Received: 4 February 2016 / Accepted: 20 July 2016
Thermodynamic formalism, i.e., the formalism of equilibrium statistical physics, was adapted to the study of dynamical systems in the classical works of [Ruelle1972], [Ruelle1978], [Sinai1968], [Sinai1972], and [Bowen1970], [Bowen1974], [Bowen2008]. It provides an ample collection of methods for constructing invariant measures with strong statistical properties. In particular, this includes constructing a certain "physical" measure known as the SRB measure (for Sinai-Ruelle-Bowen).
The general ideas can be given as follows. Let $ (X,d)$ be a compact metric space and $ f \colon X \to X$ a continuous map of finite topological entropy. Fix a continuous function $ \ph\colon X\to \RR$, which we will refer to as a potential . Denote by $ \MMM(f)$ the space of all $ f$-invariant Borel probability measures on X. Given $ \mu\in \MMM(f)$, the free energy of the system with respect to $ \mu$ is \begin{eqnarray*} E_\mu(\ph) := -\left(h_\mu(f) + \int_X \ph\,d\mu\right), \end{eqnarray*} where $ h_\mu(f)$ is the Kolmogorov-Sinai (measure-theoretic) entropy of $ (X,f,\mu)$. Optimizing over all invariant measures gives the topological pressure \begin{eqnarray*} P(\ph) := -\inf_{\mu\in \MMM(f)} E_\mu(\ph) = \sup_{\mu\in \MMM(f)} \left(h_\mu(f) + \int_X \ph\,d\mu\right), \end{eqnarray*} and a measure achieving this extremum is called an equilibrium measure (or equilibrium state ). Note that it suffices to take the infimum (supremum) over the space $ \MMM^e(f) \subset \MMM(f)$ of ergodic measures.
The variational principle relates the definition of pressure as an extremum over invariant measures to an alternate definition in terms of growth rates. Given $ \eps> 0$ and $ n\in \NN$, a set $ E\subset X$ is $ (n,\eps)$- separated if points in $ E$ can be distinguished at a scale $ \eps$ within $ n$ iterates; more precisely, if for every $ x,y\in E$ with $ x\ne y$, there is $ 0\le k\le n$ such that $ d(f^kx,f^ky)\ge\eps$. Then one has
\begin{eqnarray}\label{eqn:Pspsgrowth} P(\ph)=\lim_{\eps\to 0}\limsup_{n\to\infty}\frac1n\log\sup_{\substack{E\subset X \\ (n,\eps)\text{-sep.}}}\sum_{x\in E} e^{S_n\ph(x)}, \end{eqnarray} | (1.1) |
\begin{eqnarray}\label{ind:poten} S_n\ph(x):=\sum_{k=0}^{n-1}\ph(f^kx). \end{eqnarray} | (1.2) |
Thermodynamic formalism is most useful when the system possesses some degree of hyperbolic behavior, so that orbit complexity increases exponentially. The most complete results are available when $ f$ is uniformly hyperbolic; we discuss these in Sect. 1.2. In this article we focus on non-uniformly hyperbolic systems, and we discuss the general picture in Sect. 1.3. Our emphasis will be on general techniques rather than on specific examples. In particular, we discuss Markov models (including Young towers) in Sects. 2–4, coarse-graining techniques (based on expansivity and specification) in Sect. 5, and push-forward (geometric) approaches (based on newly introduced standard pairs approach) in Sect. 6.
We refer the reader to ([Katok and Hasselblatt1995], [Brin and Stuck2002]) for fundamentals of uniform hyperbolicity theory and to ([Bowen2008], [Parry and Pollicott1990]) for a complete description of thermodynamic formalism for uniformly hyperbolic systems. Consider a compact smooth Riemannian manifold $ M$ and a $ C^1$ diffeomorphism $ f\colon M\to M$. A compact invariant set $ \Lambda\subset M$ is called hyperbolic if for every $ x\in\Lambda$ the tangent space $ T_xM$ admits an invariant splitting $ T_xM=E^s(x)\oplus E^u(x)$ into stable and unstable subspaces with uniform contraction and expansion: this means that there are numbers $ c> 0$ and $ 0< \lambda< 1$ such that for every $ x\in\Lambda$:
Moving from the tangent bundle to the manifold itself, for every $ x\in\Lambda$ one can construct local stable $ V^s(x)$ and unstable $ V^u(x)$ manifolds (also called leaves ) through $ x$ which are tangent to $ E^s(x)$ and $ E^u(x)$ respectively and depend Hölder continuously on $ x$ ([Katok and Hasselblatt1995], Sect. 6.2). In particular, there is $ \varepsilon> 0$ such that for any $ x,y\in\Lambda$ for which $ d(x,y)\le\varepsilon$ one has that the intersection $ V^s(x)\cap V^u(y)$ consists of a single point (here $ d(x,y)$ denotes the distance between points $ x$ and $ y$ induced by the Riemannian metric on $ M$). We denote this point by $ [x,y]$.
A hyperbolic set $ \Lambda$ is called locally maximal if there is a neighborhood $ U$ of $ \Lambda$ such that for any invariant set $ \Lambda'\subset U$ we have that $ \Lambda'\subset\Lambda$. In other words, $ \Lambda=\bigcap_{n\in\mathbb{Z}}\,f^n(U)$. One can show that a hyperbolic set $ \Lambda$ is locally maximal if and only if for any $ x,y\in\Lambda$ which are sufficiently close to each other, the point $ [x,y]$ lies in $ \Lambda$ ([Katok and Hasselblatt1995], Sect. 6.4).
Given a locally maximal hyperbolic set and a Hölder continuous potential function, thermodynamic formalism produces unique equilibrium measures with strong ergodic properties: before stating the theorem we recall some notions from ergodic theory for the reader's convenience. Let $ (X,\mu)$ be a Lebesgue space with a probability measure $ \mu$ and $ T\colon X\to X$ an invertible measurable transformation that preserves $ \mu$.
Before stating the formal result, we point out that uniformly hyperbolic systems (and many non-uniformly hyperbolic ones) satisfy various other statistical properties, which we do not discuss in detail in this survey. These include large deviations principles ([Orey and Pelikan1988], [Young1990], [Kifer1990], [Pfister and Sullivan2005], [Melbourne and Nicol2008], [Rey-Bellet and Young2008], [Climenhaga et al.2013]), Borel–Cantelli lemmas ([Chernov and Kleinbock2001], [Dolgopyat2004], [Kim2007], [Gouëzel2007], [Gupta et al.2010], [Haydn et al.2013]), the almost sure invariant principle ([Denker and Philipp1984], [Melbourne and Nicol2005], [Melbourne and Nicol2009]), and many more besides.
Recall also that a finite partition $ \mathcal{R}=\{R_1,\dots,R_p\}$ of $ \Lambda$ is a Markov partition if the following are true.
Markov partitions allow one to obtain a symbolic representation of the map $ f|\Lambda$ by subshifts of finite type. More precisely, let $ \mathcal{R}=\{R_1,\dots,R_p\}$ be a finite Markov partition of $ \Lambda$. Consider the subshift of finite type $ (\Sigma_A,\sigma)$ with the transition matrix $ A$ whose entries are given by $ a_{ij}=1$ if $ f(\mathrm{int}\, R_i)\cap \mathrm{int}\, R_j\ne\emptyset$ and $ a_{ij}=0$ otherwise. One can show that for every $ \omega=(\omega_i)\in\Sigma_A$ the intersection $ \bigcap_{i\in\mathbb{Z}}f^{-i}(R_{\omega_i}) $ is not empty and consists of a single point $ \pi(\omega)$. This defines the coding map $ \pi\colon\Sigma_A\to\Lambda$, which is characterized by the fact that $ f^i(\pi(\omega)) \in R_{\omega_i}$ for all $ i\in \ZZ$ (thus $ \omega$ "codes" the orbit of $ \pi(\omega)$).
We stress that this result (and hence Theorem 1.1) may not hold if the the potential function fails to be Hölder continuous, see [Hofbauer1977], [Sarig2001a], [Pesin and Zhang2006].
Returning from SFTs to the setting of uniformly hyperbolic smooth systems, the most significant potential function is the geometric $ t$ -potential : a family of potential functions $ \ph_t(x):= -t\log \abs{\Jac(df|E^u(x))}$ for $ t\in\RR$. Since the subspaces $ E^u(x)$ depend Hölder continuously on $ x$, the potential $ \ph_t$ is Hölder continuous for each $ t$ whenever $ f$ is $ C^{1+\alpha}$; in particular, it admits a unique equilibrium measure $ \mu_t$. Furthermore, the pressure function $ P(t) := P(\ph_t)$ is well defined for all $ t$, is convex, decreasing, and real analytic in $ t$, as in Fig. 1a.
There are certain values of $ t$ that are particularly important.
