1 Introduction
The thermodynamic formalism, developed by Sinai, Ruelle and Bowen in the 1970s and 1990s, is a part of ergodic theory that came into existence through the application of techniques and results from statistical mechanics in the realm of smooth dynamics. One of its main goals is to describe the statistical behavior of a dynamical system via invariant measures, called equilibrium states, that maximize the free energy of the system.
In the classical setting, an equilibrium state associated to a continuous transformation $T:M\rightarrow M$ defined on a compact metric space M and a continuous potential $\phi :M\rightarrow \mathbb {R}$ is an invariant probability measure $\mu _{T, \phi }$ characterized by the following variational principle:
where $P_{T}(\phi )$ is the topological pressure, $h_{\mu }(T)$ denotes the entropy and the supremum is taken over all invariant probability measures.
This theory was initiated by the pioneering work of Sinai [Reference Simmons and Urbański27] who proved the existence and uniqueness of equilibrium states for Anosov diffeomorphisms and Hölder continuous potentials. In subsequent works Bowen [Reference Bowen11] and Ruelle [Reference Ruelle25] extended the results of Sinai to uniformly hyperbolic systems and Hölder continuous potentials. Since then, important contributions for this theory in the deterministic case have been given by several authors (see, for example, [Reference Climenhaga and Thompson13, Reference Oliveira and Viana21, Reference Sarig26, Reference Varandas and Viana29]).
In the context of random dynamical systems, the study of equilibrium states is still quite far from being well understood, despite some advances in the area. Briefly, a random dynamical system is a skew-product $F(w,x)=(\theta (w),f_{w}(x))$ where the randomness is modeled by an invertible transformation $\theta $ preserving an ergodic measure $\mathbb {P}.$ We are interested in understanding the dynamics of compositions
As in the deterministic case, the random topological pressure of the system is the supremum of the entropy plus the integration of the potential among all invariant probability measures whose marginal is $\mathbb {P}$ . We refer the reader to [Reference Arnold6] for the background to and a treatment of this topic.
Having established a variational principle for random maps, it is natural to ask what kinds of random dynamical systems and potentials we can develop the theory of equilibrium states. In [Reference Kifer16] Kifer proved the existence and uniqueness of equilibrium states for random uniformly expanding maps associated to Hölder continuous potentials. In [Reference Liu18] Liu extended this result for uniformly hyperbolic random systems. Later, the thermodynamical formalism was developed by Kifer [Reference Kifer15] for random expansion in average transformations, and by Mayer, Skorulski and Urbanski [Reference Mayer, Skorulski and Urbański20] for distance expanding random mappings. In the context of random countable Markov shifts, the thermodynamic formalism was proved by Denker, Kifer and Stadlbauer in [Reference Denker, Kifer and Stadlbauer14]. The existence of equilibrium states with positive Lyapunov exponents was proved by Arbieto, Matheus and Oliveira [Reference Arbieto, Matheus and Oliveira5] for certain non-uniformly expanding maps and continuous potentials with low variation. In [Reference Birkhoff10] Bilbao and Oliveira obtained uniqueness of maximizing entropy measures in this context. Recently, Stadlbauer, Suzuki and Varandas [Reference Stadlbauer, Suzuki and Varandas28] developed the thermodynamical formalism for a wide class of random maps with non-uniform expansion and differentiable potentials at high temperature.
In this work we develop the thermodynamical formalism for a robust class of random non-uniformly expanding local homeomorphisms associated to Hölder continuous potentials with small variation. First, we prove the existence of an invariant measure absolutely continuous with respect to the leading eigenmeasures of the dual transfer operators. This invariant measure is indeed an equilibrium state for the random dynamical system and it is unique in the setting of non-uniformly expanding measures. Moreover, we show that the random topological pressure is the integral of the leading eigenvalues of the transfer operators. As an application of our techniques, we extend the results obtained in [Reference Arbieto, Matheus and Oliveira5, Reference Bilbao and Oliveira9] for Hölder continuous potentials with small variation.
Finally, we study the persistence of the equilibrium state under small perturbations of the system. In the context of Sinai–Ruelle–Bowen measures, the continuous dependence with respect to the dynamics was obtained by Alves and Viana [Reference Alves and Viana4] for maps with non-uniform expansion. Such continuity was also proved by Baladi [Reference Baladi7] and Young [Reference Young30] for random perturbations of uniformly hyperbolic systems and by Alves and Araújo in [Reference Alves and Araújo1] for random perturbations of non-uniformly expanding maps. More generally, the continuity of the equilibrium state was proved by Castro and Varandas [Reference Castro and Varandas12] for a class of non-uniformly expanding maps and potentials with small variation. This property was also obtained by Alves, Ramos and Siqueira [Reference Alves, Ramos and Siqueira3] for non-uniformly hyperbolic systems and hyperbolic potentials. Here, we deal with a family of random non-uniformly expanding maps and potentials with small variation. We prove that the non-uniformly expanding equilibrium state as well as the random topological pressure vary continuously within this family.
We organize this paper as follows. In §2 we present our setting and state the main results. In §3 we introduce basic definitions such as random topological pressure and projective metrics. In §4 we recall the definition of reference measures and prove some properties that will be useful throughout the work. In §5 we use the projective metric approach to obtain the thermodynamical formalism. In §6 we prove the existence and uniqueness of equilibrium states among non-uniformly expanding measures. In §7 we show the continuous dependence of these equilibrium states and the topological pressure as functions of the random dynamics and the potential. In the final section we describe some applications of our results.
2 Setting and main results
Let M be a compact and connected manifold with distance d and $\Omega $ the space of local homeomorphisms defined on $M.$ Consider a Polish space X (that is, a separable complete metric space) and an invertible measurable map $\theta :X\rightarrow X$ preserving an ergodic Borel measure $\mathbb {P}$ of X. We recall that a random dynamical system is a continuous map $f:X\to \Omega $ given by $w\mapsto f_{w}\in \Omega $ where $(w, x)\mapsto f_{w}(x)$ is measurable. For every $n\geq 0$ we define
The skew-product generated by the maps $f_{w}$ is the measurable transformation
In particular, $F^{n}(w, x)=(\theta ^{n}(w), f^{n}_{w}(x) )$ for every $n\in \mathbb {Z}.$
Let $\mathcal {M}_{\mathbb {P}}(X \times M)$ be the space of probability measures on $X \times M$ such that the marginal on $\Omega $ is $\mathbb {P}$ . Denote by $\mathcal {M}_{\mathbb {P}}(F)\subset \mathcal {M}_{\mathbb {P}}(X \times M)$ the set of F-invariant measures. Note that, by Rokhlin’s disintegration theorem [Reference Rokhlin24], for every $\mu \in \mathcal {M}_{\mathbb {P}}(F)$ there exists a system of sample measures $\{\mu _{w}\}_{w\in X}$ of $\mu $ such that
We say that an F-invariant measure $\mu $ is ergodic if $(F, \mu )$ is ergodic.
2.1 Hypothesis about the generating maps
For each $w\in X,$ let $f_{w}:M \rightarrow M$ be a local homeomorphism satisfying the following requirement: there exist $\delta _{w}>\delta >0$ and a continuous function $L_{w} :M\rightarrow \mathbb {R}_{+}$ such that for every $x \in M$ we can find a neighborhood $U_{x}$ where $f_{w}: U_{x} \rightarrow B_{\theta (w)}(f_{w}(x), \delta _{w})$ is invertible and
As $f_{w}$ is a local homeomorphism defined on a compact metric space, we have that the number of preimages $\deg (f_{w}):\#f_{w}^{-1}(x)$ is constant for all $x \in M$ . We assume that $\deg (F)= \sup _{w} \deg (f_{w}) < \infty $ .
Suppose that there are an open region $\mathcal {A}_{w}\subset M$ , constants $\sigma _{w}>1$ and $L_{w}\geq 1$ close enough to $1$ such that the following conditions hold.
-
(I) $L_{w}(x) \leq L_{w}$ for every $x \in \mathcal {A}_{w}$ and $L_{w}(x) < \sigma _{w}^{-1}$ for every $x \in \mathcal {A}_{w}^{c}=M\setminus \mathcal {A}_{w}$ .
-
(II) There exists a finite covering $\mathcal {U}_{w}$ of M, by open domains of injectivity for $f_{w}$ , such that $\mathcal {A}_{w}$ can be covered by $q_{w} < \deg (f_{w})$ elements of $\mathcal {U}_{w}$ .
-
(III) For every $\varepsilon>0$ we can find some positive integer $\tilde {n}=\tilde {n}(w,\varepsilon )$ satisfying $f_{\theta ^{j}(w)}^{\tilde {n}}(B_{\theta ^{j}(w)}(f^{j}_{w}(x),\varepsilon ))=M$ for any $j\geq 0.$
We observe that the continuous function $L_{w}(\cdot )$ associates the Lipschitz constant of the inverse branches. This property and the assumption $\delta _{w}>\delta >0$ imply the uniform openness $B_{\theta (w)}(f_{w}(x), \delta )\subset f_{w}(U_{x})$ , and thus for every $(w, x)\in X\times M$ there exists a unique continuous inverse branch of $f_{w}$ defined on $B_{\theta (w)}(f_{w}(x), \delta )$ sending $f_{w}(x)$ to x. Conditions (I) and (II) mean that expanding and contracting behavior may exist in M, but at least one preimage is required for every point in the expanding region. Condition (III) means that the skew-product F is topologically exact.
Next, we present the setting of potentials that will be considered. For $\alpha>0$ , consider the space $C^{\alpha }(M)$ of Hölder continuous functions $\varphi :M\to \mathbb {R}$ endowed with the seminorm
and the norm
where $\|\cdot \|_{\infty }$ stands for the $\sup $ norm. Denote by $\mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M))$ the space of all measurable functions $\phi :X\times M\rightarrow \mathbb {R}$ such that for all $w\in X$ , the fiber potential $\phi _{w}:M\rightarrow \mathbb {R}$ defined by $\phi _{w}(x):=\phi (w,x)$ is Hölder continuous and $\lVert \phi \rVert _{1}=\int _{X}\lVert \phi _{w}\rVert _{\infty }\,d\mathbb {P}(w)<+\infty $ . For $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M))$ we assume the existence of some positive $\varepsilon _{\phi }>0$ satisfying, for all $w\in X$ ,
Notice that all potentials $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M))$ in a neighborhood of zero satisfy condition (IV). In the literature this class of potentials is referred to as small variation.
