This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Thermodynamic formalism for amenable groups and countable state spaces

Elmer R. Beltrán Universidad Nacional Jorge Basadre Grohmamn, Tacna, Perú; Universidad Católica del Norte, Departamento de Matemáticas, Antofagasta, Chile [email protected] Rodrigo Bissacot Faculty of Mathematics and Computer Science, Nicolaus Copernicus University, Poland; Institute of Mathematics and Statistics, University of São Paulo, Brazil [email protected] Luísa Borsato Institute of Mathematics and Statistics, University of São Paulo, Brazil [email protected]  and  Raimundo Briceño Facultad de Matemáticas, Pontificia Universidad Católica de Chile, Santiago, Chile [email protected]
Abstract.

Given the full shift over a countable state space on a countable amenable group, we develop its thermodynamic formalism. First, we introduce the concept of pressure and, using tiling techniques, prove its existence and further properties such as an infimum rule. Next, we extend the definitions of different notions of Gibbs measures and prove their existence and equivalence, given some regularity and normalization criteria on the potential. Finally, we provide a family of potentials that non-trivially satisfy the conditions for having this equivalence and a non-empty range of inverse temperatures where uniqueness holds.

Key words and phrases:
Gibbs measure; amenable group; pressure; countable state space; thermodynamic formalism.
2010 Mathematics Subject Classification:
Primary 37D35, 82B05, 37A35; secondary 37B10, 82B20, 60B15.

1. Introduction

There are two general ways to describe a system composed of many particles: microscopically and macroscopically. The first one makes use of the exact positions of the particles, as well as their local interactions. The second one, in turn, is usually outlined by thermodynamic quantities such as energy and entropy. One could say that statistical mechanics — originated from the works of Boltzmann [10] and Gibbs [32] — is the bridge between the microscopic and the macroscopic descriptions of this kind of systems. In this connection, Gibbs measures are a central object.

It is fair to say that Gibbs measures are at the core of the “conceptual basis of equilibrium statistical mechanics” [52]. Relevant examples are the Ising model, which tries to capture the magnetic properties of certain materials; the hard-core model, that describes the distribution of gas particles in a given environment; among many others [29, 30, 31]. In these cases it is customary to consider that the many particles interacting are infinite, take a value from a state space AA (also called alphabet when AA is countable), and they are disposed in a crystalline structure. This structure and its symmetries are usually represented by a countable group GG, possibly with some Cayley graph associated with it. A particular case is the hypercubic dd-dimensional lattice, which can be understood as the Cayley graph of the finitely generated abelian group G=dG=\mathbb{Z}^{d} according to its canonical generators. Then, it is natural to represent an arrangement of particles as an element of the space of configurations X=AGX=A^{G}, the GG-full shift. Considering this, one is interested in certain measures μ\mu in the space (X)\mathcal{M}(X) of Borel probability measures supported on XX. More specifically, the measures of interest are the ones that describe these kind of systems when they are in thermal equilibrium, where the energy of configurations is given by some potential ϕ:X\phi:X\to\mathbb{R}. However, there are many mathematically consistent ways to represent that situation by choosing an appropriate measure μ(X)\mu\in\mathcal{M}(X) and, as the theory evolved, it drew the attention from different areas of expertise such as probability [51, 27] and ergodic theory [56, 14]. Consequently, the very concept of Gibbs measure started to develop in more abstract and not always equivalent directions.

We focus mainly on four conceptualizations of the idea of thermal equilibrium, namely, DLR, conformal, Bowen-Gibbs, and equilibrium measures. We now proceed to briefly describe each of them.

Dating back to the 60’s, Dobrushin [22, 23] and, independently, Lanford and Ruelle [41] proposed a concept of Gibbs measure that extended the usual Boltzmann-Gibbs formalism to the infinite particles setting. Roughly, the idea involved looking for probability distributions compatible with a family of maps — sometimes called specification — that prescribe conditional distributions inside finite subsets of GG given some fixed configuration outside. More specifically, given a collection γ=(γK)K(G)\gamma=(\gamma_{K})_{K\in\mathcal{F}(G)} of probability kernels γK:×X[0,1]\gamma_{K}\colon\mathcal{B}\times X\to[0,1], with (G)\mathcal{F}(G) the set of finite subsets of GG and \mathcal{B} the Borel σ\sigma-algebra, one is interested in finding measures μ(X)\mu\in\mathcal{M}(X) such that μγK=μ\mu\gamma_{K}=\mu for every K(G)K\in\mathcal{F}(G), where μγK\mu\gamma_{K} is a new measure (a priori, different from μ\mu) obtained from μ\mu via γK\gamma_{K}. Those distributions are called DLR measures after the above cited authors and they have received considerable attention from both mathematical physicists and probabilists (see, for example, [30, 31, 38, 52]).

Another rather classical way to define a Gibbs measure, which does not involve conditional distributions, was introduced by Capocaccia in [17]. Given a class \mathcal{E} of local homeomorphisms τ:XX\tau:X\to X and a potential ϕ:X\phi:X\to\mathbb{R}, one is interested in measures μ\mu such that d(μτ1)dμ=exp(ϕτ)\frac{d(\mu\circ\tau^{-1})}{d\mu}=\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau}) for every τ\tau\in\mathcal{E}, where ϕτ:X\phi_{*}^{\tau}:X\to\mathbb{R} is a function representing the energy difference between a configuration xx and τ(x)\tau(x) (e.g., see [38, Definition 5.2.1]). This kind of measures fits in the more general context of (Ψ,)(\Psi,\mathcal{R})-conformal measures explored in [1], where \mathcal{R} is a Borel equivalence relation and Ψ:+\Psi\colon\mathcal{R}\to\mathbb{R}_{+} is a measurable function. Then, Capocaccia’s measures, that we simply call conformal measures, can be recovered by taking a function Ψ\Psi related to the given potential and \mathcal{R} the tail relation in the space of configurations. By considering other particular Borel relations \mathcal{R} and measurable functions Ψ\Psi, one can recover other relevant notions of conformal measures, such as the ones presented in [20, 49, 53], that are mainly adapted to the one-dimensional setting, i.e., when G=G=\mathbb{Z} or, considering also semigroups, when G=G=\mathbb{N}.

A third possibility, introduced by Rufus Bowen in a one-dimensional and ergodic theoretical context [14], is to define Gibbs measures by specifying bounds for the probability of cylindrical events. More concretely, one is interested in the measures μ(X)\mu\in\mathcal{M}(X) for which there exists constants C>0C>0 and pp\in\mathbb{R} such that

C1μ([a0a1an1])exp(i=0n1ϕ(Tix)pn)C for xX.C^{-1}\leq\frac{\mu([a_{0}a_{1}\cdots a_{n-1}])}{\mathop{\textrm{\rm exp}}\nolimits(\sum_{i=0}^{n-1}\phi(T^{i}x)-pn)}\leq C\qquad\text{ for }x\in X.

As in [7], we call those measures Bowen-Gibbs measures to avoid confusion. This definition has been considered in the literature [18, 36, 38, 52] and also relaxed versions of it, such as the so-called weak Gibbs measures [58, 60], where the constant CC is replaced by a function that grows sublinearly in nn. This and further relaxations have also played a relevant role in the multi-dimensional case, this is to say, when G=dG=\mathbb{Z}^{d} and d>1d>1, for finite state spaces (e.g., see [38, Theorem 5.2.4]).

The last important definition considered in this work is the one of equilibrium measure. When XX is a finite configuration space, equilibrium measures are simply probability vectors that maximize the sum (or difference) of an entropy- and an energy-like quantity, that is, a quantity like

H(p)+p(ϕ(x1),,ϕ(xk))=i=1kpilogpi+i=1kpiϕ(xi),H(p)+p\cdot(\phi(x_{1}),\dots,\phi(x_{k}))=-\sum_{i=1}^{k}p_{i}\log p_{i}+\sum_{i=1}^{k}p_{i}\phi(x_{i}),

where k=|X|k=|X|, xiXx_{i}\in X, ϕ:X\phi:X\to\mathbb{R} is a potential, p=(p1,,pk)p=(p_{1},\dots,p_{k}) is a probability vector with pip_{i} the probability associated with xix_{i}, and H(p)H(p) is the Shannon entropy of pp. These measures were considered, for example, in [31, 38, 52]. On the other hand, when XX is an infinite configuration space and there is a robust notion of specific entropy, let’s say h(μ)h(\mu), we are interested in studying measures μ(X)\mu\in\mathcal{M}(X) that maximize the quantity h(μ)+ϕ𝑑μh(\mu)+\int\phi\,d\mu for a continuous potential ϕ:X\phi:X\to\mathbb{R}. This notion tries to capture the macroscopic behaviour of the system without making any assumption of the microscopic structure.

The problem of proving equivalences among these and other related notions has already been studied in different settings. We mention some relevant results that can be found in the literature.

In the one-dimensional case, for finite state spaces, Meyerovitch [44] proved the equivalence between conformal measures and DLR measures for some families of proper subshifts. Also, Sarig [54, Theorem 3.6] proved that any DLR measure on a mixing subshift of finite type is a conformal measure, for a different but related notion of conformal, restricted to the one-dimensional setting. In the same work, for one-sided and countably infinite state spaces, Sarig [54, Proposition 2.2] proved that conformal measures — according to his definition — are DLR measures for topological Markov shifts. In this same setting, Mauldin and Urbański [43] proved the existence of equilibrium measures and that any equilibrium measure satisfies a Bowen-Gibbs equation. Moreover, if the topological Markov shift satisfies the BIP property and the potential has summable variation, Beltrán, Bissacot, and Endo [6] proved that DLR measures and conformal measures — in the same sense as Sarig — are equivalent. Finally, for potentials with summable variation on sofic subshifts, Borsato and MacDonald [12] proved the equivalence between DLR and equilibrium measures. There are also other classes of measures in the one-dimensional case which we do not treat here, such as gg-measures [37, 59] and eigenmeasures associated with the Ruelle operator [14, 52]. When the state space is finite, it is known that the set of DLR measures and gg-measures do not contain each other [28, 9], but there is a characterization for when a gg-measure is a DLR measure [7]. In addition, eigenmeasures coincide with DLR measures for continuous potentials in the one-sided setting, as proven by Cioletti, Lopes, and Stadlbauer in [19]. Pioneering works in the one-dimensional countably infinite state space setting can be found in [33, 34].

In the multi-dimensional case, some results regarding the equivalences among the four notions of Gibbs measures have been proved for finite state spaces. A first important reference is Keller [38, Theorem 5.2.4 and Theorem 5.3.1], where it is proven that when ϕ:X\phi:X\to\mathbb{R} is regular (which includes the case of local and Hölder potentials, and well-behaved interactions), then the four definitions are equivalent. Here, by regular, we mean that

n=1nd1δn(ϕ)<,\sum_{n=1}^{\infty}n^{d-1}\delta_{n}(\phi)<\infty,

where δn(ϕ)\delta_{n}(\phi) is the oscillation of ϕ\phi when considering configurations that coincide in a specific finite box, namely, [n,n]dd[-n,n]^{d}\cap\mathbb{Z}^{d}. Other classical references in this setting are due to Dobrushin [21] and Lanford and Ruelle [41], which, combined, establish the equivalence between DLR measures and equilibrium measures for a general class of subshifts of finite type. Kimura [40] generalized the equivalence between DLR and conformal measures for subshifts of finite type, and some of the implications are true for more general proper subshifts. In the countably infinite state space setting, Muir [45, 46] obtained all equivalences for the GG-full shift when G=dG=\mathbb{Z}^{d}. In order to do this, it was required that the potential ϕ:X\phi:X\to\mathbb{R} is regular and satisfies a normalization criterion, namely, exp-summability:

aexp(supϕ([a]))<.\sum_{a\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([a])\right)<\infty.

This last condition is automatically satisfied when AA is finite.

Results proving equivalences between different kinds of Gibbs measures go beyond the amenable [55, 5, 2, 15] and even the symbolic setting to general topological dynamical systems [3, 36].

One of our main contributions is to exhibit conditions to guarantee that the four notions of Gibbs measures presented above are equivalent, when considering the state space A=A=\mathbb{N} and an arbitrary countable amenable group GG, thus extending Muir’s methods to the more general amenable case. Countable amenable groups play a fundamental role in ergodic theory [48] and include many relevant classes of groups, such as abelian (so, in particular, G=dG=\mathbb{Z}^{d}), nilpotent, and solvable groups and are closed under many natural operations, namely, products, extensions, etc. (e.g., see [42]). In the more general group and finite state space setting, the equivalence between DLR and conformal measures was extended to general subshifts over a countable discrete group GG with a special growth property by Borsato-MacDonald [13, Theorems 5 and Theorem 6]. Recently, a different proof for the equivalence between DLR and conformal measures for any proper subshift was given by Pfister in [50]. Also, in [4], a Dobrushin-Lanford-Ruelle type theorem is proven in the case that the group is amenable and a topological Markov property holds, which is satisfied, in particular, by subshifts of finite type. Here, as Muir, we focus on the GG-full shift case. We consider the configuration space X=GX=\mathbb{N}^{G}, for GG an arbitrary countable amenable group, and an exp-summable potential ϕ:X\phi:X\to\mathbb{R} with summable variation (according to some exhausting sequence). The concept of summable variation extends the one of regular potential presented before. More precisely, a potential ϕ\phi has summable variation if

m=1|Em+11Em1|δEm(ϕ)<,\sum_{m=1}^{\infty}\left|E_{m+1}^{-1}\setminus E_{m}^{-1}\right|\delta_{E_{m}}(\phi)<\infty,

where {Em}m\{E_{m}\}_{m} is an exhausting sequence for GG and δEm(ϕ)\delta_{E_{m}}(\phi) is a standard generalization of δm(ϕ)\delta_{m}(\phi).

The paper is organized as follows. First, in Section 2, we present some preliminary notions about amenable groups GG, the corresponding symbolic space G\mathbb{N}^{G}, and potentials. Later, in Section 3, we introduce the concept of pressure in our framework and we prove its existence. Also, we prove that it satisfies an infimum rule and that it can be obtained as the supremum of the pressures associated with finite alphabet subsystems. In order to achieve this, we use relatively new techniques for tilings of amenable groups [26] and, inspired by ideas for entropy from [25], we develop a generalization of Shearer’s inequality for pressure. In Section 4, we introduce spaces of permutations and Gibbsian specifications in order to pave the way for the definitions of conformal and DLR measures, respectively. Next, in Section 5, we prove the equivalence between the four notions of Gibbs measures mentioned above given some conditions on the potential, such as exp-summability and summable variation. We also prove related results involving equilibrium measures. In order to prove the equivalence between DLR and conformal measures we rely on the strategies presented on [45] for the G=dG=\mathbb{Z}^{d} case, which already considers an infinite state space. Moreover, using Prokhorov’s Theorem and relying on the existence of conformal measures in the compact setting [20], we prove the existence of a conformal (and DLR) measure in our context. We also prove that DLR measures are Bowen-Gibbs. If it is also the case that the measure is invariant under shift actions of the group, we prove that any Bowen-Gibbs measure is an equilibrium measure and that any equilibrium measure is a DLR measure. At last, in Section 6, we show how to recover previous results from ours and, inspired by the Potts model and considering a version of it with countably many states, we exhibit a family of examples for which all our results apply non-trivially and, in addition, a version of Dobrushin’s Uniqueness Theorem adapted to our setting holds, thus providing a regime where the uniqueness of a Gibbs measure is satisfied.

2. Preliminaries

2.1. Amenable groups and the space G\mathbb{N}^{G}

Let GG be a countable discrete group with identity element 1G1_{G} and \mathbb{N} be the set of non-negative integers. Consider the GG-full shift over \mathbb{N}, that is, the set G={x:G}\mathbb{N}^{G}=\{x:G\to\mathbb{N}\} of \mathbb{N}-colorings of GG, endowed with the product topology. We abbreviate the set G\mathbb{N}^{G} simply by XX. Given a set AA, denote by (A)\mathcal{F}(A) the set of nonempty finite subsets of AA.

Consider a sequence {Em}m\{E_{m}\}_{m} of finite sets of GG such that E0=E_{0}=\emptyset, 1GE11_{G}\in E_{1}, EmEm+1E_{m}\subseteq E_{m+1} for all mm\in\mathbb{N}, and mEm=G\bigcup_{m\in\mathbb{N}}E_{m}=G. We will call such a sequence an exhaustion of GG or an exhausting sequence for GG. Throughout this paper, we will consider a particular type of exhausting sequences: we will assume further that E1={1G}E_{1}=\{1_{G}\} and {Em}m\{E_{m}\}_{m} strictly increasing.

Given a fixed exhaustion {Em}m\{E_{m}\}_{m}, the topology of XX is metrizable by the metric d:X×Xd\colon X\times X\to\mathbb{R} given by d(x,y)=2inf{m:xEmyEm}d(x,y)=2^{-\inf\left\{m\in\mathbb{N}\,:\,x_{E_{m}}\neq y_{E_{m}}\right\}}, where xFx_{F} denotes the restriction of a configuration xx to a set FGF\subseteq G. Denote by XF={xF:xX}X_{F}=\{x_{F}:x\in X\} the set of restrictions of xXx\in X to FF. The sets of the form [w]={xX:xF=w}[w]=\{x\in X\colon x_{F}=w\}, for wXFw\in X_{F}, F(G)F\in\mathcal{F}(G), are called cylinder sets. The family of such sets is the standard basis for the product topology of G\mathbb{N}^{G}.

Let \mathcal{B} be the σ\sigma-algebra generated by the cylinder sets and let (X)\mathcal{M}(X) be the space of probability measures on XX. Consider also G(X)\mathcal{M}_{G}(X) the subspace of GG-invariant probability on X.

The group GG acts by translations on XX as follows: for every xXx\in X and every g,hGg,h\in G,

(gx)(h)=x(hg).(g\cdot x)(h)=x(hg).

This action is also referred, in the literature, as the shift action. Moreover, it can be verified that g[xF]=[(gx)Fg1]g\cdot[x_{F}]=[(g\cdot x)_{Fg^{-1}}], for every xXx\in X, gGg\in G, and FGF\subseteq G.

Given K,F(G)K,F\in\mathcal{F}(G) and δ>0\delta>0, we say that FF is (K,δ)(K,\delta)-invariant if |KFΔF|<δ|F||KF\Delta F|<\delta|F|, where KF={kf:kK,fF}KF=\{kf:k\in K,f\in F\}. A group GG is called amenable if for every K(G)K\in\mathcal{F}(G) and δ>0\delta>0, there exists a (K,δ)(K,\delta)-invariant set FF.

For K,F(G)K,F\in\mathcal{F}(G), define:

  • i)i)

    the KK-interior of FF as IntK(F)={gG:KgF}\mathrm{Int}_{K}(F)=\{g\in G\colon Kg\subseteq F\},

  • ii)ii)

    the KK-exterior of FF as ExtK(F)={gG:KgGF}\mathrm{Ext}_{K}(F)=\{g\in G\colon Kg\subseteq G\setminus F\}, and

  • iii)iii)

    the KK-boundary of FF as K(F)={gG:KgF,KgFc}\partial_{K}(F)=\{g\in G\colon Kg\cap F\neq\emptyset,Kg\cap F^{c}\neq\emptyset\}.

2.2. Potentials and variations

A function ϕ:X\phi\colon X\to\mathbb{R} is called a potential. Given EGE\subseteq G, the variation of ϕ\phi on EE is given by

δE(ϕ):=sup{|ϕ(x)ϕ(y)|:xE=yE}.\delta_{E}(\phi):=\sup\{|\phi(x)-\phi(y)|\colon x_{E}=y_{E}\}.

Notice that if EEE\subseteq E^{\prime}, then δE(ϕ)δE(ϕ)\delta_{E^{\prime}}(\phi)\leq\delta_{E}(\phi). If E={1G}E=\{1_{G}\}, we denote δE(ϕ)\delta_{E}(\phi) simply by δ(ϕ)\delta(\phi). We say that ϕ\phi has finite oscillation if δ(ϕ)<\delta(\phi)<\infty.

Let {Em}m\{E_{m}\}_{m} be an exhausting sequence for GG. Given a potential ϕ:X\phi\colon X\to\mathbb{R}, it is not difficult to see that ϕ\phi is uniformly continuous if and only if limmδEm(ϕ)=0\lim_{m\to\infty}\delta_{E_{m}}(\phi)=0. In this context, given F(G)F\in\mathcal{F}(G), we define the FF-sum of variations of ϕ\phi (according to {Em}m\{E_{m}\}_{m}) as

VF(ϕ):=m1|Em+11FEm1F|δEm(ϕ).V_{F}(\phi):=\sum_{m\geq 1}\left|E_{m+1}^{-1}F\setminus E_{m}^{-1}F\right|\cdot\delta_{E_{m}}(\phi).

If F={1G}F=\{1_{G}\}, we denote VF(ϕ)V_{F}(\phi) simply by V(ϕ)V(\phi). We say that ϕ:X\phi\colon X\to\mathbb{R} has summable variation (according to {Em}m\{E_{m}\}_{m}) if V(ϕ)<V(\phi)<\infty.

Remark 1.

For any exhausting sequence {Em}m\{E_{m}\}_{m} and any F(G)F\in\mathcal{F}(G), the sequence {Em+11FEm1F}m\{E_{m+1}^{-1}F\setminus E_{m}^{-1}F\}_{m} is a partition of GG. Moreover, Em+11FEm1F(Em+11Em1)FE_{m+1}^{-1}F\setminus E_{m}^{-1}F\subseteq(E_{m+1}^{-1}\setminus E_{m}^{-1})F, so

|Em+11FEm1F||Em+11Em1||F|,\left|E_{m+1}^{-1}F\setminus E_{m}^{-1}F\right|\leq\left|E_{m+1}^{-1}\setminus E_{m}^{-1}\right|\left|F\right|,

and VF(ϕ)V(ϕ)|F|V_{F}(\phi)\leq V(\phi)|F|. In particular, if ϕ\phi has summable variation, VF(ϕ)<V_{F}(\phi)<\infty for all F(G)F\in\mathcal{F}(G).

Proposition 2.1.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential such that the FF-sum of variation of ϕ\phi is finite for some F(G)F\in\mathcal{F}(G). Then ϕ\phi is a uniformly continuous potential.

Proof.

Let {Em}m\{E_{m}\}_{m} be an exhausting sequence for GG. Since, in particular, EmEm+1E_{m}\subseteq E_{m+1} for every m1m\geq 1, we have that 0δEm+1(ϕ)δEm(ϕ)0\leq\delta_{E_{m+1}}(\phi)\leq\delta_{E_{m}}(\phi) for every m1m\geq 1. Then, for every M1M\geq 1,

VF(ϕ)\displaystyle V_{F}(\phi) m=1M|Em+11FEm1F|δEm(ϕ)\displaystyle\geq\sum_{m=1}^{M}|E_{m+1}^{-1}F\setminus E_{m}^{-1}F|\cdot\delta_{E_{m}}(\phi)
δEM(ϕ)m=1M|Em+11FEm1F|\displaystyle\geq\delta_{E_{M}}(\phi)\cdot\sum_{m=1}^{M}|E_{m+1}^{-1}F\setminus E_{m}^{-1}F|
=δEM(ϕ)|EM+11FE11F|,\displaystyle=\delta_{E_{M}}(\phi)|E^{-1}_{M+1}F\setminus E_{1}^{-1}F|,

where the last line follows from Remark 1. Therefore,

0limMδEM(ϕ)limMVF(ϕ)|EM+11FE11F|=0,0\leq\lim_{M\to\infty}\delta_{E_{M}}(\phi)\leq\lim_{M\to\infty}\frac{V_{F}(\phi)}{|E^{-1}_{M+1}F\setminus E_{1}^{-1}F|}=0,

and the result follows. ∎

Definition 1.

Let φ:(G)\varphi\colon\mathcal{F}(G)\to\mathbb{R} be a function. Given LL\in\mathbb{R}, we say that φ(F)\varphi(F) converges to LL as FF becomes more and more invariant if for every ϵ>0\epsilon>0 there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that |φ(F)L|<ϵ|\varphi(F)-L|<\epsilon for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G). We will abbreviate such a fact as limFGφ(F)=L\displaystyle\lim_{F\to G}\varphi(F)=L.

A sequence {Fn}n\{F_{n}\}_{n} in (G)\mathcal{F}(G) is (left) Følner for GG if

limn|gFnFn||Fn|=0, for any gG.\lim_{n\to\infty}\frac{|gF_{n}\setminus F_{n}|}{|F_{n}|}=0,\text{ for any $g\in G$}.

For example, if G=dG=\mathbb{Z}^{d} and Fn=[n,n]ddF_{n}=[-n,n]^{d}\cap\mathbb{Z}^{d}, then {Fn}n\{F_{n}\}_{n} is a Følner sequence for d\mathbb{Z}^{d}. It is not difficult to see that if limFGφ(F)=L\lim_{F\to G}\varphi(F)=L, then limnφ(Fn)=L\lim_{n\to\infty}\varphi(F_{n})=L for every Følner sequence {Fn}n\{F_{n}\}_{n}. In particular, when G=dG=\mathbb{Z}^{d}, convergence as FF becomes more and more invariant implies convergence along dd-dimensional boxes, which is a common condition in the multi-dimensional framework. It is not difficult to see that a group is amenable if and only if it has Følner sequence. Moreover, for every amenable group GG there exists a Følner sequence that is also an exhaustion.

Proposition 2.2.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then,

limFGVF(ϕ)|F|=0.\lim_{F\to G}\frac{V_{F}(\phi)}{|F|}=0.
Proof.

Let ϵ>0\epsilon>0. Since ϕ:X\phi\colon X\to\mathbb{R} has summable variation, there exists m01m_{0}\geq 1 such that

m>m0|Em+11Em1|δEm(ϕ)<ϵ.\sum_{m>m_{0}}|E_{m+1}^{-1}\setminus E_{m}^{-1}|\cdot\delta_{E_{m}}(\phi)<\epsilon.

Then, for every F(G)F\in\mathcal{F}(G),

VF(ϕ)\displaystyle V_{F}(\phi) m=1m0|Em+11FEm1F|δEm(ϕ)+m>m0|Em+11Em1||F|δEm(ϕ)\displaystyle\leq\sum_{m=1}^{m_{0}}|E_{m+1}^{-1}F\setminus E_{m}^{-1}F|\cdot\delta_{E_{m}}(\phi)+\sum_{m>m_{0}}|E_{m+1}^{-1}\setminus E_{m}^{-1}||F|\cdot\delta_{E_{m}}(\phi)
m=1m0|Em+11FEm1F|δEm(ϕ)+|F|ϵ.\displaystyle\leq\sum_{m=1}^{m_{0}}|E_{m+1}^{-1}F\setminus E_{m}^{-1}F|\cdot\delta_{E_{m}}(\phi)+|F|\cdot\epsilon.

Due to the amenability of GG, for any given m01m_{0}\geq 1, we have that, for all mm0m\leq m_{0},

|F||Em+11F||Em0+11F|(1+ϵ)|F||F|\leq|E_{m+1}^{-1}F|\leq|E_{m_{0}+1}^{-1}F|\leq(1+\epsilon)|F|

for every (Em0+1,ϵ)(E_{m_{0}+1},\epsilon)-invariant set FF. Therefore, for every ϵ>0\epsilon>0, there exists m01m_{0}\geq 1 and K(G)K\in\mathcal{F}(G) such that for every (K,ϵ)(K,\epsilon)-invariant set FF,

VF(ϕ)m=1m0((1+ϵ)|F||F|)δEm(ϕ)+ϵ|F|=ϵ|F|m=1m0δEm(ϕ)+ϵ|F|,V_{F}(\phi)\leq\sum_{m=1}^{m_{0}}((1+\epsilon)|F|-|F|)\cdot\delta_{E_{m}}(\phi)+\epsilon\cdot|F|=\epsilon\cdot|F|\sum_{m=1}^{m_{0}}\delta_{E_{m}}(\phi)+\epsilon\cdot|F|,

so

VF(ϕ)|F|ϵC,\frac{V_{F}(\phi)}{|F|}\leq\epsilon\cdot C,

where C=1+V(ϕ)C=1+V(\phi). Since ϵ\epsilon was arbitrary, we conclude. ∎

Given a potential ϕ:X\phi\colon X\to\mathbb{R}, for each F(G)F\in\mathcal{F}(G), define ϕF:X\phi_{F}\colon X\to\mathbb{R} as ϕF(x)=gFϕ(gx)\phi_{F}(x)=\sum_{g\in F}\phi(g\cdot x) and ΔF(ϕ)=δF(ϕF)\Delta_{F}(\phi)=\delta_{F}(\phi_{F}). Notice that ΔFg(ϕ)=ΔF(ϕ)\Delta_{Fg}(\phi)=\Delta_{F}(\phi) for every gGg\in G.

Lemma 2.3.

Let {Em}m\{E_{m}\}_{m} be an exhausting sequence for GG, ϕ:X\phi\colon X\to\mathbb{R} be a potential that has finite oscillation and such that lim infmδEm(ϕ)=0\liminf_{m\to\infty}\delta_{E_{m}}(\phi)=0. Then,

(1) limFGΔF(ϕ)|F|=0.\lim_{F\to G}\frac{\Delta_{F}(\phi)}{|F|}=0.

In particular, if ϕ\phi has summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, then equation (1) holds.

Proof.

Let ϵ>0\epsilon>0. Since lim infmδEm(ϕ)=0\liminf_{m\to\infty}\delta_{E_{m}}(\phi)=0, there exists m01m_{0}\geq 1 such that δEm0(ϕ)ϵ\delta_{E_{m_{0}}}(\phi)\leq\epsilon. Denote Em0E_{m_{0}} by KK. Due to amenability, we can find KKK^{\prime}\supseteq K and 0<ϵϵ0<\epsilon^{\prime}\leq\epsilon such that if FF is (K,ϵ)(K^{\prime},\epsilon^{\prime})-invariant, we have that

|IntK(F)F|<ϵ|F|.\left|\mathrm{Int}_{K}(F)\triangle F\right|<\epsilon\cdot|F|.

Considering this, if x,yXx,y\in X are such that xF=yFx_{F}=y_{F}, we have that

|ϕF(x)ϕF(y)|\displaystyle|\phi_{F}(x)-\phi_{F}(y)| gF|ϕ(gx)ϕ(gy)|\displaystyle\leq\sum_{g\in F}|\phi(g\cdot x)-\phi(g\cdot y)|
=gIntK(F)F|ϕ(gx)ϕ(gy)|+gFIntK(F)|ϕ(gx)ϕ(gy)|\displaystyle=\sum_{g\in\mathrm{Int}_{K}(F)\cap F}|\phi(g\cdot x)-\phi(g\cdot y)|+\sum_{g\in F\setminus\mathrm{Int}_{K}(F)}|\phi(g\cdot x)-\phi(g\cdot y)|
gIntK(F)FδK(ϕ)+gFIntK(F)δ(ϕ)\displaystyle\leq\sum_{g\in\mathrm{Int}_{K}(F)\cap F}\delta_{K}(\phi)+\sum_{g\in F\setminus\mathrm{Int}_{K}(F)}\delta(\phi)
|IntK(F)|ϵ+|FIntK(F)|δ(ϕ)\displaystyle\leq|\mathrm{Int}_{K}(F)|\cdot\epsilon+|F\setminus\mathrm{Int}_{K}(F)|\cdot\delta(\phi)
|F|(1+ϵ)ϵ+|F|ϵδ(ϕ)\displaystyle\leq|F|\cdot(1+\epsilon)\cdot\epsilon+|F|\cdot\epsilon\cdot\delta(\phi)
=|F|ϵ(1+ϵ+δ(ϕ)),\displaystyle=|F|\cdot\epsilon\cdot(1+\epsilon+\delta(\phi)),

and the result follows. ∎

3. Pressure

We dedicate this section to introduce the pressure of a potential. We define and work on the setting of exp-summable potentials with summable variation on a countable alphabet. The pressure — basically equivalent to the specific Gibbs free energy — is a very relevant thermodynamic quantity that helps to capture the concept of Gibbs measure in a quantitative way.

First, we prove that the pressure, which we define through a limit over sets that are becoming more and more invariant, exists in the finite alphabet case. The definition of the pressure is often done in terms of a particular Følner sequence, which is an, a priori, less robust and less overarching approach. Existence of the limit for a particular Følner sequence {Fn}n\{F_{n}\}_{n} and the fact that it is independent on the choice of such sequence is well-known (see, for example, [57, 35, 16], in the context of absolutely summable interactions). Here, we prove something stronger: that our definition of pressure obeys the infimum rule — which is a refinement of the Ornstein-Weiss Lemma (see, for example, [39, §4.5]) —, this is to say, it can be expressed as an infimum over all finite sets of GG. In order to conclude this, we extend the results about Shearer’s inequality in [25] for topological entropy to pressure.

Now, in the countable alphabet context, we take a similar approach. First, we consider again a definition of pressure in terms of sets that are becoming more and more invariant. Next, we prove that the infimum rule still holds and, finally, we prove that the pressure can be obtained as the supremum of the pressures associated with finite alphabet subsystems. A related result was obtained by Muir in [45] for the d\mathbb{Z}^{d} group case, where the pressure was defined as a limit over a particular type of Følner sequence, namely, open boxes centered at the origin of radius nn. The existence of this limit was proven through a sub-additivity argument that exploits the property that large boxes can be partitioned into many equally sized ones, which might not be valid in more general groups. In order to generalize this idea of partitioning sets, we make use of tiling techniques introduced in [26], which, together to what is done in the finite alphabet case, allow us to prove the infimum rule for infinite alphabets over a countable amenable group. This type of result was not considered in [45].

We begin by introducing some definitions. Given a potential ϕ:X\phi\colon X\to\mathbb{R} and F(G)F\in\mathcal{F}(G), define the partition function for ϕ\phi on FF as

ZF(ϕ):=wXFexp(supϕF([w])),Z_{F}(\phi):=\sum_{w\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([w])\right),

where supϕF([w])=sup{ϕF(x):x[w]}\sup\phi_{F}([w])=\sup\{\phi_{F}(x):x\in[w]\}. We define the pressure of ϕ\phi, which we denote by p(ϕ)p(\phi), as

p(ϕ):=limFG1|F|logZF(ϕ),p(\phi):=\lim_{F\to G}\frac{1}{|F|}\log Z_{F}(\phi),

whenever such limit as FF becomes more and more invariant exists. In addition, given a finite subset A()A\in\mathcal{F}(\mathbb{N}), we define ZF(A,ϕ)Z_{F}(A,\phi) as the partition function associated with the restriction of ϕ\phi to AGA^{G}. More precisely,

ZF(A,ϕ):=wXFAFexpsup(ϕF([w]AG)).Z_{F}(A,\phi):=\sum_{w\in X_{F}\cap A^{F}}\mathop{\textrm{\rm exp}}\nolimits\sup\left(\phi_{F}\left([w]\cap A^{G}\right)\right).

Similarly, we define p(A,ϕ)p(A,\phi) as

p(A,ϕ):=limFG1|F|logZF(A,ϕ),p(A,\phi):=\lim_{F\to G}\frac{1}{|F|}\log Z_{F}(A,\phi),

whenever such limit exists.

3.1. Infimum rule for finite alphabet pressure

The main goal of this subsection is to prove the following theorem.

Theorem 3.1.

Let ϕ:X\phi\colon X\to\mathbb{R} be a continuous potential. Then, for any finite alphabet AA\subseteq\mathbb{N}, p(A,ϕ)p(A,\phi) exists and

p(A,ϕ)=infE(G)1|E|logZE(A,ϕ).p(A,\phi)=\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi).

In order to prove this result, we require some definitions. A function φ:(G)\varphi\colon\mathcal{F}(G)\rightarrow\mathbb{R} is

  • GG-invariant if φ(Fg)=φ(F)\varphi(Fg)=\varphi(F) for every F(G)F\in\mathcal{F}(G) and gGg\in G;

  • monotone if φ(E)φ(F)\varphi(E)\leq\varphi(F) for every E,F(G)E,F\in\mathcal{F}(G) such that EFE\subseteq F; and

  • sub-additive if φ(EF)φ(E)+φ(F)\varphi(E\cup F)\leq\varphi(E)+\varphi(F) for any E,F(G)E,F\in\mathcal{F}(G).

A kk-cover 𝒦\mathcal{K} of a set F(G)F\in\mathcal{F}(G) is a family {K1,K2,,Kr}(G)\left\{K_{1},K_{2},\dots,K_{r}\right\}\subseteq\mathcal{F}(G) (with possible repetitions) such that each element of FF belongs to KiK_{i} for at least kk indices i{1,,r}i\in\{1,\dots,r\}. We say that φ\varphi satisfies Shearer’s inequality if for any F(G)F\in\mathcal{F}(G) and any kk-cover 𝒦\mathcal{K} of F{F}, it holds that

φ(F)1kK𝒦φ(K).\varphi({F})\leq\frac{1}{k}\sum_{K\in\mathcal{K}}\varphi(K).

Notice that Shearer’s inequality implies sub-additivity. Considering this, we have the following key lemma.