To further study the properties of the pressure function (and $ t_0$ in particular) we recall the notion of the Lyapunov exponent. Given $ x\in\Lambda$ and $ v\in T_xM$, define the Lyapunov exponent \begin{eqnarray*} \chi(x,v)=\limsup_{n\to\infty}\,\frac1n\log\|df^nv\|. \end{eqnarray*} For every $ x\in\Lambda$ the function $ \chi(x,\cdot)$ takes on finitely many values $ \chi_1(x)\le\cdots\le\chi_d(x)$ where $ d=\dim M$. The functions $ \chi_i(x)$ are Borel and are invariant under $ f$; in particular, if $ \mu$ is an ergodic measure, then $ \chi_i(x)=\chi_i(\mu)$ is constant almost everywhere for each $ i=1,\dots, d$, and the numbers $ \chi_i(\mu)$ are called the Lyapunov exponent of the measure $ \mu$. If none of these numbers is equal to zero, $ \mu$ is called a hyperbolic measure ; It is assumed that some of these numbers are positive while others are negative. × 7 note that when $ \Lambda$ is a hyperbolic set for $ f$, every invariant measure supported on $ \Lambda$ is hyperbolic. The Margulis–Ruelle inequality (see [Ruelle1979, Barreira and Pesin2013]) says that
\begin{eqnarray}\label{mri} h_\mu(f)\le\sum_{i:\chi_i(\mu)\ge 0}\chi_i(\mu) \end{eqnarray} | (1.3) |
We consider the particular case when $ \Lambda$ is a topological attractor for $ f$. This means that there is a neighborhood $ U\supset\Lambda$ such that $ \overline{f(U)}\subset U$ and $ \Lambda=\bigcap_{n\ge 0}f^n(U)$. It is not difficult to see that for every $ x\in\Lambda$, the local unstable manifold $ V^u(x)$ is contained in $ \Lambda$; Indeed, for any $ y\in V^u(x)$ the trajectory of $ y$, $ \{f^n(y)\}_{n\in\mathbb{Z}}$ lies in $ U$ and hence, must belong to $ \Lambda$ since it is locally maximal. × 8 the same is true for the global unstable manifold through $ x$. Therefore, the attractor contains all the global unstable manifolds of its points. On the other hand the intersection of $ \Lambda$ with stable manifolds of its points is usually a Cantor set.
In the case when $ \Lambda$ is a hyperbolic attractor we have that $ t_0=1$ (see [Bowen2008]), so $ P(t)$ is as in Fig. 1b. The equilibrium state $ \mu_1$ is a hyperbolic ergodic measure for which the Margulis–Ruelle inequality (1.3) becomes equality. By [Ledrappier and Young1985], this implies that $ \mu_1$ has absolutely continuous conditional measures along unstable manifolds; that is, there is a collection $ \RRR$ of local unstable manifolds $ V^u$ and a measure $ \eta$ on $ \RRR$ such that $ \mu_1$ can be written as
\begin{eqnarray}\label{eqn:ac} \mu_1(E) = \int_\RRR \mu_{V^u}(E) \,d\eta(V^u) \end{eqnarray} | (1.4) |
A $ C^{1+\alpha}$ diffeomorphism $ f$ of a compact smooth Riemannian manifold $ M$ is non-uniformly hyperbolic on an invariant Borel subset $ S\subset M$ if there are a measurable $ df$-invariant decomposition of the tangent space $ T_xM=E^s(x)\oplus E^u(x)$ for every $ x\in S$ and measurable $ f$-invariant functions $ \varepsilon(x)> 0$ and $ 0< \lambda(x)< 1$ such that for every $ 0< \varepsilon\le\varepsilon(x)$ one can find measurable functions $ c(x)> 0$ and $ k(x)> 0$ satisfying for every $ x\in S$:
If $ \mu$ is an invariant measure for $ f$ with $ \mu(S)=1$, then by the Multiplicative Ergodic theorem, if for almost every $ x\in S$ the Lyapunov exponents at $ x$ are all nonzero, i.e., $ \mu$ is a hyperbolic measure, then $ f$ is non-uniformly hyperbolic on $ S$.
A general theory of thermodynamic formalism for non-uniformly hyperbolic maps is far from being complete, although certain examples here are well-understood. They include one-dimensional maps, where the pressure function $ P(t)=P(\ph_t)$ associated with the family of geometric potentials may behave as in the uniformly hyperbolic case, or may exhibit new phenomena such as phase transitions (points of non-differentiability where there is more than one equilibrium measure). The latter is illustrated in Fig. 1c and is most thoroughly studied for the Manneville–Pomeau map $ x\mapsto x+x^{1+\alpha} \pmod 1$, where $ \alpha\in (0,1)$ controls the degree of intermittency at the neutral fixed point. In this example one has the following behavior ([Pianigiani1980], [Thaler1980], [Thaler1983], [Lopes1993], [Pollicott and Weiss1999], [Liverani et al.1999], [Young1999], [Sarig2002], [Hu2004]).
Our goal in the rest of this paper is not to discuss these results, which rely on the specific structure of the examples being studied (or on the absence of a contracting direction); rather, we want to discuss the recently developed techniques for studying multi-dimensional non-uniformly hyperbolic systems, with particular emphasis on recent results that have the potential to be applied very generally, although they do not yet give as complete a picture as the one outlined above. These general results have been obtained in the last few years and represent an actively evolving area of research.
Before describing the general methods, we recall some basic notions from non-uniform hyperbolicity; see [Barreira and Pesin2007] for more complete definitions and properties. Let $ M$ be a compact smooth manifold and $ f\colon M\to M$ a $ C^{1+\alpha}$ diffeomorphism. Recall that a point $ x\in M$ is called Lyapunov–Perron regular if for any basis $ \{v_1,\dots, v_p\}$ of $ T_xM$, \begin{eqnarray*} \liminf_{n\to\pm\infty}\frac1n\log V(n)= \limsup_{n\to\pm\infty}\frac1n\log V(n) =\sum_{i=1}^p\,\chi_i(x,v_i), \end{eqnarray*} where $ V(n)$ is the volume of the parallelepiped built on the vectors $ \{df^nv_1,\dots, df^nv_p\}$.
Let $ \mathcal{R}$ be the set of all Lyapunov–Perron regular points. The Multiplicative Ergodic theorem claims that this set has full measure with respect to any invariant measure. Consider now the set $ \Gamma\subset\mathcal{R}$ of points for which all Lyapunov exponents are nonzero, and let $ \MMM^e(f,\Gamma)\subset \MMM^e(f)$ be the set of all ergodic measures that give full weight to the set $ \Gamma$; these are hyperbolic measures and they form the class of measures where it is reasonable to attempt to recover some of the theory of uniformly hyperbolic systems.
Let $ \ph$ be a measurable potential function; note that we cannot a priori assume more than measurability if we wish to include the family of geometric potentials, since in general the unstable subspace varies discontinuously and so $ \ph_t$ is not a continuous function. On the other hand, for surface diffeomorphisms [Sarig2013] constructed Markov partitions with countably many partition elements (see Sect. 3 below), and showed [Sarig2011] that the function $ \ph_t$ can be lifted to a function on the symbolic space that is globally well-defined and is Hölder continuous. This can be used to study equilibrium measures for this function. × 10 Consider the hyperbolic pressure defined by using only hyperbolic measures:
\begin{eqnarray}\label{eqn:PG} P_\Gamma(\ph) := -\inf_{\mu\in \MMM^e(f,\Gamma)} E_\mu(\ph). \end{eqnarray} | (1.5) |
One could also fix a threshold $ h> 0$ and consider the set $ \MMM^e(f,\Gamma,h)$ of all measures in $ \MMM^e(f,\Gamma)$ whose entropies are greater than $ h$; restricting our attention to measures from this class gives the restricted pressure Because $ \MMM^e(f,\Gamma,h)$ is not compact, the existence of an optimizing measure in (1.6) becomes a more subtle issue. Although it may happen that the value of $ P_\Gamma^h(\ph)$ is achieved by a measure $ \mu$ whose entropy may not be greater than $ h$, the restriction to measures in the class $ \MMM^e(f,\Gamma,h)$ is often made to ensure a certain "liftability" condition, which may still be satisfied by $ \mu$; see Theorem 2.3 and the discussion in that section. × 12
\begin{eqnarray}\label{eqn:PGh} P_\Gamma^h(\ph) := -\inf_{\mu\in\MMM^e(f,\Gamma,h)} E_\mu(\ph). \end{eqnarray} | (1.6) |
One could also impose a threshold in other ways. For example, one could fix a reference potential $ \psi$ and a threshold $ p < P(\psi)$, then restrict attention to the set $ \MMM^e(f,\Gamma,\psi,p)$ of all measures in $ \MMM^e(f,\Gamma)$ for which $ -E_\mu(\psi)> p$. Optimizing $ E_\mu(\ph)$ over this restricted set of measures gives another notion of thresholded equilibrium states that may be useful; again, it is often natural to take $ p=P_S(\ph)$ as the topological pressure of $ f$ on a (not necessarily invariant) subset $ S\subset M$ of bad points. Another approach would be to consider only measures whose Lyapunov exponents are sufficiently large; it may be that this is a more natural approach in certain settings. We stress that while restricting the class of invariant measures using thresholds for the topological pressure or Lyapunov exponents seem to be natural it is yet to be shown to be a working tool in effecting thermodynamic formalism.
