Hostname: page-component-7bb8b95d7b-cx56b Total loading time: 0 Render date: 2024-10-05T02:17:50.540Z Has data issue: false hasContentIssue false

Dimension estimates and approximation in non-uniformly hyperbolic systems

Published online by Cambridge University Press:  12 February 2024

JUAN WANG
Affiliation:
School of Mathematics, Physics and Statistics, Shanghai University of Engineering Science, Shanghai 201620, P.R. China (e-mail: [email protected])
YONGLUO CAO
Affiliation:
Department of Mathematics, Soochow University, Suzhou 215006, Jiangsu, P.R. China Center for Dynamical Systems and Differential Equations, Soochow University, Suzhou 215006, Jiangsu, P.R. China (e-mail: [email protected])
YUN ZHAO*
Affiliation:
Center for Dynamical Systems and Differential Equations, Soochow University, Suzhou 215006, Jiangsu, P.R. China (e-mail: [email protected]) School of Mathematical Sciences, Soochow University, Suzhou 215006, Jiangsu, P.R. China
Rights & Permissions [Opens in a new window]

Abstract

Let $f: M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$-dimensional compact smooth Riemannian manifold M and $\mu $ a hyperbolic ergodic f-invariant probability measure. This paper obtains an upper bound for the stable (unstable) pointwise dimension of $\mu $, which is given by the unique solution of an equation involving the sub-additive measure-theoretic pressure. If $\mu $ is a Sinai–Ruelle–Bowen (SRB) measure, then the Kaplan–Yorke conjecture is true under some additional conditions and the Lyapunov dimension of $\mu $ can be approximated gradually by the Hausdorff dimension of a sequence of hyperbolic sets $\{\Lambda _n\}_{n\geq 1}$. The limit behaviour of the Carathéodory singular dimension of $\Lambda _n$ on the unstable manifold with respect to the super-additive singular valued potential is also studied.

Type
Original Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1 Introduction

Hyperbolic approximation plays a fundamental role in the study of smooth dynamical systems. Roughly speaking, for a hyperbolic ergodic measure $\mu $ of positive entropy, one can always find a sequence of horseshoes $\{\Lambda _n\}_{n\ge 1}$ so that the dynamical quantities on them are close to the corresponding ones of the measure $\mu $ . Such results can be traced back to the landmark work by Katok [Reference Katok18] or Katok and Hasselblatt [Reference Katok and Hasselblatt19]. An earlier related work was obtained by Misiurewicz and Szlenk [Reference Misiurewicz and Szlenk25] for piecewise continuous and monotone maps of interval. For more results of this type, we would like to refer the reader to [Reference Avila, Crovisier and Wilkinson2, Reference Cao, Pesin and Zhao8, Reference Chung10, Reference Gelfert14, Reference Gelfert15, Reference Persson and Schmeling27, Reference Przytycki and Urbański30, Reference Wang, Qu and Cao34, Reference Yang35] and the references therein.

From the point of dimension theory of dynamical systems, it is natural and non-trivial to use Hausdorff dimension to estimate how large that part of the dynamics described by these horseshoes is. If $\mu $ is an ergodic hyperbolic Sinai–Ruelle–Bowen (SRB) measure of a surface diffeomorphism, Mendoza [Reference Mendoza24] proved that the Hausdorff dimension of the horseshoes on the unstable manifolds approaches to one. For the higher dimensional case, Sánchez-Salas [Reference Sánchez-Salas31] proved that the measure $\mu $ can be approximated in the weak topology by ergodic measures supported on the horseshoes $\{\Lambda _n\}_{n\ge 1}$ . Moreover, he established some interesting results concerning the Hausdorff dimension of the horseshoes. Using Cao, Pesin and Zhao’s ideas [Reference Cao, Pesin and Zhao8], Wang, Qu and Cao [Reference Wang, Qu and Cao34] generalized Mendoza’s result [Reference Mendoza24] for diffeomorphisms on a higher dimensional manifold. In fact, the authors proved that the Hausdorff dimension of the horseshoes $\{\Lambda _n\}_{n\ge 1}$ on the unstable manifold tends to the dimension of the unstable manifold. Furthermore, if the stable direction is one dimension, then the Hausdorff dimension of the measure $\mu $ can be approximated by the Hausdorff dimension of $\{\Lambda _n\}_{n\ge 1}$ . The first result in this paper shows that the Lyapunov dimension of $\mu $ (see equation (1) for the definition) can be approximated gradually by the Hausdorff dimension of a sequence of hyperbolic sets $\{\Lambda _n\}_{n\geq 1}$ , provided that the stable direction is one or $\mu $ satisfies the Pesin’s entropy formula in the stable direction.

The main motivation of our first result is the study of the Kaplan–Yorke conjecture [Reference Frederickson, Kaplan, Yorke and Yorke13]. To be more precise, let $f: M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact smooth Riemannian manifold M and let $\mu $ be a hyperbolic ergodic f-invariant probability measure. For $x\in M$ , the pointwise dimension of $\mu $ at x is defined by

$$ \begin{align*}d_\mu(x)=\lim_{r\to0}\frac{\log\mu(B(x,r))}{\log r},\end{align*} $$

provided the limit exists, where $B(x,r)$ denotes the ball of radius r centred at x. A measure $\mu $ is called exact dimensional if $d_\mu (x)$ is constant almost everywhere and let $\dim _H\mu $ denote the Hausdorff dimension of the measure $\mu $ (see [Reference Pesin29] for the detailed definition). Young [Reference Young36] proved that almost all the known characteristics of dimension type of a measure $\mu $ coincide if $\mu $ is exact dimensional. This indicates that it is very important to show the exactness of a measure in dimension theory of dynamical systems.

Let $\Gamma $ be the set of points which are regular in the sense of Oseledec multiplicative ergodic theorem [Reference Oseledets26]. For every $x\in \Gamma $ , denote the Lyapunov exponents of f at x by

where u and $s:=m_0-u$ are the dimension of the unstable and stable subspaces of $T_xM$ , respectively.

