BODY %%%%%%%%%%%%%%% FORMATO \magnification=\magstep1\hoffset=0.cm \voffset=1truecm\hsize=16.5truecm \vsize=21.truecm \baselineskip=14pt plus0.1pt minus0.1pt \parindent=12pt \lineskip=4pt\lineskiplimit=0.1pt \parskip=0.1pt plus1pt \def\ds{\displaystyle}\def\st{\scriptstyle}\def\sst{\scriptscriptstyle} \font\seven=cmr7 %%%%%%%%%%%%%%%% GRECO \let\a=\alpha \let\b=\beta \let\c=\chi \let\d=\delta \let\e=\varepsilon \let\f=\varphi \let\g=\gamma \let\h=\eta \let\k=\kappa \let\l=\lambda \let\m=\mu \let\n=\nu \let\o=\omega \let\p=\pi \let\ph=\varphi \let\r=\rho \let\s=\sigma \let\t=\tau \let\th=\vartheta \let\y=\upsilon \let\x=\xi \let\z=\zeta \let\D=\Delta \let\F=\Phi \let\G=\Gamma \let\L=\Lambda \let\Th=\Theta \let\O=\Omega \let\P=\Pi \let\Ps=\Psi \let\Si=\Sigma \let\X=\Xi \let\Y=\Upsilon %%%%%%%%%%%%%%% DEFINIZIONI LOCALI \let\ciao=\bye \def\fiat{{}} \def\pagina{{\vfill\eject}} \def\\{\noindent} \def\bra#1{{\langle#1|}} \def\ket#1{{|#1\rangle}} \def\media#1{{\langle#1\rangle}} \def\ie{\hbox{\it i.e.\ }} \let\ig=\int \let\io=\infty \let\i=\infty \let\dpr=\partial \def\V#1{\vec#1} \def\Dp{\V\dpr} \def\tende#1{\vtop{\ialign{##\crcr\rightarrowfill\crcr \noalign{\kern-1pt\nointerlineskip} \hskip3.pt${\scriptstyle #1}$\hskip3.pt\crcr}}} \def\otto{{\kern-1.truept\leftarrow\kern-5.truept\to\kern-1.truept}} \def\LS{Logarithmic Sobolev Inequality } \def\LSC{Logarithmic Sobolev Constant } \def\Z{{\bf Z^d}} \def\supnorm#1{\vert#1\vert_\infty} \def\grad#1#2{(\nabla_{\L_{#1}}#2)^2} \def\log#1{#1^2log(#1)} \def\logg#1{#1log((#1)^{1\over2})} %%%%%%%%%%%%%%%%%%%%% Numerazione pagine \def\data{\number\day/\ifcase\month\or gennaio \or febbraio \or marzo \or aprile \or maggio \or giugno \or luglio \or agosto \or settembre \or ottobre \or novembre \or dicembre \fi/\number\year} %%\newcount\tempo %%\tempo=\number\time\divide\tempo by 60} \setbox200\hbox{$\scriptscriptstyle \data $} \newcount\pgn \pgn=1 \def\foglio{\number\numsec:\number\pgn \global\advance\pgn by 1} \def\foglioa{A\number\numsec:\number\pgn \global\advance\pgn by 1} %\footline={\rlap{\hbox{\copy200}\ $\st[\number\pageno]$}\hss\tenrm %\foglio\hss} %\footline={\rlap{\hbox{\copy200}\ $\st[\number\pageno]$}\hss\tenrm %\foglioa\hss} % %%%%%%%%%%%%%%%%% 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 dei ... %%% si puo' scrivere qualsiasi commento; %%% per assegnare un nome simbolico ad una figura, basta scrivere %%% \geq(...); per avere i nomi %%% simbolici segnati a sinistra delle formule e delle figure 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); all'inizio del lavoro %%% bisogna porre \numfig=1 (il numero delle figure non contiene la sezione. %%% Si possono citare formule o figure 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 o figure 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\numfig \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 \equ(#1) == #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 \equ(#1) == #1 }} \def\getichetta(#1){Fig. \verafigura \SIA e,#1,{\verafigura} \global\advance\numfig by 1 \write15{\string\FU (#1){\equ(#1)}} \write16{ Fig. \equ(#1) ha simbolo #1 }} \newdimen\gwidth \def\BOZZA{ \def\alato(##1){ {\vtop to \profonditastruttura{\baselineskip \profonditastruttura\vss \rlap{\kern-\hsize\kern-1.2truecm{$\scriptstyle##1$}}}}} \def\galato(##1){ \gwidth=\hsize \divide\gwidth by 2 {\vtop to \profonditastruttura{\baselineskip \profonditastruttura\vss \rlap{\kern-\gwidth\kern-1.2truecm{$\scriptstyle##1$}}}}} } \def\alato(#1){} \def\galato(#1){} \def\veroparagrafo{\number\numsec}\def\veraformula{\number\numfor} \def\verafigura{\number\numfig} %\def\geq(#1){\getichetta(#1)\galato(#1)} \def\Eq(#1){\eqno{\etichetta(#1)\alato(#1)}} \def\eq(#1){\etichetta(#1)\alato(#1)} \def\Eqa(#1){\eqno{\etichettaa(#1)\alato(#1)}} \def\eqa(#1){\etichettaa(#1)\alato(#1)} \def\eqv(#1){\senondefinito{fu#1}$\clubsuit$#1\else\csname fu#1\endcsname\fi} \def\equ(#1){\senondefinito{e#1}eqv(#1)\else\csname e#1\endcsname\fi} \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 \footline={\rlap{\hbox{\copy200}\ $\st[\number\pageno]$}\hss\tenrm \foglio\hss} %\endinput%\input formato.tex \numsec=0\numfor=1 \tolerance=10000 \font\ttlfnt=cmcsc10 scaled 1200 %small caps \font\bit=cmbxti10 %bold italic text mode % Author. Initials then last name in upper and lower case % Point after initials % \def\author#1 {\vskip 18pt\tolerance=10000 \noindent\centerline{\bit#1}\vskip 0.7cm} % % Address % \def\address#1 {\vskip 4pt\tolerance=10000 \noindent #1} % % Abstract % \def\abstract#1 {\vskip 1cm \noindent{\bf Abstract.\ }#1\par} % \vskip 1cm \centerline{\ttlfnt Approach to Equilibrium of Glauber Dynamics } \centerline{\ttlfnt In the One Phase Region. II: The General Case}\vskip 1cm \author{F. Martinelli\ddag\hskip 0.5cm and\hskip 0.5cm E. Olivieri\dag} \address{\ddag Dipartimento di Matematica Universit\`a "La Sapienza" Roma, Italy \hfill\break{\rm e-mail: martin@mercurio.dm.unirm1.it}} \address{\dag Dipartimento di Matematica Universit\`a "Tor Vergata" Roma, Italy \hfill\break{\rm e-mail: olivieri@mat.utovrm.it}} \abstract{We develop a new method, based on renormalization group ideas (block decimation procedure), to prove, under an assumption of strong mixing in a finite cube $\L_o$, a \LS for the Gibbs state of a discrete spin system. As a consequence we derive the hypercontractivity of the Markov semigroup of the associated Glauber dynamics and the exponential convergence to equilibrium in the uniform norm in all volumes $\L$ "multiples" of the cube $\L_o$.} \vskip 1cm\noindent Work partially supported by grant SC1-CT91-0695 of the Commission of European Communities \pagina %\input formato.tex \numsec=1\numfor=1 \tolerance=10000 \vskip 1cm {\bf Section 1.}\par\noindent \centerline{\bf Preliminaries, Definitions and Results} \bigskip In this paper we analyze the problem of the approach to equilibrium for a general, not necessarily ferromagnetic, Glauber dynamics, i.e. a single spin flip stochastic dynamics reversible with respect to the Gibbs measure of a classical discrete spin system with finite range, translation invariant interaction. We prove that, if the Gibbs measure satisfies a {\it Strong Mixing Condition} on a large enough finite cube $\L_o$, then the Glauber dynamics reaches the equilibrium exponentially fast in time in the {\it uniform norm}, in any finite or infinite volume $\L$, provided that $\L$ is a "multiple" of the basic cube $\L_o$. Such a result has already been proved in our previous papers [MO1], [MO2] in the so called "attractive case" by ad hoc methods. Here we prove the result in greater generality by proving a \LS for the Gibbs measure of the system. We refer to [MO2] for a general introduction to the problem of approach to equilibrium in the one phase region for Glauber dynamics; in particular in [MO2] one finds a critical discussion of the various {\it finite volume } mixing conditions for the Gibbs state and of the role played by the shape of the volumes involved when getting near to a line of first order phase transition. We also refer the reader to the beautiful series of papers by Zegarlinski [Z1], [Z2], [Z3] and Zegarlinski and Stroock [SZ1], [SZ2], [SZ3] where the theory of the \LS for Gibbs states was developed and its role in the proof of fast convergence to equilibrium of general, not necessarily attractive, Glauber dynamics was clarified. We suggest in particular the interested reader to look at the nice review in [S].\bigskip In order to precisely state our results and for reader convenience, we recall here the model and the notation of the first paper of this series.\bigskip $\S\;1$\hskip 1cm {\bf The Model}\par \par We will consider lattice spin systems with finite single spin state space $S$. We take for simplicity $S\, =\, \{-1,+1\}$ and we will denote by $\s\,\equiv\,\s_\L$ an element of the configuration space $\Omega_\Lambda \, = \,S^\Lambda $ in a subset $\Lambda \subset \Z$. The symbol $\sigma_x $ will always denote the value of the spin at the site $x\in\Lambda $ in the configuration $\sigma$.\par The {\it energy} associated to a configuration $\sigma\in\Omega_\Lambda $ when the boundary condition outside $\Lambda $ is $\tau\in\Omega_ {\Lambda ^c}$ is given by : $$H^\tau_\Lambda (\sigma) \,=\, H_\Lambda (\sigma | \tau) \,=\, \sum _{X:X\cap\Lambda \ne\emptyset}U_{X}\, \prod_{x\in X}(\t\s )_x \Eq(1.1)$$ where, in general, $\t\s$ denotes the configuration : $$\eqalign{(\sigma \tau)_x \,&=\, \sigma_x\;,\quad x\in\Lambda\cr (\sigma \tau)_x \,&=\, \tau_x\;,\quad x\in\Lambda ^c} \Eq(1.2)$$ and the potential $U \,=\, \{U_{X},\ X \subset \subset \Z\}$, where $X \subset \subset \Z$ means that $X$ is a finite subset of $\Z$, satisfies the following hypotheses: \bigskip\indent {\bf H1}.{\it \ \ Finite range : $\exists \quad r > 0 : U_{X}\equiv 0$ if diam$X > r$ (we use Euclidean distance).} \bigskip\indent {\bf H2}.{\it \ \ Translation invariance $$\forall X \subset \subset \Z\quad\forall k\in \Z\quad U_{X+k} \,=\, U_{X}$$ } Because of the hypothesis {\bf H1}, $ H^{\tau}_{\Lambda}(\sigma)$ depends only on $\tau_x$ for $x$ in $\partial^+_r\Lambda$ : $$\partial^+_r \Lambda \,=\, \{x\not\in\Lambda : \hbox{dist} (x, \Lambda )\leq r\} \Eq(1.3)$$ With the energy function $H_\L^\t(\s )$ we construct the usual {\it Gibbs measure} in $\Lambda $ with b.c. $\tau\in\Omega_{\Lambda ^c}$ given by: $$\mu_\Lambda ^\tau(\sigma) \,=\, {\hbox{exp}(H^\tau_\Lambda (\sigma))\over Z^\tau_\Lambda } \Eq(1.4)$$ where the normalization factor, or {\it partition function}, is given by $$Z^\tau_\Lambda \,=\,\quad\sum_{\sigma\in\Omega_\Lambda } \hbox{exp}(H^\tau_\Lambda (\sigma)) \Eq(1.5)$$ If there exists a unique limiting Gibbs measure for $\Lambda \to \Z,$ independent on $\tau$, it will be denoted by $\mu$ .\bigskip {\bf Remark} Notice that, for future notation convenience, we have included the usual $-\beta$ factor in the Boltzman weight \equ(1.4) directly in the energy $H^\tau_\Lambda (\sigma)$ \bigskip Next we define the stochastic {\it jump} dynamics, given by a continuons time Markov process on $\Omega \,=\, S^{ \Z}$ that will be studied in the sequel. Discrete time versions can also be considered.