A direct application of the uniformly hyperbolic approach in the non-uniformly hyperbolic setting is hopeless in general; we cannot expect to have finite Markov partitions. Indeed, if a map possesses a Markov partition, then its topological entropy is the logarithm of an algebraic number, which should certainly not be expected in general. On the other hand, in the presence of a hyperbolic invariant measure $ \mu$ of positive entropy, there are horseshoes with finite Markov partitions whose entropy approximates the entropy of $ \mu$ [Katok1980], but these have zero $ \mu$-measure. × 17 However, in many cases it is possible to use the symbolic approach by finding a countable Markov partition , or the related tools of a Young tower or a more general inducing scheme ; these are discussed in Sects. 2–4. This approach is challenging to apply completely, but can help establish existence and uniqueness of equilibrium measures and study their statistical properties including decay of correlations and the CLT.
A second approach is to avoid the issue of building a Markov partition by adapting Bowen's specification property to the non-uniformly hyperbolic setting; this is discussed in Sect. 5. This is similar to the symbolic approach in that one uses a "coarse-graining" of the system to make counting arguments borrowed from statistical physics, but sidesteps the issue of producing a Markov structure. The price paid for this added flexibility is that while existence and uniqueness can be obtained with specification-based techniques, there does not seem to be a direct way to obtain strong statistical properties without first establishing some sort of Markov structure.
A third approach, which we discuss in Sect. 6, is geometric and is based on pushing forward the leaf volume on unstable manifolds by the dynamics. More generally, one can work with approximations to unstable manifolds by admissible manifolds and use measures which have positive densities with respect to the leaf volume as reference measures. Such pairs of admissible manifolds and densities are called standard and working with them has proven to be quite a useful technique in various problems in dynamics. This notion was introduced by [Chernov and Dolgopyat2009]. × 18 So far the geometric approach can be used to establish existence of SRB measures for uniformly hyperbolic and some non-uniformly hyperbolic attractors and one can also use a version of this method to construct equilibrium measures for uniformly hyperbolic sets, see Sect. 6; the questions of uniqueness and statistical properties using this approach as well as construction of equilibrium measures for non-uniformly hyperbolic systems are still open.
In the remainder of this paper we describe the three approaches just listed in more detail, and discuss their application to open problems in the thermodynamics of non-uniformly hyperbolic systems.
In one form or another, the use of Markov models with countably many states to study non-uniformly hyperbolic systems dates back to the late 1970s and early 1980s, when [Hofbauer1979], [Hofbauer1981a], [Hofbauer1981b] used a countable-state Markov model to study equilibrium states for piecewise monotonic interval maps. Indeed, such models for $ \beta$-transformations were studied already in 1973 by [Takahashi1973].
In [Jakobson1981] Jakobson initiated the study of thermodynamics of unimodal interval maps by constructing absolutely continuous invariant measures (acim) for the family of quadratic maps $ f_a(x) = 1-ax^2$ whenever $ a\in\Delta$, where $ \Delta$ is a set of parameters with positive Lebesgue measure. First we discuss in Sect. 2.2 the extensions of Jakobson's result to study SRB measures by what have become known as Young towers . Then in Sect. 3 we discuss the study of general equilibrium states in the setting of topological Markov chains with countably many states, which generalizes the SFT theory from Sect. 1.2. Finally, in Sect. 4 we discuss the use of inducing schemes to apply this theory to the thermodynamics of smooth examples.
Roughly speaking, a tower construction begins with a base set $ \Lambda$, a map $ G\colon\Lambda\to\Lambda$, and a height function $ R\colon\Lambda\to\NN$. Then the tower is constructed as $ \tilde\Lambda:=\{(z,n)\in\Lambda\times \{0,1,2,\dots\}:n< R(z)\}$, and a map $ g\colon\tilde\Lambda\to\tilde\Lambda$ is defined by $ g(z,n)=(z,n+1)$ whenever $ n+1< R(z)$, and $ g(z,R(z)-1)=(F(z),0)$. Typically one requires that the dynamics of the return map $ G$ can be coded by a full shift, or a Markov shift on a countable set of states. To study a dynamical system $ f\colon X\to X$ using a tower, one defines a coding map $ \pi\colon\tilde \Lambda\to X$ such that $ f\circ\pi=\pi\circ g$; this coding map is usually not surjective (the tower does not cover the entire space), and so we will ultimately need to give some "largeness" condition on the tower. It is important to distinguish between the case when $ \pi(\Lambda)$ is disjoint from $ \pi(\tilde\Lambda{\setminus}\Lambda)$, so that the height $ R$ is the first return time to the base $ \pi(\Lambda)$, and the case when $ R$ is not the first return time.
Tower constructions for which the height of the tower is the first return time to the base of the tower are classical objects in ergodic theory and were considered in works of Kakutani, Rokhlin, and others. Towers for which the height of the tower is not the first return time appeared in the paper by [Neveu1969] under the name of temps d'arret and in the context of dynamical systems in the paper by [Schweiger1975], [Schweiger1979] under the name jump transformation (which are associated with some fibered systems ; see also the paper by [Aaronson et al.1993] for some general results on ergodic properties of Markov fibered systems and jump transformations).
A tower construction is implicitly present in Jakobson's proof of existence of physical measures for quadratic maps. The first significant use of the tower approach beyond the one-dimensional setting came in the study of the Hénon map
\begin{eqnarray}\label{henon} f_{a,b}(x,y) = (1-ax^2 + y, bx), \end{eqnarray} | (2.1) |
The general structure behind these results was developed in [Young1998] and has come to be known as a Young tower , It is worth mentioning that a major achievement of [Young1998] was to establish exponential decay of correlations for billiards with convex scatterers, which is an example of a uniformly hyperbolic system with discontinuities; we will not discuss such examples further in this paper. × 19 or a Gibbs–Markov–Young structure . The principal feature of a Young tower is that the induced map on the base of the tower is conjugate to the full shift on the space of two-sided sequences over countable alphabet. This allows one to use some recent results on thermodynamics of this symbolic map to establish existence and uniqueness of equilibrium measures for the original map and study their ergodic properties.
A $ C^{1+\alpha}$ diffeomorphism $ f$ of a compact smooth manifold $ M$ is called Young diffeomorphism if it admits a Young tower . This tower has a particular structure which is characterized as follows:
A formal description of the Young tower is as follows. There are two continuous families $ {\bf V}^u=\{V^u\}$ and $ {\bf V}^s=\{V^s\}$ of local unstable and stable manifolds, respectively, with the property that each $ V^s$ meets each $ V^u$ transversely in a single point and $ \Lambda=(\bigcup V^u)\cap (\bigcup V^s)$; a union of some of the manifolds $ V^u$ is called a $ u$ -set , a union of some of the manifolds $ V^s$ is called an $ s$ -set . One asks for $ \Lambda$ to have the following properties; here $ C,\eta> 0$ and $ \beta\in (0,1)$ are constants.
Once a tower structure has been found, the strength of the conclusions one can draw depends on the rate of decay of the tail of the tower ; that is, the speed with which $ m_{V^u}\{x\in V^u\mid R(x) > T\}\to 0$ as $ T\to\infty$ for $ V^u\in {\bf V}^u$. We say that with respect to the measure $ m_{V^u}$ the tower has
\begin{eqnarray}\label{expspstails} m_{V^u}\{x\mid R(x)> T\} < Ce^{-aT}; \end{eqnarray} | (2.2) |
\begin{eqnarray}\label{fullmeasure} m_{V^u}\left(\bigcup_{i\ge 1}\bar{\Lambda}_i{\setminus}\Lambda_i\right)=0; \end{eqnarray} | (2.3) |
Note that even without the arithmetic condition one still obtains the "exponential decay up to a period" result stated earlier in Theorem 1.1 (2) .
In [Young1999], Young gave an extension of the results from [Young1998] that applies in a more abstract setting, giving existence of an invariant measure that is absolutely continuous with respect to some reference measure (not necessarily Lebesgue). She also provided a condition on the height of the tower that guarantees a polynomial upper bound for the decay of correlations. The corresponding polynomial lower bound (showing that Young's bound is optimal) was obtained by [Sarig2002] and [Gouëzel2004].
The flexibility in the reference measure makes Young's result suitable for studying existence, uniqueness and ergodic properties of equilibrium measures other than SRB measures (although this was not done in [Young1999]). In particular, this is used in the proof of Statement 2 of Theorem 2.3 below; we discuss such questions more in Sects. 3, 4.
Just as the Hénon maps can be studied as a "small" two-dimensional extension of the unimodal maps, Theorems 2.1 and 2.2 can be applied to more general 'strongly dissipative' maps that are obtained as 'small' two-dimensional extensions of one-dimensional maps; this is carried out in [Wang and Young2001], [Wang and Young2008].