Lemma 3.2 ([39, §4]).

Let φ:(G)\varphi\colon\mathcal{F}(G)\rightarrow\mathbb{R} be a non-negative monotone GG-invariant sub-additive function. Then there exists α[0,)\alpha\in[0,\infty) such that

limFGφ(F)|F|=α.\lim_{F\to G}\frac{\varphi(F)}{|F|}=\alpha.

Moreover, if φ\varphi satisfies Shearer’s inequality, then

α=infE(G)φ(E)|E|.\alpha=\inf_{E\in\mathcal{F}(G)}\frac{\varphi(E)}{|E|}.

In this last case, we say that φ\varphi satisfies the infimum rule.

Now, fix a finite alphabet A()A\in\mathcal{F}(\mathbb{N}). For a continuous potential ϕ:X\phi\colon X\to\mathbb{R}, we denote by ϕA\|\phi\|_{A} the supremum norm of ϕ\phi over the compact set XAGX\cap A^{G}, i.e., ϕA=supxXAG|ϕ(x)|\|\phi\|_{A}=\sup_{x\in X\cap A^{G}}|\phi(x)|. Next, given a set EGE\subseteq G, F(G)F\in\mathcal{F}(G), and uEXEAEu_{E}\in X_{E}\cap A^{E}, we define

ZFuE:=wFEAFEexp(supϕF([wFEuE])),Z^{u_{E}}_{F}:=\sum_{w_{F\setminus E}\in A^{F\setminus E}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([w_{F\setminus E}u_{E}])\right),

where the supremum is over xXAGx\in X\cap A^{G} and, if [v]=[v]=\emptyset, then supϕ([v])=\sup\phi([v])=-\infty and exp()=0\mathop{\textrm{\rm exp}}\nolimits(-\infty)=0. Notice that ZF=ZFuEZ_{F}=Z^{u_{E}}_{F} for E=E=\emptyset.

Now, suppose that ϕ|XAG\left.\phi\right|_{X\cap A^{G}} is non-negative. Then, it is easy to check that for any EGE\subseteq G and uEAEu_{E}\in A^{E}, the function φ~:(G)\tilde{\varphi}\colon\mathcal{F}(G)\to\mathbb{R} given by φ~(F)=ZFuE\tilde{\varphi}(F)=Z^{u_{E}}_{F} satisfies that

  1. i)i)

    φ~(F)1\tilde{\varphi}(F)\geq 1 for every F(G)F\in\mathcal{F}(G) and

  2. ii)ii)

    φ~\tilde{\varphi} is monotone, that is, if F1F2F_{1}\subseteq F_{2}, then φ~(F1)φ~(F2)\tilde{\varphi}(F_{1})\leq\tilde{\varphi}(F_{2}).

Next, consider the function φ:(G)\varphi\colon\mathcal{F}(G)\to\mathbb{R} defined as φ(F)=logZF\varphi(F)=\log Z_{F}. From the properties above and properties of the log()\log(\cdot) function, it follows that φ\varphi is non-negative and monotone. Moreover, φ\varphi is GG-invariant. The following lemma is a generalization of [25, Lemma 6.1] designed to address the pressure case instead of just the topological entropy and, in particular, it claims that φ\varphi satisfies Shearer’s inequality.

Lemma 3.3.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential and A()A\in\mathcal{F}(\mathbb{N}) such that ϕ|XAG\left.\phi\right|_{X\cap A^{G}} is non-negative. Then, for every EGE\subseteq G, uEXEAEu_{E}\in X_{E}\cap A^{E}, F(G)F\in\mathcal{F}(G), and any kk-cover 𝒦\mathcal{K} of FF, it holds that

ZFuEK𝒦(ZKuE)1/k.Z^{u_{E}}_{F}\leq\prod_{K\in\mathcal{K}}(Z^{u_{E}}_{K})^{1/k}.

In particular, φ\varphi satisfies Shearer’s inequality.

Proof.

Given a kk-cover 𝒦\mathcal{K} of FF, notice that, since ϕ|XAG\left.\phi\right|_{X\cap A^{G}} is non-negative,

ϕF(x)=gFϕ(gx)1kK𝒦gKϕ(gx)=1kK𝒦ϕK(x)\phi_{F}(x)=\sum_{g\in F}\phi(g\cdot x)\leq\frac{1}{k}\sum_{K\in\mathcal{K}}\sum_{g\in K}\phi(g\cdot x)=\frac{1}{k}\sum_{K\in\mathcal{K}}\phi_{K}(x)

for any xXAGx\in X\cap A^{G}. We proceed by induction on the size of FEF\setminus E. First, suppose that |FE|=0|F\setminus E|=0. Then, FE=F\setminus E=\emptyset and

ZFuE\displaystyle Z^{u_{E}}_{F} =exp(supϕF([uE]))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([u_{E}])\right)
exp(sup1kK𝒦ϕK([uE]))\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(\sup\frac{1}{k}\sum_{K\in\mathcal{K}}\phi_{K}([u_{E}])\right)
exp(K𝒦1ksupϕK([uE]))\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(\sum_{K\in\mathcal{K}}\frac{1}{k}\sup\phi_{K}([u_{E}])\right)
=K𝒦(expsupϕK([uE]))1/k\displaystyle=\prod_{K\in\mathcal{K}}\left(\mathop{\textrm{\rm exp}}\nolimits\sup\phi_{K}([u_{E}])\right)^{1/k}
K𝒦(wKEexpsupϕK([wKEuE]))1/k\displaystyle\leq\prod_{K\in\mathcal{K}}\left(\sum_{w_{K\setminus E}}\mathop{\textrm{\rm exp}}\nolimits\sup\phi_{K}([w_{K\setminus E}u_{E}])\right)^{1/k}
=K𝒦(ZKuE)1/k.\displaystyle=\prod_{K\in\mathcal{K}}\left(Z^{u_{E}}_{K}\right)^{1/k}.

Now suppose that ZFuEK𝒦(ZKuE)1/kZ^{u_{E}}_{F}\leq\prod_{K\in\mathcal{K}}(Z^{u_{E}}_{K})^{1/k} for every EG,uEXEAEE\subseteq G,u_{E}\in X_{E}\cap A^{E}, F(G)F\in\mathcal{F}(G) with |FE|n|F\setminus E|\leq n, and every kk-cover 𝒦\mathcal{K} of FF. We will show that the same holds for E,FE,F with |FE|=n+1|F\setminus E|=n+1. Fix gFEg\in F\setminus E and notice that |F(E{g})|=n|F\setminus(E\cup\{g\})|=n. Then,

ZFuE\displaystyle Z^{u_{E}}_{F} =aAZFaguE\displaystyle=\sum_{a\in A}Z^{a^{g}u_{E}}_{F}
aAK𝒦(ZKaguE)1/k\displaystyle\leq\sum_{a\in A}\prod_{K\in\mathcal{K}}\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}
=aAK𝒦:gK(ZKaguE)1/kK𝒦:gK(ZKaguE)1/k\displaystyle=\sum_{a\in A}\prod_{K\in\mathcal{K}:g\notin K}\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}\cdot\prod_{K\in\mathcal{K}:g\in K}\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}
K𝒦:gK(ZKuE)1/kaAK𝒦:gK(ZKaguE)1/k\displaystyle\leq\prod_{K\in\mathcal{K}:g\notin K}\left(Z^{u_{E}}_{K}\right)^{1/k}\sum_{a\in A}\prod_{K\in\mathcal{K}:g\in K}\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}
K𝒦:gK(ZKuE)1/kK𝒦:gK(aAZKaguE)1/k\displaystyle\leq\prod_{K\in\mathcal{K}:g\notin K}\left(Z^{u_{E}}_{K}\right)^{1/k}\cdot\prod_{K\in\mathcal{K}:g\in K}\left(\sum_{a\in A}Z^{a^{g}u_{E}}_{K}\right)^{1/k}
=K𝒦:gK(ZKuE)1/kK𝒦:gK(ZKuE)1/k\displaystyle=\prod_{K\in\mathcal{K}:g\notin K}\left(Z^{u_{E}}_{K}\right)^{1/k}\cdot\prod_{K\in\mathcal{K}:g\in K}\left(Z^{u_{E}}_{K}\right)^{1/k}
=K𝒦(ZKuE)1/k.\displaystyle=\prod_{K\in\mathcal{K}}\left(Z^{u_{E}}_{K}\right)^{1/k}.

Notice that the first inequality follows from the induction hypothesis and the third inequality follows from the generalized Hölder inequality. Indeed, consider p1p\leq 1 such that K𝒦:gK1k=1p\sum_{K\in\mathcal{K}:g\in K}\frac{1}{k}=\frac{1}{p} and the functions fK:Af_{K}\colon A\to\mathbb{R} given by fK(a)=(ZKaguE)1/kf_{K}(a)=\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}. By the generalized Hölder inequality,

K𝒦:gKfKpK𝒦:gKfKk,\left\|\prod_{K\in\mathcal{K}:g\in K}f_{K}\right\|_{p}\leq\prod_{K\in\mathcal{K}:g\in K}\|f_{K}\|_{k},

where

K𝒦:gKfKk\displaystyle\prod_{K\in\mathcal{K}:g\in K}\|f_{K}\|_{k} =K𝒦:gK(aA((ZKaguE)1/k)k)1/k=K𝒦:gK(aAZKaguE)1/k=K𝒦:gK(ZKuE)1/k\displaystyle=\prod_{K\in\mathcal{K}:g\in K}\left(\sum_{a\in A}((Z^{a^{g}u_{E}}_{K})^{1/k})^{k}\right)^{1/k}=\prod_{K\in\mathcal{K}:g\in K}\left(\sum_{a\in A}Z^{a^{g}u_{E}}_{K}\right)^{1/k}=\prod_{K\in\mathcal{K}:g\in K}\left(Z^{u_{E}}_{K}\right)^{1/k}

and, since p\|\cdot\|_{p} is monotonically decreasing in pp for any fixed |A||A|-dimensional vector,

K𝒦:gKfKpK𝒦:gKfK1=aAK𝒦:gK(ZKaguE)1/k.\left\|\prod_{K\in\mathcal{K}:g\in K}f_{K}\right\|_{p}\geq\left\|\prod_{K\in\mathcal{K}:g\in K}f_{K}\right\|_{1}=\sum_{a\in A}\prod_{K\in\mathcal{K}:g\in K}\left(Z^{a^{g}u_{E}}_{K}\right)^{1/k}.

Therefore, ZFuEK𝒦(ZKuE)1/kZ^{u_{E}}_{F}\leq\prod_{K\in\mathcal{K}}\left(Z^{u_{E}}_{K}\right)^{1/k}. In particular, if E=E=\emptyset, ZFK𝒦(ZK)1/kZ_{F}\leq\prod_{K\in\mathcal{K}}\left(Z_{K}\right)^{1/k}. ∎

Proof (of Theorem 3.1).

As a consequence of Lemma 3.3, we have that if ϕ|XAG\left.\phi\right|_{X\cap A^{G}} is non-negative, then φ\varphi satisfies Shearer’s inequality. Thus, by the Ornstein-Weiss lemma, p(A,ϕ)p(A,\phi) exists and it satisfies the infimum rule, i.e.,

p(A,ϕ)=infE(G)1|E|logZE(A,ϕ).p(A,\phi)=\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi).

Finally, in order to deal with the general case, it suffices to apply the previous result to ϕ+ϕ\phi+\|\phi\| and then observe that p(A,ϕ+C)=p(A,ϕ)+Cp(A,\phi+C)=p(A,\phi)+C for any constant CC. ∎

Remark 2.

Notice that the previous results (namely, Lemma 3.3 and Theorem 3.1) also hold for GG-subshifts, this is to say, any closed and GG-invariant subsets XX of G\mathbb{N}^{G}.

3.2. Tilings

Pressure is one of the most important notions in thermodynamic formalism. One key technique to properly define pressure is sub-additivity, which is based on our ability to partition a system in smaller and representative pieces. In the context of countable amenable groups, it appears to be necessary to generalize tools that are classically used in the d\mathbb{Z}^{d} case (e.g., [52, 45]). In order to do this, we will begin by exploring the concept of (exact) tilings of amenable groups and the relatively recent techniques introduced in [26].

Definition 2.

Given

  1. (1)

    a finite collection 𝒮(𝒯)\mathcal{S}(\mathcal{T}) of finite subsets of GG containing the identity 1G1_{G}, called the shapes, and

  2. (2)

    a finite collection 𝒞(𝒯)={C(S):S𝒮(𝒯)}\mathcal{C}(\mathcal{T})=\{C(S):S\in\mathcal{S}(\mathcal{T})\} of disjoint subsets of GG, called center sets (for the shapes),

the family 𝒯={(S,c):S𝒮(𝒯),cC(S)}\mathcal{T}=\{(S,c):S\in\mathcal{S}(\mathcal{T}),c\in C(S)\} is called a tiling if the map (S,c)Sc(S,c)\mapsto Sc is injective and {Sc}S𝒮(𝒯),cC(S)\{Sc\}_{S\in\mathcal{S}(\mathcal{T}),c\in C(S)} is a partition of GG. In addition, by the tiles of 𝒯\mathcal{T} (usually denoted by the letter TT) we will mean either the sets ScSc or the pairs (S,c)(S,c), depending on the context.

We say that a sequence {𝒯n}n\{\mathcal{T}_{n}\}_{n} of tilings is congruent if, for each n1n\geq 1, every tile of 𝒯n+1\mathcal{T}_{n+1} is equal to a (disjoint) union of tiles of 𝒯n\mathcal{T}_{n}. The following theorem is the main result in [26], which gives sufficient conditions so that we can guarantee the existence of such sequence with extra invariance properties.

Theorem 3.4.

([26, Theorem 5.2]) Let {ϵn}n\{\epsilon_{n}\}_{n} be a sequence of positive real numbers converging to zero and {Kn}n\{K_{n}\}_{n} be a sequence of finite subsets of GG. Then, there exists a congruent sequence {𝒯n}n\{\mathcal{T}_{n}\}_{n} of tilings of GG such that the shapes of 𝒯n\mathcal{T}_{n} are (Kn,ϵn)\left(K_{n},\epsilon_{n}\right)-invariant.

Given a tiling 𝒯\mathcal{T}, we define S𝒯=S𝒮(𝒯)SS1S_{\mathcal{T}}=\bigcup_{S\in\mathcal{S}(\mathcal{T})}SS^{-1}. Notice that S𝒯S_{\mathcal{T}} contains every shape S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}), S𝒯1=S𝒯S_{\mathcal{T}}^{-1}=S_{\mathcal{T}}, and 1GS𝒯1_{G}\in S_{\mathcal{T}}. Given a tiling, the next lemma provides a way to approximate any sufficiently invariant shape by a union of tiles.

Lemma 3.5.

Given K(G)K\in\mathcal{F}(G) and δ>0\delta>0, consider a tiling 𝒯\mathcal{T} with (K,δ)(K,\delta)-invariant shapes. Then, for any ϵ>0\epsilon>0 and any (S𝒯,ϵ)(S_{\mathcal{T}},\epsilon)-invariant set F(G)F\in\mathcal{F}(G), there exist center sets CF(S)C(S)C_{F}(S)\subseteq C(S) for S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}) such that

FS𝒮(𝒯)SCF(S)and|FS𝒮(𝒯)SCF(S)|ϵ|F|.F\supseteq\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\quad\text{and}\quad\left|F\setminus\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\right|\leq\epsilon|F|.
Proof.

Consider a tiling 𝒯\mathcal{T} made of (K,δ)(K,\delta)-invariant shapes and ϵ>0\epsilon>0. Suppose that FF is (S𝒯,ϵ)(S_{\mathcal{T}},\epsilon)-invariant. Consider the sets CF(S)=C(S)IntS(F)C_{F}(S)=C(S)\cap\mathrm{Int}_{S}(F) and C¯F(S)=C(S)S1F\overline{C}_{F}(S)=C(S)\cap S^{-1}F for S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}). Notice that, since 𝒯\mathcal{T} induces a partition, |SCF(S)|=|S||CF(S)||SC_{F}(S)|=|S||C_{F}(S)|, |SC¯F(S)|=|S||C¯F(S)||S\overline{C}_{F}(S)|=|S||\overline{C}_{F}(S)|, and

S𝒮(𝒯)SCF(S)FS𝒮(𝒯)SC¯F(S).\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\subseteq F\subseteq\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}S\overline{C}_{F}(S).

Therefore,

FS𝒮(𝒯)SCF(S)\displaystyle F\setminus\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S) S𝒮(𝒯)SC¯F(S)S𝒮(𝒯)SCF(S)\displaystyle\subseteq\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}S\overline{C}_{F}(S)\setminus\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)
=S𝒮(𝒯)S(C¯F(S)CF(S))S𝒯(F).\displaystyle=\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}S(\overline{C}_{F}(S)\setminus C_{F}(S))\subseteq\partial_{S_{\mathcal{T}}}(F).

Indeed, to check the last inclusion, notice that if gS𝒮(𝒯)S(C¯F(S)CF(S))g\in\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}S(\overline{C}_{F}(S)\setminus C_{F}(S)), then g=scg=sc, where sSs\in S and cC¯F(S)CF(S)c\in\overline{C}_{F}(S)\setminus C_{F}(S) for some S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}). Therefore, since cC¯F(S)c\in\overline{C}_{F}(S),

S𝒯gFSS1scFScF.S_{\mathcal{T}}g\cap F\supseteq SS^{-1}sc\cap F\supseteq Sc\cap F\neq\emptyset.

Similarly, since cCF(S)c\notin C_{F}(S),

S𝒯gFcSS1scFcScFc.S_{\mathcal{T}}g\cap F^{c}\supseteq SS^{-1}sc\cap F^{c}\supseteq Sc\cap F^{c}\neq\emptyset.

so that gS𝒯(F)g\in\partial_{S_{\mathcal{T}}}(F). Then,

|FS𝒮(𝒯)SCF(S)||S𝒯(F)||S𝒯FF|ϵ|F|,\left|F\setminus\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\right|\leq\left|\partial_{S_{\mathcal{T}}}(F)\right|\leq\left|S_{\mathcal{T}}F\triangle F\right|\leq\epsilon\cdot|F|,

where we have used that |K(F)||(KK1{1G})FF||\partial_{K}(F)|\leq|(K\cup K^{-1}\cup\{1_{G}\})F\triangle F| for any K(G)K\in\mathcal{F}(G) and that S𝒯1=S𝒯S_{\mathcal{T}}^{-1}=S_{\mathcal{T}} and 1GS𝒯1_{G}\in S_{\mathcal{T}}. ∎

3.3. Infimum rule for countable alphabet pressure

We say that ϕ:X\phi\colon X\to\mathbb{R} is exp-summable if Z1G(ϕ)<Z_{1_{G}}(\phi)<\infty. Notice that ZF(ϕ)Z_{F}(\phi) is sub-multiplicative, that is, if E,F(G)E,F\in\mathcal{F}(G) are disjoint, then ZEF(Φ)ZE(ϕ)ZF(ϕ)Z_{E\cup F}(\Phi)\leq Z_{E}(\phi)\cdot Z_{F}(\phi). Also, notice that ZF(ϕ)Z_{F}(\phi) is GG-invariant, namely, for any gGg\in G, ZF(ϕ)=ZFg(ϕ)Z_{F}(\phi)=Z_{Fg}(\phi). Then, in particular, ZF(ϕ)Z1G(ϕ)|F|Z_{F}(\phi)\leq Z_{1_{G}}(\phi)^{|F|}, so ϕ\phi is exp-summable if and only if ZF(ϕ)<Z_{F}(\phi)<\infty for every F(G)F\in\mathcal{F}(G). Finally, observe that if ϕ\phi is exp-summable, then it must be bounded from above.

Before stating the main result of this section, we begin by the next lemma, that guarantees that given a finite shape FF, one can approximate the partition function on FF using a finite alphabet.

Lemma 3.6.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable and uniformly continuous potential. Then, for every ϵ>0\epsilon>0 and every F(G)F\in\mathcal{F}(G) such that |F|1ϵlog(1ϵ)|F|\geq-\frac{1}{\epsilon}\log(1-\epsilon), there exists AF()A_{F}\in\mathcal{F}(\mathbb{N}) such that

ZF(AF,ϕ)(1ϵ)ZF(ϕ).\displaystyle Z_{F}(A_{F},\phi)\geq(1-\epsilon)Z_{F}(\phi).
Proof.

Let ϵ>0\epsilon>0 and F(G)F\in\mathcal{F}(G) be such that |F|1ϵlog(1ϵ)|F|\geq-\frac{1}{\epsilon}\log(1-\epsilon). For every such FF, there exists a finite set of words WFXFW_{F}\Subset X_{F} such that

wWFexp(supϕF)ZF(ϕ)1ϵ.\displaystyle\sum_{w\in W_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}\right)\geq Z_{F}(\phi)\sqrt{1-\epsilon}.

On the other hand, since ϕ\phi is uniformly continuous, there must be an index m1m\geq 1 for which

δEm(ϕ)13|F|log(11ϵ).\displaystyle\delta_{E_{m}}(\phi)\leq\frac{1}{3|F|}\log\left(\frac{1}{\sqrt{1-\epsilon}}\right).

For each wWFw\in W_{F}, pick a word wEmFw^{\prime}\in\mathbb{N}^{E_{m}F} such that wF=ww^{\prime}_{F}=w and

supϕF[w]supϕF[w]13log(11ϵ).\displaystyle\sup\phi_{F}[w^{\prime}]\geq\sup\phi_{F}[w]-\frac{1}{3}\log\left(\frac{1}{\sqrt{1-\epsilon}}\right).

In addition, for each such ww^{\prime}, pick a configuration xw[w]x_{w}\in[w^{\prime}] such that

ϕF(xw)supϕF[w]13log(11ϵ).\displaystyle\phi_{F}(x_{w})\geq\sup\phi_{F}[w^{\prime}]-\frac{1}{3}\log\left(\frac{1}{\sqrt{1-\epsilon}}\right).

Define AFA_{F} to be wWFw(EmF)\bigcup_{w\in W_{F}}w^{\prime}(E_{m}F), where w(EmF)=gEmF{w(g)}w^{\prime}(E_{m}F)=\bigcup_{g\in E_{m}F}\{w^{\prime}(g)\}. It is direct that AFA_{F} is a finite subset of \mathbb{N}. Pick y[w]AFGy\in[w^{\prime}]\cap A_{F}^{G} and notice that (gxw)Em=(gy)Em(g\cdot x_{w})_{E_{m}}=(g\cdot y)_{E_{m}} for all gFg\in F. Then, for every wWFw\in W_{F},

supϕF[[w]AFG]\displaystyle\sup\phi_{F}[[w^{\prime}]\cap A_{F}^{G}] ϕF(y)\displaystyle\geq\phi_{F}(y)
ϕF(xw)gF|ϕ(gxw)ϕ(gy)|\displaystyle\geq\phi_{F}(x_{w})-\sum_{g\in F}\left|\phi(g\cdot x_{w})-\phi(g\cdot y)\right|
ϕF(xw)|F|δEm(ϕ)\displaystyle\geq\phi_{F}(x_{w})-\left|F\right|\delta_{E_{m}}(\phi)
ϕF(xw)13log(11ϵ)\displaystyle\geq\phi_{F}(x_{w})-\frac{1}{3}\log\left(\frac{1}{\sqrt{1-\epsilon}}\right)
supϕF[w]23log(11ϵ)\displaystyle\geq\sup\phi_{F}[w^{\prime}]-\frac{2}{3}\log\left(\frac{1}{\sqrt{1-\epsilon}}\right)
supϕF[w]log(11ϵ).\displaystyle\geq\sup\phi_{F}[w]-\log\left(\frac{1}{\sqrt{1-\epsilon}}\right).

Hence,

ZF(AF,ϕ)\displaystyle Z_{F}(A_{F},\phi) =wAFFexp(supϕF[[w]AFG])\displaystyle=\sum_{w\in A_{F}^{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[[w]\cap A_{F}^{G}]\right)
wWFexp(supϕF[[w]AFG])\displaystyle\geq\sum_{w\in W_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[[w]\cap A_{F}^{G}]\right)
wWFexp(supϕF[[w]AFG])\displaystyle\geq\sum_{w\in W_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[[w^{\prime}]\cap A_{F}^{G}]\right)
wWFexp(supϕF[w]log(11ϵ))\displaystyle\geq\sum_{w\in W_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[w]-\log\left(\frac{1}{\sqrt{1-\epsilon}}\right)\right)
=1ϵwWFexp(supϕF[w])\displaystyle=\sqrt{1-\epsilon}\sum_{w\in W_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[w]\right)
(1ϵ)ZF(ϕ).\displaystyle\geq(1-\epsilon)Z_{F}(\phi).

The next proposition establishes a fundamental connection between the partition function for sufficiently invariant sets F(G)F\in\mathcal{F}(G) and the pressure for a sufficiently large finite alphabet AA.

Proposition 3.7.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable and uniformly continuous potential with finite oscillation. Then, for every 12>ϵ>0\frac{1}{2}>\epsilon>0, there exist A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G), and δ>0\delta>0 such that for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G), it holds that

(2) 1|F|logZF(ϕ)infE(G)1|E|logZE(A,ϕ)+ϵ.\displaystyle\frac{1}{|F|}\log Z_{F}(\phi)\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi)+\epsilon.
Proof.

Fix 1/2>ϵ>01/2>\epsilon>0 and an exhausting sequence {Em}m\{E_{m}\}_{m} for GG. Since ϕ\phi is uniformly continuous, we have that limFGΔF(ϕ)|F|=0\lim_{F\to G}\frac{\Delta_{F}(\phi)}{|F|}=0, by Lemma 2.3. Therefore, there exist K(G)K^{\prime}\in\mathcal{F}(G) and δ>0\delta^{\prime}>0 such that ΔF(ϕ)<ϵ|F|\Delta_{F}(\phi)<\epsilon|F| for every finite (K,δ)(K^{\prime},\delta^{\prime})-invariant set FF.

By Theorem 3.4, there exists a tiling 𝒯\mathcal{T}^{\prime} such that its shapes are (K,δ)(K^{\prime},\delta^{\prime})-invariant. Without loss of generality, by possibly readjusting KK^{\prime} and δ\delta^{\prime}, assume that |S|1ϵlog(1ϵ)|S^{\prime}|\geq-\frac{1}{\epsilon}\log(1-\epsilon) for every S𝒮(𝒯)S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime}). Therefore, by Lemma 3.6, for every S𝒮(𝒯)S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime}) there exists ASA_{S^{\prime}}\Subset\mathbb{N} such that ZS(AS,ϕ)(1ϵ)ZS(ϕ)Z_{S^{\prime}}(A_{S^{\prime}},\phi)\geq(1-\epsilon)Z_{S^{\prime}}(\phi). Define AA to be S𝒮(𝒯)AS\bigcup_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}A_{S^{\prime}}. Then, AA is a finite subset of \mathbb{N}. Moreover, since ASAA_{S^{\prime}}\subseteq A, for each S𝒮(𝒯)S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime}), we have that

(3) ZS(A,ϕ)(1ϵ)ZS(ϕ),Z_{S^{\prime}}(A,\phi)\geq(1-\epsilon)Z_{S^{\prime}}(\phi),

for every S𝒮(𝒯)S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime}).

Now, by Theorem 3.1, p(A,ϕ)=limFG1|F|logZF(A,ϕ)p(A,\phi)=\lim_{F\to G}\frac{1}{|F|}\log Z_{F}(A,\phi) exists, so we can pick K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that KKK\supseteq K^{\prime}, δ<δ\delta<\delta^{\prime}, and

(4) logZF(A,ϕ)|F|(p(A,ϕ)+ϵ)\log Z_{F}(A,\phi)\leq|F|(p(A,\phi)+\epsilon)

for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G).

Next, by Theorem 3.4, we can obtain a tiling 𝒯\mathcal{T} of (K,δ)(K,\delta)-invariant sets such that every tile in 𝒯\mathcal{T} is a union of tiles in 𝒯\mathcal{T}^{\prime}, i.e., S=S𝒮(𝒯)cCS(S)ScS=\bigsqcup_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\bigsqcup_{c^{\prime}\in C_{S}(S^{\prime})}S^{\prime}c^{\prime}. Furthermore, by Lemma 3.5, for every (S𝒯,ϵ)(S_{\mathcal{T}},\epsilon)-invariant set F(G)F\in\mathcal{F}(G), there exist center sets CF(S)C(S)𝒞(𝒯)C_{F}(S)\subseteq C(S)\in\mathcal{C}(\mathcal{T}) for S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}) such that

FTFand|FTF|ϵ|F|,F\supseteq T_{F}\quad\text{and}\quad|F\setminus T_{F}|\leq\epsilon|F|,

where TF=S𝒮(𝒯)SCF(S)T_{F}=\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S).

Furthermore, for every S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}), we have that

ZS(A,ϕ)\displaystyle Z_{S}(A,\phi) =wSASexp(supϕS([wS]AG))\displaystyle=\sum_{w_{S}\in A^{S}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{S}([w_{S}]\cap A^{G})\right)
wSASexp(infϕS([wS]AG))\displaystyle\geq\sum_{w_{S}\in A^{S}}\mathop{\textrm{\rm exp}}\nolimits\left(\inf\phi_{S}\left([w_{S}]\cap A^{G}\right)\right)
S𝒮(𝒯)cCS(S)wScAScexp(infϕSc([wSc]AG))\displaystyle\geq\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\prod_{c^{\prime}\in C_{S}(S^{\prime})}\sum_{w_{S^{\prime}c^{\prime}}\in A^{S^{\prime}c^{\prime}}}\mathop{\textrm{\rm exp}}\nolimits\left(\inf\phi_{S^{\prime}c^{\prime}}\left([w_{S^{\prime}c^{\prime}}]\cap A^{G}\right)\right)
S𝒮(𝒯)cCS(S)wScAScexp(supϕSc([wSc]AG)ΔSc(ϕ))\displaystyle\geq\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\prod_{c^{\prime}\in C_{S}(S^{\prime})}\sum_{w_{S^{\prime}c^{\prime}}\in A^{S^{\prime}c^{\prime}}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{S^{\prime}c^{\prime}}\left([w_{S^{\prime}c^{\prime}}]\cap A^{G}\right)-\Delta_{S^{\prime}c^{\prime}}(\phi)\right)
=S𝒮(𝒯)exp(ΔS(ϕ)|CS(S)|)cCS(S)ZSc(A,ϕ)\displaystyle=\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\mathop{\textrm{\rm exp}}\nolimits\left(-\Delta_{S^{\prime}}(\phi)|C_{S}(S^{\prime})|\right)\prod_{c^{\prime}\in C_{S}(S^{\prime})}Z_{S^{\prime}c^{\prime}}(A,\phi)
=S𝒮(𝒯)exp(ΔS(ϕ)|CS(S)|)ZS(A,ϕ)|CS(S)|,\displaystyle=\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\mathop{\textrm{\rm exp}}\nolimits\left(-\Delta_{S^{\prime}}(\phi)|C_{S}(S^{\prime})|\right)Z_{S^{\prime}}(A,\phi)^{|C_{S}(S^{\prime})|},

where we used that, for every gGg\in G, ZF(A,ϕ)=ZFg(A,ϕ)Z_{F}(A,\phi)=Z_{Fg}(A,\phi) and that ΔF(ϕ)=ΔFg(ϕ)\Delta_{F}(\phi)=\Delta_{Fg}(\phi). Thus,

(5) S𝒮(𝒯)ZS(A,ϕ)|CS(S)|ZS(A,ϕ)S𝒮(𝒯)exp(|CS(S)|ΔS(ϕ)).\displaystyle\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}Z_{S^{\prime}}(A,\phi)^{|C_{S}(S^{\prime})|}\leq Z_{S}(A,\phi)\cdot\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\mathop{\textrm{\rm exp}}\nolimits\left(|C_{S}(S^{\prime})|\Delta_{S^{\prime}}(\phi)\right).

Now, given a (S𝒯,ϵ)(S_{\mathcal{T}},\epsilon)-invariant set F(G)F\in\mathcal{F}(G), we have that

ZTF(ϕ)\displaystyle Z_{T_{F}}(\phi) S𝒮(𝒯)cCF(S)ZSc(ϕ)\displaystyle\leq\prod_{S\in\mathcal{S}(\mathcal{T})}\prod_{c\in C_{F}(S)}Z_{Sc}(\phi)
=S𝒮(𝒯)ZS(ϕ)|CF(S)|\displaystyle=\prod_{S\in\mathcal{S}(\mathcal{T})}Z_{S}(\phi)^{|C_{F}(S)|}
S𝒮(𝒯)(S𝒮(𝒯)cCS(S)ZSc(ϕ))|CF(S)|\displaystyle\leq\prod_{S\in\mathcal{S}(\mathcal{T})}\left(\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\prod_{c^{\prime}\in C_{S}(S^{\prime})}Z_{S^{\prime}c^{\prime}}(\phi)\right)^{|C_{F}(S)|}
=S𝒮(𝒯)S𝒮(𝒯)(ZS(ϕ)|CS(S)|)|CF(S)|.\displaystyle=\prod_{S\in\mathcal{S}(\mathcal{T})}\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}(Z_{S^{\prime}}(\phi)^{|C_{S}(S^{\prime})|})^{|C_{F}(S)|}.

Therefore, from equation (3), we obtain that

S𝒮(𝒯)\displaystyle\prod_{S\in\mathcal{S}(\mathcal{T})} S𝒮(𝒯)(ZS(ϕ)|CS(S)|)|CF(S)|\displaystyle\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}(Z_{S^{\prime}}(\phi)^{|C_{S}(S^{\prime})|})^{|C_{F}(S)|}
S𝒮(𝒯)S𝒮(𝒯)(11ϵZS(A,ϕ))|CS(S)||CF(S)|\displaystyle\leq\prod_{S\in\mathcal{S}(\mathcal{T})}\prod_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\left(\frac{1}{1-\epsilon}Z_{S^{\prime}}(A,\phi)\right)^{|C_{S}(S^{\prime})||C_{F}(S)|}
(11ϵ)|TF|S𝒮(𝒯)(ZS(A,ϕ)exp(S𝒮(𝒯)|CS(S)|ΔS(ϕ)))|CF(S)|\displaystyle\leq\left(\frac{1}{1-\epsilon}\right)^{|T_{F}|}\prod_{S\in\mathcal{S}(\mathcal{T})}\left(Z_{S}(A,\phi)\mathop{\textrm{\rm exp}}\nolimits\left(\sum_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}|C_{S}(S^{\prime})|\Delta_{S^{\prime}}(\phi)\right)\right)^{|C_{F}(S)|}
(11ϵ)|F|S𝒮(𝒯)exp(|S|(p(A,ϕ)+ϵ)+S𝒮(𝒯)|CS(S)|ΔS(ϕ))|CF(S)|,\displaystyle\leq\left(\frac{1}{1-\epsilon}\right)^{|F|}\prod_{S\in\mathcal{S}(\mathcal{T})}\mathop{\textrm{\rm exp}}\nolimits\left(|S|(p(A,\phi)+\epsilon)+\sum_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}|C_{S}(S^{\prime})|\Delta_{S^{\prime}}(\phi)\right)^{|C_{F}(S)|},

where the second inequality follows from equation (5) and the third from equation (4). Hence, if 0<ϵ<120<\epsilon<\frac{1}{2}, we have that log(11ϵ)2ϵ\log\left(\frac{1}{1-\epsilon}\right)\leq 2\epsilon, so

1|F|\displaystyle\frac{1}{|F|} logZTF(ϕ)\displaystyle\log Z_{T_{F}}(\phi)
log(11ϵ)+1|F|S𝒮(𝒯)|CF(S)|(|S|(p(A,ϕ)+ϵ)+S𝒮(𝒯)|CS(S)|ΔS(ϕ))\displaystyle\leq\log\left(\frac{1}{1-\epsilon}\right)+\frac{1}{|F|}\sum_{S\in\mathcal{S}(\mathcal{T})}|C_{F}(S)|\left(|S|(p(A,\phi)+\epsilon)+\sum_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}|C_{S}(S^{\prime})|\Delta_{S^{\prime}}(\phi)\right)
=log(11ϵ)+|TF||F|(p(A,ϕ)+ϵ)+S𝒮(𝒯)S𝒮(𝒯)|CF(S)||CS(S)||S||F|ΔS(ϕ)|S|\displaystyle=\log\left(\frac{1}{1-\epsilon}\right)+\frac{|T_{F}|}{|F|}(p(A,\phi)+\epsilon)+\sum_{S\in\mathcal{S}(\mathcal{T})}\sum_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\frac{|C_{F}(S)||C_{S}(S^{\prime})||S^{\prime}|}{|F|}\frac{\Delta_{S^{\prime}}(\phi)}{|S^{\prime}|}
2ϵ+(p(A,ϕ)+ϵ)+S𝒮(𝒯)S𝒮(𝒯)|CF(S)||CS(S)||S||F|ϵ\displaystyle\leq 2\epsilon+(p(A,\phi)+\epsilon)+\sum_{S\in\mathcal{S}(\mathcal{T})}\sum_{S^{\prime}\in\mathcal{S}(\mathcal{T}^{\prime})}\frac{|C_{F}(S)||C_{S}(S^{\prime})||S^{\prime}|}{|F|}\epsilon
=p(A,ϕ)+3ϵ+|TF||F|ϵ\displaystyle=p(A,\phi)+3\epsilon+\frac{|T_{F}|}{|F|}\epsilon
p(A,ϕ)+4ϵ.\displaystyle\leq p(A,\phi)+4\epsilon.