\par Given $\Lambda \subset \subset \Z$ let $$D (\Lambda) \,=\, \{f: \Omega \to \ R : f (\eta) \,=\, f (\sigma) \ \hbox{if}\quad \eta_x \,=\, \sigma_x \ \forall x \in \Lambda \} $$ be the set of {\it cylindrical functions} with support $\Lambda.$ The set $$ D\,=\, \cup_{\Lambda} D (\Lambda)$$ is the set of cylindrical functions and by $C(\Omega )$ we denote the set of all continuous functions on $\Omega \,=\, \Pi_x S_x$ with respect to the product topology of discrete topologies on $S$. \medskip The dynamics is defined by means of its {\it generator} $L$ which is given, for $f \in D$, by: $$L f (\sigma) \,=\, \sum_{x,a} c_x (\sigma, a) \bigl ( f (\sigma^{x,a})-f (\sigma) \bigr )\Eq(1.6)$$ where $\sigma^{x,a}$ is the configuration obtained from $\sigma$ by setting the spin at $x$ equal to the value $a$ and the non-negative quantities $c_x (\sigma, a) $ are called ``jump rates''.\par We will also consider the Markov process associated to the above described jump rates in a {\it finite volume} $\Lambda$ with boundary conditions $\tau$ outside $\Lambda.$ By this we mean the dynamics on $\Omega_\Lambda$ generated by $L^\tau_\Lambda$ defined as before starting from the jump rates $$c^{\tau, \Lambda}_x (\sigma, a) \equiv c_x (\sigma \tau, a)$$ where, given $\tau \in \Omega_{\Lambda ^c}\hbox{ and } \; \sigma \in \Omega_\Lambda \; ,\,\sigma \tau$ has been defined in \equ(1.2) . \medskip The general hypoteses on the jump rates, that we shall always assume, are the following ones. \bigskip {\bf H3.}\ Finite range $r$. This means that $\eta (y) \,=\, \sigma (y) \quad \forall\, x,y :\; |y - x| \leq r $ implies \indent \ \ $c_x (\sigma, a) \,=\, c_x (\eta , a)$ \bigskip {\bf H4.}\ Translation invariance. That is if $\eta (y) \,=\, \sigma (y+x) \quad \forall \,y\;$ then $\;c_x (\sigma, a) \,=\, c_x (\eta, a)$ \bigskip {\bf H5.}\ Positivity and boundedness. There exist two positive constants $k_1$, $k_2$ such that $$0\,<\, k_1\,\leq \, \inf_{\sigma, x, a} \ c_x (\sigma, a ) \,\leq\,\sup_{\sigma, x, a} \ c_x (\sigma, a )\,\leq \,k_2 $$ \bigskip \item{-}\ \ {\bf H6.}\ Reversibility with respect to the Gibbs measure $\mu$ (in finite or infinite volume): $$\hbox{exp}(\sum_{ X \ni x}U_X(\prod_{y\in X}(\s )_y))c_x(\sigma ,a)\;=\; \hbox{exp}(\sum_{X \ni x}U_X(\prod_{y\in X}(\s^{x,a} )_y)) c_x(\sigma^{x,a} ,\sigma_x)\quad \forall\,x\in \L\Eq(1.7)$$ A similar equation holds in finite volume $\Lambda$ with boundary conditions $\tau$, provided that we replace in \equ(1.7) $\sigma$ with the configuration $ \sigma \tau$.\par \bigskip It is immediate to check that, in finite volume, reversibility implies that the unique invariant measure of the dynamics coincides with the Gibbs measure $\mu^\tau_\Lambda$. This important fact holds also in infinite volume provided that there exists a unique Gibbs measure in the thermodynamic limit. \par It is well known (see $[L]$) that under the above conditions $L$ ($L^\tau_\Lambda$) generates a unique positive selfadjoint contraction semigroup on the space $L^2(\O ,d\mu )$ ( $L^2(\O_\L ,d\mu_\L^\t )$) that will be denoted by $T_t$ or $T_t^{\Lambda ,\tau}$ .\bigskip $\S\;2$\hskip 1cm {\bf The \LS and Hypercontractivity of $\bf T_t$}\par In order to introduce the \LS (LSI) we have to define the differentiation operator on the functions of the spin configurations. We set: $$\partial_{\s_x}f(\s)\;=\;f(\s)\,-\, {1\over 2}[f(\s^{x,+1})\,+\, f(\s^{x,-1})]\Eq(1.7bis)$$ where $\s^{x,+1}$ is the configuration obtained from $\s$ by setting the spin at x equal to +1 and similarly for $\s^{x,-1}$. Given a subset A of the lattice $\Z$ the symbol $(\nabla_A{f})^2$ will be a shorthand notation for the expression $\sum_{x\in A}(\partial_{\s_x}f(\s))^2$.\par Finally we define the "standard" \LSC $c_s(\nu)$ for an arbitrary measure $\nu$ on $\O_\L$ as the smallest number $c$ such that for any non negative function f : $\O_\L\, \to \, {\bf R}$ the following inequality holds: $$\nu(f^2log(f))\;\leq\;c\nu ((\nabla_{\L}f)^2)\;+\;\nu(f^2) log((\nu (f^2))^{1\over 2})\Eq(1.8)$$ where $\nu (f)$ denotes the average of the function f with respect to the measure $\nu$.\par In the sequel we will refer to \equ(1.8) as the "standard" \LS for $\nu$. It is very important to observe that if we denote by $${\bf {\cal E}}_\L^\t(f,f)\;=\;-\mu_\L^\t(f\,L_\L^\t\,f)$$ the Dirichlet form associated to the generator $L_\L^\t$ and we take in \equ(1.8) $\nu\;=\;\mu_\L^\t$, then the term: $$\mu_\L^\t ((\nabla_{\L}f)^2)$$ satisfies the estimate: $$(4\max_{x,a,\s}c^{\tau, \Lambda}_x (\sigma, a))^{-1}\,{\bf {\cal E}} _\L^\t(f,f)\leq\mu_\L^\t ((\nabla_{\L}f)^2)\leq (4\min_{x,a,\s}c^{\tau, \Lambda}_x (\sigma, a))^{-1}\,{\bf {\cal E}} _\L^\t(f,f)$$ Therefore, if $\mu_\L^\t$ satisfies the "standard" \LS \equ(1.8), with standard logarithmic Sobolev constant $c_s(\mu_\L^\t )$, then it also satisfies the \LS for the semigroup $T_t^{\Lambda ,\tau}$ relative to the measure $\mu_\L^\t$: $$\mu_\L^\t(f^2log(f))\;\leq\;c_{{\bf {\cal E}}}(\mu_\L^\t ){\bf {\cal E}} _\L^\t(f,f)\;+\;\mu_\L^\t(f^2) log((\mu_\L^\t (f^2))^{1\over 2})\Eq(1.10)$$ with logarithmic Sobolev constant $(4k_2)^{-1}c_s(\mu_\L^\t )\,\leq \,c_{{\bf {\cal E}}}(\mu_\L^\t )\,\leq \, (4k_1)^{-1}c_s(\mu_\L^\t )$ because of H5. Thus the "standard" \LS and the \LS for the Dirichlet form ${\bf {\cal E}}$ are equivalent and in the sequel, whenever confusion does not arise, we will call both of them the \LS. \bigskip {\bf Remark} From the above discussion it also immediately follows that, if $c^{\tau, \Lambda}_x (\sigma,a)$ and $\tilde c^{\tau, \Lambda}_x (\sigma, a)$ are two different jump rates satisfying H3...H6, and if ${\bf {\cal E}}$ and $\tilde {\bf {\cal E}}$ are the corresponding Dirichet forms, then we have: $$c_{{\bf {\cal E}}}(\mu_\L^\t )\,\leq\, \max_{x,a,\s}{c^{\tau, \Lambda}_x (\sigma, a)\over \tilde c^{\tau, \Lambda}_x (\sigma, a)} c_{\tilde {\bf {\cal E}}}(\mu_\L^\t )\Eq(1.11)$$ and $$c_{{\bf {\cal E}}}(\mu_\L^\t )\,\geq\, \min_{x,a,\s}{c^{\tau, \Lambda}_x (\sigma, a)\over \tilde c^{\tau, \Lambda}_x (\sigma, a)} c_{\tilde {\bf {\cal E}}}(\mu_\L^\t )\Eq(1.12)$$ \par {\bf Remark} In the case when the single spin space consists of N elements with $N>2$, the definition of the differentiation operator is no longer so clear. One possibility (see [SZ3]) is to set: $$\partial_{\s_x}f(\s)\;=\;f(\s)\,-\, \media {f}_x\Eq(1.12bis)$$ where $\media {f}_x$ is the average with respect to the uniform measure on $S$ of the function $f$, considered as a function of the single spin $\s_x$. Another possibility is to order the elements of $S$ as $s_1,...,s_N$ and to set: $$\partial_{\s_x}f(\s)\;=\;{f(s_{i+1})\,-\,f(s_{i})\over 2} \Eq(1.12tris)$$ if $\s_x\,=\,s_i$ with $s_{N+1}\,\equiv\,s_1$.\par Both definitions are reasonable and equivalent in the sense that: \item{i)} $\media {\partial_{\s_x}f}_x\;=\;0$ \item{ii)}there exists a finite positive constant $k_o$ (in general depending on N) such that : $$(k_o)^{-1}\,{\bf {\cal E}} _\L^\t(f,f)\leq\mu_\L^\t ((\nabla_{\L}f)^2)\leq k_o\,{\bf {\cal E}} _\L^\t(f,f)$$ where the definition of the generator $L$ and of the Gibbs measure in this more general setting is the obvious one.\par Although, in some sense, the choice of the differentiation operator reflects the choice of the dynamics for the single spin at "infinite temperature" (a uniform sampling in the first case or a symmetric random walk in the second case), because of ii) above, one is free to choose whatever definition is more suited to the methods of proof. In particular if, as we do, one wants to treat $\partial_{\s_x}$ as much as possible as a continuous derivative, then the second definition seems more suited.\par The above "ambiguity" points out that the logarithmic Sobolev constant is not intrinsically associated to the Gibbs measure but, rather, to the pair ($\mu\,,\,\nabla$).\bigskip As it is well known since the basic work by L.Gross [G1] (see also [G2] ), the \LS for the Gibbs state $\mu_\L^\t$ is strictly connected with the hypercontractivity properties of the Markov semigroup $T_t^{\L,\t}$, where: \bigskip \item{}{\it $T_t^{\L,\t}$ is {\it hypercontractive} with respect to $\mu_\L^\t$ if there exists a constant $c(\L,\t )$ such that $$\parallel T_t^{\L,\t}(f)\parallel_{L^q(\mu_\L^\t )}\,\leq \, \parallel f\parallel_{L^p(\mu_\L^\t )}\;\;\forall \,(p,q,t)\quad \hbox{with } p\,\leq \, q\,\leq\, 1\,+\,(p-1)e^{2t\over c(\L,\t )}\Eq(1.13)$$} \par More precisely, Gross' Theorem states that the constant $c(\L,\t )$ in \equ(1.13) can be taken equal to the logarithmic Sobolev constant $c_{\cal E}(\mu_\L^\t )$.\par Besides its intrinsic interest, hypercontractivity of the Markov semigroup $T_t^{\L,\t}$ or of $T_t$ is a fundamental tool to transform, in the general case, exponential convergence to equilibrium of the Glauber dynamics in the $L^2(d\mu_\L^\t )$-sense into exponential convergence to equilibrium in the $L^\infty$-sense. More exactly, for an interaction $U_X$ satisfying the general hypotheses H1, H2, one can easily prove the following theorem (see [SZ2] Lemma 2.9 and Lemma 1.8 there) :\bigskip {\bf Theorem 1.1}\par {\it Let $\Gamma$ be a, finite or infinite, class of subsets of $\Z$ and suppose that there is a constant $c_o$ such that: $$\sup_{\t, \L\in \Gamma}c_{\cal E}(\mu_\L^\t )\,\leq \,c_o$$ Then there exists a positive constant $m$ and for any cylindrical (i.e. depending on finitely many spins) $f:\O_\Z\,\to \, R$ there exists a finite constant $C_f$ such that: $$\sup_{\t, \L\in \Gamma}\parallel P_t^{\t,\L}f\, -\,\mu_\L^\t(f)\parallel_\infty\,\leq \, C_f \hbox{exp}(-mt)$$} Thus, in this way, the problem of proving exponentially fast approach to equilibrium in the uniform norm for a Glauber dynamics is reduced to proving a bound on the logarithmic Sobolev constant of the Gibbs state.