Aside from such strongly dissipative maps, Young towers have been constructed for some partially hyperbolic maps where the center direction is non-uniformly contracting ([Castro2004]) or expanding ([Alves and Pinheiro2010], [Alves and Li2015]); the latter papers are built on earlier results for non-uniformly expanding maps where one does not need to worry about the stable direction ([Alves et al.2005], [Gouëzel2006]). In both cases existence (and uniqueness) of an SRB measure was proved first ([Bonatti and Viana2000], [Alves et al.2000]) via other methods closer to the push-forward geometric approach that we discuss in Sect. 6, so the achievement of the tower construction was to establish exponential decay of correlations and the CLT. These results only cover the SRB measure and do not consider more general equilibrium states.
Let $ f$ be a $ C^{1+\alpha}$ Young diffeomorphism of a compact smooth manifold $ M$. Consider the set $ \Lambda$ with hyperbolic product structure. Let $ \Lambda_i^s$ be the collections of $ s$-sets and $ R_i$ the corresponding inducing times. Set \begin{eqnarray*} Y=\bigcup_{k\ge 0}\,f^k(\Lambda). \end{eqnarray*} This is a forward invariant set for $ f$. For every $ y\in Y$ the tangent space at $ y$ admits an invariant splitting $ T_yM=E^s(y)\oplus E^u(y)$ into stable and unstable subspaces. Thus we can consider the geometric $ t$-potential $ \varphi_t(y)$ which is well defined for $ y\in Y$ and is a Borel (but not necessarily continuous) function for every $ t\in\RR$. We consider the class $ \mathcal{M}(f,Y)$ of all invariant measures $ \mu$ supported on $ Y$, i.e., for which $ \mu(Y)=1$. It follows that $ \mu(\Lambda)> 0$, so that $ \mu$ 'charges' the base of the Young tower. Further, given a number $ h> 0$, we denote by $ \mathcal{M}(f,Y, h)$ the class of invariant measures $ \mu\in\mathcal{M}(f,Y)$ for which $ h_\mu(f)> h$.
The following result describes existence, uniqueness, and ergodic properties of equilibrium measures. Given $ n> 0$, denote by \begin{eqnarray*} S_n:=\text{Card}\{\Lambda_i^s\colon R_i=n\}. \end{eqnarray*}
\begin{eqnarray}\label{growth} S_n\le e^{hn}, \end{eqnarray} | (2.4) |
\begin{eqnarray}\label{qnu} Q_\nu=\int_\Lambda R\, d\nu \end{eqnarray} | (2.5) |
\begin{eqnarray}\label{lift} \mu(E)={\mathcal L}(\nu)(E):=\frac{1}{Q_\nu}\sum_{i\ge 0}\sum_{k=0}^{R_i-1}\nu(f^{-k}(E)\cap \Lambda_i^s). \end{eqnarray} | (2.6) |
Under the condition 2.4 every measure with entropy $ {> }h$ is liftable. In general, it is shown in [Zweimüller2005] that if $ R\in L^1(Y,\mu)$ then $ \mu$ is liftable. In particular, if the return time $ R$ is the first return time to the base of the tower, then every measure that charges the base of the tower is liftable.
We describe two examples of Young diffeomorphisms for which
Theorem 2.3
applies.
The first example is Hénon-like diffeomorphisms of the plane
at the first bifurcation parameter. For parameters $ a,b$ consider the
Hénon map $ f_{a,b}$ given by
(2.1)
. It is shown in [Bedford and Smillie2004], [Bedford et al.2006], [Cao et al.2008] that for each $ 0< b\ll 1$ there
exists a uniquely defined parameter $ a^*=a^*(b)$ such that the
non-wandering set for $ f_{a,b}$ is a uniformly hyperbolic horseshoe for
$ a > a^*$ and the parameter $ a^*$ is the first parameter value for
which a homoclinic tangency between certain stable and unstable manifolds
appears. We describe the Katok map ([Katok1979]) (see also [Barreira and Pesin2013]), which can be thought of
as an invertible and two-dimensional analogue of the
Manneville–Pomeau map. Consider the automorphism of the 2-torus
given by the matrix $ T=( \begin{array}{ll} 2 &\quad 1\\ 1 &\quad 1\end{array})$ and then choose $ 0< \alpha< 1$ and a
function $ \psi:[0,1]\mapsto[0,1]$ satisfying: We slow down trajectories of
(2.7)
by perturbing it in $ D_{r_0}$ as follows: \begin{eqnarray*} \dot{s}_1=s_1\psi({s_1}^2+{s_2}^2)\log\lambda, \quad \dot{s}_2=- s_2\psi({s_1}^2+{s_2}^2)\log\lambda. \end{eqnarray*} This generates a
local flow, whose time-1 map we denote by $ g$. The choices of
$ \psi$ and $ r_0$ guarantee that the domain of $ g$
contains $ D_{r_0}$. Furthermore, $ g$ is of class $ C^\infty$
in $ D_{r_0}$ except at the origin and it coincides with $ T$ in
some neighborhood of the boundary $ \partial D_{r_0}$. Therefore, the map
\begin{eqnarray*} G(x)=\begin{cases} T(x) & \text{if} \quad x\in\mathbb{T}^2{\setminus} D_{r_0},\\ g(x) & \text{if} \quad x\in D_{r_0} \end{cases} \end{eqnarray*} defines a homeomorphism of the torus, which is a $ C^\infty$
diffeomorphism everywhere except at the origin. The map $ G$ preserves the probability measure
$ d\nu=\kappa_0^{-1}\kappa\,dm$ where $ m$ is the area and the density $ \kappa$
is defined by \begin{eqnarray*} \kappa(s_1,s_2):=\begin{cases} (\psi({s_1}^2+{s_2}^2))^{-1} &\text{if}\,\, (s_1,s_2)\in D_{r_0},\\ 1 & \text{otherwise} \end{cases} \end{eqnarray*} and \begin{eqnarray*} \kappa_0:=\int_{\mathbb{T}^2}\kappa\,dm. \end{eqnarray*} We further perturb the map
$ G$ by a coordinate change $ \phi$ in $ \mathbb{T}^2$ to
obtain an area-preserving $ C^\infty$ diffeomorphism. To achieve this,
define a map $ \phi$ in $ D_{r_0}$ by the formula
2.3.1. A Hénon-like Diffeomorphism at
the First Bifurcation
Theorem 2.4.
([Senti and Takahasi2013], [Senti and Takahasi2016], Theorem A) For any
bounded open interval $ I\subset(-1,+\infty)$ there exists $ 0< b_0\ll 1$ such that if
$ 0\le b< b_0$ then
2.3.2. The Katok
Map.
Let $ D_r=\{(s_1,s_2): {s_1}^2+{s_2}^2\le r^2\}$ where $ (s_1,s_2)$ is the coordinate system obtained from
the eigendirections of $ T$. Consider the system of differential
equations in $ D_{r_0}$
\begin{eqnarray}\label{batata10} \dot{s}_1= s_1\log\lambda,\quad \dot{s}_2=-s_2\log\lambda, \end{eqnarray}
(2.7)
\begin{eqnarray}\label{mapshi} \phi(s_1,s_2):=\frac{1}{\sqrt{\kappa_0({s_1}^2+{s_2}^2)}} \bigg(\int_0^{{s_1}^2+{s_2}^2}\frac{du}{\psi(u)}\bigg)^{1/2} (s_1,s_2) \end{eqnarray}
(2.8)
Theorem 2.5.
(see [Pesin et al.2016a]) The following statements hold:
The thermodynamic formalism for SFTs rested on the Ruelle's
version of the Perron–Frobenius theorem for finite-state topological
Markov chains. For the class of two-step potential functions $ \ph(x) = \ph(x_0,x_1)$,
which includes the zero potential $ \ph=0$, the extension of this theory to
countable-state Markov shifts dates back to work of [Vere-Jones1962], [Vere-Jones1967], [Gurevič1969], [Gurevič1970], [Gurevič1984], and [Gurevich and Savchenko1998]; we discuss this in
Sect. 3.1.
For more general potential functions a sufficiently complete picture is
primarily due to Sarig, and we discuss these in Sect. 3.2.