The Lyapunov dimension of $\mu $ is defined as follows:

(1)

where $\ell =\max \{i: \unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _i(\mu )\geq 0\}$ . It is not difficult to show that $\dim _H\mu \leq \dim _L\mu $ , e.g., see [Reference Shu32, Proposition 4.2] for details. It was conjectured in [Reference Frederickson, Kaplan, Yorke and Yorke13] that if $\mu $ is an SRB measure, which is absolutely continuous along the unstable leaves, then generically,

(2) $$ \begin{align} \dim_H\mu=\dim_L\mu. \end{align} $$

By Young’s dimension formula in [Reference Young36], the conjecture is true if M is a surface. This paper proves the conjecture in the higher dimensional case under the assumption that the stable direction is one or $\mu $ satisfies the ‘Pesin’s entropy formula in the stable direction’. Moreover, the measure $\mu $ is exact dimensional in this case (see Theorem A).

To summarize, let $h_{\mu}(f )$ denote the metric entropy of $f$ with respect to $\mu$ (see Walters’ book [Reference Walters33] for details of metric entropy), the first result is stated as the following theorem.

Theorem A. Let $f: M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M and $\mu $ a hyperbolic ergodic SRB measure on M. Assume that either one of the following properties holds:

  1. (i) $\mu $ has a one-dimensional stable manifold;

  2. (ii) $\mu $ satisfies $h_\mu (f)=-\unicode{x3bb} _{u+1}(\mu )-\unicode{x3bb} _{u+2}(\mu )-\cdots -\unicode{x3bb} _{m_0}(\mu )$ ,

then $\dim _H\mu =\dim _L\mu $ . Furthermore, there exists a sequence of hyperbolic sets $\{\Lambda _{n}\}$ such that

$$ \begin{align*} \dim_H \Lambda_{n}\rightarrow \dim_{L}\mu~~(n\to\infty). \end{align*} $$

Example 1.1. Let $f: M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M. Assume that the volume measure $\varrho $ is f-invariant ergodic and hyperbolic. Let

denote the Lyapunov exponents of f with respect to $\varrho $ . By Pesin’s entropy formula [Reference Pesin28] (see also [Reference Mañé23] for a simple proof), one has that

where the second equality holds since f is volume-preserving. By Theorem A, there exists a sequence of hyperbolic sets $\{\Lambda _{n}\}$ such that

$$ \begin{align*} \dim_H \Lambda_{n}\rightarrow m_0~~(n\to\infty), \end{align*} $$

since $\dim _{L}\mu =m_0$ in this case.

Ledrappier [Reference Ledrappier20] proved the existence of the pointwise dimension of each SRB measure. For a hyperbolic invariant measure $\mu $ of a $C^2$ (or $C^{1+\alpha }$ ) diffeomorphism f of a smooth compact Riemannian manifold M without boundary, Ledrappier and Young [Reference Ledrappier and Young22] proved the existence of dimension of $\mu $ on stable/unstable manifolds, and that the upper pointwise dimension of $\mu $ is upper bounded by the sum of the dimension of $\mu $ on stable and unstable manifolds. Later, Barreira, Pesin and Schmeling [Reference Barreira, Pesin and Schmeling4] proved that the lower pointwise dimension of $\mu $ is also lower bounded by the sum of the dimension of $\mu $ on stable and unstable manifolds. This showed that the measure $\mu $ is exact dimensional, which finally solves the Eckmann–Ruelle conjecture.

Motivated by the work in [Reference Fang, Cao and Zhao12], where it is proved that the unique solution of the measure-theoretic pressure is exactly the dimension of an invariant measure supported on an average conformal repeller, the second result in this paper shows that the unique solution of measure-theoretic pressure gives an upper bound of the dimension of a hyperbolic ergodic measure $\mu $ on stable/unstable manifolds. To be more precise, we introduce some notation first. For each $x\in M$ and $n\geq 1$ , consider the differentiable operator ${D_xf^n: T_xM \to T_{f^n(x)}M}$ and denote the singular values of $D_xf^n$ in the decreasing order by

$$ \begin{align*}\alpha_1(x, f^n) \geq \alpha_2(x, f^n) \geq \cdots \ge \alpha_u(x, f^n)\ge \cdots \geq \alpha_{m_0}(x, f^n).\end{align*} $$

Recall that u and s are the dimension of the unstable and stable subspace of $T_xM$ , respectively. For every $t\in [0,u]$ , define

$$ \begin{align*} \phi^t(x,f^n):=\sum_{i=1}^{[t]}\log\alpha_i(x,f^n)+(t-[t])\log\alpha_{[t]+1}(x,f^n) \end{align*} $$

and

$$ \begin{align*} \psi^t(x,f^n):=\sum_{i=u-[t]+1}^{u}\log\alpha_i(x,f^n)+(t-[t])\log\alpha_{u-[t]}(x,f^n). \end{align*} $$

For every $t\in [0, s]$ , define

$$ \begin{align*}\varphi^t(x, f^n) := \sum_{i=u+1}^{u+[t]} \log\alpha_i(x,f^n) + (t-[t])\log\alpha_{u+[t]+1}(x, f^n).\end{align*} $$

Since f is smooth, the functions $x\mapsto \alpha _i(x, f^n)$ , $x\mapsto \phi ^t(x,f^n)$ , $x\mapsto \psi ^t(x,f^n)$ and $x\mapsto \varphi ^t(x,f^n)$ are continuous. It is easy to see that the sequences of functions

(3) $$ \begin{align} \Phi_f(t):=\{-\phi^t(\cdot,f^n)\}_{n\geq1} \end{align} $$

are super-additive and

(4) $$ \begin{align} \Psi_f(t):=\{-\psi^t(\cdot,f^n)\}_{n\geq1}, \quad \Xi_f(t):=\{\varphi^t(\cdot,f^n)\}_{n\geq1} \end{align} $$

are sub-additive. Ledrappier and Young [Reference Ledrappier and Young22] proved the existence of stable and unstable pointwise dimension $d_\mu ^s(x)$ , $d_\mu ^u(x)$ of a hyperbolic ergodic measure $\mu $ for $\mu $ -almost every (a.e.) x. The following theorem shows that the unique solution of the sub-additive measure-theoretic pressure equation

$$ \begin{align*}P_\mu(f, \Psi_f(t))=0 \quad (P_\mu(f,\Xi_f(t))=0)\end{align*} $$

is an upper bound for the unstable (stable) dimension of $\mu $ , see §2 for the definitions of measure-theoretic pressure and stable and unstable dimension of an invariant measure.