Let $p_{w}:=\deg f_{w}-q_{w}.$ The choice of $\varepsilon _{\phi }$ and $L_{w}$ must satisfy, for each $w\in X$ ,
2.2 Statement of results
Consider the space $C^{0}(M)$ of real continuous functions $\psi :M\rightarrow \mathbb {R}$ endowed with the uniform convergence norm. Given $w\in X$ , let $f_{w}:M\rightarrow M$ be the dynamics and $\phi _{w}:M\rightarrow \mathbb {R}$ be the potential on the fiber. The Ruelle–Perron–Frobenius operator or simply transfer operator associated to $(f_{w}, \phi _{w})$ is the linear operator $\mathcal {L}_{w}:C^{0}(M)\rightarrow C^{0}(M)$ defined by
Its dual operator $\mathcal {L}_{w}^{\ast }:[C^{0}(M)]^{\ast }\to [C^{0}(M)]^{\ast } $ acts on the space of Borel measures as follows:
In our first result we describe the thermodynamic formalism for random non-uniformly expanding maps.
Theorem A. Consider a random dynamical system $F:X\times M\to X\times M$ satisfying conditions (I), (II) and (III). For any potential $\phi :X\times M\to \mathbb {R}$ satisfying (IV) and (V) the following assertions hold.
-
(1) There exists a unique measurable family of probabilities $\{\nu _{w}\}_{w\in X}$ such that
-
(3) There exists a unique measurable family of Hölder continuous functions $\{h_{w}\}_{w\in X}$ bounded away from zero and infinity such that
-
(3) The probability measure $\mu :=\{\mu _{w}\}_{w\in X}$ where $\mu _{w}:=h_{w}\nu _{w}$ is F-invariant.
We also derive that the F-invariant family $\{\mu _{w}\}_{w\in X}$ obtained in the last theorem has an exponential decay of correlations for Hölder continuous observables.
Theorem B. There exists $0<\tau < 1$ such that for any $\varphi \in L^{1}(\mu _{\theta ^{n}(w)})$ and $\psi \in C^{\alpha }(M) $ there exists a positive constant $K(\varphi , \psi ) $ satisfying
for all $n\geq 1$ .
The weak hyperbolicity property of the generating maps allows us to prove that the F-invariant measure given by Theorem A is indeed an equilibrium state for the random dynamical system. Moreover, it is unique if we consider only the measures whose pressure is located on the expanding region. We specify the setting as follows.
Suppose that there exists $c>0$ such that for $\mathbb {P}$ -almost every $w\in X$ we can find $\tilde {L}_{w}$ close enough to $1$ and $\tilde {\sigma }_{w}>1$ satisfying for every $j\geq 0$ that
where $\rho $ is given by Lemma 4.2.
We say that a subset H of $X\times M$ is non-uniformly expanding if there exists some positive constant $c>0$ such that
A probability measure $\eta $ , not necessarily invariant, is called non-uniformly expanding with exponent c if $\eta (H)=1$ . Condition (VI) above will be used to prove that the F-invariant measure given by Theorem A is non-uniformly expanding.
Our next result states uniqueness of equilibrium states for random dynamical systems among non-uniformly expanding measures.
Theorem C. Let $F:X\times M\to X\times M$ be a random dynamical system and $\phi :X\times M\to \mathbb {R}$ be a potential function satisfying conditions (I)–(VI). There exists only one F-invariant non-uniformly expanding measure $\mu _{F, \phi }\in \mathcal {M}_{\mathbb {P}}(F)$ maximizing the variational principle
where the supremum is taken in the set $\mathcal {M}_{\mathbb {P}}(F).$ Thus, $\mu _{F, \phi }$ is the unique non-uniformly expanding equilibrium state of $(F|\theta , \phi ).$
Once we have proved uniqueness of equilibrium states, we will investigate its persistence under small perturbations of the random system and the potential.
As defined above, consider the space $\mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M))$ of integrable potentials and let $\mathcal {D}\subset \Omega $ be the space of $C^{1}$ local diffeomorphisms defined on M. We shall consider the product topology on $\mathcal {D}\times \mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M))$ . We fix an invertible transformation $\theta : X\to X$ preserving an ergodic measure $\mathbb {P}$ and consider the family $\mathcal {S}$ of skew-products generated by maps of $\mathcal {D}$ ,
where $(w, x)\mapsto f_{w}\in \mathcal {D}$ is measurable. Now we define the family
By Theorem C, each $(F, \phi )\in \mathcal {H}$ has only one non-uniformly expanding equilibrium state. Our last main result establishes the continuity in the weak star topology of such equilibria within this family; this property is called equilibrium stability. We also prove the continuity of the random topological pressure in this setting.
Theorem D. The non-uniformly expanding equilibrium state and the topological pressure vary continuously on $\mathcal {H}$ .
We point out that we are fixing an invertible transformation $\theta :X\to X$ preserving an ergodic measure $\mathbb {P}.$ However, the proof of Theorem D remains true if we vary $\theta $ in the space of continuous functions.
3 Preliminaries
In this section we state some basic definitions and results about random dynamical systems that will be used throughout the text. We also recall the notion of hyperbolic times and projective metrics.
3.1 Entropy and topological pressure
We start with the definition of entropy and topological pressure for random transformations. The reader can consult more results and properties in Kifer [Reference Kifer15] and Liu [Reference Liu18].
Let $\mu \in \mathcal {M}_{\mathbb {P}}(F)$ be an F-invariant measure. Given a finite measurable partition $\xi $ of M, we set
where $H_{\nu }(\xi )=-\sum _{P\in \xi }\nu (P)\log \nu (P)$ for a finite partition $\xi $ and $\mu _{w}$ is the sample measure of $\mu .$ The entropy of $(F|_{\theta }, \mu )$ is
where the supremum is taken over all finite measurable partitions of M.
Denote by $\mathbb {L}^{1}_{\mathbb {P}}(X,C^{0}(M))$ the space of all measurable functions $\phi :X\times M\rightarrow \mathbb {R}$ such that $\phi _{w}:M\rightarrow \mathbb {R}$ defined by $\phi _{w}(x):=\phi (w,x)$ is continuous for all $w\in X$ and $\lVert \phi \rVert _{1}=\int _{X}\lVert \phi _{w}\rVert _{\infty }\,d\mathbb {P}(w)<+\infty $ .
Fix $w\in X$ . Given $\varepsilon>0$ and an integer $n\geq 1$ , we say that a subset $F_{n}\subseteq M$ is $(w,n,\varepsilon )$ -separated if for every two distinct points $y,z\in F_{n}$ there exists some $j\in \{0,1,\ldots ,n-1\}$ such that $d(f^{j}_{w}(y),f^{j}_{w}(z))>\varepsilon $ .
For $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{0}(M))$ , $\varepsilon>0$ and $n\geq 1$ we consider
where $S_{n}\phi (w,y):=\sum ^{n-1}_{j=0}\phi _{\theta ^{j}(w)}( f^{j}_{w}(y))$ .
The random topological pressure of $\phi $ relative to $\theta $ is defined by
Thus, the following pressure map is well defined:
In particular, the topological entropy of F relative to $\theta $ is $h_{\mathrm {top}}(F|_{\theta })=P_{F|\theta }(0)$ .
The topological pressure and the entropy are related by the well-known variational principle. The reader can see a proof of this result in [Reference Liu18].
Theorem 3.1. (Variational principle)
Assume that $(X, \mathbb {P})$ is a Lebesgue space. Then for any $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{0}(M))$ we have
Moreover, when $\mathbb {P}$ is ergodic, we can consider the supremum over ergodic measures.
Motivated by the variational principle, we say that an F-invariant measure $\mu \in \mathcal {M}_{\mathbb {P}}(F)$ is an equilibrium state for $(F|_{\theta }, \phi )$ relative to $\theta $ if the supremum (1) is attained by $\mu $ , that is,
Next we define the random topological pressure via open coverings and via dynamic balls. We take as reference the deterministic case where this approach is characteristic in dimension theory. We refer the reader to [Reference Pesin22, Reference Stadlbauer, Suzuki and Varandas28] for more details.
Let $w\in X$ and denote by $\mathcal {U}$ a finite open cover of M. Denote by $\mathcal {S}_{n}(\mathcal {U}{\kern1pt})$ the set of all strings $\mathbf {U}=\{U_{i_{0}}, \ldots , U_{i_{n-1}}; U_{i_{j}}\in \mathcal {U}\}$ of length $n=n(\mathbf {U})$ and put $\mathcal {S}=\mathcal {S}(\mathcal {U}{\kern1pt})=\bigcup _{n\geq 0}\mathcal {S}_{n}(\mathcal {U}{\kern1pt})$ . Given a string $\mathbf {U}=\{U_{i_{0}}, \ldots , U_{i_{n-1}}\}\in \mathcal {S}(\mathcal {U}{\kern1pt})$ , we consider the cylinder
Let $\mathcal {F}_{(N, w)}$ be the collection of all cylinders of depth at least N, that is,
For $\beta \in \mathbb {R}$ and $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{0}(M))$ let
where $S_{n(\mathbf {U})}\phi (X_{w})=\sup _{y\in X_{w}}\sum _{j=0}^{n(\mathbf {U})-1}\phi _{\theta ^{j}(w)}(f^{j}_{w}(y))$ and the infimum is taken over all finite families $\mathcal {F}$ of $\mathcal {F}_{(N, w)}$ in order that (2) is measurable in w (see, for example, §9 of [Reference Stadlbauer, Suzuki and Varandas28]). As N goes to infinity we define
The existence of the limit above is guaranteed by the function $m_{\beta }(w, \phi , F|\theta , \mathcal {U}, N)$ to be increasing with N. Taking the infimum over $\beta $ we call
Let $|\mathcal {U}|=\max \{\mathrm {diam} U_{i}; U_{i}\subset \mathcal {U}\}$ be the diameter of the cover $\mathcal {U}$ and consider
In [Reference Pesin22, Theorem 11.1] it was showed that this quantity is well defined and does not depend on the cover $\mathcal {U}$ . Moreover, since all quantities defined above are measurable functions of $w\in X$ (see, for example, §9 of [Reference Stadlbauer, Suzuki and Varandas28]), we can define the random topological pressure of $(F|\theta , \phi )$ as
In the following we present another way to define the random topological pressure. We fix $w\in X$ and $\varepsilon>0$ . For $n\in \mathbb {N}$ , $x\in M$ , let $B_{w}(x,n,\varepsilon )$ be the dynamic ball
We denote by $G_{(N,w)}$ the collection of dynamic balls:
Let $U_{w}$ be a finite or countable family of $G_{(N,w)}$ which covers M. For every $\beta \in \mathbb {R}$ and $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X,C^{0}(M))$ let
where $S_{n}\phi (B_{w}(x,n,\varepsilon ))=\sup _{y\in B_{w}(x,n,\varepsilon )}\sum _{j=0}^{n-1}\phi _{\theta ^{j}(w)}(f^{j}_{w}(y))$ . When N goes to infinity we consider
Taking the infimum over $\beta $ , we define
Since $P_{F|\theta }(w, \phi , \varepsilon )$ is decreasing on $\varepsilon $ we can take the limit
Now if we consider a finite open cover $\mathcal {U}$ of M with Lebesgue number $\varepsilon (\mathcal {U}{\kern1pt})$ we have
which implies that
Therefore, the definitions of random topological pressure via coverings and via dynamic balls coincide.
