. In our proposed scheme, all message parts that are jointly decoded with the binning auxiliary variable at the second destination are encoded independently of the state. \begin{rem} Based on our analysis, we conjecture that splitting the common (forwarding) message further into two parts is not necessary for the CRC. In \cite{YangTIT2011, WangTIT2011}, the common message is split into two parts: one for decoding at the other destination and the other for binning. Our analysis shows that both these operations can be included in one-step binning by varying the joint distribution between the state and the auxiliary random variable. This joint distribution becomes apparent when applying to the Gaussian channel as in Section \ref{sec:FD_Gaussian} next. \end{rem} \begin{figure}[t] \centering \includegraphics[scale=0.45]{FD_Gaussian_CRC.pdf} \caption{The standard full-duplex Gaussian cognitive relay channel.} \label{fig:FD_Gaussian_CRC} \end{figure} \section{Full-duplex Gaussian CRC rate regions}\label{sec:FD_Gaussian} \subsection{Full-duplex Gaussian CRC model} In this section, we analyze the standard full-duplex Gaussian cognitive relay channel model as follows. \begin{align}\label{eq:FD_Gaussian_Y} Y_{1} &= X_1 + bX_2 + Z_1 \nonumber \\ Y_2 &= aX_1 + X_2 + Z_2 \nonumber \\ Y &= cX_1 + Z, \end{align} where $Z_1$, $Z_2$, $Z \sim \mathcal{N}(0,1)$ are independent Gaussian noises. Assume that the transmit signals $X_1$ and $X_2$ are subject to power constraints $P_1$ and $P_2$, respectively. The standard Gaussian CRC is shown in Figure \ref{fig:FD_Gaussian_CRC}. If the original channel is not in this standard form, we can always transform it into the standard form using a procedure similar to the interference channel \cite{Carleial78}. \subsection{Signaling and rates for full-duplex PDF-binning}\label{sec:signal_FD_PDF_Binning} In the Gaussian channel, the signals $T_{10}$, $U_{10}$, $U_2$, $X_1$ and $X_2$ of the PDF-binning scheme in Section \ref{sec:FD_DMC_Bin} can be represented as follows. \begin{align} T_{10} &= \alpha S_{10}'(w_{10}'),\nonumber\\ U_{10} &= \alpha S_{10}'(w_{10}') + \beta S_{10}(w_{10}), \nonumber \\ X_1 &= \alpha S_{10}'(w_{10}') + \beta S_{10}(w_{10}) + \gamma S_{11}(w_{11}), \nonumber \\ X_2 &= \mu\left(\rho S_{10}'(w_{10}')+\sqrt{1-\rho^2}S_{22}\right),\nonumber \\ U_2 &= X_2 + \lambda S_{10}' = (\mu\rho+\lambda)S_{10}' +\mu\sqrt{1-\rho^2}S_{22} .\label{eq:FD_Gaussian_PDF_Bin_signaling} \end{align} where $S_{10}'$, $S_{10}$, $S_{11}$ and $S_{22}$ are independent $\mathcal{N}(0,1)$ random variables to encode $w_{10}'$, $w_{10}$, $w_{11}$ and $w_{2}$ respectively. $U_2$ is the auxiliary random variable for binning that encodes $w_2$. $X_1$ and $X_2$ are the transmit signals of $S_1$ and $S_2$. The parameters $\alpha$, $\beta$, $\gamma$, $\mu$ are power allocation factors satisfying the power constraints \begin{align}\label{eq:FD_Gaussian_PDF_power_constraints} \alpha^2 + \beta^2 + \gamma^2 \leq P_1, \nonumber\\ \mu ^2 \leq P_2, \end{align} where $P_1$ and $P_2$ are transmit power constraints of $S_1$ and $S_2$. An important feature of the signaling design in \eqref{eq:FD_Gaussian_PDF_Bin_signaling} is $\rho$ ($-1 \leq \rho \leq 1$), the correlation factor between the transmit signal ($X_2$) and the state ($S_{10}'$) at $S_2$. In traditional dirty paper coding, the transmit signal and the state are independent. Here we introduce correlation between them, which includes dirty paper coding as a special case when $\rho=0$. This correlation allows both signal forwarding and traditional binning at the same time. $\lambda$ is the partial decode-forward binning parameter which will be optimized later. Substitute $X_1$, $X_2$ into $Y_1$, $Y_2$ and $Y$ in \eqref{eq:FD_Gaussian_Y}, we get \begin{align} Y_{1} &= (\alpha+b\mu\rho) S_{10}' + \beta S_{10} + \gamma S_{11} + b \mu\sqrt{1-\rho^2} S_{22} + Z_1, \nonumber \\ Y_{2} &= (a\alpha+\mu\rho) S_{10}' + a\beta S_{10} + a \gamma S_{11} + \mu\sqrt{1-\rho^2} S_{22} + Z_2, \nonumber \\ Y &= c \alpha S_{10}' + c \beta S_{10} + c \gamma S_{11} + Z. \end{align} \begin{cor}\label{cor:FD_Gaussian_PDF_Bin} The achievable rate region for the full-duplex Gaussian-CRC using the PDF-binning scheme is the convex hull of all rate pairs ($R_1$, $R_2$) satisfying \begin{align}\label{eq:FD_Gaussian_Bin} R_{1} &\leq C \left(\frac {c^2\beta ^2} {c^2\gamma ^2 + 1}\right) + C \left( \frac {\gamma ^2} {b^2\mu ^2(1-\rho^2) + 1} \right)\nonumber\\ R_{1} &\leq C \left( \frac {(\alpha+b\mu\rho)^2 + \beta^2 + \gamma^2} {b^2 \mu ^2(1-\rho^2) + 1} \nonumber\right)\\ R_2 &\leq C\left( \frac{\mu ^2(1-\rho^2)}{a^2\beta^2 + a^2\gamma^2 + 1} \right) \end{align} where $-1 \leq \rho \leq 1$, $C(x) = \frac{1}{2}\log (1+x)$, and the power allocation factors $\alpha$, $\beta$, $\gamma$ and $\mu$ satisfy the power constraints \eqref{eq:FD_Gaussian_PDF_power_constraints}. \end{cor} \begin{proof} Applying Theorem 1 with the signaling in \eqref{eq:FD_Gaussian_PDF_Bin_signaling}, we get the rate region in Corollary \ref{cor:FD_Gaussian_PDF_Bin}. \end{proof} \begin{rem}Maximum rates for each sender \begin{itemize} \item Setting $\rho=\pm 1$, $\mu =\rho \sqrt{P_2}$, we obtain the maximum rate for $R_1$ as in partial decode-forward relaying: \begin{align}\label{eq:FD_Gaussian_Bin_maxR1} R_{1}^{\max} = \max_{\substack{\alpha^2 + \beta^2 + \gamma^2 \leq P_1}} \min \Bigg \{&C \left( \frac{c^2 \beta^2}{c^2 \gamma^2+1} \right) + C(\gamma^2), C\left( \left(\alpha+b\sqrt{P_2}\right)^2 + \beta^2 + \gamma^2 \right) \Bigg \}. \end{align} \item Setting $\rho=0$, $\beta = \gamma = 0$ and $\mu=\sqrt{P_2}$, we obtain the maximum rate for $R_2$ as in dirty paper coding \begin{align}\label{eq:FD_Gaussian_Bin_maxR2} R_{2}^{\max} = C(P_2). \end{align} \end{itemize} \end{rem} \subsection{Optimal binning parameter for full-duplex PDF-binning} In this section, we derive in closed form the optimal binning parameter $\lambda$ for \eqref{eq:FD_Gaussian_PDF_Bin_signaling} to achieve rate region \eqref{eq:FD_Gaussian_Bin}. This optimal binning parameter is different from the optimal binning parameter in dirty paper coding, as we introduce the correlation factor $\rho$ between the transmit signal and the state. This correlation contains the function of message forwarding. For example, if we set $\rho=\pm 1$, $X_2$ will only encode $w_{10}'$ without any actual binning, hence realize the function of message forwarding. If we set $\rho=0$, PDF-binning becomes dirty paper coding without any message forwarding. For $0 < \left|{\rho} \right| < 1$, PDF-binning has both the functions of binning and message forwarding. Thus, PDF-binning generalizes dirty paper coding. \begin{thm}\label{thm:FD_Gaussian_bin_lambda} The optimal $\lambda$ for the full-duplex PDF-binning scheme is \begin{align}\label{eq:lambda_FD_Gaussian_Bin} \lambda^* = \frac{ a\alpha\mu^2(1-\rho^2) - \mu\rho(a^2\beta^2+a^2\gamma^2+1) }{a^2\beta^2 + a^2\gamma^2 + \mu^2(1-\rho^2) + 1}. \end{align} \end{thm} \begin{proof} The optimal $\lambda^*$ is obtained by maximizing both rates $R_1$ and $R_2$. In rate region \eqref{eq:FD_DMC_Bin}, through the Fourier Motzkin Elimination process, we can see that if we maximize the term $I(U_{2};Y_{2}) - I(U_{2};T_{10})$, both $R_1$ and $R_2$ are maximized simultaneously. We have \begin{align*} &I(U_{2};Y_{2}) - I(U_{2};T_{10}) \\ &= H(Y_2) - H(Y_2|U_2) - H(U_2) + H(U_2|T_{10})\\ &= H(Y_2) + H(U_2|T_{10}) - H(U_2, Y_2). \end{align*} Here $\lambda$ only affects the last term $H(U_2, Y_2)$. The covariance matrix between $U_{2}$ and $Y_{2}$ is \begin{align}\label{eq:FD_Gaussian_Bin_cov} \text{cov}(U_{2}, Y_{2}) = \begin{bmatrix} \text{var}(U_{2}) & \text{E}(U_{2},Y_{2})\\ \text{E}(U_{2},Y_{2}) & \text{var}(Y_{2}) \end{bmatrix}, \end{align} where \begin{align*} \text{var}(U_{2}) &= \mu ^2 + \lambda ^2 + 2\mu\rho\lambda, \\ \text{E}(U_{2},Y_{2})&=(\mu\rho+\lambda)(a\alpha+\mu\rho) + \mu^2(1-\rho^2),\\ \text{var}(Y_{2}) &= (a\alpha+\mu\rho)^2+ a^2\beta^2 + a^2\gamma^2 + \mu^2(1-\rho^2) + 1. \end{align*} Minimizing the determinant of the covariance matrix in \eqref{eq:FD_Gaussian_Bin_cov}, we obtain the optimal $\lambda^*$ in \eqref{eq:lambda_FD_Gaussian_Bin}. \end{proof} \begin{figure}[t] \centering \includegraphics[scale=0.45]{FD_PDFBin_effect_of_rho.pdf} \caption{Effect of the binning correlation factor $\rho$.} \label{fig:FD_PDFBin_effect_of_rho} \end{figure} \begin{rem}Effect of $\rho$: \begin{itemize} \item If $\rho=0$, $\lambda^*$ becomes the optimal $\lambda$ for traditional dirty paper coding \cite{Costa83TIT}, which achieves the maximum rate for $R_2$ as in \eqref{eq:FD_Gaussian_Bin_maxR2}. \item If $\rho=\pm 1$, $\lambda^*$ differs from the $\lambda$ in traditional dirty paper coding and achieves the maximum rate for $R_1$ as in \eqref{eq:FD_Gaussian_Bin_maxR1}. \item The effect of $\rho$ can be seen in Figure \ref{fig:FD_PDFBin_effect_of_rho}. The dashed line represents the resulting rate region using only DPC-binning ($\rho=0$), while the solid line represents the region for PDF-binning when we adapt $\rho \in [-1, 1]$. Figure \ref{fig:FD_PDFBin_effect_of_rho} illustrates that the correlation factor $\rho$ can enlarge the rate region. \end{itemize} \end{rem} \subsection{Signaling and rates for full-duplex Han-Kobayashi PDF-binning}\label{sec:FD_Gaussian_combined} In the Gaussian channel, input signals for the HK-PDF-binning scheme in Section \ref{sec:FD_DMC_Combined} can be represented as \begin{align}\label{eq:FD_Gaussian_Combined_signal} T_{10} &= \alpha S_{10}'(w_{10}'), \nonumber \\ U_{10} &= \alpha S_{10}'(w_{10}') + \beta S_{10}(w_{10}), \nonumber \\ U_{11} &= \gamma S_{11}(w_{11}),\nonumber \\ X_1 &= \alpha S_{10}'(w_{10}') + \beta S_{10}(w_{10}) + \gamma S_{11}(w_{11}) + \delta S_{12}(w_{12}), \nonumber \\ U_{21} &= \theta S_{21}(w_{21}), \nonumber \\ X_2 &= \theta S_{21}(w_{21}) + \mu\left(\rho S_{10}'(w_{10}') + \sqrt{1-\rho^2}S_{22}\right),\nonumber \\ U_{22} &= X_2 + \lambda S_{10}' = (\mu\rho+\lambda)S_{10}' + \theta S_{21}(w_{21}) + \mu\sqrt{1-\rho^2}S_{22}, \end{align} where $S_{10}'$, $S_{10}$, $S_{11}$, $S_{12}$, $S_{21}$, $S_{22}$ are independent $\mathcal{N}(0,1)$ random variables to encode $w_{10}'$, $w_{10}$, $w_{11}$, $w_{12}$, $w_{21}$, $w_{22}$, respectively. $U_{22}$ is the auxiliary random variable for binning that encodes $w_{22}$. $X_1$ and $X_2$ are the transmit signals of $S_1$ and $S_2$. $\rho$ is the correlation coefficient between the transmit signal and the binning state at $S_2$ ($-1\leq \rho \leq 1$). $\lambda$ is the PDF-binning parameter. The parameters $\alpha$, $\beta$, $\gamma$, $\delta$, $\theta$ and $\mu$ are power allocation factors satisfying the power constraints \begin{align}\label{eq:FD_Gaussian_combined_power_constraints} \alpha^2 + \beta^2 + \gamma^2 + \delta^2 &\leq P_1, \nonumber\\ \theta ^2 + \mu^2 &\leq P_2, \end{align} where $P_1$ and $P_2$ are transmit power constraints of $S_1$ and $S_2$. Substitute these variables into the Gaussian channel in \eqref{eq:FD_Gaussian_Y}, we get \begin{align} Y &= c\alpha S_{10}' + c\beta S_{10} + c\gamma S_{11} + c\delta S_{12} + Z, \nonumber \\ Y_1 &= (\alpha + b\mu \rho) S_{10} ' + \beta S_{10} + \gamma S_{11} + \delta S_{12} + b\theta S_{21} + b\mu \sqrt{1-\rho^2}S_{22}+Z_1,\nonumber \\ Y_2 &= (a\alpha + \mu \rho) S_{10} ' + a\beta S_{10} + a\gamma S_{11} + a\delta S_{12} + \theta S_{21} + \mu \sqrt{1-\rho^2}S_{22}+Z_2. \end{align} \begin{cor}\label{cor:FD_Gaussian_combined} The achievable rate region for the full-duplex Gaussian-CRC using the Han-Kobayashi PDF-binning scheme is the convex hull of all rate pairs ($R_1$, $R_2$) satisfying \begin{align}\label{eq:FD_Gaussian_Combined} R_{1} &\leq \min\{I_2 + I_5, I_6\}\nonumber \\ R_{2} &\leq I_{12}- I_1\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_7,I_8 \} + I_{13}- I_1\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_3,I_4\} + I_{14}- I_1\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_9, I_{10} \} + I_{11} - I_1 \nonumber\\ 2R_{1} + R_{2} &\leq \min\{ I_2 + I_3,I_4\} + \min\{I_2 + I_9, I_{10} \} + I_{13}- I_1\nonumber \\ R_{1} + 2R_{2} &\leq \min\{I_2 + I_7, I_8 \}+ I_{11} - I_1 + I_{14}- I_1 \end{align} where \begin{align*} I_2 &= C\left(\frac{c^2\beta^2}{c^2\gamma^2 + c^2\delta^2 + 1} \right)\\ I_3 &= C\left(\frac{\delta^2}{b^2\mu^2 (1-\rho^2)+1} \right)\\ I_4 &= C\left(\frac{\beta^2 + \delta^2 }{b^2\mu^2 (1-\rho^2)+1} \right + C\left(\frac{(\alpha + b\mu \rho)^2 }{\beta^2 + \gamma^2 + \delta^2 + b^2\theta^2 + b^2\mu^2 (1-\rho^2)+ 1} \right)\\ I_5 &= C\left(\frac{\gamma^2 + \delta^2}{b^2\mu^2 (1-\rho^2)+1} \right)\\ I_6 &= C\left(\frac{\beta^2 + \gamma^2 + \delta^2}{b^2\mu^2 (1-\rho^2)+1} \right + C\left(\frac{(\alpha + b\mu \rho)^2 }{\beta^2 + \gamma^2 + \delta^2 + b^2\theta^2 + b^2\mu^2 (1-\rho^2)+ 1} \right)\\ I_7 &= C\left(\frac{\delta^2 + b^2\theta^2}{b^2\mu^2 (1-\rho^2)+1} \right)\\ I_8 &= C\left(\frac{\beta^2 + \delta^2 + b^2\theta^2}{b^2\mu^2 (1-\rho^2)+1} \right + C\left(\frac{(\alpha + b\mu \rho)^2 }{\beta^2 + \gamma^2 + \delta^2 + b^2\theta^2 + b^2\mu^2 (1-\rho^2)+ 1} \right)\\ I_9 &= C\left(\frac{\gamma^2 + \delta^2 + b^2\theta^2}{b^2\mu^2 (1-\rho^2)+1} \right)\\ I_{10} &= C\left(\frac{(\alpha + b\mu \rho)^2 + \beta^2 + \gamma^2 + \delta^2 + b^2\theta^2}{b^2\mu^2 (1-\rho^2)+1} \right)\\ I_{11}-I_{1}&=C \left( \frac {\mu^2(1-\rho^2) } {a^2\beta^2+a^2\delta^2 + 1} \right)\\ I_{12} - I_1 &= C \left( \frac {\mu^2(1-\rho^2) } {a^2\beta^2+a^2\delta^2 + 1} \right) + C\left( \frac{\theta^2 }{(a\alpha + \mu \rho)^2 + a^2\beta^2 + a^2\delta^2 + \mu^2 (1-\rho^2)+1} \right) \end{align*} \begin{align*} I_{13} - I_1 &= C \left( \frac {\mu^2(1-\rho^2) } {a^2\beta^2+a^2\delta^2 + 1} \right + C\left( \frac{a^2\gamma^2}{(a\alpha + \mu \rho)^2 + a^2\beta^2 + a^2\delta^2 + \mu^2 (1-\rho^2)+1} \right)\\ I_{14} - I_1 &= C \left( \frac {\mu^2(1-\rho^2) } {a^2\beta^2+a^2\delta^2 + 1} \right + C\left( \frac{a^2\gamma^2+\theta^2}{(a\alpha + \mu \rho)^2 + a^2\beta^2 + a^2\delta^2 + \mu^2 (1-\rho^2)+1} \right) \end{align*} and $\alpha$, $\beta$, $\gamma$, $\delta$, $\theta$ and $\mu$ are power allocation factors satisfying the power constraints \eqref{eq:FD_Gaussian_combined_power_constraints} and $-1\leq \rho \leq 1$. \end{cor} \begin{proof} Applying Theorem \ref{thm:FD_DMC_combined} with the signaling in \eqref{eq:FD_Gaussian_Combined_signal}, we obtain the rate region in Corollary \ref{cor:FD_Gaussian_combined}. \end{proof} Note that rate region \eqref{eq:FD_Gaussian_Combined} includes both the Han-Kobayashi rate region and the PDF-binning region in \eqref{eq:FD_Gaussian_Bin}. Furthermore, the maximum rates for user 1 and user 2 are the same as in \eqref{eq:FD_Gaussian_Bin_maxR1} and \eqref{eq:FD_Gaussian_Bin_maxR2}. \subsection{Optimal binning parameter for full-duplex Han-Kobayashi PDF-binning} \begin{cor}\label{cor:FD_Gaussian_optimalLambda} The optimal $\lambda^*$ for the full-duplex HK-PDF-binning scheme is \begin{align}\label{eq:lambda_FD_Gaussian_Combined} \lambda^* &= \frac{a\alpha\mu^2(1-\rho^2)-\mu\rho(a^2\beta^2+a^2\delta^2+1)}{a^2\beta^2+a^2\delta^2+\mu^2(1-\rho^2)+1} \end{align} \end{cor} \begin{proof} $\lambda^*$ is obtained by maximizing the term $I_{11}-I_{1}$ in \eqref{eq:FD_DMC_combined}. See Appendix \ref{proof:FD_Gaussian_Combined_lambda} for details. \end{proof} Note that the optimal $\lambda^*$ in \eqref{eq:lambda_FD_Gaussian_Combined} contains both the optimal $\lambda^*$ for PDF-binning in \eqref{eq:lambda_FD_Gaussian_Bin} and the optimal $\lambda$ for DPC binning \cite{Costa83TIT} as special cases. \subsection{Numerical examples}\label{sec:FD_Outer_compare} In this section, we provide numerical comparison among the proposed PDF-binning and HK-PDF-binning schemes, the original Han-Kobayashi scheme, and an outer bound as discussed below. \subsubsection{Outer bounds for the CRC capacity} We obtain a simple outer bound for the CRC capacity by combining the capacity for the (non-causal) CIC and the ourter bound for interference channel with user cooperation (IC-UC) \cite{TandonTIT2011}. Where the CIC capacity result is not available, we use the MISO broadcast capacity. \begin{align*} \text{CRC capacity} &\subset \text{CIC capacity} \bigcap \text{IC-UC outer bound} \\ &\subset \text{MISO BC capacity} \bigcap \text{IC-UC outer bound}. \end{align*} \emph{a)} {Capacity of the CIC as an outer bound:} The capacity of the ideal CIC (with non-causal knowledge of $S_1$'s message at $S_2$) is an outer bound to the CRC rate region. The CIC capacity is known in the cases of (i) weak interference \cite{Wu2007TIT,Jovicic2009TIT} (ii) very strong interference \cite{Maric2007TIT} (iii) the primary-decode-cognitive region \cite{Rini2011TIT_Gaussian}. For strong interference, we can also use the outer bound to the CIC capacity in \cite{MaricEuroTele2008} as an outer bound to the CRC. \emph{b)} {IC-UC outer bound:} Tandon and Ulukus \cite{TandonTIT2011} obtain an outer bound for the MAC with generalized feedback based on dependence balance, which is first proposed by Hekstra and Willems \cite{HekstraTIT1989} to study outer bounds for the single-output two-way channels. The basic idea of dependence balance is that no more information can be consumed than produced. Tandon and Ulukus apply this idea to obtain a new outer bound for IC-UC. It is shown that this dependence-balance-based outer bound is strictly tighter than the cutset bound (see Section V of \cite{TandonTIT2011}). Thus, this bound can be used instead of the relay channel (RC) cutset bound for $R_1$. \emph{c)} {Gaussian Vector Broadcast Outer Bound:} Consider a $2 \times 1$ MISO broadcast system as \begin{align} Y_1 = [1 \quad b] X + Z_1, \nonumber\\ Y_2 = [a \quad 1] X + Z_2, \end{align} where $a$, $b$ are the channel gains, $Z_1$ and $Z_2$ are white Gaussian noises with identity covariance. The vector codeword $X$ consists of two independent parts: \begin{align*} X = U+V, \end{align*} where $X=\left( \begin{array}{c} X_1 \\ X_2 \\ \end{array} \right) $, $U=\left( \begin{array}{c} U_1 \\ V_1 \\ \end{array} \right) $, $V=\left( \begin{array}{c} U_2 \\ V_2 \\ \end{array} \right) $, and $U_1$, $V_1$, $U_2$, $V_2$ are zero-mean Gaussian codewords with covariances: \begin{align*} K_U = \begin{bmatrix} \alpha^2 & \rho_1 \alpha \beta \\ \rho_1 \alpha \beta & \beta^2 \end{bmatrix}; \quad K_V = \begin{bmatrix} \gamma^2 & \rho_2 \gamma \delta \\ \rho_2 \gamma \delta & \delta^2 \end{bmatrix}, \end{align*} in which the power allocation factors satisfy \begin{align} \alpha^2 + \beta^2 \leq P_1, \quad \gamma^2 + \delta^2 \leq P_2, \end{align} and the input correlation factors $\rho_1, \rho_2\in [-1,1]$. The Gaussian vector broadcast capacity region is the convex closure of $R_{o1} \bigcup R_{o2}$ \cite{Gamal10notes}, where $R_{o1}$ is the region \begin{align R_1 &\leq C\left( \frac{\alpha^2 + 2b \rho_1 \alpha \beta + b^2 \beta^2 } {\gamma^2 + 2b \rho_2 \gamma \delta + b^2 \delta^2+1} \right)\nonumber\\ R_2 &\leq C\left(a^2\gamma^2 + 2a \rho_2 \gamma \delta + \delta^2 \right) \end{align} and $R_{o2}$ is the region \begin{align R_1 &\leq C\left(\alpha^2 + 2b \rho_1 \alpha \beta + b^2 \beta^2 \right) \nonumber\\ R_2 &\leq C\left( \frac{a^2 \gamma^2 + 2a \rho_2 \gamma \delta + \delta^2} {a^2 \alpha^2 + 2a \rho_1 \alpha \beta + \beta^2+1} \right). \end{align} \subsubsection{Numerical comparison} \begin{figure}[t] \centering \includegraphics[scale=0.45]{FD_PDFBin_HKPDFBin_comparison.pdf} \caption{Rate regions for full-duplex schemes in the Gaussian cognitive relay channel. } \label{fig:FD_PDFBin_HKPDFBin_comparison} \end{figure} Figure \ref{fig:FD_PDFBin_HKPDFBin_comparison} shows the comparison in the full-duplex mode among the Han-Kobayashi scheme, PDF-binning, HK-PDF-binning, and the outer bound. We can see that the proposed HD-PDF-binning scheme contains both the Han-Kobayashi and the PDF-binning rate regions, as analyzed in Remark \ref{rem:FD_DMC_Combined_Inclusions}. Note that the outer bound is the intersection of the two bounds drawn and is loose as this bound is not achievable. However, we observe that as $b$ decreases, the HK-PDF-binning rate region becomes closer to the outer bound. \section{Half-Duplex Coding schemes} In this section, we adapt the two full-duplex schemes to the half-duplex mode. The half-duplex schemes are also based on rate splitting, superposition encoding, partial decode-forward binning and Han-Kobayashi coding. There are several differences between the half- and full-duplex cases. First, under the half-duplex constraint, no node can both transmit and receive at the same time, thus leading us to divide each transmission block into two phases. In the first phase, $S_1$ sends a message to $S_2$, $T_1$ and $T_2$, while $S_2$ only receives but sends no messages. In the second phase, both $S_1$ and $S_2$ send messages concurrently. Second, $S_1$ sends different message parts in different phases. Specifically, $S_1$ only sends one part of its message to other nodes in the first phase, but will send all message parts in the second phase. Third, there is no block Markovity in the encoding since the superposition coding can be done between 2 phases of the same block instead of between 2 consecutive blocks. Finally, both $T_1$ and $T_2$ apply joint decoding only at the end of the second phase to make use of the received signals in both phases. \subsection{Half-duplex partial decode-forward binning scheme} \begin{figure*}[t] \centering \includegraphics[scale=0.7]{HD_DM_CRC_PDF_Bin.pdf} \caption{Coding structure for the half-duplex CRC based on partial decode-forward binning.} \label{fig:HD_DM_CRC_PDF_Bin} \end{figure*} The coding structure for the half-duplex PDF-binning scheme is shown in Figure \ref{fig:HD_DM_CRC_PDF_Bin}. This scheme uses superposition encoding at the first sender, and partial decode-forward relaying and binning at the second sender. The first sender $S_1$ splits its message into two parts $(w_{10}, w_{11})$, corresponding to the forwarding and private parts. In the first phase, $S_1$ sends a codeword $X_{11}^{\tau n}$ containing the message part $w_{10}$; $S_2$ sends no information but only listens. At the end of the first phase, $S_2$ decodes $w_{10}$ from $S_1$ . Note that neither $T_1$ nor $T_2$ decodes during this phase. In the second phase, $S_1$ sends a codeword $X_{12}^{\tau n}$ containing both parts $(w_{10}, w_{11})$, in which $w_{11}$ is superimposed on $w_{10}$. $S_2$ now sends both $w_2$ and $w_{10}$ and uses Gelfand-Pinsker binning technique to bin against the codeword $X_{11}^{n}(w_{10})$ decoded from $S_1$ in the first phase. At the destinations, $T_1$ uses joint decoding to decode $(w_{10},w_{11})$ from the signals received in both phases; $T_2$ decodes $w_{2}$ using the received signal in the second phase. Specifically, at the end of the first phase, $S_2$ searches for a unique $\hat{w}_{10}$ such that \begin{align*} (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y}) \in A_\epsilon^{(\tau n)}(P_{X_{11} Y}), \end{align*} where $\mathbf{y}$ is the received signal vector at $S_2$ in the first phase. It then performs binning by looking for a $v_2$ such that \begin{align*} (x_{11}^{ \bar{\tau} n}{(\hat{w}_{10})}, u_{2}^{ \bar{\tau} n}{(w_{2},v_{2})}) \in A_\epsilon^{(\bar{\tau} n)}(P_{X_{1} U_{2}}), \end{align*} and sends $x_{22}^{ \bar{\tau} n}{(\mathbf{x}_{11}, \mathbf{u}_{2})}$ as a function of $x_{11}^{ \bar{\tau} n}$ and $u_{2}^{ \bar{\tau} n}$ in the second phase. At the end of the second phase, $T_1$ searches for a unique ($\hat{w}_{10},\hat{w}_{11}$) such that \begin{align*} ( x_{11}^{\bar{\tau}n}{(\hat{w}_{10})}, x_{12}^{\bar{\tau}n}(\hat{w}_{11}|\hat{w}_{10}), \mathbf{y_{12}})&\in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} X_{12} Y_{12}})&\\ \text{and} \quad (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y_{11}}) &\in A_\epsilon^{(\tau n)}(P_{X_{11} Y_{11}}),& \end{align*} where $\mathbf{y_{11}}$ and $\mathbf{y_{12}}$ indicate the received vectors at $T_1$ during the first and second phases, respectively. $T_2$ treats the codeword $X_{11}^n$ as the state and decodes $w_{2}$. It searches for a unique $\hat{w}_2$ for some $\hat{v}_{2} $ such that \begin{align*} (u_2^{\bar{\tau}n}(\hat{w}_{2} ,\hat{v}_{2}), \mathbf{y_{22}}) \in A_\epsilon^{(\bar{\tau}n)}(P_{U_2 Y_{22}}), \end{align*} where $\mathbf{y_{22}}$ is the received vector at $T_2$ in the second phase. \begin{thm} \label{thm:HD_DMC_Bin} The convex hull of the following rate region is achievable for the half-duplex cognitive relay channel using PDF-binning: \begin{align}\label{eq:DMC_PDF_Bin} \bigcup_{\substack{P_{3}}} \left\{\begin{array}{ll} R_{1} &\leq \tau I(X_{11};Y) + \bar{\tau} I(X_{12};Y_{12}|X_{11})\\ R_{1} &\leq \tau I(X_{11}; Y_{11}) + \bar{\tau} I(X_{11}, X_{12};Y_{12})\\ R_2 &\leq \bar{\tau} I(U_2;Y_{22})- \bar{\tau} I( U_2; X_{11}) \end{array}\right. \end{align} where \begin{align*} P_{3} = &p(x_{11})p(x_{12}|x_{11})p(u_2|x_{11})p(x_{22}|x_{11},u_2 p_c(y_{11},y_{21},y,y_{12},y_{22}|x_{11},x_{12},x_{22}), \end{align*} and $p_c$ is given in \eqref{eq:p_c}, $\bar{\tau} = 1 - \tau, 0\leq \tau \leq 1$. \end{thm} \begin{proof} See Appendix \ref{proof:HD_DMC_Bin} for the detailed proof. \end{proof} \begin{rem}The maximum rate for each user. \begin{itemize} \item The first user $S_1$ achieves the maximum rate of half-duplex partial decode-forward relaying if we set $U_2=\emptyset$. \begin{align}\label{eq:HD_DM_Bin_maxR1} R_{1}^{\max} = \max_{\substack{0\leq \tau \leq 1 \\p(x_{11},x_{12})}}\min\{&\tau I(X_{11};Y) + \bar{\tau} I(X_{12};Y_{12}|X_{11}), \tau I(X_{11}; Y_{11}) + \bar{\tau} I(X_{11}, X_{12};Y_{12})\}. \end{align} This half-duplex $R_{1}^{\max}$ is slightly smaller than in the full-duplex case of \eqref{eq:FD_DMC_Bin_maxR1}. \item The second user $S_2$ achieves the maximum rate of Gelfand-Pinsker's binning if we set $\tau=0$, $X_{12}=X_{11}$. \begin{align}\label{eq:HD_DM_Bin_maxR2} R_{2}^{\max} = \max_{\substack{p(x_{11},u_2)p(x_{22}|x_{11},u_2)}} \{I(U_2;Y_{22})- I( U_2; X_{11})\}. \end{align} This half-duplex $R_{2}^{\max}$ is the same as in the full-duplex case of \eqref{eq:FD_DMC_Bin_maxR2}. Even though this equality seems somewhat surprising, it is indeed the case in the limit of $\tau \rightarrow 0$, given that user 1 sends just enough information for $S_2$ to be able to decode completely in the first phase and then bin against it in the second phase. At $\tau = 0$ and $X_{12} = X_{11} = \emptyset$, $S_2$ can achieve the interference-free rate. \end{itemize} \end{rem} \subsection{Half-duplex Han-Kobayashi PDF-binning scheme}\label{sec:HD_DMC_Combined} \begin{figure*}[t] \centering \includegraphics[scale=0.7]{HD_DM_CRC_Combined.pdf} \caption{Coding structure for the half-duplex CRC based on Han-Kobayashi partial decode-forward binning.} \label{fig:HD_DM_CRC_Combined} \end{figure*} The first half-duplex coding scheme utilizes PDF-binning at the second sender and achieves the maximum possible rates for both user $1$ and user $2$. But it does not include the Han-Kobayashi scheme for the interference channel. In this section, we extend this scheme to combine with the Han-Kobayashi scheme by further splitting the messages in the second phase. The coding structure for half-duplex HK-PDF-binning is shown in Figure \ref{fig:HD_DM_CRC_Combined}. The encoding and decoding procedure in the first phase is the same as that of half-duplex PDF-binning. The major difference is in the second phase. Message $w_1$ of the first sender $S_1$ is split into three parts ($w_{10},w_{11},w_{12}$), corresponding to the forwarding, public and private parts. Message $w_2$ is split into 2 parts ($w_{21}, w_{22}$), corresponding to the public and private parts. We generate independent codewords for messages $w_{10}$ and $w_{11}$ and superimpose $w_{12}$ on both of them. In the first phase, $S_1$ sends a codeword containing $w_{10}$, while $S_2$ does not send any message. At the end of the first phase, $S_2$ decode $\tilde{w}_{10}$ using the received signal vector $\mathbf{y}$ and then bins its private part $w_{22}$ against the decoded message $\tilde{w}_{10}$, conditionally on knowing the public part $w_{21}$. In the second phase, $S_1$ sends a codeword containing ($w_{10}, w_{11}, w_{12}$) while $S_2$ sends the binned signal containing ($w_{10}, w_{21}, w_{22}$). At the end of the second phase, $T_1$ uses joint decoding across both phases and searches for a unique triple ($\hat{w}_{10},\hat{w}_{11},\hat{w}_{12}$) for some $w_{21}$. $T_2$ also uses joint decoding based on the received signal in the second phase and searches for a unique pair $(\hat{w}_{21},\hat{w}_{22})$ for some $\hat{w}_{11}$. Specifically, in the first phase, $S_1$ sends $x_{11}^{ \tau n}(w_{10})$; $S_2$ does not transmit. In the second phase, $S_1$ sends $x_{12}^{\bar{\tau}n}(w_{12}|w_{10},w_{11})$; $S_2$ searches for some $v_{22}$ such that \begin{align}\label{eq:HD_DMC_Bin_binStep} (x_{11}^{\bar{\tau}n}{(w_{10})},&u_{21}^{\bar{\tau}n}( w_{21}), u_{22}^{\bar{\tau}n}( w_{22}, v_{22}|w_{21}) \in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} U_{22}|U_{21}}), \end{align} and then sends $x_{2}^{ \bar{\tau} n}(w_{10},w_{21},w_{22},v_{22})$. For decoding, at the end of the first phase, $S_2$ searches for a unique $\hat{w}_{10}$ such that \begin{align} (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y}) \in A_\epsilon^{(\tau n)}(P_{X_{11} Y}). \end{align} At the end of the second phase, $T_1$ searches for a unique ($\hat{w}_{10},\hat{w}_{11},\hat{w}_{12}$) for some $\hat{w}_{21}$ such that \begin{align}\label{eq:HD_DMC_Bin_jointDec1} ( x_{11}^{\bar{\tau}n}{(\hat{w}_{10})},u_{11}^{\bar{\tau}n}{(\hat{w}_{11})}, x_{12}^{\bar{\tau}n}(\hat{w}_{12}|\hat{w}_{10},\hat{w}_{11}) u_{21}^{\bar{\tau}n}{(\hat{w}_{21})}, \mathbf{y_{12}})&\in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} U_{11} X_{12} U_{21} Y_{12}})\nonumber\\ \text{and} \quad x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y_{11}}) &\in A_\epsilon^{(\tau n)}(P_{ X_{11} Y_{11}}). \end{align} $T_2$ searches for a unique $(\hat{w}_{21},\hat{w}_{22})$ for some $(\hat{w}_{11},\hat{v}_{22})$ such that \begin{align}\label{eq:HD_DMC_Bin_jointDec2} (u_{11}^{\bar{\tau}n}(\hat{w}_{11} ),&u_{21}^{\bar{\tau}n}(\hat{w}_{21} ), u_{22}^{\bar{\tau}n}(\hat{w}_{22},\hat{v}_{22}|\hat{w}_{21} ),\mathbf{y_{22}}) \in A_\epsilon^{(\bar{\tau}n)}(P_{U_{11} U_{21} U_{22} Y_{22}}). \end{align} Note that similar to the full-duplex scheme in Section \ref{sec:FD_DMC_Combined}, we use conditional binning in step \eqref{eq:HD_DMC_Bin_binStep}, and joint decoding at both destinations in steps \eqref{eq:HD_DMC_Bin_jointDec1} and \eqref{eq:HD_DMC_Bin_jointDec2} (see Remark \ref{rem:FD_DMC_Combined_features}). \begin{thm} \label{thm:HD_DMC_combined} The convex hull of the following rate region is achievable for the half-duplex cognitive relay channel using the HK-PDF-binning scheme: \begin{align}\label{eq:HD_DMC_combined} \bigcup_{\substack{P_{4}}} \left\{\begin{array}{ll} R_{1} &\leq \min\{I_2 + I_5, I_6\}\\ R_{2} &\leq I_{12}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_7,I_8 \} + I_{13}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_3,I_4\} + I_{14}- I_1 \\ R_{1} + R_{2} &\leq \min\{I_2 + I_9, I_{10} \} + I_{11} - I_1 \\ 2R_{1} + R_{2} &\leq \min\{ I_2 + I_3,I_4\} + \min\{I_2 + I_9, I_{10} \} + I_{13}- I_1\\ R_{1} + 2R_{2} &\leq \min\{I_2 + I_7, I_8 \} + I_{11} - I_ + I_{14}- I_1 \end{array}\right. \end{align} where \begin{align} P_{4} = &p(x_{11})p(u_{11})p(x_{12}|u_{11},x_{11})p(u_{21})p(u_{22}|u_{21},x_{11})\nonumber\\ &p(x_{22}|x_{11},u_{21},u_{22} p_c(y_{11},y_{21},y,y_{12},y_{22}|x_{11},x_{12},x_{22}), \end{align} with $p_c$ as given in \eqref{eq:p_c} and \begin{align} I_1 &= \bar{\tau} I( U_{22}; X_{11}|U_{21}) \nonumber \\ I_2 &= \tau I(X_{11};Y) \nonumber \\ I_3 &= \bar{\tau} I(U_{12};Y_{12}|X_{11},U_{11},U_{21}) \nonumber \\ I_4 &= \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},X_{12};Y_{12}|U_{11},U_{21}) \nonumber \\ I_5 &= \bar{\tau} I(U_{11},X_{12};Y_{12}|X_{11},U_{21}) \nonumber \\ I_6 &= \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},U_{11},X_{12};Y_{12}|U_{21}) \nonumber \\ I_7 &= \bar{\tau} I(X_{12},U_{21};Y_{12}|X_{11},U_{11})\nonumber\\ I_8 &= \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},X_{12},U_{21};Y_{12}|U_{11}) \nonumber \end{align} \begin{align} I_9 &= \bar{\tau} I(U_{11},X_{12},U_{21};Y_{12}|X_{11}) \nonumber\\ I_{10} &= \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},U_{11},X_{12},U_{21};Y_{12}) \nonumber\\ I_{11} &= \bar{\tau} I(U_{22};Y_{22}|U_{21},U_{11}) \nonumber\\ I_{12} &= \bar{\tau} I(U_{21},U_{22};Y_{22}|U_{11}) \nonumber\\ I_{13} &= \bar{\tau} I(U_{11},U_{22};Y_{22}|U_{21}) \nonumber\\ I_{14} &= \bar{\tau} I(U_{11},U_{21},U_{22};Y_{22}) \end{align} \end{thm} where $\bar{\tau} = 1 - \tau, 0\leq \tau \leq 1$. \begin{proof} See Appendix \ref{proof:HD_DMC_Combined} for the details. \end{proof} \begin{rem}Inclusion of half-duplex PDF-binning and Han-Kobayashi schemes. \begin{itemize} \item The half-duplex HK-PDF-binning scheme becomes half-duplex PDF-binning if $U_{11}=U_{21}=\emptyset$. \item The half-duplex HK-PDF-binning scheme becomes the Han-Kobayashi scheme if $\tau=0$, $X_{11}=\emptyset$ and $X_{22} = U_{22}$. \item The maximum rates for $S_1$ and $S_2$ are the same as in \eqref{eq:HD_DM_Bin_maxR1} and \eqref{eq:HD_DM_Bin_maxR2}. \end{itemize} \end{rem} \begin{figure*}[t] \centering \includegraphics[scale=0.60]{HD_Gaussian_CRC_model.pdf} \caption{The half-duplex Gaussian cognitive relay channel model.} \label{fig:model_HD_CR} \end{figure*} \section{Half-duplex Gaussian CRC rate regions} \subsection{ Half-duplex Gaussian CRC model} The Gaussian model for the half-duplex cognitive relay channel is shown in Figure \ref{fig:model_HD_CR}. The input-output signals can be represented as \begin{align} \text{First phase :}\quad \;\, Y &= c X_{11} + Z, \nonumber\\ Y_{11} &= X_{11} + Z_{11}, \nonumber\\ Y_{21} &= a X_{11} + Z_{21};\label{eq:channel_first_phase}\\ \text{Second phase :} \quad Y_{12} &= X_{12} + b X_{22} + Z_{12}, \nonumber\\ Y_{22} &= a X_{12} + X_{22} + Z_{22},\label{eq:channel_second_phase} \end{align} where $X_{11}$ is the transmit signal of $S_1$ in the first phase, $X_{12}$ and $X_{22}$ are the transmit signals of $S_1$ and $S_2$ in the second phase, respectively. $Y$, $Y_{11}$ and $Y_{21}$ are the received signals at $S_2$, $T_1$ and $T_2$ in the first phase. $Y_{21}$ and $Y_{22}$ are the received signals at $T_1$ and $T_2$ in the second phase. $a$, $b$, and $c$ are the channel gains where the direct links are normalized to $1$ as in the standard interference channel\cite{Carleial78}. $Z$, $Z_{11}$, $Z_{21}$, $Z_{12}$, and $Z_{22}$ are independent white Gaussian noises with unit variance. In the following section, we only provide the analysis for the half-duplex Gaussian HK-PDF-binning scheme and omit the analysis for half-duplex PDF-binning, which is a special case of HK-PDF-binning. \subsection{Signaling and rates for the half-duplex HK-PDF-binning}\label{sec:HD_Gaussian_Combined} In a Gaussian channel, input signals for the HK-PDF-binning scheme as in Section \ref{sec:HD_DMC_Combined} can be represented as \begin{align}\label{eq:HD_Gaussian_Combined_signaling} X_{11} &= \alpha_1 S_{10}(w_{10}), \\ X_{12} &= \alpha_2 S_{10}(w_{10}) + \beta_2 S_{11}(w_{11}) + \gamma_2 S_{12}(w_{12}), \nonumber \\ X_{22} &= \theta S_{21}(w_{21}) + \mu\left(\rho S_{10}(w_{10}) +\sqrt{1-\rho^2}S_{22}\right),\nonumber \\ U_{22} &= X_{22} + \lambda S_{10} =(\mu\rho+\lambda)S_{10}+\theta S_{21} +\mu\sqrt{1-\rho^2}S_{22},\nonumber \end{align} where $S_{10}$, $S_{11}$, $S_{12}$, $S_{21}$ and $S_{22}$ are independent $\mathcal{N}(0,1)$ random variables that encode $w_{10}$, $w_{11}$, $w_{12}$, $w_{21}$ and $w_{22}$ respectively, $U_{22}$ is the Gelfand-Pinsker binning variable that encodes $w_{22}$. The parameter $\rho$ is the correlation factor between the transmit signal $X_{22}$ and the state $X_{11}$, similar to that in Section \ref{sec:signal_FD_PDF_Binning}. $\lambda$ is a parameter for binning. Parameters $\alpha_1$, $\alpha_2$, $\beta_2$, $\gamma_2$, $\theta$ and $\mu$ are the corresponding power allocations that satisfy the power constraints \begin{align}\label{eq:combined_power_cons} \tau \alpha_1^2 + \bar{\tau} (\alpha_2^2 + \beta_2^2 + \gamma_2^2) &\leq P_1, \nonumber\\ \bar{\tau} (\mu ^2 + \theta^2) &\leq P_2, \end{align} where $\tau$ and $\bar{\tau}=1-\tau$ are the time duration for the two phases. Substitute $X_{11}$, $X_{12}$, $X_{22}$ into $Y$, $Y_{11}$, $Y_{21}$, $Y_{12}$, $Y_{22}$ in \eqref{eq:channel_first_phase} and \eqref{eq:channel_second_phase}, we get \begin{align} Y &= c \alpha_1 S_{10} + Z, \nonumber \\ Y_{11} &= \alpha_1 S_{10} + Z_{11},\nonumber \\ Y_{21} &= a\alpha_1 S_{10} + Z_{21}, \nonumber \\ Y_{12} &= (\alpha_2+b\mu\rho) S_{10} + \beta_2 S_{11} + \gamma_2 S_{12} + b\theta S_{21} + b\mu\sqrt{1-\rho^2}S_{22} + Z_{12},\nonumber \\ Y_{22} &= (a\alpha_2+\mu\rho) S_{10} + a\beta_2 S_{11} + a\gamma_2 S_{12}+ \theta S_{21} + \mu\sqrt{1-\rho^2}S_{22} + Z_{22}. \end{align} \begin{cor}\label{cor:HD_Gaussian_Combined} The achievable rate region for the half-duplex cognitive relay channel using Han-Kobayashi PDF-binning is the convex hull of all rate pairs ($R_1$, $R_2$) satisfying \begin{align}\label{eq:HD_Gaussian_Combined} R_{1} &\leq \min\{I_2 + I_5, I_6\}, \nonumber\\ R_{2} &\leq I_{12}- I_1,\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_7,I_8 \} + I_{13}- I_1,\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_3,I_4\} + I_{14}- I_1,\nonumber \\ R_{1} + R_{2} &\leq \min\{I_2 + I_9, I_{10} \} + I_{11} - I_1, \nonumber\\ 2R_{1} + R_{2} &\leq \min\{ I_2 + I_3,I_4\} + \min\{I_2 + I_9, I_{10} \} + I_{13}- I_1,\nonumber \\ R_{1} + 2R_{2} &\leq \min\{I_2 + I_7, I_8 \}+ I_{11} - I_1 + I_{14}- I_1, \end{align} where \begin{align*} I_2 &= \tau C\left(c^2 \alpha_1^2 \right),\\ I_3 &= \bar{\tau} C\left( \frac{\gamma_2^2}{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_4 &= \tau C\left(\alpha_1^2 \right) + \bar{\tau} C\left( \frac{(\alpha_2+b\mu\rho)^2 + \gamma_2^2 }{b^2\mu^2(1-\rho^2) + 1}\right), \\ I_5 &= \bar{\tau} C\left( \frac{\beta_2^2 + \gamma_2^2}{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_6 &= \tau C\left(\alpha_1^2 \right) + \bar{\tau} C\left( \frac{(\alpha_2+b\mu\rho)^2 + \beta_2^2 + \gamma_2^2 }{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_7 &= \bar{\tau} C\left( \frac{\gamma_2^2 + b^2\theta^2}{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_8 &= \tau C\left(\alpha_1^2 \right) + \bar{\tau} C\left( \frac{(\alpha_2+b\mu\rho)^2 + \gamma_2^2 + b^2\theta^2 }{b^2\mu^2(1-\rho^2) + 1}\right),\\ I_9 &= \bar{\tau} C\left( \frac{\beta_2 ^2 + \gamma_2 ^2 + b^2\theta^2}{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_{10} &= \tau C\left(\alpha_1^2 \right) + \bar{\tau} C\left( \frac{(\alpha_2+b\mu\rho) ^2 + \beta_2 ^2 + \gamma_2 ^2 + b^2\theta ^2 }{b^2\mu^2(1-\rho^2) + 1} \right),\\ I_{11}-I_{1} &=\bar{\tau} C \left( \frac {\mu^2(1-\rho^2) } {a^2\gamma_2^2 + 1} \right),\\ I_{12} - I_1 &= \bar{\tau} C \left( \frac {\mu^2(1-\rho^2) } {a^2\gamma_2^2 + 1} \right) + \bar{\tau} C\left( \frac{\theta^2}{(a\alpha_2+\mu\rho)^2 + a^2\gamma_2 ^2 + \mu^2(1-\rho^2) + 1} \right), \end{align*} \begin{align*} I_{13} - I_1 &= \bar{\tau} C \left( \frac {\mu^2(1-\rho^2) } {a^2\gamma_2^2 + 1} \right) + \bar{\tau} C\left( \frac{ a^2\beta_2 ^2}{(a\alpha_2+\mu\rho)^2 + a^2\gamma_2 ^2 + \mu^2(1-\rho^2) + 1} \right),\\ I_{14} - I_1 &= \bar{\tau} C \left( \frac {\mu^2(1-\rho^2) } {a^2\gamma_2^2 + 1} \right) + \bar{\tau} C\left( \frac{ a^2\beta_2^2 + \theta^2 }{(a\alpha_2+\mu\rho)^2 + a^2\gamma_2 ^2 + \mu^2(1-\rho^2) + 1} \right), \end{align*} and $C(x)=0.5\log_2(1+x)$; $\tau \in{[0,1]}$ and $\tau+\bar{\tau} = 1$; $\rho \in{[-1,1]}$ is the correlation factor between $S_2$'s transmit signal $X_{22}$ and the state $X_{11}$; and the power allocations $\alpha_1$, $\alpha_2$, $\beta_2$, $\gamma_2$, $\theta$ and $\mu$ satisfy the power constraints \eqref{eq:combined_power_cons}. \end{cor} \begin{proof} Applying the signaling in \eqref{eq:HD_Gaussian_Combined_signaling} to Theorem \ref{thm:HD_DMC_combined}, we obtain the rate region in Corollary \ref{cor:HD_Gaussian_Combined}. \end{proof} \begin{rem}{Inclusion of half-duplex PDF-binning and Han-Kobayashi schemes.} \begin{itemize} \item If we set $\tau=0$, $\alpha_1=\alpha_2=0$, $\rho=0$, rate region \eqref{eq:HD_Gaussian_Combined} becomes the Han-Kobayashi region \cite{Han}. \item If we set $\beta_2=\theta=0$, rate region \eqref{eq:HD_Gaussian_Combined} becomes the half-duplex PDF-binning region. \item The half-duplex PDF-binning region is the convex hull of all rate pairs ($R_1$, $R_2$) satisfying \begin{align} R_{1} &\leq \tau C \left( { c^2 \alpha_1^2 } \right) + \bar{\tau} C \left( \frac { \gamma_2^2 } {b^2 \mu^2(1-\rho^2)+ 1} \right) , \nonumber \\ R_{1} &\leq \tau C \left( { \alpha_1^2 } \right) + \bar{\tau} C \left( \frac {(\alpha_2+b\mu\rho)^2 +\gamma_2^2 } {b^2 \mu^2(1-\rho^2)+ 1} \right),\nonumber\\ R_2 &\leq \bar{\tau} C \left( \frac {\mu^2(1-\rho^2) } {a^2 \gamma_2^2 + 1} \right), \end{align} where the power allocations $\alpha_1$, $\alpha_2$, $\gamma_2$ and $\mu$ satisfy the power constraints \begin{align}\label{eq:bin_power_cons} \tau \alpha_1^2 + \bar{\tau} (\alpha_2^2 + \gamma_2^2) &\leq P_1, \nonumber\\ \bar{\tau} \mu ^2 &\leq P_2. \end{align} \item The maximum rate for $S_1$ is achieved by setting $\beta_2=\theta=0$, $\rho = \pm 1$ and $\mu =\rho \sqrt{P_2}$ as \begin{align}\label{eq:HD_Gaussian_Bin_maxR1} R_{1}^{\max} = &\max_{\substack{\tau \alpha_1^2 + \bar{\tau} (\alpha_2^2 + \gamma_2^2) \leq P_1}} \min \Bigg \{\tau C ({c^2 \alpha_1^2}) + \bar{\tau} C(\gamma_2^2), \tau C(\alpha_1^2) + \bar{\tau} C\left( \left(\alpha_2+b\sqrt{P_2}\right)^2 + \gamma_2^2 \right) \Bigg \}. \end{align} A solution for this optimization problem is available in \cite{HostTIT2005}. Note that in the half-duplex mode, partial decode-forward achieves a strictly higher rate than pure decode-forward for the Gaussian channel. \item The maximum rate for $S_2$ is achieved by setting $\tau=0$, $\rho=0$, $\alpha_1=\alpha_2=\beta_2=\gamma_2=\theta=0$, and $\mu=\sqrt{P_2}$ as \begin{align}\label{eq:HD_Gaussian_Bin_maxR2} R_{2}^{\max} = C(P_2). \end{align} \end{itemize} \end{rem} \begin{rem}{The optimal binning parameter can be found similarly to the full-duplex case as follows.} \begin{cor} The optimal parameter $\lambda$ for the half-duplex Han-Kobayashi partial decode-forward binning scheme is \begin{align} \lambda^* &= \frac{a\alpha_2\mu^2(1-\rho^2)-\mu\rho(a^2 \gamma_2^2+1)}{a^2 \gamma_2^2+\mu^2(1-\rho^2)+1}. \end{align} \end{cor} \begin{proof} Similar approach to the proof of Corollary \ref{cor:FD_Gaussian_optimalLambda}. \end{proof} \end{rem} \subsection{Performance Comparison}\label{sec:numerical_sections} \subsubsection{Existing results} Very few results currently exist for the CRC. We can only find two results for the half-duplex mode. Next we briefly discuss each of these results. Devroye, Mitran and Tarokh \cite{Devroye2006TIT} propose four half-duplex protocols with rate region as the convex hull of the four regions. One protocol is the Han-Kobayashi scheme for the interference channel, and the other three are 2-phase protocols in which $S_2$ obtains $S_1$'s message causally in the first phase as in a broadcast channel, then transmits cognitively in the second phase. All these 3 protocols have $T_1$ decode at the end of both phases instead of only at the end of the second phase, hence they are suboptimal. Protocol 2 has the idea of decode-forward by keeping the same input distribution at $S_1$ in both phases, but because in the second phase, it reduces rate at $S_1$, thus it does not achieve the rate of decode-forward relaying. Thus, even though the rate region includes the Han-Kobayashi region (in protocol 3), it does not include partial decode-forward relaying. Chatterjee, Tong and Oyman \cite{ChatterjeeISIT2010} propose an achievable rate region for the half-duplex CRC based on rate-splitting, block Markov encoding, Gelfand-Pinsker binning and backward decoding. The transmission is performed in $B$ blocks, each is divided into two phases. In each phase, each user splits its message into two parts, one common and one private. The primary user ($S_1$) superimposes its messages in both phases of the current block on the messages in the first phase of the previous block. The cognitive user ($S_2$) only transmits in the second phase and bins both its message parts against the private message of $S_1$ in the first phase of the previous block. Backward decoding is then used to decode the messages after $B$ blocks. We have several comments on this scheme: \begin{itemize} \item Block Markovity is not necessary in half-duplex mode. We can superimpose the second-phase signal on the first-phase signal of the same block, instead of superimposing both phase signals on the first-phase signal of the previous block and using backward decoding as in \cite{ChatterjeeISIT2010}. Such a half-duplex block Markovity incurs unnecessarily long decoding delay and also wastes power to transmit the first-phase information of the current block, which is decoded backwardly. \item Joint decoding of both the state and the binning auxiliary random variable at $T_1$ is not valid (similar to \cite{CaoISIT2007,YangTIT2011}). The rate region is thus larger than possible, but can be corrected in this step. \item This scheme only covers half-duplex decode-forward relaying (when there is no binning) instead of partial decode-forward relaying and hence achieves a maximum rate for $R_1$ smaller than \eqref{eq:HD_Gaussian_Bin_maxR1}. \end{itemize} \subsubsection{Numerical Examples} In this section, we provide numerical results to compare the two proposed schemes with the Han-Kobayashi and other known coding schemes \cite{Devroye2006TIT,ChatterjeeISIT2010} for the half-duplex CRC. Figure \ref{fig:HD_PDFBin_HKPDFBin_comparison} shows the comparison between half-duplex PDF-binning, HK-PDF-binning and the Han-Kobayashi scheme. It can be seen that although PDF-binning has a larger maximum rate for $R_1$ than the Han-Kobayashi scheme, it is not always better. But the half-duplex HK-PDF-binning rate region encompasses both the Han-Kobayashi and the PDF-binning regions. \begin{figure}[t] \centering \includegraphics[scale=0.45]{HD_PDFBin_HKPDFBin_comparison.pdf} \caption{Comparison of four coding schemes (HD = half-duplex, FD = full-duplex).} \label{fig:HD_PDFBin_HKPDFBin_comparison} \end{figure} \begin{figure}[t] \centering \includegraphics[scale=0.45]{HD_HKPDFBin_Devroye_comparison.pdf} \caption{Comparison of the HK-PDF-binning schemes with existing schemes.} \label{fig:HD_HKPDFBin_Devroye_comparison} \end{figure} In Figure \ref{fig:HD_HKPDFBin_Devroye_comparison}, we compare the HK-PDF-binning schemes with existing half-duplex schemes for the CRC in \cite{Devroye2006TIT, ChatterjeeISIT2010}. We can see that HK-PDF-binning is strictly better than all existing schemes. Furthermore, the proposed scheme is more comprehensive than the protocols in \cite{Devroye2006TIT} and simpler than the scheme in \cite{ChatterjeeISIT2010}. These figures also show that the gap in achievable rates by the HK-PDF-binning scheme in the half- and full-duplex modes is quite small. Thus, the rate loss caused by the half-duplex constraint appears to be insignificant \section{Conclusion} In this paper, we have proposed two new coding schemes for both the full- and half-duplex cognitive relay channels. These two schemes are based on partial decode-forward relaying, Gelfand-Pinsker binning and Han-Kobayashi coding. The half-duplex schemes are adapted from the full-duplex schemes by sending different message parts in different phases, removing the block Markov encoding and applying joint decoding across both phases. When applied to Gaussian channels, different from the traditional binning in dirty paper coding, in which the transmit signal is independent of the state, here we introduce a correlation between the transmit signal and the state, which enlarges the rate region by allowing both binning and forwarding. We also derive the optimal binning parameter for each coding scheme. Results show that the proposed binning schemes achieve a higher rate than all existing schemes for user $1$ by allowing user $2$ to also forward a part of the message of user $1$. Furthermore, the Han-Kobayashi PDF-binning scheme contains both the Han-Kobayashi scheme and partial decode-forward relaying and outperforms all existing schemes by achieving a larger rate region for both users. Numerical results also suggest that the difference in achievable rates between the half- and full-duplex modes for the CRC is small. \appendices \section{Proof of the optimal binning parameter $\lambda^*$ for full-duplex HK-PDF-binning}\label{proof:FD_Gaussian_Combined_lambda} To simultaneously maximize $R_1$ and $R_2$ in region \eqref{eq:FD_DMC_combined}, we can simply maximize the term $I_{11}-I_1$ as follows. \begin{align*} &I(U_{22};Y_{2}|U_{21},U_{11}) - I(U_{22};T_{10}|U_{21}) \\ &= H(Y_{2}|U_{21},U_{11}) - H(Y_{2}|U_{21},U_{22},U_{11}) - H(U_{22}|U_{21}) + H(U_{22}|T_{10},U_{21})\\ &= H(Y_{2}') - H(Y_{2}'|U_{22}') - H(U_{22}') + H(U_{22}|T_{10},U_{21})\\ &= H(Y_{2}') + H(U_{22}|T_{10},U_{21}) - H(U_{22}',Y_{2}'), \end{align*} where \begin{align*} Y_{2}' &= Y_{2}|U_{21},U_{11} = (a\alpha + \mu \rho) S_{10} ' + a\beta S_{10} + a\delta S_{12} + \mu \sqrt{1-\rho^2}S_{22} + Z_2\\ U_{22}' &= U_{22}|U_{21},U_{11} = (\mu\rho+\lambda)S_{10}' + \mu\sqrt{1-\rho^2}S_{22}. \end{align*} Note that $\lambda$ only affects the last term $H(U_{22}',Y_{2}')$. The covariance matrix between $U_{22}'$ and $Y_{2}' $ is \begin{align}\label{eq:FD_Gaussian_Combined_cov} \text{cov}(U_{22}', Y_{2}') = \begin{bmatrix} \text{var}(U_{22}') & \text{E}(U_{22}',Y_{2}')\\ \text{E}(U_{22}',Y_{2}') & \text{var}(Y_{2}') \end{bmatrix}, \end{align} where \begin{align*} \text{var}(U_{22}') &= \mu ^2 + \lambda ^2 + 2\mu\rho\lambda, \\ \text{E}(U_{22}',Y_{2}')&=(\mu\rho+\lambda)(a\alpha+\mu\rho) + \mu^2(1-\rho^2),\\ \text{var}(Y_{2}') &= (a\alpha+\mu\rho)^2+ a^2\beta^2 + a^2\delta^2+ \mu^2(1-\rho^2) + 1. \end{align*} Minimizing the determinant of the matrix in \eqref{eq:FD_Gaussian_Combined_cov} leads to the optimal $\lambda$ as in \eqref{eq:lambda_FD_Gaussian_Combined}. \section{Proof of Theorem \ref{thm:HD_DMC_Bin} (Half-duplex PDF-binning) }\label{proof:HD_DMC_Bin} We use random codes and fix a joint probability distribution \begin{align*} p(x_{11})p(x_{12}|x_{11})p(u_2|x_{11})p(x_{22}|x_{11},u_2). \end{align*} \subsection{Codebook generation} \begin{itemize} \item Independently generate $2^{nR_{10}}$ sequences $x_{11}^n \sim \prod_{k=1}^{n}$ $p(x_{11k})$. Index these codewords as $x_{11}^n(w_{10})$, $w_{10}\in [1,2^{nR_{10}}]$. \item For each $x_{11}^n(w_{10})$, independently generate $2^{nR_{11}}$ sequences $x_{12}^n \sim \prod_{k=1}^{n}p(x_{12k}|x_{11k}$). Index these codewords as $x_{12}^n(w_{11}|w_{10})$, $w_{11} \in [1,2^{nR_{11}}]$, $w_{10} \in [1,2^{nR_{10}}]$. \item Independently generate $2^{n(R_{2}+ R_{2}')}$ sequences $u_2^n \sim \prod_{k=1}^{n}p(u_{2k})$. Index these codewords as $ u_2^n( w_{2}, v_{2})$, $w_{2} \in [1,2^{nR_{2}}]$ and $v_{2} \in [1,2^{nR_{2}'}]$. \item For each $x_{11}(w_{10})$ and $u_2^n(w_{2}, v_{2})$, generate one $x_{22}^n \sim \prod_{k=1}^{n}p(x_{22i}|x_{11k}, u_{2k})$. Index these codewords as $x_{22}^n(w_{10}, w_{2}, v_2)$, $w_{2} \in [1,2^{nR_{2}}]$, $v_{2} \in [1,2^{nR_{2}'}]$. \end{itemize} \subsection{Encoding} \begin{itemize} \item In the first phase, $S_1$ sends the codewords $x_{11}^{ \tau n}(w_{10})$. $S_2$ does not send anything. \item In the second phase, $S_1$ sends $x_{12}^{\bar{\tau}n}(w_{11}|w_{10})$. For $S_2$, it searches for a $v_{2}$ such that \begin{align} (x_{11}^{\bar{\tau}n}{(w_{10})}, u_2^{\bar{\tau}n}( w_{2}, v_{2}) ) \nonumber \in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} U_2}). \end{align} Such $v_{2}$ exists with high probability if \begin{align} R_{2}' \geq \bar{\tau} I( U_2; X_{11}). \end{align} $S_2$ then transmits $x_{22}^{\bar{\tau}n}(w_{10}, w_{2}, v_{2})$. \end{itemize} \subsection{Decoding} \begin{itemize} \item At the end of the first phase, $S_2$ searches for a unique $\hat{w}_{10}$ such that \begin{align*} (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y}) \in A_\epsilon^{(\tau n)}(P_{X_{11} Y}). \end{align*} We can show that the decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align}\label{eq:sp_bin_begin} R_{10} &\leq \tau I(X_{11};Y). \end{align} \item At the end of the second phase, $T_1$ searches for a unique ($\hat{w}_{10},\hat{w}_{11}$) such that \begin{align*} ( x_{11}^{\bar{\tau}n}{(\hat{w}_{10})}, x_{12}^{\bar{\tau}n}(\hat{w}_{11}|\hat{w}_{10}), \mathbf{y_{12}})&\in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} X_{12} Y_{12}})&\\ \text{and} \quad (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y_{11}}) &\in A_\epsilon^{(\tau n)}(P_{X_{11} Y_{11}}).& \end{align*} Here $\mathbf{y_{11}}$ and $\mathbf{y_{11}}$ indicate the received vectors at $T_1$ during the first and second phases. The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{11} &\leq \bar{\tau} I(X_{12};Y_{12}|X_{11}) \nonumber \\ R_{10} + R_{11} &\leq \bar{\tau} I(X_{11}, X_{12};Y_{12}) + \tau I(X_{11}; Y_{11}). \end{align} \item $T_2$ treats the codeword $X_{11}^{\bar{\tau}n}$ from $S_1$ as the state and decodes $w_{2}$. It searches for a unique $\hat{w}_2$ for some $\hat{v}_{2} $ such that \begin{align*} (u_2^{\bar{\tau}n}(\hat{w}_{2} ,\hat{v}_{2}), \mathbf{y_{22}}) \in A_\epsilon^{(\bar{\tau}n)}(P_{U_2 Y_{22}}). \end{align*} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} \label{eq:sp_bin_end} R_2 + R_2' &\leq \bar{\tau} I(U_2;Y_{22}). \end{align} \end{itemize} Combine all the above rate constraints, we get \begin{align} R_{2}' &\geq \bar{\tau} I( U_2; X_{11})\nonumber \\ R_{10} &\leq \tau I(X_{11};Y)\nonumber \\ R_{11} &\leq \bar{\tau} I(X_{12};Y_{12}|X_{11}) \nonumber \\ R_{10} + R_{11} &\leq \bar{\tau} I(X_{11}, X_{12};Y_{12}) + \tau I(X_{11}; Y_{11}) \nonumber \\ R_2 + R_2' &\leq \bar{\tau} I(U_2;Y_{22}). \end{align} Let $R_1 = R_{10} + R_{11}$, apply Fourier-Motzkin Elimination, we get region \eqref{eq:DMC_PDF_Bin}. \section{Proof of Theorem \ref{thm:HD_DMC_combined} (Half-duplex HK-PDF-binning) }\label{proof:HD_DMC_Combined} We use random codes and fix a joint probability distribution \begin{align*} &p(x_{11})p(u_{11})p(x_{12}|x_{11},u_{11})p(u_{21} p(u_{22}|u_{21},x_{11})p(x_{22}|x_{11},u_{21},u_{22}). \end{align*} \subsection{Codebook generation} \begin{itemize} \item Independently generate $2^{nR_{10}}$ sequences $x_{11}^n \sim \prod_{k=1}^{n}p(x_{11k})$. Index these codewords as $x_{11}^n(w_{10})$, $w_{10}\in [1,2^{nR_{10}}]$. \item Independently generate $2^{nR_{11}}$ sequences $u_{11}^n \sim \prod_{k=1}^{n}p(u_{11k}$). Index these codewords as $u_{11}^n(w_{11})$, $w_{11} \in [1,2^{nR_{11}}]$. \item For each $x_{11}^n(w_{10})$ and $u_{11}^n(w_{11})$, independently generate $2^{nR_{12}}$ sequences $x_{12}^n \sim \prod_{k=1}^{n}p(x_{12k}|x_{11k},u_{11k}$). Index these codewords as $x_{12}^n(w_{12}|w_{10},w_{11})$, $w_{12} \in [1,2^{nR_{12}}]$. \item Independently generate $2^{n(R_{21})}$ sequences $u_{21}^n \sim \prod_{k=1}^{n}p(u_{21k})$. Index these codewords as $u_{21}^n( w_{21})$, $w_{21} \in [1,2^{nR_{21}}]$. \item For each $u_{21}^n( w_{21})$, independently generate $2^{n(R_{22}+R_{22}')}$ sequences $u_{22}^n \sim \prod_{k=1}^{n}p(u_{22k}|u_{21k}$). Index these codewords as $u_{22}^n(w_{22},v_{22}|w_{21})$, $w_{22} \in [1,2^{nR_{22}}]$, $v_{22} \in [1,2^{nR_{22}}]$. \item For each $x_{11}(w_{10})$, $u_{21}^n(w_{21})$ and $u_{22}^n(w_{22}, v_{22}|w_{21})$, generate one $x_{22}^n \sim \prod_{k=1}^{n}p(x_{22k}|u_{22k},u_{21k},x_{11k})$. Index these codewords as $x_{22}^n(w_{10}, w_{21},w_{22}, v_{22})$. \end{itemize} \subsection{Encoding} \begin{itemize} \item In the first phase, $S_1$ sends the codewords $X_{11}^{ \tau n}(w_{10})$. $S_2$ does not send anything. \item In the second phase, $S_1$ sends $x_{12}^{\bar{\tau}n}(w_{12}|w_{10},w_{11})$. For $S_2$, it searches for some $v_{22}$ such that \begin{align*} (x_{11}^{\bar{\tau}n}{(w_{10})},u_{21}^{\bar{\tau}n}( w_{21}), &u_{22}^{\bar{\tau}n}( w_{22}, v_{22}|w_{21}) \in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} U_{22}|U_{21}}). \end{align*} Such $v_{22}$ exists with high probability if \begin{align} R_{22}' &\geq \bar{\tau} I( U_{22}; X_{11}|U_{21}). \end{align} $S_2$ then transmits $x_{22}^{\bar{\tau}n}(w_{10},w_{21},w_{22}, v_{22})$. \end{itemize} \subsection{Decoding} \begin{itemize} \item At the end of the first phase, $S_2$ searches for a unique $\hat{w}_{10}$ such that \begin{align*} (x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y}) \in A_\epsilon^{(\tau n)}(P_{X_{11} Y}). \end{align*} We can show that the decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align}\label{eq:hd_bin_begin} R_{10} &\leq \tau I(X_{11};Y). \end{align} \item At the end of the second phase, $T_1$ searches for a unique ($\hat{w}_{10},\hat{w}_{11},\hat{w}_{12}$) for some $\hat{w}_{21}$ such that \begin{align*} ( x_{11}^{\bar{\tau}n}{(\hat{w}_{10})},u_{11}^{\bar{\tau}n}{(\hat{w}_{11})}, x_{12}^{\bar{\tau}n}(\hat{w}_{12}|\hat{w}_{10},\hat{w}_{11}) u_{21}^{\bar{\tau}n}{(\hat{w}_{21})},\mathbf{y_{12}})&\in A_\epsilon^{(\bar{\tau}n)}(P_{X_{11} U_{11} X_{12} U_{21} Y_{12}})\\ \text{and} \quad x_{11}^{ \tau n}{(\hat{w}_{10})}, \mathbf{y_{11}}) &\in A_\epsilon^{(\tau n)}(P_{ X_{11} Y_{11}}). \end{align*} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{12} &\leq \bar{\tau} I(X_{12};Y_{12}|X_{11},U_{11},U_{21}) \nonumber\\ R_{10} + R_{12} &\leq \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},X_{12};Y_{12}|U_{11},U_{21}) \nonumber \\ R_{11} + R_{12} &\leq \bar{\tau} I(U_{11},X_{12};Y_{12}|X_{11},U_{21}) \nonumber \\ R_{10} + R_{11} + R_{12} &\leq \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},U_{11},X_{12};Y_{12}|U_{21}) \nonumber \\ R_{12} + R_{21} &\leq \bar{\tau} I(X_{12},U_{21};Y_{12}|X_{11},U_{11}) \nonumber \\ R_{10} + R_{12} + R_{21} &\leq \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},X_{12},U_{21};Y_{12}|U_{11}) \nonumber \\ R_{11} + R_{12} + R_{21} &\leq \bar{\tau} I(U_{11},X_{12},U_{21};Y_{12}|X_{11}) \nonumber \\ R_{10} + R_{11} + R_{12} + R_{21} &\leq \tau I(X_{11};Y_{11}) + \bar{\tau} I(X_{11},U_{11},X_{12},U_{21};Y_{12}). \end{align} \item $T_2$ uses jointly decoding to decode $(w_{11},w_{21},w_{22})$. It searches for a unique $(\hat{w}_{21},\hat{w}_{22})$ for some $(\hat{w}_{11},\hat{v}_{22})$ such that \begin{align*} (u_{11}^{\bar{\tau}n}(\hat{w}_{11} ),&u_{21}^{\bar{\tau}n}(\hat{w}_{21} ), u_{22}^{\bar{\tau}n}(\hat{w}_{22},\hat{v}_{22}|\hat{w}_{21} ),\mathbf{y_{22}}) \in A_\epsilon^{(\bar{\tau}n)}(P_{U_{11} U_{21} U_{22} Y_{22}}). \end{align*} The decoding error probability goes to 0 as $n\rightarrow \infty$ if \begin{align} R_{22} + R_{22}' &\leq \bar{\tau} I(U_{22};Y_{22}|U_{21},U_{11}) \nonumber\\ R_{21} + R_{22} + R_{22}' &\leq \bar{\tau} I(U_{21},U_{22};Y_{22}|U_{11}) \nonumber\\ R_{11} + R_{22} + R_{22}' &\leq \bar{\tau} I(U_{11},U_{22};Y_{22}|U_{21}) \nonumber\\ R_{11} + R_{21} + R_{22} + R_{22}' &\leq \bar{\tau} I(U_{11},U_{21},U_{22};Y_{22}). \end{align} \end{itemize} Let $R_1=R_{10}+R_{11}+R_{12}$ and $R_2=R_{21}+R_{22}$, apply Fourier-Motzkin Elimination on the above constraints, we get region \eqref{eq:HD_DMC_combined}. \bibliographystyle{IEEEtran}
\section{Introduction} The first realization of the laser in 1960 \cite{Maiman_1960} is one of the most important technological breakthroughs. Nowadays lasers are indispensable tools for investigating physical processes in different areas ranging from atomic and plasma physics to nuclear and high-energy physics. This has been possible mainly due to the continuous progress made along two specific directions: decrease of the laser pulse duration and increase of the laser peak intensity \cite{Mourou_2011}. On the one hand, multiterawatt laser systems with a pulse duration in the femtosecond time scale are readily available nowadays and different laboratories have succeeded in the generation of single attosecond pulses. Physics at the attosecond time scale has been the subject of the recent review \onlinecite{Krausz_2009}. In this review it has been pointed out how pulses in the attosecond domain allow for the detailed investigation of the electron motion in atoms and during molecular reactions. The production of ultrashort pulses is strongly connected with the increase of the laser peak intensity. This is not only because temporal compression evidently implies an increase in intensity at a given laser energy, but also because higher intensities allow, in general, for controlling faster physical processes, which in turn can be exploited for generating correspondingly shorter light pulses. Not long after the invention of the laser, available intensities were already sufficiently high to trigger nonlinear optical effects like second harmonic generation. It is, however, only after the experimental implementation of the Chirped Pulse Amplification (CPA) technique \cite{Strickland_1985} that it has been possible to reach the intensity threshold of $10^{14}\text{-} 10^{15}\;\text{W/cm$^2$}$ corresponding to electric field amplitudes of the same order as the Coulomb field in atoms. At such intensities the interplay between the laser and the atomic field significantly alters the electron's dynamics in atoms and molecules and this can be exploited, for example, for generating high-frequency radiation in the extreme-ultraviolet (XUV) and soft-x-ray regions (high-order harmonic generation or HHG) \cite{Protopapas_1997,Agostini_2004}. HHG as well as atomic processes in intense laser fields have been recently reviewed in \onlinecite{Winterfeldt_2008,Teubner_2009,Fennel_2010}, with specific emphasis on the control of high-harmonic spectra by spatio-temporal shaping of the driving pulse \cite{Winterfeldt_2008}, on harmonic generation in laser-plasma interaction \cite{Teubner_2009} and on the dynamics of clusters in strong laser fields \cite{Fennel_2010}. By increasing the optical laser intensity to the order of $10^{17}\text{-} 10^{18}\;\text{W/cm$^2$}$, another physically important regime in laser-matter interaction is entered: the relativistic regime. In such intense electromagnetic fields an electron reaches relativistic velocities already within one laser period, the magnetic component of the Lorentz force becomes of the same order of magnitude of the electric one, and the electron's motion becomes highly nonlinear as a function of the laser's electromagnetic field. Although the increasing influence of the magnetic force causes a suppression of atomic HHG in the relativistic domain, the highly nonlinear motion of the electrons in such strong laser fields is at the origin of numerous new effects as relativistic self-focusing in plasma and laser wakefield acceleration \cite{Mulser_b_2010}. In the recent reviews \onlinecite{Salamin_2006,Mourou_2006,Ehlotzky_2009}, different processes occurring at relativistic laser intensities are discussed. In particular, in \onlinecite{Ehlotzky_2009}, QED processes like Compton, Mott and M\o ller scattering in a strong laser field are covered, in \onlinecite{Mourou_2006}, technical aspects and new possibilities of the CPA techniques are reviewed together with relativistic effects in laser-plasma interaction as, for example, self-induced transparency and wakefield generation, while in \onlinecite{Salamin_2006}, spin-effects as well as relativistic multiphoton and tunneling recollision dynamics in laser-atom interactions are reviewed. Also in the same year another review was published on nonlinear collective photon interactions, including vacuum-polarization effects in a plasma \cite{Marklund_2006}. Whereas, the physics of plasma-based laser electron accelerators is the main subject covered in \onlinecite{Esarey_2009,Malka_2011}, with a special focus on the different phases involved (electron injection and trapping, and pulse propagation) and on the role of plasma instabilities in the acceleration process. Finally, in \onlinecite{Ruffini_2010_b} different processes related to electron-positron ($e^+\text{-}e^-$) pair production are reviewed with special emphasis on those occurring in the presence of highly-charged ions and in astrophysical environments. In the present article we address physical processes that mainly occur at optical laser intensities mostly larger than $10^{21}\;\text{W/cm$^2$}$, i.e., well exceeding the relativistic threshold. After reporting on the latest technological progress in optical and x-ray laser technology (Sec. \ref{NRS}), we review some basic results on the classical and quantum dynamics of an electron in a laser field (Sec. \ref{FED}). Then, we bridge to lower-intensity physics by reviewing more recent advances in relativistic ionization and HHG in atomic gases (Sec. \ref{RD}). The main subject of the review, i.e., the response of fundamental systems like electrons, photons and even the vacuum to ultra-intense radiation fields is covered in Secs. \ref{TCS}-\ref{MPP}. As will be seen, such high laser intensities represent a unique tool to investigate fundamental processes like multiphoton Compton scattering (Sec. \ref{TCS}), to clarify conceptual issues like radiation reaction in classical and quantum electrodynamics (Sec. \ref{RR}) and to investigate the structure of the quantum vacuum (Sec. \ref{VPEs}). Also, other fundamental quantum-relativistic phenomena like the transformation of pure light into massive particles as electrons, muons and pions (and their corresponding antiparticles) can become feasible and can even limit the attainability of arbitrarily high laser intensities (Secs. \ref{PP}-\ref{MPP}). Finally, we will also review recent suggestions on employing novel high-frequency lasers and laser-accelerated particle beams to directly trigger nuclear and high-energy processes (Secs. \ref{NP}-\ref{Part_Phys}). The main conclusions of the article will be presented in Sec. \ref{Conlc}. Units with $\hbar=c=1$ and the space-time metric $\eta^{\mu\nu}=\text{diag}(+1,-1,-1,-1)$ are employed throughout this review. \section{Novel radiation sources} \label{NRS} In this section we review the latest technical and experimental progress in laser technology. We will discuss optical and x-ray laser systems separately. The latter are especially useful for $e^+\text{-}e^-$ pair production, for direct laser-nucleus interaction, as well as as probes for vacuum-polarization effects (see in particular Secs. \ref{VPEs}, \ref{PP} and \ref{NP}). For overviews of feasible accelerators also of relevance for the present review see, e.g., \onlinecite{Wilson_b_2001,Esarey_2009,Malka_2011,PDG_2010} and the relevant original literature as quoted in the respective sections. \subsection{Strong optical laser sources} \label{L_O} As has been mentioned in the Introduction, since the invention of the CPA technique \cite{Strickland_1985} laser peak intensities have been boosted by several orders of magnitude. Another amplification technique called Optical Parametric Chirped Pulse Amplification (OPCPA), based on the nonlinear interaction among laser beams in crystals, was suggested almost at the same time as the CPA and proved to be promising as well \cite{Piskarskas_1986}. As a result of the increase in available laser intensities, exciting perspectives have been envisaged in different fields spanning from atomic to plasma and even nuclear and high-energy physics \cite{Gerstner_2007,Feder_2010,Mourou_2010,Tajima_2010}. The group of G. Mourou at the University of Michigan (Michigan, USA) holds the record so far for the highest laser intensity ever achieved of $2\times 10^{22}\;\text{W/cm$^2$}$ \cite{Yanovsky_2008}, while no experiments have been performed at this intensity yet. This record intensity has been reached when the HERCULES laser was upgraded to become a $300\;\text{TW}$ Ti:Sa system, amplified via CPA and capable of a repetition rate of $0.1\;\text{Hz}$. The 4-grating compressor allowed for a pulse duration of about $30\;\text{fs}$ and adaptive optics together with a $f/1$ parabola enabled to focus the beam down to a diameter of about $1.3\;\text{$\mu$m}$. This experimental achievement on the laser intensity pushed the capabilities of a multiterawatt laser almost to the limit. The $1\text{-PW}$ threshold has been already reached and even exceeded in various laboratories. For example, the Texas Petawatt Laser (TPL) at the University of Texas at Austin (Texas, USA) has exceeded the Petawatt threshold thanks to the OPCPA technique, by compressing an energy of $186\;\text{J}$ in a pulse lasting only $167\;\text{fs}$ \cite{TPL}. The TPL has been employed for investigating laser-plasma interactions at extreme conditions, particularly relevant for astrophysics. Also, the two laser systems Vulcan \cite{Vulcan} and Astra Gemini \cite{AG} at the Central Laser Facility (CLF) in the United Kingdom provide powers of the order of $1\;\text{PW}$. The Vulcan facility can deliver an energy of $500\;\text{J}$ in a pulse lasting $500\;\text{fs}$. It is a Nd:YAG laser system amplified via CPA and can provide intensities up to $10^{21}\;\text{W/cm$^2$}$. Whereas, the Astra Gemini laser consists of two independent Ti:Sa laser beams of $0.5\;\text{PW}$ each (energy of $15\;\text{J}$ and a pulse duration of $30\;\text{fs}$), with a maximum focused intensity of $10^{22}\;\text{W/cm$^2$}$. The particular layout of the Astra Gemini laser renders this system especially versatile for unique applications in strong-field physics, where two ultrastrong beams are required. Two laser systems are likely to be updated to the Petawatt level in Germany. The first one is the Petawatt High-Energy Laser for heavy Ion eXperiments (PHELIX) Nd:YAG laser at the Gesellschaft f\"{u}r Schwerionenforschung (GSI) in Darmstadt, capable now of delivering an energy of $120\;\text{J}$ in about $500\;\text{fs}$ \cite{PHELIX}. The 1-PW threshold should be reached by increasing the pulse energy to $500\;\text{J}$. Combined with the highly-charged ion beams at GSI, the PHELIX facility can be attractive for experimental investigations in strong-field quantum electrodynamics (QED). The second system to be updated to $1\;\text{PW}$ is the Petawatt Optical Laser Amplifier for Radiation Intensive Experiments (POLARIS) laser in Jena \cite{Hein_2010}. At the moment, a power of about $100\;\text{TW}$ (energy of $10\;\text{J}$ for a pulse duration of $100\;\text{fs}$) has been reached and the goal of $1\;\text{PW}$ power should be achieved by compressing $120\;\text{J}$ in about $120\;\text{fs}$. The Scottish Centre for the Application of Plasma-based Accelerators (SCAPA) research center is one of the main initiatives within the Scottish Universities Physics Alliance (SUPA) project dedicated to the high-power laser interaction with plasmas. A laser system will be developed, which will generate pulses of 5-7 J energy and of 25-30 fs duration at a repetition rate of 5 Hz, corresponding to a peak power of 200-250 TW, with potential for future upgrades to the petawatt level \cite{SCAPA_2012}. The 1-PW threshold has been also exceeded in Ti:Sa laser systems like those described in \onlinecite{Sung_2010} (energy of $34\;\text{J}$ for a pulse duration of $30\;\text{fs}$) and in \onlinecite{Wang_2011} (energy of $32.3\;\text{J}$ for a pulse duration of $27.9\;\text{fs}$) and constructed at the Advanced Photonics Research Institute (APRI) at Gwangju (Republic of Korea) and at the Beijing National Laboratory for Condensed Matter Physics in Beijing (China), respectively. The BELLA (Berkeley Lab Laser Accelerator) is a Ti:Sa laser system under construction at the Lawrence Berkeley National Laboratory (LBNL) at Berkeley (California, USA) which will also reach the 1-PW threshold by compressing $40\;\text{J}$ in $40\;\text{fs}$ at a repetition rate of $1\;\text{Hz}$ \cite{BELLA}. All the above systems require laser energies larger than $10\;\text{J}$ and this limits the repetition rate of existing petawatt lasers in the best situation to about $0.1\;\text{Hz}$ \cite{Sung_2010}. The Ti:Sa Petawatt Field Synthesizer (PFS) system under development in Garching (Germany) aims to be the first high-repetition rate petawatt laser system with an envisaged repetition rate of $10\;\text{Hz}$ \cite{PFS}. By adopting the OPCPA technique the PFS should reach the petawatt level by compressing an energy of about $5\;\text{J}$ in $5\;\text{fs}$ \cite{Major_2010}. For a recent review on petawatt-class laser systems, see \onlinecite{Korzhimanov_2011}. Finally, we also want to mention other high-power lasers, mainly devoted to fast ignition and characterized by relatively long pulses of the order of $1\;\text{ps}\text{-} 1\;\text{ns}$. Among others we mention the OMEGA EP system at Rochester (New York, USA) (energy of $1\;\text{kJ}$ for a pulse duration of $1\;\text{ps}$) \cite{OMEGA_EP} and the National Ignition Facility (NIF) at the Lawrence Livermore National Laboratory (LLNL) at Livermore (California, USA) (energy of $2\;\text{MJ}$ distributed in 192 beams with a pulse duration of about $3\text{-}10\;\text{ns}$) \cite{NIF}. Another high-power laser facility is the PETawatt Aquitaine Laser (PETAL) in Le Barp close to Bordeaux (France), which is a multi-petawatt laser, generating pulses with energy up to 3.5 kJ and with a duration of 0.5 to 5 ps \cite{PETAL}. The PETAL facility is planned to be coupled to the Laser M\'{e}gaJoule (LMJ) under construction in Bordeaux (France). In the LMJ a total energy of $1.8\;\text{MJ}$ is distributed in a series of 240 laser beamlines, collected into eight groups of 30 beams with a pulse duration of $0.2\text{-} 25\;\text{ns}$ \cite{LMJ_2011}. \subsubsection{Next-generation $10\;\text{PW}$ optical laser systems} The possibility of building a $10\;\text{PW}$ laser system is under consideration in various laboratories. At the CLF in the United Kingdom the $10\;\text{PW}$ upgrade of the Vulcan laser has already started \cite{Vulcan_10PW}. The new laser will provide beams with an energy of $300\;\text{J}$ in only $30\;\text{fs}$ via the OPCPA. A $10\;\text{PW}$ laser system is in principle capable of unprecedented intensities larger than $10^{23}\;\text{W/cm$^2$}$ if the beam is focused to about $1\;\text{$\mu$m}$. The front-end stage of the Vulcan $10\;\text{PW}$ is already completed and it delivers pulses with about $1\;\text{J}$ of energy at a central wavelength of $0.9\;\text{$\mu$m}$, with sufficient bandwidth to support a pulse with duration less than 30 fs. Another $10\;\text{PW}$ laser project is the ILE APOLLON to be realized at the Institut de Lumi\'{e}re Extreme (ILE) in France \cite{Chambaret_2009}. The laser pulses are expected to deliver an energy of $150\;\text{J}$ in $15\;\text{fs}$ at the last stage of amplification after the front end (energy of $100\;\text{mJ}$ in less than $10\;\text{fs}$), with a repetition rate of one shot per minute. Laser intensities of the order of $10^{24}\;\text{W/cm$^2$}$ are envisaged at the ILE APOLLON system, entering the so-called ultrarelativistic regime, where also ions (rest energy of the order of $1\;\text{GeV}$) become relativistic within one laser period of such an intense laser field. We mention here also the PEtawatt pARametric Laser (PEARL-10) project at the Institute of Applied Physics of the Russian Academy of Sciences in Nizhny Novgorod (Russia), which is an upgrade of the present 0.56-PW laser employing the OPCPA technique, to $10\;\text{PW}$ ($200\;\text{J}$ of energy compressed in $20\;\text{fs}$) \cite{Korzhimanov_2011}. \subsubsection{Multi-Petawatt and Exawatt optical laser systems} The Extreme Light Infrastructure (ELI) \cite{ELI} (see Fig. \ref{Megalaser}), the Exawatt Center for Extreme Light Studies (XCELS) \cite{XCELS} and the High Power laser Energy Research (HiPER) facility at the CLF in the United Kingdom \cite{HiPER} are envisaged laser systems with a power exceeding the $100\;\text{PW}$ level. ELI is a large-scale laser facility consisting of four ``pillars'' (see Fig. \ref{Megalaser}): one devoted to nuclear physics, one to attosecond physics, one to secondary beams (photon beams, ultrarelativistic electron and ion beams) and one to high-intensity physics. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig01.eps} \end{center} \caption{(Color online) Summary of the four pillars of ELI. A power value of $10(\times2)\;\text{PW}$ indicates the availability of two laser systems each with $10\;\text{PW}$ power. Reprinted with permission from \onlinecite{Feder_2010}. Copyright 2010, American Institute of Physics.} \label{Megalaser} \end{figure} This last one is of relevance here and it is supposed to comprise ten beams each with a power of $10\text{-} 20\;\text{PW}$ that, when combined in phase, should deliver a single beam of about $100\text{-} 200\;\text{PW}$ at a repetition rate of one shot per minute. The relatively high repetition rate is obtained by compressing in each beam alone $0.3\text{-} 0.4\;\text{kJ}$ of energy in a pulse of $15\;\text{fs}$. We mention that one of the aims of the ILE APOLLON system is to provide a prototype of the $10\text{-} 20\;\text{PW}$ beams, that will be then employed at ELI. In the high-field pillar of ELI ultrahigh intensities exceeding $10^{25}\;\text{W/cm$^2$}$ are envisaged, which are well above the ultrarelativistic regime. At such intensities, it will be possible to test different aspects of fundamental physics for the first time. The XCELS infrastructure is planned to be built in Nizhny Novgorod (Russia) and it will consist of 12 beams each with energy of $300\text{-}400\;\text{J}$ and with duration of $20\text{-}30\;\text{fs}$. The pulse resulting from the superposition of these beams is expected to have a power of $200\;\text{PW}$, a pulse duration of about $25\;\text{fs}$, and divergence less than 3 diffraction limits (at a central wavelength of $0.91\;\text{$\mu$m}$). Apart from aiming to overcome the $100\;\text{PW}$ threshold, the main priorities of XCELS are the creation of sources of attosecond and subattosecond, X-ray and $\gamma$-ray pulses, the development of laser based electron and ion accelerators with electron and ion energies exceeding 100 GeV and up to 10 GeV, respectively, the realization in laboratory of astrophysical and early-cosmological conditions and the investigation of the structure of the quantum vacuum. The main goal of the other large-scale facility HiPER is the first demonstration of laser-driven fusion, or fast ignition. To this end HiPER will deliver: 1) an energy of about $200\;\text{kJ}$ distributed in 40 beams with a pulse duration of several nanoseconds and a photon energy of $3\;\text{eV}$ in the compression side; 2) an energy of about $70\;\text{kJ}$ distributed in 24 beams with a pulse duration of $15\;\text{ps}$ and a photon energy of $2\;\text{eV}$ in the ignition side. Employing HiPER for high-intensity physics would imply a feasible reconfiguration of the ignition side to deliver $10\;\text{kJ}$ in only $10\;\text{fs}$ via the OPCPA technique. This would render HiPER a laser facility with Exawatt ($10^{18}\;\text{W}$) power and with a potential intensity of $10^{26}\;\text{W/cm$^2$}$. Finally, we briefly mention the GEKKO EXA facility conceptually under design in Osaka (Japan) \cite{Gekko_EXA}. This facility is expected to deliver pulses of $2\;\text{kJ}$ energy and of $10\;\text{fs}$ duration corresponding to $200\;\text{PW}$ and with an intensity up to $10^{25}\;\text{W/cm$^2$}$. \subsection{Brilliant x-ray laser sources} \label{x_ray} Strong optical laser systems are sources of coherent radiation at wavelengths of the order of $1\;\text{$\mu$m}$, corresponding to photon energies of the order of $1\;\text{eV}$. Considerable efforts have been devoted in the past few years to develop coherent radiation sources at photon energies larger than $100\;\text{eV}$. The discovery of the Self-Amplified Spontaneous Emission (SASE) regime \cite{Bonifacio_1984} has opened the possibility of employing Free Electron Lasers (FELs) to generate coherent light at such short wavelengths. In a FEL relativistic bunches of electrons pass through a spatially-periodic magnetic field (undulator) and emit high-energy photons. In the SASE regime the interaction of the electron bunch with its own electromagnetic field ``structures'' the bunch itself into slices (microbunches) each one emitting coherently even at wavelengths below $1\;\text{nm}$ (FELs at such small wavelengths are dubbed X-ray Free Electron Lasers (XFELs)). The Free-Electron Laser in Hamburg (FLASH) facility at the Deutsches Elektronen-SYnchrotron (DESY) in Hamburg (Germany) \cite{FLASH} is one of the most brilliant operating FEL's. It delivers short pulses (duration of about $10\text{-} 100\;\text{fs}$) of coherent radiation in the extreme ultraviolet-soft x-ray regime (fundamental wavelength from $60\;\text{nm}$ down to $6.5\;\text{nm}$ corresponding to photon energies from $21$ to $190\;\text{eV}$) at a repetition rate of $100\;\text{kHz}$. The intense electron beams available at FLASH (total charge of $0.5\text{-} 1\;\text{nC}$ at an energy of $1\;\text{GeV}$) allow for peak brilliance of the photon beam of about $10^{29}\text{-} 10^{30}\;\text{photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth)}$ (see Fig. \ref{XFEL}), exceeding the peak brilliance of conventional synchrotron light sources by several orders of magnitude. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig02.eps} \end{center} \caption{(Color online) Comparison among the peak brilliances of the three facilities FLASH, LCLS and European XFEL as a function of the laser photon energy. An envisaged peak brilliance of $5\times 10^{33}$\;\text{photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth)} at a photon energy of $12.4\;\text{keV}$ for the SACLA facility is reported in \onlinecite{XFEL}. See also \onlinecite{XFEL}.} \label{XFEL} \end{figure} The Linac Coherent Light Source (LCLS) at Stanford (California, USA) uses the electron beams generated by the Stanford Linear ACcelerator (SLAC) at the National Accelerator Laboratory to generate flashes of coherent x-ray radiation of unprecedented brilliance \cite{LCLS} (see also \onlinecite{Emma_2010}). Since the electron beam energy can be varied from $4.5\;\text{GeV}$ to $14.4\;\text{GeV}$, accordingly the wavelength of LCLS can be tuned from $1.5\;\text{nm}$ to $0.15\;\text{nm}$ (corresponding to photon energies from $0.8\;\text{keV}$ to $8\;\text{keV}$). The peak brilliance of the LCLS is of the order of $10^{32}\text{-} 10^{33}$\;\text{photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth)} (see Fig. \ref{XFEL}), the pulse duration is typically of about $40\text{-}80\;\text{fs}$ up to $500\;\text{fs}$, which can be decreased to $10\;\text{fs}$ in the low-charge electron beam mode, and the repetition rate is $120\;\text{Hz}$. Another XFEL in operation is the SPring-8 Angstrom Compact free electron LAser (SACLA) at the RIKEN Harima Institute in Japan \cite{SACLA}. The electron accelerator, based on a conducting C-band high-gradient radiofrequency acceleration system, and the short-period undulator allow for a relatively compact facility of around 700 m in length (compared, for example, with the about 2 km of the LCLS). SACLA employs the 8 GeV electron beam of the Super Photon Ring - 8 GeV (SPring-8) accelerator \cite{SPring8} and has generated x-ray beams with $0.08\;\text{nm}$ wavelength (corresponding to a photon energy of $15.5\;\text{keV}$) at a repetition rate of $60\;\text{Hz}$ (see also the caption of Fig. \ref{XFEL}). The European XFEL is under development at DESY in Hamburg (Germany) \cite{XFEL}. It is expected to deliver x-ray pulses with a peak brilliance up to about $5\times 10^{33}\;\text{photons/(s\,mrad$^2$\,mm$^2$\;0.1\% bandwidth)}$ at the unprecedented repetition rate of $27\;\text{kHz}$. The electron accelerator provides an electron beam with maximal energy of $17.5\;\text{GeV}$ able to generate laser pulses with a central wavelength of $0.05\;\text{nm}$, which corresponds to a photon energy of $24.8\;\text{keV}$ and with a pulse duration of $100\;\text{fs}$. Moreover, the European XFEL will be a versatile machine consisting of three photon beamlines: the SASE-1 and SASE-2, with linearly polarized photons with energy in the range $3.1\text{-} 24.8\;\text{keV}$ and the SASE-3, with linearly or circularly polarized photons of energy in the range $0.26\text{-} 3.1\;\text{keV}$. We also mention that coherent attosecond pulses of XUV radiation (photon energy of the order of 100 eV) have been generated employing HHG in a gaseous medium \cite{Agostini_2004}. This technique allows for the production of beams with central photon energy up to several keVs \cite{Sansone_2006,Popmintchev_2009}, though with intensities several orders smaller than XFELs. Less stable sources of coherent soft x-rays are the so-called x-ray lasers, which are based on the amplification of spontaneous emission by multiply ionized atoms in dense plasmas created by intense laser pulses \cite{Zeitoun_2004,Wang_2008,Suckewer_2009}. \section{Free electron dynamics in a laser field}\label{FED} In this Section we review, for the benefit of the reader, some important basic results on the dynamics of a free electron in a laser field (see also the review \onlinecite{Eberly_1969}) and link them to recent investigations on the subject. Results in the realm of classical and quantum electrodynamics are considered separately. Radiation-reaction and electron self-interaction effects are not included here and their discussion is developed in Sec. \ref{RR}. \subsection{Classical dynamics} \label{FED_C} The motion of a charged particle in a laser field is usually associated to an oscillation along the laser polarization direction. This is pertinent to the non-relativistic regime, while the charge dynamics in the relativistic domain is enriched by new features like the drift along the laser propagation direction and other non-dipole effects (like the well-known figure-8 trajectory), as well as by the sharpening of the trajectory at those instants where the velocity along the polarization direction reverses. As a consequence, laser-driven relativistic free electrons also emit high harmonics of the laser frequency (see Sec. \ref{TCS}). The classical motion of an electron (electric charge $e<0$ and mass $m$) in an arbitrary external electromagnetic field $F^{\mu\nu}(x)$ is determined by the Lorentz equation $mdu^{\mu}/ds=eF^{\mu\nu}u_{\nu}$, where $u^{\mu}=dx^{\mu}/ds$ is the electron four-velocity and $s$ its proper time \cite{Landau_b_2_1975}. If the external field is a plane wave, the field tensor $F^{\mu\nu}(x)$ depends only on the dimensional phase $\phi=(n_0x)$, where $n_0^{\mu}=(1,\bm{n}_0)$, with $\bm{n}_0$ being the unit vector along the propagation direction of the wave. In this case, for an arbitrary four-vector $v^{\mu}=(v^0,\bm{v})$ it is convenient to introduce the notation $v_{\parallel}=\bm{n}_0\cdot\bm{v}$, $\bm{v}_{\perp}=\bm{v}-v_{\parallel}\bm{n}_0$ and $v_-=(n_0v)=v^0-v_{\parallel}$. The four-vector potential of the wave can be chosen in the Lorentz gauge as $A^{\mu}(\phi)=(0,\bm{A}(\phi))$, with $A_-(\phi)=-A_{\parallel}(\phi)=0$. We indicate as $p^{\mu}=(\varepsilon,\bm{p})=mu^{\mu}$ the (kinetic) four-momentum of the electron. Since a plane-wave field depends only on $\phi$, the canonical momenta $\bm{p}_{\perp}(\phi)+e\bm{A}(\phi)$ and $p_-(\phi)$ are conserved as they are the conjugated momenta to the cyclic coordinates $\bm{x}_{\perp}$ and $t+x_{\parallel}$, respectively. For $p^{\mu}(\phi_0)=p_0^{\mu}=(\varepsilon_0,\bm{p}_0)=m\gamma_0(1,\bm{\beta}_0)$ being the initial condition for the electron's four-momentum at a given phase $\phi_0$, the above-mentioned conservation laws already allow to write the electron's four-momentum at an arbitrary phase $\phi$ as \cite{Landau_b_2_1975} \begin{align} \label{Free_Sol_0} \begin{split} \varepsilon(\phi)=&\varepsilon_0-e\frac{\bm{p}_{0,\perp}\cdot[\bm{A}(\phi)-\bm{A}(\phi_0)]}{p_{0,-}}\\ &+\frac{e^2}{2}\frac{[\bm{A}(\phi)-\bm{A}(\phi_0)]^2}{p_{0,-}}, \end{split}\\ \label{Free_Sol_1_2} \bm{p}_{\perp}(\phi)=&\bm{p}_{0,\perp}-e[\bm{A}(\phi)-\bm{A}(\phi_0)],\\ \label{Free_Sol_3} \begin{split} p_{\parallel}(\phi)=&p_{0,\parallel}-e\frac{\bm{p}_{0,\perp}\cdot[\bm{A}(\phi)-\bm{A}(\phi_0)]}{p_{0,-}}\\ &+\frac{e^2}{2}\frac{[\bm{A}(\phi)-\bm{A}(\phi_0)]^2}{p_{0,-}}, \end{split} \end{align} where the on-shell condition $\varepsilon(\phi)+p_{\parallel}(\phi)=[\bm{p}^2_{\perp}(\phi)+m^2]/p_{0,-}$ was employed. For the paradigmatic case of a linearly polarized monochromatic plane wave, it is $A^{\mu}(\phi)=A_0^{\mu}\cos(\omega_0\phi)$, with $A_0^{\mu}=(0,E_0\bm{u}/\omega_0)$, where $E_0$ is the electric field amplitude, $\omega_0$ the angular frequency and $\bm{u}$ the polarization direction (perpendicular to $\bm{n}_0$). The above analytical solution indicates that even if an electron is initially at rest, it becomes relativistic within one laser period $T_0=2\pi/\omega_0$ if the parameter \begin{equation} \label{xi_0} \xi_0=\frac{|e|\sqrt{-A_0^2}}{m}=\frac{|e|E_0}{m\omega_0} \end{equation} is of the order of or larger than unity. In the relativistic regime the magnetic component of the Lorentz force, which depends on the electron's velocity, becomes comparable to the electric one and the electron's dynamics becomes highly nonlinear in the laser-field amplitude. Thus, the parameter $\xi_0$ is known as classical nonlinearity parameter. An heuristic interpretation of the parameter $\xi_0$ is as the work performed by the laser field on the electron in one laser wavelength $\lambda_0=T_0$ in units of the electron mass, which clearly explains why relativistic effects become important at $\xi_0\gtrsim 1$. Alternatively, Eqs. \eqref{Free_Sol_0}-\eqref{Free_Sol_3} indicate that the figure-8 trajectory has a longitudinal (transverse) extension of the order of $\lambda_0\xi_0^2$ ($\lambda_0\xi_0$), implying that the electron trajectory deviates from the unidirectional oscillating one and becomes nonlinear in the field amplitude at $\xi_0\gtrsim 1$. Note that numerically $\xi_0=6.0\sqrt{I_0[10^{20}\;\text{W/cm$^2$}]}\lambda_0[\mu\text{m}]=7.5\sqrt{I_0[10^{20}\;\text{W/cm$^2$}]}/\omega_0[\text{eV}]$, where $I_0=E_0^2/4\pi$ is the wave's peak intensity, and that $\xi_0$ is gauge- and Lorentz-invariant: the gauge invariance has to be intended with respect to gauge transformations which do not alter the dependence of the four-vector potential on $\phi$ (see \onlinecite{Heinzl_2009} for a thorough analysis of this issue). The solution in Eqs. \eqref{Free_Sol_0}-\eqref{Free_Sol_3} also indicates that in the ultrarelativistic regime at $\xi_0\gg 1$, the electron acquires a drift momentum along the propagation direction of the laser field which is proportional to $\xi_0^2$, in contrast to the transverse momentum which is proportional to $\xi_0$. In the case of an electron initially at rest, for example, the momentum $\bm{p}(\infty)$ of the electron after the laser pulse has been switched off ($\bm{A}(\infty)=\bm{0}$) has the components $\bm{p}_{\perp}(\infty)=e\bm{A}(\phi_0)$ and $p_{\parallel}(\infty)=e^2\bm{A}^2(\phi_0)/2m$. Realistic laser pulses, as those produced in laboratories, have a more complicated structure than a plane wave, essentially because they are spatially focused on the transverse planes and the area of the focusing spot changes along the laser's propagation axis. Generally speaking, if the radius of the minimal focusing area (spot radius) is much larger than the central wavelength of the laser pulse, then the pulse can be reasonably approximated by a plane wave. A Gaussian beam in the paraxial approximation offers a more accurate analytical description of a realistic laser pulse, which shows a Gaussian profile in the transverse planes \cite{Salamin_2002a}. The dynamics of an electron in such a field cannot be derived analytically and a numerical solution of the Lorentz equation is required \cite{Salamin_2002b}. \subsection{Quantum dynamics} \label{FED_Q} In the realm of relativistic quantum mechanics, i.e., when $e^+\text{-}e^-$ pair production is negligible (see also Sec. \ref{PP}) and the single-particle quantum theory is applicable, the dynamics of an electron in an external electromagnetic field with four-vector potential $A^{\mu}(x)$ is described by the Dirac equation \begin{equation} \label{Dir_Eq} \{\gamma^{\mu}[i\partial_{\mu}-eA_{\mu}(x)]-m\}\Psi=0, \end{equation} where $\gamma^{\mu}$ are the Dirac matrices and where $\Psi(x)$ is the four-component electron bi-spinor \cite{Landau_b_4_1982}. Analogously to the classical case, if the external field is a plane wave, the Dirac equation can be solved exactly. If $p_0^{\mu}=(\varepsilon_0,\bm{p}_0)$ and $\sigma_0/2=\pm 1/2$ are the electron's four-momentum and spin at $\phi\to -\infty$ and if $A^{\mu}(-\infty)=0$, the positive-energy ($\varepsilon_0>0$) solution $\Psi_{p_0,\sigma_0}(x)$ of Eq. \eqref{Dir_Eq} reads \cite{Volkov_1935,Landau_b_4_1982} \begin{equation} \label{V_S} \Psi_{p_0,\sigma_0}(x)=\left[1+\frac{e}{2p_{0,-}}\hat{n}_0\hat{A}(\phi)\right]\frac{u_{p_0,\sigma_0}}{\sqrt{2V\varepsilon_0}}e^{iS_{p_0}}, \end{equation} where in general $\hat{v}=\gamma^{\mu}v_{\mu}$ for a generic four-vector $v^{\mu}$, where $u_{p_0,\sigma_0}$ is a positive-energy free bi-spinor \cite{Landau_b_4_1982}, $V$ is the quantization volume, and where \begin{equation} S_{p_0}(x)=-(p_0x)-\int_{-\infty}^{\phi}d\phi'\left[\frac{e(p_0A(\phi'))}{p_{0,-}}-\frac{e^2A^2(\phi')}{2p_{0,-}}\right] \end{equation} is the classical action of an electron in the plane wave \cite{Landau_b_2_1975}. The above electron states are known as positive-energy Volkov states. The negative-energy states $\Psi_{-p_0,-\sigma_0}(x)$ can be formally obtained by the replacements $p_0^{\mu}\rightarrow-p_0^{\mu}$ and $\sigma_0\rightarrow -\sigma_0$ in Eq. \eqref{V_S} except for the energy in the square root (the resulting bi-spinor $u_{-p_0,-\sigma_0}$ is the corresponding negative-energy free bi-spinor \footnote{We point out that the discussed Volkov states $\Psi_{\pm p_0,\pm\sigma_0}(x)$ are the so-called Volkov in-states, as they transform into free-states in the limit $t\to-\infty$ \cite{Fradkin_b_1991}. Volkov out-states, which transform into free-states in the limit $t\to\infty$, can be derived analogously and differ from the Volkov in-states only by an inconsequential constant phase factor (recall that $\bm{A}(\infty)=\bm{0}$).} \cite{Landau_b_4_1982}). Although it has been shown long ago that positive- and negative-energy Volkov states form a complete set of orthogonal states on the hypersurfaces $\phi=\text{const}$ \cite{Ritus_1985}, the corresponding property on the hypersurfaces $t=\text{const}$ is not straightforward and it has been proved only recently (see \onlinecite{Ritus_1985,Zakowicz_2005}, and \onlinecite{Boca_2010} for a proof of the orthogonality and of the completeness of the Volkov states, respectively). Since the Volkov states form a basis of the space of the solutions of Dirac equation in a plane wave, they can be employed to build electron wave packets and study their evolution. A pedagogical example of laser-induced Dirac dynamics is displayed in Fig. \ref{El_dyn} for a plane wave with peak intensity of $6.3\times 10^{23}\;\text{W/cm$^2$}$ and central wavelength of $2\;\text{nm}$. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig03.eps} \end{center} \caption{(Color) Free wave packet evolution in a plane wave field. The solid gray line indicates the center of mass trajectory, coinciding essentially with the classical trajectory, and the laser pulse travels from left to right. The blue regions indicate the copropagating self-adaptive numerical grid. Time and space coordinates are given in ``atomic units'', with $1\;\text{a.u.}=24\;\text{as}$ and $1\;\text{a.u.}=0.05\;\text{nm}$, respectively. From \onlinecite{Bauke_2011b}.} \label{El_dyn} \end{figure} The figure shows the drift of the wave packet in the propagation direction of the wave, its spreading and its shearing due to non-dipole effects. In \onlinecite{Fillion-Gourdeau_2012} an alternative method of solving the time-dependent Dirac equation in coordinate space is presented, which explicitly avoids the fermion doubling, i.e., the appearance of unphysical modes when the Dirac equation is discretized. As in the classical case, we shortly mention here the paradigmatic case of a monochromatic, linearly polarized plane-wave field $A^{\mu}(\phi)=A_0^{\mu}\cos(\omega_0\phi)$. In this case the action $S_{p_0}(x)$ can be written in the form $S_{p_0}(x)=-(q_0x)+\text{``oscillating terms''}$, with \cite{Ritus_1985} \begin{equation} \label{q} q_0^{\mu}=p_0^{\mu}+\frac{m^2\xi_0^2}{4p_{0,-}}n_0^{\mu}. \end{equation} The four-vector $q_0^{\mu}$ plays the role of an ``effective'' four-momentum of the electron in the laser field and it is indicated as electron ``quasimomentum''. The corresponding electron ``mass'' $\sqrt{q_0^2}=m^*=m\sqrt{1+\xi_0^2/2}$ is known as electron's dressed mass. The results for the quasimomentum $q_0^{\mu}$ and the dressed mass $m^*$ in the case of a circularly polarized laser field with the same amplitude and frequency is obtained from the above ones with the replacement $\xi_0^2\rightarrow 2\xi_0^2$. The quasimomentum coincides classically with the average momentum of the electron in the plane wave. Correspondingly, the mass dressing depends only on the classical nonlinearity parameter $\xi_0$ and it is an effect of the quivering motion of the electron in the monochromatic wave (see also the recent review \onlinecite{Ehlotzky_2009}). As we will see in Sec. \ref{TCS_General}, it is important that conservation laws in QED processes in the presence of a monochromatic plane-wave field involve the quasimomentum $q_0^{\mu}$ for the incoming electrons rather than the four-momentum $p_0^{\mu}$. The question of the electron dressed mass in pulsed laser fields has been investigated in \onlinecite{Heinzl_2010b,Mackenroth_2011}. In the realm of QED the parameter $\xi_0$ can also be heuristically interpreted as the work performed by the laser field on the electron in the typical QED length $\lambda_C=1/m\approx 3.9\times 10^{-11}\;\text{cm}$ (Compton wavelength) in units of the laser photon energy $\omega_0$ (see Eq. \eqref{xi_0}). This qualitatively explains why multiphoton effects in a laser field become important at $\xi_0\gtrsim 1$, such that the laser field has to be taken into account exactly in the calculations \cite{Ritus_1985}. In the framework of QED this is achieved by working in the so-called Furry picture \cite{Furry_1951}, where the $e^+\text{-}e^-$ field $\Psi(x)$ is quantized in the presence of the plane-wave field. This amounts essentially in employing the Volkov (dressed) states and the corresponding Volkov (dressed) propagators \cite{Ritus_1985} instead of free particle states and free propagators to compute the amplitudes of QED processes. In the Furry picture the effects of the plane wave are accounted for exactly and only the interaction between the $e^+\text{-}e^-$ field $\Psi(x)$ and the radiation field $\mathcal{F}^{\mu\nu}(x)\equiv\partial^{\mu}\mathcal{A}^{\nu}(x)-\partial^{\nu}\mathcal{A}^{\mu}(x)$ is accounted for by means of perturbation theory. The complete evolution of the system ``$e^+\text{-}e^-$ field+radiation field'' is obtained by means of the $S$-matrix \begin{equation} \label{S} S=\mathcal{T}\left[\text{exp}\left(-ie\int d^4x\bar{\Psi}\gamma^{\mu}\Psi\mathcal{A}_{\mu}\right)\right], \end{equation} where $\mathcal{T}$ is the time-ordering operator and $\bar{\Psi}(x)=\Psi^{\dag}(x)\gamma^0$. For an initial state containing only a single electron with four-momentum $p_0^{\mu}$, the quantitative description of the interaction between the electron, the laser field and the radiation field involves in particular the gauge- and Lorentz-invariant quantum parameter \begin{equation} \label{chi_0} \chi_0=\frac{|e|\sqrt{-(F_{0,\mu\nu}p_0^{\nu})^2}}{m^3}=\frac{p_{0,-}}{m}\frac{E_0}{F_{\text{cr}}}, \end{equation} where $F_{\text{cr}}=m^2/|e|=1.3\times 10^{16}\;\text{V/cm}=4.4\times 10^{14}\;\text{G}$ being the critical electromagnetic field of QED \cite{Ritus_1985}. The definition of $F_{\text{cr}}$ indicates that a constant and uniform electric field with strength of the order of $F_{\text{cr}}$, provides an $e^+\text{-}e^-$ pair with an energy of the order of its rest energy $2m$ in a distance of the order of the Compton wavelength $\lambda_C$, implying the instability of the vacuum under $e^+\text{-}e^-$ pair creation in the presence of such a strong field \cite{Sauter_1931,Heisenberg_1936,Schwinger_1951}. In Eq. \eqref{chi_0} we considered the case of a linearly polarized plane wave of the form $A^{\mu}(\phi)=A_0^{\mu}\psi(\phi)$, with $\psi(\phi)$ being an arbitrary function with $\max|d\psi(\phi)/d\phi|=1$ and we introduced the tensor amplitude $F_0^{\mu\nu}=k_0^{\mu}A_0^{\nu}-k_0^{\nu}A_0^{\mu}$, with $k_0^{\mu}=\omega_0n_0^{\mu}$ . For an ultrarelativistic electron initially counterpropagating with respect to the plane wave it is $\chi_0=5.9\times 10^{-2}\varepsilon_0[\text{GeV}]\sqrt{I_0[10^{20}\;\text{W/cm$^2$}]}$. The parameter $\chi_0$ can be interpreted as the amplitude of the electric field of the plane wave in the initial rest-frame of the electron in units of the critical field of QED and it controls the magnitude of pure quantum effects like the photon recoil in multiphoton Compton scattering and spin effects. This is why it is known as ``nonlinear quantum parameter''. Since the probability $dP_e/dVdt$ per unit volume and unit time of a quantum process is a gauge- and Lorentz invariant quantity, for those processes in a plane-wave field involving an incoming electron, as, e.g., multiphoton Compton scattering, it can depend only on the two parameters $\xi_0$ and $\chi_0$ \cite{Ritus_1985}. For an electromagnetic field $F^{\mu\nu}(x)=(\bm{E}(x),\bm{B}(x))$ either constant or slowly-varying, the quantity $dP_e/dVdt$, calculated in the latter case in the leading order with respect to the fields' derivatives, can in principle also depend on the two field invariants \begin{align} \mathscr{F}(x)&=\frac{1}{4}F^{\mu\nu}(x)F_{\mu\nu}(x)=-\frac{1}{2}[E^2(x)-B^2(x)],\\ \mathscr{G}(x)&=\frac{1}{4}F^{\mu\nu}(x)\tilde{F}_{\mu\nu}(x)=-\bm{E}(x)\cdot\bm{B}(x) \end{align} which identically vanish for a plane wave. In the second equation $\tilde{F}_{\mu\nu}(x)=\epsilon_{\mu\nu\alpha\beta}F^{\alpha\beta}(x)/2$ is the dual field of $F^{\mu\nu}(x)$ and $\epsilon^{\mu\nu\alpha\beta}$ is the four-dimensional completely anti-symmetric tensor with $\epsilon^{0123}=+1$ (since $\mathscr{G}(x)$ is actually a pseudo-scalar function, the probability $dP_e/dVdt$ can only depend on $\mathscr{G}^2(x)$). Note, however, that if $|\mathscr{F}(x)|,|\mathscr{G}(x)|\ll \min(1,\chi^2(x))F_{\text{cr}}^2$, with $\chi(x)=|e|\sqrt{|(F_{\mu\nu}(x)p_0^{\nu})^2|}/m^3$, then the dependence of $dP_e/dVdt$ on $\mathscr{F}(x)$ and $\mathscr{G}(x)$ can be neglected. In this case the probability $dP_e/dVdt$ essentially coincides with the analogous quantity calculated for a constant crossed field $F_0^{\mu\nu}$, with the replacement $F_0^{\mu\nu}\rightarrow F^{\mu\nu}(x)$ \cite{Ritus_1985}. For a monochromatic plane wave with angular frequency $\omega_0$ this occurs if $\xi_0\gg 1$. As will be seen in Sec. \ref{TCS_General}, this condition corresponds, e.g., to the formation time of multiphoton Compton scattering ($\sim m/|e|E_0$) being much shorter than the laser period $T_0$. As has been mentioned, the $S$-matrix in Eq. \eqref{S} describes all possible electrodynamical processes among electrons, positrons and photons. The above considerations can be easily adapted for discussing processes involving an initial positron. Whereas, the probability $dP_{\gamma}/dVdt$ of a quantum process in a plane-wave field involving an incoming photon, as, e.g., multiphoton $e^+\text{-}e^-$ pair production, depends on the parameters $\xi_0$ and \begin{equation} \label{kappa_0} \varkappa_0=\frac{|e|\sqrt{-(F_{0,\mu\nu}k^{\nu})^2}}{m^3}=\frac{k_-}{m}\frac{E_0}{F_{\text{cr}}}, \end{equation} where $k^{\mu}=(\omega,\bm{k})$ is the four-momentum of the incoming photon (see \onlinecite{Ritus_1985} and Secs. \ref{VPEs}-\ref{Cascade}). For a photon counterpropagating with respect to the plane wave it is $\varkappa_0=5.9\times 10^{-2}\omega[\text{GeV}]\sqrt{I_0[10^{20}\;\text{W/cm$^2$}]}$. In the case of multiphoton $e^+\text{-}e^-$ pair production, the parameter $\varkappa_0$ can be interpreted as the amplitude of the electric field of the plane wave in units of the critical field $F_{\text{cr}}$ in the center-of-mass system of the created electron and positron \cite{Ritus_1985}. The above remarks on processes occurring in a constant or slowly-varying background field $F^{\mu\nu}(x)$ and involving an incoming electron, also apply to the case of an incoming photon once one replaces $\chi(x)$ with $\varkappa(x)=|e|\sqrt{|(F_{\mu\nu}(x)k^{\nu})^2|}/m^3$. \section{Relativistic atomic dynamics in strong laser fields} \label{RD} When super-intense infrared laser pulses, as those described in Sec. \ref{L_O}, impinge on an atom, the latter is immediately partly or fully ionized \cite{Protopapas_1997,Becker_2002,Keitel_2001}. The ejected electrons experience the typical ``zig-zag'' motion of a free electron in both laser polarization and propagation directions (see Eqs. \eqref{Free_Sol_1_2}-\eqref{Free_Sol_3} and Fig. \ref{El_dyn}) and will not, in general, return to the ionic core. With an enhanced binding force on the remaining electrons, the ionization dynamics becomes increasingly complex and may experience subtle relativistic and correlation effects. When the binding force of the ionic core and that of the applied laser field eventually become comparable, the electrons may in special cases return to and interact with the parent ion (rescattering \cite{Kuchiev_1987,Corkum_1993,Schafer_1993}). This interaction leads, for example, to the ejection of other electrons, to the absorption of energy in a scattering process or to the emission of high-harmonic photons in case of recombination. \subsection{Ionization} \label{RD_I} Previously, atomic or molecular ionization was studied with laser pulses of intensity below $10^{16}$ W/cm$^2$, and the relativistic laser-matter interaction was dominated by the plasma community. The pioneering experiment reported in \onlinecite{Moore_1999} on the ionization behavior of atoms and ions in interaction with a laser with intensity of $3\times 10^{18}$ W/cm$^2$ has thus attracted considerable interest. The laser magnetic field component was shown to alter the direction of ionization characteristically (see Sec. \ref{FED_C}). This is because in the relativistic regime the ionized electron in a laser field acquires a large momentum along the laser propagation direction (see Eq. \eqref{Free_Sol_3}) and photoelectrons are emitted mostly in that direction within a characteristic opening angle $\theta$: $\tan\theta\sim p_{\perp}(\infty)/p_{\parallel}(\infty)\sim 2/\xi_0$ (see the discussion below Eq. \eqref{xi_0}). With highly-charged ions becoming more easily available in a wide range of charges, e.g., via super-strong laser fields or by passing the ion beams through metallic foils, relativistic laser-induced ionization has been further studied \cite{Chowdhury_2001,Dammasch_2001,Yamakawa_2003,Yamakawa_2004,Gubbini_2005,DiChiara_2008,Palaniyappan_2008,DiChiara_2010}. In this situation rescattering is generally suppressed and multiple ionization of atoms and ions takes place mostly via direct ionization, especially including tunneling. On the theoretical side attention was then focused on the relativistic generalization \cite{Popov_1997,Milosevic_2002,Popov_2004,Popov_2006} of the so-called Perelomov-Popov-Terent'ev (PPT) theory or Ammosov-Delone-Krainov (ADK) model \cite{Perelomov_1967,Ammosov_1986}, which describes atomic ionization in the quasistatic tunneling regime. While the common intuitive interpretation of the laser induced tunneling fails in the relativistic regime \cite{Reiss_2008}, a revised picture has been proposed in \onlinecite{Klaiber_2012}. The Strong Field Approximation (SFA) \cite{Keldysh_1965,Faisal_1973,Reiss_1980}, which treats in a universal way both the multiphoton and the tunneling regimes of strong-field ionization, has also been extended to the relativistic regime \cite{Reiss_1990,Reiss_1990_JOSAB}. Both the PPT theory and the SFA assume that the direct ionization process occurs as a single-electron phenomenon and thus neglects atomic structure effects. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig04.eps} \end{center} \caption{(Color) The total Rb$^{10+}$ (black lines) and Rb$^{11+}$ (red lines) ion populations in a gaseous target as a function of the peak intensity of a linearly-polarized laser field with a wavelength of $0.8\;\text{$\mu$m}$ and with a pulse duration of $5\;\text{fs}$. The solid lines display two-electron inelastic tunneling, the dashed lines one-electron inelastic tunneling and the dashed-dotted lines the results via the PPT theory. Adapted from \onlinecite{Zon_2009}.} \label{Zon} \end{figure} When the tunneling process proceeds very fast, multi-electron correlation effects can occur due to the so-called shake-up processes. Thus, the detachment of one electron from the atom or ion via tunneling modifies the self-consistent potential sensed by the remaining electrons and may result, consequently, in the excitation of the atomic core (inelastic tunneling). A strong excitation may also trigger the simultaneous escape of several electrons from the bound state through the potential barrier (collective tunneling). These effects were known to occur also in the nonrelativistic regime \cite{Zon_1999,Zon_2000} but were concealed by competing rescattering effects. In \onlinecite{Zon_2009}, it was shown that, in the relativistic regime, the role of inelastic and collective tunneling can significantly increase and the relativistic PPT rate has been generalized in this respect. In a linearly polarized field, the rate $R_{\text{coll}}^{(N)}$ of inelastic collective tunneling of $N$ equivalent electrons from the outer shell of an ion is described by the following universal formula \cite{Zon_2009}: \begin{equation} \begin{split} R_{\text{coll}}^{(N)}=&\frac{m\sqrt{6}}{\alpha^{3(N-1)}} C^{2N}_{\kappa l}\frac{M!(l+1/2)^N}{2^MN^{M+1}\sqrt{\pi}}\prod _{j=1}^N\frac{(l+m_j)!}{(m_j!)^2(l-m_j)!}\\ &\times \mathcal{I}^2\kappa^{3N-1}\left(\frac{2E_a}{E_0}\right)^{2(\nu-1)N-M+1/2}\text{e}^{-2NE_a/3E_0}, \end{split} \end{equation} where $\alpha=e^2\approx 1/137$ is the fine-structure constant, $m_1,\ldots,m_N$ are the magnetic quantum numbers of the bound electrons, $M=\sum_{j=1}^N m_j$, $l$ is the orbital quantum number of the electrons, $\kappa=\sqrt{2I_p^{(N)}/mN}$, $I_p^{(N)}=\sum_{j=1}^N(I_{p,j}^{(0)}-\Delta_j)$, $I_{p,j}^{(0)}$ is the $j$th ionization potential of a parent ion, $\Delta_j$ is the energy of the core excitation, $E_a=\kappa^3F_{\text{cr}}$ is the atomic field, $Z|e|$ is the charge of the residual ion, $\mathcal{I}$ is the adimensional overlap integral (see \onlinecite{Zon_2009} for its precise definition) and $C_{\kappa l}\approx(2/\nu)^{\nu}/\sqrt{2\pi\nu}$, with $\nu=Z\alpha/\kappa$. According to the calculations in \onlinecite{Zon_2009}, inelastic and collective tunneling effects contribute significantly to the relativistic ionization dynamics at intensities larger than $10^{18}$ W/cm$^2$, thus changing the ionization probability by more than one order of magnitude (see Fig. \ref{Zon}). Spin effects of bound systems in strong laser fields were shown to moderately alter the quantum dynamics and its associated radiation via spin-orbit coupling in highly-charged ions already at an intensity of $\sim 10^{17}\;\text{W/cm$^2$}$ \cite{Hu_1999,Walser_2002}. More recently a nonperturbative relativistic SFA theory has been developed, describing circular dichroism and spin effects in the ionization of helium in an intense circularly polarized laser field \cite{Faisal_2011}. Here, two-photon ionization has been studied in the nonrelativistic intensity range $10^{13}\text{-} 10^{15}$ W/cm$^2$ with a photon energy of 45 eV, yielding small relative spin-induced corrections of the order of $10^{-3}$. A series of experiments has been devoted to the measurement of atomic multi-electron effects in relativistically strong laser fields. In \onlinecite{DiChiara_2008}, the energy distribution of the ejected electrons and the angle-resolved photoelectron spectra for atomic photoionization of argon at $I_0\sim 10^{19}$ W/cm$^2$ have been investigated experimentally. Here, isolation of the single-atom response in the multicharged environment has been achieved by measuring photoelectron yields, energies, and angular distributions as functions of the sample density. Ionization of the entire valence shell along with several inner-shell electrons was shown at $I_0\sim 10^{17}\text{-}10^{19}$ W/cm$^2$. A typical spectrum in the case of linear polarization is displayed in Fig. \ref{walker}. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig05.eps} \end{center} \caption{(Color) (a) Experimental photoelectron spectra for argon at $I_0=1.2\times 10^{19}$ W/cm$^2$ and at an angle of $62^{\circ}$ from the laser propagation direction. Analytical results are shown for all photoelectrons (continuous line) and for the L-shell (dashed line). The angular distributions are at an electron energy of (b) 60 keV, (c) 400 keV, and (d) 770 keV. From \onlinecite{DiChiara_2008}.} \label{walker} \end{figure} An extended plateau-like structure appears in the spectrum due to the electrons originating from the L-shell and the longitudinal component of the focused laser field. A surprising feature is observed in the energy-resolved angular distribution. In contrast to the nonrelativistic case with increasing rescatterings and, thus, angular-distribution widths at high energies, here azimuthally isotropic angular distributions are observed at low energies ($\sim 60$ keV in Fig. \ref{walker}), which become narrower for high-energy photoelectrons. The authors attribute the anomalous broad angular distribution for low-energy electrons to electron-correlation effects. A similar experiment on the energy- and angle-resolved photoionization was later reported for xenon at a laser intensity of $10^{19}$ W/cm$^2$ \cite{DiChiara_2010}. For energies below 0.5 MeV, the yield and the angular distribution were shown not to be described by a one-electron strong-field model, but rather involve most likely multielectron and high-energy atomic excitation processes. A further experiment on relativistic ionization of the methane molecule at $I_0\sim 10^{18}\text{-} 10^{19}$ W/cm$^2$ \cite{Palaniyappan_2008} indicated that molecular mechanisms of ionization play no role, and that C$^{5+}$ ions are produced at these intensities mostly via the cross-shell rescattering atomic ionization mechanism. All these experimental results still await an accurate theoretical description. On the computational side, various numerical methods have been developed to describe the laser-driven relativistic quantum dynamics in highly-charged ions. A Fast-Fourier-Transform split-operator code was implemented in \onlinecite{Mocken_2008} for solving the Dirac equation in 2+1 dimensions by employing adaptive grid and parallel computing algorithms. Another method has been developed in \onlinecite{Selsto_2009} to solve the 3D Dirac equation by expanding the angular part of the wave-function in spherical harmonics. The latter was applied to hydrogenlike ions in intense high-frequency laser pulses with emphasis on investigating the role of negative-energy states. In \onlinecite{Bauke_2011}, the classical relativistic phase-space averaging method, generalized to arbitrary central potentials, and the enhanced time-dependent Dirac and Klein-Gordon numerical treatments are employed to investigate the relativistic ionization of highly-charged hydrogenlike ions in short intense laser pulses. For ionization dynamics beyond the tunneling regime, quantum mechanical and classical methods give similar results, for laser wavelengths from the near-infrared region to the soft x-ray regime. Furthermore a useful procedure has been developed, which employs the over-the-barrier ionization yields for highly-charged ions, to determine the peak laser field strength of short ultrastrong pulses in the range $I_0\sim 10^{18}\text{-} 10^{26}$ W/cm$^2$ \cite{Hetzheim_2009}. In addition, in this article the ionization angle of the ejected electrons is investigated by the full quantum mechanical solution of the Dirac equation and the laser field strength is shown to be also linked to the electron emission angle. The magnetic field-induced tilt in the lobes of the angular distributions of photoelectrons in laser-induced relativistic ionization has also been discussed in \onlinecite{Klaiber_2007}. There are also several new theoretical results for the ionic quantum dynamics in strong high-frequency laser fields, in the so-called stabilization regime, where the ionization rate decreases or remains constant also with increasing laser intensity. An unexpected nondipole effect has been reported in \onlinecite{Foerre_2006} via numerically solving the Schr\"odinger equation for a hydrogenic atom beyond the dipole approximation. For this purpose the Kramers-Henneberger transformation \cite{Kramers_1956,Henneberger_1968} has been employed, i.e., the transformation to the instantaneous rest-frame of a classical free electron in the laser field, and the terms $\sim \xi_0^2$ have been neglected in the Hamiltonian (the value of $\xi_0$ considered was approximately $0.14$). In Fig. \ref{Forre}, the resulting angular distribution of the ejected electrons in the nondipole regime of stabilization displays a third unexpected lobe anti-parallel to the laser propagation direction, together with the two expected lobes along the laser polarization direction. As a classical explanation, a drift along the laser propagation direction was identified for the bound electron wave packet in the nondipole case (see the middle panel of Fig. \ref{Forre}). Inside the laser field the electron has a velocity component along the positive $z$ axis but this velocity tends to zero at the end of the pulse. Thus, the electromagnetic forces alone do not change the electron momentum along the propagation direction at the end of the pulse. The net effect of the Coulomb forces on the electron wave packet is consequently a momentum component along the negative $z$ axis: the electron, which is most probably situated in the upper hemisphere over the pulse, undergoes a momentum kick in the negative $z$ direction each time it passes close to the nucleus. A similar effect has been reported for molecules \cite{Forre_2007}. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig06.eps} \end{center} \caption{(Color online) Dipole (left) and nondipole (right) probability densities of the Kramers-Henneberger wave-function in the $x\text{-}z$ plane for a $x$-polarized, 10-cycle sin-like pulse propagating in the positive $z$ direction (upward), with $E_0=1.5\times 10^{11}\;\text{V/cm}$ and $\omega_0=54\;\text{eV}$. The snapshots are taken at $t=0$, $t=T_0/2$, and $t=1.8\,T_0$ from top to bottom. The length of the horizontal line corresponds to about $50\,a_B\approx 2.7\;\text{nm}$, with $a_B=\lambda_C/\alpha\approx 5.3\times 10^{-9}\;\text{cm}$ being the Bohr radius . Note that the scale is logarithmic with four contours per decade. From \onlinecite{Foerre_2006}.} \label{Forre} \end{figure} Radiative recombination, being the time-reversed process of photoionization, of a relativistic electron with a highly-charged ion in the presence of a very intense laser field has been considered in \onlinecite{Mueller_2009c}. It was shown that the strong coupling of the electron to the laser field may lead to a very broad energy spectrum of emitted recombination photons, with pronounced side wings, and to characteristic modifications of the photon angular distribution. Specific features of nondipole quantum dynamics in strong and ultrashort laser pulses have also been investigated employing the so-called Magnus approximation \cite{Dimitrovski_2009}. The dominant nondipole effect is found to be a shift of the entire wave-function towards the propagation direction, inducing a substantial population transfer into states with similar geometry. The recent experiment reported in \onlinecite{Smeenk_2011} addresses the question of how the photon momenta are shared between the electron and ion during laser-induced multiphoton ionization. Theoretically, this problem requires a nondipole treatment, even in the nonrelativistic case, to take into account explicitly the laser photon momentum. Energy-conservation of $\ell$-photon ionization here means that $\ell\omega_0=I_p+U_p+K$, where $I_p$ is the ionization energy of the atom, $U_p=e^2E_0^2/4m\omega_0^2$ is the ponderomotive energy and $K$ is the electron's kinetic energy. The experimental results in \onlinecite{Smeenk_2011}, obtained using laser fields with wavelength of $0.8\;\text{$\mu$m}$ and $1.4\;\text{$\mu$m}$ in the intensity range of $10^{14}\text{-}10^{15}\;\text{W/cm$^2$}$ has show that the fraction of the momentum, corresponding to the number of observed photons needed to overcome the ionization energy $I_p$, is transferred to the created ion rather than to the photoelectron. The electron carries only the momentum corresponding to the kinetic energy $K$, while the ponderomotive energy and the corresponding portion of the momentum are transferred back to the laser field. This experiment shows that the tunneling concept for the ionization dynamics is only an approximation. In fact, the quasistatic tunneling provides no mechanism to transfer linear momentum to the ion, a conclusion that agrees with recent concerns in \onlinecite{Reiss_2008}. \subsection{Recollisions and high-order harmonic generation} \label{HHG} Tunneling in the nonrelativistic regime is generally followed by recollisions with the parent ion along with various subsequent effects \cite{Kuchiev_1987,Corkum_1993,Schafer_1993}. A characteristic feature of strong-field processes in the relativistic regime is the suppression of recollisions due to the magnetically induced relativistic drift of the ionized electron in the laser propagation direction (see Sec. \ref{FED_C}). Although relativistic effects become significant when the parameter $\xi_0$ exceeds unity, signatures of the drift in the laser propagation direction can be observed already in the weakly relativistic regime $\xi_0\lesssim 1$. The drift will have a significant impact on the electron's rescattering probability if, at the instant of recollision, the drift distance $d_{\parallel}$ in the laser propagation direction is larger than the electron's wave packet size $a_{wp,\parallel}$ in that direction \cite{Walker_2006}. The drift distance is approximately given by $d_{\parallel}\sim \lambda_0\xi_0^2/2$ (see, e.g., Eq. \eqref{Free_Sol_3}). Instead, the wave packet size $a_{wp,\parallel}$ can be estimated from $a_{wp,\parallel}\sim v_{\parallel} \Delta t$, where $v_{\parallel}$ is a typical electron velocity along the laser propagation direction and $\Delta t$ is the excursion time of the electron in the continuum. The velocity $v_{\parallel}$ can be related to the tunneling time $\tau_{\text{tun}}$ via the time-energy uncertainty: $mv_{\parallel}^2/2\sim 1/\tau_{\text{tun}}$. In turn, one can estimate the tunneling time $\tau_{\text{tun}}$ as $\tau_{\text{tun}}\sim l_{\text{tun}}/v_b$, where $l_{\text{tun}}\sim I_p/|e|E_0$ is the tunneling length and $v_b\sim \sqrt{2I_p/m}$ the velocity of the bound electron. In the above estimate, it was assumed that the work carried out by the laser field along the tunneling length equals $I_p$. Thus, at the rescattering moment $\Delta t\sim T_0$, the wave packet size $a_{wp,\parallel}$ is of the order of $\lambda_0\sqrt{|e|E_0/\sqrt{m^3I_p}}$ and the role of the drift can be characterized by means of the parameter $r=(d_{\parallel}/a_{wp,\parallel})^2$ as estimated by \begin{equation} r \sim \xi_0^3\frac{\sqrt{2mI_p}}{16\omega_0}. \label{drift} \end{equation} The condition $r\gtrsim 1$ determines the parameter region over which the signature of the drift becomes conspicuous. As an alternative view on the relativistic drift, the ionized electron here misses the ionic core when it is ionized with zero momentum. Nevertheless, the recollision will occur if the electron is ionized with an appropriate initial momentum $p_d$ ($\sim m\xi_0^2/4$, see Eq. \eqref{Free_Sol_3}), opposite to the laser propagation direction. The probability $P_i(p_d)$ of this process is exponentially damped, though, due to the nonzero momentum $p_d$ (see, e.g., \onlinecite{Salamin_2006}): \begin{equation} P_i(p_d)\sim \exp \left[ -\frac{2}{3}\frac{(2mI_p)^{3/2}}{m|e|E_0}\left(1+\frac{p_d^2}{4mI_p}\right)\right]. \label{TI_pd} \end{equation} The drift term in the exponent proportional to $p_d^2$ will be important if $\sqrt{2mI_p}p_d^2/m|e|E_0\gtrsim 1$, which is equivalent to the condition $r\gtrsim 3$. At near-infrared wavelengths ($\omega_0\approx 1\;\text{eV}$) and for the ionization energy $I_p=13.6\;\text{eV}$ of atomic hydrogen, it becomes relevant at laser intensities $I_0$ approximately above $3\times10^{16}$ W/cm$^2$. Then, HHG and other recollision phenomena are suppressed. The attainability of relativistic recollisions would, however, be very attractive for ultrahigh HHG \cite{Kohler_2012} as well as for the realization of laser controlled high-energy \cite{Hatsagortsyan_2006} and nuclear processes \cite{Milosevic_2004,Chelkowski_2004}. Various methods for counteracting the relativistic drift have been proposed, such as by utilizing highly-charged ions \cite{Hu_2001,Hu_2002} which move relativistically against the laser propagation direction \cite{Mocken_2004,Chirila_2004}, by employing Positronium (Ps) atoms \cite{Henrich_2004}, or through preparing antisymmetric atomic \cite{Fischer_2007} and molecular \cite{Fischer_2006} orbitals. Here the impact of the drift of the ionized electron is reduced by the increase of the laser frequency in the system's center of mass, an equally strong drift via two constituents with equal mass or via appropriate initial momenta from antisymmetric orbitals, respectively. On the other hand, the laser field can also be modified to suppress the relativistic drift by employing tightly focused laser beams \cite{Lin_06}, two counter-propagating laser beams with linear polarization \cite{Keitel_1993,Kylstra_2000,Taranukhin_2000,Taranukhin_2001,Taranukhin_2002} or equal-handed circular polarization \cite{Milosevic_2004}. In the first two cases, the longitudinal component in the tightly focused laser beam may counteract the drift, or the Lorentz force may be eliminated in a small area near the antinodes of the resulting standing wave, respectively. In the third case involving circularly polarized light, the relativistic drift is eliminated because the electron velocity is oriented in the same direction as the magnetic field. This setup is well suited for imaging attosecond dynamics of nuclear processes but not for HHG because of the phase-matching problem \cite{Liu_2009}. In the weakly relativistic regime the Lorentz force may also be compensated by a second weak laser beam polarized along the direction of propagation of the strong beam \cite{Joachain_2002}. Furthermore, the relativistic drift can be significantly reduced by means of special tailoring of the driving laser pulse, which strongly reduces the time when the electron's motion is relativistic with respect to a sinusoidal laser pulse \cite{Klaiber_2006,Klaiber_2007}. Two consecutive laser pulses \cite{Verschl_2007a} or a single laser field assisted by a strong magnetic field can also be used to reverse the drift \cite{Verschl_2007b}. In addition two strong Attosecond Pulse Trains (APTs) \cite{Hatsagortsyan_2008} or an infrared laser pulse assisted by an APT \cite{Klaiber_2008} have been employed to enhance relativistic recollisions. In fact, due to the presence of the APT the ionization can be accomplished by one XUV photon absorption and the relatively large energy $\omega_X$ of the XUV-photon with $\omega_X=I_p+p_d^2/2m$ can compensate the subsequent momentum drift $p_d\sim m\xi_0^2/4$ in the infrared laser field. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig07.eps} \caption{(Color) The HHG setup with two counterpropagating APTs. After ionization by the laser pulse 1, the ejected electron is driven in the same pulse (light blue), propagates freely after the pulse 1 has left (gray dashed) and is driven back to the ion by the laser pulse 2 (dark blue). From \onlinecite{Kohler_2011}.} \label{setup} \end{center} \end{figure} The main motivation for the realization of relativistic recollisions is the extension of HHG towards the hard x-ray regime with obvious benefits for time-resolved high-resolution imaging. In the past couple of decades, nonrelativistic atomic HHG \cite{Corkum_1993,Lewenstein_1994} has been developed as a reliable source of coherent XUV radiation and attosecond pulses \cite{Agostini_2004} opening the door for attosecond time-resolved spectroscopy \cite{Krausz_2009}. Nonrelativistic HHG in an atomic gas medium allows already to generate coherent x-ray photons up to keV energies \cite{Sansone_2006} and to produce XUV pulses shorter than 100 as \cite{Goulielmakis_2008}. The most favorable conversion efficiency for nonrelativistic keV harmonics is anticipated with mid-infrared driving laser fields \cite{Popmintchev_2009,Chen_2010}. However, progress in this field has slowed down, especially because of the inhibition, alluded to above, of recollisions due to optical driving-field intensity above $3\times10^{16}$ W/cm$^2$. This indicates the limit on the cut-off frequency $\omega_{c}$ of nonrelativistic HHG to $\omega_{c}\approx 3.17 U_p\sim 10$ keV. Another factor hindering HHG at high intensities is the less favorable phase-matching. In strong laser fields, outer-shell electrons are rapidly ionized and produce a large free electron background causing a phase mismatch between the driving laser wave and the emitted x-rays. The feasibility of phase-matched relativistic HHG in a macroscopic ensemble was first investigated in \onlinecite{Kohler_2011}. Here, the driving fields are two counterpropagating APTs consisting of 100 as pulses with a peak intensity of the order of $10^{19}$ W/cm$^2$ (see Fig. \ref{setup}). The electron is driven to the continuum by the laser pulse 1 in Fig. \ref{setup}, followed by the usual relativistic drift. Thereafter, the laser pulse 2 overtakes the electron, reverses the drift and imposes the rescattering, yielding a much higher HHG signal than for a conventional laser field at the same cut-off energy. Here phase-matching can be fulfilled due to an additional intrinsic phase specific to this setup, depending on the time delay between the pulses and on the pulse intensity. The latter, being unique for this laser setup, mainly affects the electron excursion time and varies along the propagation direction. The phase-matching is achieved by modifying the laser intensity along the propagation direction and by balancing the phase slip due to dispersion with the indicated intrinsic phase. Note, however, that HHG in the relativistic regime has been observed experimentally rather efficiently in laser plasma interactions \cite{Dromey_2006}. \section{Multiphoton Thomson and Compton scattering} \label{TCS} In this section we discuss one of the most fundamental processes in QED in a strong laser field: the emission of radiation by an accelerated electron. After reporting on recent theoretical investigations on this process, we discuss its possible applications for producing high-energy photon beams. \subsection{Fundamental considerations} \label{TCS_General} When an electron is wiggled by an intense laser wave, it emits electromagnetic radiation. This process occurs with absorption of energy and momentum by the electron from the laser field and it is named as multiphoton Thomson scattering or multiphoton Compton scattering, depending on whether quantum effects, like photon recoil, are negligible or not. Multiphoton Thomson and Compton scattering in a strong laser field have been studied theoretically since a long time (see \onlinecite{Sarachik_1970,Salamin_1998,Sengupta_1949} for multiphoton Thomson scattering and \onlinecite{Goldman_1964,Nikishov_1964_a,Brown_1964} for multiphoton Compton scattering). The classical calculation of the emitted spectrum is based on the analytical solution in Eqs. \eqref{Free_Sol_0}-\eqref{Free_Sol_3} of the Lorentz equation in a plane wave and the substitution of the corresponding electron trajectory in the Li\'{e}nard-Wiechert fields \cite{Jackson_b_1975,Landau_b_2_1975}. Whereas, as we have discussed in Sec. \ref{FED_Q}, the quantum calculation of the amplitude of the process is performed in the Furry picture of QED. As a result, the total emission probability depends only on the two Lorentz- and gauge-invariant parameters $\xi_0$ (see Eq. \eqref{xi_0}) and $\chi_0$ (see Eq. \eqref{chi_0}). The parameter $\xi_0$ has already been discussed in Sec. \ref{FED_C}. In the contest of multiphoton Compton scattering this parameter controls in particular the effective order $\ell_{\text{eff}}$ of the emitted harmonics, which, for an ultrarelativistic electron, can be estimated in the following way. In order to effectively emit a frequency $\omega'$, the formation length $l_f$ of the process must not exceed the coherence length $l_{\text{coh}}$, because, otherwise, interference effects would hinder the emission. Since an electron with instantaneous velocity $\bm{\beta}$ and energy $\varepsilon=m\gamma=m/\sqrt{1-\beta^2}\gg m$ mainly emits along the direction of $\bm{\beta}$, within a cone with apex angle $\vartheta\sim 1/\gamma\ll 1$, the formation length $l_f$ can be estimated from $l_f\sim \varrho/\gamma$, with $\varrho$ being the instantaneous radius of curvature of the electron trajectory \cite{Jackson_b_1975}. On the other hand, $l_{\text{coh}}=\pi/\omega'(1-\beta \cos \vartheta)\sim \gamma^2/\omega'$ \cite{Jackson_b_1975,Baier_b_1998}. By requiring that $l_f\lesssim l_{\text{coh}}$, we obtain the following estimate for the largest-emitted frequency (cut-off frequency) $\omega'_c$: $\omega'_c\sim \gamma^3/\varrho$. Now, in the average rest-frame of the electron, i.e., in the reference frame where the average electron velocity along the propagation direction of the laser vanishes (see Eq. \eqref{Free_Sol_3}), it is $\gamma^{\star}\sim \xi_0$ (corresponding to the energy $\varepsilon^{\star}\sim m\xi_0$) and $l_f^{\star}\sim \lambda_0^{\star}/\xi_0$, where the upper index $^{\star}$ indicates the variable in this frame. Consequently, $\omega_c^{\prime\,\star} \sim \xi_0^3 \omega^{\star}_0$ and the effective order of the emitted harmonics is $\ell_{\text{eff}}\sim \xi_0^3$ (note that $\ell_{\text{eff}}$ is a Lorentz scalar). As the order of the emitted harmonics corresponds quantum mechanically to the number of laser photons absorbed by the electron during the emission process, the parameter $\xi_0$ is also said to determine the ``multiphoton'' character of the process. On the other hand, the nonlinear quantum parameter $\chi_0$ (see Eq. \eqref{chi_0}) in the contest of multiphoton Compton scattering controls the importance of quantum effects as the recoil of the emitted photon. In fact, we can estimate classically the importance of the emitted photon recoil from the ratio $\omega'_c/\varepsilon$ and our considerations above exactly indicate that $\omega'_c/\varepsilon\sim \xi_0^2 \omega^{\star}_0/m\sim \chi_0$. Thus, multiphoton Thomson scattering is characterized by the condition $\chi_0\ll 1$, while multiphoton Compton scattering by $\chi_0\gtrsim 1$. This result can also be obtained in the case of a monochromatic laser wave starting from the energy-momentum conservation relation \begin{equation} \label{Cons_law} q_0^{\mu}+\ell k_0^{\mu}=q^{\prime\mu}+k^{\prime\mu} \end{equation} in the case in which $\ell$ laser photons are absorbed in the process \cite{Ritus_1985}. Here $q_0^{\mu}$ and $q^{\prime\mu}$ are the quasimomenta of the initial and final electron (see Eq. \eqref{q}) and $k^{\prime\mu}=\omega'n^{\prime\mu}$ is the four-momentum of the produced photon ($n^{\prime 2}=0$). From this expression it is easy to obtain the energy $\omega'$ of the emitted photon as \begin{equation} \label{omega_p} \omega'=\frac{\ell\omega_0p_{0,-}}{(n'p_0)+\left(\ell\omega_0+\frac{m^2\xi_0^2}{4p_{0,-}}\right)n'_-}. \end{equation} By reminding that $\ell_{\text{eff}}\sim \xi_0^3$ and by estimating the typical emission angle of the photon \cite{Mackenroth_2011}, it is possible to show that $\omega'\sim \chi_0\varepsilon_0$ at $\xi_0\gg 1$. As it has also been throughroughly investigated analytically and numerically in \onlinecite{Boca_2011,Seipt_2011}, multiphoton Compton and Thomson spectra coincide in the limit $\chi_0\to 0$, although in \onlinecite{Seipt_2011} differences have been observed numerically in the detailed structure of the classical and quantum spectra also for $\chi_0\ll 1$. The most important difference between classical and quantum spectra is certainly the presence of a sharp cut-off in the latter as an effect of the photon recoil: the energy of the photon emitted in a plane wave is limited by the initial energy of the electron\footnote{In the case of a plane-wave background field, this limitation rather concerns the quantity $k'_-$ of the emitted photon, as $k'_-=p_{0,-}-p'_-<p_{0,-}$. However, for an ultrarelativistic electron with $p_{0,-}\gg m\xi_0$ and initially counterpropagating with respect to the laser field, it is $p_{0,-}\approx 2\varepsilon_0$, $k'_-\approx 2\omega'$ and $p'_-\approx 2\varepsilon'$ \cite{Baier_b_1998}.}. This does not occur classically, as there the frequency of the emitted radiation does not have the physical meaning of photon energy. The dependence of the energy cut-off on the laser intensity has been recently recognized as a possible experimental signature of multiphoton Compton scattering \cite{Harvey_2009}. First calculations on multiphoton Thomson and Compton scattering have mainly focused on the easiest case of a monochromatic background plane wave, either with circular or linear polarization. The main results of these investigations, like the dependence of the emitted frequencies on the laser intensities have been recently reviewed in \onlinecite{Ehlotzky_2009}. The complete description of the multiphoton Compton scattering process with respect to the polarization properties of the incoming and the outgoing electrons, and of the emitted photon in a monochromatic laser wave has been presented in \onlinecite{Ivanov_2004}. Recently significant attention has been devoted to the investigation of multiphoton Thomson and Compton scattering in the presence of short and even ultrashort plane-wave pulses (we recall that such pulses have still an infinite extension in the directions perpendicular to the propagation direction). In \onlinecite{Boca_2009} multiphoton Compton scattering has been considered in the presence of a pulsed plane wave. The angular-resolved spectra are practically insensitive to the precise form of the laser pulse for $\omega_0\tau_0\ge 20$, with $\tau_0$ being the pulse duration. The main differences with respect to the monochromatic case are: 1) a broadening of the lines corresponding to the emitted frequencies; 2) the appearance of sub-peaks, which are due to interference effects in the emission at the beginning and at the end of the laser pulse. On the one hand, the continuous nature of the emission spectrum in a finite pulse in contrast to the discrete one in the monochromatic case has a clear mathematical counterpart. In both cases, in fact, the total transverse momenta $\bm{P}_{\perp}$ with respect to the laser propagation direction and the total quantity $P_-$ are conserved in the emission process (see Sec. \ref{FED}). However, in the monochromatic case the following additional conservation law holds (for a linearly polarized plane wave, see Eq. \eqref{Cons_law}) \begin{equation} \varepsilon_0+p_{0,\parallel}+\frac{m^2\xi_0^2}{2p_{0,-}}+2\ell\omega_0=\omega'+k'_{\parallel}+\varepsilon'+p'_{\parallel}+\frac{m^2\xi_0^2}{2p'_-}, \end{equation} so that the resulting four-dimensional energy-momentum conservation law allows only for the emission of the discrete frequencies in Eq. \eqref{omega_p}. On the other hand, the appearance of sub-peaks has been in particular investigated in \onlinecite{Heinzl_2010}, where it has been found that the number $N_{\text{s-p}}$ of sub-peaks within the first harmonic scales linearly with the pulse duration $\tau_0$ and with $\xi_0^2$: $N_{\text{s-p}}=0.24\,\xi_0^2\tau_0[\text{fs}]$. In this paper the effects of spatial focusing of the driving laser pulse are also discussed. The authors investigate in particular the dependence of the deflection angle $\alpha_{\text{out}}$ undergone by the electron after colliding head-on with a Gaussian focused beam as a function of the impact parameter $b$ (see Fig. \ref{Heinzl}). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig08.eps} \caption{(Color online) Deflection $\alpha_{\text{out}}$ of an electron initially counterpropagating with respect to a laser field with an energy of 3 J and a pulse duration of $20$ fs, as a function of the impact parameter $b$ for different laser waist radii $w_0$. From \onlinecite{Heinzl_2010}. } \label{Heinzl} \end{center} \end{figure} By further decreasing the laser pulse duration, it has been argued that effects of the relative phase between the pulse profile and the carrier wave (the so-called Carrier Envelope Phase (CEP)) should become visible in multiphoton Thomson and Compton scattering. In \onlinecite{Boca_2009} the case of ultrashort pulses with $\omega_0\tau_0\gtrsim 4$ has been discussed and effects of the CEP on the harmonic yield in specific frequency ranges have been observed. In \onlinecite{Mackenroth_2010} the dependence of the angular distribution of the emitted radiation in multiphoton Thomson and Compton scattering on the CEP of few-cycles pulses has been exploited to propose a scheme to measure the CEP of ultrarelativistic laser pulses (intensities larger than $10^{20}\;\text{W/cm$^2$}$). The method is essentially based on the high directionality of the photon emission by an ultrarelativistic electron, because the trajectory of the electron, in turn, also depends on the laser's CEP. Accuracies in the measurement of the CEP of the order of a few degrees are theoretically envisaged. Multiphoton Compton scattering in one-cycle laser pulses has been considered in \onlinecite{Mackenroth_2011} and a substantial broadening of the emission lines with respect to the monochromatic case has been observed. The high-directionality of radiation emitted via multiphoton Thomson scattering has also been employed as a diagnostic tool in \onlinecite{Har-Shemesh_2012}, where a new rather precise method has been proposed to measure the peak intensity of strong laser fields (intensities between $10^{20}\;\text{W/cm$^2$}$ and $10^{23}\;\text{W/cm$^2$}$) from the angular aperture of the photon spectrum. The study of multiphoton Thomson and Compton scattering in short laser pulses has also stimulated the investigation of scaling laws for the photon spectral density \cite{Heinzl_2010,Seipt_2011,Seipt_2011b,Boca_2011b}. For example, in \onlinecite{Heinzl_2010} a scaling law has been found for backscattered radiation in the case of head-on laser-electron collisions, which simplifies the averaging over the electron-beam phase space. A more general scaling law has been determined in \onlinecite{Seipt_2011b}, which relaxes the previous assumptions on head-on collision and on backscattered radiation employed in \onlinecite{Heinzl_2010}. Moreover, in \onlinecite{Seipt_2011} a simple relation is determined between the classical and the quantum spectral densities. Finally, in \onlinecite{Boca_2011b} it is found that in the ultrarelativistic case $\gamma_0\gg 1$ the angular distribution of the emitted radiation, integrated with respect to the photon energy, only depends on the ratio $\xi_0/\gamma_0$ and not on the independent values of $\xi_0$ and $\gamma_0$ (see also \onlinecite{Mackenroth_2010}). In the above-mentioned publications the spectral properties of the emitted radiation in the classical and quantum regimes have been considered. In \onlinecite{Zhang_2008,Kim_2009}, instead, the temporal properties of the emitted radiation in multiphoton Thomson scattering have been investigated. In both papers the feasibility of generating single attosecond pulses is discussed. Photoemission by a single-electron wave packet via Thomson scattering in a strong laser field has been discussed in \onlinecite{Peatross_2008}. It was shown that the partial emissions from the individual electron momentum components do not interfere when the driving field is a plane wave. In other words, the size of the electron wave packet, even when it spreads to the scale of the wavelength of the driving field, does not affect the Thomson emission. Finally, we shortly mention that multiphoton effects in Thomson and Compton scattering have been measured in various laboratories. The second-harmonic radiation was first observed in the collision of a 1 keV electron beam with a Q-switched Nd:YAG laser, although the laser intensity was such that $\xi_0\approx 0.01$ \cite{Englert_1983}, and then in the interaction of a mode-locked Nd:YAG laser ($\xi_0=2$) with plasma electrons \cite{Chen_1998}. Multiphoton Thomson scattering of laser radiation in the x-ray domain has been reported in \onlinecite{Babzien_2006} (see \onlinecite{Pogorelsky_2000} for a similar proof-of-principle experiment). Single-shot measurements of the angular distribution of the second harmonic (photon energy $6.5\;\text{keV}$) at various laser polarizations have been carried out by employing a 60 MeV electron beam and a subterawatt CO$_2$ laser beam with $\xi_0=0.35$. In the prominent SLAC experiment \cite{Bula_1996} multiphoton Compton emission was detected for the first time. In this experiment an ultrarelativistic electron beam with energy of about $46.6\;\text{GeV}$ collided with a terawatt Nd:glass laser with an intensity of $10^{18}\;\text{W/cm$^2$}$ ($\xi_0\approx 0.8$ and $\chi_0\approx 0.3$) and four-photon Compton scattering has been observed indirectly via a nonlinear energy shift in the spectrum of the outcoming electrons. \subsection{Thomson- and Compton-based sources of high-energy photon beams} \label{TC_Sources} The single-particle theoretical analysis presented above indicates that high-energy photons can be emitted via multiphoton Thomson and Compton scattering of an ultrarelativistic electron. For example, an electron with initial energy $\varepsilon_0\gg m$ colliding head-on with an optical laser field ($\omega_0\approx 1\;\text{eV}$) of moderate intensity ($\xi_0\lesssim 1$) is barely deflected by the laser field ($\varrho\sim \lambda_0\gamma_0/\xi_0\gg \lambda_0$) and potentially emits photons with energies $\omega[\text{keV}]\lesssim 3.8\times 10^{-3}\,\varepsilon_0^2[\text{MeV}]$. This feature has boosted the idea of so-called Thomson- and Compton-based sources of high-energy photons as a valid alternative to conventional synchrotron sources, the main advantages of the former being the compactness, the wide tunability, the shortness of the photon beams in the femtosecond scale and the potential for high brightness. Unlike the experiments on multiphoton Thomson and Compton scattering where laser systems with $\xi_0\gtrsim 1$ are generally employed, Thomson- and Compton-based photon sources preferably require lasers with $\xi_0\lesssim 1$, such that multiphoton effects are suppressed and shorter bandwidths of the photon beam are achieved. On the other hand, the electron beam quality is crucial for Thomson- and Compton-based radiation sources. In particular, the brightness of the photon beam scales inversely quadratically with the electron beam emittance, and linearly with the electron bunch current density. Proof-of-principle experiments have demonstrated Thomson- and Compton-based photon sources by crossing a high-energy laser pulse with a picosecond relativistic electron beam from a conventional linear electron accelerator \cite{Ting_1995,Ting_1996,Schoenlein_1996,Leemans_1996,Pogorelsky_2000,Chouffani_2002,Sakai_2003}. We also mention the benchmark experiment carried out at LLNL, where photons with an energy of $78\;\text{keV}$ have been produced with a total flux of $1.3\times 10^6\;\text{photons/shot}$, by colliding an electron beam with an energy of $57\;\text{MeV}$ with a Ti:Sa laser beam with an intensity of about $10^{18}\;\text{W/cm$^2$}$ ($\xi_0\approx 0.5$) \cite{Gibson_2004}. Another achievement in the development of Thomson- and Compton-based photon sources has been the experimental realization of a compact all-optical setup, where the electrons are accelerated by an intense laser. In the first experiment with an all-optical setup \cite{Schwoerer_2006}, x-ray photons in the range of 0.4 keV to 2 keV have been generated. In this experiment the electron beam was produced by a high-intensity Ti:Sa laser beam ($I_0\approx 2\times 10^{19}\;\text{W/cm$^2$}$) focused into a pulsed helium gas jet. The characteristic feature of the all-optical setup is that the electron bunches and, consequently, the generated x-ray photon beams have an ultrashort duration ($\sim 100\;\text{fs}$) and a linear size of the order of $10\;\mu$m. Another advantage is that the electrons can be precisely synchronized with the driving laser field. In order to further improve the all-optical setup, design parameters for a proof-of-concept experiment have been analyzed in \onlinecite{Hartemann_2007}. For the calculation of the Compton scattering parameters, a 3D Compton scattering code has been used, which was extensively tested for Compton scattering experiments performed at LLNL \cite{Hartemann_2005,Brown_2004,Rosenzweig_2004,Hartemann_2004} (see \onlinecite{Sun_2011} for an alternative numerical simulation scheme). It is shown that x-ray fluxes exceeding $10^{21}$ s$^{-1}$ and a peak brightness larger than $10^{19}$ photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth) can be achieved at photon energies of about 0.5 MeV. A few years later the Compton-based photon source MonoEnergetic Gamma-ray (MEGa-ray) has been designed at LLNL \cite{Gibson_2010}. Production of gamma-rays ranging from 75 keV to 0.9 MeV has been demonstrated with a peak spectral brightness of $1.5\times 10^{15}$ photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth) and with a flux of $1.6\times 10^5$ photons/shot. An experimental setup for high-flux gamma-ray generation has been constructed in the Saga Light-Source facility in Tosu (Japan), by colliding a 1.4 GeV electron beam with a CO$_2$ laser (wavelength $10.6\;\text{$\mu$m}$) \cite{Kaneyasu_2011}. A flux of about $3.2\times 10^7\;\text{photons/s}$ gamma photons with energy larger than 0.5 MeV has been obtained. In the basic setups of Thomson- and Compton-based photon sources the electrons experience the intense laser field for a time-interval much shorter than that needed to cross the whole laser beam, the former being of the order of the laser's Rayleigh length divided by the speed of light. Thus, the quest for a more intense laser pulse at a given power in order to increase the photon yield implies a tighter focusing and therefore a shorter effective interaction time, which in turn causes a broadening of the photon spectrum. In \onlinecite{Debus_2010} the Traveling-Wave Thomson Scattering (TWTS) setup is proposed, which allows the electrons to stay in the focal region of the laser beam during the whole crossing time (see Fig. \ref{Debus1}). This is achieved by employing cylindrical optics to focus the laser field only along one direction (red lines in Fig. \ref{Debus1}) and, depending on the angle between the initial electron velocity and the laser wave-vector, by tilting the laser pulse front. As a result, an interaction length $\sim 1\;\text{cm}\text{-}1\;\text{m}$ can be achieved and, correspondingly very large photon fluxes (up to $5\times 10^{10}$ photons/shot at 20 keV). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig09.eps} \end{center} \caption{(Color) Schematic setup of TWTS with the red lines indicating the laser focal lines. In the notation of \onlinecite{Debus_2010} $\phi$ is the angle between the initial electrons' velocity and the laser's wave vector and $\alpha$ is the angle between the laser pulse front and the laser propagation direction. Adapted from \onlinecite{Debus_2010}.} \label{Debus1} \end{figure} An alternative way of reaching longer effective laser-electron interaction times has been proposed in \onlinecite{Karagodsky_2010}, where a planar Bragg structure is employed to guide the laser pulse and realize Thomson/Compton scattering in a waveguide. In this way, the yield of x-rays can be enhanced by about two orders of magnitude with respect to the conventional free-space Gaussian-beam configuration at given electron beam and injected laser power in both configurations. However, there are two constraints specific to this setup. On the one hand, the electron beam has to have a small angular spread in order to be injected into the planar Bragg structure without causing wall damage. On the other hand, the laser field strength has to be such that $\xi_0 \lesssim 8\times 10^{-4}$ to avoid surface damage. Finally, in \onlinecite{Hartemann_2008} a setup has been proposed to obtain bright GeV gamma-rays via Compton scattering of electrons by a thermonuclear plasma. In fact, a thermonuclear deuterium-tritium plasma produces intense blackbody radiation with a temperature $\sim 20$ keV and a photon density $\sim 10^{26}$ cm$^{-3}$ \cite{Tabak_1994}. When a thermal photon with energy $\omega \sim 1$ keV counterpropagates with respect to a GeV electron ($\gamma_0\sim 10^3$), a Doppler-shifted high-energy photon $\omega' \sim \gamma_0^2 \omega\sim 1$ GeV can be emitted on axis, i.e., in the same direction of the incoming electron. Since $\omega'$ has to be smaller than the initial electron energy $\varepsilon_0$, a kinematical photon pileup is induced in the emitted photon spectrum at $\varepsilon_0$ \cite{Zeldovich_1969} (see Fig. \ref{Hartemann_pileup}). This results in a quasimonochromatic GeV gamma-ray beam with a peak brightness $\gtrsim 10^{30}$ photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth), comparable with that of the FLASH (see Sec. \ref{x_ray}). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig10.eps} \end{center} \caption{(Color online) Normalized on-axis brightness for different values of the rapidity $\rho=\cosh^{-1}\gamma_0$ at a plasma temperature of 20 keV. See \onlinecite{Hartemann_2008} for the meaning of the blue circles at $\rho=6$. Adapted from \onlinecite{Hartemann_2008}.} \label{Hartemann_pileup} \end{figure} The unique features of Thomson- and Compton-based photon sources render them a powerful experimental device. For example, they can be employed for medical radio-isotope production and photo-fission, and for studying nuclear resonance fluorescence for \textit{in situ} isotope detection \cite{Albert_2010}. In this respect, such photon sources will represent the main experimental tool for nuclear-physics investigations at the Romanian pillar of ELI (see Fig. \ref{Megalaser}). Other possible applications, at photon energies beyond the MeV threshold, include the production of positron beams \cite{Omori_2003,Hugenschmidt_2012} as well as the investigations of high-energy processes occurring in gamma-gamma and gamma-lepton collisions \cite{Telnov_1990}. \section{Radiation reaction} \label{RR} The issue of ``radiation reaction'' (RR) is one of the oldest and most fundamental problems in electrodynamics. Classically it corresponds to the determination of the equation of motion of a charged particle, an electron for definiteness, in a given electromagnetic field $F^{\mu\nu}(x)$. In fact, the Lorentz equation $mdu^{\mu}/ds=eF^{\mu\nu}u_{\nu}$ (see Sec. \ref{FED_C}) does not take into account that the electron, while being accelerated, emits electromagnetic radiation and loses energy and momentum in this way. The first attempt of taking into account the reaction of the radiation emitted by the electron on the motion of the electron itself (from here comes the expression ``radiation reaction''), was accomplished by H. A. Lorentz in the nonrelativistic regime \cite{Lorentz_b_1909}. Starting from the known Larmor formula $\mathcal{P}_L=(2/3)e^2\bm{a}^2$ for the power emitted by an electron with instantaneous acceleration $\bm{a}$, Lorentz argued that this energy-loss corresponds to a ``damping'' force $\bm{F}_R=(2/3)e^2d\bm{a}/dt$ acting on the electron. The expression of the damping force was generalized to the relativistic case by M. Abraham in the form \cite{Abraham_b_1905} \begin{equation} \label{F_R} F^{\mu}_R=\frac{2}{3}e^2\left(\frac{d^2u^{\mu}}{ds^2}+\frac{du^{\nu}}{ds}\frac{du_{\nu}}{ds}u^{\mu}\right). \end{equation} In order to solve the problem of a radiating electron self-consistently, P. A. M. Dirac suggested in \onlinecite{Dirac_1938} to start from the coupled system of Maxwell and Lorentz equations \begin{equation} \label{M_L} \left\{ \begin{aligned} &\partial_{\mu}F^{\mu\nu}_T=4\pi j^{\nu}\\ &\partial_{\lambda}F_{T,\mu\nu}+\partial_{\mu}F_{T,\nu\lambda}+\partial_{\nu}F_{T,\lambda\mu}=0\\ &m_0\frac{du^{\mu}}{ds}=eF_T^{\mu\nu}u_{\nu}, \end{aligned} \right. \end{equation} where $F_T^{\mu\nu}(x)=F^{\mu\nu}(x)+F_S^{\mu\nu}(x)$, with $F_S^{\mu\nu}(x)$ being the ``self'' electromagnetic field generated by the electron four-current $j^{\mu}(x)=e\int ds\delta(x-x(s))u^{\mu}$ and where the meaning of the symbol $m_0$ for the electron mass will be clarified below. In order to write an ``effective'' equation of motion for the electron which includes RR, one ``removes'' the degrees of freedom of the electromagnetic field \cite{Teitelboim_1971}. This is achieved in \onlinecite{Landau_b_2_1975} at the level of the Lagrangian of the system electron+electromagnetic field and an interesting connection of the RR problem with the derivation of the so-called Darwin Lagrangian is indicated. By working at the level of the equations of motion \eqref{M_L}, one first employs the Green's function method and formally determines the retarded solution $F_{T,\text{ret}}^{\mu\nu}(x)$ of the (inhomogeneous) Maxwell's equations: $F_{T,\text{ret}}^{\mu\nu}(x)=F^{\mu\nu}(x)+F_{S,\text{ret}}^{\mu\nu}(x)$ \cite{Teitelboim_1971}. Substitution of $F_{T,\text{ret}}^{\mu\nu}(x)$ in the Lorentz equation eliminates the electromagnetic field's degrees of freedom, but it is not straightforward because $F_{T,\text{ret}}^{\mu\nu}(x)$ has to be calculated at the electron's position, where the electron current diverges. This difficulty is circumvented by modeling the electron as a uniformly charged sphere of radius $a$ tending to zero. After performing the substitution, and by neglecting terms which vanish in the limit $a\to 0$, one obtains the equation $(m_0+\delta m)du^{\mu}/ds=eF^{\mu\nu}u_{\nu}+F_R^{\mu}$ with $\delta m=(4/3)e^2/a$ being formally diverging. However, it is important to note that the only diverging term in the limit $a\to 0$ is proportional to the electron four-acceleration. At this point a sort of ``classical renormalization principle'' is employed, saying that what one measures experimentally as the physical electron mass $m$ is the overall coefficient of the four-acceleration $du^{\mu}/ds$. Therefore, one sets $m=m_0+\delta m$ and obtains the so-called Lorentz-Abraham-Dirac (LAD) equation: \begin{equation} \label{LAD} m\frac{du^{\mu}}{ds}=eF^{\mu\nu}u_{\nu}+\frac{2}{3}e^2\left(\frac{d^2u^{\mu}}{ds^2}+\frac{du^{\nu}}{ds}\frac{du_{\nu}}{ds}u^{\mu}\right). \end{equation} We point out that renormalization in quantum field theory is based on the fact that the bare quantities, like charge and mass, appear in the Lagrangian density of the theory, which is not an observable physical quantity. On the other hand, the bare electron mass $m_0$, which is formally negatively diverging for $m$ to be finite, appears here in the system of equations \eqref{M_L}, which should ``directly'' provide classical physical observables, like the electron trajectory. On the other hand, it is also known that the LAD equation is plagued with physical inconsistencies like, for example, the existence of the so-called ``runaway'' solutions with an exponentially-diverging electron acceleration, even in the absence of an external field (see the books \onlinecite{Hartemann_b_2001,Rohrlich_b_2007} for reviews on these issues). It was first shown in \onlinecite{Landau_b_2_1975} that in the nonrelativistic limit the RR force given by Eq. \eqref{F_R} is much smaller than the Lorentz force, if the typical wavelength $\lambda$ and the typical field-amplitude $F$ of the external electromagnetic field fulfill the two conditions \begin{align} \label{Cond_LL} \lambda\gg \alpha\lambda_C, && F\ll \frac{F_{\text{cr}}}{\alpha}, \end{align} where $F_{\text{cr}}$ is the critical electromagnetic field of QED (see Sec. \ref{FED_Q}). This allows for the reduction of order in the LAD equation, i.e., for the substitution of the electron acceleration in the RR force via the Lorentz force divided by the electron mass. In order to perform the analogous reduction of order in the relativistic case, the conditions \eqref{Cond_LL} have to be fulfilled in the instantaneous rest-frame of the electron \cite{Landau_b_2_1975}. The result is the so-called Landau-Lifshitz (LL) equation \begin{equation} \label{LL} \begin{split} m\frac{d u^{\mu}}{ds}=&eF^{\mu\nu}u_{\nu}+\frac{2}{3}e^2\left[\frac{e}{m}(\partial_{\alpha}F^{\mu\nu})u^{\alpha}u_{\nu}\right.\\ &\left.-\frac{e^2}{m^2}F^{\mu\nu}F_{\alpha\nu}u^{\alpha}+\frac{e^2}{m^2}(F^{\alpha\nu}u_{\nu})(F_{\alpha\lambda}u^{\lambda})u^{\mu}\right]. \end{split} \end{equation} The LL equation is not affected by the shortcomings of the LAD equation: for example, it is evident that if the external field vanishes, so does the electron acceleration. Most importantly, the conditions \eqref{Cond_LL} in the instantaneous rest-frame of the electron have always to be fulfilled in the realm of classical electrodynamics, i.e., if quantum effects are neglected. In order for this to be true, in fact, the two weaker conditions $\lambda\gg \lambda_C$ and $F\ll F_{\text{cr}}$ have to be fulfilled in the instantaneous rest-frame of the electron: the first guarantees that the electron's wave function is well localized and the second ensures that pure quantum effects, like photon recoil or spin effects are negligible \cite{Landau_b_4_1982,Baier_b_1998,Ritus_1985} (see also Sec. \ref{FED_Q}). This observation led F. Rohrlich to state recently that the LL equation is the ``physically correct'' classical relativistic equation of motion of a charged particle \cite{Rohrlich_2008}. Rohrlich's statement is also supported by the findings in \onlinecite{Spohn_2000}, where it is shown that the physical solutions of the LAD equation, i.e., those which are not runaway-like, are on the critical manifold of the LAD equation itself and are governed there exactly by the LL equation. On the other hand, since the LL equation is derived from the LAD equation, one may still doubt on its rigorous validity, due to the application in the latter equation of the ``suspicious'' classical mass-renormalization procedure. However, this procedure is avoided in \onlinecite{Gralla_2009} by employing a more sophisticated zero-size limiting procedure, where also the charge and the mass of the particle are sent to zero but in such a way that their ratio remains constant. The authors conclude that at the leading-order level the LL equation represents the self-consistent perturbative equation of motion for a charge without electric and magnetic moment. The motion of a continuous charge distribution interacting with an external electromagnetic field is also investigated by a self-consistent model and at a more formal level in \onlinecite{Burton_2007}. From the original derivation of the LL equation from the LAD equation in \onlinecite{Landau_b_2_1975}, it is expected that the two equations should predict the same electron trajectory, possibly with differences smaller than the quantum effects. This conclusion has been recently confirmed by analytical and numerical investigations in \onlinecite{Hadad_2010} for an external plane-wave field with linear and circular polarization and in \onlinecite{Bulanov_2011} for different time-dependent external electromagnetic field configurations. An effective numerical method to calculate the trajectory of an electron via the LL equation, which explicitly maintains the relativistic covariance and the mass-shell condition $u^2=1$, has been advanced in \onlinecite{Harvey_2011}. An alternative numerical method for determining the dynamics of an electron including RR effects has been proposed in \onlinecite{Mao_2010}. We should emphasize that the LL equation is not the only equation which has been suggested to overcome the inconsistencies of the LAD equation. A list of alternative equations can be found in the recent review \cite{Hammond_2010_b} (see also \onlinecite{Seto_2011}). A phenomenological equation of motion, including RR and quantum effects related to photon recoil, has been suggested in \onlinecite{Sokolov_2009,Sokolov_2010} (see also Sec. \ref{QRR}). The authors write the differential variation of the electron momentum as due to two contributions: one arising from the external field and one corresponding to the recoil of the emitted photon. The resulting equation can be written as the system \begin{equation} \label{Sok_Eq} \left\{ \begin{aligned} &m\frac{dx^{\mu}}{d\tau}=p^{\mu}+\frac{2}{3}e^2\frac{\mathcal{I}_{\text{QED}}}{\mathcal{I}_L}\frac{eF^{\mu\nu}p_{\nu}}{m^2}\\ &\frac{dp^{\mu}}{d\tau}=eF^{\mu\nu}\frac{dx_{\nu}}{d\tau}-\mathcal{I}_{\text{QED}}\frac{p^{\mu}}{m}, \end{aligned} \right. \end{equation} where $\tau$ is the time in the ``momentarily comoving Lorentz frame'' of the electron where the spatial components of $p^{\mu}$ instantaneously vanish, $\mathcal{I}_{\text{QED}}$ is the quantum radiation intensity \cite{Ritus_1985} and $\mathcal{I}_L=(2/3)\alpha\omega_0^2\xi_0^2$. The expression of $\mathcal{I}_{\text{QED}}$ in the case of a plane wave is employed, which is valid only for an ultrarelativistic electron in the presence of a slowly-varying and undercritical otherwise arbitrary external field (see Sec. \ref{FED_Q}). It has also to be stressed that the original LAD equation is still the subject of extensive investigation (the first study of the LAD equation in a plane-wave field was performed in \onlinecite{Hartemann_1996}). In \onlinecite{Ferris_2011}, for example, the origin of the Schott term in the RR force, i.e., the term proportional to the derivative of the electron acceleration (see Eq. \eqref{F_R}), is thoroughly investigated and in \onlinecite{Kazinski_2011} the asymptotics of the physical solutions of the LAD equation at large proper times are obtained. Whereas, in \onlinecite{Noble_2011} a kinetic theory of RR is proposed, based on the LAD equation and applicable to study systems of many particles including RR (this last aspect is also considered in \onlinecite{Rohrlich_b_2007}). Enhancement of RR effects due to the coherent emission of radiation by a large number of charges is discussed in \onlinecite{Smorenburg_2010}. In this respect, we mention here that, in order to investigate strong laser-plasma interactions at intensities exceeding $10^{23}\;\text{W/cm$^2$}$, RR effects have been also implemented in Particle-In-Cell (PIC) codes \cite{Zhidkov_2002,Tamburini_2010,Tamburini_2011} by modifying the Vlasov equation for the electron distribution function according to the LL equation. Specifically, in \onlinecite{Zhidkov_2002} it is shown that in the collision of a laser beam with intensity $I_0=10^{23}\;\text{W/cm$^2$}$ with an overdense plasma slab, about $35\%$ of the absorbed laser energy is converted into radiation and that the effect of RR amounts to about $20\%$. One-dimensional \cite{Tamburini_2010} and three-dimensional \cite{Tamburini_2011} PIC simulations have shown that RR effects strongly depend on the polarization of the driving field: while for circular polarization they are negligible even at $I_0\sim 10^{23}\;\text{W/cm$^2$}$, at those intensities they are important for linear polarization. The simulations also show the beneficial effects of RR in reducing the energy spread of ion beams generated via laser-plasma interactions (see also Sec. \ref{LA}). A different beneficial effect of RR on ion acceleration has been found in \onlinecite{Chen_2011} for the case of a transparent plasma: RR strongly suppresses the backward motion of the electrons, cools them down and increases the number of ions to be bunched and accelerated. Finally, the system in Eq. \eqref{Sok_Eq} has been implemented in a 3D PIC code in \onlinecite{Sokolov_2009} showing that a laser pulse with intensity $10^{22}\;\text{W/cm$^2$}$ loses about $27\%$ of its energy in the collision with a plasma slab. The same system of equations has been employed to study the penetration of ultra-intense laser beams into a plasma in the hole-boring regime \cite{Naumova_2009} and to investigate the process of ponderomotive ion acceleration at ultrahigh laser intensities in overcritical bulk targets \cite{Schlegel_2009}. \subsection{The classical radiation dominated regime} As it was already observed in \onlinecite{Landau_b_2_1975}, the fact that the RR force in the LL equation has to be much smaller than the Lorentz force in the instantaneous rest-frame of the electron does not exclude that some components of the two forces can be of the same order of magnitude in the laboratory system. This occurs if the condition $\alpha\gamma^2 F/F_{\text{cr}}\sim 1$ is fulfilled at any instant, with $\gamma$ being the relativistic Lorentz factor of the electron and $F$ the amplitude of the external electromagnetic field. For an ultrarelativistic electron this condition can be fulfilled also in the realm of classical electrodynamics (quantum recoil effects are negligible if $\gamma F/F_{\text{cr}}\ll 1$) and it characterizes the so-called Classical Radiation-Dominated Regime (CRDR). The CRDR has been investigated in \onlinecite{Shen_1970} for a background constant and uniform magnetic field. In \onlinecite{Koga_2005} an equivalent definition of the CRDR in the presence of a background laser field has been formulated, as the regime where the average energy radiated by the electron in one laser period is comparable with the initial electron energy. By estimating the radiated power $\mathcal{P}_L$ from the relativistic Larmor formula $\mathcal{P}_L=-(2/3)e^2(du_{\mu}/ds)(du^{\mu}/ds)$ \cite{Jackson_b_1975} with $du^{\mu}/ds\to(e/m)F^{\mu\nu}u_{\nu}$, one obtains that $\mathcal{P}_L\sim \alpha\chi_0\xi_0\varepsilon_0$. Therefore, the conditions of being in the CRDR are \begin{align} \label{R_C} R_C=\alpha\chi_0\xi_0\approx 1, && \chi_0\ll 1, \end{align} where, as has been seen in Sec. \ref{FED_Q}, the second condition ensures in particular that the quantum effects like photon recoil are negligible. The same condition $R_C\approx 1$ has been obtained in \onlinecite{Di_Piazza_2008_a} by exactly solving the LL equation \eqref{LL} for a general plane-wave background field. The analytical solution shows in fact that for an ultrarelativistic electron the main effect of RR is due to the last term in Eq. \eqref{LL}. As a consequence, while the quantity $u_-(\phi)$ is constant if the equation of motion is that due to Lorentz, it decreases here with respect to $\phi$ as $u_-(\phi)=u_{0,-}/h(\phi)$, where $u_0^{\mu}$ is the four-velocity at an initial $\phi_0$ and \begin{equation} \label{h} h(\phi)=1+\frac{2}{3}\frac{R_C}{\omega_0}\int_{\phi_0}^{\phi}d\varphi \left(\frac{d\psi(\varphi)}{d\varphi}\right)^2, \end{equation} where the four-potential of the wave has been assumed to have the form $A^{\mu}(\phi)=A_0^{\mu}\psi(\phi)$ (see Sec. \ref{FED_Q}, below Eq. \ref{chi_0}). This effect has been recently suggested in \onlinecite{Harvey_2011b} as a possible signature to measure RR (see also \onlinecite{Lehmann_2011}). The two conditions in Eq. \eqref{R_C} are, in principle, compatible for sufficiently large values of $\xi_0$. For example, for an optical ($\omega_0=1\;\text{eV}$) laser field with an average intensity of $10^{24}\;\text{W/cm$^2$}$ and for an electron initially counterpropagating with respect to the laser field with an energy of $20\;\text{MeV}$, it is $\chi_0=0.16$ and $R_C=1.3$. This example shows that, in general, it is not experimentally easy to enter the CRDR at least with presently-available laser systems. In \onlinecite{Di_Piazza_2009} a different regime has been investigated, which is parametrically less demanding than the CRDR but in which the effects of RR are still large. In this regime the change in the longitudinal (with respect to the laser field propagation direction) momentum of the electron due to RR in one laser period is of the order of the electron's longitudinal momentum itself in the laser field. As a result, it is found that in the ultrarelativistic case and for a few-cycle pulse, if the conditions \begin{equation} \label{cond_domin2} R_C\gtrsim \frac{4\gamma_0^2-\xi_0^2}{2\xi_0^2}>0 \end{equation} are fulfilled, then the electron is reflected in the laser field only if RR is taken into account (see also \onlinecite{Harvey_2011c} for a recent investigation of the electron's dynamics in the two complementary regimes $2\gamma_0\lessgtr \xi_0$ including RR effects). This can have measurable effects if one exploits the high directionality of the radiation emitted by an ultrarelativistic electron (see Sec. \ref{TCS_General}). The results in \onlinecite{Di_Piazza_2009} show in fact that the apex angle of the angular distribution of the emitted radiation, with and without RR effects included, may differ by more than $10^{\circ}$ already at an average optical laser intensity of $5\times 10^{22}\;\text{W/cm$^2$}$ ($\xi_0\approx 150$) and at an initial electron energies of $40\;\text{MeV}$ ($2\gamma_0\approx 156$) for which $R_C\approx 0.08$. Small RR effects on photon spectra emitted by initially bound electrons had already been predicted via numerical integration of the LL equation in \onlinecite{Keitel_1998} well below the CRDR. \subsection{Quantum radiation reaction} \label{QRR} The shortcomings of the classical approaches to the problem of RR suggest that it can be fully understood only at the quantum level. In the seminal paper \onlinecite{Moniz_1977} the origin of the classical inconsistencies, like the existence of runaway solutions of the LAD equation, were clarified in the nonrelativistic case. The authors first show that such inconsistencies are also absent in classical electrodynamics if one considers charge distributions with a typical radius larger than the classical electron radius $r_0=\alpha\lambda_C\approx 2.8\times 10^{-13}\;\text{cm}$. Going to the nonrelativistic quantum theory and by analyzing the Heisenberg equations of motion of the electron in an external time-dependent field, the authors conclude that the quantum theory of a pointlike particle does not admit any runaway solutions, provided that the external field varies slowly along a length of the order of $\lambda_C$ (this is an obvious assumption in the realm of nonrelativistic theory, as time-dependent fields with typical wavelengths of the order of $\lambda_C$ would in principle allow for $e^+\text{-}e^-$ pair production, see Sec. \ref{PP}). From this point of view a classical theory of RR has only physical meaning as the classical limit ($\hbar\to 0$) of the corresponding quantum theory and the authors indicate that the resulting equation of motion is the nonrelativistic LL equation with the bare mass $m_0$ (it is shown that the electrostatic self-energy of a point charge vanishes in nonrelativistic quantum electrodynamics). On the other hand, if one considers the quantum equations of motion of a charge distribution and performs the classical limit before the pointlike limit, then the classical equations of motion of the charge distribution are, of course, recovered and, once the point-like limit is then performed, runaway solutions appear again. The nonrelativistic form of the LL equation has been also recovered from quantum mechanics in \onlinecite{Krivitskii_1991} by including radiative corrections to the time-dependent electron momentum operator in the Heisenberg representation, and by calculating the time-derivative of the average momentum in a semiclassical state. The situation in the relativistic theory is less straightforward because relativistic quantum electrodynamics, i.e., QED, is a field theory fundamentally different from classical electrodynamics. The first theory of relativistic quantum RR goes back to W. Heitler and his group \cite{Heitler_b_1984,Jauch_b_1976}. However, the evaluations of the QED amplitudes in Heitler's theory involve the solution of complicated integral equations and it has given a practical result only in the calculation of the total energy emitted by a nonrelativistic quantum oscillator, with and without RR. At first sight one would say that RR effects are automatically taken into account in QED, because the electromagnetic field is treated as a collection of photons that take away energy and momentum, when they are emitted by charged particles. However, photon recoil is always proportional to $\hbar$, making it a purely quantum quantity with no classical counterpart. Moreover, if one calculates the spectrum of multiphoton Compton scattering in an external plane-wave field, for example, and then performs the classical limit $\chi_0\to 0$, one obtains the corresponding multiphoton Thomson spectrum calculated via the Lorentz equation and not via the LAD or the LL equation (see also Sec. \ref{TCS_General}). Finally, it has also been seen that in classical electrodynamics the RR effects may not be a small perturbation on the Lorentz dynamics and they cannot be obtained as the result of a single limiting procedure. Otherwise they would always appear as a small correction. In order to understand what RR is in QED, it is more convenient to go back to Eq. \eqref{M_L} and to notice that the LAD, namely the LL, equation is equivalent to the coupled system of Maxwell and Lorentz equations. If one determines the trajectory of the electron via the LL equation and then calculates the total electromagnetic field $F_T^{\mu\nu}(x)$ via the Li\'{e}nard-Wiechert four-potential \cite{Landau_b_2_1975}, one has solved completely the classical problem of the radiating electron in the given electromagnetic field. As has been discussed in Sec. \ref{FED_Q}, the solution of the analogous problem in strong-field QED would correspond to completely determine the $S$-matrix in Eq. \eqref{S}, as well as the asymptotic state $|t\to+\infty\rangle$ for the given initial state $|t\to-\infty\rangle=|e^-\rangle$, which represents a single electron. The first-order term in the perturbative expansion of the $S$-matrix corresponds to the process of multiphoton Compton scattering and then, classically, to the Lorentz dynamics. Whereas, all high-order terms give rise to radiative corrections and to high-order coherent and incoherent (cascade) processes, and determine what we call ``quantum RR''. Here, by high-order coherent processes is meant those involving more than one basic QED process (photon emission by an electron/positron or $e^+\text{-}e^-$ photoproduction) but all occurring in the same formation region. Analogously, in higher-order incoherent or cascade processes each basic QED process occurs in a different formation region. Now, in the case of a background plane wave at $\xi_0\gg 1$ and $\chi_0\lesssim 1$, the quantum effects are certainly important but the radiative corrections and higher-order coherent processes scale with $\alpha$ and can be neglected \cite{Ritus_1972}. Also, if $\chi_0$ does not exceed unity then the photons emitted by the electron are mainly unable, by interacting again with the laser field, to create $e^+\text{-}e^-$ pairs, as the pair production probability is exponentially suppressed (see also Secs. \ref{PP} and \ref{Cascade}). Therefore, it can be concluded that at $\xi_0\gg 1$ and $\chi_0\lesssim 1$, RR in QED corresponds to the overall photon recoil experienced by the electron when it emits many photons consecutively and incoherently \cite{Di_Piazza_2010}. A qualitative understanding of the above conclusion can be attained by assuming that $\chi_0\ll 1$ and by estimating the average number $N_{\gamma}$ of photons emitted by an electron in one laser period at $\xi_0\gg 1$. Since the probability of emitting one photon in a formation length is of the order of $\alpha$ and since one laser period contains about $\xi_0$ formation lengths \cite{Ritus_1985} then $N_{\gamma}\sim \alpha\xi_0$. Also, the typical energy $\omega'$ of a photon emitted by an electron is of the order $\omega'\sim \chi_0\varepsilon_0$, then the average energy $\mathcal{E}$ emitted by the electron is $\mathcal{E}\sim \alpha\xi_0\chi_0\varepsilon_0=R_C\varepsilon_0$. This estimate is in agreement with the classical result obtained from the LL equation. In other words, the classical limit of RR in this regime corresponds to the emission of a higher and higher number of photons all with an energy much smaller than the electron energy, in such a way that even though the recoil at each emission is almost negligible, the cumulative effect of all photon emissions may have a finite nonnegligible effect. Note that $\omega'$ and $N_{\gamma}$ are both pure quantum quantities and only their product $\mathcal{E}$ has a classical analogue in the limit $\chi_0\to 0$. These considerations have allowed for the introduction in \onlinecite{Di_Piazza_2010} of the Quantum Radiation-Dominated Regime (QRDR), which is characterized by multiple emission of photons already in one laser period. This regime is then characterized by the conditions \begin{align} \label{R_Q} R_Q=\alpha\xi_0\approx 1, && \chi_0\gtrsim 1. \end{align} Quantum photon spectra have been calculated numerically in \onlinecite{Di_Piazza_2010} without RR, i.e., by including only the emission of one photon (four-momentum $k^{\prime\mu}$), and with RR, i.e., by including multiple-photon emissions (and by integrating with respect to all the four-momenta of the emitted photons except one indicated as $k^{\prime\mu}$). The results show that in the QRDR the effects of RR are essentially three (see Fig. \ref{QRR_Spectrum}): 1) increase of the photon yield at low photon energies; 2) decrease of the photon yield at high photon energies; 3) shift of the maximum of the photon spectrum towards low photon energies. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig11.eps} \end{center} \caption{(Color online) Quantum photon spectra as a function of $\varpi'=k'_-/p_{0,-}$ calculated with (solid, black line) and without (long dashed, red line) RR and the corresponding classical ones with (short dashed, blue line) and without (dotted, magenta line) RR. The error bars in the quantum spectrum with RR stem from numerical uncertainties in multidimensional integrations. The numerical parameters in our notation are: $\varepsilon_0=1\;\text{GeV}$, $\omega_0=1.55\;\text{eV}$ and $I_0=10^{23}\;\text{W/cm$^2$}$ ($R_Q=1.1$ and $\chi_0=1.8$). Adapted from \onlinecite{Di_Piazza_2010}. } \label{QRR_Spectrum} \end{figure} Figure \ref{QRR_Spectrum} also shows that the classical treatment of RR (via the LL equation) artificially overestimates the above effects, the reason being that quantum corrections decrease the average energy emitted by the electron with respect to the classical value \cite{Ritus_1985}. However, at $\chi_0\ll 1$, i.e., when the recoil of each emitted photon is much smaller than the electron energy, then the quantum spectra converge into the corresponding classical ones. As mentioned in Sec. \ref{RR}, a semiclassical phenomenological approach to RR in the quantum regime has been proposed in \onlinecite{Sokolov_2009,Sokolov_2010b}. Finally, the quantum modifications induced by the electron's self-field onto the Volkov states (see Eq. \eqref{V_S}) have been recently investigated in \onlinecite{Meuren_2011}. It is found that the classical expression of the electron quasimomentum $q_0^{\mu}$ in a linearly polarized plane wave (see Eq. \eqref{q}) admits a correction depending on the quantum parameter $\chi_0$ and also that self-field effects induce a peculiar dynamics of the electron spin. \section{Vacuum-polarization effects} \label{VPEs} QED predicts that photons interact with each other also in vacuum \cite{Landau_b_4_1982}. Effects arising from this purely quantum interaction are referred to as vacuum polarization effects. This is in contrast to classical electrodynamics where the linearity of Maxwell's equations in vacuum forbids self-interaction of the electromagnetic field in the vacuum itself. The possibility of photon-photon interaction in vacuum, within the framework of QED, can be understood qualitatively by observing that a photon may locally ``materialize'' into an $e^+\text{-}e^-$ pair which, in turn, interacts with other photons. For the same reason a background electromagnetic field can influence photon propagation (see Fig. \ref{VP} and \onlinecite{Landau_b_4_1982}). \begin{figure} \begin{center} \includegraphics[width=5cm]{Di_Piazza_fig12.eps} \end{center} \caption{Vacuum polarization diagram in an external background electromagnetic field. The thick electron lines indicate electron propagators calculated in the Furry picture, accounting exactly for the presence of the background field.} \label{VP} \end{figure} In the latter case the extension $l_f$ of the region where this transformation occurs, i.e., its formation length, depends, in principle, on the structure of the background field \cite{Baier_2005}. However, in some cases it can be estimated qualitatively via the Heisenberg uncertainty principle from the typical momentum $p$ flowing in the $e^+\text{-}e^-$ loop in Fig. \ref{VP}. We consider, for example, a constant background electromagnetic field (or a slowly-varying one, at leading order in the space-time derivatives of the field itself). In this case, if the energy $\omega$ of the incoming photon (see Fig. \ref{VP}) is at most of the order of $m$, then the momentum $p$ flowing in the $e^+\text{-}e^-$ loop is of the order of $m$ and $l_f\sim 1/p\sim\lambda_C$. If $\omega\gg m$ the analysis is more complicated and the formation length strongly depends on the structure of the background field. From the theoretical point of view it is convenient to distinguish between low-energy vacuum-polarization effects if $\omega\ll m$ and high-energy ones if $\omega\gtrsim m$. \subsection{Low-energy vacuum-polarization effects} \label{Diff_Low} The scattering in vacuum of a real photon by another real photon is possibly the most fundamental vacuum-polarization process \cite{Landau_b_4_1982} and it has not yet been observed experimentally. The total cross section of the process depends only on the Lorentz-invariant parameter $\eta=(k_1k_2)/m^2$, with $k_1^{\mu}$ and $k_2^{\mu}$ being the four-momenta of the colliding photons or, equivalently, on the energy $\omega^*$ of the two colliding photons in their center-of-momentum system ($\eta=2\omega^{*\,2}/m^2$). This process has been investigated in \onlinecite{Euler_1936_a} in the low-energy limit $\eta\ll 1$ and then in \onlinecite{Akhiezer_1937} in the high-energy limit $\eta\gg 1$. The complete expression of the cross section $\sigma_{\gamma\gamma\to\gamma\gamma}$ was calculated in \onlinecite{Karplus_1951} and can also be found in \onlinecite{Landau_b_4_1982} (see also Fig. \ref{Gamma_Gamma}). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig13.eps} \end{center} \caption{Cross section of real photon-photon scattering as a function of the energy of the colliding photons in their center-of-momentum system in units of the electron mass. Adapted from \onlinecite{Landau_b_4_1982}. Copyright Elsevier (1982).} \label{Gamma_Gamma} \end{figure} In the low-energy limit $\eta\ll 1$ the cross section $\sigma_{\gamma\gamma\to\gamma\gamma}$ is given by \cite{Landau_b_4_1982} \begin{align} \sigma_{\gamma\gamma\to\gamma\gamma}=\frac{973}{81000\pi}\alpha^4\lambda_C^2\eta^3 && \eta\ll 1 \end{align} and in the high-energy one $\eta\gg 1$ by \cite{Landau_b_4_1982,Baier_1974} \begin{align} \begin{split} \sigma_{\gamma\gamma\to\gamma\gamma}=&\frac{1}{\pi}\left[\frac{108}{5}+\frac{13}{2}\pi^2-8\pi^2\zeta(3)+\frac{148}{225}\pi^4\right.\\ &-24\zeta(5)\Big]\alpha^4\lambda_C^2\frac{1}{\eta}, \end{split}&& \eta\gg 1 \end{align} where $\zeta(x)$ is the Riemann zeta function \cite{NIST_b_2010}. In terms of the center-of-momentum energy $\omega^*$, the cross section becomes $\sigma_{\gamma\gamma\to\gamma\gamma}[\text{cm$^2$}]=7.4\times 10^{-66}(\omega^*[\text{eV}])^6$ at $\omega^*\ll m$ and $\sigma_{\gamma\gamma\to\gamma\gamma}[\text{cm$^2$}]=5.4\times 10^{-36}/(\omega^*[\text{GeV}])^2$ at $\omega^*\gg m$. The steep dependence of $\sigma_{\gamma\gamma\to\gamma\gamma}$ on $\eta$ for $\eta\ll 1$ is the main reason why real photon-photon scattering has, so-far, eluded experimental observation (see the reviews \onlinecite{Salamin_2006,Marklund_2006} for experiments and experimental proposals until 2005 aiming to observe real photon-photon scattering in vacuum). However, various proposals have been put forward recently in order to observe this process by colliding strong laser beams which contain a large number of photons. A common theoretical starting point of all these proposals is the ``effective-Lagrangian'' approach \cite{Dittrich_b_1985,Dittrich_b_2000}. In this approach the interaction among photons in vacuum is described via an effective Lagrangian density of the electromagnetic field. By starting from the total Lagrangian density of the classical electromagnetic field and of the quantum $e^+\text{-}e^-$ Dirac field, one integrates out the degrees of freedom of the latter field and is left with a Lagrangian density depending only on the electromagnetic field. As has been seen above, the formation region of photon-photon interaction at low energies is of the order of $\lambda_C$, therefore if the classical electromagnetic field $F^{\mu\nu}(x)=(\bm{E}(x),\bm{B}(x))$ comprises only wavelengths much larger than $\lambda_C$, the interaction is approximately pointlike and the effective Lagrangian density is accordingly a local quantity. Also, since the effective Lagrangian density is a Lorentz invariant, it can depend only on the electromagnetic field invariants $\mathscr{F}(x)$ and $\mathscr{G}^2(x)$ already introduced in Sec. \ref{FED_Q}. The complete expression of the effective Lagrangian density was reported for the first time in \onlinecite{Heisenberg_1936,Weisskopf_1936} (see also \onlinecite{Schwinger_1951}) and it is known as the Euler-Heisenberg Lagrangian density. Here we are interested only in the experimentally-relevant low-intensity limit $|\mathscr{F}(x)|,|\mathscr{G}(x)|\ll F_{\text{cr}}^2$ and the leading-order Euler-Heisenberg Lagrangian density $\mathscr{L}_{EH}(x)$ reads \cite{Dittrich_b_1985,Dittrich_b_2000} \begin{equation} \label{L_EH} \mathscr{L}_{EH}(x)=-\frac{1}{4\pi}\mathscr{F}(x)+\frac{\alpha}{360\pi^2}\frac{4\mathscr{F}^2(x)+7\mathscr{G}^2(x)}{F^2_{cr}}. \end{equation} Different experimental observables have been suggested to detect low-energy vacuum-polarization effects. Those will be reviewed next. \subsubsection{Experimental suggestions for direct detection of photon-photon scattering} The most direct way to search for photon-photon scattering events in vacuum by means of laser fields is to let two laser beams collide and to look for scattered photons. However, by employing a third ``assisting'' laser beam, if one of the final photons is kinematically allowed to be emitted along this beam with the same frequency and polarization, then the number of photon-photon scattering events can be coherently enhanced \cite{Varfolomeev_1966}. In this laser-assisted setup the ``signal'' of photon-photon scattering is, of course, the remaining outgoing photon. In \onlinecite{Lundstroem_2006,Lundin_2007} an experiment has been suggested to observe laser-assisted photon-photon scattering with the Astra-Gemini laser system (see Sec. \ref{L_O}). The authors found a particular ``three-dimensional'' setup, which turns out to be especially favorable for the observation of the process (see Fig. \ref{Gamma_Gamma_Lundstroem}). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig14.eps} \end{center} \caption{(Color) Schematic ``three-dimensional'' setup for laser-assisted photon-photon scattering involving two incoming beams (in blue), an assisting one (in red) and a scattered one (in violet). From \onlinecite{Lundstroem_2006}.} \label{Gamma_Gamma_Lundstroem} \end{figure} The number $N_{\gamma}$ of photons scattered in one shot for an optimal choice of the geometrical factors and of the polarization angles between the incoming and the assisting beams is found to be \begin{equation} N_{\gamma}\approx 0.25\frac{P_1[\text{PW}]P_2[\text{PW}]P_3[\text{PW}]}{(\lambda_4[\text{$\mu$m}])^3}. \end{equation} Here, $P_1$ and $P_2$ are the powers of the incoming beams, $P_3$ is the power of the assisting beam and $\lambda_4$ is the wavelength of the scattered wave to be measured. By plugging in the feasible values for Astra-Gemini $P_1=P_2=0.1\;\text{PW}$ and $P_3=0.5\;\text{PW}$, one obtains $N_{\gamma}\approx 0.07$, i.e., roughly one photon scattered every 15 shots (for the parameters of Astra Gemini the wavelength of both incoming beams is chosen as $0.4\;\text{$\mu$m}$, that of the assisting beam as $0.8\;\text{$\mu$m}$, so that the wavelength $\lambda_4=0.276\;\text{$\mu$m}$ of the scattered photon is different from those of the incoming and assisting beams). The quantum interaction among photons in vacuum has been exploited in \onlinecite{King_2010} to propose, for the first time, a double-slit setup comprised only of light (see also \onlinecite{Marklund_2010}). In this setup two strong parallel beams collide head-on with a counterpropagating probe pulse. The photons of the probe have the choice to interact either with one or with the other strong beam, and, when scattered, they are predicted to build an interference pattern with alternating minima and maxima typical of double-slit experiments (see Fig. \ref{DS}). Also, if one of the slits is closed, i.e., if the probe collides with only one strong beam, the interference fringes disappear. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig15.eps} \end{center} \caption{(Color) Intensity $I_d$ of the vacuum-scattered wave for a probe beam propagating along the positive $y$ direction and colliding with two strong beams aligned along the $x$ axis. The crosses correspond to coordinates $x_n$ according to the classical prediction $x_n=(n+1/2)\lambda_pd/D$, where $n$ in an integer number, $\lambda_p$ is the wavelength of the probe field, $d$ is the distance between the interaction region and the observation screen and $D$ is the distance between the centers of the two strong beams. The numerical values of the parameters can be found in \onlinecite{King_2010}. Adapted from \onlinecite{King_2010}.} \label{DS} \end{figure} The key idea behind this setup is that the vacuum-scattered beam (intensity $I_d$), although propagating along the same direction as the probe, has a much wider angular distribution than the latter, offering the possibility of detecting vacuum-scattered photons outside the focus of the undiffracted probe beam. For a strong-field intensity of $I_0\approx 5\times 10^{24}\;\text{W/cm$^2$}$ that may be in the near future be available at ELI or at HiPER and for a probe beam with wavelength $\lambda_p=0.527\;\text{$\mu$m}$ and intensity $I_p=4\times 10^{16}\;\text{W/cm$^2$}$, it is predicted that about four photons per shot would contribute to build up the interference pattern in the observable region (it is the region outside the circle in Fig. \ref{DS}, where $I_d>100\,I_p$). The diffraction of a probe beam in vacuum by a single focused strong laser pulse is also investigated in \onlinecite{Tommasini_2010} in the case of almost counterpropagating beams. For optimal lasers parameters and at a strong-laser power of $100\;\text{PW}$ the diffracted vacuum signal is predicted to be measurable in a single shot. The effects on photon-photon scattering of the temporal profile of the laser pulses have been recently investigated in \onlinecite{King_2012}, showing a suppression of the number of vacuum scattered photons with respect to the infinite-pulse (monochromatic) case. The concept of Bragg scattering has been exploited in \onlinecite{Kryuchkyan_2011} to observe the scattering of photons by a modulated electromagnetic-field structure in vacuum. If a probe wave passes through a series of parallel strong laser pulses and if the Bragg condition on the impinging angle is fulfilled, the number of diffracted photons can be strongly enhanced. At a fixed intensity for each strong beam the enhancement factor with respect to laser-assisted photon-photon scattering is equal to the number of beams in the periodic structure. However, in experiments usually the total energy of the laser beams is fixed and an enhancement by a factor of two is predicted. By considering $N_G$ equal Gaussian pulses propagating along the $x$ direction and their centers separated by a distance $D > 2w_z$ from each other, the resulting photon-photon scattering probability will be proportional to the phase-matching factor $\mathscr{P}$, with \begin{equation} \mathscr{P}=\frac{\sin^2(\delta k_z N_GD/2)}{\sin^2(\delta k_z D/2)}, \end{equation} where the vector $\delta \bm{k}=\bm{k}_2-\bm{k}_1$ is the difference between the wave vectors of the reflected and incident waves. The Bragg condition is satisfied for $\delta k_z=2\pi l/D$, with $l$ being an integer. By employing ten optical laser beams with a wavelength of $1\;\text{$\mu$m}$ and each with an intensity of $2.3\times 10^{23}\;\text{W/cm$^2$}$, about five vacuum-scattered photons are predicted per shot. Finally, an enhancement of vacuum-polarization effects in laser-laser collision has been predicted in \onlinecite{Monden_2011} by employing strong laser beams with large angular aperture. For example, it is predicted that the number of vacuum-radiated photons will be enhanced by two orders of magnitude, if the angular aperture of the colliding beams is increased from $53^{\circ}$ to $103^{\circ}$. Other experimental suggestions to measure photon-photon scattering in vacuum can be found in \onlinecite{Eriksson_2004,Tommasini_2008}. \subsubsection{Polarimetry-based experimental suggestions} The expression of the Euler-Heiseberg Lagrangian density in Eq. \eqref{L_EH}, suggests to interpret a region where only an electromagnetic field is present as a material medium characterized by a polarization $\bm{P}_{EH}=\partial \mathscr{L}_{EH}/\partial\bm{E}-\bm{E}/4\pi$ and a magnetization $\bm{M}_{EH}=\partial \mathscr{L}_{EH}/\partial\bm{B}+\bm{B}/4\pi$ \cite{Jackson_b_1975} given by \begin{align} \label{P_EH} \bm{P}_{EH}&=\frac{\alpha}{180\pi^2 F_{\text{cr}}^2}\left[2(E^2-B^2)\bm{E}+7(\bm{E}\cdot\bm{B})\bm{B}\right],\\ \label{M_EH} \bm{M}_{EH}&=\frac{\alpha}{180\pi^2 F_{\text{cr}}^2}\left[2(B^2-E^2)\bm{B}+7(\bm{E}\cdot\bm{B})\bm{E}\right]. \end{align} Note that an arbitrary single plane wave cannot ``polarize'' the vacuum, as in this case $\bm{P}_{EH}$ and $\bm{M}_{EH}$ identically vanish. Equations \eqref{P_EH} and \eqref{M_EH} indicate that the presence of an electromagnetic field in the vacuum alters the vacuum's refractive index. The situation is even more complicated because of the vectorial nature of the background electromagnetic field which polarizes the vacuum and introduces a privileged direction in it. As a result, the vacuum's refractive index is altered in a way that depends, in general, on the mutual polarizations of the probe electromagnetic field and of the background field: the polarized vacuum behaves as a birefringent medium. For example, in the case of an arbitrary constant electromagnetic field $(\bm{E},\bm{B})$, the refractive indices $n_{EH,1/2}$ of a wave propagating along the direction $\bm{n}$ and polarized along one of the two independent directions $\bm{u}_1=\bm{\mathcal{E}}/|\bm{\mathcal{E}}|$ and $\bm{u}_2=\bm{\mathcal{B}}/|\bm{\mathcal{B}}|$, with $\bm{\mathcal{E}}=\bm{E}-(\bm{n}\cdot\bm{E})\bm{E}+\bm{n}\times\bm{B}$ and $\bm{\mathcal{B}}=\bm{B}-(\bm{n}\cdot\bm{B})\bm{B}-\bm{n}\times\bm{E}$ are given by \cite{Dittrich_b_2000} \begin{align} \label{n_EH_1} n_{EH,1}&=1+\frac{4\alpha}{90\pi}\frac{(\bm{n}\times\bm{E})^2+(\bm{n}\times\bm{B})^2-2\bm{n}\cdot(\bm{E}\times\bm{B})}{F^2_{\text{cr}}},\\ \label{n_EH_2} n_{EH,2}&=1+\frac{7\alpha}{90\pi}\frac{(\bm{n}\times\bm{E})^2+(\bm{n}\times\bm{B})^2-2\bm{n}\cdot(\bm{E}\times\bm{B})}{F^2_{\text{cr}}}, \end{align} respectively. The birefringence of the polarized vacuum is exploited in \onlinecite{Heinzl_2006} to show that if a linearly-polarized probe x-ray beam (wavelength $\lambda_p$) propagates along a strong optical standing wave, then it emerges from the interaction elliptically polarized with ellipticity $\epsilon$ given by \begin{equation} \epsilon=\frac{2\alpha}{15}\kappa\frac{l_{0,R}}{\lambda_p}\frac{I_0}{I_{\text{cr}}}, \end{equation} where $\kappa\sim 1$ is a geometrical factor and $l_{0,R}$ is the Rayleigh length of the intense laser beam. If this beam is generated by a laser like ELI ($I_0\sim 10^{25}\;\text{W/cm$^2$}$), values of the ellipticities of the order of $10^{-7}$ are predicted at $\lambda_p=0.1\;\text{nm}$. Recent advances on x-ray polarimetry allow for measurement of ellipticities of the order of $10^{-9}$ at a wavelength of $0.2\;\text{nm}$ \cite{Marx_2011}. In \onlinecite{Ferrando_2007} a phase-shift has been found theoretically to be induced by vacuum-polarization effects when two laser beams cross in the vacuum, which is predicted to be measurable at laser intensities available at ELI or at HiPER. When an electromagnetic wave with wavelength $\lambda$ impinges upon a material body, the features of the scattered radiation depend on the so-called diffraction parameter $\mathscr{D}=l_{\perp}^2/\lambda d$ \cite{Jackson_b_1975}. Here, $l_{\perp}$ is the spatial dimension of the body perpendicular to the propagation direction of the incident wave and $d$ is the distance of the screen, where the radiation is detected, from the interaction region. The near region, $\mathscr{D}\gg 1$, is known as the ``refractive-index limit'', because the effects of the presence of the body can be described as if the wave propagates through a medium with a given refractive index. However, if $\mathscr{D}\lesssim 1$ then the diffraction effects become important and description of the wave-body interaction only in terms of a refractive index is in general not possible. This aspect has been pointed out in \onlinecite{Di_Piazza_2006} within the context of light-light interaction in vacuum (see also \onlinecite{Di_Piazza_2007b}). Tight focusing required to reach high intensities usually renders the interaction region so small that diffraction effects may become substantial at typical experimental conditions. In some cases diffractive effects reduce by an order of magnitude the values of the ellipticity calculated via the refractive-index approach and also induce a rotation of the main axis of the polarization ellipse with respect to the initial polarization direction of the probe field \cite{Di_Piazza_2006}. Tight focusing of the strong polarizing beam requires quite a detailed mathematical description employing a realistic focused Gaussian beam, while a simpler description was employed for the usually weakly-focused probe beam. This prevented the applicability of the results in the so-called far region where $\mathscr{D}\ll 1$ and where the spatial spreading of the probe field is also important. This assumption has been recently removed in \onlinecite{King_2010_a}, where it was also pointed out that by considering the diffraction of a probe beam by two separated beams instead of that by a single standing wave, an increase in the ellipticity and in the rotation of polarization angle by a factor 1.5 is expected. A different method based on the phase-contrast Fourier imaging technique has been suggested in \onlinecite{Homma_2011} to detect vacuum birefringence. This technique provides a very sensitive tool to measure the absolute phase shift of a probe beam when it crosses an intense laser field. Numerical simulations demonstrate the feasibility of measuring vacuum birefringence also by employing an optical probe field and a 100-PW strong laser beam. Photon ``acceleration'' in vacuum due to vacuum polarization has been studied in \onlinecite{Mendonca_2006}. This effect corresponds to a shift of the photon frequency when it passes through a strong electromagnetic wave. If $k^{\mu}_p$ is the four-momentum of a probe photon with energy $\omega_p$ when it enters a region where a strong laser beam is present, then, due to vacuum-polarization effects, $\omega_p$ becomes sensitive to the gradient of the intensity of the strong beam. As a result, a frequency up-shift (down-shift) is predicted at the rear (front) of the strong beam. According to Eqs. \eqref{n_EH_1} and \eqref{n_EH_2} the phase velocity of light in vacuum is smaller than unity. This circumstance has been exploited in \onlinecite{Marklund_2005}, where Cherenkov radiation by ultrarelativistic particles moving with constant velocity in a photon gas has been predicted, if the speed of the particle exceeds the phase velocity of light. Finally, in \onlinecite{Zimmer_2012} an induced electric dipole moment of the neutron has been proposed as a signature of the polarization of the QED vacuum. \subsubsection{Low-energy vacuum-polarization effects in a plasma} Equations \eqref{n_EH_1} and \eqref{n_EH_2} indicate that vacuum-polarization effects elicited by a plane wave with intensity $I_0$ would alter the vacuum refractive index by an amount of the order of $(\alpha/45\pi)(I_0/I_{\text{cr}})$. It was first realized in \onlinecite{Di_Piazza_2007_a} that this aspect can be in principle significantly improved in a plasma. For the sake of simplicity the case of a cold plasma was considered and the vacuum-polarization effects were implemented in the inhomogeneous Maxwell's equations as an additional ``vacuum four-current'' \cite{Di_Piazza_2007_a}. Now, unlike in the vacuum, the field invariant $\mathscr{F}(x)$ for a single monochromatic circularly polarized plane wave does not vanish in a plasma. Thus, vacuum-polarization effects in a plasma already arise in the presence of a single traveling plane wave. In \onlinecite{Di_Piazza_2007_a} the vacuum-corrected refractive index $n$ of a two-fluid electron-ion plasma (ion mass, density and charge number given by $m_i$, $n_i$ and $Z$, respectively) in the presence of a circularly polarized plane wave has been found as \begin{equation} \label{n} n=\sqrt{n_0^2+\frac{2\alpha}{45\pi}\frac{I_0}{I_{\text{cr}}}(1-n_0^2)^2}, \end{equation} where \begin{equation} n_0=\sqrt{1-\frac{4\pi e^2n_i}{m\omega_0^2}\left(\frac{1}{\sqrt{1+\xi_0^2}}+\frac{Z}{\sqrt{(m_i/m)^2+Z^2\xi_0^2}}\right)}, \end{equation} is the refractive index of the plasma without vacuum-polarization effects \cite{Mulser_b_2010}. Equation \eqref{n} already indicates the possibility of enhancing the effects of vacuum polarization by working at laser frequencies $\omega_0$ such that $n_0\ll 1$, i.e., close to the effective plasma critical frequency. This region of parameters is in general complex to investigate, due to the arising of different instabilities. However, the idealized situation investigated in \onlinecite{Di_Piazza_2007_a} shows, at least in principle, the possibility of enhancing the vacuum-polarization effects by an order of magnitude at a given intensity $I_0$ with respect, for example, to the results in \onlinecite{Di_Piazza_2006}. The effects of the presence of an additional strong constant magnetic field have been analyzed in \onlinecite{Lundin_2007}. The general theory presented in this paper covers different waves propagating in a plasma as Alfv\'{e}n modes, whistler modes, and large-amplitude laser modes. We also mention the recent paper \onlinecite{Bu_2010}, in which the photon acceleration process has been investigated in a cold plasma and the reference \onlinecite{Brodin_2007}, where vacuum-induced photon splitting in a plasma is studied. Finally, we mention the possibility of testing nonlinear vacuum QED effects in waveguides. In \onlinecite{Brodin_2001} the generation of new modes in waveguides due to vacuum-polarization effects has been predicted. More recently signatures of nonlinear QED effects in the transmitted power along a waveguide have been analyzed in \onlinecite{Ferraro_2010}. \subsection{High-energy vacuum-polarization effects} \label{Diff_High} Generally speaking the treatment of vacuum-polarization effects for an incoming photon with energy $\omega$ in the presence of a background electromagnetic field with a typical angular frequency $\omega_b$ cannot be performed in an effective Lagrangian approach if $\omega\omega_b/m^2\gtrsim 1$: the incoming photon ``sees'' the nonlocality of its interaction with the background electromagnetic field through its local ``transformation'' into an $e^+\text{-}e^-$ pair (see Fig. \ref{VP}). The technical difficulty in treating vacuum-polarization effects at high energies arises from the fact that the interaction between the virtual $e^+\text{-}e^-$ pair and the background field has to be accounted for exactly. This has been accomplished for the background field of a nucleus with charge number $Z$ such that $Z\alpha\sim 1$ and Delbr\"{u}ck scattering and photon splitting in such a field have also been observed experimentally (see the reviews \onlinecite{Milstein_1994,Lee_2003}). As we have recalled in Sec. \ref{FED_Q}, the Dirac equation in a background plane wave described by the four-vector potential $A^{\mu}(\phi)$ can be solved exactly and analytically. Accordingly, the exact electron (Volkov) propagator $G(x,y|A)$ in the same background field has also been determined (see, e.g., \onlinecite{Ritus_1985}). The so-called ``operator technique'', developed in \onlinecite{Baier_1976_a,Baier_1976_b}, turns out to be very convenient for investigating vacuum-polarization effects at high energies (the operator technique for a constant background field was developed in \onlinecite{Schwinger_1951,Baier_1975_a,Baier_1975_b}). In this technique a generic electron state $\Psi(x)$ in a plane-wave field and the propagator $G(x,y|A)$ are intended as the configuration representation of an abstract state $|\Psi\rangle$ and of an operator $G(A)$ such that $\Psi(x)=\langle x|\Psi\rangle$ and $G(x,y|A)=\langle x|G(A)|y\rangle$. In particular, since the propagator $G(x,y|A)$ is the solution of the equation $\{\gamma^{\mu}[i\partial_{\mu}-eA_{\mu}(\phi)]-m\}G(x,y|A)=\delta(x-y)$, then the abstract operator $G(A)$ is simply \begin{equation} G(A)=\frac{1}{\gamma^{\mu}[P_{\mu}-eA_{\mu}(\phi)]-m}, \end{equation} with $P^{\mu}$ being the four-momentum operator. Evaluation via the operator technique of the matrix element corresponding to a generic vacuum-polarization process is then carried out by manipulating abstract operators, which is easier than by working with the corresponding quantities in configuration space. In \onlinecite{Di_Piazza_2007} the operator technique has been employed to calculate the rate of photon splitting in a strong laser field for an incoming photon with four-momentum $k^{\mu}$. This was the first investigation of a QED process involving three Volkov propagators. The calculated rate is valid for an arbitrary plane-wave field, provided that radiative corrections can be neglected, i.e., at $\alpha\varkappa_0^{2/3}\ll 1$, with $\varkappa_0=(k_-/m)(E_0/F_{\text{cr}})$, in the most unfavorable regime $\xi_0,\varkappa_0\gg 1$ \cite{Ritus_1972}. In fact, as we have mentioned in Sec. \ref{FED_Q}, QED processes involving the collision of a photon and an intense plane wave are controlled by the two Lorentz- and gauge-invariant parameters $\xi_0$ and $\varkappa_0$. In \onlinecite{Di_Piazza_2007} it turned out to be more convenient to perform a parametric study of the photon-splitting rate by varying the two parameters $\xi_0$ and $\eta_0=\omega_0k_-/m^2$ (note that $\varkappa_0=\eta_0\xi_0$). By employing the Furry theorem \cite{Landau_b_4_1982}, it is shown that photon splitting in a laser field only occurs with absorption of an odd number of laser photons. In particular, if the strong field is circularly polarized and if it counterpropagates with respect to the incoming photon, conservation of the total angular momentum along the propagation direction of the beams implies that for $\eta_0\ll 1$ photon splitting can occur only via absorption of one or of three laser photons. In \onlinecite{Di_Piazza_2008} a physical scenario has been advanced in which nonperturbative vacuum-polarization effects can be in principle observed (here ``nonperturbative'' means ``high-order'' in the quantum nonlinearity parameter, see below). In this scenario a high-energy proton collides head-on with a strong laser field. The quantum interaction of the Coulomb field of the proton with the laser field allows for a merging of laser photons into a single high-energy photon. The use of a proton, instead of an electron, for example, is required in order to suppress the background process of multiphoton Thomson/Compton scattering, where again many photons of the laser can be directly absorbed by the proton and converted into a single high-energy photon (see Sec. \ref{TCS_General}). In fact, the kinematics of the two processes, vacuum-mediated laser photon merging and multiphoton Thomson/Compton scattering is the same, except that in the treatment in \onlinecite{Di_Piazza_2008} only an even number of laser photons can merge in the vacuum-mediated process. The probability of $\ell$-photon Thomson/Compton scattering of a particle with charge $Q$ and mass $M$ depends on the parameter $\xi_{0,c}=|Q|E_0/M\omega_0$ and scales as $\xi_{0,c}^{2\ell}$ at $\xi_{0,c}\ll 1$. Thus, the use of a ``heavy'' particle like a proton (mass $m_p=1.8\times 10^3\,m=938\;\text{MeV}$) is essential to suppress this background process. Note that in order to have $\xi_{0,p}=|e|E_0/m_p\omega_0\approx 1$ for a proton, a laser field with intensity of the order of $10^{24}\;\text{W/cm$^2$}$ is required. It is found that for an optical background field such that $\xi_0\gg 1$, the amplitude of the $(2\ell)$-photon merging process depends only on the nonlinear quantum parameter \begin{equation} \chi^{(2\ell)}_{0,p}=\frac{E_0}{F_{\text{cr}}}\frac{2\ell(1+v_p)\omega_0}{m}\frac{1-\cos\vartheta}{1+v_p\cos\vartheta}, \end{equation} where $v_p$ is the proton velocity and $\vartheta$ is the angle between the direction of the emitted photon and the propagation direction of the plane wave. By colliding a proton beam, of energy available at the Large Hadron Collider (LHC) of the order of $7\;\text{TeV}$ \cite{LHC} with an optical laser beam of intensity of $3\times 10^{22}\;\text{W/cm$^2$}$, it is demonstrated that the vacuum-mediated merging of two laser photons and the analogous two-photon Thomson/Compton scattering have comparable rates, implying that the inclusive signal should be twice the one expected without vacuum-polarization effects. By accounting for the details of the proton beams available at LHC and of the laser system PFS (see Sec. \ref{L_O}), about 670 two-photon merging events and about 5 four-photon merging events are expected per hour. In the discussed setup most of the photons are emitted almost in the direction of the proton velocity ($\vartheta\approx \pi$) in which $\chi^{(2)}_{0,p}\sim 1$. It is also indicated in \onlinecite{Di_Piazza_2008} that the use of the perturbative expression of the laser-photon merging rate at leading order in $\chi^{(2)}_{0,p}\ll 1$ would lead to an error of about $30\%$. Other setups for observing vacuum-polarization effects in laser-proton collisions have been discussed in \onlinecite{Di_Piazza_2008d}, involving, for example, XFEL or single intense XUV pulses. Finally, the process of Delbr\"{u}ck scattering in a combined Coulomb and laser field has been studied in \onlinecite{Di_Piazza_2008c}. Here an incoming photon is scattered by the Coulomb field of a nucleus and by a strong laser field. While the presence of the laser field is taken into account exactly in the calculations, only leading-order effects in the nuclear parameter $Z\alpha$ are accounted for. Analogously to \onlinecite{Di_Piazza_2008}, it is found that high-order nonlinear corrections in the parameter $\varkappa_0$ to the cross section of the process already become important at $\varkappa_0\approx 0.2$. For example, these corrections amount to about $50\%$ at $\varkappa_0=0.35$. \section{Electron-positron pair production} \label{PP} One of the most important predictions of QED has been the possibility of transforming light into matter \cite{Dirac_1928}. If two photons with four-momenta $k_1^{\mu}$ and $k_2^{\mu}$ collide at an angle such that the parameter $\eta=(k_1k_2)/m^2$ exceeds two, the creation of an $e^+\text{-}e^-$ pair becomes kinematically allowed (Breit-Wheeler $e^+\text{-}e^-$ pair production \cite{Breit_1934}). Shortly after the realization of the laser in 1960, theoreticians started to study possibilities for the creation of $e^+\text{-}e^-$ pairs from vacuum by very strong laser fields \cite{Reiss_1962, Nikishov_1964_a, Yakovlev_1966}. Because of constraints from energy-momentum conservation, a single plane-wave laser field cannot create pairs from vacuum, no matter how intense it is. For a single plane wave, in fact, all the photons propagate along the same direction and the parameter $\eta$ vanishes identically for any pair of photons in the plane wave. Thus, an additional source of energy is therefore required to trigger the process of pair production in a plane wave. There are essentially three different possibilities (see also Fig. \ref{PP_13}): \begin{itemize} \item[(i)] pair production by a high-energy photon propagating in a strong laser field (multiphoton Breit-Wheeler pair production); \item[(ii)] pair production by a Coulomb field in the presence of a strong laser field; \item[(iii)] pair production by two counterpropagating strong laser beams forming a standing light wave. \end{itemize} \begin{figure} \centering \includegraphics[width=0.8\linewidth]{Di_Piazza_fig16.eps} \caption{Feynman diagrams corresponding to processes (i) (part a)), (ii) (part b)) and (iii) (part c)), respectively. The thick continuous lines in parts a) and b) indicate Volkov positive- and negative-energy states. The crossed vertex in part b) stands for the Coulomb electromagnetic field. The diagram in part c) is related to the vacuum current $j_{\text{vac}}^{\mu}(x)$, that one has to determine in order to calculate the $e^+\text{-}e^-$ pair yield \cite{Dittrich_b_2000}. The double line indicates the electron propagator calculated in the Furry picture including exactly the background standing wave.} \label{PP_13} \end{figure} These processes share the common feature to possess different interaction regimes which are mainly characterized by the value of the parameter $\xi_0$. When $\xi_0\ll 1$ the presence of the laser field can be taken into account perturbatively and this yields a pair production rate $R$ of the form $R_{e^+\text{-}e^-}\sim m\xi_0^{2\ell_m}$, where $\ell_m$ is the minimum integer number which kinematically allows the process. For the process (i) it is $\ell_m(k_0k)> 2m^2$, with $k_0^{\mu}$ and $k^{\mu}$ being the four-momentum of the laser photon and the incoming photon, respectively. For the process (ii) it is $\ell_m\omega^{\star}_0>2m$, where $\omega^{\star}_0=\omega_0 u_{c,-}$ is the laser angular frequency in the rest-frame of the charge which produces the Coulomb field and which has four-velocity $u_c^{\mu}$. Finally, for the process (iii) it is $\ell_m\omega_0>2m$. Due to the specific dependence of the pair production rate on $\xi_0$, this regime of pair production is called multiphoton regime. In contrast, when $\xi_0\gg 1$ the presence of the laser field has to be taken into account exactly by performing the calculations in the Furry picture. As the condition $\xi_0\gg 1$ is realized for vanishing laser frequencies at a fixed laser amplitude, this regime is called quasistatic regime and the pair-production rate here is governed by a different parameter which depends on the process at hand. For process (i), for example, the form of the rate depends on the physical parameter $\varkappa_0$ introduced in Sec. \ref{FED_Q} and it scales as $\sim m\varkappa^{3/2}_0\exp(-8/3\varkappa_0)$ if $\varkappa_0\ll 1$ and as $\sim m\varkappa_0^{2/3}$ if $\varkappa_0\gg 1$ \cite{Reiss_1962,Nikishov_1964_a}. For process (ii), we distinguish the case in which the incoming particle is an electron, from that in which it is a heavier particle like a nucleus with charge number $Z$. In the first case the pair production rate depends on the parameter $\chi_0$, already introduced in Sec. \ref{FED_Q}, and the recoil due to the pair creation on the electron has to be taken into account (see also Sec. \ref{PP_L}). In the second case the motion of the nucleus is usually assumed not to be altered by the pair creation process and the nucleus itself is described as a background Coulomb field (see also Sec. \ref{BH}). The pair-production rate depends on the parameter $\chi_{0,n}=u_{n,-}(E_0/F_{\text{cr}})$, with $u^{\mu}_n$ being the four-velocity of the nucleus, and on the nuclear parameter $Z\alpha$. Specifically, the pair-production rate scales as $m(Z\alpha)^2\exp(-2\sqrt{3}/\chi_{0,n})$ if $\chi_{0,n}\ll 1$ and as $m(Z\alpha)^2\chi_{0,n}\ln \chi_{0,n}$ if $\chi_{0,n}\gg 1$ \cite{Yakovlev_1966,Milstein_2006}. Finally, for process (iii), the rate $R_{e^+\text{-}e^-}$ has been derived mainly by approximating the standing wave as an oscillating electric field (see also Sec. \ref{PP_SW}). It is found that $R_{e^+\text{-}e^-}$ depends on the ratio $\Upsilon_0=E_0/F_{\text{cr}}$, with $E_0$ being the amplitude of the standing wave in the (fixed) laboratory frame, and that it scales as $m\Upsilon_0^2\exp(-\pi/\Upsilon_0)$ if $\Upsilon_0\ll 1$ and as $m\Upsilon_0^2$ if $\Upsilon_0\gg 1$ \cite{Brezin_1970,Popov_1971,Popov_1972}. As expected, these scalings coincide with the corresponding ones in a constant electric field $E_0$ \cite{Schwinger_1951}. The physical meaning of the three parameters $\varkappa_0$, $\chi_{0,n}$ and $\Upsilon_0$ can be qualitatively understood in the following way. For process (i) the dressing of the electron and positron mass (see Sec. \ref{FED_Q}) modifies the threshold of $e^+\text{-}e^-$ pair production at $\xi_0\gg 1$ according to $\ell_m(k_0k)\gtrsim 2m^2\xi_0^2$. Now, analogously to multiphoton Thomson and Compton scattering (see Sec. \ref{TCS_General}), the typical number of laser photons absorbed in pair production via photon-laser collision is of the order of $\xi_0^3$ \cite{Nikishov_1964_a} and the threshold condition becomes $\varkappa_0\gtrsim 1$. Concerning the process (ii), the appearance of the parameter $\chi_{0,n}$ in the quasistatic limit $\xi_0\gg 1$ can be understood by noting that the quantity $u_{n,-}E_0$ is the amplitude of the laser field in the rest-frame of the nucleus and that a constant and uniform electric field with strength of the order of $F_{\text{cr}}=m^2/|e|$ supplies a $e^+\text{-}e^-$ pair with its rest energy $2m$ along the typical length scale of QED $\lambda_C=1/m$ (see also Sec. \ref{FED_Q}). This last observation also demonstrates the presence of the parameter $\Upsilon_0$ for process (iii). The typical exponential scaling of the pair production rate for $\xi_0\gg 1$, and at $\varkappa_0\ll 1$, $\chi_{0,n}\ll 1$ and $\Upsilon_0\ll 1$ for processes (i), (ii) and (iii), respectively, is reminiscent of a quantum tunneling process. Thus, one refers to this regime also as tunneling pair production (note, however, that the notion of ``tunneling'' in laser-induced processes should be regarded with special care beyond the dipole approximation \cite{Reiss_2008,Klaiber_2012}). In all processes discussed above, the laser field is always participating directly in the pair creation step and fundamental properties of the quantum vacuum under extreme high-field conditions are probed. However, as will be seen shortly, lying at the border of experimental feasibility, the expected pair yields are generally rather small. It is worth mentioning here that lasers can also be applied for abundant generation of $e^+\text{-}e^-$ pairs \cite{Chen_2009, Chen_2010}. When a solid target is irradiated by an intense laser pulse, a plasma is formed and electrons are accelerated to high energies. They may emit radiation by bremsstrahlung which efficiently converts into $e^+\text{-}e^-$ pairs through the Bethe-Heitler process. The laser field plays an indirect role in the pair production here by serving solely as a particle accelerator. The prolific amount of antimatter generated this way may lead to interesting applications in various fields of science \cite{NaturePhot}. Abundant production of $e^+\text{-}e^-$ pairs and of high-energy photons in the collision of a multipetawatt laser beam and a solid target has been recently investigated in \onlinecite{Nakamura_2011} and \onlinecite{Ridgers_2012}. In particular, in \onlinecite{Nakamura_2011} it has been shown that almost all the laser pulse energy is converted after the collision into a well collimated high-power gamma-ray flash. Whereas, the numerical simulations in \onlinecite{Ridgers_2012} indicate that about $35\;\%$ of the energy of a $10\;\text{PW}$ laser pulse after the laser-target interaction is converted into a gamma-ray burst and that simultaneously a pure $e^+\text{-}e^-$ plasma is produced with a maximum positron density of $10^{26}\;\text{m$^{-3}$}$. \subsection{Pair production in photon-laser and electron-laser collisions} \label{PP_L} Among the pair-production processes mentioned above, only laser-induced pair production for $\xi_0 < 1$ has been observed experimentally. Its feasibility has been shown, in fact, in the pioneering E-144 experiment at SLAC \cite{SLAC_PP,SLAC_PRD} (see \onlinecite{Reiss_1971} for a corresponding theoretical proposal). The experiment relied on collisions of the 46.6 GeV electron beam from SLAC's linear accelerator with a counterpropagating intense laser pulse of photon energy of $\omega_0=2.4\;\text{eV}$ and intensity $1.3\times 10^{18}$\;W/cm$^2$ ($\xi_0\approx 0.4$). In the rest-frame of the electrons, the laser intensity and frequency are largely Doppler up-shifted to the required level and the pair generation probability is effectively enhanced. In principle both reactions (i) and (ii) contribute to pair production in this kind of collisions. Based on separate simulations of both production channels, reaction (i) was found to dominate. The high-energy photon originates from multiphoton Compton backscattering of a laser photon off the electron beam. Despite the significance of the SLAC experiment, a unified description of pair creation in electron-laser collisions has been presented only recently, treating the competing mechanisms (i) and (ii) within the same formalism \cite{Hu_2010}. Good agreement with the experimental results has been obtained. Moreover, it was shown that the SLAC study observed the onset of nonperturbative pair creation dynamics, which adds even further significance to this benchmarking experiment (see also \onlinecite{Reiss_2009}). A formal treatment of the process has also been given in \onlinecite{Ilderton_2011}, where special emphasis is put on effects stemming from the finite duration of the laser pulse. Figure \ref{fig:2} shows a survey of various combinations of incoming electron energies and optical laser intensities which give rise to an observable pair yield. It covers the range from the perturbative few-photon regime ($\xi_0\approx 0.1$ at $I_0\approx 10^{17}\;\text{W/cm$^2$}$) to the highly nonperturbative domain ($\xi_0\approx 10$ at $I_0\approx 10^{21}\;\text{W/cm$^2$}$), where the contributions from thousands of photon absorption channels need to be included. We note that few-GeV electron beams can be produced today using compact laser-plasma accelerators \cite{Leemans_2006} (see also Sec. \ref{LAP}). Future pair creation studies may therefore rely on all-optical setups, where a laser-generated electron beam collides with a counterpropagating laser pulse. Another all-optical setup for pair creation by a seed electron exposed to two counterpropagating laser pulses was put forward in \onlinecite{Bell_2008}, which will be discussed in Sec. \ref{Cascade}. \begin{figure} \centering \includegraphics[width=0.8\linewidth]{Di_Piazza_fig17.eps} \caption{(Color online) Transition from the perturbative to the fully nonperturbative regimes of $e^+\text{-}e^-$ pair creation in electron-laser collisions. The laser photon energy is $2.4\;\text{eV}$. Adapted from \onlinecite{Hu_2010}.} \label{fig:2} \end{figure} Pair creation studies could also be conducted as a nonstandard application of the 17.5 GeV electron beam at the upcoming European XFEL beamline at DESY \cite{XFEL}, which will normally serve to generate coherent x-ray pulses. However, in combination with a table-top 10-TW optical laser system, it would also be very suitable to probe the various regimes of pair production. In particular, the production channel (ii) could be investigated by a suitable choice of beam parameters \cite{Hu_2010}. Other aspects of pair creation by a high-energy photon and a strong laser field have been investigated in recent years. In \onlinecite{Heinzl_2010b}, process (i) was considered in the case where the laser pulse has finite duration. It was found that the finite pulse duration is imprinted on the spectra of created particles. Pair production by a high-energy photon and an ultrashort laser pulse was also considered in \onlinecite{Tuchin_2010}. Quantum interference effects can arise in photon-induced pair creation in a two-mode laser field of commensurate frequencies \cite{Narozhny_2000}. In addition, the fundamental process (i) may allow for applications as a novel tool in ultrashort pulse spectrometry. A corresponding detection scheme for the characterization of short gamma-ray pulses of GeV photons down to the zeptosecond scale, called Streaking at High Energies with Electrons and Positrons (SHEEP), has been proposed in \onlinecite{Ipp_2011}. The basic concept of SHEEP is based on $e^+\text{-}e^-$ pair production in vacuum by a photon of the test pulse, assisted by an auxiliary counter-propagating intense laser pulse. In contrast to conventional streak imaging, two particles with opposite charges, electron and positron, are created in the same relative phase within the third streaking pulse that co-propagates with the test pulse. By measuring simultaneously the energy and momentum of the electrons and the positrons originating from different positions within the test pulse, its length and, in principle, even its shape can be reconstructed. The time resolution of SHEEP for different classes of tests, streaking and strong pulses can range from femtosecond to zeptosecond duration. \subsection{Pair production in nucleus-laser collisions} \label{BH} While in electron-laser collisions the contribution of reaction (ii) to pair production is in general small, it becomes accessible to experimental observation when the projectile electrons are replaced by heavier particles such as protons or other nuclei. The two-step production process via multiphoton Compton scattering will then be strongly suppressed by the large projectile mass. The recent commissioning of the LHC at CERN has stimulated substantial activities on pair production in combined laser and nuclear Coulomb fields, which may be viewed as a generalization of the well-known Bethe-Heitler process to strong fields (multiphoton Bethe-Heitler pair production). The large Lorentz factors $\gamma_n$ of the ultrarelativistic nuclear beams lead to efficient enhancement of the laser parameters in the projectile rest-frame. Indeed, when a proton beam with Lorentz factor $\gamma_p\approx 3000$, as presently available at LHC, collides head-on with a superintense laser beam of intensity $I_0\approx 10^{22}$\;W/cm$^2$, the Lorentz-boosted laser field strength approaches the critical value $F_{\text{cr}}$. This circumstance motivated the first calculations of nonperturbative pair production in collisions of a relativistic nucleus with a superintense near-optical laser beam \cite{MVG2003NIMB,MVG2003PRA}. Smaller projectile Lorentz factors may be sufficient, when ultrastrong XFEL pulses are employed \cite{Avetissian_2003}. The calculations were based on an S-matrix treatment and assumed laser fields of circular polarization. Later on, also the case of linear field polarization was studied \cite{MVG2004PRA,Sieczka_2006, Kaminski_2006, Krajewska_2006}. This case is rendered more involved due to the appearance of generalized Bessel functions, which are of very high order when $\xi_0\gg 1$. The underlying S-matrix element is generally of the form \begin{equation} \label{Sff} S_{p_+,\sigma_+,p_-,\sigma_-} = -ie \int d^4x \Psi^{\dag}_{p_-,\sigma_-}(x)V_n(r) \Psi_{-p_+,-\sigma_+}(x). \end{equation} It describes the transition of an electron from the negative-energy Volkov state $\Psi_{-p_+,-\sigma_+}(x)$ to a positive-energy Volkov state $\Psi^{\dag}_{p_-,\sigma_-}(x)$, which is mediated by the Coulomb potential $V_n(r)=Z|e|/r$ of the projectile nucleus. An alternative approach to the problem based on the polarization operator in a plane electromagnetic wave has been developed in \onlinecite{Milstein_2006}. It allows to obtain total production rates analytically. Both approaches rely on the strong-field approximation and include the laser field exactly to all orders, whereas the nuclear field is treated at leading order in $Z\alpha$. Since the high-intensity Bethe-Heitler process has not been observed in experiment yet, in recent years physicists have proposed scenarios which may allow to realize the various interaction regimes of the process by present-day technology. Few-photon Bethe-Heitler pair production in the perturbative domain could be realized in collisions of the LHC proton beam with an XUV pulse of angular frequency $\omega_X\approx 100$\;eV and of moderate intensity $I_X\sim 10^{14}$\;W/cm$^2$ \cite{Muller2009PLB}. Corresponding radiation sources of table-top dimension are available nowadays in many laboratories. They are based on HHG from atomic gas jets or solid surfaces (see Sec. \ref{HHG}). The rate $R_{e^+\text{-}e^-}$ of pair creation by two-photon absorption close to the energetic threshold (i.e., $\omega^{\star}_X= u_{n,-}\omega_X\gtrsim m$, for the angular frequency $\omega_X^{\star}$ of the XUV pulse in the rest-frame of the nucleus) is given by \cite{Milstein_2006} \begin{equation} \label{BH2lin} R_{e^+\text{-}e^-} = \frac{1}{4^{3-j}}(Z\alpha)^2\xi_0^4\omega_X^{\star}\bigg(\frac{\omega_X^{\star}}{m}-1\bigg)^{j+2}, \end{equation} with $j=0$ for linear polarization and $j=2$ for circular polarization. In the quasistatic regime of the process sizable pair yields require superintense laser fields from a petawatt source in conjunction with an LHC proton beam \cite{MVG2003PRA,Sieczka_2006}. Such experiments will become feasible when petawatt laser pulses are made available by high-power devices of table-top size, rather than by immobile large-scale facilities as they exist at present. A method to enable tunneling pair production with more compact multiterawatt laser systems has been proposed in \onlinecite{DiPiazza_Lotstedt_2009,DiPiazza_Lotstedt_2010}. It relies on the application of an additional weak XUV field, which is superimposed on a powerful optical laser wave. In this two-color setup, the energy threshold for pair creation can be overcome by the absorption of one photon from the high-frequency field and several additional photons from the low-frequency field. As a result, by choosing the XUV frequency $\omega_X^{\star}$ such that the parameter $\delta=(2m - \omega_X^{\star})/m$ fulfills the conditions $0<\delta\ll 1$, the tunneling barrier can be substantially lowered and even controlled. The pair production rate in the quasistatic regime for $0<\delta\ll 1$ depends essentially only on the parameters $\delta$ and $\chi_{0,n}$, and on the classical nonlinearity parameter $\xi_X$ of the XUV field. For a circularly polarized strong laser field it becomes \cite{DiPiazza_Lotstedt_2009} \begin{equation} R_{e^+\text{-}e^-}=\frac{1}{64\sqrt{\pi}}m(Z\alpha)^2\xi_X^2\chi_{0,n}^2\sqrt{\zeta_0}\exp\left(-\frac{2}{3}\frac{1}{\zeta_0} \right) \label{terawatt} \end{equation} for $\zeta_0=\chi_{0,n}/2\delta^{3/2}\ll 1$, to be compared with the usual scaling $\sim(-2\sqrt{3}/\chi_{0,n})$ in the absence of the XUV field. A related process is laser-assisted Bethe-Heitler pair creation, where the high-frequency photon energy satisfies $\omega_X^{\star}>2m$. A pronounced channeling of the $e^+\text{-}e^-$ pair due to the forces exerted by the laser field after their creation was found \cite{Lotstedt_2008,Lotstedt_2009}. Multiphoton Bethe-Heitler pair creation in a two-color laser wave was investigated in \onlinecite{Roshchupkin_2001}. Analytical formulas for positron energy spectra and angular distributions in the tunneling regime of the process were obtained in \onlinecite{KuchievPRA}. For pair production at $\xi_0\sim 1$ no analytical expressions are known because of the intermediate nature of this parameter regime. However, by performing a fitting procedure to numerically obtained results, a total pair production rate scaling as $m(Z\alpha)^2\exp(-3.49/\chi_{0,n})$ was obtained in \onlinecite{AboveThreshold}, which closely resembles the tunneling exponential behaviour $m(Z\alpha)^2\exp(-2\sqrt{3}/\chi_{0,n})$. While in Eq.~(\ref{Sff}) the influence of the projectile is described by an external Coulomb field, the projectile can also be treated as a quantum particle which allows to study nuclear recoil effects \cite{SarahPRD,Krajewska_2010,Krajewska_2011}. Besides, in laser-nucleus collisions, bound-free pair creation can occur where the electron is created in a deeply bound atomic state of the nucleus. The process was studied first for circular laser polarization \cite{MVG2003PRL,Matveev_2005} and later on also for linear polarization \cite{CarlusPRA}, including contributions from the various atomic subshells. \subsection{Pair production in a standing laser wave} \label{PP_SW} Purely light-induced pair production can occur when two noncopropagating laser waves are superimposed. The simplest field configuration consists of two counterpropagating laser pulses of equal frequency and intensity. The resulting field is a standing wave which is inhomogeneous both in space and time and a theoretical treatment of the process is very challenging. In order to render the problem tractable and since the production process mainly occurs where the electric field component of the background field is stronger than the magnetic one \cite{Dittrich_b_2000}, in the standard approach the resulting standing light wave is approximated by a purely electric field oscillating in time. This approximation is expected to be justified for a strong ($I_0> 10^{20}\;\text{W/cm$^2$}$), optical laser field where the typical spatial scale of the field variation $\lambda_0\sim 1\;\text{$\mu$m}$ is much larger than the pair formation length $m/|e|E_0=\lambda_C F_{\text{cr}}/E_0\approx 2.6\times 10^{-2}\;\text{$\mu$m}/\sqrt{I_0[10^{20}\;\text{W/cm$^2$}]}$ \cite{Ritus_1985}. Note the analogy between the formation length $m/|e|E_0$ for pair production and the tunneling length $l_{\text{tun}}\sim I_p/|e|E_0$ in atomic ionization (see Sec. \ref{HHG}), with $I_p$ being the ionization potential energy. Pair production in an oscillating electric field is a generalization of the Schwinger mechanism \cite{Schwinger_1951} to time-dependent fields and it has been considered by many theoreticians, starting from the seminal works \onlinecite{Brezin_1970} and \onlinecite{Popov_1971,Popov_1972} (for a comprehensive list of references until 2005, see \onlinecite{Salamin_2006}). While in the laser-electron and laser-nucleus collisions of the previous subsections the Doppler boost of the laser parameters due to a highly relativistic Lorentz factor could be exploited, in laser-laser collisions this is not possible so that high field strengths $E_0$ or high frequencies $\omega_0$ are required in the laboratory frame. Theoreticians are therefore aiming to find ways for enhancing the pair production probability in order to render the process observable in the foreseeable future. A first possibility to facilitate the observability of pair creation in a standing optical laser wave is to superimpose an x-ray photon (or any other high-frequency component) onto the high-field region \cite{Schutzhold_2008,Dunne_2009,Monin_2010}. In this way, the Schwinger mechanism is catalyzed so that the usual exponential suppression $\sim\exp(-\pi/\Upsilon_0)$ is significantly lowered. For example, in the limit when the x-ray energy approaches the threshold value $2m$, the pair production rate $R_{e^+\text{-}e^-}$ becomes \begin{equation} R_{e^+\text{-}e^-} \sim m\exp\left(-\frac{\pi -2}{\Upsilon_0}\right) \label{catalyzed} \end{equation} assuming that the x-ray propagation direction is perpendicular to the electric field vector of the strong optical field. An overview of the pair production enhancement effect due to the x-ray assistance is shown in Fig.~\ref{Dunne}. \begin{figure} \centering \includegraphics[width=0.8\linewidth]{Di_Piazza_fig18.eps} \caption{(Color) Number of pairs produced by the x-ray assisted Schwinger mechanism for two different values of the x-ray angular frequency $\omega_X$, indicated as $\omega$ in the figure, (blue and black solid lines) and the ratio of these catalyzed pairs to those produced by the standard Schwinger mechanism (blue and black dashed lines), both as functions of the optical field strength in units of $F_{\text{cr}}$. Adapted from \onlinecite{Dunne_2009}.} \label{Dunne} \end{figure} Another proposal to enhance the pair yield is the application of multiple colliding laser pulses instead of only two \cite{Bulanov_2010_a}. It has been demonstrated that the threshold laser energy necessary to produce a single pair, decreases when the number of colliding pulses is increased. The results are summarized in Table~\ref{Bulanov}. Pair production exceeds the threshold when eight laser pulses, with a total energy of 10 kJ, are simultaneously focused on one spot. Doubling (tripling) the number of pulses leads to an enhancement by two (six) orders of magnitude. The threshold energy drops from 40 kJ for two pulses to 5.1 kJ for 24 pulses, clearly indicating that the multiple-pulse geometry is strongly favorable. Besides, it was noticed that the pre-exponential volume factor in the pair creation probability can be very large and partially compensate for the exponential suppression \cite{Narozhny_2006}. \begin{table} \caption{Number $N_{e^+\text{-}e^-}$ of $e^+\text{-}e^-$ pairs produced by different numbers $n$ of laser pulses, with a total energy $W$ of 10 kJ. The threshold value total energy $W_{\rm th}$ needed to produce one $e^+\text{-}e^-$ pair is shown in the third column. The precise collision geometry and the pulse parameters can be found in \onlinecite{Bulanov_2010_a}. Adapted from \onlinecite{Bulanov_2010_a}.} \begin{ruledtabular} \begin{tabular}{ccc} $n$ & $N_{e^+\text{-}e^-}$ at $W=10$\;kJ & $W_{\rm th}$ [kJ]\\ \hline 2 & $< 10^{-18}$ & 40\\ 4 & $< 10^{-8}$ & 20\\ 8 & 4.0 & 10\\ 16 & $1.8\times 10^3$ & 8 \\ 24 & $4.2\times 10^6$ & 5.1\\ \end{tabular} \end{ruledtabular} \label{Bulanov} \end{table} Fine details of pair production in a time-dependent oscillating electric field are being studied nowadays because they might serve as characteristic signatures to discriminate the process of interest from potentially stronger background processes. For example, it was found that the momentum spectrum of the created particles is highly sensitive to a subcycle structure of the field \cite{Hebenstreit_2009} and that in the presence of an alternating-sign time-dependent electric field, coherent interference effects are observed in the Schwinger mechanism \cite{Akkermans_2012}. The observation in \onlinecite{Hebenstreit_2009} found an elegant mathematical explanation via the Stokes phenomenon \cite{Dumlu_PRL2010}. Further effects stemming from the precise shape of the external field were analyzed in \onlinecite{Dumlu_PRD2010,Dumlu_2011}. Also, the oscillating dynamics of the $e^+\text{-}e^-$ plasma created by a uniform electric field, including backreaction effects, was investigated \cite{Ruffini_2010, Ruffini_2011, Apostol_2011} (see also \onlinecite{Kim_2011,Bialynicki-Birula_2011}). In addition to pair creation in superstrong laser pulses of low frequency, the process is also extensively discussed in connection with the upcoming XFEL facilities (see, e.g., \onlinecite{Ringwald_2001,Alkofer_2001}). Here the question arises as to what extent the spatial field dependence may influence the pair creation process, both in terms of total probabilities and particle momentum distributions. According to Noether's theorem, pair production in a time-dependent oscillating electric field occurs with conservation of the total momentum, as well as of the total spin. The problem therefore reduces effectively to a two-level system since the field couples negative and positive-energy electron states of same momentum and spin only. The production process exhibits resonance when the energy gap is an integer multiple of the laser frequency, leading to a characteristic Rabi flopping between the negative and positive-energy Dirac continua \cite{Popov_1971}. Due to the electron dressing by the oscillating field, the resonant laser frequencies are determined by the equation $\ell\omega_0 =2\langle \varepsilon\rangle$, where $\langle \varepsilon\rangle$ is the time-averaged electron energy in the time-dependent oscillating electric field. Accordingly, when the particle momentum is varied, several resonances occur corresponding to different photon numbers $\ell$. This gives rise to a characteristic ring structure in the momentum distribution \cite{Mocken_2010}. \begin{figure} \centering \includegraphics[width=0.8\linewidth]{Di_Piazza_fig19.eps} \caption{(Color online) Probability spectrum of $e^+\text{-}e^-$ pair production in two counterpropagating laser pulses, with the laser magnetic field included (black triangles) and neglected (red crosses). In the first case, the labeling $(\ell_r$-$\ell_l)$ signifies the number of absorbed photons from the right-left propagating wave; in the second case, the peak labels denote the total photon number ($\ell$). A vanishing initial momentum (i.e., positron momentum) and $\xi_0=1$ have been assumed. Adapted from \onlinecite{Ruf_2009}.} \label{resonances} \end{figure} Modifications of these well-established properties of the pair creation process, when the spatial field dependence and, thus, the laser magnetic-field component are accounted for, have been revealed in \onlinecite{Ruf_2009}. Utilizing an advanced computer code for solving the corresponding Dirac equation numerically, it was shown that the positions of the resonances are shifted, several new resonances occur, and the resonance lines are split due to the influence of the spatial field dependence (see Fig.~\ref{resonances}). The basic reason for these effects is that, in contrast to a uniform oscillating electric field, the photons in the counterpropagating laser pulses carry momentum along the beam axis. Therefore not only the total number $\ell$ of absorbed laser photons matters, but also how many of them have originated from the laser pulse travelling to the right and left, respectively. For example, for the multiphoton order $\ell=5$ two different resonance frequencies exist now, corresponding to $\ell_r=3$, $\ell_l=2$ on the one hand, and to $\ell_r=4$, $\ell_l=1$, on the other. Due to the photon momentum, the former two-level scheme is also broken into a $V$-type three-level scheme. This causes a splitting of the resonance lines, in analogy with the Autler-Townes effect known from atomic physics. For another numerical approach to space-time dependent problems in quantum field theory, we refer to the review \onlinecite{Cheng_2010}. Moreover, Schwinger pair production in a space-time dependent electric field pulse has been treated very recently within the Wigner formalism \cite{Hebenstreit_2011}. Here, a self-bunching effect of the created particles in phase space, due to the spatiotemporal structure of the pulse, was found. Finally, we mention that, unlike a plane-wave field, a spatially focused laser beam is capable to produce $e^+\text{-}e^-$ pairs from vacuum and this process has been investigated in \onlinecite{Narozhny_2004} for different field polarizations. Spontaneous pair production may, in principle, also occur in a nuclear field for charge numbers $Z$ exceeding a critical value $Z_c$, which depends on the nuclear model. For example, $Z_c=173$ for a uniformly charged sphere with radius $1.2\times 10^{-12}\;\text{cm}$ \cite{Landau_b_4_1982}. See the reviews \onlinecite{Zeldovich_1972,Baur_2007} for more detailed information also on $e^+\text{-}e^-$ pair production in heavy ion collisions. \subsection{Spin effects and other fundamental aspects of laser-induced pair creation} A particularly interesting aspect of tunneling pair creation is the electron and positron spin-polarization. In general, pronounced spin signatures in a field-induced process may be expected when the background field strength approaches the critical value $F_{\text{cr}}$ \cite{Uggerhoj_2001,Walser_2002}. This indeed coincides with the condition for a sizable yield of tunneling pair production. Studies of spin effects in pair production by a high-energy photon and a strong laser field were performed in \onlinecite{Tsai_1993} and \onlinecite{Serbo_2005}, based on considerations on helicity amplitudes and on the spin-polarization vector, respectively. Characteristic differences between fermionic and bosonic particles have been revealed with respect to pair creation in an oscillating electric field \cite{Popov_1972} and in recent studies of the Klein paradox \cite{Grobe_Spin2004,Grobe_Spin2010,Cheng_2009b}. In the latter case it was shown that the existence of a fermionic (bosonic) particle in the initial state leads to suppression (enhancement) of the pair production probability due to the different quantum statistics. The enhancement in the bosonic case may be even exponential due to an avalanche process \cite{Wagner_Spin2010}. Concerning tunneling pair creation in combined laser and nuclear Coulomb fields, it has been shown that the internal spin-polarization vector is proportional in magnitude to $\chi_{0,n}$ and, to leading order, directed along the transverse momentum component of the electron \cite{DiPiazza_Spin2010}. A helicity analysis of pair production in laser-proton collisions revealed that: 1) right-handed leptons are emitted in the laboratory frame under slightly smaller angles with respect to the proton beam than left-handed ones; 2) the rate of pair creation of spin-$1/2$ particles exceeds by almost one order of magnitude the corresponding quantity for spin-0 particles \cite{Muller_Spin2011}. Other fundamental aspects of laser-induced pair creation have been recently investigated as well. They comprise various kinds of $e^+\text{-}e^-$ correlations \cite{Krekora_2005,Fedorov_2006,Krajewska_2008}, multiple pair creation \cite{Cheng_2009}, questions of locality \cite{Cheng_2008} and vacuum decay times \cite{Labun_2009}, and consistency restrictions on the maximum laser field strength to guarantee the validity of the external-field approximation \cite{Gavrilov_2008}. \section{QED cascades} \label{Cascade} As it was discussed in the previous section, the E-144 experiment at SLAC is the only one, so far, where laser-driven multiphoton $e^+\text{-}e^-$ pair production has been observed. Considering that about 100 positrons have been detected in $22000$ shots, each comprising the collision of about $10^7$ electrons with the laser beam, the process results to be rather inefficient. One could attribute this to the relatively low intensity $I_0$ of the laser system of $1.3\times 10^{18}\;\text{W/cm$^2$}$ ($\xi_0\approx 0.4$ as the laser photon energy was $\omega_0=2.4\;\text{eV}$). However, the extremely high energy $\varepsilon_0$ of the electron beam (about $46.6\;\text{GeV}$) ensured that that the nonlinear quantum parameter $\chi_0$ was about unity ($\chi_0\approx 0.3$). A recent investigation \cite{Sokolov_2010} has pointed out in general that in the mentioned setup, i.e., an electron beam colliding with a strong laser pulse, RR effects prevent the development of a cascade or avalanche process with an efficient, prolific production of $e^+\text{-}e^-$ pairs even at much larger laser intensities such that $\xi_0\gg 1$. By an avalanche or cascade process we mean here a process in which the incoming electrons emit high-energy photons in the laser field, which can interact with the field itself generating $e^+\text{-}e^-$ pairs, which, in turn, emit photons again and so on (of course a cascade process may also be initiated by a photon beam rather than by an electron beam). The above result has been obtained by numerically integrating the kinetic equations, which describe the evolution of the electron, the positron and the photon distributions in a plane-wave background field from a given initial electron distribution, and by accounting for the two basic processes that couple these distributions, i.e., multiphoton Compton scattering and multiphoton Breit-Wheeler pair production. The physical reason why an avalanche process cannot develop in a single plane-wave field can be understood in the following way. In the ultrarelativistic case $\xi_0\gg 1$, the above-mentioned basic processes in a plane-wave field are essentially controlled by the parameters $\chi_0$ and $\varkappa_0$, respectively (see also Secs. \ref{FED_Q}, \ref{TCS_General} and \ref{PP}). Now, at the $j^{\text{th}}$ step in which an electron/positron emits a photon or a photon transforms into an $e^+\text{-}e^-$ pair, the initial quantity $p^{(j)}_{0,-}$ or $k^{(j)}_{0,-}$ is conserved and it is distributed over the two final particles (an electron/positron and a photon in multiphoton Compton scattering and an $e^+\text{-}e^-$ pair in multiphoton Breit-Wheeler pair production). Thus, both resulting particles at each step will have a value of their own parameter $\chi^{\prime (j)}_0=(p^{\prime (j)}_{0,-}/m)(E_0/F_{\text{cr}})$ or $\varkappa^{\prime (j)}_0=(k^{\prime (j)}_{0,-}/m)(E_0/F_{\text{cr}})$ smaller than that of the incoming particle. Moreover, due to the special symmetry of the plane-wave field, the quantities $p^{(j)}_{0,-}$ or $k^{(j)}_{0,-}$ are also rigorously conserved between two steps (see also Sec. \ref{FED_C}). Then, the avalanche ends when the parameters $\chi^{(k)}_{0,i}$ and $\varkappa^{(k)}_{0,i}$ at a certain step $k$ are smaller than unity for $i\in [1,\ldots,N_k]$, with $N_k$ being the number of particles at that step. The question arises as to whether other field configurations exist, where an avalanche process can be efficiently triggered (see the review \onlinecite{Aharonian_2003} for the development of QED cascades in matter, photon gas and magnetic field). A positive answer to this question has first been given in \onlinecite{Bell_2008}: even the presence of a single electron initially at rest in a standing wave generated by two identical counterpropagating circularly polarized laser fields can prime an avalanche process already at field intensities of the order of $10^{24}\;\text{W/cm$^2$}$. We note that in the presence of a single plane wave the same process would require an intensity of the order of $I_{\text{cr}}$, because for an electron initially at rest $\chi_0=E_0/F_{\text{cr}}$. From this point of view, the authors of \onlinecite{Bell_2008} explain qualitatively the advantage of employing two counterpropagating laser beams by means of an analogy taken from accelerator physics: a collision between two particles in their center-of-momenta is much more efficient than if one of the particles is initially at rest, because much more of the initial energy can be transferred, for example, to create new particles. The authors approximated the standing wave by a rotating electric field (see Sec. \ref{PP_SW}). In such a field and for an ultrarelativistic electron the controlling parameter is $\tilde{\chi}_0=(p_{\perp}/m)(E_0/F_{\text{cr}})$, where $p_{\perp}$ is the component of the electron momentum perpendicular to the electric field. By estimating $p_{\perp}\sim m\xi_0$ (see also Eq. \eqref{Free_Sol_1_2}), one obtains $\tilde{\chi}_0\approx I_0[10^{24}\;\text{W/cm$^2$}]/\omega_0[\text{eV}]$ (here $\omega_0$, $E_0$ and $I_0$ are the standing-wave's angular frequency, electric field amplitude and intensity). The investigation in \onlinecite{Bell_2008} is based on the analysis of the trajectory of the electron in the rotating electric field including RR effects via the LL equation. Since the momentum of the electron oscillates around a value of the order of $m\xi_0$, the electron emits high-energy photons efficiently that can in turn trigger the cascade (see Fig. \ref{Bell_Cascade}). The possibility of describing the evolution of the electron via its classical trajectory, can be justified as follows. When an electron interacts with a background electromagnetic field like that of a laser, quantum effects are essentially of two kinds \cite{Baier_b_1998}: the first one is associated with the quantum nature of the electron motion and the second one with the recoil undergone by the electron when it emits a photon. For an ultrarelativistic electron it can be shown that, while the first kind of quantum effects is negligible, the second kind is large and has to be taken into account \cite{Baier_b_1998}. Thus, the basic assumption is that, since the background laser field is strong, the electron is promptly accelerated to ultrarelativistic energies, the motion between two photon emissions is essentially classical and, if necessary, only the emissions have to be treated quantum mechanically by including the photon recoil. On the other hand, photons are assumed to propagate in the field along straight lines. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig20.eps} \end{center} \caption{(Color online) The number of $e^+\text{-}e^-$ pairs ($N_{\pm}$) and the number of photons ($N_{\gamma}$) created by an initial single electron in a rotating electric field as a function of the field intensity. The other plotted quantities are described in \onlinecite{Bell_2008}. From \onlinecite{Bell_2008}.} \label{Bell_Cascade} \end{figure} The model employed in \onlinecite{Bell_2008} was improved in \onlinecite{Kirk_2009} by considering colliding pulsed fields with finite time duration and a realistic representation for the synchrotron spectrum emitted by a relativistic electron. The results in \onlinecite{Bell_2008} were essentially confirmed and numerical simulations with linearly polarized beams have shown a general insensitivity of the cascade development to the polarization of the beams. Another interesting finding in \onlinecite{Kirk_2009} is that the electrons in the standing wave tend to migrate to regions where the electric field vanishes and then they do not contribute to the pair-production process anymore. In both papers \onlinecite{Bell_2008,Kirk_2009} the emission of radiation by the electron was treated classically, i.e., the electron was supposed to loose energy and momentum continuously although in \onlinecite{Kirk_2009} the damping force-term in the LL equation was evaluated by employing the total emitted power calculated quantum mechanically. The stochastic nature of the emission of a photon has been taken into account in \onlinecite{Duclous_2011}. Analogously, the energy of the emitted photon is chosen randomly following the synchrotron spectral distribution and the momentum of the photon is always chosen to be parallel to that of the emitting electron. By contrast, in the pair production process by a photon, since the photon is not deflected by the laser field, it is assumed that after it has propagated one wavelength in the field, it decays into an $e^+\text{-}e^-$ pair. The main result in \onlinecite{Duclous_2011} is that at relatively low intensities of the order of $10^{23}\;\text{W/cm$^2$}$ the pair production rate is increased if the quantum nature of the photon emission is taken into account. The reason is that, due to the stochastic nature of the emission process, the electron can propagate for an unusually large distance before emitting. In this way it may gain an unusually large amount of energy and consequently emit a high-energy photon, that can be more easily converted into an $e^+\text{-}e^-$ pair. Moreover, the discontinuous nature of the (curvature of the) electron trajectory is shown to slow down the tendency of the electrons to migrate to regions where the electric field vanishes. The intensity threshold of the avalanche process in a rotating electric field has also been investigated in \onlinecite{Fedotov_2010}. Denoting by $t_{\text{acc}}$ the time an electron needs to reach an energy corresponding to $\chi_0=1$ starting from rest in the given field, by $t_e$ ($t_{\gamma}$) the electron (photon) lifetime under photon emission ($e^+\text{-}e^-$ pair production) and by $t_{\text{esc}}$ the time after which the electron escapes from the laser field, the authors give the following conditions for the occurrence of the avalanche process: $t_{\text{acc}}\lesssim t_e,t_{\gamma}\ll t_{\text{esc}}$. Estimates based on the classical electron trajectory without including RR effects, lead to the simple condition $E_0\gtrsim \alpha F_{\text{cr}}$ for the avalanche to be primed in an optical field. The above estimate corresponds to an intensity of about $2.5\times 10^{25}\;\text{W/cm$^2$}$, i.e., one order of magnitude larger than what was found in \onlinecite{Bell_2008}. However, the main result of \onlinecite{Fedotov_2010} concerns the limitation, brought about on the maximal laser intensity that can be produced in the discussed field configuration by the starting of the avalanche process. In fact, the energy to accelerate the electrons and the positrons participating in the cascade has to come from the background electromagnetic field. By assuming an exponential increase of the number of electrons and positrons, it is found that already at laser intensities of the order of $10^{26}\;\text{W/cm$^2$}$ the created electrons and positrons have an energy which exceeds the initial total energy of the laser beams. This hints at the fact that at such intensities the colliding laser beams are completely depleted due to the avalanche process. The results obtained from qualitative estimates in \onlinecite{Fedotov_2010} have been scrutinized in \onlinecite{Elkina_2011} by means of more realistic numerical methods based on kinetic or cascade equations. In general, if $f_{\mp}(\bm{r},\bm{p},t)$ ($f_{\gamma}(\bm{r},\bm{k},t)$) is the electron/positron (photon) distribution function (upper and lower sign for electron and positron, respectively) in the phase-space $(\bm{r},\bm{p})$ ($(\bm{r},\bm{k})$) and $\varepsilon=\sqrt{m^2+\bm{p}^2}$ ($\omega=|\bm{k}|$), their evolution in the presence of a given classical electromagnetic field $(\bm{E}(\bm{r},t),\bm{B}(\bm{r},t))$ is described by the kinetic equations \begin{align} \label{f_ep} \begin{split} &\left[\frac{\partial}{\partial t}+\frac{\bm{p}}{\varepsilon}\cdot\frac{\partial }{\partial \bm{r}}\pm \bm{F}_L(\bm{r},\bm{p},t)\cdot\frac{\partial }{\partial \bm{p}}\right]f_{\mp}(\bm{r},\bm{p},t)\\ &=\int d\bm{k}\,w_{\text{rad}}(\bm{r},\bm{p}+\bm{k}\to\bm{k},t)f_{\mp}(\bm{r},\bm{p}+\bm{k},t)\\ &\quad-f_{\mp}(\bm{r},\bm{p},t)\int d\bm{k}\,w_{\text{rad}}(\bm{r},\bm{p}\to\bm{k},t)\\ &\quad+\int d\bm{k}\,w_{\text{cre}}(\bm{r},\bm{k}\to\bm{p},t)f_{\gamma}(\bm{r},\bm{k},t), \end{split}\\ \label{f_gamma} \begin{split} &\left(\frac{\partial}{\partial t}+\frac{\bm{k}}{\omega}\cdot\frac{\partial }{\partial \bm{r}}\right)f_{\gamma}(\bm{r},\bm{p},t)\\ &=\int d\bm{p}\,w_{\text{rad}}(\bm{r},\bm{p}\to\bm{k},t)[f_+(\bm{r},\bm{p},t)+f_-(\bm{r},\bm{p},t)]\\ &\quad-f_{\gamma}(\bm{r},\bm{k},t)\int d\bm{p}\,w_{\text{cre}}(\bm{r},\bm{k}\to\bm{p},t), \end{split} \end{align} where $\bm{F}_L(\bm{r},\bm{p},t)=e[\bm{E}(\bm{r},t)+(\bm{p}/\varepsilon)\times\bm{B}(\bm{r},t)]$ and where $w_{\text{rad}}(\bm{r},\bm{p}\to\bm{k},t)$ ($w_{\text{cre}}(\bm{r},\bm{k}\to\bm{p},t)$) is the local probability per unit time and unit momentum that an electron/positron (photon) with momentum $\bm{p}$ ($\bm{k}$) emits (creates) at the space-time point $(t,\bm{r})$ a photon with momentum $\bm{k}$ (an $e^+\text{-}e^-$ pair with the electron/positron having a momentum $\bm{p}$). It is worth pointing out here a connection between the development of a QED cascade and the quantum description of RR. In fact, as has been discussed in Sec. \ref{RR}, from a quantum point of view, RR corresponds to the multiple recoils experienced by the electron in the incoherent emission of many photons. Thus, in the kinetic approach RR is described by those terms in Eqs. \eqref{f_ep} and \eqref{f_gamma}, which do not involve $e^+\text{-}e^-$ pair production. In fact, in \onlinecite{Elkina_2011} it has been shown in the ultrarelativistic case that the equation of motion for the average momentum of the electron distribution, as derived from Eq. \eqref{f_ep}, coincides with the LL equation in the classical regime $\chi_0\ll 1$ (note that pair production is exponentially suppressed at $\chi_0\ll 1$). In \onlinecite{Elkina_2011} the background field is approximated as a uniform, rotating electric field $\bm{E}(t)$ and $f_{\pm}(\bm{r},\bm{p},t)\to f_{\pm}(\bm{p},t)$ ($f_{\gamma}(\bm{r},\bm{k},t)\to f_{\gamma}(\bm{k},t)$). Analogously to \onlinecite{Duclous_2011}, it is assumed that the momentum of the photon emitted in multiphoton Compton scattering is parallel to that of the emitting electron (positron); in the same way, the momenta of the electron and positron created in multiphoton Breit-Wheeler pair production are assumed to be parallel to that of the creating photon. The evolution of the electron, positron and photon distributions has been investigated by numerically integrating the resulting cascade equations via a Monte Carlo method. Whereas, the instants of radiation and pair production have been randomly generated. In particular, the exponential increase of the number of $e^+\text{-}e^-$ pairs and the qualitative estimate, for example, of the typical energy of the electron at the moment of the photon emission carried out in \onlinecite{Fedotov_2010} have been confirmed (apart from discrepancies within one order of magnitude). In both papers \onlinecite{Fedotov_2010,Elkina_2011} only the case of a rotating electric field was considered. In \onlinecite{Bulanov_2010} it is pointed out that the limitation on the maximal laser intensity reachable before the cascade is triggered, strongly depends on the polarization of the laser beams which create the standing wave. The paradigmatic cases of a rotating electric field and of an oscillating electric field are compared. The estimates presented in \onlinecite{Bulanov_2010} for the case of a rotating electric field essentially confirm that the avalanche starts at laser intensities of the order of $10^{25}\;\text{W/cm$^2$}$. The main physical reason why the cascade process in a circularly polarized standing wave starts at such an intensity is that in a rotating electric field the electron emits photons with typical energies of the order of $0.29\,\omega_0\gamma^3$, i.e., proportional to the cube of the Lorentz factor of the emitting electron $\gamma$ \cite{Bulanov_2010}. Whereas, in an oscillating electric field the typical emitted energy scales as $\gamma^2$, such that in order to radiate a hard photon with a given energy, a much more energetic electron is needed. Hence, the authors of \onlinecite{Bulanov_2010} conclude that in an oscillating electric field RR and quantum effects do not play a fundamental role at laser intensities smaller than $I_{\text{cr}}$ and that avalanche processes do not constitute a limitation. It is crucial however, for the conclusion in \onlinecite{Bulanov_2010} that the collision of the laser beams occurs in vacuum, i.e., the seed electrons and positrons which would trigger the cascade are supposed to be created in the collision itself. The question of the occurrence of the avalanche for two colliding linearly polarized pulses has also been addressed in \onlinecite{Nerush_2011}, where a detailed description of the system under investigation has been provided. In fact, previous models had assumed the background electromagnetic field as given, neglecting in this way the field generated by the electrons and positrons. The approach followed in \onlinecite{Nerush_2011} exploits the existence of two energy scales for the photons: one is that of the external laser field and of the plasma fields which is much smaller than $m$, and the other is that of the photons produced by the high-energy electrons which is, by contrast, much larger than $m$. The evolution of the low-energy photons is described by means of Maxwell's equations which are solved with a PIC code, i.e., the photons are treated as a classical electromagnetic field. Whereas, the production of hard photons as well as the creation of $e^+\text{-}e^{-}$ pairs is described as a stochastic process employing a Monte Carlo method. Unless low-energy ones, hard photons are treated as particles and their evolution is described via a distribution function. It is surprising that by considering a single seed electron initially at rest at a node of the magnetic field of a linearly polarized standing wave, an avalanche process is observed in the numerical simulation already for a laser intensity $I_0=3\times 10^{24}\;\text{W/cm$^2$}$ and a laser wavelength $\lambda_0=0.8\;\text{$\mu$m}$ (see Fig. \ref{Nerush_Cascade}). \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig21.eps} \end{center} \caption{(Color online) Snapshot of the normalized electron density $\rho_e$ (part a)), of the normalized photon density $\rho_{\gamma}$ (part b)), and of the normalized laser intensity $\rho_l$ (part c)) $25.5$ laser periods after the two laser counterpropagating beams collided. The normalized density of positrons is approximately the same as that of the electrons. Also, the intensity of each colliding beam is $I_0=3\times 10^{24}\;\text{W/cm$^2$}$ and the common wavelength is $\lambda_0=0.8\;\text{$\mu$m}$. Adapted from \onlinecite{Nerush_2011}.} \label{Nerush_Cascade} \end{figure} The figure clearly shows the formation of an overdense $e^+\text{-}e^-$ plasma in the central region $|x|\lesssim \lambda_0$. In the same numerical simulations it is found that after about 20 laser periods almost half of the initial energy of the laser field has been transferred to the plasma. Disagreement with the predictions in \onlinecite{Bulanov_2010} is stated to be due to the formation of the avalanche in regions between the nodes of the electric and magnetic field, where the simplified analysis of the electron motion carried out in \onlinecite{Bulanov_2010} is not valid. Further analytical insight into the formation of the cascade has been reported in \onlinecite{Nerush_2011_b} by analyzing approximate solutions of the cascade equations in the presence of a rotating electric field. \section{Muon-antimuon and pion-antipion pair production} \label{MPP} The production of $e^+\text{-}e^-$ pairs in strong laser fields has been discussed in Sec. \ref{PP}. In view of the ongoing technical progress the question arises as to whether also heavier particles such as muon-antimuon ($\mu^+\text{-}\mu^-$) or pion-antipion ($\pi^+\text{-}\pi^-$) pairs can be produced with the emerging near-future laser sources. The production of $\mu^+\text{-}\mu^-$ pairs from vacuum in the tunneling regime appears rather hopeless, though, since the required field needs to be close to $F_{\text{cr},\mu}=\varrho_{\mu}^2F_{\text{cr}}=5.6\times 10^{20}\;{\rm V/cm}$, with the ratio $\varrho_{\mu}=m_\mu/m\approx 207$ between the muon mass $m_{\mu}$ and the electron mass $m$. Even by boosting the effective laser fields with the Lorentz factors ($\sim 10^5$) of the most energetic electron beams available \cite{SLAC_PRD}, the value of $F_{\text{cr},\mu}$ seems out of reach. The tunneling production of $\pi^+\text{-}\pi^-$ pairs is even more difficult as $\varrho_{\pi}=m_{\pi}/m\approx 273$. However, $\mu^+\text{-}\mu^-$ and $\pi^+\text{-}\pi^-$ production can occur in microscopic collision processes in laser-generated or laser-driven plasmas, as well as by few-photon absorption from a high-frequency laser wave. \subsection{Muon-antimuon and pion-antipion pair production in laser-driven collisions in plasmas} \label{ML} Energetic particle collisions in a plasma can in principle drive $\mu^+\text{-}\mu^-$ and $\pi^+\text{-}\pi^-$ production. The plasma may consist either of electrons and ions or of electrons and positrons. Both kinds of plasmas can be produced by intense laser beams interacting with a solid target. With respect to $e^+\text{-}e^-$ plasmas, this has been predicted by \onlinecite{Liang_1998}. As has been mentioned in Sec. \ref{PP}, abundant amounts of $e^+\text{-}e^-$ pairs have been recently produced in this manner at LLNL with pair densities of the order of $10^{16}$\;cm$^{-3}$ \cite{Chen_2009,Chen_2010} and much higher densities of the order of $10^{22}$\;cm$^{-3}$ have been also predicted \cite{Meyer-ter-Vehn_2001}. Theoreticians have therefore started to investigate the properties and time evolution of relativistic $e^+\text{-}e^-$ plasmas \cite{Thoma_RMP,Thoma_EPJD,Kuznetsova_2008,Ruffini_2007,Mustafa_2009,Hu_2011,Kuznetsova_2012}. In particular, it has been shown \cite{Kuznetsova_2008,Thoma_EPJD,Thoma_RMP} that in an $e^+\text{-}e^-$ plasma of 10 MeV temperature, $\mu^+\text{-}\mu^-$ pairs, $\pi^+\text{-}\pi^-$ pairs as well as neutral $\pi^0$ can be created in $e^+\text{-}e^-$ collisions. The required energy stems from the high-energy tails of the thermal distributions. Also cold $e^+\text{-}e^-$ plasmas of high density can be generated nowadays due to dedicated positron accumulation and trapping techniques \cite{Cassidy_2005b}. When such a nonrelativistic low-energy plasma interacts with a superintense laser field, $\mu^+\text{-}\mu^-$ pair production can occur as well \cite{Muller_2006,Muller_2008b}. In this case, the plasma particles acquire the necessary energy by strong coupling to the external field which drives the electrons and positrons into violent collisions. The minimum laser peak intensity to ignite the reaction $e^+e^-\to\mu^+\mu^-$ amounts to about $7\times 10^{22}$\;W/cm$^2$ at a typical optical laser photon energy of $\omega_0=1\;\text{eV}$, corresponding to $\xi_{0,\rm min}=\varrho_{\mu}\approx 207$. The rate $R_{e^+e^-\to\mu^+\mu^-}$ of the process in the presence of a linearly polarized field reads \begin{equation} \label{Rfree} R_{e^+e^-\to\mu^+\mu^-} \approx\frac{1}{2^3\pi^2} \frac{\alpha^2}{m^2\xi_0^4} \sqrt{1-\frac{\xi_{0,\rm min}^2}{\xi_0^2}}\frac{N_+N_-}{V}, \end{equation} with the number $N_{\pm}$ of electrons/positrons and the interaction volume $V$, which is determined by the laser focal spot size. Equation (\ref{Rfree}) may be made intuitively meaningful by introducing the invariant cross section $\sigma_{e^+e^-\to\mu^+\mu^-}$ of $\mu^+\text{-}\mu^-$ production in an $e^+\text{-}e^-$ collision in vacuum \cite{Peskin_1995}: \begin{equation} \label{sigma} \sigma_{e^+e^-\to\mu^+\mu^-}=\frac{4\pi}{3}\frac{\alpha^2}{\varepsilon^{*\,2}}\sqrt{1-\frac{4m_{\mu}^2}{\varepsilon^{*\,2}}}\left(1+\frac{2m_{\mu}^2}{\varepsilon^{*\,2}}\right), \end{equation} where the upper index $^*$ indicates quantities in the center-of-mass system of the colliding electron and positron, as, e.g., the common electron and positron energy $\varepsilon^*$. Now, by exploiting the fact that the quantity $R_{e^+\text{-}e^-}/V$ is a Lorentz invariant and that in the present physical scenario $\varepsilon^*$ can be estimated as $m\xi_0$, Eq. \eqref{Rfree} implies the usual relation $R^*_{e^+e^-\to\mu^+\mu^-}/V^*\sim \sigma_{e^+e^-\to\mu^+\mu^-} n_+^* n_-^*$ between the number of events per unit volume and per unit time $R^*_{e^+e^-\to\mu^+\mu^-}/V^*$ and the cross section $\sigma_{e^+e^-\to\mu^+\mu^-}$; $n_\pm^*=N_\pm/V^*$ denote here the particle densities. The process $e^+e^-\to\mu^+\mu^-$ in the presence of an intense laser wave has also been considered in \onlinecite{Roshchupkin_2009}. In laser-produced electron-ion plasmas resulting from intense laser-solid interactions, $\mu^+\text{-}\mu^-$ and $\pi^+\text{-}\pi^-$ pairs can be generated by the cascade mechanism via energetic bremsstrahlung, like in the case of $e^+\text{-}e^-$ pair production mentioned in Sec. \ref{PP}. Assuming a laser-generated few-GeV electron beam, several hundreds to thousands of $\mu^+\text{-}\mu^-$ pairs arise from bremsstrahlung conversion in a high-$Z$ target material \cite{Kampfer_2009}. The production of $\pi^+\text{-}\pi^-$ pairs by laser-accelerated protons was considered in \onlinecite{Bychenkov_2001}, where a threshold laser intensity of $10^{21}$\;W/cm$^2$ for the process to occur was determined. \subsection{Muon-antimuon and pion-antipion pair production in high-energy XFEL-nucleus collisions} In this subsection another mechanism of $\mu^+\text{-}\mu^-$ and $\pi^+\text{-}\pi^-$ pair creation by laser fields will be pursued, which is based on the collision of an x-ray laser beam with an ultrarelativistic nuclear beam. This setup is similar to the one of Sec. \ref{BH}. In the case of $\mu^+\text{-}\mu^-$ pair creation, by considering an x-ray photon energy of $\omega_0=12$\;keV and a nuclear relativistic Lorentz factor of $\gamma_n=7000$, the photon energy in the rest-frame of the nucleus amounts to $\omega^{\star}_0\approx 2\gamma_n \omega_0 = 168$\;MeV. The energy gap of $2m_\mu$ for $\mu^+\text{-}\mu^-$ pair production can thus be overcome by two-photon absorption \cite{Deneke_2008a,LPHYS2008}. Note that because of pronounced recoil effects, the Lorentz factor which would be required for two-photon $\mu^+\text{-}\mu^-$ production by a projectile electron is much larger: $\gamma\gtrsim 10^6$, corresponding to a currently unavailable electron energy in the TeV range. At first sight, $e^+\text{-}e^-$ and $\mu^+\text{-}\mu^-$ pair production in combined laser and Coulomb fields seem to be very similar processes since the electron and muon only differ by their mass (and lifetime). In this picture, the corresponding production probabilities would coincide when the laser field strength and frequency are scaled in accordance with the mass ratio $\varrho_{\mu}$, i.e., $P_{\mu^+\text{-}\mu^-}(E_{0,\mu},\omega_{0,\mu})=P_{e^+\text{-}e^-}(E_0,\omega_0)$ for $E_{0,\mu}=\varrho_{\mu}^2E_0$ and $\omega_{0,\mu}=\varrho_{\mu}\omega_0$. This simple scaling argument does not apply, however, as the large muon mass is connected with a correspondingly small Compton wavelength $\lambda_{C,\mu}=\lambda_C/\varrho_{\mu}\approx 1.86\;$fm ($1\;\text{fm}=10^{-13}$ cm), which is smaller than the radius of most nuclei. As a result, while the nucleus can be approximately taken as pointlike in $e^+\text{-}e^-$ pair production ($\lambda_C\approx 386\;\text{fm}$), its finite extension must be taken into account in $\mu^+\text{-}\mu^-$ pair production. Pronounced nuclear size effects have also been found for $\mu^+\text{-}\mu^-$ production in relativistic heavy-ion collisions \cite{Baur_2007}. Muon pair creation in XFEL-nucleus collisions can be calculated via the amplitude in Eq. \eqref{Sff}, with the nuclear potential $V_n(r)$ arising from an extended nucleus. It leads to the appearance of a nuclear form factor $F(\mathfrak{q}^2)$ which depends on the recoil momentum $\mathfrak{q}$. For example, $F(\mathfrak{q}^2) = \exp(-\mathfrak{q}^2a^2/6)$ for a Gaussian nuclear charge distribution of root-mean-square radius $a$. Since the typical recoil momentum is $\mathfrak{q}\sim m_\mu$, the form factor leads to substantial suppression of the process. The fully differential production rate $dR_{\mu^+\text{-}\mu^-}=dR^{(\text{el})}_{\mu^+\text{-}\mu^-}+dR^{(\text{inel})}_{\mu^+\text{-}\mu^-}$ may be split into an elastic and an inelastic part, depending on whether the nucleus remains in its ground state or gets excited during the process. They read $dR^{(\text{el})}_{\mu^+\text{-}\mu^-}=dR^{(0)}_{\mu^+\text{-}\mu^-} Z^2 F^2(\mathfrak{q}^2)$ and $dR^{(\text{inel})}_{\mu^+\text{-}\mu^-}\approx dR^{(0)}_{\mu^+\text{-}\mu^-} Z [1-F^2(\mathfrak{q}^2)]$, respectively, with $dR^{(0)}_{\mu^+\text{-}\mu^-}$ being the production rate for a pointlike proton. Figure \ref{MuonTotal} shows total $\mu^+\text{-}\mu^-$ production rates $R^{\star}_{\mu^+\text{-}\mu^-}$ in the rest-frame of the nucleus for several nuclei colliding with an intense XFEL beam. For an extended nucleus, the elastic rate increases with its charge but decreases with its size. This interplay leads to the emergence of maximum elastic rates for medium-heavy ions. Figure \ref{MuonTotal} also implies that the total rate $R_{\mu^+\text{-}\mu^-}=R^{(\text{el})}_{\mu^+\text{-}\mu^-}+R^{(\text{inel})}_{\mu^+\text{-}\mu^-}$ in the laboratory frame saturates at high $Z$ values since $R^{(\text{inel})}_{\mu^+\text{-}\mu^-}$ increases with nuclear charge. For highly-charged nuclei the main contribution stems from the inelastic channel where the protons inside the nucleus act incoherently ($R^{(\text{inel})}_{\mu^+\text{-}\mu^-}\propto Z$). This implies that despite the high charges, high-order Coulomb corrections in $Z\alpha$ are of minor importance. In the collision, also $e^+\text{-}e^-$ pairs are produced by single-photon absorption in the nuclear field. However, this rather strong background process does not deplete the x-ray beam. \begin{figure} \begin{center} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig22.eps} \caption{\label{MuonTotal} Total rates for $\mu^+\text{-}\mu^-$ pair creation by two-photon absorption from an intense XFEL beam ($\omega_0=12$\;keV, $I_0=2.5\times 10^{22}$\;W/cm$^2$) colliding with various ultrarelativistic nuclei ($\gamma_n=7000$). The triangles show elastic rates, whereas the squares indicate total (``elastic + inelastic'') rates. The numerical data are connected by fit curves. The dotted line holds for a pointlike nucleus. The production rates are calculated in the rest-frame of the nucleus. Adapted from \onlinecite{Deneke_2008a}.} \end{center} \end{figure} In XFEL-proton collisions, $\pi^+\text{-}\pi^-$ pairs can be generated as well. A corresponding calculation has been reported in \onlinecite{Anis_2011}, which includes both the electromagnetic and hadronic pion-proton interactions. The latter was described approximately by a phenomenological Yukawa potential. It was shown that, despite the larger pion mass, $\pi^+\text{-}\pi^-$ pair production by two-photon absorption from the XFEL field largely dominates over the corresponding process of $\mu^+\text{-}\mu^-$ pair production in the Doppler-boosted frequency range of $\omega^{\star}_0\approx 150\text{-} 210\;\text{MeV}$. This dominance is due to the much larger strength of the strong (hadronic) force compared with the electromagnetic force. As a consequence, in this energy range $\mu^+\text{-}\mu^-$ pairs are predominantly produced indirectly via two-photon $\pi^+\text{-}\pi^-$ production and subsequent pion decay, $\pi^+\to\mu^+ + \nu_\mu$ and $\pi^-\to\mu^- + {\bar \nu}_\mu$. In relativistic laser-nucleus collisions, $\mu^+\text{-}\mu^-$ or $\pi^+\text{-}\pi^-$ pairs can also be produced indirectly within a two-step process \cite{Kuchiev_2007}. First, upon the collision, an $e^+\text{-}e^-$ pair is created via tunneling pair production. Afterwards, the pair, being still subject to the electromagnetic forces exerted by the laser field, is driven by the field into an energetic $e^+\text{-}e^-$ collision. If the collision energy is large enough, the reaction $e^+e^-\to\mu^+\mu^-$ may be triggered. This two-step mechanism thus represents a combination of the processes considered in Sec. \ref{BH} and \ref{ML}. Besides, it may be considered as a generalization of the well-established analogy between strong-field ionization and pair production (see Sec. \ref{PP}) to include also the recollision step. \section{Nuclear physics} \label{NP} Influencing atomic nuclei with optical laser radiation is, in general, a difficult task because of the large nuclear level spacing $\Delta\mathcal{E}$ of the order of $1\;\text{keV}\text{-} 1\;\text{MeV}$, which exceeds typical laser photon energies $\omega_0\sim 1$ eV by orders of magnitude \cite{Matinyan_1998}. Also the laser's electric work performed over the tiny nuclear extension $r_n\sim 1\text{-} 5$ fm is usually too small to cause any sizable effect. In fact, the requirement $|e|E_0r_n\sim \Delta\mathcal{E}$ can only be satisfied for at laser-field amplitudes at least close to $F_{\text{cr}}$. Direct laser-nucleus interactions have therefore mostly been dismissed in the past. On the other hand, laser-induced secondary reactions in nuclei have been explored especially in the late 1990s. Via laser-heated clusters and laser-produced plasmas, various nuclear reactions have been ignited, such as fission, fusion and neutron production. In all these cases, the interaction of the laser field with the target first produces secondary particles such as photo-electrons or bremsstrahlung photons which, in a subsequent step, trigger the nuclear reaction. For a recent review on this subject we refer to \onlinecite{Ledingham_2003,Ledingham_2010}. In recent years, however, the interest in direct laser-nucleus coupling has been revived by the ongoing technological progress towards laser sources of increasingly high intensities as well as frequencies. Indeed, when suitable nuclear isotopes are considered, intense high-frequency fields or superstrong near-optical fields may be capable of affecting the nuclear structure and dynamics. \subsection{Direct laser-nucleus interaction} \subsubsection{Resonant laser-nucleus coupling} There are several low-lying nuclear transitions in the keV range, and even a few in the eV range. Examples of the latter are $^{229}$Th ($\Delta\mathcal{E}\approx 7.6$ eV) and $^{235}$U ($\Delta\mathcal{E}\approx 76$ eV) \cite{Beck_2007}. These isotopes can be excited by the 5th harmonic of a Ti:Sa laser ($\omega_0=1.55\;\text{eV}$) and by pulses envisaged at the ELI attosecond source \cite{ELI}, respectively. Even higher frequencies can be attained by laser pulse reflection from relativistic flying mirrors of electrons extracted from an underdense plasma \cite{Bulanov_1994} or possibly also from an overdense plasma \cite{Habs_2008}. Otherwise keV-energy photons are generated by XFELs, which, as we have seen in Sec. \ref{x_ray}, are presently emerging as large-scale facilities, e.g., at SLAC \cite{LCLS} and DESY \cite{XFEL}, and which could be employed with focusing \cite{Mimura_2010} and reflection devices \cite{Shvydko_2011}. In addition, also XFEL facilities of table-top size \cite{Gruner_2007,Kneip_2010} and even fully coherent XFEL sources are envisaged such as the future XFEL Oscillator (XFELO) \cite{Kim_2008} or the Seeded XFEL (SXFEL) \cite{Feldhaus_1997,Linac_2011}. Brilliant gamma-ray beams with spectra peaked between 20 KeV and 150 KeV have been recently produced from resonant betatron motion of electrons in a plasma wake \cite{Cipiccia_2011}. A new material research center, the Matter-Radiation Interactions in Extremes (MaRIE) \cite{MaRIE_2011} is planned, allowing for both fully coherent XFEL light with photon energy up to 100 keV and accelerated ion beams. The photonuclear pillar of ELI to be set up near Bucharest (Romania) is planned to provide a compact XFEL along with an ion accelerator aiming for energies of $4\text{-}5$ GeV \cite{ELI}. In addition, coherent gamma-rays reaching few MeV energies via electron laser interaction are envisaged at this facility \cite{ELI,Habs_2009}. With these sources of coherent high-frequency pulses, driving electric dipole (E1) transitions in nuclei is becoming feasible \cite{Buervenich_2006}. Table \ref{nuclei} displays a list of nuclei with suitable E1 transitions. Along with an appropriate moderate nucleus acceleration, resonance may be induced due to the Doppler shift via the factor $(1+v_n)\gamma_n$ in the counterpropagating setup, with $v_n$ and $\gamma_n$ being the nucleus velocity and its Lorentz factor, respectively \cite{Buervenich_2006}. For example, with $^{223}\mathrm{Ra}$ and an XFEL frequency of $12.4$ keV a factor of $(1+v_n)\gamma_n=4$ would be sufficient. In general such moderate pre-accelerations of the nuclei would be of great assistance since they increase the number of possible nuclear transitions for the limited number of available light frequencies. Note that the electric field strength of the laser pulse transforms analogously in the rest-frame of the nucleus, such that the applied laser field in the laboratory frame may correspondingly be weaker for a counterpropagating setup. \begin{table} \caption{The transition energy $\Delta\mathcal{E}$, the dipole moment $\mu$, the life-time $\tau_g$ ($\tau_e$) of the ground (excited) state of few relevant nuclear systems and E1 transitions~\cite{Aas_1999,NuclearDataBase}. Adapted from \onlinecite{Buervenich_2006}.} \begin{ruledtabular} \begin{tabular}{cccccc} Nucleus & Transition & $\Delta\mathcal{E}$[keV] & $\mu$[$e$ fm]& $\tau_g$ & $\tau_e$[ps] \\ \hline $^{153}$Sm & $3/2^- \! \to \! 3/2^+$ & 35.8 & $>0.75$\footnotemark[1] & 47 h & $< 100$ \\ $^{181}$Ta & $9/2^- \! \to \! 7/2^+$ & 6.2 & 0.04\footnotemark[1] & stable & $6\times 10^6$ \\ $^{225}$Ac & $3/2^+ \! \to \! 3/2^-$ & 40.1 & 0.24\footnotemark[1] & 10.0 d & 720 \\ $^{223}$Ra & $3/2^- \! \to \! 3/2^+$ & 50.1 & 0.12 & 11.435 d & 730 \\ $^{227}$Th & $3/2^- \! \to \! 1/2^+$ & 37.9 & \footnotemark[2] & 18.68 d & \footnotemark[2] \\ $^{231}$Th & $5/2^- \! \to \! 5/2^+$ & 186 & 0.017 & 25.52 h & 1030 \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{Estimated via the Einstein A coefficient from $\tau_e$ and $\Delta \mathcal{E}$} \footnotetext[2]{Not listed in the National Nuclear Data Center (NNDC) \cite{NuclearDataBase}} \label{nuclei} \end{table} For optimal coherence properties of the envisaged high-frequency facilities, subsequent pulse applications were shown to yield notable excitation of nuclei \cite{Buervenich_2006}. In addition, many low-energy electric quadrupole (E2) and magnetic dipole (M1) transitions are available. Here it is interesting to note that certain E2 or M1 transitions can indeed be competitive in strength with E1 transitions \cite{Palffy_2008,Palffy2_2008}. While the majority of transitions is available for high frequencies in the MeV domain (requiring substantial nucleus accelerations), Fig. \ref{Palffy1} displays also numerous suitable nuclear transitions below 12.4 keV along with the excitation efficiencies from realistic laser pulses. \begin{figure} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig23.eps} \caption{(Color online) Number $N_{\gamma}$ of signal photons per nucleus per laser pulse for several isotopes with first excited states below 12.4~keV (green squares) and above 12.4~keV (red crosses). The results are plotted versus the atomic number $Z$. The considered European XFEL has a pulse duration of 100~fs and an average brilliance of 1.6$\times 10^{25}$ photons/(s\,mrad$^2$\,mm$^2$\,0.1\% bandwidth) \cite{XFEL}. Adapted from \onlinecite{Palffy_2008}.} \label{Palffy1} \end{figure} While indeed experimental challenges are high, resonant direct interactions of laser radiation with nuclei is expected to pave the way for nuclear quantum optics. Especially control in exciting and deexciting certain long-living nuclear states would have dramatic implications for nuclear isomer research \cite{Walker_1999,Aprahamian_2005,Palffy_2007}. As an obvious application this would be of relevance for nuclear batteries \cite{Walker_1999,Aprahamian_2005}, i.e., for controlled pumping and release of energy stored in long-lived nuclear states. In atomic physics, the STimulated Raman Adiabatic Passage (STIRAP) technique has proven to be highly efficient in controlling populations robustly with high precision \cite{Bergmann_1998}. On the basis of currently envisaged accelerators and coherent high-frequency laser facilities, it has been recently shown that such an efficient coherent population transfer will also be feasible in nuclei \cite{Liao_2011}. Most recently, a nuclear control scheme with optimized pulse shapes and sequences has been developed in \onlinecite{Wong_2011}. Serious challenges are certainly imposed by the nuclear linewidths that may either be too narrow to allow for sufficient interaction with the applied laser pulses or inhibit excitations and coherences due to large spontaneous decay. Decades of research in atomic physics allowing now for shaping atomic spectra via quantum interference \cite{Kiffner_2010,Evers_2002,Postavaru_2011} raise hopes that such obstacles may be overcome in the near future as well. Direct photoexcitation of giant dipole resonances with few-MeV photons via laser-electron interaction was shown to be feasible \cite{Weidenmueller_2011} based on envisaged experimental facilities such as ELI. Finally, care has to be taken to compare the laser-induced nuclear channels with competing nuclear processes via, for example, bound electron transitions or electron captures in the atomic shells \cite{Palffy_2010,Palffy_2007}. \subsubsection{Nonresonant laser-nucleus interactions} \label{NR} Already decades ago a lively debate was started on whether nuclear $\beta$-decay may be significantly affected by the presence of a strong laser pulse or not \cite{Nikishov_1964_b,Reiss1_1983,Becker2_1984,Akhmedov_1983} and a conclusive experimental answer to this issue is still to come. Most recently the notion of affecting nuclear $\alpha$-decay with strong laser pulses has been discussed showing that moderate changes of such nuclear reactions with the strongest envisaged laser pulses are indeed feasible \cite{Castaneda_2011,Castaneda_2012}. When the laser intensity is high enough ($I_0 > 10^{26}$ W/cm$^2$) low-frequency laser fields are able to influence the nuclear structure without necessarily inducing nuclear reactions. In such ultrastrong fields, low-lying nuclear levels get modified by the dynamic (AC-) Stark shift \cite{Buervenich2_2006}. These AC-Stark shifts are of the same order as in typical atomic quantum optical systems relative to the respective transition frequencies. At even higher, supercritical intensities ($I_0 > 10^{29}$ W/cm$^2$) the laser field induces modifications to the proton root-mean-square radius and to the proton density distribution \cite{Buervenich2_2006}. \subsection{Nuclear signatures in laser-driven atomic and molecular dynamics} Muonic atoms represent traditional tools for nuclear spectroscopy by employing atomic physics techniques. Due to the large muon mass compared to that of the electron, $m_\mu\approx 207\, m$, and because of its correspondingly small Bohr radius $a_{B,\mu}=\lambda_{C,\mu}/\alpha\approx 255\;\text{fm}$, the muonic wave-function has a large overlap with the nucleus. Precise measurements of x-ray transitions between stationary muonic states are therefore sensitive to nuclear-structure features such as finite size, deformation, surface thickness, or polarization. When a muonic atom is subjected to a strong laser field, the muon becomes a \emph{dynamic} nuclear probe which is periodically driven across the nucleus by the field. This can be inferred, for example, from the high-harmonic radiation emitted by such systems \cite{Shahbaz_2007,Shahbaz_2010}. Figure~\ref{muonicHHG} compares the HHG spectra from muonic hydrogen versus muonic deuterium subject to a very strong XUV laser field. Such fields are envisaged at the ELI attosecond source (see Fig. \ref{Megalaser}). Due to the different masses $M_n$ of the respective nuclei ($M_n=m_p$ for a hydrogen nucleus and $M_n\approx m_p+m_n$ for a deuterium nucleus by neglecting the binding energy, with $m_{p/n}$ being the proton/neutron mass), muonic hydrogen gives rise to a significantly larger harmonic cut-off energy. The reason can be understood by inspection of the ponderomotive energy \begin{equation} \label{UP} U_{p}=\frac{e^{2}E_0^{2}}{4\omega_0^{2}M_{r}}=\frac{e^{2}E_0^{2}}{4\omega_0^{2}}\left(\frac{1}{m_\mu}+\frac{1}{M_{n}}\right), \end{equation} which depends in the present case on the reduced mass $M_r=m_\mu M_{n}/(m_\mu+ M_{n})$ of the muon-nucleus system. The reduced mass of muonic hydrogen ($\approx 93\;\text{MeV}$) is smaller than that of muonic deuterium ($\approx 98\;\text{MeV}$) and this implies a larger ponderomotive energy and an enlarged \textit{plateau} extension. The influence of the nuclear mass can also be explained by the separated motions of the atomic binding partners. The muon and the nucleus are driven by the laser field into opposite directions along the laser's polarization axis. Upon recombination their kinetic energies sum up as indicated on the right-hand side of Eq. (\ref{UP}). Within this picture, the larger cut-off energy for muonic hydrogen results from the fact that, due to its smaller mass, the proton is more strongly accelerated by the laser field than the deuteron. \begin{figure} \includegraphics[width=0.64\linewidth,angle=270]{Di_Piazza_fig24.eps} \caption{(Color) HHG spectra emitted from muonic hydrogen (black) and muonic deuterium (red) in a laser field of intensity $I_0 \approx 10^{23}$ W/cm$^{2}$ and photon energy $\omega_0\approx 60$ eV. ``Arb. u.'' stands for ``arbitrary units''. From \onlinecite{Shahbaz_2007}.} \label{muonicHHG} \end{figure} Due to the large muon mass, very high harmonic cut-off energies can be achieved via muonic atoms with charge number $Z$ in the nonrelativistic regime of interaction. Since the harmonic-conversion efficiency as well as the density of muonic atom samples are rather low, it is important to maximize the radiative signal strength. A sizable HHG signal requires efficient ionization on the one hand, and efficient recombination on the other. The former is guaranteed if the laser's electric field amplitude $E_0$ lies just below the border of over-barrier ionization, \begin{equation} \label{OBI} E_0\lesssim \frac{M_{r}^{2}}{Q_{\rm eff}}\frac{(Z\alpha)^{3}}{16}. \end{equation} Here, $Q_{\rm eff} = |e|(Z/M_n+1/m_\mu)M_r$ represents an effective charge \cite{Shahbaz_2010,Reiss_1979}. Efficient recollision is guaranteed if the magnetic drift along the laser propagation direction can be neglected. Equation \eqref{drift} indicates that this is the case here, provided \begin{equation} \label{RP} \left(\frac{Q_{\rm eff}E_0}{M_{r}\omega_0}\right)^3\lesssim \frac{16\omega_0}{\sqrt{2M_{r}I_p}}. \end{equation} The two above inequalities define a maximum laser intensity and a minimum laser angular frequency which are still in accordance with the conditions imposed. At these laser parameters, the maximum harmonic cut-off energies are attained and an efficient ionization-recollision process is guaranteed. For muonic hydrogen the corresponding lowest frequency lies in the Vacuum Ultraviolet (VUV) range ($\omega_0\approx 27$\;eV) and the maximum field intensity amounts to $1.6\times 10^{23}$\;W/cm$^{2}$. At these values, the harmonic spectrum extends to a maximum energy of approximately $0.55$ MeV. For light muonic atoms with nuclear charge number $Z>1$, the achievable cut-off energies are even higher, reaching several MeVs. This holds prospects for the production of coherent ultrashort gamma-ray pulses (see also \onlinecite{Xiang_2010}). In principle, also nuclear size effects arise in the HHG signal from muonic atoms. This has been shown qualitatively in 1D numerical simulations, where a 50\% enhancement of the harmonic plateau height has been obtained for muonic hydrogen compared with muonic deuterium \cite{Shahbaz_2007,Shahbaz_2010}. This has been attributed to the enhanced final muon acceleration towards the hydrogenic core. For more precise predictions, 3D calculations are desirable. Existing results also indicate that muonic atoms in high-intensity, high-frequency laser fields can be utilized to dynamically gain structure information on nuclear ground states via their high-harmonic response. Besides, the laser-driven muonic charge cloud, causing a time-dependent Coulomb field, may lead to nuclear excitation \cite{Shahbaz_2009}. The excitation probabilities are quite small, however, because of the large difference between the laser photon energy and the nuclear transition energy. Nuclear excitation can also be triggered by intense laser-induced recollisions of field-ionized high-energy electrons \cite{Mocken_2004,Milosevic_2004,Kornev_2007}. Finally, muonic molecules are of particular interest for nuclear fusion studies. Modifications of muon-catalyzed fusion in strongly laser-driven muonic D$_2^+$ molecules have been investigated \cite{Chelkowski_2004,Paramonov_2007}. It was found that applied field intensities of the order of $10^{23}$\;W/cm$^{2}$ can control the molecular recollision dynamics by triggering the nuclear reaction on a femtosecond time scale. Similar theoretical studies have recently been carried out on aligned (electronic) HT molecules, i.e., involving a Tritium atom \cite{Zhi_2009}. \section{Laser colliders} \label{LAP} The fast advancement in laser technology is opening up the possibility of employing intense laser beams to efficiently accelerate charged particles and to make them collide for eventually even initiating high-energy reactions. \subsection{Laser acceleration} \label{LA} Strong laser fields provide new mechanisms for particle acceleration alternative to conventional accelerator technology \cite{Tajima_1979}. With presently-available laser systems an enormous electron acceleration gradient $\sim 1$ GeV/cm can be achieved, which exceeds by three orders of magnitude that of conventional accelerators and which raises prospects for compact accelerators \cite{Mangles_2004,Geddes_2004,Faure_2004,Clayton_2010,Hafz_2008}. Different schemes of laser-electron acceleration have been proposed. These include the laser wakefield accelerator, the plasma beat wave accelerator, the self-modulated laser wakefield accelerator and plasma waves driven by multiple laser pulses (see the recent reviews \onlinecite{Esarey_2009,Malka_2011}). High-gradient plasma wakefields can also be generated with an ultrashort bunch of protons \cite{Caldwell_2009}, allowing electron acceleration to TeV energies in a single stage. The achievable current and emittance of presently-available laser-accelerated electron beams is sufficient to build synchrotron radiation sources or even to aim at compact XFEL lasers \cite{Schlenvoigt_2008}. Laser acceleration of ions provides quasimonoenergetic beams with energy of several MeVs per nucleon \cite{Hegelich_2006,Schwoerer_2006,Toncian_2006,Fuchs_2007,Haberberger_2012}. It mostly employs the interaction of high-intensity lasers with solid targets. One of the main goals of laser-ion acceleration is to create low-cost devices for medical applications, such as for hadron cancer therapy \cite{Combs_2009}. Several regimes have been identified for laser-ion acceleration (see also the forthcoming review \onlinecite{Macchi_2012}). For laser intensities in the range $10^{18}\text{-} 10^{21}$ W/cm$^2$ and for solid targets with a thickness ranging from a few to tens of micrometers, the so-called target-normal-sheath acceleration is the main mechanism \cite{Fuchs_2006}. A further laser-ion interaction process is the skin-layer ponderomotive acceleration \cite{Badziak_2007}. By contrast, the radiation-pressure acceleration regime operates when the target thickness is decreased (see \onlinecite{Esirkepov_2004} and \onlinecite{Macchi_2009} for the so-called ``laser piston'' and ``light sail'' regimes, respectively). In \onlinecite{Galow_2011} a chirped ultrastrong laser pulse is applied to proton acceleration in a plasma. Chirping of the laser pulse ensures optimal phase synchronization of the protons with the laser field and leads to efficient proton energy gain from the field. In this way, a dense proton beam (with about $10^7$ protons per bunch) of high energy (250 MeV) and good quality (energy spread $\sim 1\%$) can be generated. An alternative promising way for particle acceleration is direct laser acceleration in a tightly focused laser beam \cite{Salamin_2002a} or in crossed laser beams \cite{Salamin_2003}. Especially efficient accelerations \cite{Salamin_2006b,Gupta_2007,Bochkarev_2011} can be achieved in a radially polarized axicon laser beam \cite{Dorn_2003}. For example, the generation of mono-energetic GeV electrons from ionization in a radially polarized laser beam is theoretically demonstrated in \onlinecite{Salamin_2007b,Salamin_2010}. A setup for direct laser acceleration of protons and bare carbon nuclei is considered in \onlinecite{Salamin_2008}. It has been shown that laser pulses of $0.1\text{-} 10$ PW can accelerate the nuclei directly to energies in the range required for hadron therapy. Simulations in \onlinecite{Galow_2010} and further optimization studies in \onlinecite{Harman_2011} indicate that protons stemming from laser-plasma processes can be efficiently post-accelerated employing single and crossed pulsed laser beams, focused to spot radii of the order of the laser wavelength. The protons in the resulting beam have kinetic energies exceeding 200 MeV and small energy spreads of about 1\%. The direct-acceleration method has proved to be efficient also for other applications. In \onlinecite{Salamin_2011} it is shown that 10 keV helium and carbon ions, injected into 1 TW-power crossed laser beams of radial polarization, can be accelerated in vacuum to energies of hundreds of keV necessary for ion lithography. \subsection{Laser-plasma linear collider} Laser-electron accelerators have already entered the GeV energy domain where the realm of particle physics starts \cite{Leemans_2006}. In fact, the strong interaction comes into play at distances $d$ of the order of $d \sim 1$ fm, which for relativistic processes corresponds to energies $\varepsilon \sim 1/d \sim 1$ GeV. Thus, a laser-based collider is in principle suitable for performing particle-physics experiments. However, in order to initiate high-energy reactions with sizable yield, not only GeV energies are required but also collision luminosities ${\mathcal L}$ at least as high as $10^{26}\text{-} 10^{27}$ cm$^{-2}$s$^{-1}$. Meanwhile, for the ultimate goal of being competitive with the next International Linear Collider \cite{ILC}, energies on the order of 1 TeV and luminosities of the order of $10^{34}$ cm$^{-2}$s$^{-1}$ are required \cite{Ellis_2001}. The potential of the Laser-Plasma Accelerator (LPA) scheme to develop a laser-plasma linear collider is discussed in \onlinecite{Schroeder_2010}. Two LPA regimes are analyzed which are distinguished by the relationship between the laser beam waist size $w_0$ and the plasma frequency $\omega_p=\sqrt{4\pi n_pe^2/m}$, with $n_p$ being the plasma density: 1) the quasilinear regime at large radius of the laser beam $\omega_p^2w_0^2>2\xi_0^2/\sqrt{1+\xi_0^2/2}$ ($\xi_0\sim 1$); 2) the bubble regime, at $\omega_p w_0\lesssim 2\sqrt{\xi_0}$ ($\xi_0> 1$). The latter is less suitable for the collider application because the bubble cavity leads to defocusing for positrons; besides, the focusing forces and accelerating forces are not independently controllable here as both depend on the plasma density. Whereas, in the quasilinear regime this is possible due to the existence of a second control parameter given by the laser beam waist size. In the following, we will discuss qualitatively the scaling properties of the quasilinear regime. For more accurate expressions the reader is referred to \onlinecite{Schroeder_2010}. In the standard LPA scheme, the electron plasma wave is driven by an intense laser pulse with duration $\tau_0$ of the order of the plasma wavelength $\lambda_p=2\pi/\omega_p$, which accelerates the electrons injected in the plasma wave by wave breaking \cite{Malka_2008,Esarey_2009,Leemans_2009}. The accelerating field $E_p$ of the plasma wave can be estimated from \begin{equation} E_p\sim \frac{m\omega_p}{|e|}\propto n_p^{1/2}. \label{wakefield_E} \end{equation} In fact, in the plasma wave, the charge separation occurs on a length scale of the order of $\lambda_p$, producing a surface charge density $\sigma_p\sim |e|n_p\lambda_p$ and a field $E_p\sim 4\pi \sigma_p$, which corresponds to Eq. \eqref{wakefield_E}. The number of electrons $N_e$ that can be accelerated in a plasma wave is given approximately by the number of charged particles required to compensate for the laser-excited wakefield, having a longitudinal component $E_{\parallel}$. From the relation $E_{\parallel}\sim 4\pi N_e|e|/\pi w_0^2$, it follows that \begin{equation} N_e\sim \frac{\pi n_p}{\omega_p^3}\propto n_p^{-1/2}, \label{wakefield_N} \end{equation} because $w_0\omega_p\sim 1$ in the quasilinear regime. The interaction length of a single LPA stage is limited by laser-diffraction effects, dephasing of the electrons with respect to the accelerating field, and laser-energy depletion. Laser-diffraction effects can be reduced by employing a plasma channel, and plasma-density tapering can be utilized to prevent dephasing. Therefore, the LPA interaction length will be determined by the energy depletion length $L_d$. We can estimate the latter by equating the energy spent for accelerating the $N_e$ electrons along $L_d$ ($\sim N_e|e|E_pL_d$) to the energy of the laser pulse ($\sim E_0^2\pi w_0^2\tau_0/8\pi$). Recalling that $w_0\omega_p\sim 1$, this yields \begin{equation} L_d\sim \frac{\omega_0^2}{\omega_p^2}\lambda_p\propto n_p^{-3/2}. \label{wakefield_Ld} \end{equation} A staging of LPA is required to achieve high current densities along with high energies. The electron energy gain $\Delta\varepsilon_s$ in a single-stage LPA is \begin{equation} \Delta\varepsilon_s\sim |e|E_pL_d\sim m\frac{\omega_0^2}{\omega_p^2}\propto n_p^{-1}. \label{wakefield_Ws} \end{equation} Therefore, the number of stages $N_s$ to achieve a total acceleration energy $\varepsilon_0$ is $N_s=\varepsilon_0/ \Delta\varepsilon_s\sim (\varepsilon_0/m)(\omega_p/\omega_0)^2\propto n_p$. It corresponds to a total collider length $L_c$ of \begin{equation} L_c\sim N_sL_d\sim \frac{\varepsilon_0}{m}\lambda_p\propto n_p^{-1/2}. \label{wakefield_Ns} \end{equation} When two identical beams each with $N$ particles and with horizontal (vertical) transverse beam size $\sigma_x$ ($\sigma_y$) collide with a frequency $f$, the luminosity $\mathcal{L}$ is defined as $\mathcal{L}=N^2f/4\pi\sigma_x\sigma_y$. In LPA $N=N_e$ and $f$ is the laser repetition rate, then \begin{equation} {\mathcal L}\sim\frac{1}{64\pi}\frac{1}{\omega_p^2\sigma_x\sigma_y}\frac{f}{r_0^2}. \label{wakefield_Lum} \end{equation} The laser energy $W_s$ required in a single stage in the LPA collider is approximately given by $W_s\sim m(\lambda_p/r_0)(\omega_0/\omega_p)^2$ at $\xi_0\sim 1$ and the total required power $P_T$ amounts to $P_T \sim N_sW_sf\sim \varepsilon_0f(\lambda_p/r_0)$. The above estimates show that, although the number $N_e$ of electrons in the bunch as well as the single-stage energy gain $\Delta\varepsilon_s$ increase at low plasma densities, the accelerating gradient $\Delta\varepsilon_s/L_d$ nevertheless decreases because the laser depletion length $L_d$ and the overall collider length $L_c$ increase as well. Limiting the total length of each LPA in a collider to about 100 m will require a plasma density $n_p\sim 10^{17}$ cm$^{-3}$ to provide a center-of mass energy of $\sim 1$ TeV for electrons, with $\sim 10$ GeV energy gain per stage \cite{Schroeder_2010}. For a number of electrons per bunch of $N_e\sim 10^9$, a laser repetition rate of $15\;\text{kHz}$ and a transverse beam size of about $10\;\text{nm}$ would be required to reach the goal-value of $10^{34}$ cm$^{-2}$s$^{-1}$ for the accelerator luminosity. At the usual condition $w_0\sim 1/\omega_p\sim 10\;\text{$\mu$m}$, instead, the luminosity amounts to ${\mathcal L}\sim 10^{27}$ cm$^{-2}$s$^{-1}$. In the above conditions each acceleration stage would be powered by a laser pulse with an energy of $30$ J corresponding to an average power of about $0.5\;\text{MW}$ at the required repetition rate of $15\;\text{kHz}$. Such high-average powers are beyond the performance of present-day lasers. The future hopes for high-average-power lasers are connected with diode-pumped lasers and new amplifier materials. Further challenges of LPA colliders are their complexity, as they involve plasma channels and density tapering, and, most importantly, the problem of how to accelerate positrons. \subsection{Laser micro-collider} \label{MC} We turn now to another scheme for a laser collider, which is based on principles quite different than those of the LPA scheme discussed above. In the LPA scheme, the electron is accelerated due to its synchronous motion with the propagating field. In this way the symmetry in the energy exchange process between the electron and the oscillating field is broken, as required by the Lawson-Woodward theorem \cite{Woodward_1947,Lawson_1979}. Another way to exploit the energy gain of the electron in the oscillating laser field is to initiate high-energy processes \textit{in situ}, i.e., inside the laser beam \cite{McDonald_1999}. In this case the temporary energy gain of the electron during interaction with a half cycle of the laser wave is used to trigger some processes during which the electron state may change (in particular, the electron may annihilate with a positron) and the desired asymmetry in the energy exchange can be achieved. In fact, this approach is widely employed in the nonrelativistic regime via laser-driven recollisions of an ionized electron with its parent ion (see Sec. \ref{HHG}). The question arises also as to whether the temporary energy gain of the electron in the laser beam can also be employed in the relativistic regime at ultrahigh energies. As pointed out in Sec. \ref{RD}, an extension of the established recollision scheme with normal atoms into the relativistic regime is hindered by the relativistic drift. However, the drift will not cause any problem when positronium (Ps) atoms are used because its constituent particles, electron and positron, have the same absolute value of the charge-to-mass ratio (see Sec. \ref{FED_C}). \begin{figure} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig25.eps} \caption{(a) In conventional $e^+\text{-}e^-$ colliders bunches of accelerated electrons and positrons are focused to collide head-on-head {\it incoherently}, i.e., the bunches collide head-on-head but electrons and positrons in the bunch do not. (b) In the recollision-based collider, the electron and positron originating from the same Ps atom may collide head-on-head {\it coherently} \cite{Henrich_2004}. From \onlinecite{Hatsagortsyan_2006}.} \label{microcollis} \end{figure} A corresponding realization of high-energy $e^+\text{-}e^-$ recollisions in the GeV domain aiming at particle reactions has been proposed in \onlinecite{Hatsagortsyan_2006}. It relies on (initially nonrelativistic) Ps atoms exposed to super-intense laser pulses. After almost instantaneous ionization of Ps in the strong laser field, the free electron and positron oscillate in opposite directions along the laser electric field and experience the same ponderomotive drift motion along the laser propagation direction. In this way, the particles acquire energy from the field and are driven into periodic $e^+\text{-}e^-$ collisions \cite{Henrich_2004}. Provided that the applied laser intensity is large enough, elementary particle reactions like heavy lepton-pair production can be induced in these recollisions. The common center-of-mass energy $\varepsilon^*$ of the electron and the positron at the recollision time arises mainly from the transversal momentum of the particles and it scales as $\varepsilon^*\sim m\xi_0$. A basic particle reaction which could be triggered in a laser-driven collider is $e^+\text{-}e^-$ annihilation with production of a $\mu^+\text{-}\mu^-$ pair, i.e., $e^+e^-\to\mu^+\mu^-$. The energy threshold for this process in the center-of-mass system is $2m_{\mu}\approx 210\,$MeV. It can be reached with a laser field such that $\xi_0 \sim 200$, corresponding to laser intensities of the order of $10^{22}$ W/cm$^2$, currently within reach. In addition, the proposed recollision-based laser collider can yield high luminosities compared to conventional laser accelerators. In the latter, bunches of electrons and positrons are accelerated and brought into head-on-head collision. However, the particles in the bunch are distributed randomly such that each microscopic $e^+\text{-}e^-$ collision is not head-on-head but has a mean impact parameter $b_i\sim a_b$ determined by the beam radius $a_b$, characterizing the collision as {\it incoherent} (see Fig. \ref{microcollis}). Instead, in the recollision-based collider the electron and the positron stem from the same Ps atom with initial coordinates being confined within the range of one Bohr radius $a_B\approx 5.3\times 10^{-9}\;\text{cm}$. Since they are driven coherently by the laser field, they can recollide with a mean impact parameter $b_c\sim a_{wp}$ of the order of the electron wave packet size $a_{wp}$ (see Fig. \ref{microcollis}). Consequently, the luminosity contains a {\it coherent} component \cite{Hatsagortsyan_2006}: \begin{equation} \mathcal{L}=\left[\frac{N_p(N_p-1)}{b_i^2}+\frac{N_p}{b_c^2}\right]f, \label{L} \end{equation} where $N_p$ is the number of particles in the bunch, and $f$ is the bunch repetition frequency. The coherent component $(N_p/b_c^2)f$ can lead to a substantial luminosity enhancement in the case when the particle number is low and the particle's wave packet spreading is small, $N_pa_{wp}^2<a_b^2$. Note that the reaction ${\rm Ps}\to\mu^+\mu^-$ arising in a strongly laser-driven $e^+\text{-}e^-$ plasma may be considered as the coherent counterpart of the incoherent process $e^+e^-\to\mu^+\mu^-$ discussed in Sec. \ref{ML}. Rigorous quantum-electrodynamical calculations for $\mu^+\text{-}\mu^-$ pair production in a laser field have been performed in \onlinecite{Muller_2006,Muller_2008a,Muller_2008b}. In agreement with Eq.~(\ref{L}), they enabled the development of a simple-man's model in which the rate of the laser-driven process can be expressed via a convolution of the rescattering electron wave packet with the field-free cross section $\sigma_{e^+e^-\to\mu^+\mu^-}$ (see Eq. \ref{sigma}). The latter attains the maximal value $\sigma^{(\text{max})}_{e^+e^-\to\mu^+\mu^-}\sim \alpha^2\lambda_{C,\mu}^2\sim 10^{-30}\;\text{cm$^2$}$ at $\varepsilon^*\approx 260\;\text{MeV}$ \cite{Peskin_1995}. However, when the field driving the Ps atoms is a single laser wave, the $e^+\text{-}e^-$ recollision times are long and the $\mu^+\text{-}\mu^-$ production process is substantially suppressed by extensive wave packet spreading. This obstacle can be overcome when two counterpropagating laser beams are employed. The role of the spreading of the electron wave packet in counterpropagating focused laser beams of circular and linear polarization has been investigated in detail in \onlinecite{Liu_2009}. The advantage of the circular-polarization setup is the focusing of the recolliding electron wave packet. However, this advantage is reduced by a spatial offset in the $e^+\text{-}e^-$ collision when the initial coordinate of the Ps atom deviates from the symmetric position between the laser pulses. The latter imposes a severe restriction on the Ps gas size along the laser propagation direction. Thus, the linear-polarization setup is preferable when the offset at the recollision is very small and the wave packet size at the recollision is within acceptable limits. Results from a Monte Carlo simulation of the $e^+\text{-}e^-$ wave-packet dynamics in counterpropagating linearly polarized laser pulses are shown in Fig. \ref{MCcrossed}. \begin{figure} \includegraphics[width=0.8\linewidth]{Di_Piazza_fig26.eps} \caption{(Color) The coordinate-space distributions of the electron and the positron wave-packets at the recollision time in focused counterpropagating pulses along the $z$ direction with $w_0 =10\;\mu$m, $\lambda_0=0.8\;\mu $m and with $I_0 = 4.7\times10^{22}$W/cm$^2$ (parts (a) and (c)) and $I_0 = 1.4\times10^{23}$W/cm$^2$ (parts (b) and (d)). The Ps atom is initially located at the origin. Spatial coordinates are given in ``atomic units'', with $1\;\text{a.u.}=0.05\;\text{nm}$. Adapted from \onlinecite{Liu_2009}.} \label{MCcrossed} \end{figure} The luminosity ${\mathcal L}$ and the number of reaction events ${\mathcal N}$ for the recollision-based collider with counter-propagating laser pulses can be estimated as: \begin{align} {\mathcal L} \sim& N_{\text{Ps}}\frac{1}{a_{wp}^3 }\tau_r f,\\ {\mathcal N} \sim& \frac{\sigma^{(\text{max})}_{e^+e^-\to\mu^+\mu^-}}{a_{wp}^3}\tau_r N_{\text{Ps}}N_L, \label{Ps_Lum} \end{align} respectively, where $N_{\text{Ps}}$ is the number of Ps atoms, $\tau_r$ the recollision time of the order of the lasers period, $N_L$ the number of laser pulses, and $f$ the laser repetition rate. Taking $N_{\text{Ps}}\approx 10^8$ \cite{Cassidy_2005}, $f=1$ Hz and the spatial extension of the $e^+\text{-}e^-$ pair from Fig. \ref{MCcrossed}, one estimates a luminosity of ${\mathcal L}\sim 10^{27}$ cm$^{-2}$s$^{-1}$ and about one $\mu^+\text{-}\mu^-$ pair production event every $10^3$ laser shots at a laser intensity of $4.7\times 10^{22}$ W/cm$^2$. In conclusion, the scheme of the recollision-based laser collider allows to realize high-energy and high-luminosity collisions in a microscopic setup. However, it is not easily scalable to the parameters of the ILC, namely, to TeV energies and luminosities of the order of $10^{34}$ cm$^{-2}$s$^{-1}$. \section{Particle physics within and beyond the Standard Model} \label{Part_Phys} The sustained progress in laser technology towards higher and higher field intensities raises the question as to what extent ultrastrong laser fields may develop into a useful tool for particle physics beyond QED. Below we review theoretical predictions regarding the influence of super-intense laser waves on electroweak processes and their potential for probing new physics beyond the Standard Model. \subsection{Electroweak sector of the Standard Model} The energy scale of weak interactions is set by the masses of the $W^\pm$ and $Z^0$ exchange bosons, $m_W\approx m_Z\sim 100$\,GeV. Therefore, the influence of external laser fields, even if strong on the scale of QED, is generally rather small. An overview of weak interaction processes in the presence of intense electromagnetic fields has been given in \onlinecite{Kurilin_1999}. Various weak decay processes in the presence of intense laser fields have been considered. They can be divided into two classes: 1) laser-assisted processes which also exist in the absence of the field but may be modified due to its presence; 2) field-induced processes which can only proceed when a background field is present, providing an additional energy reservoir. With respect to processes from the first category, $\pi\to\mu +\nu$ and $\mu^- \to e^- + \bar{\nu}_e + \nu_\mu$ have already been examined \cite{Ritus_1985}. Laser-assisted muon decay has also been revisited recently \cite{Narozhny_2008, Dicus_2009, Farzinnia_2009}. $W^\pm$ and $Z^0$ boson decay into a fermion-antifermion pair was calculated in \onlinecite{Kurilin_2004, Kurilin_2009}. In all cases, the effect of the laser field was found to be small. As a general result, the presence of the laser field modifies the field-free decay rate $R_{M,0}$ of a particle with mass $M$ to $R_M = R_{0,M}(1+\Delta)$, with the correction $\Delta$ being of the order of $\chi_{0,M}^2$ in the range of parameters $\xi_{0,M}=\xi_0 m/M > 1$ and $\chi_{0,M}=\chi_0(m/M)^3\ll 1$. The elastic scattering of a muon neutrino and an electron in the presence of a strong laser field has been considered in \onlinecite{Bai_2012} and multiphoton effects in the cross section are predicted. External fields can also induce decay processes, which are energetically forbidden otherwise. In \onlinecite{Kurilin_1999}, the field-induced lepton decay $l^-\to W^- + \nu_l$ was considered. Since the mass of the initial-state particle is smaller than the mass of the decay products, the process is clearly impossible in vacuum. The presence of the field does allow for such an exotic decay, but the probability $P_{l^-\to W^- + \nu_l}$ remains exponentially suppressed, i.e., $P_{l^-\to W^- + \nu_l}\sim \exp(-1/\chi_{0,W})$, where $\chi_{0,W}=\chi_0(m/m_W)^3$. Finally, the production of an $e^+\text{-}e^-$ pair by high-energy neutrino impact on a strong laser pulse has been calculated in \onlinecite{Tinsley_2005}. The setup is similar to the $e^+\text{-}e^-$ pair production processes in QED discussed in Secs. \ref{PP_L} and \ref{BH}. However, as it was shown in \onlinecite{Tinsley_2005}, the laser-induced process $\nu \to \nu + e^+ + e^-$ is extremely unlikely. At a field intensity of about $I_0=3\times 10^{18}\;\text{W/cm$^2$}$, the production length is on the order of a light year, even for a neutrino energy of 1 PeV. \subsection{Particle physics beyond the Standard Model} Recently, a lot of attention has been devoted to the possibility of employing intense laser sources to test aspects of physical theories which go even beyond the Standard Model. In \onlinecite{Heinzl_2010c}, for example, it is envisaged that effects of the noncommutativity of space-time modify the kinematics of multiphoton Compton scattering by inducing a nonzero photon mass. We recall that in noncommutative quantum field theories operators $X^{\mu}$ are associated to spacetime coordinates $x^{\mu}$, which do not commute, but rather satisfy the commutation relations $[X^{\mu},X^{\nu}]=i\Theta^{\mu\nu}$, with $\Theta^{\mu\nu}$ being an antisymmetric constant tensor \cite{Douglas_2001}. On a different side, one of the still open problems of the Standard Model is the so-called strong CP problem \cite{Kim_2010}. The nontrivial structure of the vacuum, as predicted by Quantum Chromodynamics (QCD), allows for the violation within QCD of the combined symmetry of charge conjugation (C) and parity (P). This implies a value for the neutron's electric dipole moment which, however, is already many orders of magnitude larger than experimental upper limits. One way of solving this problem was suggested in \onlinecite{Peccei_1977} which required the existence of a massive pseudoscalar boson, called axion. The axion has never been observed experimentally although some of its properties can be predicted on theoretical grounds: it should be electrically neutral and its mass should not exceed $1\;\text{eV}$ in order of magnitude. Although being electrically neutral, the axion is predicted to couple to the electromagnetic field $\mathcal{F}^{\mu\nu}(x)$ through a Lagrangian-density term \begin{equation} \label{L_ag} \mathscr{L}_{a\gamma}(x)=\frac{g}{4}a(x)\mathcal{F}^{\mu\nu}(x)\tilde{\mathcal{F}}_{\mu\nu}(x), \end{equation} with $g$ being the photon-axion coupling constant and $a(x)$ the axion field. The photon-axion Lagrangian density in Eq. \eqref{L_ag} has mainly two implications: 1) the existence of axions induces a change in the polarization of a light beam passing through a background electromagnetic field; 2) a photon can transform into an axion (and vice versa) in the presence of a background electromagnetic field. The first prediction has been tested in experiments like the Brookhaven-Fermilab-Rochester-Trieste (BFRT) \cite{BFRT} and the Polarizzazione del Vuoto con LASer (PVLAS) \cite{PVLAS}, where a linearly polarized probe laser field crossed a region in which a strong magnetic field was present of $3.25\;\text{T}$ and $5\;\text{T}$ at BFRT and at PVLAS, respectively. Testing the second prediction is the aim of the so-called ``light shining through a wall'' experiments like the Any Light Particle Search (ALPS) \cite{ALPS}, the CERN Axion Solar Telescope (CAST) \cite{CAST} and Gamma to milli-eV particle search (GammeV) \cite{GammeV} (see also the detailed theoretical analysis in \onlinecite{Adler_2008}). In the GammeV experiment, for example, the light of a Nd:YAG laser passes through a region in which a $5\;\text{T}$ magnetic field is present. A mirror is positioned behind that region in order to reflect the laser light. The axions which would eventually be created in the magnetic-field region pass through the mirror undisturbed and can be reconverted to photons by means of a second magnetic field, activated after the mirror itself. So far these experiments have given negative results. An interesting experimental proposal has been put forward in \onlinecite{Rabadan_2006}, where the high-energy photon beam delivered by an XFEL facility has been suggested as a probe beam to test regions of parameters (like the axion mass $m_a$ or the photon-axion coupling constant $g$) which are inaccessible via optical laser light. The perspective for reaching ultrahigh intensities at future laser facilities has stimulated new proposals for employing such fields to elicit the photon-axion interaction \cite{Gies_2009}. In fact, an advantage of using strong uniform magnetic fields is that they can be kept strong for a macroscopically long time (of the order of hours) and on a macroscopic spatial region (of the order of $1\;\text{m}$) (for QED processes occurring in a strong magnetic field, see the standard review \onlinecite{Erber_1966} and the very current overview paper \onlinecite{Dunne_2012} for recent progresses in the field). On the other hand, laser beams deliver fields much stronger than those employed in the mentioned experiments (the magnetic field strength of a laser beam with the available intensity of $10^{22}\;\text{W/cm$^2$}$ amounts to about $6.5\times 10^5\;\text{T}$) but in a microscopic space-time region. However, it has been first realized in \onlinecite{Mendonca_2007} that envisaged ultrahigh intensities at future laser facilities may compensate for the tiny space-time extension of the laser spot region. In \onlinecite{Mendonca_2007} the coupled equations of the electromagnetic field $\mathcal{F}^{\mu\nu}(x)$ and the axion field $a(x)$ \begin{equation} \label{a_F} \left\{ \begin{aligned} &\partial_{\mu}\partial^{\mu}a+m_a^2a=\frac{1}{4}g\mathcal{F}^{\mu\nu}\tilde{\mathcal{F}}_{\mu\nu}\\ &\partial_{\mu}\mathcal{F}^{\mu\nu}=g(\partial_{\mu}a)\tilde{\mathcal{F}}^{\mu\nu} \end{aligned} \right. \end{equation} are solved approximately. It is shown that if a probe laser field propagates through a strong plane-wave field, the axion field ``grows'' at the expense mainly of the probe field itself, whose intensity should be observed to decrease. Laser powers of the order of $1\;\text{PW}$ have already been shown to provide stronger hints for the presence of axions than magnetic-field-based experiments like the PVLAS. More realistic Gaussian laser beams are considered in \onlinecite{Dobrich_2010} where the starting point is also represented by Eq. \eqref{a_F}. The suggested experimental setup assumes a probe electromagnetic beam with angular frequency $\omega_{p,\text{in}}$ passing through a strong counterpropagating Gaussian beam with angular frequency $\omega_{0,\parallel}$ and another strong Gaussian beam propagating perpendicularly and with angular frequency $\omega_{0,\perp}$. By choosing $\omega_{0,\perp}=2\omega_{0,\parallel}$, it is found that after a photon-axion-photon double conversion, photons are generated with angular frequencies $\omega_{p,\text{out}}=\omega_{p,\text{in}}\pm \omega_{0,\parallel}$. The amplitudes of these processes are shown to be peaked at specific values of the axion mass $m_{a,\pm}=2\sqrt{\omega_{p,\text{in}}\omega_{0,\parallel}+\omega_{0,\parallel}^2(1\pm 1)/2}$. Since the optical photon energies are of the order of $1\;\text{eV}$, this setup allows for the investigation of values of the axion mass in this regime. This is very important because such a region of the axion mass is inaccessible to experiments based on strong magnetic fields, which can probe regions at most in the meV range. In addition to electrically neutral new particles such as axions, yet unobserved particles with nonzero charge may also exist. The fact that they have so far escaped detection implies that they are either very heavy (rendering them a target for large-scale accelerator experiments), or that they are light but very weakly charged. In the latter case, these so-called minicharged particles, i.e., particles with absolute value of the electric charge much smaller than $|e|$, are suitable candidates for laser-based searches \cite{Gies_2009}. Let $m_\epsilon$ and $Q_\epsilon=\epsilon e$, with $0<\epsilon \ll 1$, denote the minicharged particle mass and charge, respectively. Then, the corresponding critical field scale $F_{\text{cr},\epsilon}=m_\epsilon^2/|Q_\epsilon|$ can be much lower than $F_{\text{cr}}$. As a consequence, vacuum nonlinearities associated with minicharged particles may be very pronounced in an external laser field with intensity much less than $I_{\text{cr}}\sim 10^{29}\;\text{W/cm$^2$}$. Moreover, even at optical photon energies $\sim 1\;\text{eV}$ the effective Lagrangian approach might become inappropriate to describe the relevant physics if $m_{\epsilon}\lesssim 1\;\text{eV}$ (see Sec. \ref{Diff_Low}). In \onlinecite{Gies_2006}, vacuum dichroism and birefringence effects due to the existence of minicharged particles were analyzed when a probe laser beam with $\omega_p>2m_\epsilon$ traverses a magnetic field. It was shown that polarization measurements in this setup would provide much stronger constraints on minicharged particles in the mass range below 0.1~eV than in previous laboratory searches. \section{Conclusion and outlook} \label{Conlc} The fast development of laser technology has been paving the way to employ laser sources for investigating relativistic, quantum electrodynamical, nuclear and high-energy processes. Starting with the lowest required intensities, relativistic atomic processes are already within the reach of available laser systems, while the proposed methods to compensate the deteriorating effects of the relativistic drift still have to be tested experimentally. Moreover, a fully consistent theoretical interpretation of recent experimental results on correlation effects in relativistic multielectron tunneling is still missing. Concerning the interaction of free electrons with intense laser beams, we have seen that experiments have been performed to explore the classical regime. Only the E-144 experiment at SLAC has so far been realized on multiphoton Compton scattering, although presently available lasers and electron beams would allow for probing this regime in full detail. We have also pointed out that at laser intensities of the order of $10^{22}\text{-} 10^{23}\;\text{W/cm$^2$}$, RR effects come into play at electron energies of the order of a few GeV. It is envisaged that the quantum radiation-dominated regime, where quantum and RR effects substantially alter the electron dynamics, could be one of the first extreme regimes of light-matter interaction to be probed with upcoming petawatt laser facilities. On the theoretical side, most of the calculations have been performed by approximating the laser field as a plane wave, as the Dirac equation in the presence of a focused background field cannot be solved analytically. Certainly, new methods have to be developed to calculate photon spectra including quantum effects and spatio-temporal focusing of the laser field in order to be able to quantitatively interpret upcoming experimental results. Nonlinear quantum electrodynamical effects have been shown to become observable at future multipetawatt laser facilities, as well as at ELI and HiPER. Here the main challenges concern the measurability of tiny effects on the polarization of probe beams and on the detection of a typically very low number of signal photons out of large backgrounds. Similar challenges are envisaged to detect the presence of light and weakly-interacting hypothetical particles like axions and minicharged particles. The physical properties (mass, coupling constants, etc.) of such hypothetical particles are, of course, unknown. In this respect intense laser fields may be employed here to set bounds on physical quantities like the axion mass and, in particular, to scan regions of physical parameters, which are inaccessible to conventional methods based, for example, on astrophysical observations. Different schemes have been proposed to observe $e^+\text{-}e^-$ pair production at intensities below the Schwinger limit, which seems now to be feasible in the near future, at least from a theoretical point of view. Corresponding studies would complement the results of the pioneering E-144 experiment and deepen our understanding of the QED vacuum in the presence of extreme electromagnetic fields. This is also connected with the recent investigations on the development of QED cascades in laser-laser collisions. In addition to being intrinsically interesting, the development of QED cascades is expected to set a limit on the maximal attainable laser intensity. However, the study of quantum cascades in intense laser fields has started relatively recently and is still under vivid development. More advanced analytical and numerical methods are required in order to describe realistically and quantitatively such a complex system as an electron-positron-photon plasma in the presence of a strong driving laser field. Nuclear quantum optics is also a new exciting and promising field. Since the energy difference between nuclear levels is typically in the multi-keV and MeV range, high-frequency laser pulses, especially in combination with accelerators, are preferable in controlling nuclear dynamics. As pointed out, especially table-top highly coherent x-ray light beams, envisaged for the future, open up perspectives for exciting applications including nuclear state preparation and nuclear batteries. Finally, we want to point out that most of the considered processes have not yet been observed or tested experimentally. This is in our opinion one of the most challenging aspects of upcoming laser physics, not only from an experimental point of view, but also from the point of view of theoretical methods. Experimentally, the main reason is that in order to test, for example, nonlinear quantum electrodynamics or to investigate nuclear quantum optics, high-energy particle beams (including photon beams) are required to be available in the same laboratory as the strong laser. On the one hand, the combined expertise from different experimental physical communities is required to perform such complex but fundamental experiments. On the other hand, the fast technological development of laser-plasma accelerators is very promising and exciting, as this seems the most feasible way towards the realization of stable, table-top high-energy particle accelerators. Combining such high-energy probe beams with an ultrarelativistic laser beam in a single all-optical setup will certainly result in a unique tool for advancing our understanding of intense laser-matter interactions. \begin{acknowledgments} We are grateful to many students, colleagues and collaborators for inspiring discussions, ideas, and suggestions, in particular in joint publications during the here most relevant last six years, to H. Bauke, T. J. B{\"u}rvenich, A. Dadi, C. Deneke, J. Evers, B. Galow, R. Grobe, Z. Harman, H. Hu, A. Ipp, U. D. Jentschura, B. King, M. Klaiber, M. Kohler, G. Yu. Kryuchkyan, W.-T. Liao, C. Liu, E. L\"{o}tstedt, A. Macchi, F. Mackenroth, S. Meuren, A. I. Milstein, G. Mocken, S. J. M\"{u}ller, T.-O. M\"{u}ller, A. P\'{a}lffy, F. Pegoraro, M. Ruf, Y. I. Salamin, M. Tamburini, M. Verschl, and A. B. Voitkiv. \end{acknowledgments} \section*{List of frequently-used symbols} \begin{tabular}{ll} \hspace{-15pt} $E_0$ & laser field amplitude\\ \hspace{-15pt} $F_{\text{cr}}=m^2/|e|$ & critical electromagnetic field \\ &($\approx 1.3\times 10^{16}\;\text{V/cm}$)\\ &($\approx 4.4\times 10^{14}\;\text{G}$)\\ \hspace{-15pt} $I_0=E_0^2/4\pi$ & laser peak intensity\\ \hspace{-15pt} $I_{\text{cr}}=F_{\text{cr}}^2/4\pi$ & critical laser intensity \\ & ($\approx 4.6\times 10^{29}\;\text{W/cm$^2$}$)\\ \hspace{-15pt} $Z$ & nuclear charge number\\ \hspace{-15pt} $e$ \hspace{105pt} & electron charge \\ &($\approx -1/\sqrt{137}\approx -0.085$)\\ \hspace{-15pt} $k^{\mu}=(\omega,\bm{k})=\omega n^{\mu}$ & initial or incoming photon \\ &four-momentum\\ \hspace{-15pt} $k^{\prime\mu}=(\omega',\bm{k}')=\omega' n^{\prime\mu}$ & final or outgoing photon \\ &four-momentum\\ \hspace{-15pt} $k_0^{\mu}=\omega_0n_0^{\mu}=\omega_0(1,\bm{n}_0)$ & laser photon four-momentum\\ \hspace{-15pt} $m$ & electron mass\\ &($\approx 0.511\;\text{MeV}$)\\ \hspace{-15pt} $p_0^{\mu}=(\varepsilon_0,\bm{p}_0)=m\gamma_0(1,\bm{\beta}_0)$ & initial or incoming electron \\ &four-momentum\\ \hspace{-15pt} $\alpha=e^2$ & fine-structure constant \\ & ($\approx 1/137\approx 7.3\times 10^{-3}$)\\ \hspace{-15pt} $\chi_0=(p_{0,-}/m)(E_0/F_{\text{cr}})$ & nonlinear electron quantum \\ &parameter \\ \hspace{-15pt} $\varkappa_0=(k_-/m)(E_0/F_{\text{cr}})$ & nonlinear photon quantum \\ &parameter \\ \hspace{-15pt} $\lambda_0=T_0=2\pi/\omega_0$ & laser wavelength and period \\ \hspace{-15pt} $\lambda_C=1/m$ & Compton wavelength \\ & ($\approx 3.9\times 10^{-11}\;\text{cm}$)\\ \hspace{-15pt} $\xi_0=|e|E_0/m\omega_0$ & classical relativistic \\ & parameter \\ \hspace{-15pt} $\omega_0$ \hspace{105pt} & laser angular frequency\\ \end{tabular} \bibliographystyle{apsrmp}
\section*{Supporting Online Material} SOM Text \\ Figs. S1 to S4\\ Table S1\\ References (39 --- 57)\\
