% BW5.tex : Discounted additive functionals of Markov processes % \input mssymb \magnification=\magstep1 \vsize=9.5truein \parskip=\medskipamount % Font selections \font\sc=cmcsc10 %\baselineskip=18pt % directives for increased leading %\lineskiplimit=2pt %\lineskip=4pt % Start of definitions \def\cl{\centerline} \def\down{\downarrow} \def \up{\uparrow} % some symbols \def\leq{\leqslant} \def\geq{\geqslant} \def\brak#1#2{\langle#1,#2\rangle} \def\diag{\mathop{\rm diag}\nolimits} % new roman operators \def\spect{\mathop{\rm spect}\nolimits} \def\Leb{\mathop{\rm Leb}\nolimits} \def\supp{\mathop{\rm supp}\nolimits} \def\Re{\mathop{\rm Re}} \def\Int{\mathop{\rm Int}} \def\tendsd{\mathrel{\mathop\rightarrow^{\cal D}}} \def\wsim{\mathrel{\displaystyle\mathop\sim^{\rm w}}} \def\eqd{\mathrel{\mathop=^{\cal D}}} \def\Var{\mathop{\rm Var}\nolimits} \def\a{\alpha} \def\b{\beta} \def\g{\gamma} \def\d{\delta} % greek letters \def\e{\epsilon} \def\l{\lambda} \def\phi{\varphi} \def\s{\sigma} \def\t{\theta} \def\Al{A_\l} \def\elt{e^{-\l t}} \def\L{{\cal L}} \def\D{\triangle} \def\trho{\triangle\rho} \def\ind{I_{(0,\infty)}} \def\half{\raise1pt\hbox{$\scriptstyle{1 \over 2}\displaystyle$}} \def\endbox{\hfill$\square$\par} % Macros % % AUTOMATIC EQUATION NUMBERING 29January1992 (c)MWBAXTER % % cno = chapter number : should be set by user % sno = section number : initialised (0) at chapter start % eno = equation number : initialised (0) at section start % pno = proposition number : initialised (0) at section start % % \enum{fred} will number equation, eg (1.23), and set \fred={(1.23)} % \lnum{fred} is as \enum but without the \eqno, for use in \eqalignno % \lnuma{fred} will number as (1.23a) with \fred={(1.23)} \freda={(1.23a)} % \lnumb{fred} will number as (1.23b) with \fredb={(1.23b)} % % usage: "equation \fred\ suggests that.." (extra "\ " is trailing space) % % WARNING : DO NOT USE TEX or USER MACRO NAMES AS EQUATION NAMES % \newcount\cno \newcount\sno \newcount\eno \newcount\pno \def\addone{\global\advance\eno by1\edef\cs{(\the\eno)}} % x.x FORMAT \def\lnum#1{\addone\global\expandafter\let\csname#1\endcsname=\cs\cs} \def\enum#1{\eqno{\lnum{#1}}} \def\lnuma#1{\addone\edef\csa{(\the\eno{\rm a})}% x.x FORMAT \global\expandafter\let\csname#1\endcsname=\cs% \global\expandafter\let\csname#1a\endcsname=\csa\csa} \def\lnumb#1{\edef\csb{(\the\eno{\rm b})}% x.x FORMAT \global\expandafter\let\csname#1b\endcsname=\csb\csb} % % \pnum{Proposition}{fred} will number proposition/theorem/etc, % eg Proposition 1.1, and set \fred={Proposition 1.1} \def\paddone#1{\global\advance\pno by 1 \edef\cs{#1\ \the\pno}} % ADAPTED FOR x.x FORMAT \def\pnum#1#2{\paddone{#1}% \global\expandafter\let\csname#2\endcsname=\cs\cs} \def\proc#1{\noindent{\bf#1\ \ }} \def\proof#1{\noindent{\bf Proof of #1\ \ }} % % END OF AUTOMATIC NUMBERING MACROS \sno=0 \eno=0 \pno=0 \def\newsect#1{\bigbreak\bigskip\global\advance\sno by 1 \cl{\bf\the\sno. #1}\nobreak\bigskip\nobreak} % \def\newpage{\mark{}\vfill\eject} % % References \newcount\rno \newcount\rflag \rflag=0 \input ihp_ref % % Begining of text \cl{\bf Discounted additive functionals of Markov processes} \bigskip \cl{{\bf M}ARTIN {\bf B}AXTER} \medskip %\cl{\it Pembroke College, Cambridge CB2 1RF} %\smallskip \cl{\it Statistical Laboratory, University of Cambridge, Cambridge CB2 1SB} \bigskip\bigskip \centerline{\hbox{\vbox{\halign{#.\quad&#\hfil\cr 1&Introduction\cr 2&Large Deviations\cr 3&More exact results for Markov chains\cr}}}} \newsect{Introduction}% Much study has been made of the time averages of random processes. Most of this effort has been directed towards the Ces\`aro average which weights times uniformly up to a finite horizon. In this paper, we shall derive some results about more general averages. The initial motivation was the exponential discount (we will call this the Abel average), which appears frequently in the contexts of, among others, systems control and models of financial markets. The techniques developed, however, extend easily to other shapes of discount. In this paper we will push forward two strands of inquiry that have been developed in proceeding work. The generality that we will obtain should strictly encompass the existing material and can be read on its own, though reference will be made to earlier papers when a proof is essentially the same as one before. For general results about the large-deviations behaviour of discounted occupation times of processes, in Section 2 we will build on Section (a) of \bwii\ and on \baxiii. Previously we knew that the large-deviation property held for the Abel discounted average of a general finite-state Markov chain, and we will fully extend this to completely general discounts of chains and partially extend further to a wide class of Markov processes. We discover that although the large-deviation rate function of the discounted average can be written in terms of that for the Ces\`aro, the rate is often different for a different discount. We also derive results about the smoothness