\par The real breakthrough in this direction was made few years ago by Zegarlinski [Z1], [Z2], [Z3], who proved that, if the Gibbs state satisfies a weak coupling condition similar to the old uniqueness Dobrushin's condition (single site) [D1], then the logarithmic Sobolev constant in a set $\L$ with boundary conditions $\t$ is finite uniformly in $\L$ and $\t$. This result was then extended and generalized by Stroock and Zegarlinski [SZ1],[SZ2] ,[SZ3] to spin systems with single spin space given either by a compact Riemaniann manifold or by a finite discrete set. The main result of the above mentioned works is the equivalence of the existence of a finite logarithmic Sobolev constant and the Dobrushin-Shlosman complete analyticity [DS1], [DS2] of the Gibbs state $\mu$. Although the Stroock-Zegarlinski result was a very important progress in the general problem of relating the fast convergence to equilibrium of the dynamics to "mixing" properties of the Gibbs measure, it cannot, in general, be used concretely in order to establish results close to a line of a first order phase transition since it requires good mixing properties of the measure $\mu_\L^\t$ in volumes $\L$ of {\it arbitrary shape}. As we have discussed in detail in the first work of this series [MO2], if one is willing to produce results really close to a line of a first order phase transition, one should consider the Gibbs measure only on "fat" volumes like cubes or parallelipipeds with large enough shortest side. This is exactly the subject of the present work.\medskip $\S\;3$\hskip 1cm {\bf The Results}\par In order to present the new results of the present work, we need to recall the {\it finite volume} mixing condition that already played a important role in the first paper of this series [MO2].\medskip We say that a Gibbs measure $\mu_\Lambda $ on $\Omega_\Lambda $ satisfies a {\it strong mixing} condition with constants $C, \gamma$ if for every subset $\Delta \subset \Lambda $: $$ \sup_{\tau,\tau^{(y)} \in \Omega _{\Lambda^c}} Var(\mu_{\Lambda , \Delta}^\tau\ \ \mu_{\Lambda , \Delta}^{\tau^{(y)}})\leq C e^{-\gamma \hbox{dist}(\Delta, y)} \Eq(1.14)$$ where $\mu_{\Lambda , \Delta}^\tau$ denotes the {\it relativization} (or projection) of the measure $\mu_\Lambda^\t$ on $\O_\D$, $Var$ is the variation distance and $\tau^{(y)}_x = \tau_x$ for $ x\ne y$ and .\medskip We denote this condition by $SM(\Lambda, C,\gamma)$\medskip In [MO2] a lower bound on the gap in the spectrum of the generator $L_\L^\t$ was derived (see Theorem 1.2 below) under the assumption $SM(\Lambda_o, C,\gamma)$, where $\L_o$ is a cube of side $L_o$, provided that, given $C$ and $\gamma$, $L_o$ is sufficiently large.\par In what follows $\Gamma$ will denote the class of all subsets of $\bf Z^d$ given by the union of translates of the cube $\Lambda_o$ such that their vertices lay on the rescaled lattice $L_o{\bf Z^d}$ . The constants $C$ and $\gamma$ appearing in our mixing condition will be fixed once and for all. \bigskip {\bf Theorem 1.2}\par {\it There exists a positive constant $\bar L$ depending only on the range of the interaction and on the dimension d such that if {\it SM($L_o$,C,$\gamma$)} holds with $L_o\,\geq \, \bar L$ then: \medskip \item{i)} there exists a positive constant $m_o$ such that for any $\Lambda\,\in \,\Gamma$ and for any function f in $L^2(d\mu_{\Lambda}^{\tau})$ : $$\vert\vert T^{\Lambda ,\, \tau}_t(f)\, -\, \mu_{\Lambda}^{\tau} (f)\vert\vert_{L^2(d\mu_{\Lambda}^{\tau})} \; \leq \; \vert\vert\,f\,-\,\mu_{\Lambda}^{\tau} (f)\,\vert \vert_{L^2(d\mu_{\Lambda}^{\tau})}\hbox{exp}(-m_ot)$$ \medskip \item{ii)} There exist constants $C'$ and $\gamma '$ such that for any $\Lambda\,\in \,\Gamma$, $SM(\Lambda, C',\gamma')$ holds} \bigskip {\bf Corollary 1.1}\par {\it There exists a positive constant $\bar L$ depending only on the range of the interaction and on the dimension d such that if {\it SM($L_o$,C,$\gamma$)} holds with $L_o\,\geq \, \bar L$ then there exists a positive constant $m$ such that for any pair of cylindrical functions $f$ and $g$ with supports $S_f$ and $S_g$ and for any $\Lambda\,\in \,\Gamma$, with $S_f,\,S_g\, \subset \,\L$, one has: $$\mu_\L^\t(f;g)\,\leq \,\vert f\vert _\infty\, \vert g\vert _\infty\,\vert S_f\vert \,\vert S_g\vert _\infty \hbox{exp}(-m\, dist(S_f,S_g))$$} where $\vert X\vert $ denotes the cardinality of the set $X\subset \Z$.\par Here we considerably strengthen part i) of Theorem 1.2 by proving the following theorem:\bigskip {\bf Theorem 1.3}\par {\it There exists a positive constant $\bar L$ depending only on the range of the interaction and on the dimension d, such that if {\it SM($L_o$,C,$\gamma$)} holds with $L_o\,\geq \, \bar L$ then there exists a positive constant $c_o$ such that for any $\Lambda\,\in \,\Gamma$ and any boundary condition $\t$ the logarithmic Sobolev constant $c_{{\cal E}_\L^\t}$ is bounded by $c_o$.\par Moreover there exists a positive constant $m$ such that for any cylindrical function $f$ there exists a positive constant $C_f$, depending only on $f$, such that: $$\parallel P_t^{\t,\L}f\, -\,\mu_\L^\t(f)\parallel_\infty\,\leq \, C_f \hbox{exp}(-mt)$$ }\bigskip The second part of the above theorem proves Theorem 4.2 of [MO2] in the non attractive case.\medskip $\S\;4$\hskip 1cm {\bf The Strategy of the Proofs}\par We conclude this introduction by describing the ideas behind the proof of Theorem 3.1 and by comparing them with the Stroock-Zegarlinski's approach.\par Our proof is divided into two distinct parts: \item{1)} In this first part (see section 2) we show that any Gibbs measure $\nu$ on a set $\L$ which is the (finite or infinite) union of certain "blocks" $\L_1\,\dots\,\L_j\dots$ (e.g. cubes of side $l$ or single sites of the lattice $\Z$) has a logarithmic Sobolev constant which is not larger than a suitable constant depending on the maximum size of the blocks provided that the interaction (not necessarily of finite range) {\it between} the blocks is very weak in a suitable sense. A simple example of such a situation is represented by a Gibbs state at high temperature, but the result is more general since we do not assume that the interaction {\it inside} each block is weak.\par The result is a perturbative one since, as it is well known [G1], if there is no interaction between the blocks then the logarithmic Sobolev constant of $\nu$ is not larger than $$\sup_{j,\t}c_s(\nu_{\L_j}^\t )$$ \item{2)} In the second part (see section 3) of our approach we use renormalization group, in the form known as decimation (i.e. integration over a certain subset of the variables $\s_x$), to show that, in the assumption of the theorem, the Gibbs state $\mu_\L^\t$ after a finite ($\leq \,2^d$) number of decimations becomes a new Gibbs measure exactly of the type discussed in the part 1). It is then a relatively easy task to derive the boundedness of the logarithmic Sobolev constant of $\mu_\L^\t$.\par As it is well known since the work of Olivieri [O] and Olivieri and Picco [OP], the mixing condition $SM(\L_o,C,\gamma)$ implies that if the decimation is done over blocks of a sufficiently large size (see for instance [EFS] for pathologies that may occur if the size is not large enough) then it is possible to control, e.g. by a converging cluster expansion, the effective potential of the renormalized measure and to show that it satisfies the weak coupling condition needed in part 1). This is however more than what it is actually needed, since the hypotheses of part 1 are fulfilled by the renormalized measure as soon as the truncated two point correlation functions of the {\it original} Gibbs measure $\mu_\L^\t$ decay exponentially fast. This is exactly the content of corollary 1.1 above; therefore the method can avoid the lengthy computations of the cluster expansion.\par We want to notice at this point a difference in the role played by the DLR structure of Gibbs measures in our approach and in the one used by Zegarlinski and Stroock-Zegarlinski. For simplicity let us consider the case when the Dobrushin's uniqueness condition holds true. In [Z1],[Z2],[Z3] Zegarlinski uses in a crucial way the following property of the Gibbs local specification operator ${\bf E}_\L$ : $$\lim_{n\to \infty}{\bf E}_{i_n},\dots ,{\bf E}_{i_1}f\;=\; \mu (f)$$ where $\{i_k\in \Z\}_{k\in \bf N}$ is a suitable sequence going infinitely many times through each site of the lattice and $\mu$ is the unique infinite volume Gibbs state. A similar property is used in [SZ3] in the case when the Dobrushin-Shlosman complete analitycity condition holds true.\par In the present paper, on the contrary, we use the following simple general property valid for any probability measure $\nu$ on a finite space $\Omega$: $$\nu (A_1\cap\dots\cap A_n)\;=\;\nu (A_1|A_2\cap\dots\cap A_n) \nu (A_2|A_3\cap\dots\cap A_n)\dots\nu (A_{n-1}|A_n)\nu (A_n)$$ Clearly, if the measure $\nu$ is a Gibbs measure corresponding to a given potential, then the DLR property enters in the explicit computation of the conditionals probabilities $\nu (A_i|A_{i+1}\cap\dots\cap A_n)$.\par After the completion of this work we learned that also Lu and Yau, in their work on the gap for the Kawasaki dynamics for the Ising model [LY], obtained, by martingale techniques, a uniform bound on the Logarithmic Sobolev Constant of $\mu_\L^\t$, where $\L$ is an arbitrary cube of the lattice, under the assumption that $SM(\L,C,\gamma )$ holds for all finite cubes $\L$. \vskip 2cm %\input formato.tex \numsec=2\numfor=1 {\bf Section 2}\bigskip \centerline{LOGARITHMIC SOBOLEV INEQUALITY} \centerline{FOR WEAKLY COUPLED GIBBS MEASURES} \bigskip In this section we prove two results that will play a crucial role in the derivation of the logarithmic Sobolev inequality (LSI) for Gibbs measures satisfying a finite volume mixing condition. In order to present our results we need to precisely define the setting of the problem and the notation that we will adopt in the sequel. We warn the reader that in this section we {\it do not assume} the finite range condition (H1); in this situation we will denote the potential by $\F$ instead of $U$. \bigskip $\S\;1$\hskip 1cm {\bf The Setting of the Problem}\par Let $\L$ be a finite subset of the lattice $\Z$ such that $\L\,=\,\L_1\,\cup\,\L_2\,\cup\,,\dots ,\,\cup\,\L_N$ with $\L_i\,\cap\,\L_j\;=\;0$ if $i\,\neq \,j$ and let $\O_{\L}$ be the space of configurations $\O_{\L}\;=\;\{-1,+1\}^{\L}$. Let also $\F_X\;\; ,X\, \subset\, \Z$, be a "potential" such that for any "boundary" configuration $\h \,\in\,\O_{\Z\setminus \L}$ and any configuration $\s \,\in\,\O_{\L}$ the following Hamiltonian is finite: $$H^\h(\s)\;=\;\sum_{X\cap\L\,\neq\,0}\F_X (\prod_{x\in X}(\s\h )_x ) \Eq(2.1)$$ If for convenience we denote by $H^\h_{\L_i}(\s_{\L_i})$ the sum: $$H^\h_{\L_i}(\s_{\L_i})\;=\; \sum_{\{X\cap\L\,\neq\,0\;;\; X\cap\L_i\,= \,X\cap\L\}}\,\F_X (\prod_{x\in X}(\s\h )_x ) \Eq(2.2)$$ then the total Hamiltonian can be written as: $$H^\h(\s )\;=\;\sum_i\,H^\h_{\L_i}(\s_{\L_i})\;+\; W^\h(\s ) \Eq(2.3)$$ where the term $W^\h(\s )$ represents now the interaction between the sets $\L_i$ i=1...N .