Recall the form of Theorem 1.3 on existence of a unique MME for SFTs: Existence of a stationary vector $ \pi=(\pi_i)$ with $ \pi P=\pi$ is
determined by the recurrence
properties of the shift [Vere-Jones1962], [Vere-Jones1967]. Suppose we start our random
walk at a vertex $ a$; one can show that the probability that we
return to $ a$ infinitely many times is either 0 or 1. If the
probability of returning infinitely many times is 1, then the walk is recurrent
. Recurrence is necessary in order to have a stationary probability vector
$ \pi$, but it is not sufficient; one must distinguish between the case
when our expected return time is finite ( positive
recurrence
) and when it is infinite ( null recurrence
). If the walk is positive recurrent then there is a stationary probability vector
$ \pi$; if it is null recurrent then one can still find a vector
$ \pi$ such that $ \pi P=\pi$, but one has $ \sum_i\pi_i=\infty$, so
$ \pi$ cannot be normalized to a probability vector. In fact, the trichotomy between transience, null recurrence, and
positive recurrence is the key to generalizing all of Theorem 1.3 to the
countable-state case [Pesin2014]. The recurrence conditions can be
formulated in terms of the number of loops in the graph $ G$. Fixing
a vertex $ a$, let $ Z_n^*$ be the number of simple
loops of length $ n$ based at $ a$ (first returns to
$ a$) and $ Z_n$ be the number of
all
loops of length $ n$ based at $ a$ (including loops which
return more than once). In the next section when we consider non-zero
potentials, we will have to count the loops with weights coming from the
potential. × 28 It is instructive to note that once a distinguished vertex
$ a$ is fixed as the starting point of the loops, one can view the first
return map to $ [a]$ as a Young tower, and then the summability
condition in positive recurrence is equivalent to the condition that the tails of
the tower are integrable, which was the existence criterion in Theorem 2.1.
In discussing the extension to non-zero potentials on countable-state
topological Markov chains, we will follow the notation, terminology, and
results of [Sarig1999], [Sarig2001b], [Sarig2001a], although the contributions of [Gurevič1984], [Gurevich and Savchenko1998], [Mauldin and Urbański1996], [Mauldin and Urbański2001], [Aaronson and Denker2001], and of [Fiebig et al.2002] should also be mentioned. Sarig
adapted transience, null recurrent, and positive recurrence for non-zero
potential functions. The summability criterion for positive recurrence is
exactly as above, except that now $ Z_n$ represents the total weight of
all loops of length $ n$ and $ Z_n^*$ represents the total weight
of simple loops of length $ n$ where weight is computed with respect
to the potential function; more precisely \begin{eqnarray*} Z_n=Z_n(\varphi,a)=\sum_{\sigma^n(x)=x}\exp(\Phi_n(x))1_{[a]}(x) \end{eqnarray*} and \begin{eqnarray*} Z_n^*=Z_n^*(\varphi,a)=\sum_{\sigma^n(x)=x}\exp(\Phi_n(x))1_{[\varphi_a=n]}(x), \end{eqnarray*} where
$ \Phi_n(x)=\sum_{k=0}^{n-1}\varphi(f^kx)$. Furthermore, the Gurevich entropy $ h_G(\sigma)$ is replaced
with the Gurevich-Sarig pressure
$ P_{GS}(\sigma,\ph)$, which is the exponential growth rate of $ Z_n$, i.e.,
\begin{eqnarray*} P_{GS}(\sigma,\ph)=\lim_{n\to\infty}\frac1n \log Z_n. \end{eqnarray*} For Markov shifts with finite topological entropy, [Buzzi and Sarig2003] proved that an equilibrium
measure exists if and only if the shift is positive recurrent. A good summary of
the theory can be found in [Sarig2015]. For our purposes the main result is the
following. Using Pesin theory, Sarig recently carried out a version of the
construction of Markov partitions for non-uniformly hyperbolic
diffeomorphisms in two dimensions. Recall that for a uniformly hyperbolic
diffeomorphism $ f\colon M\to M$, one obtains an SFT $ \Sigma$ and a
coding map $ \pi\colon \Sigma\to M$ such that The analogous result to Theorem 3.2 for
three-dimensional flows was proved by [Lima and Sarig2014]. In both cases this can be
used to deduce Bernoullicity up to finite rotations of ergodic positive entropy
equilibrium states ([Sarig2011], [Ledrappier et al.2016]). However, these general
results do not give any information on the recurrence properties of the
countable state shift, or the tail of the resulting tower, and in particular they
do not provide a mechanism for verifying decay of correlations and the CLT.
This is of no surprise, since at this level of generality, one should not expect to
get exponential decay (or any other particular rate).
3.1. Recurrence Properties for
Random Walks.
In the countable-state setting, existence of eigenvectors and stationary vectors
is a more subtle question (although once these are found, the proof of
uniqueness goes through just as in the finite-state case). The general story is
well-illustrated by just considering the last step above: suppose we are given a
stochastic matrix $ P_{ij}$ with countably many entries. This
corresponds to a directed graph $ G$ with countably many vertices,
whose edges are given weights as follows: the weight of the edge from
$ i$ to $ j$ is $ P_{ij}$. Then one can consider the
Markov chain described by $ P$ as a
random walk
on $ G$.
3.2. Non-zero
Potentials.
Theorem 3.1.
Let $ \Sigma$ be a topologically mixing
countable-state Markov shift with finite topological entropy, and let
$ \ph\colon X\to\RR$ be a Hölder continuous In fact Theorem 3.1 holds
for the more general class of potentials with
summable variations
, but Hölder continuity is needed for the statistical properties mentioned
below. × 30 function such that $ P_{GS}(\ph)< \infty$.
Then $ \ph$ is positive recurrent if and only if there are $ \lambda> 0$,
a positive continuous function $ h$, and a conservative measure
$ \nu$ (i.e., a measure that allows no nontrivial wandering sets) which
is finite and positive on cylinders, such that $ \LLL_\ph h=\lambda h$, $ \LLL_\ph^*\nu=\lambda\nu$, and
$ \int h\,d\nu=1$. In this case the following are true.
3.3. Countable-State Markov Partitions for Smooth
Systems
In non-uniform hyperbolicity one must replace the SFT with a countable-state
Markov shift, and also weaken some of the conclusions. Theorem 3.2.
[Sarig2013] Let $ M$ be a compact smooth
surface and $ f\colon M\to M$ a $ C^{1+\alpha}$ diffeomorphism of positive
topological entropy. Fix a threshold $ 0< \chi < \htop(f)$. Then there is a
countable-state topological Markov shift $ \Sigma_\chi$ and a coding map
$ \pi_\chi \colon \Sigma_\chi \to M$ such that
The study of SRB measures via Young towers generalizes to the study of equilibrium states via inducing schemes , which use the tower approach to model (a large part of) the system by a countable-state Markov shift, and then apply the thermodynamic results from Sect. 3. The concept of an inducing scheme in dynamics is quite broad and applies to systems which may be invertible or not, smooth or not differentiable. Every inducing scheme generates a symbolic representation by a tower which is well adapted to constructing equilibrium measures for an appropriate class of potential functions using the formalism of countable state Markov shifts. The projection of these measures from the tower are natural candidates for the equilibrium measures for the original system.
In order to use this symbolic approach to establish existence and to study equilibrium states, some care must be taken to deal with the liftability problem as only measures that can be lifted to the tower can be 'seen' by the tower.
One may consider inducing schemes of expanding type, or of hyperbolic type. The former were introduced in [Pesin and Senti2008] and apply to study thermodynamics of non-invertible maps (e.g., non-uniformly expanding maps) while the latter were introduced in [Pesin et al.2016b] and are used to model invertible maps (e.g,, non-uniformly hyperbolic maps). In this paper we only consider inducing schemes of hyperbolic types and we follow [Pesin et al.2016b].
Let $ f\colon X\to X$ be a homeomorphism of a compact metric space $ (X, d)$. We assume that $ f$ has finite topological entropy $ \htop(f)< \infty$. An inducing scheme of hyperbolic type for $ f$ consists of a countable collection of disjoint Borel sets $ S=\{J\}$ and a positive integer-valued function $ \tau\colon S\to\mathbb{N}$; the inducing domain of the inducing scheme $ \{S,\tau\}$ is $ W=\bigcup_{J\in S}J$, and the inducing time $ \tau\colon X\to\mathbb{N}$ is defined by $ \tau(x)=\tau(J)$ for $ x\in J$ and $ \tau(x)=0$ otherwise. We require several conditions.
Set $ Y= \{f^k(x)\mid x\in W, 0\le k\le \tau(x)-1\}$. Note that $ Y$ is forward invariant under $ f$. This can be thought of as the region of $ X$ that is 'swept out' as $ W$ is carried forward under the dynamics of $ f$; in particular, it contains all trajectories that intersect the base $ W$.