and finiteness of the rate function which are used in the next Section to prove a central limit theorem. Finally we shall go beyond the limited approximation precision of the large-deviation property and give an asymptotic expansion of the density of the distribution itself, following from the density studies of Section (b) of \bwii. This again is now performed for general discounted averages of finite-state Markov chains. We also notice a pattern in the differential equations we worked with, and hypothesize about their full solutions and other generalizations. The main objects studied are the Ces\`aro and Abel averages of a process $X$. Both are random measures of unit mass on the set $S$, the state space of $X$. The former is defined as $$ C_t(F):={1\over t}\int_0^t I_F(X_s)\,ds,\qquad \hbox{\rm for $F$ measurable in $S$, $t>0$.} $$ For any density $m$ on $[0,\infty)$, that is for $m$ in $L_1^+(\Bbb R^+)$, the Abel average $\Al$ is $$ \Al(F):=\l\int_0^{\infty} m(\lambda t)I_F(X_t)\,dt,\qquad \hbox{\rm for $F$ in $S$, $\l>0$.} $$ The two main results of the paper are: \proc{Theorem A.}{\sl Suppose that $X$ is an irreducible Markov chain on a finite state-space $S$, with Q-matrix $Q$. Then both $C_t$ and $\Al$ obey the large-deviation principle with rate functions $I$ and $K$ respectively. That is that $$ \eqalign{\limsup_{t\rightarrow\infty}t^{-1}\log\Bbb P(C_t\in F) &\leq-\inf_{\nu\in F}I(\nu),\cr \limsup_{\l\downarrow0}\l\log\Bbb P(\Al\in F)&\leq-\inf_{\nu\in F}K(\nu),\cr} \qquad\eqalign{\liminf_{t\rightarrow\infty}t^{-1}\log\Bbb P(C_t\in G) &\geq-\inf_{\nu\in G}I(\nu),\cr \liminf_{\l\downarrow0}\l\log\Bbb P(\Al\in G) &\geq-\inf_{\nu\in G}K(\nu),\cr} $$ for $F$ and $G$ respectively closed and open subsets of $M:=M_1(S)$. Additionally, $K$ is related to $I$ by the equation $$ K^*(v)=\int_0^\infty I^*\bigl(m(t)v\bigr)\,dt,\qquad \hbox{\sl $v$ in $\Bbb R^S$,} $$ where $I^*$ and $K^*$ are the Legendre transforms (convex conjugates) of $I$ and $K$, satisfying}%end of sl $$ \eqalign{I^*(v)&=\sup_{x\in M}\bigl(\brak{v}{x}-I(x)\bigr),\cr K^*(v)&=\sup_{x\in M}\bigl(\brak{v}{x}-K(x)\bigr),\cr}\qquad \eqalign{I(x)&=\sup_{v\in \Bbb R^S}\bigl(\brak{v}{x}-I^*(v)\bigr),\cr K(x)&=\sup_{v\in \Bbb R^S}\bigl(\brak{v}{x}-K^*(v)\bigr).\cr} $$ \proc{Theorem B.}{\sl Additionally, if $m$ is of bounded variation, then the density of $\Al$ on $M$ under the law starting $X$ at $i$, $f_i^\l$ can be written as $$ f^\l_i(x)=e^{-K(x)/\l}(2\pi\l)^{-(n-1)/2}\bigl(\det H_K(x)\bigr)^{1/2} z_i(x)r^\l_i(x), $$ where $H_K$ is the Hessian of $K$ taken with respect to $M$, $z(x)$ is the positive eigenvector of $Q+\diag(m_0\nabla K(x))$, and the residue term $r^\l$ goes to 1 in the sense that $r^\l_i(x)\,dx$ converges weakly to $dx$ on $\Int(M)$.} \newsect{Large Deviations}% We will work with general discount shapes and positive recurrent processes. Let $X$ be a stochastic process with state-space $E$ and invariant distribution $\pi$. In the Ces\`aro case we would expect some sort of ergodic theorem such as $$ C_t(F):={1\over t}\int_0^t I_F(X_s)\,ds\rightarrow\pi(F)\qquad \hbox{\rm as $t\rightarrow\infty$},\enum{zdba} $$ for all measurable subsets $F$ of $E$. Then $C_t$ takes values in $M_1(E)$, the space of probability measures on $E$. We might also have a large-deviation result, which we can think of for the moment as the slogan $$ \hbox{\rm ``}\Bbb P(C_t\in H)\approx\exp\bigl(-t\inf_{\nu\in H}I(\nu) \bigr)\hbox{\rm\quad as $t\rightarrow\infty$ ''}\qquad H\subset M_1(E), $$ for some rate function $I$, with $I(\pi)=0$. The space $M_1(E)$ and the continuous bounded functions on $E$, $C_b(E)$ are in duality via the bracket, $\brak{v}{\nu}=\int_E v(x)\nu(dx)$. A related slogan is that of the Laplace transform $$ \hbox{\rm ``}\Bbb E\exp\bigl(\brak{v}{C_t}t\bigr)\approx\exp\bigl(t\d(v) \bigr)\hbox{\rm\quad as $t\rightarrow\infty$ ''}\qquad v\in C_b(E), $$ where $\d$ and $I$ are related by Legendre transformation (convex conjugation), in that $$ \eqalignno{ I(\nu)&=\d^*(\nu):=\sup_{v\in C_b(E)}\bigl(\brak{v}\nu-\d(v)\bigr),& \lnum{zdbb}\cr \d(v)&=I^*(v):=\sup_{\nu\in M_1(E)}\bigl(\brak{v}\nu-I(\nu)\bigr),& \lnum{zdbc}\cr} $$ Our program will be to study, for any discount density $m$ in $L_1^+(\Bbb R^+)$, the average $$ A_\l(F):=\int_0^\infty \l m(\l t)I_F(X_t)\,dt.\enum{zdbd} $$ We will show that $A_\l\rightarrow\pi$ as $\l$ goes to 0, and that the large-deviation principle holds with rate $K$ whose Legendre transform $\eta$ is given by the equation $$ \eta(v)=\int_0^\infty\d\bigl(m(t)v\bigr)\,dt.\enum{zdbe} $$ This is actually the same as the $\eta$-equation at (1.16) in \bwii\ (with discount $m_t=e^{-t}$) and at Theorem~C in \baxiii\ (with discount $m_t=(1-t)^\g$), but \zdbe\ is a more natural formulation. \noindent{\bf Standard set-up.