INSTRUCTIONS The text between the lines BODY and ENDBODY is made of 1342 lines and 53152 bytes (not counting or ) In the following table this count is broken down by ASCII code; immediately following the code is the corresponding character. 31455 lowercase letters 2069 uppercase letters 1089 digits 6674 ASCII characters 32 30 ASCII characters 33 ! 6 ASCII characters 34 " 119 ASCII characters 35 # 1284 ASCII characters 36 $ 509 ASCII characters 37 % 25 ASCII characters 38 & 101 ASCII characters 39 ' 667 ASCII characters 40 ( 666 ASCII characters 41 ) 36 ASCII characters 42 * 96 ASCII characters 43 + 828 ASCII characters 44 , 197 ASCII characters 45 - 491 ASCII characters 46 . 23 ASCII characters 47 / 77 ASCII characters 58 : 44 ASCII characters 59 ; 23 ASCII characters 60 < 220 ASCII characters 61 = 45 ASCII characters 62 > 4 ASCII characters 63 ? 3 ASCII characters 64 @ 58 ASCII characters 91 [ 3318 ASCII characters 92 \ 57 ASCII characters 93 ] 274 ASCII characters 94 ^ 1001 ASCII characters 95 _ 6 ASCII characters 96 ` 707 ASCII characters 123 { 240 ASCII characters 124 | 707 ASCII characters 125 } 3 ASCII characters 126 ~ BODY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%% PAPER %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \magnification = 1200 \font\bfb=cmbx10 scaled\magstep1 \font\bfp=cmbx7 \font\rmb=cmr10 scaled\magstep1 \font\rmg=cmr10 scaled\magstep2 \font \fBbb=msbm10 \def\vec#1{{\bf #1}} \def\noi{\noindent} \def\lan{\langle} \def\ran{\rangle} \def\dd{\partial} \def\ti{\tilde} \def\de{\delta} \def\ga{\gamma} \def\ka{\kappa} \def\si{\sigma} \def\a {\alpha} \def\b {\beta} \def\e{\varepsilon} \def\l {\lambda} \def\o {\omega} \def\Ga {\Gamma} \def\L {\Lambda} \def\O {\Omega} \def\Te{\Theta} \def\p {\varphi} \def\t {\theta} \def\E {{I\!\! E\,}} \def\PP {{I\!\! P\,}} \def\D {{\cal D}} \def\G {{\cal G}} \def\ec {\widetilde {\E}} \def\ve {\varepsilon} \def\wt {\widetilde} \def\wh {\widehat} \def\Cal #1{{\cal #1}} \def\Bbb #1{{\hbox{\fBbb #1}}} \def\RR{I\!\! R} \def\NN{I\!\! N} \def\Z{Z\!\!\!\!Z\!\!\!\!Z} \def\ZZ{Z\!\!\!\!Z} \def\sqr#1#2{{\vcenter{\vbox{\hrule height.#2pt \hbox{\vrule width.#2pt height#1pt \kern#1pt \vrule width.#2pt}\hrule height.#2pt}}}} \def\qed{\mathchoice\sqr64\sqr64\sqr{2.1}3\sqr{1.5}3} \def\missing{{\rmg missing}} \def\equiv{:=} %%%%%the following macro marks the footnote with a raising number%%%%% \newcount\footnotenumber \newtoks\everyfootnote \def\nfootnote{\global\advance\footnotenumber by 1 \footnote{$^{\number\footnotenumber}$}}% %%%%%end macro for nfootnotes%%%%%%%%%%%%%%%%%%%%%%% %Je t'envois ci apres une nouvelle version de %numerotation de formules qui place une etiquette %a cote des theoremes aussi. %%%%%%%%%%%%%%%%% EQUAZIONI CON NOMI SIMBOLICI%%% %%% Per assegnare un nome simbolico ad una equazione basta %%% scrivere \Eq(...) o, in \eqalignno, \eq(...) o, %%% nelle appendici, \Eqa(...) o \eqa(...); %%% dentro le parentesi e al posto di ... si puo' scrivere qualsiasi %%% commento; %%% per avere i nomi simbolici segnati a sinistra delle formule si deve %%% dichiarare il documento come bozza, iniziando il testo con %%% \BOZZA. Sinonimi: \Eq,\EQ,\EQS; \eq,\eqs; \Eqa,\Eqas;\eqa,\eqas. %%% All' inizio di ogni paragrafo si devono definire il %%% numero del paragrafo e della prima formula dichiarando %%% \numsec=... \numfor=... (brevetto Eckmannn). %%% Si possono citare formule seguenti; le corrispondenze fra nomi %%% simbolici e numeri effettivi sono memorizzate nel file %%% \jobname.aux, che %%% viene letto all'inizio, se gia' presente. E' possibile citare anche %%% formule che appaiono in altri file, purche' sia presente il %%% corrispondente file .aux; basta includere all'inizio l'istruzione %%% \include{nomefile} %%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \global\newcount\numsec \global\newcount\numfor \global\newcount\numtheo \global\advance\numtheo by 1 \gdef\profonditastruttura{\dp\strutbox} \def\senondefinito#1{\expandafter\ifx\csname#1\endcsname\relax} \def\SIA #1,#2,#3 {\senondefinito{#1#2}% \expandafter\xdef\csname #1#2\endcsname{#3}\else \write16{???? ma #1,#2 e' gia' stato definito !!!!} \fi} \def\etichetta(#1){(\veroparagrafo.\veraformula)% \SIA e,#1,(\veroparagrafo.\veraformula) % \global\advance\numfor by 1% \write15{\string\FU (#1){\equ(#1)}}% \write16{ EQ #1 ==> \equ(#1) }} \def\letichetta(#1){\veroparagrafo.\verotheo \SIA e,#1,{\veroparagrafo.\verotheo} \global\advance\numtheo by 1 \write15{\string\FU (#1){\equ(#1)}} \write16{ Sta \equ(#1) == #1 }} \def\tetichetta(#1){\veroparagrafo.\veraformula %%%%copy four lines \SIA e,#1,{(\veroparagrafo.\veraformula)} \global\advance\numfor by 1 \write15{\string\FU (#1){\equ(#1)}} \write16{ tag #1 ==> \equ(#1)}} \def\FU(#1)#2{\SIA fu,#1,#2} \def\etichettaa(#1){(A\veroparagrafo.\veraformula)% \SIA e,#1,(A\veroparagrafo.\veraformula) % \global\advance\numfor by 1% \write15{\string\FU (#1){\equ(#1)}}% \write16{ EQ #1 ==> \equ(#1) }} \def\BOZZA{\def\alato(##1){% {\rlap{\kern-\hsize\kern-1.4truecm{$\scriptstyle##1$}}}}% \def\aolado(##1){% {%\vtop to \profonditastruttura {%\baselineskip%\profonditastruttura\vss \rlap{\kern-1.4truecm{$\scriptstyle##1$}}}}} } \def\alato(#1){} \def\aolado(#1){} \def\veroparagrafo{\number\numsec} \def\veraformula{\number\numfor} \def\verotheo{\number\numtheo} \def\verafigura{\number\numfig} \def\Eq(#1){\eqno{\etichetta(#1)\alato(#1)}} \def\eq(#1){\etichetta(#1)\alato(#1)} \def\leq(#1){\leqno{\aolado(#1)\etichetta(#1)}} %%%%%this line for \leqno \def\teq(#1){\tag{\aolado(#1)\tetichetta(#1)\alato(#1)}} %%%%%this line for \tag \def\Eqa(#1){\eqno{\etichettaa(#1)\alato(#1)}} \def\eqa(#1){\etichettaa(#1)\alato(#1)} \def\eqv(#1){\senondefinito{fu#1}$\clubsuit$#1 \write16{#1 non e' (ancora) definito}% \else\csname fu#1\endcsname\fi} \def\equ(#1){\senondefinito{e#1}\eqv(#1)\else\csname e#1\endcsname\fi} %next six lines by paf (no responsibilities taken) \def\Lemma(#1){\aolado(#1)Lemma \letichetta(#1)}% \def\Theorem(#1){{\aolado(#1)Theorem \letichetta(#1)}}% \def\PProposition(#1){\aolado(#1){Proposition \letichetta(#1)}}% \def\Corollary(#1){{\aolado(#1)Corollary \letichetta(#1)}}% \def\Remark(#1){{\noindent\aolado(#1){\bf Remark \letichetta(#1).}}}% \def\Definition(#1){{\noindent\aolado(#1){\bf Definition \letichetta(#1)$\!\!$\hskip-1.6truemm}}} \def\Example(#1){\aolado(#1) Example \letichetta(#1)$\!