Theorem B. Suppose $f: M\rightarrow M$ is a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M and $\mu $ is a hyperbolic ergodic measure on M. Then one has

$$ \begin{align*} d_\mu^u(x) \leq t_u^{*}\quad \text{and}\quad d_\mu^s(x) \leq t_s^* \,\, \mu\text{-a.e.}\,x, \end{align*} $$

where $t_u^*$ and $t_s^*$ are the unique solutions of the equations $P_\mu (f, \Psi _f(t))=0$ and $P_\mu (f,\Xi _f(t))=0$ , respectively.

For each hyperbolic ergodic measure $\mu $ of positive entropy, there exists a sequence of hyperbolic sets $\{\Lambda _n\}_{n\geq 1}$ such that the dynamical quantities on $\Lambda _n$ gradually approach to those of the measure $\mu $ (see Theorem 2.4). Since the hyperbolic sets $\{\Lambda _n\}_{n\geq 1}$ are non-conformal, it is difficult to compute their Hausdorff dimension. Following the approach described in [Reference Cao, Pesin and Zhao8], this paper introduces the concept of Carathéodory singular dimension of a hyperbolic set on unstable manifolds (see §2 for the detailed definition). The third result of this paper shows that the zero of the super-additive/sub-additive measure-theoretic pressure $P_\mu (f,\Phi _f(t))/P_\mu (f,\Psi _f(t))$ gives a lower/upper bound of the Carathéodory singular dimension of $\Lambda _n$ on the unstable manifold. In addition, if $\mu $ is an SRB measure, then the Carathéodory singular dimension of $\Lambda _n$ on the unstable manifold tends to the dimension of the unstable manifold, and the Lyapunov dimension of $\mu $ is exactly the sum of $t_s^*$ and the dimension of the unstable manifold, where $t_s^*$ is the unique root of the equation $P_\mu (f,\Xi _f(t))=0$ .

Theorem C. Let $f: M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M, and let $\mu $ be a hyperbolic ergodic measure on M. Then there exists a sequence of hyperbolic sets $\{\Lambda _{\varepsilon }\}_{\varepsilon \ge 0}$ such that the following properties hold:

  1. (i) ${\liminf _{\varepsilon \to 0}\dim _C^{\Phi _f} (\Lambda _\varepsilon \cap W^u_{\mathrm {loc}}(x,f)) \geq t_{u*}}$ for every $x\in \Lambda _\varepsilon $ , where $t_{u*}$ is the unique root of the equation $P_\mu (f, \Phi _f(t))=0$ ;

  2. (ii) ${\limsup _{\varepsilon \to 0}\dim _C^{\Psi _f} (\Lambda _\varepsilon \cap W^u_{\mathrm {loc}}(x,f)) \leq t_{u}^*}$ for every $x\in \Lambda _\varepsilon $ , where $t_{u}^*$ is the unique root of the equation $P_\mu (f, \Psi _f(t))=0$ .

Furthermore, if $\mu $ is an SRB measure, then $\dim _L\mu =u+t_s^*$ and

$$ \begin{align*}\lim_{\varepsilon\to 0}\dim_C^{\Phi_f} (\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f))=u\end{align*} $$

for every $x\in \Lambda _\varepsilon $ , where u is the dimension of the unstable manifold and $t_s^*$ is the unique root of the equation $P_\mu (f,\Xi _f(t))=0$ .

The paper is organized as follows. Section 2 gives some basic notions and properties, including Hausdorff dimension, hyperbolic set, pressure and singular dimension. All the proofs of the main results will be given in §3.

2 Preliminaries

In this section, we will recall some definitions and preliminary results which are used in the proofs of the main results.

2.1 Hyperbolic set

Let f be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M. We say an f-invariant compact subset $\Lambda \subset M$ is a hyperbolic set if for any $x\in \Lambda $ , the tangent space admits a decomposition ${T_{x}M=E^{s}(x)\oplus E^{u}(x)}$ such that the following properties hold:

  1. (1) the splitting is $Df$ -invariant, that is, for every $x\in \Lambda $ , $D_{x}f E^{\sigma }(x)=E^{\sigma }(f(x))$ for $\sigma =s,u$ ;

  2. (2) the stable subspace $E^{s}(x)$ is uniformly contracting and the unstable subspace $E^{u}(x)$ is uniformly expanding in the sense that there are constants $C\geq 1$ and $0<\chi <1$ such that for every $n\geq 0$ and $v^{\sigma }\in E^{\sigma }(x)$ ( $\sigma =s$ or u), we have

    $$ \begin{align*}\|D_{x}f^{n}v^{s}\|\leq C\chi^{n}\|v^{s}\|\quad \text{and} \quad \|D_{x}f^{-n}v^{u}\|\leq C\chi^{n}\|v^{u}\|.\end{align*} $$

Recall that a hyperbolic set $\Lambda $ is locally maximal if there exists an open neighbourhood U of $\Lambda $ such that $\Lambda =\bigcap _{n\in \mathbb {Z}}f^{n}(U)$ , and a diffeomorphism f is called topologically transitive on $\Lambda $ if for every two non-empty (relative) open subsets $U,V\subset \Lambda $ , there exists $n>0$ such that $f^n(U)\cap V\neq \emptyset $ . Given a point $x\in \Lambda $ , for each small $\beta>0$ , the local stable and unstable manifolds at the point x are defined as follows:

$$ \begin{align*} W^{s}_{\mathrm{loc}}(x, f) = \{y\in M:d(f^{n}(x),f^{n}(y))\leq \beta \text{ for all } n\geq 0\}, \end{align*} $$

and

$$ \begin{align*} W^{u}_{\mathrm{loc}}(x, f) = \{y\in M:d(f^{-n}(x),f^{-n}(y))\leq \beta \text{ for all } n\geq 0\}. \end{align*} $$

The global stable and unstable sets of $x\in \Lambda $ are given as follows:

$$ \begin{align*} W^{s}(x, f) = \bigcup_{n\geq 0}f^{-n}(W^{s}_{\mathrm{loc}}(f^{n}(x), f)), \quad W^{u}(x, f) = \bigcup_{n\geq 0}f^{n}(W^{u}_{\mathrm{loc}}(f^{-n}(x), f)). \end{align*} $$

Let $d^s/d^u$ be the metric induced by the Riemannian structure on the stable/unstable manifold $W^s/W^u$ .