} Let $X$ be an ergodic Feller-Dynkin Markov process on a locally compact Polish space $E$, with generator $L$. We define the Ces\`aro average $C_t$ and the general average $\Al$ by \zdba\ and \zdbd\ respectively. Then $C_t$ and $\Al$ converge to $\pi$, the invariant distribution of $X$, with respect to the weak topology on $M_1(E)$, that is, in the sense of \zdba. A sufficient condition for the former limit is that, as in 8.11.2 of \bingham, $\pi$ is a limiting distribution of the transition semigroup $(P_t)$. The latter limit follows from the former by a similar $L_1$-continuity argument to that which will be used in the proof of \pnum{Theorem}{zxdbone}. \deustroo\ show that under an assumption of uniform ergodicity the large-deviation property holds for $C_t$ with rate function $I$ defined on $M_1(E)$. That is that $$ \eqalign{\limsup_{t\rightarrow\infty}t^{-1}\log\Bbb P(C_t\in F) &\leq-\inf_{\nu\in F}I(\nu),\cr \hbox{\rm and}\qquad\liminf_{t\rightarrow\infty}t^{-1}\log\Bbb P(C_t\in G) &\geq-\inf_{\nu\in G}I(\nu),\cr}\enum{zdbf} $$ for $F$ and $G$ respectively closed and open subsets of $M_1(E)$. We learn from 4.2.17 of \deustroo\ that $$ \lim_{t\rightarrow\infty}t^{-1}\log\Bbb E\exp\int_0^tv(X_s)\,ds=\d(v), \enum{zdbg} $$ where $\d$ and $I$ are convex functions satisfying \zdbb\ and \zdbc. Further there are, by 4.2.27 and 4.2.38 of \deustroo, explicit expressions for $\d$ and $I$ as $$ \eqalignno{\d(v)&=\lim_{t\rightarrow\infty}t^{-1}\log\|P_t^v\|_{\rm op}, \quad\hbox{\rm where $P_t^vf(x):=\Bbb E_x\bigl(\exp(\int_0^tv(X_s)\,ds)f(X_t)\bigr)$,} &\lnum{zdbh}\cr I(\nu)&=\sup\biggl\{-\int_E{Lf(x)\over f(x)}\,\nu(dx): f\geq1,\ f\in\mathop{\rm Dom}(L)\,\biggr\}.&\lnum{zdbi}\cr} $$ As in \bwii\ we shall be particularly interested in the case where $X$ is a Markov chain on a finite state-space $S$ with Q-matrix $Q$. Then $\d(v)=\sup\{\Re(z):z\in\spect(Q+V)\}$, where $V$ denotes the diagonal matrix $\diag(v)$ and $\spect(\cdot)$ denotes spectrum (here the set of eigenvalues). This expression for $\d$ also holds in the general Markov process setting, if the generator $L$ is $\pi$-symmetric. We begin by proving a result whose first part is similar to one remarked by \kifer\ in the context of the large-deviations of the averages of dynamical systems, but it is the second part which will be more useful in our further work. In earlier papers we derived a differential equation by the self-similarity of discount shapes such as $e^{-t}$, but it is enough to study the shifts of the discount along the time-axis, which provides a useful one-dimensional parameterisation. \noindent{\bf\zxdbone} {\sl Suppose that $X$ is an FD Markov process on a space $E$, with generator $L$, and $m$ is any density on $[0,\infty)$, and for $x$ in $E$ and $v$ in $C_b(E)$ the limit $\d(v)=\lim_\l\l\log\Bbb E_x\exp\int_0^{1/\l}v(X_s)\,ds$ exists uniformly in $x$ on $E$. If we define $\phi$ by $$ \phi(x,t,\l,v)=\Bbb E_x\exp\int_0^\infty\t_t m(\l s)v(X_s)\,ds, \eqno\hbox{\rm\lnum{zdbj}} $$ where $\t_t$ is the shift operator $\t_tf(s)=f(t+s)$, then $$ \lim_{\l\downarrow0}\l\log\phi(x,0,\l,v)=\eta(v):= \int_0^\infty\d(m_tv)\,dt.\eqno\hbox{\rm\lnum{zdbk}} $$ Further, $\phi(\cdot,t,\l,v)$ is in the domain of $L$ and $\phi(x,\cdot,\l,v)$ is differentiable and} $$ -{\partial\phi\over\partial t}=\l^{-1}(L+m_tV)\phi.\enum{zdbl} $$ \proof{\zxdbone} We prove the first part using continuity arguments. If we define $$ H(\l,x,\a):=\l\log\Bbb E_x\exp\int_0^{1/\l}\a v(X_s)\,ds,\enum{zdbm} $$ then $\sup_x|H(\l,x,\a)-\d(\a v)|$ goes to 0 as $\l$ does. We start by proving the general average limit for $m$ of the form $$ m=\sum_{i=1}^nc_iI(a_i,b_i), $$ where $\{(a_i,b_i)\}$ are disjoint intervals in $\Bbb R^+$ and $c_i>0$. Set $$ Y^\l_i:=\exp\left(\int_{a_i/\l}^{b_i/\l}c_iv(X_s)\,ds -\l^{-1}(b_i-a_i)\d(c_iv)\right),\enum{zdbn} $$ and define $y^\l_i(x):=\Bbb E_xY^\l_i$. Then for $\l$ sufficiently small $|\l\log y^\l_i(x)|<\e$ uniformly in $x$. Thus $$ \eqalign{{\Bbb E_x\exp\int_0^\infty m(\l t)v(X_t)\,dt\over \exp\l^{-1}\int_0^\infty\d(m_tv)\,dt}&= \Bbb E_x(Y^\l_1\ldots Y^\l_n)= \Bbb E_x(Y^\l_1\ldots Y^\l_{n-1}y^\l_n(X_{a_n/\l}))\cr &\leq e^{\e/\l}\Bbb E_x(Y^\l_1\ldots Y^\l_{n-1}) \leq\ldots\leq e^{n\e/\l},\cr} $$ and so we have the right upper bound. Similarly we have the lower bound. Let us now define $I_\l(m):=\int_0^\infty m(\l t)v(X_t)\,dt$, $L_\l(m):=\l\log\Bbb E\exp I_\l(m)$, and $J(m):=\int_0^\infty\d(m_tv)\,dt$. Then $\l|I_\l(m_1)-I_\l(m_2)|$ is bounded by $\|v\|_\infty\|m_1-m_2\|_1$ uniformly in $\omega$, and hence $|L_\l(m_1)-L_\l(m_2)|$ has the same bound. Because $|\d(v_1)-\d(v_2)|\leq|v_1-v_2|$, the same bound also dominates $|J(m_1)-J(m_2)|$. This $L_1$-continuity and (careful) application of monotone class theorems let us generalize firstly to all bounded $m$ of compact support and then to all $m$ in $L_1^+(\Bbb R^+)$. For the differential equation, we need only apply the Feynman-Ka\u c formula to the space-time process $Y_t:=(X_t,\tau_t)$, where $\tau_t=\tau_0+t$, which has generator $L+\partial_t$. Taking $\l=1$ for simplicity, for $v$ in $C_b(E)$, we define $v_Y$ on $C_b(E\times\Bbb R^+)$ by $v_Y(x,t):=m_tv(x)$, and $$ A_t:=\int_0^t v_Y(Y_s)\,ds=\int_0^t\theta_{\tau_0}m(s)v(X_s)\,ds. \enum{zdbo} $$ Without loss of generality