In addition,

ZF(ϕ)ZTF(ϕ)ZFTF(ϕ)ZTF(ϕ)Z1G(ϕ)|FTF|ZTF(ϕ)Z1G(ϕ)ϵ|F|,\displaystyle Z_{F}(\phi)\leq Z_{T_{F}}(\phi)Z_{F\setminus T_{F}}(\phi)\leq Z_{T_{F}}(\phi)Z_{1_{G}}(\phi)^{|F\setminus T_{F}|}\leq Z_{T_{F}}(\phi)Z_{1_{G}}(\phi)^{\epsilon|F|},

so, considering that p(A,ϕ)=infE(G)1|E|logZE(A,ϕ)p(A,\phi)=\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi) by Theorem 3.1, we have that

1|F|logZF(ϕ)infE(G)1|E|logZE(A,ϕ)+4ϵ+ϵlogZ1G(ϕ).\displaystyle\frac{1}{|F|}\log Z_{F}(\phi)\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi)+4\epsilon+\epsilon\cdot\log Z_{1_{G}}(\phi).

We conclude that, for every 0<ϵ<120<\epsilon<\frac{1}{2}, there exist A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G), and δ>0\delta>0 such that for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G),

1|F|logZF(ϕ)infE(G)1|E|logZE(A,ϕ)+ϵC\displaystyle\frac{1}{|F|}\log Z_{F}(\phi)\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi)+\epsilon\cdot C

where C=4+logZ1G(ϕ)C=4+\log Z_{1_{G}}(\phi). Since ϵ\epsilon was arbitrary, we conclude the result. ∎

Now we can prove the following generalization of Theorem 3.1.

Theorem 3.8.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable and uniformly continuous potential with finite oscillation. Then, p(ϕ)p(\phi) exists and p(ϕ)=infE(G)1|E|logZE(ϕ)p(\phi)=\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(\phi). Moreover, p(ϕ)=supA()p(A,ϕ)p(\phi)=\sup_{A\in\mathcal{F}(\mathbb{N})}p(A,\phi).

Proof.

By Proposition 3.7, for every 12>ϵ>0\frac{1}{2}>\epsilon>0, there exist A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G),

1|F|logZF(ϕ)infE(G)1|E|logZE(A,ϕ)+ϵ.\displaystyle\frac{1}{|F|}\log Z_{F}(\phi)\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi)+\epsilon.

Therefore, for every such FF,

infE(G)1|E|logZE(ϕ)\displaystyle\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(\phi) 1|F|logZF(ϕ)\displaystyle\leq\frac{1}{|F|}\log Z_{F}(\phi)
infE(G)1|E|logZE(A,ϕ)+ϵ\displaystyle\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(A,\phi)+\epsilon
infE(G)1|E|logZE(ϕ)+ϵ.\displaystyle\leq\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(\phi)+\epsilon.

Thus, limFG1|F|logZF(ϕ)=infE(G)1|E|logZE(ϕ)\lim_{F\to G}\frac{1}{|F|}\log Z_{F}(\phi)=\inf_{E\in\mathcal{F}(G)}\frac{1}{|E|}\log Z_{E}(\phi), p(ϕ)p(\phi) exists, and there exists A()A\in\mathcal{F}(\mathbb{N}) such that

p(ϕ)p(A,ϕ)+ϵp(ϕ)+ϵ,\displaystyle p(\phi)\leq p(A,\phi)+\epsilon\leq p(\phi)+\epsilon,

so p(ϕ)=supA()p(A,ϕ)p(\phi)=\sup_{A\in\mathcal{F}(\mathbb{N})}p(A,\phi). ∎

4. Permutations and specifications

In order to define conformal and DLR measures it will be crucial to introduce coordinate-wise permutations and specifications. We begin by describing and exploring some properties of coordinate-wise permutations.

4.1. Coordinate-wise permutations

Let SS_{\mathbb{N}} be the set of all permutations of \mathbb{N}. Following [38, 45], we now introduce a class of local maps on XX. Given an exhausting sequence {Em}m\{E_{m}\}_{m}, this class will allow us to understand how ϕEm(x)\phi_{E_{m}}(x) behaves if xx is changed at finitely many sites and it will be central when defining conformal measures in §5.

Definition 3.

Given K(G)K\in\mathcal{F}(G), denote by K\mathcal{E}_{K} the set of all maps τ:XX\tau\colon X\to X such that

τ(x)g={τg(xg),if gK;xg,if gK;\tau(x)_{g}=\begin{cases}\tau_{g}(x_{g}),&\text{if }g\in K;\\ x_{g},&\text{if }g\notin K;\end{cases}

where τgS\tau_{g}\in S_{\mathbb{N}}. We usually denote τ\tau by τK\tau_{K} to emphasize the set KK.

Let =K(G)K\mathcal{E}=\bigcup_{K\in\mathcal{F}(G)}\mathcal{E}_{K} and notice that there is a natural action of GG on \mathcal{E} given by

(gτK)(x)=gτK(g1x),(g\cdot\tau_{K})(x)=g\cdot\tau_{K}(g^{-1}\cdot x),

where gGg\in G, xXx\in X, K(G)K\in\mathcal{F}(G), τKK\tau_{K}\in\mathcal{E}_{K}, and gτKKg1g\cdot\tau_{K}\in\mathcal{E}_{Kg^{-1}}. In order to avoid ambiguity, we will denote gτKg\cdot\tau_{K} by τKg1\tau_{Kg^{-1}} and that will be enough for our purposes.

We can also restrict ourselves to permutations over a finite alphabet. More explicitly, for A()A\in\mathcal{F}(\mathbb{N}) and K(G)K\in\mathcal{F}(G), define

K,A={τK:hK,τh|Ac=Id|Ac}.\mathcal{E}_{K,A}=\{\tau\in\mathcal{E}_{K}:\forall h\in K,\,\tau_{h}|_{A^{c}}=\textnormal{Id}_{\mathbb{N}}|_{A^{c}}\}.

Notice that \mathcal{E} is a group with the composition generated by single-site permutations τg\tau_{g}, where K\mathcal{E}_{K} and K,A\mathcal{E}_{K,A} are subgroups. Moreover, observe that if ghg\neq h, then τgτh=τhτg\tau_{g}\tau_{h}=\tau_{h}\tau_{g}.

We will also consider a particular type of permutations, which are defined below.

Definition 4.

Given K(G)K\in\mathcal{F}(G) and w,wXKw,w^{\prime}\in X_{K}, let τw,w:XX\tau_{w,w^{\prime}}\colon X\to X be the map defined as

τw,w(x)={wxKc,if xK=w;wxKc,if xK=w;x,otherwise.\tau_{w,w^{\prime}}(x)=\begin{cases}wx_{K^{c}},&\text{if }x_{K}=w^{\prime};\\ w^{\prime}x_{K^{c}},&\text{if }x_{K}=w;\\ x,&\text{otherwise}.\end{cases}

It is clear that τw,wK\tau_{w,w^{\prime}}\in\mathcal{E}_{K}, τw,w=τw,w\tau_{w,w^{\prime}}=\tau_{w^{\prime},w} and that τw,w\tau_{w,w^{\prime}} is an involution, that is, it is its own inverse. Moreover, there exists A()A\in\mathcal{F}(\mathbb{N}), namely A=w(K)w(K)A=w(K)\cup w^{\prime}(K), such that τw,wK,A\tau_{w,w^{\prime}}\in\mathcal{E}_{K,A}. For τ\tau\in\mathcal{E} and F(G)F\in\mathcal{F}(G), define ϕFτ:X\phi^{\tau}_{F}\colon X\to\mathbb{R} as

(6) ϕFτ(x)=ϕFτ1(x)ϕF(x).\phi^{\tau}_{F}(x)=\phi_{F}\circ\tau^{-1}(x)-\phi_{F}(x).

Notice that, for τK\tau\in\mathcal{E}_{K},

ϕFτ(x)\displaystyle\phi^{\tau}_{F}(x) =gFϕ(gτK1(x))ϕ(gx)\displaystyle=\sum_{g\in F}\phi(g\cdot\tau_{K}^{-1}(x))-\phi(g\cdot x)
=gFϕ(gτK1(g1(gx)))ϕ(gx)\displaystyle=\sum_{g\in F}\phi(g\cdot\tau_{K}^{-1}(g^{-1}\cdot(g\cdot x)))-\phi(g\cdot x)
=gFϕ(τKg11(gx)))ϕ(gx).\displaystyle=\sum_{g\in F}\phi(\tau_{Kg^{-1}}^{-1}(g\cdot x)))-\phi(g\cdot x).
Lemma 4.1.

Let K(G)K\in\mathcal{F}(G) and ϕ:X\phi\colon X\to\mathbb{R} be a potential. Then, for every τKK\tau_{K}\in\mathcal{E}_{K} and every E,F(G)E,F\in\mathcal{F}(G) with FEF\subseteq E,

ϕEτKϕFτKgGFϕτKg11ϕ.\|\phi_{E}^{\tau_{K}}-\phi_{F}^{\tau_{K}}\|_{\infty}\leq\sum_{g\in G\setminus F}\left\|\phi\circ\tau_{Kg^{-1}}^{-1}-\phi\right\|_{\infty}.
Proof.

Let K,E,F(G)K,E,F\in\mathcal{F}(G) and τKK\tau_{K}\in\mathcal{E}_{K} be as in the statement of the Lemma. Then, it is easy to verify that, for any xXx\in X, (ϕEτKϕFτK)(x)=gEF[ϕ(τKg11(gx))ϕ(gx)](\phi^{\tau_{K}}_{E}-\phi^{\tau_{K}}_{F})(x)=\sum_{g\in E\setminus F}\left[\phi\left(\tau_{Kg^{-1}}^{-1}(g\cdot x)\right)-\phi(g\cdot x)\right]. Thus,

ϕEτKϕFτK\displaystyle\|\phi_{E}^{\tau_{K}}-\phi_{F}^{\tau_{K}}\|_{\infty} =supxX|ϕEτK(x)ϕFτK(x)|\displaystyle=\sup_{x\in X}\left|\phi_{E}^{\tau_{K}}(x)-\phi_{F}^{\tau_{K}}(x)\right|
=supxX|gEF[ϕ(τKg11(gx))ϕ(gx)]|\displaystyle=\sup_{x\in X}\left|\sum_{g\in E\setminus F}\left[\phi\left(\tau_{Kg^{-1}}^{-1}(g\cdot x)\right)-\phi(g\cdot x)\right]\right|
supxXgEF|ϕ(τKg11(gx))ϕ(gx)|\displaystyle\leq\sup_{x\in X}\sum_{g\in E\setminus F}\left|\phi\left(\tau_{Kg^{-1}}^{-1}(g\cdot x)\right)-\phi(g\cdot x)\right|
gEFsupxX|ϕ(τKg11(gx))ϕ(gx)|\displaystyle\leq\sum_{g\in E\setminus F}\sup_{x\in X}\left|\phi\left(\tau_{Kg^{-1}}^{-1}(g\cdot x)\right)-\phi(g\cdot x)\right|
=gEFϕτKg11ϕ\displaystyle=\sum_{g\in E\setminus F}\left\|\phi\circ\tau_{Kg^{-1}}^{-1}-\phi\right\|_{\infty}
gGFϕτKg11ϕ.\displaystyle\leq\sum_{g\in G\setminus F}\left\|\phi\circ\tau_{Kg^{-1}}^{-1}-\phi\right\|_{\infty}.

Given a potential ϕ:X\phi\colon X\to\mathbb{R} with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, the next theorem tells us that the asymptotic behaviour of ϕEm(x)\phi_{E_{m}}(x) is essentially independent of the value of the configuration xx at finite sets K(G)K\in\mathcal{F}(G). The reader can compare the next result with [38, Lemma 5.1.6].

Theorem 4.2.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, given any (possibly different) exhausting sequence {E~m}m\{\tilde{E}_{m}\}_{m}, for all K(G)K\in\mathcal{F}(G) and for all τKK\tau_{K}\in\mathcal{E}_{K}, the limit

ϕτK:=limmϕE~mτK\phi_{*}^{\tau_{K}}:=\lim_{m\to\infty}\phi^{\tau_{K}}_{\tilde{E}_{m}}

exists uniformly on XX and on K\mathcal{E}_{K}. Moreover, such limit does not depend on the exhausting sequence.

Proof.

First, suppose that KK is a singleton {h}\{h\} for some hGh\in G and let ϵ>0\epsilon>0. Since ϕ\phi has summable variation according to {Em}m\{E_{m}\}_{m}, there exists m0m_{0}\in\mathbb{N} such that mm0|Em+11Em1|δEm(ϕ)<ϵ\sum_{m\geq m_{0}}|E_{m+1}^{-1}\setminus E_{m}^{-1}|\delta_{E_{m}}(\phi)<\epsilon. Now, consider {E~m}m\{\tilde{E}_{m}\}_{m} another (possibly different) exhausting sequence. Then, there exists m1m0m_{1}\geq m_{0} such that Em01hE~mE^{-1}_{m_{0}}h\subseteq\tilde{E}_{m}, for all mm1m\geq m_{1}. On the other hand, since {Em}m\{E_{m}\}_{m} is an exhausting sequence, for every mm1m\geq m_{1}, there exists kmk_{m}\in\mathbb{N} such that for all kkmk\geq k_{m}, E~mEk\tilde{E}_{m}\subseteq E_{k}. Therefore, by Lemma 4.1, for every mm1m\geq m_{1} and every kkmk\geq k_{m},

ϕEkτhϕE~mτhgGE~mϕτhg11ϕ.\displaystyle\left\|\phi^{\tau_{h}}_{E_{k}}-\phi^{\tau_{h}}_{\tilde{E}_{m}}\right\|_{\infty}\leq\sum_{g\in G\setminus\tilde{E}_{m}}\left\|\phi\circ\tau^{-1}_{hg^{-1}}-\phi\right\|_{\infty}.

Moreover, since Em01hE~mE_{m_{0}}^{-1}h\subseteq\tilde{E}_{m}, we obtain that GE~mGEm01hG\setminus\tilde{E}_{m}\subseteq G\setminus E_{m_{0}}^{-1}h, so that

gGE~mϕτhg11ϕ\displaystyle\sum_{g\in G\setminus\tilde{E}_{m}}\left\|\phi\circ\tau^{-1}_{hg^{-1}}-\phi\right\|_{\infty} gGEm01hϕτhg11ϕ\displaystyle\leq\sum_{g\in G\setminus E_{m_{0}}^{-1}h}\left\|\phi\circ\tau^{-1}_{hg^{-1}}-\phi\right\|_{\infty}
=gGEm01ϕτg11ϕ\displaystyle=\sum_{g\in G\setminus E_{m_{0}}^{-1}}\left\|\phi\circ\tau^{-1}_{g^{-1}}-\phi\right\|_{\infty}
=mm0gEm+11Em1ϕτg11ϕ\displaystyle=\sum_{m\geq m_{0}}\sum_{g\in E_{m+1}^{-1}\setminus E_{m}^{-1}}\left\|\phi\circ\tau_{g^{-1}}^{-1}-\phi\right\|_{\infty}
mm0gEm+11Em1δEm(ϕ)\displaystyle\leq\sum_{m\geq m_{0}}\sum_{g\in E_{m+1}^{-1}\setminus E_{m}^{-1}}\delta_{E_{m}}(\phi)
=mm0|Em+11Em1|δEm(ϕ)<ϵ.\displaystyle=\sum_{m\geq m_{0}}\left|E_{m+1}^{-1}\setminus E_{m}^{-1}\right|\delta_{E_{m}}(\phi)<\epsilon.

Therefore, for every ϵ>0\epsilon>0, there exists m1m0m_{1}\geq m_{0} such that for every mm1m\geq m_{1}, there exists kmk_{m} such that for every kkmk\geq k_{m},

ϕEkτhϕE~mτh<ϵ.\left\|\phi^{\tau_{h}}_{E_{k}}-\phi^{\tau_{h}}_{\tilde{E}_{m}}\right\|_{\infty}<\epsilon.

Notice that, in the particular case that {E~m}m\{\tilde{E}_{m}\}_{m} is the same as {Em}m\{E_{m}\}_{m}, one just need to take km=mk_{m}=m and the same inequality would follow. This proves that {ϕE~mτh}m\{\phi^{\tau_{h}}_{\tilde{E}_{m}}\}_{m} is a Cauchy sequence for any τh{h}\tau_{h}\in\mathcal{E}_{\{h\}}, which implies that the uniform limit ϕτh=limmϕEmτh\phi_{*}^{\tau_{h}}=\lim_{m\to\infty}\phi^{\tau_{h}}_{E_{m}} exists. On the other hand, if {E~m}m\{\tilde{E}_{m}\}_{m} is another exhausting sequence, this proves that ϕτh=limmϕE~mτh\phi_{*}^{\tau_{h}}=\lim_{m\to\infty}\phi^{\tau_{h}}_{\tilde{E}_{m}}, i.e., the limit is independent of the exhausting sequence provided ϕ\phi has summable variation according to some exhausting sequence.

Now, let’s consider a general K(G)K\in\mathcal{F}(G) and write K={h1,,h|K|}K=\{h_{1},\dots,h_{|K|}\}. Then, for each mm\in\mathbb{N},

ϕEmτK1ϕEm=i=0|K|1(ϕEmτ{h1,,hi+1}1ϕEmτ{h1,,hi}1)=i=0|K|1ϕEmτhi+1τ{h1,,hi}1,\displaystyle\phi_{E_{m}}\circ\tau^{-1}_{K}-\phi_{E_{m}}=\sum_{i=0}^{|K|-1}\left(\phi_{E_{m}}\circ\tau^{-1}_{\{h_{1},\dots,h_{i+1}\}}-\phi_{E_{m}}\circ\tau^{-1}_{\{h_{1},\dots,h_{i}\}}\right)=\sum_{i=0}^{|K|-1}\phi^{\tau_{h_{i+1}}}_{E_{m}}\circ\tau^{-1}_{\{h_{1},\dots,h_{i}\}},

where we regard τ\tau_{\emptyset} as the identity, so the first equality follows from the fact that the considered sum is telescopic. Therefore, by considering the uniform convergence for singletons,

limmϕEmτK\displaystyle\lim_{m\to\infty}\phi^{\tau_{K}}_{E_{m}} =limmi=0|K|1ϕEmτhi+1τ{h1,,hi}1\displaystyle=\lim_{m\to\infty}\sum_{i=0}^{|K|-1}\phi^{\tau_{h_{i+1}}}_{E_{m}}\circ\tau^{-1}_{\{h_{1},\dots,h_{i}\}}
=i=0|K|1limmϕEmτhi+1τ{h1,,hi}1\displaystyle=\sum_{i=0}^{|K|-1}\lim_{m\to\infty}\phi^{\tau_{h_{i+1}}}_{E_{m}}\circ\tau^{-1}_{\{h_{1},\dots,h_{i}\}}
=i=0|K|1ϕτhi+1τ{h1,,hi}1,\displaystyle=\sum_{i=0}^{|K|-1}\phi^{\tau_{h_{i+1}}}_{*}\circ\tau^{-1}_{\{h_{1},\dots,h_{i}\}},

which concludes the result. ∎

Corollary 4.3.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for all K(G)K\in\mathcal{F}(G) and for all τKK\tau_{K}\in\mathcal{E}_{K},

ϕτK(gx)=ϕτKg(x),\phi_{*}^{\tau_{K}}(g\cdot x)=\phi_{*}^{\tau_{Kg}}(x),

for all gGg\in G and xXx\in X.

Proof.

Notice that, given gGg\in G and xXx\in X, we have that τK1(gx)=gτKg1(x)\tau_{K}^{-1}(g\cdot x)=g\cdot\tau^{-1}_{Kg}(x), so that

ϕτK(gx)=limmϕEmτK(gx)=limmϕEmgτKg(x)=ϕτKg(x),\phi_{*}^{\tau_{K}}(g\cdot x)=\lim_{m\to\infty}\phi_{E_{m}}^{\tau_{K}}(g\cdot x)=\lim_{m\to\infty}\phi_{E_{m}g}^{\tau_{Kg}}(x)=\phi_{*}^{\tau_{Kg}}(x),

since {Emg}m\{E_{m}g\}_{m} is also an exhausting sequence.

Proposition 4.4.

Let ϕ:X\phi\colon X\to\mathbb{R} be a potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for every F(G)F\in\mathcal{F}(G) and τ\tau in F\mathcal{E}_{F},

ϕτϕFτVF(ϕ).\|\phi_{*}^{\tau}-\phi^{\tau}_{F}\|_{\infty}\leq V_{F}(\phi).
Proof.

Let F(G)F\in\mathcal{F}(G). From Lemma 4.1, we know that

ϕEmτFϕFτFgGFϕτFg11ϕ,\left\|\phi^{\tau_{F}}_{E_{m}}-\phi^{\tau_{F}}_{F}\right\|_{\infty}\leq\sum_{g\in G\setminus F}\left\|\phi\circ\tau_{Fg^{-1}}^{-1}-\phi\right\|_{\infty},

for every mm\in\mathbb{N} such that FEmF\subseteq E_{m}. Therefore, by Theorem 4.2,

ϕτFϕFτF=limmϕEmτFϕFτFgGFϕτFg11ϕ.\left\|\phi_{*}^{\tau_{F}}-\phi^{\tau_{F}}_{F}\right\|_{\infty}=\lim_{m\to\infty}\left\|\phi^{\tau_{F}}_{E_{m}}-\phi^{\tau_{F}}_{F}\right\|_{\infty}\leq\sum_{g\in G\setminus F}\left\|\phi\circ\tau_{Fg^{-1}}^{-1}-\phi\right\|_{\infty}.

Now, given mm\in\mathbb{N}, notice that g(Em1F)cFg1Em=g\in(E_{m}^{-1}F)^{c}\iff Fg^{-1}\cap E_{m}=\emptyset, so that ϕτFg11ϕδEm(ϕ)\left\|\phi\circ\tau_{Fg^{-1}}^{-1}-\phi\right\|_{\infty}\leq\delta_{E_{m}}(\phi). Considering this, we have that

gGFϕτFg11ϕ\displaystyle\sum_{g\in G\setminus F}\left\|\phi\circ\tau_{Fg^{-1}}^{-1}-\phi\right\|_{\infty} =m=1gEm+11FEm1FϕτFg11ϕ\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m+1}^{-1}F\setminus E_{m}^{-1}F}\left\|\phi\circ\tau_{Fg^{-1}}^{-1}-\phi\right\|_{\infty}
m=1|Em+11FEm1F|δEm(ϕ)\displaystyle\leq\sum_{m=1}^{\infty}|E^{-1}_{m+1}F\setminus E^{-1}_{m}F|\cdot\delta_{E_{m}}(\phi)
=VF(ϕ).\displaystyle=V_{F}(\phi).

4.2. Specifications

This section tackles results about specifications, a concept related to DLR measures. More precisely, DLR measures can be defined using a special kind of specifications, but here we begin by presenting some more general results.

Let \mathcal{B} be the Borel σ\sigma-algebra, that is, the σ\sigma-algebra generated by the cylinder sets, and, for each K(G)K\in\mathcal{F}(G), let K\mathcal{B}_{K} be the σ\sigma-algebra generated by cylinder sets [w][w], with wXKw\in X_{K}. Now, a specification in our context, will mean a family γ=(γK)K(G)\gamma=(\gamma_{K})_{K\in\mathcal{F}(G)} of maps γK:×X[0,1]\gamma_{K}\colon\mathcal{B}\times X\to[0,1] such that

  1. i)

    for each xXx\in X, the map BγK(B,x)B\mapsto\gamma_{K}(B,x) is a probability measure on (X)\mathcal{M}(X);

  2. ii)

    for each BB\in\mathcal{B}, the map xγK(B,x)x\mapsto\gamma_{K}(B,x) is Kc\mathcal{B}_{K^{c}}-measurable;

  3. iii)

    (proper) for every BB\in\mathcal{B} and CKcC\in\mathcal{B}_{K^{c}}, γK(BC,)=γK(B,)𝟙C\gamma_{K}\left(B\cap C,\cdot\right)=\gamma_{K}(B,\cdot)\mathbbm{1}_{C}; and

  4. iv)

    if FKF\subseteq K, then γKγF=γK\gamma_{K}\gamma_{F}=\gamma_{K}, where γKγF(B,x)=γK(dy,x)γF(B,y),\gamma_{K}\gamma_{F}(B,x)=\int\gamma_{K}(dy,x)\gamma_{F}(B,y), for BB\in\mathcal{B} and xXx\in X.

In other words, γ\gamma is a particular family of proper probability kernels that satisfies consistency condition (iv). An element γK\gamma_{K} in the specification maps each μ(X)\mu\in\mathcal{M}(X) to μγK(X)\mu\gamma_{K}\in\mathcal{M}(X), where

μγK(B)=γK(B,x)𝑑μ(x),\mu\gamma_{K}(B)=\int\gamma_{K}(B,x)d\mu(x),

and each \mathcal{B}-measurable function h:Xh:X\to\mathbb{R} to a Kc\mathcal{B}_{K^{c}}-measurable function γKh:X\gamma_{K}h:X\to\mathbb{R} given by

γKh(y)=h(y)γK(dy,x)𝑑μ(x).\gamma_{K}h(y)=\int h(y)\gamma_{K}(dy,x)d\mu(x).

It can be checked that (μγK)(h)=μ(γKh)(\mu\gamma_{K})(h)=\mu(\gamma_{K}h). The probability measures on the set

𝒢(γ)={μ(X):μ(B|Kc)=γK(B,)μ-a.s., for all B and K(G)}\displaystyle\mathscr{G}(\gamma)=\{\mu\in\mathcal{M}(X):\mu\left(B\,|\,\mathcal{B}_{K^{c}}\right)=\gamma_{K}\left(B,\cdot\right)\,\mu\text{-a.s., for all }B\in\mathcal{B}\text{ and }K\in\mathcal{F}(G)\}

are said to be admitted by the specification γ\gamma.

Lemma 4.5.

[30, Remark 1.24] Let γ\gamma be a specification and μ(X)\mu\in\mathcal{M}(X). Then, μ𝒢(γ)\mu\in\mathscr{G}(\gamma) if and only if μγK=μ\mu\gamma_{K}=\mu, for all K(G)K\in\mathcal{F}(G).

Now, we restrict ourselves to a particular kind of specification. Namely, given an exhausting sequence of finite sets {Em}m\{E_{m}\}_{m} and ϕ:X\phi\colon X\to\mathbb{R} an exp-summable potential with summable variation according to {Em}m\{E_{m}\}_{m}, consider γ=(γK)K(G)\gamma=(\gamma_{K})_{K\in\mathcal{F}(G)} the specification coming from ϕ\phi, where each γK:×X[0,1]\gamma_{K}\colon\mathcal{B}\times X\to[0,1] is given by

(7) γK(B,x):=limmwXKexp(ϕEm(wxKc))𝟙{wxKcB}vXKexp(ϕEm(vxKc)),\gamma_{K}(B,x):=\lim_{m\to\infty}\frac{\sum_{w\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})\right)},

for each BB\in\mathcal{B} and xXx\in X. The collection γ\gamma is a (Gibbsian) specification. The expression in equation (7) is well-defined due to the following proposition.

Proposition 4.6.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If K(G)K\in\mathcal{F}(G), the limit

γK([w],x)=limmexp(ϕEm(wxKc))vXKexp(ϕEm(vxKc))\gamma_{K}([w],x)=\lim_{m\to\infty}\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})\right)}

exists for each wXKw\in X_{K}, uniformly on XX. Furthermore, for every BB\in\mathcal{B} and every xXx\in X, it holds that

(8) γK(B,x)=wXKγK([w],x)𝟙{wxKcB}.\gamma_{K}(B,x)=\sum_{w\in X_{K}}\gamma_{K}\left([w],x\right)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}.

In order to prove Proposition 4.6, we require two lemmas, which we state and prove next.

Lemma 4.7.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for any K(G)K\in\mathcal{F}(G) and for any mm\in\mathbb{N} such that KEmK\subseteq E_{m},

|ϕEmτw,v(wxKc)(supϕK[v]supϕK[w])|ΔK(ϕ)+VK(ϕ)\left|\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}})-(\sup\phi_{K}[v]-\sup\phi_{K}[w])\right|\leq\Delta_{K}(\phi)+V_{K}(\phi)

for every v,wXKv,w\in X_{K} and xXx\in X.

Proof.

Let K(G)K\in\mathcal{F}(G) and x,yXx,y\in X be such that xGK=yGKx_{G\setminus K}=y_{G\setminus K}. Notice that for any gGg\in G, (gx)GKg1=(gy)GKg1(g\cdot x)_{G\setminus Kg^{-1}}=(g\cdot y)_{G\setminus Kg^{-1}}. In addition, given mm\in\mathbb{N}, we have that g(Em1K)cKg1Em=g\in\left(E_{m}^{-1}K\right)^{c}\iff Kg^{-1}\cap E_{m}=\emptyset. In particular, if g(Em1K)cg\in\left(E_{m}^{-1}K\right)^{c}, we have that |ϕ(gx)ϕ(gy)|δEm(ϕ)|\phi(g\cdot x)-\phi(g\cdot y)|\leq\delta_{E_{m}}(\phi). Considering this, we obtain that

gGK|ϕ(gx)ϕ(gy)|\displaystyle\sum_{g\in G\setminus K}|\phi(g\cdot x)-\phi(g\cdot y)| =m=1gEm+11KEm1K|ϕ(gx)ϕ(gy)|\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m+1}^{-1}K\setminus E_{m}^{-1}K}|\phi(g\cdot x)-\phi(g\cdot y)|
m=1gEm+11KEm1KδEm(ϕ)\displaystyle\leq\sum_{m=1}^{\infty}\sum_{g\in E_{m+1}^{-1}K\setminus E_{m}^{-1}K}\delta_{E_{m}}(\phi)
=m=1|Em+11KEm1K|δEm(ϕ)\displaystyle=\sum_{m=1}^{\infty}|E_{m+1}^{-1}K\setminus E_{m}^{-1}K|\cdot\delta_{E_{m}}(\phi)
=VK(ϕ).\displaystyle=V_{K}(\phi).

Now, let m0m_{0}\in\mathbb{N} be the smallest index such that KEm0K\subseteq E_{m_{0}}. Then, for every mm0m\geq m_{0}, every xXx\in X, and every v,wXKv,w\in X_{K}, we have that

ϕEmτw,v(wxKc)\displaystyle\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}}) =ϕEm(vxKc)ϕEm(wxKc)\displaystyle=\phi_{E_{m}}(vx_{K^{c}})-\phi_{E_{m}}(wx_{K^{c}})
ϕK(vxKc)ϕK(wxKc)+gGK|ϕ(g(vxKc))ϕ(g(wxKc))|\displaystyle\leq\phi_{K}(vx_{K^{c}})-\phi_{K}(wx_{K^{c}})+\sum_{g\in G\setminus K}|\phi(g\cdot(vx_{K^{c}}))-\phi(g\cdot(wx_{K^{c}}))|
ϕK(vxKc)ϕK(wxKc)+VK(ϕ)\displaystyle\leq\phi_{K}(vx_{K^{c}})-\phi_{K}(wx_{K^{c}})+V_{K}(\phi)
supϕK[v]supϕK[w]+ΔK(ϕ)+VK(ϕ),\displaystyle\leq\sup\phi_{K}[v]-\sup\phi_{K}[w]+\Delta_{K}(\phi)+V_{K}(\phi),

and, similarly,

ϕEmτw,v(wxKc))\displaystyle\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}})) supϕK[v]supϕK[w]ΔK(ϕ)VK(ϕ),\displaystyle\geq\sup\phi_{K}[v]-\sup\phi_{K}[w]-\Delta_{K}(\phi)-V_{K}(\phi),

so we conclude that

|ϕEmτw,v(wxKc)(supϕK[v]supϕK[w])|ΔK(ϕ)+VK(ϕ).|\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}})-(\sup\phi_{K}[v]-\sup\phi_{K}[w])|\leq\Delta_{K}(\phi)+V_{K}(\phi).

Lemma 4.8.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for any K(G)K\in\mathcal{F}(G) and wXKw\in X_{K},

0<vXKexp(ϕτw,v(wxKc))=limmvXKexp(ϕEmτw,v(wxKc)),0<\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau_{w,v}}(wx_{K^{c}}))=\lim_{m\to\infty}\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits(\phi^{\tau_{w,v}}_{E_{m}}(wx_{K^{c}})),

uniformly on XX.

Proof.

Given K(G)K\in\mathcal{F}(G), wXKw\in X_{K}, and xXx\in X, consider the sequence of functions fm:XKf_{m}\colon X_{K}\to\mathbb{R} given by fm(v):=exp(ϕEmτw,v(wxKc)).f_{m}(v):=\mathop{\textrm{\rm exp}}\nolimits(\phi^{\tau_{w,v}}_{E_{m}}(wx_{K^{c}})). By Theorem 4.2, we have that {fm}m\{f_{m}\}_{m} converges pointwise (in vv) to exp(ϕτw,v(wxKc))\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau_{w,v}}(wx_{K^{c}})), uniformly on XX. In addition, by Lemma 4.7, there exist m0m_{0}\in\mathbb{N} and a constant C=exp(ΔK(ϕ)+VK(ϕ))>0C=\mathop{\textrm{\rm exp}}\nolimits(\Delta_{K}(\phi)+V_{K}(\phi))>0 such that for every mm0m\geq m_{0} and for every vXKv\in X_{K},

C1h(v)fm(v)Ch(v),C^{-1}\cdot h(v)\leq f_{m}(v)\leq C\cdot h(v),

where h(v):=exp(supϕK[w])exp(supϕK[v])h(v):=\mathop{\textrm{\rm exp}}\nolimits(-\sup\phi_{K}[w])\cdot\mathop{\textrm{\rm exp}}\nolimits(\sup\phi_{K}[v]). Notice that

vXKh(v)=exp(supϕK[w])ZK(ϕ),\sum_{v\in X_{K}}h(v)=\mathop{\textrm{\rm exp}}\nolimits(-\sup\phi_{K}[w])\cdot Z_{K}(\phi),

so hh (and therefore, ChC\cdot h) is integrable with respect to the counting measure in XKX_{K}. Therefore, by the Dominated Convergence Theorem, if follows that

vXKexp(ϕτw,v(wxKc))\displaystyle\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,v}}(wx_{K^{c}})\right) =vXKlimmexp(ϕEmτw,v(wxKc))\displaystyle=\sum_{v\in X_{K}}\lim_{m}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}})\right)
=limmvXKexp(ϕEmτw,v(wxKc))\displaystyle=\lim_{m}\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}^{\tau_{w,v}}(wx_{K^{c}})\right)
limmvXKC1h(v)\displaystyle\geq\lim_{m}\sum_{v\in X_{K}}C^{-1}\cdot h(v)
=C1exp(supϕK[w])ZK(ϕ)>0.\displaystyle=C^{-1}\cdot\mathop{\textrm{\rm exp}}\nolimits(-\sup\phi_{K}[w])\cdot Z_{K}(\phi)>0.

Proof (of Proposition 4.6).

First, note that for any given K(G)K\in\mathcal{F}(G),

vXKexp(ϕEm(vxKc))>0.\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})\right)>0.

for all mm\in\mathbb{N} and, due to Lemma 4.8, the left-hand side is bounded away from zero uniformly in mm. Furthermore, for each wXKw\in X_{K},

exp(ϕEm(wxKc))vXKexp(ϕEm(vxKc))\displaystyle\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})\right)} =1vXKexp(ϕEm(vxKc)ϕEm(wxKc)))=1vXKexp(ϕEmτw,v(wxKc)).\displaystyle=\frac{1}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})-\phi_{E_{m}}(wx_{K^{c}}))\right)}=\frac{1}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits(\phi^{\tau_{w,v}}_{E_{m}}(wx_{K^{c}}))}.