2.2 Dimension

Let X be a compact Riemannian manifold with a Riemannian metric. Given a subset Z of X, for $s\geq 0$ and $\delta>0$ , define

$$ \begin{align*} \mathcal{H}_{\delta}^{s}(Z) := \inf \bigg\{\!\sum_{i}|U_i|^s: \ Z\subset \bigcup_{i}U_i,~|U_i|\leq \delta \text{ for all } i\bigg\}, \end{align*} $$

where $|\cdot |$ denotes the diameter of a subset. The quantity

$$ \begin{align*}\mathcal{H}^{s}(Z) := \lim\limits_{\delta\rightarrow 0}\mathcal{H}_{\delta}^{s}(Z)\end{align*} $$

is called the s-dimensional Hausdorff measure of Z. It is easy to show that there is a jump-up value

$$ \begin{align*} \dim_H Z :=\inf \{s:\ \mathcal{H}^{s}(Z)=0\}=\sup \{s: \mathcal{H}^{s}(Z)=\infty\}, \end{align*} $$

which is called the Hausdorff dimension of Z.

Given a Borel probability measure $\mu $ on X, the Hausdorff dimension of the measure  $\mu $ is defined as

$$ \begin{align*}\dim_{H}\mu=\inf\{\dim_{H}Y:Y\subset X ,~\mu(Y)=1\}.\end{align*} $$

The lower and upper pointwise dimension of $\mu $ at point $x\in X$ are defined respectively by

$$ \begin{align*} \underline{d}_\mu(x)=\liminf_{r\to 0}\frac{\log\mu(B(x,r))}{\log r}\quad \text{and}\quad \overline{d}_\mu(x)=\limsup_{r\to 0}\frac{\log\mu(B(x,r))}{\log r}, \end{align*} $$

where $B(x,r)$ denotes the ball of radius r centred at x. If $\underline {d}_\mu (x)=\overline {d}_\mu (x)$ , then we denote the common value by $d_\mu (x)$ . In particular, Barreira and Wolf [Reference Barreira and Wolf5] proved that

(5) $$ \begin{align} \dim_H\mu=\text{ess sup}\{\underline{d}_\mu(x): x\in X\}, \end{align} $$

where the essential supremum is taken with respect to $\mu $ . The following well-known result gives the relation between the Hausdorff dimension and the lower pointwise dimension.

Proposition 2.1. The following properties hold:

  1. (1) if $\underline {d}_\mu (x)\geq \alpha $ for $\mu $ -a.e. $x\in X$ , then $\dim _H\mu \geq \alpha $ ;

  2. (2) if $\underline {d}_\mu (x)\leq \alpha $ for every $x\in Z\subseteq X$ , then $\dim _HZ\leq \alpha $ .

Let $f:X\rightarrow X$ be a $C^{1+\alpha }$ diffeomorphism on an $m_{0}$ -dimensional compact Riemannian manifold X, and let $\mu $ be a hyperbolic ergodic measure on X. Let $\Gamma $ be the set of points which are regular in the sense of Oseledets [Reference Oseledets26]. A measurable partition $\xi ^u$ / $\xi ^s$ of X is said to be subordinate to the unstable/stable manifold if for $\mu $ -almost every x, $\xi ^u(x)\subset W^u(x,f)$ / $\xi ^s(x)\subset W^s(x, f)$ and contains an open neighbourhood of x in $W^u(x,f)$ / $W^s(x, f)$ . Let $\{\mu _x^u\}$ and $\{\mu _x^s\}$ be the collections of conditional measures associated with $\xi ^u$ and $\xi ^s$ , respectively. For every $x\in \Gamma $ , Ledrappier and Young [Reference Ledrappier and Young22] proved the existence of the following limits:

(6) $$ \begin{align} d_\mu^u(x):=\lim_{r\to0}\frac{\log \mu_x^u(B^u(x,r))}{\log r} \quad \text{and} \quad d_\mu^s(x):=\lim_{r\to0}\frac{\log \mu_x^s(B^s(x,r))}{\log r}, \end{align} $$

which are called the stable and unstable dimension of the measure $\mu $ , respectively. Here $B^\sigma (x,r):=\{y\in W^\sigma (x,f): d^\sigma (x,y)<r\}$ with $\sigma \in \{u,s\}$ . Since we consider the limit $r\to 0$ in equation (6), the definition of $d_\mu ^u(x)$ will remain unchanged if we consider the global metric d in the dynamical ball $B^\sigma (x,r)$ instead.

2.3 Pressure

Let $(M,f)$ be a topological dynamical system (TDS for short), that is, $f:M\rightarrow M$ is a continuous map on a compact metric space M equipped with the metric d. Denote by $\mathcal {M}_{\mathrm {inv}}(f|_M)$ and $\mathcal {M}_\mathrm{erg}(f|_M)$ the set of all f-invariant and ergodic Borel probability measures on M, respectively. Given $n\in \mathbb {N}$ and $x,y\in M$ , let

$$ \begin{align*} d_{n}(x,y)=\max\{d(f^{k}(x),f^{k}(y)):0\leq k<n\}. \end{align*} $$

Given $\varepsilon>0$ , denote by $B_{n}(x,\varepsilon )=\{y:d_{n}(x,y)<\varepsilon \}$ the Bowen’s ball of radius $\varepsilon $ centred at x of length n. A subset $E\subset M$ is called $(n,\varepsilon )$ -separated if $d_n(x,y)>\varepsilon $ for any two distinct points $x,y\in E$ . A sequence of continuous functions $\Psi =\{\psi _n\}_{n\geq 1}$ on M is called sub-additive if

$$ \begin{align*} \psi_{m+n}\leq\psi_n+\psi_m\circ f^n\quad \text{for all } m,n\geq1. \end{align*} $$

Similarly, one calls a sequence of continuous functions $\Phi =\{\phi _n\}_{n\geq 1}$ on M super-additive if $-\Phi =\{-\phi _n\}_{n\geq 1}$ is sub-additive.