we can assume that $v$ is non-negative, because if \zdbk\ and \zdbl\ hold for some $v$, then they hold for all vectors of the form $v+\a{\bf 1}$, where $\bf 1$ is the constant vector $(1,1,\ldots,1)$. This shifting identity follows from the fact that $\d(v+\a{\bf 1})=\d(v)+\a$, a property which $\eta$ inherits. Then the semigroup $P^v$ defined by $$ P^v_tf(y)=\Bbb E_{Y_0=y}\bigl(e^{A_t}f(Y_t)\bigr) $$ has generator $L^v:=L+\partial_t+m_tV$, as seen in, for example, III.39 of \williams. Then if we set $\phi((x,t)):=\phi(x,t,1,v)$, which is continuous in $t$, we have that $$ P_t^v\phi(Y_0)=\Bbb E_{Y_0}\left(e^{A_t}\phi(Y_t)\right) =\Bbb E_{Y_0}\left(\Bbb E(\exp A_\infty|{\cal F}_t)\right)=\phi(Y_0). \enum{zdbp} $$ Thus $t^{-1}(P^v_t-I)\phi=0$, implying that $\phi$ is in the domain of $L^v$ and is annihilated by it. The equation $L^v\phi=0$ is exactly \zdbl. \endbox We note that in the case of $X$ a standard Brownian motion and the exponential discount $m_t=e^{-t}$ and $v(x)=I(x>0)$, then \zdbl\ is equation (3.5) of \bwi. \proc{\pnum{Corollary}{zxdbthree}} {\sl Suppose that $X$ is an irreducible Markov chain on a finite state-space $S$, with Q-matrix $Q$, and $m$ is any density on $[0,\infty)$, and $\Al$ is defined by \zdbd, then the large-deviation property analogue of \zdbf\ holds for $\Al$ with rate function $K$, $$ \limsup_{\l\rightarrow0}\l\log\Bbb P(\Al\in F) \leq-\inf_{\nu\in F}K(\nu),\quad \hbox{\sl and}\quad\liminf_{\l\rightarrow0}\l\log\Bbb P(\Al\in G) \geq-\inf_{\nu\in G}K(\nu),\enum{zdbextra} % \eqalign{\limsup_{\l\rightarrow0}\l\log\Bbb P(\Al\in F) % &\leq-\inf_{\nu\in F}K(\nu),\cr % \hbox{\sl and}\qquad\liminf_{\l\rightarrow0}\l\log\Bbb P(\Al\in G) % &\geq-\inf_{\nu\in G}K(\nu),\cr}\enum{zdbextra} $$ for $F$ and $G$ respectively closed and open subsets of $M$. The rate function $K$ relates to the $\eta$ of \zdbk\ through the following equations: $$ \eqalignno{K(x)&=\sup_{v\in\Bbb R^S}\brak{v}{x}-\eta(v),&\lnum{zdbq}\cr \eta(v)&=\sup_{x\in M}\brak{v}{x}-K(x),&\lnum{zdbr}\cr} $$ where $M:=M_1(S)=\{(x_i)_{i=1}^n:\sum_i x_i=1,\ x\geq0\}$.} \proof{\zxdbthree} We are in the context of \zxdbone\ because $X$ will satisfy condition ($\tilde{\rm U}$) of 4.2.7 of \deustroo, which is sufficient for the limit $\d$ to exist as required by the theorem. The large-deviation property and \zdbq\ come from theorem II.2 of \ellis. In his language, $t$ is our $v$, $Y_n$ is our $\Al$, $c_n(\cdot)$ is our $\l\log\phi(x,0,\l,\cdot)$, and $c(\cdot)$ is our $\eta(\cdot)$. As $\eta$ is defined and differentiable on the whole of $\Bbb R^S$, it meets Ellis' `steep' hypothesis. From \zdbk, $\eta$ inherits the (strict) convexity and differentiability of $\d$, which gives \zdbr.\endbox We complete this Section with a pair of results about the large-deviation rate function $K$. The former of these is in the spirit of Proposition~D of \bwii\ and identifies the points where the various suprema in Legendre transforms \zdbq\ and \zdbr\ are achieved. This leads to central limit results and the major result of the next Section. \proc{\pnum{Proposition}{zxdbfive}} {\sl Under the conditions of \zxdbthree, $K$ is finite, twice differentiable and strictly convex on $\Int(M)$, and the supremum of \zdbq\ is attained uniquely (up to multiples of $\bf 1$) at $v=\nabla K(x)$, and the supremum of \zdbr\ is attained uniquely at $x=\nabla\eta(v)$.} \proof{\zxdbfive} It is immediate from its definition that $\d(v)/\|v\|_\infty\rightarrow1$ as $\|v\|_\infty\rightarrow\infty$ with $v\geq0$. But as also $|\d(v)|\leq\|v\|_\infty$, the Dominated Convergence theorem gives us that $\eta(v)/\|v\|_\infty$ goes to 1 as well. Take $x\in\Int(M)$ and suppose there exists a sequence of vectors $(v_n)$ such that $$ \brak{v_n}{x}-\eta(v_n)\rightarrow\infty. $$ Without loss of generality we can replace $(v_n)$ by $(v_n-(\min_iv_n(i))\bf 1)$, because $\eta(v+\a{\bf 1})=\eta(v)+\a$, and thus assume that the $(v_n)$ are positive, with at least one zero co-ordinate. The sequence must still get infinitely large, but $$ \brak{v_n}{x}-\eta(v_n)\leq\|v\|_\infty\left((1-\min_ix_i)- \eta(v_n)/\|v_n\|_\infty\right),\enum{zdbs} $$ which is large and negative for large $n$, contradicting our supposition of $K(x)=\infty$. As remarked in \zxdbthree, $\eta$ inherits the smoothness and the strict convexity on ${\bf 1}^\bot$ of $\d$. Its continuity means that the supremum must be attained at some finite point $\hat v(x)$, and the convexity gives the uniqueness. The differentiability shows that the maximizing $\hat v$ will be the solution of $\nabla\eta(v)=x$. We can expand $\hat v$ around an $x$ as $\hat v(x+\e)=\hat v(x)+H_\eta^{-1}\e$ to see that $$ K(x+\e)=K(x)+\brak{\hat v(x)}{\e}+\half\e^\top H_\eta^{-1}\e+o(\e^2). \enum{zdbt} $$ Thus $K$ is twice differentiable, $\nabla K(x)-\hat v(x)$ is a multiple of $\bf 1$, and $K$ is locally (and hence globally) strictly convex. (Technical note: we are regarding $H_\eta$, the Hessian of $\eta$, as an automorphism of ${\bf 1}^\bot$.) By the above $x=\nabla\eta(v)$ is a solution of \zdbr, and the strict convexity of $K$ shows it is unique.