\par Given the Hamiltonian $H^\h(\s)$, we will denote by $\m^\h$ the corresponding Gibbs measure.\par In the sequel, together with the measure $\m^\h$, we will need also other Gibbs measures that are obtained from $\m^\h$ by integrating out one by one the variables $\s_{\L_j}$ j=1,2..(decimation procedure). More precisely for any i=1...N we define a new Gibbs measure $\m_{\geq i}^\h$ on the space $$\O_{\L}^{(\geq i)}\;=\; \{-1,+1\}^{\L_i\,\cup\,\L_{i+1}\,\cup\,,\dots ,\,\cup\,\L_N}$$ as the relativization to $\O_{\L}^{(\geq i)}$ of the measure $\m^\h$ : $$\m_{\geq i}^\h(\s_{\L_i},\dots ,\s_{\L_N})\;=\; \sum_{\s_{\L_1},\dots ,\s_{\L_{i-1}}} \m^\h(\s_{\L_1}, \dots ,\s_{\L_N})\Eq(2.5)$$ Obviously the measure $\m_{\geq i}^\h$ is also a Gibbs measure with a new Hamiltonian: $$\hat H^\h_{(\geq i)}(\s_{\L_{i}},\dots ,\s_{\L_{N}})\;=\; log(Z^{\h,\s_{\L_{i}},\dots ,\s_{\L_{N}}}_{\L_1,\dots ,\L_{i-1}})\Eq(2.5bis)$$ where $Z^{\h,\s_{\L_{i}},\dots ,\s_{\L_{N}}}_{\L_1,..\L_{i-1}}$ is the partition function in $\L_1\cup..\cup\L_{i-1}$ with boundary conditions $\h,\s_{\L_{i}},\dots ,\s_{\L_N}$.\par Finally, we will denote by $\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}$ the measure on $\O_{\L_i}$ obtained from $\m_{\geq i}^\h$ by conditioning to the event that the spin configurations $\s_{\L_{i+1}},\dots ,\s_{\L_N}$ are equal to the spin configurations $\t_{i+1},\dots ,\t_{N}$. We will write $\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}$ as : $$\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}\;=\; {\hbox{exp}(W^\h_{(i)}(\s_{\L_i};\t_{\L_{i+1}},\dots ,\t_{\L_N}))\over Z_i}\Eq(2.5tris)$$ where $Z_i\,\equiv \,Z_i(\t_{\L_{i+1}},\dots ,\t_{\L_N},\h)$ is a normalization factor and the "effective" interaction $$W^\h_{(i)}(\s_{\L_i};\t_{\L_{i+1}},\dots ,\t_{\L_N})\,\equiv\,W^\h_{(i)}(\s )$$ is given by: $$W^\h_{(i)}(\s_{\L_i};\t_{\L_{i+1}},\dots ,\t_{\L_N})\;=\; \hat H^\h_{(\geq i)}(\s_{\L_i};\t_{\L_{i+1}},\dots ,\t_{\L_N})\;-\;\hat H^\h_{(\geq i)}(\tilde \t_{\L_{i}},\t_{\L_{i+1}}\dots ,\t_{\L_N})\Eq(2.5iv)$$ where $\tilde \t_{\L_{i}}$ is a given reference configuration in $\L_i$ (e.g. all spins up).\par \bigskip $\S \;2$\hskip 1cm {\bf Assumptions and Results}\par We are now in a position to precisely state the hypotheses on the "potential" $\F_X$ that we need in order to prove the main results of this section. \bigskip {\bf Assumptions} \item{\bf a)} There exists a positive constant $\e$ such that: $$\sup_{(\h ,i,N)} \sum_{j=1}^{i-1}\sup_{x\in\L_i}\supnorm{\partial_{\s_x} W_{(j)}^\h(\s)}\;\leq\;\e \Eq(2.6)$$ and $$\sup_{(\h ,k,N)} \sum_{j=k+1}^{N}\sup_{x\in\L_j}\supnorm{\partial_{\s_x} W_{(k)}^{\h}(\s)}\;\leq\;\e \Eq(2.6bis)$$ \item{\bf b)} There exists a positive constant $k_o$ such that: $$\sup_{(\h ,N)} \sum_{j=1}^{N}\sup_{x\in\Z\setminus\L}\supnorm{\partial_{\h (x)}W_{(j)}^\h(\s)}\;\leq\;k_o\Eq(2.7) $$ and $$\sup_{(\h ,N)}\sup_k \sum_{x\in \Z\setminus\L}\supnorm{\partial_{\h (x)}W_{(k)}^\h(\s)}\;\leq\;k_o\sup_i|\L_i| $$ \item{\bf c)} There exists a positive constant $k_1$ such that for any N, any $\h$, any i=1,\dots ,N, any f $\in\; L^2(\O_{\L_i},d\n_i^{\h ,\t_{i+1},\dots ,\t_{N}})$ and any value of the conditioning spins $\t_{i+1},\dots ,\t_{}$ one has: $$\sum_{\s_{\L_i},\hat\s_{\L_i}} \n_i^{\h ,\t_{i+1},\dots ,\t_{N}}(\s_{\L_i}) \n_i^{\h ,\t_{i+1},\dots ,\t_{N}}(\hat\s_{\L_i}) [f(\s_{\L_i})\,-\,f(\hat\s_{\L_i})]^2\;\leq$$ $$\leq k_1^2\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}((\grad{i}{f})) \Eq(2.8)$$ {\bf Remark}\par Assumptions a) and b) are clearly expressing some decay property of the effective interaction of the Gibbs measures $\m_{\geq i}^\h$ . The reason why in this section we {\it do not } require finite range of the interaction is that after few steps of our decimation procedure, even a Gibbs state corresponding to a finite range interaction will be transformed into a new Gibbs measure corresponding to an effective interaction with unbounded range at least in some directions.\par Assumption c) looks somewhat more mysterious but nevertheless plays an important role. In some sense c) is a hypothesis of rapid approach to equilibrium for a heat bath or Metropolis dynamics in $\L_i$, reversible with respect to the Gibbs measure $\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}$. Using the arguments of section 1 $\S \,2$, the constant $k_1^2$ becomes in fact proportional to the inverse of the gap of the generator of the dynamics, i.e. $k_1^2$ is proportional to the relaxation time in the "block" $\L_i$. In the perturbation argument given below, the constant $\epsilon$, which expresses the weak coupling between the blocks $\L_1\dots \L_N$, will always appear multiplied by the costant $k_1^2$ and therefore the true "small" parameter of the analysis becomes the "coupling among the blocks $\times$ the relaxation time in a single block".\par We will see in the next section that all the above assumptions follow from the finite volume mixing condition $SM(C,\gamma ,L)$ defined in section 1 provided that $L$ is large enough and that the set $\L$ consists of a union of sufficiently "fat" subsets of the lattice $\Z$.\bigskip Under the above assumptions the following two theorems hold.\bigskip {\bf Theorem 2.1} \par {\it Given the constant $k_1$, for any $\d\,>\,0$ there exists $\e_o\,=\,\e_o(k_1,\d)$ such that if $\e\sup_i|\L_i|\,\leq \,\e_o$ then: $$\sup_\h c(\m^\h)\,\leq\,(1\,+\d)\, \sup_i\sup_{\h,\t_{i+1},\dots ,\t_{N}}c(\n_i^{\h ,\t_{i+1},\dots ,\t_{N}})$$} \bigskip {\bf Theorem 2.2} \par {\it Given the constants $k_o$ and $k_1$ there exists $\e_o\,=\,\e_o(k_0,k_1)$ such that if $\e\sup_i|\L_i|\,\leq \,\e_o$ then there exist two constants $k_2$, $k_3$ depending on $k_o$ and $k_1$ such that for any function f : $\O_\Z\, \to \, R$ the following inequality holds: $$(\nabla_{\Z\setminus\L}(\m^\h(f^2))^{1\over 2})^2\; \leq\; k_2\m^\h((\nabla_{\Z\setminus\L}f)^2)\;+\; k_3\sup_i|\L_i|\m^\h((\nabla_{\L}f)^2)$$}\bigskip {\bf Remark } Notice that in theorem 2.2 the function f is a function from $\O_\Z$ to $R$ and thus it may depend also on the boundary spins $\h$. Therefore the expression $(\m^\h(f^2))^{1\over 2}$ may depend on the spins $\h$ which are involved in the derivatives $\nabla_{\Z\setminus\L}$ in two ways: through the Gibbs measure $\m^\h$ and through the function f.\bigskip $\S \;3$\hskip 1cm {\bf Proof of Theorem 2.1}\par If there was no interaction among the sets $\L_i$ then the total Gibbs measure $\m^\h$ would have been a product measure and the proof of the theorem would be a very simple exercise . However our hypotheses say that the mutual interaction among the sets $\L_i$ is very weak in a suitable sense and it is therefore natural to try to make some perturbation theory around the non interacting case. Because of the structure of the LSI, we found convenient first of all to write the average $\m^\h(f)$ of an arbitrary function f in a form that resembles as much as possible that of the average of f over a product measure. This form is as follows : $$\m^\h(f)\;=\; \n_N^\h(\n_{N-1}^{\h ,\t_N}(,\dots ,(\n_1^{\h ,\t_2,\dots ,\t_N}(f)..)\Eq(2.9)$$ If we now apply this representation of $\m^\h(f)$ to the function $\log{f}$ we get : $$\m^\h(\log{f})\;=\; \n_N^\h(\n_{N-1}^{\h ,\t_N}(,\dots ,(\n_1^{\h ,\t_2,\dots ,\t_N}(\log{f})..)\;\leq $$ $$\leq \; c_1\n_N^\h(\n_{N-1}^{\h ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}( \grad{1}{f}))\;+\;$$ $$\n_N^\h(\n_{N-1}^{\h ,\t_N}(....(\n_2^{\h ,\t_3,\dots ,\t_N}( \logg{\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)}))\Eq(2.10)$$ where : $$c_1\;=\;\sup_{\t_2,\dots ,\t_N,\,\h}c(\n_1^{\h ,\t_2,\dots ,\t_N})$$ Next we define the new function $$g_1\,=\,(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)) ^{1\over 2}$$ and, more generally: $$g_i\,=\,(\n_i^{\h ,\t_{i+1},\dots ,\t_N}(g_{i-1}^2))^{1\over 2}\Eq(2.11)$$ With these notation if we iterate \equ(10) we obtain : $$\m^\h(\log{f})\;\leq\;\sum_ic_i\m^\h(\grad{i}{g_{i-1}})\;+\; \logg{\m^\h(f^2)}\Eq(2.12)$$ where $$c_i\;=\;\sup_{\t_{i+1},\dots ,\t_N,\,\h}c(\n_i^{\h ,\t_{i+1},\dots ,\t_N})$$ We are thus left with the estimate of the term: $$\sum_ic_i\m^\h(\grad{i}{g_{i-1}})\Eq(2.13)$$ This is done in the next proposition where we show that, by paying a small price if the constant $\e$ is small enough, one can safely replace in \equ(2.13) the functions $g_{i-1}$ with the function f.\bigskip {\bf Proposition 2.1}\par {\it Given the constant $k_1$, for any $\d\,>\,0$ there exists $\e_o\,=\,\e_o(k_1,\d)$ such that if $\e\sup_i|\L_i|\,\leq \,\e_o$ then: $$\sum_i\m^\h(\grad{i}{g_{i-1}})\;\leq\;(1\,+\,\d) \sum_i\m^\h(\grad{i}{f})$$}\bigskip Before giving the proof of the proposition let us finish the proof of the theorem. If we use the result of the propostion we see that for any $\d\,>\,0$ there exists $\e_o\,=\,\e_o(k_1,\d)$ such that if $\e\sup_i|\L_i|\,\leq \,\e_o$ then the r.h.s. of \equ(2.12) can be bounded above by: $$(1\,+\,\d)\sup_ic_i\,\sum_i\m^\h(\grad{i}{f})\;+\; \logg{\m^\h(f^2)}\Eq(2.14)$$ \ie $$\sup_\h c(\m^\h)\,\leq\,(1\,+\d)\, \sup_i\sup_{\h,\t_{i+1},\dots ,\t_{N}}c(\n_i^{\h ,\t_{i+1},\dots ,\t_{N}})$$ and the theorem follows.