\!$\hskip-1.6truemm} \let\ppclaim=\plainproclaim \def\include#1{\openin13=#1.aux \ifeof13 \relax \else \input #1.aux \closein13 \fi} \openin14=\jobname.aux \ifeof14 \relax \else \input \jobname.aux \closein14 \fi \openout15=\jobname.aux \let\EQS=\Eq\let\EQ=\Eq \let\eqs=\eq \let\Eqas=\Eqa \let\eqas=\eqa %%%%%%%%%%%%%%%%%% Numerazione verso il futuro ed eventuali paragrafi %%%%%%% precedenti non inseriti nel file da compilare \def\include#1{\openin13=#1.aux \ifeof13 \relax \else \input #1.aux \closein13 \fi} \openin14=\jobname.aux \ifeof14 \relax \else \input \jobname.aux \closein14 \fi \openout15=\jobname.aux %%%%%%%%%%%%%%%%%%%%%%%%%%%% %\BOZZA \hfuzz 16pt \centerline{\bfb Reversibility in Infinite Hamiltonian Systems} \centerline{\bfb with Conservative Noise} \vskip.5cm \centerline{\bf In Memoriam Roland Dobrushin} \vskip1cm \centerline{{\rmb J\'ozsef Fritz}$^1$, {\rmb Carlangelo Liverani}$^2$ and {\rmb Stefano Olla}$^3$} \footnote{}{$^1$~Department of Probability and Statistics, E\"otv\"os Lor\'and University of Sciences, H-1088 Budapest, M\'uzeum krt. 6-8, Hungary. E-mail: jofri@cs.elte.hu} \footnote{}{$^2$~ II Universit\'a di Roma ``Tor Vergata'', Dipartimento di Matematica, 00133 Roma, Italy. E-mail: liverani@mat.utovrm.it} \footnote{}{$^3$~ Universit\'e de Cergy--Pontoise, D\'epartement de Math\'ematiques, 2 avenue Adolphe Chauvin, Pontoise 95302 Cergy--Pontoise Cedex, France and Centre de Math\'ematiques Appliqu\'ees, Ecole Polytechnique, 91128 Palaiseau Cedex, France.\hfill\break E-Mail: olla@paris.polytechnique.fr} \footnote{}{\bf\noindent Work partially supported by grants CIPA-CT92-4016 and CHRX-CT94-0460 of the Commission of the European Community, and by grant T 16665 of the Hungarian NSF. Two of us (J.F. and C.L.) acknowledge hospitality of Ervin Schr\"odinger Institute.} \vskip.5cm \noindent{\bf Abstract:} {\sl The set of stationary measures of an infinite Hamiltonian system with noise is investigated. The model consists of particles moving in $\RR^3$ with bounded velocities and subject to a noise that does not violate the classical laws of conservation, see [OVY]. Following [LO] we assume that the noise has also a finite radius of interaction, and prove that translation invariant stationary states of finite specific entropy are reversible with respect to the stochastic component of the evolution. Therefore the results of { [LO]} imply that such invariant measures are superpositions of Gibbs states.} \vskip.5cm {\bf 0. INTRODUCTION} \vskip .25cm \numsec=0\numfor=1\numtheo=1 Let $\O$ denote the space of locally finite configurations $\o=(q_\a,p_a)_{\a\in I}$ indexed by a countable set $I\,,$ that is $q_\a,p_\a\in\RR^3$ are the position and momentum of particle $\a\in I\,;$ the set $\{q_\a\}_{\a\in I}$ has no limit points in $\RR^3$ by assumption. The classical dynamics of the system is governed by a formal Hamiltonian $\Cal H\,,$ $$ {\cal H}(\o)=\sum_{\a\in I}\phi(p_\a)+{1\over2} \sum_{\a\in I}\sum_{\b\ne\a}V(q_\a-q_\b)\,, $$ where the kinetic energy $\phi:\RR^3\mapsto\RR$ is strictly convex with bounded derivatives, and $V:\RR^3\mapsto\RR$ is a symmetric and superstable pair potential of finite range. The associated Liouville operator will be denoted by $L\,,$ $$ L\p=\sum_{\a\in I}\quad\bigl\lan{\dd\Cal H\over\dd p_\a},{\dd\p\over\dd q_\a}\bigr\ran -\bigl\lan{\dd\Cal H\over\dd q_\a},{\dd\p\over\dd p_\a}\bigr\ran\,, $$ where $\lan\cdot,\cdot\ran$ denotes the usual scalar product in $\RR^3$. Almost nothing is known on the ergodic properties of such infinite systems. In fact, very few results are available even for finite systems of this type (e.g., [KSS], [DL], [BLPS], [LW]). To ensure a proper ergodic behavior of the system we add some noise whereby obtaining stochastic equations of motion; these equations read as $$ \eqalign{ dq_\a&=\phi'(p_\a)\,dt\,,\cr dp_\a&=-\sum_{\b\neq\a}V'(q_\a-q_\b)\,dt +b_\a(\o)\,dt+\sum_{\t=1}^d\sum_{\b\neq\a} \sigma^\t_{\a,\b}(\o)\,dw^\t_{\a,\b} ,} \Eq(stochdyn) $$ where $w^\t_{\a,\b}$ is a family of independent one-dimensional Wiener processes for $\t=1,2,...,d$ and $\a\ne\b$ such that $w^\t_{\a,\b}=-w^\t_{\b,\a}\,;$ $\phi'$ and $V'$ denote the gradient of $\phi$ and $V\,,$ respectively. The coefficients $b_\a,\sigma^\t_{\a,\b}:\O\mapsto\RR^3$ are smooth local functions to be specified in the next section in such a way that total energy and momentum are both preserved by the randomized evolution (0.1). In addition, any Gibbs state $\PP$ with energy $\Cal H$ will be a reversible measure for the stochastic part of the evolution: $$ \int \p(\o)\wh L\psi(\o)\,\PP(d\o)=\int\psi(\o)\wh L\p(\o)\,\PP(d\o) \Eq(reversibility)$$ for all smooth local functions $\p,\psi:\O\mapsto\RR\,,$ where $$ \wh L\psi=\sum_{\a\in I} \bigl\lan b_\a,{\dd\psi\over\dd p_\a}\bigr\ran +{1\over4}\sum_{\t=1}^d\sum_{\a\in I}\sum_{\b\ne\a} \bigl\lan\sigma^\t_{\a,\b},(D^2_{\a,\b}\psi)\sigma^\t_{\a,\b}\bigr\ran\,, \Eq(stochgen)$$ and $D^2_{\a,\b}\psi$ is the matrix of second derivatives obtained by applying $D_{\a,\b}=\dd/\dd_{p_\a}-\dd/\dd_{p_\b}$ twice to $\psi\,.$ Since the Liouville operator is antisymmetric with respect to Gibbs distributions, the full generator, $\wt L=L+\wh L$ also satisfies the stationary Kolmogorov equation, $\smallint \wt L\psi\,d\PP=0$ for a wide class of test functions $\psi$ and any Gibbs state $\PP\,.$ The converse problem is much more complex. In our basic reference [LO] it is shown that if a translation invariant measure $Q$ with finite specific entropy satisfies the stationary Kolmogorov equation and (0.2), together with some other technical conditions, then $Q$ enjoys the Gibbs property. Let us remark that finiteness of specific entropy is a fairly natural and effective condition in the theory of hydrodynamics limits (see [OVY]). On the contrary, condition (0.2) looks rather restrictive and, at least in general, not particularly natural. The main purpose of this paper is to show that condition (0.2) of reversibility is superfluous (i.e., it follows from the stationarity of the measure).\nfootnote{Note that in applications to hydrodynamics the reversibility (0.2) is insured by construction (see [OVY] lemma 4.4), hence the present paper does not add to hydrodynamics type problems for which the results in [LO] suffice. The focus here is on the classification of stationary translation invariant measures.