Therefore, uniformly on XX,

limmexp(ϕEm(wxKc))vXKexp(ϕEm(vxKc))=1limmvXKexp(ϕEmτw,v(wxKc))\displaystyle\lim_{m\to\infty}\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(vx_{K^{c}})\right)}=\frac{1}{\lim_{m\to\infty}\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits(\phi^{\tau_{w,v}}_{E_{m}}(wx_{K^{c}}))} =1vXKexp(ϕτw,v(wxKc)),\displaystyle=\frac{1}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau_{w,v}}(wx_{K^{c}}))},

again due to Lemma 4.8. Now, let BB\in\mathcal{B} and xXx\in X. Then, uniformly on XX,

wXK\displaystyle\sum_{w\in X_{K}} γK([w],x)𝟙{wxKcB}\displaystyle\gamma_{K}\left([w],x\right)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}
=wXKlimmexp(ϕEm(wxKc))𝟙{wxKc[w]}𝟙{wxKcB}vXKexp(ϕEm(vxKc))\displaystyle=\sum_{w\in X_{K}}\lim_{m\to\infty}\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)\mathbbm{1}_{\{wx_{K^{c}}\in[w]\}}\mathbbm{1}_{\{wx_{K^{c}}\in B\}}}{\sum_{v^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(v^{\prime}x_{K^{c}})\right)}
=limmwXKexp(ϕEm(wxKc))𝟙{wxKc[w]}𝟙{wxKcB}vXKexp(ϕEm(vxKc))\displaystyle=\lim_{m\to\infty}\frac{\sum_{w\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)\mathbbm{1}_{\{wx_{K^{c}}\in[w]\}}\mathbbm{1}_{\{wx_{K^{c}}\in B\}}}{\sum_{v^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(v^{\prime}x_{K^{c}})\right)}
=limmwXKexp(ϕEm(wxKc))𝟙{wxKcB}vXKexp(ϕEm(vxKc))\displaystyle=\lim_{m\to\infty}\frac{\sum_{w\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}}{\sum_{v^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(v^{\prime}x_{K^{c}})\right)}
=γK(B,x),\displaystyle=\gamma_{K}(B,x),

where the exchange of the limit and the sum follows from Lemma 4.8. ∎

Proposition 4.9.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for every K(G)K\in\mathcal{F}(G), wXKw\in X_{K}, and xXx\in X, the equation

γK([w],x)=exp(ϕτw,v(vxKc))wXKexp(ϕτw,v(vxKc))\gamma_{K}([w],x)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})\right)}

holds for every vXKv\in X_{K}.

Proof.

Let K(G)K\in\mathcal{F}(G), wXKw\in X_{K} and xXx\in X. Then, for any vXKv\in X_{K},

limmexp(ϕEm(wxKc))wXKexp(ϕEm(wxKc))\displaystyle\lim_{m\to\infty}\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}(wx_{K^{c}})\right)}{\sum_{{w^{\prime}}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}({w^{\prime}}x_{K^{c}})\right)} =limmexp(ϕEmτw,v1ϕEm)(vxKc)wXKexp(ϕEmτw,v1ϕEm)(vxKc)\displaystyle=\lim_{m\to\infty}\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}\circ\tau_{w,v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})}{\sum_{{w^{\prime}}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}\circ\tau_{w^{\prime},v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})}
=limmexp(ϕEmτw,v1ϕEm)(vxKc)wXKlimmexp(ϕEmτw,v1ϕEm)(vxKc)\displaystyle=\frac{\lim_{m\to\infty}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}\circ\tau_{w,v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})}{\sum_{{w^{\prime}}\in X_{K}}\lim_{m\to\infty}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{E_{m}}\circ\tau_{w^{\prime},v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})}
=exp(limm(ϕEmτw,v1ϕEm)(vxKc))wXKexp(limm(ϕEmτw,v1ϕEm)(vxKc))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\lim_{m\to\infty}\left(\phi_{E_{m}}\circ\tau_{w,v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})\right)}{\sum_{{w^{\prime}}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\lim_{m\to\infty}\left(\phi_{E_{m}}\circ\tau_{w^{\prime},v}^{-1}-\phi_{E_{m}}\right)(vx_{K^{c}})\right)}
=exp(ϕτw,v(vxKc))wXKexp(ϕτw,v(vxKc)),\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})\right)},

where the last equality follows from Theorem 4.2. Also, if m0m_{0}\in\mathbb{N} is such that KEm0K\subseteq E_{m_{0}}, the exchange of limit and sum in the denominator from the first to the second line follows from Lemma 4.8.

Corollary 4.10.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for every K(G)K\in\mathcal{F}(G), γK\gamma_{K} is GG-invariant, that is, for every wXKw\in X_{K}, xXx\in X, and gGg\in G, it holds that

γKg1(g[w],gx)=γK([w],x).\gamma_{Kg^{-1}}(g\cdot[w],g\cdot x)=\gamma_{K}([w],x).
Proof.

Let K(G)K\in\mathcal{F}(G). Given vXKv\in X_{K}, let yvXy^{v}\in X be arbitrary and such that yKv=vy^{v}_{K}=v. Then,

γKg1(g[w],gx)\displaystyle\gamma_{Kg^{-1}}(g\cdot[w],g\cdot x) =γKg1([(gyw)Kg1],gx)\displaystyle=\gamma_{Kg^{-1}}([(g\cdot y^{w})_{Kg^{-1}}],g\cdot x)
=exp(ϕτKg1((gyw)Kg1(gx)(Kg1)c))wXKexp(ϕτKg1((gyw)Kg1(gx)(Kg1)c))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{Kg^{-1}}}((g\cdot y^{w})_{Kg^{-1}}(g\cdot x)_{(Kg^{-1})^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{Kg^{-1}}}((g\cdot y^{w^{\prime}})_{Kg^{-1}}(g\cdot x)_{(Kg^{-1})^{c}})\right)}
=exp(ϕτKg1(g(yKwxKc)))wXKexp(ϕτKg1(g(yKwxKc)))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{Kg^{-1}}}(g\cdot(y^{w}_{K}x_{K^{c}}))\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{Kg^{-1}}}(g\cdot(y^{w^{\prime}}_{K}x_{K^{c}}))\right)}
=exp(ϕτK(yKwxKc))wXKexp(ϕτK(yKwxKc))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{K}}(y^{w}_{K}x_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{K}}(y^{w^{\prime}}_{K}x_{K^{c}})\right)}
=γK([w],x),\displaystyle=\gamma_{K}([w],x),

where we have used the property of ϕτ\phi_{*}^{\tau} from Corollary 4.3. ∎

Definition 5.

A potential h:Xh\colon X\to\mathbb{R} is local if hh is K\mathcal{B}_{K}-measurable for some K(G)K\in\mathcal{F}(G). For each K(G)K\in\mathcal{F}(G), denote by K\mathcal{L}_{K} the linear space of all bounded K\mathcal{B}_{K}-measurable potentials and =K(G)K\mathcal{L}=\bigcup_{K\in\mathcal{F}(G)}\mathcal{L}_{K}.

A potential h:Xh\colon X\to\mathbb{R} is quasilocal if there exists a sequence {ϕn}n\{\phi_{n}\}_{n} of local potentials such that limnhhn=0\lim_{n\to\infty}\|h-h_{n}\|_{\infty}=0. Note that ¯\overline{\mathcal{L}} is the linear space of all bounded quasilocal potentials, where ¯\overline{\mathcal{L}} is the uniform closure of \mathcal{L} on the linear space of bounded \mathcal{B}-measurable potentials.

Remark 3.

[30, Remark 2.21] A potential h:Xh\colon X\to\mathbb{R} is quasilocal if and only if for all exhausting sequences of finite subsets {Em}m\{E_{m}\}_{m} of GG, limmsupx,yXxEm=yEm|h(x)h(y)|=0\lim_{m\to\infty}\sup_{\begin{subarray}{c}x,y\in X\\ x_{E_{m}}=y_{E_{m}}\end{subarray}}|h(x)-h(y)|=0.

Definition 6.

A specification γ=(γK)K(G)\gamma=\left(\gamma_{K}\right)_{K\in\mathcal{F}(G)} is quasilocal if, for each K(G)K\in\mathcal{F}(G) and h¯h\in\overline{\mathcal{L}}, it holds that γKh¯\gamma_{K}h\in\overline{\mathcal{L}}, where

γKh(x)=wXKγK(w,x)h(wxKc).\gamma_{K}h(x)=\sum_{w\in X_{K}}\gamma_{K}(w,x)h(wx_{K^{c}}).
Remark 4.

In order to verify that a specification is quasilocal it suffices to prove that γKh¯\gamma_{K}h\in\overline{\mathcal{L}}, for K(G)K\in\mathcal{F}(G) and hh\in\mathcal{L} (see [30], page 32).

Theorem 4.11.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If γ={γK}K(G)\gamma=\{\gamma_{K}\}_{K\in\mathcal{F}(G)} is defined as in equation (7), then γ\gamma is quasilocal.

Proof.

Let hh\in\mathcal{L} and let ϵ>0\epsilon>0. Given any K(G)K\in\mathcal{F}(G), first notice that

|γKh(x)|wXKγK(w,x)|h(wxKc)|hwXKγK(w,x)=h,|\gamma_{K}h(x)|\leq\sum_{w\in X_{K}}\gamma_{K}(w,x)|h(wx_{K^{c}})|\leq\|h\|_{\infty}\sum_{w\in X_{K}}\gamma_{K}(w,x)=\|h\|_{\infty},

so γKhh\|\gamma_{K}h\|_{\infty}\leq\|h\|_{\infty}. In addition, if x,yXx,y\in X are such that xEn=yEnx_{E_{n}}=y_{E_{n}} for nn to be determined, we have that

|γK\displaystyle|\gamma_{K} h(x)γKh(y)|\displaystyle h(x)-\gamma_{K}h(y)|
wXK|γK(w,x)h(wxKc)γK(w,y)h(wyKc)|\displaystyle\leq\sum_{w\in X_{K}}|\gamma_{K}(w,x)h(wx_{K^{c}})-\gamma_{K}(w,y)h(wy_{K^{c}})|
=wXKγK(w,x)|h(wxKc)γK(w,y)γK(w,x)h(wyKc)|\displaystyle=\sum_{w\in X_{K}}\gamma_{K}(w,x)\left|h(wx_{K^{c}})-\frac{\gamma_{K}(w,y)}{\gamma_{K}(w,x)}h(wy_{K^{c}})\right|
wXKγK(w,x)|h(wxKc)e±2ϵh(wyKc)|\displaystyle\leq\sum_{w\in X_{K}}\gamma_{K}(w,x)\left|h(wx_{K^{c}})-e^{\pm 2\epsilon}h(wy_{K^{c}})\right|
wXKγK(w,x)|h(wxKc)h(wyKc)|+wXKγK(w,x)(1e±2ϵ)|h(wyKc)|\displaystyle\leq\sum_{w\in X_{K}}\gamma_{K}(w,x)\left|h(wx_{K^{c}})-h(wy_{K^{c}})\right|+\sum_{w\in X_{K}}\gamma_{K}(w,x)(1-e^{\pm 2\epsilon})\left|h(wy_{K^{c}})\right|
wXKγK(w,x)|h(wxKc)h(wyKc)|+wXKγK(w,x)(1e±2ϵ)h\displaystyle\leq\sum_{w\in X_{K}}\gamma_{K}(w,x)\left|h(wx_{K^{c}})-h(wy_{K^{c}})\right|+\sum_{w\in X_{K}}\gamma_{K}(w,x)(1-e^{\pm 2\epsilon})\left\|h\right\|_{\infty}
supx,y:xEn=yEn|h(x)h(y)|+(1e±2ϵ)h.\displaystyle\leq\sup_{x^{\prime},y^{\prime}:x^{\prime}_{E_{n}}=y^{\prime}_{E_{n}}}|h(x^{\prime})-h(y^{\prime})|+(1-e^{\pm 2\epsilon})\left\|h\right\|_{\infty}.

To justify the second inequality, first observe that, for every w,vXKw,v\in X_{K}, ϕτv,w\phi_{*}^{\tau_{v,w}} is uniformly continuous, since it is a uniform limit of uniformly continuous potentials, namely, ϕEm\phi_{E_{m}}. Then, there exists n0n_{0}\in\mathbb{N} such that for every nn0n\geq n_{0}, every w,vXKw,v\in X_{K}, and every x,yXx,y\in X with xEn=yEnx_{E_{n}}=y_{E_{n}},

|ϕτv,w(wxKc)ϕτv,w(wyKc)|<ϵ,|\phi_{*}^{\tau_{v,w}}(wx_{K^{c}})-\phi_{*}^{\tau_{v,w}}(wy_{K^{c}})|<\epsilon,

so

γK(w,y)=exp(ϕτw,w(wyKc))vXKexp(ϕτv,w(wyKc))exp(ϕτw,w(wxKc)+ϵ)vXKexp(ϕτv,w(wxKc)ϵ)=e2ϵγK(w,x).\gamma_{K}(w,y)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,w}}(wy_{K^{c}})\right)}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{v,w}}(wy_{K^{c}})\right)}\leq\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,w}}(wx_{K^{c}})+\epsilon\right)}{\sum_{v\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{v,w}}(wx_{K^{c}})-\epsilon\right)}=e^{2\epsilon}\gamma_{K}(w,x).

Now, since hh is local, we have that limnsupx,yXxEn=yEn|h(x)h(y)|=0\lim_{n\to\infty}\sup_{\begin{subarray}{c}x,y\in X\\ x_{E_{n}}=y_{E_{n}}\end{subarray}}|h(x)-h(y)|=0, so that there exists n1n_{1}\in\mathbb{N} such that for all nn1n\geq n_{1}, supx,yXxEn=yEn|h(x)h(y)|<ϵ\sup_{\begin{subarray}{c}x,y\in X\\ x_{E_{n}}=y_{E_{n}}\end{subarray}}|h(x)-h(y)|<\epsilon. Taking n=max{n0,n1}n=\max\{n_{0},n_{1}\}, we obtain that

|γKh(x)γKh(y)|ϵ+(1e±2ϵ)h,|\gamma_{K}h(x)-\gamma_{K}h(y)|\leq\epsilon+(1-e^{\pm 2\epsilon})\left\|h\right\|_{\infty},

and since ϵ\epsilon was arbitrary, we conclude. ∎

5. Equivalences of different notions of Gibbs measures

In this section, we introduce the four notions of Gibbs measures to be considered, namely, DLR, conformal, Bowen-Gibbs, and equilibrium measures, and prove the equivalence among them provided extra conditions. We mainly assume that GG is a countable amenable group, the configuration space is X=GX=\mathbb{N}^{G}, and ϕ:X\phi:X\to\mathbb{R} is an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}.

We proceed to describe the content of each subsection: in §5.1, we provide a rigorous definition of each kind of measure and results about entropy and pressure; in §5.2, we establish that the set of DLR measures and the set of conformal measures coincide; in §5.3, we prove that every DLR measure is a Bowen-Gibbs measure; in §5.4, we show the existence of a conformal measure; in §5.5, we prove that a GG-invariant Bowen-Gibbs measure with finite entropy is an equilibrium measure; finally, in §5.6, we prove that if a measure is an equilibrium measure, then it is also a DLR measure.

Below, we provide a diagram of the main results of this section, including extra assumptions needed.

DLR measureTheorem 5.6Theorem 5.8Conformal measureBowen-Gibbs measure+ GG-invariance and H(μ)<H(\mu)<\inftyTheorem 5.13Equilibrium measure Theorem 5.18G-invariance\scriptstyle{\begin{subarray}{c}\text{\hskip 5.69046pt Theorem \ref{thm:equilibrium-implies-DLR}}\\ \text{+ $G$-invariance}\end{subarray}}
Remark 5.

We are not aware whether it is possible to prove that a Bowen-Gibbs measure is necessarily a DLR measure without the finite entropy assumption. In fact, we do not know if GG-invariance is a necessary assumption for that implication.

5.1. Definitions of Gibbs measures

We start by giving the definitions of DLR, conformal, and Bowen-Gibbs measures.

Definition 7.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. A measure μ(X)\mu\in\mathcal{M}(X) is a DLR measure (for ϕ\phi) if

μ(B|Kc)(x)=γK(B,x)μ(x)-a.s.,\mu\left(B\,|\,\mathcal{B}_{K^{c}}\right)(x)=\gamma_{K}(B,x)\qquad\mu(x)\text{-a.s.,}

for every K(G)K\in\mathcal{F}(G), BB\in\mathcal{B}, and xXx\in X, where γK\gamma_{K} is defined as in equation (7). We denote the set of DLR measures for ϕ\phi by 𝒢(ϕ)\mathcal{G}(\phi).

Definition 8.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. A measure μ(X)\mu\in\mathcal{M}(X) is a conformal measure (for ϕ\phi) if

(9) d(μτ1)dμ=exp(ϕτ)μ(x)-a.s.\frac{d(\mu\circ\tau^{-1})}{d\mu}=\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau})\qquad\mu(x)\text{-a.s.}

for every A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G), and τK,A\tau\in\mathcal{E}_{K,A}.

Definition 9.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. A measure μ(X)\mu\in\mathcal{M}(X) is a Bowen-Gibbs measure (for ϕ\phi) if there exists pp\in\mathbb{R} such that, for every ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that, for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) and xXx\in X,

(10) exp(ϵ|F|)μ([xF])exp(ϕF(x)p|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\mu([x_{F}])}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)-p|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).
Remark 6.

Notice that, in Definition 9, we can replace ϕF(x)\phi_{F}(x) by supϕF([xF])\sup\phi_{F}([x_{F}]) in an equivalent way, so that we have

exp(ϵ|F|)μ([xF])exp(supϕF([xF])p|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\mu([x_{F}])}{\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([x_{F}])-p|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).
Proposition 5.1.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, if μ\mu is a Bowen-Gibbs measure for ϕ\phi, the constant pp is necessarily p(ϕ)p(\phi).

Proof.

Indeed, given ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 so that

exp(ϵ|F|)exp(ϕF(x))μ([xF])exp(p|F|)exp(ϵ|F|)exp(ϕF(x))\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)\right)\leq\mu([x_{F}])\mathop{\textrm{\rm exp}}\nolimits\left(p|F|\right)\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)\right)

for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) and every xXx\in X. Since xx is arbitrary, we have that

exp(ϵ|F|)exp(supϕF[xF])μ([xF])exp(p|F|)exp(ϵ|F|)exp(supϕF[xF]),\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[x_{F}]\right)\leq\mu([x_{F}])\mathop{\textrm{\rm exp}}\nolimits\left(p|F|\right)\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right)\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}[x_{F}]\right),

and, since μ\mu is a probability measure, adding over all xFXFx_{F}\in X_{F}, we get

exp(ϵ|F|)ZF(ϕ)exp(p|F|)exp(ϵ|F|)ZF(ϕ).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)Z_{F}(\phi)\leq\mathop{\textrm{\rm exp}}\nolimits\left(p|F|\right)\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right)Z_{F}(\phi).

Then, if we take logarithms and divide by |F||F|, we obtain that

ϵ+logZF(ϕ)|F|plogZF(ϕ)|F|+ϵ,-\epsilon+\frac{\log Z_{F}(\phi)}{|F|}\leq p\leq\frac{\log Z_{F}(\phi)}{|F|}+\epsilon,

so, taking the limit as FF becomes more and more invariant, we obtain that

ϵ+p(ϕ)pp(ϕ)+ϵ,-\epsilon+p(\phi)\leq p\leq p(\phi)+\epsilon,

and since ϵ\epsilon was arbitrary, we conclude that p=p(ϕ)p=p(\phi). ∎

Consider the canonical partition of XX given by {[a]}a\{[a]\}_{a\in\mathbb{N}}. This is a countable partition that generates the Borel σ\sigma-algebra \mathcal{B} under the shift dynamic. Given a measure ν(X)\nu\in\mathcal{M}(X), the Shannon entropy of the canonical partition associated with ν\nu is given by

H(ν):=aν([a])logν([a]).H(\nu):=-\sum_{a\in\mathbb{N}}\nu([a])\log\nu([a]).

Now, for each F(G)F\in\mathcal{F}(G), let {[w]}wXF\{[w]\}_{w\in X_{F}} be the FF-refinement of the canonical partition and consider its corresponding Shannon entropy, which is given by

HF(ν):=wXFν([w])logν([w]).H_{F}(\nu):=-\sum_{w\in X_{F}}\nu([w])\log\nu([w]).

We have the following proposition.

Proposition 5.2.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable and continuous potential with finite oscillation. If ν(X)\nu\in\mathcal{M}(X) is such that ϕ𝑑ν>\int\phi d\nu>-\infty, then H(ν)<H(\nu)<\infty. Furthermore, if ν\nu is GG-invariant, then, for every F(G)F\in\mathcal{F}(G), HF(ν)<H_{F}(\nu)<\infty.

Proof.

Let {An}n\{A_{n}\}_{n} an exhausting sequence of finite alphabets and F(G)F\in\mathcal{F}(G). Consider XF,n={xX:xFAnF}FX^{F,n}=\{x\in X:x_{F}\in A_{n}^{F}\}\in\mathcal{B}_{F}. Since ϕ\phi is exp-summable, then it is bounded from above. Without loss of generality, suppose that it is bounded from above by 0. Thus, so is ϕF\phi_{F}. Define

ϕF,n(x)={ϕF(x),xXF,n;0,otherwise.\phi^{F,n}(x)=\begin{cases}\phi_{F}(x),&x\in X^{F,n};\\ 0,&\text{otherwise}.\end{cases}

Notice that, for every xXx\in X, ϕF(x)=limnϕF,n(x)\phi_{F}(x)=\lim_{n\to\infty}\phi_{F,n}(x) and, for every nn\in\mathbb{N}, ϕF(x)ϕF,n+1(x)ϕF,n(x)\phi_{F}(x)\leq\phi^{F,n+1}(x)\leq\phi^{F,n}(x). Therefore, by the Monotone Convergence Theorem, we can conclude that

ϕF𝑑ν=limnϕF,n(x)𝑑ν.\int\phi_{F}d\nu=\lim_{n\to\infty}\int\phi^{F,n}(x)d\nu.

For each nn\in\mathbb{N}, let HF,n(ν)=wAnFν([w])logν([w])H_{F,n}(\nu)=-\sum_{w\in A_{n}^{F}}\nu([w])\log\nu([w]). Then, limnHF,n(ν)=HF(ν)\lim_{n\to\infty}H_{F,n}(\nu)=H_{F}(\nu). Also, for each nn\in\mathbb{N} and F(G)F\in\mathcal{F}(G), notice that ϕF,nwAnF1[w]supϕF([w])\phi^{F,n}\leq\sum_{w\in A_{n}^{F}}1_{[w]}\sup\phi_{F}([w]). Therefore, for every nn\in\mathbb{N} and F(G)F\in\mathcal{F}(G),

HF,n(ν)+ϕF,n𝑑ν\displaystyle H_{F,n}(\nu)+\int\phi^{F,n}d\nu =wAnFν([w])logν([w])+ϕF,n𝑑ν\displaystyle=-\sum_{w\in A^{F}_{n}}\nu([w])\log\nu([w])+\int\phi^{F,n}d\nu
wAnFν([w])logν([w])+wAnFν([w])supϕF([w])\displaystyle\leq-\sum_{w\in A_{n}^{F}}\nu([w])\log\nu([w])+\sum_{w\in A_{n}^{F}}\nu([w])\sup\phi_{F}([w])
=wAnFν([w])log(exp(supϕF([w]))ν([w]))\displaystyle=\sum_{w\in A_{n}^{F}}\nu([w])\log\left(\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([w])\right)}{\nu([w])}\right)
log(wAnFexpsupϕF([w]))\displaystyle\leq\log\left(\sum_{w\in A_{n}^{F}}\mathop{\textrm{\rm exp}}\nolimits\sup\phi_{F}([w])\right)
=logZF(An,ϕ),\displaystyle=\log Z_{F}(A_{n},\phi),

where we assume that all the sums involved are over cylinder sets with positive measure. The second inequality follows from Jensen’s inequality. In addition, notice that, in the case that ν\nu is GG-invariant, it follows that

HF(ν)\displaystyle H_{F}(\nu) =limnHF,n(ν)\displaystyle=\lim_{n\to\infty}H_{F,n}(\nu)
limn(logZF(An,ϕ)ϕF,n𝑑ν)\displaystyle\leq\lim_{n\to\infty}\left(\log Z_{F}(A_{n},\phi)-\int\phi^{F,n}d\nu\right)
=logZF(ϕ)ϕF𝑑ν\displaystyle=\log Z_{F}(\phi)-\int\phi_{F}d\nu
|F|(logZ1G(ϕ)ϕ𝑑ν),\displaystyle\leq|F|\left(\log Z_{1_{G}}(\phi)-\int\phi d\nu\right),

where we have used that logZF(ϕ)|F|logZ1G(ϕ)\log Z_{F}(\phi)\leq|F|\log Z_{1_{G}}(\phi) and ϕF𝑑ν=|F|ϕ𝑑ν\int\phi_{F}d\nu=|F|\int\phi d\nu. Therefore, HF(ν)<H_{F}(\nu)<\infty and, in particular, H(ν)=H{1G}(ν)<H(\nu)=H_{\{1_{G}\}}(\nu)<\infty. ∎

Through a standard argument (for example, for the case G=G=\mathbb{Z}, see [24]; the general case is analogous), it can be justified that if the canonical partition has finite Shannon entropy, the Kolmogorov-Sinai entropy of ν\nu can be written as

h(ν)=limFG1|F|HF(ν).h(\nu)=\lim_{F\to G}\frac{1}{|F|}H_{F}(\nu).

The next proposition is based on [45, Lemma 4.9] and gives us an upper bound in terms of the pressure for the specific Gibbs free energy of a given measure with respect to some potential. Sometimes this fact is known as the Gibbs inequality.

Proposition 5.3.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable and uniformly continuous potential with finite oscillation. If μ(X)\mu\in\mathcal{M}(X) is GG-invariant and ϕ𝑑μ>\int\phi d\mu>-\infty, then h(μ)+ϕ𝑑μp(ϕ)h(\mu)+\int\phi d\mu\leq p(\phi).

Proof.

Since ϕ:X\phi\colon X\to\mathbb{R} is an exp-summable and uniformly continuous potential with finite oscillation, due to Theorem 3.8, the pressure p(ϕ)p(\phi) exists. Then,

h(μ)+ϕ𝑑μ=limFG1|F|HF(μ)+ϕ𝑑μlimFG1|F|logZF(ϕ)=p(ϕ).\displaystyle h(\mu)+\int\phi d\mu=\lim_{F\to G}\frac{1}{|F|}H_{F}(\mu)+\int\phi d\mu\leq\lim_{F\to G}\frac{1}{|F|}\log Z_{F}(\phi)=p(\phi).

Definition 10.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. A measure μG(X)\mu\in\mathcal{M}_{G}(X) is an equilibrium measure (for ϕ\phi) if ϕ𝑑μ>\int{\phi}d\mu>-\infty and

(11) h(μ)+ϕ𝑑μ=sup{h(ν)+ϕ𝑑ν:νG(X),ϕ𝑑ν>}.h(\mu)+\int{\phi}d\mu=\sup\left\{h(\nu)+\int{\phi}d\nu\colon\nu\in\mathcal{M}_{G}(X),\int{\phi}d\nu>-\infty\right\}.

Notice that it is not clear whether the supremum in equation (11) is achieved. The answer to this problem is intimately related to the concept of Gibbs measures in its various forms and their equivalences, which we address throughout this section.

Remark 7.

Notice that, in light of Proposition 5.2, any measure νG(X)\nu\in\mathcal{M}_{G}(X) such that the ϕ𝑑ν>\int\phi d\nu>-\infty has finite entropy, that is, h(ν)<h(\nu)<\infty, provided that ϕ\phi is exp-summable and has finite oscillation. Thus, in the particular case that ϕ\phi is an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, we obtain that h(ν)<h(\nu)<\infty.

5.2. Equivalence between DLR and conformal measures

This section is dedicated to proving that the notions of DLR measure and conformal measure coincide in the full shift with countable alphabet over a countable amenable group context. Nevertheless, before proving this major result, notice that for BB\in\mathcal{B}, K(G)K\in\mathcal{F}(G), and xXx\in X,

(12) μ(B|Kc)(x)=wXKμ([w]|Kc)(x)𝟙{wxKcB}.\mu(B\,|\,\mathcal{B}_{K^{c}})(x)=\sum_{w\in X_{K}}\mu([w]\,|\,\mathcal{B}_{K^{c}})(x)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}.

Indeed, it can be checked that 𝟙{wxKcB}(x)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}(x) is Kc\mathcal{B}_{K^{c}}-measurable, so μ(x)\mu(x)-a.s.,

wXKμ([w]|Kc)(x)𝟙{wxKcB}(x)\displaystyle\sum_{w\in X_{K}}\mu([w]\,|\,\mathcal{B}_{K^{c}})(x)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}(x) =μ(wXK𝟙[w]𝟙{wxKcB}|Kc)(x)\displaystyle=\mu\left(\sum_{w\in X_{K}}\mathbbm{1}_{[w]}\mathbbm{1}_{\{wx_{K^{c}}\in B\}}\middle|\mathcal{B}_{K^{c}}\right)(x)
=μ(wXK𝟙[w]𝟙B|Kc)(x)\displaystyle=\mu\left(\sum_{w\in X_{K}}\mathbbm{1}_{[w]}\mathbbm{1}_{B}\middle|\mathcal{B}_{K^{c}}\right)(x)
=μ(B|Kc)(x).\displaystyle=\mu\left(B|\mathcal{B}_{K^{c}}\right)(x).

This observation will allow us to reduce our calculations from arbitrary Borel sets BB\in\mathcal{B} to cylinder sets of the form [w][w]. Next, we have the following result.

Corollary 5.4.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. A measure μ(X)\mu\in\mathcal{M}(X) is a DLR measure for ϕ\phi if, and only if, for every K(G)K\in\mathcal{F}(G), wXKw\in X_{K} and xXx\in X, it holds that

(13) μ([w]|Kc)(x)=exp(ϕτw,v(vxKc))wXKexp(ϕτw,v(vxKc))μ(x)-a.s.,\mu([{w}]\,|\,\mathcal{B}_{K^{c}})(x)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})\right)}\qquad\mu(x)\text{-a.s.},

for every vXKv\in X_{K}.

Proof.

If μ\mu is a DLR measure for ϕ\phi, then for every K(G)K\in\mathcal{F}(G), BB\in\mathcal{B}, and xXx\in X,

μ(B|Kc)(x)=γK(B,x)μ(x)-a.s.\mu\left(B\,|\,\mathcal{B}_{K^{c}}\right)(x)=\gamma_{K}(B,x)\quad\mu(x)\text{-a.s.}

Thus, in particular, if wXKw\in X_{K}, it holds that

μ([w]|Kc)(x)=γK([w],x)μ(x)-a.s.,\mu([{w}]\,|\,\mathcal{B}_{K^{c}})(x)=\gamma_{K}([w],x)\qquad\mu(x)\text{-a.s.},

and the result follows from Proposition 4.9.

On the other hand, if we assume that for every K(G)K\in\mathcal{F}(G), wXKw\in X_{K}, and xXx\in X, equation (13) holds μ(x)\mu(x)-almost surely for every vXKv\in X_{K}, then, from equation (12) and Proposition 4.9, μ(x)\mu(x)-a.s it holds that

μ(B|Kc)(x)\displaystyle\mu(B|\mathcal{B}_{K^{c}})(x) =wXKμ([w]|Kc)(x)𝟙{wxKcB}=wXKγK([w],x)𝟙{wxKcB}=γK(B,x).\displaystyle=\sum_{w\in X_{K}}\mu([w]|\mathcal{B}_{K^{c}})(x)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}=\sum_{w\in X_{K}}\gamma_{K}([w],x)\mathbbm{1}_{\{wx_{K^{c}}\in B\}}=\gamma_{K}(B,x).

In order to relate the functions γK\gamma_{K} that appear in the definition of DLR measures with the permutations involved in the definition of conformal measures, we have the following lemma.

Lemma 5.5.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for every K(G)K\in\mathcal{F}(G), v,wXKv,w\in X_{K} and τK\tau\in\mathcal{E}_{K} such that τ1([v])=[w]\tau^{-1}([v])=[w],

γK([w],x)=exp(ϕτ(vxKc))γK([v],x).\gamma_{K}([w],x)=\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau}(vx_{K^{c}})\right)\gamma_{K}([v],x).
Proof.

Indeed, by Proposition 4.9, for every xXx\in X,

γK([w],x)\displaystyle\gamma_{K}([w],x) =exp(ϕτ(vxKc))wXKexp(ϕτw,v(vxKc))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})\right)}
=exp(ϕτv,v(vxKc))wXKexp(ϕτw,v(vxKc))exp(ϕτ(vxKc))exp(ϕτv,v(vxKc))\displaystyle=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{v,v}}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})\right)}\cdot\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau}(vx_{K^{c}})\right)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{v,v}}(vx_{K^{c}})\right)}
=γK([v],x)exp(ϕτ(vxKc)ϕτv,v(vxKc)).\displaystyle=\gamma_{K}([v],x)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau}(vx_{K^{c}})-\phi_{*}^{\tau_{v,v}}(vx_{K^{c}})\right).

Now, notice that ϕτ(vxKc)ϕτv,v(vxKc)=ϕτ(vxKc)\phi_{*}^{\tau}(vx_{K^{c}})-\phi_{*}^{\tau_{v,v}}(vx_{K^{c}})=\phi_{*}^{\tau}(vx_{K^{c}}), and the result follows. ∎

Now we can prove the main result of this subsection. The proof is a slight adaptation of the proof of [46, Theorem 3.3] and we include it here for completeness.

Theorem 5.6.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, a measure μ(X)\mu\in\mathcal{M}(X) is a DLR measure for ϕ\phi if and only if μ\mu is a conformal measure for ϕ\phi.

Proof.

Suppose first that μ(X)\mu\in\mathcal{M}(X) is a conformal measure for ϕ\phi and let K(G)K\in\mathcal{F}(G). Begin by noticing that if BKcB\in\mathcal{B}_{K^{c}} and, for some wXKw\in X_{K} and xXx\in X, wxKcBwx_{K^{c}}\in B, then vxKcBvx_{K^{c}}\in B, for every vXKv\in X_{K}. As a consequence, we have that, for all τK\tau\in\mathcal{E}_{K} and all BKcB\in\mathcal{B}_{K^{c}}, B=τ1(B)B=\tau^{-1}(B).

For w,vXKw,v\in X_{K}, consider τw,vK\tau_{w,v}\in\mathcal{E}_{K}. Thus, τw,v1([v])=[w]\tau_{w,v}^{-1}([v])=[w] and, for every BKcB\in\mathcal{B}_{K^{c}}, τw,v1([v]B)=τw,v1([v])τw,v1(B)=[w]B\tau_{w,v}^{-1}([v]\cap B)=\tau_{w,v}^{-1}([v])\cap\tau_{w,v}^{-1}(B)=[w]\cap B. Furthermore,

B𝟙[w](x)𝑑μ(x)\displaystyle\int_{B}\mathbbm{1}_{[w]}(x)\,d\mu(x) =B𝟙[v](x)d(μτw,v1)(x)\displaystyle=\int_{B}\mathbbm{1}_{[v]}(x)\,d(\mu\circ\tau_{w,v}^{-1})(x)
=B𝟙[v](x)expϕτw,v(x)dμ(x)\displaystyle=\int_{B}\mathbbm{1}_{[v]}(x)\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(x)\,d\mu(x)
=B𝟙[v](x)expϕτw,v(vxKc)dμ(x)\displaystyle=\int_{B}\mathbbm{1}_{[v]}(x)\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\,d\mu(x)
=Bμ(𝟙[v](x)expϕτw,v(vxKc)|Kc)(x)dμ(x)\displaystyle=\int_{B}\mu\left(\mathbbm{1}_{[v]}(x)\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\,\middle|\,\mathcal{B}_{K^{c}}\right)(x)\,d\mu(x)
=Bμ(𝟙[v]|Kc)(x)expϕτw,v(vxKc)dμ(x).\displaystyle=\int_{B}\mu\left(\mathbbm{1}_{[v]}\,\middle|\,\mathcal{B}_{K^{c}}\right)(x)\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\,d\mu(x).

On the other hand,

B𝟙[w](x)dμ(x)=Bμ(𝟙[w]|Kc)(x)dμ(x).\int_{B}\mathbbm{1}_{[w]}(x)\,d\mu(x)=\int_{B}\mu\left(\mathbbm{1}_{[w]}\,\middle|\,\mathcal{B}_{K^{c}}\right)(x)\,d\mu(x).

Therefore, for any w,vXKw,v\in X_{K}, μ(x)\mu(x)-almost surely it holds that

(14) μ(𝟙[v]|Kc)(x)expϕτw,v(vxKc)=μ(𝟙[w]|Kc)(x).\mu\left(\mathbbm{1}_{[v]}\,\middle|\,\mathcal{B}_{K^{c}}\right)(x)\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})=\mu\left(\mathbbm{1}_{[w]}\,|\,\mathcal{B}_{K^{c}}\right)(x).

Now, let A()A\in\mathcal{F}(\mathbb{N}) be a finite alphabet and vAKv\in A^{K}. For any wAKw^{\prime}\in A^{K}, we have that τw,vK,A\tau_{w^{\prime},v}\in\mathcal{E}_{K,A}. Summing equation (14) over all wAKw^{\prime}\in A^{K}, we obtain that μ(x)\mu(x)-almost surely it holds that

(15) μ(𝟙AK×XKc|Kc)(x)\displaystyle\mu\left(\mathbbm{1}_{A^{K}\times X_{K^{c}}}\middle|\mathcal{B}_{K^{c}}\right)(x) =wAKμ(𝟙[w]|Kc)(x)\displaystyle=\sum_{w^{\prime}\in A^{K}}\mu\left(\mathbbm{1}_{[w^{\prime}]}\middle|\mathcal{B}_{K^{c}}\right)(x)
(16) =μ(𝟙[v]|Kc)(x)wAKexpϕτw,v(vxKc).\displaystyle=\mu\left(\mathbbm{1}_{[v]}\middle|\mathcal{B}_{K^{c}}\right)(x)\sum_{w^{\prime}\in A^{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}}).