\endbox In the simple example studied in Section (d) of \bwii, the rate function was calculated exactly as $K(x)=\sum\pi_i\log(\pi_i/x_i)$, which is infinite on the boundary of $M$, whilst the Ces\`aro rate function $I$ is finite everywhere. Note that we can see that $I$ is finite in the general \zxdbthree\ situation by considering equation \zdbi. We shall have further remarks about this example in the next Section, but for the moment we derive a necessary and sufficient condition for $K$ to be everywhere finite or infinite. \proc{\pnum{Proposition}{zxdcsix}} {\sl Under the conditions of \zxdbthree, the rate function $K$ is either everywhere finite or everywhere infinite on the boundary of $M$ according as to whether the support of the discount function $m$ is of finite or infinite (Lebesgue) length.} \proof{\zxdcsix} Firstly let us define $V_+$ to be the space of elements of $(\Bbb R^+)^n$ which have at least one zero component. We note that $\l V_+=V_+$ for any positive $\l$, which is a feature we shall use later. For $x$ in $\Int(M)$, we take $v_x$ to be the unique choice in $V_+$ of the $v=\nabla K(x)$ in \zxdbfive. In fact the pair $(\nabla K,\nabla\eta)$ represents a homeomorphism between $\Int(M)$ and $V_+$. Then $x=\nabla\eta(v_x)$, so by taking the gradient of \zdbk, we can write $x$ as $$ x=\int_0^\infty m_t\nabla\d(m_tv_x)\,dt. $$ Then because $\brak{v}{\nabla\d(v)}=\d(v)+I(\nabla\d(v))$ for all $v$ in $\Bbb R^n$, $$ \brak{v_x}{x}=\int_0^\infty\Bigl(\d(m_tv_x)+I\bigl(\nabla\d(m_tv_x)\bigr) \Bigr)\,dt=\eta(v_x)+\int_0^\infty I\bigl(\nabla\d(m_tv_x)\bigr)\,dt. $$ As $v_x$ is the optimal $v$ in \zdbq, we can express $K(x)$ as $$ K(x)=\int_0^\infty I\bigl(\nabla\d(m_tv_x)\bigr)\,dt.\enum{zdbspecial} $$ Thus, for an upper bound, $$ K(x)\leq\sup_{y\in M}I(y)\,\int_0^\infty I_{\{m_t>0\}}\,dt= \sup_{y\in M}I(y)\,\Leb\supp(m), $$ and so $K$ is bounded on all of $M$ if the support of $m$, $\supp(m)$, is compact. The rate function $I$ is only 0 in $M$ at $\pi$, and $\nabla\d$ only takes the value $\pi$ in $V_+$ at 0. Thus from \zdbspecial\ we have the lower bound $$ K(x)\geq\Leb\bigl\{t:m_t\geq\|v_x\|_\infty^{-1}\,\bigr\}\, \inf\bigl\{I\bigl(\nabla\d(v)\bigr):v\in V_+,\ \|v\|_\infty\geq1\,\bigr\}>0. $$ Now as $x$ tends towards $\partial M$, the boundary of $M$, the vector $v_x$ tends to infinity in $V_+$. So if $m$ has unbounded support, then $K(x)$ tends to infinity as $x$ tends to $\partial M$. The intuition, of course, is that $X$ can with positive probability avoid hitting a certain state for all times in a finite length set but not for all times in an infinite length set. \endbox \newsect{More exact results for Markov chains}% Our aim is to obtain a sharper version of \zdbk\ for finite Markov chains, and then to derive more terms of the asymptotic expansion of the density of $\Al$. The initial case studied in \bwii\ was of a symmetrizable (reversible) Markov chain and a smooth discount density $m$. It turns out that $m$ need only be of bounded variation (see below), but for technical ease we shall give the proof first in the case where $m$ is also absolutely continuous. More interestingly, the symmetrizability is seen now to have only been needed to make one of the eigenvalues of $Q$ real and its corresponding eigenvector orthogonal to the others. This in fact happens automatically because every (non-diagonal) element of $Q$ is non-negative (we say that $Q$ is {\it essentially non-negative}). The following theorem collects all the facts about non-negative matrices that we will need. \proc{\pnum{Theorem}{zxdcone}} {\sl Let $R$ be an essentially non-negative $n\times n$ matrix. Let $\d$ be its principal eigenvalue (the one with greatest real part). Then $\d$ is itself real, and its corresponding eigenvector is non-negative and no other is positive. If, in addition, $R$ is irreducible (in the stochastic sense), then $\d$ is simple, its eigenvector is strictly positive and no other is non-negative, and there exists a real diagonal matrix $F$ with positive elements such that $S:=\d I-F^{-1}RF$ has a simple eigenvalue zero, with an orthogonal eigenprojection $P$, and that, for some positive $\sigma$} $$ \brak{Sx}{x}\geq\sigma\|(I-P)x\|^2,\qquad\hbox{\sl for all $x\in\Bbb R^n$.} \enum{zdca} $$ (Where $\brak{\ }{\ }$ and $\|\ \|$ are the standard inner product and its norm on $\Bbb R^n$, and an orthogonal projection $P$ satisfies $P^\top=P^2=P$.) \proof{\zxdcone} For the first parts see the Perron-Frobenius theorem in, for example, theorem 1.5 of \seneta\ or theorems I.7.5 and I.7.10 of \kato. For the existence of $F$ see Theorem~B of \baxiii, which itself is adapted from theorem I.7.13 of \kato.\endbox We recall that the {\it variation} of a measurable function $x:[0,\infty)\rightarrow\Bbb R$ on $[a,b]$ is defined as $$ V_x(a,b):=\sup\sum_{i=1}^n|x(t_i)-x(t_{i-1})|,\enum{zdcb} $$ where the supremum is taken over all partitions: $a=t_0t}(1+|\D y_s|)\exp V_{y^c}(t,\infty).