} To obtain such a result we are forced to prove the existence of a semigroup defined by (0.1); its regularity (locality) will play a crucial role in the argument. This problem may not seem to be a very difficult one since $\phi'(p_\a)\,,$ the velocity of particle $\a$, is bounded by assumption. However, the evolution must be defined for a very large set $\bar\O\subset\O$ of initial configurations: we need $Q(\bar\O)=1$ for any probability measure $Q$ of finite specific entropy. On the other hand, to obtain the necessary regularity properties of the dynamics we have to restrict the configuration space by excluding extremely high values of particle density. We shall see that the desired construction fails unless the dimension of the space is less than four, cf. {[FD]} and {[S]}. \vskip .5cm {\bf 1. NOTATIONS AND RESULTS} \vskip.25cm \numsec=1\numfor=1\numtheo=1 Configurations can be interpreted as $\sigma$-finite integer valued measures on $\RR^3\times\RR^3\,;$ sometimes we write $\o=(q,p)$ with $q=(q_\a)_{\a\in I}$ and $p=(p_\a)_{\a\in I}\,,$ and if $\L\subset\RR^3$ then $\o_\L$ denotes the restriction of $\o$ to $\L\,,$ i.e. $\o_\L=(q_\a,p_\a)_{q_a\in\L}\,,$ while $|\o_\L|$ is the cardinality of this set. The centered cubic box of side $r>0$ will be denoted by $\Lambda_r\,.$ Referring to functionals $\o(\p):=\sum_{\a\in I} \p(q_\a,p_\a)\,,$ where $\p:\RR^3\times\RR^3\mapsto\RR$ is continuous with compact support, we equip $\O$ with the associated weak topology and Borel structure, and $C_0(\Omega)$ denotes the space of cylinder (local) functions $\Psi(\o)=f(\o(\p_1),\o(\p_2),...,\o(\p_n))$ such that $f\in C(\RR^n)\,.$ Since all sets $\Sigma(\delta)$, defined for an increasing sequence $\delta=(\delta_1,\delta_2,...)$ such that $\delta_1\ge 1$ by $$ \Sigma(\delta):=\bigl\{\o\in\O\,:\,|\o_{\Lambda_n}|\le\delta_n \hbox{ and } |p_\a|\le\delta_n\hbox{ if } q_\a\in\Lambda_n\bigr\} $$ are compact, we need not worry too much about topology. Indeed, in the forthcoming considerations we always do have an a priori bound allowing us to restrict calculations to some compact $\Sigma(\delta)\,.$ In this situations all reasonable topologies coincide, moreover any continuous function can be uniformly approximated by elements of $C_0(\Omega)$ in view of the Stone-Weierstrass Theorem. \vskip 5pt \noindent{\bf Interaction} \vskip 2pt We consider a repelling pair potential $V$ of finite range such that $V(x)=V(-x)$ is twice continuously differentiable, $V(0)>0$ but $V(x)=0$ if $|x|>R_0\,,$ finally $\lan x,V'(x)\bigr\ran\le0$ for all $x\in\RR^3\,.$ These conditions imply that $V$ is superstable, see [R]: for each cubic box or ball $\L\subset\RR^3$ there exist some constants $A_\L\ge0$ and $B_\L>0$ such that, for any configuration, we have $$ \sum_{\a:q_\a\in\Lambda}\sum_{\b\ne\a} V(q_\a-q_\b) \ge B_\L |\o_\Lambda|^2- A_\L |\o_\Lambda| . \Eq (superstability) $$ \vskip 5pt \noindent{\bf Kinetic energy}\vskip 2pt We assume that $\phi$ has bounded second derivatives, and velocities are also bounded, i.e. $|\phi'(y)|\le \bar c<+\infty$ for all $y\in\RR^3\,.$ To define Gibbs measures we need a lower bound: $\lim\inf_{|y|\to\infty}\phi(y)/|y|\ge \underline c>0\,.$ When results of [LO] are applied and extra technical condition on $\phi$ is needed. For simplicity one can consider the case in which $\phi (y)=\sum_{i=1}^3 \phi_0(y_i)$ for $y=(y_1,y_2,y_3)\,,$ where $\phi_0\in C^{\infty}(\RR)$ is strictly convex and $$ {1\over 2}{d^2\over du^2}(\phi_0''(u))^2= \phi_0'''(u)^2+\phi_0^{iv}(u)\phi''(u)\neq 0 $$ apart from, at most, finitely many points (see [LO], section two, ``Condition on the Noise" for the general condition that $\phi$ must satisfy). Notice that if ${d^2\over du^2}(\phi''_0(u))^2= 0$ for each $u$ and if we require the natural condition $\phi_0(u)=\phi_0(-u)$ then $\phi_0''$ is a constant, which is the classical case of a quadratic kinetic energy function. \vskip 5pt \noindent{\bf Stochastic perturbation of classical dynamics} \vskip 2pt There are several ways to select the coefficients of the stochastic perturbation, we set $$ b_\a(\o)=\sum_{\b\ne\a}\gamma_{\a,\b}(q)F(p_\a,p_\b)\hbox{ and } \sigma^\t_{\a,\b}(\o)=\sqrt{\gamma_{\a\b}(q)}G_\theta(p_\a,\,p_\b)\,, \Eq (coeff)$$ where, as in { [LO]},\nfootnote{In fact, in [LO], the functions $\gamma_{\a\b}$ depend only on the variables $q_\a$ and $q_\b$; yet all is done there applies without changes to the situation described here.} $\ga_{\a,\b}(q)=\ga_{\b,\a}(q)\ge 0$ is continuously differentiable, $\gamma_{\a,\b}(q)>0$ if $|q_\a-q_\b|< R_1$ and it is zero for $|q_\a-q_\b|\ge R_1\,,$ i.e. $R_1>R_0$ is the radius of stochastic interaction. The functions $F,G_\t:\RR^6\mapsto \RR^3$ are infinitely differentiable and bounded together with their derivatives; they are chosen in such a way that the stochastic interaction also preserves the total momentum and energy of an interacting couple of particles. Moreover, $\{G_\theta\}_{\theta=1}^d$ spans, at each point, all $\RR^3$. It is natural to assume that $\ga_{\a,\b}$ depends only on the interparticle distances, and it does not depend on a coordinate $q_\de$ if $|q_\a-q_\de|>R_2$ or $|q_\b-q_\de|>R_2\,,$ where $R_2>2R_1$ is a constant. Therefore the stochastic interaction is also translation invariant and has a finite range $R_3:=R_1+R_2\,.$ A new feature of the present model is that $\gamma_{\a,\b}$ vanishes when the number of particles near $q_\a$ or $q_\b$ tends to infinity. For convenience, we set $\ga_{\a,\b}(q)=\si(q_\a-q_\b)\Te_{\a,\b}(q)\,,$ where $$ \Te_{\a,\b}(q):=\Bigl(1+\sum_{\de\in I}\chi(q_\a-q_\de) +\sum_{\de\in I}\chi(q_\b-q_\de)\Bigr)^{-1} . \Eq(gamma)$$ In (1.3) $\sigma,\chi:\RR^3\mapsto[0,1]$ are twice continuously differentiable, $\sigma(x)=\sigma(-x)>0$ if $|x|< R_1$ and it is zero for $|x|\ge R_1\,.$ Similarly, $\chi(x)=\chi(-x)>0$ if $|x|\le2R_1$ and $\chi(x)=0$ if $|x|>R_2$ with some $R_2>2R_1\,.