\section{Introduction} The game of Babylon begins with $n$ chips of various colors, spread out as ``stacks'' of height 1. Players Alice and Bob alternate combining two stacks which have the same number of chips or the same color on top (or both). The last player to make a legal move wins. The game was introduced in 2003 by French designer Bruno Faidutti \cite{F} and is available in commercial form, with 12 chips in four colors, three of each color. (Of course, the game can be played with ordinary poker chips.) Computer-aided analysis \cite{B,G,M} has shown that with best play the aforementioned initial configuration of the commercial game is a win for the second player. The simplicity and attractiveness of Babylon has led researchers to try to solve the game in general, and in particular the cases where the tokens are of only two colors. Goadrich and Schlatter \cite{GS} show that the first player (Alice) can always win if there are only one or two chips of the minority color, or if $n$ is odd; they conjectured that Bob can win in the remaining, most difficult cases, when $n$ is even and there are three or more chips of the minority color. We prove their conjecture, completing the winning-player classification for two-color Babylon. Our strategy for Bob is explicit and easy to implement in practice. \section{General Observations} Alice wins any game of Babylon if the total number of moves is odd; Bob, if even. Since the number of stacks decreases by one at each player's turn, the number of moves is at most $n{-}1$. In two-color Babylon, the only other possibility is that there are $n{-}2$ moves, ending with two stacks of different height and color (the color of a stack is naturally defined to be the color of its top chip; the colors of its other chips cannot affect the game any more). The one-color game is of course trivial, with Alice winning when $n$ is even and Bob when $n$ is odd. If $n$ is even and there are only one or two chips of the minority color (which, continuing the convention of \cite{GS}, is red), Alice wins easily by obliterating the red chips; if there's only one she covers it with a blue chip, and if there are two she stacks them and at her next move covers them with the blue 2-stack that Bob was forced to make at his first move. If $n$ is odd and there are one or two red chips, Alice needs to keep them alive; this is easily done by stacking red on blue (if there is one red chip) or stacking the two red chips, then stacking the result on Bob's blue 2-stack. Thereafter if Bob ever threatens the red stack by building a blue one of the same height, Alice piles her red stack on top of the challenging blue one. Thus the size of the red stack is always a power of 2, and at least 4. There is never any danger that Alice will herself be forced to make a blue stack of the same height as the red one, because this can only happen if all blue stacks are exactly half the height of the red one; but that is impossible because $n$ is odd. Thus, Alice wins all two-color games that begin with one or two red chips. When $n$ is odd and there are three or more red chips, Alice again needs to preserve both colors; and again, she can easily do so, as shown in \cite{GS}. Thus the critical remaining case is when there are three or more red chips and $n$ is even. This time it is Bob who needs to keep both colors alive. The difficulty of this case, relative to the odd-$n$ case, is twofold: (1) Bob lacks the power of first move, and (2) Bob must take care not to create, or permit Alice to create, a stack of height $n/2$ that is the only stack of its color. We will show that Bob can overcome these difficulties and win in all cases, completing the classification for two-color Babylon. \section{Notation} We use the symbol $b$ to denote a blue chip, $r$ a red chip; $x$ is a chip of either color and $y$ a chip of the {\em other} color; $z$ stands for a chip of any color. We will be using strings of the above symbols to describe stacks, with the left-most symbol standing for the chip on top. A {\em hill} is a stack of height at least two. When there is no possibility of confusion we often denote a hill by the capitalized form of the notation for its color, thus we could write $R$ instead of $rbb$ and $X$ instead of $xxyz$. In this notation a {\em move} by either player consists of appending one string (stack) to another, always with the restriction that the two stacks were of the same color or the same height. Since the other cases are solved, we will be concerned exclusively with games in which there are initially an even number $2m$ of chips, each its own stack, and at least three of the minority color (red). The parity of the number of stacks at any state of the game tells us who is next to move; Alice moves next from any even state, Bob from any odd state. A state will be termed {\em safe} if it leads to a win by Bob, with best play. Equivalently, we may say that an even state in which Alice cannot move (i.e., a two-stack state consisting of red and blue hills of different heights) is safe, that an even state is safe if any move of Alice produces a safe state, and an odd state is safe if there is some move by Bob leading to another safe state. Our objective is to show that the initial state is safe. \section{Main Results} \begin{theorem} \label{thm:main} Two-color Babylon with $p$ chips of one color and $q$ of the other, $1 \le p \le q$, is a win for the second player if and only if $p{+}q$ is even and $p \ge 3$. \end{theorem} \begin{proof} As noted in the introduction, owing to the results in \cite{GS} we need only provide a winning strategy for Bob when $p{+}q=2m$ and $p \ge 3$. We will show that Bob can win by creating even states in which there is at most one hill of each color, until an ``end game'' is reached in which he has to be a bit more careful. A {\em singleton} is a stack that is not a hill, that is, one consisting of just one chip. A state with $j$ red singletons, $k$ blue singletons, red hills of heights $u_1,u_2,\dots$ and blue hills of heights $v_1,v_2,\dots$ will be denoted by $\langle j,k;u_1,u_2,\dots;v_1,v_2,\dots \rangle$. The initial configuration is thus $\langle p,q;; \rangle$. Even states with exactly one hill of each color, each of even height, and at least four singletons (stacks of height one) of each color, will be called {\em target} states. Thus, a target state with $j$ red singletons, $k$ blue singletons, a red hill of height $2u$ and a blue hill of height $2v$ will be denoted by $\langle j,k;2u;2v \rangle$. Alice's options at a target state, and our notations for them, are: \begin{enumerate} \item place a singleton on top of another singleton: $xz$ \item place a singleton on top of the hill of its color (or vice-versa): $xX$ \item place one hill on top of the other: $XY$. \end{enumerate} Of course, option (2) requires that a hill exist, and (3) that there are two hills and that they are of the same height. Bob normally answers $xz$ with $xzX$, $xX$ with $xX$ and $XY$ with $yy$, always creating a new even state in which there is again exactly one hill of each color, each of even height. We call this kind of play by Bob the {\em even-hill-strategy}. In particular, $\langle j,k;u;v \rangle$ transforms to either $\langle j{-}2,k;2u{+}2;2v \rangle$, $\langle j,k{-}2;2u;2v{+}2 \rangle$, $\langle j{-}1,k{-}1;2u{+}2;2v \rangle$, $\langle j{-}1,k{-}1;2u;2v{+}2 \rangle$, $\langle j{-}2,k;2;2v{+}2u \rangle$, or $\langle j,k{-}2;2u{+}2v;2 \rangle$. Since the number of singletons of one or both colors may have been reduced below four, the new state might not be a target state. Observe that if $p \ge 6$ then Bob can create a target state at his first move: if Alice opens with $xx$ he replies with $yy$, and if she opens with $xy$ he replies with $yx$. Either way the result is target state $\langle p{-}2,q{-}2;2,2 \rangle$. Before proceeding further it will be useful to understand what happens when there is only one stack left of one of the colors. \begin{lemma} \label{lemma:1} Any even state with just one stack $X$ of color $x$ and at least one of color $y$ is safe, provided the height of $X$ exceeds $m$. \end{lemma} \begin{proof} Nothing can ever be combined with $X$, so both colors survive and Bob wins. \end{proof} \begin{lemma} \label{lemma:2} Any odd state with just one stack $X$ of color $x$ is safe, provided the height $u$ of $X$ is less than $m$, not all $y$-stacks are of height $u/2$, and at most two are of height $u$. \end{lemma} \begin{proof} We proceed by induction on the (even) number $t$ of $y$-stacks. If there are two $y$-stacks of height $u$, Bob stacks them; if one, he stacks it with any other $y$-stack. All $y$-stacks of height $u$ are thus killed. In all other cases, Bob makes the largest $y$-stack possible of height $u' \not= u$. (Since the $y$-stacks are not all of height $u/2$, either the tallest two or the shortest two will combine to make a stack of height $\not=u$; thus $u'$ is well-defined.) Now Alice cannot change the height of $X$, nor can she suddenly make three $y$-stacks of height $u$. Can she cause all the $y$-stacks to become of height $u/2$? No: if there had been a $y$-stack of height $u/2$ before we would now have $u' > u/2$. So we are reduced to the cases where Alice has created a first or second $y$-stack of height $u/2$. Since Alice leaves an even number of $y$ stacks, and $u + u/2 + u/2 < 2u < 2m = n$, there must be a stack of height different than $u/2$ after her move. \end{proof} Let us return to the proof of the main theorem. Bob's strategy is to continue applying the even-hill-strategy as long as it leaves a target state for Alice, i.e., it leaves at least four singletons in each of the colors. Now consider the first position $S$ faced by Bob such that playing from $S$ according to the even-hill-strategy would leave less than four singletons in one of the colors. If following the even-hill-strategy would yield a position with 3 singletons in one color and at least 3 in the other color (which happens if $S=\langle 3,k;2u,2;2v \rangle$ or $S=\langle 3,k;2u;2v,2 \rangle$ with $k \ge 3$, Alice having just created a stack of height 2, or if $S=\langle 5,k;0;2v\rangle$, Alice having just played $BR$), then Bob wins by the following lemma \begin{lemma} \label{lemma:4} Any even state of the form $\langle 3,k;2u;2v \rangle$ with $k\geq 3$ and $u+v\geq 3$ is safe. \end{lemma} \begin{proof} \renewcommand{\labelenumi}{(\roman{enumi})} We analyze all possible replies by Alice. In each case Bob must be careful not to allow Alice to create a solo stack of height $m$. His moves must thus take into account the value of $2u$ (of concern when at or just below $m$) and the relative values of $2u$ and $2v$ (of concern when they are both around $m/2$). \begin{enumerate} \item Alice replies $rb$; By \Lr{lemma:3} Bob wins by $rbR$ unless $2u+4=m$ and $k>3$. If the latter is the case, then Alice threatens to create a single red stack containing half the chips. But Bob can reply with $br$, and it is now easy for him to make sure that both colors survive; see Lemmas \ref{lemma:1} and \ref{lemma:2}. \item Alice replies $br$; Arguing as in the previous case, we can play $brB$ unless $2u+2=m$ and $k>3$. Note that in the latter case $u>1$ must hold for $u+v\geq 3$. Bob can then play $br$, threatening $rR$ in his next move which would finish the game by \Lr{lemma:2}. Alice can try to prevent this by $rb$, but then $brrb$ wins for Bob by \Lr{lemma:2} again. \item Alice replies $rr$; If $2u+3=m$ then $br$ is an easy win for Bob. If $u=v$ then $BR$ is easily seen to be safe using \Lr{lemma:2}. If neither of the above is the case, then $rrr$ is safe as it allows Bob to create a single red stack on his next move. \item Alice replies $rR$; If $2u+3=m$, or $2u+3=m+2$, or $2u+3=m-2$ (the latter two possibilities are dangerous if and only if $v=1$) then $br$ leads an easy win for Bob. Otherwise $rr$ wins. \item Alice replies $RB$; If the height $H$ of the new hill is $m{-}3$ or $m{-}2$ then $br$ is safe. If it is $m{-}1$ then $rr$ is safe. The other cases are easy. \item Alice replies $BR$; Bob easily wins by $rr$ by making sure that red survives. \item Alice replies $bB$ or $bb$; We have $k>3$ without loss of generality, for otherwise we can switch colors. Thus in both cases, $bbB$ creates a position of the same form as the original one, and we proceed by induction. \end{enumerate} \end{proof} Thus, it only remains to consider positions $S$ from which the even-hill-strategy would yield a position with exactly 2 singletons in one color and $S$ has not been preceeded by a position with exactly 3 singletons in that color, for we have already seen that such positions are won by Bob. Since Alice's last state was a target state, such an $S$ must have one of the following forms (or its red-blue reflection): \renewcommand{\labelenumi}{\arabic{enumi}.} \begin{enumerate} \item $\langle 3,k;2u{+}1;2v \rangle$ with $k \ge 4$ (if Alice just played $rR$); \item $\langle 2,k;2u,2;2v \rangle$ with $k \ge 4$ and even (if Alice just played $rr$). \item $\langle 4,k;0;2v \rangle$ with $k \ge 4$ and even (if Alice just played $BR$). \end{enumerate} These cases will be easily solved using the following lemma. \begin{lemma} \label{lemma:3} Any even state of the form $\langle 2,2s;2u,2v \rangle$ with $u,v \geq 1$ is safe unless $2u+2=m$ and $s>1$. \end{lemma} \begin{proof} We analyze all possible moves of Alice from this state: \renewcommand{\labelenumi}{(\roman{enumi})} \begin{enumerate} \item Alice plays $BR$; (Provided $u=v$.) Bob plays $rr$ and can make sure that the $rr$ stack survives to the end by immediately obliterating any blue stack of size 2 Alice creates. \item Alice plays $RB$; If the height $2u+2v$ of the new stack $X$ is greater than $m$ or less than $m-2$ then Bob plays $rr$ and wins easily by \Lr{lemma:1} or \Lr{lemma:2}. Otherwise we have to be slightly more careful to avoid a red stack of height $m$: if $2u+2v=m$ play $rX$, if $2u+2v=m-1$ then play $rr$, and if $2u+2v=m-2$ play $br$ and counter $rb$ with $brrb$. \item Alice plays $bb$; We may assume $s=1$ without loss of generality. If $2v+2\neq m$ then $bbB$ wins. If $2v+2= m$, in which case $u=v$ holds, we win by $BR$. \item Alice plays $bB$; We may assume $s=1$ without loss of generality. If $2v+2\neq m$ then $bbB$ as above. If $2v+2= m$ then $rb$ wins. \item Alice plays $br$; If $2u= m-1$, then $u>1$ because we have at least 8 chips in total. It is now easy to see that $br$ wins. If $2u\neq m-1$ then $rR$ wins by \Lr{lemma:1} or \Lr{lemma:2}. \item Alice plays $rb$; Bob easily wins by playing $rrb$, unless $2u=m-3$ in which case he can safely play $br$. \item Alice plays $rR$; Bob easily wins by playing $rrb$, unless $2u=m-2$ in which case he can safely play $br$. \item Alice plays $rr$; This is the only case in which, under certain circumstances, Bob can lose. Indeed, if $2u+2=m$, then Alice threatens to combine all red chips in a stack of height $m$, winning the game. The only way for Bob to defend is to capture the red hill of height $2u$ with the blue hill, and he can do so \iff\ $u=v$ which implies $s=1$. If $2u+2\neq m$ then playing $rrR$ reaches a safe position by \Lr{lemma:1} or \Lr{lemma:2}. \end{enumerate} \end{proof} We can now analyze the positions 1-3 above. For position 1 note that, by \Lr{lemma:3}, $rR$ is safe unless $2u+4=m$. If the latter is the case, then $br$ will easily allow Bob to create a single red stack of height different than $m$, winning by Lemmas \ref{lemma:1} and \ref{lemma:2}. Position 2 is very similar and we leave the details to the reader. Finally, in position 3 Bob can play $rr$ and win by \Lr{lemma:3} since $m>4$. \medskip This completes the proof in all cases when $p\geq 6$, and so Bob can create a target state at his first move as already mentioned. It remains to consider the simple cases when $p=3,4$ or $5$. When $p=5$, Bob has to deviate from his general strategy only slightly. If Alice opens with $bb$ or $rb$ then he replies with $rb$ and $bb$ respectively. This creates a target state unless the number of blue chips $q$ is also 5. But in the latter case \Lr{lemma:3} applies, with blue being the minority color now. If Alice opens with $br$ then $rr$ wins by \Lr{lemma:3} again. Finally, if Alice opens with $rr$ then $rr$ will quickly allow Bob to form a single red stack, winning by \Lr{lemma:1} or \Lr{lemma:2}. When $p=4$ then Bob can play his first move according to his 2-hill strategy for the general case unless $q=4$; this immediately reaches a position $\langle 2,2s;2;2 \rangle$, which is a win by \Lr{lemma:3}. If $q=4$ then Bob can force the position $\langle 1,3;2,2; \rangle$ after the first move, and it is easy to win that position. Finaly, the case $p=3$ is an easy excersise left for the entertainment of the reader. \end{proof} \section*{Acknowledgement}} \newcommand{\section*{Acknowledgements}}{\section*{Acknowledgements}} \comment{ \begin{lemma}\label{} \end{lemma} \begin{proof} \end{proof} \begin{theorem}\label{} \end{theorem} \begin{proof} \end{proof} } \newtheorem{proposition}{Proposition}[section] \newtheorem{definition}[proposition]{Definition} \newtheorem{theorem}[proposition]{Theorem} \newtheorem{corollary}[proposition]{Corollary} \newtheorem{lemma}[proposition]{Lemma} \newtheorem{observation}[proposition]{Observation} \newtheorem{conjecture}{{\color{red}Conjecture}}[section] \newtheorem*{noConjecture}{{\color{red}Con\-jecture}} \newtheorem{problem}[conjecture]{{\color{red}Problem}} \newtheorem*{noProblem}{{\color{red}Problem}} \newtheorem*{noTheorem}{Theorem} \newtheorem{examp}[proposition]{Example \newtheorem{claim}{Claim} \newtheorem*{noclaim}{Claim} \newtheorem{vermutung}{{\color{red}Vermutung}}[section] \newtheorem*{noVermutung}{{\color{red}Vermutung}} \newcommand{\example}[2]{\begin{examp} \label{#1} {{#2}}\end{examp}} \newcommand{\kreis}[1]{\mathaccent"7017\relax #1} \newcommand{0}{0} \ifnum 0 = 1 \newcommand{\note}[1]{ \ {\color{blue} \hspace*{-60pt} NOTE: \color{Turquoise}{\small \tt \begin{minipage}[c]{1.1\textwidth} #1 \end{minipage} \ignorespacesafterend }} \ } \else \newcommand{\note}[1]{} \fi \newcommand{\afsubm}[1]{ \ifnum 0 = 1 {\mymargin{#1}} \fi} \ifnum 0 = 1 \newcommand{\ensuremath{\color{red} \bowtie \bowtie \bowtie\ }}{\ensuremath{\color{red} \bowtie \bowtie \bowtie\ }} \else \newcommand{\ensuremath{\color{red} \bowtie \bowtie \bowtie\ }}{} \fi \ifnum 0 = 1 \newcommand{\fig}[1]{Figure ``{#1}''} \else \newcommand{\fig}[1]{Figure~\ref{#1}} \fi \ifnum 0 = 1 \newcommand{\figs}[1]{Figures ``{#1}''} \else \newcommand{\figs}[1]{Figures~\ref{#1}} \fi \ifnum 0 = 1 \usepackage[notref,notcite]{showkeys} \fi \ifnum 0 = 0 \renewcommand{\color}[1]{} \fi \newcommand{Figure}{Figure} \newcommand{\showFigu}[3]{ \begin{figure}[htbp] \centering \noindent \epsfxsize=0.8\hsize \epsfbox{#1.eps} \caption{\small #3} \label{#2} \end{figure} } \newcommand{\showFig}[2]{ \begin{figure}[htbp] \centering \noindent \epsfbox{#1.eps} \caption{\small #2} \label{#1} \end{figure} } \newcommand{\showFigI}[1]{ \begin{figure}[htbp] \centering \noindent \epsfbox{#1.eps} \label{#1} \end{figure} } \newcommand{\ensuremath{\mathbb N}}{\ensuremath{\mathbb N}} \newcommand{\ensuremath{\mathbb R}}{\ensuremath{\mathbb R}} \newcommand{\ensuremath{\mathbb Z}}{\ensuremath{\mathbb Z}} \newcommand{\ensuremath{\mathbb Q}}{\ensuremath{\mathbb Q}} \newcommand{\ensuremath{\mathcal A}}{\ensuremath{\mathcal A}} \newcommand{\ensuremath{\mathcal B}}{\ensuremath{\mathcal B}} \newcommand{\ensuremath{\mathcal C}}{\ensuremath{\mathcal C}} \newcommand{\ensuremath{\mathcal D}}{\ensuremath{\mathcal D}} \newcommand{\ensuremath{\mathcal E}}{\ensuremath{\mathcal E}} \newcommand{\ensuremath{\mathcal F}}{\ensuremath{\mathcal F}} \newcommand{\ensuremath{\mathcal G}}{\ensuremath{\mathcal G}} \newcommand{\ensuremath{\mathcal H}}{\ensuremath{\mathcal H}} \newcommand{\ensuremath{\mathcal I}}{\ensuremath{\mathcal I}} \newcommand{\ensuremath{\mathcal J}}{\ensuremath{\mathcal J}} \newcommand{\ensuremath{\mathcal K}}{\ensuremath{\mathcal K}} \newcommand{\ensuremath{\mathcal L}}{\ensuremath{\mathcal L}} \newcommand{\ensuremath{\mathcal M}}{\ensuremath{\mathcal M}} \newcommand{\ensuremath{\mathcal N}}{\ensuremath{\mathcal N}} \newcommand{\ensuremath{\mathcal O}}{\ensuremath{\mathcal O}} \newcommand{\ensuremath{\mathcal P}}{\ensuremath{\mathcal P}} \newcommand{\ensuremath{\mathcal Q}}{\ensuremath{\mathcal Q}} \newcommand{\ensuremath{\mathcal S}}{\ensuremath{\mathcal S}} \newcommand{\ensuremath{\mathcal R}}{\ensuremath{\mathcal R}} \newcommand{\ensuremath{\mathcal T}}{\ensuremath{\mathcal T}} \newcommand{\ensuremath{\mathcal U}}{\ensuremath{\mathcal U}} \newcommand{\ensuremath{\mathcal V}}{\ensuremath{\mathcal V}} \newcommand{\ensuremath{\mathcal X}}{\ensuremath{\mathcal X}} \newcommand{\ensuremath{\mathcal Y}}{\ensuremath{\mathcal Y}} \newcommand{\ensuremath{\mathcal Z}}{\ensuremath{\mathcal Z}} \newcommand{\ensuremath{\mathcal W}}{\ensuremath{\mathcal W}} \newcommand{\ensuremath{\omega}}{\ensuremath{\omega}} \newcommand{\ensuremath{\Omega}}{\ensuremath{\Omega}} \newcommand{\ensuremath{\alpha}}{\ensuremath{\alpha}} \newcommand{\ensuremath{\beta}}{\ensuremath{\beta}} \newcommand{\ensuremath{\gamma}}{\ensuremath{\gamma}} \newcommand{\ensuremath{\Gamma}}{\ensuremath{\Gamma}} \newcommand{\ensuremath{\delta}}{\ensuremath{\delta}} \newcommand{\ensuremath{\epsilon}}{\ensuremath{\epsilon}} \newcommand{\ensuremath{\kappa}}{\ensuremath{\kappa}} \newcommand{\ensuremath{\lambda}}{\ensuremath{\lambda}} \newcommand{\ensuremath{\sigma}}{\ensuremath{\sigma}} \newcommand{\ensuremath{\alpha}}{\ensuremath{\alpha}} \newcommand{\ensuremath{\ell}}{\ensuremath{\ell}} \newcommand{\ensuremath{\mathcal C}}{\ensuremath{\mathcal C}} \newcommand{\ensuremath{\mathcal C}}{\ensuremath{\mathcal C}} \newcommand{\ensuremath{\mathcal C(G)}}{\ensuremath{\mathcal C(G)}} \newcommand{\ensuremath{|G|}}{\ensuremath{|G|}} \newcommand{\mathbb 0}{\mathbb 0} \newcommand{\backslash}{\backslash} \newcommand{\upharpoonright}{\upharpoonright} \newcommand{\qed}{\qed} \newcommand{\cong}{\cong} \newcommand{\mathbb Z_2}{\mathbb Z_2} \newcommand{\ensuremath{{n\in\N}}}{\ensuremath{{n\in\ensuremath{\mathbb N}}}} \newcommand{\ensuremath{{i\in\N}}}{\ensuremath{{i\in\ensuremath{\mathbb N}}}} \newcommand{\ensuremath{[0,1]}}{\ensuremath{[0,1]}} \newcommand{\limf}[1]{\ensuremath{\lim \inf(#1)}} \newcommand{\sgl}[1]{\ensuremath{\{#1\}}} \newcommand{\pth}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~path} \newcommand{\topth}[2]{topological \ensuremath{#1}\text{--}\ensuremath{#2}~path} \newcommand{\topths}[2]{topological \ensuremath{#1}\text{--}\ensuremath{#2}~paths} \newcommand{\pths}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~paths} \newcommand{\arc}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~arc} \newcommand{\arcs}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~arcs} \newcommand{\seq}[1]{\ensuremath{(#1_n)_{n\in\ensuremath{\mathbb N}}}} \newcommand{\sseq}[2]{\ensuremath{(#1_i)_{i\in #2}}} \newcommand{\susq}[2]{\ensuremath{(#1_{#2_i})_{\ensuremath{{i\in\N}}}}} \newcommand{\oseq}[2]{\ensuremath{(#1_\ensuremath{\alpha})_{\ensuremath{\alpha}< #2}}} \newcommand{\fml}[1]{\ensuremath{\{#1_i\}_{i\in I}}} \newcommand{\ffml}[2]{\ensuremath{\{#1_i\}_{i\in #2}}} \newcommand{\ofml}[2]{\ensuremath{\{#1_\ensuremath{\alpha}\}_{\ensuremath{\alpha}< #2}}} \newcommand{\flo}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~flow} \newcommand{\flos}[2]{\ensuremath{#1}\text{--}\ensuremath{#2}~flows} \newcommand{\ensuremath{G\ }}{\ensuremath{G\ }} \newcommand{\ensuremath{G}}{\ensuremath{G}} \newcommand{\Gn}[1]{\ensuremath{G[#1]_n}} \newcommand{\Gi}[1]{\ensuremath{G[#1]_i}} \newcommand{s}{s} \newcommand{\frac{1}{2}}{\frac{1}{2}} \newcommand{\ceil}[1]{\ensuremath{\lceil #1 \rceil}} \newcommand{\ensuremath{|G|_\ell}}{\ensuremath{|G|_\ell}} \newcommand{\ensuremath{\ell-TOP}}{\ensuremath{\ell-TOP}} \newcommand{\ensuremath{\partial^\ell G}}{\ensuremath{\partial^\ell G}} \newcommand{\ensuremath{\ell: E(G) \to \R_{>0}}}{\ensuremath{\ell: E(G) \to \ensuremath{\mathbb R}_{>0}}} \newcommand{\ensuremath{\sum_{e \in E(G)} \ell(e) < \infty}}{\ensuremath{\sum_{e \in E(G)} \ell(e) < \infty}} \newcommand{\ensuremath{\sum_{e \in E(G)} \ell(e) = \infty}}{\ensuremath{\sum_{e \in E(G)} \ell(e) = \infty}} \newcommand{Kirchhoff's node law}{Kirchhoff's node law} \newcommand{Kirchhoff's cycle law}{Kirchhoff's cycle law} \newcommand{non-elusive} %cut respecting}{non-elusive} \newcommand{Non-elusive} %cut respecting}{Non-elusive} \newcommand{\vec{e}}{\vec{e}} \newcommand{\vec{E}}{\vec{E}} \newcommand{Dirichlet problem}{Dirichlet problem} \newcommand{Brownian Motion}{Brownian Motion} \newcommand{harmonic function}{harmonic function} \newcommand{\mathbb E}{\mathbb E} \renewcommand{\Pr}{\mathbb{P}} \newcommand{\Lr}[1]{Lemma~\ref{#1}} \newcommand{\Tr}[1]{Theorem~\ref{#1}} \newcommand{\Sr}[1]{Section~\ref{#1}} \newcommand{\Prr}[1]{Pro\-position~\ref{#1}} \newcommand{\Prb}[1]{Problem~\ref{#1}} \newcommand{\Cr}[1]{Corollary~\ref{#1}} \newcommand{\Cnr}[1]{Con\-jecture~\ref{#1}} \newcommand{\Or}[1]{Observation~\ref{#1}} \newcommand{\Er}[1]{Example~\ref{#1}} \newcommand{\Dr}[1]{De\-fi\-nition~\ref{#1}} \newcommand{locally finite}{locally finite} \newcommand{locally finite graph}{locally finite graph} \newcommand{finite graph}{finite graph} \newcommand{infinite graph}{infinite graph} \newcommand{basic open set}{basic open set} \newcommand{Cayley graph}{Cayley graph} \newcommand{non-locally-finite}{non-locally-finite} \newcommand{non-locally-finite graph}{non-locally-finite graph} \newcommand{Hamilton cycle}{Hamilton cycle} \newcommand{Hamilton circle}{Hamilton circle} \newcommand{Euler tour}{Euler tour} \newcommand{Euler tours}{Euler tours} \renewcommand{\iff}{if and only if} \newcommand{for every}{for every} \newcommand{For every}{For every} \newcommand{for each}{for each} \newcommand{For each}{For each} \newcommand{such that}{such that} \newcommand{so that}{so that} \newcommand{there is}{there is} \newcommand{there holds}{there holds} \newcommand{well-defined}{well-defined} \newcommand{without loss of generality}{without loss of generality} \newcommand{by the definition of}{by the definition of} \newcommand{By the definition of}{By the definition of} \newcommand{by the construction of}{by the construction of} \newcommand{By the construction of}{By the construction of} \newcommand{with respect to}{with respect to} \newcommand{finitely many}{finitely many} \newcommand{infinitely many}{infinitely many} \newcommand{is straightforward to check}{is straightforward to check} \newcommand{is easy to see}{is easy to see} \newcommand{locally connected}{locally connected} \newcommand{path connected}{path connected} \newcommand{induction hypothesis}{induction hypothesis} \newcommand{Freudenthal compactification}{Freudenthal compactification} \newcommand{topological cycle space}{topological cycle space} \newcommand{topological circle}{topological circle} \newcommand{topological path}{topological path} \newcommand{topological Euler tour}{topological Euler tour} \newcommand{\labequ}[2]{ \begin{equation} \label{#1} #2 \end{equation} } \newcommand{\labtequ}[2]{ \begin{equation} \label{#1} \begin{minipage}[c]{0.9\textwidth} #2 \end{minipage} \ignorespacesafterend \end{equation} } \newcommand{\labtequc}[2]{ \begin{equation} \label{#1} \text{ #2 } \end{equation} } \newcommand{\labtequstar}[1]{ \begin{equation*} \begin{minipage}[c]{0.9\textwidth} #1 \end{minipage} \ignorespacesafterend \end{equation*} } \newcommand{\mymargin}[1] \marginpar{% \begin{minipage}{\marginparwidth}\small% \begin{flushleft}% {\color{blue}#1}% \end{flushleft}% \end{minipage}% }% }% \newcommand{\assign}{ \mathrel{\mathop{:}}= } \newcommand{\mySection}[2]{} \newcommand{Diestel and K\"uhn}{Diestel and K\"uhn} \newcommand{Bruhn and Stein}{Bruhn and Stein} \newcommand{Kirchhoff}{Kirchhoff} \newcommand{Diestel}{Diestel} \newcommand{\cite[Section~8.5]{diestelBook05}}{\cite[Section~8.5]{diestelBook05}} \newcommand{\cite{diestelBook05}}{\cite{diestelBook05}} \newcommand{\cite[p.~208]{ElemTop}}{\cite[p.~208]{ElemTop}} \newcommand{\LemArc}[1]{ \begin{lemma}[\cite[p.~208]{ElemTop}] \label{#1} The image of a topological path with endpoints $x,y$ in a Hausdorff space $X$ contains an arc in $X$ between $x$ and $y$. \end{lemma} } \newcommand{\LemKonig}[1]{ \begin{lemma}[K\"onig's Infinity Lemma \cite{InfLemma}] \label{#1} Let $V_0,V_1,\dotsc$ be an infinite sequence of disjoint non-empty finite sets, and let $G$ be a graph on their union. Assume that every vertex $v$ in a set $V_n$ with $n\ge 1$ has a neighbour in $V_{n-1}$. Then $G$ contains a ray $v_0v_1\dotsm$ with $v_n\in V_n$ for all $n$. \end{lemma} } \newcommand{\LemCombStarC}{\cite[Lemma~8.2.2]{diestelBook05}} \newcommand{\LemCombStar}[1]{ \begin{lemma}[\LemCombStarC] \label{#1} Let $U$ be an infinite set of vertices in a connected graph $G$. Then $G$ contains either a ray $R$ and infinitely many pairwise disjoint \pths{U}{R} or a subdivided star with infinitely many leaves in $U$. \end{lemma} } \newcommand{\LemCycDecC}{\cite[Theorem~8.5.8]{diestelBook05}} \newcommand{\LemCycDec}[1]{ \begin{lemma}[\LemCycDecC] \label{#1} Every element of $\ensuremath{\mathcal C(G)}$ is a disjoint union of circuits. \end{lemma} } \newcommand{\ThmComp}[1]{ \begin{theorem} \label{#1} Let $K$ be an infinite set of propositional formulas, every finite subset of which is satisfiable. Then $K$ is satisfiable. \end{theorem} } \newcommand{\ThmCompC}{\cite{LMCS}} \newcommand{\LemCutCrit}[1]{ \begin{lemma} \label{#1} Let $F\subseteq E(G)$. Then $F\in \mathcal C(G)$ if and only if $F$ meets every finite cut in an even number of edges. \end{lemma} } \newcommand{\LemCutCritC}{\cite[Theorem~8.5.8]{diestelBook05}} \newcommand{\LemFinCut}[1]{ \begin{lemma} \label{#1} Let $G$ be a connected locally finite graph, $\{ X, Y \}$ a partition of $V(G)$, and $F := E(X, Y )$. \begin{enumerate} \item $F$ is finite if and only if $\overline{X} \cap \overline{Y}=\emptyset$. \item If $F$ is finite, there is no arc in $|G|\backslash \kreis{F}$ with one endpoint in $X$ and the other in $Y$. \end{enumerate} \end{lemma} } \newcommand{\LemFinCutC}{\cite[Theorem~8.5.5]{diestelBook05}} \newcommand{\LemHP}[1]{ \begin{lemma} \label{#1} Let $G$ be a locally finite graph, let $x,y \in V(G)$ and let $(\tau_i)_{i\in \ensuremath{\mathbb N}}$ be a sequence of topological $x$--$y$~paths in $|G|$. Then, there is an infinite subsequence $(\tau_{a_i})_{i\in \ensuremath{\mathbb N}}$ of $(\tau_i)$ and a topological $x$--$y$~path $\sigma$ in $|G|$ such that $E(\sigma) \subseteq \limf{E(\tau_{a_i})}$. Moreover, if no $\tau_i$ traverses any edge more than once then $E(G) \setminus E(\sigma) \subseteq \limf{E(G) \setminus E(\tau_{a_i})}$, no edge is traversed by $\sigma$ more than once, and for every finite subset $F$ of $E(\sigma)$ there is an $\ensuremath{{n\in\N}}$ such that the linear ordering of $F$ induced by $\sigma$ coincides with that induced by $\tau_{a_i}$ for every $a_i>n$. \end{lemma} } \newcommand{\LemHPC}{\cite{hotchpotch}} \newcommand{\LemInjHomC}{\cite{armstrong}} \newcommand{\LemInjHom}[1]{ \begin{lemma}[\LemInjHomC] \label{#1} A continuous bijection from a compact space to a Hausdorff space is a homeomorphism. \end{lemma} } \newcommand{\cite{maclane37}}{\cite{maclane37}} \newcommand{\ThmMacLane}[1]{ \begin{theorem}[\cite{maclane37}] \label{#1} A finite graph is planar if and only if its cycle space \ensuremath{\mathcal C(G)}\ has a simple generating set. \end{theorem} } \newcommand{\LemHeineCaC}{\cite{}} \newcommand{\LemHeineCa}[1]{ \begin{lemma}[Heine-Cantor Theorem] \label{#1} Let $M$ be a compact metric space, and let $f: M \to N$ be a continuous function, where $N$ is a metric space. Then $f$ is uniformly continuous. \end{lemma} }
\section{What makes an incomplete-block design good for experiments?} \label{sec:intro} Experiments are designed in many ways: for example, Latin squares, block designs, split-plot designs. Combinatorialists, on the other hand, have a much more specialized usage of the term ``design'', as we remark later. We are concerned here with incomplete-block designs, more special than the statistician's designs and more general than the mathematician's. To a statistician, a \emph{block design} has two components. There is an underlying set of experimental units, partitioned into $b$~blocks of size~$k$. There is a further set of $v$~treatments, and also a function~$f$ from units to treatments, specifying which treatment is allocated to which experimental unit; that is, $f(\omega)$ is the treatment allocated to experimental unit~$\omega$. Thus each block defines a subset, or maybe a multi-subset, of the treatments. In a \emph{complete-block design}, we have $k=v$ and each treatment occurs once in every block. Here we assume that blocks are \emph{incomplete} in the sense that $k<v$. We assume that the purpose of the experiment is to find out about the treatments, and differences between them. The blocks are an unavoidable nuisance, an inherent feature of the experimental units. In an agricultural experiment the experimental units may be field plots and the blocks may be fields or plough-lines; in a clinical trial the experimental units may be patients and the blocks hospitals; in process engineering the experimental units may be runs of a machine that is recalibrated each day and the blocks days. See \cite{rabbook} for further examples. In all of these situations, the values of $b$, $k$ and~$v$ are given. Given these values, not all incomplete-block designs are equally good. This chapter describes some criteria that can be used to choose between them. For example, Fig.~\ref{fig:queen} shows two block designs with $v=15$, $b=7$ and $k=3$. We use the convention that the treatments are labelled $1$, \ldots,~$v$, that columns represent blocks, and that the order of the entries in each column is not significant. Where necessary, we use the notation $\Gamma_j$ to refer to the block which is shown as the $j$th column, for $j=1$, \ldots, $b$. \begin{figure} \begin{center} \begin{tabular}{c@{\qquad\qquad}c} $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 1 & 2 & 3 & 4 & 5 & 6\\ 2 & 4 & 5 & 6 & 10 & 11 & 12\\ 3 & 7 & 8 & 9 & 13 & 14 & 15\\ \hline \end{array} $ & $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 1 & 1 & 1 & 1 & 1 & 1\\ 2 & 4 & 6 & 8 & 10 & 12 & 14\\ 3 & 5 & 7 & 9 & 11 & 13 & 15\\ \hline \end{array} $\\ \\ (a) & (b) \end{tabular} \end{center} \caption{Two block designs with $v=15$, $b=7$ and $k=3$} \label{fig:queen} \end{figure} The \emph{replication} $r_i$ of treatment $i$ is defined to be $\card{f^{-1}(i)}$, which is the number of experimental units to which it is allocated. For the design in Fig.~\ref{fig:queen}(a), $r_i \in \setof{1,2}$ for all~$i$. As we see later, statisticians tend to prefer designs in which all the replications are as equal as possible. If $r_i=r_j$ for $1\leq i<j\leq v$ then the design is \emph{equireplicate}: then the common value of $r_i$ is usually written as~$r$, and $vr=bk$. The design in Fig.~\ref{fig:queen}(b) is a \emph{queen-bee design} because there is (at least) one treatment that occurs in every block. Scientists tend to prefer such designs because they have been taught to compare every treatment to one distinguished treatment, which may be called a \emph{control treatment}. \begin{figure}[htbp] \begin{center} \begin{tabular}{c@{\qquad\qquad}c} $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 1 & 1 & 1 & 2 & 2 & 2\\ 2 & 3 & 3 & 4 & 3 & 3 & 4\\ 3 & 4 & 5 & 5 & 4 & 5 & 5\\ \hline \end{array} $ & $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 1 & 1 & 1 & 2 & 2 & 2\\ 1 & 3 & 3 & 4 & 3 & 3 & 4\\ 2 & 4 & 5 & 5 & 4 & 5 & 5\\ \hline \end{array} $\\ \\ (a) & (b) \end{tabular} \end{center} \caption{Two block designs with $v=5$, $b=7$ and $k=3$} \label{fig:binary} \end{figure} Fig.~\ref{fig:binary} shows two block designs with $v=5$, $b=7$ and $k=3$. The design in Fig.~\ref{fig:binary}(b) shows a new feature: treatment~$1$ occurs on two experimental units in block~$\Gamma_1$. A block design is \emph{binary} if $f(\alpha)\ne f(\omega)$ whenever $\alpha$ and $\omega$ are experimental units in the same block. The design in Fig.~\ref{fig:binary}(a) is binary. It seems to be obvious that binary designs must be better than non-binary ones, but we shall see later that this is not necessarily so. However, if there is any block on which $f$ is constant, then that block provides no information about treatments, so we assume from now on that there are no such blocks. \begin{figure}[htbp] \begin{center} \begin{tabular}{c@{\qquad\qquad}c} $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 2 & 3 & 4 & 5 & 6 & 7\\ 2 & 3 & 4 & 5 & 6 & 7 & 1\\ 4 & 5 & 6 & 7 & 1 & 2 & 3\\ \hline \end{array} $ & $\begin{array}{|c|c|c|c|c|c|c|} \hline 1 & 2 & 3 & 4 & 5 & 6 & 7\\ 2 & 3 & 4 & 5 & 6 & 7 & 1\\ 3 & 4 & 5 & 6 & 7 & 1 & 2\\ \hline \end{array} $ \\ \\ (a) & (b) \end{tabular} \end{center} \caption{Two block designs with $v=7$, $b=7$ and $k=3$} \label{fig:fano} \end{figure} Fig.~\ref{fig:fano} shows two equireplicate binary block designs with $v=7$, $b=7$ and $k=3$. A binary design is \emph{balanced} if every pair of distinct treatments occurs together in the same number of blocks. If that number is $\lambda$, then $r(k-1) = (v-1)\lambda$. Such designs are also called \emph{$2$-designs} or \emph{BIBDs}. The design in Fig.~\ref{fig:fano}(a) is balanced with $\lambda=1$; the design in Fig.~\ref{fig:fano}(b) is not balanced. Pure mathematicians usually assume that, if they exist, balanced designs are better than non-balanced ones. (Indeed, many do not call a structure a `design' unless it is balanced.) As we shall show in Section~\ref{sec:bibd}, this assumption is correct for all the criteria considered here. However, for given values of $v$ and $k$, a non-balanced design with a larger value of $b$ may produce more information than a balanced design with a smaller value of $b$. \section{Graphs from block designs} \subsection{The Levi graph} A simple way of representing a block design is its \emph{Levi graph}, or \textit{incidence graph}, introduced in \cite{levi}. This graph has $v+b$ vertices, one for each block and one for each treatment. There are $bk$ edges, one for each experimental unit. If experimental unit $\omega$ is in block $j$ and $f(\omega) = i$, then the corresponding edge~$\tilde e_\omega$ joins vertices $i$ and~$j$. Thus the graph is bipartite, with one part consisting of block vertices and the other part consisting of treatment vertices. Moreover, the graph has multiple edges if the design is not binary. Fig.~\ref{fig:levibin} gives the Levi graph of the design in Fig.~\ref{fig:binary}(b). \begin{figure}[hbtp] \begin{center} \setlength{\unitlength}{8mm} \begin{picture}(12,3.5)(0,0.5) \multiput(0,3)(2,0){7}{\circle*{0.4}} \multiput(2,1)(2,0){5}{\circle*{0.4}} \put(0,3.3){\makebox(0,0)[b]{$\Gamma_1$}} \put(2,3.3){\makebox(0,0)[b]{$\Gamma_2$}} \put(4,3.3){\makebox(0,0)[b]{$\Gamma_3$}} \put(6,3.3){\makebox(0,0)[b]{$\Gamma_4$}} \put(8,3.3){\makebox(0,0)[b]{$\Gamma_5$}} \put(10,3.3){\makebox(0,0)[b]{$\Gamma_6$}} \put(12,3.3){\makebox(0,0)[b]{$\Gamma_7$}} \put(2,0.7){\makebox(0,0)[t]{1}} \put(4,0.7){\makebox(0,0)[t]{2}} \put(6,0.7){\makebox(0,0)[t]{3}} \put(8,0.7){\makebox(0,0)[t]{4}} \put(10,0.7){\makebox(0,0)[t]{5}} \put(2,0.9){\line(-1,1){2}} \put(2,1.1){\line(-1,1){2}} \put(4,1){\line(-2,1){4}} \put(2,1){\line(0,1){2}} \put(6,1){\line(-2,1){4}} \put(8,1){\line(-3,1){6}} \put(2,1){\line(1,1){2}} \put(6,1){\line(-1,1){2}} \put(10,1){\line(-3,1){6}} \put(2,1){\line(2,1){4}} \put(8,1){\line(-1,1){2}} \put(10,1){\line(-2,1){4}} \put(4,1){\line(2,1){4}} \put(6,1){\line(1,1){2}} \put(8,1){\line(0,1){2}} \put(4,1){\line(3,1){6}} \put(6,1){\line(2,1){4}} \put(10,1){\line(0,1){2}} \put(4,1){\line(4,1){8}} \put(8,1){\line(2,1){4}} \put(10,1){\line(1,1){2}} \end{picture} \end{center} \caption{The Levi graph of the design in Fig.~\ref{fig:binary}(b)} \label{fig:levibin} \end{figure} We regard two block designs as the same if one can be obtained from the other by permuting the experimental units within each block. Since the vertices of the Levi graph are labelled, there is a bijection between block designs and their Levi graphs. Let $n_{ij}$ be the number of edges from treatment-vertex $i$ to block-vertex~$j$; that is, treatment $i$ occurs on $n_{ij}$ experimental units in block $j$. The $v\times b$ matrix $\mathbf{N}$ whose entries are the $n_{ij}$ is the \emph{incidence matrix} of the block design. If the rows and columns of $\mathbf{N}$ are labelled, we can recover the block design from its incidence matrix. \subsection{The concurrence graph} In a binary design, the \emph{concurrence} $\lambda_{ij}$ of treatments~$i$ and $j$ is $r_i$ if $i=j$ and otherwise is the number of blocks in which $i$ and $j$ both occur. For non-binary designs we have to count the number of occurrences of the pair $\setof{i,j}$ in blocks according to multiplicity, so that $\lambda_{ij}$ is the $(i,j)$-entry of $\boldsymbol{\Lambda}$, where $\boldsymbol{\Lambda} = \mathbf{N} \mathbf{N}^\top$. The matrix $\boldsymbol{\Lambda}$ is called the \emph{concurrence matrix} of the design. The \emph{concurrence graph} of the design has the treatments as vertices. There are no loops. If $i\ne j$, then there are $\lambda_{ij}$ edges between vertices $i$ and $j$. Each such edge corresponds to a pair $\setof{\alpha,\omega}$ of experimental units in the same block, with $f(\alpha)=i$ and $f(\omega)=j$: we denote this edge by $e_{\alpha\omega}$. (This edge does not join the experimental units $\alpha$ and $\omega$; it joins the treatments applied to these units.) It follows that the degree $d_i$ of vertex $i$ is given by \begin{equation} d_i = \sum_{j\ne i} \lambda_{ij}. \label{eq:valency} \end{equation} Figs.~\ref{fig:cgqueen} and~\ref{fig:cgbin} show the concurrence graphs of the designs in Figs.~\ref{fig:queen} and~\ref{fig:binary}, respectively. \begin{figure}[htbp] \begin{center} \begin{tabular}{c@{\qquad\qquad\qquad}c} \setlength{\unitlength}{8mm} \begin{picture}(5,6)(0.8,0) \put(0.4,4.6){\circle*{0.4}} \put(0.4,4.9){\makebox(0,0)[b]{13}} \put(2.4,4.6){\circle*{0.4}} \put(2.4,4.9){\makebox(0,0)[b]{7}} \put(3.6,4.6){\circle*{0.4}} \put(3.6,4.9){\makebox(0,0)[b]{8}} \put(5.6,4.6){\circle*{0.4}} \put(5.6,4.9){\makebox(0,0)[b]{11}} \put(0.4,3.4){\circle*{0.4}} \put(0.4,3.1){\makebox(0,0)[t]{10}} \put(2.4,3.4){\circle*{0.4}} \put(2.1,3.1){\makebox(0,0)[rt]{1}} \put(3.6,3.4){\circle*{0.4}} \put(3.9,3.1){\makebox(0,0)[tl]{2}} \put(5.6,3.4){\circle*{0.4}} \put(5.6,3.1){\makebox(0,0)[t]{14}} \put(1.2,1.4){\circle*{0.4}} \put(0.9,1.4){\makebox(0,0)[r]{12}} \put(2.4,1.4){\circle*{0.4}} \put(2.6,1.2){\makebox(0,0)[tl]{6}} \put(3.6,1.4){\circle*{0.4}} \put(3.9,1.4){\makebox(0,0)[l]{9}} \put(1.8,0.4){\circle*{0.4}} \put(1.8,0.1){\makebox(0,0)[t]{15}} \put(1.4,4.0){\circle*{0.4}} \put(1.4,4.3){\makebox(0,0)[b]{4}} \put(4.6,4.0){\circle*{0.4}} \put(4.6,4.3){\makebox(0,0)[b]{5}} \put(3.0,2.4){\circle*{0.4}} \put(3.3,2.4){\makebox(0,0)[l]{3}} \put(1.2,1.4){\line(1,0){2.4}} \put(1.8,0.4){\line(3,5){1.2}} \put(1.8,0.4){\line(-3,5){0.6}} \put(2.4,1.4){\line(3,5){1.2}} \put(3.6,1.4){\line(-3,5){1.2}} \put(2.4,3.4){\line(1,0){1.2}} \put(2.4,3.4){\line(0,1){1.2}} \put(2.4,3.4){\line(-5,3){2.0}} \put(0.4,3.4){\line(0,1){1.2}} \put(0.4,3.4){\line(5,3){2.0}} \put(5.6,3.4){\line(0,1){1.2}} \put(5.6,3.4){\line(-5,3){2.0}} \put(3.6,3.4){\line(5,3){2}} \put(3.6,3.4){\line(0,1){1.2}} \end{picture} & \setlength{\unitlength}{8mm} \begin{picture}(5,6)(0,0) \put(2.5,2.5){\circle*{0.4}} \put(2.5,2.2){\makebox(0,0)[t]{1}} \put(2.5,2.5){\line(5,1){2.5}} \put(2.5,2.5){\line(3,2){2.1}} \put(2.5,2.5){\line(2,3){1.4}} \put(2.5,2.5){\line(1,5){0.5}} \put(2.5,2.5){\line(-5,1){2.5}} \put(2.5,2.5){\line(-3,2){2.1}} \put(2.5,2.5){\line(-2,3){1.4}} \put(2.5,2.5){\line(-1,5){0.5}} \put(2.5,2.5){\line(-3,-2){2.1}} \put(2.5,2.5){\line(-2,-3){1.4}} \put(2.5,2.5){\line(-5,-1){2.5}} \put(2.5,2.5){\line(5,-1){2.5}} \put(2.5,2.5){\line(3,-2){2.1}} \put(2.5,2.5){\line(2,-3){1.4}} \put(5,3){\circle*{0.4}} \put(5.3,3){\makebox(0,0)[l]{12}} \put(4.6,3.9){\circle*{0.4}} \put(4.9,3.9){\makebox(0,0)[l]{11}} \put(3.9,4.6){\circle*{0.4}} \put(3.9,4.9){\makebox(0,0)[b]{10}} \put(3,5){\circle*{0.4}} \put(3,5.3){\makebox(0,0)[b]{9}} \put(0,3){\circle*{0.4}} \put(-0.3,3){\makebox(0,0)[r]{5}} \put(0.4,3.9){\circle*{0.4}} \put(0.1,3.9){\makebox(0,0)[r]{6}} \put(1.1,4.6){\circle*{0.4}} \put(1.1,4.9){\makebox(0,0)[b]{7}} \put(2,5){\circle*{0.4}} \put(2,5.3){\makebox(0,0)[b]{8}} \put(4.6,1.1){\circle*{0.4}} \put(4.9,1.1){\makebox(0,0)[l]{14}} \put(3.9,0.4){\circle*{0.4}} \put(3.9,0.1){\makebox(0,0)[t]{15}} \put(5,2){\circle*{0.4}} \put(5.3,2){\makebox(0,0)[l]{13}} \put(0,2){\circle*{0.4}} \put(-0.3,2){\makebox(0,0)[r]{4}} \put(0.4,1.1){\circle*{0.4}} \put(0.1,1.1){\makebox(0,0)[r]{3}} \put(1.1,0.4){\circle*{0.4}} \put(1.1,0.1){\makebox(0,0)[t]{2}} \put(3.9,0.4){\line(1,1){0.7}} \put(5,2){\line(0,1){1}} \put(4.6,3.9){\line(-1,1){0.7}} \put(2,5){\line(1,0){1}} \put(1.1,0.4){\line(-1,1){0.7}} \put(0,2){\line(0,1){1}} \put(0.4,3.9){\line(1,1){0.7}} \end{picture} \\ \\ (a) & (b) \end{tabular} \end{center} \caption{The concurrence graphs of the designs in Fig.~\ref{fig:queen}} \label{fig:cgqueen} \end{figure} \begin{figure} \begin{center} \begin{tabular}{c@{\qquad\qquad\qquad}c} \setlength{\unitlength}{8mm} \begin{picture}(4,3.6)(-2,0) \put(-1,0){\circle*{0.4}} \put(1,0){\circle*{0.4}} \put(-2,2){\circle*{0.4}} \put(2,2){\circle*{0.4}} \put(0,3){\circle*{0.4}} \put(-1,-0.1){\line(1,0){2}} \put(-1,0.1){\line(1,0){2}} \put(-2,2){\line(1,0){4}} \put(-2,2){\line(2,1){2}} \put(2,2){\line(-2,1){2}} \put(-1.1,0.1){\line(1,3){1}} \put(-0.9,-0.1){\line(1,3){1}} \put(1.1,0.1){\line(-1,3){1}} \put(0.9,-0.1){\line(-1,3){1}} \put(-1.1,-0.1){\line(-1,2){1}} \put(-0.9,0.1){\line(-1,2){1}} \put(1.1,-0.1){\line(1,2){1}} \put(0.9,0.1){\line(1,2){1}} \put(-2.1,2.1){\line(2,1){2}} \put(-1.9,1.9){\line(2,1){2}} \put(2.1,2.1){\line(-2,1){2}} \put(1.9,1.9){\line(-2,1){2}} \put(-1.1,0.1){\line(3,2){3}} \put(-0.9,-0.1){\line(3,2){3}} \put(1.1,0.1){\line(-3,2){3}} \put(0.9,-0.1){\line(-3,2){3}} \put(0,3.3){\makebox(0,0)[b]{3}} \put(2.3,2){\makebox(0,0)[l]{2}} \put(1.3,0){\makebox(0,0)[l]{5}} \put(-2.3,2){\makebox(0,0)[r]{1}} \put(-1.3,0){\makebox(0,0)[r]{4}} \end{picture} & \setlength{\unitlength}{8mm} \begin{picture}(4,3)(-2,0) \put(-1,0){\circle*{0.4}} \put(1,0){\circle*{0.4}} \put(-2,2){\circle*{0.4}} \put(2,2){\circle*{0.4}} \put(0,3){\circle*{0.4}} \put(-1,-0.1){\line(1,0){2}} \put(-1,0.1){\line(1,0){2}} \put(-2,2.1){\line(1,0){4}} \put(-2,1.9){\line(1,0){4}} \put(-1.1,0.1){\line(1,3){1}} \put(-0.9,-0.1){\line(1,3){1}} \put(1.1,0.1){\line(-1,3){1}} \put(0.9,-0.1){\line(-1,3){1}} \put(-1.1,-0.1){\line(-1,2){1}} \put(-0.9,0.1){\line(-1,2){1}} \put(1.1,-0.1){\line(1,2){1}} \put(0.9,0.1){\line(1,2){1}} \put(-2.1,2.1){\line(2,1){2}} \put(-1.9,1.9){\line(2,1){2}} \put(2.1,2.1){\line(-2,1){2}} \put(1.9,1.9){\line(-2,1){2}} \put(-1.1,0.1){\line(3,2){3}} \put(-0.9,-0.1){\line(3,2){3}} \put(1.1,0.1){\line(-3,2){3}} \put(0.9,-0.1){\line(-3,2){3}} \put(0,3.3){\makebox(0,0)[b]{3}} \put(2.3,2){\makebox(0,0)[l]{2}} \put(1.3,0){\makebox(0,0)[l]{5}} \put(-2.3,2){\makebox(0,0)[r]{1}} \put(-1.3,0){\makebox(0,0)[r]{4}} \end{picture} \\ \\ (a) & (b) \end{tabular} \end{center} \caption{The concurrence graphs of the designs in Fig.~\ref{fig:binary}} \label{fig:cgbin} \end{figure} \begin{table}[hbtp] \begin{center} \begin{tabular}{c@{\quad\qquad}c} $\left[ \begin{array}{rrrrr} 8 & -1 & -3 & -2 & -2\\ -1 & 8 & {-3} & -2 & -2\\ -3 & {-3} & 10 & -2 & -2\\ -2 & -2 & -2 & 8 & -2\\ -2 & -2 & -2 & -2 & 8 \end{array} \right]$ & $\left[ \begin{array}{rrrrr} 8 & {-2} & -2 & -2 & -2\\ {-2} & 8 & {-2} & -2 & -2\\ -2 & {-2} & 8 & -2 & -2\\ -2 & -2 & -2 & 8 & -2\\ -2 & -2 & -2 & -2 & 8 \end{array} \right] $ \\ \\ (a) & (b) \end{tabular} \end{center} \caption{The Laplacian matrices of the concurrence graphs in Fig.~\ref{fig:cgbin}} \label{tab:binary} \end{table} If $k=2$, then the concurrence graph is effectively the same as the block design. Although the block design cannot be recovered from the concurrence graph for larger values of $k$, we shall see in Section~\ref{sec:crit} that the concurrence graphs contain enough information to decide between two block designs on any of the usual statistical criteria. They were introduced as \emph{variety concurrence graphs} in \cite{hdperw}, but are so useful that they may have been considered earlier. \subsection{The Laplacian matrix of a graph} Let $H$ be an arbitrary graph with $n$ vertices: it may have multiple edges, but no loops. The \emph{Laplacian matrix} $\mathbf{L}$ of $H$ is defined to be the square matrix with rows and columns indexed by the vertices of $H$ whose $(i,i)$-entry $L_{ii}$ is the valency of vertex~$i$ and whose $(i,j)$-entry $L_{ij}$ is the negative of the number of edges between vertices $i$ and $j$ if $i\ne j$. Then $L_{ii} = \sum_{j\ne i} L_{ij}$ for $1\leq i \leq n$, and so the row sums of $\mathbf{L}$ are all zero. It follows that $\mathbf{L}$ has eigenvalue~$0$ on the all-$1$ vector; this is called the \emph{trivial eigenvalue} of~$\mathbf{L}$. We show below that the multiplicity of the zero eigenvalue is equal to the number of connected components of $H$. Thus the multiplicity is~$1$ if and only if $H$ is connected. Call the remaining eigenvalues of $\mathbf{L}$ \emph{non-trivial}. They are all non-negative, as we show in the following theorem (see~\cite{bcc09}). \begin{thm} \label{thm:ledge} \begin{enumerate} \item If~$\mathbf{L}$ is a Laplacian matrix, then $\mathbf{L}$~is positive semi-definite. \item If $\mathbf{L}$ is a Laplacian matrix of order $n$ and $\mathbf{x}$ is any vector in $\mathbb{R}^n$, then \[ \mathbf{x}^\top \mathbf{L x} = \sum_{\mathrm{edges\ }ij} (x_i - x_j)^2. \] \item If $\mathbf{L}_1$ and $\mathbf{L}_2$ are the Laplacian matrices of graphs $H_1$ and $H_2$ with the same vertices, and if $H_2$ is obtained from $H_1$ by inserting one extra edge, then $\mathbf{L}_2- \mathbf{L}_1$ is positive semi-definite. \item If $\mathbf{L}$ is the Laplacian matrix of the graph $H$, then the multiplicity of the zero eigenvalue of $\mathbf{L}$ is equal to the number of connected components of~$H$. \end{enumerate} \end{thm} \begin{pf} Each edge between vertices $i$ and $j$ defines a $v\times v$ matrix whose entries are all $0$ apart from the following submatrix: \[ \begin{array}{cc} & \begin{array}{cc} \hphantom{-}i & \hphantom{-}j \end{array}\\ \begin{array}{c}i\\j\end{array} & \left[ \begin{array}{rr} 1 & -1\\ -1 & 1\rlap{\qquad .} \end{array} \right] \end{array} \] The Laplacian is the sum of these matrices, which are all positive semi-definite. This proves (a), (b) and~(c). From (b), the vector $\mathbf{x}$ is in the null space of the Laplacian if and only if $\mathbf{x}$ takes the same value on both vertices of each edge, which happens if and only if it takes a constant value on each connected component. This proves~(d). \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} Theorem~\ref{thm:ledge} shows that the smallest non-trivial eigenvalue of a connected graph is positive. This eigenvalue is sometimes called the \textit{algebraic connectivity} of the graph. The statistical importance of this is shown in Section~\ref{sec:crit}. In Section~\ref{sec:estvar} we shall need the Moore--Penrose generalized inverse of $\mathbf{L}^-$ of $\mathbf{L}$ (see \cite{geninv}). Put $\mathbf{P}_0 = n^{-1}\mathbf{J}_n$, where $\mathbf{J}_n$ is the $n\times n$ matrix whose entries are all $1$, so that $\mathbf{P}_0$ is the matrix of orthogonal projection onto the space spanned by the all-$1$ vector. If $H$ is connected then $\mathbf{L}+\mathbf{P}_0$ is invertible, and \[ \mathbf{L}^- = \left(\mathbf{L}+\mathbf{P}_0 \right)^{-1} -\mathbf{P}_0, \] so that $\mathbf{L}\mathbf{L}^- = \mathbf{L}^-\mathbf{L} = \mathbf{I}_n -\mathbf{P}_0$, where $\mathbf{I}_n$ is the identity matrix of order~$n$. \subsection{Laplacians of the concurrence and Levi graphs} There is a relationship between the Laplacian matrices of the concurrence and Levi graphs of a block design $\Delta$. Let $\mathbf{N}$ be the incidence matrix of the design, and $\mathbf{R}$ the diagonal matrix (with rows and columns indexed by treatments) whose $(i,i)$ entry is the replication $r_i$ of treatment $i$. If the design is equireplicate, then $\mathbf{R}=r\mathbf{I}_v$, where $r$ is the replication number. For the remainder of the paper, we will use $\mathbf{L}$ for the Laplacian matrix of the concurrence graph $G$ of $\Delta$, and $\tilde{\mathbf{L}}$ for the Laplacian matrix of the Levi graph $\tilde{G}$ of $\Delta$. Then it is straightforward to show that \[\mathbf{L}=k\mathbf{R}-\mathbf{NN}^\top,\qquad \tilde{\mathbf{L}}=\left[ \matrix{\mathbf{R} & -\mathbf{N}\cr -\mathbf{N}^\top & k\mathbf{I}\cr} \right] . \] The Levi graph is connected if and only if the concurrence graph is connected; thus $0$ is a simple eigenvalue of $\tilde\mathbf{L}$ if and only if it is a simple eigenvalue of $\mathbf{L}$, which in turn occurs if and only if all contrasts between treatment parameters are estimable (see Section~\ref{sec:estvar}). A block design with this property is itself called \emph{connected}: we consider only connected block designs. In the equireplicate case, the above expressions for $\mathbf{L}$ and $\tilde\mathbf{L}$ give a relationship between their Laplacian eigenvalues, as follows. Let $\mathbf{x}$ be an eigenvector of $\mathbf{L}$ with eigenvalue $\phi\ne rk$. Then, for each of the two solutions $\theta$ of the quadratic equation \[rk-\phi = (r-\theta)(k-\theta),\] there is a unique vector $\mathbf{z}$ in $\mathbb{R}^b$ such that $[\begin{array}{cc} \mathbf{x}^\top & \mathbf{z}^\top \end{array}]^\top$ is an eigenvector of $\tilde{\mathbf{L}}$ with eigenvalue $\theta$. Conversely, any eigenvalue $\theta\ne k$ of $\tilde{\mathbf{L}}$ arises in this way. The Laplacian matrices of the concurrence graphs in Fig.~\ref{fig:cgbin} are shown in Table~\ref{tab:binary}. \section{Statistical issues} \subsection{Estimation and variance} \label{sec:estvar} As part of the experiment, we measure the response $Y_\omega$ on each experimental unit~$\omega$. If $\omega$ is in block $\Gamma$, then we assume that \begin{equation} Y_\omega = \tau_{f(\omega)} + \beta_\Gamma + \varepsilon_\omega; \label{eq:linmod} \end{equation} here, $\tau_i$ is a constant depending on treatment~$i$, $\beta_\Gamma$ is a constant depending on block~$\Gamma$, and $\varepsilon_\omega$ is a random variable with expectation $0$ and variance~$\sigma^2$. Furthermore, if $\alpha\ne\omega$, then $\varepsilon_\alpha$ and $\varepsilon_\omega$ are uncorrelated. It is clear that we can add a constant to every block parameter, and subtract that constant from every treatment parameter, without changing~(\ref{eq:linmod}). It is therefore impossible to estimate the individual treatment parameters. However, if the design is connected, then we can estimate all \emph{contrasts} in the treatment parameters: that is, all linear combinations of the form $\sum_i x_i \tau_i$ for which $\sum_i x_i = 0$. In particular, we can estimate all the simple treatment differences $\tau_i - \tau_j$. An \emph{estimator} is a function of the responses $Y_\omega$, so it is itself a random variable. An estimator of a value is \emph{unbiased} if its expectation is equal to the true value; it is \emph{linear} if it is a linear function of the responses. Amongst linear unbiased estimators, the \emph{best} one (the so-called BLUE), is the one with the least variance. Let $V_{ij}$ be the variance of the BLUE for $\tau_i - \tau_j$. If all the experimental units form a single block, then the BLUE of ${\tau_1 - \tau_2}$ is just the difference between the average responses for treatments $1$ and~$2$. It follows that \[ V_{12} =\left(\frac{1}{r_1} + \frac{1}{r_2}\right)\sigma^2. \] When $v=2$, this variance is minimized (for a given number of experimental units) when $r_1=r_2$. Moreover, if the responses are normally distributed then the length of the $95$\% confidence interval for $\tau_1 - \tau_2$ is proportional to $\mathrm{t}(r_1 + r_2 -2,0.975)\sqrt{V_{12}}$, where $\mathrm{t}(d,p)$ is the $100p$-th percentile of the $\mathrm{t}$~distribution on $d$ degrees of freedom. The smaller the confidence interval, the more likely is our estimate to be close to the true value. This length can be made smaller by increasing $r_1 + r_2$, decreasing $\card{r_1 - r_2}$, or decreasing $\sigma^2$. However, matters are not so simple when $k<v$ and $v>2$. The following result can be found in any statistical textbook about block designs (see the section on further reading for recommendations). \begin{thm} \label{thm:var} Let $\mathbf{L}$ be the Laplacian matrix of the concurrence graph of a connected block design. If $\sum_i x_i=0$, then the variance of the BLUE of $\sum_ix_i\tau_i$ is equal to $(\mathbf{x}^\top \mathbf{L}^{-} \mathbf{x})k\sigma^2$. In particular, the variance~$V_{ij} $ of the BLUE of the simple difference $\tau_i - \tau_j$ is given by $V_{ij} = \left(L_{ii}^- +L_{jj}^{-} -2L_{ij}^{-}\right)k\sigma^2$. \end{thm} \subsection{Optimality criteria} \label{sec:crit} We want all of the $V_{ij}$ to be as small as possible, but this is a multi-dimensional problem if $v>2$. Let $\bar V$ be the average of the variances $V_{ij}$ over all treatments $i$, $j$ with $i\ne j$. Theorem~\ref{thm:var} shows that, for each fixed~$i$, \begin{eqnarray*} \sum_{j\ne i} V_{ij} &=& \sum_{j\ne i} (L^-_{ii} + L^-_{jj} - 2L^-_{ij}) k\sigma^2\\ & = & [(v-1)L_{ii}^- + (\mathop{\mathrm{Tr}}(\mathbf{L}^-) - L^-_{ii}) + 2L_{ii}^-]k\sigma^2\\ & = & [vL^-_{ii} + \mathop{\mathrm{Tr}}(\mathbf{L}^{-})]k\sigma^2, \end{eqnarray*} because the row sums and column sums of $L$ are all $0$. It follows that $\bar V = 2k\sigma^2 \mathop{\mathrm{Tr}}(\mathbf{L}^-)/(v-1)$. Let $\theta_1$, \ldots, $\theta_{v-1}$ be the non-trivial eigenvalues of $\mathbf{L}$, now listed according to multiplicity and in non-decreasing order. Then \[ \mathop{\mathrm{Tr}}(\mathbf{L}^-) = \frac{1}{\theta_1} + \cdots + \frac{1}{\theta_{v-1}}, \] and so \[ \bar V = 2k\sigma^2 \times \frac{1}{\mbox{harmonic mean of }\theta_1, \dots, \theta_{v-1}}. \] A block design is defined to be \emph{A-optimal} (in some given class of designs with the same values of $b$, $k$ and $v$) if it minimizes the value of $\bar V$; here `A' stands for `average'. Thus a design is A-optimal if and only if it maximizes the harmonic mean of $\theta_1$, \ldots, $\theta_{v-1}$. For $v>2$, the generalization of a confidence interval is a confidence ellipsoid centered at the point $(\hat \tau_1, \ldots, \hat \tau_v)$ which gives the estimated value of $(\tau_1, \ldots, \tau_v)$ in the $(v-1)$-dimensional subspace of $\mathbb{R}^v$ for which $\sum \tau_i = 0$. A block design is called \emph{D-optimal} if it minimizes the volume of this confidence ellipsoid. Since this volume is proportional to $\sqrt{\det(\mathbf{L}^- + \mathbf{P}_0)}$, a design is D-optimal if and only if it maximizes the geometric mean of $\theta_1$, \ldots, $\theta_{v-1}$. Here `D' stands for `determinant'. Rather than looking at averages, we might consider the worst case. If all the entries in the vector $\mathbf{x}$ are multiplied by a constant~$c$, then the variance of the estimator of $\sum x_i\tau_i$ is multiplied by $c^2$. Thus, those contrast vectors $\mathbf{x}$ which give the largest variance relative to their own length are those which maximize $\mathbf{x}^\top \mathbf{L}^- \mathbf{x}/\mathbf{x}^\top \mathbf{x}$; these are precisely the eigenvectors of $\mathbf{L}$ with eigenvalue $\theta_1$. A design is defined to be \emph{E-optimal} if it maximizes the value of $\theta_1$; here `E' stands for `extreme'. More generally, for $p$ in $(0, \infty)$, a design is called \emph{$\Phi_p$-optimal} if it minimizes \[ \left(\frac{\sum_{i=1}^{v-1} \theta_i^{-p}}{v-1}\right)^{1/p} \] Thus A-optimality corresponds to $p=1$, D-optimality corresponds to the limit as $p\rightarrow 0$, and E-optimality corresponds to the limit as $p\rightarrow \infty$. Let $\mathbf{L}_1$ and $\mathbf{L}_2$ be the Laplacian matrices of the concurrence graphs of block designs $\Delta_1$ and $\Delta_2$ for $v$ treatments in blocks of size~$k$. If $\mathbf{L}_2 - \mathbf{L}_1$ is positive semi-definite, then $\Delta_2$ is at least as good as $\Delta_1$ on all the $\Phi_p$-criteria. Theorem~\ref{thm:ledge}(c) shows that adding an extra block to a design cannot decrease its performance on any $\Phi_p$-criterion. There are even more general classes of optimality criteria (see \cite{harman} and \cite{ss} for details). Here we concentrate on A-, D- and E-optimality. \subsection{Questions and an example} A first obvious question to ask is: do these criteria agree with each other? Our optimality properties are all functions of the concurrence graph. What features of this graph should we look for if we are searching for optimal, or near-optimal, designs? Symmetry? (Nearly) equal degrees? (Nearly) equal numbers of edges between pairs of vertices? Distance-regularity? Large girth (ignoring cycles within a block)? Small numbers of short cycles (ditto)? High connectivity? Non-trivial automorphism group? Is it more useful to look at the Levi graph rather than the concurrence graph? \begin{eg} \label{eg:cube} Fig.~\ref{fig:cube} shows the values of the A- and D-criteria for all equi\-replicate block designs with $v=8$, $b=12$ and $k=2$: of course, these are just regular graphs with $8$ vertices and degree~$3$. The harmonic mean is shown on the $A$-axis, and the geometric mean on the $D$-axis. (Note that this figure includes some designs that were omitted from Figure~3 of \cite{rabpaper}.) The rankings on these two criteria are not exactly the same, but they do agree at the top end, where it matters. The second-best graph on both criteria is the cube; the best is the M\"obius ladder, whose vertices are the elements of $\mathbb{Z}_8$ and whose edges are $\setof{i,i+1}$ and $\setof{i,i+4}$ for $i$ in $\mathbb{Z}_8$. These two graphs are so close on both criteria that, for practical purposes, they can be regarded as equally good. \begin{figure} \begin{center} \setlength{\unitlength}{2cm} \begin{picture}(5.4,3.2)(0.4,2.6) \put(1,3){\line(1,0){4.5}} \put(1,3){\line(0,1){2.5}} \xlabel{1}{0.6} \xlabel{2}{1.2} \xlabel{3}{1.8} \xlabel{4}{2.4} \xlabel{5}{3.0} \ylabel{3}{1.8} \ylabel{4}{2.4} \ylabel{5}{3.0} \put(5.7,2.8){\makebox(0,0){$A$}} \put(0.8,5.7){\makebox(0,0){$D$}} \put(4.83,5.25){\makebox(0,0){$\boldsymbol{\times}$}} \put(2.98,4.31){\makebox(0,0){$\boldsymbol{+}$}} \put(3.35,4.66){\makebox(0,0){$\boldsymbol{+}$}} \put(3.32,4.51){\makebox(0,0){$\boldsymbol{+}$}} \put(3.89,4.95){\makebox(0,0){$\boldsymbol{+}$}} \put(4.88,5.26){\makebox(0,0){$\boldsymbol{\times}$}} \put(4.54,5.15){\makebox(0,0){$\boldsymbol{\times}$}} \put(4.70,5.21){\makebox(0,0){$\boldsymbol{\times}$}} \put(1.75,3.92){\makebox(0,0){$\boldsymbol{\circ}$}} \put(3.60,4.81){\makebox(0,0){$\boldsymbol{+}$}} \put(3.43,4.63){\makebox(0,0){$\boldsymbol{+}$}} \put(4.01,4.91){\makebox(0,0){$\boldsymbol{+}$}} \put(2.31,4.30){\makebox(0,0){$\boldsymbol{\circ}$}} \put(4.08,4.95){\makebox(0,0){$\boldsymbol{+}$}} \put(2.30,4.20){\makebox(0,0){$\boldsymbol{\circ}$}} \put(2.44,4.45){\makebox(0,0){$\boldsymbol{\circ}$}} \put(3.71,4.77){\makebox(0,0){$\boldsymbol{+}$}} \put(4.19,4.99){\makebox(0,0){$\boldsymbol{+}$}} \put(4.9,5.4){\makebox(0,0)[l]{M\"obius ladder}} \put(4.8,5.1){\makebox(0,0)[l]{cube}} \put(4.18,4.75){\circle*{0.07}} \put(4.3,4.7){\makebox(0,0)[l]{$K_{2,6}$}} \end{picture} \end{center} \caption{Values of two optimality criteria for all equireplicate block designs with $v=8$, $b=12$, and $k=2$, and for $K_{2,6}$} \label{fig:cube} \end{figure} The plotting symbols show the edge-connectivity of the graphs: edge-connectivity $3$, $2$, $1$ is shown as $\boldsymbol{\times}$, $\boldsymbol{+}$, $\boldsymbol{\circ}$ respectively. This does suggest that the higher the edge-connectivity the better is the design on the A- and D- criteria. This is intuitively reasonable: if $k=2$, then the edge-connectivity is the minimum number of blocks whose removal disconnects the design. In this context, it has been called \emph{breakdown number}: see \cite{mahbub}. The four graphs with edge-connectivity $3$ have no double edges, so concurrences differ by at most~$1$. The only other regular graph with no double edges is ranked eighth (amongst regular graphs) by the A-criterion. This suggests that (near-)equality of concurrences is not sufficient to give a good design. The symbol {\setlength{\unitlength}{2cm}\raisebox{1mm}{\circle*{0.07}}} shows the non-regular graph $K_{2,6}$, which also has eight vertices and twelve edges. It is not as good as the regular graphs with edge-connectivity~$3$, but it beats many of the other regular graphs. This pattern is typical of the block designs investigated by statisticians for most of the 20th century. The A- and D-criteria agree closely at the top end. High edge-connectivity appears to show good designs. Many of the best designs have a high degree of symmetry. \end{eg} \section{Highly patterned block designs} \subsection{Balanced incomplete-block designs} \label{sec:bibd} BIBDs are intuitively appealing, as they seem to give equal weight to all treatment comparisons. They were introduced for agricultural experiments by Yates in \cite{fy:ibd} In \cite{ksh}, Kshirsagar proved that, if there exists a BIBD for given values of $v$, $b$ and $k$, then it is A-optimal. Kiefer generalized this in \cite{kiefer} to cover $\Phi_p$-optimality for all $p$ in $(0,\infty)$, including the limiting cases of D- and E-optimality. The core of Kiefer's proof is as follows: binary designs maximize $\mathop{\mathrm{Tr}}(\mathbf{L})$, which is equal to $\sum_{i=1}^{v-1} \theta_i$; for any fixed value $T$ of this sum of positive numbers, $\sum \theta_i^{-p}$ is minimized at $[T/(v-1)]^{-p}$ when $\theta_1 = \cdots = \theta_{v-1} = T/(v-1)$; and $T^{-p}$ is minimized when $T$ is maximized. \subsection{Other special designs} Of course, it frequently occurs that the values of $b,v,k$ available for an experiment are such that no BIBD exists. (Necessary conditions for the exist\-ence of a BIBD include the well-known divisibility conditions $v\mid bk$ and $v(v-1)\mid bk(k-1)$, which follow from the elementary results in Section~\ref{sec:intro}, and \emph{Fisher's inequality} asserting that $b\ge v$.) In the absence of a BIBD, various other special types of design have been considered, and some of these have been proved optimal. Here is a short sample. A design is \emph{group-divisible} if the treatments can be partitioned into ``groups'' all of the same size, so that the number of blocks containing two treatments is $\lambda_1$ if they belong to the same group and $\lambda_2$ otherwise. Ch\^eng~\cite{cheng1,cheng2} showed that if there is a group-divisible design with two groups and $\lambda_2=\lambda_1+1$ in the class of designs with given $v$, $b$, $k$, then it is $\Phi_p$-optimal for all $p$, and in particular is A-, D- and E-optimal. A \emph{regular-graph design} is a binary equireplicate design with two possible concurrences $\lambda$ and $\lambda+1$. It is easily proved that, in such a design, the number of treatments lying in $\lambda+1$ blocks with a given treatment is constant; so the graph $H$ whose vertices are the treatments, two vertices joined if they lie in $\lambda+1$ blocks, is regular. Now Ch\^eng~\cite{cheng2} showed that a group-divisible design with $\lambda_2 = \lambda_1 + 1 = 1$, if one exists, is $\Phi_p$-optimal in the class of regular-graph designs for all~$p$. Cheng and Bailey~\cite{cheng_bailey} showed that a regular-graph design for which the graph is \emph{strongly regular} (see~\cite{pjc:sr}) and which has singular concurrence matrix is $\Phi_p$-optimal, for all~$p$, among binary equireplicate designs with given $v$, $b$, $k$. Designs with the property described here are particular examples of \emph{partially balanced designs} with respect to an association scheme: see Bailey~\cite{rab:as}. Another class which has turns out to be optimal in many cases, but whose definition is less combinatorial, consists of the \emph{variance-balanced designs}, which we consider later in the chapter. \section{Graph concepts linked to D-optimality \label{sec:span} \subsection{Spanning trees of the concurrence graph} Let $G$ be the concurrence graph of a connected block design, and let $\mathbf{L}$ be its Laplacian matrix. A \emph{spanning tree} for $G$ is a spanning subgraph which is a tree. Kirchhoff's famous Matrix-tree theorem in \cite{old} states the following: \begin{thm} \label{thm:span} If $G$ is a connected graph with $v$~vertices and Laplacian matrix~$\mathbf{L}$, then the product of the non-trivial eigenvalues of~$\mathbf{L}$ is equal to $v$ multiplied by the number of spanning trees for $G$. \end{thm} Thus we have a test for D-optimality: \begin{quote} \textit{A design is D-optimal if and only if its concurrence graph has the maximal number of spanning trees.} \end{quote} Note that Theorem~\ref{thm:span} gives an easy proof of Cayley's theorem on the number of spanning trees for the complete graph $K_v$. The non-trivial eigenvalues of its Laplacian matrix are all equal to $v$, so Theorem~\ref{thm:span} shows that it has $v^{v-2}$ spanning trees. If $G$ is sparse, it may be much easier to count the number of spanning trees than to compute the eigenvalues of~$\mathbf{L}$. For example, if $G$ has a single cycle, which has length~$s$, then the number of spanning trees is~$s$, irrespective of the remaining edges in $G$. In the context of optimal block designs, Gaffke discovered the importance of Kirchhoff's theorem in \cite{gaff}. Cheng followed this up in papers such as \cite{cheng1,cheng2,cheng3}. Particularly intriguing is the following theorem from \cite{cmw}. \begin{thm} Consider block designs with $k=2$ (connected graphs). For each given~$v$ there is a threshold $b_0$ such that if $b\geq b_0$ then any D-optimal design for $v$ treatments in $b$ blocks of size~$2$ is \emph{nearly balanced} in the sense that \begin{itemize} \item no pair of replications differ by more than $1$; \item for each fixed $i$, no pair of concurrences $\lambda_{ij}$ differ by more than $1$. \end{itemize} \end{thm} In fact, there is no known example with $b_0 >v-1$, which is the minimal number of blocks required for connectivity. \subsection{Spanning trees of the Levi graph} In \cite{gafLevi} Gaffke stated the following relationship between the numbers of spanning trees in the concurrence graph and the Levi graph. \begin{thm} \label{thm:gaff} Let $G$ and $\tilde G$ be the concurrence graph and Levi graph for a connected incomplete-block design for $v$ treatments in $b$ blocks of size $k$. Then the number of spanning trees for $\tilde G$ is equal to $k^{b-v+1}$ times the number of spanning trees for $G$. \end{thm} Thus, an alternative test for D-optimality is to count the number of spanning trees in the Levi graph. For binary designs, the Levi graph has fewer edges than the concurrence graph if and only if $k\geq 4$. \section{Graph concepts linked to A-optimality} \subsection{The concurrence graph as an electrical network} \label{sec:concelec} We can consider the concurrence graph $G$ as an electrical network with a 1-ohm resistance in each edge. Connect a $1$-volt battery between vertices $i$ and $j$. Then current flows in the network, according to these rules. \begin{description} \item[Ohm's Law:] In every edge, the voltage drop is the product of the current and the resistance. \item[Kirchhoff's Voltage Law:] The total voltage drop from one vertex to any other vertex is the same no matter which path we take from one to the other. \item[Kirchhoff's Current Law:] At each vertex which is not connected to the battery, the total current coming in is equal to the total current going out. \end{description} We find the total current from $i$ to $j$, and then use Ohm's Law to define the effective resistance $R_{ij}$ between $i$ and $j$ as the reciprocal of this current. It is a standard result of electrical network theory that the linear equations implicitly defined above for the currents and voltage differences have a unique solution. Let $\mathcal{T}$ be the set of treatments and $\Omega$ the set of experimental units. Current flows in each edge $e_{\alpha\omega}$, where $\alpha$ and $\omega$ are experimental units in the same block which receive different treatments; let $I(\alpha,\omega)$ be the current from $f(\alpha)$ to $f(\omega)$ in this edge. Thus $I$ is a function $I \colon \Omega \times \Omega \mapsto \mathbb{R}$ such that \begin{enumerate} \item $I(\alpha,\omega) = 0$ if $\alpha=\omega$ or if $f(\alpha) = f(\omega)$ or if $\alpha$ and $\omega$ are in different blocks. \item $I(\alpha,\omega) = -I(\omega,\alpha)$ for $(\alpha,\omega)$ in $\Omega\times \Omega$. \end{enumerate} This defines a further function $I_\mathrm{out} \colon \mathcal{T} \mapsto \mathbb{R}$ by \[ I_\mathrm{out} (l) = \sum_{\alpha: f(\alpha)=l} \ \sum_{\omega\in\Omega} I(\alpha,\omega) \quad \mbox{for $l$ in $\mathcal{T}$}. \] Voltage is another function $V\colon \mathcal{T} \mapsto \mathbb{R}$. The following two conditions ensure that Ohm's and Kirchhoff's Laws are satisfied. \begin{enumerate} \addtocounter{enumi}{2} \item If there is any edge in $G$ between $f(\alpha)$ and $f(\omega)$, then \[I(\alpha,\omega) = V(f(\alpha)) - V(f(\omega)).\] \item If $l\notin\setof{i,j}$, then $I_{\mathrm{out}}(l)=0$. \end{enumerate} If $G$ is connected and different voltages $V(i)$ and $V(j)$ are given for a pair of distinct treatments $i$ and $j$, then there are unique functions $I$ and $V$ satisfying conditions (a)--(d). Moreover, $I_\mathrm{out}(j) = -I_{\mathrm{out}} (i) \ne 0$. Then $R_{ij}$ is defined by \[ R_{ij} = \frac{V(i) - V(j)}{I_{\mathrm{out}}(i)}. \] It can be shown that the value of $R_{ij}$ does not depend on the choice of values for $V(i)$ and $V(j)$, so long as these are different. In practical examples, it is usually convenient to take $V(i)=0$ and let $I$ take integer values. What has all of this got to do with block designs? The following theorem, which is a standard result from electrical engineering, gives the answer. \begin{thm} \label{thm:resist} If $\mathbf{L}$ is the Laplacian matrix of a connected graph $G$, then the effective resistance $R_{ij}$ between vertices $i$ and $j$ is given by \[R_{ij} = \left(L_{ii}^- +L_{jj}^{-} -2L_{ij}^{-}\right).\] \end{thm} Comparing this with Theorem~\ref{thm:var}, we see that $V_{ij} = R_{ij} \times k\sigma^2$. Hence we have a test for A-optimality: \begin{quote} \textit{A design is A-optimal if and only if its concurrence graph, regarded as an electrical network, minimizes the sum of the pairwise effective resistances between all pairs of vertices.} \end{quote} Effective resistances are easy to calculate without matrix inversion if the graph is sparse. \begin{figure} \begin{center} \setlength{\unitlength}{5.8mm} \begin{picture}(15,14)(-6,-7) \put(0,6.67){\circle*{0.4}} \put(0,-6.67){\circle*{0.4}} \put(-6,-3.33){\circle*{0.4}} \put(-2,-3.33){\circle*{0.4}} \put(6,-3.33){\circle*{0.4}} \put(2,-3.33){\circle*{0.4}} \put(-6,3.33){\circle*{0.4}} \put(-2,3.33){\circle*{0.4}} \put(6,3.33){\circle*{0.4}} \put(2,3.33){\circle*{0.4}} \put(-4,0){\circle*{0.4}} \put(4,0){\circle*{0.4}} \put(-6,-3.33){\line(1,0){12}} \put(-6,3.33){\line(1,0){12}} \put(-6,-3.33){\line(3,5){6}} \put(6,-3.33){\line(-3,5){6}} \put(-6,3.33){\line(3,-5){6}} \put(6,3.33){\line(-3,-5){6}} \put(0,-3.33){\vector(1,0){0}} \put(0,-2.9){\makebox(0,0){14}} \put(4,-3.33){\vector(1,0){0}} \put(4,-2.9){\makebox(0,0){7}} \put(-4,-3.33){\vector(-1,0){0}} \put(-4,-2.9){\makebox(0,0){5}} \put(-1,-5){\vector(4,-3){0}} \put(-1.4,-5.2){\makebox(0,0){7}} \put(-3,-1.67){\vector(-4,3){0}} \put(-2.6,-1.4){\makebox(0,0){10}} \put(-5,1.67){\vector(-4,3){0}} \put(-5.4,1.7){\makebox(0,0){5}} \put(1,-5){\vector(4,3){0}} \put(1.4,-5.2){\makebox(0,0){7}} \put(3,-1.67){\vector(4,3){0}} \put(2.6,-1.4){\makebox(0,0){14}} \put(5,1.67){\vector(4,3){0}} \put(5.4,1.7){\makebox(0,0){19}} \put(-5,-1.67){\vector(4,3){0}} \put(-5.4,-1.4){\makebox(0,0){5}} \put(-3,1.67){\vector(4,3){0}} \put(-2.6,1.4){\makebox(0,0){10}} \put(-1,5){\vector(4,3){0}} \put(-1.4,5.2){\makebox(0,0){5}} \put(0,3.33){\vector(1,0){0}} \put(0,2.9){\makebox(0,0){10}} \put(4,3.33){\vector(1,0){0}} \put(4,2.9){\makebox(0,0){17}} \put(-4,3.33){\vector(1,0){0}} \put(-4,2.9){\makebox(0,0){5}} \put(5,-1.67){\vector(-4,3){0}} \put(5.4,-1.4){\makebox(0,0){7}} \put(3,1.67){\vector(-4,3){0}} \put(2.6,1.4){\makebox(0,0){2}} \put(1,5){\vector(4,-3){0}} \put(1.4,5.2){\makebox(0,0){5}} \put(-1.7,-2.83){\makebox(0,0){$i$}} \put(-2.6,-4){\makebox(0,0){$[0]$}} \put(2.6,-4){\makebox(0,0){$[-14]$}} \put(-2.6,4){\makebox(0,0){$[-20]$}} \put(2.6,4){\makebox(0,0){$[-30]$}} \put(-5.1,0){\makebox(0,0){$[-10]$}} \put(5.1,0){\makebox(0,0){$[-28]$}} \put(5.4,2.9){\makebox(0,0){$j$}} \put(6.6,4){\makebox(0,0){$[-47]$}} \put(0,-7.33){\makebox(0,0){$[-7]$}} \put(0,7.33){\makebox(0,0){$[-25]$}} \put(-6.6,-4){\makebox(0,0){$[-5]$}} \put(-6.6,4){\makebox(0,0){$[-15]$}} \put(6.6,-4){\makebox(0,0){$[-21]$}} \end{picture} \end{center} \caption The current between vertices $i$ and $j$ in a concurrence graph} \label{fig:star} \end{figure} Figure~\ref{fig:star} shows the concurrence graph of a block design with $v=12$, $b=6$ and $k=3$. Only vertices $i$ and $j$ are labelled. Otherwise, numbers beside arrows denote current and numbers in square brackets denote voltage. It is straightforward to check that conditions (a)--(d) are satisfied. Now $V(i) - V(j) = 47$ and $I_{\mathrm{out}}(i) = 36$, and so $R_{ij} = 47/36$. Therefore $V_{ij} = (47/12)\sigma^2$. Moreover, for graphs consisting of $b$~triangles arranged in a cycle like this, it is clear that average effective resistance, and hence the average pairwise variance, can be calculated as a function of $b$. \subsection{The Levi graph as an electrical network} \label{sec:levielec} The Levi graph $\tilde G$ of a block design can also be considered as an electrical network. Denote by $\mathcal{B}$ the set of blocks. Now current is defined on the ordered edges of the Levi graph. Recall that, if $\omega$ is an experimental unit in block $\Gamma$, then the edge $\tilde{e}_{\omega}$ joins $\Gamma$ to $f(\omega)$. Thus current is defined on $(\Omega \times \mathcal{B}) \cup (\mathcal{B} \times \Omega)$ and voltage is defined on $\mathcal{T} \cup \mathcal{B}$. Conditions (a)--(d) in Section~\ref{sec:concelec} need to be modified appropriately. The next theorem shows that a current--voltage pair $(I,V)$ on the concurrence graph $G$ can be transformed into a current--voltage pair $(\tilde I, \tilde V)$ on the Levi graph $\tilde G$. In $\tilde G$, the current $\tilde I(\alpha,\Gamma)$ flows in edge $\tilde{e}_\alpha$ from vertex $f(\alpha)$ to vertex $\Gamma$, where $\alpha\in\Gamma$. Hence the pairwise variance $V_{ij}$ can also be calculated from the effective resistance $\tilde R_{ij}$ in the Levi graph. \begin{thm} \label{thm:tjur} Let $G$ be the concurrence graph and $\tilde G$ be the Levi graph of a connected block design with block size~$k$. If\/ $i$ and $j$ are two distinct treatments, let $R_{ij}$ and $\tilde R_{ij}$ be the effective resistance between vertices $i$ and $j$ in the electrical networks defined by $G$ and $\tilde G$, respectively. Then $\tilde R_{ij} = kR_{ij}$, and so $V_{ij} = \tilde R_{ij} \sigma^2$. \end{thm} \begin{pf} Let $(I,V)$ be a current--voltage pair on $G$. For $(\alpha,\Gamma) \in \Omega \times \mathcal{B}$, put \[ \tilde I(\alpha,\Gamma) = -\tilde I(\Gamma,\alpha) = \sum_{\omega\in\Gamma} I(\alpha,\omega) \] if $\alpha\in\Gamma$; otherwise, put $\tilde I(\alpha,\Gamma) = \tilde I(\Gamma,\alpha) = 0$. Put $\tilde V(i) = kV(i)$ for all $i$ in~$\mathcal{T}$, and \[ \tilde V(\Gamma) = \sum_{\omega\in\Gamma} V(f(\omega)) \] for all $\Gamma$ in $\mathcal{B}$. It is clear that $\tilde I$ satisfies the analogues of conditions (a) and~(b). If $\alpha\in\Gamma$, then \begin{eqnarray*} \tilde I(\alpha,\Gamma) &=& \sum_{\omega\in\Gamma} I(\alpha,\omega)\\ & = & \sum_{\omega\in\Gamma}[V(f(\alpha)) - V(f(\omega))]\\ & = & kV(f(\alpha)) - \tilde V(\Gamma) = \tilde V(f(\alpha)) - \tilde V(\Gamma), \end{eqnarray*} so the analogue of condition (c) is satisfied. If $\Gamma\in\mathcal{B}$, then \[ \tilde I_{\mathrm{out}}(\Gamma) = \sum_{\alpha\in\Gamma} \tilde I(\Gamma,\alpha) = -\sum_{\alpha\in\Gamma} \sum_{\omega\in\Gamma} I(\alpha,\omega) = 0, \] because $I(\alpha,\alpha)=0$ and $I(\alpha,\omega) = -I(\omega,\alpha)$. If $l\in \mathcal{T}$ then \[ \tilde I_{\mathrm{out}}(l) = \sum_{\alpha: f(\alpha)=l} \ \sum_{\Gamma\in\mathcal{B}} \tilde I(\alpha,\Gamma) = \sum_{\alpha: f(\alpha)=l}\ \sum_{\omega\in\Omega} I(\alpha,\omega) = I_{\mathrm{out}}(l). \] In particular, $\tilde I_{\mathrm{out}}(l) = 0$ if $l\notin\setof{i,j}$, which shows that the analogue of condition (d) is satisfied. It follows that $(\tilde I,\tilde V)$ is the current--voltage pair on $\tilde G$ defined by $\tilde V(i)$ and $\tilde V(j)$. Now \[ \tilde R_{ij} = \frac{\tilde V(i) - \tilde V(j)}{\tilde I_{\mathrm{out}}(i)} = \frac{k(V(i)-V(j))}{I_{\mathrm{out}}(i)} = kR_{ij}. \] Then Theorems~\ref{thm:var} and~\ref{thm:resist} show that $V_{ij} = R_{ij}\sigma^2$. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} When $k=2$ it seems to be easier to use the concurrence graph than the Levi graph, because it has fewer vertices, but for larger values of $k$ the Levi graph may be better, as it does not have all the within-block cycles that the concurrence graph has. Fig.~\ref{fig:levistar} gives the Levi graph of the block design whose concurrence graph is in Fig.~\ref{fig:star}, with the same two vertices $i$ and $j$ attached to the battery. This gives $\tilde R_{ij} = 47/12$, which is in accordance with Theorem~\ref{thm:tjur}. \begin{figure}[htbp] \begin{center} \setlength{\unitlength}{8mm} \begin{picture}(10,6)(-0.5,0) \multiput(0,1)(2,0){6}{\circle*{0.4}} \multiput(0,3)(2,0){6}{\circle*{0.4}} \multiput(0,5)(2,0){6}{\circle*{0.4}} \put(-0.5,1){\makebox(0,0)[r]{treatments}} \put(-0.5,3){\makebox(0,0)[r]{blocks}} \put(-0.5,5){\makebox(0,0)[r]{treatments}} \multiput(0,1)(2,0){6}{\line(0,1){4}} \put(0,1){\line(5,1){10}} \multiput(2,1)(2,0){5}{\line(-1,1){2}} \multiput(2,2.1)(2,0){3}{\vector(0,1){0}} \put(6,4.4){\vector(0,1){0}} \put(0,1.9){\vector(0,-1){0}} \put(8,1.9){\vector(0,-1){0}} \put(10,1.9){\vector(0,-1){0}} \put(3.1,1.9){\vector(1,-1){0}} \put(4.7,2.3){\vector(1,-1){0}} \put(0.9,2.1){\vector(-1,1){0}} \put(6.9,2.1){\vector(-1,1){0}} \put(8.9,2.1){\vector(-1,1){0}} \put(4.7,1.95){\vector(4,1){0}} \put(0,0.7){\makebox(0,0)[t]{$[-10]$}} \put(2,0.7){\makebox(0,0)[t]{$[0]$}} \put(2,0){\makebox(0,0)[t]{$i$}} \put(4,0.7){\makebox(0,0)[t]{$[-14]$}} \put(6,0.7){\makebox(0,0)[t]{$[-28]$}} \put(8,0.7){\makebox(0,0)[t]{$[-30]$}} \put(10,0.7){\makebox(0,0)[t]{$[-20]$}} \put(0.7,3.3){\makebox(0,0){$[-5]$}} \put(2.7,3.3){\makebox(0,0){$[-7]$}} \put(4.7,3.3){\makebox(0,0){$[-21]$}} \put(6.7,3.3){\makebox(0,0){$[-35]$}} \put(8.7,3.3){\makebox(0,0){$[-25]$}} \put(10.7,3.3){\makebox(0,0){$[-15]$}} \put(0,5.3){\makebox(0,0)[b]{$[-5]$}} \put(2,5.3){\makebox(0,0)[b]{$[-7]$}} \put(4,5.3){\makebox(0,0)[b]{$[-21]$}} \put(6,5.3){\makebox(0,0)[b]{$[-47]$}} \put(6,6){\makebox(0,0)[b]{$j$}} \put(8,5.3){\makebox(0,0)[b]{$[-25]$}} \put(10,5.3){\makebox(0,0)[b]{$[-15]$}} \put(-0.3,2){\makebox(0,0){5}} \put(0.7,2){\makebox(0,0){5}} \put(1.7,2){\makebox(0,0){7}} \put(2.7,2){\makebox(0,0){7}} \put(3.7,2){\makebox(0,0){7}} \put(4.6,1.6){\makebox(0,0){5}} \put(4.8,2.5){\makebox(0,0){7}} \put(6.3,2){\makebox(0,0){7}} \put(7.3,2){\makebox(0,0){5}} \put(8.3,2){\makebox(0,0){5}} \put(9.3,2){\makebox(0,0){5}} \put(10.3,2){\makebox(0,0){5}} \put(5.5,4.3){\makebox(0,0){12}} \end{picture} \end{center} \caption{Current between $i$ and $j$ for the Levi graph corresponding to the concurrence graph in Fig.~\ref{fig:star}} \label{fig:levistar} \end{figure} Here is another way of visualizing Theorem~\ref{thm:tjur}. From the block design we construct a graph $G_0$ with vertex-set $\mathcal{T} \cup \Omega \cup\mathcal{B}$: the edges are $\setof{\alpha,\Gamma}$ for $\alpha\in\Gamma \in \mathcal{B}$ and $\setof{\alpha,f(\alpha)}$ for $\alpha\in\Omega$. Let $(I_0,V_0)$ be a current--voltage pair on $G_0$ for which both battery vertices are in $\mathcal{T}$. We obtain the Levi graph $\tilde G$ from $G_0$ by becoming blind to the vertices in~$\Omega$. Thus the resistance in each edge of $\tilde G$ is twice that in each edge in $G_0$, so this step multiplies each effective resistance by $2$. Because none of the battery vertices is in~$\mathcal{B}$, we can now obtain $G$ from $\tilde G$ by replacing each path of the form $(i,\Gamma,j)$ by an edge $(i,j)$. There is no harm in scaling all the voltages by the same amount, so we can obtain $(I,V)$ on $G$ from $(\tilde I, \tilde V)$ on $\tilde G$ by putting $V(i) = \tilde V(i)/k$ for $i$ in $\mathcal{T}$, and $I(\alpha,\omega) = V(f(\alpha)) - V(f(\omega))$ for $\alpha$, $\omega$ in the same block. If $\Gamma$ is a block, then \[ 0 = \sum_{\alpha\in\Gamma} \tilde I(\alpha,\Gamma) = \sum_{\alpha\in\Gamma}[\tilde V(f(\alpha)) - \tilde V(\Gamma)] = k\sum_{\alpha\in\Gamma} V(f(\alpha)) - k\tilde V(\Gamma), \] and so $\tilde V(\Gamma) = \sum_{\alpha\in\Gamma} V(f(\alpha))$. Also, if $\alpha\in\Gamma$, then \begin{eqnarray*} \sum_{\omega\in\Gamma} I(\alpha,\omega)& =& \sum_{\omega\in\Gamma}[V(f(\alpha)) - V(f(\omega))]\\ &=& kV(f(\alpha)) - \sum_{\omega\in\Gamma}V(f(\omega)) \\ &=& \tilde V(f(\alpha)) - \tilde V (\Gamma)\\& =& \tilde I(\alpha,\Gamma). \end{eqnarray*} Therefore, this transformation reverses the one used in the proof of Theorem~\ref{thm:tjur}. There is yet another way of obtaining Theorem~\ref{thm:tjur}. If we use the responses $Y_\omega$ to estimate the block parameters $\beta_\Gamma$ in (\ref{eq:linmod}) as well as the treatment parameters $\tau_i$, then standard theory of linear models shows that, if the design is connected, then we can estimate linear combinations of the form $\sum_{i=1}^v x_i \tau_i + \sum_{j=1}^b z_j\beta_j$ so long as $\sum x_i = \sum z_j$. Moreover, the variance of the BLUE of this linear combination is \[ [\begin{array}{cc}\mathbf{x}^\top & \mathbf{z}^\top\end{array}] \mathbf{C}^- \left[ \begin{array}{c}\mathbf{x} \\ \mathbf{z}\end{array} \right] \sigma^2, \qquad\mbox{where} \qquad \mathbf{C} = \left[ \begin{array}{lc} \mathbf{R} & \mathbf{N}\\ \mathbf{N}^\top & k\mathbf{I}_b \end{array} \right] \] and $\mathbf{R}$ is the diagonal matrix of replications. If we reparametrize equation~(\ref{eq:linmod}) by replacing $\beta_j$ by $-\gamma_j$ for $j=1$, \ldots, $b$, then the estimable quantities are the contrasts in $\tau_1$, \ldots, $\tau_v$, $\gamma_1$, \ldots, $\gamma_b$. The so-called \emph{information matrix} $\mathbf{C}$ must be modified by multiplying the last $b$~rows and the last $b$~columns by $-1$: this gives precisely the Laplacian $\tilde \mathbf{L}$ of the Levi graph $\tilde G$. Just as for~$\mathbf{L}$, but unlike $\mathbf{C}$, the null space is spanned by the all-$1$ vector. \subsection{Spanning thickets} We have seen that the value of the D-criterion is a function of the number of spanning trees of the concurrence graph~$G$. It turns out that the closely related notion of a spanning thicket enables us to calculate the A-criterion; more precisely, the value of each pairwise effective resistance in $G$. A \emph{spanning thicket} for the graph is a spanning subgraph that consists of two trees (one of them may be an isolated vertex). \begin{thm} \label{thm:thick} If $i$ and $j$ are distinct vertices of $G$ then \[ R_{ij} = \frac{\mbox{\normalfont{number of spanning thickets with $i$, $j$ in different parts}}}{\mbox{\normalfont{number of spanning trees}}} \ . \] \end{thm} This is also rather easy to calculate directly when the graph is sparse. Summing all the $R_{ij}$ and using Theorem~\ref{thm:thick} gives the following result from \cite{shap}. \begin{thm If $F$ is a spanning thicket for the concurrence graph~$G$, denote by $F_1$ and $F_2$ the sets of vertices in its two trees. Then \[ \sum_{i< j} R_{ij} =\frac{\displaystyle \sum_{\mathrm{spanning\ thickets}\ F} \left| F_1 \right| \left| F_2\right|}{\mbox{\normalfont{number of spanning trees}}} \] \end{thm} \subsection{Random walks and electrical networks} It was first pointed out by Kakutani in 1945 that there is a very close connection between random walks and electrical networks. In a simple random walk, a single step works as follows: starting at a vertex, we choose an edge containing the vertex at random, and move along it to the other end. This definition accommodates multiple edges, and is easily adapted to graphs with edge weights (where the probability of moving along an edge is proportional to the weight of the edge). If we are thinking of an edge-weighted graph as an electrical network, we take the weights to be the conductances of the edges (the reciprocals of the resistances). The connection is simple to state: \begin{thm} Let $i$ and $j$ be distinct vertices of the connected edge-weighted graph $G$. Apply voltages of $1$ at $i$ and $0$ at $j$. Then the voltage at a vertex $l$ is equal to the probability that the random walk, starting at $l$, reaches $i$ before it reaches $j$. \end{thm} From this theorem, it is possible to derive a formula for the effective resistance between two vertices. Here are two such formulas. Given two vertices $i$ and $j$, let $P_\mathrm{esc}(i\to j)$ be the probability that a random walk starting at $i$ reaches $j$ before returning to $i$; and let $S_i(i,j)$ be the expected number of times that a random walk starting at $i$ visits $i$ before reaching $j$. Then the effective resistance between $i$ and $j$ is given by either of the two expressions \[ \frac{1}{d_iP_\mathrm{esc}(i\to j)}\qquad\mbox{and}\qquad \frac{S_i(i,j)}{d_i}, \] where $d_i$ is the degree of $i$. (If the edge resistances are not all $1$, then the term $d_i$ should be replaced by the sum of the reciprocals of the resistances of all edges incident with vertex $i$.) The random walk approach gives alternative proofs of some of the main results about electrical networks. We discuss this further in the guide to the literature. \subsection{Foster's formula and generalizations} In 1948, Foster~\cite{foster} discovered that the sum of the effective resistances between all \emph{adjacent} pairs of vertices of a connected graph on $v$ vertices is equal to $v-1$. Thirteen years later, he found a similar formula for pairs of vertices at distance~$2$: \[\sum_{i\sim h\sim j}\frac{R_{ij}}{d_h} = v-2.\] Further extensions have been found, but require a stronger condition on the graph. The sum of resistances between all pairs of vertices at distance at most~$m$ can be written down explicitly if the graph is \emph{walk-regular up to distance $m$}; this means that the number of closed walks of length $k$ starting and finishing at a vertex $i$ is independent of $i$, for $k\le m$. The formula was discovered by Emil Vaughan, to whom this part of the chapter owes a debt. In particular, if the graph is distance-regular (see \cite{bcn}), then the value of the A-criterion can be written down in terms of the so-called \emph{intersection array} of the graph. \subsection{Distance} At first sight it seems obvious that pairwise variance should decrease as concurrence increases, but there are many counter-examples to this. However, the following theorem is proved in \cite{rab:as}. \begin{thm} If the Laplacian matrix $\mathbf{L}$ has precisely two distinct non-trivial eigenvalues, then pairwise variance is a decreasing linear function of concurrence. \end{thm} It does appear that effective resistance, and hence pairwise variance, generally increases with distance in the concurrence graph. In \cite[Question 5.1]{bcc09} we pointed out that this is not always exactly so, and asked if it is nevertheless true that the maximal value of $R_{ij}$ is achieved for some pair of vertices $\setof{i,j}$ whose distance apart in the graph is maximal. Here is a counter-example. \begin{eg} Let $k=2$, so that the block design is the same as its concurrence graph. Take $v=10$ and $b=14$. The graph consists of a cube, with two extra vertices $1$ and $2$ attached as leaves to vertex $3$. The vertex antipodal to~$3$ in the cube is labelled~$4$. It is straightforward to check (either using an electrical network, or by using the fact that the cube is distance-regular) that the effective resistance between a pair of cube vertices is $7/12$, $3/4$ and $5/6$ for vertices at distances $1$, $2$ and $3$. Hence $R_{1j}\leq 11/6$ for all cube vertices~$j$, while $R_{12}=2$. On the other hand, the distance between vertices $1$ and $2$ is only~$2$, while that between either of them and vertex~$4$ is $4$. \end{eg} There are some `nice' graphs where pairwise variance does indeed increase with distance. The following result is proved in \cite{rab:dgh}. Biggs gave the equivalent result for effective resistances in \cite{biggselec}. \begin{thm} Suppose that a block design has just two distinct concurrences, and that the pairs of vertices corresponding to the larger concurrence form the edges of a distance-regular graph $H$. Then pairwise variance increases with distance in $H$. \end{thm} \section{Graph concepts linked to E-optimality} \subsection{Measures of bottlenecks} A `good' graph (for use as a network) is one without bottlenecks: any set of vertices should have many edges joining it to its complement. So, for any subset $S$ of vertices, we let $\partial(S)$ (the \emph{boundary} of $S$) be the set of edges which have one vertex in $S$ and the other in its complement, and then define the \emph{isoperimetric number} $\iota(G)$ by \[ \iota(G) = \min\left\{\frac{|\partial S|}{|S|}: S\subseteq V(G),\ 0<|S|\le \frac{v}{2} \right\}. \] The next result shows that the isoperimetric number is related to the E-criterion. It is useful not so much for identifying the E-optimal designs as for easily showing that large classes of designs cannot be E-optimal: any design whose concurrence graph has low isoperimetric number performs poorly on the E-criterion. \begin{cutset} Let $G$ have an edge-cutset of size~$c$ whose removal separates the graph into parts $S$ and $G\setminus S$ with $m$ and $n$ vertices respectively, where $0 < m\leq n$. Then \[ \theta_1\le c\left(\frac{1}{m}+\frac{1}{n}\right) \leq \frac{2\left|\partial S\right|}{\left|S\right|}. \] \end{cutset} \begin{pf} We know that $\theta_1$ is the minimum of $\mathbf{x}^\top \mathbf{L}\mathbf{x}/\mathbf{x}^\top \mathbf{x}$ over real vectors $\mathbf{x}$ with $\sum_i x_i = 0$. Put \[x_i= \left\{\begin{array}{rl}n & \mbox{ if $i\in S$}\\ -m & \mbox{otherwise.} \end{array}\right.\] Then $\mathbf{x}^\top \mathbf{x} = nm(m+n)$ and \[\mathbf{x}^\top \mathbf{L} \mathbf{x} = \sum_{\mathrm{edges\ }ij} (x_i-x_j)^2 = c(m+n)^2.\] Hence \[\theta_1 \leq \frac{\mathbf{x}^\top \mathbf{L} \mathbf{x}}{\mathbf{x}^\top \mathbf{x}} = \frac{c(m+n)^2}{nm(m+n)} = c\left(\frac{1}{m} + \frac{1}{n}\right) \leq \frac{2c}{m} = \frac{2\left|\partial S\right|}{\left|S\right|}. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \] \end{pf} \begin{cor} \label{thm:iso} Let $\theta_1$ be the smallest non-trivial eigenvalue of the Laplacian matrix $\mathbf{L}$ of the connected graph $G$. Then $\theta_1 \leq 2 \iota(G)$. \end{cor} There is also an upper bound for the isoperimetric number in terms of $\theta_1$, which is loosely referred to as a `Cheeger-type inequality'; for details, see the further reading. We also require a second cutset lemma, phrased in terms of vertex cutsets. \begin{cutset} Let $G$ have a vertex-cutset~$C$ of size $c$ whose removal separates the graph into parts $S$ and $T$ with $m$, $n$ vertices respectively (so $nm>0$). Let $m'$ and $n'$ be the number of edges from vertices in $C$ to vertices in $S$, $T$ respectively. Then \[ \theta_1 \leq \frac{m'n^2 + n'm^2}{nm(m+n)}. \] In particular, if there are no multiple edges at any vertex of $C$ then $\theta_1\leq c$, with equality if and only if every vertex in $C$ is joined to every vertex in $S\cup T$. \end{cutset} \begin{pf} Put \[x_i= \left\{\begin{array}{rl}n & \mbox{ if $i\in S$}\\ -m & \mbox{ if $i\in T$}\\ 0 & \mbox{ otherwise.} \end{array}\right.\] Then $\mathbf{x}^\top \mathbf{x} = nm(m+n)$ and $\mathbf{x}^\top \mathbf{L}\mathbf{x} = m'n^2+n'm^2$, and so \[ \theta_1 \leq \frac{m'n^2 + n'm^2}{nm(m+n)}. \] If there are no multiple edges at any vertex in $C$ then $m'\leq cm$ and $n'\leq cn$ and the result follows. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} \subsection{Variance balance} A block design is \emph{variance-balanced} if all the concurrences $\lambda_{ij}$ are equal for $i\ne j$. In such a design, all of the pairwise variances $V_{ij}$ are equal. Morgan and Srivastav proved the following result in \cite{ms}. \begin{thm} If the constant concurrence $\lambda$ of a variance-balanced design satisfies $(v-1)\lambda = \lfloor(bk/v)\rfloor (k-1)$ then the design is E-optimal. \end{thm} A block with $k$ different treatments contributes $k(k-1)/2$ edges to the concurrence graph. Let us define the \emph{defect} of a block to be \[ \frac{k(k-1)}{2} - \mbox{the number of edges it contributes to the graph.} \] The following result is proved in \cite{bcc09}. \begin{thm} \label{thm:defect} If $k<v$, then a variance-balanced design with $v$ treatments is E-optimal if the sum of the block defects is less than $v/2$. \end{thm} Table~\ref{tab:binary}(b) shows that the design in Fig.~\ref{fig:binary}(b) is variance-balanced. Block $\Gamma_1$ has defect~$1$, and each other block has defect $0$, so the sum of the block defects is certainly less than $5/2$, and Theorem~\ref{thm:defect} shows that the design is E-optimal. It is rather counter-intuitive that the non-binary design in Fig.~\ref{fig:binary}(b) can be better than the design in Fig.~\ref{fig:binary}(a); in fact, in his contribution to the discussion of Tocher's paper \cite{tocher}, which introduced this design, David Cox said \begin{quote} I suspect that \ldots\ balanced ternary designs are of no practical value. \end{quote} Computation shows that the design in Fig.~\ref{fig:binary}(a) is $\Phi_p$-better than the one in Fig.~\ref{fig:binary}(b) if $p<5.327$. In particular, it is A- and D-better. \section{Some history} As we have seen, if the experimental units form a single block and there are only two treatments then it is best for their replications to be as equal as possible. Statisticians know this so well that it is hard for us to imagine that more information may be obtained, about \emph{all} treatment comparisons, if replications differ by more than~$1$. In agriculture, or in any area with qualitative treatments, A-optimality is the natural criterion. If treatments are quantities of different substances, then D-optimality is preferable, as the ranking on this criterion is invariant to change of measurement units. Thus industrial statisticians have tended to prefer D-optimality, although E-optimality has become popular among chemical process engineers. Perhaps the different camps have not talked to each other as much as they should have. For most of the 20th century, it was normal practice in field experiments to have all treatments replicated three or four times. Where incomplete blocks were used, they typically had size from $3$ to $20$. Yates introduced his square lattice designs with $v=k^2$ in \cite{fy:lattice}. He used uniformity data and two worked examples to show that these designs can give lower average pairwise variance than a design using a highly replicated control treatment, but both of his examples were equireplicate with $r\in\{3,4\}$. In the 1930s, 1940s and 1950s, analysis of the data from an experiment involved inverting the Laplacian matrix without a computer: this is easy for BIBDs, and only slightly harder if the Laplacian matrix has only two distinct non-trivial eigenvalues. The results in \cite{kiefer} and \cite{ksh} encouraged the beliefs that the optimal designs, on all $\Phi_p$-criteria, are as equireplicate as possible, with concurrences as equal as possible, and that the same designs are optimal, or nearly so, on all of these criteria. Three short papers in the same journal in 1977--1982 demonstrate the beliefs at that time. In \cite{JAJM:rgd}, John and Mitchell did not even consider designs with unequal replication. They conjectured that, if there exist any regular-graph designs for given values of $v$, $b$ and $k$, then the A- and D-optimal designs are regular-graph designs. For the parameter sets which they had examined by computer search, the same designs were optimal on the A- and D-criteria. In \cite{BJJAE}, Jones and Eccleston reported the results of various computer searches for A-optimal designs without the constraint of equal replication. For $k=2$ and $b=v\in\{10,11,12\}$ (but not $v=9$) their A-optimal design is almost a queen-bee design, and their designs are D-worse than those in \cite{JAJM:rgd}. The belief in equal replication was so ingrained that some readers assumed that there was an error in their program. John and Williams followed this with the paper \cite{yyy} on conjectures for optimal block designs for given values of $v$, $b$ and $k$. Their conjectures included: \begin{itemize} \item the set of regular-graph designs always contains one that is optimal without this restriction; \item among regular-graph designs, the same designs are optimal on the A- and D-criteria. \end{itemize} They endorsed Cox's dismissal of non-binary designs, strengthening it to the statement that they ``are inefficient'', and declared that the three unequally replicated A-optimal designs in \cite{BJJAE} were ``of academic rather than of practical interest''. These conjectures and opinions seemed quite reasonable to people who had been finding good designs for the sizes needed in agricultural experiments. At the end of the 20th century, there was an explosion in the number of experiments in genomics, using microarrays. Simplifying the story greatly, these are effectively block designs with $k=2$, and biologists wanted A-optimal designs, but they did not know the vocabulary `block' or `A-optimal', `graph' or `cycle'. Computers were now much more powerful than in 1980, and researchers in genomics could simply undertake computer searches without the benefit of any statistical theory. In 2001, Kerr and Churchill \cite{KCh} published the results of a computer search for A-optimal designs with $k=2$ and $v=b\leq 11$. For $v\in\{10,11,12\}$, their results were completely consistent with those in \cite{BJJAE}, which they did not cite. They called cycles \textit{loop designs}. Mainstream statisticians began to get involved. In 2005, Wit, Nobile and Khanin published the paper \cite{Wit} giving the results of a computer search for A- and D-optimal designs with $k=2$ and $v=b$. The results are shown in Fig.~\ref{fig:wit}. The A-optimal designs differ from the D-optimal designs when $v\geq 9$, but are consistent with those found in \cite{KCh}. \begin{figure} \begin{center} \begin{tabular}{|c|c|c|} \hline \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(42.6,73.7){\circle*{7}} \put(-42.6,73.7){\circle*{7}} \put(-85.1,0){\circle*{7}} \put(-42.6,-73.7){\circle*{7}} \put(42.6,-73.7){\circle*{7}} \qbezier(85.1,0)(63.85,36.85)(42.6,73.7) \qbezier(42.6,73.7)(0,73.7)(-42.6,73.7) \qbezier(-85.1,0)(-63.85,36.85)(-42.6,73.7) \qbezier(85.1,0)(63.85,-36.85)(42.6,-73.7) \qbezier(42.6,-73.7)(0,-73.7)(-42.6,-73.7) \qbezier(-85.1,0)(-63.85,-36.85)(-42.6,-73.7) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(53.1,66.5){\circle*{7}} \put(-18.9,83){\circle*{7}} \put(-76.7,36.9){\circle*{7}} \put(-76.7,-36.9){\circle*{7}} \put(-18.9,-83){\circle*{7}} \put(53.1,-66.5){\circle*{7}} \qbezier(85.1,0)(69.1,33.25)(53.1,66.5) \qbezier(53.1,66.5)(17.1,74.75)(-18.9,83) \qbezier(-18.9,83)(-47.8,59.95)(-76.7,36.9) \qbezier(-76.7,36.9)(-76.7,0)(-76.7,-36.9) \qbezier(-18.9,-83)(-47.8,-59.95)(-76.7,-36.9) \qbezier(53.1,-66.5)(17.1,-74.75)(-18.9,-83) \qbezier(85.1,0)(69.1,-33.25)(53.1,-66.5) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(60.2,60.2){\circle*{7}} \put(0,85.1){\circle*{7}} \put(-60.2,60.2){\circle*{7}} \put(-85.1,0){\circle*{7}} \put(-60.2,-60.2){\circle*{7}} \put(0,-85.1){\circle*{7}} \put(60.2,-60.2){\circle*{7}} \qbezier(85.1,0)(72.75,30.1)(60.2,60.2) \qbezier(60.2,60.2)(30.1,72.75)(0,85.1) \qbezier(-85.1,0)(-72.75,30.1)(-60.2,60.2) \qbezier(-60.2,60.2)(-30.1,72.75)(0,85.1) \qbezier(85.1,0)(72.75,-30.1)(60.2,-60.2) \qbezier(60.2,-60.2)(30.1,-72.75)(0,-85.1) \qbezier(-85.1,0)(-72.75,-30.1)(-60.2,-60.2) \qbezier(-60.2,-60.2)(-30.1,-72.75)(0,-85.1) \end{picture} \\ \hline \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(65.2,54.7){\circle*{7}} \put(14.8,83.8){\circle*{7}} \put(-42.6,73.7){\circle*{7}} \put(-80,29.1){\circle*{7}} \put(-80,-29.1){\circle*{7}} \put(-42.6,-73.7){\circle*{7}} \put(14.8,-83.8){\circle*{7}} \put(65.2,-54.7){\circle*{7}} \qbezier(85.1,0)(75.15,27.35)(65.2,54.7) \qbezier(65.2,54.7)(40,69.25)(14.8,83.8) \qbezier(14.8,83.8)(-13.9,78.75)(-42.6,73.7) \qbezier(-42.6,73.7)(-62.3,51.4)(-80,29.1) \qbezier(-80,29.1)(-80,0)(-80,-29.1) \qbezier(85.1,0)(75.15,-27.35)(65.2,-54.7) \qbezier(65.2,-54.7)(40,-69.25)(14.8,-83.8) \qbezier(14.8,-83.8)(-13.9,-78.75)(-42.6,-73.7) \qbezier(-42.6,-73.7)(-62.3,-51.4)(-80,-29.1) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-31.2) \put(50,0){\circle*{7}} \put(80.9,95.1){\circle*{7}} \put(0,153.9){\circle*{7}} \put(-80.9,95.1){\circle*{7}} \put(-50,0){\circle*{7}} \put(0,-16.3){\circle*{7}} \put(80.9,42.5){\circle*{7}} \put(50,137.6){\circle*{7}} \put(-50,137.6){\circle*{7}} \put(-80.9,42.5){\circle*{7}} \qbezier(50,0)(65.45,21.25)(80.9,42.5) \qbezier(-50,0)(-65.45,21.25)(-80.9,42.5) \qbezier(80.9,95.1)(65.45,118.35)(50,137.6) \qbezier(-80.9,95.1)(-65.45,118.35)(-50,137.6) \qbezier(0,153.9)(-25,145.75)(-50,137.6) \qbezier(0,153.9)(25,145.75)(50,137.6) \qbezier(-80.9,95.1)(-80.9,68.8)(-80.9,42.5) \qbezier(80.9,95.1)(80.9,68.8)(80.9,42.5) \qbezier(0,-16.3)(25,-8.15)(50,0) \qbezier(0,-16.3)(-25,-8.15)(-50,0) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(71.6,46){\circle*{7}} \put(35.4,77.4){\circle*{7}} \put(-12.1,84.2){\circle*{7}} \put(-55.7,64.3){\circle*{7}} \put(-81.7,24){\circle*{7}} \put(71.6,-46){\circle*{7}} \put(35.4,-77.4){\circle*{7}} \put(-12.1,-84.2){\circle*{7}} \put(-55.7,-64.3){\circle*{7}} \put(-81.7,-24){\circle*{7}} \qbezier(85.1,0)(78.35,23)(71.6,46) \qbezier(71.6,46)(53.5,61.7)(35.4,77.4) \qbezier(35.4,77.4)(11.65,80.8)(-12.1,84.2) \qbezier(-12.1,84.2)(-33.9,74.25)(-55.7,64.3) \qbezier(-55.7,64.3)(-68.7,44.15)(-81.7,24) \qbezier(-81.7,24)(-81.7,0)(-81.7,-24) \qbezier(85.1,0)(78.35,-23)(71.6,-46) \qbezier(71.6,-46)(53.5,-61.7)(35.4,-77.4) \qbezier(35.4,-77.4)(11.65,-80.8)(-12.1,-84.2) \qbezier(-12.1,-84.2)(-33.9,-74.25)(-55.7,-64.3) \qbezier(-55.7,-64.3)(-68.7,-44.15)(-81.7,-24) \end{picture} \\ \hline \multicolumn{3}{c}{\rule[-20pt]{0pt}{40pt}(a) D-optimal designs}\\ \hline \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(42.6,73.7){\circle*{7}} \put(-42.6,73.7){\circle*{7}} \put(-85.1,0){\circle*{7}} \put(-42.6,-73.7){\circle*{7}} \put(42.6,-73.7){\circle*{7}} \qbezier(85.1,0)(63.85,36.85)(42.6,73.7) \qbezier(42.6,73.7)(0,73.7)(-42.6,73.7) \qbezier(-85.1,0)(-63.85,36.85)(-42.6,73.7) \qbezier(85.1,0)(63.85,-36.85)(42.6,-73.7) \qbezier(42.6,-73.7)(0,-73.7)(-42.6,-73.7) \qbezier(-85.1,0)(-63.85,-36.85)(-42.6,-73.7) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(53.1,66.5){\circle*{7}} \put(-18.9,83){\circle*{7}} \put(-76.7,36.9){\circle*{7}} \put(-76.7,-36.9){\circle*{7}} \put(-18.9,-83){\circle*{7}} \put(53.1,-66.5){\circle*{7}} \qbezier(85.1,0)(69.1,33.25)(53.1,66.5) \qbezier(53.1,66.5)(17.1,74.75)(-18.9,83) \qbezier(-18.9,83)(-47.8,59.95)(-76.7,36.9) \qbezier(-76.7,36.9)(-76.7,0)(-76.7,-36.9) \qbezier(-18.9,-83)(-47.8,-59.95)(-76.7,-36.9) \qbezier(53.1,-66.5)(17.1,-74.75)(-18.9,-83) \qbezier(85.1,0)(69.1,-33.25)(53.1,-66.5) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(200,200)(-100,-100) \put(85.1,0){\circle*{7}} \put(60.2,60.2){\circle*{7}} \put(0,85.1){\circle*{7}} \put(-60.2,60.2){\circle*{7}} \put(-85.1,0){\circle*{7}} \put(-60.2,-60.2){\circle*{7}} \put(0,-85.1){\circle*{7}} \put(60.2,-60.2){\circle*{7}} \qbezier(85.1,0)(72.75,30.1)(60.2,60.2) \qbezier(60.2,60.2)(30.1,72.75)(0,85.1) \qbezier(-85.1,0)(-72.75,30.1)(-60.2,60.2) \qbezier(-60.2,60.2)(-30.1,72.75)(0,85.1) \qbezier(85.1,0)(72.75,-30.1)(60.2,-60.2) \qbezier(60.2,-60.2)(30.1,-72.75)(0,-85.1) \qbezier(-85.1,0)(-72.75,-30.1)(-60.2,-60.2) \qbezier(-60.2,-60.2)(-30.1,-72.75)(0,-85.1) \end{picture} \\ \hline \setlength{\unitlength}{0.4pt} \begin{picture}(130,130)(-65,-65) \put(0,0){\circle*{7}} \put(-50,0){\circle*{7}} \put(50,0){\circle*{7}} \put(-50,-50){\circle*{7}} \put(0,-50){\circle*{7}} \put(0,50){\circle*{7}} \put(43.3,25){\circle*{7}} \put(43.3,-25){\circle*{7}} \put(25,43.3){\circle*{7}} \qbezier(0,0)(-25,0)(-50,0) \qbezier(0,0)(0,-25)(0,-50) \qbezier(0,0)(0,25)(0,50) \qbezier(0,0)(25,0)(50,0) \qbezier(-50,-50)(-50,-25)(-50,0) \qbezier(-50,-50)(-25,-50)(0,-50) \qbezier(0,0)(21.65,-12.5)(43.3,-25) \qbezier(0,0)(21.65,12.5)(43.3,25) \qbezier(0,0)(12.5,21.65)(25,43.3) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(130,130)(-65,-65) \put(0,0){\circle*{7}} \put(-50,0){\circle*{7}} \put(50,0){\circle*{7}} \put(-50,-50){\circle*{7}} \put(0,-50){\circle*{7}} \put(0,50){\circle*{7}} \put(43.3,25){\circle*{7}} \put(43.3,-25){\circle*{7}} \put(25,43.3){\circle*{7}} \put(-25,43.3){\circle*{7}} \qbezier(0,0)(-25,0)(-50,0) \qbezier(0,0)(0,-25)(0,-50) \qbezier(0,0)(0,25)(0,50) \qbezier(0,0)(25,0)(50,0) \qbezier(-50,-50)(-50,-25)(-50,0) \qbezier(-50,-50)(-25,-50)(0,-50) \qbezier(0,0)(21.65,-12.5)(43.3,-25) \qbezier(0,0)(21.65,12.5)(43.3,25) \qbezier(0,0)(12.5,21.65)(25,43.3) \qbezier(0,0)(-12.5,21.65)(-25,43.3) \end{picture} & \setlength{\unitlength}{0.4pt} \begin{picture}(130,130)(-65,-65) \put(0,0){\circle*{7}} \put(-50,0){\circle*{7}} \put(50,0){\circle*{7}} \put(-50,-50){\circle*{7}} \put(0,-50){\circle*{7}} \put(0,50){\circle*{7}} \put(43.3,25){\circle*{7}} \put(43.3,-25){\circle*{7}} \put(25,43.3){\circle*{7}} \put(-25,43.3){\circle*{7}} \put(-43.3,25){\circle*{7}} \qbezier(0,0)(-25,0)(-50,0) \qbezier(0,0)(0,-25)(0,-50) \qbezier(0,0)(0,25)(0,50) \qbezier(0,0)(25,0)(50,0) \qbezier(-50,-50)(-50,-25)(-50,0) \qbezier(-50,-50)(-25,-50)(0,-50) \qbezier(0,0)(21.65,-12.5)(43.3,-25) \qbezier(0,0)(21.65,12.5)(43.3,25) \qbezier(0,0)(12.5,21.65)(25,43.3) \qbezier(0,0)(-12.5,21.65)(-25,43.3) \qbezier(0,0)(-21.65,12.5)(-43.3,25) \end{picture} \\ \hline \multicolumn{3}{c}{\rule[-0pt]{0pt}{20pt}(b) A-optimal designs}\\ \end{tabular} \end{center} \caption{D-and A-optimal designs with $k=2$ and $6\leq v=b\leq 11$} \label{fig:wit} \end{figure} What is going on here? Why are the designs so different when $v\geq 9$? Why is there such a sudden, large change in the A-optimal designs? We explain this in the next section. \section{Block size two} \subsection{Least replication} \label{sec:least} If $k=2$, then the design is the same as its concurrence graph, and connectivity requires that $b\geq v-1$. If $b=v-1$, then all connected designs are trees, such as those in Fig.~\ref{fig:tree}. Theorem~\ref{thm:span} shows that the D-criterion does not differentiate between them. In a tree, the effective resistance $R_{ij}$ is just the length of the unique path between vertices $i$ and $j$. Theorems~\ref{thm:var} and~\ref{thm:resist} show that the only A-optimal designs are the stars, such as the graph on the right of Fig.~\ref{fig:tree}. \begin{figure} \begin{center} \begin{tabular}{c@{\qquad\qquad}c} \setlength{\unitlength}{0.8pt} \begin{picture}(100,100) \put(50,50){\circle*{7}} \put(50,100){\circle*{7}} \put(50,0){\circle*{7}} \put(0,50){\circle*{7}} \put(100,50){\circle*{7}} \put(83.3,83.3){\circle*{7}} \put(16.7,83.3){\circle*{7}} \put(83.3,16.7){\circle*{7}} \put(16.7,16.7){\circle*{7}} \put(50,0){\line(-2,1){33.3}} \put(0,50){\line(1,0){50}} \put(0,50){\line(1,-2){16.7}} \put(100,50){\line(-1,-2){16.7}} \put(100,50){\line(-1,2){16.7}} \put(50,100){\line(-2,-1){33.3}} \put(16.7,83.3){\line(1,-1){67}} \end{picture} & \setlength{\unitlength}{0.8pt} \begin{picture}(100,100) \put(50,50){\circle*{7}} \put(50,100){\circle*{7}} \put(50,0){\circle*{7}} \put(0,50){\circle*{7}} \put(100,50){\circle*{7}} \put(85,85){\circle*{7}} \put(15,85){\circle*{7}} \put(85,15){\circle*{7}} \put(15,15){\circle*{7}} \put(0,50){\line(1,0){100}} \put(50,0){\line(0,1){100}} \put(15,15){\line(1,1){70}} \put(15,85){\line(1,-1){70}} \end{picture} \end{tabular} \end{center} \caption{Two trees with $v=9$, $b=8$ and $k=2$} \label{fig:tree} \end{figure} In a star with $v$~vertices, the contrast between any two leaves is an eigenvector of the Laplacian matrix~$\mathbf{L}$ with eigenvalue~$1$, while the contrast between the central vertex and all the other vertices is an eigenvector with eigenvalue~$v$. If $v\geq5$ and $G$ is not a star then there is an edge whose removal splits the graph into two components of sizes at least $2$ and $3$. Cutset Lemma~1 then shows that $\theta_1 \leq 5/6 <1$. The only other tree which is not a star is the path of length $3$, for which direct calculation shows that $\theta_1 = 2 - \sqrt{2} <1$. Hence the E-optimal designs are also the stars. \subsection{One fewer treatment} \label{sec:loop} If $b=v$ and $k=2$, then the concurrence graph $G$ contains a single cycle: such graphs are called \emph{unicyclic}. Let $s$ be the length of the cycle. All the remaining vertices are in trees attached to various vertices of the cycle. Fig.~\ref{fig:move} shows two unicyclic graphs with $v=12$ and $s=6$. \begin{figure}[htbp] \begin{center} \begin{tabular}{@{}c@{\qquad\qquad}c@{}} \setlength{\unitlength}{8mm} \begin{picture}(8,7.5)(-2,-2) \put(2,0){\circle*{0.4}} \put(-2,0){\circle*{0.4}} \put(1,1.67){\circle*{0.4}} \put(3,1.67){\circle*{0.4}} \put(2,3.33){\circle*{0.4}} \put(4,3.33){\circle*{0.4}} \put(4,5){\circle*{0.4}} \put(6,3.33){\circle*{0.4}} \put(-1,1.67){\circle*{0.4}} \put(-2,3.33){\circle*{0.4}} \put(1,-1.67){\circle*{0.4}} \put(-1,-1.67){\circle*{0.4}} \put(0.7,1.37){\makebox(0,0){{{$6$}}}} \put(-0.7,1.37){\makebox(0,0){{{$1$}}}} \put(-1.6,0){\makebox(0,0){{$2$}}} \put(-0.7,-1.37){\makebox(0,0){{$3$}}} \put(0.7,-1.37){\makebox(0,0){{$4$}}} \put(1.6,0){\makebox(0,0){{$5$}}} \put(-1.4,3.33){\makebox(0,0){{$12$}}} \put(3.4,1.67){\makebox(0,0){{{$7$}}}} \put(1.6,3.33){\makebox(0,0){{{$8$}}}} \put(4,2.83){\makebox(0,0){{$9$}}} \put(3.4,5){\makebox(0,0){{$10$}}} \put(6,2.83){\makebox(0,0){{$11$}}} \put(1,1.67){\line(-1,0){2}} \put(1,-1.67){\line(-1,0){2}} \put(3,1.67){\line(-1,0){2}} \put(4,3.33){\line(-1,0){2}} \put(6,3.33){\line(-1,0){2}} \put(4,3.33){\line(0,1){1.63}} \put(-1,1.67){\line(-3,5){1}} \put(-1,-1.67){\line(-3,5){1}} \put(2,0){\line(-3,5){1}} \put(1,1.67){\line(3,5){1}} \put(1,-1.67){\line(3,5){1}} \put(-2,0){\line(3,5){1}} \end{picture} & \setlength{\unitlength}{8mm} \begin{picture}(6,7.5)(-2,-2) \put(2,0){\circle*{0.4}} \put(-2,0){\circle*{0.4}} \put(1,1.67){\circle*{0.4}} \put(3,1.67){\circle*{0.4}} \put(2,3.33){\circle*{0.4}} \put(4,3.33){\circle*{0.4}} \put(2,5){\circle*{0.4}} \put(0,3.33){\circle*{0.4}} \put(-1,1.67){\circle*{0.4}} \put(-2,3.33){\circle*{0.4}} \put(1,-1.67){\circle*{0.4}} \put(-1,-1.67){\circle*{0.4}} \put(0.7,1.37){\makebox(0,0){{{$6$}}}} \put(-0.7,1.37){\makebox(0,0){{{$1$}}}} \put(-1.6,0){\makebox(0,0){{$2$}}} \put(-0.7,-1.37){\makebox(0,0){{$3$}}} \put(0.7,-1.37){\makebox(0,0){{$4$}}} \put(1.6,0){\makebox(0,0){{$5$}}} \put(-1.4,3.33){\makebox(0,0){{$12$}}} \put(3.4,1.67){\makebox(0,0){{{$7$}}}} \put(0.4,3.33){\makebox(0,0){{{$8$}}}} \put(2,2.83){\makebox(0,0){{$9$}}} \put(1.4,5){\makebox(0,0){{$10$}}} \put(4,2.83){\makebox(0,0){{$11$}}} \put(1,1.67){\line(-1,0){2}} \put(1,-1.67){\line(-1,0){2}} \put(3,1.67){\line(-1,0){2}} \put(4,3.33){\line(-1,0){2}} \put(2,3.33){\line(0,1){1.63}} \put(-1,1.67){\line(-3,5){1}} \put(-1,-1.67){\line(-3,5){1}} \put(2,0){\line(-3,5){1}} \put(1,1.67){\line(3,5){1}} \put(1,1.67){\line(-3,5){1}} \put(1,-1.67){\line(3,5){1}} \put(-2,0){\line(3,5){1}} \end{picture} \\ \\ (a) & (b) \end{tabular} \end{center} \caption{Two unicyclic graphs with $b=v=12$ and $s=6$} \label{fig:move} \end{figure} As we remarked in Section~\ref{sec:span}, the number of spanning trees in a unicyclic graph is equal to the length of the cycle. Hence, Theorem~\ref{thm:span} gives the following result. \begin{thm} If $k=2$ and $b=v\geq3$, then the D-optimal designs are precisely the cycles. \end{thm} For A-optimality, we first show that no graph like the one in Fig.~\ref{fig:move}(a) can be optimal. If vertex~$12$ is moved so that it is joined to vertex~$6$, instead of vertex~$1$, then the sum of the variances $V_{i,12}$ for $i$ in the cycle is unchanged and the variances $V_{i,12}$ for the remaining vertices $i$ are all decreased. This argument shows that all the trees must be attached to the same vertex of the cycle. Now consider the tree on vertices $6$, $8$, $9$, $10$ and $11$ in Fig.~\ref{fig:move}(a). If the two edges incident with vertex~$8$ are modified to those in Fig.~\ref{fig:move}(b), then the set of variances between these five vertices are unchanged, as are all others involving vertex~$8$, but those between vertices $9$, $10$, $11$ and vertices outside this tree are all decreased. This argument shows that, for any given length~$s$ of the cycle, the only candidate for an A-optimal design has $v-s$ leaves attached to a single vertex of the cycle. The effective resistance between a pair of vertices at distance $d$ in a cycle of length~$s$ is $d(s-d)/s$, while that between a leaf and the cycle vertex to which it is attached is $1$. Hence a short calculation shows that the sum of the pairwise effective resistances is equal to $g(s)/12$, where \[g(s) = -s^3 +2vs^2 + 13s -12sv +12v^2 -14v.\] Now $\bar V/\sigma^2 = g(s)/[3v(v-1)]$ and we seek the minimum of $g(s)$ for integers $s$ in the interval $[2,v]$. \begin{figure} \begin{center} \setlength{\unitlength}{1.9cm} \begin{picture}(7,3.6)(-0.6,1.8) \put(0,2){\line(1,0){5.5}} \put(0,2){\line(0,1){3.2}} \xxxlabel{1}{3} \xxxlabel{2}{6} \xxxlabel{3}{9} \xxxlabel{4}{12} \xxxlabel{5}{15} \yyylabel{2}{2} \yyylabel{3}{3} \yyylabel{4}{4} \yyylabel{5}{5} \put(5.8,2){\makebox(0,0){$s$}} \put(0,5.35){\makebox(0,0){$\bar V/\sigma^2$}} \put(0.67,3){\makebox(0,0){$\boldsymbol{\star}$}} \put(1,2.8){\makebox(0,0){$\boldsymbol{\star}$}} \put(1.33,2.67){\makebox(0,0){$\boldsymbol{\star}$}} \put(1.67,2.53){\makebox(0,0){$\boldsymbol{\star}$}} \put(2,2.33){\makebox(0,0){$\boldsymbol{\star}$}} \put(0.67,3.14){\makebox(0,0){$\boldsymbol{\odot}$}} \put(1,2.98){\makebox(0,0){$\boldsymbol{\odot}$}} \put(1.33,2.90){\makebox(0,0){$\boldsymbol{\odot}$}} \put(1.67,2.86){\makebox(0,0){$\boldsymbol{\odot}$}} \put(2,2.79){\makebox(0,0){$\boldsymbol{\odot}$}} \put(2.33,2.67){\makebox(0,0){$\boldsymbol{\odot}$}} \put(0.67,3.25){\makebox(0,0){$\boldsymbol{\ast}$}} \put(1,3.12){\makebox(0,0){$\boldsymbol{\ast}$}} \put(1.33,3.07){\makebox(0,0){$\boldsymbol{\ast}$}} \put(1.67,3.07){\makebox(0,0){$\boldsymbol{\ast}$}} \put(2,3.083){\makebox(0,0){$\boldsymbol{\ast}$}} \put(2.33,3.07){\makebox(0,0){$\boldsymbol{\ast}$}} \put(2.67,3){\makebox(0,0){$\boldsymbol{\ast}$}} \put(0.67,3.33){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(1,3.22){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(1.33,3.19){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(1.67,3.22){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(2,3.28){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(2.33,3.33){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(2.67,3.36){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(3,3.33){\makebox(0,0){$\boldsymbol{\diamond}$}} \put(0.67,3.4){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(1,3.3){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(1.33,3.29){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(1.67,3.33){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(2,3.41){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(2.33,3.51){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(2.67,3.6){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(3,3.66){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(3.33,3.67){\makebox(0,0){$\boldsymbol{\bullet}$}} \put(0.67,3.45){\makebox(0,0){$\boldsymbol{\circ}$}} \put(1,3.37){\makebox(0,0){$\boldsymbol{\circ}$}} \put(1.33,3.36){\makebox(0,0){$\boldsymbol{\circ}$}} \put(1.67,3.42){\makebox(0,0){$\boldsymbol{\circ}$}} \put(2,3.52){\makebox(0,0){$\boldsymbol{\circ}$}} \put(2.33,3.64){\makebox(0,0){$\boldsymbol{\circ}$}} \put(2.67,3.76){\makebox(0,0){$\boldsymbol{\circ}$}} \put(3,3.88){\makebox(0,0){$\boldsymbol{\circ}$}} \put(3.33,3.96){\makebox(0,0){$\boldsymbol{\circ}$}} \put(3.67,4){\makebox(0,0){$\boldsymbol{\circ}$}} \put(0.67,3.5){\makebox(0,0){$\boldsymbol{\times}$}} \put(1,3.424){\makebox(0,0){$\boldsymbol{\times}$}} \put(1.33,3.424){\makebox(0,0){$\boldsymbol{\times}$}} \put(1.67,3.48){\makebox(0,0){$\boldsymbol{\times}$}} \put(2,3.59){\makebox(0,0){$\boldsymbol{\times}$}} \put(2.33,3.73){\makebox(0,0){$\boldsymbol{\times}$}} \put(2.67,3.88){\makebox(0,0){$\boldsymbol{\times}$}} \put(3,4.03){\makebox(0,0){$\boldsymbol{\times}$}} \put(3.33,4.17){\makebox(0,0){$\boldsymbol{\times}$}} \put(3.67,4.27){\makebox(0,0){$\boldsymbol{\times}$}} \put(4,4.33){\makebox(0,0){$\boldsymbol{\times}$}} \put(0.67,3.54){\makebox(0,0){$\boldsymbol{+}$}} \put(1,3.419){\makebox(0,0){$\boldsymbol{+}$}} \put(1.33,3.47){\makebox(0,0){$\boldsymbol{+}$}} \put(1.67,3.54){\makebox(0,0){$\boldsymbol{+}$}} \put(2,3.65){\makebox(0,0){$\boldsymbol{+}$}} \put(2.33,3.79){\makebox(0,0){$\boldsymbol{+}$}} \put(2.67,3.96){\makebox(0,0){$\boldsymbol{+}$}} \put(3,4.14){\makebox(0,0){$\boldsymbol{+}$}} \put(3.33,4.31){\makebox(0,0){$\boldsymbol{+}$}} \put(3.67,4.46){\makebox(0,0){$\boldsymbol{+}$}} \put(4,4.59){\makebox(0,0){$\boldsymbol{+}$}} \put(4.33,4.67){\makebox(0,0){$\boldsymbol{+}$}} \put(4.83,3.73){\makebox{$\begin{array}{cc} \boldsymbol{+} & v=13\\ \boldsymbol{\times} & v=12\\ \boldsymbol{\circ} & v=11\\ \boldsymbol{\bullet} & v=10\\ \boldsymbol{\diamond} & v=9\\ \boldsymbol{\ast} & v=8\\ \boldsymbol{\odot} & v=7\\ \boldsymbol{\star} & v=6\\ \end{array}$}} \end{picture} \end{center} \caption{Average pairwise variance, in a unicylic graph with $v$~vertices, as a function of the length~$s$ of the cycle} \label{fig:cubic} \end{figure} Fig.~\ref{fig:cubic} plots $g(s)/[3v(v-1)]$ for $s$ in $[2,v]$ and $6\leq v\leq 13$. When $v\leq 7$, the function $g$ is monotonic decreasing, so it attains its minimum on $[2,v]$ at $s=v$. For all larger values of $v$, the function~$g$ has a local minimum in the interval $[3,5]$: when $v\geq 9$, the value at this local minimum is less than $g(v)$. This change from the upper end of the interval to the local minimum explains the sudden change in the A-optimal designs. Detailed examination of the local minimum gives the following result. \begin{thm} \label{thm:aopt} If $k=2$ and $b=v\geq3$ then the A-optimal designs are: \begin{itemize} \item a cycle, if $v\leq 8$; \item a square with $v-4$ leaves attached to one vertex, if $9\leq v \leq 11$; \item a triangle with $v-4$ leaves attached to one vertex, if $ v \geq 13$; \item either of the last two, if $v=12$. \end{itemize} \end{thm} What about E-optimality? The smallest eigenvalue of the Laplacian matrix of the triangle with one or more leaves attached to one vertex is 1, as is that of the digon with two or more leaves attached to one vertex. We now show that almost all other unicyclic graphs have at least one non-trivial eigenvalue smaller than this. Suppose that vertex~$i$ in the cycle has a non-empty tree attached to it, so that $\{i\}$ is a vertex cutset. If $s\geq 3$ then there are no double edges, so Cutset Lemma~2 shows that $\theta_1 <1$ unless all vertices are joined to~$i$, in which case $s=3$. If $s=2$ and there are trees attached to both vertices of the digon, then applying Cutset Lemma~2 at each of these vertices shows that $\theta_1<1$ unless $v=4$ and there is one leaf at each vertex of the digon: for this graph, $\theta_1 = 2-\sqrt{5} <1$. A digon with leaves attached to one vertex is just a star with one edge doubled. The cycle of size~$v$ is a cyclic design. The smallest eigenvalue of its Laplacian matrix is $2(1 - \cos(2\pi/v))$, which is greater than $1$ when $v\leq 5$, is equal to $1$ when $v=6$, and is less than~$1$ when $v\leq 7$. When $v=3$ it is equal to $3$, which is greater than $3-\sqrt{3}$, which is the smallest Laplacian eigenvalue of the digon with one leaf. Putting all of this together proves the following result. \begin{thm} \label{thm:eopt} If $k=2$ and $b=v\geq3$, then the E-optimal designs are: \begin{itemize} \item a cycle, if $v\leq 5$; \item a triangle with $v-3$ leaves attached to one vertex, or a star with one edge doubled, if $v\geq 7$; \item either of the last two, if $v=6$. \end{itemize} \end{thm} Thus, for $v\geq 9$, the ranking on the D-criterion is essentially the opposite of the ranking on the A- and E-criteria. The A- and E-optimal designs are far from equireplicate. The change is sudden, not gradual. These findings were initially quite shocking to statisticians. \subsection{More blocks} What happens when $b$ is larger than $v$ but still has the same order of magnitude? The following theorems show that the A- and E-optimal designs are very different from the D-optimal designs when $v$ is large. The proofs of Theorems~\ref{thm:thresh} and \ref{thm:eleaf} are in \cite{rabpaper} and \cite{bcc09} respectively. \begin{thm} Let $G$ be the concurrence graph of a connected block design $\Delta$ with $k=2$ and $b\geq v$. If $\Delta$ is D-optimal then $G$ does not contain any bridge (an edge cutset of size one): in particular, $G$ contains no leaves. \end{thm} \begin{pf} Suppose that $\{i,j\}$ is an edge-cutset for $G$. Let $H$ and $K$ be the parts of $G$ containing $i$ and $j$, respectively. Since $G$ is not a tree, we may assume that $H$ is not a tree, and so there is some edge $e$ in $H$ that is not in every spanning tree for $H$. Let $n_1$ and $n_2$ be the numbers of spanning trees for $H$ that include and exclude $e$, respectively, and let $m$ be the number of spanning trees for $K$. Every spanning tree for $G$ consists of spanning trees for $H$ and $K$ together with the edge $\{i,j\}$. Hence $G$ has $(n_1+n_2)m$ spanning trees. Let $\ell$ be a vertex on $e$ with $\ell \ne i$. Form $G'$ from $G$ by removing edge $e$ and inserting the edge $e'$, where $e' = \{\ell,j\}$. Let $T$ and $T'$ be spanning trees for $H$ and $K$ respectively. If $T$ does not contain $e$ then $T \cup \{\{i,j\}\} \cup T'$ and $T \cup \{e'\} \cup T'$ are both spanning trees for $G'$. If $T$ contains $e$ then $(T\setminus \{e\}) \cup \{\{i,j\}\} \cup \{e'\} \cup T'$ is a spanning tree for $G'$. Hence the number of spanning trees for $G'$ is at least $(2n_2 +n_1)m$, which is greater than $(n_1+n_2)m$ because $n_2\geq 1$. Hence $G$ does not have the maximal number of spanning trees and so $\Delta$ is not D-optimal. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} \begin{thm} \label{thm:thresh} Let $c$ be a positive integer. Then there is a positive integer $v_c$ such that if $b-v=c$ and $v\geq v_c$ then all A-optimal designs with $k=2$ contain leaves. \end{thm} \begin{thm} \label{thm:eleaf} If $20 \leq v \leq b\leq 5v/4$ then the concurrence graph for any E-optimal design with $k=2$ contains leaves. \end{thm} Of course, to obtain a BIBD when $k=2$, $b$ needs to be a quadratic function of $v$. What happens if $b$ is merely a linear function of $v$? In \cite{bcc09} we conjectured that if $b=cv$ for some constant $c$ then there is a threshold result like the one in Theorem~\ref{thm:thresh}. However, current work by Robert Johnson and Mark Walters \cite{JRJMW} suggests something much more interesting---that there is a constant $C$ with $3<C<4$ such that if $b\geq Cv$ and $k=2$ then all A-optimal designs are (nearly) equireplicate, and that random such graphs (in a suitable model) are close to A-optimal with high probability. On the other hand if $b\leq Cv$ then a graph consisting of a large almost equireplicate part (all degrees 3 and 4 with average degree close to $Cv$) together with a suitable number of leaves joined to a single vertex is strictly better than any queen-bee design. \subsection{A little more history} The results on D- and A-optimality in Sections~\ref{sec:least} and~\ref{sec:loop} were proved in \cite{rabpaper}, partly to put to rest mutterings that the results of \cite{BJJAE,KCh,Wit} found by computer search were incorrect. The results on E-optimality are in \cite{bcc09}. In spite of the horror with which these results were greeted, it transpired that they were not new. The D- and E-optimal designs for $b= (v-1)/(k-1)$ were identified in \cite{bap} in 1991. The A-optimal designs for $k=2$ and $b=v-1$ had been given in \cite{mandal} in 1991. Also in 1991, Tjur gave the A-optimal designs for $k=2$ and $b=v$ in \cite{tj1}: his proof used the Levi graph as an electrical network. A fairly common response to these unexpected results was `It seems to be just block size~$2$ that is a problem.' Perhaps those of us who usually deal with larger blocks had simply not thought that it was worth while to investigate block size $2$ before the introduction of microarrays. However, as we sketch in the next section, the problem is not block size~$2$ but very low average replication. The proofs there are similar to those in this section; they are given in more detail in \cite{rabalia, alia}. Once again, it turns out that these results are not all new. The D-optimal designs for $v/(k-1)$ blocks of size~$k$ were given by Balasubramanian and Dey in \cite{baldey} in 1996---but their proof uses a version of Theorem~\ref{thm:gaff} with the wrong value of the constant. The A-optimal designs for $v/(k-1)$ blocks of size~$k$ were published by Krafft and Schaefer in \cite{krafft} in 1997---but those authors are not blameless either, because they apparently had not read \cite{tj1}! Our best explanation is that agricultural statisticians are so familiar with average replication being at least $3$ that when we saw these papers we decided that they had no applicability and so forgot them. \section{Very low average replication} In this section we once again consider general block size~$k$. A block design is connected if and only if its Levi graph is connected. The Levi graph has $v+b$ vertices and $bk$ edges, so connectivity implies that $bk \geq b+v-1$; that is, $b(k-1) \geq v-1$. \subsection{Least replication} If $b(k-1) = v-1$ and the design is connected, then the Levi graph $\tilde G$ is a tree and the concurrence graph $G$ looks like those in Fig.~\ref{fig:cgqueen}. Hypergraph-theorists do not seem to have an agreed name for such designs. For both D- and A-optimality, it turns out to be convenient to use the Levi graph. Since all the Levi graphs are trees, Theorem~\ref{thm:gaff} shows that the D-criterion does not distinguish among connected designs. By Theorem~~\ref{thm:tjur}, $V_{ij} = \tilde R_{ij} \sigma^2$. When $\tilde G$ is a tree, $\tilde R_{ij}=2$ when $i$ and $j$ are in the same block; otherwise, $\tilde R_{ij} = 4$ if any block containing~$i$ has a treatment in common with any block containing~$j$; and otherwise, $\tilde R_{ij} \geq 6$. The queen-bee designs are the only ones for which $\tilde R_{ij}\leq 4$ for all $i$ and $j$, and so they are the A-optimal designs. The non-trivial eigenvalues of a queen-bee design are $1$, $k$ and $v$, with multiplicities $b-1$, $b(k-2)$ and $1$, respectively. If the design is not a queen-bee design, then there is a treatment~$i$ that is in more than one block but not in all blocks. Thus vertex~$i$ forms a cutset for the concurrence graph $G$ which is not joined to every other vertex of $G$. Cutset Lemma~2 shows that $\theta_1<1$. Hence the E-optimal designs are also the queen-bee designs. \subsection{One fewer treatment} If $b(k-1)=v$, then the Levi graph $\tilde G$ has $bk$ edges and $bk$ vertices, and so it contains a single cycle, which must be of some even length~$2s$. If $2\leq s \leq b$, then the design is binary; if $s=1$, then there is a single non-binary block, whose defect is~$1$. In this case, $k\geq 3$, because each block must have more than one treatment. For $2\leq s \leq b$, let $\mathcal{C}(b,k,s)$ be the class of designs constructed as follows. Start with a loop design for $s$ treatments. Insert $k-2$ extra treatments into each block. The remaining $b-s$ blocks all contain the same treatment from the loop design, together with $k-1$ extra treatments. Figs.~\ref{fig:star} and~\ref{fig:levistar} show the concurrence graph and Levi graph, respectively, of a design in $\mathcal{C}(6,3,6)$. For $k\geq 4$, the designs in $\mathcal{C}(b,k,1)$ have one treatment which occurs twice in one block and once in all other blocks, with the remaining treatments all replicated once. The class $\mathcal{C}(b,3,1)$ contains all such designs, and also those in which the treatment in every block is the one which occurs only once in the non-binary block. \begin{thm} If $b(k-1)=v$, then the D-optimal designs are those in $\mathcal{C}(b,k,b)$. \end{thm} \begin{pf} The Levi graph $\tilde G$ is unicyclic, so its number of spanning trees is maximized when the cycle has maximal length. Theorem~\ref{thm:gaff} shows that the D-optimal designs are precisely those with $s=b$. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} \begin{thm} \label{thm:abig} If $b(k-1)=v$ then the A-optimal designs are those in $\mathcal{C}(b,k,s)$, where the value of $s$ is given in Table~\ref{tab:sval}. \end{thm} \begin{table}[htbp] \[ \begin{array}{cc|ccccccccccccc} k & b & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10& 11 & 12 & 13\\ \hline 2 & & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 4 & 4 & 4 & 3 \mbox{ or } 4& 3\\ 3 & & 2 & 3 & 4 & 5 & 6 & 3 & 3 & 3 & 3 & 3 & 2 & 2\\ 4 & & 2 & 3 & 4 & 5 & 3 & 2 & 2 & 2 & 2 & 2 & 2 & 2 \\ 5 & & 2 & 3 & 4 & 5 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2\\ 6 & & 2 & 3 & 4 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2 \end{array} \] \caption{Value of $s$ for A-optimal designs for $b(k-1)$ treatments in $b$ blocks of size $k$: see Theorem~\ref{thm:abig}} \label{tab:sval} \end{table} \begin{pf} The Levi graph $\tilde G$ has one cycle, whose length is $2s$, where $1\leq s\leq b$. A similar argument to the one used at the start of the proof of Theorem~\ref{thm:aopt} shows that this cannot be A-optimal unless the design is in $\mathcal{C}(b,k,s)$. If $s\geq 2$ or $k\geq 4$, then each block-vertex in the cycle has $k-2$ treatment-vertices attached as leaves; all other block-vertices are joined to the same single treatment-vertex in the cycle, and each has $k-1$ treatment vertices attached as leaves. In $\mathcal{C}(b,3,1)$ the first type of design has a Levi graph like this, and the other type has the same multiset of effective resistances between treatment-vertices, because their concurrence graphs are identical. The following calculations use the first type. Let $\mathcal{V}_1$ be the set of treatment-vertices in the cycle, $\mathcal{V}_2$ the set of other treatment-vertices joined to blocks in the cycle, and $\mathcal{V}_3$ the set of remaining treatment-vertices. For $1\leq i\leq j\leq 3$, denote by $\mathcal{R}_{ij}$ the sum of the pairwise resistances between vertices in $\mathcal{V}_i$ and $\mathcal{V}_j$. Put \[ R_1 = \sum_{d=1}^{s-1}\frac{2d(2s-2d)}{2s} = \frac{s^2-1}{3} \] and \[ R_2 = \sum_{d=0}^{s-1}\frac{(2d+1)(2s-2d-1)}{2s} = \frac{2s^2+1}{6}. \] Then $\mathcal{R}_{11} = sR_1/2$, $\mathcal{R}_{12} = s(k-2)(R_2+s)$, $\mathcal{R}_{13} = (b-s)(k-1)(R_1+2s)$, $\mathcal{R}_{22} = s(k-2)(k-3) + s(k-2)^2[R_1 + 2(s-1)]/2$, $\mathcal{R}_{23} = (b-s)(k-1)(k-2)(R_2+3s)$, and $\mathcal{R}_{33} = (b-s)(k-1)(k-2) + 2(b-s)(b-s-1)(k-1)^2$. Hence the sum of the pairwise effective resistances between treatment-vertices in the Levi graph is $g(s)/6$, where \[ g(s) = -(k-1)^2s^3 +2b(k-1)^2 s^2 -[6bk(k-1) -4k^2 +2k -1]s +c \] and $c=b(k-1)[12b(k-1) -5k-4]$. If $s=1$ then the design is non-binary. However, \[g(1)-g(2) = (3k-9+6b)(k-1) -3,\] which is positive, because $k\geq 2$ and $b\geq 2$. Therefore the non-binary designs are never A-optimal. Direct calculation shows that $g(2) > g(3)$ when $b=3$, and that $g(2)> g(3)> g(4)$ when $b=4$. These inequalities hold for all values of $k$, even though $g$~is not decreasing on the interval $[2,4]$ for large $k$ when $b=4$. If $b=5$ and $k\geq 6$, then $g(3)>g(2)$ and $g(5)>g(2)$. Thus the local minimum of $g$ occurs in the interval $(1,3)$ and is the overall minimum of $g$ on the interval $[1,5]$. Differentiation gives \[ g'(b) = b(k-1)[(b-6)(k-1) - 6] +4k^2 - 2k +1. \] If $g'(b)>0$ then $g$ has a local minimum in the interval $(1,b)$. If, in addition, $g(3)>g(2)$, then the minimal value for integer $s$ occurs at $s=2$. These conditions are both satisfied if $k=3$ and $b\geq 12$, $k= 4$ and $b\geq 8$, $k\geq 5$ and $b\geq 7$, or $k\geq 9$ and $b\geq 6$. Given Theorem~\ref{thm:aopt}, there remain only a finite number of pairs $(b,k)$ to be checked individually to find the smallest value of $g(s)$. The results are in Table~\ref{tab:sval}. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} \begin{thm} If $b(k-1)=v$, $b\geq 3$ and $k\geq 3$, then the E-optimal designs are those in $\mathcal{C}(b,k,b)$ if $b\leq4$, and those in $\mathcal{C}(b,k,2)$ and $\mathcal{C}(b,k,1)$ if $b\geq 5$. \end{thm} \begin{pf} If $2<s<b$ then the concurrence graph $G$ has a vertex which forms a vertex-cutset and which is not joined to all other vertices; moreover, $G$ has no multiple edges. Thus Cutset Lemma~2 shows that $\theta_1<1$. Direct calculation shows that $\theta_1=1$ if $s=1$ or $s=2$. For $k\geq 4$, all contrasts between singly replicated treatments in the same block are eigenvectors of the Laplacian matrix~$L$ with eigenvalue~$k$. When $k\geq 3$ and $s=b$ the contrast between singly and doubly replicated treatments has eigenvalue $2(k-1)$. For $s=b$, a straightforward calculation shows that the remaining eigenvalues of $L$ are \[ k - \cos \left ( \frac{2\pi n}{b} \right) \pm \sqrt{(k-1)^2 - \sin^2 \left( \frac{2\pi n}{b}\right)} \] for $1\leq n\leq b-1$. The smallest of these is \[k - \cos(2\pi/b) - \sqrt{(k-1^2) - \sin^2(2\pi/b)}:\] this is greater than $1$ if $b=3$ or $b=4$, but less than $1$ if $k\geq 3$ and $b\geq 5$. \protect\nolinebreak\mbox{\quad\rule{1ex}{1ex}} \end{pf} \section{Further reading} The Laplacian matrix of a graph, and its eigenvalues, are widely used, especially in connection with network properties such as connectivity, expansion, and random walks. A good introduction to this material can be found in the textbook by Bollob\'as~\cite{bollobas}, especially Chapters II (electrical networks) and IX (random walks). Connection between the smallest non-zero eigenvalue and connectivity is described in surveys by de Abreu~\cite{survey} and by Mohar~\cite{mohar}. In this terminology, a version of Theorem~\ref{thm:eopt} is in \cite{algconn}. The basic properties of electrical networks can be found in textbooks of electrical engineering, for example Balabanian and Bickart~\cite{wag}. A treatment connected to the multivariate Tutte polynomial appears in Sokal's survey~\cite{sokal}. Bollob\'as describes several approaches to the theory, including the fact (which we have not used) that the current flow minimises the power consumed in the network, and explains the interactions between electrical networks and random walks in the network. See also Deo~\cite{deo}. The connection with optimal design theory was discussed in detail by the authors in their survey~\cite{bcc09}. Further reading on optimal design can be found in John and Williams~\cite{JAJERW:book}, Schwabe~\cite{Schwab}, or Shah and Sinha~\cite{ss}. For general principles of experimental design, see Bailey~\cite{rabbook}. \paragraph{Acknowledgement} This chapter was written at the Isaac Newton Institute for Mathematical Sciences, Cambridge, UK, during the 2011 programme on Design and Analysis of Experiments.
\section{\@startsection {section}{1}{\z@}% {-\bigskipamount}% {\medskipamount}% {\normalsize\bfseries\mathversion{bold}\raggedright\pretolerance=\@M \tolerance=\@M \hyphenpenalty=\@M \exhyphenpenalty=\@M}} \renewcommand\subsection{\@startsection {subsection}{2}{\z@}% {-\medskipamount}% {\medskipamount}% {\normalsize\pretolerance=\@M \tolerance=\@M \hyphenpenalty=\@M \exhyphenpenalty=\@M}} \renewcommand\subsubsection{\@startsection{subsubsection}{3}{\z@}% {-\medskipamount}% {\medskipamount}% {\itshape\raggedright}} \renewcommand\paragraph{\@startsection{paragraph}{4}{\z@}% {\medskipamount}% {-10pt}% {\bfseries}} \renewcommand\subparagraph{\@startsection{subparagraph}{5}{\parindent}% {0.1pt}% {-1em}% {\itshape}} \def\section@numbersep{.} \def\subsection@numbersep{.} \def\subsubsection@numbersep{.} \def\paragraph@numbersep{.} \def\subparagraph@numbersep{.} \def\@seccntformat#1{{\csname #1@prefix\endcsname\csname the#1\endcsname\csname#1@numbersep\endcsname\enspace}} \overfullrule=5pt \let\abstractnamefont=\bfseries \def\normalsize\normalfont{\normalsize\normalfont} \let\abstractdot\relax \def\raggedrightmargin#1{% \let\\\@centercr\@rightskip #1 plus 1fil \rightskip\@rightskip \leftskip\z@skip \parindent\z@\pretolerance=\@M \tolerance=\@M \hyphenpenalty=\@M \exhyphenpenalty=\@M} \def\raggedrightmargin{10mm}{\raggedrightmargin{10mm}} \def\vglue-\topskip\hrule \@height 0\p@ \vskip40\p@{\vglue-\topskip\hrule \@height 0\p@ \vskip40\p@} \def\fontsize{13}{15}\selectfont\bfseries\mathversion{bold}{\fontsize{13}{15}\selectfont\bfseries\mathversion{bold}} \def\normalsize\bfseries\raggedrightmargin{42mm}\let\sep=~{\mathversion{bold}\ensuremath{\cdot}}\ {\normalsize\bfseries\raggedrightmargin{42mm}\let\sep=~{\mathversion{bold}\ensuremath{\cdot}}\ } \def~{\mathversion{bold}\ensuremath{\cdot}}\ {~{\mathversion{bold}\ensuremath{\cdot}}\ } \def\email@prefix{e-mail:~} \def\@formatauthors{\begingroup \normalsize\bfseries\raggedrightmargin{42mm}\let\sep=~{\mathversion{bold}\ensuremath{\cdot}}\ \leavevmode \gdef\surname##1{##1}% \gdef\email##1{ \hbox{(\email@prefix{\sffamily \lowercase{##1}})}}% \cnt@authors=0 \def\@k@p##1{\advance\cnt@authors by 1}\@curauths \def\@k@p##1{\advance\cnt@authors by -1 \ifnum\cnt@authors>1 \@formatname{##1}{,}\penalty0\ \fi \ifnum\cnt@authors=1 \@formatname{##1}{} \authorand \penalty0\ \fi \ifnum\cnt@authors<1 \@formatname{##1}{}\par\fi}\@curauths \vskip \afterauthorskip \endgroup} \def\footnotesize\normalfont{\footnotesize\normalfont} \def\@formatinstitute{\insert\footins{\footnotesize\normalfont \institutecase{\@curinst}\par}} \renewcommand{\fnum@figure}{{\bfseries\figurename~\thefigure.\/}} \renewcommand\thetable{\@arabic\c@table} \renewcommand{\fnum@table}{{\bfseries\tablename~\thetable.\/}} \def\acknowledgements{\paragraph*{\acknowledgementsname}% \message{\acknowledgementsname}} \newenvironment{ack}[1][Acknowledgement]{\footnotesize\paragraph*{#1}}{} \newenvironment{acks}[1][Acknowledgements]{\footnotesize\paragraph*{#1}}{} \def\@jname{\ } \def\@doi{\@doihead \@thedoi} \def\@doihead{ } \def\@thedoi{ } \def\ps@opening{% \def\@oddhead{\parbox[t]{\textwidth}{\footnotesize\@jname\\\@doi}}% \let\@evenhead\@oddhead \def\@oddfoot{\idline\hfill \def\@evenfoot{\hfill\@gobble\idline}} \def\ps@headings{% \def\@oddfoot{\idline\hfil }% \let\@evenfoot\@oddfoot \def\@evenhead{\rh@rule\hbox{}\@gobble{\rlap{\footnotesize\thepage}}\hfil \@markfont\mymyleftmark}% \def\@oddhead{\rh@rule\@markfont\mymyrightmark\hfill\@gobble{\llap{\footnotesize\thepage}}}% } \def\rh@rule{\leavevmode\lower6pt\hbox to0pt{\vrule height1pt width\textwidth\hss}} \pagestyle{headings} \def\@coprtyear{ } \def\@volume{0} \def\@@firstpage{\ } \def\@@lastpage{\ } \if@optionalrh \def\mymyleftmark{\@runningauthor}% \def\mymyrightmark{\@runningtitle}% \else \def\mymyleftmark{\@jname\@gobble{\ (\@coprtyear) \@volume:\@firstpage--\@lastpage}} \let\mymyrightmark\mymyleftmark \fi \def\gdef\@copyright{\copyright@size\copyright@text\vskip2\baselineskip}{\gdef\@copyright{\copyright@size\copyright@text\vskip2\baselineskip}} \def\copyright@text{\ } \let\copyright@size\footnotesize \def\@maketitle{% \@arttype \@title \@subtitle \@authorsandinstitutes \@date \@copyright \@abstract \@keywords \@abbreviations \@classification \@nomenclature \@translation \@dedication \@motto} \gdef\@copyright{\copyright@size\copyright@text\vskip2\baselineskip} \def\kapenumargs{% \topsep \smallskipamount \partopsep \z@ \@plus 1pt \itemsep \z@ \@plus \z@ \parsep \z@ \@plus 1pt \if@margspec \else \leftmargini \z@ \fi \if@margspec \else \leftmarginii 1em \fi \if@margspec \else \leftmarginiii 1em \fi \if@margspec \else \leftmarginiv 1em \fi \if@margspec \leftmargin\csname leftmargin\romannumeral\@enumdepth\endcsname \labelwidth\leftmargin \advance\labelwidth-\labelsep \fi \rightmargin \z@ \listparindent \z@ \itemindent \z@ } \def\enumerate{\@ifnextchar[% {\kap@enumerate}% {\if@margspec \kap@enumerate[]\else \kap@enumerate[0]\fi }} \if@solaenum \def\@roman\c@enumi {\@roman\c@enumi} \def\textit{\theenumi}) {\textit{\@roman\c@enumi})} \fi \def\kapitemargs{% \itemsep \z@% \parsep \z@% \leftmargini\z@% \itemindent\z@% } \def\textbullet {\textbullet} \def\textendash {\textendash} \def\labelitemiii{\textasteriskcentered} \def{\footnotesize +} {{\footnotesize +}} \AtBeginDocument{\@ifundefined{urlstyle}{}{\urlstyle{sf}}} \DeclareMathAlphabet {\mathbfit}{OML}{cmm}{b}{it} \def\TODAY@JWL{\number\day\space\ifcase\month\or January\or February\or March\or April\or May\or June\or July\or August\or September\or October\or November\or December\fi \space\number\year} \def\idline{\if@noid\else \rlap{\smash{\vtop to \id@boxheight{% \vfil\hbox to\textwidth{% \hfil\footnotesize\tt ~\TODAY@JWL;~p.~\thepage}}}}% \fi} \if@loadnatbib \let\bibhang\relax \let\citeauthoryear\relax \RequirePackage{natbib} \@ifundefined{newblock}{\def\hskip .11em plus .33em minus .07em}}{{\hskip .11em plus .33em minus .07em}}{} \@ifundefined{@listctr}{\newcounter{start}\setcounter{start}{0}\def\@listctr{start}}{} \let\inlinecite\citet \let\opencite\citealp \let\shortcite\citeyearpar \let\bibfont\footnotesize \gdef\NAT@bibsetup#1{\kapbib@list} \fi \def\citefix{% \def\NAT@citex% [##1][##2]##3{% \NAT@sort@cites{##3}% \let\@citea\@empty \@cite{\let\NAT@nm\@empty\let\NAT@year\@empty \@for\@citeb:=\NAT@cite@list\do {\edef\@citeb{\expandafter\@firstofone\@citeb}% \if@filesw\immediate\write\@auxout{\string\citation{\@citeb}}\fi \@ifundefined{b@\@citeb\@extra@b@citeb}{\@citea% {\reset@font\bfseries ?}\NAT@citeundefined \PackageWarning{natbib}% {Citation `\@citeb' on page \thepage \space undefined}\def\NAT@date{}}% {\let\NAT@last@nm=\NAT@nm\let\NAT@last@yr=\NAT@year \NAT@parse{\@citeb}% \ifNAT@longnames\@ifundefined{bv@\@citeb\@extra@b@citeb}{% \let\NAT@name=\NAT@all@names \global\@namedef{bv@\@citeb\@extra@b@citeb}{}}{}% \fi \ifNAT@full\let\NAT@nm\NAT@all@names\else \let\NAT@nm\NAT@name\fi \ifNAT@swa\ifcase\NAT@ctype \if\relax\NAT@date\relax \@citea\NAT@nmfmt{\NAT@nm}% \hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@date\hyper@natlinkend \else \ifx\NAT@last@nm\NAT@nm\NAT@yrsep \ifx\NAT@last@yr\NAT@year \hyper@natlinkstart{\@citeb\@extra@b@citeb}\NAT@exlab \hyper@natlinkend \else\unskip\ \hyper@natlinkstart{\@citeb\@extra@b@citeb}\NAT@date \hyper@natlinkend \fi \else \@citea\NAT@nmfmt{\NAT@nm}% \NAT@aysep\ \hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@date\hyper@natlinkend \fi \fi \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@nmfmt{\NAT@nm}\hyper@natlinkend \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@date\hyper@natlinkend \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@alias\hyper@natlinkend \fi \def\@citea{\NAT@sep\ }% \else\ifcase\NAT@ctype \if\relax\NAT@date\relax \@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@nmfmt{\NAT@nm}\hyper@natlinkend \else \ifx\NAT@last@nm\NAT@nm\NAT@yrsep \ifx\NAT@last@yr\NAT@year \hyper@natlinkstart{\@citeb\@extra@b@citeb}\NAT@exlab \hyper@natlinkend \else\unskip\ \hyper@natlinkstart{\@citeb\@extra@b@citeb}\NAT@date \hyper@natlinkend \fi \else \@citea\NAT@nmfmt{\NAT@nm}% \ \NAT@@open\if*##1*\else##1\ \fi \hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@date\hyper@natlinkend\fi \fi \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@nmfmt{\NAT@nm}\hyper@natlinkend \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@date\hyper@natlinkend \or\@citea\hyper@natlinkstart{\@citeb\@extra@b@citeb}% \NAT@alias\hyper@natlinkend \fi \if\relax\NAT@date\relax\def\@citea{\NAT@sep\ }% \else\def\@citea{\NAT@@close\NAT@sep\ }\fi \fi }}\ifNAT@swa\else\if*##2*\else\NAT@cmt##2\fi \if\relax\NAT@date\relax\else\NAT@@close\fi\fi}{##1}{##2}} \let\@citex\NAT@citex } \if@linksfromyear \if@loadnatbib \citefix \else \def\cite@@label{unknown}% \let\save@document\document \def\document{ \@ifpackageloaded{hyperref}{% \gdef\cite@year@hook##1{\hyper@@link[cite]{}{cite.\cite@@label\@extra@b@citeb}{##1}}% \gdef\cite@author@hook##1{\hyper@@link[cite]{}{cite.\cite@@label\@extra@b@citeb}{##1}}% \gdef\bibcite##1##2{% \@newl@bel{b}{##1\@extra@binfo}{% \def\cite@@label{##1}% ##2% }% }% }% {}% \save@document } \fi \fi \def\label#1{\ilabel{\thearticle #1}}% \def\ref#1{\iref{\thearticle #1}}% \def\pageref#1{\ipageref{\thearticle #1}}% \renewenvironment{article}{% \def\@definecounter##1{\art@intdefinecounter{##1}% \@addtoreset{##1}{article}}% \renewcommand{\thearticle}{\roman{article}}% \refstepcounter{article}% \message{Article \number\c@article}% \gdef\@firstpage{\the\c@page}% \@addtoreset{equation}{article}% \ifx\sectioncmd\section \@addtoreset{section}{article}% \else \@addtoreset{chapter}{article}% \fi \@addtoreset{endnote}{article}% \@addtoreset{table}{article}% \@addtoreset{figure}{article}% \@addtoreset{algorithm}{article}% \setlastpage \global\inarticletrue }{\make@ao \writelastpage \clearpage \if@openright \ifodd \c@page \else ~\thispagestyle{empty}\newpage \fi \fi \gdef\@dedication{}\gdef\@translation{}% \gdef\@title{}\gdef\@subtitle{}% \gdef\@arttype{}\gdef\@keywords{}\gdef\@classification{}% \gdef\@nomenclature{}\gdef\@abbreviations{}\gdef\@abstract{}% \gdef\@kapidenthead{}\gdef\@kapidentfoot{}% \gdef\@barcode{}\gdef\@firstpage{\thepage}% \gdef\@crline{}% \global\lastpagegivenfalse \global\inarticlefalse } \endinput
\section{Introduction} Let $K$ be a 2-bridge link in $S^3$ and let $S$ be a 4-punctured sphere in $S^3-K$ obtained from a $2$-bridge sphere of $K$. The present paper is a continuation of \cite{lee_sakuma_2} and \cite{lee_sakuma_3} in a series of papers which give a necessary and sufficient condition for two essential simple loops on $S$ to be homotopic in $S^3-K$. Ahead of this series, the authors~\cite{lee_sakuma} gave a complete characterization of those essential simple loops in $S$ which are null-homotopic in $S^3-K$. The first paper~\cite{lee_sakuma_2} of the series treated the case of a $2$-bridge link of slope $1/p=[p]$, and the second paper~\cite{lee_sakuma_3} treated the case of a 2-bridge link of slope $n/(2n+1)=[2,n]$ or slope $(n+1)/(3n+2)=[2,1,n]$. On the other hand, the first half of the present paper treats the case of a 2-bridge link of slope $n/(mn+1)=[m,n]$ or slope $(n+1)/((m+1)n+m)=[m,1,n]$, where $m \ge 3$ is an arbitrary integer. These five families play special roles in our project in the sense that the treatment of these links form a base step of an inductive proof of the main theorem for a $2$-bridge link of general slope $[m_1,m_2, \dots,m_k]$ to which the second half of the present paper contributes, where the induction uses $k \ge 1$ as the parameter. In the present paper, we also give a complete characterization of those simple loops in the $2$-bridge sphere of a hyperbolic $2$-bridge link to be peripheral or imprimitive in the link complement (see Theorems~\ref{main_corollary} and \ref{main_corollary2}). This paper is organized as follows. In Section~\ref{sec:main_result}, we describe the main results of this paper (Main Theorem~\ref{main_theorem} and Theorems~\ref{main_corollary} and~\ref{main_corollary2}). In Section~\ref{sec:technical_lemmas}, we establish technical lemmas used for the proofs in Sections~\ref{sec:proof_of_main_theorem_1} and \ref{sec:proof_of_main_theorem_3}. Two special cases of Main Theorem~\ref{main_theorem} namely, the cases of a 2-bridge link of slope $n/(mn+1)=[m,n]$ and a $2$-bridge link of slope $(n+1)/((m+1)n+m)=[m,1,n]$, where $m,n \ge 3$, are treated in Sections~\ref{sec:proof_of_main_theorem_1} and \ref{sec:proof_of_main_theorem_3}, respectively. For the case of a $2$-bridge link of general slope, we start with preliminary results in Section~\ref{sec:technical_lemmas_general} and perform transformation so that we may apply the induction as discussed in Section~\ref{sec:transformation}. In Section~\ref{sec:result_for_induction}, we prove key results for the induction, and finally the proof of Main Theorem~\ref{main_theorem} for the general cases is contained in Section~\ref{sec:proof_for_general_2-bridge_links}. In Section~\ref{sec:proof_of_main_corollary}, we prove Theorems~\ref{main_corollary} and \ref{main_corollary2}. \section{Main result} \label{sec:main_result} This paper, as a continuation of \cite{lee_sakuma_2} and \cite{lee_sakuma_3}, uses the same notation and terminology as in \cite{lee_sakuma_2} and \cite{lee_sakuma_3} without specifically mentioning. We begin with the following question, providing whose answer is the purpose of this series of papers. \begin{question} \label{question} Consider a $2$-bridge link $K(r)$ with $r\ne \infty$. For two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$, when are the unoriented loops $\alpha_s$ and $\alpha_{s'}$ homotopic in $S^3-K(r)$? \end{question} By Schubert's classification of $2$-bridge links~\cite{Schubert}, we may assume that $r$ is a rational number with $0 \le r \le 1/2$ or $r=\infty$. If $r=0$ or $\infty$, then $G(K(r))$ is the infinite cyclic group or the rank $2$ free group accordingly, and we can easily obtain an answer to Question~\ref{question} (see~\cite[Paragraph after Question~2.2]{lee_sakuma_2}). So we may assume $0< r \le 1/2$. In the first paper~\cite{lee_sakuma_2} and the second paper~\cite{lee_sakuma_3} of this series, we gave a complete answer to the question, respectively, for $r=1/p$ with $p \ge 2$ and for $r=n/(2n+1)=[2,n]$ or $r=(n+1)/(3n+2)=[2,1,n]$ with $n \ge 2$ (see \cite[Main Theorem~2.7]{lee_sakuma_2} and \cite[Main Theorems~2.2 and 2.3]{lee_sakuma_3}). Since there exist a homeomorphism from $(S^3,K(n/(2n+1)))$ to $(S^3, K(2/(2n+1)))$ and a homeomorphism from $(S^3,K((n+1)/(3n+2)))$ to $(S^3, K(3/(3n+2)))$ both of which send the upper/lower tangles to lower/upper tangles, we obtain, from \cite[Main Theorems~2.2 and 2.3]{lee_sakuma_3}, an answer to Question~\ref{question} for $K(r)$ with $r=2/(2n+1)=[n, 2]$ or $r=3/(3n+2)=[n, 1, 2]$. In the present paper, we solve Question~\ref{question} for the remaining cases. \begin{main theorem} \label{main_theorem} Suppose that $r$ is a rational number with $0 < r \le 1/2$ such that $r \neq 1/n$, $r \neq n/(2n+1)$, $r \neq 2/(2n+1)$, $r \neq (n+1)/(3n+2)$ and $r \neq 3/(3n+2)$, where $n \ge 2$ is an integer. Then, for any two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$, the unoriented loops $\alpha_s$ and $\alpha_{s'}$ are never homotopic in $S^3-K(r)$. \end{main theorem} This theorem together with \cite[Main Theorem~2.7]{lee_sakuma_2} and \cite[Main Theorems~2.2 and 2.3]{lee_sakuma_3} implies the following complete answer to Question~\ref{question}. \begin{theorem} \label{summary_theorem} Suppose that $r$ is a rational number such that $0 < r \le 1/2$. For distinct $s, s' \in I_1(r)\cup I_2(r)$, the unoriented loops $\alpha_s$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$ if and only if one of the following holds. \begin{enumerate}[\indent \rm (1)] \item $r=1/p$, where $p \ge 2$ is an integer, and $s=q_1/p_1$ and $s'=q_2/p_2$ satisfy $q_1=q_2$ and $q_1/(p_1+p_2)=1/p$, where $(p_i, q_i)$ is a pair of relatively prime positive integers. \item $r=3/8$, namely $K(r)$ is the Whitehead link, and the set $\{s, s'\}$ equals either $\{1/6, 3/10\}$ or $\{3/4, 5/12\}$. \end{enumerate} \end{theorem} The proof of the main theorem together with \cite[Theorems~2.5 and 2.6]{lee_sakuma_3} implies the following theorems, which give a complete characterization of those simple loops in the $2$-bridge sphere of a hyperbolic $2$-bridge link to be peripheral or imprimitive in the link complement. \begin{theorem} \label{main_corollary} Suppose that $r$ is a rational number with $0 < r \le 1/2$ such that $r \neq 1/n$ for any integer $n \ge 2$. For a rational number $s \in I_1(r) \cup I_2(r)$, the loop $\alpha_s$ is peripheral if and only if one of the following holds. \begin{enumerate}[\indent \rm (1)] \item $r=2/5$ and $s=1/5$ or $s=3/5$. \item $r=n/(2n+1)=[2, n]$ for some integer $n \ge 3$, and $s=(n+1)/(2n+1)$. \item $r=2/(2n+1)=[n,2]$ for some integers $n \ge 3$, and $s=1/(2n+1)$. \end{enumerate} \end{theorem} \begin{theorem} \label{main_corollary2} Suppose that $r$ is a rational number with $0 < r \le 1/2$ such that $r \neq 1/n$, where $n \ge 2$ is an integer. Then, for a rational number $s \in I_1(r) \cup I_2(r)$, the free homotopy class $\alpha_s$ is primitive with the following exceptions. \begin{enumerate}[\indent \rm (1)] \item $r=2/5$, and $s=2/7$ or $s=3/4$. In this case, $\alpha_s$ is the third power of some primitive element in $G(K(r))$. \item $r=3/7$ and $s=2/7$. In this case, $\alpha_s$ is the second power of some primitive element in $G(K(r))$. \item $r=2/7$ and $s=3/7$. In this case, $\alpha_s$ is the second power of some primitive element in $G(K(r))$. \end{enumerate} \end{theorem} We prove the above main theorem by interpreting the situation in terms of combinatorial group theory. In other words, we prove that two words representing the free homotopy classes of $\alpha_s$ and $\alpha_{s'}$ are never conjugate in the $2$-bridge link group $G(K(r))$ for any two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$. The key tool used in the proof is small cancellation theory, applied to two-generator and one-relator presentations of $2$-bridge link groups. \section{Technical lemmas for $r=[m, n]$ or $r=[m, 1, n]$ with $m, n \ge 3$} \label{sec:technical_lemmas} Throughout the remainder of the this paper, we simply write Hypotheses~A, B and C instead of \cite[Hypothesis~A]{lee_sakuma_3}, \cite[Hypothesis~B]{lee_sakuma_3} and \cite[Hypothesis~C]{lee_sakuma_3}, respectively. Throughout this section, we assume that Hypothesis~A holds. Then by \cite[Lemma~3.3]{lee_sakuma_3}, either \cite[Lemma~3.3(1)]{lee_sakuma_3} or \cite[Lemma~3.3(2)]{lee_sakuma_3} holds, that is, either Hypothesis~B or Hypothesis~C holds. Accordingly as Hypothesis~B or Hypothesis~C holds, we shall establish several technical lemmas used for the proof of Main Theorem~\ref{main_theorem} for $r=[m, n]$ and $r=[m, 1, n]$ in Sections~\ref{sec:proof_of_main_theorem_1} and \ref{sec:proof_of_main_theorem_3}, respectively. \subsection{The case when Hypothesis~B holds} We first assume that Hypothesis~B holds. We begin with the following remark before introducing technical lemmas concerning the cyclic sequence $CS(\phi(\alpha))=CS(u_s)=CS(s)$. \begin{remark} \label{rem:(1)holds} {\rm (1) If $r=[m, n]$, where $m, n \ge 3$ are integers, then, by \cite[Lemma~3.16(3)]{lee_sakuma_2}, $CS(r)=\lp m+1, (n-1) \langle m \rangle, m+1, (n-1) \langle m \rangle \rp$, where $S_1=(m+1)$ and $S_2=((n-1) \langle m \rangle)$. So, in Hypothesis~B, both $S(\phi(\partial D_i^+))$ and $S(\phi(\partial D_i^-))$ are exactly of the form $(\ell_1, n_1 \langle m \rangle, m+1, n_2 \langle m \rangle, \ell_2)$, where $0 \le \ell_1, \ell_2 \le m-1$ and $0 \le n_1, n_2 \le n-1$ are integers such that if $n_j=n-1$ then $\ell_j$ is necessarily $0$ for $j=1, 2$. In particular, $S(y_{i,b})=(\ell)$ with $1 \le \ell \le m$, unless $y_i$ is an empty word. The same is true for $S(z_{i,e})$, $S(y_{i,b}')$ and $S(z_{i,e}')$. (2) If $r=[m, 1, n]$, where $m, n\ge 3$ are integers, then, by \cite[Lemma~3.16(1)]{lee_sakuma_2}, $CS(r)=\lp n \langle m+1 \rangle, m, n \langle m+1 \rangle, m \rp$, where $S_1=(n \langle m+1 \rangle)$ and $S_2=(m)$. So, in Hypothesis~B, both $S(\phi(\partial D_i^+))$ and $S(\phi(\partial D_i^-))$ are exactly of the form $(\ell_1, n \langle m+1 \rangle, \ell_2)$, where $0 \le \ell_1, \ell_2 \le m$ are integers. In particular, $S(w_{i,b})=S(w_{i,e})=(m+1)$ and $S(w_{i,b}')=S(w_{i,e}')=(m+1)$. } \end{remark} \begin{lemma} \label{lem:case1-1(a)} Let $r=[m,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, suppose that $v$ is a subword of the cyclic word represented by $\phi(\alpha) \equiv y_1 w_1 z_1 y_2 w_2 z_2 \cdots y_t w_t z_t$ such that $v$ corresponds to a term of $CS(\phi(\alpha))=CS(s)$. Then, after a cyclic shift of indices, $v$ is equal to one of the following subwords: \[ z_{0, e}w_1w_2\cdots w_qy_{q+1, b}, \quad z_{0, e}w_1w_2\cdots w_q, \quad w_1w_2\cdots w_qy_{q+1, b}, \quad w_1w_2\cdots w_q, \] where $q\in \ZZ_+\cup\{0\}$ in the first three cases and $q\in \ZZ_+$ in the last case. In each of the above, the ``intermediate subwords'' are empty; to be precise, when we say that $z_{0, e}w_1w_2\cdots w_qy_{q+1, b}$, for example, is a subword of $(u_s)$, we assume that $y_1$, $z_iy_{i+1}$ ($1\le i\le q-1$) and $z_q$ are empty words. \end{lemma} \begin{proof} The proof can be done in the same way as in \cite[Lemma~3.6]{lee_sakuma_3}. \end{proof} Throughout the remainder of this paper, we will assume the following convention. \begin{convention} \label{con:figure} {\rm In Figures~\ref{fig.lemma-1-1}--\ref{fig.general_proof_1}, except for Figure~\ref{fig.converging}, the change of directions of consecutive arrowheads represents the change from positive (negative, resp.) words to negative (positive, resp.) words, and a dot represents a vertex whose position is clearly identified. Also an Arabic number represents the length of the corresponding positive (or negative) word. In Figures~\ref{fig.Lemma7_3_1}--\ref{fig.Lemma7_3_2} and \ref{fig.S_1_occurs}--\ref{fig.S_1_occurs(c)}, the label $S_1$ or $S_2$ on an oriented segment means that the $S$-sequence of the corresponding word is equal to $S_1$ or $S_2$ accordingly.} \end{convention} \begin{lemma} \label{lem:case1-1(b)} Let $r=[m,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item $S(z_{i, e}y_{i+1, b}) \ne (m+d)$ for any integer $d$ with $1 \le d \le m-1$. \item $S(w_iz_iy_{i+1, b}) \ne (m+1+d)$ and $S(z_{i, e}y_{i+1}w_{i+1}) \ne (m+1+d)$ for any integer $d$ with $1 \le d \le m-1$. \item $S(w_iz_iy_{i+1, b}) \neq (2m+1)$ and $S(z_{i, e}y_{i+1}w_{i+1}) \neq (2m+1)$. \item $S(w_iz_iy_{i+1}w_{i+1}) \ne (2m+2)$. \item $S(w_iz_iy_{i+1}w_{i+1}) \ne (m+1, m+1)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that the assertion does not hold. Without loss of generality, we may assume that $S(z_{1,e}y_{2,b})=(m+d)$ for some $1 \le d \le m-1$. Then, since $0 \le |z_{1,e}|, |y_{2,b}| \le m$, we have $|z_{1,e}|=k$ and $|y_{2,b}|=m+d-k$ for some $k$ with $d \le k \le m$. Here, we assume $d\le k\le m-1$ and hence $|z_{1,e}'|=m-k \ne 0$. (The case when $k=m$ and hence $|z_{1,e}'|= 0$ can be treated similarly.) Then $J$ is locally as illustrated in Figure~\ref{fig.lemma-1-1}(a) which follows Convention~\ref{con:figure}. The numbers $k$ and $m+d-k$ near the upper boundary represent the lengths of the words $z_{1,e}$ and $y_{2,b}$, respectively, whereas the numbers $m-k$ and $k-d$ near the lower boundary represent the lengths of the words $z_{1,e}'$ and $y_{2,b}'$, respectively, and the change of directions of consecutive arrowheads represents the change from positive (negative, resp.) words to negative (positive, resp.) words. Suppose first that $J=M$. Then we see from Figure~\ref{fig.lemma-1-1}(a) that $CS(\phi(\delta^{-1}))=CS(u_{s'}^{\pm 1})=CS(s')$ involves a term $(m-k)+(k-d)=m-d$. Moreover, $CS(s')$ also involves a term of the form $m+1+c$ with $c \in \ZZ_+ \cup \{0\}$, because $S(w_1')=S_1=(m+1)$. Since $m-d \le m-1$, this is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2} which says that either $CS(s')$ is equal to $\lp \ell,\ell \rp$ or $CS(s')$ consists of $\ell$ and $\ell+1$ for some $\ell \in \ZZ_+$. Suppose next that $J \subsetneq M$. Since none of $S(\phi(e_{1}'))$ and $S(\phi(e_{2}'))$ contains $S_1=(m+1)$ as a subsequence (see \cite[Lemma~3.1(1)]{lee_sakuma_3}), we see that the initial vertex of $e_2'$ lies in the interior of the segment of $\partial D_1^-$ corresponding to $S_1=(m+1)$. Similarly, the terminal vertex of $e_3'$ lies in the interior of the segment of $\partial D_2^-$ corresponding to $S_1=(m+1)$. Hence, we see from Figure~\ref{fig.lemma-1-1}(b) that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m-d, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. This yields that a term $m-d$ occurs in $CS(\phi(\partial D_1'))=CS(r)=\lp m+1, (n-1)\langle m \rangle, m+1, (n-1)\langle m \rangle\rp$, which is obviously a contradiction. \begin{figure}[h] \includegraphics{figure_general_lemma-1-1.eps} \caption{ Lemma~\ref{lem:case1-1(b)}(1) where $S(z_{1,e}y_{2,b})=(k+(m+d-k))$} \label{fig.lemma-1-1} \end{figure} (2) Suppose on the contrary that $S(w_1z_1y_{2, b})=(m+1+d)$ for some $1 \le d \le m-1$. (The other case is treated similarly.) Then, since $w_1$ and $z_1$ have different signs when $z_1$ is nonempty and since $|w_1|=m+1$ and $0 \le |y_{2,b}| \le m$, the only possibility is that $|z_1|=0$, $|y_{2, b}|=d$ and $S(w_1y_{2,b})=(m+1+d)$. If $J=M$, then we see from Figure~\ref{fig.lemma-1-2}(a) that $CS(s')$ involves both a term $m-d$ and a term of the form $m+1+c$ with $c \in \ZZ_+ \cup \{0\}$. Since $m-d \le m-1$, this is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then we see, by using \cite[Lemma~3.1(1)]{lee_sakuma_3} as in the proof of Lemma~\ref{lem:case1-1(b)}(1), that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m-d, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ (see Figure~\ref{fig.lemma-1-2}(b)). This implies that $CS(\phi(\partial D_1'))=CS(r)$ has a term $m-d$, a contradiction. \begin{figure}[h] \includegraphics{figure_general_lemma-1-2.eps} \caption{ Lemma~\ref{lem:case1-1(b)}(2) where $S(w_1z_1y_{2, b})=((m+1)+0+d)$} \label{fig.lemma-1-2} \end{figure} (3) Suppose on the contrary that $S(z_{1,e}y_2w_2)=(2m+1)$. (The other case is treated similarly.) Then $|z_{1,e}|=m$ and $|y_2|=0$. We note that $z_1=z_{1,e}$. Otherwise, $S(w_1z_1)$ contains a subword $(m+1,m,m)$ and hence $CS(\phi(\alpha))=CS(s)$ contains a term $m$ and $2m+1$, a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. Hence, we see that $CS(w_1'z_1'y_2'w_2')$ contains a term $2m$ (see Figure~\ref{fig.lemma-1-3}(a)). If $J=M$, then we see from Figure~\ref{fig.lemma-1-3}(a) that $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $m$ and a term $2m$. Since $m \ge 3$ implying that $m+2 \le 2m$, this gives a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then we see, by using \cite[Lemma~3.1(1)]{lee_sakuma_3} as in the proof of Lemma~\ref{lem:case1-1(b)}(1), that $S(\phi(e_2'e_3'))$ is of the form $(\ell_1, 2m, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. This implies that a term $2m$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. (4) This can be proved by an argument parallel to the proof of (3) (see Figure~\ref{fig.lemma-1-3}(b)). \begin{figure}[h] \includegraphics{figure_general_lemma-1-3.eps} \caption{ (a) Lemma~\ref{lem:case1-1(b)}(3) where $S(z_{1,e}y_2w_2)=(m+0+(m+1))$, and (b) Lemma~\ref{lem:case1-1(b)}(4) where $S(w_1z_1y_2w_2)=((m+1)+0+0+(m+1))$} \label{fig.lemma-1-3} \end{figure} (5) Suppose on the contrary that $S(w_1z_1y_2w_2)=(m+1, m+1)$. Then $|z_1|=|y_2|=0$ and $S(w_1w_2)=(m+1, m+1)$. If $J=M$ (see Figure~\ref{fig.lemma-1-8}(a)), then $CS(s')$ contains both a subsequence $((n-1) \langle m \rangle)$ and a term of the form $m+1+c$ with $c \in \ZZ_+ \cup \{0\}$. Here, if $c=0$, then $s' \notin I_1(r) \cup I_2(r)$ by \cite[Proposition~3.19(1)]{lee_sakuma_2}, contradicting the hypothesis of the theorem, while if $c>0$, then we have a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$ (see Figure~\ref{fig.lemma-1-8}(b)), then we see, by using \cite[Lemma~3.1(1)]{lee_sakuma_3} as in the proof of Lemma~\ref{lem:case1-1(b)}(1), that $S(\phi(e_2'e_3'))$ is of the form $(\ell_1, 2(n-1) \langle m \rangle, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. This implies that a subsequence $(2(n-1) \langle m \rangle)$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. \end{proof} \begin{figure}[h] \includegraphics{figure_general_lemma-1-8.eps} \caption{ Lemma~\ref{lem:case1-1(b)}(5) where $S(w_1z_1y_2w_2)=(m+1, m+1)$} \label{fig.lemma-1-8} \end{figure} \begin{lemma} \label{lem:case1-1} Let $r=[m,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, the following hold. \begin{enumerate}[\indent \rm (1)] \item No two consecutive terms of $CS(s)$ can be $(m+1, m+1)$. \item No term of $CS(s)$ can be of the form $m+1+d$ except $2m$, where $d \in \ZZ_+$. \item No two consecutive terms of $CS(s)$ can be $(2m, 2m)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that $CS(\phi(\alpha))=CS(u_s)=CS(s)$ contains $(m+1, m+1)$ as a subsequence. Let $v=v'v''$ be a subword of the cyclic word $(u_s)$ corresponding to a subsequence $(m+1, m+1)$, where $S(v')=S(v'')=(m+1)$. By using Lemma~\ref{lem:case1-1(a)} and the facts that $0 \le |z_{i,e}|, |y_{i,b}| \le m$ and $|w_i|=m+1$, we see that one of the following holds after a shift of indices. \begin{enumerate}[\indent \rm (i)] \item $(v',v'')=(z_{1,e}y_{2,b},w_2)$, where $S(z_{1,e}y_{2,b})=(m+1)$. \item $(v',v'')=(w_1,z_{1,e}y_{2,b})$, where $S(z_{1,e}y_{2,b})=(m+1)$. \item $(v',v'')=(w_1,w_2)$. \end{enumerate} However, (i) and (ii) are impossible by Lemma~\ref{lem:case1-1(b)}(1), and (iii) is impossible by Lemma~\ref{lem:case1-1(b)}(5). (2) Suppose on the contrary that $CS(s)$ contains a term $m+1+d$ except $2m$. Let $v$ be a subword of the cyclic word $(u_s)$ corresponding to a term $m+1+d$ except $2m$. By using Lemma~\ref{lem:case1-1(a)} and the facts that $0 \le |z_{i,e}|, |y_{i,b}| \le m$ and $|w_i|=m+1$, we see that one of the following holds after a cyclic shift of indices. \begin{enumerate}[\indent \rm (i)] \item $v=z_{0, e}y_{1, b}$ with $|z_{0, e}|, |y_{1, b}| \neq 0$. \item $v$ contains $z_{0, e}w_1$ with $|z_{0, e}| \neq 0$. \item $v$ contains $w_1y_{2, b}$ with $|y_{2, b}| \neq 0$. \item $v$ contains $w_1w_2$ with $|z_1|=|y_2|=0$. \end{enumerate} However, (i) is impossible by Lemma~\ref{lem:case1-1(b)}(1). By Lemma~\ref{lem:case1-1(b)}(2), $|z_{0, e}|=m$ provided (ii) occurs, and $|y_{2, b}|=m$ provided (iii) occurs. Hence (ii) and (iii) are impossible by Lemma~\ref{lem:case1-1(b)}(3). Also (iv) is impossible by Lemma~\ref{lem:case1-1(b)}(4). (3) Suppose on the contrary that $CS(s)$ contains $(2m, 2m)$ as a subsequence. Let $v=v'v''$ be a subword of the cyclic word $(u_s)$ corresponding to a subsequence $(2m, 2m)$, where $S(v')=S(v'')=(2m)$. If $v'$ contains $w_i$ for some $i$, then we see, by using Lemma~\ref{lem:case1-1(a)} and the identity $|w_i|=m+1$, that either $v'=w_iz_iy_{i+1,b}$ with $(|z_i|,|y_{i+1,b}|)=(0,m-1)$ or $v'=z_{i-1,e}y_iw_i$ with $(|z_{i-1}|,|y_i|)=(m-1,0)$. However both cases are impossible by Lemma~\ref{lem:case1-1(b)}(2). Thus $v'$ cannot contain $w_i$. Since $S(v')=(2m)$ is a term of $CS(s)$, this implies that $v'$ is disjoint from $w_i$ for every $i$. The same conclusion also holds for $v''$, and hence for $v=v'v''$. Thus $v$ is a subword of $z_iy_{i+1}$ for some $i$. But then $S(v)$ contains a term $2m$ at most once, a contradiction. \end{proof} \begin{corollary} \label{cor:case1-1} Let $r=[m,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, $CS(s)$ consists of $m$ and $m+1$. \end{corollary} \begin{proof} By \cite[Lemma~3.8]{lee_sakuma_2}, either $CS(s)=\lp \ell, \ell \rp$ or $CS(s)$ consists of $\ell$ and $\ell+1$ with $\ell \in \ZZ_+$. By Hypothesis~B together with Remark~\ref{rem:(1)holds}(1), $\phi(\alpha)$ involves a subword $w_i$ whose $S$-sequence is $(m+1)$, so $CS(\phi(\alpha))=CS(u_s)=CS(s)$ must contain a term of the form $m+1+c$, where $c \in \ZZ_+ \cup \{0\}$. First consider the case where $CS(s)=\lp \ell, \ell \rp$. Since $CS(s)$ contains a term $m+1+c$, we have $\ell \ge m+1$. Here, by Lemma~\ref{lem:case1-1}(1), $\ell$ is not equal to $m+1$. By Lemma~\ref{lem:case1-1}(2), $\ell$ is not equal to $m+1+d$ for any $d \in \ZZ_+$ except $2m$. However, Lemma~\ref{lem:case1-1}(3) implies that $\ell$ is not equal to $2m$, so that there remains no possibility for $\ell$. Next consider the case where $CS(s)$ consists of $\ell$ and $\ell+1$. By Lemma~\ref{lem:case1-1}(2), none of $\ell$ and $\ell+1$ is equal to $m+1+d$ for any $d \in \ZZ_+$ except $2m$. So $\ell \le m$. On the other hand, since $CS(s)$ contains a term $m+1+c$, we have $\ell+1 \ge m+1$. Therefore $\ell=m$. \end{proof} Next, we study the case where $r=[m, 1, n]$ with $m, n \ge 3$. Recall from Remark~\ref{rem:(1)holds}(2) that $CS(r)=\lp n \langle m+1 \rangle, m, n \langle m+1 \rangle, m \rp$, where $S_1=(n \langle m+1 \rangle)$ and $S_2=(m)$. Recall also $S(w_{i,b})=S(w_{i,e})=(m+1)$ and $S(w_{i,b}')=S(w_{i,e}')=(m+1)$ for every $i$. \begin{lemma} \label{lem:case3-1_easy} Let $r=[m,1,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, $CS(s)$ contains $m+1$ as a term. \end{lemma} \begin{proof} The assertion immediately follows from the fact that $\phi(\alpha)=CS(u_s)=CS(s)$ involves a subword $w_i$ whose $S$-sequence is $(n \langle m+1 \rangle)$ with $n \ge 3$. \end{proof} The following lemma is a counterpart of Lemma~\ref{lem:case1-1(b)}. \begin{lemma} \label{lem:case3-1(b)} Let $r=[m,1,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item $S(z_iy_{i+1}) \neq (m+2)$. \item $S(w_{i,e}z_iy_{i+1}) \neq (m+2)$ and $S(z_iy_{i+1}w_{i+1,b}) \neq (m+2)$. \end{enumerate} \end{lemma} \begin{proof} The proofs of (1) and (2) are parallel to those of Lemma~\ref{lem:case1-1(b)}(1) and (2), respectively. \end{proof} \begin{lemma} \label{lem:case3-1} Let $r=[m,1,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, no term of $CS(s)$ can be of the form $m+2$. \end{lemma} \begin{proof} Suppose on the contrary that $CS(s)$ contains a term of the form $m+2$. Let $v$ be a subword of the cyclic word $(u_s)$ corresponding to a term $m+2$. Without loss of generality, we may assume that \begin{enumerate}[\indent \rm (i)] \item $v=z_0y_1$ with $|z_0|, |y_1| \neq 0$; \item $v=z_0w_{1, b}$ with $|z_0| \neq 0$; or \item $v=w_{1, e}y_2$ with $|y_2| \neq 0$. \end{enumerate} However, (i) is impossible by Lemma~\ref{lem:case3-1(b)}(1), and (ii) and (iii) are impossible by Lemma~\ref{lem:case3-1(b)}(2). \end{proof} \begin{corollary} \label{cor:case2-1} Let $r=[m,1,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~B, either $CS(s)=\lp m+1, m+1 \rp$ or $CS(s)$ consists of $m$ and $m+1$. \end{corollary} \begin{proof} By Lemma~\ref{lem:case3-1_easy}, $CS(s)$ contains a term $m+1$. Also by Lemma~\ref{lem:case3-1}, $CS(s)$ does not contain a term $m+2$. Hence, by \cite[Lemma~3.8]{lee_sakuma_2}, we obtain the desired result. \end{proof} \subsection{The case when Hypothesis~C holds} We next assume that Hypothesis~C holds. We also begin with the following remark before introducing two technical lemmas concerning the sequence $S(z_iy_{i+1})$ accordingly as $r=[m, n]$ and $r=[m, 1, n]$. \begin{remark} \label{rem:(2)holds} {\rm (1) If $r=[m, n]$, where $m, n \ge 3$ are integers, then $CS(r)=\lp m+1, (n-1) \langle m \rangle, m+1, (n-1) \langle m \rangle \rp$, where $S_1=(m+1)$ and $S_2=((n-1) \langle m \rangle)$. So, in Hypothesis~C, both $S(\phi(\partial D_i^+))$ and $S(\phi(\partial D_i^-))$ are exactly of the form $(\ell_1, (n-1) \langle m \rangle, \ell_2)$, where $1 \le \ell_1, \ell_2 \le m$ are integers. (2) If $r=[m, 1, n]$, where $m, n \ge 3$ are integers, then $CS(r)=\lp n \langle m+1 \rangle, m, n \langle m+1 \rangle, m \rp$, where $S_1=(n \langle m+1 \rangle)$ and $S_2=(m)$. So, in Hypothesis~C, both $S(\phi(\partial D_i^+))$ and $S(\phi(\partial D_i^-))$ are exactly of the form $(\ell_1, n_1 \langle m+1 \rangle, m, n_2 \langle m+1 \rangle, \ell_2)$, where $0 \le \ell_1, \ell_2 \le m$ and $0 \le n_1, n_2 \le n-1$ are integers such that a pair $(\ell_j, n_j)$ cannot be $(0, 0)$ for $j=1,2$. In particular, $S(y_{i,b})=(\ell)$ with $1 \le \ell \le m+1$. The same is true for $S(z_{i,e})$, $S(y_{i,b}')$ and $S(z_{i,e}')$. } \end{remark} \begin{lemma} \label{lem:case1-2} Let $r=[m,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~C, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item $S(z_iy_{i+1})$ is not equal to $(m-1)$ nor $ (m)$. \item $S(z_iy_{i+1})$ is not equal to $(m-1, m)$, $(m, m-1)$ nor $(m, m)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that $S(z_1y_2)=(m)$. (The other case is treated analogously.) Since $|z_1|,|y_2|>0$, we have $|z_1|=d$ and $|y_2|=m-d$ for some $1 \le d \le m-1$. Here, if $J=M$ (see Figure~\ref{fig.II-lemma-1-1}(a)), then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $m$ and a term $m+2$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, let $J \subsetneq M$. Since none of $S(\phi(e_j'))$ contains $S_2$ in its interior (see \cite[Lemma~3.1(2)]{lee_sakuma_3}), wee see that the initial vertex of $e_2'$ lies in the (central) segment of $\partial D_1^-$ corresponding to $S_2=((n-1) \langle m \rangle)$ and that the terminal vertex of $e_3'$ lies in the (central) segment of $\partial D_2^-$ corresponding to $S_2=((n-1) \langle m \rangle)$. Thus we see that $CS(\phi(\partial D_1'))=CS(r)$ contains a term of the form $m+2+c$ with $c \in \ZZ_+ \cup \{0\}$, as illustrated in Figure~\ref{fig.II-lemma-1-1}(b), a contradiction. \begin{figure}[h] \includegraphics{figure_general_II_lemma-1-1.eps} \caption{ Lemma~\ref{lem:case1-2}(1) where $S(z_1y_2)=(d+(m-d))$} \label{fig.II-lemma-1-1} \end{figure} (2) Suppose on the contrary that $S(z_1y_2)=(m-1, m)$. (The other cases are treated similarly.) Then $|z_1|=m-1$ and $|y_2|=m$. Here, if $J=M$ (see Figure~\ref{fig.II-lemma-1-2}(a)), then $CS(\phi (\delta^{-1}))=CS(s')$ contains both a term $1$ and a term $m$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$ (see Figure~\ref{fig.II-lemma-1-2}(b)), then $|\phi(e_2')|=2$ and $|\phi(e_3')|=1$, for otherwise we see, by using \cite[Lemma~3.1(2)]{lee_sakuma_3} as in the proof of (1), that a subsequence of the form $(\ell_1, 1, \ell_2)$ or of the form $(\ell_1, 2, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ would occur in $S(\phi(e_2'e_3'))$ which implies that $CS(\phi(\partial D_1'))=CS(r)$ would contain a term $1$ or a term $2$, a contradiction. Assuming that $e_2', e_3', {e_3''}^{-1}, {e_2''}^{-1}$ is a boundary cycle of $D_1'$, this implies that $\phi(e_2''e_3'')$ contains a subword $w$ such that $S(w)$ contains $(S_1,S_2)=(m+1, (n-1) \langle m \rangle)$ as a proper initial subsequence or $(S_2,S_1)=((n-1) \langle m \rangle,m+1)$ as a proper terminal subsequence. But this is impossible by \cite[Corollary~3.25(2)]{lee_sakuma_2}. \end{proof} \begin{figure}[h] \includegraphics{figure_general_II_lemma-1-2.eps} \caption{ Lemma~\ref{lem:case1-2}(2) where $S(z_1y_2)=(m-1, m)$} \label{fig.II-lemma-1-2} \end{figure} \begin{lemma} \label{lem:case3-2} Let $r=[m,1,n]$, where $m, n \ge 3$ are integers. Under Hypothesis~C, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item $S(z_{i,e}y_{i+1,b})$ is not equal to $(m-1)$ nor $(m)$. \item $S(z_{i,e}y_{i+1,b})$ is not equal to $(m-1, m)$, $(m, m-1)$ nor $(m, m)$. \item $S(z_{i,e}y_{i+1,b})$ is not equal to $(m-1, m-1)$. \item $S(z_{i,e}y_{i+1,b})$ is not equal to $(m, m+1)$ nor $(m+1, m)$. \end{enumerate} \end{lemma} \begin{proof} The proofs of (1) and (2) are parallel to those of Lemma~\ref{lem:case1-2}(1) and (2), respectively. (3) Suppose on the contrary that $S(z_{1,e}y_{2,b})=(m-1, m-1)$. Then $|z_{1,e}|=|y_{2,b}|=m-1$. Moreover, we must have $z_1=z_{1,e}$ and $y_2=y_{2,b}$. To see this, suppose $z_1\ne z_{1,e}$. (The other case is treated similarly.) Then, since $S_1=(n\langle m+1\rangle)$, we see that $S(w_1z_1)$ is of the form $(S_2,m+1,*)$, where $*$ is nonempty. So, $CS(\phi(\alpha))=CS(s)$ contains a term $m+1$. This is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}, because $CS(\phi(\alpha))=CS(s)$ also contains $m-1$ by the assumption. Here, if $J=M$ (see Figure~\ref{fig.II-lemma-3-3}(a)), then $CS(\phi(\delta^{-1}))=CS(s')$ involves both a term $2$ and a term $m+1$. Since $m+1 \ge 4$, we obtain a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then by \cite[Lemma~3.1(2)]{lee_sakuma_3} the initial vertex of $e_2'$ lies in the segment of $\partial D_1^-$ corresponding to $S_2=(m)$ and that the terminal vertex of $e_3'$ lies in the segment of $\partial D_2^-$ corresponding to $S_2=(m)$. This implies that a subsequence of the form $(\ell_1, 2, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ occurs in $S(\phi(e_2'e_3'))$ (see Figure~\ref{fig.II-lemma-3-3}(b)), so in $CS(\phi(\partial D_1'))=CS(r)$ contains a term $2$, a contradiction. \begin{figure}[h] \includegraphics{figure_general_II_lemma-3-3.eps} \caption{ Lemma~\ref{lem:case3-2}(3) where $S(z_{1,e}y_{2,b})=(m-1, m-1)$} \label{fig.II-lemma-3-3} \end{figure} (4) Suppose on the contrary that $S(z_{1,e}y_{2,b})=(m, m+1)$. (The other case is treated analogously.) Then $|z_{1,e}|=m$ and $|y_{2,b}|=m+1$. Here, if $J=M$ (see Figure~\ref{fig.II-lemma-3-4}), then $CS(\phi(\delta^{-1}))=CS(s')$ involves both a term $m$ and a term $m+2$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then we see, by using \cite[Lemma~3.1(2)]{lee_sakuma_3} as in the proof of (3), that a term of the form $m+2+c$ with $c \in \ZZ_+ \cup \{0\}$ occurs in $S(\phi(e_2'e_3'))$, so in $CS(\phi(\partial D_1'))=CS(r)$ contains a term $m+2+c$, a contradiction. \end{proof} \begin{figure}[h] \includegraphics{figure_general_II_lemma-3-4.eps} \caption{ Lemma~\ref{lem:case3-2}(4) where $S(z_{1,e}y_{2,b})=(m, m+1)$} \label{fig.II-lemma-3-4} \end{figure} \section{Proof of Main Theorem~\ref{main_theorem} for $r=[m,n]$ with $m, n\ge 3$} \label{sec:proof_of_main_theorem_1} Suppose $r=[m,n]$, where $m, n \ge 3$ are integers. For two distinct elements $s, s' \in I_1(r) \cup I_2(r)$, suppose on the contrary that the unoriented loops $\alpha_s$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$, namely suppose that Hypothesis~A holds. We will derive a contradiction in each case to consider. By \cite[Lemma~3.3]{lee_sakuma_3}, there are two big cases to consider. \medskip \noindent {\bf Case 1.} {\it Hypothesis~B holds.} \medskip By Corollary~\ref{cor:case1-1}, $CS(s)$ consists of $m$ and $m+1$. Without loss of generality, we may assume that a term $m$ occurs in $S(z_1y_2)$. There are three possibilities: \begin{enumerate}[\indent \rm (i)] \item $S(z_1y_2)$ consists of only $m$, where $S(z_1)=(n_1\langle m \rangle)$, $S(y_2)=(n_2\langle m \rangle)$, and $S(z_1y_2)=((n_1+n_2)\langle m \rangle)$ with $n_1,n_2\in\ZZ_+\cup\{0\}$; \item $S(z_1y_2)$ consists of only $m$, where $S(z_1)=(n_1\langle m \rangle,d)$, $S(y_2)=(m-d,n_2\langle m \rangle)$, and $S(z_1y_2)=((n_1+n_2+1)\langle m \rangle)$ with $n_1,n_2\in\ZZ_+\cup\{0\}$ and $d \in \{1, 2, \dots, m-1\}$; \item $S(z_1y_2)$ consists of $m$ and $m+1$. \end{enumerate} First assume that (i) occurs. Then $S(z_1'y_2')=((n_1'+n_2')\langle m \rangle)$ where $n_1'=(n-1)-n_1$ and $n_2'=(n-1)-n_2$. So $n_1'+n_2'=2(n-1)-(n_1+n_2)$ and hence either $S(z_1y_2)$ or $S(z_1'y_2')$ contains $n-1$ consecutive $m$'s. If $J=M$, then this implies that either $s \notin I_1(r) \cup I_2(r)$ or $s' \notin I_1(r) \cup I_2(r)$ by \cite[Proposition~3.19(1)]{lee_sakuma_2}, contradicting the hypothesis of the theorem. On the other hand, if $J \subsetneq M$, then the above observation implies that either $S(z_1y_2)$ contains $n-1$ consecutive $m$'s and so $s \notin I_1(r) \cup I_2(r)$, or otherwise $S(z_1'y_2')$ contains $n$ consecutive $m$'s. The former case is impossible by the assumption. In the latter case, we see, by an argument using \cite[Lemma~3.1(1)]{lee_sakuma_3} as in the last step of the proof of Lemma~\ref{lem:case1-1(b)}(1), that a subsequence of the form $(\ell_1, n \langle m \rangle, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ occurs in $S(\phi(e_2'e_3'))$, so in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. Next assume that (ii) occurs. Then $S(z_1'y_2')=((n_1'+n_2'+1)\langle m \rangle)$, where $n_1'=(n-2)-n_1$ and $n_2'=(n-2)-n_2$. By using the identity $n_1'+n_2'+1=2(n-1)-(n_1+n_2+1)$, this case is treated as in the case when (i) occurs. Finally assume that (iii) occurs. Then we must have $S(z_{1, e}y_{2, b})=(m+1)$, but this is a contradiction to Lemma~\ref{lem:case1-1(b)}(1). \medskip \noindent {\bf Case 2.} {\it Hypothesis~C holds.} \medskip By Remark~\ref{rem:(2)holds}(1), $CS(s)$ contains $((n-1)\langle m\rangle)$ as a proper subsequence. Thus \cite[Lemma~3.8]{lee_sakuma_2} implies that $CS(s)$ consists of $\{m-1,m\}$ or $\{m,m+1\}$. Moreover, since $n\ge 3$, this together with \cite[Lemma~3.8]{lee_sakuma_2} implies that $CS(s)$ does not contain $(m-1,m-1)$ nor $(m+1,m+1)$ as a subsequence. \medskip \noindent {\bf Case 2.a.} {\it $CS(s)$ consists of $m-1$ and $m$.} \medskip In this case, a term $m-1$ should occur in $S(z_iy_{i+1})$ for some $i$. Since $CS(s)$ does not contain $(m-1,m-1)$ and since each of $S(z_i)$ and $S(y_i)$ has length $1$, we see that $S(z_iy_{i+1})$ is equal to $(m-1)$, $(m,m-1)$ or $(m-1,m)$. But, this is impossible by Lemma~\ref{lem:case1-2}. \medskip \noindent {\bf Case 2.b.} {\it $CS(s)$ consists of $m$ and $m+1$.} \medskip In this case, $CS(s)$ contains both $S_1=(m+1)$ and $S_2=((n-1)\langle m\rangle)$ as subsequences. Hence, by \cite[Proposition~3.19(1)]{lee_sakuma_2}, $s \notin I_1(r) \cup I_2(r)$, contradicting the hypothesis of the theorem. \qed \section{Proof of Main Theorem~\ref{main_theorem} for $r=[m,1,n]$ with $m, n \ge 3$} \label{sec:proof_of_main_theorem_3} Suppose $r=[m,1,n]$, where $m, n \ge 3$ are integers. For two distinct elements $s, s' \in I_1(r) \cup I_2(r)$, suppose on the contrary that the unoriented loops $\alpha_s$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$, namely suppose that Hypothesis~A holds. We will derive a contradiction in each case to consider. By \cite[Lemma~3.3]{lee_sakuma_3}, there are two big cases to consider. \medskip \noindent {\bf Case 1.} {\it Hypothesis~B holds.} \medskip By Corollary \ref{cor:case2-1}, we have the following two subcases. \medskip \noindent {\bf Case 1.a.} {\it $CS(s)=\lp m+1, m+1 \rp$.} \medskip Since $\phi(\alpha)$ involves a subword $w_i$ whose $S$-sequence is $(n \langle m+1 \rangle)$ and since $n\ge 3$, this is impossible. \medskip \noindent {\bf Case 1.b.} {\it $CS(s)$ consists of $m$ and $m+1$.} \medskip In this case, we see that $CS(s)$ contains a subsequence $S_1=(n \langle m+1 \rangle)$. Since it also contains a subsequence $S_2=(m)$ by the assumption, we see by \cite[Proposition~3.19(1)]{lee_sakuma_2} that $s \notin I_1(r) \cup I_2(r)$, contradicting the hypothesis of the theorem. \medskip \noindent {\bf Case 2.} {\it Hypothesis~C holds.} \medskip By Remark~\ref{rem:(2)holds}(2), $CS(s)$ contains the term $m$. So, by \cite[Lemma~3.8]{lee_sakuma_2}, Case~2 is reduced to the following three subcases: $CS(s)=\lp m, m \rp$, $CS(s)$ consists of $\{m-1, m\}$, or $CS(s)$ consists of $\{m, m+1\}$. \medskip \noindent {\bf Case 2.a.} {\it $CS(s)=\lp m, m \rp$.} \medskip In this case, there is only one possibility: $J$ consists of one $2$-cell, namely $CS(\phi(\alpha))=CS(\phi(\partial D_1^+))=\lp S(z_1y_1), S(w_1) \rp= \lp m, m \rp$. Then $S(z_1y_1)=S(z_{1,e}y_{1,b})=(m)$, contradicting Lemma~\ref{lem:case3-2}(1). \medskip \noindent {\bf Case 2.b.} {\it $CS(s)$ consists of $m-1$ and $m$.} \medskip In this case, a term $m-1$ must occur in $S(z_{i,e}y_{i+1,b})$ for some $i$. But by Lemma~\ref{lem:case3-2}(1), (2) and (3), this is impossible. \medskip \noindent {\bf Case 2.c.} {\it $CS(s)$ consists of $m$ and $m+1$.} \medskip Without loss of generality, we may assume that $m+1$ occurs in $S(z_1y_2)$. There are three possibilities: \begin{enumerate}[\indent \rm (i)] \item $S(z_1y_2)$ consists of only $m+1$, where $S(z_1)=(n_1\langle m+1 \rangle)$, $S(y_2)=(n_2\langle m+1 \rangle)$, and $S(z_1y_2)=((n_1+n_2)\langle m+1 \rangle)$ with $n_1,n_2\in\ZZ_+\cup\{0\}$; \item $S(z_1y_2)$ consists of only $m+1$, where $S(z_1)=(n_1\langle m+1 \rangle,d)$, $S(y_2)=(m+1-d,n_2\langle m+1 \rangle)$, and $S(z_1y_2)=((n_1+n_2+1)\langle m+1 \rangle)$ with $n_1,n_2\in\ZZ_+\cup\{0\}$ and $d \in \{1, \dots, m\}$; \item $S(z_1y_2)$ consists of $m$ and $m+1$. \end{enumerate} However, by an argument as in Case~1(i) and (ii) in Section~\ref{sec:proof_of_main_theorem_1}, we can see that neither (i) nor (ii) can happen. (In the above, we need to appeal to \cite[Lemma~3.1(2)]{lee_sakuma_3} instead of \cite[Lemma~3.1(1)]{lee_sakuma_3}.) So only (iii) can occur. But then a term $m$ must occur in $S(z_{1,e}y_{2,b})$, a contradiction to Lemma~\ref{lem:case3-2}(1), (2) and (4). \qed \section{Preliminary results for the general cases} \label{sec:technical_lemmas_general} The remainder of this paper is devoted to the proof of Main Theorem~\ref{main_theorem} when $r$ is {\it general}, namely either $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$, or $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. In the remainder of this paper, we assume that $r$ is general. \begin{remark} \label{rem:general_decomposition} {\rm (1) Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Then $\tilde{r}=[m_2-1, m_3, \dots, m_k]$ by \cite[Lemma~3.11]{lee_sakuma_2}, and so, by \cite[Proposition~3.12]{lee_sakuma_2}, $CS(\tilde{r})=\lp T_1,T_2,T_1,T_2 \rp$, where $T_1$ begins and ends with $m_2$ and $T_2$ begins and ends with $m_2-1$. Thus by \cite[Lemma~3.16(4)]{lee_sakuma_2}, we obtain that $CS(r)=\lp S_1, S_2, S_1, S_2 \rp$, where $S_1$ begins and ends with $(m+1, (m_2-1) \langle m \rangle, m+1)$, and $S_2$ begins and ends with $(m_2 \langle m \rangle)$. (2) Let $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Then $\tilde{r}=[m_3, \dots, m_k]$ by \cite[Lemma~3.11]{lee_sakuma_2}, and so, by \cite[Proposition~3.12]{lee_sakuma_2}, $CS(\tilde{r})=\lp T_1,T_2,T_1,T_2 \rp$, where $T_1$ begins and ends with $m_3+1$ and $T_2$ begins and ends with $m_3$. Thus by \cite[Lemma~3.16(2)]{lee_sakuma_2}, we obtain that $CS(r)=\lp S_1, S_2, S_1, S_2 \rp$, where $S_1$ begins and ends with $((m_3+1) \langle m+1 \rangle)$, and $S_2$ begins and ends with $(m, m_3 \langle m+1 \rangle, m)$. } \end{remark} The aim of this section is to prove the following proposition. \begin{proposition} \label{prop:general_m_and_m+1} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$, or $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Then under Hypothesis~A, both $CS(s)$ and $CS(s')$ consist of $m$ and $m+1$. Moreover $CS(s)$ contains $S_1$ or $S_2$ as a subsequence accordingly as Hypothesis~B or Hypothesis~C is satisfied. \end{proposition} In the following, we prove the proposition only for $CS(s)$. By applying the same argument to the annular diagram reversing the outer and inner boundaries, we see that the assertion also holds for $CS(s')$. \subsection{The case when Hypothesis~B holds} In this subsection, we study the case when Hypothesis~B holds. Suppose $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Then $S_1$ begins and ends with $(m+1, (m_2-1) \langle m \rangle, m+1)$ by Remark~\ref{rem:general_decomposition}(1). Thus Hypothesis~B implies that $CS(\phi(\alpha))=CS(s)$ contains a term $m$ and a term $m+1+\ell$ for some $\ell \ge 0$. Hence $CS(s)$ must consist of $m$ and $m+1$ by \cite[Lemma~3.8]{lee_sakuma_2}. Moreover, since $S_1$ begins and ends with $m+1$, this observation together with Hypothesis~B implies that $CS(s)$ contains $S_1$ as a subsequence. Thus Proposition~\ref{prop:general_m_and_m+1} holds in this case. Suppose $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. The following lemma is needed for the proof of Lemma~\ref{lem:general_prelim_3-1}, by which we prove Proposition~\ref{prop:general_m_and_m+1} for this type of $r$. \begin{lemma} \label{lem:general_prelim_3-1(b)} Let $r=[2, 1, m_3, \dots, m_k]$, where $k \ge 4$. Under Hypothesis~B, the following hold. \begin{enumerate}[\indent \rm (1)] \item $S(z_{i, e}y_{i+1, b}) \neq (4)$. \item $S(w_{i, e}z_iy_{i+1,b}) \neq (4)$ and $S(z_{i-1, e}y_iw_{i, b}) \neq (4)$. \item $S(w_{i, e}z_iy_{i+1,b}) \neq (5)$ and $S(z_{i-1, e}y_iw_{i, b}) \neq (5)$. \item $S(w_{i, e}z_iy_{i+1}w_{i+1, b}) \neq (6)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that $S(z_{1, e}y_{2, b})=(4)$. Then we must have that $z_1=z_{1, e}$ and $y_2=y_{2, b}$. To see this, suppose that $z_1\ne z_{1, e}$. (The other case is treated similarly.) Then, since $S_2$ begins with $(2, m_3 \langle 3 \rangle, 2)$, we see that $S(w_1z_1)$ is of the form $(S_1,2,*)$, where $*$ is nonempty. So, $CS(\phi(\alpha))=CS(s)$ contains a term $2$. This is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}, because $CS(\phi(\alpha))$ also contains a term $4$ by the assumption. Hence, $S(z_1)=S(z_{1, e})=(2)$ and $S(y_2)=S(y_{2, b})=(2)$. Since $S_2$ begins and ends with $(2, m_3 \langle 3 \rangle, 2)$, $S(z_1'y_2')$ contains a subsequence of the form $(\ell_1, 6, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ (see Figure~\ref{fig.Lemma7_3_1}(a)). So if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $2$ and a term $6$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. Suppose $J \subsetneq M$. Then the above fact together with \cite[Lemma~3.1(1)]{lee_sakuma_3} implies that $S(\phi(e_2'e_3'))$ contains the subsequence $(\ell_1, 6, \ell_2)$ (see Figure~\ref{fig.Lemma7_3_1}(b)). Thus, a term $6$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. \begin{figure}[h] \includegraphics{Figure_Lemma7_3_1.eps} \caption{ Lemma~\ref{lem:general_prelim_3-1(b)}(1) where $S(z_{1,e}y_{2,b})=(2+2)$} \label{fig.Lemma7_3_1} \end{figure} (2) Suppose on the contrary that $S(w_{1, e}z_1y_{2,b})=(4)$. (The other case is treated similarly.) Then $|z_1|=0$ and $|y_{2, b}|=1$, because $|w_{1, e}|=3$. Furthermore $y_2=y_{2, b}$ as in the proof of (1). By using the fact that $S_2$ begins and ends with $(2, m_3 \langle 3 \rangle, 2)$ and \cite[Lemma~3.1(1)]{lee_sakuma_3}, we see that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, 2, 1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ (see Figure~\ref{fig.Lemma7_3_2}(a)). So, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $1$ and a term $3$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then a term $1$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. (3) Suppose on the contrary that $S(w_{1, e}z_1y_{2,b})=(5)$. (The other case is treated similarly.) Then $|z_1|=0$ and $|y_{2, b}|=2$. Furthermore $y_2=y_{2, b}$ as in the proof of (1). So $S(y_2w_2)$ begins with a subsequence $(2, (m_3+1) \langle 3 \rangle)$, which implies that $CS(\phi(\alpha))=CS(s)$ has a term $3$. Since $CS(s)$ has a term $5$ by assumption, this is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. (4) Suppose on the contrary that $S(w_{1, e}z_1y_2w_{2, b})=(6)$. Then $|z_1|=|y_2|=0$ and $S(w_{1, e}w_{2, b})=(6)$. By using the fact that $S_2$ begins and ends with $(2, m_3 \langle 3 \rangle, 2)$ and \cite[Lemma~3.1(1)]{lee_sakuma_3}, we see that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, 4, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$ (see Figure~\ref{fig.Lemma7_3_2}(b)). So, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $2$ and a term $4$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then a term $4$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. \end{proof} \begin{figure}[h] \includegraphics{Figure_Lemma_7_3_2.eps} \caption{ (a) Lemma~\ref{lem:general_prelim_3-1(b)}(2) where $S(w_{1, e}z_1y_{2,b})=(3+0+1)$, and (b) Lemma~\ref{lem:general_prelim_3-1(b)}(4) where $S(w_{1, e}z_1y_2w_{2, b})=(3+0+3)$} \label{fig.Lemma7_3_2} \end{figure} \begin{lemma} \label{lem:general_prelim_3-1(b)_addition} Let $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 3$ and $k \ge 4$. Under Hypothesis~B, the following hold. \begin{enumerate}[\indent \rm (1)] \item $S(z_{i, e}y_{i+1, b}) \neq (m+d)$ for any $d \in {\mathbb Z}_+$ with $2 \le d \le m$. \item $S(w_{i, e}z_iy_{i+1,b}) \neq (m+1+d)$ and $S(z_{i-1, e}y_iw_{i, b}) \neq (m+1+d)$ for any $d \in {\mathbb Z}_+$ with $1 \le d \le m$. \item $S(w_{i, e}z_iy_{i+1}w_{i+1, b}) \neq (2m+2)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m+d)$ with $2 \le d \le m$. As in the proof of Lemma~\ref{lem:general_prelim_3-1(b)}(1), we must have $z_1=z_{1, e}$ and $y_2=y_{2, b}$, for otherwise $CS(\phi(\alpha))=CS(s)$ would contain both a term $m$ and a term $m+d$ contradicting \cite[Lemma~3.8]{lee_sakuma_2}. So if $2 \le d\le m-1$, then the proof is parallel to that of Lemma~\ref{lem:case1-1(b)}(1). Now let $d=m$. Then $|z_1|=|y_2|=m$, and so $CS(\phi(\alpha))=CS(s)$ has a term $2m$. Moreover, $S(y_2w_2)$ begins with $(m, (m_3+1) \langle m+1 \rangle)$, and hence $CS(s)$ has a term $m+1$, as shown in Figure~\ref{fig.lemma-3-2}. But since $m \ge 3$, we have $2m > m+2$. This gives a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. \begin{figure}[h] \includegraphics{figure_general_lemma-3-2.eps} \caption{ Lemma~\ref{lem:general_prelim_3-1(b)_addition}(1) where $S(z_{1,e}y_{2,b})=(m+m)$} \label{fig.lemma-3-2} \end{figure} (2) Suppose on the contrary that $S(w_{1, e}z_1y_{2,b})=(m+1+d)$. (The other case is treated analogously.) Then $|z_1|=0$ and $|y_{2, b}|=d$. Furthermore $y_2=y_{2, b}$ as in the proof of (1). So if $1 \le d\le m-1$, then the proof is parallel to that of Lemma~\ref{lem:case1-1(b)}(2). Now let $d=m$. Then $|y_2|=m$ and $CS(s)$ has a term $2m+1$. Also, as shown in the proof of (1), $CS(s)$ has a term $m+1$. But since $m \ge 3$, we have $2m+1 > m+2$. This gives a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. (3) Suppose on the contrary that $S(w_{1,e}z_1y_2w_{2,b})=(2m+2)$. Then $|z_1|=|y_2|=0$ and $S(w_{1, e}w_{2, b})=(2m+2)$. Assume first that $J=M$. Then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $m+1$ and a term $2m$ as illustrated in Figure~\ref{fig.lemma-3-5}(a). Since $m \ge 3$, this gives a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. Assume next that $J \subsetneq M$. Then by \cite[Lemma~3.1(1)]{lee_sakuma_3}, none of $S(\phi(e_{1}'))$ and $S(\phi(e_{2}'))$ contains a subsequence $S_1$ which begins and ends with $((m_3+1) \langle m+1 \rangle)$. This means that the initial vertex of $e_2'$ lies in the interior of the segment of $\partial D_1^-$ corresponding to $S_1$. Similarly, the terminal vertex of $e_3'$ lies in the interior of the segment of $\partial D_2^-$ corresponding to $S_1$. Hence, we see from Figure~\ref{fig.lemma-3-5}(b) that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, 2m, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. This implies that a term $2m$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, which is a contradiction. \end{proof} \begin{figure}[h] \includegraphics{figure_general_lemma-3-5.eps} \caption{ Lemma~\ref{lem:general_prelim_3-1(b)_addition}(3) where $S(w_{1, e}z_1y_2w_{2, b})=((m+1)+0+0+(m+1))$} \label{fig.lemma-3-5} \end{figure} \begin{lemma} \label{lem:general_prelim_3-1} Let $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Under Hypothesis~B, no term of $CS(s)$ can be of the form $m+1+d$ with $d \in \ZZ_+$. \end{lemma} \begin{proof} Suppose on the contrary that $CS(s)$ contains a term $m+1+d$. Let $v$ be a subword of the cyclic word $(u_s)$ corresponding to a term $m+1+d$. Without loss of generality, we may assume that \begin{enumerate}[\indent \rm (i)] \item $v$ contains $z_{1, e} y_{2, b}$ with $|z_{1, e}|, |y_{2, b}| \neq 0$; \item $v$ contains $w_{1, e}y_{2, b}$ with $|y_{2, b}| \neq 0$; \item $v$ contains $z_{0, e}w_{1, e}$ with $|z_{0, e}| \neq 0$; or \item $v$ contains $w_{1, e}w_{2, b}$ with $|z_1|=|y_2|=0$. \end{enumerate} However, (i) is impossible by Lemma~\ref{lem:general_prelim_3-1(b)}(1) or Lemma~\ref{lem:general_prelim_3-1(b)_addition}(1) accordingly as $m=2$ or $m \ge 3$. Also (ii) and (iii) are impossible by Lemma~\ref{lem:general_prelim_3-1(b)}(2)--(3) or Lemma~\ref{lem:general_prelim_3-1(b)_addition}(2) accordingly as $m=2$ or $m \ge 3$. Finally (iv) is impossible by Lemma~\ref{lem:general_prelim_3-1(b)}(4) or Lemma~\ref{lem:general_prelim_3-1(b)_addition}(3) accordingly as $m=2$ or $m \ge 3$. \end{proof} \begin{corollary} \label{cor:general_prelim_3-1} Let $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Under Hypothesis~B, the conclusion of Proposition~\ref{prop:general_m_and_m+1} holds. \end{corollary} \begin{proof} By \cite[Lemma~3.8]{lee_sakuma_2}, either $CS(s)=\lp \ell, \ell \rp$ or $CS(s)$ consists of $\ell$ and $\ell+1$, for some $\ell \in \ZZ_+$. Since $CS(\phi(\alpha))=CS(s)$ must contain a term of the form $m+1+c$ with $c \in \ZZ_+ \cup \{0 \}$, we have $\ell \ge m+1$ in the first case and $\ell \ge m$ in the second case. First, if $CS(s)=\lp \ell, \ell \rp$, then $\ell=m+1$ by Lemma~\ref{lem:general_prelim_3-1}, namely $CS(s)=\lp m+1, m+1 \rp$. This happens only when $J$ consists of only one $2$-cell with $CS(\phi(\partial D_1^+))=\lp S_1 \rp =\lp m+1, m+1\rp$. Then $CS(\phi(\partial D_1^-))=\lp S_2, S_1, S_2\rp$, and so $S(\phi(e_2'e_1'))$ contains a subsequence of the form $(\ell_1, S_2, S_2, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. So, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains two consecutive $m$'s and two consecutive $m+1$'s, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then a subsequence $(S_2, S_2)$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. Thus $CS(s) \neq \lp \ell, \ell \rp$, and so $CS(s)$ consists of $\ell$ and $\ell+1$. By Lemma~\ref{lem:general_prelim_3-1}, we have $\ell+1 \le m+1$, so that $\ell=m$, as desired. Hence, $CS(s)$ consists of $m$ and $m+1$. As already observed in the beginning of this subsection, this implies that $CS(s)$ contains $S_1$ as a subsequence. \end{proof} \subsection{The case when Hypothesis~C holds} We next assume that Hypothesis~C holds. In this case, $CS(s)$ contains $S_2$ as a subsequence. Suppose $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Then $S_2$ begins and ends with $(m, m_3 \langle m+1 \rangle, m)$ by Remark~\ref{rem:general_decomposition}(2). Hence $CS(\phi(\alpha))=CS(s)$ contains terms $m$ and $m+1$, and therefore $CS(\phi(\alpha))=CS(s)$ consists of $m$ and $m+1$ by \cite[Lemma~3.8]{lee_sakuma_2}. So, Proposition~\ref{prop:general_m_and_m+1} holds in this case. On the other hand, for $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$, we prove the following lemmas, by which we prove Proposition~\ref{prop:general_m_and_m+1} for this type of $r$. \begin{lemma} \label{lem:general_prelim_1-2_easy} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Under Hypothesis~C, $CS(s)$ consists of at least three terms including $m$. \end{lemma} \begin{proof} The assertion immediately follows from the fact that $CS(\phi(\alpha))=CS(s)$ properly contains $S_2$ which begins and ends with $(m_2 \langle m \rangle)$ (see Remark~\ref{rem:general_decomposition}(1)). \end{proof} \begin{lemma} \label{lem:general_prelim_1-2} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Under Hypothesis~C, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item $S(z_{i, e}y_{i+1, b}) \neq (m-1)$. \item $S(z_{i, e}y_{i+1, b}) \neq (m-1, m-1)$. \item $S(z_{i, e}y_{i+1, b}) \neq (m-1, m)$ and $S(z_{i, e}y_{i+1, b}) \neq (m, m-1)$. \end{enumerate} \end{lemma} \begin{proof} (1) Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m-1)$. Then we have $z_1=z_{1, e}$ and $y_2=y_{2, b}$, for otherwise $CS(\phi(\alpha))=CS(s)$ contains both a term $m-1$ and a term $m+1$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. By using \cite[Lemma~3.1(2)]{lee_sakuma_3} as in the proof of Lemma~\ref{lem:case1-2}(1), we see that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m+3, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. So, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $m$ and a term $m+3$, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then a term $m+3$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. (2) Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m-1, m-1)$. Then $CS(\phi(\alpha))=CS(s)$ involves two consecutive $m-1$'s. On the other hand, since $CS(s)$ contains $S_2$, which begins and ends with $(m_2 \langle m \rangle)$, we see that $CS(s)$ also contains two consecutive $m$'s. This is a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. (3) Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m-1, m)$. (The other case is treated similarly.) As in the proof of (1), $|z_{1, e}|=m-1$, $|y_{2, b}|=m$, $z_1=z_{1, e}$ and $y_2=y_{2, b}$. By using \cite[Lemma~3.1(2)]{lee_sakuma_3} as in the proof of Lemma~\ref{lem:case1-2}(2), we see that $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, 2, 1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. Hence, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains both a term $1$ and a term $m+1$. Since $m+1 \ge 3$, we have a contradiction to \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if $J \subsetneq M$, then a term $1$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. \end{proof} \begin{corollary} \label{cor:general_prelim_1-2} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Under Hypothesis~C, the conclusion of Proposition~\ref{prop:general_m_and_m+1} holds. \end{corollary} \begin{proof} By Lemma~\ref{lem:general_prelim_1-2_easy}, $CS(s)$ consists of at least three terms including $m$. Also, Lemma~\ref{lem:general_prelim_1-2} shows that no term of $CS(s)$ can be $m-1$. Hence by \cite[Lemma~3.8]{lee_sakuma_2}, $CS(s)$ must consist of $m$ and $m+1$. Moreover, we have already observed that $CS(s)$ contains $S_2$ as a subsequence. \end{proof} Thus we have proved Proposition~\ref{prop:general_m_and_m+1}. \section{Transformation of diagrams for the general cases} \label{sec:transformation} We first introduce a concept for a vertex of $M$ to be converging, diverging or mixing. To this end, we subdivide the edges of $M$ so that the label of any oriented edge in the subdivision has length $1$. We call each of the edges in the subdivision a {\it unit segment} in order to distinguish them from the edges in the original $M$. \begin{definition} \label{def:vertex_type} {\rm (1) A vertex $x$ in $M$ is said to be {\it converging} (resp., {\it diverging}) if the set of labels of incoming unit segments of $x$ is $\{a, b\}$ (resp., $\{a^{-1}, b^{-1}\}$). See Figure~\ref{fig.converging} and its caption for description. (2) A vertex $x$ in $M$ is said to be {\it mixing} if the set of labels of incoming unit segments of $x$ is $\{a, a^{-1},b, b^{-1}\}$. See Figure~\ref{fig.impossible_tsequence} and its caption for description. } \end{definition} \begin{figure}[h] \includegraphics{figure_converging_2.eps} \caption{ Orient each of the unit segment so that the associated label is equal to $a$ or $b$. Then a vertex $x$ is (a) converging (resp., (b) diverging) if all unit segments incident on $x$ are oriented so that they are converging into $x$ (resp., diverging from $x$).} \label{fig.converging} \end{figure} \begin{figure}[h] \includegraphics{figure_impossible_tsequence.eps} \caption{ A vertex $x$ is mixing if it looks like as in the above when we orient the segments as in Convention~\ref{con:figure}, namely, the change of directions of consecutive arrowheads represents the change from positive (negative, resp.) words to negative (positive, resp.) words.} \label{fig.impossible_tsequence} \end{figure} As we declared in the previous section, we assume that $r$ is general. The main purpose of this section is to show that, under Hypothesis~A, we may modify $M$ so that every vertex $x$ in $M$ with $\deg_M (x)=4$ is either converging or diverging (see Corollary~\ref{cor:general_critical(aa)}). \subsection{The case when vertices lie in the outer boundary layer} We first treat a vertex $x$ in the outer boundary layer $J$ with $\deg_J (x)=4$. To do this, we need several lemmas. \begin{lemma} \label{lem:converging} Assume that $r$ is general. Under Hypothesis~B, suppose that the vertex between $D_i$ and $D_{i+1}$ is either converging or diverging. Then none of the following occurs. \begin{enumerate}[\indent \rm (1)] \item $S(\phi(\partial D_i^+))$ ends with $S_1$. \item $S(\phi(\partial D_{i+1}^+))$ begins with $S_1$. \item $S(\phi(\partial D_i^-))$ ends with $S_1$. \item $S(\phi(\partial D_{i+1}^-))$ begins with $S_1$. \end{enumerate} \end{lemma} \begin{proof} We may assume that $i=1$ and that the vertex is diverging. Suppose on the contrary that (1) occurs, namely suppose that $S(\phi(\partial D_1^+))$ ends with $S_1$. Then $S(\phi(\partial D_1^-))$ ends with $(S_1,S_2)$. Since the vertex is diverging, $S(\phi(\partial D_1^-)\phi(\partial D_2^-))$ contains a subsequence $(S_1,S_2,d)$ for some $d\in\ZZ_+$. Suppose $J=M$. Then $CS(\phi(\delta^{-1}))=CS(s')$ contains $S_2$. (See Figure~\ref{fig.S_1_occurs}(a), keeping in mind Convention~\ref{con:figure}.) Moreover, the subsequence $S_1$ of $S(\phi(\partial D_1^-))$ also forms a subsequence of $CS(\phi(\delta^{-1}))=CS(s')$, because $CS(s')$ consists of $m$ and $m+1$ (Proposition~\ref{prop:general_m_and_m+1}) and $S_1$ begins and ends with $m+1$. Thus $CS(s')$ contains both $S_1$ and $S_2$, yielding that $s' \notin I_1(r) \cup I_2(r)$ by \cite[Proposition~3.19(1)]{lee_sakuma_2}, a contradiction. Suppose $J \subsetneq M$. Then we see in the following that the two $2$-cells $D_1$ and $D_1'$ in Figure~\ref{fig.S_1_occurs}(b) form a reducible pair, contradicting that $M$ is a reduced diagram. To see this, let $\gamma$ (resp., $\gamma'$) be the boundary cycle of $D_1$ (resp., $D_1'$) which goes around the boundary in clockwise (resp., counter-clockwise) direction starting from the vertex between $D_1$ and $D_2$. Let $\gamma_0$ be the common initial segment of $\gamma$ and $\gamma'$ such that $S(\phi(\gamma_0))=S_2$. Then we see by using \cite[Lemma~3.1(1)]{lee_sakuma_3} that the terminal point of $\gamma_0$ ($=$ the initial point of $e_2'$) is contained in the interior of the segment of $\partial D_1^-$ corresponding to the segment $S_1$ (see Figure~\ref{fig.S_1_occurs}(b)). Thus, in both reduced words $\phi(\gamma)$ and $\phi(\gamma')$, there are ``sign changes'' just before and after $\phi(\gamma_0)$. Hence $S(\phi(\gamma_0))=S_2$ is a subsequence of both $CS(\phi(\gamma))=CS(r)$ and $CS(\phi(\gamma'))=CS(r)$. By \cite[Proposition~3.12]{lee_sakuma_2}(2) and the fact that $\phi(\gamma)$ and $\phi(\gamma')$ shares $\phi(\gamma_0)$ as a common initial word, we must have $\phi(\gamma) \equiv \phi(\gamma')$; so $D_1$ and $D_1'$ form a reducible pair. Thus we have proved that (1) does not occur. Similarly, we can prove that (2) does not occur. \begin{figure}[h] \includegraphics{figure_S_1_occurs.eps} \caption{ Lemma~\ref{lem:converging}(1) where $S(\phi(\partial D_1^+))$ ends with $S_1$} \label{fig.S_1_occurs} \end{figure} Suppose on the contrary that (3) occurs, i.e., $S(\phi(\partial D_1^-))$ ends with $S_1$. Then $S(\phi(\partial D_1^+))$ ends with $(S_1,S_2)$. Since the vertex is diverging, this subsequence $S_2$ of $S(\phi(\partial D_1^+))$ is also a subsequence of $CS(\phi(\alpha))=CS(s)$. Moreover, we can see as in the proof of (1) that the subsequence $S_1$ of $S(\phi(\partial D_1^+))$ forms a subsequence of $CS(\phi(\alpha))$. Thus $CS(\phi(\alpha))$ contains both $S_1$ and $S_2$ as subsequences and so $s \notin I_1(r) \cup I_2(r)$ by \cite[Proposition~3.19(1)]{lee_sakuma_2}, a contradiction. The assertion (4) is proved similarly. \end{proof} \begin{lemma} \label{lem:general_critical(a)} Assume that $r$ is general. Under Hypothesis~B, we may assume that the following hold for every face $D_i$ of $J$. \begin{enumerate}[\indent \rm (1)] \item $S(\phi(\partial D_i^+))$ contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \item $S(\phi(\partial D_i^-))$ contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \end{enumerate} To be precise, we can modify the reduced annular diagram $M$ into a reduced annular diagram $M'$ keeping the outer and inner boundary labels unchanged so that every $2$-cell of the outer boundary layer of $M'$ satisfies the above conditions. \end{lemma} \begin{proof} Suppose that the assertion does not hold. Then one of the four (prohibited) conditions in Lemma~\ref{lem:converging} holds. In particular, the vertex between $D_i$ and $D_{i+1}$ is not converging nor diverging. Suppose that condition (1) in Lemma~\ref{lem:converging} occurs. Then we may assume that $S(\phi(\partial D_1^+))$ ends with $S_1$ and so $|z_1|=0$. Then $S(z_1')=S_2$, and so $S(z_{1, e}')=(m)$. Hence the sequence $S(z_{1, e}'y_2'w_2')$ begins with a subsequence of the form either $(m, d)$ or $(m+d)$, where $d \in \ZZ_+$. Suppose that $S(z_{1, e}'y_2'w_2')$ begins with a subsequence of the form $(m, d)$. Then the vertex between $D_1$ and $D_2$ is either converging or diverging, a contradiction. (In fact, since $CS(s)=CS(\phi(\alpha))$ consists of $m$ and $m+1$ by Proposition~\ref{prop:general_m_and_m+1} and since $S(\phi(\partial D_1^+))$ ends with $m+1$, there is a ``sign change'' between $\phi(\partial D_1^+)$ and $\phi(\partial D_2^+)$. Thus $J$ is locally as illustrated in Figure~\ref{fig.S_1_occurs}(a) up to simultaneous change of the edge orientations.) So $S(z_{1, e}'y_2'w_2')$ must begin with a subsequence of the form $(m+d)$. Since $S(\phi(\partial D_1^+))$ ends with a term $m+1$ and since $CS(\phi(\alpha))=CS(s)$ consists of $m$ and $m+1$, the only possibility is that $d=1$ and $S(w_{1, e} y_{2, b})=(m+1, m)$. Then, as illustrated in Figure~\ref{fig.S_1_occurs(b)}, we may transform $M$ so that $S(\phi(\partial D_1^+))$ ends with $(S_1, m)$. To be precise, we cut $J$ at the black vertex in the left figure in Figure~\ref{fig.S_1_occurs(b)} and then identify the two white vertices. The resulting diagram is illustrated in the right figure in Figure~\ref{fig.S_1_occurs(b)}, where the black vertex is the image of the white vertices. It should be noted that the boundary labels of $J$ are unchanged by this operation and so we can glue $M-J$ to this new $J$. This modification does not change the boundary labels of $M$ and the new vertex of $J$ is converging or diverging. Hence we see by Lemma~\ref{lem:converging} that none of conditions (1)--(4) occurs at the vertex between $D_1$ and $D_2$ in this new diagram. Thus we have shown that condition (1) in Lemma~\ref{lem:converging} may be assumed not to occur. Similarly, we can show that condition (2) in Lemma~\ref{lem:converging} may be assumed not to occur. \begin{figure}[h] \includegraphics{figure_S_1_occurs_b.eps} \caption{ Lemma~\ref{lem:general_critical(a)} where $S(z_{1, e}'y_2'w_2')$ begins with $(m+d)$} \label{fig.S_1_occurs(b)} \end{figure} Suppose that condition (3) in Lemma~\ref{lem:converging} occurs. Then we may assume $S(\phi(\partial D_1^-))$ ends with $S_1$ and so $|z_1'|=0$. Then $S(z_1)=S_2$, and so $S(z_{1, e})=(m)$. Hence the sequence $S(z_{1, e}y_2w_2)$ begins with a subsequence of the form either $(m, d)$ or $(m+d)$, where $d \in \ZZ_+$. If $S(z_{1, e}y_2w_2)$ begins with a subsequence of the form $(m, d)$, then we can see as in the previous case that the vertex between $D_1$ and $D_2$ is converging or diverging, a contradiction. So $S(z_{1, e}y_2w_2)$ must begin with a subsequence of the form $(m+d)$. Since $S(\phi(\partial D_1^+))$ ends with a term $m$ and since $CS(\phi(\alpha))=CS(s)$ consists of $m$ and $m+1$, the only possibility is that $d=1$ and $S(w_{1, e}' y_{2, b}')=(m+1, m)$. Then, as illustrated in Figure~\ref{fig.S_1_occurs(c)}, we may transform $M$ so that $S(\phi(\partial D_1^-))$ ends with $(S_1, m)$. Since the new vertex is either converging or diverging, we see by Lemma~\ref{lem:converging} that none of conditions (1)--(4) occurs at the common vertex of $D_1$ and $D_2$ in this new diagram. Thus we have shown that condition (3) in Lemma~\ref{lem:converging} may be assumed not to occur. Similarly, we can show that condition (4) in Lemma~\ref{lem:converging} may be assumed not to occur. This completes the proof of Lemma~\ref{lem:general_critical(a)}. \end{proof} \begin{figure}[h] \includegraphics{figure_S_1_occurs_c.eps} \caption{ Lemma~\ref{lem:general_critical(a)} where $S(z_{1, e}y_2w_2)$ begins with $(m+d)$} \label{fig.S_1_occurs(c)} \end{figure} In the remainder of this section, when we assume Hypothesis~B, we always assume that the two conditions in Lemma~\ref{lem:general_critical(a)} hold, namely the words $y_i$, $z_i$, $y_i'$ and $z_i'$ which appear in the expressions of $\phi(\partial D_i^+)$ and $\phi(\partial D_i^-)$ are nonempty. Note that when we assume Hypothesis~C, the same conditions always hold by the hypothesis. \begin{lemma} \label{lem:general_critical(b)} Assume that $r$ is general. Under Hypothesis~B or Hypothesis~C, the following hold for every $i$. \begin{enumerate}[\indent \rm (1)] \item If $S(z_{i, e} z_{i, e}'^{-1})=S(y_{i+1, b}'^{-1}y_{i+1, b})=(m)$, then $S(z_{i, e}y_{i+1, b}) \neq (m+1)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=S(y_{i+1, b}'^{-1}y_{i+1, b})=(m+1)$, then $S(z_{i, e}y_{i+1, b}) \neq (m)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=(m)$ and $S(y_{i+1, b}'^{-1}y_{i+1, b})=(m, m)$, then $S(z_{i, e}y_{i+1, b}) \neq (m+1)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=(m)$ and $S(y_{i+1, b}'^{-1}y_{i+1, b})=(m+1, m)$, then $S(z_{i, e}y_{i+1, b}) \neq (m+1)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=S(y_{i+1, b}'^{-1}y_{i+1, b})=(m+1)$, then $S(z_{i, e}y_{i+1, b}) \neq (m, m)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=(m, m+1)$ and $S(y_{i+1, b}'^{-1}y_{i+1, b})=(m+1)$, then $S(z_{i, e}y_{i+1, b}) \neq (m, m)$. \item If $S(z_{i, e} z_{i, e}'^{-1})=(m+1, m+1)$ and $S(y_{i+1, b}'^{-1}y_{i+1, b})=(m+1)$, then $S(z_{i, e}y_{i+1, b}) \neq (m+1, m)$. \end{enumerate} \end{lemma} \begin{proof} (1) Let $S(z_{1, e} z_{1, e}'^{-1})=S(y_{2, b}'^{-1}y_{2, b})=(m)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m+1)$ (see Figure~\ref{fig.general_critical_b1}(a)). Then $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m-1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. Here, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains a term $m-1$, a contradiction to Proposition~\ref{prop:general_m_and_m+1}. On the other hand, if $J \subsetneq M$, then a term $m-1$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction (cf. Proof of Lemma~\ref{lem:case1-1(b)}(1) for the case $J \subsetneq M$). (2) Let $S(z_{1, e} z_{1, e}'^{-1})=S(y_{2, b}'^{-1}y_{2, b})=(m+1)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m)$ (see Figure~\ref{fig.general_critical_b1}(b)). Then $S(\phi(e_2'e_3'))$ contains a term $m+2$. Here, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains a term $m+2$, a contradiction to Proposition~\ref{prop:general_m_and_m+1}. On the other hand, if $J \subsetneq M$, then a term of the form $m+2+c$ with $c \in \ZZ_+ \cup \{0\}$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. \begin{figure}[h] \includegraphics{figure_general_critical_b1.eps} \caption{ (a) Lemma~\ref{lem:general_critical(b)}(1), and (b) Lemma~\ref{lem:general_critical(b)}(2)} \label{fig.general_critical_b1} \end{figure} (3) Let $S(z_{1, e} z_{1, e}'^{-1})=(m)$ and $S(y_{2, b}'^{-1}y_{2, b})=(m, m)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m+1)$ (see Figure~\ref{fig.general_critical_b3}(a)). Then $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m-1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. Here, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains a term $m-1$, a contradiction to Proposition~\ref{prop:general_m_and_m+1}. On the other hand, if $J \subsetneq M$, then a term $m-1$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. (4) Let $S(z_{1, e} z_{1, e}'^{-1})=(m)$ and $S(y_{2, b}'^{-1}y_{2, b})=(m+1, m)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m+1)$ (see Figure~\ref{fig.general_critical_b3}(b)). Then $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, m-1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. So, arguing as in the proof of (3), we reach a contradiction. \begin{figure}[h] \includegraphics{figure_general_critical_b3.eps} \caption{ (a) Lemma~\ref{lem:general_critical(b)}(3), and (b) Lemma~\ref{lem:general_critical(b)}(4)} \label{fig.general_critical_b3} \end{figure} (5) Let $S(z_{1, e} z_{1, e}'^{-1})=S(y_{2, b}'^{-1}y_{2, b})=(m+1)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m, m)$ (see Figure~\ref{fig.general_critical_b5}(a)). Here, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains a term $1$, a contradiction to Proposition~\ref{prop:general_m_and_m+1}. On the other hand, if $J \subsetneq M$, then $S(\phi(e_2'e_3'))$ contains a subsequence of the form $(\ell_1, 1, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$, for otherwise $S(\phi(\partial {D_1'}^+))=S(\phi(e_2'e_3'))=(1, 1)$ which contains neither $S_1$ nor $(\ell, S_2, \ell')$ with $\ell, \ell' \in \ZZ_+$, contradicting \cite[Lemma~3.2]{lee_sakuma_3}. It the follows that a term $1$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. (6) Let $S(z_{1, e} z_{1, e}'^{-1})=(m, m+1)$ and $S(y_{2, b}'^{-1}y_{2, b})=(m+1)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m, m)$ (see Figure~\ref{fig.general_critical_b5}(b)). Then $S(\phi(e_2'e_3'))$ contains a term $m+2$. Here, if $J=M$, then $CS(\phi(\delta^{-1}))=CS(s')$ contains a term $m+2$, a contradiction to Proposition~\ref{prop:general_m_and_m+1}. On the other hand, if $J \subsetneq M$, then a term of the form $m+2+c$ with $c \in \ZZ_+ \cup \{0\}$ occurs in $CS(\phi(\partial D_1'))=CS(r)$, a contradiction. \begin{figure}[h] \includegraphics{figure_general_critical_b5.eps} \caption{ (a) Lemma~\ref{lem:general_critical(b)}(5), and (b) Lemma~\ref{lem:general_critical(b)}(6)} \label{fig.general_critical_b5} \end{figure} (7) Let $S(z_{1, e} z_{1, e}'^{-1})=(m+1, m+1)$ and $S(y_{2, b}'^{-1}y_{2, b})=(m+1)$. Suppose on the contrary that $S(z_{1, e}y_{2, b})=(m+1, m)$. Then $S(\phi(e_2'e_3'))$ contains a term $m+2$. So, arguing as in the proof of (6), we reach a contradiction. \end{proof} We are now ready to prove the following. \begin{proposition} \label{prop:general_critical} Assume that $r$ is general. Under Hypothesis~B or Hypothesis~C, we may assume that the following hold. \begin{enumerate}[\indent \rm (1)] \item For every face $D_i$ of $J$, the following hold. \begin{enumerate}[\rm (a)] \item If $S(\phi(\partial D_i^+))$ contains $S_1$, then it contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \item If $S(\phi(\partial D_i^-))$ contains $S_1$, then it contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \end{enumerate} \item Every vertex $x$ in $J$ with $\deg_{J}(x)=4$ is either converging or diverging. \end{enumerate} \end{proposition} \begin{proof} By Lemma~\ref{lem:general_critical(a)}, we may assume that (1) is satisfied. We prove that we can modify the annular diagram $M$ so that it satisfies (2). Then the resulting annular diagram satisfies both (1) and (2), because Lemma~\ref{lem:converging} guarantees that if $M$ satisfies (2) then it also satisfies (1). Since (1) is satisfied, the subwords $y_i$ and $z_i$ of $\phi(\partial D_i^+)$ and the subwords $y_i'$ and $z_i'$ of $\phi(\partial D_i^-)$ are nonempty. Suppose on the contrary that there is a vertex $x \in J$ with $\deg_J(x)=4$ such that $x$ is neither converging nor diverging. We may assume $x$ is the vertex between $D_1$ and $D_2$. Then $x$ has one of the five types as depicted in Figure~\ref{fig.vertex_type}, where $c_i$ and $d_i$ ($i=1,2$) are positive integers, up to simultaneous reversal of the edge orientations and up to the reflection in the vertical edge passing through the vertex $x$. To see this, let $L$ be the set of labels of incoming unit segments of $x$, and orient each of the unit segment so that the associated label is equal to $a$ or $b$ as in Figure~\ref{fig.converging}. If $L=\{a^{\pm 1}, b^{\pm 1}\}$, then we obtain the situation (a) or (b) in Figure~\ref{fig.vertex_type}. If $L$ consists of three elements, then we may assume that $a$ and $a^{-1}$, respectively, appear as the label of the upper left and lower right incoming unit segments and that $b$ or $b^{-1}$ does not belong to $L$. Then we obtain the situation (c) or (d) in Figure~\ref{fig.vertex_type}. If $L$ consists of two elements, then we may assume both the upper left and lower right incoming unit segments have label $a$, and both the upper left and lower right incoming unit segments have label $b^{-1}$, because $x$ is not converging nor diverging. In this case, we have the situation (e) in Figure~\ref{fig.vertex_type}. \begin{figure}[h] \includegraphics{figure_vertex_type_2.eps} \caption{ The five possible types of a vertex $x \in J$ with $\deg_J (x)=4$ such that $x$ is neither converging nor diverging} \label{fig.vertex_type} \end{figure} Assume that $x$ is as depicted in Figure~\ref{fig.vertex_type}(a). Then, for each $i=1,2$, $c_i$ is a term of $CS(\phi(\partial D_i))=CS(r)$ and so is equal to $m$ or $m+1$. Hence the term, $c_1+c_2$, of $CS(\phi(\alpha))=CS(s)$ is at least $2m$. This is a contradiction to Proposition~\ref{prop:general_m_and_m+1}. Assume that $x$ is as depicted in Figure~\ref{fig.vertex_type}(b). Then $(c_1,c_2)$ is a subsequence of $CS(\phi(\alpha))=CS(s)$. Since $CS(\phi(\alpha))=CS(s)$ consists of $m$ and $m+1$ by Proposition~\ref{prop:general_m_and_m+1}, the only possibility is that $c_1=c_2=m$ and $d_1=d_2=1$. So, we must have $S(z_{1, e} z_{1, e}'^{-1})=S(y_{2, b}'^{-1}y_{2, b})=(m+1)$ and $S(z_{1, e}y_{2, b}) = (m, m)$. But this is impossible by Lemma~\ref{lem:general_critical(b)}(5). Assume that $x$ is as depicted in Figure~\ref{fig.vertex_type}(c). Then $(c_1,c_2)$ is a subsequence of $CS(\phi(\alpha))=CS(s)$ and hence each of $c_1$ and $c_2$ is either $m$ or $m+1$. Moreover, $c_2+d_2$ is a term of $CS(r)$ and hence it is either $m$ or $m+1$. So, we have the following four possibilities: \begin{enumerate}[\indent \rm (i)] \item $c_1=m$, $c_2=m$, $d_1=m$, $d_2=1$; \item $c_1=m$, $c_2=m$, $d_1=m+1$, $d_2=1$; \item $c_1=m+1$, $c_2=m$, $d_1=m$, $d_2=1$; \item $c_1=m+1$, $c_2=m$, $d_1=m+1$, $d_2=1$. \end{enumerate} However, (ii) and (iv) are impossible by Lemma~\ref{lem:general_critical(b)}(6) and Lemma~\ref{lem:general_critical(b)}(7), respectively. If (i) or (iii) happens, then $c_2=d_1$. So we can transform $M$ so that $x$ is diverging as in Figure~\ref{fig.transformation_2} (cf. the argument in the proof of Lemma~\ref{lem:general_critical(a)} appealing to Figure~\ref{fig.S_1_occurs(b)}). \begin{figure}[h] \includegraphics{figure_transformation_2.eps} \caption{ The transformation of Figure~\ref{fig.vertex_type}(c) when $c_2=d_1$ so that $x$ is diverging} \label{fig.transformation_2} \end{figure} Assume that $x$ is as depicted in Figure~\ref{fig.vertex_type}(d). Then $c_1+c_2$ is a term of $CS(\phi(\alpha))=CS(s)$ and hence it is either $m$ or $m+1$. Moreover, $c_1+d_1$ is a term of $CS(r)$ and hence it is either $m$ or $m+1$. So, we have the following four possibilities: \begin{enumerate}[\indent \rm (i)] \item $c_1=1$, $c_2=m$, $d_1=m-1$, $d_2=m$; \item $c_1=1$, $c_2=m$, $d_1=m-1$, $d_2=m+1$; \item $c_1=1$, $c_2=m$, $d_1=m$, $d_2=m$; \item $c_1=1$, $c_2=m$, $d_1=m$, $d_2=m+1$. \end{enumerate} However, (i) and (ii) are impossible by Lemma~\ref{lem:general_critical(b)}(3) and Lemma~\ref{lem:general_critical(b)}(4), respectively. If (iii) or (iv) happens, then $c_2=d_1$. So we can transform $M$ so that $x$ is converging as in Figure~\ref{fig.transformation_1}. \begin{figure}[h] \includegraphics{figure_transformation_1.eps} \caption{ The transformation of Figure~\ref{fig.vertex_type}(d) when $c_2=d_1$ so that $x$ is converging} \label{fig.transformation_1} \end{figure} Assume that $x$ is as depicted in Figure~\ref{fig.vertex_type}(e). Then $c_1+c_2$ is a term of $CS(\phi(\alpha))=CS(s)$ and hence it is either $m$ or $m+1$. Moreover, for each $i=1,2$, $c_i+d_i$ is a term of $CS(r)$ and hence it is either $m$ or $m+1$. So, we have the following eight possibilities: \begin{enumerate}[\indent \rm (i)] \item $c_1+c_2=m$, $c_1+d_1=m$, $c_2+d_2=m$; \item $c_1+c_2=m$, $c_1+d_1=m$, $c_2+d_2=m+1$; \item $c_1+c_2=m$, $c_1+d_1=m+1$, $c_2+d_2=m$; \item $c_1+c_2=m$, $c_1+d_1=m+1$, $c_2+d_2=m+1$; \item $c_1+c_2=m+1$, $c_1+d_1=m$, $c_2+d_2=m$; \item $c_1+c_2=m+1$, $c_1+d_1=m$, $c_2+d_2=m+1$; \item $c_1+c_2=m+1$, $c_1+d_1=m+1$, $c_2+d_2=m$; \item $c_1+c_2=m+1$, $c_1+d_1=m+1$, $c_2+d_2=m+1$. \end{enumerate} However, (iv) and (v) are impossible by Lemma~\ref{lem:general_critical(b)}(2) and Lemma~\ref{lem:general_critical(b)}(1), respectively. If (i), (ii), (vii) or (viii) happens, then $c_2=d_1$. Thus, as illustrated in Figure~\ref{fig.transformation_1}, we may transform $M$ so that $x$ is converging (cf. the argument in the proof of Lemma~\ref{lem:general_critical(a)} appealing to Figure~\ref{fig.S_1_occurs(b)}). If (iii) or (vi) happens, then $c_1=d_2$. So we can transform $M$ so that $x$ is diverging as in Figure~\ref{fig.transformation_2}. \end{proof} \subsection{The case when vertices lie in an arbitrary layer} We next treat a general vertex $x$ in $M$ with $\deg_M (x)=4$ by using induction on the number of layers of $M$. Note that the base step (i.e., the case when $M=J$) was already proved in Proposition~\ref{prop:general_critical}(2). We need several lemmas and new notations. \begin{notation} \label{not:layers} {\rm Under Hypothesis~A, suppose that the number of layers of $M$ is $p+1$ with $p \ge 0$. (Recall the characterization of $M$ in \cite[Theorem~4.11]{lee_sakuma_2}.) For each $j=0, 1, \dots, p$, we define $J_j$ as follows: $J_0=J$, and $J_{j}$ is the outer boundary layer of $M-(J_0 \cup \cdots \cup J_{j-1})$ for $j \ge 1$. Then $M=J_0 \cup J_1 \cup \cdots \cup J_p$. We also define $\alpha_j$ and $\delta_j$ to be, respectively, outer and inner boundary cycles of $J_j$ starting from a vertex lying in both the outer and inner boundaries of $J_j$. Let $\alpha_j=e_{j,1}, e_{j,2}, \dots, e_{j,2t}$ and $\delta_j^{-1}=e_{j,1}', e_{j,2}', \dots, e_{j,2t}'$ be the decompositions into oriented edges in $\partial J_j$. Then clearly for each $i=1, \dots, t$, there is a face $D_{j,i}$ of $J_j$ such that $e_{j,2i-1}, e_{j,2i}, e_{j,2i}'^{-1}, e_{j,2i-1}'^{-1}$ are consecutive edges in a boundary cycle of $D_{j,i}$. We denote the path $e_{j,2i-1}, e_{j,2i}$ by $\partial D_{j,i}^+$ and the path $e_{j,2i-1}', e_{j,2i}'$ by $\partial D_{j,i}^-$. In particular, if $J_0 \cup \cdots \cup J_j \subsetneq M$, then we may assume that $e_{j,2i}'$ and $e_{j,2i+1}'$ are two consecutive edges in $\partial D_{j+1,i} \cap \delta_j^{-1}$. } \end{notation} We can easily see that \cite[Lemmas~3.1 and 3.2]{lee_sakuma_3} holds not only for the faces of $J_0=J$ but also for every face of $M$. \begin{lemma} \label{lem:vertex_position:intermediate_layer} Under Hypothesis~A and Notation~\ref{not:layers}, both of the following hold for every face $D_{j,i}$ of $M$. \begin{enumerate}[\indent \rm (1)] \item None of $S(\phi(e_{j,2i-1}))$, $S(\phi(e_{j,2i}))$, $S(\phi(e_{j,2i}'))$ and $S(\phi(e_{j,2i-1}'))$ contains $S_1$ as a subsequence. \item None of $S(\phi(e_{j,2i-1}))$, $S(\phi(e_{j,2i}))$, $S(\phi(e_{j,2i}'))$ and $S(\phi(e_{j,2i-1}'))$ contains a subsequence of the form $(\ell, S_2, \ell')$, where $\ell, \ell' \in \ZZ_+$. \end{enumerate} \end{lemma} \begin{lemma} \label{lem:two_cases:intermediate_layer} Under Hypothesis~A and Notation~\ref{not:layers}, only one of the following holds for each face $D_{j,i}$ of $M$. \begin{enumerate}[\indent \rm (1)] \item Both $S(\phi(\partial D_{j,i}^+))$ and $S(\phi(\partial D_{j,i}^-))$ contain $S_1$ as its subsequence. \item Both $S(\phi(\partial D_{j,i}^+))$ and $S(\phi(\partial D_{j,i}^-))$ contain a subsequence of the form $(\ell, S_2, \ell')$, where $\ell, \ell' \in \ZZ_+$. \end{enumerate} \end{lemma} In fact, these lemmas are proved by using \cite[Corollary~3.25]{lee_sakuma_2} and the following facts: \begin{enumerate}[\indent \rm (i)] \item The words $\phi(e_{j,i})$ and $\phi(e_{j,i}')$ are pieces. (If $1\le j\le p-1$, this follows from the assumption that $M$ is a reduced annular diagram. If $j=0$ or $p$, this follows from \cite[Convention~4.3]{lee_sakuma_2}.) \item The words $\phi(\partial D_{j,i}^+)$ and $\phi(\partial D_{j,i}^-)$ are not pieces. (Otherwise, the cyclic word $(\phi(\partial D_{j,i}))=(u_r^{\pm1})$ becomes a product of three pieces, a contradiction to \cite[Proposition~3.22]{lee_sakuma_2}.) \end{enumerate} \begin{notation} \label{not:layers2} {\rm Under Hypothesis~A and Notation~\ref{not:layers}, Lemma~\ref{lem:two_cases:intermediate_layer} implies that we may decompose the word $\phi(\alpha_j)$ into \[ \phi(\alpha_j) \equiv y_{j,1} w_{j,1} z_{j,1} y_{j,2} w_{j,2} z_{j,2} \cdots y_{j,t} w_{j,t} z_{j,t}, \] where $\phi(\partial D_{j,i}^+) \equiv \phi(e_{j,2i-1}e_{j,2i}) \equiv y_{j,i}w_{j,i}z_{j,i}$, and where either $S(w_{j,i})=S_1$ or both $S(w_{j,i})=S_2$ and $y_{j,i}, z_{j,i}$ are nonempty words. We also have the decomposition of the word $\phi(\delta_j^{-1})$ as follows: \[ \phi(\delta_j^{-1}) \equiv y_{j,1}' w_{j,1}' z_{j,1}' y_{j,2}' w_{j,2}' z_{j,2}' \cdots y_{j,t}' w_{j,t}' z_{j,t}', \] where $\phi(\partial D_{j,i}^-) \equiv \phi(e_{j,2i-1}'e_{j,2i}') \equiv y_{j,i}'w_{j,i}'z_{j,i}'$, and where either $S(w_{j,i}')=S_1$ or both $S(w_{j,i}')=S_2$ and $y_{j,i}', z_{j,i}'$ are nonempty words. Here, the indices for the $2$-cells are considered modulo $t$, and the indices for the edges are considered modulo $2t$. } \end{notation} \begin{lemma} \label{lem:converging2} Assume that $r$ is general. Under Hypothesis~A and Notation~\ref{not:layers}, suppose that the vertex between $D_{j,i}$ and $D_{j,i+1}$ is either converging or diverging. Then none of the following occurs. \begin{enumerate}[\indent \rm (1)] \item $S(\phi(\partial D_{j,i}^+))$ ends with $S_1$. \item $S(\phi(\partial D_{j,i+1}^+))$ begins with $S_1$. \item $S(\phi(\partial D_{j,i}^-))$ ends with $S_1$. \item $S(\phi(\partial D_{j,i+1}^-))$ begins with $S_1$. \end{enumerate} \end{lemma} \begin{proof} We may assume that $i=1$ and that the vertex is diverging. If $j=0$, then the assertion is nothing other than Lemma~\ref{lem:converging}. If $j=p$, then the assertion is proved by applying the proof of Lemma~\ref{lem:converging} to the inner boundary layer of $M$. So, we may assume $1\le j\le p-1$. Suppose on the contrary that (1) occurs, namely suppose that $S(\phi(\partial D_{j,1}^+))$ ends with $S_1$. Then $S(\phi(\partial D_{j,1}^-))$ ends with $(S_1,S_2)$. Since the vertex is diverging, $S(\phi(\partial D_{j,1}^-)\phi(\partial D_{j,2}^-))$ contains a subsequence $(S_1,S_2,d)$ for some $d\in\ZZ_+$. Thus we obtain a situation as illustrated in Figure~\ref{fig.S_1_occurs}(b), where $D_{j,1}$, $D_{j,2}$ and $D_{j+1,1}$, respectively, correspond to $D_1$, $D_2$ and $D_1'$ in the figure. Then by the argument in the proof of Lemma~\ref{lem:converging} for the case where (1) occurs and $J \subsetneq M$, we see that the two $2$-cells $D_{j,1}$ and $D_{j+1,1}$ form a reducible pair, a contradiction. (Here we use Lemma~\ref{lem:vertex_position:intermediate_layer}(1) instead of \cite[Lemma~3.1(1)]{lee_sakuma_3}.) Hence (1) cannot occur. By a similar argument, we can see that (2), (3) and (4) cannot occur. \end{proof} Now we are ready to prove the following generalization of Proposition~\ref{prop:general_critical}. \begin{proposition} \label{prop:general_critical(aa)} Assume that $r$ is general. Under Hypothesis~A and Notation~\ref{not:layers}, we may assume that the following hold for every $j$. \begin{enumerate}[\indent \rm (1)] \item For every face $D_{j,i}$ of $J_j$, the following hold. \begin{enumerate}[\rm (a)] \item If $S(\phi(\partial D_{j,i}^+))$ contains $S_1$, then it contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \item If $S(\phi(\partial D_{j,i}^-))$ contains $S_1$, then it contains a subsequence of the form $(\ell, S_1, \ell')$ with $\ell, \ell' \in \ZZ_+$. \end{enumerate} \item Every vertex $x$ in $J_j$ with $\deg_{J_j}(x)=4$ is either converging or diverging. \end{enumerate} \end{proposition} \begin{proof} We simultaneously prove (1) and (2) by induction on $j \ge 0$. The base step $j=0$ is already proved in Proposition~\ref{prop:general_critical}. So fix $j \ge 1$. By the inductive hypothesis, $CS(\phi(\delta_{j-1}^{-1}))=CS(\phi(\alpha_j))$ consists of $m$ and $m+1$. (1a) Suppose on the contrary that $S(\phi(\partial D_{j,1}^+))$ ends with $S_1$ so that $|z_{j,1}|=0$ (see Notation~\ref{not:layers2}). (The other case is treated similarly.) Then $S(z_{j,1}')=S_2$ and so $S(z_{j,1,e}')=(m)$. Note also that the assumption implies that the vertex between $D_{j,1}$ and $D_{j,2}$ is not converging nor diverging by Lemma~\ref{lem:converging2}. Thus, arguing as in the proof of Lemma~\ref{lem:general_critical(a)} when Lemma~\ref{lem:converging}(1) does not hold, we see that $S(z_{j,1,e}'y_{j,2}'w_{j,2}')$ begins with a subsequence of the form $(m+d)$ with $d \in \ZZ_+$. Since $S(\phi(\partial D_{j,1}^+))=CS(\phi(\delta_{j-1}^{-1}))$ ends with a term $m+1$, and since $CS(\phi(\alpha_j))$ consists of $m$ and $m+1$ by the inductive hypothesis, the only possibility is that $d=1$ and $S(w_{j,1,e} y_{j,2,b})=(m+1, m)$. Then, as illustrated in Figure~\ref{fig.S_1_occurs(b)}, we may transform $M$ so that $S(\phi(\partial D_{j,1}^+))$ ends with $(S_1, m)$. Since the new vertex of $M$ is either converging or diverging, we see by Lemma~\ref{lem:converging2} that none of the four (prohibited) conditions in the lemma holds. Thus by repeating this argument at every degree $4$ vertex of $J_j$, we obtain the desired result. (1b) Suppose on the contrary that $S(\phi(\partial D_{j,1}^-))$ ends with $S_1$. (The other case is treated similarly.) Then $|z_{j,1}'|=0$. It follows that $S(z_{j,1})=S_2$, so that $S(z_{j,1,e})=(m)$. Hence the sequence $S(z_{j,1,e}y_{j,2}w_{j,2})$ begins with a subsequence of the form either $(m, d)$ or $(m+d)$, where $d \in \ZZ_+$. Here, if $S(z_{j,1,e}y_{j,2}w_{j,2})$ begins with a subsequence of the form $(m, d)$, then we see by an argument as in the proof of Lemma \ref{lem:converging}(1) that two $2$-cells $D_{j,1}$ and $D_{j-1,2}$ are a reducible pair, contradicting that $M$ is a reduced diagram. So $S(z_{j,1,e}y_{j,2}w_{j,2})$ must begin with a subsequence of the form $(m+d)$. Since $S(\phi(\partial D_{j,1}^+))$ ends with a term $m$, and since $S(\phi(\partial D_{j,1}^+))=CS(\phi(\delta_{j-1}^{-1}))$ consists of $m$ and $m+1$ by the inductive hypothesis, the only possibility is that $d=1$ and $S(w_{j,1,e}' y_{j,2,b}')=(m+1, m)$. Then, as illustrated in Figure~\ref{fig.S_1_occurs(c)}, we may transform $M$ so that $S(\phi(\partial D_{j,1}^-))$ ends with $(S_1, m)$. Since the new vertex of $M$ is converging or diverging, we see by Lemma~\ref{lem:converging2} that none of the four (prohibited) conditions in the lemma holds. Thus by repeating this argument at every degree $4$ vertex of $J_j$, we obtain the desired result. (2) As in the proof of Proposition~\ref{prop:general_critical}, we show that we can modify $M$ so that it satisfies (2). Then it continues to satisfy (1) by Lemma~\ref{lem:converging2}. To this end, note that $CS(\phi(\alpha_j))$ consists of $m$ and $m+1$ by the inductive hypothesis. By using this fact, we can see that the statement of Lemma~\ref{lem:general_critical(b)} holds, where $z_{i,e}$, $z_{i,e}'$, $y_{i+1,e}$ and $y_{i+1,e}'$ are replaced with $z_{j,i,e}$, $z_{j,i,e}'$, $y_{j,i+1,e}$ and $y_{j,i+1,e}'$, respectively. (The proof of Lemma~\ref{lem:general_critical(b)} works if we appeal to the fact that $CS(\phi(\alpha_j))$, instead of $CS(\phi(\alpha))$, consists of $m$ and $m+1$.) So we can follow the proof of Proposition~\ref{prop:general_critical} and show that (2) holds. \end{proof} \begin{corollary} \label{cor:general_critical(aa)} Assume that $r$ is general. Under Hypothesis~A and Notation~\ref{not:layers}, we may assume that the following hold. \begin{enumerate}[\indent \rm (1)] \item Every vertex $x$ in $M$ with $\deg_{M}(x)=4$ is either converging or diverging. \item For every face $D_{j,i}$ of $M$, one of the following hold. \begin{enumerate}[\rm (a)] \item Both $S(\phi(\partial D_{j,i}^+))$ and $S(\phi(\partial D_{j,i}^-))$ contain $(m,S_1,m)$ as a subsequence. \item Both $S(\phi(\partial D_{j,i}^+))$ and $S(\phi(\partial D_{j,i}^-))$ contain $(m+1,S_2,m+1)$ as a subsequence. \end{enumerate} \item For every $j$, both $CS(\phi(\alpha_j))$ and $CS(\phi(\delta_j^{-1}))$ consist of $m$ and $m+1$. \end{enumerate} \end{corollary} \begin{proof} (1) is nothing other than Proposition~\ref{prop:general_critical(aa)}(2). (2) follows from Lemma~\ref{lem:two_cases:intermediate_layer}, Proposition~\ref{prop:general_critical(aa)}(1) and the fact that $S_1$ (resp., $S_2$) begins and ends with $m+1$ (resp., $m$). \end{proof} \section{Key results for the induction} \label{sec:result_for_induction} In this section, we prove key results, Propositions~\ref{prop:induction_general_1} and \ref{prop:induction_general_3} for $r=[m,m_2, \dots, m_k]$ with $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$ and $r=[m,1,m_3, \dots, m_k]$ with $m \ge 2$ and $k \ge 4$, respectively, used for the inductive proof of Main Theorem~\ref{main_theorem} in Section~\ref{sec:proof_for_general_2-bridge_links}. Throughout this section, we assume that Hypothesis~A holds and that the annular diagram $M$ satisfies the conditions in Corollary \ref{cor:general_critical(aa)}. \subsection{The case for $r=[m,m_2, \dots, m_k]$ with $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$} We first establish a key result, Proposition~\ref{prop:induction_general_1}, for $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Recall from Remark~\ref{rem:general_decomposition}(1) that $CS(r)=\lp S_1, S_2, S_1, S_2 \rp$, where $S_1$ begins and ends with $(m+1, (m_2-1) \langle m \rangle, m+1)$, and $S_2$ begins and ends with $(m_2 \langle m \rangle)$. \begin{lemma} \label{lem:general_1(b)} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Under Hypothesis~A and Notation~\ref{not:layers2}, the following hold for every $i$ and $j$. \begin{enumerate}[\indent \rm (1)] \item $S(z_{j,i,e}y_{j,i+1,b}) \neq (m+1, m+1)$. \item $S(z_{j,i,e}'y_{j,i+1,b}') \neq (m+1, m+1)$. \end{enumerate} \end{lemma} \begin{proof} We prove only (1), because the proof of (2) is parallel. Suppose on the contrary that $S(z_{j,1,e}y_{j,2,b})=(m+1, m+1)$ for some $j$. First assume $j=0$. If Hypothesis~B holds, then $S(z_{0,1})$ begins with $(m_2 \langle m \rangle)$, because $S_2$ begins and ends with $(m_2 \langle m \rangle)$ (see Remark~\ref{rem:general_decomposition}(1)) whereas $S(z_{0,1,e})=(m+1)$ by assumption. This implies that $CS(\phi(\alpha_0))=CS(s)$ contains two consecutive $m$'s. So $CS(s)$ contains two consecutive $m$'s and two consecutive $m+1$'s, contradicting \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if Hypothesis~C holds, then $CS(s)$ contains two consecutive $m$'s (because it contains $S_2$) and two consecutive $m+1$'s by assumption, again contradicting \cite[Lemma~3.8]{lee_sakuma_2}. Next assume $j \ge 1$. By using Lemma~\ref{lem:vertex_position:intermediate_layer}, we can see that $S(\phi(e_{j,2}e_{j,3}))$ contains a subsequence $(m+1, m+1)$. Thus $CS(\phi(\partial D_{j-1,2}))=CS(r)$ contains two consecutive $m+1$'s, a contradiction. \end{proof} \begin{corollary} \label{cor:general_1} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Under Hypothesis~A and Notation~\ref{not:layers}, the following hold for every $j$. \begin{enumerate}[\indent \rm (1)] \item $CS(\phi(\alpha_j))$ does not contain $(m+1, m+1)$ as a subsequence. \item $CS(\phi(\delta_j^{-1}))$ does not contain $(m+1, m+1)$ as a subsequence. \end{enumerate} \end{corollary} \begin{proof} We prove only (1), because the proof of (2) is parallel. Suppose on the contrary that $CS(\phi(\alpha_j))$ contains a subsequence $(m+1, m+1)$ for some $j$. Let $v$ be a subword of the cyclic word $(\phi(\alpha_j))$ corresponding to a subsequence $(m+1, m+1)$. Note that $S(\phi(\partial D_{j,i}^+))$ does not contain $(m+1, m+1)$, because $S(r)=S(\phi(\partial D_{j,i}))$ does not. Thus the only possibility is that $v=z_{j,i,e}y_{j,i+1,b}$ for some $i$ by Corollary~\ref{cor:general_critical(aa)}. But this is impossible by Lemma~\ref{lem:general_1(b)}(1). \end{proof} \begin{proposition} \label{prop:induction_general_1} Let $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Suppose that there are two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$ such that the unoriented loops $\alpha_{s}$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$, namely suppose that Hypothesis~A holds. Let $\tilde{r}=[m_2-1, m_3, \dots, m_k]$ be as in \cite[Lemma~3.11]{lee_sakuma_2}. Then there are two distinct rational numbers $\tilde{s}, \tilde{s}' \in I_1(\tilde{r}) \cup I_2(\tilde{r})$ such that the unoriented loops $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ are homotopic in $S^3-K(\tilde{r})$. Moreover, there is a reduced nontrivial conjugacy diagram over $G(K(\tilde{r}))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ such that none of the degree $4$ vertices is mixing. \end{proposition} \begin{proof} Recall from Corollaries~\ref{cor:general_critical(aa)} and~\ref{cor:general_1} that both $CS(\phi(\alpha_j))$ and $CS(\phi(\delta_j^{-1}))$ consist of $m$ and $m+1$ without a subsequence $(m+1, m+1)$ for every $j$. In particular, both $CS(\phi(\alpha_0))=CS(s)$ and $CS(\phi(\delta_p^{-1}))=CS(s')$ consist of $m$ and $m+1$ without a subsequence $(m+1, m+1)$. This implies that if $s=[p_1, p_2, \dots, p_h]$ and $s'=[q_1, q_2, \dots, q_l]$, where $p_i, q_j \in \ZZ_+$ and $p_h, q_l \ge 2$, then $p_1=q_1=m$ and $p_2, q_2 \ge 2$. Put $\tilde{s}=[p_2-1, p_3, \dots, p_h]$ and $\tilde{s}'=[q_2-1, q_3, \dots, q_l]$ as in \cite[Lemma~3.11]{lee_sakuma_2}. \medskip {\bf Claim.} {\it Both $\tilde{s}$ and $\tilde{s}'$ belong to $I_1(\tilde{r}) \cup I_2(\tilde{r})$.} \begin{proof}{\it of Claim } Since $p_1=q_1=m$, we have \[ \tilde{s}= \cfrac{1}{ \cfrac{1}{-p_1+ \cfrac{1}{s}}-1} = \cfrac{1}{ \cfrac{1}{-m+ \cfrac{1}{s}}-1} \quad \textrm{and} \quad \tilde{s}'= \cfrac{1}{ \cfrac{1}{-q_1+ \cfrac{1}{s'}}-1} = \cfrac{1}{ \cfrac{1}{-m+ \cfrac{1}{s'}}-1}. \] Put $I_1(r)=[0,r_1]$ and $I_2(r)=[r_2,1]$. Recall from \cite[Section~2]{lee_sakuma_2} that \begin{align*} r_1 &= \begin{cases} [m, m_2, \dots, m_{k-1}] & \mbox{if $k$ is odd,}\\ [m, m_2, \dots, m_{k-1}, m_k-1] & \mbox{if $k$ is even,} \end{cases}\\ r_2 &= \begin{cases} [m, m_2, \dots, m_{k-1}, m_k-1] & \mbox{if $k$ is odd,}\\ [m, m_2, \dots, m_{k-1}] & \mbox{if $k$ is even.} \end{cases} \end{align*} Also put $I_1(\tilde{r})=[0, t_1]$ and $I_2(\tilde{r})=[t_2,1]$; then we have \begin{align*} t_1 &= \begin{cases} [m_2-1, \dots, m_{k-1}, m_k-1] & \mbox{if $k$ is odd,}\\ [m_2-1, \dots, m_{k-1}] & \mbox{if $k$ is even,} \end{cases} \\ t_2 &= \begin{cases} [m_2-1, \dots, m_{k-1}] & \mbox{if $k$ is odd,}\\ [m_2-1, \dots, m_{k-1}, m_k-1] & \mbox{if $k$ is even.} \end{cases} \end{align*} It then follows that \[ t_1= \cfrac{1}{ \cfrac{1}{-m+ \cfrac{1}{r_2}}-1} \quad \textrm{and} \quad t_2= \cfrac{1}{ \cfrac{1}{-m+ \cfrac{1}{r_1}}-1}. \] Therefore the fact $s, s' \in I_1(r) \cup I_2(r)$ yields $\tilde{s}, \tilde{s}' \in I_1(\tilde{r}) \cup I_2(\tilde{r})$. \end{proof} Let $\tilde{R}$ be the symmetrized subset of $F(a, b)$ generated by the single relator $u_{\tilde{r}}$ of the upper presentation $G(K(\tilde{r}))=\langle a, b \, | \, u_{\tilde{r}} \rangle$. Then as described below, we can construct a reduced annular $\tilde{R}$-diagram $\tilde{M}$ such that $u_{\tilde{s}}$ is an outer boundary label and $u_{\tilde{s}'}^{\pm 1}$ is an inner boundary label of $\tilde{M}$. This proves that the unoriented loops $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ are homotopic in $S^3-K(\tilde{r})$. \end{proof} We shall describe the explicit construction of a reduced annular $\tilde{R}$-diagram $\tilde{M}$ from $M$. To this end, we introduce the following definition. \begin{definition} \label{def:T-sequence_1} \rm Suppose $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$. Let $w$ be an alternating word in $\{a,b\}$, and suppose that $S(w)=(a_1,a_2,\cdots,a_k)$ is a finite sequence consisting of $m$ and $m+1$, which does not contain $(m+1,m+1)$. Then we define the $T$-sequence of $w$, denoted by $T(w)$, and the cyclic $T$-sequence, denoted by $CT(w)$, as follows. Express $S(w)$ as \[ (*, t_1\langle m\rangle, m+1, t_2\langle m\rangle, \dots,m+1, t_s\langle m\rangle,*'), \] where each of $*$ and $*'$ is either $m+1$ or $\emptyset$ and $(t_1,t_2,\dots,t_s)$ is a sequence of positive integers. Then $T(w)$ is defined to be the sequence $(t_1,\cdots, t_s)$. If precisely one of $*$ and $*'$ is $m+1$ and the other is $\emptyset$, we define $CT(w)$ to be the cyclic sequence $\lp t_1,\cdots, t_s\rp$. If this $w$ represents a reduced cyclic word $u=(w)$, then we define the cyclic sequence $CT(u)$ by $CT(w)$. \end{definition} Under Hypothesis~A and Notation~\ref{not:layers}, by Corollaries~\ref{cor:general_critical(aa)} and~\ref{cor:general_1}, both $CS(\phi(\alpha_j))$ and $CS(\phi(\delta_j^{-1}))$ consist of $m$ and $m+1$ without a subsequence $(m+1, m+1)$, so the cyclic sequences $CT(\phi(\alpha_j))$ and $CT(\phi(\delta_j^{-1}))$ are well-defined for every $j$. Recall that every vertex in $M$ of degree $4$ is assumed to be converging or diverging by Corollary~\ref{cor:general_critical(aa)}. Moreover, we can also assume, by using Corollary~\ref{cor:general_critical(aa)}(2), that every degree $2$ vertex of $M$ is also either converging or diverging. (In the new diagram, it may happen that some $\phi(e_{0,i})$ or some $\phi(e_{p,i}')$ is not a piece and so \cite[Convention~4.6]{lee_sakuma_2} is not satisfied. But this does not affect the arguments in this section.) Hence $S(\phi(e_{j,i}))$ is a subsequence of $S(\phi(\alpha_j))$, which consists of $m$ and $m+1$ and does not contain $(m+1,m+1)$ as a subsequence. Thus the $T$-sequence of $\phi(e_{j,i})$ is also well-defined for every $i$ and $j$. Similarly the $T$-sequence of $\phi(e_{j,i}')$ is also well-defined for every $i$ and $j$. By Corollary~\ref{cor:general_critical(aa)}(2), we may assume that $T(\phi(e_{j,i}))$ and $T(\phi(e_{j,i}'))$ are nonempty for every $i$ and $j$. Now we construct a reduced annular $\tilde{R}$-diagram $(\tilde M,\psi)$ from the annular $R$-diagram $(M,\phi)$ by taking $T$-sequences of the boundary labels, as follows. Take the underlying map of $\tilde{M}$ being the same as that of $M$. For every $i$ and $j$, by $\tilde{e}_{j,i}$ denote the edge of $\tilde{M}$ which corresponds to the edge $e_{j,i}$ of $M$, and assign an alternating word, $\psi(\tilde{e}_{j,i})$, in $\{a,b\}$ to $\tilde{e}_{j,i}$ in the following order. \medskip \noindent {\bf Step 1.} For each $i=1, \dots, 2t$, assign $\psi(\tilde{e}_{0,i})$ so that $\psi(\tilde{e}_{0,1} \cdots \tilde{e}_{0,i}):= \psi(\tilde{e}_{0,1}) \cdots \psi(\tilde{e}_{0,i})$ is an alternating word such that \[ S(\psi(\tilde{e}_{0,1} \cdots \tilde{e}_{0,i}))=T(\phi(e_{0,1} \cdots e_{0,i})). \] Once this assignment is done, we see the following. \begin{enumerate}[\indent \rm (i)] \item The word $\psi(\tilde{e}_{0,1} \cdots \tilde{e}_{0,2t})$ is cyclically alternating, because the sum of the terms of $CT(s)=CS(\tilde{s})$ is even. \item $CS(\psi(\tilde{e}_{0,1} \cdots \tilde{e}_{0,2t}))= CT(\phi(e_{0,1} \cdots e_{0,2t}))=CT(\phi(\alpha_0))=CT(s)=CS(\tilde{s})$, because $CS(\tilde{s})$ has even number of terms. \end{enumerate} \medskip \noindent {\bf Step 2.} Assign $\psi(\tilde{e}_{0,1}')$ so that $\psi(\tilde{e}_{0,1}^{-1} \tilde{e}_{0,1}')$ is an alternating word such that $S(\psi(\tilde{e}_{0,1}^{-1} \tilde{e}_{0,1}'))=T(\phi(e_{0,1}^{-1} e_{0,1}'))$, and assign $\psi(\tilde{e}_{0,2}')$ so that $\psi(\tilde{e}_{0,1}' \tilde{e}_{0,2}')$ is an alternating word such that $S(\psi(\tilde{e}_{0,1}' \tilde{e}_{0,2}'))=T(\phi(e_{0,1}' e_{0,2}'))$. Once this assignment is done, we see the following. \begin{enumerate}[\indent \rm (i)] \item The word $\psi(\tilde{e}_{0,1} \tilde{e}_{0,2} \tilde{e}_{0,2}'^{-1} \tilde{e}_{0,1}'^{-1})$ is cyclically alternating, because the sum of the terms of $CT(r)=CS(\tilde r)$ is even. \item $CS(\psi(\tilde{e}_{0,1} \tilde{e}_{0,2} \tilde{e}_{0,2}'^{-1} \tilde{e}_{0,1}'^{-1})) =CT(\psi(e_{0,1} e_{0,2}e_{0,2}'^{-1} e_{0,1}'^{-1}))=CT(r)=CS(\tilde{r})$, because $CS(\tilde{r})$ has even number of terms. \end{enumerate} \medskip \noindent {\bf Step 3.} For $i=2, \dots, t$, assign $\psi(\tilde{e}_{0,2i-1}')$ so that $\psi(\tilde{e}_{0,2i-1}^{-1} \tilde{e}_{0,2i-1}')$ is an alternating word such that $S(\psi(\tilde{e}_{0,2i-1}^{-1} \tilde{e}_{0,2i-1}'))=T(\phi(e_{0,2i-1}^{-1} e_{0,2i-1}'))$, and assign $\psi(\tilde{e}_{0,2i}')$ so that $\psi(\tilde{e}_{0,2i-1}' \tilde{e}_{0,2i}')$ is an alternating word such that $S(\psi(\tilde{e}_{0,2i-1}' \tilde{e}_{0,2i}'))=T(\phi(e_{0,2i-1}' e_{0,2i}'))$. Then we see the following. \begin{enumerate}[\indent \rm (i)] \item $S(\psi(\tilde{e}_{0,1}' \cdots \tilde{e}_{0,2i-1}'))=T(\phi(e_{0,1}' \cdots e_{0,2i-1}'))$, because of the reason described after this list. \item The word $\psi(\tilde{e}_{0,2i-1} \tilde{e}_{0,2i} \tilde{e}_{0,2i}'^{-1} \tilde{e}_{0,2i-1}'^{-1})$ is cyclically alternating, because the sum of the terms of $CT(r)=CS(\tilde r)$ is even. \item $CS(\psi(\tilde{e}_{0,2i-1} \tilde{e}_{0,2i} \tilde{e}_{0,2i}'^{-1} \tilde{e}_{0,2i-1}'^{-1})) =CT(\phi(e_{0,2i-1} e_{0,2i} e_{0,2i}'^{-1} e_{0,2i-1}'^{-1}))=CT(r)=CS(\tilde{r})$, because $CS(\tilde{r})$ has even number of terms. \item The word $\psi(\tilde{e}_{0,1}' \cdots \tilde{e}_{0,2t}')$ is cyclically alternating, because the sum of the terms of $CT(\phi(\delta_0))$ is even. \item $CS(\psi(\tilde{e}_{0,1}' \cdots \tilde{e}_{0,2t}'))= CT(\phi(e_{0,1}' \cdots e_{0,2t}'))=CT(\phi(\delta_0))$, because $CT(\phi(\delta_0))$ has even number of terms. \end{enumerate} In the above, condition (i) is verified as follows. We explain the reason when $i=2$. (The other cases can be treated similarly.) Since $CS(\phi(\alpha_0))=CS(s)$ and $CS(\phi(\partial D_1))=CS(\phi(\partial D_2))=CS(r)$ consist of $m$ and $m+1$ and do not contain $(m+1,m+1)$, we have the four possibilities around the vertex between $D_1$ and $D_2$ as described in the left figures in Figure~\ref{fig.taking_tsequence_1}, up to reflection in the vertical line passing through the vertex. In each of the right figure, we may assume without loss of generality that the upper left segment is oriented so that it is converging into the vertex. Then the orientations of the three remaining segments in each of the right figures are specified by the requirements in Steps~1 and 2 and the the new requirement $S(\psi(\tilde{e}_{0,3}^{-1} \tilde{e}_{0,3}'))=T(\phi(e_{0,3}^{-1} e_{0,3}'))$. In each case, we can check that the condition $S(\psi(\tilde{e}_{0,1}' \tilde{e}_{0,2}' \tilde{e}_{0,3}'))= T(\phi(e_{0,1}'e_{0,2}' e_{0,3}'))$ holds. \begin{figure}[h] \includegraphics{figure_taking_tsequence_1.eps} \caption{ Step~3 of the construction of $\tilde{M}$ from $M$} \label{fig.taking_tsequence_1} \end{figure} \medskip \noindent {\bf Step 4.} For each $j=1, \dots, p$, repeat Steps~2 and 3 to obtain the following. \begin{enumerate}[\indent \rm (i)] \item The word $\psi(\tilde{e}_{j,2i-1} \tilde{e}_{j,2i} \tilde{e}_{j,2i}'^{-1} \tilde{e}_{j,2i-1}'^{-1})$ is cyclically alternating for every $i=1, \dots, t$. \item $CS(\psi(\tilde{e}_{j,2i-1} \tilde{e}_{j,2i} \tilde{e}_{j,2i}'^{-1} \tilde{e}_{j,2i-1}'^{-1})) =CT(\phi(e_{j,2i-1} e_{j,2i}e_{j,2i}'^{-1}e_{j,2i-1}'^{-1}))=CT(r)=CS(\tilde{r})$ for every $i=1, \dots, t$. \item The word $\psi(\tilde{e}_{p,1}' \cdots \tilde{e}_{p,2t}')$ is cyclically alternating. \item $CS(\psi(\tilde{e}_{p,1}' \cdots \tilde{e}_{p,2t}'))=CT(\phi(e_{p,1}' \cdots e_{p,2t}')) =CT(\psi(\delta_p))=CT(s')=CS(\tilde{s}')$. \end{enumerate} It is obvious from the construction (see Figure~\ref{fig.taking_tsequence_1}) that none of the degree $4$ vertices of the diagram $\tilde M$ constructed above is mixing. Finally we show that $\tilde M$ is reduced. Suppose on the contrary that $\tilde M$ is not reduced. Then there is a pair of faces, say ${\tilde D}$ and ${\tilde D}'$, in $\tilde M$ having a common edge, say ${\tilde e}=\partial {\tilde D} \cap \partial {\tilde D}'$, such that $\psi(\delta_1) \equiv \psi(\delta_2)^{-1}$, where ${\tilde e} \delta_1$ and $\delta_2 {\tilde e}^{-1}$ are boundary cycles of ${\tilde D}$ and ${\tilde D}'$, respectively. Then we see that the corresponding faces $D$ and $D'$ in $M$ have a common edge $e=\partial D \cap \partial D'$ such that $S(\phi(e\gamma_1))=S(\phi(e\gamma_2^{-1}))$, where $e \gamma_1$ and $\gamma_2 e^{-1}$ are boundary cycles of $D$ and $D'$, respectively. So two words $\phi(e\gamma_1) \equiv \phi(e)\phi(\gamma_1)$ and $\phi(e\gamma_2^{-1}) \equiv \phi(e)\phi(\gamma_2)^{-1}$ have the same initial letter and the same associated $S$-sequence. Then by \cite[Lemma~3.5(1)]{lee_sakuma_2} $\phi(e)\phi(\gamma_1) \equiv \phi(e)\phi(\gamma_2)^{-1}$, so $\phi(\gamma_1) \equiv \phi(\gamma_2)^{-1}$, which contradicts the fact that $M$ is reduced. \subsection{The case for $r=[m,1,m_3, \dots, m_k]$ with $m \ge 2$ and $k \ge 4$} We next establish a key result, Proposition~\ref{prop:induction_general_3}, for $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Recall from Remark~\ref{rem:general_decomposition}(2) that $CS(r)=\lp S_1, S_2, S_1, S_2 \rp$, where $S_1$ begins and ends with $((m_3+1) \langle m+1 \rangle)$, and $S_2$ begins and ends with $(m, m_3 \langle m+1 \rangle, m)$. \begin{lemma} \label{lem:general_3(b)} Let $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Under Hypothesis~A and Notation~\ref{not:layers2}, the following hold for every $i$ and $j$. \begin{enumerate}[\indent \rm (1)] \item $S(z_{j,i,e}y_{j,i+1,b}) \neq (m, m)$. \item $S(z_{j,i,e}'y_{j,i+1,b}') \neq (m, m)$. \end{enumerate} \end{lemma} \begin{proof} We prove only (1), because the proof of (2) is parallel. Suppose on the contrary that $S(z_{j,1,e}y_{j,2,b})=(m, m)$ for some $j$. First assume $j=0$. If Hypothesis~B holds, then $CS(\phi(\alpha_0))=CS(s)$ contains two consecutive $m+1$'s (because it contains $S_1$) and two consecutive $m$'s by assumption. This contradicts \cite[Lemma~3.8]{lee_sakuma_2}. On the other hand, if Hypothesis~C holds, then $S(z_{0,1})$ begins with $((m_3+1) \langle m+1 \rangle)$, since $S_1$ begins and ends with $((m_3+1) \langle m+1 \rangle)$ whereas $S(z_{0,1,e})=m$ by assumption. This implies that $CS(s)$ contains two consecutive $m$'s and two consecutive $m+1$'s, again contradicting \cite[Lemma~3.8]{lee_sakuma_2}. Next assume $j \ge 1$. By using Lemma~\ref{lem:vertex_position:intermediate_layer}, we can see that $S(\phi(e_{j,2}e_{j,3}))$ contains a subsequence of the form $(\ell_1, m, m, \ell_2)$ with $\ell_1, \ell_2 \in \ZZ_+$. Hence $CS(\phi(\partial D_{j-1,2}))=CS(r)$ contains two consecutive $m$'s, a contradiction. \end{proof} \begin{corollary} \label{cor:general_3} Let $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Under Hypothesis~A and Notation~\ref{not:layers}, the following hold for every $j$. \begin{enumerate}[\indent \rm (1)] \item No two consecutive terms of $CS(\phi(\alpha_j))$ can be $(m, m)$. \item No two consecutive terms of $CS(\phi(\delta_j^{-1}))$ can be $(m, m)$ for every $j=0, \dots, p$. \end{enumerate} \end{corollary} \begin{proof} We prove only (1), because the proof of (2) is parallel. Suppose on the contrary that $CS(\phi(\alpha_j))$ contains a subsequence $(m, m)$ for some $j$. Let $v$ be a subword of the cyclic word $(\phi(\alpha_j))$ corresponding to a subsequence $(m, m)$. Note that $S(\phi(\partial D_{j,i}^+))$ does not contain $(m, m)$, because $S(r)=S(\phi(\partial D_{j,i}))$ does not. Thus the only possibility is that $v=z_{j,i,e}y_{j,i+1,b}$ for some $i$ by Corollary~\ref{cor:general_critical(aa)}. But this is impossible by Lemma~\ref{lem:general_3(b)}(1). \end{proof} \begin{definition} \label{def:T-sequence_2} \rm Suppose $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Let $w$ be an alternating word in $\{a,b\}$, and suppose that $S(w)=(a_1,a_2,\cdots,a_k)$ is a finite sequence consisting of $m$ and $m+1$, which does not contain $(m,m)$. Then we define the $T$-sequence of $w$, denoted by $T(w)$, and the cyclic $T$-sequence, denoted by $CT(w)$, as follows. Express $S(w)$ as \[ (*, t_1\langle m+1\rangle, m, t_2\langle m+1\rangle, \dots,m, t_s\langle m+1\rangle,*'), \] where each of $*$ and $*'$ is either $m$ or $\emptyset$ and $(t_1,t_2,\dots,t_s)$ is a sequence of positive integers. Then $T(w)$ is defined to be the sequence $(t_1,\cdots, t_s)$. If precisely one of $*$ and $*'$ is $m$ and the other is $\emptyset$, we define $CT(w)$ to be the cyclic sequence $\lp t_1,\cdots, t_s\rp$. If this $w$ represents a reduced cyclic word $u=(w)$, then we define the cyclic sequence $CT(u)$ by $CT(w)$. \end{definition} Under Hypothesis~A and Notation~\ref{not:layers}, by Corollaries~\ref{cor:general_critical(aa)} and~\ref{cor:general_3}, both $CS(\phi(\alpha_j))$ and $CS(\phi(\delta_j^{-1}))$ consist of $m$ and $m+1$ without a subsequence $(m, m)$, so the cyclic sequences $CT(\phi(\alpha_j))$ and $CT(\phi(\delta_j^{-1}))$ are well-defined for every $j$. Clearly we may assume that every degree $2$ vertex of $M$ is either converging or diverging. Moreover since every vertex in $M$ of degree $4$ is assumed to be converging or diverging by Corollary~\ref{cor:general_critical(aa)}, the $T$-sequence of $\phi(e_{j,i})$ is also well-defined for every $i$ and $j$. Then as in the previous case, we can construct an annular $\tilde{R}$-diagram $\tilde M$ from the annular $R$-diagram $M$ by taking $T$-sequences of the boundary labels. \begin{proposition} \label{prop:induction_general_3} Let $r=[m,1,m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Suppose that there are two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$ such that the unoriented loops $\alpha_{s}$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$, namely suppose that Hypothesis~A holds. Let $\tilde{r}=[m_3, \dots, m_k]$ be as in \cite[Lemma~3.11]{lee_sakuma_2}. Then there are two distinct rational numbers $\tilde{s}, \tilde{s}' \in I_1(\tilde{r}) \cup I_2(\tilde{r})$ such that the unoriented loops $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ are homotopic in $S^3-K(\tilde{r})$. Moreover, there is a reduced conjugacy diagram over $G(K(\tilde{r}))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ such that none of the degree $4$ vertices is mixing. \end{proposition} \begin{proof} Recall from Corollaries~\ref{cor:general_critical(aa)} and~\ref{cor:general_3} that both $CS(\phi(\alpha_j))$ and $CS(\phi(\delta_j^{-1}))$ consist of $m$ and $m+1$ without a subsequence $(m, m)$, for every $j$. In particular, both $CS(\phi(\alpha_0))=CS(s)$ and $CS(\phi(\delta_p^{-1}))=CS(s')$ consist of $m$ and $m+1$ without a subsequence $(m, m)$. This implies that if $s=[p_1, p_2, \dots, p_h]$ and $s'=[q_1, q_2, \dots, q_l]$, where $p_i, q_j \in \ZZ_+$ and $p_h, q_l \ge 2$, then $p_1=q_1=m$, $p_2=q_2=1$ and $h, l \ge 3$. Put $\tilde{s}=[p_3, \dots, p_h]$ and $\tilde{s}'=[q_3, \dots, q_l]$ as in \cite[Lemma~3.11]{lee_sakuma_2}. Let $\tilde{R}$ be the symmetrized subset of $F(a, b)$ generated by the single relator $u_{\tilde{r}}$ of the upper presentation $G(K(\tilde{r}))=\langle a, b \, | \, u_{\tilde{r}} \rangle$. Then, as in the previous case, we can construct a reduced annular $\tilde{R}$-diagram $(\tilde{M},\psi)$ such that $u_{\tilde{s}}$ is an outer boundary label and $u_{\tilde{s}'}^{\pm 1}$ is an inner boundary label of $\tilde{M}$. This proves that the unoriented loops $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ are homotopic in $S^3-K(\tilde{r})$. Moreover, we can also see that $\tilde M$ is reduced and that none of the degree $4$ vertices of the diagram $\tilde M$ constructed above is mixing. \end{proof} \section{Proof of Main Theorem~\ref{main_theorem} for the general cases} \label{sec:proof_for_general_2-bridge_links} In this section, we prove Main Theorem~\ref{main_theorem} when $r$ is general. To this end, we use the following terminology. \begin{enumerate}[\indent \rm (1)] \item A rational number $r$ with $0< r\le 1/2$ is {\it special}, if it is equal to $1/p=[p]$ with $p\ge 2$, $[m,n]$, or $[m,1,n]$ with $m,n\ge 2$. \item A rational number $r$ with $1/2< r<1$ is {\it special}, if $1-r$ is special. \item A rational number $r$ with $0< r<1$ is {\it general}, if it is not special. \end{enumerate} The following proposition forms the starting point of the inductive proof of Main Theorem~\ref{main_theorem}. \begin{proposition} \label{prop:induction_base} Let $r$ be a special rational number with $0<r<1$. Then the following is the complete list of pairs of mutually distinct elements $\{s,s'\}$ of $I_1(r)\cup I_2(r)$ such that $\alpha_{s}$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$. \begin{enumerate}[\indent \rm(1)] \item $r=1/p$ and the set $\{s, s'\}$ equals $\{q_1/p_1, q_2/p_2\}$, where $p \ge 2$ is an integer, and $s=q_1/p_1$ and $s'=q_2/p_2$ satisfy $q_1=q_2$ and $q_1/(p_1+p_2)=1/p$, where $(p_i, q_i)$ is a pair of relatively prime positive integers. \item $r=3/8=[2, 1, 2]$ and the set $\{s, s'\}$ equals either $\{1/6, 3/10\}$ or $\{3/4, 5/12\}$. \item $r=1-1/p=[1,p-1]$ and the set $\{s, s'\}$ equals $\{1-q_1/p_1, 1-q_2/p_2\}$, where $p \ge 2$ is an integer, and $q_1=q_2$ and $q_1/(p_1+p_2)=1/p$, where $(p_i, q_i)$ is a pair of relatively prime positive integers. \item $r=1-3/8=[1,1,1,2]$ and the set $\{s, s'\}$ equals either $\{1-1/6, 1-3/10\}$ or $\{1-3/4, 1-5/12\}$. \end{enumerate} Moreover, any reduced conjugacy diagram over $G(K(r))$ for $\alpha_{s}$ and $\alpha_{s'}$ contains a vertex which is mixing. \end{proposition} \begin{proof} Suppose $0<r \le 1/2$. Then by \cite[Main Theorem~2.7]{lee_sakuma_2}, \cite[Main Theorems~2.2 and~2.3]{lee_sakuma_3} and the results in Sections~\ref{sec:proof_of_main_theorem_1} and \ref{sec:proof_of_main_theorem_3}, we see that (1) or (2) holds. We prove the assertion for the conjugacy diagram in this case. Suppose that (1) holds. Then the outer boundary layer of the conjugacy diagram should be as depicted in Figure~\ref{fig.general_proof_0} by \cite[Section~5]{lee_sakuma_2}. Since all vertices of the diagram are mixing, we obtain the desired result. Suppose that (2) holds. Then, by \cite[Section~6, in particular Figures~21(a) and 23(a)]{lee_sakuma_3}, the conjugacy diagram should be as depicted in Figure~\ref{fig.general_proof_1}. Again, every vertex of degree $4$ is mixing, and so we obtain the desired result. \begin{figure}[h] \includegraphics{figure_general_proof_0.eps} \caption{ The outer boundary layer of any of the conjugacy diagrams for the case $r=1/p$} \label{fig.general_proof_0} \end{figure} \begin{figure}[h] \includegraphics{figure_general_proof_1.eps} \caption{ The conjugacy diagrams for the case $r=[2,1,2]$, where $\{s,s'\}$ is (a) $\{1/6,3/10\}$ and (b) $\{3/4,5/12\}$} \label{fig.general_proof_1} \end{figure} Suppose $1/2< r<1$. Note that there is a homeomorphism $f:(S^3,K(r))\to (S^3,K(1-r))$ preserving the bridge sphere such that $f(\alpha_s)=\alpha_{1-s}$ and that $f$ induces an isomorphism from $G(K(r))$ to $G(K(1-r))$ sending the standard generators $a$ and $b$ to $a$ and $b^{-1}$, respectively. In fact, such a homeomorphism is obtained as the composition of the natural homeomorphisms \[ (S^3,K(r)) \to (S^3,K(-r))\to (S^3,K(1-r)), \] where the latter homeomorphism is explained in \cite[the end of Section~3]{lee_sakuma}. Thus, by \cite[Main Theorem~2.7]{lee_sakuma_2}, \cite[Main Theorems~2.2 and~2.3]{lee_sakuma_3} and the results in Sections ~\ref{sec:proof_of_main_theorem_1} and \ref{sec:proof_of_main_theorem_3}, we see that (3) or (4) holds. Moreover, the conjugacy diagram over $G(K(r))$ is obtained as the isomorphic image of the conjugacy diagram over $G(K(1-r))$. Since the image of a mixing vertex by the isomorphism is again a mixing vertex, we obtain the last assertion of the proposition. \end{proof} \begin{proof}{\it of Main Theorem~\ref{main_theorem} for the case when $r$ is general } Let $r$ be a general rational number with $0<r\le 1/2$, namely either $r=[m,m_2, \dots, m_k]$, where $m \ge 2$, $m_2 \ge 2$ and $k \ge 3$, or $r=[m, 1, m_3, \dots, m_k]$, where $m \ge 2$ and $k \ge 4$. Suppose on the contrary that there exist two distinct rational numbers $s, s' \in I_1(r) \cup I_2(r)$ such that $\alpha_s$ and $\alpha_{s'}$ are homotopic in $S^3-K(r)$, namely suppose that Hypothesis~A is satisfied. Let $\tilde{r}$ be as in \cite[Lemma~3.11]{lee_sakuma_2}. Then by Propositions~\ref{prop:induction_general_1} and~\ref{prop:induction_general_3}, there are two distinct rational numbers $\tilde{s}, \tilde{s}' \in I_1(\tilde{r}) \cup I_2(\tilde{r})$ such that the unoriented loops $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ are homotopic in $S^3-K(\tilde{r})$. Moreover, there is a reduced conjugacy diagram over $G(K(\tilde{r}))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$, such that none of the degree $4$ vertices is mixing. Regardless of the type of $r$, we put $\tilde{r}=[n_1, \dots, n_t]$, where $t \ge 2$, each $n_i \in \ZZ_+$ and $n_t\ge 2$. We proceed the proof by induction on $t \ge 2$. \medskip \noindent {\bf Case 1.} {\it $t=2$, i.e., $\tilde{r}=[n_1,n_2]$.} Then $\tilde r$ is special, and hence Proposition~\ref{prop:induction_base} shows that any reduced conjugacy diagram over $G(K(\tilde r))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ contains a mixing vertex. This contradicts to the fact observed in the above that there is a reduced conjugacy diagram over $G(K(\tilde r))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ which has no mixing vertex. \medskip \noindent {\bf Case 2.} {\it $t=3$, i.e., $\tilde{r}=[n_1,n_2,n_3]$.} If $n_1 \ge 2$ and $n_2 \ge 2$, then $\tilde{r}$ is general and the rational number $\tilde{\tilde{r}}=[n_2-1, n_3]$ is as in Case 1. So, this is impossible by the conclusion in Case 1. If $n_1 \ge 2$ and $n_2=1$, then $\tilde r$ is special and so Proposition~\ref{prop:induction_base} implies that any reduced conjugacy diagram over $G(K(\tilde r))$ for $\alpha_{\tilde{s}}$ and $\alpha_{\tilde{s}'}$ contains a mixing vertex, a contradiction. If $n_1=1$, then $\tilde{r}$ is special, because $1-\tilde{r}=[n_2+1,n_3]$ is special; so we obtain a similar contradiction by Proposition~\ref{prop:induction_base}. (To be precise, this rational number does not belong to the list in Proposition~\ref{prop:induction_base}, which is also a contradiction.) \medskip \noindent {\bf Case 3.} {\it $t=4$, i.e., $\tilde{r}=[n_1, n_2, n_3, n_4]$.} If $n_1 \ge 2$, then $\tilde{r}$ is general and the rational number $\tilde{\tilde{r}}$ is as in Case~1 or Case~2 according to whether $n_2=1$ or $n_2 \ge 2$. So, this is impossible by the conclusions in Cases~1 and 2. Hence we have $n_1=1$ and so let $\tilde{r}':=1-\tilde{r}=[n_2+1,n_3, n_4]$. If $n_3 \ge 2$, then $\tilde{r}'$ is general and the rational number $\tilde{\tilde{r}}'=[n_3-1,n_4]$ is as in Case~1. So, this is impossible by the conclusion in Case~1. If $n_3=1$, then $\tilde{r}$ is special, because $1-\tilde{r}=\tilde{r}'$ is special; so we obtain a contradiction by Proposition~\ref{prop:induction_base}. (In this case, $\tilde{r}$ is as in Proposition~\ref{prop:induction_base}(4).) \medskip \noindent {\bf Case 4.} {\it $t \ge 5$.} \medskip If $n_1\ge 2$, then $\tilde{r}$ is general, and the rational number $\tilde{\tilde{r}}$ is as in the case for $t-2$ or $t-1$ according to whether $n_2=1$ or $n_2\ge 2$. Thus this is impossible by the inductive hypothesis. Hence we have $n_1=1$ and so let $\tilde{r}':=1-\tilde{r}=[n_2+1,n_3,\cdots, n_t]$. Note that $\tilde{r}'$ is general and the rational number $\tilde{\tilde{r}}'$ is as in the case for $t-3$ or $t-2$ according to whether $n_3=1$ or $n_3\ge 2$. Thus this is impossible by the inductive hypothesis. Main Theorem~\ref{main_theorem} is now completely proved. \end{proof} \section{Proof of Theorems~\ref{main_corollary} and \ref{main_corollary2}} \label{sec:proof_of_main_corollary} Consider a hyperbolic $2$-bridge link $K(r)$ with $0<r \le 1/2$, and assume that the loop $\alpha_s$ with $s\in I_1(r) \cup I_2(r)$ is either peripheral or imprimitive. Then, by \cite[Lemma~7.2]{lee_sakuma_3}, there is a nontrivial element $w\in G(K(r))$ such that $w\not\in \langle u_s\rangle$ and $wu_sw^{-1}=u_s$. This identity cannot hold in $F(a,b)$, since $u_s$ is not a nontrivial cyclic permutation of itself. So by \cite[Lemma~7.1]{lee_sakuma_3}, the identity $wu_sw^{-1}=u_s$ in $G(K(r))$ is realized by a nontrivial reduced annular $R$-diagram, $M$, with outer and inner labels $u_s$ and $u_s^{-1}$, respectively. Then $M$ satisfies the assumption of \cite[Theorem~4.11]{lee_sakuma_2} and hence its conclusion. If $r=[2,n]$ or $[2,1,n]$ for some $n\ge 2$, then Theorems~\ref{main_corollary} and \ref{main_corollary2} are already proved in \cite[Section~7]{lee_sakuma_3}, where all possible diagrams $M$ are described. We note that we can observe that all such diagram contain a mixing vertex of degree $4$ (see \cite[Figures~13, 14(b), 15(b) and 16]{lee_sakuma_3}). If $r=[n,2]$ or $[n,1,2]$ with $n\ge 2$, then the link $K(r)$ is equivalent to the link whose slope is of the previous type, and so Theorems~\ref{main_corollary} and \ref{main_corollary2} in this case are deduced from the results in \cite[Section~7]{lee_sakuma_3}. If $r=[m,n]$ or $[m,1,n]$ with $m,n\ge 3$, then by the results in Sections~\ref{sec:proof_of_main_theorem_1} and~\ref{sec:proof_of_main_theorem_3}, we see that there are no such diagrams. So, any $\alpha_s$ with $s\in I_1(r) \cup I_2(r)$ is neither peripheral nor imprimitive in this case. Finally, suppose that $r$ is general. Then, by the proof of Propositions~\ref{prop:induction_general_1} and ~\ref{prop:induction_general_3}, we can construct from $M$ a non-trivial reduced conjugacy diagram $\tilde M$ over $G(K(\tilde r))$ with outer and inner labels $u_{\tilde s}$ and $u_{\tilde s}^{-1}$, respectively, for some $\tilde s\in I_1(\tilde r)\cup I_2(\tilde r)$, such that $\tilde M$ contains a mixing vertex of degree $4$. However, we can inductively show that this is impossible by using the preceding results. Hence, any $\alpha_s$ with $s\in I_1(r) \cup I_2(r)$ is neither peripheral nor imprimitive in this case. This completes the proof of Theorems~\ref{main_corollary} and \ref{main_corollary2}. \qed \bibstyle{plain} \bigskip
\section{Introduction} \label{sec:introduction} In their seminal paper \cite{MR84e:28003} Br\'ezis and Lieb prove a result about the decoupling of certain integral expressions, which has been used as one way to apply concentration compactness arguments in the calculus of variations. The alternative technical tool that enables concentration compactness arguments is the (measure theoretic) concentration function, in conjunction with cut-off procedures, as presented by Lions\cite{MR88e:35170,MR87e:49035a,MR87e:49035b}. To describe a special case of the Brezis-Lieb lemma, suppose that $\Omega$ is a domain in $\mathbb{R}^N$, $\mu >1$, $f(t) := \abs{t}^\mu $ for $t\in\mathbb{R}$, and $(u_n)$ a bounded sequence in $L^\mu (\Omega)$ that converges pointwise almost everywhere to some function $u$. If one denotes by $\mathcal{F}\colon L^\mu(\Omega)\to L^1(\Omega)$ the superposition operator induced by $f$, i.e., $\mathcal{F}(v)(x):= f(v(x))$, then $u\in L^\mu(\Omega)$ and \begin{equation} \label{eq:26} \mathcal{F}(u_n)-\mathcal{F}(u_n-u)\to\mathcal{F}(u) \qquad\text{in $L^1(\Omega)$, as $n\to\infty$.} \end{equation} The same conclusion is obtained in that paper for more general functions $f$, imposing conditions on $f$ that are satisfied for continuous convex functions with $f(0)=0$, and imposing additional conditions on the sequence $(u_n)$. If $\nu\ge1$ an analogous statement is also known to hold true for the map $\mathcal{F}\colon L^{\mu\nu}(\Omega) \to L^\nu(\Omega)$. Our aim is to give a similar result under a different set of hypotheses that applies to a much larger class of functions $f$ within a certain range of exponents $\mu$. In particular, we do not impose any convexity type assumptions on $f$ as was done in \cite{MR84e:28003}, nor any regularity assumptions that go further than continuity. A typical assumption of the latter type that has been used by many authors is to assume $f$ to be H{\"o}lder continuous with constants on compact intervals that have polynomially bounded growth in the correct sense. For one, even stronger, but typical, incarnation of this assumption see, e.g., \cite[Lemma~1.3]{MR1885519}. The price we pay for relaxing the hypotheses on $f$ is that we need to restrict the range of allowed growth exponents $\mu$ in comparison with \cite{MR84e:28003}, that we need to assume some type of translation invariance for $\Omega$ if it is unbounded, and that the decoupling result only applies to a smaller set of admissible sequences, namely sequences that converge weakly in $H^1(\Omega)$. Nevertheless, the numerous applications where these extra assumptions are satisfied justify the new set of hypotheses. To keep the presentation simple and highlight the main idea, we only treat the case $\Omega=\mathbb{R}^N$. It would be possible to consider other domains or superposition operators between other spaces, and we plan to do so in forthcoming work. Nevertheless, we do allow that $f$ depends on the space variable explicitly, but in a way that preserves the translation invariance. Also this condition could be weakened. In order to state our result we introduce some notation. Denote $2^*:=\infty$ if $N=1$ or $N=2$, and $2^*\coloneqq2N/(N-2)$ if $N\ge3$. Recall the continuous and locally compact embedding of the Sobolev space $H^1(\mathbb{R}^N)$ in $L^p(\mathbb{R}^N)$ for $p\in[2,2^*)$. By definition, the function $f\colon \mathbb{R}^N \times \mathbb{R}\to\mathbb{R}$ is a \emph{Caratheodory function} if $f$ is measurable and if $f(x,\cdot)$ is continuous for almost every $x\in\mathbb{R}^N$. The induced superposition operator on functions $u\colon \mathbb{R}^N\to\mathbb{R}$ is then given by $\mathcal{F}(u)(x) := f(x,u(x))$. If $A$ is a real invertible $N\times N$-matrix then $f$ is said to be $A$-periodic in its first argument if $f(x+Ak,t)=f(x,t)$ for all $x\in\mathbb{R}^N$, $k\in\mathbb{Z}^N$, and $t\in\mathbb{R}$. \begin{definition}\label{def:bl-splitting} Suppose that $X$ and $Y$ are Banach spaces. A map $\mathcal{F}\colon X\to Y$ will be called \emph{BL-splitting} (BL for Br\'ezis-Lieb) if \begin{equation*} \mathcal{F}(u_n)- \mathcal{F}(u_n-u) \to\mathcal{F}(u) \end{equation*} in $Y$ whenever $u_n\rightharpoonup u$ in $X$. \end{definition} With these preparations our main result reads: \begin{theorem} \label{thm:main} Consider $\mu>0$, $\nu\ge1$, and $C_0>0$, such that $p:=\mu\nu\in (2,2^*)$. Suppose that $f\colon\mathbb{R}^N\times\mathbb{R} \to\mathbb{R}$ is a Caratheodory function that satisfies \begin{equation} \label{eq:1} \abs{f(x,t)}\le C_0\abs{t}^\mu \qquad\text{for all } x\in\mathbb{R}^N,\ t\in\mathbb{R}, \end{equation} and which is $A$-periodic in its first argument, for some invertible matrix $A\in\mathbb{R}^{N\times N}$. Denote by $\mathcal{F} \colon L^p(\mathbb{R}^N)\to L^\nu(\mathbb{R}^N)$ the continuous superposition operator induced by $f$. Then $\mathcal{F}$ is uniformly continuous on bounded subsets of $H^1(\mathbb{R}^N)$ with respect to the $L^p$-$L^\nu$-norms and hence also with respect to the $H^1$-$L^\nu$-norms. Moreover, $\mathcal{F}\colon H^1(\mathbb{R}^N)\to L^\nu(\mathbb{R}^N)$ BL-splits. \end{theorem} \begin{remark} The result also holds in a slightly restricted sense for functions $f$ that are sums of functions as in Theorem~\ref{thm:main}, i.e., functions that satisfy merely \begin{equation*} \abs{f(x,t)}\le C_0(\abs{t}^{\mu_1}+\abs{t}^{\mu_2}) \qquad\text{for all } x\in\mathbb{R}^N,\ t\in\mathbb{R}, \end{equation*} where $\mu_i\nu\in(2,2^*)$ for $i=1,2$. In that case, $\mathcal{F}\colon H^1(\mathbb{R}^N)\to L^\nu(\mathbb{R}^N)$ is uniformly continuous on bounded subsets of $H^1(\mathbb{R}^N)$ with respect to the $H^1$-$L^\nu$ norms, and $\mathcal{F}$ BL-splits. \end{remark} \begin{remark} It seems that the uniform continuity of operators $\mathcal{F}$ on bounded subsets of $H^1(\mathbb{R}^N)$ is considered folklore in some circles. Nevertheless, we are not aware of a published proof of this fact, which is nontrivial in the generality stated in Theorem~\ref{thm:main}. Note that the uniform continuity of $\mathcal{F}\colon H^1(\Omega)\to L^\nu(\Omega)$ on bounded subsets of $H^1(\Omega)$ is trivial if $\Omega$ is bounded, by the compact Sobolev embedding $H^1(\Omega)\subseteq L^{\mu\nu}(\Omega)$. \end{remark} Theorem~\ref{thm:main} facilitates the application of the locally compact variant of the method of concentration compactness in nonlinear analysis. It works (in a certain exponent range) under weaker assumptions on the nonlinearity than known before and provides a convenient framework to avoid cumbersome cut-off arguments. Our proof of Theorem~\ref{thm:main} has similarities with the proof of \cite[Theorem~3.1]{MR2294665} but involves an intermediate cut-off step. Essentially, we first prove a weak form of BL-splitting as in \cite[Lemma~3.2]{MR2216902}, using the concentration function in the spirit of \cite{MR87e:49035a,MR87e:49035b,MR88e:35170} and the local compactness of the Sobolev embedding $H^1\hookrightarrow L^p$ for $p\in[2,2^*)$. Then we collect the possible mass loss at infinity along subsequences with the help of Lions' Vanishing Lemma, making use of the assumption $p>2$. In this way we combine the ideas from Br\'ezis-Lieb's Lemma and from Lions' approach, which are employed mutually exclusively by most authors. To give just one immediate improvement of a known result in elliptic PDEs, consider the equation \begin{equation}\label{eq:PL} -\Delta u +V(x)u=f(x,u)\qquad u\in H^1(\mathbb{R}^N)\;. \end{equation} We assume that $f$ and $V$ are $1$-periodic in all coordinates of $x$, $V$ is continuous, $V>0$, $f(x,\cdot)$ is continuously differentiable for almost all $x$, and $\partial_2f$ is a Caratheodory function. We assume the standard Ambrosetti-Rabinowitz condition for $f$ and $\partial_2f(x,t)t^2>f(x,t)t$ for all $x\in\mathbb{R}^N$ and $t\in\mathbb{R}$. Suppose that there are $2<p_1\le p_2<2^*$ such that \begin{equation} \abs{\partial_2^kf(x,t)} \le C(\abs{t}^{p_1-k-1}+\abs{t}^{p_2-k-1})\label{eq:28} \end{equation} for all $x\in\mathbb{R}^N$, $t\in\mathbb{R}$ and $k=0,1$. Then by Theorem~\ref{thm:main} the corresponding results in \cite{MR2216902} on the existence of multibump solutions of \eqref{eq:PL} hold true. In that paper we imposed \eqref{eq:28} also for $k=2$, assuming that $f$ is twice differentiable in $t$. We could have done there with an appropriate growth bound on the H\"older constant for $\partial_2f$. Theorem~\ref{thm:main} removes the need for any regularity conditions on $\partial_2f$ apart from simple continuity and the growth bound. \section{Proof of the Theorem} \label{sec:proof-theorem} For simplicity we will only consider the case $A=I$ (the identity transformation). Denote the respective translation action of the additive group $\mathbb{Z}^N$ on functions $u\colon\mathbb{R}^N\to\mathbb{R}$ by \begin{equation*} (a\star u)(x):= u(x-a),\qquad a\in\mathbb{Z}^N,\ x\in\mathbb{R}^N. \end{equation*} Let $B_R$ denote, for $R>0$, the open ball in $\mathbb{R}^N$ with center $0$ and radius $R$. If $r\in[1,\infty]$ and if $(\Omega,\Sigma,\mu)$ is a measure space with a positive measure $\mu$ then denote by $\abs{\,\cdot\,}_r$ the norm of $L^r(\mu)$. We omit $\mathbb{R}^N$ in the notation for function spaces. Let $\scp{\cdot,\cdot}$ denote the standard scalar product in $H^1$, defined by \begin{equation*} \scp{u,v}:=\int_{\mathbb{R}^N}(\nabla u\cdot\nabla v+uv), \end{equation*} and let $\norm{\,\cdot\,}$ denote the associated norm. Also denote by $\wlim$ the weak limit of a weakly convergent sequence. We first recall a functional consequence of Lions' Vanishing Lemma, \cite[Lemma~I.1.]{MR87e:49035b}. \begin{lemma} \label{lem:lions} Suppose for a sequence $(u_n)\subseteq H^1$ that $a_n\star u_n\rightharpoonup 0$ in $H^1$ for every sequence $(a_n)\subseteq\mathbb{Z}^N$. Then $u_n\to0$ in $L^p$ for all $p\in(2,2^*)$. \end{lemma} \begin{proof} Note first that $(u_n)$ is bounded in $H^1$ since $u_n\weakto0$ in $H^1$. We claim that \begin{equation} \label{eq:2} \sup_{y\in\mathbb{R}^N}\int_{y+B_1}\abs{u_n}^2\to0 \qquad\text{as }n\to\infty. \end{equation} If the claim were not true there would exist $\varepsilon>0$ and a sequence $(y_n)\subseteq\mathbb{R}^N$ such that, after passing to a subsequence of $(u_n)$, \begin{equation*} \int_{y_n+B_1}\abs{u_n}^2\ge\varepsilon. \end{equation*} Pick $(a_n)\subseteq\mathbb{Z}^N$ such that $\abs{a_n+y_n}_\infty<1$ for all $n$. With $R:=\sqrt{N}+1$ it follows that $a_n+y_n+B_1\subseteq B_R$ and hence \begin{equation*} \int_{B_R}\abs{a_n\star u_n}^2\ge\varepsilon \end{equation*} for all $n$. We reach a contradiction since $a_n\star u_n\weakto0$ in $H^1$ and hence $a_n\star u_n\to0$ in $L^2(B_R)$ by the theorem of Rellich and Kondrakov. Therefore \eqref{eq:2} holds. The claim of the theorem now follows from \cite[Lemma~I.1.]{MR87e:49035b} with $p=q=2$. Compare also with \cite[Lemma~3.3]{MR2294665}. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:main}] The continuity of $\mathcal{F}\colon L^p\to L^\nu$ follows from \eqref{eq:1} and the theory of superposition operators, see~\cite{MR1066204}. We start by proving the uniform continuity. Let $(u^0_{i,n})_{n\in\mathbb{N}_0}$ be bounded sequences in $H^1$ for $i=1,2$ and set $C_1:=\max_{i=1,2} \limsup_{n\to\infty}\norm{u^0_{i,n}}$. Suppose for a contradiction that \begin{equation} \abs{u^0_{1,n}-u^0_{2,n}}_p\to0 \qquad\text{as } n\to\infty,\label{eq:4} \end{equation} and that there is $ C_2>0$ such that \begin{equation} \abs{\mathcal{F}(u^0_{1,n})-\mathcal{F}(u^0_{2,n})}_\nu\ge C_2 \qquad\text{for all }n.\label{eq:6} \end{equation} Successively we will define a countable infinity, indexed by $k\in\mathbb{N}_0$, of sequences $(a^k_n)_n\subseteq\mathbb{Z}^N$ and $(u^k_{i,n})_n\subseteq H^1$, $i=1,2$, with the following properties (among others, to be seen below): \begin{align} \max_{i=1,2}\limsup_{n\to\infty}\norm{u^k_{i,n}}&\le C_1,\label{eq:5}\\ \lim_{n\to\infty}\abs{u^k_{1,n}-u^k_{2,n}}_p&=0,\label{eq:15}\\[.5ex] \liminf_{n\to\infty}\abs{\mathcal{F}(u^{k}_{1,n})-\mathcal{F}(u^{k}_{2,n})}_\nu &\ge C_2,\label{eq:11}\\ \intertext{and} \wlim_{n\to\infty}(-a^\ell_n)\star u^k_{i,n}&= 0 &&\qquad\text{in $H^1$, if $\ell<k$, for $i=1,2$.}\label{eq:20} \end{align} We need to say something about the extraction of subsequences. In order to obtain $(a^{k}_n)_n$ and $(u^{k+1}_{i,n})_n$ from $(u^k_{i,n})$, we first pass to a subsequence of $(u^k_{i,n})_n$ and then use its terms in the construction. Once the new sequences are built we may remove a finite number of terms at their start, with the goal of obtaining additional properties. Beginning with the following iteration there are no more retrospective changes to the sequences already built. The act of passing to subsequences will be implicit, leaving it to the reader to complete the argument, if so desired. For $k=0$ the properties \eqref{eq:5}--\eqref{eq:20} are fulfilled by the definition of $C_1$ and by \eqref{eq:4} and \eqref{eq:6}. Assume now that \eqref{eq:5}--\eqref{eq:20} hold for some $k\in\mathbb{N}_0$. Denote by $W_k$ the set of $v\in H^1$ such that there are a sequence $(a_n)\subseteq\mathbb{Z}^N$ and a subsequence of $(u^k_{1,n})$ with $\wlim_{n\to\infty}a_n\star u^k_{1,n}=v$ in $H^1$. If $\wlim_{n\to\infty}a_n\star u^k_{1,n}=0$ in $H^1$ were true for all sequences $(a_n)\subseteq\mathbb{Z}^N$, by Lemma~\ref{lem:lions} it would follow that $\lim_{n\to\infty} u^k_{1,n}=0$ in $L^p$. Equation~\eqref{eq:15} and the continuity of $\mathcal{F}$ on $L^p$ would lead to a contradiction with \eqref{eq:11}. Therefore \begin{equation*} q_k:= \sup_{v\in W_k} \norm{v}\in(0,C_1]. \end{equation*} Pick $v^k\in W_k$ such that \begin{equation} \label{eq:7} \norm{v^k}\ge \frac{q_k}{2}>0. \end{equation} There is $(a^k_n)_n\subseteq\mathbb{Z}^N$ such that $\wlim_{n\to\infty} (-a^k_n)\star u^k_{1,n}= v^k$ in $H^1$. Since $\lim_{n\to\infty} (-a^k_n)\star u^k_{1,n} = v^k$ in $L^p_\ensuremath{\mathrm{loc}}$ by the Rellich-Kondrakov theorem, we may assume by \eqref{eq:15} that \begin{equation} \wlim_{n\to\infty}(-a^k_n)\star u^k_{i,n}= v^k\qquad\text{in $H^1$, for } i=1,2,\label{eq:9} \end{equation} i.e., the same property for both indices $i=1,2$. Let us write $z^k_n:= (-a^k_n)\star u^k_{1,n}$. For $n\in\mathbb{N}_0$ define $Q_n\colon[0,\infty)\to[0,\infty)$ by \begin{equation*} Q_n(R):=\int_{B_R}\abs{z^k_n}^{p}. \end{equation*} The functions $Q_n$ are uniformly bounded and nondecreasing. We may assume that $(Q_n)$ converges pointwise almost everywhere to a bounded nondecreasing function $Q$ \cite{MR87e:49035a}. It is easy to build a sequence $R_n\to\infty$ such that for every $\varepsilon>0$ there is $R>0$, arbitrarily large, with \begin{equation*} \limsup_{n\to\infty}(Q_n(R_n)-Q_n(R))\le\varepsilon. \end{equation*} Hence \begin{equation}\label{eq:10} \forall\varepsilon>0\ \exists R>0\colon \limsup_{n\to\infty}\int_{B_{R_n}\backslash B_R}\abs{z^k_n}^p\le\varepsilon \qquad\text{and}\qquad \int_{\mathbb{R}^N\backslash B_R}\abs{v^k}^p\le\varepsilon. \end{equation} In view of \eqref{eq:15}, and taking $R$ large enough, \eqref{eq:10} also holds if we replace $z^k_n$ by $(-a^k_n)\star u^k_{2,n}$. Consider a smooth cut off function $\eta \colon[0,\infty) \to [0,1]$ such that $\eta\equiv1$ on $[0,1]$ and $\eta\equiv0$ on $[2,\infty)$. Set $v^k_n(x):=\eta(2\abs{x}/R_n)v^k(x)$. Then \begin{equation} \label{eq:12} \lim_{n\to\infty}v^k_n= v^k\qquad\text{in $H^1$.} \end{equation} From the continuity of $\mathcal{F}$ on $L^{p}(B_R)$, $v^k_n=v^k$ on $B_R$, and $\lim_{n\to\infty} z^k_n = v^k$ in $L^p(B_R)$ we obtain \begin{multline*} \lim_{n\to\infty}\int_{B_R} \bigabs{f(x,z^k_n)-f(x,z^k_n-v^k_n)-f(x,v^k_n)}^{\nu}\,\rmd x\\ =\lim_{n\to\infty}\int_{B_R} \bigabs{f(x,z^k_n)-f(x,z^k_n-v^k)-f(x,v^k)}^{\nu}\,\rmd x=0. \end{multline*} Since $v^k_n\equiv0$ in $\mathbb{R}^N\backslash B_{R_n}$, this in turn yields for any $\varepsilon>0$ and $R$ chosen accordingly, as in \eqref{eq:10}, \begin{multline*} \limsup_{n\to\infty}\int_{\mathbb{R}^N} \abs{f(x,z^k_n)-f(x,z^k_n-v^k_n)-f(x,v^k_n)}^{\nu}\,\rmd x\\ \begin{aligned} &= \limsup_{n\to\infty}\int_{B_{R_n}\backslash B_R} \abs{f(x,z^k_n)-f(x,z^k_n-v^k_n)-f(x,v^k_n)}^{\nu}\,\rmd x\\ &\le C\limsup_{n\to\infty}\int_{B_{R_n}\backslash B_R} (\abs{z^k_n}^\mu+\abs{z^k_n-v^k_n}^\mu+\abs{v^k_n}^\mu)^{\nu}\\ &\le C\limsup_{n\to\infty}\int_{B_{R_n}\backslash B_R} (\abs{z^k_n}^{p}+\abs{v^k}^{p})\\ &\le C\varepsilon, \end{aligned} \end{multline*} where $C$ is independent of $\varepsilon$. Letting $\varepsilon$ tend to $0$ and using \eqref{eq:12} we obtain \begin{equation*} \lim_{n\to\infty}\bigabs{\mathcal{F}(z^k_n)-\mathcal{F}(z^k_n-v^k_n)-\mathcal{F}(v^k)}_\nu= 0. \end{equation*} Since \eqref{eq:10} also holds for $(-a^k_n)\star u^k_{2,n}$ instead of $z^k_n$, the same arguments yield \begin{equation*} \lim_{n\to\infty}\bigabs{\mathcal{F}((-a^k_n)\star u^k_{2,n}) -\mathcal{F}((-a^k_n)\star u^k_{2,n}-v^k_n)-\mathcal{F}(v^k)}_\nu=0. \end{equation*} Set $u^{k+1}_{i,n}:= u^k_{i,n}-a^k_n\star v^k_n$. By the equivariance of $\mathcal{F}$ and the invariance of the involved norms under the $\mathbb{Z}^N$-action, \begin{align} \label{eq:14} \lim_{n\to\infty}\bigabs{\mathcal{F}(u^k_{i,n})-\mathcal{F}(u^{k+1}_{i,n}) -\mathcal{F}(a^k_n\star v^k)}_\nu&=0 \qquad\text{for $i=1,2$,}\\ \intertext{and, since $\norm{\,\cdot\,}^2$ BL-splits,} \label{eq:16} \lim_{n\to\infty}\bigabs{\norm{u^k_{i,n}}^2-\norm{u^{k+1}_{i,n}}^2 -\norm{v^k}^2}&=0\qquad\text{for $i=1,2$.} \end{align} Equations \eqref{eq:16} and \eqref{eq:5} (for $k$) imply that \begin{equation*} \max_{i=1,2}\limsup_{n\to\infty}\norm{u^{k+1}_{i,n}}\le C_1, \end{equation*} hence \eqref{eq:5} for $k+1$. The definition of the sequences $u^{k+1}_{i,n}$ and \eqref{eq:15} (for $k$) imply that \begin{equation*} \lim_{n\to\infty}\abs{u^{k+1}_{1,n}-u^{k+1}_{2,n}}_p =\lim_{n\to\infty}\abs{u^{k}_{1,n}-u^{k}_{2,n}}_p =0, \end{equation*} hence \eqref{eq:15} for $k+1$. It follows from \eqref{eq:14} and \eqref{eq:11} (for $k$) that \begin{equation*} \liminf_{n\to\infty}\abs{\mathcal{F}(u^{k+1}_{1,n})-\mathcal{F}(u^{k+1}_{2,n})}_\nu =\liminf_{n\to\infty}\abs{\mathcal{F}(u^{k}_{1,n})-\mathcal{F}(u^{k}_{2,n})}_\nu \ge C_2, \end{equation*} hence \eqref{eq:11} for $k+1$. Last but not least, from \eqref{eq:20} (for $k$), \eqref{eq:7}, and \eqref{eq:9} it follows that \begin{equation} \label{eq:21} \lim_{n\to\infty}\abs{a^\ell_n-a^k_n}=\infty\qquad\text{if } \ell<k, \end{equation} and hence by \eqref{eq:20} (for $k$) and \eqref{eq:12} that \begin{equation*} \wlim_{n\to\infty}(-a^\ell_n)\star u^{k+1}_{i,n} =\wlim_{n\to\infty}\biglr(){(-a^\ell_n)\star u^{k}_{i,n} -(a^k_n-a^\ell_n)\star v^k_n} = 0 \qquad\text{in $H^1$, if } \ell<k. \end{equation*} Moreover, by the definition of $a^k_n$, \begin{equation*} \wlim_{n\to\infty}(-a^k_n)\star u^{k+1}_{i,n} =\wlim_{n\to\infty}\biglr(){(-a^k_n)\star u^{k}_{i,n}-v^k_n} =0 \qquad\text{in $H^1$.} \end{equation*} This proves \eqref{eq:20} for $k+1$. We now modify the sequences $(a^k_n)$, $(v^k_n)$, and $(u^{k+1}_{i,n})$ we have just built by taking away a finite number of terms at their start. By \eqref{eq:15} and \eqref{eq:11} we may arrange it so that \begin{align} \abs{u^{k+1}_{1,n}-u^{k+1}_{2,n}}_p&\le\frac1{k+1}\label{eq:24}\\ \abs{\mathcal{F}(u^{k+1}_{1,n})-\mathcal{F}(u^{k+1}_{2,n})}_\nu &\ge C_2-\frac1{k+1}\label{eq:25} \end{align} for all $n\in\mathbb{N}_0$. Since $\norm{\,\cdot\,}^2$ BL-splits, \eqref{eq:12} and \eqref{eq:21} imply for any $\ell\le k+1$ that \begin{equation*} \lim_{n\to\infty}\biggnorm{\sum_{j=\ell}^{k+1}a^j_n\star v^j_n}^2 =\sum_{j=\ell}^{k+1}\norm{v^j}^2. \end{equation*} We may therefore arrange it so that \begin{equation} \label{eq:18} \biggnorm{\sum_{j=\ell}^{k+1}a^j_n\star v^j_n}^2 \le2\sum_{j=\ell}^{k+1}\norm{v^j}^2, \qquad\text{for all } n\text{ and }\ell\le k+1. \end{equation} Note that we need not and do not modify the sequences $(u^\ell_{i,n})$, $\ell\le k$, that were built in earlier steps. Now we consider the process of constructing sequences as finished and proceed to prove properties of the whole set. Equation~\eqref{eq:16} leads to \begin{equation*} \norm{u^{k+1}_{1,n}}^2 =\norm{u^0_{1,n}}^2-\sum_{j=0}^k\norm{v^j}^2+o(1) \qquad\text{as }n\to\infty, \end{equation*} and hence $\sum_{j=0}^\infty\norm{v^j}^2\le C_1$ by \eqref{eq:5}. In view of \eqref{eq:7} this yields \begin{equation} \label{eq:17} q_k\to0\qquad\text{as }k\to\infty. \end{equation} We claim that the diagonal sequence $(u^n_{1,n})$ satisfies \begin{equation} \label{eq:19} b_n\star u^n_{1,n}\weakto0\qquad\text{in $H^1$, as $n\to\infty$, for every sequence $(b_n)\subseteq\mathbb{Z}$.} \end{equation} Note that under our convention we have the representation \begin{equation} \label{eq:22} u^n_{1,n}=u^k_{1,n}-\sum_{j=k}^{n-1}a^j_n\star v^j_n \qquad\text{if } n\ge k. \end{equation} First we show that \begin{equation}\label{eq:23} \wlim_{n\to\infty}(-a^k_n)\star u^n_{1,n}=0 \qquad\text{in $H^1$, for all $k\in\mathbb{N}_0$.} \end{equation} Fix $k\in\mathbb{N}_0$. For every $w\in H^1$ and $\varepsilon>0$ there is $\ell_0$ such that \begin{equation*} \norm{w}^2\sum_{j=\ell_0}^\infty\norm{v^j}^2\le\varepsilon^2/2. \end{equation*} Then \eqref{eq:18} and \eqref{eq:22} yield for $n\ge\ell_0$ \begin{multline*} \bigabs{\bigscp{(-a^k_n)\star u^n_{1,n},w}}\\ \begin{aligned} &\le\bigabs{\bigscp{(-a^k_n)\star u^{k+1}_{1,n},w}} +\biggabs{\biggscp{\sum_{j=k+1}^{\ell_0-1}(a^j_n-a^k_n)\star v^j_n,w}} +\norm{w}\,\biggnorm{\sum_{j=\ell_0}^{n-1}a^j_n\star v^j_n}\\ &\le\bigabs{\bigscp{(-a^k_n)\star u^{k+1}_{1,n},w}} +\biggabs{\biggscp{\sum_{j=k+1}^{\ell_0-1}(a^j_n-a^k_n)\star v^j_n,w}} +\varepsilon. \end{aligned} \end{multline*} The first term in the last expression tends to $0$ as $n\to\infty$ by \eqref{eq:20}, and the second term tends to $0$ by \eqref{eq:12} and \eqref{eq:21}. Since $\varepsilon>0$ and $w\in H^1$ were arbitrary, this proves \eqref{eq:23}. To finish the proof of \eqref{eq:19}, suppose for a contradiction that $\wlim_{n\to\infty} b_n\star u^n_{1,n}= v\neq0$ in $H^1$, for a subsequence. Equation~\eqref{eq:23} implies that $\lim_{n\to\infty} \abs{b_n+a^k_n} =\infty$, for every $k\in\mathbb{N}_0$. Pick $k\in\mathbb{N}_0$ such that $q_{k}<\norm{v}$. This is possible by \eqref{eq:17}. Then, for every $w\in H^1$, it follows from \eqref{eq:18} and \eqref{eq:22} that \begin{equation*} \bigabs{\bigscp{b_n\star u^{k}_{1,n}-v,w}} \le\bigabs{\bigscp{b_n\star u^{n}_{1,n}-v,w}} +\biggabs{\biggscp{\sum_{j=k}^{n-1}(b_n+a^j_n)\star v^j_n,w}} \to0 \end{equation*} as $n\to\infty$, similarly as above. Hence $\wlim_{n\to\infty} b_n\star u^k_{1,n}= v$ with $\norm{v}>q_k$, in contradiction with the definition of $q_k$. This proves \eqref{eq:19}. We are now in the position to finish the proof of uniform continuity of $\mathcal{F}$. Equations \eqref{eq:24} and \eqref{eq:25} imply that \begin{gather*} \lim_{n\to\infty}\abs{u^n_{1,n}-u^n_{2,n}}_p=0\\ \liminf_{n\to\infty}\abs{\mathcal{F}(u^n_{1,n})-\mathcal{F}(u^n_{2,n})}_\nu\ge C_2. \end{gather*} By Lemma~\ref{lem:lions} and \eqref{eq:19} $u^n_{1,n}\to0$ in $L^p$. This contradicts the continuity of $\mathcal{F}$ on $L^p$ and therefore proves the assertion about uniform continuity. It only remains to prove BL-splitting for $\mathcal{F}$. Suppose that $u_n\rightharpoonup v$ in $H^1$. By the same arguments we used to obtain \eqref{eq:14} there is a sequence $(v_n)\subseteq H^1$ such that $v_n\to v$ in $H^1$ and \begin{equation}\label{eq:27} \mathcal{F}(u_n)-\mathcal{F}(u_n-v_n)\to\mathcal{F}(v) \qquad\text{in } L^\nu \end{equation} as $n\to\infty$. Since $(u_n)$ and $(v_n)$ are bounded in $H^1$, and by the uniform continuity of $\mathcal{F}$ on bounded subsets of $H^1$ with respect to the $L^p$-norm (and hence also with respect to the $H^1$-norm), it follows that we may replace $v_n$ by $v$ in \eqref{eq:27}. \end{proof} \begin{acknowledgement} I would like to thank Kyril Tintarev for an informative exchange on this subject. \end{acknowledgement} \bibliographystyle{amsplain-abbrv}
\subsection{Please Capitalize the First Letter of Each Notional Word in Subsection Title} \section{Introduction} \label{sect:intro} Solar filaments show some common characteristics in a broad spectrum of scales. Along the polarity inversion zone between adjacent opposite-polarity photospheric field on the quiet Sun, miniature filaments are the small-scale analog to large-scale ones. Hermans \& Martin (1986) concluded that their eruptions appear to be the counterparts of large-scale filament eruptions, which are usually associated with two-ribbon flares and coronal mass ejections (CMEs). Wang et al. (2000) have shown that miniature filaments have a mean lifetime of only 50 minutes, and their eruptions can take varying forms but almost all are accompanied by tiny flares, with spatial patterns very similar to two-ribbon/multiribbon flares in large-scale filament eruptions. Sakajiri et al. (2004) and Ren et al. (2008) also suggested that small- and large-scale filament eruptions have common properties. Therefore, it is reasonable to expect that some small-scale filament eruptions can also relate to CMEs (Innes et al. 2009; Schrijver 2010) and should be explained in the framework of the ``standard model'' for solar eruptions, in which overlying arcade erupts, producing a flare along with an ejecting filament. However, Wang et al. (2000) showed that most miniature filaments are erupted toward nearby strong network elements, meaning that perhaps most of their mass is transported to other magnetic structures rather than ejected into corona. Therefore, questions are naturally raised: is the mass transport along pre-existing magnetic force lines? Can small-scale filament eruptions be strong enough to escape into the heliosphere? Detailed observations of small-scale filament eruptions could help to answer these questions. Strongly related dynamical phenomena that can be closely associated with CMEs are plasma ejections following open or far-reaching field lines, which show up as surges in H$\alpha$ and jets in EUV and X-ray. H$\alpha$ surges are straight or slightly curved mass ejections stretching out and away from small flares at the footpoints in the chromosphere into coronal heights (Roy 1973; Bruzek \& Durrant 1977), and it is believed that they are produced by reconnection between emerging flux and ambient open or far-reaching magnetic structures (Schmieder, Van Driel-Gesztelyi, \& Freeland 1995; Canfield et al. 1996; Jiang et al. 2007; Madjarska et al. 2007; Chifor et al. 2008; Moreno-Insertis et al. 2008; Patsourakos et al. 2008; Pariat, Antiochos, \& DeVore 2010; Jiang et al. 2011; Yang et al. 2011). Therefore, surges have quite different physical processes and scenarios from those of small filament eruptions. However, some observations have showed that both of them can be related to the same kind of eruptive phenomena, specially ``narrow CMEs'' with angular widths of about 15\degr\ or less and impulsive solar energetic particle (SEP) events. Gilbert et al. (2001) concluded that 15 narrow CMEs originate in regions of closed magnetic fields since they have a high association with filaments, but Dobrzycka et al. (2003) cannot definitively determine whether they origin from jets or filament eruptions. Liu (2008) showed that large jetlike/diffuse H$\alpha$ surges are associated with jetlike/wide-angle CMEs. It is noted that, however, their H$\alpha$ data might have inadequate spatial resolution to tell us whether the surge bases include tiny filaments or not. Meanwhile, when white-light jets can be accompanied by either EUV bright points (Paraschiv et al. 2010) or filament eruptions (Wang \& Sheeley 2002), impulsive SEP events can also be closely associated with EUV and corresponding white-light jets that involve open field lines (Wang, Pick, \& Mason 2006), as well as with motions that look like filament eruptions (Nitta et al. 2006). In order to get a clear physical picture of these eruptive phenomena, it is clear that a definitive distinction of their different origins is desired. On the other hand, some observations indeed presented evidence that plasma ejections can be physically related with some kinds of activations or eruptions of small filaments. Chae et al.(1999) first reported that H$\alpha$ surges, EUV jets, and associated microflares were preceded by a small filament eruption. Zuccarello et al.(2007) showed that the destabilization and disappearance of a small short-lived filament followed two H$\alpha$ surges pouring sequentially from part of the filament body. Consistent with previous results (Yamauchi et al. (2005), in particular, Moore et al. (2010) found that in the polar coronal holes (CHs) two thirds of X-ray jets are standard ones that fit the standard reconnection picture for coronal jets, while other one third are so-called blowout jets in which the jet-base magnetic arch, often carrying a filament, undergoes a miniature version of the blowout eruptions that produce major CMEs. Similarly, Raouafi et al. (2010) showed that coronal jets can erupt from coronal micro-sigmoids, and Nistic\`o et al. (2009) also found some micro-CME-type jet events resembling the classical CMEs. Quite recently, Hong et al. (2011) and Zheng et al. (2011) indeed showed that miniature filament eruptions can lead to blowout jets and small CMEs. Therefore, at least part of plasma ejections are originated from small-scale filament eruptions, and high-resolution observations are needed to understand the possible relation between them or to distinguish them from each other. Similar to the cases in large-scale filament eruptions, in addition, a few examples clearly suggested that flux emergence and cancelation play an important role in disturbing small filaments (Hermans \& Martin 1986; Sakajiri et al. 2004; Ren et al. 2008; Hong et al. 2011). It thus appears that small filament eruptions have similar exterior driving agents to that of surges or jets (Yoshimura, Kurokawa, \& Shine 2003; Jiang et al. 2007). Like the cases occurring in some major flares (Wang et al. 2002b; Sudol \& Harvey 2005; Petrie \& Sudol 2010; Wang \& Liu 2010), an abrupt, significant, and persistent change in photospheric magnetic field was also found in a CME-associated H$\alpha$ mass ejecta event (Uddin et al. 2004; Li et al. 2005). Therefore, it is probable that such extreme change could also take place in small-scale filament eruptions, and detailed magnetic filed observations could help in verifying this possibility. On 2001 March 27, a surge-shaped eruption occurred in AR 9401 (N21\degr, E33\degr), and was associated with a compact impulsive flare, a CME, and especially, a sudden and permanent change of magnetic field and disappearance of a small pore. High-resolution H$\alpha$ observations from Big Bear Solar Observatory (BBSO) help us to find that it was in fact a small-scale filament eruption. Due to the short filament lifetime, the complete process from its formation to eruption was observed, hence provided us with the opportunity to investigate why the eruption took a surge form and its relationship with the CME and photospheric field activity. \section{Observations} \label{sect:Obs} The eruption was well covered by full-disk H$\alpha$ line-center observations from BBSO, with a 1-minute image cadence and a pixel size of roughly $1''$. The magnetic field configuration of the eruptive region was examined using full-disk magnetograms and continuum intensity images from the Michelson Doppler Imager (MDI; Scherrer et al. 1995) aboard the {\it Solar and Heliospheric Observatory} ({\it SOHO}). The magnetograms have a 1-minute cadence and a pixel size of $2^{''}$, while continuum images only have a 96-minute cadence. The MDI data are obtained from nine filtergrams taken in five different position in the Ni {\sc i} 6768 \AA\ absorption line in right- and left-circularly polarized light, and the continuum intensity is estimated from a filtergram averaging narrowband signals from both sides outside the line. The full-disk BBSO to MDI co-alignment was done by fitting the solar limbs, with an accuracy of about $2^{''}$. This eruption was also examined by using of full-disk EUV observations from the Extreme Ultraviolet Telescope (EIT; Delaboudini\`{e}re et al. 1995) on {\it SOHO}. The 12-minute cadence 195 \AA\ images were obtained for the study with a pixel resolution of $2.6^{''}$. To identify the associated CME, we used the C2 and C3 white-light coronagraph data from the Large Angle and Spectrometric Coronagraphs (LASCO; Brueckner et al. 1995) on {\it SOHO}, as well as the CME height-time data that is available in the LASCO Web site. Finally, we used soft X-ray light curves observed by the {\it Geostationary Operational Environmental Satellite} ({\it GOES}) to track the time of the flare. \section{Results} \label{sect:results} The surge-like solar eruption was accompanied by a flare of H$\alpha$ importance 1N and X-ray class M2.2, which {\it GOES} recorded its start, peak, and end times around 16:25, 16:30, and 16:32 UT, respectively. Fig. 1 presents BBSO H$\alpha$ line-center images to show the morphological evolution of the event. To aid matching, an MDI magnetogram is given in the first frame. The H$\alpha$ flare appeared as a compact one. It started to brighten slightly prior to the {\it GOES} flare start time, increased in size, flared through the period of the {\it GOES} flare, quickly faded away, and finally disappeared after about 16:40 UT. Therefore, the flare behaved as a compact, impulsive one with a total duration of only 7 minutes. A striking characteristic of the event is that a surge-like mass ejection occurred in the course of the flare, which showed up as extended linear structures with much larger size than that of the compact flare patch. When the {\it GOES} flare started (16:25 UT), a single bright structure (indicated by the white arrow) was first ejected from the flare patch. By the {\it GOES} flare peak time (16:30 UT), however, the bright structure clearly evolved into a narrow bright loop (indicated by the white arrow) with two legs, `1' and `2' (indicated by the two black arrows). The leg 2 quickly became ambiguous, while the leg 1 largely lengthened towards the northeastern direction and showed slightly curved triangular shape, thus remarkably displayed as a surge (Kurokawa \& Kawai 1993). Beginning at about 16:37 UT after the {\it GOES} flare ended, two dark components, also labeled as `1' and `2', grew nearly along the same trajectories of the two bright legs (indicated by the two black arrows). When the dark 1 gradually grew to occupy the whole path of the bright 1, the dark 2 underwent a larger development than that of the bright one. It is possible that the dark components were caused by the cooling of earlier, brighter components since they were co-spatial and sheared the same paths. Then the two dark components gradually faded away and became invisible after about 16:55 UT. \begin{figure} \centering \includegraphics[width=\textwidth, angle=0]{ms938fig1.eps} \caption{BBSO H$\alpha$ line-center images showing the evolution of the event. The first panel is an MDI magnetogram. Accompanied by a compact flare, the surge-like mass ejection with two leg-shaped features, `1' and `2', was originated from the eruption of a tiny filament, `F', at its base. The outlines of F's axis determined from the 16:11 UT H$\alpha$ image are plotted as solid curves. The field-of-view (FOV) is $190^{''}$ $\times$ $240^{''}$. The solid/dashed box indicates the FOV of Fig. 2/Fig. 5. } \label{fig1} \end{figure} Since H$\alpha$ surges can be entirely in emission and dark surges can be preceded by bright ones (\v{S}vestka, Farnik, \& Tang, 1990), at first glance, the above morphological characteristic of the eruption gives us an impression that this was a surge activity consisting of both bright and dark components, which occurred one after another, first bright then dark. Thanks to the high-resolution H$\alpha$ observations, however, we can remark by chance that there existed a tiny filament, `F', at the flare site before the eruption (see the 16:11 UT frame in Fig. 1). Moreover, we found that the legs 1 and 2 were coincided with those of the erupting F (see the 16:30 and 16:50 UT frames in Fig. 1). Therefore, it is very likely that the surge-like ejection was come from the F eruption. This motivates us to further investigate the F evolution in details. Fortunately, BBSO's observations covered the key process of the F formation and eruption. This is shown by the close-up view of F in Fig. 2. When BBSO's observations started at 15:25 UT, F was seen as a very small dark feature nearly residing above a magnetic polarity reversal boundary between adjacent opposite-polarity photospheric field. In less than 1 hr, it grew toward the west with a curved path along the polarity reversal boundary (indicated by the white arrows), and by about 16:11 UT, reached its maximum extent with a length of about 3.86 $\times$ $10^{4}$ km, about two times as long as miniature filaments with an average projected length of 1.9 $\times$ $10^{4}$ km (Wang et al. 2000). Therefore, F was a somewhat larger miniature filament. It is of interest to note that at this time it exhibited an inverse-S shape, consistent with the preferential pattern of dextral filament in the northern hemisphere (Pevtsov et al. 2003). But soon afterwards, its westernmost part began to disappear and two bright patches appeared at its two sides just before the {\it GOES} flare started (indicated by the black arrows in the 16:20 and 16:24 UT images). This is consistent with Kahler et al. (1988) that filaments began to erupt before the starting times of associated flares. Finally, the entire F quickly became invisible along with the area increase of the flare. Since F did not recover from such disappearance in the following 2h even when the flare completely faded away, we can conclude that it really underwent an overall violent eruption rather than was simply covered and obscured by the flare emission. It is noted that there were some filament-like dark features around F that were disturbed or even partially disappeared during the eruption. By carefully examining the erupting process, however, it is found that the legs 1 and 2 did not belong to them but just were the two legs of the erupting F. This is clearly indicated by the 16:30 and 16:50 UT H$\alpha$ images, where the outlines of F's axis at 16:11 UT were superposed as dashed lines. Therefore, the surge-like ejection was simply originated from the F eruption at its base in a narrow way: when its two ends were fixed in the chromosphere, its top quickly lifted up but its axis only had a little swelling or expanding perpendicular to the rising direction (also see Fig. 1). This mean that they were the same phenomenon, i.e., a filament eruption taking a surge form (Nistic\`o et al. 2009; Moore et al. 2010; Raouafi et al. 2010; Hong et al. 2011). \begin{figure} \centering \includegraphics[width=\textwidth, angle=0]{ms938fig2.eps} \caption{Close-up view of F in BBSO H$\alpha$ images. The first panel is an MDI magnetogram. The white arrows indicate that F grew toward the west, and the black arrows indicate the initial brightenings of the flare showing two-ribbon nature. The outlines of F's axis as in Fig. 1 are plotted, and its two legs are marked as `1' and `2'. The FOV is $60^{''}$ $\times$ $60^{''}$. } \label{f2} \end{figure} \begin{figure} \centering \includegraphics[width=50mm, angle=0]{ms938fig3.eps} \caption{Schematic depiction of the production of a blowout coronal jet from Moore et al. (2010). Reproduced by permission of the AAS. Only a few representative lines are drawn. Red field lines are those that have been reconnected with reconnection-heated X-ray plasma on them, while blue field lines either have not yet been reconnected or will not be reconnected. ({\it a}) Pre-eruptive magnetic field setup. A bipolar arch emerges in negative-polarity CH field. The dashed oval is the neutral line around the positive flux of the emerging arch, the short black curve represents the current sheet, and the field line arching low along the neutral line in the middle of the base arch indicates that the core field of the arch is extremely sheared. ({\it b}) Magnetic reconnection (marked by the X) at the location of the current sheet produces the small red loop on the west side and the red field line anchored on the east side. If the emerging arch has no sheared or twisted core field, the latter forms a standard X-ray jet and then the eruption would end shortly after this time. ({\it c}) If the core field of the emerging arch is strongly sheared and twisted, its blowout eruption leads to reconnection at several locations (marked by the Xs) that produces a blowout jet with a curtain-like structure, a red flare arcade in the east of the neutral line, and in additional brightening and growth of the loop in the west. ({\it d}) At the decay phase, the X-ray jet narrows and the entire field structure begins to relax down to the configuration of ({\it a}). } \label{f3} \end{figure} The schematic shown in Figure 3, a copy of Figure 10 in Moore et al. (2010), depicts the scenario for the productions of both standard and blowout jets. The magnetic field setup for both cases is the same that a compact, low-arching bipolar field emerges into a pre-existing, high-reaching, unipolar CH field with negative polarity, thus forms a neutral line surrounding the positive base of the emerging bipole, as well as a current sheet at the interface between the ambient field and its positive-leg (Fig. 3a). If the emerging-arch field has no appreciable shear or twist in it, magnetic reconnection at the current sheet creates a miniature-flare-arcade bight point over the west-side neutral line below the reconnection X-point, as well as a standard X-ray jet with a single-strand spire whipping out of the top side of the reconnection X-point (Fig. 3b). The eruption process then finishes, the field relaxes down to Fig. 3a configuration and the emerging arch to the east of the new reconnection loop in Fig. 3b remains stable. In the blowout jet case, however, an essential difference is that the core field of the emerging bipolar arch is so strongly sheared and twisted that it often shows up as a small filament and has enough free energy to drive a blowout of the arch: the emerging arch becomes destabilized by the earlier dynamics and thus triggers a blowout eruption of the sheared core field as in a filament eruption. As depicted in Fig. 3c, reconnection then occurs at three places that can result in a blowout X-ray jet with a multi-stranded curtain-like appearance (see Moore et al. (2010) for details of the blowout jet eruption process). Significantly, the blowout jet has a substantially larger horizontal size span than that of the standard jet. Finally, the jet narrows and the field configuration relaxes down to the initial state of Fig. 3a (Fig. 3d). It is difficult to make an one-to-one comparison between the above scenario and our event because no soft X-ray or high-resolution EUV imaging observations is available and the magnetic field environment is also different from that described by Moore et al. (2010) in coronal holes. However, two factors make us believe that our H$\alpha$ surge was clearly not a standard one. One is that it was originated from the small F eruption, and the other is that its 1 and 2 components quite resembled the curtain-like blowout jet rather than single-strand spire of Moore et al. (2010). Therefore, we can introduce a new term and call it ``blowout H$\alpha$ surge''. It is interesting to note that the 1 and 2 features in our event were very similar to a pair of thin ejections at the leg locations of a spicule-associated blowout jet from an eruption of relatively cool material in the event shown by Sterling, Harra, \& Moore (2010). Taking 16:20 UT as the start time of the F eruption, F had a lifetime of about 55 minutes from its first formation (15:25 UT) through the eruption. This was well consistent with the 50-minute mean lifetime of miniature erupting filaments given by Wang et al. (2000). Similar to a case studied by Sakajiri et al. (2004), in which a miniature filament erupted 25 minutes after its first formation, we see that there was an interval of only 9 minutes between the F eruption start and its full development (16:11 UT). It is clear that the flare was preceded by such an eruption. Although the flare at the main phase was too compact to discern its two ribbons and separating motion (possibly due to the small size of F), similar to the situation in compact flares observed by Tang (1985), the small H$\alpha$ brightenings at the initial F eruption phase clearly showed a two-ribbon nature. Consistent with some previous observations (Wang et al. 2000; Sakajiri et al. 2004; Ren et al. 2008; Hong et al. 2011), therefore, we believe that the F eruption was a small-scale version of active-region or quiet-region large-scale filament eruptions and the flare was a direct result of such a small-scale eruption. \begin{figure} \centering \includegraphics[width=\textwidth, angle=0]{ms938fig4.eps} \caption{EIT 195 \AA\ (panels {\it a1-a4}) and LASCO C2 (panels {\it b1-b4}) images. Panels {\it a1}, {\it b1}, and {\it b4} are original images, {\it a2-a4} are EIT 195 \AA\ fixed-base difference images obtained by subtracting a pre-flare image ({\it a1}), and {\it b2-b3} are C2 running difference images. The black arrows indicate the erupting F, and the white arrows the eruptive direction of the CME. The outlines of F's axis as in Fig. 1 are also plotted in {\it a1}, and the plus signs mark the pre-eruptive F location. The FOV for {\it a1-a4} is $810^{''}$ $\times$ $1030^{''}$, and the black boxes indicate the FOV of Fig. 1. } \label{f4} \end{figure} The eruption can be traced going beyond the northeast limb in EIT observations, and was closely associated with a CME observed by LASCO. This is quite obvious in the EIT 195 \AA\ and LASCO C2 images presented in Fig. 4. At 16:35 UT, although there was a thick, nearly straight, jet-shaped bright EUV feature responding to the H$\alpha$ surge-like bright structure (see Fig. 1) on the solar disk (included in the black box), the top part of the erupting F had indeed ejected beyond the solar limb and showed a curved loop-like shape (indicated by the black arrow). It can still be clearly discernible at 16:47 UT but then sprayed out the field-of-view (FOV) of EIT after 16:59 UT. Therefore, the EIT observation further confirmed that this was a filament eruption event. The associated CME first appeared in the LASCO C2 FOV at 17:06 UT as a ragged ejection in the northeastern direction that traveled along a thin streamer. It had a width of 66\degr\ and a central position angle (P.A.) of 46\degr. As indicated by the white arrows, the CME's eruption direction determined from the central P.A. was along the streamer and consistent with that of the erupting F. It is noteworthy that the streamer was only slightly altered after the passage of the CME (see panels b1 and b4), which was similar to the cases of streamer-puff and over-and-out CMEs recently studied by Bemporad et al.(2005), Moore \& Sterling (2007), and Jiang et al. (2009). The height-time (H-T) plot of the CME front at P.A. 38\degr\ is shown in Fig. 5. By applying second-order polynomial fitting, back extrapolation of the CME front from the H-T plots to the eruptive location yields an estimate of the CME onset time near 16:17 UT, which is very close to the start time of the F eruption at 16:20 UT, as well as the start time of the flare at 16:25 UT. The spatial and temporal consistencies strongly suggest that the CME initiation was closely related to the F eruption. In addition, the application of first- and second-order polynomial fitting to the H-T points gives an average speed of 342 km $s^{-1}$ and an acceleration of -6.2 m $s^{-2}$. These parameters showed that the CME was a slow one, consistent with the results of that slow CMEs are more often found associated with filament eruptions (Sheeley et al. 1999; St. Cyr et al. 2000). It is noted that the CME front speed is very close to the average F speed of about 344 km $s^{-1}$ estimated from 16:20 to 16:47 UT, and two H-T points of the erupting F at 16:35 and 16:47 UT plotted in Fig. 5 are also close to the extrapolated curves of the CME front, implying the tight relationship between the F eruption and the CME. From 16:25 to 16:31 UT, however, we can exactly determine the top sites of the erupting F in H$\alpha$ observations (see Fig. 1), and an average speed (acceleration) of about 151.8 km $s^{-1}$ (1.4m $s^{-2}$) is given. In accordance with the result of Zhang et al. (2001), these values indicate that F possibly underwent a large acceleration at its early eruption phase. As compared with the very small F size, however, we would like to point out that the CME had much larger scale. Therefore, the F eruption and the H$\alpha$ surge might only serve as a trigger and occupy a small part of the CME. Further observations are necessary to clarify such possibility. \begin{figure} \centering \includegraphics[width=10cm, angle=0]{ms938fig5.eps} \caption{Heights of the CME front as the function of time, and the back extrapolations by the use of second-order polynomial fitting. The vertical dashed bar indicates the {\it GOES} flare start time, and the solid bar the extrapolated CME onset time. } \label{f5} \end{figure} By further examining MDI observations, we found that the F eruption was correlated with a distinct photospheric activity below it. Fig. 6 shows the MDI magnetograms and continuum images around the eruptive region. At first view, it is quite clear that the area of a negative-polarity flux patch at the southern side of F greatly decreased after the eruption (indicated by the white arrows). More strikingly, a small pore just under the western part of F underwent a sudden disappearance (indicated by the black arrows). The pore was clearly seen at 15:59 UT before the eruption while it almost completely disappeared in the next available MDI intensity image at 17:35 UT. Such disappearance thus took place in a timescale of less than 100 minutes. Interestingly, it showed a $\delta$ configuration by making a comparison between the 15:59 UT magnetogram and continuum image. The flux changes in an area containing the eruptive F (indicated by the white contours) and in a smaller area only including the disappearing pore (indicated by the black dashed boxes) are measured and plotted in Fig. 7, and the {\it GOES-8} 1-8 \AA\ soft X-ray flux profile is also overplotted to indicate the flare time. After the flare start, we see that the negative flux in the contour (box) area had an abrupt decrease on the order of 4.5 $\times$ $10^{19}$ Mx (1.7 $\times$ $10^{19}$Mx) within the flare duration of only 7 minutes, i.e., impulsively dropped about 20\% (34.7\%), which gives an average flux loss rate of about 1.1 $\times$ $10^{17}$ Mx$s^{-1}$ (4.0 $\times$ $10^{16}$ Mx$s^{-1}$). In the following several hours, the change of the negative flux only showed a simple tendency of continuous decrease, and had no indication of a return to the pre-flare condition, indicating that such an impulsive flux loss was permanent, not transient, and thus was not due to the flare emission. We also note that the surrounding positive flux only changed a little relative to the negative flux during the flare. Therefore, the changes of opposite-polarity flux in the contour (box) area were not simultaneous and seemingly not balanced. \begin{figure} \centering \includegraphics[width=\textwidth, angle=0]{ms938fig6.eps} \caption{MDI magnetograms ({\it a}) and continuum images ({\it b}) around the eruptive filament. The white arrows indicate the negative patch with a large reduction in area during the flare, and the black arrows the vanishing small pore. The outlines of F's axis as in Fig. 1 are plotted, and the white contours and the black boxes indicate two areas, in which the changes of magnetic flux are measured and plotted in Fig. 6. The FOV is $90^{''}$ $\times$ $90^{''}$. } \label{f6} \end{figure} \begin{figure} \centering \includegraphics[width=10cm, angle=0]{ms938fig7.eps} \caption{Changes of magnetic flux in the contour ({\it a}) and the box ({\it b}) areas in Fig. 5, and time profiles of {\it GOES-8} 1-8 \AA\ soft X-ray, which are displayed in an arbitrary unit to fit in the panel. The plus/minus sign marks the positive/negative flux. To improve clarity, the absolute values for the negative flux are plotted, and the positive flux values in panel {\it a} are shifted 1 on the vertical scale. The vertical dashed bar indicates the {\it GOES} flare start time, and the solid bar the extrapolated CME onset time. } \label{f7} \end{figure} \section{Conclusions and discussion} \label{sect:Conclusion} By means of high-resolution observations, the event was identified as a surge-like eruption of the miniature filament F, i.e., a blowout surge, instead of a standard surge according to the following observations. (1) F divided opposite-polarity photospheric magnetic field around the AR's northern periphery. (2) The eruption was followed by a compact, impulsive flare. In the early eruption phase, however, the initial flare brightenings had two ribbons. (3) The evolution timescales of F, from its formation to full development through eventual eruption, was similar to those in previous studies. (4) The two legs of the erupting F were spatially consistent with the two ends of the pre-eruption F. Our observations further showed that the blowout surge was closely associated with the CME along a preexisting streamer. Therefore, these observations strongly indicated that the blowout surge, a filament eruption taking a surge form, was a small-scale version of the large-scale filament eruptions, with the exception of the small time and space scales (Nistic\`o et al. 2009; Moore et al. 2010; Raouafi et al. 2010; Hong et al. 2011). An important aspect of the event is that, despite of the very small size of F, its eruption and the following M2.2 flare were temporally and spatially relevant to a common characteristic of photospheric activity in major flares previously observed by some authors (Kosovichev \& Zharkova 1999; Wang et al. 2002b; Meunier \& Kosovichev 2003; Li, Ding, \& Liu 2005; Liu et al. 2005; Sudol \& Harvey 2005; Wang \& Liu 2010; Li et al. 2011), i.e., the sudden, significant, and persistent decrease in negative flux on the order of 4.5 $\times$ $10^{19}$ Mx and on a timescale of 7 minutes. As suggested by Sudol \& Harvey (2005), similar changes are ubiquitous features of X-class flares and most likely resulted from that the magnetic field changes direction rather than strength, suggesting that they are consequences rather than trigger of the flares. If so, the observed changes of magnetic field in our example were probably due to that the field lines were pulled or relaxed upward by the erupting F (also see Deng et al. 2005, Wang et al. 2005, and Hudson et al. 2008). As for the puzzling signature of changes of flux imbalance in two polarities, although some possible explanations were offered and discussed (Wang et al. 2002a; Wang et al. 2002b; Liu et al. 2003), we also believe that it was mainly associated with such changes in the magnetic field direction. Another rare phenomenon in this example is that during the rapid flux changes a small pore nearly disappeared in less than 100 minutes. Only one example of Wang et al. (2002a) has demonstrated that a flare can be associated with complete disappearance of a small sunspot in short period, thus our event gives another clear example of that similar pore disappearance can also occur in a blowout surge. Another key question concerning the event is why F erupted in a surge form? Since formation, maintenance, and instability of F should be controlled by its magnetic field environment (Martin 1990), we can tentatively consider the following three factors. First, the very small size of the F is a main element, which leads it to erupt in a narrow angular extent. We see that the erupting F showed a large rising height relative to the width between its two legs. Second, the eruption is strong enough to open its overlying coronal arcade and escape out into the heliosphere, which is indicated by the occurrence of the following flare and CME. This is clearly different from the cases observed by Wang et al. (2000), in which the most miniature erupting filaments are expelled to new locations. Note that all the miniature filaments of Wang et al. (2000) were on the quiet Sun with low magnetic flux density, while in our example F was located in a region with relatively high flux density. Therefore, it is naturally expected that a small change in the stronger magnetic field might be enough to influence the topology and stability of the F field in a violent manner and lead to a strong eruption (Gaizauskas 1989). Finally, there exists a guiding magnetic field with a twofold functions to channel the F eruption and restrain its lateral swelling. The CME along the streamer suggested that the F eruption from the streamer base should be guided by its open field component. This is very similar to that occurring in standard surges, which could also be ejected along either open magnetic fields or closed large loops. It is noteworthy that, however, the physical scenarios of standard and blowout surges are different. As shown in Fig. 3, magnetic reconnection only occurs between the guiding field and the closed magnetic field of new emerging flux in the standard surge case, while in the blowout surge case reconnection might occur at different places (Moore et al. 2010). Therefore, a clear distinction between standard and blowout surges by using of high-resolution observations is necessary to understand their physical natures and the origin of associated eruptive phenomena. We can imagine that if we only have H$\alpha$ observations with poorer spatial resolution, the blowout surge in this example will be readily regarded as a standard one, and thus might lead to misunderstand the origin of the associated CME. Even though the spatial resolution of our observations is high enough, the two phenomena are so easily mistakable for each other that we should differentiate them with great care. It is anticipated that observations from the {\it Solar Dynamics Observatory} (SDO) would be great beneficial in this problem. \begin{acknowledgements} We thank an anonymous referee for many constructive suggestions and thoughtful comments that helped to improve the clarity and quality of this paper. We are grateful to the observing staff at BBSO for making good observations. We thank the {\it SOHO}/MDI, EIT and LASCO teams for data support. This work is supported by the 973 Program (2011CB811400) and by the Natural Science Foundation of China under grants 10973038 and 11173058. \end{acknowledgements}
\section{Introduction} One of the most challenging problems in physics is the quantitative understanding of a quantum system of many strongly interacting particles for which numerical simulations have played an important role. Among the variety of known systems, a dilute Fermi gas near unitarity has been noted for its pure form, which can be readily studied from theory, and for its realization by using ultra-cold atomic experiments (for a recent review, see \cite{Giorgini, Inguscio} ). Besides the intrinsic physical interest, such as universal behavior, the unitary Fermi gas might also be considered as an ideal starting point to develop numerical techniques which can be applied to low energy nuclear physics, and to attack the noise problem which typically appears in numerical simulations of many particles. In 2010 we presented a highly improved lattice method for non-relativistic fermions with four-fermion contact interactions \cite{Endres} and its applications for trapped and untrapped unitary fermions \cite{JW, Nicholson}, where the largest numbers of fermions were restricted to $N=20$ and $38$, respectively, due to a statistical overlap problem. Since then, we have devised a tuning technique with a Galilean invariant form for the four-fermion interaction, which keeps the system at unitary up to Galilean boosts. We have also implemented an external harmonic potential in a more sophisticated way. Finally, we have developed a cumulant expansion technique to extract the ground state energies from data exhibiting a distribution overlap problem \cite{Noise, Trapped}. As a result, we are able to extend our calculation of ground state energies of unpolarized fermions up to $N=70$ and up to $N=66$ unitary fermions with and without a harmonic trap, respectively. For the untrapped case, this extension leads us to the regime, $N\geq 38$, where the ground state energy $E$ in units of that for non-interacting particles $E^{(0)}$\footnote{ The energies of non-interacting untrapped and trapped fermions are given by $E_{\textrm{untrapped}}^{(0)}=(3N)^{(5/3)}\pi^{4/3}/10ML^2$ and $E_{\textrm{trapped}}^{(0)}=(3N)^{4/3}\omega/4$, respectively. Here, $\omega$ is an oscillator frequency. } is constant; this implies that we are near the thermodynamic limit and can determine the universal dimensionless parameter $\xi=E/E^{(0)}$, called the Bertsch parameter. On the other hand, for the trapped case we find that $N=70$ is insufficient to reach the thermodynamic limit. \section{Lattice construction for fermions at unitarity} \label{sec:lattice_construction} We consider simulations of $N$ non-relativistic two-component fermions $\psi=(\psi^{\uparrow},\psi^{\downarrow})$ on a $T\times L^3$ Euclidean lattice, where periodic boundary conditions on the spatial directions and an open boundary condition on the temporal direction are imposed. The four-fermion contact interaction between different species of fermions is generated by $Z_2$ auxiliary fields $\phi$ which live on the time-like links of the lattice. A consequence of this set up is that the fermion-determinant does not depend on the $\phi$ field and thus the quenched simulation is exact \cite{Endres}. The discretization errors for the energy of a single fermion are eliminated through the use of a perfect dispersion relation, while the unitarity limit may be achieved by tuning the interaction strength $C({\bf q})$ to reproduce scattering data, $p\cot\delta_0=0$, using the L\"uscher formula \cite{Endres, Trapped}. Here the operator ${\bf q}$ for $|{\bf q}|<\Lambda$ corresponds to the momentum transfer between incoming and outgoing fermions, which makes the interaction Galilean invariant. The periodic function $C({\bf q})$ is defined by \beq C({\bf q})=\frac{4\pi}{M}\sum_{n=0}^{N_{O}-1} C_{2n} O_{2n}({\bf q}), \eeq where the basis functions are \begin{eqnarray} O_{2n}({\bf q}) = M_0^n\times \left\{ \begin{array}{ll} \left(1-e^{-{\bf q}^2/M_0}\right)^n & |{\bf q}|\leq\Lambda \\ \left(1-e^{-\Lambda^2/M_0}\right)^n & |{\bf q}|>\Lambda \end{array} , \right. \end{eqnarray} for ${\bf q}$ within the first Brillouin zone and periodic from one Brillouin zone to the next. Although the parameter $M_0$ may in general be different from the fermion mass $M$, we use $M_0=M$ in this work. The coefficients $C_{2n}$ are numerically determined by matching the lowest $N_{O}$ energy eigenvalues of the two-body transfer matrix in our lattice theory with the lowest $N_{O}$ L\"uscher energy eigenvalues in the corresponding continuum theory. In \reffig{fig1} we plot the $p\cot\delta_0$ (dots), computed using the exact lattice energy eigenvalues and L\"uscher's formula, which results from tuning the first $N_O$ terms in the effective range expansion to zero. These results show an expected $\eta$ scaling represented by the dashed lines in the figure. A systematic improvement is also seen for the energies of excited states above the lowest $N_{O}$ tuned states, where the correction may be predicted by \cite{Trapped}, \beq L\left(\frac{\eta_k}{\eta_k^*}-1\right)\propto L^{2-2n}. \label{energy_eta_scaling} \eeq Here $\eta_k$ are the $k$th eigenvalues of the two-body transfer matrix with $N_O$ terms tuned, while $\eta_k^*$ are the $k$th solutions of L\"uscher's formula in the unitary limit. As an example, we plot $L(\eta_k/\eta_k^*-1)$ with respect to $L$ (dots) in \reffig{fig2} and find good agreement with \refeq{energy_eta_scaling}. \begin{figure} \bmp[t]{.48\linewidth} \centering \includegraphics[width=1.0\textwidth]{ImpliedPcotDeltaLogLog.pdf} \caption{ln-ln plot of $p\cot\delta_0$ (dots) along with expected $\eta$ scaling (dashed lines) for which $N_O=1,2,3$ and $4$ coefficients are tuned by exactly matching the first $N_O$ energy eigenvalues for two particles in a box. Data is from an $L=32$ and $M=5$ lattice.}\label{fig1} \emp \hskip .2in \bmp[t]{.48\linewidth} \includegraphics[width=1.0\textwidth]{LevelNineVolDep.pdf} \caption{ $L$ dependence of the $9$th energy eigenvalues for two particles in a box. The dashed lines are fit results of the data (dots) using Eq. 2.3. } \label{fig2} \emp \end{figure} In simulations of trapped unitary fermions, the external harmonic potential, $U=\frac{1}{2}\kappa {\bf x}^2$ with spring constant $\kappa$, has been implemented in the transfer matrix as \beq T_{trapped}&=&e^{-b_\tau {\bf p}^2/4M} e^{-b_\tau U} (1-b_\tau V) e^{-b_\tau U} e^{-b_\tau {\bf p}^2/4M} \nonumber \\ &=&e^{-b_\tau (\mathcal{H}+U)+O(b_\tau^3)}, \eeq where $V$ represents the interaction and $\mathcal{H}+U$ is the target Hamiltonian for trapped unitary fermions. As seen in this formula, temporal discrerization errors appear at $O(b_\tau^2)$.\footnote{ Throughout this paper, $b_s$ and $b_\tau$ represent the spatial and temporal lattice spacings, respectively. } \section{Measurement, overlap problem, and cumulant expansion method} The $N$-body correlators are constructed by taking determinants of the Slater matrix, where each element is obtained by evolving an initial state (source) at $\tau=0$ with a single particle propagator and projecting onto a final state (sink) at $\tau=T$. The choice of sources and sinks for untrapped and trapped unitary fermions, which gives superior wave function overlap by imposing two-particle correlations at the sinks, has been described in \cite{JW} and \cite{Nicholson}, respectively. One of the greatest difficulties in extending our study to large $N$ was the apparent upward drift of the effective mass plot at large Euclidean time. By investigating the distribution of $N$-body correlators, we found that this difficulty arises from a heavily long-tailed distribution, requiring an exponentially large number of samples before the central limit theorem becomes applicable. In other words, the path-integral probability measure has small overlap with the dominant part of the operator being estimated. On the other hand, the distribution of the log of correlators is nearly Gaussian, implying that the standard estimation of the lower moments should succeed with moderately sized ensembles. A general relation between the expectation value of $\mathcal{C}$ and that of $\ln \mathcal{C}$ is given by \cite{Noise} \beq \ln \langle \mathcal{C}(\tau) \rangle = \sum_{n=1}^\infty \frac{\kappa_n}{n!}, \label{cum_exp} \eeq where $\kappa_k$ is the $k$th cumulant of $\ln \mathcal{C}$. The generalized effective mass and the ground state energy associated with each partial sum in \refeq{cum_exp} may be written as \cite{Trapped} \beq m_{\textrm{eff}}^{N_k}(\tau)=-\frac{1}{\Delta\tau} \sum_{n=1}^{N_k}\frac{1}{n!}\left[\kappa_n(\tau+\Delta\tau)-\kappa_n(\tau) \right] ~~~\textrm{and}~~~ E_{N_k}=\lim_{\tau\rightarrow\infty}m_{\textrm{eff}}^{N_k}(\tau). \label{gen_eff_mass} \eeq Since the statistical uncertainties typically increase as $N_k$ increases, one may determine the ideal value $N_k^*$ for which the statistical uncertainties and truncation errors are comparable. \section{Ground state energy of unitary fermions} To extract the energies of the system, we perform standard bootstrap resamplings and correlated $\chi^2$ fits to the plateau region of the generalized effective mass. Fitting systematic errors are obtained by varing the end points of the fitting interval. The quoted errors represent the combination of the statistical and fitting systematic errors added in quadrature. \subsection{Untrapped unitary fermions} As a nontrivial test of our lattice method, we have computed the lowest energy of three unitary fermions in a zero total momentum state with high precision. We performed the calculation for lattice sizes $L=8, 10, 12, 14, 16$, tuning the coefficients of four $O_{2n}$ operators for the $L=8$ lattice, and five for the other lattices. Since the single particle and two particle $s$-wave sectors have been highly tuned, the leading $L$ dependence will be $L^{-3}$ due to the untuned two-derivative two-body $p$-wave operator. In \reffig{fig3} we plot the lowest energies versus $L^{-3}$. We perform two parameter fits of the $L\geq10$ data to $c_1+c_2 /L^3$ and find an infinite volume energy of $0.3735^{+0.0014}_{-0.0007}$ in units of the energy of three noninteracting fermions, which agrees with the high precision calculation by Pricoupenko and Castin \cite{Castin} within our $\sim 0.3\%$ uncertainty. The preliminary results for the ground state energies of up to $66$ unpolarized unitary fermions in a periodic box are shown in \reffig{fig4}. With the given statistics, we do not resolve any shell structure in the energy for $N\geq 38$ and believe that the system is close to the thermodynamic limit. By averaging the results obtained from a constant fit of $\xi(N)$ for $N\geq 38$ using five different ensembles, we find a preliminary result for the Bertsch parameter, $\xi=0.399\pm 0.002$. The unitary Fermi gas has been extensively studied using the Quantum Monte Carlo (QMC) technique and the most recent calculation reported is $\xi\leq0.383 (3)$ \cite{qmc_recent_1,qmc_recent_2}. Recent experimental results for this parameter are $\xi=0.39 (2)$ \cite{Duke} and $0.41 (1)$ \cite{Paris}. \begin{figure} \bmp[t]{.48\linewidth} \centering \includegraphics[width=1.0\textwidth]{N3_extrapolation_inverseL_3.pdf} \caption{ Energy of three untrapped unitary fermions in a zero total momentum eigenstate plotted versus $(b_s/L)^3$ for $L/b_s=8,10,12,14,16$. The red band represents the uncertainty in two-parameter fits of the $L/b_s\geq10$ data to the function $c_1+c_2/L^3$. while the black line is the fit to the central values. The dashed line is the result from Ref. \cite{Castin}. } \label{fig3} \emp \hskip .2in \bmp[t]{.48\linewidth} \includegraphics[width=1.0\textwidth]{bertschplot.pdf} \caption{ (Preliminary) Plot of ground state energies versus numbers of untrapped unitary fermions for $L=10, 12$ and $14$. For $L=12$ we considered three different sinks, $\beta=0.5, 0.75, 1.0$, to take into account any source dependence in our extraction of the energies. The arrows indicate the positions for which noninteracting fermions have fully occupied a given momentum shell. } \label{fig4} \emp \end{figure} \subsection{Trapped unitary fermions} In numerical simulations of the trapped unitary Fermi gas, we have additional time scale $1/\omega$ and length scale $L_0=(M\omega)^{-1/2}$ which should be chosen so that $b_s\ll L_0 \ll L$ and $b_\tau \ll 1/\omega$ in order to minimize the discretization and finite volume errors. To balance the need for small temporal discretization errors with the computational cost associated with the number of time steps required to reach the ground state, we have chosen $\omega b_\tau=0.005$. An ideal $L_0$ has been determined by scanning the parameter space of $L_0/b_s$ and $L/L_0$ and finding the region where both finite volume and spatial discretization errors are small. \reffig{fig5} presents our findings for the ground state energies of $N\leq 6$ fermions, with $L_0/b_s$ ranging from $3$ to $8$ and fixed $L/b_s=48$. For $L_0/b_s\leq7$ we find that the systematic errors increase as $L_0$ decreases, which indicates that the discretization error is not negligible. For $L_0/b_s\geq7$ we have also performed simulations with $L/b_s=64$, which showed no volume dependence; and thus we conclude that both types of systematic errors are negligible. For $6<N\leq70$, based on the results of the $L_0$ scan for $N\leq 6$, we have chosen to perform the calculation for three volumes ($L=48,54,64$) and at two values of the trap size ($L_0=7.5, 8$). We find non-negligible volume dependence and perform an infinite volume extrapolation for each $L_0$. All of these considerations are included along with statistical, fitting systematic, and truncation of the cumulant expansion errors in the final quoted errors of ground state energies. We first benchmark our method for up to $N=6$ against high-precision solutions to the many-body Schr\"odinger equation \cite{Blume}, achieving agreement at $1\%$ as shown in \reftab{trapped_few_body}. In \reffig{fig6} we plot the results of the ground state energies for $N\leq 70$ along with the results from two fixed-node calculations for comparison: a Green's function Monte Carlo (GFMC) approach \cite{GFMC} and a diffusion Monte Carlo (FN-DMC) approach \cite{DMC}, which provide upper bounds on the ground state energies. We find that our energies are consistently lower than those obtained using both of these methods. However, our results show clear shell structure which indicates that $N\sim 70$ is insufficient to reach the thermodynamic limit. \begin{table} \caption{% \label{tab:SmallN}% Results for $E_\textrm{trapped}/\omega$ for $N \leq 6$, including combined statistical and fitting systematic errors (first row). For comparison we give the exact $N=3$ result \cite{Tan} and results of Ref. \cite{Blume} (second and third rows).} \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline & 3 & 4 & 5 & 6\\ \hline this work & $4.243^{+0.037}_{-0.034} $ & $5.071^{+0.032}_{-0.075} $ & $7.511^{+0.051}_{-0.091} $ & $8.339^{+0.080}_{-0.066} $ \\ exact, Ref. \cite{Tan} & 4.2727 & - & - & - \\ from Ref. \cite{Blume} & 4.273(2) & 5.008(1) & 7.458(10) & 8.358(20) \\ \hline \end{tabular} \end{center} \label{trapped_few_body} \end{table} \begin{figure} \bmp[t]{.48\linewidth} \centering \includegraphics[width=1.0\textwidth]{L0Scans.pdf} \caption{ Ground-state energies (in units of $\omega$) as a function of $L_0/b_s$ at fixed $L/b_s = 48$ for various values of N. Dashed lines are results from Ref. \cite{Blume}. }\label{fig5} \emp \hskip .2in \bmp[t]{.48\linewidth} \centering \includegraphics[width=1.0\textwidth]{BertschCompare2.pdf} \caption{ Ground-state energies of $N$-trapped unitary fermions. } \label{fig6} \emp \end{figure} \section{Conclusion} A new statistical technique, called the cumulant expansion method, allows us to extend our studies up to $N\sim 70$ unitary fermions both with and without an external harmonic trap. We report a preliminary value of $\xi=0.399\pm0.002$ from the calculation of the ground state energy for $N\geq 38$ untrapped unitary fermions. For $N\leq 70$ trapped fermions, we find that the shell structure is pronounced, which implies that we have not reached the thermodynamic limit, but our values of the ground state energies are consistently lower than those obtained from variational QMC calculations. In order to check the validity of our lattice method, we have performed high-precision calculations for $N=3$ untrapped fermions and $N\leq 6$ trapped fermions, and found good agreement with the results in \cite{Castin} and \cite{Tan, Blume}, respectively. In the near future, we will present results for the pairing gap by calculating the ground state energies of the slightly unpolarized unitary Fermi gas ($N_\downarrow=N_\uparrow+1$) and the integrated contact density by studying the dependence of the ground state energies on the $s$-wave scattering length near the unitary regime. Our lattice method combined with the new statistical technique is applicable for a wide range of nonrelativistic many-body systems. \section{Acknowledgement} This work was supported by U. S. Department of Energy grants DE-FG02-92ER40699 (M.G.E.) and DE-FG02-00ER41132 (D.B.K., J-W. L. and A.N.N.). M.G.E is supported by the Foreign Postdoctoral Researcher program at RIKEN. This research utilized resources at the New York Center for Computational Sciences at Stony Brook University/Brookhaven National Laboratory supported by the U.S. Department of Energy under Contract No. DE- AC02-98CH10886 and by the State of New York. Computations for this work were also carried out in part on facilities of the USQCD Collaboration, which are funded by the Office of Science of the U.S. Department of Energy.