3.2 Hyperbolic times
In order to explore the non-uniform expansion of the set H we need the notion of hyperbolic times. The reader can obtain more details of this concept in [Reference Alves, Bonatti and Viana2, Reference Arbieto, Matheus and Oliveira5]. In our context, the function $L_{w}(\cdot )$ plays the role of the derivative $\|Df_{w}(\cdot )^{-1}\|.$
Definition 3.1. Let $w\in X$ and $L_{w}:X\to \mathbb {R}$ be as in §2. We say that $n\in \mathbb {N}$ is a c-hyperbolic time for $(w,x)\in X\times M$ if
It is a well-known fact that if $\eta $ is a non-uniformly expanding measure with exponent c then $\eta $ -almost every point $(w,x)\in H$ has infinitely many c-hyperbolic times. A proof of this result can be found in [Reference Alves, Bonatti and Viana2].
The uniform domination required in condition (VI) will allow us to prove that the F-invariant measure $\mu $ given by Theorem A is non-uniformly expanding. Thus, we will conclude that $\mu $ -almost every point $(w, x)\in X\times M$ has infinitely many hyperbolic times.
Lemma 3.1. Given $c>0$ , there exists $\tilde {\delta }=\tilde {\delta }(c)>0$ such that, for $\mathbb {P}$ -almost every ${w\in X}$ , if n is a c-hyperbolic time of $(w,x)$ then the dynamical ball $B_{w}(x, n,\tilde {\delta })$ around x is mapped homeomorphically onto the ball $B_{\theta ^{n}(w)}(f^{n}_{w}(x), \tilde {\delta }).$ Moreover, for $z\in B_{w}(x, n, \tilde {\delta })$ and $f^{n}_{w}(z)\in B_{\theta ^{n}(w)}(f^{n}_{w}(x), \tilde {\delta })$ , we have
for each $1\leq k\leq n$ .
We point out that the proof of the lemma above is analogous to the proof of Lemma 5.5 in [Reference Arbieto, Matheus and Oliveira5] since in the definition of hyperbolic times we just replace the function $\|Df_{w}(\cdot )^{-1}\|$ by the Lipschitz function of the inverse branches $L_{w}(\cdot )$ .
Let $\mathcal {B}$ be the Borel $\sigma $ -algebra of $M.$ We say that $\xi $ is a $\mu $ -generating partition if
The next result states that every non-uniformly expanding measure admits a generating partition. See a proof of this in [Reference Arbieto, Matheus and Oliveira5].
Lemma 3.2. Given a non-uniformly expanding measure $\eta $ with exponent $c>0$ , consider $\tilde {\delta }=\tilde {\delta }(c)>0$ as in Lemma 3.1. Then any measurable partition $\mathcal {P}$ of M with diameter less than $\tilde \delta $ is an $\eta $ -generating partition.
3.3 Projective metrics
To finish this section we present the definition of projective metrics associated to convex cones. This theory was introduced by Birkhoff [Reference Birkhoff10] and provides an interesting way to obtain spectral properties of the transfer operator (see, for instance, [Reference Baladi8, Reference Liverani19]).
Consider a Banach space V. We say that a subset $\mathcal {C}\subset V\setminus \{0\}$ is a cone in V if $\mathcal {C}\cap (-\mathcal {C}) = \{0\} $ and $\unicode{x3bb} \cdot v\in \mathcal {C}$ for all $v\in \mathcal {C}$ , $\unicode{x3bb}>0$ . Moreover, a cone $\mathcal {C}$ is convex if $v,w\in \mathcal {C}$ and $\unicode{x3bb} ,\eta>0$ and we have $\unicode{x3bb} \cdot v + \eta \cdot w\in \mathcal {C}.$ The closure of a cone $\mathcal {C}$ , denoted by $\bar {C}$ , is the set
We say that a cone $\mathcal {C} $ is closed if $\bar {\mathcal {C}}= \mathcal {C} \cup \{0\}$ .
Consider a closed convex cone $\mathcal {C}$ . Given $v, w \in \mathcal {C}$ , define
where by convention $\sup \emptyset = 0$ and $\inf \emptyset = +\infty $ . It is straightforward to check that $A(v,w)$ is finite, $B(v,w)$ is positive and $A(v,w)\leq B(v,w)$ for all $v, w \in \mathcal {C}$ . We set
From the properties of A and B it follows that $\Theta (v,w)$ is well defined and takes values in $[0,+\infty ]$ . Notice that $\Theta (v,w)= 0$ if and only if $v= tw$ for some $t>0$ . Therefore, $\Theta $ defines a pseudo-metric in the cone $\mathcal {C}$ , and so it induces a metric on a projective quotient space of $\mathcal {C}$ . This metric is called the projective metric of $\mathcal {C}$ .
It is easy to verify that the projective metric depends monotonically on the cone: if $\mathcal {C}_{1} \subset \mathcal {C}_{2}$ are two convex cones in V, then $\Theta _{2}(v,w) \leq \Theta _{1}(v,w)$ for any $v,w \in \mathcal {C}_{1}$ , where $\Theta _{1}$ and $\Theta _{2}$ are the projective metrics in $\mathcal {C}_{1}$ and $\mathcal {C}_{2}$ , respectively.
In particular, if $V_{1},V_{2}$ are complete vector spaces and $L:{V}_{1} \to {V}_{2} $ is a linear operator such that $L(\mathcal {C}_{1}) \subset \mathcal {C}_{2}$ for $\mathcal {C}_{1}, \mathcal {C}_{2}$ convex cones in ${V}_{1}, {V}_{2}$ respectively, then $\Theta _{2}(L(v),L(w)) \leq \Theta _{1}(v,w)$ for any $v,w \in \mathcal {C}_{1}$ , where $\Theta _{1}$ and $\Theta _{2}$ are the projective metrics in $\mathcal {C}_{1}$ and $\mathcal {C}_{2}$ , respectively. The next result states that L will be a strict contraction if $L(\mathcal {C}_{1})$ has finite diameter in $\mathcal {C}_{2}$ .
Theorem 3.2. Let $\mathcal {C}_{1}$ and $\mathcal {C}_{2}$ be closed convex cones in the Banach spaces ${V}_{1}$ and ${V}_{2}$ , respectively. If $L:V_{1} \to V_{2} $ is a linear operator such that $L(\mathcal {C}_{1}) \subset \mathcal {C}_{2}$ and $\Delta = \mathrm { diam}_{\Theta _{2}}(L(\mathcal {C}_{1})) < \infty $ , then
In this work we will restrict our attention to cones of locally Hölder continuous observables. We prove that, applying the last result, the transfer operator is a contraction in this setting.
We fix $\delta>0$ as in §2 and we say that a function $\varphi :M\to \mathbb {R}$ is $(C, \alpha )$ -Hölder continuous in balls of radius $\delta $ if for some constant $C>0$ it follows that
Denote by $|\varphi |_{\alpha ,\delta }$ the smallest Hölder constant of $\varphi $ in balls of radius $\delta>0$ .
The next lemma states that every locally Hölder continuous function defined on a compact and connected metric space is Hölder continuous.
Lemma 3.3. Let M be a compact and connected metric space. Given $\delta> 0$ , there exists $m \geqslant 1$ (depending only on $\delta $ ) such that if $\varphi : M \to \mathbb {R}$ is $(C,\alpha )$ -Hölder continuous in balls of radius $\delta $ then it is $(Cm, \alpha )$ -Hölder continuous.
Proof. The compactness allows us to cover M with N balls of radius $\delta $ where N depends only on $\delta $ . Moreover, since M is connected, given $x, y\in M$ there are $z_{0}=x, z_{1}\ldots z_{N+1}=y$ satisfying $d(z_{i}, z_{i+1})\leq \delta $ and $d(z_{i}, z_{i+1})\leq d(x, y)$ for all $i=0,\ldots , N$ . Since $\varphi $ is $(C, \alpha )$ -Hölder continuous in balls of radius $\delta $ we have that
which implies that $\varphi $ is $(C\cdot m, \alpha )$ -Hölder continuous for $m=N + 1$ .
Notice that the same argument used in the lemma above gives an estimate for the Hölder constant of $\varphi $ in balls of radius $(1+r)\delta $ for $0 < r \leq 1.$ Indeed, let $r\in [0, 1]$ and $x, y\in M$ with $d(x, y)<(1+r)\delta $ . Since M is connected there exists $z\in M$ such that $d(x, z)=\delta $ and $d(z, y)<rd(x, z)$ . Thus,
Therefore, we conclude that if $\varphi : M \to \mathbb {R}$ is $(C, \alpha )$ -Hölder continuous in balls of radius $\delta $ then $\varphi $ is $(C(1+r^{\alpha }), \alpha )$ -Hölder continuous in balls of radius $(1+r)\delta $ for each $0 < r \leq 1.$
For each $k>0$ we consider the convex cone of locally Hölder continuous observables defined on M by
It follows by definition that $\mathcal {C}^{k_{1}}_{\delta } \subset \mathcal {C}^{k_{2},}_{\delta }$ if $k_{1} \leq k_{2}$ .
From Lemma 3.3 and from the definition of $|\varphi |_{\alpha ,\delta }$ we have that
and thus $\sup \varphi \leq \inf \varphi \cdot (1+ m(\mathrm {diam}\,M)^{\alpha }k)$ for any $\varphi \in \mathcal {C}^{k}_{\delta }.$
In the cone $\mathcal {C}^{k}_{\delta }$ of locally Hölder continuous observables we can give a more explicit expression for the projective metric. We refer the reader to [Reference Castro and Varandas12] for its proof.