$ A technical condition, $|\chi'(x)|\le K\chi(x)^{1-\ka}\,,$ where $0<\ka<1/9\,,$ will be exploited in Lemma 2.4. Since $w^\t_{\a,\b}=-w^\t_{\b,\a}\,,$ the condition $F(p_\a,p_\b)=-F(p_\b,p_\a)$ of antisymmetry of $F$ clearly implies the conservation of total momentum. For convenience, we choose $$ F(p_\a,\,p_\b):={1\over 2}\sum_{\t=1}^d\bigl\lan G_\t(p_\a,\,p_\b), D_{\a,\b}\bigr\ran G_\t(p_\a,\,p_\b) \hbox{ and } X^\t_{\a,\b}\p:={1\over \sqrt 2}\bigl\lan G_\t(p_\a,\,p_\b),D_{\a,\b}\p\bigr\ran\,, $$ then the formal generator $\wh L$ of the random component of our process becomes\nfootnote{The future requirements (1.5) and (1.6) imply that ${X^\t_{\a,\b}}^*=-X^\t_{\a,\b}$, where the adjoint is taken with respect to the measure defined by the kinetic energy, (see[LO]).} $$ \wh L\p={1\over 2}\sum_{\t=1}^d\sum_{\a\in I}\sum_{\b\ne\a} \gamma_{\a,\b}(q) X^\t_{\a,\b}\bigl(X^\t_{\a,\b}\p\bigr). \Eq (genoise) $$ In this case the orthogonality relations $$ \bigl\langle G_\t(p_\a,\,p_\b),\phi'(p_\a) -\phi'(p_\b)\bigr\rangle=0 \Eq (orthorel) $$ imply the formal conservation of energy, see {[LO]}. To have conservation of phase volume it is also assumed that $$ \bigl\lan D_{\a,\b},G_\t(p_\a,p_\b)\bigr\ran=0\,, \Eq (divnull) $$ i.e. the operators $\gamma_{\a,\b}X^\t_{\a,\b}X^\t_{\a,\b}$ are symmetric with respect to Lebesgue measure $dp_\a dp_\b\,.$ As a consequence we shall see that the conservation laws imply the reversibility of Gibbs states with respect to $\wh L\,.$ For an explicit example of $F$ and $G_\t$ see the Appendix of [LO]. \vskip 5pt \noindent{\bf Gibbs measures}\vskip 2pt Let $\l=(\l_0,\l_1,\l_2,\l_3,\l_4)$ be a set of real parameters with $\l_4>0$ and $\l^2_1+\l^2_2+\l^2_3<\underline c^2\,,$ and denote by $\Pi$ the distribution of a Poisson process of unit intensity in $\RR^3$ . A probability measure $\PP$ on $\O$ is called a Gibbs state for $\Cal H$ with parameters $\l$ if its conditional distributions given the configuration outside of any cubic box $\L\subset\RR^3$ can be represented as $$ \PP[d\o_\L|\o_{\L^c}]= {1\over Z_\L}\exp\left[\lambda_0|\o_\L| + \sum_{\a=1}^n\sum_{i=1}^3\lambda_i p_\a^i -\lambda_4{\cal H}_{\Lambda}(\omega_\Lambda,\omega_{\Lambda^c})\right] \,\Pi(dq_\L)\,dp_\L\,, $$ where $Z_\L$ is the normalization, and a natural decomposition $\o_\L=(q_\L,p_\L)$ is used, see [D]. The local Hamiltonian, $\Cal H_\L$ is defined as $$ {\cal H}_{\Lambda}(\omega_\Lambda,\omega_{\Lambda^c})= \sum_{q_\a\in\omega_\Lambda}\left[\phi(p_\a)+{1\over 2} \sum_{q_\b\in\omega_\L;\;\alpha\neq\beta}V(q_\a-q_\b) +\sum_{q_\b\in\omega_{\L^c}}V(q_\a-q_\b)\right]\,, $$ the set of such measures will be denoted by $\Cal P_\l\,,$ see {[R]} for the existence of Gibbs states for superstable interactions. \vskip 5pt \noindent{\bf Relative entropy} \vskip 2pt Let $Q$ and $P$ be probability measures on $\Omega$, and for any $\L\subset\RR^3$ denote $\Cal F_\L$ the set of continuous and bounded functions $\psi:\O\mapsto\RR$ such that $\psi(\o)=\psi(\o_\L)$ for all $\o\in\O\,.$ The entropy of $Q$ in $\L\,,$ relative to $P\,,$ is defined by $$ H_\Lambda(Q|P)\ =\ \sup_{\psi\in{\cal F}_\L}\left\{\E^Q(\psi) -\log\E^P(e^\psi)\right\} \Eq (entropy) $$ where $\E^Q$ denotes the expectation with respect to the probability measure $Q\,.$ If $\L=\RR^3$ then the subscript $\L$ of $H_\L$ will be omitted, for properties of $H_\L$ see, for example, [OVY]. As a reference measure a distinguished, translation invariant, Gibbs state $P=\PP$ will be chosen. We say that $Q$ has finite specific entropy if there exists a constant $C$ such that $H_\L(Q|\PP)\le C(1+|\L|)$ for any cubic box $\L\,.$ If $Q$ is translation invariant with finite specific entropy, then the particle density $\rho=\rho(\o)$ is $Q$-a.s. defined as the following limit taken along any increasing sequence of cubic boxes, see [LO], $$ \rho(\o)=\lim_{|\Lambda|\to\infty}|\Lambda|^{-1}|\omega_\Lambda| . $$ Main results of [LO] for the system under considerations can be summarized as follows. \proclaim {\Theorem (old)}. Suppose that $Q$ is a translation invariant probability measure on $\Omega$ with finite specific entropy, and let $\rho_c:=3/(4\pi R^3_1)\,.$ If \item{(i)} $Q[\rho(\omega)>\rho']\;=\; 1$ for some $\rho'>\rho_c\,,$ \item{(ii)} $Q$ is invariant with respect to $\wt L=L+\wh L$ in the sense that, for any smooth local function $\psi$ we have $\E^Q(\wt L\psi)=0\,,$ \item{(iii)} $Q$ is reversible with respect to $\wh L\,,$ i.e. $\E^Q(\psi\wh L\p)\ =\ \E^Q(\p\wh L\psi)\;$ for any pair $\p,\psi$ of smooth local functions, \vskip-\medskipamount \noindent{\sl then Q is a convex combination of Gibbs states.} \medskip \vskip 5pt \noindent{\bf Statement of the result} \vskip 2pt Notice that the theorem above is stated without any reference to the existence of the infinite dynamics, properties (ii) and (iii) of invariance are purely formal. However, the extraction of local information as reversibility is usually based on a method of Liapunov functions, namely entropy and its rate of change are compared, so the first step of our argument is intrinsically related to the evolution. \proclaim {\Theorem (dynamics)}. Under the conditions on the stochastic dynamics listed above, there exists an explicitly defined set $\bar\O\subset\O$ such that $Q(\bar\O)=1$ for each $Q$ with finite specific entropy. Moreover, for each $\o_0\in\bar\O$ we have a unique strong solution $\o(t)\,,\,t\ge0$ to (0.1) such that $\o(0)=\o_0$ and $\o(t)\in\bar\O$ a.s. The solution is a measurable function of the initial configuration, and every Gibbs state $\PP\in\Cal P_\l$ with $\l_4>0$ and $\l_1=\l_2=\l_3=0$ is a stationary measure for the random evolution. This theorem is proven in the next section, solutions are defined by a limiting procedure starting from finite systems. The restriction on the parameters of a Gibbs measure in the last statement could have been removed by elaborating some technical details, but we do not need such a general assertion. Having constructed the infinite evolution we can consider stationary measures instead of simply measures formally invariant as in theorem 1.1 (ii).\nfootnote{Notice that, since the infinite dynamics satisfies equations (0.1), if $Q$ is stationary, then it satisfies (ii) of theorem 1.1.