\bigskip {\bf Proof of Proposition 2.1}\par One technical complication of working with discrete spins and discrete derivatives is that the latter do not enjoy exactly the same properties of the usual continuous derivatives like Leibniz rule and so forth.\par Therefore before entering into the details of the proof let us give some elementary results concerning discrete derivatives that will be frequently used later on. Properties a),b),c),d),e) follow only from the definition of $\partial_{\s_x}$ and are true in general whereas property f) is a consequence of assumption {\bf c)} above on the interaction. A proof can be found in the Appendix.\par In what follows $\media{f}_x$ will denote the average of the function f($\s_x$) with respect to the measure ${1\over 2}(\d_{+1}\;+\;\d_{-1})$ and, for any given x $\in \;\L_j$ with $j\,>\,i$ $$\l^x_i\;=\; \supnorm{{d\n_i^{\h ,\t_{i+1}^{x,+1},., \t_N^{x,+1}}\over d\n_i^{\h ,\t_{i+1}^{x,-1},.\t_N^{x,-1}}}}\,\vee\,\supnorm{{d\n_i^{\h ,\t_{i+1}^{x,-1},., \t_N^{x,-1}}\over d\n_i^{\h ,\t_{i+1}^{x,+1},.\t_N^{x,+1}}}}$$ Then we have: \bigskip \item{a)} $\partial_{\s_x}f^2\;=\;2\media{f}_x\partial_{\s_x}f$ \item{b)} $\n_i^{\h ,\t_{i+1},., \t_N}(\media{|f|}_x)\;\leq\;\l_i^x\media{\n_i^{\h ,\t_{i+1},., \t_N}(|f|)}_x$ \item{c)} If f is non negative then $f(\s_x)\,\leq\,2\media{f}_x$. \item{d)} Given x $\in \;\L_j$ with $j\,>\,i$: $$\vert\partial_{\t_x} [\n_i^{\h ,\t_{i+1},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f)..)]\vert\;\leq $$ $$|[\partial_{\t_x}( \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)|\;+$$ $$\prod_{j=i}^1\l^x_j \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N} (|\partial_{\t_x}f|)..) $$ where $|[\partial_{\t_x} \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)|$ is a convenient way to denote the expression: $${|\n_i^{\h ,\t_{i+1}^{x,+1},.,\t_N^{x,+1}} (.(\n_1^{\h ,\t_2^{x,+1},.,\t_N^{x,+1}}(f(\s\t)..)\,-\, \n_i^{\h ,\t_{i+1}^{x,-1},.,\t_N^{x,-1}} (.(\n_1^{\h ,\t_2^{x,-1},.,\t_N^{x,-1}}(f(\s\t)..)|\over 2}$$ \item{e)} ("Leibniz rule"). Let f be non negative. Then : $$|[\partial_{\t_x} (\n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N})](f)|\;\leq$$ $$|[\partial_{\t_x}\n_i^{\h ,\t_{i+1},\dots ,\t_N}]( \n_{i-1}^{\h ,\t_{i},\dots ,\t_N}(...(\n_1^{\h ,\t_2,\dots ,\t_N}(f)..)|\;+\;$$ $$\l^x_i\n_i^{\h ,\t_{i+1},\dots ,\t_N}( |[\partial_{\t_x}( \n_{i-1}^{\h ,\t_{i},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N})](f)| $$ \item{f)} $$|[\partial_{\t_x}\n_i^{\h ,\t_{i+1},\dots ,\t_N}](f^2)|\;\leq$$ $$2k_1\l^x_i\supnorm{\partial_{\t(x)}W_{(i)}^\h(\s\t)} [\n_i^{\h ,\t_{i+1},\dots ,\t_N}(f^2)]^{1\over 2} [\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}(\grad{i}{f})]^{1\over 2}$$ \bigskip We are now ready to start our computations.\par Since we have to estimate terms like $\grad{i}{g_{i-1}}$, we start to estimate the following quantity: $$|\partial_{\t_x}((\n_{i-1}^{\t_{i},\dots ,\t_N}(g^2_{i-2}))^{1\over 2})|$$ where x is an arbitrary site in the set $\L_i$. Following Zegarlinski and using a) above, we observe that it is sufficient to estimate $$|\partial_{\t_x}((\n_{i-1}^{\t_{i},\dots ,\t_N}(g^2_{i-2}))|$$ by : $$2\media{(\n_{i-1}^{\t_{i},\dots ,\t_N}(g^2_{i-2}))^{1\over 2}}_x[....]$$ to get that $$|\partial_{\t_x}(((\n_{i-1}^{\t_{i},\dots ,\t_N}(g^2_{i-2}))^{1\over 2})|\,\leq\, [...]\Eq(2.15)$$ Using d) and e) above with $f$ replaced by $f^2$, we get: $$\vert\partial_{\t_x} [\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)..)]\vert\;\leq\; $$ $$(\prod_{j=i-1}^1\l_j^x)\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(|(\partial_{\t_x}f^2)|)..)\;+$$ $$\sum_{k=1}^{i-1}(\prod_{j=i-1}^{k+1}\l_j^x) \n_{i-1}^{\t_{i},\dots ,\t_N}(..\n_{k+1}^{\t_{k+2},\dots ,\t_N} (|[\partial_{\t_x}\n_{k}^{\t_{k+1},\dots ,\t_N}] \n_{k-1}^{\t_{k},\dots ,\t_N} (..\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)...)|)\Eq(2.16)$$ The first term in the r.h.s of \equ(2.16), using the Schwartz inequality, and properties a) and b), is bounded from above by: $$2(\prod_{j=i-1}^1\l_j^x)^{3\over 2}\media{(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} (f^2)..)^{1\over 2}}_x (\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\t_x}f)^2)..)^{1\over 2}\Eq(2.17)$$ Using f) above, the second term in the r.h.s of \equ(2.16) is bounded from above by: $$\sum_{k=1}^{i-1}(\prod_{j=i-1}^{k+1}\l_j^x) 2k_1\l^x_k\supnorm{\partial_{\s(x)}W_{(k)}^{\h}(\s\t)}$$ $$\n_{i-1}^{\t_{i},\dots ,\t_N}(..(\n_{k+1}^{\t_{k+2},\dots ,\t_N}( [\n_k^{\t_{k+1},\dots ,\t_N}(g_{k-1}^2)]^{1\over 2} [\n_k^{\t_{k+1},\dots ,\t_{N}}((\grad{k}{g_{k-1}}))]^{1\over 2} \Eq(2.18)$$ Using the definition of the function $g_{k-1}$ we have : $$[\n_k^{\t_{k+1},\dots ,\t_N}(g_{k-1}^2)]^{1\over 2}\;=\; [\n_k^{\t_{k+1},\dots ,\t_N}(...(\n_1^{\h ,\t_2,\dots ,\t_N} (f^2)..)]^{1\over2}$$ and thus, using again a) and the Schwartz inequality, we get that \equ(2.18) is bounded above by: $$2\media{[\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} (f^2)..)]^{1\over 2}}_x\sum_{k=1}^{i-1}[(\prod_{j=i-1}^{k+1}\l_j^x) \supnorm{\partial_{\s(x)}W_{(k)}^{\h}(\s\t)} 2k_1\l^x_k]$$ $$\{\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\grad{k}{g_{k-1}}))\}^{1\over2}\Eq(2.19)$$ We now define: $$V_{i,k}\;=\;\sup_{x\in\L_i} [(\prod_{j=i-1}^{k+1}\l_j^x) \supnorm{\partial_{\s(x)}W_{(k)}^{\h}(\s\t)} 2k_1\l^x_k]$$ and $$B_{i-1}\;=\;\sup_{x\in \L_i}(\prod_{j=1}^{i-1}\l_j^x)^{3\over 2}$$ With this notation, if we put together \equ(2.17) and \equ(2.19), we obtain that : $$\vert\partial_{\t_x} [\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)..)]\vert\;\leq\;$$ $$2\media{(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} (f^2)..)^{1\over 2})}_x \{B_{i-1}(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\t_x}f)^2)..)^{1\over 2}\;+\; $$ $$+\;\sum_{k=1}^{i-1}V_{i,k}[\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\grad{k}{g_{k-1}}))]^{1\over2}\}\Eq(2.20)$$ \ie , using \equ(2.15), $$ \vert\partial_{\t_x} [\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)..)]^{1\over 2}\vert\;\leq\;$$ $$B_{i-1}(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\t_x}f)^2)..)^{1\over 2}\;+\;$$ $$+\; \sum_{k=1}^{i-1}V_{i,k}[\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\grad{k}{g_{k-1}}))]^{1\over2}\Eq(2.21)$$ Thus, using the above bound, we get that $$\sum_{x\in\L_i}\vert\partial_{\t_x} [\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)..)]^{1\over 2}\vert^2\;\leq\;$$ $$pB_{i-1}^2(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\L_i}f)^2)..)\;+\;$$ $$+\; |\L_i|q(\sum_{k=1}^{i-1}V_{i,k}) \sum_{k=1}^{i-1}V_{i,k}[\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\grad{k}{g_{k-1}}))]\Eq(2.22)$$ for any $p\,>\,1$, $q\,>\,1$ with ${1\over p}\,+\,{1\over q}\;=\;1$.\par In conclusion, by summing \equ(2.21) over the index i, we obtain for the initial expression $\sum_i\m^\h(\grad{i}{g_{i-1}})$ the bound : $$\sum_i\m^\h(\grad{i}{g_{i-1}})\;\leq\;$$ $$\sum_i\{ pB_{i-1}^2\m^\h((\grad{i}{f})^2)\;+\; |\L_i|q(\sum_{k=1}^{i-1}V_{i,k}) \sum_{k=1}^{i-1}V_{i,k}\m^\h( ((\grad{k}{g_{k-1}}))\}\Eq(2.23)$$ At this point we use our decay assumption {\bf a)} on the interaction in order to estimate the numbers $B_i$ and $V_{i,k}$.\par >From the definition of $\l_j^x$ one has immediately that for $j\,\leq \,i$ : $$\sup_{x\in \L_i}\l_j^x\;\leq \;\hbox{exp}(4\sup_{x\in \L_i} \supnorm{\partial_{\t_x}W_{(j)}^\h})\Eq(2.24)$$ Therefore : $$B_{i-1}\;\leq \;\hbox{exp}(8\sum_{j=1}^{i-1}\sup_{x\in \L_i} \supnorm{\partial_{\t_x}W_{(j)}^\h})\;\leq\; \hbox{exp}(8\e)\;\leq\;1\,+\,10\,\e\Eq(2.25)$$ if $\e$ is small enough.\par Similarly : $$\eqalign{&\sum_{k=1}^{i-1}V_{i,k}\;\leq \;C\e\cr &\sum_{i=k+1}^{N}V_{i,k}\;\leq \;C\e}$$ for a suitable constant C depending on $k_1$ and any $\e$ sufficently small.\par Thus the second term in the r.h.s. of \equ(2.23) is smaller than: $$\sup_i|\L_i|q(C\e)^2\sum_{k=1}^{N}\m^\h( ((\grad{k}{g_{k-1}}))\}\Eq(2.26)$$ Therefore if $\sup_i|\L_i|q(C\e)^2\;<\;1$ we get from \equ(2.23), \equ(2.25) and \equ(2.26) that: $$\sum_i\m^\h(\grad{i}{g_{i-1}})\;\leq\;$$ $$\leq \;{p(1\,+\,10\e)\over 1\,-\,\sup_i|\L_i|q(C\e)^2}\sum_i\{ \m^\h((\grad{i}{f}))\;\leq\;(1\,+\,\d)\sum_i\{ \m^\h((\grad{i}{f}))$$ if we choose for example $q\,=\,{1\over \e}$ and $\e$ sufficiently small. The proposition is proved. \bigskip $\S \;4$\hskip 1cm {\bf Proof of Theorem 2.2}\par We proceed very similarly to the proof of proposition 2.1. By doing the same kind of computations as in \equ(2.15),\dots ,\equ(2.21) we obtain that: $$ \{\vert\partial_{\h_x} [\n_{i-1}^{\t_{i},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f^2)..)]^{1\over 2}\vert\}^2\;\leq\;$$ $$\{2(\,\sup_{x\in \Z\setminus\L}\prod_{j=1}^N(\l_j^x)^{3\over 2}\,)^2(\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\h_x}f)^2)..)\;+\;$$ $$+\; 2(\sum_{k=1}^{N}\hat V_{x,k}[\n_{N}^{\h } (...(\n_k^{\h ,\t_{k+1},\dots ,\t_N} ((\grad{k}{g_{k-1}}))]^{1\over 2}\,)^2\,\}\Eq(2.27)$$ where $$\hat V_{x,k}\;=\; [(\prod_{j=i-1}^{k+1}\l_j^x) \supnorm{\partial_{\h(x)}W_{(k)}^{\h}(\s\t)} 2k_1\l^x_k]$$ It is important to observe at this point that in general $\hat V_{x,k}$ is not small. Assumption a) (see \equ(2.6) or \equ(2.6bis)) in fact concerns only the interaction between different blocks while in some sense $\hat V_{x,k}$ measures the interaction between one block $\L_k$ and the boundary spin $\h_x$. However, thanks to assumption b) (see \equ(2.7)) and using \equ(2.24), we have that : $$\sup_{\h ,k}\sum_{x\in\Z\setminus\L}\hat V_{x,k}\;\leq \; 2k_1k_o\sup_i|\L_i|\hbox{exp} (\sum_{j=1}^N4\supnorm{\partial_{\h(x)}W_{(j)}^{\h}(\s\t)})\;\leq$$ $$\leq \;\sup_i|\L_i|2k_1k_o\hbox{exp}(4k_o)\Eq(2.28)$$ Thus if we sum over $x\in\Z\setminus\L$ the second term in the r.h.s. of \equ(2.27), we get, after a Schwartz inequality : $$\sum_{x\in\Z\setminus\L}\,2(\sum_{k=1}^{N}\hat V_{x,k})[\sum_{k=1}^{N}\hat V_{x,k}\n_{N}^{\h } (.(\n_k^{\h ,\t_{k+1},\dots ,\t_N} ((\grad{k}{g_{k-1}}))]\;\leq \; $$ $$\leq \; \sup_i|\L_i|\,2k_o^2k_1^2\hbox{exp}(8k_o)(\sum_{k=1}^{N}\n_{N}^{\h } (...(\n_k^{\h ,\t_{k+1},\dots ,\t_N} ((\grad{k}{g_{k-1}}))\Eq(2.29)$$ Using the identity: $$\n_{N}^{\h } (...