Let $\Psi =\{\psi _n\}_{n\geq 1}$ be a sub-additive sequence of continuous potentials on M, set

$$ \begin{align*} P_n(f,\Psi, \varepsilon)=\sup\bigg\{\!\sum_{x\in E}e^{\psi_n(x)}: E \text{ is an } (n,\varepsilon) \text{-separated subset of } M\bigg\}. \end{align*} $$

The quantity

$$ \begin{align*} P(f,\Psi):=\lim_{\varepsilon\to0}\limsup_{n\to\infty}\frac1n\log P_n(f,\Psi,\varepsilon) \end{align*} $$

is called the sub-additive topological pressure of $\Psi $ .

The sub-additive topological pressure satisfies the following variational principle, see [Reference Cao, Feng and Huang6] for more details.

Theorem 2.1. Let $\Psi =\{\psi _n\}_{n\geq 1}$ be a sub-additive sequence of continuous potentials on M. Then

$$ \begin{align*} P(f,\Psi)=\sup\{h_{\mu}(f)+\mathcal{F}_*(\Psi,\mu)| ~\mu\in\mathcal{M}_{\mathrm{inv}}(f|_M), \mathcal{F}_*(\Psi,\mu)\neq-\infty\}, \end{align*} $$

where $h_\mu (f)$ is the measure theoretic entropy of f with respect to the measure $\mu $ and ${\mathcal {F}_*(\Psi ,\mu )=\lim _{n\to \infty }(1/n)\int \psi _n\,d\mu }.$

Remark 2.1. If $\Psi =\{\psi _n\}_{n\geq 1}$ is additive in the sense that $\psi _n(x)=\psi (x)+\psi (fx)+\cdots +\psi (f^{n-1}x):= S_n\psi (x)$ for some continuous function $\psi : M\to \mathbb {R}$ , we simply denote the topological pressure $P(f, \Psi )$ as $P(f, \psi )$ .

Next we recall the super-additive topological pressure introduced in [Reference Cao, Pesin and Zhao8] by the variational relation for topological pressure, although it is unknown whether the variational principle holds for super-additive topological pressure defined via separated sets. Given a sequence of super-additive continuous potentials $\Phi =\{\phi _{n}\}_{n\geq 1}$ on M, the super-additive topological pressure of $\Phi $ is defined as

$$ \begin{align*} P(f,\Phi):= \sup \{h_{\mu}(f)+\mathcal{F}_{\ast}(\Phi,\mu):\mu\in \mathcal{M}_{\mathrm{inv}}(f|_M)\}, \end{align*} $$

where

$$ \begin{align*}\mathcal{F}_{\ast}(\Phi,\mu)=\lim_{n\rightarrow \infty}\frac{1}{n}\int \phi_{n}\,d\mu=\sup_{n\in \mathbb{N}} \frac{1}{n}\int \phi_{n}\,d\mu.\end{align*} $$

The second equality is due to the standard sub-additive argument. The following result gives the relation between the sub-additive (super-additive) topological pressure and the topological pressure for additive potentials.

Proposition 2.2. Let $\Phi =\{\phi _{n}\}_{n\geq 1}$ be a sequence of continuous potentials on M. Then the following properties hold:

  1. (1) if $\Phi $ is sub-additive and the entropy map $\mu \mapsto h_\mu (f)$ is upper semi-continuous, then

    $$ \begin{align*}{P(f,\Phi )=\lim _{n\rightarrow \infty } P(f, {\phi _{n}}/{n})=\lim _{n\rightarrow \infty } ({1}/{n}) P(f^{n},\phi _{n})};\end{align*} $$
  2. (2) if $\Phi $ is super-additive, then

    $$ \begin{align*}P(f,\Phi )=\lim _{n\rightarrow \infty } P(f, {\phi _{n}}/{n})=\lim _{n\rightarrow \infty } ({1}/{n}) P(f^{n},\phi _{n}).\end{align*} $$

The first statement is proved in [Reference Ban, Cao and Hu3], where the sub-additive topological pressure is defined via separated sets, so one requires that the entropy map be upper semi-continuous. The second statement is proved in [Reference Cao, Pesin and Zhao8], and one does not need any additional condition since the super-additive topological pressure is defined via the variational relations.

Following the approach described in [Reference Pesin29], we recall the topological pressure on an arbitrary subset of unstable manifolds. Let $f: M\to M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional smooth compact Riemannian manifold M and let $\Lambda \subset M$ be a hyperbolic set. Let $\Psi =\{\psi _n\}_{n\geq 1}$ be a sub-additive sequence of continuous functions on $\Lambda $ . For every $x\in \Lambda $ , denote $Z=\Lambda \cap W^u_{\mathrm {loc}}(x,f)$ . Given $s\in \mathbb {R}$ , set

(7) $$ \begin{align} m(Z,\Psi,s,\delta):=\lim_{N\to\infty} \inf \bigg\{\!\sum_i\exp\Big(-sn_i + \sup_{y\in B_{n_i}^u(x_i,\delta)}\psi_{n_i}(y)\Big)\bigg\}, \end{align} $$

where the infimum is taken over all collections $\{B^u_{n_i}(x_i,\delta )\}$ with $x_i\in \Lambda $ , $n_i\geq N$ that cover Z, and

$$ \begin{align*}B_{n_i}^u(x_i,\delta):=\{y\in W^u(x,f): d^u(f^j(x_i), f^j(y))<\delta\ \text{for } j=0,1,\ldots,n_i-1\}.\end{align*} $$

It is easy to show that there is a jump-up value

$$ \begin{align*}P_Z(f,\Psi,\delta): = \inf\{s: m(Z,\Psi, s,\delta)=0\} = \sup\{s: m(Z,\Psi,s,\delta)=+\infty\}. \end{align*} $$

The quantity

$$ \begin{align*} P_Z(f,\Psi):=\lim_{\delta\to0} P_Z(f,\Psi,\delta) \end{align*} $$

is called the topological pressure of $\Psi $ on the subset Z. It is not difficult to show that $P_\Lambda (f, \Psi )=P(f|_\Lambda , \Psi )$ (see [Reference Cao, Feng and Huang6, Proposition 4.4]).