Lemma 3.4. The projective metric $\Theta _{k}$ in the cone $\mathcal {C}^{k}_{\delta }$ is given by
where
and
In particular, we have that
From the expression for the projective metric in the cone $\mathcal {C}^{k}_{\delta }$ one can prove that its diameter is finite for k large enough; see [Reference Castro and Varandas12].
Proposition 3.1. For $0 <\gamma < 1$ , the cone $\mathcal {C}^{\gamma k}_{\delta }$ has finite diameter in $\mathcal {C}^{k}_{\delta }$ .
4 Reference measure
For $w\in X$ , let $f_{w}:M\rightarrow M$ be the fiber dynamics and $\phi _{w}:M\rightarrow \mathbb {R}$ be the potential. Let $\mathcal {L}_{w}:C^{0}(M)\rightarrow C^{0}(M)$ be the transfer operator associated to $(f_{w}, \phi _{w})$ , defined by
Consider also its dual operator $\mathcal {L}_{w}^{\ast }:[C^{0}(M)]^{\ast }\to [C^{0}(M)]^{\ast } $ which satisfies
We say that a probability measure $\nu _{w}\in \mathcal {M}^{1}(M)$ is a reference measure associated to $\unicode{x3bb} _{w}\in \mathbb {R}$ if $\nu _{w}$ satisfies
As in the deterministic case, by applying the Schauder–Tychonoff fixed point theorem, it is straightforward to prove the existence of a system of reference measures $\{\nu _{w}\}_{w\in X}$ where $\nu _{w}$ is associated to $\unicode{x3bb} _{w}$ given by
for $\mathbb {P}$ -almost every $w\in X.$ See [Reference Mayer, Skorulski and Urbański20] for details. In what follows we derive some properties of the reference measure.
The Jacobian of a measure $\eta $ with respect to f is a measurable function $J_{\eta }f$ such that
for any measurable set A where $f|_{A}$ is injective.
Lemma 4.1. The Jacobian of $\nu _{w}$ with respect to $f_{w}$ is given by $J_{\nu _{w}}f_{w} = \unicode{x3bb} _{w}e^{-\phi _{w}}$ . Moreover, $\nu _{w}$ is an open measure. In particular, $\mathrm {supp}(\nu _{w})=M$ .
Proof. Let $A\subset M$ be a measurable set such that $f_{w}|_{A}$ is injective. Notice that, for any bounded sequence $\{\zeta _{n} \}\in C^{0}(M)$ which converges to the characteristic function $\mathcal {X}_{A}$ of A, we have
Since $\int _{M} \sum _{f_{w}(z)=y} \zeta _{n}(f^{-1}_{w}(y))\,d\nu _{\theta (w)}(y)$ converges to $\int _{M} \mathcal {X}_{A}(f^{-1}_{w}(y))\,d\nu _{\theta (w)}(y)$ and $\int _{M} \mathcal {X}_{A}(f^{-1}_{w}(y))\,d\nu _{\theta (w)}(y)=\int _{M}\mathcal {X}_{f_{w}(A)} \,d\nu _{\theta (w)}=\nu _{\theta (w)}(f_{w}(A))$ we conclude that
Moreover, by induction, we obtain for every $n\in \mathbb {N}$ that
where $\unicode{x3bb} ^{n}_{w}= \unicode{x3bb} _{\theta ^{n-1}(w)} \unicode{x3bb} _{\theta ^{n-2}(w)}\cdot \cdot \cdot \unicode{x3bb} _{\theta (w)}\unicode{x3bb} _{w}$ .
Now we prove that $\nu _{w}$ is an open measure. By contradiction, suppose the existence of some non-empty open set $U_{w}\subset M$ such that $\nu _{w}(U_{w})=0$ . By the exactness assumption, we can take $\tilde {n}\in \mathbb {N}$ such that $f^{\tilde {n}}_{w}(U_{w})=M$ . Partitioning $U_{w}$ into mensurable subsets $U_{w,1}\ldots U_{w,k}$ where $f_{w}^{\tilde {n}}|_{U_{w,j}}$ is injective for $j=1\ldots k$ , we have
which is a contradiction. This completes the proof.
In the next proposition we show that the family $\{\nu _{w}\}_{w}$ satisfies a Gibbs property at hyperbolic times.
Proposition 4.1. Let n be a hyperbolic time for $(w, x)$ . For every $0<\varepsilon \leq \tilde \delta $ there exist $K_{\varepsilon }(w)>0$ and $0<\gamma _{\varepsilon }(\theta ^{n}(w))\leq 1$ such that, for all $y\in B_{w}(x,n, \varepsilon )$ ,
where $S_{n}\phi _{w}(y)=\sum ^{n-1}_{j=0}\phi _{\theta ^{j}(w)}( f^{j}_{w}(y))$ and $\unicode{x3bb} _{w}^{n}=\unicode{x3bb} _{w}\unicode{x3bb} _{\theta (w)}\cdots \unicode{x3bb} _{\theta ^{n-1}(w)}.$
Proof. Fix $0<\varepsilon \leq \tilde \delta $ . From Lemma 3.1 and condition (IV) we get
for every $z, y\in B_{w}(x,n, \varepsilon ).$ By once again applying Lemma 3.1 we know that $f_{w}^{n}$ maps homeomorphically $B_{w}(x,n, \varepsilon )$ into the ball $B_{\theta ^{n}(w)}(f_{w}^{n}(x),\varepsilon )$ . Hence, since the Jacobian of $\nu _{w}$ is bounded away from zero and infinity we can write
where $\gamma _{\varepsilon }(\theta ^{n}(w))$ depends only on the radius $\varepsilon $ of the ball $B_{\theta ^{n}(w)}(f_{w}^{n}(x),\varepsilon )$ . Therefore, for every $y\in B_{w}(x,n, \varepsilon )$ it follows that
Applying the same argument, we have that
which completes the proof.
Remark 4.1. It is possible to obtain a lower bound for $\gamma _{\varepsilon }(\theta ^{n}(w))$ . Indeed, from hypothesis we may find $\tilde {n}=\tilde {n}(w, \varepsilon )$ such that $f_{\theta ^{n}(w)}^{\tilde {n}}(B_{\theta ^{n}(w)}(f_{w}^{n}(x),\varepsilon ))=M$ and by definition of Jacobian it follows that
Thus, $ e^{\tilde {n}\inf \phi _{\theta ^{n}(w)}-\log \unicode{x3bb} ^{\tilde {n}}_{\theta ^{n}(w)}}\leq \gamma _{\varepsilon }(\theta ^{n}(w)).$ Since $\tilde {n}$ depends only on $w\in X$ and $\varepsilon>0$ we conclude that $\gamma _{\varepsilon }(\theta ^{n}(w))$ is uniformly bounded.
Consider $c>0$ given by condition (VI). Given $w\in X$ , let $H_{w}\subset M$ be the subset of M such that $(w, x)$ has infinitely many hyperbolic times, that is,
We next prove that $\nu _{w}(H_{w})=1$ for $\mathbb {P}$ -almost every $w\in X.$
Recall that we fix $\varepsilon _{\phi }>0$ small satisfying $\varepsilon _{\phi }<\inf _{w}(\log (\deg f_{w}) - \log q_{w})$ . In view of (IV) we may find $0<\varepsilon _{0} <\varepsilon _{\phi }$ such that
Let $\mathcal {P}$ be a partition of M with cardinality $\# \mathcal {P} = k$ . We suppose without loss of generality that the set $\mathcal {A}_{w}$ is contained in the first $q_{w}$ elements of $\mathcal {P}$ for all $w\in X$ . Consider the numbers
These numbers are well defined since we assume that $\deg (F)= \sup _{w} \deg (f_{w}) < \infty $ .
For $\rho \in (0,1)$ and $n\in \mathbb {N}$ let $I(\rho , n)$ be the set of itinerates
and consider
Lemma 4.2. [Reference Varandas and Viana29, Lemma 3.1]
Given $\varepsilon> 0$ , there exists $\rho _{0} \in (0,1)$ such that $C_{\rho } < \log \hat {q} + \varepsilon $ for every $\rho \in (\rho _{0} , 1)$ .
Proof. Notice that $ \# I(\rho , n) \leq \sum _{k = [\rho n]}^{n} \binom {n}{k} q_{w} q_{\theta (w)}\cdot \cdot \cdot q_{\theta ^{k-1}(w)}p_{w} p_{\theta (w)}\cdot \cdot \cdot p_{\theta ^{n -(k-1)}(w)}. $ By applying Stirling’s formula we have
Thus, there exist $C_{1}$ and $t> 0$ such that $ \# I(\rho , n) \leq C_{1} \exp {(2t(1 - \rho )n)} \ \hat {q}^{n} \hat {p}^{(1-\rho )n}. $ Taking the limit as n goes to infinity, we have
for any $\rho $ close enough to 1.
From this lemma we can fix $\rho < 1$ such that
Recalling equation (8) and the definition of $\unicode{x3bb} _{w}$ in (6), we have that
Now, using Lemma 4.1, we obtain that
Proposition 4.2. We have $\nu _{w}(H_{w})=1$ for almost every $w\in X$ .
Proof. Given $n\in \mathbb {N}$ , denote by $B_{w}(n)$ the set of points $x\in M$ whose frequency of visits to $\{\mathcal {A}_{\theta ^{j}(w)}\}_{0\leq j \leq n-1}$ up to time n is at least $\rho $ , that is,
Let $\mathcal {P}^{(n)}$ be the partition $\bigvee _{j=0}^{n-1}(f_{w}^{j})^{-1}\mathcal {P}$ . We cover $B_{w}(n)$ by elements of $\mathcal {P}^{(n)}$ and since $f^{n}_{w}$ is injective on every $P\in \mathcal {P}^{(n)}$ , we may use (9) to obtain
Thus,
Since we can assume $\hat {q}< \bar {q}e^{\varepsilon _{0} /2}$ it follows that
Hence, the measure $\nu _{w}( B_{w}(n))$ decreases exponentially fast as n goes to infinity. Applying the Borel–Cantelli lemma, we conclude that $\nu _{w}$ -almost every point belongs to $B_{w}(n)$ for at most finitely many values of n. Then, in view of (VI) we obtain for n large enough that
which proves that $\nu _{w}$ -almost every point has infinitely many hyperbolic times.