(\n_k^{\h ,\t_{k+1},\dots ,\t_N} ((\grad{k}{g_{k-1}}))\;=\;\m^\h((\grad{k}{g_{k-1}})$$ and proposition 2.1 , we see that if $\e$ is small enough there exists a constant $k_3$ such that: $$\sum_{x\in\Z\setminus\L}\,2(\sum_{k=1}^{N}\hat V_{x,k})[\sum_{k=1}^{N}\hat V_{x,k}\n_{N}^{\h } (...(\n_k^{\h ,\t_{k+1},\dots ,\t_N} ((\grad{k}{g_{k-1}}))]\;\leq \;$$ $$\sup_i|\L_i|\,k_3\m^\h(\grad{\L}{f})\Eq(2.30)$$ Analogously if we sum over $x\in\Z\setminus\L$ the first term in the r.h.s. of \equ(2.27) we get : $$\sum_{x\in\Z\setminus\L}\,2(\,\sup_{x\in \Z\setminus\L}\prod_{j=1}^N\l_j^x\,)^2 (\n_{i-1}^{\t_{i},\dots ,\t_N} (...(\n_1^{\h ,\t_2,\dots ,\t_N} ((\partial_{\h_x}f)^2)..)\;\leq\;$$ $$\leq \;2\hbox{exp}(8k_o)\m^\h(\grad{\Z\setminus\L}{f})\Eq(2.31)$$ If we finally combine \equ(2.31) and \equ(2.30) we obtain the theorem with $k_2\;=\;2\hbox{exp}(8k_o)$.\vskip 2cm %\input formato.tex \numsec=3\numfor=1 \bigskip {\bf Section 3}\bigskip \centerline{DECIMATION APPROACH TO THE} \centerline{LOGARITHMIC SOBOLEV CONSTANT}\bigskip $\S\;1$\hskip 1cm {\bf Proof of theorem 1.1}\par In this section we prove our main result namely theorem 1.1 .\par The result holds in any dimension but for the sake of simplicity of the exposition we will discuss explicitely only the two dimensional case. As already announced in section 1, our proof is based upon ideas coming from rigorous renormalization group in classical statistical mechanics in the form known as {\it decimation}. We begin therefore by illustrating our decimation procedure.\par For any odd integer $L_o$ let us consider the renormalized lattice $\Z(L_o)\;=\;L_o\Z\,\subset \Z$ and let us define for any $x\,\in \,\Z(L_o)$ the block $Q_{L_o}(x)$ as the square in the original lattice, centered at x and of side $L_o$. We will collect the blocks $Q_{L_o}(x)$ into four different families, denoted in the sequel by the letters A, B, C, D, according to whether the coordinates of their centers x are (even, odd), (even, even), (odd, even) or (odd, odd). We will also order in some way the blocks belonging to the same family so that they will be denoted by $A_1,\,A_2\,...,A_n,...$ etc.\par Let now $\L(L_o)$ be a finite subset of $\Z(L_o)$, let $\L\,=\,\bigcup_{x\in \L(L_o)}Q_{L_o}(x)$ and let $\m^\h$ be the Gibbs state in $\L$ corresponding to the Hamiltonian (1.1) with some fixed boundary condition $\h$ outside $\L$ . Given $\m^\h$ we will consider new measures, denoted by: $$\n_{A},\;\n_{B}^{\t_A},\;\n_{C}^{\t_A,\t_B},\; \n_{D}^{\t_A,\t_B,\t_C}\Eq(3.1)$$ on the finite sets $\O_{A\cap\L}$, $\O_{B\cap\L}$, $\O_{C\cap\L}$, $\O_{D\cap\L}$. Such a measures are defined in analogy with the measures $\n_i^{\t_{i+1}..\t_N}$ of section 2 as follows: \par $\n_{D}^{\t_A,\t_B,\t_C}$ is simply obtained from the Gibbs measure $\m^\h$ by conditioning the spins in ${A\cap\L},\;{B\cap\L},\;{C\cap\L}$ to have the prescribed values $\t_A,\,\t_B,\,\t_C$. To construct $\n_{C}^{\t_A,\t_B}$ we first integrate out the spins $\s_D$ in $\m^\h$ and then we condition to the spins in ${A\cap\L},\;{B\cap\L}$ to have the prescribed values $\t_A,\,\t_B$. Similarly to construct $\n_{B}^{\t_A}$ we first integrate our the spins $\s_D$ and $\s_C$ in $\m^\h$ and then we condition to the spins in ${A\cap\L}$ to have the values $\t_A$. $\n_{A}$ is simply the relativization of $\m^\h$ to $A\cap\L$.\bigskip {\bf Remark} We observe that by construction the intersection between the family of blocks of type A with the set $\L$ consists of a finite collection of blocks say $A_1,\,A_2\,...,A_{N_A}$ and the same for the other families.\par In the notation for the measures $\n_{D}^{\t_A,\t_B,\t_C}$ etc. obtained after the decimation, we have omitted for convenience the supmaxerscript $\h$ .\bigskip We are now in a position to start our calculations. Given an arbitrary function f : $\O_\L\,\to \, R$ we write, following [OP]: $$\m^\h(\log{f})\;=\;\n_{A}(\n_{B}^{\t_A}(\n_{C}^{\t_A,\t_B}( \n_{D}^{\t_A,\t_B,\t_C}(\log{f})..)\Eq(3.1bis)$$ Let us now define $c(L_o)$ to be the largest among the logarithmic Sobolev contant (LSC) of the measures $$\n_{A},\,\n_{B}^{\t_A},\,\n_{C}^{\t_A,\t_B} ,\,\n_{D}^{\t_A,\t_B,\t_C}$$ more precisely : $$c(L_o)\;=\;\sup_{\t_A,\t_B,\t_C}\ max \{c(\n_{A})\,,\, c(\n_{B}^{\t_A})\,,\, c(\n_{C}^{\t_A,\t_B})\,,\, c( \n_{D}^{\t_A,\t_B,\t_C})\}$$ With this notation and if we apply the logarithmic Sobolev inequality to $\n_{D}^{\t_A,\t_B,\t_C}$ we obtain that the r.h.s. of \equ(3.1bis) is bounded above by: $$ c(L_o)\n_{A}(\n_{B}^{\t_A}(\n_{C}^{\t_A,\t_B} (\n_{D}^{\t_A,\t_B,\t_C}(\grad{D}{f})..)\;+\;$$ $$\n_{A}(\n_{B}^{\t_A}(\n_{C}^{\t_A,\t_B} (\logg{\n_{D}^{\t_A,\t_B,\t_C}(f^2)}))\Eq(3.2)$$ Next we define the new functions : $$\eqalign{g_D\,&=\,\n_{D}^{\t_A,\t_B,\t_C}(f^2) ^{1\over 2}\cr g_C\,&=\,\n_{C}^{\t_A,\t_B}(g_{D}^2))^{1\over 2}\cr g_B\,&=\,\n_{B}^{\t_A}(g_{C}^2))^{1\over 2}\cr g_A\,&=\,\n_{A}(g_{B}^2))^{1\over 2}}$$ With these notation, if we iterate \equ(3.2), we obtain : $$\m^\h(\log{f})\;\leq\;$$ $$\leq \;c(L_o)[\m^\h(\grad{A}{g_{B}})\;+\; \m^\h(\grad{B}{g_{C}})\;+\; \m^\h(\grad{C}{g_{D}})\;+\; \m^\h(\grad{D}{f})]\;+\; $$ $$\logg{\m^\h(f^2)}\Eq(3.3)$$ The main idea at this stage is to show that, in the hypotheses of the theorem and provided that the parameter $L_o$ is large enough, each one of the measures $$\n_{A},\,\n_{B}^{\t_A},\,\n_{C}^{\t_A,\t_B} ,\,\n_{D}^{\t_A,\t_B,\t_C}$$ satisfies the conclusions of theorems 1.2 and 2.2 of section 2. Let us state this as a theorem: \bigskip {\bf Theorem 3.1} {\it Let $\{*,\,\t^*\}$ be one of the pairs $\{A,\,\h\}$ $\{B,\,\t_A\}$ $\{C,\,\t_A\t_B\}$ $\{D,\,\t_A\t_B\t_C\}$ and let $\n_{*}^{\t_{*}}$ be the corresponding Gibbs measure. There exists a constant $\bar L$ such that if $SM(L,C,\gamma )$ holds for some $L\,\geq \,\bar L$, then there exists $\bar L_o\;>\;\bar L$ such that if $L_o\,\geq \,\bar L_o$ then there exist two constants $a_o$ and $a_1$ such that :\item{i)} $$c(L_o)\;<\;\infty$$ \item{ii)} $$ \grad{\t(x)}{(\n_{*}^{\t_{*}}(f^2))^{1\over 2}}\;\leq \; a_o\n_{*}^{\t_{*}}(\grad{\t(x)}{f})\;+\;a_1 \n_{*}^{\t_{*}}(\grad{*}{f})$$ for any x $\notin\;*$.}\bigskip Before giving the proof of the above crucial result, let us first complete the proof of the main theorem.\par If we apply ii) of theorem 3.1 to $\m^\h(\grad{A}{g_{B}})$ we get that : $$\m^\h(\grad{A}{g_{B}})\;\leq \;a_o\m^\h(\grad{A}{g_C})\;+\;a_1 \m^\h(\grad{B}{g_C})$$ We have thus succeeded in moving the gradient from the function $g_B$ to the function $g_C$. If we continue this procedure two more times we end up with all the gradients acting on the original function f. More explicitely, after three repeated applications of ii) of theorem 3.1, we have that : $$ \m^\h(\grad{A}{g_{B}})\;\leq \;a_2\m^\h((\nabla_{\L}f)^2)\Eq(3.4)$$ for a suitable constant $a_2$. The same estimate of course applies also to the terms: $$\m^\h(\grad{B}{g_{C}})\quad\hbox{and }\quad \m^\h(\grad{C}{g_{D}})$$. In conclusion we have shown that: $$c(L_o)[\m^\h(\grad{A}{g_{B}})\;+\; \m^\h(\grad{B}{g_{C}})\;+\; \m^\h(\grad{C}{g_{D}})\;+\; \m^\h(\grad{D}{f})]\;\leq $$ $$\leq\;c'(L_o) \m^\h((\nabla_{\L}f)^2)\Eq(3.5)$$ provided that SM(L,C,$\gamma$) holds for some L large enough and the size of the blocks of the decimation was also sufficiently large.\par The theorem (1.1) is proved.\bigskip $\S\;2$\hskip 1cm {\bf Proof of Theorem 3.1}\par The idea of the proof is very simple: we will verify that, in the hypotheses of the theorem, all the assumptions a), b) and c) of section 2 are satisfied for any one of the measures \equ(3.1) with constants $k_o,\;k_1$ uniformly bounded in the side $L_o$ of the blocks of the decimation and with the constant $\e,\;$ going exponentially fast to zero as $L_o\;\to\;\infty$.\par In order to verify a) and b) we first need to write in a convenient way the derivative with respect to a conditioning spin of the effective potential appearing in any of the measures \equ(3.1). One possibility is to use a cluster expansion to write down the effective potential; there is however another way in which the derivative with respect to a conditioning spin of the effective potential becomes essentially a truncated correlation function of a suitable pair of local observables computed with respect to the {\it original} Gibbs measure $\m^\h$. That is of course very convenient since (see Theorem 1.2) it has been proved that in the hypotheses of the theorem the truncated correlations functions of the measure $\m^\h$ decay exponentially fast.\par So let us discuss the second way in a rather general setting.\par Suppose that we are given a subset $\L\;=\;\L_1\,\cup\,\L_2\,\cup\,\L_3$ of the lattice $\Z$ and a Gibbs measure $\m$ on $\{-1,+1\}^{\L}$ with Hamiltonian $H(\s_{\L_1},\,\s_{\L_2},\,\s_{\L_3})$ corresponding to a interaction $\Phi$ with finite norm $$||\Phi||\;=\;\sum_{O\in X}|\Phi (X)|$$ Let $$\hat H(\s_{\L_1},\,\s_{\L_2})\;=\;\hbox{log} (Z_{\L_3}^{\s_{\L_1},\s_{\L_2}})\Eq(3.6)$$ be the effective Hamiltonian after the integration of the $\s_{3}$ variables in $\L_3$ and let $$W_1^{\s_{\L_2}}(\s_{\L_1})\;\equiv \;\hat H(\s_{\L_1},\,\s_{\L_2})\,-\,\hat H(\t_{1},\,\s_{\L_2})\Eq(3.7)$$ be the effective interaction entering in the conditional Gibbs measure of the spins $\s_{\L_1}$ given the spins $\s_{\L_2}$ after the decimation of the spins $\s_{\L_3}$ (see (2.6) ). In \equ(3.7) $\t_1$ is an arbitrary reference configuration of the spins $\s_{\L_1}$ e.g. all the spins up. Then we have:\bigskip {\bf Lemma 3.1}\par {\it For each $x\,\in\,\L_2$ and $y\in \L_1$ there exist two functions $f_x^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})$ and $g_y^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})$ with the following properties: \item{i)} f and g, as functions of the spins $\s_{\L_3}$, have support in a ball centered at x and y respectively with radius equal to the range of the interaction $\Phi$. \item{ii)}$$\sup_{\s_{\L_1},\s_{\L_2},\s_{\L_3}}|f_x^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})| \;\leq\;\exp (2||\Phi||)$$ $$\sup_{\s_{\L_1},\s_{\L_2},\s_{\L_3}}|g_y^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})| \;\leq\;\exp (2||\Phi||)$$ \item{iii)} $$\eqalign{\sup_{\s_{\L_1},\s_{\L_2}}||\partial_{\s_{\L_2}(x)}W_1^{\s_{\L_2}}(\s_{\L_1})| \;&\leq\; \sum_{y\in \L_1}\exp (4||\Phi||) \sup_{\s_{\L_1},\s_{\L_2}}|\media{f_x^{\s_{\L_1},\s_{\L_2}}\,; \,g_y^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}|\cr \;&+\;\sum_{y\in\L_1}\sup_{\s_{\L_1},\s_{\L_2}}|\hbox{\rm log} ({ \media{g_y^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}\over \media{g_y^{\s_{\L_1},\s_{\L_2}^x}}^{\s_{\L_1},\s_{\L_2}}_{\L_3} })|}$$ where $\media{f}^{\s_{\L_1},\s_{\L_2}}_{\L_3}$ is the conditional average of the observable f with respect to the original Gibbs state given that the spins in $\L_1$ and $\L_2$ are equal to $\s_{\L_1}$ and $\s_{\L_2}$ and $\media{f\,;\,g}$ denotes the usual truncated expectation.