\section{Introduction} \label{sec:intro} Matter accreted onto neutron stars (NSs) is piled up and compressed, settling towards regions of increasing density and temperature. In this process, depending on the rate at which accretion proceeds, both stable and unstable thermonuclear burning of the accreted H and He into heavier elements are expected \citep{Fujimoto81}. The main parameter thought to determine the different burning regimes is the mass accretion rate on the NS per unit surface area, $\dot{m}$ \citep[e.g.][]{Fujimoto81,Bildsten98}. When the burning layer becomes thermally unstable heat cannot be transported as fast as it is produced and a thermonuclear runaway occurs, producing a shell flash that releases $10^{38}$-$10^{39}$ erg in tens of seconds \citep[we do not discuss herein long bursts and superbursts, which are more energetic, $10^{40}$-$10^{42}$ erg, and much less common; e.g.,][]{Keek08}. Most of the energy is deposited in the outermost layers of the NS within a few seconds and radiated away thermally for tens of seconds while the photosphere cools down. Type I X-ray bursts, which feature the spectral imprint of such photospheric cooling, were discovered in low-mass X-ray binaries (LMXBs) more than 30 years ago \citep{Grindlay76,Belian76,Hoffman78} and promptly identified as thermonuclear bursts from accreting NSs \citep{Woosley76,Maraschi77,Lewin77}. \citet{Joss80} pointed out that a strong NS magnetic field can act to stabilize nuclear burning in different ways, which may explain the fact that no thermonuclear bursts have been observed to date from accreting NSs in high-mass X-ray binaries (HMXBs). \begin{figure*}[t!] \centering \resizebox{2.1\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f1.eps}}} \caption{ Overview of the burst and persistent emission evolution along the outburst of T5X2. Black lines show total 2~s time resolution light curves during one {\it RXTE} orbit for nine selected dates, as indicated (using PCU2 and the full $\sim$2--60~keV band; gray dashed line shows the approximate, not subtracted, background rate). Times have been shifted arbitrarily for display purposes. The increase in burst rate and decrease in burst brightness as the persistent flux rises is evident. The step in October 13 was produced by a lunar eclipse \citep{Strohmayer10b}. The source became unobservable for {\it RXTE} after November 19 due to Solar constraints. } \label{fig:evol} \end{figure*} Although direct observational evidence of stable thermonuclear burning on accreting NSs has been elusive, as this is outshined by the much more efficient accretion-powered ``persistent'' emission, theoretical arguments and indirect observational evidence \citep{Taam81,Fujimoto81,Paradijs88,Lewin93,Bildsten98} suggest that at very high $\dot{m}$, close to or above the Eddington mass accretion rate (Secs.~\ref{sec:data} \& \ref{sec:discussion}), thermonuclear burning of the accreted H and He proceeds only stably. Near the transition between unstable and stable burning, an oscillatory burning regime was predicted by \citet{Paczynski83}, known as marginally stable burning. \citet{Revnivtsev01} discovered millihertz quasi-periodic oscillations (mHz QPOs) in the X-ray flux of three atoll sources \citep[the sub-class of low luminosity NS-LMXBs;][]{HK89}: 4U~1636-536, 4U~1608-52 and Aql~X-1 \citep[see also][]{Strohmayer11}. They attributed this new phenomenon to marginally stable burning on the NS surface. The mHz QPO frequency in one of these systems has been found to decrease with time until a bright type I X-ray burst occurs \citep{Altamirano08d}. The persistent luminosity at which such mHz QPOs are observed has remained a puzzle, as it suggests a critical $\dot{m}$ about an order of magnitude lower than the stability boundary predicted by theory \citep[e.g.,][]{Heger07}. On 2010 October 10, an X-ray transient in the direction of the globular cluster Terzan~5 was discovered with the {\it International Gamma-ray Astrophysics Laboratory} \citep[][]{Bordas10,Chenevez10}. During the following week, {\it Rossi X-ray Timing Explorer (RXTE)} observations revealed 11~Hz pulsations \citep{Strohmayer10} and burst oscillations at the same frequency \citep{Altamirano10a,Cavecchi11}. The {\it Chandra} localization \citep{Pooley10} confirmed that this was a new NS transient, named IGR~J17480--2446 \citep[labeled CX25 or CXOGlb J174804.8--244648 by][]{Heinke06}. We refer hereinafter to IGR~J17480--2446 as T5X2, as this is the second bright X-ray source discovered in Terzan~5 (after EXO~1745--248). A 21.3~hr orbital period was measured from the Doppler shifts on the pulsar frequency \citep{Strohmayer10b,Papitto11}, and the NS magnetic field was estimated to be between 10$^{8}$--10$^{10}$~G based on the inferred magnetospheric radius \citep{Papitto11,Miller11}. This makes T5X2 the type I X-ray burst source (burster) with the slowest known NS spin and with the highest estimates of the NS magnetic field strength. Near the outburst peak T5X2 showed X-ray spectral and variability behavior typical of Z sources \citep[the sub-class of high luminosity NS-LMXBs;][]{HK89}, when it was accreting at about half of the Eddington rate \citep{Altamirano10b}. \citet{Linares10c} argued that all the X-ray bursts from T5X2 had a thermonuclear origin, based on the evolution of the burst rate. Given the lack of spectral softening along the tail of many of the bursts and their short recurrence times, \citet{Galloway10} suggested that some of the T5X2 bursts were type~II instead of type~I (i.e., accretion- instead of nuclear-powered). However, the persistent-to-burst energy ratio throughout the October--November outburst of T5X2 was typical of type~I X-ray bursts, i.e., fully consistent with the accretion-to-thermonuclear efficiency ratio \citep{Linares11b,Chakraborty11,Motta11}. Furthermore, \citet{Linares11b} measured a smooth evolution of the burst luminosity and spectral profiles and put forward a scenario to explain the lack of cooling in the faintest bursts, conclusively identifying all X-ray bursts detected from T5X2 as thermonuclear. We present a thorough analysis of the mHz QPOs from T5X2, including but not limited to the ones originally reported by \citet{Linares10c}. We study the mHz QPO frequency evolution and energy-dependent amplitude, as well as all X-ray bursts from T5X2 detected with {\it RXTE} while the persistent (accretion) luminosity varied along the outburst. Unlike previous studies \citep{Motta11,Chakraborty11}, we analyze the complete sample of {\it RXTE} bursts and compare their properties to theoretical models of thermonuclear burning, along the full range in persistent luminosity ($\sim$10--50\% of the Eddington luminosity). Section~\ref{sec:data} gives the details of the data analysis, and Section~\ref{sec:results} presents the main observational results: a smooth evolution between bursts and mHz QPOs (Figure~\ref{fig:evol}), the mHz QPO properties in detail and four different bursting regimes during the October-November 2010 outburst of T5X2. In Section~\ref{sec:discussion} we place the unique mHz QPO and bursting behavior of T5X2 in the framework of thermonuclear burning theory and discuss the possible effects of composition, NS spin and magnetic field on the observed bursting properties. Section \ref{sec:conclusions} gives our summary and conclusions. \section{Data analysis} \label{sec:data} We analyzed all {\it RXTE} observations of T5X2 during its October--December 2010 outburst: a total of 46 observations taken between 2010 October 13 and 2010 November 19 (proposal-target number 95437-01). The source became Sun constrained after that date, and was not detected by the {\it Monitor of All-sky X-ray Image (MAXI)} on 2010 December 28, indicating that the outburst finished between 2010 November 19 and 2010 December 28. We visually searched for X-ray bursts in the full dataset, using 2~s time resolution 2--30~keV lightcurves. We performed time-resolved spectroscopy of all bursts using high time resolution data (E\_125us\_64M\_0\_1s, or GoodXenon when available). We extracted dead-time corrected spectra in 2~s time bins, using a $\sim$100~s-long pre- or post-burst interval as background. We added a 1\% systematic error to all channels, grouped them to a minimum of 15 counts per channel when necessary and fitted the resulting spectra within Xspec (v 12.6.0q), using a simple blackbody model with the absorbing column density fixed to 1.2$\times$10$^{22}$ cm$^{-2}$ \citep{Heinke06}. We used a distance to T5X2 of 6.3~kpc, the highest value reported from {\it HST} photometry of Terzan~5 \citep{Ortolani07}, consistent with the distance measurement based on photospheric radius expansion bursts from another burster in the same globular cluster \citep{Galloway08}. We note, however, that recent estimates of this distance range between 4.6~kpc and 8.7~kpc \citep{Cohn02,Ortolani07,Lanzoni10}, and therefore any value of the luminosity, energy and mass accretion rate has a systematic uncertainty of a factor $\sim$3.6. We measured the burst rise time \citep[$t_\mathrm{rise}$, as defined in ][i.e., the time to go from 25\% to 90\% of the peak count rate]{Galloway08}, and the total radiated energy ($E_\mathrm{b}$) by integrating the bolometric luminosity along each burst. We defined the wait time, $t_\mathrm{wait}$, as the time elapsed between the peak of the previous burst and the peak of a given burst, available when no data gaps were present before the burst. We measured the daily-averaged burst recurrence time, $t_\mathrm{rec}$, as the total exposure time during one day (excluding those orbits where no bursts are detected) divided by the number of bursts detected on that day. Therefore, the instantaneous and daily-averaged burst rate, $\nu_\mathrm{burst}$, are given by $t_\mathrm{wait}^{-1}$ and $t_\mathrm{rec}^{-1}$, respectively. When only one burst was detected on a given day, we considered $t_\mathrm{rec}$ an approximate lower limit on the recurrence time. We also obtained daily averages of $E_\mathrm{b}$, peak burst luminosity ($L_\mathrm{peak}$), blackbody temperature (k$T_\mathrm{peak}$) and radius ($R_\mathrm{peak}$), following the same method described in \citet{Linares11b}: peak burst values correspond to a 4~s long interval around the burst peak. \begin{table*}[t!] \caption{Daily-averaged burst properties and persistent luminosity from T5X2.} \begin{minipage}{\textwidth} \begin{center} \begin{tabular}{ l c c c c c c c c} \hline\hline Date\footnote{MJD~55482 is 2010 October 13, and MJD~55519 is 2010 November 19.} & Exposure\footnote{Total daily exposure time in data segments (orbits) where bursts are detected. The total daily exposure time for segments without detected bursts is indicated between square brackets (whenever this time is larger than the burst recurrence time). Thus square brackets indicate periods of intrinsic burst cessation.} & Bursts\footnote{Total number of bursts detected per day. Square brackets indicate periods of intrinsic burst cessation, when no bursts were detected despite long enough exposure time.} & $t_\mathrm{rec}$\footnote{Daily-averaged burst recurrence time.} & $E_\mathrm{b}$\footnote{Bolometric integrated burst energy ($E_\mathrm{b}$) and peak burst luminosity ($L_\mathrm{peak}$). Persistent luminosity ($L_\mathrm{2-50}$) in the 2--50~keV energy band (see Sec.~\ref{sec:data} for bolometric correction). Peak burst blackbody radius ($R_\mathrm{peak}$) not color- or redshift-corrected. All use a distance of 6.3~kpc.} & $L_\mathrm{peak}$$^e$ & k$T_\mathrm{peak}$\footnote{Peak burst blackbody temperature.} & $R_\mathrm{peak}$$^e$ & $L_\mathrm{2-50}$$^e$ \\ (MJD) & (s) & & (s) & 10$^{39}$erg & 10$^{37}$erg/s & (keV) & (km) & 10$^{37}$erg/s \\ \hline 55482 & 3158 & 1 & 3158.0 & 1.60 $\pm$ 0.07 & 4.60 $\pm$ 0.57 & 2.27 $\pm$ 0.08 & 3.67 $\pm$ 0.23 & 2.20 $\pm$ 0.25 \\ 55483 & 20666 & 24 & 861.1 & 0.42 $\pm$ 0.01 & 2.00 $\pm$ 0.10 & 2.13 $\pm$ 0.03 & 2.70 $\pm$ 0.07 & 5.70 $\pm$ 0.28 \\ 55484 & 19709 & 42 & 469.3 & 0.28 $\pm$ 0.01 & 1.30 $\pm$ 0.07 & 2.01 $\pm$ 0.03 & 2.35 $\pm$ 0.07 & 6.60 $\pm$ 0.36 \\ 55485 & 16514 & 48 & 344.0 & 0.29 $\pm$ 0.01 & 1.30 $\pm$ 0.07 & 1.91 $\pm$ 0.02 & 2.69 $\pm$ 0.07 & 8.30 $\pm$ 0.39 \\ 55486 & [17334] & [0] & - & - & - & - & - & 9.55 $\pm$ 0.44 \\ 55487 & 14235 [6980] & 50 [0] & 284.7 & 0.36 $\pm$ 0.01 & 1.30 $\pm$ 0.08 & 1.72 $\pm$ 0.02 & 3.01 $\pm$ 0.10 & 10.00 $\pm$ 0.45 \\ 55488 & 11755 [3633] & 47 [0] & 250.1 & 0.37 $\pm$ 0.01 & 1.20 $\pm$ 0.07 & 1.69 $\pm$ 0.02 & 3.13 $\pm$ 0.09 & 9.20 $\pm$ 0.46 \\ 55489 & 3565 [3752] & 13 [0] & 274.2 & 0.31 $\pm$ 0.01 & 1.00 $\pm$ 0.11 & 1.77 $\pm$ 0.05 & 2.48 $\pm$ 0.14 & 8.40 $\pm$ 0.43 \\ 55490 & 7172 & 21 & 341.5 & 0.28 $\pm$ 0.01 & 1.10 $\pm$ 0.09 & 1.82 $\pm$ 0.03 & 2.65 $\pm$ 0.10 & 8.00 $\pm$ 0.38 \\ 55491 & 2410 & 6 & 401.7 & 0.23 $\pm$ 0.01 & 0.99 $\pm$ 0.16 & 1.60 $\pm$ 0.06 & 3.21 $\pm$ 0.26 & 7.10 $\pm$ 0.45 \\ 55492 & 6079 & 13 & 467.6 & 0.30 $\pm$ 0.01 & 1.30 $\pm$ 0.14 & 1.80 $\pm$ 0.05 & 2.75 $\pm$ 0.16 & 7.10 $\pm$ 0.37 \\ 55493 & 5780 & 10 & 578.0 & 0.24 $\pm$ 0.01 & 1.10 $\pm$ 0.13 & 1.59 $\pm$ 0.04 & 3.28 $\pm$ 0.20 & 6.50 $\pm$ 0.37 \\ 55494 & 5963 & 12 & 496.9 & 0.27 $\pm$ 0.01 & 1.20 $\pm$ 0.14 & 1.73 $\pm$ 0.05 & 2.95 $\pm$ 0.18 & 6.30 $\pm$ 0.37 \\ 55495 & 6685 & 10 & 668.5 & 0.51 $\pm$ 0.01 & 1.70 $\pm$ 0.11 & 1.90 $\pm$ 0.03 & 3.10 $\pm$ 0.11 & 6.20 $\pm$ 0.38 \\ 55496 & 6860 & 11 & 623.6 & 0.36 $\pm$ 0.01 & 1.30 $\pm$ 0.11 & 1.83 $\pm$ 0.04 & 2.79 $\pm$ 0.12 & 6.10 $\pm$ 0.36 \\ 55497 & 5964 & 8 & 745.5 & 0.37 $\pm$ 0.02 & 1.60 $\pm$ 0.15 & 1.97 $\pm$ 0.04 & 2.76 $\pm$ 0.13 & 5.90 $\pm$ 0.38 \\ 55498 & 5150 & 6 & 858.3 & 0.48 $\pm$ 0.02 & 1.60 $\pm$ 0.14 & 1.94 $\pm$ 0.04 & 2.95 $\pm$ 0.13 & 5.60 $\pm$ 0.42 \\ 55499 & 5364 & 7 & 766.3 & 0.35 $\pm$ 0.01 & 1.60 $\pm$ 0.13 & 1.96 $\pm$ 0.04 & 2.81 $\pm$ 0.11 & 5.50 $\pm$ 0.34 \\ 55500 & 5372 & 4 & 1343.0 & 0.74 $\pm$ 0.03 & 1.90 $\pm$ 0.17 & 1.98 $\pm$ 0.04 & 3.11 $\pm$ 0.14 & 5.50 $\pm$ 0.32 \\ 55501 & 5236 & 5 & 1047.2 & 0.79 $\pm$ 0.02 & 2.10 $\pm$ 0.16 & 2.02 $\pm$ 0.04 & 3.07 $\pm$ 0.11 & 5.50 $\pm$ 0.33 \\ 55502 & 6789 & 6 & 1131.5 & 0.87 $\pm$ 0.02 & 2.50 $\pm$ 0.15 & 2.06 $\pm$ 0.03 & 3.19 $\pm$ 0.10 & 5.20 $\pm$ 0.30 \\ 55503 & 5292 & 5 & 1058.4 & 0.74 $\pm$ 0.02 & 2.30 $\pm$ 0.16 & 2.11 $\pm$ 0.04 & 3.00 $\pm$ 0.10 & 5.30 $\pm$ 0.37 \\ 55504 & 6810 & 6 & 1135.0 & 0.57 $\pm$ 0.02 & 2.10 $\pm$ 0.14 & 2.15 $\pm$ 0.04 & 2.75 $\pm$ 0.09 & 5.20 $\pm$ 0.31 \\ 55505 & 5536 & 4 & 1384.0 & 0.74 $\pm$ 0.02 & 2.20 $\pm$ 0.18 & 2.12 $\pm$ 0.04 & 2.88 $\pm$ 0.12 & 5.20 $\pm$ 0.32 \\ 55506 & 6803 & 5 & 1360.6 & 0.93 $\pm$ 0.02 & 2.60 $\pm$ 0.17 & 2.14 $\pm$ 0.04 & 3.06 $\pm$ 0.10 & 5.10 $\pm$ 0.31 \\ 55508 & 6799 & 4 & 1699.8 & 1.00 $\pm$ 0.03 & 2.70 $\pm$ 0.19 & 2.18 $\pm$ 0.04 & 3.07 $\pm$ 0.10 & 4.90 $\pm$ 0.30 \\ 55509 & 6781 & 5 & 1356.2 & 1.06 $\pm$ 0.02 & 2.90 $\pm$ 0.16 & 2.21 $\pm$ 0.03 & 3.08 $\pm$ 0.08 & 4.70 $\pm$ 0.29 \\ 55510 & 4893 & 3 & 1631.0 & 1.02 $\pm$ 0.03 & 3.10 $\pm$ 0.23 & 2.08 $\pm$ 0.04 & 3.54 $\pm$ 0.13 & 4.70 $\pm$ 0.31 \\ 55511 & 6788 & 4 & 1697.0 & 1.26 $\pm$ 0.03 & 3.10 $\pm$ 0.20 & 2.07 $\pm$ 0.03 & 3.60 $\pm$ 0.11 & 4.60 $\pm$ 0.29 \\ 55512 & 5329 & 3 & 1776.3 & 1.28 $\pm$ 0.03 & 3.30 $\pm$ 0.21 & 2.17 $\pm$ 0.04 & 3.35 $\pm$ 0.11 & 4.60 $\pm$ 0.31 \\ 55513 & 5661 & 2 & 2830.5 & 1.37 $\pm$ 0.05 & 3.30 $\pm$ 0.37 & 2.30 $\pm$ 0.07 & 3.01 $\pm$ 0.17 & 4.60 $\pm$ 0.34 \\ 55514 & 6737 & 2 & 3368.5 & 1.46 $\pm$ 0.03 & 3.70 $\pm$ 0.26 & 2.25 $\pm$ 0.04 & 3.32 $\pm$ 0.12 & 4.40 $\pm$ 0.28 \\ 55515 & 10197 & 4 & 2549.2 & 1.33 $\pm$ 0.02 & 3.70 $\pm$ 0.18 & 2.23 $\pm$ 0.03 & 3.40 $\pm$ 0.08 & 4.30 $\pm$ 0.29 \\ 55516 & 6803 & 4 & 1700.8 & 1.37 $\pm$ 0.03 & 3.70 $\pm$ 0.25 & 2.16 $\pm$ 0.04 & 3.61 $\pm$ 0.12 & 4.20 $\pm$ 0.24 \\ 55517 & 3410 & 1 & 3410.0 & 1.42 $\pm$ 0.04 & 3.70 $\pm$ 0.29 & 2.21 $\pm$ 0.04 & 3.47 $\pm$ 0.13 & 4.00 $\pm$ 0.28 \\ 55518 & 3407 & 2 & 1703.5 & 1.22 $\pm$ 0.03 & 3.80 $\pm$ 0.26 & 2.31 $\pm$ 0.04 & 3.24 $\pm$ 0.11 & 4.00 $\pm$ 0.30 \\ 55519 & 6788 & 2 & 3394.0 & 1.45 $\pm$ 0.04 & 3.90 $\pm$ 0.34 & 2.19 $\pm$ 0.05 & 3.61 $\pm$ 0.16 & 3.80 $\pm$ 0.27 \\ \hline\hline \end{tabular} \end{center} \end{minipage} \label{table:dailybursts} \end{table*} Due to the smooth evolution from a series of bursts into a mHz QPO and vice versa (see Fig.~\ref{fig:evol} and Sec.~\ref{sec:results}) the distinction between ``frequent bursts'' and mHz QPOs is, to some extent, arbitrary. We searched for mHz QPOs all observations taken when the daily-averaged burst recurrence time was shorter than 350~s, which corresponds to 10 observations between MJDs~55485--55490 (around the outburst peak). Given the typical duration of a continuous {\it RXTE} observation segment (an ``orbit''), this threshold ensures that about 10 or more QPO cycles, or bursts, are observed without interruption. To study the mHz QPOs in those {\it RXTE} orbits we used 2--60 keV 1~s-bin light curves from all active PCUs combined. We then calculated a Lomb-Scargle periodogram \citep[LSP, oversampled by a factor of 3; ][]{Lomb76,Scargle82} for each light curve and measured the mHz QPO frequency, $\nu_\mathrm{QPO}$, as that frequency with the highest power in the periodogram. The corresponding period was then used to fold the background-corrected light curve (using background rates from {\it pcabackest} in 16-s steps) and produce a mHz QPO folded profile, from which we measured the fractional root-mean-squared (rms) amplitude. To investigate the energy dependence of the mHz QPO amplitude, we also produced light curves in five different energy bands (in keV: 2--3, 3--5.5, 5.5--9.5, 9.5--21 and 21--53), folded the light curve at the period found in the respective 2--60 keV range dataset and measured the ``rms spectrum'' of the mHz QPOs. We also performed 2048~s-long fast Fourier transforms (FFTs) using the 2--60~keV energy band and the same {\it RXTE} orbits, in order to constrain the mHz QPO coherence or ``quality factor''. \begin{figure}[h] \centering \resizebox{1.0\columnwidth}{!}{\rotatebox{0}{\includegraphics[]{f2.eps}}} \caption{From top to bottom panels, evolution along the outburst of {\it i)} persistent 2--50~keV luminosity; {\it ii)} burst rate as measured from $t_\mathrm{wait}$ (open gray circles), $t_\mathrm{rec}$ (filled black circles) and mHz QPO frequency (red filled squares; see Sec.~\ref{sec:data} for definitions) and {\it iii)} burst rise time and duration. Gray and black circles show individual burst measurements and daily averages, respectively. Open triangles show burst rate daily averages based on one single burst, which we consider as upper limits.} \label{fig:ob} \end{figure} In order to measure the persistent (accretion) luminosity, we extracted one dead-time-corrected spectrum per observation from Standard~2 data, excluding all bursts and subtracting the background spectrum estimated with the bright source background model and {\it pcabackest} (v. 3.8). We then fitted each persistent spectrum with a model consisting of a disk blackbody, a power law and a Gaussian line with energy fixed at 6.5~keV, correcting for absorption as above. We calculated the 2--50~keV persistent luminosity ($L_\mathrm{2-50}$) from the best fit model. Furthermore, we measured the 0.01--2~keV unabsorbed flux from a simultaneous fit to the T5X2 spectrum measured by {\it Swift}-XRT (0.5-10~keV) and {\it RXTE}-PCA (2.5-25~keV) on MJD 55501 (2010 November 01), extrapolating the {\it phabs*simpl(bbody + diskbb)} best fit model down to 0.01~keV, and found a bolometric correction factor of 1.13. This bolometric correction factor converts 2--50~keV into 0.01--50~keV flux, which we take as bolometric flux given that the persistent spectrum remains soft (photon index 2.4--3.3) throughout the outburst and we therefore do not expect sizeable emission above 50~keV (with the only exception of the first observation, when T5X2 was in the hard state and the photon index was $\sim$1.7). Using the same procedure, we found a bolometric correction factor of 1.02 for the 0.5--50~keV luminosity from Cir~X-1 reported by \citet[][used in Sec.~\ref{sec:discussion}]{Linares10d}. We applied these bolometric correction factors in order to estimate the bolometric persistent luminosity: $L_\mathrm{pers}$. We use throughout this work an Eddington luminosity of $L_\mathrm{Edd}$=$2.5\times 10^{38}$ erg s$^{-1}$ to calculate Eddington-normalized $L_\mathrm{pers}$. To convert from $L_{pers}$ to $\dot{m}$, we assume homogeneous accretion onto the NS and that the (general-relativistic) gravitational energy of the in-falling matter is fully converted into radiation at the surface of a 1.4~$M_\odot$ mass, 10~km radius NS. We define the Eddington mass accretion rate per unit area, $\dot{m}_\mathrm{Edd}$, as the mass accretion rate needed to sustain a luminosity equal to $L_\mathrm{Edd}$. With this definition $\dot{m}_\mathrm{Edd}$=$1.2\times 10^5$~g~cm$^{-2}$~s$^{-1}$. \begin{table}[b] \tiny \caption{Properties of the mHz QPOs from T5X2.} \footnotetext{For comparison, the reported values for the ``low-$L_\mathrm{pers}$ mHz QPOs'' are in the following ranges: $\nu_\mathrm{QPO}$=[7--14.3]~mHz; rms$_\mathrm{QPO}$=[0.7--1.9]\% (2--5~keV); $L_\mathrm{pers}$=[0.6--3.5]$\times$10$^{37}$erg/s \citep[Sec.~\ref{sec:msb};][]{Revnivtsev01,Altamirano08d}.} \footnotetext{\footnotemark{}Observation (from proposal-target 95437-01) where the mHz QPOs are detected, orbit numbers used indicated between square brackets.} \begin{minipage}{\textwidth} \begin{tabular}{ l l c c c c} \hline\hline Date & OBSID\footnotemark{} & $\nu_\mathrm{QPO}$ & rms$_\mathrm{QPO}$ & $L_\mathrm{pers}$/10$^{37}$ & $\dot{m}$/10$^{4}$ \\ (MJD) & & (mHz) & (\%) & (erg/s) & (g~cm$^{-2}$s$^{-1}$) \\ \hline 55485.46 & 04-00 [1] & 2.75 $\pm$ 0.2 & 1.9 $\pm$ 0.1 & 9.1 $\pm$ 0.4 & 4.5 $\pm$ 0.2 \\ 55485.63 & 04-01 [1,2,3] & 3 $\pm$ 0.5 & 2.2 $\pm$ 0.1 & 9.9 $\pm$ 0.4 & 4.9 $\pm$ 0.2 \\ 55487.43 & 06-000 [1,4] & 4.2 $\pm$ 0.2 & 1.3 $\pm$ 0.1 & 11.4 $\pm$ 0.4 & 5.6 $\pm$ 0.2 \\ 55487.62 & 06-00 [1,2] & 4.2 $\pm$ 0.2 & 1.7 $\pm$ 0.2 & 12.1 $\pm$ 0.5 & 6.0 $\pm$ 0.2 \\ 55488.26 & 07-00 [1] & 4.0 $\pm$ 0.5 & 1.4 $\pm$ 0.1 & 12.0 $\pm$ 0.5 & 5.9 $\pm$ 0.2 \\ 55489.59 & 08-00 [1] & 3.75 $\pm$ 0.2 & 2.1 $\pm$ 0.1 & 9.6 $\pm$ 0.5 & 4.8 $\pm$ 0.2 \\ 55490.63 & 09-00 [1,2] & 2.9 $\pm$ 0.2 & 2.2 $\pm$ 0.1 & 9.1 $\pm$ 0.4 & 4.5 $\pm$ 0.2 \\ \hline\hline \end{tabular} \end{minipage} \label{table:mhzqpo} \end{table} \section{Results} \label{sec:results} We found and analyzed a total of 398 X-ray bursts occurring between 2010 October 13 and 2010 November 19, including the faint and frequent bursts near the peak of the outburst that form the mHz QPOs (see below). We show in Figure~\ref{fig:evol} an overview of the joint evolution of persistent emission and burst properties along the T5X2 outburst. Bursts become more frequent and fainter when the persistent luminosity, $L_\mathrm{pers}$, increases, and they turn into brighter and more frequent bursts during the outburst decay. Most interestingly, the X-ray bursts gradually develop into the observed mHz QPO during the outburst rise, and the mHz QPO mutates into a series of bursts along the outburst decay (Figure~\ref{fig:evol}), an unprecedented phenomenon among thermonuclear bursters. This also makes T5X2 the most prolific source of thermonuclear bursts known to date (an average burst rate over more than a month of 4.9 hr$^{-1}$; 5.5 hr$^{-1}$ excluding orbits where no bursts were detected), with the shortest recurrence times between thermonuclear bursts observed to date \citep[as short as $\sim$200~s; c.f.][]{Linares09c,Keek10}. The {\it smooth metamorphosis} between bursts and mHz QPOs can be seen qualitatively in Figure~\ref{fig:evol}, and quantitatively by studying the evolution of several burst properties (Figure~\ref{fig:ob} \& Table~\ref{table:dailybursts}). Burst rate, rise time and duration as well as peak burst luminosity, $L_\mathrm{peak}$, and total radiated energy $E_\mathrm{b}$; \citep[see also][]{Linares11b}, all evolve gradually while $L_\mathrm{pers}$ changes by a factor of $\sim$5 along the outburst. Figure~\ref{fig:ob} (middle panel) also shows that the burst rate equals the mHz QPO frequency during the outburst peak, $\nu_\mathrm{burst}$=$\nu_\mathrm{QPO}$, as expected given that the mHz QPOs are simply formed by a series of faint and frequent bursts. This unique behavior makes the distinction between bursts and mHz QPOs somewhat arbitrary. As explained in Section~\ref{sec:data}, our practical definition of mHz QPO requires burst recurrence times shorter than 350~s, so that typically 10 or more QPO cycles are observed without interruption. In a few occasions we find that one burst is missing from the series of regular bursts, and the corresponding values of $\nu_\mathrm{burst} = t_\mathrm{wait}^{-1}$ for individual bursts are a factor $\sim$2 lower than the general trend, as can be seen in Figure~\ref{fig:ob} (middle panel). Interestingly, a similar behavior with sporadic ``missing bursts'' is seen in the mHz QPO simulations presented in \citet{Heger07}. \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f3.eps}}} \caption{Light curves of the mHz QPOs on October 16 (gray) and 18 (black), in the 2--60~keV energy band, using 1-s time bins and no background subtraction. One {\it RXTE} orbit is shown in each case, and time is relative to the start of that orbit (2010-10-16 14:33:27 UTC and 2010-10-18 15:09:55 UTC). The persistent luminosity was higher on October 18 ($L_\mathrm{pers}$$\simeq$0.45$L_\mathrm{Edd}$) than October 16 ($L_\mathrm{pers}$$\simeq$0.38$L_\mathrm{Edd}$). To illustrate this, count rate is normalized to the number of active PCUs.} \label{fig:qpoLC} \end{center} \end{figure} \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f4.eps}}} \caption{Lomb-Scargle periodograms of the mHz QPOs on October 16 (gray) and 18 (black), showing the QPO harmonic structure and the change in $\nu_\mathrm{QPO}$.} \label{fig:qpoLS} \end{center} \end{figure} \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f5.eps}}} \caption{Folded background-subtracted 2--60~keV light curves of the mHz QPOs on October 16 (gray) and 18 (black). Peaks are at phase 0 by definition. Two QPO cycles are shown for clarity. Two PCUs were active on October 16, yielding a higher collected rate. The dashed horizontal line shows the average count rate, and the vertical range corresponds to 10\% of that value in each case, showing that the fractional amplitude was higher on October 16 (Table~\ref{table:mhzqpo}). The net peak burst luminosity was in both cases $\simeq$1.3$\times$10$^{37}$ erg~s$^{-1}$ (Table~\ref{table:dailybursts}).} \label{fig:qpoFP} \end{center} \end{figure} \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f6.eps}}} \caption{``RMS spectrum'' of the mHz QPOs on October 16 (gray) and 18 (black). Their fractional rms amplitude increases with energy up to $\sim$20~keV. No mHz QPOs are detected above that energy, empty triangles show upper limits on their fractional rms amplitude.} \label{fig:qpoRE} \end{center} \end{figure} \subsection{mHz QPOs} \label{sec:mhzqpo} We report the discovery of several instances of mHz QPOs during the peak of the T5X2 outburst, on 2010 October 16, 18, 19, 20 and 21 \citep[MJDs~55485--55490;][for the initial report of mHz QPOs on 2010 October 18 and 19]{Linares10c}. Two examples of mHz QPO light curves are shown in Figure~\ref{fig:qpoLC}, each spanning one {\it RXTE} orbit. Table~\ref{table:mhzqpo} shows the main QPO properties: fractional rms amplitudes between 1.3\% and 2.2\% (in the 2--60~keV band) and $\nu_\mathrm{QPO}$ in the range 2.8--4.2~mHz (note that the lower end of this frequency range corresponds to our mHz QPO definition, Secs.~\ref{sec:data} \& \ref{sec:results}). Figure~\ref{fig:qpoLS} presents LSPs of two cases, on October 16 and 18, clearly showing the change in $\nu_\mathrm{QPO}$ as well as the harmonic structure (up to 4 overtones are visible in the LSPs). By inspecting the power spectra obtained from 2048-s-long FFTs, we find that the mHz QPO power is in all cases spread over one or two frequency bins, from which we derive a limit on the full-width-at-half-maximum, FWHM$\lesssim$1~mHz. For the 2.8--4.2~mHz QPO frequencies this corresponds to a lower limit on the quality factor (Q$\equiv$$\nu_\mathrm{QPO}$/FWHM) of Q$\gtrsim$3, which reveals a fairly coherent QPO (a fairly constant $t_\mathrm{wait}$). We present in Figure~\ref{fig:qpoFP} light curves folded at the mHz QPO period, showing folded burst/QPO profiles for the same two dates, October 16 and 18. In both cases the burst/QPO profile is peaked and highly non-sinusoidal (as seen also in the raw, unfolded light curves; Fig.~\ref{fig:qpoLC}), which explains the harmonic content. On October 18, when $\nu_\mathrm{QPO}$ was highest (Table~\ref{table:mhzqpo}), we find a nearly symmetric burst/QPO profile. On October 16 the burst/QPO profile is slightly asymmetric, with the rise faster than the decay. We measure a mHz QPO fractional rms amplitude in the range 1.3--2.2 \%, in the total ($\sim$2--60~keV) PCA band. Furthermore, from our measurements of the mHz QPO amplitude at different energies (Sec.~\ref{sec:data}; Fig.~\ref{fig:qpoRE}) we find that its fractional rms amplitude increases between $\sim$2 and $\sim$20~keV, from $\sim$1.4\% to $\sim$2.8\% in the two cases presented in Figure~\ref{fig:qpoRE}. At energies higher than 20~keV the mHz QPO is not visible by naked eye in the light curves, nor it is conclusively detected using LSPs, FFTs or folded light curves. The presence of red noise with variable strength gives rise to different upper limits on the fractional rms amplitude of the mHz QPO above 20~keV (between 1.4\% and 3\%; Fig.~\ref{fig:qpoRE}). As discussed in Section~\ref{sec:msb}, such increase in QPO amplitude up to 20~keV is in contrast with mHz QPOs from other bursting NSs, which showed a fractional amplitude decreasing with increasing energy between 2 and 5 keV \citep{Revnivtsev01}. \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f7.eps}}} \caption{Burst peak luminosity {\it (top)}, integrated burst energy {\it (middle)} and burst recurrence time, $t_\mathrm{rec}$ {\it (bottom)} vs. bolometric persistent luminosity ($L_\mathrm{pers}$). Gray and black circles show individual burst measurements and daily averages, respectively. Red filled squares show recurrence times measured from the inverse of the mHz QPO frequency. The top axis shows the inferred $\dot{m}$ assuming homogeneous accretion onto a 1.4~$M_\odot$, 10~km radius NS (Sec.~\ref{sec:data}).} \label{fig:lumlum} \end{center} \end{figure} \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f8.eps}}} \caption{Same as Figure~\ref{fig:lumlum}, showing also the broken power law fits to the $L_\mathrm{peak}$, $E_\mathrm{b}$, $t_\mathrm{rec}$ vs. $L_\mathrm{pers}$ relations (Table~\ref{table:power}). Three regimes are apparent and labeled along the bottom axis: slow decrease (at $L_\mathrm{pers}$/$L_\mathrm{Edd}$$\lesssim$0.2; regime A), fast drop (at 0.2$\lesssim$$L_\mathrm{pers}$/$L_\mathrm{Edd}$$\lesssim$0.3; regime B) and saturation ($L_\mathrm{pers}$/$L_\mathrm{Edd}$$\gtrsim$0.3; regime C) of the $L_\mathrm{peak}$, $E_\mathrm{b}$ vs. $L_\mathrm{pers}$ relations. The $t_\mathrm{rec}$ values follow approximate $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-3}$ and $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-1}$ relations in regimes B and C, as indicated (see Figure~\ref{fig:trec}, Table~\ref{table:regimes} and Secs.~\ref{sec:burstmdot} \& \ref{sec:discussion}).} \label{fig:lumlumlin} \end{center} \end{figure} \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f9.eps}}} \caption{Daily-averaged accretion-to-burst energy ratio ($\alpha$, top panel) and burst time scales (rise time and duration, bottom panel) vs. persistent luminosity ($L_\mathrm{pers}$) during the T5X2 outburst. Open triangles represent lower limits (one single burst detected during that day). The corresponding bursting regimes are indicated with dotted lines and labels along the bottom axis (Table~\ref{table:regimes}).} \label{fig:alpha} \end{center} \end{figure} \begin{table}[ht] \caption{Broken power law fits to the burst properties vs. persistent luminosity relations.} \begin{minipage}{\textwidth} \begin{tabular}{l c c c} \hline\hline Relation\footnote{Pairs of variables fitted with a broken power law: $L_\mathrm{peak}$, $E_\mathrm{b}$\\ and $t_\mathrm{rec}$ vs. $L_\mathrm{pers}$ (see Figures~\ref{fig:lumlum} \& \ref{fig:lumlumlin}). The $L_\mathrm{pers}$ range used in\\ each fit is quoted between brackets, in units of $L_\mathrm{Edd}$. i$_1$ and i$_2$\\ represent, respectively, the power law indices before and after the\\ break (at $L_\mathrm{break}$).} &$L_\mathrm{peak}$-$L_\mathrm{pers}$ &$E_\mathrm{b}$-$L_\mathrm{pers}$ &$t_\mathrm{rec}$-$L_\mathrm{pers}$ \\ $[$range$]$ & [0.1--0.3] & [0.1--0.3] & [0.17--0.45] \\ \hline K\footnote{Power law normalization before the break, same units as $L_\mathrm{peak}$\\ (10$^{37}$ erg~s$^{-1}$), $E_\mathrm{b}$ (10$^{39}$ erg) and $t_\mathrm{rec}$ (s).} & 2.2 $\pm$ 0.7 & 0.9 $\pm$ 0.1 & 10 $\pm$ 7 \\ i$_1$ & -0.3 $\pm$ 0.2 & -0.2 $\pm$ 0.1 & -3.2 $\pm$ 0.5 \\ $L_\mathrm{break}$($L_\mathrm{Edd}$) & 0.20 $\pm$ 0.01 & 0.20 $\pm$ 0.01 & 0.33 $\pm$ 0.04 \\ i$_2$ & -2.7 $\pm$ 0.1 & -4.4 $\pm$ 0.1 & -1.0 $\pm$ 0.9 \\ $\chi^2$ / dof& 38.5 / 25 & 930 / 25 & 6.1 / 30 \\ \hline\hline \end{tabular} \end{minipage} \label{table:power} \end{table} \subsection{Burst properties vs. accretion rate} \label{sec:burstmdot} We present in this Section the relation between burst properties and $\dot{m}$, as inferred from $L_\mathrm{pers}$, including the faintest and most frequent bursts which form the mHz QPOs (see above). Figure~\ref{fig:ob} shows $L_\mathrm{peak}$, $E_\mathrm{b}$ and $t_\mathrm{rec}$ as a function of the Eddington-normalized $L_\mathrm{pers}$ and $\dot{m}$ (Sec.~\ref{sec:data}). The overall trend is that of an anti-correlation between $L_\mathrm{peak}$, $E_\mathrm{b}$ and $t_\mathrm{rec}$ on the one hand and $L_\mathrm{pers}$ on the other (see also Fig.~1 in \citealt{Linares11b}; \citealt{Motta11}). Closer inspection of the burst properties over the full $L_\mathrm{pers}$ range (0.1--0.5~$L_\mathrm{Edd}$; see Figure~\ref{fig:lumlum}) reveals a more complex and interesting behavior, namely, four different bursting regimes. \begin{itemize} \item At the lowest persistent luminosities, 0.1$<$$L_\mathrm{pers}$/$L_\mathrm{Edd}$$<$0.2, when $L_\mathrm{pers}$ increases both $L_\mathrm{peak}$ and $E_\mathrm{b}$ (and possibly $t_\mathrm{rec}$) decrease moderately, from $L_\mathrm{peak}$$\simeq$4.6$\times 10^{37}$erg/s, $E_\mathrm{b}$$\simeq$1.6$\times 10^{39}$ erg to $L_\mathrm{peak}$$\simeq$3.7$\times 10^{37}$erg/s, $E_\mathrm{b}$$\simeq$1.3$\times 10^{39}$erg. We refer to this 0.1--0.2~$L_\mathrm{Edd}$ regime as {\it slow decrease, or regime A} (see Fig.~\ref{fig:lumlum}). \item At higher persistent luminosities, 0.2$<$$L_\mathrm{pers}$/$L_\mathrm{Edd}$$<$0.3, we find $L_\mathrm{peak}$ and $E_\mathrm{b}$ to be steeply anticorrelated with $L_\mathrm{pers}$, while $t_\mathrm{rec}$ also decreases from 200~s to 400~s with increasing $L_\mathrm{pers}$. We refer to this 0.2--0.3~$L_\mathrm{Edd}$ regime as {\it fast drop, or regime B} (see Fig.~\ref{fig:lumlum}). \item At the highest persistent luminosities ($L_\mathrm{pers}$$>$0.3~$L_\mathrm{Edd}$), when mHz QPOs are detected, burst peak luminosity and energy reach approximately constant values ({\it saturation, or regime C}): $L_\mathrm{peak}$$\simeq$1.2$\times 10^{37}$erg/s and $E_\mathrm{b}$$\simeq$0.3$\times 10^{39}$erg. The burst recurrence time, however, keeps decreasing with increasing $L_\mathrm{pers}$ from $t_\mathrm{rec}$$\simeq$400~s down to $\sim$240~s on 2010 October 18 (Fig.~\ref{fig:lumlum}; Table~\ref{table:mhzqpo}). \item Finally, during incursions into the flaring/normal branches near the outburst peak, when T5X2 showed Z source behavior \citep{Altamirano10b}, neither bursts nor mHz QPOs were detected. These few episodes without bursting activity, which we refer to as {\it regime D}, were associated with moderate ($\sim$30\%) short (hour-long) {\it drops} in $L_\mathrm{pers}$ and occurred between MJDs 55486 and 55496 (while $L_\mathrm{pers}$$\gtrsim$0.3~$L_\mathrm{Edd}$). They are noted with square brackets in Table~\ref{table:dailybursts}. \end{itemize} \begin{table*}[ht] \caption{Summary of bursting regimes in T5X2. See Section~\ref{sec:burstmdot} for details.} \begin{minipage}{\textwidth} \begin{center} \begin{tabular}{c c c c c c} \hline\hline Regime & $L_\mathrm{pers}$ & $t_\mathrm{rec}$ & $E_\mathrm{b}$ & $L_\mathrm{peak}$ & Description \\ & ($L_\mathrm{Edd}$) & (s) & (10$^{39}$ erg) & (10$^{37}$ erg~s$^{-1}$) & \\ \hline A & 0.1--0.2 & $>$2000 & 1.6--1.3 & 4.6--3.7 & \specialcell[t]{Slow decrease of $E_\mathrm{b}$ \& $L_\mathrm{peak}$\\ (and $t_\mathrm{rec}$?) with increasing $L_\mathrm{pers}$.} \\ & & & & & \\ B & 0.2--0.3 & 2000--400 & 1.3--0.3 & 3.7--1.2 & \specialcell[t]{Fast drop of $t_\mathrm{rec}$, $E_\mathrm{b}$ \& $L_\mathrm{peak}$\\ with increasing $L_\mathrm{pers}$. $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}$$^{-3}$.} \\ & & & & & \\ C & 0.3--0.5 & 400--200 & $\sim$0.3 & $\sim$1.2 & \specialcell[t]{Bursts evolve into mHz QPOs with\\ increasing frequency ($t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}$$^{-1}$).\\ $E_\mathrm{b}$ \& $L_\mathrm{peak}$ ``saturate''.} \\ & & & & & \\ D & 0.3--0.5 & - & - & - & \specialcell[t]{No bursts/mHz QPOs detected on \\ normal \& flaring branch in Z source phase.} \\ \hline\hline \end{tabular} \end{center} \end{minipage} \label{table:regimes} \end{table*} Interestingly, regimes C and D show similar $L_\mathrm{pers}$ (0.3--0.5~$L_\mathrm{Edd}$), but their variability and spectral properties differ (i.e., they constitute different ``accretion states''; see Altamirano et al. 2012, in prep.). We summarize the main bursting properties of all four regimes in Table~\ref{table:regimes}. In order to quantify the anticorrelations explained above and to characterize the bursting regimes present in T5X2, we fit the $L_\mathrm{peak}$, $E_\mathrm{b}$, $t_\mathrm{rec}$ vs. $L_\mathrm{pers}$ relations with broken power law functions. The results are shown in Figure~\ref{fig:lumlumlin} and Table~\ref{table:power}, and confirm that the two ``breaks'' or transitions occur at $L_\mathrm{pers}$$\simeq$0.2~$L_\mathrm{Edd}$ and $L_\mathrm{pers}$$\simeq$0.3~$L_\mathrm{Edd}$. Due to the large scatter some of the fits are statistically poor, yet they allow us to constrain the slope and transition luminosity of the three bursting regimes (Fig.~\ref{fig:lumlumlin}). Remarkably, the $t_\mathrm{rec}$-$L_\mathrm{pers}$ relation that we find in regime B ($L_\mathrm{pers}$$\sim$0.2--0.3~$L_\mathrm{Edd}$) is far from linear, as would be expected from a $t_\mathrm{rec}$~$\propto$~$\dot{m}^{-1}$ relation if $\dot{m}$~$\propto$~$L_\mathrm{pers}$ (See Sec.~\ref{sec:trec} for further discussion). Instead, it is close to $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-3}$: we measure a $t_\mathrm{rec}$-$L_\mathrm{pers}$ power law index in regime B of -3.2$\pm$0.5 (see further discussion in Sec.~\ref{sec:discussion}). We also calculate the mean alpha parameter ($\alpha$$\equiv$$L_\mathrm{pers}$$\times$$t_\mathrm{rec}$ / $E_\mathrm{b}$) from the daily-averaged values reported by \citet{Linares11b} and, after applying the bolometric correction we obtained from broadband X-ray spectral fits (Sec.~\ref{sec:data}), we find a mean $\alpha$ of 102, with a standard deviation of 27. Figure~\ref{fig:alpha} shows that $\alpha$ increases from $\sim$60 to $\sim$120 with increasing $L_\mathrm{pers}$ until $L_\mathrm{pers}$$\simeq$0.3~$L_\mathrm{Edd}$ (regimes A \& B), and decreases at higher luminosities (regime C) to reach again values close to 60. Figure~\ref{fig:alpha} also shows the burst duration and rise time as a function of $L_\mathrm{pers}$: the general trend is for bursts to become shorter and have slower rise when $L_\mathrm{pers}$ increases. \begin{figure}[h!] \begin{center} \resizebox{1.05\columnwidth}{!}{\rotatebox{-90}{\includegraphics[]{f10.eps}}} \caption{Magnetic field strength vs. spin period in accreting NSs. The dotted rectangle in the lower left corner (labeled 'bursters') shows the location of all previously know thermonuclear burst sources with measured spin \citep[][and references therein]{Psaltis99c,Chakrabarty03,Patruno10}. The dotted rectangle in the upper right corner (labeled 'HMXBs') shows the location of X-ray pulsars in high-mass X-ray binaries (\citealt{Bildsten97}; filled gray circles show X-ray pulsars with $B$ measured from cyclotron lines, \citealt{Caballero11}). Red double-head arrows show the location of T5X2 \citep[using the two magnetic field estimates from][]{Papitto11,Miller11}. Vertical black lines mark the location of other slow pulsars in LMXBs ($P_\mathrm{s}$$>$0.01~s), none of which has shown thermonuclear bursts to date. These are labeled with the first two characters of the source names, which follow: 2A~1822--371 \citep {Jonker01}, 4U~1626--67 \citep {Rappaport77,Orlandini98}, GRO~1744--28 \citep[the ``bursting pulsar'', BP;][] {Kouveliotou96,Finger96,Cui97,Sturner96}, Her~X-1 \citep {Truemper78,Bildsten97,Mihara90,Fiume98} and GX~1+4 \citep {Chakrabarty97b,Hinkle06,Cui97}. } \label{fig:BvS} \end{center} \end{figure} \section{Discussion} \label{sec:discussion} The bursting behavior of T5X2 strikes as surprising for several reasons. During the last three decades bursts at persistent luminosities $L_\mathrm{pers}$$\gtrsim$0.2~$L_\mathrm{Edd}$ had proven exceptional and extremely difficult to detect, even when studying a large sample of bursters \citep[e.g.][and references therein]{Cornelisse03,Galloway08}. This decrease of burst rate at high $L_\mathrm{pers}$, opposite to what standard burst theory predicts \citep{Fujimoto81,Bildsten98}, has been attributed in the literature to several effects, including: i) stable thermonuclear burning becoming more important at high $\dot{m}$ and consuming an increasing fraction of the accreted fuel \citep{Paradijs88}; ii) deflagration fronts or ``flames'' propagating on the NS surface and consuming part of the fuel \citep{Bildsten95} and iii) non-spherical accretion confined to a fraction of the NS surface that would increase with $L_\mathrm{pers}$ \citep{Bildsten00}. The large number of bursts observed from T5X2 in October--November 2010 (at an average burst rate over more than a month of 4.9~hr$^{-1}$; Sec.~\ref{sec:results}) make T5X2 the most prolific source of thermonuclear bursts known to date. Notably, such copious burst activity was observed when $L_\mathrm{pers}$ was in the range 0.1--0.5~$L_\mathrm{Edd}$, a regime where burst rate is found to decrease drastically in all other bursters. The only source that has shown (albeit sporadically) such high burst rates at similar $L_\mathrm{pers}$ is Cir~X-1: on May 2010 it featured burst recurrence times as short as 1000--2000~s, when $L_\mathrm{pers}$ was $\sim$0.2~$L_\mathrm{Edd}$ \citep[][]{Linares10d}. Two other well known sources of ``high-$\dot{m}$'' bursts, GX~17+2 and Cyg~X-2, have shown bursts much less frequently \citep[mean burst rate over more than 10~yr was 0.05 and 0.1~hr$^{-1}$, respectively;][]{Galloway08} and when $L_\mathrm{pers}$ was close to $L_\mathrm{Edd}$. Other systems have shown thermonuclear bursts at intermediate accretion rates ($L_\mathrm{pers}$$\simeq$0.1--0.5~$L_\mathrm{Edd}$; including GX~3+1, GX~13+1, Ser~X-1, 4U~1636--53, 4U~1735-44, 4U~1746-37), but only sporadically \citep[$t_\mathrm{wait}$ of at least one hour and typically much longer;][and references therein]{Galloway08}. T5X2 is therefore exceptional in that it follows the expected burst rate at high $\dot{m}$ \citep{Fujimoto81,Bildsten98} much more closely than any other known burster \citep{Cornelisse03,Galloway08}. We extend on this comparison between thermonuclear burning theory and T5X2 in Section~\ref{sec:trec}. Moreover, T5X2 is the first slow X-ray pulsar (spin period $P_\mathrm{s}$$>$10~ms) to show thermonuclear bursts. None of the slower ($P_\mathrm{s}$$\gtrsim$1~s) ``classical'' X-ray pulsars in HMXBs have shown thermonuclear bursts to date, even if their persistent luminosity varies over the same range as in bursters. This is usually attributed to the stronger dipolar NS magnetic field ($B$) in HMXBs channeling the accretion flow into a much smaller area than LMXBs, leading to stable burning of all the accreted fuel due to a very high local $\dot{m}$ (\citealt{Joss80}; see also Sec.~\ref{sec:bfield}). On the other hand, only 7 out of the more than 90 bursters known have shown X-ray pulsations in the persistent emission (i.e., 7 out of the 14 accreting millisecond pulsars have shown bursts to date). Before the discovery of T5X2 the slowest spinning burster had a spin frequency more than 20 times higher than T5X2 \citep[IGR~J17511--3057, with a spin frequency $\nu_\mathrm{s}$=245~Hz;][]{Markwardt09,Altamirano10c}. T5X2 therefore bridges the gap between ``pulsars that don't burst and bursters that don't (typically) pulse'' \citep[][and references therein]{Bildsten98}. We show this in Figure~\ref{fig:BvS} by comparing both the $B$ and $P_\mathrm{s}$ values of T5X2 to those of LMXBs (including a few peculiar LMXBs) and HMXBs. Figure~\ref{fig:BvS} clearly shows that T5X2 features values of $B$ and $P_\mathrm{s}$ intermediate between bursters and HMXBs. We discuss the consequences of such a high $B$ and long $P_\mathrm{s}$ for thermonuclear burst regimes in Sections~\ref{sec:bfield} and \ref{sec:rotation}, respectively. The shortest burst wait times known to date \citep[][5.4~and 3.8~min, respectively]{Linares09c,Keek10} were based on sets of typically two or three consecutive bursts followed by much longer periods without bursts. The bursts from T5X2 presented herein are remarkably quasi-periodic, not only in the mHz QPO phase, as witnessed by the well defined and smoothly evolving wait time in individual bursts ($t_\mathrm{wait}$; see Figs.~\ref{fig:evol} \& \ref{fig:ob}). This behavior is analogous to the ``clocked burster'' \citep[GS~1826--24;][]{Tanaka89,Ubertini99,Galloway04,Heger07b} and IGR~J17511--3057 \citep{Falanga11}, which have shown bursts at regular intervals with an approximate $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-1}$ relation, although at lower accretion rates than T5X2. For this reason, we compare in Figure~\ref{fig:trec} the burst energies and recurrence times of T5X2 to those of these two sources, as well as Cir~X-1. The bursts from T5X2 (at $\dot{m} \sim$0.1--0.5~$\dot{m}_\mathrm{Edd}$) are more frequent and less energetic than those from GS~1826--24 and IGR~J17511--3057 (at $\dot{m} \sim$0.02--0.06~$\dot{m}_\mathrm{Edd}$). Figure~\ref{fig:trec} also shows that the $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-3}$ relation that we find in regime B is unique to T5X2. The bursts from Cir~X-1, however, have again a strong resemblance to those of T5X2, with energies in the same range ($\lesssim 10^{39}$~erg) and a similar $E_\mathrm{b}$-$L_\mathrm{pers}$ relation (a break or transition is also suggested by the Cir~X-1 data; Figure~\ref{fig:trec}). \begin{figure}[h!] \begin{center} \resizebox{1.0\columnwidth}{!}{\rotatebox{-0}{\includegraphics[]{f11.eps}}} \caption{Burst energy {\it(top panel)} and recurrence time {\it(bottom panel)} vs. persistent luminosity (bottom axis) and inferred mass accretion rate per unit surface (top axis) for four bursters, as indicated on the bottom panel: T5X2 (filled black circles; this work), GS~1826--24 \citep[``the clocked burster'', open gray squares; from][]{Galloway04}, Cir~X-1 \citep[open red pentagons; from][]{Linares10d} and IGR~J17511--3057 \citep[open gray triangles; from][]{Falanga11}. The solid line shows the empiric $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-1.05}$ relation found for GS~1826--24 by \citet[][]{Galloway04}, similar to the $t_\mathrm{rec} \propto L_\mathrm{pers}^{-1.1}$ relation found in IGR~J17511--3057 \citep{Falanga11}. We also show the $E_\mathrm{b}$-$\dot{m}$ and $t_\mathrm{rec}$-$\dot{m}$ relations predicted by two ignition models from \citet[][labeled CB00]{CB00} with different accreted hydrogen fractions: $X_0$=0.7 (green dashed line) and $X_0$=0.1 (magenta dotted line). In both models the metallicity was assumed to be $Z$=0.02 and the heat flux from the crust was fixed at $Q_\mathrm{b}$=0.1~MeV~nucleon$^{-1}$.} \label{fig:trec} \end{center} \end{figure} \subsection{Bursting regimes vs. burning regimes:\\ the need for heat} \label{sec:trec} In the present Section we compare in detail the T5X2 burst properties with theoretical predictions. Models of nuclear burning in the envelope of (non-magnetic, non-rotating) accreting NSs predict four different burning regimes on a NS accreting a mixture of hydrogen (H; mass fraction $X_0$), helium (He; mass fraction $Y_0$) and heavy elements \citep[mainly CNO, mass fraction $Z$; see][and references therein]{Woosley76,Fujimoto81,Taam81,Bildsten98,CB00}. At the highest accretion rates, close to or higher than $\dot{m}_\mathrm{Edd}$, both H and He burn stably and no bursts are expected (Sec.~\ref{sec:msb}). For accretion rates \begin{equation} \dot{m} > \dot{m}_\mathrm{sHb} \simeq 0.008~\dot{m}_\mathrm{Edd}\left(\frac{0.7}{X_0}\right)\left(\frac{Z}{0.01}\right)^{1/2}, \label{eq:shb} \end{equation} \citep[][where we assumed an opacity of 0.04~cm$^2$~g$^{-1}$ and the value of $\dot{m}_\mathrm{Edd}$ given in Sec.~\ref{sec:data}]{Bildsten98}, H burns stably between bursts at a constant rate, via the so-called ($\beta$-limited) ``hot-CNO'' cycle \citep[][and references therein]{Bildsten98,CB00}. In this range all bursts are triggered when He burning at the base of the accreted layer becomes thermally unstable (He ignition). The column depth (or density) at the base of the burning layer when ignition occurs is known as ignition depth, $y_\mathrm{ign}$, and the time between bursts is simply $t_\mathrm{rec}=y_\mathrm{ign}/\dot{m}$. As H burns at a constant rate (temperature and $\dot{m}$ independent, but proportional to $Z$), the time it takes to consume all H in a sinking fluid element depends only on $X_0$ and $Z$: \begin{equation} t_\mathrm{burn}=27~\mathrm{hr}~\left(\frac{X_0}{0.7}\right)\left(\frac{0.01}{Z}\right)~(1+z)/1.31 \label{eq:tburn} \end{equation} \citep[as seen by an observer; e.g.,][]{Galloway06b}, where 1+$z$=(1-2$GM/Rc^2$)$^{-1/2}$=1.31 is the gravitational redshift on the surface of a NS with mass $M$=1.4~$M_\odot$ and radius $R$=10~km. The longest burst recurrence time that we measure in T5X2 is $t_\mathrm{rec}\sim$1~hr. For Solar abundances this implies that $t_\mathrm{rec} << t_\mathrm{burn}$, i.e., there is no time to consume all H between the T5X2 bursts unless the fuel is substantially H-poor and/or metal rich (see below). In general, for accretion rates \begin{equation} \dot{m} > \dot{m}_\mathrm{dep} \simeq~0.04~\dot{m}_\mathrm{Edd}\left(\frac{0.7}{X_0}\right)\left(\frac{Z}{0.01}\right)^{13/18} \label{eq:mdep} \end{equation} \citep[][for $M$=1.4~$M_\odot$ and $R$=10~km]{Bildsten98}, there is no time to burn the accreted H before reaching ignition conditions (i.e., $t_\mathrm{rec} < t_\mathrm{burn}$) so that He ignites in a mixture of H and He. In this regime $y_\mathrm{ign}$ does not depend sensitively on $\dot{m}$, and a simple $t_\mathrm{rec} \propto \dot{m}^{-1}$ relation is expected. For $\dot{m} < \dot{m}_\mathrm{dep}$ instead, there is enough time to deplete all H before the base of the accreted layer reaches ignition conditions. Bursts are then triggered in the absence of H (pure He ignition). In this regime $y_\mathrm{ign}$ decreases with increasing $\dot{m}$, which results in a steeper decline of the burst recurrence time as $\dot{m}$ increases, close to $t_\mathrm{rec} \propto \dot{m}^{-3}$ \citep{CB00}. This transition between pure He and mixed H/He ignition regimes can be seen in the $E_\mathrm{b}$-$\dot{m}$ and $t_\mathrm{rec}$-$\dot{m}$ relations in Figure~\ref{fig:trec}, where we show two semi-analytic models from \citet{CB00} with different compositions. Even though the transition from pure He to mixed H/He ignition is expected at $\dot{m}_\mathrm{dep}$$\sim$0.05$\dot{m}_\mathrm{Edd}$ for the case of Solar abundances, Figure~\ref{fig:trec} and Equation~\ref{eq:mdep} clearly show that changes in the accreted composition can increase $\dot{m}_\mathrm{dep}$ by a factor of at least 10. In particular, Figure~\ref{fig:trec} shows that ignition models with low H abundance, [$X_0$=0.1, $Z$=0.02], can reproduce the change in slope of both the $E_\mathrm{b}$-$L_\mathrm{pers}$ and $t_\mathrm{rec}$-$L_\mathrm{pers}$ relations that we observe in T5X2 at $L_\mathrm{pers} \simeq 0.3 L_\mathrm{Edd}$, i.e., at the transition between regimes B and C (Sec.~\ref{sec:burstmdot}, Tables~\ref{table:power} \& \ref{table:regimes}). With such low $X_0$, H can be depleted before reaching He ignition at accretion rates much higher than in the case of Solar abundances, which highlights the importance of fuel composition in the burning regimes. The average T5X2 accretion-to-burst energy ratio ($\alpha$=$Q_\mathrm{grav}$(1+$z$)/$Q_\mathrm{nuc}$), $\langle\alpha\rangle = 102$, corresponds to a total nuclear energy release during the bursts $Q_\mathrm{nuc}$$\simeq$2.8~MeV~nucleon$^{-1}$ \citep[see also][]{Motta11}. Taking $Q_\mathrm{nuc} = 1.6+4\langle X\rangle$~MeV~nucleon$^{-1}$ \citep[which assumes complete burning of the accumulated fuel and $\sim$35\% neutrino energy loss;][and references therein]{Galloway08} we find an average H mass fraction over the burning layer $\langle X\rangle$$\simeq$0.3. Such low inferred $Q_\mathrm{nuc}$ lends support to the low H fraction fuel scenario for T5X2 proposed above. In summary, the slopes of the $E_\mathrm{b}$-$L_\mathrm{pers}$ and $t_\mathrm{rec}$-$L_\mathrm{pers}$ relations as well as the $\alpha$ values observed during the T5X2 outburst, indicate a low accreted H fraction and strongly suggest a transition from the pure He ignition regime to the mixed H/He ignition regime happening during the outburst rise when $\dot{m}$ increases above $\sim$0.3~$\dot{m}_\mathrm{Edd}$. It is worth noting that the reverse transition is observed during the outburst decay when $\dot{m}$ {\it drops} below 0.3~$\dot{m}_\mathrm{Edd}$ (i.e., the outburst rise and decay tracks overlap and no hysteresis is seen in Figures~\ref{fig:lumlum} \& \ref{fig:lumlumlin}). We stress that the link between the observed T5X2 {\it bursting} regimes and the theoretical {\it burning} regimes that we put forward is based on the $t_\mathrm{rec}$-$L_\mathrm{pers}$ and $E_\mathrm{b}$-$L_\mathrm{pers}$ relations. Modeling and interpretation of the individual burst light curves is beyond the scope of this work, yet we note that different burst light curves and peak luminosities are predicted in the different ignition regimes \citep[e.g.][]{Woosley04}. Pure He bursts are expected to show faster rise times than mixed H/He bursts. This agrees qualitatively with the identification of regime B as pure He ignition: $t_\mathrm{rise}$ is shorter in regime B than in regime C (Fig.~\ref{fig:alpha}). The large change in $L_\mathrm{pers}$, however, could influence the observed timescales along the burst rise and decay, as these are measured after subtracting the persistent emission. It should also be noted that at low accretion rates (near $\dot{m}_\mathrm{sHb}~\simeq~0.08~\dot{m}_\mathrm{Edd}$ for $X_0$=0.1, $Z$=0.02) there should be a transition to H-ignited bursts, not included in the models discussed in this Section \citep{CB00}. If the link between bursting and burning regimes that we propose is correct, to our knowledge this is the first time that the transition between pure He and mixed H/He ignition is observed in a single source. There is, however, a systematic and interesting discrepancy evident in Figure~\ref{fig:trec}. Even when including compressional heating or when increasing the base heat flux from the NS crust \citep[up to 2 MeV~nucleon$^{-1}$; fixed at 0.1~MeV~nucleon$^{-1}$ in Fig.~\ref{fig:trec};][]{CB00}, ignition models predict higher burst energies and longer recurrence times than those we find in T5X2, by a factor close to 10 in most cases (i.e., larger than the distance uncertainty on $L_\mathrm{pers}$ and $E_\mathrm{b}$, Sec.~\ref{sec:data}). This large difference between the observed and predicted values of $E_\mathrm{b}$ and $t_\mathrm{rec}$, together with the lack of detailed modeling of the T5X2 burst light curves, prevents a conclusive identification of the observed bursting regimes. We propose that such discrepancy could be explained by the presence of an extra source of heat in the NS envelope not accounted for by ignition models (which typically consider only hot-CNO heating). Additional heat would act to reduce $y_\mathrm{ign}$ and thereby decrease $E_\mathrm{b}$ and $t_\mathrm{rec}$, explaining the T5X2 observations presented herein. Several interesting possibilities for the nature of this extra source of heat have been partially investigated, including: i) heating due to stable He burning \citep[triple alpha reaction rates have been recently debated; e.g.,][]{Ogata09,Dotter09,Peng10}; ii) heating due to deep burning of residual H \citep[e.g.,][who already pointed out that it can lead to weaker and more frequent bursts than expected]{Taam96}; iii) thermal inertia, ``hot ashes'' or heat from previous bursts, which could become important for T5X2 as it features $t_\mathrm{rec} \sim 200s$, the shortest recurrence times between thermonuclear bursts ever observed \citep[time-dependent simulations of a series of bursts are needed to investigate this in detail; e.g.,][]{Heger07} and iv) turbulent friction at the base of the spreading layer, which could release substantial amounts of heat \citep{Inogamov10}. It is worth noting that an independent study of the same source \citep{Degenaar11c} reached similar conclusions, suggesting the presence of a ``shallow heat source'' in T5X2 in order to reconcile quiescent observations and crust heating/cooling theory. Without regard to the exact nature of this heat source, we have shown in the present work that the burst properties of T5X2 place new constraints on the thermal properties of a fast-accreting and frequently-bursting NS. \subsection{mHz QPOs and marginally stable burning} \label{sec:msb} The mHz QPOs from T5X2 \citep[Sec.~\ref{sec:mhzqpo}; see also][]{Linares10c} have distinctive properties that clearly set them apart from the previously known mHz QPOs \citep[Sec.~\ref{sec:intro};][]{Revnivtsev01,Altamirano08d}. The persistent luminosity that we measure in T5X2 when mHz QPOs are present is about 10 times higher than that observed in previous mHz QPO sources ($L_\mathrm{pers}$ higher by a factor of 4--25 taking into account the observed ranges: 0.02--0.1~$L_\mathrm{Edd}$ in atoll sources as opposed to 0.4--0.5~$L_\mathrm{Edd}$ in T5X2; Table~\ref{table:mhzqpo}). For this reason we refer to the previously known mHz QPOs as ``low-$L_\mathrm{pers}$ mHz QPOs''. Bright bursts and low-$L_\mathrm{pers}$ mHz QPOs alternate, while remaining clearly distinguishable. Instead, in T5X2 bursts smoothly evolve into mHz QPOs and vice versa (Sec.~\ref{sec:mhzqpo}), a phenomenon never observed before. Strikingly, the same qualitative evolution from bright infrequent bursts to faint and frequent bursts, mHz QPOs and ultimately stable burning is predicted to happen as $\dot{m}$ increases by both one-zone models and detailed simulations of nuclear burning on NSs accreting near the boundary between unstable and stable He burning \citep{Heger07}. We can therefore identify with confidence the mHz QPOs from T5X2 with marginally stable burning on the NS surface. The evidence that links low-$L_\mathrm{pers}$ mHz QPOs with marginally stable burning is less conclusive, but remains valid \citep[see][for details]{Revnivtsev01,Yu02,Altamirano08d}. Analytic estimates place the threshold for stable He burning at \begin{equation} \dot{m}_\mathrm{sb} \simeq 1.1~\dot{m}_\mathrm{Edd}\left(\frac{1.7}{1+X_0}\right)^{3/4}\left(\frac{Y_0~\mu}{0.3\times0.6}\right)^{1/2}, \label{eq:sb} \end{equation} where $\mu$ is the mean molecular weight of the accreted fuel \citep[][assuming again $M$=1.4~$M_\odot$ and $R$=10~km]{Bildsten98}. The persistent luminosity where mHz QPOs are observed in T5X2 (0.4--0.5~$L_\mathrm{Edd}$) is therefore closer to the expected value of $\dot{m}_\mathrm{sb}$ than what was seen in low-$L_\mathrm{pers}$ mHz QPOs, yet still inconsistent with the value predicted by theory, which is higher by a factor of $\sim$2 (unless Terzan 5 is at $\sim$9~kpc instead of 6.3kpc, Sec.~\ref{sec:data}). Invoking H-poor fuel makes this discrepancy even larger, as it can increase the expected $\dot{m}_\mathrm{sb}$ by a factor 2 (for $X_0$=0.1, $Z$=0.01; Eq.~\ref{eq:sb}). A more massive NS can have unstable burning at higher $\dot{m}$ than a less massive star, but given the weak (M$^{1/2}$) scaling of $\dot{m}_\mathrm{sb}$ different NS masses cannot explain the large difference (factor 4--25) in $L_\mathrm{pers}$ between T5X2 and the other known sources of mHz QPOs. Another remarkable difference betwen the mHz QPOs presented herein and the low-$L_\mathrm{pers}$ mHz QPOs resides in their fractional rms spectrum (Sec.~\ref{sec:data}): we find a fractional rms amplitude that increases with energy between 2 and 20~keV (Sec.~\ref{sec:mhzqpo}), while \citet{Revnivtsev01} reported fractional amplitudes decreasing with increasing photon energy between 2 and 5 keV. The energy-averaged fractional amplitudes are, however, similar \citep[1.3--2.2 \% in T5X2, this work; 0.7--1.9 \% in low-$L_\mathrm{pers}$ mHz QPOs,][]{Revnivtsev01}. The increase of fractional amplitude with energy that we find is shallower than the linear increase predicted by models of temperature oscillations at the NS surface \citep[developed in the context of burst oscillations;][]{Piro06}. An interesting possibility is that the low-$L_\mathrm{pers}$ mHz QPOs trace the boundary of stable H burning (Eq.~\ref{eq:shb}), whereas the mHz QPOs that we discovered in T5X2, at a higher accretion rate, occur at the boundary of stable He burning (Eq.~\ref{eq:sb}). This remains speculative at present given the lack of published theoretical work on this particular topic. A specific analysis of oscillatory burning at the thermal stability boundary of H burning is needed to address this hypothesis. The oscillatory behavior at the marginally stable point is generic \citep{Paczynski83} and it could also occur when H burning stabilizes. If low-$L_\mathrm{pers}$ mHz QPOs do happen at the H burning stability boundary, the need to invoke confined accretion to explain their low $L_\mathrm{pers}$ would vanish. \citet{Heger07} found that the QPO frequency is mainly sensitive to the accreted H fraction ($X_0$) and the NS surface gravity: increasing the surface gravity or decreasing $X_0$ leads to higher mHz QPO frequencies. The highest $\nu_\mathrm{QPO}$ values that we find in T5X2 (4.2~mHz) are about a factor of 3 lower than those of the low-$L_\mathrm{pers}$ mHz QPOs \citep[Table~\ref{table:mhzqpo};][]{Revnivtsev01,Altamirano08d}. If the accreted fuel has a similar composition, and if the mHz QPOs have the same origin, this would suggest a less compact NS in T5X2 than in the low-$L_\mathrm{pers}$ mHz QPO sources (the ``atoll'' sources 4U~1636-536, 4U~1608-52 and Aql~X-1). Comparing to models of marginally stable burning, we find that if X$_0$$<$0.7 the surface gravity of the NS in T5X2 must be $g$ = $GM/R^2$$<$1.9$\times$10$^{14}$~cm~s$^{-2}$ \citep[Figure~9 in][]{Heger07}. Even though it is subject to theoretical (based on analytic one-zone model) and observational (a new outburst of T5X2 could show higher $\nu_\mathrm{QPO}$) uncertainties, this illustrates how a comparison between marginally stable burning models and mHz QPO properties can be used to constrain the NS compactness. Despite the striking similarities between the T5X2 burst properties and the general bursting behavior predicted by models of nuclear burning near the transition from unstable to stable burning \citep[cf. Fig.~\ref{fig:evol} in this work and Fig.~5 in][]{Heger07}, interesting differences remain. First, as explained above, the $\dot{m}$ where marginally stable burning is expected \citep[0.925~$\dot{m}_\mathrm{Edd}$][]{Heger07} is higher than the highest inferred $\dot{m}$ where we observe mHz QPOs in T5X2 (0.5~$\dot{m}_\mathrm{Edd}$). Second, \citet{Heger07} find a sharp transition from bursts to mHz QPOs and finally stable burning (occurring between 0.923--0.95~$\dot{m}_\mathrm{Edd}$), while we observe in T5X2 a smooth evolution from bursts to mHz QPOs, and vice versa (between 0.1--0.5~$\dot{m}_\mathrm{Edd}$; Sec.~\ref{sec:results}). Time-dependent simulations of nuclear burning tailored to T5X2 would be of high interest, in particular exploring the H-poor fuel range. Third, we find that the disappearance of the mHz QPOs in T5X2 is not linked to an increase in $\dot{m}$ but to an actual {\it drop} of $L_\mathrm{pers}$ that happens when the accretion state changes (from horizontal to normal \& flaring branches). This suggests that the geometry/configuration of the accretion flow plays a role in setting the nuclear burning stability boundary. \subsection{The role of an intermediate magnetic field} \label{sec:bfield} While most models of thermonuclear bursts on accreting NSs assume that the NS magnetic field is negligible \citep[][]{Fujimoto81,Taam82,Paczynski83,Woosley04}, \citet{Joss80} showed that the presence of a strong ($B$$\gtrsim$10$^{12}$~G) magnetic field can affect the stability of nuclear burning in different ways. A strong magnetic field reduces the (conductive and radiative) opacities, allowing for more efficient heat transport and thereby stabilizing burning. Applying disk-magnetosphere interaction models to T5X2, \citet{Papitto11} estimated $B$=2$\times$10$^8$--2.4$\times$10$^{10}$~G from the luminosity range at which pulsations were detected, whereas \citet{Miller11} further constrained $B$=(0.7--4)$\times$10$^9$~G from the Fe line profile. The fact that T5X2 shows thermonuclear bursts suggests that the magnetic field needed to suppress the thermal instability and quench thermonuclear bursts must be greater than $\sim$10$^{10}$~G, in accordance with theoretical expectations \citep{Joss80}. A strong magnetic field also affects convective heat transport and mixing \citep[e.g.][]{Joss80}. \citet{Bildsten95} proposed that even in cases where thermonuclear burning is unstable the presence of a strong magnetic field can suppress convection in the NS envelope, stalling the propagation of the burning front (diffusive heat transport being slower than convection) and preventing the fast ignition that causes type I X-ray bursts \citep[see also][]{Bildsten98}. Again, the mere presence of type I X-ray bursts in T5X2 strongly suggests that the magnetic field required to suppress convective burning fronts must be higher than $\sim$10$^{10}$~G, in agreement with analytical estimates \citep{Joss80,Bildsten97b}. Rise times are rather long though (Fig.~\ref{fig:alpha}), which could perhaps indicate that $B$ is strong enough to slow down the burning fronts \citep{Bildsten97b}. In summary, the bursting properties of T5X2 presented herein (in particular combined with refined measurements of its magnetic field strength) can place new constraints on the physics of convection and heat transport under intermediate magnetic fields (10$^8$--10$^{10}$~G). Finally, a 10$^8$--10$^{10}$~G magnetic field can also channel the accretion flow onto the magnetic polar caps, as suggested by theoretical models of disk-magnetosphere interaction \citep{Lamb73} and by the presence of X-ray pulsations in T5X2 \citep{Strohmayer10,Papitto11}. Accretion confined to the magnetic polar caps may break the spherical symmetry assumed by most thermonuclear burst models, and increase $\dot{m}$ for a given (observed) $L_\mathrm{pers}$ (Sec.~\ref{sec:data}). The impact on the burning properties depends on the size of the polar caps and on how soon the accreted fuel spreads laterally while sinking into the NS envelope under the influence of a magnetic field. \citet{Bildsten97b} found that for dipolar magnetic fields $B$$<$ (2--4)$\times 10^{10}~G$, the accreted fuel spreads before igniting and the spherically symmetric case is recovered. This suggests that the T5X2 burst properties can still be compared to spherically symmetric ignition models (Sec.~\ref{sec:trec}). As the size and ``depth'' of the polar caps are nevertheless ill-constrained quantities, we explore a simple scenario in which the fuel in T5X2 is confined into a 10\% of the NS surface down to ignition depth, $y_\mathrm{ign}$. From geometrical considerations, assuming the same $y_\mathrm{ign}$ predicted by ignition models \citep{CB00}, in this confined scenario bursts would be 10 times less energetic and 10 times more frequent than in the spherically symmetric case, which would explain the mismatch between predicted and observed $E_\mathrm{b}$ and $t_\mathrm{rec}$ (Sec.~\ref{sec:trec}). However, the accretion rate per unit area would also increase by a factor 10, which for the observed $L_\mathrm{pers}$ implies that unstable burning would operate at $\dot{m}$ as high as $\sim$5 times $\dot{m}_\mathrm{Edd}$. This is well above the predicted threshold for stable steady burning of H and He \citep[$\dot{m}_\mathrm{sb}$, Eq.~\ref{eq:sb}; see also][]{Bildsten98}, and we therefore consider this scenario unlikely. If the accreted material contains no H, however, $\dot{m}_\mathrm{sb}$ can be several times larger than $\dot{m}_\mathrm{Edd}$ \citep{Bildsten98}. \subsection{Rotation, mixing and burning} \label{sec:rotation} The unique bursting behavior in T5X2 is, in essence, much closer to what theory predicts than any other burster known to date. Theoretical burning regimes were mostly based on models that assume a non-rotating NS and radial infall of the accreted mass \citep{Fujimoto81,Bildsten98}. These are admittedly simplistic assumptions, as NSs in LMXBs can have spin frequencies in excess of 600~Hz and the orbital frequencies in the innermost accretion disk are well above 1000~Hz. The accretion of matter with angular momentum onto a spinning NS introduces a shear in its surface layers \citep[][and references therein]{Piro07,Keek09}. The main effect of rotation which may have an impact on burning regimes is turbulent mixing. In particular, \citet{Keek09} found that such rotationally induced turbulent mixing stabilizes He burning, decreasing the burning stability boundary ($\dot{m}_\mathrm{sb}$). Motivated by the stronger magnetic field (factor $\sim$10) and slower spin (factor $\gtrsim$20) of T5X2 compared to the rest of bursters, we speculate that rotationally induced turbulent mixing is what sets T5X2 apart from the rest of thermonuclear burst sources. The stronger $B$ field could channel a large fraction of the accreted matter into the NS magnetic poles and produce a more radial inflow which would then spread over the whole surface, minimizing shear and turbulent mixing. Therefore T5X2 seems to meet to a greater extent the assumptions of zero spin and radial infall mentioned above, which can explain the better agreement with theory. This in turn suggests that, as argued by \citet{Piro07} and \citet{Keek09}, the effects of rotation must be considered to explain the behavior of most bursters.\\ \section{Summary and Conclusions} \label{sec:conclusions} We have presented the discovery of mHz QPOs from the NS-LMXB and 11~Hz X-ray pulsar T5X2, as well as the full bursting properties during its 2010 outburst, when the persistent luminosity varied by about a factor of 5 (between $L_\mathrm{pers}$$\simeq$0.1--0.5~$L_\mathrm{Edd}$). T5X2 showed copious thermonuclear bursts when $L_\mathrm{pers}$$\gtrsim$0.2~$L_\mathrm{Edd}$, a regime where thermonuclear bursts had proven exceptional to date. Burst energies and recurrence times gradually decreased as $L_\mathrm{pers}$ increased, turning into a rapid series of faint bursts and smoothly evolving into a mHz QPO near the outburst peak. Only at the highest $L_\mathrm{pers}$ range bursts became undetectable during a few short intervals. This behavior is both unprecedented among bursters and remarkably similar to the nuclear burning regimes expected on an accreting NS near the boundary of stable He burning. We find four different bursting regimes when studying the relation between burst properties and $L_\mathrm{pers}$ and show that in one of such regimes the burst recurrence time decays steeply with increasing $L_\mathrm{pers}$, close to $t_\mathrm{rec}$~$\propto$~$L_\mathrm{pers}^{-3}$. By confronting the change in burst properties (in particular, the relation between $E_\mathrm{b}$, $t_\mathrm{rec}$ and $L_\mathrm{pers}$) to ignition model predictions, we find evidence of a transition between pure He and mixed H/He ignition occurring in T5X2 when $L_\mathrm{pers}$$\simeq$0.3~$L_\mathrm{Edd}$. We note that large discrepancies remain between the observed T5X2 burst properties and those predicted by theory at high mass accretion rates. We further argue that the accreted fuel is H-poor and suggest that an additional source of heat in the NS envelope is needed to reconcile the observed and predicted burst properties. We examine the properties of the mHz QPOs from T5X2 in the context of marginally stable burning models, and compare them to those of previously known mHz QPO sources. T5X2 features mHz QPOs with lower frequencies (by a factor $\sim$3), when $L_\mathrm{pers}$ is substantially higher (by a factor $\sim$4--25). Finally, we discuss the role of magnetic field and spin in setting the unique T5X2 burst and mHz QPO behavior, and speculate that the absence of rotation effects such as turbulent mixing of the accreted fuel may set T5X2 apart from the rest of bursters.\\ We thank M. van der Klis for detailed comments on the manuscript. We are grateful to the International Space Science Institute in Bern, where part of this work was completed. ML acknowledges support from the NWO Rubicon fellowship.
\section{Introduction} The 650 Myr old Hyades cluster is, in many respects, a test case for theories of stellar evolution in, and dynamical evolution of, open clusters in our Galaxy. The white dwarfs, in particular, shed light on the high-mass end of the initial mass function (IMF) in open clusters and the subsequent development of the cluster's mass function. In an influential paper, \citet{1992AJ....104.1876W} have analysed the white dwarf population in the Hyades, and concluded that the cluster should contain at least 21 white dwarfs dimmer than the 7 confirmed white dwarf members known at that time. The authors derived this finding from adopting a Salpeter IMF normalised via the 24 brightest main-sequence stars presently residing in the Hyades. From different considerations, \citet{1988AJ.....96..198G} arrived at a number between 50 and 150 white dwarfs originally present in the cluster. On the other hand, \citet{1998AJ....115.1536V} counted only 10 white dwarfs as known members in the Hyades when he investigated the contribution of white dwarfs to the masses of open clusters. To summarise, not much progress has been made to solve the discrepancy between the number of white dwarfs estimated from the IMF and the actual number of white dwarfs found in the Hyades. \citet{1992AJ....104.1876W} mention in their paper that white dwarfs initially present in the bound Hyades have left the cluster in the meantime, and the authors suspect them in the Hyades supercluster as defined by \citet{1958MNRAS.118...65E}. \citet{1992AJ....104.1876W} checked the \citet{1999ApJS..121....1M} catalogue by analysing the space motions of the white dwarfs therein, and claim to have detected a handful of stars moving within 13 degrees from the Hyades convergent point and tangential velocities within $\pm 2$ $\rm{km\,s}^{-1}$ of the cluster motion. Extending the search to $\pm 5$ $\rm{km\,s}^{-1}$, they found that about 2/5 of all nearby white dwarfs may be related to the Hyades. Unfortunately, \citet{1992AJ....104.1876W} did not publish the data of their candidates, therefore we could not compare them with our results below. As a mechanism for white dwarf loss, \citet{2003ApJ...595L..53F} proposed that white dwarfs could be expelled from their parent cluster through non-spherically symmetric mass loss during the post-main-sequence evolution, which leads to a recoil speed of a few kilometres per second for the white dwarf remnant. \citet{1992AJ....104.1876W} also discussed the possibility that missing white dwarfs can hide themselves behind the red dwarf companion in binary systems. They estimated that the missing ones could only be found, even in the $B,V$ bands, if the other component is later than spectral type G. In a previous paper \citep[][henceforth Paper I]{2011A&A...531A..92R} we have analysed the Hyades cluster and their surroundings up to 30 pc to search for main-sequence stars as member candidates. We found that the present-day tidal radius is about 9 pc, and 275~$M_\odot$ (364 stellar systems) are gravitationally bound. Outside the tidal radius we found another 100~$M_\odot $ in a volume between one and two tidal radii (halo), and another 60~$M_\odot $ up to a distance of 30 pc from the centre. From their kinematics we infer that the stars outside the tidal radius are formerly bound members that left the cluster. It is therefore appropriate to repeat a selection process similar to that in Paper I to search for white dwarfs up to 30 pc and more from the cluster centre. Compared to earlier studies of Hyades white dwarfs, which performed deep searches in limited fields-of-view, we have the advantage to be able to use the deep all-sky astrometric survey PPMXL \citep{2010AJ....139.2440R}. This paper presents candidates that have five out of six phase space parameters compatible with their Hyades origin. Once available, their true space motion in radial direction must confirm or reject them. Our list of candidates from Table \ref{table:1} can serve as an input catalogue for future observations. Therefore we are hesitant at this stage to draw far-reaching conclusions, e.g. on the IMF of the Hyades, on the problematics of cooling ages, or on the initial mass-final mass relation. The paper is structured as follows: after a short listing of the so-called ``classical'' Hyades white dwarfs in Sect. \ref{classical}, we describe our selection process in Sect. \ref{select}. Section \ref{individual} follows with comments to each individual candidate. The questions of spatial distribution and completeness of the sample are covered in Sect. \ref{complete}. Finally, the discussion in Sect. \ref{discus} completes the paper. \section{The ``classical'' Hyades white dwarfs}~\label{classical} The paper by \citet{1998AJ....115.1536V} lists 10 white dwarfs that we call henceforth the ``classical'' Hyades white dwarfs. There are seven single white dwarfs and three in binary systems. We list them with their primary identifiers in SIMBAD (Data base of the Centre de Donn\'ees astronomiques de Strasbourg, CDS). The seven single white dwarfs are EGGR 26, 29, 36, 37, 39, 42, and 316. The three stars in binary systems are HR 1358, EGGR 38, and V471 Tau. Data for these stars are given in Table~\ref{table:1} (the first ten rows). In some cases there is no consistency about actual membership in the cluster. For instance, \citet{1992AJ....104.1876W} exclude EGGR 29 from membership, because they put it at 60 pc, where its tangential velocity would be discordant from the bulk tangential velocity, whereas \citet{2009ApJ...696...12D} used it for their determination of the white dwarf age of the Hyades. \citet{1992MNRAS.257..257R} lists two additional candidates RHya 102 and RHya 145. He also examined all candidates for Hyades white dwarfs proposed by \citet{1969AJ.....74....2V}, and discarded all of them except vA54 and vA71, which were outside his field-of-view. We discuss these four objects in Sect. \ref{individual}. Throughout the paper we use the following abbreviations for star names: \[ \begin{array}{lp{0.8\linewidth}} \rm{VR} & \citet{1934MNRAS..94..508V}\\ \rm{HZ} & \citet{1947ApJ...105...85H}\\ \rm{HG7} & \citet{1962LowOB...5..257G}\\ \rm{HR} & Catalogue of bright stars, \citet{1964cbs..book.....H}\\ \rm{EGGR}& \citet{1965ApJ...141...83E}\\ \rm{vA} & \citet{1969AJ.....74....2V}\\ \rm{RHya} & \citet{1992MNRAS.257..257R}\\ \rm{WD} & \citet{1999ApJS..121....1M}, updated version 2008 \\ \rm{LB} & (Luyten, Blue), Luyten W.J., Various lists published by Luyten under the general title: A Search for Faint Blue Stars (50 papers)\\ \rm{LP} & (Luyten, Palomar obs.), Published from 1963 to 1981 in Univ. Minnesota, Minneapolis, fascicules 1 to 57\\ \rm{GJ} & CNS3, Catalogue of Nearby Stars, \citep{1991adc..rept.....G}. \end{array} \] \begin{figure*}[t!] \centering \includegraphics[bb=76 42 465 785,angle=270,width=17cm,clip]{cmd3_wd.ps} \caption{Colour-absolute-magnitude diagrams: $ M_V $ vs. $ B-V$ (left), $M_J $ vs. $J-K_s$ (middle) and $ M_J $ vs. $ r'-J$ (right). In all diagrams the black dots show the probable Hyades white dwarfs from this paper, the 10 classical white dwarfs are marked additionally with crosses. The black circles show also spectroscopically confirmed white dwarfs from \citet{1999ApJS..121....1M}, which are probably non-Hyades though. The small (red) dots in the left panel show the main and the degenerate sequences of stars from the CNS3 \citep{1991adc..rept.....G}. In the middle and the right panels the degenerate sequences are again from CNS3, while the 724 Hyades members from Paper I represent the main sequences. Note that the brightest stars have no $r'$ magnitudes (right panel) in CMC14. Spectroscopic binaries are marked by their numbers in Table~\ref{table:1}.} \label{cmd} \end{figure*} \section{The selection process}~\label{select} In Paper I we described in detail how we selected MS (main sequence) Hyades candidates from their kinematic and photometric properties in the PPMXL catalogue \citep{2010AJ....139.2440R}. More specifically, we used the Carlsberg-UCAC (CU) subset containing improved proper motions and photometry by including UCAC3 \citep{2010AJ....139.2184Z} and CMC14, Carlsberg Meridian Catalog 14, \citep{2006yCat.1304....0C}. Paper I also contains a description of the convergent point method, which served as a baseline for the selection. For our general selection process we did not permit the tangential motion to differ from the one given by the bulk motion of the cluster by more than 4 $\rm{km\,s}^{-1}$, and we considered only stars closer than 30 pc from the cluster centre. The stars had to be in the CU subset to PPMXL to ensure that they have CMC14 and/or 2MASS \citep{2006AJ....131.1163S} measurements. These kinematic selection criteria were fulfilled by 15757 stars out of the 140 million contained in the CU subset, which were shown in Fig.~1 of Paper I. Two features from the kinematic selection via the convergent point method are worth to be explained in more detail. First, for each candidate that is supposed to share the bulk space motion of the cluster, the convergent point method predicts a radial velocity that only depends on the position $\alpha, \delta$ of the star. In consequence, the radial velocity of each candidate has to be measured to confirm it is still member. Second, the convergent point method attributes (predicts) a so-called secular parallax to each candidate by minimising the difference between the space motion of the cluster centre and the space motion of a candidate star. In a very strict sense the secular parallax is not the least-squares solution in the plane perpendicular to the line-of-sight, but this difference is small in all practical cases. So, the second confirmation of membership comes from an independent measurement of the distance to the star. If the result of this independent distance determination does not confirm the prediction, this means in turn that the difference in space motion between the bulk of the cluster and the candidate increases. This may rule out a given star as a member candidate. As an independent check of the predicted distances of kinematic candidates, we can consider their location in different colour-absolute-magnitude diagrams. In Paper I we used $B,V$-photometry for bright stars; for stars fainter than $M_{K_s} = 4$ the basic photometric data were $r'$ from CMC14 and $JHK_s$. Only for 724 out of 15757 kinematic candidates, the membership was confirmed by photometric selection in Paper I. The procedure adopted in Paper I is a two-step procedure. The first step is the kinematic selection, and the second is the check of kinematically predicted distances with photometric ones, i.e. a comparison with isochrones. In principle, the same approach could be applied to find white dwarfs among kinematic candidates. This would require an all-sky, accurate, multi (at least two)-colour photometric survey. Concentrating on a 30 pc radius around the centre of the Hyades, such a survey should at least cover 12.1\% of the celestial sphere or 5000 square degrees. The only survey that fulfils this requirement is 2MASS in the near-infrared. However, even at the central distance of the Hyades (46.3 pc), white dwarfs will be at the limiting magnitude of 2MASS and also of CMC14, where the photometric quality becomes very poor. Therefore, we introduced an intermediate step, in which all kinematic candidates from Paper I were cross-matched with the white dwarf catalogues by Luyten (VizieR Online Data Catalog III/70) and \citet{1999ApJS..121....1M}, the updated version (VizieR Online Data Catalog III/235B). Each match was individually checked using the VizieR data base from the CDS. After rejecting obvious red dwarfs (from Luyten candidates), we identified 20 white dwarfs that passed the kinematic selection, among which were all 10 classical Hyades white dwarfs. Ten more stars would be added if the kinematic criteria were relaxed to 40~pc and 5 $\rm{km\,s}^{-1}$. Finally, we found six faint white dwarf candidates in the PPMXL (i.e. without CMC14 and/or 2MASS measurements). One more white dwarf was found in the Prosser \& Stauffer data base (presently available from J. Stauffer, priv. comm.), although it is not explicitly mentioned there as a white dwarf. For this star, No. 20 in Table \ref{table:1}, PPMXL gives wrong proper motions, therefore we took them from \citet{2006A&A...448.1235D}. In total, the sample of the Hyades white dwarf candidates contains 37 stars. All but three of them (12, 15, 30) are spectroscopically confirmed white dwarfs. For these 37 stars we then examined if their loci in colour-absolute magnitude diagrams are consistent with those of white dwarfs. As mentioned above, 2MASS and CMC14 are not well suited for the fainter candidates. Unfortunately, the UKIRT Infrared Deep Sky Survey \citep[UKIDSS,][]{2007MNRAS.379.1599L} provides no help for this problem, because in the region of the Hyades it is constrained to the central 292 square degrees, and only K-band photometry is available for these. Indeed, except for the 10 already known white dwarfs, only stars No. 20, 21, and 22 even have K-band photometry in UKIDSS. Since the vast majority of stars in the PPMXL and its CU subset do not have appropriate photometric data in the optical, we took the $B, V$ magnitudes for the Hyades white dwarf candidates from the VizieR data base (CDS). The sources are given as footnotes to Table~\ref{table:1}. For three white dwarfs (entries 28, 29, 35 in Table~\ref{table:1}), no original $V$ measurements are available. In \citet{1999ApJS..121....1M} they are SDSS white dwarfs, and we converted their $ugriz$ magnitudes into $V$ and $B-V$ using the transformations from \citet{2005AJ....130..873J}. Also, all 37 stars have been visually inspected on the digitised sky survey charts from IRSA (NASA/IPAC Infrared Science Archive) to avoid coarse misidentifications of the different surveys. In the following paragraphs we check if the kinematically selected stars populate allowed loci in the colour-absolute-magnitude diagrams (CMDs). In Fig. \ref{cmd} we show three CMDs, ($M_V $ vs. $ B-V $, left), ($M_J $ vs. $ J-K_s$, middle) and ($M_J $ vs. $ r'-J$, right). As references for the loci of the degenerate stars we have taken the white dwarfs from the CNS3 in all three panels (small red dots). For main sequence dwarfs the loci of the 724 Hyades from Paper I (also small red dots) are taken in the middle and right diagrams. Because in most cases we could not find precise B and V magnitudes for our sample of stars from Paper I, we took the main sequence in the left panel again from the CNS3. All 37 stars discussed in this paper are marked in black in Fig. \ref{cmd}. The field white dwarfs from the CNS3 show a relatively well-defined sequence in the optical (the left diagram), and a number of our Hyades white dwarf candidates follow this sequence. We conclude that the convergent point method has correctly predicted the distances for these stars and refer to them as probable Hyades members marked as solid dots in Fig. \ref{cmd}. The classical 10 Hyades white dwarfs are additionally indicated by crosses. In this panel we find 10 white dwarfs, marked by open circles which would be sub-luminous if set at their kinematic distances. We conclude that these stars must be at farther distances from the Sun than predicted, therefore have higher tangential velocities, and must be rated as probable non-Hyades. The best coincidence between the probable Hyades white dwarfs and the reference loci is seen in the left panel, one of the reasons being the better optical photometry available for the white dwarfs in the Johnson $B,V$ system. In the near-infrared, NIR, panel (middle) the scatter in the $ J-K_s$ colour is much larger than in $ B-V $, because the fainter stars are at the detection limit of 2MASS, especially in the $K_s$ band. To a moderate extent, this holds for the $ r'-J$ colour, too. We find five stars in the middle panel ($M_J $ vs. $ J-K_s$) that perfectly lie on the Hyades NIR main sequence. We marked these stars in Fig. \ref{cmd} with their numbers from Table~\ref{table:1}. All five are included in our sample of 724 Hyades members in Paper~I. Four of them (1, 4, 7 and 13 or HR 1358, EGGR 38, V471 Tau and WD 0217+375, respectively) are known as binaries containing a WD and a MS component. The first three belong to the classical Hyades members, whereas WD 0217+375 (13) has not be associated with the Hyades before. LP 649-0071 (12) is rated a white dwarf in Luyten's White Dwarf Catalogues, though it has a NIR-colour typical of red dwarfs. This indicates a possible binary nature of this object. We discuss its properties in more detail in section \ref{individual}. \begin{table*} \centering \caption{Hyades white dwarf candidates.} \label{table:1} \begin{tabular}{r c lrrrrrrrrc c c} \hline\hline Star & other name & SpT & D & r$_c$ & v$_\perp$ & $M_V$ & $B-V$ & $M_J$ & $J-K_s$ & $r'-J$& X &Bin.&Ref. \\ No. & & & [pc] & [pc] & $[\rm{km\,s}^{-1}]$ &[mag] &[mag] &[mag] &[mag] &[mag] & & &$B,V$\\ \hline 1 & V471 Tau &DA1.5 & 47.3 & 7.5 & 0.1 & 6.12 & 0.76 & 4.40 & 0.53 & 1.34 &y &y &1,1 \\ 2 & EGGR 26, HZ 4 &DA4 & 35.1 & 13.5 & 0.9 & 11.83 & 0.10 & 12.11 &-0.23 &-0.08 &y &- &2,2 \\ 3 & EGGR 29, LB 227 &DA3 & 51.7 & 6.6 & -0.4 & 11.78 & 0.09 & 12.13 &-0.03 &-0.21 &y &- &3,3 \\ 4 & HR 1358 &DA3 & 49.1 & 3.9 & -0.2 & 2.70 & 0.47 & 1.83 & 0.22 & --- &y &y &1,1 \\ 5 & EGGR 36, VR 7 &DA2.5 & 44.1 & 2.4 & -0.8 & 11.10 &-0.10 & 11.53 &-0.34 &-0.19 &y &- &2,2 \\ 6 & EGGR 37, VR 16 &DA & 47.7 & 1.5 & -0.5 & 10.62 &-0.09 & 11.23 &-0.04 &-0.28 &y &- &2,2 \\ 7 & EGGR 38, HZ 9 &DA2.5 & 43.2 & 3.3 & -0.9 & 10.78 & 0.33 & 7.58 & 0.84 & 3.24 &y &y &2,2 \\ 8 & EGGR 39, HZ 7 &DA2.3 & 45.7 & 3.4 & 0.3 & 10.94 &-0.09 & 11.47 &-0.03 &-0.27 &- &- &2,2 \\ 9 &EGGR 316, LP 475-242 &DB4 & 46.9 & 3.3 & -0.7 & 11.56 &-0.09 & 11.96 & 0.01 &-0.26 &- &- &2,2 \\ 10 & EGGR 42, HZ14 &DA2 & 46.0 & 5.2 & 0.2 & 10.51 &-0.15 & 11.18 &-0.10 &-0.35 &- &- &2,2 \\ \hline 11 & WD 0120-024 &DC & 39.0 & 36.3 & 0.8 & 14.52 & 0.42 & 13.50 & 0.89 & --- &- &- &7,4 \\ 12 & LP 649-0071 &--- & 34.3 & 29.1 & -1.3 & 13.85 & 0.78 & 9.01 & 0.89 & 4.76 &y &? &8,5 \\ 13 & WD 0217+375 &DA & 25.1 & 29.6 & 0.6 & 10.46 & 1.17 & 6.95 & 0.84 & 3.12 &- &y &7,7 \\ 14 & WD 0230+343 &DA & 37.8 & 24.1 & 0.8 & 12.84 & 0.36 & 13.63 & 0.84 &-0.29 &- &- &6,6 \\ 15 & LP 246-0014 &--- & 35.1 & 23.5 & 2.4 & 13.80 & 0.30 & 13.65 &-0.65 &-0.10 &? &- &8,7 \\ 16 & WD 0259+378 &DA3 & 66.0 & 33.6 & -4.0 & 11.44 &-0.03 & 12.35 & 0.64 &-0.37 &- &- &3,3 \\ 17 & WD 0312+220 &DA2.5 & 44.4 & 14.3 & 4.7 & 12.43 & 0.24 & 13.01 &-0.28 &-0.01 &y &- &7,7 \\ 18 & LP 653-0026 &DA3.5 & 28.5 & 23.0 & -1.9 & 12.93 & 0.15 & 13.20 & 0.02 &-0.10 &- &- &3,3 \\ 19 & WD 0348+339 &DA4 & 38.5 & 16.1 & -0.5 & 12.27 & 0.10 & 12.59 &-0.30 &-0.17 &- &- &3,3 \\ 20 & HG7-85 &DA & 38.4 & 9.2 & -1.2 & 12.02 & 0.16 & 12.50 & 0.09 &-0.27 &y &- &0,0 \\ 21 & WD 0433+270 &DC8 & 20.1 & 26.8 & -0.5 & 14.35 & 0.65 & 13.08 & 0.46 & 1.02 &y &- &3,3 \\ 22 & WD 0437+122 &DA & 66.9 & 21.3 & 1.6 & 13.57 & 0.19 & --- & --- & --- &- &- &9,9 \\ \hline 23 & WD 0625+415 &DA3 & 47.7 & 28.9 & -2.9 & 11.60 &-0.03 & 11.95 &-0.17 &-0.20 &- &- &3,3 \\ 24 & WD 0637+477 &DA3.6 & 36.2 & 30.6 & -1.4 & 12.00 & 0.13 & 12.25 &-0.18 &-0.06 &- &- &3,3 \\ 25 & WD 0641+438 &DA & 41.9 & 30.2 & 3.0 & 12.42 & 0.06 & 12.66 &-0.09 &-0.03 &- &- &8,4 \\ 26 & WD 0743+442 &DA5 & 33.5 & 35.6 & 1.3 & 12.24 & 0.10 & 12.60 & 0.05 &-0.14 &- &- &3,3 \\ 27 & WD 0816+376 &DA5 & 38.7 & 39.6 & -0.9 & 12.80 & 0.22 & 12.70 &-1.18 &-0.10 &- &- &3,3 \\ \hline\hline 28 & WD 0233-083.1 &DA & 53.3 & 32.5 & 1.5 & 16.13 & 0.19 & --- & --- & --- &- &- &x,x \\ 29 & WD 0300-083.1 &DA4.4 & 37.9 & 25.0 & -0.0 & 15.93 & 0.07 & --- & --- & --- &- &- &x,x \\ 30 & LP 652-0342 &--- & 31.6 & 25.2 & 0.2 & 14.25 & 0.06 & 14.27 &-0.16 & 0.15 &- &- &5,5 \\ 31 & WD 0533+322 &DA4 & 11.2 & 36.1 & 1.1 & 16.18 & 0.38 & --- & --- & --- &- &- &1,1 \\ 32 & WD 0543+436 &DA5 & 21.3 & 30.3 & -2.0 & 15.45 & 0.14 & 14.76 & 1.01 & 0.30 &- &- &3,3 \\ 33 & WD 0557+237 &DA6 & 13.2 & 34.4 & -2.1 & 16.29 & 0.35 & 15.80 & 0.48 & 0.52 &- &- &2,2 \\ 34 &1RXSJ062052.2+132436 &DA & 29.7 & 24.1 & -0.6 & 12.94 &-0.30 & 13.33 & 0.09 &-0.51 &y &- &2,2 \\ 35 & WD 0758+208 &DA & 38.7 & 36.8 & -0.8 & 17.73 & 0.33 & --- & --- & --- &- &- &x,x \\ 36 & WD 0816+387 &DA6.5 & 19.1 & 38.6 & 3.8 & 15.16 & 0.30 & 14.67 & 0.49 & 0.52 &- &- &3,3 \\ 37 & WD 0820+250.1 &DA1.5 & 28.0 & 38.1 & 3.1 & 13.68 &-0.15 & 14.20 &-0.08 &-0.42 &y &- &7,7 \\ \end{tabular} \tablefoot{Stars No. 1 to 10 are the 10 ''classical'' Hyades white dwarfs as given, e.g. by \citet{1998AJ....115.1536V}. Stars No. 11 to 22 are probably former Hyades white dwarfs that fulfil the kinematic and photometric criteria. Stars No. 23 to 27 do also fulfil the kinematic and photometric criteria, but we rate them as possible non-Hyades because of their long distance from the centre in the $Z$ direction (as was done for MS stars in Paper I). Stars No. 28 to 37 are probable non-Hyades white dwarfs that photometrically do not share the loci of white dwarfs in the colour-magnitude diagrams. The printed table gives the entries necessary for the figures in this paper; the full table with additional information is available from the CDS, Strasbourg, France.} \tablebib{ (0)~\citet{1995ApJ...448..683S}; (1)~\citet{2001KFNT...17..409K}; (2)~\citet{1999ApJS..121....1M}; (3)~\citet{1994cmud.book.....M}; (4)~\citet{2005AJ....129.2428S}; (5)~\citet{2003ApJ...582.1011S}; (6)~\citet{1987AJ.....94..501K}; (7)~\citet{2008AJ....136..735L}; (8)~\citet{2004AAS...205.4815Z}; (9)~\citet{1992MNRAS.257..257R}; (x)~converted from SDSS ugriz. } \end{table*} In Table \ref{table:1} we summarise the data for the white dwarfs discussed in this paper. Column 1 is a running number, column 2 the name(s) of the star in the SIMBAD database, column 3 the spectral type taken from \citet{1999ApJS..121....1M}. In column 4 we present the distance D of the star from the Sun as calculated from the convergent point method, whereas column 5 gives the distance r$_c$ from the cluster centre. Column 6 is the tangential velocity v$_\perp$ perpendicular to the direction to the convergent point. It is a measure of how well the motion of the star and cluster coincide. Columns 7 to 11 give $M_V$, $B-V$, $M_J$, $J-K_s$ and $r'-J$. Column 12 describes whether or not the star is detected as an x-ray source (in VizieR), column 13 whether it is a known spectroscopic binary. Finally, column 14 presents the sources of the $B$ and $V$ magnitudes. For code x,x we used the transformations from ugriz (SDSS) to $B$, $V$ as given in \citet{2005AJ....130..873J}. An extended version of Table~\ref{table:1} is published only in machine-readable form via the CDS. It contains additional entries for each star including, e.g. precise positions, proper motions, apparent magnitudes to ease the preparation of follow-up observations, as well as the velocities derived from the convergent point method. The 37 stars in Table \ref{table:1} are divided into four classes. The first ten stars of class 1 are the ``classical'' Hyades. Stars 11 to 22 form class 2 of new probable Hyades co-movers. The five stars of class 3 fulfil the kinematic and photometric criteria, but are more than 20 pc away from the centre in $Z$ direction (perpendicular to the galactic plane). Stars with these characteristics have been ruled out in Paper I, because all of them had discordant radial velocities (whenever a radial velocity measurement was available). Finally, class 4 consists of 10 stars that fulfil the kinematic criteria, but would be sub-luminous in the CMD if set at their predicted distances. \begin{table}[h!] \caption{Predicted and measured radial velocities for stars from Table \ref{table:1}} \centering \label{table:2} \begin{tabular}{rrcc@{ }r@{ }l} \hline\hline Star & RV$_{\rm{pr}}$ & RV$_{\rm{S}}$ & RV$_{\rm{V}}$ & Ref. & Comments\\ No. &$[\rm{km\,s}^{-1}]$&$[\rm{km\,s}^{-1}]$&$[\rm{km\,s}^{-1}]$& S,V & \\ \hline 1 & 34.4 & 23$\pm$10 & 37.4 & 1,2 &binary\\ 2 & 35.4 & 46.3$\pm$4.2 &46.3$\pm$4.2 & 5,5 & \\ 4 & 38.1 & 37.0$\pm$2 &38.7$\pm$0.3 & 1,6 &binary\\ 7 & 38.9 &-10.4$\pm$9.7 &36.7$\pm$1.5 & 7,3 &cpm, sep. 13\arcsec(?)\\ 8 & 39.5 & 43.5$\pm$4.0 &43.5$\pm$4.0 & 5,5 & \\ 10 & 40.2 &105 &105 & 8,8 &no redshift correction \\ 11 & 9.9 & 9.8$\pm$8.9 &9.8$\pm$8.9 & 7,7 &cpm, sep. 43\arcsec\\ 13 & 19.5 & 5.8$\pm$8.6 &5.8$\pm$8.6 & 7,7 &cpm, sep. 2\arcsec\\ 18 & 33.0 & 50.9$\pm$3.3 &49.9$\pm$4.6 & 4,5 & \\ 21 & 37.6 & - &36.3 & -,9 & \\ 25 & 36.1 &-11.7$\pm$8.3 &-11.7$\pm$8.3 & 7,7 &cpm, sep. 143\arcsec \\ 36 & 34.6 & 19.8$\pm$6.5 & 19.8$\pm$6.5 & 7,7 &cpm, sep. 34\arcsec \\ \hline \end{tabular} \tablefoot{Radial velocities from literature for the stars from Table~\ref{table:1}. The first column gives the star number, the second column (RV$_{\rm{pr}}$) the radial velocity predicted by the convergent point method. In the next three columns we show the radial velocities found in SIMBAD (RV$_{\rm{S}}$), resp. VizieR (RV$_{\rm{V}}$), and the corresponding references. The sixth column gives comments.} \tablebib{ (1)~\citet{WILSON53}; (2)~\citet{2000A&AS..142..217B}; (3)~\citet{1994A&AS..108..603B}; (4)~\citet{2003A&A...400..877P}; (5)~\citet{2006A&A...447..173P}; (6)~\citet{2004A&A...418..989N}; (7)~\citet{2002AJ....124.1118S}; (8)~\citet{1967ApJ...149..283G}; (9)~\citet{2003ApJ...596..477Z} } \end{table} We also checked if measured radial velocities of the 37 candidates were available to compare them with the predicted ones from the convergent point method. Only for 12 of the 37 candidates we found radial velocities in the literature. For six of them (the stars nos. 1, 4, 7, 8, 11, 21), the predicted and the measured radial velocities (from at least one source) agree well. EGGR 42 (10) was assumed to have Hyades radial velocity by \citet{1967ApJ...149..283G} and this was used to obtain its Einstein redshift (see also the remark on this star in the next section). WD~0816+387 (36) was already rejected as a Hyades member by the photometric criteria, so a disagreement between measured and predicted radial velocities is to be expected. On the other hand, the discordant radial velocities for the stars nos. 2, 13, 18, and 25 require a more detailed discussion. A reliable determination of space velocities of white dwarfs is a challenging task. For isolated white dwarfs the apparent radial velocities must be corrected for gravitational redshift, which requires the knowledge of the mass-radius ratios, i.e. quantities that cannot be observed directly. For stars nos. 2, 8, and 18, \citet{2006A&A...447..173P} determined radial velocities from high-resolution spectra, whereas spectroscopic distances and gravitational redshifts were computed from the fundamental parameters derived by \citet{2001A&A...378..556K}. The relatively high radial velocity for EGGR 26 (2) by \citet{2006A&A...447..173P} would reject this star as a Hyades member, though in numerous studies its membership is found to be confirmed \citep[e.g.,][] {1992AJ....104.1876W,2009ApJ...696...12D}. The discrepancy for EGGR 26 may probably be explained by underestimated uncertainties introduced when deriving the redshift corrections. The same reason for discrepancy may possibly hold for LP~653-0026 (18), too: recently, \citet{2009A&A...505..441K} published an updated version of their catalogue of the fundamental parameters of white dwarfs where two different sets of parameters were considered to be equally probable for this star. A re-calculation of the radial velocities seems to be reasonable for white dwarfs from \citet{2006A&A...447..173P}. For four of our candidates (the stars nos. 7, 11, 13, 25), radial velocities were obtained by \citet{2002AJ....124.1118S} from line-of-sight velocities of M dwarfs in common proper motion pairs (cpm), each consisting of an M-dwarf and a white dwarf. The authors assume that typical separations between the components are about 1000 AU, such that orbital motion can be neglected. For star no. 13, where measured and predicted radial velocities differ by 1.6$\sigma$, the separation is 2\arcsec, corresponding to about 50 AU, given a distance of 25 pc. Here the orbital motion cannot be neglected, and even a small correction of a few $\rm{km\,s}^{-1}$ could make the difference between measured and predicted radial velocities insignificant. On the other hand, if the separation is large in a cpm pair, the argument of a common radial velocity becomes weaker because of an increasing probability of an unphysical optical pair. This could be the case of star no.~25, WD~0641+438, where the separation between the white dwarf and its MS companion reaches 143\arcsec. The PPMXL lists more or less compatible motions for these stars ($\mu = 139$~mas/yr, $\Theta = 180$~deg; $\mu = 104$~mas/yr, $\Theta = 183$~deg), respectively. However, the 2MASS colours for the MS star indicate that this star should be a late K dwarf at a distance of at least 500~pc, which excludes them as physical binary. Finally, the reason for the discrepancy between the predicted and measured radial velocity \citep{2002AJ....124.1118S} for HZ~9 (7) seems to be a misinterpretation. \citet{2002AJ....124.1118S} regarded it as a cpm pair with another star 13\arcsec away. However, the proper motion of the latter is completely different (11.4 mas/y) from the proper motion of the white dwarf HZ~9 (115 mas/y). On the other hand, \citet{1987AJ.....94..996S} analysed the radial-velocity curve of HZ~9, confirmed its binary nature, and determined an M dwarf--white dwarf separation of less than 1~AU and a radial velocity of 36.7~$\rm{km\,s}^{-1}$ for the system, which agrees well with the predicted radial velocity. To summarise the above discussion: none of the 10 classical candidates and of the 17 new probable former Hyades white dwarfs can be unambiguously discarded on the basis of the presently measured radial velocities. Given the problematics of obtaining the true radial motion of the candidates, we can only encourage new measurements for which this paper may serve as an input catalogue. \section{Individual stars}~\label{individual} \begin{description} \item [1 = V471 Tau = WD~0347+171] is a spectroscopic binary K2V+DA \citep{2006MNRAS.367.1699H}, included in our sample of 724 Hyades members from Paper I, a strong X-ray source LX~45 = $229.6 \pm 10.0$. LX~45 is the X-ray luminosity from \citet{1995ApJ...448..683S}, derived from the assumption that the star has a heliocentric distance of 45 pc. Units are $10^{28}$ erg\,s$^{-1}$ (0.1-1.8keV). Its trigonometric parallax from Hipparcos \citep{2009A&A...497..209V} of 22.7 $\pm$ 1.5 mas agrees well with the predicted one. \item [2 = EGGR 26 = HZ 4 = WD~0352+096] is considered as a certain member in e.g. \citet{1992AJ....104.1876W} and \citet{2009ApJ...696...12D}, though seems just to leave the Hyades (the estimated distance from the cluster centre is 13.5~pc). A weak X-ray source LX~45 $<$ 1.4. \item [ 3 = EGGR 29 = LB 227 = WD~0406+169:] \citet{2009ApJ...696...12D} use it in their analysis, whereas \citet{1992AJ....104.1876W} declare it to be a non-member. \citet{1992AJ....104.1876W} reject this white dwarf as a Hyades member since the mass they derived from the surface gravity sets the star to a distance of 60~pc from the Sun, with a velocity difference to the adopted cluster velocity higher than 12~$\rm{km\,s}^{-1}$. Using the convergent point method, we obtain a distance of 52~pc from the Sun, a distance from the cluster centre of 7~pc, and the velocity difference of less than 0.5~$\rm{km\,s}^{-1}$ for EGGR 29. These results support the assumption of its Hyades membership. A weak X-ray source LX~45 $<$ 0.9. \item [ 4 = HR 1358 = HD 27483 = WD~0418+137:] this system consists of two F6V stars with orbital period of 3.05 days, and a DA3 white dwarf companion \citep{1993AJ....106.1113B}. The MS binary is included in the sample of the 724 Hyades members from Paper I, X-ray source LX~45 $= 19.9 \pm 2.9$. Its trigonometric parallax from Hipparcos \citep{2009A&A...497..209V} of 21.1 $\pm$ 0.5 mas agrees well with the predicted one. \item [ 5 = EGGR 36 = VR 7 = WD~0421+162] is considered to be a certain Hyades member in \citet{1992AJ....104.1876W} and \citet{2009ApJ...696...12D}, a weak X-ray source LX~45 $<$ 1.4. \item [ 6 = EGGR 37 = VR 16 = WD~0425+168] is used as a certain Hyades member in \citet{1992AJ....104.1876W} and \citet{2009ApJ...696...12D}, X-ray source LX~45 = 4.2. \item [ 7 = EGGR 38 = HZ 9 = WD~0429+176] is a spectroscopic binary DA2.5+dM, included in the sample of the 724 Hyades members from Paper I, X-ray source LX~45 = 2.7. \item [ 8 = EGGR 39 = HZ 7 = WD~0431+126] is used as a certain Hyades member in \citet{1992AJ....104.1876W} and \citet{2009ApJ...696...12D}, a weak X-ray source LX~45 $<$ 0.9. \item [ 9 = EGGR 316 = LP 475-242 = WD~0437+138] is accepted as a Hyades member \citep{1998AJ....115.1536V}, though \citet{1992AJ....104.1876W} do not discuss it, because they are rating the data available for the star as too uncertain. A weak X-Ray, LX45 $<$ 0.9. \item [ 10 = EGGR 42 = HZ 14 = WD~0438+108] is used as a certain Hyades member in \citet{1992AJ....104.1876W} and \citet{2009ApJ...696...12D}, a weak X-Ray LX45 $<$ 1.0. Its radial velocity has been determined by \citet{1967ApJ...149..283G}, who give an apparent radial velocity of 105 $\rm{km\,s}^{-1}$. In this case the authors did not try to determine the Einstein redshift from (M/R), but assumed that it is a Hyades member, and must therefore have the Hyades radial velocity. \item [ 11 = WD 0120-024] is one of the absolutely faintest white dwarf candidates in our Hyades sample, it is not observed in CMC14. Its predicted and observed radial velocities agree well. \item [ 12 = LP 649-0071:] at the position of this object we find the MS-star no. 8 ($M = 0.17\, M_\odot$) of our sample from Paper I, and at the same time there is the blue object LP 649-0071 ($m_{pg} = 16.9$) from Luyten's White Dwarf Catalogues. The proper motions of the blue Luyten object and our red object coincide remarkably. We took $B=17.31$ from NOMAD and $V=16.53$ from \citet{2003ApJ...582.1011S}. The red component has a parallax of 29.13~mas $\pm$ 0.41~mas. This object is found as 2XMMi J021352.1-033059 in the XMM Newton Serendipitous Source Catalogue 2XMMi-DR3, with a flux of 4.0877$\times 10^{-14}$ mW\,m$^{-2}$ (0.2-12keV). Galex \citep{2007ApJS..173..682M} finds a faint object with FUV magnitude of 24.0 and NUV magnitude of 23.1 at the position of this star, probably too faint for a white dwarf with this parallax. The binary nature of this object as well as its white dwarf nature have to be verified. \item [ 13 = WD 0217+375] is a component of a close binary with a separation of 2\arcsec, not resolved in our catalogue. According to \citet{2005AJ....129.2428S}, the spectral type is M5V+DA. A parallax of 39.8~mas is predicted by the convergent point method. This agrees well with the parallax of 40~mas given for this star in the CNS3 \citep{1991adc..rept.....G}. \item [ 14 = WD 0230+343:] 2MASS photometric flags ``ACU''. \item [ 15 = LP 246-0014] is a high proper motion ($\mu_{\alpha}\cos \delta = 228.4$~mas/yr, $\mu_{\delta} = -50.8$~mas/yr) blue star, listed in Luyten's White Dwarf Catalogues. It is located in between two brighter stars. From this region a strong X-ray emission was measured by ROSAT, but it is not clear which of these objects is an X-ray source. There are no Galex observations in the area around this star. The white dwarf nature of LP 246-0014 has to be verified. \item [ 16 = WD 0259+378] is one of the white dwarfs most distant to the Sun in our candidate sample. It has a relatively high residual velocity of 4 $\rm{km\,s}^{-1}$ with respect to the Hyades, which is probably a reason for its ``unusual'' location $X, Y = $ -54pc, 32pc in Fig.~\ref{xyz}. Its 2MASS photometry is highly uncertain, especially in the $H$ and $K_s$ bands, which have photometric flags ``UD''. \item [ 17 = WD 0312+220:] 2MASS photometric flags ``BCU''. \item [ 18 = LP 653-0026 = WD~0339-035:] 2MASS photometric flags ``ABC''. \item [ 19 = WD 0348+339:] 2MASS photometric flags ``ABD''. \item [ 20 = HG7-85 = LP 474-95?] is a white dwarf observed by \citet{2009A&A...505..441K}, HS0400+1451 (Hamburg Schmidt survey). This star, first mentioned by \citet{1962LowOB...5..257G} as a member of the Hyades, is found in the Prosser \& Stauffer data base (presently available from J. Stauffer, priv. comm.). The authors identify it with LP 474-95, a star which cannot be found in the CDS database. It is also not contained in \citet{1999ApJS..121....1M}. The proper motions in PPMXL are incorrect. Therefore, we took the proper motions from \citet{2006A&A...448.1235D}. The star is also contained in \citet{1995ApJ...448..683S}, which give an X-ray luminosity $<$ 1.1 if a distance of 45 pc is assumed. \item [ 21 = GJ 171.2 B = EGGR 40 = WD 0433+270:] this star forms a cpm pair with BD+26 730, which is included as no. 461 in Paper I. The measured trigonometric parallax and radial velocity of BD+26 730 agree well with the predicted ones. Also, the measured radial velocity of the white dwarf (Table \ref{table:2}) coincides well with the predicted one. With spectral type DC8, the white dwarf WD 0433+270 is the reddest in our candidate sample. Its possible membership in the Hyades is extensively discussed in \citet{2008A&A...477..901C}. We further discuss its importance for the sample in Sect. \ref{discus}. \item [ 22 = LP475-249 = WD 0437+122 = Reid 405] is the most distant white dwarf in our candidate sample, too faint to be measured in 2MASS and CMC14. A possible membership in the Hyades was discussed by \citet{1992MNRAS.257..257R}. Owing to a relatively high velocity with respect to the cluster, this star was excluded as a Hyades member. Based on new proper motions from PPMXL, the residual velocity v$_\perp$ turns out to be about 1.63~$\rm{km\,s}^{-1}$ which is consistent with the Hyades motion. \item [ 23 = WD 0625+415:] this white dwarf has proper motions consistent with Hyades membership. Also, its location in the $M_V $ vs. $ B-V $ diagram indicates a correct distance predicted by the convergent point method. However, at $z = 11.3$~pc WD 0625+415 is more than 20~pc above the cluster centre. In Paper~I we found that all stars with $\Delta\,z > 20$~pc should be rejected as Hyades members. Fig.~\ref{vdisp} gives an additional argument that this star is probably a field white dwarf, though the radial velocity must be measured to support this assumption. \item [ 24 = WD 0637+477] is probably a field white dwarf, the same case as star no. 23 above. \item [ 25 = WD 0641+438] is probably a field white dwarf, the same case as star no. 23 above. \item [ 26 = WD 0743+442] is probably a field white dwarf, the same case as star no. 23 above. \item [ 27 = WD 0816+376] is probably a field white dwarf, the same case as star no. 23 above. \item [ 28 = WD 0233-083.1] is rejected as a Hyades candidate because of its location in the $M_V $ vs. $ B-V $ diagram. The predicted distance seems to be underestimated. No 2MASS, CMC14 and reliable $B, V$ measurements are found. $M_{V}$ and $B-V$ are estimated from SDSS ugriz. \item [ 29 = WD 0300-083.1:] the same case as star no. 28 above. \item [ 30 = LP 652-0342] is rejected as a Hyades candidate because of its location in the $M_V $ vs. $ B-V $ diagram. The predicted distance is underestimated. This is supported by the trigonometric parallax (3.9 $\pm$ 4.2 mas) from \citet{1995gcts.book.....V}. \item [ 31 = WD 0533+322] is rejected as a Hyades candidate because of its location in the $M_V $ vs. $ B-V $ diagram. The predicted distance seems to be underestimated. No 2MASS, CMC14 measurements. \item [ 32 = WD 0543+436] is rejected as a Hyades candidate because of its location in the $M_V $ vs. $ B-V $ diagram. The predicted distance is underestimated. \item [ 33 = WD 0557+237:] the same case as star no.32 above. \item [ 34 = 1RXSJ062052.2+132436:] the same case as star no.32 above. \item [ 35 = WD 0758+208:] the same case as star no.28 above. \item [ 36 = WD 0816+387:] the same case as star no.32 above. \item [ 37 = WD 0820+250:] the same case as star no.32 above.\\ \end{description} \noindent Other stars: \begin{description} \item [vA54 = HG7-128 = LP 474-185] is an M5V star \citep{2010yCat....102023S}, has Rosat observation (LX45 $<$ 2.2). This is star number 170 of Paper I. van Altena finds in his first paper \citep{1966AJ.....71..482V} $ B-V $ = 1.82, later he corrects it to $ B-V $ = 0.90 \citep{1969AJ.....74....2V}. This is definitely a red dwarf, and there is no indication for a white dwarf companion. \item [ vA71 = EGGR 32 = WD 0412+14] is classified as sdK: by \citet{2010yCat....102023S}. \citet{1975ApJ...200L..95L} rates it as a very metal-poor subdwarf with ``K-star'' colour, but with strong, sharp hydrogen lines. The General Catalog of Trigonometric Parallaxes \citep{1995gcts.book.....V} gives a parallax of 0.003 $\pm$ 0.004 arcseconds. This star is not a member of the Hyades. No X-ray detection. \item [ RHya 102 = HG7-126:] this star fails the kinematic criterion to be included as Hyades member. Its tangential velocity v$_\perp$ is $-6.1~\rm{km\,s}^{-1}$. The convergent point method puts it at a distance D from the Sun of 60.8 pc. With this distance it has $M_V $ = 12.26; and with $ B-V = -0.08$ it perfectly fits the CMD in Fig.~\ref{cmd}. It also fits the luminosity-distance relation in Fig.~\ref{vdisp}, see Sect. \ref{discus}. We discard it, however, because of kinematic reasons. \item [ RHya 154:] the proper motions given by \citet{1992MNRAS.257..257R}, (76,$-34$) mas/y and those from the CU subset (67, $-46$) disagree. In consequence the v$_\perp$-component of the tangential velocity is $-3.1~\rm{km\,s}^{-1}$ for Reid, and $-8.1~\rm{km\,s}^{-1}$ in the CU. One would count it as a kinematic member with Reid's data, and discard it with ours. The distance from the Sun is D = 66 pc (Reid), 70.6 pc (CU). Reid gives $V$ = 16.51 and $ B-V = -0.22$, which converts into $M_V $ = 12.26, and puts it 1~mag below the white dwarf sequence. So, we count it as a non-member. \end{description} \section{Spatial distribution and completeness}~\label{complete} In Fig. \ref{xyz} we show the distribution of the candidates from Table \ref{table:1} (only stars no. 1 to 27) on top of the background of the 724 members from Paper I. We use the galactic rectangular coordinate system $X, Y, Z$ with origin in the Sun, and axes pointing to the Galactic Centre ($X$), to the direction of galactic rotation ($Y$), and to the North Galactic Pole ($Z$). All classical white dwarfs except one are located within the tidal radius of the cluster. All newly found candidates lie outside the tidal radius, hence they are no longer gravitationally bound to the cluster, but share the fate of hundreds of former main-sequence members that left the bound region. The five probable field white dwarfs (Nos. 23 to 27) are all at z~$>$~10~pc. \begin{figure}[h!] \centering \includegraphics[bb=47 109 564 640,angle=270,width=9.0cm,clip]{hwd_xyz.ps} \caption{The spatial distribution in the galactic $X, Y, Z$ coordinate system of the Hyades white dwarf candidates of this paper. They are marked by the same symbols as in Fig. \ref{cmd}. The small (red) dots in the background distribution represent the 724 stars from Paper I. Probable field white dwarfs (nos. 23 to 27) are additionally marked by open circles. The two large circles display the tidal radius (9~pc), and a radius of 30~pc, as in Paper I. } \label{xyz} \end{figure} Of particular interest here is the distribution in the $X,Y$-plane. We note that all white dwarfs (except no. 16) follow the tilted distribution of the main-sequence stars. By tidal interaction with the gravitational field of the Galaxy, stars can leave the cluster on both sides via the Lagrangian points L$_1$ and L$_2$ of the Galaxy-cluster-star system, where L$_1$ is in the direction to the Galactic centre, i.e. towards larger (less negative) $X$, the Sun-facing side of the cluster, while L$_2$ lies on the opposite side of the cluster centre. All white dwarfs outside the tidal radius (except nos. 16 and 22) populate the Sun-facing part of the cluster, and may have left it through L$_1$. The deficit of newly found candidates at longer distances from the Sun needs explanation. To investigate this we plot in Fig.~\ref{photlim} the $r'$, $J$ and $ K_s$ magnitudes as a function of the distance D from the Sun. The background points in Fig.~\ref{photlim} represent again the sample of 724 from Paper I. In the NIR distributions we note that the magnitude limit of the sample of 724 in the $J$ and $ K_s$ bands is at much brighter magnitudes than the 2MASS completeness limit of $J$ = 15.8 and $ K_s$ = 14.3 (see http://www.ipac.caltech.edu/2mass/releases/allsky/doc). For fainter red dwarfs or even brown dwarfs there was no optical counterpart in CMC14, i.e. in the CU subset to PPMXL. This is different with the white dwarfs. We see from Fig.~\ref{photlim} that the fainter, hitherto unknown Hyades white dwarfs all are well beyond the 2MASS completeness limit in $J$ and $ K_s$. The photometric accuracy in the $ K_s$ band at 16.0 typically is 0.25 mag, fainter ones have no accuracy estimate at all. The situation is somewhat better in the $J$ band. In the $r'$ band the red and white dwarfs are comparable. The faintest white dwarfs are near the completeness limit of CMC14 at $r'$ = 16.8. With these remarks it becomes clear that we can reveal new Hyades white dwarfs beyond a distance of about 50 pc from the Sun in the CU subset of PPMXL only by chance. \begin{figure}[h!] \centering \includegraphics[bb=53 56 550 585,angle=270,width=9cm,clip]{wd_mag_dist.ps} \caption{Distribution of the apparent magnitudes $r'$, $J$ and $ K_s$ over the distance D from the Sun. The stars from Table~\ref{table:1} are marked by the same symbols as in Fig. \ref{cmd}. The small (red) dots in the background distribution represent the 724 stars from Paper I. } \label{photlim} \end{figure} The PPMXL goes about 3 magnitudes deeper than its CU subset, so possibly white dwarfs with Hyades motion could be found therein. However, PPMXL photometry in optical bands is from USNO-B1.0, and therefore inappropriate for this kind of work. These distant white dwarfs could only be found by cross-matching PPMXL kinematic candidates with the catalogue from \citet{1999ApJS..121....1M} if the latter would be complete down to the limiting magnitude of PPMXL. However, except for its SDSS part, which only marginally covers the Hyades area, the McCook\&Sion catalogue is quite incomplete already at $V \approx$ 16 \citep[see, e.g.][]{2003Msngr.112...25N}. Consequently, our Table~\ref{table:1} contains only two white dwarfs farther away than 50 pc from the Sun. \section{Discussion}~\label{discus} The convergent point method supplemented by photometric selection provides five out of six phase-space parameters. Generally, this allows quite a reliable selection of open cluster members. We find that only nine ``classical'' Hyades white dwarfs reside within the tidal radius of the cluster (r$_c < 9$~pc), hence are tidally bound. Outside the tidal radius, up to a distance of 40 pc from the centre, we find 18 white dwarfs co-moving with the cluster at relatively low velocity dispersion (cf. Fig.~\ref{vdisp}, top panel). When \citet{2008A&A...477..901C} discussed WD 0433+270 (no.~21 in Table~\ref{table:1}), this star was very isolated in the $M_V $ vs. $ B-V $ colour-magnitude diagram of Fig. \ref{cmd}. There was a large gap between the reddest classical white dwarf EGGR 26 (no.~2) at $M_V $= 11.83 and $ B-V $= 0.10, and WD 0433+270 at $M_V $ = 14.35 and $ B-V $ = 0.65. With our new candidates the gap does no longer exist, and, from kinematics, we strongly infer that WD 0433+270 was in fact a former member of the Hyades with all the implications that \citet{2008A&A...477..901C} rate as tantalising. Although the question of the cooling age is most critical for WD 0433+270, if it is an ejected member of the Hyades cluster, the cooling ages of the other candidates between the dimmest classical white dwarf, EGGR 26 \citep[$3.1\times10^8$ yr, see][]{1992AJ....104.1876W} and WD 0433+270 should be re-discussed after they are confirmed by their radial velocities. Note that WD 0433+270 is the most nearby star of all our candidates at a distance of 20.1 pc from the Sun. Given the incompleteness of the catalogue by \citet{1999ApJS..121....1M}, it may not be surprising to detect other ``red" white dwarfs of the Hyades once a deep, accurate optical photometric survey like, e.g. PanSTARSS becomes available. The one-dimensional velocity dispersion in Fig. \ref{vdisp} increases with increasing distance from the cluster centre. This behaviour is similar to that of the red dwarfs as shown in Fig. 12 of Paper I. From this we infer that the white dwarfs we reveal here can leave the cluster by the same mechanism as the red dwarfs do. Once the progenitor star develops into a white dwarf and its envelope is pushed away, it will be treated within the cluster as a low-mass object such as the other low-mass stars that are preferentially ejected from the cluster compared to their higher mass brothers. This may mean that, in general, an additional mechanism as proposed by \citet{2003ApJ...595L..53F} is not needed to explain the white dwarf distribution we find. It is not ruled out that the dynamical process of \citet{2003ApJ...595L..53F} has not been active in the Hyades, but the kilometre-per-second kicks that the stars got would very probably move them away from the centre much faster, so we would be unable to find most of them with the constraints we adopted. In Paper I we found mass segregation for giants and main-sequence stars in the Hyades, i.e., a strong concentration of the most massive stars ($M > 2\,M_{\odot}$) towards the cluster centre and flatter distributions for lower mass stars. Usually, such a concentration of the massive stars is observed already in the first few $10^7$ yr of a cluster's life (within the relaxation time scale). Since these stars are the progenitors of white dwarfs, one expects to find recently formed white dwarfs in the vicinity of the cluster centre. However, once they are no longer massive, they behave like other 0.6 to 0.8~$M_{\odot}$ main-sequence stars. Owing to tidal interaction with the gravitational field of the Galaxy, the chance of evaporation from the cluster becomes higher, and it is increasing with the time passed after degeneration. Therefore, merely from the point of view of dynamical evolution, we could expect the older white dwarfs at longer distances from the cluster centre. For the white dwarfs of this paper we find that only the absolutely brightest white dwarfs still are within the tidal radius, whereas the dimmer ones left the cluster. As we already noted above, the dimmest of the classical white dwarfs, EGGR 26, has a mass of 0.62 $M_\odot$ and a cooling age of $3.1\times10^8$ yr \citep{1992AJ....104.1876W}. We find it 13.5 pc away from the cluster centre, so it is already outside the tidal radius of the cluster. In the lower panel of Fig. \ref{vdisp} we see that the absolute magnitude $M_V $ of white dwarfs increases with increasing distance from the centre, to show that possibly the more distant (from the centre) white dwarfs had more time to move away from the cluster centre and to cool down. Those must have formed earlier from more massive progenitors. This empirical luminosity-distance relation has approximately a slope of 4.5 mag in 40 pc. In Fig. \ref{vdisp} (bottom) the five stars marked with their running numbers from Table \ref{table:1} are binaries, their absolute brightness in $M_V $ is not representative for the white dwarf component. However, there is a group of six stars in the bottom panel of Fig.\ref{vdisp} that do not follow the simple luminosity-distance relation of the others. They lie roughly between 30 and 40 pc from the centre at $M_V $ between 11.5 and 13.0 mag. Five of these stars (nos. 23 to 27) are marked as possible field stars because they are far away from the centre in $Z$ direction. Kinematic and photometric main-sequence candidates have been rejected in Paper I with the same argument. The loci of stars nos. 23 to 27 in Fig.\ref{vdisp} (bottom) give additional arguments to rule them out and mark them as field white dwarfs. The sixth star, WD~0259+378 (no. 16), has low $z-z_{\rm{centre}}$, which is why we keep it as a probable member for the time being. We note that these six stars have proper motions and loci in the $M_V$ vs. $B-V$ diagram that are consistent with Hyades membership. With the radial velocities measured, one will be able to decide on their membership with more reliability. On the other hand, we should not exclude the possibility that these stars experienced an additional kick when leaving the cluster, and WD~0259+378 may be a good example of the dynamical mechanism proposed by \citet{2003ApJ...595L..53F}. As has been explained in Sect. \ref{complete}, we cannot make a claim for the completeness of our sample of Hyades white dwarf candidates. However, the probability seems to be low to detect new white dwarfs of $M_V < 12$ within 10~pc from the cluster centre (or 36~pc $<$ D $<$ 56~pc): the apparent magnitudes of these stars would clearly be $V < 16$ where the catalogue of \citet{1999ApJS..121....1M} is nearly complete in this region. Indeed, the number of white dwarfs in front of and behind the cluster centre (D = 46.3~pc) is well balanced within r$_c < 10$~pc. The detection of dimmer white dwarfs is, however, biased towards shorter distances D from the Sun. If the population of white dwarfs behind the centre were similar to that in front of it, and the empirical luminosity-distance relation is valid, we estimate that one will find in the future some 8 to 12 more white dwarfs within a radius of 40 pc from the centre of the cluster. Depending on whether the white dwarfs (nos. 23 to 27) are excluded from, or are included in the consideration, this would yield a total number of 30 to 40 white dwarfs as present-day plus former members of the Hyades. This number coincides well with the postulation of \citet{1992AJ....104.1876W}, who claimed that there should be at least 21 white dwarfs dimmer than their seven confirmed Hyades white dwarfs. \begin{figure}[h!] \centering \includegraphics[bb=40 52 480 506,angle=270,width=8cm,clip]{disp_wd.ps} \caption{As a function of the distance from the cluster centre (r$_c$) this figure shows the distribution of v$_\perp$, the velocity component perpendicular to the direction to the convergent point (upper panel); the absolute magnitudes $M_V $ derived from the secular parallaxes (lower panel). White dwarf candidates with distances D $>$ 46.3 pc from the Sun are additionally marked by larger circles. Spectroscopic binaries are marked by their numbers in Table~\ref{table:1}. The red line shows an approximate fit to an empirical magnitude-distance relation explained in the text.} \label{vdisp} \end{figure} Although we did not detect new white dwarf candidates within the tidal radius, but only in a 30 pc sphere around the centre of the Hyades, this result is similar to that of Paper I. There we found that, at present, 364 main-sequence stars (275 $M_\odot $) are gravitationally bound, and 360 stars (160 $M_\odot $) are co-moving outside the present-day tidal radius of the cluster. This is qualitatively consistent (see Fig. 8 in paper I) with N-body simulations of an open cluster comparable to the Hyades \citep{2009A&A...495..807K}. So, we can expect that white dwarfs are also subject to cluster evaporation as main-sequence stars are. As an alternative, \citet{2007A&A...461..957F} proposed that stars outside the tidal radius of the Hyades, but co-moving in space with the bulk Hyades motion, could be older field stars trapped in orbital resonance with the Hyades cluster, a mechanism already described by \citet{1998AJ....115.2384D}. With the observations we have so far we cannot decide upon the relative efficiency of the two mechanisms, evaporation or capture. Radial velocity measurements are needed to confirm the co-moving, but cannot distinguish between the two mechanisms either. Only the determination of the chemical composition (chemical tracking) of the co-moving stars outside the tidal radius will finally decide about the origin at least for the main-sequence stars. \citet{2008A&A...487..557P} found empirically that typical open clusters lose between 3 to 14 $M_\odot\,$Myr$^{-1}$ into the field in the first 260 Myr of their life. So, the 160~$M_\odot $ in the 30~pc volume around the centre can be easily explained with the Hyades lifetime of some 650 Myr. The capture mechanism must be at least as efficient as that to compete with evaporation. To summarise: Within the tidal radius of the Hyades we only find nine ``classical'' bright white dwarfs. It is very improbable that, at present, more white dwarfs brighter than $M_V $ = 12 are tidally bound in the cluster. Outside the tidal radius we find 18 white dwarfs that are co-moving with the Hyades cluster and could be former tidally bound members. As a consequence of our selection process, the sample presented here is incomplete and is essentially restricted to the Sun-facing part of the cluster. We find five white dwarfs in binary systems, three were already known as Hyades members, two are new candidates. Again, this search is incomplete, because we can reveal them only if the white dwarf nature of one of the components is already known. There is an empirical luminosity-distance (from cluster centre) relation such that the white dwarfs are dimming by about 1~mag per 10 pc distance from the centre. Given the spatial incompleteness of our sample, we estimate that some 20 to 30 white dwarfs should co-move with the bulk Hyades motion in a volume between 9 pc (tidal radius) and 40 pc from the centre. This number is consistent with an extrapolation of the present day mass function (PDMF) of the cluster (Fig. 10 of Paper I) towards white dwarf progenitors. For a full confirmation of the newly found candidates, more measurements of radial velocities are needed. At present, none of the 10 classical candidates and of the 17 new probably former Hyades white dwarfs can be excluded from membership on the basis of the available measurents of radial velocities. For white dwarfs, the measurements of apparent radial velocities must be corrected for gravitational redshift. This correction requires determinations of the mass-radius ratio for each object. This, of course, may introduce additional uncertainties in the determination of the true (kinematic) radial velocity. Once the candidates are confirmed, theories of white dwarf evolution are challenged to explain their nature and their origin. \begin{acknowledgements} We thank U. Heber for a helpful discussion. This research has made extensive use of the SIMBAD database, operated at CDS, Strasbourg, France. We have made use of the NASA/ IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. \end{acknowledgements}
\section*{Appendix \thesection\protect\indent #1} \addcontentsline{toc}{section}{Appendix \thesection\ \ \ #1} } \newcommand\encadremath[1]{\vbox{\hrule\hbox{\vrule\kern8pt \vbox{\kern8pt \hbox{$\displaystyle #1$}\kern8pt} \kern8pt\vrule}\hrule}} \def\enca#1{\vbox{\hrule\hbox{ \vrule\kern8pt\vbox{\kern8pt \hbox{$\displaystyle #1$} \kern8pt} \kern8pt\vrule}\hrule}} \newcommand\figureframex[3]{ \begin{figure}[bth] \hrule\hbox{\vrule\kern8pt \vbox{\kern8pt \vbox{ \begin{center} {\mbox{\epsfxsize=#1.truecm\epsfbox{#2}}} \end{center} \caption{#3} }\kern8pt} \kern8pt\vrule}\hrule \end{figure} } \newcommand\figureframey[3]{ \begin{figure}[bth] \hrule\hbox{\vrule\kern8pt \vbox{\kern8pt \vbox{ \begin{center} {\mbox{\epsfysize=#1.truecm\epsfbox{#2}}} \end{center} \caption{#3} }\kern8pt} \kern8pt\vrule}\hrule \end{figure} } \newcommand{\rf}[1]{(\ref{#1})} \newcommand{\eq}[1]{Eq.~(\ref{#1})} \newcommand{\rfig}[1]{fig.~\ref{#1}} \newcommand{\equ}[2]{\begin{equation}{\label{#1}}{#2}\end{equation}} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\end{equation}}{\end{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\end{eqnarray}}{\end{eqnarray}} \newcommand\eol{\hspace*{\fill}\linebreak} \newcommand\eop{\vspace*{\fill}\pagebreak} \newcommand{\hspace{0.7cm}}{\hspace{0.7cm}} \newcommand{\vspace{0.7cm}}{\vspace{0.7cm}} \renewcommand{\and}{{\qquad {\rm and} \qquad}} \newcommand{{\qquad {\rm where} \qquad}}{{\qquad {\rm where} \qquad}} \newcommand{{\qquad {\rm with} \qquad}}{{\qquad {\rm with} \qquad}} \newcommand{{\qquad {\rm for} \qquad}}{{\qquad {\rm for} \qquad}} \newcommand{{\qquad , \qquad}}{{\qquad , \qquad}} \newcommand{{\it i.e.}\ }{{\it i.e.}\ } \newcommand{{\,\rm Det}} \newcommand{{\,\rm Tr}\:}{{\,\rm Tr}\:}{{\,\rm Det}} \newcommand{{\,\rm Tr}\:}{{\,\rm Tr}\:} \newcommand{{\,\rm tr}\:}{{\,\rm tr}\:} \newcommand{{\,\rm cte}\,}{{\,\rm cte}\,} \newcommand{\mathop{\,\rm Res\,}}{\mathop{\,\rm Res\,}} \newcommand{\td}[1]{{\tilde{#1}}} \renewcommand{\l}{\lambda} \newcommand{\omega}{\omega} \newcommand{{\cal P}}{{\cal P}} \newcommand{{\mathrm{i}}}{{\mathrm{i}}} \newcommand{{\,\rm e}\,}{{\,\rm e}\,} \newcommand{\ee}[1]{{{\rm e}^{#1}}} \renewcommand{\d}{{{\partial}}} \newcommand{{{\hbox{d}}}}{{{\hbox{d}}}} \newcommand{\dmat}[2]{\mathrm{d}_{\scriptscriptstyle{#1}}[#2]} \newcommand{{\int\kern -1.em -\kern-.25em}}{{\int\kern -1.em -\kern-.25em}} \newcommand{\mathrm{Vol}}{\mathrm{Vol}} \newcommand{\mathop{\mathrm{Pol}}}{\mathop{\mathrm{Pol}}} \newcommand{{\cal Z}}{{\cal Z}} \newcommand{{\cal N}}{{\cal N}} \newcommand{\moy}[1]{\left<{#1}\right>} \renewcommand{\Re}{{\mathrm{Re}}} \renewcommand{\Im}{{\mathrm{Im}}} \newcommand{{\rm sn}}{{\rm sn}} \newcommand{{\rm cn}}{{\rm cn}} \newcommand{{\rm dn}}{{\rm dn}} \newcommand{\ssq}[1]{{\sqrt{\sigma({#1})}}} \renewcommand{\l}{\lambda} \renewcommand{\L}{\Lambda} \renewcommand{\ssq}[1]{{\sqrt{\sigma({#1})}}} \newcommand{\overline}{\overline} \renewcommand{\thesection}{\arabic{section}} \renewcommand{\theequation}{\arabic{section}-\arabic{equation}} \makeatletter \@addtoreset{equation}{section} \makeatother \newtheorem{theorem}{Theorem}[section] \newtheorem{conjecture}{Conjecture}[section] \newtheorem{remark}{Remark}[section] \newtheorem{proposition}{Proposition}[section] \newtheorem{lemma}{Lemma}[section] \newtheorem{corollary}{Corollary}[section] \newtheorem{definition}{Definition}[section] \def\begin{remark}\rm\small{\begin{remark}\rm\small} \def\end{remark}{\end{remark}} \def\begin{theorem}{\begin{theorem}} \def\end{theorem}{\end{theorem}} \def\begin{definition}{\begin{definition}} \def\end{definition}{\end{definition}} \def\begin{proposition}{\begin{proposition}} \def\end{proposition}{\end{proposition}} \def\begin{lemma}{\begin{lemma}} \def\end{lemma}{\end{lemma}} \def\begin{corollary}{\begin{corollary}} \def\end{corollary}{\end{corollary}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \newcommand{\proof}[1]{{\noindent \bf proof:}\par {#1} $\square$} \textwidth 155mm \textheight 235mm \topmargin 0pt \oddsidemargin 5mm \headheight 0pt \headsep 0pt \topskip 9mm \begin{document} \sloppy \pagestyle{empty} \hfill \addtolength{\baselineskip}{0.20\baselineskip} \begin{center} \vspace{26pt} \begin{center} {\large \bf {Non-homogenous disks in the chain of matrices}} \newline \end{center} \vspace{26pt} {\sl N.\ Orantin$^a$, A.\ Veliz-Osorio$^b$}\hspace*{0.05cm}\footnote{ E-mail: <EMAIL>, <EMAIL>}\\ \vspace{6pt} $ ^a$ CAMSD, Departamento de Matem\'{a}tica, Instituto Superior T\'{e}cnico\\ Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal\\ \bigskip $ ^b$ Departamento de F\'{i}sica, Instituto Superior T\'{e}cnico\\ Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal \end{center} \vspace{20pt} \begin{center} {\bf Abstract}: We investigate the generating functions of multi-colored discrete disks with non-homogenous boundary condition in the context of the Hermitian multi-matrix model where the matrices are coupled in an open chain. We show that the study of the spectral curve of the matrix model allows to solve a set loop equations to get a recursive formula computing mixed trace correlation functions to leading order in the large matrix limit. \end{center} \vspace{26pt} \pagestyle{plain} \setcounter{page}{1} \section{Introduction} The theory of random matrices has many ramifications in different fields of mathematics and physics. In the recent years, following the work of Eynard \cite{E1MM}, matrix models have seen significant advances especially through their applications to enumerative geometry. The enumeration of surfaces can indeed be addressed by the evaluation of the result of a saddle point approximation on integrals defined over a set of random matrices as the size of the matrices gets very large. To be more precise, the result of such an approximation can be seen as the generating function of a set of fat graphs composed of ribbons glued together along vertices \cite{BIPZ}. This set of fat graphs is in bijection with the set of discrete surfaces, or maps as denoted by combinatorists, i.e. surfaces composed of polygons glued by their edges. The matrix integrals can thus be seen, in this large matrix limit, as the generating function of a set of maps with a weight depending on the probability measure put on the set of matrices under study \cite{ambjornrmt,David,Kazakov}. Such objects are of prime interest not only to mathematicians but also to physicists for investigating the behavior of statistical systems on random surfaces (or lattices) as well as for understanding quantization of two dimensional gravity or topological string theories from a discrete point of view. Discrete surfaces also provide a very good toy model for developing some intuition to be applied later to string and gauge theories. From this perspective, a lot of progress have been made by Eynard et al. leading to powerful universal technics for the computation of generating functions of maps of different types. Among the results obtained, one finds: \begin{itemize} \item The enumeration of maps of arbitrary topology, i.e. arbitrary genus and number of boundaries, has been achieved in \cite{ACKM,ec1loopF,E1MM}; \item The enumeration of bicolored maps of arbitrary genus and arbitrary boundary conditions has been completed in the series of papers \cite{eynm2m,EO1,CEO,EKK,EOsymmetry,EOallmixed}; \item The enumeration of bicolored maps drawn on non-orientable surfaces has also been solved in \cite{ChEynbeta,EM,CEM1,CEM2,BEMPF} with homogenous boundary conditions; \item The enumeration of maps whose faces are colored with more than two colors has been partly solved in \cite{Echain,EPF}. In particular, the generating functions of maps with homogenous boundary conditions are known whatever their genus is. \end{itemize} One big outcome of these works is the universality of the solution found for all these problems. Indeed, for each of the aforementioned enumerative problems, the generating function of maps with homogenous boundary conditions are given by a unique universal inductive formula on the Euler characteristic of the surfaces enumerated. Even better is the fact that this inductive formula, sometimes called topological recursion, goes beyond the field of random matrices and enumeration of maps. The topological recursion \cite{EOinvariants,EOrev} seems to be a universal solution to different problems of enumerative geometry consisting in the computation of the volume of the moduli space of Riemann surfaces with respect to different measures opening the way for new insights in mathematics and topological string theories even when this measure does not localize on discrete surfaces. The topological recursion was inspired by the computation of generating functions of open surfaces with homogenous boundary conditions in the different matrix models encountered. The next step in fully understanding the enumeration of maps in general is to understand the action of boundary operator, that is compute the generating function of surfaces with boundaries along which the boundary condition changes. This has been understood in the hermitian two matrix model case where it has been possible to compute the generating function of surfaces composed of polygons of two different colors with arbitrary boundary conditions \cite{EOallmixed}. The result points towards a generalization of the topological recursion for these new observables. Going further requires the study of the so-called chain of hermitian matrix model which generates maps colored with an arbitrary number of colors. Even if the homogenous boundary condition case is understood \cite{EPF}, there are only a few result concerning changes of boundary conditions \cite{Echain}. In these notes, we present a recursive method, similar to the topological recursion, which allows to compute the generating function of discs with mixed boundary conditions. In theorem \ref{threc} and corollary \ref{correc} we prove that the corresponding correlation function $\left<{\,\rm Tr}\: \left[ {1 \over x_1-M_1} {1 \over x_2-M_2} \dots {1 \over x_{\cal N}-M_{\cal N}} \right]\right>$ can be computed to leading order in the large matrix limit in terms of the spectral curve of the model. The formula takes a very nice form which seems to follow from the all possible degeneracies of the surfaces generated. From the two matrix model experience, it is probably the building block for all the other generating functions and probably one of the most important objects of the theory. We expect to be able to use the techniques developed in this paper to compute any observable of this model. The computation of these disk amplitudes has thus many interesting features. First of all, it gives an efficient method for computing the generating functions of discs composed of colored polygons with mixed-boundary conditions. In addition, we are convinced that the result as well as the methods used in its derivation are generalizable to the computation of the generating function of surfaces of arbitrary topology. From this perspective, these notes should provide the building block for the computation of all observable of the chain of matrices model. This result is also important for better understanding conformal field theories. Indeed, when going to a particular limit in the parameter space of matrix models, one reaches critical regimes described by minimal models, a particular type of rational conformal field theories. The knowledge of boundary operators in the matrix model setup thus also gives access to boundary operator in such rational conformal field theories. Finally, the form of the solution points towards a generalization outside of the matrix model's setup generalizing the topological recursion formalism leading to the knowledge of new open amplitudes in topological string theories or may be new relative Gromov-Witten invariants. \bigskip This paper is organized as follows: \begin{itemize} \item In section 2, we introduce the model as well as the notations required in the following and present the result of these notes; \item In section 3, we remind how algebraic geometry appears in this context through the so-called loop equations and introduce all the notations and basic knowledge further needed. \item In section 4, we derive and solve a closed set of loop equation satisfied by the leading order mixed correlations functions. \item In section 5, we specialize this result to the chain of 3 matrices and give explicit examples. \item Section 6 is devoted to the conclusion and a short discussion on the perspectives. \end{itemize} \section{The open chain of matrices} \subsection{The model} We want to generate discrete surfaces, i.e. surfaces composed of polygons glued by their edges. In addition, we want to include some additional information on these random surfaces such as a spin structure or some matter. For this purpose, each polygon carries one color labeled by an index ranging from 1 to ${\cal N}$. Generating functions for such discrete colored surfaces can be obtained as correlation functions of a multi-matrix model. The latter studies the distribution of matrix elements with respect to a probability measure $d\mu$ on the set $\left(H_N\right)^{\cal N}$ of ${\cal N}$-uple of hermitian matrices of respective size $N \times N$ defined as \begin{equation} d\mu\left(M_1,\dots,M_{\cal N}\right) := dM_1 dM_2 \dots dM_{{\cal N}} \exp\left(- {N \over T} {\,\rm Tr}\:\left[\sum_{k=1}^{{\cal N}} V_k(M_k) - \sum_{k=2}^{{\cal N}} c_{k-1,k} M_{k-1} M_k\right] \right) \end{equation} where $M_i$ is a hermitian matrix of size $N$, $dM_i$ is the product of Lebesgue measures of the real components of $M_i$ and \begin{equation} V_k(x) = \sum_{i=2}^{d_k+1} {g_i^{(k)} \over i} x^{i} \end{equation} are polynomial potentials of respective degrees $d_k+1$. The partition function of this model is defined as \begin{equation}\label{defpartition} {\cal Z} := \int_{\left(H_N\right)^{\cal N}} d\mu\left(M_1,\dots,M_{\cal N}\right). \end{equation} In this paper, we interpret this random matrix integral as a generating function of discrete surfaces. This interpretation is exact if one considers a formal version of this integral, which is defined as the result of a saddle point approximation of this one\footnote{This formal integral is defined as a formal series in $T$ and does not need to converge. More precisions about this definition can be found in \cite{Emap}}. In the following, all the matrix integrals encountered are formal in this sense. In a nutshell, the combinatorial interpretation of such formal matrix integral is obtained by expanding the non quadratic part of the action \begin{equation} {\cal S} := \sum_{k=1}^{{\cal N}} V_k(M_k) - \sum_{k=2}^{{\cal N}} M_{k-1} M_k \end{equation} around the extremum $\left\{M_i = 0\right\}_{i=1}^{\cal N}$ and represent the result by Feynman graphs whose edges are ribbons which cannot be twisted. In the expansion of the action around this saddle, each factor ${\,\rm Tr}\: M_i^k$ is represented by a $k$-valent vertex of color $i$ whose legs are ribbons. This peculiarity of the edges puts an ordering on the legs around every vertex. One can thus replace every vertex by a face (or polygon) whose number of edges equals the valence of the initial vertex. Then, each graph is dual to a map, or discrete surface. Hence, the partition function of the matrix model is defined as the generating function of closed surfaces made of polygons colored with ${\cal N}$ different colors and glued together following the prescriptions: \begin{itemize} \item the polygons of color $k$ have at most $d_k+1$ edges ; \item two polygons of respective colors $i$ and $j$ can be glued by their edges only. \end{itemize} In this representation, the partition function reads \begin{equation} {\cal Z} = \sum_{v=0}^\infty T^v \sum_{g=0}^\infty N^{2-2g} \sum_{S \in \mathbb{M}_g(v)} {1 \over \# \hbox{Aut}(S)} \prod_{i,k} \left(g_i^{(k)}\right)^{n_{i,k}(S)} \prod_{i,j} \left(\left[C^{-1}\right]_{i,j}\right)^{\widetilde{n}_{i,j}(S)} \end{equation} where $\mathbb{M}_g(v)$ is the set of orientable genus $g$ maps with $v$ vertices, $\hbox{Aut}(S)$ is the number of automorphism of $S$, $n_{i,k}(S)$ the number of faces of color $k$ with $i$ edges in $S$, $\widetilde{n}_{i,j}(S)$ the number of edges between a face of color $i$ and a face of color $j$ in $S$ and $\left[C^{-1}\right]$ the inverse of the matrix of the bilinear terms of the action \begin{equation} C :=\left( \begin{array}{cccccc} g_2^{(1)}&-c_{1,2}&&&&0\cr -c_{1,2}&g_2^{(2)}&-c_{2,3}&&&\cr &-c_{2,3}&g_2^{(3)}&-c_{3,4}&&\cr &&\ddots&\ddots&\ddots&\cr &&&-c_{{\cal N}-2,{\cal N}-1}&g_2^{({\cal N}-1)}&-c_{{\cal N}-1,{\cal N}}\cr 0&&&&-c_{{\cal N}-1,{\cal N}}&g_2^{({\cal N})}\cr \end{array} \right) . \end{equation} The free energy \begin{equation} {\cal F}:= \ln {\cal Z} \end{equation} is the generating function of the connected surfaces contributing to ${\cal Z}$. For any function $f: \left(H_N\right)^{\cal N} \to \mathbb{C}$, one defines a corresponding correlation function by \begin{equation} \moy{f\left(M_1, \dots,M_{\cal N}\right)}:= {1 \over {\cal Z}} \int_{\left(H_N\right)^{\cal N}} f\left(M_1, \dots,M_{\cal N}\right) d\mu\left(M_1,\dots,M_{\cal N}\right). \end{equation} The correlation functions of this model can also be interpreted as the generating functions of open surfaces built with the same prescription and the same measure as the closed ones, i.e. surfaces with boundaries and prescribed boundary conditions on it. Let us briefly recall how to define generating functions of such surfaces. Consider a correlation function of the form: \begin{equation} \left< {\,\rm Tr}\: \left[\left(M_{i_1}\right)^{j_1} \left(M_{i_2}\right)^{j_2} \dots \left(M_{i_k}\right)^{j_k} \right]\right> := {1 \over {\cal Z}} \int_{\left(H_N\right)^{\cal N}} {\,\rm Tr}\: \left[\left(M_{i_1}\right)^{j_1} \left(M_{i_2}\right)^{j_2} \dots \left(M_{i_k}\right)^{j_k} \right] d\mu\left(M_1,\dots,M_{\cal N}\right) \end{equation} where $(i_1, i_2, \dots, i_k) \in [1,{\cal N}]^k$ and $(j_1, \dots, j_k)\in \left(\mathbb{N}^*\right)^k$ are arbitrary positive integers. The factor ${\,\rm Tr}\: \left[\left(M_{i_1}\right)^{j_1} \left(M_{i_2}\right)^{j_2} \dots \left(M_{i_k}\right)^{j_k} \right]$ in the integrand constrains all the generated graphs to contain one vertex with ${\displaystyle \sum_{l = 1}^k i_l}$ colored legs organized so that the sequence of colors coincide with the sequence of matrices in $\left(M_{i_1}\right)^{j_1} \left(M_{i_2}\right)^{j_2} \dots \left(M_{i_k}\right)^{j_k}$. Thus, this correlation function is the generating function of all possible fat graphs that can be glued to this vertex, i.e. all fat graphs with one boundary constrained by the condition that it must be glued to this vertex. In the dual representation where k-valent vertices are replaced by k-gones of the same color, this function generates all surfaces with one boundary constrained by the same type of gluing condition. \begin{figure} \begin{center} \includegraphics[width=5cm,angle=00]{discretefig}\\ \end{center} \caption{ Example of map contributing to $\left<{\,\rm Tr}\: M_1^4 M_2^3 M_3^2\right>$ where the type 1 condition (resp. type 2 and type 3) is depicted in yellow (resp. red and blue). It is important recalling that the boundary condition corresponds to the color of the outer face which has been removed and not the color of the inner polygons touching the boundary. The dual fat graph is depicted in dotted lines. }\label{discretefig} \end{figure} In order to deal with all the possible exponents $j_l$ at once one introduces the resolvents \begin{equation}\begin{array}{rcl} W_{i_1,i_2, \dots, i_k}(x_1,x_2, \dots, x_k)&:=& {1 \over N} {\displaystyle \sum_{j_1, \dots, j_k=1}^\infty} {\left< {\,\rm Tr}\: \left[\left(M_{i_1}\right)^{j_1} \left(M_{i_2}\right)^{j_2} \dots \left(M_{i_k}\right)^{j_k} \right] \right> \over x_1^{j_1+1} x_2^{j_2+1} \dots x_k^{j_k+1} }\cr &:=& {1\over N} \left< {\,\rm Tr}\: {1 \over x_1 - M_{i_1}} {1 \over x_2 - M_{i_2}} \dots {1 \over x_k - M_{i_k}} \right>\cr \end{array} \end{equation} as formal power series in its parameters $x_1,x_2, \dots, x_k$. It is very important to note that these resolvents are asymptotic series {\em defined only when there arguments are large!}. Indeed, we will show that these correlation functions are multivalued functions of their variables $x$ in the complex plane\footnote{This property can be seen as the result of the finiteness of the radius of convergency of this series.} but can be promoted to monovalued functions on an Riemann surface. However, one must always keep in mind that the physical quantities are the coefficients of the expansion of these correlation functions around a very particular point. Using the observation of 't Hooft for QCD \cite{thooft}, one can also select the genus of the generated surfaces. In order to do so, one can show that these correlation functions are series in ${1\over N^2}$ and that the order of ${1\over N^2}$ is genus of the generated surface, i.e. one can write: \begin{equation} W_{i_1,i_2, \dots, i_k}(z_1,z_2, \dots, z_k) = \sum_{g=0}^{\infty} {1 \over N^{2g}} W_{i_1,i_2, \dots, i_k}^{(g)}(z_1,z_2, \dots, z_k) \end{equation} where $W_{i_1,i_2, \dots, i_k}^{(g)}(z_1,z_2, \dots, z_k)$ is the generating function of genus $g$ surfaces with one boundary with the boundary condition induced by $W_{i_1,i_2, \dots, i_k}(z_1,z_2, \dots, z_k)$. In this paper, we will be interested in the large $N$ limit of these correlation functions, that is to say, the generating functions of discs with boundary operators of the form $W_{i_1,i_2, \dots, i_k}^{(0)}(z_1,z_2, \dots, z_k)$. \subsection{Main result: disk amplitudes} The main result of this paper is the computation of the mixed disk amplitudes with the use of a recursion relation explained in th.\ref{threc}: \begin{equation} {\cal W}_{1,j,j+1,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N}) = \mathop{\,\rm Res\,}_{q \to p_1,p_j^{+i,j}} {\cal K}_j(p_1,q_j,p_j) {\cal W}_{1,j+1,\dots, {\cal N}}(q,p_{j+1},\dots, p_{\cal N}) \end{equation} where the ingredients are explained in the following of the paper. In particular, this formula makes a heavy use of the different sheeted structures of the spectral curve of the model under study. Another important feature of these notes are the explicit computation of these amplitudes as well as their expansion in the 3-matrix model. In section \ref{sec3matrix}, we apply the spectral curve approach as well as th.\ref{threc} to explicitly compute the number of tri-colored discs with given mixed boundary conditions. \subsection{More notations} In this section, we present the notations used in the following of these notes. We used the notations of \cite{EPF} as long as it was appropriate so that the reader interested in both aspects of this matrix model can go from one paper to the other without being confused. First of all, let us remind the correlation functions to be computed: \begin{equation} W_{i_1,i_2, \dots, i_k}(x_1,x_2, \dots, x_k):= {T \over N} \left< {\,\rm Tr}\: {1 \over x_1 - M_{i_1}} {1 \over x_2 - M_{i_2}} \dots {1 \over x_k - M_{i_k}} \right>. \end{equation} We also define the generating function of surfaces with two boundaries by \begin{equation} W_{\vec{i};\vec{j}}(x_1,x_2, \dots, x_k;y_1,\dots,y_l):= \left< {\,\rm Tr}\: \left({1 \over x_1 - M_{i_1}} \dots {1 \over x_k - M_{i_k}}\right) {\,\rm Tr}\: \left({1 \over y_1 - M_{j_1}}\dots {1 \over y_l - M_{j_l}}\right) \right>_c \end{equation} where the index $c$ stands for the connected part and \begin{equation} \vec{i} = (i_1,i_2,\dots,i_k) {\qquad , \qquad} \vec{j} = (j_1,j_2,\dots,j_l) \end{equation} are sequences of integers in $[1,{\cal N}]$. Following \cite{Echain}, let us now define a set of polynomials $f_{i,j}(x_i, \dots, x_j)$ by induction: \begin{equation} \left\{\begin{array}{l} f_{i,j} := 0 \qquad \hbox{if} \qquad j< i-1 \cr f_{i,i-1}:=1 \cr f_{i,i}(x_i) := {V_i'(x_i) \over c_{i,i+1}} \cr c_{j,j+1} f_{i,j}(x_i,\dots,x_{j}):= V_{j}'(x_{j}) f_{i,j-1}(x_i,\dots,x_{j-1}) - c_{j-1,j} \, x_{j-1} \, x_{j} \, f_{i,j-2}(x_i, \dots, x_{j-2}) \cr \end{array} \right. . \end{equation} For $1\leq k \leq l \leq {\cal N}$, let us then define \begin{equation} P_{k,l;j_1, \dots, j_m}(x_1,\dots,x_{\cal N};y_1,\dots,y_m):= \mathop{\mathrm{Pol}}_{x_k,\dots, x_l} f_{k,l}(x_k,\dots,x_l) W_{1,\dots,{\cal N};j_1,j_2, \dots, j_m}(x_1,\dots,x_{\cal N};y_1,\dots,y_m) \end{equation} which is a polynomial in the variables $x_i$, $i = k,\dots,l$. Given a pair of complex numbers $(x_1,x_2)$, we define a set of polynomial functions $\left\{\hat{x}_i(x_1,x_2)\right\}_{i=3}^{\cal N}$ by the recursion relation \begin{equation}\label{constrhat} c_{i,i+1}\hat{x}_{i+1}:= V_i'(\hat{x}_i) - c_{i-1} \hat{x}_{i-1} \end{equation} for all $i=2, \dots,{\cal N}$. The case $i=2$ is slightly different and we define \begin{equation} Y(x_1) := V_1'(x_1)-W_1(x_1). \end{equation} It should be noted that the latter function has a $1/N^2$ expansion. Finally, it is convenient to denote any observable whose variables satisfy the constraints \eq{constrhat} by: \begin{equation} \widehat{{\cal{O}}}(x_1,x_2):= {\cal{O}}(x_1,x_2,\hat{x}_3(x_1,x_2),\hat{x}_4(x_1,x_2),\dots,\hat{x}_{\cal N}(x_1,x_2)). \end{equation} All the observables of the model admit a topological expansion \cite{thooft}, i.e. they can be written as a formal power series in ${1 \over N^2}$: \begin{equation} {\cal{O}}= \sum_{g=0}^{\infty} N^{-2g} {\cal{O}}^{(g)}. \end{equation} \section{Master loop equation and spectral curve} Before going into the explicit computation of correlation functions, let us recall the algebraic structure underlying the hermitian multi-matrix model studied by Eynard in \cite{Echain}. We will particularly focus on the sheeted structure of the spectral curve since it is the most important ingredient needed in the following. The interested reader can find a longer and more complete presentation of this structure in \cite{Echain} or more recently in \cite{EPF}. \subsection{Loop equations} The partition function and the correlation functions, seen as generating functions of random surfaces, are constrained by an infinite set of equations prescribing how the weight of a surface is changed when one remove one edge from it. These equations where first obtained by Tutte \cite{tutte,tutte2} for counting triangulated surfaces and later introduced under the name of loop equation in the context of matrix models by Migdal \cite{Migdal}. These equations proved to be some very powerful tools since they allowed the computation of many observables in different types of hermitian matrix models through the arising of algebraic geometry. Practically, these equations can be obtained by writing the invariance of the integral in \eq{defpartition} under particular infinitesimal change of variables: \begin{equation} \begin{array}{l} Z= \int dM_1 \dots dM_j \dots dM_{{\cal N}} e^{- {N \over T} {\,\rm Tr}\:\left[{\displaystyle \sum_{k=1}^{{\cal N}} V_k(M_k) - \sum_{k=2}^{{\cal N}} c_{k-1,k} M_{k-1} M_k}\right] } =\cr \;\;\;\; = \int dM_1 \dots d\tilde{M}_j \dots dM_{{\cal N}} \, J(M_j,\tilde{M}_j) \, \times \cr \qquad \quad e^{- {N \over T} {\,\rm Tr}\:\left[{\displaystyle \sum_{k\neq j} V_k(M_k) + V_j(\tilde{M}_j)- \sum_{k\neq j, j+1} c_{k-1,k} M_{k-1} M_k - c_{j-1,j} M_{j-1} \tilde{M}_j - c_{j,j+1} \tilde{M_j} M_{j+1} }\right] } \cr \end{array} \end{equation} where $\tilde{M}_j := M_j + \epsilon \, \delta M_j$ is an infinitesimal deformation of $M_j$. The loop equation just corresponds to saying that the variation of the action is compensated by the Jacobian of this change of variable to first order in $\epsilon$. The set of loop equations corresponding to the changes of variables \begin{equation} M_k \to M_k + \epsilon \, {1\over x_{k+1}-M_{k+1}} {1\over x_{k+2}-M_{k+2}} \dots {1\over x_{{\cal N}}-M_{{\cal N}}} {1\over x_{1}-M_{1}} \end{equation} supplemented with some algebra allows to show that \begin{theorem} For $0<k<{\cal N}$, the correlation functions satisfy \begin{eqnarray}\label{masterloop1} \left(x_k - \hat{x}_k(x_1)\right) W_{1,k,\dots,{\cal N}}(x_1,x_k,\dots,x_{\cal N}) &=& W_{1,k+1,\dots,{\cal N}}(x_1,x_{k+1},\dots,x_{\cal N}) \cr && - P_{1,k-1}(x_1,\hat{x}_2,\dots,\hat{x}_{k-1},x_k,\dots,x_{\cal N})\cr && + {T \over N^2} P_{2,k-1;1}(x_1,\hat{x}_2,\dots,\hat{x}_{k-1},x_k,\dots,x_{\cal N})\cr \end{eqnarray} where \begin{equation} P_{i,k}(x_1,\dots,x_{\cal N}):= \mathop{\mathrm{Pol}}_{x_i,\dots,x_{k}}f_{i,k}(x_1,\dots,x_{k-1}) W_{1,\dots,{\cal N}}(x_1,\dots,x_{\cal N}) \end{equation} is a polynomial in the variables $x_j$, $j = i, \dots, k$ and the polynomials $f_{i,k}(x_i,\dots,x_k)$ are defined by induction through \begin{equation} f_{i,k}(x_i,\dots,x_{k-1}) = {V_k'(x_k) f_{i,k-1} - c_{k-1,k} x_{k-1} x_k f_{i,k-2}\over c_{k,k+1}}. \end{equation} For $k = {\cal N}$, one has \begin{equation}\label{masterloopN} \left(V_{\cal N}'(\hat{x}_{\cal N}) - \hat{x}_{{\cal N}-1}\right) W_{1}(x_1) = P_{1,{\cal N}}(x_1,\hat{x}_2,\dots,\hat{x}_{\cal N}) + {T^2 \over N^2} P_{2,{\cal N};1}(x_1,\hat{x}_2,\dots,\hat{x}_{\cal N};x_1) . \end{equation} \end{theorem} This theorem was proved in \cite{Echain,EPF}. For completeness, it is derived for the present model with the notations in hand in appendix A. \subsection{Leading order and master loop equation} By plugging $Y(x) = V_1'(x) - T W_1(x)$ into \eq{masterloop1}, the leading order in the $1/N^2$ expansion of this equation reads \begin{equation}\label{masterloop} \hat{E}(x_1,y(x_1)) = 0. \end{equation} where \begin{equation} y(x) = Y^{(0)}(x) \end{equation} is the leading order of the ${1 \over N^2}$ expansion of $Y(x)$ and \begin{equation} E(x_1,x_2, \dots, x_{\cal N}) = (V_1'(x_1) - x_2)(V_{\cal N}(x_{\cal N}) - x_{{\cal N}-1}) - P_{1,{\cal N}}^{(0)}(x_1, \dots, x_{\cal N}) \end{equation} is a polynomial in all its variables. Consider \eq{masterloop} as an equation with unknown $x_1$ and $y(x_1)$. On can see that this is a polynomial equation of degree ${\displaystyle \prod_{i = 1}^{{\cal N}}} d_i$ in $x_1$ and $1 + {\displaystyle \prod_{i = 2}^{{\cal N}}} d_i$ in $y(x_1)$. Thus, it defines an algebraic curve ${\cal E}(x,y)=0$ in $C^2$. \subsection{Description of the spectral curve} Let us now consider the algebraic equation \begin{equation} \hat{E}(x,y) = 0. \end{equation} It defines an algebraic curve in $\mathbb{C}^2$ which encodes most of the combinatorial properties of the matrix model: it is the spectral curve associated to the model. In this section, we study the properties the algebraic curve $\hat{E}(x,y) = 0$ as a compact Riemann surface ${\cal{L}}$ equipped with two meromorphic functions $x$ and $y$ such that \begin{equation} \forall \, p \in {\cal L} \, , \; \, \hat{E}(x(p),y(p)) = 0. \end{equation} A point $p\in {\cal L}$ is equivalent to a pair $(x(p),y(p))\in \mathbb{C}^2$ satisfying $\hat{E}(x(p),y(p)) = 0$. Since, given a pair $(x,y)\in \mathbb{C}^2$, all the $\hat{x}_i(x,y)$ are unambiguously fixed, one can see them as monovalued meromorphic functions $z_i$ on ${\cal{L}}$ such that: \begin{equation} \forall p \in {\cal{L}} \, , \; \; z_i(p) = \hat{x}_i(x(p),y(p)). \end{equation} By definition, one has \begin{equation} x(p) = z_1(p) {\qquad , \qquad} y(p) = z_2(p). \end{equation} Using these functions, one can write: \begin{equation} \forall p \in {\cal{L}} \; , \; \; E(z_1(p), z_2(p), z_3(p), \dots , z_{\cal N}(p)) = 0 \end{equation} and one can choose to fix any of the functions $z_i(p)$ as parameter to study the curve. Depending on which parameter we chose we obtain different descriptions that we can now characterize. \subsubsection{Sheeted structure and points at infinity} Let us first study this curve in terms of the ``physical" variable $x:=z_1$. Since $\hat{E}(x,y)$ is a polynomial in $y$ of degree $1 + {\displaystyle \prod_{i = 2}^{{\cal N}}} d_i$, $y(x)$ is a $1 + {\displaystyle \prod_{i = 2}^{{\cal N}}} d_i$-valued functions of $x$. This means that, for a generic value of $x$, there exist $1 + {\displaystyle \prod_{i = 2}^{{\cal N}}} d_i$ points $p^j$ on ${\cal{L}}$ such that: \begin{equation} \forall i = 0, \dots, {\displaystyle \prod_{i = 2}^{{\cal N}}} d_i \; , \;\; z_1(p^i) = z_1(p). \end{equation} Hence ${\cal{L}}$ can be seen as $1+s_1$ copies of the Riemann sphere, called $z_1$-sheets, glued together by cuts in such a way that $x(p) = z_1(p)$ is injective in each sheet. This sheeted structure reflects the multi-valuedness of the resolvent $W_1^{(0)}(x)$ as a function of $x$. Indeed, once the value of this correlation function is fixed, $y(p) = z_2(p)$ is fixed and all the $z_i(p) = \hat{x}_i(x(p),y(p))$'s are determined. Thus, this fixes one unique point in ${\cal{L}}$. How can we extract the right value of $y(x)$ giving access to the generating functions of random surfaces? In order to distinguish the different $x$-sheets, one can look at the pre-images of $\infty$ by $x(p)$, i.e. the points $\infty^j \in {\cal{L}}$ such that $x(\infty^j)= \infty$. From its definition as an asymptotic series, one knows that the physical solution of the algebraic equation should satisfy\footnote{We call this solution physical because it is the branch whose expansion in $x$ gives rise to the generating function of surfaces.}: \begin{equation} y(p) \sim_{x(p) \to \infty} V'(x(p)) + {1 \over x(p)} + O(x^{-2}(p)). \end{equation} It means that there exists at least one pre-image of infinity, denoted $\infty^1 \in {\cal L}$, such that \begin{equation} z_2(p) \sim_{p \to \infty^1} V'_1(z_1(p)) + {1 \over z_1(p)} + O(z_1^{-2}(p)) \end{equation} and more generally \begin{equation} \forall k = 2, \dots , {\cal N} \; , \; \; z_k(p) \sim_{p \to \infty^1} O(z_1^{s_k}) \end{equation} where \begin{equation} s_k = {\displaystyle \prod{i=1}^{k-1}} d_i \end{equation} One could symmetrically choose $z_{\cal N}$ as a variable and use the physical condition that \begin{equation} z_{{\cal N}-1}(p) \sim_{z_{\cal N}(p) \to \infty} V'_{\cal N}(z_{{\cal N}}(p)) + {1 \over z_{\cal N}(p)} + O(z_{\cal N}^{-2}(p)). \end{equation} This implies that there exist at least one pre-image $\infty^{\cal N}$ of infinity by $x$ such that: \begin{equation} z_1(p) \sim_{p \to \infty^{{\cal N}}} O(z_{\cal N}(p)^{r_1}) \end{equation} where \begin{equation} r_k = {\displaystyle \prod{i=k+1}^{{\cal N}}} d_i. \end{equation} This means that $r_1$ different $z_1$-sheets merge at $\infty^{\cal N}$. Since $r_1+1$ is exactly the number of $z_1$-sheets, there exist no other pre-image of infinity by $x = z_1$ and the fiber above infinity reads: $x^{-1}(\infty) = \left\{\infty^1,\infty^{\cal N}\right\}$. One can summarize the structure of $\hat{E}$ in terms of $z_1$ as follows\footnote{An example is depicted in fig.\ref{sheetedfig}.}: \begin{itemize} \item ${\cal{L}}$ is composed of $1+r_1$ $z_1$-sheets; \item $z_1(p)$ has two poles in ${\cal{L}}$: one simple pole noted $\infty^1$ and one pole of degree $r_1$ noted $\infty^{\cal N}$; \item $\infty^1$ belongs to only one sheet called the physical sheet since it corresponds to physical solutions. All other sheets merge at $\infty^{\cal N}$. \end{itemize} \begin{figure} \begin{center} \includegraphics[width=5cm,angle=00]{sheetedspcurve}\\ \end{center} \caption{ Example of sheeted structure of ${\cal L}$ where one has represented a genus 0 spectral curve as the Riemann sphere, the red dot represents the pole $\infty^1$ while the black dot is the other pole $\infty^{\cal N}$. The different sheets are represented by the connected components separated by continuous lines. In this case, $r_1 = 4$ of them join at $\infty^{\cal N}$ and only one contains $\infty^1$.}\label{sheetedfig} \end{figure} The unique $z_1$-sheet containing $\infty^1$ is only referred to as the physical sheet. Indeed, in order to recover the physically meaningful quantities ${\,\rm Tr}\: M_1^k$, one has to expand $y(z)$ around $z \to \infty^1$: \begin{equation} y(z) \sim_{z \to \infty^1} V_1'(x(z)) + \sum_{k = 0}^\infty {1 \over x(z)^{k+1}} \, \lim_{N \to \infty}{1 \over N} \left<{\,\rm Tr}\: M_1^k\right>. \end{equation} The expansion of $y(z)$ around $\infty^{\cal N}$ doesn't have such a simple combinatorial interpretation. One can also fix $z_k(p)$ and look at the $z_k$-sheeted structure of ${\cal{L}}$. With the same arguments, one can see that: \begin{itemize} \item There exists $r_k+s_k$ $z_k$-sheets where $z_k(p)$ is injective; \item $z_k(p)$ has two poles. One pole of degree $s_k$ in $\infty^1$ where $s_k$ $z_k$-sheets meet and one pole of degree $r_k$ in $\infty^{\cal N}$ where the other $r_k$ $z_k$-sheets merge. \item One notes $p^{+j,k}$ (resp. $p^{-j,k}$) for $j = 1, \dots , s_k$ (resp. $j = 1, \dots , r_k$) the pre-images of $z_k(p)$ belonging to the different sheets which merge at $\infty^1$ (resp. $\infty^{\cal N}$). \end{itemize} In particular, one can see that there exists a neighborhood of $\infty^1$ where $z_1$ is injective in the $z_k$-sheets. This means that: \begin{equation}\label{injectivite} z_k(p^{+i,k}) = z_k(q^{+j,k}) \Rightarrow x(p^{+i,k}) = x(q^{+j,k}) \end{equation} in some neighborhood of $\infty^1$. \section{Correlation functions} \subsection{State of the art} Let us first recall the results of \cite{Echain,EPF}. In \cite{EPF}, a very efficient inductive method has been found for computing the whole topological expansion of the free energy as well as the correlation functions involving only the first matrix of the chain. This means that the free energy as well as the correlation functions of the type $\left<{\displaystyle \prod_{i=1}^k} {\,\rm Tr}\: {1 \over x_k - M_1}\right>$ are known to any order in the large $N$ expansion. In \cite{Echain}, some other observable were computed. First of all, the leading order in the large $N$ expansion of any two point correlation function of the type $\left< {\,\rm Tr}\:{1 \over x_1-M_i} {\,\rm Tr}\: {1 \over x_2-M_j}\right>_c$ was computed and expressed in terms of a fundamental one form on ${\cal L} \times {\cal L}$ often called the Bergman kernel. More important to us is the derivation of the simplest disk amplitude with mixed boundary conditions, that is the large $N$ limit of the correlation function $\left<{\,\rm Tr}\: {1 \over x_1-M_1} {1 \over x_2 - M_{\cal N}}\right>$. \subsection{Mixed correlation functions} With this description of the algebraic curve in hand, we are ready to compute the correlation functions. Let us first promote them to mono-valued differentials ${\cal W}_{i_1, \dots i_k}(p_1, \dots, p_k)$ on ${\cal L}^k$ defined by: \begin{equation} {\cal W}_{i_1, \dots i_k}(p_1, \dots, p_k) = W_{i_1, \dots i_k}(z_{i_1}(p_1), \dots, z_{i_k}(p_k)) \, dz_{i_1}(p_1), \dots, dz_{i_k}(p_k). \end{equation} We use the same notation for any other function considered in the previous section: the curly letters represent differentials on the spectral curve. From now on, we will consider only the leading order of the correlation functions in the $1/N^2$ expansion. We thus abusively denote \begin{equation} W_{i_1,i_2, \dots, i_k}(x_1,x_2, \dots, x_k):=W_{i_1,i_2, \dots, i_k}^{(0)}(x_1,x_2, \dots, x_k) \end{equation} and \begin{equation} P_{k,l}(x_1, \dots , x_{\cal N}):=P_{k,l}^{(0)}(x_1, \dots , x_{\cal N}) \end{equation} the leading orders of the observables we are studying. Let us also remind that \begin{equation} y(x_1):=Y^{(0)}(x_1) = V_1'(x_1)-W_1^{(0)}(x_1). \end{equation} In terms of differentials on the spectral curve, to leading order in the $1/N^2$ expansion, the loop equation \eq{masterloop1} can then be written: \begin{eqnarray}\label{loop2} (z_j(p_j)-z_j(p_1)) {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N}) &=& {\cal W}_{1,j+1, \dots , {\cal N}}(p_1,p_{j+1}, \dots, p_{{\cal N}}) dz_j(p_j)\cr && - {\cal P}_{1,j-1}(p_1,z_2(p_1), \dots , z_{j-1}(p_1), p_j, \dots, p_{\cal N})\cr \end{eqnarray} where \begin{equation} {{\cal P}_{1,j-1}(p_1,z_2(p_1), \dots , z_{j-1}(p_1), p_j, \dots, p_{\cal N}) \over {\displaystyle \prod_{i=j}^{{\cal N}}} dz_i(p_i)} = P_{1,j-1}(z_1(p_1), \dots , z_{j-1}(p_1), z_j(p_j), \dots, z_{\cal N}(p_{\cal N})) \end{equation} is a polynomial in $z_1(p_1)$ of degree $r_{j}-1$. The main result of these notes is that one can solve this equation by induction on $j$. Indeed, \eq{loop2} expresses ${\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ in terms of ${\cal W}_{1,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ provided that one knows the polynomial ${\cal P}_{1,j-1}(p_1,z_2(p_1), \dots , z_{j-1}(p_1), p_j, \dots, p_{\cal N})$. The polynomiality of the later actually allows one to compute it by using only \eq{loop2} and get \begin{theorem}\label{threc} {\bf Recursion relation for the disk amplitudes.} The mixed disk amplitudes satisfy the recursion relation \begin{equation} {\cal W}_{1,j,j+1,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N})= \mathop{\,\rm Res\,}_{q \to p_1,p_j^{+i,j}} {{\cal W}_{1,j+1,\dots,{\cal N}}(q,p_{j+1},\dots,p_{\cal N}) {\displaystyle \prod_{k=1}^{s_{j}}}\left[z_{1}(p_1)-z_{1}(p_j^{+k,j})\right]\over \left[z_{1}(q)-z_{1}(p_1)\right] \left[z_{j}(p_j)-z_j(p_1)\right] {\displaystyle \prod_{k=1}^{s_{j}}} \left[z_{1}(q)-z_{1}(p_j^{+k,j})\right]}. \end{equation} \end{theorem} Since the initial value \begin{equation} {\cal W}_{1,{\cal N}}(p_1,p_{\cal N}) = \sum_{k=1}^{\cal N}\frac{E\left(Z_1,...,Z_k,\tilde{Z}_{k+1},...,\tilde{Z}_{\mathcal{N}}\right)}{\left(Z_k-\tilde{Z}_k\right)\left(Z_{k+1}-\tilde{Z}_{k+1}\right)} \end{equation} \begin{equation} Z_k=z_k\left(p_1^{+,1}(z_1)\right)\:\: \tilde{Z}_k=z_k\left(p_{\cal N}^{-,{\cal N}}(z_{\cal N})\right) \end{equation} was computed in \cite{Echain}, this theorem allows to compute ${\cal W}_{1,j,j+1,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N})$ for any $j=2,\dots,{\cal N}-1$ and in particular the complete mixed correlation function: \begin{corollary} \label{correc} The disk amplitudes read, for $j=2,\dots, {\cal N}-1$: \begin{equation} {\cal W}_{1,j,j+1,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N}) = \mathop{\,\rm Res\,}_{q_j} \mathop{\,\rm Res\,}_{q_{j+1} } \dots \mathop{\,\rm Res\,}_{q_{{\cal N}-1} } \prod_{\alpha =j}^{{\cal N}-1} {\cal K}_\alpha(p_1,q_\alpha,p_\alpha) {\cal W}_{1,{\cal N}}(q_{{\cal N}-1},p_{\cal N}) \end{equation} where one defines the recursion kernel \begin{equation} {\cal K}_\alpha(p,q,r) = {{\displaystyle \prod_{k=1}^{s_{\alpha}}}\left[z_{1}(p)-z_{1}(r^{+k,\alpha})\right]\over \left[z_{1}(q)-z_{1}(p)\right] \left[z_{\alpha}(r)-z_\alpha(p)\right] {\displaystyle \prod_{k=1}^{s_{\alpha}}} \left[z_{1}(q)-z_{1}(r^{+k,\alpha})\right]} \end{equation} and the residue sign stands for \begin{equation} \forall i = j, \dots , {\cal N}-1 \, , \; \mathop{\,\rm Res\,}_{q_i} = \mathop{\,\rm Res\,}_{q_i \to p_1} + \sum_{k=1}^{s_i} \mathop{\,\rm Res\,}_{q_i \to p_{i}^{+k,i}}. \end{equation} \end{corollary} In turn, this gives access to any disk amplitude of the form ${\cal W}_{i_1,\dots,i_k}$ with $1\leq i_1<i_2<\dots<i_k\leq i_{\cal N}$ by considering the leading order of the expansion the complete disk amplitude as some $p_i \to \infty^1$. \bigskip \subsection{Proof of the recursion relation} For proving th.\ref{threc}, one first needs a simple technical lemma. \begin{lemma} $(z_j(p_j)-z_j(p_1)) {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ vanishes when $p_1 = p_j^{+i,j}$, for $i=1, \dots, s_k$. \end{lemma} \proof{ The proof follows from the combinatorial interpretation of $ {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ when $p_1$ lies in the physical sheet. Indeed, for $p_1 \to \infty^1$ (and not in any other patch of the spectral curve), the definition of ${\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ reads \begin{eqnarray} && N \, (z_j(p_j)-z_j(p_1)) \, {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N}) = \cr && \qquad \qquad = (z_j(p_j)-z_j(p_1)) \left< {\,\rm Tr}\: {1 \over z_1(p_1) -M_1} {1 \over z_j(p_j) -M_j} \dots {1 \over z_{\cal N}(p_{\cal N}) -M_{\cal N}} \right>. \end{eqnarray} One can thus write \begin{eqnarray} && N (z_j(p_j)-z_j(p_1)) {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N}) \cr &=& \left< {\,\rm Tr}\: {1 \over z_1(p_1) -M_1} {1 \over z_{j+1}(p_{j+1}) -M_{j+1}} \dots {1 \over z_{\cal N}(p_{\cal N}) -M_{\cal N}} \right> \cr && - \left< {\,\rm Tr}\: {1 \over z_1(p_1) -M_1} {z_j(p_1) - M_j \over z_j(p_j) -M_j} \dots {1 \over z_{\cal N}(p_{\cal N}) -M_{\cal N}} \right>\cr \end{eqnarray} which vanishes when $z_j(p_j) = z_j(p_1)$, i.e. $p_1$ belongs to the same $z_j$-fiber as $p_j$. Since this formula is valid only for $p_1$ in the $z_1$ physical sheet, this additional constraints implies that $(z_j(p_j)-z_j(p_1)) {\cal W}_{1,j,j+1, \dots , {\cal N}}(p_1, p_j, \dots , p_{\cal N})$ vanishes for $p_1 = p_j^{+i,j}$.} \bigskip {\bf Proof of theorem \ref{threc}} Thus, one knows that the left hand side of \eq{loop2}, vanishes for $p_1 = p_j^{+i,j}$. On the other hand, the second term of the right hand side, $P_{1,j-1}$, is a polynomial in $z_1(p_1)$ of degree $s_j-1$. Thanks to the vanishing of the left hand side, one knows its value in the $d_1 \dots d_{j-1}$ points, $z_1(p_j^{+i,j})$ for $j=1, \dots, s_j$ : \begin{equation} P_{1,j-1}(z_1(p_j^{+i,j}),z_2(p_j^{+i,j}),\dots ,z_{j-1}(p_j^{+i,j}),z_{j}(p_j)\dots,,z_{\cal N}(p_{{\cal N}})) =W_{1,j+1,\dots,{\cal N}}(p_j^{+i,j},p_{j+1},\dots,p_{\cal N}). \end{equation} By Lagrange interpolation, one gets an explicit expression for this polynomial which one can plug back into \eq{masterloop1} to get \begin{eqnarray} &&(z_{j}(p_j)-z_j(p_1)) {\cal W}_{1,j,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N})= {\cal W}_{1,j+1,\dots,{\cal N}}(p_1,p_{j+1},\dots,p_{\cal N}) \cr &&- \sum_{i=1}^{s_j} {{\cal W}_{1,j+1,\dots,{\cal N}}(p_j^{+i,j},p_{j+1},\dots,p_{\cal N}) {\displaystyle \prod_{k=1}^{s_j}}\left[z_{1}(p_1)-z_{1}(p_j^{+k,j})\right]\over \left[z_{1}(p_1)-z_{1}(p_j^{+i,j})\right] {\displaystyle \prod_{k\neq i}} \left[z_{1}(p_j^{+i,j})-z_{1}(p_j^{+k,j})\right]}. \end{eqnarray} This last equation can be written under the form of a residue formula \begin{equation} {\cal W}_{1,j,\dots,{\cal N}}(p_1,p_{j},\dots,p_{\cal N})= \mathop{\,\rm Res\,}_{q \to p_1,p_j^{+i,j}} {{\cal W}_{1,j+1,\dots,{\cal N}}(q,p_{j+1},\dots,p_{\cal N}) {\displaystyle \prod_{k=1}^{s_j}}\left[z_{1}(p_1)-z_{1}(p_j^{+k,j})\right]\over \left[z_{1}(q)-z_{1}(p_1)\right] \left[z_{j}(p_j)-z_j(p_1)\right] {\displaystyle \prod_{k=1}^{s_j}} \left[z_{1}(q)-z_{1}(p_j^{+k,j})\right]}. \end{equation} $\square$ \section{Example of application: the 3-matrix model} \label{sec3matrix} In this section, we perform explicit computations in the case ${\cal N} = 3$ resulting in the computation of three colored discs with mixed boundary conditions. \subsection{Genus 0 spectral curve} For enumerating maps, the spectral curve has to be a rational curve, i.e. its genus has to be vanishing. Hence, there exists a global coordinate $p$ such that the functions $z_i$ are rational functions \cite{Emap} \begin{equation} \forall i = 1,\dots,3 \, , \; z_i(p) = \sum_{k= -s_i}^{r_i} \alpha_{i,k}\: p^k \end{equation} satisfying \begin{equation}\label{condi} V'_1(z_1(p))-z_2(p) \sim_{p \to \infty} {T \over p\:\alpha_{1,1} } + O\left(p^{-2}\right) \end{equation} \begin{equation} V'_3(z_3(p))-z_2(p) \sim_{p \to 0} {Tp \over \alpha_{3,-1} } + O\left(p^{2}\right) \end{equation} \begin{equation} V'_2(z_2(p))=z_1(p)+z_3(p), \end{equation} where $\alpha_{1,1}=\alpha_{3,-1}=\gamma$. The limits of the sums describing the parametrizations are obtained from the degrees of the potentials, and are given by $s_1=r_3=1$, $r_1=d_2d_3$, $s_2=d_1$, $r_2=d_3$ and $s_3=d_1d_2$. The solution branch giving rise to the spectral curve is the one where the coefficients $\gamma$ and $\alpha_{i,k}$ are algebraic functions of $T$ such that $\gamma \to 0$ as $T \to 0$. With this global parameterization $\infty_1$ is located in $p = \infty$ whereas $\infty_3$ is $p = 0$. Once the spectral curve, i.e. the collection of rational functions $z_i(p)$, is known, one can study its different sheeted structures. The $z_i$-fiber over a point $Z_i$ is given by the set of solution $p^{\pm j,i}(Z_i)$ satisfying \begin{equation} z_i\left(p^{\pm j,i}(Z_i)\right) = Z_i. \end{equation} Among these solutions, $p^{+j,i}(Z)\to \infty$ and $p^{-j,i}(Z)\to 0$ as $Z \to \infty$. \begin{remark}\rm\small In the setup of enumerative geometry, the exponent of $T$ is the generating functions is the number of vertices in the maps considered. Thus, as long as one is interested in sufficiently small maps, the preceding equations need being solved only up to a given order in the $T$ expansion. One can thus compute the fibers above a point $Z$ through its small $T$ expansion. \end{remark} \subsection{The complete one loop function} \subsubsection{Loop equations and analytic solution} Since the 3-matrix model is the simplest model which carries the whole complexity of the chain of matrices, let us use it as a toy model for deriving more carefully the loop equations involved in the computation of the complete mixed correlation function. Let us consider the following change of variables \begin{equation} M_1\rightarrow M_1+\epsilon \delta M_1 \end{equation} where \begin{equation} \delta M_1=\frac{1}{z_2-M_2}\frac{1}{z_3-M_3}\frac{1}{z_1-M_1}. \end{equation} In order to keep the path-integral invariant under such a change, we demand that the contributions from the Jacobian of the change of variables and the contributions from the variation of the action cancel each other. To order $\epsilon$, we have \begin{equation} \moy{Tr\left(\frac{\partial \delta M_1}{\partial M_1}\right)}=-N\left(\moy{Tr(V'_1(M_1)\delta M_1)}-\moy{Tr(\delta M_1M_2)}\right) \end{equation} The left hand side can be written in terms of the connected correlators \begin{equation} \moy{Tr(\delta M_1)Tr\left(\frac{1}{z_1-M_1}\right)}=\moy{Tr(\delta M_1)}_c\moy{Tr\left(\frac{1}{z_1-M_1}\right)}_c+\moy{Tr(\delta M_1)Tr\left(\frac{1}{z_1-M_1}\right)}_c, \end{equation} therefore we can write the left hand side in terms of correlators dependent on the points $p,q,r\in\Sigma$ as \begin{equation} N^2{\cal W}_{1,2,3}(p,q,r){\cal W}_{1}(p)+{\cal W}_{1,2,3;1}(p,q,r;p) \end{equation} where \begin{equation} z_1(p) = z_1{\qquad , \qquad} z_2(q) = z_2 {\qquad , \qquad} z_3(r) = z_3. \end{equation} On the other hand the right hand side can be expressed as \begin{equation} -N\moy{Tr((V'_1(M_1)-V'_1(z_1))\delta M_1)}+N^2(V'(z_1(p))+z_2(q)){\cal W}_{1,2,3}(p,q,r)-N^2{\cal W}_{1,3}(p,r). \end{equation} By collecting the terms and considering only the order $N^0$ terms, we have \begin{equation} (z_2(q)-z_2(p)) {\cal W}_{1,2,3}(p,q,r) = {\cal W}_{1,3}(p,r) - \mathop{\mathrm{Pol}}_{z_1(p)} V_1'(z_1(p)) {\cal W}_{1,2,3}(p,q,r), \end{equation} where \begin{equation} {\cal W}_{1,3}(p,r) =\frac{1}{Z_2-\tilde{Z}_2}\left(\frac{E(z_1,\tilde{Z}_2,z_3)}{z_1-\tilde{Z}_1}-\frac{E(z_1,Z_2,z_3)}{z_3-Z_3}\right) \end{equation} is known from \cite{Echain}, and where we used the notation: $Z_2=V'_1(z_1)-W_1(z_1)$, $\tilde{Z}_2=V'_3(z_3)-W_3(z_3)$, $Z_3=V'_2(Z_2)-z_1$ and $\tilde{Z}_1=V'_2(\tilde{Z}_2)-z_3$. This equation can be solved and its solution is given by th.\ref{threc} under the form \begin{equation}\label{3mm} {\cal W}_{1,2,3}(p_1,p_2,p_3) =-Res_{Q\rightarrow p_1,p_3^{+i,3}}\left( \frac{{\cal W}_{1,3}(Q,p_3){\displaystyle\prod_{k=1}^{s_2}(z_1(p_1)-z_1(p_2^{+k,2}))}}{(z_1(Q)-z_1(p_1))(z_3(p_3)-z_3(p_1))){\displaystyle\prod_{k=1}^{s_2}(z_1(Q)-z_1(p_2^{+k,2}))}}\right). \end{equation} \subsubsection{Triangulations} In this section we present some explicit computations for the three matrix model. We assume that all of the matrices are subject to cubic potentials of the form \begin{equation} V_i(x)=\frac{1}{2}g_2^{(i)}x^2+\frac{1}{3}g_3^{(i)}x^3. \end{equation} Furthermore, for simplicity we assume $g_2^{(1)}=g_2^{(3)}=1$ and $g_2^{(2)}=3$ in order to obtain a convenient form for the propagator matrix: \begin{equation} C^{-1} :=\left( \begin{array}{ccc} 2&-1&1\cr -1&1&-1\cr 1&-1&2\cr \end{array} \right) . \end{equation} \begin{figure} \begin{center} \includegraphics[width=12cm,angle=00]{2vertex}\\ \end{center} \caption{ Maps contributing to ${\cal T}_{1,0,0}^2$. The type 1 (resp. 2 and 3) boundary condition is represented by the color white (resp. red and blue). The outer face is the marked face of type 1.}\label{2vertex} \end{figure} The conditions \ref{condi}, can be used to express the $z_i$'s as a $\sqrt{T}$-expansion whose coefficients are functions on the spectral curve that also depend on the coefficients of the potentials. \begin{equation} z_i(p)=\sum_{n=1}^\infty z_i^{(n/2)}(g_3^{(1)},g_3^{(2)},g_3^{(3)};p)T^{n/2} \end{equation} For example, the first few orders of $z_1(p)$ are given by \begin{equation} z_1^{(1/2)}\left(g_3^{(j)};p\right)=-\left(p+\frac{2}{p}\right), \end{equation} which can be recognized as the gaussian contribution, \begin{equation} z_1^{(1)}\left(g_3^{(j)}; p\right)=\frac{g_3^{(2)}+3g_3^{(3)}}{p^2}-8g_3^{(1)}-2g_3^{(2)}-4g_3^{(3)}, \end{equation} \begin{eqnarray*} z_1^{(3/2)}\left(g_3^{(j)}; p\right)&=& p \left(-16(g_3^{(1)})^2-7g_3^{(1)}g_3^{(2)}-2(g_3^{(2)})^2-13g_3^{(1)}g_3^{(3)}-7g_3^{(2)}g_3^{(3)}-16(g_3^{(3)})^2\right) \\ &+&\frac{2g_3^{(2)}g_3^{(3)}}{p^3}+\frac{-32g_3^{(1)}-6g_3^{(1)}g_3^{(2)}-2g_3^{(1)}g_3^{(3)}+16(g_3^{(3)})^2}{p}, \end{eqnarray*} \begin{eqnarray*} z_1^{(2)}\left(g_3^{(j)}; p\right)&=& -384(g_3^{(1)})^3-192g_3^{(2)}(g_3^{(1)})^2-60g_3^{(1)}(g_3^{(2)})^2-12(g_3^{(2)})^3-288(g_3^{(1)})^2g_3^{(3)}- \\&-& 162g_3^{(1)}g_3^{(2)}g_3^{(3)}-48(g_3^{(2)})^2g_3^{(3)}-216g_3^{(1)}(g_3^{(3)})^2-120g_3^{(2)}(g_3^{(3)})^2+192(g_3^{(3)})^3\\ &+&\frac{1}{p^4}g_3^{(2)}(g_3^{(3)})^2+\frac{1}{p^2}\left(32(g_3^{(1)})^2g_3^{(2)} 14g_3^{(1)}(g_3^{(2)})^2+4(g_3^{(2)})^3+96(g_3^{(1)})^2g_3^{(3)}\right), \end{eqnarray*} and so on. The same treatment can be given to the other meromorphic parametrizations and obtain similar expressions : \begin{equation} z_2(p)=-\sqrt{T}\left(p+\frac{1}{p}\right)+T\left(\frac{g_3^{(3)}} {p^2}+g_3^{(1)}p^2-4g_3^{(1)}-2g_3^{(2)}-4g_3^{(3)}\right) +... \end{equation} \begin{equation} z_3(p)=-\sqrt{T}\left(2p+\frac{1}{p} \right)+T\left(\left(3g_3^{(1)}+g_3^{(2)}\right)p^2-8g_3^{(1)}-2g_3^{(2)}-4g_3^{(3)}\right)+... \end{equation} \begin{figure} \begin{center} \includegraphics[width=12cm,angle=00]{3vertex}\\ \end{center} \caption{ Maps contributing to ${\cal T}_{1,2,0}^2$. The outer face is the marked face whose boundary has type (1,2,2).}\label{3vertex} \end{figure} It is important to notice that these coefficients have to be found only once and thereafter they can be used to compute the $T$- expansions for the correlation functions without any further modification. Using these parametrizations and \eq{3mm}, we compute order by order the terms in the formal $T$-expansion for the disk amplitude with mixed boundary conditions i.e. \begin{equation} W_{1,2,3}(z_1,z_2,z_3)=\sum_{v=0}^\infty\sum_{\vec{n}\in\mathbb{N}^3}\frac{\mathcal{T}_{\vec{n}}^v(\vec{g_3})}{z_1^{n_1+1}z_2^{n_2+1}z_3^{n_3+1}}T^v. \end{equation} It is important to recall that \begin{equation} \sum_{v=0}^\infty\mathcal{T}_{\vec{n}}^v(\vec{g_3})T^v=\lim_{N\to\infty} {1 \over N} \moy{Tr\left(M_1^{n_1}M_2^{n_2}M_3^{n_3}\right)}. \end{equation} Here $\mathcal{T}^v_{\vec{n}}(\vec{g_3})$ denotes the contributions to the correlation function due to disks with $v$ vertices and whose boundary conditions are given by an ordered sequence $(n_1,n_2, n_3)$ of boundaries of types $1,2$ and $3$ respectively. Since $g_2^{(1)}=g_2^{(3)}$ the symmetry \begin{equation} \mathcal{T}^v_{(n_1,n_2,n_3)}(g^{(1)}_3,g^{(2)}_3,g^{(3)}_3)=\mathcal{T}^v_{(n_3,n_2,n_1)}(g^{(3)}_3,g^{(2)}_3,g^{(1)}_3) \end{equation} holds. Moreover, the $\mathcal{T}^v_{\vec{n}}(\vec{g_3})$'s are given by power expansions in the couplings of the cubic terms in the potentials, i.e \begin{equation} \mathcal{T}^v_{\vec{n}}(\vec{g_3})=\sum_{\vec{m}\in\mathbb{N}^3}\mathcal{T}^v_{{\vec{n}},{\vec{m}}}(g_3^{(1)})^{m_1}(g_3^{(2)})^{m_2}(g_3^{(3)})^{m_3}, \end{equation} here $\mathcal{T}^v_{{\vec{n}},{\vec{m}}}$ has to be understood as the sum of the contributions to $\mathcal{T}^v_{{\vec{n}}}$ containing $m_i$ triangles of type $i=1,2,3$. It is worth seeing this explicitly for a couple of simple examples: \begin{equation} \mathcal{T}_{(1,0,0)}^2(\vec{g_3})=-\left(4g_3^{(1)}+g_3^{(2)}+2g_3^{(3)}\right) \end{equation} \begin{equation} \mathcal{T}_{(1,2,0)}^3(\vec{g_3})=-\left(10g_3^{(1)}+4g_3^{(2)}+7g_3^{(3)}\right) \end{equation} The explicit enumeration of maps giving rise to the coefficients of these numbers is presented in fig.\ref{2vertex} and fig.\ref{3vertex}. The contributions to $W_{1,2,3}$ up to three vertices (omitting terms that can be obtained by symmetry) are displayed in the following table. For clarity we use the simplified notation $g_3^{(1)}=g_1$, $g_3^{(2)}=g_2$ and $g_3^{(3)}=g_3$. \begin{center} \begin{tabular}{ | l || p{2cm} | p{3cm}| p{4.5cm} | } \hline Boundaries $\backslash$ Vertices & 1 & 2 &3\\ \hline $(1,0,0)$&1&$-4g_3-g_2-2g_3$&$-128g_1^3$ $-64g_1^2g_2$ $-20g_1g_2^2$ $-4g_2^3$ $-96g_1^2g_3$ $-54g_1g_2g_3$ $-16g_2^2g_3$ $-72g_1g_3^2$ $-40g_2g_3^2$ $-64g_3^3$\\ \hline $(2,0,0)$&0&2&$64g_1^2$ $+24g_1g_2$ $+4g_2^2$ $+40g_1g_3$ $+12g_2g_3$ $+16g_3^2$\\ \hline $(3,0,0)$&0&0&$-32g1$ $-7g_2$ $-13g_3$\\ \hline $(4,0,0)$&0&0&$8$\\ \hline $(0,1,0)$&1&$-2g_1-g_2-2g_3$& $-64g_1^3$ $-40g_1^2g_3^2$ $-16g_1g_2^2$ $-4g_2^3$ $-56g_1^2g_3$ $-42g_1g_2g_3$ $-16g_2^2g_3$ $-56g_1g_3^2$ $-40g_2g_3^2$ $-64g_3^2$\\ \hline $(0,2,0)$&0&1&$16g_1^3 $ $+12g_1g_2$ $+4g_2^2$ $+18g_1g_3$ $+12g_2g_3$ $+16g_3^3$\\ \hline $(0,3,0)$&0&0&$-7g_1$ $-4g_2$ $-7g_3$\\ \hline $(0,4,0)$&0&0&2\\ \hline $(1,1,0)$&0&1&$32g_1^2$ $+17g_1g_2$ $+4g_2^2$ $+27g_1g_3$ $+12g_2g_3$ $+16g_3^2$\\ \hline $(1,2,0)$&0&0&$-10g_1$ $-4g_2$ $-7g_3$\\ \hline $(1,3,0)$&0&0&2\\ \hline $(2,1,0)$&0&0&$-16g_1$ $-5g_2$ $-9g_3$\\ \hline $(3,1,0)$&0&0&4\\ \hline $(2,2,0)$&0&0&1\\ \hline $(1,0,1)$&0&0&$-32g_1^2$ $-19g_1g_2$ $-5g_2^2$ $-41g_1g_3$ $-19g_2g_3$ $-32g_3^2$\\ \hline $(2,0,1)$&0&-1&$16g_1$ $+7g_2$ $+14g_3$\\ \hline $(3,0,1)$&0&0&-4 \\ \hline $(2,0,2)$&0&0&7\\ \hline $(1,1,1)$&0&0&$-2g_1$ $-g_2$ $-4g_3$\\ \hline $(2,1,1)$&0&0&1\\ \hline $(1,2,1)$&0&0&2\\ \hline \end{tabular} \end{center} \bigskip \section{Conclusion and prospects} In these notes, we found a recursive formula giving access to the so called complete disk amplitude in the multi-matrix model where an arbitrary number of matrices are coupled in chain. As in the two matrix model, this recursive formula only involves the spectral curve of the matrix model. Once the latter is known by the computation of the non-mixed disk amplitude, the recursive procedure presented in this paper explains how to get a disk with mixed boundary conditions by simple series expansion in terms of local variables on the spectral curve. This result proves a conjecture of Eynard \cite{Echain} and is the first step towards the generalization of the topological recursion formalism for the computation of mixed amplitude in the mutli-matrix setup. From the two matrix model experience \cite{EOallmixed}, it was also anticipated that the mixed correlation functions should follow from recursion relations taking into account all possible degenerations of the maps enumerated. This work obviously calls for generalizations and the computation of arbitrary mixed amplitude to any order in the $1/N^2$ expansion will probably be accessible by a recursive procedure similar to this one. This problem will be addressed in a forthcoming paper. On the other hand, a direct combinatorial understanding of these recursion relations is still missing and would deserve further investigations. Can we give any meaning to the so-called non-physical sheets of the spectral curve? Can we get a universal factorization formula for arbitrary disk amplitude as in \cite{EObethe}? These problems are definitely of high interest, not only for combinatorial motivations but also to investigate the generic structure underlying the mixed amplitudes in matrix models. This last point might be fundamental to understand the possible applications of these amplitudes to fields other than random matrix theories, such as topological string theories. \bigskip \bigskip \bigskip \noindent {\Large \bf Aknowledgement} N. O. would like to thank Bertrand Eynard and Aleix Prats-Ferrer for fruitful discussions on the subject. The work of the author is founded by the Funda\c{c}\~ao para a Ci\^{e}ncia e a Tecnologia through the fellowship SFRH/BPD/70371/2010 for N.O and SFRH / BD / 64446 / 2009 for A. V.-O. \newpage \setcounter{section}{0}
\section{Introduction} Let $\Omega\subset\mathbb{R}^n$ be an open set. A \emph{$C^m$ defining function}, $m\geq 1$, for $\Omega$ is a real-valued $C^m$ function $\rho$ defined on a neighborhood $U$ of $\partial\Omega$ such that $\set{x\in U:\rho(x)<0}=\Omega\cap U$ and $\nabla\rho\neq 0$ on $\partial\Omega$. If $\Omega$ has a $C^m$ defining function, we say that $\Omega$ is a \emph{$C^m$ domain}. For many applications on unbounded domains, the preceding definition is inadequate. For example, to work in local coordinates that are adapted to the boundary, it is necessary to work in a neighborhood whose size depends on the $C^2$ norm of the defining function. If the $C^2$ norm is not uniformly bounded, then such neighborhoods may need to be arbitrarily small, which means that a partition of unity subordinate to these neighborhoods might not have uniform bounds on the derivatives. Other problems might arise in constructions which involve choosing a constant large enough to bound quantities depending on derivatives of the defining function. Typical results on $C^m$ domains will require the following: \begin{defn} Let $\Omega\subset\mathbb{R}^n$, and let $\rho$ be a $C^m$ defining function for $\Omega$ defined on a neighborhood $U$ of $\partial\Omega$ such that \begin{enumerate} \item $\dist(\partial\Omega,\partial U)>0$, \item $\pnorm{\rho}_{C^m(U)}<\infty$, \item $\inf_U|\nabla\rho|>0$. \end{enumerate} We say that such a defining function is \emph{uniformly $C^m$}. If $\rho$ on $U$ is uniformly $C^m$ for all $m\in\mathbb{N}$, we say $\rho$ is \emph{uniformly $C^\infty$}. \end{defn} On bounded domains, compactness of the boundary implies that every bounded $C^m$ domain has a uniformly $C^m$ defining function. On unbounded $C^m$ domains with noncompact boundaries, these properties may not hold. For example, consider $\Omega\subset\mathbb{R}^3$ defined by $\Omega=\set{z<x y^2}$. This is a $C^{\infty}$ domain, and any $C^2$ defining function $\rho$ for $\Omega$ will take the form $\rho(x,y,z)=h(x,y,z)(z-x y^2)$ for a $C^1$ function $h$ satisfying $h>0$ on $\partial\Omega$. If we restrict to the line $\ell=\set{y=z=0}\subset\partial\Omega$, we see that $|\nabla\rho||_\ell=h$ and $\frac{\partial^2\rho}{\partial y^2}|_\ell=-2x h$. If $|\nabla\rho|>C_1>0$ on $U$ then $h>C_1$ on $\ell$, but if $\pnorm{\rho}_{C^2(U)}<C_2$ then $2\abs{x}h<C_2$. This is impossible if $\abs{x}\geq\frac{C_2}{2 C_1}$, so no defining function for $\Omega$ is uniformly $C^2$, even though the domain itself is $C^\infty$. The natural choice for a defining function is the signed distance function. For $\Omega\subset\mathbb{R}^n$ with $C^m$ boundary, define the signed distance function for $\Omega$ by \[ \tilde\delta(x)=\begin{cases}d(x,\partial\Omega) & x\notin\Omega\\-d(x,\partial\Omega) & x\in\overline\Omega\end{cases}. \] Note that the distance function $\delta(x) := d(x,\partial\Omega)=|\tilde\delta(x)|$ for any $x\in\mathbb{R}^n$. Let $\unp(\partial\Omega) = \set{ x\in\R^n : \text{ there exists a unique point }y\in \partial\Omega \text{ such that }\delta(x) = |y-x|}$. The following concepts were introduced in \cite{Fed59}. \begin{defn}If $y\in \partial\Omega$, then define the \emph{reach} of $\partial\Omega$ at $y$ by \[ \rea(\partial\Omega,y) = \sup\set{r\geq 0: B(y,r)\subset \unp(\partial\Omega)} \] and the reach of $\partial\Omega$ to be \[ \rea(\partial\Omega) = \inf\set{ \rea(\partial\Omega,y): y\in \partial\Omega}. \] \end{defn} Our main result is the following: \begin{thm}\label{thm:main thm} Let $\Omega\subset\mathbb{R}^n$ be a $C^m$ domain, $m\geq 2$. Then the following are equivalent: \begin{enumerate} \item \label{item:uniform} $\Omega$ has a uniformly $C^m$ defining function. \item \label{item:delta} $\partial\Omega$ has positive reach, and for any $0<\epsilon<\rea(\partial\Omega)$, the signed distance function satisfies $\|\tilde\delta\|_{C^m(U_\ep)}<\infty$ on $U_\epsilon=\set{x\in\mathbb{R}^n:\delta(x)<\epsilon}$. \item \label{item:rho} There exists a $C^m$ defining function $\rho$ for $\Omega$ and a constant $C>0$ such that for every point $p\in\partial\Omega$ with local coordinates $\set{y_1,\ldots,y_n}$ satisfying $\frac{\partial\rho}{\partial y_j}(p)=0$ for $1\leq j\leq n-1$, we have \[ \abs{\nabla\rho(p)}^{-1}\abs{\frac{\partial^k\rho(p)}{\partial y_I\partial y_n^j}}<C \] where $I$ is a multi-index of length $k-j$ with $n\notin I$ for any integers $2\leq k\leq m$ and $0\leq j\leq\min\set{m-k,k}$. \end{enumerate} \end{thm} \begin{rem} An important consequence of this theorem is that our definition of uniformly $C^\infty$ is not too strong. If for every $m\in\mathbb{N}$ there exists a defining function $\rho_m$ on $U_m$ such that $\rho_m$ is uniformly $C^m$ on $U_m$, then there exists a uniformly $C^\infty$ defining function $\rho$, and we can take $\rho$ to be the signed distance function. \end{rem} \begin{rem}\label{rem:Krantz/Parks extension} In \cite{KrPa81}, Krantz and Parks show that if $\Omega$ is a $C^m$ domain, $m\geq 2$, then there exists a neighborhood $U\supset \partial\Omega$ on which $\tilde\delta$ is $C^m$. Part \eqref{item:delta} of Theorem \ref{thm:main thm} extends their result by showing that $\tilde\delta$ is $C^m$ up to $\rea(\partial\Omega)$. \end{rem} \begin{proof} That \eqref{item:delta} implies \eqref{item:uniform} and \eqref{item:uniform} implies \eqref{item:rho} are immediate from the definitions. That \eqref{item:rho} implies \eqref{item:delta} will follow from Lemmas \ref{lem:rho_delta} and \ref{lem:delta_neighborhood}, proved in Section \ref{sec:basic_results}. \end{proof} When studying the asymptotic behavior of a domain, it is natural to consider the domain after embedding $\mathbb{R}^n\subset\mathbb{RP}^n$, and we will do so in Section \ref{sec:projective_space}. Our theorem will make it easy to check that any $C^m$ domain in $\mathbb{R}^n$ which can be extended to a $C^m$ domain in $\mathbb{RP}^n$ under this embedding will have a uniformly $C^m$ defining function. However, we will also show that there are examples which are not even $C^1$ in $\mathbb{RP}^n$ but still have uniformly $C^m$ defining functions. We conclude the paper in Section \ref{sec:applications} with two specific applications of uniformly $C^m$ defining functions. The first is the construction of weighted Sobolev spaces on unbounded domains, and the second is a brief example from several complex variables to illustrate the advantages of uniformly $C^m$ defining functions in generalizing some well-known constructions. Over the course of several papers, we will study domains $\Omega$ that admit a uniformly $C^m$ defining function, build weighted Sobolev spaces on them, and develop the elliptic theory associated to the Sobolev spaces \cite{HaRa13s}. We will then be in a position to investigate the the $\dbar$-Neumann and $\dbarb$-problems in weighted $L^2$ on $\Omega\subset\C^n$. Gansberger has obtained compactness results for the $\dbar$-Neumann operator in weighted $L^2$ \cite{Gan10}, but (at the time) there was neither the elliptic theory nor suitable Sobolev space theory to study the $\dbar$-Neumann problem in $H^s$ or facilitate the passage from the $\dbar$-Neumann operator at the Sobolev scale $s=1/2$ to the complex Green operator on $\partial\Omega$ in weighted $L^2$. There are other results about solution operators to $\dbar$ in the unbounded setting but for the case $\Omega = \C^n$, rendering any boundary discussion moot \cite{HaHe07,Gan11}. \section{Basic Results} \label{sec:basic_results} To handle rigorously multi-indices with possibly repeated indices, we identify functions with sets of ordered pairs and define a multi-index of length $k$ to be a function $I:S\rightarrow\set{1,\ldots,n}$ defined on a subset $S$ of the natural numbers such that $\abs{I}=\abs{S}=k$. If $S=\set{s_1,\ldots,s_k}$ where $\set{s_j}$ is an increasing sequence, we write $I_j=I(s_j)$. Hence, $\frac{\partial}{\partial x_I}=\frac{\partial}{\partial x_{I_1}}\cdots\frac{\partial}{\partial x_{I_k}}$. We will identify $I$ with its range, and write $n\in I$ to mean $n\in\range(I)$. The set of all increasing multi-indices is defined by \[ \mathcal{I}_k=\set{I:\set{1,\ldots,k}\rightarrow\set{1,\ldots,n},I\text{ is an increasing function.}}. \] By the identification of a function with a set of ordered pairs, all set theoretic operations are defined for multi-indices. Below, we will take the $C^k$ norm of a function on $\partial\Omega$. We take an extrinsic view, and for a $C^k$ function $f$ defined on a neighborhood of $\partial\Omega$, we set \[ \|f\|^2_{C^k(\partial\Omega)} = \sup_{p\in\partial\Omega} \sum_{j=0}^k \sum_{I\in\I_j} \Big| \frac{\p^j f(p)}{\p x_I}\Big|^2=\inf_{U\supset\partial\Omega}\pnorm{f}^2_{C^k(U)}. \] The intrinsic $C^k$ norm of a defining function is always zero, hence our use of the extrinsic norm. For $p\in\partial\Omega$, let $\set{y_1,\ldots,y_n}$ be orthonormal coordinates such that $\nabla\tilde\delta(p)=(0,\ldots,0,1)$. For functions $f$ defined in a neighborhood of $p$, we define a family of special $C^k$ norms that is adapted to the boundary. For any integer $k\geq 0$, define \[ \abs{f}^2_{C^{k}_b(p)}=\sum_{k'=0}^{k}\sum_{j'=0}^{\min\set{k-k',k'}}\sum_{I\in\mathcal{I}_{k'-j'},n\notin I}\abs{\frac{\partial^{k'} f(p)}{\partial y_I\partial y_n^{j'}}}^2. \] The $C^k_b$ norms provide a balance between computability (derivatives are only with respect to $\set{y_j}$) and theoretical elegance (intrinsic tangential derivatives and the normal). In particular, terms in the $C^k_b$ norm agree with terms in the expansion of a $k$-fold composition of tangential differential operators with respect to local coordinates. For the purposes of induction, we also define for any integers $k\geq 1$ and $k\geq 2j\geq 0$ \[ \abs{f}^2_{C^{k,j}_b(p)}=\abs{f}^2_{C^{k-1}_b(p)}+\sum_{j'=j}^{\lfloor k/2\rfloor}\sum_{I\in\mathcal{I}_{k-2j'},n\notin I}\abs{\frac{\partial^{k-j'} f(p)}{\partial y_I\partial y_n^{j'}}}^2. \] The $C^{k,j}_b$ are intermediate norms between $C^{k}_b$ and $C^{k-1}_b$. In particular, $\abs{f}^2_{C^{k,0}_b(p)}=\abs{f}^2_{C^k_b(p)}$. Also, when $k$ is even $\abs{f}_{C^{k,k/2}_b(p)}^2 = \abs{f}_{C^{k-1}_b(p)}^2 + \abs{\frac{\p^{k/2}f(p)}{\p y_n^{k/2}}}^2$ and when $k$ is odd $\abs{f}_{C^{k,(k-1)/2}_b(p)}^2 = \abs{f}_{C^{k-1}_b(p)}^2 + \sum_{\ell=1}^{n-1}\abs{\frac{\p^{(k+1)/2}f(p)}{\partial y_\ell\p y_n^{(k-1)/2}}}^2$. In general, if $I$ is a multi-index such that $n\notin I$ and $j\geq 0$ is an integer then \begin{equation} \label{eq:special_norm_estimate} \abs{\frac{\partial^{\abs{I}+j} f(p)}{\partial y_I\partial y_n^j}}\leq\abs{f}_{C^{\abs{I}+2j,j}_b(p)}\leq\abs{f}_{C^{\abs{I}+2j}_b(p)}. \end{equation} The utility of this norm can be seen from the following lemma. \begin{lem} \label{lem:rho_delta} Let $\Omega\subset\mathbb{R}^n$ have $C^m$ boundary, $m\geq 2$. Let $\rho$ be a $C^m$ defining function for $\Omega$ and let $h$ be the positive $C^{m-1}$ function defined in a neighborhood of $\partial\Omega$ by $\tilde\delta=h\rho$. Then \[ \sup_{p\in\partial\Omega}\frac{\abs{\rho}_{C^m_b(p)}}{\abs{\nabla\rho(p)}}<\infty. \] if and only if \[ \|\tilde\delta\|_{C^m(\partial\Omega)}<\infty\textrm{ and }\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{m-2}_b(p)}}{h(p)}<\infty. \] \end{lem} \begin{rem} When $m=2$ the statement about $h$ is trivial, so the conditions on $\rho$ and $\tilde\delta$ are equivalent. We will see in \eqref{eq:delta_C2} that something stronger is true in this case. \end{rem} \begin{proof} Since $|\nabla\tilde\delta|^2=1$ on a neighborhood of $\partial\Omega$ (see \cite{KrPa81} and Theorem 4.8 (3) in \cite{Fed59}), for $I\in\mathcal{I}_k$ with $1\leq k\leq m-1$, we can differentiate this equality by $\frac{\p^k}{\p x_I}$ to obtain \[ \sum_{j=1}^n\sum_{J\subseteq I}\frac{\partial}{\partial x_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial x_J}\right)\frac{\partial}{\partial x_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial x_{I\backslash J}}\right)=0. \] on $\partial\Omega$. For fixed $p\in\partial\Omega$, choose coordinates $(y_1,\ldots,y_n)$ so that $p=0$ and $\nabla_y\tilde\delta(p)=(0,\ldots,0,1)$. In these coordinates, \begin{equation} \label{eq:delta_normal_derivatives} \sum_{j=1}^n\sum_{J\neq\emptyset, J\subsetneq I}\frac{\partial}{\partial y_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial y_J}\right)\frac{\partial}{\partial y_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial y_{I\backslash J}}\right)(p)+2\frac{\partial}{\partial y_n}\left(\frac{\partial^{k}\tilde\delta}{\partial y_{I}}\right)(p)=0. \end{equation} From this, we conclude that \begin{equation} \label{eq:delta_normal_estimate} \abs{\frac{\partial}{\partial y_n}\left(\frac{\partial^{k}\tilde\delta}{\partial y_{I}}\right)(p)}\leq C_1\|\tilde\delta\|_{C^k(\partial\Omega)}^2. \end{equation} for some constant $C_1>0$ and for any $I\in\mathcal{I}_k$ with $1\leq k\leq m-1$. Since $h$ is $C^{m-1}$, we may differentiate $\tilde\delta=h\rho$ in a neighborhood of $\partial\Omega$ by $\frac{\p^k}{\p x_I}$ for $I\in\mathcal{I}_k$ to obtain \[ \frac{\partial^k\tilde\delta}{\partial x_{I}}=\sum_{J\subseteq I}\frac{\partial^{\abs{J}}h}{\partial x_J}\frac{\partial^{k-\abs{J}}\rho}{\partial x_{I\backslash J}}. \] This can not be differentiated directly again if $k=m-1$ because $h$ is only $C^{m-1}$, but we may form a difference quotient at $p\in\partial\Omega$ and take the limit to obtain \begin{equation} \label{eq:k+1_derivatives} \frac{\partial^{k+1}\tilde\delta(p)}{\partial x_{I}}=\sum_{J\subsetneq I}\frac{\partial^{\abs{J}}h(p)}{\partial x_J}\frac{\partial^{k+1-\abs{J}}\rho(p)}{\partial x_{I\backslash J}} \end{equation} for any $I\in\mathcal{I}_{k+1}$, since $\rho(p)=0$. In our special coordinates note that $\frac{\partial\rho}{\partial y_j}(p)=0$ if $j\neq n$, so if $n\notin I$ in these coordinates, all of the terms with first derivatives of $\rho$ will also vanish, leaving us with \begin{equation} \label{eq:delta_tangent_estimate} \abs{\frac{\partial^{k+1}\tilde\delta(p)}{\partial y_{I}}}\leq C_2\abs{h}_{C^{k-1}_b(p)}\abs{\rho}_{C^{k+1}_b(p)} \end{equation} for some constant $C_2>0$. For $0\leq j\leq k'$ and $I\in\mathcal{I}_{k'-j}$ with $n\notin I$, we obtain from \eqref{eq:k+1_derivatives} the equation \begin{multline*} \frac{\partial^{k'+1}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j+1}} =\sum_{\ell=0}^{j}\sum_{J\subseteq I}\binom{j+1}{\ell}\frac{\partial^{\abs{J}+\ell}h(p)}{\partial y_J\partial y_n^\ell}\frac{\partial^{k'-\abs{J}+1-\ell}\rho(p)}{\partial y_{I\backslash J}\partial y_n^{j+1-\ell}}\\ +\sum_{J\subset I,\abs{J}\leq k'-j-2}\frac{\partial^{\abs{J}+j+1}h(p)}{\partial y_J\partial y_n^{j+1}}\frac{\partial^{k'-j-\abs{J}}\rho(p)}{\partial y_{I\backslash J}}. \end{multline*} Subtracting the highest order terms in $h$ (with respect to the $C^k_b$ norm), we can use \eqref{eq:special_norm_estimate} to estimate the remainder by \begin{multline*} \abs{\frac{\partial^{k'+1}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j+1}}-(j+1)\frac{\partial^{k'}h(p)}{\partial y_I\partial y_n^j}\frac{\partial\rho(p)}{\partial y_n} -\sum_{J\subset I,\abs{J}=k'-j-2}\frac{\partial^{k'-1}h(p)}{\partial y_J\partial y_n^{j+1}}\frac{\partial^{2}\rho(p)}{\partial y_{I\backslash J}}}\\ \leq C_3\abs{h}_{C^{k'+j-1}_b(p)}\abs{\rho}_{C^{k'+j+2}_b(p)}. \end{multline*} for some constant $C_3>0$ and integers $k'\geq j+2$. If $j+2>k'\geq j$, we have simply \[ \abs{\frac{\partial^{k'+1}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j+1}}-(j+1)\frac{\partial^{k'}h(p)}{\partial y_I\partial y_n^j}\frac{\partial\rho(p)}{\partial y_n}} \leq C_3\abs{h}_{C^{k'+j-1}_b(p)}\abs{\rho}_{C^{k'+j+2}_b(p)}. \] Suppose that $0\leq j\leq\frac{k-1}{2}$ and set $k'=k-j-1$ (so that $I\in \mathcal{I}_{k-2j-1}$). Note that $\frac{\partial\rho}{\partial y_n}(p)=|\nabla\rho(p)|$ and $\frac{\partial\tilde\delta}{\partial y_n}(p)=1$, so by \eqref{eq:k+1_derivatives} with $k=0$ we have \begin{equation}\label{eqn:h nabla rho =1} h(p)|\nabla\rho(p)| = 1. \end{equation} Thus we have \begin{multline} \label{eq:h_estimate} (j+1)\abs{\frac{\partial^{k-j-1}h(p)}{\partial y_I\partial y_n^j}}h(p)^{-1}\\ \leq \abs{\frac{\partial^{k-j}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j+1}}} +\sum_{J\subset I,\abs{J}=k-2j-3}\abs{\frac{\partial^{k-j-2}h(p)}{\partial y_J\partial y_n^{j+1}}\frac{\partial^{2}\rho(p)}{\partial y_{I\backslash J}}}+C_3\abs{h}_{C^{k-2}_b(p)}\abs{\rho}_{C^{k+1}_b(p)} \end{multline} if $k\geq 2j+3$, and \begin{equation} \label{eq:h_estimate_0} (j+1)\abs{\frac{\partial^{k-j-1}h(p)}{\partial y_I\partial y_n^j}}h(p)^{-1} \leq \abs{\frac{\partial^{k-j}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j+1}}}+C_3\abs{h}_{C^{k-2}_b(p)}\abs{\rho}_{C^{k+1}_b(p)}. \end{equation} if $k<2j+3$. We now proceed by induction. Assume $\sup_{p\in\partial\Omega}\frac{\abs{\rho}_{C^m_b(p)}}{\abs{\nabla\rho(p)}}<\infty$. Suppose that for some $m-1\geq k\geq 1$, $\|\tilde\delta\|_{C^{k}(\partial\Omega)}<\infty$ and $\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{k-2}_b(p)}}{h(p)}<\infty$. When $k=1$, this is clear since $\|\tilde\delta\|_{C^1(\partial\Omega)}=1$ and the condition on $h$ is vacuous. Using $j=\lfloor\frac{k-1}{2}\rfloor$ with \eqref{eq:h_estimate_0} and the induction hypothesis we can show that $\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{k-1,\lfloor(k-1)/2\rfloor}_b(p)}}{h(p)}<\infty$. Suppose that for some $0\leq j\leq\frac{k-3}{2}$ we know that $\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{k-1,j+1}_b(p)}}{h(p)}<\infty$. Using \eqref{eq:h_estimate}, we know now that $\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{k-1,j}_b(p)}}{h(p)}<\infty$ since $|h|_{C^{k-1,j}_b(p)} =|h|_{C^{k-1,j+1}_b(p)} + \sum_{\atopp{I\in \I_{k-2j-1}}{n\not\in I}} | \frac{\p^{k-j-1} h(p)}{\p y_I \p y_n^j} |$. Proceeding by downward induction on $j$ we have $\sup_{p\in\partial\Omega}\frac{\abs{h}_{C^{k-1}_b(p)}}{h(p)}<\infty$. Using \eqref{eq:delta_normal_estimate} and \eqref{eq:delta_tangent_estimate}, we conclude $\|\tilde\delta\|_{C^{k+1}(\partial\Omega)}<\infty$. The result follows by induction on $k$. For the converse, we simply subtract the highest degree term in $\rho$ from \eqref{eq:k+1_derivatives} to obtain for $0\leq j\leq k'+1$ and $I\in\mathcal{I}_{k'+1-j}$ with $n\notin I$ \[ \abs{\frac{\partial^{k'+1}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j}}-h(p)\frac{\partial^{k'+1}\rho(p)}{\partial y_{I}\partial y_n^{j}}}\leq C_4\abs{h}_{C_b^{k'+j-1}(p)}\abs{\rho}_{C_b^{k'+j}(p)}, \] for some constant $C_4>0$. If we set $k'=k-j$ then for any $0\leq j\leq \frac{k+1}{2}$ and $I\in\mathcal{I}_{k-2j+1}$ with $n\notin I$ we have \[ \abs{\nabla\rho(p)}^{-1}\abs{\frac{\partial^{k-j+1}\rho(p)}{\partial y_{I}\partial y_n^{j}}}\leq\abs{\frac{\partial^{k-j+1}\tilde\delta(p)}{\partial y_{I}\partial y_n^{j}}}+C_4\abs{h}_{C_b^{k-1}(p)}\abs{\rho}_{C_b^{k}(p)}. \] The result follows by induction on $k$. \end{proof} Although Lemma \ref{lem:rho_delta} may not apply to all $C^m$ defining functions, it will suffice to prove the main theorem. However, the inductive procedure used to prove this lemma may also be used to construct a system of boundary invariants for any defining function. We illustrate by considering the $m=2$ and $m=3$ cases. By \eqref{eq:delta_normal_estimate}, it will suffice to consider derivatives in tangential directions. Fix $1\leq j,k,\ell\leq n-1$. In the special coordinates of Lemma \ref{lem:rho_delta} at $p$ we apply \eqref{eq:k+1_derivatives} repeatedly to obtain \begin{gather*} 1=h(p)|\nabla\rho(p)|,\qquad \frac{\partial^2\tilde\delta(p)}{\partial y_j\partial y_k}=h(p)\frac{\partial^2\rho(p)}{\partial y_j\partial y_k},\qquad \frac{\partial^2\tilde\delta(p)}{\partial y_j\partial y_n}=h(p)\frac{\partial^2\rho(p)}{\partial y_j\partial y_n}+\frac{\partial h(p)}{\partial y_j}|\nabla\rho(p)|,\\ \frac{\partial^3\tilde\delta(p)}{\partial y_j\partial y_k\partial y_\ell}=h(p)\frac{\partial^3\rho(p)}{\partial y_j\partial y_k\partial y_\ell} +\frac{\partial h(p)}{\partial y_j}\frac{\partial^2\rho(p)}{\partial y_k\partial y_\ell} +\frac{\partial h(p)}{\partial y_k}\frac{\partial^2\rho(p)}{\partial y_j\partial y_\ell}+\frac{\partial h(p)}{\partial y_\ell}\frac{\partial^2\rho(p)}{\partial y_j\partial y_k}. \end{gather*} By \eqref{eq:delta_normal_derivatives}, $\frac{\partial^2\tilde\delta(p)}{\partial y_j\partial y_n}=0$, so we may use (\ref{eqn:h nabla rho =1}) and the previous equalities to conclude \begin{equation} \label{eq:delta_C2} \frac{\partial^2\tilde\delta(p)}{\partial y_j\partial y_k}=|\nabla\rho|^{-1}\frac{\partial^2\rho(p)}{\partial y_j\partial y_k}, \end{equation} and \begin{multline} \label{eq:delta_C3} \frac{\partial^3\tilde\delta(p)}{\partial y_j\partial y_k\partial y_\ell}=|\nabla\rho|^{-1}\frac{\partial^3\rho(p)}{\partial y_j\partial y_k\partial y_\ell}\\ -|\nabla\rho|^{-2}\left(\frac{\partial^2\rho(p)}{\partial y_j\partial y_n}\frac{\partial^2\rho(p)}{\partial y_k\partial y_\ell}+\frac{\partial^2\rho(p)}{\partial y_k\partial y_n}\frac{\partial^2\rho(p)}{\partial y_j\partial y_\ell} +\frac{\partial^2\rho(p)}{\partial y_\ell\partial y_n}\frac{\partial^2\rho(p)}{\partial y_j\partial y_k}\right). \end{multline} Once we have completed the proof of the main theorem, we can derive necessary and sufficient conditions for the existence of uniformly $C^2$ (resp.\ $C^3$) defining functions by checking the boundedness of \eqref{eq:delta_C2} (resp.\ \eqref{eq:delta_C2} and \eqref{eq:delta_C3}). Higher order conditions can be derived as well, but these will be progressively more complicated. To facilitate formulas without special coordinates, we define \[ T_p(\partial\Omega)=\set{t\in\mathbb{R}^n:\sum_{j=1}^n t_j\frac{\partial\tilde\delta}{\partial x_j}(p)=0}. \] We also use the notation $y = (y',y_n)$ for $y'\in\R^{n-1}$ and $y_n\in\R$. \begin{lem} Let $\Omega\subset\mathbb{R}^n$ have a $C^2$ boundary. Then for any $C^2$ defining function $\rho$ we have \begin{equation} \label{eq:C2_reach} \sup_{p\in\partial\Omega}\sup_{\atopp{t\in T_p(\partial\Omega)}{ \abs{t}=1}}|\nabla\rho|^{-1}\abs{\sum_{j,k=1}^n t^j\frac{\partial^2\rho}{\partial x_j\partial x_k}(p)t^k}<\infty \end{equation} if and only if $\partial\Omega$ has positive reach, and \begin{equation} \label{eq:C2_reach_computed} \rea(\partial\Omega)=\left(\sup_{p\in\partial\Omega}\sup_{\atopp{t\in T_p(\partial\Omega)}{ \abs{t}=1}}|\nabla\rho|^{-1}\abs{\sum_{j,k=1}^n t^j\frac{\partial^2\rho}{\partial x_j\partial x_k}(p)t^k}\right)^{-1}. \end{equation} \end{lem} \begin{proof} For $p\in\partial\Omega$, choose local coordinates $(y_1,\ldots,y_n)$ so that $p=0$ and $\nabla\tilde\delta(p)=(0',1)$. Suppose that for some $r>0$, $B((0',r),r)\subset\Omega^c$ and $B((0',-r),r)\subset\Omega$. Then for $y\in\partial\Omega$, $|y-(0',\pm r)|^2 \geq r^2$, so $\abs{y}^2\mp 2 y_n r\geq 0$. Hence $\frac{\abs{y}^2}{2r}\geq \abs{y_n}$. By Theorem 4.18 in \cite{Fed59}, this can be accomplished at every $p\in\partial\Omega$ if and only if $\rea(\partial\Omega)\geq r$. Since the boundary is $C^2$, this is possible at each point if and only if \[ \sum_{j,k=1}^n \abs{t_j\frac{\partial^2\tilde\delta(p)}{\partial x_j\partial x_k}t_k}\leq\frac{\abs{t}^2}{r} \] on $\partial\Omega$ for any vector $t\in T_p(\partial\Omega)$. By \eqref{eq:delta_C2}, this is equivalent to \[ |\nabla\rho|^{-1}\abs{\sum_{j,k=1}^n t_j\frac{\partial^2\rho(p)}{\partial x_j\partial x_k}t_k}\leq\frac{\abs{t}^2}{r} \] for all $p\in\partial\Omega$ and $t\in T_p(\partial\Omega)$. If we take the supremum over all possible $r>0$ satisfying these inequalities, the result follows. \end{proof} \begin{lem} \label{lem:delta_neighborhood} Let $\Omega\subset\mathbb{R}^n$ have a $C^{m}$ boundary for some $m\geq 2$ and suppose that the signed distance function for $\Omega$ satisfies $\|\tilde\delta\|_{C^m(\partial\Omega)}<\infty$. Then for any $0<\epsilon<\rea(\partial\Omega)$ the signed distance function satisfies $\|\tilde\delta\|_{C^m(U)}<\infty$ on $U_\epsilon=\set{x\in\mathbb{R}^n:\delta(x)<\epsilon}$. \end{lem} \begin{proof} By the previous lemma, $\partial\Omega$ has positive reach, so $\tilde\delta$ is a $C^{m}$ function on a neighborhood $U'\supset \p\Omega$ \cite{KrPa81}. Note that the result of Krantz and Parks is essentially local, so it is possible that $d(\partial U',\partial\Omega)=0$ if $\partial\Omega$ is not compact. Set \[ U=\set{x\in\mathbb{R}^n:\delta(x)<\rea(\partial\Omega)}. \] By Theorem 4.8 (3) and (5) in \cite{Fed59}, for any $x\in U$ we have $\nabla\tilde\delta(x)=\nabla\tilde\delta(\pi(x))$, where $\pi(x)=x-\tilde\delta(x)\nabla\tilde\delta(x)\in\partial\Omega$ is the unique boundary point nearest to $x$. This is differentiable, and solving the derivative for $\nabla^2\tilde\delta$ gives us \begin{equation} \label{eq:delta_second_derivatives_neighborhood} \frac{\partial^2\tilde\delta(x)}{\partial x_j\partial x_\ell}=\sum_{\ell'=1}^n\frac{\partial^2\tilde\delta(\pi(x))}{\partial x_j\partial x_{\ell'}}\left(Id+\tilde\delta(x)\nabla^2\tilde\delta(\pi(x))\right)^{-1}_{\ell'\ell} \end{equation} for $x\in U$, where $Id$ is the identity matrix (see \cite{Wei75} and \cite{HeMc10}; see also \eqref{eq:hessian_eigenvalues} below). Note that $Id+\tilde\delta(x)\nabla^2\tilde\delta(x)$ is invertible on $U$ by \eqref{eq:delta_C2} and \eqref{eq:C2_reach_computed}. This formula shows that $\tilde\delta$ is $C^2$ on $U$ (we already know that $\tilde\delta$ is $C^2$ near $\p\Omega$ and $\pi(x)\in\partial\Omega$). Since this formula relates derivatives away from $\partial\Omega$ to derivatives on $\partial\Omega$ (which exist since $\partial\Omega\subset U'$), we may continue to differentiate and use induction to show that $\tilde\delta$ is $C^m$ on $U$. Fix $p\in\partial\Omega$ and choose new coordinates $(y_1,\ldots,y_n)$ so that $p=0$, $\nabla\tilde\delta(p)=(0',1)$, and $\nabla^2\tilde\delta(p)$ is diagonalized with eigenvalues $\kappa_1,\ldots,\kappa_n$. By Theorem 4.8 (3) in \cite{Fed59}, when $y'=0'$ and $\abs{y_n}<\rea(\partial\Omega)$, we have $\tilde\delta(y)=y_n$ and $\nabla\tilde\delta(y)=(0',1)$. Differentiating $|\nabla\tilde\delta|^2=1$ once demonstrates that $\kappa_n=0$. For $m\geq 3$, differentiating $|\nabla\tilde\delta|^2=1$ twice yields \[ 2\sum_{\ell=1}^n\left(\frac{\partial\tilde\delta}{\partial x_\ell}\frac{\partial^3\tilde\delta}{\partial x_\ell\partial x_j\partial x_k} +\frac{\partial^2\tilde\delta}{\partial x_\ell\partial x_j}\frac{\partial^2\tilde\delta}{\partial x_\ell\partial x_k}\right)=0 \] on $U$. From \eqref{eq:delta_second_derivatives_neighborhood}, we can see that eigenvectors of $\nabla^2\tilde\delta$ are preserved along the normal direction. Rewriting the above equation in our $y$-coordinates, when $j=k$ we have $2\left(\frac{\partial\kappa_j}{\partial y_n}+\kappa_j^2\right)(y)=0$ on $U$ when $y'=0'$. The unique solution to this equation is given by \begin{equation} \label{eq:hessian_eigenvalues} \kappa_j(y)=\frac{\kappa_j(0)}{1+y_n\kappa_j(0)} \end{equation} (see also Lemma 14.17 in \cite{GiTr01}, but with the opposite sign convention). Since $\rea(\partial\Omega)\leq\abs{\kappa_j(0)}^{-1}$ (see \eqref{eq:delta_C2} and \eqref{eq:C2_reach_computed}) for all $1\leq j\leq n-1$ with $\kappa_j\neq 0$, $\kappa_j$ will be uniformly bounded on $U_\epsilon$. For $3\leq k\leq m-1$, let $I\in\mathcal{I}_k$. Then differentiating $|\nabla\tilde\delta|^2=1$ gives us \[ \sum_{j=1}^n\sum_{J\subseteq I}\frac{\partial}{\partial x_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial x_J}\right)\frac{\partial}{\partial x_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial x_{I\backslash J}}\right)=0. \] on $U$. In our diagonalized coordinates, we can evaluate terms involving only first or second derivatives separately to obtain \[ 2\frac{\partial^{k+1}\tilde\delta}{\partial y_n\partial y_{I}}+2\sum_{j=1}^k\kappa_{I_j}\frac{\partial^k\tilde\delta}{\partial y_{I}} +\sum_{j=1}^n\sum_{J\subset I,2\leq\abs{J}\leq k-2}\frac{\partial}{\partial y_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial y_J}\right)\frac{\partial}{\partial y_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial y_{I\backslash J}}\right)=0 \] on $U$ when $y'=0'$. If we set \[ \mu_{I}(y_n)=\prod_{j=1}^k(1+y_n\kappa_{I_j}(0)), \] then $\mu_{I}(y_n)$ solves the initial value problem \[ \frac{\partial\mu_{I}}{\partial y_n}(y_n)=\mu_{I}(y_n)\sum_{j=1}^k\kappa_{I_j}(0',y_n)\text{ and }\mu_{I}(0)=1, \] so \[ 2\frac{\partial}{\partial y_n}\left(\mu_{I}\frac{\partial^{k}\tilde\delta}{\partial y_{I}}\right)+\mu_{I}\sum_{j=1}^n\sum_{J\subset I,2\leq\abs{J}\leq k-2}\frac{\partial}{\partial y_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial y_J}\right)\frac{\partial}{\partial y_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial y_{I\backslash J}}\right)=0 \] on $U$ when $y'=0'$. Hence, we may integrate to obtain \begin{multline} \label{eq:delta_derivatives} \frac{\partial^{k}\tilde\delta}{\partial y_{I}}(0',y_n)=\frac{1}{\mu_{I}(y_n)}\frac{\partial^{k}\tilde\delta}{\partial y_{I}}(0)\\ -\frac{1}{2\mu_{I}(y_n)}\int_0^{y_n}\mu_{I}(t)\sum_{j=1}^n\sum_{J\subset I,2\leq\abs{J}\leq k-2}\frac{\partial}{\partial y_j}\left(\frac{\partial^{\abs{J}}\tilde\delta}{\partial y_J}\right)\frac{\partial}{\partial y_j}\left(\frac{\partial^{k-\abs{J}}\tilde\delta}{\partial y_{I\backslash J}}\right)(0',t)dt. \end{multline} Since $\mu_{I}(y_n)$ is uniformly bounded below on $U_\epsilon$ and the terms in the integral are differentiated at most $k-1$ times, we may use induction on $k$ to obtain uniform bounds on $\frac{\partial^{k}\tilde\delta}{\partial y_{I}}$ on $U_\epsilon$ for all $I$ with $3\leq k\leq m-1$. Now, we wish to differentiate our formulas for $k=m-1$ to show that they also hold for $k=m$. By differentiating \eqref{eq:delta_second_derivatives_neighborhood}, we can obtain formulas for the first $m$ derivatives of $\tilde\delta$ on $U$ in terms of derivatives restricted to $\partial\Omega$. By formal manipulations, these must be equivalent to those obtained in \eqref{eq:delta_derivatives}, and hence the first $m$ derivatives remain uniformly bounded on $U_\epsilon$. \end{proof} \section{Examples in Projective Space} \label{sec:projective_space} A large class of examples of domains with uniformly $C^m$ defining functions can be found by considering $\mathbb{R}^n\subset\mathbb{RP}^n$. Recall that $\mathbb{RP}^n=(\mathbb{R}^n\backslash\set{0})/\sim$ under the equivalence relation $x\sim y$ if $x=\lambda y$ for $\lambda\in\mathbb{R}\backslash\set{0}$. If we denote coordinates on $\mathbb{RP}^n$ by $[x_1:\ldots:x_{n+1}]$, the canonical embedding of $\mathbb{R}^n$ in $\mathbb{RP}^n$ is given by $(x_1,\ldots,x_n)\mapsto[x_1:\ldots:x_n:1]$. Every unbounded domain $\Omega$ in $\mathbb{R}^n$ can be extended to a bounded domain $\tilde\Omega$ in $\mathbb{RP}^n$ with respect to this embedding. Conversely, from a domain $\tilde\Omega\subset\R\P^n$, we can canonically produce a (possibly unbounded) domain $\Omega\subset\R^n$ under the mapping $[x_1:\ldots:x_n:1] \mapsto (x_1,\dots,x_n)$. \begin{cor} Let $\tilde\Omega\subset\mathbb{RP}^n$ be a $C^m$ domain. Then the domain $\Omega\subset\mathbb{R}^n$ obtained by pulling back along the canonical embedding has a uniformly $C^m$ defining function. \end{cor} \begin{proof} For $S^n\subset\mathbb{R}^{n+1}$, we may define $\tilde\Omega$ by a $C^m$ defining function $\tilde\rho:S^n\rightarrow\mathbb{R}$ such that $\tilde\rho(-x)=\tilde\rho(x)$. Extend $\tilde\rho$ to all of $\mathbb{R}^{n+1}\backslash\set{0}$ by $\tilde\rho(x)=\tilde\rho\left(\frac{x}{\abs{x}}\right)$. Since we have $\tilde\rho(x)=\tilde\rho(\lambda x)$ for any $\lambda\in\mathbb{R}\backslash\set{0}$, we also obtain $\nabla^k\tilde\rho(x)=\lambda^k(\nabla^k\tilde\rho)(\lambda x)$. If we assume that $\abs{\nabla^k\tilde\rho}<C_k$ and $\abs{\nabla\tilde\rho}>C_0$ on $\partial\tilde\Omega\cap S^n$ for any integer $1\leq k\leq m$ and some constants $C_0,C_k>0$, then we have in general $\abs{\nabla^k\tilde\rho(x)}<C_k\abs{x}^{-k}$ and $\abs{\nabla\tilde\rho(x)}>C_0\abs{x}^{-1}$ whenever $\tilde\rho(x)=0$. A defining function $\rho$ for $\Omega\subset\mathbb{R}^n$ can now be obtained by considering $\rho(x_1,\ldots,x_n)=\tilde\rho(x_1,\ldots,x_n,1)$. Thus $\abs{\nabla^k\rho(x)}<\frac{C_k}{(1+\abs{x}^2)^{k/2}}$ and $\abs{\nabla\rho(x)}>\frac{C_0}{(1+\abs{x}^2)^{1/2}}$ on $\partial\Omega$, so \[ \frac{\abs{\nabla^k\rho}}{|\nabla\rho|}(x)<\frac{C_k}{C_0(1+\abs{x}^2)^{(k-1)/2}} \] on $\partial\Omega$ for all $1\leq k\leq m$. By our main theorem, this implies that $\Omega$ has a uniformly $C^m$ defining function. \end{proof} Note that this proof can still be used if $\tilde\rho$ is $C^m$ when $x_{n+1}\neq 0$ and $\abs{\nabla^k\tilde\rho(x)}<C_k x_{n+1}^{1-k}$ for $x\in S^n$ with $x_{n+1}\neq 0$, so a uniformly $C^m$ defining function in $\mathbb{R}^n$ covers a much larger class of examples than those given by $C^m$ domains in $\mathbb{RP}^n$. For example, consider the domain $\Omega_1\subset\mathbb{R}^2$ defined by \[ \Omega_1=\set{y<x^{-1}\sin x,x\neq 0}\cup\set{y<1,x=0}. \] Then $\Omega_1$ is a $C^\infty$ domain. Let $\rho_1(x,y)=y-x^{-1}\sin x$ when $x\neq 0$ and $\rho_1(0,y)=y-1$. By considering the Maclaurin series of $\sin x$ we can see that $\rho_1$ is real-analytic (hence smooth) in a neighborhood of the set where $x=0$. When $x\neq 0$, we have $\nabla\rho_1=(x^{-2}\sin x-x^{-1}\cos x,1)$, so $\abs{\nabla\rho_1}$ is uniformly bounded above and away from zero. Differentiating $m$ times, we have $\abs{\frac{\partial^m\rho_1}{\partial x^m}}\leq O(x^{-1})$, so this is also uniformly bounded. Hence $\rho_1$ is a uniformly $C^m$ defining function for any integer $m$. In $\mathbb{RP}^2$, this defining function can be written $\rho_1([x:y:z])=\frac{y}{z}-\frac{z}{x}\sin\left(\frac{x}{z}\right)$. On the coordinate patch where $x\neq 0$, this can be written $\rho_1(y,z)=\frac{y}{z}-z\sin(1/z)$. To normalize this near $z=0$, we use $\tilde\rho_1(y,z)=y-z^2\sin(1/z)$. Note that for fixed $y$ this is the classic example of a function which is differentiable at $z=0$ but not $C^1$ in a neighborhood of $z=0$. We conclude that $\tilde\Omega_1\subset\mathbb{RP}^2$ is not a $C^1$ domain. On the other hand, consider $\Omega_2\subset\mathbb{R}^2$ defined by \[ \Omega_2=\set{y<x^{-2}\sin x^2,x\neq 0}\cup\set{y<1,x=0}. \] Let $\rho_2(x,y)=y-x^{-2}\sin x^2$ when $x\neq 0$ and $\rho_2(0,y)=y-1$. Again, the Maclaurin series will show that all derivatives are uniformly bounded near $x=0$, so we focus on $x\neq 0$. Since $\nabla\rho_2=(2x^{-3}\sin x^2-2x^{-1}\cos x^2,1)$, we define $D_1=\frac{1}{\abs{\nabla\rho_2}}\frac{\partial}{\partial x}-\frac{2x^{-3}\sin x^2-2x^{-1}\cos x^2}{\abs{\nabla\rho_2}}\frac{\partial}{\partial y}$ and $D_2=\frac{2x^{-3}\sin x^2-2x^{-1}\cos x^2}{\abs{\nabla\rho_2}}\frac{\partial}{\partial x}+\frac{1}{\abs{\nabla\rho_2}}\frac{\partial}{\partial y}$ to represent the tangent and normal directions on the boundary. Since $\frac{\partial^2\rho_2}{\partial x^2}=4\sin x^2+O(x^{-2})$, $\rho_2$ is a uniformly $C^2$ defining function for $\Omega$, and hence $\partial\Omega$ has positive reach. However, $\frac{\partial^3\rho_2}{\partial x^3}=8x\cos x^2+O(x^{-1})$. If we fix $p = (p_x,p_y)\in\partial\Omega$ with $p_x\neq 0$, then \eqref{eq:delta_C3} tells us that \[ \begin{split} (D_1|_p)^3\tilde\delta(p)&=\abs{\nabla\rho_2}^{-1}(D_1|_p)^3\rho_2(p)-3\abs{\nabla\rho_2}^{-2}((D_1|_p)(D_2|_p)\rho_2(p))((D_1|_p)^2\rho_2(p))\\ &=8x\cos x^2+O(x^{-1}). \end{split} \] This is not uniformly bounded, so there does not exist a uniformly $C^3$ defining function for $\partial\Omega$. Hence, positive reach does not suffice to extend $C^m$ defining functions as uniformly $C^m$ defining functions. Finally, let $h(x,y)=e^{x^2}$. Then $\sup\frac{\abs{h}_{C_b^1}}{h}=\infty$ with respect to either of the previous two examples (since $\frac{\partial}{\partial x}$ is asymptotically the tangential direction in these examples). By Lemma \ref{lem:rho_delta}, the defining function $\rho^h_1(x,y)=(y-x^{-1}\sin x)h(x,y)$ fails to satisfy $\sup\frac{\abs{\rho^h_1}_{C_b^3}}{\abs{\nabla\rho^h_1}}<\infty$ even though this defines a domain with a uniformly $C^3$ defining function. Hence, not all defining functions need satisfy the conditions of Lemma \ref{lem:rho_delta}. Turning to our other example, $\rho_2^h(x,y)=(y-x^{-2}\sin x^2)h(x,y)$ still satisfies $\sup\frac{\abs{\rho_2^h}_{C_b^2}}{\abs{\nabla\rho_2^h}}<\infty$ even though $\pnorm{\rho_2^h}_{C^2(\partial\Omega)}=\infty$. Thus, it is helpful to consider the special $C_b^k$ norm in place of the standard extrinsic $C^k$ norm. \section{Applications of uniformly $C^m$ defining functions}\label{sec:applications} In \cite{HaRa13s}, we define weighted Sobolev spaces on the boundaries of unbounded domains. From the standpoint of the present paper, the weight function is irrelevant. However, it seems difficult to obtain elliptic regularity results without a weight, so for the sake of defining a meaningful space of functions we will include the weight. The weight functions that we use satisfy a number of technical hypotheses (similar to those in \cite{HaHe07,Gan10,Gan11}) all satisfied by $\vp_t(x) = t|x|^2$, $t\in\R\setminus\{0\}$. Let $\Omega\subset\R^n$ have a $C^m$ boundary, $m\geq 2$, that admits a uniformly $C^m$ defining function. We define weighted Sobolev spaces both on the boundary and near the boundary. Suppose $\Omega\subset\mathbb{R}^n$ admits a uniformly $C^2$ defining function. By Theorem \ref{thm:main thm}, $\partial\Omega$ has positive reach, so for $\frac{1}{2}\rea(\partial\Omega)>\ep>0$ we set \[ \Omega_\ep = \{x\in\Omega:\delta(x)<\ep\}. \] Since $\|\tilde\delta\|_{C^2(\Omega_{2\epsilon})}<\infty$, there exists a radius $\frac{1}{4}\rea(\partial\Omega)>r>0$ such that whenever $B(p,r)\cap\Omega_\epsilon\neq\emptyset$, there exist coordinates on $B(p,r)$ such that the level curves of $\tilde\delta$ can be written as a graph. Hence, there exists an orthonormal basis $L_1,\dots,L_{n-1}$ of the tangent space to the level curves of $\tilde\delta$ on $B(p,r)$. We also let $L_n = \nu$ be the unit outward normal to the level curves of $\tilde\delta$. For $1\leq j \leq n$, set \[ T_j = L_j - L_j(\vp_t). \] We call a first order differential operator $T$ \emph{tangential} if the first order component of $T$ is tangential. Note that we use $T_j$ instead of $L_j$ for technical reasons involving integration by parts in weighted norms, but these are not relevant for the present paper. Let $\set{p_j}$ be an enumeration of all points in $\mathbb{R}^n$ whose coordinates are integral multiples of $\frac{r}{\sqrt{n}}$. Then $\set{\overline{B(p_j,r/2)}}$ is a locally finite cover of $\mathbb{R}^n$, with a uniform upper bound on the number of sets covering each point. If $\chi\in C^\infty_0(B(0,r))$ satisfies $\chi=1$ on $B(0,r/2)$ and $1\geq\chi\geq 0$, we can construct a partition of unity subordinate to $\set{B(p_j,r)}$ by using $\chi_j(x)=\chi(x-p_j)/\left(\sum_k\chi(x-p_k)\right)$. Because there is a uniform upper bound on the number of nonzero terms in the denominator, we have a uniform bound on $\|\chi_j\|_{C^m}$ for any $m\geq 0$. Let $\set{U_j}$ be a restriction of this cover to include only those sets covering $\Omega_\epsilon$, with the corresponding modification to $\chi_j$. For any distribution $v$ on $\Omega_\epsilon$, we set $v_j = v \chi_j$, so $v = \sum_{j=1}^\infty v_j$. If $\Omega$ admits a uniformly $C^m$ defining function, $m\geq 2$, we define the weighted Sobolev space $W^{k,p}(\Omega_\ep,\vp_t,\nabla\vp_t)$, $0\leq k\leq m$, as the space of distributions $v$ on $\Omega_\ep$ whose partial derivatives up to order $k$ agree with functions and have the following norm finite: \[ \norm v \norm_{W^{k,p}(\Omega_\ep,\vp_t,\nabla\vp_t)}^p = \sum_{j=1}^\infty \sum_{|\alpha|\leq k} \| T^\alpha v_j \|_{L^p(\Omega_\ep,\vp_t)}^p \] where $T_j = L_j - L_j(\vp_t)$ is well-defined on $U_j$ and the composition $T^\alpha = T_{\alpha_1}\cdots T_{\alpha_{|\alpha|}}$. For the boundary Sobolev space, set \begin{multline*} W^{k,p}(\partial\Omega,\vp_t,\nabla\vp_t) \\ = \{ f \in L^p(\partial\Omega,\vp_t): T^\alpha f \in L^p(\partial\Omega,\vp_t),\ |\alpha|\leq k \text{ and }T_{\alpha_j} \text{ is tangential for } 1\leq j \leq |\alpha|\}. \end{multline*} Choosing uniform neighborhoods with good local coordinates only makes sense on domains with positive reach, and compositions of derivatives would be extremely difficult to control without a uniformly $C^m$ defining function. When $p=2$, we define fractional Sobolev spaces via interpolation and can prove many of the standard Sobolev space results. We also provide an example from several complex variables. The following theorem is well known in the bounded case (see for example Theorem 3.4.4 in \cite{ChSh01}). \begin{thm} Let $\Omega\subset\mathbb{C}^n$ be a domain with a $C^2$ defining function $r$ and a constant $C>0$ satisfying \begin{equation} \label{eq:strict_pseudoconvexity} \sum_{j,k=1}^n \frac{\partial^2r}{\partial z_j\partial\bar{z}_k}t_j\bar{t}_k\geq C\abs{\nabla r}\sum_{j=1}^n\abs{t_j}^2 \end{equation} on $\partial\Omega$ for $t\in T_p(\partial\Omega)$. \begin{enumerate} \item If $\Omega$ admits a uniformly $C^2$ defining function, then there exists a defining function which is strictly plurisubharmonic on $\partial\Omega$. \item If $\Omega$ admits a uniformly $C^3$ defining function, then there exists a defining function which is plurisubharmonic on $\Omega$ and strictly plurisubharmonic on $\{z\in\Omega : \delta(z) < \varepsilon\}$ for some $\varepsilon>0$. \end{enumerate} \end{thm} \begin{rem} The assumption \eqref{eq:strict_pseudoconvexity} implies that $\Omega$ is strictly pseudoconvex, but the uniform lower bound on the Levi-form is not true for all strictly pseudoconvex domains in the unbounded case. For an example that satisfies our condition, consider the tube in $\mathbb{C}^n$ defined by the defining function $r(z)=\abs{z'}^2+(\im z_n)^2-1$. \end{rem} \begin{rem} The second statement is not sharp in the bounded case, where \eqref{eq:strict_pseudoconvexity} alone (without the $C^3$ assumption) will guarantee the existence of a strictly plurisubharmonic defining function. We require $C^3$ to govern the decay rate of \eqref{eq:strict_pseudoconvexity} off of $\partial\Omega$, and our resulting function is merely plurisubharmonic because we can not use $\abs{z}^2$ to obtain strict plurisubharmonicity in the interior (it is no longer a bounded function). \end{rem} \begin{proof} Since \eqref{eq:strict_pseudoconvexity} is independent of the choice of defining function, it will be satisfied by the signed distance function. For $\lambda>0$ to be determined later, define \[ \rho(z)=e^{\lambda\tilde\delta(z)}-1 \] for $z$ in a small neighborhood of $\partial\Omega$. For $v:\Omega_\epsilon\rightarrow\mathbb{C}^n$, we may decompose $v=\tau+\nu$, where $\sum_{j=1}^n\frac{\partial\tilde\delta}{\partial z_j}\tau_j=0$ and $\nu$ is a scalar multiple of $\left(\frac{\partial\tilde\delta}{\partial \bar{z}_1},\ldots,\frac{\partial\tilde\delta}{\partial \bar{z}_n}\right)$. Then since $\left|\sum_{j=1}^n\frac{\partial\tilde\delta}{\partial z_j}v_j\right|=\frac{1}{2}\sqrt{\sum_{j=1}^n\abs{\nu_j}^2}$ we have \[ \sum_{j,k=1}^n \frac{\partial^2\rho}{\partial z_j\partial\bar{z}_k}v_j \bar{v}_k=\lambda e^{\lambda\tilde\delta}\sum_{j,k=1}^n \frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}v_j \bar{v}_k+\lambda^2 e^{\lambda\tilde\delta}\frac{1}{4}\sum_{j=1}^n\abs{\nu_j}^2. \] Since $\tilde\delta$ satisfies \eqref{eq:strict_pseudoconvexity} on $\partial\Omega$, we have \begin{multline*} \sum_{j,k=1}^n \frac{\partial^2\rho}{\partial z_j\partial\bar{z}_k}v_j \bar{v}_k\geq\lambda C\sum_{j=1}^n\abs{\tau_j}^2+\lambda \sum_{j,k=1}^n 2\re\left(\frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}\tau_j \bar{\nu}_k\right)\\ +\lambda \sum_{j,k=1}^n \frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}\nu_j \bar{\nu}_k+\lambda^2 \frac{1}{4}\sum_{j=1}^n\abs{\nu_j}^2 \end{multline*} on $\partial\Omega$. Since $\tilde\delta$ is uniformly $C^2$, there exists a constant $C_2>0$ such that $\abs{\frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}}\leq C_2$ on $\partial\Omega$. Hence, the Cauchy-Schwarz inequality gives us \[ \sum_{j,k=1}^n \frac{\partial^2\rho}{\partial z_j\partial\bar{z}_k}v_j \bar{v}_k\geq\lambda C\abs{\tau}^2-2\lambda C_2\abs{\tau}\abs{\nu} -\lambda C_2\abs{\nu}^2+\lambda^2 \frac{1}{4}\abs{\nu}^2. \] This is strictly positive provided that $C\left(\frac{1}{4}\lambda-C_2\right)> 4C_2^2$. Hence, we may choose $\lambda$ sufficiently large so that $\rho$ is strictly plurisubharmonic on $\partial\Omega$. If we assume that $\tilde\delta$ is uniformly $C^3$, then from \eqref{eq:strict_pseudoconvexity} there exists some uniform neighborhood $U$ of $\partial\Omega$ on which \[ \sum_{j,k=1}^n \frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}\tau_j\bar{\tau}_k\geq \frac{1}{2}C\sum_{j=1}^n\abs{\tau_j}^2. \] We may assume $\abs{\frac{\partial^2\tilde\delta}{\partial z_j\partial\bar{z}_k}}\leq C_2$ on $U$, so that \[ \sum_{j,k=1}^n \frac{\partial^2\rho}{\partial z_j\partial\bar{z}_k}v_j \bar{v}_k\geq\lambda e^{\lambda\tilde\delta}\frac{1}{2}C\abs{\tau}^2+2\lambda e^{\lambda\tilde\delta}C_2\abs{\tau}\abs{\nu} +\lambda e^{\lambda\tilde\delta}C_2\abs{\nu}^2+\lambda^2 e^{\lambda\tilde\delta}\frac{1}{4}\abs{\nu}^2. \] This is positive provided that $\frac{1}{2}C\left(\frac{1}{4}\lambda-C_2\right)\geq 4C_2^2$, so we may again choose $\lambda$ sufficiently large so that $\rho$ is plurisubharmonic on $\partial\Omega$. To extend $\rho$ to all of $\Omega$, let $A=\sup_{\Omega\backslash U}\tilde\delta$. Since we were able to choose a uniform neighborhood $U$, $A<0$. $\hat\rho=\max\set{\rho,A}$ will be a Lipschitz plurisubharmonic defining function for $\Omega$, and a smooth convex approximation to $\max$ can be used to obtain a smooth plurisubharmonic defining function for $\Omega$. \end{proof} \bibliographystyle{alpha}
\section{Introduction} Galaxy redshift surveys are powerful probes of cosmology. The main observables able to constrain cosmological parameters are the overall shape of the galaxy power spectrum at wavenumbers $k<0.1\, \mathrm{Mpc}^{-1}$ and the baryonic acoustic oscillations within it. These are treated as proxies for the matter power spectrum for which we can make robust theoretical predictions. Galaxies, however, are biased tracers of the cosmic mass distribution and many features appearing in their power spectrum depend on how a specific observational sample was selected. To reconstruct the matter power spectrum we thus need an accurate bias model whose free coefficients should be used as nuisance parameters and marginalized over. In the era of precision cosmology, where measurements of the matter power spectrum with per cent accuracy are required, this task is particularly demanding. Bias models can be divided into two broad classes. Eulerian biasing schemes relate the galaxy density contrast, $\delta_{\rm g}({\bf x},t)$, to the matter density distribution, $\delta$, evaluated at the same time $t$ (but not necessarily at the same spatial location). After smoothing the fields on large scales, so that $|\delta|$ is typically much smaller than unity, one can write \citep{b27} \begin{eqnarray} \label{nonloceul} \delta_{\rm g}({\bf x})&=&B_0+\int d^3x_1\,B_1({\bf x}-{\bf x_1})\,\delta({\bf x_1})+\\ &+&\frac{1}{2} \int d^3x_1\,d^3x_2\,B_2({\bf x}-{\bf x_1},{\bf x}-{\bf x}_2)\,\delta({\bf x_1})\,\delta({\bf x_2})+ \nonumber\\ &+& \dots \,, \nonumber \end{eqnarray} where all fields are evaluated at the same time $t$ and the details of the bias model are specified by the kernel functions, $B_{\rm i}$. If one further assumes that biasing is local (i.e. that all kernels can be written as products of Dirac delta distributions), this reduces to \begin{equation} \delta_{\rm g}({\bf x})=b_0+b_1\,\delta({\bf x})+\frac{b_2}{2}\,\delta^2({\bf x})+\dots \,, \label{loceul} \end{equation} where now the bias coefficients $b_{\rm i}$ are real numbers. In Lagrangian bias models, on the other hand, one considers the regions in the initial conditions that will collapse to form galaxies (or their hosting dark-matter haloes) at time $t$ and writes their density contrast, $\delta^{\rm L}_{\rm g}({\bf q})$, in terms of the linear density contrast, $\delta_0({\bf q})$. Large-scale expansions analogous to eqs. (\ref{nonloceul}) and (\ref{loceul}) can also be written in this case. As a second step, one must determine the final position, ${\bf x}({\bf q},t)$, of a fluid element initially located at ${\bf q}$, and compute $\delta_{\rm g}({\bf x}({\bf q},t),t)$ out of $\delta^{\rm L}_{\rm g}({\bf q})$. This Lagrangian-to-Eulerian mapping (LEM) accounts for gravitationally induced motions that determine the final position of the objects. Local Eulerian and Lagrangian bias schemes are not equivalent; they generate a different shape for the galaxy bispectrum \citep{b28} and are not compatible within the framework of perturbation theory \citep{b29}. In fact, \citet{b30} have shown that a local Lagrangian biasing scheme generates a non-linear, non-local and stochastic bias in Eulerian space. On the other hand, non-local Eulerian and Lagrangian schemes are equivalent and can be seen as different mathematical representations of the same physical process \citep{b29}. Due to its simplicity, the local Eulerian model is by far the most widely used in practical applications, such as perturbative calculations. However, it is purely phenomenological and does not have a strong theoretical motivation. Detailed comparison with numerical simulations has also evidenced its limited validity (e.g. \citealt{b31}, \citealt{b32}). Physical models of bias are generally given in the Lagrangian framework as conditions on galaxy (halo) formation are more easily imposed onto the linear density field using some model for the collapse of density perturbations. \citet[][hereafter MW]{b22} used a Press-Schechter-like argument \citep{b34} to compute the bias coefficients of a local Lagrangian scheme as a function of halo mass. The same authors also showed how these parameters can be combined to calculate the bias coefficients of a local Eulerian scheme assuming that large-scale density perturbations follow the spherical collapse model. The effect of non-linear shear on the LEM was discussed by \citet{b30} using the Zel'dovich approximation \citep{b33}. For halo masses $M>M_*(z)$ (where $M_*(z)$ is the characteristic mass for collapse at redshift $z$) the MW formula for $b_1$ is in good agreement with the predictions of N-body simulations \citep{b26}. At lower masses, however, the agreement rapidly deteriorates \citep{b35}. \citet{b23} and \citet{b36} showed that the discrepancy between the N-body simulations and the analytical predictions is already present in Lagrangian space and should thus be attributed to the limitations of the Press-Schechter formalism rather than to the approximated treatment of the LEM. The Lagrangian bias emerging from the extended Press-Schechter model \citep{b38} was first derived by \citet{b37}, rediscussed in \citet{b39} and tested against simulations by \citet{b40}. This approach follows correlated trajectories of $\delta_0$ at different Lagrangian locations as a function of the smoothing scale and looks for correlations in the first-crossing scales of a density threshold. According to the peak-background-split argument (\citealt{b2}; \citealt{b21}), long-wavelength density fluctuations modulate halo formation by modifying the collapse time of localized short-wavelength perturbations. This makes it possible to generalise the calculation of the Lagrangian MW bias coefficients to any model for the halo mass function (\citealt{b41}, \citealt{b43}, \citealt{b44}) and to improve the agreement with N-body simulations (\citealt{b45}, \citealt{b42}, \citealt{b46}, \citealt{b6}). \citet{b15} explained the strong clustering of Abell clusters by assuming that they originate from the regions above a density threshold in the (suitably smoothed) linear density field. Following this line of reasoning, it is common to assume that dark-matter haloes form out of linear density peaks, as an alternative to the Press-Schechter approach. Tests against N-body simulations have shown that this is a well justified assumption, especially for massive haloes (\citealt{b8}, see also \citealt{b51}). The statistical properties of the local maxima in a Gaussian random field have been extensively studied by \citet{b2} (see also \citealt{b16} and \citealt{b17}). \citet{b26} introduced peaks theory in the MW formalism, while \citet{b24} evaluated the level of stochasticity in the Lagrangian clustering of density extrema. Recently, \citet{b7} and \citet[][hereafter DS]{b1} showed that the correlation of density peaks in real and redshift space can be interpreted in terms of a simple Lagrangian biasing scheme. Due to the peak constraint, the effective Lagrangian peak density at a given point not only depends on the local value of the mass density but also on its Laplacian. The first-order peak bias depends on the mass and height of the peaks, and on the matter power spectrum. For high peaks, this reduces to the results by \citet{b24}. Moreover, although peaks move with the matter at their positions, DS infer the existence of a statistical velocity bias due to the fact that local maxima can only exist at special locations. This also leads to a bias between the linear velocity spectra of peaks and matter which is predictable in quantitative terms. In spite of the fact that the DS model provides the initial conditions for sophisticated models of the (Eulerian) halo distribution where the LEM is based either on resummed perturbation theory \citep{b48} or on the Zel'dovich approximation \citep{b47}, its predictions for the Lagrangian clustering and velocities of the regions that will form collapsed structures have never been thoroughly tested against numerical simulations. This paper provides such a test, which is necessary if we are to use advanced bias models to extract useful information on the cosmological parameters through a comparison with observations. The structure of the paper is as follows. In Section \ref{mod} we review the bias model first presented in DS. The details of our N-body simulations and analysis techniques are outlined in Section \ref{numerics}, and a comparison between the model and numerical results is presented in Section \ref{tre}. Finally, we summarize our main results in Section \ref{sum}. \section{The DS model} \label{mod} In this section we summarize, for completeness, the peaks model described in Section 2 of DS. In order to do so, we first introduce and define some quantities relevant for peak statistics. The spectral moments of the matter power spectrum are defined as \begin{equation} \sigma_{\rm n}^2 (R_{\rm s},z) = \frac{1}{2 \pi^2} \int_0^{\infty} \mathrm{d}k\,k^{2(n+1)}\,P(k,z)\,W(k,R_{\rm s})^2\,, \label{a} \end{equation} where $P(k,z)$ is the linear matter power spectrum at redshift $z$, and $W(k,R_{\rm s})$ is a smoothing kernel of characteristic length $R_{\rm s}$. In terms of the moments we define the spectral parameters: \begin{equation} \gamma_{\rm n} \equiv \frac{\sigma_{\rm n}^2}{\sigma_{\mathrm{n-1}}\sigma_{\mathrm{n+1}}}\,. \label{b} \end{equation} There are two common choices for $W(k,R_{\rm s})$: the Gaussian filter \begin{equation} W_{\rm G}(k,R_{\rm s}) = \mathrm{e}^{-\frac{k^2 R_{\rm s}^2}{2}}\,, \label{c} \end{equation} and the top-hat filter \begin{equation} W_{\rm TH}(k,R_{\rm s}) = \frac{3 [\sin(k R_{\rm s})-k R_{\rm s} \cos(k R_{\rm s})]}{k^3 R_{\rm s}^3}\,. \label{d} \end{equation} The mass contained within the comoving length $R_{\rm s}$ in these cases is \begin{equation} M_{\rm G}(R_{\rm s}) = (2\pi)^{3/2} \bar{\rho} R_{\rm s}^3 \textrm{, and } \qquad M_{\rm TH}(R_{\rm s}) = \frac{4\pi}{3} \bar{\rho} R_{\rm s}^3\,, \label{e} \end{equation} where $\bar{\rho}$ is the mean matter density of the Universe. We will often characterize peaks in terms of their dimensionless peak height, $\nu$, defined \begin{equation} \nu(R_{\rm s},z_{\rm c}) = \frac{\delta_{p}}{\sigma_0(R_{\rm s},z_{\rm c})}\,. \label{h} \end{equation} Here $\delta_{\rm p}$ is the smoothed overdensity at the peak location linearly extrapolated to $z=0$, and $\sigma_0(R_{\rm s},z_{\rm c})$ is the linear rms mass fluctuation in spheres of radius $R_{\rm s}$. DS linked density maxima of height $\nu(R_{\rm s},z)$ to dark matter haloes of mass $M_{\rm s}$ collapsing at redshift $z_{\rm c}$, assuming that $\delta_{\rm p}$ coincides with the threshold for collapse, $\delta_{\rm c}$. \citet{b2} and DS computed the cross-correlation between peaks and the underlying density field which also corresponds to the average density profile around density maxima. Similarly, \citet{b7} evaluated the leading order expressions (on large spatial separations) for the peak auto-correlation function and the line-of-sight mean streaming for pairs of discrete local maxima of height $\nu$. \citet{b7} and DS showed that the full expression for the cross-correlation and the large-scale asymptotic of the auto-correlation function are consistent with -- and thus can be thought of as arising from -- an effective bias relation. In the DS model, the number density and the velocity of peaks, $\delta n_{\rm pk}$ and ${\bf v}_{\rm pk}$, are related to the dark matter density contrast and velocity fields, linearly extrapolated to $z=0$, via: \begin{equation} \delta n_{\rm pk} ({\bf x}|z_{\rm c}) = b_{\nu}\delta_{\rm S}({\bf x})-b_{\zeta} \nabla^2\delta_{\rm S}({\bf x}) , \label{ia} \end{equation} and \begin{equation} {\bf v}_{\rm pk} ({\bf x}|z_{\rm c}) = {\bf v}_{\rm S} ({\bf x}) - \frac{\sigma_0^2}{\sigma_1^2} \nabla \delta_{\rm S}({\bf x}) \,. \label{ib} \end{equation} The subscript ``S'' indicates that the fields are smoothed on the scale $R_{\rm s}$, and the bias parameters, $b_{\nu}$ and $b_{\zeta}$, are given by \begin{equation} b_{\nu} = \frac{1}{\sigma_0} \left(\frac{\nu-\gamma_1 \bar{u}}{1-\gamma_1^2} \right)\,, \label{ja} \end{equation} and \begin{equation} b_{\zeta} = \frac{1}{\sigma_2} \left(\frac{\bar{u}-\gamma_1 \nu}{1-\gamma_1^2} \right)\,. \label{jb} \end{equation} Here $\bar{u}$ is the mean curvature of the peaks, which can be approximated by \citep{b2} \begin{equation} \bar{u}=\gamma_1 \nu + \frac{3(1-\gamma_1^2)+(1.216-0.9\gamma_1^4) \exp \left[-\frac{\gamma_1}{2} \left(\frac{\gamma_1 \nu}{2} \right)^2 \right] }{\left[3(1-\gamma_1^2)+0.45+\left(\frac{\gamma_1 \nu}{2} \right)^2 \right]^\frac{1}{2} + \left(\frac{\gamma_1 \nu}{2} \right) } \,. \label{k} \end{equation} Note that $b_{\nu}$ coincides with the peak bias factor found by \citet{b2} after neglecting the derivatives of the density correlation function. Since, by definition, the gradient of the density field vanishes at peak locations, eq. (\ref{ib}) suggests that the peak and dark matter velocity fields must be coincident there. By construction, {\em peaks move with the dark matter flow}, yet the spatial bias induces a statistical velocity bias. We will consider the scaled velocity divergence, $\theta({\bf{x}})=\nabla \times {\bf v}({\bf{x}})/(aHf)$, rather than the velocity field. Here $a$ is the scale factor, $H$ the Hubble parameter, $f=d\ln D/d\ln a$, with $D$ the linear growth factor. In these units, both $\theta({\bf{x}})$ and $\delta({\bf{x}})$ are dimensionless quantities. With these changes, eqs. (\ref{ia}) and (\ref{ib}) can be rewritten in Fourier space as \begin{equation} \delta n_{\rm pk} ({\bf k}) = (b_{\nu} +b_{\zeta} k^2)\, \delta({\bf k})\, W(k,R_{\rm s}) \label{la} \end{equation} and \begin{equation} \theta_{\rm pk} ({\bf k}) = \left(1 - b_{\sigma} k^2 \right) \theta ({\bf k})\, W(k,R_{\rm s}) = b_{\rm vel}(k)\theta ({\bf k})\,, \label{lb} \end{equation} where we have defined \begin{equation} b_{\sigma} = \frac{\sigma_0^2}{\sigma_1^2}\,. \label{m} \end{equation} In the limit of high peaks ($\nu \gg 1$) it can be shown that the bias parameters obey the following asymptotic relations: $b_{\nu} \to \nu/\sigma_0$ and $b_{\zeta} \to 0$. This implies that the highest peaks are linearly biased tracers of the underlying matter field. This is consistent with the predictions of the peak-background split \citep{b26} and DS showed that, indeed, $b_{\nu}$ is the appropriate generalization of the constant, large-scale bias for low $\nu$. Unlike the density bias factors, $b_{\sigma}$ does not depend on $\nu$. In order to test this model against N-body simulations, we will make use of the cross-spectra between the peak and dark matter densities and velocities (denoted as $P_{\mathrm{mp}}$ and $\mathcal{P}_{\mathrm{mp}}$, respectively) and of the corresponding peak auto-spectra ($P_{\mathrm{p}}$ and $\mathcal{P}_{\mathrm{p}}$). From eqs. (\ref{la}) and (\ref{lb}) we obtain \begin{eqnarray} P_{\mathrm{mp}}(k) &=& (b_{\nu} +b_{\zeta} k^2)\,P(k)\,W(k,R_{\rm s}) \,, \nonumber \\ P_{\mathrm{p}}(k) &\simeq& (b_{\nu} +b_{\zeta} k^2)^2\,P(k)\,W^2(k,R_{\rm s}) \,, \label{n} \end{eqnarray} \begin{eqnarray} \mathcal{P}_{\mathrm{mp}}(k) &\simeq& (1-b_{\sigma} k^2)\,\mathcal{P}(k)\,W(k,R_{\rm s}) \,, \nonumber \\ \mathcal{P}_{\mathrm{p}}(k) &\simeq& (1-b_{\sigma} k^2)^2\,\mathcal{P}(k)\,W^2(k,R_{\rm s}) \,, \label{nb} \end{eqnarray} where $P(k)$ and $\mathcal{P}(k)$ are the matter density and velocity divergence auto-spectra, respectively. We remind the reader that the expression for $P_{\rm p}(k)$ is only valid to first order in $P(k)$ as $k \to 0$, and that higher order corrections should be included to improve its accuracy (see e.g. \citealt{b47}). On the contrary, the expression for the cross-spectrum $P_{\rm mp}$ is exact, as shown in \citet{b2} and in the Appendix A of DS. Eq. (\ref{n}) has the same functional form as eq. (57) in \citet{b24} who studied the clustering of density extrema. The DS coefficients $b_{\nu}$ and $b_{\zeta}$ match those in \citet{b24} only in the limit $\nu \gg 1$, for which nearly all extrema are density maxima. \section{Numerical Issues} \label{numerics} In this section we provide a brief description of the main numerical issues relevant for this work. This includes a brief description of our numerical simulations in Section \ref{num}, our main analysis techniques in Section \ref{analysis}, and a characterization of halo collapse barriers in Section \ref{hei}. \subsection{N-body simulations} \label{num} Our analysis focuses on two high-resolution N-body simulations of structure formation in the standard LCDM cosmology. The cosmological parameters for our runs were chosen to be consistent with the fifth-year WMAP data release \citep{b4}. These are $h=0.701$, $\sigma_8=0.817$, $n_{\rm s}=0.96$, $\Omega_{\rm m}=0.279$, $\Omega_{\rm b}=0.0462$ and $\Omega_{\Lambda}=1-\Omega_{\rm m}=0.721$. Each simulation was run with a lean version of the Tree-PM code \textsc{Gadget}-2 \citep{b3}, and followed the dark matter using $1024^3$ collisionless particles. One simulation had a box side-length of $L_{\rm box}=1200\, h^{-1}$ Mpc and a particle mass of $m_{\rm part}=1.246\cdot 10^{11} h^{-1} M_{\odot}$; the other used $L_{\rm box}=150\, h^{-1}$ Mpc and had $m_{\rm part}=2.433\cdot 10^{8} h^{-1} M_{\odot}$. The initial redshifts of the simulations were $z_{\rm in}=50$ and $z_{\rm in}=70$ for the larger and smaller box, respectively. Using these simulations we are able to probe a wide range of halo masses, spanning $8\cdot 10^{10} h^{-1} M_{\odot} < M_{\rm h} < 10^{14} h^{-1} M_{\odot} $. These simulations were first studied in \citet{b6}, and later by \citet{b8}, and we refer the reader to those papers for further details. Haloes were identified at $z=0$ using a friends-of-friends (FOF) algorithm \citep{b25} with a linking length of 0.2 times the mean interparticle distance. Protohaloes were identified by tracing backward to the initial redshift of all subsets of the particles belonging to $z=0$ FOF haloes. We use the centre of mass of each protohalo as a proxy for its spatial location; the mass-weighted linear velocity provides an estimate of the protohalo's motion. As a test of the sensitivity of our results to the adopted halo finder, we also generated a spherical overdensity (SO) halo catalogue with an overdensity threshold of 200 times the critical density, $\rho_{\rm c}$. For a fixed halo mass, the two halo-finders produce results consistent within $10$ per cent (in terms of all the bias coefficients) and so, in what follows, we will focus on results obtained for the FOF haloes, and only consider those containing at least 100 particles. Haloes in each simulation are split into four separate mass bins in order to preserve their peculiar clustering properties. These bins are referred to as 1S to 4S for the small box, and 1L to 4L for the large one. To asses the impact of shot noise in the analysis of our small-box simulation we consider an additional mass bin, labeled bin0S, which includes all haloes with $N\geq 100$. The mass ranges and total number of haloes in each bin are given in Table \ref{tab:1}. We note that bins with label `S' refer to masses $M<M_{*}$ for which dark-matter haloes are not expected to be in one-to-one correspondence with linear peaks \citep{b8}. \subsection{Analysis} \label{analysis} We construct protohalo density and momentum fields using cloud-in-cell grid assignment on a $512^3$ mesh. Velocity fields are obtained by taking the ratio of the momentum and density fields, as described in \citet{b5}. In the case of haloes, these distributions are smoothed to preclude the existence of empty cells; the smoothing scales used are $R_{\rm f} = 7 h^{-1}$ Mpc for the large box and $R_{\rm f} = 1.8 h^{-1}$ Mpc for the small one. These values are chosen in order to mask the effects of the grid, but we have explicitly verified that our results are not significantly affected by them. All power spectra have been computed using a fast Fourier transform technique. The discreteness of dark-matter particles and haloes gives rise to a shot-noise component in the spectra. For the density fields, the estimated power spectrum, $\hat{P}$, includes a shot-noise term which is inversely proportional to the number density of objects $\bar{n}$ (assuming Poisson sampling): \begin{equation} \hat{P}=P_{\mathrm{true}} + \frac{1}{\bar{n}}\,. \label{oo} \end{equation} Shot noise is therefore negligible for the matter spectra but may be significant for that of the haloes. The issue is more severe in Lagrangian space, because fluctuations in the initial conditions are small. We will consider two alternative estimates of the protohalo bias; one is determined from the shot-noise corrected auto-spectrum, \begin{equation} b(k) \equiv \sqrt{\frac{P_{\mathrm{h}}(k)}{P(k)}}\,, \label{p} \end{equation} and the other from the cross-spectrum, \begin{equation} b_{\mathrm{eff}}(k) \equiv \frac{P_{\mathrm{mh}}(k)}{P(k)}\,. \label{q} \end{equation} Here the subscript ``h'' indicates the halo fields, and ``m'' the matter field (the analogous fields for peaks are indicated with the subscript ``p''.) The relation between the two is \begin{equation} b_{\mathrm{eff}}(k) = b(k) \cdot r(k)\,, \label{r} \end{equation} where $r$ is the linear correlation coefficient, defined as \begin{equation} r(k)=\frac{P_{\mathrm{mh}}(k)}{\sqrt{P(k) P_{\mathrm{h}}(k)}}\,. \label{rr} \end{equation} These relations tell us that the two definitions of the bias are equivalent \emph{only if} the bias is purely deterministic in Fourier space, i.e. $r=~1$. At leading order, in the model presented by DS, $b_{\mathrm{eff}}(k)= b_{\nu}+b_{\zeta}\,k^2$ and $b(k)=|b_{\mathrm{eff}}(k)|$ (neglecting the filter function). Any stochasticity (represented by the higher order corrections in $P_{\rm p}$) will degrade the correlation, yielding different estimates for the bias. Because of this, we will consider both the cross- and auto-spectra to check for a potential stochastic element of the bias. \footnote{Although the effective first-order bias model by DS is deterministic in Fourier space, it is stochastic in configuration space (\citealt{b24}, \citealt{b1}).} As for the density, we can define two estimates for the velocity bias, that we denote $b_{\theta}$ and $b_{\theta\mathrm{eff}}$, with a correlation coefficient $r_{\theta}$. There is no conclusive way to subtract the shot noise for the velocity divergence spectrum. Hence we used our simulations to gain some insight into this issue. We introduced an artificial shot noise in the matter spectrum $\mathcal{P}$ by randomly drawing a fraction of the particles and found that $\mathcal{P}_{\rm shot}(k) \simeq A \cdot k^2$ asymptotically for large $k$. Therefore we fitted the amplitude factor $A$ to obtain shot-noise corrected power spectra. Note however that other terms could be important at smaller wavenumbers; in this work we only consider data for which $\mathcal{P}_{\rm shot}(k) < 0.1 \hat{\mathcal{P}}$. \\ The DS model describes the biasing of linear density peaks that are expected to form haloes of a given mass at a specified redshift according to some collapse model (which determines the value for $\delta_{\rm c}$). However, we analyse the density and velocity fields for the actual FOF and SO protohaloes. These are the quantities of physical interest for studying galaxy clustering in terms of halo occupation models (e.g. \citealt{b49}). \citet{b8} showed that the vast majority of haloes in our N-body simulations can be unambiguously associated with linear peaks in the initial conditions when smoothed on the mass scale of the halo. For example, $\lower.5ex\hbox{\gtsima} 70$ per cent of all haloes can be matched with similar-mass peaks in $\delta_{\rm S}(\mathbf{x})$, and $\lower.5ex\hbox{\gtsima} 90$ per cent for haloes with $M \lower.5ex\hbox{\gtsima} 10^{14} h^{-1} M_{\odot}$. Note, however, that the correspondence between the DS peaks and protohaloes will not be perfect.\footnote{The measured abundance of haloes in the simulations is of the same order of magnitude as (albeit a bit higher than) the number of peaks with $\nu > 1.68/\sigma_0$ computed as in \citet{b2}.} On a given mass scale, some peaks with overdensities at the collapse threshold will evolve into substructures contained within larger virialized haloes (the so-called cloud-in-cloud problem), or to haloes of significantly different mass. On top of this, numerical simulations have shown that the collapse threshold $\delta_{\rm c}$ for haloes of a given mass and redshift has a broad probabilistic distribution rather than a fixed value (\citealt{b50}; \citealt{b10}) and possibly also depends on the form of the adopted smoothing kernel. We consider some of these issues in the following section. \begin{figure} \includegraphics[width = 3.0in,keepaspectratio=true]{fig1} \caption{ Mass-dependence of the linear overdensities measured at protohalo centres in the initial conditions of our simulations. The shaded regions and contours show the full halo distribution after smoothing with a Gaussian filter; connected circles highlight the median trend. All values of $\delta_h$ have been linearly extrapolated to $z=0$. Open and filled points correspond to our $150\, h^{-1}$ Mpc and $1200\, h^{-1}$ Mpc boxes, respectively. For comparison, we also show, using boxes, the median trends obtained after smoothing with a top-hat kernel. In both cases, the median trends are well described by eq. (\ref{o}): the dot-dashed curve shows the SMT result (note that this is not a fit to our simulation data), and the dashed curve shows the result for slightly different values of the free parameters: $\delta_{*}=1.15$, $\alpha=0.37$ and $\beta=0.515$. The top panel shows the mass dependence of the scatter about the median trend for the Gaussian-filtered case, which is well approximated by a simple power-law, $\Sigma = 0.4 \ \sigma_0^{6/5}$.} \label{fig1} \end{figure} \subsection{Barrier heights for top-hat and Gaussian filters} \label{hei} The peak model described in Section ~\ref{mod} requires well-defined spectral moments. However, due to its sharp boundary in real space, the top-hat filter decays very slowly in Fourier space so that the integral defining $\sigma_2^2$ is divergent in a LCDM model. Since eqs. (\ref{ja}), (\ref{jb}) and (\ref{k}) depend on $\sigma_2$, DS instead adopted a Gaussian filter and assumed $\delta_{\rm c}=\delta_{\mathrm{sc}}=1.68$. Here, $\delta_{\rm c}$ corresponds to the critical density for the collapse of a spherical top-hat perturbation in an otherwise unperturbed EdS universe, and this does not necessarily apply to peaks in a smoothed Gaussian random field. Another issue is that the validity of the simple spherical collapse model is questionable, at best; the probability of a protohalo or a peak being spherical is null, since it would require the three eigenvalues of the tidal tensor being equal. \citet{b9} showed that the barrier height in the more general ellipsoidal collapse model \citep{b14} can be approximated by \begin{equation} \delta_{\mathrm{ec}}(M,z_{\rm c})=\delta_*\, \left\{ 1+ \alpha \cdot \left[\frac{\sigma_0^2(R_{\rm s}(M),z_{\rm c})}{\delta_*^2} \right]^\beta \right\}\,, \label{o} \end{equation} where $\delta_*=\delta_{\rm sc}$ is taken from the spherical collapse model, and $\alpha=0.47$ and $\beta=0.615$ are determined from fits to the model results. The presence of the dispersion, $\sigma_0$, in eq. ~(\ref{o}) results in a mass-dependent barrier height: lower mass haloes require, on average, higher overdensities for collapse since they must hold themselves together against larger tidal forces. It should be emphasized that eq. (\ref{o}) describes the {\em mean} barrier height; the scatter about the mean can be approximated by $\Sigma=0.3 \ \sigma_0$ \citep{b10}. These values are valid only for the top-hat filter. What is the appropriate barrier height corresponding to a Gaussian filter? In Figure \ref{fig1} we show the linear overdensities measured at protohalo centres of mass after smoothing with a Gaussian filter on the halo mass scale. The shaded regions show the halo data, and connected circles the medians of the distribution. Open points correspond to results from our $150\, h^{-1}$ Mpc box simulation, and solid points to our $1200\, h^{-1}$ Mpc box run. Squares show the median $\delta_{\rm c}(M)$ for the same sample of haloes, but after smoothing the linear density field with a top-hat filter instead (note the good agreement with the result of \citet{b9}, shown as a dot-dashed line in Figure \ref{fig1}). All values have been linearly extrapolated to $z=0$. \begin{table*} \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|} \hline Bin & Mass range & $\#$ haloes & $\bar{M}$ & \multicolumn{2}{|c|}{$b_{\nu}$} & \multicolumn{2}{|c|}{$b_{\zeta}$} & $b_{\sigma}$ & $R_{\rm s}$ \\ \hline & ($10^{12} h^{-1}\,M_{\odot}$) & & ($10^{12} h^{-1}\,M_{\odot}$) & \multicolumn{2}{|c|}{} & \multicolumn{2}{|c|}{($h^{-2}\,\mathrm{Mpc}^2$)} & ($h^{-2}\,\mathrm{Mpc}^2$) & ($h^{-1}$ Mpc) \\ \hline & & & & SG & EG & SG & EG & & \\ \hline 0S & $0.08-0.8$ & $106746$ & $0.2986$ & -0.31 & -0.23 & 0.63 & 0.61 & 0.92 & 0.62\\ 1S & $0.08-0.1$ & $21990$ & $0.08961$ & -0.30 & -0.21 & 0.26 & 0.25 & 0.45 & 0.42\\ 2S & $0.1-0.15$ & $30610$ & $0.1235$ & -0.30 & -0.22 & 0.33 & 0.32 & 0.55 & 0.46\\ 3S & $0.15-0.3$ & $32322$ & $0.2165$ & -0.31 & -0.23 & 0.49 & 0.48 & 0.76 & 0.56\\ 4S & $0.3-0.8$ & $21824$ & $0.4934$ & -0.31 & -0.24 & 0.92 & 0.90 & 1.24 & 0.74\\ \hline 1L & $12.46-16$ & $146839$ & $14.15$ & 0.05 & -0.08 & 11.3 & 11.5 & 8.75 & 2.26\\ 2L & $16-25$ & $182006$ & $20.25$ & 0.18 & -0.003 & 14.7 & 15.2 & 10.8 & 2.54\\ 3L & $25-40$ & $121146$ & $32.05$ & 0.40 & 0.12 & 20.5 & 21.5 & 14.0 & 2.96\\ 4L & $40-100$ & $114367$ & $65.99$ & 0.92 & 0.44 & 34.5 & 37.1 & 21.2 & 3.77\\ \hline \end{tabular} \caption{Mass range, number of haloes, average mass and bias parameters (both weighted by halo counts) for the nine different mass bins used in our study. The first five belong to the small-box simulation; the last four to the large-box simulation. The bias parameters are computed according to the two models for the peak height described in the text: the spherical and ellipsoidal collapse models. All values have been obtained adopting a Gaussian filter.} \label{tab:1} \end{table*} \begin{figure*} \includegraphics[width = 3.0in,keepaspectratio=true]{dens_l} \includegraphics[width = 3.0in,keepaspectratio=true]{only} \caption{ \textit{Left}: Stars and open circles plot, respectively, $b(k)$ and $b_{\rm eff}(k)$ for the four separate mass bins in our large simulation box. Dotted lines show the predictions obtained for the spherical collapse model, and dashed lines those of the ellipsoidal collapse model adopting a Gaussian filter. Solid lines are the best-fits to the data points. \textit{Right}: Stars and open circles correspond to $b(k)$ and $|b_{\rm eff}(k)|$, respectively. As in the left panels, dotted and dashed line show to the SG and EG models, and solid lines are the best-fits to the data. Thick lines indicate negative values of the bias. In all cases, errors are propagated assuming that the uncertainty in the power-spectrum is $\sigma(k)=\sqrt{2/N(k)}\cdot P(k)$, where $N(k)$ is the number of modes in each $k$ bin.} \label{fig4} \end{figure*} It is clear that adopting a Gaussian filter entails lower values of $\delta_{\rm c}$ at all masses\footnote{At any mass scale, $M$, the smoothing length of a Gaussian filter exceeds that of a top-hat filter by a factor of $(3\sqrt{\pi /2})^{1/3}\approx 1.55$.}. Nonetheless, our results are still accurately described by eq. (\ref{o}), albeit with slightly different values for the model parameters: $\delta_{*}=1.15$, $\alpha=0.37$ and $\beta=0.515$. We show this explicitly in Figure \ref{fig1} using a dashed line. The one-sigma dispersion about the mean trend is well described by a simple power-law, $\Sigma = 0.4 \ \sigma_0^{6/5}$, as seen in the upper panel of the plot. \section{Results} \label{tre} In this section we test the predictions of the DS model against the measured density and velocity bias of dark matter protohaloes. In order to do so, we calculate the (model-predicted) bias parameters $b_{\nu}$, $b_{\zeta}$ and $b_{\sigma}$ for each mass bin in both of our simulations. Since the values of $b_{\nu}$ and $b_{\zeta}$ depend on peak height (and hence smoothing scale) we adopt two models for the collapse barrier: one assumes the spherical collapse model (hereafter SG) and the other the ellipsoidal collapse model (hereafter EG). A Gaussian filter is used in both cases. Table 1 lists the bias parameters predicted by the DS model for these choices of collapse barrier. \begin{table*} \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline Bin & \multicolumn{4}{|c|}{$b_{\nu}$} & \multicolumn{4}{|c|}{$b_{\zeta}$ ($h^{-2}\,\mathrm{Mpc}^2$)} \\ \hline &\multicolumn{2}{|c|}{$P_{\rm mh}$} & \multicolumn{2}{|c|}{$P_{\rm h}$} & \multicolumn{2}{|c|}{$P_{\rm mh}$} & \multicolumn{2}{|c|}{$P_{\rm h}$}\\ \hline &FOF &SO &FOF &SO &FOF &SO &FOF &SO \\ \hline 0S & -0.25 & -0.24 & -0.31 & -0.30 & 0.86 & 0.81 & 0.14 & 0.11\\ 1L & 0.19 & 0.22 & 0.27 & 0.37 & 8.8 & 7.9 & 12.6 & 12.0\\ 2L & 0.29 & 0.31 & 0.40 & 0.39 & 10.3 & 9.3 & 9.5 & 8.8\\ 3L & 0.43 & 0.47 & 0.52 & 0.54 & 13.9 & 13.1 & 6.7 & 5.9\\ 4L & 0.73 & 0.75 & 0.73 & 0.73 & 21.8 & 20.9 & 6.4 & 5.4\\ \hline \end{tabular} \caption{Best-fit values for the density bias parameters from the cross-spectra and the auto-spectra. Uncertainties are always at the few per cent level. Haloes are identified either with the FOF or the SO algorithms.} \label{tab:2} \end{table*} \begin{figure*} \includegraphics[width = 3.0in,keepaspectratio=true]{b1} \includegraphics[width = 3.0in,keepaspectratio=true]{b2} \caption{ Density bias parameters, $b_{\nu}$ (\textit{left}) and $b_{\zeta}$ (\textit{right}), plotted as a function of halo mass. Orange circles with error bars correspond to the best-fit values obtained from the cross-spectra of haloes identified with the FOF algorithm. The corresponding values for SO haloes are shown with purple stars. The solid and dashed lines show, respectively, the predictions of the DS model for the spherical collapse model, and for the ellipsoidal model with a Gaussian filter. The vertical dashed black lines mark $M_*$.} \label{fig5} \end{figure*} \subsection{Density spectra} In Figure \ref{fig4} we plot $b(k)$ and $b_{\mathrm{eff}}(k)$ extracted from our simulations,\footnote{The values directly measured from the simulations are rescaled by the growth factor $D(z_{\rm in})$ to match the theoretical estimates, where $\delta$ is linearly extrapolated to $z=0$. Note that the actual Lagrangian bias is a factor $D^{-1}\simeq 40-50$ larger than the values reported in the figure.} along with the model prediction (eq. (\ref{n})). Although the model is not expected to work for very small scales, we nonetheless show the results for each box up to $k\simeq 1/R_{\rm f}$ (we remind that $R_{\rm f}$ is the smoothing scale needed for the velocity field). Overall, the model expression for the initial peak bias is able to describe the simulation results reasonably well. This can be seen from the solid lines in Figure \ref{fig4}, where $b_{\nu}$ and $b_{\zeta}$ have been treated as free parameters and fitted to both the cross- \emph{and} the auto spectra. However, it is clear that the values for the fit parameters are different in the two cases (see Table \ref{tab:2}), implying $b(k)\neq |b_{\mathrm{eff}}(k)|$ or, equivalently, $r(k) \neq 1$. In particular, for the range of masses in Bin 0S, $b_{\mathrm{eff}}$ is negative, i.e. $r<0$. The predictions from the SG and EG barriers, as apparent from the dashed and dotted lines in the right panel in Figure \ref{fig4}, are also negative. Hence the bias model matches more closely $b_{\mathrm{eff}}$ rather than $b$. This is completely expected, because eq. (\ref{n}) is exact only for the cross-correlation between peaks and matter while it neglects higher-order corrections for the peak autocorrelation. However, neither the SG nor the EG barrier provide the appropriate values for the coefficients. In particular, the EG barrier performs better for low masses (bin 0S), while the situation is reversed for the higher mass bins. Figure \ref{fig4} suggests that the stochasticity is more of a problem for haloes with $M<M_*$, and on smaller scales. Since high-mass haloes are highly correlated with density peaks in the initial conditions \citep{b8} and peaks follow eq. (\ref{la}), the relationship between $\delta_{\rm h} ({\bf k})$ and $\delta({\bf k})$ is likely to be more deterministic for massive haloes. Note, however, that the estimate of $r$ is affected by the shot-noise correction, which is large in our samples. Therefore we cannot draw definitive conclusions regarding stochasticity. \begin{figure*} \includegraphics[width = 3.0in,keepaspectratio=true]{vel_l2} \includegraphics[width = 3.0in,keepaspectratio=true]{vel_s2} \caption{ Stars and circles show $b_{\theta}(k)$ and $b_{\theta \mathrm{eff}}(k)$, respectively, for haloes in our large (\textit{left}) and small (\textit{right}) simulation boxes. The solid line is the DS model prediction (eq. (\ref{lb}) with $R_s$ given in Table \ref{tab:1}), which is independent of peak height. Errors are propagated as in Fig. \ref{fig4}. Note that we only show $b_{\theta}(k)$ on scales over which the shot-noise is sub-dominant, as stated in Section \ref{analysis}.} \label{fig6} \end{figure*} The performance of the model is characterized in another way in Figure \ref{fig5}, where we compare directly the mass dependence of the bias parameters $b_{\nu} (M)$ and $b_{\zeta} (M)$ predicted by the SG and EG models to the best-fit values obtained using the parameterization of the cross-spectrum in eq. (\ref{n}). We do not show the corresponding results from the auto-spectrum because of the uncertain shot-noise subtraction. Figure \ref{fig5} reveals that overall the $b_{\nu}$ values derived from the SG model are closer to the fitted ones. In contrast, there is almost no difference between the two models for $b_{\zeta}$. For both parameters the mass dependence follows a different trend when we compare the fits to the models. In particular, the prediction for $b_{\zeta}$ and $M<M_*$ becomes more and more inaccurate with decreasing halo mass (up to a factor of $\sim 3$ for $M\simeq M_*/15$). On the other hand, in the mass range for which the model was developed ($M>M_*$), $b_{\zeta}$ is overpredicted by a factor of $\sim 1.5$, independent of halo mass. Figure \ref{fig4} shows that the disagreement is even worse when $b_{\nu}$ and $b_{\zeta}$ are obtained from $P_{\rm h}$. All of this demonstrates that eqs. (\ref{ja}) and (\ref{jb}) cannot accurately describe the mass dependence of the Lagrangian density bias, especially for $M<M_*$. We will discuss possible reasons for this in Section \ref{sum}, but first turn our attention to the velocity spectra. \subsection{Velocity spectra} Figure \ref{fig6} shows $b_{\theta}(k)$ and $b_{\theta\mathrm{eff}}(k)$ for bins 1S-4S and 1L-4L; the theoretical model, $b_{\rm vel}(k)$, is also plotted. The data have been ``de-smoothed", i.e. they have been divided by a filter function with smoothing scale $R_{\rm f}$, the same used to smooth the velocity field in the first place. It is apparent that on large scales the two estimates of the bias coincide, i.e. $r_{\theta}\simeq1$, indicating a strong correlation between the fields. Overall, we can conclude that the velocity bias is deterministic to a good approximation. There is an excellent agreement (better than $10$ per cent) between the simulation results and the model predictions for all mass bins at scales $k<0.1 \ h\, \mathrm{Mpc}^{-1}$ for the large box and $k<0.3 \ h\, \mathrm{Mpc}^{-1}$ for the small one. Therefore, the peak model provides a faithful description of the velocity bias. \section{Summary} \label{sum} We have investigated the Lagrangian bias of dark-matter haloes by testing the theoretical model proposed by DS against N-body simulations of structure formation. The model assumes that haloes form from density peaks and predicts a scale-dependent bias for both the density and velocity fields. Our main results can be summarized as follows. \begin{itemize} \item When averaged over a spherical Lagrangian volume containing the appropriate mass, the linear density contrast measured at protohalo centres depends sensitively on the choice of smoothing kernel. For a Gaussian kernel, for example, the resulting density contrasts are systematically lower than those computed using a top-hat filter. This is because, at fixed mass, the smoothing length of a Gaussian kernel exceeds that of a top-hat filter by a factor of about $1.55$ resulting in systematically lower density estimates. Nonetheless, the median barrier height computed with a Gaussian kernel can be accurately parameterized by the same fitting formula first advocated by \citet{b9} for the case of a top-hat filter, albeit with different values for the numerical parameters. We use this result to approximate the collapse threshold for dark matter halo formation when adopting a Gaussian filter. \item The functional forms for the density and velocity bias relations derived by DS - our eqs. (\ref{n}) and (\ref{nb}) - accurately describe the results obtained from our simulations, provided the parameters of the model are allowed to vary with respect to the model-predicted values. In both cases, the Lagrangian bias is characterized by a constant term that dominates on very large scales, and a scale-dependent term proportional to $k^2$. \item Quantitatively, the velocity bias predicted by the DS model is able to reproduce the measured protohalo velocity bias in our simulations to better than $10$ per cent, provided we limit ourselves to quasi-linear scales ($k\leq 0.3 \ h\, \mathrm{Mpc}^{-1}$). The predicted density bias (eqs. (\ref{ja}) and (\ref{jb})), on the other hand, does not match the Lagrangian density bias extracted from the simulations. This is likely due to the more complex nature of the density bias, which depends additionally on peak height and on the exact definition of a protohalo. These results are independent of whether one adopts a barrier height consistent with either the spherical or ellipsoidal collapse model. \item We have measured the mass dependence of the density bias, $b_{\nu}$ and $b_{\zeta}$, by fitting the density power-spectra obtained for haloes in several different mass bins in each of our two simulations. The most massive haloes identified in our simulations are the most strongly biased, and have characteristic overdensities that correlate with the underlying matter density. In contrast, the distribution for low mass haloes ($M_{\rm h} < 10^{12} h^{-1}\,M_{\odot}$) exhibits an anti-correlation. A similar trend is also predicted by the DS model, although with noticeable differences (see Figure \ref{fig5}). We emphasize that the best-fit values for $b_{\nu}$ and $b_{\zeta}$ obtained in our analysis apply for a LCDM cosmogony (with our adopted cosmological parameters) and only over the limited mass range probed by our simulations. Future work should consider how the Lagrangian density bias depends on the underlying cosmology. \item Our comparison of the auto- and cross-spectra in Section \ref{tre} suggests the presence of a stochasticity in the Fourier-space density bias of dark matter protohaloes, but none for the velocity bias. This is in disagreement with the DS model for the bias at leading order, and seems to corroborate the need for higher-order terms in the expression for the auto-spectrum of the protohaloes, as predicted in \citet{b47}. However, we cannot draw firm conclusions regarding stochasticity in the bias estimates due to uncertainties in the shot-noise subtraction. \end{itemize} We have tested the Lagrangian bias model of DS against a pair of high resolution N-body simulations of structure formation and found that it is able to accurately reproduce the velocity bias, but not that of the density. Our analysis focused on protohaloes, the high redshift progenitors of $z=0$ dark-matter haloes, whereas the model describes the biasing of density peaks. One possible explanation for the differences in the model predictions and simulation results comes form the differences between the expressions for the bias parameters: $b_{\nu}$, $b_{\zeta}$ and $b_{\sigma}$. The latter of the three is solely determined by the linear matter power spectrum and the smoothing scale corresponding to a given mass, $M$. The density bias parameters, however, depend additionally on explicit properties of the peaks, such as their height, $\nu$, mean curvature, $\bar{u}$, and on an assumed collapse threshold for their identification as haloes of a given mass. Figure \ref{fig1} shows that there is a large halo-to-halo variation in $\nu$ at any given mass scale; characterizing the collapse threshold as a single mass-dependent value may, therefore, be too simplistic. This added complexity introduces a significant margin for error in the model's estimates of the density bias. It remains to be seen whether more realistic models of the collapse barrier - such as those that account for the statistical scatter in the linear over-densities of protohaloes at a given mass - will improve the model's predictive power. Another possibility for the discrepancies stems from the fact that we are identifying linear density peaks with protohaloes in our simulation initial conditions. Although the majority of our dark matter protohaloes form in the vicinity of linear density peaks of the same characteristic mass, the fate of {\em all} peaks with the same mass and overdensity is unclear. Uncertainties associated with the identification of protohaloes within the linear density field may have adverse effects on the predictive power of the density bias model. A more detailed understanding of how protohaloes map onto linear density peaks, and vice versa, will, no doubt, provide valuable insight into the mechanisms behind halo biasing. \section*{Acknowledgments} AE acknowledges financial support from the SFB-Transregio 33 ``The Dark Universe" by the Deutsche Forschungsgemeinschaft (DFG) and a scholarship from the Bonn-Cologne Graduate School (BCGS). We acknowledge interesting discussion on halo finders with Steffen Knollmann and Andrea Macci\`o. We would like to thank Jeremy Tinker for providing us with his SO halo finder, and for help applying it to our simulations. We are very grateful to the referee, Vincent Desjacques, for useful suggestions and valuable insights.
\section{Introduction} Let $G$ be a finite group, $g$ an element of $G$ of order $n$ and $k$ and $m$ positive integers so that $k^m \equiv 1 \mod n$. Then $$\Bass{k}{m}{g}=(1+g+\dots + g^{k-1})^{m}+\frac{1-k^m}{n}(1+g+g^2+\dots+g^{n-1})$$ is a unit of the integral group ring $\Z G$. The units of this form were introduced by Bass in \cite{bass1966} and are known as Bass units or Bass cyclic units. Bass proved that if $G$ is a cyclic group then the Bass units of $\Z G$ generate a subgroup of finite index in $\U(\Z G)$. (The group of units of a ring $R$ is denoted $\U(R)$.) Bass and Milnor extended this result to finite abelian groups. In this paper we will refer to this result as the Bass-Milnor Theorem while the Bass Theorem refers to the result for cyclic groups. The Bass Theorem also provides a basis consisting of Bass units for a free abelian subgroup of finite index in $\U(\Z G)$, provided $G$ is cyclic. As far as we know, no basis consisting of Bass units for a free abelian subgroup of finite index is known for an arbitrary abelian group $G$. It is worth mentioning that in more recent work Marciniak and Sehgal \cite{2005MarciniakSehgal,2006MarciniakSehgal} investigate the so called group of generic units $\textnormal{Gen}(G)$ in $\U(\Z G)$, for $G$ a finite abelian group. It turns out that $\textnormal{Gen}(G)$ is generated by the Hoechsmann units \cite{Hoechsmann1992} (for the definition see \cite[p. 34]{Sehgal1993}), and some power of a Hoechsmann unit is a product of two Bass units. Furthermore, the group generated by the Bass units $\Bass{k}{\varphi(|G|)}{g}$ is contained in $\textnormal{Gen}(G)$. For other results, we refer the reader to \cite{2005MarciniakSehgal,2006MarciniakSehgal}. In the remainder of the paper we assume that $G$ is a finite abelian group. If $u\in \U(\Z G)$ then, by the Bass-Milnor Theorem, some power of $u$ is a product of Bass units. Unfortunately the proofs by Bass and Bass-Milnor do not provide a method to express some power of $u$ as a product of Bass units. The aim of this paper is to obtain such a method for some relevant units. To do so we first obtain a new proof of the Bass-Milnor Theorem which follows a different approach than the proofs of Bass and Bass-Milnor. In order to have a feeling of the features of this new proof first we recall the main steps of the Bass and Bass-Milnor proofs and outline those of our proof. Let $\Gamma$ be a finitely generated abelian group. The rank of all free abelian subgroups of finite index in $\Gamma$ is an invariant of the group, called the rank of $\Gamma$. Assume that $\Gamma$ has rank $r$ and let $u_1,\dots,u_r\in \Gamma$. Then $\GEN{u_1,\dots,u_r}$ has finite index in $\Gamma$ if and only if $u_1,\dots,u_r$ are multiplicatively independent. For example, $\U(\Z G)$ is finitely generated abelian and has rank $r=\frac{1+k_2+|G|-2c}{2}$, where $c$ is the number of cyclic subgroups of $G$ and $k_2$ is the number of elements of $G$ of order $2$. This result is due to \cite{Higman1940} (for the above formula for $r$ see also \cite{1969AyoubAyoub}). Bass took a concrete list of $r$ Bass units of $\Z G$ for $G$ cyclic and proved that they are multiplicatively independent. To do so, Bass used the Bass Independence Theorem which in turn uses the Franz Independence Lemma (see \cite[11.3, 11.8]{Sehgal1993} for details). Bass and Milnor proved, using K-Theory, that the group generated by the units of integral group rings of cyclic subgroups of $G$ has finite index in $\U(\Z G)$. However, both the independency and the K-Theory arguments are useless when trying to write a given unit as a product of Bass units. Alternatively, assume that $\Gamma$ is a subgroup of finite index in a finitely generated abelian group $\Lambda$ and that we know a subset $X$ of $\Lambda$ which generates a subgroup of finite index in $\Lambda$. Let $Y$ be a subset of $\Gamma$. Then $\GEN{Y}$ has finite index in $\Gamma$ if and only if for every $x\in X$ there is a positive integer $m$ such that $x^m \in \GEN{Y}$. In our proof we take $\Gamma=\U(\Z G)$, $Y$ the set of Bass units of $\Z G$, $\Lambda=\U(\O)$, where $\O$ is the unique maximal order of $\Q G$, and $X$ the set of cyclotomic units of $\Q G$, which are defined as follows. Let $\zeta_n$ denote a complex root of unity of order $n$. If $n>1$ and $k$ is an integer coprime to $n$ then $$\eta_{k}(\zeta_n)=\frac{1-\zeta_n^k}{1-\zeta_n}=1+\zeta_n+\zeta_n^2+\dots+\zeta_n^{k-1}$$ is a unit of $\Z[\zeta_n]$, the $n$-th cyclotomic ring of integers. We extend this notation by setting $$\eta_k(1)=1.$$ The units of the form $\eta_k(\zeta_n^j)$, with $j,k$ and $n$ integers such that $\gcd(k,n)=1$, are called the cyclotomic units of $\Q(\zeta_n)$. A classical result, which goes back to Kummer, states that the cyclotomic units of $\Q(\zeta_n)$ generate a subgroup of finite index in $\U(\Z[\zeta_n])$ \cite{1982Washington}. By a well-known theorem of Perlis and Walker, $\Q G$ is isomorphic to a direct product of cyclotomic fields. For simplicity, we consider this isomorphism as an equality $\Q G=\prod_{i=1}^k \Q(\zeta_{n_i})$. Then $\Z G$ is an order of $\Q G$ and $\O=\prod_{i=1}^k \Z[\zeta_{n_i}]$ is the unique maximal order of $\Q G$. In particular, $\Z G \subseteq \O$. Moreover, $\Gamma=\U(\Z G)$ has finite index in $\Lambda=\U(\O)$ (see \cite[Lemma~4.6]{Sehgal1993}). The cyclotomic units of $\Q G$ are, by definition, the elements of $\Q G$ which project to a cyclotomic unit of $\Q(\zeta_{n_i})$ for some $i=1,\dots,k$ and project to $1$ in the remaining components. Having in mind that the cyclotomic units of $\Q(\zeta_{n_i})$ generate a subgroup of finite index in $\U(\Z[\zeta_{n_i}])$, we conclude that the set $X$ of cyclotomic units of $\Q G$ generates a subgroup of finite index in $\Lambda$. Thus the Bass-Milnor Theorem is equivalent to the following proposition. \begin{proposition}\label{OneComponentBass} Let $G$ be a finite abelian group and let $\Q G = \prod_{i=1}^k \Q(\zeta_{n_i})$, the realization of Perlis-Walker Theorem. Then for every cyclotomic unit $u$ of $\Q G$ there is a positive integer $m$ such that $u^m$ is a product of Bass units of $\Z G$. \end{proposition} The proof of Proposition~\ref{OneComponentBass} (actually of Lemma~\ref{BassProjectionsL}) is constructive and avoids K-Theory and independence arguments. As a result the proof provides an algorithm that for a cyclotomic unit $\eta$ as input, returns $m$ and an expression of $\eta^m$ as a product of Bass units (see Algorithm~\ref{alg1}). This is the first result of the paper. The second result consists in giving a concrete basis $B$ formed by Bass units for a free abelian subgroup of finite index in $\U(\Z G)$. The proof of the fact that $B$ is a basis also provides an algorithmic method to write some power of any arbitrary Bass unit as a product of elements of $B$ (see Section~\ref{SectionBasis}). Observe that our result provides an algorithmic method to write some power of a given $u\in \U(\O)$ as a product of Bass units, or even of elements of $B$, as far as we can write some power of each projection of $u$ in the simple components of $\Q G$ as a product of cyclotomic units. However we do not know of any algorithmic method that starting from an element $u\in \U(\Z[\zeta_n])$ as input, returns an expression of some power of $u$ as a product of cyclotomic units. \section{A new proof of the Bass-Milnor Theorem} Throughout the rest of the paper $G$ is a finite abelian group. We denote the group generated by the Bass units of $\Z G$ by $B(G)$. In this section we prove Proposition~\ref{OneComponentBass}. This provides a new proof of the Bass-Milnor Theorem which states that $B(G)$ has finite index in $\U(\Z G)$. First of all we obtain a precise realization of Perlis and Walker Theorem which states that there is an isomorphism $f:\Q G \rightarrow \prod_d \Q(\zeta_d)^{k_d}$, where $k_d$ denotes the number of cyclic subgroups of $G$ of order $d$. This isomorphism is realized as follows. Let $\Hc=\Hc(G)$ denote the set of subgroups $H$ of $G$ such that $G/H$ is cyclic. For every subgroup $H\in \Hc$ we fix a linear representation $\rho_H$ of $G$ with kernel $H$. We also denote by $\rho_H$ the linear extension of $\rho_H$ to $\Q G$. If $d=[G:H]$ then $\rho_H(\Q G)=\Q(\zeta_d)$, where $\zeta_{d}$ denotes a primitive $d$-th root of unity. Then $$f=\prod_{H\in \Hc} \rho_H:\Q G \longrightarrow \prod_{H\in \Hc} \Q(\zeta_{[G:H]})$$ is an isomorphism. This isomorphism is the same as the one of Perlis and Walker \cite{1950PerlisWalker} (see for example also \cite{2002SehgalMilies}), since $k_d$ equals the number of subgroups $H\in \Hc$ such that $[G:H]=d$. We will use the following equalities (the order of an element $g\in G$ is denoted by $|g|$) \begin{eqnarray}\label{BassCyclotomic} \rho_H(\Bass{k}{m}{g}) &=& \eta_k(\rho_H(g))^m, \quad (H\in \Hc, g\in G, k^m\equiv 1 \mod |g|), \\ \prod_{i=0}^{n-1} (1-X\zeta_n^i) &=& 1-X^n. \label{1-Xn} \end{eqnarray} Let $\xi$ be a root of unity and assume that $k$ is coprime to $n$ and the order of $\xi$. If $\xi^n\ne 1$ then, using (\ref{1-Xn}), we obtain \begin{eqnarray*} \prod_{i=0}^{n-1} \eta_k(\xi \zeta_n^i) = \prod_{i=0}^{n-1} \frac{1-\xi^k \zeta_n^{ki}}{1-\xi{\zeta_n}^i} = \frac{\prod_{i=0}^{n-1} (1-\xi^k \zeta_n^i)}{\prod_{i=0}^{n-1} (1-\xi{\zeta_n}^i)} = \frac{1-\xi^{kn}}{1-\xi^n} = \eta_k(\xi^n). \end{eqnarray*} Otherwise, i.e. if $\xi^n=1$, then $\xi\zeta_n^j=1$ for some $j=0,1,\dots,n-1$. Then, using that $k$ is coprime to $n$, we deduce that \begin{eqnarray*} \prod_{i=0}^{n-1} \eta_k(\xi \zeta_n^i) = \prod_{i=0,i\neq j}^{n-1} \frac{1-\zeta_n^{k(i-j)}}{1-\zeta_n^{i-j}} = \frac{\prod_{i=1}^{n-1} (1-\zeta_n^{ki})}{\prod_{i=1}^{n-1} (1-\zeta_n^i)} = 1 = \eta_k(\xi^n). \end{eqnarray*} This proves the following equality for every primitive $l$-th root of unity $\xi$ \begin{equation}\label{ProductCyclotomic} \prod_{i=0}^{n-1} \eta_k(\xi \zeta_n^i) = \eta_k(\xi^n) \quad (\gcd(k,nl)=1). \end{equation} \begin{lemma}\label{ProductBassCyclic} Let $g\in G$, $H\in \Hc$ and $K$ be an arbitrary subgroup of $G$. Let $h=|H\cap K|$, $t=[K:H\cap K]$ and let $k$ and $m$ be positive integers such that $(k,t)=1$ and $k^m \equiv 1 \mod |gu|$ for every $u\in K$. Then $$\prod_{u\in K} \rho_H(\Bass{k}{m}{gu}) = \eta_k(\rho_H(g)^t)^{mh}.$$ \end{lemma} \begin{proof} As $H=\ker(\rho_H)$, if $u$ runs through the elements of $K$ then $\rho_H(u)$ runs through the $t$-th roots of unity and each $t$-th root of unity is obtained as $\rho_H(u)$ for precisely $h$ elements $u$ of $K$. Therefore $$\prod_{u\in K} \rho_H(\Bass{k}{m}{gu}) = \left( \prod_{u\in K} \eta_k(\rho_H(g)\rho_H(u)) \right)^{m} = \left(\prod_{i=0}^{t-1} \eta_k(\rho_H(g)\zeta_t^i) \right)^{mh} = \eta_k(\rho_H(g)^t)^{mh}$$ as desired. We have used (\ref{ProductCyclotomic}) in the last equality. \end{proof} Proposition~\ref{OneComponentBass} is a consequence of the following stronger lemma. \begin{lemma}\label{BassProjectionsL} Let $H\in \Hc$ with $d=[G:H]$ and let $k,j\in \N$ be such that $k$ is coprime to $d$. Set $\eta=\eta_k(\zeta_d^j)$ and let $B_k(G)$ be the subgroup of $\U(\Z G)$ generated by the Bass cyclic units of the form $\Bass{k}{m}{g}$ with $g\in G$ and $k^m \equiv 1 \mod |g|$. Then there is a positive integer $m$ and $b\in B_k(G)$ such that $\rho_H(b)=\eta^m$ and $\rho_K(b)=1$ for every $K\in \Hc\setminus \{H\}$. \end{lemma} \begin{proof} Without loss of generality, we may assume that $k$ is coprime to $n=|G|$. Indeed, by an easy Chinese Remainder argument there is an integer $k'$ coprime to $n$ such that $k\equiv k'\mod d$. Then clearly $\eta_k(\zeta_d^j)=\eta_{k'}(\zeta_d^j)$. We argue by a double induction, first on $n$ and second on $d$. The cases $n=1$ and $d=1$ are trivial. We denote by $P(G,H)$ the statement of the lemma for a finite abelian group $G$ and an $H\in \Hc(G)$. Hence the induction hypothesis includes the following statements: \begin{itemize} \item[(IH1):] $P(M,Y)$ holds for every proper subgroup $M$ of $G$ and any $Y\in \Hc(M)$. \item[(IH2):] $P(G,H_{1})$ holds for every $H_{1}\in \Hc(G)$ with $[G:H_{1}]<[G:H]=d$. \end{itemize} We consider two cases, depending on whether $j$ is coprime to $d$ or not. \textsc{Case 1}: $j$ is not coprime to $d$. Let $p$ be a common prime divisor of $d$ and $j$. Then $H$ is contained in a subgroup $S$ of $G$ with $[G:S]=p$ and $\zeta_{d}^{j}=\zeta_{d'}^{j/p}$ with $d'=[S:H]$. For every $K\in \Hc(G)$, let $\lambda_K$ denote the restriction of $\rho_K$ to $\Q S$. Clearly $\lambda_K$ is the $\Q$-linear extension of a linear representation of $S$ with kernel $S\cap K$. Since $S/(S\cap K)\cong KS/K$ and $KS/K$ is a subgroup of $G/K$ we deduce that $S/(S\cap K)$ is cyclic. Thus $K\rightarrow S\cap K$ defines a map $\Hc(G)\rightarrow \Hc(S)$. This map is surjective, but maybe not injective. Indeed, let $K_1\in \Hc(S)$. If $K_1\in \Hc(G)$ then clearly the map associates $K_1$ with $K_1$. Otherwise $p$ divides $[G:K_1]$ and $S/K_1$ is a cyclic subgroup of $G/K_1$ of maximal order. This implies that $G/K_1=S/K_1\times L/K_1$ for some subgroup $L$ of $G$ containing $K_1$ and so that $[L:K_1]=p$. Then $G/L\cong S/K_1$, so that $L\in \Hc(G)$, and $L\cap S=K_1$. Therefore $\Hc(S)=\{K\cap S : K\in \Hc(G)\}$. For every $Y\in \Hc(S)$ we choose a $K_Y\in \Hc(G)$ such that $K_Y\cap S=Y$ in such a way that $K_Y=Y$ if $Y\in \Hc(G)$. Then $$\prod_{Y\in \Hc(S)} \lambda_{K_Y} : \Q S \rightarrow \prod_{Y\in \Hc(S)} \Q (\zeta_{[S:Y]})$$ is an isomorphism of algebras. By the first induction hypothesis (IH1) there is $b\in B_k(S)$ such that $\lambda_H(b)=\eta^m$ for some positive integer $m$ and $\rho_K(b)=\lambda_K(b)=1$ if $K\in \Hc(G)$ with $K\cap S\ne H$. If $K\in \Hc(G)$ satisfies $K\cap S=H$ then either $K=H$ or $K=H_1$, where $H_1/H$ is the only subgroup of $G/H$ of order $p$, since $G/H$ is cyclic. Moreover $\rho_{H_1}(b)$ is a product of cyclotomic units, by (\ref{BassCyclotomic}). By the second induction hypothesis (IH2) there is $c\in B_k(G)$ such that $\rho_K(c)=1$ for every $K \in \Hc(G)\setminus \{H_1\}$ and $\rho_{H_1}(c)=\rho_{H_1}(b)^{m_1}$ for some positive integer $m_1$. Therefore $\rho_H(b^{m_1}c\inv)=\eta^{mm_1}$ and $\rho_K(b^{m_1}c\inv)=1$ for every $K\in \Hc(G)\setminus \{H\}$. This finishes the proof for this case. \textsc{Case 2}: $j$ is coprime to $d$. Then $G=\GEN{a,H}$ and $\rho_H(a)=\zeta_d^j$ for some $a\in G$. As $k$ is coprime to $n$, there is a positive integer $m$ such that $k^m\equiv 1 \mod |au|$ for every $u\in H$. Hence $$\eta^m = \rho_H(\Bass{k}{m}{a}),$$ by (\ref{BassCyclotomic}). Let $$b=\prod_{h\in H} \Bass{k}{m}{ah}.$$ For every $K\in \Hc(G)$, set $$d_K=[G:K],\; d'_K=[G:\GEN{a,K}],\; h_K=|H\cap K| \mbox{ and } t_K=[H:H\cap K].$$ Then, by Lemma \ref{ProductBassCyclic}, $$\rho_K(b) = \eta_k(\rho_K(a)^{t_K})^{m\, h_K} = \eta_k(\zeta_{d_K}^{t_Kd'_Ku_K})^{m\, h_K},$$ for some integer $u_K$ coprime to $d_K$. If $t_Kd'_K$ is not coprime to $d_K$ then, by Case 1, there is $b_K\in B_k(G)$ such that $\rho_K(b_K)=\rho_K(b)^{m_K}$ for some integer $m_K$ and $\rho_{K_{1}}(b_{K})=1$ for $K_{1}\in \Hc(G)\setminus \{ K \}$. By (IH2), the same holds if $d_K<d$. Let $$\Hc'=\{K\in \Hc(G): t_Kd'_K \text{ is not coprime to }d_K \text{ or } d_K<d\}.$$ For each $K\in \Hc'$ fix $b_K\in B_k(G)$ and $m_K\in \Z$ as above and $m_1=\lcm(m_K:K\in \Hc')$ and $$b_1=\left(\prod_{K\in \Hc'}b_K^{-\frac{m_1}{m_K}}\right)b^{m_1}.$$ Then $b_1\in B_k(G)$, $\rho_K(b_1)=1$ if $K\in \Hc'$ and $\rho_K(b_1)\in \GEN{\eta_k(\rho_K(a)^{t_K})}$ if $K\in \Hc(G)\setminus \Hc'$. Observe that $t_H=d'_H=u_H=1$ and hence $H\not\in \Hc'$. Therefore, $\rho_H(b_1)\in \GEN{\eta}$, because $\eta_k(\rho_H(a)^{t_H})=\eta$. To finish the proof we prove that $\Hc'=\Hc(G)\setminus \{H\}$. Suppose the contrary, that is, assume $K\in \Hc(G)\setminus\{H\}$ with $d_K\ge d$ and $\gcd(t_Kd'_K,d_K)=1$. The latter implies that $d'_K=1$, or equivalently $G=\GEN{a,K}$, and $t_K=[KH:K]$ is coprime to $d_K=[G:K]$. Consequently, $t_K=1$, or equivalently $H\subseteq K$. Hence, the assumption $d_K=[G:K]\ge [G:H]=d$ implies that $H=K$, a contradiction. \end{proof} As it was mentioned in the introduction, the Bass-Milnor Theorem is equivalent to Proposition~\ref{OneComponentBass} and the well known fact that the cyclotomic units generate a subgroup of finite index of $\U(\Z[\zeta])$ for every root of unity $\zeta$. We include a proof for completeness. \begin{theorem}[Bass-Milnor]\label{BassAbelianFiniteIndexT} If $G$ is a finite abelian group then the group generated by the Bass units of $\Z G$ has finite index in $\U(\Z G)$. \end{theorem} \begin{proof} Let $\Hc=\Hc(G)$, $f=\prod_{H\in \Hc} \rho_H:\Q G\rightarrow \prod_{H\in \Hc} \Q(\zeta_{[G:H]})$, $B=B(G)$ and $V$ be the subgroup of $\prod_{H\in \Hc} \U(\Z(\zeta_{[G:H]}))$ generated by the units that project on one component to a cyclotomic unit and project on all other components to 1. As the cyclotomic units of each ring of cyclotomic integers $\Z[\zeta_n]$ generate a subgroup of finite index in $\U(\Z[\zeta_n])$, $V$ has finite index in $\prod_{H\in \Hc} \U(\Z(\zeta_{[G:H]}))$. Hence, by Lemma \ref{BassProjectionsL}, $f(B)$ has finite index in $\prod_{H\in \Hc} \U(\Z(\zeta_{[G:H]}))$ and therefore $B$ has finite index in $\U(\Z G)$, since $B \subseteq \U(\Z G) \subseteq f\inv(\prod_{H\in \Hc} \U(\Z(\zeta_{[G:H]})))$. \end{proof} We now present Algorithm \texttt{CyclotomicAsProductOfBass} which, for a cyclotomic unit $\eta$ of $\Q G$ as input, returns a power of $\eta$ as product of Bass units. More precisely, the input is formed by a finite abelian group $G$, a subgroup $H\in \Hc(G)$, a list of linear characters $(\rho_K:K\in \Hc(G))$ of $G$ with $\ker \rho_K=K$ and two integers $k$ and $j$ with $\gcd(k,[G:H])=1$. The output is formed by an integer $m$, an integer $k'$ coprime to $|G|$ such that $k'\equiv k\mod [G:H]$ , and a list $((g_i,m_i,p_i):i\in I)$ with $g_i\in G$ and $m_i$ and $p_i$ integers such that $\gcd(k',|g_i|)=1$, $k'^{m_i}\equiv 1 \mod |g_i|$ and $b=\prod_{i\in I} \Bass{k'}{m_i}{g_i}^{p_i}$ is so that $\rho_H(b)=\eta_k(\zeta_{[G:H]}^j)^m$ and $\rho_K(b)=1$ for every $K\in \Hc(G)\setminus \{H\}$. Observe that each linear character $\rho$ with kernel $K$ is determined by $\rho(a)$, where $a$ is an element of $G$ such that $G=\GEN{a,K}$. Moreover, for such an $a$ there is a unique linear character $\rho$ with kernel $K$ such that $\rho(a)=\zeta_{[G:K]}$. Thus we describe the list of characters $\rho_K$ by a list $A=(a_K:K\in \Hc(G))$ of elements of $G$ such that $G=\GEN{a_K,K}$ for every $K\in \Hc(G)$. In other words, in the input of \texttt{CyclotomicAsProductOfBass} we replace the list of characters by the list $A$. Then $\rho_K$ is the unique linear character of $G$ with kernel $K$ such that $\rho_K(a_K)=\zeta_{[G:K]}$. \texttt{CyclotomicAsProductOfBass} follows the structure of the proof of Lemma~\ref{BassProjectionsL}. Notice that this is an inductive proof and henceforth the structure of the Algorithm is recursive. That is, the algorithm includes calls to itself. In these calls to itself the first argument, the ambient group, is replaced by a subgroup $S$ of $G$. In principle the algorithm could start by calculating $\Hc(G)$. However, this would make the algorithm inefficient because each call to itself should calculate $\Hc(S)$, which is not necessary because \begin{equation}\label{HcSHcG} \Hc(S)=\{S\cap K:K\in \Hc(G)\}, \end{equation} as was observed in the proof of Lemma~\ref{BassProjectionsL}. Thus we include $\Hc(G)$ as part of the input of the algorithm and, before each recursive call of the algorithm for a subgroup $S$ of $G$, one calculates $\Hc(S)$ from $\Hc(G)$ using (\ref{HcSHcG}). Of course one should filter the list to eliminate repetitions in $\Hc(S)$. Furthermore, from the list $\{a_K:K\in \Hc(G)\}$, satisfying $G=\GEN{a_K,K}$ for each $K\in \Hc(G)$, one can calculate another list $\{b_Y:Y\in \Hc(S)\}$, satisfying $S=\GEN{b_Y,Y}$ for each $Y\in \Hc(S)$. In fact, the proof of Lemma~\ref{BassProjectionsL} tell us that for each $Y\in \Hc(S)$ one can select a $K_Y\in \Hc(G)$ such that $Y=S\cap K_Y$, so that if $Y\in \Hc(G)$ then $K_Y=Y$ and the linear representation $\lambda_Y$ of $S$ with kernel $Y$ chosen is the restriction of $\rho_{K_Y}$ to $S$. Thus, we need to take $b_Y$ such that $\rho_K(b_Y)=\zeta_{[S:Y]}$. In fact in each recursive call of the algorithm either $S=G$ or $[G:S]$ is prime. If $S=G$, then $K_Y=Y$ and we may take $b_Y=a_Y$ for each $Y$. Assume otherwise that $[G:S]=p$, a prime integer. If $Y\in \Hc(G)$ then $Y=K_Y$, $[S:Y]=\frac{[G:K_Y]}{p}$ and $\rho_{K_Y}(a_Y^p)=\zeta_{[G:K_Y]}^p = \zeta_{[S:Y]}$, so we may take $b_Y=a_Y^p$. Otherwise, i.e. if $Y\not\in \Hc(G)$, then $[K_Y:Y]=p$ and $G/Y=S/Y\times K_Y/Y$. In particular, $p$ divides $[G:Y]$ and $[S:Y]=[G:K_Y]$. Then $b_Y$ is an element of $a_{K_Y}K_Y\cap S$. We use the following notation for $m,n\in \Z$ and $A$, $B$ and $A_l$ ($l\in L$) lists. \begin{eqnarray*} m \mod n &=& \text{Remainder of }m \text{ modulo }n; \\ O_n(m) &=& \text{Multiplicative order of } m \text{ modulo }n \text{ (if } \gcd(m,n)=1);\\ (\_) &=& \text{Empty list};\\ A\sqcup B &=& \text{ Concatenation of the lists } A \text{ and } B; \\ \bigsqcup_{i=1}^l A_i &=& \text{ Concatenation of the lists } A_1,\dots,A_l. \end{eqnarray*} {\small \begin{algorithm}[h!] \caption{\texttt{CyclotomicAsProductOfBass}($G,\Hc(G),A,H,k,j$)} \label{alg1} \begin{algorithmic}[1] \Require $G$ a finite abelian group, $\Hc(G)=(H\leq G : G/H \mbox{ cyclic})$, $A=(a_K : K\in\Hc(G))$, a list of group elements $a_K$ of $G$ such that $G=\GEN{a_K,K}$, $H\in \Hc(G)$, $j,k\in \N$ with $\gcd(k,[G:H])=1$. \Comment{$\rho_K$ linear representation of $G$: $\ker \rho_K = K$; $\rho_K(a_K)=\zeta_{[G:K]}$.} \Ensure $(m, k', B=((g_i,m_i,q_i) : i=1,\dots,r)) \in \Z^2 \times (G\times \Z^{2})^{r}$. \Comment{ $\rho_K\left(\prod_{i=1}^{r}\Bass{k'}{m_i}{g_i}^{q_i}\right)= \left\{\begin{array}{ll} \eta_k(\zeta_{[G:H]}^j)^m, & \text{if } K=H;\\ 1, &\text{if } K\in \Hc(G)\setminus \{H\}. \end{array}\right.$} \State $d := [G:H]$; \State $k':=$ an integer coprime to $|G|$ congruent to $k$ modulo $d$. \Comment{Chinese Remainder Theorem.} \State $j:=(j \mod d);$ \If {$d=1$} \State $m:=1;\;B=(\_)$; \ElsIf {$\gcd(j,d)\neq 1$} \State $p :=$ common prime divisor of $j$ and $d$; \State $S:=\GEN{a_H^p,H}$ \Comment{Unique subgroup of $G$ of index $p$ containing $H$.} \State $\Hc(S) := (\_); \; A_S=(\_)$; \For {$K \in \Hc(G)$ with $K\subseteq S$} \State Add $K$ to $\Hc(S)$; \State Add $a_{K}^p$ to $A_S$; \EndFor \For {$K\in \Hc(G)$ with $K\not\subseteq S$} \State $Y:=K\cap S$; \If {$Y \not\in \Hc(S)$} \Comment{To exclude redundancy} \State Add $Y$ to $\Hc(S)$; \State Add an element of $S\cap a_K K$, to $A_S$; \Comment{$Y=S \cap K=\ker \rho_K|_Y, \rho_K(b_Y)=\zeta_{[S:Y]}$.} \EndIf \EndFor \State $(m,k',B=((s_i,m_i,q_i):i\in L)):=\texttt{CyclotomicAsProductOfBass}(S,\Hc(S),A_S,H,k',j/p)$; \State $H_1:=\GEN{a_{H}^{d/p},H}$; \Comment{Only subgroup of $G$ containing $H$ with $[H_1:H]=p$} \For{$i\in L$} \State Determine $j_i$ such that $s_i\in a_{H_1}^{j_i}H_1$; \State $(n_i,k',C_i=((g_u,m_u,r_u) : u\in L_i)) := \texttt{CyclotomicAsProductOfBass}(G,\Hc(G),A,H_1,k',j_i)$ \EndFor \State $m'$ := $\lcm(n_i:i\in L)$; \State $B := ((s_i,m_i,q_im'):i\in L) \sqcup \bigsqcup_{i\in L} \left(\left(g_u,m_u,-\frac{r_uq_im_im'}{n_i}\right): u\in L_i\right)$; \State $m$ := $mm'$; \Else \Comment{if $\gcd(j,d)=1$} \State $m := \lcm (O_{|a_H^jh|}(k') : h\in H)$; \State $B := ((a_H^jh,m,1): h\in H)$; \For{$K\in \Hc(G)\setminus\{H\}$} \State $d_K := [G:K]; \; h_K := |H\cap K|; \; t_K := [H:H\cap K]$; \State $v_K := \text{ integer such that } a_HK=a_K^{v_K}K$; \Comment{$\rho_K\left(\prod_{h\in H} \Bass{k'}{m}{a_{H}^jh}\right) = \eta_k(\zeta_{d_K}^{t_Kv_K})^{mh_{H}}$.} \State \begin{tabular}{rcl} $B_K$&$=$&$(m_K,k',((g_{i_K},m_{i_K},r_{i_K}) : i_K \in I_K))$ \\ &:=& \texttt{CyclotomicAsProductOfBass}($G,\Hc(G),A,K,k',t_Kv_K$); \end{tabular} \EndFor \State $n$ := $\lcm(m_K:K\in \Hc(G)\setminus\{H\})$; \State $B:= ((a_H^jh,m,n) : h\in H) \sqcup \bigsqcup_{K\in \Hc(G)\setminus \{H\}} \left(\left(g_{i_K},m_{i_K},-\frac{r_{i_K}mh_Kn}{m_K}\right): i_K \in I_K\right)$; \State $m$ := $mn|H|$; \EndIf \Return $(m,k',B)$; \end{algorithmic} \end{algorithm}}\clearpage This algorithm is suitable for implementation in an appropriate programming language such as GAP \cite{GAP4} and could be added to existing GAP packages such as Wedderga \cite{Wedderga}, which also deals with computations in group rings. \section{A basis of Bass units}\label{SectionBasis} Bass proved that if $G=\GEN{g}$, a cyclic group of order $n$, and $m$ is a multiple of $\varphi(n)$ then $\left\{\Bass{k}{m}{g^d}:d\mid n,1<k< \frac{n}{2d}, (k,\frac{n}{d})=1\right\}$ is a basis of a free abelian subgroup of finite index in $\U(\Z G)$ \cite{bass1966}. (Here $\varphi$ stands for the Euler totient function.) In this section we generalize this result and obtain a basis of Bass units for a subgroup of finite index in the integral group ring of an arbitrary abelian group $G$. Moreover, the proof provides, for an arbitrary Bass unit $b$, an algorithm to express a power of $b$ as a product of a trivial unit and powers of at most two units in this basis of Bass units. \begin{theorem}\label{BasisBassT} Let $G$ be a finite abelian group. For every cyclic subgroup $C$ of $G$ choose a generator $a_C$ of $C$ and for every $k$ coprime to the order of $C$ choose an integer $m_{k,C}$ with $k^{m_{k,C}}\equiv 1 \mod |C|$. Then $$\left\{\Bass{k}{m_{k,C}}{a_C}:C \text{ cyclic subgroup of } G, 1<k< \frac{|C|}{2}, \gcd(k,|C|)=1\right\}$$ is a basis for a free abelian subgroup of finite index in $\U(\Z G)$. \end{theorem} \begin{proof} The proof is based on the following equalities (\cite[Lemma 3.1]{2006GoncalvesPassman}): \begin{eqnarray} \label{Basseq1} \Bass{k}{m}{g}&=&\Bass{k_1}{m}{g}, \mbox{ if } k\equiv k_1 \mod |g|, \\ \label{Basseq2} \Bass{k}{m}{g}\Bass{k}{m_1}{g}&=&\Bass{k}{m+m_1}{g},\\ \label{Basseq3} \Bass{k}{m}{g}\Bass{k_1}{m}{g^k}&=&\Bass{kk_1}{m}{g},\\ \label{Basseq4} \Bass{1}{m}{g}&=&1 \text{ and } \\ \label{Basseq5} \Bass{|g|-1}{m}{g} &=& (-g)^{-m}, \end{eqnarray} for $g\in G$ and $k^m\equiv k_1^m \equiv k^{m_1} \equiv 1 \mod |g|$. By (\ref{Basseq2}) we have \begin{equation}\label{Basseq6} \Bass{k}{m}{g}^{h} = \Bass{k}{mh}{g} \end{equation} and from (\ref{Basseq1}), (\ref{Basseq3}) and (\ref{Basseq5}) we deduce \begin{equation}\label{Basseq7} \Bass{|g|-k}{m}{g} = \Bass{k(|g|-1)}{m}{g} = \Bass{k}{m}{g} \Bass{|g|-1}{m}{g^k} = \Bass{k}{m}{g}g^{-km} \end{equation} provided $(-1)^m\equiv 1 \mod |g|$. By Theorem~\ref{BassAbelianFiniteIndexT}, the Bass units generate a subgroup of finite index in $\U(\Z G)$. Let $t=\varphi(|G|)$. We first prove that $B_1=\left\{\Bass{k}{t}{a_C}\mid 1<k<\frac{|C|}{2}, \gcd(k,|C|)=1\right\}$ generates a subgroup of finite index in $\U(\Z G)$. To do so we ``sieve'' gradually the list of Bass units, keeping the property that the remaining Bass units still generate a subgroup of finite index in $\U(\Z G)$, until the remaining Bass units are the elements of $B_1$. By equation (\ref{Basseq1}), to generate $B(G)$ it is enough to use the Bass units of the form $\Bass{k}{m}{g}$ with $g\in G$, $1\le k < |g|$ and $k^m \equiv 1 \mod |g|$. By (\ref{Basseq6}), for every Bass unit $\Bass{k}{m}{g}$ we have $\Bass{k}{m}{g}^u = \Bass{k}{t}{g}^v$ for some positive integers $u$ and $v$. Thus the Bass units of the form $\Bass{k}{t}{g}$ with $1\le k < |g|$ and $\gcd(k,|g|)=1$ generate a subgroup of finite index in $\U(\Z G)$. By (\ref{Basseq2}) and (\ref{Basseq3}), we can reduce further the list of generators by taking only those with $g=a_C$ for some cyclic group $C$ of $G$. By (\ref{Basseq4}) and (\ref{Basseq5}) we can exclude the Bass units with $k=\pm 1$ and still generate a subgroup of finite index in $\U(\Z G)$ with the remaining elements. Finally, $\Bass{k}{t}{g}\inv \Bass{|g|-k}{t}{g}=\Bass{|g|-1}{t}{g^k}=g^{-kt}$, by (\ref{Basseq5}) and (\ref{Basseq7}). Thus $\Bass{k}{t}{g}\inv \Bass{|g|-k}{t}{g}$ has finite order. Therefore the units $\Bass{k}{t}{g}$ with $k>\frac{|g|}{2}$ can be excluded. The remaining units are exactly the elements of $B_1$. Thus $\GEN{B_1}$ has finite index in $\U(\Z G)$, as desired. Let $\mathcal{C}$ be the set of cyclic subgroups of $G$ and $B=\{\Bass{k}{m_{k,C}}{a_C} : C\in \mathcal{C}, 1<k< \frac{C}{2}, \gcd(k,|C|)=1\}$. Using (\ref{Basseq6}) once more, we deduce that $\GEN{B}$ has finite index in $\U(\Z G)$, since so does $\GEN{B_1}$. To finish the proof we need to prove that the elements of $B$ are multiplicatively independent. To do so, it is enough to show that the rank of $\U(\Z G)$ coincides with the cardinality of $B$. For this, first observe that the cardinality of $B$ is $\sum_d k_d t_d$, where $d$ runs through the divisors of $|G|$, $k_d$ is the number of cyclic subgroups of $G$ of order $d$ and $t_d$ is the cardinality of $\{k : 1<k< \frac{d}{2}, \gcd(d,k)=1\}$. Obviously $t_1=t_2=0$ and $t_d=\frac{\varphi(d)}{2}-1$ for every $d>2$. Therefore, $|B|=\sum_{d>2} \left(\frac{k_d\varphi(d)}{2}-k_d\right) = \frac{1+k_2+\sum_d h_d}{2}-\sum_d k_d = \frac{1+k_2+|G|-2c}{2}$, where $h_d$ denotes the number of elements of $G$ of order $d$ (so that $h_1=1$ and $h_2=k_2$) and $c$ is the number of cyclic subgroups of $G$. By a Theorem of Higman \cite{Higman1940}, this number coincides with the rank of $\U(\Z G)$ and the proof is finished. \end{proof} In Theorem~\ref{BasisBassT} one can choose, for example, $m_{k,C}=\varphi(|G|)$, $m_{k,C}=\varphi(|C|)$ or $m_{k,C} = O_{|C|}(k)$. Observe that the Bass Theorem is the specialization of Theorem~\ref{BasisBassT} to $G=\GEN{g}$, $a_{\GEN{g^d}}=g^d$ for $d$ dividing $|g|$, and $m_{k,\GEN{g^d}}$ a fixed multiple of $\varphi(|g|)$. On the other hand, using the Bass and the Bass-Milnor Theorem one can easily prove that the units of Theorem~\ref{BasisBassT} generate a subgroup of finite index in $\U(\Z G)$. Indeed, by the Bass-Milnor Theorem, the Bass units generate a subgroup of finite index of $\U(\Z G)$. On the other hand, by the Bass Theorem, if $\GEN{g}=C=\GEN{a_C}$, then a power of $\Bass{k}{m}{g}$ belongs to the group generated by the units of the form $\Bass{l}{\varphi(|a_C^j|)}{a_C^j}$, with $1 < l < \frac{|a_C^j|}{2}$ and $\gcd(l,|a_C^j|)=1$. Thus another power belongs to the group generated by the units of the form $\Bass{l}{m_{l,C}}{a_C^j}$. The advantage of the new proof, with respect to the proofs of Bass and Bass-Milnor, is that it provides a way to express some power of any given Bass unit $\Bass{k}{m}{g}$ as a product of a trivial unit and powers of at most 2 elements from $$B=\left\{\Bass{k}{m_{k,C}}{a_C}:1<k\le\frac{|C|}{2}, \gcd(k,|C|)=1, C\in \mathcal{C}\right\}$$ for any given choice of generators $a_C$ of cyclic subgroups and integers $m_{k,C}$ as in Theorem~\ref{BasisBassT}. This is obtained as follows: calculate $\bullet$ $n:=|g|; \; C:=\GEN{g};$ $\bullet$ $k'_1:=$ the unique integer $0\le k'_1 < n$ such that $g=a_C^{k'_1};$ $\bullet$ $k'_0:= kk'_1 \mod n$; $\bullet$ for $i=0,1: \; k_i := \min(k'_i,n-k'_i); \; h_i := \left\{\begin{array}{ll} 1, & \text{if } k_i=k'_i; \\ a_C^{k'_i}, & \text{otherwise}. \end{array}\right.$ $\bullet$ $M:=\lcm\left(m,m_{k_0,C},m_{k_1,C}\right); \; c := \frac{M}{m};$ Then, by (\ref{Basseq3}), (\ref{Basseq6}) and (\ref{Basseq7}) we have \begin{eqnarray*} \Bass{k}{m}{g}^c &=& \Bass{k}{M}{a_C^{k'_1}} = \Bass{k'_0}{M}{a_C} \Bass{k'_1}{M}{a_C}^{-1} = \Bass{k_0}{M}{a_C} h_0^M \Bass{k_1}{M}{a_C}^{-1} h_1^{-M} \\ &= & (h_0h_1\inv)^M \Bass{k_0}{m_{k_0,C}}{a_C}^{\frac{M}{m_{k_0,C}}} \Bass{k_1}{m_{k_1,C}}{a_C}^{-\frac{M}{m_{k_1,C}}}. \end{eqnarray*} We summarize this result in the following corollary. \begin{corollary}\label{BassInBasis} Let $G$ be a finite abelian group. For every cyclic subgroup $C$ of $G$ choose a generator $a_C$ of $C$ and for every $k$ coprime to the order of $C$ choose an integer $m_{k,C}$ with $k^{m_{k,C}}\equiv 1 \mod |C|$. Then $$\left\{\Bass{k}{m_{k,C}}{a_C}:C \text{ cyclic subgroup of } G, 1<k< \frac{|C|}{2}, \gcd(k,|C|)=1\right\}$$ is a basis for a free abelian subgroup of finite index in $\U(\Z G)$. Moreover, for any Bass unit $\Bass{k}{m}{g}$ in $\Z G$ we have $$\Bass{k}{m}{g}^{c}=h\; \Bass{k_0}{m_{k_0,C}}{a_C}^{n_0}\; \Bass{k_1}{m_{k_1,C}}{a_C}^{n_1},$$ for $C=\GEN{g}$, an element $h\in G$ and integers $c,n_0,n_1,k_0,k_1$ so that $1\le k_0,k_1 \le \frac{|C|}{2}$, $g=a_C^{\pm k_1}$ and $k_0\equiv \pm kk_1 \mod |C|$. \end{corollary} Observe that the factor $\Bass{k_i}{m_{k_i,C}}{a_C}$ in Corollary~\ref{BassInBasis} is 1 if $k_i=1$. Otherwise it is a basis element. The integers $k_0$ and $k_1$ are uniquely determined by the conditions imposed, namely $1\le k_0,k_1 \le \frac{|C|}{2}$, $g=a_C^{\pm k_1}$ and $k_0\equiv \pm kk_1 \mod |C|$. A possible choice for the integers $c,n_0$ and $n_1$ is $c=\frac{M}{m}$, $n_0=\frac{M}{m_{k_0,C}}$ and $n_1=-\frac{M}{m_{k_1,C}}$ with $M=\lcm(m,m_{k_0,C},m_{k_1,C})$. For some particular choices of the $m_{k,C}$'s one can take the same exponent $c$ for every Bass unit $\Bass{k}{m}{g}$. For example, if $n$ is the exponent of $G$ then one could choose a divisor of $\varphi(n)$ for each $m_{k,C}$. With this choice one can take $c=\varphi(n)$, $n_0= m\frac{\varphi(n)}{m_{k_0,C}}$ and $n_1= -m\frac{\varphi(n)}{m_{k_1,C}}$. \section{A basis formed by products of Bass units which are powers of cyclotomic units} In this section $p$ is a prime integer. We obtain a set of multiplicatively independent units that generate a subgroup of finite index in $\U(\Z G)$, for $G$ a $p$-group which is either elementary abelian or cyclic. Each of these units is both a product of Bass units and a power of a cyclotomic unit. In both cases we will use the fact that $\{\eta_k(\zeta_{p^n})\mid 1<k< \frac{p^n}{2}, p\nmid k\}$ generates a free abelian subgroup of finite index in $\U(\Z[\zeta_{p^n}])$ for each $n\ge 1$ (see \cite[Theorem 8.2]{1982Washington}). \begin{proposition}\label{ElementaryAbelian} Let $p$ be a prime integer and $G$ an elementary abelian $p$-group. For every $H\in \Hc(G)$ fix an $a_H\in G$ with $G=\GEN{a_H,H}$. Then the set $$\left\{x_{k,H}=\prod_{h\in H}\Bass{k}{O_p(k)}{a_Hh} : 1<k<\frac{p}{2}, H\in \Hc(G)\setminus\{G\}\right\}$$ is multiplicatively independent, generates a free abelian subgroup of finite index in $\U(\Z G)$ and each of its elements $x_{k,H}$ is a power of a cyclotomic unit of $\Q G$ and it belongs to $B(G)$. \end{proposition} \begin{proof} Clearly $\rho_G(\Z G)=\Z$ and if $H\in \Hc(G)\setminus \{G\}$ then $\rho_H(\Z G) = \Z[\zeta_p]$. Let $K\in \Hc(G)$. By Lemma~\ref{ProductBassCyclic}, $\rho_K(x_{k,H}) = \eta_k(\rho_K(a_H)^{[K:H\cap K]})^{O_p(k)|H\cap K|}$. Furthermore, $\rho_H(a_H)=\zeta_p$ and, if $K\ne H$ then $[K:H\cap K]=p$ and hence $\rho_K(a_H)^{[K:H\cap K]}=1$. Thus $$\rho_K(x_{k,H}) = \left\{\begin{array}{ll} \eta_k(\zeta_p)^{O_p(k)|H|}, & \text{if } K=H;\\ 1, & \text{otherwise}. \end{array}\right.$$ Since $\eta_k(\zeta_p):1<k<\frac{p}{2}$ are multiplicatively independent and generate a subgroup of finite index in $\U(\Z[\zeta_p])$, the result follows. \end{proof} \begin{proposition}\label{CyclicpGroup} Let $G=\GEN{a}$ be a cyclic group of order $p^n$ with $p$ prime. For every $i=0,1,\dots,n$, let $G_i=\GEN{a^{p^{n-i}}}$, the subgroup of $G$ of order $p^i$. Let $k$ be a positive integer coprime to $p$. For every $0\le j \le s \le n$ we construct recursively the following products of Bass units of $\Z G$: $$b_s^s(k)=1,$$ and, for $0\le j\le s-1$, $$b_j^s(k)= \left(\prod_{h\in G_j}\Bass{k}{O_{p^n}(k)}{a^{p^{n-s}}h}\right)^{p^{s-j-1}} \left( \prod_{i=j+1}^{s-1}b_i^s(k)\inv \right) \left(\prod_{i=0}^{j-1} b_i^{s+i-j}(k)\inv\right).$$ Then $$\left\{b_j^n(k) : 1<k<\frac{p^{n-j}}{2}, 0\le j\le n, p\nmid k\right\}$$ is multiplicatively independent, generates a free abelian subgroup of finite index in $\U(\Z G)$ and each $b_j^{n}(k)$ is a power of a cyclotomic unit of $\Q G$ and belongs to $B(G)$. \end{proposition} \begin{proof} Following the same strategy as in the proof of Proposition~\ref{ElementaryAbelian} it is enough to prove the following \begin{equation}\label{rhobs} \rho_{G_{j_1}}(b_j^s(k)) = \left\{ \begin{array}{ll} \eta_k(\zeta_{p^{s-j}})^{O_{p^n}(k)p^{s-1}}, & \text{if } j=j_1;\\ 1, & \text{if } j\neq j_1. \end{array}\right. \end{equation} for every $0\le j,j_1\le s\le n$. Again we use a double induction, first on $s$ and, for fixed $s$, on $s-j$ (roughly said, $s$ is a replacement for (IH1) and $s-j$ for (IH2)). The minimal cases ($s=0$ and $s=j$) are obvious, so assume that $s>1$, $j<s$ and (\ref{rhobs}) holds when $s$ is replaced by a smaller value $s_1$ and any $0\le j,j_1\le s_1$ and when, with $s$ fixed, $j$ is replaced by a bigger value. Let $u(k)=\prod_{h\in G_j}\Bass{k}{O_{p^n}(k)}{a^{p^{n-s}}h}$. By Lemma~\ref{ProductBassCyclic}, if $0\le i \le j$ then \begin{equation}\label{CyclicpGroup1} \rho_{G_i}(u(k)) = \eta_k(\rho_{G_i}(a^{p^{n-s}})^{p^{j-i}})^{O_{p^n}(k)p^i} = \eta_k(\zeta_{p^{s-i}}^{p^{j-i}})^{O_{p^n}(k)p^i}, \end{equation} and if $j\le i\le n$ then \begin{equation}\label{CyclicpGroup1b} \rho_{G_i}(u(k)) = \eta_k(\rho_{G_i}(a^{p^{n-s}}))^{O_{p^n}(k)p^j} = \eta_k(\zeta_{p^{s-i}})^{O_{p^n}(k)p^j}. \end{equation} By (\ref{CyclicpGroup1}) and the induction hypothesis on $s$, we have \begin{equation}\label{CyclicpGroup2} \rho_{G_{j_1}}(b_i^{s+i-j}(k)) = \left\{ \begin{array}{ll} \eta_k(\zeta_{p^{s-j}})^{O_{p^n}(k)p^{s+i-j-1}} = \rho_{G_{j_1}}(u(k))^{p^{s-j-1}},& \text{if } j_1=i<j;\\ 1, & \text{if } j_1\ne i<j. \end{array}\right. \end{equation} By (\ref{CyclicpGroup1b}) and the induction hypothesis on $s-j$, we have \begin{equation}\label{CyclicpGroup3} \rho_{G_{j_1}}(b_i^s(k)) = \left\{ \begin{array}{ll} \eta_k(\zeta_{p^{s-i}})^{O_{p^n}(k)p^{s-1}}=\rho_{G_{j_1}}(u(k))^{p^{s-j-1}}, & \text{if } j<i=j_1;\\ 1, & \text{if } j<i\neq j_1 \end{array}\right. \end{equation} Now, combining (\ref{CyclicpGroup1}), (\ref{CyclicpGroup2}) and (\ref{CyclicpGroup3}) we have $$\rho_{G_{j_1}}(b_j^s(k)) = \left\{ \begin{array}{ll} \rho_{G_j}(u(k))^{p^{s-j-1}} = \eta_k(\zeta_{p^{s-j}})^{O_{p^n}(k)p^{s-1}}, & \text{if } j=j_1;\\ \rho_{G_{j_1}}(u(k))^{p^{s-j-1}} \rho_{G_{j_1}}(b_{j_1}^s(k))^{-1} = 1,& \text{if } j< j_1;\\ \rho_{G_{j_1}}(u(k))^{p^{s-j-1}} \rho_{G_{j_1}}(b_{j_1}^{s+j_1-j}(k))^{-1} = 1,& \text{if } j> j_1; \end{array} \right.$$ as desired. \end{proof} \renewcommand{\bibname}{References} \bibliographystyle{amsalpha}
\section{Introduction} In spontaneous parametric down-conversion (SPDC), photon pairs are generated with several degrees of freedom quantum correlated. In particular, the down-converted photons are spatially entangled. Due to energy and momentum conservation, the sum of the transverse momenta and the difference of the transverse positions of the photons can be well defined, even though the position and momentum of each photon are undefined \cite{BoydEPR,ShiEPR}. This type of Einstein-Podolsky-Rosen (EPR) correlation \cite{EPR}, has been used as a resource for fundamental studies of quantum mechanics \cite{BoydEPR,ShiEPR,FonsecaBROGLIE,Yarnall,PR}, for quantum imaging \cite{Pittman,Gatti}, and experiments of quantum information \cite{LeoGen,BoydQu,BoydORBITAL}. The process of SPDC has been studied extensively in the past \cite{Mandel,Klyshko,Monken1}, and much of the recent effort has been put in the quantification of the spatial entanglement for a given experimental geometry. Traditionally, this has been done through the technique of Schmidt decomposition of the two-photon wave function, which gives the Schmidt number, $K$, and the Schmidt modes allowed for each photon \cite{Eberly,Monken2,Torres,Kulik2}. While this approach describes important properties of the spatial correlation, it basically gives no information of the form of the spatial joint distributions, therefore giving no information about the Gaussianity of the spatial two-photon state of SPDC. Moreover, it is well known that the state of the down-converted photons depends on the phase-matching conditions. Nevertheless, due to its complex structure, the phase-matching functions are usually approximated by Gaussian functions \cite{Torres,Kulik2,Ether,Paulao1,Paulao2,LeoCAM}. While the approach being adopted seems to be reasonable, it has already been shown that fine (new) structures in the spatial distributions of these photons can be observed due to (the manipulation of) the phase-matching conditions \cite{ExterNF1}. In this work we study the non-Gaussianity of the spatial two-photon state of SPDC by properly taking into account the phase-matching conditions. We start by showing how the variances of the near- and far-field conditional probability distributions are affected by the phase-matching functions. Then, we analyze the role of the EPR-criterion \cite{EPR,Reid} regarding the non-Gaussianity and entanglement detection of this state. Even though it has been demonstrated that higher order separability criteria can be used for the entanglement detection of spatial non-Gaussian entangled states \cite{SteveNG}, we show that a proper consideration of the phase-matching function reveals, precisely, when the \emph{simpler} EPR-criterion can still be used for the spatial entanglement detection. We also show that (and when) the EPR-criterion can be used as a witness for the non-Gaussianity of this state. Furthermore, we introduce a statistical measure, based on the \emph{negentropy} \cite{Negentropy} of the near- and far-field joint distributions, which allows for the quantification of the non-Gaussianity of the spatial two-photon state of SPDC. This measure does not correspond to a quantum mechanical generalization of the negentropy, such as the non-Gaussianity measure based on the quantum relative entropy (QRE) \cite{Banaszek,Paris}, and so does not require the knowledge of the full density matrix. Only the moments associated with the diagonal sub-matrices of the covariance matrix need to be measured. Thus, it is experimentally more accessible \cite{Eisert}. Moreover, for most of the configurations used so far, we show that its value can be estimated from the (easier to measure) marginal and conditional distributions. We also demonstrate that it has common properties with previous introduced measures of non-Gaussianity \cite{Banaszek,Paris}. The quantification of the non-Gaussianity of a quantum state has important applications for quantum information \cite{SteveNG,Banaszek,Paris}, and thus the practicality our measure shall be relevant for new applications of the spatial correlations of SPDC in this field. \section{The phase-matching conditions and the variances of the conditional probabilities} We consider the process of quasi-monochromatic SPDC, in the paraxial regime, for configurations with negligible Poynting vector transverse walk-off, that can be obtained using non-critical phase-matching techniques \cite{ExterNF1}. In the momentum representation, the two-photon state is given by \cite{Mandel,Klyshko,Monken1} \begin{equation} \label{EstMon} |\Psi\rangle\propto \int\!\!\!\int d\brm{q}_1 d\brm{q}_2\ e^{-\frac{\zeta}{4}|\brm{q}_1+\brm{q}_2|^2} \ensuremath{\mbox{\hspace{1.3pt}sinc}\,} \frac{L|\brm{q}_1 -\brm{q}_2|^2}{4k_p} \ket{\brm{q}_1,\brm{q}_2}, \end{equation} where $\ket{\brm{q}_1,\brm{q}_2}$ represents a two-photon state in plane-wave modes whose transverse wave vectors are $\brm{q}_1$ and $\brm{q}_2$. $L$ is the crystal length, $k_p$ is the pump beam wave number, $\zeta=w_0^2-2iz/k_p$, and $w_0$ is the pump beam waist, which is located at $z=0$. This state may be rewritten in the coordinate space as \cite{ExterNF1,ExterNF2} \begin{equation} \label{EstPos} |\Psi\rangle\propto\int\!\!\!\int d\boldsymbol{\rho}_1 d\boldsymbol{\rho}_2\ e^{-\frac{1}{4\zeta}|\brm{\rho}_1+\brm{\rho}_2|^2} \ensuremath{\mbox{\hspace{1.3pt}sint}\,} \frac{k_p|\brm{\rho}_1 -\brm{\rho}_2|^2}{4L} \ket{\brm{\rho}_1,\brm{\rho}_2}, \end{equation} where we define the function $\ensuremath{\mbox{\hspace{1.3pt}sint}\,}$ as $\ensuremath{\mbox{\hspace{1.3pt}sint}\,}(x)\equiv \frac{2}{\pi}\int_x^{\infty} dt\,\ensuremath{\mbox{\hspace{1.3pt}sinc}\,} t\equiv 1-\frac{2}{\pi}\ensuremath{\mbox{\hspace{1.3pt}Si}\,}(x)$, $\ensuremath{\mbox{\hspace{1.3pt}Si}\,}(x)$ being the sine integral function. The functions $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}(b\,|\brm{q}|^2/2)$ and $\frac{\pi}{2b}\ensuremath{\mbox{\hspace{1.3pt}sint}\,}(|\brm{\rho}|^2/2b)$ form a Fourier transform pair. The $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}$ and $\ensuremath{\mbox{\hspace{1.3pt}sint}\,}$ functions arise from the phase-matching conditions, and due to the difficulty of dealing analytically with them, they are usually approximated by Gaussian functions. The approach that has been adopted consists in approximating the function $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}(bx^2)$ by $e^{-\alpha b x^2}$. Sometimes it is used that $\alpha=1$ \cite{Eberly,Ether}, and in other cases the value of $\alpha$ is chosen such that both functions coincide at $1/e^2$ \cite{Torres,Paulao1,Paulao2} or at $1/e$ \cite{LeoCAM} of their peak. While this approximation seems to be reasonable for the entanglement quantification \cite{Eberly}, there has been no investigation to determine how precise it is for describing the distribution of the momentum correlations. Besides, it should be noticed that once a Gaussian approximation is adopted for the $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}$ function, the corresponding approximation for the $\ensuremath{\mbox{\hspace{1.3pt}sint}\,}$ function is already defined by the Fourier transform. Therefore, it is also not clear that such approximation is indeed good for describing the position distributions. Thus, it is not clear whether the SPDC two-photon state can indeed be written as a two-mode Gaussian state. \begin{figure}[t] \centering \includegraphics[width=0.7\textwidth]{ngaussFig1.eps} \caption{The far- and near-field conditional probability distributions, while considering the state of SPDC ($p_{FF}^{S}$ and $p_{NF}^{S}$) and when some Gaussian approximations are assumed ($p_{FF,\alpha_{i}}^{G}$ and $p_{NF,\alpha_{i}}^{G}$). In ($a$)-($d$) the curves $p_{FF}^{S}$ and $p_{NF}^{S}$ (red solid lines) are compared with $p_{FF,\alpha_{2}}^{G}$ and $p_{NF,\alpha_{2}}^{G}$ for $\alpha_1=0.45$ (green dashed line), $\alpha_2=0.72$ (blue dot-dashed line), and for some specific values of $P$. In ($e$) and ($f$) the variances of the momentum and position conditional probabilities are plotted in terms of $P$, respectively. $\alpha_3=1$.} \label{Fig1} \end{figure} To investigate this point we study how the variances of the momentum (far-field) and position (near-field) conditional probability distributions are affected by the phase-matching function, and compare the obtained results with the cases where Gaussian approximations are considered. Due to the symmetry of the two-photon wave functions, there is no loss of generalization if we work in one dimension (i.e., $y_1=y_2=0$ and $q_{y1}=q_{y2}=0$). To simplify our analysis we define the following dimensionless variables: $\tilde{x}_j={x}_j/w_0$ and $\tilde{q}_j=w_0{q}_{j}$, $j=1,2$. The probabilities for coincidence detection at the far- and near-field planes are \begin{eqnarray} p_{FF}^S({\tilde{q}}_1,{\tilde{q}}_2)&\propto& e^{-\frac{1}{2}({\tilde{q}}_1+{\tilde{q}}_2)^2}\, \left(\ensuremath{\mbox{\hspace{1.3pt}sinc}\,} \frac{P^2({\tilde{q}}_1 -{\tilde{q}}_2)^2}{4}\right)^2 ,\label{pff}\\ p_{NF}^S({\tilde{x}}_1,{\tilde{x}}_2)&\propto& e^{-\frac{1}{2\sigma^2}({\tilde{x}}_1+{\tilde{x}}_2)^2}\left(\ensuremath{\mbox{\hspace{1.3pt}sint}\,}\frac{({\tilde{x}}_1-{\tilde{x}}_2)^2}{4P^2}\right)^2\!,\label{pnf} \end{eqnarray} where $\sigma^2=(z_0^2+z^2)/z_0^2$, $P=\sqrt{L/(2z_0)}$, and $z_0=k_pw_0^2/2$ is the diffraction length of the pump beam. The joint probabilities are related with the conditional (and marginal) probabilities $p_{FF}^S({\tilde{q}}_1|{\tilde{q}}_2)$ [$p_{FF}^S({\tilde{q}}_j$)] and $p_{NF}^S({\tilde{x}}_1|{\tilde{x}}_2)$ [$p_{NF}^S({\tilde{x}}_j$)], through the rule: $p({\xi}_1|{\xi}_2)=\frac{p({\xi}_1,{\xi}_2)}{p({\xi}_2)}$. In the case of the Gaussian approximations discussed above, the probabilities of coincidence detection are \begin{eqnarray} p_{FF,\alpha_{i}}^{G}({\tilde{q}}_1,{\tilde{q}}_2) &\propto& e^{-\frac{1}{2}({\tilde{q}}_1+{\tilde{q}}_2)^2}\,e^{-\alpha_{i}\frac{P^2}{2}({\tilde{q}}_1-{\tilde{q}}_2)^2},\\ p_{NF,\alpha_{i}}^{G}({\tilde{x}}_1,{\tilde{x}}_2) &\propto& e^{-\frac{1}{2\sigma^2}({\tilde{x}}_1+{\tilde{x}}_2)^2}e^{-\frac{1}{2\alpha_{i}P^2}({\tilde{x}}_1-{\tilde{x}}_2)^2}, \end{eqnarray} where different values of $\alpha_i$ represent distinct Gaussian approximations for the $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}$ function. In Figure \ref{Fig1}(a) [(b)] we compare the curves $p_{FF}^{S}$ and $p_{FF,\alpha_{i}}^{G}$ ($p_{NF}^{S}$ and $p_{NF,\alpha_{i}}^{G}$) considering $\tilde{x}_2,\tilde{q}_2=0$, $\sigma=1$ (and the crystal centered at $z=0$), for the case where the dimensionless parameter $P=0.1$. This parameter has been used in the study of the quantification of the spatial entanglement \cite{Eberly,Monken2}, and it brings universality to the theory since a certain value of $P$ can be reached in three different ways. Here we considered the values of $L$ and $k_p$ as fixed parameters such that $P$ varies with $w_0$. In this case we find $\alpha_1=0.45$ ($\alpha_2 =0.72$) for the case where the $\ensuremath{\mbox{\hspace{1.3pt}sinc}\,}$ and Gaussian functions coincide at $1/e$ ($1/e^2$). From Figure \ref{Fig1}(a)-(b) one can see that for a small value of $P$, the Gaussian approximation only describes properly the momentum conditional distribution. The position conditional distribution is barely described by the approximation. In Figure \ref{Fig1}(c) and Figure \ref{Fig1}(d) we have the same type of analysis but now for a larger value of $P$. In this case, the Gaussian approximation is useful only for describing the position conditional distribution. The overall behavior of the Gaussian approximations is showed in Figure \ref{Fig1}(e) and Figure \ref{Fig1}(f), where the normalized variances of the far- [$(\Delta q_1|_{q_2})^2 L/k_p$] and near-field [$(\Delta x_1|_{x_2})^2 k_p/L$] conditional distributions are plotted in terms of $P$. \section{The EPR-Criterion as a witness for the non-Gaussianity of the spatial two-photon state} The near and far-field conditional probabilities can be used for implementing the EPR-paradox \cite{BoydEPR,EPR,Paulao2,Reid}. This is done by observing the violation of the inequality $(\Delta x_1|_{x_2})^2 (\Delta q_1|_{q_2})^2 \geq \frac{1}{4}$, which certifies the quantum nature of the spatial correlations of the down-converted photons. Since we have determined how the phase-matching function affects the variances $(\Delta x_1|_{x_2})^2$ and $(\Delta q_1|_{q_2})^2$, we can also look for its effect on the EPR-criterion. This is showed with the red (solid) line in Figure \ref{Fig2}, which was calculated for the values of $\tilde{x}_2$ and $\tilde{q}_2$ at the origin. For smaller values of $P$, the conditional variances are independent of the $\tilde{x}_2$ and $\tilde{q}_2$ values \cite{BoydEPR}. Whenever $P$ increases, the variances become dependent on $\tilde{x}_2$ and $\tilde{q}_2$. Nevertheless for smaller values of $\tilde{x}_2$ and $\tilde{q}_2$, which are of most experimental relevance, the red (solid) curve shown in Figure \ref{Fig2} captures the overall behavior of the product of $(\Delta x_1|_{x_2})^2$ and $(\Delta q_1|_{q_2})^2$ for the state of Eq~(\ref{EstMon}). As we can see, for values of $P$ smaller than $0.56$ or greater than $2.58$, the EPR-criterion can safely be used for detecting the spatial entanglement of the two-photon state of SPDC. Besides of being useful as a entanglement witness, the EPR-criterion can also be used as a witness for the non-Gaussianity of the spatial state of SPDC. This emphasizes another application for this criterion, which has been related already with other quantum information tasks \cite{Reid}. To observe this, we note that for a pure two-mode Gaussian state it is possible to show that $(\Delta x_1|_{x_2})^2 (\Delta q_1|_{q_2})^2 = \frac{1}{4K}$ [see Appendix \ref{SuplementoA}], such that $\frac{1}{4}$ is an upper bound for the EPR-criterion with these states. Since it has been demonstrated in \cite{Eberly} that the two-photon state of Eq.~(\ref{EstMon}) is always entangled, we can say whenever the product of variances is greater than $\frac{1}{4}$, that it witnesses the non-Gaussianity of the entangled spatial two-photon state of SPDC. As one can see in Figure \ref{Fig2}, this happens for $0.56 \leq P \leq 2.58$. The Schmidt decomposition of the state wave function, used together with the EPR-criterion value, reveals the non-Gaussianity of a two-mode entangled state. Such observation does not necessarily hold true when other second-order moments criteria are considered. This is shown in Appendix \ref{SuplementoB} for the criterion of Ref. \cite{Mancini}, and for the states considered in Figure \ref{Fig2}. \begin{figure}[h] \centering \includegraphics[width=0.7\textwidth]{ngaussFig2.eps} \caption{The EPR-criterion plotted in terms of $P$. The red (solid) curve corresponds to the EPR values of the two-photon state generated in the SPDC [Eq.~(\ref{EstMon})]. The green (dotted), blue (dot-dashed) and black (dotted) curves describe the EPR-criterion for the two-photon Gaussian states defined in terms of $\alpha_1$, $\alpha_2$ and $\alpha_3$, respectively.} \label{Fig2} \end{figure} \section[Quantifying the non-Gaussianity of the spatial two-photon state...]{Quantifying the non-Gaussianity of the spatial two-photon state of SPDC} From our previous analysis it is clear that the spatial state of SPDC can not be seen as a two-mode Gaussian entangled state, even when it is generated with a Gaussian pump beam and in the case of perfect phase-matching. We now proceed to quantifying the non-Gaussianity of this state. First, we introduce the concept of \emph{negentropy} which is the base of our approach \cite{Negentropy}. The negentropy of a probability density function $p(\xi_1,\xi_2)$ is defined as $ N\equiv H[p^{\tilde{G}}(\xi_1,\xi_2)]-H[p(\xi_1,\xi_2)],$ where $p^{\tilde{G}}(\xi_1,\xi_2)$ is a Gaussian distribution with the same expected values and covariance matrix of $p(\xi_1,\xi_2)$. The function $H[p(\xi_1,\xi_2)]$, called \emph{differential entropy}, is defined as $H[p(\xi_1,\xi_2)]\equiv-\!\int\! d\xi_1 d\xi_2\,p(\xi_1,\xi_2)\log_2p(\xi_1,\xi_2)$ \cite{Negentropy2}. The advantage of using negentropy is that it can be seen as the optimal estimator of non-Gaussianity, as far as density probabilities are involved. This is due to the properties that it is always non-negative, and that it is zero only for Gaussian distributions. Besides, it is invariant under invertible linear transformations \cite{Comon}. Motivated by these properties we define the total non-Gaussianity of the spatial two-photon state of SPDC as \begin{equation} \label{ngT} nG^T\equiv N[p_{FF}^S({\tilde{q}}_1,{\tilde{q}}_2)]+ N[p_{NF}^S({\tilde{x}}_1,{\tilde{x}}_2)], \end{equation} where $N[p_{FF}^S({\tilde{q}}_1,{\tilde{q}}_2)]$ and $N[p_{NF}^S({\tilde{x}}_1,{\tilde{x}}_2)]$ are the negentropies of the far- and near-field joint distributions of Eq.~(\ref{pff}) and Eq.~(\ref{pnf}), respectively. In Appendix \ref{SuplementoC} we give, explicitly, the calculations of $N[p_{FF}^S({\tilde{q}}_1,{\tilde{q}}_2)]$ and $N[p_{NF}^S({\tilde{x}}_1,{\tilde{x}}_2)]$. It is possible to observe that $nG^T=0$ if and only if $|\Psi\rangle$ is a two-mode Gaussian state [see Appendix \ref{SuplementoD}]. Otherwise $nG^T>0$. We obtain that $N[p_{FF}^S({\tilde{q}}_1,{\tilde{q}}_2)] \approx 0.15$ and that $N[p_{NF}^S({\tilde{x}}_1,{\tilde{x}}_2)] \approx 0.22$. Thus, the total non-Gaussianity of the spatial two-photon state of SPDC [Eq~(\ref{EstMon})] is $nG^T\approx0.37$. It is interesting to note that it does not depend on $P$. This was expected since the phase-matching functions do not change their functional form when $P$ varies. The calculations performed for these negentropies can be adapted for different experimental geometries, or used for the proper determination of other quantum information quantities related with the differential entropy of the spatial joint distributions \cite{Howell}. The fact that $nG^T\neq0$, emphasizes that spatial Gaussian approximations should be taken carefully due to the \textit{extremality} of Gaussian states \cite{Cirac}. For comparison purposes, we calculated in Appendix \ref{SuplementoE} the value of $\delta_B$, which is the measure of non-Gaussianity based on the QRE \cite{Banaszek,Paris}. We obtain that $\delta_B=1.08$ and such result also highlight the non-Gaussian character of the state of Eq.~(\ref{EstMon}). Furthermore, it has the same behavior of $nG^T$, since it does not depend on the parameter $P$. In analogy with Eq.~(\ref{ngT}), we define the non-Gaussianity of the conditional and marginal distributions as: $nG^C\equiv N[p^S_{FF}(q_i|q_j)]+N[p^S_{NF}(x_i|x_j)]$ and $nG^M \equiv N[p^S_{FF}(q_i)]+N[p^S_{NF}(x_i)]$ with $i,j=1,2$ and $i \neq j$. According to these definitions, one can observe that the non-Gaussianity of the spatial state of SPDC decreases under partial trace, such that $nG^T > nG^M$; and that it is additive when the composite system is represented by a product state, i.e., if $\ket{\Psi}$ is a product state, then $nG^T=2 nG^M$ [see Appendix \ref{SuplementoD}]. These are common properties with the QRE measure of \cite{Banaszek,Paris}. In Figure \ref{Fig3}(a) we plot the negentropies $N[p^S_{FF}(q_1|q_2)]$ and $N[p^S_{NF}(x_1|x_2)]$ as a function of $P$. It is interesting to note that these curves quantify the idea already presented in Figure \ref{Fig1}(e)-(f). As larger the value of $P$ is, the less the conditional momentum distribution can be approximated by a Gaussian function. On the other hand, the conditional position probabilities tends to a normal distribution when $P$ increases. In Figure \ref{Fig3}(b) we plot the negentropies of the near- and far-field marginal probabilities. One can see that they have a different dependence on $P$ in comparison with the conditional probabilities. Now, the near-field distribution tends to a Gaussian function for larger values of $P$, and the far-field one for smaller values of $P$. In Figure \ref{Fig3}(c) and Figure \ref{Fig3}(d) we have $nG^C$ and $nG^M$ plotted in terms of $P$. The insets of these figures show the corresponding near- and far-field distributions at the points of minimum, indicated with red circles. \begin{figure}[t] \centering \includegraphics[angle=-90,width=0.8\textwidth]{Fig3.eps} \caption{In (a)[(b)] the negentropies of the near- and far-field conditional (marginal) distributions are plotted in terms of $P$. In (c) [(d)] we show the non-Gaussianity of the conditional (marginal) distributions.} \label{Fig3} \end{figure} As it is shown in Appendix \ref{SuplementoF} [See also Figure \ref{Fig3}(c) and Figure \ref{Fig3}(d)], in the limit of $P\ll1$, one can use the relation between the joint and conditional density probabilities to decompose $nG^T$ as the sum of $nG^C$ and $nG^M$: \begin{equation} \label{ngt2} nG^T \approx nG^C+nG^M \ \ (\textrm{iff}\ P \ll 1). \end{equation} In a typical experimental configuration for SPDC, where the pump beam spot size is around $1$~mm at the crystal plane, the value of $P$ can be smaller than $0.05$. Thus, in general, the total non-Gaussianity of the spatial two-photon state of SPDC can be estimated in terms of the (easier to measure) near- and far-field negentropies of the conditional and marginal distributions. This simplify the measurement of $nG^T$, since there is no need to scan the whole transverse planes associated with the near- and far-field joint distributions. \section{Conclusion} We have investigated the spatial distributions of the entangled down-converted photons by proper considering the phase-matching conditions. By understanding how the near- and far-field conditional distributions are affected by the phase-matching function, we could show that the EPR-criterion \cite{EPR} can be used as a witness for the non-Gaussianity of the spatial state of the SPDC. The work culminated in the quantification of the non-Gaussianity of this state, which was based in a new and very experimentally accessible measure. As it has been discussed in \cite{Banaszek,Paris}, the quantification of the non-Gaussianity of a quantum state has many applications in the area of continuous variables quantum information processing. Thus, we envisage the use of our result for new applications of the spatial correlations of SPDC.