If {An}n\{A_{n}\}_{n} is an exhausting sequence of finite alphabets, then n1(AnK×XKc)c=\bigcap_{n\geq 1}\left(A_{n}^{K}\times X_{K^{c}}\right)^{c}=\emptyset. Moreover, for each nn\in\mathbb{N},

(1𝟙AnK×XKc)2𝑑μ\displaystyle\int\left(1-\mathbbm{1}_{A_{n}^{K}\times X_{K^{c}}}\right)^{2}d\mu =|1𝟙AnK×XKc|𝑑μ=𝟙(AnK×XKc)c𝑑μ.\displaystyle=\int\left|1-\mathbbm{1}_{A_{n}^{K}\times X_{K^{c}}}\right|d\mu=\int\mathbbm{1}_{\left(A_{n}^{K}\times X_{K^{c}}\right)^{c}}\,\,d\mu.

Therefore, (1𝟙AnK×XKc)2𝑑μ0 as n\int\left(1-\mathbbm{1}_{A_{n}^{K}\times X_{K^{c}}}\right)^{2}\,d\mu\longrightarrow 0\text{ as }n\to\infty. Since conditional expectation given Kc\mathcal{B}_{K^{c}} is a continuous linear operator on L2(μ)L^{2}(\mu), we have μ(𝟙AnK×XKc|Kc)μ(1|Kc)\mu\left(\mathbbm{1}_{A_{n}^{K}\times X_{K^{c}}}\middle|\mathcal{B}_{K^{c}}\right)\longrightarrow\mu\left(1\middle|\mathcal{B}_{K^{c}}\right), μ(x)\mu(x)-almost surely in L2(μ)L^{2}(\mu) as nn\to\infty. Therefore, for any fixed vXKv\in X_{K}, there exists n0n_{0}\in\mathbb{N} be such that vAn0Kv\in A^{K}_{n_{0}} and, consequently, vAnKv\in A_{n}^{K}, for all nn0n\geq n_{0}. Therefore, μ(x)\mu(x)-almost surely it holds that

1\displaystyle 1 =μ(1|Kc)(x)\displaystyle=\mu\left(1\,|\,\mathcal{B}_{K^{c}}\right)(x)
=limnμ(𝟙AnK×XKc|Kc)(x)\displaystyle=\lim_{n\to\infty}\mu\left(\mathbbm{1}_{A_{n}^{K}\times X_{K^{c}}}\,|\,\mathcal{B}_{K^{c}}\right)(x)
=limnμ(𝟙[v]|Kc)(x)wAnKexpϕτw,v(vxKc)\displaystyle=\lim_{n\to\infty}\mu\left(\mathbbm{1}_{[v]}\,|\,\mathcal{B}_{K^{c}}\right)(x)\sum_{w\in A_{n}^{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(v\,x_{K^{c}})
=μ(𝟙[v]|Kc)(x)limnwAnKexpϕτw,v(vxKc)\displaystyle=\mu\left(\mathbbm{1}_{[v]}\middle|\mathcal{B}_{K^{c}}\right)(x)\lim_{n\to\infty}\sum_{w^{\prime}\in A_{n}^{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})
=μ(𝟙[v]|Kc)(x)wXKexpϕτw,v(vxKc).\displaystyle=\mu\left(\mathbbm{1}_{[v]}\middle|\mathcal{B}_{K^{c}}\right)(x)\sum_{w\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}}).

Moreover, equation (14) yields that, for any wXKw\in X_{K}, μ(x)\mu(x)-almost surely

1=μ(𝟙[v]|Kc)(x)wXKexpϕτw,v(vxKc)=μ(𝟙[w]|Kc)(x)expϕτw,v(vxKc)wXKexpϕτw,v(vxKc),\displaystyle 1=\mu\left(\mathbbm{1}_{[v]}\middle|\mathcal{B}_{K^{c}}\right)(x)\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})=\frac{\mu\left(\mathbbm{1}_{[w]}\middle|\mathcal{B}_{K^{c}}\right)(x)}{\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})}\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}}),

so that, for any wXKw\in X_{K}, μ(x)\mu(x)-almost surely it holds that

μ([w]|Kc)(x)=exp(ϕτw,v(vxKc))wXKexpϕτw,v(vxKc).\displaystyle\mu\left([w]\middle|\mathcal{B}_{K^{c}}\right)(x)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{w,v}}(vx_{K^{c}})\right)}{\sum_{w^{\prime}\in X_{K}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{K^{c}})}.

Therefore, due to Corollary 5.4, μ\mu is a DLR measure.

Conversely, suppose that μ(X)\mu\in\mathcal{M}(X) is a DLR measure for ϕ\phi and let A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G), and τK,A\tau\in\mathcal{E}_{K,A}. For any vXKv\in X_{K} and w=[τ1([v])]Kw=\left[\tau^{-1}([v])\right]_{K}, due to Lemma 5.5, we obtain

μτ1([v])=μ([w])\displaystyle\mu\circ\tau^{-1}([v])=\mu([w]) =μ([w]|Kc)(x)dμ(x)\displaystyle=\int\mu\left([w]\middle|\mathcal{B}_{K^{c}}\right)(x)\,d\mu(x)
=expϕτ(vxKc)μ([v]|Kc)(x)dμ(x)\displaystyle=\int\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(vx_{K^{c}})\mu\left([v]\middle|\mathcal{B}_{K^{c}}\right)(x)\,d\mu(x)
=μ(expϕτ(vxKc)𝟙[v]|Kc)(x)dμ(x)\displaystyle=\int\mu\left(\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(vx_{K^{c}})\mathbbm{1}_{[v]}\middle|\mathcal{B}_{K^{c}}\right)(x)\,d\mu(x)
=expϕτ(vxKc)𝟙[v](x)dμ(x)\displaystyle=\int\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(vx_{K^{c}})\mathbbm{1}_{[v]}(x)\,d\mu(x)
=[v]expϕτ(vxKc)dμ(x),\displaystyle=\int_{[v]}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(vx_{K^{c}})\,d\mu(x),

which concludes the result. ∎

5.3. DLR measures are Bowen-Gibbs measures

This subsection is dedicated to proving that, provided some conditions, any DLR measure for a potential ϕ\phi is a Bowen-Gibbs measure for ϕ\phi.

Proposition 5.7.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If μ(X)\mu\in\mathcal{M}(X) is a DLR measure for ϕ\phi, then, for every F(G)F\in\mathcal{F}(G), wXFw\in X_{F} and yXy\in X, it holds μ(x)\mu(x)-almost surely that

exp(2VF(ϕ)3ΔF(ϕ))μ([w]|Fc)(x)exp(ϕF(wyFc)logZF(ϕ))exp(2VF(ϕ)+3ΔF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits\left(-2V_{F}(\phi)-3\Delta_{F}(\phi)\right)\leq\frac{\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(2V_{F}(\phi)+3\Delta_{F}(\phi)\right).
Proof.

Let F(G)F\in\mathcal{F}(G) and τF\tau\in\mathcal{E}_{F}. From Proposition 4.4, we have that for every xXx\in X,

(17) |ϕτ(x)ϕFτ(x)|VF(ϕ),|\phi_{*}^{\tau}(x)-\phi^{\tau}_{F}(x)|\leq V_{F}(\phi),

which, in particular, yields that, for every xXx\in X,

(18) 0<exp(VF(ϕ))expϕτ(x)expϕFτ(x)exp(VF(ϕ))expϕτ(x).0<\mathop{\textrm{\rm exp}}\nolimits(-V_{F}(\phi))\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(x)\leq\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau}_{F}(x)\leq\mathop{\textrm{\rm exp}}\nolimits(V_{F}(\phi))\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}(x).

For a fixed vXFv\in X_{F} and for every wXFw^{\prime}\in X_{F}, the map τw,v\tau_{w^{\prime},v} belongs to F\mathcal{E}_{F}. Thus, inequality (18) holds for any such τw,v\tau_{w^{\prime},v} and, summing over all those such maps, we obtain that, for every xXx\in X,

exp(VF(ϕ))wXFexpϕτw,v(x)wXFexpϕFτw,v(x)exp(VF(ϕ))wXFexpϕτw,v(x).\mathop{\textrm{\rm exp}}\nolimits(-V_{F}(\phi))\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(x)\leq\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(x)\leq\mathop{\textrm{\rm exp}}\nolimits(V_{F}(\phi))\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(x).

Therefore, for every F(G)F\in\mathcal{F}(G), vXFv\in X_{F} and xXx\in X, we have

(19) exp(VF(ϕ))wXFexpϕFτw,v(x)wXFexpϕτw,v(x)exp(VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-V_{F}(\phi))\leq\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(x)}{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(x)}\leq\mathop{\textrm{\rm exp}}\nolimits(V_{F}(\phi)).

On the other hand, inequality (17) also yields that for every F(G)F\in\mathcal{F}(G), w,vXFw,v\in X_{F} and xXx\in X,

(20) exp(VF(ϕ))expϕτw,v(x)expϕFτw,v(x)exp(VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-V_{F}(\phi))\leq\frac{\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(x)}{\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w,v}}_{F}(x)}\leq\mathop{\textrm{\rm exp}}\nolimits(V_{F}(\phi)).

Then, from inequalities (19) and (20), we obtain that, for every F(G)F\in\mathcal{F}(G), w,vXFw,v\in X_{F} and xXx\in X,

(21) exp(2VF(ϕ))wXFexpϕFτw,v(x)wXFexpϕτw,v(x)expϕτw,v(x)expϕFτw,v(x)exp(2VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-2V_{F}(\phi))\leq\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(x)}{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(x)}\cdot\frac{\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(x)}{\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w,v}}_{F}(x)}\leq\mathop{\textrm{\rm exp}}\nolimits(2V_{F}(\phi)).

So, if x[v]x\in[v], inequality (21) can be rewritten as

(22) exp(2VF(ϕ))wXFexpϕFτw,v(vxFc)wXFexpϕτw,v(vxFc)expϕτw,v(vxFc)expϕFτw,v(vxFc)exp(2VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-2V_{F}(\phi))\leq\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(vx_{F^{c}})}{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w^{\prime},v}}(vx_{F^{c}})}\cdot\frac{\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau_{w,v}}(vx_{F^{c}})}{\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w,v}}_{F}(vx_{F^{c}})}\leq\mathop{\textrm{\rm exp}}\nolimits(2V_{F}(\phi)).

Since μ\mu is a DLR measure for ϕ\phi, from Corollary 5.4 we obtain that μ(x)\mu(x)-almost surely it holds that

(23) exp(2VF(ϕ))μ([w]|Fc)(x)wXFexpϕFτw,v(vxFc)expϕFτw,v(vxFc)exp(2VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-2V_{F}(\phi))\leq\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(vx_{F^{c}})}{\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w,v}}_{F}(vx_{F^{c}})}\leq\mathop{\textrm{\rm exp}}\nolimits(2V_{F}(\phi)).

Furthermore, notice that

wXFexpϕFτw,v(vxFc)expϕFτw,v(vxFc)\displaystyle\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w^{\prime},v}}_{F}(vx_{F^{c}})}{\mathop{\textrm{\rm exp}}\nolimits\phi^{\tau_{w,v}}_{F}(vx_{F^{c}})} =wXFexpϕF(wxFc)expϕF(wxFc),\displaystyle=\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(w^{\prime}x_{F^{c}})}{\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(wx_{F^{c}})},

so that inequality (23) can be rewritten as

(24) exp(2VF(ϕ))μ([w]|Fc)(x)wXFexpϕF(wxFc)expϕF(wxFc)exp(2VF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits(-2V_{F}(\phi))\leq\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)\frac{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(w^{\prime}x_{F^{c}})}{\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(wx_{F^{c}})}\leq\mathop{\textrm{\rm exp}}\nolimits(2V_{F}(\phi)).

For F(G)F\in\mathcal{F}(G) and xXx\in X, define the following auxiliary probability measure over XFX_{F}:

πFx(w):=expϕF(wxFc)wXFexpϕF(wxFc), for wXF.\pi_{F}^{x}(w):=\frac{\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(wx_{F^{c}})}{\sum_{w^{\prime}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\phi_{F}(w^{\prime}x_{F^{c}})},\quad\text{ for }w\in X_{F}.

Thus, inequality (24) yields that μ(x)\mu(x)-almost surely it holds that

exp(2VF(ϕ))πFx(w)μ([w]|Fc)(x)exp(2VF(ϕ))πFx(w).\mathop{\textrm{\rm exp}}\nolimits(-2V_{F}(\phi))\pi_{F}^{x}(w)\leq\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)\leq\mathop{\textrm{\rm exp}}\nolimits(2V_{F}(\phi))\pi_{F}^{x}(w).

Now, given yXy\in X, notice that the tail configuration xFcx_{F^{c}} can be replaced by yFcy_{F^{c}} with a penalty of 2ΔF(ϕ)2\Delta_{F}(\phi) as follows

πFy(w)exp(2ΔF(ϕ))πFx(w)πFy(w)exp(2ΔF(ϕ)),\pi_{F}^{y}(w)\mathop{\textrm{\rm exp}}\nolimits(-2\Delta_{F}(\phi))\leq\pi_{F}^{x}(w)\leq\pi_{F}^{y}(w)\mathop{\textrm{\rm exp}}\nolimits(2\Delta_{F}(\phi)),

so that

(25) exp(2(VF(ϕ)+ΔF(ϕ)))μ([w]|Fc)(x)πFy(w)exp(2(VF(ϕ)+ΔF(ϕ))).\mathop{\textrm{\rm exp}}\nolimits\left(-2(V_{F}(\phi)+\Delta_{F}(\phi))\right)\leq\frac{\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)}{\pi^{y}_{F}(w)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(2(V_{F}(\phi)+\Delta_{F}(\phi))\right).

Moreover, it is easy to verify that

exp(ΔF(ϕ))πFy(w)exp(ϕF(wyFc)logZF(ϕ))exp(ΔF(ϕ)).\mathop{\textrm{\rm exp}}\nolimits\left(-\Delta_{F}(\phi)\right)\leq\frac{\pi_{F}^{y}(w)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\Delta_{F}(\phi)\right).

Therefore, for every wXFw\in X_{F}, yXy\in X, it holds μ(x)\mu(x)-almost surely that

μ([w]|Fc)(x)\displaystyle\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x) exp(2(VF(ϕ)+ΔF(ϕ)))exp(ϕF(wyFc)logZF(ϕ)ΔF(ϕ))\displaystyle\geq\mathop{\textrm{\rm exp}}\nolimits\left(-2\left(V_{F}(\phi)+\Delta_{F}(\phi)\right)\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)-\Delta_{F}(\phi)\right)
=exp(2VF(ϕ)3ΔF(ϕ))exp(ϕF(wyFc)logZF(ϕ))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-2V_{F}(\phi)-3\Delta_{F}(\phi)\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)\right)

and that

μ([w]|Fc)(x)\displaystyle\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x) exp(2(VF(ϕ)+ΔF(ϕ)))exp(ϕF(wyFc)logZF(ϕ)+ΔF(ϕ))\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(2\left(V_{F}(\phi)+\Delta_{F}(\phi)\right)\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)+\Delta_{F}(\phi)\right)
=exp(2VF(ϕ)+3ΔF(ϕ))exp(ϕF(wyFc)logZF(ϕ)).\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(2V_{F}(\phi)+3\Delta_{F}(\phi)\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)\right).

Thus,

exp(2VF(ϕ)3ΔF(ϕ))μ([w]|Fc)(x)exp(ϕF(wyFc)logZF(ϕ))exp(2VF(ϕ)+3ΔF(ϕ)),\mathop{\textrm{\rm exp}}\nolimits\left(-2V_{F}(\phi)-3\Delta_{F}(\phi)\right)\leq\frac{\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(wy_{F^{c}})-\log Z_{F}(\phi)\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(2V_{F}(\phi)+3\Delta_{F}(\phi)\right),

concluding the proof. ∎

We now state the main theorem of this subsection.

Theorem 5.8.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If μ\mu is a DLR measure for ϕ\phi, then, for every ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that for every (K,δ)(K,\delta)-invariant set FF and xXx\in X, it holds μ(x)\mu(x)-almost surely that

exp(ϵ|F|)μ([w]|Fc)(x)exp(ϕF(x)p(ϕ)|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)-p(\phi)\cdot|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).

In particular, μ\mu is a Bowen-Gibbs measure for ϕ\phi.

Proof.

Indeed, for every ϵ>0\epsilon>0, we obtain, from Proposition 2.2, Lemma 2.3, and Theorem 3.8, that there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that, for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G),

ΔF(ϕ)ϵ|F|,VF(ϕ)ϵ|F|, and |logZF(ϕ)p(ϕ)|F||ϵ|F|,\Delta_{F}(\phi)\leq\epsilon\cdot|F|,\leavevmode\nobreak\ V_{F}(\phi)\leq\epsilon\cdot|F|,\text{ and }\left|\log Z_{F}(\phi)-p(\phi)|F|\right|\leq\epsilon\cdot|F|,

respectively. Considering a sufficiently large KK and sufficiently small δ\delta so that the three conditions are satisfied at the same time, we obtain from Proposition 5.7 that

exp(ϵ|F|)μ([w]|Fc)(x)exp(ϕF(x)p(ϕ)|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\mu\left([w]\,|\,\mathcal{B}_{F^{c}}\right)(x)}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)-p(\phi)\cdot|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).

Integrating this inequality with respect to dμ(x)d\mu(x), it follows that μ\mu is a Bowen-Gibbs measure for ϕ\phi. ∎

5.4. Existence of conformal measures

In order to guarantee that the equivalences we prove here are non-trivial, we prove the existence of a conformal measure for an exp-summable potential with summable variation in the context of a countably infinite state space over an amenable group. The strategy is to apply a version of Prokhorov’s Theorem.

Definition 11.

A sequence of probability measures {μn}n\{\mu_{n}\}_{n} in (X)\mathcal{M}(X) is tight if for every ϵ>0\epsilon>0 there exists a compact set KϵXK_{\epsilon}\subseteq X such that

μn(Kϵ)>1ϵ for all n.\mu_{n}(K_{\epsilon})>1-\epsilon\qquad\text{ for all $n\in\mathbb{N}$.}

We now state a version of Prokhorov’s Theorem (see [8, 47]).

Theorem 5.9.

Every tight sequence of probability measures in (X)\mathcal{M}(X) has a weak convergent subsequence.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Consider AA\subseteq\mathbb{N} a finite alphabet. Then ϕ|AG\phi|_{A^{G}} is also an exp-summable potential with summable variation according to {Em}m\{E_{m}\}_{m} and the specification defined by equation (7) is quasilocal. Moreover, the set of Borel probability measures on AGA^{G} is compact. Then, following [30, Comment (4.18)], for all xAGx\in A^{G}, any accumulation point of the sequence {γEm(,x)}m\left\{\gamma_{E_{m}}(\cdot,x)\right\}_{m}, will be a DLR measure μ\mu. Finally, if we want to obtain a GG-invariant DLR measure, for each gGg\in G, let gμg\mu be given by gμ(A)=μ(g1A)g\mu(A)=\mu(g^{-1}\cdot A), for any AA\in\mathcal{B}. Notice that, for every gGg\in G, the measure gμg\mu is also a DLR measure for ϕ|AG\phi|_{A^{G}} due to the GG-invariance of γ\gamma (see Corollary 4.10). Then it suffices to consider any accumulation point of the sequence {1|Fn|gFngμ}n\left\{\frac{1}{|F_{n}|}\sum_{g\in F_{n}}g\mu\right\}_{n}, for a Følner sequence {Fn}n\{F_{n}\}_{n}.

Now, let {An}n\{A_{n}\}_{n} in ()\mathcal{F}(\mathbb{N}) be a fixed exhaustion of \mathbb{N} and, for each nn\in\mathbb{N}, denote the set of DLR measures and GG-invariant DLR measures for ϕn=ϕ|AnG\phi^{n}=\phi|_{A^{G}_{n}} by 𝒢n(ϕ)\mathcal{G}_{n}(\phi) and 𝒢nI(ϕ)\mathcal{G}_{n}^{I}(\phi), respectively. For each nn\in\mathbb{N} and each μn𝒢nI(ϕ)\mu_{n}\in\mathcal{G}_{n}^{I}(\phi), consider its extension μ~n(X)\tilde{\mu}_{n}\in\mathcal{M}(X) given by

μ~n()=μn(AnG).\tilde{\mu}_{n}(\cdot)=\mu_{n}(\cdot\cap A_{n}^{G}).

The next result establishes that {μ~n}n\{\tilde{\mu}_{n}\}_{n} is tight and the reader can compare this to [45, Lemma 5.15].

Lemma 5.10.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to some exhausting sequence {Em}m\{E_{m}\}_{m}. Then, for any sequence {μn}n\{\mu_{n}\}_{n} with μn𝒢nI(ϕ)\mu_{n}\in\mathcal{G}_{n}^{I}(\phi), for all nn\in\mathbb{N}, the sequence of extensions {μ~n}n\{\tilde{\mu}_{n}\}_{n} is tight.

Proof.

Fix some nn\in\mathbb{N}. Then, for any aa\in\mathbb{N} and any yAn{1G}cy\in A_{n}^{\{1_{G}\}^{c}}, Proposition 5.7 yields that

exp(C(ϕn))μn([a])exp(ϕn(ay)logZE1(ϕn))exp(C(ϕn)),\displaystyle\mathop{\textrm{\rm exp}}\nolimits\left(-C(\phi^{n})\right)\leq\frac{\mu_{n}([a])}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi^{n}(ay)-\log Z_{E_{1}}(\phi^{n})\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi^{n})\right),

where ϕn=ϕ|AnG\phi^{n}=\phi|_{A_{n}^{G}} and C(ϕn)=2VE1(ϕn)+3δ(ϕn)C(\phi^{n})=2V_{E_{1}}(\phi^{n})+3\delta(\phi^{n}). Furthermore, 0<ZE1(ϕn)ZE1(ϕn+1)<0<Z_{E_{1}}(\phi^{n})\leq Z_{E_{1}}(\phi^{n+1})<\infty and {ZE1(ϕn)}n\{Z_{E_{1}}(\phi^{n})\}_{n} converges monotonically to ZE1(ϕ)Z_{E_{1}}(\phi). In particular, there exists c=logZE1(ϕ1)c=-\log Z_{E_{1}}(\phi^{1}) such that clogZE1(ϕn)c\geq-\log Z_{E_{1}}(\phi^{n}), for all nn\in\mathbb{N}.

If aAna\notin A_{n}, then μ~n([a])=0\tilde{\mu}_{n}([a])=0. On the other hand, if aAna\in A_{n}, then for every yAn{1G}cy\in A_{n}^{\{1_{G}\}^{c}},

μ~n([a])=μn([a])\displaystyle\tilde{\mu}_{n}([a])=\mu_{n}([a]) exp(C(ϕn))exp(ϕn(ay)logZE1(ϕn))\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi^{n})\right)\mathop{\textrm{\rm exp}}\nolimits\left(\phi^{n}(ay)-\log Z_{E_{1}}(\phi^{n})\right)
exp(C(ϕ)+ϕ(ay)+c),\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi)+\phi(ay)+c\right),

where C(ϕ)=2VE1(ϕ)+3δ(ϕ)C(\phi)=2V_{E_{1}}(\phi)+3\delta(\phi)

Now, let ϵ>0\epsilon>0. Since ϕ\phi is exp-summable, for each mm\in\mathbb{N}, there must exist a finite alphabet Aϵ,m()A_{\epsilon,m}\in\mathcal{F}(\mathbb{N}) such that

(26) bAϵ,mexp(supx[b]ϕ(x))<ϵexp(C(ϕ)c)2m|EmEm1|.\sum_{b\in\mathbb{N}\setminus A_{\epsilon,m}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup_{x\in[b]}\phi(x)\right)<\frac{\epsilon\cdot\mathop{\textrm{\rm exp}}\nolimits\left(-C(\phi)-c\right)}{2^{m}|E_{m}\setminus E_{m-1}|}.

Let

Kϵ=Aϵ,1E1×Aϵ,2E2E1×Aϵ,3E3E2×.K_{\epsilon}=A_{\epsilon,1}^{E_{1}}\times A_{\epsilon,2}^{E_{2}\setminus E_{1}}\times A_{\epsilon,3}^{E_{3}\setminus E_{2}}\times\cdots.

By Tychonoff’s Theorem (see [47]), KϵK_{\epsilon} is compact. Moreover, notice that

Kϵ=m=1gEmEm1aAϵ,m[ag],K_{\epsilon}=\bigcap_{m=1}^{\infty}\bigcap_{g\in E_{m}\setminus E_{m-1}}\bigcup_{a\in A_{\epsilon,m}}[a^{g}],

where [ag]={xX:x(g)=a}[a^{g}]=\{x\in X:x(g)=a\}. Therefore, for each nn\in\mathbb{N},

μ~n(XKϵ)\displaystyle\tilde{\mu}_{n}\left(X\setminus K_{\epsilon}\right) =μ~n(m=1gEmEm1aAϵ,m[ag]c)\displaystyle=\tilde{\mu}_{n}\left(\bigcup_{m=1}^{\infty}\bigcup_{g\in E_{m}\setminus E_{m-1}}\bigcap_{a\in A_{\epsilon,m}}[a^{g}]^{c}\right)
m=1gEmEm1μ~n(aAϵ,m[ag]c)\displaystyle\leq\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\tilde{\mu}_{n}\left(\bigcap_{a\in A_{\epsilon,m}}[a^{g}]^{c}\right)
=m=1gEmEm1μ~n(bAϵ,m[bg])\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\tilde{\mu}_{n}\left(\bigsqcup_{b\in\mathbb{N}\setminus A_{\epsilon,m}}[b^{g}]\right)
=m=1gEmEm1bAϵ,mμ~n([bg]).\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\sum_{b\in\mathbb{N}\setminus A_{\epsilon,m}}\tilde{\mu}_{n}\left(\left[b^{g}\right]\right).

Since all the measures considered here are GG-invariant, it follows that, for any yAn{1G}cy\in A_{n}^{\{1_{G}\}^{c}},

μ~n(XKϵ)\displaystyle\tilde{\mu}_{n}\left(X\setminus K_{\epsilon}\right) m=1gEmEm1bAϵ,mμ~n([b])\displaystyle\leq\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\sum_{b\in\mathbb{N}\setminus A_{\epsilon,m}}\tilde{\mu}_{n}\left([b]\right)
m=1gEmEm1bAϵ,mexp(C(ϕ)+ϕ(by)+c)\displaystyle\leq\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\sum_{b\in\mathbb{N}\setminus A_{\epsilon,m}}\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi)+\phi(by)+c\right)
=m=1gEmEm1exp(C(ϕ)+c)bAϵ,mexp(ϕ(by))\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi)+c\right)\sum_{b\in\mathbb{N}\setminus A_{\epsilon,m}}\mathop{\textrm{\rm exp}}\nolimits(\phi(by))
<m=1gEmEm1exp(C(ϕ)+c)ϵexp(C(ϕ)c)2m|EmEm1|\displaystyle<\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\mathop{\textrm{\rm exp}}\nolimits\left(C(\phi)+c\right)\frac{\epsilon\cdot\mathop{\textrm{\rm exp}}\nolimits\left(-C(\phi)-c\right)}{2^{m}|E_{m}\setminus E_{m-1}|}
=m=1gEmEm1ϵ2m|EmEm1|\displaystyle=\sum_{m=1}^{\infty}\sum_{g\in E_{m}\setminus E_{m-1}}\frac{\epsilon}{2^{m}|E_{m}\setminus E_{m-1}|}
=ϵ,\displaystyle=\epsilon,

where the fifth line follows from estimate (26). Therefore, for all nn\in\mathbb{N}, μ~n(XKϵ)<ϵ\tilde{\mu}_{n}\left(X\setminus K_{\epsilon}\right)<\epsilon, so that μ~n(Kϵ)=1μ~n(Kϵc)>1ϵ\tilde{\mu}_{n}\left(K_{\epsilon}\right)=1-\tilde{\mu}_{n}\left(K_{\epsilon}^{c}\right)>1-\epsilon, which proves the tightness of {μ~n}n\{\tilde{\mu}_{n}\}_{n}. ∎

We have proven that for each sequence {μn}n\{\mu_{n}\}_{n} with μn𝒢nI(ϕ)\mu_{n}\in\mathcal{G}^{I}_{n}(\phi), the sequence {μ~n}n\{\tilde{\mu}_{n}\}_{n} of their extensions is tight. Then, the existence of at least one accumulation point is guaranteed by Prokhorov’s Theorem. Let’s see that an arbitrary accumulation point, which we will denote by μ~\tilde{\mu}, is conformal for ϕ\phi and, moreover, that it is GG-invariant.

Theorem 5.11.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, the set of GG-invariant DLR measures for ϕ\phi is non-empty.

Proof.

Let {μn}n\{\mu_{n}\}_{n} be such that, for each nn\in\mathbb{N}, μn\mu_{n} is a GG-invariant conformal measure for ϕn:AnG\phi^{n}:A_{n}^{G}\to\mathbb{R} (or, equivalently, μn\mu_{n} is a GG-invariant DLR measure for ϕn\phi^{n}). Thus, for each nn\in\mathbb{N}, any K(G)K\in\mathcal{F}(G), and any τK,An\tau\in\mathcal{E}_{K,A_{n}},

(27) exp((ϕn)τn)=d(μn(τn)1)dμn,\mathop{\textrm{\rm exp}}\nolimits\left((\phi^{n})^{\tau_{n}}_{*}\right)=\frac{d(\mu_{n}\circ(\tau_{n})^{-1})}{d\mu_{n}},

where τn=τ|AnG\tau_{n}=\tau|_{A^{G}_{n}}. This yields that

exp(ϕτ)=d(μ~nτ1)dμ~n.\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau})=\frac{d(\tilde{\mu}_{n}\circ\tau^{-1})}{d\tilde{\mu}_{n}}.

Indeed, let ψ:X\psi\colon X\to\mathbb{R} be a bounded continuous potential. Observe that, for τK,An\tau\in\mathcal{E}_{K,A_{n}}, (ϕn)τn=(ϕτ)|AnG(\phi^{n})^{\tau_{n}}_{*}=(\phi_{*}^{\tau})|_{A_{n}^{G}}. Moreover, for every BB\in\mathcal{B}, since τn1(AnG)=AnG\tau_{n}^{-1}(A^{G}_{n})=A^{G}_{n} and μ~n(XAnG)=0\tilde{\mu}_{n}(X\setminus A_{n}^{G})=0, we have that μnτn1~(B)=μ~n(τ1(B))\widetilde{{\mu_{n}\circ\tau_{n}^{-1}}}(B)=\tilde{\mu}_{n}(\tau^{-1}(B)). Then, we obtain

ψd(μ~nτ1)\displaystyle\int\psi d(\tilde{\mu}_{n}\circ\tau^{-1}) =ψd(μnτn1~)\displaystyle=\int\psi d(\widetilde{\mu_{n}\circ\tau_{n}^{-1}})
=ψnd(μnτn1~)\displaystyle=\int\psi^{n}d(\widetilde{\mu_{n}\circ\tau_{n}^{-1}})
=ψnd(μnτn1)\displaystyle=\int\psi^{n}d(\mu_{n}\circ\tau_{n}^{-1})
=ψnexp((ϕn)τn)dμn\displaystyle=\int\psi^{n}\mathop{\textrm{\rm exp}}\nolimits\left((\phi^{n})^{\tau_{n}}_{*}\right)d\mu_{n}
=ψnexp((ϕn|AnG)τ)dμn\displaystyle=\int\psi^{n}\mathop{\textrm{\rm exp}}\nolimits\left((\phi^{n}|_{A_{n}^{G}})^{\tau}_{*}\right)d\mu_{n}
=ψexp(ϕτ)dμ~n,\displaystyle=\int\psi\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau})d\tilde{\mu}_{n},

where ψn=ψ|AnG\psi^{n}=\psi|_{A_{n}^{G}}.

Furthermore, Lemma 5.10 guarantees that the sequence of induced measures {μ~n}n\{\tilde{\mu}_{n}\}_{n} is tight and we can apply Prokhorov’s Theorem to guarantee the existence of a limit point for some subsequence {μnk}k\{\mu_{n_{k}}\}_{k}, which we denote by μ~\tilde{\mu}. Now, we are going to prove that μ~\tilde{\mu} is a conformal measure for ϕ\phi. For that, consider a bounded continuous potential ψ:X\psi\colon X\to\mathbb{R}, A()A\in\mathcal{F}(\mathbb{N}), K(G)K\in\mathcal{F}(G), and τK,A\tau\in\mathcal{E}_{K,A}. Then,

ψd(μ~τ1)\displaystyle\int\psi\,d(\tilde{\mu}\circ\tau^{-1}) =ψτ𝑑μ~\displaystyle=\int\psi\circ\tau\,d\tilde{\mu}
=limkψτ𝑑μ~nk\displaystyle=\lim_{k\to\infty}\int\psi\circ\tau\,d\tilde{\mu}_{n_{k}}
=limkψd(μ~nkτ1)\displaystyle=\lim_{k\to\infty}\int\psi\,d(\tilde{\mu}_{n_{k}}\circ\tau^{-1})
=limkψexpϕτdμ~nk\displaystyle=\lim_{k\to\infty}\int\psi\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}\,d\tilde{\mu}_{n_{k}}
=ψexpϕτdμ~,\displaystyle=\int\psi\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau}\,d\tilde{\mu},

where the fourth equality follows from the fact that for kk large enough, AAnkA\subseteq A_{n_{k}}, and the last equality follows from weak convergence and the fact that ψexpϕτ\psi\mathop{\textrm{\rm exp}}\nolimits\phi_{*}^{\tau} is a continuous and bounded function. Indeed, first notice that ϕτ\phi_{*}^{\tau} is a uniform limit of continuous functions that are bounded from above, since ϕ\phi is exp-summable. Therefore, the same holds for ϕτ\phi_{*}^{\tau}, so that exp(ϕτ)\mathop{\textrm{\rm exp}}\nolimits(\phi_{*}^{\tau}) is continuous and bounded (from above and below). Since AA, KK, and τ\tau are arbitrary, this proves that μ~\tilde{\mu} is conformal for ϕ\phi and, therefore, DLR for ϕ\phi.

It remains to show that μ~\tilde{\mu} is GG-invariant. For that, notice that, due to the weak convergence, for any BB\in\mathcal{B},

μ~(gB)=limkμ~nk(gB)=limkμ~nk(B)=μ(B),\tilde{\mu}(g\cdot B)=\lim_{k\to\infty}\tilde{\mu}_{n_{k}}(g\cdot B)=\lim_{k\to\infty}\tilde{\mu}_{n_{k}}(B)=\mu(B),

where we have used that, for each kk\in\mathbb{N}, μ~nk\tilde{\mu}_{n_{k}} is GG-invariant due to GG-invariance of AnGA_{n}^{G} and to the fact that μnk\mu_{n_{k}} is GG-invariant. ∎

5.5. Finite entropy Bowen-Gibbs measures are equilibrium measures

Thus far, we have proven that if ϕ:X\phi\colon X\to\mathbb{R} is an exp-summable potential with summable variation, then a measure μ(X)\mu\in\mathcal{M}(X) is a DLR measure if and only if it is a conformal measure. Also, if μ\mu is a DLR measure, then μ\mu is also a Bowen-Gibbs measure. For Bowen-Gibbs measures, we begin by exploring some equivalent hypothesis to having HF(μ)<H_{F}(\mu)<\infty for every F(G)F\in\mathcal{F}(G), or, equivalently, to have finite Shannon entropy at the identity element. This will allow us to assume, indistinctly, that the energy of the potential is finite. The following lemma generalizes [43, Lemma 3.4].

Proposition 5.12.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Then, if μ(X)\mu\in\mathcal{M}(X) is a Bowen-Gibbs measure for ϕ\phi, the following conditions are equivalent:

  • i)i)

    ϕ𝑑μ>\int\phi d\mu>-\infty;

  • ii)ii)

    asupϕ([a])exp(supϕ([a]))>\sum_{a\in\mathbb{N}}\sup\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([a])\right)>-\infty; and

  • iii)iii)

    H(μ)<H(\mu)<\infty.

Proof.