} \bigskip {\bf Proof}\par Let for $x\,\in\,\L_2$ and $y\in \L_1$ $$\eqalign{f_x^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})\;&=\; \hbox{exp}(H^{\s_{\L_1},\s_{\L_2}^x}(\s_{\L_3})\,-\,H^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3}))\cr g^{\s_{\L_1},\s_{\L_2}}(\s_{\L_3})\;&=\; \hbox{exp}( H^{\s_{\L_1}^y,\s_{\L_2}^x}(\s_{\L_3})\,-\,H^{\s_{\L_1},\s_{\L_2}^x}(\s_{\L_3}))}\Eq(3.7bis)$$ If we use \equ(3.6), \equ(3.7) and the definition of $\partial_{\s(x)}$ we obtain: $$\eqalign{\sup_{\s_{\L_1},\s_{\L_2}}|\partial_{\s_{\L_2}(x)}W_1^{\s_{\L_2}}(\s_{\L_1})| \;&\leq\;\cr 2\sum_{y\in \L_1}\sup_{\s_{\L_1},\s_{\L_2}}|\partial_{\s_{\L_2}(x)} (\hat H(\s_{\L_1},\,\s_{\L_2})\,&-\,\hat H(\s_{\L_1}^y,\,\s_{\L_2}))|\;=\;\cr \sum_{y\in \L_1}\sup_{\s_{\L_1},\s_{\L_2}}|\hbox{\rm log} ( \media{f_x^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}\,&-\, \hbox{\rm log} ({\media{g_y^{\s_{\L_1},\s_{\L_2}}f_x^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}\over \media{g_y^{\s_{\L_1},\s_{\L_2}^x}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}})|}\Eq(3.8)$$ Using $\hbox{\rm log} (1\,+\,x)\;\leq \;x$ if $x\;>\;0$ we get immediately that the r.h.s. of \equ(3.8) is bounded by: $$\sum_{y\in \L_1}\hbox{exp}(4||\Phi||) \sup_{\s_{\L_1},\s_{\L_2}}|\media{f_x^{\s_{\L_1},\s_{\L_2}}\,; \,g_y^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}|\;$$ $$\;+\; \sum_{y\in \L_1}\sup_{\s_{\L_1},\s_{\L_2}}|\hbox{\rm log} ({ \media{g_y^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}\over \media{g_y^{\s_{\L_1},\s_{\L_2}^x}}^{\s_{\L_1},\s_{\L_2}}_{\L_3} })|\Eq(3.9)$$ The Lemma is proved.\bigskip One can now apply the above lemma to each one of the measures \equ(3.1) by conveniently choosing the sets $\L_1,\,\L_2,\,\L_3$. Since the discussion is the same for anyone of the measures \equ(3.1), let us treat in detail only one of them, say $\n_B^{\t_A}$.\par Thus let us suppose that we have fixed a block of type B, say $B_j$, and let us set $$\L_1\;=\;B_j\quad \L_3\;=\;C\,\cup\,D\Eq(3.9bis)$$ and $$\L_2\;=\;A\,\cup\,\cup_{i>j}B_i\Eq(3.9tris)$$ Let also x be a site in $B_i$, with $i\,>\,j$, and let us estimate: $$|\partial_{\s(x)}W^{\s_{\L_2}}_{(1)}(\s)|\Eq(3.10)$$ In order to apply Lemma 3.1, we first observe that if $L_o$ is greater than the range of the interaction then the above defined function $g_y^{\s_{\L_1},\s_{\L_2}}$ does not depend on $\s(x)$ if $y\in B_j$ and $x\in B_i$. Therefore in this case the second term in the r.h.s. of iii) of the Lemma is zero and we get: $$\sup_{\s_{\L_1},\s_{\L_2}}||\partial_{\s(x)}W_1^{\s_{\L_2}}(\s)| \;\leq\; \sum_{y\in B_j}\hbox{exp}(4||\Phi||) \sup_{\s_{\L_1},\s_{\L_2}}|\media{f_x^{\s_{\L_1},\s_{\L_2}}\,; \,g_y^{\s_{\L_1},\s_{\L_2}}}^{\s_{\L_1},\s_{\L_2}}_{\L_3}|\Eq(3.11)$$ Using now theorem 2.1 we see that there exist constants C and $m$, depending only on the norm of the original interaction and on its range $r$, such that for any sufficiently large $L_o$ the r.h.s. of \equ(3.11) is smaller than: $$C\sum_{y\in B_j}\hbox{exp}(-m\,|x\,-\,y|)\Eq(3.12)$$ which implies that: $$\sup_{(\h,\s_A,j,N_B,)} \sum_{i=1}^{j-1}\sup_{x\in B_i}\supnorm{\partial_{\s (x)}W_{(1)}^{\s_{\L_2}}(\s)}\;\leq\;\e \Eq(3.13)$$ with e.g. $\e\;=\;\hbox{exp}(-{m\over 2}\,L_o)$ for any sufficiently large $L_o$.\par In a very similar way one checks the bounds (2.7) and (2.8).\bigskip {\bf Remark} The conclusion of the above discussion is that the effective potential between any two sites x and y, defined for example as $\partial_{\s (x)}\partial_{\s (y)}\hat H$ with $\hat H$ the effective hamiltonian of anyone of the measures \equ(3.1), decays exponentially fast in the distance between x and y. This implies in particular that in the second inequality in b) of section 2 we can replace $\sup_i|B_i|\,=\,L_o^d$ with $L_o^{d-1}$.\bigskip In order to complete the proof of theorem 3.1 we are left with the problem of verifying assumption c) of section 2. The idea at this point is to show that assumption c) is equivalent to a {\it lower } bound on the gap of the generator of the Glauber dynamics defined in section 1, reversible with respect to $\nu_*^{\t_*}$. In turn, such a lower bound will follow from theorem 1.2 i) . As before we do the computations only for $\nu_B^{\t_A}$ the other cases being analogous.\par We keep the notation \equ(3.9bis), \equ(3.9tris) and we denote by $\nu_1$ the conditional distribution on $\O_{\L_1}$ of the relativization to $\L_1\cup\L_2$ of the original Gibbs measure $\mu^\h$. For simplicity, in $\nu_1$, we have omitted to specify the boundary conditions $\t_1$ since all our estimates will hold uniformly in $\t_1$. First of all we observe that, from assumption H5 on the jump rates and Theorem 1.2 i), it follows, for any function $f\in L^2(\Omega_{\L_1},d\nu_1 )$ with $\nu_1(f)\,=\,\mu^\h(f)\,=\,0$, that: $$\nu_1(f^2)\,=\,\mu^\h(f^2)\,\leq\,{k_2\over m_o}\mu^\h((\nabla_{\L_1}f)^2)\,=\,{k_2\over m_o}\nu_1((\nabla_{\L_1}f)^2)\Eq(3.15)$$ Therefore we get immediately that: $$\sum_{\s_{\L_1},\hat\s_{\L_1}} \n_1(\s_{\L_1}) \n_1(\hat\s_{\L_1}) [f(\s_{\L_1})\,-\,f(\hat\s_{\L_1})]^2\;\leq$$ $$\leq {2k_2\over m_o}\nu_1((\nabla_{\L_1}f)^2) \Eq(3.16)$$ i.e. assumption c) holds true with the constant $k_1\,=\, ({2k_2\over m_o})^{1\over 2}$.\par The Theorem is proved.\vskip 2cm %\input formato.tex \numsec=4\numfor=1 \tolerance=10000 \vskip 1cm {\bf Section 4}\par\noindent \centerline{\bf An Application to a Non-Ferromagnetic Model} \bigskip We conclude this paper by considering a non-ferromagnetic model in two dimensions, obtained by adding a small antiferromagnetic next nearest-neighbor coupling to the standard ferromagnetic Ising model with small positive external field.\par We will show that if the antiferromagnetic n.n.n. coupling is small with respect to the n.n. ferromagnetic one, then it is possible to find constants $C$ and $\gamma$ such that for all large enough $L$ and all low enough temperatures, depending on $Lo$, $C$ and $\gamma$, the system satisfies our mixing condition $SM(\L,C,\gamma)$. It is very likely (see e.g. the examples in section 1 of [MO2]) that the system, in the same range of the parameters, does not satisfies the condition of [SZ3].\par If $\L$ denotes the square of side $L$ ($L$ odd) centered at the origin of $\bf Z^2$, then our Hamiltonian reads as follows: $$ ({-1\over \beta})H(\sigma)=-{J\over 2}\sum_{\subset \Lambda}\sigma(x)\sigma(y)+{K\over 2}\sum_{<\!\!\!\!>\subset\Lambda} \sigma(x)\sigma(y)-{h\over 2}\sum_{x\in\Lambda}\sigma(x) +\,\hbox{b.c.} \Eq(4.1) $$ where $\sum_{\subset \L}$ runs over the nearest neighbors pairs in $\L$, $\sum_{<\!\!\!\!>\subset \L}$ runs over the next nearest neighbors pairs in $\L$ and b.c. contains the interaction with the boundary configuration $\t$. Notice that, in order to follow our convention (see section 1), we have inserted the factor $-\beta$ directly into the Hamiltonian.\bigskip {\bf Theorem 4.1}\bigskip {\it There exist $C$, $\gamma$, $\bar L$ such that for any $0\,\leq \,K\,<\,{J\over 4}$ and for any $L\,\geq \,\bar L$, $L$ odd, there exists $\beta_o$ such that for any $\beta \,\geq \,\beta_o$ the mixing condition $SM(\L,C,\gamma )$ holds.}\bigskip {\bf Proof} \par Let us denote by dist' the following distance on $\bf Z^2$: $$\hbox{dist'}(x,y)\;=\;\max_i|x_i\,-\,y_i|\quad x\,,\,y\;\in \Z$$ Let $$ {\bar {\cal C}} = \{ x \in \L ;\; \hbox {dist'} (x,\hbox {corners of } \L) \geq l_o\} $$ with $$ l_o = [2(J-2K) +8K] / h $$ Thus ${\bar {\cal C}}$ looks like a cross. The theorem will immediately follow (see also section 5 of [MO2]) if we can show that for any $L$ large enough, any boundary configuration $\t$ and any $0\,\leq \,K\,<\,{J\over 4}$ the ground state of $H_{\L}^\t(\s)$ is equal to $+1$ at all the sites $x$ in ${\bar {\cal C}}$ . Let in fact $C$ and $\gamma$ be such that: $$C\exp (-\gamma \sqrt 2 l_o)\,\geq\, 1\Eq(4.2)$$ If, for any boundary condition $\t$, the ground state has the structure described above, then, because of the screening effect of the plus spins in $ \bar {\cal C}$ the ground states in each connected component (square) $Q_i , \, i=1,\dots,4$ of $\L \setminus \bar {\cal C} $ is only affected by a change of a boundary spin $\t _y,\; y \in \partial^+_{\sqrt 2} Q _i \cap \partial^+_{\sqrt 2} \Lambda$. Therefore we can always estimate the quantity $$ \sup_{\tau,\tau^{(y)} \in \Omega _{\Lambda_o^c}} Var(\mu_{\Lambda_o , \Delta}^\tau\ \ \mu_{\Lambda_o , \Delta}^{\tau^{(y)}}), \quad \Delta\subset \,\L \Eq(4.3)$$ appearing in $SM(\L,C,\gamma )$ by $1$ when for some $i=1,\, \dots , 4$ $$ y \in \partial^+_{\sqrt 2} Q_i \cap \partial^+_{\sqrt 2} \Lambda ,\quad \D \cap Q_i \not= \emptyset$$ or by $2\mu_{\L}^\t(\s_x \neq \,+1 \hbox{ for some }x\in {\bar {\cal C}}\;)$ otherwise. In both cases we get an estimate smaller than: $$C\exp (-\gamma \hbox { dist} (\Delta ,y))$$ for large enough $\beta$ because of our choice of $C$ and $\gamma$ and the fact that $$\lim_{\beta \to \infty}\,2\mu_{\L}^\t(\s \neq \,\hbox{ ground state}\;)\;=\;0\Eq(4.4)$$ In order to prove the above structure of the set of all the ground states we will use the following "rules" :\bigskip \item{i)} Let $N^-$ be the number of minus spins in a ground state configuration. Then: $$N^-\;\leq \; {4(J+2K)L\,+\,16K\over h}$$ \item{ii)} In any ground state configuration there exists no horizontal or vertical segment of minus spins ( thought as a thin rctangle) surronded on two adjacent sides (one of which a "long" one) by plus spins and with at least a plus spin along a third side. \item{iii)} In any ground state configuration there exists no Peierl's contour with a horizontal or vertical segment of length $l\,\geq\, l_o$ . \item{iv)} In any ground state configuration if there exists a Peierl's contour $\gamma$ with a right angle at the site $(x_1+{1\over 2},x_2+{1\over 2})$ of the dual lattice, $x\equiv (x_1,x_2)\,\in \,\L$, such that the plus spins lay along the exterior of the angle then, starting from $(x_1+{1\over 2},x_2+{1\over 2})$, the contour $\gamma$ has to reach the vertical and horizontal boundary of $ \L$ without bending.