} \proclaim {\Theorem (reversible)}. Every translation invariant stationary measure with finite specific entropy is reversible with respect to the stochastic part $\wh L$ of the generator; that is, condition (iii) in Theorem 1.1 holds. The proof of Theorem \equ (reversible) is the content of Section 3. Combining the above results we get the final result of the paper: \proclaim {\Theorem (main)}. Let $Q$ be a translation invariant stationary measure with finite specific entropy, then condition (i) of Theorem 1.1 implies that $Q$ is a superposition of Gibbs states. \vskip .5cm {\bf 2. INFINITE DYNAMICS} \vskip.25cm \numsec=2\numfor=1\numtheo=1 We start this section by describing the set of allowed initial configurations. Although the definition is a bit technical our choice boils down to configurations for which the energy in a box does not grow too fast with respect to the size of the box. The exact meaning of this construction will become more clear later on when the desired a priori bounds for a family of partial dynamics and the requirements for the existence of a unique limiting dynamics are discussed. \vskip 5pt \noindent{\bf Initial conditions}\vskip 2pt Let $\Cal H_m(\o,r)$ denote the total energy of $\o\in\Omega$ in a ball $B_m(r)\subset \RR^3$ of center $m$ and radius $r\le\infty$, i.e. $\Cal H_m(\o,r) \equiv \Cal H(\o_{B_m(r)})$; the number of points of $q$ in $B_{q_\a}(r)$ will be denoted as $N_\a(q,r)\,.$ For $\kappa\in (0,1/9)\,,$ see \equ(gamma), and $r\ge R_3 = R_1+R_2$, define $$ \bar \Cal H_{\kappa,r}(\o) := \sup_{|m|\le r} {\Cal H_m(\o,R_3)\over 1+|m|^{3+2\kappa}}\,,\qquad \bar\Omega_{\kappa,r} (h) = \{\o\in\Omega \;:\; \ \bar \Cal H_{\kappa,r}(\o)\le h \}. $$ Let $\Cal Q_r(k)$ be the set of Borel probability measures on $\bar\Omega_{\kappa,r} (h)$ such that $Q(\Cal H_m(\o,R_3)) \le k$ for all $|m|\le r$. Now the set of all allowed configurations is defined as $$ \bar\Omega_{\kappa,\infty} = \bigcup_{h>0} \bar\Omega_{\kappa,\infty}(h) = \{ \o\in\Omega\;:\;\bar\Cal H_{\kappa,\infty}(\o) <\infty \}. $$ Remember that the level sets $\bar\Omega_{\kappa,\infty} (h)$ are compact in the weak topology of $\Omega$ and, in view of \equ(superstability), $N_\a(q,R_3)=O(\sqrt h L^{{3/2}+\ka})$ for $q_\a\in B_m(L-R_3)$ and $(q,p)\in\bar\O_{\ka,L}(h)\,.$ We shall see that the initial condition for the existence and uniqueness of the limiting dynamics could have been formulated in terms of $N_\a$ only, but a preservation of bounds on kinetic energy will be needed when we prove locality of the dynamics. \proclaim{\Lemma(initial)}. If $\kappa>0$ then for any fixed $k>0$ we have $$ \lim_{h\to\infty} \inf_{r\ge R_3} \inf_{Q\in \Cal Q_r(k)} Q(\bar\Omega_{\kappa,r} (h) ) = 1 , $$ that is $Q(\bar\O_{\ka,\infty}) =1$ if $Q\in \Cal Q_\infty:= \bigcup_{k>0}\Cal Q_\infty(k)$. \smallskip\noi{\bf Proof:} This statement is a direct consequence of the Markov inequality. In fact we have some universal $v>0$ such that (by $v \ZZ^3$ we denote the tridimensional cubic lattice of size $v$) $$ \eqalign{ Q(\bar\O_{\ka,r}(h)^c)&\le \sum_{m\in B_0(r)\cap v\Z^3} Q[\Cal H_m(\o,\,R_3)>vh(1+|m|^{3+2\ka})\cr &\le \sum_{m\in B_0(r)\cap v\Z^3}{Q(\Cal H_m)\over h(1+|m|^{3+2\ka})}\le {k\over vh}\sum_{m\in v\Z^3}{1\over (1+|m|^{3+2\ka})} } $$ which proves the statement for any $\kappa>0\,.$ $\qed$ \smallskip Observe that the entropy condition $H_\L(Q|\PP)\le C (1+|\L|)$ implies $Q\in \Cal Q_\infty$ via (1.7) and (1.1), see Lemma 3.1 of [LO]. \vskip 5pt \noindent{\bf Local dynamics}\vskip 2pt There are several ways to define a family of partial dynamics, the advantage of the following construction consists in its direct relation to Gibbs states. Let $a:\RR^3\mapsto[0,1]$ be twice continuously differentiable with compact support, we assume also $|a'(x)|\le 1$ for all $x\in\RR^3\,.$ We interpret $a$ as a smooth version of the indicator function of a ball, its concrete shape is not very important. For every such cutoff $a$ and inverse temperature $\l_4>0$ we consider a system of stochastic differential equations, $$ \eqalign{ dq_\a=&-{1\over\l_4}e^{\l_4{\cal H}_\L(\o_\L,\o_{\L^c})} {\partial\over\partial p_\a}\left(a(q_\a)e^{-\l_4 {\cal H}_\L(\o_\L,\o_{\L^c})}\right) dt\cr dp_\a=& {1\over\lambda_4} e^{\lambda_4 {\cal H}_\L(\o_\L,\o_{\L^c})}{\partial\over\partial q_\a}\left(a(q_\a) e^{-\lambda_4 {\cal H}_\L(\o_\L,\o_{\L^c})}\right)\,dt\cr &+a(q_\a)\sum_{\b\ne\a}\gamma_{\a,\b}(q)a(q_\b)F(p_\a,p_\b)\,dt\cr &+\sqrt{a(q_\a)}\sum_{\t=1}^d\sum_{\b\ne\a} \sqrt{a(q_\b)\gamma_{\a,\b}(q)}G_\t(p_\a,p_\b)\,dw_{\a,\b}^\t\,, } \Eq(dynloc)$$ where it is assumed that $\L\subset\RR^3$ is bounded and contains the support of $a$ in its interior; in such a situation the equations above do not depend on the particular choice of $\L\,.$ Notice that in a region where $a=1$ our particles follow the original equations of motion, while they are frozen outside of the support of $a\,,$ i.e. $\dot q_\a=\dot p_\a=0\,.$ Particles approaching the boundary of the support of $a$ slow down, thus we have a smooth transition between moving and frozen particles, see [F1] for a similar construction. This means that we essentially have a finite dimensional diffusion, let $P^t_{\lambda_4,a}$ denote the Markov semigroup induced by partial dynamics (2.1), i.e. $P^t_{\l_4,a}\psi(\o):=\E_w(\psi(\o(t))\,,$ where $\o(t)$ is the solution with initial condition $\o(0)=\o\,,$ $\psi:\O\mapsto\RR$ is continuous and bounded, while $\E_w$ denotes the expectation with respect to the joint distribution of our Wiener processes. By a direct application of the Ito lemma we see that the (formal) generator of $P^t_{\l_4,a}$ decomposes as $\wt L_{\l_4,a}=L_{\l_4,a}+\wh L_{a}\,,$ where $$\eqalign{ L_{\l_4,a}\psi&=-{1\over\l_4}\sum_{\a\in I} e^{\l_4{\cal H}_\L(\o_\L,\o_{\L^c})}{\partial\over\partial p_\a} \left(a(q_\a)e^{-\l_4 {\cal H}_\L(\o_\L,\o_{\Lambda^c})}\right) {\partial\psi\over\partial q_\a}\cr &+{1\over\l_4}\sum_{\a\in I} e^{\lambda_4{\cal H}_\L (\o_\L,\o_{\L^c})}{\partial\over\partial q_\a} \left(a(q_\a)e^{-\l_4 {\cal H}_\L(\o_\L,\o_{\L^c})}\right) {\partial\psi\over\partial p_\a} ,\cr \wh L_{a}\psi &={1\over 2}\sum_{\t=1}^d\sum_{\a\in I}\sum_{\b\ne\a} \gamma_{\a,\b}(q)a(q_\a)a(q_\b)X^\t_{\a,\b}(X^\t_{\a,\b}\psi). } \Eq(pargen)$$ Since the coefficients of \equ(dynloc) are bounded smooth functions, we have a differentiable dependence of solutions on initial values. Therefore a class $\Cal D_a$ of twice continuously differentiable functions forms a common core of $L_{\l_4,a}$ and $\wh L_a\,,$ e.g. in the space of continuous and bounded functions. An extension to the space $L^2(\PP_\l)$ of square integrable functions with respect to a distinguished Gibbs state $\PP_\l$ follows by the next Lemma. \proclaim{\Lemma(pargibbs)}. Let $\l_1=\l_2=\l_3=0$ while $\l_4>0\,,$ then every Gibbs state $\PP\in\Cal P_\l$ satisfies $$ \E^\PP(\psi_1 L_{\l_4,a}\psi_2)=-\E^\PP(\psi_2 L_{\l_4,a}\psi_1)\hbox{ and } \E^\PP(\psi_1 \wh L_{a}\psi_2)=\E^\PP(\psi_2 \wh L_{a}\psi_1) $$ for $\psi_1,\psi_2\in\Cal D_a\,,$ consequently $\PP$ is a stationary measure of the process $P_{\l_4,a}^t$ for each cutoff $a$. \smallskip\noi{\bf Proof:} Both symmetry relations follow from the definition of $\PP$ by integrating by parts. The property of reversibility is a direct consequence of (1.6). Integration by parts with respect to positions is possible because of the presence of the cutoff $a\,.