Begin by noticing that, since μ\mu is a Bowen-Gibbs measure for ϕ\phi, we have that, in particular, for ϵ=1\epsilon=1, there exist K(G)K\in\mathcal{F}(G), δ>0\delta>0, and a (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) with 1GF1_{G}\in F such that, for every xXx\in X, it holds that

(28) exp(|F|(1+p(ϕ))+supϕF([xF]))μ([xF])exp(|F|(1+p(ϕ))+supϕF([xF])).\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi))+\sup\phi_{F}([x_{F}])\right)\leq\mu([x_{F}])\leq\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(-1+p(\phi))+\sup\phi_{F}([x_{F}])\right).

We now prove that i)iii)ii)i)i)\implies iii)\implies ii)\implies i).

[i)iii)i)\implies iii)] Notice that, since ϕ\phi has summable variation according to {Em}m\{E_{m}\}_{m}, then, in particular, ϕ\phi has finite oscillation. Therefore, the result follows directly from Proposition 5.2, disregarding whether μ\mu is a Bowen-Gibbs measure for ϕ\phi or not.

[iii)ii)iii)\implies ii)] Begin by noticing that, due to standard properties of Shannon entropy, H(μ)HF(μ)|F|H(μ)H(\mu)\leq H_{F}(\mu)\leq|F|H(\mu). Then,

\displaystyle-\infty <HF(μ)\displaystyle<-H_{F}(\mu)
=xFXFμ([xF])logμ([xF])\displaystyle=\sum_{x_{F}\in X_{F}}\mu([x_{F}])\log\mu([x_{F}])
xFXFμ([xF])(|F|(1+p(ϕ))+supϕF([xF]))\displaystyle\leq\sum_{x_{F}\in X_{F}}\mu([x_{F}])\left(-|F|(-1+p(\phi))+\sup\phi_{F}([x_{F}])\right)
=|F|(1+p(ϕ))+xFXFμ([xF])supϕF([xF]).\displaystyle=-|F|(1+p(\phi))+\sum_{x_{F}\in X_{F}}\mu([x_{F}])\sup\phi_{F}([x_{F}]).

Thus,

\displaystyle-\infty <xFXFμ([xF])supϕF([xF])\displaystyle<\sum_{x_{F}\in X_{F}}\mu([x_{F}])\sup\phi_{F}([x_{F}])
xFXFexp(|F|(1+p(ϕ))+supϕF([xF]))supϕF([xF])\displaystyle\leq\sum_{x_{F}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(-1+p(\phi))+\sup\phi_{F}([x_{F}])\right)\cdot\sup\phi_{F}([x_{F}])
=exp(|F|(1+p(ϕ)))xFXFexp(supϕF([xF]))supϕF([xF]),\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(-1+p(\phi))\right)\sum_{x_{F}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([x_{F}])\right)\cdot\sup\phi_{F}([x_{F}]),

so that

<xFXFexp(supϕF([xF]))supϕF([xF]).-\infty<\sum_{x_{F}\in X_{F}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([x_{F}])\right)\sup\phi_{F}([x_{F}]).

Also, for each xFXFx_{F}\in X_{F},

supϕF([xF])infϕF([xF])gFinf(ϕ{g}([xF]))gFinf(ϕ{g}([xg])).\sup\phi_{F}([x_{F}])\geq\inf\phi_{F}([x_{F}])\geq\sum_{g\in F}\inf\left(\phi_{\{g\}}([x_{F}])\right)\geq\sum_{g\in F}\inf\left(\phi_{\{g\}}([x_{g}])\right).

Now, due to exp-summability, without loss of generality we can assume that ϕ(x)0\phi(x)\leq 0, for all xXx\in X, so supϕF([xF])supϕF([x1G])supϕ([x1G])0\sup\phi_{F}([x_{F}])\leq\sup\phi_{F}([x_{1_{G}}])\leq\sup\phi([x_{1_{G}}])\leq 0. Then, abbreviating ϕ{g}\phi_{\{g\}} by ϕg\phi_{g}, we obtain that

\displaystyle-\infty <xFXFsupϕF([xF])exp(supϕF([xF]))\displaystyle<\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{F}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([x_{F}])\right)
xFXFsupϕF([xF])gFexp(infϕg([xg]))\displaystyle\leq\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{F}])\prod_{g\in F}\mathop{\textrm{\rm exp}}\nolimits\left(\inf\phi_{g}([x_{g}])\right)
xFXFsupϕF([xF])gFexp(supϕg([xg])δ(ϕ))\displaystyle\leq\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{F}])\prod_{g\in F}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])-\delta(\phi)\right)
=exp(δ(ϕ)|F|)xFXFsupϕF([xF])gFexp(supϕg([xg]))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-\delta(\phi)|F|\right)\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{F}])\prod_{g\in F}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right)
=exp(δ(ϕ)|F|)xFXFsupϕF([xF])exp(supϕ([x1G]))gF{1G}exp(supϕg([xg]))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-\delta(\phi)|F|\right)\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{F}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right)
exp(δ(ϕ)|F|)xFXFsupϕF([x1G])exp(supϕ([x1G]))gF{1G}exp(supϕg([xg]))\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(-\delta(\phi)|F|\right)\sum_{x_{F}\in X_{F}}\sup\phi_{F}([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right)
=exp(δ(ϕ)|F|)x1GsupϕF([x1G])exp(supϕ([x1G]))xF{1G}gF{1G}exp(supϕg([xg]))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-\delta(\phi)|F|\right)\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi_{F}([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\sum_{x_{F\setminus\{1_{G}\}}}\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right)
exp(δ(ϕ)|F|)x1Gsupϕ([x1G])exp(supϕ([x1G]))xF{1G}gF{1G}exp(supϕg([xg])).\displaystyle\leq\mathop{\textrm{\rm exp}}\nolimits\left(-\delta(\phi)|F|\right)\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\sum_{x_{F\setminus\{1_{G}\}}}\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right).

Moreover, notice that if m=|F|1m=|F|-1 and g1,,gmg_{1},\cdots,g_{m} is an enumeration of F{1G}F\setminus\{1_{G}\}, then

xF{1G}gF{1G}exp(supϕg([xg]))\displaystyle\sum_{x_{F\setminus\{1_{G}\}}}\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right) =xg1xgmexp(supϕg1([xg1]))exp(supϕgm([xgm]))\displaystyle=\sum_{x_{g_{1}}}\cdots\sum_{x_{g_{m}}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g_{1}}([x_{g_{1}}])\right)\cdots\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g_{m}}([x_{g_{m}}])\right)
=xg1exp(supϕg1([xg1]))xgmexp(supϕgm([xgm]))\displaystyle=\sum_{x_{g_{1}}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g_{1}}([x_{g_{1}}])\right)\cdots\sum_{x_{g_{m}}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g_{m}}([x_{g_{m}}])\right)
=gF{1G}xgXgexp(supϕg([xg])),\displaystyle=\prod_{g\in F\setminus\{1_{G}\}}\sum_{x_{g}\in X_{g}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right),

so that

\displaystyle-\infty <x1Gsupϕ([x1G])exp(supϕ([x1G]))gF{1G}xgXgexp(supϕg([xg]))\displaystyle<\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\prod_{g\in F\setminus\{1_{G}\}}\sum_{x_{g}\in X_{g}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{g}([x_{g}])\right)
=x1Gsupϕ([x1G])exp(supϕ([x1G]))gF{1G}Zg(ϕ)\displaystyle=\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\prod_{g\in F\setminus\{1_{G}\}}Z_{g}(\phi)
=x1Gsupϕ([x1G])exp(supϕ([x1G]))gF{1G}Z1G(ϕ)\displaystyle=\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)\prod_{g\in F\setminus\{1_{G}\}}Z_{1_{G}}(\phi)
=x1Gsupϕ([x1G])exp(supϕ([x1G]))Z1G(ϕ)|F|1.\displaystyle=\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)Z_{1_{G}}(\phi)^{|F|-1}.

Therefore,

x1Gsupϕ([x1G])exp(supϕ([x1G]))>.\sum_{x_{1_{G}}\in\mathbb{N}}\sup\phi([x_{1_{G}}])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi([x_{1_{G}}])\right)>-\infty.

[ii)i)ii)\implies i)] Indeed,

ϕ𝑑μ\displaystyle\int\phi d\mu ainfϕ([a])μ([a])\displaystyle\geq\sum_{a\in\mathbb{N}}\inf\phi([a])\mu([a])
=ainfϕ([a])xF:xF(1G)=aμ([xF])\displaystyle=\sum_{a\in\mathbb{N}}\inf\phi([a])\sum_{\begin{subarray}{c}x_{F}:x_{F}(1_{G})=a\end{subarray}}\mu([x_{F}])
ainfϕ([a])xF:xF(1G)=aexp(|F|(1+p(ϕ))+supϕF([xF]))\displaystyle\geq\sum_{a\in\mathbb{N}}\inf\phi([a])\sum_{\begin{subarray}{c}x_{F}:x_{F}(1_{G})=a\end{subarray}}\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi))+\sup\phi_{F}([x_{F}])\right)
=exp(|F|(1+p(ϕ)))ainfϕ([a])xF:xF(1G)=aexp(supϕF([xF]))\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi))\right)\sum_{a\in\mathbb{N}}\inf\phi([a])\sum_{\begin{subarray}{c}x_{F}:x_{F}(1_{G})=a\end{subarray}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}([x_{F}])\right)
exp(|F|(1+p(ϕ)))ainfϕ([a])xF:xF(1G)=aexp(gFsupϕg([xF]))\displaystyle\geq\mathop{\textrm{\rm exp}}\nolimits(-|F|(1+p(\phi)))\sum_{a\in\mathbb{N}}\inf\phi([a])\sum_{\begin{subarray}{c}x_{F}:x_{F}(1_{G})=a\end{subarray}}\mathop{\textrm{\rm exp}}\nolimits\left(\sum_{g\in F}\sup\phi_{g}([x_{F}])\right)
exp(|F|(1+p(ϕ))ainfϕ([a])xF:xF(1G)=agFexp(supϕ[x(g)])\displaystyle\geq\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi)\right)\sum_{a\in\mathbb{N}}\inf\phi([a])\sum_{\begin{subarray}{c}x_{F}:x_{F}(1_{G})=a\end{subarray}}\prod_{g\in F}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[x(g)]\right)
=exp(|F|(1+p(ϕ))ainfϕ([a])exp(supϕ[a])xF{1G}gF{1G}exp(supϕ[x(g)]).\displaystyle=\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi)\right)\sum_{a\in\mathbb{N}}\inf\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)\sum_{x_{F\setminus\{1_{G}\}}}\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[x(g)]\right).

Notice that, due to the same argument as in the proof of [iii)ii)][iii)\implies ii)], we have that

xF{1G}gF{1G}exp(supϕ[x(g)])=Z1G(ϕ)|F|1.\sum_{x_{F\setminus\{1_{G}\}}}\prod_{g\in F\setminus\{1_{G}\}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[x(g)]\right)=Z_{1_{G}}(\phi)^{|F|-1}.

Therefore, since exp(|F|(1+p(ϕ))Z1G(ϕ)|F|1>0\mathop{\textrm{\rm exp}}\nolimits\left(-|F|(1+p(\phi)\right)Z_{1_{G}}(\phi)^{|F|-1}>0, it suffices to prove that

ainfϕ([a])exp(supϕ[a])>,\displaystyle\sum_{a\in\mathbb{N}}\inf\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)>-\infty,

but this is true since

ainfϕ([a])exp(supϕ[a])\displaystyle\sum_{a\in\mathbb{N}}\inf\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right) a(supϕ([a])δ(ϕ))exp(supϕ[a])\displaystyle\geq\sum_{a\in\mathbb{N}}(\sup\phi([a])-\delta(\phi))\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)
=asupϕ([a])exp(supϕ[a])δ(ϕ)aexp(supϕ[a])\displaystyle=\sum_{a\in\mathbb{N}}\sup\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)-\delta(\phi)\sum_{a\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)
=asupϕ([a])exp(supϕ[a])δ(ϕ)Z1G(ϕ)\displaystyle=\sum_{a\in\mathbb{N}}\sup\phi([a])\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi[a]\right)-\delta(\phi)\cdot Z_{1_{G}}(\phi)
>.\displaystyle>-\infty.

We now proceed to prove that Bowen-Gibbs measures with finite Shannon entropy at the identity are equilibrium measures.

Theorem 5.13.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If μ(X)\mu\in\mathcal{M}(X) is a GG-invariant Bowen-Gibbs measure for ϕ\phi and H(μ)<H(\mu)<\infty, then μ\mu is an equilibrium measure for ϕ\phi.

Proof.

Since μ\mu is a Bowen-Gibbs measure for ϕ\phi, for every ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0, such that for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) and xXx\in X,

(29) exp(ϵ|F|)μ([xF])exp(ϕF(x)p(ϕ)|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\mu([x_{F}])}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)-p(\phi)\cdot|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).

Moreover, notice that, for every xXx\in X and F(G)F\in\mathcal{F}(G),

(30) supϕF([xF])ϕF(x)+ΔF(ϕ)=gFϕ(gx)+ΔF(ϕ).\sup\phi_{F}([x_{F}])\leq\phi_{F}(x)+\Delta_{F}(\phi)=\sum_{g\in F}\phi(g\cdot x)+\Delta_{F}(\phi).

Therefore,

limFG1|F|supϕF([xF])dμ(x)\displaystyle\lim_{F\to G}\frac{1}{|F|}\int{\sup\phi_{F}([x_{F}])}\,d\mu(x) limFG1|F|(gFϕ(gx)+ΔF(ϕ))dμ(x)\displaystyle\leq\lim_{F\to G}\frac{1}{|F|}\int{\left(\sum_{g\in F}\phi(g\cdot x)+\Delta_{F}(\phi)\right)}\,d\mu(x)
=limFG1|F|(gFϕ(x)dμ(x))+limFGΔF(ϕ)|F|\displaystyle=\lim_{F\to G}\frac{1}{|F|}\left(\sum_{g\in F}\int\phi(x)\,d\mu(x)\right)+\lim_{F\to G}\frac{\Delta_{F}(\phi)}{|F|}
=limFG1|F|(|F|ϕdμ)\displaystyle=\lim_{F\to G}\frac{1}{|F|}\left(|F|\int{\phi}\,d\mu\right)
=ϕdμ,\displaystyle=\int{\phi}d\mu,

where the second line follows from the GG-invariance of μ\mu and the third line follows from Lemma 2.3.

On the other hand, after taking logarithm in equation (29) and dividing by |F||F|, we obtain

ϵlogμ([xF])ϕF(x)|F|+p(ϕ)ϵ.-\epsilon\leq\frac{\log\mu([x_{F}])-\phi_{F}(x)}{|F|}+p(\phi)\leq\epsilon.

Thus, for every xXx\in X and every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G),

p(ϕ)logμ([xF])+ϕF(x)|F|+ϵlogμ([xF])+supϕF([xF])|F|+ϵ.p(\phi)\leq\frac{-\log\mu([x_{F}])+\phi_{F}(x)}{|F|}+\epsilon\leq\frac{-\log\mu([x_{F}])+\sup\phi_{F}([x_{F}])}{|F|}+\epsilon.

Integrating the last equation with respect to μ\mu, we get

p(ϕ)\displaystyle p(\phi) 1|F|xFXFμ([xF])logμ([xF])+1|F|supϕF([xF])dμ+ϵ\displaystyle\leq\frac{-1}{|F|}\sum_{x_{F}\in X_{F}}\mu([x_{F}])\log\mu([x_{F}])+\frac{1}{|F|}\int{\sup\phi_{F}([x_{F}])}d\mu+\epsilon
=1|F|HF(μ)+1|F|supϕF([xF])dμ+ϵ.\displaystyle=\frac{1}{|F|}H_{F}(\mu)+\frac{1}{|F|}\int{\sup\phi_{F}([x_{F}])}d\mu+\epsilon.

Therefore, if we take limit as FF becomes more and more invariant, we have that

p(ϕ)h(μ)+limFG1|F|supϕF([xF])dμ+ϵh(μ)+ϕdμ+ϵ,\displaystyle p(\phi)\leq h(\mu)+\lim_{F\to G}\frac{1}{|F|}\int{\sup\phi_{F}([x_{F}])}d\mu+\epsilon\leq h(\mu)+\int{\phi}d\mu+\epsilon,

where the last inequality follows from inequality (30). Since ϵ>0\epsilon>0 is arbitrary, we obtain that

p(ϕ)h(μ)+ϕdμ.p(\phi)\leq h(\mu)+\int{\phi}d\mu.

The reverse inequality follows from Proposition 5.3 and this concludes the proof. ∎

5.6. Equilibrium measures are DLR measures

In §5.4, we proved that if ϕ:X\phi\colon X\to\mathbb{R} is an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, then the set of GG-invariant DLR measures for ϕ\phi is non-empty. Throughout this section, fix a GG-invariant ν𝒢(ϕ)\nu\in\mathcal{G}(\phi).

Given E(G)E\in\mathcal{F}(G) and μG(X)\mu\in\mathcal{M}_{G}(X), denote by fμ,Ef_{\mu,E} the Radon-Nikodym derivative of μ|E\mu|_{E} with respect to ν|E\nu|_{E}, where μ|E\mu|_{E} and ν|E\nu|_{E} denote the restrictions of μ\mu and ν\nu to E\mathcal{B}_{E}, respectively. More precisely, for every xXx\in X,

(31) fμ,E(x)=wXEμ([w])ν([w])𝟙[w](x).f_{\mu,E}(x)=\sum_{w\in X_{E}}\frac{\mu([w])}{\nu([w])}\mathbbm{1}_{[w]}(x).

Notice that fμ,Ef_{\mu,E} is well-defined, because any DLR measure for ϕ\phi, in our context, is fully supported. Moreover, we can understand it as the pointwise limit of the simple functions fμ,En=wXEAnGμ([w])ν([w])𝟙[w]f_{\mu,E}^{n}=\sum_{w\in X_{E}\cap A_{n}^{G}}\frac{\mu([w])}{\nu([w])}\mathbbm{1}_{[w]}, where {An}n\{A_{n}\}_{n} is a fixed exhausting sequence of finite alphabets.

Consider the function ψ:[0,)[0,)\psi\colon[0,\infty)\to[0,\infty) given by ψ(x)=1x+xlogx\psi(x)=1-x+x\log x, where 0log(0)=00\log(0)=0. Define, for each nn\in\mathbb{N} and E(G)E\in\mathcal{F}(G), the simple function Inμ,E:=wXEAnGψ(μ([w])ν([w]))𝟙[w]I^{n}_{\mu,E}:=\sum_{w\in X_{E}\cap A_{n}^{G}}\psi\left(\frac{\mu([w])}{\nu([w])}\right)\mathbbm{1}_{[w]}. Notice that 0Inμ,E(x)In+1μ,E(x)0\leq I^{n}_{\mu,E}(x)\leq I^{n+1}_{\mu,E}(x), so we can define a measurable function Iμ,EI_{\mu,E} by considering the pointwise limit IE(x):=limnInE(x)I_{E}(x):=\lim_{n\to\infty}I^{n}_{E}(x) in [0,][0,\infty].

When there is no ambiguity, we will omit the subscript μ\mu from the previous notations.

Observe that, by the Monotone Convergence Theorem,

limnInEdν=limnInEdν=IEdν[0,].\lim_{n\to\infty}\int I^{n}_{E}d\nu=\int\lim_{n\to\infty}I^{n}_{E}d\nu=\int I_{E}d\nu\in[0,\infty].

In addition,

HnE(μ|ν):=InEdν=wXEAnG(ν([w])μ([w])+μ([w])log(μ([w])ν([w]))),\displaystyle H^{n}_{E}(\mu|\nu):=\int I^{n}_{E}d\nu=\sum_{w\in X_{E}\cap A_{n}^{G}}\left(\nu([w])-\mu([w])+\mu([w])\log\left(\frac{\mu([w])}{\nu([w])}\right)\right),

so that

IEdν\displaystyle\int I_{E}d\nu =limnInEdν\displaystyle=\lim_{n\to\infty}\int I^{n}_{E}d\nu
=wXE(ν([w])μ([w])+μ([w])log(μ([w])ν([w])))\displaystyle=\sum_{w\in X_{E}}\left(\nu([w])-\mu([w])+\mu([w])\log\left(\frac{\mu([w])}{\nu([w])}\right)\right)
=wXEμ([w])log(μ([w])ν([w])).\displaystyle=\sum_{w\in X_{E}}\mu([w])\log\left(\frac{\mu([w])}{\nu([w])}\right).

We define the relative entropy of a measure μG(X)\mu\in\mathcal{M}_{G}(X) with respect to ν\nu to be

HE(μ|ν):=Iμ,Edν,H_{E}(\mu|\nu):=\int I_{\mu,E}d\nu,

when E(G)E\in\mathcal{F}(G), and 0 if E=E=\emptyset. Notice that, a priori, HE(μ|ν)[0,]H_{E}(\mu|\nu)\in[0,\infty]. Also, if μG(X)\mu\in\mathcal{M}_{G}(X), then HEg(μ|ν)=HE(μ|ν)H_{Eg}(\mu|\nu)=H_{E}(\mu|\nu) for every gGg\in G.

Lemma 5.14.

Let E,F(G)E,F\in\mathcal{F}(G) be such that EFE\subseteq F and μ(X)\mu\in\mathcal{M}(X). Then, for every nn\in\mathbb{N}, HnE(μ|ν)HnF(μ|ν)H^{n}_{E}(\mu|\nu)\leq H^{n}_{F}(\mu|\nu). Moreover, HE(μ|ν)HF(μ|ν)H_{E}(\mu|\nu)\leq H_{F}(\mu|\nu).

Proof.

Fix nn\in\mathbb{N}. First, observe that fnμ,E=ν[fnμ,F|E]f^{n}_{\mu,E}=\nu[f^{n}_{\mu,F}\,|\,\mathcal{B}_{E}]. Indeed, it suffices to prove that for any vXEv\in X_{E},

(32) [v]AnGfnEdν=[v]AnGfnFdν,\int_{[v]\cap A_{n}^{G}}f^{n}_{E}\,d\nu=\int_{[v]\cap A_{n}^{G}}f^{n}_{F}\,d\nu,

since the supports of fnEf^{n}_{E} and fnFf^{n}_{F} are contained in AnGA_{n}^{G} and E\mathcal{B}_{E} is generated by cylinder sets of this form. If vAnEv\notin A_{n}^{E}, then both sides of equation (32) are 0 and the result is proven. Otherwise, if vAnEv\in A_{n}^{E}, then

[v]AnGfnFdν\displaystyle\int_{[v]\cap A_{n}^{G}}f^{n}_{F}\,d\nu =[v]AnGwXFAnGμ([w])ν([w])𝟙[w]dν\displaystyle=\int_{[v]\cap A_{n}^{G}}\sum_{w\in X_{F}\cap A_{n}^{G}}\frac{\mu([w])}{\nu([w])}\mathbbm{1}_{[w]}d\nu
=AnGwXFEAnGμ([vwFE])ν([vwFE])𝟙[vwFE]dν\displaystyle=\int_{A_{n}^{G}}\sum_{w\in X_{F\setminus E}\cap A_{n}^{G}}\frac{\mu([vw_{F\setminus E}])}{\nu([vw_{F\setminus E}])}\mathbbm{1}_{[vw_{F\setminus E}]}d\nu
=wXFEAnGμ([vwFE])ν([vwFE])AnG𝟙[vwFE]dν\displaystyle=\sum_{w\in X_{F\setminus E}\cap A_{n}^{G}}\frac{\mu([vw_{F\setminus E}])}{\nu([vw_{F\setminus E}])}\int_{A_{n}^{G}}\mathbbm{1}_{[vw_{F\setminus E}]}d\nu
=wXFEAnGμ([vwFE])\displaystyle=\sum_{w\in X_{F\setminus E}\cap A_{n}^{G}}\mu([vw_{F\setminus E}])
=μ([v]AnG)\displaystyle=\mu([v]\cap A_{n}^{G})
=[v]AnGdμ\displaystyle=\int_{[v]\cap A_{n}^{G}}d\mu
=[v]AnGfnEdν.\displaystyle=\int_{[v]\cap A_{n}^{G}}f^{n}_{E}\,d\nu.

Thus,

HnE(μ|ν)\displaystyle H^{n}_{E}(\mu|\nu) =fnElogfnEdν\displaystyle=\int{f^{n}_{E}\log f^{n}_{E}}\,d\nu
=ν(fnF|E)logν(fnF|E)dν\displaystyle=\int{\nu(f^{n}_{F}|\mathcal{B}_{E})\log\nu(f^{n}_{F}|\mathcal{B}_{E})}d\nu
ν(fnFlogfnF|E)dν\displaystyle\leq\int{\nu(f^{n}_{F}\log f^{n}_{F}|\mathcal{B}_{E})}d\nu
=HnF(μ|ν),\displaystyle=H^{n}_{F}(\mu|\nu),

where the inequality follows from Jensen’s inequality for conditional expectations. Finally, observe that

HE(μ|ν)=limnHnE(μ|ν)limnHnF(μ|ν)=HF(μ|ν).H_{E}(\mu|\nu)=\lim_{n\to\infty}H^{n}_{E}(\mu|\nu)\leq\lim_{n\to\infty}H^{n}_{F}(\mu|\nu)=H_{F}(\mu|\nu).

Proposition 5.15.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m} and μG(X)\mu\in\mathcal{M}_{G}(X). Then, HE(μ|ν)<H_{E}(\mu|\nu)<\infty for every E(G)E\in\mathcal{F}(G). Moreover, if ϕdμ>\int\phi\,d\mu>-\infty,

h(μ|ν):=limFG1|F|HF(μ|ν)=p(ϕ)(h(μ)+ϕdμ).h(\mu\,|\,\nu):=\lim_{F\to G}\frac{1}{|F|}H_{F}(\mu|\nu)=p(\phi)-\left(h(\mu)+\int\phi\,d\mu\right).
Proof.

Let E(G)E\in\mathcal{F}(G). Since ν\nu is a DLR measure for ϕ\phi, by Theorem 5.8, ν\nu is a Bowen-Gibbs measure for ϕ\phi. Then, for every ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that for all (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G), the following conditions hold at the same time:

|h(μ)HF(μ)|F||ϵ\left|h(\mu)-\frac{H_{F}(\mu)}{|F|}\right|\leq\epsilon

and

exp(ϵ|F|)ν([xF])exp(supϕF(x)p(ϕ)|F|)exp(ϵ|F|).\mathop{\textrm{\rm exp}}\nolimits\left(-\epsilon\cdot|F|\right)\leq\frac{\nu([x_{F}])}{\mathop{\textrm{\rm exp}}\nolimits\left(\sup\phi_{F}(x)-p(\phi)\cdot|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(\epsilon\cdot|F|\right).

Observe that, by considering the lower bound of the equation above,

xFXFAnGμ([xF])log(ν([xF]))\displaystyle-\sum_{x_{F}\in X_{F}\cap A_{n}^{G}}\mu([x_{F}])\log(\nu([x_{F}])) xFXFAnGμ([xF])(supϕF([xF])p(ϕ)|F|ϵ|F|)\displaystyle\leq-\sum_{x_{F}\in X_{F}\cap A_{n}^{G}}\mu([x_{F}])\left(\sup\phi_{F}([x_{F}])-p(\phi)|F|-\epsilon|F|\right)
=(p(ϕ)+ϵ)|F|xFXFAnGμ([xF])supϕF([xF])\displaystyle=(p(\phi)+\epsilon)|F|-\sum_{x_{F}\in X_{F}\cap A_{n}^{G}}\mu([x_{F}])\sup\phi_{F}([x_{F}])
(p(ϕ)+ϵ)|F|AnGϕFdμ\displaystyle\leq(p(\phi)+\epsilon)|F|-\int_{A_{n}^{G}}\phi_{F}d\mu
=(p(ϕ)+ϵAnGϕdμ)|F|,\displaystyle=\left(p(\phi)+\epsilon-\int_{A_{n}^{G}}\phi d\mu\right)|F|,

for any (K,δ)(K,\delta)-invariant set FF. Then, we have that

HF(μ|ν)\displaystyle H_{F}(\mu|\nu) =limn(HnF(μ|ν)HnF(μ))+limnHnF(μ)\displaystyle=\lim_{n\to\infty}\left(H^{n}_{F}(\mu|\nu)-H^{n}_{F}(\mu)\right)+\lim_{n\to\infty}H^{n}_{F}(\mu)
=limnxFXFAnGμ([xF])log(ν([xF]))+HF(μ)\displaystyle=\lim_{n\to\infty}-\sum_{x_{F}\in X_{F}\cap A_{n}^{G}}\mu([x_{F}])\log(\nu([x_{F}]))+H_{F}(\mu)
limn(p(ϕ)+ϵAnGϕdμ)|F|+(h(μ)+ϵ)|F|\displaystyle\leq\lim_{n\to\infty}\left(p(\phi)+\epsilon-\int_{A_{n}^{G}}\phi d\mu\right)|F|+(h(\mu)+\epsilon)|F|
=(p(ϕ)+h(μ)ϕdμ+2ϵ)|F|+HF(μ)<,\displaystyle=\left(p(\phi)+h(\mu)-\int\phi d\mu+2\epsilon\right)|F|+H_{F}(\mu)<\infty,

where HnF(μ):=xFXFAnGμ([xF])log(μ([xF]))H^{n}_{F}(\mu):=-\sum_{x_{F}\in X_{F}\cap A_{n}^{G}}\mu([x_{F}])\log(\mu([x_{F}])).

First, observe that for any EE, we can find a (K,δ)(K,\delta)-invariant set FF such that EFE\subseteq F. Then, by Lemma 5.14, HE(μ|ν)HE(μ|ν)<H_{E}(\mu|\nu)\leq H_{E}(\mu|\nu)<\infty. Second, for any (K,δ)(K,\delta)-invariant set FF,

HF(μ|ν)|F|p(ϕ)+(h(μ)ϕdμ)+2ϵ.\frac{H_{F}(\mu|\nu)}{|F|}\leq p(\phi)+\left(h(\mu)-\int\phi d\mu\right)+2\epsilon.

Finally, by considering the upper bound given by the definition of Bowen-Gibbs measure and using a similar argument, we obtain that

HF(μ|ν)|F|p(ϕ)+(h(μ)ϕdμ)2ϵ.\frac{H_{F}(\mu|\nu)}{|F|}\geq p(\phi)+\left(h(\mu)-\int\phi d\mu\right)-2\epsilon.

Since ϵ\epsilon was arbitrary, we conclude that

limFGHF(μ|ν)|F|=p(ϕ)(h(μ)+ϕdμ).\lim_{F\to G}\frac{H_{F}(\mu|\nu)}{|F|}=p(\phi)-\left(h(\mu)+\int\phi d\mu\right).

In particular, given ϕ:X\phi\colon X\to\mathbb{R} an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, a GG-invariant measure μ\mu is an equilibrium measure for ϕ\phi if and only if h(μ|ν)=0h(\mu\,|\,\nu)=0, for some (or every) DLR measure ν\nu. The next proposition is a generalization of Step 11 in the proof of [30, Theorem 15.37].

Proposition 5.16.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m} and μG(X)\mu\in\mathcal{M}_{G}(X) be an equilibrium measure for ϕ\phi. Then, for every α>0\alpha>0 and K(G)K\in\mathcal{F}(G), there exists E(G)E\in\mathcal{F}(G) such that KEK\subseteq E and

0HE(μ|ν)HEK(μ|ν)α.0\leq H_{E}(\mu|\nu)-H_{E\setminus K}(\mu|\nu)\leq\alpha.
Proof.

Pick δ>0\delta>0 small enough so that every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) satisfies IntK(F)\mathrm{Int}_{K}(F)\neq\emptyset. Consider 0<ϵ<10<\epsilon<1 and a tiling 𝒯\mathcal{T} with (K,δ)(K,\delta)-invariant shapes, which we can do by Theorem 3.4. Then, from Lemma 3.5, for every (S𝒯,ϵ)(S_{\mathcal{T}},\epsilon)-invariant set F(G)F\in\mathcal{F}(G), there exist center sets CF(S)C(S)𝒞(𝒯)C_{F}(S)\subseteq C(S)\in\mathcal{C}(\mathcal{T}) for S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}) such that

FS𝒮(𝒯)SCF(S)and|FS𝒮(𝒯)SCF(S)|ϵ|F|.F\supseteq\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\quad\text{and}\quad\left|F\setminus\bigsqcup_{S\in\mathcal{S}(\mathcal{T})}SC_{F}(S)\right|\leq\epsilon|F|.

Since μ\mu is an equilibrium measure, h(μ|ν)=0h(\mu\,|\nu)=0. Recall that S𝒯=S𝒮(𝒯)SS1S_{\mathcal{T}}=\bigcup_{S\in\mathcal{S}(\mathcal{T})}SS^{-1}. Then, considering Lemma 5.15, pick KS𝒯K^{\prime}\supseteq S_{\mathcal{T}} and δ<ϵ\delta^{\prime}<\epsilon so that, for every (K,δ)(K^{\prime},\delta^{\prime})-invariant set F(G)F\in\mathcal{F}(G), we have

1|F|HF(μ|ν)α(1ε)maxS𝒮(𝒯)|S|.\frac{1}{|F|}H_{F}(\mu|\nu)\leq\frac{\alpha(1-\varepsilon)}{\max_{S\in\mathcal{S}(\mathcal{T})}|S|}.

Fix a (K,δ)(K^{\prime},\delta^{\prime})-invariant set F(G)F\in\mathcal{F}(G) and an arbitrary enumeration of the tiles {Sc:S𝒮(𝒯),cCF(S)}\{Sc:S\in\mathcal{S}(\mathcal{T}),c\in C_{F}(S)\}, say T1,,TMT_{1},\dots,T_{M}, where M:=S𝒮(𝒯)|CF(S)|M:=\sum_{S\in\mathcal{S}(\mathcal{T})}|C_{F}(S)|. Notice that (1ϵ)|F|S𝒮(𝒯)|S||CF(S)|MmaxS𝒮(𝒯)|S|(1-\epsilon)|F|\leq\sum_{S\in\mathcal{S}(\mathcal{T})}|S||C_{F}(S)|\leq M\max_{S\in\mathcal{S}(\mathcal{T})}|S|. Moreover, since each TiT_{i} is a (K,δ)(K,\delta)-invariant set, for every 1iM1\leq i\leq M, IntK(Ti)\mathrm{Int}_{K}(T_{i})\neq\emptyset, i.e., there exists giGg_{i}\in G such that KgiTiKg_{i}\subseteq T_{i}. Denote W(i)=j=1iTjW(i)=\bigsqcup_{j=1}^{i}T_{j} for 0iM0\leq i\leq M. Then,

01Mi=1M(HW(i)(μ|ν)HW(i)Kgi(μ|ν))\displaystyle 0\leq\frac{1}{M}\sum_{i=1}^{M}\left(H_{W(i)}(\mu|\nu)-H_{W(i)\setminus Kg_{i}}(\mu|\nu)\right) 1Mi=1M(HW(i)(μ|ν)HW(i)Ti(μ|ν))\displaystyle\leq\frac{1}{M}\sum_{i=1}^{M}\left(H_{W(i)}(\mu|\nu)-H_{W(i)\setminus T_{i}}(\mu|\nu)\right)
=1MHW(M)(μ|ν)\displaystyle=\frac{1}{M}\,H_{W(M)}(\mu|\nu)
|F|M1|F|HF(μ|ν)\displaystyle\leq\frac{|F|}{M}\frac{1}{|F|}H_{F}(\mu|\nu)
MmaxS𝒮(𝒯)|S|M(1ϵ)α(1ε)maxS𝒮(𝒯)|S|\displaystyle\leq\frac{M\max_{S\in\mathcal{S}(\mathcal{T})}|S|}{M(1-\epsilon)}\frac{\alpha(1-\varepsilon)}{\max_{S\in\mathcal{S}(\mathcal{T})}|S|}
=α,\displaystyle=\alpha,

where the first and second inequality follow from Lemma 5.14 and, the first equality, from the fact that the sum is telescopic. Consequently, there must exist an index i{1,,M}i^{\prime}\in\{1,\dots,M\} such that

HW(i)(μ|ν)HW(i)Kgi(μ|ν)α.H_{W(i^{\prime})}(\mu|\nu)-H_{W(i^{\prime})\setminus Kg_{i^{\prime}}}(\mu|\nu)\leq\alpha.

Therefore, taking E=W(i)gi1E=W(i^{\prime})g_{i^{\prime}}^{-1}, the result follows from the GG-invariance of μ\mu and ν\nu. ∎

The next Lemma is a version of Step 2 in [30, Theorem 15.37].

Lemma 5.17.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation with respect to an exhausting sequence {Em}m\{E_{m}\}_{m} and μ(X)\mu\in\mathcal{M}(X) be an equilibrium measure for ϕ\phi. Then, for every ϵ>0\epsilon>0, there exists α>0\alpha>0 such that, if EKE\supseteq K and HE(μ|ν)HEK(μ|ν)αH_{E}(\mu|\nu)-H_{E\setminus K}(\mu|\nu)\leq\alpha, then ν(|fEfEK|)ϵ\nu\left(\left|f_{E}-f_{E\setminus K}\right|\right)\leq\epsilon.

Proof.