Let $\mu $ be an f-invariant Borel probability measure on M. Given a sub-additive potential $\Phi =\{\phi _{n}\}_{n \geq 1}$ on M, for $0<\delta < 1, n \geq 1$ and $\varepsilon>0$ , a subset $F\subset M$ is called an $(n,\varepsilon ,\delta )$ -spanning set if the union $\bigcup _{x\in F}B_n(x,\varepsilon )$ has $\mu $ -measure more than or equal to $1-\delta $ . Put

$$ \begin{align*} P_{\mu}(f, \Phi, n, \varepsilon, \delta):=\inf \bigg\{\!\sum_{x \in F} \exp \Big(\sup _{y \in B_{n}(x, \varepsilon)} \phi_{n}(y)\Big) : F \text{ is an } (n, \varepsilon, \delta)\text{-spanning set}\bigg\} \end{align*} $$

and let further that

$$ \begin{align*} P_{\mu}(f, \Phi, \varepsilon, \delta)&:=\limsup _{n \rightarrow \infty} \frac{1}{n} \log P_{\mu}(f, \Phi, n, \varepsilon, \delta), \\ P_{\mu}(f, \Phi, \delta)&:=\liminf_{\varepsilon \rightarrow 0} P_{\mu}(f, \Phi, \varepsilon, \delta), \\ P_{\mu}(f, \Phi)&:=\lim_{\delta \rightarrow 0} P_{\mu}(f, \Phi, \delta), \end{align*} $$

and we call $P_{\mu }(f, \Phi )$ the sub-additive measure-theoretic pressure of $(f,\Phi )$ with respect to $\mu $ . If one considers a super-additive potential $\Phi =\{\phi _n\}_{n\ge 1}$ on M, replacing $\sup _{y \in B_{n}(x, \varepsilon )} \phi _{n}(y)$ by $\phi _{n}(x)$ in $P_{\mu }(f, \Phi , n, \varepsilon , \delta )$ , then the corresponding quantity $P_{\mu }(f, \Phi )$ is called the super-additive measure theoretic pressure of $(f,\Phi )$ with respect to  $\mu $ .

Remark 2.2.

  1. (i) It is easy to see that $P_{\mu }(f, \Phi , \delta )$ increases with $\delta $ decreasing to zero. So the limit in the last formula exists. Moreover, it is proved in [Reference Cao, Hu and Zhao7] that $P_{\mu }(f, \Phi , \delta )$ is independent of $\delta $ . Hence, the limit of $\delta \rightarrow 0$ is redundant in the definition.

  2. (ii) If $\Phi =\{\phi _n\}_{n\ge 1}$ is an additive potential on M, that is, $\phi _n(x)=\sum _{i=0}^{n-1}\phi _1(f^ix)$ for some continuous function $\phi _1$ , then we simply write $P_\mu (f,\Phi )$ as $P_\mu (f,\phi _1)$ .

In the following, we recall some properties of sub-additive/super-additive measure- theoretic pressure which are proved in [Reference Cao, Hu and Zhao7].

Theorem 2.2. [Reference Cao, Hu and Zhao7, Theorem A]

Let $(M,f)$ be a TDS and $\Phi =\{\phi _n\}_{n\geq 1}$ a sub-additive potential on M. For every $\mu \in \mathcal {M}_\mathrm{erg}(f|_M)$ with $\mathcal {F}_*(\Phi , \mu )\neq -\infty $ , we have that

$$ \begin{align*} P_\mu(f,\Phi)=h_\mu(f)+\mathcal{F}_*(\Phi, \mu). \end{align*} $$

Theorem 2.3. [Reference Cao, Hu and Zhao7, Proposition 3.2]

Let $(M,f)$ be a TDS and $\Phi =\{\phi _n\}_{n\geq 1}$ a super-additive potential on M. For every $\mu \in \mathcal {M}_\mathrm{erg}(f|_M)$ , we have that

$$ \begin{align*} P_\mu(f,\Phi)=h_\mu(f)+\mathcal{F}_*(\Phi, \mu). \end{align*} $$

Remark 2.3. In Theorem 2.2, to avoid the indeterminate form $\infty -\infty $ , the condition $\mathcal {F}_*(\Phi , \mu )\neq -\infty $ is necessary. However, we do not need this condition in Theorem 2.3. If $\Phi =\{\phi _n\}_{n\geq 1}$ is an additive potential on M, that is, $\phi _n(x)=S_n\phi (x)$ for some continuous function $\phi $ , then we have

$$ \begin{align*} P_\mu(f,\phi)=h_\mu(f)+\int\phi\,d\mu\quad\text{for all } \mu\in \mathcal{M}_\mathrm{erg}(f|_M). \end{align*} $$

The above formula is also proven in [Reference He, Lv and Zhou16].