Let $ \varphi$ be a potential function. Existence of an equilibrium measure for $ \varphi$ is obtained by first studying the problem for the induced system $ (F,W)$ and the induced potential $ \overline{\varphi}\colon W\to\mathbb{R}$ defined by (1.2) . The study of existence and uniqueness of equilibrium measures for the induced system $ (F,W)$ is carried out by conjugating the induced system to the two-sided full shift over the countable alphabet $ S$. This requires that the potential function $ \Phi:={\bar{\varphi}}\circ \pi$ be well defined on $ S^\mathbb{Z}$. To this end we require that
Let $ X$ be a compact metric space and $ f\colon X\to X$ a homeomorphism; given $ \eps> 0$ and $ x\in X$, the set
\begin{eqnarray}\label{eqn:nespsset} \Gamma_\eps(x) := \left\{y\in X \mid d(f^nx,f^ny) < \eps \text{ for all } n\in \ZZ \right\} \end{eqnarray} | (5.1) |
To show that this equilibrium state is unique , Bowen used the following specification property of uniformly hyperbolic systems: for every $ \eps> 0$ there is $ \tau\in \NN$ such that any collection of finite-length orbit segments can be $ \eps$-shadowed by a single orbit that takes $ \tau$ iterates to transition from one segment to the next. More precisely, if we associate $ (x,n) \in X\times N$ to the orbit segment $ x,f(x),\dots, f^{n-1}(x)$ and write \begin{eqnarray*} B_n(x,\eps) = \{y\in X \mid d(f^kx,f^ky) \leq \eps \text{ for all } 0\leq k < n\} \end{eqnarray*} for the Bowen ball of points that shadow $ (x,n)$ to within $ \eps$ for those $ n$ iterates, then specification requires that for every $ (x_1,n_1),\dots, (x_k,n_k)$ there is $ y\in X$ such that $ y\in B_{n_1}(x_1,\eps)$, then $ f^{n_1 + \tau}(y)\in B_{n_2}(x_2,\eps)$, and in general
\begin{eqnarray}\label{eqn:spec} f^{\sum_{i=0}^{j-1} (n_i + \tau)}(y) \in B_{n_j}(x,\eps) \text{ for all } 1\leq j\leq k. \end{eqnarray} | (5.2) |
A continuous potential $ \ph\colon X\to \RR$ satisfies the Bowen property if there is $ K\in \RR$ such that $ |S_n\ph(x) - S_n\ph(y)| < K$ whenever $ y\in B_n(x,\eps)$, where $ S_n\ph(x) = \sum_{j=0}^{n-1} \ph(f^jx)$. The following theorem summarizes the classical results due to Bowen on systems with specification [Bowen1974]. In fact, Bowen required the slightly stronger property that the shadowing point $ y$ in (5.2) be periodic, but this is only necessary for the part of his results dealing with periodic orbits, which we omit here. × 32
Various weaker versions of the specification property have been introduced in the literature. The one which is most relevant for our purposes first appeared in [Climenhaga and Thompson2012] for MMEs in the symbolic setting, and was developed in [Climenhaga and Thompson2013], [Climenhaga and Thompson2014], [Climenhaga and Thompson2016] to a version that applies to smooth maps and flows.
Given $ \eps> 0$, consider the 'non-expansive set' $ \NE(\eps) = \{x\in X \mid \Gamma_\eps(x) \neq \{x\}\}$, where $ \Gamma_\eps(x)$ is as in (5.1) . Note that $ (X,f)$ is expansive if and only if $ \NE(\eps)=\emptyset$. The pressure of obstructions to expansivity is The idea of ignoring measures sitting on $ \NE(\eps)$ was introduced earlier by [Buzzi and Fisher2013]. × 33
\begin{eqnarray}\label{eqn:pexp} \Pexp(\ph) = \lim_{\eps\to 0} \sup_{\mu \in \MMM^e(f)} \left\{h_\mu(f) + \int \ph\,d\mu \mid \mu(\NE(\eps))=1\right\}. \end{eqnarray} | (5.3) |
Let us make this more precise. A decomposition of the space of orbit segments consists of $ \PPP, \GGG, \SSS \subset X\times \NN$ and functions $ p,g,s\colon X\times \NN\to \NN\cup \{0\}$ such that $ (p+g+s)(x,n) = n$ and \begin{align*} (x,p(x,n))&\in \PPP,\\ (f^{p(x,n)}(x),g(x,n)) &\in \GGG, \\ (f^{(p+g)(x,n)}(x),s(x,n))& \in \SSS. \end{align*} The following is ([Climenhaga and Thompson2016], Theorem 5.5).
We describe two examples for which Theorem 5.2 applies. One of them is the Mañé example [Mañé1978], which was introduced as an example of a robustly transitive diffeomorphism that is not Anosov. This "derived from Anosov" example is obtained by taking a 3-dimensional hyperbolic toral automorphism with one unstable direction and performing a pitchfork bifurcation in $ E^{cs}$ near the fixed point so that $ E^c$ becomes weakly expanding in that neighborhood. One obtains a partially hyperbolic diffeomorphism with a splitting $ E^s\oplus E^c\oplus E^u$ such that $ E^c$ "contracts on average" with respect to the Lebesgue measure; this falls under the results in [Castro2004] mentioned above, and its inverse map (for which $ E^c$ "expands on average") is covered by [Alves and Pinheiro2010], [Alves and Li2015].
Now given any Hölder continuous potential $ \ph\colon \mathbb{T}^3\to \RR$, it is shown in [Climenhaga et al.2015] that there is a $ C^1$-open class of Mañé examples for which this potential has a unique equilibrium state. In particular, when $ f$ is $ C^2$, there is an interval $ (t_0,t_1) \supset [0,1]$ such that the geometric $ t$-potential $ -t\log|\det(df|E^{cu})|$ has a unique equilibrium state for every $ t\in (t_0,t_1)$, and $ \ph_1$ is the unique SRB measure.
A related second example is the Bonatti–Viana example introduced in [Bonatti and Viana2000]. Here one takes a 4-dimensional hyperbolic toral automorphism with $ \dim E^s=\dim E^u=2$, and makes two perturbations, one in the $ E^s$-direction and another one in the $ E^u$-direction. The first perturbation creates a pitchfork bifurcation as above in $ E^s$ and then "mixes up" the two directions in $ E^s$ so that there is no invariant subbundle of $ E^s$; the second perturbation does a similar thing to $ E^u$. One obtains a map with a dominated splitting $ E^{cs}\oplus E^{cu}$ that has no uniformly hyperbolic subbundles.
The same approach as above works for the Bonatti–Viana examples, which have a dominated splitting but are not partially hyperbolic; see [Climenhaga et al.2015]. In this case the presence of non-uniformity in both the stable and unstable directions makes tower constructions more difficult, and no Gibbs-Markov-Young structure has been built for these examples. Earlier results on thermodynamics of these examples (and the Mañé examples) were given in [Buzzi et al.2012], [Buzzi and Fisher2013], which proved existence of a unique MME. These results make strong use of the semi-conjugacy between the examples and the original toral automorphisms, and in particular do not generalize to equilibrium states corresponding to non-zero potentials.
Finally, the flow version of Theorem 5.2 can be applied to geodesic flow in nonpositive curvature. Geodesic flow in negative curvature is one of the classical examples of an Anosov flow [Anosov1969], and in particular it has unique equilibrium states with strong statistical properties. Although the issue of decay of correlations is more subtle because it is a flow, not a map; see [Dolgopyat1998], among others. × 35 If $ M$ is a smooth rank 1 manifold with nonpositive curvature, then its geodesic flow is non-uniformly hyperbolic. Bernoullicity of the regular component of the Liouville measure was shown by [Pesin1977]. It was shown by [Knieper1998] that there is a unique measure of maximal entropy; his proof uses powerful geometric tools and does not seem to generalize to non-zero potentials. Using non-uniform specification, Knieper's result can be extended to the geometric $ t$-potential for $ t\approx 0$, and when $ \dim M = 2$, it works for any $ t\in (-\infty,1)$, showing that the pressure function is differentiable on this interval and we recover the same picture as for Manneville–Pomeau [Burns et al.2016].
In each of the above examples, the basic idea is as follows: one identifies a "bad set" $ B\subset X$ with the properties that
Given an orbit segment $ (x,n)$, let $ G(x,n) = \frac 1n\#\{0\leq k< n \mid f^kx \notin B\}$ be the proportion of time that the orbit segment spends in the "good" part of the system. For flows one should make the obvious modifications, replacing $ \NN$ by $ [0,\infty)$ and cardinality with Lebesgue measure. × 36 A decomposition of the space of orbit segments $ X\times \NN$ is obtained by fixing a threshold $ \gamma> 0$ and taking \begin{align*} \PPP &= \SSS = \{(x,n) \mid G(x,n) < \gamma\}, \\ \GGG &= \{(x,n) \mid G(x,k) \geq \gamma, G(f^kx,n-k)\geq \gamma \text{ for all } 0\leq k \leq n \}. \end{align*} Indeed, given any $ (x,n)\in X\times \NN$, one can take $ p$ and $ s$ to be maximal such that $ (x,p) \in \PPP$ and $ (f^{n-s}x,s)\in \SSS$, and use additivity of $ G$ along orbit segments to argue that $ (f^px,n-p-s)\in \GGG$, which yields a decomposition $ X\times \NN = \PPP\GGG\SSS$. Then one makes the following arguments to apply Theorem 5.2.
Having discussed constructions of SRB and equilibrium measures via Markov dynamics (SFTs and Young towers) and via coarse-graining (expansivity and specification), we turn our attention now to a third approach, which is in some sense more natural and more simple-minded. The first two approaches addressed not just existence but also questions of uniqueness and statistical properties; the price to be paid for these stronger results is that the construction of a tower (or even the verification of non-uniform specification) may be difficult in many examples. The approach that we now describe is best suited to prove existence, rather than uniqueness or statistical properties, but has the advantage that it seems easier to verify.