\bigskip Rules i), ii), iii) are easily verified by simple energy arguments if $4K\,<\,J$. Rule iv) is slightly more complicate. The proof goes as follows.\par Without loss of generality let us suppose that the angle has the plus spins at its right and bottom and let us suppose that the contour $\gamma$ has another right angle at the site $(x_1\,-\,n\,+{1\over 2},x_2\,+\,{1\over 2})$ with $x_1\,-\,n\, \geq \,-{(L-1)\over 2}$. Using ii) the contour $\gamma $ at the new angle can only bend down; moreover, again by ii), the minus spins above $\gamma$ at the sites $(x_1-j,x_2+1)\;j=1\dots n$ have to be surrounded from the left and from above by minus spins. It is easy to see that if the spin at the site $(x_1+1,x_2+2)$ is minus then it is energetically convenient to flip to plus one all the minus spins at the sites $(x_1-j,x_2+1)\;j=1\dots n$ irrespectively of the value of the spin at the site $(x_1-n,x_2+2)$, and the same if the spin at $(x_1-n,x_2+2)$ is minus. If the spins at $(x_1-n,x_2+2)$ and at $(x_1+1,x_2+2)$ are both plus then it is energetically convenient to flip to plus all the minus spins at the sites $(x_1-j,x_2+1)\;j=1\dots n$, $(x_1-j,x_2+2)\;j=1\dots n$. In any case the original configuration was certainly not a ground state. Similar arguments cover all the other situations.\par It is easy to show, at this point, that , for every $\t$, the structure of the ground state is the one depicted above. \par If $L$ is taken large enough, $$ L^2 > 4(3l_o)^2 [(4J+2K)L +16K]/h , $$ then, using i), for any ground state it is possible to find a square $Q^*$ of side $l\,=\,3l_o$ completely filled up with pluses and strictly contained in ${ \bar {\cal C}}$ . Using ii), iii), we know that on top of each face of $Q^*$ there exists a segment of length $l_1$ larger than $l_o$ of plus spins. If $l_1 < l$ then there is at least one right angle with exterior + spins at one end of the concerned segment. If not, all the spins adjacent from the exterior to that face of $Q^*$ are plus and we can repeat the argument. Then, continuing in this way, starting from any face of $Q^*$, either we get to $\partial \L$ on a parallel segment of equal length $l$ or, at a given step, we find a right angle inside $\L$. Using iv), by further decreasing the energy, we obtain a configuration containing a cross ${\cal C}$ of plus spins centered inside $Q^*$ with width at least $l_o$ . The complement of ${\cal C}$ in $\L$ splits into four disjoint rectangles $R_1, R_2,R_3,R_4$. Each $R_i$, by construction, contains a square $ Q_i$ of side $l_o$ having a vertex coinciding with one of the four vertices of $\L$. By applying to the faces of $R_i$, internal to $\L$, a construction similar to the one leading to ${\cal C}$ it is easy to show that it is energetically convenient to fill of pluses the sets $R_i \setminus Q_i$ so that we end up with a configuration where $\bar {\cal C}$ is full of pluses. %\input formato.tex \numsec=1\numfor=1 \tolerance=10000 \vskip 1cm {\bf Appendix}\par\noindent \bigskip We prove formulae a)$\dots$ f) given at the beginning of the proof of proposition 1.2.\par The first three ones, a) b) c), are trivially verified. In order to derive d) we assume, without loss of generality, that $\t_x\,=\,+1$ and we let $g_i^x\,=\,{d\n_i^{\h ,\t_{i+1}^{x,-1},., \t_N^{x,-1}}\over d\n_i^{\h ,\t_{i+1}^{x,+1},.\t_N^{x,+1}}}$. Then we write: $$\vert\partial_{\t_x}[ \n_i^{\h ,\t_{i+1},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(f)..)]\vert\;=\;$$ $$\vert \n_i^{\h ,\t_{i+1}^{x,-1},\dots , \t_N^{x,-1}}(....(\n_1^{\h ,\t_2^{x,-1},\dots ,\t_N^{x,-1}}(\partial_{\t_x} f)..)\;-\; [\partial_{\t_x}( \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)\vert\;=$$ $$= \;\vert \n_i^{\h ,\t_{i+1},\dots , \t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}(\prod_{j=i}^1g_j^x\partial_{\t_x} f)..)\;-\; [\partial_{\t_x}( \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)|\;\leq$$ $$\leq \;|[\partial_{\t_x}( \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)|\;+$$ $$\prod_{j=i}^1\l^x_j \n_i^{\h ,\t_{i+1},\dots ,\t_N}(....(\n_1^{\h ,\t_2,\dots ,\t_N} (|\partial_{\t_x}f|)..) $$ where $|[\partial_{\t_x} \n_i^{\h ,\t_{i+1},\dots ,\t_N} (....(\n_1^{\h ,\t_2,\dots ,\t_N}](f)|$ denotes the expression: $${|\n_i^{\h ,\t_{i+1}^{x,+1},.,\t_N^{x,+1}} (.(\n_1^{\h ,\t_2^{x,+1},.,\t_N^{x,+1}}(f(\s\t)..)\,-\, \n_i^{\h ,\t_{i+1}^{x,-1},.,\t_N^{x,-1}} (.(\n_1^{\h ,\t_2^{x,-1},.,\t_N^{x,-1}}(f(\s\t)..)|\over 2}\Eqa(a1.1)$$ Clearly \equ(a1.1) proves d).\par The "Leibniz rule" e) follows by essentially the same argument.\par In order to prove f) we follow Zegarlinski [Z1] and we write: $$4|[\partial_{\t_x}\n_i^{\h ,\t_{i+1},\dots ,\t_N}](f^2)|$$ as : $$2|\{\n_i^{\h ,\t_{i+1},\dots ,\t_N}(f^2)\,-\, \n_i^{\h ,\t_{i+1},\dots ,\t_N}(g_i^xf^2)\}|\,=\,$$ $$|\n_i^{\h ,\t_{i+1},\dots ,\t_N}\times \n_i^{\h ,\t_{i+1},\dots ,\t_N}([f^2(\sigma_{\L_i})\,-\,f^2(\tilde \sigma_{\L_i})] [g_i^x(\sigma_{\L_i})\,-\,g_i^x(\tilde \sigma_{\L_i})])|\Eqa(a1.2)$$ where $\sigma_{\L_i}$ and $\tilde \sigma_{\L_i}$ are two independent replicas in $\Omega_{\L_i}$.\par If we assume, without loss of generality, that $g_i^x(\sigma_{\L_i})\,\geq\,g_i^x(\tilde \sigma_{\L_i})$ then, from the definition of $g_i^x$, we get: $$g_i^x(\sigma_{\L_i})\,-\,g_i^x(\tilde \sigma_{\L_i})\;\leq \; g_i^x(\sigma_{\L_i})4|\partial_{\t_x}W_{(i)}^\h(\s\t )|_\infty \Eqa(a1.3)$$ The Schwartz inequality gives that: $$|\n_i^{\h ,\t_{i+1},\dots ,\t_N}\times \n_i^{\h ,\t_{i+1},\dots ,\t_N}([f^2(\sigma_{\L_i})\,-\,f^2(\tilde \sigma_{\L_i})])|\;\leq $$ $$\leq \;(\n_i^{\h ,\t_{i+1},\dots ,\t_N}\times \n_i^{\h ,\t_{i+1},\dots ,\t_N}([f(\sigma_{\L_i})\,-\,f(\tilde \sigma_{\L_i})]^2))^{1\over 2}\,2 (\n_i^{\h ,\t_{i+1},\dots ,\t_N}(f^2))^{1\over 2}\Eqa(a1.4)$$ which, in turn, implies, using (2.11) and \equ(a1.3) above, that $$4|\partial_{\t_x}\n_i^{\h ,\t_{i+1},\dots ,\t_N}](f^2)|\;\leq $$ $$\leq \;8k_1\l^x_i\supnorm{\partial_{\t(x)}W_{(i)}^\h(\s\t)} [\n_i^{\h ,\t_{i+1},\dots ,\t_N}(f^2)]^{1\over 2} [\n_i^{\h ,\t_{i+1},\dots ,\t_{N}}(\grad{i}{f})]^{1\over 2}$$ i.e. the inequality f).\vskip 2cm %\input formato.tex \def\refj#1#2#3#4#5#6#7{\parindent 2.2em \item{[{\bf #1}]}{\rm #2,} {\it #3\/} {\rm #4} {\bf #5} {\rm #6} {(\rm #7)}} \tolerance=10000 \vskip 1cm {\bf Acknowledgments}\bigskip\noindent We would like to thank B. Zegarlinski for some useful discussions concerning the differences between the results and the methods of the present paper and those developed in his joint work with D. Stroock; we are grateful to H.T.Yau for informing and explaining to us his work prior to publication. \bigskip \centerline{\bf References}\bigskip\noindent \refj{D1}{ R.L. Dobrushin}{Prescribing a System of random variables by Conditional Distributions}{Theory of Prob. Appl.}{15}{453-486}{(1970)} \refj{DS1}{ R.L. Dobrushin, S.Shlosman}{Completely Analytical Gibbs Fields}{Stat.Phys. and Dyn. Syst. Birkhauser}{}{371-403}{1985} \refj{DS2}{ R.L. Dobrushin, S. Shlosman}{Completey Analytical InteractionsConstructive description}{Journ. Stat. Phys.}{46}{983-1014}{1987} \refj{EFS}{A.v. Enter,R. Fernandez, A. Sokal}{Regularity Properties and Pathologies of Position-Space Renormalization-Group Transformations}{to appear in Journ. Stat. Phys.}{}{}{1993} \refj{G1}{L.Gross}{Logarithmic Sobolev inequalities}{Am.J.Math.}{97}{553-586}{1979} \refj{G2}{L.Gross}{Logarithmic Sobolev Inequalities and Contractivity Properties of Semigroups}{Cornell Math. Dept. Notes}{}{}{1992} \refj{L}{T.Ligget}{Interacting Particle Systems}{Springer-Verlag}{}{}{1985} \refj{LY}{S.L. Lu, H.T.Yau}{Spectral Gap and Logarithmic Sobolev Inequality for Kawasaki and Glauber Dynamics}{preprint} {}{}{1992} \refj{MO1}{ F.Martinelli, E.Olivieri}{Finite Volume Mixing Conditions for Lattice Spin Systems and Exponential Approach to Equilibrium of Glauber Dynamics}{CARR 24}{}{Les Houches}{1992} \refj{MO2}{ F.Martinelli, E.Olivieri}{Approach to Equilibrium of Glauber Dynamics in the One Phase Region I: The Attractive Case}{CARR 25}{}{}{1992} \refj{O}{E.Olivieri}{On a Cluster Expansion for Lattice Spin Systems a Finite Size Condition for the Convergence}{Journ. Stat. Phys.}{50}{1179-1200}{1988} \refj{OP}{ E.Olivieri, P.Picco}{Cluster Expansion for D-Dimensional lattice Systems and Finite Volume Factorization Properties}{Journ. Stat. Phys.}{59}{221-256}{1990} \refj{S}{D.W.Stroock} {Logarithmic Sobolev Inequalities for Gibbs States}{M.I.T.}{}{}{1992} \refj{SZ1}{ D.W.Stroock, B. Zegarlinski}{The equivalence of the logarithmic Sobolev inequality and the Dobrushin-Shlosman mixing condition}{Comm. Math.Phys.}{144}{303-323}{1992} \refj{SZ2}{ D.W.Stroock, B. Zegarlinski}{The logarithmic Sobolev inequality for continuous spin systems on a lattice}{J. Funct. Anal.}{104}{299-326}{1992} \refj{SZ3}{ D.W.Stroock, B. Zegarlinski}{The Logarithmic Sobolev Inequality for Discrete Spin Systems on a Lattice}{Comm. Math.Phys.}{149}{175-194}{1992} \refj{Z1}{B. Zegarlinski}{On log-Sobolev inequalities for infinite lattice systems}{Lett. Math. Phys.}{20}{173-182}{1990} \refj{Z2}{B. Zegarlinski}{Log-Sobolev inequalities for infinite one-dimensional lattice systems}{Comm. Math.Phys.}{133}{147-162}{1990} \refj{Z3}{B. Zegarlinski}{Dobrushin uniqueness theorem and the logarithmic Sobolev inequalities}{Journal of Funct. Anal.}{105}{77-111}{1992} \end