Notice that from the last proposition, and recalling that $\nu _{w}$ is an open measure, we conclude that $H_{w}$ is dense in M.
5 Transfer operator
Here we prove Theorems A and B. We use the projective metric approach to show that the transfer operator is a contraction in some cone of locally Hölder continuous functions. This contraction implies the existence of the invariant family $\{h_{w}\}_{w}$ uniformly bounded away from zero and infinity. Recalling the reference measure $\nu _{w}$ constructed in the previous section, we define the probability measure $\mu _{w}:=h_{w}\nu _{w}$ . From the exponential approximation of functions in the cone to the family $\{h_{w}\}_{w}$ we derive that $\mu _{w}$ has an exponential decay of correlations.
5.1 Invariant family
For the construction of the invariant family $\{h_{w}\}_{w}$ we follow the ideas of Castro and Varandas [Reference Castro and Varandas12].
Let $\delta>0$ be as in §2 and consider for each $k>0$ the cone of locally Hölder continuous functions
Since the cone does not depend on w, we denote this by $ \mathcal {C}^{k}_{\delta }$ . The next proposition shows its invariance by the transfer operator.
Proposition 5.1. For every $w\in X$ , there exists $0< \gamma _{w} < 1$ such that
for some positive constant k large enough.
Proof. For any $\varphi \in \mathcal {C}^{k}_{\delta }$ we will show that
Given $y, z \in M$ satisfying $d(y, z)< \delta $ , we denote by $y_{j}, z_{j}$ , $1\leq j\leq \deg (f_{w})$ , the respective preimages under $f_{w}$ . Note that for any continuous function $\varphi $ we have
From the definition of $\mathcal {L}_{w}$ and the constant $|\mathcal {L}_{w}(\varphi )|_{\alpha , \delta }$ we obtain
By equation (11) the last inequality is less than or equal to
Recall that, by hypothesis, we can take some $x\in M$ and neighborhood $U_{x}\subset M$ such that $y, z\in B_{\theta (w)}(f_{w}(x), \delta )$ and so the inverse branches satisfy
Moreover, we are assuming that every point has $q_{w}$ preimages in the region $\mathcal {A}$ and $p_{w}$ preimages in the expanding region $M\setminus \mathcal {A}$ where $L_{w}\leq \sigma _{w}^{-1}$ . Since, by Lemma 3.3, $\varphi $ is $((1+ (L_{w}-1)^{\alpha })|\varphi |_{\alpha ,\delta }, \alpha )$ -Hölder continuous in balls of radius $L_{w}\delta $ we conclude that the sum (12) is bounded from above by
And this expression is equal to
Using inequality (5), the definition of cone and condition (IV), it follows that the sum above is less than or equal to
By hypothesis (condition (V)), there exists some positive constant $0<\gamma _{w}< 1$ such that the previous sum is bounded from above by $\gamma _{w} k$ . This finishes the proof.
From the last proposition we have the invariance of the cone $\mathcal {C}^{k}_{\delta }.$ Since this cone has finite diameter, according to Proposition 3.1, we can apply Theorem 3.2 to conclude the next result.
Proposition 5.2. For every $w\in X$ the operator $\mathcal {L}_{w}$ is a contraction in the cone $\mathcal {C}^{k}_{\delta }$ , that is, writing $\Delta _{w} = \mathrm {diam}_{\Theta _{k}}(\mathcal {C}^{\gamma _{w} k}_{\delta })>0$ , it follows that
Since we assume in condition (V) the existence of $\gamma \in (0,1)$ such that $\gamma _{w}\leq \gamma $ for all $w\in X$ we conclude that
where $\Delta = \sup _{w} (\Delta _{w})\leq \mathrm {diam}_{\Theta _{k}}(\mathcal {C}^{\gamma k}_{\delta }).$
Let $\{\nu _{w}\}_{w}$ be the family of reference measures and $\unicode{x3bb} _{w}=\nu _{\theta (w)}(\mathcal {L}_{w}(1))$ . The contraction in the cone allows us to prove the existence of the family $\{h_{w}\}_{w}$ invariant by the transfer operator.
Proposition 5.3. For almost $w\in X$ there exists a Hölder continuous function $h_{w}:M\to \mathbb {R}$ bounded away from zero and infinity satisfying $\mathcal {L}_{w}h_{w}=\unicode{x3bb} _{w} h_{\theta (w)}.$
Proof. Consider the normalized operator $\hat {\mathcal {L}}_{w}:=\unicode{x3bb} ^{-1}_{w}\mathcal {L}_{w}$ and define the sequence $(\varphi _{n})_{n}$ by $\varphi _{n}:=\hat {\mathcal {L}}^{n}_{\theta ^{-n}(w)}(\mathbf {1})$ where
for each $n\geq 0$ . By definition of conformal measure we have
Hence, each term $\varphi _{n}$ satisfies $\sup \varphi _{n}\geq 1$ and $\inf \varphi _{n}\leq 1$ . Since $\mathbf {1}\in \mathcal {C}^{k}_{\delta }$ and $\mathcal {C}^{k}_{\delta }$ is invariant, it follows that $\varphi _{n}\in \mathcal {C}^{k}_{\delta } $ and so, applying inequality (5), we obtain that the sequence $(\varphi _{n})_{n}$ is uniformly bounded away from zero and infinity by
where $R=(1 + mk \mathrm {diam}(M)^{\alpha })$ . Moreover, as $\varphi _{n}$ is C-Hölder continuous in balls of radius $\delta $ , by Lemma 3.3 we obtain that $\varphi _{n}$ is a $Cm$ -Hölder continuous function.
Next we prove that $(\varphi _{n})_{n}$ is a Cauchy sequence in the $C^{0}$ -norm. From Proposition 5.2, for every $m, l\geq n$ the projective metric satisfies
Recalling the expression for the projective metric $\Theta _{k}(\varphi _{m}, \varphi _{l})= \log ({B_{k}(\varphi _{m}, \varphi _{l})}/{A_{k}} {(\varphi _{m}, \varphi _{l})})$ , we apply Lemma 3.4 to obtain
Thus, for all $m, l\geq n$ , we have
which proves that $(\varphi _{n})_{n}$ is a Cauchy sequence. Hence, $(\varphi _{n})_{n}$ converges uniformly to a function $h_{w}:M\to \mathbb {R}$ in the cone $\mathcal {C}^{k}_{\delta }$ satisfying $\int h_{w} \,d\nu _{w} =1$ for almost every $w\in X$ . In particular, this function is Hölder continuous and uniformly bounded away from zero and infinity. To complete the proof of the proposition, we will show that $\mathcal {L}_{w}h_{w}=\unicode{x3bb} _{w} h_{\theta (w)}.$ Consider the sequence
By what we have proved above, $(\tilde {\varphi }_{n, w})$ converges uniformly to $h_{w}$ for almost every $w\in X.$ From the continuity of $\mathcal {L}_{w}$ we obtain
Since $\hat {\mathcal {L}}^{n}_{\theta ^{-n}(\theta (w))}(\mathbf {1})$ is uniformly bounded we conclude that $\hat {\mathcal {L}}_{w}(h_{w})=h_{\theta (w)}.$
From the proof of the previous proposition we conclude that the family $\{h_{w}\}_{w\in X}$ is uniquely determined. Moreover, every $h_{w}$ satisfies
where $R=(1 + mk\, \mathrm {diam}(M)^{\alpha })$ .
5.2 Measurability
In the previous subsection we proved the existence of a family $\{h_{w}\}_{w}$ invariant under the action of the transfer operator. Here we prove that this family is measurable as well as the family $\{\nu _{w}\}_{w}$ . Moreover, defining the probability measure $\mu _{w}:=h_{w}\nu _{w}$ , we also prove that $\mu _{w}$ has an exponential decay of correlations and that the family $\{\mu _{w}\}_{w}$ is F-invariant.
The next proposition states an exponential approximation of functions in the cone to the invariant family $\{h_{w}\}_{w}$ . This is the main ingredient in the proof of the exponential decay of correlations.
Proposition 5.4. For almost $w\in X$ there exist constants $K>0$ and $0<\tau <1$ such that for every $\varphi \in \mathcal {C}^{k}_{\delta }$ satisfying $\int \varphi \,d\nu _{w} = 1$ we have that
where $\hat {\mathcal {L}}_{w}=\unicode{x3bb} _{w}^{-1}\mathcal {L}_{w}$ is the normalized operator.
Proof. Given $\varphi \in \mathcal {C}^{k}_{\delta }$ satisfying $\int \varphi \,d\nu _{w} = 1$ , we have for every $n \geq 1$ ,
Since $h_{w} \in \mathcal {C}^{k}_{\delta }$ also satisfies $\int h_{w} \,d\nu _{w} = 1$ , we derive for every $n\geq 1$ that
Recalling that $\hat {\mathcal {L}}^{n}_{w}(h_{w})=h_{\theta ^{n}(w)}$ for all $n\geq 1$ , we apply the same projective metric argument used in the proof of Proposition 5.3 to obtain
Let $\mu _{w}$ be the probability measure defined by $\mu _{w}:=h_{w}\nu _{w}.$ From the last proposition we derive the proof of Theorem B.
Theorem B. For almost every $w\in X,$ the probability measure $\mu _{w}$ has exponential decay of correlations for Hölder continuous observables: there exists $0<\tau < 1$ such that for any $\varphi \in L^{1}(\mu _{\theta ^{n}(w)})$ and $\psi \in C^{\alpha }(M) $ there exists a positive constant $K(\varphi , \psi ) $ satisfying for all $n\geq 1$ that
Proof. Given $\varphi \in L^{1}(\mu _{\theta ^{n}(w)})$ and $\psi \in C^{\alpha }(M)$ , we suppose without loss of generality that $\int \psi \,d\mu _{w} = 1$ . Let $\hat {\mathcal {L}}_{w}=\unicode{x3bb} ^{-1}_{w}\mathcal {L}_{w}$ be the normalized operator. As a first case we consider $\psi \cdot h_{w}$ in the cone $\mathcal {C}^{k}_{\delta }$ for k large enough. Recalling that $\mu _{w}=h_{w}\nu _{w}$ and that $\mathcal {L}^{*}_{w}\nu _{\theta (w)}=\unicode{x3bb} _{w}\nu _{w}$ , we have
Since $\psi \cdot h_{w}\in \mathcal {C}^{k}_{\delta }$ and $\int \psi \cdot h_{w} \,d\nu _{w}=\int \psi \,d\mu _{w} = 1$ we can apply Proposition 5.4 to conclude the existence of constants $K>0$ and $0<\tau <1$ such that
For the general case we write $\psi \cdot h_{w}= g$ where
Therefore, $g^{\pm }\in \mathcal {C}^{k}_{\delta }.$ From the previous estimates on $g^{\pm }$ and by linearity the proposition holds.