2.4 Singular dimension

Let $f: M\to M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact smooth Riemannian manifold M and $\Lambda \subset M$ a hyperbolic set. Consider the sub-additive singular valued potential $\Psi _f(t)=\{-\psi ^t(\cdot , f^n)\}_{n\geq 1}$ given by equation (4). Fix $x\in \Lambda $ and let $Z=\Lambda \cap W_{\mathrm {loc}}^u(x,f)$ . Following the approach described in [Reference Cao, Pesin and Zhao8], we introduce the Carathéodory singular dimension of Z. Put

(8) $$ \begin{align} m(Z, \Psi_f(t),\delta):=\lim_{N\to\infty}\inf\bigg\{\!\sum_i\exp \Big[\sup_{y\in B_{n_i}^u(x_i,\delta)} -\psi^t(y, f^{n_i})\Big]\bigg\}, \end{align} $$

where the infimum is taken over all collections $\{B_{n_i}^u(x_i,\delta )\}$ with $x_i\in \Lambda $ , $n_i\geq N$ that cover Z. It is easy to see that there is a jump-up value

(9) $$ \begin{align} \nonumber \dim_{C,\delta}^{\Psi_f} Z := &\inf\{t: m(Z, \Psi_f(t), \delta)=0 \} \\ = &\sup\{t: m(Z, \Psi_f(t), \delta)=+\infty \}. \end{align} $$

The quantity

(10) $$ \begin{align} \dim_C^{\Psi_f} Z := \lim_{\delta\to0} \dim_{C,\delta}^{\Psi_f} Z \end{align} $$

is called the Carathéodory singular dimension of Z with respect to the sub-additive singular valued potential $\Psi _f$ .

Consider the super-additive singular valued potential $\Phi _f(t)=\{-\phi ^t(\cdot , f^n)\}_{n\geq 1}$ given by equation (3), replacing $ {\sup _{y\in B_{n_i}^u(x_i,\delta )} -\psi ^t(y, f^{n_i})}$ by $-\phi ^t(x_i, f^{n_i})$ in equation (8), one can define $m(Z, \Phi _f(t),\delta )$ and $\dim _{C,\delta }^{\Phi _f} Z$ in a similar fashion as equations (8) and (9). The corresponding quantity $\dim _C^{\Phi _f} Z$ as in equation (10) is called the Carathéodory singular dimension of Z with respect to the super-additive singular valued potential $\Phi _f$ .

2.5 Approximation of hyperbolic measures by hyperbolic sets with dominated splitting

First we recall the definition of the dominated splitting. Let $f:M\rightarrow M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_{0}$ -dimensional compact smooth Riemannian manifold M. Suppose $\Lambda \subset M$ is a compact f-invariant set. We say $\Lambda $ admits a dominated splitting if there is a continuous invariant splitting $T_{\Lambda }M=E\oplus F$ and constants $C>0,\unicode{x3bb} \in (0,1)$ such that for each $x\in \Lambda $ , $n\in \mathbb {N}$ , $0\neq u\in E(x)$ and $0\neq v\in F(x)$ , it holds that

$$ \begin{align*} \frac{\|D_{x}f^{n}(u)\|}{\|u\|}\leq C\unicode{x3bb}^{n}\frac{\|D_{x}f^{n}(v)\|}{\|v\|}. \end{align*} $$

We say F dominates E and write it as $E\preceq F$ . Furthermore, given $0<\ell \leq m_{0}$ , we say a continuous invariant splitting $T_{\Lambda }M=E_{1}\oplus \cdots \oplus E_{\ell }$ dominates if there are numbers $\chi _{1}<\chi _{2}<\cdots <\chi _{\ell }$ , constants $C>0$ and $0<\varepsilon <\min _{1\leq i\leq \ell -1}\{({\chi _{i+1}-\chi _{i}})/{100}\}$ such that for every $x\in \Lambda $ , $n\in \mathbb {N}$ and $1\leq j\leq \ell $ and each unit vector $u\in E_{j}(x)$ , it holds that

$$ \begin{align*} C^{-1}\exp[n(\chi_{j}-\varepsilon)]\leq \|D_{x}f^{n}(u)\|\leq C \exp[n(\chi_{j}+\varepsilon)]. \end{align*} $$

In particular, it is clear that $E_{1}\preceq \cdots \preceq E_{\ell }$ . We shall use the notion $\{\chi _{j}\}$ -dominated when we want to stress the dependence on the numbers $\{\chi _{j}\}$ .

Refining Katok’s approximation theory in non-uniformly hyperbolic dynamical systems [Reference Katok18], Avila, Crovisier and Wilkinson [Reference Avila, Crovisier and Wilkinson2] obtained the following approximation result.

Theorem 2.4. [Reference Avila, Crovisier and Wilkinson2]

Let f be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact smooth Riemannian manifold M, and let $\mu $ be an ergodic hyperbolic measure with $h_{\mu }(f)>0$ . Then for every $\varepsilon>0$ and weak $^{\ast }$ neighbourhood $\mathcal {V}$ of $\mu $ in the space of f-invariant probability measures on M, there exists an f-invariant compact subset $\Lambda _{\varepsilon }\subset M$ such that:

  1. (a) $\Lambda _{\varepsilon }$ is $\varepsilon $ -close to the support set of $\mu $ in the Hausdorff distance;

  2. (b) $|h_{\text {top}}(f|_{\Lambda _{\varepsilon }}) - h_{\mu }(f)| \leq \varepsilon $ ;

  3. (c) all the invariant probability measures supported on $\Lambda _{\varepsilon }$ lie in $\mathcal {V}$ ;

  4. (d) there is a $\{\chi _{j}(\mu )\}$ -dominated splitting $TM=E_{1}\oplus E_{2}\oplus \cdots \oplus E_{\ell }$ over $\Lambda _{\varepsilon }$ , where $\chi _{1}(\mu )<\cdots <\chi _{\ell }(\mu )$ are distinct Lyapunov exponents of f with respect to the measure $\mu $ .

In the second statement, the original result does not show that $h_{\text {top}}(f|_{\Lambda _{\varepsilon }})\le h_{\mu }(f)+\varepsilon $ . However, only a slight modification can give the upper bound of the topological entropy of f on the horseshoe.

3 Proofs

This section provides the detailed proofs of the main results presented in the previous section.