\section{Introduction} Nanosystems of interacting magnetic moments have attracted much interest due to ongoing technological advances in the field of magnetic data-storage devices. With decreasing system size, ferromagnetic particles on solid surfaces \cite{Wie09} or ferromagnetic molecular magnets like Fe$_8$ and Mn$_{12}$, \cite{Sch04} for example, are found in a single-domain state. The question of the stability of the ferromagnetic state against different kinds of thermal and quantum fluctuations is a crucial technological issue which at the same time provides serious challenges to a theoretical modeling and understanding. The magnetic properties of a ferromagnetic nanoparticle can be described by a Heisenberg model \begin{equation} H_J = - \frac{1}{2} \sum_{ij} J_{ij} \ff s_i \ff s_j \labeq{ham} \end{equation} with positive exchange coupling $J_{ij}>0$ between microscopic spins $\ff s_i$ and $\ff s_j$. The spins give rise to magnetic moments which are assumed to be localized at the sites $i=1,...,L$. The latter may constitute a $d$-dimensional lattice of finite size with $L$ sites in total and constant exchange $J=J_{ij}$ between nearest neighbors $i$ and $j$ only. In a ground state of the model all spins are perfectly aligned, and the particle is ferromagnetic. However, this state is unstable against thermal fluctuations. Strictly speaking, for a system of finite size, the ferromagnetic state is destroyed at any finite temperature by fluctuations originating from a coupling of the system to an external heat bath. On the other hand, anisotropic contributions to the Hamiltonian, such as a single-site anisotropy of the form \begin{equation} H_D = - D \sum_{i} s^2_{iz} \; , \labeq{hamani1} \end{equation} which for $D>0$ distinguishes the $z$-axis as the easy axis, stabilize the ferromagnetic state. Inclusion of $H_D$ leads to a model with only two degenerate ground states in which all spins are pointing into the $+z$ or $-z$ direction, respectively. The anisotropic ferromagnetic Heisenberg model $H=H_J+H_D$ in fact represents the simplest model to study magnetization reversal, i.e.\ the thermally activated transition between two ground states across an energy barrier induced by the anisotropy. \cite{Wer01} The rate of magnetization switching and thus the magnetic stability obviously crucially depends on the height of the anisotropy-energy barrier $\Delta E$. \begin{figure}[b] \includegraphics[width=0.95\columnwidth]{fig01.eps} \caption{ Two different models for magnetization reversal. A: Coherent rotation. B: Nucleation and domain-wall propagation. Figure inspired by Ref.\ \onlinecite{KHS+09}. } \label{mechs} \end{figure} For weak anisotropy $D \ll J$, it is self-evident to consider the spins as tightly coupled by the exchange interaction and forming a huge macrospin, \begin{equation} \ff S = \sum_i \ff s_i \: , \labeq{macrospin} \end{equation} the dynamics of which can be approximated by an effective model \begin{equation} H_{\rm macro} = - D_{\rm macro} S^2_{z} + \mbox{const.} \: . \labeq{hammacro} \end{equation} with an anisotropy barrier \begin{equation} \Delta E = D_{\rm macro} S^2 \: . \end{equation} In this macrospin model, magnetization reversal takes place at zero temperature by suppression of the barrier due to an external magnetic field as described by the Stoner-Wohlfarth model. \cite{SW48} At finite temperature $T$, reversal may be caused by thermal activation. For low $T$, the switching frequency is exponentially small and a ferromagnetic state is stable over macroscopically relevant times while for high $T$, in the so-called superparamagnetic state, the magnetization vanishes on the time scale of the measurement. Classical theory \cite{Nee49,Bro63} for superparamagnetic dynamics based on the Landau-Lifshitz equation \cite{LL35} and including a stochastic Langevin field to simulate a thermal bath leads to an Arrhenius law for the thermal activation rate $\Gamma \propto \exp(- \beta \Delta E)$. This Neel-Brown law again emphasizes the role of the energy barrier $\Delta E$ for the magnetic stability of the particle. For stronger anisotropy, the macrospin approximation is no longer applicable as magnetization reversal preferably takes place by nucleation and domain-wall propagation (see Fig.\ \ref{mechs}). Here, classical many-spin models \cite{KG05} are frequently used to study the reversal mechanism and in particular the transition from coherent rotation to nucleation and domain-wall propagation with increasing strength of the anisotropy. In particular, classical Monte-Carlo simulations are employed \cite{HN98b} where Monte-Carlo update steps are related to physical time propagation. As a function of the model parameters, reversal mechanisms different from coherent rotation, can be identified and ``phase boundaries'' can be found in this way. The strong influence of the shape of the nanoparticles on the switching rate, as observed experimentally, \cite{BPKW04,KHS+09} can be explained by theoretical analysis of domain-wall propagation. \cite{Bra93,Bra94} The macrospin has a high spin quantum number $S$ and may be treated to a good approximation classically. There are, however, different types of quantum effects to be considered: For example, quantum tunneling as a reversal mechanism is competing with thermally induced reversal. \cite{ADiVS92,WS99,DeRHD+02} In case of not too small systems, tunneling is significant only for systems extremely isolated from any dissipative environment at ultralow temperatures. For a macrospin coupled to a conduction band, Kondo screening is another issue. Since $S>1/2$ there is an underscreened Kondo effect, the development of which is hampered by the anisotropy as has been shown by time-dependent numerical renormalization group. \cite{RWH08} As the quantum character of the macrospin is not accounted for by simple Langevin dynamics, \cite{Bro63} different approaches \cite{ZGP06,KCT08,BL05} have been suggested to treat the macro- or many-spin dynamics in contact with bosonic baths by means of quantum master-equation approaches and to determine, e.g., the blocking temperature. The importance of a many-spin model, for example, is demonstrated by studies of the probability density function of the macrospin obtained by replacing thermal with Markov processes. \cite{BL05} Differences between the classical-spin limit and large quantum spins have been discussed. \cite{Gar08} Here we reconsider the question of competing reversal mechanisms by taking a quantum many-spin model as the starting point. The competition between coherent rotation and nucleation can be addressed in a less ambitious way by studying the {\em static} properties of an anisotropic quantum Heisenberg model $H = H_J +H_{\rm ani}$. As the single-site anisotropy (\ref{eq:hamani1}), which is frequently studied in the classical context, cannot be used for the lowest spin $s=1/2$ in a quantum model since trivially $\ff s_i^2=\mbox{const}$, we also consider a coupling anisotropy of the form \begin{equation} H_\Delta = - \frac{1}{2} \sum_{ij} \Delta_{ij} s_{iz} s_{jz} \: , \labeq{hamani2} \end{equation} i.e.\ $H_{\rm ani} = H_D$ or $H_{\rm ani} = H_\Delta$. This model will be investigated by different complementary techniques including perturbation theory in $H_{\rm ani}$, exact diagonalization and the Lanczos approach for systems with small size $L$ and classical stability analysis in the high-spin limit. There are different goals of the present paper: As the status of the macrospin approximation is less clear in the quantum case, we aim at a strict derivation of the macrospin model in the weak-anisotropy limit. In this way it should be possible to relate the effective anisotropy strength of the macrospin to the parameters of the underlying many-spin model. The dependence on the spin quantum number $s$ should exhibit quantum effects for small $s$ and recover the classical result for $s \to \infty$. Furthermore, beyond the weak-anisotropy limit, the breakdown of the macrospin approximation and the competition between the different reversal mechanisms, depending on system size, coordination, dimensionality and on $s$, will be investigated. Our goal is to understand whether or not it makes a qualitative difference for the transition from coherent rotation to nucleation with increasing anisotropy strength if starting from a classical or from a quantum spin model. Finally, the high-spin limit is addressed as a reference and for comparison with the quantum-mechanical calculations. The paper is organized as follows: In the next section \ref{sec:ed} we briefly introduce the model, the basic concepts and notations by referring to exact-diagonalization results. Sec.\ \ref{sec:pert} present the results of first-order perturbation theory for weak anisotropy and the derivation of the macrospin model. Its limitations are discussed in Sec.\ \ref{sec:lim} on the basis of calculations employing the Lanczos technique. The dependence of the anisotropy-energy barrier on the spin-quantum number $s$ and classical stability analysis in the high-spin limit is addressed in Sec.\ \ref{sec:class}. Finally, Sec.\ \ref{sec:con} concludes the paper. \section{Exact-diagonalization results for the anisotropy barrier} \label{sec:ed} Due to the SU(2) spin-rotation symmetry of the Heisenberg model, \refeq{ham}, the total (or macro) spin, \refeq{macrospin}, is a conserved quantity: $[\ff S, H]=0$. We consider the eigenstates of $H_J$: \begin{equation} H_J | n, S, M \rangle = E_n (S) | n, S, M \rangle \; . \labeq{eigen} \end{equation} The states are classified according to their total spin quantum number $S=S_{\rm min}, ... , S_{\rm max}-1 , S_{\rm max}$ and their total magnetic quantum number $M=-S,...,S-1,S$. Here, $S_{\rm max}=Ls$ and $S_{\rm min}=0$ for $s$ even or $L$ even, while $S_{\rm min}=1/2$ otherwise. Throughout of the paper, we assume an even number of sites $L$ for the sake of simplicity. Further, $n$ labels the eigenstates in the invariant subspace with fixed $S$ and $M$. For a given $S$, the SU(2) invariance implies the eigenenergies $E=E_n(S)$ to be $2S+1$-fold degenerate and independent of $M$. For the present case of ferromagnetic exchange coupling $J_{ij} > 0$, the total spin quantum number is maximal, $S = S_{\rm max} = L s$, for a ground state. It is easy to see that the fully polarized state with all spins pointing into $+z$ direction, $| F \rangle \equiv | m_1 = s \rangle \otimes \cdots \otimes | m_L = s \rangle$, is a ground state. Assuming all exchange interactions to be equal between nearest neighbors, $J_{ij}=J >0$, the ground-state energy is given by $E_0(S_{\rm max}) = - (J/2) L s^2 \overline{z}$ where $\overline{z} = \sum_{i=1}^L z_i / L$ is the average of the coordination numbers $z_i$. The $2S_{\rm max}+1 = 2Ls+1$-fold degeneracy of the ground-state energy will partly be lifted by the anisotropy $H_{\rm ani}$, \refeq{hamani1} or \refeq{hamani2}. $H_{\rm ani}$ breaks the SU(2) symmetry, $[\ff S^2, H_{\rm ani}] \ne 0$, but preserves the rotational symmetry with respect to the easy axis ($z$ axis): $[S_z,H_{\rm ani}]=0$. The anisotropic eigenstates $|m, M \rangle$ are therefore characterized by $M$ and by an additional quantum number $m$ labeling the states in the subspace with fixed $M$. The corresponding eigenenergies $E_m(M)$ will depend on $M$ with \begin{equation} E_m(M) = E_m(-M) \; , \end{equation} resulting in a symmetric anisotropy barrier with a height $\Delta E$ given by the respective ground-state energies of the full Hamiltonian in the invariant subspaces with $M=0$ and $M=S_{\rm max}$: \begin{equation} \Delta E = E_0(M=0) - E_0(M=S_{\rm max}) > 0 \: . \end{equation} \begin{figure}[t] \includegraphics[width=0.9\columnwidth]{fig02.eps} \caption{ Energy eigenvalues of the isotropic $s=1/2$ Heisenberg model \refeq{ham} on a linear chain of $L=10$ sites and open boundaries as obtained numerically by exact diagonalization. Eigenvalues are classified according to the total magnetic quantum number $M=-S_{\rm max}, ..., S_{\rm max}$ with $S_{\rm max}=5$. The nearest-neighbor exchange coupling $J=1$ sets the energy scale. } \label{spectrum} \end{figure} \begin{figure}[b] \includegraphics[width=0.9\columnwidth]{fig03.eps} \caption{ The same as Fig.\ \ref{spectrum} but on enlarged energy scale close to the ground-state energy. The total spin quantum numbers of the different spin multiplets are indicated. } \label{spectrum1} \end{figure} \begin{figure}[t] \includegraphics[width=0.9\columnwidth]{fig04.eps} \caption{ The same as Fig.\ \ref{spectrum} but with the coupling-anisotropy term, \refeq{hamani2}, included. A constant anisotropy $\Delta_{ij} = \Delta$ between nearest neighbors is assumed. Calculations have been performed for $\Delta = J$. } \label{spectrumani} \end{figure} For systems with a moderately large Hilbert space, i.e.\ for small $s$ and small number of lattice sites $L$, the energy eigenvalue problem can be solved numerically. In case of the anisotropic problem $H_J + H_{\rm ani}$, conservation of $S_z$ can be exploited only. The dimension of the invariant subspace $\ca H(M)$ is given by: \cite{Fel68,Sch04} \begin{eqnarray} \dim \ca H(M) &=& \sum_{k=0}^{[(S_{\rm max} - M )/(2s+1)]} (-1)^k \left( \begin{array}{cc} L \\ k \\ \end{array} \right) \nonumber \\ &\times& \left( \begin{array}{cc} L-1+S_{\rm max}-M-(2s+1)k\\ L-1 \\ \end{array} \right) \: \nonumber \\ \end{eqnarray} where the binomial coefficients $\left( \begin{array}{cc} a \\ b \\ \end{array} \right) = 0$ for $b>a$, and where $[x]$ denotes the greatest integer with $[x]\le x$. For $s=1/2$ full diagonalization is easily possible for systems with $L=14$ spins. A simple example for a spin-$s=1/2$ Heisenberg chain with and without coupling anisotropy $\Delta$ on $L=10$ sites with open boundaries is given by Figs.\ (\ref{spectrum}), (\ref{spectrum1}) and (\ref{spectrumani}). The perturbed ($\Delta >0$) ground-state energy in the one-dimensional $M=S_{\rm max}$ subspace is easily calculated as $E_0(M=S_{\rm max}) = - (J/2+\Delta/2) L s^2 \overline{z}$. The $M=\pm(S_{\rm max}-1)$ subspaces are reached by a single spin flip, i.e.\ excitation of a magnon. As there must be as many magnons as lattice sites, the dimension of the subspaces is given by $L$ each. Their discrete energy spectrum can be seen in Figs.\ (\ref{spectrum}) and (\ref{spectrumani}) for $\Delta=0$ and $\Delta = J$, respectively. The low-energy sector of the isotropic spectrum for arbitrary $M$ is shown in Fig.\ \ref{spectrum1}: Ground states have the maximal total spin $S=S_{\rm max}$. The ground-state energy is $2S_{\rm max}+1$-fold degenerate. Excited states with $M=\pm(S_{\rm max}-1)$, corresponding to magnon excitation, have total spin $S=S_{\rm max}-1$. Fig.\ \ref{spectrumani} demonstrates the effect of an anisotropy term in the Hamiltonian. As the energy eigenstates are superpositions of states with different $S$ (but the same $M$) in this case, their energy becomes $M$ dependent. Consequently, an anisotropy barrier develops which, for a very strong anisotropy $\Delta=J$, is large as compared to the finite-size gap between the ground states and the first excited states. Still, there is a residual degeneracy of the two ground states $|F\rangle$ and $|-F\rangle$ in the subspaces $M=\pm S_{\rm max}$ which implies that thermal fluctuations destroy ferromagnetic order. However, switching between $|F\rangle$ and $|-F\rangle$ now requires that thermal fluctuations must overcome the barrier. The anisotropy thus leads to a superparamagnetic stabilization of the magnetic state on a certain time scale as described by the N\'eel-Brown model. \cite{Nee49,Bro63} \section{Perturbative derivation of the macrospin model} \label{sec:pert} In the weak-anisotropy limit, the different microscopic spins are tightly bound together by the ferromagnetic exchange coupling and form a huge macrospin which rotates from $+z$ to $-z$ direction, i.e.\ the microspins rotate coherently. To make this intuitive argument rigorous and quantitative, we derive the macrospin model as an effective low-energy model by means of first-order perturbation theory in $H_{\rm ani}$ in the following. This will result in a linear dependence of $\Delta E$ on the anisotropy strength. Non-trivial dependencies, however, may be expected with respect to $L$, $s$ and the matrix $J_{ij}$ of exchange-coupling constants. First-order perturbation theory requires to compute the matrix element $\langle S_{\rm max}, M' | s_{iz} s_{jz} | S_{\rm max}, M \rangle$ where the states $|S,M\rangle$ form an orthonormal common eigenbasis of $\ff S^2$ and $S_z$. Note that this includes the case of coupling ($i\ne j$) and single-site ($i=j$) single-site anisotropy. To find the action of $s_{iz} s_{jz}$ onto the state $| S_{\rm max}, M \rangle$ we write \begin{equation} | S_{\rm max}, M \rangle = \frac{\sqrt{(S_{\rm max} + M)!}}{\sqrt{(S_{\rm max}-M)! (2S_{\rm max})!}} S_-^{S_{\rm max}-M} | F \rangle \labeq{spinstate} \end{equation} in terms of the fully polarized state with all spins pointing in $+z$ direction, $| F \rangle \equiv| S_{\rm max}, M=S_{\rm max} \rangle = | m_1 = s \rangle \cdots | m_L = s \rangle$. Here $S_- = S_x - i S_y$. \refeq{spinstate} is obtained by straightforward spin algebra. Using the commutator \begin{eqnarray} [s_{iz}s_{jz}, S_-^k] &=& - k S_-^{k-1} (s_{i-}s_{jz} + s_{j-} s_{iz}) \nonumber \\ &+& k \delta_{ij} S_-^{k-1} s_{i-} \nonumber \\ &+& \frac{k(k-1)}{2} S_-^{k-2} (s_{i-} s_{j-} + s_{j-}s_{i-}) \end{eqnarray} where $s_{i\pm} = s_{ix} \pm is_{iy}$, we are left with matrix elements of the form $\langle S_{\rm max}, M' | \ca O | F \rangle$ where $\ca O$ contains microspin operators $s_{...}$ on the right-hand side acting on $|F\rangle$ as well as macrospin operators $S_{...}$ on the left-hand side which are considered to act on $\langle S_{\rm max}, M |$. The effect of the former operators is trivial. The effect of the latter ones is obtained with the help of the relation \begin{eqnarray} S_+^k | S_{\rm max}, M \rangle &=& \frac{\sqrt{(S_{\rm max} + M + k)! (S_{\rm max} - M)!}}{\sqrt{(S_{\rm max} + M)! (S_{\rm max} - M - k)!}} \nonumber \\ &\times& | S_{\rm max}, M + k \rangle \: . \end{eqnarray} Using \begin{equation} \langle S_{\rm max}, M | s_{i-} | F \rangle = \frac{2s}{\sqrt{2S_{\rm max}}} \end{equation} for $M=S_{\rm max}-1$ and \begin{eqnarray} \langle S_{\rm max}, M | s_{i-} s_{j-} | F \rangle &=& \frac{8s^2}{\sqrt{4S_{\rm max} (2S_{\rm max} -1)}} \quad (i\ne j) \nonumber \\ &=& \frac{4s(2s-1)}{\sqrt{4S_{\rm max} (2S_{\rm max} -1)}} \quad (i=j) \nonumber \\ \end{eqnarray} for $M=S_{\rm max}-2$, a straightforward calculation leads to \begin{eqnarray} && \langle S_{\rm max}, M | s_{iz} s_{jz} | S_{\rm max}, M' \rangle \nonumber \\ && = \delta_{MM'} \,\left[ s^2 + (2s-\delta_{ij}) \: \frac{M^2 - L^2 s^2}{2L^2 s - L} \right] \: . \labeq{corrf} \end{eqnarray} This is a remarkably simple result which may be used to compute the anisotropy energy barrier $E_0(M)$ in different situations. To give an example, we consider the model $H=H_J+H_{\rm ani}$ where a constant exchange interaction $J_{ij}=J$ between nearest neighbors is assumed (and likewise for $\Delta_{ij}$ in case of a coupling anisotropy). This implies that geometrical properties enter via the lattice topology only. For the case of the single-site anisotropy (\ref{eq:hamani1}), $s\ne 1/2$, we get: \begin{eqnarray} E_0(M) &=& - \frac{1}{2} JLs^2 \overline{z} - D \left( \frac{2s-1}{2Ls-1} M^2 + s^2 \frac{L(L-1)}{2Ls-1} \right) \nonumber \\ &+& \ca O(D^2) \: , \end{eqnarray} while for the coupling anisotropy (\ref{eq:hamani2}) with non-zero and constant $\Delta_{ij}=\Delta$ between nearest neighbors only, we find: \begin{eqnarray} E_0(M) &=& - \frac{1}{2} JLs^2 \overline{z} - \Delta \,\overline{z} \left( \frac{s}{2Ls-1} M^2 - \frac{1}{2} \frac{Ls^2}{2Ls-1} \right) \nonumber \\ &+& \ca O(\Delta^2) \: , \labeq{e0ofm} \end{eqnarray} where $\overline{z} = \sum_i z_i / L$ is the average coordination number. This corresponds to a macrospin model \refeq{hammacro} with \begin{equation} D_{\rm macro} = \frac{2s-1}{2Ls-1} D \labeq{macro1} \end{equation} for the single-site anisotropy, and \begin{equation} D_{\rm macro} = \frac{\overline{z}s}{2Ls-1} \Delta \labeq{macro2} \end{equation} in case of the coupling anisotropy. The anisotropy energy barrier is given by $\Delta E = D_{\rm macro} S_{\rm max}^2$. Eqs.\ (\ref{eq:macro1}) and (\ref{eq:macro2}) provide an explicit and quantitative relation between the anisotropy parameter of the effective low-energy macrospin model and the microscopic model parameters. Quite generally we note that this relation is non-trivial. In the limit $L\to \infty$ we find $D_{\rm macro} \propto 1/L$ formally. However, for fixed $D$ or $\Delta$, the macrospin approximation becomes less accurate with increasing system size $L$: Eventually the finite-size gap, which controls the validity of the perturbative treatment, is of the same order of magnitude as the anisotropy strength, i.e.\ the expansion parameter. The classical limit is reached for $s \to \infty$. Here, $D_{\rm macro} = D/L$ or $D_{\rm macro} = (\overline{z}/2) \Delta/L$, respectively. The validity of the macrospin approximation for the classical model is discussed below. \section{Limitations of the macrospin model} \label{sec:lim} It is easily seen that the limits of weak and of strong anisotropy correspond to two qualitatively different ways of how to overcome the barrier. Consider the model system discussed above but with a large coupling anisotropy $\Delta \gg J$. In the extreme case $J=0$, we recover the Ising model, $H=H_\Delta$, which exhibits a highly degenerate energy spectrum. It is displayed in Fig.\ \ref{ising}. Still there are two degenerate ground states in the $M=\pm S_{\rm max}$ sectors which are separated by an energy barrier. However, the barrier is entirely flat and thus qualitatively different from the quadratic trend given by the macrospin model. The tendency towards a flat barrier is already seen in Fig.\ \ref{spectrumani} for $\Delta=J$. The energy eigenstates are constructed trivially in the Ising limit. In the ground state with $M=S_{\rm max}$ the spins are aligned ferromagnetically, i.e.\ $|S_{\rm max} , S_{\rm max} \rangle = | \uparrow ,\uparrow, ... ,\uparrow \rangle$. Its energy is $E_0(M=S_{\rm max}) = - (\Delta/2) L s^2 \overline{z} = -2.25$ for the example shown in the figure. The ground state with $M=S_{\rm max}-1$ is found by flipping a single spin, i.e.\ $|S_{\rm max} , S_{\rm max}-1 \rangle = | \downarrow ,\uparrow, ... ,\uparrow \rangle$. The corresponding excitation energy is $\Delta E = 2 \Delta s^2 = 0.5$, resulting in $E_0(M=S_{\rm max}-1) = -1.75$. Shifting the ``domain wall'' to the right, i.e.\ $|S_{\rm max} , M \rangle = | \downarrow , ... , \downarrow, \uparrow, ... ,\uparrow \rangle$, is possible without further energy cost. Hence, nucleation and domain-wall propagation is the apparent mechanism for magnetization reversal in the Ising limit. \begin{figure}[t] \includegraphics[width=0.9\columnwidth]{fig05.eps} \caption{ The same as Fig.\ \ref{spectrum} but for $J=0$ and $\Delta=1$. } \label{ising} \end{figure} The breakdown of the macrospin approximation happens for still weak anisotropies but is gradual rather than abrupt. This is demonstrated by the exact-diagonalization results for the anisotropy-energy barrier $E_0(M)$ shown in Fig.\ \ref{barrier}. For strong anisotropy $\Delta = J$ (corresponding to Fig.\ \ref{spectrumani}), the macrospin approximation \refeq{e0ofm} completely fails and strongly overestimates the barrier height $\Delta E$. With decreasing $\Delta$, the macrospin approximation becomes more and more reliable. Still, at $\Delta = J/10$ deviations from the exact-diagonalization results are visible on the scale of Fig.\ \ref{barrier}. If interpreted within the classical theory, this is still significant as the energy barrier affects the thermal activation rate exponentially strong. \begin{figure}[t] \includegraphics[width=0.65\columnwidth]{fig06.eps} \caption{ Anisotropy barrier $E_0(M)$ as obtained from the macrospin model (\refeq{e0ofm}, lines) compared to the exact-diagonalization result (points) for three different anisotropy strengths as indicated. Results for an open chain of $L=10$ spins with $s=1/2$ and nearest-neighbor exchange and coupling anisotropy. } \label{barrier} \end{figure} \begin{figure}[b] \includegraphics[width=0.75\columnwidth]{fig07.eps} \caption{ Height of the anisotropy energy barrier $\Delta E$ as a function of the (coupling) anisotropy strength $\Delta$ for open chains of $L=10$ and $L=24$ spins with $s=1/2$ as obtained by the macrospin model (blue lines, \refeq{macro2}) and by the full model (red lines). Black line: $\Delta E$ as a function of $\Delta$ for the Ising limit with $J=0$ for both, $L=10$ and $L=24$. } \label{size} \end{figure} The fact that the breakdown of the macrospin model with increasing strength of the anisotropy is gradual rather than abrupt also implies a gradual change between the corresponding reversal mechanisms. Within a quantum model, there are no well-defined phase boundaries separating coherent rotation from nucleation and domain-wall propagation. This becomes obvious again in Fig.\ \ref{size} where the anisotropy energy barrier $\Delta E$ is shown as a function of $\Delta$. The macrospin and the exact-diagonalization results start to deviate from each other at arbitrarily weak $\Delta$, i.e.\ already at second order in the anisotropy strength. Nevertheless, a typical anisotropy strength can be identified that marks the smooth crossover from coherent rotation to nucleation or the qualitative breakdown of the macrospin approximation. Fig.\ \ref{size} shows that this crucially depends on the system size $L$. For $L=10$ spins (red lines), the macrospin model is valid up to much stronger $\Delta$ as compared to the case with $L=24$ spins (blue lines). Generally, with increasing $L$, substantial deviations from the linear trend of $\Delta E (\Delta)$ are found for weaker and weaker $\Delta$. In the strong-anisotropy limit, $\Delta E (\Delta)$ approaches a linear trend again which is given by the Ising limit (black line). Note that for the one-dimensional model considered here, the domain-wall energy and thus $\Delta$ in the strong-anisotropy limit is independent of the system size while in the macrospin or weak-anisotropy limit $\Delta E$ is roughly proportional to $L$ (see \refeq{e0ofm}). Calculations for $L=24$ spins with $s=1/2$ can no longer be carried out by full diagonalization. The present results have been obtained by employing the Lanczos algorithm. \cite{Lan50,LG93} This approximates the ground-state energy $E_0(M)$ of $H$ in the invariant subspace $\ca H(M)$ by the ground-state energy of $H$ in a Krylov space \begin{equation} \ca K_n(M) = \mbox{span} \{|i,M\rangle, H |i,M\rangle,..., H^{n-1} |i,M\rangle\} \subset \ca H(M) \end{equation} of dimension $n\ll \dim \ca H(M)$ that is spanned by the states $|i,M\rangle, H |i,M\rangle,..., H^{n-1} |i,M\rangle$ where $|i,M\rangle \in \ca H(M)$ is an arbitrary initial state. Typically, $n \approx 100$ Krylov-space dimensions are fully sufficient to obtain an excellent approximation which, on the scale of the figures, cannot be distinguished from the exact result. \begin{figure}[t] \includegraphics[width=0.85\columnwidth]{fig08.eps} \caption{ Height of the anisotropy energy barrier $\Delta E$ as a function of the (coupling) anisotropy strength $\Delta$ for systems of $L=24$ spins with $s=1/2$ as obtained by the macrospin model (\refeq{macro2}) and by the full model ($\Delta E$ in the full model is always smaller). Results for different planar geometries as indicated. $\Delta E$ for the fully connected model, where the macrospin approximation is exact, is shown for comparison (black line). } \label{coord} \end{figure} The extent of applicability of the macrospin model strongly depends on the system's geometry. It turns out to become more and more reliable with increasing average coordination number $\overline{z}$. This is demonstrated with Fig.\ \ref{coord} where again macrospin and exact-diagonalization (Lanczos) results for the $\Delta$ dependence of the energy barrier are compared. Here, the system size is kept fixed to $L=24$ while the system geometry varies from a one-dimensional chain to a compact two-dimensional array, i.e.\ the average coordination number $\overline{z}$ increases. The corresponding increasingly better agreement of the macrospin approximation with the exact-diagonalization results is easily understood by consideration of the extreme case (see black line in Fig.\ \ref{coord}): For $z_i=L-1=\overline{z}$, i.e. for the fully connected model with $J_{ij}=J$ and $\Delta_{ij}=\Delta$ for arbitrary pairs $i,j$, the Hamiltonian reduces to \begin{equation} H=-J \ff S^2 - \Delta \, S_z^2 + \mbox{const} \end{equation} and thus \begin{equation} E_0(M) = - \Delta \, M^2 + \mbox{const} \; , \end{equation} i.e.\ in this limit the macrospin approximation is trivially exact. \section{Critical anisotropy strength in the classical limit} \label{sec:class} The limit of the classical anisotropic Heisenberg model should be recovered for spin quantum number $s\to \infty$. Fig.\ \ref{class} shows the anisotropy barrier $\Delta E$ as a function of the coupling anisotropy strength $\Delta$ as obtained from the Lanczos technique for a small chain of $L=4$ sites. This permits a study of the spin model with quantum numbers up to $s=10$. To compare the results for different $s$, we consider rescaled parameters, \begin{eqnarray} J_{ij} &\mapsto& J_{ij} / s(s+1) \; , \nonumber \\ \Delta_{ij} &\mapsto& \Delta_{ij} / s(s+1) \; , \end{eqnarray} which corresponds to a rescaling $\ff s_i \mapsto \ff s_i / \sqrt{s(s+1)}$ of the spin variables. Disregarding quantum fluctuations, this amounts to a classical spin of unit length $|\ff s_i|=1$ in the limit $s\to \infty$. For $s=1/2$ we again note the smooth crossover from the macrospin limit and coherent rotation at weak $\Delta$ to the Ising limit and nucleation and domain-wall propagation at strong $\Delta$. In the quantum model the breakdown of the macrospin approximation starts immediately, i.e.\ the linear-in-$\Delta$ trend of the barrier is hardly visible as is apparent from the inset of Fig.\ \ref{class}. The crossover to the Ising limit takes place quickly. For $s=1$, the range of anisotropy strengths where an almost linear trend is found, increases while the crossover to the Ising limit is seen to be delayed. With increasing spin quantum number $s$, these changes continue in a systematic way and become more and more pronounced. From the result for $s=10$, one can easily anticipate the $s\to \infty$ limit: Here the anisotropy barrier is strictly linear in $\Delta$ for weak anisotropies up to a certain critical value $\Delta_c / J=1$. At $\Delta_c$ we find a cusp, i.e.\ a discontinuous jump of $\partial^2 \Delta E / \partial \Delta^2$. Beyond this point, a fairly broad crossover region follows until finally a linear-in-$\Delta$ trend is established again in the $\Delta \to \infty$ limit. The strong-anisotropy limit is easily understood, as quantum fluctuations are suppressed anyway for all $s$. For weak anisotropy, on the other hand, the macrospin approximation is found to be exact in a finite range up to $\Delta = \Delta_c$, i.e.\ perturbative corrections of second and higher order in $\Delta$ are decreasing with increasing $s$ and finally vanish exactly in the classical limit $s\to \infty$. For all $\Delta < \Delta_c$, magnetization reversal takes place by a coherent rotation of completely correlated spins in the classical case. For any finite $s$, on the other hand, quantum fluctuations break up this perfect correlation. The magnetic moments $\langle \ff s_i \rangle$ are still aligned to the $z$ axis during the reversal -- we have $\langle \ff s_i \rangle \propto \ff e_z$ since $S_z$ is a conserved quantity -- but the effect of fluctuations can be seen in the spin-spin correlation functions: In the $M=0$ sector, for example, we find $\langle s_{iz} s_{jz} \rangle = 0$ classically ($s\to \infty$) and $\langle s_{iz} s_{jz} \rangle = s^2 (1-1/(1-1/2Ls))$ in the quantum case for sufficiently small $\Delta$ (see \refeq{corrf}), i.e.\ a small quantity of the order $\ca O(L^{-1})$ which is independent of $i$ and $j$. With increasing $\Delta$ at finite $s$, however, $\langle s_{iz} s_{jz} \rangle$ gradually becomes more and more site dependent. As a function of $j$ and fixing $i$ to one of the chain edges, $\langle s_{iz} s_{jz} \rangle$ is monotonically decreasing in the $M=0$ sector. Spins at the chain edges are oriented antiferromagnetically: $\langle s_{iz} s_{jz} \rangle < 0$ for $i=1$ and $j=L$. Hence, for any finite $\Delta$ in the quantum case there are already indications for domain-wall formation while in the classical limit there is a clear-cut distinction, given by a critical coupling $\Delta_c$, between the two reversal mechanisms. \begin{figure}[t] \includegraphics[width=0.85\columnwidth]{fig09.eps} \caption{ $\Delta$ dependence of energy barrier $\Delta E$ for open chains with $L=4$ spins and different spin quantum numbers $s$ as indicated. Exact results for the full model $H_J + H_\Delta$ after rescaling $\ff s_i \to \ff s_i / \sqrt{s(s+1)}$ for $s=1/2,1,3/2,2,5/2,3,7/2,4,5,10$. The inset shows the first derivative of $\Delta E$ with respect to $\Delta$. } \label{class} \end{figure} To understand the origin of the ``phase transition'' in the classical limit and to find a means to compute $\Delta_c$ for arbitrary system size and geometry, we first performed numerical minimizations of the total energy $E(\{ \ff s_1, ..., \ff s_L \})$ of the classical model $H = H_J + H_{\rm ani}$ for both, $H_{\rm ani} = H_\Delta$ and $H_{\rm ani}=H_D$. Arbitrary spin configurations $\{ \ff s_1, ..., \ff s_L \}$ with $|\ff s_i|=1$ and satisfying the constraint $\sum_{i=1}^L s_{iz} = M$ are considered. For $M=0$ this provides the classical barrier height and the corresponding optimal spin configuration. Calculations have at first been done for different small system sizes $L$, typically $L=4$, to reproduce the critical point anticipated from the results of Fig.\ \ref{class}. \begin{figure}[t] \includegraphics[width=0.99\columnwidth]{fig10.eps} \caption{ Sketch of the different magnetization-reversal mechanisms in the classical model for anisotropy strength $D$ smaller (macrospin model, coherent rotation) or larger (nucleation) than the critical value $D_c$ and for infinite $D$ (nucleation, vanishing domain-wall width) as verified by calculations for $L=4$ spins. } \label{mechan} \end{figure} We found the optimal spin configuration with minimal total energy for a given $M$ to be coplanar in all cases, i.e.\ there is a coordinate frame with $s_{iy}=0$ for all spins. As dictated by the symmetry of the model, configurations that transform into each other by rotations around the $z$ axis are degenerate. For a chain geometry, there is also degeneracy between configurations obtained by mirror transformations, $\ff s_i \to \ff s_{L+1-i}$, and nucleation can thus start from both edges with equal probability. Furthermore, for $M=0$ we have $s_{i,z} = - s_{L-i+1,z}$ in the optimal configuration. The results can be summarized by the sketch of the optimal spin configurations at a given $M$ in Fig.\ \ref{mechan}. Results for coupling and for single-site anisotropy are found to be qualitatively the same. As is seen from Fig.\ \ref{mechan}, the different mechanisms for magnetization reversal in the two limits of weak and strong anisotropy are reproduced from the calculations for small $L$. More important, however, the physical origin of the critical point can be identified easily: At $\Delta = \Delta_c$ (or at $D=D_c$, respectively), and approaching the critical point from the weak-anisotropy side, the optimal spin configuration exhibits an instability towards a non-collinear ordering. As sketched in Fig.\ \ref{mechan}, the instability develops simultaneously for any $M$. This observation can be exploited to calculate the critical anisotropy strength for arbitrary system size and system geometry in the following way: Using $s_{iy}=0$, we parameterize the spin variables as $\ff s_i=(\sqrt{1-m_i^2},0,m_i)$ with $-1 \le m_i \le 1$. This yields the energy functional in the form \begin{eqnarray} && E(\{m_1,...,m_L\}) = - \frac{1}{2} \sum_{ij} D_{ij} m_i m_j \nonumber \\ && - \frac{1}{2} \sum_{ij} J_{ij} \left( \sqrt{1-m_i^2}\sqrt{1-m_j^2} + m_i m_j \right) \: , \end{eqnarray} which comprises both, the coupling and the single-site anisotropy case. To find the critical point, it is sufficient to concentrate on $M=0$ and to compute the magnetic response of the system in the $M=0$ state with spin configuration $m_1= \cdots = m_L=0$. Close to the critical point, we can assume $m_i \ll 1$. Expanding the total energy in powers of $m_i$ then yields \begin{equation} E(\{m_1,...,m_L\}) = \mbox{const.} + \frac{1}{2} \sum_{ij} \chi^{-1}_{ij} m_i m_j + {\cal O}({m_i^4}) \end{equation} with the inverse susceptibility matrix \begin{equation} \chi^{-1}_{ij} = - J_{ij} + J_i \delta_{ij} - D_{ij} \: , \end{equation} and where we have defined $J_i \equiv \sum_j J_{ij}$. We assume the model parameters as real and symmetric, $J_{ij}= J_{ji}$ and $D_{ij}=D_{ji}$. Note that approaching the isotropic limit with anisotropy strength $D\to 0$, the susceptibility matrix exhibits an eigenvalue approaching zero. This corresponds to an instability of the state towards ferromagnetic ordering with all $m_i > 0$ and indicates the incipient degeneracy of states with different $M$. For finite $D$, this eigenvalue is negative and thus implies an indefinite susceptibility. This indicates that, by construction, $\chi$ refers to the thermodynamically unstable excited state with $M=0$ and $m_i=0$. The {\em local} instability of this state towards a non-collinear alignment of the magnetic moments at the critical anisotropy strength $D_c$ shows up as a zero of another eigenvalue of $\chi$. For a short chain of $L=4$ spins with coupling anisotropy this happens exactly at $\Delta_c = J$, in agreement with the Lanczos results displayed in Fig.\ \ref{class}. The eigenvector corresponding to the zero eigenvalue describes the critical spin profile which agrees with the $M=0$ spin configuration for $D>D_c$ sketched in Fig.\ \ref{mechan}. \begin{figure}[t] \includegraphics[width=0.99\columnwidth]{fig11.eps} \caption{ System-size dependence of the critical value for the anisotropy strength $D_c$ in units of $J$ for one-dimensional open chains, for two-dimensional square arrays and for three-dimensional simple cubic arrays and spheres with open boundaries. Blue lines: single-site anisotropy. Red lines: coupling anisotropy. } \label{critical} \end{figure} Results for the critical anisotropy strength as a function of the system size are shown in Fig.\ \ref{critical} for both, coupling and single-site anisotropy and for systems of different dimensions $d$. In all cases we considered systems with open boundaries. Generally, the anisotropy strength up to which the macrospin approximation is found to be valid, decreases rapidly with system size but depends only weakly on the type of the anisotropy term. There is also a weak dependence on the form of the particle as is obvious from the comparison of cubic arrays and spheres in $d=3$ dimensions. For small nanosystems, the dimension and shape dependence of $D_c$ is somewhat irregular. For a larger number of classical spins, on the other hand, the critical anisotropy strength for which the macrospin model can be applied, is the stronger the higher the dimension of the nanostructure. In fact, in the limit $L\to \infty$, the different results are found to follow a simple power law, \begin{equation} D_c(L) \propto L^{-2/d} = l^{-2} \; , \label{eq:scal} \end{equation} where $d$ is the dimensionality, $L$ the number of spins and $l = L^{1/d}$ the {\em linear} extension of the system. This can easily be explained by comparing the energy barrier for coherent rotation $\Delta E_{\rm cor.rot.}$ with the barrier for nucleation and domain-wall propagation $\Delta E_{nucl.}$. While the volume of the system $\propto l^d$ is relevant for $\Delta E_{\rm cor.rot.} \propto D \, l^d$, the $d-1$-dimensional domain wall of width $\sigma(D) \propto \sqrt{J/D}$ \cite{Kit49,Mid63} essentially determines $\Delta E \propto D \, \sigma(D) \, l^{d-1}$. The critical anisotropy is then obtained from $D_c \, l^d \propto D_c \sqrt{J/D_c} \, l^{d-1}$, resulting in the scaling law (\ref{eq:scal}). \section{Conclusions} \label{sec:con} Anisotropy terms coupled to the ferromagnetic Heisenberg model of many-spin nanosystems give rise to an energy barrier $\Delta E$ that crucially determines the thermal activation of magnetization reversal and thus the magnetic stability of nanosystems. Employing perturbation theory, exact diagonalization and Lanczos as well as classical total-energy minimization and stability analysis, different quantum effects have been revealed that substantially affect the barrier height. Quantum effects are absent in the limit of strong anisotropy $D$. For $D/J \to \infty$ the Ising model is approached, i.e.\ a classical model where there are no quantum fluctuations. The magnetization-reversal mechanism in the Ising limit is given by nucleation at the edge or surface of the system followed by the propagation of a domain wall with minimal (unit) width. With decreasing anisotropy strength, the domain-wall width increases. Finally, a critical value $D = D_c$ is reached where the reversal mechanism changes to a coherent rotation of the spins that are tightly aligned by the exchange coupling $J$ and form a huge macrospin. This critical anisotropy strength is well defined in the classical limit $s \to \infty$ (with model parameters rescaled by the factor $s(s+1)$) and can be computed for arbitrary dimension, system size and shape by an analysis of the local stability of the susceptibility matrix in the globally (thermodynamically) unstable excited state at the top of the barrier. For $D<D_c$, i.e.\ for coherent rotation, the classical anisotropy-barrier height is simply given by $\Delta E_{\rm class} = E(m_1= \cdots m_L = 0) - E(m_1= \cdots = m_L =1)$, i.e.: \begin{equation} \Delta E_{\rm class} = L \, D \; . \label{eq:deltaeclass} \end{equation} This is the same as the barrier of a classical macrospin of length $\ff S=L\ff s$, composed of microscopic spins of unit length $|\ff s|=1$, as described by a macrospin model $H=D_{\rm macro} S_z^2$ with a barrier $\Delta E = D_{\rm macro} L^2$ if $D_{\rm macro} = D / L$. Classically, and in the regime where coherent rotations takes place, the macrospin approximation is exact. The first quantum effect manifests itself in the absence of the a well-defined critical anisotropy strength. As can be seen in the spin-spin correlation function $\langle s_{iz} s_{jz} \rangle$, there is rather a smooth crossover from nucleation and domain-wall propagation to coherent rotation with decreasing $D$. This crossover becomes less and less sharp with decreasing $s$, i.e.\ as the importance of quantum fluctuations increases. Second, for any finite anisotropy strength, the macrospin approximation is no longer correct in the quantum case. Only within perturbation theory to first order in $D$, the macrospin model emerges as the effective low-energy theory of the anisotropic quantum Heisenberg model. Comparing the true anisotropy-energy barrier $\Delta E$ of the many-spin system with the one of the corresponding effective macrospin model, we find \begin{equation} \Delta E < \Delta E_{\rm macro} \: . \end{equation} In the quantum case, there is virtually never a reversal by pure coherent rotation but always some ``admixture'' of nucleation and domain-wall propagation. The latter leads to the decrease of $\Delta E$ as compared to $\Delta E_{\rm macro}$. This quantum effect becomes more and more pronounced with increasing anisotropy strength, with increasing system size, with decreasing dimensionality and with decreasing average coordination number. Third, the quantum derivation of the macrospin model by means of first-order perturbation theory provides an explicit expression for the anisotropy parameter [see Eqs. (\ref{eq:macro1}) and (\ref{eq:macro2})]. This leads to an energy barrier that is lower than the classical one: \begin{equation} \Delta E_{\rm macro} < \Delta E_{\rm class} \: . \end{equation} Consider a single-site anisotropy case as an example: After the rescaling $D \to D/s(s+1)$ we have $\Delta E_{\rm macro} = [(2s-1)/(2Ls-1)] [L^2s^2 / s(s+1) ] D$. Therewith, the classical result (\ref{eq:deltaeclass}) is recovered in the high-spin limit: $\lim_{s\to\infty} \Delta E_{\rm macro} = \Delta E_{\rm class}$. In the low-spin case, e.g.\ for $s=1$, which is the extreme case for single-site anisotropy, however, we have $\Delta E_{\rm macro} = (L/4) D$. This is smaller by a factor of 4 than the classical barrier $\Delta E_{\rm class}$. Even in the limit of large systems $L\to \infty$, there is substantial difference, \begin{equation} \Delta E_{\rm macro} = \frac{1}{2} \frac{2s-1}{s+1} L D < L D = \Delta E_{\rm class} \: , \end{equation} compared to the classical limit which is approached rather slowly, i.e.\ $\Delta E_{\rm macro} / \Delta E_{\rm class} = 1 - 3s / s + \ca O(s^{-2})$. We conclude that, in addition to quantum many-spin effects, there is a quantum correction of the macro-spin model itself which, for low spin quantum numbers, can be as important as the former one. \acknowledgments We would like to thank E.\ Y.\ Vedmedenko for helpful discussions. The work is supported by the Deutsche Forschungsgemeinschaft within the Sonderforschungsbereich 668 (project B3).
\section{Statement of the results} The subject of this paper are rational ${\mathbb Q}$-factorial Fano varieties $X$ defined over an algebraically closed field ${\mathbb K}$ of characteristic zero, see for example \cite{Is} and \cite{MoMu} for classical work. A more recent focus in this field are toric Fano varieties, where one uses the description in terms of lattice polytopes (\cite{Ba}, \cite{KaKrNi}, \cite{Ka2}, etc.). Here we study the case that $X$ comes, more generally, with an effective action of a torus $T$ of complexity one, i.e. $\dim X-\dim T=1$; by Fano varieties we mean normal projective varieties with ample anticanonical divisor $-K_X$. We continue the work of \cite{HaHeSu} where classification results for the case $\operatorname{Cl}(X)={\mathbb Z}$ were given. {In this paper we study the more general case of Picard number one, i.e.~we allow torsion in the divisor class group. A first step is Theorem \ref{Th:FiniteIndex} where we provide effective bounds for the number of deformation types of Fano varieties $X$ as above with fixed dimension $d$ and Picard index $\mu:=[\operatorname{Cl}(X):\operatorname{Pic}(X)]$. As a consequence we obtain restricting statements about the number $\delta(d,\mu)$ of different deformation types of ${\mathbb Q}$-factorial $d$-dimensional Fano varieties with a complexity-one torus action, Picard number one and Picard index~$\mu$. In the toric situation $\delta(d,\mu)$ is bounded above by $\mu^{d^2}$. For the non-toric case we get the following asymptotic results: \begin{theorem}\label{Th:asymptotic} For fixed $d\in{\mathbb Z}_{>0}$, the number $\delta(d,\mu)$ is asymptotically bounded above by $\mu^{(1+\varepsilon)\mu^2}$ for $\varepsilon>0$ arbitrarily small, and for fixed $\mu\in{\mathbb Z}_{>0}$, it is asymptotically bounded above by $d^{Ad}$ with a constant $A$ depending only on $\mu$. \end{theorem} We turn to the classification. Our approach uses the Cox ring $\mathcal R(X)$ which is defined by $$ \mathcal R(X) \ = \ \bigoplus_{D\in\operatorname{Cl}(X)}\Gamma(X,\mathcal O_X(D)). $$ Given this ring, the variety $X$ can be realized as a quotient of an open subset in $\mathrm{Spec}(\mathcal R(X))$ by the action of a diagonalizable group. According to \cite[Theorem 1.3]{HaSu} the Cox ring of a normal complete rational variety with a complexity-one torus action is finitely generated. Furthermore, every such Fano variety is uniquely determined by its Cox ring (as a $\operatorname{Cl}(X)$-graded ring). In case of Picard number one the toric varieties of this type correspond to the fake weighted projective spaces as defined in \cite{Ka} and the Cox ring is polynomial. In the subsequent theorems we list non-toric complexity-one Fanos with Picard number one in the cases where $\operatorname{Cl}(X)$ has non trivial torsion; for the non-toric results in case of $\operatorname{Cl}(X)={\mathbb Z}$ we refer to \cite{HaHeSu}. The Cox rings are described in terms of generators and relations and we specify the $\operatorname{Cl}(X)$-grading by giving the degrees of the generators. Additionally we list the degree of the Fano varieties $d_X:=(-K_X)^d$ and the Gorenstein index $\iota(X)$, i.e. the smallest positive integer such that $\iota(X)\cdot K_X$ is Cartier. \begin{theorem}\label{Theo:surfaces<=6} Let $X$ be a non-toric Fano surface with an effective ${\mathbb K}^*$-action, Picard number one, non trivial torsion in the class group and $[\operatorname{Cl}(X):\operatorname{Pic}(X)]\leq 6$. Then its Cox ring is precisely one of the following. \begin{center} \begin{longtable}[htbp]{lllccc} \multicolumn{6}{c}{\bf $[\operatorname{Cl}(X):\operatorname{Pic}(X)] = 2$} \\[1ex] \toprule No. & $\mathcal{R}(X)$ & $\operatorname{Cl}(X)$ & \emph{grading} & $d_X$ & $\iota(X)$ \\ \midrule 1 & ${\mathbb K}[{T_1,\ldots,T_4}]/ \bangle{T_1T_2^3+T_3^4+T_4^2}$ & ${\mathbb Z}\oplus {\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2\\ \overline 0 & \overline 0 & \overline 1 & \overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \bottomrule \\[2ex] \\ \\ \\ \multicolumn{6}{c}{\bf $[\operatorname{Cl}(X):\operatorname{Pic}(X)] = 3$} \\[1ex] \toprule No. & $\mathcal{R}(X)$ & $\operatorname{Cl}(X)$ & \emph{grading} & $d_X$ & $\iota(X)$ \\ \midrule 2 & ${\mathbb K}[{T_1,\ldots,T_4}]/ \bangle{T_1T_2^2+T_3^3+T_4^3}$ & ${\mathbb Z}\oplus{\mathbb Z}/3{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 1 & 1 & 1\\ \overline 1 & \overline 1 & \overline 2 & \overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \bottomrule \\[2ex] \multicolumn{6}{c}{\bf $[\operatorname{Cl}(X):\operatorname{Pic}(X)] = 4$} \\[1ex] \toprule No. & $\mathcal{R}(X)$ & $\operatorname{Cl}(X)$ & \emph{grading} & $d_X$ & $\iota(X)$ \\ \midrule 3 & ${\mathbb K}[{T_1,T_2,T_3,S_1}] / \bangle{T_1^2+T_2^2+T_3^2}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 1 & 1 & 1\\ \overline 1 & \overline 1 & \overline 0 & \overline 0\\ \overline 0 & \overline 1 & \overline 1 & \overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 4 & ${\mathbb K}[{T_1,\ldots,T_4}] / \bangle{T_1T_2+T_3^2+T_4^2}$ & ${\mathbb Z}\oplus{\mathbb Z}/4{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 1 & 1 & 1\\ \overline 1 & \overline 3 & \overline 2 & \overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 5 & ${\mathbb K}[{T_1,\ldots,T_4}] / \bangle{T_1^2T_2+T_3^2+T_4^4}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 2 & 2 & 1\\ \overline 1 & \overline 0 & \overline 1 & \overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 6 & ${\mathbb K}[{T_1,\ldots,T_4}] / \bangle{T_1T_2^2+T_3^6+T_4^2}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 2 & 2 & 1 & 3\\ \overline 0 & \overline 1 & \overline 0 & \overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 7 & ${\mathbb K}[{T_1,\ldots,T_5}] / \bangle{ \begin{smallmatrix} T_1T_2+T_3^2+T_4^2,\\ \lambda T_3^2+T_4^2+T_5^2 \end{smallmatrix}}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 1 & \overline 1 & \overline 0 & \overline 1 &\overline 0\\ \overline 0 & \overline 0 & \overline 1 & \overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \bottomrule \\[2ex] \multicolumn{6}{c}{\bf $[\operatorname{Cl}(X):\operatorname{Pic}(X)] = 6$} \\[1ex] \toprule No. & $\mathcal{R}(X)$ & $\operatorname{Cl}(X)$ & \emph{grading} & $d_X$ & $\iota(X)$ \\ \midrule 8 & ${\mathbb K}[{T_1,T_2,T_3,S_1}] / \bangle{T_1^3+T_2^3+T_3^2}$ & ${\mathbb Z}\oplus{\mathbb Z}/3{\mathbb Z}$ & $\left( \begin{smallmatrix} 2 & 2 & 3 & 1\\ \overline 1 & \overline 2 & \overline 0 & \overline 0 \end{smallmatrix} \right) $ & $2/3$ & $3$ \\ \midrule 9 & ${\mathbb K}[{T_1,\ldots,T_4}]/ \bangle{T_1T_2+T_3^3+T_4^3}$ & ${\mathbb Z}\oplus{\mathbb Z}/3{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 2 & 1 & 1\\ \overline 1 & \overline 2 & \overline 2 & \overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 10 & ${\mathbb K}[{T_1,\ldots,T_4}] / \bangle{T_1T_2+T_3^2+T_4^4}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 3 & 1 & 2 & 1\\ \overline 1 & \overline 1 & \overline 1 & \overline 0 \end{smallmatrix} \right)$ & $3$ & $1$ \\ \midrule 11 & ${\mathbb K}[{T_1,\ldots,T_4}] / \bangle{T_1T_2^5+T_3^2+T_4^8}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 3 & 1 & 4 & 1\\ \overline 1 & \overline 1 & \overline 1 & \overline 0 \end{smallmatrix} \right)$ & $1/3$ & $3$ \\ \bottomrule \end{longtable} \end{center} where the parameter $\lambda$ occuring in the second relation of surface number $7$ can be any element of ${\mathbb K}^*\setminus \{1\}$. Furthermore the Cox rings listed above are pairwise non-isomorphic as graded rings. \end{theorem} \begin{remark} Gorenstein surfaces are well known to have ADE-singularities which are in particular canonical. Consequently the surfaces of number $1$ to $5$, $7$, $9$ and $10$ are canonical. Furthermore in \cite{Su} all log-terminal Del Pezzo ${\mathbb K}^*$-surfaces of Gorenstein index up to three are classified. Comparing the surfaces listed in \cite[Theorems~4.9, 4.10]{Su} with the table above shows that number $11$ is not log-terminal. The resolution of this surface can be explicitly computed by using the method of toric ambient modification as demonstrated in \cite[Examples~3.20, 3.21]{Ha3}. \end{remark} \begin{theorem} \label{thm:3fano2} Let $X$ be a three-dimensional non-toric Fano variety with an effective two-torus action, Picard number one, non trivial torsion in the class group and $[\operatorname{Cl}(X):\operatorname{Pic}(X)]=2$. Then its Cox ring is precisely one of the following. \begin{center} \begin{longtable}[htbp]{lllcccc} \toprule No. & $\mathcal{R}(X)$ & $\operatorname{Cl}(X)$ & \emph{grading} & $d_X$ & $\iota(X)$ \\ \midrule 1 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2+T_3^2+T_4^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 1 &\overline 1 &\overline 1 &\overline 0 &\overline 0 \end{smallmatrix} \right)$ & $27$ & $1$ \\ \midrule 2 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^3+T_3^2+T_4^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 2 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 3 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^3+T_3^2+T_4^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 2 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $8$ & $1$ \\ \midrule 4 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^3+T_3^2+T_4^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 2 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 5 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^5+T_3^2+T_4^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 3 & 1 & 1\\ \overline 0 &\overline 0 &\overline 0 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 6 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^5+T_3^2+T_4^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 3 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 7 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1^2T_2^4+T_3^2+T_4^3 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 3 & 2 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $4$ & $1$ \\ \midrule 8 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1^2T_2^4+T_3^2+T_4^3 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 3 & 2 & 1\\ \overline 0 &\overline 1 &\overline 1 &\overline 0 &\overline 0 \end{smallmatrix} \right)$ & $4$ & $1$ \\ \midrule 9 & ${\mathbb K}[{T_1,\ldots,T_4,S_1}]/ \langle T_1T_2^5+T_3^3+T_4^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 2 & 3 & 1\\ \overline 1 &\overline 1 &\overline 0 &\overline 0 &\overline 0 \end{smallmatrix} \right)$ & $4$ & $2$ \\ \midrule 10 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3^2T_4^2+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 2\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 11 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3^2T_4^2+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 1 &\overline 1 &\overline 0 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 12 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3^2T_4^4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 13 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3^2T_4^4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 0 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 14 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1^2T_2^4+T_3^3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 15 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1^2T_2^4+T_3^3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 1 &\overline 0 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 16 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1^2T_2^4+T_3^5T_4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 17 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1^2T_2^4+T_3^5T_4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 1 &\overline 0 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 18 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2+T_3T_4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $27$ & $2$ \\ \midrule 19 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2+T_3T_4+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $27$ & $1$ \\ \midrule 20 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^2+T_3T_4^2+T_5^3 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 0 &\overline 1 &\overline 0 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $12$ & $1$ \\ \midrule 21 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3T_4^3+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 1 &\overline 1 &\overline 0 &\overline 0 &\overline 0 \end{smallmatrix} \right)$ & $2$ & $2$ \\ \midrule 22 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3T_4^3+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $2$ & $1$ \\ \midrule 23 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 2\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 24 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^3+T_3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 2\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 25 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3T_4^5+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 26 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3T_4^5+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 27 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3^3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 28 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^5+T_3^3T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $2$ \\ \midrule 29 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^2+T_4^2+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $8$ & $2$ \\ \midrule 30 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^2+T_4^2+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2 & 1\\ \overline 0 &\overline 0 &\overline 0 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $1$ \\ \midrule 31 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^2+T_4^2+T_5^4 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 1 \end{smallmatrix} \right)$ & $8$ & $1$ \\ \midrule 32 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^4+T_4^2+T_5^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 3 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 33 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^4+T_4^2+T_5^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 3 & 1\\ \overline 0 &\overline 0 &\overline 1 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 34 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^2T_3^3+T_4^2+T_5^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 3 & 1\\ \overline 0 &\overline 1 &\overline 0 &\overline 1 &\overline 0 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 35 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^2T_3^3+T_4^2+T_5^6 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 3 & 1\\ \overline 0 &\overline 1 &\overline 0 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $1$ & $1$ \\ \midrule 36 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2^2T_3^3+T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2 & 3\\ \overline 1 &\overline 0 &\overline 1 &\overline 0 &\overline 0 \end{smallmatrix} \right)$ & $4$ & $1$ \\ \midrule 37 & ${\mathbb K}[{T_1,\ldots,T_5}]/ \langle T_1T_2T_3^4+T_4^3+T_5^2 \rangle$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $\left( \begin{smallmatrix} 1 & 1 & 1 & 2 & 3\\ \overline 0 &\overline 0 &\overline 1 &\overline 0 &\overline 1 \end{smallmatrix} \right)$ & $4$ & $2$ \\ \midrule 38 & ${\mathbb K}[{T_1,\ldots,T_6}] / \bangle{ \begin{smallmatrix} T_1T_2+T_3T_4+T_5^2,\\ \lambda T_3T_4+T_5^2+T_6^2 \end{smallmatrix}}$ & ${\mathbb Z}\oplus{\mathbb Z}/2{\mathbb Z}$ & $ \left( \begin{smallmatrix} 1 & 1 & 1 & 1 & 1 & 1\\ \overline 0 & \overline 0 & \overline 1 & \overline 1 &\overline 1 & \overline 0\\ \end{smallmatrix} \right)$ & $16$ & $2$ \\ \bottomrule \end{longtable} \end{center} where the parameter $\lambda$ occuring in the second relation of threefold number $38$ can be any element of ${\mathbb K}^*\setminus \{1\}$. Furthermore the Cox rings listed above are pairwise non-isomorphic as graded rings. \end{theorem} Let $X$ be a normal complete rational variety coming with a complexity-one torus action of $T$. Consider the $T$-invariant open subset $X_0$ consisting of all points $x\in X$ having finite isotropy group. According to \cite[Corollary~3]{Sum} there is a geometric quotient $q\colon X_0\to X_0/T$ such that $X_0/T$ is irreducible and normal but possibly not separated. The property of the orbit space $X_0/T$ being separated is reflected in the Cox ring relations by the condition that each monomial depends on only one variable, e.g. surface number $3$ in Theorem \ref{Theo:surfaces<=6}; see \cite[Theorem 1.2]{HaSu}. Geometrically, this means, that every orbit is contained in the closure of either exactly one maximal orbit or of infinitely many maximal orbits. For such varieties we have the following general finiteness statement: \begin{theorem}\label{Theo:sepquot} The number of $d$-dimensional normal complete rational varieties of Picard number one with a complexity-one torus action of $T$ and Picard index $\mu$, such that $X_0/T$ is separated, is finite. \end{theorem} \section{Description of the Cox ring} We briefly recall from \cite{HaHe} a construction of ${\mathbb Q}$-factorial normal rational projective varieties with a complexity-one torus action. Here, we specialize to the case of Picard number one; the details are given in \cite[Proposition~2.4]{HaHe}. \begin{construction}\label{Con:R(A,n,L,m)}\label{Con:varX} For $r \ge 1$, consider a sequence $A = (a_0, \ldots, a_r)$ of pairwise linearly independent vectors in ${\mathbb K}^2$, a sequence $\mathfrak{n} = (n_0, \ldots, n_r)$ of positive integers, a non-negative integer $m$ and a family $L = (l_{ij})$ of positive integers, where $0 \le i \le r$ and $1 \le j \le n_i$. Set $$ R(A,\mathfrak n,L,m) \ := \ {\mathbb K}[T_{ij},S_k] \ / \ \bangle{g_0, \ldots, g_{r-2}}. $$ where the $T_{ij}$ are indexed by $0 \le i \le r, \; 1 \le j \le n_i$, the $S_k$ by $1\leq k\leq m$ and the relations $g_i$ are defined as follows: Set $T_i^{l_i}:=T_{i1}^{l_{i1}}\cdots T_{in_i}^{l_{in_i}}$ and $$ g_i \ := \ \det \begin{pmatrix} a_i & a_{i+1} & a_{i+2}\\ T_{i}^{l_i} &T_{i+1}^{l_{i+1}} &T_{i+2}^{l_{i+2}} \end{pmatrix}. $$ Define $n:=n_0+\ldots+n_r$ and let $K:={\mathbb Z}\oplus K^t$ be an abelian group with torsion part $K^t$. Suppose that $R(A,\mathfrak n, L,m)$ is positively $K$-graded via $$ \deg\, T_{ij} \ = \ w_{ij}\in K, \qquad \deg\, S_k \ = \ u_k\in K, $$ i.e. $w_{ij},u_k\in{\mathbb Z}_{\geq 0}\otimes K^t$, and that any $n+m-1$ of these degrees generate $K$ as a group. The $K$-grading defines a diagonal action of $H:={\rm Spec}\; {\mathbb K}[K]$ on ${\mathbb K}^{n+m}$. By construction $$ \overline X \ := \ V(g_{i};0\leq i\leq r-2) \ = \ {\rm Spec}\, R(A,\mathfrak n,L,m) $$ is invariant under this $H$-action. The open set ${\mathbb K}^{n+m}\setminus\{0\}$ allows a geometric quotient of this $H$-action which is denoted by $p\colon {\mathbb K}^{n+m}\setminus\{0\} \to Z$, where the toric variety $Z$ is a a fake weighted projective space. Furthermore we get a geometric quotient $p\colon\widehat X\to X$ of the embedded open subset $\widehat X := \overline X\setminus \{0\}$. \[ \begin{xy} \xymatrix{ & {\widehat X}\; \ar[d]^{ p} \ar@{^ {(}->}[r] &{\mathbb K}^{n+m}\setminus\{0\}\ar[d]^{p} \\ &X\; \ar@{^ {(}->} [r] &Z} \end{xy} \] The quotient space $X:=\widehat X\sslash H$ is a ${\mathbb Q}$-factorial normal projective variety of dimension $$ \dim(X) \ = \ n+m-r. $$ It has divisor class group $\operatorname{Cl}(X)=K$, Cox ring $\mathcal R(X) =R(A,\mathfrak n,L,m)$ complexity-one torus action. This torus is given by the stabilizer of $X$ under the action of the maximal torus $T_Z$ of $Z$.\\ Note that, if there is an index $0\leq i\leq r$ such that $l_{i1}=1$ and $n_i=1$, then there is at least one relation containing a linear term. In this case the ring is isomorphic to the polynomial ring that we get, if we omit the relations of this type. Consequently we may always assume $l_{i1}n_{i}\neq 1$. \end{construction} \begin{remark} Varieties with complexity-one action, as constructed in \ref{Con:varX}, can be considered as a generalized version of well-formed complete intersections in weighted projective spaces, in the sense of \cite{IaFl}. \end{remark} According to \cite[Theorem~1.5]{HaHe} every ${\mathbb Q}$-factorial normal complete rational variety $X$ with a complexity-one torus action and Picard number one has a Cox ring $R(X)$ which is isomorphic as a graded ring to some $K$-graded algebra $R(A,\mathfrak n,L,m)$ with $K\cong\operatorname{Cl}(X)$. We collect some geometric properties of the varieties $X$ just constructed. Every element $w\in K={\mathbb Z}\oplus K^t$ can be written as $w=w^0+w^t$ where $w^0\in {\mathbb Z}$ and $w^t\in K^t$. Furthermore, every $\overline x=(\overline x_{ij},\overline x_k) \in\widehat X\subseteq {\mathbb K}^{n+m}$ defines a point $x\in X$ by $x:=p(\overline x)$; the points $\overline x\in\widehat X$ are called Cox coordinates of $x$. We denote the set of all weights corresponding to a non-zero coordinate of $\overline x$ by $$ W_{\overline x} :=\{w_{ij};\;\overline x_{ij}\neq 0\} \cup \{u_{k};\;\overline x_{k}\neq 0\}. $$ Moreover, let $\operatorname{Cl}(X,x)$ denote the local divisor class group in $x$, i.e. the group of all divisor classes that are principal near $x$. \begin{proposition}\label{Prop:FanoPicard} Let $X$ be a ${\mathbb Q}$-factorial complete normal variety with complexity-one torus action and Picard number one as constructed in \ref{Con:varX} and set $\gamma_i:=\deg(g_i)$, $0\leq i\leq r$. Then the following statements hold: \begin{itemize} \item[(i)] For any $\overline{x} \in \widehat{X}$, the local divisor class group $\operatorname{Cl}(X,x)$ of $x := p(\overline{x})$ is finite and $\mathrm{gcd}(w^0; \; w\in W_{\overline x})$ always divides the order of the group. \item[(ii)] The Picard group $\operatorname{Pic}(X)$ is free and the Picard index is given by \begin{eqnarray*} [\operatorname{Cl}(X):\operatorname{Pic}(X)] & = & \mathrm{lcm}_{x \in X}(\mathrm{gcd}(w^0; \; w\in W_{\overline x}))\cdot|\operatorname{Cl}(X)^t|. \end{eqnarray*} In particular $|\operatorname{Cl}(X)^t|$ is a divisor of $[\operatorname{Cl}(X):\operatorname{Pic}(X)]$ and we have $|\operatorname{Cl}(X)^t|\leq[\operatorname{Cl}(X):\operatorname{Pic}(X)]$. \item[(iii)] For the anticanonical class $-K_X\in\operatorname{Cl}(X)$ and its self intersection number $d_X:=(-K_X)^d$ one has \begin{align*} -K_X & = \ \sum_{i=0}^r\sum_{j=1}^{n_i} w_{ij} +\sum_{k=1}^mu_k-\sum_{i=0}^{r-2}\gamma_i,\\ \quad d_X & = \ \left(\sum_{i=0}^r\sum_{j=1}^{n_i} w_{ij}^0 +\sum_{k=1}^mu_k^0-\sum_{i=0}^{r-2}\gamma_i^0\right)^d \frac{\gamma_0^0\cdots\gamma_{r-2}^0} {\prod_{i=0}^{r} \prod_{j=1}^{n_i} w_{ij}^0 \prod_{k=1}^mu_k^0\cdot|\operatorname{Cl}(X)^t|}. \end{align*} \item[(iv)] The variety $X$ is Fano if and only if the following inequality holds: \begin{eqnarray*} (r-1)\deg(g_0)^0 \ = \ \sum_{i=0}^{r-2} \deg (g_i)^0 & < & \sum_{i=0}^{r}\sum_{j=1}^{n_i} w_{ij}^0 +\sum_{i=1}^mu_k^0. \end{eqnarray*} \end{itemize} \end{proposition} \begin{proof} Let $\overline x(i,j)$ resp. $\overline x(k)$ be a point in $\widehat X$ having the $ij$-th resp. $(n+k)$-th entry one and all others zero. With $\widehat Z:={\mathbb K}^{n+m}\setminus\{0\}$ we obtain a commutative diagram \[ \begin{xy} \xymatrix{ {\widehat X} \ \ar@^{(-}[r] \ar[r] \ar[d]_{\sslash H} & {\widehat Z} \ar[d]^{\sslash H}\\ X \ \ar@^{(-}[r] \ar[r] & Z } \end{xy} \] where the induced map embeds $X$ into a toric variety $Z$ such that $\operatorname{Cl}(X)\cong\operatorname{Cl}(Z)$ and $\operatorname{Pic}(X)\cong\operatorname{Pic}(Z)$ holds; see \cite[Corollary~III.3.1.7]{ArDeHaLa}. By choice $\overline x(i,j)$ resp. $\overline x(k)$ is a toric fixed point. Consequently, the Picard group $\operatorname{Pic}(Z)$, and also $\operatorname{Pic}(X)$, is free \cite[Theorem VII. 2.16]{Ew}. According to \cite[Corollary~4.9]{Ha2} we obtain $$ \operatorname{Pic}(X) \ = \ \bigcap_{\overline x \in \widehat X} \langle w;\; w\in W_{\overline x}\rangle \ \cong \ \bigcap_{\overline x \in \widehat X} \langle w^0;\; w\in W_{\overline x}\rangle, $$ where the last equality follows from the fact that $\operatorname{Pic}(X)$ is free. This proves assertions (i) and (ii). The remaining statements are special cases of~\cite[Proposition~4.15 and Corollary~4.16]{Ha2}. The self intersection number can be easily computed by using toric intersection theory in the ambient toric variety; compare \cite[Construction~III 3.3.4]{ArDeHaLa}. \end{proof} \begin{corollary} Let $X$ be a ${\mathbb Q}$-factorial complete normal variety with complexity-one torus action and Picard number one. If $X$ is locally factorial, then the divisor class group $\operatorname{Cl}(X)$ is free. \end{corollary} The following example shows that one can use Proposition \ref{Prop:FanoPicard}(iv) to create series of Fano varieties by altering the torsion part of the divisor class group $\operatorname{Cl}(X)$: \begin{example}\label{Ex:difftorsion} Set $l_{01}=7$, $l_{02}=1$, $l_{11}=5$ and $l_{21}=2$ as well as $w_{01}^0=1$, $w_{02}^0=3$, $w_{11}^0=2$ and $w_{21}^0=5$. According to Construction \ref{Con:R(A,n,L,m)} these data define one single Cox ring relation of the form $g_0=T_{01}^7T_{02}+T_{11}^5+T_{21}^2$. Since we have $$ w_{01}^0+w_{02}^0+w_{11}^0+w_{21}^0 \ = \ 11 \ > \ 10 \ = \ \deg(g_0)^0, $$ one can use these data to create Cox rings of Fano varieties. We provide some possible $\operatorname{Cl}(X)$-gradings, given by the matrices $Q_i$, defining Del Pezzo ${\mathbb K}^*$-surfaces with fixed grading in the free part of the divisor class group and varying torsion part of the class group $\operatorname{Cl}(X)^t$: \begin{align*} &Q_1 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \end{pmatrix}, &&\operatorname{Cl}(X_1)={\mathbb Z}; \\ &Q_2 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \overline 0 & \overline 2 & \overline{1}& \overline 1 \end{pmatrix}, &&\operatorname{Cl}(X_2)={\mathbb Z}\oplus{\mathbb Z}/3{\mathbb Z}; \\ &Q_3 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \overline 2 & \overline 1 & \overline 3& \overline 3 \end{pmatrix}, &&\operatorname{Cl}(X_3)={\mathbb Z}\oplus{\mathbb Z}/9{\mathbb Z}; \\ &Q_4 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \overline 0 & \overline 1 & \overline{9}& \overline 6 \end{pmatrix}, &&\operatorname{Cl}(X_4)={\mathbb Z}\oplus{\mathbb Z}/11{\mathbb Z}; \\ &Q_5 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \overline 0 & \overline 3 & \overline{11}& \overline 8 \end{pmatrix}, &&\operatorname{Cl}(X_5)={\mathbb Z}\oplus{\mathbb Z}/13{\mathbb Z}; \\ &Q_6 \ = \ \begin{pmatrix} 1 & 3 & 2 & 5\\ \overline 0 & \overline 7 & \overline{15}& \overline{12} \end{pmatrix}, &&\operatorname{Cl}(X_6)={\mathbb Z}\oplus{\mathbb Z}/17{\mathbb Z}. \end{align*} Note that in this situation not every group of the form ${\mathbb Z}\oplus{\mathbb Z}/k{\mathbb Z}$, $k\in{\mathbb N}_{>0}$, can be realized as divisor class group. \end{example} In Example \ref{Ex:difftorsion} the numbers $\ell_i:=\mathrm{gcd}(l_{i1},\ldots,l_{in_i})$ are pairwise coprime, namely $\ell_0=1,\;\ell_1=2$ and $\ell_2=5$. This allows the case $\operatorname{Cl}(X_1)={\mathbb Z}$; see \cite[Theorem~1.9]{HaHeSu}. If the numbers $\ell_i$ are not pairwise coprime, then there is always non trivial torsion in the divisor class group as the following lemma shows. \begin{lemma}\label{ggTPicind} Set $\ell_i:=\mathrm{gcd}(l_{i1},\ldots,l_{in_i})$. Then all numbers $\mathrm{gcd}(\ell_i,\ell_j)$, where $0\leq i\neq j\leq r,$ divide $|\operatorname{Cl}(X)^t|$ and the Picard index $\mu$. In particular this holds for $\mathrm{lcm}_{j\neq i}(\mathrm{gcd}(\ell_i,\ell_j))$. \end{lemma} \begin{proof} According to \cite[Theorem~1.5]{HaHe} the divisor class group $\operatorname{Cl}(X)$ is isomorphic to ${\mathbb Z}^{n+m}/\mathrm{im}(P^*)$ where $P^*$ is dual to $P\colon{\mathbb Z}^{n+m}\to{\mathbb Z}^{n+m-1}$ given by a matrix of the form $$ P \ = \ \begin{pmatrix} -l_0 & l_1 & \ldots & 0 & 0\\ \vdots & & \ddots & \vdots &\\ -l_0 & 0 & \ldots & l_r & 0\\ d_0 & d_1 & \ldots & d_r & d' \end{pmatrix}, $$ with $l_i=(l_{i0},\ldots,l_{in_i})$ and some integral block matrices $d_i$ and $d'$. Consequently $|\operatorname{Cl}(X)^t|$ is the product of all elementary divisors of $P$ which implies that $\mathrm{gcd}(\ell_0,\ell_j)$ divides $|\operatorname{Cl}(X)^t|$. By an elementary row transformation we obtain the analogous result for $\mathrm{gcd}(\ell_i,\ell_j)$ where $0\leq i,j\leq r$, $i\neq j$. Since $|\operatorname{Cl}(X)^t|$ divides the Picard index $\mu$, the assertion follows. \end{proof} \begin{remark} One can even prove that $\mathrm{lcm}_{0\leq j\leq r}(\prod_{i\neq j}\mathrm{gcd}(\ell_i,\ell_j))$ divides $|\operatorname{Cl}(X)^t|$ (see for example surface number $3$ in Theorem \ref{Theo:surfaces<=6}). \end{remark} \section{Effective bounds} First we consider the case $n_0 = \ldots = n_r=1$, that means that each relation $g_{i}$ of the Cox ring $\mathcal{R}(X)$ depends only on three variables. Then we have $n=r+1$ and consequently $m=d-1$. Furthermore, we may write $T_i$ instead of $T_{i1}$ and $w_i$ instead of $w_{i1}$, etc.. In this setting, we obtain the following bounds for the numbers of possible varieties~$X$ (Fano or not). \begin{proposition} \label{prop:Finite3Var} For any pair $(d,\mu) \in {\mathbb Z}^2_{>0}$ there is, up to deformation equivalence, only a finite number of complete $d$-dimensional varieties with Picard number one, Picard index $[\operatorname{Cl}(X):\operatorname{Pic}(X)] = \mu$ and Cox ring of the form $$ {\mathbb K}[T_0,\ldots, T_r,S_1,\ldots, S_m] \ / \ \langle{\alpha_{i} T_i^{l_i} +\alpha_{i+1} T_{i+1}^{l_{i+1}} +\alpha_{i+2} T_{i+2}^{l_{i+2}}; \; 0 \leq i \leq r-2} \rangle. $$ In this situation we have $r <\mu+\xi(\mu)-1$ where $\xi(\mu)$ denotes the number of primes smaller than $\mu$. Moreover for $w_i^0\in{\mathbb Z}_{>0}$ and $u_k^0\in{\mathbb Z}_{>0}$ where $0 \leq i \leq r$, $1 \leq k \leq m$, and the exponents $l_i$ one has $$ l_i\leq\mu, \qquad w_i^0 \ \leq \ \mu^r, \qquad u_k^0 \ \leq \ \mu. $$ \end{proposition} \begin{proof} Consider the total coordinate space $\overline X\subseteq{\mathbb K}^{r+1+m}$ and the quotient $p\colon\widehat X\to X$ as well as the points $\overline x(k)\in\widehat X$ having the $(r+k)$-th coordinate one and all others zero. Set $x(k):=p(\overline x(k))$. Then $u_k^0$ divides the order of the local class group $\operatorname{Cl}(X,x(k))$. In particular we have $u_k^0\leq\mu$.\\ For each $0\leq i\leq r$ fix a point $\overline y(i)=(\overline y_0,\ldots,\overline y_r,0,\ldots,0)$ in $\widehat X$ such that $\overline y_i=0$ and $\overline y_j\neq 0$ for $i\neq j$, and set $y_i:=p(\overline y(i))$. Then we obtain $$ \mathrm{gcd}(w_j^0,j\neq i) \; \mid \; |\operatorname{Cl}(X,y(i))|. $$ By Lemma \ref{ggTPicind} we have $\mathrm{lcm}_{j\neq i}(\mathrm{gcd}(l_i,l_j))\mid |\operatorname{Cl}(X)^t|$. Now consider $l_i'$ such that $l_i=\mathrm{lcm}_{j \neq i}(\mathrm{gcd}(l_i,l_j))\cdot l_i'$. Then the homogeneity condition $l_iw_i^0=l_jw_j^0$ gives $l_i'\mid w_j^0$ for all $j\neq i$ and consequently $l_i'\mid\mathrm{gcd}(w_j^0,j\neq i)$. Since $l_i= l_i'\cdot\mathrm{lcm}_{j\neq i}(\mathrm{gcd}(l_i,l_j))$ we can conclude $l_i\leq\mu$ by using the formula $$ [\operatorname{Cl}(X):\operatorname{Pic}(X)] \ = \ \mathrm{lcm}_{x\in X}(\mathrm{gcd}(w^0;\; w\in W_{\overline X}))\cdot |\operatorname{Cl}(X)^t| $$ of Proposition \ref{Prop:FanoPicard}(ii). Since the $l_i'$ are pairwise coprime we obtain $l_0'\cdots l_r'\mid\gamma^0$ and $l_0'\cdots l_r'\mid\mu$ where $\gamma^0:=\deg(g_0)^0=l_iw_i^0$. From $l_iw_i^0=l_jw_j^0$ we deduce that $$ l_i=l_0\frac{w_0^0}{w_i^0}= l_0\frac{w_0^0\cdots w_{i-1}^0}{w_1^0\cdots w_i^0} =\eta_i\cdot\frac{\mathrm{gcd}(w_0^0,\ldots,w_{i-1}^0)}{\mathrm{gcd}(w_0^0,\ldots, w_i^0)} \leq \mu, $$ where $1\leq \eta_i\leq \mu$. In particular the last fraction is smaller than $\mu$. All in all this gives us \begin{align*} w_0^0 &= \frac{w_0^0}{\mathrm{gcd}(w_0^0,w_1^0)}\cdot \frac{\mathrm{gcd}(w_0^0,w_1^0)}{\mathrm{gcd}(w_0^0,w_1^0,w_2^0)}\cdot \ldots\cdot \frac{\mathrm{gcd}(w_0^0,\ldots,w_{r-2}^0)}{\mathrm{gcd}(w_0^0,\ldots,w_{r-1}^0)} \cdot \mathrm{gcd}(w_0^0,\ldots,w_{r-1}^0)\\ &\leq \mu^{r-1}\cdot\mu=\mu^r. \end{align*} Analogously we get the boundedness for all $w_i^0$. Now let $q$ be the number of $l_i'$ that are greater than one. Since all $l_i'$, $0\leq i\leq r$, are coprime $q$ is bounded by $\xi(\mu)$ the number of primes smaller than $\mu$. To avoid the toric case we assume $l_i\neq 1$ for all $0\leq i\leq r$. Consequently if $l_i'=1$, then there is at least one $0\leq j\leq r$ such that $\mathrm{gcd}(l_i,l_j)>1$. Since $\mathrm{gcd}(l_i,l_j)$ divides $\mu$ we get $r+1-q<\mu$ as a rough bound. All in all we get $ r+1 \ = \ r+1-q+q \ < \ \mu+\xi(\mu). $ \end{proof} \begin{proof}[Proof of Theorem \ref{Theo:sepquot}] Let $X$ be a variety as in Theorem \ref{Theo:sepquot}. Then each monomial of the Cox ring relations depends on only one variable, i.e. $n_i=1$ for $0\leq i\leq r$; for details see \cite[Theorem~1.2]{HaSu}. Consequently Proposition \ref{prop:Finite3Var} provides bounds for the discrete data such as the non torsion parts of the weights $w_{ij}^0$ and $u_k^0$, the exponents $l_{ij}$ and the number of Cox ring relations $r$. Since $|\operatorname{Cl}(X)^t|\leq\mu$ holds, the number of possibilities for the torsion part of the grading is also restricted which implies the assertion. \end{proof} \begin{theorem} \label{Th:FiniteIndex} Let $X$ be a Fano variety with complexity-one torus action as introduced in Construction \ref{Con:varX}. Fix the dimension $d = \dim(X)=m+n+r$ and the Picard index $\mu = [\operatorname{Cl}(X):\operatorname{Pic}(X)]$. Then the number of Cox ring relations $r$, the free part of the degree of the relations $\gamma^0$, the weights $w_{ij}^0,\; u_k^0$ and the exponents $l_{ij}^0$, where $0 \le i \le r,\;1 \le j \le n_i$ and $1\leq k\leq m$, are bounded. In particular one obtains the following effective bounds: We have $$ u_k^0 \ \le \ \mu \quad \text{for } 1 \le k \le m \quad \text{and} \quad |\operatorname{Cl}(X)^t|\leq\mu. $$ Moreover, the handling of the remaining data can be organised in five cases, where $\xi(x)$ denotes the number of primes smaller than $x$. \begin{enumerate} \item[(i)] Suppose that $r = 0,1$ holds. Then $n + m \le d+1$ holds and one has the bounds $$ w_{ij}^0 \ \le \ \mu \quad\text{for } 0 \le i \le r \text{ and } 1 \le j \le n_i, $$ and the Picard index is given by $$ \mu \ = \ \mathrm{lcm}(w_{ij}^0,u_k^0; \;0 \le i \le r, 1 \le j \le n_i, 1 \le k \le m ) \cdot|\operatorname{Cl}(X)^t|. $$ \item[(ii)] Suppose that $r \ge 2$ and $n_0=1$ hold. Then $r \le \mu+\xi(\mu)-1$ and $n=r+1$ and $m=d-1$ hold and one has $$ \qquad\qquad w_{i1}^0 \ \le \ \mu^r, \quad l_{i1} \ \mid \ \mu \qquad \text{for } 0 \le i \le r, \qquad \gamma^0 \ \le \ \mu^{r+1}, $$ and the Picard index is given by $$ \mu \ = \ \mathrm{lcm}(\mathrm{gcd}_i(w_{j1}^0; \; i\neq j),u_k^0; \; 0 \le i \le r , 1 \le k \le m ) \cdot|\operatorname{Cl}(X)^t|. $$ \item[(iii)] Suppose that $r \ge 2$ and $n_0 > n_1=1$ hold. Then we may assume $l_{11} \geq \ldots \geq l_{r1} \ge 2$, we have $r \le\mu+ \xi(6d\mu)-1$ and $n_0+m = d$ and the bounds $$ w_{01}^0,\ldots,w_{0n_0}^0 \ \le \ \mu, \qquad l_{01},\ldots,l_{0n_0} \ \leq \ 6d\mu, \qquad \gamma^0 \ < 6d\mu, $$ $$ w_{11}^0\ < \ 2d\mu, \quad w_{21}^0\ < \ 3d\mu, \qquad w_{i1}^0,\; l_{i1} \ < \ 6d\mu \quad \text{for } 1 \le i \le r, $$ and the Picard index is given by $$ \mu \ = \ \mathrm{lcm}(w_{0j}^0, \mathrm{gcd}(w_{11}^0,\ldots,w_{r1}^0), u_k^0; \; 1 \le j \le n_0, 1 \le k \le m ) \cdot|\operatorname{Cl}(X)^t|. $$ \item[(iv)] Suppose that $n_1 > n_2 = 1$ holds. Then we may assume $l_{21} \geq \ldots \geq l_{r1} \ge 2$, we have $r \le\mu+ \xi(2(d+1)\mu)-1$ and $n_0+n_1+m = d+1$ and the bounds $$ w_{ij}^0 \ \le \ \mu \quad \text{for } i=0,1 \text{ and } 1 \le j \le n_i, \qquad w_{21}^0 \ < \ (d+1)\mu, $$ $$ \gamma^0,\;w_{ij}^0,l_{ij} \ < \ 2(d+1)\mu \quad \text{for } 0 \le i \le r ,\text{ and } 1 \le j \le n_i, $$ and the Picard index is given by $$ \mu \ = \ \mathrm{lcm}(w_{ij}^0, u_k^0; \; 0 \le i\le 1, 1 \le j \le n_i, 1 \le k \le m) \cdot |\operatorname{Cl}(X)^t|. $$ \item[(v)] Suppose that $n_2 > 1$ holds and let $s$ be the maximal number with $n_{s}>1$. Then one may assume $l_{s+1,1} \geq \ldots \geq l_{r1} \ge 2$, we have $s\leq d$, $r \le\mu+ \xi((d+2)\mu)+d-1$ and $n_0+ \ldots + n_s+m = d+s$ and the bounds $$ w_{ij}^0 \ \le \ \mu, \quad \text{for } 0 \le i \le s, \qquad \gamma^0 \ < \ (d+2)\mu, $$ $$ w_{ij}^0, l_{ij} \ < \ (d+2)\mu \quad \text{for } 0 \le i \le r \text{ and } 1 \le j \le n_i, $$ and the Picard index is given by $$ \mu \ = \ \mathrm{lcm}(w_{ij}^0, u_k^0; \; 0 \le i \le s, 1 \le j \le n_i, 1 \le k \le m) \cdot |\operatorname{Cl}(X)^t|. $$ \end{enumerate} Note that assertion (i) and (ii) do not require the Fano condition. \end{theorem} The remaining part of this chapter is devoted to the proofs of the main statements of this paper. To prove Theorem \ref{Th:FiniteIndex} we need the following essential lemma. \begin{lemma} \label{Lem:1relation} Consider the ring ${\mathbb K}[T_{ij}; \; 0 \le i \le 2, \; 1 \le j \le n_i][S_1,\ldots,S_k] / \bangle{g} $ where $n_0 \ge n_1 \ge n_2 \ge 1$ holds and let $K$ be a finitely generated abelian group of the form $K={\mathbb Z}\oplus K^t$ with torsion part $K^t$. Suppose that $g$ is homogeneous with respect to the $K$-grading of ${\mathbb K}[T_{ij},S_k]$ given by $\deg \, T_{ij} = :w_{ij}=w_{ij}^0 + w_{ij}^t \in K$ with $w_{ij}^0 \in{\mathbb Z}_{>0}$ and $\deg \, S_k = :u_k=u_k^0+u_k^t \in K$ with $u_k^0 \in{\mathbb Z}_{> 0}$, and assume \begin{eqnarray*} \deg (g)^0 & < & \sum_{i=0}^2\sum_{j=1}^{n_i}w_{ij}^0 \ + \ \sum_{i=1}^m u_i^0. \end{eqnarray*} Let $\mu \in {\mathbb Z}_{>1}$, assume $w_{ij}^0 \le \mu$ whenever $n_i > 1$, $1 \le j \le n_i$ and $u_k^0 \le \mu$ for $1 \le k \le m$ and set $d := n_0+n_1+n_2+m-2$. Depending on the shape of $g$, one obtains the following bounds. \begin{enumerate} \item[(i)] Suppose that $g = \eta_0 T_{01}^{l_{01}} \cdots T_{0n_0}^{l_{0n_0}} + \eta_1 T_{11}^{l_{11}} + \eta_2 T_{21}^{l_{21}}$ with $n_0 > 1$ and coefficients $\eta_i \in {\mathbb K}^*$ holds. If we have $l_{11} > l_{21} \ge 2$ and $\mathrm{gcd}(l_{11},l_{21})\mid \mu$, then one has $$ w_{11}^0\ < \ 2d\mu, \quad w_{21}^0\ < \ 3d\mu, \quad \quad l_{22}, l_{21},\deg (g)^0 \ < \ 6d\mu. $$ If we have $l_{11}=l_{21}\geq 2$, then one has $$ l_{11} \, ,w_{11}^0\, ,l_{21}\, , w_{21}^0\, , \deg (g)^0\ \leq\ \mu. $$ \item[(ii)] Suppose that $g = \eta_0 T_{01}^{l_{01}} \cdots T_{0n_0}^{l_{0n_0}} + \eta_1 T_{11}^{l_{11}} \cdots T_{1n_1}^{l_{1n_1}} + \eta_2 T_{21}^{l_{21}}$ with $n_1 > 1$ and coefficients $\eta_i \in {\mathbb K}^*$ holds and we have $l_{21} \ge 2$. Then one has $$ \qquad\qquad w_{21}^0 \ < \ (d+1)\mu, \qquad\qquad \deg (g)^0 \ < \ 2(d+1)\mu. $$ \end{enumerate} \end{lemma} \begin{proof} We prove~(i). Set for short $c := (n_0+m)\mu = d\mu$. Then, using homogeneity of $g$ and the assumed inequality, we obtain $$ l_{11}w_{11}^0 \ = \ l_{21}w_{21}^0 \ = \ \deg (g)^0 \ < \ \sum_{i=0}^2\sum_{j=1}^{n_i}w_{ij}^0 + \sum_{i=1}^m u_i ^0 \ \le \ c+w_{11}^0+w_{21}^0. $$ First have a look at the case $l_{11} > l_{21} \ge 2$. Plugging this into the above inequalities, we arrive at $2 w_{11}^0 < c + w_{21}^0$ and $w_{21}^0 < c + w_{11}^0$. We conclude $w_{11}^0 < 2c$ and $w_{21}^0 < 3c$. Consequently we obtain $$ \deg (g)^0 \ < \ c + w_{11}^0 + w_{21}^0 \ < \ 6c=6d\mu. $$ If we have $l_{11}=l_{21}$, the homogeneity condition $l_{11}w_{11}^0=l_{21}w_{11}^0$ gives us $w_{11}^0=w_{21}^0$. Thus we have $\mathrm{gcd}(w_{11}^0,w_{21}^0)=w_{11}^0=w_{21}^0\mid\mu$ and by assumption $\mathrm{gcd}(l_{11},l_{21})=l_{21}=l_{11}\mid\mu$. Consequently $l_{11},w_{11}^0 ,l_{21} , w_{21}^0 ,\deg (g)^0\leq\mu$. \\ We prove (ii). Here we set $c := (n_0+n_1+m)\mu = (d+1)\mu$. Then the assumed inequality gives $$ l_{21}w_{21}^0 \ = \ \deg (g)^0 \ < \ \sum_{i=0}^1\sum_{j=1}^{n_i}w_{ij}^0+ \sum_{i=1}^m u_i^0+ w_{21}^0 \ \le \ c+w_{21}^0. $$ Since we assumed $l_{21} \geq 2$, we can conclude $w_{21}^0 < c$. This in turn gives us $\deg (g)^0 < 2c$. \end{proof} \begin{proof}[Proof of Theorem \ref{Th:FiniteIndex}] As before, we denote by $\overline{X} \subseteq {\mathbb K}^{n+m}$ the total coordinate space and we consider the quotient $p \colon \widehat{X} \to X$. We first discuss the case that $X$ is a toric variety. Then the Cox ring is a polynomial ring, $\mathcal R(X) = {\mathbb K}[S_1,\ldots,S_m]$. For each $1 \le k \le m$, consider the point $\overline x(k) \in \widehat{X}$ having the $k$-th coordinate one and all others zero and set $x(k) := p(\overline {x}(k))$. Then, by~Proposition~\ref{Prop:FanoPicard}, the order of the local class group $\operatorname{Cl}(X,x(k))$ is divisible by $u_k^0$. Together with Proposition \ref{Prop:FanoPicard}(ii) we get $u_k^0\le \mu$ for $1 \le k \leq m$ and $|\operatorname{Cl}(X)^t|\le\mu $ which settles assertion~(i). Now we treat the non-toric case, which means $r \ge 2$. Note that we have $n \ge 3$. The case $n_0=1$ is done in Proposition~\ref{prop:Finite3Var}, which proves assertion (ii). Hence, we are left with $n_0>1$. For every $i$ with $n_i > 1$ and every $1 \le j \le n_i$, there is the point $\overline{x}(i,j) \in \widehat{X}$ with $ij$-coordinate $T_{ij}$ equal to one and all others equal to zero, and thus we have the point $x(i,j) := p(\overline{x}(i,j)) \in X$. Moreover, for every $1 \le k \le m$, we have the point $\overline{x}(k) \in \overline{X}$ having the $k$-coordinate $S_k$ equal to one and all others zero; we set $x(k):=p(\overline{x}(k))$. Proposition~\ref{Prop:FanoPicard} provides the bounds \begin{equation}\label{ie1} w_{ij}^0 \ \le \ \mu, \quad u_k^0 \ \le \ \mu \qquad \text{for } \quad n_i > 1, \, 0\leq i\leq r, \, 1 \le j \le n_i, \, 1 \le k \le m. \end{equation} Let $0 \le s \le r$ be the maximal number with $n_{s} > 1$. Then $g_{s-2}$ is the last polynomial such that each of its three monomials depends on more than one variable. For any $t \ge s$, we have the ``cut ring'' \begin{eqnarray*} R_t & := & {\mathbb K}[T_{ij},S_k] \ / \ \bangle{g_0, \ldots,g_{t-2}} \end{eqnarray*} where $0 \le i \le t$, $1 \le j \le n_i$, $1\le k\le m$ and the relations $g_{i}$ depend on only three variables as soon as $i > s$ holds. For the free part of the degree $\gamma^0$ of the relations we have \begin{eqnarray*} (r-1)\gamma^0 & = & (t-1)\gamma^0 \ + \ (r-t)\gamma^0 \\ & = & (t-1)\gamma^0 \ + \ l_{t+1,1}w_{t+1,1}^0 + \ldots + l_{r1}w_{r1}^0 \\ & < & \sum_{i=0}^r\sum_{j=1}^{n_i}w_{ij}^0 \ + \ \sum_{i=1}^m u_i^0 \\ & = & \sum_{i=0}^t \sum_{j=1}^{n_i}w_{ij}^0 \ + \ w_{t+1,1}^0+ \ldots + w_{r1}^0 \ + \ \sum_{i=1}^m u_i^0. \end{eqnarray*} Note that the inequality is derived from the Fano condition of Proposition \ref{Prop:FanoPicard}(iv). Since $l_{i1}w_{i1}^0 > w_{i1}^0$ holds in particular for $t+1 \le i \le r$, we derive from this the inequality \begin{eqnarray}\label{ie2} \gamma^0 & < & \frac{1}{t-1} \left( \sum_{i=0}^t\sum_{j=1}^{n_i}w_{ij}^0 \ + \ \sum_{i=1}^m u_i^0 \right). \end{eqnarray} To obtain the bounds in assertions~(iii) and~(iv), we consider the cut ring $R_t$ with $t=2$ and apply Lemma~\ref{Lem:1relation} and Proposition \ref{Prop:FanoPicard}; note that we have $d = n_0+n_1+n_2+m-2$ for the dimension $d = \dim(X)$ and that $l_{21} \ge 2$ is due to the fact that $X$ is non-toric. The bounds $w_{i1}^0, l_{i1} < 6d\mu$ for $3 \le i \le r$ in assertion~(iii) follow from $\gamma^0 < 6d\mu$. Similarly $w_{ij}^0,l_{ij} < 2(d+1)\mu$ for $0\leq i\leq r$, $1\leq j\leq n_i$ in assertion~(iv) follow from $ \gamma^0 < 2(d+1) \mu$. We still have to prove the restriction for the number of relations, which means bounding $r$. Recall from Lemma \ref{ggTPicind} the definition $\ell_i:=\mathrm{gcd}(l_{i1},\ldots,l_{in_i})$ and set $\ell_i=\mathrm{lcm}_{0\leq j\neq i\leq r}(\mathrm{gcd}(\ell_i,\ell_j))\cdot \ell_i'$. Then $\ell_0',\ldots, \ell_r'$ are coprime. For $i\geq 1$ we have $n_i=1$. Thus analogously to the proof of Proposition \ref{prop:Finite3Var} we get $r+1=r+1-q+q\le\mu+\xi(6d\mu)$ where $q$ is the number of $\ell_i'$ that are greater than one and satisfy $n_i=1$. For the bound in assertion~(iv) the same argument yields $r+1=r+1-q+q\le\mu+\xi(2(d+1)\mu)$. To obtain the bounds in assertion~(v), we consider the cut ring $R_t$ with $t=s$. Using $n_i=1$ for $i \ge t+1$ and applying the inequalities \ref{ie1} and \ref{ie2}, we can derive an upper bound for the degree of the relation as follows: $$ \gamma^0 \ < \ \frac{(n_0 + \ldots + n_t + m) \mu}{t-1} \ = \ \frac{(d + t) \mu}{t-1} \ \le \ (d + 2) \mu. $$ We have $w_{ij}^0l_{ij} \le \gamma^0$ for any $0 \le i \le r$ and any $1 \le j \le n_i$, which implies that all $w_{ij}^0$ and $l_{ij}$ are bounded by $(d+2)\mu$. Since $n_0,\ldots,n_{s-1}>1$ holds, the number $s$ is bounded by $s=2s-(s-1)-1\le d$. Consequently we obtain $ r+1=r+1-s-q+s+q\le\mu+\xi((d+2)\mu)+d, $ where $q$ is defined as above. Finally, we have to express the Picard index $\mu$ in terms of the free part of the weights $w_{ij}^0$, $u_k^0$ and the torsion part $\operatorname{Cl}(X)^t$ as claimed in the assertions. This is a direct application of the formula of Proposition~\ref{Prop:FanoPicard}. \end{proof} \begin{proof}[Proof of Theorem \ref{Th:asymptotic}] Theorem \ref{Th:FiniteIndex} provides bounds for the exponents and the number of relations as well as for the free part of the weights and the torsion part of $\operatorname{Cl}(X)$. Since we have $|\operatorname{Cl}(X)^t|\leq \mu$ the possibilities for the torsion part of the weights are also restricted. One computes that the number $\delta(d,\mu)$ of different deformation types is bounded above by $$ \mu^{\mu^2+3\mu+\xi(\mu)^2+\xi(6d\mu)+5d} (6d\mu)^{2\mu+2\xi(6d\mu)+3d-2} $$ which leads to the results of Theorem \ref{Th:asymptotic}. \end{proof} \begin{proof} [Proof of Theorems \ref{Theo:surfaces<=6} and \ref{thm:3fano2}] For fixed $d$ and $\mu$ Theorem \ref{Th:FiniteIndex} bounds the number of possible data $l_{ij}$, $w_{ij}^0$, $u_k^0$, belonging to Fano varieties. We identify all these constellations by a computer based algorithm. Since $|\operatorname{Cl}(X)^t|\leq \mu$ holds, there is only a finite number of possibilities for the torsion part of the weights that we have to check. By this procedure we obtain the tables of \ref{Theo:surfaces<=6} and \ref{thm:3fano2}. We claim that any two of the listed Cox rings do not describe varieties that are isomorphic to each other. Two minimal systems of homogeneous generators of the Cox ring contain (up to reordering) the same free parts of generator degrees $w_{ij}^0$, $u_k^0\in{\mathbb Z}$. Consequently they are invariant under isomorphy. Furthermore the exponents $l_{ij}>1$ represent the orders of all finite non-trivial isotropy groups of one-codimensional orbits of the action $T$ on $X$; see \cite[Theorem~1.3]{HaSu}. Moreover, since none of the listed Cox rings is polynomial the varieties are all non-toric. This implies that every complexity-one action is maximal and consequently can be assigned to a maximal torus in ${\rm Aut}(X)$. Note that ${\rm Aut}(X)$ is also acting effectively on $X$. Since the maximal tori of ${\rm Aut}(X)$ are all conjugated the varieties with complexity-one torus action are isomorphic if and only if they are $T$-equivariantly isomorphic. Thus, running through the exponents $l_{ij}$ we see that any two of the varieties listed in Theorem \ref{Theo:surfaces<=6} are not isomorphic. In case of Theorem \ref{thm:3fano2} there is some more work to do. There are not isomorphic threefolds varying only in the torsion part of the weights, see for example number $2$, $3$ and $4$. In these cases, comparing the torsion parts of the gradings shows that it is not possible to install a $\operatorname{Cl}(X)$-graded ring isomorphism between the Cox rings of two different threefolds. As an example we consider the threefolds number $2$ and $3$: Let $D_{2}$ be a prime divisor, representing $\deg(T_2)\in\operatorname{Cl}(X)$ and let $E_1$ be a prime divisor, representing $\deg(S_1)\in\operatorname{Cl}(X)$ . Then $D_{2}$ has isotropy group of order $l_{2}=3$ and $E_1$ has infinite isotropy. In case of threefold number $2$ the term $D_{2}-E_1$ represents a non-trivial torsion element whereas in case of threefold number $3$ it is the zero element in $\operatorname{Cl}(X)$. Thus, these two varieties are not isomorphic. Analogously we proceed with all other cases to obtain finally the list of Theorem \ref{thm:3fano2}. Finally, we apply \cite[Corollary 4.9]{Ha2} to compute the Gorenstein index $\iota(X)$ for all listed varieties, i.e. we have to find the smallest integer $\iota(X)$ such that $\iota(X)\cdot K_X$ is contained in all local divisor class groups $\operatorname{Cl}(X,x)$; see also Proposition \ref{Prop:FanoPicard}. \end{proof} \begin{acknowledgements} The author would like to thank J\"urgen Hausen and the referees for their valuable comments, remarks and references that improved this article a lot. \end{acknowledgements}
\section{Introduction} The tremendous experimental progress in performing scattering experiments has evolved to the point where it is nowadays possible to study collisions between particles in the laboratory over an energy range of about 25 orders of magnitude. The collisions of highest energy are produced by modern charged-particle accelerators reaching the TeV range, while the collisions of lowest energy are studied in ultracold atomic quantum gases going all the way down to the nK regime. In both types of collision experiments scattering resonances play an important role. In high-energy collisions, a resonance in the cross section caused by the formation of an intermediate bound state is a direct way to detect previously unseen particles. In ultracold atomic scattering, the energy of an intermediate bound molecular state to be formed during the collision can sometimes be accurately tuned by applying an external magnetic field. As a result, the scattering length of low-energy $s$-wave collisions gets under full experimental control, giving rise to a unique quantum many-body environment with a completely tunable interaction parameter \cite{inguscio:08,stoof:09}. Because molecules are typically harder to manipulate than atoms and charged particles, the observations of resonances in molecular beam scattering have been limited to a few rare cases \cite{schutte:75a,toennies:79,skodje:00a,skodje:00b,qiu:06,Dong:Science327:1501}. However, in recent years rapid progress has been made in performing high-precision cold molecular scattering experiments due to the application of the Stark deceleration technique to the study of molecular collisions \cite{gilijamse:06}. A Stark decelerator operates according to the same principles as a linear charged-particle accelerator, where the dipolar or Stark force is used to decelerate neutral polar molecules with time-varying electric fields \cite{bethlem:99}. With the Stark decelerator it is possible to generate almost perfectly quantum-state selected molecular beams with a computer-controlled final velocity and a small longitudinal velocity spread. By applying this technique to the scattering of the OH radical with rare gas atoms, such as Xe \cite{gilijamse:06}, Ar \cite{scharfenberg:10}, and He \cite{kirste:10}, the threshold behavior for inelastic scattering into the first excited rotational levels of OH could be accurately determined. Excellent agreement was found with cross sections obtained from close-coupling calculations using {\it ab initio} potential energy surfaces (PESs) \cite{gilijamse:06,scharfenberg:10,scharfenberg:11}. In the same way, also cold inelastic collisions of OH radicals with D${}_2$ molecules were studied experimentally \cite{kirste:10}. In this article, we study in detail cold collisions between NH${}_3$ molecules and He atoms. The ammonia-He system is a van der Waals complex, and in general the (quasi-)bound states of such complexes are sensitive to the interaction potential. As a result, high-resolution spectroscopy on van der Waals complexes has been an important tool for increasing our understanding of intermolecular forces \cite{avoird:94,wormer:00a}. High-precision scattering experiments are a very promising additional tool for obtaining detailed information on potential energy surfaces. At higher scattering energies the short-range repulsive part of the interaction is mainly probed, while at very low collision energies the long-range part of the potential is dominant in determining the scattering behavior. Moreover, scattering resonances give important information on the energy of quasi-bound states that are sensitive to potential wells at mid-range interparticle distances. This shows that large parts of the potential energy surfaces can be accurately probed by cold collision experiments. Recent scattering experiments have indeed been able to distinguish between PESs that were only of good quality and PESs that were of excellent quality \cite{scharfenberg:10,scharfenberg:11}. A very different experiment in which the NH$_{3}$-He interaction plays an important role, is the trapping of NH$_{3}$ molecules inside He nanodroplets to perform high-resolution spectroscopy \cite{slipchenko:05}. Rotational energy transfer by cold collisions is an important process in various astrochemical environments, such as interstellar clouds and cold exoplanetary atmospheres. Since the first identification of NH$_{3}$ molecules in the interstellar medium \cite{cheung:68}, ammonia has been detected in several gas-phase astrochemical spectra. The rate coefficients of NH$_{3}$-He scattering are an important ingredient for a numerical modelling of astrochemical environments. This is one of the reasons why NH$_{3}$-He collisions have been studied experimentally \cite{oka:68, meyer:86, seelemann:88, schleipen:91, meyer:95} and theoretically \cite{green:76,billing:85,chen:97,wang:03,machin:05,yang:08} by several groups. The most recent scattering calculations have been performed with the potential energy surface of Hodges and Wheatley \cite{hodges:01}. However, in order to get agreement with experimentally determined virial coefficients, this potential had to be scaled by a rather large factor \cite{wang:03}. The same potential has also been used to theoretically study low-energy NH$_{3}$-He collisions, where strong scattering resonances were observed for various initial and final states of the NH$_{3}$ molecule \cite{yang:08}. Unfortunately, the initial state that is most suitable for Stark deceleration was not considered. This is namely the state $|j k \pm \rangle = |1 1 - \rangle$, where $j$ is the angular momentum of the ammonia molecule, $k$ is the projection on its threefold symmetry axis and $+/-$ refers to its symmetric/antisymmetric umbrella inversion tunneling state. For the energy level diagram of the ammonia molecule, see Fig.~\ref{figlevels}. Moreover, in the study of Ref.~\cite{yang:08} ammonia was treated as a rigid molecule, implying that the umbrella inversion motion of the NH$_{3}$ molecule was not considered. In this article we study all possible elastic and inelastic scattering processes at low collision energies, using $| 11 - \rangle$ as an initial state of the {\it para} NH$_{3}$ molecule. We show that particularly the inversion inelastic scattering to the $| 11 + \rangle$ state gives rise to pronounced resonant structures that are promising to be observed experimentally in crossed beam experiments. We start by introducing the theoretical framework for studying the atom-molecule collisions. After this, we present a new NH${}_3$-He potential using the most recent developments in electronic structure calculations. We describe the numerical methods to fit the potential, after which we present the calculations of the integral and differential cross sections. In both cross sections we find rapid variations as a function of energy, which are clear signs of resonant behavior. To determine the origin of these resonances we perform bound state calculations as well as reconstructions of the full scattering wavefunctions. The phase shifts and the lifetimes are also determined near resonance. Finally, we comment on the prospects of observing these scattering resonances in the NH${}_3$-He system in the near future. \begin{figure} \begin{center} \includegraphics[width=0.6\columnwidth]{figure1} \caption{\label{figlevels} Energy levels $|j k \pm \rangle$ of the ammonia molecule, where $j$ is the angular momentum of the molecule, $k$ is the projection on its threefold symmetry axis and $+/-$ refers to the symmetric/antisymmetric umbrella state. Throughout this article we use the $|1 1 - \rangle$ state as the initial state, so that we only consider {\it para} ammonia. We do not take into account hyperfine interactions. The collision energy is defined relative to the initial state. } \end{center} \end{figure} \section{Theory} To theoretically study the low-energy scattering of NH${}_3$ molecules with He atoms we briefly introduce three coordinate frames that are used in the calculations \cite{bladel:91}. These coordinate frames are discussed in more detail in the Appendix. The first frame is an orthonormal, right-handed space-fixed (`sf') laboratory frame located at the center of mass $\mathbf{Q}$ of the dimer. The coordinate $R$ is the length of the vector ${\bf R}$ that points from the center of mass $\mathbf{X}$ of the NH${}_3$ monomer to the He atom, while $\theta^{\rm sf}$ is the zenith angle of the vector $\mathbf{R}$ and $\phi^{\rm sf}$ is the azimuth angle in the space-fixed frame. The second frame is an orthonormal, right-handed body-fixed (`bf') dimer frame, also centered at the center of mass of the dimer. As explained in the Appendix, this frame is obtained by a rotation that aligns its $z$ axis with the vector ${\bf R}$. The third frame is an orthonormal, right-handed monomer-fixed (`mf') frame centered at the center-of-mass of the NH${}_3$ molecule, whose $z$ axis is aligned with the symmetry axis of the ammonia molecule. This monomer frame is obtained from the space-fixed frame by rotating over the three Euler angles $\zeta^{\rm sf} = (\alpha^{\rm sf},\beta^{\rm sf},\gamma^{\rm sf})$. Here, $\alpha^{\rm sf}$ and $\beta^{\rm sf}$ are the azimuth and zenith angles of the ammonia $C_3$ symmetry axis in the space-fixed frame, while $\gamma^{\rm sf}$ describes the rotation of the NH${}_3$ molecule about this axis. The Euler angles of the monomer frame can also be given with respect to the body-fixed frame, and are then denoted by $\zeta^{\rm bf} = (\alpha^{\rm bf},\beta^{\rm bf},\gamma^{\rm bf})$. Finally, we introduce the umbrella or inversion angle $\rho$ of the ammonia molecule, which is the angle between the $z$ axis of the monomer frame and a vector pointing from the N atom to one of the H atoms, so that $\rho = \pi/2$ corresponds to a planar NH${}_3$ geometry. The Hamiltonian of the NH${}_3$-He system can now be written as \cite{bladel:92a} \begin{eqnarray}\label{eq:ham} \hat{H} &=& \hat{H}_{\rm mol} - \frac{1}{2 \mu R} \frac{\partial^2}{\partial R^2}R + \frac{1}{2 \mu R^2}\left[ \hat{J}^2 + \hat{j}^2 - 2 \hat{\mathbf{j}}\cdot \hat{\mathbf{J}} \right] \nonumber \\ && + V_{\rm int}(R, \beta^{\rm bf}, \gamma^{\rm bf}, \rho), \end{eqnarray} where throughout the article we set $\hbar = 1$, $\hat{H}_{\rm mol}$ is the Hamiltonian of the NH${}_3$ molecule, $\mu$ is the reduced mass of the atom-molecule complex, $\hat{\mathbf{j}}$ is the angular momentum operator of the NH${}_3$ monomer with respect to the body-fixed frame, $\hat{\mathbf{J}}$ is the total angular momentum operator also with respect to the body-fixed frame, and $V_{\rm int}$ is the interaction energy. We consider the interaction potential to depend on four coordinates, which implies that we assume the N-H bond length to be fixed and NH${}_3$ to keep its threefold symmetry. The Hamiltonian of the NH${}_3$ molecule includes the monomer's rotation, as well as the kinetic and potential energy of its umbrella motion \cite{bladel:92a}, namely \begin{eqnarray} \hat{H}_{\rm mol} &=& \sum_{\lambda=x,y,x} \frac{\hat{j}^2_{\lambda}}{2I_{\lambda\lambda}(\rho)} -\frac{1}{2\sqrt{g(\rho})} \frac{\partial}{\partial \rho} \frac{\sqrt{g(\rho})}{I_{\rho\rho}(\rho)} \frac{\partial}{\partial\rho}\nonumber\\ &&+V_{\rm umb}(\rho),\label{eq:hammon} \end{eqnarray} where $I_{xx}(\rho)$, $I_{yy}(\rho)$ and $I_{zz}(\rho)$ are the moments of inertia of the threefold symmetric ammonia molecule with respect to the monomer frame axes, see e.g. Ref. \cite{bladel:92a}, while $I_{\rho\rho}(\rho) = 3 m_{\rm H}r_0^2 (\cos^2 \rho + \eta \sin^2 \rho)$ with $ m_{\rm H}$ the hydrogen mass, $r_0 = 1.9099 a_0$ the fixed N-H bond length \cite{huang:08}, $\eta=m_{\rm N}/(3 m_{\rm H} + m_{\rm N})$ and $ m_{\rm N}$ the nitrogen mass. Moreover, we have that $g(\rho) = I_{xx}I_{yy}I_{zz}I_{\rho\rho}$, while the potential energy for the umbrella motion $V_{\rm umb}(\rho)$ leads to a double well potential that we model by \begin{eqnarray} V_{\rm umb}(\rho) = \frac{k_{\rho}}{2}\left(\rho - \frac{\pi}{2}\right)^2+ a_{\rho} \exp \left[ - b_{\rho} \left(\rho - \frac{\pi}{2}\right)^2 \right] \end{eqnarray} with the parameters $k_{\rho}=90 \, 651$ cm${}^{-1}$rad${}^{-2}$, $a_{\rho} = 23 \, 229$ cm${}^{-1}$ and $b_{\rho}= 3.1846$ rad${}^{-2}$. The resulting double well potential gives rise to umbrella vibration levels of which two levels have an energy below that of the barrier at the planar ammonia geometry \cite{bladel:92a}. Moreover, each of these two vibration levels splits into a pair of energy levels due to tunneling. The parameters of the umbrella potential are chosen such that the experimental energy splitting between the two tunnel states in the vibrational ground state, as well as the experimental splittings \cite{townes:75} between the ground state and the two tunnel states of the first vibrationally excited level are accurately reproduced. To treat the Schr\"{o}dinger equation in body-fixed coordinates, we expand the scattering wavefunction in the following coupled-channel basis \begin{equation} \Psi^{\rm bf} (R)= \frac{1}{R} \sum_{\mathbf{n}} | \mathbf{n}\rangle \chi_{\mathbf{n}}(R), \end{equation} where the radial dependence of the wavefunction is given by $\chi_{\mathbf{n}}(R)$, while the body-fixed angular basis set \begin{eqnarray}\label{eq:bfbas} | \mathbf{n} \rangle &\equiv&| j, k, K, J, M_J, v^{\pm} \rangle = \left[\frac{(2 j +1) (2 J + 1)}{32 \pi^3}\right]^{1/2} \nonumber \\ &&\times \phi^{\pm}_v(\rho)D^{(j)*}_{K k}(\zeta^{\rm bf})D^{(J)*}_{M_J K}(\phi^{\rm sf},\theta^{\rm sf},0) \end{eqnarray} is used to treat the angular part of the Hamiltonian. Here, $D^{(j)*}_{m m'}(\zeta)=e^{i m \alpha}d^{(j)}_{mm'}(\beta)e^{i m' \gamma}$ with $d^{(j)}_{mm'}(\beta)$ the well-known Wigner $d$-functions, $k$ is the projection quantum number of the monomer angular momentum with eigenvalue $j$ on the monomer $z$ axis, $K$ is the projection quantum number of both the monomer angular momentum and the total angular momentum with eigenvalue $J$ on the body-fixed dimer $z$ axis, $M_J$ is the projection of the total angular momentum on the space-fixed $z$ axis, $v$ is the umbrella vibration quantum number, and the superscript $+/-$ refers to the even/odd umbrella tunneling function. As a result, our task is to solve the following second-order matrix differential equation \begin{eqnarray}\label{eq:schr} - \frac{\partial^2 \chi_{\mathbf{n}'}(R)}{\partial R^2} = \sum_{\mathbf{n}} \langle \mathbf{n}' | \hat{W} | \mathbf{n} \rangle \chi_{\mathbf{n}}(R), \end{eqnarray} where we introduced the operator $\hat{W}=2\mu(E -\hat{H} + \hat{K})$ with the kinetic energy operator $\hat{K}$ given by the second term on the right-hand side of Eq.~(\ref{eq:ham}). We note that $J$ and $M_J$ are good quantum numbers, and that the operator $\hat{W}$ is diagonal in $J$ and independent of $M_J$. Furthermore, the monomer part of the Hamiltonian, $\hat{H}_{\rm mol}$, is also diagonal in the angular basis set. The complexity of the matrix $ \langle \boldsymbol{\mathbf{n}'} | \hat{W} | \mathbf{n}\rangle$ can be further reduced by considering the symmetry properties of the NH$_{3}$-He complex. Because the Hamiltonian commutes with permutations of the three hydrogen atoms in NH${}_3$ and the operator for inversion in space $\hat{E}^*$, it is useful to adapt the basis states such that they transform as the irreducible representations of the corresponding molecular symmetry group $D_{3h}({\rm M})$ in the notation of Bunker and Jensen \cite{bunker:98}. The adapted basis states of different symmetry cannot be mixed by the Hamiltonian. The precise procedure for this adaptation is described in the Appendix. Moreover, in Ref.~\cite{avoird:94} several useful relations can be found for determining the matrix elements of the $\hat{W}$ operator in the angular basis. In order to fully solve Eq.~(\ref{eq:schr}), the wavefunctions must satisfy the appropriate scattering boundary conditions \cite{johnson:73}. These boundary conditions are directly formulated in a space-fixed frame. The exact solution of the space-fixed Schr\"{o}dinger equation at larger separations $R$, i.e., when the interaction energy has approached zero, is a linear combination of the proper spherical Bessel functions. These Bessel functions are labelled by the space-fixed end-over-end rotational quantum number $L$, which has become a good quantum number at such large separations. Therefore, the matching of the propagated wavefunction from Eq.~(\ref{eq:schr}) to spherical Bessel functions can be performed at distances where the centrifugal energy, set by $L$ and decaying as $R^{-2}$, is still large, as long as the interaction energy, decaying in our case as $R^{-6}$, has become negligibly small. To perform the matching, it is necessary to transform between the body-fixed and the space-fixed basis sets. The latter basis set is for the present case given by \begin{eqnarray}\label{eq:sfbas} &&| j, k, L, J, M_J, v^{\pm} \rangle = \left[\frac{(2 j +1) (2 L + 1)}{32 \pi^3}\right]^{1/2} \phi^{\pm}_v(\rho) \nonumber \\ &&\times \sum _{m_j,M_L} D^{(j)*}_{m_j k}(\zeta^{\rm sf})C^{L}_{M_L}(\theta^{\rm sf},\phi^{\rm sf}) \langle j m_j; L M_L | J M_J \rangle,\quad ~ \end{eqnarray} with $C^{L}_{M_L}(\theta^{\rm sf},\phi^{\rm sf})$ the Racah-normalized spherical harmonics and $M_L$ the projection of the end-over-end angular momentum on the space-fixed $z$ axis, $m_j$ the projection of the monomer angular momentum on the space-fixed $z$ axis, and $\langle j_1 m_1; j_2 m_2 | j_3 m_3 \rangle$ a Clebsch-Gordan coefficient. The transformation between the body-fixed and the space-fixed basis then becomes \cite{avoird:94} \begin{eqnarray} &&| j, k, L, J, M_J, v^{\pm} \rangle \\ &&= \sum_{K} | j, k, K, J, M_J, v^{\pm} \rangle \left(\frac{2 L + 1}{2J+1}\right)^{1/2} \langle j K; L 0 | J K \rangle.\nonumber \end{eqnarray} To end our discussion of the matching procedure, we mention the various possible open and closed channels following from the Hamiltonian of Eq.~(\ref{eq:hammon}). The vibration-tunneling states of the umbrella motion are determined by calculating the eigenstates of the Hamiltonian formed by the last two terms of Eq.~(\ref{eq:hammon}). Only the lowest four eigenstates, labeled $\phi_{v}^{\pm}(\rho)$ with vibrational quantum numbers $v=0$ and $v=1$, are kept. With these four states as a basis for the umbrella motion, we turn to the first term of the Hamiltonian of Eq. (\ref{eq:hammon}). As a result, the rotational constants $1/2I_{\lambda \lambda}(\rho)$ become $4\times4$ matrices, but the $+$ and $-$ states are not mixed. The eigenstates that result from diagonalization of the full monomer Hamiltonian in a basis containing all rotational states with $j \le 6$ and the four umbrella states, determine the open and closed channels. We label the open channels by $|j k \pm \rangle$; the vibrational quantum number $v$ is omitted from this notation because all vibrational states with $v>0$ are closed for the energy range in which we are interested. We consider the ammonia molecules to be prepared in the $| 1 1 - \rangle$ state, so that the lower lying $| 1 1 + \rangle$ state is open for all collision energies. Increasing the collision energy beyond the energy of excited monomer states opens up the corresponding channels, and inelastic scattering into these states occurs if it is allowed by symmetry. The matching procedure of the wavefunctions to the boundary conditions for scattering at large distance $R$ ultimately leads to an expression for the scattering matrix. This $S$ matrix is subsequently directly related to the differential and integral cross sections for the elastic and inelastic channels, which can be compared with the outcome of collision experiments \cite{child:74}. \section{The NH${}_3$-He potential} Before we can apply the above described formalism to solve the scattering problem, we need to determine the NH${}_3$-He interaction potential. To this end {\it ab initio} calculations were performed with {\sc molpro} \cite{molpro:09}, using the supermolecule approach with the counterpoise procedure of Boys and Bernardi \cite{boys:70}. We considered the interaction energy to be dependent on four coordinates, namely $R$, $\beta^{\rm bf}$, $\gamma^{\rm bf}$ and $\rho$. The grid for the {\it ab initio} calculations consisted of 4180 points. For $R$, in total 19 points were used. In the short and intermediate range, i.e., for $4 a_0 \le R \le 10 a_0$, we used an equidistant grid of 13 points with a separation of $0.5 a_0$, while in the long range, that is for $R > 10 a_0$, we used an approximately logarithmic grid consisting of the points $12 a_0$, $14.4 a_0$, $17.3 a_0$, $20.8 a_0$, $25 a_0$, and $30 a_0$. For $\beta^{\rm bf}$, we used a Gauss-Legendre grid consisting of 11 points for $0 \le \beta^{\rm bf} \le \pi$, while for $\gamma^{\rm bf}$, we used an equidistant Gauss-Chebyshev grid consisting of the points $\pi/24$, $3\pi/24$, $5\pi/24$, and $7\pi/24$. Finally, for the grid in $\rho$ we used an equidistant grid of 5 points, where the middle point was given by the value $\rho_{3}=0.6226 \pi$, while the distance between the points was given by $\Delta\rho=( 2 \rho_{3}-\pi)/5$. The calculations in the long range were performed with the coupled-cluster method taking into account single and double excitations and a perturbative treatment of triple excitations [CCSD(T)], using the augmented correlation-consistent polarized valence quadruple-zeta (AVQZ) basis set. For the short range we used the explicitly correlated CCSD(T)-F12 method \cite{adler:07} to account more efficiently for the strong effect of electron correlations in this regime. The CCSD(T)-F12 method was found to yield accurate results with the smaller triple-zeta basis set (AVTZ), as illustrated by Table \ref{tabf12}. In this table the interaction energies are shown for different NH${}_3$-He geometries in the short, intermediate and long range with both the CCSD(T) and the CCSD(T)-F12 method using different basis sets. Also the effect of using midbond functions \cite{koch:98} is included in this table, where the midbond orbitals were located along the vector connecting the center of mass of the ammonia molecule with the helium atom at a distance of $(r_0+R)/2$ from the ammonia center of mass. Particularly in the short and intermediate range these midbond functions improve the interaction energies, so that they were used in the calculations with F12. From Table \ref{tabf12}, we see that in the short range the CCSD(T)-F12 method with an AVTZ basis set including midbond functions performs better than the CCSD(T) method with an AV5Z basis set, although the latter calculation is much more expensive due to the large basis set. The main reason why we did not use the F12 method in the long range is that we found that the implementation of this method in {\sc molpro} \cite{molpro:09} gives rise to an incorrect $1/R$ behavior in the very long range, rather than the correct $1/R^6$ behavior for the system under consideration. This behavior is caused by the fitting of the electron density distributions, which unfortunately does not result in exactly charge neutral monomers. Although the artifical residual charges can be reduced by introducing a larger electron density fitting basis set, the $1/R$ behavior will eventually always dominate the correct $1/R^6$ behavior. Hence, we decided to use the CCSD(T) method without F12 for the long range. We used the AVQZ basis set, and we may conclude from Table \ref{tabf12} that this basis set indeed gives rise to accurate interaction energies in the long range. In order to switch smoothly between the results of the two methods, we used the switching function $s(R)$ \begin{eqnarray} s(R)=\left\{ \begin{array}{l l} 0 & {\rm if } \quad R\le a \\ 1 & {\rm if } \quad R\ge b \\ \frac{1}{2}+\frac{1}{4}\sin\frac{\pi x}{2}(3-\sin^2\frac{\pi x}{2}) & {\rm otherwise} \end{array} \right. \end{eqnarray} with $x=(2R -b-a)/(b-a)$, $a =10 a_0$ and $b=13 a_0$. The function is chosen such that the first three derivatives of $s$ at $R=a$ and $R=b$ are zero. We thus calculated the interaction energies for the angular geometries at distance $R=12 a_0$ with both methods, where the calculated value with F12 was given a weight of $1-s(12a_0)$, while the value without F12 was given a weight $s(12a_0)$. \begin{table}[t] \caption{\label{tabf12} Comparison of the interaction energy between the CCSD(T) method (abbreviated as CC) and the CCSD(T)-F12 method (abbreviated as F12) for different basis sets and different geometries as calculated with {\sc molpro} \cite{molpro:09}. We used the augmented triple zeta (AVTZ), quadruple zeta (AVQZ) and quintuple zeta (AV5Z) basis sets. We also studied the effect of midbond functions \cite{koch:98}, which are indicated in the Table by the $+$ sign, when they are added to the basis set. For the short-range geometry, indicated by ${\bf x}_{\rm s}$ in the Table, we used $R = 4.5 a_0$, $\beta^{\rm sf}= 0$, $\gamma^{\rm sf}= 0$ and $\rho =14 \pi/24$. For the mid-range geometry, indicated by ${\bf x}_{\rm m}$, we used $R = 7 a_0$, $\beta^{\rm sf}= \pi/2$, $\gamma^{\rm sf}= \pi/6$ and $\rho =15 \pi/24$. For the long-range geometry, indicated by ${\bf x}_{\rm l}$, we used $R = 15 a_0$, $\beta^{\rm sf}= \pi$, $\gamma^{\rm sf}= \pi/3$ and $\rho = 16 \pi/24$. The interaction energies are given in cm$^{-1}$. } \begin{tabular}{ l c c c c c } & AVTZ & AVTZ$+$ & AVQZ & AVQZ$+$ & AV5Z \\ \hline CC (${\bf x}_{\rm s}$) & 1446.2& 1414.2 & 1407.4 & 1397.3 & 1394.7 \\ F12 (${\bf x}_{\rm s}$)& 1393.5& 1386.0 & 1389.8 & 1386.8 & 1386.1 \\ CC (${\bf x}_{\rm m}$) &$-$21.716&$-$23.529&$-$22.733&$-$23.521&$-$23.179\\ F12 (${\bf x}_{\rm m}$)&$-$23.329&$-$23.679&$-$23.237&$-$23.510&$-$23.381\\ CC (${\bf x}_{\rm l}$) &$-$0.2506&$-$0.2583&$-$0.2521&$-$0.2538&$-$0.2526\\ F12 (${\bf x}_{\rm l}$)&$-$0.2622&$-$0.2637&$-$0.2548&$-$0.2554&$-$0.2524\\ \end{tabular} \end{table} \begin{figure} \begin{center} \includegraphics[width=0.65\columnwidth]{figure2} \caption{\label{figpot} Contour plot of the NH${}_3$-He interaction potential as a function of $R$ and $\beta^{\rm bf}$ for $\gamma^{\rm bf} = 0$ and the equilibrium umbrella angle $\rho_{\rm e} = 112.15 {}^{\circ}$. The energies of the contours are given in cm$^{-1}$. } \end{center} \end{figure} To obtain an analytic representation of the interaction potential between the NH${}_3$ molecules and the He atoms we first perform an expansion in tesseral spherical harmonics, namely \begin{equation} \label{vaniso} V_{\rm int}(R, \beta^{\rm bf}, \gamma^{\rm bf}, \rho) =\sum_{l,m} (-1)^m v_{lm}(R,\rho) S_{lm}( \beta^{\rm bf}, \gamma^{\rm bf}), \end{equation} where due to the symmetry of the dimer only terms with $m = 0,3,6,...$ are present. Because we have 11 grid points in $\beta^{\rm bf}$, the summation over $l$ is from $0 \le l \le 10$. The summation over $m$ is from 0 to the largest multiple of 3 that is smaller than or equal to the corresponding $l$ value. On all grid points $R_i$ and $\rho_j$ we determine the coefficients of the angular expansion $v_{lm}(R_i,\rho_j)$ by means of a quadrature on the {\it ab initio} grid with the appropriate Gauss-Legendre and Gauss-Chebyshev weights \cite{avoird:94}. For the resulting expansion coefficients $v_{lm}$, we distinguish between the short-range and the long-range behavior, so that $v_{lm}(R,\rho)= v^{\rm sr}_{lm}(R,\rho)+v^{\rm lr}_{lm}(R,\rho)$. Both in the short range and the long range the dependence of the coefficients $v_{lm}$ on $\rho$ is represented by a polynomial expansion in $(\rho-\pi/2)^p$, where $p$ ranges from $0$ to $9$. If $l+m$ is even, then the polynomial expansion only contains even powers in $p$, while if $l+m$ is odd only odd powers are present. In the long range, we expanded the potential in powers of $R^{-n}$, resulting in \begin{equation}\label{eqcoeflr} v^{\rm lr}_{lm}(R, \rho) =\sum_{n,p} c_{lmpn} f_{n}(a R) \left(\rho-\frac{\pi}{2}\right)^p R^{-n}. \end{equation} where the inverse powers of $R$ that are involved depend on $l$. It can be shown \cite{avoird:80} that for $l = 0,2$ the expansion starts with $n_i =6$, while for $l=1,3$ it starts with $n_i=7$. For $l \ge 4$, it starts with $n_i = l + 4$. We used the analytic long range expansion of Eq. (\ref{eqcoeflr}) only for $l \le 5$, and for each $l$ we took the leading term $R^{-n_i}$ and the next-to-leading term $R^{-n_i-2}$ into account. The Tang-Toennies damping function \begin{equation} f_n(x)=1-\left(\sum_{i=0}^n \frac{x^i}{i!} \right)e^{-x}. \end{equation} was included to avoid the singular behavior of the long-range terms in the short range \cite{tang:84}. For the value of $a$ in Eq.~(\ref{eqcoeflr}) we used the isotropic exponent in the short-range, or, to be more precise, $a= \ln[v_{00}(R_1,\rho_3)/v_{00}(R_2,\rho_3)]/\Delta R =2.088 a_0^{-1}$ with $R_1$ and $R_2$ the first two points of the $R$ grid and $\Delta R=R_2-R_1$. The expansion coefficients $c_{lmpn}$ were obtained from $v_{lm}(R_i,\rho_j)$ by performing a weighted least-squares fit using the last three points of the $R$ grid and all points of the $\rho$ grid. The three $R$ points were weighted for each $l$ by $R_i^{n_i}$, with $n_i$ the leading power of the long range decay for the considered $l$. To describe the short and intermediate range of the potential the same expansion was employed in $\beta^{\rm bf}, \gamma^{\rm bf}$ and $\rho$ as for the long range, but the behavior in $R$ was treated differently. The following procedure was used. First, for all the grid geometries, the corresponding value of the analytic long range potential was subtracted from the {\it ab initio} values. Next, after performing the expansion in tesseral harmonics and powers of $(\rho-\pi/2)$, the behavior of the resulting coefficients $v_{lmp}(R)$ was interpolated with a reproducing kernel Hilbert space (RKHS) method \cite{ho:96}. The smoothness parameter of the RKHS interpolation was set to 2, while the RKHS parameter $m$, which determines the power with which the interpolation function decays, was chosen to depend on $l$. For $l \le 5$, the parameter was set to $m = n_i+1$. Then the RKHS function decays as $R^{-n_i-2}$, which is faster for each $l$ than the leading term in the analytic fit of Eq. (\ref{eqcoeflr}). However, for $l>5$, no analytic long range fit was done, and we used $m=n_i -1$. As a result, the corresponding RKHS functions decayed for each $l$ as $R^{-n_i}$ with $n_i = l+4$, which is the correct leading long range behavior for $l > 5$ \cite{avoird:80}. We have compared the fitted potential with the {\it ab initio} values on the full grid to test the accuracy of the fit in the angles $\beta^{\rm bf}, \gamma^{\rm bf}$ and $\rho$. The quality of the fit in $R$ cannot be tested in this way, because the RKHS procedure goes by construction precisely through the points to be fitted. We calculated the RMS (root mean square) error for each grid distance $R_i$ and divided by the mean {\it ab initio} interaction energy at that distance, giving for the relative RMS error $\xi(R_i)$ that \begin{eqnarray} \xi(R_i) &=& \frac{\sqrt{\frac{1}{n}\sum_{j,k,l}[\Delta V(R_i,\beta_j^{\rm bf}, \gamma_k^{\rm bf},\rho_l)]^2}}{ \left|\frac{1}{n}\sum_{j,k,l} V_{\rm int}^{\rm abi}(R_i,\beta_j^{\rm bf}, \gamma_k^{\rm bf},\rho_l)\right| } \,100 \%,\quad \end{eqnarray} where $\Delta V(R_i,\beta_j^{\rm bf}, \gamma_k^{\rm bf},\rho_l)= V^{\rm fit}_{\rm int}(R_i,\beta_j^{\rm bf}, \gamma_k^{\rm bf},\rho_l)-V^{\rm abi}_{\rm int}(R_i,\beta_j^{\rm bf}, \gamma_k^{\rm bf},\rho_l)$, and the summations are over all $n=220$ angular grid points. For our potential fit we found that the relative error $\xi(R_i)$ is less than 0.05 \% for all $R_i$, so the fits in the angular coordinates are excellent. To also test the fit in $R$, we calculated {\it ab initio} interaction energies for an additional 495 points, that were chosen to lie about halfway between the grid points used for the fit. The relative RMS error of the values calculated from the fit compared to the new {\it ab initio} values depended quite strongly on $R$, where the largest error was found to occur in the short range. Namely, for the test points at $R=4.3 a_0$ we found with the use of 45 different angular points a relative RMS error of $3.5\%$, while for all other $R$ values we obtained a relative error of about $0.5\%$ or less. An important reason for this behavior is that we use a RKHS fit for the short range, which behaves as a power law, while the true behavior of the potential is exponential. The fitting procedure could thus have been further improved using an exponential form. However, we note that already the present fitting error is rather small. Moreover, in the present paper we use the potential to describe cold scattering with collision energies of maximally $130$ cm$^{-1}$, so that the extreme short-range behavior of the potential is not being probed. \begin{figure} \begin{center} \includegraphics[width=0.9\columnwidth]{figure3.eps} \caption{\label{figvlm} Coefficients $v_{lm}(R,\rho_e)$ of the NH${}_3$-He interaction energy as a function of center-of-mass distance $R$, evaluated at the equilibrium umbrella angle $\rho_e = 112.15 {}^{\circ}$. The isotropic $v_{00}(R,\rho_e)$ coefficient is largest. The $v_{10}$, $v_{20}$, $v_{30}$, $v_{33}$, $v_{40}$ and $v_{43}$ coeffients are shown as well.} \end{center} \end{figure} In Fig.~\ref{figpot}, we show a contour plot of the NH${}_3$-He interaction potential for $\gamma^{\rm bf} = 0$ and the equilibrium umbrella angle $\rho_e = 112.15 {}^{\circ}$. For this value of $\gamma^{\rm bf}$ and $\rho$, the minimum of the potential is given by $V_{\rm min }=-35.08$ cm$^{-1}$ for $R=6.095$ $a_0$ and $\beta^{\rm bf} = 89.0{}^{\circ}$. This may be compared to the potential of Hodges {\it et al.} \cite{hodges:01}, where the minimum of the potential for $\gamma^{\rm bf} = 0$ and $\rho_e = 112.15 {}^{\circ}$ is given by $V_{\rm min }=-33.46$ cm$^{-1}$ for $R=6.133$ $a_0$ and $\beta^{\rm bf} = 88.75{}^{\circ}$. Although this difference in the well depth is not very large, we have found that the consequences for low-energy scattering can still be quite substantial, as we will discuss in Section \ref{secresults}. Finally, we have for the leading isotropic coefficient, defined as $C_6 = -v_{00}(R,\rho_3) R^6 $ for large $R$, that in atomic units $C_6 = 39.6$ $E_h \cdot a_0^6$. The relative importance of the various $v_{l m}(R,\rho_e)$ expansion coefficients is shown in Fig.~\ref{figvlm}. The potential is available in {\sc fortran} 77 on {\sc epaps} \cite{epaps:pot}. \section{Computational aspects} Having discussed the formalism and the potential, we now turn to the numerical procedures that we used in order to obtain converged cross sections that can be compared with future cold-collision experiments. To numerically determine the four lowest lying vibration-inversion levels $\phi^{\pm}_{v}(\rho)$ of the Hamiltonian of Eq.~(\ref{eq:hammon}) for $j = 0$, we used the discrete variable representation based on sinc-functions (sinc-DVR) \cite{groenenboom:93}. The resulting eigenfunctions were used to determine the matrix elements $\langle \phi '|1/2I_{\lambda\lambda}| \phi \rangle$ with $\phi =\phi_{v}^{\pm}(\rho) $ by numerical integration. For the propagation in solving Eq.~(\ref{eq:schr}), the renormalized Numerov algorithm was used, starting at $4 a_0$ and ending at $50 a_0$, using an equidistant grid with 273 points. The renormalized Numerov method also allows for a complete reconstruction of the scattering wavefunctions. The angular basis set contained all monomer states with $j \le 6$, where we checked that the inclusion of more monomer levels resulted only in deviations of maximally 1 \% for the calculated cross sections. The maximal value for the total angular momentum $J$ that we used depended on the collision energy. For collision energies $E \le 10$ cm${}^{-1}$, we included all angular basis states with $J \le 10$, while for $10< E \le 50$ cm${}^{-1}$, we included all basis states with $J \le 20$, and for $50< E \le 130$ cm${}^{-1}$, we included all states with $J \le 30$. The convergence of the cross sections with respect to the total angular momentum $J$ is slowest for the elastic cross section. The inelastic cross sections are converged at considerably lower values of $J$ than reported here. In order to check our results and gain additional insight, we also implemented a commonly applied model to treat the ammonia umbrella motion in scattering calculations \cite{davis:78,green:80}. In this model, no vibrationally excited umbrella states are taken into account and the ground-state umbrella tunneling states are approximated as an even and odd combination of the two rigid equilibrium structures. These two states are thus written as $| \pm \rangle = [f(\rho-\rho_e)\pm f(\pi-\rho+\rho_e)]/2^{1/2}$, where $f(x)$ is a function localized around $x=0$. More precisely, the two-state model amounts to approximating the matrix elements of the potential by $\langle \pm | v_{lm}(R,\rho) | \pm\rangle = v_{lm}(R,\rho_e)$ for $l+m$ even, and $\langle \pm | v_{lm}(R,\rho) | \mp\rangle = v_{lm}(R,\rho_e)$ for $l+m$ odd. For the rotational constants we use the experimentally determined values $A_{xx} = A_{yy} = 9.9402$ cm${}^{-1}$, and $A_{zz}=6.3044$ cm${}^{-1}$ in the model. Furthermore, we include the experimental ground state splitting of $0.79$ cm$^{-1}$ \cite{townes:75} between the two tunneling states in the scattering calculations. This simple model has been implemented in the scattering program {\sc molscat} \cite{molscat:94}. We have used {\sc molscat} to double-check the results that we obtained from our own scattering program. The model was previously found to result in good agreement with more elaborate treatments of the umbrella motion for scattering at higher collision energies \cite{sanden:93}. In this article we also want to test the accuracy of the model for cold collisions, and in particular for the calculation of scattering resonances. \section{Results} \label{secresults} \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure4} \caption{\label{figcross_log} Integral cross sections for NH${}_3$-He scattering as a function of collision energy. The initial state of the NH${}_{3}$ molecule is the $|11-\rangle$ state. At low collision energies only elastic scattering (upper red curve) and inelastic scattering into the lower lying $|11+\rangle$ state (lower blue curve) can occur.} \end{center} \end{figure} In Fig.~\ref{figcross_log}, we show the integral cross sections for the scattering of NH${}_3$ molecules with He atoms for collision energies ranging from 10${}^{-4}$ cm${}^{-1}$ to 20 cm${}^{-1}$. Initially, we only have elastic scattering and inelastic scattering into the $| 1 1 + \rangle$ state, which lies $0.79$ cm${}^{-1}$ lower in energy. Figure~\ref{figcross_log} was made using the previously described elaborate treatment of the umbrella motion, however, with the use of the model treatment almost exactly the same results were obtained. We observe in the first place that, in agreement with the Wigner threshold laws \cite{wigner:48}, the elastic cross section becomes constant for very small collision energies, while the inelastic cross section decreases with $E$ as $1/\sqrt{E}$. Going more into the details of the figure, we observe two shape resonances closely together in the elastic channel at collision energies of 1.86 and 2.22 cm${}^{-1}$. In bound state calculations with the NH$_3$-He complex enclosed in a box of variable size we found continuum levels with nearly the same energies that are practically independent of the box size, so we may conclude that these peaks in the scattering cross section indeed correspond to shape or orbiting resonances caused by quasi-bound states. Such quasi-bound states may occur either in the incoming or in the outgoing scattering channel; for the specific case of elastic scattering these are the same. Looking at the dominant contributions to the cross section, the first peak was found to be mainly caused by quasi-bound states with total angular momenta $J=4$ and $J=6$, while the second peak was mainly caused by a quasi-bound state with total angular momentum $J=5$. In both cases they corresponded to an end-over-end angular momentum of $L=5$. Looking in the same energy range at the inelastic scattering into the $| 1 1 + \rangle$ state, we observe not only two similar peaks at the same collision energies, but also two additional peaks at 1.08 and 1.44 cm${}^{-1}$. These two additional shape resonances can be readily understood by noting that for inelastic scattering the resonant quasi-bound state can occur either in the incoming channel or in the outgoing channel, where the latter channel is about $0.8$ cm$^{-1}$ lower in energy. This is indeed precisely the energy with which the two additional peaks in the inelastic channel are shifted to the left in Fig.~\ref{figcross_log}. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure5} \caption{\label{figcross1} Elastic (upper red curve) and inversion inelastic (lower blue curve) integral cross sections as a function of collision energy for NH${}_3$-He scattering. The inelastic cross section is scaled with a factor of 150, so that the actual inelastic cross section is 150 times smaller than shown in the figure. At higher collision energies the $|2 2 \pm\rangle$ and the $|2 1 \pm\rangle$ channels open. As a result, Feshbach resonances are observed, which are most pronounced for the inelastic scattering into the $|11+\rangle$ state.} \end{center} \end{figure} For completeness, we note that we studied this collision energy range also with the potential of Hodges and Wheatley \cite{hodges:01}. Although the difference in the well depth between the two potentials at the equilibrium umbrella angle was only about 5\%, we still found large differences in the resonant structures at very low energies. For example, using the Hodges and Wheatley potential \cite{hodges:01}, we observed two very strong shape resonances at collision energies of 0.03 and 0.45 cm${}^{-1}$ induced by quasi-bound states with total angular momenta $J=3$ and $J=4$ and end-over-end angular momentum $L=4$. However, because our own potential is deeper, we find that these quasi-bound states have become true bound states with energies below the scattering continuum, so that they cannot cause shape resonances anymore. As a result, the first shape resonances we find with our potential are induced by quasi-bound states with total angular momenta $J=4,5$ and $6$ and $L=5$, as shown in Fig.~\ref{figcross_log}. This point also clearly shows that scattering resonances at low energy can be very sensitive to the precise shape of the potential energy surface, which means that accurate scattering experiments can be used to probe very precisely our knowledge of intermolecular interactions. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure6} \caption{\label{figwf1} (a) Contributions of various channels and partial waves to the scattering wavefunction at a collision energy of 37.28 cm${}^{-1}$, where a Feshbach resonance for inelastic scattering into the $| 1 1 + \rangle$ state occurs. We used a total angular momentum of $J=3$ and considered $E''$ symmetry. As a result, the initial state $|11-\rangle$ asymptotically corresponds to $L=3$, while there are eight open partial waves in outgoing channels. (b) Contributions of various channels to the square of the wavefunction. In the inner region, where the collision takes place, a large amplitude in the asymptotically closed $|21-\rangle$ channel is observed. } \end{center} \end{figure} In Fig.~\ref{figcross1}, we again show the integral cross sections for elastic scattering and inelastic scattering into the $| 1 1 + \rangle$ state, but now considering collision energies from 10 to 50 cm${}^{-1}$. Note that the inelastic cross is actually 150 times smaller than shown in the figure. As can be seen from Fig.~\ref{figlevels}, at a collision energy of $28.33$ cm$^{-1}$ it becomes energetically possible to excite the ammonia molecule from its $| 1 1 - \rangle$ state to its $| 2 2 + \rangle$ state, and a new scattering channel opens. At 29.12, 39.33 and 40.12 cm$^{-1}$, the $| 2 2 - \rangle$, $| 2 1 + \rangle$ and $| 2 1 - \rangle$ channels open, respectively, as also indicated in Fig.~\ref{figcross1}. The opening of the new channels is seen to have a profound effect on the inelastic cross sections to the $| 1 1 + \rangle$ state. Namely, before these new channels open a bunch of Feshbach resonances is observed. These resonant structures are called Feshbach resonances because they are caused by a molecular level that is different from the incoming and the outgoing channel. In Fig.~\ref{figcross1} we see that especially the Feshbach resonances induced by the $| 2 1 \pm \rangle$ levels at collision energies around 40 cm$^{-1}$ are strong, giving rise to almost a factor of 3 increase compared to the background inelastic cross section. These resonances seem to be particularly suited to observe in a collision experiment. We come back to this point more elaborately in Section \ref{sec:disc}. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure7} \caption{\label{figphases} Top panel: Phase shift sum as a function of collision energy for various total angular momenta $J$. Both $E'$ and $E''$ symmetries are considered, so that we have two curves for each $J$. Bottom panel: the corresponding lifetimes as a function of the collision energy. The lifetimes are obtained from the derivative of the phase shifts with respect to the collision energy (in cm$^{-1}$); they are given in units of 5.3088 ps. } \end{center} \end{figure} To understand the Feshbach resonances in more detail, we have studied the scattering wavefunctions. In Fig.~\ref{figwf1} contributions to the scattering wavefunction are shown at a collision energy of 37.28 cm${}^{-1}$. At this collision energy, there is a Feshbach resonance for inelastic scattering into the $| 1 1 + \rangle$ state, somewhat below the energy at which the $| 2 1 - \rangle$ channel opens. In Fig.~\ref{figwf1}(a), we show contributions of different open and closed channels to the scattering wavefunction. For this particular figure, we considered a total angular momentum of $J=3$ and symmetry $E''$ (see the Appendix). This means that for the incoming channel, i.e., the $|11-\rangle$ state, asymptotically only the partial wave with $L = 3$ contributed. For the four open outgoing channels, namely $|11\pm\rangle$ and $|22\pm \rangle$, in total eight open partial waves are possible for the considered $J$ and $E''$ symmetry. In the inner region also contributions corresponding to asymptotically closed channels can gain amplitude, when they are coupled to the considered incoming state and outgoing state by the interaction potential. In Fig.~\ref{figwf1}(b), we show for each channel the resulting contributions to the square of the wavefunction. From Figs.~\ref{figwf1}(a) and (b) we clearly see that in particular the closed $|21-\rangle$ channel has a very strong amplitude in the collision region, which shows that this state is responsible for the strong Feshbach resonance observed at this collision energy. A different way to study the Feshbach resonances is by looking at phase shifts in the scattering wavefunction. These phase shifts can be obtained from the eigenvalues of the scattering matrix \cite{child:74,ashton:83}. In Fig.~\ref{figphases}, we show in the top panel the sums of the phase shifts in all open channels for various total angular momenta $J$. Since we consider both symmetries $E'$ and $E''$, we have two curves for each $J$. From scattering theory it follows that when a resonance occurs, the phase shift sum rapidly increases by $\pi$ \cite{ashton:83} as a function of energy. In the top panel of Fig.~\ref{figphases}, we indeed see this happening many times at the collision energies where resonances are found in the elastic and inelastic cross sections. The derivatives of the phase shifts with respect to the energy give the lifetime of the collision complex \cite{child:74}. These lifetimes are shown in the lower panel of Fig.~\ref{figphases}. This figure shows that at the collision energies where resonances occur, we indeed have long-lived intermediate quasi-bound states. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure8} \caption{\label{figcross2} Inelastic integral cross sections for scattering into the $|2 2 \pm\rangle$ channels and the $|2 1 \pm\rangle$ channels as a function of collision energy. The inelastic cross sections for the $| 2 2 + \rangle$ and the $| 2 1 + \rangle$ channels are scaled with a factor of 2. After the various channels open, shape resonances are observed.} \end{center} \end{figure} In Fig.~\ref{figcross2}, we show the integral cross sections for inelastic scattering into the $| 2 1 \pm \rangle$ and the $| 2 2 \pm \rangle$ states, for a collision energy ranging from the energies at which these channels open, up to 50 cm${}^{-1}$. Note that the inelastic cross sections to the $| 2 2 + \rangle$ and $| 2 1 + \rangle$ states are scaled with a factor of 2. Immediately after each channel opens, we see strong resonant features, which are shape resonances, caused by quasi-bound states in the outgoing channel. In the $| 2 2 \pm \rangle$ channels we also find Feshbach resonances due to quasi-bound states of $| 2 1 \pm \rangle$ character. In the energy range from 10 to 50 cm${}^{-1}$, we also studied the scattering cross sections using the previously described model treatment of the NH${}_{3}$ umbrella motion. We found that the model calculations have the tendency to somewhat overestimate the strength of certain resonance peaks compared to the elaborate treatment of the umbrella motion. Studying this effect in more detail, we found that the differences are mainly due to the approximation of the nonzero potential matrix elements for the two tunneling states as $v_{lm}(R,\rho_e)$, rather than due to the neglect of the higher lying $\phi_1^{\pm}(\rho)$ states. Namely, by calculating the cross sections with the elaborate treatment and taking only the lowest two umbrella functions $\phi_0^{\pm}(\rho)$ into account we obtained cross sections that were nearly equal to the elaborate treatment with four umbrella functions, while they gave rise to the same differences with the model treatment. However, we note that in general the model treatment performed very satisfactory in describing the resonance structures. All resonant peaks found with the elaborate treatment were also found with the model treatment, and typically the strength of the scattering resonances differed by less than 10\%. Because the precise strength and location of the resonances are very sensitive to the the potential, we conclude that the use of the model treatment is useful in studying scattering resonances, especially in cases when the elaborate treatment is computationally too expensive. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure9} \caption{\label{figcross3} Elastic and inelastic integral cross sections for scattering into the $|1 1 +\rangle$, $|2 2 \pm\rangle$, $|2 1 \pm\rangle$ states as a function of collision energy. The inelastic cross sections for the $| 2 2 - \rangle$ and $| 2 1 - \rangle$ channels are scaled with a factor of 20, and for the $|1 1 +\rangle$, $|2 2 +\rangle$ and $|2 1 +\rangle$ channels with a factor of 40. At higher collision energies various $|3 k \pm\rangle$ and $|4 k \pm\rangle$ channels for the {\it para} ammonia molecules open. As a result, small Feshbach resonances are observed.} \end{center} \end{figure} Looking again at Fig.~\ref{figcross2}, we note that there are significant differences in the magnitudes of the inelastic cross sections for the various collision channels. For example, the transition to the $| 2 2 - \rangle$ state is seen to be much stronger than the transition to the $| 2 2 + \rangle$ state, and the same holds for the transition to the $| 2 1 - \rangle$ state compared to the $| 2 1 + \rangle$ state. The relative magnitude of the integral cross sections for the elastic and inelastic scattering channels can even be more clearly observed in Fig.~\ref{figcross3}. In this figure, we show the integral cross sections for scattering into the $| 1 1 \pm \rangle$, $| 2 1 \pm \rangle$ and $| 2 2 \pm \rangle$ states, for collision energies ranging from 80 cm${}^{-1}$ to 130 cm${}^{-1}$. Notice the scaling of the inelastic cross sections indicated in the figure. To explain the relative strengths of the transitions shown, we note that the scattering from the $| 1 1 - \rangle$ channel into different $| j k\pm \rangle$ channels is caused by different anisotropic terms in the interaction potential with coefficients $v_{lm}(R,\rho)$, cf. Eq.~(\ref{vaniso}). For example, in order to change the umbrella state of the ammonia molecule (i.e., going from the odd $-$ state to the even $+$ state) we need terms in the potential for which $l+m $ is odd, so that also the corresponding coefficient $v_{lm}(R,\rho)$ is odd in $\rho$. The various potential energy coefficients are plotted as a function of $R$ at the equilibrium umbrella angle $\rho_e$ in Fig.~\ref{figvlm}. \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure10} \caption{\label{figcross4} Inelastic integral cross sections for scattering into the $|3 2 \pm\rangle$ and the $|3 1 \pm\rangle$ states as a function of collision energy. At higher collision energies the $|4 4 \pm\rangle$ channels for the {\it para} ammonia molecules open. As a result, Feshbach resonances are observed.} \end{center} \end{figure} From this figure, we see that the isotropic coefficient $v_{00}(R,\rho_e)$ is by far the largest coefficient of all. This coefficient can only couple the initial $|1 1 - \rangle$ state to itself (see for example Refs.~\cite{avoird:94,millan:95}), causing a large elastic scattering cross section. For a transition to a different umbrella state, or to a state with different $j$, we need potential energy coefficients with $l \ge 1$. Since $v_{10}$ and $v_{30}$ are odd in $\rho$, they cause transitions from the $|1 1 - \rangle$ state to the $|2 1 + \rangle$ and the $|1 1 + \rangle$ state for example. From Fig.~\ref{figcross3}, we see that the inelastic cross sections to these two states are indeed approximately equally large. The $v_{20}$ term causes transitions to the $|2 1 - \rangle$ state, and because this expansion coefficient is relatively large, the corresponding cross section is large as well. Finally, in order to change $k$ in the collision, we need potential terms with $m \ne 0$, of which the first two are the $v_{33}$ and the $v_{43}$ coefficients. The $v_{33}$ coefficient causes $- \rightarrow -$ transitions and the $v_{43}$ coefficient causes $- \rightarrow +$ transitions. From Fig.~\ref{figvlm} we see that the $v_{33}$ coefficient is rather large, explaining the large cross sections to the $|2 2 - \rangle$ state, while the $v_{43}$ coefficient is small, explaining the small cross sections to the $|2 2 + \rangle$ state. \begin{figure*} \begin{center} \includegraphics[width=1.4\columnwidth]{figure11} \caption{\label{figdcs} Differential cross sections as a function of the zenith angle $\theta$ for collisions between NH${}_{3}$ and He at various collision energies. Upper two panels: differential cross sections for elastic scattering at a collision energy of 0.87 cm${}^{-1}$ (left) and 1.86 cm${}^{-1}$ (right). In the first case there is no resonance, the second case corresponds to a shape resonance. Middle two panels: differential cross sections for inelastic scattering into the $| 1 1 + \rangle$ state at a collision energy of 20 cm${}^{-1}$ (left panel, no resonance) and 24.36 cm${}^{-1}$ (right panel, Feshbach resonance). Lower two panels: the same but for a collision energy of 30 cm${}^{-1}$ (left panel, no resonance) and 37.28 cm${}^{-1}$ (right panel, Feshbach resonance).} \end{center} \end{figure*} In Fig.~\ref{figcross4}, we show the integral cross sections for inelastic scattering into the $| 3 2 \pm \rangle$ and the $| 3 1 \pm \rangle$ states, for a collision energy ranging from the energies at which these channels open, up to 130 cm${}^{-1}$. These small cross sections will be harder to observe experimentally. However, if these channels can be detected, they allow for the observation of pronounced shape resonances at higher collision energies. The cross sections for scattering into the $| 3 2 \pm \rangle$ states also give rise to Feshbach resonances with quasi-bound states of $| 3 1 \pm \rangle$ character. Even stronger Feshbach resonances at higher collision energies between about 120 and 125 cm${}^{-1}$ are found in the cross sections for scattering into the $| 3 1 - \rangle$ state. These Feshbach resonances are caused by the opening of the $| 4 4 \pm \rangle$ channels at a collision energy of about 125 cm${}^{-1}$. Finally, we also studied differential cross sections, where we looked in particular at the behavior of the differential cross sections as a function of energy close to resonance. In general, we found that the differential cross sections can change rapidly and dramatically close to resonance. This behavior is illustrated by Fig.~\ref{figdcs}. Here, we see in the upper two panels the differential cross sections for elastic scattering at collision energies of 0.87 and 1.86 cm${}^{-1}$. For the first of these energies there is no resonance, while for the second there is a shape resonance. For the off-resonance case we find that, apart from diffraction oscillations, there is only a forward scattering peak. On resonance there is also a strong backward peak. The lower four panels show the differential cross sections for inelastic scattering to the $| 1 1 + \rangle$ state at collision energies of 20, 24.36, 30 and 37.28 cm${}^{-1}$. At 20 and 30 cm${}^{-1}$, which are shown in the two lower plots on the left, there is no resonance and the differential cross sections look rather similar to the upper left one, giving predominantly rise to forward scattering. At 24.36 and 37.28 cm${}^{-1}$, which are shown in the two lower plots on the right, there is a Feshbach resonance present, and as a result the differential cross sections look very differently, giving again rise to significant backscattering. In general, the precise structure of the differential cross section depends on various aspects such as the lifetime and the rotational state of the intermediate collision complex. As a result, it is expected that the differential cross sections show clear changes near a resonance, but the precise way in which they change is hard to predict and can be very different for different resonances, as is also seen in Fig.~\ref{figdcs}. \section{Discussion and conclusion}\label{sec:disc} In this article, we have theoretically studied cold collisions of NH${}_3$ molecules with He atoms, where we looked in detail at shape and Feshbach scattering resonances. Prior to collision, we considered the ammonia molecules to be in their antisymmetric umbrella state with angular momentum $j=1$ and projection $k=1$, which is a suitable state for Stark deceleration. We calculated state-to-state integral and differential cross sections for collision energies ranging from 10${}^{-4}$ cm$^{-1}$ to 130 cm$^{-1}$, using fully converged quantum close-coupling calculations. We treated the umbrella motion of the ammonia molecule by solving the corresponding Hamiltonian in curvilinear coordinates and taking the resulting first four vibration-tunneling states exactly into account. We call this the elaborate treatment. We also used a common model for the umbrella motion which approximates the umbrella tunneling states as an even and odd combination of the two possible rigid equilibrium structures for ammonia. This we call the model treatment. To describe the interaction between the NH${}_3$ molecules and the He atoms accurately, we presented a new four-dimensional potential energy surface, based on a high-quality fit of 4180 {\it ab initio} points. In the short range we used the explicitly correlated CCSD(T)-F12 method with an AVTZ basis set including midbond functions, and we showed that this approach leads to excellent results in the short range. In the long range we used the CCSD(T) method with an AVQZ basis but without F12, since we found that the electron density fitting that accompanies the F12 treatment does not exactly preserve charge neutrality of the monomers and eventually leads to a dominant $1/R$ dependence of the potential at very large $R$ values. Our potential has a well depth $D_{e} = 35.08$ cm${}^{-1}$, which is to be compared with the well depth of 33.46 cm${}^{-1}$ for the potential of Hodges and Wheatley \cite{hodges:01}. Although this difference is not very large, we found that small differences in the potential can have profound consequences for the observed resonance structures at low scattering energies. We studied all open collision channels for {\it para} ammonia up to $j=3$ and in all these channels we found pronounced shape resonances right after the opening of these channels, caused by quasi-bound states in the incoming and outgoing channels. We also found Feshbach resonances that are particularly strong for the outgoing $| 1 1 + \rangle$ channel at collision energies of about 25 cm${}^{-1}$ caused by intermediate $| 2 2 \pm \rangle$ states, and at collision energies of about 35 cm${}^{-1}$ caused by intermediate $| 2 1 \pm \rangle$ states. Due to the large cross section of these inelastic resonances, namely more than 1 \AA${}^2$, they seem to be a good candidate for experimental observation. Also in the $| 3 1 - \rangle$ channel at collision energies of about 120 cm${}^{-1}$ relatively strong Feshbach resonances were seen that are due to intermediate $| 4 4 \pm \rangle$ states. We analyzed the observed resonant structures in detail by looking at the corresponding scattering wavefunctions, phase shifts and lifetimes. We also investigated the validity of using the model treatment for the ammonia umbrella motion in describing low-energy scattering resonances. We found that the model performs qualitatively very well, but on a quantitative level some resonance peaks are somewhat overestimated compared to the elaborate treatment. However, considering the sensitivity of these resonances to the interaction potential, for which even state-of-the-art {\it ab initio} methods still lead to uncertainties on the order of a percent, the model treatment seems adequate in treating low-energy resonant scattering, especially in cases when the elaborate treatment becomes computationally too expensive. The calculated integral cross sections at low collision energies can be measured using Stark-decelerated molecular beams. The NH$_3$ molecule, and its isotopologue ND$_3$, are amenable to the Stark deceleration technique, and have been employed frequently in deceleration experiments \cite{Bethlem:PRA65:053416}. The Stark decelerator provides a beam of ammonia molecules, state-selected in the upper inversion component of the $j=k=1$ level, with a velocity that is tunable between standstill and high velocities \cite{Heiner:PCCP8:2666}. In a crossed beam experiment, the Stark decelerated ammonia molecules can be collided with an atomic beam of helium. In an optimized geometry, the two beams collide at a small beam intersection angle. An intersection angle of less than $ 90^{\circ}$ reduces the attainable collision energy and improves the collision energy resolution of the experiment \cite{Scharfenberg:PCCP13:8448}. As shown in section \ref{secresults}, there are a number of scattering channels with pronounced shape and/or Feshbach resonances. The most promising prospects for the experimental observation of resonant features are found in the channels $|11+ \rangle \leftarrow |11- \rangle$ (see Fig.~\ref{figcross1}), $|22- \rangle \leftarrow |11- \rangle$, and $|21- \rangle \leftarrow |11- \rangle$ (see Fig.~\ref{figcross2}). \begin{figure} \begin{center} \includegraphics[width=1.0\columnwidth]{figure12} \caption{\label{figconv} Convoluted inversion inelastic integral cross sections as a function of the mean collision energy for NH${}_3$-He scattering. The initial state of the NH${}_{3}$ molecules is the $|11-\rangle$ state and the final state is the $|11+\rangle$ state. The figure is similar to Fig.~\ref{figcross1}, only now we have assumed a Gaussian collision energy distribution for the colliding particles to simulate more realistically what would be observed with present day experimental technology. The blue curve corresponds to a full width at half maximum (FWHM) of $1~$cm$^{-1}$, and the red curve of $3~$cm$^{-1}$. } \end{center} \end{figure} The resonant structures are found at collision energies in the 20 - 50 cm$^{-1}$ range. To simulate what would be observed in a molecular beam scattering experiment, we convoluted the integral scattering cross section for the $|11+\rangle \leftarrow |11-\rangle$ inversion inelastic channel with Gaussian collision energy distributions having both a 1~cm$^{-1}$ and a 3~cm$^{-1}$ full width at half maximum (FWHM). In the considered scattering channel bunches of Feshbach resonances are observed that are caused by the opening of the $|22\pm\rangle$ and $|21\pm\rangle$ channels, as seen in Fig.~\ref{figcross1}. The result of the convolutions are shown in Fig.~\ref{figconv}. From this figure we conclude that the details of the dense resonance structures in Fig.~\ref{figcross1} can only be resolved when an experimental collision energy spread that is much less than $1$ cm$^{-1}$ can be achieved. For an experimental resolution of 3~cm$^{-1}$, however, the bunch of scattering resonances can still be discerned from the background inelastic signal by measuring the inelastic cross section as a function of collision energy. Figure~\ref{figconv} shows that such a measurement would lead to a clear enhancement of the inelastic signal by more than a factor of two at the energies where the bunch of resonances is located. To estimate the feasibility of obtaining collision energy resolutions in this range with current experimental technology, we assume an experiment in which a Stark-decelerated packet of NH$_3$ molecules collides with a conventional beam of He atoms at a beam intersection angle of 45$^{\circ}$. We assume that the He atom beam is produced using a cryogenic source that is maintained at a temperature of about 50 Kelvin, resulting in a He atom velocity of 550 m/s. The relevant range of collision energies is then obtained when the velocity of the NH$_3$ molecules is tuned between 370 and 850 m/s. This is well within the range of state-of-the-art Stark deceleration molecular beam machines \cite{Scharfenberg:PRA79:023410}. We further assume velocity spreads of 10 m/s and 55 m/s for the NH$_3$ molecules and He atoms, respectively, and a spread in the beam intersection angle of 40 mrad due to the divergence of both beams. These are values that can realistically be obtained with current experimental techniques. With these parameters, we expect an optimum in the collision energy resolution to occur at a collision energy of 30 cm$^{-1}$, i.e., at the center of the relevant collision energy range. This maximum accuracy amounts to a spread of 3.1 cm$^{-1}$ (FWHM), while the collision energy spread increases to approximately 4 cm$^{-1}$ both for collision energies down to 20 cm$^{-1}$ and collision energies up to 50 cm$^{-1}$. These energy resolutions will not yet allow for the observation of single scattering resonances in the NH$_3$-He system, but they will certainly enable to observe the enhancement of about a factor of two in the inelastic cross section as a function of collision energy shown in Fig.~\ref{figconv}, revealing the combined effect of the underlying bunch of Feshbach resonances. An alternative and complementary approach to study scattering resonances is to measure differential cross sections. Referring back to Fig.~\ref{figdcs}, dramatic changes in the differential cross section can occur at collision energies where a resonance is observed. Feshbach resonances that give rise to strong backward scattering can be detected by measuring selectively the scattered flux in the backward direction. A similar approach has been used recently to measure partial-wave resolved resonances in the collision energy dependence of reactions between F atoms and HD molecules \cite{Dong:Science327:1501}. For inelastic scattering between NH$_3$ molecules and He atoms, differential cross sections are measured most conveniently using the velocity map imaging (VMI) technique \cite{eppink:97}. To experimentally resolve the angular dependence of the differential cross sections, large recoil velocities of the scattered molecules are advantageous. For the $|11+\rangle \leftarrow |11-\rangle$ channel, the recoil velocity of the scattered NH$_3$ molecules in the center of mass frame amounts to about 100~m/s at the most relevant collision energies. This is well within the range of velocities that can be imaged using current VMI techniques, offering interesting prospects to study the behavior of molecular scattering resonances. At the Fritz-Haber-Institute in Berlin, Germany, and the Radboud University Nijmegen, The Netherlands, we have embarked on an experimental program to study scattering resonances in both the integral and differential cross sections using Stark-decelerated molecular beams. It is the hope that the experimental study of these resonances will test our theoretical understanding of molecular interactions with unprecedented accuracy, and also will contribute to an enhancement of our ability to control the way in which molecules collide.
\section{Introduction} Understanding the scattering of cosmic rays in the interplanetary and interstellar plasmas is a problem of central importance in astrophysics. The diffusion processes of charged particles in the directions, parallel and perpendicular, to an ordered magnetic field (e.\,g., the magnetic field of the Sun) can be described by the diffusion tensor, whose components can be (1) calculated using analytical transport theories; (2) extracted from numerical test-particle simulations; and (3) obtained from heliospheric observations. Understanding of such observations is a key subject of space physics \cite{bie94:pal,dro00:rig}. An important way to test analytical theories \cite{jok66:qlt,tau06:sta} is numerical Monte-Carlo simulations \cite{gia94:mul,mic96:alf,gia99:sim,mic01:sim,tau10:pad} that operate under the same restrictions imposed on analytical calculations---stationary or static turbulence, prescribed turbulence geometry and power spectrum, no back-reaction of the particles on the turbulence field. Hence, such simulations are called ``test particle'' simulations although, using the same approach, other effects such as magnetic field line diffusion can be investigated. In this Note, the special case of \emph{isotropic} magnetostatic turbulence will be investigated, which is an important test case for both numerical and analytical approaches \cite{fis74:iso,bie88:iso,tau06:sta,sha09:hil}. It will be shown what the basic turbulence properties are and how magnetic turbulence is usually generated in numerical simulations (Sec.~\ref{iso}). Some problems will be discussed that are inherent in the basic formulation of numerical turbulence, because not all physical requirement such as vanishing magnetic divergence and isotropy in position and wavenumber space can be fulfilled at the same time. Finally, several simulation results will be compared and discussed (Sec.~\ref{res}). \section{Isotropic Turbulence}\label{iso} Isotropic turbulent magnetic fields can be thought of as a superposition of plane waves with random phase angles and random orientations. In the limit of an infinite number of plane waves, the resulting turbulence is homogeneous and isotropic \cite{bat82:tur}. There are a number of both analytical and numerical constraints and also pitfalls. Consider each in turn. \subsection{Analytical constraints} In general, homogeneous turbulence \cite{bat82:tur} is described using a stochastic approach that is based on a two-point, two-time correlation tensor \begin{equation} \bigl\langle B_l(\f x,t)\,B_m^\star(\f x',t')\bigr\rangle=\mathsf R_{lm}(\f x,\f x',t,t'), \end{equation} where $B_{l,m}$ refers to the turbulent magnetic field components with $l,m\in\{x,y,z\}$. To account for a power spectrum that is (at least partially) known in wavenumber space \cite{kol41:tur,kol91:tur}, a Fourier transform is applied, where the assumption of homogeneity leads to a delta function $\delta(\f k-\f k')$ \cite{bat82:tur,rs:rays,tau10:rag}. With the additional assumption of a time-independent turbulence field (i.\,e., magnetostatic turbulence), the result reads \begin{equation} \bigl\langle\hat B_l(\f k)\,\hat B_m^\star(\f k')\bigr\rangle=\delta(\f k-\f k')\,\mathsf P_{lm}(\f k). \end{equation} For homogeneous and isotropic turbulence, the correlation tensor has the form \cite{bat82:tur,rs:rays,sha09:nli} \begin{equation}\label{eq:iso_tens} \mathsf P_{lm}(\f k)=\frac{G(k)}{8\pi k^2}\left(\delta_{lm}+\frac{k_lk_m}{k^2}+i\sigma\epsilon_{lmn}\,\frac{k_n}{k}\right) \end{equation} with $\sigma(k)\in[-1,1]$ the magnetic helicity (usually assumed to be zero, with some noticeable exceptions) and $\epsilon_{lmn}$ the Levi-Civit\`a tensor. The normalization of the correlation tensor $\mathsf P_{lm}$ is given through the condition \cite{rs:rays} \begin{equation}\label{eq:bnorm} B^2=B_x^2+B_y^2+B_z^2=\int\ensuremath{\mathrm{d}}^3k\sum_{i=1}^3\mathsf P_{ii}(\f k), \end{equation} where $B$ corresponds to the (average) turbulent magnetic field strength. Consider now the three constraints for the isotropic turbulent magnetic field. \paragraph{Wave vectors} Isotropic turbulence means that each orientation of the wave vector has equal probability. Consider the Fourier transform (for illustration purposes in two dimensions) of an isotropic function \begin{align} \hat F(\f k)&=\int\ensuremath{\mathrm{d}}^2r\;F(r)e^{i\f k\cdot\f r}\nonumber\\ &=\int_0^\infty\ensuremath{\mathrm{d}} r\;rF(r)\int_0^{2\pi}\ensuremath{\mathrm{d}}\phi\;e^{ikr\cos(\psi-\phi)}, \end{align} with polar coordinates $\f k=(k\cos\psi,k\sin\psi)$ and $\f r=(r\cos\phi,r\sin\phi)$. Then \cite{gr:int} \begin{align} \int_0^{2\pi}\ensuremath{\mathrm{d}}\phi\;e^{ikr\cos(\psi-\phi)}&=\ensuremath{\sum_{n=-\infty}^\infty} i^nJ_n(kr)\int_0^{2\pi}\ensuremath{\mathrm{d}}\phi\;e^{in(\psi-\phi)}\nonumber\\ &=2\pi\,J_0(kr) \end{align} so that the result does not depend on the orientation of the wave vector as described through the angle $\psi$. Here, $J_n$ denotes the Bessel function of the first kind of order $n$. For the three-dimensional case, a similar calculation is slightly more involved. However, one can always transform to a new coordinate system $(r,\alpha,\beta)$ where $\alpha$ is defined through $\f k\cdot\f r=kr\cos\alpha$. Then $\beta$ is unused in the Fourier integral so that the two-dimensional case is recovered. \paragraph{Field strength} According to Eq.~\eqref{eq:bnorm}, the mean value of each individual field component is determined through\footnote{Strictly speaking, $k=0$ must be excluded because it represents a uniform magnetic field.} \begin{equation} B_i^2=\int\ensuremath{\mathrm{d}}^3k\;P_{ii}(\f k). \end{equation} Due to the fact that the isotropic turbulence tensor from Eq.~\eqref{eq:iso_tens} does not distinguish, for example, the $z$ from the $x,y$ directions it is immediately clear that, on average, $\langle B_x^2\rangle=\langle B_y^2\rangle=\langle B_z^2\rangle$. Any isotropic turbulence generator must therefore fulfill the constraint of equal amplitude field components in all three spatial directions. \paragraph{Divergence} Every magnetic field must obey Maxwell's equation of vanishing divergence, i.\,e., $\nabla\cdot\f B=0$ corresponding to the property that no magnetic monopoles exist (although, in the realm of quantum effects, the subject remains under active investigation \cite{gib11:mon}). \subsection{Numerical turbulence generation}\label{sim} Following the ideas of \cite{gia94:mul,mic96:alf,gia99:sim,mic01:sim}, the turbulence in the \textsc{Padian} code \cite{tau10:pad} is generated via a summation over $N$ plane wave modes as \begin{equation}\label{eq:dB} \f B(x,y,z)=\text{Re}\sum_{n=1}^N\hat{\f\xi}_nA(k_n)e^{i\left(k_nz'+\beta_n\right)}, \end{equation} where $\beta_n$ is the phase angle of the plane waves. The vector $\hat{\f\xi}_n$ denotes the amplitude direction of each wave and is defined as \begin{equation} \hat{\f\xi}_n=\cos(\alpha_n)\hat{\f e}_{x',n}+i\sin(\alpha_n)\hat{\f e}_{y',n}. \end{equation} where $\alpha_n$ is the polarization angle. The unit vectors $\hat{\f e}_{x',n}$ and $\hat{\f e}_{y',n}$ are given by the first and second lines, respectively, of a three-dimensional rotation matrix \begin{equation} \ensuremath{\Lambda}_{lm}= \begin{pmatrix} \;\cos\theta\cos\phi & \cos\theta\sin\phi & -\sin\theta\;\\ \;-\sin\phi & \cos\phi & 0\;\\ \;\sin\theta\cos\phi & \sin\theta\sin\phi & \cos\theta\; \end{pmatrix}. \end{equation} Physically, one can think of the angle $\alpha$ describing wave types that vary between fast-mode waves ($\alpha=0$) and Alfv\'en waves ($\alpha=\pi/2$), but since no time-dependence is considered, the analogy is limited. The direction of propagation of the plane waves, i.\,e., the $z'$ direction, results from $z'=x\,\ensuremath{\Lambda}_{31}+y\,\ensuremath{\Lambda}_{32}+z\,\ensuremath{\Lambda}_{33}$. Because the $z'$ direction is always perpendicular to $\hat{\f\xi}_n$, one immediately has $\f k_n\cdot\hat{\f\xi}_n=0$ for every mode $n$, which corresponds to $\nabla\cdot\f B=0$, thus ensuring that the turbulent magnetic field is divergence free. For each summand $n$, all angles $\theta$, $\phi$, $\alpha$, and $\beta$ are randomly generated\footnote{Note that not the angle $\theta$ but instead its cosine, $\eta=\cos\theta$, is uniformly distributed. This ensures that the density of wave directions is equal for all solid angles $\ensuremath{\mathrm{d}}\phi\,\ensuremath{\mathrm{d}}\eta$.} but are then kept fixed. Thus, the same $\f B$ results for the same set of coordinates $(x,y,z)$, corresponding to what is called a ``turbulence realization''. By constraining the angles $\theta$ and $\alpha$ \cite{gia99:sim,tau10:pad}, non-isotropic turbulence geometries such as slab and 2D can be obtained. To fulfill the additional constraint that $\f B\perp\ensuremath{\hat{\boldsymbol{e}}_z}$ \cite{gra96:sca,bie96:two}, the polarization angle has to be set to $\alpha=\pi/2$. The amplitude function $A(k_n)$ is defined through \begin{equation} A^2(k_n)=G(k_n)\ensuremath{\varDelta} k_n\left(\sum_{\nu=1}^NG(k_\nu)\ensuremath{\varDelta} k_\nu\right)^{\!-1}, \end{equation} where, for example, the turbulence spectrum $G(k_n)$ is of the form \cite{sha09:flr} \begin{equation}\label{eq:spect} G(k_n)=\frac{k_n^{\;q}}{\left(1+k_n^{\;2}\right)^{(s+q)/2}}, \end{equation} where $q$ and $s$ are the energy range and inertial range spectral indices, respectively. A logarithmic spacing of the wavenumbers is commonly used so that $\ensuremath{\varDelta} k_n/k_n$ is constant. \subsection{Normalization of the turbulence} The turbulent field as generated through Eq.~\eqref{eq:dB} should be normalized to unity.\footnote{Technically, such is due to the requirement that a unit magnetic field vector $\hat{\f e}_B$ is generated, which, in the equation of motion, is ``manually'' scaled with the factor $B/B_0$ to yield the requested turbulence strength relative to the mean magnetic field strength, $B_0$ \cite{tau10:pad}.} Therefore, it is required that, on average, \begin{equation} B^2=\m{B_x^2+B_y^2+B_z^2}=1. \end{equation} If such were not the case, the resulting transport parameters would be falsified, i.\,e., would be too large (small) if the magnetic field strength were to be smaller (greater) than unity. However, on implementing Eq.~\eqref{eq:dB} as is, one finds that: (i) the average strength of the turbulent fields is considerably smaller than unity; (ii) the $B_z$ component is (also on average) smaller than the other two components, which contradicts the requirement of isotropy. Both of these drawbacks are due to the fact that \begin{align} \m{\xi_x^2}&=\bigl\langle\abs{\cos\alpha\cos\theta\cos\phi-i\sin\alpha\sin\phi}^2\bigr\rangle=3/8\nonumber\\ \m{\xi_y^2}&=\bigl\langle\abs{\cos\alpha\cos\theta\sin\phi+i\sin\alpha\cos\phi}^2\bigr\rangle=3/8\nonumber\\ \m{\xi_z^2}&=\bigl\langle\left(-\cos\alpha\sin\theta\right)^2\bigr\rangle=1/4. \end{align} If the turbulent magnetic field components are divided by the mean values of the unit vector components, i.\,e., by $\sqrt{3/8}$ and by $1/2$, respectively, then all components of the turbulent magnetic field have equal means and the total magnetic field strength is approximately unity, as required. \subsection{Fulfilling physical constraints}\label{constr} One is faced with a choice because one has three options: \begin{itemize} \item using the original turbulence generation mechanism; but the original form for the turbulence is not in agreement with equal mean values for the three magnetic field components, as required for isotropy; \item normalizing the turbulence as described above; but the normalization factors are not compatible with the requirement that the divergence of the turbulent magnetic field be zero (note that one still has $\langle\nabla\cdot\f B\rangle=0$ on average); \item renormalizing the turbulent wave vector via $k_{x,y}\to k_{x,y}\sqrt{3/8}$ and $k_z\to k_z/2$ so that $\f k_n\cdot\hat{\f\xi}_n=0\;\forall n$ is restored; but then the wave vector is not isotropic any more. \end{itemize} \begin{figure}[tb] \centering \includegraphics[width=85mm]{isoComp_lp2} \caption{(Color online) The parallel scattering mean free path as a function of the particle speed. The errorbars with red triangles and blue dots show the simulation results from the \textsc{Padian} code \cite{tau10:pad} with and without renormalized wave vectors, respectively. Previous simulation results \cite{gia99:sim} do not use any normalization of the turbulent magnetic field (black squares). The black dashed line shows an analytical result \cite{tau08:soq}.} \label{ab:mfp} \end{figure} The important point to note is that, using the approach described by Eq.~\eqref{eq:dB}, not all three requirements can be fulfilled at the same time. Such is comparable, e.\,g., to various formulations of smoothed particle hydrodynamics \cite{vau08:sph,spr10:sph}, where conditions such as $\nabla\cdot\f B=0$ or the conservation of mass, energy, or angular momentum are frequently violated. It is therefore left to the resulting transport parameters to decide which option gives the best results. \section{Results and Conclusion}\label{res} In Fig.~\ref{ab:mfp}, the parallel mean free path is shown as resulting from two test-particle simulations in isotropic turbulence with a nominal turbulence strength $B=1$. Furthermore, the turbulent wave vector was scaled so that either $\f k_n\cdot\hat{\f\xi}_n=0$ or the orientation of $\f k$ is isotropic (see Sec.~\ref{constr}). However, such has only marginal influence on the resulting transport parameters compared to the estimated errors. In contrast, comparison of the classic results by Giacalone \& Jokipii \cite{gia99:sim}, where no turbulence renormalization had been done, with the \textsc{Padian} results shows systematical deviation, as clearly exhibited by Fig.~\ref{ab:mfp}. Therefore, the main deviation results from the fact that, in the \textsc{Padian} code, the turbulent magnetic field components have been normalized so that $B=1$ on average. Moreover, analytical results that have been derived using second-order quasi-linear theory \cite{sha05:soq,tau08:soq} agree better with test-particle simulations in a turbulent field with the correct turbulence strength. Such can be understood from the well-known fact that, as a rough estimate from classic quasi-linear theory \cite{jok66:qlt}, one has $\lambda\ensuremath{_\parallel}\propto(B/B_0)^{-2}$, thus underlining the important influence of the magnetic field strength on transport parameters. Here, $\ensuremath{\boldsymbol{B}_0}=B_0\ensuremath{\hat{\boldsymbol{e}}_z}$ denotes the mean magnetic field, which is usually assumed to be homogeneous. To conclude, using the conventional approach of superposing plane waves, it is not possible to create a strictly \emph{isotropic} turbulent magnetic field structure that obeys all physical constraints, which are (i) equal mean of all magnetic field components; (ii) isotropy of the wave vectors; and (iii) vanishing divergence of the magnetic field. Such magnetic fields are widely implemented in test-particle Monte-Carlo simulations, which are used to obtain (i) scattering mean free paths of charged particles; (ii) field line diffusion coefficients. While the turbulent magnetic field strength plays an important role for the results, such does not seem to be the case for a non-zero magnetic field divergence and/or the isotropy of the wave vectors. Future work should explore the possibility of a turbulence approach that is sufficiently simple but is fully compatible with all physical boundary conditions. \section*{Acknowledgments} The author thanks Andreas Shalchi, Ian Lerche, and Timo Laitinen for valuable comments. \input{simTurb2.bbl} \end{document} \endinput
\section{Introduction.} In this paper we study semiclassical solutions of the Cauchy problem for a time-dependent Schr\"{o}dinger equation on a metric graph. This article is, in a s certain sense, a continuation of \cite{cts}, \cite{mian} where you can find necessary definitions and formulations (nevertheless, we quote all necessary terms here, in section \ref{notation}). There (and in \cite{mz}) are also references to some articles and reviews related to the study of differential equations on metric graphs. In \cite{mian} a formula for the asymptotics of the number of Gaussian packets was written for a star graph, and a question of finding the leading coefficient for the case of an arbitrary finite compact graph was raised. In this paper that problem is solved and a general formula for that coefficient is obtained (see Theorem \ref{th_koef}). Moreover, in \cite{mian} the uniformity of distribution of packets for the special case of two vertices connected by three edges was proven. In this paper we present the proof of uniformity of the asymptotic distribution of packets over an arbitrary finite compact graph for almost all edge travel times (see Theorem \ref{th_ravnom}). In the second part (section \ref{neq}) we discuss a description of the asymptotic behavior of packets in the case of linearly dependent over $\mathbb{Q}$ edge travel times. In this situation there is no correspondence (see \cite{cts}, \cite{mian} for details) between the number of packets and the number of nodes of an integer lattice that lie on some faces of an expanding simplex, but analysis of their number is still possible. If rank of the system of edge travel times (lengthes) is equal to one then the number of packets grows only in a finite interval of time and stabilizes at a certain value, which depends on the lengths of cycles. Also we consider an example of a graph, whose rank of the system of lengthes equals two. \subsection{Acknowledgment.} V.\,L. Chernyshev thanks M.\,M. Skriganov, N.\,G. Moschevitin, P.\,B. Kurasov and O.\,V. Sobolev for useful discussions and attention to his work. Authors are grateful to A.\,I. Shafarevich for constant attention to their work. The work is done with partial finance support of grants MK-943.2010.1, RFFI 10-07-00617-a and 09-07-00327-a, RNP 2.1.1/11818 and state contract 14.740.11.0794. \subsection{Terms and Definitions.} \label{notation} Recall (see, e.g., \cite{pokkniga}, \cite{ku} and references therein), that a {\it metric graph} is a one-dimensional cell complex whose edges are parameterized curves. We denote a geometric graph by $\Gamma$, its edges by $\gamma_j$, its vertices by $a_j$. A set of all edges adjacent to the vertex $a$ we denote by $ \Gamma(a)$. We consider only finite metric graphs. Let $E$ and $V$ stand for the number of edges and the number of vertices respectively. \par Let $Q$ be an arbitrary real valued continuous function on $ \Gamma $, smooth on the edges. {\it Schr\"{o}dinger operator} \begin{equation} \widehat{H}\psi=-{h^2}\frac{d^2 \psi(x)}{d x^2} +Q(x)\psi(x)\label{shrodik} \end{equation} is defined on the set of functions from Sobolev spaces $\psi\in\oplus\sum\limits_{j}H^2(\gamma_j)$, satisfying the following boundary conditions at the vertices: \begin{enumerate} \item function $\psi$ is continuous on $\Gamma$; \item \begin{equation} \sum\limits_{\gamma_j\in\Gamma(a_m)}\alpha_j\frac{d \psi_j} {d x} (a_m)=0, \label{trans_nashe} \end{equation} in all internal vertices (i.e., the vertices of degree greater than one); \item $\psi(a_m)=0$ in all the external vertices, i.e. vertices of degree one. \end{enumerate} Here $\alpha_j=1$ for each edge emerging from the vertex, and $\alpha_j=-1$ for each incoming edge. These conditions are called {\it natural} (see \cite{kurasov}). They, in particular, ensure self-adjointness of the operator $\widehat{H}$. {\it A time-dependent Schr\"{o}dinger equation on the graph} $\Gamma$ is an equation of the form \begin{equation} \begin{array}{cc} \displaystyle ih\frac{{\partial}\psi}{{\partial}t}=\widehat{H}\psi, \label{shrodik_n} \end{array} \end{equation} where a semiclassical parameter $h\to 0$. We choose initial conditions that have the form of a narrow packet localized near the point $x_0$, which lies on the edge of the graph: \begin{equation} \psi(x,0)=h^{-1/4}K\exp\left(\frac{iS_0(x)}{h}\right). \label{nach_dannye} \end{equation} $$ S_0(x)=a(x-x_0)^2 + b(x-x_0) + c, $$ where $b$ and $c$ are real constants, and $a$ and $K$ are complex. The imaginary part of $a$ is positive. Normalization factor $h^{-1/4}$ is introduced to ensure that the initial function $\psi(x, 0)$ is of order one in the norm of $L^2(\Gamma)$. Due to the positivity of the imaginary part of $a$ the initial function is localized in a small neighborhood of $x_0$: $\psi(x, 0)=O(h^{\infty})$ with $| x-x_0 |\geq\delta>0$ ($\delta$ is independent of $h$). For simplicity we assume that there are no turning points (see, e.g., \cite{mf}, \cite{cwkb}) on the $\Gamma$ ; their presence can be accounted for in the standard way (see \cite{mf}). Asymptotic solution of the Cauchy problem (\ref{shrodik_n})--(\ref{nach_dannye}) is described in \cite{mian}, the explicit formulas are given therein. \begin{theorem} (See \cite{mian}) Solution of the Cauchy problem \ref{shrodik_n} -- \ref{nach_dannye} for $ t \in [0, T] $ ($T$ does not depend on $h$), is given by the following formula \begin{equation} \begin{array}{lr} \displaystyle \psi(x,t,h)=\sum_{j=1}^{N(t)}h^{-1/4}\varphi_j(t)e^{iS_j(x,t)/h}+O(\sqrt{h}), \label{rostok} \end{array} \end{equation} where the functions $\varphi_j(t), S_j(t)$ are explicitly expressed in terms of the solutions of two hamiltonian systems. \end{theorem} Each term in the sum (\ref{rostok}) is localized in a small neighborhood of $X_j(t)$. Here we assume that all the terms that are localized in the same point $X_j(t)$ form one {\it Gaussian packet}. Later, under the $N(t)$ we would mean the number of such packets. In \cite{mian} it is shown that in moments of penetration of the vertices of the graph a quantum packet is divided into $m$ packets ($m$ is a degree of the vertex): one is reflected and $m-1$ are ``diffused''. The amplitude of the packet is divided as $(m-2):2 $ (2 for the scattered and $(m-2)$ for the reflected packet). We consider the asymptotical behavior of function $\psi(x,\,t\,h)$ as $t{\rightarrow}\infty$. Namely, we will see how the number of Gaussian packets $N(t)$ changes in time. Note that this problem differs from the task of describing the asymptotic solution of the Schr\"{o}dinger equation at $t{\rightarrow}\infty$, as the error estimation is valid only for finite times. From a physical point view, this means that we are considering big $t$, but much smaller than $1/h$. Let $t_j$ stand for $j$-th edge travel time (this is an analog of length). \begin{lemma} Travel time of any edge of the graph depends only on the initial data and is the same for any Gaussian packet on each fixed edge. \end{lemma} \textbf{Proof.} Let us take an edge with index $j$. By construction a solution on each edge (using the method of Maslov complex germ), we find that $P^2+Q(x)=E_j$ holds (by the energy conservation law for Hamiltonian systems), where the value of $E_j$ is determined by the initial conditions. Writing $P$ as $\frac{d X}{dt}$, we obtain for edge travel time on the edge an explicit expression $t_j=\int\limits_a^b\frac{dx}{\sqrt{E_j-Q(x)}}$, where the integral is taken over the edge. The value of $E_j$ is the same for all edges, as the potential $Q(x)$ is assumed to be continuous function in the vertices, and $P$ can only change sign (as demonstrated by a construction of the solution). So the formula $E_j=P^2+Q(x)$ determines the same value $E_j=E$ for all packets. At the initial moment of time $E=\left(\frac{\partial S_0} {\partial x} \right)^2 + Q(x_0)$. Values of $t_j$ may be different, since the restriction of the potential $Q(x)$ on the edges may be different. {\bf Definition.} Let us consider the number of packets coming out of a fixed vertex to a fixed edge. The leading coefficient of the asymptotics of this number is called a radiation coefficient. Correctness of this definition will be shown in the proof of the theorem \ref{th_ravnom}. \section{The case of linearly independent over $\mathbb{Q}$ edge travel times.} \subsection{Main results.} \begin{theorem}[About uniformity of distribution] Consider a finite connected graph $\Gamma$. Suppose that for any vertex its degree is not equal to two. Suppose that there are no turning points on edges. Consider an edge $e$. Let $F_e(t)$ be a ratio of $N_e(t)$ (number of packets on the edge $e$) to $N(t)$ (total number of packets). Then for almost all incommensurable (i.e. linearly independent over $\mathbb{Q}$) numbers $t_1,\dots t_E$, a ratio of $F_e(t)$ to the length of $e$ tends to a constant $\left(\sum\limits_{j=1}^{E}t_j\right)^{-1}$ as time increases. \label{th_ravnom} \end{theorem} \begin{zam} It means that the distribution of number of packets tends to a uniform distribution as time increases. \end{zam} \textbf{Proof.} Let us choose on any edge with travel time $t_j$ a segment $dg$ with travel time $\tau$. Let us find $N_\tau(t)/N(t)$. We know (see \cite{mian}) that $N(t)=Ct^{E-1}+o(t^{E-1})$. Let us find $N_\tau(t)$. Since the number of packets changes only in vertices and there are no turning points, then: \begin{equation} N_\tau(t)=N_{{\rightarrow}d}(t)-N_{{\rightarrow}d}(t-\tau)+N_{{\rightarrow}g}(t)-N_{{\rightarrow}g}(t-\tau). \label{Ntau} \end{equation} Here $N_{{\rightarrow}d}(t)$ stands for the number of packets which arrived at the edge from point $d$. It is clear that the number of packets arrived at point $d$ at time $t$ equals the number of packets, which came out from the nearest vertex $a$ at time $t-T_1$. Here $T_1$ is a travel time from $a$ to $d$. By $N_{a{\rightarrow}d}(t)$ we denote the number of packets, which came from $a$ to $d$. We have to know asymptotics of the number of packets that come out of a vertex $a$. Packets can come out of the vertex only at times that are linear combinations (with nonnegative integer coefficients) of edge travel times. The number of release moments (when at least one packet comes out of the vertex $a$) is described by the number of set $\{n_j\}$ satisfying inequations of a kind: \begin{equation} n_1t_{l_1}+\ldots+n_mt_{l_m}{\leq}t, \label{ineq} \end{equation} where $t_j$ is a travel time of the $j$-th edge. Since the leading part of asymptotics of the number of packets is defined by the volume of a simplex defined by (\ref{ineq}), events with maximal numbers of summands happen more often. In other words, packets arrived at our vertex should visit all edges. I.e. \begin{equation} N_{a{\rightarrow}d}(t)=R^at^{E}+o(t^{E}). \label{Ra} \end{equation} For almost all $t_1,\dots t_E$ the estimation can be improved (see \cite{Skriganov}). There exists $K^a$ such that $N_{a{\rightarrow}d}(t)=R^at^{E}+K^at^{E-1}+o(t^{E-1})$. Let us show that $R^a$, which is called {\it a radiation coefficient}, does not depend on the choice of a vertex. Consider vertices $a$ and $b$. There exists a path connecting $a$ and $b$. Let $\delta$ be its travel time. Any packet coming out from $a$ to $d$ over time that does not exceed $2\delta$ generates at least one packet that come out from $b$ to $d'$. This is correct for packets coming out from $b$. We obtain inequations: $N_{a{\rightarrow}d}(t+2\delta){\geq}N_{b{\rightarrow}d'}(t)$ and $N_{b{\rightarrow}d'}(t+2\delta){\geq}N_{a{\rightarrow}d}(t)$. We know that $N_{a{\rightarrow}d}(t)=R^at^E+o(T^E)$, $N_{b{\rightarrow}d'}(t)=R^bt^E+o(T^E)$. Thus $R^at^E+o(t^E)=R^bt^E$. Hence $R^a=R^b$. Let us modify the expression for $N_\tau(t)$ $N_\tau(t)=R(t-T_1)^{E}+K^a(t-T_1)^{E-1}-R(t-T_1-\tau)^{E}-K^a(t-T_1-\tau)^{E-1}+R(t-T_2)^{E}+K^b(t-T_2)^{E-1}-R(t-T_2-\tau)^{E}-K^b(t-T_1-\tau)^{E-1}+o(t^{E-1})=2ER{\tau}t^{E-1}+o(t^{E-1})$. Thus we obtain \begin{equation} \frac{N_\tau(t)}{N(t)}{\rightarrow}\frac{2ER}{C}\tau. \label{otv} \end{equation} It remains to show that a coefficient near $\tau$ has the required form. We consequentially take edges as $dg$ and then summarize obtained expressions: \begin{equation} 1=\sum\limits_{j=1}^{E}\frac{N_{t_j}(t)}{N(t)}{\rightarrow}\frac{2ER}{C}\sum\limits_{j=1}^{E}t_j. \label{C_C1} \end{equation} Hence, \begin{equation} C=2ER\sum\limits_{j=1}^{E}t_j. \label{C_R_1} \end{equation} The proof is completed. \begin{zam} For any two vertices numbers $K^a$ and $K^b$ are related with inequation $|K^b-K^a|{\leq}E{\delta}R$. Here ${\delta}$ is a travel time for any path from $a$ to $b$. \end{zam} In the proof we obtain the following statement: \begin{sled}[Relation between coefficients $C$ and $R$] The leading coefficient for the number of packets $C$ and the radiation coefficient $R$, for almost all edge travel times, are related in the following manner: \begin{equation} C=2ER\sum\limits_{j=1}^{E}t_j. \label{C_R} \end{equation} \end{sled} \begin{theorem}[About the leading coefficient of the number of packets] Consider a finite connected compact graph $\Gamma$. Suppose that there are no vertices of degree two. Suppose that there are no turning points on edges. Then for almost all incommensurable numbers $t_1,\dots t_E$ the leading coefficient has the following form: \begin{equation} C=\frac{1}{2^{V-2}(E-1)!}\frac{\sum\limits_{j=1}^{E}t_j}{\prod\limits_{j=1}^{E}t_j}. \label{form_C} \end{equation} \label{th_koef} \end{theorem} \textbf{Proof} is based on (\ref{form_C}) and the following lemma. \begin{lemma} Let us consider a finite connected graph with incommensurable edge travel times $t_i (i=1 \ldots E)$ and a number of independent cycles $\beta$. Let B be an arbitrary vertex. Then for almost all edge travel times the number of packets arriving at B at time T asymptotically equals $$ R(T) \sim \frac{2^{\beta}}{2^E\, E!\, {\prod\limits_{j=1}^{E}t_j}} T^E. $$ \end{lemma} \begin{zam} This is equivalent to $$ R=\frac{1}{2^{V-1}E!}\frac{1}{\prod\limits_{j=1}^{E}t_j}. $$ \end{zam} \textbf{Proof.} Is is sufficient to consider paths of packets that traveled upon all edges. Only those paths give us the leading coefficient. Let A be an initial vertex. For each such path we can construct a ``code'' i.e. a sequence of coefficients of a corresponding chain with coefficients in $\mathbb{Z}_2$. It is clear that the code does not change under a path homotopy. Let us find the number of all possible codes. Consider cross connections, i.e. edges that are not in the spanning tree. Parity of passages on cross connections defines a path's homotopy class. All chain coefficients are defined by coefficients on cross connections. Thus the number of possible codes equals $2^\beta$. Now for every code $(c_1, \ldots, c_E), c_i \in \{ 0, 1 \}$ we associate times $$ \left\{ \sum_{i=1}^E t_i (c_i + 2 n_i) | n_i \in \mathbb{N} \cup \{ 0 \} \right\}. $$ At every such time (they are different) at least one packet arrives at the vertex $B$. The number of such times that are less than $T$ asymptotically equals to $$ \frac{T^E}{2^E \, E! \, t_1 \cdots t_E}. $$ Finally we summarize this over all possible codes. The proof of the lemma is finished. At the end we apply Euler's relation $\beta=E-V+1$ (see, for example, \cite{chris}). \section{The case of linearly dependent travel times.} \label{neq} In this section we assume that travel times are linearly dependent over $\mathbb{Q}$. It means that there is no one-to-one correspondence (described in \cite{cts}, \cite{mian}) between the number of packets and the number of integer lattice points in an expanding simplex. For the simplest example consider a star graph with three edges of the same length. The number of packets reaches three and does not increase. While the number of integer lattice points grows with time. \begin{statement} Consider a finite graph with edge travel times $t_1, = n_1t_0 \ldots,t_E = n_E t_0$, where $n_i \in \mathbb{N}$ and GCD$(n_1,\ldots,n_E) = 1$.\\ Then starting from a certain time, the number of packets becomes constant.\\ 1) If there exists a cycle with travel time $2 k t_0, k \in \mathbb{N}$ then $$ N(T) = 2 \sum_{i=1}^{E} n_i. $$ 2) Otherwise $$ N(T) = \sum_{i=1}^{E} n_i. $$ \end{statement} We omit the proof. \begin{statement} Consider a star graph with three edges $e_1, e_2, e_3$ with travel times $t_1 = n t_0, t_2 = m t_0, t_3$, where $n \in \mathbb{N}, m \in \mathbb{N}$ (GCD(n,m) = 1), and $t_3$ is such that rank $\{t_1, t_2, t_3\}$ over $\mathbb{Q}$ equals 2. Then the number of packets asymptotically equals $$ N(T) = \frac{T}{2}\left( \frac{ m + n }{t_3} + \frac{1}{t_0} \right) + o(T). $$ \end{statement} \textbf{Proof.} The number of packets changes only at times of a kind $2((n \alpha + m \beta)t_0 + t_3 \gamma )$, where $\alpha, \beta, \gamma$. Note that this representation is not unique. If we can present a time moment with $\alpha \ge 1, \beta \ge 1, \gamma \ge 1$, then at that moment $N(T)$ does not change (since packets arrive at a vertex from all three edges) \begin{zam} Number $n \alpha$ can be represented as $n\alpha' + m \beta'$ ($\beta' \ne 0$) if and only if $\alpha \ge m$. \end{zam} Hence to define asymptotics of $N(T)$ it is sufficient to consider the following times: \\ 1) $\gamma \ne 0, \beta = 0, 1 \le \alpha < m$. At those times $N(T)$ increases by 1. The number of such times equals the number of pairs $(\alpha, \gamma)$ such that $2(\alpha t_0 + \gamma t_3) < T$. The number asymptotically equals: $(m-1)\frac{T}{2t_3}$. \\ 2) $\gamma \ne 0, \alpha = 0, 1 \le \beta < n$. Similarly the number of pairs asymptotically equals $(n-1)\frac{T}{2t_3}$. \\ 3) $\alpha = 0, \beta = 0$. The number of such times that are less than $T$ increases as $\frac{T}{2t_3}$. Here the number of packets increases by 2. \\ 4) $\gamma = 0$. Let us show that linear combinations $\alpha n + \beta m$ contain all natural numbers that are greater than a certain fixed number. \begin{lemma} If $(m,n) = 1$ then there exists $M$ such that any $N > M$ can be presented as $N = \alpha n + \beta m$, where $\beta \ge 1, \alpha \ge 1 $. \end{lemma} \textbf{Proof.} Numbers $m, 2m, \ldots, nm$ give all residues modulo $n$: $$ \{ \beta m \}_{\beta = 1}^{n} = \{\beta_i m = i + n k_i | i \in [0,n-1], k_i \in \mathbb{Z}_{+} \}. $$ Let $k_0 = \max k_i$. Choose any $N > nk_0$. Then $N = i + nk \ (k > k_0, 0 \le i < n)$ and $N - \beta_i m = n (k - k_i)$. Thus any number that is greater than $n k_0$ can be represented as $\alpha n + \beta m$, where $1 \le \beta \le n, \alpha \ge 1 $. Asymptotics of the number of such times coincides with asymptotics of numbers $2kt_0 (k \in \mathbb{N})$ that are less than $T$. Since at such times 2 packets arrive from $e_1$ and $e_2$, then $N(T)$ increases by 1.
\section{Preliminaries}\label{crresults} Let $G$ be a split semisimple Chevalley group. By this we mean an affine algebraic group scheme over $\mathbb{Z}$, whose Lie algebra $\mathfrak{g}_{\mathbb{Z}}$ has a fixed Chevalley basis defined over $\mathbb{Z}$ corresponding to a root system $\Delta$. The elements of the Chevalley basis are the nilpotent elements $X_{\alpha}$ where $\alpha$ runs through all roots and the coroots $H_{\alpha} = [X_{\alpha}, X_{- \alpha}]$ where $\alpha$ runs through the simple positive roots. The structure constants of the Chevalley basis are all integers (cf.~{\cite{Chevalley}}). If $\alpha$ is a root, let $x_{\alpha} : \mathbb{G}_{\text{a}} \longrightarrow G$ be the one parameter subgroup tangent to $X_{\alpha}$. Let $T$ be the split maximal torus whose Lie algebra is spanned by the $H_{\alpha}$, and let $N$ (resp.~$N_-$) be the unipotent subgroup whose Lie algebra is spanned by the $X_{\alpha}$ (resp.~$X_{- \alpha}$) as $\alpha$ runs through the positive roots. Then $B = TN$ is the standard Borel subgroup of $G$. Let $F$ be a nonarchimedean local field and $\mathfrak{o}$ its ring of integers, $\mathfrak{p}$ the maximal ideal of $\mathfrak{o}$, and let $q = | \mathfrak{o} / \mathfrak{p} |$. We will denote the residue field $\mathbb{F}_q$. The group $G (F)$ has $K = G ( \mathfrak{o})$ as a maximal compact subgroup. The group $X_{\ast} (T)$ of rational cocharacters is isomorphic to $T (F) / T ( \mathfrak{o})$, where the one-parameter subgroup $\varphi \in X_{\ast} (T)$ corresponds to the coset $\varphi (\varpi) T ( \mathfrak{o})$ with $\varpi$ a prime element. If $w$ is an element of the Weyl group $W$, we will choose a fixed representative of $w$ in $G ( \mathfrak{o})$, and by abuse of notation we will denote this element also as $w$. Nothing will depend on this choice in any essential way. We will denote by $w_0$ the long Weyl group element. A character $\tau$ of $T (F)$ is called {\tmem{unramified}} if it is trivial on $T ( \mathfrak{o})$. We will let $W$ act on the right on unramified characters $\tau$, so that \[ (\tau w) (t) = \tau (wtw^{- 1}), \hspace{2em} w \in W, \hspace{2em} t \in T (F) . \] Since $\tau$ is unramified, this does not depend on the choice of representative $w$ of the Weyl group element. Let $\hat{G}$ be the connected Langlands dual group. It is an algebraic group over $\mathbb{C}$ whose root data are dual to $G$. Thus if $\Delta$ is the root system of $G$ with respect to $T$, we will denote by $\alpha \longrightarrow \alpha^{\vee}$ the bijection of $\Delta$ with system $\Delta^{\vee}$ of coroots, and $\Delta^{\vee}$ may be regarded as the root system of $\hat{G}$. If $\hat{T}$ is a maximal torus of $\hat{G}$ then the group $X_{\ast} (T)$ of rational one-parameter subgroups of $T$ is identified with the group $X^{\ast} ( \hat{T})$ of rational characters. Thus we have a homomorphism \begin{equation} \label{canmap} T (F) \longrightarrow T (F) / T ( \mathfrak{o}) \cong X_{\ast} (T) \cong X^{\ast} ( \hat{T}) . \end{equation} If $\boldsymbol{z} \in \hat{T} ( \mathbb{C})$ and $\lambda \in X^{\ast} ( \hat{T})$ we will denote by $\boldsymbol{z}^{\lambda}$ the application of the character $\lambda$ to $\boldsymbol{z}$. Also if $t \in T (F)$ we may apply the homomorphism (\ref{canmap}) to $t$ and apply the resulting rational character of $\hat{T}$ to $\boldsymbol{z}$; we will denote the result by $\tau_{\boldsymbol{z}} (t)$. Thus $\tau_{\boldsymbol{z}}$ is an unramified character of $T$ and $\boldsymbol{z} \longmapsto \tau_{\boldsymbol{z}}$ is an isomorphism of $\hat{T} ( \mathbb{C})$ with the group of unramified characters of $T (F)$. If $\lambda \in X^{\ast} ( \hat{T})$, we will denote by $a_{\lambda}$ a representative of the coset in $T (F) / T ( \mathfrak{o})$ corresponding to $\lambda$ by the isomorphism in (\ref{canmap}). Let $\tau = \tau_{\boldsymbol{z}}$ be such an unramified character. Let $M (\tau)$ be the space of the representation of $G (F)$ induced from $\tau$. This is the space of locally constant functions $f : G (F) \longrightarrow \mathbb{C}$ that satisfy \[ f (bg) = (\delta^{1 / 2} \tau) (b) f (g), \hspace{2em} b \in B (F), \] where $\delta : B (F) \longrightarrow \mathbb{R}^{\times}$ is the modular character. The standard intertwining integral $\mathcal{A}_w : M (\tau) \longrightarrow M (\tau w)$ is \[ \mathcal{A}_w^{\tau} \Phi (g) = \int_{N (F) \cap w^{- 1} N_- (F) w} \Phi (wng) \hspace{0.25em} dn. \] The integral is convergent for $\tau = \tau_{\boldsymbol{z}}$ with $| \alpha^{\vee} ( \boldsymbol{z}) | < 1$ when $\alpha \in \Delta^+$. It makes sense for other $\boldsymbol{z}$ by meromorphic continuation in a suitable sense. In order to convert between this notation (which is the same as Reeder {\cite{ReederVector}}) and that of Casselman~{\cite{Casselman}}, bear in mind that $\mathcal{A}^{\tau}_w$ is Casselman's $T_{w^{- 1}}$. Due to this difference, the Weyl group action on characters is a right action: $w : \tau \longrightarrow \tau w$. Let $J$ be the Iwahori subgroup of $G (F)$. It is the inverse image of $B ( \mathbb{F}_q)$ under the natural map $K \longrightarrow G ( \mathbb{F}_q)$. The dimension of the space $M (\tau)^J$ of $J$-fixed vectors is $|W|$. Let $\{\Phi_w^\tau\}$ and $\{\widetilde{\Phi}_w^\tau\}$ be the bases of $M (\tau)^J$ defined by (\ref{phidef}) and (\ref{phitildef}). The basis element $\Phi_w^\tau$ is denoted $\phi_w$ by Casselman~{\cite{Casselman}}. \begin{lemma} \label{wellknow}If $n \in N (F)$ and $w_0 n \in Bw_0 J$ then $n \in N ( \mathfrak{o})$. \end{lemma} \begin{proof} Using the Iwahori factorization $J = N_- ( \mathfrak{p}) T ( \mathfrak{o}) N ( \mathfrak{o})$ we may write $w_0 n = bw_0 \gamma n_1$ with $b \in B (F)$, $\gamma \in N_- ( \mathfrak{p}) T ( \mathfrak{o})$ and $n_1 \in N ( \mathfrak{o})$. Since $w_0 \gamma w_0^{- 1} \in B (F)$, this implies that $nn_1^{- 1} \in N (F) \cap w_0^{- 1} B (F) w_0$, so $n = n_1 \in N ( \mathfrak{o})$. \end{proof} \begin{lemma} \label{notdomzer}Let $\lambda \in X^{\ast} ( \hat{T})$. Then $\mathcal{W}_{\tau} \Phi^{\tau}_w (a_{\lambda}) = 0$ if $\lambda$ is not dominant. \end{lemma} \begin{proof} See Casselman and Shalika~{\cite{CasselmanShalika}} Lemma~5.1. (The proof there obviously applies to all Iwahori-fixed Whittaker functions since $J$ contains $N ( \mathfrak{o})$.) \end{proof} \begin{proposition} \label{wflongev} Given an unramified character $\tau$ of $T(F)$ we have $\Phi^{\tau}_{w_0} = \tilde{\Phi}^{\tau}_{w_0}$ and \[ \mathcal{W}_{\tau} \Phi_{w_0}^{\tau} (a_{\lambda}) = \left\{ \begin{array}{ll} \delta^{1 / 2} (a_{\lambda}) \boldsymbol{z}^{w_0 \lambda} & \text{if $\lambda$ is dominant,}\\ 0 & \text{otherwise} . \end{array} \right. \] \end{proposition} \begin{proof} The fact that $\tilde{\Phi}_{w_0} = \Phi_{w_0}$ is clear since $w_0$ is maximal in $W$. For the last assertion by Lemma~\ref{notdomzer} we may assume that $\lambda$ is dominant. Denoting $a = a_{\lambda}$, \[ \mathcal{W}_{\tau} \Phi_{w_0}^{\tau} (a) = \int_{N (F)} \Phi_{w_0}^{\tau} (w_0 na) \psi (n) \hspace{0.25em} dn. \] We make the variable change $n \longrightarrow ana^{- 1}$ whose Jacobian is $\delta (a)$, and the integral becomes \begin{eqnarray*} \delta (a) \int_{N (F)} \Phi_{w_0}^{\tau} (w_0 an) \psi (an a^{- 1}) \hspace{0.25em} dn & = & \\ \delta (a) \cdot \delta^{1 / 2} \tau (w_0 aw_0^{- 1}) \int_{N (F)} \Phi_{w_0}^{\tau} (w_0 n) \psi (ana^{- 1}) \hspace{0.25em} dn. & & \end{eqnarray*} Since $\delta^{1 / 2} (w_0 aw_0^{- 1}) = \delta^{- 1 / 2} (a)$, \[ \delta (a) \cdot \delta^{1 / 2} \tau (w_0 aw_0^{- 1}) = \delta^{1 / 2} (a) (\tau w_0) (a) = \delta^{1 / 2} (a) \boldsymbol{z}^{w_0 \lambda} . \] By Lemma~\ref{wellknow}, $\Phi_{w_0}^{\tau} (w_0 n) = 1$ if $n \in N ( \mathfrak{o})$ and $0$ otherwise. Since $\lambda$ is dominant, $\psi (ana^{- 1}) = 1$ when $n \in N ( \mathfrak{o})$, and the statement follows. \end{proof} Let $\tau = \tau_{\boldsymbol{z}}$ and define \begin{equation} \label{caltau} C_{\alpha} (\tau) = \frac{1 - q^{- 1} \boldsymbol{z}^{\alpha}}{1 - \boldsymbol{z}^{\alpha}} . \end{equation} \begin{proposition} We have \begin{equation} \label{wwithacomp} \mathcal{W}_{\tau w} \mathcal{A}_w^{\tau} = \prod_{\tmscript{\begin{array}{c} \alpha \in \Delta^+\\ w^{- 1} \alpha \in \Delta^- \end{array}}} \frac{1 - q^{- 1} \boldsymbol{z}^{- \alpha}}{1 - \boldsymbol{z}^{\alpha}} \mathcal{W}_{\tau} . \end{equation} \end{proposition} \begin{proof} This is Casselman and Shalika~{\cite{CasselmanShalika}}, Proposition~4.3. \end{proof} In (\ref{demdef}) we defined the Demazure operator $\partial_i = \partial_{\alpha_i}$ corresponding to a simple reflection $\alpha_i$. We will also make use of alternative version defined by \[ \partial_{\alpha_i}' f ( \boldsymbol{z}) = \partial_i' f ( \boldsymbol{z}) = \frac{f ( \boldsymbol{z}) - \boldsymbol{z}^{\alpha_i} f (s_i \boldsymbol{z})}{1 - \boldsymbol{z}^{\alpha_i}} = \frac{f (s_i \boldsymbol{z}) - \boldsymbol{z}^{- \alpha_i} f ( \boldsymbol{z})}{1 - \boldsymbol{z}^{-\alpha_i}} . \] Again we may define $\partial'_w = \partial_{i_1}' \cdots \partial'_{i_k}$ if $w = s_{i_1} \cdots s_{i_k}$ is a reduced decomposition. If $w = w_0$ and $\mu$ is dominant, then $\partial'_{w_0} ( \boldsymbol{z}^{w_0 \mu})$ is the character of the irreducible representation of highest weight $\mu$. \section{Whittaker functions} In this section we will prove that the Iwahori Whittaker functions $W (\Phi_w)$ are polynomials in $q^{- 1}$ whose constant terms are Demazure characters. \begin{proposition} \label{casselslick}Let $s = s_{\alpha}$ with $\alpha$ a simple root. Then \begin{equation} \label{secondid} \mathcal{A}_s \Phi^{\tau s}_w + C_{\alpha} (\tau) \Phi^{\tau}_{w_{}} = \left\{ \begin{array}{ll} \Phi^{\tau}_w + \Phi^{\tau}_{sw} & \text{if $sw < w$,}\\ q^{- 1} (\Phi^{\tau}_w + \Phi^{\tau}_{sw}) & \text{if $sw > w$.} \end{array} \right. \end{equation} \end{proposition} \begin{proof} The identities \[ \mathcal{A}_s \Phi^{\tau}_w = \left\{ \begin{array}{ll} (C_{\alpha} (\tau) - q^{- 1}) \Phi_w^{\tau s} + \Phi^{\tau s}_{sw} & \text{if $sw < w,$}\\ (C_{\alpha} (\tau) - 1) \Phi_w^{\tau s} + \Phi^{\tau s}_{sw} & \text{if $sw > w,$} \end{array} \right. \] are Casselman~{\cite{Casselman}}, Theorem 3.4. These imply that \begin{equation} \label{firstid} \mathcal{A}_s \Phi^{\tau}_w - (C_{\alpha} (\tau) - q^{- 1} - 1) \Phi^{\tau s}_{w_{}} = \left\{ \begin{array}{ll} \Phi^{\tau s}_w + \Phi^{\tau s}_{sw} & \text{if $sw < w$,}\\ q^{- 1} (\Phi^{\tau s}_w + \Phi^{\tau s}_{sw}) & \text{if $sw > w$.} \end{array} \right. \end{equation} Replacing $\tau$ by $\tau s$, noting that $C_{\alpha} (\tau s) = C_{- \alpha} (\tau)$ and using \begin{equation*} C_{\alpha} (\tau) + C_{- \alpha} (\tau) = 1 + q^{- 1} . \end{equation*} we obtain (\ref{secondid}). \end{proof} \begin{proposition} \label{demazuremi}Let $\alpha$ be a simple root, and $s = s_{\alpha}$ the corresponding reflection. Then \begin{equation} \label{altqmagic} \mathcal{W}_{\tau} ( \mathcal{A}_s \Phi^{\tau s}_w + C_{\alpha} (\tau) \Phi^{\tau}_{w_{}}) = (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial_{\alpha}' \mathcal{W}_{\tau} \Phi_w^{\tau} . \end{equation} \end{proposition} \begin{proof} By (\ref{wwithacomp}), we have \[ \mathcal{W}_{\tau s} \mathcal{A}_s^{\tau} = \frac{1 - q^{- 1} \boldsymbol{z}^{- \alpha}}{1 - \boldsymbol{z}^{\alpha}} \mathcal{W}_{\tau}, \] which we rewrite as \[ \mathcal{W}_{\tau} \mathcal{A}_s^{\tau s} = \frac{1 - q^{- 1} \boldsymbol{z}^{\alpha}}{1 - \boldsymbol{z}^{- \alpha}} \mathcal{W}_{\tau s} . \] Now the left-hand side in (\ref{altqmagic}) equals \[ \frac{1 - q^{- 1} \boldsymbol{z}^{\alpha}}{1 - \boldsymbol{z}^{- \alpha}} \mathcal{W}_{\tau s} \Phi^{\tau s}_w + \frac{1 - q^{- 1} \boldsymbol{z}^{\alpha}}{1 - z^{\alpha}} \mathcal{W}_{\tau} \Phi^{\tau}_w = (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \frac{1}{1 - \boldsymbol{z}^{\alpha}} (\mathcal{W}_{\tau} \Phi^{\tau}_w - \boldsymbol{z}^{\alpha} \mathcal{W}_{\tau s} \Phi^{\tau s}) . \] By the definition of the Demazure operator we have \[ \frac{1}{1 - \boldsymbol{z}^{\alpha}} (\mathcal{W}_{\tau} \Phi^{\tau}_w - \boldsymbol{z}^{\alpha} \mathcal{W}_{\tau s} \Phi^{\tau s}) = \partial'_{\alpha} \mathcal{W}_{\tau} \Phi^{\tau}_w, \] and we are done. \end{proof} Define operators on the ring $\mathcal{O}(\hat T)$ of functions on $\hat{T}(\mathbb{C})$ as follows. Let $s_i$ be a simple reflection corresponding to the simple root $\alpha=\alpha_i$. Define \begin{equation} \label{frakdpdef} \mathfrak{D}'_i=\mathfrak{D}'_\alpha=(1 - q^{- 1} \boldsymbol{z}^{\alpha})\partial_{\alpha}', \qquad\mathfrak{T}'_i=\mathfrak{T}'_\alpha=\mathfrak{D}'_i-1. \end{equation} \begin{proposition} \label{earlymagic} Suppose that $s = s_{\alpha}$ is a simple reflection and $s w < w$. Then \[ \mathcal{W}_{\tau} \Phi_{s w}^{\tau} = \mathfrak{T}'_\alpha\mathcal{W}_{\tau} \Phi^{\tau}_w .\] \end{proposition} \medbreak \begin{proof} This follows immediately from Propositions~\ref{demazuremi} and~\ref{casselslick}. \end{proof} The operator $\mathfrak{T}_i'$ is closely related to the Demazure-Lusztig operator. We will return to this point in Section~\ref{HeckeAlgebras}. At the moment, we make use of the Whittaker function to give a short proof that they satisfy the braid relation. Let $\mathcal{D}$ be the ring of expressions of the form $\sum_{w \in W} f_w \cdot w$ where $f_w \in \mathcal{O}( \hat{T})$, and the multiplication is defined by $(f_1 \cdot w_1) (f_2 \cdot w_2) = f_1 {^{w_1} f_2} \cdot w_1 w_2$. The $\mathfrak{D}_i$ are naturally elements of this ring. The ring $\mathcal{D}$ acts on $\mathcal{O}( \hat{T})$ in the obvious way. \begin{lemma} \label{dominantenough} Suppose that $D \in \mathcal{D}$ and that $D$ annihilates $\boldsymbol{z}^{w_0\lambda}$ for every dominant weight $\lambda$. Then $D = 0$. \end{lemma} \begin{proof} Define the {\tmem{support}} of $f \in \mathcal{O}( \hat{T})$ to be the finite set of weights with nonzero coefficients in $f$. Let $D = \sum_{w \in W} f_w \cdot w$. We may choose the dominant weight $\lambda$ so that the functions $f_w \boldsymbol{z}^{ww_0 \lambda}$ have disjoint support. Then $D\boldsymbol{z}^{w_0\lambda} = 0$ implies that each $f_w = 0$ and so $D = 0$. \end{proof} \begin{proposition}\label{whitbr} Let $s=s_i$ and $s_j$ be simple reflections. Then the operators $\mathfrak{T}'_i$ and $\mathfrak{T}'_j$ satisfy the same braid relations as $s_i$ and $s_j$. That is, if $k$ is the order of $s_is_j$ then \begin{equation} \label{braidp} \mathfrak{T}'_i\mathfrak{T}'_j\mathfrak{T}'_i\cdots = \mathfrak{T}'_j\mathfrak{T}'_i\mathfrak{T}'_j\cdots, \end{equation} where $k$ is the number of factors on both sides of this equation. \end{proposition} \begin{proof} By Lemma~\ref{dominantenough} it is enough to show that these both have the same effect on $\boldsymbol{z}^{w_0\lambda}$ where $\lambda$ is a dominant weight. By Proposition~\ref{wflongev}, $\boldsymbol{z}^{w_0\lambda}=c\mathcal{W}_\tau\Phi_{w_0}(a_\lambda)$ where the constant $c=(\delta^{1/2}\tau)(a_\lambda)$ is independent of $\boldsymbol{z}$ and hence commutes with operators in $\mathcal{D}$. Applying either side of (\ref{braidp}) to $\mathcal{W}_\tau\Phi_{w_0}(a_\lambda)$ gives $\mathcal{W}_\tau\Phi_{ww_0}(a_\lambda)$ where $w$ is the long element of the rank two Weyl group $\left<s_i,s_j\right>$. Indeed, Proposition~\ref{earlymagic} is applicable taking $(s,w)=(s_i,w_0)$, $(s,w)=(s_j,s_iw_0)$, etc.\ until we reach $ww_0$ since at each stage $s$ is a descent of $w$. \end{proof} \begin{proposition} \label{quadrel} The operators satisfy the quadratic relations \begin{equation} \label{quadrelation} (\mathfrak{D}'_i)^2=(1+q^{-1})\mathfrak{D}_i',\qquad (\mathfrak{T}'_i)^2=(q^{-1}-1)\mathfrak{T}_i'+q^{-1}. \end{equation} \end{proposition} \begin{proof} The two relations are equivalent. We prove the first. Writing $\alpha = \alpha_i$, \[ \mathfrak{D}_i^2 = (1 - q^{-1}\boldsymbol{z}^{\alpha}) \partial'_i (1 - q^{-1}\boldsymbol{z}^{\alpha}) \partial'_i = (1 - q^{-1}\boldsymbol{z}^{\alpha}) (\partial'_i)^2 - (1 - q^{-1}\boldsymbol{z}^{\alpha}) q^{-1} \partial'_i \boldsymbol{z}^{\alpha} \partial'_i \] and the quadratic relation in the form $\mathfrak{D}_i^2 = (v + 1)\mathfrak{D}_i$ follows from the properties $(\partial'_i)^2 = \partial_i'$ and $\partial'_i \boldsymbol{z}^{\alpha} \partial'_i=-\partial'_i$ of the usual Demazure operator. \end{proof} If $v$ is either an element of $\mathbb{C}$ or a field containing it, let $\mathcal{H}_v$ be the complex algebra generated by $T_i$ subject to the quadratic relations $T_i^2=(v-1)T_i+v$ together with the braid relations. We see from Propositions~\ref{quadrel} and~\ref{whitbr} that there is a representation of $\mathcal{H}_{q^{-1}}$ on the Whittaker model $\mathcal{W}_\tau M(\tau)^J$ in which $T_i$ acts by $\mathfrak{T}_i'$. Given any $w\in W$ we may construct an element $\mathfrak{T}_w'$ of $\mathcal{D}$ as follows. Let $w=s_{i_1}\cdots s_{i_k}$ be any reduced decomposition of $w$ into a product of simple reflections. Then $\mathfrak{T}_w'=\mathfrak{T}_{i_1}'\cdots\mathfrak{T}_{i_k}'$. This is well-defined by Proposition~\ref{whitbr}. Iwahori and Matsumoto~\cite{IwahoriMatsumoto} showed that the convolution ring of $J$-bi-invariant functions supported in $G(\mathfrak{o})$ is isomorphic to $\mathcal{H}_q$ and that algebra acts by right convolution on $\mathcal{W}M(\tau)^J$. The rings $\mathcal{H}_q$ and $\mathcal{H}_{q^{-1}}$ are isomorphic, and one might wonder whether the action of $\mathcal{H}_{q^{-1}}$ we have just defined is the same. It is not, since the isomorphism class of the convolution action depends on $\tau$, while the isomorphism class of the representation that we have just defined does not. \begin{theorem} \label{midmagic} Let $w\in W$ and let $\lambda$ be a dominant weight. Then \begin{equation} \mathcal{W}_\tau\Phi_{ww_0}^\tau(a_\lambda)= \delta^{1/2}(a_\lambda)\mathfrak{T}_w'\boldsymbol{z}^{w_0\lambda}. \end{equation} \end{theorem} \begin{proof} If $w=1$, this follows from Proposition~\ref{wflongev}. The general case follows by repeated applications of Proposition~\ref{earlymagic}. \end{proof} In passing from the functions $\mathcal{W}_\tau\Phi_w^\tau$ to $\mathcal{W}_\tau\widetilde\Phi_w^\tau$, the combinatorics of the Bruhat order begins to play a role. \begin{proposition} \label{deodchar}Let $s$ be a simple reflection and $w_1, w_2 \in W$. (i) Assume that $sw_1 < w_1$ and $sw_2 < w_2$. Then $w_1 \leqslant w_2$ if and only if $sw_1 \leqslant w_2$ if and only if $sw_1 \leqslant sw_2$. (ii) Assume that $sw_1 > w_1$ and $sw_2 > w_2$. Then $w_1 \geqslant w_2$ if and only if $sw_1 \geqslant w_2$ if and only if $sw_1 \geqslant sw_2$. \end{proposition} \begin{proof} Part (i) is a well-known property of Coxeter groups, called property $Z (s, w_1, w_2)$ by Deodhar~{\cite{Deodhar}}. Note that $w \longmapsto ww_0$ is an order reversing bijection of $W$. Applying this gives (ii). \end{proof} Suppose that $s = s_{\alpha}$ is a descent of $w \in W$: $sw < w$. Then we will define \[ H' (w, s) =\{u \in W|u, su \geqslant w\}. \] \begin{proposition} \label{cofinal}The set $H' (w, s)$ is cofinal in $W$ in the sense that if $u \in H' (w, s)$ and $t \geqslant u$ then $t \in H' (w, s)$. \end{proposition} \begin{proof} We have $t \geqslant u$ with both $u, su \geqslant w$. We wish to show that $t \in H' (w, s)$. We may assume without loss of generality that $su > u$. For if not, then $su < u$ so $t \geqslant su$. Thus interchanging $u$ and $su$ if necessary, we may assume that $su > u$. Also without loss of generality, $t > st$ since otherwise both $t, st$ are $\geqslant u \geqslant w$ as required. Now taking $w_1 = su$ and $w_2 = t$ in Proposition~\ref{deodchar}~(i), we see that both $t, st \geqslant u$ and so $t \in H' (w, s)$. \end{proof} Define an integer-valued function $c'_{w, s}$ on $W$ by \[ c'_{w, s} (u) = \sum_{\tmscript{\begin{array}{c} t \in H' (w, s)\\ t \leqslant u \end{array}}} (- 1)^{l (t) - l (u)} . \] \begin{theorem} \label{magictheorem}Let $\alpha$ be a simple root, and let $s = s_{\alpha}$ denote the corresponding reflection. Assume that $sw < w$. Then \begin{equation} \label{firstmagic} \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_{sw} = (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_w - q^{- 1} \sum_{u \in H' (w, s)} \mathcal{W}_{\tau} \Phi^{\tau}_u . \end{equation} Equivalently, \begin{equation} \label{secondmagic} \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_{sw} = (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_w - q^{- 1} \sum_{u \in H' (w, s)} c'_{w, s} (u) \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_u . \end{equation} \end{theorem} \begin{proof} By Proposition~\ref{demazuremi} \[ (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_w = \sum_{x \geqslant w} \mathcal{W}_{\tau} ( \mathcal{A}_s \Phi^{\tau s}_x + C_{\alpha} (\tau) \Phi^{\tau}_{x_{}}) . \] We split the sum into two parts according as $sx < x$ or $sx > x$ and use Proposition~\ref{casselslick}. We have \[ \sum_{\tmscript{\begin{array}{c} x \geqslant w\\ sx < x \end{array}}} \mathcal{W}_{\tau} ( \mathcal{A}_s \Phi^{\tau s}_x + C_{\alpha} (\tau) \Phi^{\tau}_{x_{}}) = \sum_{\tmscript{\begin{array}{c} x \geqslant w\\ sx < x \end{array}}} (\mathcal{W}_{\tau} \Phi_x^{\tau} + \mathcal{W}_{\tau} \Phi_{sx}^{\tau}) . \] By Proposition~\ref{deodchar}~(i) with $w_1 = w$ and $w_2 = x$, we see that \[ \bigcup_{\tmscript{\begin{array}{c} x \geqslant w\\ sx < x \end{array}}} \{x, sx\}=\{u \in W|u \geqslant sw\}. \] Therefore this contribution equals $\mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_{sw}$. On the other hand, let us consider the contributions from $sx > x$. By Proposition~\ref{casselslick} these contribute \[ q^{- 1} \sum_{\tmscript{\begin{array}{c} x \geqslant w\\ sx > x \end{array}}} (\mathcal{W}_{\tau} \Phi_x^{\tau} + \mathcal{W}_{\tau} \Phi_{sx}^{\tau}) = q^{- 1} \sum_{u \in H' (w, s)} \mathcal{W}_{\tau} \Phi^{\tau}_u . \] This proves (\ref{firstmagic}). By M\"obius inversion (Verma~{\cite{Verma}}, Deodhar~{\cite{Deodhar}}) we may write \[ \Phi_u^{\tau} = \sum_{t \leqslant u} (- 1)^{l (t) - l (u)} \tilde{\Phi}_t^{\tau}, \] and substituting this gives (\ref{secondmagic}). \end{proof} The function $c'_{w, s}$ has a tendency to take on only a few nonzero values. It vanishes off $H' (w, s)$. This sparseness means there are usually only a few terms on the right-hand side in (\ref{secondmagic}). For example, in $A_3$, if we consider the pairs $w,s$ where $s$ is a left descent of $w$, we find sixteen such pairs where $c_{w,s}'$ is always zero. Thus for these pairs the identity (\ref{secondmagic}) takes the form \[ \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_{sw}) = (1 - q^{- 1} \boldsymbol{z}^{- \alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} ( \tilde{\Phi}^{\tau s}_w) . \] There are seventeen pairs $(w,s)$ such that $c'_{w, s} (u) \neq 0$ for only one particular $u$. Then \[ \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_{sw}) = (1 - q^{- 1} \text{} \boldsymbol{z}^{- \alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} ( \tilde{\Phi}^{\tau s}_w) - q^{- 1} \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_u) . \] Finally, there are three cases where \[ \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_w) = (1 - q^{- 1} \text{} \boldsymbol{z}^{- \alpha}) \partial'_{\alpha} \mathcal{W}_{\tau} ( \tilde{\Phi}^{\tau s}_w) - q^{- 1} \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_u) - q^{- 1} \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_t) + q^{- 1} \mathcal{W}_{\tau s} ( \tilde{\Phi}^{\tau s}_x) . \] These are: \[ \begin{array}{|l|l|l|l|l|} \hline w & s & u & t & x\\ \hline s_2 & s_2 & s_1 s_2 & s_3 s_2 & s_1 s_3 s_2\\ \hline s_3 s_1 & s_1 & s_2 s_3 s_1 & s_3 s_2 s_1 & s_2 s_3 s_2 s_1\\ \hline s_3 s_1 & s_3 & s_2 s_3 s_1 & s_1 s_2 s_3 & s_2 s_1 s_2 s_3\\ \hline \end{array} \] The ring $\mathcal{O}( \hat{T})$ of regular functions on $\hat{T}$ is the complex group algebra over the weight lattice $\Lambda$ of $\hat{T}$. Let $v$ be an indeterminate, and let $\mathcal{R}=\mathbb{C}[v] \otimes \mathcal{O}( \hat{T})$. If $P \in \mathcal{R}$ we will denote by $P (\boldsymbol{z}, q^{- 1})$ the image of $P$ under the specialization map $\mathcal{R} \longrightarrow \mathcal{O}( \hat{T})$ that sends $v$ to~$q^{- 1}$. \begin{theorem} \label{wasthmone}Let $\lambda$ be dominant. There exists an element $P_{\lambda, w}$ of $\mathcal{R}$ such that \[ \mathcal{W}\tilde{\Phi}_w^{\tau} (a_{\lambda}) = \delta^{1 / 2} (a_{\lambda}) P_{\lambda, w} ( \boldsymbol{z}, q^{- 1}) . \] We have \[ P_{\lambda, w} ( \boldsymbol{z}, 0) = \partial_w' \boldsymbol{z}^{w_0 \lambda} . \] \end{theorem} It should be emphasized that $P_{\lambda, w}$ is independent of $q$ and of the field $F$. \begin{proof} This may be proved by induction on $l (ww_0)$. If $w = w_0$ then by Proposition~\ref{wflongev} we have $P_{\lambda, w_0} ( \boldsymbol{z}, q^{- 1}) = \boldsymbol{z}^{w_0 \lambda}$, so the assertion is true. In general, we may apply (\ref{secondmagic}) to deduce both statements for $sw$ when it is known for Weyl group elements $\geqslant w$. Indeed, by induction every term on the right-hand side is in $\mathcal{R}$, so $P_{\lambda, sw} \in \mathcal{R}$, and specializing $q^{- 1} \longrightarrow 0$ produces simply $\partial_i' P_{\lambda, w} = \partial'_i \partial_w' \boldsymbol{z}^{w_0 \lambda} = \partial'_{sw} \boldsymbol{z}^{w_0 \lambda}$. \end{proof} \section{\label{bottsam}Fibers of Partial Bott-Samelson Varieties} Let $G$ be a complex reductive group, and let $B$ be a Borel subgroup. In the application to Whittaker functions, $G$ will be the group formerly denoted $\hat{G} ( \mathbb{C})$, but we suppress the hat in this section. Let $X = G / B$ be the flag variety. If $w$ is an element of the Weyl group $W$, let $X_w^{\circ}$ be the image of $BwB$ in $X$. The closure \[ X_w = \bigcup_{u \leqslant w} X_u^{\circ} \] is the closed Schubert cell. Choose a reduced decomposition $\mathfrak{w} = (s_{h_1}, \cdots, s_{h_k})$ of $w = s_{h_1} \cdots s_{h_k}$ into a product of simple reflections. To define the Bott-Samelson variety $Z_{\mathfrak{w}}$, let $P_i$ be the parabolic subgroup generated by $B$ and $s_i$. The group $B^k$ acts on $P_{h_1} \times \cdots \times P_{h_k}$ on the right by \begin{equation} \label{borelaction} (p_1, \cdots, p_k) (b_1, \cdots, b_k) = (p_1 b_1, b_1^{- 1} p_2 b_2, \cdots, b_{r - 1} p_r b_r) . \end{equation} Then $Z_{\mathfrak{w}}$ is the quotient variety. The multiplication map $P_{h_1} \times \cdots \times P_{h_k} \longrightarrow G$ induces a rational map $Z_{\mathfrak{w}} \longrightarrow X_w$ that is a birational equivalence. Although $X_w$ may be singular $Z_{\mathfrak{w}}$ is always smooth, so this gives a resolution of the singularities of~$X_w$. If $s$ is an ascent of $w$ with respect to the Bruhat order then we have a partial Schubert variety $Z_{s, w}$ which is the quotient $(P \times X_w) / B$ where if $s = s_i$ then $P = P_i$ and $B$ acts on the right by $(p, x) \cdot b = (pb^{- 1}, bx)$. It is birationally equivalent to $X_{sw}$. We are interested in the map $\mu : Z_{s, w} \longrightarrow X_{sw}$ from the partial Bott-Samelson variety to the Schubert variety. The fiber over an open Schubert cell $Y_u$ (with $u \leqslant sw$) is either a single point or a $\mathbb{P}^1$, and we will find a combinatorial criterion for these cases. Let $T$ be a maximal torus of $G$ contained in $B$. Since the fibers of $\mu$ are constant on Schubert cells $Y_t \subset X_{sw}$ with $t \in W$, it suffices to study the fiber $\mu^{- 1} (y_t)$, where $y_t \in Y_t^T$ is the unique $T$-fixed point in the Schubert cell $Y_t$. Since the fiber $\mu^{- 1} (y_t)$ is either a single point or has dimension 1, it is determined by its Euler characteristic $\chi (\mu^{- 1} (y_t))$, and this is what we will compute. \begin{lemma} \label{fixedpointseuler}Let $Y$ be a projective complex algebraic variety with a $T$ action whose fixed point set $Y^T$ consists of isolated points. Then $\chi (Y) =\#Y^T$. \end{lemma} \begin{proof} Choosing a regular element $\lambda$ of $\tmop{Hom} ( \mathbb{C}^{\times}, T)$, it follows that $Y$ has a $\mathbb{C}^{\times}$ action with the same fixed point set, that is, $Y^{\mathbb{C}^{\times}} = Y^T$. This $\mathbb{C}^{\times}$ action defines a Bialynicki-Birula cellular decomposition of $Y$, with cells $\{U_x \}_{x \in Y^{\mathbb{C}^{\times}}}$ defined by \[ U_x =\{z \in Y \mid \lim_{\varepsilon \to 0} \varepsilon \cdot z = x\}, \] one cell for each $x$. See Bialynicki-Birula, Carrell and McGovern {\cite{BialynickiBirula}}. Since the cells are all of even real dimension, the Euler characteristic of $Y$ is simply the number of cells - that is, the number of fixed points. \end{proof} \begin{proposition} \label{fiberdet}The fiber of $\mu$ over $Y_u$ is $\mathbb{P}^1$ if and only if both $u, su \leqslant w$, and is a point otherwise. \end{proposition} \begin{proof} In view of Lemma~\ref{fixedpointseuler}, in order to compute $\chi (\mu^{- 1} (y_t))$, we need to compute the number of fixed points in the set $\mu^{- 1} (y_t)^T$. This is straightforward since the fixed point set $Z_{s, w}^T$ equals $\{(u, t) \mid u \in \left<s\right>, \; t \leq w\}$ and the map $\mu^T : Z_{s, w}^T \longrightarrow X_{sw}^T$ is multiplication $(u, t) \mapsto ut$. Discussions of these facts may be found in many places, usually for ``standard'' Bott-Samelson varieties rather than these partial ones. See, for example Brion~{\cite{Brion}}. Now, from these two facts, we compute $(\mu^T)^{- 1} (y_t)$ for $t \leqslant w$. In general, we have \[ (\mu^T)^{- 1} (y_t) =\{(u, x) \,|\, \text{$u \in \left<s\right>$, $x \leqslant w$ and $ux = t$}\}. \] Since $\left<s\right>$ has order two, there are at most two points in $(\mu^T)^{- 1} (y_t)$. One of them is the point $y_{(1, t)}$, which is the image of the affine point $(1, t)$ in $\mathbb{P}^1$. The other possibility is the $y_{(s, s t)}$; but this point is only a point of $Z_{s, w}^T$ if $st \leqslant w$. Thus we conclude that $(\mu^T)^{- 1} (y_t)$ is in bijection with the elements $z$ such that $t$ and $st$ are less than or equal to $w$. The map $\mu$ is an isomorphism over the big cell $Y_{sw}$. Thus it remains to study the fibers over the cells $Y_u$ with $u \leqslant w$. It now suffices to to show that the fiber over $u$ is a $\mathbb{P}^1$ if and only if both $u, su \leqslant w$, and is a point otherwise. Thus we must show that the Euler characteristic of $\mu^{- 1} (y_u)$ is equal to 2 if and only if both $u, su \leqslant w$, and is equal to one otherwise. By Lemma~\ref{fixedpointseuler}, we must show that the cardinality of $(\mu^T)^{- 1} (y_u)$ is equal to 2 if and only if both $u, su \leqslant w$, and is equal to 1 otherwise. But, as explained above, $(\mu^T)^{- 1} (y_u)$ is the one element set $y_{(1, u)}$ unless both $u, su \leqslant w$, in which case $(\mu^T)^{- 1} (y_u)$ is the 2 element set $\{y_{(1, u)}, y_{(s, su)} \}$. \end{proof} Let us reformulate this proposition using the notation (\ref{suggested}) from the introduction. Assuming that $sw > w$, define \[ H (w, s) =\{u \in W|u, su \leqslant w\}. \] As in Proposition~\ref{cofinal} the set $H (w, s)$ has the property that if $u \in H (w, s)$ and $t \leqslant u$ then $t \in H (w, s)$. Define \[ c_{w, s} (u) = \sum_{\tmscript{\begin{array}{c} t \in H (w, s)\\ t \geqslant u \end{array}}} (- 1)^{l (t) - l (u)} . \] \begin{proposition} \label{fibersymbolic}Let $s$ be a left ascent of $w$. Then \[ \{y \in X_{sw}\,|\, \text{\rm $\mu^{- 1} (y)$ is nontrivial} \}= \sum_{u \in H (w, s)} c_{w, s} (u) X_u . \] \end{proposition} \begin{proof} Let $y \in X$. Let $t \in W$ such that $y \in Y_t$. Then \[ \sum_{\tmscript{\begin{array}{c} u \in H (w, s)\\ y \in X_u \end{array}}} c_{w, s} (u) = \sum_{\tmscript{\begin{array}{c} u \in H (w, s)\\ t \leqslant u \end{array}}} c_{w, s} (u) . \] It follows from M\"obius inversion (Verma~{\cite{Verma}} or Deodhar~{\cite{Deodhar}}) that given $y \in X$ that this is $1$ if $t \in H (w, s)$ and $0$ otherwise. Thus the statement follows from Proposition~\ref{fiberdet}. \end{proof} \section{Proof of Theorems~\ref{thmain} and~\ref{qone}} To prove Theorem~\ref{thmain} we may specialize $v = q^{- 1}$ with $q$ the cardinality of the residue field. Let $\theta : \hat{T} ( \mathbb{C}) \longrightarrow \hat{T} ( \mathbb{C})$ be the map that sends $\boldsymbol{z} \longmapsto \boldsymbol{z}^{- 1}$. Then \begin{equation} \label{delconj} \partial_i = \theta \partial_i' \theta \end{equation} Let $\lambda$ be a dominant weight of $\hat{G} ( \mathbb{C})$. Then $- w_0 \lambda$ is also a dominant weight, and (\ref{bigxdef}) may be written \begin{equation} \label{xdefined} \boldsymbol{X}_w (\lambda) = \delta (a_{- w_0 \lambda})^{- 1 / 2} \theta \mathcal{W}_\tau\tilde{\Phi}^\tau_{ww_0} (a_{- w_0 \lambda}) , \end{equation} where $\tau=\tau_{\boldsymbol{z}}$. Similarly \begin{equation} \label{ydefined} \boldsymbol{Y}_w (\lambda) = \delta (a_{- w_0 \lambda})^{- 1 / 2} \theta \mathcal{W}_\tau \Phi_{ww_0}^\tau (a_{- w_0 \lambda}) . \end{equation} We have $u \in H (s, w)$ if and only if $uw_0 \in H' (s, ww_0)$. Note that $sww_0 < ww_0$ so that $H' (s, ww_0)$ is defined, and it is also easy to see that $c_{w, s} (u) = c_{ww_0, s}' (uw_0)$. \begin{proposition} \label{correctionterms}Let $\lambda$ be a dominant weight. Then \begin{equation} \label{wehave} \boldsymbol{X}_1 (\lambda) =\boldsymbol{Y}_1 (\lambda) =\boldsymbol{z}^{\lambda} . \end{equation} Assume that $s = s_{\alpha}$ is a simple reflection and that $sw > w$. Then \begin{equation} \label{schubid} \boldsymbol{X}_{sw} (\lambda) = (1 - q^{- 1} \boldsymbol{z}^{- \alpha}) \partial_{\alpha} \boldsymbol{X}_w (\lambda) - q^{- 1} \sum_{\tmscript{\begin{array}{c} u \in H (w, s) \end{array}}} c_{w, s} (u)\boldsymbol{X}_u (\lambda) . \end{equation} \end{proposition} \begin{proof} Equation~(\ref{wehave}) follows from Proposition~\ref{wflongev}. To prove (\ref{schubid}) we begin with (\ref{secondmagic}), with $w$ replaced by $ww_0$. Applying $\theta $ and using (\ref{delconj}) we obtain \[ \theta \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_{sww_0} = (1 - q^{- 1} \boldsymbol{z}^{- \alpha}) \partial_{\alpha} \theta \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_{ww_0} - q^{- 1} \sum_{uw_0 \in H' (ww_0, s)} c'_{ww_0, s} (uw_0) \theta \mathcal{W}_{\tau} \tilde{\Phi}^{\tau}_u . \] Now evaluating this at $a_{- w_0 \lambda}$ and multiplying by the constant $\delta (a_{- w_0 \lambda})^{- 1 / 2}$ (which depends on $\lambda$ but not $\boldsymbol{z}$) we obtain (\ref{schubid}). \end{proof} Combining Propositions~\ref{fibersymbolic} and~\ref{correctionterms} gives Theorem~\ref{thmain}. To prove Theorem~\ref{qone}, we will take $q = 1$. Then the operators $\mathfrak{D}_i$ simplify to \[ \mathfrak{D}_i f ( \boldsymbol{z}) = f (z) - \boldsymbol{z}^{- \alpha_i} f (s_i \boldsymbol{z}) . \] These operators do not satisfy the braid relation. We show that if we define $\widehat{\boldsymbol{X}}_w ( \boldsymbol{z})$ to be the right-hand side of (\ref{permutahedron}), and if $s = s_i$ with $l (sw) > l (w)$, then \begin{equation} \label{xprimerecursion} \widehat{\boldsymbol{X}}_{sw} = \mathfrak{D}_i \widehat{\boldsymbol{X}}_w - \sum_{u \in H (w, s)} c (u) \widehat{\boldsymbol{X}}_u, \end{equation} where $c (u)$ is as in Theorem~\ref{thmain}. The dominant weight $\lambda$ will be fixed and we suppress it from the notation. By comparison with Theorem~\ref{thmain}, $\boldsymbol{X}_w$ (with $q = 1$) and $\widehat{\boldsymbol{X}}_w$ both satisfy the same recursion, and agree when $w = 1$; therefore by induction on the Bruhat order, they are equal. Let $\widehat{\boldsymbol{Y}}_w = \boldsymbol{z}^{w (\rho + \lambda) - \rho}$. The identity (\ref{xprimerecursion}) is equivalent by M\"obius inversion to \begin{equation} \label{equivdc} \mathfrak{D}_i \widehat{\boldsymbol{X}}_w - \sum_{u \in H (w, s)} \widehat{\boldsymbol{Y}}_u, \end{equation} Since $\boldsymbol{z}^{- \alpha} \boldsymbol{z}^{- s \rho} = \boldsymbol{z}^{- s \rho}$, we have \[ \mathfrak{D}_i \widehat{\boldsymbol{Y}}_u = \widehat{\boldsymbol{Y}}_u - \widehat{\boldsymbol{Y}}_{su} . \] Therefore \begin{equation} \label{possdc} \mathfrak{D}_i \widehat{\boldsymbol{X}}_w = \sum_{u \leqslant w} (- 1)^{l (u)} \widehat{\boldsymbol{Y}}_u + \sum_{su \leqslant w} (- 1)^{l (u)} \widehat{\boldsymbol{Y}}_u, \end{equation} where in the second term we have replaced $u$ by $su$. By Proposition~\ref{deodchar} (ii), we have $u \leqslant sw$ if and only $\min (u, su) \leqslant w$ where the notation $\min (u, su)$ makes sense since $u$ and $su$ are always comparable in the Bruhat order. So \[ \{u|u \leqslant sw\}=\{u|u \leqslant w\} \cup \{u|su \leqslant w\} \] and we may write \[ \widehat{\boldsymbol{X}}_{sw} = \sum_{u \leqslant w} (- 1)^{l (u)} \widehat{\boldsymbol{Y}}_u + \sum_{su \leqslant w} (- 1)^{l (u)} \widehat{\boldsymbol{Y}}_u - \sum_{u, su \leqslant w} (- 1)^{l (u)} \widehat{\boldsymbol{\boldsymbol{Y}}}_u \] where we have subtracted the terms that are double counted in the first two sums. Now using (\ref{possdc}) we obtain (\ref{equivdc}). \section{\label{HeckeAlgebras}Hecke algebras} \,The space of Iwahori fixed vectors $M (\tau)^J$ for a fixed unramified character $\tau$ is isomorphic to the finite-dimensional Hecke algebra $\mathcal{H}$ of compactly supported $J$-bi-invariant functions on $G (F)$ which have support inside of $G (\mathfrak{o})$. We recall the isomorphism, following Bump and Nakasuji~{\cite{BumpNakasuji}}, who in turn follow Rogawski~{\cite{Rogawski}}. In this isomorphism, $f \in M (\tau)^J$ corresponds to the element $\varrho_{\tau} (f) \in \mathcal{H}$ where \[ \varrho_{\tau} (f) (g) = \left\{ \begin{array}{ll} f (g^{- 1}) & \text{if $g \in G (\mathfrak{o})$,}\\ 0 & \text{otherwise.} \end{array} \right. \] Both $M (\tau)^J$ and $\mathcal{H}$ are left $\mathcal{H}$-modules, where the action of $\mathcal{H}$ on $M (\tau)$ is \[ \phi \cdot f (g) = \int_{G (\mathfrak{o})} \phi (x) f (g x) \, d x. \] Then $\varrho_{\tau}$ is a homomorphism of left $\mathcal{H}$-modules. The Hecke algebra $\mathcal{H}$ has the following description by generators and relations, due to Iwahori and Matsumoto {\cite{IwahoriMatsumoto}}. If $t_w$ is the characteristic function of the double coset $J w J$ then $\mathcal{H}$ is generated by the $t_i = t_{s_i}$ where $s_i$ is a simple reflection. The generators satisfy the quadratic relation \[ t_i^2 = (q - 1) t_i + q, \] together with the braid relations. Thus $t_i=t_{s_i}$. The braid relations and the quadratic relations give a presentation of $\mathcal{H}$ which is valid even if $q$ is not a prime. Let $v$ be an indeterminate. Let $\mathcal{H}_v$ be the abstract algebra over $\mathbb{C}[v,v^{-1}]$ generated by $T_1, \cdots, T_r$ subject to the braid relations together with the condition that \[ T_i^2 = (v - 1) T_i + v. \] There is an antihomomorphism $\sigma:\mathcal{H}_v\to\mathcal{H}$ defined by $\sigma(v)=q^{-1}$ and $\sigma(T_w)=t_w^{-1}$. In particular \begin{equation} \label{sigmat} \sigma(T_i+1)=q^{-1}(t_i+1). \end{equation} The operators $\mathfrak{T}_i=\mathfrak{D}_i - 1$ are closely related to the \textit{Demazure-Lusztig operators} that were introduced in Lusztig~\cite{LusztigK1} equation (8.1). These are the operators defined by \[ \mathcal{L}_i (\boldsymbol{z}^{\lambda}) =\mathcal{L}_{i, v} (\boldsymbol{z}^{\lambda}) = \frac{\boldsymbol{z}^{\lambda} -\boldsymbol{z}^{s_i \lambda}}{\boldsymbol{z}^{\alpha_i} - 1} - v \frac{\boldsymbol{z}^{\lambda} -\boldsymbol{z}^{\alpha + s_i \lambda}}{\boldsymbol{z}^{\alpha_i} - 1} . \] Lusztig shows that these satisfy the same relations as the $T_i$ in the (finite) Hecke algebra, and we will also prove this in the equivalent form of Proposition~\ref{braidrel} below. The precise relationship between our operators and Lusztig's is as follows. Lusztig's operators $\mathcal{L}_i$ satisfy the quadratic relation \begin{equation} \label{dlrel} \mathcal{L}_i^2 = (v - 1)\mathcal{L}_i + v. \end{equation} Replacing $v$ by $v^{- 1}$ and multiplying by $- v$ let $\mathcal{L}_i' = - v\mathcal{L}_{i, v^{- 1}}$. We have \begin{equation} \label{dlprel} \mathcal{L}_i' \boldsymbol{z}^{\lambda} = \frac{- v\boldsymbol{z}^{\lambda} -\boldsymbol{z}^{\alpha_i + s_i \lambda} + v\boldsymbol{z}^{s_i \lambda} +\boldsymbol{z}^{\lambda}}{\boldsymbol{z}^{\alpha_i} - 1} . \end{equation} It follows from (\ref{dlrel}) that the $\mathcal{L}_i'$ satisfy the same braid and quadratic relations as the $\mathcal{L}_i$ and hence there is a representation of $\mathcal{H}_v$ in which $T_i \longrightarrow \mathcal{L}_i'$. We recall that if $\lambda$ is a weight, then $\zeta^\lambda$ is the operation on $\mathcal{O}(\hat{T})$ defined by multiplication by $\boldsymbol{z}^{-\lambda}$. It follows easily from (\ref{dlprel}) that \begin{equation} \label{usvsthem} \zeta^{- \rho} (\mathfrak{D}_i - 1) \zeta^{\rho} =\mathcal{L}'_i . \end{equation} Thus the following proposition is equivalent to the result of~\cite{LusztigK1}. Our proof is different, since we use Whittaker functions to prove the braid relations. \begin{proposition} \dueto{Lusztig~\cite{LusztigK1}} \label{braidrel} The operators $\mathfrak{T}_i =\mathfrak{D}_i - 1$ satisfy the braid relations and the quadratic relations of $\mathcal{H}_v$. Thus $T_i \longmapsto \mathfrak{T}_i$ is a representation of~$\mathcal{H}_v$. \end{proposition} \begin{proof} The corresponding facts for $\mathfrak{T}'_i$, which is defined by (\ref{frakdpdef}), are Proposition~\ref{quadrel} for the quadratic relation and Proposition~\ref{whitbr} for the braid relation. Since $\mathfrak{T}_i=\theta\mathfrak{T}_i'\theta$, the result follows. \end{proof} We have now defined an action of $\mathcal{H}_v$ on $\mathbb{C}[v,v^{-1}]\otimes \mathcal{O}(\hat{T})$, so Theorem~\ref{amazing} now has meaning. In order to prove it we will have define certain maps $\xi_\lambda:\mathcal{H}_v\to \mathbb{C}[v,v^{-1}]\otimes \mathcal{O}(\hat{T})$ and that is our next goal. These maps $\xi_\lambda$ also turn out to be the key to extend the action to the affine Hecke algebra $\widetilde{\mathcal{H}}_v$. \begin{lemma} \label{bylemma} Let $s = s_{\alpha}$ be a simple reflection, and let $A^{\tau s}_s : \mathcal{H} \longrightarrow \mathcal{H}$ be the homomorphism of left $\mathcal{H}$-modules that makes the following diagram commutative: \[\begin{CD} M(\tau s)@>\mathcal{A}_s^{\tau s}>>M(\tau)\\ @VV\varrho_{\tau s}V @VV\varrho_{\tau}V\\ \mathcal{H}@>A_s^{\tau s}>> \mathcal{H} \end{CD}\] Then $A_s^{\tau s}$ is right multiplication by \begin{equation} \label{righttrans} \frac{1}{q} t_i + \left( 1 - \frac{1}{q} \right) \frac{\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}}, \end{equation} where $\tau = \tau_{\boldsymbol{z}}$. \end{lemma} \begin{proof} We recall that $\varrho_\tau$ is a homomorphism of left $\mathcal{H}$-modules. The composition $\varrho_{\tau} \circ \mathcal{A}_s^{\tau s} \circ \varrho_{\tau s}^{- 1}$ is thus a homomorphism of left $\mathcal{H}$-modules $\mathcal{H} \longrightarrow \mathcal{H}$ and since $\mathcal{H}$ is a ring it must consist of right multiplication by some element. The scalar may be evaluated by applying $\varrho_{\tau} \circ \mathcal{A}_s^{\tau s} \circ \varrho_{\tau s}$ to the unit element of $\mathcal{H}$, which corresponds under $\varrho^{- 1}$ to $\Phi_1^{\tau s}$. By Casselman~{\cite{Casselman}}, Theorem 3.4 we have \[ \mathcal{A}_s^{\tau} \Phi^{\tau}_1 = \frac{1}{q} \Phi^{\tau s}_s + \left( 1 - \frac{1}{q} \right) \frac{\boldsymbol{z}^{\alpha}}{1 -\boldsymbol{z}^{\alpha}} \Phi^{\tau s}_1 \] so \[ \mathcal{A}_s^{\tau s} \varrho_{\tau s}^{- 1} (1) =\mathcal{A}_s^{\tau} \Phi^{\tau s}_1 = \frac{1}{q} \Phi^{\tau}_s + \left( 1 - \frac{1}{q} \right) \frac{\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} \Phi^{\tau}_1.\] Applying $\rho_\tau$ gives \[\frac{1}{q} t_i + \left( 1 - \frac{1}{q} \right) \frac{\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} . \] \end{proof} Let us fix a dominant weight $\lambda$ for $\hat{T}$. Then we will define a map $\xi'_{\lambda} : \mathcal{H} \longrightarrow \mathcal{O}( \hat{T})$, where $\mathcal{O}( \hat{T})$ is the ring of rational functions on $\hat{T}$, which is the group algebra of the weight lattice $\Lambda$ of $\hat{T}$. We regard a rational function on $\hat{T}$ as a function on the complex points, and to specify $\xi_{\lambda} (\phi)$ for $\phi \in \mathcal{H}$, it is enough to specify its value $\xi_{\lambda} (\phi) (\boldsymbol{z})$ for $\boldsymbol{z} \in \hat{T} (\mathbb{C})$. Let $\tau = \tau_{\boldsymbol{z}}$ and $\Phi = \varrho_{\tau}^{- 1} (\phi) \in M (\tau)$. Define $\xi_{\lambda}' (\phi)$ by \[ \xi'_{\lambda} (\phi) (\boldsymbol{z}) = \delta^{- 1 / 2} (a_{\lambda}) \, \mathcal{W}_{\tau} \Phi (a_{\lambda}) . \] \begin{proposition} The following diagram is commutative for any dominant weight $\lambda$ of $\hat{T}$. Let $s = s_i$ be a simple reflection, and let $\alpha = \alpha_i$ be the corresponding simple root. \begin{equation} \label{primehocd} \begin{CD} \mathcal{H} @>\xi'_{\lambda}>> \mathcal{O}( \hat{T})\\ @VV\times q^{-1}(1+t_i)V @VV(1-q^{-1}\boldsymbol{z}^\alpha)\partial_\alpha' V\\ \mathcal{H} @>\xi'_{\lambda}>> \mathcal{O}( \hat{T}) \end{CD} \end{equation} where the left vertical arrow is right multiplication by $q^{-1} (1 + t_i)$. Moreover \begin{equation} \label{xprimeev} \xi'_{\lambda} (t_{w^{- 1}}) = \delta^{-1/2}(a_\lambda)\,\mathcal{W}_{\tau} \Phi^{\tau}_w (a_{\lambda}) . \end{equation} \end{proposition} \medbreak \begin{proof} Let us consider both ways of traversing the commutative diagram (\ref{primehocd}) as applied to $t_{w^{- 1}}\in\mathcal{H}$. Then with $\tau = \tau_{\boldsymbol{z}}$ we have $\varrho_{\tau}^{- 1} t_{w^{- 1}} = \Phi^{\tau}_w$ and so we obtain (\ref{xprimeev}). Furthermore by Proposition~\ref{demazuremi} \[ \delta^{1 / 2} (a_{\lambda}) (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial'_{\alpha} \xi_{\lambda}' (t_{w^{- 1}}) = \mathcal{W}_{\tau} (\mathcal{A}_s^{\tau s} \Phi^{\tau s}_w + C_{\alpha} (\tau) \Phi^{\tau}_w) (a_{\lambda}) . \] By Lemma~\ref{bylemma}, \[ \mathcal{A}_s^{\tau s} \Phi^{\tau s}_w =\mathcal{A}_s^{\tau s} \varrho_{\tau s}^{- 1} t_{w^{- 1}} = \varrho_{\tau}^{- 1} \left( t_{w^{- 1}} \left( \frac{1}{q} t_i + (1 - q^{- 1}) \frac{\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} \right) \right) . \] We have \[ \left( 1 - \frac{1}{q} \right) \frac{\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} + C_{\alpha} (\tau) = \frac{1}{q} . \] Therefore \[ \mathcal{A}_s^{\tau s} \Phi^{\tau s}_w + C_{\alpha} (\tau) \Phi^{\tau}_w = \varrho_{\tau}^{- 1} \left( t_{w^{- 1}} q^{- 1} (t_i + 1) \right) . \] Applying $\mathcal{W}_{\tau}$ and evaluating at $a_{\lambda}$, we see that \[ (1 - q^{- 1} \boldsymbol{z}^{\alpha}) \partial'_{\alpha} \xi_{\lambda}' (t_{w^{- 1}}) = \xi_{\lambda}' (t_{w^{- 1}} q^{- 1} (t_i + 1)), \] which is the required commutativity. \end{proof} If $\lambda$ is dominant, we will define $\xi_\lambda$ as follows. By Theorem~\ref{qone}, if $\phi\in\mathcal{H}_v$ then $\theta\xi'_{-w_0\lambda}(\sigma\phi)$ is a polynomial in $q^{-1}$, so there exists an element $\xi_\lambda(\phi)$ of $\mathbb{C}[v,v^{-1}] \otimes \mathcal{O}(\hat{T})$ that corresponds to it. Thus $\xi_{\lambda} : \mathcal{H}_v \longrightarrow \mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T})$ is the map such that \[ \hbox{ev}\circ\xi_\lambda = \theta\circ\xi'_{-w_0\lambda}\circ\sigma\;.\] where $\hbox{ev}:\mathbb{C}[v,v^{-1}]\otimes\mathcal{O}(\hat{T})\to \mathcal{O}(\hat{T})$ is the evaluation map that sends $v$ to~$q^{-1}$. \begin{proposition} \label{commdomwt}Let $\lambda$ be a weight, and let $\alpha=\alpha_i$ be a simple root The following diagram is commutative: \begin{equation} \label{xicomdiag} \begin{CD} \mathcal{H}_v @>\xi_{\lambda}>> \mathbb{C}[v,v^{-1}]\otimes \mathcal{O}(\hat{T})\\ @VV (1+T_i)\times\cdot V @VV\mathfrak{D}_i V\\ \mathcal{H}_v @>\xi_{\lambda}>> \mathbb{C}[v,v^{-1}]\otimes \mathcal{O}(\hat{T}) \end{CD} \end{equation} where the left vertical arrow is left multiplication by $1 + T_i$. Moreover \begin{equation} \xi_{\lambda} (T_{(w w_0)^{- 1}}^{- 1}) =\boldsymbol{Y}_w (\lambda) . \label{xilambdat} \end{equation} \end{proposition} \begin{proof} We observe that (\ref{xilambdat}) follows from the definitions of $\xi_{\lambda}$ and $\boldsymbol{Y}_w (\lambda)$ and (\ref{xprimeev}). As for (\ref{xicomdiag}), consider the diagram \begin{equation*} \begin{CD} \mathcal{H}_v@>{\qquad\sigma\qquad}>>\mathcal{H}@>\xi'_{-w_0\lambda}>>\mathcal{O}(\hat T)@>{\qquad\theta\qquad}>>\mathcal{O}(\hat T)\\ @V(1+T_i)\times VV @V q^{-1}(1+t_i)VV @VV(1-q^{-1}\boldsymbol{z}^\alpha)\partial'_\alpha V @VV(1-q^{-1}\boldsymbol{z}^{-\alpha})\partial_\alpha V\\ \mathcal{H}_v@>{\qquad\sigma\qquad}>>\mathcal{H}@>\xi'_{-w_0\lambda}>>\mathcal{O}(\hat T)@>{\qquad\theta\qquad}>>\mathcal{O}(\hat T)\\ \end{CD} \end{equation*} The commutativity of the left square follows from (\ref{sigmat}). The commutativity of the middle square is (\ref{primehocd}). The commutativity of the right square is the fact that $\theta \partial_\alpha \theta = \partial_\alpha'$. The composition on the top is $\hbox{ev}\circ\xi_\lambda$ and the statement follows. \end{proof} The (extended) affine algebra $\tilde{\mathcal{H}}_v$ is generated by $\mathcal{H}_v$ and another commutative subalgebra $\zeta^{\Lambda}$ isomorphic to the weight lattice $\Lambda$. If $\lambda \in \Lambda$ let $\zeta^{\lambda}$ be the corresponding element of $\zeta^{\Lambda}$. To complete the presentation of $\tilde{\mathcal{H}}$ we have the relation \begin{equation} \label{bernstein} T_i \zeta^{\lambda} - \zeta^{s_i \lambda} T_i = \zeta^{\lambda} T_i - T_i \zeta^{s_i \lambda} = \left( \frac{v - 1}{1 - \zeta^{- \alpha_i}} \right) (\zeta^{\lambda} - \zeta^{s_i \lambda}) . \end{equation} \begin{lemma} Let $\lambda$ run through the dominant weights, and for each $\lambda$ let $w$ run through a set of coset representatives for $W / W_{\lambda}$ where $W_{\lambda}$ is the stabilizer of $\lambda$. Then $\boldsymbol{Y}_w (\lambda)$ runs through a basis of the $\mathbb{C}(v)$-vector space $\mathbb{C}(v) \otimes \mathcal{O}( \hat{T})$. Moreover \begin{equation} \label{xidsdecomp} \mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T}) = \bigoplus_{\text{$\lambda$ dominant}} \xi_{\lambda} (\mathcal{H}_v) . \end{equation} \end{lemma} \begin{proof} Every weight may be written as $w \lambda$ with $\lambda$ dominant. Here $\lambda$ is uniquely determined, and if there is more than one choice for $w$ we take the one of smallest length. We make a partial order on the weights by $w_1 \lambda_1 \preccurlyeq w_2 \lambda_2$ with either $\lambda_1 < \lambda_2$ in the usual partial order on weights, or $\lambda_1 = \lambda_2$ and $w_1 \leqslant w_2$ in the Bruhat order. Then $\boldsymbol{Y}_w (\lambda) =\boldsymbol{z}^{w \lambda}$ plus terms that are lower in the partial order, so these are a basis. The direct sum decomposition follows from~(\ref{xilambdat}). \end{proof} \begin{theorem} \label{cherednik} There is a right action of $\tilde{\mathcal{H}}_v$ on $\mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T})$ such that $\xi_{\lambda}$ is a homomorphism of right $\mathcal{H}_v$-modules. In this action \begin{equation} \label{cheredaction} \zeta^{\lambda} \cdot \boldsymbol{z}^{\mu} =\boldsymbol{z}^{\mu - \lambda}, \hspace{2em} (1 + T_i) \cdot \boldsymbol{z}^{\mu} =\mathfrak{D}_i (\boldsymbol{z}^{\mu}) = \left( \frac{1 - v\boldsymbol{z}^{- \alpha_i}}{1 -\boldsymbol{z}^{- \alpha_i}} \right) (\boldsymbol{z}^{\mu} -\boldsymbol{z}^{s_i \mu - \alpha_i}) . \end{equation} \end{theorem} This action of $\tilde{\mathcal{H}}_v$ appeared previously in Lusztig~\cite{LusztigK1}. The interpretation here in terms of Whittaker functions appears to be new. \medbreak \begin{proof} We make use of (\ref{xidsdecomp}). Each summand may be given a right $\mathcal{H}_v$-module structure such that $\xi_{\lambda}$ is a right $\mathcal{H}_v$-module homomorphism. Thus $\mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T})$ becomes an $\mathcal{H}_v$-module. By the commutativity of (\ref{xicomdiag}) \[ (1 + T_i) \cdot \xi_{\lambda} (\phi) = \xi_{\lambda} ((1 + T_i) \phi) =\mathfrak{D}_i \xi_{\lambda} (\phi), \] which gives the second identity in (\ref{cheredaction}). As for the first, we make this a definition and then must check compatibility with the relation (\ref{bernstein}) in the equivalent form \begin{equation} \label{equivbernstein} (T_i + 1) \zeta^{\lambda} - \zeta^{s_i \lambda} (T_i + 1) = \left( \frac{v - \zeta^{- \alpha_i}}{1 - \zeta^{- \alpha_i}} \right) (\zeta^{\lambda} - \zeta^{s_i \lambda}) . \end{equation} We will write $s = s_i$ and $\alpha = \alpha_i$. Applying the left-hand side of (\ref{equivbernstein}) to $\boldsymbol{z}^{\mu}$ gives \[ \left( \frac{1 - v\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} \right) (\boldsymbol{z}^{\mu - \lambda} -\boldsymbol{z}^{- \alpha + s \mu - s \lambda}) - \left( \frac{1 - v\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} \right) (\boldsymbol{z}^{\mu - s \lambda} -\boldsymbol{z}^{- \alpha + s \mu - s \lambda}) \] This equals \[ \left( \frac{1 - v\boldsymbol{z}^{- \alpha}}{1 -\boldsymbol{z}^{- \alpha}} \right) (\boldsymbol{z}^{\mu - \lambda} -\boldsymbol{z}^{\mu - s \lambda}) =\boldsymbol{z}^{\mu} \left( \frac{v -\boldsymbol{z}^{\alpha}}{1 -\boldsymbol{z}^{\alpha}} \right) (\boldsymbol{z}^{- \lambda} -\boldsymbol{z}^{- s \lambda}) . \] which equals the right-hand side of (\ref{equivbernstein}) using the action $\zeta^{\lambda} \cdot \boldsymbol{z}^{\mu} =\boldsymbol{z}^{\mu - \lambda}$, as desired. \end{proof} The maps $\xi_\lambda$ have two applications in this paper. The first is that they allow one to extend the action of $\mathcal{H}_v$ to all of $\widetilde{\mathcal{H}}_v$ as in the previous theorem. The second application is Theorem~\ref{amazing}. \begin{proof}[Proof of Theorem~\ref{amazing}] Since $l (w_0) = l (w) + l (w_0 w^{- 1})$ we have $T_{(w w_0)^{- 1}} T_w = T_{w_0^{- 1}}$ or $T_{(w w_0)^{- 1}}^{- 1} = T_w T_{w_0^{- 1}}^{- 1} .$ Using (\ref{xilambdat}) and the fact that $\xi_{\lambda}$ is $\mathcal{H}_v$-equivariant, \[ \boldsymbol{Y}_w (\lambda) = \xi_{\lambda} (T_{(w w_0)^{- 1}}^{- 1}) = T_w \cdot \xi_{\lambda} (T_{w_0^{- 1}}^{- 1}) . \] But using Proposition~\ref{correctionterms} and (\ref{xilambdat}) again, $\xi_{\lambda} (T_{w_0}^{- 1}) =\boldsymbol{Y}_1 (\lambda) =\boldsymbol{z}^{\lambda}$. \end{proof} \section{Complements} It is easy to see that $\partial_i \boldsymbol{z}^{- \rho} = 0$ for each Demazure operator $\partial_i$. As a consequence, $\mathfrak{D}_i \boldsymbol{z}^{- \rho} = 0$ and so in the action of Theorem~\ref{cherednik} we have \[ T_i \cdot \boldsymbol{z}^{- \rho} = -\boldsymbol{z}^{- \rho} . \] Hence this vector spans a one-dimensional $\mathcal{H}_v$-invariant subspace affording the sign representation of $\mathcal{H}_v$. \begin{theorem} The $\tilde{\mathcal{H}}_v$-module $\mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T})$ is isomorphic to the representation of $\tilde{\mathcal{H}}_v$ induced from the sign representation of $\mathcal{H}_v$. \end{theorem} \begin{proof} The module $\mathbb{C}[v,v^{-1}] \otimes \mathcal{O}( \hat{T})$ is generated by this vector $\boldsymbol{z}^{- \rho}$ since each coset of $\mathcal{H}_v$ has as a coset representation $\zeta^{\lambda}$ for a unique weight $\lambda$, which maps $\boldsymbol{z}^{- \rho}$ to a basis vector $\boldsymbol{z}^{- \lambda - \rho}$. The statement is thus clear. \end{proof} Finally, let us define certain elements $\boldsymbol{C}_w (\lambda)$ that resemble the $\boldsymbol{X}_w (\lambda)$ but may in some sense be more natural. We begin by recalling the Kazhdan-Lusztig basis $C_w'$ for the (finite) Hecke algebra $\mathcal{H}_v$. It is uniquely characterized by the following properties. First, $\overline{C_w}' = C'_w$, where $x \longmapsto \bar{x}$ is the Kazhdan-Lusztig involution that sends $q \rightarrow q^{- 1}$ and $T_w \rightarrow T_{w^{- 1}}^{- 1}$; and second, $v^{l (w) / 2} C'_w = \sum_{u \leqslant w} P_{u, w} (v) T_u$, where $P_{u, w} (v) \in \mathbb{Z} [v]$ of degree $\leqslant \frac{1}{2} (l (w) - l (u) - 1)$ for $y < w$ and $P_{w, w} = 1$. The $P_{u, w}$ are the Kazhdan-Lusztig polynomials. (Note that $P$ is used differently in other parts of our paper.) Let $\boldsymbol{C}_w (\lambda) = v^{l (w) / 2} C'_w \cdot \boldsymbol{z}^{\lambda}$, where the action is that of Theorem~\ref{cherednik}. \begin{proposition} Let $\lambda$ be a dominant weight and $w \in W$. We have $\boldsymbol{X}_w (\lambda) =\boldsymbol{C}_w (\lambda)$ if and only if the Schubert variety $X_w$ is rationally smooth in the sense of Kazhdan and Lusztig~{\cite{KL}}. \end{proposition} \begin{proof} Each Kazhdan-Lusztig $P_{u, w}$ with $u \leqslant w$ is a polynomial in $v$ with constant term equal to $1$, so $\boldsymbol{X}_w (\lambda) =\boldsymbol{C}_w (\lambda)$ if and only if $P_{u, w} = 1$ for all $u \leqslant w$. By Theorem~A2 of {\cite{KL}}, this is true if and only if $X_w$ is rationally smooth. \end{proof}
\section{Introduction} Our starting point is the following conjecture of Liebeck, Nikolov and Shalev \cite{lns2}. \begin{conj}\label{c: 2} There exists an absolute constant $c$ such that if $G$ is a finite simple group and $S$ is a subset of $G$ of size at least two, then $G$ is a product of $N$ conjugates of $S$ for some $N\leq c \log|G|/ \log |S|$. \end{conj} Note that we must have $N\ge\log|G|\big/\log|S|$ by order considerations, and so the bound above is best possible up to the value of the constant $c$. The conjecture is an extension of a deep (and widely applied) theorem of Liebeck and Shalev. Indeed, the main result of \cite{lieshal} states that the above conjecture holds when $S$ is a conjugacy class or, more generally, a normal subset (that is, a union of conjugacy classes) of $G$. In \cite{lns2} Conjecture~\ref{c: 2} is also proved for sets of bounded size. Somewhat earlier Liebeck, Nikolov and Shalev \cite{lns} posed the following (still unproved) weaker conjecture. \begin{conj}\label{c: 1} There exists an absolute constant $c$ such that if $G$ is a finite simple group and $H$ is any nontrivial subgroup of $G$, then $G$ is a product of $N$ conjugates of $H$ for some $N\leq c \log|G|/ \log |H|$. \end{conj} Conjecture \ref{c: 1} itself represents a dramatic generalization of a host of earlier work on product decompositions of finite simple groups, most of which prove Conjecture \ref{c: 1} for particular subgroups $H$. For instance, in \cite{liepy} it is proved that a finite simple group of Lie type in characteristic $p$ is a product of 25 Sylow $p$-subgroups (see also \cite{bnp} for a recent improvement from 25 to 5). Further positive evidence for Conjecture \ref{c: 1} is provided by \cite{lns3}, \cite{lub} and \cite{nikolov} (when $H$ is of type $SL_n$). Certain results of this type are essential to prove that finite simple groups can be made into expanders (see the announcement \cite{kln}). The main purpose of this note is to prove Conjecture \ref{c: 2} for finite simple groups of Lie type of bounded rank. Put another way, we prove a version of Conjecture \ref{c: 2} in which the constant $c$ depends on the rank of the group $G$. Our main result follows. \begin{thm}\label{t: gps} Fix a positive integer $r$. There exists a constant $c=c(r)$ such that if $G$ is a finite simple group of Lie type of rank $r$ and $S$ is a subset of $G$ of size at least two then $G$ is a product of $N$ conjugates of $S$ for some $N\leq c \log|G|/ \log |S|$. \end{thm} In \cite{lns2} a weaker bound of the form $N\leq\big(\log|G|/ \log |S|\big)^{c(r)}$ is obtained. Also, in \cite{lns}, Theorem~\ref{t: gps} is proved when $S$ is a maximal subgroup of $G$. As a byproduct of our proof we obtain two results of independent interest. In these results, and throughout the paper, we denote by $S^g$ the conjugate $g^{-1}Sg$ of a subset $S$ of a group $G$ by an element $g$ of $G$, and, given a positive integer $m$, we denote by $S^m$ the product $SS\dotsb S$ of $m$ copies of $S$. There should be no confusion between these two similar notations because the type of the exponent will always be given. \begin{thm}\label{t: 2} Fix a positive integer $r$. There exists a positive constant $\varepsilon=\varepsilon(r)$ such that if $G$ is a finite simple group of Lie type of rank $r$ and $S$ is a subset of $G$ then for some $g$ in $G$ we have $\left|S S^g\right|\ge|S|^{1+\varepsilon}$ or $S^3=G$. \end{thm} The next theorem is similar, but concerns only normal subsets, in which case we obtain absolute constants. \begin{thm}\label{t: normal subsets} There exists $\varepsilon>0$ and a positive integer $b$ such that if $G$ is a finite simple group and $S$ is a normal subset of $G$ then $\left|S^2\right|\ge|S|^{1+\varepsilon}$ or $S^b=G$. \end{thm} Theorem~\ref{t: normal subsets} relates to a result of Shalev \cite[Theorem 7.4]{shalev}, which we strengthen in Section~\ref{s: shalev}. Note that the theorem would not be true were we to consider sets that are not normal. For instance, take $S$ to be a maximal parabolic subgroup in $G=PSL_n(q)$ with index $\frac{q^n-1}{q-1}$. Clearly $S^b=S$ for all positive integers $b$; on the other hand, for any positive number $\varepsilon$, and any $g$ in $G$, we have $|S S^g|\leq |G|\leq |S|^{1+\varepsilon}$ once $n$ is large enough. We conclude that neither of the given options can hold in this more general situation. Theorems~\ref{t: 2} and \ref{t: normal subsets}, and the remarks of the previous paragraph, lead us to make the following conjecture. \begin{conj}\label{c: 3} There exists $\varepsilon>0$ and a positive integer $b$ such that if $S$ is a subset of a finite simple group $G$ then for some $g$ in $G$ we have $\left|S S^g\right|\ge|S|^{1+\varepsilon}$ or $G$ is the product of $b$ conjugates of $S$. \end{conj} Note that, by Theorems~\ref{t: gps}~and~\ref{t: 2}, Conjectures~\ref{c: 2},~\ref{c: 1}~and~\ref{c: 3} hold for all exceptional simple groups. Note too that all three conjectures could be phrased in terms of \emph{translates} of the set $S$, rather than conjugates. This follows from the simple fact that a product of translates of $S$ is equal to a translate of a product of conjugates of $S$. Similarly a product of conjugates of a translate of $S$ is equal to a translate of a product of conjugates of $S$, a fact which will be useful in its own right. It is possible that Conjecture~\ref{c: 3} actually holds with $b=3$. When $b=2$ counterexamples are given by large non-real conjugacy classes (see the final section of \cite{shalev} for some related issues). Further counterexamples are given by certain families of maximal subgroups (see for example \cite[Corollary 2]{lps}, which states that large enough simple unitary groups of odd dimension cannot be decomposed into the product of two proper subgroups). We derive Theorems~\ref{t: gps} and \ref{t: 2} as consequences of the recent Product theorem for finite simple groups, proved independently by Breuillard, Green and Tao \cite{bgt2}, and Pyber and Szab\'o \cite{ps2} (see Section~\ref{s: proof2}). Theorem~\ref{t: normal subsets} follows from a version of Conjecture~\ref{c: 2} for normal subsets due to Liebeck and Shalev \cite{lieshal} and an extension of Pl\"unnecke's theorem \cite[Theorem 6.27]{taovu} to normal subsets of nonabelian groups (see Section~\ref{s: doubling}). In the final section we use a result of Petridis \cite{petridis} to derive an analogue of the classical Doubling lemma, a special case of Pl\"unnecke's theorem. We refer to the new result as the Skew doubling lemma; it can be thought of as a nonabelian version of the classical Doubling lemma. The Skew doubling lemma is applied to prove that Conjecture~\ref{c: 2} implies Conjecture~\ref{c: 3}. In the other direction, a standard argument (similar to the proof of Corollary~\ref{c: polynomial}) shows that Conjecture~\ref{c: 3} implies that a simple group $G$ is a product of $\left(\log|G|/\log|S|\right)^c$ conjugates of $S$, a weaker version of Conjecture~\ref{c: 2}. \section{Proof of Theorem~\ref{t: 2}}\label{s: proof2} We begin with a result of Petridis \cite[Theorem 4.4]{petridis}, which extends work of Helfgott, Ruzsa and Tao \cite{helfgott3,Ru2009,Ru2010,tao}. It relates to the Doubling lemma for abelian groups, which we return to in Section~\ref{s: doubling}. \begin{lem}\label{l: petridis} Let $S$ be a finite subset of a group $G$. Suppose that there exist positive numbers $J$ and $K$ such that $|S^2|\leq J|S|$ and $|SgS|\leq K|S|$ for each $g$ in $S$. Then $|S^3| \leq J^7K|S|$. \end{lem} Suppose now that $G$ is a finite group, and let $\minclass(G)$ denote the size of the smallest nontrivial conjugacy class in $G$. Let $\minclass(S,G)$ denote the size of the smallest nontrivial conjugacy class in $G$ that intersects $S$, and let $\mindeg(G)$ denote the dimension of the smallest nontrivial complex irreducible representation of $G$. As observed in \cite{npy}, a result of Gowers \cite{gowers} implies the following. \begin{prop}\label{p: gowers} Let $G$ be a finite group and let $k=\mindeg(G)$. Take $S\subseteq G$ such that $|S| \geq \frac{|G|}{\sqrt[3]k}.$ Then $G=S^3$. \end{prop} Now let $G=G_r(q)$ be a simple group of Lie type of rank $r$ over $\Fq$, the finite field of order $q$. We need some facts about $G$. The first result can be deduced, for example, from \cite[Tables 5.1 and Theorem 5.2.2]{kl}. \begin{prop}\label{p: conjugacy size} We have $q^r\leq \minclass(G)<|G|\leq q^{8r^2}$. \end{prop} \begin{prop}\label{p: landseitz} Let $k=\mindeg(G)$. Then $|G|<k^{8r^2}$. \end{prop} \begin{proof} We use the lower bounds on projective representations given by Landazuri and Seitz \cite{landseitz}, allowing for the slight errors corrected in \cite[Table 5.3.A]{kl}. For $G\neq PSL_2(q)$, we see that $k\geq q$, and so the result follows from Proposition~\ref{p: conjugacy size}. Now suppose that $G=PSL_2(q)$; then $|G|<q^3$ and $r=1$. For $q\geq 5$ and $q\neq 9$, $k=\frac1{(2,q-1)}(q-1)$ and it is clear that $k^8>q^3$. When $q=4$ we have $k=2$ and the result follows; likewise when $q=9$ we have $k=3$ and the result follows. \end{proof} The next result was obtained independently in \cite{guralnick-kantor} and \cite{stein}. \begin{prop}\label{p: stein} Each finite simple group $G$ is $\frac32$-generated; that is, for any nontrivial element $g$ of $G$ there exists $h$ in $G$ such that $\langle g, h\rangle=G$. \end{prop} \begin{cor}\label{c: 32} Let $G$ be a finite simple group and let $S$ be a subset of $G$ of size at least two. Then some translate of $S$ generates $G$. \end{cor} \begin{proof} Let $u$ and $v$ be distinct elements of $S$. Since $G$ is $\frac32$-generated, there exists $x$ in $G$ such that $\langle vu^{-1}, x\rangle=G$. Therefore the translate $Su^{-1}x$, which contains $x$ and $vu^{-1}x$, generates $G$. \end{proof} The next result, the Product theorem, is our primary tool for proving Theorems~\ref{t: gps} and \ref{t: 2}. Versions of this result can be found in \cite{bgt2,ps2}. It was first proved by Helfgott for the groups $PSL_2(p)$ and $PSL_3(p)$ in \cite{helfgott2, helfgott3}. \begin{thm}\label{t: generating sets} Fix a positive integer $r$. There exists a positive constant $\eta=\eta(r)$ such that, for $G$ a finite simple group of Lie type of rank $r$ and $S$ a generating set of $G$, either $S^3=G$ or $|S^3|\geq |S|^{1+\eta}.$ \end{thm} We can now prove Theorem~\ref{t: 2}. \begin{proof}[Proof of Theorem~\ref{t: 2}] Given a positive integer $r$, let $\eta$ be the constant from Theorem~\ref{t: generating sets}. It suffices to prove Theorem~\ref{t: 2} for sets $S$ of size larger than some constant $L>1$ that depends only on $\eta$, because if $|S|<L$, and $S^3\neq G$, then, by the simplicity of $G$, there is an element $g$ of $G$ such that $|SS^g|\geq |S|+1$, and $|S|+1\geq |S|^{1+\delta}$, where $\delta=\log(L+1)/\log L - 1$. In particular, we assume that $|S|\geq 8^{\frac{2}{\eta}}$, and we define $\varepsilon = \tfrac{1}{16}\min\left\{\eta,\tfrac{1}{24r^2}\right\}$. Since $G$ is $\frac32$-generated, there exists an element $g$ of $G$ such that the set $T=S\cup \{g\}$ generates $G$. We can apply Theorem~\ref{t: generating sets} to $T$ to conclude that either $T^3=G$ or $|T^3|\geq |S|^{1+\eta}$. Now, $T^3$ is the union of the eight sets $SSS$, $SSg$, $SgS$, $gSS$, $Sgg$, $gSg$, $ggS$ and $\{ggg\}$. Suppose that $|T^3|\geq |S|^{1+\eta}$. By the pigeon-hole principle at least one of the eight sets is larger than $\frac18 |S|^{1+\eta}$. We assumed earlier that $|S|\geq 8^{\frac{2}{\eta}}$, from which it follows that $\frac18 |S|^{1+\eta} > |S|^{1+\frac{\eta}{2}}$. Therefore one of the first seven of the eight sets is larger than $|S|^{1+\frac{\eta}{2}}$. All of these seven sets except $SSS$ are equal to a translate of the product of one or two conjugates of $S$, so if any of these have size at least $|S|^{1+\frac{\eta}{2}}$ then $|SS^h|\geq |S|^{1+\frac{\eta}{2}}$ for some element $h$ of $G$. If, on the other hand, $|SSS|\geq|S|^{1+\frac{\eta}{2}}$, then Lemma~\ref{l: petridis} (with $J=K=|S|^{\frac{\eta}{16}}$) implies that there is an element $h$ of $S\cup \{1\}$ with $|SS^h|\geq |S|^{1+\frac{\eta}{16}}$. Thus in both cases there is an element $h$ with $|SS^h|\geq |S|^{1+\varepsilon}$. The remaining possibility is that $T^3=G$. If $S^3\neq G$ then Proposition~\ref{p: gowers} implies that $|S|\leq |G|/\sqrt[3]{k}$ where $k=\mindeg(G)$. But Proposition~\ref{p: landseitz} gives that $|S|\leq |G|^{1-\frac{1}{24r^2}}$, and this implies, in particular, that $|T^3|=|G|\geq |S|^{1+\frac{1}{24r^2}}$. The argument of the previous paragraph applies again, to give $|SS^h|\geq |S|^{1+\varepsilon}$ for some element $h$. \end{proof} Note that we can immediately deduce the following result of \cite{lns} (which we will use later). \begin{cor}\label{c: polynomial} Fix a positive integer $r$. There exists a constant $d$ such that if $G$ is a finite simple group of Lie type of rank $r$ and $S$ is a subset of $G$ of size at least two then $G$ is a product of $N$ conjugates of $S$ for some $N\leq 3 (\log|G|/ \log |S|)^d$. \end{cor} \begin{proof} Let $\varepsilon$ be the constant from Theorem~\ref{t: 2}, and define $d=\log_{1+\varepsilon}2$. Let $M$ be the integer part of $\log_{1+\varepsilon} \frac{\log|G|}{\log|S|}$. Theorem~\ref{t: 2} implies that $G$ is the product of $3\cdot 2^M$ conjugates of $S$, and \[ 3\cdot 2^M \leq 3\left(\frac{\log|G|}{\log|S|}\right)^{d}. \] \end{proof} The results in this section motivate a common generalisation of the Product theorem, and Conjecture~\ref{c: 3}, for groups of Lie type. \begin{conj}\label{c: 4} There exists $\varepsilon>0$ and a positive integer $b$ such that the following statement holds. For each integer $r$ there is a positive integer $c(r)$ such that if $G$ is a finite simple group of Lie type of rank $r$ and $S$ a generating set of $G$, then either $|SS^g|\ge|S|^{1+\varepsilon}$ for some $g\in S^{c(r)}$, or else $G$ is the product of $b$ conjugates $S^{g_1},\dots,S^{g_b}$, where $g_1,\dots,g_b\in S^{c(r)}$. \end{conj} It would be interesting to prove Conjecture~\ref{c: 3} in the case when $S$ is a subgroup of $G$. A rather general qualitative result in this direction was obtained by Bergman and Lenstra \cite{bergman-lenstra}. They show that if $H$ is a subgroup of a group $G$ satisfying $\big|H H^g\big|\leq K|H|$ for all $g$ in $G$, then $H$ is ``close to'' some normal subgroup $N$ of $G$, in the sense that $\big|H:H\cap N\big|$ and $\big|N:H\cap N\big|$ are both bounded in terms of $K$. \section{Proof of Theorem~\ref{t: gps}}\label{s: proof1} Given an element $g$ of a group $G$ we define \[ g^G=\{g^h\,:\, h\in G\}, \] and, for a subset $Z$ of $G$, \[ Z^G = \{Z^h\,:\, h\in G\}. \] We begin the proof of Theorem~\ref{t: gps} with a simple combinatorial lemma, which enables us to deal with ``small'' sets. \begin{lem}\label{l: get a big set} Let $S$ be a subset of a finite group $G$. There exist a positive integer $m$ and $m$ conjugates of $S$ such that their product $X$ satisfies $$|X|= |S|^m\geq \frac{\sqrt{\minclass(SS^{-1},G)}}{|S|}\geq \frac{\sqrt{\minclass(G)}}{|S|}.$$ \end{lem} \begin{proof} Define $X_1=S$ and, if possible, choose an element $g$ of $G$ such that $X_1^{-1}X_1\cap gSS^{-1}g^{-1}=\{1\}$. Define $X_2=X_1gSg^{-1}$. Notice that if $x_L, x_R\in X_1, s_L, s_R\in S$, and $x_Lgs_Lg^{-1}=x_Rgs_Rg^{-1}$, then $x_R^{-1}x_L=gs_Rs_L^{-1}g^{-1}$. Hence $x_R^{-1}x_L\in X_1^{-1}X_1\cap gSS^{-1}g^{-1}$, and so $x_L=x_R$ and $s_L=s_R$. It follows that $|X_2|=|X_1||S|$. Now repeat this process with $X_2$ replacing $X_1$, and so on. The process terminates with a set $X$ of size $|S|^m$, which is a product of $m$ conjugates of $S$, and such that $|X^{-1}X\cap gSS^{-1}g^{-1}|\geq 2$ for all $g$ in $G$. Let $T$ be a set of smallest possible size that intersects every conjugate of $Z=SS^{-1}$ nontrivially, and write $t=|T|$. Let $n=|G:N_G(Z)|$, the number of $G$-conjugates of $Z$. By the pigeonhole principle there exists an element $g$ of $Z$ that lies in at least $\frac{n}{t}$ different conjugates of $Z$. Let us count the set \[ \Omega= \big\{(g',Z') \in g^G\times Z^G \, \big| \, g'\in Z'\big\} \] in two different ways. First, since every conjugate of $g$ lies in the same number of conjugates of $Z$, we know that $|g^G| \frac{n}{t}\leq |\Omega|.$ On the other hand it is clear that $|\Omega|\leq n|Z|$. Putting these together we obtain that $|g^G|\frac{n}{t}\leq n|Z|$. Therefore $$t\geq \frac{|g^G|}{|Z|}\ge\frac{\minclass(SS^{-1},G)}{|S|^2}$$ and using $|X|^2\geq |X^{-1}X|\geq t$ our statement follows. \end{proof} \begin{remark} Lemma \ref{l: get a big set} and Proposition~\ref{p: conjugacy size} imply that if $G$ is a simple group of Lie type of rank $r$ and $S$ a subset of size less that $q^{r/4}$ then we have $\big|SS^g\big|=|S|^2$ for some $g$ in $G$. \end{remark} We are now ready to prove Theorem~\ref{t: gps}. \begin{proof}[Proof of Theorem~\ref{t: gps}] As observed above, a product of conjugates of a translate of $S$ is equal to the translate of a product of conjugates of $S$. By Corollary~\ref{c: 32}, a translate of $S$ generates $G$. Therefore we assume that $S$ generates $G$. Suppose that $|S| \geq \big|\minclass(G)\big|^{1/4}$; then $|G|<|S|^{32r}$ by Proposition~\ref{p: conjugacy size}. Now Corollary~\ref{c: polynomial} implies that $G$ is a product of fewer than $3(32r)^d$ conjugates of $S$. The theorem holds in this case with $c=3(32r)^d$. Suppose instead that $|S| < |\minclass(G)|^{1/4}$. By Lemma~\ref{l: get a big set} we can choose conjugates $S_1,\dots,S_m$ of $S$ such that the set $X=S_1\dotsb S_m$ satisfies $|X|=|S|^m$ and \[ |X|\geq \frac{\sqrt{|\minclass(G)|}}{|S|} \ge \big|\minclass(G)\big|^{1/4} \,. \] It follows from the first part of the proof that $G$ is a product of fewer than $c\log|G|/\log|X|$ conjugates of $X$. Therefore $G$ is a product of fewer than $mc\log|G|/\log|X|$ conjugates of $S$ and, since $\log|X|=m\log|S|$, the result follows. \end{proof} \section{Pl\"unnecke-Ruzsa estimates for nonabelian groups}\label{s: doubling} The following basic result in additive combinatorics is due to Pl\"unnecke \cite{plun, plun2} (see also \cite[Section~6.5]{taovu}). \begin{thm}\label{t: doubling} Let $A$ and $B$ be finite sets in an abelian group $G$ and suppose that $|A B|\leq K|A|$ where $K$ is a positive number. Then for any positive integer $m$ there exists a nonempty subset $X$ of $A$ such that \[ |X B^m|\leq K^m|X|. \] In particular, $|B^2|\leq K|B|$ implies that $|B^m|\leq K^m|B|$ for $m=1,2,\dotsc$. \end{thm} The last statement (``In particular\ldots'') is called the Doubling lemma; it does not hold for nonabelian groups, however, as we saw in Lemma~\ref{l: petridis}, there are useful analogues in this context due to Helfgott, Petridis, Ruzsa and Tao \cite{helfgott3,petridis,Ru2009,Ru2010,tao}. Petridis also proved the following lemma \cite[Proposition 2.1]{petridis}. \begin{lem}\label{l: petridis 2} Let $X$ and $B$ be finite sets in a group. Suppose that \[ \frac{|XB|}{|X|} \leq \frac{|ZB|}{|Z|} \] for all $Z\subseteq X$. Then, for all finite sets $C$, \[ |CXB|\leq \frac{|CX| |XB|}{|X|}. \] \end{lem} Using this lemma we can extend Pl\"unnecke's theorem to normal subsets of nonabelian groups. The statement and proof mimic \cite[Theorem 3.1]{petridis}, which is a stronger version of Theorem~\ref{t: doubling}. \begin{thm}\label{t: normal doubling} Let $A$ and $B$ be finite sets in a group $G$ with $B$ normal in $G$. Suppose that $|A B|\leq K|A|$ for some positive number $K$. Then there exists a nonempty subset $X$ of $A$ such that \[ |X B^m|\leq K^m|X| \] for $m=1,2,\dotsc$. In particular, $|B^2|\leq K|B|$ implies that $|B^m|\leq K^m|B|$ for $m=1,2,\dotsc$ \end{thm} \begin{proof} We proceed by induction on $m$. First choose $X\subseteq A$ such that \[ \frac{|X B|}{|X|} \leq \frac{|Z B|}{|Z|} \] for all $Z\subseteq A$. Then \[ |X B|\leq |X| \frac{|A B|}{|A|}\leq K|X|, \] so the result is true for $m=1$. Now suppose that $|X B^{m}|\leq K^{m}|X|$ for some positive integer $m$. Normality of $B$ implies that $|X B^{m+1}|=|B^{m} X B|$, and then Lemma~\ref{l: petridis 2} gives \[ |X B^{m+1}| = |B^{m} X B|\leq \frac{|B^mX||XB|}{|X|} \leq K^{m+1}|X|. \] This verifies the inductive step, and completes the proof of the theorem. \end{proof} Following an argument of Petridis (see the proof of \cite[Theorem 1.2]{petridis}) we observe that the Pl\"unnecke-Ruzsa estimates \cite[Corollary 6.29]{taovu} can also be generalised using Theorem~\ref{t: normal doubling}. \begin{cor} Suppose that $A$ and $B$ are subsets of a group $G$, with $B$ normal in $G$, and $|AB|\leq K|A|$. Then \[ |B^mB^{-n}|\leq K^{m+n}|A| \] for all positive integers $m$ and $n$. \end{cor} Theorem~\ref{t: normal doubling} suggests that certain techniques in additive combinatorics concerning subsets of abelian groups can be applied to normal subsets of nonabelian groups. The next example -- which is a consequence of Pl\"unnecke's theorem, and generalises \cite[Corollary 2.4]{Ru2009} -- supports this suggestion. \begin{thm}\label{t: plun} Let $A$ and $B$ be subsets of a group $G$ with $B$ normal in $G$, and suppose that $|AB^j|\leq K|A|$ for some positive integer $j$. If $m\geq j$ then \[ |B^m| \leq K^{\frac{m}{j}}|A|. \] \end{thm} \begin{proof}[Sketch of proof] We use the notation of \cite[Section~6.5]{taovu}. Construct the $m$-tuple of directed bipartite graphs $$(G_{A,B}, G_{A B, B}, \dots, G_{A B^{m-1}, B}).$$ This $m$-tuple is a Pl\"unnecke graph. Now Pl\"unnecke's theorem \cite[Theorem 6.27]{taovu} yields the result immediately. \end{proof} \section{Proof of Theorem~\ref{t: normal subsets}}\label{s: shalev} In this section we prove Theorem~\ref{t: normal subsets} and generalise some related results of Shalev. We will need the following theorem of Liebeck and Shalev \cite{lieshal}. \begin{thm}\label{t: normal} There exists an absolute positive constant $a$ such that, if $G$ is a finite simple group and $S$ is a nontrivial normal subset of $G$, then $G=S^m$, where $m\leq a\frac{\log|G|}{\log|S|}$. \end{thm} \begin{proof}[Proof of Theorem~\ref{t: normal subsets}] Let $a$ be the absolute constant from Theorem~\ref{t: normal}. Choose a positive integer $b$ larger than $2a$. Suppose first that $|S|\geq \sqrt{|G|}$. Then Theorem \ref{t: normal} implies that $G=S^m$ where \[ m\leq \frac{a\log|G|}{\log|S|} \leq 2a \leq b, \] and hence $S^b=G$. Now suppose that $|S|\leq \sqrt{|G|}$. Then \[ \frac{\log|S|}{a\log|G|} \geq \frac{\log|S|}{2a(\log|G|-\log|S|)}=\frac{\log|S|}{2a(\log(|G|/|S|)}. \] Theorem \ref{t: normal} implies, once again, that for some $m\leq\frac{a\log|G|}{\log|S|}$ we have $G=S^m$. Hence, applying Theorem~\ref{t: normal doubling} to the normal subset $S$, we see that \[ \frac{|S^2|}{|S|} \geq \left(\frac{|S^m|}{|S|}\right)^{\frac{1}{m}} \geq \left(\frac{|G|}{|S|}\right)^{\frac{\log|S|}{a\log|G|}} \geq \left(\frac{|G|}{|S|}\right)^{\frac{\log|S|}{2a(\log(|G|/|S|)}} =|S|^{\frac{1}{2a}}\geq |S|^\frac{1}{b}, \] and this completes the proof. \end{proof} The next result is a strengthening of \cite[Theorem 7.4]{shalev}. \begin{prop}\label{p: shalev} For every $\delta>0$ there exists $\varepsilon>0$ such that for any finite simple group $G$ and subsets $A$ and $B$ of $G$ with $B$ normal in $G$ and $|A|\leq |G|^{1-\delta}$ we have \[ |AB|\geq |A||B|^{\varepsilon}. \] \end{prop} \begin{proof} We assume that $A$ is nonempty and $B$ is nontrivial, otherwise the result is immediate. By Theorem~\ref{t: normal}, $G=B^m$, where $m\leq a\frac{\log|G|}{\log|B|}$. Let $K=|AB|/|A|$. Then, by Theorem~\ref{t: normal doubling}, there is a nonempty subset $X$ of $A$ such that $|XB^m|\leq K^m|X|$. It follows that \[ |G|=|B^m|=|XB^m| \leq K^m|X|\leq K^m|A|. \] Since $|A|\leq |G|^{1-\delta}$ and $m\leq a\frac{\log|G|}{\log|B|}$ we can rearrange this inequality to give \[ |G|^\delta \leq K^{a\frac{\log|G|}{\log|B|}}. \] This is equivalent to $|B|^\frac{\delta}{a}\leq K$, which, with $\varepsilon=\frac{\delta}{a}$, is the required result. \end{proof} Proposition~\ref{p: shalev} constitutes the expansion result for $B^2$ that was partially proven in \cite[Proposition 10.4]{shalev}. Furthermore it goes some way towards a proof of \cite[Conjecture 10.3]{shalev} although what remains is the more difficult part of the conjecture. We can strengthen \cite[Proposition 10.4]{shalev} in a different direction as follows. \begin{prop}\label{p: shalev 2} For every $\delta>0$ and positive integer $r$ there exists $\varepsilon>0$ such that for any finite simple group $G$ of Lie type of rank $r$ and any set $S\subseteq G$ such that $|S|\leq |G|^{1-\delta}$, there exists $g$ in $G$ such that $$|SS^g|\geq |S|^{1+\varepsilon}.$$ \end{prop} \begin{proof} Given $\delta>0$ and a positive integer $r$, let $\varepsilon$ be the positive constant from Theorem~\ref{t: 2}. Now choose any subset $S$ of $G$ such that $|S|\leq |G|^{1-\delta}$. According to Theorem~\ref{t: 2}, either $|S S^g|\geq |S|^{1+\varepsilon}$ or else $S^3=G$. In the former case the result is proven. In the latter case we apply Lemma~\ref{l: petridis} with $J=K=(|S^3|/|S|)^{1/10}$ to deduce the existence of an element $g$ of $G$ with $|SgS|> K|S|$. Then, using $S^3=G$ and $|G|\geq |S|^{1+\delta}$, it follows that \[ |SgS| > \left(\frac{|S^3|}{|S|}\right)^{\frac{1}{10}}|S|\geq |S|^{1+\frac{\delta}{10}}. \] Provided that $\varepsilon$ is chosen to be smaller than $\tfrac{\delta}{10}$, the inequality $|S S^g|\geq |S|^{1+\varepsilon}$ is again satisfied. \end{proof} \section{The Skew doubling lemma}\label{s: skew doubling} The next result is another analogue of the Doubling lemma for nonabelian groups, which we call the {\it Skew doubling lemma}. \begin{lem}[Skew doubling lemma]\label{l: skew doubling} If $S$ is a finite subset of a group $G$ such that, for some positive number $K$, $|S S^g|\leq K|S|$ for every conjugate $S^g$ of $S$, then \[ |S_1\dotsb S_m|\leq K^{14(m-1)}|S| \] for $m=1,2,\dotsc$, where each of $S_1,\dots,S_m$ is any conjugate of either $S$ or $S^{-1}$. \end{lem} To prove Lemma~\ref{l: skew doubling} we will use Lemma~\ref{l: petridis} and the following result, Ruzsa's triangle inequality \cite{ruzsa} (see also \cite[Section~2.3]{taovu}). \begin{lem}\label{l: ruzsa} Let $U$, $V$ and $W$ be finite subsets of a group $G$. Then \[ \frac{|VW^{-1}|}{|U|} \leq \frac{|UV^{-1}|}{|U|} \frac{|UW^{-1}|}{|U|}. \] \end{lem} First we prove a special case of Lemma~\ref{l: skew doubling}. \begin{lem}\label{l: tao} Let $S$ be a finite subset of a group $G$. Suppose that $K$ is a positive number such that $|SS^g| \leq K|S|$ for each $g$ in $G$. Then $|S_1S_2S_3| \leq K^{14}|S|$, where each of $S_1$, $S_2$ and $S_3$ is any conjugate of either $S$ or $S^{-1}$. \end{lem} \begin{proof} Choose elements $a$ and $b$ of $G$. We can apply Lemma \ref{l: petridis} with $J=K$ to obtain \[ |S^3|\leq K^8|S|. \] Using this inequality and Lemma~\ref{l: ruzsa} (with $U=S^{-1}$, $V=SS$ and $W=S$) we obtain \[ \frac{|SSS^{-1}|}{|S|}\leq \frac{|S^{-1}S^{-1}S^{-1}|}{|S|}\frac{|S^{-1}S^{-1}|}{|S|}=\frac{|S^3|}{|S|}\frac{|S^2|}{|S|}\leq K^{9}. \] Using this inequality and Lemma~\ref{l: ruzsa} (with $U=S$, $V=S^{-1}$ and $W=SS^{-1}$) we obtain \[ \frac{|S^{-1}SS^{-1}|}{|S|}\leq \frac{|SS|}{|S|}\frac{|SSS^{-1}|}{|S|}\leq K^{10}. \] Using this inequality and Lemma~\ref{l: ruzsa} (with $U=S^{-1}$, $V=SS^{-1}$ and $W=Sa$) we obtain \[ \frac{|SS^{-1}a^{-1}S^{-1}|}{|S|}\leq \frac{|S^{-1}SS^{-1}|}{|S|}\frac{|S^{-1}a^{-1}S^{-1}|}{|S|}\leq K^{11}. \] Using this inequality and Lemma~\ref{l: ruzsa} (with $U=S$, $V=SaS$ and $W=S^{-1}b^{-1}$) we obtain \begin{equation}\label{e1} \frac{|SaSbS|}{|S|}\leq \frac{|SS^{-1}a^{-1}S^{-1}|}{|S|}\frac{|SbS|}{|S|}\leq K^{12}. \end{equation} Using this inequality and Lemma~\ref{l: ruzsa} (with $U=S$, $V=S^{-1}$ and $W=S^{-1}b^{-1}S^{-1}a^{-1}$) we obtain \begin{equation}\label{e2} \frac{|S^{-1}aSbS|}{|S|}\leq \frac{|SS|}{|S|}\frac{|SaSbS|}{|S|}\leq K^{13}. \end{equation} Finally, using this inequality and Lemma~\ref{l: ruzsa} (with $U=S^{-1}$, $V=S^{-1}aSb$ and $W=S$) we obtain \begin{equation}\label{e3} \frac{|S^{-1}aSbS^{-1}|}{|S|}\leq \frac{|S^{-1}b^{-1}S^{-1}a^{-1}S|}{|S^{-1}|}\frac{|S^{-1}S^{-1}|}{|S^{-1}|} = \frac{|S^{-1}aSbS|}{|S|}\frac{|SS|}{|S|} \leq K^{14}. \end{equation} Equations \eqref{e1}, \eqref{e2} and \eqref{e3} imply that, given any conjugates $S_1$, $S_2$ and $S_3$ of either $S$ or $S^{-1}$, we have $|S_1S_2S_3|/|S| \leq K^{14}$, as required. \end{proof} We need the following proposition. \begin{prop}\label{p: top} If $A$ and $B$ are finite subsets of a group $G$ such that, for some positive number $K$, $|BB^g|\leq K|B|$ for every conjugate $B^g$ of $B$, then \[ |AB_1B_2|\leq K^{14}|AB_3|, \] where each of $B_1$, $B_2$ and $B_3$ is any conjugate of $B$ or $B^{-1}$. \end{prop} \begin{proof} By Lemma~\ref{l: tao} we have \[ \frac{|B_3^{-1}B_1B_2|}{|B_3|} \leq K^{14}, \] where each of $B_1$, $B_2$ and $B_3$ is any conjugate of $B$ or $B^{-1}$. Applying Lemma~\ref{l: ruzsa} with $U=B_3^{-1}$, $V=A$ and $W=B_2^{-1}B_1^{-1}$ we obtain \[ \frac{|AB_1B_2|}{|AB_3|}=\frac{|AB_1B_2|}{|B_3^{-1}A^{-1}|}\leq \frac{|B_3^{-1}B_1B_2|}{|B_3|}\leq K^{14}, \] as required. \end{proof} We can finally prove Lemma~\ref{l: skew doubling}. \begin{proof}[Proof of the Skew doubling lemma] The result holds trivially when $m=1$ and $m=2$. Suppose that $m\geq 3$. Apply Proposition~\ref{p: top} with $B=S$, $A=S_1\dotsb S_{n-2}$, $B_1=B_3=S_{n-1}$ and $B_2=S_{n}$ to see that \[ \frac{|S_1\dotsb S_{n}|}{|S_1\dotsb S_{n-1}|} \leq K^{14} \] for $n=3,4,\dots,m$. It follows that \begin{align*} \frac{|S_1\dotsb S_{m}|}{|S|} &= \left(\frac{|S_1\dotsb S_{m}|}{|S_1\dotsb S_{m-1}|}\right)\left(\frac{|S_1\dotsb S_{m-1}|}{|S_1\dotsb S_{m-2}|}\right)\dotsb \left(\frac{|S_1S_2S_3|}{|S_1S_2|}\right)\left(\frac{|S_1S_2|}{|S_1|}\right) \\ &\leq (K^{14})^{m-2}K \\ & \leq K^{14(m-1)}, \end{align*} as required. \end{proof} Using the Skew doubling lemma we can derive Conjecture~\ref{c: 3} from Conjecture~\ref{c: 2}. The proof is similar to the proof of Theorem~\ref{t: normal subsets}. \begin{proof}[Proof that Conjecture \ref{c: 2} implies Conjecture \ref{c: 3}] Let $c$ be the absolute constant from Conjecture~\ref{c: 2}. We define $b$ to be a positive integer greater than $2c$, and $\varepsilon=1/(28c)$. Suppose first that $|S|\geq \sqrt{|G|}$. Then Conjecture~\ref{c: 2} implies that $G=S_1\dotsb S_N$, for conjugates $S_1,\dots,S_N$ of $S$, where \[ N\leq \frac{c\log|G|}{\log|S|} \leq 2c < b, \] and hence $G$ is certainly the product of $b$ conjugates of $S$. Now suppose that $|S|\leq \sqrt{|G|}$. Then \[ \frac{\log|G|-\log|S|}{c\log|G|-\log|S|} \geq \frac{\log|G|-\log|S|}{c\log|G|} \geq \frac{1}{2c}. \] In particular observe that \[c\log|G|-\log|S| \leq 2c(\log|G|-\log|S|) = 2c\log(|G|/|S|).\] Conjecture~\ref{c: 2} implies, once again, that for some $N\leq\frac{c\log|G|}{\log|S|}$ we have $G=S_1\dotsb S_N$, for conjugates $S_1,\dots,S_N$ of $S$. Using the Skew doubling lemma, Lemma~\ref{l: skew doubling}, we see that there is an element $g$ of $G$ for which \[ \frac{|SS^g|}{|S|} \geq \left(\frac{|S_1\dots S_N|}{|S|}\right)^{\frac{1}{14(N-1)}} \geq \left(\frac{|G|}{|S|}\right)^{\frac{\log|S|}{14(c\log|G|-\log|S|)}} \geq \left(\frac{|G|}{|S|}\right)^{\frac{\log|S|}{28c(\log(|G|/|S|))}} \geq |S|^{\frac{1}{28c}}, \] and this completes the proof. \end{proof} \bibliographystyle{plain}
\section{Introduction} \label{sec:intro} An infinitesimal deformation problem asks, given some structure $\xi$ over a scheme $S$ and an infinitesimal extension $S'$ of $S$, whether it is possible to extend $\xi$ to a structure of the same type over $S'$. In any reasonable example, $\xi$ may be viewed as a section of a stack $X$, and the deformation problem can be viewed as one of extending a map: \begin{equation} \label{eqn:20} \xymatrix{ S \ar[r]^\xi \ar[d] & X \\ S' \ar@{-->}[ur]_{\xi'} . } \end{equation} Such a problem may be ``solved'' by describing an obstruction group, $\Obs(S,J)$, depending in a functorial manner on $S$ and $J$, but not on the specific extension $S'$, and an obstruction class $\omega \in \Obs(S,J)$ that does depend on all of the data and is also functorial in $S$ and $J$. Should $\omega$ vanish, the dashed arrows completing Diagram~\eqref{eqn:20} form a torsor under a second group, $\Def(S, J)$, again depending only on $S$ and $J$. While there is a canonical description of the group $\Def(S,J)$, there may be many different different choices for $\Obs(S,J)$. It was shown by Fantechi and Manetti~\cite{FM} that there is at least a \emph{minimal} example of an obstruction theory, but the minimal example is rarely the most natural or convenient to use; moreover, it is necessary to make use of other obstruction theories to define the virtual fundamental class~\cite{LT,BF} for use in virtual enumerative geometry. Despite the utility of obstruction theories in the construction of the virtual fundamental class and in Artin's criteria for the algebraicity of stacks, it remains unclear just what an obstruction theory \emph{is}. There are at least three different definitions in use, all of which are likely inequivalent. The purpose of this paper is to propose a definition of an obstruction theory that, at least in the author's opinion, captures all of the properties one would want from an obstruction theory without sacrificing any of the generality of the definitions already available. We will also compare this definition with the others. The first of these is Artin's, which was give to streamline his criteria for algebraicity of stacks \cite[Section~(2.6)]{versal}. Fantechi and Manetti generalized Artin's definition to a relative setting~\cite{FM}.\footnote{In fact, Fantechi and Manetti gave a definition that applies even to functions whose ``tangent spaces'' are not even required to be vector spaces! The definition given here can likely be generalized that far as well, but I have not pursued it since I do not know of an application.} Obstruction theories usually arise cohomologically, and from this point of view the definition of Artin--Fantechi--Manetti definition is phrased in terms of the cohomology groups of an inexplicit cohomology theory. The finer structure of a cohomology theory itself proved essential in the definition of the virtual fundamental class. To capture this finer structure, two roughly dual definitions of a perfect obstruction theory are available. They were introduced at about the same time by Li--Tian \cite{LT} and Behrend--Fantechi \cite{BF}. The Li--Tian definition is given in deformation-theoretic terms very close in spirit to Artin's original definition; the essential refinement \cite[Definition~1.3]{LT} is the introduction of a perfect complex whose cohomology groups are the tangent and obstruction spaces. Behrend and Fantechi also work crucially with perfect complexes but they work dually, relying elegantly on the existence of a cotangent complex $\bL_X$. From their point of view, an obstruction theory is simply a morphism of complexes $\bE_\bullet \rightarrow \bL_{X}$ whose cone is concentrated in degrees $\leq -2$ \cite[Definition~4.4]{BF}. The deformation-theoretic content of an obstruction theory is neatly encapsulated in its relation to the cotangent complex, via the cotangent complex's celebrated role in deformation theory \cite{Ill1,Ols-cc,Ols-def}. An obstruction theory in the sense of Behrend--Fantechi induces one in the sense of Li--Tian and these were shown to give equivalent virtual fundamental classes~\cite{KKP}. A Li--Tian obstruction theory induces an obstruction theory in the sense of Artin--Fantechi--Manetti: \begin{equation*} \begin{pmatrix} \text{Behrend} \\ \text{Fantechi} \end{pmatrix} \Rightarrow \begin{pmatrix} \text{Li} \\ \text{Tian} \end{pmatrix} \Rightarrow \begin{pmatrix} \text{Artin} \\ \text{Fantechi} \\ \text{Manetti} \end{pmatrix} . \end{equation*} The definition presented here lies between those of Behrend--Fantechi and Li--Tian and our main result shows that, with a sufficient number of adjectives, it is equivalent to the former: \begin{theorem*} An obstruction theory, in the sense presented here, is induced from a perfect obstruction theory in the sense of Behrend--Fantechi if and only if it is \emph{locally of finite presentation} and \emph{quasi-perfect} (see Definition~\ref{def:lfp}). \end{theorem*} \noindent The theorem is proved as Proposition~\ref{prop:perf} in Section~\ref{sec:perf}. While we believe our definition is inequivalent to Li's and Tian's, it is nonetheless close to it in spirit, and the theorem therefore gives a means by which to show that Li--Tian obstruction theories are induced from Behrend--Fantechi obstruction theories. This application was the original motivation for this paper. The reason for the disconnect between the definitions of Behrend--Fantechi and Li--Tian comes down to the cotangent complex. It is a powerful tool for the construction and manipulation of obstruction theories (see, for example, \cite{KKP,Cos,Man}), but it can also be a liability when it must be used in situations not foreseen by Illusie in~\cite{Ill1,Ill2}. One such situation arose in the theories of relative and degenerate stable maps. These theories give a good definition of stable maps into mildly singular schemes by restricting the class of deformations with the ``pre-deformability condition'' \cite{Li1}. This condition is not open in the space of all deformations, and it therefore obscures the relationship between pre-deformable deformations and the cotangent complex of their moduli space. Because the definition of Li and Tian is more closely related to deformation theory, it becomes easier (but by no means easy!) to define an obstruction theory in the sense of Li--Tian. This was accomplished by Li in \cite{Li2}. In order to compare Li's obstruction theory to others defined more easily using the Behrend--Fantechi formalism in \cite{AMW,CMW} (and to a lesser extent, \cite{ACW}), it was necessary to find a definition that incorporates the strengths of both approaches. Another situation where reliance on the cotangent complex proved a hindrance is the theory of logarithmic stable maps, which was held back for several years by the lack of the logarithmic cotangent complex needed to imitate the Behrend--Fantechi construction of the obstruction theory for stable maps. The situation was eventually remedied when Olsson defined the logarithmic cotangent complex, and there are now two different moduli spaces of logarithmic stable maps \cite{Kim,Chen,AC,GS}. However, the construction of the obstruction theory for stable maps in Section~\ref{sec:stable-maps} makes no reference to the cotangent complex, and therefore may be adapted virtually unchanged to the logarithmic setting as well. The definition given here is based on that of a \emph{left exact, additively cofibered category}, as given by Grothendieck~\cite{ccct} during his early investigation of the cotangent complex. Additively cofibered categories are an efficient way of describing \emph{additive} $2$-functors from an additive category into the $2$-category of categories. Infinitesimal deformations naturally form left exact additively cofibered categories, making this a convenient setting in which to treat infinitesimal deformation problems. Modulo the many functoriality properties that an obstruction theory must satisfy, our definition is the following. Consider the relative version of Diagram~\eqref{eqn:20}: \begin{equation} \label{eqn:25} \xymatrix{ S \ar[r] \ar[d] & X \ar[d]^h \\ S' \ar@{-->}[ur] \ar[r] & Y . } \end{equation} Suppose that $S = \Spec A$ and $S' = \Spec A'$ are affine and let $J$ be the kernel of the homomorphism $A' \rightarrow A$. We write $\Exal_X(A, J)$ for the category of commutative diagrams~\eqref{eqn:25}, including the dashed arrow, in which the ideal of $S$ in $S'$ is $J$. Let $h^\ast \Exal_Y(A,J)$ be the category of such diagrams, excluding the dashed arrow. An obstruction theory is an additively cofibered category $\sE(A, -)$ on the category of $A$-modules such that the diagrams \begin{equation} \label{eqn:19} \xymatrix{ \Exal_X(A,J) \ar[r] \ar[d] & e(A,J) \ar[d] \\ h^\ast \Exal_Y(A,J) \ar[r] & \sE(A,J) } \end{equation} are cartesian, with $e(A,J)$ denoting the zero object of $\sE(A,J)$. We can understand the diagram as saying that for any $\xi \in h^\ast \Exal_Y(A,J)$---that is, any commutative diagram of solid lines~\eqref{eqn:25}---there is an associated obstruction $\omega \in \sE(A,J)$ obtained as the image of $\xi$ under the lower horizontal arrow of~\eqref{eqn:19}. Furthermore, the isomorphisms between $\omega$ and $e(A,J)$ are in bijection, in a specified way, with the lifts of $\xi$ to $\Exal_X(A,J)$---the dashed arrows completing Diagram~\eqref{eqn:18}. In particular, $\omega$ is \emph{isomorphic} to zero if and only if a solution to the deformation problem exists. \subsection{Summary of the paper} The context for the definitions presented here is that of \emph{homogeneous stacks}. This is generalization of the homogeneous groupoids of Rim \cite{Rim}. We define homogeneous stacks in Section~\ref{sec:schl} and explain some of their basic properties. The definition of an obstruction theory is given in Section~\ref{sec:def}, after some background material on additively cofibered categories. The main results of this paper are contained in Section~\ref{sec:rep}, where we relate our definition to the cotangent complex in~\ref{sec:cc} and to the definition of Behrend--Fantechi in Corollaries~\ref{cor:BF1} and~\ref{cor:BF2}. Proposition~\ref{prop:perf} gives criterion for an obstruction theory to be representable by a complex of perfect amplitude in degress $[-1,0]$, which is useful in \cite{CMW}. In Section~\ref{sec:obgrp} we explain how to recover an obstruction sheaf from an obstruction theory as defined here. This permits us to relate our definition to Artin's and to Li's and Tian's. In Section~\ref{sec:compat}, we define compatible obstruction theories and relate this notion to the one defined by Manolache~\cite{Man}. Finally, in Section~\ref{sec:ex}, we give some familiar examples of obstruction theories, viewed from the perspective presented here. Perhaps the most interesting example of an obstruction theory is the one for degenerate stable maps that motivated this paper. It is too long to give a thorough treatment here, and will be described in a separate paper \cite{predef}. Many details of a related obstruction theory are given in \cite{CMW} as well. The appendix contains some background material on bilinear maps of vector bundle stacks. \subsection{Notational conventions} As in \cite{versal}, we have tried here to work systematically with covariant functors and cofibered categories over the category of rings instead of contravariant functors and fibered categories over the category of schemes. In company with \cite{versal} we have not been entirely successful in this respect. The reason we have used the algebraic language instead of the geometric one is our systematic reliance on the category $\ComRngMod$ of pairs $(A,I)$ where $A$ is a commutative ring and $I$ is an $A$-module. The opposite of this category, which might most naturally be described as the category of paris $(S, F)$ where $S$ is an affine scheme and $F$ is an object of the \emph{opposite category} of quasi-coherent sheaves on $S$, is a syntactic atrocity. Furthermore, we will find various categories that have are contravariant with respect to $\ComRng$, but whose natural extensions to schemes are only covariant with respect to \emph{affine} morphisms. Focusing only on affine schemes allows us to avoid this technicality and shorten a number of statements. All rings and algebras in this paper are commutative and unital, but we often emphasize the commutativity anyway. If $A$ is a commutative ring and $J$ is an $A$-module then we write $\tJ$ for the associated sheaf on the small \'etale site of $A$, which is denoted $\et(A)$. Likewise, $\tA$ denotes the structure sheaf on $\et(A)$. We employ the same notation for complexes of $A$-modules. In this paper, the distinction between chain complexes and cochain complexes is purely a matter of notation: every complex $\bE$ is simultaneously a chain complex $\bE_\bullet$ and a cochain complex $\bE^\bullet$ with $\bE_i = \bE^{-i}$. When referring to the degrees of a complex, chain or cochain, we always use cohomological degrees. Therefore to say that a chain complex $\bE_\bullet$ is concentrated in degrees $[-1,0]$ means that $\bE_i = 0$ for $i \not= 0,1$, or, equivalently, that $\bE^i = 0$ for $i \not= -1, 0$. Finally, note that we often use the following convention: if $\sE$ is cofibered category over $X\hAlgMod$ (defined in Section~\ref{sec:mod}) then we write $\usE(A,I)$ for the cofibered category over the category $A\hEtAlg$ of \'etale $A$-algebras. We use the same notation for the corresponding fibered category on the small \'etale site of $\Spec A$. We also extend this notation to apply to sheaves. \section{Homogeneous stacks} \label{sec:schl} We introduce the stacks for which we will describe obstruction theories later in the paper. Although we are ultimately interested in obstruction theories for moduli spaces that are Deligne--Mumford stacks, these obstruction theories can often be constructed by including these stacks in much larger stacks that cannot be algebraic; moreover, one application of obstruction theories is in demonstrating that stacks are algebraic, so we must have a definition that applies to stacks that are not known a priori to be algebraic. The stacks we work with will be called \emph{homogeneous}, expanding the terminology employed by \cite{Rim} to a more general setting. Let $\ComRng$ be the category of commutative rings. A cofibered category over $\ComRng$ is called a \emph{stack} if it satisfies descent in the \'etale topology. Note that our stacks are cofibered over commutative rings instead of fibered over schemes, as is more conventional, and the fibers of our stacks are \emph{not} required to be groupoids. We will say that a cofibered category $F \rightarrow \ComRng$ is \emph{homogeneous} or \emph{satisfies Schlessinger's conditions} if, whenever \begin{equation*} \xymatrix{ A' \ar[r] \ar[d] & B' \ar[d] \\ A \ar[r] & B } \end{equation*} is a cartesian diagram in $\ComRng$ and the kernel of $B' \rightarrow B$ is nilpotent, the natural map \begin{equation*} F(A') \rightarrow F(A) \fp_{F(B)} F(B') \end{equation*} (which is defined up to unique isomorphism) is an equivalence of categories. Following \cite[Definition~2.5]{Rim}, this can be put somewhat more intrinsically: \renewcommand{\labelenumi}{(\roman{enumi})} \begin{enumerate} \item if $\eta' \rightarrow \eta$ and $\xi \rightarrow \eta$ are morphism in $F$ such that $p(\eta') \rightarrow p(\eta)$ is an infinitesimal extension, then a fiber product $\xi \fp_\eta \eta'$ exists, and \item the projection $p : F \rightarrow \ComRng$ preserves such fiber products. \end{enumerate} Rim's definition of homogeneity \cite{Rim} efficiently implies all of Schlessinger's criteria \cite[Theorem~2.11]{Schless} simultaneously. The study of homogeneous groupoids has been taken up again more recently by Osserman (see \cite[Definition~2.7]{Oss}, where ``deformation stacks'' are analogous to homogeneous stacks). If $X$ is a cofibered category over $\ComRng$, we also write $X\hAlg$ for the underlying category of $X$. (If $X = \Spec A$ is an affine scheme then $X\hAlg$ is the category of $A$-algebras.) Accordingly, we refer to the objects of $X\hAlg$ as $X$-algebras. If $B$ is a commutative ring then we also refer to the objects of $X(B)$ as $X$-algebra structures on $B$. The following proposition is well-known. \begin{proposition} Any algebraic stack is homogeneous. \end{proposition} \begin{proof} Let $X$ be an algebraic stack, suppose that $A' \rightarrow A$ is an infinitesimal extension of $X$-algebras and $B \rightarrow A$ is a homomorphism of $A$-algebras. We wish to construct a fiber product of $X$-algebras $B' = B \fp_A A'$. Let $\oA$, $\oA'$, and $\oB$ denote the underlying commutative rings of $A$, $A'$, and $B$, respectively. Let $\oB'$ be the fiber product of $\oA'$ and $\oB$ over $\oA$. The problem is to define a canonical $X$-algebra structure on $\oB'$. Since $X$ is a stack in the \'etale topology, this is an \'etale-local problem in $\oB'$, and since the \'etale site of $\oB'$ is the same as the \'etale site of $\oB$, it is also an \'etale local problem in $\oB$. We can therefore assume that the $X$-algebra structure on $\oB$ is induced from a $C$-algebra structure on $\oB$ for some \emph{smooth} $X$-algebra $C$. By the infinitesimal criterion for smoothness, we can extend the induced $C$-algebra structure on $\oA$ to a $C$-algebra structure on $\oA'$. From this lift we obtain a \emph{canonical} $C$-algebra structure on $\oB'$ because $C$ is representable by a commutative ring. By composition, this induces an $X$-algebra structure on $\oB'$. We must argue that this construction does not depend on the choice of the smooth $X$-algebra $C$. Indeed, suppose we had two $X$-algebra structures $B'$ and $B''$ on $\oB'$ that induce the given $X$-algebra structures on $\oA$, $\oA'$, and $\oB$. Then we obtain an $X \times X$-algebra structure on $\oB'$ such that the induced $X \times X$-algebra structures over $\oA$, $\oA'$, and $\oB$ are compatible induced from the diagonal map $X \rightarrow X \times X$. But $X$ is an algebraic stack so the diagonal is representable by algebraic spaces. Therefore the proposition is reduced to the corresponding statement for algebraic spaces. Repeating the argument again with $X$ being an algebraic space (and in particular with the diagonal of $X$ being an embedding), we discover that we can assume $X$ is representable by embeddings of algebraic spaces. If we repeat the argument once more, we can assume that $X$ is representable by isomorphisms (the diagonal of an embedding being an isomorphism), in which case there is nothing to prove. \end{proof} \subsection{Modules over commutative rings} \label{sec:mod} Let $\ComRngMod$ be the category of pairs $(A, I)$ where $A$ is a commutative ring and $I$ is an $A$-module. A map $(A, I) \rightarrow (B, J)$ in $\ComRngMod$ is a ring homomorphism $\varphi : A \rightarrow B$ and an $A$-module homomorphism $I \rightarrow J_{[\varphi]}$. This category is bifibered (i.e., fibered and cofibered) over $\ComRng$ via the projection sending $(A,I)$ to $A$. The base change functor, with respect to a morphism $\varphi : B \rightarrow A$ sends the pair $(A,I)$ to $I_{[\varphi]}$ (which is in fact a splitting) and the co-base change functor sends $(B, J)$ to $(A, J \tensor_B A)$. The fibers of $\ComRngMod$ over $\ComRng$ are abelian categories, and $\ComRngMod$ forms a stack over $\ComRng$ in the fpqc topology. If $A$ is an $X$-algebra then by definition an $A$-module is a module under the underlying commutative ring of $A$. We write $X\hAlgMod$ for the category of pairs $(A, I)$ where $A \in X\hAlg$ and $I$ is a $A$-module. The morphisms in $X\hAlgMod$ are the evident ones: a map $(A, I) \rightarrow (B,J)$ is a morphism of $X$-algebras $\varphi : A \rightarrow B$ and a morphism of $A$-modules $I \rightarrow J_{[\varphi]}$. In more sententious terms, $X\hAlgMod$ is the strict fiber product $X\hAlg \fp_{\ComRng} \ComRngMod$. \subsection{Tangent vectors} \label{sec:tang} Suppose that $A$ is a commutative ring and $I$ is an $A$-module. We may form a new commutative ring $A + I$ whose underlying abelian group is $A \times I$ and multiplication is defined by \begin{equation*} (a, x) . (a', x') = (a a', a x' + a' x). \end{equation*} So, in particular, $I \subset A + I$ is an ideal that squares to zero. This is the \emph{trivial square-zero extension} of $A$ by $I$. Recall that if $\varphi : B \rightarrow A$ is a homomorphism of commutative rings, then the derivations of $B$ into $I_{[\varphi]}$ can be identified with the lifts of the diagram \begin{equation*} \xymatrix{ & A + I \ar[d] \\ B \ar@{-->}[ur] \ar[r]^{\varphi} & A . } \end{equation*} In other words, the $B$-algebra structures on $A + I$ lifting the specified $B$-algebra structure on $A$ can be identified with the $I$-valued tangent vectors of $\Spec B$ at the $A$-point $\Spec A \rightarrow \Spec B$. This is a commutative group, which we denote $\sT_B(A,I)$. The purpose of this section is to extend this definition to the case where $B$ is a homogeneous stack $X$, and make explicit the various functoriality properties of $\sT_X$ that will imply the categories $\sT_X(A, I)$ are abelian $2$-groups. The construction $(A, I) \mapsto A + I$ gives a new functor from $\Phi : \ComRngMod \rightarrow \ComRng$ (recall that we also have the projection which forgets $I$). This second functor is faithful since a morphism $(A, I) \rightarrow (B, J)$ can be recovered from the induced map $A + I \rightarrow B + J$, but it is not full. Let $X$ be a cofibered category over $\ComRng$. We may define a cofibered category $\sT_X$ to be the strict fiber product $X\hAlg \fp_{\ComRng} (\ComRngMod, \Phi)$. The objects of $\sT_X$ are triples $(A, I, \varphi)$ where $(A, I) \in X\hAlgMod$ and $\varphi$ is an $X$-algebra structure on $A + I$. A morphism in $\sT_X$ from $(A, I, \varphi)$ to $(B, J, \psi)$ is a morphism $u : (A, I) \rightarrow (B, J)$ and a morphism of $X$-algebras $(A + I, \varphi) \rightarrow (B + J, \varphi)$ lifting the map $A + I \rightarrow B + J$ induced from $u$. \begin{proposition} \label{prop:tang-cof} Let $X$ be a cofibered category over $\ComRng$. \begin{enumerate}[label=(\roman{*}), ref=(\roman{*})] \item $\sT_X$ is cofibered over $X\hAlgMod$ via the projection sending $(A, I, \varphi)$ to $(A, I)$. \label{der:1} \item $\sT_X$ is cofibered over $X\hAlg$ via the projection sending $(A, I, \varphi)$ to $A$. \label{der:2} \end{enumerate} Assume that $X$ is also homogeneous. \begin{enumerate}[label=(\roman{*}), ref=(\roman{*}), resume] \item $\sT_X$ is fibered over $X\hAlg$ via the projection sending $(A, I, \varphi)$ to $A$. \label{der:3} \item For each $A \in X\hAlg$, the category $\sT_X(A)$ is left exact and additively cofibered over $A\hMod$. \label{der:4} \item For any $(A, I) \in X\hAlgMod$, the category $\sT_X(A, I)$ is naturally equipped with the structure of an abelian $2$-group. \label{der:5} \end{enumerate} \end{proposition} \begin{proof} \ref{der:1} The functor $\sT_X(A, I) \rightarrow \sT_X(B, J)$ associated to a morphism $u : (A, I) \rightarrow (B, J)$ in $X\hAlgMod$ sends $(A, I, \varphi)$ to $(B, J, v_\ast \varphi)$ where $v : A + I \rightarrow B + J$ is the morphism induced from $u$. \ref{der:2} $X\hAlgMod$ is cofibered over $X\hAlg$ and $\sT_X$ is cofibered over $X\hAlgMod$ by Part~\ref{der:1}. \ref{der:3} Let $u : A \rightarrow B$ be a homomorphism of commutative rings, and let $J$ be a $B$-module. Suppose that $\xi \rightarrow \eta$\marg{maybe notation for $\xi$ and $\eta$ should be changed} is a morphism of $X$-algebras lifting $u$. By homogeneity, the map \begin{equation*} X(A + J_{[u]}) \rightarrow X(A) \fp_{X(B)} X(B + J) \end{equation*} is an equivalence. Restricting, via the projection to $X(A)$, to the fiber over $\xi$, we get an equivalence \begin{equation*} \sT_X(\xi, J_{[u]}) \rightarrow \sT_X(\eta, J) . \end{equation*} An inverse to this functor gives the base change functor $\sT_X(\eta, J) \rightarrow \sT_X(\xi, J_{[u]})$. \ref{der:4} Let $A$ be a commutative ring. Suppose that \begin{equation*} 0 \rightarrow I \rightarrow J \rightarrow K \rightarrow 0 \end{equation*} is an exact sequence of $A$-modules. Then we have a cartesian diagram of rings \begin{equation*} \xymatrix{ A + I \ar[r] \ar[d] & A + J \ar[d] \\ A \ar[r] & A + K . } \end{equation*} We therefore deduce that the diagram \begin{equation*} \xymatrix{ \sT_X(A, I) \ar[r] \ar[d] & \sT_X(A, J) \ar[d] \\ X(A) \ar[r] & \sT_X(A, K) } \end{equation*} is cartesian by homogeneity. Restricting to the fiber over $\xi \in X(A)$, we get a cartesian diagram \begin{equation*} \xymatrix{ \sT_X(\xi, I) \ar[r] \ar[d] & \sT_X(\xi, J) \ar[d] \\ \set{ \xi } \ar[r] & \sT_X(\xi, K) } \end{equation*} which means that the sequence \begin{equation*} 0 \rightarrow \sT_X(\xi, I) \rightarrow \sT_X(\xi, J) \rightarrow \sT_X(\xi, K) \end{equation*} is exact. \ref{der:5} An immediate consequence of \ref{der:4} (in view of \cite[Section~1.4]{ccct} or \cite{gst}). \end{proof} \subsection{Square-zero extensions} \label{sec:sq-zero} Recall that a square-zero extension of a commutative algebra $A$ by an $A$-module $I$ is a surjection of commutative algebras $A' \rightarrow A$ whose kernel consists of elements that square to zero, and an isomorphism between the kernel and $I$, with its $A'$-module structure induced by the projection. Let $\Exal$ be the category of triples $(A, I, A')$ where $A$ is a commutative ring, and $A'$ is a square-zero extension of $A$ by $I$. The projection $\Exal \rightarrow \ComRng$ makes $\Exal$ into a fibered category over the category of commutative rings. There is also a projection $\Exal \rightarrow \ComRngMod$ sending $(A, I, A')$ to $(A, I)$. If $X$ is a cofibered category over $\ComRng$, define $\Exal_X$ to be the category of triples $(A, I, A')$ where $A$ is an $X$-algebra and $A'$ is a square-zero $X$-algebra extension of $A$ by $I$. A morphism from $(A, I, A')$ to $(B, J, B')$ is a pair of compatible morphisms of $X$-algebras $\varphi : A \rightarrow B$ and $A' \rightarrow B'$ (which induce a morphism of $A$-modules $I \rightarrow J_{[\varphi]}$. The following is a more precise form of this definition: let $\Phi : \Exal \rightarrow \ComRng$ be the projection sending $(A, A')$ to $A'$. If $X$ is a cofibered category over $\ComRng$, then $\Exal_X = X \fp_{\ComRng} (\Exal, \Phi)$. \begin{proposition} \label{prop:def-cof} Let $X$ be a cofibered category over $\ComRng$. \begin{enumerate}[label=(\roman{*}), ref=(\roman{*})] \item $\Exal_X$ is cofibered over $\Exal$. \label{exal:5} \item $\Exal_X$ is cofibered over \'etale morphisms in $X\hAlgMod$. \label{exal:1} \end{enumerate} Assume that $X$ is also homogeneous. \begin{enumerate}[label=(\roman{*}), ref=(\roman{*}), resume] \item $\Exal_X$ is fibered over $\ComRng$ and the map $\Exal_X \rightarrow \ComRngMod$ is cartesian. \label{exal:2} \item For each $A \in \ComRng$, the category $\Exal_X(A)$ is left exact and additively cofibered over $A\hMod$. \label{exal:3} \item For any $(A, I) \in \ComRngMod$, the category $\Exal_X(A, I)$ is naturally equipped with the structure of an abelian $2$-group. \label{exal:4} \end{enumerate} \end{proposition} \begin{remark} Although $\Exal_X$ and $\sT_X$ have some similar functoriality properties, the forms of Propositions~\ref{prop:tang-cof} and~\ref{prop:def-cof} are not identical. This is because $\sT_X(A,I)$ is naturally related to the deformations of $X$, while $\Exal_X(A,I)$ is naturally related to the deformations of $A$. This can be seen using the cotangent complex, provided one is available: we can identify the isomorphism classes in $\sT_X(A,I)$ with $\Ext^0(\bL_X, I)$ and we can identify the isomorphism classes in $\Exal_X(A,I)$ with $\Ext^1(\bL_{A/X}, I)$. \end{remark} \begin{proof}[Proof of Proposition~\ref{prop:def-cof}] \ref{exal:5} This follows immediately from the fact that $X\hAlg$ is cofibered over $\ComRng$. \ref{exal:1} We have a commutative diagram \begin{equation*} \xymatrix{ \Exal_X \ar[r] \ar[d] & \Exal \ar[d] \\ X\hAlgMod \ar[r] & \ComRngMod . } \end{equation*} To show that $\Exal_X$ is cofibered over \'etale maps in $X\hAlgMod$, it suffices to show that $\Exal_X$ is cofibered over \'etale maps in $\ComRngMod$, since $X\hAlgMod$ is cofibered over $\ComRngMod$. Furthermore, $\Exal_X$ is cofibered over $\Exal$ by~\ref{exal:5}, so it is sufficient to show that $\Exal$ is cofibered over \'etale maps in $\ComRngMod$. One can factor a morphism in $\ComRngMod$ in a unique way into a morphism that is cocartesian over $\ComRng$ followed by a morphism in $A\hMod$ for some commutative ring $A$. Pushout of exact sequences immediately demonstrates that $\Exal(A)$ is cofibered over $A\hMod$, so the problem reduces to showing that $\Exal$ is cocartesian over \'etale morphisms of $\ComRng$. Let $A \rightarrow B \rightarrow C$ be a sequence of morphisms in $\ComRng$ with $A \rightarrow B$ \'etale, and let $\xi \rightarrow \omega$ be a morphism of $\Exal$ over $A \rightarrow C$. We must show there is a morphism $\xi \rightarrow \eta$ in $\Exal(A \rightarrow B)$, depending only on $\xi$ and the map $A \rightarrow B$, through which the map $\xi \rightarrow \omega$ factors uniquely. We can represent $\xi$ by a square-zero extension $A'$ of $A$, and $\omega$ by a square-zero extension $C'$ of $C$. By \cite[Th\'eor\`eme~1.1]{sga4-VIII}, the tensor product induces an equivalence from \'etale $A'$-algebras to \'etale $A$-algebras. Therefore there is an \'etale $A'$-algebra $B'$, determined uniquely up to unique isomorphism, and a cocartesian diagram \begin{equation*} \xymatrix{ A' \ar[r] \ar[d] & B' \ar[d] \\ A \ar[r] & B. } \end{equation*} To check the universality of $B'$, we must show that the commutative diagram of solid lines \begin{equation*} \xymatrix{ A' \ar@/^10pt/[drr] \ar[d] \ar[dr] \\ A \ar@/^10pt/[drr] \ar[dr] & B' \ar@{-->}[r] \ar[d] & C' \ar[d] \\ & B \ar[r] & C } \end{equation*} can be completed in a unique way by a dashed arrow. This is equivalent to showing that the diagram \begin{equation*} \xymatrix{ C & \ar[l] \ar@{-->}[dl] B' \\ C' \ar[u] & A' \ar[l] \ar[u] } \end{equation*} can be completed in a unique way by a dashed arrow. But the map $A' \rightarrow B'$ is \'etale and $C' \rightarrow C$ is an infinitesimal extension, so the dashed arrow and its uniqueness are guaranteed by the infinitesimal criterion for being \'etale \cite[D\'efinition~(19.10.2)]{ega4-1}. \ref{exal:2} Note first that $\Exal$ is fibered over $\ComRng$ by pullback of exact sequences. This proves the claim when $X$ is a point. To prove the general case, we must show that, given any sequence of morphisms $\xi \rightarrow \eta \rightarrow \omega$ in $X$ and square-zero extensions $\xi' \rightarrow \xi$ and $\omega' \rightarrow \omega$, there is a square-zero extension $\eta' \rightarrow \eta$, independent of $\xi$, and a unique factorization of $\xi' \rightarrow \omega'$ through $\eta'$: \begin{equation*} \xymatrix{ \xi' \ar@/^10pt/[drr] \ar@{-->}[dr] \ar[d] \\ \xi \ar@/^10pt/[drr] \ar[dr] & \eta' \ar[d] \ar[r] & \omega' \ar[d] \\ & \eta \ar[r] & \omega } \end{equation*} The diagram illustrates that we are searching for a fiber product $\eta' = \eta \fp_{\omega} \omega'$, which is provided by the homogeneity of $X$. \ref{exal:3} First we show that if $\xi$ is an $X$-algebra with underlying commutative ring $A$ then $\Exal_X(\xi)$ is additively cofibered over $A\hMod$. We must show that $\Exal_X(\xi, 0) = 0$, which amounts to noting that $\Exal_X(A, 0) \rightarrow X(A)$ is an equivalence. We must also check that \begin{equation} \label{eqn:3} \Exal_X(\xi, I \oplus J) \rightarrow \Exal_X(\xi, I) \times \Exal_X(\xi, J) \end{equation} is an equivalence. We note now that if $\xi' \in \Exal_X(\xi, I \oplus J)$ and $\eta' \in \Exal_X(A, I)$ and $\omega' \in \Exal_X(A, J)$ are the objects induced from $\xi'$, then there is a cartesian diagram \begin{equation*} \xymatrix{ \xi' \ar[r] \ar[d] & \eta' \ar[d] \\ \omega' \ar[r] & \xi } \end{equation*} in $X\hAlg$. Therefore $\xi'$ can be recovered up to unique isomorphism as $\eta' \fp_{\xi} \omega'$, which implies that the morphism displayed in Equation~\eqref{eqn:3} is an equivalence. To demonstrate left exactness, suppose that \begin{equation*} 0 \rightarrow I \rightarrow J \rightarrow K \rightarrow 0 \end{equation*} is an exact sequence of $\xi$-modules for an $X$-algebra $\xi$ with underlying commutative ring $A$. We must show that the sequence \begin{equation*} 0 \rightarrow \Exal_X(\xi, I) \rightarrow \Exal_X(\xi, J) \rightarrow \Exal_X(\xi, K) \end{equation*} is exact. Recall that the exactness of this diagram is equivalent to the following diagram's being cartesian: \begin{equation} \label{eqn:2} \xymatrix{ \Exal_X(A, I) \ar[r] \ar[d] & \Exal_X(A, J) \ar[d] \\ X(A) \ar[r] & \Exal_X(A, K) } \end{equation} where the maps $\Exal_X(A, I) \rightarrow X(A) \rightarrow \Exal_X(A, K)$ come from the identification $X(A) \simeq \Exal_X(A, 0)$. An object of the fiber product can be identified with a diagram of solid lines \begin{equation} \label{eqn:1} \xymatrix{ \xi' \ar@{-->}[r] \ar@{-->}[d] & \eta' \ar[d] \\ \xi \ar[r] & \omega' } \end{equation} inside $X\hAlg/\xi$ where $\eta' \in X(A, J)$, $\omega' \in X(A, K)$, and $\eta' \rightarrow \omega'$ is cocartesian over $A\hMod$. By homogeneity, this diagram has a fiber product $\xi'$ completing the diagram above. Let \begin{equation*} \xymatrix{ A'_1 \ar[r] \ar[d] & A'_2 \ar[d] \\ A \ar[r] & A'_3 } \end{equation*} be the image of the cartesian diagram~\eqref{eqn:1} in $\ComRng$. By homogeneity, this diagram is cartesian. Furthermore, $A'_2$ is a square-zero extension of $A$ by $J$ and $A'_3$ is a square-zero extension of $A$ by $K$. It follows that $A'_1$ is a square-zero extension of $A$ by $I$. This demonstrates an inverse to the natural map $\Exal_X(A, I) \rightarrow \Exal_X(A, J) \fp_{\Exal_X(A, J)} X(A)$ and therefore shows that Diagram~\eqref{eqn:2} is cartesian. \ref{exal:4} An immediate consequence of~\ref{exal:3} and \cite{ccct} or \cite{gst}. \end{proof} \section{Obstruction theories} \label{sec:def} \subsection{Additively cofibered categories} \label{sec:acc} Let $X$ be a stack. A \emph{stack of cofibered categories}\marg{not sure if this is the right name} on $X$ is a cofibered category $\sE$ over $X\hAlgMod$ that is a stack over $X\hAlg$ in the \'etale topology. We call it \emph{left exact} and \emph{additively cofibered} if, for each $X$-algebra $A$, the cofibered category $\sE(A)$ over $A\hMod$ has the corresponding property. Note that this implies that each $\sE(A, I)$ possesses a zero object $e(A, I)$, well-defined up to unique isomorphism, and these can be combined into a cocartesian section $e$ of $\sE$ over $X\hAlgMod$. We will call a stack of additively cofibered categories \emph{cartesian} if it is also cartesian over the arrows of $X\hAlgMod$ that are cartesian over $X\hAlg$. Suppose that $\sE$ is a stack of additively cofibered categories on $X\hAlgMod$. We call $\sE$ \emph{exact} if it is left exact and, for any $X$-algebra $A$, and any exact sequence \begin{equation*} 0 \rightarrow I \rightarrow J \rightarrow K \rightarrow 0, \end{equation*} the map $\sE(A,J) \rightarrow \sE(A,K)$ is \emph{locally} essentially surjective. This means that for any section $\xi$ of $\sE(A,K)$, the collection of $A$-algebras $B$ such that $\xi \tensor_A B$ can be lifted, up to isomorphism, to $\sE(B, J \tensor_A B)$ constitutes an \'etale cover of $A$. Let $v : (A,I) \rightarrow (B, J)$ be a morphism of $X\hAlgMod$ that is cartesian over $X\hAlg$. We have a morphism of $X$-algebras $u : A \rightarrow B$ and an induced isomorphism $I \simeq J_{[u]}$. Then $v$ induces functors $\Phi : \sE(B,J) \rightarrow \sE(A,I)$ and $\Psi : \sE(A,I) \rightarrow \sE(B,J)$, each well-defined up to unique isomorphism, that must be \emph{adjoint}: for each $x \in \sE(A,I)$ and $y \in \sE(B,J)$, we have \begin{equation*} \Hom_{\sE(A,I)}(x, \Phi(y)) = \Hom_f(x,y) = \Hom_{\sE(B,J)}(\Psi(x), y) . \end{equation*} But both $\sE(A,I)$ and $\sE(B,J)$ are groupoids by \cite[1.5~a)]{ccct}, so $\Phi$ and $\Psi$ must be mutually inverse functors. \begin{proposition} \label{prop:cart} Let $\sE$ be a cofibered category over $X\hAlgMod$. The following conditions are equivalent. \begin{enumerate}[label=(\roman{*}), ref=(\roman{*})] \item $\sE$ is cartesian over the morphisms of $X\hAlgMod$ that are cartesian over $X\hAlg$. \label{cart:1} \item $\sE$ is cartesian over $X\hAlg$. \label{cart:4} \item For every map $(A,I) \rightarrow (B,J)$ of $X\hAlgMod$ that is cartesian over $X\hAlg$, the functor $\sE(A,I) \rightarrow \sE(B,J)$ is an equivalence. \label{cart:2} \item For every $X$-algebra homomorphism $f : A \rightarrow B$ and every $B$-module $J$, the functor $\sE(A,J_{[f]}) \rightarrow \sE(B,J)$ is an equivalence. \label{cart:3} \end{enumerate} \end{proposition} \begin{proof} The implication \ref{cart:1} \implies \ref{cart:2} was demonstrated before the statement of the proposition and \ref{cart:2} \implies \ref{cart:3} is obvious. The equivalence \ref{cart:1} \iff \ref{cart:4} holds because $X\hAlgMod$ is cartesian over $X\hAlg$. We prove \ref{cart:3} \implies \ref{cart:4}. Let $f : A \rightarrow B$ be a homomorphism of $X$-algebras and let $J$ be a $B$-module. Let $y$ be an object of $\sE(B,J)$. Let $I = J_{[f]}$. We show that there is an object $x \in \sE(A,J_{[f]})$ and a morphism $x \rightarrow y$ of $\sE$ that is cartesian above $X\hAlg$. Since the co-base change functor $\Psi : \sE(A,I) \rightarrow \sE(B,J)$ is an equivalence, there is an object $x \in \sE(A,I)$ such that $\Psi(x)$ is isomorphic to $y$. Choosing such an $x$ and such an isomorphism, we get an object of $\Hom_{\sE(B,J)}(\Psi(x), y) = \Hom_f(x,y)$, which gives the required map $\varphi : x \rightarrow y$. It remains to demonstrate that $\varphi$ is cartesian over $X\hAlg$. Let $g : C \rightarrow A$ be a homomorphism of $X$-algebras and suppose that $z \in \sE(C,K)$ for some $C$-module $K$. We must show that the natural map \begin{equation} \label{eqn:26} \Hom_{g}(z,x) \rightarrow \Hom_{fg}(z,y) \end{equation} is a bijection. Associated to any $\alpha \in \Hom_g(z,x)$ we have a map $(C,K) \rightarrow (B,J)$, hence a homomorphism of $C$-modules $K \rightarrow J_{[fg]} = I_{[f]}$. Hence we can decompose~\eqref{eqn:26} as \begin{equation*} \coprod_{u : K \rightarrow I_{[f]}} \Hom_{(g,u)}(z,x) \rightarrow \coprod_{u : K \rightarrow J_{[fg]}} \Hom_{(fg, u)}(z,y) . \end{equation*} It suffices to demonstrate that the map is a bijection on each component. Now, because $\sE$ is cocartesian over $X\hAlgMod$, there must be a morphism $z \rightarrow w$ of $\sE$ that is cocartesian over the map $(g,u) : (C,K) \rightarrow (C,I_{[f]})$. The horizontal arrows in the commutative diagram \begin{equation*} \xymatrix{ \Hom_{(g,\id_{I_{[f]}})}(w,x) \ar[r] \ar[d] & \Hom_{(g,u)}(z,x) \ar[d] \\ \Hom_{(fg,\id_{I_{[f]}})}(w,y) \ar[r] & \Hom_{(fg,u)}(z,y) }\end{equation*} are both bijections, so it will now suffice to demonstrate that the vertical arrow on the left is a bijection. Use the notation \begin{gather*} \Psi_f : \sE(A,I) \rightarrow \sE(B,J) \\ \Psi_g : \sE(C,I_{[f]}) \rightarrow \sE(A,I) \\ \end{gather*} for the co-base change functors. We have \begin{align*} \Hom_{fg}(w,y) & = \Hom_{\sE(B,J)}(\Psi_{f} \Psi_g(w), y) & & \text{because $w \rightarrow \Psi_f \Psi_g(w)$ is cocartesian} \\ & = \Hom_{\sE(A,I)}(\Psi_{g} (w), x) & & \text{because $\Psi_f$ is an equivalence} \\ & & & \qquad \text{and $\Psi_f(x) \rightarrow y$ is an isomorphism} \\ & = \Hom_g(w,x) & & \text{because $w \rightarrow \Psi_g(w)$ is cocartesian} . \end{align*} \end{proof} \begin{definition} \label{def:obs} Let $h : X \rightarrow Y$ be a morphism of homogeneous stacks over $\ComRng$. An obstruction theory for $X$ over $Y$ is a cartesian stack of left exact additively cofibered categories $\sE$ over $X$ equipped with a cartesian diagram \begin{equation} \label{eqn:6}\xymatrix{ \Exal_X \ar[r] \ar[d] & e \ar[d] \\ h^\ast \Exal_Y \ar[r] & \sE } \end{equation} in which the horizontal arrows preserve \begin{enumerate} \item arrows that are cocartesian over \'etale maps in $\ComRngMod$, and \item arrows that are cartesian over $\ComRng$. \end{enumerate} \end{definition} \subsection{The relative intrinsic normal stack} \label{sec:ins} To specify Diagram~\eqref{eqn:6} should be the same as to give a morphism \begin{equation*} h^\ast \Exal_Y / \Exal_X \rightarrow \sE , \end{equation*} provided one can give a satisfactory definition of the source. We indicate how this can be done when $X$ and $Y$ are algebraic stacks and $X \rightarrow Y$ is of \emph{Deligne--Mumford type}, meaning that for any $Y$-algbra $A$, the stack $X_A$ on $A$-algebras is a Deligne--Mumford stack. We make some abbreviations in this section. If $A$ is an $X$-algebra, then we write $\Spec A$ for the spectrum of its underlying commutative ring, equipped with a map to the stack $X$. We also write $\et(A)$ in place of $\et(\Spec A)$ for the small \'etale site of $\Spec A$. If $J$ is an $A$-module, we write $\tJ$ for the associated sheaf on $\et(A)$. Thus $\tA$ is the structure sheaf on $\et(A)$. Assume first that both $X$ and $Y$ are Deligne--Mumford stacks. Then the map $h : X \rightarrow Y$ can be represented by a morphism of \'etale locally ringed topoi $(h, \varphi) : (\et(X), \cO_X) \rightarrow (\et(Y), \cO_Y)$. \begin{definition} \label{def:ins} Suppose that $X \rightarrow Y$ is a morphism of Deligne--Mumford stacks. Define $\sN_{X/Y}$ to be the category of triples $(C, J, \sA)$ where $(C,J) \in X\hAlgMod$, corresponding to a morphism $f : \et(C) \rightarrow \et(X)$ and a homomorphism $\varphi : f^\ast \cO_X \rightarrow \tC$ of sheaves of rings, and $\sA$ is a square-zero extension of $f^\ast \cO_X$ by $\tJ_{[\varphi]}$ as a $f^\ast h^\ast \cO_Y$-algebra. \end{definition} Let $X\hEtAlg$ be the category of \'etale $X$-algebras. Then we can restrict the projections $\sN_{X/Y} \rightarrow X\hAlg$ and $h^\ast \Exal_Y \rightarrow X\hAlg$ via the inclusion $X\hEtAlg \rightarrow X\hAlg$. \begin{lemma} The restrictions of $\sN_{X/Y}$ and $h^\ast \Exal_Y$ to $X\hEtAlg$ are equivalent. \end{lemma} \begin{proof} Suppose that $C$ is an \'etale $X$-algebra. Let $i : \et(C) \rightarrow \et(X)$ be the canonical map. Then the sheaves $\tC$ and $i^\ast \cO_X$ on $\et(C)$ are canonically isomorphic. Moreover, if $J$ is a $C$-module, then extensions of $\tC$ by $\tJ$ as a $i^\ast h^\ast \cO_Y$-algebra are equivalent by taking global sections to extensions of $C$ by $J$ as a $Y$-algebra. \end{proof} Let $(B, I, \sB)$ and $(C, J, \sC)$ be objects of $\sN_{X/Y}$ with $u : \et( B) \rightarrow \et(X)$ and $v : \et( C) \rightarrow \et(X)$ the corresponding morphisms of \'etale sites. A morphism $(B, I, \sB) \rightarrow (C, J, \sC)$ of $\sN_{X/Y}$ is a morphism $f : (B,I) \rightarrow (C,J)$ of $X\hAlgMod$, corresponding to a map $w : \et( C) \rightarrow \et( B)$, and a morphism of extensions $w^\ast \sC \rightarrow \sB$ that is compatible with the $Y$-algebra structures and the maps $w^\ast \tC \rightarrow \tB$ and $w^\ast \tJ \rightarrow \tI$ induced from $f$.\marg{is it clear enough what that means?} There is a projection $\sN_{X/Y} \rightarrow X\hAlgMod$ sending a triple $(C, J, \sA)$ to $(C,J)$, as in Definition~\ref{def:ins}. \begin{proposition} \begin{enumerate}[label=(\roman{*}), ref=(\roman{*})] \item \label{ins:1} $\sN_{X/Y}$ is cocartesian over $X\hAlgMod$. \item \label{ins:2} $\sN_{X/Y}$ is cartesian over $X\hAlg$ and the projection $\sN_{X/Y} \rightarrow X\hAlgMod$ preserves morphisms that are cartesian over $X\hAlg$. \item \label{ins:3} For each $X$-algebra $C$, the cofibered category $\sN_{X/Y}(C)$ over $C\hMod$ is additively cofibered and left exact. \end{enumerate} \end{proposition} \begin{proof} For the first two assertions we limit ourselves to describing the pushforward and pullback functors. Their universal properties are not difficult to verify. \ref{ins:1} Suppose that $(B, I) \rightarrow (C, J)$ is a morphism in $X\hAlgMod$ whose corresponding morphism of \'etale sites is $w : \et(C) \rightarrow \et(B)$ and $(B,I,\sB) \in \sN_{X/Y}(B,I)$. Let $u : \et(B) \rightarrow \et(X)$ and $v : \et(C) \rightarrow \et(X)$ be the morphisms of \'etale sites. Then $w^\ast$ is exact so $w^\ast \sB$ is an extension of $w^\ast u^\ast \cO_X \simeq v^\ast \cO_X$ by $w^\ast \tI$. Pushing out this extension via the map $w^\ast \tI \rightarrow \tJ$, we get the desired extension of $v^\ast \cO_X$ by $\tJ$. \ref{ins:2} Suppose $\sC \in \sN_{X/Y}(C)$ for some $C \in X\hAlg$, and $u : \et(C) \rightarrow \et(X)$ is the corresponding map of \'etale sites. Let $\varphi : B \rightarrow C$ be an $X$-algebra homomorphism and let $w : \et(C) \rightarrow \et(B)$ be the corresponding map of \'etale sites. Write $v$ for the map of \'etale sites $\et(B) \rightarrow \et(X)$. Then $\sC$ is an extension of $u^\ast \cO_X$ by $\tJ$ for some $C$-module $J$. Using the fact that $w$ is affine, so $R^1 w_\ast \tJ = 0$, it follows that $w_\ast \sC$ is an extension of $w_\ast u^\ast \cO_X$ by $w_\ast \tJ = \tJ_{[\varphi]}$. Pulling this extension back via the map $v^\ast \cO_X \rightarrow w_\ast u^\ast \cO_X$ gives the extension of $v^\ast \cO_X$ by $\tJ_{[\varphi]}$ that we seek. \ref{ins:3} The proof is very similar to that of Proposition~\ref{prop:def-cof}~\ref{exal:3} and we omit it. A proof of the left exactness may be found in \cite[Section~7.1.11]{ccct}. \end{proof} Now we construct a cartesian diagram~\eqref{eqn:6} (with $\sE = \sN_{X/Y}$), demonstrating that $\sN_{X/Y}$ is an obstruction theory for $X$ over $Y$. First we construct the map $h^\ast \Exal_Y \rightarrow \sN_{X/Y}$. An object of $h^\ast \Exal_Y$ is representable by an $X$-algebra $C$ and a square-zero extension $C'$ of $C$ as a $Y$-algebra. Let $I$ denote the ideal of $C$ in $C'$ and let $f$ denote the map of \'etale topoi $\et( C) \rightarrow X$ coming from the $X$-algebra structure on $C$. Since the \'etale site of $C'$ is the same as that of $C$, we obtain a square-zero extension $\tC'$ of $\tC$ by $\tI$ on $\et( C)$ as a $f^\ast h^\ast \cO_Y$-algebra. Pulling back via the map $f^\ast \cO_X \rightarrow \tC$, we get a square-zero extension $\sC = \tC' \fp_{\tC} f^\ast \cO_X$ of $f^\ast \cO_X$ on $\et(C)$ as a $f^\ast h^\ast \cO_Y$-algebra: \begin{equation} \label{eqn:24} \xymatrix{ 0 \ar[r] & \tI \ar[r] \ar[d] & \sC \ar[r] \ar[d] & f^\ast \cO_X \ar[r] \ar[d] & 0 \\ 0 \ar[r] & \tI \ar[r] & \tC' \ar[r] & \tC \ar[r] & 0. } \end{equation} This gives an object of $\sN_{X/Y}(C, I)$. To verify that Diagram~\eqref{eqn:6} is cartesian, we check that isomorphisms between the object constructed above and the trivial extension of $f^\ast \cO_X$ by $\tI$ are in bijection with the $X$-algebra structures on $\tC'$ lifting the given one on $\tC$. But to give an $X$-algebra structure on $\tC'$ is the same as to lift the diagram \begin{equation*} \xymatrix{ & & & f^\ast \cO_X \ar@{-->}[dl] \ar[d] \\ 0 \ar[r] & \tI \ar[r] & \tC' \ar[r] & \tC \ar[r] & 0 } \end{equation*} which is of course the same as to split the extension in the first row of Diagram~\eqref{eqn:24}. \begin{proposition} \label{prop:ins} Let $\sE$ be an obstruction theory for $X$ over $Y$. Then there is a uniquely determined (up to unique isomorphism) fully faithful functor $\sN_{X/Y} \rightarrow \sE$ inducing the cartesian diagram~\eqref{eqn:6} from the corresponding diagram for $\sN_{X/Y}$ (up to a unique isomorphism). \end{proposition} \begin{proof} Since both $\sE$ and $\sN_{X/Y}$ are stacks on $X\hAlg$, it is sufficient to describe the value of the map $\sN_{X/Y} \rightarrow \sE$ on $(C,J)$ \'etale locally in $C$. Since $X$ is a Deligne--Mumford stack, this means that we can assume there is an \'etale $X$-algebra $D$ and an $X$-algebra homomorphism $\varphi : D \rightarrow C$. Let $g : \et(C) \rightarrow \et(D)$ be the induced morphism of \'etale sites. The pushforward maps \begin{gather*} \sN_{X/Y}(D, J_{[\varphi]}) \rightarrow \sN_{X/Y}(C, J) \\ \sE(D, J_{[\varphi]}) \rightarrow \sN_{X/Y}(C, J) \end{gather*} are both equivalences by Proposition~\ref{prop:cart}. Moreover, if the map $D \rightarrow C$ factors through a map $\psi : D' \rightarrow C$, where $D'$ is another \'etale $X$-algebra, then the diagrams \setlength{\hoffset}{-0.5in} \begin{equation*} \xymatrix@C=0pt{ \sF_{X/Y}(D, J_{[\varphi]}) \ar[rr] \ar[dr] & & \sF_{X/Y}(D', J_{[\psi]}) \ar[dl] \\ & \sF_{X/Y}(C,J) } \end{equation*} (with $\sF = \sN$ or $\sF = \sE$) commute in a canonical way, and a similar statement holds for the tetrahedra coming from a sequence $D \rightarrow D' \rightarrow D'' \rightarrow C$. Therefore, to define the map over all $X$-algebras $C$, it is sufficient to provide a definition when $f : \et(C) \rightarrow X$ is \'etale. In this case, the map $f^\ast \cO_X \rightarrow \tC$ is an isomorphism, so $\sN_{X/Y}(C) = h^\ast \Exal_Y(C)$ and we have been given a map $h^\ast \Exal_Y(C) \rightarrow \sE(C)$ by hypothesis. To check that the cartesian diagram~\eqref{eqn:6} can be recovered from the one for $\sN_{X/Y}$, it is sufficient to verify this over all \'etale $X$-algebras, in which case it is true by definition. \end{proof} The proposition demonstrates that $\sN_{X/Y}$ is the \emph{universal} obstruction theory for $X$ over $Y$. We can extend this construction to the situation where $Y$ is an algebraic stack and $X$ is only assumed to be of Deligne--Mumford type over $Y$. As before, it is enough to define $\sN_{X/Y}(C, J)$ \'etale locally in $C$, so we can assume that there is a smooth $Y$-algebra $B$ and a $Y$-algebra homomorphism $B \rightarrow C$. We then take $\sN_{X/Y}(C,J) = \sN_{X_B/B}(C,J)$. To demonstrate that this is well-defined, we must prove \begin{lemma}[cf.\ {\cite[Lemma~5.3]{def2}}] If $A \rightarrow B$ is a flat homomorphism of commutative rings, $X_A$ is a Deligne--Mumford stack over $A$, and $X_B = X \tensor_A B$, then for any $X_B$-algebra $C$ and any $C$-module $J$, the natural map $\sN_{X_B/B}(C,J) \rightarrow \sN_{X_A/A}(C,J)$ is an equivalence. \end{lemma} \begin{proof} We describe an inverse. Let $f : \et(C) \rightarrow \et(X_B)$ and $g : \et(C) \rightarrow \et(X_A)$ be the maps induced from the $X_A$- and $X_B$-algebra structures on $C$. Suppose that $J$ is a $C$-module and $\sC$ is an extension of $g^\ast \cO_{X_A}$ by $\tJ$. Then $\sC \tensor_A B$ is an extension of $g^\ast \cO_{X_A} \tensor_A B = f^\ast \cO_{X_B}$ by $\tJ \tensor_A B$. There is a canonical map $\tJ \tensor_A B \rightarrow \tJ$ coming from the bilinear map $B \times J \rightarrow J$ induced from the $C$-action on $J$ and the map $B \rightarrow C$. Pushing out the extension by this map gives the required extension of $f^\ast \cO_{X_B}$ by $\tJ$. \end{proof} The verification that this is an obstruction theory for $X$ over $Y$ is a local problem and reduces to the verification that $\sN_{X_B/B}$ is an obstruction theory for $X_B$ over $B$, which was shown above. \begin{corollary} \label{cor:ins} An obstruction theory for a morphism $X \rightarrow Y$ of Deligne--Mumford type may be specified by giving a cartesian, left-exact, additively cofibered category $\sE$ over $X\hAlgMod$ and a fully faithful morphism $\sN_{X/Y} \rightarrow \sE$ respecting morphisms that are cocartesian over $X\hAlgMod$ and morphisms that are cartesian over $X\hAlg$. \end{corollary} \begin{remark} Given a suitable definition of a $2$-category of obstruction, the corollary may be rephrased to say that the $2$-category of obstruction theories is equivalent to the $2$-category of fully faithful morphisms $\sN_{X/Y} \rightarrow \sE$. \end{remark} \subsection{Global obstructions} \label{sec:global} Let $h : X \rightarrow Y$ be a morphism of homogeneous stacks over $\ComRng$ and suppose that $\sE$ is an obstruction theory for $X$ over $Y$. Consider a lifting problem \begin{equation} \label{eqn:4} \xymatrix{ S \ar[r] \ar[d] & X \ar[d]^h \\ S' \ar[r] \ar@{-->}[ur] & Y } \end{equation} in which $S$ and $S'$ are algebraic spaces and $S'$ is a square-zero extension of $S$. This implies that the ideal $J$ of $S$ in $S'$ is a quasi-coherent $\cO_S$-module on the small \'etale site of $S$ (which we identify with the small \'etale site of $S'$). Let $\usE$ be the stack on the small \'etale site of $S$ whose category of sections over $U$ is $\sE(U, J_U)^\circ$ when $U$ is affine and \'etale over $S$. This is a legitimate definition since $\sE$ was assumed to be a stack in the \'etale topology. If it becomes necessary to specify the dependence of $\usE$ on $S$ and $J$, we will write $\usE(S, J)$. Define $\uDef_X(S,\sI)$ to be the stack on the small \'etale site of $S$ whose sections are commutative diagrams \begin{equation*} \xymatrix{ S \ar[r] \ar[d] & X \\ S' \ar[ur] } \end{equation*} where $S'$ is a square-zero extension of $S$ by the quasi-coherent sheaf of ideals $\sI$. Defining $\uDef_Y$ likewise, so that $h^\ast \uDef_Y(S,\sI)$ is the collection of commutative diagrams \begin{equation*} \xymatrix{ S \ar[r] \ar[d] & X \ar[d] \\ S' \ar[r] & Y } \end{equation*} where $S'$ is a square-zero extension of $S$ by $\sI$, we obtain a map $\uDef_X(S,\sI) \rightarrow h^\ast \uDef_Y(S,\sI)$. Note that if $S = \Spec A$ is affine, we have an equivalence \begin{gather*} \uDef_X(S, \tI) = \Exal_X(A, I)^\circ \\ h^\ast \uDef_Y(S, \tI) = h^\ast \Exal_Y(A, I)^\circ \end{gather*} and these equivalences are functorial in $A$. When $S$ and $\sI$ are fixed, we omit them from the notation and write $\uDef_X$ for $\uDef_X(S, \sI)$. Since $\sE$ is an obstruction theory and $\uDef_X$, $\uDef_Y$, and $\usE$ are all stacks, we have a cartesian diagram \begin{equation} \label{eqn:5} \xymatrix{ \uDef_X \ar[d] \ar[r] & e \ar[d] \\ h^\ast \uDef_Y \ar[r]^\omega & \usE . } \end{equation} The commutativity of the diagram above induces a map \begin{equation*} \uDef_X \rightarrow \set[(\xi, \varphi)]{\xi \in \uDef_Y, \varphi : \omega(\xi) \simeq e} \end{equation*} from $\uDef_X$ to the category of pairs $(\xi, \varphi)$ where $\xi$ is a section of $\uDef_Y$ and $\varphi$ is an isomorphism between the image $\omega(\xi)$ in $\usE$ and the zero section $e$ of $\usE$. Note now that $h^\ast \uDef_Y$ can be identified with the collection of diagrams of solid lines~\eqref{eqn:4} and that to solve the lifting problem indicated in~\eqref{eqn:4} is precisely to lift a section of $h^\ast \uDef_Y$ to one of $\uDef_X$. The cartesian diagram~\eqref{eqn:5} then provides us with a bijection between lifts of $\xi \in h^\ast \uDef_Y$ and isomorphisms between $\omega(\xi)$ and $e$. In other words, we may think of $\omega(\xi)$ as the obstruction to the existence of a lift of $\xi$: it is isomorphic to $e$ if and only if a lift exists and such isomorphisms correspond exactly to the lifts. \section{Representability} \label{sec:rep} \subsection{Additively cofibered categories associated to complexes} Let $\sC$ be an abelian category. Suppose that $\bF_\bullet$ is a chain complex. Following \cite[Section~3.1]{biext1}, we define an additively cofibered category $\Psi_{\bF_\bullet}$ of diagrams \begin{equation*} \xymatrix{ & & & \bF_2 \ar[d] \ar@/_10pt/[ddl]_0 \\ & & & \bF_1 \ar[d] \ar[dl] \\ 0 \ar[r] & J \ar[r] & X \ar[r] & \bF_0 \ar[r] & 0 } \end{equation*} in which the sequence at the bottom is exact. A morphism in $\Psi_{\bF_\bullet}$ from $(X, I, \alpha) \in \Psi_{\bF_\bullet}(I)$ to $(Y,J,\beta) \in \Psi_{\bF_\bullet}(J)$ is a commutative diagram \begin{equation*} \xymatrix@R=10pt{ 0 \ar[r] & I \ar[r] \ar[d] & X \ar[r] \ar[d] & \bF_0 \ar@{=}[d] \ar[r] & 0 \\ 0 \ar[r] & J \ar[r] & Y \ar[r] & \bF_0 \ar[r] & 0 } \end{equation*} such that the map $X \rightarrow Y$ carries $\alpha$ to $\beta$. Objects of $\Psi_{\bF_\bullet}(J)$ are called \emph{extensions of $\bF_\bullet$ by $J$}. We also write $\Ext(\bF_\bullet, J)$ for $\Psi_{\bF_\bullet}(J)$. \begin{example} \begin{enumerate}[label=(\roman{*})] \item If $\bF_\bullet = \bF_1[1]$, we have $\Psi_{\bF_\bullet}(J) = \Hom(\bF_\bullet, J)$. \item If $\bF = \bF_0[0]$ then $\Psi_{\bF_\bullet}(J)$ is the category of extensions of $\bF_0$ by $J$. \end{enumerate} \end{example} \begin{definition} \label{def:rep-loc} An additively cofibered category $\sE$ over an abelian category $\sC$ is said to be \emph{representable} if there exists a $2$-term complex $\bF_\bullet$ in $\sC$, concentrated in degrees $[-1,0]$ and an isomorphism $\sE \simeq \Psi_{\bF_\bullet}$. \end{definition} \begin{remark} Note that if $\bE_\bullet$ is any chain complex, $\Psi_{\bE_\bullet} = \Psi_{\tau_{\geq -1} \bE_\bullet}$, so the definition is equivalent to the one that would be obtained by suppressing the restriction on the degrees in which $\bE_\bullet$ is concentrated. \end{remark} The definition of $\Psi_{\bF_\bullet}$ can be extended to cochain complexes. Suppose that $\bE^\bullet$ is a cochain complex (in non-negative degrees). Following \cite[Section~3.3.2]{biext1}, we define $\Psi_{\bF_\bullet}(\bE^\bullet)$ to be the category of diagrams \begin{equation*} \xymatrix{ & & & \bF_2 \ar[d] \ar@/_10pt/[ddl]_0 \\ & & & \bF_1 \ar[d] \ar[dl] \\ 0 \ar[r] & \bE^0 \ar[r] \ar[d] & X \ar[r] \ar[dl] \ar@/^10pt/[ddl]^0 & \bF_0 \ar[r] & 0 \\ & \bE^1 \ar[d] \\ & \bE^2 } \end{equation*} in which the middle row is exact and the maps \begin{gather*} \bF_1 \rightarrow X \rightarrow \bE^1 \\ \bF_2 \rightarrow X \\ X \rightarrow \bE^2 \end{gather*} are all zero. We note that these are abelian examples of \emph{butterflies} \cite{Ald,AN1,AN2}. The construction above is a special case of a more general construction. Suppose that $\sE$ is an additively cofibered category over an abelian category $\sC$. Let $\bF^\bullet$ be a cochain complex. Define $\sE(\bF^\bullet)$ to be the category of pairs $(X, \alpha)$ where $X \in \sE(\bF^0)$ and $\alpha : d_\ast X \simeq 0_{\sE(\bF^1)}$ is an isomorphism inducing the zero isomorphism \begin{equation*} 0_{\sE(\bF^2)} \simeq d_\ast d_\ast X \xrightarrow{d_\ast \alpha} d_\ast 0_{\sE(\bF^1)} \simeq 0_{\sE(\bF^2)} . \end{equation*} This determines an additively cofibered category over the category of cochain complexes. Moreover, if $\sE$ is a left exact and $\bF^\bullet \rightarrow \bG^\bullet$ is a quasi-isomorphism then by \cite[Proposition~2.7]{ccct}, the induced map $\sE(\bF^\bullet) \rightarrow \sE(\bG^\bullet)$ is an equivalence. Suppose that $X$ is a stack in groupoids over $\ComRng$ and $\bF_\bullet$ is a chain complex of sheaves of flat modules\footnote{In fact, the definition and many of the results are valid when only $\bF_0$ is required to be flat.} on $X\hAlg$. That is, for each $X$-algebra $A$, we have a chain complex $\bF_\bullet(A)$ of sheaves of $\tA$-modules on $\et(A)$, concentrated in degrees $\leq 0$, with $\bF_i(A)$ flat over $\tA$, equipped with maps $u^\ast \bF_\bullet(A) \rightarrow \bF_\bullet(B)$ for each morphism $A \rightarrow B$ of $X$-algebras corresponding to $u : \et(B) \rightarrow \et(A)$. These maps are required to satisfy the expected compatibility condition and to induce isomorphisms $u^\ast \bF_\bullet(A) \tensor_{u^\ast \tA} \tB \xrightarrow{\sim} \bF_\bullet(B)$. Define $\Psi_{\bF_\bullet}$ to be the category of triples $(A, J, X)$ where $A$ is an $X$-algebra, $J$ is an $A$-module, and $X$ is an extension of $\bF_\bullet(A)$ by~$\tJ$. If $\bE^\bullet$ is a chain complex of $A$-modules then we can define $\Psi_{\bF_\bullet}(A, \bE^\bullet) = \Psi_{\bF_\bullet(A)}(\tbE^\bullet)$. \begin{lemma} Suppose that $X$ is a stack in groupoids on $\ComRng$. Let $\bF_\bullet$ be a chain complex of flat modules on $X$. The category $\Psi_{\bF_\bullet}$ defined above is a left exact, additively cofibered, cartesian stack over $X$. \end{lemma} \begin{proof} We describe the pushforward functor associated to morphisms of $X\hAlgMod$. Let $Z$ be an extension of $\bF_\bullet(A)$ by $\tI$ for some $X$-algebra $A$ and $A$-module $I$. Let $(A,I) \rightarrow (B,J)$ be a homomorphism in $X\hAlgMod$ and let $u : \et(B) \rightarrow \et(A)$ be the corresponding morphism of \'etale sites. Then $u^\ast Z \tensor_{u^\ast \tA} \tB$ is an extension of $u^\ast \bF_\bullet(A) \tensor_{\tA} \tB = \bF_\bullet(B)$ by $u^\ast \tI \tensor_{u^\ast \tA} \tB$ (by the flatness of $\bF(A)_0$). Pushing out this extension by the map $u^\ast \tI \tensor_{u^\ast \tA} \tB \rightarrow \tJ$, we get the extension we are looking for. We also check that $\Psi_{\bF_\bullet}$ is fibered over $X\hAlg$. Let $\varphi : A \rightarrow B$ be a morphism of $X$-algebras and $u : \et(B) \rightarrow \et(A)$ the corresponding morphism of \'etale sites. Suppose that $Z$ is an extension of $\bF_\bullet(B)$ by $\tJ$ for some $B$-module $J$. Then since $\tJ$ is quasi-coherent, $u_\ast Z$ is an extension of $u_\ast \bF_\bullet(B)$ by $u_\ast \tJ = \tJ_{[\varphi]}$. Pulling this extension back via the map $\bF_\bullet(A) \rightarrow u_\ast \bF_\bullet(B)$ gives the desired object of $\Psi_{\bF_\bullet}(A,J_{[\varphi]})$. To see that $\Psi_{\bF_\bullet}$ forms a stack, it's enough to remark that $\bF_\bullet$ forms an \'etale sheaf of complexes of $\cO_X$-modules and extensions of $\cO_X$-modules satisfy \'etale descent. The left exactness is well-known and we omit the proof. \end{proof} \begin{definition} \label{def:rep-glob} If $\sE$ is an additively cofibered stack over $X$ then we call $\sE$ \emph{representable} if it is isomorphic to a stack of the form $\Psi_{\bF_\bullet}$ for some chain complex of flat modules on $X$. \end{definition} Note that if $\sE$ is represented by the complex $\bF_\bullet$ in the sense of Definition~\ref{def:rep-glob} then for each $X$-algebra $A$, the additively cofibered category $\sE(A)$ on $A\hMod$ is represented by $\bF_\bullet(A)$, in the sense of Definition~\ref{def:rep-loc}. \subsubsection*{Morphisms of complexes and additively cofibered categories} Let $\bE_\bullet \rightarrow \bF_\bullet$ be a morphism of $2$-term complexes. This induces a map $\Psi_{\bF_\bullet} \rightarrow \Psi_{\bE_\bullet}$ by pullback. The following lemma says that the functor which takes a complex to its associated additively cofibered category is $2$-categorially fully faithful. It is essentially an example of the $2$-categorical Yoneda lemma. \begin{lemma} \label{lem:morphs} Let $\bE_\bullet$ and $\bF_\bullet$ be $2$-term chain complexes in degrees $[-1,0]$ in an abelian category $\sC$. Let $\sE = \Psi_{\bE_\bullet}$ and $\sF = \Psi_{\bF_\bullet}$ be the associated additively cofibered categories over $\sC$. \begin{enumerate}[label=(\roman{*}), ref=Lemma~\ref{lem:morphs}~(\roman{*})] \item \label{lem:adj} There is an equivalence of categories \begin{equation*} \Hom(\sF, \sE) \simeq \Psi_{\bE_\bullet}(\bF_\bullet[-1]). \end{equation*} \item There is a bijection \begin{equation*} \Hom_{D(\sC)}(\bE_\bullet, \bF_\bullet) \rightarrow \Hom(\sF, \sE) / \text{isom.} \end{equation*} \item \label{lem:psi-equiv} If $\bE_\bullet \rightarrow \bF_\bullet$ is a quasi-isomorphism then the induced map $\Psi_{\bF_\bullet} \rightarrow \Psi_{\bE_\bullet}$ is an equivalence. \end{enumerate} \end{lemma} \begin{proof} See \cite[Corollaire~6.13]{ccct}. \end{proof} Let $u, v : \bE_\bullet \rightarrow \bF_\bullet$ be two homomorphisms of complexes in an abelian category $\sC$ and let $h$ be a chain homotopy from $u$ to $v$. Then $h$ induces an isomorphism between the two induced maps $\Psi_{\bF_\bullet} \rightarrow \Psi_{\bE_\bullet}$. As an object of $\Psi_{\bE_\bullet}(\bF_\bullet[-1])$, the map $u$ can be described by a diagram \begin{equation*} \xymatrix{ & & & \bE_1 \ar[d]^d \ar[dl]_{\bigl( \begin{smallmatrix} d \\ u_1 \end{smallmatrix} \bigr)} \\ 0 \ar[r] & \bF_1 \ar[d]_{-d} \ar[r]^{\bigl( \begin{smallmatrix} 0 \\ -\id \end{smallmatrix} \bigr)} & \bE_0 \times \bF_1 \ar[dl]^{( \begin{smallmatrix} u_0 & d \end{smallmatrix} )} \ar[r]_{( \begin{smallmatrix} \id & 0 \end{smallmatrix} )} & \bE_0 \ar[r] & 0 . \\ & \bF_0 } \end{equation*} Note the sign on the differential $\bF_1 \rightarrow \bF_0$ because of the shift. There is of course a similar diagram with $u$ replaced by $v$ describing the object of $\Psi_{\bE_\bullet}(\bF_\bullet[-1])$ induced from $v$. Let $\mu$ and $\nu$ be the corresponding objects of $\Psi_{\bE_\bullet}(\bF_\bullet[-1])$. The chain homotopy $h$ gives us a map \begin{equation*} \bigl( \begin{smallmatrix} \id & 0 \\ h & \id \end{smallmatrix} \bigr) : \bE_0 \times \bF_1 \rightarrow \bE_0 \times \bF_1 . \end{equation*} One checks that this gives a morphism of commutative diagrams, and therefore gives an isomorphism between $\mu$ and $\nu$. It follows immediately from the definition that the composition of chain homotopies (by addition) is carried by this construction to the composition of morphisms in $\Psi_{\bE_\bullet}(\bF_\bullet[-1])$. \subsection{The cotangent complex} \label{sec:cc} Assume that $X \rightarrow Y$ is of Deligne--Mumford type and $Y$ is algebraic. There is therefore a relative cotangent complex $\bL_{X/Y}$ and we can identify $\sN_{X/Y}(C, J) = \Ext(f^\ast \bL_{X/Y}, J)$,\marg{reference} where $f : \et(C) \rightarrow \et(X)$ is the morphism of \'etale sites induced from the $X$-algebra structure on $C$. We likewise have $\Exal_X(C, f, J) = \Ext(\bL_{C/X}, J)$ and $h^\ast \Exal_Y(C, f, J) = \Ext(\bL_{C/Y}, J)$. Moreover, the exact sequence \begin{equation*} 0 \rightarrow \Exal_X(C, J) \rightarrow h^\ast \Exal_Y(C, J) \rightarrow \sN_{X/Y}(C, J) \end{equation*} coincides via these identifications with the exact sequence \begin{equation*} 0 \rightarrow \Ext(\bL_{C/X}, J) \rightarrow \Ext(\bL_{C/Y}, J) \rightarrow \Ext(f^\ast \bL_{X/Y}, J) \end{equation*} associated to the exact triangle \begin{equation*} f^\ast \bL_{X/Y} \rightarrow \bL_{C/Y} \rightarrow \bL_{C/X} \rightarrow f^\ast \bL_{X/Y}[1] . \end{equation*} \begin{proposition} Suppose $\bE_\bullet \rightarrow \bL_{X/Y}$ is a morphism of chain complexes that induces an isomorphism $H_0(\bE_\bullet) \rightarrow H_0(\bL_{X/Y})$ and a surjection $H_1(\bE_\bullet) \rightarrow H_1(\bL_{X/Y})$. Then the induced map $\sN_{X/Y} \rightarrow \sE$, where $\sE = \Ext(\bE_\bullet, -)$, is an obstruction theory. \end{proposition} \begin{proof} We only need to check that it is fully faithful. Since these are cofibered categories whose fibers are abelian $2$-categories, it's enough to check that the map induces an isomorphism on the automorphism group of the identity elements and is injective on isomorphism classes. But the map on automorphism groups of the identity elements (resp.\ on isomorphism classes) can respectively be identified with the maps \begin{gather*} u_p : \Ext^p(\bL_{X/Y}, J) \rightarrow \Ext^p(\bE_\bullet, J) . \end{gather*} for $p = 0$ (resp.\ for $p = 1$). Let $\bF_\bullet$ be the cone of $\bE_\bullet \rightarrow \bL_{X/Y}$. This is quasi-isomorphic to a chain complex concentrated in degrees $2$ and higher. Therefore $\Hom(\bF_\bullet, J) = \Ext^1(\bF_\bullet, J) = 0$, so by the long exact sequence \begin{multline*} \Hom(\bF_\bullet, J) \rightarrow \Hom(\bL_{X/Y}, J) \rightarrow \Hom(\bE_\bullet, J) \rightarrow \\ \rightarrow \Ext^1(\bF_\bullet, J) \rightarrow \Ext^1(\bL_{X/Y}, J) \rightarrow \Ext^1(\bE_\bullet, J) \end{multline*} we deduce that $u_p$ is an isomorphism for $p = 0$ and injective for $p = 1$. \end{proof} \begin{corollary} \label{cor:BF1} Any obstruction theory in the sense of Behrend--Fantechi gives rise to an obstruction theory in the sense of Definition~\ref{def:obs}. \end{corollary} \begin{corollary}[of Lemma~\ref{lem:morphs}] \label{cor:BF2} Suppose that $X \rightarrow Y$ is a morphism of Deligne--Mumford type and $\sN_{X/Y} \rightarrow \sE$ is an obstruction theory for $X$ over $Y$. If the additively cofibered stack $\sE$ is representable by a flat chain complex $\bE_\bullet$ on $X$ with coherent cohomology then the obstruction theory $\sN_{X/Y} \rightarrow \sE$ is induced from an obstruction theory $\bE_\bullet \rightarrow \bL_{X/Y}$ in the sense of Behrend--Fantechi. \end{corollary} \begin{proof} Lemma~\ref{lem:morphs} guarantees that the map $\sN_{X/Y} \rightarrow \sE$ is induced from a map $\bE_\bullet \rightarrow \bL_{X/Y}$ where $\bE_\bullet$ is concentrated in degrees $\leq 0$. Since we have assumed the cohomology of $\bE_\bullet$ is coherent, $\bE_\bullet$ satisfies Condition~($\star$) of \cite[Definition~2.3]{BF}. We must check that the map $\bE_\bullet \rightarrow \bL_{X/Y}$ induces an isomorphism on $H^0$ and a surjection on $H^{-1}$. This is a local condition, so we assume that $X = \Spec A$ is affine and replace the complex of sheaves $\bL_{X/Y}$ with a complex of $A$-modules that is quasi-isomorphic to $\bL_{X/Y}$ and do the same for $\bE_\bullet$. The condition we wish to verify is equivalent to the condition that for every $A$-modules $J$, the map \begin{equation} \label{eqn:27} \Ext^p(\bL_{X/Y}, J) \rightarrow \Ext^p(\bE_\bullet, J) \end{equation} be an isomorphism for $p = 0$ and an injection for $p = 1$ (by, for example, \cite[Lemma~A.1.1]{ACW}). But we may identify $\Ext^0(\bL_{X/Y}, J)$ (resp.\ $\Ext^0(\bE_\bullet, J)$) as the automorphism group of the identity in $\sN_{X/Y}(A, J)$ (resp.\ $\sE(A,J)$) and $\Ext^1(\bL_{X/Y}, J)$ (resp.\ $\Ext^1(\bE_\bullet, J)$) as the group of isomorphism classes in $\sN_{X/Y}(A,J)$ (resp.\ $\sE(A,J)$). The full faithfulness of $\sN_{X/Y} \rightarrow \sE$ implies the required facts about the maps~\eqref{eqn:27}. \end{proof} \subsubsection*{Virtual fundamental classes} \label{sec:vfc} Let $\sE$ be an additively cofibered, left exact, cartesian category over $X$. For each $X$-algebra $A$, let $M_A$ be the $A$-module $A$ itself. The collection of pairs $(A, M_A)$ constitutes a cocartesian section of $X\hAlgMod$ over $X\hAlg$, and therefore by pullback via this section we obtain a stack $\fE$ on $X$. This is an $A$-module stack (the analogue of an abelian group stack, or ``Picard stack,'' for $A$-modules). In particular, it has an action of $\bA^1$, in the sense of \cite[Definition~1.5]{BF}. If $X \rightarrow Y$ is a morphism of Deligne--Mumford type, write $\fN_{X/Y}$ for the stack obtained in this way from $\sN_{X/Y}$. Suppose that $\sE$ is an obstruction theory for $X$ over $Y$. Then we obtain a map $\fN_{X/Y} \rightarrow \fE$ from the map $\sN_{X/Y} \rightarrow \sE$. This map is an embedding, by Corollary~\ref{cor:ins}. Recall that there is a canonical embedding of the intrinsic normal cone $\fC$ in $\fN$ \cite[Definition~3.10]{BF}. This induces an embedding $\fC \subset \fE$. If $\fE$ is a vector bundle stack then we obtain a virtual fundamental class $[X/Y]^{\vir}_{\fE}$ by intersecting $\fC$ with the zero locus in $\fE$. \subsection{Perfection} \label{sec:perf} \begin{definition} \label{def:lfp} \begin{enumerate} \item We say that a left exact additively cofibered category $\sE$ over $X\hAlgMod$ is \emph{locally of finite presentation} if the natural map \begin{equation*} \varinjlim_j \sE(A_j, I_j) \rightarrow \sE(A, I) \end{equation*} is an equivalence whenever $\set{(A_j, I_j)}_j$ is a filtered system of objects of $X\hAlgMod$. \item A left exact additively cofibered category $\sE$ over $X\hAlgMod$ is called \emph{quasi-perfect} if the stack of $\cO_X$-modules $\fE$, whose value on an $X$-algebra $A$ is $\sE(A, A)$, is a vector bundle stack \cite[Definition~1.9]{BF} over $X$. \item A left exact additively cofibered category $\sE$ is called \emph{perfect} if it is locally of finite presentation and quasi-perfect. \end{enumerate} \end{definition} \begin{lemma} Suppose that $\fE$ is a vector bundle stack on a Deligne--Mumford stack $X$. Then there is a morphism of sheaves of flat $\cO_X$-modules $\bF_1 \rightarrow \bF_0$ such that $\fE = [\bF_0 / \bF_1]$. \end{lemma} \begin{proof} By definition, there is, \'etale-locally in $X$, a surjective linear map from a vector bundle onto $\fE$. Choose an \'etale cover of $X$ such that on each $U$ in the cover there is a surjection $V_U \rightarrow \fE_U$. Take $\bF_0$ to be the direct sum $\sum_{i : U \rightarrow X} i_! V_U$ and let $\bF_1 = \bF_0 \fp_{\fE} e$ be the kernel of the map $\bF_0 \rightarrow \fE$. Then $\bF_0$ and $\bF_1$ are sheaves of $\cO_X$-modules and $\fE = [\bF_0 / \bF_1]$ since $\bF_0$ surjects onto $\fE$. To check that $\bF_0$ is flat, it suffices to check that each $i_! V$ is flat if $V$ is. We need to see that $W \mapsto i_! V \tensor W$ is left exact. But $i_! V \tensor W = i_!(V \tensor i^\ast W)$ and $W \mapsto i_!(V \tensor i^\ast W)$ is the composition of the exact functors $i^\ast$, $V \tensor (-)$, and $i_!$. We check that $\bF_1$ is also flat by showing that $\uTor_1(\bF_1, J) = 0$ for all sheaves of $\cO_X$-modules $J$. The complex $\bF_\bullet$ has perfect amplitude in $[-1,0]$, so $\uTor_i(\bF_\bullet, -)$ vanishes for $i \not\in [0,1]$. On the other hand, we have an exact sequence \begin{equation*} \uTor_2(\bF_\bullet, J) \rightarrow \uTor_1(\bF_1, J) \rightarrow \uTor_1(\bF_0, J) \end{equation*} for any sheaf of $\cO_X$-modules $J$. As just noted, the first term of this sequence vanishes because $\bF_\bullet$ has perfect amplitude in $[-1,0]$ and the last term vanishes because $\bF_0$ was demonstrated above to be flat. \end{proof} \begin{proposition} \label{prop:vb} If $\fE$ is a vector bundle stack on $X$ then there is a complex $\bE_\bullet$ of flat $\cO_X$-modules in degrees $[-1,0]$ and equivalences $\fE(A) \simeq \Psi_{\bE_\bullet}(A)$ for each $X$-algebra $A$, compatible with the covariance of both terms with respect to $A$. \end{proposition} \begin{proof} By Corollary~\ref{cor:vbdual}, $\fE^\vee$ is also a vector bundle stack. Therefore by the lemma there is a $2$-term complex of flat $\cO_X$-modules $\bE_\bullet$ in degrees $[-1,0]$ such that $\fE^\vee = [\bE_0 / \bE_1]$ and we have $(\fE^\vee)^\vee \simeq \fE$ (Corollary~\ref{cor:vbdd}). On the other hand, a section of $(\fE^\vee)^\vee$ over $\Spec A$, for $A$ an $X$-algebra, can be identified with a morphism of $\tA$-module stacks $\fE^\vee \rightarrow B \tA$. By left exactness, that is the same as a morphism $\bE_0 \tensor_{\cO_X} \tA \rightarrow B \tA$ and an isomorphism between the induced map $\bE_1 \tensor_{\cO_X} \tA \rightarrow B \tA$ and the zero map. In other words, it is a commutative diagram \begin{equation*} \xymatrix{ & & & \bE_1 \tensor_{\cO_X} \tA\ar[d] \ar[dl] \\ 0 \ar[r] & \tA \ar[r] & Z \ar[r] & \bE_0 \tensor_{\cO_X} \tA \ar[r] & 0 } \end{equation*} in which the second row is exact (cf.\ the proof of Lemma~\ref{lem:vbcx} in the appendix). But to give such a diagram is the same as to give an object of $\Psi_{\bE_\bullet}(A, A)$, by definition. \end{proof} \begin{proposition}\label{prop:perf} Suppose that $X$ is a Deligne--Mumford stack and $\sE$ is a cartesian, left exact, additively cofibered category on $X\hAlgMod$ and that the associated abelian cone stack $\fE$ of $\sE$ is a vector bundle stack. Assume also that $\sE$ is locally of finite presentation and $X$ is locally of finite presentation over $Y$. Then there is a perfect complex $\bE_\bullet$, of perfect amplitude in $[-1,0]$, and an equivalence of fibered categories \begin{equation*} \sE \simeq \Psi_{\bE_\bullet} \end{equation*} that is determined, up to unique isomorphism, by $\fE$. \end{proposition} \begin{proof} The first step will be to construct the map $\Psi_{\bE_\bullet} \rightarrow \sE$. Then we will show it is an equivalence. By Proposition~\ref{prop:vb}, for $\fE$ to be a vector bundle stack means that the cofibered category obtained by restricting $\sE$ to pairs $(A, I)$ where $I$ is isomorphic to $A$ is equivalent to the cofibered category obtained by restricting $\Psi_{\bE_\bullet}$ to the same subcategory, for some chain complex $\bE_\bullet$ of $\cO_X$-modules of perfect amplitude in $[-1,0]$. Since both $\Psi_{\bE_\bullet}$ and $\sE$ are additively cofibered, this equivalence can be extended in an essentially unique way to the category of pairs $(A,I)$ where $I$ is a free $A$-module. Since both $\Psi_{\bE_\bullet}$ and $\sE$ are stacks over $X\hAlg$, this equivalence can be further extended in an essentially unqiue way to the category of $(A,I)$ where $I$ is a \emph{locally free} $A$-module. Since both are left exact, this equivalence extends even to the cofibered categories $\Psi_{\bE_\bullet}$ and $\sE$ on pairs $(A,\bI^\bullet)$ where $\bI^\bullet$ is a cochain complex of locally free $A$-modules. The proposition asserts that the isomorphism between $\sE$ and $\Psi_{\bE_\bullet}$ is unique, which means that if the proposition is proved locally, the local statements will automatically glue to a global statement. That is, to construct the maps $\Psi_{\bE_\bullet}(A,I) \xrightarrow{\sim} \sE(A, I)$ is a local problem on the \'etale site of $A$, provided that we also demonstrate they are uniquely determined and compatible with further localization. We can therefore work locally and assume (using \cite[Corollary~4.3]{perf}) that there is a quasi-isomorphism $\tbF_\bullet \rightarrow \bE_\bullet$ where $\bF_\bullet$ is a $2$-term complex of \emph{free} $A$-modules, with $F_i = 0$ for $i \not= 0,1$, and $\tbF_\bullet$ is its associated complex of sheaves. The perfect complex $\bF_\bullet$ is not unique, however, so we will to argue later that our constructions depend on $\bF_\bullet$ only up to unique isomorphism. Now that we are working locally, say over a ring $A$, we can use the fact that $\sE$ and $\Psi_{\bE_\bullet}$ are cartesian to reduce the problem to constructing an equivalence between their restrictions to the category of $A$-modules. Indeed, if $\varphi : A \rightarrow B$ is a homomorphism of $X$-algebras and $J$ is a $B$-module, then $\Ext_B(\bE_\bullet(B), J) \rightarrow \Ext_A(\bE_\bullet(A), J_{[\varphi]})$ is an equivalence, as is $\sE(B,J) \rightarrow \sE(A,J_{[\varphi]})$. Accordingly, we now drop reference to the ring $A$ from the notation, and simply write $\Psi_{\bE_\bullet}(I)$, $\Ext(\bE_\bullet, \tI)$ and $\sE(I)$ in place of $\Psi_{\bE_\bullet}(A,I)$, $\Ext_A(\bE_\bullet(A), \tI)$ and $\sE(A,I)$ for an $A$-module $I$. There is an object $\xi \in \Psi_{\bE_\bullet}(\bF_\bullet) = \Ext(\bE_\bullet, \tbF_\bullet[-1])$, well-defined up to unique isomorphism, corresponding to the identity map $\id_{\bF_\bullet}$ via the equivalence (\ref{lem:psi-equiv}): \begin{equation*} \Psi_{\bE_\bullet}(\bF_\bullet[-1]) \rightarrow \Psi_{\bF_\bullet}(\bF_\bullet[-1]) . \end{equation*} Let $\eta$ be the image of $\xi$ via the equivalence \begin{equation*} \Psi_{\bE_\bullet}(\bF_\bullet[-1]) \rightarrow \sE(\bF_\bullet[-1]). \end{equation*} By~\ref{lem:adj}, this extends, uniquely up to unique isomorphism, to a cocartesian section of $\sE$ over $\Psi_{\bF_\bullet}$. We have therefore constructed a morphism $\Psi_{\bF_\bullet} \rightarrow \sE$. By composition with the equivalence $\Psi_{\bE_\bullet} \rightarrow \Psi_{\bF_\bullet}$, we obtain a map \begin{equation}\label{eqn:7} \Psi_{\bE_\bullet} \rightarrow \sE \end{equation} again determined uniquely up to unique isomorphism from $\bF_\bullet \rightarrow \bE_\bullet$. In the above discussion, every choice was determined uniquely up to unique isomorphism except for the map $\tbF_\bullet \rightarrow \bE_\bullet$. In fact, this map is uniquely determined up to contractible ambiguity in a $2$-category of $2$-term chain complexes, but instead of using this we will verify explicitly that the definition of the map $\Psi_{\bE_\bullet} \rightarrow \sE$ depends on the choice of the quasi-isomorphism $\tbF_\bullet \rightarrow \bE_\bullet$ only up to unique isomorphism. Suppose that we are given two quasi-isomorphisms $u : \tbF_\bullet \rightarrow \bE_\bullet$ and $v : \tbF'_\bullet \rightarrow \bE_\bullet$ where $\bF_\bullet$ and $\bF'_\bullet$ are chain complexes of free $A$-modules, concentrated in degrees $[-1,0]$. Then by \cite[4.2]{perf}, there is a quasi-isomorphism $w : \bF_\bullet \rightarrow \bF'_\bullet$ and a chain homomotpy $h$ between $u$ and the induced map $v \tw : \tbF_\bullet \rightarrow \bE_\bullet$ (recall that $\tw : \tbF_\bullet \rightarrow \tbF'_\bullet$ is the morphism of complexes of sheaves associated to $w$). Suppose that $w' : \bF_\bullet \rightarrow \bF'_\bullet$ is another morphism and that $h'$ is a chain homotopy connecting $v\tw'$ and $u$. Then $h - h'$ is a chain homotopy between $v(\tw - \tw')$ and the zero map. \begin{lemma} Suppose that $w : \bF_\bullet \rightarrow \bF'_\bullet$ is a morphism of $2$-term complexes of free $A$-modules and $v : \tbF'_\bullet \rightarrow \bE_\bullet$ is a quasi-isomorphism of $\tA$-modules on $\et(A)$. Then any chain homotopy $v\tw \simeq 0$ is induced from a \emph{unique} chain homotopy $w \simeq 0$. \end{lemma} \begin{proof} First we note that chain homotopies between $w$ and zero form a pseudo-torsor under $\Hom(\bF_\bullet, \bF'_\bullet[-1]) = \Hom(H_0(\bF_\bullet), H_1(\bF'_\bullet))$. Analogously, chain homotopies between $v \tw$ and zero form a torsor under $\Hom(\tbF_\bullet, \bE_\bullet[-1]) = \Hom(H_0(\bF_\bullet)^\sim, H_1(\bE_\bullet))$. The map \begin{equation*} \Hom(H_0(\bF_\bullet), H_1(\bF'_\bullet)) \rightarrow \Hom(H_0(\bF_\bullet)^\sim, H_1(\bE_\bullet)) \end{equation*} is a bijection because $v$ is a quasi-isomorphism. Therefore if $h : vw \simeq 0$ is a chain homotopy in $\Hom(\bF_\bullet, \bE_\bullet)$, there is at most one lift of $h$ to a chain homotopy between $v$ and $0$ in $\Hom(\bF_\bullet, \bF'_\bullet)$. Now we argue that a chain homotopy lifting $h$ must exist. Since $v$ is a quasi-isomorphism and $vw$ induces the zero map on homology, $w$ also induces the zero map on homology. We therefore have a commutative diagram with exact rows: \begin{equation*} \xymatrix{ 0 \ar[r] & H_1(\bF_\bullet) \ar[r] \ar[d]_0 & \bF_1 \ar[r]^d \ar[d]_{w_1} & \bF_0 \ar@{-->}[dl]_{g} \ar[d]^{w_0} \ar[r] \ar[dr]^0 & H_0(\bF_\bullet) \ar[r] \ar[d]^0 & 0 \\ 0 \ar[r] & H_1(\bF'_\bullet) \ar[r] & \bF'_1 \ar[r]_d & \bF'_0 \ar[r] & H_0(\bF'_\bullet) \ar[r] & 0 } \end{equation*} Since $\bF_0$ is free, there is certainly a dashed arrow $g$ such that $d g = w_0$. The failure of this map to be a chain homotopy is measured be the difference $g d - w_1$. However, $d g d - dw_1 = w_0 d - w_0 d = 0$, so the map $g d - w_1$ factors through a map $\bF_1 \rightarrow H_1(\bF'_\bullet)$. Moreover, this map vanishes on $H_1(\bF_\bullet) \subset \bF_1$, so it factors through a map $\bF_1 / H_1(\bF_\bullet) \rightarrow H_1(\bF'_\bullet)$. We would like to extend this map to a map $\bF_0 \rightarrow H_1(\bF'_\bullet)$, for then we could subtract this map from $g$ to get the chain homotopy we are looking for. The obstruction to the existence of such a lift lies in $\Ext^1(H_0(\bF_\bullet), H_1(\bF'_\bullet))$ and does not depend on the choice of the original map $g$. But the map \begin{equation*} \Ext^1(H_0(\bF_\bullet), H_1(\bF'_\bullet)) \rightarrow \Ext^1(H_0(\tbF_\bullet), H_1(\bE_\bullet)) \end{equation*} is an isomorphism (because $v$ is a quasi-isomorphism) and by construction, it sends the obstruction to the existence of a chain homotopy $w \simeq 0$ to the obstruction to the existence of a chain homotopy $vw \simeq 0$. The latter obstruction is zero by assumption, and therefore so is the former. \end{proof} By the lemma, there is a \emph{unique} chain homotopy $w \simeq w' : \bF_\bullet \rightarrow \bF'_\bullet$. This induces a uniquely determined isomorphism $w_\ast \xi \simeq w'_\ast \xi \in \Psi_{\bE_\bullet}(\bF'_\bullet[-1])$ and therefore a uniquely determined isomorphism between the induced objects of $\sE(\bF'_\bullet)$. This proceeds to give an isomorphism between the induced cocartesian sections of $\sE$ over $\Psi_{\bF'_\bullet}$, and therefore between the induced maps $\Psi_{\bE_\bullet} \rightarrow \sE$. \vskip .5cm For the rest of this proof, we write $\sF$ for the additively cofibered stack $\Psi_{\bE_\bullet}$. To demonstrate that the map $\sF \rightarrow \sE$ is an equivalence is a local problem, so we are free to assume that $X$ is representable by a commutative ring $A$ and $\bE_\bullet$ is representable by a $2$-term complex of free $A$-modules. We note that because $X$ is assumed to be locally of finite presentation, any $X$-algebra $A$ admits a map from an $X$-algebra $A_0$ whose underlying commutative ring is finitely generated over $\bZ$. Likewise, the local finite presentation of $\sE$ implies that any object of $\sE(A, I)$ is induced from some object of $\sE(A_1, I_1)$ where $A_1$ is an $A_0$-algebra of finite type and $I_1$ is an $A_1$-module of finite type. Moreover, by the same argument, any two such representations of an object of $\sE(A,I)$ as objects over finite type objects of $\ComRngMod$ can be compared over a finite type object. It follows that to demonstrate $\sF \rightarrow \sE$ is an equivalence, it suffices to check that $\sF(A, I) \rightarrow \sE(A,I)$ is an equivalence whenever $A$ is an $X$-algebra whose underlying commutative ring is of finite type over $\bZ$ and $I$ is an $A$-module of finite type. In particular, we can assume that $A$ is noetherian and therefore that $I$ admits a filtration of finite length whose associated graded module is a direct sum of modules $A/\frkp$ where $\frkp$ is a prime ideal of $A$ \cite[Theorem~6.4]{M2}. We note furthermore that $\sF$ is an \emph{exact} additively cofibered category. Applying the following lemma inductively completes the proof. \begin{lemma} Suppose that $\sF \rightarrow \sE$ is a morphism of left exact additively cofibered categories over an abelian category $\sC$ and that \begin{equation*} 0 \rightarrow I \rightarrow J \rightarrow K \rightarrow 0 \end{equation*} is an exact sequence in $\sC$ such that $\sF(I) \rightarrow \sE(I)$ and $\sF(K) \rightarrow \sE(K)$ are isomorphisms. Assume also that $\sF$ is right exact. Then $\sF(J) \rightarrow \sE(J)$ is also an isomorphism. \end{lemma} We have to show that $\sF(J) \rightarrow \sE(J)$ is fully faithful and essentially surjective. For an additively cofibered category $\sG$, write $\sG_1(L)$ and $\sG_0(L)$, respectively, for the isomorphism classes in $\sG(L)$ and the automorphism group of the identity object of $\sG(L)$. We use the exact sequence \cite[(2.5.2)]{ccct} associated to $\sF$ and $\sE$ (note, however, that our notation differs: in loc.\ cit.\ $\sF^0$ denotes the isomorphism group of the identity and $\sF^1$ denotes the set of isomorphism classes). The rows of the commutative diagram \begin{equation*} \xymatrix@R=10pt{ 0 \ar[r] & \sF_1(I) \ar[r] \ar[d] & \sF_1(J) \ar[r] \ar[d] & \sF_1(K) \ar[r] \ar[d] & \sF_0(I) \ar[d] \\ 0 \ar[r] & \sE_1(I) \ar[r] & \sE_1(J) \ar[r] & \sE_1(K) \ar[r] & \sE_0(I) } \end{equation*} are exact, so the $5$-lemma implies the faithfulness; the exactness of the rows of \begin{equation*} \xymatrix@R=10pt{ \sF_1(K) \ar[r] \ar[d] & \sF_0(I) \ar[r] \ar[d] & \sF_0(J) \ar[r] \ar[d] & \sF_0(K) \ar[d] \ar[r] & 0 \\ \sE_1(K) \ar[r] & \sE_0(I) \ar[r] & \sE_0(J) \ar[r] & \sE_0(K) } \end{equation*} and a version of the $5$-lemma implies that the functor is full and essentially surjective. Note that the surjectivity of $\sF_0(J) \rightarrow \sF_0(K)$ comes from the exactness of $\sF$. \end{proof} \section{Obstruction groups} \label{sec:obgrp} Let $\sE$ be an obstruction theory for a morphism of homogeneous stacks $h : X \rightarrow Y$. For each $X$-algebra $A$ and each $A$-module $I$, write $\sE_0(A, I)$ for the abelian group of isomorphism classes in $\sE(A,I)$ and $\sE_1(A,I)$ for the abelian group of automorphisms of the identity section of $\sE(A,I)$ (cf.\ \cite[Section~1.6]{ccct}, but note the difference in the indexing). \begin{lemma} Suppose that $A_0$ is an $A$-algebra such that the $A$-module $I$ is induced from an $A_0$-module $I_0$. Then the $A$-module structures on $\sE_0(A,I)$ and $\sE_1(A,I)$ is induced from the $A_0$-module structures on $\sE_0(A_0,I_0)$ and $\sE_1(A_0,I_0)$. \end{lemma} \begin{proof} One may use the fact that $\sE_i(A,I) \simeq \sE_i(A_0,I_0)$ because $\sE(A,I) \simeq \sE(A_0,I_0)$ (see Proposition~\ref{prop:cart}~\ref{cart:3}). Alternatively, note that the $A$-module structure on $\sE_i(A,I)$ is induced from the map $A \rightarrow \End_A(I) \rightarrow \End(\sE_i(A,I))$ by the functoriality of $\sE_i(A,I)$ with respect to $A$-module homomorphisms. The map $A \rightarrow \End_A(I)$ factors through $A_0$ by hypothesis. \end{proof} Suppose that $A$ is an $X$-algebra and $A'$ is a square-zero $Y$-algebra extension of $A$ with ideal $I$. The corresponding object $\xi$ of $h^\ast \Exal_Y$ induces an obstruction $\omega \in \sE(A,I)$, whose isomorphism class in $\sE_0(A,I)$ we denote by $\oomega$. Then $\oomega$ is zero if and only if $\xi$ is induced from some object of $\Exal_X$, meaning that there is an $X$-algebra structure on $A'$ inducing both the given $X$-algebra structure on $A$ and the $Y$-algebra structure on $A'$. In geometric terms, $\oomega$ functions as an obstruction to the existence of a solution to the lifting problem \begin{equation*} \xymatrix{ \Spec A \ar[r] \ar[d] & X \ar[d] \\ \Spec A' \ar[r] \ar@{-->}[ur] & Y . } \end{equation*} This implies that the obstruction groups $\sE_0(A,I)$ satisfy \cite[(2.6)~(ii)]{versal}. The obstruction groups $\sE_0(A,I)$ are also functorial in $(A,I)$ ``in the obvious sense'' (loc.\ cit.) but they need not take finite modules to finite modules (as in \cite[(2.6)~(i)]{versal}) without an extra hypothesis. Should $\oomega$ vanish, the lifts of $\xi$ to $\Exal_X$ are in bijection with the isomorphisms between $\omega$ and a (fixed) zero section of $\sE(A,I)$. Such isomorphisms form a torsor under the automorphism group of the zero section, namely $\sE_1(A,I)$. \subsection{Obstruction sheaves} Suppose that $X \rightarrow Y$ is a morphism of homogeneous stacks and $\sE$ is an obstrution theory for $X$ over $Y$. For each $X$-algebra $A$ and each $A$-module $I$, define $\usE(A,I)$ to be the fibered category on $\et(A)$ whose value on an \'etale $A$-algebra $B$ is $\usE(B, I \tensor_A B)$. Of course, $\usE(A,I)$ is a stack on $\et(A)$ because it is the restriction of a stack from the large \'etale site. Define $\usE_1(A,I)$ and $\usE_0(A,I)$ analogously. \begin{proposition} \label{prop:qcoh} Suppose $X$ is of Deligne--Mumford type over $Y$, let $A$ be an $X$-algebra, and let $I$ be an $A$-module. Then the presheaf $\usE_1(A,I)$ is a quasi-coherent sheaf. \end{proposition} \begin{proof} It is clear that $\sE_1$ is a sheaf since it is the presheaf of automorphisms of a section of a stack. It is therefore a local problem in $A$ to demonstrate that $\sE_1(A,I)$ is quasi-coherent. By definition, we have \begin{equation*} \sE_1(A,I) = e(A,I) \fp_{\sE(A,I)} e(A,I) = e(A,I) \fp_{\Exal_Y(A,I)} \Exal_X(A,I) \end{equation*} where $e$ is a zero section of $\sE$. But $e(A,I) \rightarrow \Exal_Y(A,I)$ is representable by the trivial $Y$-algebra extension of $A$ by $I$. Therefore $\sE_1(A,I)$ consists of the $X$-algebra structures on $A + I$ lifting the trivial $Y$-algebra structure. In geometric terms, it is the set of lifts of \begin{equation*} \xymatrix{ S \ar[r] \ar[d] & X \ar[d] \\ S[\tI] \ar@{-->}[ur] \ar[r] & Y } \end{equation*} where $S = \Spec A$ and $S[\tI] \rightarrow Y$ is the map induced from the map $S \rightarrow Y$ and the retraction $S[\tI] \rightarrow S$. To prove the proposition, we must show that $\sE_1(A, I) \tensor_A B \simeq \sE_1(B, I \tensor_A B)$ whenever $B$ is an \'etale $A$-algebra. By base change, we can assume that $Y$ is representable by a commutative ring $C$ and therefore that $X$ is a Deligne--Mumford stack. If $U$ is \'etale over $X$ and the $X$-algebra structure on $A$ is induced from a $U$-algebra structure on $A$, then $\Exal_X(A,I) = \Exal_U(A,I)$. Since we are free to work locally in $A$, we can therefore replace $X$ by $U$ and assume that $X$ is representable by a $C$-algebra $D$. Let $\varphi : D \rightarrow A$ denote the map giving the $X$-algebra structure on $A$ and let $\psi : D \rightarrow B$ be the induced map. Now we have \begin{equation*} \sE_1(B, I \tensor_A B) = \Hom_D(\Omega_{D/C}, (I \tensor_A B)_{[\psi]}) . \end{equation*} Since $B$ is flat over $A$, we have \begin{equation*} \Hom_D(\Omega_{D/C}, (I \tensor_A B)_{[\psi]}) = \Hom_D(\Omega_{D/C}, I_{[\varphi]}) \tensor_A B = \sE_1(A, I) \tensor_A B \end{equation*} which implies the proposition. \end{proof} Although $\usE_1(A,I)$ is always a sheaf, $\usE_0(A,I)$ generally will not be. However, we have \begin{proposition} The functor $B \mapsto \sE_0(B, I \tensor_A B)$ is a separated presheaf on the \'etale site of $A$. \end{proposition} \begin{proof} In effect, we must show that if $\omega$ is an object of $\sE(A, I)$ that is locally isomorphic to zero then it is globally isomorphic to zero. But if $\omega$ is locally isomorphic to zero, the isomorphisms between $\omega$ and zero form a torsor on the \'etale site of $A$ under the sheaf of abelian groups $\usE_1(A,I)$. By Proposition~\ref{prop:qcoh}, $\usE_1(A,I)$ is quasi-coherent and therefore a torsor under $\usE_1(A,I)$ on $\et(A)$ admits a global section. \end{proof} \begin{corollary} \begin{enumerate}[label=(\roman{*})] \item The map from $\usE_0(A,I)$ to its associated sheaf is injective. \item If $\omega \in \usE_0(A,I)$ is an obstruction to a lifting problem then its image in the associated sheaf of $\usE_0(A,I)$ is also an obstruction to the same lifting problem. \end{enumerate} \end{corollary} \subsection{Li--Tian obstruction theories} Let $M_A$ denote the $A$-module structure on $A$ itself. For each $A$-module $I$, there is a natural map of presheaves of $\tA$-modules on $\et(A)$, \begin{equation} \label{eqn:17} \usE_0(A, M_A) \tensor_A I \rightarrow \usE_0(A, I) , \end{equation} induced by the $A$-linear map $A \rightarrow I$ associated to each element of $I$. The linearity with respect to $I$ follows from the additivity of $\sE_0(A,-)$. The functor $I \mapsto \sE_0(A,M_A) \tensor_A I$ is right exact, so if~\eqref{eqn:17} is an isomorphism for all $I$ then $\sE_0(A, -)$ is right exact. \begin{proposition} If the maps~\eqref{eqn:17} are isomorphisms for all $I$ then $\sE$ is an \emph{exact} additively cofibered category. If $\sE$ is exact and locally of finite presentation then the maps~\eqref{eqn:17} are isomorphisms. \end{proposition} \begin{proof} If $J \rightarrow K$ is a surjection of $A$-modules, then right exactness of $\usE_0(A,-)$ implies that $\usE(A,J) \rightarrow \usE(A,K)$ is locally surjective on isomorphism classes, which is what it means to be right exact. This proves the first claim. Conversely, suppose that $\sE$ is exact, so $\sE_0$ is right exact. Then for any exact sequence \begin{equation} \label{eqn:18} 0 \rightarrow I \rightarrow J \rightarrow K \rightarrow 0 \end{equation} of $A$-modules, we have a commutative diagram \begin{equation*} \xymatrix@R=12pt{ \usE_0(A,M_A) \tensor_A I \ar[r] \ar[d] & \usE_0(A,M_A) \tensor_A J \ar[r] \ar[d] & \usE_0(A,M_A) \tensor_A K \ar[r] \ar[d] & 0 \\ \usE_0(A,I) \ar[r] & \usE_0(A,J) \ar[r] & \usE_0(A,K) \ar[r] & 0 } \end{equation*} with exact rows. By additivity and local finite presentation, the map $\sE_0(A,M_A) \tensor_A J \rightarrow \sE_0(A,J)$ is an isomorphism if $J$ is a free $A$-module. If $K$ is given, we may choose an exact sequence~\eqref{eqn:18} in which $J$ is free. Then the commutative diagram above implies the surjectivity of $\sE_0(A,M_A) \tensor_A K \rightarrow \sE_0(A,K)$. This applied with $K$ replaced by $I$ gives the surjectivity of the leftmost vertical arrow, which is enough to imply that the map $\sE_0(A,M_A) \tensor_A K \rightarrow \sE_0(A,K)$ is a bijection, by the $5$-lemma. \end{proof} Let $\usE_0(A,I)^+$ denote the associated sheaf of $\usE_0(A,I)$ on $\et(A)$. \begin{proposition} If $\sE$ is an exact obstruction theory for $X$ over $Y$ then it induces an obstruction theory for $X$ over $Y$ in the sense of \cite[Definition~1.2]{LT}, such that sheaf of obstruction groups is $\usE_1(A,I)^+$. \end{proposition} \begin{proof} We recall the situation of \cite[Definition~1.2]{LT} (but change some of the letters). Suppose that $S_0 \rightarrow S \rightarrow S'$ is a sequence of closed embeddings of schemes with $S_0 = \Spec A_0$ affine. Assume that the ideal of $S_0$ in $S$ annihilates the ideal of $S$ in $S'$ and write $I$ for the $A_0$-module corresponding to the ideal of $S$ in $S'$. Given a lifting problem \begin{equation*} \xymatrix{ S \ar[r] \ar[d] & X \ar[d] \\ S' \ar@{-->}[ur] \ar[r] & Y } \end{equation*} there is an obstruction to the existence of a lift in $\sE_0(A_0, I)$. Since $\sE_0(A_0, I) \subset \Gamma(\et(A), \usE_0(A,I)^+)$ we can also view the obstruction as living in $\usE_0(A,I)^+$. Since $\sE$ is exact and locally of finite presentation, $\sE_0(A_0, I) = \sE_0(A_0, M_{A_0}) \tensor_{A_0} I$. Since sheafification preserves tensor product, we also get \begin{equation*} \usE_0(A_0, I)^+ = \usE_0(A_0, M_{A_0})^+ \tensor_{A_0} I . \end{equation*} We determine that there is an obstruction class in $\usE_0(A_0, M_{A_0})^+ \tensor_{A_0} I$, as required by~\cite[Definition~1.2]{LT}. \end{proof} Note that an obstruction theory that is represented by a $2$-term complex of free $A$-modules is exact. Therefore any perfect obstruction theory, in the sense of \cite{BF} induces an obstruction theory in the sense of \cite{LT} (which is perfect in the sense of Li--Tian). Moreover, this is the same process by which a Li--Tian obstruction theory is associated to a Behrend--Fantechi obstruction theory in \cite[Section~3]{KKP}. Provided that the complex representing the obstruction theory is globally representable by a $2$-term complex of vector bundles (in order that the Li--Tian obstruction class be defined), \cite[Corollary~1]{KKP} implies that the virtual fundamental class defined by Li and Tian coincides with that defined by Behrend and Fantechi (as in Section~\ref{sec:vfc}). \section{Compatibility} \label{sec:compat} \subsection{Compatible obstruction theories for morphisms of Deligne--Mumford type} \label{sec:compat-dm} We recall that an exact sequence of left exact, additively cofibered categories \begin{equation*} 0 \rightarrow \sE \rightarrow \sF \rightarrow \sG \end{equation*} is a specified cartesian diagram \begin{equation*} \xymatrix{ \sE \ar[r] \ar[d] & \sF \ar[d] \\ e \ar[r] & \sG . } \end{equation*} We say that the sequence is right exact, and we affix an arrow $\sG \rightarrow 0$ at the end of the sequence, to mean that the map $\sF \rightarrow \sG$ is locally essentially surjective. Let $X \xrightarrow{f} Y \xrightarrow{g} Z$ be a sequence of morphisms of homogeneous stacks. We will specify a notion of compatibility between relative obstruction theories for $f$, $g$, and $gf$ that reduces to a familiar notion for obstruction theories in the sense of Behrend--Fantechi. It is easiest to describe compatibility for morphisms of Deligne--Mumford type, so we do that first; then we explain how this can be generalized. \begin{definition} \label{def:compat} Let $X \xrightarrow{f} Y \xrightarrow{g} Z$ be a sequence of morphisms of Deligne--Mumford type. A compatible sequence of relative obstruction theories for this sequence of maps is a commutative diagram \begin{equation} \label{eqn:11}\xymatrix{ 0 \ar[r] & \sN_{X/Y} \ar[r] \ar[d] & \sN_{X/Z} \ar[r] \ar[d] & f^\ast \sN_{Y/Z} \ar[d] \\ 0 \ar[r] & \sE_{X/Y} \ar[r] & \sE_{X/Z} \ar[r] & f^\ast \sE_{Y/Z} \ar[r] & 0 } \end{equation} in which the first row is the canonical sequence of relative instrinsic normal stacks (whose left exactness we leave to the reader), the second row is an exact sequence of left exact additively cofibered stacks, and the map $f^\ast \sN_{Y/Z} \rightarrow f^\ast \sE_{Y/Z}$ is pulled back from an obstruction theory $\sN_{Y/Z} \rightarrow \sE_{Y/Z}$ for $Y$ over $Z$. The map $\sN_{X/Y} \rightarrow \sE_{X/Y}$ is required to be isomorphic to the one induced from the commutative square on the right by the exactness of the two rows (provided it exists the isomorphism is unique). \end{definition} \begin{proposition} Suppose we have a compatible sequence of obstruction theories as in Definition~\ref{def:compat} and the obstruction theories in question are all \emph{perfect}, hence give rise to relative Gysin pullback functors. Then \begin{gather*} (gf)^! = f^! g^! \qquad \text{and} \\ f^! [Y/Z]^{\vir} = [X/Z]^{\vir} . \end{gather*} \end{proposition} \begin{proof} This reduces immediately to \cite[Theorem~4]{Man}. \end{proof} \subsection{Compatible obstruction theories in general} Suppose that $X \xrightarrow{f} Y \xrightarrow{g} Z$ is a sequence of morphisms of homogeneous stacks. We say that relative obstruction theories $\sE_{X/Y}$, $\sE_{X/Z}$, and $\sE_{Y/Z}$ are \emph{compatible} if we are given a commutative diagram \begin{equation} \label{eqn:22} \xymatrix{ \Exal_X \ar[r] \ar[d] & f^\ast \Exal_Y \ar[r] \ar[d] & f^\ast g^\ast \Exal_Z \ar[d] \\ e \ar[r] & \sE_{X/Y} \ar[r] \ar[d] & \sE_{X/Z} \ar[d] \\ & e \ar[r] & f^\ast \sE_{Y/Z} . } \end{equation} in which \begin{enumerate} \item the square in the upper left is the cartesian square associated to the obstruction theory $\sE_{X/Y}$, \item the upper large rectangle is the cartesian square associated to the obstruction theory $\sE_{X/Z}$, \item the large rectange on the right is the cartesian square associated to the obstruction theory $\sE_{Y/Z}$, pulled back via $f^\ast$, \item the square in the lower right is cartesian, and \item the map $\sE_{X/Z} \rightarrow f^\ast \sE_{Y/Z}$ is locally essentially surjective. \end{enumerate} \begin{proposition} Suppose that $X \rightarrow Y \rightarrow Z$ are morphisms of Deligne--Mumford type. If $\sE_{X/Y}$, $\sE_{X/Z}$, and $\sE_{Y/Z}$ are compatible relative obstruction theories in the sense described above then they are also compatible relative obstruction theories in the sense of Section~\ref{sec:compat-dm}. \end{proposition} \begin{proof} The commutative diagram~\eqref{eqn:22} induces a commutative diagram \begin{equation*} \xymatrix@R=10pt{ f^\ast \Exal_Y \ar[r] \ar[d] & f^\ast g^\ast \Exal_Z \ar[r] \ar[d] & g^\ast \Exal_Z \ar[d] \\ \sE_{X/Y} \ar[r] \ar[d] & \sE_{X/Z}\ar[d] \ar[r] & \sE_{Y/Z} \ar[d] \\ X \ar@{=}[r] & X \ar[r] & Y . } \end{equation*} Recall that the restriction of $f^\ast \Exal_Y$ to the small \'etale site of $X$ is the same as the restriction of $\sN_{X/Y}$ to the small \'etale site of $X$. Likewise $f^\ast g^\ast \Exal_Z$ and $\sN_{X/Z}$ have the same restriction to the small \'etale site of $X$ and $g^\ast \Exal_Z$ has the same restriction as $\sN_{Y/Z}$ to the small \'etale site of $Y$. Restricting the maps above to the small sites and then pulling back to the big sites, we obtain a commutative diagram \begin{equation*} \xymatrix@R=10pt{ \sN_{X/Y} \ar[r] \ar[d] & \sN_{X/Z} \ar[r] \ar[d] & \sN_{Y/Z} \ar[d] \\ \sE_{X/Y} \ar[r] \ar[d] & \sE_{X/Z} \ar[r] \ar[d] & \sE_{Y/Z} \ar[d] \\ X \ar@{=}[r] & X \ar[r] & Y } \end{equation*} which induces the required diagram~\eqref{eqn:11}. \end{proof} \subsection{A construction for compatible obstruction theories} Suppose that $X \xrightarrow{f} Y \xrightarrow{g} Z$ is a sequence of morphisms of homogeneous stacks and $\sE_{X/Z}$ and $\sE_{Y/Z}$ are relative obstruction theories. We therefore have cartesian diagrams \begin{equation*} \xymatrix{ \Exal_X \ar[r] \ar[d] & e \ar[d] \\ f^\ast g^\ast \Exal_Z \ar[r] & \sE_{X/Z} } \qquad \xymatrix{ f^\ast \Exal_Y \ar[r] \ar[d] & e \ar[d] \\ f^\ast g^\ast \Exal_Z \ar[r] & \sE_{Y/Z} . } \end{equation*} Suppose that these diagrams can be fit together in a commutative diagram \begin{equation} \label{eqn:23} \xymatrix{ f^\ast \Exal_Y \ar[r] \ar[dr] & f^\ast g^\ast \Exal_Z \ar[r] \ar[dr] & \sE_{X/Z} \ar[d] \\ \Exal_X \ar[u] \ar[ur] \ar[r] & e \ar[r] \ar[ur] & f^\ast \sE_{Y/Z} } \end{equation} over $X\hAlgMod$. Given such a diagram, define $\sE_{X/Y}$ to be the kernel of the map $\sE_{X/Z} \rightarrow f^\ast \sE_{Y/Z}$. That is, define $\sE_{X/Y} = \sE_{X/Z} \fp_{f^\ast \sE_{Y/Z}} e$. \begin{lemma} If $\sE_{X/Z}$ and $\sE_{Y/Z}$ are relative obstruction theories for $X/Z$ and $Y/Z$, respectively, fitting into a commutative diagram~\eqref{eqn:23}, and $\sE_{X/Y}$ is defined as above, then $\sE_{X/Y}$ is naturally equipped with the structure of an obstruction theory for $X$ over $Y$. \end{lemma} \begin{proof} The functoriality properties of $\sE_{X/Y}$ are all deduced from those of $\sE_{X/Z}$ and $f^\ast \sE_{Y/Z}$. The only thing to verify is that $\sE_{X/Y}$ comes equipped with a diagram~\eqref{eqn:6}. Following the upper horizontal arrow across~\eqref{eqn:23}, we get a map $f^\ast \Exal_Y \rightarrow \sE_{X/Z}$ and an isomorphism between the induced map $f^\ast \Exal_Y \rightarrow f^\ast \sE_{Y/Z}$ and the zero map. It therefore induces (uniquely up to unique isomorphism) a map $f^\ast \Exal_Y \rightarrow \sE_{X/Y}$. We therefore obtain a commutative diagram~\eqref{eqn:22}. The lower right square is cartesian by definition and the large rectangle on the right is cartesian by hypothesis; therefore the upper right square is cartesian. The large upper rectangle is also cartesian by hypothesis, so the square in the upper left is cartesian as well, as was required. \end{proof} \section{Examples} \label{sec:ex} \subsection{Smooth morphisms} Recall the definition of $\sT_{X/Y}$ from Section~\ref{sec:tang}. We write $B \sT_{X/Y}$ for the additively cofibered category of $\sT_{X/Y}$-torsors: an object of $B \sT_{X/Y}(A,I)$ is a torsor on $\et(A)$ under the sheaf of groups $\usT_{X/Y}(A,I)$. \begin{proposition} \label{prop:smooth} Suppose that $X$ and $Y$ are algebraic stacks and $X \rightarrow Y$ is a smooth morphism of Deligne--Mumford type. Then $\sN_{X/Y} \simeq B \sT_{X/Y}$. \end{proposition} \begin{proof} It is sufficient to assume that $Y$ is affine (or at least of Deligne--Mumford type) and therefore that $X$ is a Deligne--Mumford stack. We can construct the construct isomorphism on the small \'etale site of $X$. Note in that case that a section of $\sN_{X/Y}(A)$, when $A$ is an \'etale $X$-algebra, is a square-zero extension $A'$ of $A$ as a $Y$-algebra. The lifts of the $Y$-algebra structure on $A'$ to an $X$-algebra structure form a torsor on $\Spec A$ under the quasi-coherent sheaf of groups $T_{X/Y} \tensor_X \tJ$ where $J$ is the ideal of $A$ in $A'$. But $\Gamma(A, T_{X/Y} \tensor_X \tJ) = \sT_{X/Y}(A, J)$, so this gives the desired map $\sN_{X/Y} \rightarrow B \sT_{X/Y}$. It is well-known that this map induces a bijection on morphisms. Since both $\sN_{X/Y}(A,J)$ and $B \sT_{X/Y}(A,J)$ are gerbes, this implies that the map is an equivalence. \end{proof} \subsection{Local complete intersection morphisms} Let $X$ and $Y$ be algebraic stacks and let $X \rightarrow Y$ be a morphism of Deligne--Mumford type. Then $\sN_{X/Y}$ is a relative obstruction theory for $X$ over $Y$, as we saw in Section~\ref{sec:ins}. \begin{proposition} If $X \rightarrow Y$ is a local complete intersection morphism, then $\sN_{X/Y}$ is perfect. \end{proposition} \begin{proof}[First proof.] Remark that $\sN_{X/Y}(A, J) = \Ext(\bL_{X/Y} \tensor_X A, J)$ and $\bL_{X/Y}$ is perfect in degrees $[-1,0]$. \end{proof} \begin{proof}[Second proof.] First note that $X$ is locally of finite presentation over $Y$ and $\sN_{X/Y}$ is locally of finite presentation as an obstruction theory. We can therefore use the criterion of Proposition~\ref{prop:perf}: it suffices to see that the abelian cone stack $\fN_{X/Y}$ is a vector bundle stack. It is sufficient to assume that $Y$ is an affine scheme by base change, and the problem is \'etale local in $X$, so we can also assume that $X$ is affine. We can therefore assume that $X \rightarrow Y$ factors as a closed complete intersection embedding $i : X \rightarrow Z$ followed by a smooth map $Z \rightarrow Y$. We have an exact sequence \begin{equation*} 0 \rightarrow \fN_{X/Z} \rightarrow \fN_{X/Y} \rightarrow i^\ast \fN_{Y/Z} \end{equation*} and $\fN_{X/Z}$ is a vector bundle and $i^\ast \fN_{Y/Z} = i^\ast B T_{Y/Z}$ (Proposition~\ref{prop:smooth}) is a vector bundle stack . Therefore by \cite[Compl\'ement~I.4.11]{perf}, $\fN_{X/Y}$ is a vector bundle stack. \end{proof} \subsection{Stable maps} \label{sec:stable-maps} Let $X \rightarrow V$ be a \emph{smooth}, Deligne--Mumford type morphism of algebraic stacks. Consider the stack $\fM(X/V)$ whose $S$-points are commutative diagrams \begin{equation} \label{eqn:9} \xymatrix{ C \ar[r] \ar[d]_\pi & X \ar[d]^q \\ S \ar[r] & V } \end{equation} where $C / S$ is a family of twisted Deligne--Mumford pre-stable curves. Let $\fM$ be the stack of twisted Deligne--Mumford pre-stable curves. There is a projection $h : \fM(X/V) \rightarrow \fM \times V$. We describe a relative obstruction theory. For the description of the obstruction theory itself, there is no requirement that $C$ be a pre-stable curve; that hypothesis is only necessary to demonstrate that the obstruction theory is perfect. We will pass to the opposite category here and define the relative obstruction theory as a fibered category over the category of affine schemes over $\fM(X/V)$. Begin by defining $\sT_{X/V}^\circ$ as in Section~\ref{sec:tang} (but passing to the opposite category as we work geometrically in this section). If $C$ is an $X$-scheme and $J$ is a quasi-coherent sheaf on $C$ then an object of $\sT_{X/V}^\circ(C,J)$ is a completion of the diagram \begin{equation*} \xymatrix{ C \ar[r] \ar[d] & X \ar[d] \\ C[J] \ar[r] \ar@{-->}[ur] & V } \end{equation*} where $C[J]$ is the trivial square-zero extension of $C$ by $J$ and the map $C[J] \rightarrow V$ is the zero tangent vector (the map factoring through the retraction of $C[J]$ onto $C$). Recall that $B \sT^\circ_{X/V} = \sN^\circ_{X/V}$. Given $S$-point~\eqref{eqn:9} of $\fM(X/V)$ and a quasi-coherent sheaf $J$ on $S$, we obtain a stack $\usN^\circ_{X/V}(C,\pi^\ast J)$ by restricting $\sN^\circ_{X/V}$ to $\et(C)$. Define $\usE^\circ(S,J) = \pi_\ast \usN^\circ_{X/V}(C, \pi^\ast J)$. To show that $\usE^\circ(S,J)$ is the (opposite of the) restriction of an obstruction theory $\sE^\circ$ to $S$ and $J$, we must describe how its functoriality with respect to the variation of $S$ and $J$. It must vary both covariantly and contravariantly with $S$ and covariantly with $J$ (the contravariance in $S$ and covariance in $J$ combine to $\sE$'s being a cofibered category over $\ComRngMod$ and the contravariance with $S$ corresponds to being fibered over $\ComRng$). For the contravariance in $S$ and the covariance in $J$, consider a commutative diagram \begin{equation} \label{eqn:28} \xymatrix{ C_S \ar@/^12pt/[rr]^{\varphi_S} \ar[r]_g \ar[d] & C_T \ar[d] \ar[r]_{\varphi_T} & X \ar[d] \\ S \ar[r]_f & T \ar[r] & V } \end{equation} corresponding to a morphism in $\fM(X/V)$, and a morphism of $\cO_S$-modules $f^\ast I \rightarrow J$. We assume that $S \rightarrow T$ is affine. Since $\sN^\circ_{X/V}$ is an obstruction theory there is a map $\sN^\circ_{X/V}(C_T, \pi_T^\ast I) \rightarrow \sN^\circ_{X/V}(C_S, \pi_S^\ast J)$. Allowing $C_S$ and $C_T$ to vary in their \'etale sites gives a map $\usN^\circ_{X/V}(C_T, \pi_T^\ast I) \rightarrow g_\ast \usN^\circ_{X/V}(C_S, \pi_S^\ast J)$. Then pushing forward by $\pi_T$, we get $\usE^\circ(T, I) \rightarrow f_\ast \usE^\circ(S,T)$ and passing to global sections gives the map $\sE^\circ(T,I) \rightarrow \sE^\circ(S,J)$. For the covariance in $S$, consider again the diagram~\eqref{eqn:28}. We only need to consider the case where $S$ and $T$ are both affine (by definition), but it is actually enough to assume only that $f$ is an affine map. Therefore $g$ is also affine. We have given a $V$-extension of $\varphi_S^\ast \cO_X$ by $\pi_S^\ast J$, we may push forward by $g$ to obtain a $V$-extension of $g_\ast \varphi_S^\ast \cO_X$ by $g_\ast \pi_S^\ast J = \pi_T^\ast f_\ast J$. Pulling back via the map $\varphi_T^\ast \cO_X \rightarrow g_\ast \varphi_S^\ast \cO_X = g_\ast g^\ast \varphi_T^\ast \cO_X$, we get an extension of $\varphi_T^\ast \cO_X$ by $\pi_T^\ast f_\ast J$, which is a section of $\sE^\circ(T, f_\ast J)$, as needed. \begin{remark} In fact, one can demonstrate more generally that if $\sE^\circ$ is an obstruction theory then its natural extension to all schemes (Section~\ref{sec:global}) behaves covariantly with respect to affine morphisms. \end{remark} Finally, we have to construct the cartesian diagram~\eqref{eqn:6} (or really, given that we are working now with the opposite category, a cartesian diagram \begin{equation*} \xymatrix{ \Def_{\fM(X/V)} \ar[r] \ar[d] & e \ar[d] \\ h^\ast \Def_{\fM \times V} \ar[r]^<>(0.5){\omega} & \sE^\circ , } \end{equation*} using the notation of Section~\ref{sec:global}). An object of $h^\ast \Def_{\fM \times V}$ corresponds to a commutative diagram of solid lines \begin{equation} \label{eqn:29} \xymatrix{ C \ar[r] \ar[d] \ar@/^15pt/[rr]^{\varphi} & C' \ar[d] \ar@{-->}[r] & X \ar[d]^q \\ S \ar[r] & S' \ar[r] & V } \end{equation} in which $S'$ is a square-zero extension of $S$, and to lift it to $\Def_{\fM(X/V)}$ means to construct a dashed arrow making the diagram commutative. This immediately translates to the lifting problem \begin{equation} \label{eqn:30} \xymatrix{ C \ar[r] \ar[d] & X \ar[d] \\ C' \ar@{-->}[ur] \ar[r] & V , } \end{equation} which we know is obstructed by the obstruction theory $\sN^\circ_{X/V}$. This gives the map $\Def_{\fM \times V} \rightarrow \sN^\circ_{X/V}$ and it is immediate from the fact that $\sN^\circ_{X/V}$ is an obstruction theory for $X$ over $V$ that $\sE^\circ$ is an obstruction theory for $\fM(X/V)$ over $\fM \times V$. This can be put somewhat more precisely as follows. We obtain the map $h^\ast \uDef_{\fM \times V}(S,J) \rightarrow \usE^\circ(S,J) = \pi_\ast \usN^\circ_{X/V}(C, \pi^\ast J)$ as the composition \begin{equation*} h^\ast \uDef_{\fM \times V}(S,J) \rightarrow \pi_\ast q^\ast \uDef_V(C,\pi^\ast J) \rightarrow \pi_\ast \usN^\circ_{X/V} . \end{equation*} The first arrow sends a map $S' \rightarrow \fM \times V$ extending the one induced from $S \rightarrow \fM(X/V)$ to the diagram~\eqref{eqn:30}. Since $\sN^\circ_{X/V}$ is an obstruction theory for $X$ over $V$, we have \begin{equation*} \Def_X(C, \pi^\ast J) = q^\ast \Def_V(C, \pi^\ast J) \fp_{\usN^\circ_{X/V}(C, \pi^\ast J)} e . \end{equation*} Pushforward is exact, so we get \begin{equation*} \uDef_{\fM \times V}(S,J) \fp_{\usE^\circ(S,J)} e = \uDef_{\fM \times V}(S,J) \fp_{\pi_\ast q^\ast \uDef_V(C, \pi^\ast J)} \uDef_X(C, \pi^\ast J) . \end{equation*} Unwound, the last term is exactly the collection of completions of Diagram~\eqref{eqn:29}. This is exactly what we needed to show that $\sE$ is an obstruction theory. We can also demonstrate that the obstruction theory is perfect using the criterion for perfection proved above, provided $X$ is locally of finite presentation over $V$. In effect, it is sufficient to remark that in this case $\sE$ is locally of finite presentation, and therefore it is enough to prove that $\pi_\ast \varphi^\ast \fN_{X/V}$ ($ = \pi_\ast \usN^\circ_{X/V}(C, \cO_{C})$) is a vector bundle stack. But now we use the facts that $C$ has relative dimension $1$ over $C$, that $\usN^\circ_{X/V}(C,\cO_C) = B \usT^\circ_{X/V}(C, \cO_C)$ is representable by $\varphi^\ast T_{X/V}[1]$, and that $T_{X/V}$ is flat over $S$ to show that $R \pi_\ast \varphi^\ast T_{X/V}[1]$ is perfect in degrees $[-1,0]$.