We start by discussing SRB measure, which for dissipative systems plays the role of Lebesgue measure in conservative systems and is the most natural measure from the physical point of view. So in trying to find an SRB measure, it is natural to start with Lebesgue measure itself; while it may not be invariant, we will follow the standard Bogolubov–Krylov procedure of taking a non-invariant measure $ m$, average it under the dynamics to produce the sequence
\begin{eqnarray}\label{evol} \mu_n = \frac1n\sum_{k=0}^{n-1}f_*^km, \end{eqnarray} | (6.1) |
At an intuitive level, this approach is consistent with Viana's conjecture ([Viana1998]) that nonzero Lyapunov exponents imply existence of an SRB, since this should be exactly the setting in which the iterates of Lebesgue spread out along the unstable manifolds and converge in average to a measure that is absolutely continuous in the unstable direction. Now we describe how it can be made precise.
In the uniformly hyperbolic setting, this approach can be carried out as follows. Let $ \RRR$ be the set of all standard pairs $ (W,\rho)$, where $ W$ is a small piece of unstable manifold and $ \rho\colon W\to (0,\infty)$ is integrable with respect to $ m_W$, the leaf volume on $ W$. Let $ \Mac$ be the set of all (not necessarily invariant) probability measures $ \mu$ on the manifold $ M$ that can be expressed as
\begin{eqnarray}\label{eqn:Mac} \mu(E)=\int_\RRR\int_{W\cap E} \rho(x)\,dm_W(x)\,d\zeta(W,\rho) \end{eqnarray} | (6.2) |
In order to pass to the limit and obtain $ \mu\in\Mac$ one needs a little more control. Fixing $ K> 0$, let $ \RRR_K$ be the set of all standard pairs $ (W,\rho)$ such that $ W$ has size at least $ 1/K$, and $ \rho\colon W\to [1/K,K]$ is Hölder continuous with constant $ K$. Then defining $ \Mac_K$ using (6.2) with $ \RRR_K$ in place of $ \RRR$, one can show that $ \Mac_K$ is weak* compact and is $ f_*$-invariant for large enough $ K$. This is basically a consequence of the Arzelà–Ascoli theorem and the fact that $ f$ uniformly expands unstable manifolds; in particular it relies strongly on the uniform hyperbolicity assumption. Then $ \mu_n\in \Mac_K$ for all $ n$ by invariance, and by compactness, $ \mu=\lim \mu_{n_j}\in\Mac_K\cap\MMM(f)$ is an SRB measure. Thus we have the following statement.
In the situation where the centre-unstable direction $ E^{cu}$ is only non-uniformly expanding more care must be taken with the above approach because $ \Mac_K$ may no longer be $ f_*$-invariant: even if $ W$ is a "sufficiently large" local unstable manifold, its image $ f(W)$ may be smaller than $ 1/K$, and similarly the Hölder constant of the density $ \rho$ can get worse under the action of $ f_*$ if $ W$ is contracted by $ f$.
The solution is to use hyperbolic times , which were introduced by [Alves2000]. Roughly speaking, a time $ n$ is hyperbolic for a point $ x$ if $ df^k|E^u(f^{n-k}x)$ is uniformly expanding for every $ 0\leq k\leq n$. If $ W$ is a local unstable manifold around $ x$ and $ n$ is a hyperbolic time for $ x$, then $ f^n(W)$ contains a large neighborhood of $ f^n(x)$, and the density $ \rho$ behaves well under $ f_*^n$. Thus from the point of view of the construction above, the key property of hyperbolic times is that if $ H_n$ is the set of all points $ x$ for which $ n$ is a hyperbolic time, then the measures
\begin{eqnarray}\label{eqn:nun} \nu_n := \frac 1n \sum_{k=0}^{n-1} f_*^k(m|H_k) \end{eqnarray} | (6.3) |
One can also construct the SRB measure by beginning "within the attractor": instead of using Lebesgue measure on $ M$ as the starting point for the sequence (6.1) , one can let $ m^u$ be leaf volume along a local unstable manifold and then consider the sequence
\begin{eqnarray}\label{evol1} \nu_n(x)=\frac1n\sum_{k=0}^{n-1}\,f_*^km^u(x). \end{eqnarray} | (6.4) |
Several results are available that establish existence (and in some cases uniqueness) of SRB measures under some additional requirements on the action of the system along the central direction $ E^c$ or central-unstable direction $ E^{cu}$. For example the case of systems with mostly contracting central directions was carried out in [Bonatti and Viana2000], [Burns et al.2008] and with mostly expanding central directions in [Alves et al.2000]. A more general case of systems whose central direction is weakly expanding was studied in [Alves et al.2014].
In these settings one at least has a dominated splitting, which gives the system various uniform geometric properties, even if the dynamics is non-uniform. To extend this approach to settings where the geometry is non-uniform (no dominated splitting, stable and unstable directions vary discontinuously) some new tools are needed. An important observation (which holds in the uniform case as well) is that for many purposes we can replace $ V^u(x)$ itself with a local manifold passing through $ x$ that is $ C^1$-close to $ V^u(x)$. Such a manifold is called admissible , and in the next section will develop the machinery of standard pairs, the class of measures $ \Mac$, and the sequences of measures (6.4) using admissible manifolds in place of unstable manifolds.
The difficulties encountered in the geometrically non-uniform setting can be overcome by the machinery of 'effective hyperbolicity' from [Climenhaga and Pesin2016], [Climenhaga et al.2016]. This approach has the advantage that the requirements on the system appear weaker, and much closer to the Viana conjecture. The drawback of this approach is that it is currently out of reach to use it to establish exponential (or even polynomial) decay of correlations and the CLT.
Let $ U$ be a neighborhood of the attractor $ \Lambda$ for a $ C^{1+\epsilon}$ diffeomorphism, and consider a forward invariant set $ S\subset U$ on which there are two measurable cone families $ K^s(x)=K^s(x,E^s(x),\theta)$ and $ K^u(x)=K^u(x,E^s(x),\theta)$ that are
\begin{eqnarray}\label{eqn:lambdaspsx} \lambda(x) := \min(\lambda^u(x) - \Delta(x), - \lambda^s(x)); \end{eqnarray} | (6.5) |
With the above notions in mind, we consider the following set of points: \begin{eqnarray*} S' = \left\{x\in S \mid \llim_{n\to\infty} \frac 1n \sum_{k=0}^{n-1} \lambda(f^kx) > 0 \text{ and } \lim_{\ba\to 0} \rho_{\ba}(x) = 0 \right\}. \end{eqnarray*} Thus $ S'$ contains points for which the average asymptotic rate of effective hyperbolicity is positive, and for which the asymptotic frequency with which the angle between the cones degenerates can be made arbitrarily small. Then we have the following result, which is a step towards Viana's conjecture.
The construction of an SRB measure in the setting of Theorem 6.2 follows the same averaging idea as in Sects. 6.1.1, 6.1.2: if $ \mu_n$ is the sequence of measures given by (6.1) , then one wish to show that a uniformly large part of $ \mu_n$ lies in the set of "uniformly absolutely continuous" measures $ \Mac_K$.
In this more general setting the definition of $ \Mac_K$ is significantly more involved. Broadly speaking, in the definition of $ \RRR$ we must replace unstable manifolds $ W$ with admissible manifolds ; an admissible manifold $ W$ through a point $ x\in S$ is a smooth submanifold such that $ T_xW \subset K^u(x)$ and $ W$ is the graph of a function $ \psi\colon B^u(r)\subset E^u(x) \to E^s(x)$, such that $ D\psi$ is uniformly bounded and is uniformly Hölder continuous. The Hölder constant for $ D\psi$ can be thought of as the "curvature" of $ W$.
When the geometry is uniform as in the previous setting, the image of an admissible manifold $ W$ is itself admissible; this is essentially the classical Hadamard–Perron theorem. In the more general case this is no longer true; although $ f^n(W)$ contains an admissible manifold, its size and curvature may vary with time $ n$, with the size becoming arbitrarily small and the curvature arbitrarily large. In this setting a version of the Hadamard–Perron theorem was proved in [Climenhaga and Pesin2016] that gives good bounds on $ f^n(W)$ when $ n$ is an effective hyperbolic time for $ x\in W$; that is, when \begin{eqnarray*} \sum_{j=k}^{n-1}\lambda(f^jx)\ge\chi(n-k) \end{eqnarray*} for every $ 0\le k < n$, where $ \chi> 0$ is a fixed rate of effective hyperbolicity .
The set of effective hyperbolic times is a subset of the set of hyperbolic times; the extra conditions in the definition of effective hyperbolic time guarantee that we can control the dynamics of $ f$ on the manifold itself, not just the dynamics of $ df$ on the tangent bundle. In the uniform geometry setting from earlier, this extension came for free for hyperbolic times.