$ $\qed$\smallskip Since (2.1) violates the law of momentum conservation in regions where $a$ is not a constant, Lemma 2.2 is not true for general Gibbs measures. \vskip 5pt \noindent{\bf Construction of the infinite dynamics}\vskip 2pt First we derive an a priori bound for local dynamics; we show that the initial condition is preserved for all $t>0$ and the related bound does not depend on the particular choice of the cutoff function $a\,.$ \proclaim{\Lemma(apriori)}. There exists a constant $c_1$ depending only on $\l_4$ and on the parameters of the infinite system (0.1) such that $$ \E_w\left(\Cal H_m(\o_a(t),R_3)\right)\le (c_1+c_1t)(1+\Cal H_m(\o_a(0),R_3+\bar ct) $$ for all $m,t$ and $a\,,$ where $\o_a(t)$ is any solution to (2.1). \smallskip\noi{\bf Proof:} Since all velocities are bounded by $\bar c\,,$ we have $|q_\a(t)-q_\a(0)|\le \bar ct\,,$ whence $N_\a(q(t),r)\le N_\a(q(0),r+\bar ct)\,,$ which yields an explicit deterministic bound for the potential energy via superstability (1.1). On the other hand, $$ |b_\a(\o)|+\sum_{\t=1}^d\sum_{\b\ne\a}|\sigma^\t_{\a,\b}(\o)|^2 \le c_1' N_{\a}(q,R_1)\,, $$ and the same bound holds true for the corresponding coefficients of (2.1), from the stochastic equations by the Schwarz inequality we get $$ \E_w\left(|p_\a(t)-p_\a(0)|^2\right)\le c_1''t(1+t) N_{\a}(q(0),R_1+\bar ct)^2\,, \Eq(papr)$$ which completes the proof. Indeed, as $\phi(y)\le\phi(0)+\bar c|y|\,,$ taking the square root of both sides and summing for $\a\in I$ such that $|q_\a(0)-m|\le R_3+\bar ct\,,$ we get a bound for $\Cal H_m(\o(t),R_3)\,;$ the square of the number of points at time zero is estimated again by superstability. $\qed$\smallskip To prove the existence of limiting solutions when the cutoff is removed we have to compare different partial solutions. Let $A_L$ denote the set of twice continuously differentiable $a:\RR^3\mapsto [0,1]$ with compact support such that $|a'(x)|\le1$ for all $x$ and $a(x)=1$ if $|x|\le L\,.$ For $a,\bar a\in A_L$ let $\o(t)=(p(t),q(t))$ and $\bar\o(t)=(\bar q(t),\bar p(t))$ denote the corresponding solutions to \equ(dynloc) with a common initial value $\o(0)=\bar\o(0)=(\xi,\eta)\,.$ Supposing $|x|\le L-2\bar ct$, for $|x-\xi_\a|\le R_0+2\bar ct$ we get $\dd_t|q_\a-\bar q_\a|\le K_0|p_\a-\bar p_\a|\,,$ whence $$ |q_\a(t)-\bar q_\a(t)|^2\le 2K_0\int_0^t |q_\a(s)-\bar q_\a(s)||p_\a(s)-\bar p_\a(s)|\,ds\,; \Eq(qb)$$ here and in what follows, $K_0,K_1,K'_1...$ denote constants depending only on the coefficients of the infinite system. The case of the momentum variables is more complex. If $a=\bar a=1$ can be assumed as before, then by Ito's formula we get $$ \E_w\left(|p_\a(t)-\bar p_\a(t)|^2\right)=\E_w\left(\int_0^t\bigl(J_{\a,1}(s)+J_{\a,2}(s) +J_{\a,3}(s)\bigr)\,ds\right)\,, $$ where $$ \eqalign{ J_{\a,1}&=-2\sum_{\a\ne\b}\bigl\lan p_\a-\bar p_\a, V'(q_\a-q_\b)-V'(\bar q_\a-\bar q_\b)\bigr\ran\,,\cr J_{\a,2}&=2\sum_{\b\ne \a}\bigl\lan p_\a-\bar p_\a, \gamma_{\a,\b}(q)F(p_\a,p_\b)-\gamma_{\a,\b}(\bar q) F(\bar p_\a,\bar p_\b)\bigr\ran\cr J_{\a,3}&=\sum_{\t=1}^d\sum_{\b\ne \a} \bigl|\sqrt{\gamma_{\a,\b}(q)}G_\t(p_\a,p_\b) -\sqrt{\gamma_{\a\b}(\bar q)}G_\theta(\bar p_\a,\bar p_\b)\bigr|^2 . } $$ Introduce now $\Ga_i$ for $i=1,2$ and $r,t\ge0$ by $$ \Ga_1(\xi,\eta,a,\bar a;r,t):=\sum_{\a:|\xi_\a|\le r} |q_\a(t)-\bar q_a(t)|^2\,; $$ in the definition of $\Ga_2$ the variables $q_\a$ and $\bar q_a$ should be replaced by $p_\a$ and $\bar p_\a\,,$ respectively. Our main tool is the following: \proclaim{\Lemma(converg)}. Suppose that $\ka<1/9\,,$ $2\bar cT0$ and $i=1,2$ we have $$ \lim_{L\to\infty}\sup_{a,\bar a\in A_L}\;\sup_{(\xi,\eta)\in\bar \O_{\ka,L}(h)}\;\E_w\left(\Ga_i(\xi,\eta,a,\bar a;r,t)\right)=0\,, $$ and the convergence is uniform on the time interval $[0,\,T]$. \smallskip\noi{\bf Proof:} The idea of the proof is to define and to estimate a ``distance" (based on $\Gamma_i$) among different partial dynamics in boxes of radius $r0$ implies $|\ti q^{\a,\b}_{\a} -\ti q^{\a,\b}_{\b}|\le R_1+2\bar c t\le 2R_1\,,$ i.e. $N_\a(q,R_1) \le N_\a(\ti q^{\a,\b},2R_1)\,.$ This means that $N_\a(q,R_1)\le 1/\Te_{\a,\b}(\ti q^{\a,\b})\,,$ consequently $$ \eqalign{ J_{\a,2}(t)&\le K_2'\bar N^\ka_{L,T}\ti J_{\a,1}(t) +K'_2\ti J_{\a,2}(t)\,;\cr \ti J_{\a,2}(t)&:=\sum_{\b\ne\a}\ga_{\a,\b}(q) \bigl(|p_\a-\bar p_\a|^2+|p_\b-\bar p_\b|^2\bigr). } \Eq(pb2) $$ In a similar way we get $$ \eqalign{ J_{\a,3}(t)&\le K_3\ti J_{\a,2}(t)+K_3\ti J_{\a,3}(t)\cr &+K_3\sum_{\b\ne\a}{\si(q_\a-q_\b)\over\Te_{\a,\b}(\ti q^{\a,\b})} \Bigl(\sum_{\de\in I}|\dd_\de \Te_{\a,\b}(\ti q^{\a,\b})| |q_\de-\bar q_\de|\Bigr)^2\,;\cr &\ti J_{\a,3}(t):=\sum_{\de\in I}\ti\ga_{\a,\b}(t) \bigl(|q_\a-\bar q_\a|^2+|p_\de-\bar p_\de|^2\bigr) } $$ whence by the Cauchy inequality and \equ(dga) $$ J_{\a,3}(t)\le K'_3\ti J_{\a,2}+K'_3\bar N^{2\ka}_{L,T}\ti J_{\a,3}(t) . \Eq(pb3) $$ Introduce now $d(r,t):=\E_w\left(\Ga_2(\xi,\eta,a,\bar a;r,t)\right) +\bar N_{L,T}\E_w\left(\Ga_1(\xi,\eta,a,\bar a;r,t)\right)$ for $t0$ depending only on $R_3$ and $T$, while $\bar N_{L,T}= O(\sqrt hL^{3/2+\ka})\,.$ Using $|q_\a(t)-\xi_\a|\le \bar ct$ and the second a priori bound \equ(papr), we see that the right hand side of \equ(expl) vanishes as $L\to+\infty$ because $\ell!=O\bigl((\ell/e)^\ell\bigr)$ and $({1/2}+\ka)({3/2}+\ka)<1$ by hypothesis. $\qed$\smallskip Now we are in a position to prove the existence and uniqueness of limiting solutions to (0.1). Let us consider a sequence of partial solutions $\o_n=\o_n(t)\,,\,n\in\NN$ of \equ(dynloc) with a common initial value $\o_n(0)=(\xi,\eta)\in\bar\O_{\ka,\infty}\,;$ the corresponding cutoff $a_n:\RR^3\mapsto\RR$ is assumed to be a decreasing smooth function of $|x|$ such that $a_n(x)=1$ if $|x|\le n$ and $a_n(x)=0$ if $|x|>n+1\,.