\section{The NA62 Experiment} The decays $K^+\to\pi^+\nu\bar{\nu}$ and $K_L\to\pi^0\nu\bar{\nu}$ are flavor-changing neutral-current processes whose amplitudes are dominated by $Z$-penguin and box diagrams. Because there are no contributions from long-distance processes with intermediate photons and because the hadronic matrix elements can be obtained from rate and form factor measurements for common $K\to\pi\ell\nu_\ell$ decays, the branching ratios (BRs) for the $K\to\pi\nu\bar{\nu}$ decays can be calculated in the Standard Model (SM) with minimal intrinsic uncertainty (see \Ref{C+11:kaonRev} for a recent review). The BRs for these decays are therefore a sensitive probe of the SM flavor sector and provide constraints on the CKM unitarity triangle that are complementary to those from measurements of $B$-meson decays. On the other hand, the tiny BRs for these decays are notoriously difficult to measure, not least because of the three-body kinematics with two undetectable neutrinos in the final state. At present, the experimental value of the BR for the decay $K^+\to\pi^+\nu\bar{\nu}$ is \EA{1.73}{1.15}{1.05}{-}{10} on the basis of seven detected candidate events \cite{E949+08:Kpnn2}. The goal of NA62, an experiment at the CERN SPS, is to detect $\sim$100 $K^+\to\pi^+\nu\bar{\nu}$ decays with a S/B ratio of 10:1 in two years of data taking beginning in 2013. The experiment is fully described in \Ref{NA62+10:TDD}. The experimental layout is illustrated in \Fig{fig:expt}. \begin{figure} \centering \includegraphics[width=0.9\linewidth]{valencia_fig_expt} \caption{The NA62 experimental layout.} \label{fig:expt} \end{figure} NA62 will make use of a 75-GeV unseparated positive secondary beam with a total rate of nearly 800~MHz, of which $\sim$50 MHz is $K^+$'s. The $K^+$'s are identified by the CEDAR differential Cerenkov counter in the beamline. All 800~MHz of beam particles are tracked by three silicon-pixel tracking detectors (the Gigatracker) located at the achromat just upstream of the vacuum decay volume, providing event-by-event measurements of the $K^+$ trajectory and momentum. The decay volume is evacuated to $10^{-6}$ mbar in order to minimize background between interactions and residual gases. It begins $\sim$100~m downstream of the production target, is $\sim$110~m long, and consists of segments of increasing diameter, from $\sim$2~m upstream to $\sim$3~m downstream. 5~MHz of kaon decays are observed in the 65-m fiducial decay region, in the upstream part of the vacuum tank. The ring-shaped large-angle photon vetoes (LAVs) are placed at 11 stations along the vacuum volume and provide full coverage for decay photons with $8.5~{\rm mrad}<\theta<50~{\rm mrad}$. The last 35~m of the vacuum volume host a dipole spectrometer with four straw-tracker stations operated in the vacuum. At the exit of the vacuum region, a ring-imaging Cerenkov counter (RICH) 4~m in diameter by 17~m in length helps to identify charged decay secondaries. Downstream of the RICH, a number of photon vetoes provide hermeticity, including principally the NA48 liquid-krypton calorimeter (LKr) to veto forward ($1~{\rm mrad}<\theta<8.5~{\rm mrad}$), high-energy photons. The 12th LAV station provides downstream coverage at large angles ($8.5~{\rm mrad}<\theta<50~{\rm mrad}$), while a ring-shaped shashlyk calorimeter (IRC) about the beamline provides coverage for photons with $\theta<1~{\rm mrad}$. Further downstream, a muon veto detector (MUV) provides additional rejection for $K\to\mu\nu$ events, and a small-angle shashlyk calorimeter (SAC) around which the beam is deflected completes the coverage for very-small-angle photons that would otherwise escape via the beam pipe. Assuming an acceptance for signal events of about 10\%, the experiment must be able to reject background from the dominant $K^+$ decays such as $K^+\to\pi^+\pi^0$ at the level of $10^{12}$. Kinematic cuts on the $K^+$ and $\pi^+$ tracks (as reconstructed in the Gigatracker and straw chambers, respectively) provide a rejection factor of $10^4$ and ensure that the photons from the $\pi^0$ have 40~GeV of energy. There is a kinematic correlation between photon energy and angle of emission with respect to the beam axis; the forward photons that are intercepted by the LKr calorimeter, IRC, and SAC have much higher energies than those intercepted by the LAVs. Nevertheless, the photons from $K^+ \to \pi^+\pi^0$ intercepted by the LAVs may have energies from a few tens of MeV to several GeV. In order to detect the $\pi^0$ with an inefficiency of $\leq 10^{-8}$, the maximum tolerable inefficiency in the LAV detectors for photons with energies as low as 200~MeV is $10^{-4}$. In addition, the LAV detectors must have good time resolution ($\sim$1 ns) to allow signals from incident particles to be identified with the correct event, and good energy resolution ($\sim$10\% at 1 GeV) for precise vetoing and use in full-event reconstruction. The system must also be sensitive to minimum-ionizing particles. Finally, the LAV detectors must be compatible with operation in a vacuum of $10^{-6}$ mbar. \section{The Large-Angle Veto System} The NA62 LAV detectors make creative reuse of lead glass blocks recycled from the OPAL electromagnetic calorimeter barrel\cite{OPAL+91:NIM}, which became available in 2007, when various technologies were under consideration for the construction of the LAV detectors. Other solutions considered included a lead/scintillating tile design originally proposed for use in the (later canceled) CKM experiment at Fermilab, and a lead/scintillating-fiber design based on the electromagnetic calorimeter for the KLOE experiment. Prototype instruments based on each of the three technologies were obtained or constructed, and tested with the electron beam at the Frascati Beam-Test Facility. These tests demonstrated that all three technologies are suitable for use in NA62 \cite{A+07:Veto}. In particular, the inefficiency for the detection of single, tagged electrons with the OPAL lead glass modules was measured to be \EA{1.2}{0.9}{0.8}{-}{4} at 203 MeV and \EA{1.1}{1.9}{0.7}{-}{5} at 483 MeV. Basing the construction of the LAV system on the OPAL lead glass modules provides significant economic advantages. \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_block} \caption{A module from the OPAL calorimeter, without wrapping and with reinforcement plates at the interface between the glass and the steel flange.} \label{fig:block} \end{figure} The modules from the central part of the OPAL electromagnetic calorimeter barrel consist of blocks of Schott SF57 lead glass. This material is about 75\% lead oxide by weight and has a density $\rho = 5.5~{\rm g}/{\rm cm}^3$ and a radiation length $X_0 = 1.50$~cm; its index of refraction is $n \approx 1.85$ at $\lambda = 550~{\rm nm}$ and $n \approx 1.91$ at $\lambda = 400~{\rm nm}$. Electromagnetic showers in the lead glass are detected by virtue of the Cerenkov light produced; our measurements indicate that, averaged over modules, minimum ionizing particles produce about 0.34~p.e. per MeV of deposited energy. The front and rear faces of the blocks measure about $10\times10$ cm$^2$ and $11\times11$ cm$^2$, respectively; the blocks are 37~cm in length. (The precise geometry depends slightly on the ring of the OPAL calorimeter from which each block is extracted; blocks of uniform geometry are used in the construction of each ring of the LAV system.) Each block is read out at the back side by a Hamamatsu R2238 76-mm PMT, which is optically coupled via a 4-cm long cylindrical light guide of SF57 of the same diameter as the PMT. The rear face of the glass block is glued to a 1-cm thick stainless steel flange featuring the holes that allow mounting hardware to be attached and a circular cutout for the light guide. A mu-metal shield surrounding the PMT and light guide is also glued to the flange. A complete module (block plus PMT) is a monolithic assembly; the block and PMT cannot be independently replaced. Figure~\ref{fig:block} shows a picture of a complete module. \begin{figure} \centering \includegraphics[width=0.45\linewidth]{valencia_fig_A1study} \hspace{0.10\linewidth} \includegraphics[width=0.39\linewidth]{valencia_fig_A1} \caption{Design study (left) and completed prototype A1 veto station (right) making use of the OPAL lead glass calorimeter elements.} \label{fig:station} \end{figure} A LAV station is made by arranging these blocks around the inside of a segment of vacuum tank, with the blocks aligned radially to form an inward-facing ring. Multiple rings are used in each station in order to provide the desired depth for incident particles. The blocks in successive rings are staggered in azimuth; the rings are spaced longitudinally by about 1 cm. \begin{table} \renewcommand{\arraystretch}{1.3} \caption{Parameters of LAV Stations} \label{tab:param} \centering \begin{tabular}{@{}llllll@{}} \hline\hline Station & Diameter [mm] & \multicolumn{2}{l}{Block radius [mm]} & Layers & Blocks\\ & Outer wall & Inner & Outer & & \\ \hline A1--A5 & \phantom{$\sim$}2168 & \phantom{1}537 & \phantom{1}907 & 5 & 160\\ A6--A8 & \phantom{$\sim$}2662 & \phantom{1}767 & 1137 & 5 & 240\\ A9--A11 & \phantom{$\sim$}3060 & \phantom{1}980 & 1350 & 4 & 240\\ A12 & $\sim$3250 & 1072 & 1442 & 4 & 256\\ \hline\hline \end{tabular} \end{table} The LAV system consists of a total of 12 stations, the diameter of which increases with distance from the target. The geometry of the LAV stations is summarized in \Tab{tab:param}. For stations A1--A5, A6--A8, and A9--A11, apart from the different sizes and block configurations, the designs are conceptually similar. The geometry of A12 is not yet final; this station is operated in air and its design is different from that of the other stations. Since the spaces between the blocks are significantly smaller in the larger-diameter vessels, fewer layers are necessary. As a result of the staggering scheme, particles incident on any station are intercepted by blocks in at least three rings, for a total minimum effective depth of 21 radiation lengths. The vast majority of incident particles are intercepted by four or more blocks (27\,$X_0$). The stations with five layers (A1--A8) are 1.55~m in length, while those with four layers (A9--11) are 1.43~m in length. Station A1 was constructed as a prototype during the first half of 2009. It was installed in the NA62 beamline at CERN and tested with electrons and muons in October 2009. The design study and completed station are shown in \Fig{fig:station}. On the basis of experience gained during the test beam, various improvements were made. Station A2 was then constructed and tested with an unseparated, low-energy positive beam in the T9 beamline at the CERN PS in August 2010. A1 was subsequently rebuilt to incorporate the design improvements. Construction of the remaining stations was then commenced and is now underway. \section{LAV Construction} The LAV vacuum vessels are constructed from 25-mm thick steel plate, rolled into a cylinder and arc welded along the seam. Structural steel (S275JR) is used for stations A1 to A8 and A11. Stations A9 and A10 are mounted astride the spectrometer magnet, so non-magnetic ($\mu/\mu_0 \approx 1.01$) 304L stainless steel is used for these stations. After rolling, seam welding, and welding of the end flanges, the interior is turned on a vertical lathe and the vessel is vacuum tested. Five 200-mm ISO-F flanges for HV, signal, and calibration connections are welded via nozzles 20-cm long to one side of the vessel. One 630-mm ISO-K flange is attached to the opposite side for use as a manhole; the pumps for the NA62 vacuum system are also mounted on these flanges. Bolt holes on the flanges and in the interior walls of the tank are drilled with a CNC milling machine. A final vacuum test is performed, the residual magnetic field is mapped, and the chamber is cleaned and the outside is painted. A passivator that does not interfere with high-vacuum operation is applied to the inside walls of the vessel. The contract for the construction of vessels A1 to A8 and A11 was awarded to Fantini SpA (Anagni, Italy). Construction of A1 to A7 has been completed; construction of A8 is in progress. The OPAL detector modules (lead glass block plus PMT) were manufactured by Hamamatsu during the mid-1980s. Their recycling requires substantial care throughout the assembly procedure. After the storage area in which the modules were kept was inundated during a flash flood in spring 2008, the modules were subject to an extensive sorting and clean-up effort carried out at CERN by an industrial recovery firm. They were subsequently shipped to Frascati for use in LAV assembly. The interface between the stainless steel flange and lead glass block is fragile, and is found to be critically damaged in a few percent of the modules upon first examination. Typically, the epoxy between the glass and steel is found to have become delaminated in some areas, while in others, the glass is fractured at the interface with pieces still adhering to the flange. This is attributed to thermally induced stress from the differing expansion coefficients of the steel and of the glass and/or thermal shock in the glass due to the conductivity of the flange. Some modules are found to have the glass block completely separated from the steel flange. Since the modules are suspended by the flange in the LAV system, the fragility of the interface is a structural vulnerability and a safety risk. The first step in the processing of the modules at Frascati is to therefore to reinforce the interface. Using epoxy resin, 20~cm$^2 \times 0.5$-mm thick stainless steel plates are attached across the glass-steel interface on all four sides of the block. Calculations indicate and static tests confirm that the reinforced bond is several times stronger than the original bond. To maintain the design vacuum of $\sim$10$^{-6}$~mbar, all components of the LAV detectors must be carefully cleaned. A variety of techniques are used, including pickling, solvent degreasing, and ultrasonic cleaning. The lead glass blocks require special care. After reinforcement of the glass/steel interface, the original wrapping is removed and the blocks are thoroughly cleaned with acetone or isopropyl alcohol. They are then wrapped with a new laser-cut and heat-welded Tyvek cover. During the A1 test beam, ringing of the analog signal was observed to lead to errors in charge reconstruction using the time-over-threshold technique discussed below. This problem was traced to a small parasitic inductance in the PMT dynodes and solved by replacing the original OPAL HV dividers soldered to the PMTs with new dividers of our own design. The new divider features additional resistors on the last three dynodes and anode to damp out the oscillation, storage capacitors for the last three dynode stages to improve response linearity for large signals, and a decoupling resistor between the HV and signal grounds to decrease noise. \begin{figure} \centering \includegraphics[width=0.725\linewidth]{valencia_fig_oven} \caption{Test stand used for the characterization of lead glass detector modules. The top and bottom rows of lead glass blocks form a telescope for the selection of vertical cosmic rays. Here, the front panel is open; it is closed during operation.} \label{fig:oven} \end{figure} After the divider is replaced, the blocks are tested and characterized 12 at a time using a test station featuring an LED pulser and cosmic-ray telescope (\Fig{fig:oven}). The PMT gains are measured first, by varying the intensity of the light pulses from the LEDs as well as the PMT HV settings and mapping out the response for each block. Using the gain curves so obtained, the PMTs are then set to a reference value of the gain ($9\times10^5$ or $1\times10^6$) and the response to cosmic rays selected by the telescope is measured; the photoelectron yield for the block (p.e./MeV) is then obtained assuming that vertically incident cosmic rays leave 77 MeV in each block. As noted above, photoelectron yields of 0.34~p.e./MeV are typical. Finally, using the gain curves and the measurements of photoelectron yield, the PMT voltages are set to the values expected to produce a common output charge level of 4.5 pC for cosmic ray events. The response is measured and the HV setting is validated. Thus, at the end of a 12-hour cycle, which is fully automated using LabView, we have PMT gain and photoelectron yield measurements as well as the operational HV settings for 12 modules. Additional data (current-draw measurements, dark-count rates) are also collected using the test station. The OPAL design features an optical port at the base of each module. Blue LEDs are installed in these optical ports as part of the calibration and monitoring system, and will allow monitoring of the operational status and relative timing for each block. A low-capacitance LED was chosen to minimize the rise and fall times of the light pulse; this is important for use with the time-over-threshold-based readout system discussed in the following section. In principle, the LED system should allow in-situ gain measurement as well. \begin{figure} \centering \includegraphics[width=0.725\linewidth]{valencia_fig_assy} \caption{Installation of lead glass modules into the steel vacuum vessel.} \label{fig:assy} \end{figure} After testing and characterization, the blocks are arranged in groups of four in an aluminum mounting bracket. For the installation, the vacuum vessel is turned on end. The aluminum mounting bracket with four modules is lowered into the upended vessel by overhead crane and bolted to the wall, as illustrated in \Fig{fig:assy}. The HV, signal, and LED cables are then routed to the flanges along the cable grille visible at the top of the photo. In order to preserve uniformity of signal shapes and timing across channels, all signal (and LED) cables have the same length, with cable slack taken up on the grille. For HV cables, the excess length is cut. As noted above, the signal, HV, and LED connections are made on five 200-mm ISO-F flanges. Flanges with eight or ten DB37 connectors are used both for signal and LED feed-through, 16 channels per connector, with separate ground connections for each signal channel. Flanges with eight 32-pin MIL-C-26482 connectors are used for HV feed-through, 32 channels per connector, with a common ground for all 32 channels on a separate MHV feedthrough. \section{Front-End Electronics} As noted above, the LAVs must furnish time and energy measurements over a large range of incident photon energies. For reasons of cost and simplicity, we have decided on a readout scheme based on the time-over-threshold (ToT) technique. This scheme is implemented using a dedicated front-end ToT discriminator board of our own design \cite{A+11:ToT} and a digital readout board (TEL62) used by various NA62 detector subsystems \cite{A+11:TEL62}. The ToT discriminator converts the analog signals from the detector to low-voltage differential signal (LVDS) pulses, with width equal to the duration of the analog signal from the detector above a specified threshold. The signal from each PMT is compared to two different thresholds, a low threshold of about 5~mV and a high threshold expected to be about 50~mV, corresponding to two different LVDS outputs. (For comparison, the expected noise level is 2~mV under normal operating conditions.) The TEL62 is based on the design of the TELL1 readout board developed for the LHCb experiment. On the TEL62, TDC mezzanines measure the leading and trailing times of the LVDS pulses. Using a time-to-charge calibration parameterization, the FPGA on board the TEL62 calculates the time corrected for slewing, as well the charge for each hit as reconstructed from the pulse width above threshold. This information is sent to the subsequent DAQ stages. Level-0 trigger primitives are also calculated on board the TEL62 and sent to the level-0 trigger processor. \begin{figure} \centering \includegraphics[width=0.9\linewidth]{valencia_fig_fee} \caption{Conceptual schematic of the ToT discriminator board.} \label{fig:fee} \end{figure} A conceptual schematic of the ToT discriminator board is presented in \Fig{fig:fee} to illustrate the signal processing for a single channel. Since each channel is discriminated against two different thresholds, the board has 32 input channels and 64 output channels. While the amplitude of the PMT signal from a minimum ionizing particle is about 20~mV, signals from 20-GeV showers may be as large as 10~V. For protection, the input signal is clamped at 0.6~V by a circuit that maintains the timing of the rising and falling edges of the pulse. The clamped signal is then passively split. One copy of the signal is summed with the signals from other channels to form diagnostic analog outputs as discussed below. The other copy is amplified $\times3$ by a low-noise, high-bandwidth, high-speed amplifier and passively split into two copies. Each copy is used as input to a high-speed comparator with an LVDS driver. The threshold for comparison is provided by a programmable DAC in the board controller. To reduce double pulses from the comparator due to noise in the input signal, 3 mV of hysteresis is also provided through a feedback resistor, so that the output signal is extended until the input signal falls to 3 mV below the threshold. The other copy of the clamped analog signal is summed with the signals from three adjacent channels (and clamped at 2~V) for diagnostic purposes. The resulting analog sum is made available via a front-panel LEMO connector. These sums of four are in turn summed four at a time to produce sums of 16. For 32 input channels, there are a total of 10 front-panel analog outputs: eight sums of four and two sums of 16. A compact way to perform single channel diagnostics is to pulse the blocks one at a time using the LED system and read out the analog signal from these sums. The readout board is implemented on a 9U VME card and has a modular structure, with many of the important functions on mezzanines, to reduce costs and simplify repairs. The ToT discriminators are implemented on 16 mezzanines, each with 4 discriminator circuits. The analog sums are handled on 10 mezzanines (one per front-panel output). A pulse-generator mezzanine allows pulses of programmable width and amplitude to be sent to a pattern of channels on the input connectors. The board controller mezzanine handles communication with the experiment's slow-control system to allow the thresholds to be set and read, to control the test-pulse generator, and to monitor the board status. \section{Test-Beam Performance} The A2 LAV station was tested using the positive secondary beam in the T9 area at the CERN PS in August 2010. Data were collected at various beam momenta over the interval 0.3--10~GeV. The composition of the T9 beam changes as a function of the momentum setting. At 0.3~GeV, the beam is roughly 70\% $e^+$ and 30\% $\pi^+$ (including decay $\mu^+$), while above 6~GeV the $e^+$ component drops off sharply and at 10~GeV the beam is roughly a 50\% mix of $\pi^+/\mu^+$ and $p$. Two threshold Cerenkov counters in the beamline using CO$_2$ at adjustable pressure allowed samples enriched $e^+$ and $\mu^+$ to be selected. The T9 detectors also included two scintillation paddles and a beam wire chamber for beam counting and diagnostics. The beam focus was 2~m downstream of the beam wire chamber, and A2 was positioned so that the beam focus coincided approximately with the first layer of lead glass blocks. Two additional scintillator paddles ($60\times85$~mm$^2$) were placed in a cross configuration at beam entrance and one larger paddle was placed at beam exit. Requirements on the signal amplitude for both entrance counters significantly enriched the sample in events with a single beam particle incident on a single column of lead-glass blocks, while requiring in addition a hit on the exit counter significantly enriched the sample in events in which the beam particle was a $\mu^+$. A prototype of the ToT discriminator board (with one discriminator per channel, rather than two) was tested together with A2. For much of the test, our board was used together with commercial QDCs and TDCs, but in dedicated runs, the times were also measured with a prototype setup using a TELL1 board from LHCb and an early version of the TDC mezzanine. \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_qtot} \caption{Scatter plot of signal charge measured using the QDCs vs.\ signal time over threshold, for electrons of various energies at the T9 test beam.} \label{fig:qtot} \end{figure} \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_lin} \caption{Equivalent charge-integrated signal from ToT vs.\ incident electron energy, demonstrating linearity of response.} \label{fig:lin} \end{figure} \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_eres} \caption{Energy resolution obtained using the ToT technique compared with that obtained using the QDCs, as a function of incident electron energy.} \label{fig:eres} \end{figure} Figure~\ref{fig:qtot} shows a scatter plot of the signal charge measured using the QDCs vs.\ the signal time over a threshold of 4~mV, for electrons. The data are summed for all blocks on which the beam was incident and over runs of different energies. For small signals, a small increase in the integrated charge corresponds to a large increase in the time over threshold; the ToT measurement provides more sensitivity for small signals than does the QDC measurement. This is a desirable property for the LAV detectors, since high detection efficiency is required for low-energy photons. Parameterizations of the type illustrated in \Fig{fig:qtot} are used to convert the time over threshold to an effective charge measurement. Figure~\ref{fig:lin} demonstrates that good linearity of response is obtained using this method. The plot shows the equivalent charge in pC from the ToT measurement (with 4~mV threshold) for electrons of energy 0.3, 0.5, 1, and 2~GeV. Linearity to within 1\% is observed. The energy resolution obtained using the ToT technique is compared with that obtained using the QDCs in \Fig{fig:eres}. As expected from the form of the curve in \Fig{fig:qtot}, at low energies, the resolution obtained with the ToT technique is better than that obtained with the QDCs. The fits in \Fig{fig:eres} give \begin{eqnarray*} {\rm QDC:}&\hspace{1mm}& \sigma_E/E = 8.6\%/\sqrt{E~{\rm[GeV]}} \oplus 13\%/E, \\ {\rm ToT:}&\hspace{1mm}& \sigma_E/E = 9.2\%/\sqrt{E~{\rm[GeV]}} \oplus 5\%/E \oplus 2.5\%. \end{eqnarray*} While, as expected, the statistical contribution to the energy resolution is about the same with either readout scheme, the contribution from noise (term proportional to $1/E$) appears to be significantly smaller with the ToT technique. The presence of the constant term with the ToT technique may be due to small differences in the charge vs.\ ToT curves from block to block, and if so, can potentially be reduced. \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_tres} \caption{Time resolution obtained using the ToT technique as a function of incident electron energy.} \label{fig:tres} \end{figure} To study the time resolution, it is first necessary to correct for slewing. The event time reference is obtained from the scintillator paddles and Cerenkov counters, and the curve of signal time vs.\ integrated charge from ToT is parameterized assuming that the signal shape is well described by the form $V(t) \propto t^a e^{-bt}$. The width of the signal time distribution in slices of charge gives a measurement of the time resolution. A fit to the measurements of $\sigma_t$ vs. charge from ToT (4~mV threshold) for a single block gives $\sigma_t = 220~{\rm ps}/\sqrt{E~{\rm[GeV]}} \oplus 140~{\rm ps}$, where the constant term is assumed to be due to trigger jitter. This assumption can be tested by measuring the width of the distribution of signal time differences for two successive blocks. Assuming that the two blocks have the same intrinsic time response and that there is no common-mode contribution to the resolution, $\sigma_{t_1} = p/\sqrt{E_1}$ and $\sigma_{t_2} = p/\sqrt{E_2}$, so that we expect $\sigma_{\Delta t} = p/[E_1 E_2/(E_1 + E_2)]^{1/2}$, with no constant term. Figure~\ref{fig:tres} shows the measurements of $\sigma_{\Delta_t}$ in slices of the weighted charge measurements from ToT, $Q = Q_1Q_2/(Q_1 + Q_2)$ (4~mV threshold). The overlaid curve has $p = 210$~ps and no constant term, which is indicative of the intrinsic time resolution of the detector. \begin{figure} \centering \includegraphics[width=0.85\linewidth]{valencia_fig_sig} \caption{Simulation of PMT signals. The red histogram shows the number of photoelectrons produced in bins of time (scale at right). The black curve is the resulting PMT signal (scale at left).} \label{fig:sig} \end{figure} \begin{figure*} \centering \includegraphics[width=0.75\linewidth]{valencia_fig_sim} \caption{Comparison of charge vs.\ ToT curves for MC (9.5 mV threshold) and data (various thresholds). Apart from a small discrepancy in the value of the threshold, the simulation accurately reproduces the shape of the charge vs.\ ToT curve.} \label{fig:sim} \end{figure*} The results obtained at the beam test provide a point of comparison for the Monte Carlo (MC) simulation of the LAV system. The Geant4-based NA62 MC includes a detailed description of the LAV geometry and materials. Two sets of routines are available for shower simulation and the creation and propagation of Cerenkov photons: the standard Geant4 routines, with complete tracking of the optical photons, and our own routines, which make use of a response matrix obtained from the full simulation. In either case, the simulation produces the number of Cerenkov photons that arrive at the photocathode, together with their arrival times, as shown by the histogram in red in \Fig{fig:sig}. A complete simulation of the PMT uses this information to generate an output signal, taking into account the PMT gain and transit-time fluctuations, the capacitance of the PMT, and dispersion in the readout cable. The resulting signal, which is illustrated by the black curve in \Fig{fig:sig}, is input to a full simulation of the ToT discriminator, including the comparator hysteresis. The complete simulation thus yields a description of the PMT signal, the integrated charge, and the ToT response. Figure~\ref{fig:sim} shows charge vs.\ ToT curves for test-beam muons for four different threshold settings (4, 5, 6, and 8~mV). The red curve is from the simulation, with a threshold at 9.5~mV and all other adjustable parameters (e.g., tube gain, interdynode time fluctuations, PMT capacitance) set to typical or measured values. The simulation with a threshold of 9.5~mV accurately reproduces the shape of the measured charge vs.\ ToT curve with threshold 8~mV. Apart from this slight discrepancy, which is easily attributed to the accuracy of the manual threshold adjustment on the prototype ToT discriminator board, these results confirm our detailed understanding of the detector and the readout chain. \section{Status and Outlook} \begin{figure} \centering \includegraphics[width=0.8\linewidth]{valencia_fig_A6} \caption{The first intermediate diameter LAV station (A6) upon completion at Frascati in November 2011.} \label{fig:A6} \end{figure} Construction of the five small-diameter stations (including updating of the A1 prototype) was completed with all stations delivered to CERN and awaiting installation on the beamline in July 2011. Installation of these five LAVs is scheduled for December 2011. As of November 2011, construction of the remaining stations is in progress; the first of the intermediate diameter stations was recently completed (\Fig{fig:A6}). Serial production of the front-end electronics boards is beginning. The installed detectors will be read out in a dry run in mid 2012, while a technical run with beam is scheduled for late 2012 and will include at least eight of the twelve LAVs. Data taking with the remaining detectors is planned for 2013. \section*{Acknowledgments} We warmly thank C. Capoccia and A. Cecchetti of the Experimental Apparatus Design Service (SPAS) at the INFN Frascati Laboratories, and S. Bianucci of INFN Pisa, for their collaboration on the mechanical design of the LAV system, including the steel vacuum vessels. We also thank our mechanical design team for their assistance in oversight of the construction of the steel vessels and of the transport of the completed detectors to CERN. We thank G. Corradi, D. Tagnani, and C. Paglia of the Electronics Service (SELF) at INFN Frascati for their collaboration on the design of the new HV dividers and front-end electronics boards. Construction was made possible by the contributions from our INFN technicians: E. Capitolo, R. Lenci, V. Russo, M. Santoni, T. Vassilieva, and S. Valeri (Frascati); F. Cassese and L. Roscilli (Napoli); L. Berretta and G. Petragnani (Pisa); and F. Pellegrino (Roma). Finally, We thank V. Lollo and P. Chimenti of the Vacuum Service at INFN Frascati for their assistance with vacuum-related issues. \bibliographystyle{IEEEtran}
\section{\textbf{Introduction}} With the immense advance of modern nanotechnology these decades, it has become possible to confine the charge carriers in nanostructures, such as quantum dots (QDs), quantum wires, quantum wells and superlattices, with well controlled shape size and other properties [1-5]. Since QDs are created mainly by imposing a lateral confinement potential to electrons in a very narrow quantum well, the features which should be mainly considered in a QD are geometric shape, size and confining potential. As we know the physical properties can not be extrapolated from behavior at large sizes because of the prevalence of quantum effects. So the investigation of QDs has attracted enormous theoretical and experimental interest [6-10]. Further, the distinctive electrons and optical properties of QDs show a considerable potential application in electronic optoelectronic device. It is especially necessary to investigate the optical properties of low-dimensional semiconductor because the study of optical transitions yields important information about the energy spectrum, the Fermi surface of electrons, and the value of the electronic effective mass. A great deal of work has been done on the linear and nonlinear optical properties of low-dimensional semiconductor systems, and numerous of important new information is discovered. For instance, nonlinear optical properties of an off-center donor in a QD in presence of applied magnetic field is investigated by Wenfang Xie [11], the results show the ion position, the applied magnetic field, the confinement strength and the incident have an important influence on the nonlinear optical properties of off-center donors. Therefore, some special needs can be met by modulating these physical quantities in designing electro-optical devices. In addition, the researches on isotropic parabolic potential have drawn great attention, and lots of corresponding reports have been published. Also the effects of the parabolic on the energy and optical properties have been widely investigated. In many studies, an isotropic harmonic oscillator potential are used as the lateral confinement of electrons to do the optical research for the reasons that the parabolic confinement is more appropriate when the QD is fabricated and the parabolic gives a quadratic Hamiltonian, the spectrum of which can be obtained by a purely algebraic method. As far as we know, a number of important optical information has been obtained and many nonlinear problems have been resolved by this method [12-14]. In recent years, the theory for boundary value problems for anisotropic quantum systems have aroused significant interest of the researchers because the new and useful effects have been discovered in such quantum systems [15-21]. For example, the resonance Raman scattering in the anisotropic QDs in the presence of arbitrarily directed magnetic field and polarization vector have theoretically investigated by Shorokhov and Margulis [15], the results reveal that the QD subjected in a magnetic field can be used as the detector of phonon modes. Hybrid resonances in the optical absorption of a three-dimensional anisotropic quantum well has been studied by Geyler $et$ $ al.$ [16], the results show that in the general case there are three resonance frequencies on the absorption curve, and in the limit $T\rightarrow0$ ( the case of a degenerate gas), the absorption curve as a function of the radiation frequency contains kinks. In addition, hybrid-impurity resonances in anisotropic QDs has been discussed by Margulis $et$ $ al.$ [17], and they reported that the scattering of electrons by impurities leads to the resonance absorption even if we have only one impurity in the QD; the position of the hybrid-impurity resonances depends strongly on the magnetic field and the characteristic frequencies of the parabolic confinement. Moreover, magnetoexcitons in quantum wires with an anisotropic parabolic potential has been calculated by Tanaka $et$ $ al.$ [18]. Their results show that the experimental results on magnetophotoluminescence (magneto-PL) of GaAs QWR's can be well explained by this model, demonstrating the importance of the exciton effect to understand the magneto-PL properties of QWR's. All these results indicate that the optical properties are strongly dependent on the anisotropy of the material structure. In this paper, we present a theoretical study of the linear and nonlinear optical absorption coefficients (ACs) and refractive index (RI) changes in a three-dimensional QD subjected to a uniform magnetic field directed with respect to the $z-$axis. To simplify the calculations, the linear canonical transformations of the phase space has been adopted [16,17]. By means of these transformations, one can find a new coordinate system such that in the new phase coordinates the wave functions have the simplest form. From the consideration of device applications, engineering the electric structure of materials by means of characteristic frequency, magnetic field, incident optical intensity and the dot radius, offers the possibility of tailoring the energy spectrum to produce desirable nonlinear optical properties. \section{\textbf{Model and theory}} \subsection{Electronic state in an anisotropic quantum dot } Within the framework of the effective mass approximation, the Hamiltonian description of an electron confined in an anisotropic parabolic QD in the presence of magnetic field $\vec{B}$ has the form \begin{equation} H=\frac{1}{2m^*}(\vec{p}-\frac{e}{c}\vec{A})^2+\frac{m^*}{2}(\Omega_{x}^2x^2+\Omega_{y}^2y^2+\Omega_{z}^2z^2), \end{equation} where $m^*$ is the electronic effective mass, $\vec{A}$ is the vector potential of magnetic field $\vec{B}$, $\Omega_{i}$ $(i=x,y,z)$ are the characteristic frequencies of parabolic potential. Here we choose the following gauge for the vector potential: \begin{equation} \vec{A}=(\frac{1}{2}B_{y}z-B_{z}y,0,B_{x}y-\frac{1}{2}B_{y}x). \end{equation} It is surely a complicate problem to calculate directly the matrix elements of electron-photon interaction. To resolve this problem, a method of canonic transformation of the phase space has been adopted via using only simple calculations from linear algebra, and which has been successfully used to resolve many problems about the anisotropic systems [16,18-21]. Within the linear canonic transformation of the phases pace, the Hamiltonian can be represented with canonic form in the new phase coordinates ($\vec{P}, \vec{Q}$) \begin{equation} H(\vec{P},\vec{Q})=\frac{1}{2m^*}(P_{1}^2+P_{2}^2+P_{3}^2)+\frac{m^*}{2}(\omega_{1}^2Q_{1}^2+ \omega_{2}^2Q_{2}^2+\omega_{3}^2Q_{3}^2), \end{equation} where $\omega_{i}$ $(i=1, 2, 3)$ are the hybrid frequencies related to the magnitude and direction of the magnetic field $\vec{B}$, and the frequencies are obtained from six-order algebraic equation [22]. Hence, the spectrum of Hamiltonian can be obtained as \begin{equation} E_{nml}=\hbar\omega_{1}(n+\frac{1}{2})+\hbar\omega_{2}(m+\frac{1}{2})+\hbar\omega_{3}(l+\frac{1}{2}), \quad n,m,l,=0, 1, 2,\ldots \end{equation} and the corresponding wave functions in the new phase coordinates have the form \begin{equation} \Phi_{nml}(Q_{1},Q_{2},Q_{3})=\phi_{n}(Q_{1})\phi_{m}(Q_{2})\phi_{l}(Q_{3}) \end{equation} where $\phi_{j}(Q)$ are the oscillator functions, and $j=n,m,l$.\\ \subsection{Calculation of linear and third-order nonlinear optical absorption coefficients and refractive index changes} In the present work, the compact-density approach to calculate absorption coefficient and refractive index for a three-dimensional anisotropic QD. For simplicity, suppose the system is excited by a monochromatic electromagnetic field as \begin{equation} E(t)=E_{0}\cos{(\omega t)}=\widetilde{E}e^{i\omega t} +\widetilde{E}e^{-i\omega t}, \end{equation} where $\omega$ is the frequency of the external field incident with a polarization vector normal to the QD along $z$-direction. The time-dependent equation of the density matrix operator $\varrho$ is given by \begin{equation} \frac{\partial\varrho_{ij}}{\partial t}=\frac{1}{i\hbar}[H_{0}-qzE(t),\varrho]_{ij}-\Gamma_{ij}(\varrho-\varrho^{(0)})_{ij}, \end{equation} where $H_{0}$ is the Hamiltonian of the system without the incident field $E_{(t)}$, $q$ is the electronic charge. $\varrho^{(0)}$ is the unperturbed density matrix and $\Gamma_{ij}=\frac{1}{\tau_{ij}}$ is the relaxation rate representing the damping due to the electron-phonon interaction. Using the iterative method, the time-dependent density matrix operator can be solved as \begin{equation} \varrho(t)=\sum_{n}\varrho^{(n)}(t), \end{equation} with \begin{equation} \frac{\partial\varrho_{ij}^{(n+1)}} {\partial t}=\frac{1}{i\hbar}\{[H_{0},\varrho^{(n+1)}]_{ij}-i\hbar\Gamma_{ij}\varrho^{(n+1)}_{ij}\} \frac{1}{i\hbar}[qz-\varrho^{(n)}]_{ij}E(t). \end{equation} The electronic polarization $\vec{P}(t)$ and susceptibility $\chi(t)$ are defined by the dipole operator $M$, and the density matrix $\varrho$, respectively, \begin{equation} \vec{P}(t)=\epsilon_{0}\chi_{\omega}\widetilde{E}e^{i\omega t}+\epsilon_{0}\chi_{-\omega}\widetilde{E}e^{-i\omega t} =\frac{1}{V}Tr(\varrho M), \end{equation} where $\epsilon_{0}$ is the permittivity of free space, $V$ is volume of the system. $Tr$ denotes the trace or the summation over the diagonal elements of the matrix $\varrho M$, the analytic expressions of the linear and third-order nonlinear susceptibilities can be obtained as follows: for the linear term \begin{equation} \epsilon_{0}\chi^{(1)}(\omega) =\frac{\rho|M_{fi}|^2}{E_{fi}-\hbar\omega-i\hbar\Gamma_{fi}}, \end{equation} and for the third-order nonlinear term \begin{eqnarray} \epsilon_{0}\chi^{(3)}(\omega)&=&-\frac{\rho|M_{fi}|^2\widetilde{E}^2}{E_{fi}-\hbar\omega-i\hbar\Gamma_{if}} [\frac{4|M_{fi}|^2}{(E_{fi}-\hbar\omega)^2+(\hbar\Gamma_{fi})^2}\nonumber\\ &-&\frac{(M_{ff}-M_{ii})^2}{(E_{fi}-i\hbar\Gamma_{fi})(E_{fi}-\hbar\omega-i\hbar\Gamma_{if})}], \end{eqnarray} where $\rho$ is carriers density. $E_{fi}=E_{f}-E_{i}$ is the energy interval of the two level system. $M_{fi}=e\langle \Phi_{f}|x|\Phi_{i}\rangle$ denotes the electric dipole moment of the transition from $\Phi_{i}$ state to $\Phi_{f}$ state. $\Gamma$ is phenomenological operator. Non-diagonal matrix element $\Gamma_{if}$ $(i\neq f)$ of operator $\Gamma$, which is called as relaxation rate of $f$th state, is the inverse of the relaxation time $\tau_{if}$. The susceptibility $\chi(\omega)$ is related to the absorption coefficient $\alpha(\omega)$ and the changes in the refractive index $\Delta n(\omega)/n_{r}$ as follows \begin{equation} \alpha(\omega)=\omega\sqrt{\frac{\mu}{\epsilon_{R}}}Im(\epsilon_{0}\chi(w)), \end{equation} \begin{equation} \frac{\Delta n(\omega)}{n_{r}}=Re(\frac{\chi(\omega)}{2n_{r}}), \end{equation} where $\mu$ is the permeability of the material, $\epsilon_{R}=n_{r}^2\epsilon_{0}$ is the real part of the permittivity. The linear and third-order nonlinear absorption coefficients are obtained as \begin{equation} \alpha^{(1)}(\omega)=\omega\sqrt{\frac{\mu}{\epsilon_{R}}}\frac{\rho|M_{fi}|^2\hbar\Gamma_{if}} {(E_{fi}-\hbar\omega)^2+(\hbar\Gamma_{fi})^2}, \end{equation} \begin{eqnarray} \alpha^{(3)}(\omega,I)&=&-\omega\sqrt{\frac{\mu}{\epsilon_{R}}}(\frac{I}{2\epsilon_{0}n_{r}c}) \frac{\rho|M_{fi}|^2\hbar\Gamma_{if}} {(E_{fi}-\hbar\omega)^2+(\hbar\Gamma_{fi})^2}[4|M_{fi}|^2\nonumber\\ &-&\frac{|M_{ff}-M_{ii}|^2[3E_{fi}^2-4E_{fi}\hbar\omega+\hbar^2(\omega^2-\Gamma_{if}^2)]}{E_{fi}^2+ (\hbar\omega_{if})^2}]. \end{eqnarray} Here $I$ is the intensity of incident radiation, and the total absorption coefficient $\alpha(\omega,I)$ is given by \begin{equation} \alpha(\omega,I)=\alpha^{(1)}(\omega)+\alpha^{(3)}(\omega,I). \end{equation} The linear and third-order nonlinear refractive index changes are given by \begin{equation} \frac{\Delta n^{(1)}(\omega)}{n_{r}}=\frac{\rho|M_{fi}|^2}{2n_{r}^2\epsilon_{0}}\frac{E_{fi}-\hbar\omega} {(E_{fi}-\hbar\omega)^2+(\hbar\Gamma_{fi})^2}, \end{equation} and \begin{eqnarray} \frac{\Delta n^{(3)}(\omega)}{n_{r}}&=&-\frac{\rho|M_{fi}|^2}{4n_{r}^3\epsilon_{0}} \frac{\mu c I}{[(E_{fi}-\hbar\omega)^2+(\hbar\Gamma_{fi})^2]^2}\nonumber\\ & &\times[4(E_{fi}-\hbar\omega)|M_{fi}|^2-\frac{|M_{ff}-M_{ii}|^2}{E_{fi}^2+(\hbar\omega)^2} ((E_{fi}-\hbar\omega)\nonumber\\ & &[E_{fi}(E_{fi}-\hbar\omega)-(\hbar\Gamma_{if})^2]-(\hbar\Gamma_{if})^2(2E_{fi}-\hbar\Gamma_{if}))]. \end{eqnarray} Therefore, the total refractive index change can be obtained as \begin{equation} \frac{\Delta n(\omega)}{n_{r}}=\frac{\Delta n^{(1)}(\omega)}{n_{r}}+\frac{\Delta n^{(3)}(\omega)}{n_{r}}. \end{equation} \section{\textbf{Result and Discussions}} In this study, the linear, the third-order nonlinear and the total ACs and RI changes of the three-dimensional anisotropic $GaAs/AlGaAs$ QD subjected to the uniform magnetic field directed with respect to the $z-$axis have been numerically investigated. The physical parameters for calculations are used as follows [23-25]: $\rho=5.0\times10^{22} m^{-3}$, $\tau=0.14 s$, $n_{r}=3.2$ and $m^*=0.067 m_{0}$, where $m_{0}$ is the free electronic mass. The other two characteristic frequencies are set as $\Omega_{y}=2.0\times10^{12}s^{-1}$ and $\Omega_{z}=1.5\times10^{11} s^{-1}$. The results of our calculations are presented in figures 1-8. Fig. 1 presents the linear, the third-order nonlinear and the total ACs of a three-dimensional anisotropic QD as a function of the incident photon energy $\hbar\omega$ for three different values of characteristic frequency $\Omega_{x}=0.9\times10^{13} s^{-1}$, $1.1\times10^{13} s^{-1}$ and $1.3\times10^{13} s^{-1}$, respectively. In Fig. 1, the applied magnetic field $B$ is set to be $1.0 T$, and the incident optical intensity is set to be $1.0 Mw/cm^{2}$. From Fig. 1, it can be easily found that the ACs as a function of $\hbar\omega$ in the three-dimensional anisotropic parabolic potential QD share the similar physical features with that in the isotropic parabolic potential QD [26]. Moreover, it can be noted that the linear, third-order nonlinear and total ACs are not monotonic functions of $\hbar\omega$. For each $\Omega_{x}$, the linear, third-order nonlinear, and total optical ACs as a function of $\hbar\omega$ has a prominent peak, respectively, at some position because of the one-photon resonance enhancement. And when the magnitude of $\hbar\omega$ is larger than $30 meV$, all the ACs will be extremely slow approaching zero as $\hbar\omega$ further increases. Additionally, it is can be found the linear ACs are monotonic functions of the characteristic frequency $\Omega_{x}$, and the values of which are monotonously increased as $\Omega_{x}$ increases. While the nonlinear terms are monotonously decreased due to the negative sign. Also, the increase of the $\Omega_{x}$ shifts the peak position of the corresponding AC to higher frequencies. It indicates a strong confinement-induced blue shift of an anisotropic parabolic resonance in semiconductor QDs. This physical behavior is originated from an augment of the energy difference between the energy levels with the enhancement of $\Omega_{x}$. So it can be concluded that the ACs of the anisotropic QD are strongly dependent on the characteristic frequency. The RI changes are always thought to be another important parameters in optical studies of QDs. The linear, third-order nonlinear and total RI changes of the three-dimensional anisotropic QD as a function of $\hbar\omega$ are illustrated in Fig. 2 with the same parameters as taken in Fig. 1. As can be seen from this figure, the total RI changes are remarkably reduced by the third-order nonlinear RI changes because the nonlinear terms are opposite in sign of the linear ones. It is indicated that the third-order nonlinear term has an important contribution in the variation of total RI change, although the change in the total RI is mainly caused by the linear term. Moreover, one can observe that the RI changes are not monotonic functions of $\hbar\omega$. At beginning, the linear and total RI changes increase steadily with photon energy, and then reach the maximum values at some $\hbar\omega$. However, as $\hbar\omega$ approaches the threshold energy, the linear and total RI changes decrease quickly and reach the minimum values. Note that, in this region, the dispersion for any frequency of the incident photon changes its sign. So this region is called as an absorption band due to the strong optical absorption [11]. On the other hand, it can be easily seen that, with increase of $\Omega_{x}$, the RI peaks will move to right side, which shows a blue shift of the resonance in QD. This physical behavior can be explained as that the energy interval between the initial state and the final state in an anisotropic QD becomes wider with increasing $\Omega_{x}$. Obviously, the RI changes are greatly affected by the characteristic frequency. In Fig. 3 and 4, the linear, third-order nonlinear and total optical ACs and RI changes are displayed, respectively, as a function of the incident photon energy $\hbar\omega$ for three different values of magnetic field with $\Omega_{x}=1.2\times10^{13} s^{-1}$, and $I=1.0 Mw/cm^2$. The magnetic field effect on the optical properties has been clearly shown in these two figures. One can see that the ACs and RI changes are also not monotonic functions of $\hbar\omega$ for the different magnetic field. In addition, it is found that the ACs and RI changes as a function of $\hbar\omega$ are very sensitive to the magnetic field. Just as shown in the figures, the strong absorption can be obtained at a weak magnetic field, and the peak values of ACs are rapidly increased by the increasing $B$. Moreover, the linear ACs and RI changes are increased with increase of $B$, while the third-order nonlinear ACs and RI changes are significantly reduced because of their negative sign. Besides, the increase of $B$ enhances the total terms and moves all total optical absorption peaks to higher energy. Note that all these behaviors are originated from the effect of the magnetic field on the anisotropic parabolic potential. The increase of applied magnetic field causes an enhancement in the confinement of electron in the dot and leads to a larger separation between the initial energy level and the final energy level. From Eqs. (16) and (19), it is theoretically shown that the nonlinear terms strongly depend on the incident photon intensity, so the contribution of nonlinear terms in ACs and RI changes is enhanced by an increase of incident photon intensity. In order to illustrate this effect, the linear, third-order nonlinear and total optical ACs and RI changes are plotted as a function of the incident photon energy $\hbar\omega$ for different values of incident optical intensity $I$ with $\Omega_{x}=1.2\times10^{13} s^{-1}$, and $B=1.0 T$ in the Fig. 5 and 6. As expected, the linear terms can not be changed by the variation of $I$, but the third-order nonlinear terms are monotonously decreased by the increasing $I$, where the decrease is originated from the negative sign of the third-order nonlinear terms. As a result, the total ACs and the total RI changes are reduced. Moreover, from Fig. 5, one can see that the absorption is strongly bleached at sufficiently high incident optical intensities, and the strong absorption begins to occur at around $I=2.0 Mw/cm^2$ in this work. When the incident optical intensity exceeds this value, the total absorption peak will be evidently spilt into two peaks, which is a consequence of the strong absorption. In addition, there is no shift at the resonance peak position with the variation of the incident optical intensity because the energy difference can not be affected by $I$. On the other hand, as can be seen from Fig. 6, the linear RI changes do not change with the intensity, while the third-order nonlinear terms are enhanced by increasing intensity. And that since the third-order term is opposite in sign of the linear one, any enhancement in the magnitude of nonlinear term will lead to a reduction in the difference between the linear term and the corresponding third-order nonlinear term. As a result, the total RI change is reduced. Therefore, to obtain a larger RI change, a weaker incident optical intensity should be applied. This result is in good agreement with the available refs. [10,11]. It is well know that the electron effective mass is strongly affected by its position, and the relationship between them has been reported by Peter $et$ $ al. $ [27] as $\frac{1}{m(\vec{r})}=\frac{1}{m^*}+[1-\frac{1}{m^*}]\exp{[-\xi r]}$, where $r$ denotes the dot radius, and $\xi$ is a constant which is chosen to be $0.01 a.u.$ This choice is selected taking into account that as $r\rightarrow0$, the particle is strongly bound within a $\delta$- function well, whereas as $r\rightarrow\infty$, the system is three-dimensional and characterized by the conduction band effective mass $m^*$ [27,28]. By using this expression, it can be easily explained that there is no appreciable change of the effective mass for a large dot radius, which also shows that mass variations are unimportant for large dots. Take into account the position-dependent effective mass, the linear, the third-order nonlinear and the total optical ACs and RI changes are plotted, respectively, in Fig. 7 and 8 as a function of the incident photon energy $\hbar\omega$ for three different values of dot radius with $\Omega_{x}=1.2\times10^{13} s^{-1}$, $B=1.0 T$ and $I=1.0 Mw/cm^{2}$. From Fig. 7, it can be seen that the linear ACs are monotonously increased with the increase of dot radius $r$, as opposed to the third-order nonlinear ones. The physical origin is that the electron effective mass is strongly dependent on its position. So it is clear that the mass variation with position plays an important role in the three-dimensional anisotropic QDs. On the other hand, from Fig. 8, it can be readily observed that the increase of dot radius leads to an increase of the maximum values and a decrease of the minimum values of the corresponding RI changes, respectively. This again shows the importance of mass variation with position for the three-dimensional anisotropic QDs. Such a strong dependence of the linear, third-order nonlinear and total ACs and RI changes on the dot radius would be important for many device application. \section{\textbf{Summary}} In this paper, we have presented a detail study of the linear and nonlinear optical ACs and RI changes in a three-dimensional anisotropic QD subjected to a uniform magnetic field directed with respect to the $z-$axis within the framework of effective-mass approximation. Using density matrix theory and iterative method, the expressions of linear and third-order nonlinear optical ACs and RI changes have been obtained. The dependence of the optical absorption and refractive index change on the characteristic frequency $\Omega_{x}$ of the parabolic potential, on the magnitude of the magnetic field $B$, on the incident optical intensity $I$, and on the dot radius $r$ has been investigated. It has been revealed that the nonlinear optical properties of the three-dimensional anisotropic QD are strongly dependent on these factors. For example, it is shown that the peak positions of linear and nonlinear terms shift to higher photon energy as the characteristic frequency increases; the ACs and RI changes will be enhanced with increasing magnetic field; the nonlinear part of ACs and RI changes are increased by increasing the incident optical intensity. And also the increase of incident optical intensity leads to a reduction of the total ACs and RI changes. Furthermore, the linear, third-order nonlinear and total ACs and RI changes are obviously affected by the dot radius. The present results are useful for further understanding the nonlinear optical properties of a three-dimensional anisotropic QD, and we hope that this theoretical study can make a significant contribution to experimental studies and practical applications. \section{\textbf{Acknowledgement}} This work is supported by National Natural Science Foundation of China (under Grant No. 11074055). \newpage
\section{Introduction} Study of the electromagnetic (EM) decays of the hadronic states is of interest for understanding the structure of hadronic matter and for revealing the fundamental mechanisms for the interactions of photons and hadrons~\cite{landsberg1982,landsberg1985}. The EM Dalitz decays $V \rightarrow P l^+l^-$ of light unflavored vector mesons ($\rho$, $\omega$ or $\phi$) are specially interesting for probing the electromagnetic structure arising at the vertex of the transition $V$ to the pseudoscalar. In table~\ref{lightdecays}, we summarize the experimental results for the EM Dalitz decays of the light vector mesons. The ratios of the Dalitz decays to the corresponding radiative decays of vector mesons are suppressed by two orders of magnitude, especially, for $V\rightarrow P e^+e^-$ modes. Assuming point-like particles, the decay rate of this process vs. $m_{l^+l^-}$ can be exactly described by QED~\cite{kroll1955} in the standard model (SM). However, the rate is strongly modified by the dynamic transition form factor $F_{V P}(q^2)$ ($q^2 = m^2_{l^+l^-}$), which can be estimated based on models of QCD~\cite{sov1992,faessler2000,klingl,kopp}. \begin{table}[htbp] \caption{The experimental results on the light vector Dalitz decays , $V\rightarrow P l^+l^-$ ($V= \rho$, $\omega$ or $\phi$), and ratios of the Dalitz decays to radiative decays of the vector mesons. These data are from PDG2010~\cite{pdg2010}.}\label{lightdecays} \begin{tabular}{c|c|c}\hline Decay mode & Experimental results & $\frac{\Gamma(V\rightarrow P l^+l^-)}{\Gamma(V\rightarrow P\gamma)}$ \\ \hline $\rho^0 \rightarrow \pi^0 e^+ e^-$& $<1.2 \times 10^{-5} $ (90\% C.L.) & $< 2.0\times 10^{-2}$ \\ \hline $\omega \rightarrow \pi^0 e^+ e^- $ & $(7.7\pm 0.6) \times 10^{-4}$ & $(0.93 \pm 0.08) \times 10^{-2}$ \\ \hline $\omega \rightarrow \pi^0 \mu^+ \mu^-$& $(1.3\pm 0.4) \times 10^{-4}$ & $(0.16\pm 0.05) \times 10^{-2}$\\\hline $\omega \rightarrow \eta e^+e^-$ & $<1.1\times 10^{-5}$ (90\% C.L.) & $<2.4\times 10^{-2}$ \\ \hline $\phi \rightarrow \eta e^+ e^-$& $(1.15\pm 0.10) \times 10^{-4}$& $(0.88 \pm 0.08)\times 10^{-2}$ \\ \hline $\phi \rightarrow \eta \mu^+ \mu^-$& $<9.4 \times 10^{-6}$(90\% C.L.) & $<0.07 \times 10^{-2}$\\ \hline \end{tabular} \end{table} Experimentally, $|F_{VP}(q^2)|^2$ is directly accessible by comparing the measured invariant mass spectrum of the lepton pairs from the Dalitz decays with the point-like QED prediction. A comprehensive review of the topic is contained in reference~\cite{landsberg1985}. Recently, high quality data from NA60 experiment measured the $q^2$-dependent form factor of the Dalitz decay $\omega \rightarrow \pi^0 \mu^+\mu^-$~\cite{na60-2009}. Using the usual pole approximation $F(q^2) = 1/(1-\frac{q^2}{\Lambda^2})$ for the form factor, the $\Lambda^{-2}$ has been found to be $2.24 \pm 0.06$ GeV$^{-2}$, which strongly deviates from the expectation of vector meson dominance(VMD)~\cite{iva2011,le2010,faessler2000,klingl,kopp}. The form factor showed a relative increase close to the kinematic cut-off by a factor of $\sim$ 10~\cite{na60-2009}. For the decay of $\phi \rightarrow \eta e^+e^-$, the SND experiment has looked for the $m_{e^+e^-}$ invariant mass distribution with 213 events~\cite{snd-phi}, and measured the form factor slope $\Lambda^{-2}$ to be $3.8\pm 1.8$ GeV$^{-2}$. Most recently, KLOE-2 has selected 7000 $\phi \rightarrow \eta e^+e^-$ events with $\eta \rightarrow \pi^+\pi^-\pi^0$ using a sample of 739 pb$^{-1}$ on the $\phi$ peak~\cite{kloe-hadron11}. A preliminary fit to the $m_{e^+e^-}$ indicates the possibility to reach a 5\% error on the form factor slope~\cite{kloe-hadron11}. These theoretical and experimental investigations of the EM Dalitz decays of light vector mesons motivate us to study the rare charmonium decays $J/\psi \rightarrow P l^+l^-$ ($P= \pi^0$, $\eta$ or $\eta^\prime$). The measurements of the $q^2$-dependent form factors will provide useful information on the interaction of the charmonium states with electromagnetic field. In particular, with more phase space, the transition between $J/\psi$ and the pseudoscalar can be explored over a large region of momentum transfer, which will be used to test QCD prediction cleanly. The decay rates of the $J/\psi \rightarrow \gamma \pi^0$, $\gamma \eta$ and $\gamma \eta^\prime$ are $(3.49^{+0.33}_{-0.30})\times 10^{-5}$, $(1.104\pm0.034)\times 10^{-3}$ and $(5.28\pm0.15)\times 10^{-3}$~\cite{pdg2010}, respectively. From a direct estimation according to the ratios of $\omega$ and $\phi$ Dalitz decays to the corresponding radiative decays, the expected Dalitz decay rates could reach $\sim 10^{-7}$ for $J/\psi \rightarrow \pi^0 e^+e^-$ and $\sim 10^{-5}$ for $J/\psi \rightarrow \eta e^+e^-$ and $\eta^\prime e^+e^-$, respectively. It is also interesting to search for dark photon, the light $U$ boson, in $J/\psi \rightarrow P U$ and $U\rightarrow l^+l^-$, in which the virtual photon is replaced by the light on-shell dark photon $U$. The light U boson may couple to the SM charged particles with a much suppressed coupling which has been considered in various contexts~\cite{dark-1,dark-2,dark-3, dark-4,dark-5,dark-6}. We consider the new Abelian gauge group $U(1)_d$ which has a gauge-invariant kinetic mixing with the SM hypercharge $U(1)_Y$~\cite{u-1,u-2,u-3}. After electroweak symmetry breaking, we have the Lagrangian: \begin{equation} {\cal L} = {\cal L}_{SM} + \epsilon_Y F^{Y,\mu\nu}F^d_{\mu\nu} + m^2_{U} A^{d, \mu}A^d_{\mu}, \label{mixingl} \end{equation} where ${\cal L}_{SM}$ is the SM Lagrangian, $F^Y_{\mu\nu}$ and $F^d_{\mu\nu}$ are the field strength for the SM gauge boson $B$ and $U$ boson, respectively, $A^d$ is the gauge field of a massive dark $U(1)_d$ gauge group~\cite{u-2}. The second term in Eq.(\ref{mixingl}) is kinetic mixing operator, and $\epsilon \sim 10^{-8} - 10^{-2}$ is generated at any scale by loops of heavy fields charged under both $U(1)$s. In a supersymmetry theory, the kinetic mixing operator induces a mixing between the $D$-terms associated with $U(1)_d$ and $U(1)_Y$. The hypercharge $D$-term gets a vacuum expectation value from the electroweak symmetry breaking and induces a weak-scale effective Fayet-Iliopoulos term for $U(1)_d$. Consequently, the $U(1)_d$ symmetry breaking scale is suppressed by loop factors or by $\sqrt{\epsilon}$, leading to MeV to GeV-scale $U$ boson mass~\cite{u-1,u-3}. The parameters of concern in this paper are $\epsilon$ and $m_{U}$. In the BESIII experiment, more than 1 billion $J/\psi$ ($\psi^\prime$) sample will be collected next year, the first look of these decay modes will be accessible~\cite{haibo-hadron11}. It will shed light on probing new physics beyond the standard model, such as possible $U$ boson with mass less than 300 MeV range~\cite{fayet1,fayet2,fayet3,li2010,kloe}, which may contribute to the process by replacing virtual photon. By looking at the spectrum of the dilepton, with huge data set in the BESIII experiment, one may see possible new physics contribution which will modify the shape of the di-lepton spectrum and total decay rate of the Dalitz decay process. In this paper, the full angular distribution and $q^2$-dependent rate are derived in the SM framework in Section~\ref{dalitz}. Using the pole approximation, the decay rates for the EM Dalitz decays of $J/\psi$ are estimated for the first time. An interesting consequence is that the dark photon can be probed in the low energy BESIII experiment. In Section~\ref{darkphoton}, we estimate the reach of the $U$ boson search in the $J/\psi \rightarrow P U$ and $U\rightarrow l^+l^-$ decay. \section{$q^2$-dependent decay rate of $J/\psi \rightarrow P l^+l^-$} \label{dalitz} The amplitude of the Dalitz decay $\psi \rightarrow P l^+l^-$ (hereafter, $\psi$ denotes $J/\psi$) has the Lorentz-invariant form: \begin{eqnarray} T = 4\pi \alpha f_{\psi P}\epsilon^{\mu\nu\rho\sigma} p_{\mu} q_{\nu} \epsilon_{\rho} \frac{1}{q^2} \bar{u}_1 \gamma_{\sigma}v_2, \label{eq:invariance} \end{eqnarray} where $f_{\psi P}$ is the form factor of the $\psi \rightarrow P$ transition; $q_\nu$ is the 4-momentum of the virtual photon or the total 4-momentum of $l^+l^-$ ($l= e$, $\mu$) system ; $q^2 = m^2_{l^+l^-}$ is the effective mass squared of the lepton pair; $p_\mu$ is the 4-momentum of the pseudoscalar meson; $\epsilon_\rho$ is the polarization 4-vector of the $\psi$; $\epsilon^{\mu\nu\rho\sigma}$ is the totally antisymmetric unity tensor. It is straightforward to obtain the magnitude of the amplitude squared as below: \begin{widetext} \begin{eqnarray} |T|^2 = 16\pi^2 \alpha^2 \frac{|f_{\psi P}(q^2)|^2}{q^4} \left[ 8(p\cdot q)^2 m_l^2 - 8p^2q^2m^2_l - 2p^2q^4 -8 (k_1 \cdot p)(k_2\cdot p)q^2+4(p\cdot q)^2q^2\right], \label{eq:ampl} \end{eqnarray} where $m_l$ is the lepton mass; $q=k_1 + k_2$, and $p$, $k_1$ and $k_2$ are 4-momenta of particles $P$, $l^+$ and $l^-$, respectively. The angular distribution of the differential decay width can be obtained as \begin{eqnarray} \frac{d\Gamma(\psi \rightarrow P l^+l^-)}{d q^2} &=&\frac{1}{3} \frac{\alpha^2}{256\pi^3 m_{\psi}^3} \frac{|f_{\psi P}(q^2)|^2}{q^2}\left(1-\frac{4m^2_l}{q^2}\right)^{1/2} \left[(m^2_{\psi}-m^2_P+q^2)^2-4m^2_{\psi}q^2\right]^{3/2}\times \nonumber \\ && \int d\Omega_3 d\Omega^*_1 \left[\left(1+\frac{4m^2_l}{q^2}\right) + \left(1-\frac{4m^2_l}{q^2} \right ) \mbox{cos}^2\theta^*_1 \right], \label{eq:dgammaan} \end{eqnarray} where $m_{\psi}$ and $m_P$ are the masses of the initial charmonium state and pseudoscalar meson; $d\Omega_3 = d\phi_3d(\mbox{cos}\theta_3)$ is the solid angle of $P$ in the rest frame of $\psi$ and $d\Omega^*_1 = d\phi^*_1 d(\mbox{cos}\theta^*_1)$ is the solid angle of one of the lepton pair in the rest frame of $l^+l^-$ system (the $z$ direction is defined as the momentum direction of $l^+l^-$ in the $\psi$ system); $\theta^*_1$ is the helicity angle of $l^+l^-$ system, which is defined as the angle between momentum direction of one of the lepton pair and direction of the $P$ meson in the rest frame of $l^+l^-$ system. By integrating the solid angles in Eq.~\ref{eq:dgammaan}, one can obtain the $q^2$-dependent differential decay width: \begin{eqnarray} \frac{d\Gamma(\psi \rightarrow P l^+l^-)}{d q^2} =\frac{1}{3} \frac{\alpha^2}{24\pi m_{\psi}^3} \frac{|f_{\psi P}(q^2)|^2}{q^2}\left(1-\frac{4m^2_l}{q^2}\right)^{1/2} \left(1+\frac{2m^2_l}{q^2}\right) \left[(m^2_{\psi}-m^2_P+q^2)^2-4m^2_{\psi}q^2\right]^{3/2}. \label{eq:dgamma} \end{eqnarray} \end{widetext} For the corresponding radiative decay of $\psi \rightarrow P \gamma$, the decay width can be obtained as : \begin{eqnarray} \Gamma(\psi \rightarrow P \gamma ) = \frac{1}{3} \frac{\alpha(m^2_{\psi} - m^2_P)^3} {8 m^3_{\psi}}|f_{\psi P}(0)|^2. \label{eq:gammap} \end{eqnarray} From Eqs.~(\ref{eq:dgamma}) and~(\ref{eq:gammap}) the $q^2$-dependent differential decay width in the $\psi \rightarrow P l^+l^-$ decay normalized to the width of the corresponding radiative $\psi \rightarrow P \gamma$ is derived : \begin{widetext} \begin{eqnarray} \frac{d\Gamma(\psi \rightarrow P l^+l^-)}{d q^2 \Gamma (\psi \rightarrow P \gamma)} &=& \frac{\alpha}{3\pi} \left|\frac{f_{\psi P}(q^2)}{f_{\psi P}(0)}\right|^2 \frac{1}{q^2} \left(1-\frac{4m^2_l}{q^2}\right)^{1/2} \left(1+\frac{2m^2_l}{q^2}\right) \left[\left(1+\frac{q^2}{m^2_{\psi}-m^2_P}\right)^2 - \frac{4m^2_{\psi}q^2}{(m^2_{\psi}-m^2_P)^2}\right]^{3/2}\nonumber \\ &=& |F_{\psi P}(q^2)|^2 \times [\mbox{QED}(q^2)], \label{eq:dgamman} \end{eqnarray} \end{widetext} where the normalized form factor for the $\psi \rightarrow P$ transition is defined as $ F_{\psi P}(q^2)\equiv f_{\psi P}(q^2)/f_{\psi P}(0) $, and the normalization is $F_{\psi P}(0) =1$. The form factor defines the electromagnetic properties of the region in which $\psi$ is converted into pseudoscalar. By comparing the measured spectrum of the lepton pairs in the Dalitz decay with QED calculations for point-like particles, it is possible to determine experimentally the transition form factor in the time-like region of the momentum transfer~\cite{landsberg1982,landsberg1985}. Namely, the form factor can modify the lepton spectrum as compared with that obtained for point-like particles. For the decays accompanied by the production of the electron-positron pair, we should note that the radiative corrections proportional to $\alpha \mbox{ln}^2(q^2/m^2_l)$ will be important. We will not discuss the high order QED corrections in this analysis since the data sample in the BESIII experiment is still small, and BESIII is expected to see the first signal for the effect at leading order. In addition to that, the external conversion of the $\gamma$ from the radiative decay of $\psi \rightarrow P \gamma$ will make the analysis more complicated, however, at the BESIII the external conversion rate could be up to 2\%, and the invariant mass of the $m_{e^+e^-}$ will form a narrow peak at 20-40 MeV, which will not really affect the slope shape of the dilepton. For the decays accompanied by the production of the muon pairs the radiative corrections and external radiation effects are negligibly small. To estimate the order of magnitude, one may use the Vector Dominance Model (VDM), in which the hadronic EM current is proportional to vector meson fields~\cite{gellman61,bauer78}. Hence the VDM predicts a growth of the transition form factors with increasing dilepton mass. The form factor may be parameterized in the simple pole approximation as \begin{eqnarray} F_{\psi P}(q^2) = \frac{1}{1-\frac{q^2}{\Lambda^2}}, \label{eq:ff} \end{eqnarray} where the pole mass $\Lambda$ should be the mass of the vector resonance near the energy scale of the decaying particle according to the VDM model. In $\psi$ decay the pole mass could be the mass of $\psi^\prime$. By assuming the pole approximation and taking $\Lambda = m_{\psi^\prime}$, in Fig.~\ref{fig:dd} we show the differential decay rates for $\psi \rightarrow \pi^0 l^+ l^-$, $\eta l^+l^-$ and $\eta^\prime l^+l^-$, respectively. The decay rates for $\psi \rightarrow \pi^0 l^+ l^-$, $\eta l^+l^-$ and $\eta^\prime l^+l^-$ are estimated and presented in table \ref{t2}. To study the dependence of the decay rates on the value of the pole mass, we varied the pole mass. The results for $J/\psi \to\eta\mu^+\mu^-$ are shown in Figs.~\ref{poledepen1} and \ref{poledepen2} as an example. The cases for the other decay modes considered in this work are similar. Both the differential and total decay rates are not sensitive to the value of the pole mass if the pole mass is in the range $\Lambda^2>q^2_{\rm max}=(m_\psi-m_p)^2$ in the Dalitz decay process of $J/\psi$. The reason can be well understood. The dominant contribution to the decay rate comes from the region of small value of $q^2$. For the pole mass with large value, $q^2/\Lambda ^2$ is small, therefore this term cannot give large effect. This is different from the case for the light vector meson Dalitz decays. Because the decay rates are not sensitive to the pole mass in the form factor, the predicted decay rates in Table \ref{t2} are more reliable. Comparing these predictions with experimental measurement should make sense. In the BESIII experiment, more than 1 billion $J/\psi$ ($\psi^\prime$) sample will be collected in year 2012, the first look of these decay modes will be accessible~\cite{haibo-hadron11}. \begin{center} \begin{figure}[ht] \scalebox{0.6}{\epsfig{file=fig1.eps}} \caption{The differential decay rates for $\psi \rightarrow P l^+ l^-$, where the solid curve is for $\psi \rightarrow \pi^0 l^+ l^-$, the dashed curve for $\psi \rightarrow \eta l^+ l^-$ and the dotted for $\psi \rightarrow \eta^\prime l^+ l^-$. (a) is for the case that the lepton pair is $e^+e^-$, while (b) for the case $\mu^+\mu^-$ . In (a), only part of the phase space for $m_{e^+e^-}$ is shown for demonstration purpose since most of the phase space is accumulated near small $m_{e^+e^-}$ region. } \label{fig:dd} \end{figure} \end{center} \begin{center} \begin{table}[h] \caption{ The estimated decay rates for $\psi \rightarrow \pi^0 l^+ l^-$, $\eta l^+l^-$ and $\eta^\prime l^+l^-$ based on Eq.(~\ref{eq:dgamman}) by assuming pole approximation and $\Lambda = m_{\psi^\prime}$. The error on the decay rate is from the measured error of $\psi \rightarrow P \gamma$ which is used as normalization. } \label{t2} \begin{tabular}{c|c|c}\hline Decay mode & $e^+e^-$ & $\mu^+\mu^-$ \\ \hline $\psi \rightarrow \pi^0 l^+ l^-$ & $(3.89^{+0.37}_{-0.33}) \times 10^{-7}$ & $(1.01^{+0.10}_{-0.09} )\times 10^{-7}$ \\ \hline $\psi \rightarrow \eta l^+ l^-$ & $(1.21\pm 0.04 )\times 10^{-5}$ & $(0.30\pm 0.01) \times 10^{-5}$ \\ \hline $\psi \rightarrow \eta^\prime l^+ l^-$ & $(5.66\pm 0.16)\times 10^{-5}$ & $(1.31\pm 0.04) \times 10^{-5}$ \\ \hline \end{tabular} \end{table} \end{center} \begin{center} \begin{figure}[ht] \scalebox{0.6}{\epsfig{file=fig2.eps}} \caption{The differential decay rate for $\psi \rightarrow \eta \mu^+ \mu^-$, where the solid curve is for the pole mass taken as $\Lambda=m_{\psi\prime}$, the dashed curve for $\Lambda=3.0\;{\rm GeV}$ and the dotted for $\Lambda=4.0\;{\rm GeV}$.} \label{poledepen1} \end{figure} \end{center} \begin{center} \begin{figure}[ht] \scalebox{0.6}{\epsfig{file=fig3.eps}} \caption{The decay rate for $\psi \rightarrow \eta \mu^+ \mu^-$ with the variation of the pole mass $\Lambda$.} \label{poledepen2} \end{figure} \end{center} \section{Reach of U-boson search in $J/\psi \rightarrow P U$, $U\rightarrow l^+l^-$ decay.} \label{darkphoton} In this section, we discuss the constraints and discovery potential for the $U$ boson at Beijing electron-positron collider II (BEPCII) ~\cite{bepcii}. Since the $U$ boson couples mainly to the SM electromagnetic current~\cite{wang}. Its production at BEPCII is the same as that of photon, although with a much suppressed rate. Therefore, the process $J/\psi \rightarrow P \gamma^* \rightarrow Pl^+l^-$ will have a chance to look for U boson as well. The on-shell U boson will decay to a pair of leptons in $J/\psi \rightarrow PU$, leading to a signal of $Pl^+l^-$. The SM background $J/\psi \rightarrow P \gamma^* \rightarrow Pl^+l^-$, although large for this process, is not a severe problem as the kinematics of the signal are quite distinct. The invariant mass of the lepton pair is just within a single bin due to the tiny width of the vector $U$ boson and can be distinguished from the SM background. It will be interesting to look for a low mass, up to GeV scale, U boson in these modes. From Eq.~(\ref{eq:dgamman}), the number of background events in the window of $\delta q^2$ (resolution of the $q^2$) around $q^2 = m^2_{l^+l^-} = m^2_{U}$ is about \begin{eqnarray} N_B &=& N_{\psi}\int^{q^2_{\text{max},i}}_{q^2_{\text{min}, i}} \frac{d\Gamma(\psi \rightarrow P l^+l^-)}{d q^2 \Gamma (\psi \rightarrow P \gamma)}dq^2 \times BR(\psi \rightarrow P\gamma) \nonumber \\ &\approx& N_{\psi}\frac{d\Gamma(\psi \rightarrow P l^+l^-)}{d q^2 \Gamma (\psi \rightarrow P \gamma)}\delta q^2 \times BR(\psi \rightarrow P\gamma) \nonumber \\ &=& N_{\psi}|F_{\psi P}(q^2)|^2 \times [\mbox{QED}(q^2)] \delta q^2 \times BR(\psi \rightarrow P\gamma),\nonumber \\ \label{eq:bgk} \end{eqnarray} where $q^2_{\text{max},i}$ ($q^2_{\text{min}, i}$) is the upper (lower) value of the $i$-$th$ $q^2$ bin; $N_{\psi}$ is the total number of $\psi$ decay events; and $BR(\psi \rightarrow P\gamma)$ is the branching fraction from PDG~\cite{pdg2010}. The size of the bin is the window size $\delta q^2$ which is obtained from the resolution functions of the $m_{l^+l^-}$ in Eqs. (4) and (5) in Ref~\cite{li2010} based on the BESIII Monte Carlo simulations. By replacing the photon by $U$ boson in $\psi\rightarrow P\gamma$, the signal rate can be estimated to be $BR(\psi \rightarrow PU) \approx \epsilon^2 BR(\psi \rightarrow P\gamma)$, where $\epsilon$ was defined in Eq.~(\ref{mixingl}). Thus, the expected number of signal events is about \begin{eqnarray} N_S = N_{\psi}\times \epsilon^2 BR(\psi \rightarrow P\gamma)BR(U\rightarrow l^+l^-), \label{eq:sig} \end{eqnarray} where we assume $BR(U\rightarrow l^+l^-)=1$ in this study. Combining Eqs.~(\ref{eq:bgk}) and (\ref{eq:sig}), We estimate the expected numerical results based on the significance \begin{eqnarray} \frac{S}{\sqrt{B}} &=& \frac{N_S}{\sqrt{N_B}} \nonumber \\ &=& \sqrt{N_{\psi}} \frac{\epsilon^2 \sqrt{BR(\psi \rightarrow P\gamma)}BR(U\rightarrow l^+l^-)}{\sqrt{|F_{\psi P}(q^2)|^2 \times [\mbox{QED}(q^2)] \delta q^2}} \label{eq:ratio} \end{eqnarray} Therefore, with 1 billion $\psi$ events, the reach for U-boson searching can be $\epsilon \sim 10^{-2}$ - $10^{-3}$ in the $\psi \rightarrow PU$ decays. In Figs.~\ref{eep} and \ref{mumup}, we show the reach of the parameter $\epsilon$ by defining $S/\sqrt{B} = 5$ for different mass of $U$ boson. \begin{figure}[htb] \includegraphics[width=0.8\columnwidth]{eu_Pee_10b.eps} \caption{Illustrative plot of the reach of vector boson at BES-III in the channel of $\psi \rightarrow PU$ (solid curve for $P=\pi^0$; dot-dashed curve for $P=\eta$ and dashed curve for $P=\eta^\prime$, respectively), where $U$ decay into $e^+e^-$. } \label{eep} \end{figure} \begin{figure}[htb] \includegraphics[width=0.8\columnwidth]{eu_Pmumu_10b.eps} \caption{The same as Fig.~\ref{eep} with $U$ decay into $\mu^+ \mu^-$.} \label{mumup} \end{figure} \section{Summary} In summary, the EM Dalitz decays of $\psi \rightarrow P l^+l^-$ are studied in this paper. We demonstrate the differential decay rate as both $q^2$-dependent rate and angular dependent rate explicitly. By assuming simple pole approximation the decay rates for $\psi \rightarrow P l^+l^-$ are estimated for the first time. The estimated Dalitz decay rates could reach $\sim 10^{-7}$ for $\psi \rightarrow \pi^0 l^+l^-$ and $\sim 10^{-5}$ for $\psi \rightarrow \eta l^+l^-$ and $\eta^\prime l^+l^-$, respectively. They will be accessible in the BESIII experiment with data sample of 1 billion $\psi$ decay events. Especially, the $q^2$-dependent differential decay rate can be measured by looking at the invariant mass of the lepton pairs. In the BESIII experiment, these measurements will be important for us to understand the interaction of vector charmonium states with photon, as well as to probe new physics beyond the standard model. we have investigated the signatures of a hidden $U(1)_d$ sector at the BESIII experiment in $\psi\rightarrow PU$ decays, and find that the BESIII should have an intrinsic sensitivity to the kinetic mixing parameter $\epsilon$ in the range of $10^{-2}$ - $10^{-4}$, which depends on the mass values of the $U$ boson. One of the authors (H.~B.~Li) would like to thank Jianping Ma, J\'er\^ome Charles and S\'ebastien Descotes-Genon for helpful discussions. This work is supported in part by the National Natural Science Foundation of China under contract Nos. 11125525, 10575108, 10975077, 10735080, and by the Fundamental Research Funds for the Central Universities No. 65030021.
\section{Introduction} Accretion is one of the most important physical processes in astrophysics. It is widely accepted that the accreting matter toward the central object is a source of power active galactic nuclei (AGN), and galactic X-ray sources (see for a review e.g. Frank et al. 2002). This idea is also well-applicable to interpret many observations of astrophysical phenomena, such as, prototype stellar objects (Chang \& Choi 2002), symbiotic stars (Lee \& Park 1999), gamma-ray bursts (Brown et al. 2000). There are some types of accretion flows that a significant fraction of the generated heat by dissipation processes retains in the fluid rather than being radiated away, have been the subject of considerable attention in recent years (Ogilvie 1999; Chang 2005, Akizuki \& Fukue 2006, Khesali \& Faghei 2009, Faghei 2011). These advection dominated accretion flows (ADAF) place an intermediate position between the spherically symmetric accretion flow of non-rotating fluid (Bondi 1952) and the cool thin disc of classical accretion disc theory (e. g. Pringle 1981). These types of accretion flows have been widely applied to explain observations of the galactic black hole candidate (e. g. Narayan et al. 1996, Hameury et al. 1997), the spectral transition of Cyg x-1 (Esin 1996) and multi-wavelength spectral properties of Sgr A$^*$ (Narayan \& Yi 1995; Manmoto et al. 2000; Narayan et al. 1997). The X-ray observations of black holes imply that they are capable of accreting gas under a variety of flow configurations. In particular, observational evidences confirm existence of hot accretion flow, contrasted with the classical cold and thin accretion disc scenario (Shakura \& Sunyaev 1973). Hot accretion flow can be found in the population of supermassive black holes in galactic nuclei and during quiescent of accretion onto stellar-mass black holes in X-ray transients (e.g., Narayan et al. 1998a; Lasota et al. 1996; Di Matteo et al. 2000; Esin et al. 1997, 2001; Menou et al. 1999; see Narayan et al. 1998b; Melia \& Falcke 2001; Narayan 2002; Narayan \& Quataert 2005 for reviews). \textit{Chandra} observations provide tight constraints on the density and temperature of gas at or near the Bondi capture radius in Sgr A$^*$ and several other nearby galactic nuclei. Tanaka \& Menou (2006) used these constraints (Loewenstein et al. 2001; Baganoff et al. 2003; Di Matteo et al. 2003; Ho et al. 2003) to calculate the mean free path for the observed gas. They suggested that accretion in these systems will be proceeded under the weakly collisional condition. Furthermore, they suggested that thermal conduction can be as a possible mechanism by which the sufficient extra heating is provided in hot advection dominated accretion flows. Generally, semi-analytical studies of hot accretion flows with thermal conduction have been related to steady state models (e. g. Tanaka \& Menou 2006; Johnson \& Quataert 2007; Shadmehri 2008; Abbassi et al. 2008, 2010; Ghanbari et al. 2009), and dynamics of such systems have been studied in simulation models (e. g. Sharma et al. 2008; Wu et al. 2010). For example, Tanaka \& Menou (2006) have carried out a related analysis and found the accretion flow can spontaneously produce thermal outflows driven in part by conduction. Their analysis is two-dimensional but self-similar in radius. Their assumption of self-similarity enforces a density profile that varies as $r^{-3/2}$, whereas simulations of ADAFs consistently find density profiles shallower than this (e.g., Stone et al. 1999; Igumenshchev \& Abramowicz 1999; Stone \& Pringle 2001; Hawley \& Balbus 2002; Igumenshchev et al. 2003). Johnson \& Quataert (2007) studied the effects of electron thermal conduction on the properties of hot accretion flows, under the assumption of spherical symmetry. Since, electron heat conduction is important for low accretion rate systems, thus their model is applicable for Sgr A$^*$ in the Galactic centre. They show that heat conduction leads to supervirial temperatures, implying that conduction significantly modifies the structure of the accretion flow. Their model similar to Tanaka \& Menou (2006) was the steady state, but they solved their equations numerically. As mentioned, semi-analytical studies of hot accretion flows with thermal conduction have been in a steady state. Thus, it will be interesting to study \textit{dynamics} of such systems. Ogilvie (1999) by the unsteady self-similar method studied time-dependence of quasi-spherical accretion flow without thermal conduction. The solutions of Ogilvie (1999) provided some insight into the dynamics of quasi-spherical accretion and avoided many of the limits of the steady self-similar solution. In this research, we want to explore how thermal conduction can affect the dynamics of a rotating and accreting viscous gas. We answer this question by solving Ogilvie (1999) model that is affected by thermal conduction. This paper is organized as follows. In Section 2, we define the general problem of constructing a model for hot quasi-spherical accretion flow. In Section 3, we use the unsteady self-similar method to solve the integrated equations that govern the dynamical behaviour of the accreting gas, and numerical study of the model is brought in this section, too. We will present a summary of the model in Section 4. \section{Basic Equations} We start with the approach adopted by Ogilvie (1999), who studied quasi-spherical accretion flows without thermal conduction. Thus, we derive the basic equations that describe the physics of accretion flow with thermal conduction. We use the spherical coordinates $(r, \theta, \phi)$ centred on the accreting object and make the following standard assumptions: \begin{enumerate} \item The gravitational force on a fluid element is characterized by the Newtonian potential of a point mass, $\Psi=-G M_* / r$, with $G$ representing the gravitational constant and $M_*$ standing for the mass of the central star. \item The written equations in spherical coordinates are considered in the equatorial plane $\theta=\pi/2$ and terms with any $\theta$ and $\phi$ dependence are neglected, hence all quantities will be expressed in terms of spherical radius $r$ and time $t$. \item For simplicity, self-gravity and general relativistic effects have been neglected. \end{enumerate} Under these assumptions, the dynamics of accretion flow describes by the following equations: \\ the continuity equation \begin{equation}\label{a1} \frac{\partial\rho}{\partial t} +\frac{1}{r^{2}}\frac{\partial}{\partial r}(r^{2}\rho v_{r})=0, \end{equation} the radial force equation \begin{equation}\label{a2} \frac{\partial v_{r}}{\partial t}+v_{r}\frac{\partial v_{r}}{\partial r}=r(\Omega^{2}-\Omega_K^{2})-\frac{1}{\rho}\frac{\partial p}{\partial r}, \end{equation} the azimuthal force equation \begin{equation}\label{a3} \rho\left[\frac{\partial}{\partial t}(r^{2}\Omega)+v_{r}\frac{\partial}{\partial r}(r^{2}\Omega)\right]= \frac{1}{r^{2}}\frac{\partial}{\partial r}\left[\nu\rho r^{4}\frac{\partial \Omega}{\partial r}\right], \end{equation} the energy equation \begin{eqnarray} \nonumber \frac{1}{\gamma-1}\left[\frac{\partial p}{\partial t}+v_r\frac{\partial p}{\partial r}\right]+ \frac{\gamma}{\gamma-1}\frac{p}{r^2}\frac{\partial}{\partial r}\left(r^2 v_r\right)=\\ Q_{vis}-Q_{rad}+Q_{cond}. \end{eqnarray} Here $\rho$ the density, $v_r$ the radial velocity, $\Omega$ the angular velocity, $\Omega_K[=(G M_*/r^3)^{1/2}]$ Keplerian angular velocity, $p$ the gas pressure, $\gamma$ is the adiabatic index, $\nu$ the kinematic viscosity coefficient and it is given as in Narayan \& Yi (1995a) by an $\alpha$-model \begin{equation} \nu = \alpha \frac{p_{gas}}{\rho\Omega_{K}}. \end{equation} The parameter of $\alpha$ is assumed to be a constant less than unity. The terms on the right-hand side of the energy equation, $Q_{vis}$ is the heating rate of the gas by the viscous dissipation, $Q_{rad}$ represents the energy loss through radiative cooling, and $Q_{cond}$ is the transported energy by thermal conduction. For the right-hand side of the energy equation, we can write \begin{equation} Q_{adv}=Q_{vis}-Q_{rad}+Q_{cond} \end{equation} where $Q_{adv}$ is the advective transport of energy. We employ the advection factor, $f=1-Q_{rad}/Q_{vis}$, that describes the fraction of the dissipation energy which is stored in the accretion flow and advected into the central object rather than being radiated away. The advection factor of $f$ in general depends on the details of the heating and radiative cooling mechanism and will vary with position (e.g. Watari 2006, 2007; Sinha et al. 2009). However, we assume a constant $f$ for simplicity. Clearly, the case $f = 1$ corresponds to the extreme limit of no radiative cooling and in the limit of efficient radiative cooling, we have $f = 0$. In a collisional plasma, mean free path for electron energy exchange, $\lambda$, is shorter than temperature scale height, $L_T=T/|\nabla T|$, and thus the heat flux due to thermal conduction can be written as \begin{equation} F_{cond}=-\kappa \nabla T, \end{equation} where $\kappa$ is the thermal conductivity coefficient. Thermal conductivity in a dense, fully ionized gas is given by the Spitzer (1962) formula, \begin{equation} \kappa=\frac{1.84 \times 10^{-5} T_e^{5/2}}{\ln \Lambda}, \end{equation} where $T_e$ is the electron temperature ($T_e=T$ for a one-temperature plasma) and $\ln \Lambda$ is Coulomb logarithm that for $T > 4.2 \times 10^5 K $ is \begin{equation} \ln \Lambda=29.7+\ln n^{-1/2} (T_e/10^6 K). \end{equation} The heat is conducted by the electron, and equation (8) includes the effect of the self-consistent electric required to maintain the electric current at zero; this reduces $\kappa$ by a factor of about $0.4$ from the value it would otherwise have (Cowie \& McKee 1977, hereafter CM77). As noted in the introduction, the inner regions of hot accretion flows are, in many cases, collisionless with electron mean free path due to Coulomb collision larger than the radius (e. g. Tanaka \& Menou 2006). When the mean free path of an electron becomes comparable to or larger than the temperature gradient scale $\lambda \gtrsim T/|\nabla T|$, equation (7) for the heat flux is no longer valid; CM77 described this effect as \textit{saturation}. The maximum heat flux in a plasma can be expressed as $(3/2) n_e k T_e v_{char}$, where $v_{char}$ is a characteristic velocity which one might expect to be the order of the electron thermal velocity (Parker 1963). Assuming Maxwellian distribution for heat source, the characteristic velocity can be written as (Williams 1971; CM77) \begin{equation} V_{char}=(\frac{8}{9\,\pi})^{1/2}\,(\frac{k\,T}{m_e})^{1/2}. \end{equation} Similar to CM77, we assume that the heat flux is reduced by the same factor of $0.4$ in the saturated case as in the classical (collisional) case so that the saturated heat flux is \begin{equation} F_{sat}=0.4\,n_e k T_e\,\sqrt{\frac{2 k T_e}{\pi m_e}}. \end{equation} CM77 showed that the saturated heat flux is significantly less than conjectured by Parker (1963). Thus, in order to explicitly allow for uncertainty in the estimate of $F_{sat}$, they introduced a factor of $\phi_s$, which was less than unity and rewrote equation (11) as \begin{equation} F_{sat}=5 \phi_s \rho c_s^3=5 \phi_s p \sqrt{\frac{p}{\rho}}, \end{equation} where $c_s$ is sound speed, which is defined as $c_s^2=p / \rho$. The factor of $\phi_s$ is called as saturation constant (CM77). Now, the viscous heating rate and the energy transport by thermal conduction are expressed as \begin{equation} Q_{vis}=\nu\rho r^2 \left(\frac{\partial \Omega}{\partial r}\right)^2 \end{equation} \begin{equation} Q_{cond}=-\frac{1}{r^2}\frac{\partial}{\partial r} \left(r^2 F_{sat}\right) \end{equation} By using equations (13) and (14) for the advective transport of energy, we can write \begin{equation} Q_{adv}=f\nu\rho r^2 \left(\frac{\partial \Omega}{\partial r}\right)^2-\frac{1}{r^2}\frac{\partial}{\partial r} \left(r^2 F_{sat}\right) \end{equation} The mass accretion rate in a qausi-spherical accretion flow can be written as \begin{equation} \dot{M}(r,t)=-4\pi r^2 \rho v_r. \end{equation} We will use this quantity in the next section, and will investigate effects of saturation constant and viscous parameter on it. \section{Self-Similar Solutions} \subsection{analysis} Tanaka \& Menou (2006) solved essentially equations (1)-(4) for the case of a steady, radially self-similar flow. Here, we will try to find unsteady self-similar solutions for these equations. Thus, we introduce a similarity variable $\xi$ and assume that each physical quantity is given by the following form: \begin{equation} \xi=r(G M_* t^2)^{-1/3}, \end{equation} \begin{equation} \rho(r,t)=R(\xi) (\dot{M}_0 / G M_*) t^{-1}, \end{equation} \begin{equation} p(r,t)=\Pi(\xi) (\dot{M}_0 / (G M_*)^{1/3}) t^{-5/3}, \end{equation} \begin{equation} v_r(r,t)=V(\xi) (G M_*)^{1/3} t^{-1/3}, \end{equation} \begin{equation} \Omega(r,t)=\omega(\xi) t^{-1}, \end{equation} \begin{equation} \dot{M}(r,t)=\dot{M}_0\dot{m}(\xi), \end{equation} where $\dot{M}_0$ is a constant and its value can be obtained by typical values of the system. In addition, we assumed that $\dot{M}(r,t)$ under similarity transformations is a function of $\xi$ only (Khesali \& Faghei 2008, 2009). Substitution of above transformations into the basic equations (1)-(4), yields dimensionless equations below, \begin{equation}\label{a20} \left(V-\frac{2\xi}{3}\right)\frac{dR}{d\xi}-R=-\frac{R}{\xi^2} \frac{d}{d\xi}\left(\xi^2V\right), \end{equation} \begin{eqnarray}\label{a21} \left(V-\frac{2\xi}{3}\right)\frac{dV}{d\xi}-\frac{V}{3}= \xi(\omega^{2}-\xi^{-3})-\frac{1}{R}\frac{d\Pi}{d\xi}, \end{eqnarray} \begin{eqnarray}\label{a22} \nonumber R\left[\left(V-\frac{2\xi}{3} \right)\frac{d}{d\xi}\left(\xi^{2}\omega\right) +\frac{1}{3}\left(\xi^{2}\omega\right)\right]~~~~~~~~~~~~~~~~~~~~~\\ =\frac{\alpha}{\xi^{2}}\frac{d}{d\xi} \left[\Pi\xi^{11/2}\frac{d\omega}{d\xi}\right], \end{eqnarray} \begin{eqnarray} \nonumber\frac{1}{\gamma-1}\left[\left(V-\frac{2\xi}{3}\right)\frac{d\Pi}{d\xi}-\frac{5}{3}\Pi\right]+ \frac{\gamma}{\gamma-1}\frac{\Pi}{\xi^2}\frac{d}{d\xi}\left(\xi^2V \right)\\ = \alpha f \Pi \xi^{7/2}\left(\frac{d\omega}{d\xi}\right)^2 -\frac{5 \phi_s}{\xi^2} \frac{d}{d \xi} \left(\xi^2 \Pi \sqrt{\frac{\Pi}{R}} \right). \end{eqnarray} These equations provide a fourth-order system of non-linear ordinary differential equations that must be solved numerically. \subsection{Inner limit} An appropriate asymptotic solution as $\xi\rightarrow0$ is the form as \begin{equation} R(\xi) \sim \xi^{-3/2} (R_{0}+R_{1} \xi + \cdot \cdot \cdot), \end{equation} \begin{equation} \Pi(\xi) \sim \xi^{-5/2} (\Pi_{0}+\Pi_{1} \xi + \cdot \cdot \cdot), \end{equation} \begin{equation} V(\xi) \sim \xi^{-1/2} (V_{0}+V_{1} \xi + \cdot \cdot \cdot), \end{equation} \begin{equation} \omega(\xi) \sim \xi^{-3/2} (\omega_{0}+\omega_{1} \xi + \cdot \cdot \cdot), \end{equation} in which undetermined coefficients of $R_0$, $R_1$, $\Pi_0$, and etc must be specified. By substituting above relations in equations (23)-(26) and choosing the significant sentences, we can write \begin{equation} V_0^2+\frac{5 \Pi_0}{R_0}-2+2\,\omega_0^2\approx 0, \end{equation} \begin{equation} 2\,R_0\,V_0 +3\,\alpha\,\Pi_0\approx0, \end{equation} \begin{equation} 6\,V_0\,(\gamma-\frac{5}{3})+(\gamma-1)\left[20\phi_s \sqrt{\frac{\Pi_0}{R_0}}-9 \alpha f \omega_0^2 \right]\approx0. \end{equation} Also, the dimensionless mass accretion rate, $\dot{m}(\xi)$, under the above asymptotic solution becomes \begin{equation} \dot{m}_{in}\approx-4 \pi R_0 V_0, \end{equation} where $\dot{m}_{in}$ is the value of $\dot{m}$ at $\xi_{in}$, where $\xi_{in}$ is a point near to the centre. After algebraic manipulations for equations (31)-(34), we obtain an algebraic equation for $R_0$: \begin{eqnarray} \nonumber R_0^2 -\frac{10}{27} \frac{\phi_s}{\alpha f} \sqrt{\frac{6 \dot{m}_{in}}{\alpha \pi}}\, R_0^{3/2} -\frac{5}{12} \frac{\dot{m}_{in}}{\alpha \pi} \times \\ \left( 1-\frac{2}{5 f} \frac{\gamma-5/3}{\gamma-1} \right) R_0 -\frac{1}{32} \left(\frac{\dot{m}_{in}}{\pi}\right)^2\approx0, \end{eqnarray} and the rest of the physical variables are \begin{equation} \Pi_0\approx\frac{\dot{m}_{in}}{6\pi\alpha}, \end{equation} \begin{equation} V_{{0}}\approx-\frac{\dot{m}_{in}}{4\pi R_0}, \end{equation} \begin{equation} \omega_0^2\approx1-\frac{5}{12} \frac{\dot{m}_{in}}{\pi\alpha R_0} \left( 1+\frac{3}{40} \frac{\alpha \dot{m}_{in}}{\pi R_0} \right). \end{equation} Without thermal conduction, $\phi_s=0$, equation (35) can be solved analytically. Since, we want to consider systems with non-zero saturation constant, $\phi_s\neq 0$, we will solve this equation numerically. \input{epsf} \begin{figure*} \centerline { {\epsfxsize=7.5cm\epsffile{./Fig.1/R.eps} }{\epsfxsize=7.5cm\epsffile{./Fig.1/T.eps} } } \centerline { {\epsfxsize=7.5cm\epsffile{./Fig.1/V.eps} }{\epsfxsize=7.5cm\epsffile{./Fig.1/W.eps} } } \caption{Time-dependent self-similar solution for $\gamma=1.3$, $\alpha=0.1$, $f=1.0$, and $\dot{m}_{in}=0.001$. The solid, the dashed, and the short-dashed lines represent $\phi_s=0$, $0.05$, and $0.1$, respectively. } \end{figure*} \input{epsf} \begin{figure*} \centerline { {\epsfxsize=7.5cm\epsffile{./Fig.2/R.eps} }{\epsfxsize=7.5cm\epsffile{./Fig.2/T.eps} } } \centerline { {\epsfxsize=7.5cm\epsffile{./Fig.2/V.eps} }{\epsfxsize=7.5cm\epsffile{./Fig.2/W.eps} } } \caption{Same as Figure 1, but $\phi_s=0.03$. The solid, dahsed, and short-dashed lines represent $\alpha=0.05$, $0.1$, and $0.2$, respectively.} \end{figure*} \subsection{Numerical solution} If the value of $\xi_{in}$ is guessed, i.e. we take a point very near to the centre, the equations (23)-(26) by Runge-Kutta-Fehlberg fourth-fifth order method can be integrated from this point outward by the above expansion [(27)-(30)]. Examples of such solutions are presented in Figs 1-5. \subsubsection{The influences of saturation constant and viscous parameter on physical quantities} The delineated quantity of $\Pi/R$ in Figs 1 and 2 is the sound speed square in self-similar flow, which is rescaled in the course of time and represents the flow temperature. The profiles of $\Pi/R$ in Fig. 1 show the flow temperature decreased by adding saturation constant, $\phi_s$. Because, the generated heat by viscous dissipation can be transfered by thermal conduction. Furthermore, this temperature decrease is qualitatively consistent with simulation results of Sharma et al. (2008) and Wu et al. (2010). We know the viscous dissipation of the flow increases by adding $\alpha$ parameter. Thus, the temperature increased by viscous parameter confirmed by the temperature profiles in Fig 2. The density profiles show the gas density increased by adding the $\phi_s$ parameter. It can be due to temperature fall of fluid. Increasing density by adding saturation constant is another consistency of our results with simulations of Wu et al. (2010). Also, the density profiles show that it decreases by adding the viscous parameter. This also could be result of the temperature rising. The viscous turbulence in this paper is proportional to the gas temperature ($\nu \propto c_s^2 \propto T$). Thus, increase or decrease of temperature will affect dynamics of the accreting gas. As we said the temperature decreased by saturation constant implies that the viscous turbulence decreases, too. The decreasing of viscous turbulence reduces the effect of negative viscous torque in angular momentum equation. Thus, we expect the flow rotates faster by adding saturation constant that confirmed by the angular velocity profiles in Fig. 1. Since, the efficiency of the angular momentum transport decreases by adding the saturation constant, we expect the decrease of radial infall velocity that the radial velocity profiles in Fig. 1 confirm it. The efficiency of angular momentum transport increases by adding the viscous parameter of $\alpha$. Thus, we expect the flow rotates slower and accretes faster by adding the viscous parameter. The profiles of radial and angular velocities in Fig. 2 confirm them. \input{epsf} \begin{figure*} \centerline { {\epsfxsize=8.5cm\epsffile{./Fig.3/MdotPhiChange.eps}}{\epsfxsize=8.5cm\epsffile{./Fig.3/MdotAlphaChange.eps} } } \caption{Time-dependent self-similar solution of mass accretion rate. The input parameters in \textit{left panel} are same as Figure 1, but the solid, the dashed, and the short-dashed lines represent $\phi_s=0.01$, $0.02$, and $0.03$, respectively. The input parameters in \textit{right panel} are same as Figure 2. but $\phi_s=0.01$. } \end{figure*} \subsubsection{Mass accretion rate} The behaviour of mass accretion rate as a function of similarity variable $\xi$ for several values of the viscous parameter and saturation constant are plotted in Fig. 3. In the present model, the mass accretion rate is reduced by radius. While, the mass accretion rate in steady hot accretion flows is a constant (Tanaka \& Menou 2006). There are some researches in steady hot accretion flows that have studied power-law function of mass accretion rate (Shadmehri 2008; Abbassi et al. 2008). However, the mass accretion rate in their models is not dependent on important parameters such as saturation constant and viscous parameter. The profiles of mass accretion rate in Fig. 3 show that it is reduced by adding the saturation constant. This property is qualitatively consistent with numerical results of Johnson \& Quataert (2007). Also, the profiles of mass accretion rate imply that it increases by adding the viscous parameter of $\alpha$. This property is qualitatively consistent with previous works in accretion flows (e. g. Park 2009). \input{epsf} \begin{figure*} \centerline { {\epsfxsize=8.5cm\epsffile{./Fig.4/MuPhiChange.eps} }{\epsfxsize=8.5cm\epsffile{./Fig.4/MuAlphaChange.eps} } } \caption{Time-dependent self-similar solution of Mach number. The input parameters in \textit{left panel} are same as Figure 1, but $\alpha=0.5$ and the solid, the dashed, and the short-dashed lines represent $\phi_s=0.0$, $0.2$, and $0.3$, respectively. The input parameters in \textit{right panel} are same as Figure 1, but $\phi_s=0.05$ and the solid, the dashed, and the short-dashed lines represent $\alpha=0.2$, $0.3$, $0.4$, respectively. } \end{figure*} \subsubsection{Mach number} Here, it will be interesting to investigate the existence of the transonic point in hot accretion flow. The transonic point occurs in place that the amount of \textit{Mach number} becomes equal to unity. The Mach number referring to the reference frame is defined as (Gaffet \& Fukue 1983; Fukue 1984) \begin{equation}\label{a36} \mu\equiv\frac{v_{r}-v_{F}}{c_{s}}=\frac{V-n\xi}{S} \end{equation} where \begin{equation}\label{a37} v_{F}=\frac{dr}{dt}=n\frac{r}{t} \end{equation} is the velocity of the reference frame which is moving outward as time goes by, which the sound speed can be subsequently expressed as \begin{equation}\label{a35} c_{s}^2 \equiv \frac{p}{\rho}= S^2 (G M_*/ t)^{2/3} \end{equation} and, $S=\left(\Pi/R\right)^{1/2}$ the sound speed in self-similar flow is rescaled in the course of time. The Mach number introduced so far, represents the \emph{instantaneous} and \emph{local} Mach number of the unsteady self-similar flow. In steady self-similar solution (e. g. Tanaka \& Menou 2006), the Mach number does not vary by radii and is a constant. While, the Mach number in unsteady self-similar varies by radii (see Fig. 4). As seen in Fig. 4, there is a transonic point ($|\mu|=1$). The dependency of transonic point to saturation constant shows that this point moves inward by adding the parameter of $\phi_s$. Because, thermal conduction transfers the heat to larger radii, so the sound speed/temperature decreases by adding the saturation constant. In other words, whatever the radii smaller, the radial velocity relative to the sound speed larger. Also, the Mach number profiles show the transonic point decreasing by adding the viscous parameter which can be due to increase of radial velocity along with adding the viscous parameter. \input{epsf} \begin{figure*} \centerline { {\epsfxsize=8.5cm\epsffile{./Fig.5/R.eps} }{\epsfxsize=8.5cm\epsffile{./Fig.5/T.eps} } } \centerline { {\epsfxsize=8.5cm\epsffile{./Fig.5/V.eps} }{\epsfxsize=8.5cm\epsffile{./Fig.5/W.eps} } } \caption{Same as Figure 1, but $\phi_s=0.1$. The solid, the dashed, and the short-dashed lines represent $\gamma=1.3$, $1.4$, and $1.5$, respectively.} \end{figure*} \subsubsection{Comparison of steady and unsteady self-similar solutions} As a comparison between steady and unsteady self-similar solution, the physical quantities in unsteady self-similar solutions are divided into their radial dependence in steady self-similar solution then plotted in terms of $\xi$ in Fig. 5. Also, the effect of adiabatic index on hot accretion flow is investigated in Fig. 5. The delineated quantities ($R\, \xi^{3/2}$, $V\,\xi^{1/2}$, $\cdot$ $\cdot$ $\cdot$) in Fig 5 are constant in steady self-similar solutions of hot accretion flows (Tanaka \& Menou 2006; Shadmehri 2008; Abbassi et al. 2008; Ghanbari et al. 2009). While they vary with position in this research. Fig. 5 represents the density profile varies shallower than $\xi^{-3/2}$, that this property is qualitatively consistent with simulation results (e.g., Stone et al. 1999; Igumenshchev \& Abramowicz 1999; Stone \& Pringle 2001; Hawley \& Balbus 2002; Igumenshchev et al. 2003). The radial dependency of the temperature and radial velocity in unsteady self-similar solution show that they vary deeper than the steady self-similar solution. Also, the study of the angular velocity show that it varies shallower than $\xi^{-3/2}$. Thus, the physical quantities in the present model avoid the limits of the steady self-similar solution. The physical quantities profiles in Fig. 5 show that their radial dependency in the unsteady self-similar limited to the steady self-similar solution by adding adiabatic index $\gamma$. \section{Summary and Discussion} In hot accretion flows, the collision timescale between ions and electrons is longer than the inflow timescale. Thus, the inflow plasma is collisionless, and the transfer of energy by thermal conduction can be dynamically important. The low collisional rate of the gas is confirmed by direct observation, particularly in the case of the Galactic centre (Quataert 2004; Tanaka \& Menou 2006) and in the intracluster medium of galaxy clusters (Sarazin 1986). Here, we have investigated how thermal conduction affects dynamics of hot quasi-spherical accretion flows. We adopted the presented solutions by Ogilvie (1999) and Tanaka \& Menou (2006). Thus, we assumed that angular momentum transport is due to viscous turbulence and the $\alpha$-prescription is used for the kinematic coefficient of viscosity. We also assumed the flow does not have a good cooling efficiency and so a fraction of energy accretes along with matter on to the central object. The effect of thermal conduction is studied by a saturation form of it introduced by Cowie \& McKee (1977). To solve the equations that govern the dynamical behaviour of hot accretion flow, we have used unsteady self-similar solution. The effect of saturation constant and the viscous parameter on the present model is investigated. The solutions show that with the increase of conductivity, the equatorial density becomes denser and the temperature becomes lower. These results are qualitatively consistent with simulation results of Wu et al. (2010). Furthermore, the solutions show that by adding the saturation constant, the angular velocity becomes larger and the radial velocity decreases. The mass accretion rate is reduced by adding the saturation constant that is qualitatively consistent with the result of Johnson \& Quataert (2007). The solutions imply that the viscous parameter has opposite effects in comparison to saturation constant on physical quantities of the system. Also, the study of physical quantities of the present model in comparison to steady self-similar solution show that our results deviate from steady self-similar solution and do not have its limits. Here, we studied dynamical behaviour of hot accretion flow in one-dimensional approach ignored by latitudinal dependence of physical quantities. Although, some authors have shown that latitudinal dependence of physical quantities is important for structure and dynamics of hot accretion flow (Tanaka \& Menou 2006; Ghanbari et al. 2009; Wu et al. 2010). Thus, latitudinal behaviour of the present model can be investigated in other studies. \section*{Acknowledgments} I would like to thank the anonymous referee for very useful comments that helped me to improve the initial version of the paper.
\section{Introduction} The accurate prediction of the breakup of a many-particle system into multiple fragments is one of the most challenging problems in quantum mechanics. Not only the relative motion of the particles needs to be modeled, but also the internal structure of the target and the products need to be described accurately. This leads in many cases, to a high--dimensional Schr\"odinger equation posed on a huge domain. For example, in breakup or collisions of nuclear clusters the cross section depends on a delicate interplay of the forces that hold the clusters together and the forces between the clusters. The quantum state that describes the internal structure of a bound many particle system is often represented as linear combination of eigenstates of the quantum mechanical oscillator. These states form an $L^2$-basis which reduces the problem to finding the correct expansion coefficients. Examples are the correlated many electron state of a quantum dot \cite{PhysRevB.47.2244,sako2003confined,PhysRevB.80.045321} and the many nucleon state in nuclear physics \cite{filippov1980use,Vasilevsky2001a}. The various applications of the oscillator representation are discussed in \cite{dineykhan1995oscillator}. Such a representation is very efficient for tightly bound ground state or lowest excited states as any smooth potential well is close to parabolic shape near it's bottom. As a result, if the oscillator parameter is optimized to match this local parabolic potential, the lowest states of a system can be efficiently represented with only a few oscillator functions \cite{navratil}. However, the representation is inefficient to describe scattering or break-up processes. Scattering states are not square integrable and many oscillator states are required to represent the interaction and asymptotic region. Furthermore, many potential matrix elements need to be calculated and this results in a dense linear system that has a complexity of $N^3$ to arrive at the solution, where $N$ is the number of oscillator states used in the representation, omitting the cost of calculating the potential matrix elements. The $J$-matrix method offers a way of calculating cross sections and other scattering observables in the oscillator representation. It was proposed by Heller and Yamani in the seventies \cite{heller1974new,heller} and mainly applied to atomic problems. The method exploits the tridiagonal structure of the kinetic energy operator. The $J$-matrix has been under constant development since its inception and a review of the recent developments can be found in \cite{jmatrix}. For problems with Coulomb interactions the Coulomb-Sturmian basis is preferred over the oscillator representation. Recently the Coulomb-Sturmians have been used to describe multiphoton single and double-ionization \cite{foumouo2006theory} and electron impact ionization \cite{PhysRevA.82.022708}. At the same time, and mostly independent, the Algebraic Resonating Group Method was developed for nuclear scattering problems \cite{filippov1980use,filippov1981,Arickx1994,okhrimenko1984allowance, PhysRevA.55.265}. This method exploits the same principles as the $J$-matrix method for description of nuclear cluster systems, where the oscillator representation is efficiently used to describe the internal structure of clusters. If the same representation is used for intercluster degrees of freedom, then the nucleon symmetrization rules become straightforward. One shortcoming of the methods based on an $L^2$ basis is that the asymptotic solutions need to be known explicitly before a system for the wave function in the inner region and the scattering observables can be written down. This is a serious limitation since it is hard to find the asymptotic wave function for breakup reactions with multiple fragments. The Complex Scaling method, which scales the full domain into the complex plane, can be easily implemented in the oscillator representation by taking a complex valued oscillator strength. This has been used to calculate energy spectrum or extract the resonances \cite{csoto}, but the calculation of cross sections and other scattering observables may be quite difficult. In recent decades significant progress has been made in the numerical solution of scattering processes described by the Helmholtz equation. In contrast to many particle systems, where the potential $V$ is often non-local, the wave number $k(x)$ in the Helmholtz equation depends only on the local material parameters such as the speed of sound in acoustics or the electric permittivity and magnetic permeability for electromagnetic scattering. For these problems grid based representations such as finite difference, finite element \cite{NME:NME1620380303,Babuska1995325}, or Discontinuous Galerkin \cite{Farhat20016455,Farhat20031389} are preferred since they lead to sparse matrices that can be easily solved by preconditioned Krylov subspace methods \cite{vandervorst, erlangga2006comparison}. Another technique that has found widespread application is the use of absorbing boundary conditions. These boundary conditions allow a scattering calculation without prior knowledge about the asymptotic wave form. Exterior Complex Scaling (ECS) and Complex Scaling (CS) are widely used in atomic and molecular physics \cite{McCurdy2004,moiseyev1998quantum,rescigno1999collisional,vanroose2005complete}. Perfectly matched layers (PML) are used for electromagnetic and acoustic scattering \cite{berenger1994perfectly}, which can also be interpreted as a complex stretching transformation \cite{chew}. There are many other excellent absorbing boundary conditions \cite{givoli,antoine, nissen2010perfectly}. In \cite{PhysRevC.82.064603} the JM-ECS method was introduced that combines the $J$-Matrix method with a grid based ECS. The method describes the scattering solution in the interior region with an oscillator representation and in the exterior region with finite differences. The two representations are matched through a low order asymptotic formula with an error that scales as $\mathcal{O}(N^{-1/2})$, where $N$ is the size of the oscillator basis describing the inner region. Once the grid and oscillator representation are matched, it is easy to introduce an absorbing boundary layer since the grid representation can be easily extended with an ECS absorbing layer or any other absorbing boundary condition. The resulting method was illustrated for one-- and two--dimensional model problems representative for real scattering problems with local interactions. Furthermore, the representation was used for nuclear $p$-shell scattering. However, the accuracy of the calculations was unsatisfactory due the low--order matching condition. While the grid representation on the exterior has an accuracy of $\mathcal{O}(N^{-1})$, the matching is only accurate to order $\mathcal{O}(N^{-1/2})$. The main contribution of the current paper is to increase the accuracy of the asymptotic formula that allows a better matching of the grid and oscillator representations. The better asymptotic formula takes function values and its third derivative into account. It can bring the matching error down to the level of the accuracy of the grid representation. A higher--order asymptotic approximation of the oscillator representation was already discussed by S. Igashov in \cite{igashov_jmatrix}. However, the formula was not used to increase the accuracy of scattering calculations. This paper is outlined as followed. In Section \ref{sec:reviewscattering} we shortly describe the process of scattering calculations in the oscillator representation. In Section \ref{sec:higherorder} we derive a higher--order asymptotic formula that takes into account the behavior of the function in the classical turning points in coordinate and Fourier space. In Section 4 and 5 we use this asymptotic formula to solve scattering problems. \section{Review of scattering calculations in the oscillator representation} \label{sec:reviewscattering} In this section we discuss the most important properties of the oscillator representation and how they can be used to perform scattering calculations with the $J$-matrix method. We also recall the working of the hybrid $J$-matrix and ECS method proposed in \cite{PhysRevC.82.064603}. \subsection{The radial scattering equation} The aim is to solve the Schr\"odinger equation in atomic units ($m=1$, $\hbar=1$) that describes a scattering process of two particles for an energy $E\in \mathbb{R}$. This equation written in relative coordinates is \begin{equation}\label{eq:scattering3d} \left[-\frac{1}{2}\Delta + V(\mathbf{r}) - E \right] \Psi(\mathbf{r}) = F(\mathbf{r}), \quad \forall \mathbf{r} \in \mathbb{R}^3 \end{equation} where $F(\mathbf{r})$ is the function describing the initial state or the source term and $V(\mathbf{r})$ is the potential. The coordinate $\mathbf{r}$ can be written in spherical coordinates $(\rho, \theta, \phi)$. In case of spherically symmetric potential ($V(\mathbf{r})=V(\rho)$), Eq.~\eqref{eq:scattering3d} can be reduced to a one-dimensional radial equation using partial wave decomposition $$ \Psi(\mathbf{r})=\Psi(\rho, \theta, \phi) = \sum_{l,m} \frac{\psi_l(\rho)}{\rho} Y_{l,m}(\theta, \phi), \quad F(\mathbf{r}) = F(\rho, \theta, \phi) = \sum_{l,m} \frac{f_l(\rho)}{\rho} Y_{l,m}(\theta, \phi), $$ where $Y_{l,m}(\theta, \phi)$ is a spherical harmonic, $l=0,1,2, \ldots$ is the orbital angular momentum of the relative motion, $m$ is the projection of this angular momentum. The resulting reduced radial equation becomes \begin{equation}\label{eq:scattering} \left[-\frac{1}{2} \frac{d^2}{d \rho^2} + \frac{l(l+1)}{2 \rho^2} + V(\rho) - E \right] \psi_l(\rho) = f_l(\rho). \end{equation} For the problem we are interested in there is a range $a > 0 $ such that for all $\rho>a$ both $V(\rho)$ and $\chi(\rho)$ are zero. We solve the Eq.~\eqref{eq:scattering} for $E>0$ and extract from the solution $\psi_l$ scattering observables such as the cross sections or phase shift. In order to solve the Eq.~\eqref{eq:scattering} we represent the solution as \begin{equation} \label{eq:expansion} \psi_l(\rho) = \sum_{n=0}^{\infty} c_{n,l} \varphi_{n,l}(\rho), \end{equation} where $\varphi_{n,l}(\rho)$ are orthogonal $L^2$ functions, in particular we will use reduced oscillator functions, whose properties will be explained in the next section. After projection of~\eqref{eq:expansion} on $\varphi_{n,l}$, Eq.~\eqref{eq:scattering} results in an infinite linear system \begin{equation}\label{eq:discretescattering} \sum_{n=0}^\infty \left(T^{(l)}_{kn} + V^{(l)}_{kn} - E \right)c_{n,l} =b_{k,l}, \end{equation} where $T_{kn}^{(l)}$ denotes the elements of the kinetic energy matrix and $V_{kn}$ the elements of the potential energy matrix. They are the integrals \[T^{(l)}_{kn} = \int_0^\infty \varphi_{k,l}(\rho) \left(-\frac{1}{2} \frac{d^2}{d \rho^2} + \frac{l(l+1)}{2\rho^2} \right) \varphi_{n,l}(\rho) \; d\rho,\] \[V^{(l)}_{kn} = \int_0^\infty \varphi_{k,l}(\rho) V(\rho)\varphi_{n,l}(\rho) \; d\rho \] and $b_{k,l} = \int_0^\infty \varphi_{k,l}(\rho) \chi_l(\rho) \; d\rho$. As the considered problem is effectively one-dimensional we will use $x$ instead of $\rho$ to denote radial relative coordinate in all following sections devoted to two-body problem. \subsection{The oscillator representation} Before we explain the strategy to solve Eq.~\eqref{eq:discretescattering}, we repeat the main properties of the oscillator representation and the function $\varphi_{n,l}$ that will be used in the expansion. The reduced radial equation analogous to~\eqref{eq:scattering} for the quantum harmonic oscillator with an oscillator strength $\omega$ is \begin{equation}\label{eq:oscil_rad} \left[-\frac{1}{2} \frac{d^2}{d x^2} +\frac{1}{2}\frac{l(l+1)}{x^2} + \frac{1}{2}\omega^2 x^2 \right] \varphi_{n,l}(x) = E_{n,l} \varphi_{n,l}(x) \end{equation} with $x \in [0, \infty[$ a radial coordinate. The boundary conditions are $\varphi_{n,l}(0)=0$ and $\lim_{x\rightarrow \infty} \varphi_{n,l}(x) = 0$. The eigenvalues are \begin{equation} \label{eq:osc_spectrum} E_{n,l} = \left(2 n + l + \frac{3}{2}\right)\omega, \end{equation} and eigenstates \begin{equation}\label{eq:oscillatorfunction} \varphi_{n,l}(x) = (-1)^n N_{n,l} b^{-1/2} \left(\frac x b\right)^{l+1} \exp\left(-\frac{x^2}{2b^2}\right) L_n^{l+1/2}\left(\frac{x^2}{b^2}\right), \end{equation} where $L_n^{l+1/2}$ are Laguerre polynomials. The normalization is $N_{n,l} = \sqrt{2n!/\Gamma(n+l+3/2)}$, where $n \in \{0,1,\ldots,\}$ and oscillator length is defined as $b=\sqrt{1/\omega}$. The classical turning point associated with each state is $R_{n,l} = b \sqrt{4n + 2l + 3}$, defined as the point where the potential energy equals the total energy of the system. The functions (\ref{eq:oscillatorfunction}) form a complete orthonormal basis and any wave function $\psi_l(x)$ that behaves as $x^l$ in $x=0$ can be represented using the infinite sum: \begin{equation} \psi_l(x) = \sum_{n=0}^{\infty} c_{n,l} \varphi_{n,l}(x), \quad \text{where} \quad c_{n,l} = \int_0^\infty \varphi_{n,l}(x) \psi_l(x) dx. \label{eq:osc_rep} \end{equation} In the following section we will use that the radial oscillator equation, \eqref{eq:oscil_rad}, can be rewritten in terms of $b$, the oscillator length, and $R_{n,l}$, the classical turning point as \begin{equation}\label{eq:oscillator} \left[-\frac12 \frac{d^2}{dx^2} + \frac{l(l+1)}{x^2} + \frac{x^2}{2 b^4}- \frac{R_{n,l}^2}{2 b^4}\right] \varphi_{n,l}(x)=0. \end{equation} An important property that forms the basis of the results in this paper is that the $n$-th oscillator state has $n$ oscillations between the origin and the classical turning point $R_{n,l}=\sqrt{4n+2l+3}$. Beyond this turning point the function is exponentially decaying without additional oscillations. This means that as $n$ increases, the frequency of the oscillation between the origin and $R_{n,l}$ grows proportional to $\sqrt{n}$. This property will be used to derive the asymptotic formula in Section \ref{sec:higherorder}. \begin{figure} \begin{center} \includegraphics[width=0.5\linewidth]{oscil10}\includegraphics[ width=0.5\linewidth]{oscil30} \end{center} \caption{The reduced oscillator state $\varphi_{n,l}(x)$ with a classical turning point $R_{n,l}$ for $l=0$ and two values of $n$: $n=10$ (left) and $n=30$ (right). The function oscillates $n$ times on the interval $[0,R_{n,l}]$. Since $R_{n,l}$ only grows with $\mathcal{O}(\sqrt{n})$ the state becomes rapidly oscillating as $n$ grows.} \label{fig:oscil} \end{figure} The Bessel transform $\tilde{F_l}(k)$ of a function $F_l(x)$ is defined as \begin{equation} \tilde F_l(k) := \sqrt{\frac2{\pi}} \int_0^\infty F_l(x) \hat j_l(kx) dx. \end{equation} where $\hat j_l(kx)$ is a Riccati--Bessel function, the regular solution of the radial Schr\"odinger equation without potential, \eqref{eq:scattering}. It is connected with the ordinary Bessel function in a following way $\hat j_l(x) = \sqrt{\pi x/2}\,J_{l+1/2}(x)$. The Bessel transform of an oscillator state is again an oscillator state \begin{equation} \tilde \varphi_{n,l}(k) = (-1)^n b\, \varphi_{n,l}(kb^2), \end{equation} with a classical turning point $K_{n,l} = R_{n,}/b^2 = \sqrt{4n+2l+3}/b$. This relation that can easily be derived using {\em e.g.} formula (7.421.4) from \cite{gradshtein}. The oscillator states therefore form an orthonormal basis of $L^2([0,\infty[)$ that diagonalizes the Bessel transform. Similar properties hold for the Cartesian oscillator state based on Hermite polynomials, see, for example, \cite{PhysRevB.80.045321}. An other important property that will be used in the remainder of the text is the following: \begin{proposition}\label{lem:interchange} Let $\psi_l$ a function that behaves as $x^l$ in $x=0$. The projection on the oscillator state $\varphi_{n,l}$ can be calculated with either $\psi_l$ or its Bessel transform $\tilde{\psi}_l$ as: \begin{equation} c_{n,l} = \int_0^\infty \varphi_{n,l}(x) \psi_l(x) dx = (-1)^n b \int_0^\infty \varphi_{n,l}(x) \tilde{\psi}_l(x) dx. \end{equation} \end{proposition} \begin{proof} Since Parseval's theorem holds we can calculate $c_n$ either with $\varphi_{n,l}$ or its Bessel transform \begin{equation} c_{n,l} = \int_0^\infty \varphi_{n,l}(x) \psi_l(x) dx = \int_0^\infty \tilde{\varphi}_{n,l}(k) \tilde{\psi}_l(k) dk. \end{equation} But since $\tilde{\varphi}_{n,l}(k)= (-1)^n b \varphi_{n,l}(k b^2)$, the oscillator state in the latter integral is the same as in the first integral, after substitution of variables and up to a constant.\end{proof} This result will be used in the next section to derive the asymptotic formula for $c_{n,l}$ with large $n$. This symmetry between $\psi_l(x)$ and its Bessel transform $\tilde{\psi_l}(k)$ should also hold in the asymptotic formula. The kinetic energy operator $T_{i,j}^l$ is a tri-diagonal matrix. For the oscillator basis the non-zero elements are \begin{equation} T^l_{i,j} = \int_0^\infty \varphi_{i,l}(x) \left(-\frac{1}{2}\frac{d^2}{dx^2} + \frac{l(l+1)}{2x^2}\right) \varphi_{j,l}(x) dx = \left\{\!\!\begin{array}{c@{\quad}l} \left(2i+l+\frac32 \right)\frac{\omega}{2} & \text{for } j=i,\\ -\sqrt{i\,\left(i+l+\frac12\right)}\frac{\omega}{2} & \text{for } j=i-1,\\ -\sqrt{(i+1)\left(i+l+\frac32 \right)}\frac{\omega}{2}& \text{for } j=i+1. \end{array}\right. \label{eq:kinetic_en} \end{equation} For the remainder of the text we will consider only the case of zero angular momentum and drop $l$ from the notation. All results, however, are valid for arbitrary $l$. \subsection{The asymptotic formula} In \cite{Vanroose2001} an asymptotic formula was derived for the projection of a smooth radial function $\psi$ on the state $\varphi_n$. The derivation uses stationary phase arguments to exploit the increase of oscillations as $n\rightarrow \infty$. We repeat here the derivation of the results. There are two main contributions to the value of the integral \begin{equation} c_n = \int_0^\infty \varphi_n(x)\psi(x) dx \approx I_0 + I_{R_n}. \end{equation} There is a first contribution, denoted $I_0$, of the left integration boundary near zero. There is a second contribution, denoted $I_{R_n}$, from the point of stationary phase, which is the classical turning point where the oscillations stop (see Figure \ref{fig:oscil}). Here we provide only general steps of calculating this integral in the asymptotic region. More details can be found in \cite{Vanroose2001}. The contribution from the classical turning point can be calculated by approximating the oscillator state $\varphi_n$ near the classical turning point by an Airy function as \begin{equation} \varphi_n(x) \approx \frac{2}{b}\left(\frac{b^4}{2R_n}\right)^{1/6} \mathrm{Ai}\left[ \left(\frac{2R_n}{b}\right)^{1/3}(x-R_n)\right] \quad \text{if} \quad |x-R_n| \ll 1. \end{equation} The integral representation of this Airy function and the stationary phase approximation leads to the contribution of the turning point to the integral \begin{equation} I_{R_n} \approx b\sqrt{\frac{2}{R_n}} \psi(R_n). \end{equation} The contribution of the left integration boundary, $I_0$, can be derived by approximating the oscillator state near the origin by a Riccati--Bessel function \begin{equation} \varphi_n(x) \approx (-1)^n \frac{\sqrt{2}}{b}\sqrt{\frac{2}{\pi K_n}} \hat j_0(K_n x), \end{equation} where $K_n$ is the classical turning point of the oscillator state in the momentum space. Then the contribution from the origin becomes \begin{equation} I_0 \approx (-1)^n\frac{1}{b}\sqrt{\frac{2}{K_n}} \tilde{\psi}(K_n). \end{equation} And the resulting asymptotic approximation of the oscillator coefficients becomes \begin{equation}\label{eq:prl_firstorder} c_n \approx (-1)^n \frac{1}{b} \sqrt{\frac{2}{K_n}} \tilde{\psi}(K_n) + b \sqrt{\frac{2}{R_n}} \psi(R_n) \quad \text{if} \quad n \gg 1. \end{equation} This relation has a contribution from the turning points in the coordinate space, $R_n$ and the Fourier space, $K_n$. Note that it satisfies the symmetry observed in \ref{lem:interchange} above. Note that Eq.~\eqref{eq:prl_firstorder} as well as the similar equation in \cite{Vanroose2001} does not provide the order of the approximation. This is one of the shortcomings that will be addressed in the current paper. \subsection{Scattering calculations in the oscillator representation} We now discuss how a finite linear system can be obtained that solves for the wave function in the interaction region and the phase shift, describing the solution in the asymptotic region. The presentation here is based on the asymptotic formula and differs from how the method was derived historically. For more detail about the formulation of the original $J$-matrix method we refer to \cite{heller} and \cite{Arickx1994}. Let $\psi(x)$ be a smooth two-body radial scattering state with a bounded energy, depending on one spatial coordinate (can be always reduced using center-of-mass relative coordinates and spherical coordinates). Since it is a scattering state, the function does not go to zero as $x\rightarrow \infty$. However, its Bessel transform $\tilde{\psi}(k)$ goes to zero as $k \rightarrow \infty$. This means that, as $n$ grows, the only contribution to the expansion coefficient $c_n$ comes from the classical turning point in coordinate space. So, for a scattering state $\psi$ with total energy $E=k^2/2$ the solution is asymptotically $\hat{j}_l(kr) + \tan (\delta_l) \hat{n}_l(kr)$ where $\delta_l$ is the phase shift~\cite{Tay83} holds that \begin{equation} c_n \approx b \sqrt{\frac{2}{R_n}} \psi(R_n) = b\sqrt{\frac{2}{R_n}}\left(\hat{j}_l(kR_n) + \tan (\delta_l) \hat{n}_l(kR_n)\right), \label{eq:as_coeff} \end{equation} where $\hat{j}_l$ and $\hat{n}_l$ are the regular and irregular solutions of the free-particle equation. The $c_n$ becomes the representation of $\psi$ on the grid of classical turning points $R_n$. The aim of a one-dimensional radial scattering calculation is to find a numerical approximation to $\tan (\delta_l)$ for a given potential $V$. This can be achieved by writing the solution in the oscillator representation as \begin{equation}\label{eq:asymptotic} c_{n,l} = \begin{cases} c^0_{n,l} + j_{n,l} + \tan (\delta_l) \, n_{n,l} & n<N \\ j_{n,l} + \tan(\delta_l)\, n_{n,l} & n \geq N, \end{cases} \end{equation} where $j_{n,l}$ is the regular solution of the homogeneous three-term recurrence relation \begin{equation} T_{n,n-1}j_{n-1,l} + (T_{n,n}-E)j_{n,l} + T_{n,n+1}j_{n+1,l} = 0, \end{equation} where as $n \gg 1$ holds that $j_{n,l} =b\sqrt{2/R_n}\,\hat{j}_l(kR_n)$. The $n_{n,l}$ is the irregular solution of the recurrence relation that goes as $n_{n,l} = b\sqrt{2/R_n}\,\hat{n}_l(kR_n)$ when $n$ becomes large. With the form \eqref{eq:asymptotic} we reduce the infinite linear system to a finite dimensional problem with a set of $N+1$ unknowns $\{c^0_{i,l}, \tan \delta_l \}$. Once solved we simultaneously obtain the wave function of the system, $\{ c^0_{i,l}\}$, and the scattering information, $\tan (\delta_l)$. A detailed description of the construction of this linear system, known as the $J$-matrix, can be found in \cite{Arickx1994}. \subsection{The hybrid $J$-matrix and ECS method} In \cite{PhysRevC.82.064603} the asymptotic formula \eqref{eq:prl_firstorder} was used to introduce a hybrid oscillator and grid representation for the Schr\"odinger equation that is useful for scattering calculations where an asymptotic form such as~\eqref{eq:as_coeff} is not explicitly known. Such a situation appears in breakup reactions of three or more particles. Let $\psi(x)$ be again a smooth one-dimensional radial scattering state with a bounded energy such that \begin{equation} c_n \approx b \sqrt{\frac{2}{R_n}} \psi(R_n). \label{eq:matching} \end{equation} holds. The $c_n$ becomes the representation of $\psi$ on the grid of classical turning points $R_n$. The grid distance between these points becomes smaller when $n$ increases since \begin{equation} h_n = R_{n}-R_{n-1} \approx b^2\frac2{R_n}. \label{eq:vgridsize} \end{equation} However, the asymptotic $\psi$ does not necessarily need to be represented on the grid of turning points $R_n$. Another option is, for example, a regular grid of equally spaced points. The hybrid JM-ECS method represents the one-dimensional radial wave function as a vector ${\sf \Psi}$ in $\mathbb{C}^{N+K}$, where \begin{equation} \label{eq:hybridvector} {\sf \Psi} = (c_0,c_1, \ldots, c_{N-1},\psi(R_{N}),\psi(R_{N}\! +\! h),\ldots,\psi(R_{N}\! +\! (K\!-\!1)h) ). \end{equation} The first $N$ elements represent the wave function in the oscillator representation. While the remaining $K$ elements represent $\psi$ on an equidistant grid that starts at $R_N$, the $N$-th classical turning point, and runs up to $R_N + (K-1)h$ with a grid distance $h$ equal to the difference of the last two turning points of the oscillator representation $h = R_{N}-R_{N-1}$. It is assumed that the matching point that connects the oscillator to finite--difference representation corresponds to a large index $N$ such that the asymptotic formula, \eqref{eq:matching}, applies. Again, the kinetic energy operator in this hybrid representation is tridiagonal since in both finite--difference and oscillator representations it is tridiagonal. One should only be careful in matching both representations. To obtain the kinetic energy in the last point of the oscillator representation, the tridiagonal kinetic energy formula~\eqref{eq:kinetic_en} is used. It involves a recurrence relation connecting the three terms $c_{N-2}$, $c_{N-1}$ and $c_{N}$. The latter, the coefficient $c_N$, is unknown. Only $\psi(R_N)$ is available on the grid. Using the asymptotic relation~\eqref{eq:matching}, however, it is possible to calculate the required matrix element as follows: \[\begin{split} (Tc)_{N-1} = T_{N-1,N-2} c_{N-2} + T_{N-1,N-1} c_{N-1} + T_{N-1,N} b \sqrt{2/ R_{N}} \psi_l(R_{N}). \end{split}\] To calculate the kinetic energy in the first point of the finite difference grid, the second derivative of the wave function has to be known. To approximate the latter with a finite difference formula, one needs the wave function in the grid points $R_{N-1}$, $R_{N}$ and $R_N + h$. Again it is possible to apply~\eqref{eq:matching} to obtain $\psi(R_{N-1})$ in terms of $c_{N-1}$: \[ \psi^{\prime\prime}(R_N) \approx \frac{c_{N-1}/( b \sqrt{2/ R_{N-1}}) - 2 \psi(R_{N}) + \psi(R_{N}+h)}{h^2}. \] The coupling between both representations around the matching point is sketched in Figure \ref{fig:coupling}, together with the terms involved to determine the correct matching. \begin{figure}[ht] \begin{center} \hspace{10pt} Oscillator \hspace{30pt} Finite Differences $$ \overbrace{\vphantom{\rule[5pt]{5pt}{5pt}}c_{n-2} \;\; \quad c_{N-1} \qquad \psi(R_N)}^{T c_{N-1}} \quad \psi(R_{N}+h) \makebox[0pt][r]{$\underbrace{ \phantom{c_{N-1} \qquad \psi(R_N) \quad \psi(R_{N+1}) \vphantom{\rule[-20pt]{20pt}{20pt}} }}_{\psi''(R_N)}$} $$ \ifx\JPicScale\undefined\def\JPicScale{0.6}\fi \unitlength \JPicScale mm \begin{picture}(120,30)(-1,-32) \linethickness{0.3mm} \put(0,20){\line(1,0){120}} \linethickness{0.3mm} \put(90,20){\circle*{3}} \linethickness{0.3mm} \put(20,20){\circle{3}} \linethickness{0.3mm} \put(40,20){\circle{3}} \linethickness{0.3mm} \put(67,20){\circle*{3}} \linethickness{0.1mm} \put(53,-6){\line(0,1){50}} \end{picture} \end{center} \vspace{-60pt} \caption{Illustration on the calculation of the kinetic energy matrix elements that are calculated in the last point $R_{N-1}$ of the oscillator representation and in the first point $R_N$ of the finite difference representation. To calculate $T$ applied to a solution vector we need to translate the oscillator representation to the grid. Vice versa for the application of the finite difference stencil approximation to the second derivative. \label{fig:coupling}} \end{figure} However, in the next section we will show that the asymptotic matching condition that is used in \cite{PhysRevC.82.064603} to couple both representations, is only accurate to $\mathcal{O}(R_N^{-1})$, while the outer grid is accurate to order $\mathcal{O}(h^2)=\mathcal{O}(R_N^{-2})$. Therefore the largest error is made in the matching condition that couples both representations. Note that introducing an absorbing layer is easy once the oscillator representation is coupled to a grid representation. For example ECS is implemented by extending the grid with a complex scaled part \cite{McCurdy2004}. \section{Higher--order asymptotic formula} \label{sec:higherorder} In this section we derive a higher--order asymptotic formula. It not only takes into account the function values in the turning point $R_n$ but also the third derivative in this point. A similar asymptotic formula was derived by S. Igashov in \cite{igashov_jmatrix}, however, it did not include the contributions from the Fourier space. \begin{proposition} \label{lemma1} Let $\psi(x)$ be a regular scattering state that behaves as $x^l$ in $x=0$, that is infinitely differentiable and $\varphi_n(x)$ is a reduced oscillator state, solution of~\eqref{eq:oscillator}. Then the projection of the scattering state on the oscillator state can be approximated as \begin{equation}\label{eq:first_corr} c_{n} = \int_0^\infty \, \varphi_n(x)\, \psi(x) dx = b \sqrt{\frac2{R_{n}}} \left( \psi(R_{n}) + \frac{b^4}{6 R_{n}} \psi^{\prime\prime\prime}(R_{n}) \right) + \mathcal{O}(R_{n}^{-5/2}). \end{equation} This expresses the expansion coefficient in terms of the wave function and its derivatives in the classical turning point $R_{n}$. \end{proposition} \begin{proof} The oscillator Eq.~\eqref{eq:oscillator} can be rewritten as \begin{equation} \left[-\frac{d^2}{dx^2} + \frac{2 R_n (x-R_n)}{b^4} + \frac{(x-R_n)^2}{b^4}\right] \varphi_n(x)=0. \label{eq:osc_rewritten} \end{equation} For $R_n \gg |x-R_n|$ (which means either $n\gg1$ or $|x-R_n| \approx 0$), we can neglect the quadratic term in this equation and get the Airy equation with the solution \begin{equation} \varphi_n(x) \approx \varphi_n^{(0)}(x) = \frac 2 b \left( \frac{b^4}{2 R_{n}} \right)^{1/6} \mathrm{Ai} \left[ \left(\frac{2 R_{n}}{b^4}\right)^{1/3}(x-R_{n})\right], \label{eq:airy_1st} \end{equation} where the normalization is chosen such that it coincides with the oscillator state near the classical turning point $R_n$. This can be further improved by writing $\varphi_n(x) = \left(1 + u_n(x)\right)\varphi^{(0)}_n$ and inserting this in the oscillator equation. The equation for $u_n(x)$ becomes \begin{equation} \label{eq:33} -\frac12 u_n^{\prime\prime} \varphi_n^{(0)} - u_n^\prime {\varphi_n^{(0)}}^\prime -\frac12 u_n\,{\varphi_n^{(0)}}^{\prime \prime} + \left(R_n(x-R_n) + \frac{1}{2}(x-R_n)^2\right) u_n \varphi_n^{(0)} = -\frac{1}{2}(x-R_n)^2 \varphi_n^{(0)}. \end{equation} Using that $${\varphi_n^{(0)}}^\prime= \left(2 R_n/b^4\right)^{1/3} \mathrm{Ai}^\prime\left[\left(2 R_{n}/b^4\right)^{1/3}(x-R_{n})\right]$$ and $${\varphi_n^{(0)}}^{\prime\prime}= \left(2 R_n/b^4\right)^{2/3} \mathrm{Ai}^{\prime\prime}\left[\left(2 R_{n}/b^4\right)^{1/3}(x-R_{n})\right]$$ the equation~\eqref{eq:33} becomes \begin{align} &-\frac12 u_n^{\prime\prime} \varphi_n^{(0)} - u^\prime \left(\frac{2R_n}{b^4}\right)^{1/3} \mathrm{Ai}^\prime\left[\left(\frac{2 R_{n}}{b^4}\right)^{1/3}(x-R_{n})\right]\nonumber\\ &-\frac12 u_n \, \left(\frac{2 R_n}{b^4}\right)^{2/3} \mathrm{Ai}^{\prime\prime}\left[\left(\frac{2 R_{n}}{b^4}\right)^{1/3}(x-R_{n})\right] + \left(R_n(x-R_n) + \frac{1}{2}(x-R_n)^2\right) u_n \, \varphi_n^{(0)} = -\frac{1}{2}(x-R_n)^2 \varphi_n^{(0)}. \end{align} In the limit $n \rightarrow \infty$, the term with $R_n(x-R_n)$ dominates compared to $(1/2)(x-R_n)^2$ and the other terms since both the first and second derivative of $\mathrm{Ai}(x)$ around zero are bounded. We find that $u_n(x) \approx -(x-R_n)/(2 R_n)$ is a solution of the remaining equation. At this time we will not consider the contributions to the integral for the boundary at zero. This contribution will be equal to the contribution from the Fourier space as in Eq.~\eqref{eq:prl_firstorder} and for scattering states and large $n$ this contribution is negligible. So we can lower the integration boundary and write \begin{equation} c_{n} \approx \frac{2\sqrt{a_n}}{b}\int_{-\infty}^{\infty} \psi(x+R_{n}) \mathrm{Ai} (x/a_n) \left(1-\frac{x}{2R_n}\right) dx, \label{eq:airy_approx} \end{equation} where $a_n:=(b^4/2R_{n})^{1/3}$ and we have substituted the integration variable $x\rightarrow x+R_n$. The value of this integral will be determined by the behavior around $R_n$, the point of stationary phase. Since the function $\psi$ is infinitely differentiable we can Taylor expand \begin{equation} c_{n} \approx \frac{2\sqrt{a_n}}{b}\int_{-\infty}^{\infty} \sum_{m=0}^{\infty} \frac{x^m}{m!}\psi^{(m)}(R_{n}) \left(1-\frac{x}{2R_n}\right) \mathrm{Ai} (x/a_n) dx \end{equation} All integrals in these series can be calculated explicitly using specific properties of Airy function (see page 52 of \cite{airy}). \begin{equation} \int_{-\infty}^{\infty} \mathrm{Ai}(x) x^{3k} dx = \frac{(3k)!}{3^k k!} \quad \text{and} \quad \int_{-\infty}^{\infty} \mathrm{Ai}(x) x^{3k+1} dx = \int_{-\infty}^{\infty} \mathrm{Ai}(x) x^{3k+2} dx = 0. \end{equation} We finally get \begin{equation}\label{eq:ho_matching} \begin{split} c_{n} &\approx \frac{2 a_n^{3/2}}{b} \sum_{k=0}^\infty \frac{a_n^{3k}}{3^k k!}\left( \psi^{(3k)}(R_{n}) - \frac{a_n^3}{2R_n}\psi^{(3k+2)}(R_{n})\right) \\ &= b \sqrt{\frac2{R_{n}}} \sum_{k=0}^{\infty} \frac{1}{k!}\left(\frac{b^4}{6 R_{n}}\right)^k \left(\psi^{(3k)}(R_{n}) - \frac{b^4}{4R_n^2}\psi^{(3k+2)}(R_{n})\right). \end{split} \end{equation} In lowest order (in terms of $1/R_n$) this gives us exactly the initial relation (\ref{eq:matching}). Including the next order correction we get a relation (\ref{eq:first_corr}). For testing purposes we also include terms of the next order $1/R_n^2$ \begin{equation} \begin{split} c_{n} &= b \sqrt{\frac2{R_{n}}} \left(\psi(R_{n}) + \frac{b^4}{6 R_{n}} \psi^{\prime\prime\prime}(R_{n})\right. \left. +\frac1{R_{n}^2}\left(\frac{b^8}{72}\psi^{(6)}(R_{n}) - \frac{b^4}{4}\psi^{\prime\prime}(R_{n})\right) + \mathcal{O}(\frac1{R_n^3})\right) \end{split} \label{eq:second_corr} \end{equation} \end{proof} Note that a similar result is readily obtained for Cartesian harmonic oscillator states that are based on the Hermite polynomials. There, however, there is a contribution from both turning points, one from $R_n=\sqrt{2n + 1}$ and $R_n = -\sqrt{2n+1}$. \begin{corollary} Let $\psi(x)$ be a function that behaves as $x^l$ in $x=0$ that is infinitely differentiable, then the asymptotic expansion coefficient is \begin{align} c_{n} = &\int_0^\infty \,\varphi_n(x)\, \psi(x) dx = b \sqrt{\frac2{R_{n}}} \left( \psi(R_{n}) + \frac{b^4}{6 R_{n}} \psi^{\prime\prime\prime}(R_{n}) \right)\nonumber \\ & + \frac{(-1)^n}b \sqrt{\frac2{K_{n}}} \left( \tilde{\psi}(K_{n}) + \frac{1}{6 b^4 K_{n}} {\tilde{\psi}}^{\prime\prime\prime}(K_{n}) \right) + \mathcal{O}(n^{-5/4}). \label{eq:withfourier} \end{align} \end{corollary} \begin{proof} The asymptotic formula should have the same result when we interchange $\psi(x)$ with its Bessel transform $\tilde{\psi}(k)$ as a result of Proposition \ref{lem:interchange}. Indeed, the integral \[ c_n = (-1)^n b \int_0^\infty \varphi_n(k)\tilde{\psi}(k)dk \] will have a contribution from integral boundary near 0 and a contribution from the turning point $K_n$ in Fourier space. The latter is, with the help of the previous results, $(-1)^n b\sqrt{2/K_n}(\tilde{\psi}(K_n) + \tilde{\psi}^{\prime\prime\prime}(K_n)/6b^4K_n)$. While the contribution from boundary become the contribution from the turning point $R_n$ in coordinate space. \end{proof} \begin{example} We illustrate the convergence of the asymptotic formula \eqref{eq:withfourier} with an application to the function $\psi(x) = \exp(-a x)\sin(x)$. We give results for two choices of $a$. As the scale of the representation is defined by the oscillator length $b$, the result will change with its choice. If the product $a b$ is large, the Fourier components will dominate in the expansion coefficient. When $a b$ is small the function $\psi$ resembles a scattering state, on a scale defined by $b$, and the coordinate approximation will dominate. However, as Figure~\ref{fig:fourier} illustrates the combined formula always gives the correct result. We see an overall convergence with $n^{-5/4}$. This is smaller than $\mathcal{O}(h_n^2)$ which goes as $\mathcal{O}(1/n)$. In general, it is possible to construct a function, for which both coordinate and Fourier terms fail to represent the exact oscillator coefficient, but the combined formula remains valid even in this case (see Figure~\ref{fig:fourier2}). \begin{figure} \begin{center} \begin{tabular}{cc} \includegraphics[width=0.5\linewidth]{general_error_loa}&\includegraphics[width=0.5\linewidth]{general_error_hia} \end{tabular} \end{center} \caption{ Convergence of the general asymptotic formula (\ref{eq:withfourier}) with only coordinate terms (crosses), only Fourier terms (circles), both coordinate and Fourier terms (solid line) for test function $\psi(x) = \exp(-a x)\sin(x)$ with different values of $a$: left figure --- $a=0.2$, right figure --- $a=1.5$ ($b=1$ in both cases)} \label{fig:fourier} \end{figure} \begin{figure} \begin{center} \includegraphics[width=0.5\linewidth]{general_error_cos} \end{center} \caption{ Convergence of the general asymptotic formula (\ref{eq:withfourier}) for the test function $\psi(x) = \exp(-a x)\cos(x)$ with $a=0.2$, for which neither coordinate (crosses) nor Fourier terms (circles) give accurate result, while the combined formula (solid line) does.} \label{fig:fourier2} \end{figure} \end{example} \begin{example} We have verified the asymptotic relations derived in the previous section with a numerical experiment using the wave function $\psi(x)=\sin(kx)$. We calculate oscillator coefficients directly and compare them against the values obtained with asymptotic formulas of different order derived previously (see Fig.~\ref{fig:coeff_as}). \begin{figure} \begin{center} \includegraphics[width=0.5\linewidth]{ho_coef_conv}\includegraphics[ width=0.5\linewidth]{inverse} \end{center} \caption{Convergence of the asymptotic formula (\ref{eq:second_corr}) (left) and the inverse asymptotic formula (\ref{eq:first_corr_inv}) (right) with different number of terms included (crosses --- one term, solid line --- two terms, circles --- three terms). The test function has the form $\psi(x)=\sin(kx)$ with $k=1.5$ and the oscillator length $b=0.8$} \label{fig:coeff_as} \end{figure} \end{example} \subsection{Inverse relation} The asymptotic expansion coefficient is an approximation with the help of the functions values of $\psi$ evaluated at certain grid points. It is also useful to derive in inverse relation that allows us to construct the function value in certain grid points given the expansion coefficients. To approximate $c_n$ with Eq.~\eqref{eq:first_corr} the function value and its third derivative need to be calculated at the turning point $R_n$. When only the function values of $\psi$ are available at the turning points $R_n$, the third derivative can be approximated by finite differences. The coefficient is then calculated as \begin{equation} c_{n} = b \sqrt{\frac2{R_{n}}} \left( \psi(R_{n}) + \frac{b^4}{6 R_{n}} \sum_{k} D^{(3)}_{k} \psi(R_{n+k}) \right) + \mathcal{O}(R_n^{-5/2}) + R_n^{-3/2} \epsilon_n, \label{eq:first_corr_fd} \end{equation} where $D^{(3)}$ is the matrix with coefficients that approximates the third derivative and $k$ indicates the stencil points and $\epsilon_n$ is the error term of this approximation. In case of 5-point finite difference approximation $k\in\{-2,-1,0,1,2\}$. Note that the grid of turning points $R_n$ is an irregular grid and the coefficients will depend on the local distances between the neighboring grid points. The error of the approximation $\epsilon_n$ will also depend on the local grid distances as $\mathcal{O}(h^{p}) \sim \mathcal{O}(R_n^{-p})$, where $p$ is the order of the approximation. The resulting transformation can be represented by a banded sparse matrix $U$ \begin{equation} c_{n} = \sum_{k} U_{nk} \psi(R_k) + \mathcal{O}(R_n^{-5/2}). \end{equation} Note that the matrix elements of $U$ will only depend on the values of $R_n$. Indeed, The differential operator $D^{(3)}$ only depends on the grid distances and so do the coefficients of the asymptotic expression. The relation \eqref{eq:first_corr_fd} translates $\psi$ on the grid of turning points to corresponding $c_n$. It is also useful to derive the inverse relation that gets $\psi(R_n)$ from known values of $c_n$. We can not use a direct inversion of Eq.~\eqref{eq:first_corr} since it involves the third derivative, a dense operator in the oscillator representation. However, an approximate inverse relation can be obtained by rearranging terms in (\ref{eq:first_corr_fd}) as \begin{equation} \psi(R_{n}) = \frac1b \sqrt{\frac{R_{n}}2} c_{n} - \frac{b^4}{6 R_{n}} \sum_{k} D^{(3)}_{k} \psi(R_{n+k}) + \mathcal{O}(R_n^{-2}) + R_n^{-1} \epsilon_n \end{equation} and replacing values of the wave function in the right--hand side by oscillator coefficients using only the first term of (\ref{eq:matching}), i.e. $\psi(R_n) = (1/b)\sqrt{R_n/2}\,c_n + \mathcal{O}(R_n^{-1})$. This only introduces an error of the order of $\mathcal{O}(R_n^{-1})$ but combined with the $1/R_n$ this gives the approximate inverse relation \begin{equation} \psi(R_{n}) = \frac1b \sqrt{\frac{R_{n}}2} c_{n} - \frac{b^3}{6 R_{n}} \sum_{k} D^{(3)}_{k} \sqrt{\frac{R_{n+k}}2} c_{n+k} + \mathcal{O}(R_n^{-2}) \label{eq:first_corr_inv} \end{equation} that is accurate to $\mathcal{O}(R_n^{-2})$. An example of the resulting numerical accuracy of this inverse asymptotic relation is shown on the right panel of figure~\ref{fig:coeff_as}. Also this transformation can be presented as a sparse banded matrix multiplication \begin{equation}\label{eq:w_matrix} \psi(R_n) = \sum_{n} W_{nk} c_{k} + \mathcal{O}(R_n^{-2}). \end{equation} Again the matrix elements of $W$ only depend on the values of the turning points $R_n$. Combining the two relations leads to an approximate partition of unity: \begin{equation} U W = I + \mathcal{O}(R_n^{-2}). \end{equation} It is important to note that since these transformation matrices only depend on values of $R_n$ they can, in principle, both be defined for an arbitrary grid. \subsection{Approximate discretization of Operators} With the help of these two transformation matrices $U$ and $W$ we can now build an approximate oscillator representation of an operator $Q$ using its finite difference representation. Let $Q^{(fd)}$ be the finite difference representation of $Q$ on the grid $R_n$. Then the application of the $Q$ on $\psi$ can be written as \begin{equation*} (Q \psi)^{(osc)}_n = \sum_{m}Q^{(osc)}_{nm} c_m = \sum_{m}\check Q^{(osc)}_{nm} c_m + \mathcal{O}(n^{-1}), \qquad \text{where} \quad \check Q^{(osc)}_{nm} := [U Q^{(fd)} W]_{nm}. \end{equation*} We illustrate the accuracy of this relation on a second derivative operator $D^(2)$, as we intend to apply the constructed representation to Helmholtz-type equations. \begin{equation} \check D^{(2)(osc)}=U D^{(2)(fd)} W, \label{eq:d2trans} \end{equation} where $D^{(2)(fd)}$ is a finite difference matrix of the second derivative. To analyze the accuracy of this approximation Figure~\ref{fig:app_der} shows the result of the error operator defined as $(D^{(osc)}-\check D^{(osc)})$ acting on the vector of oscillator coefficients of the test function $\psi(x)=\sin(kx)$. For comparison we show the error of the approximate identity operator, that can be defined as $(I - \check I) = (I - U W)$. We see that both considered error operators decay as $\mathcal{O}(n^{-1})$ in high-$n$ region. This means that our approximate operators are asymptotically equal to the exact ones and we can expect the same order of convergence in a solution of the scattering problem. \begin{figure} \begin{center} \includegraphics[width=0.5\linewidth]{d2_conv} \end{center} \caption{Convergence of the approximate second derivative operator $\check D = U D^{(fd)} W$ (crosses) and the approximate identity operator $\check I = U W$ (circles) with a test function of the form $\psi(x)=\sin(kx)$ ($k$ and $b$ are the same as in Figure~\ref{fig:coeff_as})} \label{fig:app_der} \end{figure} It is important to note that all matrices in (\ref{eq:d2trans}) can be built for any arbitrary spatial grid, though it will not be an approximate oscillator representation. But we can modify this grid only in the asymptotic region ({\em e.g.} to implement the absorbing boundary with ECS). Then the coefficients $c_n$ have no physical meaning as oscillator coefficients but the coordinate wave function can sill be reconstructed with (\ref{eq:w_matrix}). \section{High-order hybrid representation for scattering problems} The aim is now to build a hybrid representation for scattering calculations where the oscillator basis is used in the internal region and finite differences in the outer region. The matching of the two should use the higher--order asymptotic formula \eqref{eq:first_corr}. However, the strategy displayed on Fig.~\ref{fig:coupling} cannot be easily applied since the third derivative of the wave function needs to be estimated from several neighboring points symmetrically distributed on both sides of the matching point. But in the matching point we only have the required function values on one side of the matching point. This arguments holds both for the oscillator and for the finite difference representation. We therefore present a matching strategy that differs from the proposal of \cite{PhysRevC.82.064603}. Consider a arbitrary grid $\bar{R}_n$ in $[0,\infty]$ with $n=1,2,\ldots$ that differs from the grid of turning points $R_n$. On this grid we can construct the differential operator $D^{(3)}$ and construct the operators $\bar{U}$ and $\bar{W}$ by replacing every appearance of $R_n$ in equations \eqref{eq:first_corr_fd} and \eqref{eq:first_corr_inv} by $\bar{R}_n$. Similarly, for a function $\psi$ sampled in the grid points $\bar{R}_n$ we can calculate \begin{equation} \bar{c}_n = \sum_{k} \bar{U}_{nk} \psi(\bar{R}_k). \end{equation} These coefficients $\bar{c}_n$ are not the expansion coefficients of the function $\psi$ in the oscillator function basis. Only when $\bar{R}_n$ equals $R_n$, the oscillator turning points, then the $\bar{c}_n$ are approximations to the oscillator expansion coefficients $c_n$. To build the hybrid representation, we choose the grid $\bar{R}_n$ such that first $N+1$ points correspond to the turning points $R_n$. The remaining points of $\bar{R}_n$, for $n > N+1$, are chosen to correspond to points on a equidistant finite difference grid. To solve a scattering problem the equation \eqref{eq:scattering} needs to be discretized with finite differences on the grid $\bar{R}_n$ and then transformed with the help of $\bar{U}$ and $\bar{W}$ into \begin{equation} \sum_j \left( \bar{U}H^{fd}\bar{W}-E \bar{U}\bar{W}\right)_{ij} \bar{c}_j = (\bar{U}f)_i \end{equation} to arrive at an equation for $\bar{c}_i$. The first $N$ coefficients of $\bar{c}_n$ correspond now to approximations to oscillator expansion coefficients $c_n$. However, because $n$ is low, they are only a poor approximation. The idea is now to replace the first $N$ coefficients with the exact coefficients. At the same time we replace the first $N$ rows of the matrix with the exact operator in the oscillator representation. The linear system then becomes {\footnotesize \begin{equation} \left(\begin{array}{ccc|cc} H^{(osc)}_{00} - E & \hdots& H^{(osc)}_{0N} &H^{(osc)}_{0N+1} &\hdots\\ H^{(osc)}_{10} & \hdots & H^{(osc)}_{1N} & H^{(osc)}_{1N+1} & \hdots \\ \vdots & & \vdots & \vdots &\\ H^{(osc)}_{N0} & \hdots & H^{(osc)}_{NN}-E & H^{(osc)}_{NN+1} & \hdots \\ \hline [\bar{U}H^{(fd)}\bar{W}]_{N+1,0} & \hdots & [\bar{U}H^{(fd)}\bar{W}]_{N+1,N} & [\bar{U}(H^{(fd)}-E)\bar{W}]_{N+1,N+1} & \hdots\\ \vdots &&\vdots&\vdots& \end{array} \right) \left(\begin{array}{c} \vphantom{H^{(N)}_N} c_{0}\\ \vphantom{H^{(N)}_N} c_{1}\\ \vdots\\ \vphantom{H^{(N)}_N} c_{N}\\ \hline \vphantom{H^{(N)}_N} \bar{c}_{N+1}\\ \vdots \end{array} \right) = \left(\begin{array}{c} f^{(osc)}_{0}\\ f^{(osc)}_{1}\\ \vdots\\ f^{(osc)}_{N}\\ \hline \sum_j\bar{U}_{N+1j} f(x_j)\\ \vdots \end{array} \right), \label{eq:combined} \end{equation} } where $H^{(osc)}$ is the representation of the Hamiltonian in the oscillator representation and $H^{(fd)}$ in the finite difference representation. We emphasize the difference with \cite{PhysRevC.82.064603}. Here we do not match two regions by using the asymptotic formula in one point. Now the representation in the asymptotic region is a fairly good approximation of the oscillator representation. Therefore the structure of the discretized wave function is simpler than \eqref{eq:hybridvector}. Now we have $$ \bar{\sf \Psi} = \left(c_0, \ldots c_N, \bar{c}_{N+1}, \ldots \bar{c}_{K}\right), $$ where in the initial version of hybrid method, the latter elements of the vector are the function values of $\psi$ in the grid points. Now the internal region is covered by an exact oscillator representation, and the asymptotic region is covered by approximate representation which is based on finite differences and includes the ECS transformation. \subsection{Numerical Illustration} We illustrate the method for a one-dimensional radial example. After the solution of (\ref{eq:combined}), we first reconstruct the coordinate wave function using (\ref{eq:w_matrix}). Outside the range of the potential $V$ the solution can be written as a linear combination $\psi_{sc} = A \hat{h}_l^+(kx) + B \hat{h}_l^-(kx)$, where $\hat{h}_l^\pm$ are the in- and outgoing Riccati-Hankel functions. The coefficient $A$ is then extracted with \begin{equation} A = W\left(\psi_\text{sc}(x),\hat{h}_l^-(kx)\right)/W\left(\hat{h}_l^+(kx),\hat{h}_l^-(kx)\right), \end{equation} where $x$ is outside the range of the potential but still on the real part of the ECS domain. The Wronskian is calculated as $W(u,v) = u^\prime v-v^\prime u$, where the derivatives can be implemented with finite differences. From the Wronskians for $A$ and $B$ we can extract the phase shift of the solution. It is important to note that the accuracy of the method does not directly depend on the size of the asymptotic region as long as this region is located outside the range of the potential ($V=0$ in the considered part of $\check H^{(osc)}$). In this region we use a grid of 150 equidistant points and then the ECS layer that spans 10 dimensionless length units and applies a complex rotation of $45^{\circ}$ to the coordinate axis. We first consider a model problem with a attractive potential in Gaussian form, $V(x) = -\exp(-x^2)$, and we limit our consideration to the case of zero total angular momentum $l=0$. Nevertheless, all the conclusions are applicable to higher angular momenta as well. The right panel of Figure \ref{fig:pshift} shows the error in the scattering phase shift of the considered problem calculated with the original JM-ECS and the new approach. To obtain the reference phase shift we use the highly accurate variable phase approach (VPA). We see that for all energies the new method gives much more accurate and less oscillatory results. The convergence as a function of the number of oscillator states in the inner region of both methods is shown on Figure \ref{fig:pshift_conv}. We see that we arrived at the desired the convergence rate to $N^{-1}$ in both low-energy and high-energy regions. \begin{figure} \includegraphics[width=0.41\linewidth]{pshift} \includegraphics[width=0.59\linewidth]{pserror} \caption{ Scattering phase shift of the model problem of Section 4 with a Gaussian potential (left) and the absolute error of the scattering phase shift calculated with JM-ECS (dashed line) and our new approach (solid line) depending on the energy of the system (right). Both calculations were made with 100 functions in the oscillator basis with the oscillator length $b=0.7$.} \label{fig:pshift} \end{figure} \begin{figure} \includegraphics[width=0.5\linewidth]{conv_newho_e02} \includegraphics[width=0.5\linewidth]{conv_newho} \caption{Accuracy of the scattering phase shift calculated with JM-ECS (crosses) and the new method (circles) depending on the size of the oscillator basis. Calculations were made for two values of energy: $E=0.2$, where $\delta=43.73 ^{\circ}$ (left) and $E=3$, where $\delta=19.67 ^{\circ}$ (right)} \label{fig:pshift_conv} \end{figure} The Gaussian shape of the interaction potential was chosen as the main model problem as it is particularly well adapted to the harmonic oscillator basis and $J$-matrix method for this potential converges for any oscillator length $b$. However we can also test our approach against other short-range potentials. But if the asymptotic behavior of the potential is not Gaussian then the results will converge only for specific values of the oscillator length, that match the range of the potential. Also, if the potential has a special point in the origin (like Yukawa potential) then the convergence will be much slower due to the importance of the Fourier contribution in the potential matrix elements. For additional convergence tests we have chosen two potentials with non Gaussian asymptotics and different behavior in the origin: Morse potential $V(x) = \exp(-2 |x/b|)-\exp(- |x/b|)$ and Yukawa potential $V(x) = -\exp(-|x/b|)/x$. In both potentials the potential range is chosen to match the oscillator length of the basis. Figure \ref{fig:pshift_conv_yukava} shows the convergence of the phase shift for these two potentials. We see that in high-energy region the convergence pattern is mostly similar for all potentials as the kinetic energy is much higher than the potential energy in this case. The behavior of the error is much less clear for low energies. The convergence rate appears to be the same here for both methods with Morse potential, for Yukawa potential the convergence rate is blurred due to high oscillations which clearly indicates the importance of Fourier terms. Though we can generally conclude that the new approach always gives better and generally less oscillatory result. \begin{figure} \includegraphics[width=0.5\linewidth]{morse_02} \includegraphics[width=0.5\linewidth]{morse_3} \includegraphics[width=0.5\linewidth]{yukava_02} \includegraphics[width=0.5\linewidth]{yukava_3} \caption{Same as Figure \ref{fig:pshift_conv} but for the Morse potential (two upper panels) and Yukawa potential (two lower panels)} \label{fig:pshift_conv_yukava} \end{figure} \section{Extension to systems with more particles} The results of the previous sections can be generalized to systems of three and more particles. Let $\mathbf{r}_1$ and $\mathbf{r}_2$ be two relative coordinates describing a three--body problem. The 6D wave function can be expanded in partial waves \begin{equation} \label{eq:2dexpansion} \Psi(\mathbf{r}_1,\mathbf{r}_2)= \Psi(\rho_1,\rho_2,\theta_1,\theta_2,\phi_1,\phi_2) = \sum_{l_1,m_1,l_2,m_2} \frac{\psi_{l_1,l_2,m_1,m_2}(\rho_1,\rho_2)}{\rho_1 \rho_2}Y_{l_o1,m_1}(\theta_1, \phi_1)Y_{l_2,m_2}(\theta_2, \phi_2). \end{equation} Such an expansion is used for example to describe double ionization processes in atomic physics \cite{vanroose2006double} Let $\psi(x,y)$ now be an infinitely differentiable two--dimensional radial scattering wave function that is expanded in the bi-oscillator basis as \begin{equation} \psi(x,y) = \sum_{n=0}^{\infty} \sum_{m=0}^{\infty} c_{nm} \varphi_n(x) \varphi_m(y), \label{eq:bioscillator} \end{equation} where $x$ and $y$ should be interpreted as two radial coordinates. The expansion coefficient is then calculated as a double integral that can be approximated by applying the asymptotic relation twice. First in the $x$-direction and then in the $y$-direction \begin{equation} \begin{aligned} \label{eq:2derrors} c_{nm}=& \int_0^\infty \int_0^\infty \varphi_n(x)\varphi_m(y) \psi(x,y) \,dx dy \\ =&\int_0^\infty \varphi_n(x)\left[ \sqrt{2} \, R_m^{-1/2}\psi(x,R_m) + (b^4/6) R_m^{-3/2} \partial_{yyy} \psi(x,R_m) + \mathcal{O}(R_m^{-5/2})\right]dx \\ =& 2\, (R_m R_n)^{-1/2}\psi(R_n,R_m) + \frac{b^4 \sqrt{2}}{6} R_n^{-3/2} R_m^{-1/2} \partial_{xxx} \psi(R_n,R_m) + \frac{b^4 \sqrt{2}}{6} R_n^{-1/2}R_m^{-3/2}\psi_{yyy}(R_n,R_m) \\ & + \frac{b^8}{36}R_n^{-3/2}R_m^{-3/2}\partial_{xxx,yyy} \psi(R_n,R_m) + \mathcal{O}(R_n^{-1/2}R_m^{-5/2}) + \mathcal{O}(R_m^{-1/2}R_n^{-5/2}). \end{aligned} \end{equation} It is important to note that the accuracy depends on both indices $n$ and $m$. For example, when the low order approximation is used for both integrals, there will be error terms of the order $R_m^{-1/2}R_n^{3/2}$ and $R_n^{-1/2}R_m^{3/2}$. Along the diagonal, i.e. when $R_m = R_n$, it is accurate up to order $R_n^{-1/2}R_m^{-3/2} = R_n^{-2} = \mathcal{O}(n^{-1})$, which is of the order $h_n^2$, with $h_n$ the distance between the classical turning points. For the higher--order approximation the error term along the diagonal, where $R_n=R_m$, will be $R_n^{-3} = \mathcal{O}(n^{-3/2})$. In the left panel of Figure \ref{fig:coeff_2d} we compare the exact expansion coefficients with the first order and second order approximation along the diagonal for $n=m$. The results show an improvement with the higher--order expression over the first order approximation. However, when one of the indices remains small, for example $m$, then the error becomes asymptotically $\mathcal{O}(R_n^{-1/2})$, as expected from \eqref{eq:2derrors}. As $n$ becomes large, all error terms with a higher--order accuracy decay until the term $O(R_n^{-1/2} R_m^{-5/2})$ dominates for the higher--order formula, or $O(R_n^{-1/2} R_m^{-3/2})$ for the low order accuracy. This is shown in the right panel of Figure \ref{fig:coeff_2d}. \begin{figure} \begin{center} \includegraphics[width=0.45\linewidth]{2d_coeff_conv_diag} \includegraphics[width=0.45\linewidth]{2d_coeff_conv_axis} \end{center} \caption{Convergence of the asymptotic formula for two-dimensional radial functions. The error is calculated by comparing the exact expansion coefficient with the asymptotic approximations. The function $f(x,y) = \sin(kx) \sin(ky)$ along the diagonal (n=m, left) and along the line with fixed m=20 (right). Crosses show the first order and solid lines correspond to the second order asymptotic expansion coefficient ($k=1$, $b=1$). } \label{fig:coeff_2d} \end{figure} \subsection{Three--body scattering problems} In a similar way as in the previous sections we can use the asymptotic formulate to build a hybrid representation that can solve the radial equation of two radial variables that arises after an expansion of a full three--body scattering problem equation in spherical harmonics \eqref{eq:2dexpansion}. This typically leads to a set of coupled two--dimensional radial partial waves. A diagonal block of this coupled set reads \begin{equation} \left(-\frac{1}{2}\frac{d^2}{dx^2} - \frac{1}{2}\frac{d^2}{dy^2} + \frac{l_1(l_1+1)}{2x^2} + \frac{l_2(l_2+1)}{2y^2} + V(x,y) - E\right) \psi(x,y) = \chi(x,y) \label{eq:2d_scattering} \end{equation} with boundary conditions $\psi(x,0)=\psi(0,y)=0$. The solution of this equation can be represented in a bi-oscillator basis, (\ref{eq:bioscillator}), using two indices $n$ and $m$. Then similar to the two--body case we use approximate oscillator coefficients in the asymptotic region. But there are now three asymptotic regions: first, the region where $n$ is large, second, the region where $m$ is large and finally, the region where both $n$ and $m$ are large. For each of these regions we define approximate oscillator coefficients that take the form: \begin{align} c_{nm} \approx \grave c_{nm} := & \sum_{i} U_{ni} \int_0^{\infty} \varphi_m(y) \psi(R_i,y) \,dy + \mathcal{O}(R_n^{-5/2}) \quad \text{where} \quad n \gg 1\\ c_{nm} \approx \acute c_{nm} := & \sum_{i} U_{mi} \int_0^{\infty} \varphi_n(x) \psi(x,R_i) \,dx + \mathcal{O}(R_m^{-5/2}) \quad \text{where} \quad m \gg 1\\ c_{nm} \approx \check c_{nm} := & \sum_{i,j} U_{ni} U_{mj} \psi(R_i,R_j) + \mathcal{O}(R_n^{-1/2}R_m^{-5/2}) + \mathcal{O}(R_m^{-1/2}R_n^{-5/2}) \quad \text{where} \quad n,m \gg 1, \end{align} where the error terms in the last expression are explained in (\ref{eq:2derrors}). Then the coefficient matrix $c_{nm}$ of hybrid representation is divided into four blocks corresponding to different regions \begin{equation} c_{nm} = \left(\begin{array}{lll|lll} c_{0\,0} & \ldots &c_{0\,M-1} & \acute c_{0\,M} & \ldots &\acute c_{0\,{K'}}\\ c_{1\,0} & \ldots &c_{1\,M-1} & \acute c_{1\,M} & \ldots &\acute c_{1\,{K'}}\\ \vdots & & \vdots& \vdots & &\vdots\\ c_{N-1\,0} & \ldots &c_{N-1\,M-1} & \acute c_{N-1\,M} & \ldots & \acute c_{N-1\,K'} \\ \hline \grave c_{N\,0} & \ldots &\grave c_{N\,M-1} & \check c_{N\,M} & \ldots &\check c_{N\,K'}\\ \grave c_{N+1\,0} & \ldots &\grave c_{N+1\,M-1} & \check c_{N+1\,M} & \ldots &\check c_{N+1\,K'}\\ \vdots & &\vdots & \vdots & &\vdots\\ \grave c_{K\,0} & \ldots &\grave c_{K\,M-1} & \check c_{K\,M} & \ldots &\check c_{K\,K'}\\ \end{array}\right), \end{equation} where we define $N$ and $M$ as sizes of the oscillator bases in each direction and $K$ and $K'$ are the total number of variables in each direction. To build a linear system corresponding to (\ref{eq:2d_scattering}), we reshape the matrix $c_{nm}$ to a vector. Then the Hamiltonian matrix of a 2D problem is constructed as a Kronecker sum of the two-body Hamiltonian's constructed as in (\ref{eq:combined}) and a two-body potential matrix. In the case of our approximate oscillator representation this takes the form \begin{equation} \check H^{(osc)}_{2D} = \check H^{(osc)}_{1} \otimes (U^y W^y) + (U^x W^x) \otimes \check H^{(osc)}_{2} + (U^x \otimes U^y) V_{12}^{(fd)} (W^x \otimes W^y) \end{equation} so the Kronecker sum contains the approximated unity operator instead of the real one. Where $U^x$ and $W^x$ are the transformation matrices for the $x$ coordinate and similarly for the $y$ coordinate. The total size of the Hamiltonian matrix is $K \times K^\prime$. \subsection{Numerical results} We first evaluate the method with the help of a model Helmholtz problem with a constant wave number or energy $E=k^2/2$ and for $l_1=0$ and $l_2=0$. The equation is then \begin{equation} \left(-\frac{1}{2} \Delta - E\right) \psi(x,y) = \varphi_0(x)\varphi_0(y), \label{eq:simple_2d_problem} \end{equation} with boundary conditions $\psi(x,0)=0$ and $\psi(0,y)=0$. The form of the right--hand side here was chosen for simplicity as it results in a vector $(1,0,0,\dots 0)$ in bi-oscillator representation. This problem is exactly solvable with the help of the Greens function \begin{align} G(x,y;x',y') = & \frac{i}{2} \left( H_{0}^{(1)}(k\sqrt{(x-x')^2+(y-y')^2}) - H_{0}^{(1)}(k\sqrt{(x+x')^2+(y-y')^2}) \right. \\ -&\left. H_{0}^{(1)}(k\sqrt{(x-x')^2+(y+y')^2}) + H_{0}^{(1)}(k\sqrt{(x+x')^2+(y+y')^2}) \right), \end{align} where $H_0^{(1)}$ is a 0-th order Hankel function of the first kind. The scattering solution is then \begin{equation} \psi(x,y) = \int_0^\infty \!\!\int_0^\infty \varphi_0(x')\varphi_0(y') G(x,y;x',y')\; x'y'\; dx'\; dy' \label{eq:green_integral} \end{equation} For simplicity we are not going to compare any scattering information extracted from the wave function as it involves additional operations like surface integration, which can lead to additional loss of accuracy and need to be studied separately. We will compare the values of the wave function in a fixed spatial point. As the spatial grid is different for every size of the oscillator basis we can not ensure that any fixed spatial point lies exactly on the grid point at every calculation, so we need to interpolate. We have used a cubic spline interpolation, and from comparison to the other interpolation methods we can expect that the additional error introduced by this operation is negligible compared to the errors of the method (of course, that is only if the function is smooth enough, that means not very high values of $E$). As we do not have any potential in this model problem, the convergence of the proposed hybrid oscillator representation will depend mainly on the accuracy of the asymptotic relations used. This means that we can choose the size of the spatial domain on which we solve the two--dimensional radial problem as small as possible but not smaller than the region spanned by the exact oscillator basis. As the maximal size of the oscillator basis we use in this calculations is $350$ and the oscillator length was chosen as $b=0.7$ ($R_N \approx 26$) the size of the spatial domain was chosen as $30$ dimensionless length units of the real grid and additional $10$ units of the ECS layer. The value of the wave function was extracted at the point $(28,28)$ when we increase the basis size simultaneously and at $(28,12)$ when we keep the basis size fixed in the $y$-dimension. In Figure \ref{fig:conv_wf_2d} we compare the numerical solution of the linear system with the one obtained from~\eqref{eq:green_integral} which we consider as exact. We see that, similar to Figure~\ref{fig:coeff_2d}, if we expand the basis only in one direction the convergence is of the order $N^{-1/2}$. \begin{figure} \begin{center} \includegraphics[width=0.45\linewidth]{conv_wf2d_diag} \includegraphics[width=0.45\linewidth]{conv_wf2d_fix} \end{center} \caption{Convergence of the two-dimensional scattering wave function for the problem in Eq.~\eqref{eq:simple_2d_problem} at the fixed spatial point. Left figure represents the case where the basis was increased in two dimensions simultaneously ($N = M$), right figure shows the convergence with a fixed $M = 50$. Both calculations are made with $b=0.7$ and $E=2$.} \label{fig:conv_wf_2d} \end{figure} Next we test the method on a model potential scattering problem described by (\ref{eq:2d_scattering}) with zero partial angular momenta and the Gaussian interaction potential in the form \begin{equation} V(x,y) = -3\mathrm{e}^{-x^2} - 3\mathrm{e}^{-y^2} + 10 \mathrm{e}^{-x^2-y^2} \label{eq:2d_pot} \end{equation} The one-body potential, here $V_1(x) = -3\exp(-x^2)$, can support one bound state. This means that using the potential \eqref{eq:2d_pot} we can model a three-body breakup problem using the right--hand side in the form \begin{equation} \chi(x,y) = -(V(x,y)-V_1(x))f_0(x)\hat j_0(y), \end{equation} where $f_0(x)$ is the wave function of the bound state, $\hat j_0(y)$ is the Riccati-Bessel function and together they represent the initial state of the system. For the considered problem there are no analytical results to compare with, so on the Figure \ref{fig:surf_2d} we present the solution of the linear system for $c_{nm}$ and the spatial wave function, reconstructed from $c_{nm}$ by applying the transformation matrix $W \otimes W$. On both figures we can clearly recognize the patterns of elastic scattering (the plane waves along the axes) and the breakup (the radial waves with lower frequency as part of the energy was spent to break the bound state). The in-depth analysis of the three--body scattering results is the subject of future work. \begin{figure} \begin{center} \includegraphics[width=0.49\linewidth]{surf_c} \includegraphics[width=0.49\linewidth]{surf_psi} \end{center} \caption{The solution of three--body scattering problem in oscillator representation (left figure) and the reconstructed spatial wave function (right figure). The calculation is made with $b=1$ and $E=1$. The problem has Dirichlet boundaries on all sides. } \label{fig:surf_2d} \end{figure} \section{Conclusions and outlook} This paper focuses on scattering calculations in the oscillator representation, where the solution is expanded in the eigenstates of the harmonic oscillator. The oscillator representation is not the most natural representation to describe scattering processes since it involves a basis set that is designed to describe bound states. This leads to a rather low convergence and may result in a linear system with very large dense matrix. It is often more natural to describe a scattering process with the help of a grid representation. These grid--based calculations have been very successful in describing scattering and breakup processes in atomic and molecular physics. The Helmholtz equation is also efficiently solved on these grids. However, internal structure of the nuclear clusters and other many--particle systems are efficiently described by such an oscillator basis, since the bottom of the potential can be mimicked by the oscillator potential leading to an efficient description of the internal structure. In this paper we combine the advantages of grid--based calculations with those of the oscillator representation. The method was originally proposed in \cite{PhysRevC.82.064603}, as a further development of the $J$-matrix or algebraic method for scattering. There the method combined the grid and oscillator representation with a low--order asymptotic formula. In this paper we have improved this matching with a higher--order approximation and this brings the overall error down to the level of the discretization error of the finite--difference grid. Although a similar asymptotic formula appeared earlier in the work of S. Igashov \cite{igashov_jmatrix}, we believe that the asymptotic formula presented in the current paper is more generally applicable. Furthermore, we have used the asymptotic formulas to improve the accuracy and convergence of the hybrid simulation method. The convergence of the method is illustrated with various examples from two--body and three--body scattering. In preparation of this papers, our initial efforts involved application of the strategy showed in Figure \ref{fig:coupling} with the higher--order formula. However, this required the use of a forward and backward stencils to estimate the third derivative in the first grid point of the finite difference representation. This strategy, however, gives rise to eigen modes with a negative energy localized near the interface. This destroys the positive definiteness of the Laplace operator. These modes are avoided in the proposed method that uses a symmetric stencil around the matching point. This gives satisfactory convergence. We have applied the asymptotic technique to radial oscillator state that are based on Laguerre polynomials. Similar results can be derived for the 1D oscillator states that are based on the Hermite polynomials. These Hermite polynomials are closely related to Gaussian Type Orbitals (GTO) that are frequently used in computational quantum chemistry \cite{mcmurchie1978one} . It is worth to explore if the asymptotic formulas make it possible to combine GTO's with grid based calculations to describe molecular scattering processes. In the future further research is necessary on asymptotic expressions of products of two functions. This will involve convolution integrals. Better expressions for products will also help to take into account asymptotic behavior of the potentials in the grid representation. \section{Acknowledgments} We acknowledge fruitful discussions with F. Arickx and J. Broeckhove and we are grateful to B. Reps and P. K{\l}osiewicz for reading the manuscript. We also thank the anonymous referees for their suggestions. We are thankful for support from FWO--Flanders with project number G.0120.08. \bibliographystyle{elsarticle-num-names}
\section*{Abstract} We present a numerically efficient method to reconstruct a disordered network of thin biopolymers, such as collagen gels, from three-dimensional (3D) image stacks recorded with a confocal microscope. Our method is based on a template matching algorithm that simultaneously performs a binarization and skeletonization of the network. The size and intensity pattern of the template is automatically adapted to the input data so that the method is scale invariant and generic. Furthermore, the template matching threshold is iteratively optimized to ensure that the final skeletonized network obeys a universal property of voxelized random line networks, namely, solid-phase voxels have most likely three solid-phase neighbors in a $3\times3$ neighborhood. This optimization criterion makes our method free of user-defined parameters and the output exceptionally robust against imaging noise. \section*{Introduction} Many biological materials, such as the cytoskeleton or the extracellular matrix, self-organize into complex networks by the polymerization of protein molecules into fibrils (Fig.\,\ref{fig:network}). If the thickness of the fibrils is negligible compared to the pore size, the resulting structure can be mathematically described as a disordered line network. In general, the functional properties of these networks, such as their mechanical stiffness on the macroscopic scale, or their permeability for diffusing particles and for actively migrating cells on a microscopic scale, depend on the geometrical details of the microscopic network structure. In order to study the relationship between structure and function, it is therefore important to extract, or reconstruct, the 3D network structure from image stacks. One aspect of the reconstruction is the binarization of the intensity values of the image stack, so that each voxel is assigned one of two possible values, corresponding either to the solid phase (1, collagen fibers) or the liquid phase (0, surrounding medium). Another aspect of the reconstruction is the skeletonization, so that the optically broadened fibers are reduced to their central (medial) axis, with a width of only one voxel. \begin{figure}[!htb] \centering \includegraphics[width=0.8\linewidth]{figures/NETWORK_CUBE.pdf} \caption{\label{fig:network} \textbf{A cube of collagen gel.} Dimensions $(32 \times 32 \times 34) \mu m^3$, concentration 1.2\,$\frac{\rm{mg}}{\rm{ml}}$, recorded with confocal reflection microscopy and without any image processing. (a) Top view, i.e. in z-direction. (b) Side view. The lateral (x-, y-direction) resolution of the fibers is considerably better than the vertical (z-direction) resolution, due to the anisotropic point spread function. In addition, only fiber segments that run in small angles to the imaging plane are visible, the so-called blind spot effect.} \end{figure} While most of the standard reconstruction methods realize the two aspects of reconstruction (i.e. binarization and skeletonization) in a two-step process, our template matching method achieves both aspects in a single step. This new method avoids the problem of chosing an arbitrary intensity threshold for the binarization. Instead, the template matching algorithm automatically adapts to the input data such that within the reconstructed fraction of solid-phase voxels the most probable number of next neighbors equals three. This represents a universal property of voxelized line networks. \paragraph{Criteria for reconstruction methods} We regard it essential to define the following criteria for our reconstruction method: (1) The method needs to be free of user-adjustable parameters, and (2) be insensitive to variations in the input data quality. To test this criterion we image collagen networks under a wide range of different confocal microscope settings such as amplifyer gain and laser outlet power. The method (3) must be able to correctly reconstruct known networks. To simulate realistic conditions, these networks are convoluted with a point spread function of the imaging system, and different levels of noise are added. \paragraph{Existing reconstruction methods} A vast variety of methods can be used for the reconstruction problem. A large class of these methods works with two separate steps of binarization and skeletonization \cite{pud98, ma96, lee94, pro07, ren09, wan07}. The simplest way to binarize an image stack is by comparing each individual voxel's grayscale with a threshold value $\theta$ and to assign all voxels that are brighter than $\theta$ to the solid phase. As we shall demonstrate below, this method naturally leads to binarized arrays with many artifacts, i.e. false positive and false negative voxels. This includes the simple removal of isolated solid-phase voxels, which can result from noise or dirt particles in the medium. More demanding, it requires the thinning of the broadened binarized fibers to their medial axis of one voxel diameter. Binarization can also lead to the disintegration of fibers, so that closing methods, consisting of dilatation with subsequent erosion steps \cite{soi99, sze10}, have to be applied as well. Another class of methods is based on edge detection with convolution kernels, using, for example, Laplace filters or Sobel operators \cite{soi99, sze10}. This class of methods is also plagued with the production of artifacts that have to be removed afterwards. Finally, there are the class of learning algorithms, such as vector clustering methods \cite{gan07} and neural networks \cite{bis96}, for example the k-means algorithm \cite{bra00} or RBF networks \cite{how07}. A significant advantage of such methods is their ability to automatically adapt to the specific properties of the data at hand. Our proposed template matching algorithm generates its template patterns automatically from the data and can therefore be considered as a learning algorithm as well. \paragraph{Problems with existing methods} A detailed summary and comparison of all reconstruction methods is beyond the scope of this paper. Instead, we shall briefly consider the simple example of global threshold binarization and discuss some of its fundamental shortcomings. This will be useful to highlight the advantages of the template matching method proposed later. We start with an image stack recorded by reflection microscopy. Let us assume that the grayscales of the image stack are coded with 8 bits, i.e. all brightness values $B$ are in the range $B \in [0,255]$, with $B=0$ corresponding to completely dark (black) and $B=255$ to maximum bright (white) voxels. In our setup (Leica SP5X confocal microscope), a typical distribution $p(B)$ of brightness values has a sharp peak around $B=15\pm5$ and a very flat tail towards large values (Fig.\,\ref{fig:SURROGATE}a). The reasonable range of binarization thresholds $\theta$ is located somewhere within this tail. However, the distribution $p(B)$ itself offers no hint as to where the optimum threshold point should be set. To characterize different network reconstruction methods, we use artificially generated image stacks. This requires realistic models of both, the line network itself and its transformation into cross-sectional images by the microscope. As described in more detail in the Methods section, we use a "Mikado" model for the line network, where straight lines of fixed lengths and isotropic orientations are distributed throughout the volume with a homogeneous density \cite{met11}. To model the imaging process, we take into account the broadening (simulated by a convolution with a point spread function), the blind spot effect (i.e. a gradual darkening of steep fibers) \cite{jaw09} and the addition of random noise. The resulting image stacks have statistical properties almost indistinguishable from measured image stacks (Fig.\,\ref{fig:SURROGATE}), but with the advantage that the underlying mathematical line network is known precisely. Is it possible to perfectly reconstruct the original line network using global threshold binarization? This would require the existence of a threshold $\theta$, such that all fluid-phase voxels have brightnesses below and all solid-phase-voxels brightnesses above this threshold. However, when we use our synthetic stack and plot the brightness distributions $p_S(B)$ and $p_F(B)$ of the two phases separately, we find in general two peaks with a significant overlap (Fig.\,\ref{fig:brightnesshistoseperate}). This means that no global threshold can be found, even in principle, for separating the two phases, without also producing some false positive and false negative voxels. The voxels with brightnesses in the overlap interval include, for example, isolated bright points due to noise. It would be relatively easy to remove such cases in a subsequent post-processing step. More problematic is that the overlap interval also includes liquid phase voxels from the narrow gaps between two fibers, which have been raised in brightness beyond the threshold by the superposition of the fibers' point spread functions. This effect would lead to a merging of the two close-by fibers in the binarized image and would require a much more sophisticated procedure to be repaired. Finally, the overlap region includes voxels of fiber segments that are more vertically oriented, and therefore too dark, to exceed the threshold, because of the blind spot problem \cite{jaw09}. We note that a human observer could still recognize such dark fiber segments quite easily. Taken together, the threshold binarization has some fundamental limitations. To a certain extent, the method can be improved by using variable thresholds, which take into account the local brightness conditions in the environment of each voxel to be binarized. This, however, can already be viewed as a first step towards a template matching method that will be discussed in the following. \paragraph{Template matching in line networks} Template matching methods recognize specific image parts within larger image stacks by comparing features, e.g. the brightness patterns, of small sub volumes of the stack with the brightness pattern of pre-defined templates. The templates incorporate the a-priori-knowledge about the features to be found. In the case of line networks, the templates would contain short line segments, that are oriented in arbitrary directions. The number of required templates turns out to be impractically large in 3D. However, the situation is much simpler when only 2D cross-sections are used for the template matching: The vertical cross-section of a broadened line segment with a plane is a elliptical spot of finite size that can be easily recognized by 2D template matching (Fig.\,\ref{fig:SLICES}a). The shape of the spot will vary slightly as the angle of intersection becomes less than 90 degrees. For angles less than 45 degrees, the distortion of the spot can become too large to match the template (Fig.\,\ref{fig:SLICES}b), but in this case the same line segment can be easily recognized by its intersection with a perpendicular plane. Therefore, all line segments (solid voxels) can be detected by sequentially scanning through the x-, y- and z-planes of the sample volume. As shown below this binarization method turns out to be much more reliable and robust than the simple threshold method. The similarity between the cross-sectional brightness pattern and the template will be largest if the template is located exactly at the center of the finite spot. Therefore, the medial axes of the broadened line segments can be identified as local minima of the mismatching measure. In this way, the 2D template matching simultaneously achieves a skeletonization of the broadened fibers. We note that this method meets the design criteria imposed before. In order to eliminate all internal parameters (1), we have implemented an automatic template generator, which is entirely based on the input image stack and requires no user intervention. We will demonstrate in the Results section that our method is also robust with respect to the quality of the input data (2) and yields reconstructions that reproduce simulated line networks almost perfectly (3). The algorithm is implemented in C++ to achieve fast execution times (about 12 minutes on a standard notebook for reconstructing a $(512\times512\times597)$-voxel stack) and is available as a supplement to this publication \section*{Methods} \paragraph{Generation of surrogate data sets} Since the true shape of the underlying fiber network of real grayscale data sets is unknown, we artificially generated surrogate data sets to validate the performance of our algorithm. Firstly we created idealized line networks using a "Mikado" model. Straight lines of fixed lengths and isotropic orientations are distributed throughout the volume with a homogeneous density \cite{met11}. Binary surrogate data sets are derived from these parameterized networks by a voxelization operation. Then we simulate the imaging process to transform these binary data sets into grayscale data sets. Initially $N$ sets of parameters representing $N$ lines are randomly generated. Each parameter set contains uniformly distributed Cartesian coordinates of a line's center point $(x_c,y_c,z_c)$, uniformly distributed azimuthal angle $\varphi \in \left[-\pi,\pi\right]$ with $p(\varphi)=\frac{1}{2\pi}$ and polar angle $\vartheta \in \left[0,\pi\right]$ with $p(\vartheta)=\frac{\sin\vartheta}{2}$. This is an efficient way of representing a line network, since for each arbitrary point within the volume one can unambiguously determine whether or not the point is located on any line. To derive a binary data set from the parameterized network the whole volume is divided into distinct volume elements representing the voxels. Initially all voxels within the 3D-array are set to value 0 (fluid phase). Starting at the first center point, the line is traced in opposite directions according to its orientation $(\varphi,\vartheta)$ until the half length of the line is reached for each direction. While tracing the line, every voxel corresponding to a touched volume element is set to value 1 (solid phase). This process is repeated for all $N$ sets of line parameters. To convert the binary data set into a 8-bit-grayscale data set, we apply a process called numeric blurring which simulates the imaging process. Numeric blurring is done in five subsequent steps: \begin{enumerate} \item The binary 3D-array of voxels is converted into a grayscale 3D-array by setting all voxels with value 1 to brightness values $B \in [0,255]$ according to the polar angle $\vartheta$ of the line to which they belong. This procedure replicates the blind spot effect (i.e. a gradual darkening of steep fibers) as described in \cite{jaw09}. \item A small number of dark voxels is randomly selected and set to brightness values $>200$. This simulates dirt particles within the fluid phase that appear as isolated bright fluctuations in the original microscope images. \item The 3D-array of voxels is convoluted with an anisotropic Gaussian to simulate the point spread function: \begin{align} B(x,y,z) &:= (B \ast psf)(x,y,z) \\ (B \ast psf)(x,y,z) &= \iiint\limits_{}B(x',y',z')psf(x-x',y-y',z-z')dx' \, dy' \, dz' \end{align} where $B(x,y,z) \in \left[0,255\right]$ is brightness of voxel $(x,y,z)$. The point spread function is defined as \\ $psf(x-x',y-y',z-z') = \exp\left(-\left(\frac{(x-x')^{2}}{\sigma_{x}^{2}}+\frac{(y-y')^{2}}{\sigma_{y}^{2}}+\frac{(z-z')^{2}}{\sigma_{z}^{2}}\right)\right)$ with $\sigma_x=\sigma_y<\sigma_z$ according to the characteristics of confocal microscopy. \item A Gaussian distributed random variable is added to each voxel to simulate noise\footnote{We are aware that photon shot noise would not be normally distributed. However, as shown below, the resulting statistical properties of the surrogate stacks agree almost perfectly with measured data.}. \item Finally, all brightness values are rescaled to values from 0 to 255 by affine transformation. \end{enumerate} We have analyzed the statistical properties of the resulting grayscale data sets\footnote{It is possible to determine the direction vector of a short fiber segment, even from its voxelized representation, by treating the brightness distribution as a mass distribution, computing and diagonalizing the moment of inertia tensor, and finding the principal component axis of mimimal inertia. This principal component corresponds to the direction of the locally straight line segment. Several conditions have to be met when analyzing a given small test volume: First, the fluid background of the fiber segment in the test volume should have a much smaller (ideally zero) brightness/mass than the fiber itself. It is therefore advisable to use already reconstructed stacks for this analysis. Second, the test volume should contain enough solid voxels to clearly define a single fiber segment. Third, the test volume should not be so large to contain several line-like objects with different directions. We have therefore used the following method: A number of PMX=$10^5$ spherical test volumes of radius r=3.0 vu (voxel unit, i.e. linear size of a single voxel) have been chosen randomly throughout the reconstructed stack, ensuring that each test volume contains at least NMIN=5 solid voxels. Inside each sphere, the voxels were treated as mass points, located at the voxel centers, and with constant mass m=1 for all solid and m=0 for all liquid voxels. After determining the easy axis of the inertia tensor, the corresponding unit direction vector was computed. Note that this vector does not depend on the exact position of the test sphere's center, as long as the same solid voxels are enclosed. From this Cartesian vector, the azimuthal angle $\varphi$ and the polar angle $\vartheta$ in spherical coordinates were computed. Finally, histograms were generated for $\varphi$ and $\vartheta$.}. They are almost indistinguishable from real data sets imaged with confocal reflection microscopy (Fig.\,\ref{fig:SURROGATE}), but with the advantage that the underlying mathematical line network is known in detail. \begin{figure} [!htb] \begin{center} \includegraphics[width=1.0\linewidth]{figures/SURROGATE.pdf} \end{center} \caption{\textbf{Statistical properties of real and surrogate image stacks.} (a) Comparison of the brightness distributions in the real and surrogate image stacks. Both distributions are similar. (b) and (c) show angular distributions of the fiber segments. (b) Typical distributions of azimuthal angles $\varphi$ in a real and a surrogate data set. The distributions are almost indistinguishable. The peaks are a result of voxelization. The principal directions, corresponding to the x- and y-direction, as well as the principal diagonals are over-represented in short fiber segments and lead to maxima at $\varphi=0, \pm\frac{\pi}{4}, \pm\frac{\pi}{2}, \pm\frac{3\pi}{4}, \pm\pi$. (c) Typical distributions of polar angles $\vartheta$ in a real and a surrogate data set. Again, the distributions are similar. Compared to an ideal isotropic network with $p(\vartheta)\propto \sin(\vartheta)$, polar angles smaller than $\frac{\pi}{2}$ are increasingly suppressed due to the blind spot effect of confocal reflection microscopy \cite{jaw09}.} \label{fig:SURROGATE} \end{figure} When we plot the brightness distributions $p_S(B)$ and $p_F(B)$ of the two phases (solid and fluid) separately, we find in general two peaks with a significant overlap (Fig.\,\ref{fig:brightnesshistoseperate}). This means that no global threshold can be found, even in principle, for separating the two phases, without also producing some false positive and false negative voxels. \begin{figure} \begin{center} \includegraphics[width=0.5\linewidth]{figures/brightnesshistoseperate.pdf} \end{center} \caption{\textbf{Brightness distributions $p_S(B)$ and $p_F(B)$ of the solid and the fluid phase.} The two peaks show a significant overlap. This clearly points out that no global threshold can be found, even in principle, for separating the two phases, without also producing some false positive and false negative voxels.} \label{fig:brightnesshistoseperate} \end{figure} \paragraph{Preprocessing} Some data sets show a z-dependence of mean grayscale. The average brightness of each z-slice $\mu(z)$ is not constant in all layers, but rather decreases for deeper located slices in the stack. This effect is caused by scattering and absorption of light by collagen fibers and ambient medium. For compensation the data set is firstly normalized to an equal global mean grayscale, disregarding statistical fluctuations: \begin{align} B_{xyz} &:= B_{xyz}-\mu(z) & \forall \ x,y,z \\ \intertext{where $ \mu(z)=\frac{1}{N_x N_y}\sum\limits_{i=1}^{N_x}\sum\limits_{j=1}^{N_y} B_{ijz}$} \intertext{and secondly rescaled to values $\left[0,255\right]$ by affine transformation:} B_{xyz} &:= 255\cdot\frac{B_{xyz} - \min\left\{B_{ijk}\right\}} {\max\left\{B_{ijk}\right\} - \min\left\{B_{ijk}\right\}} & \forall \ x,y,z,i,j,k \end{align} \paragraph{Automatic template generation} The number of required templates turns out to be impractically large in 3D. However, the situation becomes much easyer when only 2D cross-sections are used for the template matching: The vertical cross-section of a broadened line segment with an image plane is a ellitpical spot of finite size that can be easily recognized by 2D template matching (Fig.\,\ref{fig:SLICES}). \begin{figure} [!htb] \begin{center} \includegraphics[width=0.35\linewidth]{figures/SLICES.pdf} \end{center} \caption{\textbf{2D cross-section of a 3D image stack.} (a) The vertical cross-section of a broadened line segment with a plane is a elliptical spot of finite size that can be easily recognized by 2D template matching. (b) The shape of the spot is varying slightly as the angle of intersection becomes less than 90 degrees. For angles less than 45 degrees, the distortion of the spot can become too large to match the template.} \label{fig:SLICES} \end{figure} The templates are derived automatically from the gray scale data set. The described process requires neither any knowledge of the point spread function's characteristics nor any user interactions. Since fiber detection is performed in three directions (x, y and z) and the point spread function may be anisotropic, three different templates $\mathbf T_x, \mathbf T_y, \mathbf T_z$ are required. One for each direction. \begin{enumerate} \item A sufficient large sample of voxels ($10^5$) is picked out randomly. In order to minimize computation time only voxels with intensities greater than the global mean intensity $\mu$ were accepted. \item Each voxel is the center of three small 2D images ($I_x, I_y, I_z$), where the first image is located in the yz-, the second in the xz- and the third in the xy-plane. The images' height $H$ and width $W$ are initially set to arbitrary small values. The adaptive resizing process is described in a following section. \item These small images are represented as $H\times W$ matrices $\mathbf P_1,\mathbf P_2,\ldots,\mathbf P_N$, with matrix entries corresponding to voxel brightnesses. Hence three sets of training patterns $\left\{\mathbf P^x_i\right\}, \left\{\mathbf P^y_i\right\}, \left\{\mathbf P^z_i\right\}$ are obtained to compute the three different templates. \item From all training patterns belonging to a set, the weighted average pattern $\mathbf A$ is computed: \begin{equation} \mathbf{A}:=\frac{\sum\limits_i w_i\mathbf P_i}{\sum\limits_i w_{i}} \end{equation} where $w_i$ is the central entry of matrix $\mathbf P_i$. \item To become independent from absolute brightnesses the global mean gray scale $\mu$ is subtracted from each matrix entry. \item Finally the matrices $\mathbf{A_x}, \mathbf{A_y}, \mathbf{A_z}$ are normalized to obtain the templates $\mathbf T_x, \mathbf T_y, \mathbf T_z$: \begin{equation} \mathbf{T}:=\frac{1}{\left\|\mathbf{A}\right\|} \, \mathbf{A} \end{equation} \end{enumerate} \paragraph{Fiber detection process} As previously mentioned the fiber detecting process in the gray scale input data set is performed subsequently for all three directions x, y and z. During these three cycles three temporary binary output data sets $\mathbf O_x, \mathbf O_y, \mathbf O_z$ are obtained, having each voxel labeled with one of two possible values 0 (fluid phase) or 1 (collagen phase). The similarity between a given cross-sectional brightness pattern and the corresponding template will be largest if the template is located exactly at the center of the finite spot. Therefore, the medial axes of the broadened line segments can be identified as local minima of the mismatching measure. The fiber detection process includes the following steps: \begin{enumerate} \item Initially all voxels of the first x-slice (yz-plane) with brightnesses larger than the mean gray scale $\mu$ are investigated. This restriction is only made to decrease computation time. Without this restriction the fraction of additionally detected fiber voxels is less than $10^{-5}$. This small fraction is negligible since it does not significantly affect the resulting network properties on the one hand. On the other hand, performance time would be increased drastically by a factor of 5, because 80\% of all voxels are darker than the mean brightness. Consequently all voxels with gray scales smaller than $\mu$ are set to 0 in the binarized output data set $\mathbf O_x$. \item The voxels to be investigated are taken as centers of small 2D-images located in the yz-plane. These images are represented as matrices with the same size as the corresponding template. These matrices are the unknown patterns to be compared with the template. We call them search patterns $\mathbf S_n$. \item From each entry of $\mathbf S_n$ the local mean value $\mu_n$ is subtracted. \item The search patterns are normalized: \begin{equation} \mathbf{S_n}:=\frac{1}{\left\|\mathbf{S_n}\right\|} \, \mathbf{S_n} \end{equation} \item Now the mismatch $d_n$ between search patterns $\mathbf{S_{n}}$ and template pattern $\mathbf{T_{x}}$ for detection in x-direction is computed. The smaller $d_n$, the more probable the central voxel of the search pattern belongs to the collagen phase. As mismatch measuring metric we chose the Euclidean distance in feature space. Hence we call $d_n$ the matching distance. \begin{equation} d_{n}:=\left\|\mathbf{S_{n}}-\mathbf{T_x}\right\|=\sqrt{\sum\limits_{i}\sum\limits_{j}\left(s_{ij}^{n}-t_{ij}^{x}\right)^{2}} \end{equation} Since all matrices are normalized, the range of $d_n$ is $\left[0,2\right]$ where $d_n=0$ means perfect matching. \item The calculated matching distances $d_n$ are compared with the threshold $\theta_d^x$. This threshold defines whether the search pattern is sufficiently similar to the template, so the pattern's central voxel may be a fiber voxel. For $d_{n}>\theta_d^x$ the corresponding voxels in $\mathbf O_x$ are set to 0. Note that all thresholds $\theta_d^x, \theta_d^y, \theta_d^z$ are not chosen arbitrarily, but derived adaptively from the input data. The detailed method of defining these thresholds will be described in a later section. \item To all voxels being considered ($d_{n}<\theta_d^x$) a local minimum filter is applied, labeling only local best matching voxels as belonging to collagen phase by assigning the value 1, while all others are set to 0. After having finished this step the first x-slice (yz-plane) of the input data set is completely binarized and stored in the temporary binary output data set $\mathbf{O_x}$. \item All previous steps are repeated for all other x-slices resulting the first binarized data set $\mathbf{O_x}$. \item Some fibers with angles less than 45 degrees to any yz-plane (x-slice) may be not well detected since their cross-sections, appearing as roundish spots of finite size, are distorted and hence are not sufficiently similar to the template. Therefore the complete detection process is repeated in perpendicular xz- and xy-planes (y- and z-slices) of the sample volume. Having all y- and z-slices binarized two more binary data sets $\mathbf{O_y}$ and $\mathbf{O_z}$ are obtained. \item Finally the three temporary reconstruction results are combined to one binary data set using logical OR: \begin{equation} O_{ijk} := O_{ijk}^x \vee O_{ijk}^y \vee O_{ijk}^z \ \ \ \ \forall \ i,j,k \end{equation} \item As a post-processing step all isolated voxels (i.e. solid phase voxels with all 26 neighbors belonging to fluid phase) are converted to fluid phase. \end{enumerate} \paragraph{Define optimal template sizes} The right size $H\times W$ of the template $\mathbf{T} \in \mathds R^{H\times W}$ is crucial. On the one hand the template must not be too small, because otherwise the pattern is not completely covered. On the other hand templates should not be too large to limit computation time. A good indication is given by the algebraic sign of template's matrix entries. Since all entries are reduced by subtracting the mean gray scale $\mu$, positive entries correspond to gray scales brighter than and vice versa negative entries to gray scales darker than the average. Taking into account that the patterns to be detected are resulting from convolution of bright fibers with the point spread function it is reasonable to consider positive matrix entries as belonging to the pattern's foreground, while negative entries represent the surrounding background. Hence, if the template is sized in a kind that all outer entries are negative and at the same time all inner entries are positive it is warranted that the pattern is completely covered by the template. Furthermore using both, positive and negative matrix entries, implies better use of the complete domain of definition and contrast enhancement since matching distances $d_n$ only will be minimal if the template is perfectly centered at the patterns being investigated. Initially height $H$ and width $W$ are set to arbitrary, small values, e.g. $H=15$ and $W=9$. Then the templates are iteratively resized and recalculated until the optimal size is found. Since computing the template takes only a few seconds the complete runtime is not affected significantly by these iterations. \paragraph{Define optimal matching thresholds} The optimal thresholds $\theta_d^x$, $\theta_d^y$ and $\theta_d^z$ are also defined iteratively. It is clear that it depends on the thresholds how many voxels are labeled as fiber voxels. Because both templates and search matrices are normalized the maximum range of matching distances $d_n$ is $\left[0,2 \right]$ and consequently the optimal thresholds are in the same range as well. \newline If matrices are treated as vectors in high-dimensional feature space, three special cases can be distinguished: \begin{description} \item[Identity] \noindent\hspace*{20mm} $d_n=0$ \noindent\hspace*{10mm} $\Rightarrow$ \noindent\hspace*{10mm} $\mathbf{S}_n \ \equiv \ \ \mathbf{T}$ \item[Orthogonality] \noindent\hspace*{9.5mm} $d_n=\sqrt{2}$ \noindent\hspace*{7mm} $\Rightarrow$ \noindent\hspace*{10mm} $\mathbf{S}_n\ \ \bot \ \ \ \mathbf{T}$ \item[Inversion] \noindent\hspace*{18mm} $d_n=2$ \noindent\hspace*{10mm} $\Rightarrow$ \noindent\hspace*{10mm} $\mathbf{S}_n \ =-\mathbf{T}$ \end{description} Where orthogonality means a maximum dissimilarity between template and search matrix. Inversion implies identical absolute values of corresponding matrix entries with inverted algebraic signs. Hence $\left[\sqrt{2},2 \right]$ is no expedient range for the thresholds which rather must be within $\left[0,\sqrt{2} \right]$. Since reconstructed fibers should be skeletonized, the most probably number of direct fiber voxel neighbors $E_{mode}$ in a $3^3$-neighborhood of a central fiber voxel turns out to be a good criterion. Extensive evaluations of simulated line networks showed $E_{mode}=3$ in case of perfect skeletonization. Obviously there is not only a single value for thresholds that achieves $E_{mode}=3$, but rather a range, where the optimal thresholds would be the top of this range, because this causes a maximum number of detected fibers while simultaneously the constraint $E_{mode}=3$ is fulfilled. \newline The threshold for each reconstruction direction x, y and z is defined in two subsequent steps. Firstly a threshold that fulfills $E_{mode}=4$ is found using a binary search algorithm. And secondly the found threshold is reduced step by step until it fulfills $E_{mode}=3$. \subparagraph{Binary Search} Binary search is an efficient standard algorithm for searching a specified value by halving the number of items to check with each iteration \cite{cor09, aro09}. Initially the threshold is set to the middle of the range to be searched $\left[0,\sqrt{2} \right]$ \begin{equation} \theta_d^{(0)}:=\frac{0+\sqrt{2}}{2}=\frac{1}{\sqrt{2}} \end{equation} and the bounds are defined \begin{align} \theta_d^{max(0)} & := \sqrt{2} \\ \theta_d^{min(0)} & := 0 \end{align} \newline Then the following steps are repeated until the stop criterion is fulfilled: \begin{enumerate} \item The fiber detection algorithm is performed with recent threshold $\theta_d^{(n)}$. To limit computation time not the complete data set (containing $512 \times 512 \times 597$ voxels) is used but a smaller sub set containing only $150 \times 150 \times 150$ voxels. \item The most probable number of fiber voxel neighbors $E_{mode}$ is evaluated. \item The threshold and the bounds of the searching range are updated: \\ If $E_{mode}<4$ then: \begin{align} \theta_d^{max(n+1)} & := \theta_d^{max(n)} \\ \theta_d^{min(n+1)} & := \theta_d^{(n)} \\ \theta_d^{(n+1)} & := (\theta_d^{max(n+1)}+\theta_d^{min(n+1)})/2 \end{align} If $E_{mode}>4$ then: \begin{align} \theta_d^{max(n+1)} & := \theta_d^{(n)} \\ \theta_d^{min(n+1)} & := \theta_d^{min(n)} \\ \theta_d^{(n+1)} & := (\theta_d^{max(n+1)}+\theta_d^{min(n+1)})/2 \end{align} If $E_{mode}=4$ then binary search is stopped. \end{enumerate} \subparagraph{Reduction of threshold} The threshold is now iteratively reduced until $E_{mode}=3$: \begin{enumerate} \item The fiber detection algorithm is performed using the recent threshold $\theta_d^{(n)}$. Again due to limit computation time not the complete data set is used but a sub set. However containing now more voxels (i.e. $250 \times 250 \times 250$) than previously used for binary search to increase accuracy. \item The most probably number of fiber voxel neighbors $E_{mode}$ is evaluated. \item If $E_{mode}=4$ then the recent threshold is reduced by subtracting a small $\varepsilon$. \\ If $E_{mode}=3$ then the optimal threshold is found and the iteration loop is stopped. \\ Note that the smaller $\varepsilon$ is chosen the more exact the best threshold is found on the one hand but on the other hand the more iteration steps are required. As a good compromise to achieve both high accuracy and runtime limitation we found $\varepsilon=0.01$. \end{enumerate} \paragraph{Distribution of nearest obstacle distances} The geometric properties of line networks, such as the pore sizes, are good indicators to estimate the similarity between different networks. The pore sizes of a network can be quantified in different ways, for instance by placing within each pore a sphere of the maximum possible size and then analyzing the size distribution of these spheres \cite{Mic08}. In this paper, we choose another, yet equivalent approach: We compute the distribution $p(r_{\rm{no}})$ of nearest obstacle distances in the binarized network. This is done by selecting a set of random points within the stack, computing the distance from each test point to its closest solid state obstacle (i.e. fiber segment) and then finding the distribution of these distances \cite{met11, ME-Paper}. \paragraph{Quality measures} To evaluate the validity of our algorithm we defined two quality measures based on Pearson product-moment correlation coefficient using the surrogate data sets. The \emph{sample correlation coefficient of local averaged voxel arrays} $r(\overline{S},\overline{R}) \in [-1,1]$ and the \emph{sample correlation coefficient of the distributions of nearest obstacle distances} $r(p_S,p_R) \in [-1,1]$. A value of $r(\overline{S},\overline{R})=1$ indicates a complete linear dependence between the two data sets and hence implies a perfect reconstruction of the network, while $r(\overline{S},\overline{R})=0$ corresponds to linear independence, i.e. both networks are entirely different. In a similar manner, $r(p_S,p_R)=1$ indicates that both distributions are identic. The reconstructed fibers, derived from grayscale surrogate data sets, are not exactly straight but rather smoothly fluctuating. That means that for a given solid phase voxel in binary surrogate data set the position of the corresponding solid phase voxel in the reconstructed data set may differ. However, the range of difference does not exceed one voxel size in each direction. Taking into account that such small fluctuations do not affect the network's global properties, we do not compare the binary data sets (i.e. binary surrogate and binary reconstruction) itselves, but local averaged voxel arrays $\overline{S}$ and $\overline{R}$. Therefore both binary data sets to be compared are converted by setting each voxel to the average value of itself and its 26 direct neighbors. The resulting arrays provide some significant advantages: The information of local solid voxel density is preserved since larger average values are corresponding to a larger number of solid phase voxels within a $3^3$-neighborhood. Furthermore, independency against small differences from exact positions is achieved and the global solid voxel distribution, i.e. the network morphology, is also preserved. After having converted the data sets to be compared, the sample correlation coefficient $r(\overline{S},\overline{R})$ of voxel values is calculated. To compare the distributions of nearest obstacle distances $p_S(r_{\rm{no}})$ and $p_R(r_{\rm{no}})$ in binary surrogate and reconstruction result, firstly both distributions are evaluated and secondly the empirical correlation coefficient $r(p_S,p_R)$ is calculated. \section*{Results} In the following, we show the results of binarizing grayscale image stacks with the template matching algorithm, i.e. the insensitivity to variations in the input data quality and the correct reconstruction of known networks. A typical reconstruction result of original microscope image stacks can be seen in Fig.\,\ref{fig:recresults}. \begin{figure} [!htb] \begin{center} \includegraphics[width=0.5\linewidth]{figures/recresults.pdf} \end{center} \caption{\textbf{Reconstruction result.} Three subsequent original microscope images (left) and the corresponding binarizations (right) generated by the template matching algorithm.} \label{fig:recresults} \end{figure} \subsection*{Insensitivity to variations in the input data quality} We first showed that our binarization algorithm was widely independent from the setup parameters of the imaging process, such as the laser outlet power of the confocal microscope that defines the global picture brightness and the gain of the photomultiplier tubes (PMTs) which mostly determines the signal-to-noise-ratio (SNR) of the images. Therefore, both parameters were changed systematically in the available range. The laser outlet power was shifted from 10\,\% to 90\,\% of the laser's produced beam, the gain being constantly fixed to an apropriate value for 50\,\% laser outlet power. Furthermore, the gain was varied about 50\,V around the fitting value for a fixed laser power of 50\,\%. The results of a sequence of changed parameters can be seen in Fig.\,\ref{fig:INDIPEND}. \begin{figure} [!htb] \begin{center} \includegraphics[width=0.8\linewidth]{figures/INDIPEND.pdf} \end{center} \caption{\textbf{Insensitivity of the algorithm to variations in the input data quality.} The algorithm produces stable results in a wide range of photomultiplier gain and laser outlet power. Hence it is insensitive to variations in the input data quality. (a) Evaluated pore size as a function of photomultiplier gain (signal-to-noise-ratio). (b) Evaluated pore size as a function of laser outlet power (image brightness). The data in (a) and (b) correspond to two collagen gels that have been fabricated under identical conditions. The slight differences in the observed pore sizes reflect natural sample-to-sample fluctuations.} \label{fig:INDIPEND} \end{figure} \subsection*{Correct reconstruction of known networks} We calculated $r(\overline{S},\overline{R})$ to evaluate the similarity between the surrogate and the reconstructed networks (compare Methods Section). To highlight the performance of the template matching algorithm (PM), we also calculated $r(\overline{S},\overline{R})$ for a simple threshold binarization algorithm (TH). We used a threshold $\theta=\mu+2\sigma$, with mean grayscale $\mu$ and standard deviation of grayscales $\sigma$, since extensive tests have shown, that this rough rule of thumb provides a quite fair value for the threshold. Typical results are: \begin{quote} $r_{PM}(\overline{S},\overline{R})=0.84$ \ \ and \ \ $r_{TH}(\overline{S},\overline{R})=0.46$ \end{quote} The algorithm's correct reconstruction was proved by using surrogate data sets. We compared the distributions $p_S(r_{\rm{no}})$ and $p_B(r_{\rm{no}})$ of nearest obstacle distances for about 100 surrogate and reconstructed networks and calculated the correlation coefficient. As can be seen in the example of Fig.\,\ref{fig:quality}, the distributions are almost identical with a sample correlation coefficient of $0.93$. \begin{figure} [!htb] \begin{center} \includegraphics[width=0.4\linewidth]{figures/quality.pdf} \end{center} \caption{\textbf{Distributions of nearest obstacle distances $p_S(r_{\rm{no}})$ and $p_B(r_{\rm{no}})$ in binary surrogate data set and reconstruction result.} Both distributions are, disregarding statistical fluctuations, almost identic.} \label{fig:quality} \end{figure} \subsection*{Summary and Outlook} In this paper we have presented a fast, robust and objectively tested method to reconstruct disordered fiber networks from confocal image stacks. In the original stacks, visible fiber segments appear as "cylindrical clouds" with a bright core, surrounded by a broad "halo" with slowly decaying gray level. After reconstruction, the fiber segments are represented by contiguous voxels with value 1, while all background voxels are assigned the value 0. These reconstructed traces, due to the remaining voxelization, "wiggle" around the smooth space curve of the fiber's medial axis. This level of reconstruction is sufficient for many purposes, such as the statistical evaluation of the distribution of nearest obstacle distances in the fiber network. However, other statistical investigations, for example evaluating the distribution of curvatures along the fibers, would benefit from a parametrized description of each visible fiber segment $s$ in terms of a space curve $\vec{R}_s(t)$. This could be achieved by a suitable post-processing of the present, voxelized representation. Alternatively, our template matching algorithm could be extended to sub-voxel accuracy. In this case the 2D position of the fiber center would be treated as a continuous variable within each of the cross sectional planes. The local mismatch minimum can then be found, with arbitrary spatial resolution, using standard continuum optimization techniques. Once the fiber centers are determined in each cross sectional plane, they can be connected by straight line segments or spline-interpolated to obtain the space curves $\vec{R}_s(t)$. \section*{Acknowledgments} We are grateful for financial support by the German Research Foundation (DFG). \clearpage \newpage \section*{References} \renewcommand\refname{}
\section{Introduction} \bigskip Fibonacci numbers, Lucas numbers and Perrin numbers ar \begin{equation*} F_{n}=F_{n-1}+F_{n-2}\text{ for\ }n>2\text{ \ and }F_{1}=F_{2}=1, \end{equation* \begin{equation*} L_{n}=L_{n-1}+L_{n-2}\text{ \ for }n>1\text{ \ and }L_{0}=2,\text{ }L_{1}=1, \end{equation* \begin{equation*} R_{n}=R_{n-2}+R_{n-3}\text{ \ for }n>3\text{ \ and }R_{0}=3,\text{ }R_{1}=0 \text{ }R_{2}=2 \end{equation* respectively. There are large amount of studies on these sequences. In addition, generalization of these sequences have been studied by many researchers. Miles [10] defined generalized order-$k$ Fibonacci numbers(GO$k$F) as \begin{equation} f_{k,n}=\sum\limits_{j=1}^{k}f_{k,n-j}\ \end{equation for $n>k\geq 2$, with boundary conditions: $f_{k,1}=f_{k,2}=f_{k,3}=\cdots =f_{k,k-2}=0$ and $f_{k,k-1}=f_{k,k}=1.$\newline \bigskip Er [2] defined $k$ sequences of generalized order-$k$ Fibonacci numbers ($k SO$k$F) as; for $n>0,$ $1\leq i\leq k \begin{equation} f_{k,n}^{\text{ }i}=\sum\limits_{j=1}^{k}c_{j}f_{k,n-j}^{\text{ }i}\ \ \label{mil} \end{equation with boundary conditions for $1-k\leq n\leq 0,$ \begin{equation*} f_{k,n}^{\text{ }i}=\left\{ \begin{array}{l} 1\text{ \ \ \ \ \ if \ }i=1-n, \\ 0\text{ \ \ \ \ \ otherwise, \end{array \right. \end{equation* where $c_{j}$ $(1\leq j\leq k)$ are constant coefficients, $f_{k,n}^{\text{ i}$ is the $n$-th term of $i$-th sequence of order $k$ generalization. For c_{j}=1$, $k$-th sequence of this generalization involves the Miles generalization(\ref{mil}) for $i=k,$ i.e \begin{equation} f_{k,n}^{k}=f_{k,k+n-2}. \label{fib} \end{equation} Kili\c{c} and Ta\c{c}c\i \lbrack 3] defined $k$ sequences of generalized order-$k$ Pell numbers ($k$SO$k$P) as; for $n>0,$ $1\leq i\leq k \begin{equation} p_{k,n}^{\text{ }i}=2p_{k,n-1}^{\text{ }i}+p_{k,n-2}^{\text{ }i}+\cdots +\ p_{k,n-k}^{\text{ }i}\ \label{pell} \end{equation with initial conditions for $1-k\leq n\leq 0,$ \begin{equation*} p_{k,n}^{\text{ }i}=\left\{ \begin{array}{l} 1\text{ \ \ \ \ \ if \ }i=1-n, \\ 0\text{ \ \ \ \ \ otherwise, \end{array \right. \end{equation* where $p_{k,n}^{\text{ }i}$ is the $n$-th term of $i$-th sequence of order k $ generalization. \bigskip MacHenry [7] defined generalized Fibonacci polynomials $(F_{k,n}(t))$, Lucas polynomials $(G_{k,n}(t))$, where $t_{i}$ $(1\leq i\leq k)$ are constant coefficients of the core polynomia \begin{equation} P(x;t_{1},t_{2},\ldots ,t_{k})=x^{k}-t_{1}x^{k-1}-\cdots -t_{k}, \label{core} \end{equation which is denoted by the vector \begin{equation} t=(t_{1},t_{2},\ldots ,t_{k}). \label{vkt} \end{equation $F_{k,n}(t)$ is defined inductively by \begin{eqnarray} F_{k,n}(t) &=&0,\text{ }n<1 \label{fippol} \\ F_{k,1}(t) &=&1 \notag \\ F_{k,2}(t) &=&t_{1} \notag \\ F_{k,n+1}(t) &=&t_{1}F_{k,n}(t)+\cdots +t_{k}F_{k,n-k+1}(t). \notag \end{eqnarray} $G_{k,n}(t_{1},t_{2},\ldots ,t_{k})$ is defined b \begin{eqnarray} G_{k,n}(t) &=&0,\text{ }n<0 \label{lucpol} \\ G_{k,0}(t) &=&\text{ }k \notag \\ G_{k,1}(t) &=&t_{1}\text{ } \notag \\ G_{k,n+1}(t) &=&t_{1}G_{k,n}(t)+\cdots +t_{k}G_{k,n-k+1}(t). \notag \end{eqnarray} Moreover in [8,9], MacHenry obtained some properties of these polynomials. In [9] MacHenry gave a relation between generalized Fibonacci and Lucas polynomials as; \ \begin{equation*} G_{k,0}(t)=k,G_{k,n}(t)=F_{k,n+1}(t)+\sum_{j=1}^{k-1}jt_{j+1}F_{k,n-j}(t). \end{equation*} Equivalently this relation can be written as \begin{equation*} G_{k,0}(t)=k,\text{ }G_{k,n}(t)=\sum_{j=1}^{k}jt_{j}F_{k,n-j+1}(t). \end{equation*} \bigskip Kaygisiz and \c{S}ahin [4] definied generalized Perrin polynomials by using generalized Lucas Polynomials. For $k\geq 3;$ \begin{eqnarray} R_{k,0}(t) &=&k \notag \\ R_{k,1}(t) &=&0 \notag \\ R_{k,2}(t) &=&2t_{2} \notag \\ R_{k,3}(t) &=&t_{2}R_{k,1}(t)+3t_{3} \label{perpol} \\ R_{k,4}(t) &=&t_{2}R_{k,2}(t)+t_{3}R_{k,1}(t)+4t_{4} \notag \\ &&\vdots \notag \\ R_{k,k-1}(t) &=&t_{2}R_{k,k-3}(t)+\cdots +t_{k-1}R_{k,1}(t)+kt_{k} \notag \end{eqnarray and for $n\geq k$ \begin{equation*} R_{k,n}(t)=\sum\limits_{i=2}^{k}t_{i}R_{k,n-i}(t). \end{equation*} \begin{remark} Let $f_{k,k+n-2},$ $f_{k,n}^{\text{ }i},$ $p_{k,n}^{\text{ }i}$, F_{k,n}(t), $ $G_{k,n}(t)$ and $R_{k,n}(t)$ be GO$k$F(\ref{mil}), $k$SO$k$F \ref{fib}), $k$SO$k$P(\ref{pell}), generalized Fibonacci polynomials(\re {fippol}), generalized Lucas polynomials(\ref{lucpol}) and generalized Perrin polynomials(\ref{perpol}) respectively. Then \newline $i)$ substituting $c_{i}=t_{i}$ for $1\leq i\leq k$ in (\ref{fib}) and (\re {fippol}), we obtain the equality \begin{equation*} f_{k,n-1}^{\text{ }1}=F_{k,n}(t), \end{equation* $ii)$ substituting $t_{1}=2$ and $t_{i}=1$ for $2\leq i\leq k$ in (\re {fippol}), we obtain the equalit \begin{equation*} p_{k,n-1}^{\text{ }1}=F_{k,n}(t), \end{equation* $iii)$ substituting $t_{1}=0$ in (\ref{lucpol}), we obtain the equality \begin{equation*} R_{k,n}(t)=G_{k,n}(t), \end{equation* $iv)$ substituting $t_{i}=1$ in (\ref{fippol}), we obtain the equality \begin{equation*} f_{k,k+n-2}=F_{k,n}(t), \end{equation* $v)$ substituting $t_{i}=1$ and $k=2$ in (\ref{lucpol}), we obtain the equalit \begin{equation*} L_{n}=G_{k,n}(t), \end{equation* $vi)$ substituting $t_{1}=0$ and $t_{i}=1$ for $2\leq i\leq k$ and $k=3$ in \ref{lucpol}), we obtain the equalit \begin{equation*} R_{n}=G_{k,n}(t). \end{equation*} \end{remark} \bigskip This Remark shows that $F_{k,n}(t)$ and $G_{k,n}(t)$ are general form of all sequences mentioned above. Therefore, any result obtained from the polynomial $F_{k,n}(t)$ and $G_{k,n}(t)$ are valid for other sequences. Many researchers studied on determinantal and permanental representations of $k$ sequences of generalized order-$k$ Fibonacci and Lucas numbers. For example, Minc [11] defined an $n\times n$ (0,1)-matrix $F(n,k),$ and showed that the permanents of $F(n,k)$ is equal to the generalized order-$k$ Fibonacci numbers. In [5] and [6] the authors defined two (0,1)-matrices and showed that the permanents of these matrices are the generalized Fibonacci and Lucas numbers. \"{O}cal [12] gave some determinantal and permanental representations of $k$-generalized Fibonacci and Lucas numbers and obtained Binet's formulas for these sequences. Y\i lmaz and Bozkurt [15] derived some relationships between Pell and Perrin sequences, and permanents and determinants of a type of Hessenberg matrices. In [13] and [14] the authors give some relation between determinant and permanent. In this paper we give some determinantal and permanental representations of Generalized Lucas Polynomials by using various Hessenberg matrices. In addition we show under what conditions that the determinants of the Hessenberg matrix becomes its permanent. \section{The determinantal representations} \bigskip An $n\times n$ matrix $A_{n}=(a_{ij})$ is called lower Hessenberg matrix if a_{ij}=0$ when $j-i>1$ i.e. \begin{equation} A_{n}=\left[ \begin{array}{ccccc} a_{11} & a_{12} & 0 & \cdots & 0 \\ a_{21} & a_{22} & a_{23} & \cdots & 0 \\ a_{31} & a_{32} & a_{33} & \cdots & 0 \\ \vdots & \vdots & \vdots & & \vdots \\ a_{n-1,1} & a_{n-1,2} & a_{n-1,3} & \cdots & a_{n-1,n} \\ a_{n,1} & a_{n,2} & a_{n,3} & \cdots & a_{n,n \end{array \right] \label{an} \end{equation} \begin{thrm} \label{cahil}$\bigskip \lbrack 1]$ $A_{n}$ be the $n\times n$ lower Hessenberg matrix for all $n\geq 1$ and define $\det (A_{0})=1,$ then \begin{equation*} \det (A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation} \det (A_{n})=a_{n,n}\det (A_{n-1})+\sum\limits_{r=1}^{n-1}((-1)^{n-r}a_{n,r}\pro \limits_{j=r}^{n-1}a_{j,j+1}\det (A_{r-1})). \label{det} \end{equation} \end{thrm} \bigskip \begin{thrm} \label{t1}Let $k\geq 2$ be an integer, $G_{k,n}(t)$ be the generalized Lucas Polynomials and $C_{k,n}=(c_{rs})$ be an $n\times n$ Hessenberg matrix, wher \begin{equation*} c_{rs}=\left\{ \begin{array}{l} i^{\left\vert r-s\right\vert }.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ }s\neq 1\text{\ and }-1\leq r-s<k\text{ }, \\ i^{\left\vert r-s\right\vert }.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}.(r-s+1)\text{ \ \ \ \ \ \ \ if \ }s=1\text{\ and }-1\leq r-s<k\text{ },\text{\ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation*} Q_{k,n}=\left[ \begin{array}{cccccc} t_{1} & it_{2} & 0 & 0 & \cdots & 0 \\ 2i & t_{1} & it_{2} & 0 & \cdots & 0 \\ 3i^{2}\frac{t_{3}}{t_{2}^{2}} & i & t_{1} & it_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ ki^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & i^{k-4}\frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & \ddots & it_{2} \\ 0 & 0 & 0 & \cdots & i & t_{1 \end{array \right] . \end{equation* The \begin{equation*} \det (C_{k,n})=G_{k,n}(t) \end{equation* where $t_{0}=1$ and $i=\sqrt{-1}.$ \end{thrm} \bigskip \begin{proof} \bigskip Proof is by mathematical induction on $n$. The result is true for n=1$ by hypothesis. Assume that it is true for all positive integers less than or equal to $m,$ that is $\det (C_{k,m})=G_{k,m+1}(t).$ Using Theorem \ref{cahil} we have \begin{eqnarray*} \det (C_{k,m+1}) &=&q_{m+1,m+1}\det (C_{k,m})+\sum\limits_{r=1}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (C_{k,r-1})\right) \\ &=&t_{1}\det (C_{k,m})+\sum\limits_{r=1}^{m-k+1}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (C_{k,r-1})\right) \\ &&+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (C_{k,r-1})\right) \\ &=&t_{1}\det (C_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}q_{m+1,r}\prod\limits_{j=r}^{m}q_{j,j+1}\det (C_{k,r-1})\right) \\ &=&t_{1}\det (C_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r \frac{t_{m-r+2}}{t_{2}^{(m-r+1)}}\prod\limits_{j=r}^{m}it_{2}\det (C_{k,r-1})\right) \\ &=&t_{1}\det (C_{k,m}) \\ &&+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r}\frac{t_{m-r+2}} t_{2}^{(m-r+1)}}.i^{m+1-r}.t_{2}^{(m-r+1)}\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}\left( (-1)^{m+1-r}.i^{m+1-r}t_{m-r+2}.i^{m+1-r}.\det (Q_{k,r-1})\right) \\ &=&t_{1}\det (Q_{k,m})+\sum\limits_{r=m-k+2}^{m}t_{m-r+2}\det (Q_{k,r-1}) \\ &=&t_{1}\det (Q_{k,m})+t_{2}\det (Q_{k,m-1})+\cdots +t_{k}\det (Q_{k,m-(k-1)}). \end{eqnarray* From the hypothesis and the definition of generalized Lucas polynomials we obtai \begin{equation*} \det (C_{k,m+1})=t_{1}G_{k,m}(t)+t_{2}G_{k,m-1}(t)+\cdots +t_{k}G_{k,m-(k-1)}(t)=G_{k,m+1}(t). \end{equation* Therefore, the result is true for all possitive integers. \end{proof} \bigskip \begin{exam} We obtain $6$-th Generalized Lucas polynomials for $k=5,$ i.e. $G_{5,6}(t),$ by using Theorem \ref{t1} \begin{eqnarray*} \det (C_{5,6}) &=&\det \left[ \begin{array}{cccccc} t_{1} & -it_{2} & 0 & 0 & 0 & 0 \\ 2i & t_{1} & -it_{2} & 0 & 0 & 0 \\ 3\frac{-t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & 0 & 0 \\ 4\frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & 0 \\ 5\frac{t_{5}}{t_{2}^{4}} & \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}} t_{2}^{2}} & i & t_{1} & -it_{2} \\ 0 & \frac{t_{5}}{t_{2}^{4}} & \frac{-it_{4}}{t_{2}^{3}} & \frac{-t_{3}} t_{2}^{2}} & i & t_{1 \end{array \right] \\ &=&6t_{1}t_{5}+6t_{2}t_{4}+12t_{1}t_{2}t_{3}+2t_{2}^{3}+3t_{3}^{2} \allowbreak t_{1}^{6}+6t_{1}^{2}t_{4}+6t_{1}^{3}t_{3}+6t_{1}^{4}t_{2}+9t_{1}^{2}t_{2}^{2}\allowbreak \\ &=&G_{5,6}(t). \end{eqnarray*} \end{exam} \bigskip \begin{thrm} \label{t2}\bigskip Let $k\geq 2$ be an integer$,$ $G_{k,n}$ be the generalized Lucas Polynomial and $B_{k,n}=(b_{ij})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} b_{ij}=\left\{ \begin{array}{l} -t_{2}\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ \ \ j=i+1, \\ \frac{t_{i-j+1}}{t_{2}^{(i-j)}}\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if\ \ }i\neq 1\text{\ \ and }0\leq i-j<k\text{,} \\ \frac{t_{i-j+1}}{t_{2}^{(i-j)}}.(i-j+1)\text{\ \ \ \ \ \ \ \ if\ \ \ }i= \text{\ \ and }0\leq i-j<k\text{,} \\ 0\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise \end{array \right. \end{equation* i.e. \begin{equation*} B_{k,n}=\left[ \begin{array}{cccccc} t_{1} & -t_{2} & 0 & 0 & \cdots & 0 \\ 2 & t_{1} & -t_{2} & 0 & \cdots & 0 \\ 3\frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ k\frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \ddots & \cdots & 0 \\ 0 & \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & \ddots & -t_{2} \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] . \end{equation* The \begin{equation*} \det (B_{k,n})=G_{k,n}(t). \end{equation* where $t_{0}=1.$ \end{thrm} \bigskip \begin{proof} \bigskip \bigskip Proof is similar to the proof of Theorem \ref{t1} using Theorem \ref{cahil}. \end{proof} \begin{exam} \bigskip We obtain $5$-th generalized Lucas polynomial for $k=4$, i.e. G_{4,5}(t),$ by using Theorem \ref{t2} \begin{eqnarray*} B_{4,5} &=&\left[ \begin{array}{ccccc} t_{1} & -t_{2} & 0 & 0 & 0 \\ 2 & t_{1} & -t_{2} & 0 & 0 \\ 3\frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} & 0 \\ 4\frac{t_{4}}{t_{2}^{3}} & \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & -t_{2} \\ 0 & \frac{t_{4}}{t_{2}^{3}} & \frac{t_{3}}{t_{2}^{2}} & 1 & t_{1 \end{array \right] \\ &=&5t_{1}t_{4}+5t_{2}t_{3}+t_{1}^{5}+5t_{1}t_{2}^{2}+5t_{1}^{2}t_{3} \allowbreak 5t_{1}^{3}t_{2} \\ &=&G_{4,5}(t). \end{eqnarray*} \end{exam} \begin{cor} \bigskip \bigskip\ If we rewrite Theorem \ref{t1} and Theorem \ref{t2} for t_{i}=1$ and $k=2,$ we obtai \begin{equation*} \det (C_{k,n})=L_{n}^{\text{ }} \end{equation* an \begin{equation*} \det (B_{k,n})=L_{n}^{\text{ }} \end{equation* respectively, where $L_{n}^{\text{ }}$ are the ordinary Lucas numbers. \end{cor} \bigskip \begin{cor} \bigskip \bigskip If we rewrite Theorem \ref{t1} and Theorem \ref{t2} for t_{1}=0$ for $1\leq i\leq k,$ we obtai \begin{equation*} \det (C_{k,n})=R_{k,n}(t) \end{equation* an \begin{equation*} \det (B_{k,n})=R_{k,n}(t) \end{equation* respectively, where $R_{k,n}(t)$ are the generalized Perrin polynomials. \end{cor} \bigskip \begin{cor} If we rewrite Theorem \ref{t1} and Theorem \ref{t2} for $t_{1}=0$ and t_{i}=1$ for $2\leq i\leq k$ and $k=3$ we obtai \begin{equation*} \det (C_{k,n})=R_{n} \end{equation* an \begin{equation*} \det (B_{k,n})=R_{n} \end{equation* respectively, where $R_{n}$ are the ordinary Perrin numbers. \end{cor} \bigskip \section{The permanent representations} \bigskip Let $A=(a_{i,j})$ be an $n\times n$ square matrix over a ring R. The permanent of $A$ is defined b \begin{equation*} \text{per}(A)=\sum\limits_{\sigma \in S_{n}}\prod\limits_{i=1}^{n}a_{i,\sigma (i)} \end{equation* where $S_{n}$ denotes the symmetric group on $n$ letters. \bigskip \begin{thrm} \label{ocal}$\left[ 10\right] $Let $A_{n}$ be an $n\times n$ lower Hessenberg matrix for all $n\geq 1$ and define per$(A_{0})=1.$ Then \begin{equation*} \text{per}(A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation} \text{per}(A_{n})=a_{n,n}\text{per}(A_{n-1})+\sum\limits_{r=1}^{n-1}(a_{n,r \prod\limits_{j=r}^{n-1}a_{j,j+1}\text{per}(A_{r-1})). \label{per} \end{equation} \end{thrm} \bigskip \begin{thrm} \label{t3}\bigskip \bigskip Let $k\geq 2$ be an integer, $G_{k,n}(t)$ be the generalized Lucas Polynomials and $H_{k,n}=(h_{rs})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} h_{rs}=\left\{ \begin{array}{l} i^{(r-s)}.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if \ }s\neq 1\text{\ and}-1\leq r-s<k, \\ i^{(r-s)}.\frac{t_{r-s+1}}{t_{2}^{(r-s)}}.(r-s+1)\text{ \ \ \ \ \ \ \ if \ s=1\text{\ and }-1\leq r-s<k\text{ },\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation} H_{k,n}=\left[ \begin{array}{cccccc} t_{1} & -it_{2} & 0 & 0 & \cdots & 0 \\ 2i & t_{1} & -it_{2} & 0 & \cdots & 0 \\ 3i^{2}\frac{t_{3}}{t_{2}^{2}} & i & t_{1} & -it_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ ki^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & i^{k-4}\frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & i^{k-1}\frac{t_{k}}{t_{2}^{k-1}} & i^{k-2}\frac{t_{k-1}}{t_{2}^{k-2}} & i^{k-3}\frac{t_{k-2}}{t_{2}^{k-3}} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] . \end{equation The \begin{equation*} \text{per}(H_{k,n})=G_{k,n}(t) \end{equation* where $t_{0}=1$ and $i=\sqrt{-1}.$ \end{thrm} \bigskip \begin{proof} \bigskip Proof is similar to the proof of Theorem \ref{t1} by using Theorem \ref{ocal}. \end{proof} \bigskip \begin{exam} \bigskip We obtain $3$-th Generalized Lucas polynomials for $k=4$, by using Theorem \ref{t3 \begin{eqnarray*} \det (H_{4,3}) &=&\det \left[ \begin{array}{ccc} t_{1} & -it_{2} & 0 \\ 2i & t_{1} & -it_{2} \\ 3\frac{-t_{3}}{t_{2}} & i & t_{1 \end{array \right] \\ &=&3t_{3}+3t_{1}t_{2}+t_{1}^{3}. \end{eqnarray*} \end{exam} \bigskip \begin{thrm} \label{t4}Let $k\geq 2$ be an integer$,$ $G_{k,n}(t)$ be the generalized Lucas Polynomials and $L_{k,n}=(l_{ij})$ be an $n\times n$ lower Hessenberg matrix such tha \begin{equation*} l_{ij}=\left\{ \begin{array}{l} \begin{array}{l} \frac{t_{i-j+1}}{t_{2}^{(i-j)}}\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if\ \ }j\neq 1\text{\ \ and }0\leq i-j<k\text{,} \\ \frac{t_{i-j+1}}{t_{2}^{(i-j)}}.(i-j+1)\text{\ \ \ \ \ \ \ \ if\ \ \ }j= \text{\ \ and }0\leq i-j<k\text{, \end{array} \\ 0\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ otherwise\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation*} L_{k,n}=\left[ \begin{array}{cccccc} t_{1} & t_{2} & 0 & 0 & \cdots & 0 \\ 2 & t_{1} & t_{2} & 0 & \cdots & 0 \\ 3\frac{t_{3}}{t_{2}^{2}} & 1 & t_{1} & t_{2} & \cdots & 0 \\ \vdots & \vdots & \vdots & \vdots & & \vdots \\ k\frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \frac{t_{k-3}}{t_{2}^{k-4}} & \cdots & 0 \\ 0 & \frac{t_{k}}{t_{2}^{k-1}} & \frac{t_{k-1}}{t_{2}^{k-2}} & \frac{t_{k-2}} t_{2}^{k-3}} & \cdots & 0 \\ & \vdots & \vdots & \vdots & \ddots & \\ 0 & 0 & 0 & \cdots & \cdots & t_{1 \end{array \right] \end{equation* where $t_{0}=1.$ The \begin{equation*} \text{per}(L_{k,n})=G_{k,n}(t). \end{equation*} \end{thrm} \bigskip \begin{proof} \bigskip Proof of the theorem is similar to the proof of Theorem \ref{t1} using Theorem \ref{ocal}. \end{proof} \bigskip \begin{cor} If we rewrite Theorem \ref{t3} and Theorem \ref{t4} for $t_{i}=1$ $(1\leq i\leq k)$ and $k=2$ $,$ we obtai \begin{equation*} \text{per}(H_{k,n})=L_{n} \end{equation* an \begin{equation*} \text{per}(L_{k,n})=L_{n} \end{equation* respectively, where $L_{n}$ are the ordinary Lucas numbers. \end{cor} \bigskip \begin{cor} \bigskip If we rewrite Theorem \ref{t3} and Theorem \ref{t4} for $t_{1}=0$ and $t_{i}=1$ $(2\leq i\leq k),$ we obtain \begin{equation*} \text{per}(H_{k,n})=R_{k,n}(t) \end{equation* an \begin{equation*} \text{per}(L_{k,n})=R_{k,n}(t) \end{equation* respectively, where $R_{k,n}(t)$ are the generalized Perrin polynomials. \end{cor} \bigskip \begin{cor} \bigskip If we rewrite Theorem \ref{t3} and Theorem \ref{t4} for $t_{1}=0$, t_{i}=1$ $(2\leq i\leq k)$ and $k=3$ we obtain \begin{equation*} \text{per}(H_{k,n})=R_{n} \end{equation* an \begin{equation*} \text{per}(L_{k,n})=R_{n} \end{equation* respectively, where $R_{n}$ are the ordinary Perrin numbers. \end{cor} \subsection{Determinat and Permanent of a Hessenberg Matrix} In this section we give a relation between the determinant and the permanent of a Hessenberg matrix. \begin{thrm} Let $A_{n}$\bigskip\ be the Hessenberg matrix in (\ref{an}) and B_{n}=(b_{ij})$ be an $n\times n$ Hessenberg matrix such tha \begin{equation*} b_{ij}=\left\{ \begin{array}{l} 0\text{\ \ \ \ \ \ \ \ \ \ \ \ if \ }j-i>1, \\ -a_{ij}\text{ \ \ \ \ \ \ \ if \ \ }j-i=1\text{ },\text{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } \\ a_{ij}\text{ \ \ \ \ \ \ \ \ \ otherwise.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \end{array \right. \end{equation* i.e. \begin{equation*} B_{n}=\left[ \begin{array}{ccccc} a_{11} & -a_{12} & 0 & \cdots & 0 \\ a_{21} & a_{22} & -a_{23} & \cdots & 0 \\ a_{31} & a_{32} & a_{33} & \cdots & 0 \\ \vdots & \vdots & \vdots & & \vdots \\ a_{n-1,1} & a_{n-1,2} & a_{n-1,3} & \cdots & -a_{n-1,n} \\ a_{n,1} & a_{n,2} & a_{n,3} & \cdots & a_{n,n \end{array \right]. \end{equation* The \begin{equation*} \det B_{n}=\text{per}A_{n} \end{equation* o \begin{equation*} \det A_{n}=\text{per}B_{n}. \end{equation*} \end{thrm} \bigskip \begin{proof} \bigskip We know from (\ref{det}) \begin{equation*} \det (A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation*} \det (A_{n})=a_{n,n}\det (A_{n-1})+\sum\limits_{r=1}^{n-1}((-1)^{n-r}a_{n,r}\pro \limits_{j=r}^{n-1}a_{j,j+1}\det (A_{r-1})) \end{equation* and from (\ref{per}) \begin{equation*} \text{per}(A_{1})=a_{11} \end{equation* and for $n\geq 2 \begin{equation*} \text{per}(A_{n})=a_{n,n}\text{per}(A_{n-1})+\sum\limits_{r=1}^{n-1}(a_{n,r \prod\limits_{j=r}^{n-1}a_{j,j+1}\text{per}(A_{r-1})). \end{equation* Using mathematical induction on $n$, we prove this theorem by using (\ref{det}) and (\ref{per}). The result is true for $n=1$ by hypothesis. Assume that it is true for all positive integers less than or equal to $m,$ namely $\det B_{m}=$per$A_{m}.$ For $n\geq 2$ \begin{eqnarray*} \det (B_{m+1}) &=&a_{m+1,m+1}\det (B_{m})+\sum\limits_{r=1}^{m}((-1)^{m+1-r}a_{m+1,r}\pro \limits_{j=r}^{m}b_{j,j+1}\det (B_{r-1})) \\ &=&a_{m+1,m+1}\text{per}(A_{m})+\sum\limits_{r=1}^{m}((-1)^{m+1-r}a_{m+1,r \prod\limits_{j=r}^{m}(-a_{j,j+1})\text{per}(A_{r-1})) \\ &=&a_{m+1,m+1}\text{per}(A_{m})+\su \limits_{r=1}^{m}((-1)^{m+1-r}a_{m+1,r}(-1)^{m+1-r}\pro \limits_{j=r}^{m}a_{j,j+1}\text{per}(A_{r-1})) \\ &=&a_{m+1,m+1}\text{per}(A_{m})+\sum\limits_{r=1}^{m}(a_{m+1,r}\pro \limits_{j=r}^{m+1}a_{j,j+1}\text{per}(A_{r-1})) \\ &=&\text{per}(A_{m+1}). \end{eqnarray* Therefore, the result is true for all possitive integers. \end{proof} \bigskip
\section{Introduction}\label{sect1} Sequence E variables are ellipsoidal binaries which follow a loose period-luminosity relation \citep{1999IAUS..191..151W,2004AcA....54..347S}. In the LMC, they make up 0.5 to 2\% of luminous red giant stars. Most of the sequence E stars lie on the first giant branch in the case of low mass stars \citep{1999IAUS..191..151W}, and some of them also evolve to the AGB \citep{2004AcA....54..347S,2007ApJ...660.1486S}. While all the sequence E stars show ellipsoidal variations, about 7\% of them show eclipses in addition to the ellipsoidal light variations \citep{2004AcA....54..347S}. In the ellipsoidal binary systems, the primary star is a red giant substantially filling its Roche lobe, and the secondary star is an unseen companion orbiting the primary star. Due to the tidal interaction, the primary star is distorted by the companion, causing an ellipsoid or pear-like shape. As the system orbits around, the change in the apparent surface area gives rise to the ellipsoidal light variation. Sequence E stars are likely the precursors of close binary Planetary Nebulae (PNe). In the ellipsoidal binary systems, the red giant primary star has already substantially filled its Roche lobe. As the star evolves, it will experience a Roche lobe overflow and then faces a common envelope (CE) event. The primary star will terminate its red giant phase by ejecting the CE and it will produce a close binary PN. At the present time the most favoured theory explaining the nonspherical PNe is the binary hypothesis \citep{1990ApJ...355..568B, 1993ApJ...418..794Y, 1997ApJS..112..487S, 2000ASPC..199..115B, 2007BaltA..16...79Z, 2009PASP..121..316D}. Theoretical considerations suggest that a binary with CE will eject the entire envelope to produce an asymmetrical PN, with elliptical or bipolar shape. In fact, PNe with close binary nuclei have been confirmed by observations \citep{1987fbs..conf..221B,1992IAUS..151..517B,1994ASPC...56..179B,2000ASPC..199..115B, 2009A&A...496..813M}. However, about their fraction, little is known. Bond and collaborators found about 10-15\% of PNe with close binary central stars, while \citet{2009A&A...496..813M} found 12-21\% of Planetary Nebulae Nuclei (PNNe) are close post-CE binaries. Sequence E stars are likely to produce close binary PNe, and their fraction is known. In the LMC, \citet{1999IAUS..191..151W} found 0.5\% of the red giants on the top 1 magnitude of the RGB in the LMC are sequence E variables. Similarly, \citet{2004AcA....54..347S} and \citet{2007ApJ...660.1486S} found 1--2\% (we use 1.5\%) of red giants in the LMC show ellipsoidal variations. Moreover, for these fractions, we found the detectability limit of the full light curve amplitude is $\sim$0.05 mag for Macho Red band ($M_R$), and $\sim$0.025 mag for OGLE $I$ band. It is our aim to use the observed fraction of sequence E stars among all the red giants to estimate the fraction of PNe with close binary central stars that will be produced by a CE event. \section{Monte Carlo Simulation}\label{sect2} A Monte Carlo simulation is made to predict the fraction of close binary PNe. One million red giant binaries are initially generated by using the observed orbital elements distributions. All the binaries are evolved up to the RGB and AGB and their evolutionary fates are examined to find which stars would produce close binary PNe. The fraction of close binary PNe is estimated by using the observed fraction of sequence E stars on the top 1 magnitude of the RGB. \subsection{Generating the binary systems}\label{sect2.1} The simulation requires as input the orbital elements distributions of the binary systems. However, these are poorly known in the LMC, so we adopt the distributions from binaries in the solar vicinity. Input from LMC sources are used when available. (1) We consider the Initial Mass Function (IMF) as well as the star formation history to get the initial mass distribution of the primary star. The IMF is assumed to follow \citet{1955ApJ...121..161S}'s power law. The LMC star formation history is adopted from \citet{1992ApJ...388..400B}: a star burst begins $\sim$4 Gyrs ago and ceases $\sim$0.5 Gyrs ago, with the ratio of burst to quiescent star formation rate being 10. According to the evolutionary tracks of \citet{2000A&AS..141..371G}, a 4 Gyrs old star has a mass of $\rm{1.3~M_{\odot}}$, and a 0.5 Gyrs old star has a mass of $\rm{3.0~M_{\odot}}$. The mass range is set from $\rm{0.9~M_{\odot}}$ to $\rm{1.85~M_{\odot}}$ for RGB stars, where $\rm{0.9~M_{\odot}}$ is the mass for red giants born in the early universe of age 13.7 Gyrs \citep{2003ApJS..148..175S}, and $\rm{1.85~M_{\odot}}$ is the upper limiting mass for red giants with electron degenerate helium cores on the first giant branch. For AGB stars, a full mass range from 0.9 to $\rm{3.0~M_{\odot}}$ is allowed. (2) The mass ratio of the binary systems is drawn from the distribution of \citet{1991A&A...248..485D}, which follows a Gaussian-type relation, with its peak at 0.23. (3) The orbital period of the binary systems is drawn from the distribution of \citet{1991A&A...248..485D}, which follows a Gaussian-type relation, with its peak at 173 years. (4) We assume the eccentricity is zero. (5)The orbital separation is calculated by using the equation of the orbital motion. (6) The orbital inclination is obtained assuming a random orientation of the orbital pole. \subsection{The properties of the binary systems}\label{sect2.2} \subsubsection{Mass loss}\label{sect2.2.1} We consider mass loss from the RGB stars via a stellar wind. We use the empirical formulation by \citet{1975MSRSL...8..369R} to calculate the mass loss rate, but with the rate multiplied by a parameter $\eta$ which is set equal to 0.33 \citep{1983ARA&A..21..271I,2005A&A...441.1117L}. In order to know the amount of mass lost per magnitude of evolution up the red giant branch, we need to calculate the evolution rate. According to the evolutionary track of \citet{2000A&AS..141..371G}, the evolution rate for RGB stars is set to d$M_{\rm{bol}}$/d$t$ $\approx$~0.15 mag/Myr. Due to the mass loss from the primary red giant, the binary system loses its orbital angular momentum. The orbital evolution of the system is calculated with equation (20) in \citet{2002MNRAS.329..897H}. AGB mass loss driven by the superwind is not considered explicitly. At the AGB tip, stars will lose their mass by a superwind, causing the termination of the AGB evolution by ejecting the envelope, just as happens for single stars, producing single PNe or wide binary PNe. We treat the AGB superwind simply as terminating AGB evolution at a specific luminosity. \subsubsection{Stellar radius}\label{sect2.2.2} For an ellipsoidal binary system containing a red giant, the requirement for the detection of the ellipsoidal variability is the minimum fractional filling of the Roche lobe. In order to know the minimum filling factor of the Roche lobe, we use the light curve generator \textsc{nightfall}\footnote{http://www.hs.uni-hamburg.de/DE/Ins/Per/Wichmann/Nightfall.html} to model the observed ellipsoidal light variations of partial-Roche lobe filling systems as a function of binary parameters. The input binary parameters are typical of sequence E stars. We find the relation between the light curve amplitude ($M_R$) and Roche lobe filling factor `$f$' is: \begin{equation} \label{ramp} \Delta{M_R}=(0.221f^{4}+0.005)\times(1.44956q^{0.25}-0.44956)\times{sin^2i}~~, \end{equation} for $0.5<f<0.9,~ 0.1<q<1.5,~0 < i < \frac{\pi}{2}$. We also find the relation between $\Delta{M_R}$ and $\Delta{I}$ is well approximated by: \begin{equation} \label{iamp} \Delta{I}=0.87\times{\Delta{M_R}}~~, \end{equation} where $\Delta{I}$ is the full light curve amplitude in the $I$ band. Given the minimum amplitude for detectable light variations (0.05 mag in the $M_R$ band or 0.025 in the $I$ band), the minimum radius filling factor $f_{\rm min}$ of a red giant in a binary system with detectable ellipsoidal light variation can be determined from equations (\ref{ramp}) and (\ref{iamp}). With minimum Roche lobe filling factor $f_{\rm min}$ , the just-detectable stellar radius is express as: \begin{equation} \label{rmin} R_{\rm{min}}=R_{\rm{L}}\times{f_{\rm min}}~~, \end{equation} where $R_{\rm{L}}$ is the equivalent radius of the Roche lobe \citep{1983ApJ...268..368E}. When the Roche lobe is filled, the maximum stellar radius is: \begin{equation} \label{rmax} R_{\rm{max}}=R_{\rm{L}}~~. \end{equation} \subsubsection{Effective temperature and luminosity}\label{sect2.2.3} The effective temperature and luminosity of the primary red giants are calculated separately for O-rich and C-rich stars. To obtain these, we use the intermediate age LMC globular cluster NGC 1978 as a template for LMC red giant properties. For O-rich stars, the giant branch slope is obtained by using the data of \citet{2010MNRAS.408..522K}. With the evolutionary tracks of \citet{2000A&AS..141..371G}, we make a mass correction for the effective temperature. The zero point of the effective temperature $T_{\rm eff} (M_{\rm bol},m)$ is estimated by fitting the observed HR diagram of sequence E stars. For C-rich stars, we use the giant branch slope obtained from \citet{2010MNRAS.408..522K}. The zero point of the $T_{\rm eff} (M_{\rm bol},m)$ relation for C-rich stars is obtained by equating the temperatures of the O-rich and C-rich stars at the transition luminosity from the O-rich to C-rich stars. The transition luminosity is calculated by using the data of \citet{1990ApJ...352...96F} and \citet{1994ApJS...92..125V}, with the distant modulus of the LMC equal to 18.54 \citep{2006ApJ...642..834K}. With $L=4\pi \sigma R^{2}T_{\rm{eff}}^{4}$, $\rm{log~{\emph{L}/L_{\odot}}=-0.4{(\emph{M}_{bol}-4.75)}}$ and the $T_{\rm eff} (M_{\rm bol},m)$ from above, we derive the bolometric magnitude for O-rich and C-rich stars as a function of $m$ and $R$. If we substitute $R$ with $R_{\rm{min}}$ or $R_{\rm{max}}$, then we get the minimum or maximum bolometric magnitude of the red giants with just-detectable light variations or with Roche lobes that are just full, respectively. \subsection{Scenario for finding close binary PNe}\label{sect3} An ellipsoidal variable almost filling its Roche lobe will experience a Roche lobe overflow as the star evolves. As a result of the Roche lobe overflow, the binary system will suffer a CE event, leading to the ejection of the entire envelope. A close binary PN will be produced if the primary star is on the AGB, or a post-RGB binary system will be formed if the primary star is on the RGB\footnote{We did not consider the coalescence of the two companions.}. For RGB binaries, since the evolution is too slow after envelope ejection, the system will be unable evolve to high effective temperature quickly enough to ionize the ejected envelope before it disperses. On the other hand, post-RGB stars will evolve slowly through the instability strip where they will be seen as Population II Cepheids or RV Tauris. We note that some Population II Cepheids and RV Tauris do have luminosities below the RGB tip luminosity \citep{1998AJ....115.1921A}. \section{Results}\label{sect4} With the model described in Sect. \ref{sect2}, we calculated the fraction of PNNe that are close binaries. The fraction of sequence E stars in the LMC was used to normalize the calculation. In order to get the observed fraction of sequence E stars, we added single stars until the measured fraction is equal to the observed. \begin{figure} \begin{center} \includegraphics[angle=0,width=0.49\textwidth,height=0.50\hsize]{f1.1.eps} \includegraphics[angle=0,width=0.49\textwidth,height=0.50\hsize]{f1.2.eps} \caption{The distributions of the orbital period (left) and full velocity amplitude (right) of sequence E stars.\label{distribution} Solid line denotes distribution from the observation (left: OGLE II; right: \citet{2010MNRAS.405.1770N}), dotted line denotes distribution from the model by using Wood's frequency, short dashed line denotes distribution from the model by using Soszy\'nski's frequency.} \end{center} \end{figure} The predicted fraction of close binary PNe is $\sim$10 or 16\%, depending on whether we use Wood's (0.5\%) or Soszy\'nski's (1.5\%) frequency. It indicates that short-period post-CE binaries are a small fraction of PNe central stars. To test the simulation, we predicted the distributions of the period and velocity amplitude of sequence E stars and compared them with the observations. Fig. \ref{distribution} (left panel) is the period distribution from the model and observation (OGLE II). It shows that the simulation reproduces the observation well. The period ranges from $\sim$30 to 1000 days, with the peak located at $\sim$250 days. We also predicted the full velocity amplitude distribution (right panel in Fig. \ref{distribution}). The distribution from the model shows a good consistency with the observational results of \citet{2010MNRAS.405.1770N} although the observed numbers are small. An updated and improved version of the results presented here is currently being prepared for publication (Nie et al. 2011, in preparation). \acknowledgements The authors have been partially supported during this work by Australian Research Council Discovery Project DP1095368.
\section{Introduction} Tycho (G120+1.4, 3C 10) is the bright spherical supernova remnant (SNR) of SN 1572 \cite{kle79} with a diameter of $\approx 8^\prime$. The distance $d$ to the source is not well-established, but is in the range $d \approx 1.5$ -- 4 kpc (e.g., \citep{smi91}). The narrow $\theta_{\rm rim} \approx 4^{\prime\prime}$ spherical rim of hard X-ray emission is coincident with the outer edge of the radio emission \citep{hwa02} and is explained by synchrotron emission of relativistic electrons accelerated to TeV energies \citep{war05} by a forward shock moving with speed $v_{\rm sh} \approx 4600 \,(d/2.3\,\rm kpc) \, \, km \, s^{-1}$ \citep{Hug00}. Infrared observations from AKARI \citep{ish10} between 9 and 160 $\mu$ reveal thermal dust emission in the NE and NW shells. The soft X-rays also reveal a rich line structure implying significant thermal radiation \citep{bad06}. Gamma-ray fluxes from Tycho between $\approx 1$ and 10 TeV have been recently reported by the VERITAS collaboration \citep{acc11}. The Fermi LAT Collaboration announced a detection in the energy range from $\approx 0.4$ to 60~GeV \citep{gio11}, with energy spectral index $\alpha \simeq 1.3$. The total Fermi and VERITAS spectra are described by a single power law with $\alpha \approx 1.1\! - \! 1.2$ from $\approx 500$ MeV to 10 TeV. Nonthermal emission at lower frequencies, from radio through X-rays, is universally attributed to synchrotron radiation, which is a signiture of primary electron acceleration. In leptonic models, the $\gamma$ rays could result from electron bremsstrahlung and/or Compton-scattering processes. In hadronic models the $\gamma$ rays arise from the decay of, predominantly, $\pi^0$ mesons produced in interactions of cosmic-ray protons with gas and dust in the remnant. By modeling its broadband radiation fluxes, \citet{mc11} argue that the reported $\gamma$-ray fluxes provide strong evidence for hadronic acceleration in Tycho as the leptonic models seem to fail. A similar conclusion was derived by \citet{vbk08} on the basis of the then-existing upper limits of the GeV-TeV $\gamma$-rays. Because of the fundamental importance to the theory of SNR origin of Galactic cosmic rays, here we undertake a detailed analysis of the robustness of those claims. Implicit in the previous modeling is that the detected synchrotron and Compton emissions are both produced by the very same population of relativistic electrons. This is the working assumption of the commonly used single-zone model approach. In a SNR environment with spatially non-uniform magnetic fields a more realistic description of the nonthermal emission should allow at least two zones \citep{ato00a,ato00b}, where zone 1 represents a strong magnetic-field region in the shock vicinity where acceleration is strong, and zone 2 adjacent (further downstream ) to zone 1, with larger volume but lower magnetic field, into which the zone 1 electrons escape by convection with plasma and/or by diffusion. Because the synchrotron emissivity in zone 2 is significantly lower than in zone 1, while the target photon field in both zones is basically the same, the Compton $\gamma$-ray fluxes in the two-zone model can greatly exceed the predictions of the single-zone model. Within this framework, we show that a leptonic model alone can explain well the reported radio through $\gamma$-ray data of Tycho. The TeV $\g$ rays are made primarily by Compton emission, and bremsstrahlung produces most of the radiation at GeV energies. The two-zone model is described in Section 2, model parameters and interpretation of the Tycho's broadband spectrum are given in Section 3. As concluded in Section 4, the most important confirmation of the hadronic cosmic ray hypothesis remains the detection of the $\pi^0$-decay feature. \section{Two-Zone Model for SNRs} It is widely believed that cosmic rays are accelerated through Fermi processes by SNR shocks. Both first-order (shock) and second-order (stochastic) processes rely on differences of magnetic field in different regions over which particle diffuse. Strong magnetic fields favored for efficient acceleration can be produced in the vicinity of the strong forward shock through non-linear amplification of the magnetic turbulence/Alfven waves by the cosmic-ray ions accelerated at the shock (e.g., \citep{bl01,bell04}). The narrow $\theta_{\rm rim} \approx 4^{\prime \prime}$ spherical rim of hard X-ray emission \citep{hwa02} is evidence for such processes operating at the forward shock of Tycho. Interpretation of the rim width $h\approx 1.93\, 10^{-2} d_{\rm kpc} \, \rm pc$, at the source distance $d_{\rm kpc}\equiv d/1 \, \rm kpc$, as due to fast synchrotron cooling of $\gtrsim 10 \,\rm TeV$ electrons implies a high magnetic field in the rim, $B\gtrsim 400 \,\rm \mu G$ \citep{vbk08}. In effect, this explanation leads to a single-zone model where all the TeV $\gamma$-rays and the X-rays are assumed to be produced by the same electrons. However, the rim could also be explained as the region with strongly enhanced magnetic field behind the shock. The rim width then reflects the length scale of damping the magnetic turbulence rather than cooling of TeV electrons \citep{pyl05} (see also the discussion in \citep{rgb11}). Relativistic electrons accelerated at the rim escape from zone 1 (the rim) into the zone 2 by diffusion and/or convection with the fluid. The theory of cosmic-ray transport is highly developed, and involves at least the spectrum and anisotropic character of the turbulence \citep[e.g.,][]{gs97,mlp06}. Nevertheless, reasonably accurate general estimates can be made. For Tycho the convective escape time $\tau_{\rm c} =h/v_{\rm fl}^\prime \approx 40 \, \rm yr$, where $v_{\rm fl}^\prime\approx v_{\rm sh}/4 \approx 500 \, d_{\rm kpc} \,\rm km \, s^{-1}$ is the fluid speed in the forward shock frame. The diffusive escape time from zone 1 is $\tau_{\rm dif} \approx h^2/(2\kappa)$ where $\kappa = \lambda_{\rm sc} c/3$ is the diffusion coefficient. The mean scattering length $\lambda_{\rm sc}$ of relativistic particles is equal to the gyroradius $r_{\rm gyr}$ in the Bohm diffusion limit, which is attained if the ratio of the turbulent magnetic field fluctuations to the mean field is $\eta = |\delta B|/B \simeq 1$. A value $\eta < 1$ is probably more realistic in general. In this case $\lambda_{\rm sc}\approx r_{\rm gyr}/\eta^2$ (e.g., \citep{bell04}). The diffusive escape time of electrons with Lorenz factor $\gamma$ from the rim can be estimated as $\tau_{\rm dif} \approx 1.1 \times 10^4 \, \eta^2 \, B_{\rm \mu G} d_{\rm kpc}^2 \, \gamma^{-1} \, \rm yr$. Comparing this with the synchrotron cooling time $t_{\rm syn} \approx 2.45\times 10^{13}\, B_{\rm \mu G}^{-2} \gamma^{-1}\,$yr, one finds that accelerated electrons (of any energy) could cool in the rim before escaping downstream only if $B \geq 1.31 \,\eta^{-2/3} \rm mG$. Even in the limit of $\eta = 1$ this field is high. For any smaller magnetic field, electron escape from the acceleration region before cooling cannot be prevented. We note that this very effect explains the detection of non-thermal X-rays not only from the rim but also from a wider region interior to the rim. Weaker magnetic fields interior to the thin rim could be the result of dissipation of the magnetic turbulence in the thermal plasma downstream of the forward shock on timescales $\tau_{\rm c}$ yr. Therefore the region of the shell interior to the rim forms zone 2. This zone extends at least down to the contact discontinuity at $r_{\rm CD}/r_{\rm sh}\approx 0.927$ \citep{war05}, in which case the volume $V_2$ of zone 2 is larger by a factor $\approx 3.3$ than the zone 1 volume $V_1\approx 0.05\,V_{\rm SNR} $. In a more elaborate model, the region between the contact discontinuity and the reverse shock at $\theta \approx 183^{\prime \prime}$ would be interpreted as another zone with a different set of source parameters. However, because the prime goal of this paper is to establish the need for multi-zone modeling, which is appropriate for spatially inhomogeneous sources and can significantly relax the constraints on the synchrotron and Compton fluxes compared to the single-zone model, here we limit our calculations within the framework of two zones only. This assumes that zone 2 includes most of the shell between the thin X-ray rim (the zone 1) and the reverse shock. This is qualitatively justified as the contact discontinuity is apparently porous, with both thermal gas and relativistic particles penetrating through the contact discontinuity further down, e.g., through Rayleigh-Taylor instabilities as evident from detection of thermal X-ray filaments \citep[e.g.,][]{hwa02,bad06}. The implied volume filling factors of the two zones in the remnant are $\zeta_1=V_1/V_{\rm SNR}\approx 0.05$ and $\zeta_2=V_2/V_{\rm SNR}\lesssim 0.5$. These values can accomodate both interpretations of the X-ray stripes in Tycho \cite{2011ApJ...728L..28E} as part of either zone 1 (as further inhomogeneities in the rim \cite{2011ApJ...728L..28E}) or zone 2 (inhomogeneities in the shell). The system of coupled equations describing the energy distribution functions $N_1(E,t)$ and $N_2(E,t)$ of electrons (or protons) in a two-zone model was derived in \citep{ato00a}: \begin{equation} \frac{\partial N_1}{\partial t} = \frac{\partial (P_1\,N_1 )} {\partial E} - \frac{N_1}{\tau_{\rm c}} - \left( \frac{N_1}{V_1} -\frac{N_2}{V_2}\right)\! \frac{V_1}{\tau_{\rm dif}} +\, Q_1 \end{equation} \begin{equation} \frac{\partial N_2}{\partial t} = \frac{\partial ( P_2\,N_2 )} {\partial E} + \frac{N_1}{\tau_{\rm c}} + \left( \frac{N_1}{V_1} -\frac{N_2}{V_2}\right)\! \frac{V_1}{\tau_{\rm dif}} +\, Q_2 \end{equation} Here $P_{1} \! \equiv \! P_1 (E,t)$ and $P_2 \! \equiv \! P_2 (E,t)$ represent the energy-loss rates $(-d E/ d t)$ of particles with energy $E$ in zones 1 and 2 respectively. The second term in the right side describes convective escape of electrons from zone 1 into zone 2, and the 3$^{rd}$ term describes the diffusive escape of electrons from the zone 1 into zone 2 and vice versa proportional to the {\it difference} of the spatial densities ({\it the gradient}) of their energy distributions in the transition region between the zones. This leads to vanishing of the diffusive exchange of particles between the zones at very high energies when the energy densities in the two zones become equal. The timescale $\tau_{\rm dif}(E)$ depends primarily on the diffusion rate in zone 1 because diffusion their is much slower (see \citep{ato00a} for details). $Q_1\equiv Q_1(E,t)$ and $Q_2 \equiv Q_2(E,t)$ are the acceleration rates of particles in zones 1 and 2, respectively. The integral form of the solution for this type of equationsegions of can be used to find a self-consistent solution for the system using the numerical method of iterations \citep{ato00b}, neglecting the diffusive influx of particles from zone 2 to zone 1 in the first iteration. Calculations show that $\lesssim 50$ iterations result in a stable (converged) solution for $N_1(E,t)$ and $N_2(E,t)$. \section{Model Parameters and Radiation Fluxes for Tycho SNR} Fluxes of $\gamma$-rays produced by relativistic electron bremsstrahlung or by hadronic $pp$-interactions are proportional to the target gas density $n_{\rm p}$ (in terms of hydrogen atoms) in the source. The gas density $n_1$ in zone 1 (the rim) is proportional to the ambient gas density $n_0$ upstream of the forward shock, and is about $n_1 \approx 4 \, n_0$ for a strong shock. Modeling of the X-ray emission lines suggests \citep{bad06} the best fit ambient gas density $\rho_0=m_{\rm p} n_0 \simeq 2 \times 10^{-24} \, \rm g\, cm^{-3}$, or $n_0\approx 1.2 \,\rm cm^{-3}$ (see also \cite{hss86}). Smaller values, in the range $n_0 \simeq (0.3- 1)\,\rm cm^{-3}$ for distances $d \simeq (3-2.35) \,\rm kpc$, respectively, have been inferred from hydrodynamic modeling of the X-ray emission of forward-shocked material \citep{cas07}. The mean gas density $n_2$ in zone 2 (the shell) can be derived from the estimate of the total gas mass accumulated in zone 2, $M_{2}\! \approx \! m_p n_2 \zeta_2 \, V_{\rm SNR}$, where $\zeta_2$ is the volume filling factor. Note that in general this mass is contributed not only from $n_0$, but also from the mass of the pre-supernova wind swept up by the forward shock, and the mass $M_0$ of the SN ejecta. To calculate $M_2$ we note that the measured speed of the forward shock is significantly smaller than the initial $v_0\approx 10^4\, (E_{\rm SN.51})^{1/2} \, M_{0.\odot}^{-1/2}\, \rm km \, s^{-1}$. For a typical Type Ia SN ejecta with ejecta mass $M_{0.\odot} = (M_0/M_\odot) \approx 1 $ and kinetic energy $E_{\rm SN.51}=E_{\rm SN}/ 10^{51} \rm erg \approx 1 $ this implies up to several Solar masses residing in the shell of Tycho. Indeed, most of the initial explosion energy should still be in kinetic form as both the total irradiated energy and the energy accumulated in accelerated particles are much smaller than $E_{\rm SN}$. The total energy of the thermal electrons with temperature $k T_{\rm e} \sim 1\,\rm keV$ deduced from X-ray observations \citep{hwa02, war05, bad06} is only $E_{\rm T,e} = (3k T_{\rm e}/2) M_{2}/m_{\rm p} \approx 2.5 \times 10^{48} (M_2/M_\odot) \,\rm erg$. To contribute substantially to the total energy budget in the shell, thermal proton plasma should have a temperature of order $T_{\rm p} \gtrsim 10^{9} \,\rm K$. This is unlikely given fast heating of the electrons through Coulomb interactions with protons in such two-temperature plasma. Thus, the kinetic energy of the fluid moving with $v_{\rm f} \simeq 3v_{\rm sh}/4$ behind the forward shock in the shell should be still close to the initial energy, $E_{\rm kin} = M_{2} v_{\rm f}^2/2 \simeq 10^{51}E_{51}\, \rm erg$. For $v_{\rm sh}$ $\approx 4600 \, (d/2.3) \,\rm kpc$ \citep{Hug00} the implied gas density in zone 2 \begin{equation} n_2 \approx 1.05 \,\, \zeta_2^{-1} \, (d /3 \, {\rm kpc})^{-5} E_{51} \; \rm cm^{-3} \, . \end{equation} For $\zeta_2 = 0.5$ this leads to $n_2$ in the range from $ 5 \, \rm cm^{-3}$ for $d = 2.5\,\rm kpc$ to $ \approx 0.9 \, \rm cm^{-3} $ for $d = 3.5 \,\rm kpc$. In calculations here we use $n_2=3\,\rm cm^{-3}$ corresponding to $d\approx 2.8\,\rm kpc$. For the Compton radiation we take into account not only the CMB target photons, but also the FIR photons produced by thermal dust in the NE and NW parts of Tycho \citep{ish10}. This component can contribute up to 30\% to the total Compton flux at TeV energies. The particle injection rate $Q_1$ is in general time-dependent, reaching the maximum at the transition time to Sedov phase, and gradually declining afterwards. However, taking into account that Tycho is still in its early Sedov phase, we assume a stationary injection rate $ Q_1(E,t)\propto \, E^{-\alpha} e^{-E/E_{\rm cut}}$ with $\alpha=2.3$ and $E_{\rm cut}=36 \,\rm TeV$. Given that the reverse shock in Tycho is not a bright non-thermal X-ray source, injection of electrons in zone 2 at the reverse shock is presumably much smaller than at the forward shock. To keep the model simple, we neglect electron acceleration in zone 2, i.e. $Q_2 = 0$. \begin{figure} {\includegraphics[width=\columnwidth]{fig1.ps}} \caption[]{ The synchrotron fluxes from radio through X-rays in the two-zone model. Dashed and dot-dashed lines show the fluxes from zone 1 and zone 2, respectively, and the total flux is shown by the solid line. Calculations assume density $n_2\approx 3 \,\rm cm^{-3}$ at $d_{\rm kpc} =2.8$, and $n_1 \approx n_2$. Other model parameters are $B_1= 100 \,\rm \mu G$ and $B_2= 32 \,\rm \mu G$, $\eta=0.3$, $\alpha = 2.3$ and $E_{\rm cut}=36 \,\rm TeV$. Also shown are $\lesssim\,$GeV bremsstrahlung fluxes produced by relativistic electrons in zones 1 and 2. } \label{Fig1} \end{figure} \begin{figure} \includegraphics[width=\columnwidth]{fig2.ps} \caption{ The $\gamma$-ray fluxes produced in the two-zone model. The heavy solid line shows the total flux of leptonic origin. The total bremsstrahlung and Compton radiation fluxes are shown by dashed and solid (thin) lines, respectively. Both are predominantly produced in zone 2. For comparison, the Compton flux contribution from zone 1 is also shown (dot-dashed line). The full open dots show a possible fit to data by a hadronic model (see text). } \label{Fig2} \end{figure} Figure 1 shows the synchrotron fluxes from radio through X-rays calculated assuming $B_1=100\,\rm \mu G$ and $B_2 = 32 \,\rm \mu G$ in zone 1 and zone 2, respectively. In the radio to sub-mm wavelengths the total flux is contributed mostly by zone 2, while in non-thermal X-rays zone 1 contributes more than zone 2. This picture is in qualitative agreement with observations that the narrow rim is much brighter in X-rays than in radio. Figure 2 shows the $\gamma$-ray fluxes. The total flux of leptonic origin is dominated by bremsstrahlung up to $\sim 100\, \rm GeV$, most of which is produced in zone 2 where most of electrons accelerated in zone 1 eventually reside. The assumed mean gas density $n_2\approx 3 \,\rm cm^{-3}$ in zone 2, at $d=2.8\,\rm kpc$, implies about $3\,M_\odot$ of gas accumulated in the shell of Tycho. The gas density $n_1$ in zone 1 doesn't have any significant impact on the calculated fluxes. A density $n_1$ in the rim about the same as the average $n_2$ in the shell would imply an ambient gas density $n_0\approx n_1/4 = 0.75 \,\rm cm^{-3}$. Even lower gas densities $n_0$ cannot, however, be excluded taking into account that the mass currently accumulated in the shell may have a composition dominated not by the swept up ambient gas but by the presupernova wind. The total energy of the electrons in zones 1 and 2 are $E_{e.1}=4.4 \times 10^{47} \,\rm erg$ and $ E_{e.2}=4.3 \times 10^{48} \,\rm erg$, respectively. The thin solid line shows the total Compton flux. Because the target photon field (CMBR + FIR) penetrates easily through the entire source and thus is same in both zones, the Compton contribution from zone 1 is insignificant. The detected $\gamma$-ray fluxes can also be explained by relativistic protons. Full open dots in Figure 2 show the flux from $\pi^0$ decay $\gamma$-rays produced by protons with the power-law index $\alpha_{\rm p}= 2.3$ and the total energy $E_{p}=3\times 10^{49} \,\rm erg$. We emphasize that the {\rm essential} difference between the hadronic and leptonic fluxes occurs only at energies below $\approx 500 \,\rm MeV$. Therefore only by detecting the characteristic $\pi^0$ decay cutoff in the energy spectrum of $\gamma$-rays at $E\simeq (100-300)$ MeV will it be possible to claim that cosmic ray hadrons are the origin of the $\gamma$-rays from Tycho. \section{Summary and Discussion} Clear identification of the $\pi^0$ decay bump peaking at $E = m_{\pi^o}/2$ in a photon spectrum was proposed long ago as a promising way to establish acceleration sites of cosmic-ray protons and nuclei \citep{gs64,hay69}. The Large Area Telescope on {\it Fermi} provides a new database for tackling this question, and the Type Ia SNR Tycho presents one of the best opportunities to search for signatures of hadronic production. Any realistic treatment of the acceleration and radiation must consider at least two zones of different magnetic field strength. Electrons in the thin acceleration zone behind the shock where the field is enhanced are bright in synchrotron X-rays. Electrons in the region with the weaker field (zone 2) occupying the bulk of the shell volume will have dim synchrotron emission, and can be essentially missed in single-zone analyses. This zone can, however, provide most of the Compton and bremsstrahlung flux. This effect significantly weakens limitations inferred from a single-zone analysis of the observed synchrotron X-ray and GeV-TeV $\gamma$-ray fluxes. With a more realistic two-zone model, we find that a good fit to the Fermi and VERITAS data from the Tycho SNR can be made by a purely leptonic model, with bremsstrahlung emission making, primarily, the GeV $\gamma$ rays, and TeV photons coming mainly from Compton-scattered soft radiation fields. No hadrons are required in this model. The underlying densities and volume filling factors of the two zones are not uniquely parameterized, our solution represents only one of a range of possibilities. Figure 2 shows that hadronic models can fit the data just as well as a leptonic model. The two models are quite different only at low energies, with the hadronic model exhibiting the $\pi^0$-decay feature, shifted up to a few hundred MeV in a $\nu F_\nu$ representation (e.g. \citep{der86}). The absence of this this low energy feature would rule out a hadronic origin, while its detection would support an origin from cosmic-ray protons and ions. Establishing the flux of Tycho at low energies with the {\it Fermi}-LAT is difficult because of the reduction in the effective area and the larger point spread function below 1 GeV \citep{atw09}. Moreover, background modeling of the diffuse $\gamma$-ray emission from the Galaxy disk has to be carefully treated. Progress in low-energy analysis and longer exposure times will help improve the quality of the {\it Fermi}-LAT spectrum of Tycho. The kinematic $\pi^0$ flux reduction is, however, a generic feature of cosmic-ray hadron origin from the GeV cosmic rays that carry most of the energy. Identification of this feature in Tycho or other SNRs will finally provide experimental confirmation of the SNR origin of the Galactic cosmic rays. \vskip0.1in \noindent We thank J.\ Ballet, M.\ Laming, J.\ Finke, S.\ Funk, S.\ Razzaque, and S.\ Reynolds for discussions. AA appreciates the support and hospitality of the NRL Gamma and Cosmic Ray Astrophysics Branch during his visit when this work was initiated. The work of C.D.D.\ is supported by the Office of Naval Research.
\section{Introduction} The idea that black hole quasi-normal modes carry important information about the black hole area quantization appeared for the first time in the remarkable work of Hod \cite{Hod}. Later Kunstatter \cite{Kunstatter} used the Bohr-Sommerfeld quantization condition for the adiabatic invariants to add more support for Hod's ideas. It can be shown that for a classical system that oscillates with a characteristic frequency $\omega$ the quantity $I=\int dE/\omega$ becomes an adiabatic invariant. The Bohr-Sommerfeld quantization condition then suggests that $I$ has to be quantized in integers, as: \begin{equation}\label{B-S} I=\int \frac{dE}{\omega}=n\cdot\hbar. \end{equation} Hod's original suggestion was to take for the black hole's characteristic frequency the asymptotic real part of the quasi-normal frequencies ($\omega_{n}=\omega_{nR}+i\cdot\omega_{nI}$), hence the limit \begin{equation} \lim_{n\to\infty}\omega_{nR}. \end{equation} For the Schwarzschild black hole this suggestion leads through the first law of black hole mechanics to a remarkable result, since in Planck units it gives the horizon's area quantization in the form: \begin{equation}\label{statistical} A_{n}=4\ln{(k)}\cdot n,~~~~~k\in\mathbb{Z_{+}}. \end{equation} This form of area quantization was, (for statistical reasons), suggested before by Bekenstein and Mukhanov \cite{Bekenstein, Mukhanov}. In particular, Hod's suggestion leads to the value $k=3$ and has an interpretation also in terms of Loop Quantum Gravity \cite{Dreyer}. However over the years the Hod's original suggestion led to many difficulties, (for the details see for example \cite{Kon-review, Maggiore}). The various objections to the Hod's proposal were largely answered by Maggiore's modification of the Hod's original suggestion \cite{Maggiore}. Maggiore argued that black hole perturbations are in fact given by a collection of damped oscillators and their energy levels are given by \begin{equation} \hbar\cdot\omega_{0n}=\hbar\cdot\sqrt{\omega_{nR}^{2}+\omega_{nI}^{2}}, \end{equation} rather than by ~$\hbar\cdot\omega_{nR}$. The semiclassical information is then, again, recovered in the limit $n\to\infty$. If the real part of the quasi-normal frequencies is bounded and the imaginary part is (upper) unbounded, then we obtain the black hole's characteristic frequency as: \begin{equation}\label{infty} \lim_{n\to\infty}\Delta_{(n,n-1)} \left(\sqrt{\omega_{nR}^{2}+\omega_{nI}^{2}}\right)=\lim_{n\to\infty}\Delta_{(n,n-1)}~\omega_{nI}~~. \end{equation} This leads in Planck units to the following Schwarzschild black hole horizon's area spectrum: \begin{equation}\label{spectrum} A_{n}=8\pi\cdot n~. \end{equation} (This area spectrum was also suggested in many works before.) The spectrum \eqref{spectrum} is not of the form \eqref{statistical}, however Maggiore offered convincing arguments that \eqref{statistical} does not necessarily have to be expected in the semi-classical limit \cite{Maggiore}. Maggiore's conjecture has been later successfully applied also to the Garfinkle-Horowitz-Strominger black hole \cite{Wei} , the Reissner-Nordstr\"om (R-N) black hole \emph{in the small charge limit} \cite{Lopez-Ortega}, and in the same paper \cite{Lopez-Ortega} to the uncharged Dirac field perturbations of arbitrary non-extremal Reissner-Nordstr\"om (R-N) black hole, always leading to the spectrum \eqref{spectrum}. (For the Kerr black hole the asymptotic QNM spectrum was both, mumerically and analytically calculated in \cite{Yoshida, Keshet, Neitzke, Kao} and some attempts to derive the Kerr area spectrum were made in \cite{Vagenas, Medved}.) The area spectrum of the form \eqref{spectrum} was predicted also by the use of different methods and ideas for the non-extremal Reissner Nordstr\"om (R-N) black hole \cite{Barvinsky1, Barvinsky2}. (Arbitrary non-extremal Reissner-Nordstr\"om black hole, not only in the small charge limit.) If the results for the area spectra of the non-extremal Reissner-Nordstr\"om black hole are correct, and so is Maggiore's modification of the Hod's original conjecture, the results should be confirmed through the analysis of the asymptotic QNM frequencies of the electromagnetic-gravitational and scalar perturbations of the R-N black hole. This is the main purpose of our investigations in this paper. The basic questions are the following: Does the limit \eqref{infty} generally exist in case of quasinormal frequencies of electromagnetic-gravi\-tatio\-nal and scalar perturbations of the non-extremal Reissner-Nordstr\"om black hole? If it exists, what is its value? If it exists only for some specific values of the parameters $M, Q$, what constrains does it put on those parameters? We believe that after a short introduction the reader already understands what is the importance of these questions in terms of Maggiore's conjecture. There are at least two different suggestions about how to write the adiabatic invariant for the Reissner-Nordstr\"om black hole \cite{Lopez-Ortega, Kwon}. However, if the limit \eqref{infty} exists, its value answers the question of what is the characteristic frequency that has to be substituted to either of the suggestions for the quantity with equally spaced spectrum. This eventually leads to the derivation of the area spectrum for the non-extremal Reissner-Nordstr\"om black hole. It has been shown in the previous papers \cite{Visser, Skakala}, that (also) for the R-N black hole the following holds: If and only if the ratio of the surface gravities (of the outer and inner horizons) $\kappa_{+}/\kappa_{-}$ is a rational number, the asymptotic quasi-normal frequencies can be split into (generally) multiple equispaced families. By an equispaced family of QNM frequencies we mean a set of QNM frequencies of the form \begin{equation}\label{equispaced} \omega_{an}=\omega_{a}+ i(\hbox{gap})\cdot n,~~~ (n\in\mathbb{N}). \end{equation} Here the $\omega_{a}$ frequencies we call the \emph{base} frequencies and we define them by the condition $0<\omega_{aI}\leq(\hbox{gap})$. Furthermore, we have proven \cite{Visser, Skakala} that for each of the families the constant gap in the spacing between the frequencies is the same and it is given as $(\hbox{gap})=\kappa_{*}=|\kappa_{\pm}|n_{\pm}$. (Here $\kappa_{+}/|\kappa_{-}|=n_{-}/n_{+}$ and $n_{+},n_{-}$ are relatively prime integers. Hence $\kappa_{*}$ is the lowest common multiplier of the numbers $\kappa_{+},~|\kappa_{-}|$.) (The cited papers contain very general results, valid for large classes of formulas for the asymptotic frequencies, but particularly for the Reissner-Nordstr\"om black hole these properties were already discovered in \cite{Andersson}.) These results provide the basis for our analysis of the existence (and value) of the characteristic frequency of the R-N black hole. (The characteristic frequency understood in the sense of Maggiore's conjecture by the limit \eqref{infty}.) The analysis in this paper leads to the following (unfortunate) conclusion: the limit \eqref{infty} can exist only for strongly constrained values of the surface gravities. Particularly, assuming the validity of the standardly accepted formulas for the asymptotic frequencies of the non-extremal R-N black hole \cite{Motl, Andersson}, we can prove that the limit \eqref{infty} does not exist unless the ratio of the two surface gravities is rational and, moreover, given as a ratio of such relatively prime $n_{+}, n_{-}\in\mathbb{N}$, that their product ~$n_{+}\cdot n_{-}$~ is an even number. But even in case the rational ratio is given by the relatively prime integers whose product is an even number we cannot guarantee the existence of the limit \eqref{infty}. On the other hand, if it exists, we automatically know its value. Before we prove anything, let us intuitively demonstrate why the limit \eqref{infty} does not exist for the rational ratio of the two surface gravities, given as a ratio of two relatively prime \emph{odd} numbers. We decided to demonstrate it on the example where the surface gravities fulfil the rational ratio condition, as in this particular case it is easy to get an insight into the logic of what is happening. As we already mentioned, for the rational ratios of the surface gravities the frequencies (in general) split in multiple equispaced families. For the case ~$n_{+}\cdot n_{-}$~ being odd (non-extremal, hence $n_{+}\cdot n_{-}>1$), we can prove that if the frequencies exist, they must be split at least into two equispaced families generated by at least two base frequencies with \emph{different} imaginary parts. Let us for the sake of simplicity assume that there are only two equispaced families generated from the two base frequencies with different imaginary parts. If there existed only one equispaced family, then the limit \eqref{infty} would trivially exist and would be exactly equal to the constant $``(\hbox{gap})$''. (This is the case of the Schwarzschild black hole.) Unfortunately here the situation is more complicated: in order to investigate the existence of the limit \eqref{infty}, we have to monotonically order all the frequencies (the union of all the frequencies from the two families) with respect to their imaginary part. But one can immediately observe, that if the difference between the imaginary parts of the two base frequencies ~$\omega_{a}$~ is other than $``(\hbox{gap})/2$'', then the monotonic ordering of all the frequencies gives a structure with periodically changing gap (with the period 2), rather than an equispaced sequence of frequencies. To see it more clearly let us say that the gap between the imaginary parts of the two base frequencies is ~$\hbox{(gap)}/4$,~ (the constant ~$``\hbox{(gap)}$''~ is again the gap in the spacing between the frequencies within each of the two families). Then the gap in the spacing between the frequencies of the (with respect to the imaginary part) monotonically ordered union of the two families oscillates between ~$\hbox{(gap)}/4$~ and ~$3\hbox{(gap)}/4$.~ This means in such case the sequence ~$\Delta_{(n,n-1)}\omega_{nI}$~ oscillates with the period 2 and hence the limit \eqref{infty} does not exist. We will prove that for the surface gravities rational ratios and the product $n_{+}\cdot n_{-}$ odd: a) there are always at least two families with base frequencies having \emph{different} imaginary parts, b) the base frequencies cannot be equispaced in such way that the collection of all the frequencies forms one equispaced sequence. This means the limit \eqref{infty} cannot for the product $n_{+}\cdot n_{-}$ odd exist! \section{Asymptotic quasi-normal frequencies of the electromagnetic-\-gravita\-tional/scalar perturbations of the non-extremal R-N black hole} The metric of the Reissner-Nordstr\"om black hole can be expressed as: \begin{equation} ds^{2}=-f(r)dt^{2}+f(r)^{-1}dr^{2}+r^{2}d\Omega^{2}~, \end{equation} where in Planck units \begin{equation} f(r)=1-\frac{2M}{r}+\frac{Q^{2}}{r^{2}}~. \end{equation} The equation for the asymptotic quasi-normal frequencies for the non-extremal Reissner-Nordstr\"om black hole was originally derived by \cite{Motl} and the result was later independently confirmed by \cite{Andersson}. (It is also fully consistent with the numerical results obtained by \cite{Kokkotas}.) The equation can be for both, scalar and electromagnetic-gravitational perturbations written as \cite{Andersson}: \begin{equation}\label{AndHowls} \exp\{8\pi\omega M\}=-3-2\exp\left\{-\frac{2\pi\omega M(1-\kappa)^{2}}{\kappa}\right\}~. \end{equation} Here ~$\kappa=\sqrt{1-\frac{Q^{2}}{M^{2}}}$. This formula is not guaranteed to hold for the extremal case. (Also one has to be careful with the order of limits when recovering the Schwarzschild black hole case, the limit $Q\to 0$ within this formula does not lead to the Schwarzschild spectrum. For the discussion of this issue see \cite{Motl, Kunstatter2}.) Take the surface gravities of the two horizons ~$\kappa_{\pm}=\frac{1}{2}f'(r_{\pm})$.~ (Here $r_{+}$ is the radial coordinate of the outer horizon and $r_{-}$ is the radial coordinate of the inner, Cauchy horizon.) Then the condition for the asymptotic QNM frequencies of the R-N black hole can be written as: \begin{equation}\label{basicEQ} \exp\left(\frac{2\pi\omega}{\kappa_{+}}\right) + 3\exp\left(\frac{2\pi\omega}{|\kappa_{-}|}\right) + 2=0 ~~. \end{equation} This can be easily shown: Rewrite the equation \eqref{AndHowls}, so it becomes: \begin{equation}\label{Andersson} \exp\left\{2\pi\omega\left(4M+\frac{M(1-\kappa)^{2}}{\kappa}\right)\right\}+3\exp\left\{2\pi\omega\frac{M(1-\kappa)^{2}}{\kappa}\right\}+2=0. \end{equation} Then one can write $|\kappa_{\pm}|$ as: \begin{equation} |\kappa_{\pm}|=\pm\frac{f'(r_{\pm})}{2}=\pm\frac{M^{2}\kappa(\kappa\pm 1)}{r_{\pm}^{3}}. \end{equation} Then, since ~$r_{\pm}=M(1\pm\kappa)$,~ we can conclude the following: \begin{equation}\label{RNsurface} |\kappa_{\pm}|=\pm\frac{\kappa(\kappa\pm1)}{M(1\pm\kappa)^{3}}. \end{equation} For ~$1/|\kappa_{-}|$~ this gives \begin{equation} \frac{1}{|\kappa_{-}|}=\frac{M(1-\kappa)^{2}}{\kappa}, \end{equation} ~ and for ~$1/|\kappa_{+}|=1/\kappa_{+}$~ it gives \begin{equation} \frac{1}{\kappa_{+}}=\frac{M(1+\kappa)^{2}}{\kappa}=4M+\frac{M(1-\kappa)^{2}}{\kappa}, \end{equation} both being the expressions in the exponents in the equation \eqref{Andersson}. The quasi-normal frequencies are the solutions of the equation \eqref{basicEQ}, such that they fulfil the condition ~$\omega_{R}\geq 0$,~ where again ~$\omega=\omega_{R}+i\cdot\omega_{I}$.~ (The quasi-normal frequencies are symmetrically spaced with respect to the imaginary axis, so the frequencies with the negative real part are then obtained from such frequencies by a simple transformation ~$\omega_{R}\to -\omega_{R}$.)~ Call the solutions of \eqref{basicEQ} with ~$\omega_{R}\geq 0$~ to be the ``\emph{relevant} solutions''. From the ~$\omega_{R}\geq 0$~ condition and from the equation \eqref{basicEQ} one can immediately see that the real part of the modes is both, upper and lower bounded. This means that if the imaginary part is \emph{not} (upper) bounded necessarily holds the equation \eqref{infty}. The periodic functions in the equation \eqref{basicEQ} suggest that the $\omega_{nI}$ is (upper) unbounded. This intuition can be explicitly proven for the rational ratio of the surface gravities, continuity of $\omega_{n}$ as a function of $\kappa_{\pm}$ and numerical results suggest that such intuition should work for arbitrary $\kappa_{\pm}$. In this paper the basic assumption is that the $\omega_{nI}$ spectrum is (upper) \emph{unbounded} so the Maggiore's conjecture makes a good sense. \section{The existence theorems} Take the rational ratio of the surface gravities $\kappa_{+}/|\kappa_{-}|=n_{-}/n_{+}$, where $n_{\pm}$ are relatively prime integers. As we already mentioned in the introductory part, in such case we have \cite{Visser, Skakala} proven that quasi-normal frequencies split into (in general) multiple equispaced families with the same gap in the spacing between the frequencies. Then, if we define by $z=\exp\left(\frac{2\pi\omega}{\kappa_{*}}\right)$, ~(to remind the reader $\kappa_{*}=\kappa_{\pm}n_{\pm}$), the equation \eqref{basicEQ} turns into the following equation \begin{equation}\label{polynomial} z^{n_{+}}+3z^{n_{-}}+2=0~~. \end{equation} By the ''\emph{relevant} roots'' of the polynomial \eqref{polynomial} call such roots, that they generate the relevant solutions of \eqref{basicEQ}. Since $n_{+}>n_{-}$, (as a result of the fact that always $\kappa_{+}<|\kappa_{-}|$), the base frequencies that generate the different equispaced families of the form \eqref{equispaced} are the relevant ($\omega_{R}\geq 0\to |z|\geq 1$) roots of this $n_{+}$-degree polynomial \cite{Andersson, Visser}. Let us first prove the following theorem: \newtheorem{one}{Theorem}[section] \begin{one} Assume that $\kappa_{+}/\kappa_{-}$ is an irrational number. Then the limit \begin{equation} \lim_{n\to\infty} [\omega_{nI}-\omega_{n-1I}] \end{equation} does not exist. \end{one} \begin{proof} In \cite{Visser} we have proven that if there exists an equispaced family of quasi-normal frequencies the ratio of surface gravities is rational. Let us prove now a much stronger statement: If there exists a subset of relevant solutions of \eqref{basicEQ} of the type \begin{equation}\label{set} \omega_{n}=\omega_{nR}+i.\left[\omega_{I}+(\hbox{gap})\cdot n+\beta(n)\right],~~~~n\in\mathbb{N}, \end{equation} where $\beta(n)$ is some sequence that goes to 0 as $n\to\infty$, the ratio of the two surface gravities must be rational. Write the equation \eqref{basicEQ} as the following two equations: \begin{equation}\label{first} C_{(n)+}\cos\{\alpha_{(n)+}\}+3C_{(n)-}\cos\{\alpha_{(n)-}\}=-2 \end{equation} and \begin{equation}\label{second} C_{(n)+}\sin\{\alpha_{(n)+}\}+3C_{(n)-}\sin\{\alpha_{(n)-}\}=0 ~. \end{equation} Here ~$C_{(n)\pm}=\exp\left(\frac{2\pi\omega_{nR}}{|\kappa_{\pm}|}\right)$~ and ~$\alpha_{(n)\pm}=\frac{2\pi\omega_{nI}}{|\kappa_{\pm}|}$. If we take the squares of each side of each of the two equations and add them, then after some trivial algebra (including a basic trigonometric identity) we obtain the following equation: \begin{equation}\label{condition1} \cos\{\alpha_{(n)+}-\alpha_{(n)-}\}=\frac{4-[C_{(n)+}^{2}+9C_{(n)-}^{2}]}{6C_{(n)+}C_{(n)-}}~. \end{equation} But for the relevant $C_{(n)\pm}\geq 1$, related as $C_{(n)-}=C_{(n)+}^{R}$ (where $0<R=\kappa_{+}/|\kappa_{-}|<1$), always holds the following: There exists a positive valued function ~$A(R)$,~ ($A(R)>0$~ for ~$\forall R$), such that \begin{equation}\label{bound} F(C_{+},R)=\frac{4-(C_{+}^{2}+9C_{+}^{2R})}{6C_{+}^{R+1}}<-A(R)~. \end{equation} This statement follows from the fact that $F$ is for ~$C_{+}\geq 1$~ continuous, furthermore ~$F(1,R)=-1$, ~for ~$C_{+}\geq 1$~ ($R\in (0,1)$~) holds that ~ $F(C_{+},R)<0$,~ and also for ~$C_{+}\to\infty$~ the function $F$ behaves as ~$F(C_{+},R)\to -\infty$.~ Now consider the sequence of the type: \begin{equation}\label{set2} \omega_{n}=\omega_{nR}+i\cdot\left[\omega_{I}+(\hbox{gap})\cdot n\right],~~~~n\in\mathbb{N}. \end{equation} If the frequencies are equispaced in their imaginary part as in the case of \eqref{set2}, then there is a constant phase shift in the argument of the cosinus in the equation \eqref{condition1}. The only way to prevent the cosinus becoming after a finite number of phase shifts in its argument positive and contradicting \eqref{condition1} together with \eqref{bound} is if the phase shift in the argument of the cosinus is ~$\Delta_{(n,n-1)}[\alpha_{(n)+}-\alpha_{(n)-}]=const.=2\pi K$, ~$K\in\mathbb{Z}$. But in such case the sequence \eqref{set} is allowed to converge to a sequence of the form \eqref{set2}, only if \eqref{set2} fulfills the condition ~$\Delta_{(n,n-1)}[\alpha_{(n)+}-\alpha_{(n)-}]=2\pi.K$,~ $K\in\mathbb{Z}$. The reason for this is the following: Since cosinus is a uniformly continuous function, \eqref{set} is allowed to be a set of solutions of the equation \eqref{basicEQ} only if the following holds: \begin{eqnarray} \lim_{n\to\infty}~\Bigg(\cos\left\{2\pi\left[\frac{1}{\kappa_{+}}-\frac{1}{|\kappa_{-}|}\right].\left[\omega_{I}+(\hbox{gap}).n+ \beta(n)\right]\right\}~~~~~~~~~~~~~\nonumber\\ - ~\cos\left\{2\pi\left[\frac{1}{\kappa_{+}}-\frac{1}{|\kappa_{-}|}\right].\left[\omega_{I}+(\hbox{gap}).n\right]\right\}\Bigg)=0. \end{eqnarray} This can be fulfilled only in the case the constant phase shift ~(in the argument of the cosinus within the equation \eqref{condition1}) is in the asymptotic limit ~$2\pi K$,~ $K\in\mathbb{Z}$,~ since otherwise the sequence~~ \[\cos\left\{2\pi\left[\frac{1}{\kappa_{+}}-\frac{1}{|\kappa_{-}|}\right].\left[\omega_{I}+(\hbox{gap}).n+ \beta(n)\right]\right\}\] is, as we have seen, upper bounded by a negative number and the sequence \[\cos\left\{2\pi\left[\frac{1}{\kappa_{+}}-\frac{1}{|\kappa_{-}|}\right].\left[\omega_{I}+(\hbox{gap}).n\right]\right\}\] becomes (almost) periodically positive. So in case the asymptotic phase shift is other than $2\pi K$, there cannot be any solutions of \eqref{basicEQ} of the form given by \eqref{set}. But this is a necessary condition for the limit \eqref{limit} to exist. Hence for such case we have proven that the sequence $\Delta_{(n,n-1)}\omega_{nI}$ does not converge to any limit. Now let us explore the case in which the \emph{asymptotic} phase shift in the argument of the cosinus in the equation \eqref{condition1} is $2\pi K$, $K\in\mathbb{Z}$. In such way the cosinus in the equation \eqref{condition1} asymptotically approaches a constant value. Let us factorize the arguments in the trigonometric functions by $2\pi$ and denote the factorized arguments by $\tilde\alpha_{(n)\pm}$. Then it is quite clear that for a given $R$ there exists only finite set of $\left(\tilde\alpha_{(b)\pm},C_{(b)\pm}\right)$, such that they fulfill the equations \eqref{first}, \eqref{second} and give \emph{the} particular asymptotically approached constant value of the cosinus in the equation \eqref{condition1}. Then there necessarily exists such ~$\left(\tilde\alpha_{(\infty)\pm}, C_{(\infty)\pm}\right)$~ which is (up to an integer multiple of $2\pi$) obtained as the limit ~$n\to\infty$~ of some infinite subsequence of the sequence ~$(\alpha_{(n)\pm},C_{(n)\pm})$. Take such an infinite subsequence of the sequence ~$\left(\alpha_{(n)\pm},C_{(n)\pm}\right)$.~ Let the subsequence be labelled by $m\in\mathbb{N}, ~m=1,2,3...$. Then if $M(m)$ is an infinite monotonically growing sequence of the natural numbers the subsequence can be always written in the following form: \begin{equation} \alpha_{(m)\pm}=\tilde\alpha_{(\infty)\pm}+M(m).\left[\Delta\alpha +2\pi K_{1\pm}\right]+\gamma_{\pm}(m),~~~~~K_{1\pm}\in\mathbb{Z}. \end{equation} Here \begin{equation}\label{period} [\Delta_{(m,m-1)} M(m)].[\Delta\alpha+2\pi K_{1\pm}]=2\pi K_{2\pm}(m),~~~~K_{2\pm}(m)\in\mathbb{Z}, \end{equation} and ~$\gamma_{\pm}(m)$~ go to 0 as ~$m\to\infty$.~Furthermore ~$0\leq\Delta\alpha<2\pi$~ is a constant phase shift factorized by ~$2\pi$~ and common to $\alpha_{(n)\pm}$.~ From the equation \eqref{period} necessarily follows that: \begin{equation} \Delta\alpha+2\pi K_{1\pm}=\pi\left(2 K_{2\pm}(m)/[\Delta_{(m,m-1)} M(m)]\right)=\pi u, ~~u\in\mathbb{Q}. \end{equation} But then the following holds \begin{equation} \Delta_{(m,m-1)}\alpha_{(n)\pm}=2\pi K_{2\pm}(m)+\Delta_{(m,m-1)}\gamma_{\pm}(m). \end{equation} This also means the following: \begin{eqnarray} \Delta_{(m,m-1)}\alpha_{(m)+}=~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\nonumber \\ \Delta_{(m,m-1)}\alpha_{(m)-}+2\pi[\Delta_{(m,m-1)}M(m)]\cdot[\Delta_{\pm}K_{1\pm}]+\Delta_{\pm}[\Delta_{(m,m-1)}\gamma_{\pm}(m)].~~~~~~~~~~~\nonumber \end{eqnarray} From this follows: \begin{eqnarray}\label{ratio} \frac{\Delta_{(m,m-1)}\alpha_{(m)+}}{\Delta_{(m,m-1)}\alpha_{(m)-}}=\frac{|\kappa_{-}|}{\kappa_{+}}=~~~~~~ ~~~~~~~~~~~~~ ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\nonumber\\ 1+\frac{2\pi[\Delta_{(m,m-1)}M(m)]\cdot[\Delta_{\pm}K_{1\pm}]+\Delta_{\pm}[\Delta_{(m,m-1)}\gamma_{\pm}(m)] }{2\pi K_{2-}(m)+\Delta_{(m,m-1)}\gamma_{-}(m)}.~~~~~ \end{eqnarray} It can be easily seen that both ~$\Delta_{\pm}[\Delta_{(m,m-1)}\gamma_{\pm}(m)]$, ~and~ $\Delta_{(m,m-1)}\gamma_{-}(m)$ go to 0 as $m\to\infty$. ~From the equation \eqref{ratio} necessarily follows \begin{equation}\label{final} \frac{2\pi[\Delta_{(m,m-1)}M(m)].[\Delta_{\pm}K_{1\pm}]+\Delta_{\pm}[\Delta_{(m,m-1)}\gamma_{\pm}(m)] }{2\pi K_{2-}(m)+\Delta_{(m,m-1)}\gamma_{-}(m)}=const.~. \end{equation} The constant can be obtained by taking the limit $m\to\infty$. The limit ~$m\to\infty$~ then gives the following: \begin{equation} const.=\frac{2\Delta_{\pm}K_{1\pm}}{u}\in\mathbb{Q}. \end{equation} But then the ratio of the two surface gravities is proven to be rational. This contradicts our assumption that the surface gravity ratio is an irrational number. It means that also in the case ~$\Delta_{(n,n-1)}[\alpha_{(n)+}-\alpha_{(n)-}]=2\pi.K$, ~$K\in\mathbb{Z}$, there cannot be a set of solutions to the equations \eqref{basicEQ} of the form \eqref{set}. This is a necessary condition for the limit \eqref{limit} to exist. As a result of this fact the theorem is proven. \end{proof} \newtheorem{two}[one]{Theorem} \begin{two} Assume that $\kappa_{+}/|\kappa_{-}|=(n_{-}/n_{+})\in\mathbb{Q}$, where $n_{+},n_{-}$ are relatively prime integers. Then, if there exist relevant solutions of the equation \eqref{basicEQ}, (hence quasi-normal frequencies), the limit \begin{equation}\label{limit} \lim_{n\to\infty}\Delta_{(n,n-1)}\omega_{nI}=\lim_{n\to\infty} [\omega_{nI}-\omega_{n-1I}] \end{equation} does not exist, unless the polynomial in \eqref{polynomial} has only real relevant roots. (This also means that the limit \eqref{limit} does not exist in case the product ~$n_{+}\cdot n_{-}$~ is an odd number.) \end{two} \begin{proof} By investigating whether the limit \eqref{limit} exists, we need to order and label all the quasi-normal frequencies monotonically with respect to their imaginary part. Now consider the fact that if the frequencies exist, there are in general multiple families of them generated by multiple relevant roots. Then if there exist at least two $\omega_{a}$ base frequencies with \emph{distinct} imaginary parts, we obtain (in general) a structure with frequencies ordered with periodically changing gap in their spacing. (Two base frequencies with the same imaginary parts and different real parts can be interpreted as being in the asymptotic sense identical: asymptotically they give the same damped oscillator energy levels.) The periodicity is (in general) given by the number of different families given by the $\omega_{a}$-s with \emph{distinct} imaginary parts. If the gap in the spacing between the frequencies periodically changes, the limit \eqref{limit} does not exist, since ~$\Delta_{(n,n-1)}\omega_{nI}$~ periodically oscillates. The following statement holds: Take the case in which the quasi-normal frequencies split into $N\geq 2$ families, such that are generated from $\omega_{a}$ base frequencies with \emph{distinct} imaginary parts. Then the limit \eqref{limit} can exist, if and only if the imaginary part of the base modes $\omega_{a}$, such that they generate the different families, is equispaced with the gap in the spacing given as $(\hbox{gap})/N$. (Here $(\hbox{gap})$ is the gap in the spacing between the frequencies within each of the families.) Now take the equation \eqref{condition1} and consider that it can be, similarly to the previous case, fulfilled only if ~$\Delta_{(n,n-1)}[\alpha_{(n)+}-\alpha_{(n)-}]=2\pi K$, ~$K\in\mathbb{Z}$.~ But, considering that $(\hbox{gap})=\kappa_{*}$, this means \begin{equation} \Delta_{(n,n-1)}[\alpha_{(n)+}-\alpha_{(n)-}]=\frac{2\pi(n_{+}-n_{-})}{N}=2\pi K . \end{equation} This implies that \begin{equation}\label{condition2} n_{+}-n_{-}= K\cdot N . \end{equation} This cannot be fulfilled for example in case both $N$ and the product ~$n_{+}\cdot n_{-}$~ are even. (If the product $n_{+}\cdot n_{-}$ is even, then the difference between $n_{\pm}$ is necessarily odd.) Note that, since \eqref{polynomial} has real coefficients, then the complex conjugate of a root of \eqref{polynomial} is a root of \eqref{polynomial}. Furthermore the complex conjugate of a \emph{relevant} root is a \emph{relevant} root and the imaginary parts of the base frequencies $\omega_{a}$ given by two distinct complex conjugate roots are always distinct. This implies that in case $n_{+}\cdot n_{-}$ is even, the frequencies can form equispaced structure only if there exist real relevant roots of \eqref{polynomial}, (since otherwise is $N$ due to the complex conjugation always even). The imaginary parts of the base frequencies ~$\omega_{1}...\omega_{N}$~ are ordered as ~$0<\omega_{1I}<...<\omega_{NI}\leq(\hbox{gap})$ ~and must be given by the formula \begin{equation} \omega_{lI}=\omega_{1I}+\frac{(l-1)(\hbox{gap})}{N}~~. ~~l=1,...,N. \end{equation} But note that in fact ~$\omega_{1I}$~ and ~$-\omega_{1I}=\omega_{ NI}=\omega_{1I}+\frac{(N-1)(\hbox{gap})}{N}$~ belong to the complex conjugate roots. This means that ~$\omega_{1I}-(-\omega_{ 1I})=\frac{(\hbox{gap})}{N}$~ and hence ~$\omega_{ 1I}=\frac{(\hbox{gap})}{2N}$.~ That means the imaginary parts of base frequencies (and all the following frequencies) must be given as: \begin{equation} \omega_{nI}=\frac{(2n-1)(\hbox{gap})}{2N}=\frac{(2n-1)\kappa_{*}}{2N}~~~~n\in\mathbb{N} . \end{equation} (For $N$ odd the imaginary part of the real roots is given by the choice ~$n_{R}=(N+1)/2+N$.)~ The complex phases of the relevant roots are then given as: \begin{equation}\label{phaseeq} \phi_{n}=\frac{\pi (2n-1)}{N}~. \end{equation} The equation \eqref{condition2} implies together with \eqref{phaseeq} the following: \begin{equation} \phi_{n}(n_{+}-n_{-})=(2n-1)K\pi~. \end{equation} Plug this into the equation \eqref{basicEQ} and obtain: \begin{equation} e^{i\phi_{n}n_{+}}\left(|z_{n}|^{n_{+}}+3(-1)^{-(2n-1)K}|z_{n}|^{n_{-}}\right)+2=\nonumber\\ = e^{i\phi_{n}n_{+}}\left(|z_{n}|^{n_{+}}\pm 3|z_{n}|^{n_{-}}\right)+2=0. \end{equation} This necessarily implies that \begin{equation} e^{i\phi_{n}n_{+}}=\pm 1~~~~~\to~~~~~\phi_{n}n_{+}=K'_{n}\pi,~~~~K'_{n}\in\mathbb{Z}. \end{equation} This means that for arbitrary $n$ holds the following \begin{equation} \frac{(2n-1)n_{+}}{N}=K'_{n}~~~~~\to~~~~~~\frac{n_{+}}{N}=K'_{1}\in\mathbb{Z}. \end{equation} Then from \eqref{condition2} follows that also ~$n_{-}/N=(K'_{1}-K)\in\mathbb{Z}$.~ But then if ~$N\geq 2$,~ $n_{\pm}$~ are \emph{not} relatively prime, since $N$ is their common divisor, which contradicts their basic definition. So the only possible case is $N=1$ and this means there exist only base frequencies given by real relevant roots of \eqref{polynomial}. (As we argued before, if there exist non-real relevant roots then necessarily $N\geq 2$.) The relevant roots must be all negative, as there exist \emph{no} positive real \emph{relevant} roots of \eqref{polynomial}. This is due to the fact that for $z\geq 1$ necessarily holds the following: \begin{equation}\label{noneq} z^{n_{+}}+3z^{n_{-}}\geq 4 >2 ~~. \end{equation} Note, that the inequality \eqref{noneq} can be for $n_{+}\cdot n_{-}$ odd, and ~$|z|\geq1$, ~$z\in\mathbb{R}$~ written as: \begin{equation}\label{noneq2} |z^{n_{+}}+3z^{n_{-}}|\geq 4 >2 ~~, \end{equation} which means that for the product ~$n_{+}\cdot n_{-}$~ odd there do \emph{not} exist any real relevant roots, neither positive, nor negative. So the negative real relevant root can exist only in the case $n_{+}\cdot n_{-}$ is even. This proves all the statements of the theorem. (It can be easily shown that a negative real relevant root exists if $n_{+}$ is even and $n_{-}$ odd, since then the relevant $z=-1$ trivially solves the equation \eqref{polynomial}, giving the zero value of ~$\omega_{R}$.) \end{proof} \section{Discussion} In this work we have proven the following: If one takes the largely accepted formulas for the asymptotic QNM frequencies of the non-extremal Reissner-Nordstr\"om black hole \cite{Motl, Andersson}, (for electromagnetic-gravitational/scalar perturbations), then the limit \[\lim_{n\to\infty}\Delta_{(n,n-1)}\omega_{nI}\] for ``most'' of the values of the two horizon's surface gravities does \emph{not} exist. Only in the special case of the surface gravities ratio being rational, hence $n_{+}/n_{-}$ with $n_{\pm}$ being relatively prime integers, and the product $n_{+}\cdot n_{-}$ being an even number there is a \emph{possibility} that the sequence ~$\Delta_{(n,n-1)}\omega_{nI}$~ converges to some limit. (In such case it is very hard either to disprove that it converges, or to prove that it does converge.) On the other hand, if in such case the limit exists, we automatically know that it is equal to $\kappa_{*}=\kappa_{\pm}.n_{\pm}$. These results were certainly not expected, as Maggiore's conjecture was used with a great success in the case of Schwarzschild black hole, Dirac field perturbations of the non-extremal Reissner-Nordstr\"om black hole and also for the Reissner-Nordstr\"om black hole in a \emph{small charge limit}. Moreover the limit \eqref{limit} seems to trivially exist also in case of Kerr black hole, see e.g. \cite{Kao}. (The results for the Reissner-Nordstr\"om black hole in the small charge limit suggest, that although the sequence ~$\Delta_{(n,n-1)}\omega_{nI}$~ does not generally converge, in a small charge limit it shows increasingly ``decent'' behavior. This is trivially expected. From the equation \eqref{AndHowls} one can easily observe, that for the limit $Q\to 0$ one recovers the Schwarzschild result for the asymptotic gap in the spacing between the imaginary parts of the quasi-normal frequencies.) It is natural to expect that all the ``difficulties'' with Maggiore's conjecture discovered in this work will transfer to the case of \emph{rotating, charged} black hole (Kerr-Newman black hole). (For a very interesting paper concerning the problem of quasi-normal modes of Kerr-Newman black holes see \cite{BertiKokkotas}.) The proofs of the theorems from this paper use specific features and simplicity of the formula \eqref{basicEQ}. On the other hand there are great similarities between the behavior of the asymptotic frequencies of the non-extremal Reissner-Nordstr\"om case and some of the other spherically symmetric multi-horizon spacetimes, (like the Schwarzschild-deSitter spacetime and the Reissner-Nord\-str\"om-deSitter spacetime). (For details see \cite{Visser, Skakala, Visser2}.) The similarities are so significant, that one can expect that it is to possible to generalize (at least) most of the results obtained in this paper also to these other cases. The theorems we have proven can be viewed in two different ways, a positive and a negative way. The negative view one might take is that they might be understood as a disproof of the validity of Maggiore's conjecture (at least if the conjecture is claimed to hold (minimally) whenever we have a better idea about black hole thermodynamical variables). On the other hand let us assume that there exists an infinite number of special subcases of the case when the surface gravities ratio is rational and the product $n_{+}\cdot n_{-}$ is being even, such that in their case the limit exists. In such case one might adapt a different, positive\footnote{It is fair to mention that such a positive view would still contradict some of the results for the quasi-normal frequencies of the Dirac field perturbations of the non-extremal R-N black hole \cite{Lopez-Ortega}.} view: Maggiore's conjecture might carry some information about some quantum restrictions on the values of the two surface gravities (some information about quantization of the surface gravities). Such a view might not be completely unreasonable: If we could extend our results to the case of Schwarzschild-deSitter black hole, the restriction on the values of the surface gravities given by Maggiore's conjecture might largely coincide with the results derived in the paper \cite{Padmanabhan}. Here the authors came to the conclusion that in the case of Schwarzschild-deSitter (S-dS) spacetime there exists a concept of global temperature (hence global thermodynamical equilibrium of spacetime) if and only if the ratio of the two surface gravities is rational. So future generalizations of our results to the Schwarzschild-deSitter (S-dS) case could, through the Maggiore's conjecture, offer a much deeper understanding of the thermodynamics of S-dS spacetime and quantum black holes in general. But also there is a ``weaker'' sense in which the behavior of quasi-normal frequencies matches the results of \cite{Padmanabhan}. As we have seen the case of surface gravities rational ratios is special in the sense that it asymptotically gives, (in general), a set of black hole characteristic frequencies. (The characteristic frequencies are given by the values between which the sequence $\Delta_{(n,n-1)}\omega_{nI}$ asymptotically oscillates.) This might start to have a lot of new meaning, if one finds a new formulation for the Maggiore's conjecture, such that it works with a set of black hole's characteristic frequencies, rather than with \emph{the} black hole's characteristic frequency. In case of the R-N black hole there is another indicator for such (general) surface gravity rational ratio condition. In \cite{Barvinsky1, Barvinsky2} the authors used suitable boundary conditions to obtain an exact quantization of the spherically symmetric charged black holes. They derived the horizon's area spectrum as: \begin{equation}\label{Barv.quant} A_{BH}=8\pi[n+2p]\cdot l_{p}^{2}+4\pi\cdot l_{p}^{2}, ~~~n, p\in\mathbb{N}. \end{equation} Here $n$ determines the excitation of the black hole above extremality and $p$ is the number determining the charge of the black hole as ~$Q^{2}=\hbar\cdot p$.~ From the equation \eqref{RNsurface} one can easily conclude that the ratio of the two surface gravities of the R-N black hole is given as\footnote{Specifically here the author wants to thank one of the anonymous referees for making some very useful suggestions.}: \begin{equation}\label{discussion} \frac{\kappa_{+}}{|\kappa_{-}|}=\frac{M^{2}(1-\kappa)^{2}}{M^{2}(1+\kappa)^{2}}=\frac{r^{2}_{-}}{r^{2}_{+}}=\frac{A_{-}}{A_{+}}=\frac{r_{+}^{2}r_{-}^{2}}{r_{+}^{4}}=\frac{p^{2}}{(2n+1+p)^{2}}. \end{equation} (In the derivation of the equation \eqref{discussion} we used the fact that in Planck units ~$r_{+}\cdot r_{-}=Q^{2}$.~) The equation \eqref{discussion} means that the rational ratios of the two surface gravities might be seen as a consequence of the quantization of the horizon's area suggested by \cite{Barvinsky1, Barvinsky2}. \bigskip \acknowledgments{The author is thankful to the referees for interesting and useful comments that improved the content of the present paper. This research was supported by Fundac\~ao de Amparo a Pesquisa do Estado de S\~ao Paulo.}
\section{Introduction} Amphiphilic molecules in solution form bilayers for a wide range of molecular architectures and experimental conditions \cite{zasadzinski_rev}. These structures may form both from lipids \cite{zidovska,sorre} and from block copolymers \cite{jain_bates,discher,discher_eisenberg}, and are of great importance in a number of scientific disciplines. For example, lipid bilayers are an integral component of cells, where they form the outer membrane and also play a role in transport processes \cite{sorre}. The bilayer vesicles that form from block copolymers, on the other hand, are longer-lived and less permeable than their lipid counterparts \cite{discher} making them promising candidates as vehicles for drug delivery \cite{smart}. The encapsulation of hydrophobic molecules between the two leaflets of the bilayer has been a recurrent issue, and has been discussed in a variety of contexts, including drug carrier design, vesicle formation from inverse phase methods, and lipid distribution in cell membranes. In particular, it is often important to know whether the hydrophobic molecules form droplets or are spread more evenly throughout the bilayer center. For example, lipid cell membranes are often found to contain a second species of lipid \cite{hamilton,hamilton2,hamilton3}, in surprisingly high concentrations \cite{khandelia}. The question then arises of whether this is due to the presence of small lipid domains within the bilayer, the existence of which has also been suggested to explain the formation of the lipid droplets seen in the center of cells \cite{ohsaki,zanghellini}. Observations of such bilayer domains have indeed been reported in the biophysics literature \cite{ferretti,may}, and these structures have recently been seen in molecular dynamics simulations \cite{khandelia}. However, their probable small size and the existence of other lipid domains in the cell complicates the interpretation of experiments, and the issue is not fully resolved \cite{hakumaki}. The formation of oil droplets in amphiphilic bilayers is also a problem of current interest in microfluidics, and has been observed in recent experiments on block copolymer systems \cite{hayward}. The aim of this research \cite{hayward,shum,thiele} was to produce aqueous solutions of monodisperse vesicles from block copolymers in water-oil-water double emulsions by evaporating the oil. Here, the presence of oil droplets in the bilayers is undesirable, as it leads to unevenness in the vesicle wall. In addition, the encapsulation of hydrophobic molecules in bilayers is of importance in the delivery of drugs using block copolymer vesicles \cite{onaca,meng,li,mueller}. Although it might at first appear more natural to encapsulate a hydrophobic substance in the core of a spherical micelle, copolymer vesicles can offer certain advantages over these smaller structures. In particular, they can encapsulate both hydrophobic and hydrophilic compounds \cite{qin,zasadzinski}. Furthermore, faster release of the hydrophobic compound can be obtained from vesicles \cite{chen}. In this paper, we investigate how much information about hydrophobic droplets in bilayers can be obtained from a simple mean-field model of oil and amphiphile in solution. First, we will study the shape of the droplet and to what extent this varies with its size. We will then relate our results to capillarity arguments. Next, we will investigate the effect of the strength $\chi_\text{BO}$ of the repulsion between the hydrophobic liquid and the amphiphile tails on the droplet shape. This question is of relevance to several of the situations described above, and our results will give some guidance as to how robust the phenomenon of droplet formation is expected to be in experiments. In addition, understanding the role of the interaction strength might allow an oil to be chosen to encourage or discourage \cite{hayward} the formation of well-defined droplets. Finally, with similar objectives in mind, we will study the effect of the length of the hydrophobic molecules on the droplet shape and stability. The paper is organized as follows. In the following section, we introduce the theoretical technique to be used, self-consistent field theory. We then present and discuss our theoretical results, and give our conclusions in the final section. \section{Self-consistent field theory}\label{scft} Self-consistent field theory (SCFT) \cite{edwards} has been used with success to investigate the equilibrium structures formed in melts and blends of polymers \cite{maniadis,drolet_fredrickson,matsen_book}, and may also be used to study metastable structures, \cite{duque,katsov1} and amphiphiles in solution \cite{cavallo,schuetz}. It can be applied to a wide range of amphiphilic molecules, including simple homopolymers \cite{werner}, more complex copolymers \cite{mueller_gompper,wang} and any given mixture of these \cite{denesyuk}. SCFT requires less computational power than simulation methods such as Monte Carlo, yet often provides comparably accurate predictions of the form of individual structures \cite{cavallo,wijmans_linse,leermakers_scheutjens-shape}. Furthermore, as a coarse-grained model, with a simple description of the polymer molecules, it will allow us to capture the basic phenomenology of the system clearly. We now give a brief introduction to SCFT, and refer the reader to reviews \cite{matsen_book,fredrickson_book,schmid_scf_rev} for a fuller presentation. A complete description of our calculations for amphiphiles in solution is given in a recent publication \cite{gg}, and we present details only when our current system differs from that described there. SCFT models individual molecules as random walks in space, and so neglects fine details of their structure and packing \cite{schmid_scf_rev}. An ensemble of many such molecules is considered. The interactions between the molecules are modeled by assuming that the blend is incompressible and by introducing contact potentials between the molecules \cite{matsen_book}. The strengths of the potentials between the various species are specified by the Flory parameters $\chi_{ij}$ \cite{jones_book}. The computational difficulty of the problem is then sharply reduced by making a mean-field approximation \cite{matsen_book}; that is, by neglecting fluctuations. This approximation is quantitatively accurate when the molecules are long \cite{cavallo,fredrickson_book,matsen_book}. In addition, SCFT can provide considerable qualitative insight into systems containing smaller molecules, such as lipid bilayers \cite{katsov1,gg} and aqueous solutions of copolymer \cite{schuetz,schuetz2}. We now discuss the application of SCFT to our system of amphiphile and oil in a solvent, which we model by a mixture of block copolymer with two incompatible homopolymers that represent the oil and the solvent respectively. Although such a mixture of polymers may appear quite simple, models of this level of complexity have been used to study a wide range of lipid and copolymer systems \cite{katsov1,gbm_jcp}, and can capture broad phenomenology more clearly than more complicated theories. We take the copolymer to have a mean-squared end-to-end distance of $a^2N$, where $a$ is the monomer length and $N$ is the degree of polymerization \cite{matsen_book}. One half of the monomers in this polymer are hydrophilic (type A) and the other half are hydrophobic (type B), so that the degrees of polymerization $N_\text{A}$ and $N_\text{B}$ for the A and B blocks are equal. We choose the same value of $a^2N$ for the A homopolymer solvent as for the copolymers. Together with the values of $N_\text{A}$ and $N_\text{B}$, this ensures that the amphiphile preferentially forms flat bilayer structures \cite{gg} for the interaction strength we will consider here. The degree of polymerization $N_\text{O}\equiv \alpha N$ of the oil will be varied between $N/4$ and $2N$. In this paper, we keep the amounts of copolymer and homopolymer in the simulation box fixed; that is, we work in the canonical ensemble. This will make it easier for us to access more complex structures such as droplets. Such structures are more difficult to stabilize in ensembles where the system is able to relax by varying the amount of the various species, and can require constraints to be imposed on the density profile \cite{katsov1}. For concreteness and to introduce the appropriate notation, we note that the SCFT approximation to the free energy of our system has the form \begin{align} \lefteqn{\frac{FN}{k_\text{B}T\rho_0V}=\frac{F_\text{h}N}{k_\text{B}T\rho_0V}}\nonumber\\ & -(1/V)\int\mathrm{d}\mathbf{r}\,[\chi N_\text{AB} (\phi_\text{A}(\mathbf{r})+\phi_\text{S}(\mathbf{r})-\overline{\phi}_\text{A}-\overline{\phi}_\text{S})(\phi_\text{B}(\mathbf{r})-\overline{\phi}_\text{B}) \nonumber\\ & +\chi N_\text{AO}(\phi_\text{A}(\mathbf{r})+\phi_\text{S}(\mathbf{r})-\overline{\phi}_\text{A}-\overline{\phi}_\text{S})(\phi_\text{O}(\mathbf{r})-\overline{\phi}_\text{O}) \nonumber\\ & +\chi N_\text{BO}(\phi_\text{B}(\mathbf{r})-\overline{\phi}_\text{B})(\phi_\text{O}(\mathbf{r})-\overline{\phi}_\text{O})] \nonumber\\ & -(\overline{\phi}_\text{A}+\overline{\phi}_\text{B})\ln (Q_\text{AB}/V)-\overline{\phi}_\text{S}\ln(Q_\text{S}/V) -(\overline{\phi}_\text{O}/\alpha)\ln (Q_\text{O}/V) \label{FE} \end{align} where the $\overline{\phi}_i$ are the mean volume fractions of the various components. The $\phi_i(\mathbf{r})$ are the local volume fractions, with $i=\text{A}$ for the hydrophilic blocks, $i=\text{B}$ for the hydrophobic blocks, $i=\text{O}$ for the oil, and $i=\text{S}$ for the solvent. The first Flory parameter, $\chi_\text{AB}$, is set to $50/N$, so that sharp, well-defined bilayers form. The other Flory parameters will be varied to study the effect of the nature of the oil on the droplet shape. $V$ is the total volume of the system, $1/\rho_0$ is the volume of a monomer, and $F_\text{h}$ is the SCFT free energy of a homogeneous system containing the same components. The details of the individual polymers enter through the single-chain partition functions $Q_i$. These are calculated \cite{matsen_book} from integrals over the propagators $q$ and $q^\dagger$, which are also used to compute the polymer density profiles \cite{matsen_book,fredrickson_book}. Reflecting the fact that the molecules are modeled as random walks, the propagators satisfy modified diffusion equations with a field term that describes the polymer interactions. These equations are solved using a finite difference method \cite{num_rec} with step size of $0.04\,aN^{1/2}$. We assume that the droplet is cylindrically symmetric and forms at the center of the system, and hence consider an effectively two-dimensional problem in a cylindrical calculation box. Reflecting boundary conditions are imposed at the edges of the system. The derivation of the mean-field free energy $F$ also generates a set of simultaneous equations linking the values of the fields and densities. In order to calculate the SCFT density profiles for a given set of polymer concentrations, we begin by making an initial guess for the fields $w_i(\mathbf{r})$ and solve the diffusion equations to calculate the propagators and then the densities corresponding to these fields. The new $\phi_i(\mathbf{r})$ are then substituted into the simultaneous equations to calculate new values for the $w_i$ \cite{matsen2004}. The procedure is repeated until convergence is achieved. We have checked that the algorithm converges to the same solution from different initial states and with different iteration speeds. To form the structure we wish to study, a bilayer in the $z=0$ plane encapsulating a droplet at its center, a suitable initial guess for the $w_i$ must be made. It is important to note that it is not necessary to include any detailed information about the shape of the droplet or bilayer in this ansatz. Although, for the sake of speed, we will often use the final self-consistent fields corresponding to one bilayer-droplet system as initial guesses for a subsequent calculation, the initial form of the fields can be very simple. Specifically, to encourage the formation of the bilayer, it is sufficient to start the SCFT iteration with a simple square well for the hydrophobic block field $w_\text{B}$. If $z_0$ is the approximate width of the bilayer, we set $w_\text{B}$ to a low value for $-z_0/2<z<z_0/2$, and a higher value elsewhere. The initial value of the hydrophilic block field $w_\text{A}$ can simply be set to zero, as the A and B blocks are connected and the above ansatz for $w_\text{B}$ is enough for an AB bilayer to form. The iteration for the field corresponding to the oil, $w_\text{0}$, can similarly be initiated with a square well potential. The potential is set to a low value in a cylindrical region at the center of the system ($-z_0'<z<z_0'$, $r<r_0$), with $z_0'<z_0$ and $r_0$ much smaller than the radius of the simulation box. We set the tension of a bilayer with no oil to zero, as this corresponds most closely to the experimental situation of a vesicle in solution. To find the zero-tension bilayer, we proceed as follows \cite{gbm_jcp}. First, we calculate the free-energy density of a (one-dimensional) box containing an infinite planar aggregate in solvent. The volume of the simulation box is then varied in the direction perpendicular to the bilayer surface, keeping the volume fraction of copolymer fixed (at $10\%$), until the box size with the minimum free-energy density is found. This scheme was introduced to mimic the behavior of a system of many aggregates \cite{gbm_jcp,gbm_macro}, which minimizes its free energy by varying the number of aggregates and hence the volume (`box size') occupied by each. Equivalently, this procedure allows us to prepare a bilayer under zero tension. Decreasing the box size at constant copolymer volume fraction reduces the amount of amphiphile in the system and so thins the bilayer, which corresponds to stretching it parallel to its surface. Conversely, increasing the box size thickens the bilayer, which is physically equivalent to compressing it. The bilayer found for the box size where the free energy is at a minimum is one that is neither too stretched nor too compressed and has no contributions to the free energy from polymer chains that are forced into unfavorable configurations. This calculation is then used to fix the size of the cylindrical box in the $z$-direction to $17.6\,aN^{1/2}$, so that $-Z<z<Z$, where $Z=8.8$. The radius of the box is set to $R=16\,aN^{1/2}$, to allow droplets of a wide range of sizes to be studied. \section{Results and discussion}\label{results} In this section, we study the shape and size of the droplets in detail for a single set of system parameters. Then we relate the droplet shape to the bilayer and monolayer tensions and amphiphile concentrations. Finally, we investigate how the droplet shape and stability depends on the nature and size of the oil molecules. \subsection{Droplet shape.} To begin, we calculate the density profiles for bilayer-encapsulated droplets of various sizes. We focus on a system with oil molecules that are half the size of the amphiphiles, so that $N_\text{O}=N/2$. The interaction strength between the hydrophobic B-block of the amphiphile and the oil is set to $\chi_\text{BO}=5/N$. The effect of varying this quantity will be investigated later. Given that the strength of the repulsion between the A and B blocks of the amphiphile has already been set to $\chi_\text{AB}=50/N$, we no longer have complete freedom in our choice of the final Flory parameter, $\chi_\text{AO}$. If we assume that $\chi_{ij}$ is related to the polarizabilities $\alpha_{i,j}$ of the two polymer species by $\chi_{ij}=\kappa(\alpha_i-\alpha_j)^2$, where $\kappa$ is a constant of proportionality \cite{boudenne_book}, we find that $\chi_\text{AO}$ is given in terms of the other two interaction strengths by \begin{equation} \chi_\text{AO}=\chi_\text{AB}\left(1\pm\sqrt{\frac{\chi_\text{BO}}{\chi_\text{AB}}}\right)^2 \label{flory_eqn} \end{equation} and is therefore set to $\chi_\text{AO}=23.4/N$, where we choose the negative root to give a moderate incompatibility between the oil and the solvent. Starting with the tensionless bilayer described above, we compute the density profiles of bilayer-encapsulated droplets for a range of oil volume fractions between $\overline{\phi}_\text{O}=0.01$ and $\overline{\phi}_\text{O}=0.1$. The first of these values corresponds to the smallest droplet that could be stabilized in our calculations. In Figure \ref{dropcomb_fig}, we show the density profiles for (a) the amphiphile and (b) the oil for an intermediate-sized droplet with $\overline{\phi}_\text{O}=0.04$. Figure \ref{dropcomb_fig}a clearly shows the splitting of the amphiphile bilayer into two thin monolayers to incorporate the droplet. In Figure \ref{dropcomb_fig}b, we plot the density profile of this lens-shaped droplet, and see also that a significant amount of oil remains between the two leaflets of the bilayer in the region surrounding the drop. This feature appears as horizontal gray lines on either side of the droplet in the density plot of Figure \ref{dropcomb_fig}b, and is a result of the relatively weak repulsion $\chi_\text{BO}$ between the hydrophobic block and the oil. \begin{figure} \includegraphics[angle=270,width=\linewidth]{dropcomb} \caption{\label{dropcomb_fig} Density plots of (a) amphiphile (including both hydrophilic and hydrophobic blocks) and (b) oil in a droplet-bilayer system with $\overline{\phi}_\text{O}=0.04$. Cylindrical polar coordinates are used, and dark regions indicate high volume fraction.} \end{figure} To help visualize the droplets, we show a ray-traced plot of the droplet surface (defined as the locus of points where $\phi_\text{O}(\mathbf{r})=0.5$) in Figure \ref{lathe_fit_bw_fig}. The overall lens shape of the droplet is clearly visible, as is the slight rim where the edge of the droplet meets the oil remaining in the bilayer. \begin{figure} \begin{center} \includegraphics[width=0.4\linewidth]{lathe_fit_bw} \end{center} \caption{\label{lathe_fit_bw_fig} Ray-traced plot of the surface of the droplet shown in Figure \ref{dropcomb_fig}b.} \end{figure} \begin{figure} \includegraphics[width=\linewidth]{circlefits} \caption{\label{circlefits_fig} Cuts through the droplet surface for a range of droplet sizes with $\overline{\phi}_\text{O}=0.01$, $0.02$, $0.04$, $0.06$, $0.08$, and $0.1$. Circles show the data points from our SCFT calculations. Although all points are used in the fits, only every fifth point is plotted for clarity. Solid lines show fits to the data using sections of a circle. The inset shows the contact angle $\theta$ and the surface tensions $\gamma_1$ and $\gamma_2$, for later reference.} \end{figure} \subsection{Surface concentrations and tensions.} In order to gain more detailed insight into the droplet shapes, we plot cuts through the droplet surface for a range of droplet sizes from the smallest to the largest (see Figure \ref{circlefits_fig}). If we assume that no long-range forces act in the system, that the bending rigidity of the membrane can be neglected, and that the pressure inside the droplet is constant, both the upper and lower halves of the droplet will have a spherical cap shape \cite{leger} in order to obey the Laplace law \cite{rowlinson_widom,landau_lifshitz}. The expected shape for these cuts shown in Figure \ref{circlefits_fig} is then a section of a circle, with the contact angle $\theta$ (see inset) determined by the mechanical equilibrium of the surface tensions along the contact line \cite{leger}. We indeed find that this shape gives a very good fit to the data for all droplet shapes (Figure \ref{circlefits_fig}), with a slight deviation in the rim region shown in Figure \ref{lathe_fit_bw_fig}, where the droplet spreads out slightly between the two amphiphile leaflets instead of forming a perfect cusp. In addition, the contact angle calculated from the fits is the same for all droplet sizes to the accuracy of the calculations, and is given by $\theta\approx 51^\circ$. To understand how the presence of the bilayer leads to the formation of these lens-shaped droplets, and to illustrate some other features of the density profiles plotted in Figure \ref{dropcomb_fig}, we now plot a series of cuts through the density profiles of the various species. In Figure \ref{centres_edges_fig}a, we show a cut in the $z$-direction (perpendicular to the bilayer) through the density profiles of all species at the edge of the system containing the smallest droplet studied. This plot shows the state of the bilayer as far away from the droplet as possible. In Figure \ref{centres_edges_fig}b, we show the corresponding plot for the bilayer in the system with the largest droplet. First, we note that the two plots are very similar and that the bilayer is not strongly distorted by the presence of the droplet. We will return to this point in a more quantitative fashion later on. Second, we see that a small but significant amount of oil remains in the center of the bilayer. Furthermore, the density profile of the oil shows a clear peak at the interface between the hydrophilic A and hydrophobic B blocks. This is because both $\chi_\text{AO}$ and $\chi_\text{BO}$ are smaller than $\chi_\text{AB}$, the repulsion between the two blocks of the amphiphile. A thin film of oil therefore forms in this region to protect these two strongly incompatible species from each other. \begin{figure} \includegraphics[width=\linewidth]{centres_edges} \caption{\label{centres_edges_fig} Cuts through the density profiles perpendicular to the bilayer at the edge of the system and at the center of the droplet. The hydrophilic A-blocks are shown with thick dashed lines, the hydrophobic B-blocks with thick full lines, the oil with thin full lines, and the solvent with thin dashed lines. (a) Bilayer at the edge of a system containing the smallest droplet studied ($\phi_\text{O}=0.01$). (b) Bilayer at the edge of a system containing the largest droplet ($\phi_\text{O}=0.1$). (c) Monolayer covering a droplet at the center of the $\phi_\text{O}=0.01$ system. (d) Monolayer covering a droplet a the center of the $\phi_\text{O}=0.1$ system. Note the change in $z$-axis between (c) and (d). } \end{figure} A clear contrast is seen between the density profiles of the bilayer a long way from the droplet, and those of the monolayer that covers the droplet. These latter profiles, calculated at the center of the system ($r=0$), are plotted in Figure \ref{centres_edges_fig}c (for the smallest droplet) and d (for the largest droplet). We note that, in both these cases, the maximum values of the density profiles of the A and B blocks of the amphiphile are lower than in Figure \ref{centres_edges_fig}a and b. This is because the monolayer has to be stretched and thinned to cover the droplet, lowering the surface amphiphile concentration. In addition, the difference between the profiles for the smallest and largest droplets is much more marked than in the case of the bilayer. In particular, the amphiphile concentration in the monolayer covering the largest droplet (Figure \ref{centres_edges_fig}d) is noticeably lower than in that covering the smallest droplet (Figure \ref{centres_edges_fig}c), showing that the monolayer must be further stretched to encapsulate more oil. Furthermore, the monolayer in Figure \ref{centres_edges_fig}d is more symmetric with respect to its inner and outer leaflets than that in Figure \ref{centres_edges_fig}c, and the peak values of the A- and B-block concentrations are much closer. The reason for this is that the surface of the larger droplet is flatter, and the monolayer that encloses it is quite close to that which would form at a planar oil-solvent interface. \begin{figure} \includegraphics[width=\linewidth]{centre_vs_edge} \caption{\label{centre_vs_edge_fig} (a) Surface density of the monolayer covering the droplet plotted against that of (half) the bilayer at the edge of the system. (b) Normalized free energy density of the bilayer plotted against its surface density. } \end{figure} We now present a more quantitative discussion of the amphiphile density profiles. To begin, we integrate the bilayer and monolayer density profiles (Figure \ref{centres_edges_fig}) in the $z$-direction, including both hydrophilic and hydrophobic blocks. Specifically, we calculate \begin{align} \lefteqn{\Gamma_1=\int^Z_0\,\mathrm{d}z\left(\phi_\text{A}(R,z)+\phi_\text{B}(R,z)\right)} \nonumber\\ & \Gamma_2=\int^Z_0\,\mathrm{d}z\left(\phi_\text{A}(0,z)+\phi_\text{B}(0,z)\right) \label{surface_densities} \end{align} $\Gamma_1$ is then the surface density of (half) the bilayer at the edge of the system ($r=R$), while $\Gamma_2$ is the monolayer surface density at the center of the system ($r=0$). The amount of amphiphile remaining in the bulk is low, so that its contribution to the surface densities is very small. $\Gamma_1$ and $\Gamma_2$ are calculated for all droplet sizes studied and are plotted against each other in Figure \ref{centre_vs_edge_fig}a. As would be expected from the profiles shown in Figure \ref{centres_edges_fig}, the monolayer surface density varies over a wider range than the corresponding quantity for the bilayer, as the amphiphilic molecules are spread out more and more thinly in the monolayer as the droplet size increases. Furthermore, $\Gamma_1$ and $\Gamma_2$ are linearly related for a wide range of droplet sizes, with deviations from linearity only being seen for the higher surface densities corresponding to very small droplets. To understand this behavior, we relate $\Gamma_1$ and $\Gamma_2$ to the corresponding surface energies. First, we note that the amphiphile in our system acts as a surfactant, separating the solvent from the oil in the droplet and bilayer. Then, we assume that adding amphiphile linearly reduces the surface tension from its value in the absence of amphiphile, $\gamma_0$, so that $\gamma_i=\gamma_0-\delta\Gamma_i$. Here, $\delta$ is a constant of proportionality and $i=1$ for the tensions and densities at the system edge and $2$ for those at the droplet surface. Balancing these two tensions at the rim where the droplet meets the bilayer, as shown in the inset to Figure \ref{circlefits_fig}, we find also that $\gamma_1=\gamma_2\cos\theta$. Combining this with our expressions for $\gamma_1$ and $\gamma_2$, we have \begin{equation} \Gamma_2=\frac{\Gamma_1}{\cos\theta}+\frac{\gamma_0}{\delta}\left(1-\frac{1}{\cos\theta}\right) \label{g1g2} \end{equation} From the slope of the straight line in Figure \ref{centre_vs_edge_fig}a, we find that $\cos\theta\approx 0.627$, so that $\theta\approx 51^\circ$, in excellent agreement with our independent measurement of $\theta$ from the cross-sections in Figure \ref{circlefits_fig}. This shows the validity of the force balance argument, and also confirms our use of the same proportionality constant $\delta$ in our expressions for $\gamma_1$ and $\gamma_2$. To check our linear formula for $\gamma_1$, we have also calculated, using Equation \ref{FE}, the free energy density of a flat oil-containing bilayer with the profile shown in Figure \ref{centres_edges_fig}a. Similar calculations are performed for all values of $\overline{\phi}_\text{O}$. We then plot the quantity $f=FN/k_\text{B}T\rho_0V-F_\text{h}N/k_\text{B}T\rho_0V$ (the free energy density measured with respect to that of the homogeneous solution with the same composition) as a function of the surface concentration $\Gamma_1$. As can be seen from Figure \ref{centre_vs_edge_fig}b, $f$ decreases linearly with $\Gamma_1$ for all but the very smallest droplets, confirming our simple model for the surface energy $\gamma_1$. We note that the free energy density as calculated from Equation \ref{FE} includes a contribution from the solvent region as well as from the bilayer itself. However, this is likely to have a relatively small effect on the variation of $f$, as the bulk amphiphile concentration changes very little with droplet size. Both plots show some deviation from linearity for the smallest two or three droplet sizes considered. For the plot of the two surface concentrations in Figure \ref{centre_vs_edge_fig}a, the deviation comes from the increasing relative importance of the rim around the edge of the droplet (Figure \ref{lathe_fit_bw_fig}), which means that a simple force balance argument based on a well-defined contact angle is less valid. The slight breakdown of linearity in Figure \ref{centre_vs_edge_fig}b may be due to the fact that the droplet is nearing its lower size limit and the free energy density $f$ is dropping more rapidly as the bilayer relaxes towards the flat state. \subsection{Nature and size of the oil molecules.} Having established a basic picture of droplet formation in our system, we now turn our attention to the question of how the nature and size of the oil molecules affects the shape and stability of the droplets. First, we investigate the effect of changing the Flory parameter $\chi_\text{BO}$ determining the strength of the interaction between the oil and the hydrophobic blocks of the amphiphile. This corresponds to changing the chemical nature of the oil while keeping the length of the oil molecules constant. We keep the same type of amphiphiles as used in the preceding section, so that $N$ and $\chi_\text{AB}$ are unchanged. However, as noted earlier, the three Flory parameters cannot be varied completely independently \cite{boudenne_book}, and $\chi_\text{AO}$ must be recalculated according to Equation \ref{flory_eqn} for each value of $\chi_\text{BO}$. The volume fraction of oil is set to $\overline{\phi}_\text{O}=0.04$. In Figure \ref{oil_chi_fig}, we plot the outlines of the droplets formed as $\chi_\text{BO}N$ is decreased from $5$ (the value used in our earlier calculations) to $1$ in steps of $1$. The droplet with $\chi_\text{BO}N=5$ is that with the most rounded shape and the greatest thickness in the $z$-direction. As $\chi_\text{BO}N$ is lowered to $4$, the droplet spreads outward slightly into the bilayer and becomes thinner, as the oil becomes more compatible with the hydrophobic blocks of the amphiphile. The spreading effect here is rather small, suggesting that, above a certain $\chi_\text{BO}N$, the droplet shape is relatively insensitive to the nature of the oil and retains its characteristic lens form. As $\chi_\text{BO}N$ is reduced further, the droplet continues to spread. However, the amount by which the droplet thickness falls due to each reduction of $\chi_\text{BO}N$ by $1$ gradually increases, until we obtain an almost flat structure at $\chi_\text{BO}N=1$. It is interesting to note that, even for this very weak repulsion between the oil and the hydrophobic blocks, the droplet remains at least metastable. \begin{figure} \includegraphics[width=\linewidth]{oil_chi} \caption{\label{oil_chi_fig} Droplet outline for a range of values of $\chi_\text{BO}N$ from $1$ (flattest drop) to $5$ (roundest drop). The oil volume fraction $\overline{\phi}_\text{O}$ is fixed to $0.04$. } \end{figure} The results shown in Figure \ref{oil_chi_fig} differ somewhat from the classical problem of the spreading of a single droplet \cite{leger}, as our droplet is in equilibrium with a film of oil in the bilayer, which grows in thickness, taking material from the droplet, as the oil becomes less incompatible with the hydrophobic B-blocks. This is particularly clear for the lowest value of $\chi_\text{BO}N$ considered. Here, the oil concentration in the bilayer is so high that our definition of the droplet surface as the locus of points at which $\phi_\text{O}(\mathbf{r})=0.5$ now includes the oil film as an extension of the droplet. Finally, we study the effect of the size of the oil molecules on the droplet shape. Returning to our original set of Flory parameters, with $\chi_\text{BO}N=5$, we consider the following values of the oil polymerization index: $N_\text{O}=2N$, $N$, $0.5N$ (the original value), and $0.25N$, and plot the droplet outlines in Figure \ref{oil_size_fig}. As above, $\overline{\phi}_\text{O}=0.04$. For the largest three values of $N_\text{O}$, the droplet shape changes rather little. It simply shrinks slightly as $N_\text{O}$ is lowered, as this change reduces the repulsion between the oil and the hydrophobic sections of the amphiphile so that more material leaks out of the droplet into the oil film. \begin{figure} \includegraphics[width=\linewidth]{oil_size} \caption{\label{oil_size_fig} Droplet outline for a range of values of $N_\text{O}$ from $2N$ (outermost flat drop), through $N$ and $0.5N$, to $0.25N$ (round drop). } \end{figure} However, as $N_\text{O}$ is lowered further, to $0.25N$, a sharp change occurs in the droplet shape. The thickness of the droplet in the $z$-direction is now significantly greater, while its radius is smaller. It is difficult to interpret this result in terms of the simple force balance arguments used earlier. This is because the rim feature, which was previously a small perturbation on droplets whose shape could be represented by two joined spherical caps \cite{leger}, is now a much more significant part of the droplet, since the small oil molecules penetrate more effectively into the bilayer. We speculate that the elongated shape of this droplet may be a precursor to its eventual splitting into two smaller monolayer-wrapped droplets, such as those recently studied by Kusumaatmaja and Lipowsky in the context of membranes in contact with several fluids \cite{kusumaatmaja}. In any case, it certainly seems that the simple single-droplet solution to SCFT becomes unstable around $N_\text{O}=0.25N$. If we reduce $N_\text{O}$ below this value, no solution to the SCFT equations can be found using our current methods, and our algorithm slows down considerably even for $N_\text{O}=0.25N$. \section{Conclusions}\label{conclusions} Using a coarse-grained mean-field approach (self-consistent field theory) we have modeled several aspects of the structure of hydrophobic droplets encapsulated between the two leaflets of an amphilic bilayer. First, we have found that droplets of a range of sizes have the same simple lens shape that would be expected from simple capillarity arguments. We have explained both this shape and the amphiphile concentrations in different regions of the system by considering the balance of the surface tensions around the edge of the droplet. Next, we studied the effect of the oil parameters on the droplet shape. We found that, although reducing the incompatibility $\chi_\text{BO}$ causes the droplet to flatten and spread outwards into the bilayer, it remains at least metastable even for very low $\chi_\text{BO}$. There appears to be no clear threshold value of $\chi_\text{BO}$ for droplet solutions to SCFT to exist. The droplets are also relatively insensitive to changes in the oil molecule length. In fact, provided these molecules are longer than the hydrophobic part of the amphiphile, their length has little effect on the droplet shape. These observations provide some evidence that droplet formation is a relatively robust effect and may therefore be a reasonable explanation for phenomena such as the inclusion of significant amounts of a second lipid species in lipid bilayers \cite{khandelia}. Furthermore, the formation of droplets even in our simple model indicates that this might be quite a general phenomenon and that hydrophobic domains such as those seen in the molecular dynamics simulations of Khandelia et al.\ \cite{khandelia} might be observed in a variety of systems. The current work opens a number of interesting perspectives that could be discussed within the framework of self-consistent field theory. First, our free energy calculations could be extended, to find the parameter range where the oil will form a droplet rather than spreading. Second, the question of whether an optimum droplet size exists could be addressed, perhaps by using a range of system sizes or by considering the stability of a droplet with respect to two smaller droplets. We could also study how likely the droplet is to split off from the bilayer, for example by comparing the free energies of the bilayer-encapsulated droplet and a system of a droplet covered by a monolayer in coexistence with a bilayer. Finally, droplets formed from a second species of amphiphile could be studied, to bring our calculations closer to the problem of lipid domains in bilayers \cite{khandelia,hakumaki}. \section{Acknowledgements} M.J.G. gratefully acknowledges funding from the EU under an FP7 Marie Curie fellowship.