Notice that, for each ϵ>0\epsilon>0, there exists rϵ>0r_{\epsilon}>0 such that

(33) |x1|rϵψ(x)+ϵ2,|x-1|\leq r_{\epsilon}\psi(x)+\dfrac{\epsilon}{2},

where ψ(x)=1x+xlogx\psi(x)=1-x+x\log x.

For a given ϵ>0\epsilon>0, consider α=ϵ2rϵ\alpha=\frac{\epsilon}{2r_{\epsilon}}, and let E,K(G)E,K\in\mathcal{F}(G) be such that KEK\subseteq E and HE(μ|ν)HEK(μ|ν)αH_{E}(\mu|\nu)-H_{E\setminus K}(\mu|\nu)\leq\alpha, which we can do by Proposition 5.16. Let B={xX:fEK(x)0}B=\{x\in X:f_{E\setminus K}(x)\neq 0\}. Notice that BEKB\in\mathcal{B}_{E\setminus K}

𝟙XBfEdν=XBfEdν=XBν(fE|EK)dν=XBfEKdν=0.\int{\mathbbm{1}_{X\setminus B}f_{E}}d\nu=\int_{X\setminus B}f_{E}d\nu=\int_{X\setminus B}\nu(f_{E}|\mathcal{B}_{E\setminus K})d\nu=\int_{X\setminus B}f_{E\setminus K}d\nu=0.

Then, since fE(x)0f_{E}(x)\geq 0, we obtain that fE(x)=0f_{E}(x)=0 ν(x)\nu(x)-almost surely on XBX\setminus B. Next, notice that

BfElog(fEfEK)dν\displaystyle\int_{B}f_{E}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\nu =Blog(fEfEK)dμ\displaystyle=\int_{B}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\mu
=BlogfEdμBlogfEKdμ\displaystyle=\int_{B}\log f_{E}\,d\mu-\int_{B}\log f_{E\setminus K}\,d\mu
=BfElogfEdνBfEKlogfEKdν\displaystyle=\int_{B}f_{E}\log f_{E}d\nu-\int_{B}f_{E\setminus K}\log f_{E\setminus K}\,d\nu
=fElogfEdνfEKlogfEKdν\displaystyle=\int f_{E}\log f_{E}d\nu-\int f_{E\setminus K}\log f_{E\setminus K}\,d\nu
=HE(μ|ν)HEK(μ|ν),\displaystyle=H_{E}(\mu\,|\,\nu)-H_{E\setminus K}(\mu\,|\,\nu),

where, making an abuse of notation, we just write μ\mu and ν\nu, ignoring the restrictions. Thus,

HE(μ|ν)HEK(μ|ν)=BfElog(fEfEK)dν.H_{E}(\mu\,|\,\nu)-H_{E\setminus K}(\mu\,|\,\nu)=\int_{B}f_{E}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\nu.

Furthermore, in BB, observe that

ψ(fEfEK)=1fEfEK+fEfEKlog(fEfEK),\psi\left(\frac{f_{E}}{f_{E\setminus K}}\right)=1-\frac{f_{E}}{f_{E\setminus K}}+\frac{f_{E}}{f_{E\setminus K}}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right),

so that

fEKψ(fEfEK)=fEKfE+fElog(fEfEK).\displaystyle f_{E\setminus K}\psi\left(\frac{f_{E}}{f_{E\setminus K}}\right)=f_{E\setminus K}-f_{E}+f_{E}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right).

Therefore,

BfEKψ(fEfEK)dν\displaystyle\int_{B}f_{E\setminus K}\psi\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\nu =B(fEKfE)dν+BfElog(fEfEK)dν.\displaystyle=\int_{B}\left(f_{E\setminus K}-f_{E}\right)\,d\nu+\int_{B}f_{E}\log\left(\frac{f_{E}}{f_{E\setminus K}}\right)\,d\nu.

Since fEK=ν(fE|EK)f_{E\setminus K}=\nu(f_{E}\,|\,\mathcal{B}_{E\setminus K}), we have that B(fEKfE)dν=0\int_{B}\left(f_{E\setminus K}-f_{E}\right)\,d\nu=0, so that we can rewrite

HE(μ|ν)HEK(μ|ν)=BfEKψ(fEfEK)dν.H_{E}(\mu\,|\,\nu)-H_{E\setminus K}(\mu\,|\,\nu)=\int_{B}f_{E\setminus K}\psi\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\nu.

Therefore, from inequality (33), it follows that

ν(|fEfEK|)\displaystyle\nu(|f_{E}-f_{E\setminus K}|) =B|fEfEK|dν+XB|fEfEK|dν\displaystyle=\int_{B}|f_{E}-f_{E\setminus K}|d\nu+\int_{X\setminus B}|f_{E}-f_{E\setminus K}|d\nu
=B|fEfEK|dν\displaystyle=\int_{B}\left|f_{E}-f_{E\setminus K}\right|d\nu
=B|fEfEK1|fEKdν\displaystyle=\int_{B}\left|\frac{f_{E}}{f_{E\setminus K}}-1\right|f_{E\setminus K}d\nu
rϵBfEKψ(fEfEK)dν+ϵ2BfEKdν\displaystyle\leq r_{\epsilon}\int_{B}f_{E\setminus K}\psi\left(\frac{f_{E}}{f_{E\setminus K}}\right)d\nu+\frac{\epsilon}{2}\int_{B}f_{E\setminus K}d\nu
=rϵ(HE(μ|ν)HEK(μ|ν))+ϵ2Bdμ\displaystyle=r_{\epsilon}(H_{E}(\mu\,|\,\nu)-H_{E\setminus K}(\mu\,|\,\nu))+\frac{\epsilon}{2}\int_{B}d\mu
rϵα+ϵ2=ϵ.\displaystyle\leq r_{\epsilon}\alpha+\frac{\epsilon}{2}=\epsilon.

Theorem 5.18.

Let ϕ:X\phi\colon X\to\mathbb{R} be an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. If μG(X)\mu\in\mathcal{M}_{G}(X) is an equilibrium measure for ϕ\phi, then μ\mu is a DLR measure for ϕ\phi.

Proof.

Since μ\mu is an equilibrium measure, then h(μ|ν)=0h(\mu|\nu)=0. The strategy is to prove that, for every K(G)K\in\mathcal{F}(G), μγK=μ\mu\gamma_{K}=\mu, where γ\gamma is the Gibbsian specification defined by equation (7). Then, by Lemma 4.5, it will follow that μ\mu is a DLR measure for ϕ\phi.

Let h:Xh\colon X\to\mathbb{R} be a bounded local function and ϵ>0\epsilon>0. Since γ\gamma is a quasilocal specification (see Theorem 4.11), then γKh\gamma_{K}h is a bounded quasilocal Kc\mathcal{B}_{K^{c}}-measurable function. Thus, there exists a bounded local Kc\mathcal{B}_{K^{c}}-measurable function h~:X\tilde{h}\colon X\to\mathbb{R} such that γKhh~<ϵ\left\|\gamma_{K}h-\tilde{h}\right\|_{\infty}<\epsilon. Since h~\tilde{h} is a local potential, there exists B(G)B\in\mathcal{F}(G), BKB\supseteq K, such that h~\tilde{h} is a BK\mathcal{B}_{B\setminus K}-measurable. Also, since hh is local, we can assume, without loss of generality, that hh is B\mathcal{B}_{B}-measurable.

Consider α\alpha as in Lemma 5.17, that is, whenever EBE\supseteq B and HE(μ|ν)HEB(μ|ν)αH_{E}(\mu|\nu)-H_{E\setminus B}(\mu|\nu)\leq\alpha, then ν(|fEfEB|)ϵ\nu\left(\left|f_{E}-f_{E\setminus B}\right|\right)\leq\epsilon. Now, using Proposition 5.16, fix a set E(G)E\in\mathcal{F}(G) such that EBE\supseteq B and HE(μ|ν)HEB(μ|ν)αH_{E}(\mu|\nu)-H_{E\setminus B}(\mu|\nu)\leq\alpha. Therefore, by the monotonicity of the relative entropy, we obtain that HE(μ|ν)HEK(μ|ν)αH_{E}(\mu|\nu)-H_{E\setminus K}(\mu|\nu)\leq\alpha, so that ν(|fEfEK|)ϵ\nu\left(\left|f_{E}-f_{E\setminus K}\right|\right)\leq\epsilon.

We now compute |μγK(h)μ(h)|\left|\mu\gamma_{K}(h)-\mu(h)\right|. First observe that since h~\tilde{h} is BK\mathcal{B}_{B\setminus K}-measurable and BEB\subseteq E, then h~\tilde{h} is EK\mathcal{B}_{E\setminus K}-measurable. Therefore, recalling that μγK(h)=μ(γKh)\mu\gamma_{K}(h)=\mu(\gamma_{K}h),

|μγK(h)μ(h)|\displaystyle\left|\mu\gamma_{K}(h)-\mu(h)\right| |μ(γKh)μ(h~)|+|μ(h~)ν(fEKh~)|+|ν(fEKh~)ν(fEK(γKh))|\displaystyle\leq\left|\mu(\gamma_{K}h)-\mu(\tilde{h})\right|+\left|\mu(\tilde{h})-\nu(f_{E\setminus K}\tilde{h})\right|+\left|\nu(f_{E\setminus K}\tilde{h})-\nu(f_{E\setminus K}(\gamma_{K}h))\right|
+|ν(fEK(γKh))ν(fEKh)|+|ν(fEKh)ν(fEh)|+|ν(fEh)μ(h)|\displaystyle\quad+\left|\nu(f_{E\setminus K}(\gamma_{K}h))-\nu(f_{E\setminus K}h)\right|+\left|\nu(f_{E\setminus K}h)-\nu(f_{E}h)\right|+\left|\nu(f_{E}h)-\mu(h)\right|
μ(|γKhh~|)+0+ν(fEK|h~γKh|)+0+hν(|fEKfE|)+0.\displaystyle\leq\mu\left(\left|\gamma_{K}h-\tilde{h}\right|\right)+0+\nu\left(f_{E\setminus K}\left|\tilde{h}-\gamma_{K}h\right|\right)+0+\left\|h\right\|_{\infty}\nu\left(\left|f_{E\setminus K}-f_{E}\right|\right)+0.

We begin by justifying the terms that vanished from the first inequality to the second. Notice that |μ(h~)ν(fEKh~)|=0\left|\mu(\tilde{h})-\nu(f_{E\setminus K}\tilde{h})\right|=0 and |ν(fEh)μ(h)|=0\left|\nu(f_{E}h)-\mu(h)\right|=0, because h~\tilde{h} is EK\mathcal{B}_{E\setminus K}-measurable and because hh is E\mathcal{B}_{E}-measurable. We also have that |ν(fEK(γKh))ν(fEKh)|=0\left|\nu\left(f_{E\setminus K}(\gamma_{K}h)\right)-\nu\left(f_{E\setminus K}h\right)\right|=0, because fEKf_{E\setminus K} is Kc\mathcal{B}_{K^{c}}-measurable and γ\gamma is proper, so ν(fEK(γKh))=ν(γK(fEKh))\nu(f_{E\setminus K}(\gamma_{K}h))=\nu(\gamma_{K}(f_{E\setminus K}h)) and, in addition, since ν\nu is a DLR measure, we have that

ν(γK(fEKh))=(νγK)(fEKh)=ν(fEKh).\nu(\gamma_{K}(f_{E\setminus K}h))=(\nu\gamma_{K})(f_{E\setminus K}h)=\nu(f_{E\setminus K}h).

We now have to deal with the three other terms. Notice that

μ(|γKhh~|)<ϵ and ν(fEK|h~γKh|)<ϵ,\mu\left(\left|\gamma_{K}h-\tilde{h}\right|\right)<\epsilon\quad\text{ and }\quad\nu\left(f_{E\setminus K}\left|\tilde{h}-\gamma_{K}h\right|\right)<\epsilon,

because γKhh~<ϵ\left\|\gamma_{K}h-\tilde{h}\right\|_{\infty}<\epsilon. Lastly, since ν(|fEKfE|)ϵ\nu\left(\left|f_{E\setminus K}-f_{E}\right|\right)\leq\epsilon, it follows that

|μγK(h)μ(h)|<2ϵ+hϵ.\left|\mu\gamma_{K}(h)-\mu(h)\right|<2\epsilon+\left\|h\right\|_{\infty}\epsilon.

Since ϵ>0\epsilon>0 and h:Xh\colon X\to\mathbb{R} are arbitrary, we obtain that, μγK=μ\mu\gamma_{K}=\mu, which concludes the result. ∎

6. Final considerations

In this section we consider the case when the group is finitely generated, which includes the well-studied case G=dG=\mathbb{Z}^{d} and show that our approach generalizes previous ones. Next, we present a version of Dobrushin’s Uniqueness Theorem adapted to our framework and we apply it to a concrete class of examples of potentials defined in the GG-full shift for any countable amenable group GG.

6.1. The finitely generated case

We now restrict ourselves to the case that GG is a finitely generated group. The main goal is to prove that our definition of a Bowen-Gibbs measure (Definition 9) for a given exp-summable potential with summable variation according to an exhausting sequence is related to the standard — but more restrictive — way to define Bowen-Gibbs measures (e.g., [45, 38]). For that, we will prove that the bounds in Definition 9 can be replaced by a bound which involves the size of the boundary of invariant sets.

Suppose that GG is finitely generated and let SS be a finite and symmetric generating set. Without loss of generality, suppose that 1GS1_{G}\in S. In this context, it is common to implicitly consider an exhausting sequence Em+1=SmE_{m+1}=S^{m}. For example, if G=dG=\mathbb{Z}^{d} and SS is the set of all elements sds\in\mathbb{Z}^{d} with s1\|s\|_{\infty}\leq 1, the sequence {Em}m\{E_{m}\}_{m} recovers the notion of “boxes” with sides of length 2m+12m+1 centered at the origin, which is the most usual in the literature. In particular, one recovers the more standard definition of summable variation for a potential ϕ:X\phi\colon X\to\mathbb{R}, which is given by

m1|Em+11Em1|δEm(ϕ)=m0|Sm+1Sm|δSm(ϕ)=m0|B(1G,m)|δB(1G,m)(ϕ),\sum_{m\geq 1}|E_{m+1}^{-1}\setminus E_{m}^{-1}|\cdot\delta_{E_{m}}(\phi)=\sum_{m\geq 0}|S^{m+1}\setminus S^{m}|\cdot\delta_{S^{m}}(\phi)=\sum_{m\geq 0}|\partial B(1_{G},m)|\cdot\delta_{B(1_{G},m)}(\phi),

where B(1G,m)=SmB(1_{G},m)=S^{m} denotes the ball of radius mm (according to the word metric), F:=SFF\partial F:=SF\setminus F denotes the “(exterior) boundary” of a set FF, and |B(1G,m)||\partial B(1_{G},m)| is proportional to md1m^{d-1} in the d\mathbb{Z}^{d} case. Usually, potentials that have summable variation according to this particular exhausting sequence are called regular (see, for example, [38]).

Notice that when {Em}m\{E_{m}\}_{m} is an exhausting sequence of the form SmS^{m}, we have that

|(SmF)|=|S(SmF)SmF|=|Sm+1FSmF||Sm+1Sm||intF|,|\partial(S^{m}F)|=|S(S^{m}F)\setminus S^{m}F|=|S^{m+1}F\setminus S^{m}F|\leq|S^{m+1}\setminus S^{m}||\partial_{\rm int}F|,

where intF=Fc\partial_{\rm int}F=\partial F^{c} denotes the “interior boundary” of FF. Indeed, if gSm+1FSmFg\in S^{m+1}F\setminus S^{m}F, there must exist hintFh\in\partial_{\rm int}F such that dS(g,h)=m+1d_{S}(g,h)=m+1, where dSd_{S} denotes the word metric. In addition, we also have that |intF||S||F||\partial_{\rm int}F|\leq|S||\partial F|, so

|(SmF)|=|Sm+1FSmF||Sm+1Sm||S||F|.|\partial(S^{m}F)|=|S^{m+1}F\setminus S^{m}F|\leq|S^{m+1}\setminus S^{m}||S||\partial F|.

From this, it is direct that

VF(ϕ)=m0|Sm+1FSmF|δSm(ϕ)m0|Sm+1Sm||S||F|δSm(ϕ)=V(ϕ)|S||F|.\displaystyle V_{F}(\phi)=\sum_{m\geq 0}|S^{m+1}F\setminus S^{m}F|\cdot\delta_{S^{m}}(\phi)\leq\sum_{m\geq 0}|S^{m+1}\setminus S^{m}||S||\partial F|\cdot\delta_{S^{m}}(\phi)=V(\phi)|S||\partial F|.

On the other hand, if x,yXx,y\in X are such that xF=yFx_{F}=y_{F}, we have that

|ϕF(x)ϕF(y)|\displaystyle|\phi_{F}(x)-\phi_{F}(y)| gF|ϕ(gx)ϕ(gy)|\displaystyle\leq\sum_{g\in F}|\phi(g\cdot x)-\phi(g\cdot y)|
=m0gIntSm(F)IntSm+1(F)|ϕ(gx)ϕ(gy)|\displaystyle=\sum_{m\geq 0}\sum_{g\in\mathrm{Int}_{S^{m}}(F)\setminus\mathrm{Int}_{S^{m+1}}(F)}|\phi(g\cdot x)-\phi(g\cdot y)|
m0|IntSm(F)IntSm+1(F)|δSm(ϕ).\displaystyle\leq\sum_{m\geq 0}|\mathrm{Int}_{S^{m}}(F)\setminus\mathrm{Int}_{S^{m+1}}(F)|\cdot\delta_{S^{m}}(\phi).

Notice that if gIntSm(F)IntSm+1(F)g\in\mathrm{Int}_{S^{m}}(F)\setminus\mathrm{Int}_{S^{m+1}}(F), then dS(g,F)=m+1d_{S}(g,\partial F)=m+1, i.e., gSm+1FSmFg\in S^{m+1}\partial F\setminus S^{m}\partial F, so

|IntSm(F)IntSm+1(F)|\displaystyle|\mathrm{Int}_{S^{m}}(F)\setminus\mathrm{Int}_{S^{m+1}}(F)| |Sm+1FSmF|\displaystyle\leq|S^{m+1}\partial F\setminus S^{m}\partial F|
|Sm+1Sm||int(F)|\displaystyle\leq|S^{m+1}\setminus S^{m}||\partial_{\rm int}(\partial F)|
|Sm+1Sm||S||(F)|\displaystyle\leq|S^{m+1}\setminus S^{m}||S||\partial(\partial F)|
|Sm+1Sm||S|2|F|\displaystyle\leq|S^{m+1}\setminus S^{m}||S|^{2}|\partial F|

and

|ϕF(x)ϕF(y)|m0|Sm+1Sm||S||F|δSm(ϕ)=V(ϕ)|S|2|F|.\displaystyle|\phi_{F}(x)-\phi_{F}(y)|\leq\sum_{m\geq 0}|S^{m+1}\setminus S^{m}||S||\partial F|\cdot\delta_{S^{m}}(\phi)=V(\phi)|S|^{2}|\partial F|.

Therefore, we conclude that ΔF(ϕ)V(ϕ)|S|2|F|\Delta_{F}(\phi)\leq V(\phi)|S|^{2}|\partial F|.

We now provide an alternative way of proving Proposition 2.2 and Lemma 2.3. Begin by noticing that a finitely generated group is amenable if and only if limFG|F||F|=0\lim_{F\to G}\frac{|\partial F|}{|F|}=0 (indeed, given ϵ>0\epsilon>0, we have that |F||SFF|<ϵ|F||\partial F|\leq|SF\triangle F|<\epsilon\cdot|F| for every (S,ϵ)(S,\epsilon)-invariant set FF). Therefore, if ϕ\phi has summable variation, it follows that

0limFGVF(ϕ)|F|V(ϕ)|S|limFG|F||F|=00\leq\lim_{F\to G}\frac{V_{F}(\phi)}{|F|}\leq V(\phi)|S|\lim_{F\to G}\frac{|\partial F|}{|F|}=0

and, similarly,

0limFGΔF(ϕ)|F|V(ϕ)|S|2limFG|F||F|=0.0\leq\lim_{F\to G}\frac{\Delta_{F}(\phi)}{|F|}\leq V(\phi)|S|^{2}\lim_{F\to G}\frac{|\partial F|}{|F|}=0.

In particular, in this context, we could alternatively have defined a Bowen-Gibbs measure as follows: if GG is a finitely generated amenable group with generating set SS and ϕ:X\phi\colon X\to\mathbb{R} is an exp-summable potential with summable variation according to {Sm}m\{S^{m}\}_{m}, a measure μ(X)\mu\in\mathcal{M}(X) is a Bowen-Gibbs measure for ϕ\phi if for every ϵ>0\epsilon>0, there exist K(G)K\in\mathcal{F}(G) and δ>0\delta>0 such that for every (K,δ)(K,\delta)-invariant set F(G)F\in\mathcal{F}(G) and xXx\in X,

exp(C|F|)μ([xF])exp(ϕF(x)p(ϕ)|F|)exp(C|F|),\mathop{\textrm{\rm exp}}\nolimits\left(-C|\partial F|\right)\leq\frac{\mu([x_{F}])}{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{F}(x)-p(\phi)\cdot|F|\right)}\leq\mathop{\textrm{\rm exp}}\nolimits\left(C|\partial F|\right),

where C>0C>0 is a constant that we can choose to be

C:=5V(ϕ)|S|22V(ϕ)|S|+3Δ(ϕ)|S|2.C:=5V(\phi)|S|^{2}\geq 2V(\phi)|S|+3\Delta(\phi)|S|^{2}.

This recovers the more standard definition of Bowen-Gibbs measure in terms of boundaries. Furthermore, with this choice of CC, it is not difficult to check that we could mimic the proofs of Proposition 5.7, Theorem 5.8, and Theorem 5.13, thus providing all the implications involving Bowen-Gibbs measures.

6.2. Dobrushin’s Uniqueness Theorem

From §5.4, we know that if ϕ:X\phi\colon X\to\mathbb{R} is an exp-summable potential with summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}, then the set of GG-invariant DLR measures for ϕ\phi is non-empty. One natural question that may arise is under which conditions we have uniqueness of the DLR measure. When a specification is a Gibbsian specification, the Dobrushin’s Uniqueness Theorem (see [30]) addresses this question. For a detailed proof of a version of this theorem adapted to our setting, see [11].

Let 22^{\mathbb{N}} be the set of all subsets of \mathbb{N}, which is a σ\sigma-algebra, and (,2)\mathcal{M}(\mathbb{N},2^{\mathbb{N}}) be the set of probability measures on (,2)(\mathbb{N},2^{\mathbb{N}}). For A2A\in 2^{\mathbb{N}}, wXw\in X, and gGg\in G, denote

γ0{g}(A,w)(η)=γ{g}(A×G{g},x),\gamma^{0}_{\{g\}}(A,w)(\eta)=\gamma_{\{g\}}\left(A\times\mathbb{N}^{G\setminus\{g\}},x\right),

where γ\gamma is a specification, notice that, for each xXx\in X, γg0(,x)(,2)\gamma_{g}^{0}(\cdot,x)\in\mathcal{M}(\mathbb{N},2^{\mathbb{N}}). Now, for each hGh\in G, the whw_{h}-dependence of γ0{g}(,w)\gamma^{0}_{\{g\}}(\cdot,w) is estimated by the quantity

ρgh(γ)=supw,ηXwG{h}=ηG{h}γ{g}0(,η)γ{g}0(,w),\rho_{gh}(\gamma)=\sup_{\begin{subarray}{c}w,\eta\in X\\ w_{G\setminus\{h\}}=\eta_{G\setminus\{h\}}\end{subarray}}\left\|\gamma_{\{g\}}^{0}(\cdot,\eta)-\gamma_{\{g\}}^{0}(\cdot,w)\right\|,

where, for any given μ,μ~(,2)\mu,\tilde{\mu}\in\mathcal{M}(\mathbb{N},2^{\mathbb{N}}), μμ~=maxA|μ(A)μ~(A)|\|\mu-\tilde{\mu}\|=\max_{A\in\mathcal{E}}|\mu(A)-\tilde{\mu}(A)| (see [30, §8.1]).

The infinite matrix ρ(γ)=(ρgh(γ))g,hG\rho(\gamma)=(\rho_{gh}(\gamma))_{g,h\in G} is called Dobrushin’s interdependence matrix for γ\gamma. When there is no ambiguity, we will omit the parameter γ\gamma from the notation.

Remark 8.

Notice that ρgg=0\rho_{gg}=0, for all gGg\in G.

Definition 12.

Let γ\gamma be a specification. We say that γ\gamma satisfies the Dobrushin’s condition if γ\gamma is quasilocal and

c(γ):=supgGhGρgh<1.c(\gamma):=\sup_{g\in G}\sum_{h\in G}\rho_{gh}<1.
Theorem 6.1 (Dobrushin’s Uniqueness Theorem).

If γ\gamma is a specification that satisfies the Dobrushin’s condition, then there is at most one measure that is admitted by the specification γ\gamma.

We now present an example of a potential inspired by the Potts model [29, 27] such that, under some conditions to be presented, is exp-summable and has summable variation according to an exhausting sequence {Em}m\{E_{m}\}_{m}. Moreover, this potential will also satisfy that, if μ\mu is a Bowen-Gibbs measure, ϕdμ>\int\phi d\mu>-\infty. Another important property of this potential is that it is non-trivial, in the sense that it depends on every coordinate of GG. We will also explore conditions on β>0\beta>0 such that the potential βϕ\beta\phi satisfies Dobrushin’s condition.

6.3. Main example

Given a countable amenable group GG, consider the potential ϕ:X\phi\colon X\to\mathbb{R} given by

(34) ϕ(x):=gGc(g,x(1G))𝟙{x(1G)=x(g)},\phi(x):=-\sum_{g\in G}c(g,x(1_{G}))\mathbbm{1}_{\{x(1_{G})=x(g)\}},

with c:G×[0,)c\colon G\times\mathbb{N}\to[0,\infty) such that, given an exhausting sequence {Em}m\{E_{m}\}_{m} of GG, it holds that

  1. (1)

    m1|Em+1Em|gGEmC(g)<\sum_{m\geq 1}|E_{m+1}\setminus E_{m}|\sum_{g\in G\setminus E_{m}}C(g)<\infty, with C(g):=supnc(g,n)C(g):=\sup_{n}c(g,n) for g1Gg\neq 1_{G}; and

  2. (2)

    for all M>0M>0, there exists n0n_{0}\in\mathbb{N} such that for all nn0n\geq n_{0}, Mlog(n)c(1G,n)M\log(n)\leq c(1_{G},n).

Lemma 6.2.

If the potential ϕ:X\phi\colon X\to\mathbb{R} given by ϕ(x)=gGc(g,x(1G))𝟙{x(1G)=x(g)}\phi(x)=-\sum_{g\in G}c(g,x(1_{G}))\mathbbm{1}_{\left\{x(1_{G})=x(g)\right\}} satisfies conditions (1) and (2), then, for every β>0\beta>0, the potential βϕ\beta\phi is well-defined, has summable variation according to the exhausting sequence {Em}m\{E_{m}\}_{m}, is exp-summable, and ϕdμβ>\int\phi d\mu_{\beta}>-\infty for any Bowen-Gibbs measure μβ(X)\mu_{\beta}\in\mathcal{M}(X) for βϕ\beta\phi.

Proof.

Notice that condition (1) implies that

0gGg1GC(g)=m1gEm+1EmC(g)m1|Em+1Em|gGEmC(g)<.0\leq\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}C(g)=\sum_{m\geq 1}\sum_{g\in E_{m+1}\setminus E_{m}}C(g)\leq\sum_{m\geq 1}|E_{m+1}\setminus E_{m}|\sum_{g\in G\setminus E_{m}}C(g)<\infty.

Now, for any xXx\in X,

|ϕ(x)|\displaystyle|\phi(x)| =c(1G,x(1G))+gGg1Gc(g,x(1G))c(1G,x(1G))+gGg1GC(g)<,\displaystyle=c(1_{G},x(1_{G}))+\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}c(g,x(1_{G}))\leq c(1_{G},x(1_{G}))+\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}C(g)<\infty,

so ϕ(x)\phi(x) is well-defined and therefore βϕ\beta\phi is well-defined, too.

Next, notice that, for every mm\in\mathbb{N} and x,yXx,y\in X such that xEm=yEmx_{E_{m}}=y_{E_{m}}, we have that

|ϕ(x)ϕ(y)|\displaystyle|\phi(x)-\phi(y)| =|gGc(g,x(1G))(𝟙{y(1G)=y(g)}𝟙{x(1G)=x(g)})|\displaystyle=\left|\sum_{g\in G}c(g,x(1_{G}))\left(\mathbbm{1}_{\left\{y(1_{G})=y(g)\right\}}-\mathbbm{1}_{\left\{x(1_{G})=x(g)\right\}}\right)\right|
gGEmc(g,x(1G))|𝟙{y(1G)=y(g)}𝟙{x(1G)=x(g)}|\displaystyle\leq\sum_{g\in G\setminus E_{m}}c(g,x(1_{G}))\left|\mathbbm{1}_{\left\{y(1_{G})=y(g)\right\}}-\mathbbm{1}_{\left\{x(1_{G})=x(g)\right\}}\right|
gGEmC(g).\displaystyle\leq\sum_{g\in G\setminus E_{m}}C(g).

Therefore, for any mm\in\mathbb{N}, δEm(ϕ)gGEmC(g)\delta_{E_{m}}(\phi)\leq\sum_{g\in G\setminus E_{m}}C(g), so that

V(ϕ)=m=1|Em+11Em1|δEm(ϕ)m=1|Em+11Em1|gGEmC(g).\displaystyle V(\phi)=\sum_{m=1}^{\infty}|E_{m+1}^{-1}\setminus E_{m}^{-1}|\cdot\delta_{E_{m}}(\phi)\leq\sum_{m=1}^{\infty}|E_{m+1}^{-1}\setminus E_{m}^{-1}|\sum_{g\in G\setminus E_{m}}C(g).

Thus, due to condition (1), we have that ϕ\phi has summable variation according to {Em}m\{E_{m}\}_{m}. In addition, observe that V(βϕ)=βV(ϕ)V(\beta\phi)=\beta V(\phi), so βϕ\beta\phi has also summable variation.

Pick 0<α=β2<β0<\alpha=\frac{\beta}{2}<\beta. This determines n1n_{1} such that

n1βc(1G,n)exp(βc(1G,n))\displaystyle\sum_{n\geq 1}\beta c(1_{G},n)\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n)) =n<n1βc(1G,n)exp(βc(1G,n))\displaystyle=\sum_{n<n_{1}}\beta c(1_{G},n)\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))
+nn1βc(1G,n)exp(βc(1G,n))\displaystyle\quad+\sum_{n\geq n_{1}}\beta c(1_{G},n)\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))
=C0+βnn1exp(αc(1G,n))exp(βc(1G,n))\displaystyle=C_{0}+\beta\sum_{n\geq n_{1}}\mathop{\textrm{\rm exp}}\nolimits(\alpha c(1_{G},n))\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))
=C0+βnn1exp(β2c(1G,n)),\displaystyle=C_{0}+\beta\sum_{n\geq n_{1}}\mathop{\textrm{\rm exp}}\nolimits\left(-\frac{\beta}{2}c(1_{G},n)\right),

where

C0:=n<n1βc(1G,n)exp(βc(1G,n))<.C_{0}:=\sum_{n<n_{1}}\beta c(1_{G},n)\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))<\infty.

It remains to bound nn1exp(β2c(1G,n))\sum_{n\geq n_{1}}\mathop{\textrm{\rm exp}}\nolimits\left(-\frac{\beta}{2}c(1_{G},n)\right). Now, for any ϵ>0\epsilon>0 and M=2β(1+ϵ)>0M=\frac{2}{\beta}(1+\epsilon)>0, condition (2) implies that there exists n0n_{0}\in\mathbb{N} (maybe n0>n1n_{0}>n_{1}) such that nn0\forall n\geq n_{0}, Mlog(n)c(1G,n)M\log(n)\leq c(1_{G},n), so 2β(1+ϵ)log(n)c(1G,n)-\frac{2}{\beta}(1+\epsilon)\log(n)\geq-c(1_{G},n). Therefore,

nn1exp(β2c(1G,n))\displaystyle\sum_{n\geq n_{1}}\mathop{\textrm{\rm exp}}\nolimits\left(-\frac{\beta}{2}c(1_{G},n)\right) =n1n<n0exp(β2c(1G,n))+nn0exp(β2c(1G,n))\displaystyle=\sum_{n_{1}\leq n<n_{0}}\mathop{\textrm{\rm exp}}\nolimits\left(-\frac{\beta}{2}c(1_{G},n)\right)+\sum_{n\geq n_{0}}\mathop{\textrm{\rm exp}}\nolimits\left(-\frac{\beta}{2}c(1_{G},n)\right)
C1+nn0exp((1+ϵ)log(n))\displaystyle\leq C_{1}+\sum_{n\geq n_{0}}\mathop{\textrm{\rm exp}}\nolimits\left((1+\epsilon)\log(n)\right)
=C1+nn01n1+ϵ<,\displaystyle=C_{1}+\sum_{n\geq n_{0}}\frac{1}{n^{1+\epsilon}}<\infty,

where C1:=n1n<n0exp(β2c(1G,n))<C_{1}:=\sum_{n_{1}\leq n<n_{0}}\mathop{\textrm{\rm exp}}\nolimits(-\frac{\beta}{2}c(1_{G},n))<\infty. So, from Proposition 5.12 we have that ϕdμβ>\int\phi d\mu_{\beta}>-\infty for every β>0\beta>0.

Later, choosing a n3n_{3}\in\mathbb{N}, great enough

Z1G(βϕ)\displaystyle Z_{1_{G}}(\beta\phi) =n1exp(βc(1G,n))\displaystyle=\sum_{n\geq 1}\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))
=n<n3exp(βc(1G,n))+nn3c(1G,n)exp(βc(1G,n))<.\displaystyle=\sum_{n<n_{3}}\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))+\sum_{n\geq n_{3}}c(1_{G},n)\mathop{\textrm{\rm exp}}\nolimits(-\beta c(1_{G},n))<\infty.

Therefore, the potential βϕ\beta\phi is exp-summable, for all β>0\beta>0. ∎

Remark 9.

The set of functions c:G×[0,)c\colon G\times\mathbb{N}\to[0,\infty) satisfying conditions (1) and (2) is non-vacuous. For example, given an exhausting sequence {Em}m\{E_{m}\}_{m}, consider c:G×[0,)c\colon G\times\mathbb{N}\to[0,\infty) and some constant L0L\geq 0 such that

  • (a)

    for every m1m\geq 1, 0c(g,n)L2m1|Em+1|20\leq c(g,n)\leq\frac{L2^{-m-1}}{|E_{m+1}|^{2}} for every gEm+1Emg\in E_{m+1}\setminus E_{m}; and

  • (b)

    any c(1G,n)c(1_{G},n) of polynomial order will satisfy condition (2).

We now prove that any such cc satisfies the previous conditions (1) and (2). Due to (a) and the fact that {Em}m\{E_{m}\}_{m} is nested, for every m1m\geq 1,

gGEmC(g)=mgE+1EL21|E+1|2=mL21|E+1|L2m|Em+1|<.\sum_{g\in G\setminus E_{m}}C(g)\leq\sum_{\ell=m}^{\infty}\sum_{g\in E_{\ell+1}\setminus E_{\ell}}\frac{L2^{-\ell-1}}{|E_{\ell+1}|^{2}}\leq\sum_{\ell=m}^{\infty}\frac{L2^{-\ell-1}}{|E_{\ell+1}|}\leq\frac{L2^{-m}}{|E_{m+1}|}<\infty.

Then,

gGg1GC(g)=gGE1C(g)L2|E2|<\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}C(g)=\sum_{g\in G\setminus E_{1}}C(g)\leq\frac{L}{2|E_{2}|}<\infty

and

m=1|Em+1Em|gGEmC(g)m=1|Em+1Em|L2m|Em+1|m=1L2m=L<,\sum_{m=1}^{\infty}|E_{m+1}\setminus E_{m}|\sum_{g\in G\setminus E_{m}}C(g)\leq\sum_{m=1}^{\infty}|E_{m+1}\setminus E_{m}|\frac{L2^{-m}}{|E_{m+1}|}\leq\sum_{m=1}^{\infty}L2^{-m}=L<\infty,

so condition (1) is satisfied. Now, due to (b), we have that condition (2) is satisfied.

Our next goal is to study under which conditions we have uniqueness of Gibbs measures for βϕ\beta\phi, where β\beta can be interpreted as the inverse of the temperature of the system. For that, we use the Dobrushin’s Uniqueness Theorem (Theorem 6.1). In order to obtain explicit conditions on β\beta, we divide the rational into claims.

Claim 1.