Now we state the measurability of the families $\{\nu _{w}\}_{w}$ and $\{h_{w}\}_{w}$ . We start by observing that for almost every $w\in X$ and every continuous function $g\in C^{0}(M)$ we have
where $\hat {\mathcal {L}}$ is the normalized operator $\hat {\mathcal {L}}_{w} = \unicode{x3bb} ^{-1}_{w} \mathcal {L}_{w}.$ Indeed, we can suppose that the function $g_{w}$ is Hölder continuous because any continuous function is approximated by such functions. Moreover, following the proof of Theorem 5.2, we can just consider the case $ g \cdot h_{w} \in \mathcal {C}^{k}_{\delta }$ for k large enough. We have that
By Proposition 5.4 the convergence in (13) follows. We use this on the next result.
Lemma 5.1. Let $\unicode{x3bb} _{w}=\nu _{\theta (w)}(\mathcal {L}_{w}(1)).$ The family $\{\nu _{w}\}$ is uniquely determined by
Moreover, the map $w\mapsto \nu _{w}(g_{w})$ is measurable for any $g\in \mathbb {L}^{1}_{\mathbb {P}}(X, C^{0}(M))$ .
Proof. Fix $w\in X$ and let $(x_{n})$ be a sequence of points in M. Define the probability
Since $\nu _{w}$ satisfies the condition $\mathcal {L}_{w}^{\ast }\nu _{\theta (w)}=\unicode{x3bb} _{w}\nu _{w} $ we can apply the convergence (13) to conclude that for almost every $w\in X$ and any continuous function $g_{w}$ we have
The convergence of $\nu _{w,n}\stackrel {w*}\longrightarrow \nu _{w}$ therefore follows. Since the sequence $(x_{n})$ was arbitrary the uniqueness of the family $\{\nu _{w}\}$ is proved. Moreover, the equality
implies the measurability of $w\mapsto \nu _{w}(g_{w})$ since the transfer operator is measurable.
The lemma above enables us to define the probability measure $\nu $ on the Borel sets of $X\times M$ by
Let $c>0$ be given by condition (VI) and let $H\subset X\times M$ be the non-uniformly expanding set defined in (⋆). As in §4, for $w\in X$ consider
By Proposition 4.2 for almost every $w\in X$ we have $\nu _{w}(H_{w})=1$ . Thus, we conclude that $\nu (H)=1$ , that is, $\nu $ is a non-uniformly expanding measure.
Notice that since $\unicode{x3bb} _{w}=\nu _{\theta (w)}(\mathcal {L}_{w}(1))$ we have that the map $w\mapsto \unicode{x3bb} _{w}\in \mathbb {R}$ is measurable. Recalling that for almost every $w\in X$ the function $h_{w}$ is given by
we deduce the measurability of the map $(w, y)\mapsto h_{w}(y)$ from the measurability of $\unicode{x3bb} _{w}$ and the transfer operator. From this result the probability measure $\mu _{F, \phi }\in \mathcal {M}_{\mathbb {P}}(X\times M)$ is well defined by the formula
In order to prove the F-invariance of $\mu _{F, \phi }$ we first observe that
Therefore, for every integrable function we get
This finishes the proof of Theorem A.
6 Equilibrium states
In this section we prove that the measure $\mu _{F, \phi }$ constructed above is an equilibrium state for $(F|_{\theta }, \phi )$ . Moreover, we show that any ergodic non-uniformly expanding equilibrium state has disintegration absolutely continuous with respect to the system of reference measures. From this we derive the uniqueness.
As a first step, in the next proposition we obtain an upper bound for the topological pressure of the random dynamical system.
Proposition 6.1. For any potential $\phi $ satisfying condition (IV) we have that
Proof. Fix $w\in X$ such that $H_{w}$ is not empty (see definition in §4) and let $\varepsilon>0$ be small. Since every point $x\in H_{w}$ has infinitely many hyperbolic times, for $N>1$ large enough we have
where $H_{n}=H_{n}(w)$ denotes the set of points that have n as a hyperbolic time. From Lemma 3.1 each $f^{n}(B_{w}(x,n, \varepsilon ))$ is the ball $B_{\theta ^{n}(w)}(f_{w}^{n}(x),\varepsilon )$ in M, thus by applying the Besicovitch covering lemma it is straightforward to check that there exists a countable family $F_{n}\subset H_{n}$ such that every point $x\in H_{n}$ is covered by at most $d=d(\dim (M))$ dynamical balls $B_{w}(x,n, \varepsilon )$ with $x\in F_{n}.$ Therefore,
is a countable open covering of $H_{w}$ by dynamic balls with diameter less than $\varepsilon>0$ . Recalling that $H_{w}$ is dense on M and the closure of $B_{w}(x,n, \varepsilon )$ is a subset of $B_{w}(x,n, 2\varepsilon )$ , we have that
Hence,
is a countable open covering of M by dynamic balls with diameter less than $2\varepsilon $ .
Let $\beta> \int \log \unicode{x3bb} _{w}\,d\mathbb {P}(w).$ By the definition of topological pressure given in §3 and by applying Lemma 4.1 to each element in $\mathcal {G}_{N}$ we obtain
As from Remark 4.1 the variable $\gamma ^{-1}_{2\varepsilon }(\theta ^{n}(w))$ is uniformly bounded and by ergodicity we have $\lim _{n\to \infty }({1}/{n})\sum _{i=0}^{n-1}\log \unicode{x3bb} _{\theta ^{i}(w)}=\int \log \unicode{x3bb} _{w}\,d\mathbb {P}(w)$ for almost every $w\in X$ , we obtain as N goes to infinity that
for any $\beta>\int \log \unicode{x3bb} _{w}\,d\mathbb {P}(w)$ and $\varepsilon>0$ small. Thus, we necessarily have
Since this is true for $\mathbb {P}$ -almost every $w\in X$ , we prove the proposition.
In the next subsection we will prove that $P_{F|_{\theta }}(\phi )=\int \log \unicode{x3bb} _{w}\,d\mathbb {P}(w).$
6.1 Existence
Consider the family $\{\nu _{w}\}_{w\in X}$ of reference measures and $\{h_{w}\}_{w\in X}$ as in Theorem A. For each $w\in X$ let $\mu _{w}$ be the probability measure defined by $\mu _{w}=h_{w}\nu _{w}$ . Recalling that the Jacobian of $\nu _{w}$ is $J_{\nu _{w}}f_{w}=\unicode{x3bb} _{w}e^{-\phi _{w}}$ , it is easy to verify that the Jacobian of $\mu _{w}$ relative to $f_{w}$ is given by
Consider the probability measure $\mu _{F, \phi }$ whose disintegration is $\{\mu _{w}\}_{w}$ , that is,
for every continuous function $g:X\times M\to \mathbb {R}$ . As we saw in §5.2, $\mu _{F,\phi }$ is F-invariant. In the next proposition we show its ergodicity.
Proposition 6.2. The probability measure $\mu _{F, \phi }$ is ergodic.
Proof. Given a F-invariant set $A\subset X\times M$ , for each $w\in X$ denote by $A_{w}$ the set $A_{w}=\{z\in M; (w, z)\in A\}.$ The F-invariance of A implies that $f_{w}^{-1}(A_{\theta (w)})=A_{w}.$ Consider $X_{0}=\{w\in X; \mu _{w}(A_{w})>0\}$ . It is straightforward to check that $X_{0}$ is a $\theta $ -invariant subset of X. Since $\theta $ is ergodic with respect to $\mathbb {P}$ , we will obtain the ergodicity of $\mu _{F, \phi }$ by showing that for almost every $w\in X_{0}$ we have $\mu _{w}(A_{w})=1$ , if $\mathbb {P}(X_{0})>0.$
Let $\varphi _{w}$ be the characteristic function of $A_{w}$ , that is, $\varphi _{w}=\mathbf {1}_{A_{w}}$ . Notice that $\varphi _{\theta ^{n}(w)}\circ f_{w}^{n}=\varphi _{w}$ holds $\mathbb {P}$ -almost everywhere. Given $\psi _{w}\in L^{1}(\mu _{w})$ such that $\int \psi _{w}\, d\mu _{w}=0$ , it follows from the decay correlation property of $\mu _{w}$ (Theorem B) that
And thus $\int _{A_{w}} \psi _{w}\, d\mu _{w}=0$ for any $\psi _{w}\in L^{1}(\mu _{w})$ satisfying $\int \psi _{w}\, d\mu _{w}=0$ . This proves that $\mu _{w}(A_{w})=1$ for $\mathbb {P}$ -almost every $w\in X_{0}$ which finishes the proof.
In §5.2 we observe that the measure $\nu $ defined by
is non-uniformly expanding. Since $\mu _{F, \phi }$ is absolutely continuous with respect to it, we have that $\mu _{F, \phi }$ is also a non-uniformly expanding measure. In particular, by Lemma 3.2, it admits a generating partition. Thus, we can use the random Rokhlin formula to express the entropy of $\mu _{F, \phi }$ in terms of its Jacobian.
Theorem 6.1. (Random Rokhlin formula)
Let $\mu \in \mathcal {M}_{\mathbb {P}}(F)$ be an ergodic measure which admits a $\mu $ -generating partition. Then
where $J_{\mu _{w}}f_{w}$ denotes the Jacobian of $f_{w}$ relative to $\mu _{w}$ .
The reader can consult [Reference Przytycki and Urbanski23, Theorem 1.9.7] for a proof of the last result. Now we are ready to prove that $\mu _{F, \phi }$ is an equilibrium state for $(F|\theta , \phi )$ . We have
From the F-invariance of $\mu _{F, \phi }$ we derive that
Thus, we can write
Applying the variational principle (1) and Proposition 6.1, it follows that
which implies, in particular, that $P_{ F|\theta }(\phi )=\int _{X}\log \unicode{x3bb} _{w}\, d\mathbb {P}(w)$ and so $\mu _{F, \phi }$ is an equilibrium state.
6.2 Uniqueness
So far we have proved the existence of an equilibrium state for $(F|_{\theta }, \phi )$ . Here we prove uniqueness in the set of non-uniformly expanding measures.