\enum{zdcea} $$ Another useful result follows from integration by parts, in that $$ \half d\|x_t\|^2=\brak{x_t}{dx_t}-\half\|\D x_t\|^2.\enum{zdceb} $$ \proc{\pnum{Theorem}{zxdcextra}} {\sl \zxdctwo\ remains true if $m$ is a discontinuous non-negative density of bounded variation.} \proof{\zxdcextra} We follow the proof of \zxdctwo\ exactly down to \zdcg, except that we take $R$, $F$ $S$ and $P$ to be functions of $t$ rather than $v$. We write $G:=F^{-1}$, $w:=Fy$ and $w^*:=Gy$, and take the normalisation that $\|y\|=1$ ($\iff\brak{w}{w^*}=1$) and $\brak{w}{dw^*}=0$ ($\iff\brak{dw}{w^*_-}=0$). Note that all these functions are BV. Then \zdch\ becomes $$ d\chi=\l^{-1}S\chi\,dt+dG\,F\chi.\enum{zdcec} $$ $$ \hbox{\rm So}\quad\half d\|\chi_t\|^2\geq\brak{dG\,F\chi}{\chi}_t -\half\|\D G F\chi\|^2_t\geq K_1\|\chi_t\|^2\|dG\|,\enum{zdced} $$ by \zdceb\ and \zdcea, for some constant $K_1$. Hence $\|\chi_t\|$ is uniformly bounded in $t$ by some constant $K_\chi$. Now using \zdcec\ and \zdceb\ we again work with the components of $\chi$ orthogonal to $y$, $$ \eqalign{ \half d\|\chi_-\|_t^2&\geq{\sigma_0\over\l}\|\chi_-\|_t^2\,dt +\brak{dG\,F\chi-dP\,\chi+\D P\D\chi}{\chi_-}_t-\half\|\D(P\chi)\|^2_t \cr &\geq{\sigma_0\over\l}\|\chi_-\|_t^2\,dt-K_2(\|dG_t\|+\|dP_t\|),\cr} $$ for some constant $K_2$. So a result of the same form as \zdcl\ holds. Finally we find that $$ d\xi=\xi\brak{w}{dw^*}+\brak{F\chi_-}{dw^*}, $$ and we have a bound similar to that of \zdcn, because $\brak{w}{dw^*}=0$. Explicitly, $w_t$ is the positive eigenvector of $Q+m_t V$ with the normalisation that $w_\infty={\bf 1}$ and $dw_t$ is orthogonal to the positive eigenvector of $Q^\top+m_{t-}V$.\endbox We can calculate an exact expression for $w$. \proc{\pnum{Lemma}{zxdcestwo}} {\sl Let $y(v)$ be the positive eigenvector of $Q+V$ of constant norm, with $y(0)={\bf 1}$. If $w^0_t:=y(m_tv)$ and $w^*_t$ is the positive eigenvector of $Q^\top+m_tV$ satisfying $\brak{w^0_t}{w^*_t}=1$, then $$ w_t=w^0_t\prod_{s>t}(1+\brak{\D w^0_s}{w^*_{s-}})\exp\int_t^\infty \brak{dw^{0,c}_s}{w^*_s}. $$ Further $w_t=w_t(v,m)$ is continuous in $v$.} \proof{\zxdcestwo} If we set $w_t=r_tw^0_t$, then $$ dw_t=dr^c_t\,w^0_t+r_t\,dw^0_t+\D r_tw^0_{t-}\quad \hbox{\rm so}\quad\brak{dw_t}{w^*_{t-}}=dr_t+r_t\brak{dw^0_t}{w^*_{t-}}. $$ An application of \zdcea\ gives the expression for $w$. Elementary perturbation results, in for example Kato~[7], tell us that $y$ is smooth in $v$ and so $$ dw^0_t(v)=\brak{v}{\nabla}y(m_tv)\,dm^c_t+\D(w^0_t(v)). $$ Thus the difference $dw^0_t(v)-dw^0_t(u)$ can be written as $$ \displaylines{ \Bigl(\brak{v-u}{\nabla}y(m_tv) +\brak{u}\nabla(y(m_tv)-y(m_tu))\Bigr)\,dm^c_t+\D(y(m_tv)-y(m_tu)).\cr \rlap{\hbox{\rm Hence}}\hfill\quad|dw^0_t(v)-dw^0_t(u)|\leq K_1\|v-u\|\, |dm^c_t| +K_2\|v-u\|\,|\D m_t|,\hfill\cr} $$ where $K_1:=\sup_{K_V}\|\nabla y(v)\|+V_K\sup_{K_V}\|v\|\,\|\nabla y(v)- \nabla y(u)\|/\|v-u\|$ and some constant $K_2$. Hence $w_t$ is (Lipschitz) continuous in $v$.\endbox \proc{\pnum{Theorem}{zxdcfour}} {\sl Let $X$ be an irreducible continuous-time Markov chain on a finite set $S$, with Q-matrix $Q$. Let $m$ be a non-negative density on $[0,\infty)$ of bounded variation. Then where $f_i^\l$ is the density of $A_\l$ on $M$ under the law starting $X$ at $i$, the $(f_i^\l)$ can be written as $$ f^\l_i(x)=e^{-K(x)/\l}(2\pi\l)^{-(n-1)/2}\bigl(\det H_K(x)\bigr)^{1/2} z_i(x)r^\l_i(x),\enum{zdcp} $$ where $K$ is as defined by \zdbq, $H_K$ denotes its Hessian taken with respect to $M$, $z(x)$ is the positive eigenvector of $Q+\diag(m_0\nabla K(x))$, and the residue term $r^\l$ goes to 1 as $\l$ goes to 0, in the sense that $$ \limsup_{\l\downarrow0}\int_F r_i^\l(x+\sqrt\l y)\,dy\leq|F|,\quad \hbox{\sl and}\quad\liminf_{\l\downarrow0}\int_G r_i^\l(x+\sqrt\l y)\,dy \geq|G|,\enum{zdcq} $$ for all $x$ in $\Int(M)$, and for $F$ and $G$ respectively closed and open bounded subsets of ${\bf 1}^\bot$.} \noindent{\bf Notes:} (1) We take the Hessian regarding $K$ as a function on an open subset of $\Bbb R^{n-1}$, that is $K(x_1,\ldots,x_{n-1},1-\sum_i^{n-1}x_i)$. See the example at the end of this paper. \noindent (2) Unfortunately we would really like to prove the result that $$ \Bbb P_i(A_\l\in H)\sim\int_H e^{-K(x)/\l} (2\pi\l)^{-(n-1)/2}\bigl(\det H_K(x)\bigr)^{1/2}z_i(x)\,dx,\enum{zdcr} $$ for suitable $H$, as $\l$ goes to 0. This could be proved if the integrand in our control of $r^\l$ was $r_i^\l(x+\l y)$ rather than $r_i^\l(x+\sqrt\l y)$. \proof{\zxdcfour} \zxdctwo\ can be taken as saying that for $v$ in $\Bbb R^S$, $$ f_i^{v,\l}(x):=\exp\Bigl(\l^{-1}\bigl(\brak{v}{x}-\eta(v)\bigr)\Bigr) f_i^\l(x)/w_i(v)\enum{zdcs} $$ is (asymptotically) a density on $M$, where $w_i(v)$ is the $w_0(v,m)(i)$ of \zxdcestwo, and our $z_i(x)$ will be $w_i(\nabla K(x))$. If $A_\l^v$ under $\Bbb P_i$ has the law $f_i^{v,\l}$, then we can derive a central limit result by considering $Z_\l^v:=(A_\l^v-\nabla\eta(v))/\sqrt\l$. We see that for $u$ in $\Bbb R^S$, $$ \eqalignno{\relax\Bbb E_i\bigl(\exp\brak{u}{Z_\l^v}\bigr) &=\exp\Bigl(\l^{-1}\bigl(\eta(v+\sqrt\l u)-\eta(v)- \brak{\sqrt\l u}{\nabla\eta(v)}\,\bigr)\Bigr)\cr &\qquad\ldots\Bigl(w_i(v+\sqrt\l u)+o(1)\Bigr)/w_i(v), &\lnum{zdct}\cr} $$ using the local uniformity in $v$ of the convergence of $o(1)$. As $\eta$ inherits the smoothness of $\d$, we can expand it about $v$ as $$ \eta(v+\sqrt\l u)=\eta(v)+\brak{\sqrt\l u}{\nabla\eta(v)}+ \half\l u^\top H_\eta(v)u+o(\l),\enum{zdcu} $$ and hence deduce that $$ \lim_{\l\downarrow0}\Bbb E_i\bigl(\exp\brak{u}{Z_\l^v}\bigr)= \exp\bigl(\half u^\top H_\eta(v)u\bigr).