\section{\label{Intro} Introduction} In \cite{Fleischer:2010sq} we have worked out an algebraic method to present one-loop tensor integrals in terms of scalar one-loop $1$-point to $4$-point functions. The tensor integrals are defined as \begin{eqnarray} \label{definition} I_{n}^{\mu_1\cdots\mu_R} &=&~\int \frac{d^dk}{i {\pi}^{d/2}}~~\frac{\prod_{r=1}^{R} k^{\mu_r}}{\prod_{j=1}^{n}c_j}, \end{eqnarray} with denominators $c_j$, \begin{eqnarray}\label{propagators} c_j &=& (k-q_j)^2-m_j^2 +i \epsilon. \end{eqnarray} For the tensor decomposition we use Davydychev's approach~\cite{Davydychev:1991va}, recursion relations as given in \cite{Fleischer:1999hq} and make detailed use of \emph{modified Cayley determinants} introduced for this purpose in \cite{Melrose:1965kb}. For these techniques and details of definitions we ask the reader to to consult \cite{Fleischer:2010sq}. The following linear combinations of the chords have proven as particularly useful: \begin{eqnarray} Q_s^{\mu}&=&\sum_{i=1}^{5} q_i^{\mu} \frac{{s\choose i}_5}{\left( \right)_5},~~~ s=0 \cdots 5, \label{Q6} \\\label{3.11} Q_s^{t,\mu}&=&\sum_{i=1}^{5} q_i^{\mu} \frac{ {st\choose it}_5}{{t\choose t}_5},~~~ s,t=1 \cdots 5. \end{eqnarray} \section{\label{Examples} Explicit examples} According to \cite{Davydychev:1991va} we write the tensor of rank $2$ as ($[d+]^l=d+2 l, ~d= 4-2\varepsilon$) \begin{eqnarray} \label{tensor2} I_{5}^{\mu\, \nu}= \sum_{i,j=1}^{4} \, q_i^{\mu}\, q_j^{\nu} \, {\nu}_{ij} \, \, I_{5,ij}^{[d+]^2} -\frac{1}{2} \, g^{\mu \nu} \, I_{5}^{[d+]} . \end{eqnarray} Inserting \begin{eqnarray} \frac{1}{2} g^{\mu\, \nu}=\sum_{i,j=1}^{5} {\frac{{i\choose j}_5}{\left( \right)_5}} \, q_i^{\mu}\, q_j^{\nu} \label{gmunu} \end{eqnarray} and using the recursion relation \begin{eqnarray} \label{A522} {\nu}_{ij} I_{5,ij}^{[d+]^2}&=&-\frac{{0\choose j}_5}{\left( \right)_5} I_{5,i}^{[d+]} + \sum_{s=1,s \ne i}^{5} \frac{{s\choose j}_5}{\left( \right)_5} I_{4,i}^{[d+],s} + {\frac{{i\choose j}_5}{\left( \right)_5}} I_{5}^{[d+]} , \end{eqnarray} we see that the $g^{\mu \nu}$ terms in \eqref{tensor2} cancel with the result \begin{eqnarray} \label{twoteetwo} I_{5}^{\mu\, \nu}=I_{5}^{\mu} Q_0^{\nu}-\sum_{s=1}^{5} \left\{Q_0^{s,\mu}I_4^s-\sum_{t=1}^{5} Q_t^{s,\mu}I_3^{st}\right\}Q_s^{\nu}, \end{eqnarray} \begin{eqnarray} I_{5}^{\mu}=E Q_0^{\mu}-\sum_{s=1}^{5} I_4^s Q_s^{\mu},~~~~~~~~~~~~~~ E=\frac{1}{{0\choose 0}_5}\sum_{s=1}^{5} {s\choose 0}_5 I_4^s. \end{eqnarray} A more {direct approach} is the use of the $5$-point recursion in terms of $4$-point functions \cite{Diakonidis:2009fx}, \begin{eqnarray} { I_5^{\mu_1 \dots \mu_{R-1} \mu} =I_5^{\mu_1 \dots \mu_{R-1}} Q_0^{\mu} - \sum_{s=1}^{5} I_4^{\mu_1 \dots \mu_{R-1},s } Q_s^{\mu}}. \label{tensor5general} \end{eqnarray} {This formula in general takes into account some cancellations of the $g_{\mu \nu}$}. Only {$g_{\mu \nu}$-contributions from $4$-points have still to be dealt with - and they are simpler to handle}. This we demonstrate for the tensor of degree $3$. As a special case of \eqref{tensor5general} we have \begin{eqnarray}\label{3.138} I_5^{\mu \nu \lambda}=I_5^{\mu\nu} \cdot Q_0^{\lambda} -\sum_{s=1}^{5} I_4^{\mu\nu, s} \cdot Q_s^{\lambda} . \end{eqnarray} The corresponding $4$-point function reads ($q_5=0$): \begin{eqnarray} \label{TwoTy} I_4^{\mu \nu ,s}&=&\sum_{i,j=1}^{4} q_i^{\mu} q_j^{\nu} {\nu}_{ij} I_{4,ij}^{[d+]^2,s}-\frac{1}{2} g^{\mu \nu}I_{4}^{[d+],s}, \nonumber \\ \label{TwoTyTwo} {\nu}_{ij} I_{4,ij}^{[d+]^2,s}&=&-\frac{ {0s\choose js}_5}{{s\choose s}_5} I_{4,i}^{[d+],s}+ \frac{ {is\choose js}_5}{{s\choose s}_5}I_{4}^{[d+],s}+ \sum_{t=1}^{5} \frac{ {ts\choose js}_5}{{s\choose s}_5} I_{3,i}^{[d+],st} , \end{eqnarray} and with \begin{eqnarray} \frac{ {is\choose js}_5}{{s\choose s}_5}&=&\frac{ {i\choose j}_5}{{\left( \right)}_5} -\frac{ {s\choose i}_5 {s\choose j}_5 }{{\left( \right)}_5 {s\choose s}_5} \label{trick} \end{eqnarray} we again observe the possibility to cancel $g^{\mu \nu}$ with the result \begin{eqnarray} I_4^{\mu \nu,s}&=&Q_0^{s,\mu} Q_0^{s,\nu} I_4^s-\frac{{\left( \right)}_5}{{s\choose s}_5}Q_s^{\mu} Q_s^{\nu} I_4^{[d+],s} -\sum_{t=1}^5 \left\{Q_t^{s,\mu} Q_0^{s,\nu} I_3^{st}+Q_t^{s,\nu} I_3^{\mu , st} \right\}, \nonumber \\ I_3^{\mu ,st}&=& - \sum_{i=1}^4 q_i^{\mu} I_{3,i}^{[d+],st} = Q_0^{st,\mu} I_3^{st}-\sum_{u=1}^5 Q_u^{st,\mu} I_2^{stu}. \label{I4munu} \end{eqnarray} \section{\label{5to4}Contracting the tensor integrals} Scalar expressions, contracting with chords, are \begin{eqnarray} \label{defini1} q_{i_1\mu_1}\cdots q_{i_R\mu_R} ~~I_{5}^{\mu_1\cdots\mu_R}& =& ~\int\frac{ d^d k~}{i {\pi}^{d/2}}~\frac{\prod_{r=1}^{R} (q_{i_r} \cdot k)}{\prod_{j=1}^{5}c_j}, \\ \label{defini2} g_{\mu_1,\mu_2} q_{i_1\mu_3}\cdots q_{i_R\mu_R} ~~I_{5}^{\mu_1\cdots\mu_R} &=& ~\int\frac{k^2 d^d k~}{i {\pi}^{d/2}}~\frac{\prod_{r=3}^{R} (q_{i_r} \cdot k)}{\prod_{j=1}^{5}c_j}, \end{eqnarray} etc. Eqns. \eqref{defini1} and \eqref{defini2} define the contraction of all tensor indices with chords and the direct contraction of two tensor indices, respectively. These are obtained in realistic matrix element calculations by constructing projection operators or by constructing scalar differential cross sections (Born $\times$ 1-loop) before loop integration. As a result of these contractions the $1 / \left( \right)_5$ cancels already. To begin with, we have a look at \begin{eqnarray} q_{a \mu} q_{b \nu} I_5^{\mu \nu} =(q_a \cdot I_5) (q_b \cdot Q_0) -\sum_{s=1}^{5} \left\{ (q_a \cdot Q_0^s) I_4^s -\sum_{t=1}^{5} (q_a \cdot Q_t^s) I_3^{st} \right\} (q_b \cdot Q_s). \end{eqnarray} For $q_n=0,~~ a=1, \dots , n-1,~~s=1, \dots n$ \begin{eqnarray} \label{Zcalar1} (q_a \cdot Q_0) &=&\sum_{j=1}^{n-1} (q_a \cdot q_j) \frac{{0\choose j}_n} {{\left(\right)}_n}=-\frac{1}{2}\left( Y_{an}-Y_{nn} \right), \\ (q_a \cdot Q_s) &=&\sum_{j=1}^{n-1} (q_a \cdot q_j) \frac{{s\choose j}_n} {{\left( \right)}_n}=~~~\frac{1}{2} \left({\delta}_{as}-{\delta}_{ns}\right) , \label{Zcalar2} \end{eqnarray} and \begin{eqnarray} (q_a \cdot I_5)= E (q_a \cdot Q_0)-\sum_{s=1}^5 I_4^s (q_a \cdot Q_s). \end{eqnarray} Further \begin{eqnarray} \label{contrs} (q_a \cdot Q_0^s) = \frac{1}{{s\choose s}_5} \Sigma_a^{2,s}, ~~~ (q_a \cdot Q_t^s) = \frac{1}{{s\choose s}_5} \Sigma_a^{1,st}, \end{eqnarray} where the sums $\Sigma_a^{2,s}$ and $\Sigma_a^{1,st}$ are given in \cite{Fleischer:2011nt}. Both are linear combinations of ${s\choose s}_5$ and Kronecker-$\delta$'s, $\left({\delta}_{as}-{\delta}_{5s}\right) $. Indeed the fact that there is no inverse $\left( \right)_5$ anymore is due to relations \eqref{Zcalar1} and \eqref{Zcalar2}. The second scalar which can be constructed from the tensor of degree $2$ is $g_{\mu \nu} I_5^{\mu \nu}$. Due to \eqref{twoteetwo} we need to evaluate the following scalar products: \begin{eqnarray} (Q_0 \cdot Q_0)&=&\frac{1}{2}\left[\frac{{0\choose 0}_5}{\left( \right)_5}+Y_{55} \right], \nonumber \\ (Q_0 \cdot Q_s)&=&\frac{1}{2}\left[\frac{{s\choose 0}_5}{\left( \right)_5}-{\delta}_{s5} \right], \nonumber \\ (Q_0^s \cdot Q_s)&=&-\frac{1}{2}{\delta}_{s5}, \nonumber \\ (Q_t^s \cdot Q_s)&=&~~~0. \label{scalpr} \end{eqnarray} In this case the terms with ${1} / {\left( \right)_5}$ cancel and, not surprisingly, the result finally is \begin{eqnarray} \label{Nosurprise} g_{\mu \nu} I_5^{\mu \nu}=\frac{Y_{55}}{2} E + I_4^5. \end{eqnarray} To calculate $g_{\mu \nu} I_5^{\mu \nu \lambda}$ we need $g_{\mu \nu} I_4^{\mu \nu, s }$ and thus further scalar products, see \eqref{I4munu}: \begin{eqnarray} (Q_0^s \cdot Q_0^s)&=&\frac{1}{2 {s\choose s}_5}\left[{0s\choose 0s}_5+2 {s\choose 0}_5 {\delta}_{s5}\right]+ \frac{1}{2} Y_{55} , \nonumber \\ (Q_s \cdot Q_s)&=&\frac{1}{2}\frac{{s\choose s}_5}{\left( \right)_5}, \nonumber \\ (Q_t^s \cdot Q_0^s)&=&\frac{1}{2 {s\choose s}_5}\left[{ts\choose 0s}_5-{s\choose s}_5 {\delta}_{t5}+{s\choose t}_5 {\delta}_{s5}\right], \nonumber \\ (Q_t^s \cdot Q_0^{st})&=&\frac{1}{2 {s\choose s}_5}\left[~~~~~~~~~~~-{s\choose s}_5 {\delta}_{t5}+{s\choose t}_5 {\delta}_{s5}\right], \nonumber \\ (Q_t^s \cdot Q_u^{st})&=&0, \end{eqnarray} which yields \begin{eqnarray} \label{nottriv} g_{\mu \nu} I_4^{\mu \nu, s }= \frac{ Y_{55}}{2} I_4^s + I_3^{s5} + \frac{{\delta}_{s5}}{{s\choose s}_5} \left[{s\choose 0}_5 I_4^s - \sum_{t=1}^5 {s\choose t}_5 I_3^{st}\right] \end{eqnarray} and finally \begin{eqnarray} \label{twocon} g_{\mu \nu} I_5^{\mu \nu \lambda}=-\frac{Y_{55}}{2} \sum_{s=1}^5 I_4^s~ Q_s^{0,\lambda} + I_4^5 Q_0^{5,\lambda}-\sum_{t=1}^4 I_3^{5t} Q_t^{5,\lambda}. \end{eqnarray} It is remarkable that \eqref{nottriv} is trivial again for $s \ne 5$. For $s=5$, however, the standard cancelation of propagators does not work and for this case \eqref{nottriv} is indeed a useful result. For further contraction of \eqref{twocon} with a vector $q_{\lambda}$ again \eqref{contrs} can be applied. \section{\label{Avoi}Avoiding inverse $4$-point Gram determinants} While in the above approach of taking scalar products of the tensors with chords the inverse $\left( \right)_5$ Gram determinant cancels already, there still remains the inverse ${s\choose s}_5$ sub-Gram determinant of the $4$-point functions. Therefore we have to choose a different approach for the case if the latter becomes small. This approach consists in avoiding the inverse $\left( \right)_5$ already in the $5$-point tensors from the very beginning and keeping only $4$-point integrals in higher dimensions (i.e. integrals with only powers 1 of the scalar propagators), which for small ${s\choose s}_5$ should be evaluated in a different manner than by standard recursion, see \cite{Fleischer:2010sq}. If one does not want to reintroduce the inverse $\left( \right)_5$ to see the cancellation of the $g^{\mu \nu}$, one can explicitely see its cancelation also after taking contractions. For the tensor of degree $2$ we refer to \cite{Fleischer:2011nt} for the contraction with two chords. For the self-contraction of the tensor indices no simpler result than \eqref{Nosurprise} can be achieved anyway. For the tensor of degree $3$ we present new results for the contraction with three chords and a self-contraction. \vspace{0.5cm} The tensor can be written as follows (see \cite{Fleischer:2010sq} (4.35)-(4.37)): \begin{eqnarray} I_{5}^{\mu\, \nu\, \lambda}&&= \sum_{i,j,k=1}^{5} \, q_i^{\mu}\, q_j^{\nu} \, q_k^{\lambda} E_{ijk}+\sum_{k=1}^5 g^{[\mu \nu} q_k^{\lambda]} E_{00k}, \label{Exyz0} \end{eqnarray} with \begin{eqnarray} \label{Exyz1} E_{00k} &=& \sum_{s=1}^5 \frac{1}{{0\choose 0}_5} \left[\frac{1}{2} {0s\choose 0k}_5 I_4^{[d+],s}- \frac{d-1}{3} {s\choose k}_5 I_4^{[d+]^2,s} \right] , \\ E_{ijk} &=&- \sum_{s=1}^5\frac{1}{{0\choose 0}_5} \left\{ \left[{0j\choose sk}_5 I_{4,i}^{[d+]^2,s}+ (i \leftrightarrow j)\right]+{0s\choose 0k}_5 {\nu}_{ij} I_{4,ij}^{[d+]^2,s} \right\}. \label{Exyz2} \end{eqnarray} \vspace{0.5cm} Contraction of the tensor with three chords yields: \begin{eqnarray} \label{Proj3} &&q_{a \mu} q_{b \nu} q_{c \lambda}I_{5}^{\mu\, \nu\, \lambda}= \sum_{i,j,k=1}^4 (q_a \cdot q_i) (q_b \cdot q_j) (q_c \cdot q_k)E_{ijk}~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ \nonumber \\ &&~~~~+~ \sum_{k=1}^4 \left[(q_a \cdot q_b) (q_c \cdot q_k) +(q_a \cdot q_c) (q_b \cdot q_k)+(q_b \cdot q_c) (q_a \cdot q_k)\right] E_{00k} . \end{eqnarray} \vspace{0.5cm} Introducing ${{\rm{Y}}}_a =Y_{a5}-Y_{55}$, ${{\rm{D}}}_a^s ={\delta}_{as}-{\delta}_{5s}$ and the following kinematical objects, \begin{eqnarray} \label{P1I4d} P_{I4}&=&\frac{1}{8}\frac{1}{{0s\choose 0s}_5} \left\{ {s\choose s}_5 \left[ {s\choose 0}_5 {{\rm{Y}}}_a {{\rm{Y}}}_b {{\rm{Y}}}_c+ {0\choose 0}_5 \left( {{\rm{Y}}}_a {{\rm{Y}}}_b {{\rm{D}}}_c^s + {{\rm{Y}}}_a {{\rm{Y}}}_c {{\rm{D}}}_b^s + {{\rm{Y}}}_b {{\rm{Y}}}_c {{\rm{D}}}_a^s \right) \right] \right. \nonumber \\ && \left. +{0\choose 0}_5 \left[ {s\choose 0}_5 \left( {{\rm{Y}}}_a {{\rm{D}}}_b^s {{\rm{D}}}_c^s + {{\rm{Y}}}_b {{\rm{D}}}_a^s {{\rm{D}}}_c^s + {{\rm{Y}}}_c {{\rm{D}}}_a^s {{\rm{D}}}_b^s \right) + {0\choose 0}_5 {{\rm{D}}}_a^s {{\rm{D}}}_b^s {{\rm{D}}}_c^s \right] \right\}, \\ \label{P3I4d} P_{Z4}&=&P_{I4}-\frac{1}{12} {\left( \right)}_5 \left\{{{\rm{Y}}}_a {{\rm{Y}}}_b {{\rm{D}}}_c^s +{{\rm{Y}}}_a {{\rm{Y}}}_c {{\rm{D}}}_b^s + {{\rm{Y}}}_b {{\rm{Y}}}_c {{\rm{D}}}_a^s \right\}, \\ P_{I3}&=&\frac{1}{24}\frac{{\left( \right)}_5 }{{0s\choose 0s}_5}\left\{ \left[ {{\rm{Y}}}_a {{\rm{D}}}_b^s+{{\rm{Y}}}_b {{\rm{D}}}_a^s \right] \left[{0s\choose 0s}_5 {{\rm{D}}}_c^t-{0s\choose 0t}_5 {{\rm{D}}}_c^s \right]+ (a \leftrightarrow c) + (b \leftrightarrow c) \right\} , \label{C3remain} \nonumber \\ \end{eqnarray} \vspace{0.5cm} we can write \newpage \begin{eqnarray} q_{a \mu} q_{b \nu} q_{c \lambda} I_{5}^{\mu\, \nu\, \lambda} = && \frac{d-2}{8} \frac{{\left(\right)}_5}{{0s\choose 0s}_5} ({\delta}_{ab}{\delta}_{ac}{\delta}_{as}-{\delta}_{5s}) (d-1)I_{4}^{[d+]^2,s} - \frac{1}{{0\choose 0}_5} \left\{ P_{I4}~ I_4^{[d+],s} \nonumber \right. \\ && \left. -P_{Z4} ~Z_4^{[d+],s}+P_{I3}~ I_3^{[d+],st} +\frac{1}{3}\left[ {\Sigma}_{c}^{1,s} R_{ab} + {\Sigma}_{b}^{1,s} R_{ac} +{\Sigma}_{a}^{1,s} R_{cb} \right] \right\} \nonumber \\ && + F_{abc}^s, \end{eqnarray} symmetric in the indices $a,b,c$ and summation over $s,t$ assumed. Here, the \begin{eqnarray} R_{ab} =&& -~\frac{1}{{0s\choose 0s}_5} \left\{\frac{1}{{0s\choose 0s}_5}{\Sigma}^{2,s}_{b}{\Sigma}^{2,st}_{a}(d-2)I_3^{[d+],st} \right. \nonumber \\ &&+~ \left. \frac{1}{{0st\choose 0st}_5}{\Sigma}^{2,st}_{b} \left[{\Sigma}^{3,st}_{a}(d-2)I_3^{[d+],st}-\sum_{u=1}^{5}{\Sigma}^{2,stu}_{a}I_2^{stu} \right] \right\} \end{eqnarray} contains only $3$-point functions and no inverse ${s\choose s}_5$. Further, \begin{eqnarray} F_{abc}^s=-\frac{1}{24} \frac{{\left(\right)}_5}{{0s\choose 0s}_5} \frac{{s\choose 0}_5 }{{0\choose 0}_5 } \left[{{\rm{Y}}}_c {{\rm{D}}}_a^s {{\rm{D}}}_b^s+{{\rm{Y}}}_b {{\rm{D}}}_a^s {{\rm{D}}}_c^s+{{\rm{Y}}}_a {{\rm{D}}}_b^s {{\rm{D}}}_c^s \right] \label{term2} \end{eqnarray} is a rational term obtained from an $\varepsilon$-expansion. The fact that no scalar products from \eqref{Proj3} remain demonstrates that the $g^{\mu \nu}$ term has canceled. \vspace{0.5cm} For the selfcontracted tensor we obtain \begin{eqnarray} \label{resultkkv} && q_{a \lambda}I_{5,\mu}^{~~~~\mu\, \lambda} = -\frac{1}{{0\choose 0}_5}\sum_{s=1}^5\left\{{\Sigma}_a^{1,s} \left[1+\frac{1}{2} \frac{1}{{0s\choose 0s}_5} \left({s\choose s}_5 Y_{55}+ 2 {s\choose 0}_5 {\delta}_{s5} \right)\right] \right. \nonumber \\ && \left. ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~+ \left( \right)_5 \left(Y_{a5}-Y_{55}\right) {\delta}_{s5} \right\} I_4^{[d+],s} \nonumber \\ &&-\frac{1}{{0\choose 0}_5}\sum_{s,t=1}^5 {\Sigma}_a^{1,s} \frac{{ts\choose 0s}_5}{{0s\choose 0s}_5} \frac{Y_{55}}{{0st\choose 0st}_5} \left[\frac{d-2}{2}{st\choose st}_5 I_3^{[d+],st}+\frac{1}{2} \sum_{u=1}^5 {st0\choose stu}_5 I_2^{stu} \right]\nonumber \\ &&+\frac{{\delta}_{s5}}{{0\choose 0}_5}\sum_{t=1}^5 {\Sigma}_a^{1,s}\frac{1}{{0s\choose 0s}_5 {0st\choose 0st}_5}\left\{\left[{0s\choose 0t}_5{st\choose st}_5~~-{ts\choose 0s}_5{st\choose 0t}_5 ~~\right]\frac{d-2}{2}I_3^{[d+],st} \nonumber \right. \\ && \left.~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~+ \sum_{u=1}^5\left[{0s\choose 0t}_5{st0\choose stu}_5-{ts\choose 0s}_5{st0\choose 0tu}_5\right] \frac{1}{2}I_2^{stu} \right\}\nonumber \\ &&-\frac{{\delta}_{t5}}{{0\choose 0}_5}\sum_{s=1}^5 {\Sigma}_a^{1,s}\frac{1}{{0s\choose 0s}_5 {0st\choose 0st}_5}\left\{\left[{0s\choose 0s}_5{st\choose st}_5~~+{ts\choose 0s}_5{ts\choose 0s}_5 ~~\right]\frac{d-2}{2}I_3^{[d+],st} \nonumber \right. \\ && \left.~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~+ \sum_{u=1}^5\left[{0s\choose 0s}_5{st0\choose stu}_5+{ts\choose 0s}_5{ts0\choose 0su}_5\right] \frac{1}{2}I_2^{stu} \right\}\nonumber \\ &&+\frac{1}{{0\choose 0}_5}\sum_{s,t=1}^5 {\Sigma}_a^{1,s}\frac{1}{{0s\choose 0s}_5} \left\{ \left[{s\choose t}_5{\delta}_{s5}-{s\choose s}_5{\delta}_{t5}\right]\frac{d-2}{2}I_3^{[d+],st} - {ts\choose 0s}_5\frac{1}{2}I_2^{st5} \right\}\nonumber \\ &&+\frac{1}{{0\choose 0}_5}\sum_{s,t=1}^5 \left[{t\choose 0}_5 \left({\delta}_{as}-{\delta}_{5s} \right)- {s\choose 0}_5 \left({\delta}_{at}-{\delta}_{5t} \right) \right] I_3^{[d+],st}. \end{eqnarray} The first two lines of \eqref{resultkkv} contain complete double sums while the remaining terms contribute only for specific values of $s,t,u$. \section*{Acknowledgements} J.F. thanks DESY for kind hospitality. Work is supported in part by Sonderforschungsbereich/Trans\-re\-gio SFB/TRR 9 of DFG ``Com\-pu\-ter\-ge\-st\"utz\-te Theoretische Teil\-chen\-phy\-sik" and European Initial Training Network LHCPHENOnet PITN-GA-2010-264564. \small
\section{Introduction} Asymptotic frequentist properties of Bayesian non-parametric methods have received a lot of attention in recent years. It is now recognized that a single, fully automatic Bayesian model can offer adaptive, optimal rates of convergence for large collections of {\it true} data generating distributions, ranging over several smoothness classes. In a seminal work, \citet{vandervaart&vanzanten09} establish adaptability of rescaled Gaussian process models for non-parametric regression, classification and density estimation. \cite{rousseau10} discusses adaptive density estimation with finite beta mixtures with a hierarchical prior on the number of mixture components. \citet{kruijer.10} and \cite{dejonge.zanten10} derive similar results for finite location-scale mixture models, respectively, in density estimation and regression, again with a prior on the number of mixture components. Quite interestingly, adaptability has not yet been established for Dirichlet process (DP) mixture of normals models for density estimation. Even rates of convergence of these models remain to be derived beyond the univariate case. This is surprising because these models are the most studied of all Bayesian non-parametric models, and have been among the firsts for which positive results about convergence of the posterior were established \citep{ghosal99, ghosal&vandervaart01, ghosal&vandervaart07}. The main challenge in establishing adaptability of DP mixture models and to derive rates of convergence in higher dimensions lies in constructing a suitable {\it low-entropy, high-mass sieve} on the space of non-parametric mixture densities. Such sieve constructions are an integral part of the current technical machinery for deriving rates of convergence. The sieves that have been used to study DP mixture models \citep[e.g., in][]{ghosal&vandervaart07} do not scale to higher dimensions and lack adaptability to smoothness classes \citep{wughosal10}. The main import of this article is to plug this gap. It is demonstrated that a novel sieve construction proposed by this author \citep[reported earlier in an yet unpublished paper][]{pati.etal.11} give the desired dimension-scalability and smoothness-adaptability. This sieve utilizes the well known stick-breaking representation of a DP \citep{sethuraman94} and can be potentially useful for studying a large class of stick-breaking mixture models beyond the DP mixtures \citep[e.g.,][]{dunsonpark08, chungdunson09, rodriguez11}. This sieve paves way to the following results. For independent and identically distributed observations $X_1, \cdots, X_n$ from an unknown density $p$ on $\mathbb{R}^d$, posterior convergence rates are derived for a simple DP location mixture model at a true data generating density $p_0$ belonging to either a class of infinitely differentiable densities or a class of compactly supported densities with two continuous derivatives. The derived rates are minimax optimal for these classes (up to logarithmic factors), and adapt to these two classes without requiring any user intervention to select or estimate any tuning parameters. The two classes considered here form two extremes of the classes of smooth densities. Finer rate adaptability results can be derived by looking at the intermediate classes of H\"older smooth densities. These classes have well defined minimax optimal rates associated with them. It is demonstrated that the new sieve works for all H\"older classes. However, we stop short of deriving precise rates of convergence for these classes. This derivation requires an additional calculation of prior thickness rates for a $p_0$ belonging to these classes, which is a challenging and interesting problem but is tangential to the focus of this article. Interested readers are referred to some recent developments reported in \citet{kruijer.10}. \section{A simple DP location mixture model} \label{model} Let $\phi_\sigma$ denote the density of the $d$-variate normal distribution with mean zero and variance $\sigma^2 I$. For any probability measure $F$ on $\mathbb{R}^d$, use $p_{F, \sigma}$ to denote the mixture density \begin{equation} \label{def mix} p_{F,\sigma}(x) = \int \phi_\sigma(x - z) dF(z),\;\; x \in \mathbb{R}^d. \end{equation} Assign $p$ a prior distribution $\Pi$ given by the law of the random density $p_{F, \sigma}$ when $(F, \sigma^{-d}) \sim \textit{DP}(\alpha) \times \textit{Ga}(a, b)$ where $\textit{DP}(\alpha)$ denotes the Dirichlet process distribution \citep{ferguson73} with base measure $\alpha$ and $\textit{Ga}(a,b)$ denotes the gamma distribution with shape $a$ and rate $b$. It is useful to recall two different characterizations of DP distributions, the original characterization by \citet{ferguson73} through a consistent system of Dirichlet distributions over measurable partitions and the later stick-breaking interpretation due to \citet{sethuraman94}. The first approach characterizes an $F \sim \textit{DP}(\alpha)$, with $\alpha$ a finite measure on $\mathbb{R}^d$, as: \begin{equation} \label{dp1} (F(B_1), \cdots, F(B_k)) \sim \textit{Dir}(\alpha(B_1), \cdots, \alpha(B_k)). \end{equation} for any Borell measurable partition $B_1, \cdots, B_k$ of $\mathbb{R}^d$. The stick breaking characterization says an \begin{equation} \label{dp2} F = \sum_{h = 1}^\infty \pi_h \delta_{Z_h}, \;\; \pi_h = V_h \prod_{j < h}(1 - V_j),\;\;\delta_x = \mbox{ Dirach measure at }x, \end{equation} has a $\textit{DP}(\alpha)$ distribution if $\{V_h, h \ge 1\}$ are independent $\textit{Be}(1, |\alpha|)$ random variables with $|\alpha| = \alpha(\mathbb{R}^d)$, $\{Z_h, h \ge 1\}$ are independently distributed according to the probability measure $\bar \alpha = \alpha / |\alpha|$ and these two sets of random variables are mutually independent. The base measure $\bar\alpha$ gives the mean of $F$, and also determines its support. The only assumptions we make on $\bar\alpha$ are that it admits a Lebesgue density that is strictly positive over the whole of $\mathbb{R}^d$ and that for some constant $b_1$, $\bar \alpha([-a,a]^d) \lesssim \exp(-b_1 a^2)$, where $f(a) \lesssim g(a)$ means $f(a) \le K g(a)$ for all $a$, for some fixed constant $K$. \section{Posterior convergence rates and adaptability} \label{prelim} Consider modeling $d$-variate measurements $X_1, X_2, \cdots$ as independent observations from a density $p$, which is assigned a prior distribution $\Pi$. Here $\Pi$ is a probability measure on the space $\mathcal{P}$ of Lebesgue probability densities, equipped with the Borel $\sigma$-field under a metric $\rho$, usually taken to be the $L_1$ metric $\rho(p, q) = \|p - q\|_1 = \int_{\mathbb{R}^d} |p(x) - q(x)|dx$ or the Hellinger metric $\rho(p, q) = h(p, q) = [\int_{\mathbb{R}^d}\{p^{1/2}(x) - q^{1/2}(x)\}^2]^{1/2}$. Let $\Pi_n(\cdot|X_1, \cdots, X_n)$ denote the posterior distribution of $p$ based on the first $n$ measurements, defined for every measurable $B \subset \mathcal{P}$ as $$\Pi_n(B | X_1, \cdots, X_n) = \frac{\int_B \prod_{i = 1}^n p(X_i)\Pi(dp)}{\int_\mathcal{P} \prod_{i = 1}^n p(X_i)\Pi(dp)}.$$ Let $\{\epsilon_n\}_{n \ge 1}$ be a sequence of positive numbers with $\lim_{n \to \infty} \epsilon_n = 0$. For any $p_0 \in \mathcal{P}$ we say the posterior convergence rate at $p_0$ is (not slower than) $\epsilon_n$ if for some finite constant $M$ \begin{equation} \label{def rate} \lim_{n \to 0} \Pi(\{p: \rho(p_0, p) > M\epsilon_n\} | X_1, \cdots, X_n) = 0 \end{equation} almost surely whenever $X_1, X_2,\cdots$ are independent and identically distributed (iid) with density $p_0$. Although \eqref{def rate} only establishes $\{\epsilon_n\}_{n \ge 1}$ as a bound on the convergence rate, it serves as a useful calibration of the method induced by $\Pi$ for classes of true densities $p_0$ for which optimal estimation rates are known. For example, for various classes of infinitely differentiable densities the optimal rate is known to be $n^{-1/2}(\log n) ^k$ for some $k \ge 0$ \citep{ibkhas83}, whereas for the class of compactly supported, twice continuously differentiable densities, the optimal rate is known to be $n^{-2/(4 + d)}$ \citep{huang04}. A method is considered adaptive if it provides convergence rates that are within a power of $\log n$ of these optimal rates. Along this line, we present the following results. \begin{theorem} \label{post rate} Let $\Pi$ be the DP mixture prior of Section \ref{model}. \begin{enumerate} \item If $p_0$ equals $p_{F_0, \sigma_0}$ for some probability measure $F_0$ on $\mathbb{R}^d$ and some $\sigma_0 > 0$, then \eqref{def rate} holds with $\epsilon_n = n^{-1/2} (\log n)^{(d + 1 + s )/2}$ for every $s > 0$. Such a $p_0$ will be called a super-smooth density. \item If $p_0$ is compactly supported and twice continuously differentiable then \eqref{def rate} holds with $\epsilon_n = n^{-2/(4 + d)}(\log n)^{(4d + 2)/(d + 4) + s}$ for every $s > 0$. Such a $p_0$ will be called an ordinary-smooth density. \end{enumerate} \end{theorem} These results are proved in Sections \ref{s sieve} and \ref{s conc}. The main tool needed to establish \eqref{def rate} is a set of sufficient conditions proposed in \citet[Theorem 2.1]{ghosal&etal00}. We present here a slightly modified version adapted from \citet[Theorem 2.1]{ghosal&vandervaart01}. In the following, for any two probability densities $p$ and $q$ and any positive number $\epsilon$, we denote $K(p, q) = \int_{\mathbb{R}^d} p(x)\log \{p(x)/q(x)\} dx$, $V(p, q) = \int_{\mathbb{R}^d} p(x)[\log \{p(x)/q(x)\}]^2 dx$, $B(\epsilon; p) = \{q \in \mathcal{P}: K(p, q) \le \epsilon^2, V(p, q) \le \epsilon^2\}$. For any $\mathcal{Q} \subset \mathcal{P}$, its $\epsilon$-covering number $N(\epsilon, \mathcal{Q}, \rho)$ is defined to be the minimum number of balls of radius $\epsilon$ (in the metric $\rho$) needed to cover $\mathcal{Q}$; with $\log N(\epsilon, \mathcal{Q}, \rho)$ referred to as the $\epsilon$-entropy of $\mathcal{Q}$. \begin{theorem} \label{suff} Let $\rho$ be the Hellinger metric on $\mathcal{P}$. Suppose for positive sequences $\tilde\epsilon_n$, $\bar\epsilon_n \to 0$ with $n\min(\tilde\epsilon^2_n,\bar \epsilon^2_n) \to \infty$, there exist positive constants $c_1, c_2, c_3, c_4$ and sets $\mathcal{P}_n \subset \mathcal{P}$, $n \ge 1$, such that for all large $n$ \begin{align} \log N(\bar \epsilon_n, \mathcal{P}_n, \rho) & \le c_1 n\bar\epsilon_n^2, \label{gv1}\\ \Pi(\mathcal{P}_n^c) & \le c_3 e^{-(c_2 + 4)n\tilde \epsilon_n^2} ,\label{gv2}\\ \Pi(B(\tilde\epsilon_n; p_0)) & \ge c_4e^{-c_2 n\tilde\epsilon_n^2} \label{gv3}. \end{align} Then \eqref{def rate} holds with $\epsilon_n = \max(\tilde\epsilon_n, \bar\epsilon_n)$. \end{theorem} \begin{remark} If (\ref{def rate}) holds with $\rho = $ the Hellinger metric then it holds with $\rho =$ the $L_1$ metric, because for any two probability density $\|p - q\|_1 \le 2 h(p, q)$. \end{remark} It is common to call the sequence $\{\mathcal{P}_n\}_{n \ge 1}$ a sieve on $\mathcal{P}$. The first two conditions require existence of a low-entropy, high mass sieve. The third condition requires a quantitative bound on the thickness of the prior $\Pi$ at the true density $p_0$. We first take up the more challenging task of sieve construction for the DP mixture prior of Section \ref{model}, followed by prior thickness calculations. \section{Sieve construction} \label{s sieve} \subsection{The basic construct} The chief novelty of the sieve proposed in \citet{pati.etal.11} lies in exploiting the stick-breaking representation of a DP distribution. A high-mass, low-entropy subset of $\mathcal{P}$ can be obtained by considering densities $p_{F,\sigma}$, with $F$ as given in \eqref{dp2} with limited tail mass $\sum_{h > H} \pi_h$. A precise statement is given below. \begin{theorem} \label{basic sieve} Fix reals $\epsilon, a, \underline{\sigma} > 0$ and integers $M, H \ge 1$. Define \begin{equation} \label{eff} \mathcal{Q} = \bigg\{p_{F, \sigma}: F = \sum_{h = 1}^\infty \pi_h \delta_{z_h}: z_h \in [-a,a]^d, h \le H; \sum_{h > H} \pi_h < \epsilon; 1 < \frac{\sigma}{\underline{\sigma}} < (1 + \epsilon)^M\bigg\}. \end{equation} Then, for some positive constants $b_1, b_2$ and $b_3$, \begin{enumerate} \item $\log N(\epsilon, \mathcal{Q}, \rho) \lesssim dH \log\frac{a}{\underline{\sigma}\epsilon} + H \log \frac{1}{\epsilon} + \log M$, where $\rho$ is either the $L_1$ or the Hellinger metric. \item If $\Pi$ is the DP mixture prior of Section \ref{model}, then $\Pi(\mathcal{Q}^c) \lesssim He^{-b_1a^2} + e^{-b_2\underline{\sigma}^{-d}} + \underline{\sigma}^{-b_3d}(1 + \epsilon)^{-b_3dM} + \{(e|\alpha|/H)\log(1/\epsilon)\}^H$. \end{enumerate} \end{theorem} \begin{proof} Let $R^*$ be a $(\underline{\sigma}\epsilon)$-net of $[-a,a]^d$ and let $S^*$ be an $\epsilon$-net of the $H$-simplex $\S_H = \{p = (p_1, \cdots, p_H): p_h \ge 0, \sum_h p_h = 1\}$. It is well known that the size of $R^*$ is $\lesssim \{a/(\underline{\sigma}\epsilon)\}^d$ and that of $S^*$ is $\lesssim (1 / \epsilon)^H$. For any $p_{F, \sigma} \in \mathcal{Q}$, with $F = \sum_{h = 1}^\infty z_h \delta_{z_h}$, find $z^*_1, \cdots, z^*_H \in R^*$, $\pi^* = (\pi^*_1, \cdots, \pi^*_H) \in S^*$ and $m^* \in \{1, \cdots, M\}$ such that \begin{align} & \max_{1 \le h \le H}\|z_h - z^*_h\| < \underline{\sigma} \epsilon,\\ & \sum_{h = 1}^H |\tilde \pi_h - \pi^*_h| < \epsilon,\mbox{ where }\tilde \pi_h = \frac{\pi_h }{ 1-\sum_{l > H} \pi_l}, 1 \le h \le H, \mbox{ and}\\ & \sigma^* = \underline{\sigma}(1 + \epsilon)^{m^*}\mbox{ satisfies }1 < \sigma/\sigma^* < 1 + \epsilon. \end{align} Then, with $F^* = \sum_{h = 1}^H \pi^*_h \delta_{z^*_h}$, we have, \begin{align*} \|p_{F, \sigma} - p_{F^*, \sigma^*}\|_1 & \le \|p_{F, \sigma} - p_{F, \sigma^*}\|_1 + \|p_{F, \sigma^*} - p_{F^*, \sigma^*}\|_1\\ & \le \frac{\sigma - \sigma}{\sigma^*} + \sum_{h > H} \pi_h + \sum_{h = 1}^H \pi_h \|\phi_{\sigma^*}(\cdot - z_h) - \phi_{\sigma^*}(\cdot - z^*_h)\|_1 + \sum_{h = 1}^H |\pi_h - \pi^*_h|. \end{align*} Each of the first three terms above is smaller than or equal to $\epsilon$. The last term is smaller than or equal to $(1 - \sum_{h > H}\pi_h)\sum_{h = 1}^H|\tilde \pi_h - \pi^*_h| + \sum_{h > H}\pi_h\sum_{h = 1}^H \pi^*_h \le 2\epsilon$ Thus a $5\epsilon$-net of $\mathcal{Q}$, in the $L_1$ topology, can be constructed with $p^* = p_{F^*, \sigma^*}$ as above. The total number of such $p^*$ is $\lesssim (\frac{a}{\underline{\sigma}\epsilon})^{dH}(\frac{1}{\epsilon})^HM$. This proves the first assertion of the theorem with $\rho = \|\cdot\|_1$; the constant multiplication by 5 can be absorbed in $\lesssim$ form of the bound. The same obtains for $\rho =$ the Hellinger metric because it is bounded by the square-root of the $L_1$ metric. Now with $\Pi$ denoting the DP mixture prior of Section \ref{model}, we have a stick-breaking representation of a random $p\sim \Pi$ given by $p = p_{F, \sigma} = \sum_{h = 1}^\infty \pi_h \phi_\sigma(\cdot - Z_h)$ with $\pi_h$ and $Z_h$ as described in \eqref{dp2} and the paragraph that follows, and $\sigma^{-d} \sim \textit{Ga}(a, b)$. Therefore, \begin{equation} \Pi(\mathcal{Q}^c) \le H \bar\alpha([-a,a]^d) + \Pr(\sigma^2 \not\in (\underline{\sigma}^2, \underline{\sigma}^2(1 + \epsilon)^{2M})) + \Pr\left(\sum_{h > H}\pi_h > \epsilon\right). \label{pc1} \end{equation} The first term is $\lesssim H\exp(-b_1a^2)$, by assumption on $\alpha$. The second term equals $\Pr(\sigma^{-d} \ge \underline{\sigma}^{-d}) + \Pr(\sigma^{-d} \le \underline{\sigma}^{-d}(1 + \epsilon)^{-Md}) \lesssim \exp(-1/b_2 \underline{\sigma}^{-d}) + (\underline{\sigma}^d(1 + \epsilon)^{Md})^{-b_3}$ because $\sigma^{-d} \sim \textit{Ga}(a, b)$. To bound the last term in \eqref{pc1}, note that $W = -\sum_{h = 1}^H\log(1 - V_h) \sim \textit{Ga}(H, |\alpha|)$, and therefore the last term equals $$\Pr(W < \log(1 / \epsilon)) \le (|\alpha|\log \frac{1}{\epsilon})^H / \Gamma(H + 1) \le \bigg(\frac{e|\alpha|}{H}\log\frac{1}{\epsilon}\bigg)^H$$ by Stirling's formula. This proves the second assertion. \end{proof} \subsection{Sieves for Theorem \ref{post rate}} The subset $\mathcal{Q}$ of Theorem \ref{basic sieve} can be easily adapted to form sieves targeted for different rates of convergence. Below we show this for the nearly parametric, super-smooth rate and also for the slower rates associated with H\"older classes of finitely differentiable functions. All this is done for any arbitrary dimension $d \ge 1$. \begin{proposition}[Super-smooth rate] \label{smooth sieve} Fix any $s > 0$. For $\tilde \epsilon_n = n^{-1/2}(\log n)^{(d + 1)/2}$ and $\bar \epsilon_n = \tilde\epsilon_n (\log n)^{s/2}$, there is a sequence of sets $\mathcal{P}_n$ such that $\log N(\bar \epsilon_n, \mathcal{P}_n, \rho) \lesssim n\bar\epsilon_n^2$ and $\Pi(\mathcal{P}_n^c) \lesssim \exp(-cn\tilde\epsilon_n^2)$ for every $c > 0$, where $\rho$ is either the $L_1$ or the Hellinger metric. \end{proposition} \begin{proof} Let $\mathcal{P}_n$ be defined as $\mathcal{Q}$ of \eqref{eff} with $\epsilon = \bar\epsilon_n = n^{-1/2}(\log n)^{(d + 1 + s)/2}$, $H = n\bar\epsilon_n^2 / \log n = (\log n)^{d + s}$, and $M = a^2 = \underline{\sigma}^{-d} = n$. Then, by Theorem \ref{basic sieve}, \begin{align*} \log N(\bar \epsilon_n, \mathcal{P}_n, \rho) & \lesssim d (\log n)^{d + s + 1} + (\log n)^{d + s + 1} + \log n\\ & \lesssim (\log n)^{d + s + 1} = n\bar \epsilon_n^2 \end{align*} which proves the first assertion. Also, \begin{equation} \label{step1} \Pi(\mathcal{P}_n^c) \lesssim (\log n)^{d + s}e^{-b_1n} + e^{-b_2 n} + n^{b_3}e^{-b_3 dn\log(1 + \bar \epsilon_n)} + (\log n)^{-(d + s - 1)(\log n)^{d + s}}. \end{equation} For any $c > 0$, the first, second and fourth terms on the right hand side of \eqref{step1} are clearly bounded by $C\exp(-c(\log n)^{d + s})$ for some constant $C$. The third term, too, is bounded by the same, possibly with different $C$ because $n\log(1 + \bar\epsilon_n) \gtrsim n\bar\epsilon_n^2 = (\log n)^s(\log n)^{d + 1} > c(\log n)^{d + s}$. And therefore $\Pi(\mathcal{P}_n^c) \lesssim \exp(-c(\log n)^{d + 1})$. This proves the second assertion of the theorem. \end{proof} \begin{proposition}[H\"older-smooth sieve] \label{ord sieve} Fix any $\beta \in (0, 1/2)$, $q \ge 0$ and $s > 0$. For $\tilde \epsilon_n = n^{-\beta}(\log n)^{q}$, $\bar \epsilon_n = \epsilon_n (\log n)^s$, there is a sequence of sets $\mathcal{P}_n$ such that $\log N(\epsilon_n, \mathcal{P}_n, \rho) \lesssim n\bar \epsilon_n^2$ and $\Pi(\mathcal{P}_n^c) \lesssim \exp(-cn\tilde\epsilon_n^2)$ for every $c > 0$, where $\rho$ is either the $L_1$ or the Hellinger metric. \end{proposition} \begin{proof} Let $\mathcal{P}_n$ be defined as on the right hand side of \eqref{eff} with $\epsilon = \bar \epsilon_n = n^{-\beta}(\log n)^{q + s}$, $H = n\bar\epsilon_n^2 / \log n = n^{1 - 2\beta}(\log n)^{2(q + s) - 1}$, $M = a^2 = \underline{\sigma}^{-d} = n$. Then by Theorem \ref{basic sieve}, $\log N(\epsilon_n, \mathcal{P}_n, \rho) \lesssim n^{1 - 2\beta} (\log n)^{2(q + s)}$ and for every $c > 0$, \begin{align*} \label{step1b} \Pi(\mathcal{P}_n^c) & \lesssim n^{1 - 2\beta}(\log n)^{2(q + s) - 1}e^{-b_1n} + e^{-b_2 n} + n^{b_3}e^{-b_3 dn\log(1 + \bar\epsilon_n)} + n^{-(1 - 2\beta)n^{1 - 2\beta}(\log n)^{2(q + s)-1}}\\ & \lesssim e^{-(1 - 2\beta)n^{1 - 2\beta}(\log n)^{2(q + s)}} \lesssim e^{-cn^{1 - 2\beta}(\log n)^{2q}}. \end{align*} \end{proof} The ordinary-smooth rate corresponds to $\beta = 2 / (4 + d)$, and more generally, a H\"older class of functions with continuous derivatives up to order $k$ corresponds to $\beta = k / (2k +d)$. \section{Prior Thickness} \label{s conc} With sieve conditions (\ref{gv1}), (\ref{gv2}) taken care of, a proof of Theorem \ref{post rate} requires establishing the prior thickness property (\ref{gv3}) of $\Pi$ for each of the two classes of densities. Below we show that for a $p_0$ from either class, $\Pi(B(A\tilde \epsilon_n; p_0)) \gtrsim e^{-cn\tilde\epsilon_n^2}$ for some constants $A > 0, c > 0$ ,with $\tilde \epsilon_n$ as in Proposition \ref{smooth sieve} or Proposition \ref{ord sieve} as appropriate (with $\beta = 2 / (4 + d)$). This immediately leads to $\Pi(B(\tilde \epsilon_n; p_0)) \gtrsim e^{-c_2 n\tilde \epsilon_n^2}$ for some finite number $c_2 > 0$ and completes a proof of Theorem \ref{post rate}, with $\epsilon_n = \bar \epsilon_n$, because Propositions \ref{smooth sieve} and \ref{ord sieve} hold for all constants $c > 0$, including, $c = c_2 + 4$, as needed by Theorem \ref{suff}. We will first tackle prior thickness at ordinary-smooth densities $p_0$ which present a bigger challenge than the super-smooth ones. Our proof closely follows the calculations presented in \citet{ghosal&vandervaart07} with some minor adaptation needed to handle higher dimensions. For this reason, most of the results are presented in the Appendix, with proofs given only for those where some adaptation is needed. However, we present the main argument below, because a similar argument presented in \citet[Section 9]{ghosal&vandervaart07} leaves some gaps (pun intended). \begin{proposition}[Ordinary-smooth thickness] \label{ordinarysmooth} Suppose $p_0$ is compactly supported and $$\int (\|\nabla p_0\| / p_0)^4 p_0d\lambda < \infty,\;\; \int (\|\nabla^2 p_0\|_2 / p_0)^2 p_0 d\lambda < \infty,$$ where $\|A\|_2$ denotes the spectral norm of a matrix $A$. Then $ \Pi(B(A\tilde \varepsilon_n; p_0)) \gtrsim e^{-cn\tilde\varepsilon_n^2}$ with $\tilde \epsilon_n = n^{-2/(4 + d)}(\log n)^{(4d + 2)/(d + 4)}$ for some constants $A > 0, c > 0$. \end{proposition} \begin{proof} Fix a $\sigma^2 \in \tilde\epsilon_n \{\log (1/\tilde\epsilon_n)\}^{-2} \cdot (1/2, 1)$. Find a $b > 1$ such that $\tilde\epsilon_n^b \{\log(1/\tilde\epsilon)\}^{9/4} \le \tilde\epsilon_n$. Let $P_0$ denote the probability measure associated with the density $p_0$. By Corollary \ref{cor disc approx}, there is a discrete probability measure $F_\sigma = \sum_{j = 1}^{N} p_j \delta_{z_j}$ with at most $N \lesssim \sigma^{-d} \log( 1 / \tilde\epsilon_n)^d$ support points in $[-a,a]^d$, with at least $\sigma \tilde\epsilon_n^{2b}$ separation between any $z_i \ne z_j$, such that $$\|p_{P_0, \sigma} - p_{F_\sigma,\sigma}\|_\infty \lesssim \varepsilon_n^{2b} / \sigma^{d + 1}~\mbox{and}~\|p_{P_0, \sigma} - p_{F_\sigma,\sigma}\|_1 \lesssim \varepsilon_n^{2b}\{\log(1/\varepsilon_n)\}^{1/2}.$$ Place disjoint balls $U_j$ with centers at $z_j$, $j = 1, \cdots, N$ with diameter $\sigma\varepsilon_n^{2b}$ each. Extend $\{U_1, \cdots, U_N\}$ to a partition $\{U_1, \cdots, U_K\}$ of $[-a, a]^d$ such that each $U_j$, $j = N + 1, \cdots, K$, has diameter smaller than or equal to $\sigma$. This can be done with $K \lesssim \sigma^{-d}\{\log(1/\tilde\epsilon_n)\}^d$. Further extend this to a partition $U_1, \cdots, U_M$ of $\mathbb{R}^d$ such that $(\sigma\tilde\epsilon_n^{2b})^d \lesssim \alpha(U_j) \le 1$ for all $j = 1, \cdots, M$. We can still have $M \lesssim \sigma^{-d}\{ \log(1 / \tilde\epsilon_n)\}^d \lesssim \tilde\epsilon_n^{-d/2}\{\log(1/\tilde\epsilon_n)\}^{2d}$. Define $p_j = 0$, $j = N + 1, \cdots, M$. Let $\mathcal{P}_\sigma$ denote the set of probability measures $F$ on $\mathbb{R}^d$ with $\sum_{j = 1}^M |F(U_j) - p_j| \le 2\tilde\epsilon_n^{2db}$ and $\min_{1 \le j \le M} F(U_j) \ge \tilde\epsilon_n^{4db} / 2$. Then, by Lemma \ref{lem diff} (with $V_i = U_i$, $i = 1,\cdots, N$, $V_0 = \cup_{j > N} V_j$) for any $F \in \mathcal{P}_\sigma$, $\|p_{F_\sigma,\sigma} - p_{F, \sigma}\|_\infty \lesssim \tilde\epsilon_n^{2b}/\sigma^d$, $\|p_{F_\sigma,\sigma} - p_{F, \sigma}\|_1 \lesssim \tilde\epsilon_n^{2b}$ and hence, by Lemma \ref{lem dist} and Lemma \ref{lem p0}, \begin{align*} h(p_0, p_{F, \sigma}) & \le h(p_0, p_{P_0, \sigma}) + h(p_{P_0, \sigma}, p_{F_\sigma, \sigma}) + h(p_{F_\sigma, \sigma}, p_{F, \sigma})\\ & \lesssim \sigma^2 + \tilde\epsilon_n^b \{\log(1/\tilde\epsilon_n)\}^{1/4} + \tilde\epsilon_n^b\\ & \lesssim \sigma^2 + \tilde\epsilon_n^b \{\log(1/\tilde\epsilon_n)\}^{1/4} \end{align*} Also, for any such $F$, for every $x \in [-a,a]^d$ with $J(x)$ denoting the $j \in \{1, \cdots, K\}$ such that $x \in U_j$, $$p_{F,\sigma}(x) \ge \int_{\|z - x\| \le \sigma} \phi_\sigma(x - z) dF(z) \gtrsim \frac{1}{\sigma^d} \int_{\|x - z\| \le \sigma} dF(z) \ge \frac{1}{\sigma^d}F(U_{J(x)}) \gtrsim \frac{\tilde\epsilon_n^{4db}}{\sigma^d}$$ because, $U_{J(x)}$, with diameter no larger than $\sigma$, must be a subset of the ball of radius $\sigma$ around $x$. So $F \in \mathcal{P}_\sigma$ implies $\log \|p_0 / p_{F, \sigma}\|_\infty \lesssim \log (1/\tilde\epsilon_n)$ and therefore, by Lemma \ref{lem dist}, $K(p_0, p_{F, \sigma}) \le A^2 \tilde\epsilon_n^2$ and $V(p_0 , p_{F, \sigma}) \le A^2 \tilde\epsilon_n^2$, for a universal constant $A > 0$ that does not depend on $\sigma$. Note that $M\tilde\epsilon_n^{2db} \lesssim \tilde\epsilon_n^{2db - d/2}\{\log(1/\tilde\epsilon_n)\}^{2d} \le 1$ and for some large constant $a_1 > 0$, $\tilde\epsilon_n^{2db} \lesssim a_1 \{\min_{1 \le j \le M} \alpha(U_j)\}^{2/3}$. So, by Lemma \ref{lem dir}, $\Pr(F \in \mathcal{P}_\sigma) \ge C \exp(-c M \log 1 / \tilde\epsilon_n) \gtrsim C\exp(-c\tilde\epsilon_n^{-d/2}\{\log(1/\tilde\epsilon_n)\}^{2d + 1})$, for some constants $C, c$ that depend on $\alpha(\mathbb{R}^d), a, d$ and $b$. Therefore, \begin{align*} \Pi(B(A\tilde\epsilon_n; p_0)) & \gtrsim \exp(-c\tilde\epsilon_n^{-d/2}\{\log(1/\tilde\epsilon_n)\}^{2d + 1})\Pr(\sigma^2 \in \tilde\epsilon_n \{\log(1/\tilde\epsilon_n)\}^{-2}\cdot (1/2,1))\\ & = \exp(-c\tilde\epsilon_n^{-d/2}\{\log(1/\tilde\epsilon_n)\}^{2d + 1})\Pr(\sigma^d \in \tilde\epsilon_n^{d/2} \{\log(1/\tilde\epsilon_n)\}^{-d}\cdot (1/2^d,1))\\ & \gtrsim \exp(-c\tilde\epsilon_n^{-d/2}\{\log(1/\tilde\epsilon_n)\}^{2d + 1}) \end{align*} because $\sigma^{-d}$ has a gamma distribution. From this the result follows if $\tilde\epsilon_n^{-d/2}\{\log(1 / \tilde\epsilon_n)\}^{2d + 1} \le n\tilde\epsilon_n^2$. With $\tilde\epsilon_n = n^{-2 / (4 + d)}(\log n)^{q}$, we get $n\tilde\epsilon_n^2 = n^{d/(4 + d)}(\log n)^{2q}$ and $\tilde\epsilon_n^{-d/2}\{\log (1/\tilde\epsilon_n)\}^{2d + 1} < n^{d/(4 + d)}(\log n)^{2d + 1 - dq/2}$ and hence the condition is satisfied if $2d + 1 - dq / 2 \le 2q$, i.e., if $q \ge (4d + 2) / (d + 4)$. \end{proof} Prior thickness calculation at a super-smooth $p_0$ follows along the same line, but is simpler because we can bypass the first step in the proof of Proposition \ref{ordinarysmooth} of approximating $p_0$ by a $p_{F,\sigma}$. In fact, this approximation is the main driver of the slower thickness rate $\tilde\epsilon_n$, the recent developments in \citet{kruijer.10} are about refining this approximation for densities that have higher order derivatives. \begin{proposition}[Super-smooth thickness] \label{supersmooth} If $p_0 = p_{F_0, \sigma_0}$ for some $F_0$ supported on $[-a,a]^d$, then $\Pi(B(A\tilde \varepsilon_n; p_0)) \gtrsim e^{-cn\tilde\varepsilon_n^2}$ with $\tilde \epsilon_n = n^{-1/2}(\log n)^{(d + 1)/2}$ for some constants $A, c > 0$. \end{proposition} \begin{proof} Fix a $\sigma \in \sigma_0 \cdot (1 - \tilde\epsilon_n\{\log(1/\tilde\epsilon_n)\}^{-2}, 1)$. Fix $b > 1$ such that $\tilde\epsilon_n^b\{\log(1/\tilde\epsilon_n)\}^{9/4} \le \tilde\epsilon_n$. Construct $\mathcal{P}_\sigma$ as before, but with $p_{F_0, \sigma}$ instead of $p_{P_0, \sigma}$. Because $\sigma$ is bounded from below by $\sigma_0/2$, this can be constructed with an $M \lesssim \{\log(1/\tilde\epsilon_n)\}^d$ and hence $\Pr(F \in \mathcal{P}_\sigma) \gtrsim \exp(-c\{\log(1/\tilde\epsilon)\}^{d + 1})$ for some constant $c$. Note that \begin{align*} \|p_0 - p_{F_\sigma, \sigma}\|_1 & \le \|p_0 - p_{F_0, \sigma}\|_1 + \|p_{F_0, \sigma} - p_{F_\sigma, \sigma}\|_1\le 1 - \sigma / \sigma_0 + \tilde\epsilon_n^{2b}\{\log(1/\tilde\epsilon_n)\}^{1/2}\\ & \le \tilde\epsilon_n \{\log(1 / \tilde\epsilon_n)\}^{-2} + \tilde\epsilon_n^{2b}\{\log(1/\tilde\epsilon_n)\}^{1/2} \end{align*} and therefore, $F \in \mathcal{P}_\sigma$ implies $K(p_0 , p_{F, \sigma}) \le A^2 \tilde\epsilon_n^2$ and $V(p_0, p_{F,\sigma})\le A^2 \tilde\epsilon_n^2$ for some universal constant $A > 0$ that does not depend on $\sigma$. Now, because $\Pr(\sigma \in \sigma_0(1 - \tilde\epsilon_n\{\log(1/\tilde\epsilon_n)\}^{-2}, 1)) \gtrsim \tilde\epsilon_n\{\log(1/\tilde\epsilon_n)\}^{-2} \gtrsim \exp(-\{\log(1/\tilde\epsilon_n)\}^{d + 1})$ we have $p_n \gtrsim \exp(-c\{\log(1/\tilde\epsilon_n)\}^{d + 1})$. From this the result follows if $\{\log(1/\tilde\epsilon_n)\}^{d + 1} \le n\tilde\epsilon_n^2$, which is satisfied with $\tilde\epsilon_n = n^{-1/2} (\log n)^q$ for $2q \ge d + 1$. \end{proof}
\section{Introduction} During solar flares, particles are believed to be accelerated, and plasma heated as a result of magnetic reconnection at an X-point or neutral sheet in the corona \citep{Kopp1976}. The accelerated electrons stream down to the footpoints of coronal magnetic loops, producing hard X-ray bremsstrahlung as they are thermalized by Coulomb collisions in the dense lower corona or chromosphere. The directly heated plasma already in the loop and the ablated chromospheric material produce hot loops below the reconnection site, with temperature that can be 20 MK or higher, and densities as high as 10$^{11}$ cm$^{-3}$. These loops are visible in soft X-rays, and later, as they cool down, become visible in EUV and in H$\alpha$. The reconnection site gradually moves upwards and continues to release energy, even as the X-ray flux diminishes. This translates into the appearance of higher and higher hot loops, and cooler loops at the lower altitudes \citep[see e.g.][]{Svetska1987}. Coronal mass ejections (CMEs) are often associated with flares. One of the models invoked in their creation is the catastrophe or flux-rope model \citep[see e.g.][]{LinForbes2000}, in which a current sheet (CS) is thought to extend from the top of the reconnected loop system to the plasma bubble that surrounds the expelled flux rope. A CS is supposed to be so thin as to make direct observation quite difficult. However, there have recently been reports of CS detection in the extended corona from observations acquired in the wake of CMEs by the {\it Ultraviolet Coronograph Spectrometer} \citep[UVCS;][]{Kohl1995}, in the form of narrow, very hot (several MK) features, most prominently in the Fe$^{17+}$ line: \citet{Ciaravella2002,Ko2003,Raymond2003,Lin2005, Bemporad2006, Bemporad2008, Ciaravella2008}. In particular, \citet{Ciaravella2008} have firmly established that the CS thickness (for one event, at least) to be between 0.04 and 0.08 $R_S$, far larger than classical ($\lesssim$100 m) or anomalous (a few 10s' of km) resistivity would predict. In a re-analysis of previous results, \citet{Bemporad2008} have explained these observations with the existence of many ($\sim$10$^{-11}$ to 10$^{-17}$ m$^{-3}$) microscopic CSs of small sizes ($\approx$10--10$^4$ m) that, through non-thermal turbulent broadening, can justify not only the high CS temperatures but also the large observed thicknesses of macroscopic CSs. \citet{Bemporad2006} examine one such event, that lasted at least 2.3 days. The CME to which these observations pertain started at around 17:00 UT on 2002 November 26 on the western limb of the Sun. \citet{Bemporad2006} discusses in detail UVCS, {\it Large Angle and Spectrometric Coronograph} \citep[LASCO;][]{Brueckner1995} and {\it Extreme Ultraviolet Imager} \citep[EIT;][]{Delaboudiniere1995} (instruments on board SOHO, the {\it Solar and Heliospheric Observatory}) observations of this event. Our paper will concentrate on examining the concurrent X-ray emission, with data from {\it GOES} (Geostationary Environmental Satellite), and the {\it Ramaty High Energy Solar Spectroscopic Imager} \citep[RHESSI][]{Lin2002}, which was launched a few months prior to this event. Section~\ref{sect:obs} will briefly summarize the observations reported by \citet{Bemporad2006}, then complement them with X-rays observations: lightcurves, spectra and imaging. Section~\ref{sect:discussion} will then discuss interpretations of these observations, and the possibility that accelerated non-thermal particles provide the energy required to power the X-ray source, and perhaps the CS. \section{Observations} \label{sect:obs} \subsection{Brief summary of previously reported observations:} \citet{Bemporad2006} have reported observing a current sheet (CS) in the wake of a CME that started on 2002 November 26 around 17:00 UT. That conclusion was mainly supported by UVCS observations (starting at 18:39 UT) of a hot (initially well beyond 8 MK) plasma above the western limb of the Sun ($\approx 25^{\circ}$ north latitude), in the same radial direction as the CME, at an altitude of about 0.7 $R_S$ above the solar photosphere, directly above a loop system observed with EIT. This hot plasma had a width of $\approx$100 Mm perpendicularly to the radial direction from the Sun and to our line of sight. It cooled to 3.5 MK after 2.3 days, at which point UVCS observations stopped. \citet{Bemporad2006} also estimated that adiabatic heating is insufficient to explain the hot plasma, at least initially, and that reconnection must be the source of the thermal energy. In his re-analysis, \citet{Bemporad2008} further strengthens that hypothesis. The remainder of this section will concentrate on complementing the aforementioned study with X-ray observations from RHESSI and GOES. \subsection{X-ray lightcurves and imaging} \begin{figure*}[ht!] \centering \includegraphics[width=16.6cm]{f1.ps} \caption{ GOES lightcurves: {\it black} is the 1--8$\AA$ flux, the {\it gray} line is the 0.5--4$\AA$ flux. The data has been smoothed using a 2-hour smoothing window. The {\it dashed line} represents a constant flux at 4.1$\times$10$^{-7}$ W m$^{-2}$. The {\it red line} represents RHESSI 4--8 keV flux {\it from the coronal source only}. } \label{fig:goes} \end{figure*} As can be seen in Figure~\ref{fig:goes} and~\ref{fig:tp}, on 2002 November 26, around 13:40 UT, the GOES ``baseline'' in both channels increased suddenly, and stayed fairly high until 2002 November 29 $\approx$00:00 UT. During this time interval, several flares occurred at different positions on the solar disc, revealed as individual peaks in the GOES lightcurves (which are spatially integrated) of Figure~\ref{fig:tp}. \begin{figure*}[ht!] \centering \includegraphics[height=15cm]{f2.ps} \caption{ {\it First (top) plot:} GOES lightcurves ({\it black:} 1--8 {\AA}, {\it gray:} 0.5--4 {\AA}). {\it Second plot:} EIT lightcurves ({\it black:} full-Sun, {\it light gray:} ROI is 600'' square centered around [1050,450], i.e. encompassing slightly more than Figure~\ref{fig:compositeimg}, {\it dark gray:} ROI is 200''x300'' rectangle centered at [1050,450]), shown as a green box in Figure~\ref{fig:compositeimg}. {\it Third plot:} RHESSI 4--8 keV flux from imaging with ROI being a 256'' square centered around [1100,400]. {\it Fourth plot:} Diamonds: source altitude in the 4--8 keV band, from RHESSI imaging using subcollimator 8. Dashed line: northern $EIT$ loop system altitude \citep[from][]{Bemporad2006}. Dotted line: southern $EIT$ loop system altitude \citep[from][]{Bemporad2006}. {\it Fifth plot:} Source azimuthal distance from an arbitrary radial, using RHESSI's subcollimator 8 in the 4--8 keV band. In the RHESSI plots (last three plots), the displayed information include the initial flare at $\approx$12:00 UT. Beyond 13:00 UT, the plots only display the information pertaining to the coronal source, removing any disc flares. } \label{fig:tp} \end{figure*} \begin{figure}[ht!] \centering \includegraphics[width=7cm]{f3.ps} \caption{ SOHO/EIT 304 {\AA} image taken at 23:12:10 on 2002 November 26, with RHESSI 6--12 keV contours (50\% level) at different times (5 minute exposures centered around 12:15:58, 13:46:04, \& 23:24:46 UT). Using values from Figure 6 and from Figure 9 (top left) of \citet{Bemporad2006}, we have drawn arrows that delimit the angular extend of the region where UVCS observed hot plasma at 1.7 $R_S$. The {\it long} arrows delimit the region of plasma above 5 MK, while the {\it short} arrows delimit plasma above 3 MK. The green box is the ROI used to compute the {\it dark gray} lightcurve in the second plot of Figure~\ref{fig:tp}. } \label{fig:compositeimg} \end{figure} At about 12:00 on November 26, a solar flare started (Figure~\ref{fig:tp}, first and third plots). It was observed with RHESSI on the western limb of the Sun, at about 25 degrees of north latitude. RHESSI observed it until $\approx$12:42, at which time it entered Earth's shadow. Very shortly (3--4 minutes) after the rise in GOES fluxes at $\approx$13:40, RHESSI came out of Earth's shadow, and imaged an X-ray source at the same solar latitude, but about 80 Mm above the limb (Figure~\ref{fig:compositeimg}). This high altitude coronal X-ray source (hereafter HACXS) remains observable by RHESSI until $\approx$01:10 the next day (2002 November 27), i.e. for almost 12 hours. During that 12-hours period, several disc flares occurred, and with their much higher fluxes, often drowned the HACXS when attempting imaging. From its start at about 13:45, to about 16:00, the HACXS moved mostly radially outward from the Sun, at about 1.6 km/s (Figure~\ref{fig:tp}). During that time, the HACXS flux increased and then decreased. At about 16:15, the source seemingly ``jumps'' in altitude, by about 60 Mm. It could be argued that our initial source actually dimmed, and that this is a new, different source that appears at higher altitude. The RHESSI coverage between $\approx$16:00 and $\approx$17:00 is spotty: during that time interval, a flare occurred on the eastern limb of the Sun (introducing noise in images of our region-of-interest, ROI), the spacecraft was initially in the South Atlantic Anomaly (with detectors turned off), and also spent time in Earth's shadow. The 5-minute image that shows a source midway between the two sites (at $\approx$16:20 in Figure~\ref{fig:tp}) suggest we might indeed have had a single exciter that jumped across 60 Mm in about 70 minutes ($\approx$14 km/s velocity). Between about 16:00 and 18:45, the X-ray source moved progressively faster towards the solar equator (azimuthal velocity close to 5 km/s), then stopped just as a flare at the footpoint of the loop system appears (position [900,400] in Figure~\ref{fig:compositeimg}). EIT images show the rise of a filament-like feature at around 16:12, from the same active region, and a cusp-like feature and expelled material at around 17:12--17:24, the latter two in the same direction as the CME, starting below the HACXS altitude, and just a few tens of arcseconds northward of it\footnote{http://sprg.ssl.berkeley.edu/$\sim$shilaire/movies/20021126\_js/}. The X-ray source then settled on a mostly radial course at $\approx$2 km/s, before ceasing to be observed by RHESSI around 01:10 on 2002 November 27. This velocity is in very good agreement with the velocity of the rising post-CME loop systems observed with $EIT$ (Figure~\ref{fig:tp}). The HACXS stays well above ($\sim$0.1 $R_S$) the EUV loop system throughout the observations. \subsubsection{X-ray source size and shape:} \begin{figure*}[ht!] \centering \includegraphics[height=16cm]{f4.ps} \caption{High altitude coronal X-ray source characteristics at different energies, derived from RHESSI visibilities accumulated over five minute intervals: flux, altitude above photosphere, distance to $25^{\circ}$ radial; 2D Gaussian FWHM, eccentricity and orientation with respect to solar equator. For clarity, error bars were omitted, but the scatter of the points is a good approximation. Information on the 8--9 and 9--10 keV bands has been omitted before 18:00 and 20:00 UT, repectively, because of their weak fluxes. } \label{fig:visfittingtimeprof} \end{figure*} Using RHESSI visibilities \citep[a new software method akin to radio visibilities, see e.g.][]{Schmahl2003}, source size, shape, and position were determined and are displayed in Figure~\ref{fig:visfittingtimeprof} at different energies. Higher energies tend to be at higher altitude, suggesting that the hottest plasma is at higher altitude, as would be expected if the reconnection X-point already flew past our region of interest (see Section~\ref{sect:discussion} and Appendix~\ref{appendix:epslocation}). Source size does not vary remarkably during the 12-hours interval. The shape of the HACXS stays generally elongated, with the higher energies having the tendency for higher eccentricities. \subsubsection{Spectroscopy:} \label{sect:spectroscopy} \begin{figure*}[ht!] \centering \includegraphics[width=16.6cm]{f5.ps} \caption{ RHESSI spectra obtained using different methods. {\it Black data points}: RHESSI imaging spectroscopy with SC 8, accumulated between 2002/11/26 20:30:24 and 20:35:24 UT, with vertical error bars. The horizontal error bars actually correspond to the bin widths. {\it Red data points}: Average of all 5-minute imaging spectroscopy spectra (using SC 8) from 2002/11/26 20:00 to 24:00 UT. {\it Blue data points}: Imaging spectroscopy with SC 8, using the sum of all 5-minute images between 2002/11/26 20:00 and 24:00 UT. Each image has been shifted in accordance with the source motion. {\it Purple data points:} Spatially-integrated spectroscopy between 2002/11/26 20:30:24 and 20:35:24 UT. Isothermal fitting yields 9.5 MK and 2.3$\times$10$^{47}$ cm$^{-3}$. {\it Green data points:} visibility-derived (shifted phase-centers) between 2002/11/26 20:10 and 2002/11/27 01:10 UT. Most methods yield unreliable values beyond $\approx$15 keV ({\it black vertical dashed line}), where background countrate is typically an order of magnitude above the source countrate. } \label{fig:spectroscopy} \end{figure*} \begin{figure}[ht!] \centering \includegraphics[width=8cm]{f6.ps} \caption{Five-hour long RHESSI image, from 2002/11/26 20:10 to 2002/11/27 01:10 UT, in the 4--8 keV energy band. RHESSI visibilities were generated in five minute accumulations, then bundled together after shifting their phases to remove source motion effects (relative to the source position at 20:09:48 UT). } \label{fig:longexp} \end{figure} Spectroscopy of our high-altitude X-ray source was done. In Figure~\ref{fig:spectroscopy}, full-Sun spectroscopy {\it (purple)} was done using the OSPEX {\it Solarsoft} suite of routines. Background selection and subtraction is a delicate process, particularly for that event, as disc flares occurred during the 12-hours time interval that the HACXS was observed. Hence, imaging spectroscopy was also employed: It is less sensitive than full-Sun spectroscopy, but does provide the inherent ability of removing background effects. The best spectrum from imaging was obtained by making 5-minute long images over five hours (2002 November 26 20:10 to 2002 November 27 01:10 UT), adding them together (rebinning and shifting for source motion), and determining fluxes at different energies (i.e. spectrum) using back-projected maps with subcollimator 8 \citep{Hurford2002}: the noise level is typically 4 times smaller than from a spectrum obtained by simply adding together the spectra from each 5-minute accumulations. As an additional check, visibility-based imaging spectroscopy was also employed: the visibilities were phase-shifted to remove smearing from the source motion over five hours. The results are displayed as green data points in Figure~\ref{fig:spectroscopy}, and a long-integration image of our event is shown in Figure~\ref{fig:longexp}. The spectra displays no clear non-thermal (power-law) emission at high energies. Fitting an isothermal component below 15 keV to the 5-hours back-projected data (blue data points in Figure~\ref{fig:spectroscopy}) yield a temperature of $T$=11.4 MK and emission measure $EM$=1.4$\times$10$^{47}$ cm$^{-3}$. For comparison, 5-minutes long accumulation on 2002 November 26 around 20:30 UT yields $T$=11.5 MK and 1.2$\times$10$^{47}$ cm$^{-3}$ with imaging and $T$=9.5 MK and 2.3$\times$10$^{47}$ cm$^{-3}$ with spatially-integrated spectroscopy (Figure~\ref{fig:spectroscopy}, {\it black} and {\it purple} data points). The lower temperature and higher emission measures obtained from RHESSI full-Sun spectroscopy are probably due to the presence of a low-energy flux component with large spatial extend, the likely residuals from previous disc flares. The even lower temperatures and higher emission measures measured by GOES ($T$=7.5 MK, $EM$=10$^{48}$ cm$^{-3}$) at the same time and throughout this event (Figure~\ref{fig:tem}) are typically attributed to the GOES response \citep[see e.g.][]{Holman2003}, which is more sensitive to lower temperature plasmas. \begin{figure*}[ht!] \centering \includegraphics[width=10cm]{f7.ps} \caption{ Black crosses: RHESSI full sun spectroscopy. Red squares: RHESSI imaging spectroscopy (SC 8). Solid black line: GOES temperature and emission measure measurements. The short duration ($\sim$minutes) peaks observed by GOES are due to disc flares, unrelated to the HACXS. } \label{fig:tem} \end{figure*} \section{Discussion} \label{sect:discussion} \subsection{X-ray source timeline, position, and morphology} The HACXS first appeared some 80 Mm above the footpoints of a loop system, about 1.5 hours after a flare located in these footpoints erupted. It progressed generally outwards, its intensity rising and then decreasing over the course of $\approx$2 hours. From $\approx$17:00 to $\approx$19:30, EIT observed material being formed (e.g. a cusp feature and a filament feature) and expelled (e.g. a flux rope, and other ejecta). These ejections seem to have disturbed the HACXS: it jumped about 60 Mm in altitude, and its flux started increasing again. The height and velocity profiles of the HACXS and the EUV loop system support the picture of a looptop (or ``above the looptop'') reconnection point that moves upwards, heating the local plasma to X-ray emitting temperatures, before they cool down and are later seen in EUV, giving the impression that the EUV loops trail the X-ray source in space, when in fact they are trailing in time. Apart from its very long duration, the source altitude profile of the HACXS is very similar to the observations reported by \citet{Gallagher2002}, including the higher energies being located at slightly higher altitudes than the lower energies. This further supports the scenario that hotter plasma is located at higher altitude (see Appendix~\ref{appendix:epslocation} for a simple justification, and Section~\ref{sect:epslocationmodeling} for an attempt at modeling it), consistent with the \citet{Kopp1976} model and the \citet{Svetska1987} observations. After $\approx$21:00, the high-eccentricity (elongation) and orientation of the 2D Gaussians fitted to the HACXS (as shown in the two bottom plots of Figure~\ref{fig:visfittingtimeprof}) are consistent with the geometry of a CS-like feature that extends radially outward. The HACXS ceased to be observed by RHESSI around 01:10 UT on 2002 November 27. This is due to both it having decreased in intensity to near or below RHESSI's sensitivity and the intense flaring activity that started at that time and lasted several hours. As observed in Figure~\ref{fig:goes}, the GOES baseline (i.e. non-flaring level) after 2002 November 27 $\approx$01:10 UT is ill-determined, because of the intense flaring activity. But it is conceivable that the HACXS remains present until 2002 November 29 $\approx$00:00, as the GOES X-ray flux in both GOES channels never drops back to pre-event levels until then. \subsection{Energy-position relationship} \label{sect:epslocationmodeling} The peak emissions at different energies are slightly displaced (Fig.~\ref{fig:visfittingtimeprof}). In fact, \citet{Sui2003} and \citet{Liu2008} have observed similar behavior with RHESSI X-ray data: they observed the centroid position of successively higher energies to be located at higher altitude, and, then the trend reversed. They have attributed this behavior to a hot CS located at the position where the trend reversed. Another possibility to explain the spatially-displaced energies is the presence accelerated particles which, much as in \citet{Brown2002}, are being stopped at larger distances (column densities) the larger their initial energies are. Although the greater elongation observed at high energies supports that scenario \citep[see e.g. Appendix A of][]{PSH2009}, the absence of non-thermal radiation (discussed in Section~\ref{sect:nth}) in the spatially-integrated spectrum clearly dispels that hypothesis. We have considered different models of temperature and emission measure profiles (see Appendix~\ref{appendix:epslocation}) to model the emission of different energies at different positions. We have attempted to fit our RHESSI data (visibilities accumulated from 2002/11/26 20:10 to 2002/11/27 01:10, and phase-shifted to remove source motion smearing) with the last two models mentioned in Appendix~\ref{appendix:epslocation}: the first one, using exponential profiles (with altitude) for both temperature and emission measure, yielded very poor results. The second one, where the temperature profile was assumed Gaussian, and the emission measure profile remained exponentially decreasing with altitude, yielded better results: The best-fitting parameters were $T_0$=119 MK, $H_T$=38.1'', $H_{EM}$=657'', $z_0$=1221.7'', and reduced $\chi^2$=0.33 (Figure~\ref{fig:tgrad2}, {\it black}). Such a high temperature seems highly unlikely. The synthetic X-ray spectrum computed from a plasma with such a temperature distribution (Figure~\ref{fig:tgrad2}, {\it} right) is clearly not observed. The height scales are loosely compatible with the typical density height scales found in the corona (about 100''). Given that our error bars are rather large (partly explaining the good chi-squared despite the obviously too-high temperature), we have tried for comparison to fix the temperature $T$ at 10 MK, and redo the fitting process. We found $H_T$=4115'', $H_{EM}$=64525'', $z_0$=1232.8'', and reduced $\chi^2$=0.9 (Figure~\ref{fig:tgrad2}, {\it gray}). The corresponding synthetic spectrum is more in accordance with our observations, but these new rather large scales heights are somewhat unexpected, and would mean that the densities along the HACXS (and possibly the CS) change very slowly with altitude. A full exploration of the model space and other fitting techniques are beyond the scope of this paper, but will be addressed in a subsequent one. \begin{figure*}[ht!] \centering \includegraphics[height=7cm]{f8.ps} \caption{ {\it \bf Left:} {\it Crosses:} Positions of emission at different energies, with error bars (vertical lines) and bin widths (horizontal lines). {\it Solid black line:} Fitting to the Gaussian $T$ and exponential $EM$ profiles. {\it Solid gray line:} Same, with $T$ fixed to 10 MK. {\it \bf Right:} {\it Crosses:} Synthetic X-ray spectra generated using the fitting parameters found. {\it Solid line:} Fit to the crosses in the 4--10 keV band. Color scheme same as in the plot on the left. See text for more details. } \label{fig:tgrad2} \end{figure*} \subsection{Energetics} \subsubsection{Thermal energy in RHESSI source} \label{Sect:th} We now want to estimate how much power and energy are needed to maintain the X-ray source at such high temperature for such a long time. On 2002 November 26, around 20:30, RHESSI observations indicate the source has an emission measure of about 2$\times$10$^{47}$ cm$^{-3}$, and a source size of $\approx$100'' FWHM, leading to a source volume $V$=2$\times$10$^{29}$ cm$^3$ (assuming HACXS has a spherical shape). Using $n=\sqrt{EM/V}$, one obtains an electron density $n$=10$^9$ cm$^{-3}$, and a total number of electrons $nV$=2$\times$10$^{38}$ electrons (incidentally, a typical number for total flare-accelerated electrons found in HXR footpoints of large flares). With the temperature $T \approx$10.5 MK (average of the 9.5 MK and 11.5 MK found in Section~\ref{sect:spectroscopy}), this leads to a radiative loss timescale of $\tau_{rad}\approx$8$\times$10$^{4}$ s. The half loop-length from the chromospheric footpoint to the HACXS is $L\approx$$\pi/2 \times H$, with the source height $H$=220'', from which the conductive loss timescale $\tau_{cond} \approx$8$\times$10$^2$ s can be derived \citep[see e.g.][]{AschwandenBook2004} (we assumed energy is lost mainly to the chromospheric heat sink, and not to interplanetary space via open field lines). Conductive losses are more important than radiative losses, and we will hence use the former to estimate the power $P$ required to maintain the temperature of the high altitude X-ray source at around 10 MK, using: \begin{eqnarray} P &=& \frac{E_{th}}{\tau_{cond}} = \frac{3kTnV}{\tau_{cond}} \\ &\approx& 6 \times 10^{35} keV/s \approx 1 \times 10^{27} erg/s, \end{eqnarray} where $k$ is Boltzmann's constant, and $E_{th}$ is the thermal energy content of the source ($\approx$9$\times$10$^{29}$ erg around 20:30 UT). I.e. over 12 hours, about 4$\times$10$^{31}$ ergs must have been deposited in the source. \subsubsection{Non-thermal energy} \label{sect:nth} If the power $P$ calculated in the previous paragraph came exclusively from accelerated particles as they dump all their energy into heating the plasma, could it be that their associated non-thermal emission is so weak as to be unobservable by RHESSI? A photon power-law with spectral index $\gamma \approx$8 and flux at 10 keV $F_{10} \approx$3$\times$10$^{-2}$ photons s$^{-1}$ cm$^{-2}$ keV$^{-1}$ is an order of magnitude below the observed (thermal) X-ray emission (Figure~\ref{fig:spectroscopy}), and could be concealed by it. The characteristics of the electron distribution corresponding to such a hypothetical non-thermal photon emission can be determined assuming either {\it thick-target} or {\it thin-target} assumptions: The column density traversed $N$=$n L$=2.5$\times$10$^{19}$ cm$^{-2}$ stops injected electrons with start energy below 11 keV. $N$= $n \sqrt[3]{V}$= 6.3$\times$10$^{18}$ cm$^{-2}$ stops injected electrons with start energy below 5.6 keV. I.e. we are very near a thick target at the energies where we have observed emission. Using Eq.~\ref{eq:nth:thick:2} of Appendix~\ref{appendix:brm} to obtain the low-energy cutoff value $E_C$ that equates thermal and non-thermal energies, we obtain $E_C$=6.3 keV. A plausible value, although flares have never been observed to go that low \citep[partly because thermal emission usually blocks any attempts at such observations, see e.g.][]{Holman2003,Kontar2008}. We find the total number of injected electrons in this case to be $F_{tot}=8.5 \times 10^{34}$ electrons per second above 6.3 keV. This injection rate is at least an order of magnitude below typical large flare values \citep{Holman2003}. Assuming accelerated electrons escape (which need not be the case), this rate implies the HACXS must be replenished every $\approx$40 minutes, far from the ``flare number problem'', where the acceleration region is estimated to be replenished sometimes as fast as every few tens of seconds \citep[see e.g.][]{Miller1997}. A photon spectral index of $\gamma \approx$8 is not very flare-like, although not unlike what \citet{Liu2008} have found in coronal sources during the impulsive phase of a flare. Smaller $\gamma$ are permissible, but will decrease $E_C$ correspondingly, in order to conserve $P$ to the same required amount. $E_C$ cannot be below $\approx$1 keV, the thermal temperature of the plasma: at these energies, electrons would essentially be indistinguishable from the local thermal plasma, and not contribute energy to a non-thermal beam of electrons \citep{Emslie2003}. For $E_C$=1 keV, a $\gamma$=3.2 non-thermal power-law would conserve the injected non-thermal power and still be concealed below the thermal emission. While it is to be noted that low-energy cutoffs below 10 keV have so far never been reliably observed, the heating of the HACXS over 12 hours purely by accelerated electrons cannot be firmly contradicted by our observations. \subsection{Connection between RHESSI observations and UVCS observations} There is a small overlap in time when both RHESSI observes the bottom of the CS (at an altitude of about 0.3 $R_S$), and UVCS can make a reliable temperature diagnostic of the CS at 0.7 $R_S$: On 2002 November 27, between 00:00 and 01:10, both RHESSI and UVCS sources indicate temperatures of $\approx$8 MK. If one infers that this temperature is constant between these two altitudes, one can estimate the total amount of (thermal) energy contained within this region. Assuming a rectangular sheet with $\approx$100" width, $\approx$0.4 $R_S$ length, $\approx$10$^4$ km thickness \citep[as assumed by][]{Bemporad2006}, then one gets a volume $\approx$2$\times$10$^{29}$ cm$^3$. Assuming the average electron density to be the geometric mean between what is found by RHESSI at 0.3 $R_S$ ($\approx$10$^9$ cm$^{-3}$) and what is found by UVCS at 0.7 $R_S$ \citep[$\approx$ 7$\times$10$^7$ cm$^{-3}$, according to][]{Bemporad2006}, one finds a total of $\approx$6$\times$10$^{37}$ electrons, for a total thermal energy content of $\approx$4$\times$10$^{37}$ keV, or $\approx$6$\times$10$^{28}$ erg, i.e about 7\% of the instantaneous thermal energy found in the HACXS (and about 0.1\% of the total energy that must have been injected in the HACXS over 12 hours). Hence, it is conceivable that the energy that powers the CS comes from the HACXS region, e.g. via heat conduction, and not only via magnetic reconnection in the CS. On the other hand, the fact that the HACXS starts before the CME ($\approx$13:40 vs. $\approx$17:00), the fact that \citet{Bemporad2008}'s non-thermal turbulent reconnection model explains well the UVCS observations, and the fact that the expelled material observed in EUV appears to flow beside the EUV loop system and the HACXS, still leaves the question open as to whether the CME/CS and the HACXS are significantly tied together. \section{Summary and Conclusion} UVCS observations in the wake of a CME that started around 2002 November 26 17:00 UT show hot plasma (initially well over 8 MK) at 0.7$R_S$ above the photosphere, for 2.3 days (at which point it had cooled down below $\approx$ 3.5 MK and was no longer observed). This hot plasma was interpreted as the signature of current sheet material. X-ray observations during the same time interval show enhanced X-ray emissivity throughout that period, albeit at lower altitudes ($\lesssim$0.3 $R_S$). For 12 hours, RHESSI observes a thermal coronal source that is near the base of the curent sheet, and, as it has at least an order of magnitude more thermal energy, it could technically provide the heat for the CS (i.e. in this scenario, reconnection and plasma heating occur mostly near the looptops rather that in the CS). On the other hand, heat and energy transport through a turbulent environment (as is probably the CS) can be quite complex and slow (usual plasma coefficients must be replaced by effective ones: anomalous heat conductivity), and as turbulent reconnection in the CS \citep{Bemporad2008} explains elegantly the CS temperature, it is possible that the coronal X-ray source and EIT looptop system are only weakly related to the CS/CME. We used novel long-accumulation imaging spectroscopy techniques to better estimate the photon spectrum and we have fitted it with an isothermal component. The RHESSI source temperature peaked at 10--11 MK. The emission measure of this source essentially increased during this whole 12-hours period, reaching above 5$\times$10$^{47}$ cm$^{-3}$. We have also observed an energy vs. position displacement in the emission from this HACXS, consistent with a plasma that has a Gaussian profile for its temperature distribution with altitude. Because of the lack of observed non-thermal emission, it appears unlikely, though not impossible, that the heating in the HACXS is due to particles being accelerated in it.
\section{Motivation} Launched into a low earth orbit in 2008, the \emph{Fermi} Gamma-Ray Space Telescope consists of two instruments, the Large Area Telescope (LAT)~\cite{atwood} which uses a pair-production system of detection and the Gamma-Ray Burst Monitor (GBM)~\cite{meegan} which consists of 14 scintillation detectors. Observing the entire unocculted sky, the GBM has an energy range of 8 keV-40 MeV which overlaps with the lower end of the LAT's range, 20 MeV-300 GeV. \emph{Fermi} has an inclination of 26.5$^{\circ}$, an altitude of $\sim$565 km and a period of $\sim$ 96 minutes. The primary observation mode of is Sky Survey Mode, this optimises the sky coverage of the LAT whilst maintaining near uniform exposure. In this mode the satellite rocks about the zenith ($\pm$50$^{\circ}$ formerly $\pm$35$^{\circ}$) such that the entire sky is observed for $\sim$30 minutes every 2 orbits ($\sim$ 3 hours). In addition to this rocking the satellite pointing alternates between the northern and southern hemispheres each orbit. Due to the fact that \emph{Fermi}'s instruments are deactivated in the South Atlantic Anomaly (SAA), which is primarily in the southern hemisphere, there is an exposure differential of $\sim$15 \% between observations in the north and south hemispheres. The GBM was optimised for the study of the prompt emission from GRBs, which is characterised by impulsive peaks with sharp rises, often highly structured, and easily distinguishable against instrumental backgrounds. The timescale on which this emission typically occurs is usually short enough that the background can be modelled as a polynomial of order 0-4. However, this method is not suited to resolving smoother long lived emission. \begin{figure*}[t] \centering \includegraphics[scale=0.32]{./lc_1416.pdf} \includegraphics[scale=0.32]{./lc_30.pdf} \caption{Estimated background and source rates for detector NaI B ($\sim$35-1000 keV) determined using $\pm$14,16 (left panel) and $\pm$30 (right panel) orbits for region A. In each panel the upper plot is the source (yellow) and background (blue) rates and the lower is the residual rates. In both cases the estimated rates agree well with the source rates.} \label{lc} \end{figure*} The LAT has observed long lived emission on the order of ks from GRBs, which follows the prompt phase and usually decays as a power law (e.g. for GRB090926A following the prompt emission in the LAT ($\sim$25 s) the flux decayed as power low ($t^{-1.7}$)~\cite{ackerman11}). To investigate whether such emission is detectable in GBM and also to confidently determine the background for solar flares, we have implemented a method for estimating the background which uses the rates from adjacent days, when the satellite is at the same geographical coordinates, to estimate the background at the time of interest. This project has been motivated by the work of Connaughton~\cite{val_tail}, who employed a similar technique with BATSE. \section{Method} The orbit of \emph{Fermi} is such that it will be at approximately the same geographical coordinates every 15 orbits ($\sim$ 24 hrs). Therefore, it would seem that the background at time $T_0$ could be approximated by averaging the rates at times $T_0$ $\pm$15 orbits. However, this is not possible as the rocking angle of the spacecraft in sky survey mode is the same every 2 orbits and as a result detectors which are pointed towards the source at $T_0$ will not be pointed towards it at $T_0$ $\pm$15 orbits. One solution to this is to use the rates from $T_0\pm$30 orbits. An alternative is to use the average of the rates from $T_0\pm$14 orbits and $T_0\pm$16 orbits to approximate the rates from $T_0$ $\pm$15 orbits. An obvious limitation of this technique is that it cannot be employed to investigate GRBs for which the satellite accepted an Autonomous Repoint Request (ARR). When an ARR is trigged the telescope will slew so that the GBM calculated position is within the LAT FOV. A natural consequence of this is that the periodic pointing is interrupted for the duration of the ARR ($\sim$2 hours, formerly 5 hours). This is unfortunate as it means that in general we cannot use our method to search for extended emission in GRBs with LAT detections where there is extended emission in the MeV-GeV range, as these will usually trigger an ARR. An additional issue is the passage of the satellite through the SAA which can cause elevated rates in the detectors due to activation, particularly in the BGO. The time spent in the SAA varies from orbit to orbit due to the precession of \emph{Fermi} ($\sim$52 days). This can lead to a systematic deviation between the estimated background rates and the source rates (see Section~\ref{saa_sec}). \section{Blank Sky Tests} In order to test the validity of the method, a blank sky test was performed. Between May 2009 and April 2011 four regions with triggerless periods of $\sim$4 days were selected and used as pseudo sources (see Table \ref{tab_details}). An additional criteria in the selection of the blank sky regions was that the regions offset $\pm$14,15,30 orbits did not contain and were not preceded by SAA passages. \begin{table}[h] \begin{center} \caption{Blank Sky test details: The four regions were selected with the criteria that they correspond to a period of $\sim$ 4 triggerless days and that the regions offset $\pm$14,15,30 orbits from the zero time did not contain SAA passages.} \begin{tabular}{c c c c } \hline \textbf{Region} & \textbf{Date} & \textbf{Zero Time (MET)} & \\ A & 09/07/28 & 270470679 & \\ B &10/08/23 &302922999 & \\ C &10/08/10 &324800173 & \\ D &11/04/18 &304200725 &\\ \hline \end{tabular} \label{tab_details} \end{center} \end{table} In order to determine which temporal selection provided the background most similar to the actual rate, the background was estimated from $T_0\pm$30 orbits and the average of $T_0\pm$14,16 orbits for all regions. The background was estimated for a duration of 3500 s using the continuous CSPEC data (128 energy channels and temporal resolution of 4.096 s). Sample lightcurves for detector NaI b for region A can be seen in Figure~\ref{lc}. \begin{figure*}[t] \centering \includegraphics[scale=0.3]{./nai_residuals.pdf} \includegraphics[scale=0.3]{./bgo_residuals.pdf} \caption{Residual Rates for all NaI (left panel) and all BGO (right panel) detectors in all regions for $\pm$30 (blue) and $\pm$14,16 orbits (grey). } \label{hist} \end{figure*} \begin{figure*}[t] \centering \includegraphics[scale=0.32]{./b1_1416-saa_activation.pdf} \includegraphics[scale=0.32]{./b1_30-saa_activation.pdf} \caption{Estimated background and source rates for detector B1 (0.1-50 MeV) determined using $\pm$14,16 (left panel) and $\pm$30 (right panel) orbits. In each panel the upper plot is the source (yellow) and background (blue) rates and the lower is the residual rates. The systematic offset between the source and background rates arises due to activation following SAA passage. This offset is much lessened for $\pm$30 orbits. } \label{activation_lc} \end{figure*} The systematic difference between the estimated rates and the actual rates was examined by histogramming the rate residuals from all the regions separately for the NaI and BGO. In order to limit the effect of interfering sources, the low energy channels of the NaI ($<$25 keV) were not included. Excluding also the overflow channel gives an energy range of $\sim$25-1000 keV. For the BGO only the first and overflow channels were discarded, corresponding to an effective energy range of $\sim$0.1-45 MeV. The results can be seen in Figure~\ref{hist}. For each offset ($\pm$30 and $\pm$14,16) the histogrammed residuals were fit with gaussians, the parameters of which can be seen in Table~\ref{fit_details}. The fit parameters are quite similar for $\pm$30 and $\pm$14,16 for both NaI and BGO which implies that both offsets are equally valid for estimating the background. \begin{table}[t] \begin{center} \scriptsize \caption{Parameters from gaussian fit to histogrammed residual rates, all variables have units of counts/s. The fit parameters are quite similar for both $\pm$30 and $\pm$14,16 orbits. } \begin{tabular}{c c c c c cc} \hline & & $\pm$30 & & &$\pm$14,16& \\ &\textbf{Amplitude} & \textbf{Mean} & \textbf{Sigma} &\textbf{Amplitude} & \textbf{Mean} & \textbf{Sigma} \\ NaI &3087 & -0.45 & 15 & 3970 & -0.9 & 14.6 \\ BGO &399 & -0.26& 38 & 518 & -2.5 & 37 \\ \hline \end{tabular} \label{fit_details} \end{center} \end{table} \subsection{Effect of SAA passage}\label{saa_sec} Inspection of the residual lightcurves for a region preceded by an SAA passage showed that the background estimated from $\pm$30 orbits more closely matched the observed rates following an SAA passage, than that derived from $\pm$14,16. This is demonstrated in Figure~\ref{activation_lc}. This is to be expected as the time spent in the SAA for the $\pm$30 orbit offset will be a closer match for the passage time for the source region. For regions where the effect of the SAA passage is negligible the rates from $\pm$30 and $\pm$14,16 orbits closely match the observed rates. This is beneficial as for these regions they can be used interchangeably, allowing the background to be estimated with $\pm$30 orbits if there is an interfering source in $\pm$14,16 orbits and vice versa. \section{Conclusion} Attempting to study long-lived or non-impulsive smooth emission in a background limited instrument in GBM is challenging. In order to do so a method using the rates from adjacent days to estimate the background has been developed. For the 4 triggerless periods studied this method generates a background which closely matches the source rates. Our study has shown that the rates from $\pm$30 and $\pm$14,16 orbits can be used interchangeably to estimate the background at the time of interest unless there is an SAA passage exit close to the time of interest, in which case the rates from $\pm$30 should be used. \subsection{Acknowledgments} \label{Ack} \begin{acknowledgments} This work has been suppourted by a Marie Curie European Reintegration Grant within the 7th Program under contract number PERG04-GA-2008-239176 and by the European Space Agency/Enterprise Ireland. \end{acknowledgments}
\section{Introduction} Most of the mass in the Interstellar Medium (ISM) is in neutral form. Photo-dissociation regions (PDRs) are neutral regions of the ISM at the interface between the stars and opaque cores of molecular clouds, where the heating and chemistry are regulated by the penetrating far ultraviolet (FUV) photons of the ionising source(s) \citep[e.g.][]{hol99, ber05}. The gas in the PDRs cools mainly via the emission of far-infrared (FIR) fine structure lines such as \ion{[C}{ii]}, \ion{[O}{i]}, and molecules (e.g. CO, H$_2$). The intensity of these lines depends on local conditions in the gas, and when compared to PDR models can be used to trace the spatial evolution of the physical conditions of the gas across the PDRs. Due to its proximity \citep[415pc,][]{men07} and edge-on orientation, the Orion Bar has been subject to extensive studies in the literature, where it is usually adopted as the prototypical PDR template in the study of high mass star forming regions, and is often used to test PDR models \citep[e.g.,][]{pel09}. It is one of the brightest PDRs with an FUV radiation field at the ionisation front of G$=$1-4$\times$10$^{4}$ G$_{\rm 0}$ \citep{tie85,mar98}, where G$_{\rm 0}$ is the mean interstellar radiation field \citep[1.6$\times$10$^{-6}$ W~m$^{-2}$,][]{hab68}. \citet{tie93} showed how the Bar presented a layered distribution of Polycyclic Aromatic Hydrocarbons (PAHs), H$_2$, and CO. This distribution is thought to be the result of an extended gas component seen edge-on, of average gas density between 10$^4$-10$^5$ cm$^{-3}$. Evidence of clumps have also been suggested to explain high density and temperature tracers \citep[e.g.,][]{bur90,vdw96}. Recently, \citet{rub11} presented mid-IR Spitzer observations that extend beyond the Bar, from 2.1 to up to 12\hbox{$^\prime$}.1 from the ionising star $\theta$$^1$ Ori C. They detect ionised material (Ne$^{+2}$, IP of 41eV) all the way up to the boundary of their observations. They also find evidence for a decrease in electron density and an increase of the PDR tracers ([Si II] 34.8 $\mu$m, [Fe II] 26.0 $\mu$m, and molecular hydrogen) as the distance from $\theta$$^1$ Ori C increases. In the FIR, \citet{her97} mapped the Orion region with a 22$\hbox{$^{\prime\prime}$}$ to 55$\hbox{$^{\prime\prime}$}$ beam using the Kuiper Airborne Observatory (KAO). Their derived column densities in the Bar for O$^0$ and C$^+$ were in agreement with an edge-on geometry. In this paper, we present the first maps of the \ion{[C}{II]} (158$\mu$m), \ion{[O}{I]} (63 and 145$\mu$m), and \ion{[N}{II]} (122$\mu$m) lines of the Orion Bar region from the recently launched Herschel Space Observatory \citep[HSO,][]{pil10}. With its access to FIR and superb spatial resolution at these wavelengths, the HSO allows us to study the spatial distribution of these lines in unprecedented detail and so improve our understanding of the ionised and neutral interface. These observations enable us to trace the spatial evolution and excitation conditions of the gas across the illuminated interface of the Bar. This paper is part of a Herschel study of the Orion Bar by our group and is complemented by Herschel/SPIRE studies of the CO and CH$^+$ line emission \citep{hab10,hab11,nay10}, and dust emission \citep{ara11}. The paper is organized as follows. The observations, data reduction and line measurements are given in Sect.~2. In the following section (Sect.~3) the spatial morphology of the lines is discussed. In Sect.~4 the distribution and correlation of the lines are analysed. Sect.~5 presents a model tailored to reproduce the distribution of the lines. Discussions on the origin of the \ion{[C}{II]} emission and the cooling are presented in Sect.~6 and Sect.~7. Finally, the conclusions are summarized in the last section. \section{Observations and Data Reduction} \begin{figure}[!t] \begin{center} \includegraphics[width=8cm]{./f1.eps} \end{center} \caption{Overlay of the \ion{[O}{I]} 63$\mu$m PACS observation on a Spitzer/IRAC 8$\mu$m image of the Orion Bar. A 4$\times$4 raster map - 16 overlapping footprints - was performed (see Sect. 2). An example of a footprint with its 5$\times$5 {\em spaxels} is illustrated on the top left of the figure together with their dimensions. The map covers the region before and behind the Bar (toward $\theta$$^1$ Ori C).} \end{figure} The observations were taken using the PACS instrument \citep{pog10} on board the HSO on 24 February 2010. These observations are part of the {\em Evolution of the Interstellar Medium} guaranteed time key project (observation ID=1342191152) \citep{abe10}. Four fine-structure lines were targeted: \ion{[C}{II]} at 158$\mu$m, \ion{[O}{I]} at 63 and 145$\mu$m, and \ion{[N}{II]} at 122$\mu$m. At these wavelengths, and in increasing order, PACS provides a spectral resolution of 3420, 1050, 1170, and 1270 respectively (v$\sim$90-285 km~s$^{-1}$). The observations were taken during Science Demonstration Phase (SDP) in the - now decommissioned - wavelength switching mode. To trace the PDR and H\,II region interface a 4x4 raster map was performed. Each raster position (or footprint) is composed by 5$\times$5 spatial pixels referred to as {\em spaxels}. For each spaxel the line is observed in 16 different spectral scans, each with an up and down scan. The configuration at the time of the observation is shown in Figure~1, where the raster map at the epoch of observation is overlaid on top of an 8~$\mu$m IRAC image of the Orion Bar. We note that an additional map in the direction of the Trapezium stars and overlapping the current observation is currently scheduled for observation and will be presented in a future paper. Given the brightness of the lines, the minimum exposure configuration of one cycle and repetition per line was performed. For these maps the best sampling is achieved using Nyquist sampling, which we adopted for the lines measured in the red channel (\ion{[C}{II]} at 158$\mu$m, \ion{[O}{I]} 145$\mu$m, and \ion{[N}{II]} at 122$\mu$m). This consists of a raster point and raster line step of 24\hbox{$^{\prime\prime}$}.0 and 22\hbox{$^{\prime\prime}$}.0 respectively. This results approximately in steps of 2/3 and 1/2 of the slit size along both directions. For the \ion{[O}{I]} 63$\mu$m line located in the red channel, Nyquist sampling is achieved with 16\hbox{$^{\prime\prime}$}.0~ and 14\hbox{$^{\prime\prime}$}.5 raster point and line steps. However, to minimise exposure time, and given the strength of the line, the same step size as the lines for the red channel was adopted. An {\em off} position, about 40\hbox{$^\prime$}~from the center of the map, was taken at the beginning and the end of the observation (not shown in the figure). This observation is not subtracted from the data as the purpose of the wavelength switching mode is to cancel out the background by determining a differential line profile. The {\em off} observation, however, confirmed that there was no background contamination to any of the lines studied. The data were processed using version 6.1.0 of the reduction and analysis package HIPE \citep{ott10}. From level 0.5 to level 1 the standard procedures were followed. From this level on the cubes were further processed, using proprietary tools, to correct for drifting and for flux misalignments between scans. Drifting is caused by temperature deviations in the telescope which can cause the signal to be modified over time. They can be easily identified by comparing the up and down scans. To correct for this effect we first passed a median filter in the time domain to avoid glitches. The continuum was then fitted using a linear fit\footnote{In our data the signal was either linear or slightly curved. A polynomial fit of second degree produced very similar results.} to characterise and remove the drift. The flux misalignment between the spectral scans can be due to improper dark subtraction or flatfield correction. The former will have an additive effect, and the latter a multiplicative one. In our case, the scans could be well aligned using an additive value which indicates that in our observations the cause of the flux misalignments is the dark subtraction. We note that the lines are very strong and these corrections were minor. At this point the cubes are exported and we used the IDL-based software PACSman \citep{leb11} to measure the line fluxes (by fitting a Gaussian) and create the final map. In this code, the line fluxes are measured for all the spaxels independently. The lines are not spectrally resolved. To create the final map, PACSman recreates an oversampled pixelated grid of the observations (with 3$\hbox{$^{\prime\prime}$}$ pixel resolution) and calculates the average fractional contribution of the given spaxels to the relevant position. The code calculates the statistical uncertainties on the fly, which include the dispersion in the reduction process and the RMS of the fit. These uncertainties are small and usually amount to less than 4\%, 1\%, 3\% for the \ion{[C}{II]} and \ion{[O}{I]} lines. For the weaker \ion{[N}{II]} 122$\mu$m line, these are higher and oscillate between 5-13\%. The relative accuracy between spaxels given in the manual is 10\%\footnote{From the PACS spectroscopy performance and calibration manual. This can be found at http://herschel.esac.esa.int/twiki/bin/view/Public/PacsCalibrationWeb}, and for the remaining of the paper we have adopted the higher of this value or the uncertainty given by PACSman. We note that absolute flux calibration is quoted to be 30\% of the peak-to-peak accuracy (scatter around the expected flux densities). The final maps for each of the four lines are shown in Figure~2. We have compared our \ion{[C}{II]} and \ion{[O}{I]} line fluxes with those measured by \cite{her97} with the Fabry-P\'erot interferometer FIFI and integrated over a 55\hbox{$^{\prime\prime}$}~beam. Their measurements are in good agreement (within $\sim$30\%) with those reported in this paper when convolved to their larger beam size. \label{obs_s} \begin{figure*}[!ht] \begin{center} \includegraphics[width=18cm]{./f2.ps} \end{center} \caption{PACS contour maps of the \ion{[C}{II]} (158$\mu$m), \ion{[O}{I]} (63 and 145$\mu$m), and \ion{[N}{II]} (122$\mu$m) lines in units of 10$^{-6}$ W~m$^{-2}$~sr$^{-1}$. In increasing order of wavelength the beams sizes are 4\hbox{$^{\prime\prime}$}.5, 8\hbox{$^{\prime\prime}$}.8, 10\hbox{$^{\prime\prime}$}, and 11\hbox{$^{\prime\prime}$}. Note the offset from zero in the intensity scale to emphasize the detail.} \end{figure*} \begin{figure*} \begin{center} \includegraphics[width=16.cm]{./f3.ps} \end{center} \caption{Image and contour plots for different combination of lines and 8$\mu$m band as indicated in the title of each panel. All maps have been convolved to the PSF at 158$\mu$m and are in flux units of 10$^{-6}$ W~m$^{-2}$~sr$^{-1}$ (except for the IRAC image which is in 10$^{3}$MJy~sr$^{-1}$. Here the intensity scales start at zero.} \end{figure*} \section{Spatial Distribution} The observed spatial distribution of the lines is shown in Figure~2, where the ionising star $\theta^1$ Ori C is illuminating the region from the top right, outside the maps (see also Fig.4a for a sketch of the different regions). The beams sizes are represented in the top right of each map and correspond, in increasing order of wavelength, to 4\hbox{$^{\prime\prime}$}.5, 8\hbox{$^{\prime\prime}$}.8, 10\hbox{$^{\prime\prime}$}, and 11\hbox{$^{\prime\prime}$}. At the distance of the Bar (415~pc), 10$\hbox{$^{\prime\prime}$}$ corresponds to a physical scale of 0.02~pc. The Bar\footnote{For the remainder of the paper, the Bar is defined as the region where the emission is higher than 75\% of the peak emission for carbon. This threshold falls between the third and fourth highest contours in Figure~3 and is represented by the solid line in Figure~4a. Similarly, we use the same definition for the discussion in Sect. 6 about the \ion{[N}{II]} peak.} is resolved in all four lines and we detect emission all over the region probed, including detection of \ion{[N}{ii]} (ionised gas) behind the Bar. The range of intensities is not large, with variations by factors 4 (\ion{[C}{II]} and \ion{[N}{II]}) to 8 (\ion{[O}{I]}) from the peak (Bar) to the fainter (outer) regions. The overall morphology of the region can be best discussed from the \ion{[O}{I]} 63$\mu$m map because its smaller point spread function (PSF) provides the highest detail (Figure~2). Even if this line is more affected by self-absorption, the gain in resolution bests this caveat. The emission peaks at about 123$\hbox{$^{\prime\prime}$}$ SE of $\theta^1$ Ori C. The most striking feature is the presence of several knots of enhanced emission resulting in small scale structures, which suggests a clumpy distribution within the Bar. These knots, three in the south-west and the two in the north-east, are bridged by weaker emission, $\sim$16\% lower in flux than the knots. In the \ion{[O}{I]} 63$\mu$m image, the smaller knots have diameters between $\sim$6$\hbox{$^{\prime\prime}$}$ and $\sim$10$\hbox{$^{\prime\prime}$}$ which, at the distance of the Orion nebula, correspond to 0.01-0.02 pc. The smaller ones (6\hbox{$^{\prime\prime}$}) are marginally resolved and could result from the superposition of even smaller clumps. These knots could be clumps being photo-evaporated by the intense FUV penetrating the PDR. \citet{gor02} studied the effects of the radiation field on the evolution of clumps in PDRs. Using their relation between the size of the clumps and the column density of the FUV heated region (N$_0$=2$\times$10$^{21}$ cm$^{-2}$, their Eq.~34), we derive a density of the gas at the base of the photo-evaporating flow of 6-10$\times$10$^{4}$ cm$^{-3}$. This value is consistent with the inter-clump density we use in Sect. 5 for the modelling. Knowing that the density inside the clumps is higher than this, and assuming that the observed size is that of an isolated clump, we estimate that the photo-evaporation timescale of these clumps should be higher than $\sim$4\,500-7\,500 yr, depending on size. This morphology is somewhat mimicked in the \ion{[O}{I]} 145$\mu$m and \ion{[C}{II]} 158$\mu$m maps, where the lower resolution has washed out the clumps of emission. As we move away from the Bar (both in front and behind) the emission decreases gradually. There is, however, a confined region of brighter emission in the north of the map (middle top of the \ion{[O}{I]} 63$\mu$m map), probably due to an increase in density in the PDR which lies behind the HII region. Alternatively this could be the result of a geometrical effect where the inclination of the background PDR is steeper. This region is also revealed in the other three maps, albeit with less contrast. The \ion{[O}{I]} 63$\mu$m map also points to an increased emission at the western edge of the Bar and extending north. This excess was also detected in CO emission from the ground \citep[e.g.,][]{lis98}. The morphology of the \ion{[C}{II]} line follows very well that of the \ion{[O}{I]} lines (especially that of \ion{[O}{I]} 145$\mu$m) with the Bar peaking at the same position in the three maps. This is better seen in Figures~3c and 3d where the images have been convolved to the same resolution as the \ion{[C}{II]}158$\mu$m map (largest PSF), and are shown with over-plotted contours. In Figure~3a we also see that the PAH emission, as traced by the 8$\mu$m IRAC band, is close to the \ion{[C}{II]} peak but slightly shifted towards the ionising star $\theta$$^1$ Ori C. The peak of \ion{[C}{ii]} emission in the top-left corner of the Bar is, however, not followed as well by the PAH emission. In Figure~3d we can see that the different intensity levels in the 145$\mu$m map delineate the Bar very well, whereas in the 63$\mu$m map the spread in the levels is broader. This is probably the result of the higher optical depth on the 63$\mu$m line, where changes in column density have a less pronounced effect than for the 145$\mu$m line. The peak of the \ion{[N}{II]} emission is displaced with respect to the other lines by about 12$\hbox{$^{\prime\prime}$}$~towards to $\theta^1$ Ori C (Fig. 2 and Fig. 3b). This is expected because with an ionisation potential of 14.5~eV the \ion{[N}{II]} line comes from the ionised phase and is representative of the ionisation front. There is a small region of overlap between the \ion{[N}{II]} and \ion{[C}{II]} lines but this could be due to an orientation effect (Bar slightly tilted). We even note that the north-east tip (top-left) of the Bar that is seen in the \ion{[C}{II]} 158$\mu$m and \ion{[O}{I]} 145$\mu$m maps (and also 8$\mu$m IRAC map, Figs. 1 and 2), delineates the boundary of emission of the \ion{[N}{II]} line in that region. On the western side of the Bar, the \ion{[N}{II]} emission is somewhat delineated by the above-mentioned emission extending north in the \ion{[O}{I]} 63$\mu$m map and also detected in CO and dust emission. Finally, the gap in the center of the Bar, seen in the \ion{[O}{I]} and \ion{[C}{II]} lines, coincides with the maximum emission of the \ion{[N}{II]} 122$\mu$m line in front of it (Fig. 3b). This could indicate that the radiation field in this specific spot is higher (or more intense) and has been able to photo-ionise the gas. Alternatively this could be due to variations in the column density. \section{Line Correlations and Ratios} \begin{figure*} \begin{center} \includegraphics[width=9.3cm,angle=90]{./f4.ps} \end{center} \caption{Intensity plots for different combination of lines fluxes (in 10$^{-6}$ W~m$^{-2}$~sr$^{-1}$) and ratios from the convolved maps. The different regions are indicated in the first panel and are labelled in the third, where blue represents points in front of the Bar and red behind the Bar. Plus symbols indicate points within the Bar(CII). The square data point is the position where the cooling is calculated (see Sect. 7). The reference points for the adopted cut (Sect. 5) are plotted in circles and include error bars to give an indication of the uncertainties at different regions. In panels {\em c} and {\em g} the solid lines represent respectively a linear fit ($y=-2.28+1.06x$), and a polynomial fit for the fainter region outside the Bar ($y=6.37-1.53x+0.12x^2$).} \end{figure*} \begin{figure*} \begin{center} \includegraphics[width=16cm]{./f5.ps} \end{center} \caption{Map ratios for different combination of lines as indicated in the title of each panel.} \end{figure*} For the purpose of the discussion we have divided the map into several regions which are represented in Figure~4a and annotated in panel {\em c} of the same figure. Summarizing: data points in front of and behind the Bar are colour-coded in blue and red, respectively; plus symbols indicate points that trace the Bar in \ion{[C}{II]} (those inside the solid contour line), and dots points outside the Bar in \ion{[C}{II]}. In addition, selected points along an adopted cut (see \S5) have been labeled (orange) and include error bars. Figure~4b plots the \ion{[N}{II]} versus the \ion{[C}{II]} line fluxes. The points before and behind the Bar (blue and red respectively) occupy different regions but there are no clear trends. \citet{abe05} hinted, from their models, at a correlation between these two lines in the ionised medium. The intricate behaviour in our figure does not suggest that. On the other hand, Figure~4c illustrates that the \ion{[C}{II]} 158$\mu$m line follows very well the \ion{[O}{I]} 145$\mu$m line. A linear fit gives a slope of 1.059$\pm$0.005. Variation of the \ion{[O}{I]} 145$\mu$m line is dominated by the column density of the gas, and so this correlation indicates that most of the \ion{[C}{II]} comes from the PDR and not the HII region (see Sect. 6). The \ion{[O}{I]} 63$\mu$m and \ion{[C}{II]} 158$\mu$m lines show also a good correlation (Fig.~4d) but it seems that the correlation outside the Bar(CII) differs with respect to the other regions. This is also seen in Figure~4e where the two oxygen lines are plotted against each other. The points behind the Bar (red dots) are displaced compared to those before the Bar (blue dots), and there is a change of slope between points in the Bar (plus signs) and outside the Bar (dots). The excitation conditions of these lines are similar and therefore this behaviour could reflect the effect of opacities that more strongly affect the \ion{[O}{I]}63$\mu$m line (as it becomes self-absorbed at rather low column densities). In the following panels we plot the ratio of several lines which, in an optically thin environment, allows us to remove the effect of column density in the trends: \begin{itemize} \item[$\bullet$] Figure~4f shows that the \ion{[C}{II]}(158$\mu$m)/\ion{[N}{II]}(122$\mu$m) ratio is different before and behind the Bar. A map of this line ratio is illustrated Figure~5b where we clearly distinguish two regions, with the points before the Bar, which are more representative of the ionised gas, having a lower ratio. We find that a threshold of 15 for this ratio could be used to distinguish between emission from the neutral and ionised region. \item[$\bullet$] The \ion{[C}{II]}158$\mu$m/\ion{[O}{I]}145$\mu$m ratio in Figure~4g decreases as we approach the Bar. The trend in the figure of the points outside the Bar (dots) can be approximated by a polynomial fit. This behaviour is again evident in the map ratio shown in Figure~5a where points before and behind the Bar show the similar values. This suggests similar physical conditions in these two regions because the excitation parameters for both lines are different. This is not expected for a single PDR and suggests contamination of a background PDR (see Sect. 5) Another important point is that within the Bar the ratio is very homogeneous (within 16\%) but it differs from the regions outside the Bar. We can use the \ion{[C}{II]}158$\mu$m/\ion{[O}{I]}145$\mu$m ratio to put a constraint on the densities by comparing with the predicted values from PDR models \citep{fer98,kau99}. We find densities that vary from 10$^4$-10$^5$cm$^{-3}$ in the Bar, and 10$^3$-10$^4$cm$^{-3}$ outside it. \item[$\bullet$] In the last panel Figure~4h, it is clear that the line ratio \ion{[O}{I]}63$\mu$m/\ion{[O}{I]}145$\mu$m splits before and behind the Bar (being lower before the Bar). Given that the excitation conditions in these regions must be similar (Fig. 4g), we conclude that the observed difference is the result of \ion{[O}{I]} 63$\mu$m becoming self-absorbed; in the Bar, as it is expected, the ratio is smaller because the column density is higher than outside the Bar. We note that the gas in the region where the \ion{[O}{I]}63$\mu$m line is optically thick must still be warm to excite the \ion{[O}{I]}145$\mu$m line (E$_u=$326~K). From Figure~5c we see that the ratio is indeed not homogeneous within the Bar (as it was for the \ion{[C}{II]} and \ion{[O}{I]} 145$\mu$m, Fig.~5a), indicating that opacity effects are present. \end{itemize} \section{Modelling} \begin{figure} \begin{center} \includegraphics[width=9cm,angle=90]{./f6.ps} \end{center} \caption{Observed (plus symbols) and modelled (solid and dotted lines) profiles along the cut shown in Figure~4 (see Sect. 5) measured with respect to the distance to the exciting star $\theta$$^1$ Ori C. Observed and modelled fluxes are convolved to the 158$\mu$m PSF, where the dotted line includes an offset due to background emission (see text). The first four panels compare for each line the observed convolved profiles with a Cloudy model, where this has been scaled to the peak of the \ion{[O}{I]} (145 $\mu$m) emission. The bottom panel compares the normalized model profiles. In this panel, the solid dots represent the position of the reference points in the adopted cut (Fig. 4a).} \end{figure} We have modelled the line emission across the Bar using the radiative transfer code Cloudy \citep{fer98}. This code computes the chemistry and radiative transfer at the surface of a molecular cloud (in our case assumed to be a plane-parallel semi-infinite slab) illuminated by FUV photons from the ionising star(s). We adopt a Kurucz model for the star at 39\,600~K \citep{pel09}, and ISM abundances \citep{sav96,mey97,mey98}. In order to reproduce the spatial stratification we adjust the starting point of the HII region and the density. Namely, the ionised region is adjusted to start at 0.134~pc (67$\hbox{$^{\prime\prime}$}$) from the star. The adopted gas density is set to a constant value of 3200~cm$^{-3}$ \citep{pel09} in the ionised region, and is then coupled with a profile density for the PDR as described in \citet{ara11}, with a density scaled to 6$\times$10$^4$ cm$^{-3}$ at the peak of the \ion{[O}{I]} and \ion{[C}{II]} and emission (projected distance of 0.246~pc). No tilting is assumed for the Bar and the depth of the PDR along the line of sight is adjusted to 0.35~pc. To illustrate the comparison between the observed fluxes and the model (convolved to the 158$\mu$m PSF) we assume the cut illustrated in Figure~4a. This cut is also adopted for the discussion in our accompanying papers by \citet{hab11} and \citet{ara11}. It is made to pass through the exciting star $\theta^1$ Ori C (5h 35m 16.46s, -5d 23m 23.17s) and an arbitrary point in the Bar (5h 35m 21.82s, -5d 24m 59.18s). The chosen point in the Bar minimizes effects of other contaminating stars in the direction of the cut that can affect the dust \citep{ara11}, and allows us to sample different conditions in the region. In Figure~6 the solid line represents the output of the model, and the dotted line the model plus an offset (see point 3.) From the comparison in Figure~6 we find that: \begin{enumerate} \item This simple model does a very good job in reproducing the observed spatial stratification of the peak positions of the \ion{[N}{II]} line at 111.5$\hbox{$^{\prime\prime}$}$, and the \ion{[C}{II]} and \ion{[O}{I]} lines at 123.5$\hbox{$^{\prime\prime}$}$ (top and middle panels). This stratification is the result of the attenuation of the incident radiation field across the Bar and is sensitive to the density profile. \item The model also shows a gradual increase of the \ion{[N}{II]} line as we approach the Bar (better seen in the bottom panel) which is also seen in the observations, while in contrast the oxygen lines show a more abrupt increase. This is the result of the ionisation structure. The model also shows a gradual increase for the \ion{[C}{II]} line but this is not clear from the observations. This could be the result of it being more easily excited (low temperature and density) compared to the oxygen lines and/or contribution of the \ion{[C}{II]} line from the ionised phase. \item It is striking that the observed profiles do not fall to zero as the model (solid lines) does both before and after the Bar. This background emission is significant, amounting to about $\sim$18\% of the peak emission for the \ion{[O}{I]} 145$\mu$m line and $\sim$38\% for the other three lines. These offsets have been added to the model emission as the dotted lines in the Figure. This suggests the presence of another PDR(s), probably from the background cavity in the molecular cloud, which must produce the base of emission that we measure. This PDR is probably not entirely face-on because the strong intensity observed, which may result from a limb brightening effect, cannot be reproduced by a face-on PDR model. \item The model line profiles are narrower than the observations. This could be due to the Bar being tilted to the observer ($\lesssim$10$\hbox{$^\circ$}$), an effect already inferred in other studies \citep[e.g.,][]{pel09}. Alternatively, a highly structured medium would allow FUV radiation to permeate the region and heat gas on larger spatial scales. \item In order to reproduce the absolute flux, and assuming that the base emission from the additional PDR(s) has an additive effect on top of the emission of the Bar (dotted line), we need to assume a depth of the PDR in the model of 0.3~pc. The depth given above is larger than, but comparable to, the width of the Bar ($\sim$0.06~pc) as expected for a tilted edge-on Bar. \item In this model we take into account optical depth effects across the PDR\footnote{We assume both thermal and micro-turbulence line width with a turbulence velocity of 3~km~s$^{-1}$.} but not along the line of sight. However, we can see that the model (with offset) reproduces the peak of emission of the \ion{[N}{II]} 122$\mu$m, \ion{[O}{I]} 145$\mu$m fairly well, and the \ion{[C}{II]} 158$\mu$m within factor 1.2. This indicates that, except for the \ion{[O}{I]} 63$\mu$m line, the emission of these lines are not dominated by optical effects. \end{enumerate} A more detailed modelling of the region including additional PDR/s (even the effects of clumpiness and tilting) and dust emission is outside our scope here but will be presented in a future paper. Still, it is encouraging to see how a simple model (single edge-on PDR) can be used to at least reproduce with a fair degree of success the observations of such a complex environment. \section{[CII] from the PDR and HII region} In the ISM, the \ion{[C}{II]}158$\mu$m line is important in the study of the cooling and chemistry of PDRs. In extragalactic studies it is important for redshift determinations, and for the extent to which its luminosity is a measure of the star formation rate \citep[SFR,][]{sta91, mei07,luh03}. With a low ionisation potential of 11.3~eV, this line can originate both in the PDR and in the HII region. It is thus important to characterise in detail its contribution in different environments. Because of the high radiation field and high density in the Orion nebula, one expects that most of the \ion{[C}{II]} should originate from the PDR. The PACS maps show in great detail the distribution of the \ion{[C}{II]} and \ion{[N}{II]} lines where we can investigate this issue. \begin{figure} \begin{center} \includegraphics[width=7.0cm,angle=90]{./f7.ps} \end{center} \caption{Line intensity ratios for different combinations of the \ion{[N}{II]} (122 and 205$\mu$m) and \ion{[C}{II]} 158$\mu$m lines. Theoretical curves are shown by the solid, dotted, and dashed lines. Grey areas indicate the range of observed ratios, and the enclosing boxes represent the range of uncertainties in the ratios. The square point is the ratio of the \ion{[N}{II]} 122/205 lines.} \end{figure} We have therefore quantified the contribution of \ion{[C}{II]} 158$\mu$m and \ion{[N}{II]} 122$\mu$m to the Bar as traced by the \ion{[C}{II]} emission (Bar(CII)), and as traced by the \ion{[N}{II]} emission (Bar(NII)). These are represented as areas enclosed by the solid and dashed lines, respectively, in Figure~4a. This combined region, which we name Bar(CII$+$NII), encompasses the peak emission of the PDR and ionised region. We find that 76\% of the \ion{[C}{II]} emission is coming from the PDR region (Bar(CII)) relative to the total emission in the Bar(CII+NII). The remaining 31\% comes from the Bar(NII) region\footnote{These quantities do not add to 100\% because there is an overlapping region as it is explained in Sect. 3.}. This 31\% contribution to the Bar(NII) region could include emission from the ionised region or from the additional PDR (see Sect. 5). To estimate this, and ignoring confusion from tilting effects, we can use the close relation of the \ion{[C}{II]} line with the \ion{[O}{I]}145$\mu$m line. For a broad range of densities between 10$^3$ and 10$^5$ (typical of PDRs), the \ion{[C}{II]}/\ion{[O}{I]}145$\mu$m ratio varies between 0.5 and 2 \citep{fer98,kau99,lep06}. The observed \ion{[O}{I]}145$\mu$m in the Bar(NII) is about 2$\times$10$^{-6}$ W~m$^{-2}$~sr$^{-1}$, from which we predict a PDR \ion{[C}{II]} emission between 1 to 4$\times$10$^{-6}$ W~m$^{-2}$~sr$^{-1}$. We can then compare this range to our observed \ion{[C}{II]} emission of 4$\times$10$^{-6}$ W~m$^{-2}$~sr$^{-1}$ in this region. Thus, from the original 31\% of \ion{[C}{II]} in this region, we deduce that either all could come from the additional PDR, or the HII region could contribute to at most 24\%. Following the analysis by \citet{obe11} in the Carina nebula, we have compared the expected \ion{[C}{II]}/\ion{[N}{II]} ratio in the ionised medium with the observed values in the Bar(NII) region. This can be used to get an additional estimate of the contribution from the ionised region to the \ion{[C}{II]} line. To do this we have complemented our observations with the SPIRE-FTS data of the \ion{[N}{II]}205$\mu$m line \citep{hab11}, and have calculated the theoretical curves for the \ion{[N}{II]}122$\mu$m/\ion{[N}{II]}205$\mu$m, \ion{[C}{II]}158$\mu$m/\ion{[N}{II]}205$\mu$m, and \ion{[C}{II]}158$\mu$m/\ion{[N}{II]}122$\mu$m ratios at a temperature of 9000~K \citep{nie11}. The \ion{[C}{II]}/\ion{[N}{II]} line ratios depend on the relative ionic abundance of CII and NII. We use the carbon and nitrogen abundances adopted in Sect.~5, coupled with the ionisation fraction of \ion{C}{II} and \ion{N}{II} for the HII conditions in the Orion nebula as given by \citet{rub85}, to derive a relative ionic abundance of \ion{[C}{II}/\ion{N}{II]}$\sim$1.6. These theoretical curves are shown in Figure 7\footnote{The \ion{[C}{II]}158$\mu$m/\ion{[N}{II]} (122, 205$\mu$m) curves can be applied to any other object by simply scaling them to the ratio between the relative abundance of \ion{[C}{II}/\ion{N}{II]} in the object and that assumed here (1.6).}. The \ion{[N}{II]}122$\mu$m/\ion{[N}{II]}205$\mu$m ratio can be used to determine the electron density. The observed ratio (square in the figure) falls in the non-linear regime of the theoretical curve (solid line) but can serve to obtain an upper limit in the density of 1000 cm$^{-3}$, consistent with our adopted value of 3200 cm$^{-3}$ \citep{pel09}. The observed \ion{[C}{II]}158$\mu$m/\ion{[N}{II]}205$\mu$m, and \ion{[C}{II]}158$\mu$m/\ion{[N}{II]}122$\mu$m line ratios, grey areas in the figure, are clearly above the theoretical curves (dotted and dashed lines respectively). Taking into account the uncertainties in the line fluxes (enclosing boxes in the figure), the observed ratios indicate that no more than 9\% of the \ion{[C}{II]}158$\mu$m comes from the ionised region. Allowing for an uncertainty of about a factor 2 in the relative abundance of \ion{C}{II}/\ion{N}{II} and collisional strengths used to calculate the theoretical curves, we can safely conclude that in the Bar less than 18\% of the \ion{[C}{II]} emission originates in the ionised medium. This is a slightly lower threshold than our previous estimate. \citet{abe05} made some modifications to the Cloudy code to, among others things, estimate the contribution of several PDR lines from the H\,II region. They apply their model to the starburst galaxy NGC\,253 where they find that about 30\% of the \ion{[C}{II]} comes from the ionised medium. More recently, \citet{moo11} using PACS data have mapped the FIR emission of the \ion{[C}{II]} 158$\mu$m line in spiral galaxy M33. They find that between 20 to 30\% of this emission comes from the ionised medium. Their values are slightly higher than our determinations, which is not surprising since we are probing dense PDR in the Orion Bar while in these galaxies lower density HII regions could dominate the emission. We note, however, that this comparison is hampered by the difference between the larger scales (where different regions are mixed) that these studies probe, and the small scales probed in our study. \section{Cooling} In order to establish the role of the \ion{[C}{II]} and \ion{[O}{I]} lines in the cooling of the PDR we have compared their contribution to that of other relevant cooling lines in the mid- and FIR. We have included in this calculation the H$_2$ rotational lines from the ISO/SWS, and $^{12}CO$ rotational lines (J=4-3 to J=21-20), $^{13}CO$ (J=5-4 to J=14-13), H$_2$O, and CH$^+$ from the PACS and SPIRE instruments \citep{hab10,job11} . To make this comparison we use the same position (5h35m21s, -5\hbox{$^\circ$}~25\hbox{$^\prime$}~18\hbox{$^{\prime\prime}$}) of the H$_2$ measurements by ISO to derive the contribution of each line (or cascade of lines). This position is indicated by the white square in Figure~4a. For this comparison we convolve all the data to the largest beam size (SPIRE 40\hbox{$^{\prime\prime}$}). We find that in this region the \ion{[C}{II]} and \ion{[O}{I]} lines studied in this paper contribute 90\% of the total power emitted by all these lines, with the \ion{[O}{I]} 63$\mu$m line contributing 72\% of this emission. The CO, H$_2$, and CH$^+$ contribute, respectively, 5, 4, and less than 1\%. These estimates accentuates the importance of the \ion{[O}{I]} 63$\mu$m line to the cooling budget from the gas lines in these regions. This fact is also highlighted by the relative strength of this line compared to the other lines in the region (see Fig. 6). \section{Summary and Conclusions} We have presented the first HSO observations of the \ion{[C}{II]}158$\mu$m, \ion{[O}{I]}63$\mu$m and 145$\mu$m, and \ion{[N}{II]}122$\mu$m lines of the Orion Bar. Its angular resolution has allowed us to map the spatial distribution of these lines in unprecedented detail. The \ion{[C}{II]} and \ion{[O}{I]} maps peak at the same position and fall close to the peak emission of PAHs (as traced by the Spitzer/IRAC 8$\mu$m band). The \ion{[N}{II]} peaks slightly closer to $\theta$$^1$ Ori C with a small region of overlap with respect to the other PDR lines. Within the Bar we can distinguish between knots (clumps) of emission, about 0.01-0.02~pc in size (6 to 10\hbox{$^{\prime\prime}$}), and which are 16\% higher in flux than the interclump medium. These clumps could be photo-evaporated by the FUV inside the PDR. These knots of emission are better seen in the \ion{[O}{I]}63$\mu$m map as this line offers the best resolution (having the smallest PSF), it is also seen in the \ion{[O}{I]}145$\mu$m and \ion{[C}{II]}158$\mu$m maps. The \ion{[C}{II]}158$\mu$m correlates very well with the \ion{[O}{I]} 145$\mu$m emission. The \ion{[C}{II]}158$\mu$m line does not correlate with the \ion{[N}{II]} in the ionised region as some studies have suggested. The combined information on the \ion{[N}{II]} and \ion{[O}{I]} lines provides a great diagnostic to estimate the emission from the \ion{[C}{II]}158$\mu$m line, and to distinguish between an origin in an ionised or neutral region. The ratio between the \ion{[O}{I]} 145$\mu$m/63$\mu$m lines show the effect of the opacities, where the \ion{[O}{I]} 63$\mu$m line becomes self-absorbed at high column densities. We have modelled the emission of the lines with the photo-ionisation code Cloudy and reproduce the relative position of the lines. The emission profiles reveal that, in addition to the Bar, there is a significant background emission all over the region (present for all four lines). This points to the presence of additional PDR(s). This should be follow up with detail radiative transfer models to infer the physical conditions and coupled to what has been learned from the emission of dust. The \ion{[C}{II]} line can come from the neutral and ionised medium. We have made different estimations of its contribution and find that most of the [CII] emission originates in the PDR(s) ($>$82\%). Using ancillary ISO and Herschel data we have calculated the total power emitted by the most relevant cooling lines from the mid- to FIR (\ion{[O}{I]}, \ion{[C}{II]}, CO, H$_2$O, CH$^+$). We show that the power emitted by the atomic \ion{[C}{II]} 158$\mu$m, and \ion{[O}{I]} 63$\mu$m and 145$\mu$m lines account for 90\% of the power emitted in the region by all of the cooling lines considered, with the \ion{[O}{I]} 63$\mu$m line contributing 72\% of the total. This emphasizes the predominant role of the latter in the cooling process from emission lines of these regions. \begin{acknowledgements} We would like to thank the referee, David Hollenbach for his comments and suggestions. JBS wishes to acknowledge the support from a Marie Curie Intra-European Fellowship within the 7th European Community Framework Program under project number 272820. HCSS, HSpot, and HIPE are joint developments by the Herschel Science Ground Segment Consortium, consisting of ESA, the NASA Herschel Science Center, and the HIFI, PACS and SPIRE consortia. JBS thanks E. Romano-Diaz for help with contour manipulation. We thank V. Lebouteiller for the use of the PACSman software and discussion on data reduction. \end{acknowledgements}
\section{Introduction} There has recently been interest in studying quantum field theories with a nontrivial infrared fixed point (IRFP), both in continuum and on the lattice. Under the renormalization group evolution, the coupling of these theories shows asymptotic freedom at small distances, analogously to QCD, but flows to a fixed point at large distances where the theory looks conformal. Such theories have applications in beyond Standard Model model building. These include unparticles, i.e. an infrared conformal sector coupled weakly to the Standard Model \cite{Georgi:2007ek,Georgi:2007si,Cheung:2007zza,Sannino:2008nv}, and extended technicolor scenarios, that explain the masses of the Standard Model gauge bosons and fermions via strong coupling gauge theory dynamics \cite{TC,Eichten:1979ah,Hill:2002ap,Sannino:2008ha}. In addition to direct applications to particle phenomenology, the phase diagrams of gauge theories, as a function of the number of colours, $N$, flavours $N_f$ and fermion representations, are interesting from the purely theoretical viewpoint of understanding the nonperturbative gauge theory dynamics from first principles. In figure \ref{phasediagram} we show a sketch of such phase diagram for SU($N$) gauge theory \cite{Sannino:2004qp}. In addition to the fundamental representation, the figure shows the phase structure in the cases of two-index (anti)symmetric and adjoint representations. The shaded regions in the figure depict the conformal windows for each of these fermion representations; below the conformal window the theory is in the chiral symmetry breaking and confining phase, while above the conformal window the theory is in the non-Abelian QED-like Coulomb phase. The upper boundary in each case corresponds to the loss of asymptotic freedom, i.e. to the value of $N_{f}$ where the one-loop coefficient $\beta_0=11/3 N-4/3 N_{f} T(R)$ of the $\beta$-function vanishes. The group theory factor $T(R)$ is defined for each representation as \begin{equation} \textrm{Tr}(T^aT^b)=T(R)\delta^{ab}. \end{equation} For fundamental (F), two-index symmetric (2S), two-index antisymmetric (2AS) and adjoint (A), the concrete values are, respectively, $T(F)=1/2$, $T(2S)=(N - 2)/2$, $T(2AS)=(N + 2)/2$ and $T(A)=N$. \begin{figure} \centering \includegraphics[scale=0.6]{SDphasediagram.eps} \caption{The phase diagram of SU($N$) gauge theory as a function of the number of colours, flavours and fermion representations (F = Fundamental, 2A = 2-index antisymmetric, 2S = 2-index symmetric, Adj = Adjoint). The shaded bands indicate the estimated conformal windows.} \label{phasediagram} \end{figure} For values of $N$ and $N_f$ inside the conformal window, the theory is expected to have a non-trivial IRFP. Near the upper boundary one expects the value of the coupling at the fixed point, $\alpha^\ast$, to be small and perturbation theory to be applicable \cite{Banks:1981nn}. However, as one moves deeper into the conformal window, by e.g. lowering $N_f$ at fixed $N$, the fixed point coupling grows, and nonperturbative methods are required for the analysis. In figure \ref{phasediagram} we have used the traditional estimate of evaluating the critical coupling for chiral symmetry breaking in the ladder approximation and setting it equal to the fixed point value of the two-loop coupling. In other words, at the lower boundary of the conformal window the fixed point coupling becomes super critical with respect to the critical coupling for the onset of chiral symmetry breaking. Explicitly, at two loops we have \bea \beta(g)\equiv\mu\frac{dg}{d\mu} &=&-\frac{\beta_0}{16\pi^2}g^3 - \frac{\beta_1}{(16\pi^2)^2}g^5, \nonumber\\ \beta_0 &=& \frac{11}{3}N-\frac{4}{3}T(R)N_f,\nonumber \\ \beta_1 &=& \frac{34}{3}N^2-\frac{20}{3}N\,T(R) N_f-4 C_2(R)T(R) N_f, \end{eqnarray} which implies that IRFP is at \begin{equation} \alpha^\ast=-\frac{\beta_0}{\beta_1}(4\pi). \end{equation} Above, $C_2(R)$ is the quadratic Casimir operator in representation $R$ and, for the representations we consider, it assumes values $C_2(F)=(N^2-1)/(2N)$, $C_2(2(A)S)=(N\pm 1)(N\mp 2)/N$ and $C_2(A)=N$. The estimated critical coupling on the other hand is $\alpha_c=\pi/(3C_2(R))$ \cite{Appelquist:1986an}. Solving $ \alpha^\ast=\alpha_c$, one obtains the lower boundary of the conformal window, denoted by dashed lines in figure \ref{phasediagram}. The phase diagram in the figure \ref{phasediagram} is therefore a conjecture which must be checked by nonperturbative analysis. Since currently the lattice simulations provide the only robust nonperturbative method for non-supersymmetric four dimensional gauge theories, such a check provides an interesting challenge for the lattice community. In this work we study SU(2) gauge field theory with $N_f =4, 6$ and $10$ massless flavours of fermions in the fundamental representation on the lattice using $O(a)$ improved Wilson-clover fermions. On the basis of two-loop beta-function six and ten flavour SU(2) theories may have a fixed point and be within the conformal window, while the four flavour theory is expected to confine. The model with ten fermions is chosen because it is close to the upper edge of the conformal window ($N_f=11$, where asymptotic freedom is lost) and is expected to have a fixed point at a rather small coupling. Thus, in this case we can compare the results from lattice simulations with those obtained from perturbative computations. The universal 2-loop $\beta$-function has a zero at $g^2 \approx 2.90$, and at 3 or 4 loops in the MS-scheme it is at $g^2 \approx 2.47$ or $2.52$, respectively \cite{vanRitbergen:1997va}. While the MS-scheme results cannot be directly compared with the Schr\"odinger functional results, the convergence indicates that the perturbative result should be reasonably accurate. In the model with six fermions, however, the fixed point determined by the two loop beta function is at much larger coupling, where we cannot expect the perturbation theory to hold. Indeed, the location of the fixed point varies significantly at different loop levels: using the MS-scheme $\beta$-function at 2, 3 and 4 loops, the location is $g_{\rm FP}^2 \approx 140$, $21$ and $30$, respectively. Such a large value of the fixed point coupling may be super critical with respect to the critical coupling for the onset of spontaneous chiral symmetry breaking; this is explicitly illustrated in figure \ref{phasediagram}, where the two colour and six flavour theory is already outside the conformal window. However, due to the nonperturbative nature of the lower boundary of the conformal window, one cannot exclude the possibility that the six flavour theory is very close or even within the conformal window. In addition, if we treat $N_f$ as a continuously variable parameter, the fixed point vanishes from the 4-loop $\beta$-function when $N_f$ is lowered only slightly from 6 to 5.945. In the proximity of this value the 4-loop $\beta$-function obtains a shape typical of walking coupling, i.e. the $\beta$-function has a local maximum with a small negative value. Clearly, these results imply that the six flavour case can be resolved only with non-perturbative calculations. On the other hand, the four flavour case is expected to be QCD-like, with confinement and chiral symmetry breaking. It is included here for comparison with the larger $N_f$ cases. On the lattice, SU(2) gauge field theory with $N_f > 2$ fundamental representation fermions has been recently studied in Refs.~\cite{Bursa:2010xr,Ohki:2010sr,Voronov}.% \footnote{% In related work, the existence of the infrared fixed point in SU(2) gauge theory with two adjoint representation fermions (which is of interest for technicolor model building) has been studied in \cite{Catterall:2007yx,Hietanen:2008mr, DelDebbio:2008zf,Catterall:2008qk,Hietanen:2009az, Bursa:2009we,DelDebbio:2009fd,DelDebbio:2010hx, DelDebbio:2010hu,Bursa:2011ru,DeGrand:2011qd}, and SU(3) gauge with various fermion representations and numbers of flavours in \cite{Damgaard:1997ut,Appelquist:2007hu,Appelquist:2009ty, Fodor:2009wk,Deuzeman:2008sc,Deuzeman:2009mh,Itou:2010we, Jin:2010vm,Hayakawa:2010yn,Hasenfratz:2010fi,Hasenfratz:2009ea, Shamir:2008pb,DeGrand:2008kx,DeGrand:2010na, Fodor:2009ar,Kogut:2010cz}.% } The study by Bursa {\em et al.} \cite{Bursa:2010xr} of SU(2) with six fundamental flavours suggested a possibility of an IRFP at much smaller coupling than expected from perturbation theory. However, this work used unimproved Wilson fermions which can be expected to be subject to large discretization errors. Indeed, as described below, the leading order perturbative analysis reveals that the running coupling measurement (step scaling using Schrödinger functional) using unimproved Wilson action can have finite cutoff effects of order 30-60\%, whereas the improved action reduces these to a few percent level using lattices presently within computational reach. In our simulations using $O(a)$ improved Wilson-clover fermions we indeed observe large deviations from the unimproved results. We measure the running coupling using the Schr\"odinger functional method in theories with four, six and ten flavours, and the mass anomalous dimension in the phenomenologically most interesting six flavour case. Unfortunately, in the six flavour case we are not able to fully resolve whether the fixed point exists, but the possible locations of the fixed point moves to significantly larger coupling ($g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 10$) than the unimproved lattice action results indicate. This paper is structured as follows: In section \ref{model} we introduce the model and describe the Schr\"odinger functional method, and present the perturbative step scaling function results. The lattice simulations and results are presented in section \ref{measurements} and in section \ref{conclusions} we conclude. \section{The model and the Schr\"odinger functional method} \label{model} The theory is defined by the action \begin{equation} S=S_G + S_F, \label{eq:action} \end{equation} where $S_G$ is the standard Wilson single plaquette gauge action for the SU(2) Yang Mills theory \begin{equation} S_G = \beta_L \sum_{x;\mu<\nu} \left (1 - \frac12 \textrm{Tr}\, [U_\mu(x) U_\nu(x+a\hat\mu) U^\dagger_\mu(x+a\hat\nu) U^\dagger_\nu(x) ] \right), \end{equation} with $\beta_L=4/g_0^2$. The part $S_F$ is the clover improved Wilson fermion action \begin{equation} S_F = a^4\sum_{\alpha=1}^{N_f} \sum_x \left [ \bar{\psi}_\alpha(x) ( i D + m_0 ) \psi_\alpha(x) + a c_{\rm{sw}} \bar\psi_\alpha(x)\frac{i}{4}\sigma_{\mu\nu} F_{\mu\nu}(x)\psi_\alpha(x) \right ], \end{equation} where $D$ is the standard Wilson-Dirac operator \begin{equation} D=\frac12 [\gamma_\mu(\nabla_\mu^* + \nabla_\mu ) - a\nabla_\mu^*\nabla_\mu], \end{equation} with the gauge covariant forward and backward lattice derivatives \begin{equation} \nabla_\mu\psi(x) =1/a[U_\mu(x)\psi(x+a\hat\mu) - \psi(x)], ~~~~ \nabla_\mu^*\psi(x) =1/a[\psi(x) - U^\dagger_\mu(x-a\hat\mu)\psi(x-a\hat\mu)]. \end{equation} The clover improvement term contains the usual symmetrized field strength tensor. We set the improvement coefficient $c_{\rm{sw}}$ to the $N_f$-independent perturbative value \cite{Luscher:1996vw} \begin{align} c_{\rm{sw}} = 1 + 0.1551(1) g_0^2 + O(g_0^4). \label{pertcsw} \end{align} While $c_{\rm{sw}}$ can be determined nonperturbatively \cite{Luscher:1996ug}, in our tests we observed that at strong lattice coupling the nonperturbatively determined $c_{\rm{sw}}$ is close to the perturbative one at $N_f=6$ and $10$. The situation is very different at $N_f=2$ \cite{Karavirta:2010ym,Karavirta:2011mv}, where $c_{\rm{sw}}$ diverges as $g_0^2$ increases. Thus, while the perturbative result (\ref{pertcsw}) does not give full cancellation of the $O(a)$ discretization effects, it is a concrete recipe which we expect to eliminate most of the $O(a)$ effects. We also include the perturbative improvement at the Schr\"odinger functional boundaries as described in \cite{Karavirta:2010ef,Karavirta:2011mv}. We measure the running coupling using the Schr\"odinger functional method \cite{Luscher:1992an,Luscher:1992ny,Luscher:1993gh,DellaMorte:2004bc}. The coupling is defined as a response to the change of the background field and the scale is set by the finite size of the lattice. We consider a lattice of volume $V=L^4=(Na)^4$. The spatial links at the $t=0$ and $t=L$ boundaries are fixed to \cite{Luscher:1992ny} \begin{align} U_\mu(\bar x,t=0) &= e^{-i\eta \sigma_3 a/L} \\ U_\mu(\bar x,t=L) &= e^{-i(\pi - \eta) \sigma_3 a/L} \end{align} with $\sigma_3$ the third Pauli matrix. The spatial boundary conditions are periodic for the gauge field. The fermion fields are set to vanish at the $t=0$ and $t=L$ boundaries and have twisted periodic boundary conditions to spatial directions: $\psi(x + L\hat i) = \exp(i\pi/5)\psi(x)$. At the classical level the boundary conditions generate a constant chromoelectric field and the derivative of the action with respect to $\eta$ can be easily calculated: \begin{align} \frac{\partial S^{\textrm{cl.}}}{\partial \eta} = \frac{k}{g^2_0}, \end{align} where $k$ is a function of $N=L/a$ and $\eta$ \cite{Luscher:1992ny}. At the full quantum level the coupling is defined by \begin{align} \ev{ \frac{\partial S}{\partial \eta} } = \frac{k}{g^2}. \end{align} To quantify the running of the coupling we use the step scaling function $\Sigma(u,s,L/a)$ introduced in \cite{Luscher:1992an}. It characterizes the change of the measured coupling when the linear size of the system is changed from $L$ to $sL$ while keeping the bare coupling $g_0^2$ (and hence lattice spacing) constant: \begin{align} &\Sigma(u,s,L/a) = \left. g^2(g_0^2,sL/a) \right |_{g^2(g_0^2,L/a)=u} \label{stepscaling}\\ &\sigma(u,s) = \lim_{a/L\rightarrow 0} \Sigma(u,s,L/a) \end{align} In this work we choose $s=2$. To obtain the continuum limit $\sigma(u,s)$ we calculate $\Sigma(u,s,L/a)$ at $L/a=6$ and $8$. Since we expect the discretization errors to be (mostly) removed to the first order in the improved action, we use quadratic extrapolation to find the limit $a\rightarrow 0$. The step scaling function is related to the $\beta$-function by \begin{align} -2\ln(s) = \int_u^{\sigma(u,s)} \frac{dx}{\sqrt x \beta(\sqrt x)}. \end{align} Close to the fixed point, where the running is slow and $|\beta|$ small, we can approximate the $\beta$-function by \begin{align} \beta(g) \approx \beta^*(g) = \frac{g}{2\ln(s)} \left ( 1 - \frac{\sigma(g^2,s)}{g^2} \right ). \label{eq:beta*} \end{align} The estimating function $\beta^*(g)$ is exact at a fixed point but deviates from the actual $\beta$-function as $|g-g^\ast |$ becomes large. We also measure the mass anomalous dimension $\gamma=d\ln m_q/d\ln\mu$ of the theory with 6 fermion flavours using the pseudoscalar density renormalization constant. The mass anomalous dimension is of interest for this case, because the possible infrared fixed point is at non-perturbative value of the coupling. It is defined on the lattice as \cite{DellaMorte:2005kg} \begin{align} Z_P(L) = \frac{\sqrt{3} f_1}{f_P(L/2)}, \end{align} where $f_1$ and $f_P$ are correlation functions of the pseudoscalar density \begin{align} f_1 &= \frac{-1}{12 L^6} \int d^3u d^3v d^3y d^3z \ev{\bar \zeta'(u)\gamma_5\lambda^a\zeta'(v)\bar\zeta(y)\gamma_5\lambda^a\zeta(z)},\\ f_P(x_0) &= \frac{-1}{12 L^6} \int d^3y d^3z \ev{\bar \psi(x_0)\gamma_5\lambda^a\psi(x_0)\bar\zeta(y)\gamma_5\lambda^a\zeta(z)}. \end{align} For these measurements the boundary matrices at $t=0$ and $t=L$ are set to unity; thus, separate simulations are needed. The mass step scaling function is then defined as in \cite{Capitani:1998mq}: \begin{align} \Sigma_P(u,s,L/a) &= \left. \frac {Z_P(g_0,sL/a)}{Z_P(g_0,L/a)} \right |_{g^2(g_0,L/a)=u} \label{Sigmap}\\ \sigma_P(u,s) &= \lim_{a/L\rightarrow 0} \Sigma_P(u,s,L/a), \end{align} and we choose again $s=2$. We find the continuum step scaling function $\sigma_P$ by measuring $\Sigma_P$ at $L/a=6,8$ and $10$ and doing a quadratic extrapolation. The mass step scaling function is related to the anomalous dimension by (see \cite{DellaMorte:2005kg}) \begin{equation} \sigma_P(u,s) = \left ( \frac{u}{\sigma(u,s)} \right ) ^{d_0/(2b_0)} \exp \left [-\int_{\sqrt u}^{\sqrt{\sigma(u,s)}} dx \left ( \frac{\gamma(x)}{\beta(x)} - \frac{d_0}{b_0 x} \right ) \right ], \label{massstep} \end{equation} where $b_0=\beta_0/(16\pi^2)$ in terms of the one-loop coefficient $\beta_0=22/3-2 N_f/3$ of the beta function and $d_0=3 C_2(F) g^2/(8\pi^2)=9/(32\pi^2)$ is the corresponding one-loop coefficient for the anomalous dimension, $\gamma=-d_0 g^2$. Close to the fixed point (\ref{massstep}) simplifies considerably. Denoting the function estimating the anomalous dimension $\gamma(u)$ by $\gamma^*(u)$, we have \begin{align} \log \sigma_P(g^2,s) &= -\gamma^*(g^2)\int_\mu^{s\mu} \frac{d\mu'}{\mu'} = -\gamma^*(g^2)\log s , \\ &\Rightarrow \gamma^*(g^2) = -\frac{\log \sigma_P(g^2,s)}{\log s }. \label{gammastar} \end{align} The estimator $\gamma^*(g^2)$ is exact only at a fixed point where $\beta(g^2)$ vanishes and deviates from the actual anomalous dimension when $\beta$ is large. The Schr\"odinger functional boundary conditions enable one to run at vanishing quark mass. The Wilson fermion action breaks the chiral symmetry explicitly and allows additive renormalization of the quark mass. We therefore find the quark mass from the PCAC relation \begin{align} aM(x_0) = \frac14 \frac{(\partial_0^* + \partial_0) f_A(x_0)}{f_P(x_0)} + c_A \, \frac a 2 \frac{\partial_0^* \partial_0 f_P(x_0)}{f_P(x_0)}. \end{align} We have used here the improved axial current \begin{align} f_A^I &= f_A + a \, c_A \, \frac{1}{2} (\partial_\mu^* + \partial_\mu) f_P, \\ f_A(x_0) &= \frac{-1}{12 L^6} \int d^3y d^3z \ev{\bar \psi(x_0)\gamma_0\gamma_5\lambda^a\psi(x_0) \bar\zeta(y)\gamma_0\gamma_5\lambda^a\zeta(z)}. \end{align} In this work use the perturbative expansion for the improvement coefficient $c_A$ \cite{Luscher:1996vw}: \begin{equation} c_A = -0.00567(1) C_2(F) g_0^2 + \mathcal{O}(g_0^4), ~~~~ C_2(F) = 3/4. \end{equation} The Schr\"odinger functional boundary conditions remove the zero modes that would normally make it impossible to run simulations at zero mass. We define $\kappa_c$ as the value of the parameter $\kappa$ where the mass $aM(L/2)$ vanishes. To find $\kappa_c$ we measure the mass at 3 to 7 values of $\kappa$ on lattices of size $L/a=16$ and interpolate to find the point where the mass becomes zero. We then use the same value of $\kappa_c$ on all lattice sizes. The values of $\kappa_c$ used in the simulations are given in table \ref{table:kappa}. In practice we achieve $|aM| < 0.01$. \begin{table} \centering \begin{tabular}{|l|l|l| c |l|l|l| c |l|l|l|} \cline{1-3} \cline{5-7} \cline{9-11} \multicolumn{3}{|c|}{$N_f=4$} & & \multicolumn{3}{|c|}{$N_f=6$} & & \multicolumn{3}{|c|}{$N_f=10$} \\ \cline{1-3} \cline{5-7} \cline{9-11} $\beta_L$ & $\kappa_c$ & $N_{traj}$ & & $\beta_L$ & $\kappa_c$ & $N_{traj}$ & & $\beta_L$ & $\kappa_c$ & $N_{traj}$ \\ \cline{1-3} \cline{5-7} \cline{9-11} 1.8 & 0.14162 & 108755 & & 1.39 & 0.144351377 & 215119 & & 1 & 0.14199 & 216356 \\ \cline{1-3} \cline{5-7} \cline{9-11} 1.9 & 0.139914 & 89064 & & 1.4 & 0.139914 & 221273 & & 1.3 & 0.13922 & 232495 \\ \cline{1-3} \cline{5-7} \cline{9-11} 2 & 0.138638 & 55031 & & 1.44 & 0.14350583 & 209124 & & 1.5 & 0.13762 & 100476 \\ \cline{1-3} \cline{5-7} \cline{9-11} 2.2 & 0.136636 & 13294 & & 1.5 & 0.142446 & 233273 & & 1.7 & 0.13627 & 99792 \\ \cline{1-3} \cline{5-7} \cline{9-11} 2.4 & 0.135205 & 23359 & & 1.8 & 0.1385229 & 25886 & & 2 & 0.13466 & 213290 \\ \cline{1-3} \cline{5-7} \cline{9-11} 3 & 0.132548 & 64035 & & 2 & 0.1367336 & 54710 & & 3 & 0.13146 & 99918 \\ \cline{1-3} \cline{5-7} \cline{9-11} 4 & 0.130326 & 58879 & & 2.4 & 0.1342875 & 33818 & & 4 & 0.12981 & 100328 \\ \cline{1-3} \cline{5-7} \cline{9-11} & & & & 3 & 0.132115 & 41221 & & 6 & 0.1282 & 98042 \\ \cline{1-3} \cline{5-7} \cline{9-11} & & & & 4 & 0.13014328 & 47419 & & 8 & 0.12739 & 99255 \\ \cline{1-3} \cline{5-7} \cline{9-11} & & & & 5 & 0.1290368 & 41440 & & & & \\ \cline{1-3} \cline{5-7} \cline{9-11} & & & & 8 & 0.1274578 & 7116 & & & & \\ \cline{1-3} \cline{5-7} \cline{9-11} \end{tabular} \caption{Parameter $\kappa$ used in the simulations at each $\beta_L=4/g_0^2$ and the number of measurements performed on the largest lattice.} \label{table:kappa} \end{table} \section{Improvement in perturbative analysis} The degree of improvement obtained with the clover term and boundary counter terms can be quantified with one loop order perturbative analysis of the step scaling function \nr{stepscaling}. Because the gauge action is identical with improved and unimproved fermions, to this order it is sufficient to consider only the fermion contribution to the step scaling. To one loop order the step scaling function \nr{stepscaling} can be expanded as \begin{eqnarray} \Sigma(u,s,L/a) &=& g^2(g_0,sL/a)\vert_{g^2(g_0,L/a)=u} \nonumber \\ &=& u+(\Sigma_{1,0}+\Sigma_{1,1} N_f)u^2. \end{eqnarray} The fermion contribution is denoted by $\Sigma_{1,1}$. To evaluate these perturbative contributions we use the methods in \cite{Sint:1995ch,Sommer:1997jg}, and choose $s=2$. The continuum limit of $\Sigma_{1,1}$ is given by the fermionic contribution to the one loop coefficient $b_{0}=\beta_0/(16\pi^2)$ of the beta function, i.e. \begin{equation} \lim_{L/a\rightarrow 0}\Sigma_{1,1}=2 b_{0,1}\ln{2}, \end{equation} where $b_{0,1}=1/(24\pi^2)$. In figure \ref{pertstep} we show $\delta\equiv \Sigma_{1,1}/(2 b_{0,1}\ln{2})$ both for unimproved Wilson fermions and with ${\cal{O}}(a)$ improvement. One immediately observes that without improvement, $\Sigma_{1,1}$ depends strongly on $L/a$ and approaches the continuum limit only for presently impractically large lattices, while with improvement the large lattice artefacts are absent. While this level of improvement is probably not preserved at higher orders, this nevertheless strongly motivates the use improved actions in the lattice studies of these theories with Wilson fermions. \begin{figure} \centering \includegraphics[scale=0.35]{Stepscaling_presentation.eps} \caption{Contribution of a massless Wilson quark to the step scaling function normalized to its continuum value at one loop order in perturbation theory. The top three curves show the result for gauge groups SU(4), SU(3) and SU(2) (from top to bottom) for unimproved Wilson fermions, while the lower three curves show the result after ${\cal{O}}(a)$ improvement has been taken into account.} \label{pertstep} \end{figure} \section{Simulations} \label{measurements} We use hybrid Monte Carlo (HMC) simulation algorithm with 2nd order Omelyan integrator \cite{Omelyan,Takaishi:2005tz} and chronological initial values for the fermion matrix inversions \cite{Brower:1995vx}. The trajectory length is 1, and the step length is tuned to have acceptance rate larger than 80\%. \begin{table} \centering \begin{tabular}{|l|l|l|l|l|} \hline $\beta_L$ & $L/a=6$ & $L/a=8$ & $L/a=12$ & $L/a=16$ \\ \hline 4 &1.2394(18)& 1.263(3) & 1.300(3) & 1.32(1) \\ 3 & 1.832(5) & 1.882(5) & 1.971(18)& 2.02(2) \\ 2.4 & 2.629(7) & 2.767(15)& 2.94(2) & 3.17(4) \\ 2.2 & 3.113(7) & 3.29(2) & 3.58(3) & 3.88(9) \\ 2 & 3.93(2) & 4.24(3) & 4.77(8) & 5.18(11)\\ 1.9 & 4.65(2) & 4.95(5) & 5.48(7) & 6.9(3) \\ 1.8 & 5.78(4) & 6.43(7) & 8.15(17) & 9.0(5) \\ \hline \end{tabular} \caption{The measured values of $g^2$ at each $\beta_L=4/g_0^2$ and $L/a$ with $4$ flavours of fermions.} \label{table:couplingnf4} \end{table} \begin{table} \centering \begin{tabular}{|l|l|l|l|l|l|} \hline $\beta_L$ & $L/a=6$ & $L/a=8$ & $L/a=10$ & $L/a=12$ & $L/a=16$ \\ \hline 8 & 0.5207(8) & 0.5222(9)& & 0.5274(13)& 0.528(4) \\ 5 & 0.8585(15) & 0.868(3) & & 0.875(3) & 0.889(4) \\ 4 & 1.095(3) & 1.109(2) & 1.112(4)& 1.122(8) & 1.135(7) \\ 3 & 1.535(8) & 1.555(10)& & 1.587(10) & 1.623(15) \\ 2.4 & 2.030(8) & 2.087(16)& & 2.19(3) & 2.25(4) \\ 2 & 2.655(15) & 2.84(6) & 2.76(3) & 2.95(5) & 3.1(2) \\ 1.8 & 3.25(3) & 3.33(4) & 3.45(5) & 3.47(4) & 3.57(11) \\ 1.5 & 5.40(6) & 5.59(6) & 5.57(11)& 5.75(11) & 6.12(13) \\ 1.44 & 7.21(10) & 7.11(15) & 7.2(3) & 7.3(3) & 7.5(2) \\ 1.4 & 9.74(13) & 9.82(13) & 10.2(3) & 9.8(3) & 10.4(4) \\ 1.39 & 11.48(16) & 13.4(3) & & 13.5(6) & 13.5(8) \\ \hline \end{tabular} \caption{The measured values of $g^2$ at each $\beta_L$ and $L/a$ with $6$ flavours of fermions.} \label{table:couplingnf6} \end{table} \begin{table} \centering \begin{tabular}{|l|l|l|l|l|} \hline $\beta_L$ & $L/a=6$ & $L/a=8$ & $L/a=12$ & $L/a=16$ \\ \hline 8 & 0.4700(2)& 0.4706(4) & 0.4705(5) & 0.4707(10) \\ 6 & 0.6148(3)& 0.6159(5) & 0.6180(9) & 0.6181(19) \\ 4 & 0.8897(9)& 0.8897(13)& 0.895(4) & 0.895(3) \\ 3 &1.1528(16)& 1.156(3) & 1.150(2) & 1.146(4) \\ 2 & 1.651(4) & 1.653(5) & 1.637(6) & 1.624(13) \\ 1.7 & 1.924(4) & 1.907(5) & 1.905(11) & 1.896(13) \\ 1.5 & 2.183(3) & 2.137(7) & 2.116(11) & 2.10(2) \\ 1.3 & 2.542(8) & 2.473(9) & 2.382(11) & 2.37(2) \\ 1 & 4.03(2) & 3.55(2) & 3.23(3) & 3.09(4) \\ \hline \end{tabular} \caption[a]{The measured values of $g^2$ at each $\beta_L$ and $L/a$ with $10$ flavours of fermions.} \label{table:couplingnf10} \end{table} The measured values of the running coupling squared in theories with 4, 6 and 10 fundamental representation fermions are given, respectively, in tables \ref{table:couplingnf4}, \ref{table:couplingnf6} and \ref{table:couplingnf10} and shown in figures \ref{fig:couplingnf46} and \ref{fig:couplingnf10}. The measurements are taken after every trajectory, and the number of trajectories for each point is up to 230,000 (for $N_f=6$, volume $16^4$ and largest $\beta_L$-values). For 6 and 10 flavours the strongest lattice couplings (smallest $\beta_L$) are very close to the strongest practical values for our action; at smaller $\beta_L$ the simulations become either too slow or unstable, possibly signalling a proximity of a bulk phase transition. These transitions are a lattice artefact and limit the range of allowed lattice couplings. \begin{figure} \centering \includegraphics[height=0.5\textwidth]{nf4g2.eps} \includegraphics[height=0.5\textwidth]{nf6g2.eps} \caption[a]{ The measured values of $g^2(g_0^2,L/a)$ against $a/L$ with 4 and 6 flavours of fermions. The black dashed line gives an example of the running in 2-loop perturbation theory at modest coupling, normalized so that it matches the measurement at $L/a=6$. } \label{fig:couplingnf46} \end{figure} \begin{figure} \centering \includegraphics[height=0.5\textwidth]{nf10g2.eps} \caption[a]{ The measured values of $g^2(g_0^2,L/a)$ against $a/L$ with 10 flavours of fermions. The black dashed line gives an example of the running in 2-loop perturbation theory. } \label{fig:couplingnf10} \end{figure} One can immediately recognize the main features from figures \ref{fig:couplingnf46} and \ref{fig:couplingnf10}: at $N_f=4$, the coupling becomes stronger as lattice size increases, and the running becomes faster at stronger coupling. This agrees with the expected QCD-like behaviour. In contrast, with 10 flavours we observe basically no running at all within our statistical errors, except with the strongest lattice coupling ($\beta_L=1$) used, where $g^2$ becomes smaller as volume increases. This is the expected behaviour if we are at the strong coupling side of an infrared fixed point, where the $\beta$-function is positive. However, at $\beta_L=1$ the data differ qualitatively from the larger $\beta_L$ measurements, possibly a signal of contamination from finite lattice spacing effects at large bare coupling. In the theory with 6 fermions the running remains small and slightly positive in the studied range of couplings. To illustrate the running we show the scaled step scaling function $\Sigma(g^2,2,L/a)/g^2 = g^2(g_0^2,2L/a)/g^2(g_0^2,L/a)$ in figure \ref{fig:step_nf6} at $L/a=6$ and $8$. The running is compatible with the perturbation theory at small $g^2$, but deviates from it at large coupling. The scaled step scaling measurements decrease at $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 4$ and may reach unity at $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 10$, indicating an ultraviolet fixed point. However, there appears to be a systematic difference between the $L/a=6$ and $8$ points at large $g^2$. Thus, proper continuum limit is necessary. We also note that at $L/a=6$ the very largest coupling $\beta_L=4/g_0^2 = 1.39$ point deviates substantially from other points. This is caused by the finite size effects at the smallest volume, clearly visible in figure \ref{fig:couplingnf46}. \begin{figure} \centering \includegraphics[height=0.5\textwidth]{nf6Sigmalat.eps} \caption[stepscale]{The $N_f=6$ scaled lattice step scaling function $\Sigma(g^2,2,L/a)/g^2 = g^2(g_0^2,2L/a)/g^2(g_0^2,L/a)$ for $N_f=6$ calculated directly from the data in table \ref{table:couplingnf6}. The black dashed line is the continuum 2-loop perturbative result for $\sigma(g^2,2)$. } \label{fig:step_nf6} \end{figure} \begin{figure} \centering \includegraphics[scale=0.42]{betainterpolate} \includegraphics[scale=0.42]{betaint2} \caption[Interpolations in beta]{The interpolating functions for six fermions. The plot on the left shows the polynomial interpolation and the plot of the right shows the rational interpolation. For clarity, the graphs are displaced up by $3$,$2$,$1$ and $0$ and left by $0.3$,$0.2$,$0.1$ and $0$ for lattice sizes $L/a = 6$,$8$,$12$ and $16$ respectively. } \label{fig:betainterpolate} \end{figure} \begin{table} \centering \begin{tabular}{|l|l|l|l|l|l|l|l|l|} \hline $N_f$ & function & $L/a=6$ & $L/a=8$ & $L/a=12$ & $L/a=16$ & d.o.f & $L/a=10$ & d.o.f \\ \hline 4 & Polynomial & 1.565 & 6.063 & 18.961 & 3.854 & 3 & & \\ & Rational & 2.512 & 3.623 & 5.175 & 3.771 & 3 & & \\ \hline 6 & Polynomial & 35.25 & 72.76 & 30.20 & 16.11 & 6 & & \\ & Rational & 8.030 & 5.897 & 6.968 & 4.661 & 7 & 3.545 & 2\\\hline 10 & Polynomial & 13.744 & 8.087 & 13.802 & 5.732 & 5 & & \\ & Rational & 9.896 & 6.271 & 12.704 & 6.301 & 6 & & \\ \hline \end{tabular} \caption[$\chi^2$ values]{$\chi^2$ values and free degrees of freedom in the polynomial fits \ref{eq:g2function} and the rational fits \ref{eq:g2function2}.} \label{table:chi2} \end{table} The continuum limit of the step scaling function $\Sigma(g^2,2,L/a)$ has to be evaluated at constant Schrödinger functional coupling $g^2$. However, the measurements were performed at selected fixed values of $g_0^2=4/\beta_L$, which do not correspond to the same $g^2$-values at $L/a=6$ and $8$. Therefore, it is necessary to shift the measurements so that $g^2$-values match. This could be done by performing new simulations so that $g^2$ at $L/a=6$ and $8$ match, or by reweighting in $g_0^2$ and $\kappa$. However, a much more economical and convenient way to achieve this is to interpolate the actual measurements of $g^2(g_0^2,L/a)$ at each lattice size $L/a$ by fitting to a function of $g_0^2$. This results in a ``measurement'' of $g^2(g_0^2,L/a)$ over a continuous range of $g_0^2$. It is necessary that the measurements cover the interpolated range densely enough; otherwise the form of the fitted function ansatz may influence the final results. Obviously, the fit must also be statistically good in order to describe the true behaviour of the data. The interpolating function also somewhat averages out fluctuations in individual data points. We have done the interpolation using polynomial and rational interpolating functions. Polynomial function is a power series in the bare coupling, and it is commonly used in the literature \cite{Appelquist:2009ty,Bursa:2010xr}: \begin{align} \frac{1}{g^2(g_0^2,L/a)} = \frac{1}{g_0^2} \left[ 1 + \sum_{i=1}^n c_i g_0^{2i}\right]. \label{eq:g2function} \end{align} In the case of 4 and 10 flavours we truncate the series at $n=4$ and in the case of 6 flavours, where more $\beta_L$-values are available, at $n=5$. These values were chosen to optimise the confidence levels of the fits, while keeping the number of the fit parameters tolerably small. We also kept the same number of terms for each $L/a$. The $\chi^2$-values of the fits are shown in table \ref{table:chi2}, and, for $N_f=6$, the fit is shown in figure \ref{fig:betainterpolate}, left. It is clear that the polynomial fits do not fit the data very well, especially at the strongest bare couplings where the measured $g^2$ increases very rapidly. This leads us to try a rational function interpolation: \begin{align} \frac{1}{g^2(g_0^2,L/a)} = \frac{1}{g_0^2} \left[\frac{ 1 + \sum_{i=1}^n a_i g_0^{2i}}{1 + \sum_{i=1}^m b_i g_0^{2i}}\right]. \label{eq:g2function2} \end{align} For 4 and 6 fermions the number of terms were chosen to be $n=m=2$ and for 10 fermions $n=1$ and $m=2$.\footnote{% For $N_f=10$ choosing $n=2$ produces a singularity within the fitting range, which is not acceptable.} Again, these parameters were chosen to optimize the confidence level without producing singularities within the range of $g_0^2$ of the measurements. The stability of the fits was checked by varying $n$ or $m$. The rational interpolation captures the rapid increase at strong bare couplings better, and the confidence levels of the fits are better in all cases, as shown in table \ref{table:chi2} and figure \ref{fig:betainterpolate}. Thus, we perform the subsequent analysis using the rational interpolation functions. The interpolating functions are used to calculate $\Sigma(u,s,L/a)$ with $L/a=6,8$ in a continuous range of values of $u=g^2$. This enables us to perform a continuum extrapolation. Since we expect most of the order $a$ errors to have been removed, we fit the data at $L/a=6$ and $L/a=8$ with a function of the form \begin{align} \Sigma(u,2,L/a) = \sigma(u,2) + c(u) \left ( L/a \right )^{-2} \label{continuum} \end{align} independently at each value of $u$. This extrapolation is shown for $N_f=6$ for three selected values of $u$ in figure \ref{sigmacontNf6}. To estimate the systematic errors from the extrapolation we compare with the result obtained by simply taking the largest volume step scaling, $\sigma(u,2) = \Sigma(u,2,8)$. Because we only have step scaling data at $L/a=6$ and $8$, we are naturally not able to verify that the first order in $L/a$ -term is indeed small or that the higher order contributions can be neglected. However, we note that using a first order only extrapolation would give results roughly comparable with the extrapolation (\ref{continuum}), but with larger statistical errors and further away from the largest volume result. The $L/a=10$ results with $N_f=6$ are not used for the measurement of the step scaling for the running coupling, but are needed for the measurement of the anomalous dimension. \begin{figure} \centering \includegraphics[scale=0.45]{sigmacontNf6} \caption[Continuum extrapolation of sigma.]{ The step scaling function extrapolated to the continuum limit with $N_f=6$ and for three chosen values of $u=g^2$. } \label{sigmacontNf6} \end{figure} The final results of the step scaling functions are shown in figures \ref{sigmanf4}, \ref{sigmanf6} and \ref{sigmanf10} for 4, 6 and 10 fermion flavours respectively. The errors shown include only the statistical errors from the measurements, fits and extrapolation. A rough measure of the systematic errors can be obtained from the comparison of the infinite volume extrapolated result and the result using only the largest volume step scaling without extrapolation. The error propagation is calculated using jackknife blocking throughout the whole analysis. \begin{figure} \centering \includegraphics[scale=0.5]{nf4.eps} \caption[The step scaling function with 4 fermions.]{ (colour online) The scaled step scaling function $\sigma(g^2,2)/g^2$ with 4 fermions. The thick red line corresponds to the continuum extrapolation, \eq{continuum}, and the hashed band to the statistical errors of the extrapolation. The thick dashed line with the shaded error band is the largest volume step scaling function without extrapolation. The thin dashed line is the 2-loop perturbative value of $\sigma(g^2,2)/g^2$. } \label{sigmanf4} \end{figure} \begin{figure} \centering \includegraphics[scale=0.5]{nf6.eps} \caption[The step scaling function with 6 fermions.]{ As in figure \ref{sigmanf4} but with 6 fermion flavours.} \label{sigmanf6} \end{figure} \begin{figure} \centering \includegraphics[scale=0.5]{nf10.eps} \caption[The step scaling function with 10 fermions.]{ As in figure \ref{sigmanf4} but with 10 fermion flavours.} \label{sigmanf10} \end{figure} \begin{figure} \centering \includegraphics[width=0.48\textwidth]{nf4bf.eps} \includegraphics[width=0.48\textwidth]{nf6bf.eps}\\ \includegraphics[width=0.48\textwidth]{nf10bf.eps} \caption[The estimators for the $\beta$-function.]{ The estimators for the $\beta$-function in equation \eq{eq:beta*}. } \label{fig:beta*} \end{figure} We have also calculated the estimator for the $\beta$-function, $\beta^*$, defined in \eq{eq:beta*}. These are shown in figure \ref{fig:beta*}, together with the 2-loop $\beta$-function and $\beta^*$ obtained from the 2-loop perturbative step scaling function. Thus, the difference between the perturbative curves gives a measure of the error made in the approximation in \eq{eq:beta*}. What can we conclude from the results? From figure \ref{sigmanf4} we can observe that, in the range of couplings studied, the $N_f=4$ case behaves as expected: $\beta$-function is negative and becomes smaller as the coupling increases, and agrees overall with the 2-loop perturbative $\beta$-function. Naturally deviations from the perturbation theory are expected at larger couplings. However, for our purposes it is sufficient to verify the QCD-like behaviour and we did not attempt to reach smaller couplings. For $N_f=6$, the results in figure \ref{sigmanf6} show that the largest volume step scaling deviates from the perturbative result already at $g^2 \sim 4$, after which it decreases. However, the continuum extrapolation remains close to the perturbative curve up to $g^2\sim 7$--$8$. After this the curve turns sharply downwards; this behaviour is caused by the anomalous $L/a=6$, $\beta_L=1.39$ point visible in figures \ref{fig:couplingnf46} and \ref{fig:step_nf6}. The largest volume step scaling (shaded band) can be compared to the corresponding $L/a=8$ uninterpolated step scaling in figure \ref{fig:step_nf6}. The large difference between the largest volume and extrapolated step scaling functions is naturally due to the fact that the step scaling $\Sigma(g^2,L/a)$ at $L/a=6$ and $L/a=8$ differ substantially, and the lever arm to the continuum ($a/L\rightarrow 0$) is long. The variation between these two curves gives an estimate of the systematic errors in the extrapolation. The largest values of the Schr\"odinger functional coupling we achieved was around $g^2 \approx 12$ ($\alpha \approx 0.96$). Unfortunately, neither the coupling is strong enough nor the errors are sufficiently small in order to unambiguously distinguish between an infrared fixed point at $g^2\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 11$ and ``walking'' behaviour, where the step scaling function starts to increase again at stronger coupling. With our choice of the action we cannot reach stronger couplings because simulations become rapidly impractical at smaller $\beta_L$ than 1.39, our strongest lattice coupling at $N_f=6$. Obviously, calculating the step scaling at volumes $L/a=10$ and $20$ (or larger) would be needed in order to stabilize the continuum extrapolation. Finally, for $N_f=10$ the measured $\beta$-function is essentially compatible with zero or with the perturbative $\beta$-function at $g^2 < 2.5$. In this case the 2-loop perturbative $\beta$-function is expected to be relatively accurate (higher order perturbative corrections in MS-scheme are small \cite{vanRitbergen:1997va}). The value of the $\beta$-function is very small, and clearly, the accuracy of our results falls far short from being able to resolve the behaviour of the $\beta$-function in this range. Nevertheless, theoretically we know that the $\beta$-function must be negative at small coupling. At $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2.5$ the measured $\beta$-function deviates significantly upwards from the perturbative one. Combined with the knowledge that the $\beta$-function is negative at small couplings, this indicates the presence of an infrared fixed point. The onset for this behaviour is very close to the fixed point in the 2-loop $\beta$-function. This deviation is in practice caused by the strongest coupling $\beta_L=1$ data from the lattice: from figure \ref{fig:couplingnf10} we can see that only at $\beta_L=1$ the measured coupling $g^2 > 2.5$. At these couplings we can expect strong lattice artefacts; thus, we believe that the deviation from the perturbative value is caused by unaccounted for systematic errors in the continuum extrapolation. We still emphasize that in absolute numbers the deviation is still rather small; in order to resolve very slow running we would need extremely accurate measurements with correspondingly very small systematic errors. This does not appear to be doable at $N_f=10$ with the methods used here. \begin{table} \centering \begin{tabular}{|l|l|l|l|l|l|l|} \hline $\beta_L$ & $L/a=6$ & $L/a=8$ & $L/a=10$ & $L/a=12$ & $L/a=16$ & $L/a=20$ \\ \hline 2.4 & 0.9666(8) & 0.9353(9) & 0.9171(19)& 0.9014(19)& 0.8870(15)& 0.865(4) \\ 2 & 0.8953(10)& 0.857(3) & 0.838(2) & 0.823(3) & 0.793(4) & 0.766(6) \\ 1.5 & 0.702(4) & 0.669(3) & 0.646(4) & 0.618(6) & 0.586(5) & 0.573(6) \\ 1.44 & 0.636(3) & 0.610(3) & 0.588(4) & 0.572(4) & 0.548(4) & 0.517(6) \\ 1.4 & 0.543(5) & 0.547(5) & 0.539(5) & 0.534(5) & 0.508(6) & 0.480(8) \\ 1.39 & 0.508(5) & 0.515(7) & 0.520(6) & 0.517(6) & 0.488(8) & 0.476(9) \\ \hline \end{tabular} \caption{The measured values of $Z_P$ at each $\beta_L$ and $L$ with $6$ flavours of fermions.} \label{table:zpnf6} \end{table} We measure the anomalous dimension for the interesting case of SU(2) with 6 flavours of fermions using the mass step scaling method described in section \ref{model}. This method is much less noisy than the measurement of the coupling, and we can now reliably use lattices of size $20^4$. The measurements of $Z_p$ are listed in table \ref{table:zpnf6}. \begin{table} \centering \begin{tabular}{|l|l|l|l|l|l|l|} \hline $L/a=6$ & $L/a=8$ & $L/a=10$ & $L/a=12$ & $L/a=16$ & $L/a=20$ \\ \hline 25.15 & 6.48 & 0.28 & 0.84 & 3.00 & 1.50 \\ \hline \end{tabular} \caption[$\chi^2$ values]{$\chi^2$ values (2 degrees of freedom) for the $Z_P$ fits for $N_f=6$ using \eq{eq:zpfunction}. } \label{table:zpchi2} \end{table} We start the analysis by finding an interpolating function to the measured values of $Z_P$ by fitting with a power series \begin{align} Z_P(\beta_L,L/a) = 1 + \sum_{i=1}^n c_i g_0^{2i}, \label{eq:zpfunction} \end{align} where we have truncated the series at $n=4$. The $\chi^2$ values for the fits are given in table \ref{table:zpchi2}. The fit at $L/a=6$ has bad $\chi^2/$ value; however, its effect to the final extrapolation is small. We check the systematic errors of the interpolation by also truncating at $n=3$; the results remain essentially the same with somewhat increased statistical errors. From the interpolated $Z_p(\beta_L,L/a)$ we obtain the mass step scaling function $\Sigma_P(u,s,L/a)$ at $L/a=6,8$ and $10$ using \eq{Sigmap}, using $u=g^2$ from rational fit in \eq{eq:g2function2}. The continuum extrapolation is then done by fitting to the extrapolating function \begin{align} \Sigma_P(u,2,L/a) = \sigma_P(u,2) + c(u) \left ( L/a \right )^{-2}. \end{align} The fit is shown in figure \ref{sigmapcontNf6}. We check the systematic effects in the extrapolation by leaving out the smallest volume step scaling at $L/a=6$. The results remain compatible with each other, but with naturally much smaller final statistical errors when $L/a=6$ is included. Step scaling function is converted to the estimate of the anomalous dimension using \eq{gammastar}, and the results are shown in figure \ref{figgamma}. In this case the non-trivial continuum extrapolation is essential in order to find the monotonic growth of $\gamma^*(g^2)$. If we would take e.g. $L/a=8$ step scaling alone to estimate the continuum behaviour, $\gamma^*$ would decrease at $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 6$. This emphasizes the importance of the controlled continuum limit. \begin{figure} \centering \includegraphics[scale=0.42]{sigmapcont} \caption[a]{The mass step scaling function extrapolated to the continuum limit, using $N_f=6$ data and shown for three chosen values of $g^2$. } \label{sigmapcontNf6} \end{figure} \begin{figure} \centering \includegraphics[width=0.5\textwidth]{gamma.eps} \caption[The mass anomalous dimension with 6 fermions.]{The estimate of the mass anomalous dimension, $\gamma^*(g^2)$, with $N_f=6$. Shown are the continuum extrapolation using the mass step scaling $\Sigma(L/a,2,g^2)$ with all three volumes $L/a=6,8,10$ (hashed band) and only with $L/a = 8,10$ (shaded band).} \label{figgamma} \end{figure} The mass anomalous dimension we measure is somewhat smaller than the perturbative one at strong coupling. It remains small at all measured values of the coupling, and if there is a fixed point at $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 12$, the anomalous dimension is $\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.25$. The behaviour of the anomalous dimension is in contrast with the results obtained with an unimproved Wilson action \cite{Bursa:2010xr}, where $\gamma^*$ significantly larger than the perturbative one was seen already at $g^2 \approx 5$, albeit with large errors. Both of the measurements were done using the Schr\"odinger functional scheme, and thus should yield identical results in the continuum limit. This again points to the importance of controlling the cutoff effects and the continuum extrapolation. On the other hand, the behaviour we measure is qualitatively similar to the one observed in SU(2) gauge theory with adjoint fermions using HYP-smeared improved fermions \cite{DeGrand:2011qd}. For the $N_f=10$ case we did not perform a measurement of the mass anomalous dimension. Since our results for the coupling constant in the $N_f=10$ case are compatible with perturbative results, we expect this to hold also for the mass anomalous dimension. In other words, we expect that $\gamma^\ast=0.08$ corresponding with a perturbative fixed point at $g^2\simeq 2.90$. \section{Conclusions} \label{conclusions} In this paper we have presented the results of a lattice study of SU(2) gauge theory with $N_f = 4$, $6$ or $10$ flavours of Wilson fermions in the fundamental representation of the gauge group. The numbers of fermion flavours were chosen according to the expected boundaries of the conformal window: $N_f=10$ is the largest value where the theory is still asymptotically free, and hence it has a fixed point at small coupling where perturbation theory is expected to be accurate. On the other hand, $N_f=6$ is expected to be close to the lower boundary of the conformal window, whereas $N_f=4$ should be safely below it and hence QCD-like. We have measured the $\beta$-functions of the above theories using the Schr\"odinger functional scheme. We also measure the mass anomalous dimension $\gamma$ for the $N_f=6$ theory. In our analysis we have used perturbatively improved Wilson-clover action, together with the improvement of the boundary terms in the Schr\"odinger functional approach. With this improvement we expect most of the $O(a)$ errors to be eliminated. Overall our findings agree with the expectations. For $N_f=10$ we observe a very small $\beta$-function at small couplings, and a positive $\beta$-function at larger ($g^2\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$) couplings. Together with the fact that due to the asymptotic freedom the $\beta$-function must be negative at small enough coupling, this indicates that the $\beta$-function has a zero somewhere at $g^2 \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$. This agrees with the perturbative 2-loop $\beta$-function which vanishes at $g^2 \approx 2.90$. At the interesting case of six flavours we observe step scaling (or $\beta$-function) behaviour compatible with the perturbation theory up to $g^2 \approx 5$. Above this the $\beta$-function starts to approach zero, with a possible fixed point around $g^2 \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 12$. However, our statistical errors are too large in order to fully resolve the behaviour of the $\beta$-function: instead of the fixed point, the $\beta$-function could as well start to decrease again at stronger couplings. This kind of behaviour is characteristic for the walking coupling. It would be very interesting to resolve the behaviour in this case: as discussed in the introduction, in the ladder approximation the chiral symmetry breaking sets in around the critical coupling $g^2_c \sim 4\pi^2/(3C_2(R)) \approx 17$. The estimated provisional fixed point is close to this value (although computed with a different schema), and thus it is indeed plausible that either the fixed point or the chiral symmetry breaking is realized. Obviously, simulation at much larger volumes (closer to the continuum) would be required to fully pin down the behaviour of the $\beta$-function. However, it is just barely possible to reach these values of the Schr\"odinger functional scheme coupling with the action we use: in order to reach strong Schr\"odinger functional coupling one has to use strong bare lattice coupling, and the couplings used in this work are already near the largest values which can be used in practice. Clearly, it is important to develop an action which can be used at stronger couplings. We also measured the mass anomalous dimension using six fermion flavours. The dimension is seen to grow slowly as the coupling increases; somewhat slower than the perturbative result at strong couplings, but overall the result in this case is as expected. Finally, for four flavours the observed behaviour is unsurprising: the measurements indicate a smoothly decreasing $\beta$-function and QCD-like behaviour. \acknowledgments This work is supported by the Academy of Finland grant 114371. JR acknowledges the support from Väisälä foundation and TK from Magnus Ehrnrooth foundation. The simulations were performed at the Finnish IT Center for Science (CSC), Espoo, Finland, and at EPCC, University of Edinburgh. Parts of the simulation program have been derived from the MILC lattice simulation program \cite{MILC}.
\section{Introduction} The stability of general relativity (GR) predictions against small graviton mass has been a persistent challenge of classical field theory. The simplest ghost-free extension of GR with a linear mass term \cite{Fierz:1939ix} suffers from the van Dam-Veltman-Zakharov discontinuity, giving rise to different predictions for the classical tests in the vanishing mass limit \cite{vanDam:1970vg, Zakharov:1970cc}. Although this problem can be alleviated by nonlinear terms \cite{Vainshtein:1972sx}, the cost is the emergence of the Boulware-Deser (BD) ghost \cite{Boulware:1973my}, which is generically unavoidable due to six degrees of freedom in the metric instead of the five of the massive spin 2 field. Adopting an effective field theory approach in the decoupling limit, the source of all the above issues can be traced back to the helicity 0 mode of the graviton \cite{ArkaniHamed:2002sp}. In this perspective, an analogue of the cancellation of the BD ghost in the linear theory by a specific choice of the mass term can be performed, rendering the nonlinear theory ghost-free up to quartic order by tuning the coefficients \cite{Creminelli:2005qk}. More recently, a two parameter theory of nonlinear massive gravity was developed \cite{deRham:2010ik, deRham:2010kj}. In this construction, terms to each order are chosen to remove the additional degree of freedom (would-be BD ghost) at the decoupling limit and it has the potential to be free of the BD ghost at fully nonlinear level \cite{Hassan:2011hr, deRham:2011rn, deRham:2011qq}. The more general construction with dynamical auxiliary metrics was also claimed to be free from the BD ghost at fully nonlinear order \cite{Hassan:2011tf, Hassan:2011ea}. The massive extensions of GR are known to allow self-accelerating solutions \cite{Salam:1976as, Damour:2002wu, deRham:2010tw, Koyama:2011xz, Nieuwenhuizen:2011sq, Koyama:2011yg, Chamseddine:2011bu, D'Amico:2011jj, Gumrukcuoglu:2011ew, Koyama:2011wx, Comelli:2011wq, Volkov:2011an, vonStrauss:2011mq, Comelli:2011zm, Berezhiani:2011xx}. This provides an application opportunity for the recent formulations of massive gravity, as an alternative approach to account for the current accelerated expansion. On the other hand, to model the accelerating universe in the scope of this infrared modified gravity theory, a cosmological solution is necessary. In the construction of \cite{deRham:2010ik, deRham:2010kj} with Minkowski fiducial metric, flat Friedmann-Robertson-Walker (FRW) universe cannot be realized \cite{D'Amico:2011jj}, although cosmological solutions with negative spatial curvature exist \cite{Gumrukcuoglu:2011ew}. For more general constructions with a nondynamical \cite{Tanahashi-progress} and dynamical \cite{vonStrauss:2011mq, Comelli:2011zm} auxiliary metrics, maximally symmetric FRW with either curvature may be allowed. Given the nontrivial cosmological solutions which self-accelerate, it is thus a necessity to understand their properties against perturbations. In the light of the construction which removes the additional degree (would-be BD ghost), the expectation is to have five degrees of freedom associated with the five polarizations of the massive graviton. On the other hand, the absence of BD ghost does not guarantee that the theory is safe; one still needs to determine the conditions under which the helicity 0 and 1 modes, which are pure gauge in the massless theory, are stable \cite{Higuchi:1986py}. Furthermore, these additional degrees should not be in conflict with observations. For instance, because of the emergence of a new scalar degree in the gravity sector, the Newtonian potential may acquire modifications from couplings between the matter sector and the helicity 0 graviton. If this is the case, the parameters of the theory can be restricted by e.g. the solar system tests. The primary goal of the present paper is to address these questions. We consider the cosmological solutions in massive gravity as background and in the presence of a generic matter content, we present a gauge invariant formulation of perturbations. Although as we argued above, we expect the gravity sector to contain $5$ dynamical degrees of freedom, at the level of the quadratic action, we show that the helicity 0 and 1 graviton modes have vanishing kinetic terms but finite masses.\footnote{ A similar situation was noted in the decoupling limit \cite{deRham:2010tw,Koyama:2011wx}.} As a result, in addition to the matter perturbations, the only dynamical degrees of freedom are the two tensor polarizations of gravity waves. This is exactly the same number of degrees of freedom as in GR. For the action quadratic in perturbations around the cosmological backgrounds, the mass terms of the potentially ghost-free construction turn out to be completely decoupled from the standard part (Einstein-Hilbert and matter terms) for scalar and vector modes. Thus, these modes evolve identically to their counterparts in GR, except for the additional cosmological constant contributed by the graviton mass. The only non trivial effect on the dynamics occurs in the tensor modes, which acquires a time dependent mass term determined by the fiducial metric of the theory. The paper is organized as follows. In Section II, we review the setup with a fiducial metric that is Minkowski and summarize the only cosmological solution \cite{Gumrukcuoglu:2011ew} allowed in this case. In Section III, we give a detailed study of the complete quadratic action for perturbations in the presence of a general fiducial metric of the FRW form, with an ansatz for the physical background metric also of FRW type with arbitrary spatial curvature. As an example, we specify to perturbations around the open universe solution \cite{Gumrukcuoglu:2011ew} driven by a single scalar field matter in Section IV. We conclude with Section V, where we summarize our results. The paper is supplemented by a number of Appendices, where the details of calculations are presented. \section{Open FRW solution with Minkowski fiducial metric} \label{setup} In this section, we review the open FRW universe solution \cite{Gumrukcuoglu:2011ew} in nonlinear massive gravity \cite{deRham:2010kj} coupled to general matter content. The covariant action for the gravity sector is constructed out of the four dimensional metric $g_{\mu\nu}$ and the four scalar fields $\phi^a\,(a=0,1,2,3)$ called {\it St\"uckelberg fields}. The action respects the Poincare symmetry in the field space, i.e. invariance under the constant shift of each of $\phi^a$ and the Lorentz transformation mixing them: \begin{equation} \phi^a \to \phi^a + c^a, \quad \phi^a \to \Lambda^a_b\phi^b. \label{eqn:Poincare-tr} \end{equation} The following line element in the field space is invariant under these transformations. \begin{equation} \eta_{ab}d\phi^ad\phi^b = -(d\phi^0)^2 + \delta_{ij}d\phi^id\phi^j. \label{eqn:modulispacemetric} \end{equation} Indeed, this is the unique geometrical quantity in the field space of $\phi^a$. Thus the action can depend on $\phi^a$ only through the spacetime tensor \begin{equation} f_{\mu\nu} \equiv \eta_{ab}\partial_{\mu}\phi^a\partial_{\nu}\phi^b. \end{equation} In this language, general covariance is spontaneously broken by the vacuum expectation value (vev) of $f_{\mu\nu}$. By assumption, matter fields propagate on the physical metric $g_{\mu\nu}$, but are not coupled to $f_{\mu\nu}$ directly. The tensor $f_{\mu\nu}$, constructed from the invariant line element in the field space, is often called a {\it fiducial metric}. On the other hand, the spacetime metric $g_{\mu\nu}$, on which matter fields propagate, is often called a {\it physical metric}. The gravity action is the sum of the Einstein-Hilbert action (with the cosmological constant $\Lambda$) $I_{EH,\Lambda}$ for the physical metric $g_{\mu\nu}$ and the graviton mass term $I_{mass}$ specified below. Adding the matter action $I_{matter}$, the total action is \begin{equation} I = I_{EH,\Lambda}[g_{\mu\nu}] + I_{mass}[g_{\mu\nu},f_{\mu\nu}] + I_{matter}[g_{\mu\nu},\sigma_I], \end{equation} where \begin{eqnarray} I_{EH,\Lambda}[g_{\mu\nu}] & = & \frac{M_{Pl}^2}{2}\int d^4x\sqrt{-g}(R-2\Lambda), \\ I_{mass}[g_{\mu\nu},f_{\mu\nu}] & = & M_{Pl}^2m_g^2\int d^4x\sqrt{-g}\, ( {\cal L}_2+\alpha_3{\cal L}_3+\alpha_4{\cal L}_4), \label{eqn:Imass} \end{eqnarray} and $\{\sigma_I\}$ ($I=1,2,\cdots$) represent matter fields. Demanding the absence of ghost at least in the decoupling limit~\cite{deRham:2010kj}, each contribution in the mass term $I_{mass}$ is constructed as \begin{eqnarray} {\cal L}_2 & = & \frac{1}{2} \left(\left[{\cal K}\right]^2-\left[{\cal K}^2\right]\right)\,, \nonumber\\ {\cal L}_3 & = & \frac{1}{6} \left(\left[{\cal K}\right]^3-3\left[{\cal K}\right]\left[{\cal K}^2\right]+2\left[{\cal K}^3\right]\right), \nonumber\\ {\cal L}_4 & = & \frac{1}{24} \left(\left[{\cal K}\right]^4-6\left[{\cal K}\right]^2\left[{\cal K}^2\right]+3\left[{\cal K}^2\right]^2 +8\left[{\cal K}\right]\left[{\cal K}^3\right]-6\left[{\cal K}^4\right]\right)\,, \label{lag234} \end{eqnarray} where the square brackets denote trace operation and \begin{equation} {\cal K}^\mu _\nu = \delta^\mu _\nu - \left(\sqrt{g^{-1}f}\right)^{\mu}_{\ \nu}\,. \label{Kdef} \end{equation} The square-root in this expression is the positive definite matrix defined through \begin{equation} \left(\sqrt{g^{-1}f}\right)^{\mu}_{\ \rho} \left(\sqrt{g^{-1}f}\right)^{\rho}_{\ \nu} = f^{\mu}_{\ \nu}\ (\equiv g^{\mu\rho}f_{\rho\nu}). \label{eqn:square-root} \end{equation} As already stated above, a vev of the tensor $f_{\mu\nu}$ breaks general covariance spontaneously. Thus, in order to find FRW cosmological solutions in this theory, we should adopt an ansatz in which not only $g_{\mu\nu}$ but also $f_{\mu\nu}$ respects the symmetry of the FRW universes~\cite{Gumrukcuoglu:2011ew}. Since the tensor $f_{\mu\nu}$ is the pullback of the Minkowski metric in the field space to the physical spacetime, construction of such an ansatz is equivalent to finding a flat, closed, or open FRW coordinate system for the Minkowski line element. It is well known that the Minkowski line element does not admit a closed FRW chart but allows an open FRW chart. For this reason, in order to find open FRW solutions \cite{Gumrukcuoglu:2011ew}, we first perform the field redefinition from $\phi^a$ to new fields $\varphi^a$ so that $f_{\mu\nu}$ written in terms of $\varphi^a$ manifestly has the symmetry of open FRW universes as \begin{equation} f_{\mu\nu} = -n^2(\varphi^0) \partial_{\mu}\varphi^0\partial_{\nu}\varphi^0 + \alpha^2(\varphi^0) \Omega_{ij}(\varphi^k)\partial_{\mu}\varphi^i\partial_{\nu}\varphi^j, \label{eqn:fmunu} \end{equation} where $i,j=1,2,3$, and \begin{equation} \Omega_{ij}(\varphi^k) = \delta_{ij} + \frac{K\delta_{il}\delta_{jm}\varphi^l\varphi^m} {1-K\delta_{lm}\varphi^l\varphi^m} \label{eqn:Omegaij} \end{equation} is the metric of the maximally symmetric space with the curvature constant $K$ ($<0$). Concretely, this is achieved by \begin{equation} \phi^0 = f(\varphi^0)\sqrt{1-K\delta_{ij}\varphi^i\varphi^j} \,,\qquad \phi^i = \sqrt{-K}f(\varphi^0)\varphi^i\,, \label{eqn:field-redefinition} \end{equation} and \begin{equation} n(\varphi^0) = |\dot{f}(\varphi^0)|, \qquad \alpha(\varphi^0) = \sqrt{-K}|f(\varphi^0)|, \label{eqn:n-alpha-Minkowski} \end{equation} where $f$ is a function to be determined and $\dot{f}$ represents its derivative. We then adopt the ``unitary gauge'' \begin{equation} \varphi^0 = t, \quad \varphi^i = x^i, \label{eqn:unitary-gauge} \end{equation} so that \begin{equation} f_{\mu\nu}dx^{\mu}dx^{\nu} = - (\dot{f}(t))^2\,dt^2 + |K|\,(f(t))^2\,\Omega_{ij}(x^k)dx^idx^j\,. \label{etaline} \end{equation} This is nothing but the Minkowski line element in the open chart. For the physical metric, we adopt the open FRW ansatz \begin{equation} ds^2 = -N(t)^2dt^2+a(t)^2\,\Omega_{ij}(x^k) dx^idx^j\,. \label{eqn:FRWbackground} \end{equation} Hereafter, we assume that $N > 0$ and $a>0$, without loss of generality. The background action now yields, up to boundary terms, \begin{equation} I = M_{Pl}^2 \int dt\,d^3x Na^3 \sqrt{\Omega} \left(L_{EH}[N,a]+m_g^2L_{mass}[N,a,f] \right) + I_{matter}[N,a,\sigma_I]\,, \label{totactin} \end{equation} consisting of the Einstein-Hilbert part \begin{equation} L_{EH} = \frac{3\,K}{a^2}-\frac{3\,\dot{a}^2}{a^2\,N^2}\,, \end{equation} and the contribution from the mass term \begin{eqnarray} L_{mass} &=& \left(1-\frac{\sqrt{-K}|f|}{a}\right) \left[6+4\,\alpha_3+\alpha_4 - \frac{\sqrt{-K}|f|}{a} \left(3+5\,\alpha_3 +2\,\alpha_4\right) - \frac{K\,|f|^2}{a^2}\,\left(\alpha_3+\alpha_4\right)\right] \nonumber\\ && +{\mathrm sgn}(\dot{f}/f)\frac{|f|\,\dot{a}}{N\,a}\nonumber\\ && \quad\times \left[3\left(3+3\,\alpha_3+\alpha_4\right) - \frac{3\,\sqrt{-K}|f|}{a} \left(1+2\,\alpha_3 +\alpha_4\right) - \frac{K\,|f|^2}{a^2}\,\left(\alpha_3+\alpha_4\right)\right]\,. \end{eqnarray} Hereafter, an overdot represents derivative w.r.t. the time $t$. Varying the action (\ref{totactin}) with respect to $f$ yields the following constraint \begin{eqnarray} & & \left[H-{\mathrm sgn}(\dot{f}/f)\frac{\sqrt{-K}}{a} \right] \nonumber\\ & & \quad \times \left[3+3\,\alpha_3 +\alpha_4 -\frac{2\,\sqrt{-K}\,|f|}{a}\left(1+2\,\alpha_3+\alpha_4\right)-\frac{K\,|f|^2}{a^2}\,\left(\alpha_3+\alpha_4\right)\right]=0\,, \label{eqn:constraintf} \end{eqnarray} where the Hubble expansion rate of the physical metric is defined as \begin{equation} H \equiv \frac{\dot{a}}{N\,a}\,. \label{eqn:def-Hubble} \end{equation} Out of the three solutions of the constraint (\ref{eqn:constraintf}), the trivial solution $\dot{a} = {\mathrm sgn}(\dot{f}/f)\sqrt{-K}\,N$ corresponds to the Minkowski spacetime in open chart. The remaining two branches of solutions are given by \cite{Gumrukcuoglu:2011ew} \begin{equation} \alpha(t) = X_{\pm}a(t)\,, \qquad X_{\pm} \equiv \frac{1+2\,\alpha_3+\alpha_4 \pm\sqrt{1+\alpha_3+\alpha_3^2-\alpha_4}} {\alpha_3+\alpha_4}\ (>0)\,, \label{eq:fsolcosmo} \end{equation} and describe FRW cosmologies with $K<0$.\footnote{Note that $X_{\pm}$ are positive by definition since $\alpha(t)>0$ and we assumed $a(t)>0$. If we instead assumed $a(t)<0$ then the corresponding solutions would be $\alpha(t)=-X_{\pm}a(t)$ with the same $X_{\pm}$ and we would conclude $X_{\pm}>0$ again. The essential reason for the positivity of $X_{\pm}$ is that the square-root in (\ref{Kdef}) is the positive one.\label{footnote:Xpositive}} In the present paper we will focus only on these nontrivial cosmological solutions. Using the above constraint and varying the action (\ref{totactin}) with respect to $N$ and $a$, we obtain the remaining background equations \begin{eqnarray} 3\,H^2 +\frac{ 3\,K}{a^2} & = & \Lambda_\pm +\frac{1}{M_{Pl}^2}\rho\,, \nonumber\\ -\frac{2\dot{H}}{N} +\frac{ 2\,K}{a^2} & = & \frac{1}{M_{Pl}^2}(\rho+P), \end{eqnarray} where $\rho$ and $P$ are the energy density and the pressure of matter fields calculated from $I_{matter}$, and \begin{equation} \Lambda_\pm \equiv -\frac{m_g^2}{\left(\alpha_3+\alpha_4\right)^2}\left[\left(1+\alpha_3\right)\left(2+\alpha_3 +2\,\alpha_3^2-3\,\alpha_4\right) \pm 2\,\left(1+\alpha_3 +\alpha_3^2-\alpha_4\right)^{3/2}\right]\,. \label{lambdapm} \end{equation} Thus, for the cosmological solutions (\ref{eq:fsolcosmo}), the contribution from the graviton mass term $I_{mass}$ at the background level mimics a cosmological constant with the value $\Lambda_\pm$. For $\alpha_4=(3+2\,\alpha_3+3\,\alpha_3^2)/4$ and $\pm (1+\alpha_3)>0$, the effective cosmological constant $\Lambda_\pm$ vanishes, and the background solution reduces to the open FRW universe solution of GR. On the other hand, both $X_{\pm}$ and $\Lambda_\pm$ diverge for $\alpha_4 =-\alpha_3$ and $\pm(1+\alpha_3)>0$. In Figure~\ref{fig:Lambdarange}, we show the sign of $\Lambda_\pm$ in the $(\alpha_3, \alpha_4)$ space. Note that $X_{\pm}$ are restricted to be positive by definition, as explained in footnote \ref{footnote:Xpositive}. Except for the restriction due to the positivity of $X_{\pm}$, these are in agreement with the analogous region plots presented in Ref.\cite{Koyama:2011wx}~\footnote{Substituting $\alpha_3 \to 3\,\alpha_3$, $\alpha_4 \to 12\,\alpha_4$, and switching the positive and negative branch definitions, our expression (\ref{lambdapm}) recovers Eq.(6.6) of Ref.\cite{Koyama:2011wx}. However, note that $f_{\mu\nu}$ in the solution of \cite{Koyama:2011wx} does not respect the FRW symmetry. }. \begin{figure}[ht] \begin{center} \includegraphics[width=0.48\textwidth]{LRangep.eps}~~ \includegraphics[width=0.48\textwidth]{LRangem.eps} \end{center} \caption{Sign of the effective cosmological constant $\Lambda_\pm$ in the positive (left panel) and negative (right panel) branches. In the red (green) region with $+45^\circ$ ($-45^\circ$) lines, $\Lambda_{\pm}$ is positive (negative). The white region and the dotted squared region correspond to $1+\alpha_3+\alpha_3^2-\alpha_4<0$ and $X_{\pm}<0$, respectively, and are excluded since the cosmological solutions (\ref{eq:fsolcosmo}) do not exist there. Along the dotted black line (defining the boundary between the red and green regions), $\Lambda_\pm =0$ and the background solution reduces to the GR one. The solid line corresponds to $X_{\pm}=0$ and thus defines one of the boundaries between the allowed (red or green) and excluded (dotted squared) regions. Along the dashed line, both $X_{\pm}$ and $\Lambda_{\pm}$ diverge, and it defines another boundary between the allowed (red or green) and excluded (dotted squared) regions.} \label{fig:Lambdarange} \end{figure} \section{Perturbations in general setup} \label{sec:perturbations} In this section we consider the graviton mass term $I_{mass}[g_{\mu\nu},f_{\mu\nu}]$ defined by (\ref{eqn:Imass})-(\ref{Kdef}) and (\ref{eqn:fmunu})-(\ref{eqn:Omegaij}), but with an arbitrary value of $K$ and arbitrary functions $n(\varphi^0)$ and $\alpha(\varphi^0)$. We shall develop a formalism to analyze perturbations of this generalized system around flat ($K=0$), closed ($K>0$) and open ($K<0$) FRW universes. Cosmological implications of this type of generalized massive gravity will be discussed in future publication. For the background we adopt the physical metric of the FRW form (\ref{eqn:FRWbackground}) with (\ref{eqn:unitary-gauge}), but with general $K$, $n(\varphi^0)$ and $\alpha(\varphi^0)$. Without loss of generality, we assume that $N>0$, $n>0$, $a>0$ and $\alpha>0$ at least in the vicinity of the time of interest, where $N$ and $a$ are the lapse function and the scale factor of the background FRW physical metric. (Otherwise, we consider $|N|$, $|n|$, $|a|$ and $|\alpha|$ and rename them as $N$, $n$, $a$ and $\alpha$.) As reviewed in the previous section for open universes in the case of the Minkowski fiducial metric and as shown in Appendix~\ref{app:gravitonmass} for general cases with arbitrary $K$, $n(\varphi^0)$ and $\alpha(\varphi^0)$, the background equation of motion for the St\"uckelberg fields $\varphi^a$ has three branches of solutions. One of them does not allow nontrivial cosmologies and thus is not of our interest. The other two branches of solutions allow nontrivial cosmologies and are given by (\ref{eq:fsolcosmo}) even for general $K$, $n(\varphi^0)$ and $\alpha(\varphi^0)$. In this section we then consider perturbations of the physical metric and the St\"uckelberg fields around the FRW solutions in these nontrivial branches. \subsection{Exponential map and Lie derivative} Since the fiducial metric $f_{\mu\nu}$ is defined as in (\ref{eqn:fmunu}) without referring to the physical metric $g_{\mu\nu}$, let us begin with perturbations of the St\"uckelberg fields $\varphi^a$. We define perturbations $\pi^a$ of $\varphi^a$ through the so-called exponential map. Actually, since the action will be expanded only up to the quadratic order, we can truncate the exponential map at the second order. We thus define $\pi^a$ by \begin{equation} \varphi^a = x^a + \pi^a + \frac{1}{2}\pi^b\partial_b\pi^a + O(\epsilon^3),\label{eqn:exponentialmap} \end{equation} or equivalently, \begin{equation} \pi^a = (\varphi^a-x^a) - \frac{1}{2}(\varphi^b-x^b)\partial_b(\varphi^a-x^a) + O(\epsilon^3). \end{equation} Here, $\epsilon$ is a small number counting the order of perturbative expansion: $\pi^a=O(\epsilon)$ and $\varphi^a-x^a=O(\epsilon)$. By substituting the expansion (\ref{eqn:exponentialmap}) to the definition of the fiducial metric \begin{equation} f_{\mu\nu} = \bar{f}_{ab}(\varphi^c) \partial_{\mu}\varphi^a\partial_{\nu}\varphi^b, \end{equation} where \begin{equation} \bar{f}_{00}(\varphi^c) = -n^2(\varphi^0), \quad \bar{f}_{0i}(\varphi^c) = \bar{f}_{i0}(\varphi^c) = 0, \quad \bar{f}_{ij}(\varphi^c) = \alpha^2(\varphi^0)\Omega_{ij}(\varphi^k), \label{eqn:fbarab} \end{equation} we obtain \begin{equation} f_{\mu\nu} = \bar{f}_{\mu\nu}(x^{\rho}) + {\cal L}_{\pi}\bar{f}_{\mu\nu}(x^{\rho}) + \frac{1}{2}\left({\cal L}_{\pi}\right)^2 \bar{f}_{\mu\nu}(x^{\rho}) + O(\epsilon^3). \label{eqn:fmunu-expansion} \end{equation} Here, ${\cal L}_{\pi}$ represents the Lie derivative along $\pi^{\mu}$. Actually, this formula is not restricted to (\ref{eqn:fbarab}) but holds for any $\bar{f}_{ab}(\varphi^c)$. \subsection{St\"uckelberg fields and gauge invariant variables} \label{subsec:gauge-invariant-variables} We define perturbations $\phi$, $\beta_i$ and $h_{ij}$ of the physical metric by \begin{eqnarray} g_{00} & = & -N^2(t)\left[1+2\phi\right], \nonumber\\ g_{0i} & = & N(t)a(t)\beta_i\nonumber\\ g_{ij} & = & a^2(t)\left[\Omega_{ij}(x^k)+h_{ij}\right]. \end{eqnarray} We suppose that $\phi,\beta_i,h_{ij}=O(\epsilon)$. Under the linear gauge transformation \begin{equation} x^{\mu} \to x^{\mu} + \xi^{\mu}, \quad (\xi^{\mu}=O(\epsilon)) \label{eqn:gaugetr} \end{equation} each variable transforms as \begin{eqnarray} \pi^0 & \to & \pi + \xi^0, \nonumber\\ \pi_i & \to & \pi_i + \xi_i, \nonumber\\ \phi & \to & \phi + \frac{1}{N}\partial_t(N\xi^0), \nonumber\\ \beta_i & \to & \beta_i - \frac{N}{a}D_i\xi^0 + \frac{a}{N}\dot{\xi}_i, \nonumber\\ h_{ij} & \to & h_{ij} + D_i\xi_j + D_j\xi_i + 2NH\xi^0\Omega_{ij}, \end{eqnarray} where $H$ is the Hubble expansion rate as defined in (\ref{eqn:def-Hubble}), \begin{equation} \pi_i \equiv \Omega_{ij}\pi^j, \quad \xi_i \equiv \Omega_{ij}\xi^j, \end{equation} and $D_i$ is the spatial covariant derivative compatible with $\Omega_{ij}$. We then define gauge invariant variables \begin{eqnarray} \phi^{\pi} & \equiv & \phi - \frac{1}{N}\partial_t(N\pi^0), \nonumber\\ \beta^{\pi}_i & \equiv & \beta_i + \frac{N}{a}D_i\pi^0 - \frac{a}{N}\dot{\pi}_i, \nonumber\\ h^{\pi}_{ij} & \equiv & h_{ij} - D_i\pi_j - D_j\pi_i - 2NH\pi^0\Omega_{ij}. \label{eqn:phipi-betapi-hpi} \end{eqnarray} For later convenience, let us decompose $\beta^{\pi}_i$ and $h^{\pi}_{ij}$ as \begin{eqnarray} \beta^{\pi}_i & = & D_i\beta^{\pi} + S^{\pi}_i, \nonumber\\ h^{\pi}_{ij} & = & 2\psi^{\pi}\Omega_{ij} + \left(D_iD_j-\frac{1}{3}\Omega_{ij}\triangle\right)E^{\pi} + \frac{1}{2}(D_iF^{\pi}_j+D_jF^{\pi}_i) + \gamma_{ij}, \label{eqn:decompose-hpi} \end{eqnarray} where $S^{\pi}_i$ and $F^{\pi}_i$ are transverse, and $\gamma_{ij}$ is transverse and traceless: \begin{equation} D^iS^{\pi}_i = D^iF^{\pi}_i = 0, \quad \quad D^i\gamma_{ij} = 0, \quad \Omega^{ij}\gamma_{ij} = 0, \end{equation} and $D^i\equiv\Omega^{ij}D_j$. \subsection{Graviton mass term} At the FRW background level, the graviton mass term acts as an effective cosmological constant $\Lambda_{\pm}$ shown in (\ref{lambdapm}). The proof of this statement is presented in Appendix~\ref{app:gravitonmass} for arbitrary $K$, $n(\varphi^0)$ and $\alpha(\varphi^0)$. Thus, calculations are expected to be simplified if we add $M_{Pl}^2\int d^4x\sqrt{-g}\Lambda_{\pm}$ to $I_{mass}$ before performing perturbative expansion. For this reason, we define \begin{equation} \tilde{I}_{mass}[g_{\mu\nu},f_{\mu\nu}] \equiv I_{mass}[g_{\mu\nu},f_{\mu\nu}] +M_{Pl}^2\int d^4x\sqrt{-g}\, \Lambda_{\pm}, \label{eqn:def-Itildemass} \end{equation} and expand it instead of $I_{mass}$ itself. As shown explicitly in Appendix~\ref{app:gravitonmass}, upon using the background equation of motion for the St\"uckelberg fields but without using the background equation of motion for the physical metric, the graviton mass term can be expanded up to the quadratic order as \begin{eqnarray} \tilde{I}_{mass} & = & \tilde{I}_{mass}^{(0)} + \tilde{I}_{mass}^{(2)}[h^{\pi}_{ij}] + O(\epsilon^3), \nonumber\\ \tilde{I}_{mass}^{(2)}[h^{\pi}_{ij}] & = & \frac{M_{Pl}^2}{8}\int d^4x Na^3\sqrt{\Omega}\, M_{GW}^2 \left[ (h^{\pi})^2-h_{\pi}^{ij}h^{\pi}_{ij}\right], \end{eqnarray} where the zero-th order part $\tilde{I}_{mass}^{(0)}$ is independent of perturbations, \begin{eqnarray} M_{GW}^2 & \equiv & \pm (r-1)m_g^2\, X_{\pm}^2\sqrt{1+\alpha_3+\alpha_3^2-\alpha_4},\label{eqn:MGW2}\\ r & \equiv & \frac{na}{N\alpha} = \frac{1}{X_{\pm}}\frac{H}{H_f}, \quad H \equiv \frac{\dot{a}}{Na}, \quad H_f \equiv \frac{\dot{\alpha}}{n\alpha}, \nonumber \end{eqnarray} $X_{\pm}$ is given by (\ref{eq:fsolcosmo}), and \begin{equation} h^{\pi} \equiv \Omega^{ij}h^{\pi}_{ij}, \quad h_{\pi}^{ij} \equiv \Omega^{ik}\Omega^{jl}h^{\pi}_{kl}. \end{equation} With the decomposition of $h^{\pi}_{ij}$ in (\ref{eqn:decompose-hpi}), the quadratic mass term is expanded as \begin{eqnarray} \tilde{I}_{mass}^{(2)} & = & M_{Pl}^2\int d^4x Na^3\sqrt{\Omega}\, M_{GW}^2 \nonumber\\ & & \times \left[3(\psi^{\pi})^2 - \frac{1}{12}E^{\pi}\triangle(\triangle+3K)E^{\pi} + \frac{1}{16}F_{\pi}^i(\triangle+2K)F^{\pi}_i - \frac{1}{8}\gamma^{ij}\gamma_{ij} \right], \label{eqn:Imass2} \end{eqnarray} where \begin{equation} F_{\pi}^i \equiv \Omega^{ij}F^{\pi}_{j}, \quad \gamma^{ik} \equiv \Omega^{jl}\Omega^{jl}\gamma_{kl}. \end{equation} What is important here is that the quadratic part $\tilde{I}_{mass}^{(2)}$ is gauge-invariant and depends only on $h^{\pi}_{ij}$, or equivalently ($\psi^{\pi}$, $E^{\pi}$, $F^{\pi}_i$, $\gamma_{ij}$). In particular, it does not contribute to the equations of motion for $\phi$ and $\beta_i$. We note that $M_{GW}^2$ vanishes or diverges for some special values of the parameters ($\alpha_3$, $\alpha_4$): \begin{eqnarray} \alpha_4 = -3\,(1+\alpha_3),\quad \pm(\alpha_3 +2)>0 & \Longrightarrow & {M}^2_{GW} =0\,,\nonumber\\ \alpha_4 = 1+\alpha_3 +\alpha_3^2 \quad & \Longrightarrow & {M}^2_{GW} =0\,, \label{MGWzero}\\ \alpha_4\to -\alpha_3, \quad \pm(1+\alpha_3)>0 \quad & \Longrightarrow & |{M}^2_{GW}| \to \infty, \nonumber \end{eqnarray} where the $\pm$ signs are for the $\pm$ branches, respectively. In the following we suppose that the parameters ($\alpha_3$, $\alpha_4$) take generic values away from the special values shown in (\ref{MGWzero}). \subsection{Matter perturbations and gauge-invariant variables} Let us divide matter fields $\sigma_I$ ($I=1,2,\cdots$) into the background values $\sigma^{(0)}_I$ and perturbations as \begin{equation} \sigma_I = \sigma^{(0)}_I + \delta\sigma_I. \label{eqn:def-delta-sigma} \end{equation} We suppose that $\{\sigma_I\}$ forms a set of mutually independent physical degrees of freedom. Otherwise, we consider a subset of the original $\{\sigma_I\}$ consisting of independent physical degrees of freedom and rename it as $\{\sigma_I\}$. We can construct gauge-invariant variables $Q_I$ from $\delta\sigma_I$ and metric perturbations, without referring to the St\"uckelberg fields. For illustrative purpose let us decompose $\beta_i$, $h_{ij}$ and $\xi_i$ as \begin{eqnarray} \beta_i & = & D_i\beta + S_i, \nonumber\\ h_{ij} & = & 2\psi\Omega_{ij} + \left(D_iD_j-\frac{1}{3}\Omega_{ij}\triangle\right)E + \frac{1}{2}(D_iF_j+D_jF_i) + \gamma_{ij}, \nonumber\\ \xi_i & = & D_i\xi + \xi^T_i, \end{eqnarray} where $S_i$, $F_i$ and $\xi^T_i$ are transverse, and $\triangle$ is the Laplacian associated with $\Omega_{ij}$: \begin{equation} D^iS_i = D^iF_i = D^i\xi^T_i = 0, \quad \triangle \equiv D^iD_i. \end{equation} Under the gauge transformation (\ref{eqn:gaugetr}), each component of the physical metric perturbation transforms as \begin{eqnarray} \phi & \to & \phi + \frac{1}{N}\partial_t(N\xi^0), \nonumber\\ \beta & \to & \beta -\frac{N}{a}\xi^0 + \frac{a}{N}\dot{\xi}, \nonumber\\ \psi & \to & \psi + NH\xi^0 + \frac{1}{3}\triangle\xi, \nonumber\\ E & \to & E + 2\xi, \nonumber\\ S_i & \to & S_i + \frac{a}{N}\dot{\xi}^T_i, \nonumber\\ F_i & \to & F_i + 2\xi^T_i, \nonumber\\ \gamma_{ij} & \to & \gamma_{ij}. \end{eqnarray} Noting that the vector $Z^{\mu}$ defined by \begin{eqnarray} Z^0 = -\frac{a}{N}\beta + \frac{a^2}{2N^2}\dot{E}, \quad Z^i = \frac{1}{2}\Omega^{ij}(D_jE+F_j) \end{eqnarray} transforms as \begin{equation} Z^{\mu} \to Z^{\mu} + \xi^{\mu}\,, \end{equation} we can construct the following gauge-invariant variables out of matter perturbations and physical metric perturbations: \begin{eqnarray} Q_I & \equiv & \delta\sigma_I-{\cal L}_Z\sigma^{(0)}_I, \nonumber\\ \Phi & \equiv & \phi - \frac{1}{N}\partial_t(NZ^0), \nonumber\\ \Psi & \equiv & \psi - NHZ^0 - \frac{1}{6}\triangle E, \nonumber\\ B_i & \equiv & S_i - \frac{a}{2N}\dot{F}_i\,, \end{eqnarray} and $\gamma_{ij}$ is gauge-invariant by itself. In the above, ${\cal L}_Z$ is the Lie derivative along $Z^{\mu}$. Those gauge-invariant variables defined here and in Subsection~\ref{subsec:gauge-invariant-variables}, i.e. $\{Q_I$, $\Phi$, $\Psi$, $B_i$, $\gamma_{ij}$, $\phi^{\pi}$, $\beta^{\pi}$, $S^{\pi}_i$, $\psi^{\pi}$, $E^{\pi}$, $F^{\pi}_i\}$, are not independent. Indeed, it is easy to show that \begin{eqnarray} \phi^{\pi} & = & \Phi + \frac{1}{N}\partial_t \left[\frac{1}{H} \left(\psi^{\pi}-\Psi-\frac{1}{6}\triangle E^{\pi}\right) \right], \nonumber\\ \beta^{\pi} & = & -\frac{1}{aH}\left(\psi^{\pi}-\Psi-\frac{1}{6}\triangle E^{\pi}\right) + \frac{a}{2N}\dot{E}^{\pi}, \nonumber\\ S^{\pi}_i & = & B_i + \frac{a}{2N}\dot{F}^{\pi}_i. \end{eqnarray} There are no more independent relations among gauge-invariant variables defined here and in Subsection~\ref{subsec:gauge-invariant-variables}.\footnote{ See also the sentence just after (\ref{eqn:def-delta-sigma}).} Therefore, we have the following set of independent gauge-invariant variables. \begin{equation} \{Q_I,\ \Phi,\ \Psi,\ B_i, \ \gamma_{ij}, \ \psi^{\pi}, \ E^{\pi}, \ F^{\pi}_i\}. \label{eqn:independent-variables} \end{equation} Based on their origins, we can divide this set of independent gauge-invariant variables into two categories as \begin{equation} \{Q_I,\ \Phi,\ \Psi,\ B_i, \ \gamma_{ij}\} \quad \mbox{and} \quad \{\psi^{\pi}, \ E^{\pi}, \ F^{\pi}_i\}. \end{equation} The first category consists of those gauge-invariant variables that originate from the physical metric $g_{\mu\nu}$ and the matter fields $\{\sigma_I\}$. Thus, those in the first category already exist in GR coupled to the same matter content. On the other hand, those in the second category are physical degrees of freedom associated with the four St\"uckelberg fields $\varphi^a$. \subsection{Structure of total quadratic action} Let us now define \begin{equation} \tilde{I}[g_{\mu\nu},\sigma_I] \equiv I_{EH,\tilde{\Lambda}}[g_{\mu\nu}] + I_{matter}[g_{\mu\nu},\sigma_I],\quad \tilde{\Lambda} \equiv \Lambda+\Lambda_{\pm}, \end{equation} so that \begin{equation} I = \tilde{I}[g_{\mu\nu},\sigma_I] + \tilde{I}_{mass}[g_{\mu\nu},f_{\mu\nu}]. \label{eqn:I=Itilde+Itildemass} \end{equation} Since $\tilde{I}_{mass}$ was already shown to be gauge-invariant up to the quadratic order, (\ref{eqn:I=Itilde+Itildemass}) implies that $\tilde{I}$ is also gauge-invariant up to that order. Thus the quadratic part $\tilde{I}^{(2)}$ of $\tilde{I}$ can be written in terms of gauge-invariant variables constructed solely from perturbations of the physical metric perturbations ($\phi$, $\beta_i$, $h_{ij}$) and matter perturbations $\delta\sigma_I$, i.e. $\{Q_I$, $\Phi$, $\Psi$, $B_i$, $\gamma_{ij}\}$. Therefore, the total quadratic action has the following structure. \begin{equation} I^{(2)} = \tilde{I}^{(2)}[Q_I, \Phi, \Psi, B_i, \gamma_{ij}] + \tilde{I}^{(2)}_{mass} [\psi^{\pi},\ E^{\pi}, \ F^{\pi}_i, \gamma_{ij}], \end{equation} where the explicit form of $\tilde{I}^{(2)}_{mass}$ is shown in (\ref{eqn:Imass2}). As already stated, gauge-invariant variables listed in (\ref{eqn:independent-variables}) are independent from each other. Note that $\psi^{\pi}$, $E^{\pi}$ and $F^{\pi}_i$ do not have kinetic terms but have nonvanishing masses, provided that the parameters ($\alpha_3$, $\alpha_4$) take generic values away from the special values shown in (\ref{MGWzero}). Thus, we can integrate them out: their equations of motion lead to \begin{equation} \psi^{\pi} = E^{\pi} = 0, \quad F^{\pi}_i = 0, \end{equation} and then \begin{equation} I^{(2)} = \tilde{I}^{(2)}[Q_I, \Phi, \Psi, B_i, \gamma_{ij}] - \frac{M_{Pl}^2}{8}\int d^4x Na^3\sqrt{\Omega} M_{GW}^2\gamma^{ij}\gamma_{ij}, \end{equation} where $M_{GW}^2$ is given by (\ref{eqn:MGW2}). For scalar and vector modes, this quadratic action is exactly the same as that in GR with the matter content $\{\sigma_I\}$. \subsection{Gravitational waves with time-dependent mass} The total quadratic action for the tensor sector is \begin{equation} I^{(2)}_{tensor} = \frac{M_{Pl}^2}{8}\int d^4x Na^3\sqrt{\Omega} \left[\frac{1}{N^2}\dot{\gamma}^{ij}\dot{\gamma}_{ij} + \frac{1}{a^2}\gamma^{ij}(\triangle-2K)\gamma_{ij} -M_{GW}^2\gamma^{ij}\gamma_{ij}\right], \label{eqn:I2tensor} \end{equation} provided that there is no tensor-type contribution from the quadratic part of $I_{matter}$. In this way the dispersion relation of gravitational waves is modified. The squared mass of gravitational waves $M_{GW}^2$ is given by (\ref{eqn:MGW2}) and is time-dependent. If $M_{GW}^2$ is negative then long wavelength gravity waves exhibit linear instability. For generic values of parameters ($\alpha_3$, $\alpha_4$) away from the special values shown in (\ref{MGWzero}), we see from the formula (\ref{eqn:MGW2}) that the sign of $M_{GW}^2$ is the same as the sign of the combination $\pm(r-1)m_g^2$, where $\pm$ signs correspond to $\pm$ branches, respectively. \section{An example: scalar matter field and Minkowski fiducial} \label{sec:example-scalar} In the previous section we have analyzed quadratic action for perturbations around nontrivial FRW backgrounds, with a general FRW fiducial metric and a general matter content. For scalar and vector modes, we have shown that the quadratic action is exactly the same as that in GR with the matter content. For tensor modes, on the other hand, we have seen that gravitational waves obtain a time-dependent mass. In this section, in order to illustrate these results a little more explicitly, we shall consider a simple example consisting of the massive gravity with Minkowski fiducial metric, coupled to a canonical scalar matter field with potential $V$. The total action of this system is \begin{equation} I = \int d^4x\sqrt{-g} \left[M_{Pl}^2\left(\frac{R}{2} + m_g^2( {\cal L}_2+\alpha_3{\cal L}_3+\alpha_4{\cal L}_4) \right) -\frac{1}{2}\,\partial_\mu \sigma \partial^\mu \sigma- V(\sigma)\right]\,, \end{equation} where ${\cal L}_{1,2,3}$ are the graviton mass terms defined in (\ref{lag234}). In the case of the Minkowski fiducial, the only nontrivial FRW background is the open FRW solution found in \cite{Gumrukcuoglu:2011ew} and reviewed in Sec.~\ref{setup}. Thus, in this section the curvature constant $K$ is set to be negative and the form of the fiducial metric is specified by (\ref{eqn:fmunu})-(\ref{eqn:n-alpha-Minkowski}). For the FRW background (\ref{etaline})-(\ref{eqn:FRWbackground}) with $K<0$ and $\sigma=\sigma^{(0)}(t)$, the equations of motion read \begin{eqnarray} &&3\,H^2 +\frac{ 3\,K}{a^2} = \Lambda_\pm+\frac{1}{M_{Pl}^2}\left[\frac{(\dot{\sigma}^{(0)})^2}{2\,N^2}+V(\sigma^{(0)})\right]\,, \nonumber\\ && -\frac{2\dot{H}}{N} +\frac{ 2\,K}{a^2} = \frac{(\dot{\sigma}^{(0)})^2}{M_{Pl}^2\,N^2} \nonumber\\ && \frac{1}{N}\partial_t\left(\frac{\dot{\sigma}^{(0)}}{N}\right) + \frac{3H}{N}\dot{\sigma}^{(0)} + V'(\sigma^{(0)}) = 0. \label{bgeqs} \end{eqnarray} We now introduce perturbations to the metric $g_{\mu\nu}$, the four St\"uckelberg fields $\varphi^a$ and the scalar matter field $\sigma$. We will be developing the perturbation theory without specifying a gauge and in the end we will switch to gauge invariant perturbations, as we have already done in the previous section in a more general setup. The total quadratic action before switching to gauge invariant perturbations is presented in Appendix~\ref{app:acttot}. We then adopt the decomposition of the form \begin{eqnarray} &&\qquad\quad \pi_i= D_i \pi + \pi_i^T \,, \qquad \beta_i = D_i\beta + S_i\,, \nonumber\\ h_{ij} &=& 2\,\psi \,\Omega_{ij} + \left(D_i D_j -\frac{1}{3}\,\Omega_{ij} \triangle\right) E + \frac{1}{2}\left(D_i F_j + D_j F_i\right)+\gamma_{ij}\,, \label{decomp} \end{eqnarray} where $\pi^T_i$, $S_i$ and $F_i$ are transverse: \begin{equation} D^i \pi_i^T = D^i S_i = D^i F_i = 0\,. \end{equation} \subsection{Tensor sector} We start by considering the tensor sector. We use the decomposition (\ref{decomp}) in the total action (\ref{totact})-(\ref{totlag}), keeping only the transverse traceless mode $\gamma_{ij}$. Since the matter sector has no tensor degrees, the action is the same as the one given in (\ref{eqn:I2tensor}) with (\ref{eqn:MGW2}) and for the Minkowski fiducial, we have \begin{equation} r = \frac{aH}{\sqrt{-K}}. \end{equation} Notice that in the accelerating universe, $r$ and thus $M^2_{GW}$ grow at late time. Switching to conformal time $d\eta \equiv N\,dt / a$ and defining the canonical fields \begin{equation} \bar{\gamma}_{ij} \equiv \frac{M_{Pl}\,a}{2}\,\gamma_{ij}\,, \end{equation} the tensor action takes the form \begin{equation} I^{(2)}_{tensor} = \frac{1}{2}\int d^3x\,d\eta\,\sqrt{\Omega}\, \left[\bar{\gamma}'_{ij}\bar{\gamma}^{\prime\,ij} + \bar{\gamma}^{ij} \left(\triangle-2K+\frac{a''}{a}\right)\bar{\gamma}_{ij} - a^2M_{GW}^2\bar{\gamma}^{ij}\bar{\gamma}_{ij} \right]\,, \end{equation} where a prime denotes differentiation with respect to conformal time. Next, we use harmonic expansion through \begin{equation} \bar{\gamma}_{ij} = \int k^2 dk \,\bar{\gamma}_{\vec{k}} \, Y_{ij}(\vec{k},\vec{x})\,, \end{equation} where $Y_{ij}(\vec{k},\vec{x})$ is the tensor harmonic satisfying \begin{equation} \left(\triangle +\vec{k}^2\right)Y_{ij} = 0\,,\quad D^i Y_{ij} = 0, \quad \Omega^{ij}Y_{ij} =0, \quad (\vec{k}^2 \equiv \Omega_{ij}\vec{k}^i\vec{k}^j) \end{equation} with $\vec{k}^2 \ge |K|$ taking continuous values. Suppressing the momentum index, we obtain the equation of motion \begin{equation} \bar{\gamma}'' + \left(\vec{k}^2 - \frac{a''}{a} +2\,K + a^2\,M^2_{GW}\right) \bar{\gamma} =0\,. \end{equation} As we showed in the previous section for a generic setup, the tensor mode acquires a mass contribution whose time dependence is determined by the fiducial metric (which is Minkowski in the present example). \subsection{Vector sector} We now move on to the vector sector. Keeping only the transverse vector modes in the decomposition (\ref{decomp}), the total action (\ref{totact})-(\ref{totlag}) reduces to \begin{equation} I^{(2)}_{vector}=\frac{M_{Pl}^2}{8} \int d^4xNa^3\sqrt{\Omega}\, {\cal L}_{vector} \label{d2Ldef} \end{equation} with \begin{eqnarray} {\cal L}_{vector} &=& \frac{1}{2\,N^2} \left(D^i \dot{F}^j D_i \dot{F}_j - 2\,K \dot{F}^i \dot{F}_i\right) -\frac{2}{N\,a} \left(D^i \dot{F}^j D_i S_j-2\,K \dot{F}^i S_i\right) \nonumber\\ && -\frac{1}{2}\,M^2_{GW}\left(D^i F^j D_i F_j - 2\,K F^i F_i\right) + \frac{2}{a^2}\,D^i S^j D_iS_j -\frac{4\,K}{a^2}\,S_iS^i \nonumber\\ && -2\,M_{GW}^2 \left(\triangle\pi^T_i+2\,K\,\pi^T_i\right)\left(F^i-\pi_T^i\right) \,, \end{eqnarray} where \begin{equation} S^i \equiv \Omega^{ij}S_j, \quad F^i \equiv \Omega^{ij}F_j, \quad \pi_T^i \equiv \Omega^{ij}\pi^T_j. \end{equation} We then switch to gauge invariant perturbations, as defined in (\ref{eqn:phipi-betapi-hpi})-(\ref{eqn:decompose-hpi}), and obtain \begin{eqnarray} {\cal L}_{vector} &=& \frac{1}{2\,N^2} \left(D^i \dot{F}_{\pi}^j D_i \dot{F}^{\pi}_j-2\,K \dot{F}_{\pi}^i \dot{F}^{\pi}_i\right) -\frac{2}{N\,a} \left(D^i \dot{F}_{\pi}^j D_i S^{\pi}_j-2\,K \dot{F}_{\pi}^i S^{\pi}_i\right) \nonumber\\ && -\frac{1}{2}\,M_{GW}^2 \left(D^i F_{\pi}^j D_i F^{\pi}_j - 2\,K F_{\pi}^i F^{\pi}_i\right) + \frac{2}{a^2}\,D^i S_{\pi}^j D_iS^{\pi}_j -\frac{4\,K}{a^2}\,S_{\pi}^i S^{\pi}_i\,, \end{eqnarray} which is manifestly gauge invariant. Varying this action with respect to $S^{\pi}_i$ yields an algebraic equation for $S^{\pi}_i$, which can be solved by \begin{equation} S^{\pi}_i = \frac{a}{2\,N}\,\dot{F}^{\pi}_i \,. \end{equation} Using this solution back in the action, we get \begin{equation} I^{(2)}_{vector} =\frac{M_{Pl}^2}{16}\int d^4xNa^3\sqrt{\Omega}\, M^2_{GW} \,F_{\pi}^i(\triangle+2K)F^{\pi}_i\,. \end{equation} This clearly shows that the kinetic term for vector perturbation vanishes at quadratic order. Provided that the parameters ($\alpha_3$, $\alpha_4$) take generic values away from the special values shown in (\ref{MGWzero}), the equation of motion for $F^{\pi}_i$ leads to $F^{\pi}_i=0$ and then $I^{(2)}_{vector}=0$. \subsection{Scalar sector} Finally, we consider the scalar perturbations. Using the decomposition (\ref{decomp}) in the total action (\ref{totact})-(\ref{totlag}), we obtain the action for the scalar sector as \begin{equation} I^{(2)}_{scalar} = \frac{M_{Pl}^2}{2} \int d^4x Na^3\sqrt{\Omega}\, {\cal L}_{scalar}\,, \end{equation} with \begin{eqnarray} {\cal L}_{scalar}&=& \frac{1}{M_{Pl}^2}\left[\frac{1}{N^2}\,\delta\dot{\sigma}^2 - \frac{1}{a^2}\,D_i \delta \sigma D^i \delta \sigma-V''\,\delta\sigma^2 -\frac{2\,\dot{\sigma}}{N^2}\left(\phi-3\,\psi\right)\delta\dot{\sigma} - 2\,V'\,\left(\phi+3\,\psi\right)\delta\sigma\right] \nonumber\\&& +\frac{1}{N^2}\left(\frac{1}{6} (\triangle \dot{E})^2 - \frac{K}{2} D^i \dot{E} D_i \dot{E}-6\,\dot{\psi}^2 \right) +\frac{12\, H}{N}\,\phi\,\dot{\psi}-\frac{12\,K}{a^2}\phi\,\psi \nonumber\\ &&- \frac{1}{6} \left( M^2_{GW}+\frac{K}{a^2}\right)(\triangle E)^2 +\frac{K}{2} M^2_{GW}D_iE\,D^iE -\left(6\, H^2 -\frac{\dot{\sigma}^2}{M_{Pl}^2\,N^2}\right)\phi^2 \nonumber\\&& -\frac{2\,K}{a^2}D_i\beta\,D^i\beta +\frac{2}{a\,N}\,\left(2\,\triangle\dot{\psi}-\frac{1}{3}\,\triangle^2 \dot{E}-K\,\triangle\dot{E} - 2\,N H\triangle\phi +\frac{\dot{\sigma}}{M_{Pl}^2}\,\triangle\delta\sigma \right)\beta \nonumber\\&& -\frac{2}{a^2}\,\left(2\,\triangle\psi-\frac{1}{3}\triangle^2E - K\,\triangle E \right)\phi+6\,\left( M_{GW}^2 - \frac{K}{a^2}\right)\psi^2 \nonumber\\&& +\frac{1}{a^2}\,\left(2\,D_i\psi D^i\psi +\frac{2}{3}\,\triangle \psi \triangle E + \frac{1}{18}\,D_i \triangle E \,D^i \triangle E - 2\, K \,D_i\psi\,D^i E\right) \nonumber\\&& +2\, M^2_{GW}\,\left(K\left(D_i\pi\,D^i\pi-D_iED^i\pi\right)+\frac{1}{3}\triangle E\,\triangle \pi-2\,\psi\,\triangle\pi-6\, H\,N\,\psi\,\pi^0\right) \nonumber\\&& +2\, H\,N^2\, M^2_{GW}\,\left(\frac{2}{N}\,\triangle\pi+3\, H\pi^0\right)\pi^0\,. \end{eqnarray} Hereafter in this subsection, $\sigma$ represents the background value $\sigma^{(0)}(t)$. Next, we switch to gauge invariant variables defined in (\ref{eqn:phipi-betapi-hpi})-(\ref{eqn:decompose-hpi}) and carry out harmonic expansion. We then obtain the equations of motion for the nondynamical degrees as \begin{eqnarray} \phi^{\pi} &=& \frac{1}{H}\,\left(\frac{K}{a}\, \beta^{\pi} +\frac{\dot{\psi}^{\pi}}{N}+\frac{k^2-3\,K}{6\,N}\,\dot{E}^{\pi}+\frac{\dot{\sigma}}{2\,M_{Pl}^2N}\,\delta\sigma^{\pi}\right)\,,\nonumber\\ \beta^{\pi} &=& -\frac{a}{k^2\,H}\,\left[\left(-3\,H^2+\frac{\dot{\sigma}^2}{2\,M_{Pl}^2N^2}\right)\phi^{\pi} + \frac{k^2-3\,K}{a^2}\,\psi^{\pi} + \frac{3\,H}{N}\,\dot{\psi}^{\pi} + \frac{k^2(k^2-3\,K)}{6\,a^2}\, E^{\pi} \right. \nonumber\\ &&\left.\qquad\qquad\qquad-\frac{V'}{2\,M_{Pl}^2}\,\delta\sigma^{\pi} -\frac{\dot{\sigma}}{2\,M_{Pl}^2N^2}\delta\dot{\sigma}^{\pi}\right]\,, \label{nondyn} \end{eqnarray} where \begin{equation} \delta\sigma^{\pi} \equiv \delta\sigma - \dot{\sigma}\pi^0, \end{equation} and substitute their solutions into the action. Defining the analogue of the Sasaki-Mukhanov variable \begin{equation} Q\equiv \delta\sigma^{\pi}-\frac{\dot{\sigma}}{N\,H}\,\left(\psi^{\pi} +\frac{k^2}{6}\,E^{\pi}\right), \label{SasMuk} \end{equation} the resulting quadratic action reads, up to boundary terms, \begin{equation} I^{(2)}_{scalar}=\frac{M_{Pl}^2}{2} \int d^3k\,dt\, a^3\,N \,\left[6\,M_{GW}^2|\psi^{\pi}|^2-\frac{1}{6}\,M_{GW}^2k^2\left(k^2-3\,K\right)| E^{\pi}|^2 + {\cal L}_Q\right]\,, \end{equation} where \begin{equation} {\cal L}_Q\equiv\frac{2\, H^2(k^2-3\,K)}{2\, H^2M_{Pl}^2\,(k^2-3\,K)+K\,\frac{\dot{\sigma}^2}{N^2}}\left(\frac{1}{N^2}\,|\dot{Q}|^2-M_Q^2\,|Q|^2\right), \label{Qact} \end{equation} and \begin{equation} M^2_Q \equiv V'' + \frac{k^2}{a^2}\left(1-\frac{a^2\,\dot{\sigma}^2}{K M_{Pl}^2\,N^2}\right)+\frac{2\,\left(k^2\, H \,\frac{\dot{\sigma}}{N}+K\,V'\right)\,\left[\left(k^2\, H^2-\frac{K^2}{a^2}\right)\frac{\dot{\sigma}}{N}+K\, H\,V'\right] }{K\,M_{Pl}^2\, H\left(2\, H^2\,(k^2-3\,K)+K\frac{\dot{\sigma}^2}{M_{Pl}^2N^2}\right)}\,. \end{equation} Provided that the parameters ($\alpha_3$, $\alpha_4$) take generic values away from the special values shown in (\ref{MGWzero}), the equation of motion for $\psi^{\pi}$ and $E^{\pi}$ lead to $\psi^{\pi}=E^{\pi}=0$ and then \begin{equation} I^{(2)}_{scalar}=\frac{M_{Pl}^2}{2} \int d^3k\,dt\, a^3\,N {\cal L}_Q\,. \label{Qac} \end{equation} As shown in Appendix~\ref{app:comparison}, the action (\ref{Qac}) agrees with the standard results in GR coupled to the same scalar matter field $\sigma$. To summarize, the scalar sector consists of a dynamical degree which evolves exactly like the standard Sasaki-Mukhanov variable in GR, and two more degrees which have infinite mass. \section{Summary and discussions} In the context of the potentially ghost-free, nonlinear massive gravity \cite{deRham:2010kj} and for a general fiducial metric of FRW form, we analyzed linear perturbations around self-accelerating cosmological solutions of arbitrary spatial curvature, populated by generic matter content that is minimally coupled to gravity. By constructing a gauge invariant formulation of perturbations, we found that massive graviton modes in the scalar and vector sectors have vanishing kinetic terms but nonzero mass terms. By integrating them out, we showed that the part of the action quadratic in scalar and vector perturbations is exactly the same as in GR with the same matter content. In other words, the dynamical degrees in the gravity sector comprise only the two gravity wave polarizations. We also found that these acquire a mass whose time dependence is set by the fiducial metric. In Fierz-Pauli theory in de Sitter background, it has been known that the scalar mode among five degrees of freedom of massive spin-$2$ graviton becomes ghost for $2H^2>m_{FP}^2$, where $H$ is the Hubble expansion rate and $m_{FP}$ is the graviton mass~\cite{Higuchi:1986py}. This conclusion does not hold in the nontrivial cosmological branches of the nonlinear massive gravity. Indeed, as stated above, the scalar and vector modes have vanishing kinetic terms and nonzero mass terms for any FRW background. This sharp contrast to the linear (Fierz-Pauli) massive gravity stems from a peculiar structure of the graviton mass term expanded up to the quadratic order in perturbations: it depends only on the ($ij$)-components of metric perturbations and thus are independent of ($00$) and ($0i$)-components. This Lorentz-violating structure is possible because the vev of $f_{\mu\nu}$ in the cosmological branches spontaneously breaks diffeomorphism invariance in a nontrivial way. It is still fair to say that the nature of the cancellation of the kinetic terms in the quadratic action of scalar and vector sectors is not well understood. This may be an indication that these sectors exhibit an infinitely strong coupling. If this is the case then we cannot properly describe the scalar and vector sectors without knowledge of a UV completion. On the other hand, since the modes have non vanishing masses, it may be possible that these are just infinitely heavy modes without low-energy dynamics, as we have assumed in the main part of the present paper. In this case we can safely integrate out those extra modes and trust the resulting low energy effective theory. In order to establish the fate of those extra degrees of freedom, i.e to judge whether they are strongly-coupled or non-dynamical, the linear perturbation theory is not sufficient and nonlinear methods are needed. This study is beyond the scope of the present paper and is left for a future work. On the other hand, the modification in the tensor sector may leave a signature in the stochastic gravitational wave spectrum. The additional term in the mass of the tensor modes is time dependent, while its sign is determined by both the fiducial metric and the cosmological evolution. A positive but large contribution may give rise to a suppression of the gravity waves and a null signal in the large scale tensor-to-scalar ratio. However, this deviation from scale invariance may allow the signal at small scales to be potentially observable in the space-based gravity wave observatories such as DECIGO \cite{DECIGO}, BBO \cite{BBO} and LISA \cite{Danzmann:1997hm}. It is important to note that the present analysis is purely classical and special care is needed when discussing the evolution of cosmological perturbations which start off in quantum mechanical vacuum. In order to address quantum issues such as the radiative stability of the structure of the effective theory describing cosmological perturbations, one of the first important steps is to identify the strong coupling scale below which the nonlinear massive gravity theory can be trusted. In the branch described by the trivial solution to the St\"uckelberg equation of motion (\ref{eqn:constraintf}), the so called decoupling limit has been useful for this purpose. However, since this trivial branch is not compatible with FRW cosmologies, in the present paper we have considered other two branches described by non-trivial solutions (\ref{eq:fsolcosmo}). In these cosmological branches, the usual decoupling limit is not applicable, at least apparently. We thus need to develop a new technique to identify the strong coupling scale, or directly analyze nonlinear dynamics of the whole system. This is certainly one of the most important issues in the future research. \begin{acknowledgments} The authors thank G.~D'Amico, C.~de Rham, N.~Kaloper, K.~Koyama, N.~Tanahashi and A.~J.~Tolley for useful discussions. This work was supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. S.M. also acknowledges the support by Grant-in-Aid for Scientific Research 17740134, 19GS0219, 21111006, 21540278, by Japan-Russia Research Cooperative Program. \end{acknowledgments}
\section{Introduction} Binaries influence the evolution of open clusters and are catalysts for the formation of anomalous stars like blue stragglers (BSs). Stellar dynamicists have long recognized that encounters involving binary stars can supply the inevitable energy flow out of the cores of star clusters. These complex dynamical dances also lead to stellar exchanges, binary orbit evolution, close stellar passages, mass transfer, mergers, and stellar collisions, all of which lead to an array of new stellar evolution paths and products. Observations of the old \citep[7 Gyr;][and see also \citealt{mei09}]{sar99} open cluster NGC 188 reveal a rich binary population and a wide variety of such interaction-product candidates, including BSs, sub-subgiants and X-ray sources \citep[e.g.,][]{bel98,gon05,gel08b,gel09}. This abundance of anomalous stars may be a direct result of the dynamical evolution of the single and binary stars as they interacted over the lifetime of the cluster. The field binaries are an important comparison population to open cluster binaries, as we would not expect field binaries to undergo dynamical encounters due to the low stellar density of the Galactic field. Therefore comparisons between open cluster and field binary populations may provide insights into the impact of stellar dynamics on shaping a binary population and creating exotic stars. There have been a number of studies of the field dwarf binaries in the past \citep[notably, ][]{abt76,duq91}. Currently the most comprehensive and complete survey of field dwarf binaries is that of \citet[][hereafter R10]{rag10}. They analyze the multiplicity of a complete sample of 454 $\sim$F6 - K3 dwarf and subdwarf stars within 25 pc of the Sun, finding a frequency of multiple systems of 46\% (34\% $\pm$ 2\% strictly binaries). The orbital period distribution is fit by a log-normal curve centered on log($P$ [days] ) = 5.03. For binaries with periods longer than the circularization period, R10 find that the orbital eccentricity distribution is roughly flat out to $e \sim 0.6$. The mass-ratio distribution for binaries is found to be roughly uniform between 0.2 and 0.95, with an additional peak at a mass ratio of unity (commonly termed ``twins''). \citet[][hereafter Paper 2]{gel09} present \orbm~solar-type binary orbits in NGC 188 derived from our ongoing radial-velocity (RV) survey of the cluster. These binaries are of similar spectral type to those studied by R10, and thus provide a comparison between field binaries and those that have evolved within the dynamical environment of an open cluster. This work is an integral part of the WIYN\footnote{\footnotesize The WIYN Observatory is a joint facility of the University of Wisconsin - Madison, Indiana University, Yale University, and the National Optical Astronomy Observatories.} Open Cluster Study \citep[WOCS;][]{mat00}. Our observations cover a complete sample of stars within a magnitude range of \magn~(1.1 - 0.9 M$_{\odot}$) and extending to 30 arcmin (17 pc in projection or roughly 13 core radii)\footnote{We adopt a core radius of 1.3 pc \citep{bon05} at a distance of 1.9 kpc, which corresponds to 2.35 arcminutes on the sky.} in radius, essentially out to the tidal radius of the cluster \citep[21 $\pm$ 4 pc;][]{bon05}. Our magnitude limits include solar-type main-sequence (MS) stars, subgiants, giants, and BSs. For most binaries, our observations span $\sim$11 years, while for some binaries our time baseline covers up to 35 years. Our detectable binaries have periods ranging from a few days to of order 10$^4$ days. A full description of our stellar sample, observations, data reduction routine, and precision, as well as a summary table listing relevant information on each star in our sample, can be found in \citet{gel08b} (hereafter, Paper 1). The binary orbital parameters, including estimates for the primary masses, can be found in Paper 2. In this third paper in the series, we study the characteristics of the NGC 188 binary population. We first perform a detailed analysis of our observational completeness (Section~\ref{incomp}), and use the results from this section to study the global hard-binary frequency (Section~\ref{freq}). In Section~\ref{SelogP} we discuss the eccentricity - log period ($e - \log(P)$) diagram. We then analyze the hard-binary distributions in eccentricity (Section~\ref{Sefreq}), period (Section~\ref{SPfreq}), mass ratio and secondary mass (for MS binaries in Section~\ref{Sqm2freq} and BS binaries in Section~\ref{Sqm2freqBS}). Throughout the paper, we compare the MS, giant and BS binary populations. Additionally, we compare the NGC 188 binary population to similar binaries in the Galactic field. In Section~\ref{Hcomp}, we also compare these results to the $N$-body open cluster simulation of \citet{hur05}, and we build on the work of \citet{mat09} and \citet{gel11} to discuss the possible origins of the NGC 188 BS population. Finally, in Section~\ref{summary} we provide a brief summary and conclusions. \section{Completeness in Binary Detection and Orbital Elements} \label{incomp} There are 630 stars with non-zero proper-motion membership probabilities from \citet{pla03} within the magnitude and spatial domain that define our complete stellar sample. Updating the results of Paper 1, we have obtained $\geq$3 RV measurements for 623 of these stars\footnote{We require at least three observations before we categorize a given star as either single or binary. We use the term ``single'' to identify stars with no significant RV variation; a star is termed single if the standard deviation of its RV measurements is less than four times our precision. Certainly some of these stars are also binaries, although generally with longer periods and/or lower total mass than the binaries identified in this study.}. Thus, we have derived RV membership probabilities and characterized the RV variability of 99\% of this complete sample. Two of the seven stars that don't have three observations are too blue to derive reliable RVs with our current observing setup and two are rotating too rapidly (both are W UMa's). The stars with nonzero proper-motion and RV memberships define the stellar sample used in the subsequent analysis, totaling 487 primary stars. Here, we characterize our completeness in binary detection and orbital solutions within this sample. We use a Monte Carlo approach to produce simulated observations of a set of simulated binaries with characteristics of the Galactic field binary population of R10. Specifically we choose the binary periods from a log-normal distribution, centered on log($P$ [days] ) = 5.03 with $\sigma_{log P} = 2.28$. We fit a wide Gaussian to the observed eccentricity distribution in R10 that is centered on $e = 0.39$ with $\sigma_e = 0.31$ ($\chi^2_{red} = 0.78$) and choose the eccentricities from this distribution. (A flat eccentricity distribution fit to the R10 data results in $\chi^2_{red} = 2.28$.) We use 1 M$_{\odot}$~MS primary stars and choose mass ratios ($q$) from a flat distribution such that the secondary stars have masses between 0.08~-~1 M$_{\odot}$. We do not include the peak in the observed field distribution at mass ratios of unity as we detect such binaries as double-lined systems in our spectra and identify them as binaries directly. Once detected, we observe double-lined systems on a more frequent basis than the majority of the binaries in our survey (see Paper 1). The orbital inclinations and phases of the binaries are chosen randomly. To each of the simulated observations of a given simulated binary, we add a random error generated from a Gaussian distribution centered on zero and with a standard deviation equal to our single-measurement precision of 0.4 km s$^{-1}$. As defined in Paper 1, an observed star is considered to be in a binary if the standard deviation of its RV measurements is greater than four times our precision of 0.4 km s$^{-1}$~(or $e/i$~$>$~4). In our survey, we initially followed \citet{mat83} in requiring three observations over the course of at least one year before attempting to categorize a star as being either single or binary. Here, we use our Monte Carlo method to re-evaluate this procedure. We determine our binary detection completeness by analyzing simulated observations of the simulated binary population covering time intervals defined by our actual observations of NGC 188. \begin{figure*}[!t] \plotone{f1.eps} \caption{\footnotesize Observational completeness curves for NGC 188 binaries resulting from our Monte Carlo analysis. We plot our completeness as a function of period (left), eccentricity ($P_{circ} < P \leq 3000$ d; center) and mass ratio (P$\leq 3000$ d; right). In all plots we show our completeness in binary detection with the solid line and in orbital solutions with the dashed line. } \epsscale{1.0} \label{incfig} \end{figure*} Specifically, we have observed the stars in our NGC 188 WIYN sample over the course of about fourteen years. The distribution of the time separation between our first two observations is bimodal with two equally strong peaks, one at about two months and another at about nine months. The distribution of the time separation between our first and third observations is also bimodal, with a strong peak at about eight months and a smaller peak at about 2.5 years. These distributions reflect the approximate time between individual observing runs and the number of stars that we can observe during each observing run (see Paper 1 for a full description). We use these distributions to define time intervals of simulated observations in our Monte Carlo analysis in order to determine our completeness in binary detections. This analysis shows that with our first two observations we should detect 78\% of the binaries with periods less than 10$^3$ days. If we include a third observation, this value increases significantly to 88\%. If we include a fourth observation, our completeness only increases to 90\%. Thus to optimize our efficiency and completeness in detected binaries, we require three observations spanning roughly one year before classifying a star as either single or binary, confirming the results of \citet{mat83}. We plot our binary detection completeness as functions of binary orbital elements in the solid lines in Figure~\ref{incfig}, and we use this analysis to correct our binary frequencies in Section~\ref{freq}. To estimate our incompleteness in binary orbital solutions, we simulate observations of the simulated binaries on the actual observation dates of each of the NGC 188 binaries, respectively. For each combination of a simulated binary with a set of true observation dates, we attempt to derive an orbital solution using an automated version of the same code used in Paper 2. This code iteratively minimizes the residuals between the RV data and the orbital fit to return the best fitting orbital solution. An acceptable orbital solution has RVs covering at least one period, errors on $P$, $e$ and the orbital amplitude $K$ of less than 30\% of the derived values, respectively, an RMS residual velocity of less than 1 km s$^{-1}$, and a range in RVs covering at least 75\% of the orbital amplitude (mainly applicable for highly eccentric binaries). The resulting orbits define our completeness in orbital solutions, and each binary in NGC 188 produces a unique completeness curve in period, eccentricity, etc. The completeness resulting from all combinations of actual observing dates and simulated binaries is plotted in the dashed lines in Figure~\ref{incfig}. These histograms show the fraction of all binaries (within the specified period limits) for which we expect to derive orbital solutions, and hence obtain the orbital parameters. We use this incompleteness analysis to correct the observed distributions in period, eccentricity, secondary mass and mass ratio. Our completeness in detected binaries and orbital solutions decreases dramatically with increasing period (see the left panel of Figure~\ref{incfig}). Longer-period binaries generally have a decreased amplitude of RV variations and require a longer time baseline of observations in order to detect and fully characterize the orbit. We detect 88\% of binaries with periods less than 1000 days, 78\% of binaries with periods less than 3000 days, and 63\% of binaries with periods less than 10$^4$ days. Only a negligible fraction of binaries with periods greater than 10$^4$ days is detectable with our current data. We therefore use the completeness in detected binaries with $P < 10^4$ days in our analysis of the binary frequency in Section~\ref{freq}. Only two of our MS binaries with orbital solutions have periods greater than 3000 days. Therefore for the following analysis of the distributions of orbital parameters, we choose to limit our investigation to binaries with $P < 3000$ days. We show our completeness as a function of eccentricity for binaries with $P_{circ} < P < 3000$ days \citep[$P_{circ}$~=~15.0 days,][]{mat04} in the middle panel of Figure~\ref{incfig}. The completeness in detected binaries is much less sensitive to orbital eccentricity than it is to period. However, our completeness in orbital solutions drops off rapidly at the highest eccentricities. High eccentricity binaries require well-placed observations to define the orbital amplitude and eccentricity, greatly increasing the difficulty in deriving reliable orbital solutions. In the right panel of Figure~\ref{incfig}, we plot our completeness in the mass ratio ($q = M_2/M_1$) for binaries with $P<3000$ days. Because we use a single primary mass of 1 M$_{\odot}$~in this Monte Carlo analysis, the completeness in secondary mass is identical to the completeness in mass ratio shown here. Therefore, this curve also shows our decreasing completeness with decreasing secondary mass, as expected from spectroscopic surveys of this type. The completeness curves in Figure~\ref{incfig} are not independent. Our detectability is most sensitive to period, since shorter-period binaries generally have higher-amplitude orbits. Hence this first bin in period shows the highest completeness, but the completeness is still less than 100\% due to short-period binaries with very high eccentricity and/or very low-mass secondaries. Likewise no bin in eccentricity or mass ratio shows 100\% completeness due in part to long-period binaries found at all mass ratios and eccentricities. We use this analysis to correct our observed MS and giant binary frequencies and distributions of orbital parameters for incompleteness. However we do not apply a completeness correction to the BS sample since we have generally observed the BSs more frequently than the majority of the ``normal'' cluster stars, and the distributions of orbital parameters for the BSs may be significantly different than the field dwarfs and subdwarfs surveyed by R10. There is only one detected binary BS for which we do not have an orbital solution (WOCS ID 8104).\footnote{\footnotesize There is also one proper-motion member in the BS region of the color-magnitude diagram for which we have been unable to derive a reliable RV membership due to rapid rotation and RV variability (WOCS ID 4230). The mean RV is outside of the cluster distribution, but we cannot be certain about membership without an orbital solution (and therefore a center-of-mass RV). We do not include this system in the following analysis due to it's uncertain membership; although we mention it here because it is an X-ray source \citep[S27,][]{gon05} and therefore warrants further observations.} The orbital period for this binary is longer than the baseline of our observations. This system is accounted for in our analysis of the BS binary frequency, but not in the distributions of BS orbital parameters. Finally we note that there are only five non-RV-variable BSs in our sample. If we coarsely estimate their primary masses by their proximity to evolutionary tracks \citep{mar08}, and use these primary masses along with the true observations of these BSs in our Monte Carlo analysis (simplistically applying the same distributions of orbital parameters as observed for field dwarfs), we find only a 24\% chance of having an undetected binary with a period $<10^4$ days in this sample. Moreover we would only expect to have missed at most one BS binary in our data with $P<10^4$ days. Therefore any incompleteness correction must necessarily be minimal. \section{Hard-Binary Frequency} \label{freq} A hard binary is defined as having an internal energy that is much greater than the energy of the relative motion, with respect to the binary, of a single star moving at the velocity dispersion of the cluster \citep{heg75}. Using the method of \citet{gel10} to correct for our measurement precision and undetected binaries, we derive a one-dimensional velocity dispersion in NGC 188 of 0.41~$\substack{+0.09 \\ -0.10}$~km s$^{-1}$~for MS single members of NGC 188. (This value is consistent with that found in Paper 1.) Therefore, for solar-type stars in NGC 188, binaries with periods of less than $\sim$10$^6$ days are hard binaries. As shown in Section~\ref{incomp}, nearly all of our detectable binaries have periods less than 10$^4$ days. Therefore we detect only hard binaries, and in the following, we will use the term ``hard binary'' to refer specifically to the detectable hard binaries in NGC 188. These hard binaries are catalysts for the production of anomalous stars like BSs and X-ray sources \citep[e.g.,][]{bai95,hur05,iva06}, and as such the hard-binary frequency may directly influence the number of such anomalous stars produced within an open cluster. In this section, we compare the hard-binary frequencies of the NGC 188 MS, giant and BS populations. We find 375 MS members including 90 MS binary members\footnote{Here, as in Paper 1, we use our binary member (BM) and binary likely member (BLM) classes to define our binary member sample.}. This observed MS hard-binary frequency is slightly inflated due to the faint magnitude limit defining our stellar sample. The combined light of two MS stars in a binary system can be up to 0.75 magnitudes brighter than the primary star alone. Thus, the faint magnitude limit of our sample includes binary systems with primary stars that would lie below our magnitude limit. Here we discard the four binaries that, after correcting for secondary light, have primaries below the faint limit. We find an observed $V$ magnitude corrected MS hard-binary frequency of 23~$\pm$~2~\% (86/371). As shown in Section~\ref{incomp}, we detect 63\% of the MS binaries with periods of less than 10$^4$ days. The Monte Carlo analysis assumes that these binaries are detectable from the first three observations alone. However continued observations of some stars in our sample have allowed us to detect certain (generally long-period) binaries after further observations, that did not show significant RV variability initially and therefore were not detected as binaries after only the first three observations. Therefore, in order to correct this observed MS binary frequency for our incompleteness, we consider only the 67 binaries that show significant RV variability (i.e., $e/i \geq 4$, as in Paper 1) in their first three observations. When this sample is corrected for incompleteness, we find that the true MS binary frequency among binaries with periods of less than 10$^4$ days is 29~$\pm$~3~\%. \citet{sol10} estimate a MS core binary frequency for NGC 188 of $58.2 \pm 13.5$ \% (over all mass ratios) through analysis of the CMD. To compare with this result, we first limit our sample to cover a similar spatial extent as that of \citet{sol10}. Specifically, we select all stars within 17 arcminutes from the cluster center. (We note that \citet{sol10} included lower-mass primaries than are included in our RV survey, but we do not attempt to correct for this difference in primary mass range.) Within our limited sample, we find a solar-type MS binary frequency of $30 \pm 4$ \% for binaries with periods $< 10^4$ days. If the NGC 188 solar-type MS binary period distribution is consistent with that of similar binaries in the Galactic field, then we can use the R10 field binary period distribution to extend this NGC 188 binary frequency out to the hard-soft boundary of the cluster, of $\sim 10^6$ days. This analysis yields a MS binary frequency of $62 \pm 8$\%, within the given spatial domain and out to the cluster hard-soft boundary, in excellent agreement with the binary frequency derived by \citet{sol10}. The NGC 188 hard-binary frequency also agrees with values derived for other open clusters of a wide range of ages. \citet{mat90} find M67 (4 Gyr) to have a frequency of binaries with periods less than 10$^3$ days of 9\% to 15\%. Matching this period range, we derive a binary frequency of 22~$\pm$~3~\% out to a period of 10$^3$~days in NGC 188, somewhat higher than M67, but only at the margin of significance. The similar WOCS study of the intermediate-aged (2.5 Gyr) open cluster NGC 6819 \citep{hol09} report an observed binary frequency of 17\%, all of which have periods less than 10$^4$ days. We note that this observed binary frequency in NGC 6819 has not yet been corrected for incompleteness. The WOCS study of the young (150 Myr) open cluster M35 finds an incompleteness-corrected MS binary frequency of 24~$\pm$~3\% for binaries with periods $<10^4$ days \citep{gel10}. Additionally, \citet{mer08} find a spectroscopic binary frequency for stars of FGK spectral types of 20\% in Blanco 1 ($\sim$100 Myr). Their observations have a similar time baseline to our NGC 188 survey though with fewer epochs, suggesting a slightly lower detection frequency as found here. Remarkably, these clusters span nearly 7 Gyr in age without a dramatic change in binary frequency. Out of the 454 primary stars in the R10 sample, there are 87 companions with periods less than 10$^4$ days, yielding a field binary frequency within this period range of 19~$\pm$~2~\%. The NGC 188 hard-binary frequency out to the same period cutoff is marginally higher than that of the field. This higher binary frequency in NGC 188 may be a dynamical signature, as $N$-body models predict that single stars will be preferentially lost from a cluster through evaporation as compared to the generally higher total mass binaries \citep{hur05}. Thus NGC 188 may have started with a lower solar-type binary frequency, perhaps very similar to that of the field, that increased throughout the cluster lifetime. Furthermore, the hard-binary frequency in the halo of the cluster (see Figure~\ref{cbf}), where the dynamical relaxation time is long and stellar densities are low, is in good agreement with that of the field. We find 70 giants, 19 of which are in binaries, resulting in an observed giant hard-binary frequency of 27~$\pm$~6~\%. (See Paper 1 for our definition of the giant region in the color-magnitude diagram, CMD.) 14 giant binaries would be detected after the first 3 observations alone, and are therefore applicable to our incompleteness correction. As a star evolves along the giant branch, its increasing physical radius will preclude more and more companions at close separations and short periods. \citet{mer07} find a minimum period of 41.5 days in their sample of red giant binaries in open clusters. Using this as our minimum period in the incompleteness analysis, we find a giant binary frequency of 37~$\pm$~10~\% for binaries with $P<10^4$ days. The minimum period for the giants in our NGC 188 sample is $\sim$11.5 days. Using this as the minimum period results in a binary frequency of 34~$\pm$~9~\% for binaries with $P<10^4$ days. The giant hard-binary frequency derived using either minimum period is consistent with that of the MS within the same period range, and with too large an error to see evolutionary differences in the frequency. \begin{figure}[!t] \plotone{f2.eps} \caption{\footnotesize Incompleteness corrected hard-binary frequency as a function of projected radius, including the main-sequence and giant populations. We remove faint main-sequence binaries with primary stars that would likely fall below our observational limit of $V=16.5$. The vertical error bars show the one sigma Poisson errors on the binary frequencies. The horizontal bars show the bin sizes of 5, 10 and 15 arcminutes, respectively. For main-sequence binaries we use our standard incompleteness correction, and for the giant binaries we apply a minimum period of 11.5 days in our incompleteness correction. We observe a central concentration of binaries as compared to the single stars in the cluster.} \label{cbf} \end{figure} In Figure~\ref{cbf} we plot the incompleteness corrected hard-binary frequency for the MS and giant populations as a function of projected radius from the cluster center. Again, we remove the four faint MS binaries with primaries that likely fall below our $V$ magnitude limit. We exclude the BSs because their spatial distribution may be tied to their formation mechanism(s) rather than dynamical (energy equipartition) effects. We divide our data into three bins of increasing size, covering 5, 10 and 15 arcminutes, respectively, and denoted by the horizontal bars associated with each data point. As shown in Figure~\ref{cbf} the binary frequency clearly decreases with radius from the cluster center. The binary frequency in the core is significantly higher than that of the halo; the first and last bins in Figure~\ref{cbf} are distinct with $>$99\% confidence. This increase in binary frequency reflects the central concentration of the solar-type binaries with respect to the solar-type single stars in NGC 188 (as we also show in Paper 1). In marked contrast to the MS and giant populations, \citet{mat09} find 16 of the 21 BSs to have binary companions. (In Paper 1 we show the region on the CMD from which we have defined our BS population.) These 16 BSs are secure binaries as we have derived orbital solutions for all but one; the remaining, 8104, is unquestionably a binary, with a longer period than our time baseline of observations for that star. Thus, we find a BS hard-binary frequency of 76~$\pm$~19~\%, more than three times and distinct with $>$99\% confidence from the observed MS hard-binary frequency \citep{mat09}. Furthermore, the five BSs that appear to be single may prove to be in longer-period binaries that currently lie outside of our spectroscopic detection limits. As discussed in \citet{mat09}, the large BS binary frequency is not sensitive to our selection criteria. For example, if we take only bright BSs ($V < 15$), the binary frequency is $76 \pm 21$~\% (13/17). Alternatively if we include only those BSs with proper-motion membership probabilities $\geq$90\% the binary frequency is $81 \pm 23$\% (13/16). We also note that there are two stars that lie above the MS turnoff and just fainter than our cutoff for BS selection (IDs 5101 and 8899). Our selection criteria excludes any stars that could be members of normal MS binaries. Thus, normal stars found just fainter than our BS region of the CMD are expected to be double-lined (SB2) binaries. To date we have not detected the flux from any companions in our spectra for either of these two stars, and their RVs do not vary above our threshold for binary detection. Therefore both of these stars are considered single cluster members (SM, see Paper 1), and perhaps should be included in the BS sample. If we include these two stars in the sample the BS binary frequency is $70 \pm 17$~\% (16/23), still significantly higher than the observed MS binary frequency. In Paper 1 we find that the BSs appear to occupy a bimodal spatial distribution, with a centrally concentrated population of 14 BSs and a halo population of 7 BSs. We compare the binary frequencies of the core and halo BS populations, divided at 10 arcmin from the cluster center, and find the core BSs to have a binary frequency of 79~$\pm$~24~\% (11/14) and the halo group to have a binary frequency of 71~$\pm$~32~\% (5/7). The binary frequencies of these two subsets of NGC 188 BSs agree to within their large uncertainties. \section{The \textit{e} - log(\textit{P}) Diagram} \label{SelogP} Tidal circularization for the NGC 188 MS binaries has been well studied by \citet{mat04} and \citet{mei05}. Here we reintroduce the eccentricity versus log period ($e - \log(P)$) diagram to present our new data, and to compare the MS, giant and BS distributions. We show the diagram in Figure~\ref{elogP} for the MS (top), giant (middle) and BS (bottom) populations. Here, as for all such figures in this paper, we differentiate the MS, giant and BS binaries by plotting them in light gray, dark gray and black (green, red and blue in the online version), respectively. For reference, in the MS plot we show with the dotted line the circularization cutoff period, $P_{circ}$ = 15.0 days, found by \citet{mat04} and consistent with the circularization period of 14.5 days found by \citet{mei05}. \begin{figure}[!t] \plotone{f3.CMYK.eps} \caption{\footnotesize Eccentricity plotted against the logarithm of the period for the main-sequence (top), giant (middle) and blue straggler (bottom) populations. For the main-sequence binaries, we plot the tidal circularization cutoff period $P_{circ}$ = 15.0 days \citep{mat04} with the dotted line. For the blue straggler binaries, we identify the core (triangles) and halo (squares) populations.} \label{elogP} \end{figure} The MS and giant $e - \log(P)$ diagrams are of a similar form to those of solar-type binaries in the Galactic field (R10) and open clusters of a wide range in age \citep[see][]{mei05}. The binaries are distributed across a large range in periods, with the short-period binaries ($P<P_{circ}$) having mainly circular orbits and longer-period binaries having a range in non-zero eccentricities. A two-dimensional Kolmogorov-Smirnov (K-S) test (using the technique of \citealt{fas87}) comparing the MS and giant distributions reveals no statistically significant difference, although the expected lack of short-period orbits among the giants is present. However the BS binaries have a significantly different $e - \log(P)$ distribution than the MS (as also discussed in \citealt{mat09}). The majority of the BS binaries have periods on the order of 10$^3$ days. A two-dimensions K-S test results in 99\% confidence that the MS and BS binaries are drawn from different parent populations. This concentration of BS binaries near periods of 10$^3$ days is also observed in the M67 BSs \citep{lat07} as well as in metal-poor field BS binaries \citep{car05}, and may provide significant insights into the origins of the BSs. The mean eccentricity for long-period ($P > 500$ day) BS binaries is 0.28 $\pm$ 0.06, while the mean eccentricity for similar MS binaries is significantly higher, at 0.44 $\pm$ 0.06 \citep{mat09}. Furthermore, three of these twelve long-period BS binaries have eccentricities within two standard deviations of zero, while none of the long-period MS binaries have such low eccentricities. Additionally, in the bottom panel of Figure~\ref{elogP} we divide our BS binaries to compare the core and halo BS populations (again divided at 10 arcmin from the cluster center). We plot the eleven core BS binaries with triangles and the four halo BS binaries with squares. We find no statistically significant distinction in either one- or two-dimensional K-S tests comparing the period and eccentricity distributions of these two populations, though these results are drawn from small sample sizes. Finally, we note that none of our binaries with orbital solutions have periods of less than one day. Indeed our shortest-period MS binary has a period of 3.8 days, and our shortest-period giant binary has a period of 11.5 days. We can understand this lack of short-period giant binaries as an evolutionary effect due to increasing radius. For two solar-mass stars, a period of 10 days corresponds to an orbital semi-major axis of $\sim$25 R$_{\odot}$, which is smaller than the radii of the NGC 188 giants \citep{van99,gir02}. On the other hand the R10 period distribution predicts that we should observe $\sim$6-8 MS binaries with periods below 4 days. It is well known that NGC 188 contains a large number of W UMa systems \citep[e.g.][]{bal85,zha02,kaf03}. There are four known W UMa's within our sample that have non-zero proper-motion memberships, each having a photometric period of less than one day. We have been unable to obtain kinematic orbital solutions for these systems, as their rapid rotation hinders our ability to derive precise RVs. This lack of detached short-period binaries combined with the high frequency of W UMa systems in NGC 188 suggests that we may be observing an evolutionary depletion of close binaries. Perhaps this is a result of orbital angular momentum loss from magnetic braking in stellar winds that has slowly decreased the semi-major axes of the closest binaries \citep{bal85}. Alternatively, short-period binaries in triple systems may have been driven together through the Kozai mechanism and tidal friction \citep{egg06} or dynamical encounters. (We have not yet detected tertiary companions to any of the W UMa's in our sample, though kinematic detection of tertiaries is impeded by the rapid rotation of the W UMa.) \section{Eccentricity Distribution} \label{Sefreq} Here we discuss the eccentricity distribution for binaries with periods between $P_{circ}$ and 3000 days, where $P_{circ} = 15.0$ days \citep{mat04}. In Figure~\ref{efreq} we plot the eccentricity distributions of NGC 188 binaries in both histogram and cumulative distribution form. We plot the detected binaries with orbital solutions in filled histograms. The incompleteness-corrected distributions (for MS and giant binaries) are plotted with the solid-lined histograms above the observed distributions. We choose to bin the MS data with bins of 0.1 and the giant and BS data with bins of 0.15 to reduce the scatter and larger errors associated with the relatively smaller number of giant and BS binaries. In the bottom panel of Figure~\ref{efreq}, we plot the cumulative distributions of these NGC 188 binary populations. \begin{figure}[!t] \plotone{f4.CMYK.eps} \caption{\footnotesize Eccentricity distributions for NGC 188 binaries with $P_{circ} < P < 3000$ days. We show the observed main-sequence, giant and blue straggler binaries in light gray, dark gray and black (green, red and blue in the online version), respectively. The top three plots show the distributions of eccentricity in histogram form. We correct the main-sequence and giant distributions for our incompleteness and plot the corrected histograms with the solid lines above the observed distributions. In each histogram plot, we also show a Gaussian fit to the \citet{rag10} field distribution in the dashed line and a thermal distribution in the dotted line, both normalized to the respective sample size using the field binary frequency within this period range of 11.6\%. The bottom plot shows the cumulative distribution functions for the three populations in the same gray-scale (color in the online version) coding as for the histograms. For clarity, we plot the main-sequence binaries with circles, giant binaries with diamonds, and we distinguish the core and halo blue straggler populations with the triangles and squares, respectively. As a comparison, we also plot the field and thermal distributions corrected for our observational completeness in orbital solutions in the dashed and dotted lines, respectively.} \label{efreq} \end{figure} We correct the MS and giant distributions for incompleteness by applying the results of Section~\ref{incomp}. (Again, we do not account for evolutionary effects in our Monte Carlo incompleteness study; e.g., to account for the increased radius of a giant star). As discussed in Section~\ref{incomp}, we do not attempt to correct the BS distributions for incompleteness as the masses and distributions of orbital parameters for these primaries are likely significantly different from the 1~M$_{\odot}$~primaries and field dwarf binary orbital parameters used in our incompleteness analysis. Again note that there are only six additional BSs in NGC 188, only one of which is detected as a binary. Thus any incompleteness correction for the BS binaries must be minimal. We also show a Gaussian fit to the R10 field eccentricity distribution and a thermal distribution ($f(e) = 2e$) as the dashed and dotted lines, respectively, both normalized by the respective sample sizes using the R10 field binary frequency of 11.6\% in this period range. In the cumulative distribution plot, we show the R10 and thermal distributions modified by our observational completeness (in the corresponding line styles). To produce these curves we multiply the R10 and thermal distributions, respectively, by our incompleteness distributions in orbital solutions (from Figure~\ref{incfig}), effectively limiting these distributions to reflect our completeness level as a function of eccentricity. These completeness-corrected distributions are then converted into the cumulative distributions shown in Figure~\ref{efreq}. \begin{figure}[!t] \plotone{f5.CMYK.eps} \caption{\footnotesize Eccentricity as a function of projected radius from the cluster center. We plot the main-sequence binaries in light-gray circles, giant binaries in dark-gray diamonds and blue straggler binaries in black triangles (core) and squares (halo). (We use green, red and blue for the main-sequence, giant and blue straggler binaries, respectively, in the online version.) Note that we have not derived orbits for any binaries with $e > 0.5$ outside of 5 core radii from the cluster center. When we divide the NGC 188 binaries at 5 core radii ($\sim13$ arcminutes), we find the inner binaries to have a distribution shifted to higher eccentricities at the 97\% confidence level.} \label{evr} \end{figure} The NGC 188 MS eccentricity distribution has a roughly Gaussian form centered on $e = 0.35$. We find no distinction between the MS NGC 188 eccentricity distribution and either the Gaussian fit to the R10 eccentricity distribution or a flat eccentricity distribution. The NGC 188 MS eccentricity distribution also agrees with that of similar samples in open clusters of a wide range in age. The mean eccentricity of similar binaries in both M67 (4 Gyr) and the Pleiades (150 Myr) is $\sim$0.4 \citep{mat90,mer92}. Likewise, the eccentricity distribution of M35 ($\sim$150 Myr) is well approximated by a Gaussian with a mean eccentricity of $\sim$ 0.3 - 0.4 \citep{mat08}. Conversely a K-S test shows that the NGC 188 MS eccentricity distribution is not drawn from a thermal distribution at the $>$99\% confidence level. We find no statistically significant distinction between the giant and MS eccentricity distributions. However a K-S test comparing the BS and MS eccentricity shows the two distributions to be distinct at the 97\% confidence level, with the BS distribution shifted to lower eccentricities. This result agrees with that found by \citet{mat09} and also discussed in Section~\ref{SelogP}, where we show that the BS binaries with $P>500$ days have significantly lower eccentricities than MS binaries of similar periods. We also investigate the dependence of eccentricity on projected radius from the cluster center. In Figure~\ref{evr} we show all binaries with orbital solutions, no longer limiting by period. As is clear from this figure, we have not derived orbital solutions for any high eccentricity ($e > 0.5$) binaries at $r \gtrsim 5$ core radii ($r \gtrsim 13$ arcminutes), while we do have such binaries in the inner-cluster region. If we divide our sample of all binaries at 5 core radii, we find the inner binaries to have a mean eccentricity of 0.31~$\pm$~0.03, and the outer binaries to have a mean eccentricity of 0.16~$\pm$~0.04. These two mean values can be distinguished at the 97\% confidence level by a Student's T-test. Likewise a K-S test shows that the inner binaries have a distribution shifted to higher eccentricities at the 97\% confidence level. This distinction may be the result of dynamical encounters that are expected to occur most frequency in the denser core of a cluster. In star clusters, fly-by interactions are the most common type of encounter \citep{bac96}, and can perturb a binary causing an increase in orbital eccentricity \citep{heg96}. Furthermore close three- and four-body interactions involving lower eccentricity binaries have a high likelihood of producing high-eccentricity products \citep{fre04,gie03}. However, we caution the reader that our completeness in orbital solutions drops off dramatically for very high eccentricity binaries (Figure~\ref{incfig}). \begin{figure}[!t] \plotone{f6.CMYK.eps} \caption{\footnotesize Period distributions for the NGC 188 binaries with $P < 3000$ days. We show the observed main-sequence, giant and blue straggler binaries in light gray, dark gray and black (green, red and blue in the online version), respectively. The top three plots show the period distributions in histogram form, while the bottom plot shows the cumulative distributions. The last bin extends to periods of 3000 days and is normalized to reflect the different bin size. We correct the main-sequence and giant binary period distributions for incompleteness and plot the corrected histograms with the solid line, above the observed distributions. In each histogram plot, we also show the \citet{rag10} Galactic field log-normal distribution (dashed line) normalized to the respective sample sizes using the field binary frequency within this period range of 14.1\%. In the bottom plot, we show the cumulative distributions of our three NGC 188 samples in the same gray-scale (color in the online version) coding as for the histograms. Again, we plot the main-sequence binaries in circles, giant binaries in diamonds and blue stragglers binaries in triangles (core) and squares (halo). Here we plot the field distribution corrected for our observational completeness in orbital solutions with the dashed line.} \label{Pfreq} \end{figure} \pagebreak \section{Period Distribution} \label{SPfreq} \begin{figure}[!t] \plotone{f7.CMYK.eps} \caption{\footnotesize The logarithm of the period as a function of projected radius from the cluster center. We plot the main-sequence binaries in light-gray circles, giant binaries in dark-gray diamonds and blue straggler binaries in black triangles (core) and squares (halo). (We use green, red and blue for the main-sequence, giant and blue straggler binaries, respectively, in the online version.) We find no trend in period with radius. (Note the errors in period, when plotted in the log, lie within the symbols used for plotting the data points.)} \label{Pvr} \end{figure} In Figure~\ref{Pfreq} we plot our observed MS, giant and BS period distributions in both histogram and cumulative distribution form for binaries with $P < 3000$ days. Additionally we plot the R10 period distribution (dashed line) normalized to the respective sample sizes using the R10 binary frequency within this period range of 14.1\%. For the MS and giant distributions we use our incompleteness study of Section~\ref{incomp} to correct our observed values. In the bottom panel of Figure~\ref{Pfreq}, we plot the cumulative distributions of the NGC 188 binaries as compared to the R10 distribution modified by our observational completeness in the same manner as in Figure~\ref{efreq}. The incompleteness-corrected NGC 188 MS log-period distribution rises towards our completeness limit, in good agreement with the log-normal distribution found by R10 for solar-type field binaries. A K-S test comparing the MS NGC 188 and R10 cumulative distributions results in a 67\% probability that they are drawn from different parent distributions. This value is as large as 67\% due to the lack of observed binaries with periods less than $\sim$4 days, as discussed in Section~\ref{SelogP}. Thus we find that the MS binary period distribution is indistinguishable from the R10 field distribution over our period range. A similar analysis returns $>$99\% confidence that the present day NGC 188 MS binary population is \textit{not} drawn from a flat distribution in log($P$), a common choice for initializing binaries in $N$-body models. Additionally the flat period distribution is not supported by observations of the young ($\sim$150 Myr) open cluster M35 \citep{mat08}. A K-S test comparing the cumulative distributions of the MS and BS binary populations results in $>$99\% confidence that these two samples are drawn from distinct parent populations \citep{mat09}. The majority of the BS binaries are found at periods near 1000 days, as is also clear from Figure~\ref{elogP}. A similar test comparing the giant and MS binaries results in an 88\% probability that they are drawn from different parent populations. This marginal distinction between the giants and MS binaries is partly due to the lack of short-period giant binaries. This loss of short-period binaries is expected as the primary evolves off the MS and expands to become a red giant. If we exclude MS binaries with $P < 11.5$ days (the minimum orbital period found amongst the giant binaries in NGC 188), the probability that the MS and giant distributions are different is reduced to $<$50\%. We find no dependence in period with projected radius from the cluster center (Figure~\ref{Pvr}). The core and halo BS binary populations are seen clearly in Figure~\ref{Pvr}. There is no distinction between the period distributions of these two subsets of the BSs, even to the extent of each having one short-period SB2 BS binary. \section{Solar-Type Main-Sequence Secondary-Mass and Mass-Ratio Distributions for Hard Binaries} \label{Sqm2freq} The secondary-mass and mass-ratio distributions for most binary populations are less well known than, for instance, the period and eccentricity distributions. Kinematic orbital solutions alone do not reveal the component masses for spectroscopic binaries. For SB2 binaries we derive the mass ratio directly; however the individual masses obtained from the kinematic orbital solutions are dependent on the inclination angle. For single-lined (SB1) binaries, we only derive the mass function in which the primary and secondary masses are entangled and also dependent on the inclination. In Paper 2 we used a photometric deconvolution technique to estimate primary masses for all of the MS binaries with orbital solutions. Here we use these primary mass estimates and the dynamical mass functions within a statistical algorithm \citep{maz92} to derive the secondary-mass and mass-ratio distributions. In Figure~\ref{qm2freq} we plot the resulting MS mass-ratio and secondary-mass distributions for the NGC 188 binaries. As in previous figures, all detected binaries with orbital solutions are shown in the filled histograms, while the incompleteness corrected distributions are plotted in the thick solid lines above the observed distributions. The horizontally-hatched regions in Figure~\ref{qm2freq} show the SB2 binaries and the vertically-hatched regions show the SB1 binaries. \begin{figure*}[!t] \begin{center} \epsscale{1.0} \plottwo{f8a.CMYK.eps}{f8b.CMYK.eps} \caption{\footnotesize Secondary-mass (left) and mass-ratio (right) distributions for the NGC 188 solar-type main-sequence hard binaries. The detected binaries with orbital solutions are plotted in the filled histograms, with SB2s in the horizontally-hatched histograms and SB1s in the vertically-hatched histograms. For the SB1 binaries we utilize the statistical inversion technique of \citet{maz92} to derive the secondary-mass and mass-ratio distributions from the observed mass functions. The incompleteness-corrected distributions are plotted as the thick solid lines above the observed distributions. The error bars are derived through a Monte Carlo analysis utilizing the \cite{maz92} algorithm to convert the Poisson errors on the mass function distribution into uncertainties on the secondary-mass and mass-ratio distributions, respectively, and we show the 95\% confidence intervals. For comparison, in the secondary-mass panel we show the \citet{kro01a} IMF, and in the mass-ratio panel we plot the distribution obtained from choosing random partners from this IMF, both with the solid curves. The curves are normalized to contain the same number of binaries as in our incompleteness-corrected NGC 188 sample. Finally we plot the secondary-mass distribution in the more familiar log-log format in the upper-right inset. We show this figure for illustrative purposes, but focus our statistical analysis on the observed mass-function distribution shown in Figure~\ref{fmfig}. } \label{qm2freq} \end{center} \epsscale{1.0} \end{figure*} \begin{figure}[!t] \begin{center} \plotone{f9.eps} \caption{\footnotesize Cumulative mass-function distribution for the NGC 188 solar-type main-sequence binaries. We show the observed mass functions with the black points, and circle the double-lined binaries (for reference). With the solid line we plot the mass function distribution derived by drawing companions from the \citet{kro01a} IMF. To derive this curve we assume the inclination angles are distributed isotropically and draw primary masses from a Gaussian fit to the observed NGC 188 main-sequence primary-mass distribution. This theoretical distribution is then corrected for our incompleteness in orbital solutions (out to $P=3000$ days) in the same manner as Figures~\ref{efreq} and \ref{Pfreq}. A K-S test shows that the observed mass-function distribution is statistically indistinguishable from the IMF. } \label{fmfig} \end{center} \end{figure} For the SB1 binaries we use the iterative algorithm of \citet{maz92} to convert our distributions of mass functions into distributions of mass ratio and secondary mass, using our primary mass estimates and assuming the inclination angles to be randomly distributed. Note that this technique (as well as other statistical inversion methods) does not reveal the secondary mass or mass ratio for a given binary, but instead derives the respective distributions for a sample of binaries. We correct the SB1 binaries for our incompleteness using the analysis of Section~\ref{incomp} after applying the statistical algorithm to derive the observed distributions. We do not use this same incompleteness correction for the SB2 binaries as these systems (once detected) are given the highest priority in our observations. As such, we have only one confirmed SB2 binary without an orbital solution (ID 9119), and we add in this binary by hand. The turnoff in NGC 188 is at $\sim$1.1 M$_{\odot}$, and our incompleteness correction covers binaries with $P < 3000$ days that have stellar companions down to 0.08 M$_{\odot}$~(excluding brown dwarfs and other sub-stellar objects). In the secondary-mass plot, we also show the ``universal IMF'' from \citet{kro01a} with the solid curve. In the mass-ratio plot, we show the distribution obtained from choosing random partners from the IMF in the solid curve, limiting the primary to be within our observable mass range and with the mass ratio always less than one. The curves are normalized to contain the same number of binaries as in our incompleteness corrected NGC 188 sample. The \citet{kro01a} IMF was not intentionally derived to model secondary stars in binaries. Still the shape of the MS secondary-mass distribution is consistent with the IMF in that it rises towards lower masses. Likewise the mass-ratio distribution shows a similar form to the curve derived by choosing random partners from this IMF. We also observe an excess of binaries with masses between 0.85 and 1.0 M$_{\odot}$~(and mass ratios $>0.9$) as compared to the IMF. This excess of near equal-mass binaries remains if we instead use the mass functions for all binaries (including the SB2s) and derive the secondary-mass and mass-ratio distributions for the full binary population with the \citet{maz92} algorithm; therefore this result is not sensitive to the method used here. This may be analogous to the abundance of ``twins'' observed in some studies, generally at shorter periods \citep[e.g.][R10]{tok00,fis05}. We have investigated the mass-ratio distribution as a function of period, but find that the decreased sample sizes make these results highly uncertain. To check formally for consistency between the observed companion masses and the IMF we turn to the observed mass functions. In Figure~\ref{fmfig}, we compare the observed cumulative mass-function distribution (black points, with SB2s circled) to a theoretical distribution (solid line) derived by drawing companions from the IMF, primary masses from a Gaussian fit to our observed NGC 188 MS primary mass distribution, and assuming an isotropic inclination distribution. We then correct this theoretical distribution for our incompleteness in orbital solutions using the same method as in Figures~\ref{efreq} and \ref{Pfreq}. A K-S test shows that we cannot formally distinguish the observed mass-function distribution from the IMF. (We also note that the mass-function distribution cannot formally be distinguished from a distribution derived by choosing binaries from a uniform mass-ratio distribution.) \begin{figure*}[!t] \begin{center} \epsscale{1.0} \plottwo{f10a.CMYK.eps}{f10b.CMYK.eps} \caption{\footnotesize Secondary-mass (left) and mass-ratio (right) distributions for the NGC 188 blue straggler binaries. The SB2s are plotted with the horizontally-hatched histograms and the SB1s are plotted with the vertically-hatched histograms. For SB1s, the primary masses are estimated in relation to evolutionary tracks of \citet{mar08}, and we then use the \citet{maz92} algorithm to derive the secondary-mass and mass-ratio distributions from the observed mass functions. The masses for the two SB2 blue stragglers are discussed in detail in Section~\ref{Sqm2freqBS}, and see also \citet{mat09}. The uncertainties shown here are derived in the same manner as in Figure~\ref{qm2freq}. } \label{qm2freqBS} \end{center} \epsscale{1.0} \end{figure*} There is some disagreement in the literature about the secondary-mass and mass-ratio distributions for field dwarfs. R10 find their sample of field dwarfs to have a uniform mass-ratio distribution between $0.2 < q < 0.95$, with evidence for an abundance of twins at $q > 0.95$. \citet{maz03} find a mass-ratio distribution that is formally consistent with a uniform distribution. They also note a possible rise in the distribution towards lower mass-ratio values, but with large uncertainties. \citet{duq91} and \citet{gol03} find distributions that rise towards lower-mass secondaries. (Strictly \cite{gol03} find this result for disk binaries, and not for the halo binaries.) We conclude that the companions to the NGC 188 MS binaries are consistent with those of the field dwarf binaries, but given the uncertainty in the literature about the field population this conclusion lacks precision. \section{Blue Straggler Secondary-Mass and Mass-Ratio Distributions} \label{Sqm2freqBS} The masses and evolutionary states of the companion stars to the BSs in binaries are a powerful tool to investigate the origins of these systems. Knowledge of whether the secondaries are normal MS stars or white dwarfs (WDs) can be particularly useful in distinguishing between possible formation mechanisms \citep{gel11}. For instance, the Case B or Case C mass transfer mechanisms predict a BS binary with a WD companion \citep{mcc64,pac71}, while a merger of an inner binary of a triple system \citep{per09} or a stellar collision that leaves a bound companion \citep{hur05} would more likely produce MS secondaries. In order to derive the observed BS secondary-mass and mass-ratio distributions, we must first estimate the BS primary masses. To do so we compare the observed magnitudes and colors of the BSs to a coarse grid of evolutionary tracks \citep{mar08} extending from the ZAMS to 7 Gyr, and for stars with masses ranging from the cluster turnoff mass of 1.1~M$_{\odot}$~to twice this mass (in steps of 0.1~M$_{\odot}$). For the SB1 BSs, the secondaries are at least 2.5 magnitudes (in $V$) fainter than the BS primaries (Paper 1). For a given BS, deconvolving a secondary star at this limiting magnitude from the combined light would result in a primary of $\lesssim 0.1$ $V$ magnitudes fainter than the combined $V$ magnitude. This minor difference in $V$ magnitude is below the mass resolution of the evolutionary tracks (e.g., see Figure~\ref{CMD5078}). Therefore we do not attempt to deconvolve the secondary light from these BSs, and estimate their masses based on the proximity of the combined magnitude and color to the evolutionary tracks. We find the NGC 188 SB1 BSs to have masses between 1.2 - 1.6 M$_{\odot}$. We then use the \citet{maz92} algorithm in the same manner as for the MS binaries to derive the BS binary secondary-mass and mass-ratio distributions shown in Figure~\ref{qm2freqBS} (with SB1s in the vertically cross-hatched histograms and SB2s in the horizontally cross-hatched histograms). The BS SB1 secondary-mass distribution is narrow and peaks near a mass of $\sim$0.5~M$_{\odot}$. This result, when considered in light of the periods near 1000 days and low eccentricities, suggests an origin through Case C mass transfer, which predicts carbon-oxygen WD companions of masses between $\sim$0.5 - 0.6 M$_{\odot}$~with orbital periods of order 1000 days \citep{mcc64,pac71,hur02,che08}. However the secondary-mass distribution alone does not rule out MS companions, as we would not detect the flux from a 0.5~M$_{\odot}$~MS (or WD) star in our BS spectra due to the large difference in luminosities. We discuss this result in detail in \cite{gel11} and build upon these results in Section~\ref{Hcomp}. For the two short-period SB2 BS binaries, the secondary adds significant light to the system, and thus for each SB2 BS we deconvolve the secondary from the combined light before estimating the primary mass. Additionally we utilize a grid of synthetic spectra to estimate the effective temperatures of both components of these systems, which helps to further constrain the masses. We discuss these results briefly in \citet{mat09} and expand upon them here. \begin{figure}[!t] \begin{center} \plotone{f11.eps} \caption{\footnotesize Color-magnitude diagram showing the loci of possible locations for the components of the blue straggler binary ID 5078. Cluster members are shown in the small black points, and 5078 is circled. We also show a ZAMS and a 7 Gyr isochrone in the thin solid lines, and evolutionary tracks for 1.2~M$_{\odot}$, 1.3~M$_{\odot}$, 1.4~M$_{\odot}$~and 1.5~M$_{\odot}$~stars in the dotted lines, respectively \citep{mar08}. The thick gray line shows the locus of potential secondary stars, and the thick black line shows the locus of potential primary stars. The derived effective temperature of 5850 K for the secondary implies a mass of 1.02~M$_{\odot}$; we plot the location of this secondary star, and the location of the associated primary, with the light-gray crosses. However, given the kinematic mass ratio of $q=0.678\pm0.009$ and this secondary mass, the blue straggler mass would be 1.5~M$_{\odot}$. Yet the blue straggler is significantly less luminous than a normal 1.5~M$_{\odot}$~star at any point in its evolution. } \label{CMD5078} \end{center} \end{figure} BS ID 7782 has a kinematic mass ratio of $q=1.005\pm0.013$ (Paper 2). A spectral analysis yields a luminosity ratio\footnote{Due to an update in our TODCOR analysis we find a slightly different luminosity ratio from \citet{mat09}, who gave a value of $0.75 \pm 0.01$.} of $0.871 \pm 0.024$, and effective temperatures for the primary and secondary stars of 6500 K and 6325 K, respectively. Both stars are hotter than the MS turnoff effective temperature of 5900 K and more luminous than turnoff stars. Evidently 7782 comprises \textit{two} BSs. The location of the system in the cluster halo suggests that a close dynamical encounter was involved in the formation of the binary, perhaps the final of several encounters near the cluster core that exchanged the BSs into the system or formed one of them collisionally and ejected the binary into the cluster halo. The differing temperatures and luminosities are unexpected given the nearly identical masses. Possibly the two BSs did not form at the same time and thus have different evolutionary ages. We are able to model the system with two 1.25 M$_{\odot}$~stars having an age difference of $\sim$2 Gyr \citep{mar08}, although this solution is not unique given the present data. More detailed spectral analysis of the two stars is merited. Nonetheless, given the large bin sizes used in constructing the secondary-mass distribution (Figure~\ref{qm2freqBS}), we can confidently place the secondary of 7782 in the 1.2 - 1.4 M$_{\odot}$~bin. BS ID 5078 has a kinematic mass ratio of $q=0.678\pm0.009$ (Paper 2). From our analysis of the spectra, we derive an effective temperature for the primary of 6500 K and for the secondary of 5850 K. The effective temperature of the secondary is consistent with a star on the upper MS. Assuming the secondary is a MS star, we deconvolve the combined light using secondary light contributions derived from a 7 Gyr \citet{mar08} isochrone. We limit the possible secondaries such that the effective temperature is within 250 K ($\sim 2\sigma$) of 5850 K, and we limit the possible BS primaries such that the BS is on or redder than the ZAMS. The resulting loci of possible CMD locations for the primary and secondary are shown in the thick black and gray lines in Figure~\ref{CMD5078}, respectively. We find a range in possible masses for the secondary of 5078 of 0.94~M$_{\odot}$~to~1.08~M$_{\odot}$. The nominal effective temperature for the secondary of 5850 K corresponds to a 1.02~M$_{\odot}$~MS star, and we use this secondary mass in Figure~\ref{qm2freqBS}. The location of the associated BS primary lies on or near a 1.2~M$_{\odot}$~evolutionary track. We note though that the effective temperature of 6500 K derived for the BS primary suggests a $(\bv)$ color that is $\sim$0.1 mag bluer than inferred from this analysis. The range in secondary masses and the kinematic mass ratio yields a range for the BS primary star mass between 1.39~M$_{\odot}$~and~1.59~M$_{\odot}$. However, the locus of potential CMD locations for the primary clearly falls below such evolutionary tracks, and even the combined light of 5078 is too faint. The standard stellar evolution tracks underestimate the dynamically informed mass of the BS primary star by 15\% to 30\%. This potential systematic error does not meaningfully effect the results shown in Figure~\ref{qm2freqBS}. If we simply assign the same mass to each BS within the observed range of 1.2 - 1.6 M$_{\odot}$, the resulting secondary-mass and mass-ratio distributions stay within the uncertainties on the distributions shown in Figure~\ref{qm2freqBS}. Assigning all BSs a mass of 1.3 M$_{\odot}$, which is approximately the average BS mass that we derive, results in distributions that are nearly identical to those shown here. If we take the extreme case of setting all BSs to twice the turnoff mass (2.2 M$_{\odot}$), the general forms of the distributions remain the same, but the peak in the secondary-mass distribution shifts to the 0.6 - 0.8~M$_{\odot}$~bin. Again we do not attempt to correct these BS distributions for our incompleteness and show only the BS binaries with orbital solutions. We remind the reader that there are only six additional NGC 188 BSs, only one of which is detected as a binary but does not yet have an orbital solution. \section{Comparison to a Sophisticated $N$-body Open Cluster Simulation} \label{Hcomp} \subsection{The Simulation} The open cluster $N$-body simulation of M67 by \citet{hur01,hur05} was able to track the evolution and dynamics of realistic initial numbers of single stars \textit{and} binaries, and included detailed prescriptions for mass loss, tidal processes, mass transfer, mergers and collisions. Through their model, Hurley et al.~were able to match the observed M67 color-magnitude diagram remarkably well. They produced similar populations of anomalous stars as are observed in M67, including BSs. Importantly, they were able to track the dynamical evolution of these anomalous stars and pinpoint their origins with great detail, many of which involved interactions with binary stars. Simulations such as these have the power to revolutionize our understanding of the evolution of a binary population and the origins of anomalous stars, like BSs, in star clusters. Though Hurley et al.~designed their simulations to model M67, their results can also be compared to similar open clusters, such as NGC 188. Both NGC 188 and M67 are old open clusters at 4 and 7 Gyr respectively. The current dynamical relaxation time for NGC 188 is 180 Myr \citep{chu10} and for M67 is 100 Myr \citep{mat86}. As shown in Section~\ref{freq}, NGC 188 and M67 have comparable binary frequencies, and their binary orbital period and eccentricity distributions are of a similar form \citep{lat07}. Furthermore, the BS populations of these two clusters show strong similarities. \citet{lat96} observed $\sim$60\% of the M67 BS stars to be members of hard-binary systems, consistent with the 76~$\pm$~19~\% binary frequency amongst the NGC 188 BSs observed here. Also, the M67 BS binaries show a similar tendency towards periods of order 1000 days \citep{lat96}. Thus, in this section we compare the NGC 188 hard-binary population to that of the \citet{hur05} simulation. The simulation started with a 50\% binary frequency over all periods and masses, in which binary orbital elements were chosen from a flat period distribution and a thermal eccentricity distribution. The primary stars were chosen to follow a \citet{kro91} IMF, with masses $> 0.1$ M$_{\odot}$, and the mass ratios for binaries were chosen from a uniform distribution in $q$. The binaries were then modified according to the \citet{kro95b} tidal evolution prescription, and the resulting initial stellar population was evolved for 4 Gyr. We compare the MS and giant populations of the simulation at this final age to NGC 188. We only analyze MS and giant stars that would be within our observable period, spatial and magnitude domain. Specifically we consider only those binaries with $P < 10^4$ days to be detectable, and assume that we would only obtain orbital solutions for binaries with $P < 3000$ days. Our analysis only includes the inner 17 pc (radius in projection along the $z$ axis) of the cluster. We use the NGC 188 distance modulus of $(M-m)_V$=11.44 and the reddening $E(\bv)$=0.09 \citep{sar99} to define the lower-mass (faint) limit of the MS. The giants are all within our observable magnitude domain; thus we only limit the giants by radius from the cluster center. We choose not to impose magnitude or spatial limits on the BSs created within the simulation. There are only four BSs created within the simulation throughout the entire cluster lifetime that at some moment reside outside of our observable spatial region. Each of these were formed near the core, and later ejected from the cluster, spending only one time step ($\sim$60 Myr) outside of 17 pc (in projection) from the cluster center before traveling beyond the cluster's tidal radius. In the following analysis, we use an "integrated sample" of all BSs that existed between 2 and 4 Gyr. The ``integrated sample'' comprises the union of every BS at each time step in the simulation during this range in time. There are 57 individual BSs present at some point during this interval, while the integrated sample contains multiple realizations of these BSs over these 33 time steps. In so doing, the integrated sample effectively weights results by the lifetime of each individual BS in a certain binary configuration. For example, there were 26 BSs created in binaries between 2 - 4 Gyr, 11 of which underwent an encounter after formation that significantly changed the orbital parameters. Such changes in orbital parameters are accounted for, in the integrated sample, as separate systems with lifetimes that are equal to the duration at each orbital configuration. To close, Hurley et al.~did not have access to RV data as comprehensive as ours for either the M67 or NGC 188 binary population when constructing their simulation or analyzing their results. Therefore, we are able to investigate the accuracy of their model both in more detail and, perhaps more importantly, in observational dimensions not available to them. Many of the comparisons discussed next truly test \textit{a priori} predictions of their theory. \begin{figure}[!t] \plotone{f12.CMYK.eps} \caption{\footnotesize Period distribution for binaries in the \citet{hur05} simulation. We compare the main-sequence (top), giant (middle) and blue straggler (bottom) populations from the simulation, using the 4 Gyr main-sequence and giant binaries and the integrated blue straggler sample. Here we include all binaries with $P < 3000$ days, and the last bin is normalized to account for the different bin size. In the main-sequence plot, we also show the initial period distribution of the simulation within the same mass, period and spatial domain with the gray line. All simulated distributions are normalized by the respective NGC 188 sample sizes, and we show the true frequencies on the right y-axes. The respective NGC 188 distributions are plotted in the thick-lined histograms. For comparison we also show the Galactic field period distribution of \citet{rag10} with the dashed lines, normalized to the NGC 188 sample size using the field binary frequency within this period range. The simulation contains a large overabundance of short-period main-sequence binaries as compared to observations of both NGC 188 and the field. \label{HPfreq} } \end{figure} \subsection{Comparison of the Main-Sequence Hard-Binary Populations in NGC 188 and in the Simulation} \label{ShMS} \subsubsection{Frequency} We find a MS solar-type hard-binary frequency at 4 Gyr in the simulation of 64~$\pm$~4~\% (230/361) for $P<10^4$ days. (Here the uncertainty is the expected counting error given multiple realizations of the simulation.) We specify the binaries defining this frequency in the same manner as for the MS NGC 188 binaries in Section~\ref{freq}. The MS hard-binary frequency in the simulation has grown from its initial value of 47.9~$\pm$~2.5~\% (for the same mass, spatial and period domain). This increase in binary frequency is a result of the preferential evaporation of the solar-type single stars over their higher total-mass solar-type binary counterparts. The 4 Gyr MS solar-type binary frequency is twice that of the NGC 188 MS hard-binary frequency of 29~$\pm$~3~\%, and these two frequencies can be distinguished at the $>$99\% confidence level. The simulated giant hard-binary frequency of 52~$\pm$~16~\% (11/21) is also higher than for the incompleteness-corrected NGC 188 giant hard-binary frequency of 31~$\pm$~7~\%, although given the large uncertainties these two values are not distinct at high confidence. These differences in binary frequency are directly linked to the high initial hard-binary frequency in the simulation, and the small subsequent change in frequency during the cluster lifetime. \begin{figure}[!t] \plotone{f13.CMYK.eps} \caption{\footnotesize $e - \log(P)$ diagram for binaries in the \citet{hur05} simulation. We compare the main-sequence (top), giant (middle) and blue straggler (bottom) populations from the simulation. We show the 4 Gyr main-sequence and giant binaries. For the blue stragglers, we show the integrated (2 - 4 Gyr) sample, where each point represents a given blue straggler binary configuration during one time step. If a dynamical encounter alters the orbital period or eccentricity of a given blue straggler in a binary, the result is shown by individual (often partially overlapping) points at each orbital configuration, respectively. The circularization period $P_{circ} =15.0$ days \citep{mat04} is shown in the main-sequence $e - \log(P)$ plot by the dotted line. Note the unphysical ``envelope'' to the main-sequence plot that is not seen in the NGC 188 observations (or those of other open clusters of a wide range in age, \citealt{mei05}). \label{HelogP} } \end{figure} \begin{figure}[!t] \plotone{f14.CMYK.eps} \caption{\footnotesize Eccentricity distribution for binaries in the \citet{hur05} simulation. We compare the main-sequence (top), giant (middle) and blue straggler (bottom) populations from the simulation, using the 4 Gyr main-sequence and giant binaries and the integrated blue straggler sample. Here we show only binaries with $P_{circ} < P < 3000$ days, where the circularization period $P_{circ} = 15.0$ days \citep{mat04}. In the main-sequence plot, we also show the initial eccentricity distribution of the simulation within the same mass, period and spatial domain with the gray line. All simulated distributions are normalized by the respective NGC 188 sample sizes, and we show the true frequencies on the right y-axes. The respective NGC 188 distributions are plotted in the thick-lined histograms. Finally, we show the \citet{rag10} distribution in the dashed lines, normalized to the NGC 188 sample size using the field binary frequency within this period range. \label{Hefreq} } \end{figure} \subsubsection{Orbital Period} In Figure~\ref{HPfreq} we plot the period distribution for the simulation. We use the same gray-scale (color in the online version) coding for the simulated MS, giant and BS populations as we have done throughout the paper for NGC 188. We normalize the simulated output by the respective NGC 188 sample sizes and show the true frequencies on the right y-axes. For reference, we also include the NGC 188 distributions as the thick histograms and the normalized R10 field distributions as the dashed lines. As in Figure~\ref{Pfreq} we show binaries out to $P = 3000$~days in Figure~\ref{HPfreq}, and normalize the last bin to account for the different bin size. \begin{figure*}[!ht] \begin{center} \plottwo{f15a.CMYK.eps}{f15b.CMYK.eps} \caption{\footnotesize Log period (top) and eccentricity (bottom) plotted against projected radius from the cluster center for the solar-type initial (left) and final (at 4 Gyr; right) binary populations. We show the main-sequence, giant and blue straggler binaries in light-gray circles, dark-gray diamonds and black triangles (green, red and blue in the online version), respectively. (In the initial plots all binaries are on the main-sequence.) On the top x-axes we convert parsecs to core radii using a core radius of 0.64 pc \citep{hur05}. Despite becoming more centrally concentrated, the overall form of the period distribution does not change over 4 Gyr of evolution. On the other hand, we notice a population of high-eccentricity binaries in the core of the cluster at 4 Gyr that were not present in the initial population. Their higher eccentricities are the result of dynamical encounters.} \label{Hpevr} \epsscale{1.0} \end{center} \end{figure*} The MS period distribution of the simulation is significantly different from that observed in NGC 188 (Figure~\ref{Pfreq}). In particular, the simulation contains a large overabundance of short-period MS binaries as compared to both NGC 188 and the Galactic field. If we normalize the MS period distribution from the simulation to contain the same number of binaries as in the NGC 188 MS sample, a $\chi^2$ test shows that the two distributions are distinct at the $>$99\% confidence level. The distribution of binary periods has not changed significantly over the 4 Gyr of simulated evolution. A K-S test relating the initial and final hard-binary period distributions (Figure~\ref{HPfreq}) shows no distinction. Thus this discrepancy between the simulated and observed populations is a result of the flat initial distribution in log($P$) used in the simulation. \subsubsection{Orbital Eccentricity} In Figure~\ref{HelogP} we show the $e - \log(P)$ diagram for the \citet{hur05} simulation. The $e - \log(P)$ distribution of the simulation MS binaries is strikingly different from the observed $e - \log(P)$ distribution of MS binaries in NGC 188 (top panel, Figure~\ref{elogP}), and indeed from observations of many similarly evolved open clusters and the Galactic field \citep[see][R10]{mei05}. The differences in the simulation include: \begin{enumerate} \item The upper envelope on orbital eccentricity, rising with increasing period. This derives directly from the thermal initial conditions and the Kroupa tidal evolution prescription. The observed distribution of orbital eccentricities longer than the circularization period is better matched with a wide Gaussian in eccentricity at all periods, as found in the young cluster M35 \citep{mat08} and in the field (R10), and as used in the modeling of detached binary evolution by \citet{mei05}. \item The large number of circular binaries with periods greater than the observed circularization period of 15.0 days. Many of these systems had circular orbits initially due to the Kroupa pre-evolution process. An additional small population of long-period circular binaries was created through unstable mass transfer (common-envelope) events involving initially non-circular MS binaries. \item The large number of eccentric binaries with periods shorter than the circularization period. This suggests that the tidal energy dissipation rate is underestimated in the \citet{hur05} model. Indeed \citet{bel08} suggest that the convective tidal damping should be increased by a factor of 50 to match observations of M67. \end{enumerate} Similarly, in Figure~\ref{Hefreq} we show the eccentricity distributions from the simulation. Here we use the same period limits of $P_{circ} < P < 3000$ days as in Section~\ref{Sefreq} to define our sample, where $P_{circ}$ = 15.0 days (for NGC 188 found by \citealt{mat04}). The marked excess in the number of circular orbits compared to NGC 188 (Figure~\ref{efreq}) and the field (R10) is evident. This same excess is seen in comparison with the young open cluster M35 \citep{mat08}. Most of these circular orbits are in short-period binaries, and as such play a critical role in the formation of BSs in the simulation. \begin{figure*}[!t] \begin{center} \plottwo{f16a.CMYK.eps}{f16b.CMYK.eps} \caption{\footnotesize Secondary-mass (left) and mass-ratio (right) distributions for solar-type main-sequence binaries in the \citet{hur05} simulation (filled histograms) and in NGC 188 (thick-lined black histograms), both for binaries with $P < 3000$ days. We also show the initial populations from the model within the same mass and period range in the gray lines. We normalize the simulated distributions by the NGC 188 sample size, and show the true frequencies on the right y-axes. As the simulation only included stars with masses $>$0.1 M$_{\odot}$~in the initial population, we show the secondary-mass distribution from 0.10 - 1.15 M$_{\odot}$~and the mass-ratio distribution from 0.1 - 1.} \label{Hqm2freq} \epsscale{1.0} \end{center} \end{figure*} Finally, in Figure~\ref{Hpevr} we show the initial and final distributions of eccentricity as functions of radius. Over the course of the cluster evolution, a population of higher-eccentricity binaries developed within $\sim$5 core radii from the cluster center. (This population can also be seen in the $e - \log(P)$ diagram of Figure~\ref{HelogP}.) No similarly high eccentricity binaries are found in the cluster halo. The increases in the orbital eccentricities of these binaries can be traced to dynamical encounters. These high-eccentricity binaries in the model are likely analogues of the high-eccentricity binaries observed in the core of NGC 188 (discussed in Section~\ref{Sefreq}), suggestive of a dynamical origin for some or all of the NGC 188 high-eccentricity binaries. \subsubsection{Secondary-Mass and Mass-Ratio Distributions} Unlike our observed NGC 188 sample where we require a statistical method to derive the secondary-mass and mass-ratio distributions, within the simulation we know the secondary masses and mass ratios for every binary. In Figure~\ref{Hqm2freq} we show the 4 Gyr secondary-mass (left) and mass-ratio (right) distributions for the simulated MS binaries (filled histograms), and compare to the respective NGC 188 MS distributions (thick-lined black histograms) as well as the initial distributions in the simulation (thin-lined gray histograms). The excess in hard-binary frequency - both initial and at 4 Gyr - in the simulation compared to NGC 188 is again clearly evident. However here we focus on the shapes of the distributions. Both the secondary-mass and mass-ratio distributions of the simulation are essentially flat at 4 Gyr, except for the highest mass secondaries which are few in the simulation. This flat shape is in marked contrast to that found in NGC 188, which rises to lower masses and mass ratios. The NGC 188 distribution also shows some evidence of a peak at comparable primary and secondary masses ("twins"), which is not seen in the simulation. In fact, the initial secondary-mass and mass-ratio distributions of the simulation do rise to lower secondary mass and look much like the current NGC 188 distributions. However, by 4 Gyr they flatten with subsequent dynamical evolution of the cluster. This would suggest that the NGC 188 initial distribution may itself have been even more peaked to lower masses than currently observed. However the simulation is silent with respect to the creation or evolution of a population of twins. The 4 Gyr mass-ratio distribution from the simulation is quite similar to that of the field binaries of R10, in that the distribution is consistent with being uniform over mass ratios of $\sim$0.2 - 0.95. However the simulation does not reproduce the abundance of twins ($q \sim 1$), again suggesting that this feature seen in both the field and open clusters is likely linked to the processes of binary formation rather than dynamics. \subsubsection{Summary} The MS hard-binary population predicted by the simulation is significantly different from the MS hard-binary population of NGC 188. Along some dimensions this reflects on the adopted initial conditions of the simulation and along others it reflects on the physics of the simulation itself. The initial binary frequency and period distribution of the simulation are largely maintained throughout the cluster lifetime, and so the marked differences with the NGC 188 frequency and period distribution are the result of inaccurate initial conditions. Specifically, the flat initial period distribution is distinct from the log normal period distribution of the observed young ($\sim$150 Myr) open cluster M35 \citep{mat08}, and the initial hard-binary frequency integrated over period is a factor two higher than found in such clusters. Given that the initial binary period distribution imprints itself on the cluster throughout its evolution, it is imperative that the properties of the initial binaries used in future $N$-body simulations be guided by observations of such young clusters. The differences in the simulated and observed orbital eccentricity and $e - \log(P)$ distributions are also in large part the result of inaccurate initial conditions. Again, the eccentricity distributions in observed open clusters (e.g., M35, \citealt{mat08}; and see \citealt{mei05}) argue that a Gaussian-like distribution in eccentricity, with a weak dependence on period above the initial circularization cutoff, is a better choice than a thermal distribution. In contrast, the large frequency of eccentric binaries with periods shorter than the circularization period of M67 and NGC 188 indicates that the physics of tidal circularization incorporated in the simulation is not accurate. Arguably the tidal energy dissipation rate is too small, as suggested by \citet{bel08}. This can be empirically tested with future simulations. On the other hand, the Kroupa (1995) tidal evolution prescription applied to the initial binary population evidently produced too many initial circular orbits at longer periods. The presence in the simulation of a central concentration of binaries and especially binaries in the core with dynamically produced high orbital eccentricities is in agreement with the binary population of NGC 188 and is encouraging with respect to the dynamical encounter physics in the simulation. We have not attempted a specific comparison of rates and frequencies, in part because \citet{hur05} derive a core radius for the simulation of 0.64 pc at 4 Gyr, smaller than observed in NGC 188 (1.3 pc, \citealt{bon05}; 2.1 pc, \citealt{chu10}). This higher central concentration will change evolution timescales in ways that likely are not easily compensated for accurately. The correct initial secondary-mass and mass-ratio distributions remain poorly known, and comparison with NGC 188 provides only a rather distant lever on the question. Based on the simulation and NGC 188 distributions, it would appear that a secondary distribution rising to smaller masses is appropriate, and perhaps more so than used by Hurley et al. (2005). More intriguing is whether twins are present initially, given the hint of an excess of them existing both in NGC 188 and the field. The CMD was the primary interface with observation used by Hurley et al. (2005), and they were very successful in matching the observed CMD of M67. In particular they reproduced well the BS population of M67. Importantly, the BS population of the simulation derived largely from mergers of close binaries, which we find to be over-represented in the simulation due to the adopted initial binary population. A plausible hypothesis is that reducing the binary frequency by a factor of two and reducing the frequency of the closest binaries relative to those of longer period would substantially reduce the BS production rate and under-predict the BS population compared to M67. Indeed, \citet{hur05} note that their $N$-body models using initial period distributions such as \citet{duq91} and \citet{kro95b} produce significantly fewer BSs than their M67 model (which used a flat initial period distribution). Thus such models may require an offsetting change in the physics of BS formation or the relative frequency of BS formation channels in order to recover agreement with M67. We begin discussion of this issue in the next section. \subsection{Comparison of the Blue Straggler Populations in NGC 188 and in the Simulation} \label{ShBS} \subsubsection{Comparison of the Blue Stragglers in Binaries} Using the integrated sample of BSs, we derive a BS hard-binary ($P<10^4$ days) frequency of 29~$\pm$~2~\% (168/584). This hard-binary frequency for the BSs in the simulation is significantly lower than the NGC 188 BS hard-binary frequency of 76~$\pm$~19~\%. In Figure~\ref{M67hBSfreq}, we plot the \citet{hur05} BS hard-binary frequency as a function of cluster age (no longer limited to ages $>$2 Gyr). Over most of the lifetime of the simulated cluster the BS hard-binary frequency is less than half that of NGC 188 (and M67). Thus, although the simulation has an overabundance of hard MS binaries, they are decidedly lacking amongst the simulated BS population. \begin{figure}[!t] \begin{center} \plotone{f17.eps} \caption{\footnotesize Blue straggler hard-binary ($P < 10^4$ days) frequency as a function of time in the \citet{hur05} $N$-body simulation. To reduce the uncertainties and scatter we average every four adjacent time steps, so that each point represents $\sim$240 Myr in time. We show the region in time covered by each point in the gray horizontal bars. For comparison, we also show the NGC 188 BS binary frequency of 76~$\pm$~19~\% with the light gray band.} \label{M67hBSfreq} \end{center} \end{figure} We estimate from the CMD location that the most massive NGC 188 BS in a binary with an orbital solution has a mass of $\sim$1.5 times the MS turnoff mass of 1.1 M$_{\odot}$. The MS turnoff mass in the simulation is 1.7 M$_{\odot}$~at 2 Gyr and decreases to 1.3 M$_{\odot}$~by 4 Gyr. Roughly 44\% of the BSs in binaries ($P < 3000$ days) in the simulation (between 2 - 4 Gyr) have masses $>$1.5 times the respective simulation turnoff mass. Thus, on average the BSs in the simulation are more massive than those in NGC 188. In Figure~\ref{Hqm2freqBS} we compare the NGC 188 BS secondary-mass and mass-ratio distributions to those in the \citet{hur05} simulation. The BSs in the simulation do not show the strong peak at $\sim 0.5$~M$_{\odot}$~that is observed in the NGC 188 BSs. Instead the distribution is roughly uniform for secondary masses $>$0.2~M$_{\odot}$. Interestingly the observed and simulated mass-ratio distributions for the BSs appear to be quite similar in form. This agreement between the mass-ratio distributions is the result of dividing high-mass companions by the high-mass BSs, both of which individually are inconsistent with the NGC 188 BSs. The binaries in the integrated BS sample of the simulation also have significantly different periods and eccentricities than the BS binaries of NGC 188. Specifically, the simulation produces BS binaries with shorter periods and a higher frequency of large eccentricities. A 2D K-S test comparing the observed and simulated BS $e - \log(P)$ distributions shows that the two distributions are drawn from different parent populations at the $>$99\% confidence level. Likewise 1D K-S tests comparing the observed and simulated eccentricity and period distributions return distinctions at the 99\% and $>$99\% confidence levels, respectively. \begin{figure*}[!t] \begin{center} \plottwo{f18a.CMYK.eps}{f18b.CMYK.eps} \caption{\footnotesize Secondary-mass (left) and mass-ratio (right) distributions comparing the blue straggler binaries in the \citet{hur05} simulation to those in NGC 188. The thick-lined histograms reproduce the results shown in Figure~\ref{qm2freqBS} for the NGC 188 blue stragglers. For the simulation, we use the integrated sample of all simulated BSs from 2 - 4 Gyr. The filled histograms show the simulated distributions normalize by the NGC 188 sample size. The true frequencies are shown on the right y-axes. For comparison of distribution shapes we also re-normalize the distributions from the simulation to contain the same number of blue straggler binaries as in NGC 188, shown with the dashed lines.} \label{Hqm2freqBS} \epsscale{1.0} \end{center} \end{figure*} Thus the BS sample from the simulation has significantly different binary properties from those of the NGC 188 BS population. These discrepancies are almost certainly linked, at least in part, to the inaccurate initial binary population used in the simulation. \subsubsection{Comparison of the Blue Straggler Spatial Distributions} \label{SBSspat} The spatial distribution of the BSs in the simulation is also quite different from what we observe in NGC 188 (Figure~\ref{BSspatdist} and see also Figure 9 from Paper 1). In Paper 1 we show that the NGC 188 BSs have a seemingly bimodal spatial distribution, with a population of 14 centrally concentrated BSs in the core and 7 BSs in the halo. The full NGC 188 BS population is not significantly more centrally concentrated than the solar-type single stars in the cluster. In the simulation, we do not observe this relatively large population of BSs towards the cluster halo. Instead the full BS population is centrally concentrated with respect to the single MS stars (at the 98\% confidence level; also see Figure~\ref{Hpevr}). Furthermore, if we only analyze the 4 Gyr BS population, all but one BS is within 4.5 pc from the cluster center; the remaining BS is at a radius of 19.38 pc which is beyond the spatial extent of our NGC 188 survey. Other than the difference in number of BSs, the core and halo BS populations in NGC 188 are statistically indistinguishable. Both have consistent binary frequencies and indistinguishable distributions of orbital eccentricities, periods and mass functions. Furthermore both the core and halo populations contain one of the two short-period SB2 BSs, respectively. The distributions of $v \sin i$ rotational velocities \citep{mat09} are also consistent between the core and halo populations. \begin{figure}[!t] \begin{center} \plotone{f19.eps} \caption{\footnotesize Cumulative spatial distribution of stellar populations in the \citet{hur05} $N$-body model. The blue stragglers from the integrated sample are shown in the solid line. This BS spatial distribution is compared to the single main-sequence and giant stars, shown in the dashed and dotted lines, respectively. The blue stragglers are more centrally concentrated than the single main-sequence stars (at the 98\% confidence level). } \label{BSspatdist} \end{center} \end{figure} Similar bimodal BS spatial distributions have also been observed in globular clusters \citep[e.g.,][]{fer97,zag97,fer04,sab04,lan07}. One interpretation for the origins of these peculiar spatial distributions has been the presence of two distinct formation mechanisms, with mass transfer dominating in the halos and collisions dominating in the core \citep{map06}. Alternatively the distribution may be a dynamical signature where encounters kick BSs out of the core leaving them to either relax back towards the center or remain in the halo depending on the local relaxation time \citep{sig94}. \begin{figure*}[!t] \begin{center} \plottwo{f20a.eps}{f20b.eps} \caption{\footnotesize Cumulative orbital period distributions comparing the NGC 188 blue stragglers to those in the \citet{hur05} simulation. The NGC 188 blue straggler orbital periods are shown in the black points, with the SB2s circled. The period distributions from the simulation are divided by formation channel, with mass transfer in the solid line, mergers in the dotted line and collisions in the dashed line. In the left panel we show only blue straggler binaries in the integrated sample from the model with $P < 3000$ days, while on the right we no longer limit by period. In the right panel the gray-filled region marks periods beyond our observational completeness limit. The simulation predicts that merger products will either be found in contact binaries (in the process of merging) or in wide binaries that result from exchange encounters, both of which are inconsistent with the NGC 188 blue straggler periods. Collision products that remain in binaries are predicted to have a significantly lower frequency of binaries with $P<3000$ days than observed for the NGC 188 blue stragglers. The mass-transfer products, on the other hand, all have periods within our completeness limit.} \label{BSPcum} \end{center} \end{figure*} The similarities in the observed properties between the core and halo NGC 188 BS samples suggest that in NGC 188 this spatial distribution is not the result of two distinct formation mechanisms. Instead, as discussed in \citet{gel11}, the majority of the NGC 188 BSs are consistent with an origin in mass transfer. Dynamical encounters in the core similar to those suggested by \citet{sig94} (though not resulting in collisions) may be responsible for the BS spatial distribution. Alternatively the majority of NGC 188 BSs may have formed with minimal dynamical disturbance following a very similar spatial distribution to the binaries in the cluster. The ``gap'' now seen in the spatial distribution between radii of $\sim$5 and $\sim$15 arcmin may then be a result of mass segregation processes, and may mark the region where the relaxation time approaches the age of the BSs. (A similar mechanism has also been suggested by other authors, e.g., \citealt{map04} and \citealt{lei11}, to explain the bimodal BS spatial distributions observed in some globular clusters.) \subsubsection{Testing Blue Straggler Formation Channels in the Cluster Simulation} \label{BSformch} There are three mechanisms for forming BSs in the \citet{hur05} $N$-body model: mergers, mass-transfer processes and stellar collisions. Here we investigate the binary populations for BSs formed from each mechanism as they appear in the integrated sample, and compare their binary characteristics to those observed for the NGC 188 BSs. In addition to completed mergers, we include in the merger category any binaries that are currently in contact or undergoing mass transfer and eventually merge. (In the \citealt{hur05} model, mergers result from Case A and/or B mass transfer. The model never contains a large enough population of triple stars to test BS formation through the recent Kozai-based merger mechanism of \citealt{per09}.) Between 2 and 4 Gyr in the simulation, 75\% (439/584) of the BSs in the integrated sample formed through mergers, 20\% (117/584) of the BSs formed through collisions and 5\% (28/584) of the BSs formed through mass transfer (that did not result in a merger). The relative rates of BSs created through each of these processes in the simulation are closely linked to the initial binary population. The dominant formation channel for the BSs in the simulation is mergers, a direct consequence of the over-abundance of short-period binaries included in the primordial population. We've shown that the initial binary population used in the simulation is not appropriate for NGC 188; thus the absolute and relative BS formation \textit{rates} cannot be expected to agree with the NGC 188 BSs. However the predictions for the BS binary \textit{properties} of each formation channel (e.g. distributions of binary orbital parameters) are only minimally sensitive to the initial binary population, and therefore still provide a valuable means to investigate the origins of the NGC 188 BSs. Here we use the \citet{hur05} models to define theoretical expectations for properties of the BSs formed by each channel and compare to the observed properties of the NGC 188 BSs. Our goal is to use these properties to begin to determine which mechanisms played significant roles in the formation of the NGC 188 BSs. We begin first with the BS binary frequency. As expected, the merger mechanism predicts a low BS binary frequency ($P<10^4$ days) of only 13~$\pm$~2~\% (61/439), much lower than the observed NGC 188 BS binary frequency. Furthermore, the majority of these binaries are contact systems, which we do not observe among the NGC 188 BSs. The collision mechanism predicts a total BS binary frequency of 74~$\pm$~8~\% (86/117). However, the vast majority (74\%; 64/86) of the BSs in binaries produced by collisions have periods longer than 3000 days, our completeness limit for orbital solutions in NGC 188 (see Section~\ref{incomp}). 67\% of the NGC 188 BSs have periods shorter than 3000 days, which contradicts the simulation result of only 19\% (22/117) binaries with periods of less than 3000 days from the collisional formation mechanism. The mass-transfer mechanism predicts a 100\% BS binary frequency, all of which are detached and have periods of less than 3000 days. Next we consider the distributions of orbital parameters predicted by the simulation for binaries formed by each mechanism, beginning with the period distributions. The predicted cumulative period distributions for the three mechanism are shown in Figure~\ref{BSPcum}; the left panel is limited to BS binaries with periods below our completeness limit while the right panel shows all BS binaries from the simulation. Two populations of merger products are evident in Figure~\ref{BSPcum}: a grouping of contact systems at short periods and a grouping at longer periods where the BSs were subsequently exchanged into binaries. The period distributions of both of these merger populations are inconsistent with the period distribution of the NGC 188 BSs. The distribution of collisionally produced BS binaries with periods of less than 3000 days is consistent with the shape of the observed distribution of NGC 188 BS binaries. However, as already noted, the frequency of such binaries in the simulation is much too small. \begin{figure}[!t] \begin{center} \plotone{f21.eps} \caption{\footnotesize Cumulative orbital eccentricity distributions comparing the NGC 188 blue stragglers to those in the \citet{hur05} simulation, for binaries with $P < 3000$ days. The NGC 188 blue straggler orbital eccentricities are shown in the black points, with the SB2s circled. The eccentricity distributions from the simulation are divided by formation channel, with mass transfer in the solid line, mergers in the dotted line and collisions in the dashed line. The $N$-body model predicts that collision products will have significantly higher eccentricities than observed in the NGC 188 blue straggler binaries. Both mass-transfer and merger products are predicted to have nearly all circular orbits. (The few non-circular orbits in these samples are the result of dynamical encounters.) The assumption of circular orbits for contact binaries is likely valid. However the assumptions used in the \citet{hur05} model of rapid circularization during mass transfer may not be valid in all cases \citep{sok00,bon08,sep09}. Thus the accuracy of the predicted eccentricity distribution for mass-transfer products from the model is uncertain.} \label{BSecum} \end{center} \end{figure} All of the mass-transfer BSs in the \citet{hur05} model have periods within our detection limit, and are split between one population near 100 days (a result of Case B mass transfer from a red giant) and a second population near 1000 days (a result of Case C mass transfer from an asymptotic giant).\footnote{\footnotesize Though Figure~\ref{BSPcum} appears to suggest that the \citet{hur05} model favors Case B over Case C mass transfer, the difference in frequency is instead due to two longer lived Case B products rather than a significantly larger number of BSs produced through Case B mass transfer.} Similarly all but one of the NGC 188 SB1 BSs are found with periods near 1000 days; the remaining SB1 BS has a period of $\sim$120 days. The theoretical and observed distributions are not formally distinguishable. The NGC 188 SB2 BS binaries both have very short periods around 5 days. Interestingly, the only formation mechanism in the M67 model that produces BSs with such short periods is the merger mechanism. However both NGC 188 BSs are in detached binaries. Thus the M67 model struggles to produce BSs with similar characteristics to the two SB2 BS binaries in NGC 188. \begin{figure}[!t] \begin{center} \plotone{f22.eps} \caption{\footnotesize Cumulative blue straggler mass distribution comparing the NGC 188 blue stragglers to those in the \citet{hur05} simulation. We normalize all masses by the respective main-sequence turnoff mass to facilitate the comparison. The integrated sample of blue stragglers from the simulation is limited here to only include binaries with $P < 3000$ days to reflect our observational completeness in orbital solutions. Our NGC 188 blue straggler mass estimates are shown in the black points, with the SB2s circled. The primary mass distributions from the simulation are divided by formation channel, with mass transfer in the solid line, mergers in the dotted line and collisions in the dashed line. The simulation predicts that collision products are the most massive, followed by mergers and then mass-transfer products. Both the collision and merger products are significantly more massive than the NGC 188 blue stragglers, while products of mass-transfer have a mass distribution that is statistically indistinguishable from that of the NGC 188 blue stragglers.} \label{BSM1cum} \end{center} \end{figure} The observed distribution of orbital eccentricities is a challenge for all of the formation mechanisms. Collision products that retain a binary companion are predicted to have significantly higher eccentricities than observed for the NGC 188 BSs; the simulation result is rejected at the $>$99\% confidence level (Figure~\ref{BSecum}). Conversely, the mass-transfer and merger mechanisms both predict that essentially all of the BSs in binaries should have circular orbits. In the model, any non-zero orbital eccentricity prior to the BS formation is quickly circularized by tides during the mass-transfer or contact binary stages. The few non-circular orbits in both of these samples are the results of subsequent dynamical encounters. Three of the SB1 BSs in NGC 188 do have orbits consistent with being circular. Theoretical main-sequence tidal circularization timescales for binaries at $\sim$1000-day periods are prohibitively long and cannot explain these circular orbits \citep{mei05}. Indeed, there are no circular MS binaries with periods near 1000 days in NGC 188. Thus these BS circular orbits may be evidence of a mass transfer origin. However, the other 10 long-period BS binaries in NGC 188 have measurable eccentricities, one as high as 0.8. Recently, observational and theoretical evidence has begun to suggest that mass transfer, particularly involving an asymptotic-giant star, will not always lead to rapid circularization. Examples of eccentric post-mass-transfer systems may include barium-star systems \citep{jor98}, post-AGB binaries \citep[][and references therein]{van99}, and possibly also the long-period eccentric BSs in the galactic halo \citep{car05}. Theoretical work has suggested mechanisms that can yield ``eccentricity-pumping'' \citep[e.g.][]{sok00,bon08,sep09}, which were not included in the \citet{hur05} model. Thus at this time the implications of the observed eccentricity distribution for the mass-transfer channel remains uncertain. The eccentricities of the two short-period SB2's are interesting. The double BS binary 7782 found in the cluster halo has a circular orbit, while the BS-MS pair 5078 has a small eccentricity of 0.17. Both are discussed in more detail in \citet{mat09}. 7782 is likely the result of a dynamical encounter and exchange that may have ejected it to the halo; presumably after such an encounter it would have at least initially had a highly eccentric orbit. On the other hand, 5078 has managed to sustain a small eccentricity against a theoretically short tidal circularization timescale. In the previous section we found that the BSs in binaries within the simulation are generally more massive than those in NGC 188, where all are $\lesssim$1.5 times the turnoff mass. If we divide the integrated BS binary sample from the simulation by formation mechanism, we find that the mean masses from mergers, collisions and mass transfer are $1.43 \pm 0.05$, $1.72 \pm 0.10$ and $1.3 \pm 0.03$ times the turnoff mass, respectively. Collision products in binaries are predicted to be the most massive BSs, followed by merger products and finally mass-transfer products. In Figure~\ref{BSM1cum} we compare the mass distributions from the simulated BSs to that of the NGC 188 BS binaries (with orbital solutions). K-S tests show that the NGC 188 BSs are significantly less massive than predicted by the collisional and merger hypotheses, at the $>$99\% and 95\% confidence levels, respectively. This finding is consistent with that of \citet{gle08}, who rule out recent formation of massive blue stragglers through collisions in NGC 188. Thus both collisions and mergers generally produce more massive BSs than those observed in NGC 188. On the other hand, a similar comparison between the predicted BS masses for mass-transfer products to the NGC 188 BS masses shows no significant distinction. Finally, and for completeness, we note that \citet{gel11} present an analysis of the secondary-mass distributions of the SB1 BS binaries of NGC 188 (e.g., Figure~\ref{qm2freqBS}) and those formed in a new $N$-body simulation of NGC 188, created using a very similar code to that used by \citet{hur05}, but with a more realistic initial binary population. They find that the observed secondary-mass distribution, narrowly peaked around a mean mass of 0.53 M$_{\odot}$, is not consistent at $>$99\% confidence with the theoretical secondary-mass distribution of BS binaries formed by collisions, which have a mean mass nearly twice that observed. On the other hand, the observed secondary-mass distribution is fully consistent with a population of carbon-oxygen WDs with masses between 0.5~M$_{\odot}$~and 0.6 M$_{\odot}$. They also find that formation of BSs in hierarchical triples via the Kozai-induced merger mechanism is not favored by the observed distribution, but is only rejected at the 98\% confidence level. \subsubsection{Summary} To summarize, our comparisons show that BSs formed by collisions that retain a companion will have a significantly lower binary frequency ($P < 3000$ days), longer orbital periods, higher orbital eccentricities, higher masses and higher-mass companions than are observed for the NGC 188 SB1 BS binaries. BSs formed by mergers as modeled by \citet{hur05} are also highly inconsistent with the observations, most notably having a significantly lower binary frequency, shorter periods, and higher masses than are observed. Thus we can confidently rule out both mergers within isolated close binaries and collisions as the dominant formation channel for the NGC 188 BS binaries. Predictions for mass-transfer products are consistent with most of the binary properties of the NGC 188 SB1 BSs, which comprise two-thirds of the NGC 188 BSs. We are unable to investigate the Kozai-induced merger mechanism within the \citet{hur05} model, as this model never contains sufficient triples. Importantly the Kozai-induced merger and mass transfer hypotheses make additional observable predictions that can distinguish between the two mechanisms. Mass transfer predicts WD companions while the Kozai-induced merger mechanism predicts predominantly MS companions. Relatively young ($\lesssim 0.4$ Gyr) WD companions are hot enough to contribute a significant excess in UV emission to the combined spectrum of the BS binary, and can be detected through HST observations. Additionally mass-transfer products are expected to show a higher depletion of lithium than merger and collision products \citep{lom95,san97,nor97,gle10}. (The field BS binaries for which lithium abundances are available all show lithium depletion; \citealt{car01}.) Mass-transfer processes may also produce modified CNO abundances \citep{she00} and enhancements in $s$-process elements \citep{car01} from the processed material of the giant star donor. Such observations of the NGC 188 BS population are currently underway and will soon help to further constrain the origins of the NGC 188 BSs. However mass transfer cannot explain the companion masses or orbital periods of the NGC 188 SB2 BSs. These two BS binaries along with S0182 of M67 \citep{san03} represent intriguing case-studies for future $N$-body simulations. We suspect that they are among the most direct observational evidence for the close dynamical encounters of binaries that $N$-body studies have predicted for decades. Additionally there are five BSs in NGC 188 that aren't detected as binaries. The M67 model suggests that these BSs likely did not all form through mass transfer, while both collisions yielding very wide companions and mergers predict such BSs. Finally, we anticipate that shifting the MS binary period distribution away from short-period binaries (to better match the observations of MS binaries) will shift the dominant formation mechanism away from mergers and thereby change the properties of the resultant BS population, perhaps in better agreement with the binary observations. But recall that the \citet{hur05} simulation successfully matched the number and CMD distribution of the M67 BSs. To retain this agreement in BS frequency with a reduced number of initial short-period binaries may require modification of the physics and efficacy of the formation channels. We will explore this in a future paper discussing a more accurate $N$-body simulation of NGC 188. \section{Summary and Conclusions} \label{summary} Through a series of papers (Papers 1,2, \citealt{mat09} and \citealt{gel11}) of which this is the most recent contribution, we study the dynamical state of the old (7 Gyr) open cluster NGC 188. In previous papers we identify a complete sample of single and binary cluster members among the solar-type stars in NGC 188, present orbital solutions for the majority of our detected binaries, and study the BSs in detail. Here we analyze the hard-binary population of the cluster, including the hard-binary frequencies and distributions of binary orbital parameters of the MS, giant and BS populations. We compare these results in detail to the observed Galactic field binary population of R10 as well as the simulated binary population in the \citet{hur05} $N$-body open cluster simulation to study the impact that relaxation processes and close dynamical encounters may have on the cluster's binary population. We find the properties of the solar-type MS binaries in NGC 188 to be consistent with similar binaries in the Galactic field. The NGC 188 MS binaries have a log-period distribution that increases towards our detection limit and is consistent with the log-normal distribution found by R10 for solar-type binaries in the Galactic field. The MS eccentricity distribution has a roughly Gaussian form (for $P_{circ} < P < 3000$ days), also consistent with R10. Likewise the MS binary secondary-mass and mass-ratio distributions are indistinguishable from those of similar binaries in the Galactic field. In NGC 188 both the secondary-mass and mass-ratio distributions appear to rise towards lower masses and mass ratios, as also observed by \citet{maz03} and \citet{duq91} in the field. Both NGC 188 distributions also show evidence for a secondary peak at mass ratios near unity, as observed by R10, \citet{tok00} and \citet{fis05} in the field. The observed mass-function distribution for the NGC 188 MS binaries is formally consistent with a secondary-mass distribution where secondaries are drawn from either a standard \citet{kro01a} IMF or a uniform mass-ratio distribution. The observed global MS hard-binary frequency in NGC 188 is 23~$\pm$~2~\%, which when corrected for our incompleteness results in a true MS hard-binary frequency of 29~$\pm$~3~\% for $P<10^4$ days. This hard-binary frequency is marginally higher than that of the R10 field within the same period range, of 19~$\pm$~2~\%. This somewhat higher hard-binary frequency in NGC 188 may be the result of dynamical relaxation processes where lower total mass single stars are preferentially ejected from the cluster, as is seen in $N$-body models such as \citet{hur05}. Two-body relaxation is evident in the central concentration of binaries in the cluster. Additional evidence for dynamical encounters may be seen in the population of high-eccentricity binaries in the core of the cluster but not in the halo (Figure~\ref{evr}). High-eccentricity binaries such as these are created within the \citet{hur05} model in the cluster core via dynamical interactions throughout the cluster lifetime. Importantly, the NGC 188 BS binaries are significantly different from the NGC 188 MS solar-type binaries, providing some of the most distinguishing characteristics of the BS population found to date. The observed BS hard-binary frequency in NGC 188 is 76~$\pm$~19~\%, three times the observed solar-type MS hard-binary frequency. The BS and MS hard binary frequencies are distinguishable at the $>$99\% confidence level. The BS binary $e - \log(P)$ distribution is distinct from that of the MS at the 99\% confidence level. The BS SB1 binaries in NGC 188 have particularly remarkable orbital characteristics. All but one have periods near 1000 days, most have modest eccentricities (with three in circular orbits), and their secondary-mass distribution is narrow and peaked, with a mean mass of $\sim$0.5~M$_{\odot}$~(Figure~\ref{qm2freqBS}). The two short-period ($P<10$ days) BS binaries are both observed as SB2s. These SB2 BSs have the most massive companions to any of the BSs in NGC 188. Indeed BS ID 7782 has a mass ratio of unity and is composed of \textit{two} BSs. We argue that these short-period BS binaries likely have a dynamical origin \citep{mat09}. SB2 BS, ID 5078, has a companion star on the upper MS. Normal stellar evolution tracks \citep{mar08} predict a mass for the BS primary star that is 15\% to 30\% lower than the mass we derive using the kinematic mass ratio. Thus normal stellar evolution tracks may underestimate the masses of BSs. Comparisons to the BS population in the \citet{hur05} $N$-body open cluster model show that we can rule out the hypothesis that the NGC 188 SB1 BS binaries have a collisional origin. Collision products that retain binary companions are predicted to have significantly higher eccentricities, higher masses, higher-mass companions, and a lower frequency of binaries with $P < 3000$ days than observed for the NGC 188 SB1 BS binaries. Likewise mergers within isolated binaries will not reproduce the binary properties of the NGC 188 BSs. BSs produced by isolated mergers are too massive, have a binary frequency that is much too low, and the few binaries have too short of periods. Predictions from the mass transfer mechanism are consistent with most of the binary properties of the NGC 188 SB1 BSs, which comprise two-thirds of the NGC 188 BSs. Mass transfer products are predicted to have a very high binary frequency, with long orbital periods, and companions of about 0.5~M$_{\odot}$, all closely consistent with the NGC 188 BSs (as also discussed in \citealt{gel11}). The \citet{hur05} model never contains a sufficient frequency of triples to investigate the Kozai-induced merger mechanism for BS formation \citep{per09}. \citet{gel11} find that the observations do not favor this formation mechanism, but the data only reject the hypothesis at the 98\% confidence level. Spectroscopic observations to determine the abundances of the SB1 BSs as well as HST observations aimed at detecting the FUV flux from the WDs predicted by the mass transfer mechanism are currently underway, and will further constrain the origins of the NGC 188 BSs. Comparing detailed observations of open cluster binaries with predictions from sophisticated $N$-body open cluster models is a powerful method for investigating the dynamical evolution of a binary population and the formation of the BSs. Comparisons with observations, like those presented in this paper, are also essential to determine the validity of the models. In this paper we show that accurately defining the initial binary population is critical. We suggest that this is best accomplished through comparisons with detailed observations of binary populations in young open clusters, although every indication to date is that the field binary population is a reasonable proxy. $N$-body simulations modeling NGC 188 and using an observationally defined initial binary population are underway and will soon allow for a detailed study of the dynamical evolution of the old open cluster NGC 188. \acknowledgments The authors express their gratitude to the staff of the WIYN Observatory without whom we would not have been able to acquire the thousands of superb stellar spectra upon which this work rests. We also thank the many undergraduate and graduate students who have helped to obtain these spectra over the years at WIYN for this project. We owe many thanks to J.~Hurley for access to the output of his simulations and for many insightful conversations. Thanks to T.~Mazeh for the helpful discussion on the mass-ratio and secondary-mass distributions. This work was funded by the National Science Foundation grant AST-0908082, the Wisconsin Space Grant Consortium, and the Lindheimer Fellowship at Northwestern University. Facilities: \facility{WIYN}, \facility{DAO:1.22m}, \facility{Hale} \bibliographystyle{apj}
\section{Introduction} The electronic structure of lanthanide and actinide compounds has a number of distinctive features that are manifestations of atomic $f$-electron physics in bulk solids, including strong on-site correlations and relativistic spin-orbit effects. The effects of chemical environment on the ground states and excitation spectra of $f$-electrons are particularly interesting, since they are responsible for splitting the otherwise $2J+1$-fold degenerate free-ion ground state $^{2S+1}L_{J}$, giving rise to rich physics and applications. Recently, actinide dioxides in the cubic fluorite structure have attracted renewed theoretical interest in the context of their use as nuclear fuels \cite{Petit2010PRB45108, Nakamura2010PRB155131,Zhang2010PRB144110, Meredig2010PRB195128, Alexandrov2010PRB174115, Suzuki2010PRB241103, Andersson2009PRB60101, Dorado2009PRB235125,Tiwary2009PRB174302, Geng2008PRB180101}. The crystal field (CF) method is a well established tool for describing the ligand environment of localized electrons. In its conventional form, which requires spectroscopic information for fitting the CF parameters, the CF method has been applied to $f$-electron compounds with considerable success \cite{Wybourne1965Spectroscopic, Newman2000, Liu2005Spectroscopic}, including numerous characterizations of actinide oxides.\cite{Rahman1966JPCS1833, Amoretti1989PRB1856, Amoretti1992JPCM3459, Kern1999PRB104,Gajek2004JMMM415,Magnani2005PRB54405, Magnani2007JPCS2020, Santini2000PRL2188, Kern1984SSC295, Boothroyd2001PRL2082, Magnani2008PRB104425} Using first-principles density functional theory (DFT) approaches, Divis and co-workers calculated the crystal field in praseodymium oxides \cite{ Divis2005JMMM1015,Novak2007PSSB3168}, and Colarieti-Tosti and co-workers studied PuO$_{2}$ \cite{Colarieti-Tosti2002PRB195102}. However, since the CF splitting is much weaker that the Coulomb repulsion and spin-orbit coupling (SOC), DFT-based calculations are often plagued by various technical issues, such as lack of a fully self-consistent treatment of the $f$-charge density or explicit consideration of electronic correlation. Recently Gaigalas and co-workers \cite{Gaigalas2009LJP403} calculated CF levels of actinide dioxides with relativistic quantum chemical methods. We have recently developed a fully self-consistent method of calculating the CF parameters, which combines an improved nonspherical self-interaction free LDA+$U$ scheme \cite{Zhou2009PRB125127} with a model on-site Hamiltonian including Coulomb, spin-orbit, and CF terms.\cite{Zhou2011PRB85106} Our approach utilizes the existence of multiple local minima in the LDA+$U$ total energy functional and uses the corresponding $f$-electron wavefunctions and total energies to extract CF parameters. Good agreement with experiment was obtained in terms of the predicted UO$_{2}$ CF excitation spectrum (within about $10$ to $20$ meV) and magnetic properties of UO$_{2}$.\cite{Zhou2011PRB85106} In this paper, we extend this method to calculate the CF parameters of other $f$-element dioxides MO$_{2}$ in the fluorite structure with the $f^{n}$ configuration, including $n=1$ (PaO$_{2}$, PrO$_{2}$), $n=3$ (NpO$_{2}$), and $n=4$ (PuO$_{2}$). Some results for UO$_{2}$ ($n=2$) are included for completeness. Other $f$-elements are not considered either because they have no valence $f$ electrons (CeO$_{2}$, ThO$_{2}$), no stable dioxides (heavier lanthanides), or no suitable pseudopotential presently available to us (AmO$_{2}$, CmO$_{2}$). \section{Method} The CF of MO$_{2}$ in the fluorite structure is given by: \begin{eqnarray} H_{\mathrm{CF}}&=& \frac{16\sqrt{\pi}}{3} V_{4} (Y_{4}^{0} + \sqrt{\frac{10}{7}} \mathfrak{Re} Y_{4}^{4}) \nonumber \\ &+& 32 \sqrt{\frac{\pi}{13}} V_{6} (Y_{6}^{0} - \sqrt{14} \mathfrak{Re} Y_{6}^{4}), \label{eq:CF} \end{eqnarray} where $V_{4}$ and $V_{6}$ are CF parameters of the cubically coordinated metal ion, related to other common CF notations \cite{Newman2000} by \begin{eqnarray*} V_{4}&=& B_{4}/8, \\ V_{4}&=& B_{6}/16. \end{eqnarray*} In addition, free-ion parameters $F^{k}$ ($k=2,4,6$) and $\zeta$ describe the Coulomb and SOC terms, respectively, in the total Hamiltonian: \begin{eqnarray} H &=& H_{\mathrm{CF}} + \hat{V}_{\mathrm{ee}} + \zeta \hat{\boldsymbol l} \cdot \hat{\boldsymbol s}. \label{eq:CI-Hamiltonian} \end{eqnarray} where $\hat{V}_{\mathrm{ee}}$ designates the Coulomb repulsion between $f$-electrons. Since the Slater integrals $F^{k}$ ($k=2,4,6$) in $\hat{V}_{\mathrm{ee}}$ are heavily correlated \cite{Carnall1992JCP8713}, the following approximation has been adopted:\cite{Berry1988CP105} \begin{eqnarray} F^2&=&F^4/0.668=F^6/0.494, \label{eq:Fk-to-J} \end{eqnarray} eliminating free parameters $F^{4}$ and $F^{6}$. The Hamiltonian of Eq.~(\ref{eq:CI-Hamiltonian}) is diagonalized with $f^{n}$ basis wavefunctions, which are chosen in this work as $n$-body Slater determinants constructed from 14 $f^{1}$ spin-orbitals $\{ Y_{3}^{m} \sigma \}$ ($m=-3 ,\cdots ,3$, $\sigma=\uparrow, \downarrow$). Therefore there are $C^{n}_{14}$ basis wavefunctions to expand an $f^{n}$ state. \begin{table}[htbp] \begin{ruledtabular} \begin{tabular}{|c|ccccc|} &PrO$_{2}$ &PaO$_{2}$ &UO$_{2}$ &NpO$_{2}$ &PuO$_{2}$\\ \hline $a$ (\AA)\cite{Villars1991PearsonHandbook} &5.386 &5.505 &5.470 &5.433 &5.396 \\ $f$-conf. & $f^{1}$ &$f^{1}$ &$f^{2}$ &$f^{3}$ &$f^{4}$\\ Ion GS. &$^{2}F_{5/2}$ &$^{2}F_{5/2}$ &$^{3}H_{4}$ &$^{4}I_{9/2}$ &$^{5}I_{4}$ \\ CF GS &$\Gamma_{8}$(4) &$\Gamma_{8}$(4) &$\Gamma_{5}$(3) &$\Gamma_{8}^{(1)}$(4) &$\Gamma_{1}$(1) \\ $J$ (eV) &0.78 &0.58 & 0.60 & 0.58 & 0.55 \\ \end{tabular} \end{ruledtabular} \caption{List of studied $f$-element oxides MO$_{2}$ (M=Pr, Pa, U, Np, Pu), including the lattice constants $a$, number of localized $f$-electrons $n$, free-ion and crystal field ground states (GS) and their multiplicities (in parentheses), and the $J$ parameter used in LDA+$U$ calculations.\cite{Zhou2009PRB125127} \label{tab:input}} \end{table} All DFT calculations were carried out using the same computational settings as in our previous work \cite{Zhou2011PRB85106}. Input parameters for the LDA+$U$ \cite{Anisimov1991PRB943, Zhou2009PRB125127} corrections are chosen as $U=6$ eV, $c=0.5$ and the $J$ parameter for exchange interactions are determined by the requirement of numerical degeneracy of degenerate ionic states \cite{Zhou2009PRB125127}. For each compound, 50 calculations with randomly initialized $f^{n}$ wavefunctions were carried out at the experimental lattice parameters (Table~\ref{tab:input}). The magnetization axis for analyzing the energy eigenstates is chosen along $z$. More details of our technical approach can be found in Ref.~\onlinecite{Zhou2011PRB85106}. \section{Results and discussions} The calculated model parameters are summarized in Table~\ref{tab:summary}. For comparison, we also give the free-ion parameters $F^{k}$ and $\zeta$ of tetravalent actinides in the corresponding fluorides.\cite{Carnall1989JLM201} The Coulomb interactions $F^{k}$ do not enter the Hamiltonian of the $f^{1}$ compounds PrO$_{2}$ and PaO$_{2}$. $F^{k}$ is found to be slightly smaller in NpO$_{2}$ than UO$_{2}$ and PuO$_{2}$, in agreement with the trend observed in MF$_{4}$. The SOC parameters $\zeta_{5f} \approx 0.2$ to $0.3$ eV of the heavier actinides are found substantially larger than lanthanide, since relativistic effects are more pronounced in heavier elements. The calculated $\zeta_{5f}$ of PaO$_{2}$ is almost twice as large as the corresponding $\zeta_{4f}=0.115$ eV in the rare earth compound PrO$_{2}$. $\zeta$ is predicted to increase over the actinide series, in agreement with experiment. However, our calculated $\zeta_{5f}$ values are overestimated by $5$ to $15\%$. Higher localization of the $4f$ states explains the smaller CF parameters $V_{4}$ and $V_{6}$ in PrO$_{2}$ compared to $5f$ actinides. The 4th-order CF parameter $V_{4}$ is significantly larger than $V_{6}$ for all the dioxides, in agreement with results obtained from fitting experimental spectra \cite{Amoretti1989PRB1856,Magnani2005PRB54405}. \begin{table}[htbp] \begin{ruledtabular} \begin{tabular}{|c|ccccc|} &PrO$_{2}$ &PaO$_{2}$ &UO$_{2}$ &NpO$_{2}$ &PuO$_{2}$\\ \hline $F^{2}$ & & &5.649 & 5.004 &6.147 \\ $\zeta$ &0.115 & 0.210 &0.230 &0.293 &0.304 \\ $V_{4}$ &-0.067 & -0.113 &-0.093 &-0.082 &-0.099 \\ $V_{6}$ &0.005 & 0.015 &0.016 &0.014 &0.017 \\ & \multicolumn{5}{c|}{Free-ion parameters in MF$_{4}$ from Ref.~\onlinecite{Carnall1989JLM201}} \\ $F^{2}$ & & &5.86 &5.55 &5.88 \\ $\zeta$ & & &0.22 &0.25 &0.28 \\ \hline & \multicolumn{5}{c|}{Excited CF levels} \\ State & $\Gamma_{7}(2)$ & $\Gamma_{7}(2)$ & $\Gamma_{3}(2)$ &$\Gamma_{8}^{(2)}(4)$ &$\Gamma_{4}(3)$ \\ & & &$\Gamma_{4}(3)$, $\Gamma_{1}(1)$ & $\Gamma_{6}(2)$ & $\Gamma_{3}(2)$, $\Gamma_{5}(3)$ \\ Pred. & 0.129 &0.186 &0.126, &0.034 &0.097 \\ & & &0.158, 0.176 & 0.125 & 0.195, 0.204 \\ \cline{2-6} Expt. &0.131 & n/a &0.150, & 0.055 & 0.123 \\ & & & 0.158, 0.170 & n/a & n/a \\ Ref. &\onlinecite{Kern1984SSC295}, \onlinecite{Boothroyd2001PRL2082} & & \onlinecite{Nakotte2010JPCS12002} & \onlinecite{Amoretti1992JPCM3459} & \onlinecite{ Kern1999PRB104} \\ \cline{2-6} Calc. & & & & & 0.099 \\ (Ref.~\onlinecite{Colarieti-Tosti2002PRB195102}) & & & & & 0.162, 0.208 \\ \cline{2-6} Calc. & 0.082 & & 0.167 & 0.056 & 0.112 \\ (Ref.~\onlinecite{Magnani2005PRB54405}) & & & 0.187, n/a & n/a & n/a \\ \cline{2-6} Calc. & & &0.155 & 0.034 & 0.064 \\ (Ref.~\onlinecite{Gaigalas2009LJP403}) & & & 0.161, 0.189 & 0.099 & 0.103, 0.127 \\ \cline{2-6} \end{tabular} \end{ruledtabular} \caption{Calculated model parameters and energy eigenvalues in eV. $F^{4,6}$ may be derived from eq.~\ref{eq:Fk-to-J}. Excited CF levels are labeled with the corresponding degeneracy in parentheses and compared with reported measured or calculated values. \label{tab:summary}} \end{table} Using the parameters given in Table~\ref{tab:summary}, the crystal field eigenstates are obtained by diagonalizing the effective Hamiltonian in Eq.~\ref{eq:CI-Hamiltonian}; the resulting wave functions are visualized in Fig.~\ref{fig:eigenstates}. Following the procedure of Refs.~\onlinecite{Sievers1982ZPB289,*Walter1986ZPB299}, the radius $R(\Omega=\theta,\phi)$ of the spherical plots of the charge distribution is $$ R(\Omega)= (\rho(\Omega)- \bar\rho )^{1/3}, $$ where $\rho(\Omega)$ is the spherical part of the charge distribution centered at the metal ion, and $\bar\rho$ is an appropriate amount of monopole subtracted from $\rho(\Omega)$ to emphasize its asymmetric character. \begin{figure}[htbp] \includegraphics[width=0.99 \linewidth]{eigenstates} \caption{The $f^{n}$ ($n=1,\cdots, 4$) eigenstates under the fluorite cubic crystal field. Only the $2J+1$ lowest states are shown.} \label{fig:eigenstates} \end{figure} The predicted low-energy CF excitation levels are shown in Table~\ref{tab:summary} alongside available experimental data. Qualitatively, the correct ground states and ordering of the excited states are obtained in all cases. Quantitatively, good agreement with the measured spectrum has been obtained, with the errors in the excitation energies being within 10 to 20 meV. Reasonable agreement has also been found with previous theoretical calculations reported in Refs.~\onlinecite{Colarieti-Tosti2002PRB195102, Magnani2005PRB54405, Gaigalas2009LJP403}. Next, we discuss each compound in detail (except UO$_{2}$). \subsection{PrO$_{2}$ and PaO$_{2}$} The $f^{1}$ compounds PrO$_{2}$ and PaO$_{2}$ differ from $n>1$ cases in that the multiple local-minima issues that plague calculations for multi-$f$ electron systems are less severe. Out of the 50 random wave function initializations, approximately 5\% with highly unstable starting states failed to converge within 100 electronic steps and were discarded. The rest exhibited a relatively small energy spread and were all within $0.2$ to $0.3$ eV from the CF ground state, compared to the spread of about 2 eV observed in UO$_{2}$.\cite{Zhou2011PRB85106} This shows that the many-body interaction is a main reason for the existence of many local-minimum solutions, and without this obstacle the $f^{1}$ calculations can find the $j=5/2$ Russell-Sanders ground state, even though they may fail in locating the CF ground state. The original LDA+$U$ \cite{Anisimov1991PRB943} scheme was tested for PrO$_{2}$ and PaO$_{2}$ and found to increases significantly the energy spread of the local-minimum solutions due to orbital-dependent self-interaction errors \cite{Zhou2009PRB125127}. The predicted $\Gamma_{8} \rightarrow \Gamma_{7}$ excitation energy for PrO$_{2}$ is 129 meV, in excellent agreement with the measured value of 131 meV \cite{Boothroyd2001PRL2082} and more accurate than our previous rough estimation of $73$ to $142$ meV in Ref.~\onlinecite{Zhou2009PRB125127}, showing that our method based on Eq.~(\ref{eq:CI-Hamiltonian}) leads to significant error cancellation in the calculated CF energies. Predictions for the higher CF levels of $J=7/2$ are 0.376 ($\Gamma'_{6}$), 0.433 ($\Gamma'_{8}$) and 0.622 eV ($\Gamma'_{7}$), respectively, compared with observed values of 0.320, 0.390 and 0.580 eV from Ref.~\onlinecite{Boothroyd2001PRL2082}. The relative splitting within the $J=7/2$ manifold agrees very well with experiment, showing the validity of our predicted CF parameters, while the center of these levels are 11\% too high, due to the over-estimated spin-orbit coupling (our $\zeta=0.115$ eV compared to 0.1 eV of Ref.~\onlinecite{Boothroyd2001PRL2082}). Experiments on PaO$_{2}$ are relatively scarce. The only available number of 140 meV for the $\Gamma_{8} \rightarrow \Gamma_{7}$ transition cited in Ref.~\onlinecite{Konings2004JCT121} is based on private communications, which we inquired about but could not confirm. Our prediction of 186 meV for $\Gamma_{8} \rightarrow \Gamma_{7}$ in PaO$_{2}$ is substantially larger than the corresponding value for PrO$_{2}$, in agreement with the trends in CF parameters in Table~\ref{tab:summary}. \begin{table}[tbp] \newlength{\graphheight} \graphheight 0.6cm \begin{ruledtabular} \begin{tabular}{|l|cc|cc|} \multicolumn{1}{|c|}{State} & \multicolumn{2}{c|}{PrO$_{2}$} & \multicolumn{2}{c|}{PaO$_{2}$} \\ &$\mu_{S}$ & $\mu$ &$\mu_{S}$ & $\mu$ \\ \hline (1) $\Gamma_{8}$ \includegraphics[height=\graphheight]{f1-mag001-1}& -0.64& 1.49 & -0.57 & 1.54 \\ (2) $\Gamma_{8}$ \includegraphics[height=\graphheight]{f1-mag001-2}& -0.03 & 0.47 & -0.10 & 0.45 \\ (3) $\Gamma_{8}$ \includegraphics[height=\graphheight]{f1-mag001-3}& 0.03 & -0.47 & 0.10 & -0.45\\ (4) $\Gamma_{8}$ \includegraphics[height=\graphheight]{f1-mag001-4}& 0.64 & -1.49 & 0.57 & -1.54 \\ (5) $\Gamma_{7}$ \includegraphics[height=\graphheight]{f1-mag001-5}& 0.06 & 0.73 & 0.10 & 0.70\\ (6) $\Gamma_{7}$ \includegraphics[height=\graphheight]{f1-mag001-6}& -0.06 & -0.73 & -0.10 & -0.70 \\ \end{tabular} \end{ruledtabular} \caption{The $\Gamma_{8}$ ground state quartet and $\Gamma_{7}$ excited doublet of PrO$_{2}$ and PaO$_{2}$ ($f^{1}$) in the fluorite structure, and the corresponding spin and total magnetic moment in $\mu_{B}$.} \label{tab:f1} \end{table} \subsection{UO$_{2}$} Recent measurement by \textcite{Nakotte2010JPCS12002} of the crystal field levels in UO$_{2}$ provides updated information than \textcite{Amoretti1989PRB1856}: the excitation peak at 180 meV is spurious. \subsection{NpO$_{2}$} NpO$_{2}$ has the $5f^{3}$ configuration. The excitation energy between the CF ground state $\Gamma_{8}^{(2)}$ and the first excited $\Gamma_{8}^{(1)}$ state has been measured to be 55 meV \cite{Amoretti1992JPCM3459}. Our prediction of 34 meV is a reasonable under-estimation. Note that a recent quantum chemical calculation \cite{Gaigalas2009LJP403} for NpO$_{2}$ predicted the same value as ours. Two estimated values for the second excited $\Gamma_{6}$ energy level (145 meV and 274 meV) are given in Ref.~\onlinecite{Amoretti1992JPCM3459}, and only the first value scales over the actinide dioxides series \cite{Magnani2005PRB54405}. We predict an excitation energy of 125 meV for $\Gamma_{8}^{(2)} \rightarrow \Gamma_{6}$, which agrees with the latter assessment. \begin{table}[tbp] \graphheight 0.6cm \begin{ruledtabular} \begin{tabular}{|c|ccc|} State &$\mu_{S}$ & $\mu$ & Proj. \\ \hline $a$ \includegraphics[height=\graphheight]{f3-mag001-1}& -1.00& 1.72 & $0.68(1,2,3)+0.57(1,5,6) $\\ $b$ \includegraphics[height=\graphheight]{f3-mag001-2}& -0.37 & 0.57 & $0.68(1,2,4)+0.57(2,5,6) $\\ $c$ \includegraphics[height=\graphheight]{f3-mag001-3}& 0.37 & -0.57 & $0.68(1,3,4)+0.57(3,5,6) $\\ $d$ \includegraphics[height=\graphheight]{f3-mag001-4}& 1.00 & -1.72 & $0.68(2,3,4)+0.57(4,5,6) $\\ \end{tabular} \end{ruledtabular} \caption{The $\Gamma_{8}^{(2)}$ ground state quartet of NpO$_{2}$ ($5f^{3}$), along with their magnetic moment and projection onto Slater determinants (only a few leading terms shown) composed of $f^{1}$ eigenstates $\Gamma_{8}$ and $\Gamma_{7}$. One-electron orbitals (1) through (6) are defined in Table~\ref{tab:f1}.} \label{tab:f3} \end{table} Since the $f^{1}$ configuration in the fluorite structure is split by the crystal field into the $\Gamma^{8}$ quartet and $\Gamma_{7}$ doublet (see Table~\ref{tab:f1}), the multi-electron configurations $f^{n}$ of UO$_{2}$, NpO$_{2}$ and PuO$_{2}$ are sometimes interpreted within a picture where the added electrons gradually fill the CF levels, in analogy to the well-known scenario of $d$-electrons filling the $t_{2g}$ and $e_{g}$ CF levels in transition metal compounds. As we have shown previously \cite{Zhou2011PRB85106}, this picture fortuitously holds for the $\Gamma_{5}$ ground state of UO$_{2}$. However, transition metal CF splittings are usually several eV, while for $f$ the CF splittings (on the order of 0.1 eV) is much smaller than the effective Coulomb interactions ($\sim$ eV). Hence, the $f^{n}$ eigenstates are in general multi-configurational. According to Table \ref{tab:f3}, the $\Gamma_{8}^{(2)}$ ground states of NpO$_{2}$ are composed of multiple determinants, including ones with substantial projections onto not only the $\Gamma_{8}$ CF ground states, but also the $\Gamma_{7}$ excited states of $f^{1}$. In other words, the $f^{3}$ ground state $\Gamma_{8}^{(2)}$ occupies both $\Gamma_{8}$ and $\Gamma_{7}$ orbitals in order to lower its electrostatic energy at the expense of a slightly increased CF energy. \subsection{PuO$_{2}$} PuO$_{2}$ has the $5f^{4}$ configuration. The crystal field was measured by Kern and co-workers using inelastic neutron scattering (INS) \cite{Kern1999PRB104}. Our calculated $\Gamma_{1} \rightarrow \Gamma_{4}$ excitation energy of 97 meV agrees reasonably well with the measured value of 123 meV \cite{Kern1999PRB104} and a previous calculation of 99 meV by Colarieti-Tosti {\it et al.\/}\cite{Colarieti-Tosti2002PRB195102}. Note that the splitting was under-estimated in all the calculations, including this work and Refs.~\onlinecite{Colarieti-Tosti2002PRB195102, Magnani2005PRB54405, Gaigalas2009LJP403}. The non-magnetic $5f^{4}$ ground state $\Gamma_{1}$ with $\mu_{S}=\mu=0$, \begin{eqnarray*} \mathrm{\includegraphics[height=6mm]{f4-mag001-1}} = 0.7(1,2,3,4)+ 0.32 (1,4,5,6) + 0.32 (2,3,5,6) + \ldots, \end{eqnarray*} is sometimes referred to as four fully filled $\Gamma_{8}$ $f^{1}$ orbitals. We obtained $|\langle \Gamma_{1}| 1,2,3,4 \rangle|^{2}$ = 0.49, showing that such a simplified picture of 4 filled $\Gamma_{8}$ orbitals is not entirely valid and multi-electron correlations account for more than 50\% of the ground state wave function. \begin{figure}[tbp] \includegraphics[width=0.8 \linewidth]{entropy} \caption{Predicted electronic entropy of MO$_{2}$ from the calculated CF levels compared to experimental estimation.} \label{fig:entropy} \end{figure} As a simple application of the CF calculations, Fig.~\ref{fig:entropy} shows the calculated electronic entropy $$ S_{\mathrm{e}} = - \sum_{i} p_{i} \ln p_{i}, $$ where $p_{i}= e^{-E_{i}/k_{B}T}/ \sum_{j} e^{-E_{j}/k_{B}T}$ is the Boltzmann probability of the electronic eigenstate $i$. As shown by Konings,\cite{Konings2004JCT121} the vibrational contribution to the total entropy of actinide oxides varies smoothly across the elemental series, while electronic contributions, which depend delicately on the CF excitation energies, cannot be interpolated over the series. To accurately predict thermodynamic properties of actinide oxides, the electronic entropy cannot be ignored. Our predicted $S_{\mathrm{e}}$ (solid curves) agree reasonably well with the results of Ref.~\onlinecite{Konings2004JCT121} (crosses) at $T=298.15$ K. In conclusion, we have calculated the CF levels of PrO$_{2}$, PaO$_{2}$, NpO$_{2}$, and PuO$_{2}$. The $f$-electron charge density and on-site correlations are calculated fully self-consistently within a version of LDA+$U$ that removes orbital-dependent self-interaction energies. Good agreement with experimental CF levels and a consistent trend across the actinide series have been achieved. In both NpO$_{2}$ and PuO$_{2}$, substantial contributions of the $\Gamma_{7}$ one-electron excited state are found in the multi-electron crystal field ground states. This work was supported by the U.S. Department of Energy, Nuclear Energy Research Initiative Consortium (NERI-C) under grant No.\ DE-FG07-07ID14893, and used resources of the National Energy Research Scientific Computing Center, which is supported by the DOE Office of Science under Contract No.\ DE-AC02-05CH11231.