Let $\eta $ be an ergodic non-uniformly expanding equilibrium state for $(F|_{\theta }, \phi )$ . We prove that the disintegration of $\eta $ is absolutely continuous to the reference measure. For this we use the following remark from the basic calculus.
Remark 6.1. Jensen’s inequality
Given positive numbers $p_{i}>0$ and $q_{i}>0$ , $i=1, \ldots , n$ , such that $\sum _{i=1}^{n}p_{i}=1$ , we have that $\sum _{i=1}^{n}p_{i}\log q_{i}\leq \log (\sum _{i=1}^{n}p_{i} q_{i})$ , with equality holding if and only if the $q_{i}$ are equal.
Proposition 6.3. Consider any ergodic non-uniformly expanding equilibrium state $\eta \in \mathcal {M}_{\mathbb {P}}(F)$ of $(F|_{\theta }, \phi )$ and let $(\eta _{w})_{w}$ be its disintegration. Then, for almost $w\in X$ , $\eta _{w}$ is absolutely continuous with respect to $\nu _{w}$ .
Proof. We begin by proving that for almost every $w\in X$ the Jacobian of $\eta _{w}$ is given by
Indeed, since $\eta $ is an ergodic non-uniformly expanding equilibrium state we can apply the Rokhlin formula to obtain that
Recalling that $h_{w}$ is a bounded function and that the Jacobian of $\mu _{w}$ is given by $J_{{\mu }_{w}}{f_{w}}\;=\unicode{x3bb} _{w} e^{-{\phi _{w}}}\cdot {h_{\theta (w)}}\circ {f_{w}}/{h_{w}}$ , we have
From the definition of Jacobian we can write
Take $p_{i}=J_{\eta _{w}}f_{w}^{-1}(z_{i})$ and $q_{i}=J_{\eta _{w}}f_{w}(z_{i})/J_{\mu _{w}}f_{w}(z_{i})$ where the $z_{i}$ are the preimages of y. Since $(f_{w})_{\ast }\eta _{w}=\eta _{\theta (w)}$ we have $\sum _{i=1}^{\deg (f_{w})}{p_{i}}=\sum _{z=f_{w}^{-1}(y)}J_{\eta _{w}}f_{w}^{-1}(z)=1$ for $\eta _{\theta (w)}$ -almost every $y\in M.$ Therefore, we can apply Remark 6.1 to conclude that
for $\eta _{\theta (w)}$ -almost every $y\in M$ . Recalling the inequality (15), we obtain that
Thus, from the second part of Remark 6.1, the values $q_{i}=J_{\eta _{w}}f_{w}(z_{i})/J_{\mu _{w}}f_{w}(z_{i})$ must be the same for all $z_{i}\in f_{w}^{-1}(y)$ in a full $\eta _{\theta (w)}$ -measure set. In other words, for every $y\in M$ on the preimage of a full $\eta _{\theta (w)}$ -measure set, we have
To finish the proof of the proposition we observe that ${1}/{h_{w}}\cdot \eta _{w}$ is a reference measure associated to $\unicode{x3bb} _{w}$ for the dual transfer operator:
From the uniqueness given by Theorem A we conclude that ${1}/{h_{w}}\cdot \eta _{w}$ is equivalent to $\nu _{w}$ and thus $\eta _{w}$ is absolutely continuous to the latter.
Finally, we prove the uniqueness of the equilibrium state associated to $(F|_{\theta }, \phi ).$ Suppose that there exist two ergodic equilibrium states $\mu $ and $\eta $ . Let $(\mu _{w})_{w\in X}$ and $(\eta _{w})_{w\in X}$ be the disintegration of $\mu $ and $\eta $ , respectively.
By the proposition above we have that $\mu _{w}$ and $\eta _{w}$ are equivalent measures. From the Radon–Nikodym theorem we know that there exists a mensurable function $q_{w}: M\rightarrow \mathbb {R}$ such that $\mu _{w}=q_{w}\eta _{w}$ for every $w\in X$ . Consider $q: X\times M \rightarrow \mathbb {R}$ , defined by $q(w,x)=q_{w}(x)$ . Given a measurable set $E\subset X\times M$ , consider $E_{w}\subset M$ the intersection $E_{w}=E\cap M$ . Then we have that
Moreover, it follows from the F-invariance of $\mu $ and $\eta $ that
Since the Radon–Nikodym derivative is essentially unique, we conclude that $q=q\circ F$ at $\eta $ -almost every point. By ergodicity we have that q is constant everywhere and thus $\mu =\eta .$
6.3 Positive Lyapunov exponents
The main tool in the proof of Proposition 6.3 is the existence of a generating partition for the equilibrium state. In the context of random dynamical systems generated by non-uniformly expanding maps, the existence of generating partitions for ergodic measures with Lyapunov exponents bounded away from zero was proved by Bilbao and Oliveira in [Reference Bilbao and Oliveira9]. Therefore, in this setting we can also apply our Proposition 6.3 to obtain uniqueness of equilibrium states. This is our goal now.
Consider a compact and connected Riemann manifold $M^{d}$ of dimension d. Let $F:X\times M\to X\times M$ be the skew-product $(\theta (w), f_{w}(x))$ generated by $C^{1}$ local diffeomorphisms $f_{w}: M\to M$ satisfying conditions (I)–(III). For $1\leq k\leq d-1$ , define
where $\Lambda ^{k}$ is the kth exterior product. We suppose that, for some $\varepsilon>0$ ,
For potentials $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X, C^{\alpha }(M))$ satisfying conditions (IV) and (V) such that for almost every $w\in X$ the inequality $\sup \phi _{w}-\inf \phi _{w}<\varepsilon \int \log \deg (f_{w})\,d \mathbb {P}(w)$ holds, we obtain the following result.
Corollary 6.1. There exists only one equilibrium state associated to $(F|_{\theta }, \phi )$ .
Proof. Let $\eta \in \mathcal {M}_{\mathbb {P}}(F|\theta )$ be an ergodic equilibrium state for $(F|_{\theta }, \phi )$ . Denote by $\unicode{x3bb} _{1}(w,x)\leq \cdots \leq \unicode{x3bb} _{d}(w,x)$ the Lyapunov exponents of $\eta $ at $(w,x)$ . We claim that they are bigger than $\beta (F)>0$ . If not, by applying the random version of the Marquis–Ruelle inequality [Reference Liu18, Theorem 2.4] we have
Thus, for potentials such that $\sup \phi _{w}-\inf \phi _{w}<\varepsilon \int \log \deg (f_{w})\,d \mathbb {P}(w)$ it follows that
which is a contradiction. Therefore, the Lyapunov exponents of $\eta $ are bigger than $\beta (F)$ , and so $\eta $ admits generating partitions with small diameter. Applying the proof of Proposition 6.3, we have that $\eta $ is absolutely continuous with respect to $\nu $ , and thus the uniqueness is proved.
7 Equilibrium stability
Consider a sequence $(F_{k}, \phi _{k})$ in $\mathcal {H}$ converging to $(F, \phi )$ . For each $k\in \mathbb {N}$ , let $\mu _{k}$ be the non-uniformly expanding equilibrium state of $(F_{k}, \phi _{k})$ . We will prove that any accumulation point $\mu $ of the sequence $(\mu _{k})$ is the non-uniformly expanding equilibrium state of $(F, \phi ).$
For each $k\in \mathbb {N}$ , consider the disintegration $\{\mu _{k, w}\}_{w\in X}$ of $\mu _{k}$ . From Theorem A, we know that $\mu _{k, w}=h_{k, w}\nu _{k, w}$ where $h_{k, w}$ and $\nu _{k, w}$ satisfy
We point out that for any $\psi \in C^{\alpha }(M)$ we have $\mathcal L_{k, w}(\psi )$ converging to $\mathcal L_{w}(\psi )$ in $C^{0}$ -norm, a proof of which is given in [Reference Alves, Ramos and Siqueira3].
Let $\unicode{x3bb} _{w}$ , $\nu _{w}$ and $h_{w}$ be as in Theorem A applied to $(F, \phi ).$ The main step in the proof of the equilibrium stability is the following proposition.
Proposition 7.1. For almost all $w\in X$ we have the convergence
Proof. Recalling that for each $k\in \mathbb {N}$ we have
the sequence $(\unicode{x3bb} _{k, w})$ admits some accumulation point $\bar {\unicode{x3bb} }_{w}.$ Moreover, taking subsequences, if necessary, there exist probability measures $\bar {\nu }_{w}$ and $\bar {\nu }_{\theta (w)}$ such that $\nu _{k, w}\stackrel {w*}\longrightarrow \bar {\nu }_{w}$ and $\nu _{k, \theta (w)}\stackrel {w*}\longrightarrow \bar {\nu }_{\theta (w)}$ . We will prove that $\mathcal {L}_{w}^{\ast }\bar {\nu }_{\theta (w)}=\bar {\unicode{x3bb} }_{w}\bar {\nu }_{w}.$
For any $\psi \in C^{\alpha }(M)$ we can write
From the convergence of $\nu _{k, \theta (w)}$ to $\bar {\nu }_{\theta (w)}$ we have
Since $\nu _{k, \theta (w)}$ is a reference measure, the last equality can be rewritten as
Thus, $\mathcal {L}_{w}^{\ast }\bar {\nu }_{\theta (w)}(\psi )=\bar {\unicode{x3bb} }_{w} \bar {\nu }_{w} (\psi )$ for any $\psi \in C^{\alpha }(M)$ . Because $C^{\alpha }(M)$ is dense in $C^{0}(M)$ we conclude that $\mathcal {L}_{w}^{\ast }\bar {\nu }_{\theta (w)}=\bar {\unicode{x3bb} }_{w}\bar {\nu }_{w}.$
Now we will verify that $\bar {\unicode{x3bb} }_{w}=\unicode{x3bb} _{w}$ . Therefore, from the uniqueness given by Theorem A it follows that $\bar {\nu }_{w}=\nu _{w}.$
Given $\varepsilon>0$ small and $n\in \mathbb {N}$ , consider a $(w, n, \varepsilon )$ -separated set $F_{n}$ . Let $\mathcal {U}$ be the open cover of M defined by $\mathcal {U}:=\{\bigcap _{j=0}^{n-1}f^{-j}_{w}(B(f_{w}^{j}(x), \varepsilon )); x\in F_{n}\}$ . Because $(\mathcal {L}^{n}_{w})^{\ast }\bar {\nu }_{\theta ^{n}(w)}=\bar {\unicode{x3bb} }^{n}_{w}\bar {\nu }_{w}$ it follows that
Thus, $\sum _{j=0}^{n-1}(\log \bar {\unicode{x3bb} }_{\theta ^{j}(w)}-|\phi _{\theta ^{j}(w)}|_{\alpha }\varepsilon )\leq \log P_{F|_{\theta }}(w, n, \varepsilon ).$ As $\mathbb {P}$ is ergodic we obtain
for every $\varepsilon>0$ small. Hence, $\int \!\log \bar {\unicode{x3bb} }_{w} \,d\mathbb {P}(w)\leq P_{F}(\phi )$ . On the other hand, since $\bar {\nu }_{w}$ is a reference measure, it satisfies a Gibbs property on hyperbolic times (Proposition 4.1). Thus, we can apply the proof of Proposition 6.1 to obtain $P_{F}(\phi )\leq \int \log \bar {\unicode{x3bb} }_{w}\,d\mathbb {P}(w)$ . Recalling that $P_{F}(\phi )=\int \log \unicode{x3bb} _{w}\, d\mathbb {P}(w)$ , we have
Since this is constant for $\mathbb {P}$ -almost every $w\in X$ , we have proved that $\bar {\unicode{x3bb} }_{w}=\unicode{x3bb} _{w}$ .