With the notion of effective hyperbolic times, the approach outlined in Sects. 6.1.1, 6.1.2 can be carried out. One must add some more conditions to the collection $ \RRR$; most notably, one must fix $ n\in \NN$ and then consider only admissible manifolds $ W$ for which \begin{eqnarray*} d(f^{-k}(x),f^{-k}(y))\le Ce^{-\chi k}d(x,y)\quad \text{ for all} \quad 0\le k\le n \quad \text{ and} \quad x,y\in W, \end{eqnarray*} and then define $ \Mac_{K,n}$ using only this class of admissible manifolds. In addition to size of $ W$ and regularity of $ \rho$, the constant $ K$ must also be chosen to govern the curvature of $ W$, but we omit details here. The point is that the set $ \Mac_{K,n}$ is compact, but not $ f_*$-invariant, and so the proof of Theorem 6.2 can be completed via the following steps.
A natural next step is to extend the above procedure to study general equilibrium states, and not just SRB measures. The direct analogue of the previous section has not yet been fully developed, and we describe instead a related approach that is also based on studying how densities transform under the dynamics.
First consider the case of a piecewise expanding interval map, and the question of finding an SRB measure. In this case there is no stable direction, and so we do not have to keep track of the "shape" of unstable manifolds, or admissible manifolds; indeed, a local unstable manifold is just a small piece of the interval, and an SRB measure is just an invariant measure that is absolutely continuous with respect to Lebesgue. Thus the entire problem is reduced to the following question: given a (not necessarily invariant) absolutely continuous measure $ \mu\ll m$, how is the density function of its image $ f_*\mu$ related to the density function of $ \mu$? One ends up defining a transfer operator $ \mathcal{L}$ with the property that if $ d\mu = h\,dx$, then $ d(f_*\mu)=(\mathcal{L} h)\,dx$. Questions about the existence of an absolutely continuous invariant measure, and its statistical properties, can be reduced to questions about the transfer operator $ \LLL$.
The central issue in studying $ \LLL$ is the problem of finding a Banach space $ \BBB$ (of functions) on which $ \LLL$ acts "with good spectral properties". Generally speaking this means that 1 is a simple eigenvalue of $ \LLL$ (so there is a unique fixed point $ h=\LLL h$, which corresponds to the unique absolutely continuous invariant measure) and the rest of the spectrum of $ \LLL$ lies inside a disc of radius $ r< 1$, which guarantees exponential decay of correlations and other statistical properties.
For piecewise expanding interval maps, this was accomplished by [Lasota and Yorke1974], and the approach can be adapted to equilibrium states for other potential functions by considering a transfer operator that depends on the potential in an appropriate way. A thorough account of this approach is given in [Baladi2000].
The mechanism that drives this approach is that the expansion of the dynamics acts to "smooth out" the density function; irregularities in the function $ h$ are made milder by passing to $ \LLL h$. (The precise meaning of this statement depends on the particular choice of Banach space $ \BBB$, and is encoded by the Lasota–Yorke inequality, which we do not pursue further here.) But this means that one runs into problems when going from expanding interval maps to hyperbolic diffeomorphisms, where there is a non-trivial stable direction; the contracting dynamics in the stable direction make irregularities in the function worse!
In the classical approach to uniformly hyperbolic systems, this was dealt with by passing to a symbolic coding by an SFT (as described after Theorem 1.1) and then replacing the two-sided SFT $ \Sigma \subset A^\ZZ$ by its one-sided version $ \Sigma^+ \subset A^\NN$. As described after Theorem 1.3, the transfer operator $ \LLL$ has an eigenfunction $ h\in C(\Sigma^+)$, and its dual $ \LLL^*$ has an eigenmeasure $ \nu \in C(\Sigma^+)^*$; combining them gives the equilibrium state $ d\mu = h\,d\nu$. Note that positive indices of an element of $ \Sigma$ code the future of a trajectory, while negative indices code the past, and so dynamically, passing from $ \Sigma$ to $ \Sigma^+$ can be interpreted as "forgetting the past". Geometrically, this means that we conflate points lying on the same local stable manifold; taking a quotient in the stable direction eliminates the problem described in the previous paragraph, where contraction in the stable direction exacerbates irregularities in the density function.
More recent work has shown that this problem can be addressed without the use of symbolic dynamics. The key is to consider a Banach space $ \BBB$ whose elements are not functions, but are rather objects that behave like functions in the unstable direction, and like distributions in the stable direction. For SRB measures, this was carried out in [Blank et al.2002], [Gouëzel and Liverani2006], [Baladi and Tsujii2007]. A further generalization to equilibrium states for other potential functions was given in [Gui and Li2008]; as with expanding interval maps, this requires working with a transfer operator $ \LLL$ that depends on the potential. Moreover, instead of distributions along the stable direction, one must consider a certain class of "generalized differential forms". We refer the reader to ([Gui and Li2008], Sect. 7) for a comparison of this approach to equilibrium states and other related approaches, including the technique of "standard pairs".
It remains an open problem to extend this approach to the non-uniformly hyperbolic setting.
An important open question is to study uniqueness and statistical
properties of the SRB measure produced in Theorem 6.2, or of any
equilibrium states that may be produced by an analogous result for other
potentials. One potential approach is to study the standard pairs
$ (W,\rho)$ and derive statistical properties via coupling
techniques, as was done by Chernov and Dolgopyat in another setting ([Chernov and Dolgopyat2009]). Coupling
techniques are also at the heart of Young's tower results for subexponential
mixing rates ([Young1999]). ×
38 One might also hope to adapt the
functional analytic approach from Sect. 6.2 into
the non-uniformly hyperbolic setting and obtain statistical properties this way.
For now, though, we only mention results on Bernoullicity and hyperbolic
product structure.
By a result of [Ledrappier1984], a hyperbolic SRB measure has at
most countably many ergodic components and every hyperbolic SRB measure
is Bernoulli up to a finite period. It follows that there may exist at most
countably many ergodic SRB measures on $ \Lambda$. One way to ensure
uniqueness of SRB measures is to show that its every ergodic component is
open (mod 0) in the topology of $ \Lambda$ and that $ f|\Lambda$ is
topologically transitive.
Let $ \mu$ be a hyperbolic ergodic measure for a
$ C^{1+\alpha}$ diffeomorphism $ f$. Given $ \ell> 0$,
consider the regular set $ \Gamma_\ell$, which consists of points
$ x\in \Gamma$ whose local stable $ V^s(x)$ and unstable $ V^u(x)$
manifolds have size at least $ 1/\ell$. For $ x\in \Gamma_\ell$ and some
sufficiently small $ r> 0$ let $ R_\ell(x,r)=\bigcup_{y\in A^u(x)}V^s(y)$ be a rectangle
at $ x$, where $ A^u(x)$ is the set of points of intersection of
$ V^u(x)$ with local stable manifolds $ V^s(z)$ for $ z\in\Gamma_\ell\cap B(x,r)$.
We denote by If true, this would imply that $ \mu$ has some "nice" ergodic
properties; for example, it has at most countably many ergodic components.
Similar results have recently been established (using the symbolic approach)
for two-dimensional diffeomorphisms and three-dimensional flows ([Sarig2011], [Ledrappier et al.2016]). We conclude with a conjecture on the relationship between effective
hyperbolicity (from Sect. 6.1) and decay of correlations. Suppose that
$ \Lambda$ is an attractor with trapping region $ U$, and that
we have invariant measurable transverse cone families defined Lebesgue-a.e. in
$ U$, with the property that there is $ \chi> 0$ for which
\begin{eqnarray*} S'=\left\{x\in U\mid\llim_{n\to\infty}\frac1n\sum_{k=0}^{n-1}\lambda(f^kx)> \chi\text{ and } \lim_{\ba\to 0}\rho_{\ba}(x)=0\right\} \end{eqnarray*} has full Lebesgue measure in $ U$. Consider for each
$ N\in \NN$ the set \begin{eqnarray*} X_N=\left\{x\in U\mid\sum_{k=0}^{n-1}\lambda(f^kx)> \chi n\text{ for all } n> N\right\}, \end{eqnarray*} and note that the assumption on
$ S'$ guarantees that $ m(U{\setminus} X_N)\to 0$ as $ N\to\infty$. Some support for this conjecture is provided by the fact that the
analogous result for partially hyperbolic attractors with mostly expanding
central direction was proved in [Alves and Li2015].
6.3.1. SRB Measures
6.3.2. Equilibrium Measures
Say that $ \mu$ has a direct product
structure
if the holonomy map is absolutely continuous with the Jacobian uniformly
bounded away from 0 and $ \infty$ on $ R_\ell(x,r)$. Conjecture 6.3.
If $ \mu$ is a hyperbolic ergodic
equilibrium measure for the geometric $ t$-potential for a
$ C^{1+\alpha}$ diffeomorphism $ f$, then $ \mu$ has a
direct product structure. Conjecture 6.4.
If $ m(U{\setminus} X_N)$ decays exponentially in
$ N$, then the SRB measure $ \mu$ produced by Theorem
6.2 has
exponential decay of correlations.
V. Climenhaga is partially supported by NSF Grant DMS-1362838. Y. Pesin is partially supported by NSF Grant DMS-1400027