$ In view of Lemma 2.4 $\o_n$ converges in probability to some limit $\o(t)$ for each $t0\,.$ In view of Lemma 2.1 and Lemma 2.3 we know that $QP^t\in\Cal Q_{\infty}\,,$ too. While $P^t_n$ has fairly good regularity properties, semigroup theory does not apply directly to the limiting case. Nevertheless, all we need in the next section is summarized as follows. \proclaim{\Lemma(locality)}. Suppose that $\psi:\bar\O_{\ka,\infty} \mapsto\RR$ is a continuous and bounded local function, i.e. $\psi(\o) \equiv\psi(\o_{B_0(r)})$ for some $r>0\,,$ then $$ \lim_{\ell\to\infty}\sup_{n>\ell+r}\sup_{Q\in\Cal Q_n(k)} |QP^t_n\psi-QP^t\psi|=0 $$ for all $t,k>0\,,$ and the convergence is uniform on compact time intervals. \smallskip\noi{\bf Proof:} The a priori bound of Lemma 2.3 extends immediately to the limiting dynamics, thus for any $\e,T>0$ we have some $\bar k>k$ and $h>\bar k$ such that $Q(\bar\O_{\ka,r}(h))\ge1-\e\,,$ $QP^t_n(\bar\O_{\ka,r}(h))\ge1-\e\,,$ and $QP^t(\bar\O_{\ka,r}(h))\ge1-\e$ whenever $tr+R_3+2\bar cT$ and $Q\in\Cal Q_n(k)\,.$ Since $\bar\O_{\ka,r}(\bar k)$ is compact, there exists also an $\e'>0$ such that $|q_\a-\bar q_\a|+|p_\a-\bar p_\a|<\e'$, for all $\a\in I$ with $|q_\a|,|\bar q_\a|\le r$, implies $|\psi(\o)-\psi(\bar\o)|\le\e$ for $\o,\bar \o\in\bar\O_{\ka,r}(\bar k)\,.$ Therefore the statement follows from Lemma 2.4 and Chebishev inequality by a $3\e$ argument. $\qed$\smallskip The final statement of Theorem \equ (dynamics) on stationarity of certain Gibbs states is now a direct consequence of Lemma 2.2. \vskip .5cm {\bf 3. AN ENTROPY ARGUMENT}\vskip.25cm \numsec=3\numfor=1\numtheo=1 In this section we extend a familiar argument by Holley [H] to the present more complex situation. For a probability measure $Q$ on $\O$, let $H(Q|\PP_\lambda)$ denote the entropy relative to a distinguished Gibbs state $\PP_\lambda$ with $\l_1=\l_2=\l_3=0$, as defined by \equ(entropy) with $\Lambda = \RR^3$. The family of partial dynamics \equ(dynloc) has been chosen such that $\PP_\l$ is a common stationary measure of each local dynamics $P_n^t=P_{\l_4,a_n}$ introduced in Section 2. Therefore $P_n^t$ is a strongly continuous semigroup in $L^2(\PP_\lambda)\,,$ and smooth cylinder functions form a core for its generator $\wt L_n = L_n +\wh L_n\,,$ see \equ(pargen). Remember that $L_n\equiv L_{\l_4,a_n}$, the Hamiltonian part, is antisymmetric in $L^2(\PP_\l)$ while the symmetric (reversible) component is just $\wh L_n\equiv \wh L_{a_n}\,.$ If $\G$ is a generator in $L^2(\PP_\l)$ then the corresponding Donsker-Varadhan rate function is defined as $$ D(Q|\G) \ =\ \sup_\psi\Bigl\{ - \int {\G \psi\over \psi} dQ \;:\; \psi\in \hbox{Dom }\G,\,\inf \psi > 0\Bigr\} . $$ If $\G$ is self-adjoint and $\G<0$, then we can apply the following result due to Donsker and Varadhan (cf. [DV], Theorem 5) \proclaim{\Theorem (donsker-var)}. $D (Q|\G )<+\infty$ if and only if $Q<<\PP_\l$ and $g:=\sqrt{dQ/d\PP_\l}\in\hbox{Dom }\sqrt{-\G}$; moreover $$ D(Q|\G) \ =\ \int \left(\sqrt{-\G} g\right)^2 d\PP_\l . \Eq(dirichlet) $$ Our main tool consists of the following entropy inequality. \proclaim{\PProposition (H-th)}. Let $\bar Q_n^t := (1/t) \int_0^t QP_n^s ds$. If $H(Q|\PP_\l) <\infty$ then $$ H(QP_n^t|\PP_\l) + 2t D(\bar Q_n^t|\hat L_n) \le H(Q|\PP_\l) . \Eq(H-ineq) $$ \smallskip\noi{\bf Proof:} Let $P^{*t}_n$ be the adjoint semigroup with respect to $\PP_\l$, which is again a diffusion with formal generator $\tilde L_n^* = -L_n+\hat L_n$. Both forward and backward diffusion are essentially finite dimensional with smooth coefficients, thus twice continuously differentiable functions form a common core $\Cal D_n$ of $L_n$ and $L^*_n\,.$ This suffice to justify the following computations. Observe first that, as an easy consequence of Jensen's inequality, we have $$ H(Q'P_n^{\tau}|Q''P_n^{\tau}) \le H(Q'|Q'') \Eq (H-increase) $$ for any two measures $Q',\,Q''$. For any strictly positive $\psi\in\Cal D_n$ with $\PP_\l(\psi)=1$ define $Q''$ by $dQ''=\psi d\PP_\l\,.$ Since $$ {dQ''P_n^{\tau}\over d\PP_\l} = P^{*{\tau}}_n \psi $$ we have $$ H(Q'|Q'') = H(Q'|\PP_\l) - Q'(\log \psi) $$ and $$ H( Q' P_n^{\tau} | Q''P_n^{\tau}) = H(Q'P_n^{\tau}|\PP_\l) - Q'P_n^{\tau}(\log P^{*{\tau}}_n\psi). $$ Accordingly, by \equ(H-increase), $$ H(Q'|\PP_\l) - H(Q'P_n^{\tau}|\PP_\l) \ge Q'(\log \psi) - Q'P_n^{\tau}(\log P^{*{\tau}}_n\psi), $$ whence, by the concavity of the logarithm and the inequality $\log(x+1) \le x$, $$ H(Q'|\PP_\l) - H(Q'P_n^{\tau}|\PP_\l) \ge Q'(\log \psi) - Q'(\log P_n^{\tau} P^{*{\tau}}_n\psi)\ge \int {\psi - P_n^{\tau} P^{*{\tau}}_n\psi\over \psi} dQ' . $$ Remembering that $P_n^t$ and $P^{*t}_n$ are both Feller semigroup and $\psi$ belongs to the common core of $\tilde L_n$ and $\tilde L_n^*$, we have, for small $\tau$, $$ \psi - P_n^{\tau} P^{*\tau}_n\psi = \psi - P_n^{\tau} \psi+ \psi - P^{*\tau}_n\psi + P_n^{\tau}\left(\psi - P^{*\tau}_n\psi\right)- \left(\psi - P^{*\tau}_n\psi\right) = -2\tau \hat L_n \psi+o(\tau). $$ Therefore, by dividing the given interval $[0,t]$ into $m$ small pieces, with $\tau=t/m$ and $Q'=QP_n^{{it\over m}}$, we get $$ \eqalign{H(Q|\PP_\l) - H(QP_n^t|\PP_\l)&= \lim_{m\to\infty} {1\over m}\sum_{i=0}^{m-1}\left[H(QP_n^{{i\over m}t}|\PP_\l) - H(QP_n^{{i+1\over m}t}|\PP_\l)\right]\cr &\ge -2\int_0^t ds \int {\hat L_n \psi \over \psi}\ dQP_n^s .} $$ By taking the supremum over all $\psi$ considered we conclude the proof. $\qed$\smallskip Observe now that if $D (Q|\hat L_n)< \infty$, then by theorem \equ(donsker-var) it can be written as a sum, namely, if $g = \sqrt{dQ/d\PP_\l}$, \nfootnote{To see this, since $g\in \hbox{Dom }\sqrt{-\hat L_n}$ (hence $g\in \hbox{Dom }\sqrt{a(q_\a)a(q_\b)\gamma_{\a\b}(q)}X_{\a\b})$), one can approximate it by local smooth functions, then use the closability of the Dirichlet form $D$.