\enum{zdcv} $$ In other words $$ Z_\l^v\tendsd N(0,H_\eta(v)). $$ We can think of $f_i^{v,\l}$ as the distribution of $\Al$ conditioned in some way to converge to $\nabla\eta(v)$, but we do not make this formal. \zxdbfive\ provides the interpretation of $\nabla\eta(v)$ as the maximizing $x$ in the Legendre transform. Recall that a sequence of laws $(\nu_n)$ on a Polish space $E$ converges to a law $\nu$ with respect to the weak topology on $M_1(E)$ if $\brak{v}{\nu_n}\rightarrow\brak{v}\nu$ for all $v$ in $C_b(E)$. \billingsley, 2.1, shows that this is equivalent to each of the following $$ \eqalign{\limsup_{n\rightarrow\infty}\nu_n(F)&\leq\nu(F)\qquad \hbox{\rm $F$ closed in $E$,}\cr \liminf_{n\rightarrow\infty}\nu_n(G)&\geq\nu(G)\qquad \hbox{\rm $G$ open in $E$,}\cr \hbox{\rm and}\qquad\lim_{n\rightarrow\infty}\nu_n(H)&=\nu(H)\qquad \hbox{\rm $H$ in $E$ with $\nu(\partial H)=0$.}\cr} $$ Setting $x=\nabla\eta(v)$, we recall from \zxdbfive\ that $\nabla K(x)$ is $v$, up to a multiple of~$\bf 1$. The asymptotics of the density of $Z_\l^v$ are given by $$ f_i^{Z,\l}(y):=\l^{(n-1)/2}f_i^{v,\l}(x+\sqrt\l y) \sim(2\pi)^{-(n-1)/2}\det H_K(x)^{\half}e^{-\half y^\top H_K(x)y}\, r_i^\l(x+\sqrt\l y), $$ because $$ \eqalign{&\brak{v}{x+\sqrt\l y}-\eta(v)-K(x+\sqrt\l y)\cr &\quad=\brak{v}{x}-\eta(v)-K(x)+\sqrt\l\brak{v-\nabla K(x)}{y} -\half\l y^\top H_K(x)y+o(\l)\cr &\quad=\l(-\half y^\top H_K(x)y+o(1)).\cr} $$ The normal distribution $N(0,H_\eta(v))$ itself has density $$ f(y)=(2\pi)^{-(n-1)/2}\det H_K(x)^{\half} e^{-\half y^\top H_K(x)y}, $$ as \zxdbfive\ tells us that $H_K=H_\eta^{-1}$ on $M$. By Lemma I.45.1 of \williams, if $H$ is bounded and $|\partial H|=0$ then $$ \int I_H(y)(f(y))^{-1}f_i^{Z,\l}(y)\,dy\rightarrow|H|,\qquad \hbox{\rm or}\qquad \int_H r_i^\l(x+\sqrt\l y)\,dy\rightarrow|H|.\enum{zdcw} $$ Hence by the equivalence of the above expressions for weak convergence, the result is proved.\endbox The following Corollary is intended in the way of a remark, and was the original statement of \zxdcfour, but is now seen to be weaker, although perhaps a more natural formulation. \proc{\pnum{Corollary}{zxdcfive}} {\sl Under the conditions of \zxdcfour, $$ \limsup_{\l\downarrow0}\int_F r_i^\l(x)\,dx\leq|F| \qquad\hbox{\sl and} \qquad\liminf_{\l\downarrow0}\int_G r_i^\l(x)\,dx\geq|G|,\enum{zdcx} $$ for $F$ closed in $\Int(M)$ and $G$ open in $\Int(M)$. In other words, $r_i^\l(x)\,dx$ converges weakly to $dx$ on $\Int(M)$.} \proof{\zxdcfive} Take $G$ open in $\Int(M)$, $\d$ small and positive with $G_\d:=\{y\in G:B(y,\d)\subseteq G\}$, and $B$ a ball around 0, then by Fatou's lemma and Fubini's theorem $$ \eqalign{|G_\d|\,|B|&=\int_{G_\d}\left(\liminf_{\l\downarrow0} \int_B r_i^\l(x+\sqrt\l y)\,dy\right)\,dx\cr &\leq\liminf_{\l\downarrow0}\int_{G_\d}\int_B r_i^\l(x+\sqrt\l y)\,dy\,dx \leq\left(\liminf_{\l\downarrow0}\int_G r_i^\l(x)\,dx\right)\,|B|.\cr} $$ Letting $\d$ tend to 0, we have one of our bounds. For some $F$ closed in $\Int(M)$, we need $\int_B r_i^\l(x+\sqrt\l y)\,dy$ to be uniformly bounded on $F$ and for $\l$ near 0. It is, and the bound is $$ \sup_{x\in F}(2\pi)^{(n-1)/2}\det H_K(x)^{-\half}\sup_{y\in B} e^{\half y^\top H_K(x)y}<\infty. $$ Working with this $F$ and with $F^\d:=\{y\in M:d(y,F)\leq\d\}$, we can show in a similar way that $$ \limsup_{\l\downarrow0}\int_F r_i^\l(x)\,dx\leq|F^\d|, $$ and hence we are home.\endbox \noindent{\bf Some remarks.} \noindent{\bf (a)} Were the $r_i^\l$ to be equicontinuous (or some such condition) we would have that $r_i^\l(x)\rightarrow1$ for all $x$ and hence that $f_i^\l(x)/f_j^\l(x)\rightarrow z_i(x)/z_j(x)$ and $\lim-(Qf^\l)_i/f^\l_i$ differs from $m_0\nabla K(x)$ only by a multiple of $\bf 1$, as in Section (c) of \bwii, where the choice of $\nabla K(x)$ in $\ker(\d)$ was called $g(x)$. \noindent{\bf (b)} Note that the proof of \zxdcfour\ gives us a central limit theorem for $\Al$ as $$ Z_\l:=(A_\l-\pi)/\sqrt\l\tendsd N(0,H_\eta(0)).\enum{zdcy} $$ Taking a Taylor expansion of $\d$ about 0 and integrating we discover that $H_\eta(0)=\sigma^2 H_\d(0)$, where $\sigma=\|m\|_{L_2}$ which is finite because $m$ is in both $L_1$ and $L_\infty$. \noindent{\bf Example.