To finish the proof of the proposition it remains to prove the convergence $h_{k,w}\to h_{w}$ . Since $(F_{k}, \phi _{k}) \in \mathcal {H}$ we can assume that the transfer operator $\mathcal {L}_{k, w}$ preserves the same cone $\mathcal {C}^{\hat {k}}_{\delta }$ for $\hat {k}$ large enough. Then, recalling the proof of Proposition 5.3, we have each $h_{k, w}\in \mathcal {C}^{\hat {k}}_{\delta }$ with $\int h_{k,w}\,d\nu _{k, w}=1$ . Moreover, it satisfies
Therefore, $(h_{k,w})$ is a discontinuous and uniformly bounded sequence. From the Arzelà–Ascots theorem there exists some accumulation point $\bar {h}_{w}$ . Notice that $\bar {h}_{w}$ is Hölder continuous and $\int \bar {h}_{w}\,d\nu _{w}=1$ because $\nu _{k, w}\stackrel {w*}\longrightarrow \nu _{w}$ . Moreover, $\bar {h}_{w}$ satisfies
By the uniqueness of Theorem A we obtain $\bar {h}_{w}=h_{w}$ almost everywhere.
From the previous result we obtain for almost $w\in X$ that the sequence $(\mu _{k, w})$ converges to $\mu _{w}$ defined by $\mu _{w}=h_{w}\nu _{w}.$ Therefore, the sequence $(\mu _{k})$ converges to the probability measure $\mu $ whose disintegration is $\{\mu _{w}\}_{w\in X}$ . As in §6, we have that $\mu $ is the non-uniformly expanding equilibrium state of $(F|_{\theta }, \phi ).$ Moreover, since in the family $\mathcal {H}$ we have that $P_{F|_{\theta }}(\phi )=\int \log \unicode{x3bb} _{w}\,d\mathbb {P}(w)$ , we obtain
which proves that the random topological pressure varies continuously in the family. This finishes the proof of Theorem D.
8 Applications
In this section we present some classes of systems which satisfy our results. We start by describing a robust class of local diffeomorphisms which contains an open set of non-uniformly expanding maps that are not uniformly expanding. This class was studied in the deterministic case by several authors [Reference Alves and Araújo1, Reference Alves, Bonatti and Viana2, Reference Castro and Varandas12, Reference Varandas and Viana29]. The first example is a one-dimensional version of this class.
Example 8.1. Let $g:\mathbb {S}^{1}\to \mathbb {S}^{1}$ be a $C^{1} $ -local diffeomorphism defined on the unit circle. Fix $\varepsilon>0$ small, $\sigma <1$ and consider a covering $\mathcal Q$ of $\mathbb {S}^{1}$ by injectivity domains of g and a region $\mathit{A}\subset \mathbb {S}^{1}$ covered by q elements of $\mathcal Q$ with $q<\deg (g)$ such that
-
(H1) $\|Dg^{-1}(x)\|\leq 1+\varepsilon $ , for every $x\in \mathit{A}$ ;
-
(H2) $\|Dg^{-1}(x)\|\leq \sigma $ , for every $x\in M\setminus \mathit{A}$ .
Let us consider a suitable $C^{1}$ perturbation of g to produce an open set $\mathcal {F}$ of $C^{1}$ -local diffeomorphisms satisfying conditions (I)–(II). The perturbation can be chosen small enough to guarantee the uniform openness property in the family $\mathcal {F}$ . We also assume that every $g\in \mathcal {F}$ is topologically exact and its degree $\deg g$ is constant. Notice that $\mathcal {F}$ may contain expanding maps, perturbations of expanding maps and intermittent maps.
Let $\theta : \mathbb {S}^{1}\to \mathbb {S}^{1}$ be any invertible function preserving an ergodic measure $\mathbb {P}$ on $\mathbb {S}^{1}$ . Thus, any random dynamical system $g=(g_{w})_{w}$ generated by maps $g_{w}\in \mathcal {F}$ satisfies the hypotheses of our theorems. For potentials $\phi \in \mathbb {L}_{\mathbb {P}}^{1}(\mathbb {S}^{1}, C^{\alpha }(\mathbb {S}^{1}))$ satisfying (IV) and (V) we can apply our results to obtain the thermodynamic formalism in this class and the existence of only one equilibrium state on the set of non-uniformly expanding measures.
Moreover, if the potential also satisfies the condition $\sup \phi < P_{\phi }(f)$ then the equilibrium state is unique in the class of ergodic measures. Indeed, using the random versions of Oseledets’s theorem and Ruelle’s inequality (see [Reference Liu18]), for the equilibrium state $\mu $ the Lyapunov exponent $\unicode{x3bb} (\mu )$ satisfies
Therefore, $\unicode{x3bb} (\mu )$ is positive and bounded away from zero. In dimension one this implies that the equilibrium state is non-uniformly expanding.
The second example is a generalization of the previous one in higher dimension. The existence of equilibrium state for random transformations given by maps in this setting was considered by Arbieto, Matheus and Oliveira [Reference Arbieto, Matheus and Oliveira5].
Example 8.2. Let $M^{l}$ be a compact l-dimensional Riemannian manifold and $\mathcal {D}$ the space of $C^{2}$ local diffeomorphisms on M. Let $(\Omega , T, \mathbb {P})$ be a measure-preserving system where $\mathbb {P}$ is ergodic. Define the skew-product by
where the map $f(w)\in \mathcal {D}$ varies continuously on $w\in \Omega $ . Fix positive constants $\delta _{0}, \delta _{1}, \delta _{2}$ small and $p, q \in \mathbb {N}$ , satisfying for every $f(w)\in \mathcal {D}$ the following properties.
-
(H0) There exist $\delta _{w}>\delta _{0}>0$ such that for every $x \in M$ we can find a neighborhood $U_{x}$ where $f(w): U_{x} \rightarrow B_{T(w)}(f(w)(x), \delta _{w})$ is invertible.
-
(H1) There exists a covering $B_{1}\ldots B_{p}\ldots B_{p+q}$ of M by injectivity domains such that:
-
• $\|Df(x)^{-1}\|\leq (1+\delta _{2})^{-1}$ for every $x\in B_{1} \cup \cdots \cup B_{p}$ ;
-
• $\|Df(x)^{-1}\|\leq (1+\delta _{1})$ for every $x\in M$ .
-
-
(H2) f is everywhere volume expanding: $|\det Df(x)| \geq \sigma _{1}$ with $\sigma _{1}> q$ .
-
(H3) There exists $A_{0}$ such that $|\log \|f\|_{C^{2}} | \le A_{0}$ for any $f\in \mathcal {F} \subset \mathcal {D}$ .
Adding other technical hypotheses, the authors in [Reference Arbieto, Matheus and Oliveira5] showed the existence of equilibrium states for potentials with small variation. They also proved that these measures are non-uniformly expanding. Now, for potentials satisfying conditions (IV) and (V), we can apply our results to obtain the thermodynamic formalism and the uniqueness of equilibrium state for this class. Moreover, by considering the set $\mathcal {S}$ of skew-products generated by maps of $\mathcal {D}$ where $T:\Omega \to \Omega $ is fixed,
we define the family $\mathcal {H}=\{(F, \phi )\in \mathcal {S}\times \mathbb {L}^{1}_{\mathbb {P}}(X,C^{\alpha }(M)); (F, \phi )\, \mbox {satisfying~(I)--(VI)}\}$ and observe that $\mathcal {H}$ satisfies the hypothesis of Theorem D. Thus, the equilibrium state and the random topological pressure vary continuously within this family.
Next we present an application of our Corollary 6.1. This example appears in [Reference Bilbao and Oliveira9] in the context of maximizing entropy measures. Here we prove uniqueness of equilibrium states for potentials with small variation.
Example 8.3. Let $f_{0},f_{1}: M\rightarrow M$ be $C^{1}$ local diffeomorphisms of a compact and connected manifold M satisfying our conditions (I)–(III). For $1\leq k < \dim M=d$ suppose that $\log \lVert \Lambda ^{k} Df_{1}\rVert < \log \deg f_{1}$ and consider
Let $\mathbb {P}_{\alpha }$ be the Bernoulli measure on the sequence space $X=\{0,1\}^{\mathbb {Z}}$ such that $\mathbb {P}_{\alpha }([1])= \alpha $ . In [Reference Bilbao and Oliveira9] Birkhoff proved the existence of $\alpha \in (0, 1)$ close to $1$ such that
for every $x\in M.$ Therefore, for some $\varepsilon>0$ we have
which means that the hypothesis of Corollary 6.1 was verified.
Thus, for potentials $\phi \in \mathbb {L}^{1}_{\mathbb {P}}(X, C^{\alpha }(M))$ satisfying (IV) and (V) with variation $\sup \phi _{w}-\inf \phi _{w}<\varepsilon \int \log \deg (f_{w}) \,d \mathbb {P}_{\alpha }$ we conclude uniqueness of equilibrium states.
Acknowledgements
We would like to thank P. Varandas for useful suggestions and encouragement. We are grateful to the referee for their valuable comments which helped us to correct and improve the manuscript. RB thanks to K. Oliveira for many conversations and to IM-UFAL for the hospitality. VR thanks FAPEMA-Brazil for its financial support.