} $$ D (Q|\hat L_n ) = {1\over 2} \int \sum_{\t, \a,\b} a_n(q_\a) a_n(q_\b) \gamma_{\a\b}(q) (X^\t_{\a\b} g)^2 \ d\PP_\l . $$ Let $a_{n,1}(x), a_{n,2}(x),\dots, a_{n,j}(x)$ be smooth non-negative functions with compact support, and assume that their supports are disjoint. Furthermore, assume that $a_n(x) \ge a_{n,1}(x) + \dots + a_{n,j}(x)$, then $$ D (Q|\hat L_n ) \ge D (Q|\hat L_{a_{n,1}} ) + \dots + D (Q|\hat L_{a_{n,j}} ). $$ Therefore, from \equ(H-ineq), we have $$ H(QP_n^t|\PP_\l) + 2t \sum_{i=1}^jD(\bar Q_n^t|\hat L_{a_{n,i}}) \le H(Q|\PP_\l). $$ Thus we can choose strictly positive and smooth functions $\psi_0, \psi_1,\dots,\psi_j$ such that $$ QP_n^t (\psi_0) - \log \PP_\l(e^{\psi_0}) - 2t \sum_{i=1}^j \bar Q_n^t \left({\hat L_{a_{n,i}}\psi_i\over \psi_i}\right) \le H(Q|\PP_\l) . $$ This inequality extends by continuity to the infinite dynamics (cf. lemma \equ(locality) and note that $Q\in \Cal Q_\infty$) $$ QP^t (\psi_0) - \log \PP_\l(e^{\psi_0}) - 2t \sum_{i=1}^j \bar Q^t \left({\hat L_{a_{n,i}}\psi_i\over \psi_i}\right) \le H(Q|\PP_\l) . \Eq(ready) $$ Now we are in a position to take the thermodynamic limit and conclude the main result of this section. \proclaim{\PProposition (loc-sta)}. If $Q_*$ is a translation invariant stationary measure of the infinite system \equ(stochdyn), and $Q_*$ has finite specific entropy with respect to $\PP_\l$, then $D( Q_*|\wh L_{\bar a}) = 0$ for all smooth functions $\bar a\le 1$ of compact support. \smallskip\noi{\bf Proof}: We are going use \equ(ready) with $Q=Q_{*n}$ where $Q_{*n}$ is defined by $$ Q_{*n}(\psi)\ =\ \int \PP_\l(\psi|\Cal F_{\L_n})\ dQ_* $$ and $\L_n$ denotes the centered cubic box of size $n$. Of course, $H(Q_{*n}|\PP_\l) = H_{\L_n}(Q_*|\PP_\l)\,,$ thus $$ \bar H(Q_*|\PP_\l):= \lim_{n\to \infty}{H(Q_{*n}|\PP_\l)\over |\L_n|} = \sup_{\psi}\left( Q_*(\psi) - \bar F(\psi)\right)\ , \Eq(H-density) $$ where $\psi$ are the local, bounded and continuous functions; in addition, $$ \bar F(\psi) := \lim_{n\to \infty} {1\over |\L_n|}\log \int \exp \left(\sum_{k\in \L_n\cap \Z^3} \vec s^{k} \psi\right) d\PP_\l , \Eq(free) $$ and $\vec s^{k}$ denotes the shift in $\RR^3$ by $k \in \RR^3$, i.e. $\vec s^{k}\psi(p,q) = \psi( p,\vec s^{k}q)$ and $\vec s^{k}q = \{q_\a +k\}$ if $q = \{q_\a\}$. The proof of the existence of \equ(H-density) and \equ(free) can be found in [OVY]. Now we set $$ \psi_0 = \sum_{k\in \L_n\cap \Z^3}\vec s^{k} \psi $$ for some local bounded continuous function $\psi$. Without loss of generality we can suppose $n$ so large that $\L_n$ contains the support of $\bar a$, and define $a_{n,i}(x) = \bar a (x+k_i)$, $k_i \in J_n$, and $J_n$ is a discrete subset of $\L_n$ such that the $a_{n,i}$ have the disjoint supports contained in $\L_n$, and ${|J_n|\over n^3 }\ge \bar J_0$, for some fixed constant $\bar J_0$.\nfootnote{This can be done as to ensure that $a_n\geq\sum_{i=1}^ja_{n,i}$.} Correspondingly we choose $\psi_i =\vec s^{k_i} \bar\psi,\ \vec k_i\in J_n$, for a local bounded continuous function $\bar\psi$. Substituting in \equ(ready) and dividing by $|\L_n|$, it remains to prove that $$ \lim_{n\to \infty} {1\over |\L_n|} \sum_{k\in \L_n\cap \Z^3} Q_{*n} P^t(\vec s^{k} \psi) = Q_*(\psi) \Eq(unif1) $$ and $$ \lim_{n\to \infty} {1\over |J_n|} \sum_{k_i\in J_n} \bar Q_{*n}^{t} \left(\vec s^{k_i}{\wh L_{\bar a}\bar\psi \over\bar \psi}\right) = Q_* \left({\wh L_{\bar a}\bar\psi\over \bar\psi} \right) . \Eq(unif2) $$ Indeed, then \equ(ready), \equ(H-density), \equ(free), \equ(unif1) and \equ(unif2) imply $$ Q_* (\psi) - \bar F(\psi) - 2t J_0 Q_* \left({\hat L_{\bar a}\bar\psi\over\bar\psi}\right) \le \bar H(Q_*|\PP_\l), $$ and taking the supremum over all $\psi$ and $\bar\psi$ considered we obtain the wanted result. To prove \equ(unif1), observe first that the rate of convergence in Lemma 2.5 depends only on the magnitude and the modulus of continuity of the underlying function. In the present situation all functions are translates of each other, thus the convergence is uniform on such functions. Therefore, for $k\in\L_{n-\sqrt n}$ we approximate $P^t$ with the local dynamics $\vec s^kP^t_{\sqrt n}$ in the ball $B_k(\sqrt n)\,,$ otherwise we use simply the uniform bound of $\psi\,.$ The proof of \equ(unif2) is similar.$\qed$ As it is well known, see [DV], $D(Q_*|\wh L_{\bar a})=0$ implies the reversibility of $Q_*$ with respect to $\wh L_{\bar a}\,,$ which completes the proof of Theorem 1.3, whereby proving Theorem 1.4 as well, by a direct argument. \vskip .5cm {\bf REFERENCES.}\vskip.25cm \frenchspacing \item{[BLPS]} L. Bunimovich, C. Liverani, A. Pellegrinotti, Y. Suhov: Ergodic Systems of n Balls in a Billiard Table. Commun. Math. Phys. {\bf 146} (1992), 357--396. \item{[D]} R.L. Dobrushin: Gibbsian random fields for particles without hard core. (in Russian) Teor. Mat. Fiz. {\bf 4} (1969), 101--118. \item{[DL]} V. Donnay, C. Liverani: Potential on the Two-Torus for which the Hamiltonian Flow is ergodic. Commun. Math. Phys. {\bf 135} (1991), 267--302. \item{[DV]} M.D. Donsker, S.R.S. Varadhan: Asymptotic Evaluation of Certain Markov Process Expectations for Large Time. I, Commun. Math. Phys. {\bf 28} (1975), 1--47. \item{[FD]} J. Fritz, R.L. Dobrushin: Non-equilibrium dynamics of two-dimensional infinite particle systems with a singular interaction. Commun. Math. Phys. {\bf 57} (1977), 67--81. \item{[FFL]} J. Fritz, T. Funaki, J.L. Lebowitz: Stationary States of Random Hamiltonian Systems. Probab. Theory Related Fields {\bf 99} (1994), 211--236. \item{[F1]} J. Fritz: Gradient dynamics of infinite point systems. Ann. Prob. {bf 15} (1987), 478--514. \item{[F2]} J. Fritz: Stationary States of Hamiltonian Systems with Noise. In: On Three Levels, 203--214, M. Fannes, Ch. Maes, A. Verbeure Eds, Plenum, New York, 1994. \item{[H]} R. Holley: Free energy in a Markovian model of a lattice spin system. Commun. Math. Phys. {\bf 23} (1971), 87-99. \item{[KSS]} A. Kr\'amli, N. Simanyi, D. Sz\'asz: The K property for Four Billiard Balls. Commun. Math. Phys. {\bf 144} (1992), 107--148. \item{[LO]} C. Liverani, S. Olla: Ergodicity in Infinite Hamiltonian Systems with Conservative Noise. Probab. Theory Related Fields, {\bf 106}, 3, (1996), 401--445. \item{[LW]} C. Liverani, M. Wojtkowski: Ergodicity in Hamiltonian Systems, Dynamics Reported {\bf 4} (1995), 130--202. \item{[OVY]} S. Olla, S.R.S. Varadhan, H. T. Yau : Hydrodynamics Limit for a Hamiltonian System with Weak Noise. Commun. Math. Phys. {\bf 155} (1993), 523--560. \item{[R]} D. Ruelle: Superstable interactions in classical statistical mechanics, Commun. Math. Phys. {\bf 18} (1970), 127-159. \item{[S]} R. Siegmund-Schultze: On nonequilibrium dynamics of multidimensional infinite particle systems. Commun. Math. Phys. {\bf 100} (1985), 245-265. \bye