} (This case was first studied in Section (d) of \bwii.) Suppose we have a Markov chain which is symmetric and space-homogeneous, with Q-matrix $q_{ij}:=\pi_j-\d_{ij}$, where $\pi$ is a distribution on a finite set $S$. The Ces\`aro large-deviation rate function is $I(x)=1-(\sum\sqrt{\pi_ix_i})^2$, and the exponentially discounted large-deviation rate is $K(x)=\sum\pi_i\log(\pi_i/x_i)$. We found then that $\d(v)$ is the unique root $\d$ in $(\max_i(v_i-1),\infty)$ of $$ \sum_{i\in S}{\pi_i\over\d+1-v_i}=1, $$ and that $\eta$ is given by $\eta(v)=\d(v)-\sum_i\pi_i\log(\d(v)+1-v_i)$. We find now that $$ \eqalign{\nabla_i I(x)&=1-\sqrt{\pi_i\over x_i}\Bigl(\sum_{j\in S} \sqrt{\pi_j x_j}\Bigr),\cr \nabla_i\d(v)&={\pi_i\over(\d(v)+1-v_i)^2}\Big/\sum_{j\in S} {\pi_j\over(\d(v)+1-v_j)^2},\cr \nabla_iK(x)&=1-{\pi_i\over x_i},\cr \hbox{\rm and}\quad\nabla_i\eta(v)&={\pi_i\over\d(v)+1-v_i}.\cr} $$ Here we chose $\nabla I$ and $\nabla K$ to be in the kernel of $\d$. The distribution of $\Al$ can be calculated explicitly to be a multidimensional $\b$-distribution with density $$ f_i^\l(x)={x_i\over\pi_i}\Gamma(\l^{-1})\prod_{j\in S}{x_j^{(\pi_j/\l)-1} \over\Gamma(\pi_j/\l)}. $$ Note that the Hessian of $K$ on $M$ is not the same as that derived from the extension of $K$ to $\Bbb R^S$, but by using any of the following co-ordinate schemes: $$ \eqalign{ K_i : \Bbb R^{S\backslash\{i\}}&\longrightarrow\Bbb R \qquad\hbox{\rm for each $i\in S$}\cr \hbox{\rm where}\qquad(x_j)_{j\neq i}&\longmapsto K(x_1,\ldots,x_{i-1},1-\sum_{j\neq i}x_j,x_{i+1},\ldots,x_n)\cr \noalign{\smallskip} \hbox{\rm or}\qquad K_0 : \Bbb R^n&\longrightarrow\Bbb R\cr \hbox{\rm where}\qquad x&\longmapsto K(x+(1-{\bf 1}^\top x){\bf1}/n)+ \half({\bf1}^\top x)^2.\cr} $$ What is happening here is that our choice of basis for evaluating the Hessian corresponds to our choice of basis for integrating which was made back at the start of Section (b) of \bwii. The $K_0$ representation projects onto $M$ and adds a strictly convex term which is perpendicular to $M$. This representation is more natural, though cumbersome to calculate with, and can be shown equivalent to any of the others by verifying that the change of basis matrix has determinant one. Thus the Hessian (in the $K_n$ realisation) and its determinant are given by $$ H_{K_n}(x)_{ij}={\pi\over x_i^2}\d_{ij}+{\pi_n\over x_n^2}\qquad \hbox{\rm and}\qquad\det(H_K(x))= \Bigl(\prod_{i\in S}{\pi_i\over x_i^2}\Bigr)\sum_{j\in S}{x_j^2\over\pi_j}. $$ The normalisation of the eigenvector $z_i(x)$ is that $\|z\|_\pi=1$, so it is given by $$ z_i(x)={x_i\over\pi_i}\Bigl(\sum_{j\in S}{x_j^2\over\pi_j}\Bigr)^{-1/2}. $$ (The corresponding vector for the Ces\`aro case is $\sqrt{x_i/\pi_i}(\sum_j\sqrt{\pi_jx_j})$.) We can now calculate the residual functions using Stirling's formula $$ \Gamma(x)=\sqrt{2\pi}x^{x-\half}e^{-x}(1+O(x^{-1})), $$ where $|O(x^{-1})|\leq K/x$ as $x\rightarrow\infty$ for some constant $K$. It is thus discovered that the residual functions $r_i^\l(x)$ can be calculated and are found to be independent of both $i$ and $x$, and are of size $1+O(\l)$. \proc{\pnum{Hypothesis}{pxchyp}} We recall from Theorem~C of \bwii\ that in the set-up of \zxdcfour\ with the exponential discount ($m_t=e^{-t}$), the density $f^\l$ satisfies the vector differential equation $$ \L f^\l=-\l^{-1}Qf^\l,\enum{zdcha} $$ where $\L$ is the matrix differential operator $\L=\diag(\sum_{j\neq i} (\partial_j-\partial_i)x_j)_{i\in S}$. Here we have changed the domain of $f^\l$ from a subset of $\Bbb R^{n-1}$ equivalent to $M$, to a neighbourhood of $M$ in $\Bbb R^n$ by extension. The operator $\L$ is invariant to the extension chosen. If we discount $f^\l$ by the known large-deviation rate function $K$, that is by defining $g^\l$ by $$ \displaylines{f^\l(x)=e^{-K(x)/\l}(2\pi\l)^{-(n-1)/2}g^\l(x)\cr \rlap{\hbox{\rm then}}\hfill \L g^\l=\l^{-1}R\bigl(\nabla K(x)\bigr)g^\l. \hfill\cr} $$ This compares with equation \zdce\ which said that $$ \partial_t\psi=\l^{-1}R(m_tv)\psi, $$ where $\psi$ is the discount of $\phi$ as defined by \zdcd. The matrix $R(v)$ has a simple eigenvalue 0 and all other eigenvalues have positive real part. We saw that $\psi$ tended to a multiple of the 0-eigenvector of $R(m_tv)$ as $\l$ went to 0, and also that $g^\l$ tended (in some sense) to $z(x)$, which was the 0-eigenvector of $R(\nabla K(x))$. We can formulate an analogue of \zdcha\ for the general discount case as follows. Let us write $A_{\l,t}$ for $\l M_t^{-1}\int_0^\infty\t_tm(\l s)\d_{X_s}\,ds$, where $M_t:=\int_0^\infty\t_tm(s)\,ds$, and $f^{\l,t}_i$ for the density of $A_{\l,t}$ if $X$ starts in state $i$. Then $A_{\l,t}$ will satisfy the large-deviation property with rate function $K_t$, where $$ K_t=\eta_t^*\quad\hbox{\rm where}\quad\eta_t(v):=\int_0^\infty \d(M_t^{-1}\t_tm(s)v)\,ds, $$ and we write $f^{\l,t}$ as $$ \displaylines{ f^{\l,t}(x)=e^{-K_t(x)/\l}(2\pi\l)^{-(n-1)/2}g^{\l,t}(x).\cr \rlap{\hbox{\rm Then}}\hfill M_t^{-1}m_t\L g^{\l,t}+\partial_t g^{\l,t} =\l^{-1}R\bigl(M_t^{-1}m_t\nabla K_t(x)\bigr)g^{\l,t}.\hfill\cr} $$ Again \zxdcfour\ tells us that $g^{\l,t}$ tends (in some sense) to the 0-eigenvector, $z$, of the matrix $R$. We hypothesize that the convergence is in fact pointwise. \newpage \centerline{\bf REFERENCES} \vskip0.7truein \input ref \end % End of Text