\section{introduction} Recent discoveries pointing to novel ground states and unusual behaviors in field and temperature dependent magnetic and thermodynamic properties have led to renewed interest in the physics of frustrated spin systems. Delafossite structure PdCrO$_2$ has attracted recent attention in this context. Importantly, like the sister compound PdCoO$_2$, it can be grown in high quality form facilitating experimental investigation of its physical properties. \cite{takatsu,takatsu2,takatsu3} This material can be described as a stacking of 2D CrO$_2$ layers, separated by Pd, with the Cr ions arranged on a triangular lattice. In a standard nearest neighbor Heisenberg picture, the ground state would be expected to be a non-collinear 120$^\circ$ spin structure below the Neel temperature, $T_N$=37.5 K, \cite{mekata} but the actual magnetic structure of PdCrO$_2$ could be more complex. \cite{rastelli} Takatsu and co-workers find that the magnetic Bragg peaks in neutron scattering are rather broad and half-integer $l$ (1/3,1/3,$l$) and (2/3,2/3,$l$) peaks are found in addition to the integer $l$ peaks. \cite{takatsu} In any case, PdCrO$_2$ is a particularly interesting system because like its sister compound PdCoO$_2$ it is conducting, \cite{wichainchai} and has a very large conductivity anisotropy. \cite{takatsu10} This offers an opportunity to study the interplay between frustrated magnetism and charge carriers in a near two dimensional metal. Furthermore, recent experiments have revealed an unconventional anomalous Hall effect in this material. \cite{takatsu2} Doumerc and co-workers reported susceptibility data up to $\sim$ 450 K. \cite{doumerc} They obtained an effective moment of 4.1 $\mu_B$ and a Weiss temperature of -500 K, in a Curie-Weiss fit, although they qualify these values as approximate. The data show a poor fit except over a relatively small temperature range above $\sim$ 300 K, in contrast to the other compounds they report: CuCrO$_2$, AgCrO$_2$ and CuFeO$_2$, which show very good Curie-Weiss behavior over a wide temperature range extending to below 100 K. In fact, in their data, there is still an apparent upward curvature to $\chi^{-1}(T)$ at the highest $T$, suggesting that the Weiss temperature could be lower in magnitude or that the system may have itinerant character. Takatsu and co-workers show more details of the susceptibility below 350 K. \cite{takatsu} There is no plausible Curie-Weiss fit to the data, but they do state that their data are consistent with the values extracted by Doumerc and co-workers. Interestingly, the data shows a broad maximum above $T_N$. However, in contrast to strongly 2D magnetic systems, $\chi$ just above $T_N$ is only slightly lower than the maximum value, and there are strong signatures of the ordering in $\chi(T)$ as well as resistivity and specific heat. This is not really consistent with expectations for a strongly 2D Heisenberg system. On the other hand, Takatsu and co-workers do observe a sublinear resistivity, $\rho(T)$ above $T_N$, which they attribute to short range magnetic correlations, and an unusual behavior in the specific heat also above $T_N$. Motivated by this we performed density functional calculations of electronic structure and energetics of PdCrO$_2$ in order to address the nature and role of the metallic conduction electrons in PdCrO$_2$ and the extent to which the system is a realization of a 2D nearest neighbor Heisenberg magnet. We find in contrast to previous assumptions that (1) there is a considerable interplay between the Cr moments and the metallic electrons, and (2) there are strong magnetic interactions along the $c$-axis direction, so that even though PdCrO$_2$ is very two dimensional as an electron gas, it is a very three dimensional magnetic system. Finally, there is metal-metal bonding, consistent with previous studies of delafossite compounds. \cite{seshadri} \section{approach} We performed density functional calculations with the Perdew, Burke, Ernzerhof (PBE) generalized gradient approximation, \cite{pbe} using the all electron linearized augmented planewave (LAPW) method \cite{singh-book} as implemented in the WIEN2k code, \cite{wien} similar to our previous calculations for PdCoO$_2$. \cite{ong} We carefully tested the convergence of our results against the various parameters, including tests with different sphere radii and different choices of the augmentation including both LAPW and so-called APW+lo basis sets. \cite{sjostedt} We found the results to be stable. We used the experimental lattice parameters, $a$=2.923 \AA{} and $c$=18.087 \AA{} (delafossite structure, space group $R\bar{3}m$, Pd at ($\frac{1}{2}$,$\frac{1}{2}$,$\frac{1}{2}$), Cr at (0,0,0), O at ($\pm z$,$\pm z$,$\pm z$), in rhombohedral coordinates). \cite{shannon} Refinement of the internal parameter corresponding to the O height above the Cr plane has not been reported in literature to our knowledge. As such, we calculated this parameter for a ferromagnetic ordering and used it in the other calculations. We obtained $z$=0.6111, which yields a Cr-O distance of 1.962 \AA. This is in accord with the expected bond length for high spin Cr$^{3+}$ (the Shannon radii are 1.22 \AA{} for O$^{2-}$ and 0.755 \AA{} for octahedral Cr$^{4+}$, summing to 1.975 \AA). We also note that the Pd-Pd distance of $a$=2.923 \AA{} is not much longer than the Pd-Pd distance of 2.75 \AA{} in Pd metal. The Cr-O-Cr bond angles in the layers are 96.3$^\circ$. Considering the Cr lattice, the O atoms directly connect the nearest neighbor Cr atoms in plane. There are no direct O connections to second neighbors. Secondly, in the out-of-plane direction, the shortest hopping paths are Cr-O-Pd-O-Cr. The Cr-O-Pd bond angles along this path are 120.7$^\circ$ with this O position. \section{moment formation and electronic structure} We find that PdCrO$_2$ is strongly unstable against Cr moment formation regardless of the arrangement of the moments with an energy of $\sim$1 eV/Cr. We also find all arrangements are metallic in accord with experiment. Therefore, PdCrO$_2$ should be described as a metal containing Cr local moments. We start our description with the calculated electronic structure with ferromagnetic alignment of these moments. The calculated energy is 0.92 eV lower than the non-spin-polarized (no Cr moments) case, and the magnetization is 2.87 $\mu_B$ per formula unit. The magnetization inside a Cr LAPW sphere (radius 2.05 bohr) is 2.57 $\mu_B$. \begin{figure} \includegraphics*[height=0.95\columnwidth,angle=270]{dos-f.ps} \caption{(color online) Density of states and projections on Cr and Pd LAPW spheres of radius 2.05 bohr for ferromagnetic ordering} \label{dos-f} \end{figure} \begin{figure} \includegraphics*[width=0.45\columnwidth,angle=0]{fs-f-up.eps} \includegraphics*[width=0.45\columnwidth,angle=0]{fs-f-dn.eps} \caption{(color online) Fermi surface for ferromagnetic ordering. Left panel is majority spin, right panel is minority. Note that these are hole cylinders centered at $\Gamma$.} \label{fs-f} \end{figure} The calculated electronic density of states (DOS) is given in Fig. \ref{dos-f}. It clearly shows crystal field split Cr $d$ states with the majority spin $t_{2g}$ states occupied, and the $e_g$ unoccupied, as are all the minority spin Cr states. The DOS at $E_F$ is Pd derived. These states provide a simple very two dimensional Fermi surface (Fig. \ref{fs-f}), consistent with the observed strong conductivity anisotropy and reminiscent of PdCoO$_2$, which is also a highly anisotropic metal. \cite{ong} As discussed previously, the states giving rise to the Fermi surface arise from Pd-Pd metal-metal bonding as in other delafossites. \cite{seshadi} Additionally, two additional features are seen: (1) The Fermi surface has very little rounding, which is the condition for maximizing nesting (but note that density of states is Pd derived and low, $N(E_F)$=0.7 eV$^{-1}$, per formula unit on a both spins basis for this ferromagnetic ordering), and (2) While both the majority and minority spin Fermi surfaces are similar in shape, they are not identical, implying an interplay between the metallic, largely Pd derived, conduction band, and the Cr derived magnetic moments. In fact, the minority spin Fermi surface is 13\% larger in volume than the majority spin surface. This interplay between Cr moments and the conduction electrons provides a mechanism for spin scattering, that will increase the resistivity as $T$ disorders the spins and also a mechanism for long range Cr-Cr interactions in-plane, as well as interactions in the $c$-axis direction. \section{magnetic interactions} \begin{figure} \includegraphics*[width=\columnwidth,angle=0]{hexagons.eps} \caption{(color online) Structures and energetics of small supercells, in meV per Cr, relative to the ferromagnetic ordered configuration. The top panel is for ferromagnetic layers, stacked along the $c$-axis either ferromagnetically, or antiferromagnetically, while the bottom panel shows three in-plane antiferromagnetic orderings. The unit cells are indicated by dotted black lines. The three numbers in parentheses below each structure denote the numbers of opposite spin neighbors, at the the nearest neighbor position (out of six), the in-plane next nearest Cr neighbor position (out of six) and the out-of-plane nearest neighbor position (out of six), respectively.} \label{hex} \end{figure} We studied the magnetic interactions through a series of supercell calculations with different magnetic orderings. These calculations and their energetics are summarized in Fig. \ref{hex}. One may immediately note from the top panel of the figure that the out-of-plane magnetic interactions are large. Specifically, an antiferromagnetic stacking of ferromagnetic layers is found to be 18 meV/Cr lower in energy than a ferromagnetic stacking. This is large compared with the ordering temperature, $T_N$= 38 K, i.e. $kT_N$= 3.3 meV. \begin{figure} \includegraphics[width=0.80\columnwidth,angle=0]{fs-af-plot.eps} \caption{(color online) Fermi surface viewed down the $c$-axis for the lowest energy AF-2 configuration (see text), shown in an extended zone scheme. The white lines show the supercell Brillouin zone and the highlight sketches the non-reconstructed Fermi surfaces of Fig. \ref{fs-f}.} \label{fs-af} \end{figure} Of the small collinear cells that we studied, the lowest energy belongs to AF-2 (see Fig. \ref{hex} for notation), which corresponds to the ordering in CuFeO$_2$. \cite{terada,ye,nakajima} The electronic structure for the various the orderings remains very anisotropic at the Fermi energy, where the states are Pd derived. However, as mentioned above, there is a strong interplay between the Fermi surface and the Cr moments. For example, the Fermi surface for the lowest energy magnetic configuration is shown in Fig. \ref{fs-af}. The figure shows overlapped hexagonal sections, as expected for the folded zone, but also sizable splittings at the intersections. This interplay between magnetic order and the conduction electrons is also seen in the calculated densities of states. As mentioned, for the ferromagnetic order, we obtain $N(E_F)$=0.70 eV$^{-1}$ on a per formula unit both spins basis. For the AF-z (see Fig. \ref{hex}) order, we obtain $N(E_F)$=0.69 eV$^{-1}$ on the same basis, while we obtain $N(E_F)$= 0.81 eV$^{-1}$, for AF-1, $N(E_F)$= 0.76 eV$^{-1}$, for AF-2, and $N(E_F)$= 0.77 eV$^{-1}$, for AF-3. The bare specific heat coefficient $\gamma$ inferred from these values is 1.6-1.9 mJ/mol K$^2$, comparable to the specific heat value of Takatsu and co-workers of 1.4$\pm$0.2 mJ/mol K$^{2}$ in the ground state. \cite{takatsu} This leaves little room for any renormalization at low $T$. The variability of the electronic structure at $E_F$ depending on magnetic state provides a qualitative framework for understanding why the resistivity of high quality crystals of PdCrO$_2$ is so much higher than that of PdCoO$_2$, even though the Fermi surface, structure and band character at the Fermi surface are very similar, \cite{ong,ong2} and also why the resistivity shows a strong signature of the magnetic ordering, with a pronounced decrease in resistance as $T$ is lowered through $T_N$. \cite{takatsu} Specifically, this connection between magnetic order and the electrons at the Fermi surface indicates that strong spin-fluctuation scattering is expected above the ordering temperature, freezing out as $T$ is lowered below $T_N$. Materials with this type of coupling between magnetism and electrons at the Fermi surface tend to be the ones that display unusual properties, such as superconductivity (as in e.g. the Fe-pnictides where the coupling is very strong) when the magnetism is suppressed. Therefore, this coupling of spin fluctuations to electrons at the Fermi surface suggests that it will be of considerable interest to experimentally examine what happens as the magnetic order is suppressed, e.g. by pressure or alloying. We now turn to the energetics (Fig. \ref{hex}) in more detail. To do this we write a short range model with interactions to nearest and next-nearest neighbors in-plane and to the nearest out-of-plane neighbor: \begin{equation} E= E_{(F)} + j_1 N_{1\downarrow} + j_2 N_{2\downarrow} + j_z N_{z\downarrow} \end{equation} \noindent where $E$ is the energy per Cr, $E_{(F)}$ is the energy of the ferromagnetic ordering, $N_{1\downarrow}$ is the average number of opposite spin first nearest neighbors in the structure, and similarly for next nearest neighbors and neighbors along the $c$-axis direction. While this could be converted into a three neighbor Heisenberg model, we write it this way because the moments in the Cr LAPW spheres vary between the different configurations (in the range 2.48 $\mu_B$ to 2.57 $\mu_B$). In any case, if we match the energies of the ordered structures, F, AF-z, AF-1 and AF-3, we obtain $j_1$= -16 meV, $j_2$= +2 meV and $j_z$= -3 meV. The energies cannot be reasonably reproduced without all three parameters (note that each shell has 6 atoms, so e.g. the difference between ferromagnetic and AF-z is 6$j_z$=-18 meV, which is not small, etc.) Use of these values, yields the energy for the AF-2 state to within 1 meV / Cr atom. While the model is perhaps too simple to capture all aspects of the magnetic interactions in PdCrO$_2$, it does clearly show that the material cannot be reasonably understood as a simple 2D nearest neighbor Heisenberg model. \section{discussion and conclusions} Considering the ordering temperature $kT_N$=3.3 meV, it would seem that none of these interactions can be neglected, i.e. at least first and second in-plane neighbors and crucially c-axis interactions need to be included when discussing the magnetism of PdCrO$_2$. This conclusion may seem unexpected considering the very two dimensional electronic structure at the Fermi energy, as found both in our calculations and as is clear from the experimental resistivity anisotropy. \cite{takatsu10} However, we note that a similar behavior is found in the layered cobaltate Na$_x$CoO$_2$. \cite{bayrakci,helme} That materials has a related crystal structure, that is similar to the delafossite, except that the layer stacking differs and bridges between planes are O neighbors along $c$ instead of two O separated by a Pd (i.e. in Na$_x$CoO$_2$ the O in neighboring layers are directly on top of each other). The three dimensional magnetic behavior in Na$_x$CoO$_2$ arises in part because of this bonding topology, which provides many superexchange exchange paths between the neighboring CdI$_2$ structure CoO$_2$ layers. \cite{johannes} Within this framework, the c-axis magnetic interactions and the coupling to the conduction electrons are inter-related since they are both mediated by or through Pd. In any case, we note that the Kosterlitz-Thouless type of suppression of the ordering temperature is logarithmic in the anisotropy of the interactions. As such, very strong anisotropy is needed to obtain two dimensional, as opposed to three dimensional, magnetism. We can conclude that PdCrO$_2$ is a three dimensional frustrated antiferromagnet with an interesting interplay between magnetism and conduction electrons. \acknowledgements Work at ORNL was supported by the Department of Energy, Basic Energy Sciences, Materials Sciences and Engineering Division. Work at IHPC was supported by the Singapore Agency for Science Technology and Research (A*STAR). DJS is grateful for the hospitality of IHPC where a portion of this work was performed.
\section{Introduction} Prestellar cores are the earliest identified phases of star formation. They are centrally condensed, likely to be gravitationally bound and destined to ultimately form stars or clusters. They have power-law density profiles with flat inner regions, implying that no protostar has yet formed in their center \cite[e.g.][]{War1999, Bac2000}. Their linewidths are thermalized, implying no significant turbulence support \cite[e.g.][]{Mye1983, Bar1998, Goo1998, Kir2007}. They show no evidence for substantial amounts of rotation \citep{Goo1993,Bar1998,Cas2002}. Intensity ratios of different transitions show that starless cores are almost isothermal, with $T\sim 10$ K \citep{Taf1998,Taf2002,Taf2004}, although there may be small variations of the order of a few K \cite[e.g.][]{Eva2001}. Their evolution depends critically on the physical conditions of the star forming region in which they are embedded. In addition, their relatively quiescent dynamical state allows for an easier interpretation of observations because of the absence of the effects of thermal and dynamical feedback from a central protostar. For these reasons prestellar cores are ideal probes of the ultimate stellar origins: the initial conditions of star formation. Prestellar cores are observed to be magnetic \cite[see, e.g.][for a review]{Hei2005}. The magnetic field can provide support against the self-gravity of a molecular cloud or core, and thus affect its dynamical evolution. The amount of magnetic support is quantified by the mass to magnetic flux ratio $M/\Phi_B$ of the object under consideration. There is a critical value for the mass-to-flux ratio, \beq \left(\frac{M}{\Phi_B}\right)_{\rm crit} = \left(\frac{1}{63G}\right)^{1/2} \eeq \citep{Mou1976} where $G$ is the gravitational constant. If the mass-to-flux ratio of an object exceeds this value (the object is magnetically supercritical), then the magnetic field is not strong enough to support the object against its own self-gravity, and the object contracts dynamically. The opposite is true for mass-to-flux ratios below the critical value (magnetically subcritical objects). Magnetic models of cloud fragmentation and core collapse predict that prestellar cores are magnetically {\em supercritical}, dynamically collapsing fragments, formed in magnetic parent clouds. These parent clouds which can be magnetically subcritical as a whole, and the supercritical fragments are formed through the process of ambipolar diffusion. The latter is the process of neutral particle diffusion through ions and magnetic field lines toward centers of gravity which increases the mass-to-flux ratio of the fragment \cite[e.g.][]{Fie1993}. Determining the amount of magnetic support in a molecular cloud as a whole and in individual molecular cloud cores is made complicated by at least the following three effects. First of all, Zeeman measurements only trace the component of the magnetic field along the line of sight, and for this reason geometrical considerations due to unknown cloud and core orientations limit the amount of information that can be obtained on an object-to-object basis \cite[e.g.][]{Shu1999,Tro2008,MT09}. Second, in the case of Zeeman measurements, the magnetic field strength is convolved with the abundance profile of the tracer molecule, and, as a result, it is non-trivial to determine which parts of a core contribute most to a finite-beam Zeeman measurement of its magnetic field. Understanding the time evolution and spatial variation of the abundance of different species is essential. Observations show that when such measurements are made using different molecular lines, they reveal different values of the magnetic field which cannot be reconciled. A characteristic example is CN and OH Zeeman observations of the same objects yielding different results \cite[e.g.][]{Fal2008}. Similarly, \citet{Cru2000} have argued, in the case of L1544, that OH data (and the associated Zeeman measurement of the magnetic field) do not sample the small, dense core observed in dust emission, so \citet{Cru2004} concluded that their Chandrasekhar-Fermi--measured plane-of-sky magnetic field may be discrepant from the Zeeman-measured line-of-sight magnetic field of \citet{Cru2000}. This may be a result not only of geometrical projection effects, but also because different measurements correspond to different parts of the core. To verify the origin of such discrepancies, we need to understand how different molecules trace different parts of cores and thus different magnetic field values. This becomes especially important when attempting to study the variation of the magnetic field strength with density \cite[e.g.][]{Cru2010}. Third, by the time the column density contrast between core and cloud becomes large enough for the core to be detected with high statistical significance and to be studied in detail, the core has already left the quasistatic contraction phase and has entered the phase of dynamical collapse, even in models which are originally heavily magnetically supported \citep{TM04}. At this stage, the mass-to-flux ratio is already larger than critical regardless of its value at larger scales \cite[e.g.][and references therein]{MTK06}. At even higher central densities, during the very advanced stages of collapse, the dynamical evolution of cores and the resulting density and velocity profiles have been shown to be quite insensitive to the initial value of the mass-to-flux ratio \citep{TM07b}. In addition to the mass-to-flux ratio, the effect of the magnetic field on the dynamics of molecular clouds and molecular cloud cores also depends on the degree of ionization of the cloud or core. The ambipolar diffusion timescale is proportional to the degree of ionization, \begin{equation} \tau_{\rm AD} \propto \frac{n_i}{n_{\rm H_2}} \end{equation} \citep{Cio1993}. This dependence is physically straight-forward to understand: a higher degree of ionization implies a better coupling between magnetic field and matter, which results in an increased resistance of the ions as the neutrals drift past them and consequently an increased ambipolar diffusion timescale. The degree of ionization of a cloud or core is determined by the ionization rate (in dense cores the dominant ionization mechanism is cosmic ray ionization, and hence the relevant quantity is the cosmic ray ionization rate $\zeta$) and by the relevant recombination reactions. Because of its feedback effect on the dynamics, and especially in the case where the chemistry is out of equilibrium, the degree of ionization is non-trivial to obtain; ideally, it has to be self-consistently calculated through a model following both dynamics and chemistry simultaneously. For these reasons, the role of magnetic fields in the fragmentation of molecular clouds and the core formation and evolution process remains observationally uncertain and a hotly debated subject in the field. In a companion paper \cite[][hereafter Paper I]{PaperI} we discussed our models of evolving prestellar cores which couple non-equlibrium chemistry with dynamics. Our extensive parameter study included magnetic and non-magnetic dynamical models with varying initial values of the mass-to-flux ratio or collapse retardation times respectively, varying C/O ratios, cosmic-ray ionization rates, and temperatures. In Paper I we focused on the evolution of the abundances of the most abundant and commonly observed molecules, and their dependence on the various model parameters. Here, we focus on the properties of the model cores that can help elucidate the role of magnetic fields in core formation and evolution: the degree of ionization, the most common ions and the mean ion molecular weight, abundance profiles of molecules used in Zeeman observations, and how the magnetic field that would be measured through Zeeman observations in our model cores depends on the actual strength of the magnetic field at the core center. This paper is organized as follows. A brief review of the different models we consider is given in \S \ref{mod}. The results and their dependence on the various parameters we have studied are presented in \S \ref{res}. We summarize and discuss our conclusions in \S \ref{disc}. \section{Models}\label{mod} \begin{figure} \plotone{Models.eps} \caption{\label{ModVis} Line types and colors used to denote each of the models studied in this work, unless otherwise noted. Solid normal-thickness red line: ``reference'' magnetic model; solid normal-thickness blue line: ``reference'' non-magnetic model. Dotted lines: ``fast'' models; dashed lines: ``slow'' models. Brown/purple lines: magnetic/nonmagnetic models with temperatures differing from the ``reference'' models. Orange/cyan shaded areas: variation in C/O ratio. Thin/thick solid red/blue lines: lower/higher cosmic-ray ionization rate magnetic/non-magnetic models (see text for details). } \end{figure} Since in this paper we are interested not only in the measurement of the magnetic fields but also in the evolution of the degree of ionization, we consider both magnetic and non-magnetic models. The details of the dynamical and chemical models of each class is discussed in detail in \S 2 and 3 of Paper I. Here, we very briefly review the parameters we have varied for each class of models presented and discuss in the following sections. These parameters include the temperature, the C/O ratio, the cosmic ray ionization rate, and a parameter controlling the time available for chemical evolution. The latter is the mass-to-flux ratio in the case of magnetic models, and the collapse delay time in the case of non-magnetic models (an initial time period during which chemistry evolves but the core does not evolve dynamically, representing an early stage of support due to turbulence which later decays). Our ``reference'' magnetic model has a mass-to-flux ratio equal to the critical value for collapse, a temperature of 10 K, a C/O ratio of 0.4, and a cosmic ray ionization rate of $\zeta = 1.3 \times 10^{-17} {\rm \, s^{-1}}$. Our ``reference'' non-magnetic model has a collapse delay time (hereafter ``delay'') of 1Myr, and values for the C/O ratio, temperature, and $\zeta$ identical to those of the ``reference'' magnetic model. For magnetic models we examine two additional values of the initial mass to magnetic flux ratio: 1.3 times the critical value (a faster-evolving, magnetically supercritical model), and 0.7 of the critical value (a slower, magnetically subcritical model). For non-magnetic models we examine two additional values of delay: zero, and 10Myr. For each of these six dynamical models the carbon-to-oxygen ratio is varied from its ``reference'' value by keeping the abundance of C constant and changing that of O. The two other values of C/O ratio examined are 1 and 1.2. We have studied in this way a total of 18 different models (9 magnetic and 9 non-magnetic). In addition, to test the effect of the temperature, we have varied each of the six basic dynamical models by changing $T$ by a factor of $\sim 1.5$ from its reference value of 10 K and examined models with $T = 7$ K and $T = 15$ K. We have thus studied 12 models (6 magnetic and 6 non-magnetic) with temperature varied from its reference value. Finally, to test the effect of the cosmic ray ionization rate, we have studied four additional models (two magnetic and two non-magnetic), which have a ``reference'' value for the temperature, C/O ratio, and mass-to-flux ratio or delay (for magnetic and non-magnetic models respectively), but for which $\zeta$ is varied by a factor of four above ($\zeta = 5.2 \times 10^{-17}$ $s^{-1}$) and below ($\zeta = 3.3 \times 10^{-18}$ $s^{-1}$) its ``reference'' value (covering the range of observational estimates \cite[e.g.][]{McC2003,Hez2008}. These additional models bring the total of different models we have run and examined to 32. Figure \ref{ModVis} visually depicts these different models and the line type/color used to denote each one (unless explicitly noted otherwise). \begin{figure*} \plotone{plotB2abund_multi_new.eps} \caption{\label{Bmulti} Left column: upper panel: radial profiles of the $z$-component (solid lines) and $r$-component (dotted lines) of the $B-field$ at four different times $t_1-t_4$ (corresponding to the same snapshots for which the average $B_z$ traced by OH and CN is shown in Fig.~\ref{meanBz}); lower panel: radial profiles of the number density at the same times $t_1-t_4$. Right column: radial profiles of the OH (upper panel) and CN (lower panel) abundance taken at the same times as in the left column. } \end{figure*} \begin{figure*} \epsscale{0.8} \plotone{plotcolabundi3_multi_lin4.eps} \caption{\label{Colden} Column density of CO, NH$_3$, OH, and CN plotted against the fractional radius of the core, for a central density of $10^6 {\rm cm^{-3}}$ (corresponding to snapshot $t_3$ in Fig.~\ref{Bmulti}). The three red lines correspond to the different magnetic dynamical models (magnetically critical, subcritical and supercritical for the solid, dashed, and dotted lines respectively) for a reference value of the C/O ratio equal to 0.4, while the yellow line corresponds to a magnetically critical model with C/O ratio equal to 1 (see Fig.~\ref{ModVis}).} \end{figure*} \section{Results}\label{res} \subsection{Zeeman-Traced Core Magnetic Field} \begin{figure} \plotone{plotmeanBvsn.eps} \caption{\label{meanBz} Mean value of the z-component of the magnetic field of a core traced by OH (solid lines) and CN (dotted lines) for the initially magnetically subcritical (diamonds), critical ($\times$), and supercritical ($+$) cloud models. For comparison, the initial magnetic field of the model cloud is $B_z=$4.2, $B_z=$5.6, and $B_z=$7.5 for the magnetically supercritical, critical and subcritical cloud models, respectively. } \end{figure} \begin{figure*} \plotone{plotxecvsn_multi.eps} \caption{\label{degion} Evolution of central electron abundance versus central number density in magnetic and non-magnetic models. In all panels, red lines correspond to magnetic models with reference values for C/O and temperature (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and blue lines to non-magnetic models with reference values for C/O and temperature (dashed: 10Myr delay; solid: 1Myr delay; dotted: no delay). The black dashed line corresponds to Eq.\ (\ref{canonical}). Left panel: effect of a varying C/O ratio. The cyan and orange shaded areas (for non-magnetic and magnetic models respectively) correspond to a range of C/O values between $0.4$ (reference value) and $1.2$. Middle panel: effect of a varying core temperature. The thin and thick brown (purple) solid lines correspond to temperature of 7 K and 15 K respectively for magnetic (non-magnetic) models; the ``reference'' value for the temperature is 10 K. Right panel: effect of a varying cosmic-ray ionization rate. The thick and thin red (blue) solid lines correspond to $\zeta$ a factor of four above and below the ``reference'' value for magnetic (non-magnetic) models (see Fig.~\ref{ModVis}). } \end{figure*} In this section we examine how the magnetic field value that is measured through finite-beam Zeeman observations compares to the magnetic field at the center of a core. We consider two molecules commonly used for Zeeman observations, OH and CN. Figure \ref{Bmulti} shows radial profiles of the magnetic field (left column, upper panel; $z-$ and $r-$components are shown with solid and dotted lines respectively), the OH abundance (right column, upper panel) and the CN abundance (right column, lower panel) at four time instances $t_1-t_4$, each of which represents an increase in the H$_2$ number density at the center of the core of an order of magnitude. The central number density profiles at these times are shown in the lower panel of the left column. These radial profiles correspond to the magnetically critical (``reference'') dynamical model. Note that although CN and OH appear significantly depleted at the very central parts of the core, these results are consistent with observations by \citet{HBetal10}, who find that significant amounts of CN remain at densities of $\sim 3\times 10^4 {\rm \, cm^{-3}}$, where CO has already depleted. Comparing with the lower-left panel of Fig.~\ref{Bmulti}, we find that significant depletion of CN (abundance more than an order of magnitude smaller than in the outskirts of the core) only sets in at densities higher than $\sim 10^5 {\rm \, cm^{-3}}$. To facilitate comparison with such observations, we show, in Fig.~\ref{Colden}, the column density of four molecules of interest (CO, NH$_3$, OH, and CN) as a function of (linearly plotted) fractional radius of the core, for a central density of 10$^6$ cm$^{-3}$, corresponding to snapshot $t_3$ in Fig.~\ref{Bmulti}, when the central abundance of CN for example is more than five orders of magnitude lower than that in the outskirts of the core. The three red lines correspond to the different magnetic dynamical models (magnetically critical, subcritical and supercritical for the solid, dashed, and dotted lines respectively) for C/O ratio equal to 0.4, while the yellow line corresponds to a magnetically critical model with C/O ratio equal to 1. A peak at relatively large radius and then a plateau/mild decrease of the column density toward the center is a sign of depletion. In this context, CO and NH$_3$ represent extreme cases of significant and little depletion respectively. OH and CN are intermediate cases, both peaked at smaller radii than CO, in agreement with the findings of \citet{HBetal10}. If C/O is 1 rather than the fiducial 0.4 then CN exhibits a central decrease/plateau at much smaller radii, while CO and OH are significantly depleted. In Fig.~\ref{meanBz} the mean magnetic field that would be seen through Zeeman measurements by a 0.1pc beam (extending out to a radius where the neutral number density falls to about $10^4 \, {\rm cm^{-3}}$) is plotted against the actual central value of the $z-$component of the magnetic field. The mock Zeeman-obtained ``observed'' value of the field is derived as follows. We calculate the average $z$-component magnetic field in the magnetic core under consideration, weighted by the number density of the corresponding molecule (OH or CN respectively) within a radial extent of 0.1 pc (the assumed size of the beam) at four different time instances. The core is assumed to be viewed face-on (down the $z-axis$). In this way, geometrical effects do not enter our calculation of the mock Zeeman observation (the $z-$component of the magnetic field, which we examine here, is oriented exactly along the line of sight), and we can instead focus only on the effect of depletion. We then compare each of the averages derived in this way to the value of the z-component of the magnetic field at the center of the core at the same time. In practice, this means that $\langle B_z \rangle$ is calculated through weighting by the OH or the CN number density in each radial annulus within the beam, respectively, assuming optically thin lines. Three different dynamical models are shown: magnetically subcritical (diamonds), critical ($\times$) and supercritical ($+$). Solid lines are used for the magnetic field traced by OH and dashed lines for the magnetic field traced by CN. In all cases we can see that, although at low central densities the value of the Zeeman-traced magnetic field is close to the actual value of the central $B_z$, at higher densities the measured field increases only mildly and shows an overall increase by only a factor of two, whereas the actual central value of $B_z$ has increased by an order of magnitude. Most of the increase in the Zeeman-measured field strength takes place at low densities, while at higher densities the measured field tends to saturate. The reason for this behavior is seen in Fig.~\ref{Bmulti}: the abundance of OH and CN is falling at the center of the core much more rapidly with time than the density increases. As a result, the Zeeman-traced magnetic field corresponds to the outer layers of the core ``onion skin'', where OH and CN have still relatively higher abundances \subsection{Degree of ionization} \begin{figure*} \plotone{plotxecvsn_pie.eps} \caption{\label{degion_pie} Degree of ionization as a function of central number density $n_c$ for the three magnetic (upper panels) and non-magnetic (lower panels) models, with ``reference'' values for the C/O ratio and the ionization rate $\zeta$. Left column: ``fast'' models; middle column: ``reference'' models; right column: ``slow'' models. Pie charts show the relative contribution of the dominant ions at the central densities denoted by stars. } \end{figure*} Figure \ref{degion} shows the evolution of electron abundance (degree of ionization) versus central number density in magnetic and non-magnetic models. In all panels red lines correspond to magnetic models with reference values for C/O and temperature (dashed: magnetically subcritical; solid: magnetically critical; dotted: magnetically supercritical), and blue lines to non-magnetic models with reference values for C/O and temperature (dashed: 10Myr delay; solid: 1Myr delay; dotted: no delay; see also Fig.~\ref{ModVis}). The left panel shows the effect of varying the C/O ratio, the middle panel the effect of varying the temperature, and the right panel the effect of varying the degree of ionization. For comparison we overplot with the black dashed line the usually adopted scaling [obtained through fits to earlier calculations, e.g., \citet{Elm1979,Nak1979}] between degree of ionization and number density, \begin{equation}\label{canonical} \frac{n_i}{n} = K_0\left(\frac{n}{10^5{\rm cm^{-3}}}\right)^{k-1} \end{equation} \citep{BM94}, with $k=0.5$ and $K_0 = 3\times 10^{-8}$. It is obvious from Fig.\ \ref{degion} that at low central densities the out-of-equilibrium chemistry causes a variety of transitional effects, and the scaling of the degree of ionization with density cannot be expressed in a simple form such as Eq. (\ref{canonical}). For central densities higher than $\sim 10^5 {\rm \, cm ^{-3}}$ for magnetic models and $\sim 10^6 {\rm \, cm^{-3}}$ for non-magnetic models the scaling does asymptotically approach a simple power-law. The magnetic models approach power-law scaling at lower central densities because of the increased time available to the chemistry to overcome transitional effects before a specific density is reached. The slope of this scaling (-0.6, corresponding to $k=0.4$) is steeper than the value usually adopted, but well within the uncertainties quoted in \citet{BM94}, who give a range for $k$ between $0.3$ and $0.5$. Our value for the slope is also consistent with the findings of \citet{Cas2002a} for the case of L1544. The normalization of our models is also consistently lower than Eq.\ (\ref{canonical}), but the discrepancy is not larger than the uncertainty on $K_0$ (an order of magnitude). Except for the earliest evolutionary times that are dominated by transitional effects in chemistry, the slope of the scaling initially steepens (the value of $k$ in Eq.\ \ref{canonical} decreases) with increasing density, in agreement with \citet{Cio1994}. However, once $k$ reaches 0.4, the scaling of the degree of ionization with density becomes a power law, in contrast to \citet{Cio1994}, who found that $k$ continues to decrease and the scaling continues to steepen. This is likely a result of the (equilibrium) chemistry network adopted by \citet{Cio1994}, which was considerably more simplified than the (non-equilibrium) chemical model we use here, and which, as we will see below, resulted in a considerably different ion population, dominated by different species. The non-magnetic models show a qualitatively different behavior, with the scaling of the degree of ionization with density becoming more shallow before it later steepens. At sufficiently high central densities the non-magnetic models asymptotically approach the same power laws as the magnetic models with the same value of $\zeta$. The normalization of the scaling changes with $\zeta$ approximately as $\zeta^{1/2}$, consistent with the usually assumed scaling \cite[see, e.g., ][Eq.\ 24]{McK2007}. The dependence at high densities on temperature is small (the asymptotic normalization being $\propto T^{1/2}$), however at low densities it can be considerable. Similarly, the dependence on the C/O ratio can be appreciable at low densities. At sufficiently late evolutionary times and their associated high central densities magnetic and non-magnetic models with the same value of $\zeta$ and $T$ converge to the same scaling, regardless of delay time, C/O ratio, or mass-to-flux ratio. We note that the high-temperature non-magnetic model has not yet converged to its asymptotic form by the highest density displayed in Fig. \ref{degion}; for this reason, we have extended this run to even higher central densities, and we have confirmed that above $10^7 {\rm cm^{-3}}$ the degree of ionization does indeed approach the corresponding scaling of the magnetic model with the same values of $\zeta$ and $T$. \begin{figure*} \plotone{plotmuioncvsn_multi.eps} \caption{\label{muion} Mean molecular weight $\mu_i$ at the center of the core, plotted as a function of central number density $n_c$. Panels, lines, and colors as in Fig.~\ref{degion} (see also Fig.~\ref{ModVis}). } \end{figure*} \subsection{Dominant ions and mean molecular weight} Figure \ref{degion_pie} shows the degree of ionization at the center of the core as a function of central number density $n$ for the three magnetic (upper panels) and non-magnetic (lower panels) models, with ``reference'' values for the C/O ratio, temperature, and cosmic-ray ionization rate $\zeta$. ``Reference'' dynamical models are shown in the middle column, with ``fast'' and ``slow'' models in the left and right columns, respectively. The pie charts above the lines show the relative contribution of the dominant ions to the total ion population at the densities marked by stars. Due to our adopted initial condition, which assumes all C to be ionized, C$^+$ is initially the dominant ion. However, once the chemistry starts operating, this evolves to a different mixture. Non-magnetic models have a more diverse mixture of ions at early times, with significant contributions from HCO$^+$ and H$_3$O$^+$. At late evolutionary times H$_3^+$ is the dominant ion in all models, with H$^+$ and He$^+$ being the most important secondary contributors (grain growth, not considered here, may alter the relative importance of H$_3^+$ and H$^+$ in favor of H$^+$, see \citealp{Fl2005}). This is an important difference from previous studies \cite[e.g.][]{Cio1993,Des2001,TM2007,McK2010}, which have assumed that the ion population at comparable densities is dominated by HCO$^+$ or by metallic ions such as Na$^+$, both of which have much higher molecular weight, implying that the ion fluid has much higher inertia at a given number density. Interestingly, the highest-density pie chart for all magnetic models and the ``slow'' non-magnetic model are very similar, so this ion distribution appears to be the one to which the ionization profile of the core settles at late enough times. For magnetic models the ion with the next highest contribution at low densities is C$^+$. Non-magnetic models, on the other hand, feature additional substantially contributing ions, such as C$_3$H$_3^+$, H$_3$O$^+$, HCO$^+$. Figure \ref{muion} shows the evolution with central number density of the mean molecular weight of ions $\mu_i$. Panels, lines and colors correspond to various magnetic and non-magnetic models as described in Fig.~\ref{degion} (see also Fig.~\ref{ModVis}). As expected from the results of Fig.\ \ref{degion_pie}, contrary to the usually adopted values of $\mu_i$ which are in the range of 20-30, the mean molecular weight of ions generally stays below 20. At high densities $\mu_i$ asymptotically approaches 3 for all cases except the high-temperature non-magnetic model, consistent with H$_3^+$ dominating the ions in these cases (see also \citealp{Cas2002a}). The non-magnetic "slow" model appears significantly different (has a much lower mean molecular weight of ions at low densities) than all other models because before the gas starts to collapse, the chemistry has already had 10Myr to evolve, and the mixture of ions is already close to the characteristic "late time" mix of the other models. The low values of $\mu_i$ (for most of the central number density range) in the magnetic models implies that the ambipolar diffusion timescale is {\em shorter} than the one computed using the usually adopted higher value for $\mu_i$, due to the smaller inertia of the ion fluid (which is attached to the magnetic field lines at these densities) and its associated decreased ability to provide resistance to the diffusion of the neutral fluid towards centers of gravity. The C/O ratio is shown to affect the mean molecular weight appreciably. Its effect is stronger in the non-magnetic models, since oxygen and carbon-bearing molecules are more abundant in the gas phase in faster-evolving models such as the ``reference'' and ``fast'' nonmagnetic models. Indeed, the ``slow'' nonmagnetic model shows a much smaller sensitivity to the C/O ratio, comparable to that of the magnetic models. Similarly, the temperature has a significant effect on $\mu_i$, mainly through its effect on the chemistry of N-bearing molecules (see also discussion in Paper I). Specifically, we have verified that the increase in the mean molecular weight with temperature in the non-magnetic model (which persists even at high densities) is due to an increased abundance of N$_2$H$^+$. By contrast in the case of the magnetic high-temperature models, H$^+$ is the dominant ion at intermediate central densities and $\mu_i$ decreases below three. \section{Conclusions}\label{disc} We have examined the effect that non-equilibrium chemistry in dynamical models of collapsing cores has on molecular abundances and thus on measurements of the magnetic field in these cores, on the degree of ionization and on the mean molecular weight of ions. We have considered both magnetic and non-magnetic models and models with different C/O ratios, cosmic-ray ionization rates, and core temperatures. We have found that molecules usually used in Zeeman observations of the line-of-sight magnetic field (OH and CN) have an abundance that decreases toward the center of the core much faster than the density increases (see Fig.\ \ref{Bmulti}). Thus, Zeeman measurements tend to sample the outer layers of the core and consistently underestimate the core magnetic field, especially for higher-density cores with higher magnetic field strengths. The degree of ionization was found to follow a complicated dependence on the number density in early evolutionary phases, when central densities are less than $10^5 \, {\rm cm^{-3}}$ for magnetic models and $10^6 \, {\rm cm ^{-3}}$ for non-magnetic models. At higher densities the scaling approaches a power-law with a slope of -0.6 (slightly steeper than the usually assumed value of -0.5) and with a normalization that scales with the product of $\zeta$ and $T$ to the 1/2 power; however, we note that for non-magnetic models with temperatures higher than 10 K, the dependency does not attain its asymptotic form until much higher central densities. The mean molecular weight of the ions was found to be systematically lower than the usually assumed value of $20-30$, and, at high densities, to asymptotically approach a value of 3 for all models, due to the asymptotic dominance of H$_3^+$. The only exceptions are the 15 K non-magnetic models (no delay and 1Myr delay), for which N$_2$H$^+$ becomes dominant at higher densities due to the sensitivity of nitrogen chemistry to temperature (see paper I, \S 5.3). At low densities the dominating ions and the associated value of $\mu_i$ evolve with density, and exhibit sensitivity to the value of the C/O ratio, $\zeta$, and the temperature. The considerably lower value of $\mu_i$ for the magnetic models compared to the usually assumed one implies that ambipolar diffusion operates faster. \acknowledgements{We thank Paul Goldsmith and the anonymous referee for insightful and constructive comments that improved this paper. This work was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. \copyright 2012. All rights reserved.}
\section{Introduction} Perturbative QCD corrections are well-known for being the dominant contributions to the radiative inclusive penguin decay~\cite{Hurth:2010tk,Hurth:2007xa,Hurth:2003vb}. This perturbative dominance was recently reassured by a dedicated analysis~\cite{Benzke:2010js} in which non-perturbative corrections to the inclusive decay mode $\bar B \rightarrow X_s \gamma$ have been estimated to be well below $10\%$. Within a global effort, a perturbative QCD calculation to the next-to-next-to-leading-logarithmic (NNLL) level within the Standard Model (SM) has been performed and has led to the first NNLL prediction of the $\bar B \to X_s \gamma$ branching fraction~\cite{Misiak:2006zs}. Using the photon energy cut $E_0=1.6$ GeV, the branching ratio reads \begin{equation}\label{final1} {\cal B}(\bar B \to X_s \gamma)_{\rm NNLL} = (3.15 \pm 0.23) \times 10^{-4}. \end{equation} This result is based on various highly-nontrivial perturbative calculations ~\cite{Misiak:2004ew,Bobeth:1999mk,Gorbahn:2004my,Gorbahn:2005sa,Czakon:2006ss,Blokland:2005uk,Melnikov:2005bx,Asatrian:2006ph,Asatrian:2006sm,Bieri:2003ue,Misiak:2006ab}. The combined experimental data according to the Heavy Flavor Averaging Group (HFAG)~\cite{hfag} leads to \begin{equation} {\cal B}(\bar B \rightarrow X_s \gamma) = (3.55 \pm 0.24 \pm 0.09) \times 10^{-4} \,, \end{equation} where the first error is combined statistical and systematic, and the second is due to the extrapolation in the photon energy. Thus, the SM prediction and the experimental average are consistent at the $1.2 \sigma$ level. This is just one example among the impressive confirmation of the SM in all experiments in flavour physics during the last decade~\cite{Antonelli:2009ws, Buchalla:2008jp}, including the first generation of the $B$ factories at KEK (Belle experiment at the KEKB $e^+e^-$ collider)~\cite{Belle} and at SLAC (BaBar experiment at the PEP-II $e^+e^-$ collider)~\cite{Babar}, and the Tevatron $B$ physics programs (CDF~\cite{TevatronB1} and D0~\cite{TevatronB2} experiments). Also the first results of the LHCb experiment~\cite{LHCb} are in full agreement with the simple CKM theory of the SM. This feature is somehow unexpected because in principle flavour changing neutral current (FCNC) processes like $\bar B \to X_s \gamma$ offer high sensitivity to new physics (NP). Additional contributions to the decay rate, in which SM particles in the loops are replaced by new particles such as the supersymmetric charginos or gluinos are not suppressed by the loop factor $\alpha/4\pi$ relative to the SM contribution. Thus, FCNC decays provide information about the SM and its extensions via virtual effects to scales presently not accessible otherwise. This approach is complementary to the direct production of new particles at collider experiments. \section{Supersymmetric flavour problem} The experimental fact that none of the dedicated flavour experiments has observed any unambiguous sign of new physics yet, in particular no ${\cal O}(1)$ NP effects in any FCNC process, implies the famous flavour problem, namely why FCNC processes are suppressed. It has to be solved in any viable new physics model. The hypothesis of minimal flavour violation (MFV)~\cite{Chivukula:1987py,Hall:1990ac,D'Ambrosio:2002ex}, i.e.\ that the NP model has no flavour structures beyond the Yukawa couplings, solves the problem formally. However, new flavour structures beyond the Yukawa couplings are still compatible with the present data~\cite{Hurth:2009ke} because the flavour sector has been tested only at the $10\%$ level in the $b\to s$ transitions. Today supersymmetric models are often given priority in our search for NP beyond the SM. This is primarily suggested by theoretical arguments related to the well-known hierarchy problem. Supersymmetry eliminates the sensitivity for the highest scale in the theory and, thus, stabilizes the low energy theory. There are other features in supersymmetric theories which are promising like the unification of the gauge couplings and the existence of a dark matter candidate. Supersymmetry also represents the unique extension of Poincare symmetry. The precise mechanism of the necessary supersymmetry breaking is unknown. A reasonable approach to this problem is the inclusion of the most general soft breaking term consistent with the SM gauge symmetries in the so-called unconstrained minimal supersymmetric standard model (MSSM). This leads to a proliferation of free parameters in the theory. The decay $\bar B \rightarrow X_s \gamma$ is sensitive to the mechanism of supersymmetry breaking because, in the limit of exact supersymmetry, the decay rate would be just zero: \begin{equation} {\cal B} (\bar B \to X_s \gamma)_{Exact\, Susy} = 0. \end{equation} This follows from an argument first given by Ferrara and Remiddi in 1974 \cite{Ferrara}. In that work the absence of the anomalous magnetic moment in a supersymmetric abelian gauge theory was shown. In the MSSM there are new sources of FCNC transitions. Besides the CKM-induced contributions, which are brought about by a charged Higgs or a chargino, there are generic supersymmetric contributions that arise from flavour mixing in the squark mass matrices in case they are not aligned to the ones in the quark sector. Then the gluino contribution enhanced by an extra factor $\alpha_s$ instead of $\alpha_{\rm weak}$ significantly contributes to the decay rate. Thus, the general structure of the MSSM does not explain the suppression of FCNC processes, which is observed in experiments; the gauge symmetry within the supersymmetric framework does not protect the observed strong suppression of the FCNC transitions. This is the crucial point of the well-known supersymmetric flavour problem. \section{Parameter bounds from the inclusive decay $\bar B \rightarrow X_s \gamma$} Parameter bounds on NP from flavour physics is a model-dependent issue. The present data on $\bar B \to X_s \gamma$ implies a very stringent bound for example on the inverse compactification radius of the minimal universal extra dimension model (mACD) ($1/R > 600 {\rm GeV}$ at $95\%$ CL)~\cite{Haisch:2007vb}. The bound is much stronger than the ones derived from other measurements. Moreover, there is a bound induced by $\bar B \rightarrow X_s \gamma$ on the charged Higgs mass in the two Higgs-doublet model (II): $M_{H^+} > 295{\rm GeV}$ at $95\%$ CL~\cite{Misiak:2006zs}. It is based on a NLL QCD calculation within this model presented in Refs.~\cite{Ciuchini:1997xe,Borzumati:1998tg}. The latter bound is not valid in general two-Higgs doublet models, especially in supersymmetric models. However, the two-Higgs-doublet model (II) is a good approximation for gauge-mediated supersymmetric models with large $\tan\beta$, where the charged Higgs contribution dominates the other supersymmetric contributions. Simplifying assumptions about the parameters often introduce model-dependent correlations between different observables. Thus, flavour physics will also help in discriminating between the various models that will be proposed by then. In view of this, it is important to calculate the rate of the rare $B$ decays, with theoretical uncertainties as reduced as possible and general enough for generic supersymmetric models. The rare decay $\bar B \rightarrow X_s \gamma$ has already carved out large regions in the space of free parameters of most of the supersymmetric models. Once more precise data from the Super $B$ factories are available, this decay will undoubtedly gain even more efficiency in selecting the viable regions of the parameter space in the various classes of models. Constraints based on nontrivial QCD calculations within various supersymmetric extensions are heavily analyzed in the literature, see for example the Refs.\cite{Bertolini:1990if, Ciuchini:1998xy,Degrassi:2000qf, Carena:2000uj, Borzumati:1999qt, Besmer:2001cj, Ciuchini:2002uv, Ciuchini:2003rg, Okumura:2003hy, Degrassi:2006eh, Ciuchini:2007ha, Altmannshofer:2008vr, Crivellin:2009ar,Crivellin:2011jt} . Finally, model-independent analyses in the effective field theory approach without~\cite{Ali:2002jg} and with the assumption of minimal flavour violation~\cite{Hurth:2008jc} also show the strong constraining power of the $\bar B \rightarrow X_s \gamma$ branching fraction. \section{NLL calculations in supersymmetry} While in the SM, the rate for $\bar B \to X_s \gamma$ is known up to NNLL in QCD, also within supersymmetric theories higher order calculations have been pushed forward in recent years. At the LL level there are several contributions to the decay amplitude: besides the contributions solely induced by flavour mixing in the quark sector with a $W$ boson or a charged Higgs boson and a top quark in the loop, there is also a chargino contribution with an up-type squark which can be induced by the CKM matrix. If we consider also generic new sources of flavour violation induced by a disalignement of quarks and squarks, there are additional contributions from a chargino, gluino and also neutralino. The first complete analysis of the decay rate of $\bar B \rightarrow X_s \gamma$ has been presented in Ref.~\cite{Bertolini:1990if}. It is highly desirable to analyse these non-standard contributions with NLL precision: Besides the large uncertainties in the LL predictions, the step from the LL to the NLL precision is also necessary in order to check the validity of the perturbative approach in the model under consideration. Moreover, it was already shown in specific NP scenarios that bounds on the parameter space of non-standard models are very sensitive to NLL contributions. \subsection{NLL calculation in MFV} The MFV hypothesis is a formal model-independent solution to the NP flavour problem. It assumes that the flavour and the CP symmetry are broken as in the SM. Thus, it requires that all flavour- and CP-violating interactions be linked to the known structure of Yukawa couplings. A renormalization-group-invariant definition of MFV based on a symmetry principle is given in~\cite{D'Ambrosio:2002ex}; this is mandatory for a consistent effective field theoretical analysis of NP effects. The MFV hypothesis is an important benchmark. Because any measurement which is inconsistent with the general constraints and relations induced by the MFV hypothesis~\cite{Hurth:2008jc} indicates the existence of new flavour structures. This hypothesis can also be used within the MSSM. It can be implemented by assuming that the squark and quark mass matrices can be simultaneously diagonalized (alignement). In this case there are no flavour-changing interactions induced by the gluino at the tree level. The first NLL calculation of the inclusive decay $\bar B \rightarrow X_s \gamma$ in the MSSM with the MFV hypothesis includes the gluon corrections to the charged Higgs and the chargino contribution~\cite{Ciuchini:1998xy}, see Figure~\ref{Figure1}. In particular the possibility of destructive interference of the chargino and the charged Higgs contribution is studied. The analysis is done under the MFV assumption that the only source of flavour violation at the electroweak scale is that of the SM, encoded in the CKM matrix. Other flavour-changing interactions were suppressed by assuming the gluino being heavy. It is found that, in this specific supersymmetric scenario, bounds on the parameter space are rather sensitive to NLL contributions and they lead to a significant reduction of the stop-chargino mass region, where the supersymmetric contribution has a large destructive interference with the charged-Higgs boson contribution~\cite{Ciuchini:1998xy}. \begin{figure} \begin{center} \epsfig{figure=rev13.eps,width=12cm} \end{center} \caption{Example diagrams for NLL gluonic corrections to the $W$ boson, charged Higgs and chargino contribution.} \label{Figure1} \end{figure} There are also further analyses within the MFV hypothesis which try to include only the potentially large contributions beyond the leading order which are enhanced by large $\tan\beta$ factors or by large logarithms of the form $\ln(M_{\rm Susy}/M_W)$ where the masses of the supersymmmetric particles are assumed to be significantly larger than the $W$-boson mass~\cite{Degrassi:2000qf,Carena:2000uj, D'Ambrosio:2002ex}. A practically complete MFV analysis has been presented in Ref.~\cite{Degrassi:2006eh}. To LL precison this calculation includes the one-loop diagrams containing a $W$ boson and up-type quark, or a charged Higgs boson and an up-type quark, or a chargino and an up-type squark (see Figure~\ref{Figure2}). Neutralino and gluino exchange diagrams are neglected under the MFV assumption. To NLL precision the gluonic two-loop corrections to the SM and charged Higgs loops are included, also two-loop diagrams with a gluino together with a Higgs or W boson, and finally two-loop diagrams with a chargino together with a gluon or a gluino or a quartic squark coupling. As already shown in Ref.\cite{Ciuchini:1998xy}, the two-loop gluonic corrections to the chargino loops are not UV finite: in order to obtain a finite result one has to combine them with the chargino-gluino diagrams. \begin{figure} \begin{center} \epsfig{figure=rev15.eps,height=4.4cm} \epsfig{figure=rev16.eps,height=4.4cm} \end{center} \caption{Example diagrams for NLL gluino corrections to the chargino and W boson contribution.} \label{Figure2} \end{figure} However, a MFV analysis should take into account the fact that the simultaneous diagonalization of the quark and squark mass matrices can be imposed at one scale $\mu_{\rm MFV}$. The renormalization group evolution of the MSSM parameters then leads to a disalignement between the squark and quark mass matrices at scales different from $\mu_{\rm MFV}$~\cite{Degrassi:2006eh}. So if the MFV condition is imposed at a scale much larger than the superparticle mass scale $M_{\rm Susy}$, very large logarithms of $M_{\rm Susy}/\mu_{\rm MFV}$ occur in the Wilson coefficients. Then the soft Susy-breaking mass parameters -- which are assumed to be flavour-diagonal at the scale $\mu_{\rm MFV}$ -- must be evolved down to $M_{\rm Susy}$ with the help of the appropriate renormalization group equations (RGE), thus, generating some flavour violation in the squark mass matrices which gets absorbed in the couplings of the squark mass eigenstates with the gluinos and charginos. In Ref.~\cite{Degrassi:2006eh}, it is argued, that the effects of the RGE-induced flavour mixing is relatively small and, therefore, are only included to LL order, in the one-loop diagrams with gluinos and down-type squarks and in the one-loop diagrams with charginos and up-type squarks. There is a public computer code for this MFV calculation available which includes all contributions discussed above~\cite{Degrassi:2007kj}. \subsection{NLL calculation in general MSSM} Beyond minimal flavour violation, the most important role is played by the non-diagonal gluino-quark-squark vertex due to the large strong coupling which comes with this vertex. As discussed above, this flavour non-diagonal vertex is induced by squark-mixing to the extent as it is misaligned with quark mixing. It represents a new flavour structure beyond the SM Yukawa couplings. To understand these new sources of flavour violation that may be present in supersymmetric models in addition to those enclosed in the CKM matrix, one has to consider the contributions to the squark mass matrices \begin{equation} {\cal M}_f^2 \equiv \left( \begin{array}{cc} m^2_{\,f,\,LL} +F_{f\,LL} +D_{f\,LL} & \left(m_{\,f,\,LR}^2\right) + F_{f\,LR} \\[1.01ex] \left(m_{\,f,\,LR}^{2}\right)^{\dagger} + F_{f\,RL} & \ \ m^2_{\,f,\,RR} + F_{f\,RR} +D_{f\,RR} \end{array} \right) \,, \label{squarku} \end{equation} where $f$ stands for up- or down-type squarks. In the super-CKM basis, where the quark mass matrices are diagonal and the squarks are rotated in parallel to their superpartners, the $F$ terms from the superpotential and the $D$ terms from the gauge sector turn out to be diagonal $3 \times 3$ submatrices of the $6 \times 6$ mass matrices ${\cal M}^2_f$. This is in general not true for the additional terms $m^2_{f, XY}$ with $X,Y \in \{L,R\}$, originating from the soft supersymmetric breaking potential. Because the squark-quark-gluino coupling is flavour-diagonal in the super-CKM basis, the gluino vertex in the mass eigenstate basis is non-diagonal in flavour-space due to the off-diagonal elements of the soft terms $m^2_{f,LL}$, $m^2_{f,RR}$, $m^2_{f,RL}$. A complete LL analysis of the corresponding gluino contribution to the inclusive decay rate of $\bar B \rightarrow X_s \gamma$ has been presented in Ref.~\cite{Borzumati:1999qt}. The sensitivity of the bounds on the down squark mass matrix to radiative QCD LL corrections is systematically analysed, including the SM and the gluino contributions. In Ref.~\cite{Besmer:2001cj} the interplay between the various sources of flavour violation and the interference effects of SM, gluino, chargino, neutralino and charged Higgs boson contributions is studied. The bounds on simple combinations of elements of the soft part of the squark mass matrices are found to be, in general, one order of magnitude weaker than the bound on the single off-diagonal element, which was derived in previous work by neglecting any kind of interference effects. Some effects beyond LL precision like large $\tan \beta$ effects are estimated in Ref.~\cite{Okumura:2003hy} in analogy to the MFV analyses of Refs.~\cite{Degrassi:2000qf,Carena:2000uj}. Recently, the complete NLL corrections to the Wilson coefficients (at the matching scale $\mu_W$) of the various versions of magnetic and chromomagnetic operators which are induced by a squark-gluino loop have been calculated~\cite{Greub:2011ji}. In this analysis all the appearing heavy particles (which are the gluino, the squarks and the top quark) are simultaneously integrated out at the high scale. There are two classes of two-loop diagrams which have to be considered: diagrams with one gluino and a virtual gluon and diagrams with two gluinos (see Figure~\ref{Figure3}) or with one gluino and a squark-loop. The former have been presented already in Ref.~\cite{Bobeth:1999ww} and now confirmed, while the latter have been calculated for the first time~\cite{Greub:2011ji}. \begin{figure} \begin{center} \epsfig{figure=rev17.eps,height=4.4cm} \end{center} \caption{Example diagrams for NLL gluino gluino contribution.} \label{Figure3} \end{figure} Besides these NLL contributions due to the gluino vertex, there are of course more NLL corrections with non-minimal flavour violation; they involve electroweak (gaugino and higgsino) vertices. However, such contributions are in general suppressed compared to the ones related to the gluino. There are two types of such contributions at the NLL level: First, there are electroweak corrections to the non-minimal LL gluino contribution (in which the electroweak vertex is flavour-diagonal or MFV-like) which are naturally suppressed due to the smaller coupling constants and due to the CKM hierarchy. Second, there is also non-minimal flavour violation via squark-mixing in the electroweak vertices possible. But such contributions are already suppressed at the LL level compared to the gluino contribution due to the smaller coupling constant, apart from the chargino contributions in specific parts of the parameter space in which for example the trilinear coupling $A^u_{23}$ is very large. These features do not change of course when gluon- and gluino-induced NLL corrections are added to such LL contributions. Still, the leading chirally enhanced corrections can be easily calculated by inserting the effective Feynman rules of \cite{Crivellin:2011jt} into the results of \cite{Besmer:2001cj}. Summing up, the complete NLL corrections induced by the gluino vertex given in Ref.~\cite{Greub:2011ji} represent the dominant contribution beyond MFV at this order in most parts of the MSSM parameter space. They are complementary to the MFV contributions at the NLL level which are given in Ref.~\cite{Degrassi:2006eh}. The results are presented also in public computer code~\cite{Greub:2011ji}. \section*{Acknowledgments} TH thanks the organizers of the conference for the interesting and valuable meeting and the CERN theory group for its hospitality during his regular visits to CERN where part of this work was written. \\
\section{Introduction} In 1967 Aumann and Maschler presented their celebrated model for games with incomplete information, see \cite{AuMaS} and references therein. The game they consider consists in a set of, say $I$, standard discrete time two person zero-sum games. At the beginning one of these zero-sum games is picked at random according to a probability $p$. The information which game was picked is transmitted to Player 1 only, while Player 2 just knows $p$. It is assumed that both players observe the actions of the other one, so Player 2 might infer from the actions of his opponent which game is actually played. It turns out that it is optimal for the informed player to play with an additional randomness. Namely in a such a way, that he optimally manipulates the beliefs of the uninformed player The extension to two-player zero-sum stochastic differential games has recently been given by Cardalia- guet and Rainer in \cite{CaRa2}, \cite{Ca}, where the value function is characterized by the unique viscosity solution of a Hamilton Jacobi Isaacs (HJI) equation with an obstacle in the form of a convexity constraint in $p$. The HJI equation without obstacle is the one which is also found to characterize stochastic differential games in the classical work of Fleming and Souganidis \cite{FS}. The probability $p$ appears as an additional parameter in which the value function has to be convex. In Cardaliaguet \cite{Carda} an approximation scheme for the value function of deterministic differential games with incomplete information is introduced. An extension of \cite{Carda} to deterministic games with information incompleteness on both sides is given in the work of Souquiere \cite{Sou}. We consider the case where the underlying dynamic is given by a diffusion with controlled drift but uncontrolled non-degenerate volatility. In constrast to \cite{Carda} and \cite{Sou} we can work on the problem under a Girsanov transform. This transform is a well known tool to consider stochastic games with complete information in the context of backward stochastic differential equations (BSDEs) (see Hamad\`ene and Lepeltier \cite{HaLe}). An approximation of the value function of a stochastic differential game via BSDEs has been discussed in Bally \cite{Ba}. Different to \cite{Ba} our algorithm is closely related to the work of Barles and Souganidis \cite{BaSu} who consider monotone approximation schemes for fully nonlinear second order partial differential equations. The latter was also applied in the recent work of Fahim, Touzi and Warin \cite{FTW} where fully nonlinear parabolic PDEs are treated. As in \cite{FTW} we use a kind of finite difference scheme for the HIJ backwards in time and combine it with taking the convex hull in $p$ at each time step to capture the effect of the information incompleteness. Note that this rather direct ansatz using a probabilistic PDE scheme also significantly differs from the Makov chain approximation method for stochastic differential games described in Kushner \cite{Ku}. From the very beginning of the investigation of BSDEs initiated by Peng in \cite{P} the close relationship with optimal control problems and quasilinear PDEs has been exploited. Consequently, also the approximation of solutions to BSDEs and to quasilinear PDEs are closely related. For a survey on BSDEs we refer to El Karoui, Peng and Quenez \cite{ElK}, while a survey on the numerical approximation of BSDEs can be found in Bouchard, Elie and Touzi \cite{BuETou}. In this sense our result can also be interpreted as approximation of the solutions to the BSDEs which appear in the BSDE representation of the value function for stochastic differential games with incomplete information in \cite{CG}. The outline of the paper is as follows. In section 2 we describe the game and restate the results of \cite{CaRa2} and \cite{Ca} which build the basis for our investigation. In section 3 we present the approximation scheme and give some regularity proofs. Section 4 is devoted to the convergence proof. \section{Setup} \subsection{Formal description of the game} Let $\mathcal{C}([t_0,T];\mathbb{R}^d)$ be the set of continuous functions from $\mathbb{R}$ to $\mathbb{R}^d$, which are constant on $(-\infty,t_0]$ and on $[T,+\infty)$. We denote by $B_s(\omega_B)=\omega_B(s)$ the coordinate mapping on $\mathcal{C}([t_0,T];\mathbb{R}^d)$ and define $\mathcal{H}=(\mathcal{H}_s)$ as the filtration generated by $s \mapsto B_s$. We denote $\Omega_t=\{\omega\in\mathcal{C}([t,T];\mathbb{R}^d)\}$ and $\mathcal{H}_{t,s}$ the $\sigma$-algebra generated by paths up to time $s$ in $\Omega_t$. Furthermore we provide $\mathcal{C}([t_0,T];\mathbb{R}^d)$ with the Wiener measure $\mathbb{P}^0$ on $(\mathcal{H}_s)$.\\ In the following we investigate a two-player zero-sum differential game starting at a time $t\geq t_0$ with terminal time $T$. For any fixed initial data $t\in[t_0,T], x\in\mathbb{R}^d$ the two players control a diffusion on $(\mathcal{C}([t,T];\mathbb{R}^d),(\mathcal{H}_{t,s})_{s\in[t,T]}, \mathcal{H},\mathbb{P}^0)$ given by \begin{eqnarray} dX^{t,x,u,v}_s=b(s,X^{t,x,u,v}_s,u_s,v_s)ds+\sigma(s,X^{t,x,u,v}_s)dB_s\ \ \ \ X^{t,x}_{t}=x. \end{eqnarray} where we assume that the controls of the players $u$, $v$ can only take their values in some compact subsets of some finite dimensional spaces, denoted by $U$, $V$ respectively.\\ The aim of the game is to optimize \begin{itemize} \item[(i)] running costs: $(l_i)_{i\in\{1,\ldots , I\}}:[t_0,T]\times\mathbb{R}^d\times U \times V \rightarrow \mathbb{R}$ \item[(ii)] terminal payoffs: $(g_i)_{i\in\{1,\ldots , I\}}:\mathbb{R}^d\rightarrow\mathbb{R}$, \end{itemize} which are chosen according to a probability $p\in \Delta(I)$ before the game starts. At the beginning of the game this information is transmitted only to Player 1. We assume that Player 1 chooses his control to minimize, Player 2 chooses his control to maximize the expected payoff. Furthermore we assume both players observe their opponents control. So Player 2, knowing only the probability $p_{i}$ for scenario $i\in\{1,\ldots , I\}$ at the beginning, will try to guess the missing information from the behavior of his opponent.\\ The following will be the standing assumption throughout the paper.\\ {\bf Assumption (A)} \begin{itemize} \item[(i)] $b:[t_0,T]\times\mathbb{R}^d\times U \times V \rightarrow \mathbb{R}^d$ is bounded and continuous in all its variables and Lipschitz continuous with respect to $(t,x)$ uniformly in $(u,v)$. \item[(ii)] For $1\leq k,l\leq d$ the function $\sigma_{k,l}:[t_0,T]\times\mathbb{R}^d \rightarrow \mathbb{R}$ is bounded and Lipschitz continuous with respect to $(t,x)$. For any $(t,x)\in[0,T]\times\mathbb{R}^d$ the matrix $\sigma^*(t,x)$ is non-singular and $(\sigma^*)^{-1}(t,x)$ is bounded and Lipschitz continuous with respect to $(t,x)$. \item[(iii)] $(l_i)_{i\in I}:[t_0,T]\times\mathbb{R}^d\times U \times V \rightarrow \mathbb{R}$ is bounded and continuous in all its variables and Lipschitz continuous with respect to $(t,x)$ uniformly in $(u,v)$. $(g_i)_{i\in I}:\mathbb{R}^d \rightarrow \mathbb{R}$ is bounded and uniformly Lipschitz continuous. \item[(iv)] Isaacs condition: for all $(t,x,\xi,p)\in[t_0,T]\times\mathbb{R}^d\times\mathbb{R}^d\times\Delta(I)$ \begin{equation} \begin{array}{rcl} &&\inf_{u\in U}\sup_{v\in V} \left\{\langle b(t,x,u,v),\xi\rangle+\sum_{i=1}^{I}p_il_i(t,x,u,v)\right\}\\ \ \\ &&\ \ \ \ =\sup_{v\in V} \inf_{u\in U}\left\{\langle b(t,x,u,v),\xi\rangle+\sum_{i=1}^{I}p_il_i(t,x,u,v)\right\}=:H(t,x,\xi,p). \end{array} \end{equation} \end{itemize} By assumption (A) the Hamiltonian $H$ is Lipschitz continuous in $(\xi,p)$ uniformly in $(t,x)$ and Lipschitz continuous in $(t,x)$ with Lipschitz constant $c(1+|\xi|)$, i.e. it holds for all $t,t'\in[0,T]$, $x,x'\in\mathbb{R}^d$, $\xi,\xi'\in\mathbb{R}^d$, $p,p'\in\Delta(I)$ \begin{eqnarray} |H(t,x,\xi,p)|\leq c (1+|\xi|) \end{eqnarray} and \begin{eqnarray} |H(t,x,\xi,p)-H(t',x',\xi',p')|\leq c (1+|\xi|)(|x-x'|+|t-t'|)+c|\xi-\xi'|+ c |p-p'|. \end{eqnarray} \subsection{Strategies and value function} We now give the necessary definitions and the results of \cite{Ca} and \cite{CaRa2} on which we will base our investigation. \begin{defi} For any $t\in[t_0,T[$ an admissible control $u=(u_s)_{s\in[t,T]}$ for Player 1 is a progressively measurable process with respect to the filtration $(\mathcal{H}_{t,s})_{s\in[t,T]}$ with values in $U$. The set of admissible controls for Player 1 is denoted by $\mathcal{U}(t)$.\\ The definition for admissible controls $v=(v_s)_{s\in[t,T]}$ for Player 2 is similar. The set of admissible controls for Player 2 is denoted by $\mathcal{V}(t)$. \end{defi} \begin{defi} A strategy for Player 1 at time $t\in[t_0,T[$ is a map $\alpha:[t,T]\times\mathcal{C}([t,T];\mathbb{R}^d)\times L^0([t,T];V)\rightarrow U$ which is nonanticipative with delay, i.e. there is $\delta>0$ such that for all $s\in[t,T]$ for any $f,f'\in\mathcal{C}([t,T];\mathbb{R}^d)$ and $g,g'\in L^0([t,T];V)$ it holds: $f=f'$ and $g=g'$ a.e. on $[t,s]$ $\Rightarrow$ $\alpha(\cdot,f,g)=\alpha(\cdot,f',g')$ a.e. on $[t,s+\delta]$. The set of strategies for Player 1 is denoted by $\mathcal{A}(t)$.\\ The definition of strategies $\beta: [t,T]\times\mathcal{C}([t,T];\mathbb{R}^d)\times L^0([t,T];U)\rightarrow V$ for Player 2 is similar. The set of strategies for Player 2 is denoted by $\mathcal{B}(t)$. \end{defi} With Definition 2.2. it is possible to prove via a fixed point argument the following Lemma, which is a slight modification of Lemma 5.1. in \cite{CaRa2}. \begin{lem} To each pair of strategies $(\alpha,\beta)\in\mathcal{A}(t)\times\mathcal{B}(t)$ one can associate a unique couple of admissible controls $(u,v)\in\mathcal{U}(t)\times\mathcal{V}(t)$, such that for all $\omega\in\mathcal{C}([t,T];\mathbb{R}^d)$ \[\alpha(s,\omega,v(\omega))=u_s(\omega)\ \ \ \ \textnormal{ \textit{and}}\ \ \ \ \beta(s,\omega,u(\omega))=v_s(\omega)\ .\] \end{lem} A characteristic feature of games with incomplete or asymmetric information is that the players have to find a balance between acting optimally according to their information and hiding it. To this end it turns out that he will give his behavior a certain additional randomness. This effect is captured in the following definition. \begin{defi} A random strategy for Player 1 at time $t\in[t_0,T[$ is a a pair $((\Omega_\alpha,\mathcal{G}_\alpha,\mathbb{P}_\alpha),\alpha)$, where $(\Omega_\alpha,\mathcal{G}_\alpha,\mathbb{P}_\alpha)$ is a probability space in $\mathcal{I}$ and $\alpha: [t,T]\times\Omega_\alpha\times\mathcal{C}([t,T];\mathbb{R}^d)\times L^0([t,T]; V)\rightarrow U$ satisfies \begin{itemize} \item[(i)] $\alpha$ is a measurable function, where $\Omega_\alpha$ is equipped with the $\sigma$-field $\mathcal{G}_\alpha$, \item[(ii)] there exists $\delta>0$ such that for all $s\in[t,T]$ and for any $f,f'\in\mathcal{C}([t,T];\mathbb{R}^d)$ and $g,g'\in L^0([t,T];V))$ it holds: \center{$f=f'$ and $g=g'$ a.e. on $[t,s]$ $\Rightarrow$ $\alpha(\cdot,f,g)=\alpha(\cdot,f',g')$ a.e. on $[t,s+\delta]$ for any $\omega\in\Omega_\alpha$.} \end{itemize} The set of random strategies for Player 1 is denoted by $\mathcal{A}^r(t)$.\\ The definition of random strategies $((\Omega_\beta,\mathcal{G}_\beta,\mathbb{P}_\beta),\beta)$, where $\beta: [t,T]\times \Omega_\beta \times\mathcal{C}([t,T];\mathbb{R}^d)\times L^0([t,T];U)\rightarrow V$ for Player 2 is similar. The set of random strategies for Player 2 is denoted by $\mathcal{B}^r(t)$. \end{defi} \begin{rem} Again one can associate to each couple of random strategies $(\alpha,\beta)\in\mathcal{A}^r(t)\times\mathcal{B}^r(t)$ for any $(\omega_\alpha,\omega_\beta)\in\Omega_\alpha\times\Omega_\beta$ a unique couple of admissible strategies $(u^{\omega_\alpha,\omega_\beta},v^{\omega_\alpha,\omega_\beta})\in\mathcal{U}(t)\times\mathcal{V}(t)$, such that for all $\omega\in\mathcal{C}([t,T];\mathbb{R}^d)$, $s\in[t,T]$ \begin{equation*} \alpha(s,\omega_\alpha,\omega,v^{\omega_\alpha,\omega_\beta}(\omega))=u^{\omega_\alpha,\omega_\beta}_s(\omega)\ \ \ \ \textnormal{ and }\ \ \ \ \beta(s,\omega_\beta,\omega,u^{\omega_\alpha,\omega_\beta}(\omega))=v^{\omega_\alpha,\omega_\beta}_s(\omega)\ . \end{equation*} Furthermore $(\omega_\alpha,\omega_\beta)\rightarrow (u^{\omega_\alpha,\omega_\beta},v^{\omega_\alpha,\omega_\beta})$ is a measurable map, from $\Omega_\alpha\times\Omega_\beta$ equipped with the $\sigma$-field $\mathcal{G}_\alpha\otimes\mathcal{G}_\beta$ to $\mathcal{V}(t)\times\mathcal{U}(t)$ equipped with the Borel $\sigma$-field associated to the $L^1$-distance. \end{rem} For any $(t,x,p)\in[t_0,T[\times\mathbb{R}^d\times\Delta(I)$, $\bar \alpha \in (\mathcal{A}^r(t))^I$, $\beta\in\mathcal{B}^r(t)$ we set \begin{eqnarray} J(t,x,p,\bar \alpha,\beta)=\sum_{i=1}^I p_i \ \mathbb{E}_{\bar \alpha_i,\beta}\left[\int_0^Tl_i(s,X_s^{t,x,\bar \alpha_i,\beta}, (\bar\alpha_i)_s,\beta_s)ds+g_i(X_T^{t,x, \bar\alpha_i,\beta})\right], \end{eqnarray} where as in Remark 2.5. we associate to ${\bar \alpha_i},\beta$ for any $(\omega_{\bar \alpha_i},\omega_\beta)\in\Omega_{\bar \alpha_i}\times\Omega_\beta$ the couple of controls $(u^{\omega_{\bar \alpha_i},\omega_\beta},v^{\omega_{\bar \alpha_i},\omega_\beta})$. The process $X^{t,x,\bar \alpha_i,\beta}$ is then defined for any $(\omega_{\bar \alpha_i},\omega_\beta)$ as solution to the SDE (1) with the associated controls. Furthermore $\mathbb{E}_{\bar \alpha_i,\beta}$ is the expectation on $\Omega_{\bar \alpha_i}\times\Omega_\beta\times\mathcal{C}([t,T];\mathbb{R}^d)$ with respect to the probability $\mathbb{P}_{\bar \alpha_i}\otimes\mathbb{P}_\beta\otimes\mathbb{P}^0$, where $\mathbb{P}^0$ denotes the Wiener measure on $\mathcal{C}([t,T];\mathbb{R}^d).$\\ Under assumption (A) the existence of the value of the game and its characterization as a viscosity solution to an obstacle problem is shown in \cite{Ca},\cite{CaRa2}. \begin{thm} For any $(t,x,p)\in[t_0,T[\times\mathbb{R}^d\times\Delta(I)$ the value of the game with incomplete information $V(t,x,p)$ is given by \begin{equation} \begin{array}{rcl} V(t,x,p) &=& \inf_{\bar \alpha \in (\mathcal{A}^r(t))^I}\sup_{\beta\in \mathcal{B}^r(t)} J(t,x,p,\bar \alpha,\beta)\\ \ \\ &=&\sup_{\beta\in \mathcal{B}^r(t)} \inf_{\bar \alpha \in (\mathcal{A}^r(t))^I} J(t,x,p,\bar \alpha,\beta). \end{array} \end{equation} Furthermore the function $V:[0,T[\times \mathbb{R}^d\times\Delta(I)\rightarrow\mathbb{R}$ is the unique viscosity solution to \begin{equation} \min \left\{ \frac{\partial w} {\partial t}+\frac{1}{2}\textnormal{tr}(\sigma\sigma^*(t,x)D_x^2w)+H(t,x,D_xw,p),\lambda_{\min}\left(p,\frac{\partial ^2 w}{\partial p^2}\right)\right\}=0 \end{equation} with terminal condition $w(T,x,p)=\sum_{i}p_ig_i(x)$, where for all $p\in\Delta(I)$, $A\in\mathcal{S}^I$ \begin{eqnarray} \lambda_{\min}(p,A):=\min_{z\in T_{\Delta(I)(p)}\setminus\{0\}} \frac{\langle Az,z\rangle}{|z|^2}. \end{eqnarray} and $T_{\Delta(I)(p)}$ denotes the tangent cone to $\Delta(I)$ at $p$, i.e. $T_{\Delta(I)(p)}=\overline{\cup_{\lambda>0}(\Delta(I)-p)/\lambda}$ . \end{thm} \begin{rem} Unlike the standard definition of viscosity solutions (see e.g. \cite{CIL}) the subsolution property to (7) is required only on the interior of $\Delta(I)$ while the supersolution property to (7) is required on the whole domain $\Delta(I)$ (see \cite{Ca} and \cite{CaRa2}). This is due to the fact that we actually consider viscosity solutions with a state constraint, namely $p\in\Delta(I)\subsetneq\mathbb{R}^I$. For more details we refer to \cite{CaDoL}. \end{rem} \section{Approximation of the value function} \subsection{Numerical scheme} Our approximation scheme of the value function basically amounts to approximate the solution of the obstacle problem (7). In order to do so it is convenient to consider the real dynamics of the game (1) under a Girsanov transform. This technique - first applied to stochastic differential games by \cite{HaLe} - enables us to decouple the forward dynamics (1) from the controls of the players. As in \cite{Ba} where this transformation is applied in the context of numerical approximation for stochastic differential games via BSDE we will use the following approximation for the forward dynamics .\\ For $L\in\mathbb{N}$ we define a partition of $[t_0,T]$ with stepsize $\tau=\frac{T}{L}$ by $\Pi^\tau=\{t_0,t_1,\ldots ,t_L=T\}$. Then for all $k=0,\ldots ,L$, $x\in\mathbb{R}^d$, $p\in\Delta(I)$ let $(X^{t_k,x}_s)_{s\in[t_k,T]}$ denote the diffusion \begin{eqnarray} X^{t_k,x}_s=x+\int_{t_k}^s \sigma(r,X_r^{t,x})dB_r. \end{eqnarray} Furthermore we define the discrete process $(\bar X^{k,x}_{n})_{n=k,\ldots , L}$ as the standard Euler scheme approximation for (9) on $\Pi^\tau$ \begin{eqnarray} \bar X^{k,x}_{n}=x+\sum_{j=k}^{n-1} \sigma(t_{j},\bar X^{k,x}_{j})\Delta B^{j}, \end{eqnarray} where $\Delta B^{j}= B_{t_{j+1}}- B_{t_{j}}$.\\ We will approximate the value function (6) backwards in time. To do so we set for all $x\in\mathbb{R}^d$, $p\in\Delta(I)$ \begin{equation} V^\tau(t_L,x,p)=\langle p, g(x)\rangle \end{equation} and we define recursively for $k=L-1,\ldots,0$ \begin{equation} \begin{array}{rcl} V^\tau(t_{k-1},x,p)&=&\textnormal{Vex}_p\left( \mathbb{E}\left[V^\tau(t_{k},\bar X^{{k-1},x}_{k},p)\right]+\tau H(t_{k-1},x,\bar z_{k-1}(x,p),p)\right), \end{array} \end{equation} where $\bar z_{k-1}(x,p)$ is given by \begin{eqnarray} \bar z_{k-1}(x,p)=\frac{1}{\tau}\mathbb{E}\left[V^\tau(t_{k},\bar X^{{k-1},x}_{k},p) (\sigma^*)^{-1}(t_{k-1},x)\Delta B^{k-1}\right] \end{eqnarray} and $\textnormal{Vex}_p$ denotes the convex hull, i.e. the largest function that is convex in the variable $p$ and does not exceed the given function. \subsection{Some regularity properties} \subsubsection{Monotonicity} First we show that our scheme fulfills a monotonicity condition which corresponds to the one in \cite{BaSu} (2.2). It is well known that this criteria is crucial for the convergence of general finite difference schemes.\\ \begin{lem} Let $\phi:\mathbb{R}^d\rightarrow\mathbb{R}$ be a uniformly Lipschitz continuous function with Lipschitz constant $M$. Then there exists for all $x,x'\in \mathbb{R}^d$ a $\theta\in\mathbb{R}^d$ with $|\theta|\leq M$ \begin{equation*} \phi(x)-\phi(x')=\langle \theta, x-x'\rangle \end{equation*} \end{lem} \begin{proof} For $\phi\in C^1$ the result follows from partial integration with $\theta=\int_0^1D_x\phi(x+r(x'-x))dr$. For the case of general Lipschitz continuous function $\phi$ one chooses a sequence of $\mathcal{C}^1$ functions $(\phi^\epsilon)_{\epsilon>0}$ which converges uniformly to $\phi$. Since $\phi$ is uniformly Lipschitz continuous, we may assume that the absolute value of $D_x\phi^\epsilon$ and hence the corresponding $\theta^\epsilon$ are uniformly bounded by the constant $M$. Consequently, possibly passing though a subsequence, there exists a $\theta\in\mathbb{R}^d$ with $|\theta|\leq M$ such that the lemma holds. \end{proof} With the help of Lemma 3.1 we now establish: \begin{lem} Let $k\in\{0,\ldots,L-1\}$ and $\phi,\psi: \mathbb{R}^d\rightarrow\mathbb{R}$ be two Lipschitz continuous functions. Then for any $x\in\mathbb{R}$, $p\in\Delta(I)$ \begin{eqnarray*} &&\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})\right]+\tau H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k, x}_{{k+1}}) (\sigma^*)^{-1}(t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \geq \mathbb{E}\left[\psi(\bar X^{k,x}_{{k+1}})\right]+\tau H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \psi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}(t_k,x) \Delta B^{k}\right],p)-\tau \mathcal{O}(\tau), \end{eqnarray*} where $\mathcal{O}(\tau)$ is independent of $p$. \end{lem} \begin{proof} By (4) $H$ is uniformly Lipschitz continuous in $\xi$. So by Lemma 3.1. there exists a $\theta\in\mathbb{R}^d$ with $|\theta|\leq M$, where $M$ denotes the Lipschitz constant of $H$, such that \begin{eqnarray*} &&\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})\right] + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ - H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \psi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p))\bigg)\\ &&=\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})\right]+ \left \langle \tau \theta,\left(\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right]-\frac{1}{\tau} \mathbb{E}\left[ \psi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right]\right)\right\rangle\\ &&=\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})\right]+\left\langle \theta , \mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})(\sigma^*)^{-1}(t_k,x)\Delta B^k\right]\right\rangle\\ &&=\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})\left(1+\langle \theta ,(\sigma^*)^{-1}(t_k,x) \Delta B^k\rangle \right)\right]. \end{eqnarray*} Since $0\leq \phi(x)-\psi(x) \leq c$ for any $x\in\mathbb{R}$, we have \begin{eqnarray*} &&\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{{k+1}})\left(1+\langle \theta, (\sigma^*)^{-1}(t_k,x) \Delta B^k\rangle\right) \right]\\ &\geq&\mathbb{E}\left[(\phi-\psi)(\bar X^{k,x}_{t_{k+1}}) 1_{|\Delta B^k|\geq\|\theta\sigma^{-1}\|_\infty^{-1}}\langle \theta, (\sigma^*)^{-1}(t_k,x) \Delta B^k\rangle \right]\\ &\geq& - C \mathbb{E}\left[1_{|\Delta B^k|\geq {\frac{1}{C}}} |\Delta B^k|\right] \end{eqnarray*} with $C:=\|M\sigma^{-1}\|_\infty$ independent of $(t_k,x,p)$ and $\tau$. Furthermore we can explicitely calculate \begin{eqnarray*} \mathbb{E}\left[1_{|\Delta B^k|\geq \frac{1}{C}} |\Delta B^k|\right]= \ \frac{1}{(2\pi)^\frac{d}{2}(\tau)^\frac{1}{2}}\int_{|x|\geq\frac{1}{C}}^\infty |x| e^{-\frac{x^2}{2\tau}}dx =\frac{1}{2^{\frac{d}{2}-1}\Gamma(\frac{d}{2})}\tau^\frac{1}{2} e^{-\frac{1}{2 C^2 \tau}}, \end{eqnarray*} where $\Gamma$ denotes the gamma function. \end{proof} \subsubsection{Lipschitz continuity in $x$} To show that the Lipschitz continuity in $x$ is preserved under the scheme, we establish the following Lemma. \begin{lem} Let $k\in\{0,\ldots,L-1\}$ and $\phi:\mathbb{R}^d\rightarrow\mathbb{R}$ be a uniformly Lipschitz continuous function with Lipschitz constant $M$. Then for any $k\in\{0,\ldots,L-1\}$, $x,x'\in\mathbb{R}$, $p\in\Delta(I)$ \begin{eqnarray*} &&\bigg|\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})\right]+\tau H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k, x}_{{k+1}}) (\sigma^*)^{-1}(t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ -\mathbb{E}\left[\phi(\bar X^{k,x'}_{{k+1}})\right]-\tau H(t_k,x',\frac{1}{\tau} \mathbb{E}\left[\phi(\bar X^{k, x'}_{{k+1}}) (\sigma^*)^{-1}(t_k,x') \Delta B^{k}\right],p)\bigg|\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \leq C^{M,\tau} |x-x'|, \end{eqnarray*} where $C^{M,\tau}=M(1+c\tau)+c \tau$ with $c$ independent of $p$. \end{lem} \begin{proof} We fix $k\in\{0,\ldots,L-1\}$, $x,x'\in\mathbb{R}$, $p\in\Delta(I)$ and write \begin{equation} \begin{array}{rcl} &&\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}})\right] + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ - H(t_k,x',\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\bigg)\\ &&=\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}})\right]\\ &&\ \ \ + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ - H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\bigg)\\ &&\ \ \ + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ - H(t_k,x',\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\bigg). \end{array} \end{equation} Assume that $\phi\in \mathcal{C}^1$ with $|D_x\phi|\leq M$. First we consider the last term of (14). We have for $\Theta^1:=\int_0^1 D_x\phi(x'+r \sigma(t_{k},x')\Delta B^{k}) dr$ that $|\Theta^1|\leq M$ and \begin{eqnarray*} \left|\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right]\right| & =& \frac{1}{\tau}\left|\mathbb{E}\left[\phi(x') (\sigma^*)^{-1}(t_{k},x')\Delta B^{k}+\Theta^1 |\Delta B^{k}|^2\right]\right|\\ & \leq& M. \end{eqnarray*} Since by (4) the Hamiltonian $H$ is uniformly Lipschitz continuous in $x$ with Lipschitz constant $c(1+|\xi|)$ it holds \begin{eqnarray*} &&\tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)- H(t_k,x',\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\bigg)\\ &&\leq \tau c(1+M) |x-x'|. \end{eqnarray*} For the remaining terms in (14) we note that by (4) the Hamiltonian $H$ is uniformly Lipschitz continuous. So there exists as in Lemma 3.1. a $\theta^1\in\mathbb{R}^d$ with $|\theta^1|\leq c$, such that \begin{eqnarray} &&\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}})\right]\nonumber\\ &&\ \ \ + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)- H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right],p)\bigg)\nonumber\\ &&=\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}})\right] \nonumber \\ &&\ \ \ \ \ \ \ \ \ + \langle \theta^1, \left(\mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}}) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right]-\mathbb{E}\left[ \phi(\bar X^{k,x'}_{{k+1}}) (\sigma^*)^{-1}( t_k,x') \Delta B^{k}\right]\right)\rangle\nonumber\\ &&=\mathbb{E}\left[(\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}}) )(1+ \langle \theta^1, (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)\right]\\ &&\ \ \ \ \ \ \ \ + \mathbb{E}\left[\langle \theta^1, \phi(\bar X^{k,x'}_{{k+1}}) ((\sigma^*)^{-1}( t_k,x)-(\sigma^*)^{-1}( t_k,x')) \Delta B^{k}\rangle\right]\nonumber. \end{eqnarray} For the first term of (15) we have with $\Theta^2:=\int_0^1D_x\phi(\bar X^{k,x}_{{k+1}}+r(\bar X^{k,x'}_{{k+1}}-\bar X^{k,x}_{{k+1}}))dr$ \begin{eqnarray*} &&\mathbb{E}\left[(\phi(\bar X^{k,x}_{{k+1}})-\phi(\bar X^{k,x'}_{{k+1}}) )(1+ \langle \theta^1, (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)\right]\\ &&=\mathbb{E}\left[\left\langle \Theta^2, \bar X^{k,x}_{{k+1}}-\bar X^{k,x'}_{{k+1}} \right\rangle(1+ \langle \theta^1,(\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)\right]\\ &&\leq\mathbb{E}\left[\left\langle \Theta^2,(1+\langle \theta^1,(\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)(x-x')+ (\sigma(t_k,x)-\sigma(t_k,x')) \Delta B^k\right\rangle\right]\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ +c\tau|x-x'|. \end{eqnarray*} We finally use Cauchy-Schwartz (note that in the expansion of the square the $\Delta B^k$ parts vanish when taking expectation), $|\Theta^2|\leq M$ and the Lipschitz contiunity of $\sigma$ to get \begin{eqnarray*} &&\mathbb{E}\left[\left\langle \Theta^2, (1+\langle \theta^1,(\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)(x-x')+ (\sigma(t_k,x)-\sigma(t_k,x')) \Delta B^k\right\rangle\right]\\ &&\leq M \mathbb{E}\left[\left((1+\langle \theta^1,(\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)(x-x')+ (\sigma(t_k,x)-\sigma(t_k,x')) \Delta B^k\right)^2\right]^\frac{1}{2}\\ &&\leq M|x-x'| \left(\mathbb{E}\left[1+c|\Delta B^k|^2\right]\right)^\frac{1}{2}= M|x-x'| (1+c\tau)^\frac{1}{2}\leq M|x-x'| (1+\frac{c}{2}\tau). \end{eqnarray*} For the second term of (15) we use the uniform Lipschitz continuity of $(\sigma^*)^{-1}$ (by assumption (A)) to have with the $\mathbb{R}^d$-valued random variable $\Theta^3:=\int_0^1D_x\phi(\bar X^{k,x'}_{{k+1}}+r(\bar X^{k,x'}_{{k+1}}-x'))dr$ \begin{eqnarray*} &&\mathbb{E}\left[\langle \theta^1, \phi(\bar X^{k,x'}_{{k+1}}) ((\sigma^*)^{-1}( t_k,x)-(\sigma^*)^{-1}( t_k,x')) \Delta B^{k}\rangle\right]\\ &&=\mathbb{E}\left[ \langle \theta^1, \phi(\bar X^{k,x'}_{{k+1}}) ((\sigma^*)^{-1}(t_k,x)-(\sigma^*)^{-1}(t_k,x')) \Delta B^{k}\rangle\right]\\ &&=\mathbb{E}\left[\langle \theta^1, (\phi(x')+\langle \Theta^3, \sigma(t,x') \Delta B^{k}\rangle)((\sigma^*)^{-1}(t_k,x)-(\sigma^*)^{-1}(t_k,x')) \Delta B^{k}\rangle \right]\\ &&\leq c M \tau |x-x'|. \end{eqnarray*} The case of Lipschitz continuous $\phi$ follows by approximation with a sequence of $\mathcal{C}^1$ functions $(\phi^\epsilon)_{\epsilon>0}$ which converges uniformly to $\phi$. Since $\phi$ is uniformly Lipschitz continuous with constant $M$, we may assume that $|D_x\phi^\epsilon|\leq M$ for all $\epsilon>0$. \end{proof} With the previous Lemma it is easy to show the Lipschitz continuity of $V^\tau(t_{\cdot},x,p)$ in $x$. \begin{prop} $V^\tau(t_{\cdot},x,p)$ is uniformly Lipschitz continuous in $x$ with a Lipschitz constant that depends only on the constants of assumption (A). \end{prop} \begin{proof} We will show Proposition 3.5. by induction. With (A) we have that $V^\tau(t_L,x,p)$ is Lipschitz continuous in $x$ with a constant $M_L$ that depends only on the constants of assumption (A). Let $M_k$ be the Lipschitz constant for $V^\tau(t_k,\cdot,p)$ then by (12) and Lemma 3.3. and since Vex is monotonic, we have \begin{eqnarray*} |V^\tau(t_{k-1},x,p)-V^\tau(t_{k-1},x',p)|\leq M_k((1+c\tau)^\frac{1}{2}+c\tau)+c \tau)|x-x'|. \end{eqnarray*} Hence $M_{k-1}:=M_k(1+c\tau)+c \tau$ is a Lipschitz constant for $V^\tau(t_{k-1},\cdot,p)$ and $M:=M_L C e^{{CT}}$ for a $C$ independent of $\tau,x,p$ is a constant dominating the recursively defined Lipschitz constants $(M_k)_{k=0,\ldots,L}$. \end{proof} \ \\ \ \\ With the uniform Lipschitz continuity of $V^\tau$ in $x$ it follows that the value function is uniformly bounded. \begin{prop} $V^\tau(t_{\cdot},x,p)$ is uniformly bounded by a constant only depending on the constants of assumption (A). \end{prop} \begin{proof} Fix $k\in\{0,L-1\}$, $x\in\mathbb{R}^d$, $p\in\Delta(I)$. Assume first that $V^\tau$ is at $t_{k+1}$ continuously differentiable in the second variable with $|D_xV^\tau|\leq M$. Then with $\Theta:=\int_0^1 D_xV^\tau(t_{k+1},x+r \sigma(t_{k},x)\Delta B^{k},p) dr$ \begin{equation} \begin{array}{rcl} |\bar z_k(x,p)| &=& \frac{1}{\tau}\left| \mathbb{E}\left[V^\tau(t_{k+1},x+\sigma(t_k,x)\Delta B^k,p) (\sigma^*)^{-1}(t_{k},x)\Delta B^{k}\right]\right|\\ \ \\ & =& \frac{1}{\tau}\left|\mathbb{E}\left[V^\tau(t_{k+1},x,p) (\sigma^*)^{-1}(t_{k},x)\Delta B^{k}+\Theta |\Delta B^{k}|^2\right]\right|\\ \ \\ & \leq& M. \end{array} \end{equation} Since $V^\tau$ is by Lemma 3.3. uniformly Lipschitz continuous in $x$ one has (16) in the general case again by regularization.\\ By (A) $V^\tau(t_L,x,p)$ is bounded by a constant $M_L$ that depends only on the constants of assumption (A). Let $M_k$ be a bound for $|V^\tau(t_k,\cdot,p)|$ then by (3) the definition (12) and (16) we have \begin{eqnarray*} \mathbb{E}\left[V^\tau(t_{k},\bar X^{{k-1},x}_{k},p)\right]+\tau H(t_{k-1},x,\bar z_{k-1}(x,p),p)\leq M_k + c \tau(1+M) \end{eqnarray*} and $M_L + c T (1+M)$ is a constant dominating the recursively defined constants $(M_k)_{k=0,\ldots,L}$. \end{proof} \subsubsection{Lipschitz continuity in $p$} The Lipschitz continuity of $V^\tau(t_{\cdot},x,p)$ in $p$ can be shown with similar methods. \begin{lem} Let $k\in\{0,\ldots,L-1\}$ and $\phi:\mathbb{R}^d\times\Delta(I)\rightarrow\mathbb{R}$ be a uniformly Lipschitz continuous function with Lipschitz constant $M$. Then for any $k\in\{0,\ldots,L-1\}$, $x\in\mathbb{R}^d$, $p,p'\in\Delta(I)$ \begin{eqnarray*} &&\bigg|\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p)\right]+\tau H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k, x}_{{k+1}},p) (\sigma^*)^{-1}(t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ -\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p')\right]-\tau H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[\phi(\bar X^{k, x}_{{k+1}},p') (\sigma^*)^{-1}(t_k,x) \Delta B^{k}\right],p')\bigg|\\ &&\ \ \ \ \ \ \ \ \ \ \ \leq \bar C^{M,\tau} |p-p'|, \end{eqnarray*} where $\bar C^{M,\tau}=M(1+c\tau)+c \tau$. \end{lem} \begin{proof} We fix $k\in\{0,\ldots,L-1\}$, $x\in\mathbb{R}^d$, $p,p'\in\Delta(I)$. First note that by (4) the Hamiltonian is uniformly Lipschitz in $p$. Hence \begin{eqnarray*} &&\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p')\right] + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ - H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p') (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p')\bigg)\\ &&\leq \mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p')\right] + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ - H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p') (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\bigg)+c\tau|p-p'|. \end{eqnarray*} By (4) the Hamiltonian $H$ is uniformly Lipschitz continuous in $\xi$ with a constant $c$. So by Lemma 3.1. \begin{eqnarray*} &&\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p')\right] + \tau \bigg(H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\\ &&\ \ \ \ \ - H(t_k,x,\frac{1}{\tau} \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p') (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right],p)\bigg)\\ &&=\mathbb{E}\left[\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p')\right]\\ &&\ \ \ \ \ +\left \langle \theta, \mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p) (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right]-\mathbb{E}\left[ \phi(\bar X^{k,x}_{{k+1}},p') (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\right]\right\rangle\\ &&=\mathbb{E}\left[(\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p') )(1+\langle \theta, (\sigma^*)^{-1}( t_k,x) \Delta B^{k})\rangle\right]. \end{eqnarray*} Assume for now that $\phi$ is differentiable in $p$ with $|D_p\phi|\leq M$. Then with $\Theta:=\int_0^1 D_p\phi(\bar X^{k,x}_{{k+1}},p+r(p-p')) dr$ we have \begin{eqnarray*} &&\mathbb{E}\left[(\phi(\bar X^{k,x}_{{k+1}},p)-\phi(\bar X^{k,x}_{{k+1}},p') )(1+\langle \theta, (\sigma^*)^{-1}( t_k,x) \Delta B^{k})\rangle\right]\\ &&=\mathbb{E}\left[\langle \Theta,(1+ \langle \theta, (\sigma^*)^{-1}( t_k,x) \Delta B^{k}\rangle)(p-p')\rangle\right]\\ &&\leq M |p-p'| \left(\mathbb{E}\left[1+c|\Delta B^k|^2\right]\right)^\frac{1}{2}= M|p-p'| (1+c\tau)^\frac{1}{2}\leq M|p-p'| (1+\frac{c}{2}\tau), \end{eqnarray*} where for the first estimate in the last line we used again Cauchy Schwartz as in the previous Lemma. The general case follows again by regularization.\\ \end{proof} It is now easy to show the Lipschitz continuity of $V^\tau(t_{\cdot},x,p)$ in $p$ as in Proposition 3.4. \begin{prop} $V^\tau(t_{\cdot},x,p)$ is uniformly Lipschitz continuous in $p$ with a Lipschitz constant only depending on the constants of assumption (A). \end{prop} \subsubsection{H\"older continuity in $t$} Finally we use the Lipschitz continuity of $V^\tau$ in $x$ to establish the H\"older continuity in time. \begin{prop} For all $L\in\mathbb{N}$, $x\in\mathbb{R}^d$, $p\in\Delta(I)$ it holds that $(t_.,x,p)\rightarrow V^\tau(t_{.},x,p)$ is H\"older continuous in $t_.$, in the sense that for all $k\in\{1,\dots,L-1\}, l\in\{1,\ldots L-k\}$, there exists a constant $c$ only depending on the constants of assumption (A), such that \begin{equation*} |V^\tau(t_{k+l},x,p)-V^\tau(t_k,x,p)|\leq c |t_{k+l}-t_{k}|^\frac{1}{2}. \end{equation*} \end{prop} \begin{proof} We fix $(x,p)\in\mathbb{R}^d\times\Delta(I)$. By (12), (3) and the convexity of $V^\tau$ in $p$ we have \begin{eqnarray*} &&|V^\tau(t_{k+l},x,p)-V^\tau(t_k,x,p)|\\ &&\ \ \ \ =\left|V^\tau(t_{k+l},x,p)-\textnormal{Vex}_p\left(\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},p)\right]+\tau H(t_k,x,\bar z_k(x,p),p)\right)\right|\\ &&\ \ \ \ \leq \left|\mathbb{E}\left[V^\tau(t_{k+l},x,p)-V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},p)\right]\right|+c\tau (1+M), \end{eqnarray*} where we used that by (16) $|\bar z_k(x,p)|$ is bounded uniformly in $p\in\Delta(I)$ by the Lipschitz constant of $V^{\tau}$ in $x$. Note that by definition (12) \begin{eqnarray*} V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},p)=\textnormal{Vex}_p\bigg(\mathbb{E}\left[V^{\tau}(t_{k+2},\bar X^ {{k+1},{x'}}_{{k+2}},p)\right]+\tau H(r,x',\bar z_{{k+1}}(x',p),p)\bigg)\bigg|_{x'=\bar X^ {k,x}_{{k+1} }}. \end{eqnarray*} Hence by (A) and the fact that $V^\tau$ is convex in $p$ we have \begin{eqnarray*} &&\left|V^\tau(t_{k+l},x,p)-\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},p)\right]\right|\\ &&\ \ \ \ =\bigg| V^\tau(t_{k+l},x,p)\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ -\mathbb{E}\bigg[\textnormal{Vex}_p\bigg(\mathbb{E}\left[V^{\tau}(t_{k+2},\bar X^ {{k+1},{x'}}_{{k+2}},p)\right]+\tau H(r,x',\bar z_{{k+1}}(x',p),p)\bigg)\bigg|_{x'=\bar X^ {k,x}_{{k+1} }} \bigg]\bigg|\\ &&\ \ \ \ \leq \left|V^\tau(t_{k+l},x,p)-\mathbb{E}\left[V^{\tau}(t_{k+2},\bar X^ {{k+1},\bar X^ {k,x}_{{k+1} }}_{{k+2}},p)\right]\right|+c \tau (1+M)\\ &&\ \ \ \ = \left|V^\tau(t_{k+l},x,p)-\mathbb{E}\left[V^{\tau}(t_{k+2},\bar X^ {k,x}_{{k+2}},p)\right]\right|+c\tau(1+M). \end{eqnarray*} Since $l\tau=|t_{k+l}-t_k|$ repeating this now $l-2$ times gives \begin{eqnarray*} |V^\tau(t_{k+l},x,p)-V^\tau(t_k,x,p)|&\leq& \left|V^\tau(t_{k+l},x,p)-\mathbb{E}\left[V^{\tau}(t_{k+l},\bar X^ {k,x}_{{k+l}},p)\right]\right|+c(1+M)|t_{k+l}-t_k|. \end{eqnarray*} Furthermore by the Lipschitz continutity of $V^\tau$ in $x$ and (A) it holds \begin{eqnarray*} \left|V^\tau(t_{k+l},x,p)-\mathbb{E}\left[V^{\tau}(t_{k+l},\bar X^ {k,x}_{{k+l}},p)\right]\right|\leq M \mathbb{E}\left[|\bar X^ {k,x}_{{k+l}}-x|\right] \leq c |t_{k+l}-t_k|^\frac{1}{2}, \end{eqnarray*} hence \begin{eqnarray*} |V^\tau(t_{k+l},x,p)-V^\tau(t_k,x,p)|\leq M |t_{k+l}-t_k|^\frac{1}{2}+c(1+M)|t_{k+l}-t_k|. \end{eqnarray*} \end{proof} \section{Convergence} \begin{thm} Under (A) we have uniform convergence on the compact subsets of $[0,T]\times\mathbb{R}^d\times\Delta(I)$, i.e. \begin{eqnarray} \lim_{\tau\downarrow0,t_k\rightarrow t,x'\rightarrow x,p'\rightarrow p} V^\tau(t_k,x',p')=V(t,x,p). \end{eqnarray} \end{thm} Note that by Proposition 3.5. the family $(V^\tau,\tau>0)$ is uniformly bounded. Furthermore by Proposition 3.4., 3.7. and 3.8. the family $(V^\tau,\tau>0)$ is equicontinuous, hence by Arzela Ascoli compact for the topology of uniform convergence. Furthermore any candidate for the limit of $V^\tau$ as $\tau\downarrow0$ is as a limit of convex functions convex in $p$.\\ Let $w:[0,T]\times\mathbb{R}^d\times\Delta(I)\rightarrow\mathbb{R}$ be a candidate for the limit. We will show that $w$ is a viscosity solution to (7). Since this property uniquely characterizes the value function $V$ the convergence follows immediately. \subsection{One step a posteriori martingales and DPP} By construction there exists at each time step $t_k$ for any $x\in\mathbb{R}^d$ and $p\in\Delta(I)$ a linear combination of $\pi^{k,1}(x,p),\ldots ,$ $\pi^{k,I}(x,p)\in\Delta(I)$ such that \begin{eqnarray} \sum_{l=1}^I \lambda_l^k(x,p)\pi^{k,l}(x,p)=p\ \ \ \ \ \sum_{l=1}^I \lambda_l^k(x,p)=1 \end{eqnarray} and \begin{equation} \begin{array}{rcl} &&V^\tau(t_{k},x,p)\\ \ \\ &&\ \ \ \ =\sum_{l=1}^I \lambda_l^k(x,p)\left(\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},\pi^{k,l}(x,p))\right]+\tau H(t_k,x,\bar z_k(x,\pi^{k,l}(x,p)),\pi^{k,l}(x,p))\right) \end{array} \end{equation} with \begin{eqnarray} \bar z_k(x,\pi^{k,l}(x,p))=\frac{1}{\tau}\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},\pi^{k,l}(x,p)) (\sigma^*)^{-1}(t_{k},x)\Delta B^{k}\right], \end{eqnarray} where we can choose $(x,p)\rightarrow \lambda^k(x,p)\in\Delta(I)$ and $(x,p)\rightarrow \pi^{k}(x,p)\in\Delta(I)^I$ Borel measurable.\\ \begin{defi} For all $i\in I$, $k=0,\ldots,L$, $x\in\mathbb{R}^n$ and $p\in\Delta(I)$ we define the one step feedbacks $\bold{p}^{i,x,p}_{k+1}$ as $\Delta(I)$-valued random variables which are independent of $\sigma(B_s)_{s\in\mathbb{R}}$, such that \begin{itemize} \item [(i)] for $k=0,\ldots,L-1$ \begin{itemize} \item [(a)] if $p_i=0$ set $\bold{p}^{i,x,p}_{k+1}={p}$ \item [(b)] if ${p}_i>0$: $\bold{p}^{i,x,p}_{k+1}\in\{\pi^{k,1}(x,p),\ldots ,\pi^{k,I}(x,p)\}$ with probability \begin{eqnarray*} &&\mathbb{P}\left[\bold{p}^{i,x,p}_{k+1}=\pi^{k,l}(x,p)|(\bold{p}^{j,x',p'}_{l})_{j\in\{1,\ldots,I\},x'\in\mathbb{R},p'\in\Delta{I},l\in\{1,\ldots,k\}}\right] =\lambda_l^k(x,p)\frac{(\pi^{k,l}(x,p))_i}{p_i} \end{eqnarray*} \end{itemize} \item [(ii)] for $k=L$ set $\bold{p}^{i,x,p}_{L+1}=e^i$. \end{itemize} Furthermore we define one step a posteriori martingales $\bold{p}^{x,p}_{k+1}=\bold{p}^{\bold{i},x,p}_{k+1}$, where the index $\bold{i}$ is a random variable with law $p$, independent of $\sigma(B_s)_{s\in[0,T]}$ and $(\bold{p}^{j,x',p'}_{l})_{j\in\{1,\ldots,I\},x'\in\mathbb{R},p'\in\Delta{I},l\in\{1,\ldots,L\}}$. The martingale property is a direct consequence of the proof of the Lemma given below. \end{defi} The following one step dynamic programming is a direct consequence of Definition 4.2. \begin{lem} For all ${k}=0,\ldots, L-1$, $x\in\mathbb{R}^d$, $p\in\Delta(I)$ we have \begin{equation} \begin{array}{rcl} &&V^\tau(t_{k},x,p)=\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},\bold p^{x,p}_{k+1})+\tau H(t_k,x,\bar z_{k}(x,\bold p^{x,p}_{k+1}),\bold p^{x,p}_{k+1})\right] \end{array} \end{equation} with \begin{eqnarray} \bar z_{k}(x,\bold p^{x,p}_{k+1})=\frac{1}{\tau}\mathbb{E}\left[V^{\tau}(t_{k+1},\bar X^{k,x}_{{k+1}},p) (\sigma^*)^{-1}(t_{k},x)\Delta B^{k}\right]\bigg|_{p=\bold p^{x,p}_{k+1}}. \end{eqnarray} \end{lem} \begin{proof} Assume $(p)_i>0$ for all $i=1,\ldots,I$. By the construction for all suitable functions $f:\Delta(I)\rightarrow\mathbb{R}$ it holds \begin{eqnarray*} &&\mathbb{E}[f(\bold p^{x,p}_{k+1})]=\sum_{i=1}^I\mathbb{E} \left[ 1_{\{\bold i=i\}} f(\bold p^{i,x,p}_{k+1})\right] =\sum_{i=1}^I\mathbb{E} \left[ 1_{\{\bold i=i\}} \right]\mathbb{E} \left[f(\bold p^{i,x,p}_{k+1})\right]\\ &&\ \ \ \ =\sum_{i=1}^I p_i \sum_{l=1}^I\lambda_l^k(x,p)\frac{(\pi^{k,l}(x,p))_i}{p_i} f((\pi^{k,l}(x,p)))\\ &&\ \ \ \ =\sum_{l=1}^I \lambda_l^k(x,p)f(\pi^{k,l}(x,p)) \end{eqnarray*} and the Lemma follows with (19). \end{proof} \subsection{Viscosity solution property} \subsubsection{Viscosity subsolution property of $w$} \begin{prop} $w$ is a viscosity subsolution of (7) on $[0,T]\times\mathbb{R}^d\times\textnormal{Int}(\Delta(I)).$ \end{prop} \begin{proof} Let $\phi:[0,T]\times\mathbb{R}\times \Delta(I)\rightarrow\mathbb{R}$ be a test function such that $w-\phi$ has a strict global maximum at $(\bar{t},\bar x,\bar p)$, where $\bar p\in\textnormal{Int}(\Delta(I))$. We have to show, that \begin{eqnarray} \min\bigg\{\frac{\partial \phi}{\partial t}+\frac{1}{2}\textnormal{tr}(\sigma\sigma^*(t,x)D_x^2\phi)+H(t,x,D_x\phi,p), \lambda_{\min} \left(p,\frac{\partial^2 \phi}{\partial p^2}\right)\bigg\}\geq 0 \end{eqnarray} holds at $(\bar{t},\bar x,\bar p)$. As a limit of convex functions $w$ is convex in $p$ and we have since $\bar p\in\textnormal{Int}(\Delta (I))$ \[\lambda_{\min}\left(\bar p,\frac{\partial ^2 \phi}{\partial p^2}(\bar{t},\bar x,\bar p)\right)\geq 0.\]\\ So it remains to show \begin{eqnarray} \frac{\partial \phi}{\partial t}+\frac{1}{2}\textnormal{tr}(\sigma\sigma^*(t,x)D_x^2\phi)+H(t,x,D_x\phi,p)\geq 0. \end{eqnarray} Note that by standard arguments (e.g. \cite{Bar}) there exists a sequence $(\bar t_k,\bar x_k,\bar p_k)_{k\in\mathbb{N}}$ such that $\bar t_{k}=l_k \frac{T}{k}= l_k \tau \in\Pi^\tau$ converges to $\bar t$ and $(\bar x_k,\bar p_k)$ converge to $(\bar x,\bar p)$ and such that $V^\tau-\phi$ has a global maximum at $(\bar{t}_{k},\bar x_k,\bar p_k)$.\\ Define $\phi^\tau=\phi+(V^\tau(\bar{t}_{k},\bar x_k,\bar p_k)-\phi(\bar{t}_{k},\bar x_k,\bar p_k))=\phi+\Delta_\tau$. Hence for all $x\in\mathbb{R}, p\in \Delta(I)$ \begin{eqnarray*} V^\tau(\bar t_{k}+\tau,x,p)-\phi^\tau(\bar t_{k}+\tau,x,p)\leq V^\tau(\bar t_{k},\bar x_k,\bar p_k)-\phi^\tau(\bar t_{k},\bar x_k,\bar p_k)=0. \end{eqnarray*} Set \begin{eqnarray*} \bar X_{k+1}=\bar x_k+\sigma(\bar t_k,\bar x_k) \Delta B^{l_k} \end{eqnarray*} and \begin{eqnarray*} \bar z_{k}=\frac{1}{\tau}\mathbb{E}\left[V^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_{k}) (\sigma^*)^{-1}(\bar t_k,\bar x_k)\Delta B^{l_k}\right]. \end{eqnarray*} By the definition of $V^\tau$ (12) it holds \begin{eqnarray*} 0&=&\textnormal{Vex}_p\left(\mathbb{E}\left[V^{\tau}(\bar t_{k}+\tau,\bar X_{k+1}, \bar p_k)+\tau H(\bar t_{k},\bar x_k,\bar z_{k},\bar p_k)ds\right]\right)-V^{\tau}(\bar t_{k},\bar x_k,\bar p_k)\\ &\leq& \mathbb{E}\left[V^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)\right]+\tau H(\bar t_{k},\bar x_k,\bar z_{k},\bar p_k)-V^{\tau}(\bar t_{k},\bar x_k,\bar p_k). \end{eqnarray*} Hence by the monotonicity Lemma 3.2. we have for all $\tau>0$ \begin{eqnarray*} 0&\leq& \mathbb{E}\left[V^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\tau H(\bar t_{k},\bar x_k,\frac{1}{\tau}\mathbb{E}\left[V^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k) (\sigma^*)^{-1}(\bar t_k,\bar x_k) \Delta B^{l_k}\right],\bar p_k)\right]\\ &&\ \ \ \ \ \ \ \ -V^{\tau}(\bar t_{k},\bar x_k,\bar p_k)\\ &\leq& \mathbb{E}\left[\phi^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\tau H(\bar t_{k},\bar x_k,\frac{1}{\tau}\mathbb{E}\left[\phi^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k) (\sigma^*)^{-1}(\bar t_k,\bar x_k)\Delta B^{l_k}\right],\bar p_k)\right]\\ &&\ \ \ \ \ \ \ \ -\phi^{\tau}(\bar t_{k},\bar x_k,\bar p_k)+\tau \mathcal{O}(\tau). \end{eqnarray*} By expansion of the smooth function $\phi^\tau$ we have since $\phi^\tau$ is equal to $\phi$ with the linear shift $\Delta_\tau$ the inequality (24). \end{proof} \subsubsection{Viscosity supersolution property of $w$} \begin{prop} $w$ is a viscosity supersolution of (7) on $[0,T]\times\mathbb{R}^d\times\Delta(I).$ \end{prop} \begin{proof} To show that $w(t,x,p)$ is a viscosity supersolution of (7) let $\phi:[0,T]\times\mathbb{R}\times\Delta(I)$ be a test function, such that $w-\phi$ has a strict global minimum at $(\bar{t},\bar x,\bar p)$ with $w(\bar{t},\bar x,\bar p)-\phi(\bar{t},\bar x,\bar p)=0$ and such that its derivatives are uniformly Lipschitz continuous in $p$.\\ We have to show, that \begin{eqnarray} &&\min\bigg\{\frac{\partial \phi}{\partial t}+\frac{1}{2}\textnormal{tr}(\sigma\sigma^T(t,x)D_x^2\phi)+b(t,x)D_x\phi+H(t,x,D_x\phi,p), \lambda_{\min} \left(p,\frac{\partial^2 \phi}{\partial p^2}\right)\bigg\}\leq 0 \end{eqnarray} holds at $(\bar{t},\bar x,\bar p)$. Observe that, if $\lambda_{\min} \left(\frac{\partial^2 \phi}{\partial p^2}\right)\leq0$ at $(\bar{t},\bar x,\bar p)$, then (25) follows immediately. So we assume now $\lambda_{\min} \left(\frac{\partial^2 \phi}{\partial p^2}\right)>0$.\\ By standard arguments (e.g. \cite{Bar}) there exists a sequence $(\bar{t}_k,\bar x_k,\bar p_k)_{k\in\mathbb{N}}$ such that $\bar t_{k}= l_k \tau \in\Pi^\tau$ converges to $\bar t$ and $(\bar x_k,\bar p_k)$ converge to $(\bar x,\bar p)$ and such that $V^\tau-\phi$ has a global minimum at $(\bar{t}_k,\bar x_k,\bar p_k)$.\\ Define $\phi^\tau=\phi+(V^\tau(\bar{t}_k,\bar x_k,\bar p_k)-\phi(\bar{t}_k,\bar x_k,\bar p_k))=\phi+\Delta_\tau$. Since the minimum is global, we have \[V^\tau(\bar t_k+\tau,x,p)-\phi^\tau(\bar t_k+\tau,x,p)\geq V^\tau(\bar{t}_k,\bar x_k,\bar p_k)-\phi(\bar{t}_k,\bar x_k,\bar p_k)= 0.\] Note that by the assumption $\lambda_{\min} \left(\frac{\partial^2 \phi}{\partial p^2}\right)>0$ there exists $\delta,\eta>0$ such that for all $k$ great enough we have \begin{eqnarray} \langle \frac{\partial^2 \phi^\tau}{\partial p^2}(t,x,p)z,z\rangle>4\delta|z|^2\ \ \ \ \ \ \forall (x,p)\in B_\eta(\bar x_k,\bar p_k),\ \ t\in[\bar t_k,\bar t_{k}+\tau],\ \ \ z\in T_{\Delta(I)(\bar p_k)}. \end{eqnarray} Since $\phi^\tau$ is a test function for a purely local viscosity notion, one can modify it outside a neighborhood of $(\bar t_k,\bar x_k,\bar p_k)$, such that for all $(s, x)\in [\bar t_k,T]\times\mathbb{R}^d$ the function $\phi^\tau(s,x,\cdot)$ is convex on the whole convex domain $\Delta(I)$. Thus for any $p\in\Delta(I)$ it holds \begin{eqnarray} V^\tau(s,x,p)\geq \phi^\tau(s,x,p)\geq \phi^\tau(s,x,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}(s,x,p),p-\bar p_k\rangle. \end{eqnarray} We proceed in several steps. \begin{itemize} \item [(1)] First we show a local estimate which is stronger than (27) using (26). \item [(2)] In the second step we establish estimates for $\bold p_{k+1}:=\bold p_{l_k+1}^{\bar p_k,\bar x_k}$ where $\bold p_{l_k+1}^{\bar p_k,\bar x_k}$ is defined as one step martingale with initial data $(\bar t_k,\bar x_k,\bar p_k)$ as in Definition 4.2. \item [(3)] Then we use the estimates of the second step together with the monotonicity in Lemma 3.3. to conclude the viscosity supersolution property. \end{itemize} \textbf{Step 1}: We claim that there exist $\eta,\delta>0$, such that for all $\tau>0$ small enough (meaning $k$ great enough) it holds \begin{eqnarray} V^\tau(\bar t_{k}+\tau,x,p)\geq\phi^\tau(\bar t_{k}+\tau,x,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}( \bar t_{k}+\tau ,x,\bar p_k),p-\bar p_k\rangle+\delta|{p}-\bar p_k|^2. \end{eqnarray} for all $x\in B_\eta(\bar x_k)$, $p\in\Delta(I)$. By Taylor expansion in p \begin{eqnarray} \phi^\tau(t,x,p)\geq \phi^\tau(t,x,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}(t,x,p),p-\bar p_k\rangle+2\delta|p-\bar p_k|^2 \end{eqnarray} holds for $(x,p)\in B_\eta(\bar x_k,\bar p_k)$, $t\in[\bar t_k, \bar t_k+\tau]$. Hence (28) is true locally in $p$. To establish (28) for all $p\in\Delta(I)$ we set for $p\in\Delta(I)\setminus \textnormal{Int}(B_\eta(\bar p_k))$ \[ \tilde p = \bar p_k+\frac{p-\bar p_k}{|p-\bar p_k|}\eta.\] So by the convexity of $V^\tau$ in $p$ and (29) we have for a $\hat{ p}\in\partial {V^\tau}^-(\bar t_k,\bar x_k, \tilde p)$ \begin{eqnarray*} V^\tau(\bar t_{k},\bar x_k,p) &\geq& V^\tau(\bar t_k,\bar x_k,\tilde p)+\langle \hat p,p-\tilde p \rangle\\ &\geq& \phi^\tau(\bar t_{k},\bar x_k,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k),\tilde p-\bar p_k\rangle+2\delta\eta^2+\langle \hat p,p-\tilde p \rangle\\ &\geq& \phi^\tau(\bar t_{k},\bar x_k,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k), p-\bar p_k\rangle+2\delta\eta^2+\langle \hat p-\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k),p-\tilde p \rangle. \end{eqnarray*} Since $\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k)\in\partial{V^\tau}^-(\bar t_k,\bar x_k, \bar p_k)$ and $p-\tilde p=c (p-\bar p_k)$ $(c>0)$ and $V^\tau$ is convex in $p$ it holds \[\langle \hat p-\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k),p-\tilde p \rangle\geq0.\] So we have for all $p\in\Delta(I)\setminus \textnormal{Int}(B_\eta(\bar p_k))$ \begin{eqnarray} V^\tau(\bar t_{k},\bar x_k,p) \geq \phi^\tau(\bar t_{k},\bar x_k,\bar p_k)+\langle\frac{\partial \phi^\tau }{\partial p}(\bar t_{k},\bar x_k,\bar p_k), p-\bar p_k\rangle+2\delta\eta^2 \end{eqnarray} which gives in the limit for all $p\in\Delta(I)\setminus \textnormal{Int}(B_\eta(\bar p))$ \begin{eqnarray} w(\bar t,\bar x,p) \geq \phi(\bar t,\bar x,\bar p)+\langle\frac{\partial \phi }{\partial p}(\bar t,\bar x,\bar p), p-\bar p\rangle+2\delta\eta^2. \end{eqnarray} Assume now that (28) does not hold for a $p\in\Delta(I)$. Hence there exists a sequence $(\tau,x_{k_n},p_{k_n})\rightarrow(0, 0, p)$ with $\tau=\frac{T}{n}$, ${p_{k_n}}\in\Delta(I)\setminus B_\eta(\bar p_{k_n})$, such that \begin{eqnarray*} &&V^\tau(\bar t_{k_n}+\tau,\bar x_{k_n}+x_{k_n},p_{k_n})\\ &&\ \ \ \ \ \ < \phi^\tau(\bar t_{k_n}+\tau,\bar x_{k_n}+x_{k_n},\bar{p}_{k_n})+\langle\frac{\partial \phi^\tau}{\partial p}(\bar t_{k_n}+\tau,\bar x_{k_n}+x_{k_n},p_{k_n}),p_{k_n}-\bar p_{k_n}\rangle+\delta|p_{k_n}-\bar p_{k_n}|^2 \end{eqnarray*} Thus for $n\rightarrow\infty$, $p\in\Delta(I)\setminus \textnormal{Int}(B_\eta(\bar p))$ and \begin{eqnarray} w(\bar t,\bar x,p)< \phi(\bar t,\bar x,\bar p)+\langle\frac{\partial \phi }{\partial p}(\bar t,\bar x,\bar p), p-\bar p\rangle+\delta\eta^2 \end{eqnarray} which contradicts (31).\\ In the following we denote \begin{eqnarray*} \bar X_{k+1}=\bar x_k+\sigma(\bar t_k,\bar x_k) \Delta B^{l_k}. \end{eqnarray*} where $\Delta B^{l_k}= B_{\bar t_k+\tau}-B_{\bar t_k}$. With the estimate (28) we have for $\tau$ small enough for all $p\in\Delta(I)$ \begin{eqnarray*} &&\mathbb E\left[V^\tau(\bar t_{k}+\tau,\bar X_{k+1},p)\right]\\ &&=\mathbb E\left[V^\tau(\bar t_{k}+\tau ,\bar X_{k+1},p)1_{|\bar X_{k+1}-\bar x_k|< \eta}\right]+\mathbb{E}\left[V^\tau(\bar t_{k}+\tau,\bar X_{k+1},p)1_{|\bar X_{k+1}-\bar x_k|\geq \eta}\right]\\ &&\geq\mathbb E\left[\left(\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),p-\bar p_k\rangle+\delta |p-\bar p_k|^2\right)1_{|\bar X_{k+1}-\bar x_k|< \eta}\right]\\ &&\ \ \ \ \ \ +\mathbb{E}\left[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},p)1_{|\bar X_{k+1}-\bar x_k|\geq \eta}\right]\\ &&=\mathbb E\left[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),p-\bar p_k\rangle+\delta 1_{|\bar X_{k+1}-\bar x_k|< \eta} |p-\bar p_k|^2)\right]\\ &&\ \ \ \ \ \ +\mathbb{E}\bigg[1_{|\bar X_{k+1}-\bar x_k|\geq\eta}\bigg(\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},p)-\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k) \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ -\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),p-\bar p_k\rangle\bigg)\bigg]. \end{eqnarray*} Recalling that $\phi^\tau$ is convex with respect to $p$, we get for all $p\in\Delta(I)$ \begin{equation} \begin{array}{rcl} \mathbb E\left[V^\tau(\bar t_{k}+\tau,\bar X_{k+1},p)\right]&\geq&\mathbb E\bigg[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),p-\bar p_k\rangle\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +\delta 1_{|\bar X_{k+1}-\bar x_k|< \eta} |p-\bar p_k|^2)\bigg]. \end{array} \end{equation} \textbf{Step 2}: Next we establish an estimate for $\bold p_{k+1}:=\bold p_{l_k+1}^{\bar p_k,\bar x_k}$ where $\bold p_{l_k+1}^{\bar p_k,\bar x_k}$ is defined as one step martingale as in Definition 4.2. with initial data $(\bar t_k,\bar x_k,\bar p_k)$.\\ Note that by the one step dynamic programming (21) it holds \begin{eqnarray} V^\tau(\bar t_k,\bar x_{k}, \bar p_k) =\mathbb{E}\left[ V^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bold p_{k+1})+\tau H(\bar t_{k},\bar x_k,\bar z_{k}(\bar x_k,\bold p_{k+1}),\bold p_{k+1})\right]. \end{eqnarray} Together with $V^\tau(\bar t_k,\bar x_{k}, \bar p_k)=\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)$ and the estimate (33) we have for all small enough $\tau>0$ \begin{eqnarray*} \phi^\tau(\bar t_k,\bar x_{k}, \bar p_k) &\geq&\mathbb{E}\Bigg[ \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_{k})+\tau H(\bar t_{k},\bar x_k,\bar z_{k}(\bar x_k,\bold p_{k+1}),\bold p_{k+1})\\ &&\ \ \ \ \ \ \ +\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),\bold p_{k+1}-\bar p_k\rangle+\delta 1_{|\bar X_{k+1}-\bar x_k|< \eta} |\bar p_k-\bold p_{k+1}|^2\Bigg]. \end{eqnarray*} Since $\bold p_{k+1}$ and $\Delta B^{l_k}$ are independent, $\phi^\tau$ has bounded derivatives and $\bold p_{k+1}$ is a one step martingale, it holds \begin{eqnarray*} &&\mathbb{E}\left[ \langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),\bold p_{k+1}-\bar p_k\rangle\right]\\ &&\ \ \ \ =\mathbb{E}\left[ \langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar x_k+\sigma(\bar t_k,\bar x_k) \Delta B^{l_k},\bar p_k),\bold p_{k+1}-\bar p_k\rangle\right]=0. \end{eqnarray*} Furthermore by the Markovian inequality and assumption (A) we have \begin{eqnarray*} &&\mathbb{E}\left[1_{|\bar X_{k+1}-\bar x_k|< \eta} |\bold p_{k+1}-\bar p_k|^2\right]\\ &&\ \ \ \ =\mathbb{E}\left[1_{|\sigma(\bar t_k,\bar x_k) \Delta B^{l_k}|< \eta}|\bold p_{k+1}-\bar p_k|^2\right]\geq c (1-\tau^\frac{1}{2}) \mathbb{E}\left[|\bold p_{k+1}-\bar p_k|^2\right]. \end{eqnarray*} with a sufficiently small constant $c$ independent of $k$. Thus \begin{equation} \begin{array}{rcl} 0&\geq&\mathbb{E}\Bigg[ \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)-\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)+\tau H(\bar t_{k},\bar x_k,\bar z_{k}(\bar x_k,\bold p_{k+1}),\bold p_{k+1})\\ && \ \ \ \ \ \ \ \ +c \delta (1-\tau^\frac{1}{2})|\bold p_{k+1}-\bar p_k|^2\Bigg]. \end{array} \end{equation} Since $\phi^\tau$ has bounded derivatives it holds by assumption (A) \begin{eqnarray} \left|\mathbb{E}\left[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)-\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)\right]\right|&\leq& c\tau \end{eqnarray} and since $\mathbb{E}\left[|\bar z_{k}(\bar x_k,\bold p_{k+1})|\right]\leq c$ (16) it holds by (A) and H\"older \begin{equation} \begin{array}{rcl} \mathbb{E}\left[\tau H(\bar t_{k},\bar x_k,\bar z_{k}(\bar x_k,\bold p_{k+1}),\bold p_{k+1})\right]&\leq& c \tau. \end{array} \end{equation} Combining (35)-(37) we have for small enough $\tau>0$ and a generic constant $c'>0$ \begin{eqnarray*} \mathbb{E}[|\bold p_{k+1}-\bar p_k|^2]\leq \frac{c'}{c \delta(1-\tau^\frac{1}{2})}\tau, \end{eqnarray*} hence for small enough $\tau$ and a constant $c''>0$ \begin{eqnarray} \mathbb{E}[|\bold p_{k+1}-\bar p_k|^2]\leq c''\tau. \end{eqnarray} \textbf{Step 3}: Furthermore we have with (35) and the monotonicity Lemma 3.3., since $V^\tau(\bar t_k,\bar x_{k}, \bar p_k)=\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)$ \begin{eqnarray} 0 &\geq&\mathbb{E}\bigg[ \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bold p_{k+1})-\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)+\tau H(\bar t_{k},\bar x_k,\tilde z_{k}(\bar x_k,\bold p_{k+1}),\bold p_{k+1})\bigg]-\tau \mathcal{O}(\tau), \end{eqnarray} where \begin{eqnarray*} \tilde z_{k}(\bar x_k,\bold p_{k+1})=\frac{1}{\tau}\mathbb{E}\left[\phi^{\tau}(\bar t_{k}+\tau,\bar X_{k+1},p) (\sigma^*)^{-1}(\bar t_k,\bar x_k)\Delta B^{l_k}\right]\big|_{p=\bold p_{k+1}}. \end{eqnarray*} From the construction of $\bold p_{k+1}$ and the fact that $\phi^\tau$ is convex it holds with (27) \begin{equation} \begin{array}{rcl} \mathbb{E}\left[ \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bold p_{k+1})\right] &\geq&\mathbb E\left[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)+\langle \frac{\partial}{\partial p} \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k),\bold p_{k+1}-\bar p_k\rangle\right]\\ \ \\ &=&\mathbb E\left[\phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)\right]. \end{array} \end{equation} It remains to get a suitable estimate for $\tilde z_{k}(\bar x_k,\bold p_{k+1})$. Since $\phi^{\tau}$ is uniformly Lipschitz continuous in $x$, it holds by Taylor expansion in $x$ \begin{eqnarray*} \tilde z_{k}(\bar x_k,\bold p_{k+1})&=&\frac{1}{\tau}\mathbb{E}\left[\phi^{\tau}(\bar t_{k}+\tau,\bar X_{{k+1}}, p) (\sigma^*)^{-1}(\bar t_k,\bar x_k)\Delta B^{l_k}\right]\big|_{p=\bold p_{k+1}}\\ &=&\frac{1}{\tau}\mathbb{E}\left[\phi^{\tau}(\bar t_{k}+\tau,\bar x_k,p) (\sigma^*)^{-1}(\bar t_k,\bar x_k)\Delta B^{l_k}\right]\big|_{p=\bold p_{k+1}}\\ &&\ \ \ \ \ +\frac{1}{\tau}\mathbb{E}\left[D_x \phi^{\tau}(\bar t_{k}+\tau,\bar x_k,p)|\Delta B^{l_k}|^2\right]\big|_{p=\bold p_{k+1}}+ \mathcal{O}(\tau)\\ &=&\frac{1}{\tau}\mathbb{E}\left[D_x \phi^{\tau}(\bar t_{k}+\tau,\bar x_k,p)|\Delta B^{l_k}|^2\right]\big|_{p=\bold p_{k+1}}+ \mathcal{O}(\tau). \end{eqnarray*} Furthermore since $D_x\phi^{\tau}$ is Lipschitz continuous in $p$ it holds with (38) \begin{eqnarray*} &&\mathbb{E}\left[\left|D_x \phi^{\tau}(\bar t_{k}+\tau,\bar x_k,\bold p_{k+1})|\Delta B^{l_k}|^2-D_x \phi^{\tau}(\bar t_{k}+\tau,\bar x_k,\bar p_{k})|\Delta B^{l_k}|^2\right|\right]\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \leq c\mathbb{E}\left[|\bold p_{k+1}-p_k||\Delta B^{l_k}|^2\right]\leq c\tau^\frac{3}{2} \end{eqnarray*} So from (40) we have \begin{eqnarray*} 0&\geq& \mathbb{E}\Bigg[ \phi^\tau(\bar t_{k}+\tau,\bar X_{k+1},\bar p_k)-\phi^\tau(\bar t_k,\bar x_{k}, \bar p_k)+\tau H(\bar t_{k},\bar x_k, D_x \phi^{\tau}(\bar t_{k}+\tau,\bar x_k,\bar p_{k}) ,\bar p_{k})\Bigg]\\ &&\ \ \ \ -c({\tau}^\frac{3}{2}+\tau \mathcal{O}(\tau)) \end{eqnarray*} which implies (25) since $\phi^\tau$ is equal to $\phi$ up to a linear shift. \end{proof} \section{Concluding Remarks and Outlook} In this paper we gave an approximation scheme for the value function of a stochastic differential game with incomplete information. It is natural to ask whether this approximation might be used to determine optimal feedback strategies for the informed player. In the deterministic games with complete information it is well known that the answer is positive (see the step by step motions associated with feedbacks in \cite{K}). The case of deterministic games with incomplete information has been treated in \cite{Carda}.\\ The approximation of optimal strategies for stochastic differential games is a more delicate topic even in the case with complete information. \cite{Ba} - also considering the game under a Girsanov transform - gives a partwise answer under a weak Lipschitz assumption of the feedback control. The result is shown by using approximations of BSDEs however not in a completely discrete framework. In the very recent paper \cite{FH} approximately Markov strategies are constructed with an approximation that in contrast to ours takes into account the actions of the other player during the time intervals. This however makes the approximation much harder to implement.\\ In fact, if we use the approximation for the construction of optimal strategies for the informed player we are in the same situation as \cite{Ku}. For the approximation of the value function in \cite{Ku} nearly optimal policies are constructed which possess a certain optimality in the approximative discrete time games instead of the continuous time one. To the authors knowledge the problem of finding an efficient approximation of optimal strategies in stochastic differential games (with or without incomplete information) is open and poses an interesting problem for further research.
\section{Introduction} The classical kissing number problem for sphere packings is the search for an optimal upper bound on the number of $n$-dimensional (euclidean) unit spheres, pairwise disjoint in their interior, that can be tangent to a fixed unit sphere. Exact values for these numbers, commonly called kissing numbers for simplicity, are only known in a finite number of cases ($n=1-4, 8, 24$, see \cite{pfzi04} and references therein). A seemingly easier problem is to ask the same question as above, but with the restriction that the centers of the spheres lie on some euclidean lattice. This gives rise to the so-called kissing number problem for lattice sphere packings. As above, values are only known in a finite number of cases ($n=1-9, 24$, see \cite{coslbook} and references therein). In some cases the solutions to the two problems coincide (but not in general, for instance they are known to be different for $n=9$). As an example, when $n=2$, the answer is $6$ for both problems and the optimal solution is given by points lying on the hexagonal lattice. In this case, the question is equivalent to asking how many shortest (nontrivial) vectors a lattice can contain. In terms of {\it systoles} (meaning shortest non-trival curves of a manifold), how many distinct isotopy classes of systoles can a flat torus have? The kissing number problem for lattices is the natural generalization: how many isotopy classes of systoles can the underlying $n$-torus contain? Another possible generalization - proposed by Schmutz Schaller - is to hyperbolic closed surfaces. We ask the same question where genus plays the part of dimension. For hyperbolic surfaces there is a unique geodesic in a prescribed nontrivial isotopy class, so the question is as follows. {\it How many systoles can a hyperbolic surface of genus $g$ have?} By analogy with the lattice case, Schmutz Schaller defined the maximum number for each genus to be the kissing number for hyperbolic surfaces. Usually, we consider simple closed geodesics, and systoles in particular, to be unoriented objects, so morally we should be dividing lattice sphere packing kissing numbers by two to make the two problems truly analogous. As in the case of lattices, very little is known about exact values: the only known case is for genus $2$ where the answer is $12$. A related problem for systoles of hyperbolic surfaces is to the search for surfaces with systole of maximal length, again initiated by Schmutz Schaller \cite{sc931}. As before, there is a possible analogy with lattices: this is the hyperbolic surface version of the search for the optimal Hermite constants. (The Hermite constant can be thought of as the square of the maximal systole length of unit volume flat torus.) For each genus, via a compactness argument, there exists a hyperbolic surface with longest possible systole. Again, the only known exact value is for genus $2$, a result of Jenni \cite{je84}. It is realized by the same surface as for the kissing number problem - the so-called Bolza surface. The Bolza surface is also maximal among genus $2$ hyperbolic surfaces for the number of self-isometries. In the non-compact cases, arithmetic surfaces coming from principal congruence subgroups of ${\rm PSL}_2({\mathbb Z})$ are known to be maximal among finite area hyperbolic surfaces in their respective moduli spaces, a result of Schmutz Schaller \cite{sc941}, see also \cite{ad98} for another proof. For both problems, although finding exact solutions seems difficult, one can ask for upper and lower bounds, and in particular it is natural to study the asymptotic growth of how these constants vary in function of genus. Via a simple area argument, it is easy to see that the systole length of a hyperbolic surface cannot exceed $2 \log g$ ($+C$ for some constant $C>0$). Buser and Sarnak \cite{busa94} were the first to construct families of surfaces with $\sim \log g$ systole growth. More precisely, they showed $$ \limsup_{g\to \infty}\frac{ \max_{S\in {\mathcal M}_g} {\rm sys}(S)}{ \log g}\geq \frac{4}{3} $$ where ${\mathcal M}_g$ is the moduli space of all hyperbolic surfaces up to isometry and ${\rm sys}(S)$ is the length of the systole of $S$. Since then there have been other constructions, see for example \cite{kascvi07}, but all are based on arithmetic methods of some sort. The existence of such surfaces is somewhat surprising: the radius of a maximally embedded disk in a hyperbolic surface is $\sim \log g$, so a surface with $\sim \log g$ systole looks essentially everywhere like a fat disk that is pasted together in some clever way in order to avoid creating a short loop somewhere. Schmutz Schaller wrote a series of papers for the hyperbolic surface kissing number problem, in both the compact and non-compact cases, where he proved some interesting lower bounds. In first instance, one might think that it might not be possible to have a surface where the number of systoles is considerably bigger than the size of a maximal isometry group. Via Hurwitz's bound, this would imply an upper bound to the number of systoles which grows linearly in genus. Schmutz Schaller's first result on this was the existence of families of surfaces with a number of systoles that grew more than linearly in genus \cite{sc962,sc964}. The best results for closed surfaces appeared later \cite{sc97} where he showed that $$ \limsup_{g\to \infty} \frac{ \log \left(\max_{S\in {\mathcal M}_g} {\rm Kiss}(S)\right)}{\log g} \geq \frac{4}{3}-\varepsilon $$ for any $\varepsilon>0$ where ${\rm Kiss}(S)$ denotes the number of systoles of $S$. Stated otherwise, for any $\varepsilon>0$, there exists a family of surfaces, one in each genus, with more than $g^{\sfrac{4}{3}-\varepsilon}$ systoles (for large enough $g$). Again, the construction is based on arithmetic methods. Based on the intuition that one cannot do better for arithmetic surfaces and that arithmetic surfaces should be optimal for these problems, Schmutz Schaller \cite{sc98} made two conjectures, namely that the inequalities above are in fact equalities (with the right hand side equal to $\sfrac{4}{3}$). For the sake of avoiding repititious repetition, we rephrase them as follows. \begin{conjecture}\label{conj:size} There exists a constant $A>0$ such that $$ \max_{S\in {\mathcal M}_g} {\rm sys}(S) \leq \frac{4}{3} \log g +A. $$ \end{conjecture} \begin{conjecture}\label{conj:number} There exists a constant $B>0$ such that $$ \max_{S\in {\mathcal M}_g} {\rm Kiss}(S) \leq B g^{\sfrac{4}{3}}. $$ \end{conjecture} The two problems seem to run in parallel because of the arithmetic nature of the interesting known examples, but Schmutz Schaller did not establish a direct link between the two. In contrast to Conjecture \ref{conj:size}, there was no ``easy" known upper bound for Conjecture \ref{conj:number} that behaves roughly as conjectured. One might expect an easy quadratic upper bound for the number of systoles which would seem to be the counterpart to the straightforward $2 \log g$ bound. (In fact Schmutz Schaller claims such a quadratic bound in \cite{sc942} but the argument, supposedly only based on the topological condition that two systoles pairwise intersect at most once, is faulty and does not seem to be easily repairable.) The first result of this paper is to find an upper bound for ${\rm Kiss}(S)$, which depends on the length of the systole of $S$. \begin{theorem}\label{thm:mainhyperbolicintro} There exists a constant $U>0$ such that for any hyperbolic surface $S$ of genus $g$ with systole $\ell$ the following holds: $$ {\rm Kiss}(S) \leq U\, \frac{e^{\,\ell/2}}{\ell} \,g. $$ \end{theorem} Using the $2\,\log g$ bound on $\ell$ mentioned above, this has the following consequence. \begin{corollary}\label{cor:subquad} There exists a constant $U>0$ such that any hyperbolic surface of genus $g$ has at most $U \dfrac{g^2}{\log g}$ systoles. \end{corollary} More generally, if there exists an upper bound on length of systoles of type $C \log g$ for some $C>0$, then Theorem \ref{thm:mainhyperbolicintro} implies a bound of order $\sim g^{\sfrac{C}{2} + 1}$. In particular a positive answer to Conjecture \ref{conj:size} would imply that there are at most $\sim g^{\sfrac{5}{3}}$ systoles on a genus $g$ surface. Another consequence is that Conjecture \ref{conj:number} holds for all surfaces with systole bounded above by $\sfrac{2}{3} \log g$. The bound also shows that if a family of surfaces has sub-logarithmic systole growth, then the number of systoles is ``almost" at most linear. More precisely: \begin{corollary}\label{cor:linear} Let $f(g)$ be any positive function with $\lim_{g\to \infty} \sfrac{f(g)}{\log g} = 0$. Then $$ \max \{{\rm Kiss}(S) \,|\, S\in {\mathcal M}_g{\text{ with }}{\rm sys}(S)\leq f(g)\} \leq g^{1+\varepsilon}. $$ for any $\varepsilon>0$ and large enough $g$. \end{corollary} The theorem implies more accurate than the above corollary for ``intermediate" growth, but the formulation above is given for clarity. In particular, this means that any family of surfaces with ``many" systoles (by which we mean at least $g^{1+a}$ for some $a>0$), then the family has $\sim \log g$ systole growth as well. Of course it is not a priori easier to construct surfaces with many systoles, but if one does, then the large size systoles come for free. In the more general context of Riemannian metrics on surfaces, similar questions can be asked. One no longer has uniqueness of a geodesic in an isotopy class, so the appropriate question on the number of systoles is an upper bound on the number of distinct isotopy classes of simple closed curves that can simultaneously be realized as systoles for some metric. {\it How many systoles up to isotopy can a closed Riemannian surface of genus $g$ have?} We'll relate to optimal upper bounds to this problem as the kissing number problem for general surfaces. It is not known what sets of topological curves on either a hyperbolic or Riemannian surface can be realized by systoles. By a cutting and pasting argument, it is not too difficult to see that on any closed surface (not necessarily hyperbolic) a systole is necessarily a simple closed curve. Likewise, any two systoles can pairwise intersect at most once. This leads us to a related purely topological problem due to B. Farb and C. Leininger (see \cite{marith10}). {\it Up to isotopy, how many distinct curves can be realized on a surface of genus $g$ such that they pairwise intersect at most once?} It seems to be a surprisingly hard question to answer. The best known upper bound is in fact exponential, and the best lower bound quadratic \cite{marith10}. These numbers provide upper bounds for the kissing number problems, and to the best of the author's knowledge, these are the best known upper bounds for kissing numbers of general surfaces and were the best bounds even in the case of hyperbolic surfaces prior to Corollary \ref{cor:subquad}. Observe that Corollary \ref{cor:subquad} shows that optimal kissing numbers for hyperbolic surfaces cannot be the same as the numbers coming from the purely topological problem. It was already known that the topological condition was quite different from the systolic condition: in \cite{paanpesys}, there are constructions of configurations of isotopy classes curves that fail to be systoles for {\it any} Riemannian metric on the surface. For completeness, we mention that the problem of finding an upper bound on systole length in the case of variable curvature can also be made to make sense. For a genus $g$ surface with area normalized to the area of its hyperbolic counterparts, Gromov \cite{gr83,gr96} gave a $\sim \log g$ upper bound, which in light of the hyperbolic examples is roughly optimal, but the precise asymptotic growth remains completely open. Whether the asymptotic growth for both problems should be different for both the upper bound on length or the kissing numbers is also unknown. The second main result of this paper is about lower bounds for kissing number for Riemannian surfaces. \begin{theorem}\label{thm:generalkiss} There exist surfaces of genus $g>0$ with a number of systoles of order of growth at least $g^{\sfrac{3}{2}}$. \end{theorem} This shows that if Conjecture \ref{conj:number} is correct, the asymptotic growth for kissing numbers is considerably different in the case of variable curvature surfaces. Recall the best known - and conjectured optimal - lower bounds for hyperbolic surfaces are roughly $g^{\sfrac{4}{3}}$. The proof is by construction, and the geometry of the surfaces come from embeddings of complete graphs. One might wonder if one can use these techniques to find hyperbolic surfaces with the same behavior, but the surfaces are far from being hyperbolic. In fact, in striking contrast to what is possible in the hyperbolic case (Corollary \ref{cor:linear}), the systole length is quite small proportionally to area (constant systole length for area $\sim g$). \section{Bounds on numbers of systoles of hyperbolic surfaces} We denote ${\rm Kiss}(S)$ the number of systoles of a surface $S$. Before proceeding to proofs of upper bounds on ${\rm Kiss}(S)$ for hyperbolic surfaces, we begin with some observations on the geometry of systoles. \subsection{Geometric properties of systoles} The estimates needed rely essentially on trigonometric arguments in the hyperbolic plane. Recall the classical collar lemma. \begin{lemma}[Collar lemma] Let $\gamma$ be a simple closed geodesic of length $\ell$. Then there is an embedded collar of width $w(\ell)$ around $\gamma$ where $$ w(\ell)= {\,\rm arcsinh}\left(\frac{1}{\sinh(\ell/2)}\right). $$ Furthermore, any simple closed geodesic $\delta$ that enters this collar essentially intersects $\gamma$. \end{lemma} For systoles, one can do even better. \begin{lemma}[Systolic collar lemma]\label{lem:scl} Let $\alpha$ and $\beta$ be systoles of length $\ell$ that don't intersect. Then there are at a distance at least $2 r(\ell)$ where $$ r(\ell) = {\,\rm arcsinh}\left(\frac{1}{2\,\sinh(\ell/4)}\right). $$ \end{lemma} \begin{proof} Take a shortest path $c$ between $\alpha$ and $\beta$ and consider the geodesic $\gamma$ in the free homotopy class of $\alpha * c * \beta * c^{-1}$ where $\alpha,\beta$ and $c$ are oriented so that $\gamma$ is simple (see figure \ref{fig:orientedpants}). Note that $\alpha,\beta,\gamma$ form a pair of pants. \vspace{-6pt} \begin{figure}[h] \leavevmode \SetLabels \L(.37*.92) $\alpha$\\ \L(.62*.92) $\beta$\\ \L(.49*.11) $\gamma$\\ \L(.49*.6) $c$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/orientedpants.pdf,width=4.5cm,angle=0}}} \vspace{-24pt} \end{center} \caption{Orientations of $\alpha$, $\beta$ and $c$} \label{fig:orientedpants} \end{figure} Now because $\ell(\gamma)\geq \ell$, we can compute the minimal distance between $\alpha$ and $\beta$. The result follows from standard trigonometry in a pair of pants and the double $\sinh$ formula. \end{proof} From this the following corollary is immediate. \begin{corollary}\label{cor:close} If $\gamma$ and $\delta$ are two systoles of length at most $\ell$ which pass through a same disk of radius $r(\ell)$, then they essentially intersect. \end{corollary} Using only this observation, it is possible to obtain a universal polynomial bound on the number of systoles of hyperbolic surfaces, but we need to work harder to obtain our estimates. We begin by noticing that if two systoles intersect, then their angle of intersection can be bounded below by their length. \begin{lemma}\label{lem:angle} Let $\gamma$ and $\delta$ be systoles of length $\ell$ that intersect. Then $\sin \angle(\gamma,\delta) > \frac{1}{2} \left(\cosh(\ell/4)\right)^{-1}. $ \end{lemma} \begin{proof} Let $p$ be the intersection point of the two curves and consider the point $q$ of $\delta$ that is at distance exactly $\ell/2$ away from $p$. Fix one of the two paths of $\delta$, say $\delta_1$, between $q$ and $\gamma$. Among all paths freely homotopic to $\gamma_1$ with one fixed endpoint at $q$ and the other on any point of $\gamma$, there is one of minimal length which we shall denote $h$. Denote $q'$ the point of intersection of $h$ and $\gamma$. We now have a right triangle $p,q,q'$ with hypothenuse $H$ is of length $\sfrac{\ell}{2}$, the side $h$, and the basis $b$. Observe that using the other arc of $\delta$, say $\delta_2$, one obtains a symmetric situation with an isometric right angle triangle where the two triangles are linked via a rotation of angle $\pi$ around $q$ (see figure \ref{fig:anglepants}). \begin{figure}[h] \leavevmode \SetLabels \L(.13*.466) $\gamma$\\ \L(.19*.33) $p$\\ \L(.325*.49) $q$\\ \L(.26*.73) $\delta$\\ \L(.52*.69) $p$\\ \L(.67*.58) $q$\\ \L(.534*.865) $q'$\\ \L(.805*1.04) $p$\\ \L(.49*.9) $\gamma$\\ \L(.88*.937) $\gamma$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/anglepants1.pdf,height=3.5cm,angle=0}\hspace{1.0cm} \epsfig{file =Figures/anglepants2.pdf,height=3.5cm,angle=0}}} \vspace{-18pt} \end{center} \caption{Two intersecting systoles} \label{fig:anglepants} \end{figure} As the two triangles are isometric, we can concentrate on the first one. It is not difficult to see that $b$ is of length at most $\ell/4$, otherwise one can construct a shorter nontrivial path than $\delta$. Consider the angle $\theta=\angle(\gamma,\delta)$. By the sine formula for hyperbolic triangles, it satisfies $$ \frac{\sin(\theta)}{\sinh(h)} = \frac{\sin(\pi/2)}{\sinh(\ell/2)}. $$ Now because the basis is at most $\ell/4$, $h$ is strictly greater than $\ell/4$ (otherwise the same occurs in the other triangle, and we obtain $\ell(\delta)<\ell$). The inequality $$ \sin \angle(\gamma,\delta) > \frac{\sinh(\ell/4)}{\sinh(\ell/2)} $$ and one concludes by using the $\sinh$ formula for a double angle. \end{proof} \begin{notation} We denote $\theta_\ell := \arcsin \frac{1}{2 \cosh(\ell/4)}$. \end{notation} For large $g$, the behavior is roughly $e^{\ell/4}$. In particular, using the $2 \log g$ upper bound on length, this implies a $\frac{1}{\sqrt{g}}$ lower bound on the angle between systoles. Thus a collection of systoles that intersect in a single point cannot have cardinality greater than roughly $\sqrt{g}$. This is very different from the case of variable curvature where one can construct surfaces with $\sim g$ systoles that intersect in a single point, see Remark \ref{rem:gint}. Corollary \ref{cor:close} ensures that two systoles $\gamma,\delta$ with points $p_\gamma \in \gamma$, $p_\delta \in \delta$ that satisfy $d(p_\gamma,p_\delta)< d(\ell)$ must intersect somewhere on the surface. This next lemma gives a bound on how far the intersection point can be from $p_\gamma$ and $p_\delta$. \begin{lemma}\label{lem:dist} Let $\gamma,\delta$ be systoles of length $\ell$ which cross a disk of radius $r(\ell)$ with center $p$. Then the intersection point $q$ between $\gamma$ and $\delta$ satisfies $$ d(p,q) < {\,\rm arcsinh}\left(2 \coth\frac{\ell}{4}\right). $$ \end{lemma} \begin{proof} Consider points $p_\gamma \in \gamma$, $p_\delta \in \delta$ which lie in the disk centered in $p$. Consider the two angles $\angle(q,p,p_\gamma)$ and $\angle(q,p,p_\delta)$. By Lemma \ref{lem:angle}, one of the two angles must be greater than $\theta_\ell\over 2$. Without loss of generality, let us suppose that this angle is $\theta=\angle(q,p,p_\gamma)$. We note that via Lemma \ref{lem:angle} again we obtain $$ \sin(\theta) \geq \sin({\theta_\ell / 2}) = \frac{1}{2 \cos (\theta_\ell / 2)} \sin(\theta_\ell) > \frac{1}{4 \cosh(\ell/4)}. $$ We now concentrate our attention to the triangle $p,q,p_\gamma$. Denote by $\theta'$ the angle opposite the side $\overline{pq}$ (see figure \ref{fig:disktriangles}). We have the following identity: $$ \frac{\sin(\theta')}{\sinh d(p,q)}= \frac{\sin(\theta)}{\sinh d(p,p_\gamma)}. $$ \begin{figure}[h] \leavevmode \SetLabels \L(.218*.49) $q$\\ \L(.593*.693) $\theta'$\\ \L(.402*.537) $\theta$\\ \L(.63*.5) $p$\\ \L(.645*.79) $p_\gamma$\\ \L(.56*.27) $p_\delta$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/disktriangles.pdf,width=8.0cm,angle=0}}} \vspace{-18pt} \end{center} \caption{Computing the distance to an intersection point} \label{fig:disktriangles} \end{figure} Via Lemma \ref{lem:scl} and what precedes we have $$ \sinh d(p,q) < \frac{\sin(\theta')}{\sin(\theta)} \frac{1}{2\,\sinh(\ell/4)} < 2 \coth \frac{\ell}{4}$$ which concludes the proof. \end{proof} \begin{notation} For future reference, we fix the following notation: $$ R(\ell):={\,\rm arcsinh}\left(2 \coth\frac{\ell}{4}\right). $$ \end{notation} \begin{remark}\label{rem:radius} Observe that via the same argument, for two systoles that intersect in a point $q$ and both simultaneously pass through a disk of radius $\rho$ and center $p$, the following holds: $$ \sinh d(p,q) \leq \frac{\sinh \rho}{\sin(\theta_\ell/2)} < \frac{\sinh \rho}{4 \cosh(\ell/4)}. $$ \end{remark} \subsection{Proof of the upper bound} We can now proceed to the proof of our main upper bound. We begin by giving a more precise version of Theorem \ref{thm:mainhyperbolicintro}. \begin{theorem}\label{thm:mainhyperbolic} Let $S$ be a hyperbolic surface of genus $g$ with systole bounded by $\ell$. Then $$ {\rm Kiss}(S) \leq C_\ell \,(g-1) $$ where $C_\ell$ is a constant that depends on $\ell$ which can be taken to be $$ C_\ell = 100\, \frac{e^{\,\sfrac{\ell}{2}}}{\ell}. $$ \end{theorem} \begin{remark} The constant $100$ in front can easily be improved, but it is the order of growth we are interested in. \end{remark} \begin{proof} If $L\leq 2 {\,\rm arcsinh} 1$ then, by the collar lemma, systoles are disjoint, and thus, as there are at most $3g-3$ disjoint simple closed geodesics on a genus $g$ surface, we obtain $$ {\rm Kiss}(S) \leq 3g-3. $$ We can now concentrate on the case when $\ell \geq 2 {\,\rm arcsinh} 1$. The basic strategy will be the following. We begin by covering the surface $S$ by balls of radius $r(\ell)$ (where $r(\ell)$ is given by Lemma \ref{lem:scl}). The first step will be to estimate $F(S)$, an upper bound on the minimum number of these balls required to cover $S$. We'll then find an upper bound $G(S)$ on the number of systoles that can cross such a ball. Finally, if we denote $H(S)$ the minimum number of covering balls that a systole of $S$ must cross, we have $$ {\rm Kiss}(S) \leq \frac{F(S)\, G(S)}{H(S)}. $$ Let us now concentrate on finding bounds for these quantities in function of $\ell$ and $g$.\\ \noindent{\underline{The number of balls required to cover $S$}} As usual in these type of estimates, we use \begin{center} $ F(S)=\{\text{Number of balls of radius $r(\ell)$ needed to cover $S$}\}$ $\leq$ $ \{\text{Max number of balls of radius $\sfrac{r(\ell)}{2}$ that embed and are pairwise disjoint}\}.$ \end{center} Now as $$ \text{Area}(D_{\sfrac{r(\ell)}{2}}) = 2\pi(\cosh(\sfrac{r(\ell)}{2})-1) = 2\pi\left( \cosh\left(\frac{{\,\rm arcsinh}\left(\frac{1}{2\sinh(\ell/4)}\right)}{2}\right) -1 \right) $$ we deduce a bound for $F(S)$ that depends only on $g$ (coming from the area of $S$) and $\ell$: $$ F(S) \leq \frac{\text{Area}(S)}{\text{Area}(D_{\sfrac{r(\ell)}{2}})}. $$ To get an idea of the order of growth of this bound, observe that $$ \text{Area}(D_{\sfrac{r(\ell)}{2}}) > \frac{\pi}{4} e^{-\sfrac{\ell}{2}} $$ and it follows that $$ F(S) <16 (g-1) \, e^{\sfrac{\ell}{2}}. $$ \vspace{0.3cm} \noindent{\underline{The number of systoles intersecting each ball}} We now proceed to find an upper bound on the number of systoles that can intersect a ball of radius $r(\ell)$. Consider a disk $D_0$ on $S$ of radius $r(\ell)$. Any two given systoles that cross $D_0$ must intersect, and via Lemma \ref{lem:dist} we know that their intersection point $q$ lies within $R(\ell)$ of the center $p$ of $D_0$. We shall now reason in the universal cover, considering $D_0$ and the disk $D_1$ of center $p$ and radius $R(\ell)$. The geometric problem we are interested in is as follows. {\it How many (hyperbolic) lines, any two of which pairwise intersect in an angle of at least $\theta_\ell$, can intersect $D_0$?} For this consider the disk $D_2$, also of center $p$, but of radius $R(\ell)+R'$ for a given $R'>0$. Consider geodesics $\gamma_1$ and $\gamma_2$ which pass through $D_0$ and intersect in $q \in D_1$. From $q$, we consider one of the four angular sectors of angle $\theta>0$ and the two rays of $\gamma_1$ and $\gamma_2$ which bound the sector. We set $q_1,q_2$ to be the intersection points of the rays of $\gamma_1$ and $\gamma_2$ with the boundary of $D_2$ as in figure \ref{fig:3disks}. Note that $d(q,q_1)$ and $d(q,q_2) > R'$ and that $\theta> \theta_0$. \begin{figure}[h] \leavevmode \SetLabels \L(.633*.166) $p$\\ \L(.53*.28) $q$\\ \L(.65*.30) $D_0$\\ \L(.65*.67) $D_1$\\ \L(.65*.91) $D_2$\\ \L(.33*.5) $q_1$\\ \L(.412*.813) $q_2$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/threedisks.pdf,width=6.0cm,angle=0}}} \vspace{-18pt} \end{center} \caption{The disks $D_0, D_1$ and $D_2$} \label{fig:3disks} \end{figure} Considering the triangle $q,q_1,q_2$, observe the following: the distance $d(q_1,q_2)$ is strictly greater than the distance one would compute if $$d(q,q_1)=d(q,q_2) = R' {\text{ and }} \theta=\theta_0.$$ Via hyperbolic trigonometry in the $(q,q_1,q_2)$ triangle: $$ d(q_1,q_2) > 2 {\,\rm arcsinh}\left( \sinh(R') \sin(\theta_\ell)\right). $$ We now consider a maximal set of lines $\gamma_1,\hdots,\gamma_k$ that cross $D$. The lines divide $\partial D_2$ into $2k$ arcs. As such, the cardinality of the set of lines is bounded by the length of $\partial D_2$ divided by twice the minimal distance between the intersection points of the lines with $D_2$. By the above estimates, we have $$ G(S) < \frac{ \pi \sinh(R(\ell)+R')}{2 {\,\rm arcsinh}(\sinh(R') \sinh(\theta_\ell))}. $$ Setting $R' := {\,\rm arcsinh}(1)$ this becomes $$ G(S) < \frac{\pi}{2} \frac{\sinh(R(\ell) +{\,\rm arcsinh}(1))}{\arcsin \frac{1}{2 \cosh(\ell/4)}}. $$ Note that $R(\ell)$ remains bounded as $\ell$ increases so we get an upper bound on $G(S)$ which has order of growth $\sim e^{\sfrac{\ell}{4}}$. \vspace{0.5cm} \noindent{\underline{The number of balls each systole crosses}} \noindent The last estimate we need is a lower bound on the number of balls $H(S)$ in our covering of $S$ that each systole necessarily crosses. The estimate we'll use is straightforward: to cover a geodesic segment of length $\ell$ with balls of radius $r(\ell)$ one requires at least $2\, \sfrac{\ell}{r(\ell)}$ balls. Thus $$ H(S)\geq \frac{2\, \ell}{r(\ell)} = \frac{2\, \ell}{{\,\rm arcsinh}\left(\frac{1}{2\sinh(\ell/4)}\right)}. $$ Here the order of growth is roughly $\ell \,e^{\,\sfrac{\ell}{4}}$. We can now conclude by using our estimates for $F(S),G(S)$ and $H(S)$. By our above estimates, the order of growth of the bound is $$\sim \text{Area}(S) \frac{e^{\,\sfrac{\ell}{2}}}{\ell}. $$ By a few elementary considerations, for $\ell \geq 2 {\,\rm arcsinh} 1$, we have the following effective inequality $$ {\rm Kiss}(S) < 100 \,(g-1) \frac{e^{\,\sfrac{\ell}{2}}}{\ell} $$ which concludes the proof. \end{proof} The proofs of Corollaries \ref{cor:subquad} and \ref{cor:linear} in the introduction follow from elementary estimates. \section{Non-hyperbolic surfaces with many systoles} In this section, we construct non-hyperbolic surfaces with many systoles. Recall the statement from the introduction. \begin{theorem}\label{thm:completegraph} There exist surfaces of genus $g>0$ with a number of systoles of order of growth at least $g^{\sfrac{3}{2}}$. \end{theorem} \begin{remark} The statement is stated in terms of order of growth for simplicity. The result we shall in fact prove is stronger, namely that there exist surfaces with at least $$ g\sqrt{48 g - 47} +15 g +\frac{1}{3} \sqrt{48 g - 47} - \frac{41}{3} > 6 g^{\sfrac{3}{2}} $$ number of systoles, but it is the order of growth we are really interested in. \end{remark} \begin{proof} The proof is by construction. The general idea of the proof is as follows: we'll begin with a complete graph $\Gamma_n$ (with $n>3$ for this construction to work) and imitate its geometry on a surface into which it is embedded. To give a rough idea of where the numbers are coming from, recall that one can embed a complete $n$-graph into a surface of genus $g\sim n^2$. On such a graph, the any nontrivial loop has length at least $3$, so the short loops are those of length exactly $3$ and there are $\left( \begin{array}{c} n \\ 3 \end{array} \right) \sim n^3 \sim g^{\sfrac{3}{2}}$ of those. The idea is then to construct an (almost) isometric embedding so that ``many" of the nontrivial short cycles of the graph remain nontrivial on the surface, and so that there aren't any others of smaller length. Then the surface will require some tweaking so that the lengths of the short nontrivial curves are all exactly the same. Our first observation is topological: consider a minimal genus surface $S_{g_n}$ of genus $g_n$ into which $\Gamma_n$ is embedded. By Ringel's and Youngs' Theorem \cite{riyo68} we have $$ g_n < \frac{(n-3)(n-4)}{12} +1. $$ This embedding will serve as our blueprint for the construction of the surface. In fact, we will begin building the geometry of our goal surface in a neighborhood of the embedded graph before describing the metric on the full surface. Geometrically we will be thinking of our graph as a metric graph with all edges of length $1$, and obtained by pasting $n$-pods together (each $n$-pod consists of a vertex with $n$ half-edges of length $\sfrac{1}{2}$). We begin by embedding each $n$-pod in the euclidean plane as follows. The vertex is on the origin and the half-edges are euclidean segments of length $\sfrac{1}{2}$ equally distributed around the origin. \begin{figure}[h] \leavevmode \SetLabels \L(.23*.4) $\;$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/eucnpod1.pdf,width=4.0cm,angle=0}\hspace{1.5cm} \epsfig{file =Figures/eucnpod2.pdf,width=4.0cm,angle=0}}} \vspace{-18pt} \end{center} \caption{Constructing a ribbon $n$-pod} \label{fig:eucnpods} \end{figure} We now fix a small $\varepsilon>0$ and consider a {\it ribbon type $n$-pod} of width $\varepsilon>0$ around our embedded $n$-pod. The ends of the ribbons are flat as is illustrated in figure \ref{fig:eucnpods}. Formally: the ribbon type $n$-pod is the intersection of the closed $\varepsilon>0$ neighborhood of the ribbon and a collection of closed half-planes defined as follows. For each euclidean ray supporting a half-edge, we consider the lower half space delimited by the unique line perpendicular to the ray and distance $\frac{1}{2}$ from the origin. The flat ends of each individual ribbon are euclidean segments of length $2 \varepsilon$. Two half ribbons can be glued geometrically in the natural way to (locally) obtain a picture of a surface with boundary (see figure \ref{fig:ribbons}). \begin{figure}[h] \leavevmode \SetLabels \L(.23*.4) $\;$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/ribbons.pdf,height=3.0cm,angle=0}\hspace{.5cm} \epsfig{file =Figures/subsurface.pdf,height=3.0cm,angle=0}}} \vspace{-18pt} \end{center} \caption{The construction of $S_{\varepsilon.g_n}$ and its embedding in $S_{g_n}$} \label{fig:ribbons} \end{figure} Going back to the pairings induced by the original embedding, we perform this individual gluing on each of the pairings, and the result is a surface with boundary. Observe that the surface is homeomorphic to a closed neighborhood of the graph embedded in the surface. We denote this surface $S^\varepsilon_{g_n}$. We shall observe a few things about this surface. First of all, thinking of $S^\varepsilon_{g_n}$ as a topological subsurface of $S_{g_n}$, observe that all the boundary curves of the embedded subsurface $S^\varepsilon_{g_n}$ are trivial (otherwise the embedding would not have minimal genus). Via this construction, there is a natural projection (at the level of homotopy classes) from cycles in $\Gamma_n$ (and curves in $S^\varepsilon_{g_n}$) to curves in $S_{g_n}$. We are interested in minimal length cycles of $\Gamma_n$ (those of length $3$) that project to nontrivial and homotopically distinct curves on $S_{g_n}$. We now need to count them. Consider a cycle on $\Gamma_n$ and consider it on $S^\varepsilon_{g_n}$ via our original embedding. As it passes through a vertex, it enters through one half edge, exits through another and separates the ribbon type $n$-pod into $2$ pieces separating the remaining half-edges into two sets (one of which can be empty). This is illustrated in figure \ref{fig:localcurve}. \begin{figure}[h] \leavevmode \SetLabels \L(.23*.4) $\;$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/localcurve.pdf,width=8.5cm,angle=0}}} \vspace{-18pt} \end{center} \caption{A local picture of a short cycle of $\Gamma_n$ in $S_{g_n}$} \label{fig:localcurve} \end{figure} Consider the subset of all length $3$ cycles which contain a vertex on which the cycle separates the other half-edges of the associated $n$-pod into two non-empty sets. Now we claim that any such cycle projects to a {\it nontrivial} curve on $S_{g_n}$ and that any two such cycles are non-isotopic. To see this observe that any such cycle intersects another such cycle in exactly one point. Via the bigon criterion, both of these cycles are essential and in fact homologically nontrivial. Furthermore, observe that any two distinct such cycles must intersect at least one other such cycle differently, and as such are homologically, and thus non-isotopic. We can also remark that any two such cycles intersect at most once. {\it Remark.} Although this doesn't play a role in what follows (because we are concerned with lower bounds), the cycles described above constitute the full set of short cycles of $\Gamma_n$ that project to pairwise distinct nontrivial curves on $S_{g_n}$. To see this, observe that if a short cycle fails to be in this set, then via our embedding into $S_{g_n}$, it is freely homotopic to a curve that is disjoint from the embedding of $\Gamma_n$. If it were to be nontrivial on $S_{g_n}$, then by cutting along this curve, one obtains an embedding of $\Gamma_n$ into a surface a surface of smaller genus than $S_{g_n}$, a contradiction. \begin{figure}[h] \leavevmode \SetLabels \L(.23*.4) $\;$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/trivialcycle.pdf,width=5.5cm,angle=0}}} \vspace{-18pt} \end{center} \caption{A nontrivial cycle in $\Gamma_n$ which projects to a trivial curve in $S_{g_n}$} \label{fig:orientedpants} \end{figure} We can now count these cycles. Note that any two edges of a short cycle determine the cycle completely as this determines the $3$ vertices of the cycle. Around each vertex there are $\frac{n(n-3)}{2}$ choices of pairs of distinct edges joined in the vertex that separate the remaining edges of the vertex into two non-empty sets. We can make this number of choices for each vertex, and each cycle will have been counted (at most) $3$ times. The total number of cycles $N$ is thus at least $$ N\geq \frac{n^2(n-3)}{6}. $$ These cycles determine a fixed set of (free) homotopy classes of curves on $S_{g_n}$ and it is the set we are to going to find a geometric structure for which they are all represented by systoles. We denote this set of free homotopy classes by $\mathcal H$. We now return to our surface with boundary $S^\varepsilon_{g_n}$. Each of the cycles on this surface admits a minimal length geodesic in its free homotopy/isotopy class. Often one refers to the minimal length of an isotopy class as the length of an isotopic geodesic of minimal length. {\it Claim.} Each of the minimal length geodesics has length $\in [3-6\varepsilon, 3]$. {\it Proof of claim.} The upper bound is obvious as each free homotopy class is represented by a cycle of length $3$, passing through $3$ vertices $v_1,v_2,v_3$. For the lower bound, observe that the distance between two of the vertices on $S^\varepsilon_{g_n}$ is exactly $1$. Now consider points $p_1,p_2,p_3$ along a minimal geodesic $\gamma$ that are of minimal distance to $v_1,v_2,v_3$. By construction, these distances are all less than $\varepsilon$ (any curve freely homotopic to the cycle must traverse each of the three ``corridors"). Consider the three subarcs of $\gamma$ between the points $p_1,p_2,p_3$, and denote their lengths $\ell_1, \ell_2, \ell_3$. Now by concatenating the paths, and using the fact that the vertices are at distance $1$ from each other, we see that $$ 1 \leq \ell_k + 2 \varepsilon $$ for $k=1,2,3$. It follows that $$ \ell(\gamma)= \ell_1+\ell_2+\ell_3 \geq 3 - 6\varepsilon. $$ This proves the claim.\\ Our next objective is to modify the metric so that the curves in $\mathcal H$ all have length exactly $3$. We begin by choosing, one for each of our isotopy classes, a minimal length geodesic. If necessary, we can paste arbitrarily thin euclidean cylinders on the boundary of the surface, to ensure that all minimal geodesics we have chosen do not touch the boundary. Observe that any {\it other} nontrivial isotopy class on the surface has minimal length at least $4$ (minus something close to $0$ when $\varepsilon$ is small). Via \cite{frhasc82}, any two minimal geodesics intersect minimally among all representatives in their respective free homotopy classes. In particular, any two distinct and freely homotopic minimal length geodesics are disjoint. It follows that there exists a $\varepsilon'>0$ (with $\varepsilon'<\varepsilon$) such that the closed $\varepsilon'>0$ neighborhood of one our chosen minimal length geodesics only (completely) contains other minimal length geodesics that are in the same free homotopy class. As such, we can consider an $\varepsilon'>0$ neighborhood of the full set of chosen minimal length geodesics for which we are sure that any two minimal length geodesics have the same ``combinatorics" if and only if they are isotopic. By {\it combinatorics} we mean that they pass through the same $\varepsilon'$ corridors (strips of total length $\sim 1$ and of width $\varepsilon'$). We consider this new surface with boundary $S^{\varepsilon'}_{g_n}$. Observe that the number of boundary components may have increased, but the genus remains the same (topologically we've just cut disjoint disks out of $S^\varepsilon_{g_n}$ to obtain $S^{\varepsilon'}_{g_n}$). We can now proceed to the modification of the metric on $S^{\varepsilon'}_{g_n}$. Consider the set of {\it maximal length} geodesics among our set of minimal length geodesics representing $\mathcal H$. To illustrate the construction, let us suppose that there is only one curve of maximal length, say $\gamma$, of length $L \leq 3$. We can insert a euclidean cylinder of width $\omega$ and with both boundary lengths $\ell(\gamma)$ along $\gamma$ while respecting how the two copies of $\gamma$ were pasted together. (We will call this {\it grafting} along $\gamma$, although the term is generally used for this type of construction along geodesics of hyperbolic surfaces.) Observe that this process {\it increases} the minimal length of any isotopy class of a curve that crosses $\gamma$. To see this, suppose this was not the case, i.e., there is a curve $\delta$ on the modified surface of shorter length than a given minimal length geodesic on the original surface. Consider the subarc $c$ of $\delta$ that passes through the cylinder. Consider the projection $c'$ of $c$ to the basis of the cylinder as in figure \ref{fig:graft}. By construction $\ell(c)>\ell(c')$. Now we ``project" $\delta$ to the original surface, by concatenating the image of $c'$ and the arc of $\delta$ that was disjoint from the cylinder to obtain a curve $\tilde{\delta}$ on the original surface of strictly shorter length than $\ell(\delta)$, a contradiction. \begin{figure}[h] \leavevmode \SetLabels \L(.8*.38) $\gamma$\\ \L(.683*.321) $c'$\\ \L(.281*.235) $c'$\\ \L(.34*.72) $\delta$\\ \L(.279*.421) $c$\\ \L(.743*.63) $\tilde{\delta}$\\ \endSetLabels \begin{center} \AffixLabels{\centerline{\epsfig{file =Figures/graft1.pdf,width=5.0cm,angle=0}\hspace{1.0cm} \epsfig{file =Figures/graft2.pdf,width=5.0cm,angle=0}}} \vspace{-18pt} \end{center} \caption{Grafting along $\gamma$} \label{fig:graft} \end{figure} Also observe that via grafting along $\gamma$, the minimal length of the isotopy class of $\gamma$ does not change. In fact, we obtain a cylinder of minimal length geodesics parallel to the two copies of $\gamma$. Also observe, that in light of the fact that minimal geodesics intersect minimally, the minimal length of {\it any} isotopy class that does not intersect $\gamma$ remains unchanged during this process. As $\ell(\gamma)\leq 3$, and any curve going through the inserted cylinder has length strictly less than the width of the cylinder, there exists a width $\omega_0< 3$ for which there is at least one other isotopy class among our selected classes with minimal length exactly $\ell(\gamma)$. We now imitate this construction, and repeat it iteratively, but for more than one curve. Specifically, consider a set $\gamma_1,\hdots,\gamma_k$ of curves with maximal minimal length among our chosen classes on a surface with boundary $S^i$, obtained after $i$ metric modifications of $S^{\varepsilon'}_{g_n}$. We consider an $\tilde{\varepsilon}$ neighborhood around them, again as in the original construction of $S^{\varepsilon'}_{g_n}$. This gives us a sub-surface of $S^i$ with boundary curves $\delta_1,\hdots,\delta_b$. We insert a euclidean cylinder of width $\omega$ along each of the $\delta_k$. As above this process will make any curve that intersects at least of one of the $\gamma_k$ increase. And again, if there is at least one free homotopy class in our set of chosen classes that intersects of these curve, because the length of all the $\gamma_k$ is less or equal to $3$, there exists an $\omega_0$ for which a new homotopy class now has minimal length equal to $L$. We now observe the following: between any two free homotopy classes $h$ and $\tilde{h}$ in our chosen set, there exist a sequence of $h^k\in \mathcal H$ with $h^0=h$, $h^p=\tilde{h}$ and ${\mathit i} (h^i,h^{i+1})=1$. In particular, this guarantees that the above process leaves no curve in $\mathcal H$ isolated and finishes in at most $N$ steps where $N$ is the cardinality of $\mathcal H$. We can now summarize the result of the construction: via a finite number of insertions of cylinders into a surface with closed boundary, we now have a surface $\tilde{S}$ of genus $g_n$, also with closed boundary which has $N$ distinct free homotopy classes of curves of minimal length $L$ and where all other (non-peripheral) curves have length at least $\sim 4$. We now complete the description of the metric by using a well-known trick: for each of the boundary curves $\delta \subset \partial \tilde{S}$, we glue a round hemisphere of equator length $\ell(\delta)$. On the resulting surface $S$, it is easy to see that {\it any} minimal geodesic will be completely disjoint from any of the hemispheres and will lie completely in the subsurface $\tilde{S}$. In particular, the systoles we constructed in $\tilde{S}$ are systoles of $S$. As $n\sim \sqrt{g_n}$, we have $N\sim g_n^{\sfrac{3}{2}}$ and the resulting surface has the desired properties. \end{proof} \begin{remark}\label{rem:gint} There is another construction of a surface, much simpler, that also underlies the difference between hyperbolic and non-hyperbolic surfaces. Consider a single ribbon $n$-pod (for some small $\varepsilon>0$ with $n=4m$, and pair opposite edges in the obvious way. The result is a surface of genus $g=m= \sfrac{n}{4}$ with one boundary component. As above, the curves from the original $n$-pod are of length exactly $1$, and because $n$ is even, they are smooth curves on the surfaces. These curves are in fact systoles: any curve with more complicated combinatorics has length at least $\sim 2$ and any curve that stays in a single ribbon has length at least $1$. (Each of these curves has a family of parallel isotopic systoles as well.) As above, we glue a euclidean hemisphere to the single boundary curve of the surface. The interest in this construction is that in the center vertex of the embedded $n$-pod, there are exactly $n= 4g$ non-isotopic systoles that intersect in a single point. This is again in striking difference to what is possible for hyperbolic surfaces in view of Lemma \ref{lem:angle}. \end{remark} \addcontentsline{toc}{section}{References} \bibliographystyle{Hugo} \def$'${$'$}
\section{Introduction} Oeljeklaus-Toma manifolds (defined in \cite{_Oeljeklaus_Toma_}) are compact complex manifolds that are a generalization of Inoue surfaces (defined in \cite{_Inoue_}). Let us describe them in detail. \subsection{Oeljeklaus-Toma manifolds} Let $K$ be a number field (i.e. a finite extension of $\mathbb Q$), $s>0$ be the number of its real embeddings and $2t>0$ be the number of its complex embeddings. One can easily prove that for each $s$ and $t$ there exists a field $K$ which has these numbers of real and complex embeddings (see e.g. \cite{_Oeljeklaus_Toma_}). \definition The \textit{ring of algebraic integers} $O_K$ is a subring of $K$ that consists of all roots of polynomials with integer coefficients which lie in $K$. \textit{Unit group} $O^*_K$ is the multiplicative subgroup of invertible elements of $O_K$. \hfill Let $m$ be $s+t$. Let $\sigma_1, \ldots, \sigma_s$ be real embeddings of the field $K$, $\sigma_{s+1}, \ldots, \sigma_{s+2t}$ be complex embeddings such that $\sigma_{s+i}$ and $\sigma_{s+t+i}$ are complex conjugate for each $i$ from $1$ to $t$. Now we can define a map $l: O_K^*\rightarrow \mathbb R^m$ where $l(u)=(\ln|\sigma_1(u)|, \ldots, \ln|\sigma_s(u)|, 2\ln|\sigma_{s+1}(u)|, \ldots, 2\ln|\sigma_m(u)|)$. Denote $O_K^{*, +}=\{a\in O_K^*:\sigma_i(a)>0, i=1, \ldots, s\}$. Let us consider following definitions: \hfill \definition A \textit{lattice} $\Lambda$ in $\mathbb R^n$ is a discrete additive subgroup such that $\Lambda\otimes\mathbb R=\mathbb R^n$. \definition \cite{_Oeljeklaus_Toma_} The group $U\subset O_K^{*, +}$ of rank $s$ is called \textit{admissible for the field} $K$ if the projection of $l(U)$ to the first $s$ components is a lattice in $\mathbb R^s$. Consider a linear space $L=\{x\in\mathbb R^m \mid \sum_{i=1}^m x_i=0\}$. The projection of $L\subset R^m$ to the first $s$ coordinates is surjective, because $s<m$. Using the Dirichlet unit theorem (see e.g. \cite{_Milne09_}) one can prove that $l(O_K^{*, +})$ is a full lattice in $L$. Therefore there exists a group $U$ that is admissible. \hfill Let $\mathbb H=\{z\in\mathbb C \mid \operatorname{im} z>0\}$. Let $U\subset O_K^{*, +}$ be a group which is admissible for $K$. The group $U$ acts on $O_K$ multiplicatively. This defines a structure of semidirect product $U':=U\ltimes O_K$. Define the action of $U'$ on $\mathbb H^s\times\mathbb C^t$ as follows. The element $u\in U$ acts on $\mathbb H^s\times \mathbb C^t$ mapping $(z_1, \ldots, z_m)$ to $(\sigma_1(u)z_1, \ldots, \sigma_m(u)z_m)$. Since $U$ lies in $O_K^{*, +}$, the action $U$ on the first $s$ coordinates preserves $\mathbb H$. The additive group $O_K$ acts on $\mathbb H^s\times \mathbb C^t$ by parallel translations: $a\in O_K$ is mapping $(z_1, \ldots, z_m)$ to $(\sigma_1(a)+z_1, \ldots, \sigma_m(a)+z_m)$. Since the first $s$ embeddings are real, this action preserves $\mathbb H$ in the first $s$ coordinates. One can see that $(u, a)\in U\ltimes O_K$ maps $(z_1, \ldots, z_m)$ to $(\sigma_1(u)z_1+\sigma_1(a), \ldots, \sigma_m(u)z_m+\sigma_m(a))$. One can easily show that this action is compatible with the group operation in the semidirect product. \hfill \definition An \textit{Oeljeklaus-Toma manifold} is the quotient of $\mathbb H^s\times\mathbb C^t$ by the action of the group $U\ltimes O_K$, which was defined above. This quotient exists because $U\ltimes O_K$ acts properly discontiniously on $\mathbb H^s\times\mathbb C^t$. Additionally $\mathbb H^s\times\mathbb C^t/U\ltimes O_K$ is a compact complex manifold. To prove it, let $U$ be admissible for $K$. The quotient $\mathbb H^s\times\mathbb C^t/O_K$ is obviously diffeomorphic to the trivial toric bundle $(\mathbb R_{>0})^s\times (S^1)^n$. The group $U$ acts properly discontinuously on the base $(\mathbb R_{>0})^s$. Therefore it acts properly discontinuously on $\mathbb H^s\times\mathbb C^t/O_K$. Also, the groups $U$ and $O_K$ act holomorphically on $\mathbb H^s\times\mathbb C^t$. Therefore the quotient has a holomorphic structure. \section{Curves on the Oeljeklaus-Toma manifolds} In this section we shall prove that there are no complex curves on the Oeljeklaus-Toma manifolds, just as on Inoue surfaces of type $S_M$ (see \cite{_Inoue_}). \subsection{The exact semipositive (1,1)-form on the Oeljeklaus-Toma manifold} The $(1,1)$-form we will be using was previously introduced in the paper \textit{Subvarieties in Oeljeklaus-Toma manifolds} by Ornea and Verbitsky. Authors use this form to prove that Oeljeklaus-Toma manifolds with $t=1$ (that means that the corresponding number field has only two complex embeddings) do not contain any submanifolds. Our result works for all Oeljeklaus-Toma manifolds, but we consider only curves instead of submanifolds of any dimension. Later on we will explain how our method differs from the one in \cite{_Ornea_Verbitsky_} and how this implies the difference between our results. Define the notion of $(1,1)$-form. Let $M$ be a smooth complex manifold, $z_1, \ldots, z_n$ --- local complex coordinates in the open neighborhood of the point $y\in M$. \definition A \textit{$(1,1)$-form} on a complex manifold $M$ is a 2-form $\omega$, such that $\omega(Iu,v)= -\omega(u,Iv)=\sqrt{-1}\omega(u,v)$ for each $u,v\in T_yM$, where $I$ is the almost complex structure on $M$. \definition A $(1, 1)$-form $\omega$ on a complex manifold $M$ is \textit{semipositive} if $\omega(u, Iu)\geqslant 0$ for each tangent vector $u\in T_yM$. \hfill As in \cite{_Ornea_Verbitsky_}, we consider a certain semipositive $(1, 1)$-form on the Oeljeklaus-Toma manifold $M=\mathbb H^s\times\mathbb C^t/(U\ltimes O_K)$. We introduce a $(1, 1)$-form $\widetilde\omega$ on $\widetilde M=\mathbb H^s\times\mathbb C^t$ which is preserved by the action of the group $\Gamma=(U\ltimes O_K)$ and since then it would be a $(1,1)$-form on $M$. Let $(z_1, \ldots, z_m)$ be complex coordinates on $\widetilde M$. Define $\varphi(z)=\Pi_{i=1}^s \operatorname{im}(z_i)^{-1}$. Since the first $s$ components of $\widetilde M$ correspond to upper half-planes $\mathbb H\subset\mathbb C$, this function is positive on $\widetilde M$. \hfill Let us now consider the form $\widetilde\omega=\sqrt{-1}\partial\bar\partial\log\varphi$. Using standard coordinates on $\widetilde M$ one can write this form as $\widetilde\omega=\sqrt{-1}\sum_{i=1}^s\frac{dz_i\wedge d\bar z_i}{4(\operatorname{im} z_i)^2}$. Therefore $\widetilde\omega$ is a semipositive $(1,1)$-form on $\widetilde M$. Let us show that this form is $\Gamma$-invariant. \hfill The group $\Gamma$ is a semidirect product of the additive group $O_K$ and the multiplicative group $U$ The additive group acts on the first $s$ components of $\widetilde M$ (which correspond to upper half-planes $\mathbb H\subset\mathbb C$) by translations along the real line. Therefore it does not change $\operatorname{im} z_i$ for $i=1\ldots s$. Hence the function $\log\varphi$ is preserved by the action of the additive component. The multiplicative component acts on the first $s$ coordinates of $\widetilde M$ by multiplying them by a real number (since the first $s$ embeddings of the number field $K$ are real). Then every $\operatorname{im} z_i$ is multiplied by a real number and so there is a real number added to $\log (\operatorname{im} z_i)$. Since $\log\varphi(z)=-\sum_{i=1}^s\log (\operatorname{im} z_i)$, there is a real number added to $\log\varphi$. The operator $\bar\partial$ is zero on the constants, so $\widetilde\omega=\sqrt{-1}\partial\bar\partial\log\varphi$ is preserved by action of the group $\Gamma$. Since the $(1,1)$-form $\widetilde\omega$ is $\Gamma$-invariant it is the pullback of $(1,1)$-form $\omega$ on the Oeljeklaus-Toma manifold $M=\widetilde M/\Gamma$. Let us now show that the form $\widetilde\omega$ is exact on $\widetilde M$. For that we define the operator $d^c$. \definition Define the \textit{twisted differential} $d^c=I^{-1}dI$ where $d$ is a De Rham differential and $I$ is the almost complex structure. Since $dd^c=2\sqrt{-1}\partial\bar\partial$ (see \cite{_Griffits-Harris_}), one can see that $\widetilde\omega=\sqrt{-1}\partial\bar\partial\log\varphi=\frac{1}{2}dd^c\log\varphi$ and so $\widetilde\omega$ is exact as a form on $\widetilde M$. Also since the operator $d^c$ vanishes on constants the form $d^c\log\varphi$ is $\Gamma$-invariant, so $\omega$ is exact on $M$. \subsection{The $(1, 1)$-form $\omega$ and curves on the Oeljeklaus-Toma manifold} Since the form $\omega$ on the manifold $M$ is semipositive, its integral on any complex curve $C\subset M$ is nonnegative. The form $\omega$ is exact. Hence Stokes' theorem implies that its integral on any complex curve vanishes. Therefore if $C\subset M$ is a closed complex curve, $\omega$ vanishes on it. To find out on which curves $\omega$ vanishes, let us define the zero foliation of the form $\omega$. \hfill \definition An \textit{involutive distribution (or foliation)} on $M$ is a subbundle $B\subset TM$ of the tangent bundle that is closed under the Lie bracket: $[B, B]\subset B$. \definition A \textit{leaf of a foliation} $B$ is a connected submanifold of $M$ such that its dimension is equal to $\dim B$ and that is tangent to $B$ at every point. \theorem (Frobenius) Let $B\subset TM$ be an involutive distribution. Then for each point of the manifold $M$, there is exactly one leaf of this distribution that contains this point (see e.g. \cite{_Boothby_} Section IV. 8. Frobenius Theorem). \theorem Let $N\subset M$ be a connected submanifold such that its tangent space at every point lies in a foliation $F\subset TM$. Then $N$ lies in a leaf of the foliation $F$ (see e.g. \cite{_Boothby_} Section IV. 8. Theorem 8.5). \hfill \definition The \textit{zero foliation} of a semipositive $(1,1)$-form $\omega$ on $M$ is the subundle of $TM$ that consists of tangent vectors $u\in T_yM$ such that $\omega(u, Iu)=0$, where $I$ is the almost complex structure on $M$. \hfill Consider the zero foliation of $\widetilde\omega$ on $\widetilde M$. The form $\widetilde\omega$ is strictly positive on each vector $v=(z_1,\ldots,z_m)$ such that at least one of $z_i$ for $i=1,\ldots, s$ is nonzero. Such a vector cannot be tangent to a leaf of the zero foliation. Therefore on each leaf of the zero foliation of the form $\widetilde\omega$ the first $s$ coordinates are constant. Hence a leaf of the zero foliation of $\widetilde\omega$ on $\widetilde M$ is isomorphic to $\mathbb C^t$. \hfill Let us now consider the zero foliation of $\omega$ on $M$. We show that the non-trivial image of the action of $\Gamma$ on any leaf $L$ of the zero foliation of the form $\widetilde\omega$ does not intersect with $L$. One can see that $L$ is $(z_1, \ldots, z_s)\times\mathbb C^t$ for some fixed $(z_1, \ldots, z_s)$. Therefore, for any $\gamma\in\Gamma$ such that $L\cap \gamma(L)\neq\emptyset$, the first $s$ coordinates of the points in $L$ coincide with the first $s$ coordinates of the points in $\gamma(L)$. Then for such $\gamma$ have the following system of equations: $$\sigma_i(u)z_i+\sigma_i(a)=z_i,\quad i=1\ldots s,$$ where $\gamma=(u, a)$. These equations imply that $z_i=\frac{\sigma_i(a)}{1-\sigma_i(u)}$. Therefore $z_i$ are real but $\mathbb H$ does not have real elements. We showed that $L\cap \gamma(L)=\emptyset$ for every $\gamma\neq 1$ in $\Gamma$. Since $\omega$ vanishes on each compact curve $C\subset M$, each curve is contained in some leaf of the zero foliation of $\omega$. Since $\widetilde\omega$ is $\Gamma$-invariant, each leaf of the zero foliation of $\omega$ on $M$ is isomorphic to a component of the leaf of the zero foliation of $\widetilde\omega$ on $\widetilde M$. Therefore, it is isomorphic to $\mathbb C^t$. And $\mathbb C^t$ does not contain any compact complex submanifolds. We proved the following theorem: \hfill \theorem There are no compact complex curves on the Oeljeklaus-Toma manifolds. \section{Closing remarks} Let us now briefly explain the connection between our work and \cite{_Ornea_Verbitsky_}. As in \cite{_Ornea_Verbitsky_} we use the zero foliation of a certain $(1,1)$-form. The leaves of this zero foliation are $t$-dimensional complex manifolds. In \cite{_Ornea_Verbitsky_} authors consider $t=1$, and submanifolds of any dimension. We consider any $t\geq 1$, and submanifolds of dimension 1. In both works one uses the semipositivity of the $(1,1)$-form to prove that a submanifold and a leaf of the zero foliation, which contains a point in this submanifold, could not intersect transversely. In both cases, one of the manifolds is of dimension one. Non-transversality implies that it is contained in the second one. Authors of \cite{_Ornea_Verbitsky_} use the fact that the leaves of the zero foliation are Zariski dense, and therefore could not be contained in any submanifold. In our work we prove that each leaf is isomorphic to $\mathbb C^t$ and therefore could not contain any submanifolds. The arguments used in our work and \cite{_Ornea_Verbitsky_} could not be applied to the case of higher dimensions. It seems obvious that there should be Oeljeklaus-Toma manifols which contain submanifolds --- other Oeljeklaus-Toma manifolds, corresponding to smaller number fields, but we do not have a formal proof. In \cite{_Oeljeklaus_Toma_} it was proved that Oeljeklaus-Toma manifolds do not admit non-trivial meromorphic functions. Therefore all the divisors on these manifolds are fixed in their linear systems (which are 0-dimensional). It is conjectured that Oeljeklaus-Toma manifolds admit no divisors. {\small
\section{Introduction} Metamagnetism is represented by a sudden increase in magnetization with increasing an applied field. In heavy fermion (HF) systems, CeRu$_{2}$Si$_{2}$ with the tetragonal ThCr$_{2}$Si$_{2}$ structure shows the metamagnetic behavior at about 7.7~T when a magnetic field ($H$) is applied parallel to the $c$-axis. Although various experiments as well as theoretical studies have been carried out\cite{Y.Hirose_JPCS_2011,Y.Onuki_JPSJ_2004}, the mechanism is still controversial. In order to understand the metamagnetic behavior in HF systems, it might be desired to investigate new metamagnetic compounds. The iron oxypnictide CeFePO is a related material of the iron-based superconductor LaFePO\cite{Y.Kamihara_JACS_2006,Y.Kamihara_JPCS_2008}. They possess the same two-dimensional layered structure, stacking Ce(La)O and FeP layers alternately. Br${\rm \ddot{u}}$ning $et~al.$ reported that CeFePO is a magnetically nonordered HF metal with a Sommerfeld coefficient $\gamma$ = 700~mJ/(mol K$^2$)\cite{E.Buning_PRL_2008}. At present, it is difficult to synthesize large single crystals of CeFePO for NMR measurements, but $^{31}$P-NMR can probe in-plane and out-of-plane magnetic response separately using $c$-axis aligned polycrystalline samples. Here we report novel metamagnetic behavior observed in $H\perp c$, and suggest that metamagnetism of Ce-based HF compounds is driven by Kondo-breakdown (drastic reduction of $c$-$f$ hybridization) as clarified experimentally. The polycrystalline CeFePO was synthesized by solid-state reaction\cite{Y.Kamihara_JPCS_2008}. Basic properties are consistent with the previous report\cite{E.Buning_PRL_2008}. To measure anisotropic magnetic properties of CeFePO, the samples were uniaxially aligned using a magnetic field\cite{B.L.Young_RSI_2002}. The polycrystalline CeFePO was ground into powder, mixed with stycast 1266, and was rotated in the external field of 1.4~T while the stycast cures. The $c$-axis of the sample is nicely aligned, which is shown from the angle dependence of $^{31}$P-NMR spectra (see in the inset of Fig.\ref{Fig.1}), and $^{31}$P-NMR measurement was performed on the sample. \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=8cm,clip]{Fig1.eps} \end{center} \caption{(Color online)(Main panel)$T$ dependence of $H$-swept NMR spectra at 31.7~MHz for $H \perp c$ (solid line) and $H \parallel c$ (broken line). $K = 0$ was determined by reference material H$_3$PO$_4$. (Inset)Angle dependence of $H$-swept NMR spectra at 31.7 MHz measured at 20 K. $\theta$ is the angle between magnetic field and $c$-axis. Solid line is corresponding to fitting line.} \label{Fig.1} \end{figure} Figure \ref{Fig.1} shows $H$-swept NMR spectra in $H \parallel c$ and $H \perp c$ obtained at 31.4~MHz and various temperatures ($T$). The resonance peak for $H \parallel c$ is almost $T$ independent, but the peak for $H \perp c$ shows the characteristic $T$ dependence originating from $\chi (T)$. The Knight shift $K_{\perp(\parallel)}$ was determined from the peak field of the $^{31}$P NMR spectrum obtained in $H$ perpendicular (parallel) to the $c$-axis. $K = 0$ was determined by reference material H$_3$PO$_4$. $K_{i}(T, H) (i = \perp {\rm and} \parallel)$, which is the measure of the local susceptibility at the nuclear site, is defined as, \begin{align} K_{i}(T, H_{\rm res}) = \left(\frac{H_0-H_{\rm res}}{H_{\rm res}}\right)_{\omega = \omega_0} \propto \frac{M_{i}(T,H_{\rm res})}{H_{\rm res}} \end{align} where $H_{\rm res}$ are magnetic fields at resonance peaks, $H_0$ and $\omega_0$ are the resonance field and frequency of bare $^{31}$P nucleus and have the relation of $\omega_0 = \gamma_n H_0$ with gyromagnetic ratio $\gamma_n$, and $M_{i} (T,H_{\rm res})$ is the magnetization under $H_{i,\rm res} (i = \perp {\rm and} \parallel)$ at $T$. $K_{\parallel}$ is almost independent of $T$ and $H$, whereas $K_{\perp}$ shows strong $T$ dependence originating from the Curie-Weiss behavior of $\chi (T)$ above 10~K as shown in Fig.\ref{Fig.2}(a). The anisotropic Knight shift suggests that static spin properties possess $XY$-type spin anisotropy. It should be noted that $K_{\perp}$ exhibits $H$ dependence below 4~K and above 2~T, indicative of a non-linear relation between $M_{\perp}$ and $H$. Using the hyperfine coupling constant $^{31}A_{\rm hf} = 0.2~{\rm T}/\mu_{\rm B}$, which is estimated from the plot between isotropic component of $K$ and $\chi (T)$ above 10~K (not shown), we can plot $M_{i}(H)$ against $H$ in Fig.\ref{Fig.2}(b). $M_{\perp}(H)$ becomes superlinear against $H$ at 0.1~K, which is the hallmarks of metamagnetism, whereas $M_{\parallel}(H)$ is linear up to 6.2~T, which is again highly anisotropic. \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=9cm,clip]{Fig2.eps} \end{center} \caption{(Color online)(a) $T$ dependence of the Knight shift determined at the peaks of $H \perp c$ and $H \parallel c$ spectra obtained at various $H$. Strong anisotropy of Knight shift suggests that static spin properties possess $XY$-type spin anisotropy. (b) $H$ dependence of magnetization $M_i(H)$ ($i = \perp {\rm and} \parallel$) using the relation of $M_i(H) = K_i(H) H_{\rm res}/ A_{\rm hf}$. Solid and broken lines are guide to eyes. $M_{\perp}(H)$ suddenly increases with increasing $H$ and deviates from linear relation in the field range of 3 - 5~T, which is a definition of a metamagnetic behavior, while such a behavior was not observed in $M_{\parallel}(H)$ up to 6.2~T and down to 1.5~K.} \label{Fig.2} \end{figure} Next, we focus on $T$ and $H$ dependence of low-energy spin dynamics probed with the nuclear spin-lattice relaxation rate ($1/T_1$). $1/T_1$ of $^{31}$P was measured at each resonance peak by the saturation-recovery method, and was uniquely determined by a single component in whole measured range. The inset of Fig. \ref{Fig.3} shows $T$ dependence of $1/T_1T$ at low field $\mu_0H \simeq$ 0.6~T parallel and perpendicular to the $c$-axis. Below 1.5~K, $1/T_1T$ as well as $K$ along both directions becomes constant, indicative of the formation of a Fermi-liquid (FL) state of heavy electrons. In general, $1/T_1$ probes spin fluctuations perpendicular to applied $H$, and thus $1/T_1$ in $H \parallel c$ and $H \perp c$ are described as, \begin{align} (1/T_1)_{H \parallel c} &= 2(\mu_0\gamma_n)^2\sum_{q}|H_{\perp}(q,\omega_{\rm res})|^2 \notag \\ &\propto 2A^2\sum_{q}|S_{\perp}(q,\omega \sim 0)|^2, \text{~and} \notag\\ (1/T_1)_{H \perp c} &= (\mu_0\gamma_n)^2\sum_{q}\left[|H_{c}(q,\omega_{\rm res})|^2+|H_{\perp}(q,\omega_{\rm res})|^2\right] \notag\\ &\propto A^2 \sum_{q}\left[|S_{\parallel}(q,\omega \sim 0)|^2 + |S_{\perp}(q, \omega \sim 0)|^2\right]. \label{eq.3} \end{align} Here $|X(\omega)|$ denotes the power spectral density of a time-dependent random variable $X(t)$, and $A$ is assumed to be $q$-independent due to the metallic state. From these equations, we can decompose spin fluctuations along each direction as shown in the main panel of Fig. \ref{Fig.3}. $\sum_{q}|S_{\perp}(q, \omega \sim 0)|^2$ is dominant at low $T$, since $(1/T_1T)_{H \parallel c}$ is almost twice larger than $(1/T_1T)_{H \perp c}$. This indicates that the spin dynamics also possess $XY$-type anisotropy. The $XY$-type spin fluctuations have the predominance of ferromagnetic (FM) correlations as inferred from the Korringa relation between $(1/T_1T)_{H\parallel c}$ and $K_{\perp}$ in low-$T$ FL state, which is consistent with the previous $^{31}$P-NMR result\cite{E.Buning_PRL_2008} and with the experimental facts that CeFePO is close to FM instability\cite{Y.Luo_PRB_2010,C.Krellner_PRB_2007}. \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=8cm,clip]{Fig3.eps} \end{center} \caption{(Color online) (Main panel)$T$ dependence of low energy spin fluctuations parallel and perpendicular to the $c$-axis at $\simeq$ 0.6~T evaluated with $1/T_1T$ measured in $H \perp c$ and $H \parallel c$ (See eq. \ref{eq.3}). The in-plane spin fluctuations are dominant at low $T$, suggesting that the spin dynamics also possess $XY$-type anisotropy. (Inset) $T$ dependence of $1/T_1T$ at 10.3~MHz ($\simeq$ 0.6~T) for $H \perp c$ and at 10.3~MHz ($\simeq$ 0.6~T), 31.7~MHz ($\simeq$ 1.8~T), and 107.2~MHz ($\simeq$ 6.2~T) for $H \parallel c$. $(1/T_1T)_{H \parallel c}$ is independent of $H$ up to 6.2~T.} \label{Fig.3} \end{figure} \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=9cm,clip]{Fig4.eps} \end{center} \caption{(Color online)$T$ dependence of $(1/T_1T)_{H \perp c}$ below 4~T (a), and above 4~T (b). (c): $H$ dependence of $(1/T_1T)_{H \perp c}$ at low $T$. $(1/T_1T)_{H \perp c}$ shows a distinct maximum at around $H_{\rm M}$, indicating the enhancement of $N(E_{F})$ related to metamagnetic anomaly. (d) $H$ - $T$ phase diagram defined by $T_{\rm max}$ where $(1/T_1T)_{H \perp c}$ shows a maximum. Non FL behavior characterized by continuous increase in $(1/T_1T)_{H \perp c}$ with decreasing $T$ (broken lines are shown in (b)) was observed in a narrow field region intervening between low-field paramagnetic (PM) state and high-field polarized PM state above 6~T.} \label{Fig.4} \end{figure} The evolution of the spin dynamics against $H$ was investigated for both directions. Figure \ref{Fig.4} shows $T$ dependence of $(1/T_1T)_{H\perp c}$ below 4~T (approximately the metamagnetic field, $H_{\rm M}$) (a) and above 4~T (b). Although $(1/T_1T)_{H\parallel c}$ does not depend on $H$ up to 6.2~T as shown in the inset of Fig.\ref{Fig.3}, $(1/T_1T)_{H\perp c}$ changes significantly by $H$ as shown in Figs.\ref{Fig.4} (a) and (b). $H$ dependence of $(1/T_1T)_{H \perp c}$ at 0.1 and 1.5~K is shown in Fig.\ref{Fig.4} (c). $(1/T_1T)_{H \perp c}$ shows a distinct maximum at $H_{\rm M}$, suggesting that the enhancement of the density of states (DOS) is related to metamagnetic behavior. If we assume that $1/T_1T \propto N(E_{\rm F})^2$, $N(E_{\rm F})$ at $H_{\rm M}$ is almost 1.5 times larger than $N(E_{\rm F})$ at 0~T. However, it is noteworthy that non FL behavior characterized by continuous increase in $(1/T_1T)_{H \perp c}$ with decreasing $T$ was observed down to 100~mK ($(1/T_1T)_{H \perp c} \sim -\log T$) in 4.3~T $<~\mu_0H~<$ 5~T intervening between low-field paramagnetic (PM) state and high-field polarized PM state above 6~T. In the high-field polarized state, $(1/T_1T)_{H \perp c}$ decreases with increasing $H$ and $(1/T_1T)_{H \perp c}$ at 14~T shows almost the same value as (LaCa)FePO [$\simeq$ 1.5 (sK)$^{-1}$] without 4$f$ electrons\cite{Y.Nakai_PRL_2008}. The $H$ variation of $(1/T_1T)_{H \perp c}$, reflecting the evolution of DOS at the Fermi level with $H$, strongly suggests the evolution of the Fermi surfaces (FSs) by $H$. It is noted that such a significant $H$ dependence was not reported in the previous specific-heat measurement\cite{E.Buning_PRL_2008}. This would be because the magnetic field is applied in the various angles against the $c$-axis, and suggests that the metamagnetic behavior would be observed when $H$ is exactly perpendicular to the $c$-axis. \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=8cm,clip]{Fig5.eps} \end{center} \caption{(Color online)Fermi surfaces (FSs) calculated with the $ab~initio$ band structure calculation as Ce-4$f$ electrons are itinerant (a), or localized (b), where the Fermi velocity is mapped on the FSs. (c) shows the FS colored by the $j_z$ character in the $j$ = 5/2 multiplet of 4$f$ orbitals, where $j$ is the total angular momentum. (d): the partial density of states. The Fermi level corresponds to 0 eV. } \label{Fig.5} \end{figure} To investigate such evolution of FSs by $H$, we performed the $ab~initio$ band-structure calculation in the paramagnetic state of CeFePO by using the WIEN2k package\cite{wien2k}. FSs in the low-field region are composed of itinerant Ce 4$f$ electrons as shown in Fig.\ref{Fig.5} (a). The large FS shows the characteristic neck structures around $X$-$R$ at the Brillouin zone boundary, at which boundary electrons have small Fermi velocity or heavy electron mass. The orbital character is dominated by the $j_z$ = $\pm$ 1/2 component in the $j$ = 5/2 multiplet of 4$f$ orbitals, as seen in Figs.\ref{Fig.5} (c) and (d) \cite{cap}. These features of Fermi surface imply that the low-field HF state possesses the small $q$ magnetic correlations and their in-plane component is much larger than the out-of-plane component, in good agreement with the experimental results. The band calculation also shows that applied field pinches off the neck FSs around $R$ and $X$ in order, that is, ``{\it the field-induced Lifshitz transition}'' appears. This is accompanied by a drastic change in DOS at the Fermi level, which can be a driving force for the metamagnetic transition and non-FL behavior around $H_{\rm M}$\cite{T.Senthil_PRB_2008,Y.Yamaji_JPSJ_2006}. With applying higher $H$, 4$f$ electrons are localized, and thus the FSs become small. The resultant FSs are shown in Fig.\ref{Fig.5} (b), familiar in the iron-based superconductors\cite{S.Lebegue_PRB_2007}. Thus the scenario of the field-induced Lifshitz transition can link with the nature in the metamagnetic transition in CeFePO. Here, we compare the present results with CeRu$_2$Si$_2$, one of the most well-known metamagnetic compounds. Although both compounds show similar $T$ and $H$ dependence of Knight shift and $1/T_1T$ in $H$ parallel to the magnetic easy axis\cite{K.Ishida_PRB_1998}, as well as the similar $H$ - $T$ phase diagram defined by $T_{\rm max}$\cite{Y.Aoki_JMMM_1998} as shown in Fig.\ref{Fig.4} (d), magnetic properties are quite different. For example, magnetic easy axis is different between CeFePO and CeRu$_2$Si$_2$: CeFePO possesses two dimensional $XY$-type spin anisotropy, whereas CeRu$_2$Si$_2$ possesses Ising-type spin anisotropy\cite{P.Haen_JLTP_1987}. As a result, the ground states of the crystal-field level are different and their metamagnetic behavior is observed in different directions. In addition, dominant magnetic fluctuations in CeFePO differ from those in CeRu$_2$Si$_2$. It is reported that CeRu$_2$Si$_2$ is located close to antiferromagnetic instability accompanied with FM fluctuations\cite{J.Flouquet_LTP_2010,Y.Kitaoka_JPSJ_1985}. \begin{figure}[tb] \vspace*{-10pt} \begin{center} \includegraphics[width=7cm,clip]{Fig6.eps} \end{center} \caption{(Color online)(a) Values of metamagnetic fields are plotted against $T_{\rm max}$ determined from a maximum in static susceptibility (a) or inverse of Sommerfeld coefficient $\gamma$ at $H$ = 0 (b) in CeFePO, CeCu$_6$ and CeRu$_2$Si$_2$ with pressurized and La-doped system. Broken lines are guide to eyes. A linear relation holds between two quantities, although three compounds possess quite different crystal structures and magnetic properties, indicating that $H_{\rm M}$ is linked with the local Kondo singlet energy $T_{\rm max}$.} \label{Fig.6} \end{figure} Figure \ref{Fig.6} shows the relationship between the metamagnetic field $H_{\rm M}$ and the temperature where the bulk susceptibility shows a maximum $T_{\rm max}$ or inverse of Sommerfeld coefficient $\gamma$ at $H$ = 0 for CeCu$_6$, CeFePO, and CeRu$_2$Si$_2$ with doped and pressurized systems\cite{E.Buning_PRL_2008,P.Haen_JLTP_1987,T.Fujita_JMMM_1985,A.Schroder_JMMM_1992,J.M.Mignot_JMMM_1988,R.A.Fisher_LTP_1991}. It deserves to note that the linear relation holds between the two quantities, notwithstanding that the three compounds possess totally different crystal structures and magnetic properties. Since $T_{\rm max}$ is regarded as a Kondo temperature $T_{\rm K}$ and approximately, the relation of $\gamma T_{\rm K} = const.$ holds in the HF state, these facts indicate that $H_{\rm M}$ is merely related to the local Kondo singlet energy $T_{\rm max}$ and is not linked with the magnetic fluctuations originating from the intersite coupling between neighboring Ce ions and/or the nesting between the ``large'' FS. The experimental facts that $H_{\rm M}$ is linearly proportional to $T_{\rm K}$ in Fig.\ref{Fig.6} strongly suggests that the metamagnetic behavior is linked with the Kondo breakdown\cite{A.Hackl_PRB_2008}. Therefore, in the Ce-based metamagnets, the Kondo breakdown and the Fermi-surface instability accompanied by the drastic change of DOS occur almost simultaneously around $H_{\rm M}$, which can induce novel non-FL behavior. In summary, we performed $^{31}$P-NMR in the uniaxially-aligned CeFePO and found that CeFePO possesses two dimensional $XY$-type FM fluctuations, and shows metamagnetic behavior when $H$ is applied to $H \perp c$ below 5~K, accompanied with non FL behavior around metamagnetic field $H_{\rm M} \simeq$ 4~T. As far as we know, this is a first example that the metamagnetic behavior occurs in a nonmagnetic Ce-based HF compound with the $XY$-type spin anisotropy. From the band calculation and the comparison with other Ce-based metamagnets, we claim that $H_{\rm M}$ is a magnetic field breaking the local Kondo singlet, which is determined with the intrasite coupling between Ce-4$f$ and conduction electrons, and that the FSs change drastically due to the Kondo breakdown. This work was partially supported by Kyoto Univ. LTM center, the "Heavy Electrons" Grant-in-Aid for Scientific Research on Innovative Areas (No. 20102006) from The Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of Japan, a Grant-in-Aid for the Global COE Program ``The Next Generation of Physics, Spun from Universality and Emergence'' from MEXT of Japan, a grant-in-aid for Scientific Research from Japan Society for Promotion of Science (JSPS), KAKENHI (S and A) (No. 20224008 and No. 23244075) and FIRST program from JSPS. One of the authors (SK) is financially supported by a JSPS Research Fellowship.
\section{Introduction} Cones of divisors play an essential role in describing the birational geometry of a smooth complex projective variety $X$. A key feature of these cones is their interplay with cones of curves via duality statements. The dual of the nef cone and the pseudo-effective cone of divisors were determined by \cite{kleiman66} and \cite{bdpp04} respectively. We consider the third cone commonly used in birational geometry: the movable cone of divisors. \begin{defn} Let $X$ be a smooth projective variety over $\mathbb{C}$. The movable cone $\overline{\mathrm{Mov}}^{1}(X) \subset N^{1}(X)$ is the closure of the cone generated by classes of effective Cartier divisors $L$ such that the base locus of $|L|$ has codimension at least $2$. We say a divisor is movable if its numerical class lies in $\overline{\mathrm{Mov}}^{1}(X)$. \end{defn} \begin{defn} Let $X$ be a smooth projective variety over $\mathbb{C}$. We say that an irreducible curve $C$ on $X$ is movable in codimension $1$, or a $\mathrm{mov}^{1}$-curve, if it deforms to cover a codimension $1$ subset of $X$. \end{defn} It is natural to guess that a divisor $L$ is movable if and only if it has non-negative intersection with every $\mathrm{mov}^{1}$-curve. This is false, as demonstrated by \cite{payne06} Example 1. Nevertheless, Debarre and Lazarsfeld have asked whether one can formulate a duality statement for movable divisors and $\mathrm{mov}^{1}$-curves. This has been accomplished for toric varieties in \cite{payne06} and for Mori Dream Spaces in \cite{choi10} by taking other birational models of $X$ into account. Our main theorem proves an analogous statement for all smooth varieties. Before stating this theorem, we need to analyze the behavior of the movable cone under birational transformations. Suppose that $\phi: Y \to X$ is a birational map of smooth projective varieties and that $L$ is a movable divisor on $X$. It is possible that $\phi^{*}L$ is not movable -- for example, some $\phi$-exceptional centers could be contained in the base locus of $L$. The following definition from \cite{nakayama04} allows us to quantify the loss in movability. \begin{defn} Let $X$ be a smooth projective variety over $\mathbb{C}$ and let $L$ be a pseudo-effective $\mathbb{R}$-divisor on $X$. Fix an ample divisor $A$ on $X$. For any prime divisor $\Gamma$ on $X$ we define \begin{equation*} \sigma_{\Gamma}(L) = \inf\{\mathrm{mult}_{\Gamma}(L') | L' \geq 0 \textrm{ and } L' \sim_{\mathbb{R}} L + \epsilon A \textrm{ for some } \epsilon > 0 \} \end{equation*} where $\sim_{\mathbb{R}}$ denotes $\mathbb{R}$-linear equivalence. As demonstrated by \cite{nakayama04} III.1.5 Lemma, $\sigma_{\Gamma}$ is independent of the choice of $A$. \end{defn} Suppose that $E$ is an exceptional divisor for a birational map $\phi: Y \to X$. The $\mathbb{R}$-divisor $\sigma_{E}(\phi^{*}L)E$ represents the ``extra contribution'' from $E$ to the non-movability of $\phi^{*}L$. By subtracting these contributions, we can understand the geometry of the original divisor $L$. \begin{defn} \label{movtrans} Let $X$ be a smooth projective variety over $\mathbb{C}$ and let $L$ be a pseudo-effective $\mathbb{R}$-divisor on $X$. Suppose that $\phi: Y \to X$ is a birational map from a smooth variety $Y$. The movable transform of $L$ on $Y$ is defined to be \begin{equation*} \phi^{-1}_{\mathrm{mov}}(L) := \phi^{*}L - \sum_{E \, \, \phi \mathrm{-exceptional}} \sigma_{E}(\phi^{*}L)E. \end{equation*} \end{defn} Note that the movable transform is not linear and is only defined for pseudo-effective divisors. We can now state our main theorem. \begin{thrm} \label{maintheorem} Let $X$ be a smooth projective variety over $\mathbb{C}$ and let $L$ be a pseudo-effective $\mathbb{R}$-divisor. $L$ is not movable if and only if there is a $\mathrm{mov}^{1}$-curve $C$ on $X$ and a birational morphism $\phi: Y \to X$ from a smooth variety $Y$ such that \begin{equation*} \phi^{-1}_{mov}(L) \cdot \widetilde{C} < 0 \end{equation*} where $\widetilde{C}$ is the strict transform of a generic deformation of $C$. \end{thrm} There does not seem to be an easy way to translate Theorem \ref{maintheorem} into a statement involving only intersections on $X$. This is a symptom of the fact that the natural operation on movable divisors is the push-forward and not the pull-back. The proof of Theorem \ref{maintheorem} is accomplished by reinterpreting the orthogonality theorem of \cite{bdpp04} and \cite{bfj09} using the techniques of \cite{lehmann10}. \begin{exmple} For surfaces Theorem \ref{maintheorem} reduces to the usual duality of the nef and pseudo-effective cones. \end{exmple} \begin{exmple} Suppose that $X$ is a smooth Mori dream space and $L$ is an $\mathbb{R}$-divisor on $X$. By running the $L$-MMP as in \cite{hk00}, we obtain a small modification $\phi: X \dashrightarrow X'$, a morphism $f: X \to Z$, and an ample $\mathbb{R}$-divisor $A$ on $Z$ such that \begin{equation*} \phi_{*}^{-1}L \equiv f^{*}A \end{equation*} where $\phi_{*}^{-1}$ denotes the strict transform. Let $W$ be a smooth variety admitting birational maps $\psi: W \to X$ and $\psi': W \to X'$. Using \cite{nakayama04} III.5.5 Proposition, one easily verifies that \begin{equation*} \psi^{-1}_{mov}(L) \equiv \psi'^{*}(\phi_{*}^{-1}L). \end{equation*} Thus Theorem \ref{maintheorem} implies the statements of \cite{payne06} and \cite{choi10}: for a smooth toric variety or Mori Dream Space $X$, a divisor class is movable iff its strict transform class on every $\mathbb{Q}$-factorial small modification $X'$ has non-negative intersection with every $\mathrm{mov}^{1}$-curve on $X'$. \end{exmple} \begin{exmple} Suppose that $X$ is a smooth projective variety with $K_{X}$ numerically trivial. \cite{choi10} explains how to apply techniques of the minimal model program to analyze $\overline{\mathrm{Mov}}^{1}(X)$. Just as before, a divisor class is movable if and only if its strict transform class on every $\mathbb{Q}$-factorial small modification has non-negative intersection with every $\mathrm{mov}^{1}$-curve. When $X$ is hyperk\"ahler, \cite{huybrechts03} and \cite{boucksom04} show that in fact it suffices to consider small modifications that are also smooth hyperk\"ahler varieties. More generally, \cite{choi10} shows that small modifications can detect certain regions of $\overline{\mathrm{Mov}}^{1}(X)$ by using the minimal model program. \end{exmple} We will also prove a slightly stronger version of Theorem \ref{maintheorem} that involves the non-nef locus $\mathbf{B}_{-}(L)$ of $L$ (which will be defined in Definition \ref{nonnefdef}). Although the non-nef locus represents the ``obstruction'' to the nefness of $L$, it is not true that $\mathbf{B}_{-}(L)$ is covered by curves $C$ with $L \cdot C < 0$. However, Proposition \ref{mainprop} formulates a birational version of this negativity using the movable transform. Finally, we will use Proposition \ref{mainprop} to understand $k$-movability for $k>1$. Define the $k$-movable cone of $X$ to be the closure of the cone in $N^{1}(X)$ generated by effective Cartier divisors whose base locus has codimension at least $k-1$. We say that a divisor is $k$-movable if its numerical class lies in the $k$-movable cone. Note that the $1$-movable cone is just $\overline{\mathrm{Mov}}^{1}(X)$. Debarre and Lazarsfeld have asked whether there is a duality between the $k$-movable cone of divisors and the closure of the cone of irreducible curves that deform to cover a codimension $k$ subset (for $0 < k < \dim X$). Corollary \ref{maincor} constructs a birational version of this duality. Again, this generalizes results for toric varieties in \cite{payne06} and for Mori dream spaces in \cite{choi11}. \section{Background} Throughout $X$ will denote a smooth projective variety over $\mathbb{C}$. We use the notations $\sim, \sim_{\mathbb{Q}}, \sim_{\mathbb{R}}, \equiv$ to denote respectively linear equivalence, $\mathbb{Q}$-linear equivalence, $\mathbb{R}$-linear equivalence, and numerical equivalence of $\mathbb{R}$-divisors. The volume of an $\mathbb{R}$-divisor $L$ is \begin{equation*} \mathrm{vol}_{X}(L) = \limsup_{m \to \infty} \frac{h^{0}(X,\lfloor mL \rfloor)}{m^{\dim X}}. \end{equation*} \subsection{Divisorial Zariski decomposition} Let $L$ be a pseudo-effective $\mathbb{R}$-divisor on a smooth projective variety $X$. Recall that for a prime divisor $\Gamma$ on $X$ we have defined \begin{equation*} \sigma_{\Gamma}(L) = \inf\{\mathrm{mult}_{\Gamma}(L') | L' \geq 0 \textrm{ and } L' \sim_{\mathbb{R}} L + \epsilon A \textrm{ for some } \epsilon > 0 \}. \end{equation*} where $A$ is any fixed ample divisor. \cite{nakayama04} III.1.11 Corollary shows that there are only finitely many prime divisors $\Gamma$ on $X$ with $\sigma_{\Gamma}(L) > 0$, allowing us to make the following definition. \begin{defn}[\cite{nakayama04} III.1.16 Definition] \label{zardecom} Let $L$ be a pseudo-effective $\mathbb{R}$-divisor on $X$. Define \begin{equation*} N_{\sigma}(L) = \sum \sigma_{E}(L) E \qquad \qquad P_{\sigma}(L) = L - N_{\sigma}(L) \end{equation*} The decomposition $L = N_{\sigma}(L) + P_{\sigma}(L)$ is called the \emph{divisorial Zariski decomposition} of $L$. \end{defn} Note that for a birational morphism $\phi: Y \to X$ we have $\phi_{mov}^{-1}(L) = P_{\sigma}(\phi^{*}L) + \phi_{*}^{-1}N_{\sigma}(L)$ where $\phi_{*}^{-1}$ denotes the strict transform. The divisorial Zariski decomposition is closely related to the non-nef locus of $L$. \begin{defn} \label{nonnefdef} Let $X$ be a smooth projective variety and let $L$ be a pseudo-effective $\mathbb{R}$-divisor on $X$. We define the $\mathbb{R}$-stable base locus of $L$ to be the subset of $X$ given by \begin{equation*} \mathbf{B}_{\mathbb{R}}(L) = \bigcup \{ \mathrm{Supp}(L') | L' \geq 0 \textrm{ and }L' \sim_{\mathbb{R}} L \}. \end{equation*} The non-nef locus of $L$ is then defined to be \begin{equation*} \mathbf{B}_{-}(L) = \bigcup_{A \textrm{ ample }\mathbb{R}\textrm{-divisor}} \mathbf{B}_{\mathbb{R}}(L+A). \end{equation*} \end{defn} The following proposition records the basic properties of the divisorial Zariski decomposition. \begin{prop}[\cite{nakayama04} III.1.14 Proposition, III.2.5 Lemma, V.1.3 Theorem] Let $X$ be a smooth projective variety and let $L$ be a pseudo-effective $\mathbb{R}$-divisor. \begin{enumerate} \item $P_{\sigma}(L)$ is a movable $\mathbb{R}$-divisor. In particular for any prime divisor $E$ the restriction $P_{\sigma}(L)|_{E}$ is pseudo-effective. \item If $\phi: Y \to X$ is a birational morphism of smooth varieties and $\Gamma$ is a prime divisor on $Y$ that is not $\phi$-exceptional, then $\sigma_{\Gamma}(\phi^{*}L) = \sigma_{\phi(\Gamma)}(L)$. \item The union of the codimension $1$ components of $\mathbf{B}_{-}(L)$ coincides with $\mathrm{Supp}(N_{\sigma}(L))$. \end{enumerate} \end{prop} \subsection{Numerical dimension and orthogonality} Given a pseudo-effective divisor $L$, the numerical dimension $\nu(L)$ of \cite{nakayama04} and \cite{bdpp04} is a numerical measure of the ``positivity'' of $L$. There is also a restricted variant $\nu_{X|V}(L)$ introduced in \cite{bfj09}; since the definition is somewhat involved, we will only refer to a special subcase using an alternate characterization from \cite{lehmann10}. \begin{defn} \label{numdimdef} Let $L$ be a pseudo-effective divisor on $X$. Fix a prime divisor $E$ on $X$ and choose $L' \equiv L$ whose support does not contain $E$. We say $\nu_{X|E}(L) = 0$ if \begin{equation*} \liminf_{\phi} \mathrm{vol}_{\widetilde{E}}(P_{\sigma}(\phi^{*}L')|_{\widetilde{E}}) = 0 \end{equation*} where $\phi: \widetilde{X} \to X$ varies over all birational maps and $\widetilde{E}$ denotes the strict transform of $E$. \end{defn} The connection with geometry is given by the following version of the orthogonality theorem of \cite{bdpp04} and \cite{bfj09}. \begin{thrm}[\cite{bfj09}, Theorem 4.15] Let $L$ be a pseudo-effective divisor. If a prime divisor $E \subset X$ is contained in $\mathrm{Supp}(N_{\sigma}(L))$ then $\nu_{X|E}(L) = 0$. \end{thrm} \begin{proof} Fix an ample divisor $A$. Choose an $\epsilon > 0$ sufficiently small so that $\mathrm{Supp}(N_{\sigma}(L)) = \mathrm{Supp}(N_{\sigma}(L + \epsilon A))$ and apply \cite{bfj09} Theorem 4.15. The comparison between the numerical dimension of \cite{bfj09} and Definition \ref{numdimdef} is given by \cite{lehmann10} Theorem 7.1. \end{proof} \section{Proof} \begin{proof}[Proof of Theorem \ref{maintheorem}:] Suppose that $L$ is not movable. Denote by $E$ a fixed divisorial component of $N_{\sigma}(L)$. Fix a sufficiently general ample divisor $A$ on $X$ and choose $\epsilon$ small enough so that $E$ is a component of $N_{\sigma}(L+\epsilon A)$. Applying the orthogonality theorem of \cite{bdpp04}, we see there is a birational map $\phi: Y \to X$ so that \begin{enumerate} \item $\widetilde{E}$ is smooth. \item $\mathrm{vol}_{\widetilde{E}}(P_{\sigma}(\phi^{*}(L+\epsilon A))|_{\widetilde{E}}) < \mathrm{vol}_{E}(A|_{E}) = \mathrm{vol}_{\widetilde{E}}(\phi^{*}A|_{\widetilde{E}})$. \item The strict transform of every component of $N_{\sigma}(L)$ is disjoint. \end{enumerate} There is a unique expression \begin{equation*} P_{\sigma}(\phi^{*}(L+\epsilon A)) = P_{\sigma}(\phi^{*}L) + \phi^{*}A + \alpha(\epsilon)\widetilde{E} + F \end{equation*} where $\widetilde{E}$ is the strict transform of $E$, $F$ is an effective divisor with $F \leq N_{\sigma}(\phi^{*}L)$ and the support of $F$ does not contain $E$, and $\alpha(\epsilon)$ is positive and goes to $0$ as $\epsilon$ goes to $0$. By shrinking $\epsilon$ we may ensure that $\alpha(\epsilon) < \sigma_{E}(L)$. Condition (2) above, along with Lemma \ref{vollemma}, show that the restriction $(P_{\sigma}(\phi^{*}L) + \alpha(\epsilon)\widetilde{E})|_{\widetilde{E}}$ is not pseudo-effective for any $\epsilon > 0$. Since $\alpha(\epsilon) < \sigma_{E}(L)$, we also have that $(P_{\sigma}(\phi^{*}L) + \sigma_{E}(L)\widetilde{E})|_{\widetilde{E}}$ is not pseudo-effective. As the strict transform of components of $N_{\sigma}(L)$ are disjoint, the restriction of $P_{\sigma}(\phi^{*}L) + \phi_{*}^{-1} N_{\sigma}(L)$ to $\widetilde{E}$ is still not pseudo-effective. By \cite[0.2 Theorem]{bdpp04} there is a curve $\widetilde{C}$ whose deformations cover $\widetilde{E}$ such that $$(P_{\sigma}(\phi^{*}L) + \phi_{*}^{-1} N_{\sigma}(L)) \cdot \widetilde{C} < 0.$$ Since $\widetilde{E}$ is not $\phi$-exceptional, $C = \phi(\widetilde{C})$ is a $\mathrm{mov}^{1}$-curve. Conversely, if $L$ is movable, then $\phi^{-1}_{mov}(L) = P_{\sigma}(\phi^{*}L)$ is also movable for every $\phi$. Thus every movable transform has non-negative intersection with the strict transform of every $\mathrm{mov}^{1}$-curve general in its family. \end{proof} \begin{lem} \label{vollemma} Let $X$ be a smooth projective variety and let $L$ and $L'$ be pseudo-effective divisors on $X$. Then $\mathrm{vol}_{X}(L + L') \geq \mathrm{vol}_{X}(L)$. \end{lem} \begin{proof} We may assume $L$ is big since otherwise the inequality is automatic. Then for any sufficiently small $\epsilon > 0$ we have $$\mathrm{vol}_{X}(L+L') = \mathrm{vol}_{X}((1-\epsilon) L + (\epsilon L + L')) \geq (1-\epsilon)^{\dim X}\mathrm{vol}_{X}(L)$$ since $\epsilon L + L'$ is big. \end{proof} We now give an alternate formulation of Theorem \ref{maintheorem}. \begin{prop} \label{mainprop} Let $X$ be a smooth projective variety and let $L$ be a pseudo-effective $\mathbb{R}$-divisor. Suppose that $V$ is an irreducible subvariety of $X$ contained in $\mathbf{B}_{-}(L)$ and let $\psi: X' \to X$ be a smooth birational model resolving the ideal sheaf of $V$. Then there is a birational morphism $\phi: Y \to X'$ from a smooth variety $Y$ and an irreducible curve $\widetilde{C}$ on $Y$ such that \begin{equation*} \phi^{-1}_{mov}(\psi^{*}L) \cdot \widetilde{C} < 0 \end{equation*} and $\psi \circ \phi(\widetilde{C})$ deforms to cover $V$. \end{prop} \begin{proof} Let $E$ be the $\psi$-exceptional divisor dominating $V$. Since we have $E \subset \mathrm{Supp}(N_{\sigma}(\psi^{*}L))$, we may argue as in the proof of Theorem \ref{maintheorem} for $\psi^{*}L$ and $E$ to find a birational map $\phi$ such that $\phi^{-1}_{mov}(\psi^{*}L)|_{\widetilde{E}}$ is not pseudo-effective. \cite[2.4 Theorem]{bdpp04} shows that there is some curve $\widetilde{C}$ on $\widetilde{E}$ with $\phi^{-1}_{mov}(\psi^{*}L) \cdot \widetilde{C} < 0$ such that $\widetilde{C}$ deforms to cover $\widetilde{E}$ and is not contracted by any morphism from $\widetilde{E}$ to a variety of positive dimension. Choosing $\widetilde{C}$ on $\widetilde{E}$ to satisfy this stronger property, we obtain the statement of Proposition \ref{mainprop}. \end{proof} Proposition \ref{mainprop} shows that the non-nef locus is covered by $L$-negative curves in a birational sense. Alternatively, one can rephrase this result using $k$-movability. \begin{cor} \label{maincor} Let $X$ be a smooth projective variety and let $L$ be a pseudo-effective $\mathbb{R}$-divisor. Then $L$ is not $k$-movable if and only if there is a birational morphism $\psi: X' \to X$ from a smooth variety $X'$, a birational morphism $\phi: Y \to X'$ from a smooth variety $Y$, and an irreducible curve $\widetilde{C}$ on $Y$ such that \begin{equation*} \phi^{-1}_{mov}(\psi^{*}L) \cdot \widetilde{C} < 0 \end{equation*} and $\psi \circ \phi(\widetilde{C})$ deforms to cover a $k$-dimensional subset of $V$. \end{cor} \begin{proof} To say that $L$ is not $k$-movable is equivalent to saying that $\mathbf{B}_{-}(L)$ has a component of dimension at least $k$. Apply Proposition \ref{mainprop} to obtain the forward implication. The converse is immediate. \end{proof} \begin{rmk} It is unclear whether Corollary \ref{maincor} is the best formulation possible for the duality of $k$-movable divisors. For Mori Dream Spaces varieties and for $2 < k < \dim X$, \cite{payne06} Theorem 1 and \cite{choi11} Corollary 3 prove a slightly stronger statement. The essential difference is that one does not need to blow-up along top-dimensional components of $\mathbf{B}_{-}(L)$. More precisely, if $L$ is not $k$-movable, one may find a $\mathbb{Q}$-factorial small modification $f: X \dashrightarrow X'$ that is regular at the generic point of a component $V \subset \mathbf{B}_{-}(L)$ of codimension at most $k$ and a family of curves covering the strict transform of $V$ with $f_{*}L \cdot C < 0$. In contrast, Corollary \ref{maincor} may produce a birational map that is not regular at any point of $V$. \end{rmk} \nocite{*} \bibliographystyle{amsalpha}
\section{Introduction} The Fokas-Gel'fand formula for immersion associated with integrable models is a powerful tool for the construction and investigation of 2D soliton surfaces in Lie algebras \cite{FG, FGFL}. In \cite{GP2011a}, the authors described how a zero-curvature representation of integrable nonlinear partial differential equations (PDEs) can be used, via the Fokas-Gel'fand formula, to induce surfaces immersed in Lie algebras via symmetries of the compatibility conditions for a linear spectral problem (LSP). The procedure was then applied to integrable nonlinear ordinary differential equations (ODEs) admitting a Lax representation \cite{GP2011b}. The results obtained were so promising that it seemed to be worthwhile to try to extend this method and check its effectiveness for a different form of a zero-curvature condition (ZCC). This is, in short, the aim of the present paper which provides a self-contained comprehensive study of the symmetry approach to the Fokas-Gel'fand immersion formula as applied to both PDE and ODE cases. Namely, the general theory is developed for each of the following three forms of matrix LSPs: zero-curvature representation (ZCR) of PDEs, Lax representation of ODEs and ZCR of ODEs. In addition to more general considerations, we focus on the third form of a ZCC for a differential equation, that of zero-curvature representations for ODEs involving the differentiation of the potential matrices in a Lax pair with respect to the independent variable and the spectral parameter. Such a Lax pair has been used in the study of Painlev\'e equations. Each of the three types of Lax pair described above, can be thought of as a sub-case of a general ZCC and to this end, we describe infinitesimal deformations of such equations and their associated surfaces. The zero-curvature representation of the Gauss-Mainardi-Codazzi (GMC) equations is a well known tool for the analysis of many, varied forms of integrable surfaces \cite{ BavMarv2010, Bob1994, CGS1995}. The necessary and sufficient conditions for the existence of surfaces immersed in the Lie algebra $\mathfrak{g}$ whose GMC equations are equivalent to an infinitesimal deformation of the ZCC are given in terms of the determining equation for generalized symmetries of the ZCC. For this purpose, in Proposition 1, we give a complete classification of all generalized symmetries of the ZCC via an isomorphism between such symmetries and surfaces written in the moving frame defined by conjugation by the wave function in the LSP. In addition to symmetries of the ZCC, it is possible to obtain infinitesimal deformations of the ZCC from the integrable model considered, see \cite{FGFL}. In \cite{GP2011a, GP2011b}, the necessary and sufficient conditions for the existence and explicit integration of the surfaces were given in terms of the symmetry criterion for generalized vector fields. Taken together, the symmetries of the ZCC and the integrable model give a quite general form for soliton surfaces, which we refer to as the Fokas-Gel'fand formula for immersion of surfaces in Lie algebras. These surfaces include those already known, namely those constructed from a conformal symmetry of the spectral parameter (the Sym-Tafel formula for immersion \cite{Sym, Tafel}), gauge symmetry \cite{Cies1997, FGFL}, and generalized symmetries of the integrable system and its LSP \cite{FGFL}. In this paper, we give additional generalized symmetries of the ZCC and their associated soliton surfaces. In many cases, including those considered herein, the classical symmetry analysis is not a proper tool for obtaining these types of surfaces because the Lie point symmetries of the initial system, written in terms of the ZCC and its LSP, are too restrictive. To overcome this difficulty, we make use of the concept of a generalized symmetry as introduced by E. Noether \cite{Noether}. The progress in studying general properties of generalized symmetries has been subsequently developed by several authors (see e.g. \cite{Olver} and references therein). The most important element of this approach is the introduction of a symmetry formalism of generalized vector fields (defined on some extended jet space), their prolongation structure and their links with the Fr\'echet derivative. An interesting consequence of the adopted approach is that it leads to a direct connection between generalized symmetries and their associated surfaces for Painlev\'e type equations (see e.g. \cite{Incebook}). We focus here on the Painlev\'e equations P1, P2 and P3 as examples, namely \begin{eqnarray} P1: & x_{tt}=6x^2+t,\nonumber \\ P2: & x_{tt}=2x^3+tx-\alpha,\nonumber \\ P3: & x_{tt}=\frac{(x_t)^2}{x}-\frac{x_t}{t}+\frac1t(\alpha x^2+\beta)+\gamma x^3+\frac{\delta}{x}.\nonumber \end{eqnarray} The quantities $\alpha, \beta, \gamma, $ and $\delta$ in the above equations are free parameters. For specific values of these parameters, several explicit solutions of the Painlev\'e equations have been found (see e.g. for review \cite{AblClabook, Conte1999, FIKNbook, Gromak1999, Iwaetal1991, McLOlv1983}). The successive application of B\"acklund transformations (with different values of parameters) allows one to create new solutions of a Painlev\'e equation from the old ones. In this context, the symmetry analysis of Painlev\'e equations has been systematically undertaken in \cite{Noumibook} and has proven to be a useful tool since it leads to new solutions. However, the present state of solvability of the Painlev\'e equations through the so-called Painlev\'e transcendents is still not satisfactory from the point of view of symmetry groups. The progress in studying general properties of generalized symmetries of Painlev\'e equations and their LSPs opens a possibility for applying this symmetry approach to the construction of soliton surfaces. An original procedure for constructing surfaces associated with Painlev\'e equations via isomonodromic deformations was devised by A. Bobenko and U. Eitner in \cite{BobEitbook}. These equations, written in zero-curvature form involving the differentiation of the potential matrices in a Lax pair with respect to the independent variable and the spectral parameter, arise from the compatibility conditions of an LSP. The surfaces considered in this work correspond to infinitesimal deformations of such surfaces as the corresponding structural equations are infinitesimal deformations of the zero-curvature conditions of the integrable models. This paper is organized as follows. In section 2, we discuss surfaces immersed in Lie algebras whose GMC equations are given by infinitesimal deformations of the ZCC. In section 3, we discuss three different ways that the ZCC can be realized as the compatibility condition for the LSP of an integrable model and how the symmetries of this model give further surfaces. The case where the ZCC is realized by the Lax pair for an ODE is given in detail in section 4. To illustrate the theoretical considerations, section 5 contains explicit construction and geometric analysis of surfaces associated with the Painlev\'e equations P1, P2 and P3. Section 6 contains final remarks and possible future developments. \section{The zero curvature condition and its associated surfaces} \subsection{The ZCC and its jet space} Consider the following partial differential equation (PDE) which, in what follows, shall be referred to as the zero-curvature condition (ZCC), \begin{equation} \label{Delta} \Delta[u]=D_2 u^1-D_1 u^2 +\left[u^1, u^2\right]=0,\end{equation} with independent variables $\xi_{i}, \ i=1,2$ and dependent matrix variables $u^\alpha,$ $\alpha=1,2$ which take their values in a Lie algebra $\mathfrak{g}.$ Here, we write \eref{Delta} as a function on the jet space $M\equiv(\xi_1, \xi_2, u^\alpha, u^\alpha_J),$ where the derivatives of $u^\alpha$ are given by \begin{equation} \label{coordsjetspace}\frac{\partial ^n}{\partial \xi_{j_1}\ldots \partial \xi_{j_n}} u^\alpha\equiv u^\alpha_J, \qquad J=(j_1, \ldots j_n), \quad j_i=1,2, \ |J|=n.\end{equation} We define $\mathcal{A}\equiv C^\infty (M)$ to be the set of smooth functions on the jet space $M$ and use the abbreviated notation $f(\xi_1, \xi_2, u^\alpha, u^\alpha_J)\equiv f[u]\in\mathcal{A}.$ Thus, the PDE \eref{Delta} becomes a vanishing element of $\mathcal{A}.$ The derivative of a function on the jet space $M$, in the direction $\xi_i,$ is given by the total derivative $D_i$ defined as \begin{equation} D_i=\frac{\partial}{\partial \xi_i}+u^{\alpha}_{J,i}\frac{\partial}{\partial u^\alpha_J}, \qquad \alpha,i=1,2.\end{equation} The ZCC \eref{Delta} can be realized as the compatibility conditions of an associated linear spectral problem (LSP) \begin{equation}\label{LSP} D_\alpha \Psi=u^\alpha \Psi, \qquad \alpha=1,2,\end{equation} where $\Psi=\Psi([u])$ is a function from the jet space $M$ to a Lie group, G, whose algebra is $\mathfrak{g}$. Note that $\Psi$ satisfying the LSP \eref{LSP} is an immersion function for a surface immersed in the Lie group $G$ where the compatibility of the tangent vectors $D_\alpha \Psi$ is equivalent to the ZCC \eref{Delta}. Furthermore, for any (generalized) symmetry of \eref{Delta} there exists a $\mathfrak{g}$-valued immersion function, $F[u],$ defined in terms of its tangent vectors \begin{equation} \label{DF} D_1 F=\Psi^{-1} Q^1\Psi, \qquad D_2 F=\Psi^{-1} Q^2\Psi,\end{equation} where $Q^1$ and $Q^2$ satisfy an infinitesimal deformation of the ZCC \eref{Delta}, namely \begin{equation}\label{deteq} D_1 Q^2-D_2Q^1+[Q^1,u^2]+[u^1, Q^2]=0.\end{equation} In what follows, we suppose that $F$ is sufficiently smooth. The existence of such surfaces as well as the determining equation \eref{deteq} was first identified for point symmetries by Fokas and Gel'fand in \cite{FG}. Following the notation as in \cite{Olver}, a generalized symmetry of \eref{Delta} is given by a vector field, in evolutionary form, \begin{equation} \vec{v}_Q=Q^{\alpha,j}[u] \frac{\partial}{\partial u^{\alpha,j}},\end{equation} where $u^{\alpha, j}$ are the components of $u^\alpha$ in a basis for $\mathfrak{g},$ i.e. $u^{\alpha}=u^{\alpha, j}e_j$ with $j=1...n.$ Note that $Q^\alpha\equiv Q^{\alpha, j}e_j$ is an element of the Lie algebra $\mathfrak{g}$. The prolongation of $\vec{v}_Q$ is defined to be \begin{equation} pr\vec{v}_Q=\vec{v}_Q+D_J(Q^{\alpha,j}[u]) \frac{\partial}{\partial u^{\alpha,j}_J}.\end{equation} The vector field $\vec{v}_Q$ is a generalized symmetry of a nondegenerate PDE \eref{Delta} if and only if \cite{Olver} \begin{equation} pr\vec{v}_Q(\Delta[u])=0, \qquad \mbox{ whenever } \Delta[u]=0,\end{equation} holds. In fact the prolongation of $\vec{v}_Q$ acting on $\Delta[u]=0$ is exactly the determining equations \eref{deteq}. We have the following theorem. \begin{theorem}[Existence of the $\mathfrak{g}$-valued immersion function] Suppose that there exists a smooth, $G$-valued function $\Psi$ which satisfies the LSP \eref{LSP}. Suppose also that the generalized vector field, $\vec{v}_Q,$ is a symmetry of the ZCC \eref{Delta}. Then there exists a smooth $\mathfrak{g}$-valued function $F[u]$ with tangent vectors given by \eref{DF}. Further, these tangent vectors are explicitly integrated as \begin{equation} \label{F} F=\Psi^{-1} pr\vec{v}_Q\Psi\in \mathfrak{g}\end{equation} if and only if the vector field $\vec{v}_Q$ is a symmetry of the LSP \eref{LSP} in the sense that \begin{equation}\label{deteqLSP} pr\vec{v}_R\left(D_\alpha\Psi-u^\alpha \Psi\right)=0, \qquad \mbox{ whenever } D_\alpha\Psi-u^\alpha\Psi=0.\end{equation} \end{theorem} This theorem has been proved in \cite{GP2011a}. It need only be observed that $Q^\alpha=pr\vec{v}_Q(u^\alpha)$ and that the compatibility conditions for \eref{DF} are exactly the determining equations \eref{deteq}. The latter are in turn equivalent to $pr\vec{v}_Q(\Delta[u])=0$ since the total derivatives $D_\alpha$ commute with the prolongation of vector fields in evolutionary form, $pr\vec{v}_Q$, (see Lemma 5.12 page 306 in \cite{Olver}). To prove that $F$ given by \eref{F} has the appropriate tangent vectors, one need only compute the tangent vectors from \eref{F} to show that they coincide with \eref{DF} if and only if \eref{deteqLSP} holds. Note that whenever the characteristics $Q^1$ and $Q^2$ are linearly independent and the tangent vectors $D_1F$ and $D_2F$ are compatible, the differential 1-form $dF$ is closed and its integral \begin{equation}\label{dF} F=\int_\gamma \Psi^{-1}Q^1\Psi d\xi_1+\Psi^{-1} Q^2\Psi d\xi_2\in \mathfrak{g},\end{equation} depends only on the end points of the trajectory $\gamma$ in the plane $\mathbb{R}$. The integral \eref{dF} defines a mapping $F:\mathbb{R}\backepsilon (\xi_1, \xi_2)\rightarrow \mathfrak{g}$ which coincides with \eref{F} whenever \eref{deteqLSP} hold. The representation of a surface in terms of the 1-form \eref{dF} coincides with the generalized Weierstrass formula for immersion as introduced in \cite{Kono1996}. \subsection{Symmetries of the ZCC and associated surfaces}\label{symalpha} In this section, we present the symmetries of the ZCC and their associated surfaces. Note that Theorem 1 says that from a given symmetry of the ZCC it is possible to construct a surface immersed in the Lie algebra $\mathfrak{g}$. Next, we shall prove the converse. That is given a surface immersed in the Lie algebra $\mathfrak{g}$, it is possible to define symmetries of the ZCC and, as a result, to characterize all generalized symmetries of the ZCC in terms of a gauge function. Finally, one gets the following. \begin{proposition}[Extracting symmetries from a surface] Suppose that $\Psi[u]$ is a smooth, $G$-valued solution of the LSP \eref{LSP} for any solution $u^\alpha$ of the ZCC $\Delta[u]=0$. Suppose further that $F[u]$ is a smooth function on jet space $M$ taking values in the Lie algebra $\mathfrak{g}$, then the generalized vector field in evolutionary form \begin{equation} \vec{v}_Q=Q^{\alpha,j}[u] \frac{\partial}{\partial u^{\alpha,j}},\qquad \label{QsF} Q^{\alpha}=\Psi\left( D_\alpha F\right)\Psi^{-1},\quad \alpha=1,2 \ j=1,\ldots, n\end{equation} is a symmetry of the ZCC, $\Delta[u]=0.$ \end{proposition} {\bf Proof}: Since the components of $F[u]$ are smooth functions, the cross partials commute, i.e. $D_1D_2F=D_2D_1F$. Thus, $Q^1$ and $Q^2$ defined in terms of $F[u]$ as in \eref{QsF} satisfy \begin{eqnarray} D_2 Q^1&=&u^2\Psi( D_1F)\Psi^{-1} +\Psi( D_2D_1F)\Psi^{-1} -\Psi( D_1F)\Psi^{-1} u^2\nonumber \\ &=& u^2Q^1-Q^1u^2+\Psi (D_2D_1F)\Psi^{-1},\end{eqnarray} and similarly, \begin{equation} D_1Q^2=u^1Q^2-Q^2u^1+ \Psi( D_2D_1F)\Psi^{-1}.\end{equation} Hence, $Q^1$ and $Q^2$ defined as in \eref{QsF} satisfy \begin{equation} D_1 Q^2-D_2Q^1+[Q^1,u^2]+[u^1, Q^2]=pr\vec{v}_Q(\Delta[u])=0\end{equation} for any solutions of the ZCC $\Delta[u]=0.$ \qed \begin{corollary} Any generalized symmetry of the ZCC $\Delta[u]=0$ can be written in terms of a gauge function $S[u] \in \mathfrak{g}$ as \begin{equation}\label{QsS} Q^1=D_1 S+[S, u^1], \qquad Q^{2}=D_2 S+[S, u^2].\end{equation} \end{corollary} {Proof:} Since $\Psi$ is invertible, it is possible to express any $\mathfrak{g}$-valued function $F[u]$ in terms of a gauge matrix $S[u]$ \begin{equation} F[u]=\Psi^{-1} S[u] \Psi, \qquad D_\alpha F[u]=\Psi^{-1}\left(D_\alpha S[u] +[S[u], u^\alpha]\right)\Psi, \end{equation} so that the characteristic $Q^\alpha$ of the vector field $\vec{v}_Q$, defined as in \eref{QsF}, satisfies \eref{QsS}. \qed From this corollary, one can construct an infinite set of generalized symmetries. As examples we give the following symmetries, beginning with the most general \begin{equation} \fl \begin{array}{lr} \vec{v}_{Q_0}=\left(D_\alpha(S)+\left[S,u^\alpha\right]\right)^j\frac{\partial}{\partial u^{\alpha j}}&\!\!\!\!\!\!\!\! \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\mbox{ gauge transformation of the LSP }\\ \vec{v}_{Q_i}=u^{\alpha j}_i\frac{\partial}{\partial u^{\alpha j}}&\!\!\!\!\!\!\!\!\!\!\!\!\!\!\mbox{ translations in } \xi_i, \ i=1,2 \\ \vec{v}_{Q_3}=D_1\left(\xi_1 u^1\right)\frac{\partial}{\partial u^1}+\xi_1D_1\left( u^2\right)\frac{\partial}{\partial u^2}&\!\!\!\!\!\!\!\!\mbox{ dilation in } \xi_1, u^1\\ \vec{v}_{Q_4}=\xi_2 D_2\left( u^1\right)\frac{\partial}{\partial u^1}+D_2\left(\xi_2 u^2\right)\frac{\partial}{\partial u^2}&\!\!\!\! \! \!\!\!\mbox{ dilation in } \xi_2, u^2\\ \vec{v}_{Q_5}=\left(D_1^2u^1+[D_1u^1,u^1]\right)\frac{\partial}{\partial u^1}+ \left(D_1^2u^2+[D_1u^2,u^1]\right)\frac{\partial}{\partial u^2}&\!\!\! \!\!\!\!\mbox{generalized symmetry}.\end{array} \end{equation} In each of these cases, it is possible to give the explicit form of the function $F$ with tangent vectors as in \eref{DF} as was demonstrated in \cite{FG}. In what follows, we assume that the tangent vectors listed below are linearly independent. Then, $F^{i}$ for $i=1,\ldots 5$ are the immersion functions of 2D-surfaces in the Lie algebra $\mathfrak{g}$. The surface associated with the gauge symmetry (studied in \cite{Cies1997, FGFL}) is, in terms of $S[u]\in g$, \begin{eqnarray} F^{S}=\Psi^{-1}S\Psi, \\ D_\alpha F^{S}=\Psi^{-1}\left( D_\alpha S+[S, u^\alpha]\right)\Psi\nonumber.\end{eqnarray} As noted above, $F^{S}$ is in fact an arbitrary surface immersed in the Lie algebra $\mathfrak{g}$ and the gauge function $S[u]$ can be thought of as the representation of the surface in the moving frame defined by conjugation with respect to $\Psi\in G.$ Below, we give several examples of symmetries of the ZCC \eref{Delta} and the corresponding surfaces, or equivalently the gauges, to which they are associated. The surfaces associated with the symmetries $\vec{v}_{Q_1}$ and $\vec{v}_{Q_3}$ are special cases of a more general surface given, in terms of $r(\xi_1),$ an arbitrary function of $\xi_1,$ by \begin{eqnarray} F^{r}=r(\xi_1) \Psi^{-1} D_1 \Psi,\\ D_1 F^{r}=\Psi^{-1}D_1\left(r(\xi_1)u^1\right) \Psi,\qquad D_2 F^{r}=r(\xi_1)D_1(u^2)\Psi.\end{eqnarray} The gauge for this surface is \[ S^{r}[u]=\Psi F^{r}\Psi^{-1}=r(\xi_1)u^1,\] and the characteristics $Q^\alpha$ for the vector field $\vec{v}_Q$ associated with this gauge are \begin{eqnarray}\fl Q^{1}=D_1S^{r}+[S^r,u^1]=D_1\left(r(\xi_1)u^1\right),\qquad Q^{2}=D_2 S^{r}+[S^r,u^2]=r(\xi_1)D_1(u^2). \end{eqnarray} In the case that $r(\xi_1)=1$, the vector field $\vec{v}_{Q}$ reduces to $\vec{v}_{Q_1}$ and the surface is given by \[ F^{1}=\Psi^{-1} D_1 \Psi.\] If instead the function reduces to $r(\xi_1)=\xi_1$, the vector field becomes $\vec{v}_{Q_3}$ and the surface is \[ F^{3}=\xi_1 \Psi^{-1} D_1 \Psi.\] Similarly, the surfaces associated with the symmetries $\vec{v}_{Q_2}$ and $\vec{v}_{Q_4}$are special cases of the more general surface, \begin{eqnarray} F^{s}=s(\xi_2) \Psi^{-1} D_2 \Psi,\\ D_1 F^{s}=s(\xi_2)\Psi^{-1}D_2u^1 \Psi,\qquad \nonumber D_2 F^{s}=\Psi^{-1}D_2\left(s(\xi_2)u^2\right) \Psi ,\end{eqnarray} where $s(\xi_2)$ is an arbitrary function of $\xi_2.$ The gauge associated with this surface is \[ S^{s}[u]=\Psi F^{s}\Psi^{-1}=s(\xi_2)u^2,\] and the characteristics $Q^\alpha$ for the vector field $\vec{v}_Q$ which give this particular gauge are \begin{eqnarray} Q^{1}=D_1S^{s}+[S^s,u^1]=s(\xi_2)D_2\left(u^1\right),\nonumber \\ Q^{2}=D_2 S^{s}+[S^s,u^2]=D_2\left(s(\xi_2)u^2\right)\label{vecST}. \end{eqnarray} The construction of the surfaces $F^2$ and $F^4$ is similar to that above for the surfaces $F^1$ and $F^3$ with $\xi_1$ replaced with $\xi_2.$ The point symmetries associated with the vector fields $\vec{v}_{Q_1}, \ldots, \vec{v}_{Q_4}$ (translations and scalings) and the surfaces $F^{1}, \ldots , F^{4}$ were identified in \cite{FG}. The final symmetry to be considered as an example is a generalized symmetry, \begin{eqnarray} F^{5}=\Psi^{-1}D_1u^1\Psi,\\ \label{D1F6}\fl D_1 F^{5}=\Psi^{-1}\left(D_1^2u^1+[D_1u^1,u^1]\right)\Psi,\qquad D_2 F^{5}= \Psi^{-1}\left(D_1^2u^2+[D_1u^2,u^1]\right)\Psi.\end{eqnarray} It is immediate to see that the tangent vectors to $F^{5}$ satisfy \eref{D1F6} . Perhaps less obvious is that the vector field, \begin{equation} \label{v6} \vec{v}_{Q_5}=\left(D_1^2u^1+[D_1u^1,u^1]\right)\frac{\partial}{\partial u^1}+\left(D_1^2u^2+[D_1u^2,u^1]\right)\frac{\partial }{\partial u^2},\end{equation} is in fact a generalized symmetry of $\Delta[u]=0.$ Computing the action of the prolongation of the vector field $\vec{v}_{Q_5}$ on $\Delta[u]=0$, in the coordinates of the extended jet space \eref{coordsjetspace}, gives \begin{eqnarray} \fl pr\vec{v}_{Q_5}(\Delta[u])&=&u^1_{112}+[u^1_{12},u^1]+[u^1_1, u^1_2]-u^2_{111}-[u^2_{11},u^1]-[u^2_1,u^1_1]\nonumber \\ \fl && +[u^1_{11},u^2]+\left[[u^1_1,u^1],u^2\right]+[u^1,u^2_{11}]+\left[u^1, [u^2_1,u^1]\right]\nonumber \\ \fl &=&D_1^2(\Delta[u])+\left[D_1(\Delta[u]),u^1\right]-\left[\Delta[u],u^1_1\right]\nonumber \\ \fl &=& 0, \qquad \qquad \qquad \qquad \mbox{ whenever } \Delta[u]=0\nonumber.\end{eqnarray} Thus, the vector field $ \vec{v}_{Q_5}$ \eref{v6} is a generalized symmetry of $\Delta[u]=0.$ In this section, we have shown how generalized symmetries of the ZCC \eref{Delta} can be used to construct surfaces whose GMC equations are given by infinitesimal deformations of the ZCC \eref{Delta}. In the process, we were able to characterize all possible symmetries of the ZCC \eref{Delta} in terms of a gauge transformation of the associated LSP. In the next section, we shall consider cases where the ZCC \eref{Delta} is equivalent to some integrable model and the surfaces associated to such a model. In this case, the form of surfaces obtained is enlarged since the considered symmetries are not only those of the ZCC \eref{Delta} but also of the integrable model itself. In general, these two classes of symmetries are distinct. \section{Integrable equations and their Lax pairs}\label{integrable} Consider a Lax pair representation of an integrable model whose determining equation is equivalent to the ZCC \eref{Delta}. That is, suppose that we are able to parameterize the matrices \[\tau(u^\alpha)= U^\alpha[\theta]\in \mathfrak{g}\] in terms of some set of dependent variables $\theta^j$ which depend on (possibly a subset of) the independent variables $\xi^j$. The matrices $U^\alpha[\theta]$ are also allowed to depend on the independent variables themselves, $\xi^i$ and possibly some space of constants. Define $N$ to be the jet space associated with $\theta^j$ and its derivatives and $\mathcal{B}$ to be the space of smooth functions on this jet space, which depend also on a finite number, say $k,$ of constants that take values in some field $\mathbb{K}$, usually either $\mathbb{C}$ or $\mathbb{R}.$ This parameterization induces a mapping between function spaces $\mathcal{A} $ and $\mathcal{B}$ \begin{equation} \begin{array}{ccc} & \tau &\\ u^\alpha\in \mathfrak{g} &\rightarrow& U^\alpha([\theta])\in \mathfrak{g} \\ \downarrow & & \downarrow\\ & \tau &\\ \mathcal{A}=C^\infty(M) & \rightarrow &\mathcal{B}=C^\infty(N\times \mathbb{K}^k), \end{array} \end{equation} where the function $\tau$ is defined by the mapping $u^\alpha_J\rightarrow D_J U^\alpha[\theta]$ so that \begin{equation} \tau\left( f(\xi_1,\xi_2,u^\alpha, u^\alpha_J)\right)=f\left(\xi_1, \xi_2, U^\alpha, D_JU^\alpha[\theta] \right)\in \mathcal{B}.\end{equation} Under this mapping, the ZCC \eref{Delta} is converted into a system of differential equations for the dependent variables $\theta^j$ denoted \begin{eqnarray} \label{Delta'} \Omega[\theta]&=&\tau(\Delta[u])\\ &=&D_1U^2[\theta]-D_2U^1[\theta]+\left[U^1[\theta], U^2[\theta]\right]=0\nonumber.\end{eqnarray} The mapping of the LSP is straightforward. The wave functions are written $\Phi=\tau(\Psi)$ and satisfy \begin{eqnarray}\label{LSP'} D_1\Phi=\tau(u^1)\Phi=U^1[\theta]\Phi,\qquad D_2\Phi=\tau(u^2)\Phi=U^2[\theta]\Phi.\end{eqnarray} Again, for any generalized symmetry of the system of differential equations $\Omega[\theta]=0$ written in evolutionary form \begin{equation} \label{wQ} \vec{w}_R=R^j[\theta]\frac{\partial}{\partial \theta^j}, \qquad R^j[\theta]\in \mathcal{B} \end{equation} there exists a surface $F\in \mathfrak{g}$ whenever the tangent vectors given by \begin{equation} \label{DF'} D_1F=\Phi^{-1} \left(pr\vec{w}_R U^1[\theta]\right) \Phi, \qquad D_2F=\Phi^{-1}\left( pr\vec{w}_R U^2[\theta]\right) \Phi\end{equation} are linearly independent. Here we use $R$ to denote the characteristic of the vector field $\vec{w}_R$ on jet space $N$ to distinguish it from the characteristic $Q$ of the vector field $\vec{v}_Q $ on jet space $M$. The following theorem holds as was first proven in \cite{GP2011a}. \begin{theorem} Suppose that the smooth potential matrices $U^\alpha[\theta]\in \mathfrak{g}$ satisfy \eref{Delta'} and $\Phi\in G$ satisfies the LSP \eref{LSP'}. Suppose further that there exists some generalized vector field $\vec{w}_R$ which is a symmetry of \eref{Delta'}. Then, there exists a $\mathfrak{g}$-valued function $F$ whose tangent vectors satisfy \eref{DF'}, which can be written, up to a $\mathfrak{g}$-valued constant, by \begin{equation}\label{F'} F=\Phi^{-1} pr\vec{w}_R \Phi\end{equation} if and only if $\vec{w}_R$ is a symmetry of the LSP \eref{LSP'} in the sense that, for $\alpha=1,2$, \begin{equation}\label{deteqLSP'} pr\vec{w}_R\left(D_\alpha\Phi-U^\alpha[\theta]\Phi\right)=0, \qquad \mbox{ whenever } D_\alpha\Phi-U^\alpha[\theta]\Phi=0.\end{equation} \end{theorem} To prove this theorem, one need only check that the compatibility conditions for \eref{DF'} are equivalent to the determining equations for a symmetry of \eref{Delta'}. Similarly, it is a direct computation to show that the tangent vectors of $F$ given by \eref{F'} coincide with those given by \eref{DF'} if and only if \eref{deteqLSP'} holds. We can consider the parameterization of the matrix functions $U^\alpha[\theta]$ as the Lax pairs for some integrable differential equations. Below, we distinguish three possible choices of parameterizations of potential matrices $U^\alpha[\theta]\in \mathfrak{g}$: \begin{itemize} \item[ 1.]{\bf Zero-curvature representation for PDEs} In this case, suppose that the matrices $U^\alpha$ depend on both the independent variables $\xi_i=x_i$ $i=1,2$ and some dependent variables $f^j(x_1, x_2)$, as well as a spectral parameter $\lambda$ such that the ZCC $\Delta[u]=0$ is equivalent to a system of PDEs independent of the spectral parameter (see e.g. \cite{ZakSha1974, ZakSha1978}) \begin{eqnarray} \tau(\Delta[u])&=&D_2U^1([f],\lambda)-D_1U^2([f],\lambda)+\left[U^1([f],\lambda),U^2([f],\lambda)\right]\nonumber \\ &=&\Omega[f]=0.\label{Delta'f} \end{eqnarray} This was the case treated in \cite{GP2011a} where the surface considered were associated with symmetries of the ZCC $\Delta[u]=0$ and the PDE \eref{Delta'f}. \item[2.] {\bf Lax form for ODEs} Suppose now that we have only one independent variable $\xi_1=x_1$ and several dependent functions of one variable $g^j(x_1)$. The matrices $U^\alpha$ are functions on the jet space defined by $x_1$ and $g^j(x_1)$ and depend also on some spectral parameter $\lambda$ but not on $\xi_2=x_2$. The ZCC $\Delta[u]=0$ is thus equivalent to a system of ODEs which is assumed to be independent of the spectral parameter (see e.g. \cite{Lax1968}) \begin{eqnarray} \tau(\Delta[u])&=&-D_1U^2([g],\lambda)+\left[U^1([g],\lambda),U^2([g],\lambda)\right]\nonumber \\ &=&\Omega[g]=0.\label{Delta'g} \end{eqnarray} This is the case treated in \cite{GP2011b}. Here again, the surfaces considered were associated with symmetries of the ZCC $\Delta[u]=0$ and the ODE \eref{Delta'g}. \item[3.] {\bf Zero-curvature representation for ODEs} Suppose now that we have several dependent functions $x^j(t)$ which depend only on one independent variable $\xi_1=t$. The matrices $U^\alpha$ are functions on the jet space defined by $t$ and $x^j(t)$ and the other independent variable, which here takes the form of a spectral parameter $\xi_2=\lambda.$ In this case, the ZCC $\Delta[u]=0$ is equivalent to a system of ODEs which is assumed to be independent of the spectral parameter (see e.g. \cite{BobEitbook}) \begin{eqnarray} \tau(\Delta[u])&=&D_\lambda U^1([x],\lambda)-D_tU^2([x],\lambda)+\left[U^1([x],\lambda),U^2([x],\lambda)\right]\nonumber \\ &=&\Omega[x]=0,\label{Delta'x}\end{eqnarray} where $D_t$ and $D_\lambda$ represent the total derivatives with respect to $t$ and $\lambda$ respectively. This is the situation to be treated for the remainder of the paper and th surfaces are defined via symmetries of the ZCC $\Delta[u]=0$ and the ODE \eref{Delta'x} \end{itemize} Note that in the first two cases, the spectral parameter, $\lambda,$ enters as a parameter in the LSP \eref{LSP'} whereas, in the third case, it enters as an independent variable in the LSP. Thus, the Sym-Tafel (ST) formula (see e.g \cite{Dodd1998, Kono1996, Sym, Tafel}) has a different realization in these cases. In the first two cases the independent variables are $x_1$ and $x_2$ and the ST formula, give by \begin{equation} F^{ST}(x_1, x_2)=s(\lambda) \Phi^{-1} D_\lambda \Phi \label{STx}, \end{equation} is associated with a conformal symmetry in the spectral parameter of the equations \eref{Delta'f} and \eref{Delta'g} respectively. The tangent vectors of the ST formula in these cases are \begin{eqnarray} D_1F^{ST}(x_1, x_2)=s(\lambda) \Phi^{-1}D_\lambda U^1\Phi,\qquad D_2F^{ST}(x_1, x_2)=s(\lambda) \Phi^{-1}D_\lambda U^2\Phi\nonumber .\end{eqnarray} On the other hand, in the third case, the independent variables are $t$ and $\lambda$ and the immersion function (in analogy with \eref{STx}) \begin{equation} F^{ST}(t, \lambda)=s(\lambda) \Phi^{-1} D_\lambda \Phi \label{Stl}, \end{equation} is associated with a symmetry of the ZCC condition \eref{Delta}, which is a linear combination of a conformal symmetry in the spectral parameter $\xi_2=\lambda$ and a scaling of the potential matrices $u^1$ and $u^2$. The vector field $\vec{v}_Q$, associated with this symmetry, has characteristics given by \eref{vecST} and the tangent vectors of the surface are \begin{eqnarray} D_t F^{ST}(t,\lambda)=s(\lambda)\Phi^{-1}D_\lambda U^1 \Phi,\qquad \nonumber D_\lambda F^{ST}(t,\lambda)=\Phi^{-1}D_\lambda\left(s(\lambda)U^2\right) \Phi .\end{eqnarray} \section{Lax pairs for an ODE written in zero-curvature form} In this section, we shall consider specifically case 3 above; that is, ODEs with independent variable $t$ and dependent variable $x(t)$ which admit Lax pairs $U^\alpha([x],\lambda)$ that satisfy \begin{equation} D_\lambda U^1([x],\lambda)-D_tU^2([x],\lambda)+\left[U^1([x],\lambda),U^2([x],\lambda)\right]=0.\end{equation} This type of Lax pair has been considered for Painlev\'e equations by several authors including \cite{ConteMusette, FIKNbook}. Particularly relevant is \cite{BobEitbook} where the LSP was used to represent the Painlev\'e equations and derive the geometry of surfaces associated with a group-valued wave function $\Phi$ \eref{LSP} with compatibility conditions \eref{Delta}. On the other hand, the analysis contained herein is concerned with surfaces given by $\mathfrak{g}$-valued functions F with compatibility conditions \eref{deteq} that are infinitesimal deformations of \eref{Delta}. The surfaces to be considered are given by $F([x],\lambda)\in \mathfrak{g}$ with tangent vectors \begin{eqnarray}\label{FAB} D_t F=\Phi^{-1} A([x],\lambda) \Phi, \qquad D_\lambda F=\Phi^{-1} B([x],\lambda) \Phi,\end{eqnarray} where the $\mathfrak{g}$-valued matrices $A$ and $B$ satisfy \begin{equation}\label{AB'} D_\lambda A-D_t B+[A,U^2]+[U^1,B]=0.\end{equation} Recall that the group-valued function $\Phi=\tau(\Psi)$ satisfies \begin{equation}\label{LSP'lt} D_t\Phi=U^1\Phi , \qquad D_\lambda\Phi=U^2\Phi.\end{equation} The possible forms of $A$ and $B$ can be decomposed into symmetries $\vec{v}_Q$ of the ZCC $\Delta[u]=0$ and $\vec{w}_R$ of the integrable equation $\Omega[x]=0$. They are \begin{eqnarray} A([x],\lambda)=\tau\left(pr\vec{v}_Q(u^1)\right)+pr\vec{w}_R(U^1),\\ B([x],\lambda)=\tau\left(pr\vec{v}_Q(u^2)\right)+pr\vec{w}_R(U^2).\end{eqnarray} Note that these two classes of symmetries are different. For example, the ZCC admits translation symmetries with respect to the independent variables whereas the integrable equation it is equivalent to may not admit such symmetries. This will be the case for the Painlev\'e equations considered in the next section. Using a linear combination of the symmetries in subsection \ref{symalpha}, the tangent vectors can be written \begin{eqnarray} \fl A([x],\lambda)=\alpha_1D_tU^1+\alpha_2D_\lambda U^1+\alpha_3 \left(t D_t U^1+U^1\right)+\alpha_4 \lambda D_\lambda U^1+ \alpha_5\left(D_t^2U^1+[D_tU^1,U^1]\right)\nonumber \\ +\alpha_6 pr\vec{w}_R U^1\label{Aalpha}\end{eqnarray} and \begin{eqnarray} \fl\label{Balpha} B([x],\lambda)=\alpha_1D_tU^2 +\alpha_2 D_\lambda U^2+\alpha_3 t D_t U^2+\alpha_4 \left( \lambda D_\lambda U^2+U^2\right)+\alpha_5\left(D_t^2U^2+[D_tU^2,U^1]\right)\nonumber \\ +\alpha_6 pr\vec{w}_R U^2, \end{eqnarray} where the $\alpha_i$'s are constants. We define the Fokas-Gel'fand formula for immersion in Lie algebras to be the immersion function $F$ with tangent vectors given by \eref{FAB} which are constructed out of a linear combination of symmetries $\vec{v}_Q$ of $\Delta[u]=0$ and $\vec{w}_R$ of $\Omega[x]=0$, as in \eref{Aalpha} and \eref{Balpha}. In what follows we will refer to it as such. This formula contains a linear combination of the immersion functions defined by \eref{F} and \eref{F'}. The integrated form of the surface with tangents defined by \eref{Aalpha} and \eref{Balpha}, is given by \begin{equation}\label{Falpha} \fl F=\phi^{-1}\left(\alpha_2D_\lambda +\alpha_1D_t +\alpha_3t U^1+\alpha_4 \lambda U^2+\alpha_5D_tU^1 +\alpha_6 pr\vec{w}_R \right)\Phi,\end{equation} if and only if $\vec{w}_R$ is a generalized symmetry of the LSP \eref{LSP'}, in the sense of Theorem 2. Note here that, on the one hand, there exist exact integrated forms of $F$ for symmetries parameterized by the constants $\alpha_i$ for $i=1,\ldots,5.$ However, in order to check if the immersion function $F$ given by \eref{Falpha} has the form \begin{equation} F^{i}=\tau(\Psi^{-1}pr\vec{v}_{Q_i}\Psi),\end{equation} it is required to have a general solution for the wave function $\Psi$ so as to verify whether $\vec{v}_Q$ is a symmetry of the LSP \eref{LSP}. For example, only with an exact form of the wave function $\Psi$ is it possible to say whether the LSP in invariant under translations in $t$. On the other hand, an integrated form of the surface $F^{6}$ can be given only in the case that $\vec{w}_R$ is a symmetry of $\Omega[\theta]=0$ and its transformed LSP \eref{LSP'}. In this case, the integrated form is \begin{equation} \label{F6} F^{6}=\Phi^{-1}pr\vec{w}_R(\Phi).\end{equation} \section{Soliton surfaces via Painlev\'e equations P1, P2, and P3} Next, we present some examples which illustrate the theoretical considerations described in sections 2-4. Various cases of the Painlev\'e type P1, P2, and P3 equations have been chosen in order to construct associated soliton surfaces whose GMC equations are equivalent to infinitesimal deformations of the Painlev\'e equations. \subsection{On the application of the method} Here we shall briefly describe the present state of the analytic approach for finding soliton surfaces through the Fokas-Gel'fand formula for immersion associated with differential equations. In particular, we focus on difficulties arising in the attempt to explicitly construct surfaces immersed in the Lie algebra. The basic method for solving this problem requires three parts for an explicit representation of the immersion function $F$ (given by \eref{Falpha}), namely \begin{enumerate} \item[i)] A zero-curvature representation of the ODE $\Omega[x]=0.$ \item[ii)] A generalized symmetry, $\vec{w}_R$, of the ODE $\Omega[x]=0.$ \item[iii)] A solution $\Phi$ of the LSP \eref{LSP'}. \end{enumerate} Note here that (i) is always required. However, even without the remaining two conditions some analysis of the induced immersion function $F$ \eref{Falpha} can be performed. For example, without an explicit form of the generalized symmetry it is still possible to consider surfaces associated with terms $\alpha_i$ for $i=1\ldots 5$ in an integrated form \eref{Falpha}. Also, the geometry associated with the sixth term in \eref{Falpha} can be computed whenever $R$ satisfies the determining equation for a generalized symmetry of the considered model. In the third case, if we do not have an explicit form of the wave function $\Phi$ it is still possible to consider the geometric characteristics of the surfaces using the Killing form. Since the Killing form is invariant under group conjugation, the metric or pseudo-metric on the tangent vectors \eref{FAB}, with \eref{Aalpha} and \eref{Balpha}, as well as the normals, defined up to a normalization factor by \begin{equation} N=\Phi^{-1}\left[A,B\right]\Phi,\end{equation} are independent of the wave function $\Phi.$ The Killing form \cite{Helgason} is a symmetric bilinear product $\mathcal{B}(X,Y)$, given (up to a normalization factor) by \begin{equation} \label{B} \mathcal{B}(X, Y) =\frac12 tr(X\cdot Y),\qquad X,Y\in \mathfrak{g}.\end{equation} In terms of the basis $\{e_1, e_2, e_3\}$ for $sl(2,\mathbb{R})$, \begin{equation}\label{basis} e_1=\left[\begin{array}{cc}1&0\\ 0&-1 \end{array} \right], \quad e_2=\left[\begin{array}{cc}0&1\\ 1&0 \end{array} \right], \quad e_3=\left[\begin{array}{cc}0&-1\\ 1&0 \end{array} \right],\end{equation} the matrices $X$ and $Y$ can be decomposed as $X=X^je_j $ and $Y=Y^je_j$ for $j=1,2,3$ and the scalar product $\mathcal{B}$ defines a pseudo-Euclidean metric as \begin{equation} \mathcal{B}(X,Y)=X^j\mathcal{B}_{jk}Y^k, \qquad \mathcal{B}_{jk}= \left[\begin{array}{ccc}1 &0 &0\\ 0 & 1& 0 \\ 0&0&-1\end{array} \right].\end{equation} On $sl(2,\mathbb{R})$, the Killing form has signature $(2,1)$ and so induces a pseudo-Euclidean metric on the tangent vectors to the 2D-surface given by the immersion function $F\in sl(2, \mathbb{R})$. With this inner product, the surfaces are pseudo-Riemannian manifolds \cite{Eisenhart, doCarmo}. In what follows, the surfaces considered are defined by a single smooth immersion function $F$ as the independent variables range over some subset $(t, \lambda)\in \Sigma \subset \mathbb{C} \cup \{ \infty\}.$ The singularity structures of the Painlev\'e equations are transported onto the surfaces and, in particular, the surfaces may have unbounded components. \subsection{Painlev\'e P1} In this section, we give several examples of surfaces associated with the Lax pair for the first Painlev\'e equation, P1, \begin{equation} \label{P1} x_{tt}-6x^2-t=0\end{equation} An LSP for the Painlev\'e equation P1 \eref{P1} is given in terms of the potential matrices $U^1( [x],\lambda )$ and $U^2( [x],\lambda)$ taking values in the Lie algebra $ sl(2,\mathbb{R})$ \cite{JM1981, Kit1994} % \begin{eqnarray} \label{U1P1} U^1=\left[\begin{array}{cc} 0 &\lambda+2x\\ 1 & 0\end{array}\right], \\ \label{U2P1} U^2=\left[\begin{array}{cc} -x_t &2\lambda^2+2\lambda x+t+2x^2\\ 2(\lambda-x) & x_t\end{array} \right],\nonumber\end{eqnarray} % which satisfy the zero-curvature condition % \begin{equation}\label{ZCCP1} D_\lambda U^1-D_tU^2+[U^1, U^2]=(x_{tt}-6x^2-t)e_1 \nonumber.\end{equation} % Thus, there exists a wave function $\Phi$ of the LSP \eref{LSP'lt} taking values in the group $SL(2,\mathbb{R})$ if and only if $x(t)$ is a solution of the P1 equation \eref{P1}. Let us consider the surface associated with translation in the variable $t$ % \[ F^{1}= \Phi^{-1}D_t \Phi.\] % The tangent vectors to the surface are given by \eref{FAB} with \begin{eqnarray} A=D_tU^1=\left[ \begin {array}{cc} 0&2\,x_{{t}}\\\noalign{\medskip}0&0 \end {array} \right]\in sl(2,\mathbb{R}), \nonumber \\ B=D_tU^2=\left[ \begin {array}{cc} -6\,{x}^{2}-t&1+2\lambda x_{{t}}+4xx_{{t} }\\\noalign{\medskip}-2\,x_{{t}}&6\,{x}^{2}+t\end {array} \right]\in sl(2,\mathbb{R}).\nonumber\end{eqnarray} The first fundamental form for the surface is \begin{eqnarray} I(F^{1})=-4{x_{{t}}}^{2}dtd\lambda+\left(36\,{x}^{4}+12\,{x}^{2}t+{t}^{2}-2\,x_{{t}}-4\,{x_{{t}}}^{2}\lambda-8 \,{x_{{t}}}^{2}x\right)d\lambda^2.\nonumber\end{eqnarray} Note that the tangent vector $D_tF^{1}$ is an isotropic vector. The normal to the surface $F^{1}$ is % \begin{equation} N=\Phi^{-1}\left[ \begin {array}{cc} 1&-{\frac {6\,{x}^{2}+t}{x_{{t}}}} \\\noalign{\medskip}0&-1\end {array} \right]\Phi\in sl(2,\mathbb{R}),\nonumber\end{equation} % which allows for the computation of the second fundamental form and the Gaussian and mean curvatures \begin{eqnarray} II(F^{1})=2x_{{t}}dt^2+8x_{{t}} \left( \lambda-x \right)dtd\lambda\nonumber \\ \qquad +2{\frac { \left( \lambda-x \right)\left( 4{x_{{t}}}^{2}\lambda+8 {x_{{t}}}^{2}x-36{x}^{4}+x_{{t}}-{t}^{2}-12\,{x}^{2}t \right) }{x_ {{t}}}}d\lambda^2,\nonumber \\ K(F^{1})={\frac { \left( -12\,{x_{{t}}}^{2}x-x_{{t}}+36\,{x}^{4}+12\,{x}^{2}t+{ t}^{2} \right) \left( \lambda-x \right) }{{x_{{t}}}^{4}}},\nonumber \\ H(F^{1})= -{\frac {36\,{x}^{4}+12\,{x}^{2}t+{t}^{2}-2\,x_{{t}}+4\,{x_{{t}}} ^{2}\lambda-16\,{x_{{t}}}^{2}x}{4{x_{{t}}}^{3}}}.\nonumber\end{eqnarray} Note that all the points on the line $\lambda=x(t)$, where $x(t)$ is a solution of the Painlev\'e equation P1, are parabolic points ($K=0$). The umbilical points, where the principal curvatures coincide, on the surface satisfy \begin{equation} H^2-K=c_2x_t^{-2}+c_3x_t^{-3} +c_4x_t^{-4} +c_5x_t^{-5} +c_6x_t^{-6}=0,\end{equation} where \begin{eqnarray} c_2=\left( \lambda+2\,x \right) ^{2}, \qquad c_3=3x,\nonumber \\ c_4=-\left(x+\frac{\lambda}{2}\right) (6x^2+t)^2+\frac14, \nonumber \\ c_5=-\frac14 (6x^2+t)^2,\qquad c_6=\frac1{16}(6x^2+t)^4\nonumber.\end{eqnarray} It is straightforward to observe that the umbilical points on the surface $F^1$ lie on the curves \[ \lambda=-2x_t+\frac{(6x^2+t)^2}{4x_t^2}\pm \frac{1}{x_t^2}\sqrt{x_t(6x^2+t)^2-12x_t^2(xx_t-1)},\] whenever $x_t\ne 0.$ % Next we consider the surfaces $F^{2}$ associated with translation in the spectral parameter $\lambda$, \[ F^{2}=\Phi^{-1}D_\lambda \Phi.\] % The tangent vectors to the surface are defined as in \eref{FAB} with \begin{eqnarray} A=D_\lambda U^1= \left[ \begin {array}{cc} 0&1\\\noalign{\medskip}0&0\end {array} \right]\in sl(2,\mathbb{R}) , \nonumber \\ B=D_\lambda U^2= \left[ \begin {array}{cc} 0&4\,\lambda+2\,x\\\noalign{\medskip}2&0 \end {array} \right]\in sl(2,\mathbb{R}).\nonumber \end{eqnarray} The first fundamental form associated with this surface is given by \begin{equation} I(F^{2})=2dtd\lambda+4(x+2\lambda)d\lambda^2.\nonumber\end{equation} Again, the tangent vector $D_tF^{2}$ is an isotropic vector. The normal to the surface is % \begin{equation} N=\Phi^{-1}e_1\Phi\in sl(2,\mathbb{R}).\nonumber \end{equation} % Thus, the image of the surfaces $F^{2},$ written in the moving frame defined by the (nonconstant) wave function $\Phi$, lies in a plane. The second fundamental form and Gaussian and mean curvature for this surface are \begin{eqnarray} II(F^{2})=-dt^2+4(x-\lambda)dtd\lambda+2\left(4x^2+4\lambda x+t-\lambda^2\right)d\lambda^2,\\ K(F^{2})=2(6x^2+t)=2x_{tt},\\ H(F^{2})=2(2x+\lambda).\end{eqnarray} Note that the Gaussian curvature does not depend on the spectral parameter $\lambda$ and in fact the sign of the second derivative of the solution $x(t)$ of the Painlev\'e equation P1 \eref{P1} determines whether the points of the surface are hyperbolic, elliptic or parabolic. Umbilical points of the surface are determined by \[ H^2-K=4(2x+\lambda)^2-2(6x^2+t)=0,\] which are exactly the curves \[ \lambda=-2x\pm \sqrt{\frac{6x^2+t}{2}}=-2x\pm \sqrt{\frac{x_{tt}}{2}}.\] There are no umbilical points in the regimes where $x_{tt}<0.$ Let us now consider the surfaces $F^{5}$ associated with a generalized symmetry of the ZCC, $\Delta[u]=0$. The surface $F^{5}$ is given by \begin{eqnarray} F^{5}=\Phi^{-1}D_tU^1\Phi,\end{eqnarray} with tangent vectors \begin{eqnarray} D_t F^{5}=\Phi^{-1}\left(D_t^2U^1+[D_tU^1,U^1]\right)\Phi,\nonumber \\ \nonumber D_\lambda F^{5}=\Phi^{-1}\left(D_t^2U^2+[D_tU^2,U^1]\right)\Phi.\end{eqnarray} For the particular choices of potential matrices $u^\alpha=U^\alpha([x],\lambda)$ as in \eref{U1P1} and \eref{U2P1}, the tangents vectors are given as in \eref{FAB} with \begin{eqnarray} A=\left[ \begin {array}{cc} 2\,x_{{t}}&12\,{x}^{2}+2\,t \\\noalign{\medskip}0&-2\,x_{{t}}\end {array} \right]\in sl(2,\mathbb{R}), \nonumber\\ B= \left[ \begin {array}{cc} -4\,xx_{{t}}+4\,x_{{t}}\lambda&4\,{x_{{t}}} ^{2}\\\noalign{\medskip}0&4\,xx_{{t}}-4\,x_{{t}}\lambda\end {array} \right]\in sl(2,\mathbb{R}).\nonumber\end{eqnarray} Although the tangent vectors are linearly independent, the first fundamental form on the surface \begin{equation} I(F^{5})=4x_t^2\left(dt^2-2(x-\lambda)dtd\lambda +4(x-\lambda)^2d\lambda^2\right), \nonumber\end{equation} is degenerate ($det(g_{ij})=0$) since the normal to the surface is an isotropic vector. Thus the surface given by the immersion function $F^{5}$ lies in a plane in the moving frame defined by conjugation by $\Phi.$ Let us now use a generalized symmetry of the ODE \eref{P1} to induce the surface. That is, suppose that there exists a generalized vector field in evolutionary representation \begin{equation} \vec{w}_R =R[x]\frac{\partial }{\partial x}\end{equation} which is a symmetry of P1. The determining equation for $R[x]$ is obtained from the second prolongation of the vector field $\vec{w}_R$ applied to the P1 equation \begin{equation} \label{detQP1} D_t^2R-12 xR=0, \qquad \mbox{ whenever } x_{tt}-6x^2-t=0.\end{equation} Expanding the total derivative $D_t$ in terms of partial derivatives with respect to $t, x, x_t$, taken whenever P1 \eref{P1} holds, gives \begin{eqnarray}\fl D_tR&=& \frac{\partial }{\partial t} R+x_t\frac{\partial}{\partial x}R +(6x^2+t)\frac{\partial }{\partial x_t}R,\nonumber\\ \fl D_t^2R&=& \frac{\partial ^2}{\partial t^2}R +2x_t\frac{\partial^2}{\partial t \partial x}R+2(6x^2+t)\frac{\partial^2}{\partial t\partial x_t}R+x_t^2\frac{\partial ^2}{\partial x^2}R +2x_t(6x^2+t)\frac{\partial ^2}{\partial x \partial x_t}R\nonumber \\ \fl &&+(6x^2+t)^2\frac{\partial ^2}{\partial x_t^2}R+(6x^2+t)\frac{\partial }{\partial x}R+(12xx_t+1)\frac{\partial}{\partial x_t}R\nonumber.\end{eqnarray} Thus, the determining equation \eref{detQP1} becomes a linear, analytic second-order PDE for $R$ \begin{eqnarray}\fl \frac{\partial ^2}{\partial t^2}R +2x_t\frac{\partial^2}{\partial t \partial x}R+2(6x^2+t)\frac{\partial^2}{\partial t x_t}R+x_t^2\frac{\partial ^2}{\partial x^2}R +2x_t(6x^2+t)\frac{\partial ^2}{\partial x \partial x_t}R\label{detQP1exp}\\ \!\!\!\!\!\!\!\!\!\!\!\! +(6x^2+t)^2\frac{\partial ^2}{\partial x_t^2}R+(6x^2+t)\frac{\partial }{\partial x}R+(12xx_t+1)\frac{\partial ^2}{\partial x_t}R-12xR=0 \nonumber .\end{eqnarray} The construction of generalized symmetries of the Painlev\'e equation P1 requires a solution for $R$ as a function of $t, x, x_t$. For equation \eref{detQP1exp}, the Cauchy-Kovalevskaya theorem ensures that analytic solutions exist provided that the initial data is analytic (see e.g. \cite{FJon, Olver}). So, in what follows, we assume that the quantity $R$ is a given function of $t, x,$ and $ x_t$ satisfying \eref{detQP1exp}. As proven in section \ref{integrable}, if the vector field $\vec{w}_R$ is a symmetry of the P1 equation \eref{P1} then the matrices % \begin{eqnarray} A=pr\vec{w}_R(U^1)= \left[ \begin {array}{cc} 0&2\,R\\\noalign{\medskip}0&0\end {array} \right]\in sl(2,\mathbb{R}), \nonumber\\ B=pr\vec{w}_R(U^2)=\left[ \begin {array}{cc} -D_{{t}} R &R \left( 2\, \lambda+4\,x \right) \\\noalign{\medskip}-2\,R&D_{{t}} R \end {array} \right]\in sl(2,\mathbb{R}), \nonumber\end{eqnarray} % satisfy \eref{AB'} and hence there exists a surface $F^6$ with tangent vectors given by \eref{FAB}. The first fundamental form of this surface is % \begin{eqnarray} I(F^{6})=-4\,{R}^{2}dtd\lambda+\left( \left( D_{{t}}R \right) ^{2}-4\,{R}^{2}\lambda-8\,{R }^{2}x\right)d\lambda^2,\nonumber\end{eqnarray} Note that, in the surface $F^{6}$ the tangent vector in the direction $D_t$ is again an isotropic vector. The normal vector % \begin{eqnarray} N=\Phi^{-1}\left[ \begin {array}{cc} 1&-D_t\left(\ln (R)\right) \\\noalign{\medskip}0&-1\end {array} \right]\Phi,\nonumber\end{eqnarray} % allows for the construction of the second fundamental form and Gaussian and mean curvatures \begin{eqnarray} \fl II(F^{6})=2 R dt^2+8 R \left( \lambda-x \right) dtd\lambda\nonumber \\ \fl \qquad +\frac{2}{R}\left( 4{R}^{2}{\lambda}^{2}+4 {R}^{2}\lambda x+{R}^{2}t-2{R }^{2}{x}^{2}-(D_{{t}} R ) x_{{t}}R- (x+\lambda)\left( D_{{t}} R \right) \right) ^{2}d\lambda^2,\nonumber\\ \fl K(F^{6})={\frac{-1}{R^{4}}}\left({12\,{R}^{2}\lambda\,x+{R}^{2}t-6{R}^{2}{x}^{2}- \left( D_{{t}} R \right) x_{{t}}R- \left( D_{{t}} R \right) ^ {2}\lambda+ \left( D_{{t}} R \right) ^{2}x}\right),\nonumber \\ \fl H(F^{6})=\frac{-1}{4R^3} \left(\left( D_{{t}} R \right) ^{2}+4\,{R}^{2 }\lambda-16\,{R}^{2}x \right)\nonumber.\end{eqnarray} The mean and Gaussian curvatures depend on the Painlev\'e trancendent P1 and on the solution of its associated determining equation \eref{detQP1exp}. The umbilical points of the surface $F^6$ satisfy \[ H^2-K=c_{4}D_t^4R+c_2D_t^2R+c_1D_tR +c_0=0,\] where \begin{eqnarray} c_4=\frac{1}{16R^6}, \qquad c_2=-\frac{2x+\lambda}{2R^4},\nonumber \\ c_1=\frac{-x_t}{R^3}, \qquad c_0=\frac{\lambda^2+4\lambda x+10x^2+t}{R^2}\nonumber,\end{eqnarray} the solution of which lie on the curves on the surface $F^6$ \[ \lambda=\frac{1}{4}\left(D_t(\ln R)\right)^2-8x\pm \sqrt{x_t D_t(\ln R)-x_{tt}}\] \subsection{Painlev\'e P2} In this section, we give several examples of surfaces associated with the Lax pair for the second Painlev\'e equation P2 \begin{equation} \label{P2} x_{tt}=2x^3+tx-\alpha.\end{equation} A Lax pair for this equation is given by \cite{Garn1960, ItsKap1988} \begin{eqnarray} U^{1}= \left[ \begin {array}{cc} -\lambda&x\\\noalign{\medskip}x&\lambda \end {array} \right] \in sl(2,\mathbb{R}),\nonumber\\ U^{2}= \left[ \begin {array}{cc} 4\,{\lambda}^{2}-2\,{x}^{2}-t&-4\,x\lambda+ {\frac {\alpha}{\lambda}}+2\,x_{{t}}\\\noalign{\medskip}-4\,x\lambda+{ \frac {\alpha}{\lambda}}-2\,x_{{t}}&-4\,{\lambda}^{2}+2\,{x}^{2}+t \end {array} \right] \in sl(2, \mathbb{R}) ,\nonumber \end{eqnarray} First, we consider surfaces $F^{1}$ associated with translation in the $t$-direction, \begin{equation} F^{1}=\Phi^{-1}D_t \Phi\in sl(2, \mathbb{R}),.\end{equation} % The tangent vectors to the surface are defined as in \eref{FAB} % \begin{eqnarray} D_tF^{1}=\Phi^{-1}D_t U^{1}\Phi, \nonumber \\ D_\lambda F^{1}=\Phi^{-1} D_t U^{2}\Phi^{-1},\nonumber \end{eqnarray} % with matrices % \begin{eqnarray} A=D_t U^{1}=\left[ \begin {array}{cc} 0&ix_{{t}}\\\noalign{\medskip}-ix_{{t}}&0 \end {array} \right], \nonumber \\ B=D_t U^{2} =\left[ \begin {array}{cc} -i \left( 1+4\,xx_{{t}} \right) &4\,ix_{{t} }\lambda+4\,{x}^{3}-2\,{x}^{2}+2\,\alpha\\\noalign{\medskip}-4\,ix_{{t }}\lambda+4\,{x}^{3}-2\,{x}^{2}+2\,\alpha&i \left( 1+4\,xx_{{t}} \right) \end {array} \right].\nonumber \end{eqnarray} % The corresponding first fundamental form of the surface is given by \begin{eqnarray}\fl\nonumber I(F^{1})={x_{{t}}}^{2}dt^2 -8\lambda\,{x_{{t}}}^{2} dtd\lambda\\ +\left(16\left( {x}^{2}+{\lambda}^{2} \right) {x_{{t}}}^{2}+8\,xx_{{t} }- 4\left( \,\alpha-2\,{x}^{3}-\,{x}^{2} \right)^2+1 \right)d\lambda^2,\nonumber\\ \fl \nonumber det\left(g_{ij}(F^{1})\right)={x_{{t}}}^{2} \left((1+4xx_t)^2-4(\alpha-tx-2x^3)^2\right). \end{eqnarray} The (unnormalized) normal vector \begin{eqnarray} N(F^{1})&=&\Phi^{-1}\left[ \begin {array}{cc} -4\,{x}^{3}-2\,xt+2\,\alpha&1+4\,xx_{{t}} \\\noalign{\medskip}-4\,xx_{{t}}-1&4\,{x}^{3}+2\,xt-2\,\alpha \end {array} \right] \Phi ,\end{eqnarray} % is used to determine the second fundamental forms of the surface % \begin{eqnarray}\fl II(F^{1})&=& -2\lambda \left( 1+4\,xx_{{t}} \right) x_{{t}}dt^2+4\,x_{{t}} \left( -4\,x_{{t}}\alpha-2\,{x}^{2}+4\,{\lambda}^{2}-t+16\, xx_{{t}}{\lambda}^{2} \right) dtd\lambda\nonumber \\\fl && +\lambda^{-1}\Bigg[ 32\,{ { \left( {\lambda}^{2}+{x}^{2} \right) \left( -4\,x{ \lambda}^{2}+\alpha \right) {x_{{t}}}^{2}}}+8\,{ { \left( -6\,{x}^{2}{\lambda}^{2}-4\,{\lambda}^{4}+{\lambda}^{2}t+2\, \alpha\,x \right) x_{{t}}}}\nonumber \\ \fl && -2\,{ { \left( 2\,\alpha+1-2 \,xt-4\,{x}^{3} \right) \left( 2\,\alpha-1-2\,xt-4\,{x}^{3} \right) \left( -4\,x{\lambda}^{2}+\alpha \right) }}\Bigg]d\lambda^2.\nonumber \end{eqnarray} From the fundamental forms, it is possible to compute the Gaussian and mean curvatures. The umbilical points on the surface satisfy a polynomial equation in $\lambda,$ $x$ and $x_t.$ In each case, the formulas are too involved to be presented in an instructive manner. Consider next the surfaces associated with translation in the spectral parameter $\lambda$ \begin{equation} F^{2}=\Phi^{-1}D_\lambda \Phi\in sl(2, \mathbb{R}).\end{equation} % The tangent vectors to the surface are defined as in \eref{FAB}, % \begin{eqnarray} D_tF^{2}=\Phi^{-1}D_\lambda U^{1}\Phi, \nonumber \\ D_\lambda F^{2}=\Phi^{-1}D_\lambda U^{2}\Phi\nonumber,\end{eqnarray} % with matrices \begin{eqnarray} A= D_\lambda U^{1}=\left[ \begin {array}{cc} -1&0\\\noalign{\medskip}0&1\end {array} \right], \nonumber \\ B=D_\lambda U^{2} =\left[ \begin {array}{cc} 8\,\lambda&-4\,x-{{\alpha}{{\lambda}^ {-2}}}\\\noalign{\medskip}-4\,x-{{\alpha}{{\lambda}^{-2}}}&-8\, \lambda\end {array} \right].\nonumber \end{eqnarray} % The first fundamental form for this surfaces immersed in $sl(2, \mathbb{R})$ is \begin{eqnarray} I\left(F^{2}\right)=dt^2-16\lambda dtd\lambda+\left(\left(\frac{\alpha}{\lambda^2}+4x\right)^2+64\lambda^2\right)d\lambda^2,\nonumber\\ det\left(g_{ij}\left(F^{2}\right)\right)=\left(\frac{\alpha}{\lambda^2}+4x\right)^2.\label{gF2} \end{eqnarray} In the moving frame defined by conjugation by the (non-constant) wave function $\Phi$, the normal to the surface is constant and so the images of the immersion function $F^{2}$ is contained in a plane. The second fundamental form, Gaussian and mean curvatures for the surface in $sl(2, \mathbb{R})$ is given by \begin{eqnarray} II(F^{2})=2xdt^2-4(4x\lambda-\alpha\lambda^{-1})dtd\lambda\nonumber \\ \qquad\qquad +2\left(8x^3+2\alpha\lambda^{-2} x^2+(16\lambda^{2}+4t)x-\alpha\lambda^{-2}(12\lambda^2-t)\right)d\lambda^2,\nonumber \\ K(F^{2})=\frac{4\lambda^2 x_{tt}}{4x\lambda^2+\alpha},\nonumber\\ H(F^{2})=\frac{4\lambda^4+\lambda^2(6x^2+t)+\alpha x}{4x\lambda^2+\alpha}.\nonumber\end{eqnarray} Note here the sign of the second derivative of the solution $x(t)$ of the Painlev\'e equation P2 \eref{P1} as well as square root of the determinant of the metric \eref{gF2} determines whether the points of the surface are hyperbolic, elliptic or parabolic. The umbilical points of the surface $F^{2}$ satisfy \[\fl 16{\lambda}^{8}+ 8\left(6{x}^{2}+t \right) {\lambda}^{6}+ \left( 4{x}^{4}-4t {x}^{2}+24x\alpha+{t}^{2} \right) {\lambda}^ {4}+2\alpha \left( 2{x}^{3}-tx+2\alpha \right) {\lambda}^{2}+{ x}^{2}{\alpha}^{2}=0. \] Next, consider surfaces associated with generalized symmetries of the Painlev\'e P2 equation. Since we do not have explicit forms of the wave functions $\Phi$ which satisfy the LSP \eref{LSP'}, it is not possible to say whether the integrated form of the immersion function \eref{F6} holds. However, the geometry of the surfaces can be studied via the tangent vectors, defined as in \eref{FAB}, % \begin{eqnarray}\label{tanQP2a} D_tF^{6}=\Phi^{-1}\left(pr\vec{w}_R U^{1}\right)\Phi, \\ \label{tanQP2b} D_\lambda F^{6}=\Phi^{-1}\left(pr\vec{w}_R U^{2}\right)\Phi.\end{eqnarray} % Here $R$ is the characteristic of the vector field $\vec{w}_R$ written in evolutionary form \eref{wQ} and hence is required to satisfy the linear determining equation \begin{equation} D_t^2R=6x^2R+tR,\qquad \mbox{ whenever } x_{tt}=2x^3+tx-\alpha.\end{equation} % The matrices in \eref{tanQP2a} and \eref{tanQP2b} are given by \begin{eqnarray*} A=pr\vec{w}_R U^{1}=\left[ \begin {array}{cc} 0&R\\\noalign{\medskip}R&0\end {array} \right], \\ B= pr\vec{w}_R U^{2}= \left[ \begin {array}{cc} -4\,xR &-4\,\lambda\,R+2\,D_{{t}} \left( R \right) \\\noalign{\medskip}-4\,\lambda\,R-2\,D_{{t}} \left( R \right) &4\,xR\end {array} \right]. \end{eqnarray*} The first fundamental form and determinant of the induced scalar product are \begin{eqnarray} I(F^{6})={R}^{2}dt^2-8\lambda R^2dtd\lambda -4\left((D_tR)^2-4(x^2+\lambda^2)R^2\right)d\lambda^2,\nonumber \\ det\left(g_{ij}(F^{6})\right)=-4R^2\left((D_tR)^2-4x^2R^2\right).\nonumber \end{eqnarray} The second fundamental forms and Gaussian and mean curvatures are \begin{eqnarray} \fl II\left(F^{6}\right)=\frac{1}{\sqrt{(D_tR)^2-4x^2R^2}}\Bigg[4\lambda xR^2dt^2-8\,R \left( x_{{t}}D_{{t}} R +4\,xR{\lambda}^{2}-2{x}^{3}R-{x}^{2}R \right) dtd\lambda \nonumber \\ \fl \qquad + \frac1\lambda\Bigg(\left( -16\,x{\lambda}^{2}+4\,\alpha \right) \left( D_{{t}} R \right) ^{2}+16 x_{{t}}{\lambda}^{ 2}R D_{{t}}R -16\,{R}^{2}x \left( -x{\lambda}^{2}+2\,{x}^{2}{\lambda}^{2}+\alpha \,x-4\,{\lambda}^{4} \right) \Bigg)d\lambda^2\Bigg],\nonumber \\ \fl K\left(F^{6}\right)=\frac{-1}{ \left((D_tR)^2-4x^2R^2 \right) ^{2}}\Bigg( 4\left(- 4 {x}^{2}{\lambda}^{2}+\alpha x-{x_{{t}}}^{2} \right) \left( D_{{t}}R \right) ^{2}+2xR x_{{t}} \left( 2{\lambda}^{2}-2{x}^{2}-x \right) D_{{t}} R\nonumber \\ +{R}^{2}{x}^{3} \left( 4\,{x}^{3}+4\,{x}^{2}-24\,x{\lambda}^{ 2}+x+4\,\alpha-4\,{\lambda}^{2} \right) \Bigg),\nonumber \\ \fl H\left(F^{6}\right)=\frac{1}{2\lambda \left(D_t(R)^2-4x^2R^2\right)^{\frac32}}\Bigg(\left( 8\,x{\lambda}^{2}-\alpha \right) \left( D_{{t}} R \right) ^{2}+4 x_{{t}}{\lambda}^{2 }R D_{{t}} R -4{x}^{2}{R}^{2} \left( 10\,x{\lambda}^{2}+{\lambda}^{2}-\alpha \right) \Bigg).\nonumber\end{eqnarray} In particular, we can use classical solutions of P2 to find explicit forms of the soliton surfaces associated with particular solutions. For example, it has been shown in \cite{Gromak1999}, that Painlev\'e P2 has rational solutions for integer values of $\alpha$. These rational solutions have recently emerged in studying the small dispersion or semi-classical limit of the KdV and sine-Gordon equations respectively \cite{BucMil2011, ClaGra2010}. Below, we list the first two rational solutions of the Painlev\'e equation P2 (\ref{P2}) along with the characteristic of vector fields, $R,$ which satisfy the determining equation \begin{equation}\label{detQP2} D_t^2R-(6x^2+t)R=0, \end{equation} for the given integer values of $\alpha$: \begin{equation}\fl \begin{array}{llll} &\alpha=1\qquad & x=t^{-1} & R=\sqrt{t} I_\frac{5}3 (\frac{2t^\frac32}{3})\\ \fl& &\mbox{ or} & R=\sqrt{t} K_\frac{5}3 (\frac{2t^\frac32}{3})\\ \mbox{and} &\alpha=2 \qquad &x=-\frac{2(t^3-2)}{t(t^3+4)} & R=\frac{1}{t^2(t^3+4)^2}HeunC(0,-\frac53,-5, \frac49,\frac{73}{18},-\frac{t^3}{4})\\ \fl &&\mbox{ or} & R=\frac{t^3}{(t^3+4)^2}HeunC(0,\frac53,-5, \frac49,\frac{73}{18},-\frac{t^3}{4})\end{array} \nonumber\end{equation} Here $I_\nu(z), K_\nu(z)$ are modified Bessel functions \cite{Yellowbook} and $HeunC(\alpha, \beta, \gamma, \delta, \mu, t)$ is a solution of the confluent Heun equation \cite{Ronveaux}. The graphs of surfaces given by immersion functions $F^1$ and $F^2$ are displayed in figure 1 for the first two rational solutions of P2. \begin{figure}\label{gneg} \begin{center}$ \begin{array}{cc} \includegraphics[width=2.5in]{F1R1notfilledin50x.jpg}& \includegraphics[width=2.5in]{F2R1notfilledin50x.jpg} \\ F^{1}: \alpha=1 ,&F^{2}: \alpha=1,\\ \includegraphics[width=2.5in]{F1R2notfilledin50x}& \includegraphics[width=2.5in]{F2R2notfilledin50x} \\ F^{1}: \alpha=2 ,&F^{2}: \alpha=2,\end{array}$ \end{center} \caption{Surfaces $F^{1}$ and $F^{2}$ for rational solutions of P2. The domain of parameterization is chosen to be $t\in [-30,30],$ $\lambda \in [-30,30]$ except for an omitted range of $\pm .01$ around the singularities $\lambda=0$ and $t=0$ for the first rational solution and $t=0,-2^{2/3}$ for the second rational solution. The viewing window for the graphs are $[-50,50]^3$ and the axes are the components of surface in the basis \eref{basis}. } \end{figure} For the case $\alpha=\epsilon / 2$ with $\epsilon ^2=1$, a solution of P2 can be written in terms of the Airy function with \cite{AblSeg1977, Gromak1999} \begin{equation} x(t)=-\epsilon\frac{d}{dt}\ln\left(Ai\left(2^{-\frac13}t\right)\right),\qquad \epsilon^2=1,\end{equation} where $Ai(\cdot)$ is the Airy function. In this case, $x(t)$ satisfies the first-order differential equation \begin{equation} \label{P212} x_t=\epsilon x^2+\epsilon \frac t2.\end{equation} The first fundamental forms and normals for surfaces $F^1 $ and $F^6$, in the case $\alpha=\epsilon/2$, are given by \begin{eqnarray} I\left(F^{1}\right)=-(2x^2+t)^2\left(\frac{1}4dt^2-2\lambda dt d\lambda+4\lambda^2d\lambda^2\right),\\ N\left(F^{1}\right)=\Phi^{-1}\left[\begin{array}{cc} 1 &-1\\ 1 &-1\end{array} \right]\Phi,\end{eqnarray} and \begin{eqnarray} I\left(F^{6}\right)=-R^2\left(dt^2-4\lambda dt d\lambda +4\lambda^2d\lambda^2\right),\\ N\left(F^{6}\right)=\Phi^{-1}\left[\begin{array}{cc} 1 &-1\\ 1 &-1\end{array} \right]\Phi,\end{eqnarray} respectively. Here $R$ satisfies the determining equation for a symmetry of \eref{P212}, \begin{eqnarray} \label{detP212} D_tR=\epsilon 2xR.\\ \Rightarrow\frac{D_t R}{R}=2\frac{d}{dt}\ln\left(Ai\left(2^{-\frac13}t\right)\right).\end{eqnarray} The solution for this $R$ is given by \[ R =cAi(2^{-\frac13}t)^2. \] Note that a solution $R$ of \eref{detP212} is also a solution of the general determining equation for a symmetry of P2 \eref{detQP2} whenever $x(t)$ is a solution of \eref{P212}. Indeed, \begin{eqnarray} D_t^2R &=&D_t(\epsilon 2xR)=(6x^2+t)R .\end{eqnarray} \begin{figure}\label{gneg} \begin{center}$ \begin{array}{c} \includegraphics[width=2.5in]{P3rational.jpg},\end{array}$ \end{center} \caption{Surface $F^{6}$ with $\alpha=0, \beta=1, \gamma=2/5, \delta=0$ and rational solution for $x(t).$ The domain of parameterization is chosen to be $t\in [-30,30],$ $\lambda \in [-30,30]$ except for an omitted range of $\pm .01$ around the singularities $\lambda=0$ and $t=0$. The viewing window for the graph is $[-50,50]^3$ and the axes are the components of surface in the basis \eref{basis}. } \end{figure} The surfaces $F^{1}$ and $F^{6}$, obtained from different symmetries, are conformally equivalent and are contained in planes with isotropic normal vectors in the moving frame defined by conjugation by $\Phi$. Note that, with the restriction $\alpha=\epsilon/2,$ the ODE \eref{P212} and determining equation for $R$ \eref{detP212} are exactly those required so that the metrics induced by the Killing form on the tangents to the surfaces $F^{1}$ and $F^{6}$ are degenerate, i.e. the determinants of the first fundamental forms are zero. As above, the tangent vectors to the surface are linearly independent and so the immersion functions define surfaces instead of a curves. A graph of the surface $F^6$ for the symmetry $\vec{w}_{R_1}$ is given in figure 2. \subsection{Painlev\'e P3} In this section, we consider the cases of the third Painlev\'e equation P3 which admit point symmetries and their associated surfaces. In these cases, we use the known solutions of the determining equations for the vector field to obtain explicit expressions for the geometry of such surfaces. The third Painlev\'e equation, P3, is given by \begin{equation} \label{P3} x_{tt}=\frac{(x_t)^2}{x}-\frac{x_t}{t}+\frac1t(\alpha x^2+\beta)+\gamma x^3+\frac{\delta}{x}.\end{equation} The LSP is given in terms of the potential matrices $U^1(\lambda, [x]) $ and $U^2(\lambda, [x])$ taking values in the Lie algebra $sl(2, \mathbb{R})$ \cite{ConteMusette, Garn1960, Gromak1999} \begin{eqnarray}\fl U^1(\lambda, [x])=\frac12 \left[ \begin {array}{cc} \,{\frac {x_{{t}}}{x}}+\,\gamma\,x+\,{\frac {\delta}{x}}&2\lambda\\\noalign{\medskip}2\lambda&-\,{ \frac {x_{{t}}}{x}}-\,\gamma\,x-\,{\frac {\delta}{x}} \end {array} \right],\\ \fl U^2(\lambda, [x])= \left[ \begin {array}{cc} {\frac {2\,t{\lambda}^{2}x_{{t}}+2\,t{ \lambda}^{2}\gamma\,{x}^{2}+2\,t{\lambda}^{2}\delta-x\alpha\,\delta+x \beta\,\gamma}{4\lambda\,x \left( {\lambda}^{2}+\gamma\,\delta \right) }} &-{\frac {-2\,t{\lambda}^{2}+\gamma\,tx_{{t}}+\gamma\,x+{\gamma }^{2}t{x}^{2}-\gamma\,\delta\,t+\alpha\,x}{2({\lambda}^{2}+\gamma\, \delta)}}\\ \noalign{\medskip}-{\frac {-2\,t{x}^{2}{\lambda}^{2}+ \delta\,tx_{{t}}-\delta\,x+{\delta}^{2}t-\gamma\,\delta\,t{x}^{2}- \beta\,x}{2{x}^{2} \left( {\lambda}^{2}+\gamma\,\delta \right) }} &- {\frac {2\,t{\lambda}^{2}x_{{t}}+2\,t{\lambda}^{2}\gamma\,{x}^{2}+2 \,t{\lambda}^{2}\delta-x\alpha\,\delta+x\beta\,\gamma}{4\lambda\,x \left( {\lambda}^{2}+\gamma\,\delta \right) }}\end {array} \right] ,\end{eqnarray} which satisfy the zero-curvature condition \begin{eqnarray}\fl\label{ZCCP2} \Omega[x]=D_\lambda U^1-D_tU^2+[U^1, U^2]\nonumber \\ \fl \qquad =\frac12\left(x_{tt}-\frac{(x_t)^2}{x}+\frac{x_t}{t}-\frac1t(\alpha x^2+\beta)-\gamma x^3-\frac{\delta}{x}\right)\left[ \begin {array}{cc} \,{\frac {-\lambda\,t}{x \left( {\lambda }^{2}+\gamma\,\delta \right) }}&\,{\frac {\gamma\,t}{{\lambda}^{2}+ \gamma\,\delta}}\\\noalign{\medskip}\,{\frac {\delta\,t}{{x}^{2} \left( {\lambda}^{2}+\gamma\,\delta \right) }}&\,{\frac {\lambda\, t}{x \left( {\lambda}^{2}+\gamma\,\delta \right) }}\end {array} \right]=0 .\end{eqnarray} It has been shown, \cite{Gromak1999}, that one can find solutions of Painlev\'e P3 under the restrictions $\beta=\delta=0$ or $\alpha=\gamma=0$. Consider the former case, $\beta=\delta=0.$ With these restrictions (\ref{P3}) admits a point symmetry given in evolutionary form by \begin{equation} \vec{w}_{R_1}=(x+tx_t)\frac{\partial}{\partial x}.\end{equation} In the case $\alpha=\gamma=0$, the point symmetry of the reduced equation is \begin{equation} \vec{w}_{R_2}=(x-tx_t)\frac{\partial}{\partial x}.\end{equation} % Below, we consider the two surfaces defined by these symmetries, which we index by a superscript indicating the considered integrable case, \begin{eqnarray} A=pr\vec{w}_{R_1}(U^1)=\frac{{\gamma}^{2}t{x}^{2}+ \left( \gamma+\alpha \right) x+\gamma\,tx_{{t}}}2\left[ \begin {array}{cc}1 &0\\\noalign{\medskip}0&-1\end {array} \right]\in sl(2,\mathbb{R}), \nonumber \\ B= pr\vec{w}_{R_1}(U^2)=\frac{{\gamma}^{2}t{x}^{2}+ \left( \gamma+\alpha \right) x+\gamma\,tx_{{t}}}2 \left[ \begin {array}{cc} t\lambda^{-1}&-{\frac {x+tx_{{t}}+ \gamma\,t{x}^{2}}{x{\lambda}^{2}}}\\\noalign{\medskip}0&-t\lambda^{-1}\end {array} \right] \in sl(2,\mathbb{R}), \nonumber \\ I(F^{R_1})=\frac{\left({\gamma}^{2}t{x}^{2}+ \left( \gamma+\alpha \right) x+\gamma\,tx_{{t}}\right)^2}{4\lambda^2}\left(\lambda dt+td\lambda\right)^2,\nonumber \\ N(F^{R_1})=\Phi^{-1}\left[\begin{array}{cc}0 &1\\0 &0\end{array}\right]\nonumber\Phi, \end{eqnarray} and in the second case \begin{eqnarray} A=pr\vec{w}_{R_2}(U^1)=\frac{\delta\,tx_{{t}}-x\delta+\beta\,x+{\delta}^{2}t}{2x^2} \left[ \begin {array}{cc} 1&0\\\noalign{\medskip}0&-1\end {array} \right]\in sl(2,\mathbb{R}), \nonumber \\ B=pr\vec{w}_{R_2}(U^2)=\frac{\delta\,tx_{{t}}-x\delta+\beta\,x+{\delta}^{2}t}{2x^2}\left[ \begin {array}{cc}t\lambda^{-1}&0 \\\noalign{\medskip}-{\frac {tx_{{t}}-x+\delta\,t}{{\lambda}^{2}{x }}}&-t\lambda^{-1}\end {array} \right]\in sl(2,\mathbb{R}), \nonumber \\ I(F^{R_2})=\frac{(-\delta\,tx_{{t}}+x\delta+\beta\,x-{\delta}^{2}t)^2}{4x^2\lambda^2}\left(\lambda dt+td\lambda\right)^2,\nonumber \\ N(F^{R_2})=\Phi^{-1}\left[\begin{array}{cc}0 &0\\1 &0\end{array}\right]\Phi\nonumber.\end{eqnarray} As in several of the previous cases, in the moving frame defined by conjugation by the (non-constant) wave function $\Phi$, the surfaces are contained in planes with isotropic normal vectors. Note that the Killing form on the tangents to the surfaces is degenerate, i.e. the determinants of the first fundamental forms are zero, however the tangent vectors are linearly independent and so the immersion functions define surfaces instead of curves. \section{Concluding remarks} The main objective of this paper is to extend the applicability of the Fokas-Gel'fand procedure for constructing explicit soliton surfaces associated with integrable ODEs admitting zero-curvature representation. The most important advantage of this method is that it gives effective tools for constructing certain classes of soliton surfaces in a systematic way. It is much simpler and faster than traditional methods and its effectiveness has been demonstrated by the results obtained in section 5 for Painlev\'e equations P1-P3. At this point, we can summarize our approach for constructing the immersion function for 2D-surfaces immersed in Lie algebras. \begin{itemize} \item[1.] We have provided a general framework for soliton surfaces whose Gauss-Mainardi-Codazzi equations are integrable deformations of the ZCC. This characterization includes three sub-cases, namely integrable PDEs in ZCC form, integrable ODEs in Lax representation and in ZCC form (see sections 2 and 3). \item[2.] We have provided a complete classification of all admissible generalized symmetries of the ZCC using the associated LSP. As a result, we have shown that any 2D-surface immersed in a Lie algebra is associated with a generalized symmetry of the ZCC (see section 2 Proposition 1 ''Extracting symmetries from a surface"). As a corollary, we have shown that all generalized symmetries of the ZCC can be written in terms of a gauge function on jet space. \item[3.] We have considered in depth the case of ODEs admitting Lax pairs in zero-curvature form and their associated surfaces, defined either explicitly by their tangent vectors or, when possible, from an explicit integrated form. \item[4.] To illustrate these theoretical considerations, we consider Lax pairs for Painlev\'e equations in zero-curvature form whose potential matrices take values in the Lie algebra $sl(2,\mathbb{R})$. For the considered equations, we explore some geometric characteristics of the surfaces. These include the first and second fundamental forms, whose coefficients are not independent but subject to the differential constraints of the Painlev\'e equations P1, P2, and P3. The fundamental forms were then used to construct expressions for the Gaussian and mean curvatures for the surfaces; in each case the Gaussian and mean curvatures are functionally independent. We also computed the umbilical points of several of the surfaces and in each case, the points lie on curves determined by polynomial equations in the spectral parameter. It is interesting to note that several of the obtained surfaces lie in planes in the (non-constant) moving frame defined by conjugation by the wave function $\Phi.$ In the case of the P2 equation, we also consider specific rational and Airy function solutions. In the first case, for rational solutions, we were able to solve the determining equations for infinitesimal generators of symmetries and consequently the corresponding surfaces are expressed in terms of Bessel and confluent Heun functions. \item[5.] We have described the integrable cases of the P3 equation, their associated point symmetries and their soliton surfaces. We have shown that the tangent vectors to the surfaces are planes conjugated by the (non-constant) wave function $\Phi.$ \end{itemize} Finally, it is worth noting that the proposed approach for constructing 2D-surfaces associated with integrable models can be used "in reverse" to address certain physical problems. Namely, it is sometimes the case that a 2D-surface is known in terms of functions appearing in the physical system for which analytical models are not yet fully developed. Using our approach (Proposition 1), it is possible to select an appropriate surface immersed in a Lie algebra and evaluate some geometric properties such as the induced metric, mean curvature, and the corresponding Willmore functional. A variational problem of this functional allows for the straightforward computation of the class of equations describing the physical problem. These applications and further theoretical issues will be explored in our future works. \ack The authors thank Professor R. Conte for helpful discussions on this topic. The work was supported by a research grant from NSERC of Canada. S Post acknowledges a postdoctoral fellowship provided by the Laboratory of Mathematical Physics of the Centre de Recherches Math\'ematiques, Universit\'e de Montr\'eal. \section*{References}
\section{SED fitting with MCMC: a two-step process} SED fitting is the process of extracting information on the physical properties of galaxies, such as stellar population age, mass, star formation rate, dust content, metallicity, and redshift, starting from a set of templates that predict how galaxy spectra look like as a function of these properties, which are the SED fitting parameters. This process relies on the simple but powerful idea that since the properties of the models are known, if we can find models that resemble the observations we can infer the properties of the data. A more rigorous way of comparing models with observations -- in other words, deciding whether a model resembles the data or not - is the $\chi^2$ statistics. For each set of parameters, we need to compute the prediction of what the observations would be if that model was the true one. This step is conceptually simple but complicated in practice, because of the many astrophysical processes that needs to be modeled. GalMC implements this process through the sequence described in Fig. 1. GalMC is based on Bayesian statistics. Therefore, a second step of the inference process requires to reconstruct the probability distribution of the SED fitting parameters, which are treated as random variables. This is done by exploring the parameter space with a random walk biased so that the frequency of visited locations is proportional to the probability density function. This path through parameter space is the Markov Chain. Once these probabilities are known, one can compute the desired credible intervals for each of the parameters; and because of how visited locations are chosen, integrating the probability distribution functions (PDFs) becomes a simple matter of summing over the points in the chains. \section{Probability distributions, degeneracies, and the impact of systematics} Detailed results for two stacked samples of Lyman Alpha Emitting galaxies at $z \sim 3$ were presented in \cite[Acquaviva et al (2011)]{2011arXiv1101.2215A}. Here we just show that GalMC is able to capture multi-modal probability distributions, such as the double peak in the Age vs Stellar Mass distribution, which is due to the degeneracy between these two parameters. We also want to highlight the impact of the assumptions made in modeling the stellar populations, shown for the case of Stellar Mass in the right panel of Fig. 2. The different curves all refer to models commonly used in the literature: the BC03 (\cite{BC03}) and CB07 (\cite{CB07}) stellar population templates, at Solar or variable metallicity, and with or without including nebular emission. The corresponding scatter in the estimate of stellar mass (which does not include the possibility of different initial mass functions, IMFs) is a factor of $\sim$ 2.5, significantly larger than the statistical uncertainty for the same data. \begin{figure}[t] \begin{center} \includegraphics[width=\linewidth]{ChainSteps.jpg} \caption{The series of steps performed by GalMC to obtain the predicted spectrum as a function of the SED parameters. After the convolution with filters transmission curves, this quantity can be directly compared to the data to obtain a $\chi^2$ value. } \end{center} \end{figure} \section{SpeedyMC: MCMC for large galaxy catalogs} MCMC algorithms are much more efficient ways of exploring high-dimensional parameter spaces with respect to algorithms where the probability distribution is sampled at a set of fixed locations on a grid. In fact, the ``interesting" region of parameter space (the one where data and models look like each other) often occupies a small fraction of the total volume. While grid-based models need to explore all of it, Markov Chains are able to ``recognize" the interesting regions and will spend most of the time visiting (sampling) those locations. Yet, the complicated process described in Fig. 1, which leads to the computation of the $\chi^2$ value corresponding to a set of parameters, usually needs to be repeated tens of thousands of times. The computational bottlenecks in this case are the generation of a stellar population template at the right age, and the convolution with the filter transmission curves. To alleviate the first problem, GalMC uses our modified version of GALAXEV (Bruzual and Charlot 2011), developed in collaboration with the authors, which is $\sim$ 20 times faster than the official release. However, the typical time per iteration is still about 0.4 seconds on a 2.2GHz MacBook Pro laptop (for simplicity, all quoted running times will be referred to this machine), and therefore the typical chain per object takes a few hours to run. This becomes impractical for catalogs comprising thousands of objects. The basic idea of SpeedyMC is to find a different (faster) way to compute the $\chi^2$ corresponding to a certain set of parameters. To achieve this objective, we take the following four steps: \begin{figure}[t!] \begin{center} \includegraphics[width=5cm]{EmLinesRescaledBF.pdf} \includegraphics[width=4.2cm]{SED_part1.pdf} \includegraphics[width=3.56cm]{Masses.pdf} \label{sed} \caption{\small{ {\it Left}: Data and best-fit model of the z = 3.1 Lyman Alpha Emitters from \cite{2011arXiv1101.2215A}. {\it Middle}: Marginalized constraints on age and stellar mass. The contours indicate the 68\% and 95\% credible regions, while the color gradient is based on average likelihood in the binned chain. For flat priors, lack of exact overlap indicates that the posterior distribution is non-Gaussian; in this case, the contours also show a bi-modal probability distribution. Markov chains are analyzed using the public software from \cite[Lewis and Bridle 2002]{lb}. {\it Right}: Probability distribution for the Stellar Mass assuming $Z = Z_\odot$ for the BC03 (dotted-dashed, magenta) and BC07 (dashed, blue) models, then including nebular emission (thin solid, red), and varying $Z$ with a logarithmic prior (thick solid, green). Shaded regions show the constraints from \cite[Lai et al (2008)]{lai}. }} \end{center} \end{figure} \begin{enumerate} \item We compute the spectra on a grid of locations exploring the entire parameter space, saving the final product of the sequence of steps described in Fig 1 ({\it after} convolution with the filter transmission curves, so we retain only a handful of numbers corresponding to the flux densities in the observations' bands); \item We read the grid into memory; \item We run MCMC as usual, but to compute the $\chi^2$ at each location we use {\it multi-linear interpolation} between the pre-computed spectra; \item We enjoy the speed up factor of 20,000, which allows us to fit the SED of each galaxy in a few seconds (even assuming to run several chains per object). \end{enumerate} A couple of caveats are in order. First, this method doesn't have the flexibility of GalMC; because it is difficult to perform interpolation in more than three dimensions, not more than four SED fitting parameters can be used (the fourth parameter being stellar mass, which is a normalization and therefore is excluded from the interpolation process). Second, there is an ``overhead" cost in computing the grid; for 50 values of age and E(B-V), and 100 values of redshift, running the initial grid takes about 24 hours, and this need to be repeated for a different survey (since the set of utilized filters change), or to use, \eg, a different star formation history or IMF. Still, for large surveys the set of filters is fixed, the number of different modeling options one might want to try is limited, and four parameters are enough to capture the general physical properties of a population of galaxies. Finally, let us observe that the resolution of the initial grid {\it does not} correspond to the resolution with which the PDF is sampled, as is the case in grid-based models. MCMC is still free to sample any desired location in parameter space, and the accuracy of the predicted spectrum corresponds to the accuracy of the linear interpolation between the points of the grid. This is illustrated in Fig. 3. The accuracy can be improved by increasing the number of points in the grid. A test conducted on the LAEs at $z = 3.1$ revealed that 50 points in age between 0 and the age of the Universe and 50 values of E(B-V) between 0 and 1 are enough to produce the estimate and credible intervals as the original GalMC, and using 100 values rather than 50 does not produce any appreciable difference. SpeedyMC is currently being used for the analysis of data from the Cosmic Assembly Near-Infrared Deep Extragalactic Legacy Survey (CANDELS), (\cite[Grogin et al 2011]{grogin}, \cite[Koekemoer et al 2011]{koekemoer}, \cite[Acquaviva et al 2012]{Acq2012}). The algorithm is not yet public, but you are welcome to contact the author for discussion on how to implement it, starting from GalMC. \begin{figure}[t] \begin{center} \includegraphics[width=8cm]{MCMCpath.jpg} \caption{An example of the path of SpeedyMC for a two-dimensional grid. The visited locations do not need to lie at the locations where template spectra have been saved (indicated by stars); instead, the corresponding spectrum is obtained by very fast bi-linear interpolation between the four corner stars. } \end{center} \end{figure}
\section{Introduction} Pairs of differential forms on the finite dimensional manifolds $M$ and $S$ induce differential forms on the Fr\'echet manifold $\F(S,M)$ of smooth functions. More precisely, if $S$ is a compact oriented $k$--dimensional manifold, the hat pairing is: \begin{gather*} \Omega^p(M)\times\Omega^q(S)\to\Om^{p+q-k}(\F(S,M))\\ {\widehat{\om\cdot\al}=\fint_S\ev^*\om\wedge\pr^*\al}, \end{gather*} where $\ev:S\x\F(S,M)\to M$ denotes the evaluation map, $\pr:S\x\F(S,M)\to S$ the projection and $\fint_S$ fiber integration. We show that the hat pairing is compatible with the canonical $\rm{Diff}(M)$ and ${\rm Diff}(S)$ actions on $\F(S,M)$, and with the exterior derivative. As a consequence we obtain a hat pairing in cohomology. The hat (transgression) map is the hat pairing with the constant function $1$, so it associates to any form $\om\in\Om^p(M)$ the form $\widehat{\om\cdot 1}=\widehat\om=\fint_S\ev^*\om\in\Om^{p-k}(\F(S,M))$. Since $\X(M)$ acts infinitesimally transitive on the open subset $\Emb(S,M)\subset\F(S,M)$ of embeddings of the $k$--dimensional oriented manifold $S$ into $M$ \cite{H76}, the expression of $\widehat\om$ at $f\in\Emb(S,M)$ is \[ \widehat{\om}(X_1\o f,\dots, X_{p-k}\o f)=\int_Sf^*(i_{X_{p-k}}\dots i_{X_1}\om),\quad X_1,\dots,X_{p-k}\in\X(M). \] When $S$ is the circle, then one obtains the usual transgression map with values in the space of $(p-1)$-forms on the free loop space of $M$. Let $\Gr_k(M)$ be the non-linear Grassmannian of $k$--dimensional oriented submanifolds of $M$. The tilda map associates to every $\om\in\Om^p(M)$ a differential $(p-k)$-form on $\Gr_k(M)$ given by \cite{HV04} \[ \tilde\om(\tilde Y_N^1,\dotsc,\tilde Y_N^{p-k}) =\int_Ni_{Y_N^{p-k}}\cdots i_{Y_N^1}\om,\quad\forall\tilde Y_N^1,\dots,\tilde Y_N^{p-k}\in \Ga(TN^\perp)=T_N\Gr_k(M), \] for $\tilde Y_N$ section of the orthogonal bundle $TN^\perp$ represented by the section $Y_N$ of $TM|_N$. The natural map \[ \pi:\Emb(S,M)\to\Gr_k(M),\quad\pi(f)=f(S) \] provides a principal bundle with the group $\Diff_+(S)$ of orientation preserving diffeomorphisms of $S$ as structure group. The hat map on $\Emb(S,M)$ and the tilda map on $\Gr_k(M)$ are related by $\widehat\om=\pi^*\tilde\om$. This is the reason why for the hat calculus one has similar properties to those for the tilda calculus. The tilda calculus was used to study the non-linear Grassmannian of co-dimension two submanifolds as symplectic manifold \cite{HV04}. We apply the hat calculus to the hamiltonian formalism for $p$-branes and open $p$-branes \cite{AS05} \cite{BZ05}. The bar map $\bar\om=\widehat{\om\cdot\mu}$ is the hat pairing with a fixed volume form $\mu$ on $S$, so \[ \bar\om(Y^1_f,\dots,Y^p_f)=\int_S \om(Y^1_f,\dots,Y^p_f)\mu, \quad\forall Y^1_f,\dots,Y^{p}_f\in \Ga(f^*TM)=T_f \mathcal{F}(S,M). \] We use the bar calculus to study $\F(S,M)$ with symplectic form $\bar\om$ induced by a symplectic form $\om$ on $M$. The natural actions of $\Diff_{ham}(M,\om)$ and $\Diff_{ex}(S,\mu)$, the group of hamiltonian diffeomorphisms of $M$ and the group of exact volume preserving diffeomorphisms of $S$, are two commuting hamiltonian actions on $\F(S,M)$. Their momentum maps form the dual pair for ideal incompressible fluid flow \cite{MW83} \cite{GBV09}. We are grateful to Stefan Haller for extremely helpful suggestions. \section{Hat pairing}\label{calc} We denote by $\mathcal{F}(S,M)$ the set of smooth functions from a compact oriented $k$--dimensional manifold $S$ to a manifold $M$. It is a Fr\'echet manifold in a natural way \cite{KM97}. Tangent vectors at $f\in\mathcal{F}(S,M)$ are identified with vector fields on $M$ along $f$, \ie sections of the pull-back vector bundle $f^*TM$. Let $\ev:S\x\F(S,M)\to M$ be the evaluation map $\ev(x,f)=f(x)$ and $\pr:S\x\F(S,M)\to S$ the projection $\pr(x,f)=x$. A pair of differential forms $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$ determines a differential form $\widehat{\om\cdot\al}$ on $\F(S,M)$ by the fiber integral over $S$ (whose definition and properties are listed in the appendix) of the $(p+q)$-form $\ev^*\om\wedge\pr^*\al$ on $S\x\F(S,M)$: \begin{equation}\label{use} {\widehat{\om\cdot\al}=\fint_S\ev^*\om\wedge\pr^*\al} \end{equation} In this way we obtain a bilinear map called the {\it hat pairing}: \begin{equation*}\label{pair} \Om^p(M)\x\Om^q(S)\to\Om^{p+q-k}(\F(S,M)). \end{equation*} An explicit expression of the hat pairing avoiding fiber integration is: \begin{align}\label{ffff} (\widehat{\om\cdot\al})_f(Y_f^1,\dots,Y_f^{p+q-k}) =\int_Sf^*(i_{Y_f^{p+q-k}}\dots i_{Y_f^1}(\om\o f))\wedge\al, \end{align} for $Y_f^1,\dots Y_f^{p+q-k}$ vector fields on $M$ along $f\in\F(S,M)$. Here we denote by $f^*\be_f$ the "restricted pull-back" by $f$ of a section $\be_f$ of $f^*(\La^m T^*M)$, which is a differential $m$--form on $S$ given by $f^*\be_f:x\in S\mapsto(\La^mT^*_xf)(\be_f(x))\in\La^mT_x^*S$, where $T_x^*f:T^*_{f(x)}M\to T^*_xS$ denotes the dual of $T_xf$. The fact that \eqref{use} and \eqref{ffff} provide the same differential form on $\F(S,M)$ can be deduced from the identity \[ (\ev^*\om)_{(x,f)}(Y_f^1,\dots, Y_f^{p-k},X_x^1,\dots, X_x^k) =f^*(i_{Y_f^{p-k}}\dots i_{Y_f^1}(\om\o f))(X_x^1,\dots, X_x^k) \] for $Y_f^1,\dots,Y_f^{p-k}\in T_f\F(S,M)$ and $X_x^1,\dots,X_x^k\in T_xS$. Since $\X(M)$ acts infinitesimally transitive on the open subset $\Emb(S,M)\subset\F(S,M)$ of embeddings of the $k$--dimensional oriented manifold $S$ into $M$, we express $\widehat\om$ at $f\in\Emb(S,M)$ as: \begin{equation}\label{embf} (\widehat{\om\cdot\al})_f(X_1\o f,\dots, X_{p+q-k}\o f)=\int_Sf^*(i_{X_{p+q-k}}\dots i_{X_1}\om)\wedge\al. \end{equation} One uses the fact that the "restricted pull-back" by $f$ of $i_{X_{p+q-k}\o f}\dots i_{X_1\o f}(\om\o f)$ is $f^*(i_{X_{p+q-k}}\dots i_{X_1}\om)$. Next we show that the hat pairing is compatible with the exterior derivative of differential forms. \begin{theo}\label{dddd} The exterior derivative $\dd$ is a derivation for the hat pairing, \ie \begin{equation}\label{deri} \dd(\widehat{\om\cdot\al})= \widehat{(\dd\om)\cdot\al}+(-1)^{p}\widehat{\om\cdot\dd\al}, \end{equation} where $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$. \end{theo} \begin{proof} Differentiation and fiber integration along the boundary free manifold $S$ commute, so \begin{align*} \dd(\widehat{\om\cdot\al})&= \dd\fint_S\ev^*\om\wedge\pr^*\al=\fint_S\dd(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S\ev^*\dd\om\wedge\pr^*\al+(-1)^p\fint_S\ev^*\om\wedge\pr^*\dd\al =\widehat{(\dd\om)\cdot\al}+(-1)^{p}\widehat{\om\cdot\dd\al} \end{align*} for all $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$. \end{proof} The differential form $\widehat{\om\cdot\al}$ is exact if $\om$ is closed and $\al$ exact (or if $\al$ is closed and $\om$ exact). In the special case $p+q=k$ these conditions imply that the function $\widehat{\om\cdot\al}$ on $\F(S,M)$ vanishes. \begin{coro} The hat pairing induces a bilinear map on de Rham cohomology spaces \begin{equation}\label{hh} H^p(M)\x H^q(S)\to H^{p+q-k}(\F(S,M)). \end{equation} In particular there is a bilinear map $$H^p(M)\x H^q(M)\to H^{p+q-k}(\Diff(M)).$$ \end{coro} \begin{rema} The cohomology group $H^q(S)$ is isomorphic to the homology group $H_{k-q}(S)$ by Poincar\' e duality. With the notation $n=k-q$, the hat pairing \eqref{hh} becomes \[ H^p(M)\x H_n(S)\to H^{p-n}(\F(S,M)), \] and it is induced by the map $(\om,\si)\mapsto \fint_\si\ev^*\om$, for differential $p$-forms $\om$ on $M$ and $n$-chains $\si$ on $S$. \end{rema} If $S$ is a manifold with boundary, then formula \eqref{deri} receives an extra term coming from integration over the boundary. Let $i_\pa:\pa S\to S$ be the inclusion and $r_\pa:\F(S,M)\to\F(\pa S,M)$ the restriction map. \begin{prop}\label{inde} The identity \begin{equation}\label{unu} \dd(\widehat{\om\cdot\al})= \widehat{(\dd\om)\cdot\al}+(-1)^{p}\widehat{\om\cdot\dd\al} +(-1)^{p+q-k}r_\pa^*(\widehat{\om\cdot i_\pa^*\al}^\pa) \end{equation} holds for $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$, where the upper index $\pa$ assigned to the hat means the pairing $$ \Om^p(M)\x\Om^q(\pa S)\to\Om^{p+q-k+1}(\F(\pa S,M)). $$ \end{prop} \begin{proof} For any differential $n$--form $\be$ on $S\x\F(S,M)$, the identity \[ \dd\fint_S\be-\fint_S\dd\be=(-1)^{n-k}\fint_{\pa S}(i_\pa\x 1_{\F(S,M)})^*\be \] holds because of the identity \eqref{r4} from the appendix. The obvious formulas \[ \pr\o(i_\pa\x 1_{\F(S,M)})=i_\pa\o\pr_\pa,\quad \ev\o(i_\pa\x 1_{\F(S,M)})=\ev_\pa, \] for $\ev_\pa:\pa S\x\F(S,M)\to M$ and $\pr_\pa:\pa S\x\F(S,M)\to\pa S$, are used to compute \begin{align*} \dd&(\widehat{\om\cdot\al})= \dd\fint_S\ev^*\om\wedge\pr^*\al\\ &=\fint_S\dd(\ev^*\om\wedge\pr^*\al) +(-1)^{p+q-k}\fint_{\pa S}(i_\pa\x 1_{\F(S,M)})^*(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S\ev^*\dd\om\wedge\pr^*\al+(-1)^p\fint_S\ev^*\om\wedge\pr^*\dd\al +(-1)^{p+q-k}\fint_{\pa S}\ev_\pa^*\om\wedge\pr_\pa^*i_\pa^*\al\\ &=\widehat{(\dd\om)\cdot\al}+(-1)^{p}\widehat{\om\cdot\dd\al} +(-1)^{p+q-k}{r_\pa}^*(\widehat{\om\cdot i_\pa^*\al}^\pa), \end{align*} thus obtaining the requested identity. \end{proof} \paragraph{Left $\Diff(M)$ action.} The natural left action of the group of diffeomorphisms $\Diff(M)$ on $\F(S,M)$ is $\ph\cdot f=\ph\o f$. The infinitesimal action of $X\in\X(M)$ is the vector field $\bar X$ on $\F(S,M)$: \[ \bar X(f)=X\o f,\quad\forall f\in\F(S,M). \] We denote by $\bar\ph$ the diffeomorphism of $\F(S,M)$ induced by the action of $\ph\in\Diff(M)$, so $\bar\ph(f)=\ph\o f$ is the push-forward by $\ph$. \begin{prop}\label{pppp} Given $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$, the identity \begin{align}\label{star} \bar\ph^*\widehat{\om\cdot\al}&=\widehat{(\ph^*\om)\cdot\al} \end{align} and its infinitesimal version \begin{align}\label{el} L_{\bar X}\widehat{\om\cdot\al}&=\widehat{(L_X\om)\cdot\al} \end{align} hold for all $\ph\in\Diff(M)$ and $X\in\X(M)$. \end{prop} \begin{proof} Using the expression \eqref{use} of the hat pairing and identity \eqref{r1} from the appendix, we have: \begin{align*} \bar\ph^*\widehat{\om\cdot\al}& =\bar\ph^*\fint_S\ev^*\om\wedge\pr^*\al =\fint_S(1_S\x\bar\ph)^*(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S\ev^*\ph^*\om\wedge\pr^*\al =\widehat{(\ph^*\om)\cdot\al}, \end{align*} since $\pr\o(1_S\x\bar\ph)=\pr$ and $\ev\o(1_S\x\bar\ph)=\ph\o\ev$. \end{proof} A similar result is obtained for any smooth map $\et\in\F(M_1,M_2)$ and its push-forward $\bar\et:\F(S,M_1)\to\F(S,M_2)$, $\bar\et(f)=\et\o f$: \begin{equation*} \bar\et^*\widehat{\om\cdot\al}=\widehat{\et^*\om\cdot\al}, \end{equation*} for all $\om\in\Om^p(M_2)$ and $\al\in\Om^q(S)$. \begin{lemm}\label{ins} For all vector fields $X\in\X(M)$, the identity $i_{\bar X}\widehat{\om\cdot\al}=\widehat{(i_X\om)\cdot\al}$ holds. \end{lemm} \begin{proof} The vector field $0_S\x \bar X$ on $S\x \F(S,M)$ is $\ev$-related to the vector field $X$ on $M$, so \begin{align*} i_{\bar X}\widehat{\om\cdot\al}& =i_{\bar X}\fint_S \ev^*\om\wedge\pr^*\al =\fint_Si_{0_S\x \bar X}(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S\ev^*(i_X\om)\wedge\pr^*\al =\widehat{(i_X\om)\cdot\al}. \end{align*} At step two we use formula \eqref{r3} from the appendix. \end{proof} \paragraph{Right $\Diff(S)$ action.} The natural right action of the diffeomorphism group $\Diff(S)$ on $\F(S,M)$ can be transformed into a left action by $\ps\cdot f=f\o\ps^{-1}$. The infinitesimal action of $Z\in\X(S)$ is the vector field $\hat Z$ on $\F(S,M)$: \[ \widehat Z(f)=-Tf\o Z,\quad\forall f\in\F(S,M). \] We denote by $\widehat\ps$ the diffeomorphism of $\F(S,M)$ induced by the action of $\ps$, so $\widehat\ps(f)=f\o\ps^{-1}$ is the pull-back by $\ps^{-1}$. \begin{prop}\label{five} Given $\om\in\Om^p(M)$ and $\al\in\Om^q(S)$, the identity \begin{align*} \widehat\ps^*\widehat{\om\cdot\al}&=\widehat{\om\cdot\ps^*\al} \end{align*} and its infinitesimal version \begin{align*} L_{\widehat Z}\widehat{\om\cdot\al}&=\widehat{\om\cdot L_Z\al} \end{align*} hold for all orientation preserving $\ps\in\Diff(S)$ and $Z\in\X(S)$. \end{prop} \begin{proof} The obvious identities $\ev\o(1_S\x \widehat\ps)=\ev\o(\ps^{-1}\x 1_\F)$, $\pr\o(1_S\x \widehat\ps)=\pr$ and $\pr\o(\ps\x 1_\F)=\ps\o\pr$ are used in the computation \begin{align*} \widehat\ps^*\widehat{\om\cdot\al}& =\widehat\ps^*\fint_S\ev^*\om\wedge\pr^*\al =\fint_S(1_S\x\widehat\ps)^*(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S\big((\ps^{-1}\x 1_\F)^*\ev^*\om\big)\wedge\pr^*\al =\fint_S\ev^*\om\wedge(\ps\x 1_\F)^*\pr^*\al\\ &=\fint_S\ev^*\om\wedge\pr^*\ps^*\al =\widehat{\om\cdot\ps^*\al}, \end{align*} together with formula \eqref{r2} from the appendix at step four. \end{proof} \begin{lemm}\label{ins2} The identity $i_{\widehat Z}\widehat{\om\cdot\al}=(-1)^{p}\widehat{\om\cdot i_Z\al}$ holds for all vector fields $Z\in\X(S)$, if $\om\in\Om^p(M)$. \end{lemm} \begin{proof} The infinitesimal version of the first identity in the proof of proposition \ref{five} is $T\ev.(0_S\x \widehat Z)=T\ev.(-Z\x 0_{\F(S,M)})$, so we compute: \begin{align*} i_{\widehat Z}\widehat{\om\cdot\al}& =i_{\widehat Z}\fint_S \ev^*\om\wedge\pr^*\al =\fint_S i_{0_S\x \widehat Z}(\ev^*\om\wedge\pr^*\al)\\ &=\fint_S (i_{0_S\x \widehat Z}\ev^*\om)\wedge\pr^*\al =\fint_S(i_{-Z\x 0_{\F(S,M)}}\ev^*\om)\wedge\pr^*\al\\ &=\fint_Si_{-Z\x 0_{\F(S,M)}}(\ev^*\om\wedge\pr^*\al) -\fint_S(-1)^p\ev^*\om\wedge i_{-Z\x 0_{\F(S,M)}}\pr^*\al\\ &=(-1)^{p}\fint_S\ev^*\om\wedge\pr^*(i_Z\al) =(-1)^{p}\widehat{\om\cdot i_Z\al}. \end{align*} At step two we use formula \eqref{r3} from the appendix. \end{proof} \section{Tilda map and hat map}\label{s3} Let $\Gr_k(M)$ be the non-linear Grassmannian (or differentiable Chow variety) of compact oriented $k$--dimensional submanifolds of $M$. It is a Fr\' echet manifold \cite{KM97} and the tangent space at $N\in\Gr_k(M)$ can be identified with the space of smooth sections of the normal bundle $TN^\perp=(TM|_N)/TN$. The tangent vector at $N$ determined by the section $Y_N\in\Ga(TM|_N)$ is denoted by $\tilde Y_N\in T_N\Gr_k(M)$. The {\it tilda map} \cite{HV04} associates to any $p$--form $\om$ on $M$ a $(p-k)$--form $\tilde\om$ on $\Gr_k(M)$ by: \begin{equation}\label{tide} \tilde\om_N(\tilde Y_N^1,\dotsc,\tilde Y_N^{p-k}) =\int_Ni_{Y_N^{p-k}}\cdots i_{Y_N^1}\om. \end{equation} Here all $\tilde Y_N^j$ are tangent vectors at $N\in\Gr_k(M)$, \ie sections of $TN^\perp$ represented by sections $Y_N^j$ of $TM|_N$. Then $i_{Y_N^{p-k}}\cdots i_{Y_N^1}\om\in\Omega^k(N)$ does not depend on representatives $Y_N^j$ of $\tilde Y_N^j$, and integration is well defined since $N\in\Gr_k(M)$ comes with an orientation. Let $S$ be a compact oriented $k$--dimensional manifold. The {\it hat map} is the hat pairing with the constant function $1\in\Om^0(S)$. It associates to any form $\om\in\Om^p(M)$ the form $\widehat\om\in\Om^{p-k}(\F(S,M))$: \begin{equation}\label{hatm} \widehat\om=\widehat{\om\cdot 1}=\fint_S\ev^*\om. \end{equation} On the open subset $\Emb(S,M)\subset\F(S,M)$ of embeddings, formula \eqref{ffff} gives \begin{equation}\label{exte} \widehat{\om}(X_1\o f,\dots, X_{p-k}\o f)=\int_Sf^*(i_{X_{p-k}}\dots i_{X_1}\om). \end{equation} \begin{rema} The hat map induces a transgression on cohomology spaces \[ H^p(M)\to H^{p-k}(\F(S,M)). \] When $S$ is the circle, then one obtains the usual transgression map with values in the $(p-1)$-th cohomology space of the free loop space of $M$. \end{rema} Let $\pi$ denote the natural map \[ \pi:\Emb(S,M)\to\Gr_k(M),\quad\pi(f)=f(S). \] where the orientation on $f(S)$ is chosen such that the diffeomorphism $f:S\to f(S)$ is orientation preserving. The image $\pi(\Emb(S,M))$ is the manifold $\Gr_k^S(M)$ of $k$--dimensional submanifolds of $M$ of type $S$. Then $\pi:\Emb(S,M)\to\Gr_k^S(M)$ is a principal bundle over $\Gr_k^S(M)$ with structure group $\Diff_+(S)$, the group of orientation preserving diffeomorphisms of $S$. Note that there is a natural action of the group $\Diff(M)$ on the non-linear Grassmannian $\Gr_k(M)$ given by $\ph\cdot N=\ph(N)$. Let $\tilde\ph$ be the diffeomorphism of $\Gr_k(M)$ induced by the action of $\ph\in\Diff(M)$. Then $\tilde\ph\o\pi=\pi\o\bar\ph$ for the restriction of $\bar\ph(f)=\ph\o f$ to a diffeomorphism of $\Emb(S,M)\subset\F(S,M)$. As a consequence, the infinitesimal generators for the $\Diff(M)$ actions on $\Gr_k(M)$ and on $\Emb(S,M)$ are $\pi$--related. This means that for all $X\in\X(M)$, the vector fields $\tilde X$ on $\Gr_k(M)$ given by $\tilde X(N)=X|_N$ and $\bar X$ on $\Emb(S,M)$ given by $\bar X(f)=X\o f$ are $\pi$--related. \begin{prop} The hat map on $\Emb(S,M)$ and the tilda map on $\Gr_k(M)$ are related by $\widehat\om=\pi^*\tilde\om$, for any $k$--dimensional oriented manifold $S$. \end{prop} \begin{proof} For the proof we use the fact that $\X(M)$ acts infinitesimally transitive on $\Emb(S,M)$, so $T_f\Emb(S,M)=\{X\o f:X\in\X(M)\}$. With \eqref{tide} and \eqref{exte} we compute: \begin{multline*} (\pi^*\tilde\om)_f(X_1\o f,\dots, X_{p-k}\o f) =\tilde\om_{f(S)}(X_1|_{f(S)},\dots,X_{p-k}|_{f(S)})\\ =\int_{f(S)} i_{X_{p-k}}\dots i_{X_1}\om=\int_Sf^*(i_{X_{p-k}}\dots i_{X_1}\om)=\widehat\om_f(X_1\o f,\dots,X_{p-k}\o f), \end{multline*} since $\bar X$ and $\tilde X$ are $\pi$--related. \end{proof} From the properties of the hat pairing presented in proposition \ref{pppp}, lemma \ref{ins} and theorem \ref{dddd}, a hat calculus follows easily: \begin{prop}\label{cor3} For any $\om\in\Om^p(M)$, $\ph\in\Diff(M)$, $X\in\X(M)$, and $\et\in\F(M', M)$ with push-forward $\bar\et:\F(S,M')\to\F(S,M)$, the following identities hold: \begin{enumerate} \item $\bar\ph^*\widehat{\om}=\widehat{\ph^*\om}$ and $\bar\et^*\widehat\om=\widehat{\et^*\om}$ \item $L_{\bar X}\widehat\om=\widehat{L_X\om}$ \item $i_{\bar X}\widehat\om=\widehat{i_X\om}$ \item $\dd\widehat\om=\widehat{\dd\om}$. \end{enumerate} \end{prop} \begin{rema} If $S$ is a manifold with boundary, then the formula 4. above receives an extra term coming from integration over the boundary $\pa S$ as in proposition \ref{inde}: \begin{equation}\label{bond} \dd\widehat\om=\widehat{\dd\om}+(-1)^{p-k}r_\pa^*\widehat\om^\pa \end{equation} for $\om\in\Om^p(M)$. As before, $r_\pa:\F(S,M)\to\F(\pa S,M)$ denotes the restriction map on functions and $\om\in\Om^p(M)\mapsto\widehat\om^\pa\in\Om^{p-k+1}(\F(\pa S,M))$. \end{rema} Now the properties of the tilda calculus follow imediately from proposition \ref{cor3}. \begin{prop}\cite{HV04}\label{cor4} For any $\om\in\Om^p(M)$, $\ph\in\Diff(M)$ and $X\in\X(M)$, the following identities hold: \begin{enumerate} \item $\tilde\ph^*\tilde{\om}=\widetilde{\ph^*\om}$ \item $L_{\tilde X}\tilde\om=\widetilde{L_X\om}$ \item $i_{\tilde X}\tilde\om=\widetilde{i_X\om}$ \item $\dd\tilde\om=\widetilde{\dd\om}$. \end{enumerate} \end{prop} \begin{proof} We verify the identities 1. and 4. From relation 1. from proposition \ref{cor3} we get that \begin{align*} \pi^*\tilde\ph^*\tilde\om& =\bar\ph^*\pi^*\tilde\om =\bar\ph^*\widehat\om =\widehat{\ph^*\om} =\pi^*\widetilde{\ph^*\om}, \end{align*} and this implies the first identity. Using identity 4. from proposition \ref{cor3} we compute $$ \pi^*\dd\tilde\om=\dd\pi^*\tilde\om =\dd \widehat\om =\widehat{\dd\om}=\pi^*\widetilde{\dd\om}, $$ which shows the last identity. \end{proof} \subsection*{Hamiltonian formalism for $p$-branes} In this section we show how the hat calculus appears in the hamiltonian formalism for $p$-branes and open $p$-branes \cite{AS05} \cite{BZ05}. Let $S$ be a compact oriented $p$-dimensional manifold. The phase space for the $p$-brane world volume $S\x\RR$ is the cotangent bundle $T^*\F(S,M)$, where the canonical symplectic form is twisted. The twisting consists in adding a magnetic term, namely the pull-back of a closed 2-form on the base manifold, to the canonical symplectic form on a cotangent bundle \cite{MR99}. These twisted symplectic forms appear also in cotangent bundle reduction. We consider a closed differential form $H\in\Om^{p+2}(M)$. Since $\dim S=p$, the hat map \eqref{hatm} provides a closed 2-form $\widehat H$ on $\F(S,M)$. If $\pi_\F:T^*\F(S,M)\to\F(S,M)$ denotes the canonical projection, the twisted symplectic form on $T^*\F(S,M)$ is \[ \Om_H=-\dd\Th_\F+\frac{1}{2}\pi_\F^*\widehat H, \] where $\Th_\F$ is the canonical 1-form on $T^*\F(S,M)$. For the description of open branes one considers a compact oriented $p$-dimensional manifold $S$ with boundary $\pa S$ and a submanifold $D$ of $M$. The phase space is in this case the cotangent bundle $T^*\F_D(S,M)$ over the manifold \cite{M80} \[ \F_D(S,M)=\{f:S\to M|f(\pa S)\subset D\}. \] The twisting of the canonical symplectic form is done with a closed differential form $H\in\Om^{p+2}(M)$ with $i^*H=\dd B$ for some $B\in\Om^{p+1}(D)$, where $i:D\to M$ denotes the inclusion. The twisted symplectic form on $T^*\F_D(S,M)$ is \[ \Om_{(H,B)}=-\dd\Th_{\F_D}+\frac12\pi_{\F_D}^*(\widehat H-\pa^*\widehat B^\pa) \] with $\pa:\F_D(S,M)\to\F(\pa S,D)$ the restriction map and $\pi_{\F_D}:T^*\F_D(S,M)\to\F_D(S,M)$. To distinguish between the hat calculus for $\F(S,M)$ and the hat calculus for $\F(\pa S,M)$, we denote $\widehat \ ^\pa:\Om^n(M)\to\Om^{n-p+1}(\F(\pa S,M))$. The only thing we have to verify is the closedness of $\widehat H-\pa^*\widehat B^\pa$. We first notice that \eqref{bond} implies $\dd\widehat H=\widehat{\dd H}+r_\pa^*\widehat H^\pa$, where $r_\pa:\F(S,M)\to\F(\pa S,M)$ denotes the restriction map, and identity 4 from proposition \ref{cor3} implies $\widehat{\dd B}^\pa=\dd\widehat B^\pa$. On the other hand identity 1 from proposition \ref{cor3} ensures that $\widehat{i^*H} ^\pa=\bar i^*\widehat H^\pa$, with $\bar i:\F(\pa S,D)\to\F(\pa S,M)$ denoting the push-forward by $i:D\to M$. Knowing that $r_\pa=\bar i\o\pa$, we compute: \begin{align*} \dd\widehat H =\widehat{\dd H}+r_\pa^*\widehat H^\pa =\pa^*\bar i^*\widehat H^\pa=\pa^*\widehat{i^*H}^\pa =\pa^*\widehat{\dd B}^\pa=\dd\pa^*\widehat B^\pa, \end{align*} so the closed 2--form $\widehat H-\pa^*\widehat B^\pa$ provides a twist for the canonical symplectic form on the cotangent bundle $T^*\F_D(S,M)$. \subsection*{Non-linear Grassmannians as symplectic manifolds} In this subsection we recall properties of the co-dimension two non-linear Grassmannian as a symplectic manifold. \begin{prop}\cite{I96} Let $M$ be a closed $m$--dimensional manifold with volume form $\nu$. The tilda map provides a symplectic form $\tilde\nu$ on $\Gr_{m-2}(M)$ $$ \tilde\nu_N(\tilde X_N, \tilde Y _N)=\int_Ni_{ Y _N}i_{X_N}\nu, $$ for $\tilde X_N$ and $\tilde Y _N$ sections of $TN^\perp$ determined by sections $X_N$ and $Y_N$ of $TM|_N$. \end{prop} \begin{proof} The 2--form $\tilde\nu$ is closed since $\dd\tilde\nu=\widetilde{\dd\nu}$ by the tilda calculus. To verify that it is also (weakly) non-degenerate, let $X_N$ be an arbitrary vector field along $N$ such that $\int_Ni_{ Y _N}i_{X_N}\nu=0$ for all vector fields $Y_N$ along $N$. Then $X_N$ must be tangent to $N$, so $\tilde X_N=0$. \end{proof} In dimension $m=3$ the symplectic form $\tilde\nu$ is known as the Marsden--Weinstein symplectic from on the space of unparameterized oriented links, see \cite{MW83} \cite{B93}. \paragraph{Hamiltonian $\Diff_{ex}(M,\nu)$ action.} The action of the group $\Diff(M,\nu)$ of volume preserving diffeomorphisms of $M$ on $\Gr_{m-2}(M)$ preserves the symplectic form $\tilde\nu$: \[ \tilde\ph^*\tilde\nu=\widetilde{\ph^*\nu}=\tilde\nu, \quad\forall\ph\in\Diff(M,\nu). \] The subgroup $\Diff_{ex}(M,\nu)$ of exact volume preserving diffeomorphisms acts in a hamiltonian way on the symplectic manifold $(\Gr_{m-2}(M),\tilde\nu)$. Its Lie algebra is $\X_{ex}(M,\nu)$, the Lie algebra of exact divergence free vector fields, \ie vector fields $X_\al$ such that $i_{X_\al}\nu=\dd\al$ for a potential form $\al\in\Om^{m-2}(M)$. The infinitesimal action of $X_\al$ is the vector field $\tilde X_\al$. By the tilda calculus $\tilde\al\in\F(\Gr_{m-2}(M))$ is a hamiltonian function for the hamiltonian vector field $\tilde X_\al$: \[ i_{\tilde X_\al}\tilde\nu=\widetilde{i_{X_\al}\nu}=\widetilde{\dd\al}=\dd\tilde\al. \] It depends on the particular choice of the potential $\al$ of $X_\al$. A fixed continuous right inverse $b:\dd\Om^{m-2}(M)\to\Om^{m-2}(M)$ to the differential $\dd$ picks up a potential $b(\dd\al)$ of $X_\al$. The corresponding momentum map is: \begin{align*} \mathbf{J}:\M\to \X_{ex}(M,\nu)^*, \quad\langle\mathbf{J}(N),X_\al\rangle =\widetilde{b(\dd\al)}(N)=\int_N b(\dd\al). \end{align*} On the connected component $\M$ of $N\in\Gr_{m-2}(M)$, the non-equivariance of $\mathbf{J}$ is measured by the Lie algebra 2--cocycle on $\X_{ex}(M,\nu)$ \begin{align*} \si_N(X,Y)=&\langle\mathbf{J}(N),[X,Y]^{op}\rangle -\tilde\nu(\tilde X,\tilde Y)(N) =\widetilde{(b\dd i_Yi_X\nu)}(N)-\widetilde{(i_Yi_X\nu)}(N)\\ &=\widetilde{(Pi_Xi_Y\nu)}(N) =\int_NPi_Xi_Y\nu. \end{align*} Here $P=1_{\Om^{m-2}(M)}-b\o \dd$ is a continuous linear projection on the subspace of closed $(m-2)$-forms and $(X,Y)\mapsto[Pi_Yi_X\nu]\in H^{m-2}(M)$ is the universal Lie algebra 2--cocycle on $\X_{ex}(M,\nu)$ \cite{R95}. The cocycle $\si_{N}$ is cohomologous to the Lichnerowicz cocycle \begin{equation}\label{lich} \si_\et(X,Y)=\int_M\et(X,Y)\nu, \end{equation} where $\et$ is a closed 2-form Poincar\'e dual to $N$ \cite{V09}. If $\nu$ is an integral volume form, then $\si_N$ is integrable \cite{I96}. The connected component $\M$ of $\Gr_{m-2}(M)$ is a coadjoint orbit of a 1--dimen\-sional central Lie group extension of $\Diff_{ex}(M,\nu)$ integrating $\si_{N}$, and $\tilde\nu$ is the Kostant-Kirillov-Souriau symplectic form. \cite{HV04}. \section{Bar map}\label{s4} When a volume form $\mu$ on the compact $k$--dimensional manifold $S$ is given, one can associate to each differential $p$-form on $M$ a differential $p$-form on $\F(S,M)$ \[ \bar\om(Y^1_f,\dots,Y^p_f)=\int_S \om(Y^1_f,\dots,Y^p_f)\mu, \quad\forall Y^i_f\in T_f \mathcal{F}(S,M), \] where $\om(Y^1_f,\dots,Y^p_f):x\mapsto \om_{f(x)}(Y_f^1(x),\dots, Y_f^p(x))$ defines a smooth function on $S$. In this way a {\it bar map} is defined. Formula \eqref{ffff} assures that this bar map is just the hat pairing of differential forms on $M$ with the volume form $\mu$ \begin{equation}\label{barb} \bar\om=\widehat{\om\cdot\mu}=\fint_S\ev^*\om\wedge\pr^*\mu. \end{equation} From the properties of the hat pairing presented in proposition \ref{pppp}, lemma \ref{ins} and theorem \ref{dddd}, one can develop a bar calculus. \begin{prop}\label{cor2} For any $\om\in\Om^p(M)$, $\ph\in\Diff(M)$ and $X\in\X(M)$, the following identities hold: \begin{enumerate} \item $\bar\ph^*\bar{\om}=\overline{\ph^*\om}$ \item $L_{\bar X}\bar\om=\overline{L_X\om}$ \item $i_{\bar X}\bar\om=\overline{i_X\om}$ \item $\dd\bar\om=\overline{\dd\om}$. \end{enumerate} \end{prop} \subsection*{$\F(S,M)$ as symplectic manifold} Let $(M,\om)$ be a connected symplectic manifold and $S$ a compact $k$--dimensional manifold with a fixed volume form $\mu$, normalized such that $\int_S\mu=1$. The following fact is well known: \begin{prop} The bar map provides a symplectic form $\bar\om$ on $\F(S,M)$: $$ \bar\om_f(X_f,Y_f)=\int_S\om(X_f,Y_f)\mu. $$ \end{prop} \begin{proof} That $\bar\om$ is closed follows from the bar calculus: $\dd\bar\om=\overline{\dd\om}=0$. The (weakly) non-degeneracy of $\bar\om$ can be verified as follows. If the vector field $X_f$ on $M$ along $S$ is non-zero, then $X_f(x)\ne 0$ for some $x\in S$. Because $\om$ is non-degenerate, one can find another vector field $Y_f$ along $f$ such that $\om(X_f,Y_f)$ is a bump function on $S$. Then $\bar\om(X_f,Y_f)=\int_S\om(X_f,Y_f)\mu\ne 0$, so $X_f$ does not belong to the kernel of $\bar\om$, thus showing that the kernel of $\bar\om$ is trivial. \end{proof} \paragraph{Hamiltonian action on $M$.} Let $G$ be a Lie group acting in a hamiltonian way on $M$ with momentum map $J:M\to\g^*$. Then $\F(S,M)$ inherits a $G$-action: $(g\cdot f)(x)=g\cdot(f(x))$ for any $x\in S$. The infinitesimal generator is $\xi_\F=\bar\xi_M$ for any $\xi\in\g$, where $\xi_M$ denotes the infinitesimal generator for the $G$-action on $M$. The bar calculus shows quickly that $G$ acts in a hamiltonian way on $\F(S,M)$ with momentum map \[ \mathbf{J}=\bar J:\F(S,M)\to \g^*,\quad \bar J(f)=\int_S(J\o f)\mu,\quad\forall f\in\F(S,M). \] Indeed, for all $\xi\in\g$ \[ i_{\xi_\F}\bar\om=i_{\bar\xi_M}\bar\om=\overline{i_{\xi_M}\om} =\overline{\dd\langle J,\xi\rangle}=\dd\langle\bar J,\xi\rangle. \] Let $M$ be connected and let $\si$ be the $\RR$-valued Lie algebra 2--cocycle on $\g$ measuring the non-equivariance of $J$, \ie \[ \si(\xi,\et)={\langle J(x),[\xi,\et]\rangle} -{\om(\xi_M,\et_M)}(x),\quad x\in M, \] (both terms are hamiltonian function for the vector field $[\xi,\et]_M=-[\xi_M,\et_M]$). Then the non-equivariance of $\mathbf{J}=\bar J$ is also measured by $\si$: for all $f\in\F(S,M)$ \[ \langle\bar J(f),[\xi,\et]\rangle-\bar\om(\xi_\F,\et_\F)(f) =\overline{\langle J,[\xi,\et]\rangle}(f) -\overline{\om(\xi_M,\et_M)}(f) =\si(\xi,\et). \] \paragraph{Hamiltonian $\Diff_{ham}(M,\om)$ action.} The action of the group $\Diff(M,\om)$ of symplectic diffeomorphisms preserves the symplectic form $\bar\om$: \begin{equation*} \bar\ph^*\bar\om=\overline{\ph^*\om}=\bar\om,\quad \forall\ph\in\Diff(M,\om). \end{equation*} The subgroup $\Diff_{ham}(M,\om)$ of hamiltonian diffeomorphisms of $M$ acts in a hamiltonian way on the symplectic manifold $\F(S,M)$. The infinitesimal action of $X_h\in\X_{ham}(M,\om)$, $h\in\F(M)$, is the hamiltonian vector field $\bar X_h$ on $\F(S,M)$ with hamiltonian function $\bar h$. This follows by the bar calculus: \begin{gather*} \dd\bar h=\overline{\dd h}=\overline{i_{X_h}\om}=i_{\bar X_h}\bar\om. \end{gather*} The hamiltonian function $\bar h$ of $\bar X_h$ depends on the particular choice of the hamiltonian function $h$. To solve this problem we fix a point $x_0\in M$ and we choose the unique hamiltonian function $h$ with $h(x_0)=0$, since $M$ is connected. The corresponding momentum map is \begin{gather*} \mathbf{J}:\mathcal{F}(S,M)\to \X_{ham}(M,\om)^*,\quad\langle\mathbf{J}(f),X_h\rangle=\bar h(f)=\int_S(h\o f)\mu. \end{gather*} The Lie algebra 2--cocycle on $\X_{ham}(M,\om)$ measuring the non-equivariance of the momentum map is \[ \si(X,Y)=-\om(X,Y)(x_0), \] by the bar calculus \begin{align*} \si(X,Y)(f)&=\langle\mathbf{J}(f),[X,Y]^{op}\rangle -\bar\om(X_\F,Y_\F)(f)\\ &=\overline{\om(X,Y)-\om(X,Y)(x_0)}(f)-\bar\om(\bar X,\bar Y)(f) =-\om(X,Y)(x_0). \end{align*} This is a Lie algebra cocycle describing the central extension $$0\to\RR\to\F(M)\to\X_{ham}(M,\om)\to 0$$ where $\F(M)$ is enowed with the canonical Poisson bracket. A group cocycle on $\Diff_{ham}(M,\om)$ integrating the Lie algebra cocycle $\si$ if $\om$ exact is studied in \cite{ILM06}. \paragraph{Hamiltonian $\Diff_{ex}(S,\mu)$ action.} The (left) action of the group $\Diff(S,\mu)$ of volume preserving diffeomorphisms preserves the symplectic form $\bar\om$: \begin{equation*} \widehat\ps^*\bar\om=\widehat\ps^*\widehat{\om\cdot\mu}=\widehat{\om\cdot\ps^*\mu}=\widehat{\om\cdot\mu}=\bar\om,\quad\forall\ps\in\Diff(S,\mu). \end{equation*} The subgroup $\Diff_{ex}(S,\mu)$ of exact volume preserving diffeomorphisms acts in a hamiltonian way on the symplectic manifold $\F(S,M)$. The infinitesimal action of the exact divergence free vector field $X_\al\in\X_{ex}(S,\mu)$ with potential form $\al\in\Om^{k-2}(S)$ is the hamiltonian vector field $\widehat X_\al$ on $\F(S,M)$ with hamiltonian function $\widehat{\om\cdot\al} $. Indeed, from $i_{X_\al}\mu=\dd\al$ follows by the hat calculus that \begin{gather*} \dd(\widehat{\om\cdot\al})=\widehat{\dd\om\cdot\al}+\widehat{\om\cdot\dd\al}=\widehat{\om\cdot i_{X_\al}\mu}=i_{\widehat X_\al}\widehat{\om\cdot\mu}=i_{\widehat X_\al}\bar\om. \end{gather*} If the symplectic form $\om$ is exact, then the corresponding momentum map is \begin{gather*} \mathbf{J}:\mathcal{F}(S,M)\to \X_{ex}(S,\mu)^*,\quad \langle\mathbf{J}(f),X_\al\rangle =\widehat{(\om\cdot \al)}(f) =\int_Sf^*\om\wedge \al. \end{gather*} It takes values in the regular part of $\X_{ex}(S,\mu)^*$, which can be identified with $\dd\Om^1(S)$, so we can write $\mathbf{J}(f)=f^*\om$ under this identification. In general the hamiltonian function $\widehat{\om\cdot\al}$ of $\widehat X_\al$ depends on the particular choice of the potential form $\al$ of $X_\al$. To fix this problem we consider as in Section \ref{s3} a continuous right inverse $b:\dd\Om^{m-2}(M)\to\Om^{m-2}(M)$ to the differential $\dd$, so $b(\dd\al)$ is a potential for $X_\al$. The corresponding momentum map is \begin{gather*} \mathbf{J}:\mathcal{F}(S,M)\to \X_{ex}(S,\mu)^*,\quad \langle\mathbf{J}(f),X_\al\rangle =\widehat{(\om\cdot b\dd\al)}(f) =\int_Sf^*\om\wedge b(\dd\al). \end{gather*} On a connected component $\F$ of $\F(S,M)$, the non-equivariance of $\mathbf{J}$ is measured by the Lie algebra 2--cocycle \begin{align*} \si_\F(X,Y)&=\langle\mathbf{J}(f),[X,Y]\rangle -\bar\om(\hat X,\hat Y)(f)=({\om\cdot b\dd i_Yi_X\mu})\hat\ (f) -({\om\cdot i_Yi_X\mu})\hat\ (f)\\ &=({\om\cdot Pi_Xi_Y\mu})\hat\ (f) =\int_Sf^*\om\wedge Pi_Xi_Y\mu \end{align*} on the Lie algebra of exact divergence free vector fields, for $P=1-b\dd$ the projection on the subspace of closed $(m-2)$-forms. It does not depend on $f\in\F$, because the cohomology class $[f^*\om]\in H^2(S)$ does not depend on the choice of $f$. The cocycle $\si_\F$ is cohomologous to the Lichnerowicz cocycle $\si_{f^*\om}$ defined in \eqref{lich} \cite{V09}. Since $\int_S\mu=1$, the cocycle $\si_\F$ is integrable if and only if the cohomology class of $f^*\om$ is integral \cite{I96}. \begin{rema} The two equivariant momentum maps on the symplectic manifold $\F(S,M)$, for suitable central extensions of the hamiltonian group $\Diff_{ham}(M,\om)$ and of the group $\Diff_{ex}(S,\mu)$ of exact volume preserving diffeomorphisms, form the dual pair for ideal incompressible fluid flow \cite{MW83} \cite{GBV09}. \end{rema} \section{Appendix: Fiber integration} Chapter VII in \cite{GHV72} is devoted to the concept of integration over the fiber in locally trivial bundles. We particularize this fiber integration to the case of trivial bundles $S\x M\to M$, listing its main properties without proofs. Let $S$ be a compact $k$--dimensional manifold. Fiber integration over $S$ assigns to $\om\in\Om^n(S\x M)$ the differential form $\fint_S\om\in\Om^{n-k}(M)$ defined by \[ (\fint_S\om)(x)=\int_S\om_x\in\La^{n-k}T^*_xM,\quad\forall x\in M, \] where $\om_x\in\Om^k(S,\La^{n-k}T_x^*M)$ is the retrenchment of $\om$ to the fiber over $x$: \[ \langle\om_x(Z_s^1,\dots,Z_s^{n-k}),X_x^1\wedge\dots\wedge X_x^k\rangle=\om_{(s,x)}(X_x^1,\dots,X_x^k,Z_s^1,\dots, Z_s^{n-k}) \] for all $X_x^i\in T_xM$ and $Z_s^j\in T_sS$. The properties of the fiber integration used in the text are special cases of the propositions (VIII) and (X) in \cite{GHV72}: \begin{enumerate} \item Pull-back of fiber integrals: \begin{equation}\label{r1} f^*\fint_S\om=\fint_S(1_S\x f)^*\om,\quad\forall f\in\F(M',M), \end{equation} with infinitesimal version \begin{equation}\label{rr} L_X\fint_S\om=\fint_SL_{0_S\x X}\om,\quad\forall X\in\X(M). \end{equation} \item Invariance under pull-back by orientation preserving diffeomorphisms of $S$: \begin{equation}\label{r2} \fint_S(\ph\x 1_M)^*\om=\fint_S\om,\quad\forall\ph\in\Diff_+(S), \end{equation} with infinitesimal version $\fint_SL_{Z\x 0_M}\om=0,\quad\forall Z\in\X(S)$. \item Insertion of vector fields into fiber integrals: \begin{equation}\label{r3} i_X\fint_S\om=\fint_Si_{0_S\x X}\om,\quad\forall X\in\X(M). \end{equation} \item Integration along boundary free manifolds commutes with differentiation. When $\pa S$ denotes the boundary of the $k$--dimensional compact manifold $S$ and $i_\pa:\pa S\to S$ the inclusion, \begin{equation}\label{r4} \dd\fint_S\be-\fint_S\dd\be=(-1)^{n-k}\fint_{\pa S}(i_\pa\x 1_M)^*\be \end{equation} holds for any differential $n$--form $\be$ on $S\x M$. \end{enumerate}
\section{Introduction} The accumulating experimental evidence suggests that the pseudogap phase could be a key issue in understanding the underlying mechanism of high-transition-temperature superconducting (high-T$_c$) copper oxides \cite{taill}. Magnetotransport data in electron-doped copper oxide La$_{2-x}$Ce$_x$CuO$_4$ suggests that linear temperature-dependent resistivity correlates with the electron pairing and spin-fluctuating scattering of the electrons \cite{greene}. Simultaneously, in the hole-doped copper oxide YBa$_2$Cu$_3$O$_y$ a large in-plane anisotropy of the Nernst effect sets in at the boundary of the pseudogap phase and indicates that this phase breaks four-fold rotational symmetry in the a-b plane, pointing to stripe or nematic order \cite{taill1}. Hence, we study here analytically a coexistence of the stripe order and d-wave superconductivity. We found an exact solution of the Bogoliubov-de Gennes equations in the simple Hubbard t-U-V mode indicating that the Abrikosov's vortex core naturally gives rise to a stripe-ordered domain. We show that the size of the stripe domain may exceed superconducting vortex's core size $\xi_s$ and the inter-vortex distance in the Abrikosov's lattice in the limit of weak magnetic fields $H \leq H_{c1}$. As far as we know, this is the first analytic solution of such type. Previously, coexistence of the stripe-order and Abrikosov's vortices in the limit of high magnetic fields $H\sim H_{c2}$ has been investigated numerically \cite{knapp}. Calculations were limited by the size of the model cluster of 26$\times$52 sites. Hence, the numerical results covered only the case when the inter-vortex distance was less than correlation length of the static AFM order. Here we consider analytically the opposite case of weak magnetic fields, where the inter-vortex distance is mach greater than stripe-order (including AFM) correlation length. \% \% Predicted numerically stripe order \cite{Zaanen}, e.g. coupled spin- and charge-density periodic superstructure (SDW-CDW), was found in the underdoped superconducting cuprates experimentally, specifically in $La_{2-x}Ba_xCuO_4$ \cite{fujita} and $La_{1.6-x}Nd_{0.4}Sr_xCuO_4$ \cite{ich,hu}. It was shown analytically, that stripe-order may arise already in the short-range repulsive Hubbard model due to a quantum interference between backward and Umklapp scattering of electrons by SDW potential close to half-filling in the presence of the CDW order with "matching" wave-vector \cite{sm}. In the quasi-1D case analytical kink-like spin- and charge-density coupled solutions were found \cite{mm} in the normal state. A study of $La_{1.875}Ba_{0.125}CuO_4$ with angle-resolved photoemission and scanning tunneling spectroscopies \cite{valla} has found evidence for a d-wave-type gap at low temperature, well within the stripe-ordered phase but above the bulk superconducting Tc. An earlier inelastic neutron scattering data \cite{aeppli1} had shown field-induced fluctuating magnetic order with space periodicity $8a_0$ and wave vector pointing along Cu-O bond direction in the ab-plane of the optimally doped La$_{1.84}$Sr$_{0.16}$CuO$_4$ in external magnetic field of $7.5$ T below $10$ K. The applied magnetic field ($\sim 2-7$ T) imposes the vortex lattice and induces "checkerboard" local density of electronic states (LDOS) seen in the STM experiments in high-T$_c$ superconductor Bi$_2$Sr$_2$Ca Cu$_2$O$_{8+\delta}$ \cite{davis}. The pattern originating in the Abrikosov's vortex cores has $4a_0$ periodicity, is oriented along Cu-O bonds, and has decay length $\sim 30$ angstroms reaching well outside the vortex core. The existence of antiferromagnetic spin fluctuations well outside the vortex cores is also discovered by NMR \cite{nmr} in superconducting YBCO in a $13$ T external magnetic field. Theoretical predictions had also been made of the magnetic field induced coexistence of antiferromagnetic ordering phenomena and superconductivity in high-T$_c$ cuprates \cite{zhang,arovas,sachdev1,sachdev2,sachdev3} due to assumed proximity of pure superconducting state to a phase with co-existing superconductivity and spin density wave order. In these works effective Ginzburg-Landau theories of coupled superconducting-, spin- and charge-order fields were used. Alternatively, the fermionic quasi-particle weak-coupling approaches were focused on the theoretical predictions arising from the model of BCS superconductor with $d_{x^2-y^2}$ symmetry \cite{volovik}. An effect of the nodal fermions on the zero bias conductance peak in tunneling studies was predicted. However STM experiments of the vortices in high-T$_c$ compounds revealed a very different structure of LDOS \cite{davis}. In this paper we make an effort to combine both theoretical approaches and present analytical mean-field solutions of coexisting spin-, charge- and superconducting orders derived form microscopic Hubbard model in the weak-coupling approximation. The previous analytical results obtained in the quasi 1D cases \cite{mm,mm1,ma,fe} are now extended for two real space dimensions. Different analytical solutions for collinear and checkerboard stripe-phases, as well as for spin-charge density modulation inside Abrikosov's vortex core are obtained. Simultaneously, our theory provides wave-functions of the fermionic states in all considered cases. \section{ Effective hamitonian. Bogoliubov-de Gennes equations} Consider the Hamiltonian $H = H_0 + H_{sc}$ consisting of two parts: the first part is the Hubbard Hamiltonian with on-site repulsion $U > 0$ \begin{equation} H_0 =\displaystyle -t\sum_{\langle i,j\rangle, \sigma}c^{\dagger}_{i,\sigma}c_{j,\sigma}+ U\displaystyle\sum_{i} \hat{n}_{i,\uparrow} \hat{n}_{i, \downarrow} - \mu \sum_{i. \sigma} \hat{n}_{i, \sigma} , \label{hubbard} \end{equation} and the interaction part including superconducting correlations \begin{equation} H_{sc} = \sum_{<i,j>, \sigma } \Delta(i,j;\sigma)c^{\dagger}_{i,\sigma} c^{\dagger}_{j,-\sigma} + h.c., \label{hs} \end{equation} where $\sum_{<i,j>, \sigma }$ is a summation over nearest neighboring sites ${\bf r}_i$, ${\bf r}_j$ of the square lattice, and spin components $\sigma = 2 s_z = \pm 1$. In the self-consistent approximation the Hamiltonian acquires the form \begin{eqnarray} H = -t \sum_{\langle i,j\rangle \sigma}c^{\dagger}_{i,\sigma}c_{j,\sigma} +\frac{U}{2} \sum_{i,\sigma}(\rho_i c^{\dagger}_{i,\sigma}c_{i,\sigma} -\frac{\rho_i^2}{2}) \nonumber \\ -U \sum_{i, \sigma} \langle \hat{S}_z ({\bf r}_i ) \rangle \sigma c^{\dagger}_{i,\sigma}c_{i,\sigma} +U \langle \hat{S}_z ({\bf r}_i ) \rangle^2 -\mu \sum_{i, \sigma}c^{\dagger}_{i,\sigma}c_{i,\sigma} \nonumber \\ \sum_{<i,j>,\sigma } \Delta(i,j;\sigma)c^{\dagger}_{i,\sigma} c^{\dagger}_{j,-\sigma} +h.c. + \frac{|\Delta |^2}{g}, \label{H} \end{eqnarray} where we introduce similar \cite{mm} slowly varying functions for spin order parameter $m({\bf r}_i )$ and the charge density $\rho ({\bf r}_i)$ defined as \begin{eqnarray} \rho ({\bf r}) = \langle \hat{n} ({\bf r} ) \rangle, \quad (-1)^{x_i + y_i} m ({\bf r}_i ) = U \langle \hat{S}_z ({\bf r}_i ) \rangle, \nonumber \\ \Delta(i,j;\sigma)= -g \langle c_{j,-\sigma} c_{i,\sigma}\rangle. \label{selfc} \end{eqnarray} We can diagonalize the total Hamiltonian $H = H_0 + H_{sc}$ by performing a unitary Bogoliubov transformation \begin{equation} \hat{c}_{\sigma}({\bf r}) = \sum_n \gamma_{n, \sigma} u_{n, \sigma}({\bf r})-\sigma \gamma^+_{n, -\sigma} v^*_{n, -\sigma}({\bf r}) \label{tr} \end{equation} New operators $\gamma$, $\gamma^+$ satisfy the fermionic commutative relations $ \{\gamma_{n, \sigma}, \gamma^+_{m, \sigma^{\prime}}\} = \delta_{m,n} \delta_{\sigma, \sigma^{\prime}}$. The transformations (\ref{tr}) must diagonalize the Hamiltonian $H$: \begin{equation} H = E_g + \sum_{\epsilon_n >0} \epsilon_n \gamma_{n,\sigma}^+ \gamma_{n,\sigma}, \label{eg} \end{equation} where $E_g$ is the ground state energy and $\epsilon_n >0$ is the energy of the n-th excitation. Following \cite{bdg} we obtain the eigenvalue equations \begin{eqnarray} -t\sum_{{\bf \delta}} u_{\sigma }({\bf r} +{\bf \delta} )+ (\frac{U}{2} \rho({\bf r})-\mu )u_{\sigma }& +m ({\bf r} )(-1)^{x_i +y_i} \sigma u_{\sigma }({\bf r} ) \nonumber \\ +\sum_{{\bf \delta}} \Delta({\bf r},{\bf r} +{\bf \delta};\sigma )\sigma v_{\sigma } ({\bf r} +{\bf \delta} )& =\epsilon_{\sigma } u_{\sigma }({\bf r}), \label{deq1} \end{eqnarray} \begin{eqnarray} -\sum_{{\bf \delta}} \Delta^*({\bf r},{\bf r} +{\bf \delta};-\sigma )\sigma u_{\sigma } ({\bf r} +{\bf \delta} ) +t\sum_{{\bf \delta}} v_{\sigma }({\bf r} +{\bf \delta} ) \nonumber \\ - (\frac{U}{2} \rho({\bf r})-\mu )v_{\sigma } +m ({\bf r} )(-1)^{x_i +y_i} \sigma v_{\sigma }({\bf r} ) = \epsilon_{\sigma } v_{\sigma }({\bf r}), \label{deq2} \end{eqnarray} where ${\bf \delta}=\pm \hat{\bf x}, \pm \hat{\bf y}$. We suppose the $d_{x^2 -y^2}$ symmetry of the superconducting order parameter $\Delta({\bf r} ,{\bf r} \pm \hat{\bf x};\sigma )= \sigma \Delta_d ({\bf r} )$, $\Delta({\bf r} ,{\bf r} \pm \hat{\bf y};\sigma )= -\sigma \Delta_d ({\bf r} )$. The Fourier transform gives the usual dependence $\Delta_{{\bf p}} ({\bf r}) =$ $ \sigma \sum_{{\bf \delta}} \Delta_{sc}({\bf r},{\bf r}+{\bf \delta};\sigma ) \exp[-i {\bf p} {\bf \delta} ]= 2(\cos p_x - \cos p_y ) \Delta_d ({\bf r}). $ The system (\ref{deq1}) - (\ref{deq2}) can be rewritten in the continuum approximation. Consider states near the Fermi surface (FS) (see Fig.1) and use linear approximation for the quasiparticles spectrum. Since for SDW pairing components with wave vectors ${\bf p}$ and ${\bf p} - {\bf Q_+}$ (or ${\bf p}$ and ${\bf p} - {\bf Q_-}$, where ${\bf Q_+} -{\bf Q_-} = 2\pi (0,1)$ is the lattice vector for the pure system without doping, when $(-1)^{x_i + y_i} \equiv e^{\pm i {\bf Q} {\bf r}}$. ) are important (see Fig. 1), we represent the functions $u({\bf r} )$ and $v({\bf r} )$, similar to the one-dimensional case, as \begin{equation} u_{\sigma }({\bf r} ) =\sum_{{\bf p} \in FS, p_x >0}[u_{{\bf p}, \sigma } e^{i {\bf p} {\bf r}} + \sigma u_{{\bf p} - {\bf Q}, \sigma } ({\bf r} ) e^{i({\bf p} -{\bf Q} ){\bf r} }], \end{equation} where ${\bf Q} = {\bf Q}_+$ for wave vectors ${\bf p}_y > 0$ and ${\bf Q} = {\bf Q}_-$ for wave vectors $p_y < 0$, respectively. \begin{figure}[tbph] \begin{center} \includegraphics[width=3.0in]{fs.eps} \caption{ The Fermi surface } \end{center} \end{figure} For the doped case nesting vectors $ {\bf Q}_{\pm}$ are no longer equivalent. Therefore in the general case we consider vectors ${\bf Q}_{\pm}$ as independent and make the substitution $(-1)^{x_i + y_i} m ({\bf r}_i ) \to$ \[ m_+ ({\bf r}_i )\exp (i {\bf Q}_+ {\bf r}_+ ) + m_- ({\bf r}_i )\exp (i {\bf Q}_- {\bf r}_- ) + h.c. \] Eigenvalue equations (\ref{deq1}), (\ref{deq2}) take form similar to the 1D case: $\hat{H}\Psi = \epsilon \Psi$, with \begin{equation} \hat{H}=\left( \matrix{ A_{{\bf p}} & B_{{\bf p}} \cr B_{-{\bf p}} & -A_{{\bf p}}}\right), \quad B_{{\bf p}} = \left(\matrix{\Delta_{-{\bf p}} &0 \cr 0 & \Delta_{-{\bf p} +{\bf Q}}}\right), \label{HH} \end{equation} \begin{equation} A_{{\bf p}} = \left(\matrix{-iV_{{\bf p}}\nabla_{{\bf r}} +\epsilon_{{\bf p}} -\eta& m_{\pm} ({\bf r})\cr m^*_{\pm} ({\bf r}) & -iV_{{\bf p}-{\bf Q}}\nabla_{{\bf r}} +\epsilon_{{\bf p} -{\bf Q}} -\eta}\right), \label{ca} \end{equation} where $\eta({\bf r}) = \mu - \frac{U}{2}\rho({\bf r} )$, $\Psi^T = (u_{{\bf p}}, u_{{\bf p}-{\bf Q}}, v_{{\bf p}}, v_{{\bf p}-{\bf Q}}) =(u_+, u_-, v_+, v_- )$, $\epsilon_{{\bf p}} = -2t(\cos p_x + \cos p_y ) -\mu$, ${\bf V}_{{\bf p}}=2 t (\sin p_x, \sin p_y )$, and, as before, ${\bf Q} = {\bf Q}_+$ for wave vectors ${\bf p}_y > 0$ and ${\bf Q} = {\bf Q}_-$ for wave vectors $p_y < 0$.The sign in $m_{\pm}$ is taken by the same rule. In the presence of the magnetic field ${\bf H}$ functions $u, v$ becomes spin-dependent, and equations (\ref{HH}), (\ref{ca}) are changed: $\nabla \to \nabla - i \frac{e}{c} {\bf A}$, $\mu \to \mu_{\sigma} =\mu + \sigma H$. For the constant magnetic field perpendicular to the plane we have ${\bf A} = {\bf H} \times{ \bf r} /2$, so that $\partial_x \to \partial_x +i(e/c)yH/2$, $\partial_y \to \partial_y -i(e/c)xH/2$. We suppose everywhere that the magnetic field is small $H \sim H_{c_1} \ll H_{c_2}$, ($\xi \ll \lambda$), therefore ignore the effect of terms with vector-potential on the solution at distances $r \ll \lambda$. For the case $d_{x^2 -y^2}$ symmetry we consider $\Delta_{-{\bf p}}= \Delta_{{\bf p}}=-\Delta_{{\bf p} -{\bf Q}}= 2(\cos p_x - \cos p_y ) \Delta_d ({\bf r})$, which corresponds to $\sigma \Delta({\bf r},{\bf r} \pm \hat{x};\sigma ) =-\sigma \Delta({\bf r},{\bf r} \pm \hat{y};\sigma ) = \Delta_0$ in the uniform ground state. We retained the main terms in the expansion over $\Delta$. In the higher order approximation, instead of the terms $\Delta_{-{\bf p}}$ and $\Delta_{-{\bf p} + {\bf Q}}$ we would have to write $\Delta_{-{\bf p}} -i(\nabla_{{\bf p}}\Delta_{-{\bf p}})\nabla_{{\bf r}}$ and $\Delta_{-{\bf p} +{\bf Q}} - i(\nabla_{{\bf p}} \Delta_{-{\bf p} + {\bf Q}})\nabla_{{\bf r}}$. The continuum approximation is not valid for a band filling very close to the half-filled case (the number of particles per one site $\rho =1$), where the Fermi velocity tends to zero at points ${\bf p} = (0, \pm \pi), (\pm \pi, 0)$. In the homogeneous case $\rho(x)=$ const, $m , \Delta_d = const$ for coexisting spin- and superconducting order parameters, so the eigenvalue spectrum has the form \begin{equation} E^2 = (\sqrt{m^2 + \epsilon^2 ({\bf p}) } \pm \eta )^2 + \Delta_{{\bf p}}^2, \end{equation} with $\Delta_{{\bf p}} = \Delta_d (\cos p_x - \cos p_y )$. The self-consistent conditions are derived by substitution of functions $u$, $v$ into (\ref{selfc}), similar to the one-dimensional case. In the continuum approximation they read: \begin{equation} \rho({\bf r}) = 2\sum_{\epsilon} [(u_+^* u_+ + u_-^* u_- )f + (v_+^* v_+ + v_-^* v_- )(1-f)] \label{s1} \end{equation} \begin{equation} (-1)^{x_i + y_i} m ({\bf r}) = 4 U[\sum_{\epsilon} u_-^* u_+ f - \sum_{\epsilon} v_-^* v_+ (1-f)] \label{s2} \end{equation} \begin{eqnarray} \Delta_{\bf q}({\bf r}) =2g\sum_{\epsilon} (v_{+}^* u_{+} -v_{-}^* u_{-}) [(1-f)(\cos (p_x-q_x) \nonumber \\ +\cos (p_y-q_y) )- f ((\cos (p_x + q_x) +\cos (p_y + q_y ) ) ], \end{eqnarray} where $f = {1}/({\exp[\epsilon /T] +1})$. We omitted spin indices since in our representation for wave functions all equations are diagonal over spin. \section{Spin-Charge Density Wave Structures} In the low doping limit the ground state of the model is the periodic charge-spin superstructure with the absence of superconductivity: $\Delta \equiv 0$. Consider different structures, having close ground state energies. In real systems the exact ground state must be determined by taking into account real long-distance 3D interactions. \subsection{Diagonal and vertical stripes} For diagonal stripes we search the solution in the form : \[ u_{{\bf p}}({\bf r}) = {u}_{{\bf p}} (r_+ ),\quad v_{{\bf p}}({\bf r}) = {v}_{{\bf p}} (r_+ ), \] where $r_{\pm} = (x \pm y)/\sqrt{2}$, ${\bf p} \in FS $. Substituting into Eqn. (\ref{ca}) we obtain a one-dimensional eigenvalue equation \begin{eqnarray} -iV_p \frac{\partial}{\partial r_+} u_+ +\frac{U}{2}\rho (r_+)u_+ + m(r_+) u_- = E u_+, \nonumber \\ m^* (r_+) u_+ + i V_p \frac{\partial}{\partial r_+} u_- +\frac{U}{2}\rho (r_+)u_- = E u_-, \end{eqnarray} where $V_p = 2t\sin p_x$. The only difference from the considered one-dimensional model\cite{mm} is the dispersion of the velocity $V_p$. This system is exactly solvable. In the ground state, at $\rho = 1$, we have $m (r_+) = m_0$. Increased doping leads to the stripe structure. The one stripe solution has the form \begin{equation} m (r_+) = m_0 \tanh \frac{r_+}{\xi}, \label{discrete} \end{equation} where the width $\xi$ is defined from the minimum of the total energy. Solution (\ref{discrete}) corresponds to $\rho=1$ in the thermodynamic limit for number of holes per lattice site. In our case (\ref{discrete}) is valid only in the vicinity of each single stripe that enters a periodic superstructure called stripe-phase (compare \cite{schulz}). Distinct from the Peierls model, where $\xi = V_F/m_0$, $V_F = const$, the present model has a more complicated spectrum. Besides continuum bands $E^2 = V_{k}^2 k_{\parallel}^2 + m_0^2$ we find some discrete levels (for a given $p_x$) inside the gap: \begin{equation} E_n^2 = m_0^2 \lambda n(2-\lambda n), \label{En} \end{equation} where $n$ is integer number, $0\leq n \leq 1/\lambda$, and $\lambda =\lambda (p_x) =V_p /(\xi m_0)$, or $\lambda = V_p/\bar{V}$ with $\xi \equiv \bar{V}/m_0$. Each level inside a gap forms a band due to dispersion of the coefficient $\lambda (p_x)$. For $\lambda \geq 1$ we obtain only one level, $E=0$, with wave function: \begin{equation} \psi_{\pm} = (u_+ \pm u_-)/\sqrt{2} \propto \frac{1}{(\cosh r_+/\xi )^{1/\lambda }}. \label{uv} \end{equation} The wave functions of all states are described in terms of the hypergeometric function $F(a,b |c| z )$, and for local levels they have polynomial form: \begin{equation} \psi_{\pm, n} \sim\frac{1}{(\cosh x/\xi )^{1/\lambda -n}} F[\frac{2}{\lambda} +1 -n, \, -n, \, \frac{1}{\lambda}-n +1,\, \frac{1}{2}(1 + \tanh \frac{x}{\xi} )] \label{uvn} \end{equation} For $1/2 <\lambda < 1$ two levels $n=0,1$, can already exist. Allowing for equality $\lambda = V_p/\bar{V}$ we conclude that both possibilities $\lambda <1$ and $\lambda >1$ take place, each one in the proper interval of $p_x$. Similar to the 1D case \cite{mm} the appearance of the kink in the spin channel is accompanied by the local charge distribution $\rho (r_+) - \langle \rho \rangle \sim 1/\cosh^2(r_+ /\xi )$. An increase of the doping leads to the periodic spin-charge density superstructure. In the limiting case of "overdoping" ($\mid \rho - 1 \mid \gg 1/\xi$) the spin-charge structure becomes harmonic \[ m (r_+) \propto \sin (\pi \mid \rho - 1 \mid r_+), \, \rho(r_+) - \langle \rho \rangle \propto \cos(2 \pi \mid \rho -1\mid r_+). \] For vertical stripes we use an ansatz: \[ u_p({\bf r}) = u_p(x ), \quad v_p ({\bf r} ) = v_p (x) \] and obtain a system of equations with $V_p =2t\sin p_x$, which is similar to the diagonal case. For the same values $m_0$, $\xi$, the parameter $\lambda$ in the considered case is less than for diagonal stripes. Therefore the condition $n < 1/\lambda$ can be valid for larger values of $n$, resulting in additional bands inside the gap, as it is seen from numerical results. \subsection{Checkerboard structure} As we have seen the spin-charge density structure may be arranged in vertical (horizontal) or diagonal directions. Consider the solution with square symmetry. In the same approximation as before we find the solution of system (\ref{ca}) in the form $m_{\pm}({\bf r}) = m ({\bf r}_{\pm})$, $u_p({\bf r}) = {u}_p ({\bf r}_{\pm})$, $v_{p} ({\bf r} ) = {v}_p ({\bf r}_{\pm})$. Equations are decoupled and we obtain \begin{eqnarray} -iV_p\frac{d {u}_+}{dr_{\pm}} + m (r_{\pm}) {u}_- = E {u}_+\\ m^* (r_{\pm}) {u}_+ + i V_p \frac{d{u}_-}{d r_{\pm}} = E{u}_-, \end{eqnarray} with $V_p = 2t \sin p_x$, $r_{\pm}= (\pm x + y)/\sqrt{2}$. The one "cross" solution has the form \begin{equation} m_+ = m_0 \tanh \frac{r_+}{\xi},\qquad m_- = m_0 \tanh\frac{r_-}{\xi}, \end{equation} The spectrum $E$ and wave functions are found as above for the case of stripes. In the case of high doping the one kink solution is transformed to the periodic structure \[ \langle S_z ({\bf r}) \rangle \propto (-1)^{x+ y} m_0 \cos[ \pi (\sqrt{\rho} -1 )x] \cos [ \pi (\sqrt{\rho} - 1)y], \] in which we considered the squared Fermi surface approximation with electron density $\rho = |{\bf Q}|^2/ 2\pi^2$. \section{Superconductivity and spin-charge modulation} \subsection{Vortex solution} Consider pure superconducting state ($\Delta({\bf r})\equiv 0$). The BdG equations are decoupled. The first pair is \begin{equation} -i {\bf V}_{{\bf p}} \nabla_{{\bf r}} u_{{\bf p}} ({\bf r}) + \Delta_{{\bf p}}v_{{\bf p}} = \epsilon u_{{\bf p}} \label{v1} \end{equation} \begin{equation} \Delta_{{\bf p}}^* u_{{\bf p}} + i{\bf V}_{{\bf p}} \nabla_{{\bf r}}v_{{\bf p}} ({\bf r}) =\epsilon v_{{\bf p}}. \label{v2} \end{equation} When the filling $\rho$ is close to 1, the Fermi surface has nearly square form, therefore ${\bf V}_{{\bf p}} \nabla_{{\bf r}} \approx V_{{\bf p}} \partial/\partial r_{\pm}$, depending on signs $p_x, \, p_y$ In this case the system of equations (\ref{v1}), (\ref{v2}) has the following vortex solution: \begin{equation} \Delta_{{\bf p}}({\bf r} ) = \Delta_p\frac{\sinh \frac{r_+}{\xi_s} + i \sinh \frac{r_-}{\xi_s}} {\sqrt{\sinh^2 \frac{r_+}{\xi_s} + \sinh^2 \frac{r_-}{\xi_s} + 1}}, \end{equation} where $\Delta_p = \Delta_0 (\cos(p_x) - \cos(p_y)) $. For $r_- = 0$ the order parameter has a kink form $\Delta_p({\bf r}) \propto \tanh r_+ /\xi_s$. For the case $r_+ = 0$ the order parameter acquires the phase: $\Delta \propto \exp( i \pi / 2)\tanh r_- /\xi_s$. In the diagonal direction $r_+ = r_-$ the solution $\Delta_p({\bf r}) \propto \tanh r_+ /\xi /\sqrt{\tanh^2 r_+ /\xi_s + 1} \exp(i \pi /4) $ has the phase $\pi /4$. It is known that in one-dimensional case finite-band solutions of equations (\ref{v1}) - (\ref{v2}) are related to the soliton (kink) solutions of the nonlinear Schrodinger equation (NSE). Note, that along the curve $ \sinh r_- /\xi_s = \alpha \cosh r_+ /\xi_s$ the order parameter acquires the form of a general kink solution of the NES: $\Delta_p({\bf r}) \sim (i \alpha + \tanh x/\xi_s )/\sqrt{\alpha^2 + 1}$ with the localized state in the gap with the energy $E_0 = \Delta_p \alpha /\sqrt{\alpha^2 + 1}$. \subsection{Coexistence of spin-charge structure and superconductivity} Consider solutions of equations (\ref{ca}) in the superconducting region. By analogy with 1D case \cite{ma} we use ansatz: \[ v_{\pm} = \gamma_{\pm} u_{\mp} \] which takes place in the uniform case. The term $U\rho({\bf r})/2$ in equations can be eliminated by the shift of wave functions $u,v \to u,v \exp i\Phi$, ${\bf V_p} \nabla \Phi =U \rho({\bf r} )/2$. Considering $\epsilon({\bf p} ) - \mu = 0$ on the Fermi surface we obtain for the case \[ m ({\bf r} ) = \vert m({\bf r}) \vert e^{ i \varphi},\, \Delta_{{\bf p}} ({\bf r}) = \vert \Delta_{{\bf p}} ({\bf r})\vert e^{i\varphi_s},\, \varphi,\, \varphi_s = const, \] the solution $ \gamma_+ = \pm i e^{i(\varphi -\varphi_s )}, \, \gamma_- = \pm i e^{-i(\varphi + \varphi_s)}$, and the system (\ref{ca}) acquires the form \begin{eqnarray} -i{\bf V}_{{\bf p}}\nabla u_+ + \tilde{\Delta}({\bf r})u_- = E u_+\label{se0}\\ \tilde{\Delta}^*({\bf r})u_+ + i{\bf V}_{{\bf p}}\nabla u_- = Eu_- \label{se} \end{eqnarray} with $\tilde{\Delta}({\bf r}) = (\vert m({\bf r})\vert \pm i\vert\Delta_{{\bf p}}\vert) e^{i\varphi}$, and, as before, $m = m_{\pm}$, depending on the sign of $p_y$. Equations (\ref{se0}), (\ref{se}) are exact provided that phases $\varphi$, $\varphi_s$ are constant or slowly varying in space functions. We show that inhomogeneity of the superconductor order parameter leads to the origination of the antiferromagnetic order parameter. Consider a 1D geometry case: $ u = u(r_+)$, where assumption of constant phases is valid. The solution of Eqs. (\ref{se0}), (\ref{se}) describing the coexistence of superconductivity and spin-charge density ordering, compatible with self-consistent equations, has the form of two bound solitons of the nonlinear Schrodinger equation, see, for example, \cite{fadtakh} \begin{equation} \tilde{\Delta}_{1,2} = \Delta_p \frac{\cosh (2 \kappa x + c_1) + \cosh (c_2 \pm 2 \beta i)/|\lambda| } {\cosh (2 \kappa x + c_1) + \cosh (c_2 )/|\lambda|}, \end{equation} where $\pm \lambda ({\bf p})$ are positions of local levels inside the gap, $\kappa = \sqrt{\Delta_p^2 - \lambda^2}$, and $\exp i\beta = \lambda + i\kappa$. Eigenfunctions of equations (\ref{se0}), (\ref{se}) have the form \cite{bm} \begin{equation} u_{\pm}(x) \propto \sqrt{\tilde{\Delta}(x) (E^2 - \gamma^2(x))}\exp \left[\pm i\int^x \frac{\sqrt{(E^2 - \Delta_{p}^2)(E^2 - \lambda^2)}}{E^2 - \gamma^2(y)} dy \right], \label{wavefunction} \end{equation} where \begin{equation} \gamma (x) = \frac{1}{2} \frac{\partial}{\partial x} \ln \tilde{\Delta}(x). \end{equation} For superconducting and spin order parameters we obtain \begin{equation} \Delta_{sc} = \Delta_p (1 - \Gamma \tanh a (\tanh(\frac{r_+}{\xi} + \frac{a}{2}) - \tanh (\frac{r_+}{\xi}- \frac{a}{2} ))), \end{equation} \begin{equation} m = m_0 \Gamma \tanh a (\tanh(\frac{r_+}{\xi} + \frac{a}{2}) - \tanh (\frac{r_+}{\xi}- \frac{a}{2} )) \end{equation} where averaged over Fermi surface functions are defined as: $\Gamma = <\Gamma_p>=<\Delta^2_p /(\Delta_p^2 + m_0^2)>_p $, $\xi = <v_p/(\Delta_p \sqrt{\Gamma_p} \tanh a)>_p$, and we use the parametrization for the local level $\lambda$: \begin{equation} \lambda^2 = \frac{\Delta_p^2}{\Delta_p^2 + m_0^2}\left(m_0^2 + \frac{\Delta_p^2}{\cosh^2 a}\right) \end{equation} Values of $m_0$, $\Delta_{{\bf p}} = \Delta_d (\cos p_x - \cos p_y)$, $\xi$ and a dimensionless parameter $a$ are defined by the self-consistent conditions (\ref{s1}), (\ref{s2}). The solution describes the spin-charge stripe in superconducting phase. \begin{figure}[tbph] \begin{center} \includegraphics[width=3.0in]{coex4.eps} \caption{ False color plot of the coexisting superconducting (downward) and antiferromagnetic stripe-like (upward) orders. The envelope functions are plotted in real space, $x$ and $y$ coordinates are measured in units of correlation length $\xi$, $a=5$, $\Delta_d=1$, $m_0=0.2$.} \end{center} \end{figure} The spin inhomogeneity generates the charge distribution $\delta \rho ({\bf r} ) \propto m^2({\bf r} )$. Note, that two-soliton solution in the similar form was used for describing polaron-bipolaron states in the Peierls dielectrics \cite{brkir}. The superconducting correlation length is increased in comparison to clean superconductor case $\xi_{sc}$ as $\xi = \xi_{sc} \sqrt{1+ \xi^2_{sc} /\xi^2_{AF}}$. \section{Discussion} We considered a simple self-consistent 2D model on a squared lattice to describe different states, including charge-spin structures, superconductivity, and their coexistence. The origin of spin-charge periodic state (which is responsible for the pseudogap) is due to the existence of flat parallel segments of the Fermi surface (nesting) at low hole doping concentrations. Effects of commensurability lead to a pinning of stripe structure at rational filling points $|\rho - 1| = m/n$. As a result, there is an exponentially small (for large $n$) decrease in the total energy of the order $\delta E \sim \exp( - c\, n)$ at any commensurate point, stabilizing stripes, as in 1D systems. For this reason, we think, stripes are mostly observable near $n = 8$ point ($|\rho- 1|= 1/8$). An increase of doping leads to the decrease of flat segments of the Fermi surface and attenuation of spin-charge structure. We found the solution describing the coexistence of superconductivity and stripes (28), (29). The decrease (or a deviation from the homogenous value ) of the superconducting order parameter generates the spin-charge periodic structure in this region. Note, that due to symmetry of Eqs. (25), (26) (duality $\Delta \leftrightarrow i m$) we can write the same equation, describing the origin of superconducting correlations in the region of a inhomogeneity of spin-charge density. The situation is qualitatively similar to the 1D case \cite{ma}. Experimental data in underdoped high-T$_c$ cuprates LSCO \cite{aeppli1} indicates that antiferromagnetic stripe-like spin-density order can be induced by magnetic field perpendicular to the CuO planes in the interval of fields much smaller than upper critical field H$_{c2}$ . The size of the magnetically ordered domains exceeds superconducting vortex's core size $\xi_s$ and the inter-vortex distance in the Abrikosov's lattice. Our present theoretical results demonstrate that this is indeed possible in the simple Hubbard t-U-V model that we consider. In particular, the dimensionless parameter $a$ in Eqs. (28), (29) is an independent variational parameter and depends on the magnetic and superconducting coupling strengths \cite{ma}, as well as on the magnitude of the external magnetic field. Hence, the size $\sim a\times \xi$ of the antiferromagnetic domain (see Fig. 2, upward red plane bump) can exceed the superconducting (and magnetic) Ginzburg-Landau correlation length $\xi$ when $a(H)>>1$. Previously coexistence of superconducting order and slow antiferromagnetic fluctuations was studied merely on the basis of a phenomenological Ginzburg-Landau free energy functional approach in \cite{sachdev1}. We note also, that equations Eqs. (25)-(26) can be simply extended to include d-density waves (DDW). \section{Acknowledgements} This work is in part supported by RFFI grant 12-02-01018-a.