If x,yXx,y\in X are such that xG{g}=yG{g}x_{G\setminus\{g\}}=y_{G\setminus\{g\}}, for some gGg\in G, then

hG(ϕ(hx)ϕ(hy))\sum_{h\in G}\left(\phi(h\cdot x)-\phi(h\cdot y)\right)

converges absolutely. Moreover,

hG(ϕ(hx)ϕ(hy))=\displaystyle\sum_{h\in G}\left(\phi(h\cdot x)-\phi(h\cdot y)\right)= c(1G,x(g))+c(1G,y(g))\displaystyle-c(1_{G},x(g))+c(1_{G},y(g))
+hGh1G(c(h,x(hg))+c(h1,x(hg)))(𝟙{xhg=xg}+𝟙{yhg=yg}).\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left(c(h,x(hg))+c(h^{-1},x(hg))\right)\left(-\mathbbm{1}_{\{x_{hg}=x_{g}\}}+\mathbbm{1}_{\{y_{hg}=y_{g}\}}\right).
Proof of Claim 1..

Since we are summing over all hh-translations of xx and yy, for hGh\in G, we can assume, without loss of generality, that g=1Gg=1_{G}, that is, xG{1G}=yG{1G}x_{G\setminus\{1_{G}\}}=y_{G\setminus\{1_{G}\}}. Then,

hG|ϕ(hx)ϕ(hy)|\displaystyle\sum_{h\in G}\left|\phi(h\cdot x)-\phi(h\cdot y)\right| =hG|gGc(g,x(h))𝟙{xh=xgh}+gGc(g,y(h))𝟙{yh=ygh}|\displaystyle=\sum_{h\in G}\left|-\sum_{g\in G}c(g,x(h))\mathbbm{1}_{\{x_{h}=x_{gh}\}}+\sum_{g\in G}c(g,y(h))\mathbbm{1}_{\{y_{h}=y_{gh}\}}\right|
=|gGc(g,x(1G))𝟙{x1G=xg}+c(g,y(1G))𝟙{y1G=yg}|\displaystyle=\left|\sum_{g\in G}-c(g,x(1_{G}))\mathbbm{1}_{\{x_{1_{G}}=x_{g}\}}+c(g,y(1_{G}))\mathbbm{1}_{\{y_{1_{G}}=y_{g}\}}\right|
+hGh1G|gG(c(g,x(h))𝟙{x(h)=x(gh)}+c(g,y(h))𝟙{yh=ygh})|\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|\sum_{g\in G}\left(-c(g,x(h))\mathbbm{1}_{\{x(h)=x(gh)\}}+c(g,y(h))\mathbbm{1}_{\{y_{h}=y_{gh}\}}\right)\right|
=|gGc(g,x(1G))𝟙{x1G=xg}+c(g,y(1G))𝟙{y1G=yg}|\displaystyle=\left|\sum_{g\in G}-c(g,x(1_{G}))\mathbbm{1}_{\{x_{1_{G}}=x_{g}\}}+c(g,y(1_{G}))\mathbbm{1}_{\{y_{1_{G}}=y_{g}\}}\right|
+hGh1G|gGc(g,x(h))(𝟙{xh=xgh}𝟙{yh=ygh})|\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|\sum_{g\in G}-c(g,x(h))\left(\mathbbm{1}_{\{x_{h}=x_{gh}\}}-\mathbbm{1}_{\{y_{h}=y_{gh}\}}\right)\right|
|gGc(g,x(1G))𝟙{x1G=xg}+c(g,y(1G))𝟙{y1G=yg}|\displaystyle\leq\left|\sum_{g\in G}-c(g,x(1_{G}))\mathbbm{1}_{\{x_{1_{G}}=x_{g}\}}+c(g,y(1_{G}))\mathbbm{1}_{\{y_{1_{G}}=y_{g}\}}\right|
+hGh1G|gGgh1c(g,x(h))(𝟙{xh=xgh}𝟙{yh=ygh})|\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|\sum_{\begin{subarray}{c}g\in G\\ g\neq h^{-1}\end{subarray}}-c(g,x(h))\left(\mathbbm{1}_{\{x_{h}=x_{gh}\}}-\mathbbm{1}_{\{y_{h}=y_{gh}\}}\right)\right|
+hGh1G|c(h1,x(h))(𝟙{xh=x1G}𝟙{yh=y1G})|.\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|-c(h^{-1},x(h))\left(\mathbbm{1}_{\{x_{h}=x_{1_{G}}\}}-\mathbbm{1}_{\{y_{h}=y_{1_{G}}\}}\right)\right|.

Note that if h1Gh\neq 1_{G} and gh1g\neq h^{-1}, we have that x(h)=y(h)x(h)=y(h) and x(gh)=y(gh)x(gh)=y(gh), so that x(gh)=y(gh)y(h)=y(gh)x(gh)=y(gh)\iff y(h)=y(gh). Then, 𝟙{x(h)=x(gh)}𝟙{y(h)=y(gh)}=0\mathbbm{1}_{\{x(h)=x(gh)\}}-\mathbbm{1}_{\{y(h)=y(gh)\}}=0 and

hG|ϕ(hx)ϕ(hy)|\displaystyle\sum_{h\in G}\left|\phi(h\cdot x)-\phi(h\cdot y)\right| |gG(c(g,x(1G))𝟙{x(1G)=x(g)}+c(g,y(1G))𝟙{y(1G)=y(g)})|\displaystyle\leq\left|\sum_{g\in G}\left(-c(g,x(1_{G}))\mathbbm{1}_{\{x(1_{G})=x(g)\}}+c(g,y(1_{G}))\mathbbm{1}_{\{y(1_{G})=y(g)\}}\right)\right|
+hGh1G|c(h1,x(h))(𝟙{x(h)=x(1G)}𝟙{y(h)=y1G})|\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|-c(h^{-1},x(h))\left(\mathbbm{1}_{\{x(h)=x(1_{G})\}}-\mathbbm{1}_{\{y(h)=y_{1_{G}}\}}\right)\right|
c(1G,x(1G))+c(1G,y(1G))\displaystyle\leq c(1_{G},x(1_{G}))+c(1_{G},y(1_{G}))
+gGg1G|c(g,x(1G))𝟙{x(1G)=x(g)}+c(g,y(1G))𝟙{y(1G)=y(g)}|\displaystyle\quad+\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}\left|-c(g,x(1_{G}))\mathbbm{1}_{\{x(1_{G})=x(g)\}}+c(g,y(1_{G}))\mathbbm{1}_{\{y(1_{G})=y(g)\}}\right|
+hGh1G|c(h1,x(h))(𝟙{x(h)=x(1G)}𝟙{y(h)=y(1G)})|\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left|-c(h^{-1},x(h))\left(\mathbbm{1}_{\{x(h)=x(1_{G})\}}-\mathbbm{1}_{\{y(h)=y(1_{G})\}}\right)\right|
c(1G,x(1G))+c(1G,y(1G))\displaystyle\leq c(1_{G},x(1_{G}))+c(1_{G},y(1_{G}))
+gGg1G(c(g,x(1G))+c(g,y(1G))+c(g1,x(g))).\displaystyle\quad+\sum_{\begin{subarray}{c}g\in G\\ g\neq 1_{G}\end{subarray}}\left(c(g,x(1_{G}))+c(g,y(1_{G}))+c(g^{-1},x(g))\right).

Therefore,

(35) hG|ϕ(hx)ϕ(hy)|3hGh1GC(h)+c(1G,x(1G))+c(1G,y(1G)),\sum_{h\in G}\left|\phi(h\cdot x)-\phi(h\cdot y)\right|\leq 3\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}C(h)+c(1_{G},x(1_{G}))+c(1_{G},y(1_{G})),

from where it follows that, due to condition (1),

hG|ϕ(hx)ϕ(hy)|<.\sum_{h\in G}\left|\phi(h\cdot x)-\phi(h\cdot y)\right|<\infty.

Moreover, notice that, applying the same rational with no absolute values, we get that

hG(ϕ(hx)ϕ(hy))\displaystyle\sum_{h\in G}\left(\phi(h\cdot x)-\phi(h\cdot y)\right)
=c(1G,x(1G))+c(1G,y(1G))\displaystyle=-c(1_{G},x(1_{G}))+c(1_{G},y(1_{G}))
+hGh1G[𝟙{xh=x1G}(c(h,x(1G))c(h1,x(h)))+𝟙{yh=y1G}(c(h,y(1G))+c(h1,x(h)))]\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left[\mathbbm{1}_{\{x_{h}=x_{1_{G}}\}}\left(-c(h,x(1_{G}))-c(h^{-1},x(h))\right)+\mathbbm{1}_{\{y_{h}=y_{1_{G}}\}}\left(c(h,y(1_{G}))+c(h^{-1},x(h))\right)\right]
=c(1G,x(1G))+c(1G,y(1G))\displaystyle=-c(1_{G},x(1_{G}))+c(1_{G},y(1_{G}))
+hGh1G(c(h,x(h))+c(h1,x(h)))(𝟙{xh=x1G}+𝟙{yh=y1G}).\displaystyle\quad+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left(c(h,x(h))+c(h^{-1},x(h))\right)\left(-\mathbbm{1}_{\{x_{h}=x_{1_{G}}\}}+\mathbbm{1}_{\{y_{h}=y_{1_{G}}\}}\right).

Now, for a fixed bb\in\mathbb{N}, define, for each gGg\in G and zXz\in X the potential φgz:\varphi^{g}_{z}\colon\mathbb{N}\to\mathbb{R} given by

(36) φgz(a)=ϕτa,b(bzG{g}).\varphi^{g}_{z}(a)=\phi_{*}^{\tau_{a,b}}(bz_{G\setminus\{g\}}).

Notice that, from Claim 1,

φgz(a)\displaystyle\varphi^{g}_{z}(a) =ϕτa,b(bzG{g})\displaystyle=\phi_{*}^{\tau_{a,b}}(bz_{G\setminus\{g\}})
=hG[ϕ(h(azG{g}))ϕ(h(bzG{g}))]\displaystyle=\sum_{h\in G}\left[\phi(h\cdot(az_{G\setminus\{g\}}))-\phi(h\cdot(bz_{G\setminus\{g\}}))\right]
(37) =c(1G,b)c(1G,a)+hGh1G[(c(h,zhg)+c(h1,zhg))(𝟙{zhg=b}𝟙{zhg=a})].\displaystyle=c(1_{G},b)-c(1_{G},a)+\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left[(c(h,z_{hg})+c(h^{-1},z_{hg}))(\mathbbm{1}_{\{z_{hg}=b\}}-\mathbbm{1}_{\{z_{hg}=a\}})\right].

Now, pick h0gh_{0}\neq g and z,zXz,z^{\prime}\in X such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}} and define the function φgz,z:×[0,1]\varphi^{g}_{z,z^{\prime}}\colon\mathbb{N}\times[0,1]\to\mathbb{R} given by

φgz,z(a,t):=tφgz(a)+(1t)φgz(a)=φgz(a)+tΔgz,z(a),\varphi^{g}_{z,z^{\prime}}(a,t):=t\varphi^{g}_{z^{\prime}}(a)+(1-t)\varphi^{g}_{z}(a)=\varphi^{g}_{z}(a)+t\Delta^{g}_{z,z^{\prime}}(a),

with Δgz,z(a)=φgz(a)φgz(a)\Delta^{g}_{z,z^{\prime}}(a)=\varphi^{g}_{z^{\prime}}(a)-\varphi^{g}_{z}(a). Notice that φgz,z(a,0)=φgz(a)\varphi^{g}_{z,z^{\prime}}(a,0)=\varphi^{g}_{z}(a) and φgz,z(a,1)=φgz(a)\varphi^{g}_{z,z^{\prime}}(a,1)=\varphi^{g}_{z^{\prime}}(a).

Claim 2.

Let gGg\in G. Then, for every h0gh_{0}\neq g and z,zXz,z^{\prime}\in X such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}, it holds that

Δgz,z2(C(h0g1)+C(gh01)).\|\Delta^{g}_{z,z^{\prime}}\|_{\infty}\leq 2(C(h_{0}g^{-1})+C(gh_{0}^{-1})).
Proof of Claim 2.

Let g,h0Gg,h_{0}\in G, with h0gh_{0}\neq g, and z,zXz,z^{\prime}\in X be such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}. Then, due to the equation (6.3), we have that, for all aa\in\mathbb{N},

Δgz,z(a)\displaystyle\Delta^{g}_{z,z^{\prime}}(a) =hGh1G[(c(h,z(hg))+c(h1,z(hg)))(𝟙{zhg=a}+𝟙{zhg=b})]\displaystyle=\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left[(c(h,z^{\prime}(hg))+c(h^{-1},z^{\prime}(hg)))(-\mathbbm{1}_{\{z^{\prime}_{hg}=a\}}+\mathbbm{1}_{\{z^{\prime}_{hg}=b\}})\right]
hGh1G[(c(h,z(hg))+c(h1,z(hg)))(𝟙{zhg=a}+𝟙{zhg=b})].\displaystyle\quad-\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left[(c(h,z(hg))+c(h^{-1},z(hg)))(-\mathbbm{1}_{\{z_{hg}=a\}}+\mathbbm{1}_{\{z_{hg}=b\}})\right].

Notice that, if hgh0hg\neq h_{0}, then zhg=zhgz_{hg}=z^{\prime}_{hg}, so that

(c(h,z(hg))+c(h1,z(hg)))(𝟙{zhg=a}+𝟙{zhg=b})\displaystyle(c(h,z^{\prime}(hg))+c(h^{-1},z^{\prime}(hg)))(-\mathbbm{1}_{\{z^{\prime}_{hg}=a\}}+\mathbbm{1}_{\{z^{\prime}_{hg}=b\}})
=(c(h,z(hg))+c(h1,z(hg)))(𝟙{zhg=a}+𝟙{zhg=b}).\displaystyle\quad=(c(h,z(hg))+c(h^{-1},z(hg)))(-\mathbbm{1}_{\{z_{hg}=a\}}+\mathbbm{1}_{\{z_{hg}=b\}}).

This means that the only terms that remain in the sums above are the ones such that hg=h0hg=h_{0}, that is, h=h0g1h=h_{0}g^{-1}. Thus,

Δgz,z(a)\displaystyle\Delta^{g}_{z,z^{\prime}}(a) =[(c(h0g1,zh0)+c((h0g1)1,zh0))(𝟙{zh0=a}+𝟙{zh0=b})]\displaystyle=\left[(c(h_{0}g^{-1},z^{\prime}_{h_{0}})+c((h_{0}g^{-1})^{-1},z^{\prime}_{h_{0}}))(-\mathbbm{1}_{\{z^{\prime}_{h_{0}}=a\}}+\mathbbm{1}_{\{z^{\prime}_{h_{0}}=b\}})\right]
[(c(h0g1,zh0)+c((h0g1)1,z(h0)))(𝟙{z(h0)=a}+𝟙{z(h0)=b})].\displaystyle\quad-\left[(c(h_{0}g^{-1},z_{h_{0}})+c((h_{0}g^{-1})^{-1},z(h_{0})))(-\mathbbm{1}_{\{z(h_{0})=a\}}+\mathbbm{1}_{\{z(h_{0})=b\}})\right].

Therefore,

|Δgz,z(a)|\displaystyle\left|\Delta^{g}_{z,z^{\prime}}(a)\right| c(h0g1,z(h0))+c(gh01,z(h0))+c(h0g1,z(h0))+c(gh01,z(h0))\displaystyle\leq c(h_{0}g^{-1},z^{\prime}(h_{0}))+c(gh_{0}^{-1},z^{\prime}(h_{0}))+c(h_{0}g^{-1},z(h_{0}))+c(gh_{0}^{-1},z(h_{0}))
2(C(h0g1)+C(gh01)),\displaystyle\leq 2(C(h_{0}g^{-1})+C(gh_{0}^{-1})),

where the last inequality follows from the definition of C(g)C(g). ∎

Claim 3.

Let g,h0Gg,h_{0}\in G and z,zXz,z^{\prime}\in X be such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}. Then, for every AA\in\mathcal{E},

(38) γ0g(A,z)=γg(A×G{g},z)=aAexp(φgz,z(a,0))nexp(φgz,z(n,0))\gamma^{0}_{g}(A,z)=\gamma_{g}(A\times\mathbb{N}^{G\setminus\{g\}},z)=\frac{\sum_{a\in A}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,0)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,0)\right)}

and

(39) γ0g(A,z)=γg(A×G{g},z)=aAexp(φgz,z(a,1))nexp(φgz,z(n,1)),\gamma^{0}_{g}(A,z^{\prime})=\gamma_{g}(A\times\mathbb{N}^{G\setminus\{g\}},z^{\prime})=\frac{\sum_{a\in A}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,1)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,1)\right)},

where γ\gamma is the Gibbsian specification given by equation (7).

Proof of Claim 3.

Begin by noticing that, from equation (8), for any zXz\in X, we can write

γg(A×G{g},z)=aγg([ag],z)𝟙{aA}=aAγg([ag],z).\displaystyle\gamma_{g}(A\times\mathbb{N}^{G\setminus\{g\}},z)=\sum_{a\in\mathbb{N}}\gamma_{g}\left([a^{g}],z\right)\mathbbm{1}_{\{a\in A\}}=\sum_{a\in A}\gamma_{g}\left([a^{g}],z\right).

Moreover, from Proposition 4.9, we have that

γg([ag],z)=exp(ϕτa,b(bz{g}c))nexp(ϕτn,b(bz{g}c))=exp(φgz(a))nexp(φgz(n))=exp(φgz,z(a,0))nexp(φgz,z(n,0)).\displaystyle\gamma_{g}([a^{g}],z)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{a,b}}(bz_{\{g\}^{c}})\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\phi_{*}^{\tau_{n,b}}(bz_{\{g\}^{c}})\right)}=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z}(a)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z}(n)\right)}=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,0)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,0)\right)}.

This concludes the proof of equation (38). The proof of equation (39) follows from an analogous argument, by just replacing zz by zz^{\prime}. ∎

Now, let mm be the counting measure on \mathbb{N} and, for each t[0,1]t\in[0,1], gGg\in G, and aa\in\mathbb{N}, consider the measure

νt=χg(,t)dm, with χg(a,t)=exp(φgz,z(a,t))nexp(φgz,z(n,t)).\nu_{t}=\chi_{g}(\cdot,t)dm,\text{ with }\chi_{g}(a,t)=\frac{\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,t)\right)}.

For each AA\subseteq\mathbb{N}, g,h0Gg,h_{0}\in G, and z,zXz,z^{\prime}\in X such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}, from Claim 3, we obtain that

ν0(A)=aAexp(φgz,z(a,0))nexp(φgz,z(n,0))=γ0g(A,z)\displaystyle\nu_{0}(A)=\frac{\sum_{a\in A}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,0)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,0)\right)}=\gamma^{0}_{g}(A,z)

and

ν1(A)=aAexp(φgz,z(a,1))nexp(φgz,z(n,1))=γ0g(A,z).\displaystyle\nu_{1}(A)=\frac{\sum_{a\in A}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,1)\right)}{\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,1)\right)}=\gamma^{0}_{g}(A,z^{\prime}).

In order to study conditions under which Theorem 6.1 holds, we need some estimates, which we calculate now. First, notice that ν1ν0TV=12|χg(a,1)χg(a,0)|dm.\left\|\nu_{1}-\nu_{0}\right\|_{TV}=\frac{1}{2}\int\left|\chi_{g}(a,1)-\chi_{g}(a,0)\right|dm.

Claim 4.

For each aa\in\mathbb{N} and gGg\in G, the map tχg(a,t)t\mapsto\chi_{g}(a,t) is differentiable and

tχg(a,t)=χg(a,t)(Δgz,z(a)Δgz,z(b)dνt(b)).\frac{\partial}{\partial t}\chi_{g}(a,t)=\chi_{g}(a,t)\left(\Delta^{g}_{z,z^{\prime}}(a)-\int\Delta^{g}_{z,z^{\prime}}(b)\,d\nu_{t}(b)\right).
Proof.

Notice that, for each t[0,1]t\in[0,1], the function exp(φgz,z(,t))\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(\cdot,t)\right) is integrable with respect to mm. Furthermore, for each aa\in\mathbb{N}, texp(φgz,z(a,t))\frac{\partial}{\partial t}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right) exists and, more precisely,

texp(φgz,z(a,t))=Δgz,z(a)exp(φgz,z(a,t)).\frac{\partial}{\partial t}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right)=\Delta^{g}_{z,z^{\prime}}(a)\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right).

Moreover, notice that, from Claim 2, the expression Δgz,z(a)exp(φgz,z(a,t))\Delta^{g}_{z,z^{\prime}}(a)\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right) is integrable with respect to mm. Therefore, we conclude that

(40) texp(φgz,z(a,t))dm(a)=Δgz,z(a)exp(φgz,z(a,t))dm(a).\frac{\partial}{\partial t}\int\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right)\,dm(a)=\int\Delta^{g}_{z,z^{\prime}}(a)\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right)\,dm(a).

Now, fix gg and h0Gh_{0}\in G and let z,zXz,z^{\prime}\in X be such that zG{h0}=zG{h0}z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}. Then, the map tnexp(φgz,z(n,t))t\mapsto\sum_{n\in\mathbb{N}}\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(n,t)\right) is differentiable with respect to tt. This implies that the map tχg(x,t)t\mapsto\chi_{g}(x,t) is also differentiable.

From equation (40) and the fact that tχg(x,t)t\mapsto\chi_{g}(x,t) is differentiable, we have that

tχg(a,t)\displaystyle\frac{\partial}{\partial t}\chi_{g}(a,t) =t(exp(φgz,z(a,t))exp(φgz,z(b,t))dm(b))\displaystyle=\frac{\partial}{\partial t}\left(\frac{\mathop{\textrm{\rm exp}}\nolimits(\varphi^{g}_{z,z^{\prime}}(a,t))}{\int\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(b,t)\right)\,dm(b)}\right)
=Δgz,z(a)χg(a,t)(Δgz,z(b)χg(b,t)dm(b))exp(φgz,z(a,t))exp(φgz,z(b,t))dm(b)\displaystyle=\Delta^{g}_{z,z^{\prime}}(a)\chi_{g}(a,t)-\frac{\left(\int\Delta^{g}_{z,z^{\prime}}(b)\chi_{g}(b,t)\,dm(b)\right)\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(a,t)\right)}{\int\mathop{\textrm{\rm exp}}\nolimits\left(\varphi^{g}_{z,z^{\prime}}(b,t)\right)\,dm(b)}
=χg(a,t)(Δgz,z(a)Δgz,z(b)dνt(b)).\displaystyle=\chi_{g}(a,t)\left(\Delta^{g}_{z,z^{\prime}}(a)-\int\Delta^{g}_{z,z^{\prime}}(b)\,d\nu_{t}(b)\right).

Considering Claim 4, we have that

|χg(a,1)χg(a,0)|dm(a)\displaystyle\int\left|\chi_{g}(a,1)-\chi_{g}(a,0)\right|dm(a) =|01(tχg(a,t))dt|dm(a)\displaystyle=\int\left|\int_{0}^{1}\left(\frac{\partial}{\partial t}\chi_{g}(a,t)\right)\,dt\right|\,dm(a)
01|tχg(a,t)|dm(a)dt\displaystyle\leq\int_{0}^{1}\int\left|\frac{\partial}{\partial t}\chi_{g}(a,t)\right|dm(a)\,dt
=01|χg(a,t)(Δgz,z(a)Δgz,z(b)dνt(b))|dm(a)dt\displaystyle=\int_{0}^{1}\int\left|\chi_{g}(a,t)\left(\Delta^{g}_{z,z^{\prime}}(a)-\int\Delta^{g}_{z,z^{\prime}}(b)\,d\nu_{t}(b)\right)\right|dm(a)\,dt
=01|Δgz,z(a)Δgz,z(b)dνt(b)|dνt(a)dt\displaystyle=\int_{0}^{1}\int\left|\Delta^{g}_{z,z^{\prime}}(a)-\int\Delta^{g}_{z,z^{\prime}}(b)\,d\nu_{t}(b)\right|d\nu_{t}(a)\,dt
01(Δgz,z+Δgz,zdνt(b))dνt(a)dt\displaystyle\leq\int_{0}^{1}\int\left(\|\Delta^{g}_{z,z^{\prime}}\|_{\infty}+\int\|\Delta^{g}_{z,z^{\prime}}\|_{\infty}\,d\nu_{t}(b)\right)d\nu_{t}(a)\,dt
=2Δgz,z.\displaystyle=2\|\Delta^{g}_{z,z^{\prime}}\|_{\infty}.

Thus, by Claim 2 we have

ρgh012supz,zXzG{h0}=zG{h0}2Δgz,z2(C(h0g1)+C(gh01)).\displaystyle\rho_{gh_{0}}\leq\frac{1}{2}\sup_{\begin{subarray}{c}z,z^{\prime}\in X\\ z_{G\setminus\{h_{0}\}}=z^{\prime}_{G\setminus\{h_{0}\}}\end{subarray}}2\|\Delta^{g}_{z,z^{\prime}}\|_{\infty}\leq 2(C(h_{0}g^{-1})+C(gh_{0}^{-1})).

Therefore, considering that ρgg=0\rho_{gg}=0,

hGρgh\displaystyle\sum_{h\in G}\rho_{gh} 2hGhg[C(hg1)+C(gh1)]=2hGh1G[C(h)+C(h1)]=4hGh1GC(h),\displaystyle\leq 2\sum_{\begin{subarray}{c}h\in G\\ h\neq g\end{subarray}}\left[C(hg^{-1})+C(gh^{-1})\right]=2\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}\left[C(h)+C(h^{-1})\right]=4\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}C(h),

so

c(γ)=supgGhGρgh(γ)4hGh1GC(h).c(\gamma)=\sup_{g\in G}\sum_{h\in G}\rho_{gh}(\gamma)\leq 4\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}C(h).

Finally, if we consider the potential βϕ\beta\phi for β>0\beta>0, then by linearity, we have

c(γβϕ)=supgGhGρgh(γβϕ)4βhGh1GC(h),c(\gamma^{\beta\phi})=\sup_{g\in G}\sum_{h\in G}\rho_{gh}(\gamma^{\beta\phi})\leq 4\beta\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}C(h),

where γβϕ\gamma^{\beta\phi} is the specification given by equation (7) for the potential βϕ\beta\phi. Thus, if

β<(4hGh1GC(h))1,\beta<\left(4\sum_{\begin{subarray}{c}h\in G\\ h\neq 1_{G}\end{subarray}}C(h)\right)^{-1},

Dobrushin’s condition is satisfied and, by Theorem 6.1, we have at most one DLR measure for the potential βϕ\beta\phi. Furthermore, if β>0\beta>0, then the set of GG-invariant DLR measures for βϕ\beta\phi is non-empty, so that we can guarantee that if β(0,14h1GC(h))\beta\in\left(0,\frac{1}{4\sum_{h\neq 1_{G}}C(h)}\right), there exists exactly one DLR measure for βϕ\beta\phi.

Acknowledgements

Elmer R. Beltrán would like to thank to the fellow program Fondo Postdoctorado Universidad Católica del Norte No 0001, 2020. Rodrigo Bissacot is supported by CNPq grants 312294/2018-2 and 408851/2018-0, by FAPESP grant 16/25053-8, and by the University Center of Excellence “Dynamics, Mathematical Analysis and Artificial Intelligence”, at the Nicolaus Copernicus University. Luísa Borsato is supported by grants 2018/21067-0 and 2019/08349-9, São Paulo Research Foundation (FAPESP). Raimundo Briceño would like to acknowledge the support of ANID/FONDECYT de Iniciación en Investigación 11200892.

References

  • [1] J. Aaronson and H. Nakada. Exchangeable, Gibbs and equilibrium measures for Markov subshifts. Ergod. Theory Dyn. Syst., 27(2):321–339, 2007.
  • [2] A. Alpeev. The entropy of Gibbs measures on sofic groups. J. Math. Sci., 215(6):649–659, 2016.
  • [3] V. Baladi. Gibbs states and equilibrium states for finitely presented dynamical systems. J. Stat. Phys., 62(1-2):239–256, 1991.
  • [4] S. Barbieri, R. Gómez, B. Marcus, and S. Taati. Equivalence of relative Gibbs and relative equilibrium measures for actions of countable amenable groups. Nonlinearity, 33(5):2409, 2020.
  • [5] S. Barbieri and T. Meyerovitch. The Lanford–Ruelle theorem for actions of sofic groups. Trans. Am. Math. Soc., 376(02):1299–1342, 2023.
  • [6] E. R. Beltrán, R. Bissacot, and E. O. Endo. Infinite DLR measures and volume-type phase transitions on countable Markov shifts. Nonlinearity, 34(7):4819, 2021.
  • [7] S. Berghout, R. Fernández, and E. Verbitskiy. On the relation between Gibbs and gg-measures. Ergod. Theory Dyn. Syst., 39(12):3224–3249, 2019.
  • [8] P. Billingsley. Convergence of probability measures. John Wiley & Sons, 2013.
  • [9] R. Bissacot, E. O. Endo, A. C. D. van Enter, and A. Le Ny. Entropic repulsion and lack of the gg-measure property for Dyson models. Commun. Math. Phys., 363(3):767–788, 2018.
  • [10] L. Boltzmann and M. Brillouin. Leçons sur la théorie des gaz, volume 2. Gauthier-Villars, 1902.
  • [11] L. Borsato. Strong Law of Large Numbers for Bernoulli sequences and Gibbs measures on subshifts for finite and infinite alphabets. PhD thesis, University of São Paulo, 2022.
  • [12] L. Borsato and S. MacDonald. A Dobrushin-Lanford-Ruelle theorem for irreducible sofic shifts, 2020. Preprint, arXiv:2007.05862 [math.DS].
  • [13] L. Borsato and S. MacDonald. Conformal measures and the Dobrushin-Lanford-Ruelle equations. Proc. Am. Math. Soc., 149(10):4355–4369, 2021.
  • [14] R. E. Bowen. Equilibrium states and the ergodic theory of Anosov diffeomorphisms, volume 470. Springer Science & Business Media, 2008.
  • [15] R. Briceño. Kieffer-pinsker type formulas for Gibbs measures on sofic groups, 2021. Preprint, arXiv:2108.06053 [math.DS].
  • [16] A. I. Bufetov. Pressure and equilibrium measures for actions of amenable groups on the space of configurations. Sb. Math., 202(3):341, 2011.
  • [17] D. Capocaccia. A definition of Gibbs state for a compact set with v\mathbb{Z}^{v} action. Commun. Math. Phys., 48(1):85–88, 1976.
  • [18] J. R. Chazottes and E. Ugalde. Entropy estimation and fluctuations of hitting and recurrence times for Gibbsian sources. Discrete Contin. Dyn. Syst. - B, 5(3):565–586, 2005.
  • [19] L. Cioletti, A. O. Lopes, and M. Stadlbauer. Ruelle operator for continuous potentials and DLR-Gibbs measures. Discrete Contin. Dyn. Syst., 40(8):4625, 2020.
  • [20] M. Denker and M. Urbański. On the existence of conformal measures. Trans. Am. Math. Soc., 328(2), 1991.
  • [21] P. L. Dobruschin. The description of a random field by means of conditional probabilities and conditions of its regularity. Theory Probab. its Appl., 13(2):197–224, 1968.
  • [22] R. L. Dobrushin. The problem of uniqueness of a gibbsian random field and the problem of phase transitions. Funct. Anal. its Appl., 2(4):302–312, 1968.
  • [23] R. L. Dobrushin. Gibbsian random fields for lattice systems with pairwise interactions. Funct. Anal. its Appl., 2(4):292–301, 1969.
  • [24] T. Downarowicz. Entropy in dynamical systems, volume 18. Cambridge University Press, 2011.
  • [25] T. Downarowicz, B. Frej, and P.-P. Romagnoli. Shearer’s inequality and infimum rule for Shannon entropy and topological entropy. Dynamics and Numbers, pages 63–75, 2016.
  • [26] T. Downarowicz, D. Huczek, and G. Zhang. Tilings of amenable groups. J. für die Reine und Angew. Math., 2019(747):277–298, 2019.
  • [27] H. Duminil-Copin. Lectures on the Ising and Potts models on the hypercubic lattice. In PIMS-CRM Summer School in Probability, pages 35–161. Springer, 2017.
  • [28] R. Fernández, S. Gallo, and G. Maillard. Regular gg-measures are not always Gibbsian. Electron. Commun. Probab., 16(none):732–740, 2011.
  • [29] S. Friedli and Y. Velenik. Statistical mechanics of lattice systems: a concrete mathematical introduction. Cambridge University Press, 2017.
  • [30] H.-O. Georgii. Gibbs measures and phase transitions, volume 9. Walter de Gruyter, 2011.
  • [31] H.-O. Georgii, O. Häggström, and C. Maes. The random geometry of equilibrium phases. In Phase transitions and critical phenomena, volume 18, pages 1–142. Elsevier, 2001.
  • [32] J. W. Gibbs. Elementary principles in Statistical Mechanics: developed with especial reference to the rational foundations of thermodynamics. C. Scribner’s sons, 1902.
  • [33] B. M. Gurevich. Shift entropy and Markov measures in the space of paths of a countable graph. Dokl. Akad. Nauk, 192(5):963–965, 1970.
  • [34] B. M. Gurevich and S. V. Savchenko. Thermodynamic formalism for countable symbolic Markov chains. Russ. Math. Surv., 53(2):245, 1998.
  • [35] B. M. Gurevich and A. A. Tempelman. A Breiman type theorem for Gibbs measures. J. Dyn. Control Syst., 13(3):363–371, 2007.
  • [36] N. Haydn and D. Ruelle. Equivalence of Gibbs and equilibrium states for homeomorphisms satisfying expansiveness and specification. Commun. Math. Phys., 148(1):155–167, 1992.
  • [37] M. Keane. Strongly mixing gg-measures. Invent. Math., 16(4):309–324, 1972.
  • [38] G. Keller. Equilibrium States in Ergodic Theory, volume 42. Cambridge University Press, 1998.
  • [39] D. Kerr and H. Li. Ergodic theory. Springer Monographs in Mathematics. Springer, Cham, 2016.
  • [40] B. Kimura. Gibbs measures on subshifts. Master’s thesis, University of São Paulo, 2015.
  • [41] O. Lanford and D. Ruelle. Observables at infinity and states with short range correlations in statistical mechanics. Commun. Math. Phys., 13(3):194–215, 1969.
  • [42] C. Löh. Geometric Group Theory. Springer, 2017.
  • [43] R. D. Mauldin and M. Urbański. Gibbs states on the symbolic space over an infinite alphabet. Isr. J. Math., 125:93–130, 2001.
  • [44] T. Meyerovitch. Gibbs and equilibrium measures for some families of subshifts. Ergod. Theory Dyn. Syst., 33:934–953, 2013.
  • [45] S. Muir. Gibbs/equilibrium measures for functions of multidimensional shifts with countable alphabets. PhD thesis, University of North Texas, 2011.
  • [46] S. Muir. A new characterization of Gibbs measures on 𝕫d\mathbb{N}^{\mathbbm{z}^{d}}. Nonlinearity, 24(10):2933, 2011.
  • [47] J. R. Munkres. Elements of algebraic topology. CRC Press, 2018.
  • [48] D. S. Ornstein and B. Weiss. Entropy and isomorphism theorems for actions of amenable groups. J. Anal. Math., 48(1):1–141, 1987.
  • [49] K. Petersen and K. Schmidt. Symmetric Gibbs measures. Trans. Am. Math. Soc., 349(7):2775–2811, 1997.
  • [50] C.-E. Pfister. Gibbs measures on compact ultra metric spaces, 2022. Preprint, arXiv:2202.06802 [math.DS].
  • [51] F. Rassoul-Agha and T. Seppäläinen. A course on large deviations with an introduction to Gibbs measures, volume 162. American Mathematical Soc., 2015.
  • [52] D. Ruelle. Thermodynamic formalism: the mathematical structures of equilibrium statistical mechanics. Cambridge, 2nd edition, 2004.
  • [53] O. M. Sarig. Thermodynamic formalism for countable Markov shifts. Ergod. Theory Dyn. Syst., 19(6):1565–1593, 1999.
  • [54] O. M. Sarig. Lecture notes on thermodynamic formalism for topological Markov shifts. Penn State, 2009.
  • [55] C. Shriver. Free energy, Gibbs measures, and Glauber dynamics for nearest-neighbor interactions on trees, 2020. Preprint, arXiv:2011.00653 [math.PR].
  • [56] Y. G. Sinai. Gibbs measures in ergodic theory. Russ. Math. Surv., 27(4):21, 1972.
  • [57] A. A. Tempelman. Ergodic theorems for group actions: Informational and Thermodynamical Aspects, volume 78. Springer Science & Business Media, 2013.
  • [58] P. Varandas and Y. Zhao. Weak Gibbs measures: speed of convergence to entropy, topological and geometrical aspects. Ergod. Theory Dyn., 37(7):2313–2336, 2017.
  • [59] P. Walters. Ruelle’s operator theorem and gg-measures. Trans. Am. Math. Soc., 214:375–387, 1975.
  • [60] M. Yuri. Zeta functions for certain non-hyperbolic systems and topological Markov approximations. Ergod. Theory Dyn. Syst., 18(6):1589–1612, 1998.