\section{{\small INTRODUCTION}} General relativity is arguably the most successful theory of gravity to date. The theory, aside from being pleasing on grounds such as diffeomorphism invariance, has passed all major tests in the solar system. On scales larger than the solar system, general relativity, supplemented with non-standard matter fields such as dark matter or dark energy, explains the dynamics of the large-scale structure of the universe rather well. However, there has long been an interest in extended theories of gravitation, which possess general relativity in some limit (for example, see \cite{ref:faraonibook}, \cite{ref:caplau}.) These extensions are mainly motivated by the desire to eliminate the potentially ``exotic'' non-standard matter fields mentioned \cite{ref:noeg}, \cite{ref:noeg2}, or to mimic some low energy quantum gravity effects which are thought to manifest at very high curvatures \cite{ref:birrellanddavies}, \cite{ref:qeg}. Other motivations are perhaps more academic; we have only been able to study gravity effectively in its weak limit, and hence cannot be certain that Einstein gravity holds in stronger gravitational fields. Therefore it is of interest to study gravitational theories which are diffeomorphism invariant and give Einstein gravity in an appropriate limit, but deviate from Einstein gravity in some way outside of the realm where gravitational effects have commonly been observed. This, admittedly, gives us much freedom regarding the types of theories that may be successful candidates. Of particular interest are the class of theories known as $f(R)$ gravity. These theories consider gravitational fields generated by actions of the form \cite{ref:buchfr}: \begin{equation} S=\frac{1}{2\kappa}\int\,\sqrt{-g}\,f(R)\,d^{4}x + \int\,\sqrt{-g}\, \mathcal{L}_{\mbox{\tiny{mat}}}(g)\,d^{4}x \,, \label{eq:lag} \end{equation} where $f(R)$ is some function of the Ricci scalar, $R$ and $\mathcal{L}_{\mbox{\tiny{mat}}}(g)$ is the matter Lagrangian density. The equations of motion generated via varying this action with respect to the metric, and demanding that this variation vanish. These equations of motion are: \begin{equation} \tfrac{\partial f(R)}{\partial R} R^{\mu}_{\;\,\nu} -\frac{1}{2} f(R)\,\delta^{\mu}_{\;\,\nu} -\nabla^{\mu}\nabla_{\nu} \tfrac{\partial f(R)}{\partial R} +\delta^{\mu}_{\;\,\nu}\, \dalembertian \tfrac{\partial f(R)}{\partial R} = \kappa\,T^{\mu}_{\;\,\nu}\,, \label{eq:eom} \end{equation} where as usual $T^{\mu}_{\;\nu}$ is the stress-energy tensor of the matter field which comes from varying the matter action. A number of earlier generalizations to Einstein gravity fall within the paradigm of $f(R)$ theories. For example, theories with $f(R)=R+\alpha_{2}R^{2}$ have long been studied (called ``$R$-squared'' gravity. See, for example, \cite{ref:rsq1}-\cite{ref:rsqlast}). The Starobinsky inflationary theory \cite{ref:starob} may be the most popular application of $R$-squared gravity theory. In the opposite regime, where curvature is low, there has been less interest in modifying Einstein gravity. However, it is worth mentioning that $f(R)$ theory with \emph{inverse powers} of the Ricci scalar has been considered as a possible candidate to explain the observed accelerated expansion of the universe \cite{ref:o3}-\cite{ref:o6}. Inverse Ricci terms may also appear in certain sectors of string/M theory (see \cite{ref:odin}, \cite{ref:gzbr} for example). The vacuum state for inverse-$R$ theories is not Minkowski space-time, but is instead either deSitter or anti-deSitter space-time. Admittedly, these inverse theories are highly constrained from stability considerations or solar system tests but may not be completely ruled out. (See \cite{ref:dolgov}-\cite{ref:faraonisolar} and references therein for a discussion of the restrictions.) For generality we also include this sector in our study. A thorough study of solutions in $R^{n}$ cosmology may be found in \cite{ref:nodint2} as well as \cite{ref:cosdyn} and a complete discussion of $f(R)$ modifications to aid acceleration may be found in \cite{ref:lobo}. Outside of cosmology, and perhaps more relevant here, studies have been performed that have more bearing to stellar physics \cite{ref:frstars1}-\cite{ref:frstars9}. In another popular extension, namely that of $f(T)$ gravity where $T$ is the torsion scalar, such solutions have also been studied \cite{ref:s1}, \cite{ref:s2}. Given that integer powers of $R$ have arguably generated the most interest in $f(R)$ extensions, we propose here to study the properties of wormhole throats in gravity theories given by \begin{equation} f(R)={\underset{{\hspace{-0.05cm}\vspace{-0.1cm}n}}{\sum}} \alpha_{n}\,R^{n}\,, \label{eq:fofr} \end{equation} with the $\alpha_{n}$ constants. This is quite general as any $f(R)$ analytic in $R$ may be expanded as such a power series (in the positive $n$ sector) such as exponential gravity \cite{ref:o5}. For completion, we allow for both positive and negative powers of $R$ in the action (hence extending to $f(R)$ expandable in negative powers as well). Of course, this series includes the $n=1$ term (Einstein term) and can include $n=0$ (cosmological constant). The wormhole is of interest as it may provide a simple model for the space-time foam thought to be manifest at very high energies, where the corrections to Einstein gravity due to higher curvature terms may be important in encompassing some quantum gravity effects. As well, exhaustive studies have been performed on wormhole geometries in Einstein gravity and their properties (mainly in the static cases with spherical and axial symmetry) are now well known in Einstein gravity. For example, it is an interesting property that static wormhole throats necessarily must violate the weak/null energy conditions in Einstein gravity \cite{ref:WEC1}, \cite{ref:hochvis} and, furthermore, must be anisotropic. It has been of much interest in wormhole physics to find either somewhat realistic matter models which can meet these slightly exotic conditions in a way required to support a wormhole throat, or else to consider alternative gravitational theories which may allow for energy condition respecting matter. In the latter vein, due to the complexity of the resulting equations, the zero tidal force class of wormholes is most often studied ($g_{tt}=\mbox{const.}$). Even then, analytic models are hard to come by and one often resorts to numerics. The zero tidal force models are interesting due to their tractability, which allows one to study important properties of these geometries. However, they may not be very realistic from a physics perspective though, as they yield a constant frequency shift in inhomogeneous structures possessing a preferred center and, in the weak-field limit, a constant Newtonian potential. In this work we study wormhole \emph{throats} and \emph{without} the restriction of zero tidal force, although we do consider the zero tidal force cases as well for completion. The consideration of the throat is in many ways more general than considering the asymptotics, as the most salient features of a wormhole occur in the throat region. For example, in Einstein gravity, the necessary violation of energy conditions in static spherically-symmetric wormholes occurs in the neighborhood of the throat, regardless of asymptotics. In fact, in the Einstein gravity scenarios, if the matter field falls off sufficiently fast, or is patched to a vacuum or other solution, the properties far from the throat do not generally differ greatly from similar systems with trivial topology. As well, it has been argued (for example in \cite{ref:hochvis}) that the global topology is too limited a tool to study wormholes, and a local geometric analysis near the throat is generally more useful in discerning interesting properties of wormholes. In such cases, the throat does not necessarily coincide with the common definition of a wormhole, which relies on global properties, and is viewed as an interesting object in its own right, capable of describing more general scenarios than just wormholes. We therefore now concentrate on the near throat region and study the properties in this vicinity. Admittedly, in some cases, demanding flatness at infinity would place extra restrictions on the properties of the wormhole \cite{ref:GS}, \cite{ref:bronnfr} but as the asymptotics of the universe are not exactly known, and we are interested in the existence of throats only, we will not consider such restrictions here. In section 2 we briefly review some background material and present a non-standard coordinate gauge that is more suited for wormhole analysis than the standard spherical coordinate chart. We discuss the mathematical construction of the throat in this chart. This is followed by a study, which is analytic when possible but otherwise numerical, of various scenarios depending on the terms present in the gravitational action. {Of special importance, due to their physical relevance, are the non-zero tidal force models, which we examine in some detail.} We particularly concentrate on the properties of energy conditions as well as the degree of anisotropy. {In general, we find that the parameter space of Einstein gravity supplemented with terms where $n$ is positive is more favorable in both these respects than Einstein gravity alone for a large range of parameters. Supplements with negative $n$ have either a very small parameter space where they improve the energy conditions or else tend to worsen the situation when compared to Einstein gravity alone.} Finally, in section 3, we conclude the study. \section{{\small THE MODELS}} As is common in wormhole studies in Einstein gravity, we will utilize here an anisotropic fluid source whose stress-energy tensor is given by: \begin{equation} T^{\mu}_{\;\,\nu}=(\rho + p_{t})u^{\mu}u_{\nu} + p_{t}\,\delta^{\mu}_{\;\,\nu} + (p_{r} - p_{t})s^{\mu}s_{\nu}\,. \label{eq:anisoT} \end{equation} Here $\rho$, $p_{t}$ and $p_{r}$ are the energy density, the perpendicular (to the inhomogeneous direction) pressure, and the parallel pressure respectively as measured in the fluid element's rest frame. The vector $u^{\mu}$ is the fluid 4-velocity and $s^{\mu}$ is a space-like vector orthogonal to $u^{\mu}$. These vectors satisfy: \begin{equation} u^{\mu}u_{\mu}=-1,\;\;\; s^{\mu}s_{\mu}=+1,\;\;\; u^{\mu}s_{\mu}=0\,. \label{eq:fluidvecs} \end{equation} Since we are interested in the properties of the actual material generating the gravitational field, we \emph{do not} transform the system into an effective scalar-tensor system but instead keep the geometry and the material properties explicit. We impose spherical symmetry and hence may write the metric as \begin{align} ds^{2}=&-e^{\Phi(r)}\,dt^{2} + e^{\Xi(r)}\,dr^{2} + r^{2}\,d\theta^{2} + r^{2}\sin^{2}(\theta)\,d\varphi^{2}\,, \label{eq:spherechart} \\[0.2cm] =& -e^{\Phi(r)}\,dt^{2} + \left\{1+ \left[\partial_{r} P(r)\right]^{2} \right\}\,dr^{2} + r^{2}\,d\theta^{2} + r^{2}\sin^{2}(\theta)\,d\varphi^{2}\,, \nonumber \end{align} where the function $P(r)$ describes the profile of the three geometry for fixed $\theta$ and $\varphi$ (see below for details). In this chart the Ricci scalar is given by \begin{align} R=&\frac{1}{2r^{2}\left[1+\left(P^{\prime}\right)^{2}\right]^{2}}\Bigl[ 4(P^{\prime})^{4} -4r(P^{\prime})^{2}\Phi^{\prime} -4r\Phi^{\prime} +8rP^{\prime}P^{\prime\prime}+2r^{2}P^{\prime}P^{\prime\prime}\Phi^{\prime} \nonumber \\ &\qquad\qquad\qquad\quad +4(P^{\prime})^{2} -2r^{2}(P^{\prime})^{2}\Phi^{\prime\prime} -r^{2}(P^{\prime})^{2}(\Phi^{\prime})^{2} -2r^{2}\Phi^{\prime\prime} -r^{2}(\Phi^{\prime})^{2}\Bigr]\,, \label{eq:Rscal1} \end{align} where the prime denotes differentiation with respect to $r$. Although the above is the most common form of metric system to locally describe spherically symmetric gravitational fields, this coordinate system is not ideal for studies of wormhole throats and hence we use a different chart, which we now present. \subsection{{\footnotesize WORMHOLES}} We consider here static throats, which must be anisotropic in Einstein gravity, in the class of modified gravities given by $f(R)={\underset{{\hspace{-0.05cm}\vspace{-0.1cm}n}}{\sum}} \alpha_{n}\,R^{n}$ for both positive and negative $n$. As mentioned previously, it is common in the literature on wormholes to consider zero tidal force models via the condition $g_{tt}(r)=\mbox{constant}$. However here we relax this restriction and study more general scenarios. As well, although the spherical coordinate chart of (\ref{eq:spherechart}) can be utilized for wormhole throats if care is taken, it is not optimal, as $g_{rr}(r) \rightarrow \infty$ as $r \rightarrow r_{0}$ (see Figure \ref{fig:rotatechart} and equation (\ref{eq:spherechart})), and the most interesting properties of wormholes are arguably found near the throat region. Therefore, we use a different chart for the study of wormholes which essentially involves tilting the standard spherical chart by $\pi/2$. The wormhole throat is then given via the creation of a surface of revolution of the profile curve, $r=Q(x)$, as shown in Figure~\ref{fig:rotatechart} (also see figure caption). In this new chart, the space-time metric's line element may be written as \begin{equation} ds^{2}=-e^{\Phi(x)}\, dt^{2} + \left\{1+ \left[\partial_{x} Q(x)\right]^{2} \right\}\, dx^{2} + Q^{2}(x)\, d\theta^{2} + Q^{2}(x)\sin^{2}(\theta)\, d\varphi^{2}\,. \label{eq:newchartline} \end{equation} Note that in this new chart the metric is analytic at the throat since $\partial_{x} Q(x) \rightarrow 0$, and hence $g_{xx}(x) \rightarrow 1$ as one approaches the throat ($x=0$)\footnote{As an aside, in this coordinate system the exterior Schwarzschild metric is given by $Q(x)=2M+\frac{x^{2}}{8M}$. The horizon is located at $x=0$.}. Also, only one chart is required now to cover the wormhole, as opposed to two charts in the standard spherical coordinates. The function $Q(x)$ must possess the following properties: \begin{enumerate}[i)] \item $Q_{0}:=Q(0) > 0$,\vspace{-0.2cm} \item $\qzero^{\prime}:=Q^{\prime}(x)_{|x=0}=0$,\vspace{-0.2cm} \item $Q^{\prime\prime}(x) > 0$ in some neighborhood of the throat\footnote{More precisely, if $Q$'s first non-zero derivative (higher than first order) at $x=0$ is of even order, the function attains a local minimum if this derivative is positive, and hence we have a wormhole throat. If its first non-zero derivative is of odd order, it is a point of inflection and therefore does not describe a wormhole throat.}. \end{enumerate} Aside from the above properties we make the mild assumption that $Q(x)$ is analytic in some nonzero domain about $x=0$. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[bb=0 0 1205 485, scale=0.325, clip, keepaspectratio=true] {rot_prof_curve2.eps} \end{center}\vspace{-0.1cm} \caption{\small{a) In the standard spherical chart the wormhole throat is generated by the surface of revolution created by rotating the profile curve $x=P(r)=P_{-}(r) \bigcup P_{+}(r)$ but this leads to a metric singularity at the throat ($r=r_{0}$) where $\partial_{r}P(r) \rightarrow \infty$. A new chart is created in b) via rotating the coordinates by $\pi/2$, and the surface of revolution (inset) is generated by rotating the curve $r=Q(x)=P^{-1}(x)$. This does not have the metric singularity at the throat ($x=0$). Also, a single chart can now cover both regions.}}% \label{fig:rotatechart}% \end{framed} \end{figure} In the above coordinate chart, the Ricci scalar is given by \begin{align} R=&\frac{1}{2Q^{2}\left[1+\left(Q^{\prime}\right)^{2}\right]^{2}}\Bigl[4-4Q(Q^{\prime})^{3}\Phi^{\prime} + 4(Q^{\prime})^{2} -8QQ^{\prime\prime}-4QQ^{\prime}\Phi^{\prime} +2Q^{2}Q^{\prime}Q^{\prime\prime}\Phi^{\prime} \nonumber\\ &\qquad\qquad\qquad\quad\quad -2Q^{2}\Phi^{\prime\prime} -Q^{2}(\Phi^{\prime})^{2}-2Q^{2}(Q^{\prime})^{2}\Phi^{\prime\prime} -Q^{2}(Q^{\prime})^{2}(\Phi^{\prime})^{2} \Bigr]\,, \label{eq:Rscal2} \end{align} where here and subsequently the prime denotes differentiation with respect to $x$. We do not expect any serious pathologies (such as infinite tidal forces) in our analysis as, in the domain of validity of the coordinate chart (which is larger than the ``$r$'' chart, and covers the throat) the metric and its derivatives are well behaved and hence we expect the orthonormal Riemann tensor components to also be well behaved. We show later that the function $Q(x)$ describing the spatial geometry, is $C^{\infty}$ and that $\Phi(x)$ is also very smooth. One issue which is of great interest in wormhole physics is the study of energy condition violation of the matter supporting the wormhole. It is known that in Einstein gravity energy conditions \emph{must} be violated by static wormholes somewhere in the vicinity of the throat \cite{ref:WEC1}, \cite{ref:hochvis}. It is possible that in more complicated gravitational theories, one may circumvent this issue and possess a throat region which respects energy conditions. An analysis in this vein was performed in \cite{ref:furdeb} where the analysis was limited to $n=-1$ and $n=2$ extensions only. A very complete analysis may be found in \cite{ref:lobooliv}, \cite{ref:olivthesis}. In the above cited studies, however, due to the complications involved, the zero tidal force assumption was made. This simplification eliminates some of the complications but still allows one to glean a number of the interesting properties that $f(R)$ wormholes possess. However, constant redshift, and constant weak-field Newtonian potential, from various regions of an inhomogeneous gravitating object may not be particularly realistic and we wish to remove this assumption and allow for a non constant $\Phi(x)$ function. This can alter the physical properties of the throat significantly, as we shall see below. Aside from yielding more complicated equations, a non-constant $\Phi(x)$ now presents us with the problem of how to prescribe this function. For this we appeal to some physical considerations. It is more physical to demand some realistic properties on the matter fields than it is to blindly prescribe $\Phi(x)$, although the former is much more difficult than the latter, as one needs to solve a complicated differential equation for $\Phi(x)$. With this in consideration, what we initially do is prescribe some reasonable energy density profiles, $\rho(x)$, for the matter field (eg. large positive energy density with zero slope at the throat and monotonically decreasing outward towards the surface or infinity, using parameters where this is possible) and numerically solve for $\Phi(x)$ using the numerical code COLSYS \cite{ref:colsys}. This will then give us an idea of what a realistic $\Phi(x)$ function should look like, and we use similar functions for subsequent studies. (The spatial geometry must also be prescribed, and this poses no problem as we apriori know that, spatially, we are dealing with a spherically symmetric wormhole.) We simply use this method to provide an idea of what the functions $\Phi(x)$ should look like in realistic scenarios and then use similar functions for our studies. It should be noted that the function $\Phi(x)$ controls the presence of event horizons. We avoid horizons by ensuring that $\Phi(x)$ does not approach $-\infty$ anywhere in the domain of study. In the following analysis we wish to discern how the various terms in the action affect the wormhole solution. In the case of $f(R)={\sum} \alpha_{n}\,R^{n}$, each term in the sum contributes independently to the left-hand side of the field equations, and hence also contribute separately in determining the properties of $T^{\mu}_{\;\nu}$. Therefore, it is sufficient to study individual terms in the power series of $f(R)$ separately to see how they each contribute individually to the matter field. Note that in this approach, $\alpha_{n}$ is simply an overall multiplicative constant, and hence we set it equal to one. One may see the effect of changing $\alpha_{n}$ by simply rescaling the results below by whatever value of $\alpha_{n}$ one chooses (with the exception of the anisotropy studies, for which we study the effects of varying $\alpha_{n}$). Negative values of $\alpha_{n}$ also reflect the graphs about the $z=0$ plane. \subsubsection{{\normalsize $n=1$}} This case, of course, corresponds to Einstein gravity. It is worthwhile presenting this to summarize the results of Einstein gravity in this coordinate chart, and for comparison with other powers of $n$ later. Regardless of the powers present in the full action, it is expected to have an $n=1$ term. We summarize the $n=1$ results as follows: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} \tilde{\rho}:=&\kappa\,\rho=-\kappa\,T^{0}_{\;\,0}=\frac{1+(Q^{\prime})^{2}-2QQ^{\prime\prime}}{Q^{2}\left[1+\left(Q^{\prime}\right)^{2}\right]^{2}}\,, \label{eq:einstrho}\\[0.2cm] \tilde{p}_{r}:=&\kappa\,p_{r}= \kappa\,T^{1}_{\;\,1}=\frac{\Phi^{\prime}QQ^{\prime}-1}{Q^{2}\left[1+\left(Q^{\prime}\right)^{2}\right]}\,, \label{eq:einstpr}\\[0.2cm] \tilde{p}_{t}:=&\kappa\,p_{t}= \kappa\,T^{2}_{\;\,2}=\frac{1}{4Q\left[1+\left(Q^{\prime}\right)^{2}\right]^{2}} \left[ 4Q^{\prime\prime} +2 Q^{\prime}\Phi^{\prime} +2(Q^{\prime})^{3}\Phi^{\prime} +Q(Q^{\prime})^{2}(\Phi^{\prime})^{2} \right. \nonumber \\[0.1cm] &\qquad\qquad\qquad \left.-2QQ^{\prime}Q^{\prime\prime}\Phi^{\prime} +2Q(Q^{\prime})^{2}\Phi^{\prime\prime} +Q(\Phi^{\prime})^{2} + 2 Q \Phi^{\prime\prime}\right]\,. \label{eq:einstpt} \end{align}} \end{subequations} There are also the following {combinations} which occur in the energy conditions: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} \tilde{\rho}+\tilde{p}_{r}=& \frac{Q^{\prime}\Phi^{\prime}+(Q^{\prime})^{3}\Phi^{\prime}-2Q^{\prime\prime}}{Q\left[1+\left(Q^{\prime}\right)^{2}\right]^{2}}\,, \label{eq:einstecond1} \\[0.2cm] \tilde{\rho}+\tilde{p}_{t}=&\frac{1}{4Q^{2}\left[1+\left(Q^{\prime}\right)^{2}\right]^{2}} \left[4+4(Q^{\prime})^{2} -4QQ^{\prime\prime} +2 QQ^{\prime}\Phi^{\prime} +2 Q (Q^{\prime})^{3}\Phi^{\prime} \right. \nonumber \\[0.1cm] &\left.+Q^{2}(Q^{\prime})^{2}(\Phi^{\prime})^{2} - 2 Q^{2}Q^{\prime} Q^{\prime\prime} \Phi^{\prime} +2Q^{2}(Q^{\prime})^{2}\Phi^{\prime\prime} +Q^{2}(\Phi^{\prime})^{2} +2Q^{2}\Phi^{\prime\prime} \right]\,.\label{eq:einstecond2} \end{align}} \end{subequations} The above expressions, though not overly complicated, do not shed much insight into the behavior of the matter fields near the throat, where we are most interested in. We therefore expand the relevant expressions in Taylor series about the throat ($x=0$): \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} &\tilde{\rho}= \frac{1}{Q_{0}}\left[1-2 Q_{0} Q_{0}^{\prime\prime}\right] -2 \frac{Q_{0}^{\prime\prime\prime}}{Q_{0}} x + \frac{1}{Q_{0}^{3}} \left[4Q_{0}^{2} (Q_{0}^{\prime\prime})^{3} - Q_{0}^{\prime\prime} -Q_{0}^{2}Q_{0}^{\prime\prime\prime\prime}\right] x^{2} + \mathcal{O}(x^{3})\,, \label{eq:einstnearthroatrho} \\[0.2cm] &\tilde{\rho}+\tilde{p}_{r}= -2\frac{Q_{0}^{\prime\prime}}{Q_{0}} + \frac{1}{Q_{0}} \left[Q_{0}^{\prime\prime} \Phi_{0}^{\prime} -2 Q_{0}^{\prime\prime\prime}\right]x + \frac{1}{Q_{0}^{2}} \Bigl[4Q_{0}(Q_{0}^{\prime\prime})^{3} -Q_{0} Q_{0}^{\prime\prime\prime\prime} + \frac{1}{2} Q_{0} Q_{0}^{\prime\prime\prime}\Phi_{0}^{\prime} \nonumber \\[0.1cm] &\qquad\qquad\quad +Q_{0}\qzero^{\prime\prime}\Phi_{0}^{\prime\prime} + (Q_{0}^{\prime\prime})^{2}\Bigr] x^{2} +\mathcal{O}(x^{3})\,, \label{eq:einstnearthroate1} \\[0.2cm] &\tilde{\rho}+\tilde{p}_{t}= \frac{1}{4Q_{0}^{2}}\left[4+Q_{0}^{2}(\Phi_{0}^{\prime})^{2}-4Q_{0}\qzero^{\prime\prime}+2Q_{0}^{2}\Phi_{0}^{\prime\prime}\right] +\frac{1}{Q_{0}}\left[Q_{0}\Phi_{0}^{\prime}\Phi_{0}^{\prime\prime} -Q_{0}(\qzero^{\prime\prime})^{2}\Phi_{0}^{\prime}\right. \nonumber \\[0.1cm] &\qquad\qquad\quad + \left.Q_{0}\Phi_{0}^{\prime\prime\prime} -2\qzero^{\prime\prime\prime} +\qzero^{\prime\prime}\Phi_{0}^{\prime} \right] x +\mathcal{O}(x^{2})\,. \label{eq:einstnearthroate2} \end{align}} \end{subequations} where the zero subscript indicates that the quantity is evaluated at $x=0$. From the (\ref{eq:einstnearthroatrho}), it can be seen that the energy density in the throat region may be made positive. Regarding (\ref{eq:einstnearthroate1}), if $Q_{0}^{\prime\prime}$ is non-zero it must be positive (as the throat is a local minimum) and (\ref{eq:einstnearthroate1}) must therefore be negative near the throat. If $Q_{0}^{\prime\prime}$ is zero, then the condition for a local minimum implies that $Q_{0}^{\prime\prime\prime}$ is also zero, and hence the lowest order term which would contribute near the throat is the fourth derivative term, $\qzero^{\prime\prime\prime\prime}$. Since this fourth derivative must then be positive under the condition of a minimum, and it appears with a negative sign in (\ref{eq:einstnearthroate1}), this contribution is negative. In such a scenario energy conditions are (barely) met at the throat, but are violated as one moves away from the throat. Similar arguments apply to higher derivatives in case the fourth derivative vanishes at the throat (although one needs to study higher order terms in the expansion, which for brevity we did not write). It is in this way that energy conditions must be violated in the vicinity of a static wormhole throat in Einstein gravity. For comparison with some of the modified gravity results to follow, which, due to the complicated expressions they yield, must be studied via computational methods, we choose a spatial geometry governed by \begin{equation} Q(x)=A\cosh\left(\frac{x}{x_{o}}\right)\,, \label{eq:Qcosh} \end{equation} as this function possesses all the salient properties to describe a throat. The parameter $A$ represents the radius of the throat and $x_{0}$ represents the degree of ``flare-out'' of the wormhole. Both these parameters are strongly related to the degree of energy condition violation, and hence we pay particular attention to these quantities. For all cases studied in this work, when an explicit form of $Q(x)$ is required, we use the form in (\ref{eq:Qcosh}). For the function $\Phi(x)$ we appeal to physical considerations (with the aid of the numerical code COLSYS, as mentioned previously). For positive $n$ we expect $\Phi(x)$ to slowly asymptote to a constant value far away from the throat, where the geometry is expected to approach Minkowski space-time. For negative $n$ (and for generality, for all $n$) the asymptotic value should approach the deSitter or anti-deSitter value. As we are interested in the near-throat region only it is not crucial that the asymptote is manifest in the domain of consideration, but an indication of such an asymptote is desirable. We consider both scenarios where $g_{tt}=-e^{\Phi(x)}$ smoothly increases towards the asymptote, as well as scenarios where $g_{tt}$ smoothly decreases towards the asymptote. (One cannot necessarily rule out one scenario over the other in extended gravity theories.) Specifically, we choose fitting functions of the form \begin{subequations} \begin{align} g_{tt}(x)=&\mathsf{A}_{0}\frac{\mathsf{B_{0}}+x^2}{1+x^2}-\mathsf{C}_{0} & \mbox{for concave-down $g_{tt}$\,,} \label{eq:gttfuna}\\ g_{tt}(x)=& -\frac{\mathsf{A}_{0} + \mathsf{B}_{0}x^{2}}{\mathsf{C}_{0}+x^{2}} & \mbox{for concave-up $g_{tt}$\,,}\label{eq:gttfunb} \end{align} \end{subequations} Where $\mathsf{A}_{0}$, $\mathsf{B}_{0}$ and $\mathsf{C}_{0}$ are fitting constants to fit generally to the COLSYS numerical results. Since the wormhole profile chosen in (\ref{eq:Qcosh}) is symmetric about the throat, we also make similar symmetry demands on $g_{tt}$. There is no great loss of generality in doing this since if one wishes to model a non-symmetric throat, the solution presented can be viewed as being valid only on one side of the throat, and a different solution can be patched to the other side. The forms of (\ref{eq:gttfuna},b) are used throughout this manuscript in the analysis of non-zero tidal force models. In the figures below (Figures \ref{fig:einstxoconcdown} - \ref{fig:einstradiusconcupaniso}) we display the results of Einstein theory ($n=1$) for the spatial geometry governed by (\ref{eq:Qcosh}). The first set of figures (figs.~\ref{fig:einstxoconcdown}a-d) show the behavior of $g_{tt}$, along with the energy conditions (\ref{eq:einstrho}) and (\ref{eq:einstecond1},ii) as a function of the \emph{flare-out parameter}, $x_{0}$. Although the energy density is positive (as is $\tilde{\rho}+\tilde{p}_{t}$), the other energy condition is not positive. We can also see here the well-known situation that the greater the degree of flare-out (equivalent to small $x_{0}$) the more severe the energy condition violation in Einstein gravity. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_rho_c_down_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_econd1_c_down_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_econd2_c_down_vary_xo.eps}} \end{tabular} \end{center} \caption{\small{Einstein gravity, $g_{tt}$ \emph{decreasing}, throat radius=0.05, varying flare-out $x_{0}$.}} \label{fig:einstxoconcdown} \end{framed} \end{figure} Figure \ref{fig:einstradiusconcdown} displays the behavior of $g_{tt}$, along with the energy conditions (\ref{eq:einstrho}) and (\ref{eq:einstecond1},ii) as a function of the \emph{throat radius}, $A$. Note that although the energy density can be made positive near the throat if the throat is not too large, the quantity (\ref{eq:einstecond1}) is negative in the throat vicinity, confirming the analytic analysis above. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{einst_rho_c_down.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{einst_econd1_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{einst_econd2_c_down.eps}} \end{tabular} \end{center} \caption{\small{Einstein gravity, $g_{tt}$ \emph{decreasing}, $x_{0}=1$, varying throat radius.}} \label{fig:einstradiusconcdown} \end{framed} \end{figure} In Figure \ref{fig:einstradiusconcdownaniso} we display the anisotropy, $\tilde{p}_{t}-\tilde{p}_{r}$, as a function of \emph{throat radius}. Note that larger throats require less anisotropy (i.e. are ``more isotropic''). \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip] {einst_anisotropy_c_down.eps} \end{center}\vspace{-0.1cm} \caption{\small{Anisotropy of throat region in Einstein case, $g_{tt}$ \emph{decreasing}, $x_{0}=1$, varying throat radius.}} \label{fig:einstradiusconcdownaniso} \end{framed} \end{figure} \clearpage In Figures \ref{fig:einstxoconcup}, \ref{fig:einstradiusconcup} and \ref{fig:einstradiusconcupaniso} we present a similar analysis as above, except that in these figures, $g_{tt}$ is \emph{concave-up}. Note the similarity of these results to the previous analysis. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_rho_c_up_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_econd1_c_up_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{Einstein_econd2_c_up_vary_xo.eps}} \end{tabular} \end{center} \caption{\small{Einstein gravity, $g_{tt}$ \emph{increasing}, throat radius=0.05, varying flare-out $x_{0}$.}} \label{fig:einstxoconcup} \end{framed} \end{figure} \vspace{-2.85cm} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{einst_rho_c_up.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{einst_econd1_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{einst_econd2_c_up.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{Einstein gravity, $g_{tt}$ \emph{increasing}, $x_{0}=1$, varying throat radius.}} \label{fig:einstradiusconcup} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.5cm} \includegraphics[width=60mm,height=45mm, clip] {einst_anisotropy_c_up.eps} \end{center}\vspace{-0.1cm} \caption{\small{Anisotropy of throat region in Einstein case, $g_{tt}$ \emph{increasing}, $x_{0}=1$, varying throat radius.}} \label{fig:einstradiusconcupaniso} \end{framed} \end{figure} \FloatBarrier \subsubsection{{\normalsize $n=2$}} We now study the $n=2$ contribution. Adding an $\alpha_{2}R^{2}$ term to the gravitational action is arguably the most popular supplement to the Einstein-Hilbert action (see \cite{ref:rsq1}-\cite{ref:rsqlast}). For example, this modification has been utilized to drive inflation purely from the gravitational sector (the Starobinsky inflationary theory \cite{ref:starob}) or to capture some low-order quantum corrections \cite{ref:birrellanddavies}. As mentioned earlier, since in the theory given by the action (\ref{eq:lag}) each term in the series contributes to the energy conditions separately, it is more instructive to look at this contribution on its own, to see whether it helps or hinders with respect to the energy conditions. For the anisotropy analysis, we add the Einstein term to it as well to see if the modification can make the system more or less isotropic compared to Einstein gravity alone. We consider zero and non-zero tidal force cases. \paragraph{{\small Zero tidal force:}} In the case of zero tidal force ($\Phi(x)=\mbox{const.}$) we can get a handle on the behavior of the energy conditions at the throat by performing a series expansion as was done in the Einstein case. The results are summarized as: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} &\tilde{\rho}=\frac{2\alpha_{2}}{Q_{0}^{4}} \left[1+4Q_{0}^{3}\qzero^{\prime\prime\prime\prime}-16Q_{0}^{3}(\qzero^{\prime\prime})^{3}+4Q_{0}^{2}(\qzero^{\prime\prime})^{2}\right] \nonumber \\ &\qquad\quad\quad +\frac{8\alpha_{2}}{Q_{0}^{2}} \left[Q_{0}\qzero^{\prime\prime\prime\prime\prime} -25Q_{0}(\qzero^{\prime\prime})^{2}\qzero^{\prime\prime\prime} + 3\qzero^{\prime\prime}\qzero^{\prime\prime\prime}\right]\,x + \mathcal{O}(x^{2})\,, \label{eq:squaredzerorho}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{r}=\frac{8\alpha_{2}}{Q_{0}^{2}}\left[Q_{0}\qzero^{\prime\prime\prime\prime}-4Q_{0}(\qzero^{\prime\prime})^{3}+2(\qzero^{\prime\prime})^{2}\right] \nonumber \\ &\qquad\quad\quad +\frac{8\alpha_{2}}{Q_{0}^{2}} \left[Q_{0}\qzero^{\prime\prime\prime\prime\prime} -25 Q_{0}(\qzero^{\prime\prime})^{2}\qzero^{\prime\prime\prime} + 3\qzero^{\prime\prime}\qzero^{\prime\prime\prime}\right] \,x + \mathcal{O}(x^{2})\,, \label{eq:squaredzeroecond1}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{t}=\frac{4\alpha_{2}}{Q_{0}^{4}} \left[1+2Q_{0}^{2}(\qzero^{\prime\prime})^{2} -3Q_{0}\qzero^{\prime\prime}\right] +\frac{12\alpha_{2}}{Q_{0}^{3}}\qzero^{\prime\prime\prime}\left(2Q_{0}\qzero^{\prime\prime}-1\right)\,x + \mathcal{O}(x^{2})\,.\label{eq:squaredzeroecond2} \end{align}} \end{subequations} For $\alpha_{2} > 0$, which is the more physical sector \cite{ref:faraonibook}, it can be seen that the above can all be made positive at the throat (for example, by considering models where $\qzero^{\prime\prime}=0$ and $\qzero^{\prime\prime\prime\prime}>0$). By analyticity in a non-zero neighborhood, in such scenarios there therefore must be a non-zero domain about the throat for which the functions are non-negative. Therefore, in this class of models, throats (and we stress again here that we are not considering asymptotics at infinity) \emph{are allowed} which respect energy conditions. This will remain true even when adding the Einstein term to the action, as in the Einstein case the near-throat violation of energy conditions may be made arbitrarily small \cite{ref:VKD}, \cite{ref:nandi} independently of the $\alpha_{2}$ parameter. \paragraph{{\small Non-zero tidal force:}} Unfortunately, for non-zero tidal force the analytic expressions with a general $Q(x)$ and $\Phi(x)$ are very long and complicated, even as a near throat expansion, and not very revealing. We must therefore specify these functions in order to perform numerical studies. For this purpose we choose the same function as in the Einstein case so that comparisons may be easily made. That is, we choose the profile $Q(x)$ as given in (\ref{eq:Qcosh}) since, to reiterate, this function possesses all the required properties to describe a throat. Recall that the parameter $A$ represents the radius of the throat and $x_{0}$ represents the degree of ``flare-out'' of the wormhole. We also use similar $\Phi(x)$ functions as presented in the Einstein case. As mentioned previously, we set $\alpha_{2}=1$, as all results here can be rescaled by whatever value of $\alpha_{2}$ one wishes to study, including negative values, which result in a reflection about the horizontal planes in the graphs. As mentioned above, it should be noted that $\alpha_{2}>0$ is the preferred model. The first set of results are summarized in Figure~\ref{fig:rsquaredalphaconcdown} (also please refer to figure captions for details). \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_rho_c_down_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{squared_econd1_c_down_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{squared_econd2_c_down_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{2}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, and varying $x_{0}$.}} \label{fig:rsquaredalphaconcdown} \end{framed} \end{figure} Note that like the zero tidal-force case, there is a large region of the parameter space where all energy conditions can be respected. In fact, for positive $\alpha_{2}$, a large flare-out (small $x_{0}$) actually helps the condition $\tilde{\rho}+\tilde{p}_{r}>0$ near the throat. Also of interest is the anisotropy. Namely, can the presence of the extra terms in $f(R)$ gravity lessen the amount of anisotropy required to support a throat, compared to Einstein gravity alone. To study this we present Figure~\ref{fig:rsquaredalphaconcdownaniso} where the anisotropy, defined here as $\tilde{p}_{t}-\tilde{p}_{r}$, is presented for the Lagrangian $R+\alpha_{2}R^{2}$ and is plotted in the vicinity of the minimum anisotropy curve. Note from this figure that anisotropy is a minimum for $\alpha_{2}<0$, and not for the Einstein case ($\alpha_{2}=0$), although the minimum curve is very close to $\alpha_{2}=0$. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip] {R_plus_R_sq_aniso_c_down_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Minimum anisotropy region: $R+\alpha_{2}R^{2}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{2}$.}} \label{fig:rsquaredalphaconcdownaniso} \end{framed} \end{figure} In Figure~\ref{fig:rsquaredAconcdown} we present another non-zero tidal force analysis (with the same $g_{tt}(x)$ as in the previous case) where, instead of allowing $x_{0}$ to vary, we vary the throat radius. This study is interesting as even in Einstein gravity the size of the throat affects the amount of energy condition violation. We find that, generally, smaller throat radius is more favorable for respecting energy conditions. (This result is reversed for $\alpha_{2}<0$.) \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.0cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{squared_rho_c_down_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{squared_econd1_c_down_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{squared_econd2_c_down_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{2}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{decreasing} and varying throat radius. For negative $\alpha_{2}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rsquaredAconcdown} \end{framed} \end{figure} Next we consider non-zero tidal force models where $g_{tt}(x)$ is increasing instead of decreasing. The figures below (figs.~\ref{fig:rsquaredalphaconcup} and \ref{fig:rsquaredalphaconcupaniso}) summarize the results for varying $x_{0}$ and $\alpha_{2}$. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{squared_rho_c_up_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{squared_econd1_c_up_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{squared_econd2_c_up_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{2}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, and varying $x_{0}$.}} \label{fig:rsquaredalphaconcup} \end{framed} \end{figure} Note that in Figure~\ref{fig:rsquaredalphaconcup} the graphs mimic the previous scenario with decreasing $g_{tt}(x)$, indicating insensitivity to the form of the lapse function. Therefore, in this scenario, again energy conditions can be respected for the more physical sector of positive $\alpha_{2}$. We also present the anisotropy in Figure~\ref{fig:rsquaredalphaconcupaniso} for the full Lagrangian of $R+\alpha_{2}R^{2}$. Anisotropy is again minimized for $\alpha_{2}\neq 0$ and in the negative sector. \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_sq_aniso_c_up_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Minimum anisotropy region: $R+\alpha_{2}R^{2}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{2}$.}} \label{fig:rsquaredalphaconcupaniso} \end{framed} \end{figure} Finally for this sub-section, we present the results for the increasing $g_{tt}(x)$ but where the throat radius is allowed to vary. These results are summarized in Figure~\ref{fig:rsquaredAconcup}. \clearpage \begin{figure}[H] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{squared_rho_c_up_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{squared_econd1_c_up_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{squared_econd2_c_up_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{2}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{increasing}, and varying throat radius. For negative $\alpha_{2}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rsquaredAconcup} \end{framed} \end{figure} \noindent Note that again smaller throat radius is more favorable for energy conditions than larger throat radius if $\alpha_{2}$ is positive. (The result is reversed for negative $\alpha_{2}$.) Parameters exist where energy conditions may again be satisfied, even when adding these results to the energy condition violating Einstein gravity terms. \FloatBarrier \subsubsection{{\normalsize $n=3$}} The $n=3$ scenarios are also of interest and it has been shown in a cosmological setting how this case can be transformed to a system consisting of Einstein gravity and a scalar field with a self-coupling proportional to $\alpha_{3}\phi^{4}$ \cite{ref:n3a}. \paragraph{{\small Zero tidal force:}} Again for the $\Phi(x)=\mbox{const.}$ scenario we present an analytic expansion about the throat, although, due to the complexity of the coefficients, we produce only the lowest order term. The results are: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} &\tilde{\rho}= \frac{4\alpha_{3}}{Q_{0}^{6}} \left[1- 24Q_{0}^{4}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime} + 6Q_{0}\qzero^{\prime\prime} -24 Q_{0}^{4}(\qzero^{\prime\prime\prime})^{2} +12 Q_{0}^{3}\qzero^{\prime\prime\prime\prime} -12Q_{0}^{2}(\qzero^{\prime\prime})^{2}\right. \nonumber \\ &\qquad\left. -56Q_{0}^{3}(\qzero^{\prime\prime})^{3} +96Q_{0}^{4}(\qzero^{\prime\prime})^{4}\right] + \mathcal{O}(x)\,, \label{eq:cubedzerorho}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{r}=\frac{24\alpha_{3}}{Q_{0}^{5}} \left[\qzero^{\prime\prime}-4Q_{0}^{3}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime} -4Q_{0}^{3}(\qzero^{\prime\prime\prime})^{2} +2Q_{0}^{2}\qzero^{\prime\prime\prime\prime} -12Q_{0}^{2}(\qzero^{\prime\prime})^{3} \right. \nonumber \\ &\qquad\quad\quad \left. +16 Q_{0}^{3}(\qzero^{\prime\prime})^{4} \right] + \mathcal{O}(x)\,, \label{eq:cubedzeroecond1}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{t}= \frac{12\alpha_{3}}{Q_{0}^{6}} \left[1-5Q_{0}\qzero^{\prime\prime} +8Q_{0}^{2}(\qzero^{\prime\prime})^{2} -4 Q_{0}^{3} (\qzero^{\prime\prime})^{3}\right] + \mathcal{O}(x)\,.\label{eq:cubedzeroecond2} \end{align}} \end{subequations} Here, in the case of positive $\alpha_{3}$, (\ref{eq:cubedzerorho}) - (\ref{eq:cubedzeroecond2}) can in principle be made positive. The simplest models obeying energy conditions would again be those where $\qzero^{\prime\prime}=0$ (which, due to the condition for a minimum, also requires $\qzero^{\prime\prime\prime}=0$) and $\qzero^{\prime\prime\prime\prime}>0$, as was the case for the corresponding situation in $n=2$. Note that if $\qzero^{\prime\prime}=0$, the contributions from Einstein gravity, from (\ref{eq:einstnearthroatrho}) - (\ref{eq:einstnearthroate2}), exactly at the throat are non-negative (although (\ref{eq:einstnearthroate1}) becomes negative in some neighborhood away from the throat), and hence with the addition of these $\alpha_{3}R^{3}$ contributions, $f(R)=R+\alpha_{3}R^{3}$ gravity can be made to obey energy conditions (and therefore, including the previous analysis, $f(R)=R+\alpha_{2}R^{2}+\alpha_{3}R^{3}$ gravity can also be made to obey energy conditions). \paragraph{{\small Non-zero tidal force:}} The analysis of the non-zero tidal forces proceeds in a similar order as the $n=2$ case. Due to the interest in keeping the paper of reasonable length, we present all the graphs (figs.~\ref{fig:rcubedalphaconcdown} - \ref{fig:rcubedAconcup}) for this case first, and then briefly comment on the results afterward in table~1. (Please refer to figure captions for details.) \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_rho_c_down_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_econd1_c_down_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_econd2_c_down_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{3}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05 and varying $x_{0}$.}} \label{fig:rcubedalphaconcdown} \end{framed} \end{figure} \vspace{-1.4cm} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_cb_aniso_c_down_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Minimum anisotropy region: $R+\alpha_{3}R^{3}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{3}$.}} \label{fig:rcubedalphaconcdownaniso} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_rho_c_down_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_econd1_c_down_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_econd2_c_down_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{3}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{decreasing}, and varying throat radius. For negative $\alpha_{3}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rcubedAconcdown} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_rho_c_up_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_econd1_c_up_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{cubed_econd2_c_up_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{3}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, and varying $x_{0}$.}} \label{fig:rcubedalphaconcup} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_cb_aniso_c_up_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Minimum anisotropy region: $R+\alpha_{3}R^{3}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{3}$.}} \label{fig:rcubedalphaconcupaniso} \end{framed} \end{figure} \vspace{4cm} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.0cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_rho_c_up_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_econd1_c_up_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{cubed_econd2_c_up_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{ $R^{3}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{increasing}, and varying throat radius. For negative $\alpha_{3}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rcubedAconcup} \end{framed} \end{figure} \FloatBarrier We summarize the results for $n=3$ in table~1. \begin{table}[!ht] \newcommand\Ta{\rule{0pt}{2.6ex}} \newcommand\Bb{\rule[-1.2ex]{0pt}{0pt}} \begin{center} \caption{\small{Summary of $\alpha_{3}R^{3}$ contributions.}\vspace{-0.2cm}} \begin{tabular}{|c|c|c|c|} \hline \Ta \parbox[t]{2.5cm}{Parameter studied}& \parbox[t]{3.8cm}{zero tidal-force} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-down} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-up} \\[0.5cm] \hline\hline \parbox[t]{2.5cm}{$x_{0}$} & \parbox[t]{3.8cm}{{\small Possible to make energy conditions positive.}}\Ta & \parbox[t]{3.8cm}{{\small Tendency to make positive contribution to all energy conditions for all range of $x_{0}$ studied near the throat for positive $\alpha_{3}$, and negative contribution for negative $\alpha_{3}$.}} & \parbox[t]{3.8cm}{{\small Tendency to make positive contribution to all energy conditions for positive $\alpha_{3}$, and negative contribution for negative $\alpha_{3}$. Shows insensitivity to form of $g_{tt}(x)$.}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Throat radius} & \parbox[t]{3.8cm}{{\small Possible to make energy conditions positive.}}\Ta & \parbox[t]{3.8cm}{{\small Energy conditions more positive for small throat radius (with $\alpha_{3}>0$. Reverse reported results for $\alpha_{3}<0$).}} & \parbox[t]{3.8cm}{{\small Energy conditions more positive for small throat radius (with $\alpha_{3}>0$. Reverse reported results for $\alpha_{3}<0$).}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Minimum anisotropy for $R+\alpha_{3}R^{3}$} &\parbox[t]{3.8cm}{{\small Not studied.}}\Ta & \parbox[t]{3.8cm}{{\small $\alpha_{3} < 0$ minimizes anisotropy.}} & \parbox[t]{3.8cm}{{\small $\alpha_{3} < 0$ minimizes anisotropy.}} \\[0.07cm] \hline \end{tabular} \end{center} \end{table} \FloatBarrier \subsubsection{{\normalsize $n=-1$}} \paragraph{{\small Zero tidal force:}} Again we begin with an analytic analysis of the zero tidal-force model. The results are summarized as: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} &\tilde{\rho}=\frac{\alpha_{-1}Q_{0}^{2}}{4\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{4}} \left[1-2Q_{0}\qzero^{\prime\prime} + 4Q_{0}^{3}\qzero^{\prime\prime\prime\prime} +24 Q_{0}^{4} (\qzero^{\prime\prime\prime})^2 +4Q_{0}^{2}(\qzero^{\prime\prime})^{2}\right. \nonumber \\ &\qquad\quad\quad \left. -24 Q_{0}^{3}(\qzero^{\prime\prime})^{3} -8Q_{0}^{4}\qzero^{\prime\prime\prime\prime}\qzero^{\prime\prime} +32Q_{0}^{4}(\qzero^{\prime\prime})^{4}\right] + \mathcal{O}(x)\,, \label{eq:minusonezerorho}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{r}=\frac{\alpha_{-1}Q_{0}^{3}}{2\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{4}} \left[3\qzero^{\prime\prime}+2Q_{0}^{2}\qzero^{\prime\prime\prime\prime} +12 Q_{0}^{3}(\qzero^{\prime\prime\prime})^{2} -8Q_{0}(\qzero^{\prime\prime})^{2} \right. \nonumber \\ &\qquad\quad\quad \left. -4Q_{0}^{2}(\qzero^{\prime\prime})^{3} -4Q_{0}^{3}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime} + 16 Q_{0}^{3}(\qzero^{\prime\prime})^{4} \right] + \mathcal{O}(x)\,, \label{eq:minusonezeroecond1}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{t}= \frac{\alpha_{-1}Q_{0}^{2}}{4\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{2}}\left[Q_{0}\qzero^{\prime\prime}-1\right] + \mathcal{O}(x)\,.\label{eq:minusonezeroecond2} \end{align}} \end{subequations} Here it can be seen that with inverse powers of $R$ a peculiar singularity occurs. The singularity occurs when $2QQ^{\prime\prime}=1$ and corresponds to the vanishing of the Ricci scalar at these points. This is not unexpected and it is not a curvature singularity, but does herald a problem with the equations of motion when $R=0$. One common way to attempt remedy this situation in studies of inverse-$R$ gravity is to postulate that the (gravitational) vacuum state of the theory is not Minkowski space-time, but is instead deSitter or anti-deSitter space-time \cite{ref:carroll}. However, when one is not far away from sources (as in the study here) there can be curves or surfaces in the space-time on which $R=0$ and these must be excluded. Hence the parameter space studied here will not include parameters near these pathologies, which we excise. Here the vanishing of $\qzero^{\prime\prime}$ does not yield an energy condition respecting throat, unlike in the positive $n$ cases. (Either $\tilde{\rho}+\tilde{p}_{t} <0$ for $\alpha_{-1}>0$ or else $\tilde{\rho}+\tilde{p}_{r}$ and $\tilde{\rho}<0$ for $\alpha_{-1}<0$.) Note that in principle it may be possible to respect energy conditions near the throat. One can say that if $Q_{0}\qzero^{\prime\prime} >1$ and furthermore if $\qzero^{\prime\prime\prime\prime}$ is large then it may be possible to respect energy conditions, but the situation is not obvious. Hence here the numerical results are needed, which we present for the more general non-zero tidal force scenarios. \paragraph{{\small Non-zero tidal force:}} If one is considering the region far from the throat, then $g_{tt}(x)$ should, as mentioned above, asymptote to deSitter or anti-deSitter space-time. Our choices for $g_{tt}(x)$ can accommodate this, but we are only interested in the near-throat region, so strictly speaking it is not a requirement. Again, due to the interest in keeping the length reasonable, we present all the graphs for this case first (figs.~\ref{fig:rminusonealphaconcdown} - \ref{fig:rminusoneAconcup}), and then briefly comment on the results afterward in a summary table (table~2). \vspace{-0.2cm} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0,5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_rho_c_down_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_econd1_c_down_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_econd2_c_down_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-1}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, and varying $x_{0}$.}} \label{fig:rminusonealphaconcdown} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_min_one_aniso_c_down_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Anisotropy of throat region: $R+\alpha_{-1}R^{-1}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{-1}$.}} \label{fig:rminusonealphaconcdownaniso} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_rho_c_down_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_econd1_c_down_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_econd2_c_down_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-1}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{decreasing}, and varying throat radius. For negative $\alpha_{-1}$ the graphs should be flipped around the plane $z=0$. The region around the singularity has been omitted.}} \label{fig:rminusoneAconcdown} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_rho_c_up_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_econd1_c_up_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_one_econd2_c_up_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-1}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, and varying $x_{0}$.}} \label{fig:rminusonealphaconcup} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_min_one_aniso_c_up_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Anisotropy of throat region: $R+\alpha_{-1}R^{-1}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{-1}$.}} \label{fig:rminusonealphaconcupaniso} \end{framed} \end{figure} \clearpage \begin{figure}[H] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_rho_c_up_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_econd1_c_up_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_one_econd2_c_up_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-1}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{increasing}, and varying throat radius. For negative $\alpha_{-1}$ the graphs should be flipped around the plane $z=0$. The region around the singularity has been omitted.}} \label{fig:rminusoneAconcup} \end{framed} \end{figure} \FloatBarrier We summarize the results for $n=-1$ in table 2. \begin{table}[!ht] \newcommand\Ta{\rule{0pt}{2.6ex}} \newcommand\Bb{\rule[-1.2ex]{0pt}{0pt}} \begin{center} \caption{\small{Summary of $\alpha_{-1}R^{-1}$ contributions.}\vspace{-0.2cm}} \begin{tabular}{|c|c|c|c|} \hline \Ta \parbox[t]{2.5cm}{Parameter studied}& \parbox[t]{3.8cm}{zero tidal-force} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-down} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-up} \\[0.5cm] \hline\hline \parbox[t]{2.5cm}{$x_{0}$} & \parbox[t]{3.8cm}{{\small Respecting energy conditions seems difficult at best.}}\Ta & \parbox[t]{3.8cm}{{\small Little sensitivity very near the throat on variations of $x_{0}$.}} & \parbox[t]{3.8cm}{{\small Little sensitivity very near the throat on variations of $x_{0}$.}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Throat radius} & \parbox[t]{3.8cm}{{\small Respecting energy conditions seems difficult at best.}}\Ta & \parbox[t]{3.8cm}{{\small Possible to respect energy conditions near singular region for negative $\alpha_{-1}$ for a small section of parameter space. (Only on one side of the singular region.)}} & \parbox[t]{3.8cm}{{\small Seems possible to make small positive contribution to all energy conditions (see general $n$ section) for very small region of parameter space.}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Minimum anisotropy for $R+\alpha_{-1}R^{-1}$} &\parbox[t]{3.8cm}{{\small Not studied.}}\Ta & \parbox[t]{3.8cm}{{\small Magnitude of anisotropy larger than corresponding pure Einstein case regardless of $\alpha_{-1}$.}} & \parbox[t]{3.8cm}{{\small Magnitude of anisotropy larger than corresponding pure Einstein case regardless of $\alpha_{-1}$.}} \\[0.07cm] \hline \end{tabular} \end{center} \end{table} \FloatBarrier \subsubsection{{\normalsize $n=-2$}} Finally we present here the contribution from $\alpha_{-2}R^{-2}$. Much of this scenario mimics the $\alpha_{-1}R^{-1}$ contribution and hence we also summarize the results in a table (table~3) after all the graphs (figs.~\ref{fig:rminustwoalphaconcdown} - \ref{fig:rminustwoAconcup}). \paragraph{{\small Zero tidal force:}} At the throat, the relevant quantities possess the following values: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} &\tilde{\rho}= \frac{\alpha_{-2}Q_{0}^{4}}{8\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{5}}\left[56Q_{0}^{3}(\qzero^{\prime\prime})^{3} -6Q_{0}\qzero^{\prime\prime} -12Q_{0}^{3}\qzero^{\prime\prime\prime\prime} -96Q_{0}^{4}(\qzero^{\prime\prime\prime})^{2} -96Q_{0}^{4}(\qzero^{\prime\prime})^{4} \right. \nonumber \\ &\qquad\quad\quad \left. +12Q_{0}^{2}(\qzero^{\prime\prime})^{2} +24Q_{0}^{4}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime}-1\right] +\mathcal{O}(x)\,, \label{eq:minustwozerorho}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{r}=\frac{\alpha_{-2}Q_{0}^{5}}{2\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{5}}\left[8Q_{0}^{2}(\qzero^{\prime\prime})^{3} - 4\qzero^{\prime\prime} -3Q_{0}^{2}\qzero^{\prime\prime\prime\prime} -24Q_{0}^{3}(\qzero^{\prime\prime\prime})^{2} \right. \nonumber \\ &\qquad\quad\quad \left. -24Q_{0}^{3}(\qzero^{\prime\prime})^{4}+10Q_{0}(\qzero^{\prime\prime})^{2} +6Q_{0}^{3}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime}\right] + \mathcal{O}(x)\,, \label{eq:minustwozeroecond1}\\[0.2cm] &\tilde{\rho}+\tilde{p}_{t}= \frac{\alpha_{-2}Q_{0}^{4}}{4\left(2Q_{0}\qzero^{\prime\prime}-1\right)^{3}}\left[1-Q_{0}\qzero^{\prime\prime}\right] + \mathcal{O}(x)\,.\label{eq:minustwozeroecond2} \end{align}} \end{subequations} Again it can be noted that the equations of motion become singular where the Ricci scalar vanishes. As expected, the degree of the singularity has increased in comparison to the $n=-1$ case. The situation regarding energy conditions here, like for $n=-1$, does not seem promising in the zero tidal force regime. \paragraph{{\small Non-zero tidal force:}} The following figs.~\ref{fig:rminustwoalphaconcdown} - \ref{fig:rminustwoAconcup} display the results of the non-zero tidal force scenarios. \FloatBarrier \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_gtt_c_down.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_rho_c_down_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_econd1_c_down_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_econd2_c_down_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-2}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, and varying $x_{0}$. For negative $\alpha_{-2}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rminustwoalphaconcdown} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.22cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_min_two_aniso_c_down_vary_nu.eps} \end{center} \vspace{-0.1cm} \caption{\small{Anisotropy of throat region: $R+\alpha_{-2}R^{-2}$ contribution, $g_{tt}$ \emph{decreasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{-2}$.}} \label{fig:rminustwoalphaconcdownaniso}\vspace{-0.1cm} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_rho_c_down_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_econd1_c_down_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_econd2_c_down_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-2}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{decreasing}, and varying throat radius. For negative $\alpha_{-2}$ the graphs should be flipped around the plane $z=0$. The region around the singularity has been omitted.}} \label{fig:rminustwoAconcdown} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{cc} \subfloat[$g_{tt}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_gtt_c_up.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_rho_c_up_vary_xo.eps}} \\ \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_econd1_c_up_vary_xo.eps}}&\hspace{0.5cm} \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=45mm,height=45mm,clip]{minus_two_econd2_c_up_vary_xo.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-2}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, and varying $x_{0}$. For negative $\alpha_{-2}$ the graphs should be flipped around the plane $z=0$.}} \label{fig:rminustwoalphaconcup} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{0.0cm} \includegraphics[width=60mm,height=45mm, clip]{R_plus_R_min_two_aniso_c_up_vary_nu.eps} \end{center}\vspace{-0.1cm} \caption{\small{Anisotropy of throat region: $R+\alpha_{-2}R^{-2}$ contribution, $g_{tt}$ \emph{increasing}, throat radius=0.05, $x_{0}=1$, and varying $\alpha_{-2}$.}} \label{fig:rminustwoalphaconcupaniso} \end{framed} \end{figure} \clearpage \begin{figure}[H] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_rho_c_up_vary_radius.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_econd1_c_up_vary_radius.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{minus_two_econd2_c_up_vary_radius.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{-2}$ contribution for $x_{0}=1$, $g_{tt}$ \emph{increasing}, varying throat radius. For negative $\alpha_{-2}$ the graphs should be flipped around the plane $z=0$. The region around the singularity has been omitted.}} \label{fig:rminustwoAconcup} \end{framed} \end{figure} \FloatBarrier We summarize the results for $n=-2$ in table 3. \begin{table}[!ht] \newcommand\Ta{\rule{0pt}{2.6ex}} \newcommand\Bb{\rule[-1.2ex]{0pt}{0pt}} \begin{center} \caption{\small{Summary of $\alpha_{-2}R^{-2}$ contributions.}\vspace{-0.2cm}} \begin{tabular}{|c|c|c|c|} \hline \Ta \parbox[t]{2.5cm}{Parameter studied}& \parbox[t]{3.8cm}{zero tidal-force} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-down} & \parbox[t]{3.8cm}{$g_{tt}(x)$ concave-up} \\[0.5cm] \hline\hline \parbox[t]{2.5cm}{$x_{0}$} & \parbox[t]{3.8cm}{{\small Respecting energy conditions seems difficult at best.}}\Ta & \parbox[t]{3.8cm}{{\small Energy condition respecting model possible near throat. For most of positive parameter space, small values are found, which may be dwarfed by negative value of Einstein term).}} & \parbox[t]{3.8cm}{{\small Energy condition respecting model not found near the throat.}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Throat radius} & \parbox[t]{3.8cm}{{\small Respecting energy conditions seems difficult at best.}}\Ta & \parbox[t]{3.8cm}{{\small Seems possible to make positive contribution to energy conditions for very small parameter space near the singular region for $\alpha_{-2}<0$. (Only on one side of the singular region.)}} & \parbox[t]{3.8cm}{{\small Does not seem possible to respect energy conditions.}} \\[0.07cm] \hline \parbox[t]{2.5cm}{Minimum anisotropy for $R+\alpha_{-2}R^{-2}$} &\parbox[t]{3.8cm}{{\small Not studied}} \Ta & \parbox[t]{3.8cm}{{\small Magnitude of anisotropy larger than corresponding pure Einstein case regardless of $\alpha_{-2}$.}} & \parbox[t]{3.8cm}{{\small Magnitude of anisotropy larger than corresponding pure Einstein case regardless of $\alpha_{-2}$.}} \\[0.07cm] \hline \end{tabular} \end{center} \end{table} \FloatBarrier \subsubsection{{\normalsize General $n$}} Finally, for general $n$ we first quote the throat values in the zero tidal-force case. The quantities are as follows: \begin{subequations}\renewcommand{\theequation}{\theparentequation \roman{equation}} {\allowdisplaybreaks\begin{align} \tilde{\rho}=& 2^{n-1}\frac{\alpha_{n}}{Q_{0}^{2n}}\left(1-2Q_{0}\qzero^{\prime\prime}\right)^{n-3} \left[12n^{2}Q_{0}^{4} (\qzero^{\prime\prime\prime})^{2} +16n^{2}Q_{0}^{4}(\qzero^{\prime\prime})^{4} +2n^{2}Q_{0}^{3}\qzero^{\prime\prime\prime\prime} +2n^{2}Q_{0}\qzero^{\prime\prime} \right. \nonumber \\[0.1cm] &-4n^{2}Q_{0}^{2}(\qzero^{\prime\prime})^2 -8Q_{0}^{3}(\qzero^{\prime\prime})^{3} -6Q_{0}\qzero^{\prime\prime} +1 -8n^{2}Q_{0}^{3}(\qzero^{\prime\prime})^{3} +4nQ_{0}^{4}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime}\nonumber \\[0.1cm] &-4n^{3}Q_{0}^{4}(\qzero^{\prime\prime\prime})^{2} + 8nQ_{0}^{3}(\qzero^{\prime\prime})^{3} -2nQ_{0}\qzero^{\prime\prime} -16nQ_{0}^{4} (\qzero^{\prime\prime})^{4} +12Q_{0}^{2}(\qzero^{\prime\prime})^{2} \nonumber \\[0.1cm] &+4nQ_{0}^{2}(\qzero^{\prime\prime})^{2}-8nQ_{0}^{4}(\qzero^{\prime\prime\prime})^{2} -2nQ_{0}^{3}\qzero^{\prime\prime\prime\prime} -4n^{2}Q_{0}^{4}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime} \left.\right] +\mathcal{O}(x)\,,\label{eq:gennrho}\\[0.2cm] \tilde{\rho}+\tilde{p}_{r}=& 2^{n} \frac{n\alpha_{n}}{Q_{0}^{2n-1}} \left(1-2Q_{0}\qzero^{\prime\prime}\right)^{n-3} \left[2Q_{0}^{3}\qzero^{\prime\prime\prime\prime}\qzero^{\prime\prime} +6nQ_{0}^{3}(\qzero^{\prime\prime\prime})^{2} +8nQ_{0}^{3}(\qzero^{\prime\prime})^{4}\right. \nonumber \\[0.1cm] & +nQ_{0}^{2} \qzero^{\prime\prime\prime\prime} -2\qzero^{\prime\prime} +6Q_{0}(\qzero^{\prime\prime})^{2} - 2n^{2}Q_{0}^{3}(\qzero^{\prime\prime\prime})^{2} -4nQ_{0}^{2}(\qzero^{\prime\prime})^{3}+n\qzero^{\prime\prime} -8Q_{0}^{3}(\qzero^{\prime\prime})^{4} \nonumber \\[0.1cm] &-2nQ_{0}(\qzero^{\prime\prime})^{2} -4Q_{0}^{3}(\qzero^{\prime\prime\prime})^{2} -Q_{0}^{2}\qzero^{\prime\prime\prime\prime} -2nQ_{0}^{3}\qzero^{\prime\prime}\qzero^{\prime\prime\prime\prime}\left.\right] +\mathcal{O}(x)\,,\label{eq:gennecond1}\\[0.2cm] \tilde{\rho}+\tilde{p}_{t}=&2^{n-1}\frac{n\alpha_{n}}{Q_{0}^{2n}} \left[1-Q_{0}\qzero^{\prime\prime}\right]\left[1-2Q_{0}\qzero^{\prime\prime}\right]^{n-1}+\mathcal{O}(x) \,.\label{eq:gennecond2} \end{align}} \end{subequations} These conditions represent the contribution to the energy conditions for a particular value of (arbitrary) $n$. The general energy conditions would then constitute the sum of these conditions summed over the $n$ values which contribute to the gravitational action. For the non-zero tidal force scenarios we present the following results in order to show the trends as $n$ increases. The graphs (\ref{fig:rn_gtt_dec_xo})-(\ref{fig:rn_gtt_inc_A}) are plots of the energy conditions \emph{exactly at the throat} $x=0$ for $g_{tt}(x)$ concave-down and $g_{tt}(x)$ concave-up respectively. Note that from the analyticity of $Q(x)$ in a neighborhood of the throat, if a quantity is positive at the throat then there exists a non-zero neighborhood of the throat where this quantity is positive, and hence the particular energy condition can be respected. In all the following graphs, only results where the energy conditions are positive are shown. (Refer to figure captions for details.) \vspace{-0.0cm} \begin{figure}[H] \begin{framed} \begin{center} \vspace{-0.2cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_xo_rho.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_xo_econd1.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_xo_econd2.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{n}$ contribution at $x=0$ for $A=0.05$ and $\alpha_{n}=1$, $g_{tt}$ \emph{decreasing}, and varying $x_{0}$. In the far-left graph the plots correspond to $n=-3\rightarrow 5$ (bottom to top). The middle graph corresponds $n=-3,\,-2,\,-1,\,2,\,3,\,4,\,5$ (bottom to top). The right graph corresponds to parameters $n=1,\,2,\,3,\,4,\,5$ (bottom to top). Values of $n$ not displayed are negative (note though that these would be positive if the sign of $\alpha_{n}$ was negative).}} \label{fig:rn_gtt_dec_xo} \end{framed} \end{figure} \vspace{-1.25cm} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_A_rho.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_A_econd1.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_down_vary_A_econd2.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{n}$ contribution at $x=0$ for $x_{0}=1$ and $\alpha_{n}=1$, $g_{tt}$ \emph{decreasing}, and varying throat radius. In the far-left graph, the graphs correspond to $n=-3\rightarrow +5$ from bottom to top. In the middle graph, the plots correspond to $n=-3,\,-2,\,-1,\,3,\,4,\,5$ from bottom to top (the others being either negative or too close to zero to show on the scale). In the far-right plot, the curves approaching the vertical-axis correspond to $n=1\rightarrow 5$ from bottom to top. The curve that does not approach the vertical-axis corresponds to $n=-2$.}} \label{fig:rn_gtt_dec_A} \end{framed} \end{figure} \begin{figure}[!ht] \begin{framed} \begin{center} \vspace{-0.0cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_xo_rho.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_xo_econd1.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_xo_econd2.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{n}$ contribution at $x=0$ for $A=0.05$ and $\alpha_{n}=1$, $g_{tt}$ \emph{increasing}, and varying $x_{0}$. In the far-left graph the plots correspond to $n=-3\rightarrow 5$ (bottom to top). The middle graph corresponds $n=0,\,-3,\,-2,\,-1,\,2,\,3,\,4,\,5$ from bottom to top. (Note that $n=0$ should give a contribution of zero, as it corresponds to the cosmological constant only scenario. The line at $10^{-16}\approx 0$ is a numerical artefact.) The right graph corresponds to parameters $n=0,\,1,\,2,\,3,\,4,\,5$ (bottom to top, again with the $n=0$ plot being a numerical artefact.) Values of $n$ not displayed are negative (note though that these would be positive if the sign of $\alpha_{n}$ was negative).}} \label{fig:rn_gtt_inc_xo} \end{framed} \end{figure} \begin{figure}[H] \begin{framed} \begin{center} \vspace{-0.5cm} \begin{tabular}{ccc} \subfloat[$\tilde{\rho}(x)$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_A_rho.eps}} & \subfloat[$\tilde{\rho}+\tilde{p}_{r}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_A_econd1.eps}}& \subfloat[$\tilde{\rho}+\tilde{p}_{t}$]{\includegraphics[width=42mm,height=42mm,clip]{R_to_n_c_up_vary_A_econd2.eps}} \end{tabular} \end{center} \vspace{-0.1cm} \caption{\small{$R^{n}$ contribution at $x=0$ for $x_{0}=1$ and $\alpha_{n}=1$, $g_{tt}$ \emph{increasing}, and varying throat radius. In the far-left graph, the graphs correspond to $n=-3\rightarrow +5$ from bottom to top. In the middle graph, the plots correspond to $n=-3,\,-2,\,-1,\,2,\,3,\,4,\,5$ from bottom to top (the others being either negative or too close to zero to show on the scale). In the far-right plot, the curves intersecting the vertical-axis correspond to $n=1\rightarrow 5$ from bottom to top. The curves that do not approach the vertical-axis corresponds to $n=-1$ (bottom) and $n=-3$ (top).}} \label{fig:rn_gtt_inc_A} \end{framed} \end{figure} \FloatBarrier \section{{\small CONCLUDING REMARKS}} We have studied the existence of wormhole throats in modified gravity theories with gravitational actions of the form $S=\int_{M^{4}} {\underset{{\hspace{-0.05cm}\vspace{-0.1cm}n}}{\sum}} \alpha_{n}\,R^{n}\sqrt{g}\,d^{4}x$. This action includes the $n=1$ contribution of Einstein gravity. Of specific interest to our studies are energy condition violation and the degree of anisotropy required to support the throat, since in Einstein gravity alone the existence of throats implies both violation of energy conditions as well as the presence of anisotropy. We studied various terms contributing to the action separately, in order to discern which terms aid and which terms hinder the energy conditions. It would also be worthwhile to consider all terms together, but this would be more computationally involved. We have studied both the zero and non-zero tidal force solutions. {In general, we find that the parameter space for energy condition respecting solutions is much larger for positive $n$ than negative $n$, and that larger $n$ increases the energy conditions. This is due to the fact that $f(R)$ gravity has an equivalent description as a scalar-tensor theory, where the ``scalar field'' component may violate energy conditions. The positive $n$ sector also allows one to minimize the anisotropy required to support a wormhole throat, whereas the negative $n$ sector tends to make the required anisotropy much larger when compared to Einstein gravity.} It may be then that if gravitation is governed by such a Lagrangian, or one which is expandable as such, throats may exist which obey the energy conditions violated by Einstein gravity alone and require little or perhaps no anisotropy. In light of this, it may be possible that if the true gravitational action contains higher positive powers of $R$, then throats in space-time may be more likely than with no such augmentation or with negative powers. Since it is unknown (though perhaps unlikely) whether or not the topology of space-time can change, we cannot say whether throats would be more likely in the early universe (where higher powers may contribute to inflationary expansion) than in the late universe (where inverse powers may contribute to late-time acceleration) (see, for example \cite{ref:no3} for an analysis of modified gravity accommodating early and late-time accelerated expansion). However, it is interesting to speculate on this. \section*{{\small ACKNOWLEDGMENTS}} A.D. is grateful for the kind hospitality of the University of Zagreb, where some of this work was carried out. D.H. is partially supported by the Croatian Ministry of Science under the project No. 036-0982930-3144. This work was also partially supported by CompStar, a Research Networking Programme of the European Science Foundation. \linespread{0.6} \bibliographystyle{unsrt}
\section{Introduction} \label{Introduction} A subset of Interplanetary Coronal Mass Ejections (ICMEs) has simple flux rope-like magnetic fields, which are characterized by enhanced magnetic fields that rotate slowly through a large angle. Such events named magnetic clouds or MCs \citep{Burlaga1981, KleinBurlaga1982, Gosling1990} have received considerable attention because the magnetic field configuration is described by the force-free model \citep[e.g.,][and references therein]{Lepping1990, Osherovich1997}. When the MC is moving faster than the surrounding solar wind (SW), the plasma and magnetic field typically accumulate in front of it, forming a disturbed region so-called sheath region. The rapid decrease in the total SW pressure with solar distance is the main driver of the flux-rope radial expansion \citep{DemoulinDasso2009}. However, if a MC is moving slower than the surrounding solar wind \citep{KleinBurlaga1982, Zhang1988, Burlaga1988}, the sheath region is difficult to be detected and the magnetic cloud could go unnoticed by the spacecraft. The shock wave produced by the ICMEs is spatially greater than the magnetic cloud, so a spacecraft can detect only the shock wave sometimes \citep{Schwenn2006}. Near $1~\mathrm{AU}$ MCs have enormous radial sizes $(0.28~\mathrm{AU})$ with an average duration of $27~\mathrm{h}$, an average peak of the magnetic field strength of $18~\mathrm{nT}$ and an average solar wind speed of $420~\mathrm{Km/s}$ \citep{KleinBurlaga1982, Lepping2000}. As has been noted by many authors \citep[e.g.,][]{Zwickl1983, Richardson1995}, sometimes individual signatures may not be detected in all ICMEs, because they are not present or even there are data gaps. A practical criterion to detect magnetic clouds is to calculate the plasma beta, defined as the ratio of the thermal and magnetic pressure ($\beta = 2 \mu_0 p / B^2$), to find values significantly lower than one. For this calculation, magnetic field and plasma observations at spacecraft are required. Many times the temperature and density data on spacecraft have many gaps during periods in which the plasma instruments could be saturated as a result of intense particle fluxes (for example, Bastille Day in the ACE spacecraft). If this condition occurs, it makes impossible to calculate the plasma beta, but it is still possible to detect the magnetic cloud using magnetometers data \citep[e.g.][]{Huttunen2005, Nieves-Chichilla2005}. In this paper, we propose a nonlinear time series analysis as an auxiliary tool to identify magnetic clouds using only the magnetic field data. The result of this calculation can be interpreted as an index related to some dynamical features that identifies the MC. \subsection*{MC events} MC events were investigated by \citet{Huttunen2005} using the minimum variance analysis (MVA) \citep{SonnerupCargill1967} to determine if they have a flux-rope structure. They identified $73$ MCs in the period $(1997-2003)$ and $7$ cloud candidate observed by the ACE and WIND spacecraft in solar cycle $23$. The axis of a MC can have any orientation with respect to the ecliptic plane \citep{BothmerSchewenn1994, BothmerSchewenn1998}. This axis orientation, identified by $\phi_C$, the azimuthal direction in the ecliptic, and $\theta_C$, the inclination relative to the ecliptic, is also calculated by the MVA \citep{BothmerSchewenn1998}. To classify MCs, eight flux rope categories are often used: \begin{itemize}[--] \item Bipolar MCs (low inclination and flux rope-type: SWN, SEN, NES, NWS), $\theta_C \leq 45\,^{\circ}$ and \item Unipolar MCs (high inclination, and flux rope-type: WNE, ESW, ENW, WSE), $\theta_C > 45\,^{\circ}$, \end{itemize} where the meanings are S for south, N North, W west and E east. Those authors have also included seven ''cloud candidate`` events for which either the fitting with MVA was not successful (e.g. the eigenvalue ratio $<2$ or the directional change less than $30\,^{\circ}$) or there was large values of beta throughout the event. In that study the criterion to define a MC were based on the smoothness of the rotation in the magnetic field direction confined to one plane. Additionally, they required that a MC must have the average values of plasma beta less than $0.5$, the maximum value of the magnetic field at least $8~\mathrm{nT}$, and the duration at least $6~\mathrm{h}$. With the last two criteria, they aimed to remove the ambiguity of identifying the small and weak MCs. All selected events were investigated by analyzing $1~\mathrm{h}$ magnetic field data with the MVA \citep{SonnerupCargill1967}, where MCs are identified from the smooth rotation of the magnetic field vector in the plane of the maximum variance \citep{KleinBurlaga1982}. \subsection*{Nonlinear analysis} The concept of entropy is fundamental for the study of several branches of the physics like statistical mechanics and thermodynamics. Loosely interpreted, entropy is a thermodynamic quantity which describes or quantifies the amount of disorder in a physical system. Therefore we can generalize this concept to characterize or quantify the amount of information stored in more general probability distributions of a time series. This is in part what an information theory is concerned with. The aforementioned theory was developed during the $1940s$ and $1950s$ with the main contributions given by \citet{Shannon1949, Renyi1959, Kolmogorov1958}; and \citet{Sinai1959}. As the solar wind plasma generates complex fluctuations in a spacecraft-detected signals, those fluctuations can be investigated with techniques adopted from the nonlinear dynamics theory \citep{Remya2010662}. In principle, it is possible to study the MCs by analyzing the time series of interplanetary magnetic field (IMF). So, the main goal in this work is to find some nonlinear tool that helps someone identifying the MCs in the solar wind. \citet{Ojeda2005} studied time series that represent physical parameters of the solar wind (interplanetary magnetic field and velocity components in the plasma flow) recorded by the WIND spacecraft with a time resolution of $1~\mathrm{min}$. In that work they studied $20$ MCs, $17$ plasmoids (events not identified as MC), and $20$ time series of equivalent time duration of quiet solar wind. The IMF $B_z$ and solar wind $V_x$ components in a time interval of $48~\mathrm{hour}$ before each MC were studied. Under MC conditions, a feature was identified that the IMF $B_z$ has the tendency to present lower spatio-temporal entropy$\footnote{name given by Eugene Kononov's Visual Recurrence Analysis (VRA) software, not to be confused with spatio-temporal entropy image (STEI) \citep{Ma2001}}$ (STE) values than the $B_z$ in other cases, such as in plasmoids and during quiet solar wind. It seems to be a very contrasting aspect to be analyzed. Thus in this work a more detailed study of the STE in MCs is carried out. The behavior of the STE in time series of the IMF components, measured by the ACE spacecraft with a time resolution of $16~\mathrm{s}$, is explored. \section{Methodology and Dataset} \label{Dataset and Methodology} In this work, the purpose is to present a tool using the STE in a method that can identify simply and quickly the occurrence of a MC at about $1~\mathrm{AU}$. The main advantage of the approach proposed is that it takes into account only the interplanetary magnetic field data to identify the cloud. Nevertheless, this STE method does not solve the problem of identifying the boundaries of the clouds. To delimit the boundaries, if this is the case, the methodology presented in the work of \citet{Huttunen2005} is available to be used . In order to test and validate our identification methodology, MCs identified by other authors will be also used. Additionally new MCs will be identified by this tool unravelling some features hidden in those time series of data. Our methodology uses a strictly mathematical criterion from the nonlinear dynamics techniques to identify MCs in the solar wind. This methodology is indeed a novelty. \subsection{The recurrence plots in Visual Recurrence Analysis software} It is presented a summary of the ideas expressed in the Eugene Kononov's Visual Recurrence Analysis (VRA) software$\footnote{\label{VRA}VRA v4.7 http://nonlinear.110mb.com/vra/}$ about recurrence plots (RPs). In order to present the ideas, some figures are used to guide the description. Figure~\ref{fig:RPSine} shows a RP for a simple sine wave, using the data file just included in VRA software. In it organized patterns of color characteristics are shown for the periodical signal. In order to allow a comparative view, a RP of white noise is shown in Figure~\ref{fig:RPNoise}, with the data file also included. With a different result, an uniform distribution of color characteristics is noticed for the random signal. \begin{figure}[htp] \centering \includegraphics[width=0.4\textwidth]{RPSine.eps} \caption{Recurrence plot of simple sine wave, using a data file included with VRA software. In the RP organized patterns of color characteristics are shown for a periodic signal (A color version is available in the electronic version).} \label{fig:RPSine} \end{figure} \begin{figure}[htp] \centering \includegraphics[width=0.4\textwidth]{RPNoise.eps} \caption{Recurrence plot of White Noise, using data file included with VRA software. In the RP an uniform distribution of color characteristics is shown for a random signal(A color version is available in the electronic version).} \label{fig:RPNoise} \end{figure} The RP is a relatively recent technique for the qualitative assessment of time series \citep{Eckmann1987}, A technique that allow someone detects graphically hidden patterns and structural changes in data or see similarities in patterns across the time series under analysis. The fundamental assumption underlying the idea is that an observable time series (a sequence of observations) is the manifestation of some dynamic process. It has been proved mathematically that one can recreate a topologically equivalent picture of the original multidimensional system behavior by using the time series of a single observable variable \citep{Takens1981}. The basic idea is that the effect of all the other (unobserved) variables is already reflected in the series of the observed output. Furthermore, the rules that govern the behavior of the original system can be recovered from its output. In the RPs a one-dimensional time series from a data file is expanded into a higher-dimensional space, in which the dynamic of the underlying generator takes place. This is done by a technique called ''delayed coordinate embedding'', which recreates a phase space portrait of the dynamical system under study from a single (scalar) time series. To expand a one-dimensional signal into an M-dimensional phase space, one substitutes each observation in the original signal X(t) with vector $(y(i) = \{x(i), x(i - d), x(i - 2d)$, $\cdots , x(i - (m-1)d\})$, where $i$ is the time index, $m$ is the embedding dimension, $d$ is the time delay. As a result, we have a series of vectors $Y = {y(1), y(2), y(3), \cdots, y(N-(m-1)d)}$, where $N$ is the length of the original series. With such reconstruction it is possible to reproduce the original system states at each time where we have an observation of that system output. Each unknown state $Z(t)$ at time $t$ is approximated by a vector of delayed coordinates $Y(t) = { x(t), x(t - d), x(t - 2d), \cdots , x(t - (m-1)d }$. After the Euclidean distances between all vectors are calculated, they are mapped to colors from the pre-defined color map and are displayed as colored pixels in their corresponding places (see Figure~\ref{fig:RPSine}, for example). The RP is a graphical representation of a correlation integral. The important distinction (and an advantage) is that the RP, unlike the correlation integrals, preserve the temporal dependence in the time series, in addition to the spatial dependence. In RPs, if the underlying signal is truly random and has no structure, the distribution of colors is uniform and does not have any identifiable patterns (see Figure~\ref{fig:RPNoise}, for example). There is some determinism in the signal generator, which can be detected by some distinctive color distribution. For example, hot colors (yellow, red, and orange) can be associated with small distances between the vectors, while others colors (blue, black) may be used to show large distances. In this printed work colors are noticed as a grey pattern (from white to black). Therefore one can visualize and study the motion of the system trajectories and infer some characteristics of the dynamical system that generated the time series. Also, the length of diagonal line segments of the same color on the RP brings an idea about the signal predictability. But, RP is mostly a qualitative tool. For random signals, the uniform (even) distribution of colors over the entire RP is expected. The more deterministic the signal, the more structured the RP. So for the purpose of comparison in Figure~\ref{fig:RPSine} the RP of a strictly periodic signal can see and in Figure~\ref{fig:RPNoise} the RP of the white noise time series. \subsection{The entropy concepts in recurrence plot: a review} The RP is a visual tool for the investigation of temporal recurrence in phase space \citep{Takens1981}. In the literature there are some methods to calculate the entropy in the RP and the phase space \citep{Takens1981}. The method (STE) presented by Eugene Kononov's VRA software is the only one provided satisfactory results when it is applied to time series of the IMF. A brief review on the entropy concepts is presented here. The STE entropy was used to measures the image ``structuredness'' in a bidimensional representation, \textit{i.e.}, both in ``space'' and time domains. It was implemented in VRA software to quantify the order in RPs. In physical terms, this quantity compares the distribution of distances between all pairs of vectors in the reconstructed state space with that of distances between different orbits evolving in time. The result is normalized and presented as a percentage of ``maximum'' entropy (randomness). When the entropy has a value of $100\%$ it means the absence of any structure whatsoever (uniform distribution of colors, pure randomness, seen in Figure~\ref{fig:RPNoise}). On the other hand, $0\%$ of entropy implies ``perfect'' structure (distinct color patterns, perfect ``structuredness'' and predictability, seen in Figure~\ref{fig:RPSine}). Recurrence is the most important feature of chaotic systems \citep{Eckmann1987}. The popularity of RPs lies in the fact that their structures are visually appealing, and that they allow the investigation of high dimensional dynamics by means of a simple two-dimensional plot \citep{Facchini2009}. For a better understanding and quantification of the recurrences, \cite{Webber1994} have proposed a set of quantification measures, which are mainly based on the statistical distribution of the line structures in the RP. Recurrence quantification analysis (RQA) is a nonlinear technique used to quantify the information supplied by a RP \citep{Webber1994, Zbilut1992}. Recurrence variables are calculated from the upper triangular area of the recurrence plot, excluding the central diagonal, because the plot is symmetrical about the main diagonal. The RQA can be used as a tool for the exploration of bifurcation phenomena and dynamics changes also in nonstationary and short time series. The entropy (ENT) is one of the recurrence variables of the RQA method. It is the Shannon information entropy for the distribution probability of the diagonal lines. That is: \begin{equation} ENT=- \sum_{k=L_{min},p(k)\neq 0}^{L_{max}} p(k)log_2 (p(k)), \end{equation} where $L_{min}$ is the minimum length of diagonal lines in RP and \begin{equation} p(k)=\frac{\mbox{number of diagonal lines of length k in RP}}{\mbox{number of diagonal lines in RP}}. \end{equation} The ENT can be calculated using the VRA software; but it should not be confused with the STE. \cite{Little2007} developed a Recurrence Period Density Entropy (RPDE) method, first it requires the embedding of a time series in phase space, which, according to Taken's embedding theorems, can be carried out by forming time-delayed vectors for each value $x_n$ in the time series. Then, around each point in the embedded phase space, a recurrence neighbourhood of radius $\epsilon$ is created. All recurrences into this neighbourhood are tracked, and the time interval T between recurrences is recorded in a histogram. This histogram is normalized to create an estimate of the recurrence period density function p(T). The normalized entropy of this density is the RIDE value $H_{norm}$ \citep{Little2007}. \begin{equation} H{norm}=-(ln T_{max})^{-1}\sum_{t=1}^{T_{max}}p(t)ln p(t). \end{equation} The RPDE value is a scalar in the range zero to one. For purely periodic signals, $H_{norm}=0$ (STE=$0\%$) whereas for purely uniform white noise, $H_{norm}=1$ (STE=$100\%$). However, estimates obtained with this technique (RPDE) are different from those obtained with the STE. \cite{Dasan2002} report an analysis, using the tools of nonlinear dynamics and chaos theory, of the fluctuations in the stress determined from simulations of shear flow of Stokesian suspensions. They also computed the STE using VRA for the stress. The calculated values of the STE for the shear and normal stresses were nearly zero, showing perfect structure in the data. They observed definite structure in the phase-space plot of the stress components \citep{Dasan2002}. They cited the works of \cite{Peacock1983, Carr1998}. \cite{Peacock1983} presented a two-dimensional analogue of Kolmogorov-Smirnov test, useful for analysing the distribution of data in two dimensions, as is the RP. \cite{Carr1998} investigated the fluctuation phenomena in plasmas that often necessitates the analysis of spatio-temporal signals. It was shown how such signals can be analyzed using the biorthogonal decomposition, which splits them into orthogonal spatial and temporal modes. Several parameters allow one to quantify the weight distribution in the biorthogonal decomposition. The total energy of spatio-temporal signal $u(t,j)$ is found to be equal to the sum of the eigenvalues, $\alpha_m$: \begin{equation} E(u)=\sum_{m=1}^N \alpha_m^2. \end{equation} They can define the relative energy of the $m^{th}$ structure as \begin{equation} E_m(u)=\frac{\alpha_m^2}{E(u)}, \end{equation} and the entropy of the spatio-temporal signal $u(t,j)$ is defined as \begin{equation} H(u)=-\frac{1}{log N}\sum_{m=1}^{N}E_m(u)log E_m(u). \end{equation} It describes how the energy is distributed across the $N_s$ significant structures. Signal whose energy is concentrated in a single structure such that $N_s=1$ will have very low entropy $H(u)=0$, our $H(u)=1$ if the energy is distributed equally among the $N_s$ significant structures. The results will be presented in this paper shows the usefulness of STE implemented by Eugene Kononov's software. Other ways to calculate the entropy are not useful as an auxiliary tool to identify the occurrence of MCs. \subsection{Test of stationary time series using STE} \label{Trend in time series} From an intuitive point of view, a time series is said to be stationary if there is no systematic change in mean (no trend) and no systematic change in variance and if strictly periodic variations have been removed \citep{Chartfield}. Trend estimation is a statistical technique to aid in the interpretation of data \citep{Chartfield}. When a time series related to measurements of a process are treated, trend estimation can be used to make and justify statements about tendencies in the data. Given a set of data and the desire to produce some kind of "model" of those data (model, in this case, meaning a function fitted through the data), there are a variety of functions that can be chosen for the fit. However, if there is no prior understanding of the data, then the simplest function to fit is a straight line and thus this is the "default". Once it has been decided to fit a straight line, there are various ways to do so, but the most usual choice is a least-squares fit, equivalent to minimization of the $L2$ norm. If there is no global trend in time series the angle (``trend angle``) between the straight line and the positive $x$ axis must be zero, else the angle is not zero. The aim of this subsection is to calculate the STE versus the trend angle. We deal with time series with data file included in VRA to calculate the STE of each one versus the trend angle. Figure~\ref{fig:TendenciaLorentz} shows a time series plot of Lorenz data file included in VRA. We gave trends to the series through rotations about the origin. After that, we calculated the STE of each temporal series. The results were included in row~$2$ presented in Table~\ref{table:trend_series}, in which three time series with data file included in VRA version~$4.7$ are shown. The effects of the trends in every series related with STE values are quantified. \begin{figure}[htp] \centering \begin{tabular}{c} \includegraphics[width=0.7\textwidth]{Lorenz.eps} \end{tabular}\\[.7cm] \caption{Time series plot of Lorenz data file included in VRA (case with $\theta=0$). Series rotated about origin ($\theta=-0.01\mbox{ rad }$, $\theta=0.01\mbox{ rad }$, $\theta=0.0175\mbox{ rad }$) and the three resulting series also plotted. After that, we calculate the STE of each series.} \label{fig:TendenciaLorentz} \end{figure} We follow the same idea to cause a trend in time series for another cases, Sine and White Noise data file also included in VRA. The results were included in rows~$3$ and~$4$ of Table~\ref{table:trend_series}. For periodic time series (sine data file) the value of STE is always zero independently from the increasing trend. If the trend in the time series increases, the STE value decreases (see row~$2 \mbox{ and } 4$ in Table~\ref{table:trend_series}). \begin{table}[htp] \caption{STE values related to trends for three time series with data file included in VRA $4.7$.} \centering \begin{tabular}{l l l l l} \hline Series&$\theta=0\mbox{ rad}$&$\theta=-0.01\mbox{ rad}$&$\theta=0.01\mbox{ rad}$&$\theta=0.0175\mbox{ rad}$\\ \hline STE(Lorenz)&$73\%$&$30\%$&$29\%$&$0\%$\\ STE(Sine)&$0\%$&$0\%$&$0\%$&$0\%$\\ STE(White Noise)&$80\%$&$34\%$&$34\%$&$3\%$\\ \hline \end{tabular} \label{table:trend_series} \end{table} Also, the first difference applied in a time series is often enough to convert series with a trend into a stationary time series. The first-order differences of time series values $x_1,x_2,x_3,\cdots,x_N$ are given by a new series $y_1,y_2,\cdots,y_{N-1}$, where $y_{N-1}=x_N-x_{N-1}$. The operation $y_t=x_t-x_{t-1}=\nabla x_t$ is called the first difference and $\nabla$ is the difference operator. Sometimes the first differencing is not enough to remove the trend in mean. We then try further differencing. For example, the second-order difference is given by: \begin{equation} \nabla^2y_t=\nabla x_t-\nabla x_{t-1}=x_t-2x_{t-1}+x_{t-2} \end{equation} Our interest is to study variations in the STE values when a first-order differences are applied on stationary time series. In the previously studied time series (Lorenz, Sine and White Noise data file included in VRA), we calculate the first-order differences. After that, we calculated the STE of each time series and compared the results with the STE of the original series, which are shown in Table~\ref{table:Differ_series}. The STE values are similar in both of them, \textit{i.e.}, for transformed (first-order differences) and untransformed time series. Thus, when the time series has no trend, the non-linear delicate structures are not destroyed. \begin{table}[htp] \caption{STE values related to the first-order differences in time series.} \centering \begin{tabular}{l l l} \hline Series& Untransformed & first-order differences\\ \hline STE(Lorenz)&$73\%$&$75\%$\\ STE(Sine)&$0\%$&$0\%$\\ STE(White Noise)&$80\%$&$82\%$\\ \hline \end{tabular} \label{table:Differ_series} \end{table} The entropy value is small and may tend to zero in any time series with trend. The STE value for nonstationary signals is not meaningful. If there is a trend in the time series, you might want to consider removing it by differencing the original time series before calculating the STE. Keep in mind, however, that taking the first differences may destroy the delicate nonlinear structure in the time series (if there is any). Following this idea, sometimes we differentiate the series of the interplanetary magnetic field before calculating the entropy. This is done on a trial basis after calculating the entropy of the original series. The reason is to remove the trend and calculate the true value of entropy of the series, the results are shown in Figure \ref{Fig:Anisotropias}~(c) and will be discussed later. The relationship between trend and STE values in magnetic clouds is discussed in this work. \subsection{ACE spacecraft data} The plasma particle detected by ACE arrived to the Earth after approximately $30~\mathrm{min}$. The Magnetic Field Experiment (MAG) on board ACE consists of twin vector fluxgate magnetometers to measure interplanetary magnetic field. The data$\footnote{http://www.srl.caltech.edu/ACE/ASC/level2/lvl2DATA\_MAG.html}$ contains time averages of the magnetic field over time periods $1s$, $16s$, $4min$, hourly, daily and $27$ days (1 Bartels rotation). Magnetic field vectors are given in the $\mbox{RTN}$, $\mbox{GSE}$, and $\mbox{GSM}$ coordinate systems. The Solar Wind Electron, Proton, and Alpha Monitor (SWEPAM) measures the solar wind plasma electron and ion fluxes (rates of particle flow) as functions of direction and energy. The data$\footnote{http://www.srl.caltech.edu/ACE/ASC/level2/lvl2DATA\_SWEPAM.html}$ contains time averages of solar wind parameters over time periods $64s$ (ion data only), $128s$ (electron data only), hourly (all data), daily (all data) and 27 days (1 Bartels rotation) (all data). In this paper we use data from the magnetic field components with the time resolution of $16s$ in $\mbox{GSM}$ and $\mbox{GSE}$ coordinate systems, because we are interested in studying time series of magnetic clouds from $1000$ to $5500$ data records. We think that some local informations are lost in the short time series analysis. We cannot calculate the STE in the VRA software for time series with more than $5500$ data points (explained later in Subsection~\ref{subs:Table}). If we need to detect the boundaries of a magnetic cloud with MVA method then we use data with~$1h$ time resolution (similar to \citet{Huttunen2005}) for MAG and SWEPAM experiments respectively. \begin{table}[htp] \caption{Solar Wind data studied (from \citet{Huttunen2005}). } \centering \begin{tabular}{l l l l l l l l} \hline No. & Year & Shock & UT & MC, start & UT & MC, stop & UT\\ \hline 01 & 1998 & 06 January & 13:19 & 07 January & 03:00 & 08 January & 09:00\\ 02 & & 03 February & 13:09 & 04 February & 05:00 & 05 February & 14:00\\ 03 & & 04 March & 11:03 & 04 March & 15:00 & 05 March & 21:00\\ 04 & & 01 May & 21:11 & 02 May & 12:00 & 03 May & 17:00\\ 05 & & 13 June & 18:25 & 14 June & 02:00 & 14 June & 24:00\\ 06 & & 19 August & 05:30 & 20 August & 08:00 & 21 August & 18:00\\ 07 & & 24 September & 23:15 & 25 September & 08:00 & 26 September & 12:00\\ 08 & & 18 October & 19:00 & 19 October & 04:00 & 20 October & 06:00\\ 09 & & 08 November & 04:20 & 08 November & 23:00 & 10 November & 01:00\\ 10 & & 13 November & 00:53 & 13 November & 04:00 & 14 November & 06:00\\ 11 & 1999 & 18 February & 02:08 & 18 February & 14:00 & 19 February & 11:00\\ 12 & & 16 April & 10:47 & 16 April & 20:00 & 17 April & 18:00\\ 13 & & 08 August & 17:45 & 09 August & 10:00 & 10 August & 14:00\\ 14 & 2000 & 11 February & 23:23 & 12 February & 12:00 & 12 February & 24:00\\ 15 & & 20 February & 20:57 & 21 February & 14:00 & 22 February & 12:00\\ 16 & & 11 July & 11:22 & 11 July & 23:00 & 13 July & 02:00\\ 17 & & 13 July & 09:11 & 13 July & 15:00 & 13 July & 24:00\\ 18 & & 15 July & 14:18 & 15 July & 19:00 & 16 July & 12:00\\ 19 & & 28 July & 05:53 & 28 July & 18:00 & 29 July & 10:00\\ 20 & & 10 August & 04:07 & 10 August & 20:00 & 11 August & 08:00\\ 21 & & 11 August & 18:19 & 12 August & 05:00 & 13 August & 02:00\\ 22 & & 17 September & 17:00 & 17 September & 23:00 & 18 September & 14:00\\ 23 & & 02 October & 23:58 & 03 October & 15:00 & 04 October & 14:00\\ 24 & & 12 October & 21:36 & 13 October & 17:00 & 14 October & 13:00\\ 25 & & 28 October & 09:01 & 28 October & 24:00 & 29 October & 23:00\\ 26 & & 06 November & 09:08 & 06 November & 22:00 & 07 November & 15:00\\ 27 & 2001 & 19 March & 10:12 & 19 March & 22:00 & 21 March & 23:00\\ 28 & & 27 March & 17:02 & 27 March & 22:00 & 28 March & 05:00\\ 29 & & 11 April & 15:18 & 12 April & 10:00 & 13 April & 06:00\\ 30 & & 21 April & 15:06 & 21 April & 23:00 & 22 April & 24:00\\ 31 & & 28 April & 04:31 & 28 April & 24:00 & 29 April & 13:00\\ 32 & & 27 May & 14:17 & 28 May & 11:00 & 29 May & 06:00\\ 33 & & 31 October & 12:53 & 31 October & 22:00 & 02 November & 04:00\\ 34 & 2002 & 23 March & 10:53 & 24 March & 10:00 & 25 March & 12:00\\ 35 & & 17 April & 10:20 & 17 April & 24:00 & 19 April & 01:00\\ 36 & & 18 May & 19:44 & 19 May & 04:00 & 19 May & 22:00\\ 37 & & 01 August & 23:10 & 02 August & 06:00 & 02 August & 22:00\\ 38 & & 30 September & 07:55 & 30 September & 23:00 & 01 October & 15:00\\ 39 & 2003 & 20 March & 04:20 & 20 March & 13:00 & 20 March & 22:00\\ 40 & & 17 August & 13:41 & 18 August & 06:00 & 19 August & 11:00\\ 41 & & 20 November & 07:27 & 20 November & 11:00 & 21 November & 01:00\\ \hline \end{tabular} \label{table:ListaMCs} \end{table} \section{Results and Discussion} \label{Results and Discussion} In the next subsection, we compare STE values between time series corresponding to MCs and sheaths. This comparison is performed only for IMF data, which were data of the MCs detected by \citet{Huttunen2005}. From the $73$ MCs identified by those researchers we have dealt with just $41$ MCs: just for clouds preceded by shock waves. In the other subsections, we have studied time intervals of the solar wind, designated here as solar wind windows, with the purpose of validating a method for detecting MCs. Over all the windows the STE variations were studied. We named the previous methodology of "windowed spatio-temporal entropy". The criteria to select a specific solar wind window are presented in the begin of each subsection. \subsection{STE calculated for $41$ magnetic clouds between $1998 - 2003$} \label{subs:Table} We processed ACE data, from March $1998$ to December $2003$, of the three magnetic field components $(B_x,\: B_y,\: B_z)$ in GSM, with a time resolution of $16s$. We obtained the recurrence plots for $246$ time series $(3*(41~MCs~+~41~Shts) = 246)$. We have calculated the STE value for each temporal series with embedding dimension and time delay equal to one respectively. It may be noticed that STE value changes for different embedding parameters. For example, for Lorenz attractor (Lorenz data file included in VRA), STE is near its minimum when the correct embedding is used (dimension = 3, time delay = 16 for that particular data file). This and other results suggest that STE can be used to determine the optimal embedding parameters. We selected the same embedding and time delay, equal to one, to maintain equivalence in the calculation of entropy among all series. Table~\ref{table:ListaMCs} presents the events in chronological order. The STE values for the 41 MC events, presented in Table \ref{table:ListaMCs}, are shown in Figure~\ref{Fig:Anisotropias}~(a). At the top, the STE of the three IMF components $B_x$, $B_y$ and $B_z$, plotted respectively as the $"\circ"$, $"+"$ and $"\times"$, for the MCs. At the bottom, the same for the sheath regions. In Figure~\ref{Fig:Anisotropias}~(b), the histogram of STE obtained for the $B_z$ component taking into account the MCs and the sheath regions are shown. In Figure~\ref{Fig:Anisotropias}~(a), the STE values of the $246$ time series (each separated by cloud and sheath respectively) were plotted in chronological order (as appeared in column~$1$ in Table~\ref{table:ListaMCs}) earlier for MCs and bellow for sheath regions. The STE values were calculated with embedding dimension and time delay equal to one respectively. Some MCs do not have values close to zero entropy in the three components simultaneously. Then, it is possible to find components with perfect structuredness (low STE) and absence of structure (high STE) in the same magnetic cloud. If we compare the STE values between the same components for the sheath and the cloud regions (for example, $B_x$-sheath $(STE=56\%)$ with STE $B_x$-cloud $(STE=0\%)$ in the event $1$), we detect that for $5/41\: (3/41)$ cases in the $B_x$ $(B_y, B_z)$ component(s) the STE in clouds is bigger than the one in sheath respectively. We found a decrease in entropy in the clouds present in all IMF components. Someone can notice that a clear tendency of the MCs events to present STE with lower values, close to zero, as observed in \citet{Ojeda2005} and extended in this work, which is a new feature added to the usual features \citep{Burlaga1981} established to the MCs. The STE was zero in $20/41,\: 21/41,\: 25/41$ MCs to $B_x,\: B_y,\: B_z$ respectively. The three components had $STE=0\%$ value in $17/41$ MCs and $1/41$ sheath at the same time. The sheath region with $STE=0\%$ corresponds to event number $06$ in Table~\ref{table:ListaMCs}. The presence of another cloud in the sheath region will be examined later (subsection~\ref{MC20August}). \begin{figure}[htp] \centering \includegraphics[width=1.0\textwidth]{Fig4EFig5.eps}\\[0.2cm] \caption{(a)~The STE values for the 41 MC events between 1998-2003 presented in Table \ref{table:ListaMCs}. At the top, the STE for $B_x$, $B_y$ and $B_z$, the three IMF components, plotted respectively as the $"\circ"$, $"+"$ and $"\times"$. At bottom, the same as to above but for the sheath regions. (b)~The histogram of STE obtained from (a) for the $B_z$ component taking into account the cloud and the sheath regions. (c)~It is similar to (a), but with the trend removed through the first-order differences in time series. (d)~It is similar to (b), The histogram of STE obtained from (c) for the $B_z$ component taking into account the cloud and the sheath regions.} \label{Fig:Anisotropias} \end{figure} Figure~\ref{Fig:Anisotropias}~(b) shows a histogram of STE derived from (a) for the $B_z$ component of clouds (in grey) and sheaths. We have $37/41$ of MCs with STE less than $40\%$ and $37/41$ of sheaths with STE greater than $40\%$ respectively. We did some tests with the time series to explain the results above. First, if the Gaussian noise is removed from the signal and the STE calculated, the STE value tend to decrease in less than $~5\%$ from its initial value. Second, when a trend is removed of the series through a rotation (with the angle of slope line of best fit) the STE varies, but still the three components had $STE=0\%$ value in $17/41$ MCs and $1/41$ sheath at the same time in time series with more than $5500$ data points (VRA software limitation). These results are shown in Figure~\ref{fig:STEDimensionDR_Original} and will be discussed later. Third, by removing the trend through the first order difference in time series (see Figure~\ref{Fig:Anisotropias}~(c)) we still have MCs with $STE=0\%$. Figure~\ref{Fig:Anisotropias}~(c) is similar to panel~(a), but we have eliminated the trend through the first order difference in time series. In this case most of the STE values in the sheath increased to $\sim90\%$ in the three components. Figure~\ref{Fig:Anisotropias}~(d) shows a histogram of STE derived from (c) for the $B_z$ component of clouds and sheaths. If we eliminate the trend in the series then the entropy increases. Then, there is not a nonlinear structure in the time series, the signal is nondeterministic. Also the STE increases in the clouds, but there are still MCs with zero STE. We find a limitation to interpret this case, the length of the series. The calculation of the STE with the VRA software, version~$4.7$, can not be made for time series with a size bigger than $5500$ points because the STE results always in zero. It seems to be a limitation of the software by some reason not explained in its tutorial. \begin{figure}[htp] \centering \includegraphics[width=0.5\textwidth]{STEDimensionDR_Original.eps} \caption{At middle, STE values versus time series length for $B_z$ component for the $41$~MCs, where the $"\circ"$ and $"+"$ symbols corresponds to original and transformed (remove the Gaussian noise and trend through a rotation) time series respectively. At bottom STE values versus time series length for the $41$~sheaths. At top the number of events versus intervals of the time series length. The histogram helps to identify overlapping points of MC in the middle panel. The STE is zero to the right of the vertical line due to a software limitation.} \label{fig:STEDimensionDR_Original} \end{figure} In Figure~\ref{fig:STEDimensionDR_Original} at bottom, we have plotted the STE versus the length of the time series for the $B_z$ components for a sheath region. The $"\circ"$ and $"+"$ symbols correspond respectively to the original and transformed (by removing the Gaussian noise and ``trend'' through a rotation) time series. Panel at middle is similar to one at bottom but for the MC region. The numbers to the right of the vertical line show the total of events ($17$) that each point represent. At top, the histogram helps to identify overlapping points of MC in the middle panel. The STE is zero to the right of the vertical line due to a software limitation. At middle panel, we have MCs ($"\circ"$ symbols or original) with zero STE values to the left of the vertical line. When we remove the Gaussian noise and trend through a rotation the STE values ($"+"$ symbols or transformed) increase and are different from zero. In Figure~\ref{fig:STEDimensionDR_Original}, related to the top and middle plots, it is shown $STE=0\%$ in $\:17/41$ (events $1-4$, $6-10$, $13$, $16$, $27$, $30$, $33-35$, $40$ in Table~\ref{table:ListaMCs}) in the MC with more than $5500$~points, in the other $9/41$ MCs (events $5$, $12$, $21$, $23-25$, $31$, $32$, $37$ in Table~\ref{table:ListaMCs}) the STE value leaves the zero value when the time series were transformed by removing the Gaussian noise and ``trend'' through a rotation. So far, we have been seeking the causes of the large amount of MCs with $STE=0\%$. During a MC, the magnetic field strength is higher than the average, the magnetic field direction rotates smoothly through a large angle, then the periods with MCs present more trend in the magnetic behavior than the periods of sheaths or "quiet" solar wind. The trend is the principal cause of the lower STE values in MC. Also, the MCs are more structured than sheath and "quiet" solar wind \citep{Ojeda2005}. Then it establishes a new feature to be considered in analyses. \subsection{The Windowed Spatio-Temporal Entropy} \label{subs:Window} From the $41$~MCs studied the STE decreases inside the MCs in relation to the sheath. In \citet{Ojeda2005} we found MCs with lower STE values than the "quiet" solar wind and other events in the interplanetary medium not identified as MCs. Then we have decided to analyze solar wind intervals of $10$~days where we already knew about the existence of MCs, as the events reported in Table \ref{table:ListaMCs}. As a criterion we have selected data windows of $2500$ records moving forward as $200$ records at head, until the end of the time series was reached. For every data segment selected the STE was calculated, allowing to analyze the evolution of the STE at every $200*16s$ in time evolution for every $2500$ records in the time series. We select $2500$ points because this interval represents an interval of $11.11~\mathrm{h}$, and the MCs have a smooth rotation of the magnetic field vector in the order of $1$ day, where the field reaches a peak and decreases. With a temporal window size of $11.11~\mathrm{h}$ and a resolution of $16~\mathrm{s}$ it is possible to cover the entire range of trend in the most of MCs with dimensions larger than $24~\mathrm{h}$. The STE values are calculated every $0.89~\mathrm{h}$ (STE time resolution adopted) and representing $\sim 8\%$ of the size of each temporal window. So the variation of the values of STE between two continuous windows $(11.11~\mathrm{h})$ must also be in the order of $\sim \pm8\%$, a sufficient condition for our initial purpose. We will be able to study in detail the calculation of the STE by temporal windows for six solar wind intervals following in the studies ($4$ MCs represented in Table \ref{table:ListaMCs} and other $2$ cases). \subsubsection{07-January-1998 MC event} We decided to study in more details the STE variations taking into account some days before and after the event number $1$ in the Table~\ref{table:ListaMCs}. The time series of the $B_x,\: B_y$ and $B_z$ IMF components of MC has zero STE value (see, Figure~\ref{Fig:Anisotropias}~(a), the first event). We find in Figure~\ref{fig:STEDimensionDR_Original} that the time series are composed of 6753 data points and, consequently, we have zero entropy when the STE was calculated using the VRA software. Based on the features analyzed in the beginning of this section, our hypothesis is to find zero STE in the period of the magnetic cloud to identify it. Then, we have selected a time window for $03-12$ January, $1998$ and take IMF three components in the Geocentric Solar Magnetospheric (GSM) coordinate system recorded by the ACE spacecraft with time resolution of $16s$. We selected data windows of $2500$ records moving forward as $200$ records to head, until the end of the time series. The numbers of windows result in $258$ and we calculate the STE of each one. In Figure~\ref{Fig:STEJaneladaJan1998}, we show the values of STE versus time for the time series of IMF $B_x,\: B_y$ and $B_z$ components in the solar wind. Additionally, we investigated if ACE detected some other event during these ten days. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzJan1998.eps}\\ \caption{Values of spatio-temporal entropy on January $03-12$, $1998$, as a function of the time for times series of IMF $B_x,\: B_y$ and $B_z$ components in the solar wind. The thick curve represents the Entropy Index (EI) calculated over the analyzed period. The shock, the start and end of the MC are represented by three vertical dotted lines.} \label{Fig:STEJaneladaJan1998} \end{figure} \citet{CaneRichardson2003} summarized the occurrence of interplanetary coronal mass ejections (ICMEs) in the near-Earth solar wind during $1996-2002$, corresponding to the increasing and maximum phases of solar cycle $23$. In particular, they give a detailed list of such events based on in situ observations. They reported two ICME in this interval, $07$ January $01:00$ to $08$ January $22:00$ and $09$ January $07:00$ to $10$ January $08:00$. Then we have one MC, one interplanetary disturbance that was not classified as MC, and a "quiet" solar wind period. Some features related to the ICME classified as MC in January $07$ were reported by \citet{CaneRichardson2003}: for example $V_{ICME} = 400 km/s$ is the mean solar wind speed in the ICME; $V_{max} = 410 km/s$ is the maximum solar wind speed in the post disturbance region; $B= 16 nT$ is the mean field strength; $V_T = 480 km/s$ is the transit speed. The MC was also reported by \citet{Huttunen2005} but with different size (see Table. \ref{table:ListaMCs}). We used the dates reported by \citet{Huttunen2005} to represent the shock, the start and end of the MC. These dates were represented with three vertical dotted lines in Figure~\ref{Fig:STEJaneladaJan1998}. This unipolar MC has a flux-rope type ENW and the observed angular variation of magnetic field is left-handed \citep{Huttunen2005}. The MVA method gives the eigenvalue ratio $\lambda_2/\lambda_3= 48$, the angle between the first and the last magnetic field vectors $\chi=160\,^{\circ}$, and the orientation of the axis $(\phi_C,\theta_C)=(21\,^{\circ},52\,^{\circ})$. The $B_z$ component was northward almost during the whole passage of the MC, then a magnetic storm ($Dst_{min}=-77 nT$) is caused by the sheath of heated and compressed solar wind plasma piled up in front of CME ejecta \citep{Tsurutani1988}. Figure~\ref{Fig:STEJaneladaJan1998} shows the behavior of STE in ten days of solar wind. The $B_y,\: B_z$ components have a zero STE value only during the passage of the MC, approximately in the first half of it. The second minimum value of $STE = 10\%$ corresponds to $B_z$ component in January, $09$. At this date $(09\: \mbox{Jan}\: 07:00 - 10\: \mbox{Jan}\: 08:00)$ \citet{CaneRichardson2003} detected one ICME with $V_{ICME} = 450 Km/s$, $V_{max} = 500 Km/s$, $B= 6 nT$. This result seems to be very interesting because in ten days of solar wind data two magnetic components had zero value only within a MC. \citet{Huttunen2005} first performed a visual inspection of the data to find the candidate MCs. This is always the first step in any work aiming at studying magnetic clouds. Then, the calculation of STE could be used as a mathematical tool to help finding the candidate MCs and being a new feature of many MCs. To generalize this idea it is necessary to study other MCs and explain physically two questions: (1) Why STE is zero only somewhere inside the MC? (2) Why don't all the magnetic components have zero STE inside the MC? These questions will be answered later. As not all the magnetic components have zero entropy at the same time, we created a standardization index (called by us Entropy Index (EI)) to allow joining the results of the three variable $(B_x, B_y, B_z)$ based on the physical process related to the MC geometry. The index is the result of multiplying the values of entropy of the three variables at the same time $t$ normalized by $10^4$: \begin{equation} EI= \prod_{i=1}^{3}{\frac{STE}{10^4}[\%]} \end{equation} The thick curve in Figure~\ref{Fig:STEJaneladaJan1998} is the representation of the EI calculated over the analyzed period. We could think of a similarity with the Dst index, which is a geomagnetic index which monitors the world wide magnetic storm level. Negative $Dst$ values indicate that a magnetic storm is in progress, the more negative Dst, the more intense the magnetic storm. But the entropy index value can only decrease to zero somewhere inside the MC. The MCs have simple flux rope-like magnetic fields, characterized by enhanced magnetic fields that rotate slowly through a large angle. Then, the time series of IMF components have a trend and a more ordered dynamic behavior and a higher degree of correlation between its temporary neighbors \citep{Ojeda2005}. The above mentioned behavior is found only in the magnetic structures of the MCs, a necessary condition for the zero entropy index value. Conclusion, using the entropy index we find the MC occurrences. But not the frontiers. \subsubsection{20-August-1998 MC event} \label{MC20August} We take this temporal windows of ten days because we already knew that there was an MC near the center, as reported in Table~\ref{table:ListaMCs} number $6$ event. Also, this is the only event that has zero STE in the sheath. Figure~\ref{Fig:STEJaneladaAug1998} is similar to Figure~\ref{Fig:STEJaneladaJan1998} but in the period $15-25$ August $1998$. At this date $(19\: \mbox{Aug}\: 18:47-21\: \mbox{Aug}\: 20:00,\: 1998)$, \citet{CaneRichardson2003} detected one ICME classified as MC with $V_{ICME} = 300 Km/s$, $V_{max} = 340 Km/s$, $B= 14 nT$. The three numbered $(1,2,3)$ vertical dotted lines in Figure~\ref{Fig:STEJaneladaAug1998} represent the shock $(1)$, the start $(2)$ and end $(3)$ of the MC2 reported by \citet{Huttunen2005} (see event number $6$ Table \ref{table:ListaMCs}). The MVA method to MC2 gives $\lambda_2/\lambda_3= 30$, $\chi=177\,^{\circ}$ and $(\phi_C,\theta_C)=(113\,^{\circ},-16\,^{\circ})$. This bipolar MC has a flux-rope type SWN and the observed angular variation of the magnetic field is right-handed \citep{Huttunen2005}. The observed SN-type MC had $B_z<0$ for about eleven hours, with values of less than $-10nT$. In the sheath it was observed $B_z<0$ for more than seven hours and values close to $B_z=-11nT$. The $Dst$ index shows two geomagnetic storms, the first with $Dst_{min}= -42 nT$ caused by the sheath, the second with $Dst_{min}= -67 nT$ caused by the MC. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzAug1998.eps}\\ \caption{Values of Spatio-Temporal Entropy on $15-25$ August, $1998$ as a function of the time for times series of IMF $B_x,\: B_y$ and $B_z$ components in the solar wind. The thick curve represents the Entropy Index (EI) calculated over the analyzed period. We identify the MC1. EI is zero after the cloud due to the adjustment in the time window for calculating the entropy and that this cloud is small.} \label{Fig:STEJaneladaAug1998} \end{figure} Figure~\ref{Fig:STEJaneladaAug1998} shows that the $B_x,\: B_y$ components have a zero STE value during the sheath of the MC, approximately at the end half of it. After that, approximately in the center of the MC, the three components $(B_x,\: B_y,\: B_z)$ have zero STE values. The EI is zero at the end of the sheath and within the MC. We think that maybe there are two magnetic clouds, one after the other. It is very difficult to detect the two MCs when a visual inspection of the data is performed and then apply the method of minimum variation \citep[see][]{BothmerSchewenn1998}. Someone could make a mistake in the interpretation of the cloud considering it as part of the sheath of the other MC. To solve the above hypothesis, we used the same method (MVA) explained in \citet{Huttunen2005}, with success. \begin{figure}[htp] \begin{center} $\begin{array}{c@{\hspace{1in}}c} \multicolumn{1}{l}{\mbox{\bf (a)}} & \multicolumn{1}{l}{\mbox{\bf (b)}} \\ [0.1 cm] \includegraphics[width=0.41\textwidth]{MaxVarAug20MC1.eps} & \includegraphics[width=0.41\textwidth]{MinVarAug20MC1.eps}\\ [0.1cm] \end{array}$ \end{center} \caption{The rotation of the magnetic field vector in the plane of maximum variance and in the plane of minimum variance for the bipolar MC observed by ACE on 19 August at 10:00 UT to 20 August at 00:00 UT 1998. In the Figure~\ref{Fig:STEJaneladaAug1998} shown this MC (MC1), the EI help to identify the MC.} \label{Fig:MinVarAug1998MC1} \end{figure} On $19$ August $1998$, at $02:00$ UT the parameters were: proton density $N_p=3.57 cm^{-3}$, the proton temperature $T_p=7.31*10^4~\mathrm{K}$, the ratio of $alphas/protons=4.5*10^{-3}$, the proton speed $V_p= 333.9~\mathrm{km/s}$, magnetic field magnitude $B=2.4~\mathrm{nT}$ and plasma beta $\beta=1.6$. One hour later some parameters had a change to $N_p=3.98~\mathrm{cm^{-3}}$, $T_p=7.1*10^4~\mathrm{K}$, $alphas/protons = 6.1*10^{-3}$, $V_p= 331.1~\mathrm{km/s}$, $B=0.57~\mathrm{nT}$ and $\beta=29.8$. It is very difficult to identify a shock, but the plasma beta increased abruptly and we believe it was related to the arrival of an event. Two hours later, at $05:00$ UT, the plasma beta decreased to $\beta=0.58$, after that there were eleven hours with gaps. The plasma beta minimum value was $\beta=7.8*10^{-2}$ at $21:00$ UT where there was a maximum magnetic field magnitude of $B=13.3~\mathrm{nT}$. We used the magnetic field rotation confined to one plane, the plane of maximum variance $(B_x^*B_y^*)$ (see Figure~\ref{Fig:MinVarAug1998MC1}), to find the boundaries of the cloud. Finally, we identified the bipolar MC1 observed by ACE on $19$ August at $10:00$ UT to $20$ August at $00:00$ UT (see Figure~\ref{Fig:MinVarAug1998MC1}). This MC has a flux-rope type SEN and the observed angular variation of the magnetic field is left-handed. The MVA method gives to MC1 the eigenvalue ratio $\lambda_2/\lambda_3= 18$, the angle between the first and the last magnetic field vectors $\chi=142\,^{\circ}$, the orientation of the axis $(\phi_C,\theta_C)=(135\,^{\circ},19\,^{\circ})$, the direction of minimum variance $(\phi_{min},\theta_{min})=(43\,^{\circ},-6\,^{\circ})$ and eigenvalues $[\lambda_1,\lambda_2,\lambda_3]=[15.9, 12.6, 0.7]$. In Table \ref{table:ListaTwoMCs1998} we show the date of the two MCs, the time between the first and second MC is $8h$ (sheath of the second MC). The minimum speed in the first MC (MC1) was $V_{p1}=272.9~\mathrm{km/s}$ at $15:00$ UT and in the beginning of the second MC (MC2) the maximum speed was $V_{p2}=346.1~\mathrm{km/s}$ one day later, at $05:00$ UT. Then, $V_{p2}>V_{p1}$ and the two events are very close to each other. Probably there are other similar cases where a MC was confused within the sheath of the preceding MC. The entropy index (EI) can help to identify the MCs and such kinds of cases and find the boundary of MC with MVA method. \begin{table}[htp] \caption{Two continues MCs from $19-21$ August $1998$. The EI helps to identify the first MC.} \centering \begin{tabular}{l l l l l l l l l l l} \hline Year & Shock1 & MC1, start & MC1, stop & MC2, start & MC2, stop \\ \hline 1998 & 19 Aug, 02:00 & 19 Aug, 10:00 & 20 Aug, 00:00 & 20 Aug,08:00 & 21 Aug, 18:00 \\ \hline \end{tabular} \label{table:ListaTwoMCs1998} \end{table} \subsubsection{18-February-1999 MC event} A study in more details of the STE variations is presented for some days before and after the event number $11$ Tabla \ref{table:ListaMCs}. Time series of the magnetic components $B_x,\: B_y$ and $B_z$ for MC have nonzero STE values (see, Figure~\ref{Fig:Anisotropias}a, the eleventh point). Why is this happening? To answer this question we calculate the windowed STE of three magnetic components $B_x,\: B_y$ and $B_z$ in windows temporal of ten days. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzFeb1999.eps} \caption{Values of spatio-temporal entropy at $14-24$ February, $1999$ as a function of the time for times series of IMF $B_x,\: B_y \mbox{ and } B_z$ components in the solar wind. The thick curve represents the Entropy Index (EI) calculated over the analyzed period.}. \label{Fig:STEJaneladaFeb1999} \end{figure} \citet{CaneRichardson2003} detected one ICME not classified as MC at the date $13$ Feb $19:00$ - $14$ Feb $15:00$ $1999$ with $V_{ICME} = 440~\mathrm{km/s}$, $V_{max} = 470~\mathrm{km/s}$ and $B= 9~\mathrm{nT}$. Figure~\ref{Fig:STEJaneladaFeb1999} shows the results of the calculation of STE from $14$ Feb $15:00$ $1999$ and the EI has a small but non-zero value. At the date $18-21$ Feb $1999$ the ACE spacecraft detected one ICME classified as MC with $V_{ICME} = 520~\mathrm{km/s}$, $V_{max} = 700~\mathrm{km/s}$, $B= 8~\mathrm{nT}$ and $V_T = 870~\mathrm{km/s}$ was the transit speed \citep{CaneRichardson2003}. The MVA method gives $\lambda_2/\lambda_3= 8$, $\chi=83\,^{\circ}$ and $(\phi_C,\theta_C)=(96\,^{\circ},6\,^{\circ})$. This bipolar MC has a flux-rope type NWS and the observed angular variation of magnetic field is left-handed \citep{Huttunen2005}. Inside the MC represented by vertical lines in the Figure~\ref{Fig:STEJaneladaFeb1999}, one can see that the $STE=0\%$ only for the $B_x$ component. In this case, during the analyzed $10$ days the $EI=0\%$ only inside the MC. So far, we see MCs with zero values of STE only in one component, in two or three at a time. The interpretation is that it may be related to the inclination of the axis of the MC relative to the ecliptic plane and the distance from the spacecraft to the axis, but we have poor statistics and this issue remains open. However, the EI has always detected the presence of a magnetic cloud. \subsubsection{13-October-2000 MC event} It is presented a study in more details of the STE variations with some days before and after the event number $24$ Tabla \ref{table:ListaMCs}. Time series of the magnetic components $B_y$ and $B_z$ for MC have zero STE (see, Figure~\ref{Fig:Anisotropias}a the twenty-fourth point) but $B_x$ has nonzero STE. Figure~\ref{Fig:STEJaneladaOct2000} shows the results of the calculation of STE from $8-18$ October $2000$. On October $8$ the $B_x,\:B_y$ components had small values of STE but different from zero. At this date it was not reported the presence of ICME, but we saw a small magnetic structure where the field had a smooth variation. The EI detects these small structures that are unnoticed in most of the studies. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzOct2000.eps} \\ \caption{We show the values of spatio-temporal entropy on $08-18$ October, $2000$ as a function of the time for times series of IMF Bx, By and Bz components in the solar wind. The thick curve is the calculation of Entropy Index (EI) over the analyzed period.}. \label{Fig:STEJaneladaOct2000} \end{figure} \begin{figure}[htp] \begin{center} $\begin{array}{c@{\hspace{1in}}c} \multicolumn{1}{l}{\mbox{\bf (a)}} & \multicolumn{1}{l}{\mbox{\bf (b)}} \\ [0.1 cm] \includegraphics[width=0.4\textwidth]{AllFigOct07-092000.eps} & \includegraphics[width=0.4\textwidth]{AllFig01-10April2010.eps}\\ [0.2cm] \end{array}$ \end{center} \caption{Left and right part : Solar wind parameters during two MC events. Top to bottom: magnetic field strength, polar (Blat) and azimuthal (Blong) angles of the magnetic field vector in GSE coordinate system, solar wind speed and plasma beta. The first vertical line indicate the shock, the other two vertical lines indicate the interval of a MC. Left: $07-09$ October $2000$, the bipolar MC observed by ACE on $07$ October at $22:00$ UT to $08$ October at $17:00$ UT $2000$. Right: $01-11$ April $2010$, the bipolar MC observed by ACE on $05$ April at $16:00$ UT to $06$ April at $14:00$ UT $2010$.} \label{Fig:DataOct2000Apr2010} \end{figure} In the Figure~\ref{Fig:DataOct2000Apr2010}~(a) solar wind parameters during one MC event are shown. From top to bottom: magnetic field strength, polar ($B_{lat}$) and azimuthal ($B_{long}$) angles of the magnetic field vector in GSE coordinate system, solar wind speed and plasma beta. The first vertical line indicate the shock, the other two vertical lines indicate the interval of a MC. The plasma beta minimum value were $\beta=2.7 \times 10^{-3}$ on $07$ October at $23:00$ UT. We used the magnetic field rotation confined to one plane, the plane of maximum variance $(B_x^*B_y^*)$ to find the boundaries of the cloud (see Table \ref{table:ListaOct2000}). The MVA method gives the eigenvalue ratio $\lambda_2/\lambda_3= 47.5$, the angle between the first and the last magnetic field vectors $\chi=148.4\,^{\circ}$, the orientation of the axis $(\phi_C,\theta_C)=(113.8\,^{\circ},7\,^{\circ})$, the direction of minimum variance $(\phi_{min},\theta_{min})=(22\,^{\circ},-10\,^{\circ})$ and eigenvalues $[\lambda_1,\lambda_2,\lambda_3]=[124.6, 41.8, 0.9]$. This MC has a flux-rope type SEN (see panel two and three in Figure~\ref{Fig:DataOct2000Apr2010}~(a)) and the observed angular variation of the magnetic field is left-handed. \citet{CaneRichardson2003} detected one ICME classified as MC in the date $12$ October $22:28$ - $14$ October $17:00$ $2000$ with $V_{ICME} = 410 km/s$, $V_{max} = 470 km/s$, $B= 13 nT$ and $V_T = 580 km/s$. The MVA method gives $\lambda_2/\lambda_3= 4$, $\chi=62\,^{\circ}$ and $(\phi_C,\theta_C)=(33\,^{\circ},-25\,^{\circ})$. This bipolar MC has a flux-rope type NES and the observed angular variation of magnetic field is right-handed \citep{Huttunen2005}. The $B_y,\: B_z$ components have $STE=0\%$ only during the passage of the MC, approximately in the first half of it and the EI detected the presence of the MC. \begin{table}[htp] \caption{Magnetic Cloud from $07-08$ October $2000$.} \centering \begin{tabular}{l l l l l l} \hline Year & Shock & MC, start & MC, stop \\ \hline $2000$ & $07$ October, $09:00$ & $07$ October, $22:00$ & $08$ October, $17:00$ \\ \hline \end{tabular} \label{table:ListaOct2000} \end{table} \subsubsection{Interface between the interplanetary ejecta and the high-speed stream on 22 October 1999} High-speed streams, originating in coronal holes, are observed often following interplanetary coronal mass ejection (ICME) at $1$ AU \citep{KleinBurlaga1982}. \citet{Alisson2006} studied the $17-22$ October $(1999)$ solar-interplanetary event, which was associated to a very intense magnetic storm ($Dst = -237~\mathrm{nT}$). They presented an analysis of pressure balance between the ICME observed on $21-22$ October and the high-speed streams following this ICME. Close to the Earth, at $L1$, an interplanetary shock was detected by ACE magnetic field and plasma instruments on $21$ October $(1999)$, at $01:34$ UT, as shown in Figure~\ref{Fig:STEJaneladaOct1999} with the first vertical dotted line. The driver of this shock is an ICME, which can be distinguished from the normal solar wind by its intense magnetic field, of the order of $20~\mathrm{nT}$ throughout the most part of $21$ October, and its low beta ($\sim 0.1$) \citep{Alisson2006}. The start of the ejecta was at $03:58$ UT of $21$ October. Toward the end of this ejecta, an increase of the magnetic field intensity was observed, starting at $02:30$ UT of $22$ October, reaching a peak value of $37~\mathrm{nT}$ \citep{Alisson2006}. At $06:15$ UT of $22$ October (the second vertical dotted line in Figure~\ref{Fig:STEJaneladaOct1999}), the magnetic field dropped abruptly around $10~\mathrm{nT}$. \citet{Alisson2006} defined this point as the end of the ICME. They were not sure whether this ICME is an MC or not according to the criteria of \citet{Burlaga1981}, because the direction of the magnetic field does not rotate smoothly. \citet{CaneRichardson2003} also detected one ICME not classified as MC at the date $21$ October $08:00$ UT - $22$ October $07:00$ UT $1999$ with $V_{ICME} = 500~\mathrm{km/s}$, $V_{max} = 580~\mathrm{km/s}$, $B= 20~\mathrm{nT}$ and $V_T= 480~\mathrm{Km/s}$. As presented in the previous paragraph it is clear that during the period $20-26$ October $1999$ was not identified any magnetic cloud. In Figure~\ref{Fig:STEJaneladaOct1999}, we show the values of spatio-temporal entropy as a function of the time for times series of IMF $B_x,\: B_y\: \mbox{and } B_z$ components in the solar wind from $20-26$ October. All time the values of STE are different of zero, so the EI is also different from zero. This result helped to validate the EI to detect MC, because EI methodology can differentiate when an ICME is not classified as a MC, since the EI is zero only inside a MC. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzOct1999.eps}\\ \caption{Values of spatio-temporal entropy on $20-26$ October, $1999$ as a function of the time for times series of IMF $B_x,\: B_y\: \mbox{ and } B_z$ components in the solar wind. The thick curve represents the Entropy Index (EI) calculated over the period analyzed.}. \label{Fig:STEJaneladaOct1999} \end{figure} \subsubsection{Identification by windowed STE method of 05-06 April 2010 MC event} On $3$ April $2010$ the Sun launched a cloud of material, known as a coronal mass ejection (CME), in a direction that reaches the Earth. This CME was much faster than most of the CMEs seen recently, with a speed of at least $800~\mathrm{Km/s}$. The bulk of the CME passed south of Earth, but a piece of it hit the Earth's magnetosphere on April $5$, causing a geomagnetic storm ($Dst_{\mathrm{min}}=-73~\mathrm{nT}$ on $06$ April at $15:00$ UT). As a result bright auroras were seen around the world at higher latitudes. This event was very recent and we wanted to use the technique to verify any possible MC occurrence. The ACE Magnetic Field Experiment data in level $2$ (verified) were not available in April $2010$ when the data were processed. It was only possible to obtain such data for $16$ second average IMF in RTN and GSE coordinates via anonymous ftp, from Caltech \footnote{ftp://mussel.srl.caltech.edu /pub/ace/browse/MAG16sec}. For this reason, in Figure~\ref{Fig:STEJaneladaApril2010} the STE values are calculated for data in GSE coordinates. \begin{figure}[htp] \centering \includegraphics[width=.8\textwidth]{STEBxByBzApril2010.eps} \\ \caption{Values of spatio-temporal entropy on $01-11$ April, $2010$ as a function of the time for times series of IMF $B_x,\: B_y\: \mbox{ and } B_z$ components in the solar wind. The thick curve represents the Entropy Index (EI) calculated over the period analyzed. The MC is identified.}. \label{Fig:STEJaneladaApril2010} \end{figure} Figure~\ref{Fig:STEJaneladaApril2010} is similar to Figure~\ref{Fig:STEJaneladaJan1998} but for the period $01-10$ April $2010$ with GSE coordinates. We find minimum values of $STE= 1 \%$ to $B_z$ on $05$ April at $21:33:20$ and $22:26:40$ UT respectively. The $B_x$ and $B_y$ components had STE values less than $10\%$ on $05$ April at $18:00:00$ and $18:53:20$ UT respectively. Then, the EI has a minimum values less than $1\%$ on day $05$ between $18:00:00$ to $23:20:00$ UT. Thus the EI detects a structure with characteristics of a magnetic cloud. The same method (MVA) explained in \citet{Huttunen2005} was used to find the boundaries of the cloud. \begin{table}[htp] \caption{Magnetic Cloud from $05-06$ April $2010$.} \centering \begin{tabular}{l l l l l l} \hline Year & Shock & MC, start & MC, stop \\ \hline $2010$ & $05$ April, $07:00$ & $05$ April, $16:00$ & $06$ April, $14:00$ \\ \hline \end{tabular} \label{table:ListaMCs2010} \end{table} Figure~\ref{Fig:DataOct2000Apr2010}(b) is similar to (a) but from $01-11$ April $2010$. On $05$ April $2010$, at $07:00$ UT, the proton density $N_p=2.8~\mathrm{cm^{-3}}$, proton temperature $T_p=2.4*10^5~\mathrm{K}$, ratio of $alphas/protons = 7 \times 10^{-2}$, proton speed $V_p=564.5~\mathrm{km/s}$, magnetic field magnitude $B=5.2~\mathrm{nT}$ and plasma beta $\beta=0.9$. One hour later some parameters have an change $N_p=7.98~\mathrm{cm^{-3}}$, $T_p=5.5 \times 10^5~\mathrm{K}$, $alphas/protons = 1.7*10^{-2}$, $V_p= 724.8~\mathrm{km/s}$, $B=10.9~\mathrm{nT}$ and $\beta=1.3$. It is easy to identify a shock because the velocity, density and magnetic field magnitude increases abruptly and we believe it is related to the arrival of an event to ACE. Eight hours latter than the time of the shock, at $16:00$ UT, the plasma beta decreases to $\beta=3.2*10^{-2}$, and it is the beginning of the magnetic cloud. \begin{figure}[htp] \centering \begin{tabular}{c} \includegraphics[width=.7\textwidth]{All_HistKSTE_Bx.eps}\\ \includegraphics[width=.7\textwidth]{All_HistKSTE_By.eps}\\ \includegraphics[width=.7\textwidth]{All_HistKSTE_Bz.eps} \\ [1ex] \end{tabular} \caption{At top: Plot of STE versus $\theta$ to $B_x,\: B_y\: \mbox{and } B_z$ components and a parabola fitted to the data. At bottom: The corresponding histogram in which a Gaussian distribution was fitted.} \label{Fig:AnguloSTE} \end{figure} The plasma beta minimum value were $\beta=8.8 \times 10^{-3}$ on $06$ April at $12:00$ UT. We used the magnetic field rotation confined to one plane, the plane of maximum variance $(B_x^*B_y^*)$ to find the boundaries of the cloud (see Table~\ref{table:ListaMCs2010}). The MVA method gives the eigenvalue ratio $\lambda_2/\lambda_3= 16$, the angle between the first and the last magnetic field vectors $\chi=48\,^{\circ}$, the orientation of the axis $(\phi_C,\theta_C)=(33\,^{\circ},-32.6\,^{\circ})$, the direction of minimum variance $(\phi_{min},\theta_{min})=(129\,^{\circ},9\,^{\circ})$ and eigenvalues $[\lambda_1,\lambda_2,\lambda_3]=[16.8, 3.6, 0.2]$. This MC has a flux-rope type NWS (see panel two and three in Figure~\ref{Fig:DataOct2000Apr2010}~(b)) and the observed angular variation of the magnetic field is left-handed. \subsection{Correlation between the Windowed STE and the inclination of IMF} Following the same idea presented in subsection~\ref{Trend in time series}, we calculate the STE versus the trend angle. In Figure~\ref{fig:TendenciaLorentz}, we presented time series plot of Lorenz data file include in the VRA for four differents trend angles. We mentioned earlier that the time series of magnetic clouds are not stationary, there is a trend. The linear least squares fitting technique is the simplest and most commonly applied form of linear regression and provides a solution to the problem of finding the best fitting straight line through a set of points and to obtain the equation of the line. The angle $\theta$ that a line makes with the positive x axis is closely related to the slope m via the inverse tangent function $\theta=\tan^{-1}(m)$. We applied the previous methodology to all IMF time series shown in the Figures~\ref{Fig:STEJaneladaJan1998}, \ref{Fig:STEJaneladaAug1998}, \ref{Fig:STEJaneladaFeb1999}, \ref{Fig:STEJaneladaOct2000}, \ref{Fig:STEJaneladaOct1999}. In other word, we separate all pairs of points formed by the trend angle $\theta$ and STE for the data. Figure~\ref{Fig:AnguloSTE} presents for the $B_X$, $B_Y$ and $B_Z$ components, at the top of each panel, the plot of STE versus $\theta$ for each magnetic component and a parabola fitted to the data, and at the bottom of each panel, the corresponding histogram, in which a Gaussian distribution was fitted. At the first panel in the figure, for $B_X$, from the STE versus $\theta$, the correlation coefficient (a Pearson's correlation coefficient between two time series, STE vs $\theta$) found was $C=-70.8\%$. If the variables tend to go up and down in opposition, with low values of one variable associated with high values of the other, the correlation coefficient will be negative (anti-correlation), as in our cases. The anti-correlation between the two parameters (STE and $\theta$) gives the idea that the trend in the $B_X$ series is responsible for the low values of STE. But the above idea is not totally true; because we have $STE\neq0\%$ series with a larger trend, more separated points of the vertical line. The histograms help to see the distribution of the tend. One response to these kinds of problems can be fluctuations in the signal. Sometimes, the trend in the records for a sheath region is bigger than the trend for a cloud, but the sheath has rapid fluctuations that changes the sign of the signal and the STE is not zero. Then the $STE~=~0\%$ at magnetic clouds is related to the simultaneous occurrence of two factors: (1) the trend existing in the signal, (2) the analyzed variable (in this case, IMF $B_x$) has a more ordered dynamical behavior (\textit{i.e.} few fluctuations) and higher degree of correlation between its time neighbors. In Figure~\ref{Fig:AnguloSTE} the middle panel, for $B_Y$, and the bottom panel, for $B_Z$, present respectively the correlation coefficient ($C$) results of $-67.6\%$ and $-65.7\%$. From the three panel, it is possible to see that $B_z$ component has largest trend compared to the others two, but the reason is not know. \section{Conclusions} \label{Conclusions} We dealt with time series of solar wind for a group of magnetic clouds in order to analyze the dynamical behavior of the IMF~$B_x$, $B_y$ and~$B_z$ components. The tool chosen to this study was the spatio-temporal entropy (called STE) method. From methodological considerations based on physical reasons, the STE was calculated from solar wind time windows with $2500$ points corresponding to $11.11$ hours of data. The results were significant and very promissing. we found higher entropy values in the sheaths. MCs present zero STE values at least for one magnetic component. As not all the magnetic components have zero entropy at the same time, an useful technique was to create a standardization index, called Entropy Index (EI), to allow joining the STE results of the three variable $(B_x, B_y, B_z)$. We studied in more detail the STE variations taking into account some days before and after to six selected events, in order to allow understanding and validating the analysis procedure. By this approach, the EI identified three new MCs in addition to the known MCs presented in previous tables from other researchers, and after the MVA method was used to calculate the boundaries easily. The analyses developed in this work shows that: \begin{itemize} \item MCs have STE values lower than the ones for sheath region. \item The differences among the STE values for the three magnetic components in a MC give an idea about the anisotropy in the structure of some MCs. But this idea is still an open question to be better investigated. \item In MC the magnetic field strength is higher than the average, the magnetic field direction rotates smoothly through a large angle, then periods of MC have more trend than sheath region and "quiet" solar wind periods. The trend is the principal cause of the lower values of STE in the MC. Also, MCs are most structuredness than sheath and "quiet" solar wind. \item Using the EI and MVA method it is possible to identify all the magnetic clouds that arrive at spacecrafts. \item The entropy index (EI) proposed gets success as an auxiliary tool to identify magnetic cloud candidates and calculation of EI could be used in an automatic procedure of preliminary investigation or convenient warning. \end{itemize} Based on its skill to provide an objective analysis, the STE method can be seen as a valuable tool in the study of interplanetary phenomena such as the time localization of MCs. \section{Acknowledgments} This work was supported by CNPq (grants $309017/2007-6$, $486165/2006-0$, $308680/2007-3$, $478707/2003$, $477819/2003-6$, $382465/01-6$), FAPESP (grants $2007/07723-7$) and CAPES (grants $86/2010-29$). We are grateful to V. E. Menconni (FAPESP $2008/09736-1$) for their helpful computational assistance. The authors would like to thank ACE Staff and WDC-Kyoto for the datasets used in this work. Ojeda G. A. would like to thank the CAPES and CNPq for the PhD Scholarship Programme. \bibliographystyle{elsarticle-harv}
\section{#1}\eqnreset} \renewcommand{\theequation}{\thesection.\arabic{equation}} \def\medno{\medbreak\noindent} \def\text#1{\;\;\;\;{\rm \hbox{#1}}\;\;\;\;} \def\qquad{\quad\quad} \def\qqquad{\quad\quad\quad} \def\itema{\item[{\rm (a)}]} \def\itemb{\item[{\rm (b)}]} \def\itemc{\item[{\rm (c)}]} \def\itemd{\item[{\rm (d)}]} \def\iteme{\item[{\rm (e)}]} \def\itemf{\item[{\rm (f)}]} \def\itemg{\item[{\rm (g)}]} \def\itemh{\item[{\rm (h)}]} \def\itemi{\item[{\rm (i)}]} \def\msy#1{{\mathbb #1}} \def\C{{\msy C}} \def\N{{\msy N}} \def\Z{{\msy Z}} \def\R{{\msy R}} \def\D{{\msy D}} \def\ga{\alpha} \def\gb{\beta} \def\gd{\delta} \def\ge{\varepsilon} \def\gf{\varphi} \def\gg{\gamma} \def\gk{\kappa} \def\gl{\lambda} \def\gn{\nu} \def\gr{\rho} \def\gs{\sigma} \def\gx{\chi} \def\gz{\zeta} \def\gD{\Delta} \def\gF{\Phi} \def\gS{\Sigma} \def\got#1{\mathfrak #1} \def\fa{{\got a}} \def\fb{{\got b}} \def\fg{{\got g}} \def\fh{{\got h}} \def\fj{{\got j}} \def\fk{{\got k}} \def\fl{{\got l}} \def\fm{{\got m}} \def\fn{{\got n}} \def\fbn{{\bar {\n}}} \def\fp{{\got p}} \def\fq{{\got q}} \def\fs{{\got s}} \def\ft{{\got t}} \def\fv{{\got v}} \def\implies{\Rightarrow} \def\to{\rightarrow} \def\Re{{\rm Re}\,} \def\Im{{\rm Im}\,} \def\inp#1#2{\langle#1\,,\,#2\rangle} \def\hinp#1#2{\langle#1\,|\,#2\rangle} \def\Ad{{\rm Ad}} \def\Endo{{\rm End}} \def\Hom{{\rm Hom}} \def\Id{{\rm I}} \def\ad{{\rm ad}} \def\after{\,{\scriptstyle\circ}\,} \def\iso{\simeq} \def\pr{{\rm pr}} \def\tensor{\otimes} \def\implies{\Leftarrow} \def\pijl#1{{\buildrel #1 \over \longrightarrow}} \def\tr{{\rm tr}\,} \def\ideal{\,\lhd\,} \def\iq{{\rm q}} \def\iC{{\scriptscriptstyle \C}} \def\iR{{\scriptscriptstyle \R}} \def\cA{{\mathcal A}} \def\cB{{\mathcal B}} \def\cC{{\mathcal C}} \def\cD{{\mathcal D}} \def\cE{{\mathcal E}} \def\cF{{\mathcal F}} \def\cG{{\mathcal G}} \def\cH{{\mathcal H}} \def\cI{{\mathcal I}} \def\cK{{\mathcal K}} \def\cL{{\mathcal L}} \def\cM{{\mathcal M}} \def\cN{{\mathcal N}} \def\cO{{\mathcal O}} \def\cP{{\mathcal P}} \def\cR{{\mathcal R}} \def\cS{{\mathcal S}} \def\cT{{\mathcal T}} \def\cU{{\mathcal U}} \def\cV{{\mathcal V}} \def\cW{{\mathcal W}} \def\cY{{\mathcal Y}} \def\cZ{{\mathcal Z}} \mathcode`:="603 \def\col{\,:\, \def\too{\longrightarrow} \def\Ci{C^\infty} \def\Cartan{\theta} \def\Aut{{\rm Aut}} \def\fap{\fa_0} \def\Ap{A_0} \def\minparabs{\cP_0} \def\fapd{\fa_0^*} \def\fapdc{\fa_{0\iC}^*} \def\stN{N_0} \def\stM{{M_0}} \def\cAtwo{\cA_2} \def\Cci{C_c^\infty} \def\udE{{}^u\! E^*} \def\Vtau{V_\tau} \def\ev{{\rm ev}} \def\uFou{{}^u \cF} \def\tayl{{\rm tayl}} \def\supp{{\rm supp}\,} \def\uAC{{}^u\!{\rm AC}} \def\uACh{{}^u\!{\rm AC}_\cO} \def\uAAC{{}^u_{\rm A}{\rm AC}} \def\Pst{P_0} \def\dotvar{\,\cdot\,} \def\EPW{{\mathrm H}} \def\EPR{{\mathrm H}_R} \def\uPW{{}^u {\rm PW}} \def\uAPW{{}^u_{\rm A}{\rm PW}} \def\App{A^+} \def\cl{{\rm cl}} \def\embeds{\hookrightarrow} \def\uPWd{{}^u{\rm PW}^*} \def\PW{{\rm PW}} \def\dE{E^*} \def\Fou{\cF} \def\Ccmin{C^{-\infty}_c} \def\multCP{\cU_P} \def\multCPt{\cU_P^t} \def\Mer{\cM} \def\Hyp{\cH} \def\laur{{\rm laur}} \def\Lau{\cL} \def\AC{{\rm AC}} \def\Aut{{\rm Aut}} \def\uFoust{{}^u \cF_0} \def\faQp{\fa_Q^+} \def\AQp{A_Q^+} \def\ooC{{}^\circ\hspace{-0.6pt} C} \def\Ccminf{C^{- \infty}_c} \def\hol{{\rm hol}} \def\uAChol{{}^u{\rm AC}_\hol} \def\void{\varnothing} \def\mspace{\mbox{\hspace{-0.5pt}}} \def\spaceind{{}_{\scriptscriptstyle *}} \def\spgs{\spaceind\sigma} \def\spG{\spaceind\mspace G} \def\spH{\spaceind\mspace H} \def\spK{\spaceind\mspace K} \def\spfaq{{\spaceind \mspace\fa_\iq}} \def\spnE{{\spaceind \mspace E^\circ}} \def\spgl{{\spaceind \gl}} \def\spP{{\spaceind \mspace P}} \def\sppsi{{\spaceind\mspace \psi}} \def\nE{{E^\circ}} \def\sptau{{\spaceind\mspace \tau}} \def\spX{{\mathrm X}} \def\spspX{{\spaceind \spX}} \def\spf{{\spaceind f}} \def\Cartan{\theta} \def\spCartan{\spaceind\mspace\Cartan} \def\minf{{-\infty}} \def\spfp{\spaceind \fp} \def\spfg{\spaceind \fg} \def\spfq{\spaceind \fq} \def\spfh{\spaceind \fh} \def\spinp#1#2{\spaceind\inp{#1}{#2}} \def\spB{\spaceind B} \def\barB{\bar B} \def\spbarB{\spaceind \bar B} \def\spfaqd{\spaceind \fa_\iq^*} \def\spfaqdc{\spaceind \fa_{\iq\iC}^*} \def\spga{\spaceind \ga} \def\spgS{\spaceind \gS} \def\spAq{\spaceind A_\iq} \def\spPmin{\spaceind\cP_\gs^{\mathrm{min}}} \def\sprho{\spaceind\rho} \def\diag{\mathrm{diag}} \def\spcAtwo{\spaceind\cA_2} \def\spLau{\spaceind \cL} \def\spW{\spaceind W} \def\spcW{\spaceind \cW} \def\oC{{}^\circ C} \def\pspsi{\spaceind \psi} \def\spM{\spaceind M} \def\spbarP{\spaceind \bar P} \def\tpsi{\tilde \psi} \def\sptpsi{\spaceind \tilde \psi} \def\HC{{\mbox{\tiny \rm HC}}} \def\spHyp{\spaceind \cH} \def\spd{\spaceind d} \def\sppi{\spaceind \pi} \def\fapdreg{\fa_0^{*{\rm reg}}} \def\bfapd{\bar\fa_{0}^*} \def\mult{\gamma} \def\barB{\bar B} \def\inj{{\mathrm i}} \def\gL{{\Lambda}} \def\faPdc{\fa_{P\iC}^*} \def\PWpre{{\rm PW}^{\rm pre}} \def\PWpreP{{\rm PW}_P^{\rm pre}} \def\PWA{{\rm PW}_{\rm A}} \def\fadc{{\fa_\iC^*}} \def\PWPzeropre{{\rm PW}_{P_0}^{\rm pre}} \def\bfi{{\mathbf i}} \def\diffD{J} \def\Cartan{\theta} \def\oL{{}^\circ L} \def\Ind{{\rm Ind}} \def\fac{\fa_\iC} \def\data{\Sigma} \def\faPd{\fa_P^*} \def\half{{\textstyle \frac12}} \def\Vect{{\rm Vect}} \def\FDVect{{\rm FDVect}} \def\coleq{\;:\!=} \def\dM{M^\wedge} \def\dMP{M_P^\wedge} \def\dMPds{M_{P, ds}^\wedge} \def\dMQds{M_{Q, ds}^\wedge} \def\hatMPds{\widehat M_{P,ds}} \def\hatMPzero{\widehat M_{P_0}} \def\dK{K^\wedge} \def\hatM{\widehat M} \def\dstarM{{}^*M^\wedge} \def\hatotimes{\widehat\otimes} \def\holspace{\cF} \def\rmM{{\rm M}} \def\FDMOd{{\rm FDmod}} \def\FDmod{{\rm FDmod}} \def\HFG{{\rm H\rm F}_G} \def\HFGinfty{\HFG^\infty} \def\HFGinftyadm{\HFG^{\infty, {\rm adm}}} \def\HFGO{{}^\circ\HFG} \def\HFGOinfty{{}^\circ\HFGinfty} \def\HFGOinftyadm{{}^\circ\HFGinftyadm} \def\HFL{{\rm H\rm F}_L} \def\HFLinfty{\HFL^\infty} \def\HFLinftyadm{\HFL^{\infty, {\rm adm}}} \def\HFLO{{}^\circ\HFL} \def\HFLOinfty{{}^\circ\HFLinfty} \def\HFLOinftyadm{{}^\circ\HFLinftyadm} \def\obj{{\rm obj}} \def\rmi{{\rm i}} \def\fg{{\mathfrak g}} \def\HFgKadm{{\rm H\rm F}_{(\fg, K)}^{\,{\rm adm}}} \def\HFGadm{\HFG^{\rm adm}} \def\gs{\sigma} \def\Fou{{\mathfrak F}} \def\one{{\rm I}_{C^\infty(K/M)}} \def\FMod{{\rm F}{\rm M}} \def\pre{{\rm pre}} \def\bs{\backslash} \numberwithin{equation}{section} \title{A comparison of Paley--Wiener theorems for real reductive Lie groups} \author{E.~P.~van den Ban, S.~Souaifi} \begin{document} \maketitle \tableofcontents \section*{Introduction} In this paper we make a detailed comparison between the Paley--Wiener theorem of J. Arthur \cite{arthurpw}, and the one recently established by P.\ Delorme \cite{delormepw}. \par Let $G$ be a real reductive Lie group of the Harish-Chandra class and let $K$ be a maximal compact subgroup. Let $C_c^\infty(G, K)$ denote the space of smooth compactly supported functions on $G$ which behave finitely under both left and right translation by $K.$ The Paley--Wiener theorem of each of the above mentioned authors describes the image of $C_c^\infty(G,K)$ under Fourier transformation, in terms of a so called Paley--Wiener space. In this paper we will show that the two Paley--Wiener spaces are equal, without using the proof or the validity of any of the associated Paley--Wiener theorems. It thus follows that the two theorems are equivalent from an a priori point of view. \par In order to be able to be more specific about the contents of this paper, we shall first give a more detailed description of the two Paley--Wiener theorems mentioned above. \par Let $\fg = \fk \oplus \fp$ be a Cartan decomposition associated with the maximal compact subgroup $K.$ Here and in the following we use the convention to denote Lie groups by Roman capitals, and their Lie algebras by the corresponding lower case German letters. Let $\fa$ be a maximal abelian subspace of $\fp,$ and let $A = \exp \fa$ be the associated vectorial closed subgroup of $G.$ We denote by $\cP(A)$ the (finite) set of cuspidal parabolic subgroups of $G$ containing $A.$ Each parabolic subgroup $P \in \cP(A)$ has a Langlands decomposition of the form $P = M_P A_P N_P,$ with $A_P \subset A.$ Let $\dMPds$ denote the set of (equivalence classes of) discrete series representations of $M_P.$ For $(\xi,\cH) \in \dMPds$ and $\gl \in \faPdc,$ we denote by $\pi_{\xi,\gl}= \pi_{P, \xi,\gl}$ the representation of $G$ induced from the representation $\xi \otimes (\gl + \rho_P) \otimes 1$ of $P.$ The associated module of smooth vectors for $\pi_{\xi, \gl}$ has a realization on the space $C^\infty(K\col \xi)$ consisting of smooth functions $\psi: \;K \to \cH_\xi,$ transforming according to the rule: $$ \psi(mk) = \xi(m) f(k),\qquad\mbox{\rm for all} \quad k \in K, m \in K\cap M_P. $$ Accordingly, each function $f \in C^\infty_c(G)$ has an operator valued Fourier transform $$ \hat f (P, \xi, \gl) \coleq \pi_{\xi,\gl}(f) = \int_G f(x)\, \pi_{\xi, \gl} (x) \; dx \in \Endo(C^\infty(K: \xi)). $$ Moreover, the endomorphism $\hat f(P, \xi, \gl)$ depends holomorphically on the variable $\gl \in \faPdc.$ If $f$ is bi-K-finite, then $\hat f(P, \xi, \gl)$ belongs to the space $\cS(P: \xi)$ of bi-K-finite elements of $\Endo(C^\infty(K: \xi)).$ Moreover, the holomorphic maps $\gl \mapsto \hat f (P, \xi, \gl)$ are non-zero for only finitely many of the pairs $(P, \xi).$ It follows that $\hat f(P)$ may be viewed as an element of the algebraic direct sum \begin{equation} \label{e: holomorphic functions for PW} \oplus_{\xi}\;\;\cO(\faPdc) \otimes \cS(P:\xi). \end{equation} We define the pre-Paley--Wiener space $\PWpreP(G,K)$ as the space of elements $(\gf(P,\xi)\mid \xi \in \dMPds)$ in (\ref{e: holomorphic functions for PW}) for which there exists a constant $R> 0$ and for every $n > 0$ a constant $C_n > 0$ such that \begin{equation} \label{e: PW estimate} \| \gf (P, \xi, \gl) \| \leq C_n ( 1 + |\gl|)^{-n} e^{R |\Re \gl|} \end{equation} for all $\xi, \gl.$ \par We can now describe the Paley--Wiener space involved in Arthur's theorem in \cite{arthurpw}. Let $P_0$ be a fixed minimal parabolic subgroup in $\cP(A).$ Its Langlands decomposition is of the form $P_0 = M A N_0,$ where $M$ is the centralizer of $A$ in $K.$ The Arthur Paley--Wiener space \begin{equation} \label{e: PW A} \PW^A(G,K) \end{equation} is defined as the space of $\gf \in \PWPzeropre(G,K)$ satisfying all finite linear relations of the form \begin{equation} \label{e: arthurs condition} \sum_i \inp{ \gf(P_0,\xi_i, \gl_i; u_i)}{ \psi_i} = 0, \end{equation} with $\xi_i \in \dM,$ $\psi_i \in \cS(P_0: \xi_i)^*_{K \times K},$ $\gl_i \in \fadc$ and with $u_i \in S(\fa^*)$ acting as differential operators in the $\gl$-variable (see Section \ref{ss: derivation process} for notation), as soon as these relations are satisfied by all families of functions $(\gl \mapsto \pi_{P_0, \xi, \gl}(x)\mid \xi\in \dM),$ for $x \in G.$ These are the so-called Arthur-Campoli relations. In \cite{arthurpw}, Arthur defines a similar Paley--Wiener space involving all minimal parabolic subgroups from $\cP(A).$ In \cite{BSpwspace} this space is shown to be isomorphic to the one defined in (\ref{e: PW A}). \par Next, let us describe the Paley--Wiener space introduced by Delorme \cite{delormepw}. The definition of this Paley--Wiener space involves the operation of taking successive derivatives of a family $\pi_{\gl}$ of representations, depending holomorphically on a parameter $\gl \in \faPdc,$ for some $P \in\cP(A).$ Such an operation is encoded by a sequence $\eta =(\eta_1, \ldots, \eta_N)$ in $\faPdc,$ listing the directions in which the derivatives should be taken successively. The associated family $\pi^{(\eta)}_\gl$ is again a holomorphic family of representations. The operation of derivation also applies to a holomorphic family $\gl \mapsto \gf_\gl$ of continuous endomorphisms of $\pi_\gl$ and then gives a holomorphic family $\gf^{(\eta)}$ of endomorphisms of $\pi^{(\eta)}_\gl.$ Let $\cD$ be the set of all $4$-tuples $(P, \xi, \gl, \eta)$ with $P \in \cP(A),$ $\xi \in \dMPds,$ $\gl \in \faPdc$ and $\eta$ a finite sequence of linear functionals from $\faPdc$ as above. Given a datum $\delta = (P, \xi,\gl ,\eta)$ we define $\pi_\gd \coleq \pi_{P, \xi, \gl}^{(\eta)}.$ Moreover, given $\gf \in \PWPzeropre(G,K)$ we define $\gf_\gd$ in a similar fashion. Finally, given a sequence $\gd = (\gd_1, \ldots, \gd_N)$ of data from $\cD,$ we write $\pi_{\gd}\coleq \pi_{\gd_1} \oplus \cdots \oplus \pi_{\gd_N},$ and $\gf_\gd\coleq \gf_{\gd_1} \oplus \cdots \oplus \gf_{\gd_N}.$ \par Delorme's Paley--Wiener space is defined as the space $\PW^D(G,K)$ of functions $\gf \in \oplus_{P\in\cP(A)}\PWpreP(G,K)$ such that \begin{enumerate} \itema for each finite sequence $\gd \in \cD^N$ the function $\gf_\gd$ preserves all invariant subspaces of $\pi_{\gd};$ \itemb for any two finite sequences $\gd_1 \in \cD^{N_1}$ and $\gd_2 \in\cD^{N_2},$ and any two sequences of closed invariant subspaces $U_j \subset V_j$ for $\pi_{\gd_j},$ $(j=1,2),$ the induced maps $\bar \gf_{\gd_j} \in \Endo(V_j/U_j)$ are intertwined by all $G$-equivariant operators $T: V_1/U_1 \to V_2/U_2.$ \end{enumerate} \par There is a natural map $\PW^D(G,K) \to \PWPzeropre(G,K),$ given by \begin{equation}\label{e: restriction map PW} (\gf(P, \xi)\mid P \in \cP(A),\xi \in \hatMPds) \mapsto (\gf(P_0, \xi)\mid \xi \in \hatMPzero). \end{equation} In this paper we show that the map (\ref{e: restriction map PW}) is a linear isomorphism from $\PW^D(G,K)$ onto $\PW^A(G,K),$ see Theorem \ref{t: main thm}. \par To understand better the conditions involving the derivatives in the definition of the Paley--Wiener spaces, we start, in Section \ref{s: holomorphic families}, with the study of holomorphic families and their derivatives. Instead of focusing on first order derivatives, we replace a holomorphic family by the associated holomorphic section in a suitable jet bundle. This idea also occurs in W.\ Casselman's paper \cite[\S 9]{casselman}. We reformulate it slightly, by using a suitable trivialization of the jet bundle. Our construction starts with fixing a finite dimensional module $E$ for the ring $\cO_0$ of germs of holomorphic functions (of $\gl$) at zero. It then gives, for $\pi_\gl$ a holomorphic family of representations in a fixed complete space $V$ a new holomorphic family $\pi_\gl^{(E)}$ in the space $E \otimes V.$ The differentiation procedure of Delorme turns out to be a special case of this procedure, with $E$ a suitable module of dimension $2.$ In the same Section \ref{s: holomorphic families} we study how the functor $\pi_\gl \mapsto \pi_\gl^{(E)}$ behaves with respect to analytic families of intertwining operators and with respect to induction. \par In the next section, \ref{s: Arthur's PW}, we give an equivalent definition of Arthur's Paley--Wiener space by invoking the functors $\pi \mapsto \pi^{(E)},$ instead of the derivations given by elements of $S(\fa^*)$ in the Arthur-Campoli relations. These relations may then be reformulated as linear relations on differentiated families of representations. \par In Section \ref{s: Delorme's PW} we simplify the definition of ${\rm PW}^D(G,K).$ First of all, due to the intertwining relations in the definition, the space can be defined in terms of just the minimal parabolic subgroup $P_0.$ Next, the intertwining conditions (b) turn out to be a consequence of the invariant subspace conditions (a). \par In the final Section \ref{s: Hecke algebra}, we bring into play the Hecke algebra $\H(G,K)$ consisting of all bi-K-finite distributions on $G$ supported by $K.$ The importance of this algebra for representation theory is based on the fact that the category of Harish-Chandra modules is isomorphic to the category of finitely generated admissible modules for this algebra. A key lemma in this section is the following. For $(\pi,V)$ a Harish-Chandra module, let $\Endo(\pi)^\#$ denote the space of $K\times K$-finite endomorphisms $\gf$ of $V$ with the property that for every positive integer $n$ the Cartesian power $\gf^{\times n}$ preserves all invariant subspaces of $V^{\times n}.$ The mentioned key lemma asserts that $$ \Endo(\pi)^{\#} = {\rm image}(\H(G,K))\subset \Endo(V). $$ It follows from this lemma that the Paley--Wiener space $\PW^{D}(G,K)$ allows the following description in terms of the Hecke algebra. For every finite dimensional $\cO_0$-module $E$, and all finite sets $\Xi \subset \dM$ and $\gL \subset \fadc,$ we define the representation $\pi_{E, \Xi, \gL}$ to be the direct sum of the representations $\pi_{P_0, \xi, \gl}^{(E)},$ for $(\xi, \gl) \in \Xi\times \gL.$ Moreover, for $\gf \in \PWPzeropre(G,K),$ we define the endomorphism $\gf_{E, \Xi, \gL}$ of $\pi_{E, \Xi, \gL}$ by taking a similar direct sum. Then $\PW^{D}(G,K)$ maps isomorphically onto the space of $\gf \in \PWPzeropre(G,K)$ such that for all $E, \Xi, \gL$ as above, \begin{equation} \label{e: gf in Hecke} \gf_{E, \Xi, \gL} \in \pi_{E, \Xi, \gL}(\H(G,K)). \end{equation} On the other hand, it follows from its definition that Arthur's Paley--Wiener space $\PW^A(G,K)$ is equal to the space of $\gf \in \PWPzeropre(G,K)$ such that for all $E, \Xi, \gL$ as above, $ \gf_{E, \Xi, \gl} $ is annihilated by the annihilator of $\pi_{E, \Xi, \gL}(\H(G,K))$ in the contragredient module. Since this condition is equivalent to (\ref{e: gf in Hecke}), it thus follows that the map (\ref{e: restriction map PW}) is a linear isomorphism onto $\PW^A(G,K)$ (Theorem \ref{t: main thm}). Returning to the original formulation of Arthur's Paley--Wiener theorem, we finally wish to mention that the condition (\ref{e: arthurs condition}) may be replaced by the condition that for all $\xi_i, u_i, \gl_i$ as in (\ref{e: arthurs condition}), there exists a $h \in \H(G,K)$ such that for all $i,$ $$ \gf(\xi_i, \gl_i;u_i) = \pi_{P_0, \xi_i, \gl_i;u_i}(h). $$ This characterization is given in Subsection \ref{s: another useful char of PW} where it is used to derive, from Arthur's theorem, the Paley-Wiener theorem for bi-K-invariant functions, due to S.\ Helgason \cite{helgasonpw} and R.\ Gangolli \cite{gangolli}. \medno {\bf Acknowledgments:\ } We thank Pierre Baumann for a helpful discussion, which led to a simpler proof of Lemma \ref{l: dcomm}. The second named author was partially supported by a grant of The Netherlands Organization for Scientific Research, NWO, under project number 613.000.213. \section{Notation and preliminaries} Throughout this paper, $G$ will be a real reductive Lie group in the Harish-Chandra class and $K$ a maximal compact subgroup. Let $U(\g)$ be the universal enveloping algebra of the complexification $\g_\iC$ of $\g.$ We denote by $X\mapsto X^\vee$ the anti-automorphism of $U(\fg)$ which on $\fg$ is given by $X \mapsto - X.$ \par In this paper, locally convex spaces will always be assumed to be Hausdorff and defined over $\C.$ \par For any continuous representation $(\tau,V)$ of $K$ (in a quasi-complete locally convex space) and any class $\gamma$ in the unitary dual $\dK$ of $K$, the $K$-isotypic component of $(\tau,V)$ of type $\gamma$ is denoted by $V_\gg.$ The associated $K$-equivariant projection onto $V_\gg$ is denoted by $P_\gg.$ For every finite subset $\theta$ of $\dK$, we put $$ V_\theta\coleq \oplus_{\gamma\in\theta}\;V_\gamma \quad \text{and}\quad P_\theta\coleq \oplus_{\gamma\in\theta}\;P_\gamma. $$ \par For any continuous representation $(\pi,V)$ of $G$, with $V$ a quasi-complete locally convex space, let $V^\infty$ and $V_K$ denote the vector subspaces of smooth and $K$-finite elements of $V,$ respectively. The first one gives rise to a subrepresentation of $\pi$ and the second one to its underlying $(\g,K)$-module $(\pi,V_K).$ \par We say that a (continuous) $G$-representation or a $(\g,K)$-module is admissible if all its $K$-isotypic components are finite dimensional. A Harish-Chandra module is an admissible $(\g,K)$-module which is finitely generated as a $U(\g)$-module. \par The space $C^\infty(G)$ of complex valued smooth functions on $G$ is equipped with the left and right regular actions of $G;$ the subspace $C_c^\infty(G)$ of compactly supported functions is invariant for these actions. The actions are continuous for the usual locally convex topologies on these spaces and may be dualized by taking contragredients. Let $\E'(G)$ denote the space of compactly supported distributions on $G,$ i.e., the topological linear dual of $C^\infty(G).$ \par Fix a (bi-invariant) Haar measure $dx$ on $G.$ Then the linear map $$ C_c^\infty(G) \; \longrightarrow \;\E'(G),\quad f \;\longmapsto \; f\,dx $$ is an injective intertwining operator for both $G$-action. Accordingly, we will use this map to view $C_c^\infty(G)$ as a submodule of $\cE'(G).$ \par For any continuous representation $(\pi,V)$ and any $f\in C_c^\infty(G)$, let $\pi(f)$ denote the endomorphism of $V$ defined by $$\pi(f)v\coleq \int_Gf(x)\pi(x)v\,dx,\quad v\in V.$$ Then for all $v\in V^\infty$ and $\xi\in (V^\infty)^*$, the following equality holds $$ \xi(\pi(f)v)=\langle f\,dx,\xi(\pi(\,\cdot\,)v)\rangle; $$ the bracket on the right-hand side of the equation indicates the natural pairing between $\E'(G)$ and $C^\infty(G)$. \par Let $(\pi_i,V_i)$, $i=1,2$, be two $(\g,K)$-modules. The space $\Hom(V_1,V_2)$ of (linear) homomorphisms from $V_1$ to $V_2$ is naturally endowed with a $(\g\times\g,K\times K)$-module structure. Indeed, for any $T\in \Hom(V_1,V_2)$, $$ \begin{array}{l} (X_1,X_2)T=\pi_2(X_2)\circ T - T \circ \pi_1(X_1),\quad X_1,X_2\in \g,\\ (k_1,k_2)T=\pi_2(k_2)\circ T\circ \pi_1(k_1^{-1}),\quad k_1,k_2\in K. \end{array} $$ Accordingly, the subspace $\Hom_{(\g,K)}(V_1,V_2)$ of $(\g,K)$-homomorphisms consists of the elements of $\Hom(V_1,V_2)$ which are invariant under the diagonal action. \begin{lemma}\label{ext1} Let $(\pi,V)$ be an admissible representation of $G$ and let $\theta_1$ and ${\theta_2}$ be finite subsets of $\dK.$ Then the linear map $$ C_c^\infty(G) \; \longrightarrow \;\Hom(V_{\theta_1}, V_{\theta_2}),\qquad f \; \longmapsto \; P_{{\theta_2}} \pi(f) P_{{\theta_1}} $$ uniquely extends to a continuous linear map from $\E'(G)$ to $\Hom(V_{\theta_1}, V_{\theta_2}).$ \end{lemma} \begin{proof} Uniqueness of the extension follows by density of $C_c^\infty(G)\, dx$ in $\E'(G).$ Let $v\in V_{\theta_1}$ and $\xi\in V_{\theta_2}^*.$ Then by finite dimensionality of $\Hom(V_{\theta_1}, V_{\theta_2})$ it suffices to show that the linear map $$ \mathcal{L}\,:\,C_c^\infty(G) \; \longrightarrow \; \C,\quad f \; \longmapsto \;\xi(\pi(f)v) $$ extends continuously to $\E'(G).$ Define the function $m \in C^\infty(G)$ by $m(x) = \xi(\pi(x)v).$ Then $\mathcal{L}(f) = \langle f\, dx , m\, \rangle$, for $f\in C_c^\infty(G).$ Thus, $u \mapsto \langle u , m \rangle$ defines a continuous linear extension of $\cL$ to $\E'(G).$ \end{proof} We consider the convolution product $*$ on $C^\infty_c(G)$ given by $$ f * g\, (x) = \int_G f(y) g(y^{-1} x )\; dy, $$ for $f ,g \in C^\infty_c(G)$ and $x \in G.$ It defines an algebra structure on $C_c^\infty(G).$ The subspace $C_c^\infty(G,K)$ of left and right $K$-finite elements in $C_c^\infty(G)$ is closed under convolution, hence a subalgebra of $C_c^\infty(G).$ \par The convolution product has a unique extension to a separately continuous bilinear map $\E'(G) \times \E'(G) \to \E'(G),$ denoted $(u,v) \mapsto u * v.$ This turns $\E'(G)$ into an algebra. It is readily seen that the subspace $\E'(G,K)$ of left and right $K$-finite elements in $\E'(G)$ is closed under convolution, hence a subalgebra. Likewise, the subspace $\E_K'(G)$ of distributions with support in $K$ is a subalgebra, and so is the intersection \begin{equation} \label{e: Hecke algebra} \H(G,K)\coleq \E_K'(G) \cap \E'(G,K). \end{equation} The latter is also called the Hecke algebra of the pair $(G,K)$ and is sometimes denoted by $\H$ for simplicity. \par From Lemma \ref{ext1}, we obtain the continuous linear map \begin{equation} \label{e: pi on E'(G,K)} \E'(G, K) \; \to \; \mathrm{End}(V_K)_{K\times K}, \quad u \; \mapsto \; \pi(u) \end{equation} which intertwines the $(\g\times \g,K\times K)$-actions. Here the space on the right is equipped with the weakest topology for which the $K \times K$-equivariant projections of finite rank are continuous. By application of Fubini's theorem we see that $\pi$ is a morphism on the convolution algebra $C_c^\infty(G,K),$ which is a dense subalgebra of $ \E'(G,K).$ By separate continuity of $\ast$ and continuity of (\ref{e: pi on E'(G,K)}), it now follows that the map (\ref{e: pi on E'(G,K)}) is a homomorphism of algebras. \par Fix normalized Haar measure $dk$ on $K.$ Then each $\gf \in C(K)$ defines a distribution $\gf\,dk $ in $\E'_K(G),$ given by $$ \inp{\gf\, dk}{f} = \int_K \gf(k) f(k) \, dk. $$ \par For a given representation $\gamma\in\dK,$ we define the distribution $\ga_\gg \in \H$ by $\ga_\gg = \dim(\gg) \chi_{\gg^\vee}\, dk,$ where $\chi_{\gg^\vee}$ denotes the character of the contragredient $\gg^\vee$ of $\gg.$ Moreover, for $\theta \subset \dK$ a finite subset, we define the element $\ga_\theta \in \H$ by $$ \alpha_\theta\coleq \sum_{\gg\in \theta} \ga_\gg. $$ The functions $\alpha_\theta,$ viewed as elements of $\H,$ will later be seen to define an approximation of the identity in $\H.$ \par Let $\theta_1, \theta_2 \subset \dK$ be finite subsets. We agree to write $\cE'(G,K)_{\theta_1\theta_2}$ for the space of distributions $\gf \in \cE'(G,K)$ satisfying $$ \ga_{\theta_1} * \gf * \ga_{\theta_2} = \gf. $$ Then $\cE'(G,K)_{\theta_1\theta_2}$ consists of the distributions in $\cE'(G,K)$ of left $K$-type in $\theta_1$ and of right $K$-type in $\theta_2^\vee = \{\gg^\vee \mid \gg \in \theta_2\}.$ Similarly, we write \begin{eqnarray*} C^\infty_c(G,K)_{\theta_1\theta_2} & \coleq & C_c^\infty(G,K)\; \cap\; \cE'(G,K)_{\theta_1\theta_2},\;\; \text{and}\\ \H_{\theta_1\theta_2} &\coleq &\H \; \cap\; \cE'(G,K)_{\theta_1\theta_2}. \end{eqnarray*} \par so that, for all admissible $G$-representations $(\pi,V)$ and $(\tau,U)$, $$ \Hom(U, V)_{\theta_1\theta_2} \simeq \Hom(U_{\theta_2}, V_{ \theta_1}), $$ naturally. Viewing $\Hom(U, V)_{K\times K} \simeq V_K \otimes (U^*)_K$ as a $(\fg\times \fg, K \times K)$-module in a natural way, we see that $$ \Hom(U, V)_{\theta_1\theta_2} = \Hom(U, V)_{\theta_1\otimes \theta_2^\vee}. $$ In particular, it is readily seen that $$ \pi(\cE'(G,K)_{\theta_1\theta_2}) \subset \End{V}_{\theta_1\theta_2} \simeq \End{V_K}_{\theta_1\theta_2}. $$ Here we note that $\End{V_K}_{K \times K} \subset \End{V},$ naturally. \begin{prop}\label{end-hecke} For any admissible representation $(\pi,V)$ of $G$, $$ \pi(C_c^\infty(G,K))=\pi(\E'(G,K))=\pi(\H). $$ \end{prop} \begin{proof} We denote the three given subspaces of $\End{V_K}_{K\times K}$ by $E_\infty,$ $E_{\E'}$ and $E_\H$ respectively. Let $\gb$ be the $K$-equivariant bilinear form on $\End{V_K}_{K \times K}$ given by $$ \gb(A, B) = \tr (A \after B). $$ By admissibility of $\pi$ it follows that $\gb$ defines a non-degenerate pairing, which is perfect when restricted to $\End{V_K}_{\theta\theta},$ for $\theta$ any finite subset of $\dK$. Therefore, it suffices to show that the $\gb$-orthocomplements $E_\infty^\perp,$ $E_{\cE'}^\perp$ and $\H^\perp$ are equal. Thus, let $T\in \End{V_K}_{K\times K}$. Then it suffices to show that the following assertions are equivalent: \begin{enumerate} \item[(i)] $\tr{T\pi(x)}=0$, for all $x\in G$; \item[(ii)] $\tr{T\pi(f)}=0$, for all $f\in C_c^\infty(G,K)$; \item[(iii)] $\tr{T\pi(f)}=0$, for all $f\in \E'(G,K)$; \item[(iv)] $\tr{T\pi(f)}=0$, for all $f\in \H.$ \end{enumerate} Obviously, (i) implies (ii). By density of $C_c^\infty(G,K)_{\theta\theta}$ in $\cE'(G,K)_{\theta\theta},$ for every finite subset $\theta\subset \dK,$ if follows that (ii) implies (iii). Moreover, (iii) implies (iv). We will finish the proof by showing that (iv) implies (i). \par Assume (iv). For $x\in G$, we define $M(x)\coleq \tr{T\pi(x)}.$ By admissibility and $K\times K$-finiteness of $T$ it follows that $M$ is an analytic function on $G.$ From (iv) it follows that $\langle f,M\rangle=0$ for any $f\in\H.$ Fix a finite subset $\theta \subset \dK$ such that $M \in C_c^\infty(G,K)_{\theta\theta}.$ Then $M = \ga_\theta * M * \ga_\theta.$ Let $u\in U(\g)$, and set $$ f\coleq \alpha_\theta\ast(L_u\delta_e)\ast\alpha_\theta\in\H, $$ where $\delta_e$ is the Dirac measure at the unit element $e$ of $G.$ Then $$ 0=\langle f,M\rangle=\langle L_u\delta_e,M\rangle=(L_{\check{u}}M)(e). $$ By analyticity this implies that $M$ vanishes on the identity component $G_0$ of $G.$ By $K$-stability of $\H$, we deduce that $M$ vanishes on $KG_0.$ Since $G$ is of the Harish-Chandra class, this means that $M=0$ on $G.$ \end{proof} \par \begin{cor}\label{end-hecke-ktype} Let $(\pi,V)$ be an admissible representation of $G$ and assume that $\theta_1$ and $\theta_2$ are finite subsets of $\dK$. Then $$ \begin{array}{rcl} \pi(C_c^\infty(G,K)_{\theta_1\theta_2}) & = & \pi(\cE'(G,K)_{\theta_1\theta_2})\\ & = & \pi(\H_{\theta_1\theta_2})\\ & = & \pi(\H)\cap\End{V_K}_{\theta_1\theta_2} \end{array} $$ \end{cor} \begin{proof} This follows from Proposition \ref{end-hecke} by using $K$-equivariant projections. \end{proof} \section{Holomorphic families of representations and their derivatives}\label{s: holomorphic families} Let $\fv$ be a finite dimensional real linear space. For any open subset $\Omega$ of its complexification $\fv_\iC$, we denote by $\O(\Omega)$ the space of holomorphic $\C$-valued functions on $\Omega$, endowed with the topology of uniform convergence on compact subsets. \par For $\mu\in\fv_\iC,$ we denote by $\O_\mu$ the algebra of germs at $\mu$ of holomorphic functions defined on a neighborhood of $\mu.$ For any $\Omega$ as above, $\mu\in\Omega$ and $f \in \O(\Omega)$, the germ of $f$ at $\mu$ is denoted by $\gamma_\mu(f)\in\O_\mu.$ \par Let $\cP = \cP(\fv_\iC)$ denote the algebra of polynomial functions $\fv_\iC \to \C.$ Then the map $p \mapsto \gg_0(p)$ is an embedding of algebras, $\cP \hookrightarrow \cO_0.$ Accordingly, we shall view $\cP$ as a subalgebra of $\cO_0.$ \par The ring $\cO_0$ is local; its unique maximal ideal $\cM$ consists of the elements vanishing at $0.$ An ideal $\cI \ideal \cO_0$ is said to be cofinite if the quotient $\cO_0/\cI$ is finite dimensional as a vector space over $\C.$ For $k \in \N,$ let $\cP_k$ denote the space of polynomial functions $\fv_\iC \to \C$ of degree at most $k.$ Then \begin{equation} \label{e: deco cO} \cO_0 = \cP_k \oplus \cM^{k+1}. \end{equation} Therefore, the ideal $\cM^{k+1}$ is cofinite in $\cO_0.$ \begin{lemma} \label{l: cofinite ideals} Let $\cI$ be an ideal in $\cO_0.$ Then the following assertions are equivalent: \begin{enumerate} \itema the ideal $\cI$ is cofinite; \itemb there exists a $k \in \N$ such that $\cM^{k+1} \subset \cI.$ \end{enumerate} \end{lemma} \begin{proof} As $\cM^{k+1}$ is cofinite, (b) implies (a). Conversely, assume (a). The space $V = \cO_0/\cI$ is a finite dimensional vector space, and an $\cO_0$-module for left multiplication. The associated algebra homomorphism $\cO_0 \to \End{V}$ is denoted by $\gL.$ As $\cO_0$ is a commutative algebra and $V$ is finite dimensional, there exists a positive integer $p$ such that the module $V$ decomposes as a finite direct sum of generalized weight spaces $$ V_\chi\coleq \bigcap_{f \in \cO_0} \ker (\gL(f) - \chi(f) {\rm id}_V)^p $$ with $\chi \in \widehat \cO_0 \coleq \Hom(\cO_0, \C).$ Since $\cO_0$ is a local ring with maximal ideal $\cM,$ the set $\widehat \cO_0$ of characters consists of the single element $\chi_0: g \mapsto g(0).$ It follows that $V = V_{\chi_0},$ so that $(f - f(0))^p \in \cI$ for all $f \in \cO_0.$ In particular, $f^p \in \cI$ for all $f \in \cM.$ As an ideal, $\cM$ is generated by $n$ elements. Hence, $\cM^{k+1} \subset \cI,$ for $k \geq np -1.$ \end{proof} \subsection{The derivation process}\label{ss: derivation process} For each vector $X \in \fv$ we denote by $\partial_X$ the first order differential operator given by $\partial_X \gf (a) = \frac{d}{dz} \gf(a + z X)|_{z = 0}$ for $a \in \fv_\iC$ and $\gf$ a holomorphic function defined on a neighborhood of $a$ in $\fv_\iC.$ The map $X \mapsto \partial_X$ has a unique extension to an algebra isomorphism $u \mapsto \partial_u$ from the symmetric algebra $S(\fv)$ of $\fv_\iC$ onto the algebra of constant coefficient (holomorphic) differential operators on $\fv_\iC.$ We will follow Harish-Chandra's convention to write \begin{equation} \label{e: HC notation for action of Sv} \gf(a; u) \coleq \partial_u \gf (a), \end{equation} for $\gf$ a holomorphic function defined on a neighborhood of $a.$ We define the pairing $\langle\,\cdot\, ,\,\cdot\,\rangle$ between $\O_0$ and $S(\fv)$ by \begin{equation}\label{sec3} \begin{array}{rcl} \O_0\times S(\fv) & \rightarrow & \C\\ (\varphi,u) & \mapsto & \varphi(0;u). \end{array} \end{equation} For a given cofinite ideal $\cI \ideal \cO_0,$ let $S_\cI(\fv)$ denote the annihilator of $\cI$ in $S(\fv)$ relative to this pairing. For $k\in\N$, let $S_k(\fv)$ be the linear subspace of $S(\fv)$ consisting of the elements of order at most $k$. \begin{lemma}\label{sec4} Let $k \in \N.$ Then \begin{enumerate} \itema $S_k(\fv)=S_{\cM^{k+1}}(\fv);$ \itemb the pairing (\ref{sec3}) induces a perfect pairing $(\O_0/\cM^{k+1})\times S_k(\fv) \to \C.$ \end{enumerate} \end{lemma} \begin{proof} The pairing $\langle\,\cdot\, ,\,\cdot\,\rangle$, defined in \eqref{sec3}, vanishes on $\cM^{k+1}\times S_k(\fv)$. Thus, $S_k(\fv)\subset S_{\cM^{k+1}}(\fv).$ \par From the decomposition $\O_0=\cM^{k+1}\oplus \P_k$, and non-degeneracy of the pairing, it follows that $S_{\cM^{k+1}}(\fv)\hookrightarrow \P_k^*.$ In particular, the dimension of $S_{\cM^{k+1}}(\fv)$ does not exceed the dimension of $\P_k,$ which in turn equals the dimension of $S_k(\fv).$ Assertion (a) now follows. It also follows that the induced embedding $S_k(\fv) = S_{\cM^{k+1}}(\fv)\hookrightarrow \P_k^* \simeq (\cO_0/ \cM^{k+1})^*$ is an isomorphism onto. By finite dimensionality, this implies assertion (b). \end{proof} \begin{lemma}\label{sec5} Let $\cI$ be a cofinite ideal of $\O_0$. Then the pairing (\ref{sec3}) induces a linear isomorphism $S_\cI(\fv)\simeq (\O_0/\cI)^*.$ \end{lemma} \begin{proof} By Lemma \ref{l: cofinite ideals} there exists a $k \in \N$ such that $\cM^{k+1} \subset \cI.$ This inclusion induces an embedding of $(\O_0/\cI)^*$ into $(\O_0/\cM^{k+1})^*.$ In view of Lemma \ref{sec4}, it follows that the pairing induces an embedding $(\O_0/\cI)^* \hookrightarrow S(\fv).$ Its image is contained in the annihilator $S_\cI(\fv),$ by definition of the latter. On the other hand, the pairing induces an inclusion $S_\cI(\fv)\hookrightarrow\O_0^*$ and elements of $S_\cI(\fv)$ vanish on $\cI$, so that $S_\cI(\fv)\hookrightarrow (\O_0/\cI)^*.$ The result follows. \end{proof} For $\mu \in \fv_\iC$ we denote by $T_\mu$ the translation in $\fv_\iC$ given by $ \nu \mapsto \nu + \mu.$ We note that the pull-back map $T_\mu^*: \gf \mapsto \gf \after T_\mu$ induces a ring isomorphism from $\cO_\mu$ onto $\cO_0.$ Let $\cI\ideal \O_0$ be a cofinite ideal and let $\Omega$ be an open subset of $\fv_\iC$. Then for every $f\in\O(\Omega)$ and each $\mu\in\Omega$ we define \begin{equation}\label{e: JI} J_\cI f(\mu)\coleq \mathrm{pr}_\cI(\, \gamma_0( T_\mu^*f)\, )\;\in \cO_0/\cI, \end{equation} where $\mathrm{pr}_\cI$ denotes the projection of $\cO_0$ onto $\cO_0/\cI.$ In the following lemma, which is a straightforward consequence of the definitions, $\langle\, \cdot\,,\,\cdot\, \rangle$ denotes the pairing induced by (\ref{sec3}), see Lemma \ref{sec5}. \begin{lemma}\label{sec7 Let $f \in \cO(\Omega).$ Then for all $\mu \in \Omega,$ $$ \langle J_\cI f(\mu),u\rangle = f(\mu;u),\qquad u\in S_{\cI}(\fv). $$ \end{lemma} \begin{cor \label{c: sec7} The map $f\mapsto J_\cI f$ defines a continuous algebra homomorphism from $\O(\Omega)$ to $\O(\Omega,\O_0/\cI)$. \end{cor} \begin{proof} Let $u\in S(\fv).$ For every $f\in\O(\Omega)$, the function $\partial_u f = f(\cdot;u)$ belongs to $\O(\Omega).$ Moreover, the map $\partial_u$ is a continuous linear endomorphism of $\cO(\Omega).$ In view of Lemma \ref{sec7}, it follows that for each $\xi \in (\cO_0/I)^*$ the map $f \mapsto \xi \after [J_\cI(f)]$ is a continuous linear endomorphism of $\cO(\Omega).$ By finite dimensionality of $\cO_0/\cI,$ it follows that $J_\cI$ is continuous. The assertion that $J_\cI$ is an algebra homomorphism follows from combining the observations that $T_\mu^*,$ $\gg_0,$ and $\pr_{\cI}$ are algebra homomorphisms. \end{proof} \begin{ex} \rm Let $\xi\in\fv_\iC^*$. Denote by $e^\xi$ the holomorphic function on $\fv_\iC$ given by: $$ e^\xi(\mu)\coleq e^{\xi(\mu)},\quad \mu\in\fv_\iC. $$ In terms of the canonical identification of the symmetric algebra $S(\fv)$ with the algebra $\P(\fv_\iC^*)$ of polynomial functions on $\fv_\iC^*$ we have $ \partial_u e^\xi = u(\xi) e^\xi,$ for $u \in S(\fv).$ Hence, if $\cI$ is a cofinite ideal in $\cO_0,$ then for all $\mu \in \fv_\iC$ and $u \in S_\cI(\fv),$ \vspace{1mm} \begin{enumerate} \item[(i)] $ \langle\, J_\cI e^\xi(\mu) \, , \, u \, \rangle = u(\xi) \, e^{\xi(\mu)},$ \vspace{1mm} \item[(ii)] $\;J_\cI e^\xi \,(\mu) = e^{\xi(\mu)}\, \mathrm{pr}_\cI \circ\gamma_0(e^\xi).$ \end{enumerate} \end{ex} \begin{definition}\label{d: diffD \rm Let $E$ be a finite dimensional $\O_0$-module. For every $f\in\O(\Omega)$ and all $\mu\in\Omega$, we define $f^{(E)}(\mu)\in\mathrm{End}(E)$ by: $$ f^{(E)}(\mu)e\coleq \gamma_0(T_\mu^*f)\cdot e, \quad e\in E. $$ \end{definition} \begin{ex \rm If $E=\O_0/\cI$ for some cofinite ideal $\cI$ of $\O_0$, then, for any $f\in\O(\Omega),$ $$ J_\cI f(\mu)= f^{(E)}(\mu)(1+\cI), \qquad \mu\in\Omega. $$ \end{ex} \medbreak\medbreak Let $A$ be an algebra and $E$ an $A$-module. We denote by $\mathrm{ann}_A(E)$ the annihilator of $E$ in $A,$ i.e., the kernel of the natural algebra homomorphism $A\rightarrow\mathrm{End}(E)$. If $E$ is finite dimensional, $\mathrm{ann}_A(E)$ is a cofinite ideal of $A$. \begin{lemma}\label{sec13 Let $\Omega \subset \fv_\iC$ be open, and let $E$ be a finite dimensional $\cO_0$-module. Then for every $f \in \cO(\Omega)$ and all $\mu \in \Omega,$ \vspace{1mm} \begin{enumerate} \itema $ f^{(E)}(\mu)e= J_{\mathrm{ann}_{\O_0}(E)}f(\mu)\cdot e,\quad \text{for all} e \in E; $ \vspace{1mm} \itemb $ (T_\mu^* f)^{(E)} = T_\mu^* f^{(E)}. $ \end{enumerate} \end{lemma} \begin{proof These formulas follow by straightforward computation. \end{proof} \begin{lemma \label{l: dual of End E} Let $E$ be a finite dimensional $\cO_0$-module and let $\eta \in \Endo(E)^*.$ Then there exists an element $u = u_\eta \in S(\fv)$ such that $$ \eta \after f^{(E)} = \partial_u f, $$ for every open $\Omega \subset \fv_\iC$ and all $f \in \cO(\Omega).$ \end{lemma} \begin{proof} It suffices to prove this for $\eta = e^* \otimes e,$ with $e^* \in E^*$ and $e \in E.$ Then, for $f \in \cO(\Omega)$ and $\gl \in \Omega,$ \begin{equation}\label{e: linear-f as diff} \eta \after f^{(E)}(\gl) = e^*(\gg_0(T^*_\gl f) \cdot e). \end{equation} Let $\cI$ be the (cofinite) annihilator of $e$ in $\cO_0.$ Then the linear functional $L: \gf \mapsto e^*(\gf \cdot e)$ on $\cO_0$ factors through a linear map $\cO_0/\cI \to \C.$ Hence, in view of Lemma \ref{sec5}, there exists an element $u \in S_\cI(\fv)$ such that $L(\gf) = \partial_u \gf(0)$ for all $\gf \in \cO_0.$ It follows that the expression on the right-hand side of \eqref{e: linear-f as diff} equals $\partial_u f(\gl).$ \end{proof} We shall also need a kind of converse to the above lemma. \begin{lemma \label{l: E associated with u} Let $F \subset S(\fv)$ be a finite subset. Then there exists a finite dimensional $\cO_0$-module $E,$ and linear functionals $\eta_u \in \Endo(E)^*,$ for $u \in F,$ such that $$ \eta_u \after f^{(E)} = \partial_u f $$ for every $u \in F,$ every open $\Omega \subset \fv_\iC$ and all $f \in \cO(\Omega).$ \end{lemma} \begin{proof By taking direct sums of finite dimensional $\cO_0$-modules we may reduce to the case that $F$ consists of a single element $u \in S(\fv).$ Let $k$ be the order of $u.$ Then $\gf \mapsto \gf(0;u)$ defines a linear functional $e^*$ on $\cO_0/\cI,$ for $\cI = \cM^{k+1}.$ We put $E = \cO_0/\cI$ and let $e$ denote the image of $1\in\cO_0$ in $E.$ Let $\eta = e^*\otimes e$ be the linear functional on $\Endo(E)$ defined by $T \mapsto e^*(Te).$ Then for all $f \in \cO(\Omega)$ and all $\gl \in \Omega$ we have $$ \eta \after f^{(E)}(\gl) = \eta(\gg_0 (T_\gl^* f)) = e^* (\gg_0 (T_\gl^* f)\cdot e) = \partial_u (\gg_0(T_\gl^*f))(0) = \partial_u f(\gl). $$ \end{proof} We retain the assumption that $\Omega$ is an open subset of $\fv_\iC.$ \begin{cor \label{c: continuity of diffD} Let $E$ be a finite dimensional $\O_0$-module. Then $f\mapsto f^{(E)}$ is a continuous algebra homomorphism from $\O(\Omega)$ to $\O(\Omega,\mathrm{End}(E))$. \end{cor} \begin{proof The map is an algebra homomorphism by Lemma \ref{sec13} (a) and Corollary \ref{c: sec7}. The continuity is an immediate consequence of Lemma \ref{l: dual of End E}. \end{proof} We agree to use the following notation for the map of Corollary \ref{c: continuity of diffD}, \begin{equation} \label{e: defi diffD} \diffD^{(E)} : \;f \mapsto f^{(E)}, \quad \cO(\Omega) \to \cO(\Omega, \End{E}). \end{equation} The following property is an immediate consequence of the definitions; here we keep in mind that $\End{E} \oplus \End{F} \embeds \End{E \oplus F},$ naturally. \begin{property Let $E$, $F$ be two finite dimensional $\O_0$-modules. Then for every $f\in\O(\Omega),$ $$ f^{(E\oplus F)}=f^{(E)}\oplus f^{(F)}. $$ \end{property} To prepare for deriving more properties of the map $\diffD^{(E)},$ we formulate a few results on finite dimensional $\cO_0$-modules. \begin{lemma Let $E$ be a finite dimensional $\O_0$-module. Then $E$ is cyclic if and only if there exists a cofinite ideal $\cI$ of $\O_0$ such that $E\simeq\O_0/\cI$. \end{lemma} \begin{proof Straightforward. \end{proof} \begin{cor}\label{sec18 Let $E$ be a finite dimensional $\O_0$-module. Then there exist finitely many cofinite ideals $\cI_1, \ldots, \cI_n$ of $\cO_0$ such that $E$ is a quotient of the direct sum $\cO_0/ \cI_1 \oplus \cdots \oplus \cO_0 /\cI_n$ of $\cO_0$-modules. \par In particular, there exist $k,N \in \N$ such that $E$ is a quotient of the $\cO_0$-module $(\O_0/\cM^{k+1})^N$ for some $k,N \in\N.$ \end{cor} \begin{proof The first assertion results from the previous lemma. The second follows from Lemma \ref{l: cofinite ideals}. \end{proof} Besides the decomposition (\ref{e: deco cO}), we have the following decomposition of $\cP = \cP(\fv_\iC),$ for $k \in \N,$ \begin{equation} \label{e: deco cP} \P = \P_k\oplus (\cM\cap\P)^{k+1}. \end{equation} Hence, the embedding $\cP \hookrightarrow \cO_0$ induces, for each $k \in \N,$ an isomorphism of algebras \begin{equation}\label{sec23} \iota_k: \;\;\P/(\cM\cap\P)^{k+1}\stackrel{\sim}{\longrightarrow}\O_0/\cM^{k+1}. \end{equation} It follows that $\cP/(\cM \cap \P)^{k+1}$ is a local ring, with unique maximal ideal equal to $(\cM \cap \cP)/(\cM \cap \cP)^{k+1}.$ Thus, if $\widetilde \cI \subset \cP$ is an ideal with $(\cM \cap \cP)^{k+1} \subset \widetilde\cI,$ then $\widetilde \cI \subset \cM.$ \begin{lemma}\label{sec24 Let $\widetilde{\cI}$ be an ideal of $\P.$ Then the following assertions are equivalent: \begin{enumerate} \itema there exists a $k \in \N$ such that $(\cM \cap \cP)^{k + 1} \subset \widetilde \cI;$ \itemb there exists an ideal $\cI\;\lhd\;\O_0 $ of finite codimension, such that $\widetilde{\cI}=\cI\cap\P.$ \end{enumerate} If any of these conditions is fulfilled, then the ideal $\cI$ in (b) is unique and the embedding of $\cP$ into $\cO_0$ induces an isomorphism of algebras $\cP/\widetilde \cI \to \cO_0 / \cI.$ \end{lemma} \begin{proof Assume (a). The image $\widetilde \cI'$ of $\widetilde \cI$ in $\cP/(\cM \cap \cP)^{k+1}$ is an ideal. Its image $\iota_k(\tilde \cI')$ is an ideal of $\cO_0/ \cM^{k+1}.$ Let $\cI$ be the preimage of $\iota_k(\tilde\cI')$ in $\cO_0.$ Then the following diagram commutes $$ \xymatrix@1{ \cP/(\cM \cap \cP)^{k+1}\;\;\ar[d]_p\ar[r]^{\iota_k} & \;\;\ar[d]^q\cO_0/\cM^{k+1}\\ \cP/\widetilde\cI\; \ar[r]^{\bfi} & \;\; {\cO_0/\cI}. } $$ Here $\bfi$ is induced by the inclusion map $\cP\hookrightarrow \cO_0.$ The kernel of $p$ equals $\widetilde \cI'$ and the kernel of $q$ equals $\iota_k(\widetilde \cI').$ As $\iota_k$ is an isomorphism of algebras, it follows that $\bfi$ is an isomorphism of algebras, and (b) is immediate. Conversely, assume (b). Let $\cI_j \ideal \cO_0$ be ideals such that $\widetilde \cI = \cI_j \cap \cP,$ for $j =1, 2.$ We will complete the proof by showing that (a) holds and that $\cI_1 = \cI_2.$ Since $\cI_1$ and $\cI_2$ are cofinite, there exists a constant $k \in \N$ such that $\cM^{k+1}\subset \cI_j,$ for both $j =1,2.$ Therefore, $$ (\cM \cap \cP)^{k+1} \subset (\cM^{k+1} \cap \cP ) \subset \widetilde \cI $$ and (a) follows. Moreover, for each $j=1,2$ we have the following commutative diagram: $$ \xymatrix@1{ \cP/(\cM \cap \cP)^{k+1}\;\;\ar[d]_{p}\ar[r]^{\iota_k} & \;\;\ar[d]^{q_j}\cO_0/\cM^{k+1}\\ \cP/\widetilde\cI\; \ar[r]^{\bfi_j} & \;\; {\cO_0/\cI_j}. } $$ From the assumption on $\cI_j$ it follows that the map $\bfi_j$ is injective. Moreover, since $\iota_k$ and $q_j$ are surjective, it follows that $\bfi_j$ is an isomorphism of algebras, for $j =1, 2.$ This implies that $\iota_k(\ker p) = \ker (q_j).$ Since $\cI_j$ equals the preimage of $\ker(q_j)$ in $\cO_0,$ for $j =1,2,$ it follows that $\cI_1 = \cI_2.$ \end{proof} \par Let $\FMod_{\O_0}$ denote the category of finite dimensional $\O_0$-modules and $\FMod_\P$ the category of finite dimensional $\P$-modules $E$ for which there exists a $k \in \N$ such that $$ (\cM\cap \cP)^{k+1} \subset \mathrm{ann}_\P(E). $$ If $E,F$ belong to $\FMod_\P$ then the $\FMod_\P$-morphisms are defined to be the $\cP$-module homomorphisms $E \to F.$ We observe that their kernels and images belong to the category $\FMod_\P$ as well. \par If $E$ is a finite dimensional $\cO_0$-module, then its annihilator $\cI$ is cofinite. Since $\cP \hookrightarrow \O_0,$ the space $E$ is a finite dimensional $\cP$-module as well and its annihilator $\widetilde \cI \coleq \ann_\cP(E)$ in $\cP$ is given by $\widetilde \cI = \cI \cap \cP.$ Furthermore, by Lemma \ref{l: cofinite ideals} there exists a $k \in \N$ such that $$ (\cM \cap \cP )^{k+1} \subset \ann_\cP(E). $$ We conclude that there is a well-defined forgetful functor $\cF: \FMod_{\cO_0} \to \FMod_\cP.$ \begin{lemma}\label{sec25 The forgetful functor $\cF: \FMod_{\cO_0} \to \FMod_\cP $ is an isomorphism of categories. \end{lemma} \begin{proof Let $E$ be a non-trivial $\cP$-module from the category $\FMod_\cP.$ We will first show that $E$ carries a unique compatible structure of $\cO_0$-module. \par The ideal $\widetilde \cI = \mathrm{ann}_\P(E)$ satisfies condition (a) of Lemma \ref{sec24}, hence equals $\cP \cap \cI$ for a unique cofinite ideal $\cI \ideal \cO_0.$ Via the isomorphism $\cP/\widetilde \cI \simeq \cO_0/\cI$ we equip $E$ with the structure of $\cO_0$-module; this establishes existence. \par For uniqueness, assume that $E$ is equipped with a compatible structure of $\cO_0$-module. Let $\cI'$ be the annihilator of $E$ in $\cO_0.$ Then $\cI'$ is cofinite and, by compatibility, $\widetilde \cI = \cI' \cap \cP.$ By the first part of the proof, we see that $\cI' = \cI.$ Furthermore, by compatibility it follows that the diagram $$ \xymatrix{ \ar[d] \cP/\widetilde \cI \ar[rd] & \\ \cO_0/\cI \ar[r] & \End{E} } $$ commutes. The vertical arrow is induced by the inclusion $\cP \to \cO_0$ and represents an isomorphism. This establishes uniqueness. \par It follows from these considerations that each submodule of $E$ in the category $\FMod_\P$ is a submodule for the associated compatible structure of $\cO_0$-module as well. \par For $E$ a finite dimensional $\cP$-module from $\FMod_\P,$ let $\cG(E)$ denote the same space with the uniquely defined compatible structure of $\cO_0$-module. If $E_0$ is a submodule in the category $\FMod_\P,$ then by what we said before, $\cG(E_0)$ is an $\cO_0$-submodule of $\cG(E).$ The quotient $\cG(E)/\cG(E_0)$ is an $\cO_0$-module in a natural way, whose $\cO_0$-module structure is compatible with the $\cP$-module structure of $E/E_0.$ Hence, $\cG(E)/\cG(E_0) = \cG(E/E_0).$ It now follows that every morphism $f: E \to F$ in the category $\FMod_\P$ is an $\cO_0$-module homomorphism $\cG(E) \to \cG(F).$ We thus see that $\cG$ defines a functor $\FMod_\P \to \FMod_{\cO_0},$ which obviously is a two-sided inverse to $\cF.$ \end{proof} \begin{cor For every pair $E,\,F$ of finite dimensional $\cO_0$-modules, $$\mathrm{Hom}_{\O_0}(E,F)=\mathrm{Hom}_\P(E,F).$$ \end{cor} Our next objective is to consider tensor products in the categories $\FMod_\P$ and $ \FMod_{\cO_0}.$ \par Let $n \in \N^*$ and consider the $n$-fold Cartesian product $\fv_\iC^n$ of $\fv_\iC.$ Projection onto the $j$-th coordinate is denoted by $\pr_j.$ Pull-back by $\pr_j$ defines an embedding of algebras $\pr_j^*: \; p \mapsto p \after \pr_j,$ $\cP \to \cP(\fv_\iC^n).$ The multi-linear map $$ (p_1, \ldots , p_n) \mapsto \prod_{j=1}^n \pr_j^*(p_j) $$ induces an isomorphism of algebras $\cP^{\otimes n } \to \cP(\fv_\iC^n),$ via which we shall identify the elements of these spaces. Accordingly, $$ (p_1 \otimes \cdots \otimes p_n)(\mu_1, \ldots, \mu_n) = p_1(\mu_1) \cdots p_n(\mu_n), $$ for $p_j \in \cP$ and $\mu_j \in \fv_\iC.$ The maximal ideal $\cM_{\cP^{\otimes n}}$ in $\cP(\fv_\iC^n)$ consisting of the polynomials vanishing at $0$ is now given by $\cM_{\cP^{\otimes n}} = \sum_{i=1}^n \cM_{\cP, n, i},$ where $$ \cM_{\cP, n , i} \coleq \cP \otimes \cdots \otimes \cP \otimes {\buildrel i \over {\overbrace{(\cM \cap \cP)}} }\otimes \cP \otimes \cdots \otimes \cP. $$ \begin{lemma} \label{l: tensor products and cC Let $E_1, \ldots, E_n$ be finite dimensional $\cP$-modules from the category $\FMod_\P.$ Then $E_1 \otimes \cdots \otimes E_n$ is a $\cP^{\otimes n}$-module from the category $\FMod_{\P^{\otimes n}}.$ \end{lemma} \begin{proof There exists a $k \in \N$ such that $(\cM\cap\cP)^{k+1}$ annihilates each of the modules $E_i,$ for $1 \leq i \leq n.$ It is now readily checked that $$ (\cM_{\cP^{\otimes n}})^{n (k+1)}\subset \ann_{\cP^{\otimes n}}(E_1 \otimes \cdots \otimes E_n). $$ \end{proof} We consider the map $$ \alpha_n:\;\; \fv_\iC^{ n} \rightarrow \fv_\iC, \quad (\mu_1,\dots,\mu_n) \mapsto \mu_1 + \cdots +\mu_n. $$ Pull-back by $\alpha_n$ induces an algebra homomorphism \begin{equation}\label{alphaene} \alpha_n^*:\;\P\longrightarrow\P^{\otimes n}. \end{equation} \vspace{2mm} \begin{lemma The homomorphism $\ga_n^*$ maps the maximal ideal $\cM\cap \cP$ of $\cP$ into the maximal ideal $\cM_{\cP^{\otimes n}}$ of $\cP^{\otimes n}.$ \end{lemma} \begin{proof The ideal $\cM\cap \cP$ is generated by the first order polynomials $\xi \in \fv_\iC^*.$ Now \begin{equation} \label{e: ga n on xi} \ga_n^*(\xi) = \sum_{i = 1}^n 1 \otimes \cdots \otimes 1 \otimes {\buildrel i \over {{\xi}}} \otimes 1 \otimes \cdots \otimes 1, \end{equation} and the result follows. \end{proof} \begin{cor}\label{c: sec28 Let $n\in\N^*$. Via the homomorphism (\ref{alphaene}), every finite dimensional $\P^{\otimes n}$-module from the category $\FMod_{\P^{\otimes n}}$ becomes a $\cP$-module from the category $\FMod_\P.$ \end{cor} \begin{proof Let $E$ be a non-trivial module from the category $\FMod_{\P^{\otimes n}}.$ Then there exists a $k \in \N$ such that $$ (\cM_{\cP^{\otimes n}})^{k+1} \subset \ann_{\cP^{\otimes n}}(E). $$ By application of the previous lemma, we obtain $$ \ga_n^*((\cM\cap \cP)^{k+1}) \subset \ann_{\cP^{\otimes n}}(E). $$ The $\P$-module structure on $E$ is defined by: $$ \xymatrix{ \P\ar[r]^{\alpha_n^*}&\P^{\otimes n}\ar[r]&\End{E}. } $$ From this we see that $ \mathrm{ann}_\P(E) = \alpha_n^{*-1}(\ann_{\P^{\otimes n}}(E)) $ and we infer that $$ (\cM\cap \cP)^{k+1} \subset \ann_{\cP}(E). $$ \end{proof} \begin{rem} \label{r: tensor products in cC \rm If $E$ is a finite dimensional $\cP$-module, we define $$ m_E:\;\P \; \longrightarrow\; \End{E} $$ by $m_E(p)e\coleq p\cdot e,$ for $p \in \cP$ and $e\in E$. If $E_1, \cdots, E_n$ are finite dimensional modules from the category $\FMod_\P,$ then by combining Lemma \ref{l: tensor products and cC} and Corollary \ref{c: sec28} we may equip the tensor product $E_1 \otimes \cdots \otimes E_n$ with the structure of a module from the same category $\FMod_\P.$ The module structure is given by the rule \begin{equation} \label{e: rule for multiplication} m_{E_1 \otimes \cdots \otimes E_n} = (m_{E_1} \otimes \cdots \otimes m_{E_n})\after \ga_n^*. \end{equation} In view of (\ref{e: ga n on xi}) this module structure is completely determined by the rule \begin{eqnarray} \lefteqn{ m_{E_1 \otimes \cdots \otimes E_n}(\xi) =}\nonumber\\ \label{e: m on xi} & = & \sum_{i=1}^n \id_{E_1} \otimes \cdots \otimes \id_{E_{i-1}} \otimes m_{E_i}(\xi) \otimes \id_{E_{i+1}} \otimes \cdots \otimes \id_{E_n} \end{eqnarray} for $\xi \in \fv_\iC^*\subset \cP.$ \par Accordingly, if $E_1, \ldots, E_n$ are finite dimensional $\cO_0$-modules, then in particular they are $\cP$-modules from the category $\FMod_\P.$ We equip the module $E_1 \otimes \cdots \otimes E_n$ from $\FMod_\P$ with the unique compatible structure of $\cO_0$-module. \end{rem} \begin{lemma}\label{sec34 Let $E_1,\ldots, E_n$ be finite dimensional $\O_0$-modules. Then, as $\O_0$-modules: $$ E_1\otimes \cdots \otimes E_n\simeq E_1\otimes(E_2\otimes\cdots \otimes E_n). $$ \end{lemma} \begin{proof In view of Remark \ref{r: tensor products in cC}, it suffices to establish the identity as an identity of $\cP$-modules. Thus, it suffices to show that \begin{equation} m_{{E_1} \otimes \cdots \otimes {E_n}} = m_{{E_1} \otimes ({E_2\otimes\cdots \otimes E_n})}. \end{equation} Since $\fv_\iC^*$ generates $\cP,$ it suffices to check this identity on any element $\xi \in \fv_\iC^*.$ This is easily done by using the identity (\ref{e: m on xi}). \end{proof} We return to the setting of an open subset $\Omega \subset \fv_\iC$ and resume our study of the map $\diffD^{(E)}: f \mapsto f^{(E)},$ $\cO(\Omega) \to \cO(\Omega , \End{E}),$ introduced in (\ref{e: defi diffD}), for $E$ a finite dimensional $\cO_0$-module. \begin{lemma \label{l: restriction diffD to cP} The restriction to $\P$ of the map $\diffD^{(E)}$ is given by $$ \diffD^{(E)}|_\cP = (\id_\P\otimes m_E)\circ\alpha_2^*. $$ In particular, this restriction maps $\P$ into $\P\otimes \End{E}$. \end{lemma} \begin{proof Let $p\in\P$. For any $\mu\in\fv_\iC$, we have $T_\mu^*p\in\P,$ and for any $e\in E,$ $$ p^{(E)}(\mu)e = (T_\mu^*p)\cdot e = m_E(T_\mu^*p)e. $$ Hence, $$ (\ev_\mu \otimes \id_{\End{E}})\after \diffD^{(E)}|_\cP = m_E \after T_\mu^*|_\cP. $$ Viewing $\ev_\mu\otimes\id_\P$ as a $\P$-valued function on $\P\otimes\P$, we may identify the map $T_\mu^*$ with $(\ev_\mu\otimes\id_\P)\circ\alpha^*_2$ on $\P.$ Using the relation $$ m_E\circ(\ev_\mu\otimes\id_\P)=(\ev_\mu\otimes\id_{\End{E}})\circ(\id_\P\otimes m_E), $$ we obtain the assertion of the lemma. \end{proof} \begin{prop}\label{sec33 Let $E_1$ and $E_2$ be two finite dimensional $\O_0$-modules. Then, for every $f\in\O(\Omega)$, $$(\diffD^{(E_1)}\otimes \id_{\End{E_2}})f^{(E_2)}=f^{(E_1\otimes E_2)}.$$ \end{prop} \begin{proof By density of $\P$ in $\O(\Omega)$ and continuity of the maps $\diffD^{(E_1\otimes E_2)}$ and $(\diffD^{(E_1)}\otimes \id_{\End{E_2}})\circ \diffD^{(E_2)},$ see Corollary \ref{c: continuity of diffD}, it suffices to establish the validity of the identity on a fixed element $p \in \cP.$ \par By definition of $\diffD^{(E_1)},$ we have $\ev_0 \after \diffD^{(E_1)}q = m_{E_1}(q)$ for each $q \in \cP.$ Using Lemma \ref{l: restriction diffD to cP} we obtain $$ \begin{array}{rcl} (\diffD^{(E_1)}\otimes \id_{\End{E_2}})p^{(E_2)}(0) & = & (m_{E_1}\otimes \id_{\End{E_2}})p^{(E_2)}\\ & = & (m_{E_1}\otimes \id_{\End{E_2}})\circ (\id_{\P}\otimes m_{E_2})(\alpha_2^*p)\\ & = & (m_{E_1}\otimes m_{E_2})\circ\alpha_2^*(p)\\ & = & m_{E_1\otimes E_2}(p) = p^{(E_1\otimes E_2)}(0). \end{array} $$ The result now follows by translation invariance (see Lemma \ref{sec13}(b)). \end{proof} \begin{prop}\label{p: iterateder Let $E_1,\dots,E_n$ be finite dimensional $\O_0$-modules. Then, for every $f\in\O(\Omega)$, $$ \left(\diffD^{(E_1)}\circ\cdots\circ \diffD^{(E_n)}\right)(f)=f^{(E_1\otimes\cdots\otimes E_n)}. $$ \end{prop} In the formulation of this proposition we have slightly abused notation, by using the abbreviation $\diffD^{(E_k)}$ for $\diffD^{(E_k)} \otimes \id_{\End{E_{k+1}}} \otimes \cdots \otimes \id_{\End{E_n}}.$ \begin{proof In view of Lemma \ref{sec34}, the result follows by repeated application of Proposition \ref{sec33}. \end{proof} Let $\lambda\in\fv_\iC$. By $\cI_\lambda$ we denote the subset of $\O_0$ consisting of all elements $\varphi\in\cM$ satisfying \begin{equation}\label{e: idealcodim2} \partial_\lambda \varphi \in \cM. \end{equation} By Leibniz's rule, $\cI_\lambda$ is an ideal of $\cO_0$ containing $\cM^2$. \begin{lemma} \label{l: codimension two Assume $\lambda\neq 0.$ Then $\cI_{\lambda}$ is an ideal in $\cO_0$ of codimension 2. If $\xi \in \fv_\iC^*$ is such that $\xi(\gl) \neq 0,$ then $\cO_0 = \C 1 \oplus \C \xi \oplus \cI_\gl.$ \end{lemma} \begin{proof Fix a basis $\xi_1,\dots,\xi_N$ for $\fv_\iC^*$ such that $\xi_1 = \xi$ and such that $\xi_j(\gl) = 0$ for $j > 1.$ The monomials $$ \xi^\beta\coleq \xi_1^{\beta_1}\cdots\xi_N^{\beta_N}, \qquad \gb\in \N^N, $$ form a basis of the complex vector space $\cP.$ Consider a polynomial $p\in\P$ and write $ p=\sum_\beta c_\beta \xi^\beta, $ with $c_\gb \in \C.$ Then by an easy calculation we find that $$ p(0) = c_0 \quad \textrm{and}\quad \partial_\lambda p(0) = c_{(1,0,\ldots, 0)}\xi_1(\gl). $$ It follows that $p$ belongs to $\cI_\gl$ if and only if $c_\gb = 0$ for $\gb = 0$ and for $\gb = (1,0,\ldots, 0).$ This implies that $\cP = \C 1 \oplus \C \xi_1 \oplus (\cI_\gl \cap \cP).$ By Lemma \ref{sec24} the inclusion map $\cP \to \cO_0$ induces an isomorphism of algebras $\cP/(\cI_\gl\cap \cP) \simeq \cO_0/\cI_\gl.$ All assertions follow. \end{proof} Let $\widetilde \cI_1, \ldots, \widetilde \cI_n$ be a collection of ideals from $\cP$ containing $(\cM \cap \cP)^{k+1}$ for some $k \in \N.$ Then the algebra homomorphism $\ga_n^*: \cP \to \cP^{\otimes n}$ induces an algebra homomorphism $$ \bar\ga_n^* : \cP \to (\cP/\widetilde \cI_1) \otimes \cdots \otimes (\cP/\widetilde \cI_n) $$ by composition with the natural projection from $\cP^{\otimes n}$ onto the quotient algebra on the right. \begin{lemma}\label{sec40 Let $\lambda\in\fv_\iC\setminus \{0\}$, and let $\tilde{\cI} \ideal \cP$ be an ideal with $(\cM\cap\P)^{k+1}\subset\tilde{\cI}$ for some $k\in\N$. Then the kernel of $$ \bar\alpha_2^*:\;\;\P\longrightarrow \P/\tilde{\cI}\otimes(\P/\cI_\lambda\cap\P) $$ is equal to $\{p\in\tilde{\cI}\mid \, \partial_\gl p \in\tilde{\cI}\}.$ In particular, this kernel contains $\cM^{k+2} \cap \cP.$ \end{lemma} \begin{proof The last assertion follows from the first one by application of the Leibniz rule. We turn to the proof of the first assertion. \par We fix a basis $\xi_1 \ldots, \xi_N$ of $\fv_\iC^*$ as in the proof of Lemma \ref{l: codimension two} and adopt the notation of that proof. Let $X_1, \ldots, X_N$ be the dual basis of $\fv_\iC.$ We note that $\gl$ is proportional to $X_1.$ Our first goal is to obtain a suitable formula for $\ga_2^*.$ \par Let $p \in \cP$ and $\mu \in \fv_\iC.$ Using Taylor expansion at $\mu$ we find, for all $\nu\in\fv_\iC$, $$ p(\mu+\nu)=\sum_\beta\frac{p(\mu;{X}^\beta)}{\beta !}\,{\xi}^\beta(\nu) $$ (where $X^\beta\coleq X_1^{\beta_1}\cdots X_N^{\beta_N}$ and $\beta !\coleq \beta_1 !\cdots\beta_N !$). From this formula we deduce that $$ \alpha_2^*(p) =\sum_\beta\frac{p(\dotvar ;X^\beta)}{\beta !}\otimes {\xi}^\beta. $$ In view of the characterization of $\cI_\gl$ in the proof of Lemma \ref{l: codimension two}, we now see that \begin{equation}\label{eq4} \alpha^*_2(p)=p\otimes 1 + p(\dotvar ; X_1)\otimes\xi_1 \quad\mathrm{mod}\;\; \cP \otimes(\cI_\lambda\cap\P). \end{equation} Furthermore, again by Lemma \ref{l: codimension two}, we have $\cP = \C 1 \oplus \C \xi_1 \oplus (\cI_\gl \cap \cP),$ so that $\cP^{\otimes 2} = [\cP \otimes (\C 1 \oplus \C \xi_1)] \oplus [\cP \otimes (\cI_\gl \cap \cP)].$ Accordingly, the natural projection map $\cP^{\otimes 2}\to \cP/\tilde \cI \otimes \cP / (\cI_\gl \cap \cP)$ has kernel equal to \begin{equation} \label{e: kernel pr} [\widetilde\cI \otimes (\C 1 \oplus \C \xi_1)]\, \oplus\,[ \cP \otimes (\cI_\gl \cap \cP)]. \end{equation} Combining (\ref{eq4}) and (\ref{e: kernel pr}), we finally see that $\bar \ga_2^*(p) = 0$ if and only if $$ p\otimes 1 + p(\dotvar ; X_1)\otimes\xi_1 \;\; \in \widetilde \cI \otimes (\C 1 \oplus \C \xi_1), $$ which in turn is equivalent to $p \in \widetilde \cI$ and $\partial_\gl p \in \widetilde \cI.$ \end{proof} \begin{lemma} \label{sec41 Let $\lambda_1,\dots,\lambda_n\in\fv_\iC\setminus\{0\}$. If $n>1$, the kernel of $$ \bar{\alpha}_n^*:\P\longrightarrow(\P/\cI_{\lambda_1}\cap\P)\otimes\cdots\otimes(\P/\cI_{\lambda_n}\cap\P) $$ equals the kernel of $\bar{\alpha_2}^*:\P\longrightarrow(\P/\ker{\bar{\alpha}_{n-1}^*})\otimes(\P/\cI_{\lambda_n}\cap\P),$ where $\bar{\alpha}_{n-1}^*$ denotes the composition on the left of the algebra homomorphism $\alpha^*_{n-1}$ with the projection $\P^{\otimes (n-1)}\rightarrow (\P/\cI_{\lambda_1}\cap\P)\otimes\cdots\otimes(\P/\cI_{\lambda_{n-1}}\cap\P)$. \end{lemma} \par \begin{proof The map $\bar{\alpha}_{n-1}^*$ induces an embedding (denoted by the same symbol) $$ \bar{\alpha}_{n-1}^*:\P/\ker{\bar{\alpha}_{n-1}^*}\hookrightarrow (\P/\cI_{\lambda_1}\cap\P)\otimes\cdots\otimes(\P/\cI_{\lambda_{n-1}}\cap\P). $$ Moreover, the following diagram commutes \vspace{1mm} $$ \xymatrix{ \P\ar[d]_{\bar{\alpha}_2^*}\ar[r]^{\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\bar{\alpha}_n^*} & (\P/\cI_{\lambda_1}\cap\P)\otimes\cdots\otimes(\P/\cI_{\lambda_n}\cap\P)\\ (\P/\ker{\bar{\alpha}_{n-1}^*})\otimes(\P/\cI_{\lambda_n}\cap\P)\ar@{^{(}->}[ur]_{\qquad\bar{\alpha}_{n-1}^*\otimes\id_{\P/(\cI_{\lambda_n}\cap\P)}} & } $$ \vspace{2mm} Thus, $\ker{\bar{\alpha}_n^*}=\ker{\bar{\alpha}_2^*}$. \end{proof} \begin{lemma}\label{sec42 Let $\lambda_1,\dots,\lambda_n\in\fv_\iC\setminus\{0\}$. Then the kernel of $$ \bar{\alpha}_n^*:\P\longrightarrow(\P/\cI_{\lambda_1}\cap\P)\otimes\cdots\otimes(\P/\cI_{\lambda_{n}}\cap\P) $$ consists of the elements $p$ of $\cM\cap\P$ satisfying $p(\dotvar;\lambda_{j_1}\cdots\lambda_{j_l})\in\cM$, for all $1\leq l\leq n$, $1\leq j_1<\cdots<j_l\leq n.$ In particular, this kernel contains $\cM^{n +1} \cap \cP.$ \end{lemma} \begin{proof The final statement follows from the first by repeated application of the Leibniz rule. We prove the first statement by induction on $n.$ If $n=1$, then $\bar{\alpha}_1^*$ is just the projection $\P\rightarrow\P/\cI_{\lambda_1}\cap\P$, and $\ker(\bar{\alpha}_1^*)=\cI_{\lambda_1}\cap\P=\{p\in\cM\cap\P\mid \,p(\dotvar ;\lambda_1)\in\cM\}$. Let $n\geq 2$ and assume the statement to be valid for smaller values of $n.$ Using Lemmas \ref{sec40} and \ref{sec41}, we find that $$ \ker(\bar{\alpha}_n^*)= \{p\in \cP \mid p \in \ker\bar{\alpha}_{n-1}^*\;\; \textrm{and}\;\; \partial_{\gl_n} p \in\ker\bar{\alpha}_{n-1}^*\}. $$ By the induction hypothesis it follows that $\ker (\bar \ga_n^*)$ consists of all polynomials $p \in \cM \cap \cP$ such that for all $1\leq l \leq n-1$ and all $1\leq j_1 < \cdots < j_l \leq n-1$ the polynomials $p(\dotvar; \lambda_{j_1}\cdots\lambda_{j_l})$ and $p(\dotvar ; \lambda_{j_1}\cdots\lambda_{j_l}\gl_n) = \partial_{\gl_n}p(\dotvar ; \lambda_{j_1}\cdots\lambda_{j_l})$ belong to $\cM.$ This implies the result. \end{proof} \begin{prop \label{sec43}${}$ \begin{enumerate} \itema Let $E_1,\dots,E_n$ be finite dimensional $\O_0$-modules. There exist $N,\,k\in\N$ such that $E_1\otimes\cdots\otimes E_n$ is a quotient of the $\O_0$-module $\C^N\otimes \O_0/\cM^{k+1}$ (action on the second tensor component). \itemb Let $\cI$ be a cofinite ideal of $\O_0$. There exist $\gl_1,\dots,\gl_n\in\fv_\iC$ such that the algebra homomorphism $$ \bar{\alpha}_n^*:\P\longrightarrow\;(\P/\cI_{\lambda_1}\cap\P)\otimes \cdots\otimes(\P/\cI_{\lambda_n}\cap\P) $$ satisfies $\ker{\bar{\alpha}_n^*}\subset \cI.$ In particular, the $\cO_0$-module $\cO_0/\cI$ is a subquotient of $(\cO_0/\cI_{\lambda_1})\otimes \cdots\otimes(\cO_0/\cI_{\lambda_n}).$ \end{enumerate} \end{prop} \begin{proof Assertion (a) follows from Corollary \ref{sec18} and \eqref{sec23}. For the second assertion, we note that by Lemma \ref{sec25}, there exists a number $k\in\N$ such that $(\cM\cap\P)^{k+1}\subset \cI\cap\P\subset\cM$. Fix a basis $\{X_1,\dots, X_N\}$ of $\fv$. Put $n\coleq kN$ and $$ \gl_{kj + i}\coleq X_j \qquad \textrm{for}\;\;\; 0 \leq j \leq N-1\;\; \textrm{and}\;\; 1 \leq i \leq k. $$ By Lemma \ref{sec42}, the ideal $\ker{\bar{\alpha}_n^*}$ is equal to $$ \{p\in\cM\cap\P\mid \,p(\dotvar ;\lambda_{j_1}\cdots\lambda_{j_l}) \in\cM,\,1\leq l\leq n,\,1\leq j_1<\cdots<j_j\leq n\}. $$ Hence, if $p\in\ker{\bar{\alpha}_n^*}$, then $$ p(0; X^\beta)=0,\quad |\beta|\leq k. $$ This implies that $p\in (\cM\cap\P)^{k+1}$. Thus, $\ker{\bar{\alpha}_n^*}\subset \cI.$ In particular, it follows that the $\cP$-module $\cP/(\cI \cap \cP)$ is a subquotient of $(\cP/\cI_{\gl_1} \cap \cP) \otimes \cdots \otimes (\cP/\cI_{\gl_n} \cap \cP).$ The final assertion now follows by application of Lemmas \ref{sec24} and \ref{sec25}. \end{proof} \subsection{Holomorphic families of continuous representations} \label{s: holomf We retain our assumption that $\Omega$ is an open subset of the finite dimensional complex linear space $\fv_\iC.$ \par If $X$ is a locally compact Hausdorff space, and $V$ a locally convex (Hausdorff) space, then by $C(X,V)$ we denote the space of $V$-valued continuous functions on $X$, equipped with the topology of uniform convergence on compact subsets. If $V$ is quasi-complete, then $C(X, V)$ is quasi-complete as well. Let $V$ be a quasi-complete locally convex space. The subspace $\cO(\Omega,V)$ of $V$-valued holomorphic functions on $\Omega$ is closed in $C(\Omega,V)$ hence quasi-complete of its own right. For our further considerations, it is important to note that the algebraic tensor product $\cP(\fv_\iC)\otimes V$ is dense in $\O(\Omega,V),$ see Lemma \ref{l: density of P in O}. \par If $V,W$ are locally convex spaces, we write $\Hom(V,W)$ for the space of continuous complex linear maps $V \to W.$ This space, equipped with the strong operator topology is locally convex again. Moreover, if $V$ is barrelled and $W$ quasi-complete, then $\Hom(V,W)$ is quasi-complete as well, see appendix, text before Lemma \ref{l: continuity of composition}. \par As usual, we write $\Endo(V)$ for $\Hom(V,V).$ Let $U$ be a third locally convex space. Then the composition map $$ \gb: \Hom(U,V) \times \Hom(V,W) \to \Hom(U,W),\quad (A,B) \mapsto B\after A $$ is bilinear and separately continuous. Moreover, if $V$ is barrelled, then by the principle of uniform boundedness, $\gb$ is continuous relative to (i.e., when restricted to) sets of the form $\Hom(U,V) \times C,$ with $C \subset \Hom(V,W)$ compact; see appendix, Lemma \ref{l: continuity of composition}. Now assume that $U,V$ and $W$ are quasi-complete locally convex spaces. If $U$ and $V$ are barrelled, then it follows from the material in the appendix, see Lemma \ref{l: composition of holomorphic homomorphisms}, that the natural pointwise composition defines a bilinear map $$ \gb_*: \cO(\Omega,\Hom(U,V)) \times \cO(\Omega, \Hom(V,W)) \to \cO(\Omega,\Hom(U,W)). $$ \begin{definition}\label{defhfr \rm A holomorphic family of continuous representations of $G$ over $\Omega$ is a pair $(\pi,V)$ such that the following conditions are fulfilled: \begin{enumerate} \itema $V$ is a Fr\'echet space. \itemb $\pi$ is a continuous map from $G$ to $\O(\Omega,\End{V})$ satisfying: \begin{enumerate} \item[(1)] $\pi(g_1g_2)=\pi(g_1)\pi(g_2),\quad g_1,g_2\in G,$\\ $\pi(e_G)= 1_\pi$, where $1_\pi(\mu) = \id_V$ for all $\mu \in \Omega;$ \item[(2)] for every $k\in K$, the $\End{V}$-valued function $\pi(k)$ is constant on $\Omega$. \end{enumerate} \end{enumerate} A holomorphic family of smooth representations of $G$ over $\Omega$ is a family $(\pi,V)$ as above such that in addition \begin{equation} \label{e: condition smooth family} (g,\gl)\mapsto \pi(g)(\gl)v,\quad G \times \Omega \to V \end{equation} is smooth for every $v \in V.$ \end{definition} \begin{rem \rm Given $g \in G$ and $\gl \in \Omega$ we agree to write $\pi_\gl(g)\coleq \pi(g)(\gl).$ The condition that $V$ is Fr\'echet ensures that the principle of uniform boundedness is valid. By application of this principle it follows that the map \begin{equation} \label{e: holomorphic family written alternatively} G \times \Omega \times V \to V,\qquad (g,\gl,v) \mapsto \pi_\gl(g)v \end{equation} is continuous and holomorphic in $\gl.$ More generally we could have given the definition of a holomorphic family of continuous representations under the weaker assumption that $V$ be quasi-complete and barrelled. All results of the present section and the next are in fact valid under this weaker assumption. However, in \S \ref{s: parabolic induction} the assumption that $V$ is Fr\'echet will really be needed. \end{rem} \begin{definition \rm Let $(\pi,V_\pi)$ and $(\rho,V_\rho)$ be two holomorphic families of continuous representations of $G$ over $\Omega$. A holomorphic family of intertwining operators, $T$, between $(\pi,V_\pi)$ and $(\rho,V_\rho)$ is an element of $\O(\Omega,\Hom(V_\pi,V_\rho))$ satisfying: $$ T\pi(g)=\rho(g)T,\qquad g\in G. $$ \end{definition} \begin{definition} \rm The category $\HFG$ is defined as follows. \begin{enumerate} \item[(a)] The objects are the holomorphic families of continuous representations of $G$ over $\Omega$. \item[(b)] The morphisms are the holomorphic families of intertwining operators between two objects. \end{enumerate} For any $(\pi,V_\pi)\in\HFG$, the identity morphism is the holomorphic family $1_\pi$ of interwining operators defined by: $$ 1_\pi(\gl)\coleq \id_{V_\pi},\quad \gl\in\Omega. $$ The composition $T'\after T$ of two (composable) morphisms is given by pointwise composition: $(T'\after T)_\gl = T_\gl' \after T_\gl.$ It is again a holomorphic family of intertwining operators, by virtue of Lemma \ref{l: composition of holomorphic homomorphisms}. \end{definition} If $V$ and $W$ are locally convex spaces, and $E$ a finite dimensional complex linear space, then the map $(A_1, A_2) \mapsto A_1 \otimes A_2$ induces a linear isomorphism \begin{equation}\label{e: end-hom} \End{E}\otimes \Hom(V,W)\simeq \Hom(E\otimes V,E\otimes W) \end{equation} which we shall use for identifying these spaces. Accordingly, if $V$ and $W$ are quasi-complete, and $V$ barrelled, then $$ \diffD^{(E)}: T \mapsto T^{(E)} $$ defines a continuous linear map $$ \cO(\Omega, \Hom(V,W)) \to \cO(\Omega, \Hom(E \otimes V, E \otimes W)) $$ (see Remark \ref{r: J is M}). \begin{definition \rm Let $E$ be a finite dimensional $\O_0$-module. \begin{enumerate} \item[(a)] For $(\pi,V)\in\HFG$, we define $\pi^{(E)}$ to be the continuous map $$ \pi^{(E)} = \diffD^{(E)}\after \pi:\;\; g\mapsto\pi(g)^{(E)}, \quad G \to \O(\Omega,\End{E\otimes V}). $$ \item[(b)] For any morphism $T:(\pi,V_\pi)\to(\rho,V_\rho)$ of $\HFG$, we define $$ T^{(E)}\coleq \diffD^{(E)}(T) \in \O(\Omega,\Hom(E\otimes V_\pi,E\otimes V_\rho)). $$ \end{enumerate} \end{definition} \begin{prop}\label{p: the differentiation is a functor Let $E$ be a finite dimensional $\O_0$-module. Then $\diffD^{(E)}$ defines a functor from $\HFG$ to itself. \end{prop} \begin{proof Let $(\pi, V) \in \HFG.$ We will first show that $(\pi^{(E)},E\otimes V)\in \HFG$. Let $\beta$ denote the composition map in $\End{V}$. It is a bilinear map, which preserves holomorphy on $\Omega,$ by Lemma \ref{l: composition of holomorphic homomorphisms}. Moreover, by Corollary \ref{c: multiplicativity of diffD} we have, for all $g_1,g_2\in G$, $$ \begin{array}{rcl} \pi^{(E)}(g_1g_2) & = & \pi(g_1g_2)^{(E)}\\ & = & (\pi(g_1)\pi(g_2))^{(E)}\\ & = & \pi(g_1)^{(E)}\pi(g_2)^{(E)}\\ & = & \pi^{(E)}(g_1)\pi^{(E)}(g_2). \end{array} $$ Also, $\pi^{(E)}(e_G)=1_\pi^{(E)}=1_{\pi^{(E)}}$. For $k\in K$, the $\Endo(V)$-valued function $\pi(k)$ is a constant on $\Omega,$ and therefore, so is the $\Endo(E\otimes V)$-valued function $\pi^{(E)}(k) = \pi(k)^{(E)}.$ Hence (b.1) and (b.2) of Definition \ref{defhfr} are fulfilled. \par It remains to show that, for any morphism $T:(\pi,V_\pi)\to (\rho,V_\rho)$ of $\HFG$, $T^{(E)}$ is a morphism of $\HFG$ from $(\pi^{(E)}, E\otimes V_\pi)$ to $(\rho^{(E)},E\otimes V_\rho)$. For this it suffices to show that, for $g\in G$, $$ T^{(E)} \pi^{(E)}(g)=\rho^{(E)}(g)T^{(E)}. $$ This follows from Corollary \ref{c: multiplicativity of diffD}, see also the first part of the present proof. \end{proof} The category $\HFG$ has a null object, $(0,\{0\})$, and one can define a biproduct in $\HFG$ as follows. Let $(\pi,V_\pi),\,(\rho,V_\rho)\in\HFG$. Set, for any $g\in G$, $\gl\in\Omega$ and $(v,w)\in V_\pi\oplus V_\rho$, $$(\pi\oplus\rho)_\gl(g)(v,w)\coleq (\pi_\gl(g)v,\rho_\gl(g)w).$$ Then $(\pi\oplus\rho,V_\pi\oplus V_\rho)$ defines an object of $\HFG.$ \par We define the full subcategory $\HFGinfty$ of $\HFG$ by stipulating that the objects are the holomorphic families of smooth representations over $\Omega$ (the set of morphisms between objects in $\HFGinfty$ coincides with the set of morphisms between the objects viewed as objects for the bigger category $\HFG.$ Likewise, the full subcategory $\HFGinftyadm$ of $\HFGinfty$ consists of the objects $(\pi, V)$ of $\HFGinfty$ with $V_K$ admissible. \par If $(\pi, V) \in \HFG,$ then the identity morphism $1_\pi$ is constant as a $\End{V}$-valued function on $\Omega.$ This enables us to define a particular subcategory. \par \begin{definition}\label{d: sec59 \rm The subcategory $\HFGO$ of $\HFG$ is defined as follows. \begin{enumerate} \itema The objects are the objects of $\HFG.$ \itemb If $(\pi, V)$ and $(\rho,W)$ are objects of $\HFG,$ then the associated collection of $\HFGO$-morphisms consists of all $\HFG$-morphisms $T$ in $\cO(\Omega, \Hom(V,W))$ which are constant as a function on $\Omega.$ \end{enumerate} In a similar fashion, we define the subcategories $\HFGOinfty$ and $\HFGOinftyadm$ of $\HFGinfty$ and $\HFGinftyadm$, respectively. \end{definition} Note that $\HFGOinftyadm$ is a full subcategory of $\HFGOinfty,$ which in turn is a full subcategory of $\HFGO.$ \begin{rem \label{r: invariance under functor} \rm For every $E\in \FMod_{\cO_0}$, the functor $J^{(E)}: \HFG \to \HFG$ leaves all subcategories $\HFGinfty,$ $\HFGinftyadm,$ $\HFGO$, $\HFGOinfty$ and $\HFGOinftyadm$ invariant. \end{rem} \begin{lemma Let $\psi: E \to E'$ be a morphism in $\FMod_{\cO_0}$ and let $(\pi, V_\pi)$ be a holomorphic family of representations. Then $\psi \otimes 1_\pi$ intertwines the families $\pi^{(E)}$ and $\pi^{(E')}.$ \end{lemma} \begin{proof If $f \in \cO(\Omega),$ then from the definitions it readily follows that $$ \psi \after f^{(E)}(\mu) = f^{(E')}(\mu) \after \psi,\quad \mu \in \Omega. $$ From this and the identification $\End{E}\otimes\End{V_\pi}\simeq\End{E\otimes V_\pi}$, it follows that for all $g \in G$ the map $\psi \otimes 1_{\pi}: E \otimes V_\pi \to E'\otimes V_\pi$ intertwines $\pi_\mu^{(E)}(g) = \pi(g)^{(E)}(\mu)$ with $\pi_\mu^{(E')}(g) = \pi(g)^{(E')}(\mu).$ \end{proof} The above lemma justifies the following definition. \begin{definition} \rm Given any $(\pi,V_\pi)\in\obj(\HFG)$, we define the functor $X_\pi$ from $\FMod_{\O_0}$ to $\HFGO$ as follows. \begin{enumerate} \itema For an object $E\in\FMod_{\O_0},$ the associated object of $\HFGO$ is given by $$ X_\pi(E)\coleq (\pi^{(E)},E\otimes V_\pi). $$ \itemb For a morphism $\psi: E\rightarrow E'$ of $\FMod_{\O_0}$, the associated morphism of $\HFGO$ is given by $$ \quad X_\pi(\psi)\coleq \psi\otimes 1_\pi:\;\; (\pi^{(E)},E\otimes V_\pi)\,\longrightarrow\, (\pi^{(E')},E'\otimes V_\pi). $$ \end{enumerate} \end{definition} \begin{rem \rm It is readily checked that $X_\pi$ is a functor. Indeed, $X_\pi$ respects composition of morphisms, and $X_\pi(\id_E) = \id_{E\otimes V_\pi}= 1_{\pi^{(E)}}.$ \end{rem} As $\FMod_{\cO_0}$ is an abelian category, we have the usual notion of finite direct sums and exact sequences in $\FMod_{\cO_0}.$ \par The category $\Vect$ of complex vector spaces is abelian. If $T: (\pi, V) \to (\rho, W)$ is a morphism in $\HFGO,$ then there exists a unique linear map $T_0: V \to W$ such that $T(\gl) = T_0$ for all $\gl \in \Omega.$ By abuse of notation we will write $T$ for $T_0.$ We thus have a forgetful functor $\HFGO \to \Vect.$ \par The category $\HFGO$ is not abelian. Nevertheless, we may use the forgetful functor to define exact sequences \begin{definition}\rm A sequence $((\pi_k, V_k), T_k), p \leq k \leq q,$ in the category $\HFGO$, where $p, q \in \Z, p < q,$ will be called exact if its image under the forgetful functor $\HFGO \to \Vect$ is exact, i.e., the image of $T_{k-1}$ equals the kernel of $T_k$ for all $p < k \leq q.$ \end{definition} \begin{lemma} \label{l: exactness properties X pi} Let $(\pi, V_\pi) \in \obj(\HFG).$ Then the functor $X_\pi: \FMod_{\cO_0} \to \HFGO$ has the following properties. \begin{enumerate} \itema It sends every short exact sequence $0 \to E \to E' \to E'' \to 0$ to a similar short exact sequence in $\HFGO.$ \itemb It sends every exact sequence of the form $0 \to E \to E'$ in $\FMod_{\cO_0}$ to an exact sequence of similar form in $\HFGO.$ \itemc It sends every exact sequence of the form $E' \to E''\to 0$ in $\FMod_{\cO_0}$ to an exact sequence of similar form in $\HFGO.$ \itemd It sends a direct sum of the form $E = E_1 \oplus E_2$ in $\FMod_{\cO_0}$ to a similar direct sum in $\HFGO.$ \end{enumerate} \end{lemma} \begin{proof Let $\FDVect$ denote the (abelian) category of finite dimensional complex linear spaces. Then we have a forgetful functor $\cF$ from $\FMod_{\cO_0}$ to $\FDVect.$ A sequence in $\FMod_{\cO_0}$ is exact if and only if its image under $\cF$ is exact in $\FDVect.$ According to the definition above, the forgetful functor $\cF': \HFGO \to \Vect$ has a similar property. \par Given $U \in \Vect$ we define the functor $X_U: \FDVect \to \Vect$ by $X_U(E) = E \otimes U$ for an object $E$ of $\FDVect.$ A morphism $f: E \to E'$ in $\FMod_{\cO_0}$ is mapped to $X_U(f)\coleq f \otimes \id_U: E \otimes U \to E' \otimes U.$ It is readily seen that the functor $X_U$ is exact, and has the obvious properties analogous to (a) - (d). \par Since $ \cF' \after X_\pi = X_{\cF'(\pi,V)} \after \cF, $ assertions (a), (b) and (c) of the lemma follow. For assertion (d) it remains to be shown that each natural embedding ${\rm i}_j: E_j \to E$, for $j =1,2,$ is mapped to an embedding $X_\pi(\rmi_j)$ from $X_\pi(E_j)$ onto a closed subspace of $X_\pi(E).$ Let $p_j: E \to E_j$ be the natural projection. Then by exactness of the sequence $0 \to E_1 \to E \to E_2 \to 0$ it follows from the established assertion (a) that $0 \to X_\pi(E_1) \to X_\pi(E) \to X_\pi(E_2) \to 0$ is exact. This implies that $X_\pi(\rmi_1)$ has closed image in $X_\pi(E).$ Likewise, $X_\pi(\rmi_2)$ is seen to have close image. \end{proof} \begin{rem} In view of Remark \ref{r: invariance under functor}, Lemma 2.44 has an obvious generalization to objects from $\HFGinfty$ and from $\HFGinftyadm.$ \end{rem} \begin{prop} \begin{enumerate} \item[(a)] Let $E_1,\dots,E_n\in \FMod_{\O_0}$ and set $E\coleq E_1\otimes\cdots\otimes E_n.$ Then there exist $N,\, k\in\N$ such that, for any object $(\pi,V_\pi)$ in $\HFG$ (resp.~$\HFGinfty,\,\HFGinftyadm$), the family $(\pi^{(E)},E\otimes V_\pi)$ is a quotient of $$ (\id_{\iC^N}\otimes \pi^{(\O_0/\cM^{k+1})},\C^N\otimes(\O_0/\cM^{k+1}\otimes V_\pi)) $$ in the category $\HFGO$ (resp.~$\HFGOinfty,\,\HFGOinftyadm$). \item[(b)] Let $E\in\FMod_{\O_0}$. Then there exist $\lambda_1,\dots,\lambda_n\in\fv_\iC$ such that, for any object $(\pi,V_\pi)$ in $\HFG$ (resp.~$\HFGinfty,\,\HFGinftyadm$), the family $ (\pi^{(E)},E\otimes V_\pi)$ is a subquotient of $$ (\pi^{((\O_0/\cI_{\lambda_1})\otimes\cdots\otimes(\O_0/\cI_{\lambda_n}))}, (\O_0/\cI_{\lambda_1})\otimes\cdots\otimes(\O_0/\cI_{\lambda_n})\otimes V_\pi) $$ in the category $\HFGO$ (resp.~$\HFGOinfty,\,\HFGOinftyadm$). \end{enumerate} \end{prop} \begin{proof This follows from Proposition \ref{sec43} combined with Lemma \ref{l: exactness properties X pi}. \end{proof} \subsection{Holomorphic families of admissible $(\g,K)$-modules For the purpose of this paper, it is convenient to introduce the following notion of holomorphic families of admissible $(\g,K)$-modules. Recall that $\Omega$ is an open subset of the finite dimensional complex linear space $\fv_\iC.$ \par \begin{definition}\label{defhfm \rm A holomorphic family of admissible $(\g,K)$-modules over $\Omega$ is a triple $(\pi_1,\pi_2,V)$ satisfying the following conditions. \begin{enumerate} \itema $V$ is a complex vector space. \itemb $\pi_1$ is a map from $U(\g)\times \Omega$ to $\End{V}$ such that \begin{enumerate} \item[(1)] for each $\lambda\in\Omega$, the map $\pi_1(\cdot,\lambda)$ is a Lie algebra homomorphism; \item[(2)] for each $u\in U(\g)$ and every $v\in V,$ the vector subspace of $V$ generated by the $\pi_1(u,\lambda)v$, $\lambda\in\Omega$, is finite dimensional and the map $\lambda\mapsto \pi_1(u,\lambda)v$ is holomorphic from $\Omega$ into this subspace. \end{enumerate} \itemc $\pi_2$ is a Lie group homomorphism from $K$ to $\mathrm{GL}(V)$ such that: \begin{enumerate} \item[(1)]for each $v\in V,$ the vector subspace of $V$ generated by the $\pi_2(k)v$, $k\in K,$ is finite dimensional; \item[(2)] for all $ v\in V$, $\gl \in \Omega$, $u\in U(\g)$, $k\in K$ and $X\in\k,$ $$ \begin{array}{rcl} \pi_1(\mathrm{Ad}(k)u,\gl)v&=&\pi_2(k)\pi_1(u,\gl)v,\\ \frac{d}{dt}\left[\pi_2(\exp{(tX)})v\right]_{\vert t=0}&=&\pi_1(X,\gl)v; \end{array} $$ \end{enumerate} \itemd for every $\delta\in\hat{K}, $ the $K$-isotypic component $\pi_{2,\delta}$ of $\pi_2$ of type $\delta$ is finite dimensional. \end{enumerate} \end{definition} \begin{rem \rm Let $(\pi_1, \pi_2, V)$ be as in the above definition. Given $\gl \in \Omega$ we agree to write $\pi_{1\gl}$ for the map $\pi_1(\dotvar, \gl): U(\fg) \to \End{V}.$ Then $(\pi_{1\gl}, \pi_2, V)$ is an admissible $(\fg, K)$-module. \end{rem} We also need the following notion of holomorphic family of intertwining operators. \begin{definition \rm Let $(\pi_1,\pi_2,V)$ and $(\rho_1,\rho_2,W)$ be two holomorphic families of admissible $(\g,K)$-modules over $\Omega$. A holomorphic family of intertwining operators between $(\pi_1,\pi_2,V)$ and $(\rho_1,\rho_2,W)$ is a function $T: \Omega \to \Hom(V,W)$ satisfying the following conditions. \begin{enumerate} \itema $T(\lambda)\pi_1(u,\lambda)=\rho_1(u,\lambda)T(\lambda),$ for all $\lambda\in \Omega$ and $u\in U(\g)$; \itemb $T(\lambda)\pi_2(k)=\rho_2(k)T(\lambda),$ for all $\lambda\in \Omega$ and $k\in K$; \itemc for any finite dimensional subspace $\tilde{V}$ of $V$, there exists a finite dimensional subspace $\tilde{W}$ of $W$ such that $T(\lambda)(\tilde V) \subset \tilde{W}$ for all $\lambda \in \Omega,$ and the associated function $\lambda \mapsto T(\gl)|_{\tilde V}$ belongs to $\cO(\Omega, \Hom(\tilde V, \tilde W)).$ \end{enumerate} \end{definition} Let $\HFgKadm$ denote the corresponding category of holomorphic families of admissible $(\fg,K)$-modules. Let $(\pi, V)$ be an object in the category $\HFGinftyadm.$ Then for each $\gl \in \Omega,$ $u \in U(\fg)$ and $v \in V$ we may define $$ \pi(u)(\gl)v:= L_{u^\vee}(g\mapsto \pi(g)(\gl)v)|_{g =e}. $$ We put $\pi_1(u,\gl) = \pi(u)(\gl)|_{V_K}$ and $\pi_2(k) = \pi(k)|_{V_K}.$ \begin{lemma} Let $(\pi, V) \in \HFGinftyadm$ and let $\pi_1, \pi_2$ be defined as above. Then $$ (\pi, V)_K: =(\pi_1, \pi_2, V_K) $$ is a holomorphic family in $\HFgKadm.$ Moreover, $(\dotvar)_K: (\pi, V) \mapsto (\pi, V)_K$ defines a functor $\HFGinftyadm \to \HFgKadm.$ \end{lemma} \begin{proof} It follows from the smoothness of the map (\ref{e: condition smooth family}), the continuity of the map $L_{u^\vee}: C^\infty(G) \to C^\infty(G)$ and the holomorphy with respect to $\gl,$ that $\gf: \gl \mapsto \pi(u)(\gl)v$ defines a holomorphic function $\Omega \to V.$ Let now $v \in V_K$ and let $\vartheta_2 \subset \widehat K$ denote the set of $K$-types appearing in the $\pi(K)$-span of $v.$ Moreover, let $\vartheta_1$ be the set of $K$-types which appear in the $\Ad(K)$-span of $u.$ Finally, let $\vartheta$ be the union of the sets of $K$-types of $\gd_1 \otimes \gd_2,$ for $\gd_j \in \vartheta_j,$ $j =1,2.$ Then $\gf$ has image contained in the finite dimensional subspace $V_\vartheta \subset V_K.$ It follows that $\gf$ is holomorphic as a function $\Omega \to V_\vartheta.$ This shows that $(\pi_1, \pi_2 , V_K)$ satisfies condition (b)(2) of Definition \ref{defhfm}. The other conditions of that definition are pointwise in $\gl,$ and therefore consequences of the standard theory of assigning the $(\fg, K)$-module of $K$-finite vectors to an admissible smooth Fr\'echet representation, see for instance \cite[Lemma 3.3.5]{wallbook1}. The latter assignment is a functor from the category of admissible smooth Fr\'echet representations to the category of Harish-Chandra modules. This implies that $(\pi, V) \mapsto (\pi,V)_K$ has functorial properties which are pointwise in $\gl.$ This in turn is readily seen to imply that $(\dotvar)_K$ is a functor as stated. \end{proof} \medbreak We will now discuss the functor $J^{(E)}$ on the level of holomorphic families of admissible $(\fg, K)$-modules. \begin{definition} \rm Let $E\in\FMod_{\cO_0}.$ For any $(\pi_1,\pi_2,V)\in\HFgKadm$ the triple $(J^{(E)}\pi_1, J^{(E)}\pi_2, J^{(E)}V)$ (also denoted $(\pi_1^{(E)},\pi_2^{(E)},V^{(E)})$) is defined as follows. \begin{enumerate} \itema $J^{(E)}V\coleq E\otimes V.$ \itemb $J^{(E)}\pi_1$ is the map from $U(\g)\times \Omega$ to $\End{E\otimes V}$ given by $$ J^{(E)}\pi_1(u,\lambda)e\otimes v\coleq (\pi_1(u,\dotvar)v)^{(E)}(\lambda)e, $$ for $u\in U(\g),\,\lambda\in \Omega,\, e\in E,$ and $v\in V.$ \itemc $J^{(E)}\pi_2\coleq \id_E\otimes \pi_2.$ \end{enumerate} \end{definition} We now have functors $J^{(E)}$ on $\HFGinftyadm$ and $\HFgKadm.$ They are linked as follows. \begin{lemma}{\ } \begin{enumerate} \itema For any $E\in\FMod_{\cO_0}$, the assignment $J^{(E)}$ defines a functor from $\HFgKadm$ to itself. \itemb The following diagram commutes: $$ \xymatrix{ \ar[d]_{J^{(E)}}\HFGinftyadm \ar[rr]^{(\dotvar)_{K}} &&\ar[d]^{J^{(E)}}\HFgKadm \\ \HFGinftyadm \ar[rr]_{(\dotvar)_{K}} && \;\;\HFgKadm\;\;. } $$ \end{enumerate} \end{lemma} \par \begin{proof Assertion (a) follows by similar arguments as in the proof of Proposition \ref{p: the differentiation is a functor}. For (b), assume that $(\pi,V)$ is a family in $\HFGinftyadm.$ Put $(\pi,V)_K = (\pi_1, \pi_2, V_K),$ then $$ J^{(E)}((\pi, V)_K) = (\pi_1^{(E)}, \pi_2^{(E)}, (V_K)^{(E)}). $$ On the other hand, $J^{(E)}(\pi, V) = (\pi^{(E)}, V^{(E)})$ and $$ (J^{(E)}(\pi, V))_K = ((\pi^{(E)})_1, (\pi^{(E)})_2, (V^{(E)})_K) $$ Now $V^{(E)} = E\otimes V$ is equipped with the $K$-action on the second component, so that $(V^{(E)} )_K = E\otimes V_K = (V_K)^{(E)}.$ Moreover, $\pi_2^{(E)}(k) = 1_E \otimes \pi_2(k) = 1_E \otimes \pi(k,\gl)|_{V_K} = \pi^{(E)}(k,\gl)|_{E \otimes V_K} = (\pi^{(E)})_2(k).$ It remains to establish the identity \begin{equation} \label{e: identity pi 1 E} \pi_1^{(E)} = (\pi^{(E)})_1. \end{equation} Since both are representations of $U(\fg)$ in $E \times V_K,$ it suffices to check the identity on a fixed element $X \in \fg.$ Fix $e\in E$ and $v \in V.$ We first observe that $$ \pi_1^{(E)}(X)v = (\pi_1(X, \dotvar)v)^{(E)} = [\frac{d}{dt}\pi(\exp tX)(\dotvar)v]^{(E)}|_{t = 0}. $$ In view of the natural identification $\Endo (E) \otimes V \simeq \Hom(E, E \otimes V),$ we note that \begin{eqnarray*} (\pi^{(E)})_1(X, \dotvar)v &=& \frac{d}{dt}[\pi^{(E)}(\exp tX)(\dotvar)v]|_{t = 0}\\ &=& \frac{d}{dt}[(\pi(\exp tX)(\dotvar)v)^{(E)}]|_{t = 0}. \end{eqnarray*} The identity (\ref{e: identity pi 1 E}) now follows by application of the lemma below. \end{proof} \begin{lemma} Let $V$ be a quasi-complete locally convex space, and let $\gf: \R \times \Omega \to V$ be a $C^1$-map which is holomorphic in the second variable. Then \begin{equation} \label{e: d dt and (E)} \frac{d}{dt} [\gf(t, \dotvar)^{(E)}] = [\frac{\partial \gf}{\partial t}(t, \dotvar)]^{(E)}. \end{equation} \end{lemma} \begin{proof} Let $S$ be the space of $C^1$-maps $\R \times \Omega \to V,$ equipped with the usual quasi-complete topology. Let $S_0$ be the subspace consisting of functions in $S$ which are holomorphic in the second variable. Then $S_0$ is closed in $S,$ hence quasi-complete. The identity (\ref{e: d dt and (E)}) at $(t,\gl)$ can be viewed as an identity of continuous linear functionals on $S_0.$ Hence, it suffices to check the identity on the dense subspace $C^1(\R) \otimes \cO(\Omega) \otimes V.$ This amounts to checking whether $$ (\frac{d}{dt} \otimes I \otimes I) \after (I \otimes J^{(E)} \otimes I) = (I \otimes J^{(E)} \otimes I) \after (\frac{d}{dt} \otimes I \otimes I). $$ The latter is obvious. \end{proof} \subsection{Parabolic induction}\label{s: parabolic induction Let $\fg = \fk \oplus \fp$ be a Cartan decomposition associated with the maximal compact subgroup $K$ and let $\Cartan$ be the associated involution of $G.$ Let $\fa \subset \fp$ be a maximal abelian subspace, and let $A = \exp \fa.$ Let $\cP(A)$ denote the collection of parabolic subgroups of $G$ containing $A.$ Let $\cL(A)$ denote the collection of $\Cartan$-stable Levi components of parabolic subgroups from $\cP(A).$ \par Let $\fv$ be a finite dimensional real linear space, and $\Omega$ an open subset of its complexification. For $L \in \cL(A)$ we denote by $\HFL$ the category defined as in Definition \ref{defhfr}, with the group $L$ in place of $G.$ \par The parabolic induction functor from $\HFLinfty$ to $\HFGinfty$ is defined as follows. Let $(\xi,V_\xi)\in\HFLinfty$ be a holomorphic family of smooth representations of $L$ defined over $\Omega.$ Denote by $\bar{\pi}_{P,\xi_\gl}$ the right regular representation of $G$ on $C^\infty(G:P:\xi_\gl),$ where $$ \begin{array}{rl} C^\infty(G:P:\xi_\gl)\coleq \{\psi\in C^\infty(G,V_\xi): & \psi(nmg)=\xi_\gl(m)\psi(g),\\ & (g,m,n)\in G\times L\times N_P\}. \end{array} $$ Let \begin{equation} \label{e: defi CinftyKxi} \begin{array}{rl} C^\infty(K:\xi)\coleq \{\psi\in C^\infty(K,V_\xi): & \psi(mk)=\xi(m)\psi(k),\\ & (k,m)\in K\times (K\cap L)\}. \end{array} \end{equation} Restriction of functions to $K$ induces a continuous linear isomorphism between these spaces. Let $\pi_{P,\xi_\gl}$ denote the representation of $G$ on $C^\infty(K:\xi)$, obtained from $\bar{\pi}_{P,\xi_\gl}$ by transfer of structure, and set, for $g\in G$, $\gl\in \fv_\iC$, $$ \pi_{P,\xi,\gl}(g)\coleq \pi_{P,\xi_\gl}(g). $$ Then it is readily seen that $(\pi_{P,\xi},C^\infty(K:\xi))$ belongs to $\HFGinfty.$ Here, the information that $V_\xi$ is a Fr\'echet space is needed to conclude that $C^\infty(K\col \xi)$ is Fr\'echet, in particular barrelled. Let $W_1,\,W_2$ and $W_3$ be quasi-complete locally convex spaces, and let $\alpha$ be a continuous linear map from $W_2$ to $W_3$. Then the map $$ L_\alpha: \; \varphi\longmapsto \alpha\circ\varphi, \;\;\;\Hom(W_1,W_2)\longrightarrow \Hom(W_1,W_3) $$ is continuous linear. It readily follows that the map $$ \id_{\cO(\Omega)}\, \hatotimes \,L_\ga: \;\;\cO(\Omega, \Hom(W_1, W_2)) \longrightarrow \cO(\Omega, \Hom(W_1, W_3)) $$ given by $f \mapsto L_\ga \after f$ is continuous linear. The notation for this map is explained by the fact that it may be viewed as the unique continuous linear extension of $\id_{\cO(\Omega)} \otimes L_\ga.$ \begin{lemma Let $L_\ga$ be as above and let $E$ be a finite dimensional $\cO_0$-module. Then we have the following identity of maps $$ \cO(\Omega, \Hom(W_1,W_2)) \too \cO(\Omega, \End{E} \otimes \Hom(W_1, W_3)): $$ $$ \diffD^{(E)}\circ (\id_{\O(\Omega)}\, \hatotimes\, L_\alpha)= (\id_{\O(\Omega)}\,\hatotimes \,\id_{\End{E}}\,\hatotimes\, L_\alpha)\circ {\diffD}^{(E)}. $$ \end{lemma} \begin{proof By continuity of the expressions on both sides of the identity it suffices to prove the identity on the dense subspace $\O(\Omega)\otimes \Hom(W_1,W_2)$ of $\O(\Omega,\Hom(W_1,W_2)).$ But then the identity becomes obvious, in view of Remark \ref{r: J is M}, last line. \end{proof} The following result may be phrased as `derivation commutes with induction'. \begin{prop Let $E$ be a finite dimensional $\O_0$-module, $P$ a parabolic subgroup with Levi component $L$, and $(\xi,V_\xi)\in\HFL^\infty$. Then $$\pi_{P,\xi}^{(E)}=\pi_{P,\xi^{(E)}}.$$ \end{prop} \begin{proof Let $g\in G$ and $k\in K$. Put $\alpha(k)\coleq \ev_k:C^\infty(K:\xi)\to V_\xi$. Let $$L_{\alpha(k)}:\End{C^\infty(K:\xi)}\to \Hom(C^\infty(K:\xi),V_\xi)$$ be defined as in the previous lemma. Accordingly, \begin{equation}\label{comm0} (\id_{\O(\Omega)}\hatotimes\id_E\hatotimes L_{\alpha(k)})(\pi_{P,\xi}^{(E)}(g)) = \left((\id_{\O(\Omega)}\hatotimes L_{\alpha(k)})(\pi_{P,\xi}(g))\right)^{(E)}. \end{equation} Write $g=np(g)\kappa(g)$ uniquely via the decomposition $G=N_PA_P(M_P\cap\exp\p)K$, where $n\in N_P$, $p(g)\in A_P(M\cap\exp\p)$ and $\kappa(g)\in K$. We then have the following identity: $$ (\id_{\O(\Omega)}\hatotimes L_{\alpha(k)})(\pi_{P,\xi}(g)) = \xi(p(kg))\circ L_{\alpha(\kappa(kg))}. $$ Hence, it follows from \eqref{comm0} that \begin{eqnarray*} \lefteqn{ \!\!\!\!\!\!\!\!\!\!\!\!\!\! (\id_{\O(\Omega)}\hatotimes\id_E\hatotimes L_{\alpha(k)})(\pi_{P,\xi}^{(E)}(g)) \;\;\;\;\;\;\;\;\;\;\;\;\;\; \;\;\;\;\;\;\;\;\;\;\;\;\;\;}\\ &=& \left(\xi(p(kg))\circ L_{\alpha(\kappa(kg))}\right)^{(E)}\\ &=& \left(\xi^{(E)}(p(kg))\circ L_{\alpha(\kappa(kg))}\right)\\ &=& (\id_{\O(\Omega)}\hatotimes\id_E\hatotimes L_{\alpha(k)})(\pi_{P,\xi^{(E)}}(g)) \end{eqnarray*} and the statement follows. \end{proof} \section{The Arthur-Campoli relations}\label{s: Arthur's PW} Given a parabolic subgroup $P\in\cP(A)$ we denote its Langlands decomposition by $P=M_PA_PN_P.$ Let $(\xi, V_\xi)$ be a smooth, irreducible and admissible Fr\'echet representation of $M_P.$ Given $\gl \in \faPdc$ we denote by $\xi \otimes \gl$ the smooth representation of $L_P: = M_P A_P$ in $V_\xi$ defined by $$ \xi \otimes \gl\, (ma) = a^{\gl + \rho_P} \xi(m). $$ As usual, here $\rho_P \in \faPd$ is defined by $\rho_P =\half \tr(\ad(\dotvar)|_{\fn_P}).$ The associated induced representation $\bar\pi_{P,\xi \otimes \gl}$ of $G$ in $C^\infty(G\col P\col \xi \otimes \gl)$ is defined as in Section \ref{s: parabolic induction}, with $\fv = \fa^*, \Omega = \fadc$ and $\xi_\gl = \xi \otimes \gl.$ The family of these representations may be viewed as a holomorphic family $\pi_{P, \xi\otimes(\dotvar)}$ of smooth representations of $G$ over $\Omega = \fadc$ on the fixed space $C^\infty(K\col \xi),$ defined in (\ref{e: defi CinftyKxi}). As in the mentioned section, we agree to write $\pi_{P, \xi, \gl} = \pi_{P, \xi \otimes \gl}.$ If $P$ is a minimal parabolic subgroup, the representations $\pi_{P,\xi, \gl}$ just defined are called representations of the smooth minimal principal series of $G.$ Let now $P \in \cP(A)$ be arbitrary again. Then $M_P$ is a group of the Harish-Chandra class, with maximal compact subgroup $K_P:= M_P \cap K.$ Two continuous admissible $M_P$-representations of finite length in a quasi-complete locally convex space are said to be infinitesimally equivalent if their Harish-Chandra modules are equivalent as $(\fm_P, K_P)$-modules. For each equivalence class $\omega$ of irreducible unitary representations of $M_P,$ we fix a smooth admissible Fr\'echet representation $\xi = \xi_\omega$ which is infinitesimally equivalent to a representation of class $\omega,$ and which is topologically equivalent to a closed subrepresentation of a representation of the smooth minimal principal series of $M_P.$ Indeed, this is possible by the subrepresentation theorem for the group $M_P.$ The set of all these chosen representations $\xi_\omega$ is denoted by $\dMP.$ Thus, $\omega \mapsto \xi_\omega$ defines a bijection from the set of equivalence classes of irreducible unitary representations of $M_P$ onto $\dMP.$ \begin{rem \rm In view of the theory of the Casselman--Wallach globalization functor, see \cite[Section 11]{wallbook2}, the representation $\xi_\omega$ is a smooth Fr\'echet globalization of moderate growth of the Harish-Chandra module of any representative of $\omega.$ This characterization makes the choice of $\dMP$ more natural, but will not be needed in the present paper. \end{rem} We denote by $\dMPds$ the subset of $\dMP$ consisting of the representations $\xi_\omega,$ with $\omega$ a discrete series representation. In particular, if $P \in \cP(A)$ is minimal, then $M_P$ equals the centralizer $M$ of $\fa$ in $K,$ and $M_{ds}^\wedge=\dM$. For each $(\xi, V_\xi)\in \dMPds$ we put \begin{equation}\label{e: defi SSPxi} \SS(P,\xi)\coleq \End{C^\infty(K:\xi)}_{K\times K}, \end{equation} and we define the algebraic direct sum of linear spaces \begin{equation}\label{e: defi cSP} \cS(P)\coleq \oplus_{\xi \in \hatMPds} \cS(P, \xi). \end{equation} \subsection{The Arthur-Campoli relations Fix a minimal parabolic subgroup $P_0$ in $\cP(A)$ and let $P_0=MAN_0$ be its Langlands decomposition. In \cite[III, \S 4]{arthurpw}, Arthur defines a Paley-Wiener space involving all minimal parabolic subgroups containing $A.$ This definition is given in terms of on the one hand Paley-Wiener growth conditions and on the other the so-called Arthur-Campoli conditions. In \cite[Thm.~3.6]{BSpwspace} it is shown that the Arthur Paley-Wiener space is isomorphic to one defined in terms of the single minimal parabolic subgroup $P_0.$ We shall now describe the Arthur-Campoli relations in the context of the operator valued Fourier transform $f \mapsto \hat f(P_0).$ \par For $f \in C^\infty_c(G,K)$ and $\xi \in \dM,$ the Fourier transform $\hat f(P_0, \xi) \in \cO(\fadc)\otimes \SS(P_0, \xi)$ is defined by \begin{equation} \label{e: defi Fourier} \hat{f}(P_0,\xi,\gl)\coleq \int_G f(x)\pi_{P_0,\xi,\lambda}(x)\,dx,\quad \gl\in\fa_\iC^*. \end{equation} Then $f \mapsto \hat f(P_0)$ maps $C_c^\infty(G,K)$ into $\cO(\fa_\iC^*) \otimes \cS(P_0).$ \par We define $\data$ to be the set of $4$-tuples of the form $(\xi,\psi,\lambda,u)$ with $\xi \in \dM,$ $\psi \in \cS(P_0, \xi)^*_{K\times K}$, $\gl \in \fadc$ and $u \in S(\fa^*).$ An Arthur-Campoli sequence in $\data$ is defined to be a finite family $(\xi_i, \psi_i, \lambda_i, u_i)$ in $\data$ such that $$ \sum_i\langle\pi_{P_0,\xi_i,\lambda_i;u_i}(x),\psi_i\rangle=0, \qquad x \in G. $$ By integration over $x$ it follows that this condition is equivalent to the condition that \begin{equation}\label{e: acs} \sum_i \langle \hat f(P_0, \xi_i, \gl_i;u_i), \psi_i \rangle = 0, \qquad f \in C_c^\infty(G,K). \end{equation} \begin{definition} \label{d: AC relations \rm A function $\varphi \in \O(\a_\iC^*)\otimes\SS(P_0)$ is said to satisfy the Arthur-Campoli relations if $$ \sum_i\langle\varphi_{\xi_i,\lambda_i;u_i},\psi_i\rangle=0 $$ for any Arthur-Campoli sequence $(\xi_i,\psi_i,\lambda_i,u_i)$ in $\data.$ \end{definition} \subsection{Reformulation of the Arthur-Campoli relations In the following, $\O_0$ denotes the ring of germs at $0$ of holomorphic functions defined on a neighborhood of $0$ in $\a_\iC^*.$\par Let $E$ be a finite dimensional $\O_0$-module and let $\Xi \subset \dM$ and $\Lambda \subset \fa_\iC^*$ be finite sets. We define the representation $\pi_{E,\Xi,\Lambda}$ of $G$ by \begin{equation} \label{e: defi pi E Xi gL} \pi_{E,\Xi,\Lambda} = \oplus_{(\xi, \gl) \in \Xi \times \gL}\; \pi_{P_0, \xi, \gl}^{(E)}. \end{equation} Note that this representation is admissible and of finite length. Its underlying space is given by \begin{equation} \label{e: defi V E Xi gl} V_{E, \Xi, \gL} = \oplus_{(\xi, \gl)\in \Xi \times \gL} \;\; E \otimes C^\infty(K\col \xi). \end{equation} For each element $\gf \in \cO(\fa_\iC^*) \otimes \cS(P_0)$ we define the $K\times K$-finite endomorphism $\varphi_{E,\Xi,\Lambda}$ of $V_{E, \Xi, \gL}$ by taking the similar direct sum \begin{equation} \label{e: defi gf E Xi gL} \gf_{E, \Xi, \gL} \coleq \oplus_{(\xi,\gl) \in \Xi \times \gL} \;\; \gf^{(E)}_{\xi}(\gl). \end{equation} We note that $\pi_{E, \Xi, \gL}(C_c^\infty(G,K))$ is a subset of the space $\Endo(V_{E,\Xi, \gL})$ and agree to write $\pi_{E, \Xi, \gL}(C_c^\infty(G,K))^\perp$ for its annihilator in the space $$ \Endo(V_{E, \Xi, \gL})^*_{K\times K}. $$ \begin{prop}\label{p: reformulation ac Let $\phi\in \O(\a_\iC^*)\otimes\SS(P_0).$ Then the following conditions are equivalent: \begin{enumerate} \itema $\phi$ satisfies the Arthur-Campoli relations; \itemb for every finite dimensional $\O_0$-module $E,$ every pair of finite sets $\Xi \subset \dM,\,$ $\Lambda \subset \a_\iC^*$ and all $\Psi\in\pi_{E,\Xi,\Lambda}(C^\infty_c(G,K))^\perp,$ $$ \langle\phi_{E,\Xi,\Lambda},\Psi\rangle=0. $$ \end{enumerate} \end{prop} We prove the result through a number of lemmas. In the following results a complication is caused by the circumstance that $\Endo(V_{E,\Xi,\gL})_{K\times K}$ is not the direct sum of the spaces $\Endo(V_{E, \xi, \gl})_{K\times K},$ but rather that of the spaces $\Hom(V_{E, \xi_1, \gl_1}, V_{E, \xi_2, \gl_2})_{K\times K},$ for $\xi_1, \xi_2 \in \Xi$ and $\gl_1, \gl_2 \in \gL.$ For $(\xi, \gl) \in \Xi \times \gL,$ let $$ \inj_{\xi, \gl}: \; \Endo(V_{E, \xi, \gl})_{K\times K} \longrightarrow \Endo(V_{E, \Xi, \gL})_{K\times K} $$ denote the associated embedding, and let $$ \pr_{\xi, \gl}:\;\; \Endo(V_{E, \Xi, \gL})_{K\times K} \longrightarrow \Endo(V_{E, \xi, \gl})_{K\times K} $$ denote the associated projection map. \begin{lemma} \label{l: ac data associated with Psi Let $E, \Xi, \gL$ be as above. Then for each $\Psi \in \Endo(V_{E, \Xi, \gL})^*_{K\times K}$ there exists a finite sequence $(\xi_i, \psi_i, \gl_i, u_i)$ in $\data$ such that for all $\gf \in \cO(\fa^*_\iC) \otimes \cS(P_0),$ \begin{equation} \label{e: ac data associated with Psi} \inp{\gf_{E, \Xi, \gL}}{\Psi} = \sum_{i} \inp{\gf_{\xi_i, \gl_i ; u_i}}{\psi_i}. \end{equation} \end{lemma} \begin{proof Put $$ \cS(P_0, \Xi, \gL) \coleq \oplus_{(\xi,\gl)\in\Xi\times\gL}\cS(P_0,\xi). $$ We observe that $\Endo(V_{E, \Xi, \gL})^*_{K\times K}$ may be viewed as the direct sum of the $K \times K$-submodule $\Endo(E)^* \otimes \cS(P_0, \Xi, \gL)^*_{K\times K}$ and a unique $K\times K$-submodule $\End{E}^*\otimes \cT$ (consisting of the `cross terms'). Since every element of $\End{E}^*\otimes \cT$ annihilates $\gf_{E, \Xi, \gL},$ we may assume that $\Psi \in \End{E}^*\otimes\cS(P_0, \Xi, \gL)^*_{K\times K}.$ By linearity, we may then reduce to the situation that $\Xi = \{\xi\}$ and $\gL = \{\gl\}.$ Again by linearity we may assume that $\Psi$ is of the form $\eta \otimes \psi,$ with $\eta \in \Endo(E)^*$ and $\psi \in \cS(P_0, \xi)^*_{K\times K}.$ Let $u \in S(\fa^*)$ be associated with $\eta$ as in Lemma \ref{l: dual of End E}. Then for all $\gf \in \cO(\fa_\iC^*) \otimes \cS(P_0),$ $$ \inp{\gf_{E, \xi, \gl}}{\Psi} = \inp{\gf_{\xi, \gl}^{(E)}}{\eta \otimes \psi} = \inp{\gf_{\xi, \gl ; u}}{\psi}. $$ This finishes the proof. \end{proof} \begin{lemma} \label{l: E associated with ac data Let $(\xi_i, \psi_i, \gl_i, u_i)$ be a finite sequence in $\data.$ Then there exists a finite dimensional $\cO_0$-module $E$, and finite sets $\Xi \subset \dM,$ $\gL \subset \fadc$ and an element $\Psi \in \Endo(V_{E, \Xi, \gL})^*_{K\times K}$ such that for all $\gf \in \cO(\fa^*_\iC) \otimes \cS(P_0),$ \begin{equation} \label{e: E associated with ac data} \sum_{i} \inp{\gf_{\xi_i, \gl_i ; u_i}}{\psi_i} = \inp{\gf_{E, \Xi, \gL}}{\Psi}. \end{equation} \end{lemma} \begin{proof Let $E$ and $\eta_i \in \Endo(E)^*$ be associated to the finite sequence $u_i$ as in Lemma \ref{l: E associated with u}. Then $\Psi_i = \eta_i \otimes \psi_i$ belongs to $\Endo(E)^* \otimes \cS(P_0, \xi_i)^*_{K\times K} \simeq \Endo(V_{E, \xi_i,\gl_i})_{K\times K}^*.$ Let $\Xi$ be the finite set of all $\xi_i$ and let $\gL$ be the finite set of all $\gl_i.$ Let $\pr_{\xi_i, \gl_i}$ be defined as above. We define $\Psi \in \Endo(V_{E, \Xi, \gL})^*_{K \times K}$ by $$ \Psi =\sum_i \pr_{\xi_i, \gl_i}^* ( \eta_i \otimes \psi_i ). $$ Then for all $\gf \in \cO(\fadc) \otimes \cS(P_0)$ we have \begin{eqnarray*} \sum_i \inp{\gf_{\xi_i,\gl_i ; u_i}}{\psi_i} &=& \sum_i \inp{\gf^{(E)}_{\xi_i,\gl_i}}{\eta_i \otimes \psi_i}\\ & =& \sum_i \inp{\pr_{\xi_i, \gl_i}\gf_{E, \Xi, \gL}}{\eta_i \otimes \psi_i} \\ & = & \inp{\gf_{E, \Xi, \gL}}{\Psi}. \end{eqnarray*} \end{proof} \begin{proof}[Proof of Proposition \ref{p: reformulation ac} Let $\phi$ be as stated and assume (a). Let $E,\,\Xi,\, \Lambda$ and $\Psi$ be as asserted in (b). In particular, $\Psi \in \Endo(V_{E, \Xi, \gL})^*_{K\times K}.$ Let $(\xi_i, \psi_i, \gl_i, u_i)$ be a sequence in $\data,$ associated to $\Psi$ as in Lemma \ref{l: ac data associated with Psi}. Then using the relation (\ref{e: ac data associated with Psi}) with $\gf = \hat f(P_0)$ for $f \in C_c^\infty(G,K),$ we see that $(\xi_i, \psi_i, \gl_i, u_i)$ is an Arthur-Campoli sequence (see \eqref{e: acs}). Hence, $$ \inp{\phi_{E, \Xi, \gL}}{\Psi} = \sum_i \inp{ \phi_{\xi_i,\gl_i; u_i}}{\psi_i} = 0. $$ We have proved (b). \par Conversely, assume (b) and let $(\xi_i, \psi_i, \gl_i, u_i)$ be an Arthur-Campoli sequence in $\data.$ Let $E, \Xi, \gL, \Psi$ be associated with this sequence as in Lemma \ref{l: E associated with ac data}. Then it follows from \eqref{e: E associated with ac data} with $\gf = \hat f(P_0)$ for $f \in C_c^\infty(G,K)$ that $\Psi$ belongs to $\pi_{E, \Xi, \gL}(C_c^\infty(G,K))^\perp.$ This implies that $$ \sum_i \inp{\phi_{\xi_i,\gl_i;u_i}}{\psi_i} = \inp{\phi_{E, \Xi, \gL}}{\Psi} = 0. $$ Hence (a). \end{proof} \section{Delorme's interwining conditions}\label{s: Delorme's PW \subsection{Successive derivatives}\label{s: succder Let $\v$ be a finite dimensional vector space over $\R,$ let $\Omega \subset \fv_\iC^*$ be an open subset and let $V$ be a quasi-complete locally convex space. Following Delorme \cite{delormepw}, we define, for $\Phi \in \cO(\Omega, \End{V})$ and $\eta \in \fv_\iC^*,$ the holomorphic function $\Phi^{(\eta)}: \Omega \to \End{V\oplus V}$ by $$ \Phi^{(\eta)}(\lambda)(v_1,v_2)\coleq \left(\Phi(\lambda)v_1+\frac{d}{dz}(\Phi(\lambda+z\eta)v_2)_{|z=0}\;,\;\Phi(\lambda)v_2\right), $$ for $\gl \in \Omega$ and $v_1,v_2 \in V.$ Still following \cite{delormepw}, we define, for any finite sequence $\eta=(\eta_1,\dots,\eta_N)$ in $\v_\iC^*$, the iterated derivative $$ \Phi^{(\eta)} \coleq (\cdots (\Phi^{(\eta_N)})^{(\eta_{N-1})}\cdots)^{(\eta_1)} $$ of $\Phi$ along $\eta.$ Then $\Phi^{(\eta)}$ is a holomorphic function on $\Omega$ with values in $\End{V^{(\eta)}},$ where $V^{(\eta)}$ denotes the direct sum of $2^N$ copies of $V.$ \par Now assume that $V$ is Fr\'echet (or more generally, barrelled). If $\pi$ is a holomorphic family of continuous representations of $G$ in $V$ over the parameter set $\fv_\iC^*$ then it follows by application of the methods of \cite{delormepw} that for each $\gl\in \Omega,$ $$ \pi^{(\eta)}_\gl(x): = \pi(x)^{(\eta)}(\gl), \qquad x \in G, $$ defines a continuous representation of $G$ in $V^{(\eta)}.$ \subsection{The intertwining conditions} \label{s: the intertwining conditions Let $\cD$ be the set of $4$-tuples $\delta=(P,\xi,\lambda,\eta)$, with $P\in\cP(A),$ $\xi\in\dMPds$, $\lambda\in\a_{P\iC}^*$ and $\eta$ a finite sequence in $\a_{P\iC}^*.$ Given $\gd\in\cD,$ we define the representation $\pi_\gd$ of $G$ in $V_{\pi_\gd}\coleq C^{\infty}(K\col \xi)^{(\eta)}$ by $$ \pi_\gd \coleq \pi_{P, \xi, \gl}^{(\eta)} $$ For $P \in \cP(A)$ we define the space $ \holspace_P\coleq \cO(\faPdc) \otimes \cS(P), $ with $\cS(P)$ defined as in (\ref{e: defi cSP}). Furthermore, we put \begin{equation} \label{e: holspace} \holspace:= \oplus_{P\in\cP(A)} \;\; \holspace_P. \end{equation} Given $\gf\in \holspace$ and $\gd = (P, \xi, \gl, \eta) \in \cD,$ we define $\gf_\gd \in \Endo(C^\infty(K\col \xi)^{(\eta)})$ in a similar fashion as $\pi_\gd,$ by $$ \gf_\gd\coleq \gf_{P, \xi}^{(\eta)}(\gl). $$ Finally, given a sequence $\gd = (\gd_1, \ldots, \gd_N)$ of data from $\cD,$ we define $\pi_{\gd}\coleq \pi_{\gd_1} \oplus \cdots \oplus \pi_{\gd_N},$ $V_{\pi_{\gd}} \coleq V_{\pi_{\gd_1}} \oplus \cdots \oplus V_{\pi_{\gd_N}}$ and $\gf_\gd\coleq \gf_{\gd_1} \oplus \cdots \oplus \gf_{\gd_N}.$ \par \begin{definition}\label{d: Delorme's intertwining conditions \rm We say that a function $\gf \in \holspace$ satisfies Delorme's intertwining conditions (see \cite[Definition 3 (4.4)]{delormepw}) if \vspace{3pt} \begin{enumerate} \itema for every $N \in \Z_+$ and each $\gd \in \cD^N$ the function $\gf_\gd$ preserves all invariant subspaces of $\pi_{\gd};$ \itemb for all $N_1, N_2 \in \Z_+,$ all $\gd_1 \in \cD^{N_1}$ and $\gd_2 \in\cD^{N_2},$ and any two sequences of closed invariant subspaces $U_j \subset V_j$ for $\pi_{\gd_j},$ the induced maps $\bar \gf_{\gd_j} \in \Endo(V_j/U_j)$ are intertwined by all intertwining operators $T: V_1/U_1 \to V_2/U_2.$ \end{enumerate} The space of functions $\gf \in \holspace$ satisfying (a) and (b) is denoted by $\holspace(\cD).$ \end{definition} \subsection{A simplification of the intertwining conditions In this section we will show that condition (b) of Definition \ref{d: Delorme's intertwining conditions} is in fact a consequence of condition (a) of the same definition. \begin{lemma}\label{l: one Delorme condition sufficient Let $\gf \in \holspace$. Then $\gf \in \holspace(\cD)$ if and only if $\gf$ satisfies condition (a) of Definition \ref{d: Delorme's intertwining conditions}. \end{lemma} \begin{proof Assume that $\gf$ satisfies condition (a) of the mentioned definition. Then we must show that $\gf$ satisfies condition (b) as well. \par Let $\delta_j,\, \pi_{\gd_j}, \,U_j,\, V_j$ be as in (b), for $j =1,2.$ Let $T: V_1/U_1 \to V_2/U_2$ be an intertwining operator. Let $p_j: V_j \to V_j/U_j$ denote the canonical projection, for $j =1,2,$ and let $p \coleq p_1 \oplus p_2.$ Since $T$ is equivariant, its graph $W$ is an invariant subspace of $V_1/U_1 \oplus V_2/U_2.$ Hence, $p^{-1}(W)$ is an invariant subspace of $V_1 \oplus V_2.$ Since $\varphi$ satisfies (a), it follows that $p^{-1}(W)$ is a $\varphi_{\delta_1} \oplus \varphi_{\delta_2}$-invariant subspace of $V_1 \oplus V_2.$ This in turn is easily seen to imply that $T\circ \bar{\varphi}_{\delta_1} = \bar{\varphi}_{\delta_2}\circ T.$ \end{proof} \subsection{Reduction to a single minimal parabolic subgroup In this section, we will show that the space $\holspace(\cD)$ of functions satisfying Delorme's intertwining conditions is naturally isomorphic to a space of functions defined in terms of just the minimal parabolic subgroup $P_0.$ We denote by $\cD_{P_0}$ the set of $4$-tuples $(P,\xi,\lambda,\eta)$ in $\cD$ for which $P=P_0$. \begin{definition} \label{d: restricted intertwining conditions \rm We say that a function $\gf \in \holspace$ satisfies Delorme's intertwining conditions associated with $P_0$ if conditions (a) and (b) of Definition \ref{d: Delorme's intertwining conditions} are valid with everywhere $\cD$ replaced by $\cD_{P_0}.$ The space of functions $\gf \in \holspace$ satisfying all such intertwining conditions is denoted by $\cF(\cD_{P_0}).$ \end{definition} The following analogue of Lemma \ref{l: one Delorme condition sufficient} is now valid, with essentially the same proof. \begin{lemma} \label{l: one restricted intertwining condition suffices Let $\gf\in \holspace.$ Then $\gf \in \cF(\cD_{P_0})$ if and only if $\gf$ satisfies condition (a) of Definition \ref{d: Delorme's intertwining conditions} for every $N\in\Z_+$ and all $\gd\in (\cD_{P_0})^{N}.$ \end{lemma} We consider the component $$ \holspace_{P_0} \coleq \cO(\fadc) \otimes \cS(P_0) $$ of the direct sum $\holspace$ defined in (\ref{e: holspace}), and denote the natural projection $\holspace \to \holspace_{P_0}$ by $\pr.$ Moreover, we define $$ \holspace_{P_0}(\cD_{P_0}) \coleq \holspace_{P_0} \cap \holspace(\cD_{P_0}). $$ \medbreak \begin{prop} \label{p: reduction to one minimal The natural projection $\pr: \holspace \to \holspace_{P_0}$ restricts to a linear isomorphism from $\holspace(\cD)$ onto $\holspace_{P_0}(\cD_{P_0}).$ \end{prop} \begin{proof If $\gf \in \holspace$ and if $\gd$ is a finite sequence in $\cD_{P_0},$ then $\gf_\gd = (\gf_{P_0})_\gd.$ Hence $\gf \in \holspace(\cD_{P_0})$ if and only if $\pr (\gf) \in \holspace(\cD_{P_0}).$ From $\cD_{P_0} \subset \cD$ it follows that $\holspace(\cD) \subset \holspace(\cD_0).$ Therefore, $\pr$ maps $\holspace(\cD)$ into $\holspace_{P_0} \cap \holspace(\cD_{P_0}) = \holspace_{P_0}(\cD_{P_0}).$ The proof will be completed by showing that the restricted projection $\pr_{\cD} \coleq \pr|_{\holspace(\cD)}: \holspace(\cD) \to \holspace_{P_0}(\cD_{P_0})$ is a linear isomorphism. We will do this by defining a map $$ \tau: \;\;\holspace_{P_0} \too \holspace $$ which will turn out to induce a two-sided inverse for $\pr_\cD.$ Each parabolic subgroup $P \in \cP(A)$ has a Langlands decomposition $P = M_P A_P N_P.$ Here $M_P$ is a real reductive group of the Harish-Chandra class, with Cartan decomposition $M_P = (M_P\cap K) \exp(\fm_P \cap \fp).$ Moreover, ${}^*\fa\coleq \fm_P \cap \fa$ is maximal abelian in $\fm_P \cap \fp,$ and the centralizer of ${}^*\fa$ in $K_P\coleq M_P \cap K$ equals $M = Z_K(\fa).$ We select a minimal parabolic subgroup $$ {}^*Q = M\, {}^*\!A \, {}^*\!N $$ of $M_P.$ In addition, we fix a minimal parabolic subgroup $Q$ of $G$ containing $A$ such that ${}^*Q = Q \cap M_P,$ and we fix an element $w \in N_K(\fa)$ such that $P_0 = w^{-1} Q w.$ In case $P = P_0$, we agree to make the special choice $Q = P_0$ and $w =e.$ Then by the subrepresentation theorem applied to the group $M_P,$ for each $\xi \in \dMPds$ we may fix a representation $\gs \in\dM$ and an element $\mu \in {}^*\fadc$ such that $\xi$ is equivalent to a subrepresentation of the parabolically induced (smooth) representation $\pi^{M_P}_{{}^*Q, \gs, \mu}$ of $M_P.$ In addition, we fix an $M_P$-equivariant embedding $$ j_{\xi}:\; (\xi, V_\xi) \;\;\embeds \;\; (\pi^{M_P}_{{}^*Q, \gs, \mu}, C^\infty(K \cap M_P: \gs)). $$ Through induction by stages, this embedding induces an embedding \begin{equation} \label{e: j xi P} j_{\xi P}^G:\; C^{\infty}(K\col \xi) \too C^\infty(K\col \gs) \end{equation} which intertwines $\pi_{P, \xi, \gl}$ with $\pi_{Q, \gs, \mu + \gl},$ for all $\gl \in \faPdc.$ Here ${}^*\fa_{\iC}^*$ and $\fa_{P\iC}^*$ are viewed as subspaces of $\fadc$ via the direct sum decomposition $\fa = {}^*\fa \oplus \fa_P.$ Thus, (\ref{e: j xi P}) is a morphism in the category $\HFGO$, see Definition \ref{d: sec59}. In the special case $P = P_0,$ we agreed that $Q =P_0$ and $w = e.$ In this case, we have $\mu = 0,$ and as $\gs$ must be equivalent to $\xi,$ it follows that $\gs = w\cdot \gs = \xi.$ It is now readily verified that $j_{\xi P}^G$ is the identity map of $C^\infty(K\col \xi).$ Left translation by $w$ induces a topological linear isomorphism \begin{equation} \label{e: L w} L(w): \; C^\infty(K\col\gs) \too C^\infty(K\col w\gs), \end{equation} which intertwines the induced representation $\pi_{Q, \gs, \mu+ \gl}$ with $\pi_{P_0, w \gs, w(\mu + \gl)},$ for all $\gl \in \faPdc.$ Here $w\gs$ denotes the representation of $M$ in $V_\gs,$ given by $w\gs(m) =\gs(w^{-1}mw).$ We denote by $(w\cdot \gs, V_{w \cdot \gs})$ the unique element of $\dM$ which is equivalent to $(w\gs, V_\gs).$ Fix an equivalence $t_w: V_\gs \to V_{w\cdot \gs}$ and denote the induced map $C^\infty(K\col w\gs) \to C^\infty(K \col w\cdot\gs)$ by $T_w.$ Then writing $\cL_w = T_w\after L(w),$ we obtain a continuous linear injection \begin{equation} \label{e: embedding K xi to K w gs} \cL_w\after j_{\xi P}^G:\;\; C^\infty(K\col \xi) \too C^\infty(K\col w\cdot \gs) \end{equation} which intertwines the representation $\pi_{P,\xi, \gl}$ with $\pi_{P_0, w\cdot \gs, w(\mu + \gl)},$ for all $\gl \in \faPdc.$ In the special case $P = P_0,$ where $Q = P_0, w =e, \mu = 0$ and $\gs = \xi,$ we may take $t_w = \id_{V_\xi}$ and then $\cL_w$ and hence (\ref{e: embedding K xi to K w gs}) become the identity map of $C^\infty(K\col \xi).$ We are now ready to define the map $\tau.$ Let $P, \xi, \gl$ be as above, and let $\gf\in \holspace_{P_0}(\cD_0).$ Then $\gf_{P_0, w\cdot\gs, w(\mu + \gl)}$ leaves the invariant subspace ${\rm im}(\cL_w\after j_{\xi P}^G)$ for the representation $\pi_{P_0, w\cdot \gs, w(\mu +\gl)}$ invariant by condition (a) of Definition \ref{d: Delorme's intertwining conditions}, so that we may define $$ \tau(\gf)_{P, \xi,\gl}: \;\; C^\infty(K\col \xi) \to C^\infty(K\col \xi) $$ to be the unique linear map such that the following diagram commutes \begin{equation} \label{e: comm diag t gf} \xymatrix{ \ar[d]_{\tau(\gf)_{P,\xi,\gl}} C^\infty(K\col \xi) \ar[rr]^{\cL_w\after j_{\xi P}^G} &&\ar[d]^{\gf_{P_0, w\cdot\gs,w(\mu + \gl)}} C^\infty(K\col w\cdot\gs) \\ C^\infty(K\col \xi ) \ar[rr]_{\cL_w\after j_{\xi P}^G} && C^\infty(K\col w\cdot \gs). } \end{equation} Since $\gf_{P_0, w\cdot\gs} \in \cO(\fadc) \otimes \End{C^\infty(K\col w\cdot\gs)}_{K\times K},$ it is readily seen that $\tau(\gf)_{P,\xi}$ defines a holomorphic function on $\faPdc,$ with values in $\cS(P, \xi).$ Accordingly, $\tau$ defines a linear map $\holspace_{P_0} \to \holspace.$ We will now finish the proof by showing that \begin{enumerate} \item[{\rm (i)}] $\tau$ maps $\holspace_{P_0}(\cD_{P_0})$ into $\holspace(\cD),$ \item[{\rm (ii)}] $\tau$ restricts to a linear isomorphism $\tau_\cD: \holspace_{P_0}(\cD_{P_0})\to \holspace(\cD)$ which is a two-sided inverse for $\pr_\cD.$ \end{enumerate} We will first establish (i). Let $\gf \in \holspace_{P_0}(\cD_0),$ and let $N \in \Z_+$ and $\gd \in \cD^N.$ We claim that it suffices to show that there exists a $\gd' \in \cD_{P_0}^N$ and a linear embedding $j: V_{\gd} \to V_{\gd'}$ intertwining $\pi_\gd$ with $\pi_{\gd'}$ such that \begin{equation} \label{e: intertwining embedding j} \gf_{\gd'} \after j = j \after \tau(\gf)_\gd. \end{equation} Indeed assume the claim to hold, and let $W \subset V_\gd$ be an invariant subspace. Then $j(W)$ is an invariant subspace of $V_{\gd'}.$ Moreover, since $\gf \in \holspace_{P_0}(\cD_0),$ it follows that $\gf_{\gd'}$ leaves $j(W)$ invariant. As $j$ is injective, it now follows from (\ref{e: intertwining embedding j}) that $\tau(\gf)_\gd$ leaves $W$ invariant. Thus, the validity of the claim would imply that the above condition (i) holds. We turn to the proof of the claim. It clearly suffices to prove the claim in case $N =1,$ so that $\gd \in \cD.$ Thus, $\gd$ is a $4$-tuple of the form $(P, \xi, \gl_0, \eta)$ with notation as in the beginning of Section \ref{s: the intertwining conditions}. In particular, $\eta$ is a finite sequence in $\faPdc.$ Let $Q, \gs, \mu, w$ be associated with the data $P, \xi$ as in the first part of this proof, where the definition of $\tau$ was given. Then the injective linear map (\ref{e: embedding K xi to K w gs}) intertwines $\pi_{P,\xi, \gl}$ with $\pi_{P_0, w\cdot\gs, w(\mu + \gl)}$ for all $\gl \in \faPdc.$ It follows that the map $$ j:\;\; C^\infty(K\col \xi)^{(\eta)} \too C^\infty(K\col w\cdot\gs)^{(\eta)} $$ induced by (\ref{e: embedding K xi to K w gs}) interwines $\pi_\gd$ with $\pi_{\gd'},$ where $\gd' = (P_0, w\cdot\gs, w(\gl_0 + \mu), w\eta).$ Moreover, the commutativity of the diagram (\ref{e: comm diag t gf}) implies that $$ j \after \tau(\gf)_{P,\xi, \gl}^{(\eta)} = \gf_{P_0, w\cdot\gs, w(\mu + \gl)}^{(w\eta)} \after j, $$ or, abbreviated, $j \after \tau(\gf)_\gd = \gf_{\gd'} \after j.$ This establishes the claim. We now consider the induced map $\tau_\cD: \holspace_{P_0}(\cD_0) \to \holspace(\cD),$ obtained by restriction of $\tau,$ and will finish the proof by establishing (ii). Let $\gf \in \holspace_{P_0}(\cD_{P_0}).$ Moreover, let $Q, \gs, \mu, w$ be associated to $P = P_0$ as above. By the special choices we made for this particular parabolic subgroup, the diagram (\ref{e: comm diag t gf}) becomes $$ \xymatrix{ \ar[d]_{\tau(\gf)_{P_0,\xi,\gl}} C^\infty(K\col \xi) \ar[rr]^{ \id } &&\ar[d]^{\gf_{P_0, \xi ,\gl} \;\;\;\;\;\;\;\;\;\;\;\;} C^\infty(K\col \xi) \\ C^\infty(K\col \xi ) \ar[rr]_{\id } && C^\infty(K\col \xi), } $$ It follows that $\pr\after \tau (\gf) = \gf,$ and we see that $\tau_\cD$ is a right inverse to $\pr_\cD.$ We will finish the proof by showing that $\tau_\cD$ is also a left inverse. Let $\psi \in \holspace(\cD),$ and let $P \in \cP(A)$ and $\xi \in \dMPds.$ Let $Q, \gs, \mu, w$ be as above. As $\psi$ satisfies condition (b) of Definition \ref{d: Delorme's intertwining conditions}, it follows that the following diagram commutes, for all $\gl \in \faPdc,$ $$ \xymatrix{ \ar[d]_{\psi_{P,\xi,\gl}} C^\infty(K\col \xi) \ar[rr]^{\cL_w \after j_{\xi P}^G} &&\ar[d]^{\psi_{P_0, w\cdot\gs,w(\mu + \gl)} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!} C^\infty(K\col w\cdot\gs) \\ C^\infty(K\col \xi ) \ar[rr]_{\cL_w\after j_{\xi P}^G} && C^\infty(K\col w\cdot\gs). } $$ By comparison with (\ref{e: comm diag t gf}) we see that $\psi_{P,\xi} = \tau(\psi_{P_0})_{P, \xi}.$ Hence, $\tau\after\pr(\psi) = \psi,$ for $\psi \in \cF(\cD).$ Therefore, $\tau_\cD $ is a left inverse to $\pr_\cD.$ \end{proof} \subsection{Reformulation in our setting We shall now compare Delorme's derivation process with the process defined in the present paper. In the following we shall identify $V \oplus V$ with $\C^2 \otimes V$ via the map $(v_1, v_2) \mapsto (1,0) \otimes v_1 + (0,1) \otimes v_2.$ This identification induces an identification $\Endo(V \oplus V) \simeq \Endo(\C^2) \otimes \Endo(V).$ \par Let now $\eta \in \fadc$ and let $\Omega$ be an open subset of $\fa_\iC^*.$ For $\gf\in \cO(\Omega)$ we define $D^{(\eta)}\gf \in \cO(\Omega) \otimes \Endo(\C^2)$ by $$ D^{(\eta)}\gf(\gl) = \left( \begin{array}{cc} \gf(\gl) & \gf(\gl; \eta) \\ 0 & \gf(\gl) \end{array} \right). $$ Let $V$ be a quasi-complete locally convex space. Then for $\Phi \in \cO(\Omega) \otimes \Endo(V),$ Delorme's derivative $\Phi^{(\eta)}$ is given by $$ \Phi^{(\eta)} = (D^{(\eta)} \otimes \id_{\Endo(V)}) \Phi. $$ Here we note that \begin{equation} \label{e: V eta as tensor product} V^{(\eta)} = \C^2\otimes V, \end{equation} so that $\Endo(V^{(\eta)}) = \Endo(\C^2) \otimes \Endo(V).$ It follows that Delorme's differentiation map $(\dotvar)^{(\eta)}$ from $\cO(\Omega, \Endo(V)) $ to $\cO(\Omega, \Endo(V^{(\eta)}))$ is the unique continuous linear extension of the map $D^{(\eta)} \otimes \id_{\Endo(V)}.$ \par Now assume that $\eta \neq 0$ and let $\cI_\eta$ be the cofinite ideal of $\cO_0 = \cO_0(\fa_\iC^*)$ defined in (\ref{e: idealcodim2}). Then $\cI_\eta$ has codimension $2$ by Lemma \ref{l: codimension two}. More precisely, let $X \in \fac$ be such that $\eta(X) = 1.$ Then the map $ (z_1, z_2) \mapsto z_1 + z_2 X + \cI_\eta $ defines a linear isomorphism $\kappa$ from $\C^2$ onto the $\cO_0$-module $$ E_\eta\coleq \cO_0/\cI_\eta. $$ Now assume that $V$ is barrelled. Then in view of (\ref{e: V eta as tensor product}) the isomorphism $\kappa$ induces an isomorphism from $\cO(\Omega, \Endo(V^{(\eta)}))$ onto $\cO(\Omega, \Endo(E_\eta \otimes V)),$ denoted $\kappa_*.$ \begin{lemma With notation as above, the following diagram commutes $$ \xymatrix{ \O(\Omega,\End{V})\ar@{->}[d]_{(\dotvar)^{(\eta)}}\ar@{->}[drr]^{\diffD^{(E_{\eta})}}&&\\ \O(\Omega,\End{V^{(\eta)}})\ar@{->}[rr]_{\kappa_*} && \O(\Omega,\End{E_{\eta}\otimes V}) } $$ \end{lemma} \begin{proof By the same calculation as in the proof of Lemma \ref{sec40}, it follows that the multiplication action $m(\gf)$ of an element $\gf \in \cO_0$ on $E_\eta$ with respect to the basis $\bar 1, \bar X$ is given by the matrix $$ M(\gf)\coleq \left( \begin{array}{cc} \gf(0) & \gf(0; \eta) \\ 0 & \gf(0) \end{array} \right). $$ Let $\widetilde \kappa$ be the isomorphism $\Endo(\C^2) \to \Endo(E_\eta)$ induced by $\kappa.$ Then it follows that for all $\gf \in \cO(\Omega)$ and $\gl \in \Omega,$ $$ J^{(E_\eta)} \gf(\gl) = m(\gg_0(T_\gl\gf)) = \widetilde \kappa \, (\, M(\gg_0(T_\gl\gf))\, ) = \widetilde\kappa\,(\, D^{(\eta)}\gf(\gl)\,). $$ This immediately implies the commutativity of the diagram in case $V = \C.$ From this we see that the diagram commutes with the spaces $\cO(\Omega, \Endo(W))$ replaced by the subspaces $\cO(\Omega) \otimes \Endo(W),$ for $W$ equal to $V,\, V^{(\eta)}$ or $E_\eta\otimes V.$ By density of the mentioned spaces and continuity of the maps involved, the result follows. \end{proof} Given a finite sequence $\eta = (\eta_1,\ldots, \eta_N)$ of elements in $\fadc,$ we define the $\cO_0$-module $E_\eta\coleq E_{\eta_1} \otimes \cdots \otimes E_{\eta_N}.$ \begin{cor For every finite sequence $\eta$ as above, and all $\xi \in \dM$ and $\gl \in \fadc,$ $$ \pi_{P_0, \xi, \gl}^{(\eta)} \simeq \pi^{(E_\eta)}_{P_0, \xi, \gl}. $$ \end{cor} \begin{proof For $\eta$ of length one this follows from the above lemma. For arbitrary sequences it follows from the recurrent nature of the definition of $\pi^{(\eta)}$ and Proposition \ref{p: iterateder}. \end{proof} \begin{definition} \label{d: End sharp \rm Let $V$ be a Harish-Chandra module. We denote by $\Endo(V)^\#$ the space of endomorphisms $\gf \in \Endo(V)_{K\times K}$ such that for every $n \in \N$ the product map $\gf^{\times n} \in \Endo(V^{\times n})$ leaves all $(\fg, K)$-invariant subspaces of $V^{\times n}$ invariant. \par If $(\pi, V_\pi)$ is an admissible representation of $G$ of finite length, we define $\Endo(\pi)^\# \coleq \Endo((V_{\pi})_K)^\#.$ \end{definition} \par For $E$ a finite dimensional $\cO_0$-module (where $\cO_0 = \cO_0(\fadc)$), and for $\Xi \subset \dM$ and $\gL \subset \fadc$ finite subsets, we recall the definition of the representation $\pi_{E, \Xi, \gL}$ by (\ref{e: defi pi E Xi gL}). Moreover, given $\gf \in \holspace_{P_0} = \cO(\fadc) \otimes \cS(P_0),$ we recall the definition of $\gf_{E, \Xi, \gL}$ in (\ref{e: defi gf E Xi gL}). \begin{lemma}\label{l: stability of ic under subq Let $E,E'$ be finite dimensional $\cO_0$-modules such that $E$ is a subquotient of $E'.$ Let $\Xi \subset \Xi'\subset \dM$ be finite subsets and let $\gL \subset \gL'\subset \fadc$ be finite subsets. Then for all $\gf \in \holspace_{P_0},$ $$ \gf_{E', \Xi', \gL'} \in \Endo(\pi_{E', \Xi', \gL'})^\# \Longrightarrow \gf_{E, \Xi, \gL} \in \Endo(\pi_{E, \Xi, \gL})^\# $$ \end{lemma} \begin{proof The result follows from the crucial observation that the representation $\pi_{E, \Xi, \gL}$ is a subquotient of $\pi_{E', \Xi', \gL'}$ and the map $\gf_{E', \Xi', \gL'}$ induces the map $\gf_{E, \Xi, \gL}.$ Obviously, it suffices to prove this observation for $\Xi = \Xi'$ and $\gL = \gL'.$ In this case, we put, for $\gl\in\a_\iC^*,$ $$ \pi_{\gl} \coleq \oplus_{\xi \in \Xi} \;\;\pi_{P_0, \xi, \gl}. $$ Then $\pi$ is a holomorphic family of admissible smooth representations of $G$ over $\fadc,$ so that the functor $X_\pi: E \mapsto \pi^{(E)}$ has the exactness properties of Lemma \ref{l: exactness properties X pi}. Hence, $\pi^{(E)}_\gl$ is a subquotient of $\pi^{(E')}_\gl,$ and $\gf_{E, \Xi, \gl}$ is induced by $\gf_{E', \Xi, \gl},$ for all $\gl \in \gL.$ The result now follows by taking the direct sum over $\gl \in \gL.$ \end{proof} \begin{prop}\label{p: dinter Let $\gf \in \holspace_{P_0}.$ Then the following conditions are equivalent: \begin{enumerate} \itema $\gf \in \holspace_{P_0}(\cD_{P_0});$ \itemb for every finite dimensional $\O_0$-module $E$ and every pair of finite sets $\Xi \subset \dM$ and $\Lambda\subset \a_\iC^*,$ the endomorphism $\gf_{E,\Xi,\Lambda}$ belongs to $\Endo(\pi_{E,\Xi,\Lambda})^\#.$ \end{enumerate} \end{prop} \begin{proof First assume (b). Let $\gd =( \gd_1, \ldots, \gd_N) $ be a sequence of data in $\cD_{P_0}.$ Then for (a) it suffices to show that \begin{equation} \label{e: gf in sharp} \gf_\gd \in \Endo(V_{\pi_\gd})^\#, \end{equation} in view of Lemma \ref{l: one restricted intertwining condition suffices}. \par We note that, with $\gd_j = (P_0, \xi_j, \gl_j, \eta_j),$ $$ \pi_\gd = \oplus_{j=1}^n \; \pi^{(E_{\eta_j})}_{P_0, \xi_j, \gl_j}. $$ Let $E$ be the direct sum of the modules $E_{\eta_j}$, and put $\Xi \coleq \{\xi_1, \ldots, \xi_N\}$ and $\gL \coleq \{\gl_1, \ldots, \gl_N\}.$ For each $1 \leq j \leq N$ and $\gl \in \fadc,$ let $(\pi_{j})_{\gl} \coleq \pi_{P_0, \xi_j, \gl}.$ Then $\pi_j$ is a holomorphic family of admissible smooth representations of $G$ over $\fadc,$ so that Lemma \ref{l: exactness properties X pi} applies to the functors $X_{\pi_j}: E \mapsto \pi_j^{(E)}.$ It follows that $$ \pi_{\gd_j} \preceq \pi_{j \gl_j}^{(E)}, \quad \text{and}\quad \gf_{\gd_j} = \gf_{P_0, \xi_j, \gl_j}^{(E)}|_{V_{\gd_j}} $$ in a natural fashion (here $\rho \preceq \pi$ indicates that $\rho$ is a subrepresentation of $\pi$). This in turn implies that $$ \pi_\gd \preceq \pi_{E,\Xi,\gL}, \quad \text{and}\quad \gf_\gd = \gf_{E, \Xi, \gL}|_{V_\gd}. $$ Hence (\ref{e: gf in sharp}) follows. \par Conversely, assume (a). Let $E, \Xi, \gL$ be as stated in (b). First we consider the case that $E$ is of the form $E_\eta,$ with $\eta$ a finite sequence in $\fadc.$ Then $$ \pi_{E_\eta, \Xi, \gL} = \oplus_{\xi \in \Xi, \gl \in \gL} \;\; \pi^{(\eta)}_{P_0, \xi, \gl} $$ and it follows from (a) that $\gf_{E_\eta, \Xi, \gL}$ belongs to $\Endo(\pi_{E_\eta, \Xi, \gL})^\#.$ In view of Lemma \ref{l: exactness properties X pi}, this implies that $\gf_{E, \Xi, \gL}$ belongs to $\Endo(\pi_{E, \Xi, \gL})^\#$ for $E$ a direct sum of copies of $E_\eta.$ Finally, let $E$ be arbitrary. Then by Proposition \ref{sec43} the module $E$ is a subquotient of $E_\eta^N$ for a suitable finite sequence $\eta$ and a suitable $N\in \N.$ This implies, again by Lemma \ref{l: exactness properties X pi}, that $\pi_{E, \Xi, \gL}$ is a subquotient of $\pi_{E_\eta^N, \Xi, \gL}$ and that $\gf_{E, \Xi, \gL}$ is induced by $\gf_{E_\eta^N, \Xi, \gL}.$ From this and Lemma \ref{l: stability of ic under subq} it follows that $\gf_{E, \Xi, \gL}$ belongs to $\Endo(\pi_{E, \Xi, \gL})^\#.$ Hence (b). \end{proof} \section{Conditions in terms of the Hecke algebra}\label{s: Hecke algebra Let $k$ be a field and $A$ a $k$-algebra with an approximate identity $(\alpha_j)_{j\in J}$. This means that \begin{enumerate} \itema $J$ is a partially ordered set; \itemb for all $j_1,j_2$ in $J$ with $j_1\leq j_2$ we have $\alpha_{j_1}\alpha_{j_2} =\alpha_{j_2}\alpha_{j_1}=\alpha_{j_1};$ \itemc for every $a\in A$ there exists a $j\in J$ such that $\alpha_j a= a\alpha_j=a.$ \end{enumerate} The Hecke algebra $\H(G,K)$ is an example of such an algebra (see \cite[Chap.~I, \S 6]{knappvogan}). Indeed, let $(\vartheta_j)_{j \in \N}$ be an increasing sequence of finite subsets of $\dK,$ whose union is $\dK.$ Then $$ \ga_j \coleq \sum_{\gd \in \vartheta_j} \dim(\gd)\, \chi_{\gd^\vee}, \qquad j \in \N, $$ defines an approximate identity in $\H(G,K).$ We denote the opposite algebra of $A$ by $A^{opp}.$ It is readily seen that the elements $\ga_j \otimes \ga_j$ form an approximate identity for the algebra $A \otimes A^{opp},$ so that this algebra is approximately unital. \par For $j \in J,$ let $A_j$ be the set of $a \in A$ with $\ga_j a = a \ga_j = a .$ Then it is readily seen that $A_j$ is a subalgebra of $A.$ Moreover, (b) implies that $A_{j_1} \subset A_{j_2}$ whenever $j_1 \leq j_2$, and (c) implies that $A$ is the union of the subalgebras $A_j.$ \subsection{Some general facts on approximately unital $A$-modules Let $V$ be a left (or right) $A$-module (in particular, $V$ is a $k$-linear space). We say that $V$ is approximately unital if, for every $v\in V$, there exists a $j\in J$ such that $\alpha_j\cdot v= v$. \par For $j\in J$, let $$ V_j := \{ \ga_j \cdot v \mid v \in V\}. $$ Then $V_{j_1}\subset V_{j_2}$ whenever $j_1\leq j_2$. We note that $V$ is approximately unital if and only if $\cup_{j \in J}\, V_j = V.$ \par Let $\End{V}$ denote the algebra of $k$-linear endomorphisms of $V$ and let $$ \pi: \; A\longrightarrow \mathrm{End}(V) $$ denote the canonical algebra homomorphism. Then $V_j$ is the image of $\pi(\alpha_j).$ We note that $V_j$ is invariant under the action of $A_j,$ turning this space into a left $A_j$-module. Let $V^j$ denote the kernel of $\pi(\alpha_j)$. Then since $\ga_j$ is an idempotent, it follows that \begin{equation} \label{e: deco V in V j} V = V^j \oplus V_j. \end{equation}\indent For any $A$-module $V$, the submodule $V_{au}\coleq A\cdot V$ is approximately unital, and it is in fact the maximal submodule with this property. Note that $$ V_{au} = \cup_{j \in J} V_j. $$ \par We now assume that $V$ is an approximately unital $A$-module. The $k$-linear space $\Hom_k(V,k)$ has a natural $A^{opp}$-module structure. Accordingly, we define $$ V^\vee\coleq (\Hom_k(V,k))_{au}. $$ We denote by $\End{V}_0$ the image of the natural linear map $V \otimes V^\vee\to \Endo(V)$ induced by $v\otimes v'\mapsto (u \mapsto v'(u)v).$ \par We will say that an $A$-module $V$ is admissible if $V_j$ is finite dimensional for every $j \in J.$ \begin{lemma Let $V$ be an admissible approximately unital $A$-module. Then, as $A\otimes A^{opp}$-modules, $$ \End{V}_0=\mathrm{End}(V)_{au}. $$ \end{lemma} \begin{proof It is readily seen that $\End{V}_0$ is an $A\otimes A^{opp}$-submodule of $\mathrm{End}(V)$ which is approximately unital. This implies that $\End{V}_0\subset \mathrm{End}(V)_{au}.$\par Conversely, let $j \in J.$ We consider the inclusion map $\iota_j: V_j \to V$ and the epimorphism $p_j: V \to V_j$ determined by $\iota_j \after p_j = \pi(\ga_j).$ These maps induce linear maps $S_j : \End{V} \to \End{V_j}$ and $T_j: \End{V_j} \to \End{V}$ given by $$ S_j: \;f \mapsto p_j \after f \after \iota_j \quad{\rm and}\quad T_j:\; g \mapsto \iota_j \after g \after p_j. $$ It is readily checked that $$ S_j \after T_j = \id_{\End{V_j}}, \quad T_j \after S_j(f) = \pi(\ga_j) \after f \after \pi(\ga_j) = \pi_{\End{V}}(\ga_j) f. $$ Hence, $T_j$ is an injective linear map $\End{V_j} \to \End{V}$ with image $\End{V}_j.$ One now readily checks that the following diagram commutes: \def\rmi{{\rm i}} $$ \xymatrix{ {\End{V}_j} \ar[rr]^{\rmi_j} && {\End{V}} \\ {\End{V_j}} \ar[u]^{t_j} && \ar[ll] V_j \otimes (V_j)^* \ar[u]_{i_j \otimes p_j^*} } $$ Here $\rmi_j$ is the inclusion map, and $t_j$ is the map uniquely determined by $\rmi_j \after t_j = T_j.$ The map at the bottom is the canonical inclusion. We now observe that the map at the bottom is a linear isomorphism by admissibility of $V.$ Moreover, $p_j^*$ maps $(V_j)^*$ into $(V^*)_j \subset V^\vee,$ and we infer that $\End{V}_j \subset \End{V}_0.$ Finally, by taking the union over all $j,$ we conclude that $\End{V}_{au}\subset \End{V}_0$. \end{proof} \subsection{A double commutant theorem \begin{definition}\label{d: commut \rm Let $(\pi, V)$ be an approximately unital $A$-module. We define $\End{\pi}^\#$ to be the space of $\varphi \in \End{V}_0$ such that for every $n\in\N\setminus\{0\}$ and every $A$-submodule $W \subset V^{\times n}$ we have $\varphi^{\times n}(W) \subset W.$ \end{definition} \begin{lemma}\label{l: dcomm Let $(\pi,V)$ be an admissible approximately unital $A$-module. Then $$ \pi(A)=\End{\pi}^\#. $$ \end{lemma} \begin{proof Since $\pi(A)$ is an approximately unital $A\otimes A^{opp}$-submodule of $\End{V}$, it is contained in $\End{V}_{au}=\End{V}_0$. Moreover, if $a\in A$, then for every $n\in\N\setminus\{0\}$ and every invariant subspace $W$ of $V^{\times n}$ we have $\pi(a)^{\times n}(W)\subset W.$ Hence, $\pi(A)\subset \Endo(\pi)^\#.$ \par We now turn to the converse inclusion. By definition $\cE\coleq \End{\pi}^\#$ is a subspace of $\End{V}_0=\End{V}_{au}$. Moreover $\cE$ is $A\otimes A^{opp}$-invariant, hence an approximately unital $A\otimes A^{opp}$-module. Fix $\gf \in \cE.$ Then it follows that there exists a $j\in J$ such that $(\alpha_j\otimes\alpha_j)\cdot\varphi = \varphi.$ \par By admissibility, the space $V_j$ has a finite basis $u_1, \ldots, u_n$ over $k.$ Let $W$ be the $A$-submodule of $V^{\times n}$ generated by $(u_1, \ldots, u_n).$ Then by definition of $\Endo(\pi)^\#,$ the space $W$ is $\gf^{\times n}$-invariant. This implies the existence of an element $a \in A$ such that $\gf(u_k) = \pi(a) u_k$ for all $1 \leq k \leq n.$ Hence $\gf = \pi(a)$ on $V_j.$ It follows that $\gf = \pi(a \ga_j)$ on $V_j.$ On the other hand, both $\gf = (\ga_j \otimes \ga_j) \cdot \gf = \pi(\ga_j) \after \gf \after \pi(\ga_j)$ and $\pi(a\ga_j)$ vanish on $V^j.$ In view of (\ref{e: deco V in V j}) this implies that $\gf = \pi(a\ga_j)$ on $V.$ Hence $\gf = \pi(a\ga_j) \in \pi(A)$ and the proof is complete. \end{proof} \subsection{Application to Harish-Chandra modules}\label{s: hcmod Every admissible $(\fg, K)$-module $V$ is a module for the Hecke algebra $\H(G, K)$ in a natural way, and as such it is admissible and approximately unital. This assignment of an $\H(G,K)$-module to an admissible $(\fg, K)$-module defines a functor which establishes an isomorphism of categories, from the category of admissible $(\fg, K)$-modules onto the category of admissible approximately unital $\H(G,K)$-modules (see \cite[I, \S 6, Theorem 1.117]{knappvogan}). \par Accordingly, if $V$ is an admissible $(\fg, K)$-module, then the associated algebra $\End{\pi}^\#$ consists of all $K\times K$-finite endomorphisms $\gf$ of $V$ with the property that, for every positive integer $n,$ the map $\gf^{\times n}$ preserves all $(\fg, K)$-invariant subspaces of $V^{\times n}.$ In particular, the present notation is compatible with the notation introduced earlier in Definition \ref{d: End sharp}. \par \begin{cor}\label{c: dcomm Let $(\pi,V)$ be an admissible $(\fg, K)$-module. Then $$ \pi(\H(G,K))=\End{\pi}^\#. $$ \end{cor} \begin{proof This follows from Lemma \ref{l: dcomm}. \end{proof} \subsection{Proof of the main theorem} \label{s: proof of main thm} For $P \in \cP(A)$ we define $\PW^\pre_P(G,K)$ to be the subspace of $\cF_P$ (see (\ref{e: holspace})), consisting of all functions $\gf \in \cF_P$ such that there exists an $R > 0$ and for all $n\in \N$ a constant $C_n >0$ such that the estimate (\ref{e: PW estimate}) holds for all $\xi \in \dMPds$ and all $\gl \in \faPdc.$ Furthermore, we define $$ \PW^\pre(G,K) = \oplus_{P \in \cP(A)} \; \PW_P^\pre(G,K). $$ Then $\PW^\pre(G,K)$ is a subspace of $\cF.$ We recall that $P_0 \in \cP(A)$ is a fixed minimal parabolic subgroup of $G.$ \begin{definition}{\ \rm \begin{enumerate} \itema The Delorme Paley--Wiener space $\PW^D(G,K)$ is defined to be the intersection $\PW^\pre(G,K) \cap \cF(\cD).$ \itemb The restricted Delorme Paley--Wiener space $\PW_{P_0}^D(G,K)$ is defined to be the intersection $\PW_{P_0}^\pre(G,K) \cap \cF_{P_0}(\cD_{P_0}).$ \end{enumerate} \end{definition} \begin{thm}\label{t: first PW iso The natural projection $\pr: \cF \to \cF_{P_0}$ restricts to an isomorphism of $\PW^D(G,K)$ onto $\PW_{P_0}^D(G,K).$ \end{thm} \begin{proof It is clear that $\pr$ maps $\PW^\pre(G,K)$ into $\PW^\pre_{\scriptscriptstyle P_0}(G,K).$ Combining this with Proposition \ref{p: reduction to one minimal}, we see that $\pr_\cD$ maps $\PW^D(G,K)$ injectively into $\PW^D_{P_0}(G,K).$ Let $\tau: \cF_{P_0} \to \cF$ be defined as in the proof of Proposition \ref{p: reduction to one minimal}. Then it is readily checked that $\tau$ maps $\PW^\pre_{\scriptscriptstyle P_0}(G,K)$ into $\PW^\pre(G,K).$ This implies that $\tau_\cD$ maps $\PW^D_{\scriptscriptstyle P_0}(G,K)$ into $\PW^D(G,K).$ As $\pr_\cD \after \tau_\cD = \id$ on $\PW^D_{\scriptscriptstyle P_0}(G,K),$ it follows that $\pr_\cD$ maps the space $\PW^D(G,K)$ surjectively onto $\PW^D_{\scriptscriptstyle P_0}(G,K).$ \end{proof} \begin{definition \label{d: defi PWH} \rm We define $\PW^H(G,K)$ to be the space of functions $\gf \in \PW^{\rm pre}_{P_0}(G,K),$ such that for every finite dimensional $\cO_0$-module $E,$ and every pair of finite sets $\Xi \subset \dM,$ $\gL\subset \fadc,$ we have $$ \gf_{E,\Xi,\Lambda} \in \pi_{E,\Xi,\Lambda}(\H(G,K)). $$ \end{definition} \vspace{4pt} \medbreak \begin{thm} \label{t: PWD is PWH The space $\PW^D_{P_0}(G,K)$ equals $\PW^H(G,K).$ \end{thm} \begin{proof Both spaces are subspaces of $\PW^\pre_{\scriptscriptstyle P_0}(G,K).$ Let $\gf$ be an element of the latter space. Then $\gf \in \PW_{\scriptscriptstyle P_0}^D(G,K)$ is equivalent to $\gf\in \cF_{\scriptscriptstyle P_0}(\cD_{P_0}),$ which in turn is equivalent to condition (b) of Proposition \ref{p: dinter}. In view of Corollary \ref{c: dcomm}, the latter condition is equivalent to $\gf \in \PW^H(G,K).$ \end{proof} \begin{definition} \rm We define the Arthur Paley--Wiener space $\PW^A(G,K)$ to be the space of functions $\gf \in \PW^\pre_{P_0}(G,K)$ such that $\gf$ satisfies the Arthur--Campoli relations (see Definition \ref{d: AC relations}). \end{definition} \begin{thm}\label{t: PWD = PWA The space $\PW^H(G,K)$ equals $\PW^A(G,K).$ \end{thm} \begin{proof Both are subspaces of $\PW_{\scriptscriptstyle P_0}^\pre(G,K).$ Let $\gf$ be a function in the latter space. Then by Proposition \ref{p: reformulation ac} the assertion $\gf \in \PW^A(G,K)$ is equivalent to the assertion that for every finite dimensional $\cO_0$-module, and every pair of finite subsets $\Xi \subset \dM,$ $\gL \subset \fadc,$ we have \begin{equation} \label{e: gf in double perp} \gf_{E, \Xi, \gL} \in (\pi_{E, \Xi, \gL}(C_c^\infty(G,K))^{\perp})^{\perp}. \end{equation} As every function from $\PW_{P_0}^{\rm pre}(G,K)$ has values in $$ \oplus_{\xi \in \dM} \;\Endo(C^\infty(K\col \xi))_{\theta\theta} $$ for some finite subset $\theta \subset \dK,$ it follows by admissibility that (\ref{e: gf in double perp}) is equivalent to $$ \gf_{E, \Xi, \gL} \in \pi_{E, \Xi, \gL}(C_c^\infty(G,K)). $$ By Proposition \ref{end-hecke} this in turn is equivalent to $$ \gf_{E, \Xi, \gL} \in \pi_{E, \Xi, \gL}(\H(G,K)). $$ If follows that $\gf \in \PW^A(G,K) \iff \gf\in \PW^H(G,K).$ \end{proof} We now come to the main result of our paper. \begin{thm \label{t: main thm} The map $\pr: \gf \mapsto \gf_{P_0}$ defines a linear isomorphism from $\PW^D(G,K)$ onto $\PW^A(G,K).$ \end{thm} \begin{proof This follows from combining Theorems \ref{t: first PW iso}, \ref{t: PWD is PWH} and \ref{t: PWD = PWA}. \end{proof} \subsection{Another useful characterization of the Paley-Wiener space} \label{s: another useful char of PW} In this subsection we will obtain another useful characterization of the Paley-Wiener space $\PW^A(G,K).$ This will allow us to derive the Paley-Wiener theorem due to Helgason \cite{helgasonpw} and Gangolli \cite{gangolli} from Arthur's Paley-Wiener theorem. We define the Fourier transform $\Fou: \Cci(G,K)_{K \times K} \to \cO(\fadc) \otimes \cS_{P_0}$ by $\Fou(f)_\xi(\gl) = \widehat f(P_0, \xi, \gl),$ see (\ref{e: defi Fourier}). From this definition it is immediate that $\Fou$ is a $K\times K$-equivariant linear map. In terms of this Fourier transform, Arthur's Paley-Wiener theorem may be stated as follows (see \cite{arthurpw} and \cite{BSpwspace}). \begin{thm}[Arthur's Paley-Wiener theorem] The map $\Fou$ is a $K \times K$-equivariant linear isomorphism from $C_c^\infty(G,K)_{K \times K}$ onto $\PW^A(G,K).$ \end{thm} By Theorem \ref{t: PWD = PWA} the Paley-Wiener space $\PW^A(G,K)$ equals the space $\PW^H(G,K)$ introduced in Definition \ref{d: defi PWH}. We shall now give another characterization of that space. We use the notation $\PW(\fa)$ for the (Euclidean) Paley-Wiener space associated with $\fa,$ i.e., $\PW(\fa)$ is the image of the classical Fourier transform $C_c^\infty(\fa) \to \cO(\fadc).$ Then the space $\PW_{P_0}^\pre(G,K),$ introduced in the beginning of Subsection \ref{s: proof of main thm}, equals $\PW(\fa) \otimes \cS(P_0).$ \begin{prop} The space $\PW^H(G,K)$ is equal to the space of all functions $\gf \in \PW(\fa) \otimes \cS(P_0)$ with the following property. For each finite number of triples $(u_j, \xi_j, \gl_j) \in S(\fa^*) \times \dM \times \fadc,$ $1 \leq j \leq n$, there exists an element $h \in \H(G,K)$ such that $$ \gf(\xi_j, \gl_j; u_j) = \pi_{P_0, \xi_j, \gl_j; u_j}(h)\qquad \text{for all }\;\;\;1 \leq j \leq n. $$ \end{prop} \begin{proof} We first assume that $\gf \in \PW^H(G,K).$ Then $\gf \in \PW(A) \otimes \cS(P_0).$ Let a finite number of triples $(u_j, \xi_j , \gl_j)$ be given. For each $u_j$ there exists a finite dimensional $\cO_0$-module $E_j,$ and a linear functional $\eta_j \in \Endo(E_j)^*$ such that $\eta_j \after f^{(E_j)} = f(\gl_j;u_j),$ for all $f \in \cO(\fadc);$ see Lemma \ref{l: E associated with u}. Put $E = \oplus E_j,$ and let $\pr_j: E \to E_j$ denote the associated projection maps. There exists an element $h \in \H(G,K)$ such that $ \gf_{E, \Xi, \gL} = \pi_{E, \Xi, \gL}(h).$ This implies that $\gf_{E,\xi_j, \gl_j} = \pi_{E, \xi_j,\gl_j}(h).$ for each $j.$ By application of $\eta_j \after \pr_j \otimes I$ to the latter expression it follows that $\gf(\xi_j, \gl_j;u_j) = \pi_{P_0, \xi_j, \gl_j; u_j}(h),$ for all $j.$ This proves that $\PW^H(G,K)$ is included in the space described. To obtain the other inclusion, let $\gf \in \PW(\fa) \otimes \cS(P_0)$ satisfy the conditions of the space described. Let $E$ be a finite dimensional $\cO_0$-module, and let $\Xi \subset M^\wedge$ and $\gL \subset \fadc$ be finite sets. Fix a basis $\{\eta^k\}$ of $\Endo(E)^*.$ We may number the elements of $\gL$ by $\gl_j.$ Then for each $j, k$ there exists a $u_j^k \in S(\fa^*) $ such that $\eta^k f^{(E)}(\gl_j) = f(\gl_j; u_j^k)$ for all $f \in \cO(\fadc),$ see Lemma \ref{l: dual of End E}. By the assumption on $\gf,$ there exists an element $h \in \H(G,K)$ such that $\gf(\xi,\gl_j; u^k_j) = \pi_{P_0, \xi, \gl_j; u^k_j}(h)$ for all $\xi \in \Xi$ and all $j,k.$ It follows that $$ (\eta^k \otimes 1)\after \gf_{E, \xi, \gl_j} = (\eta^k \otimes 1)\after \pi_{E,\xi, \gl_j}(h), $$ for all $k, j$ and $\xi \in \Xi.$ This implies that $\gf_{E, \xi, \gl_j} = \pi_{E, \xi, \gl_j}(h)$ for all $j$ and all $\xi \in \Xi.$ Hence, $\gf_{E, \Xi, \gL} = \pi_{E,\Xi, \gL}(h),$ and we conclude that $\gf \in \PW^H(G,K).$ \end{proof} \begin{cor} \label{c: char of PWH11} The space $\PW^H(G,K)_{11}$ of $K\times K$-fixed elements in\break $\PW^H(G,K)$ consists of all $\gf \in \PW(\fa)\otimes \cS(P_0, 1)_{11}$ such that for each finite number of pairs $(u_i, \gl_i) \in S(\fa^*) \times \fadc,$ $1 \leq j \leq n$, there exists an element $h \in \H(G,K)_{11}$ such that $$ \gf(\gl_j; u_j) = \pi_{P_0, 1, \gl_j; u_j}(h)\qquad \text{for all }\;\;\;1 \leq j \leq n. $$ \end{cor} \begin{proof} Since $\cS(P_0)_{11} = \cS(P_0, 1)_{11},$ this follows from the previous result by projection onto the $K \times K$-type $(1,1).$ \end{proof} Let $P_1 \in \Endo(C^\infty(K:1))$ denote the $K$-equivariant projection onto the one-dimensional subspace of $C^\infty(M\bs K)$ consisting of the constant functions. Then we observe that \begin{equation} \label{e: char cS11} \cS(P_0,1)_{11} \simeq \Endo(C^\infty(K:1))_{11} = \C \,P_1. \end{equation} Accordingly, we may may view $\PW^H(G,K)_{11}$ as a subspace of $\PW(\fa)\otimes \C\, P_1.$ The inclusion map $\iota: K \to G$ induces a continuous linear map $\iota^*: C^\infty(G) \to C^\infty(K)$ by pull-back. The transposed of this map is an injective continuous linear map $\iota_*: \cE'(K) \to \cE'(G).$ Accordingly, we shall use this map to view $\cE'(K)$ as a subspace of $\cE'(G).$ In particular, the normalized Haar measure $dk$ will be viewed as the element of $\H(G,K)$ given by $ dk(\gf) = \int_K \gf(k)\; dk,$ for $\gf \in C^\infty(G).$ Clearly, $dk \in \H(G,K)_{11}.$ \begin{lemma} \label{l: 11 component of H} The map $u \mapsto R_u dk$ induces a linear isomorphism from $U(\fg)^K /U(\fg)^K \cap U(\fg)\fk$ onto $\H(G,K)_{11}.$ \end{lemma} \begin{proof} We define the linear map $\ga: U(\fg) \otimes \cE'(K) \to \cE'(G)$ by $$ \ga(u \otimes T) = R_u T, $$ where $R$ denotes the right regular representation on $\cE'(G).$ Then $\ga$ factors to a linear isomorphism $$ \bar \ga: U(\fg)\otimes_{U(\fk)} \cE'(K) \to \H(G,K), $$ see \cite[I, \S 6]{knappvogan}. This map intertwines the $K\times K$-actions $(\Ad \otimes R) \times (1 \otimes L)$ and $R \times L,$ hence restricts to an isomorphism $$ (U(\fg)\otimes_{U(\fk)} \cE'(K))_{11} {\buildrel \simeq\over \longrightarrow} \H(G,K)_{11} $$ The space of left $K$-invariants in $\cE'(K)$ equals $\C dk.$ As $dk$ is also right $K$-invariant, we see that $$ (U(\fg)\otimes_{U(\fk)} \cE'(K))_{11} \simeq U(\fg)^K \otimes_{U(\fk)} \C dk \simeq U(\fg)^K / U(\fg)^K \cap U(\fg)\fk. $$ The result now follows. \end{proof} In the following lemma, $W$ denotes the Weyl group of $\fa$ in $\fg.$ \begin{lemma} \label{l: image H11} The image of $\H(G,K)_{11}$ under Fourier transform $h \mapsto \widehat h$ equals the subspace $P(\fadc)^W \otimes \C P_1$ of $\cO(\fadc)\otimes \cS(P_0).$ \end{lemma} \begin{proof} Denote the image by $S.$ Then $S$ is contained in $\cO(\fadc) \otimes \cS(P_0)_{11} \simeq \cO(\fadc) \otimes \C P_1.$ Let $h \in \H(G,K)_{11}.$ Then it follows that the Fourier transform of $h$ is of the form $\psi \otimes P_1,$ with $\psi \in \cO(\fadc).$ In view of the previous result, $h = R_u(dk),$ with $u \in U(\fg)^K.$ Hence, for all $\gl \in \fadc,$ \begin{eqnarray*} \psi(\gl)1_{M \bs K} & = & (\psi(\gl)\otimes P_1)1_{M\bs K}\\ &=& \pi_{P_0,1 , \gl}(h) 1_{M \bs K} \\ & =& \pi_{P_0, 1, \gl}(u)1_{M \bs K} = \gg(u, \gl) 1_{M \bs K}, \end{eqnarray*} where $\gg$ denotes the Harish-Chandra algebra homomorphism $U(\fg)^K \to S(\fa)$ which has image $S(\fa)^W = P(\fadc)^W.$ The result now follows by application of Lemma \ref{l: 11 component of H} \end{proof} \begin{cor} \label{c: PW 11} The space $\PW^H(G,K)_{11}$ equals $\PW(\fa)^W \otimes \C P_1.$ \end{cor} \begin{proof} Let $\gf \in \PW^H(G,K)_{11}.$ Let $\gl \in \fadc$ and $w \in W.$ Then by Corollary \ref{c: char of PWH11} there exists an element $h \in \H(G,K)_{11}$ such that $\gf(\gl) = \widehat h(P_0, 1, \gl)$ and $\gf(w\gl) = \widehat h(P_0, 1, w\gl).$ As $\widehat h(P_0, 1, \gl) = \widehat h(P_0, 1, w\gl)$ by Lemma \ref{l: image H11}, we see that $\gf$ is $W$-invariant. Hence, $\gf \in \PW(\fa)^W \otimes \C P_1.$ Conversely, assume that $\gf \in \PW(\fa)^W \otimes \C P_1.$ Write $\gf = \psi \otimes P_1,$ then $\psi$ is a $W$-invariant holomorphic function. Let $(u_j, \gl_j) \in S(\fa^*) \times \fadc,$ $1 \leq j \leq n.$ Then there exists an element $p \in P(\fa^*)^W$ such that $$ \psi( \gl_j; u_j) = p(\gl_j; u_j) \qquad \text{for all }\;\;\;1 \leq j \leq n. $$ In view of Lemma \ref{l: image H11} there exists an element $h \in \H(G,K)_{11}$ such that $\widehat h(P_0, 1, \gl) = p \otimes P_1.$ Hence, $$ \gf( \gl_j; u_j) = \widehat h (P_0, 1, \gl_j; u_j) = \pi_{P_0, 1, \gl_j;u_j}(h), \qquad (1 \leq j \leq n). $$ In view of Corollary \ref{c: char of PWH11} it follows that $\gf \in \PW^H(G,K)_{11}.$ \end{proof} We now note that the spherical Fourier transform $$ \cF_{11}: C_c^\infty(G,K)_{11} \to \cO(\fadc) $$ is given by the formula $ \cF_{11}f(\gl) 1_{K/M} = \pi_{P_0, 1, \gl}(f) 1_{K/M}. $ This implies that for all $f \in C_c^\infty(G,K)_{11}$ we have $$ \cF_{11}f(\gl) \otimes P_1 = \Fou f(\gl). $$ We can now finally deduce the Paley-Wiener theorem of Helgason \cite{helgasonpw} and Gangolli \cite{gangolli}. \begin{cor} $\cF_{11}(C_c^\infty(G,K)_{11}) = \PW(\fa)^W.$ \end{cor} \begin{proof} By $K\times K$-equivariance, it follows from Arthur's Paley-Wiener theorem that $$ \cF_{11} (C_c^\infty(G,K)_{11})\otimes \C P_1 = \Fou(C^\infty_c(G,K)_{11}) = \PW^A(G,K)_{11}. $$ The latter space equals $\PW^H(G,K)_{11},$ by Theorem \ref{t: PWD = PWA}. Now apply Corollary \ref{c: PW 11}. \end{proof}