3.1 Proof of Theorem A

(i) Since $\mu $ is a hyperbolic ergodic SRB measure for a $C^{1+\alpha }$ diffeomorphism f and has a one-dimensional stable manifold, by [Reference Wang, Qu and Cao34, Lemma 15 and 25], one has

for $\mu $ -a.e. x. Barreira, Pesin and Schmeling [Reference Barreira, Pesin and Schmeling4] proved that $d_\mu (x)=d_\mu ^u(x)+d_\mu ^s(x)$ for $\mu $ -a.e. x. As a consequence, one has that

for $\mu $ -a.e. x. Hence, one has

If $\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _{m_0}(\mu )<0$ , then one can show that

since $\mu $ is an SRB measure and has a one-dimensional stable manifold. Therefore, we have that

$$ \begin{align*}\dim_H\mu=\dim_L\mu.\end{align*} $$

If $\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _{m_0}(\mu )\geq 0$ , it follows from the definition of Lyapunov dimension that $\dim _L\mu =m_0$ . Since $\mu $ is an SRB measure for f and has a one-dimensional stable manifold, one has that

This together with the fact that

implies that $({h_\mu (f)}/{-\unicode{x3bb} _{m_0}(\mu )})=1$ . This yields that $\dim _H\mu =\dim _L\mu $ .

By [Reference Wang, Qu and Cao34, Theorem B], there exists a sequence of hyperbolic sets $\Lambda _{n}$ such that

$$ \begin{align*} \dim_H \Lambda_{n}\rightarrow \dim_{L}\mu~( n\to\infty). \end{align*} $$

(ii) Since $\mu $ is a hyperbolic ergodic SRB measure for a $C^{1+\alpha }$ diffeomorphism f, by [Reference Wang, Qu and Cao34, Lemma 15], one has that

$$ \begin{align*} d_\mu^u(x)=u\quad \mu\text{-a.e. }x. \end{align*} $$

Considering $f^{-1}$ instead of f, since $h_\mu (f)=-\unicode{x3bb} _{u+1}(\mu )-\unicode{x3bb} _{u+2}(\mu )-\cdots -\unicode{x3bb} _{m_0}(\mu )$ , by [Reference Ledrappier and Young21, Theorem A], we have that the measure $\mu $ has absolutely continuous conditional measures on stable manifolds of f. Using the same arguments as the proof of [Reference Wang, Qu and Cao34, Lemma 15], we have that

$$ \begin{align*}d_\mu^s(x)=s\quad \mu\text{-a.e. }x.\end{align*} $$

Hence, $d_\mu (x)=u+s=m_0$ for $\mu $ -a.e. x, which implies that $\dim _H\mu =m_0$ . Since $\mu $ is an SRB measure for f and $h_\mu (f)=-\unicode{x3bb} _{u+1}(\mu )-\unicode{x3bb} _{u+2}(\mu )-\cdots -\unicode{x3bb} _{m_0}(\mu )$ , one can conclude that

then $\dim _L\mu =m_0$ by the definition of Lyapunov dimension. This proves that

$$ \begin{align*}\dim_L\mu=\dim_H\mu.\end{align*} $$

Finally, for each $\varepsilon>0$ , there exists a hyperbolic set $\Lambda _\varepsilon $ satisfying properties (a)–(d) in Theorem 2.4. Fix a positive integer $n\ge 1$ . Let $t_n$ be the unique root of Bowen’s equation $P(f^{2^n}|\Lambda _\varepsilon ,-\psi ^t(\cdot , f^{2^n}))=0$ and let $\mu _n^u$ be the unique equilibrium state for the topological pressure $P(f^{2^n}|\Lambda _\varepsilon ,-\psi ^{t_n}(\cdot , f^{2^n}))$ . Similarly, let $t_n^{\prime }$ be the root of Bowen’s equation $P(f^{2^n}|\Lambda _\varepsilon , \phi ^t(\cdot , f^{2^n}))=0$ and let $\mu _n^s$ be the unique equilibrium state for the topological pressure $P(f^{2^n}|\Lambda _\varepsilon , \phi ^{t_n}(\cdot , f^{2^n}))$ . As in the proof of [Reference Wang, Qu and Cao34, Theorem B], the following properties hold:

  1. (e) ${\lim _{\varepsilon \to 0}\lim _{n\to \infty }t_n=u}$ and ${\lim _{\varepsilon \to 0}\lim _{n\to \infty }t_n'=s}$ ;

  2. (f) there is a Markov partition $\mathcal {P}=\{P_1,P_2,\ldots ,P_\ell \}$ of $\Lambda _\varepsilon $ . For every $i\in \{1,2,\ldots ,\ell \}$ , there is a family of conditional measures $\{\mu _{n,x}^u\}_{x\in P_i}$ ( $\{\mu _{n,x}^s\}_{x\in P_i}$ ) of $\mu _n^u$ ( $\mu _n^s$ ) on the local unstable (stable) sets $W^u_{P_i}$ ( $W^s_{P_i}$ ) such that for every $x\in P_i$ , there is small $r_0>0$ such that for every $r\in (0, r_0)$ ,

    $$ \begin{align*}r^{u+\varepsilon} \leq \mu_{n,x}^u(B^u(x,r)) \leq r^{t_n-\varepsilon}\end{align*} $$
    and
    $$ \begin{align*}r^{s+\varepsilon} \leq \mu_{n,x}^s(B^s(x,r)) \leq r^{t_n'-\varepsilon},\end{align*} $$
    where $W_{P_i}^u(x,f):=W_{loc}^u(x,f)\cap P_i$ and $W_{P_i}^s(x,f):=W_{loc}^s(x,f)\cap P_i$ for every $x\in P_i$ .

Define a measure $\hat {\mu }_n$ on $P_i$ as follows:

$$ \begin{align*} \hat{\mu}_n(B(x,r)) = \mu_{n,x}^u(B^u(x,r)) \cdot \mu_{n,x}^s(B^s(x,r)) \end{align*} $$

for every $x\in P_i$ and each sufficiently small $r>0$ . This yields that

$$ \begin{align*} t_n+t_n'-2\varepsilon \leq \underline{d}_{\hat{\mu}_n}(x) \leq \overline{d}_{\hat{\mu}_n}(x) \leq m_0+2\varepsilon \end{align*} $$

for every $x\in P_i$ . By Proposition 2.1 and the fact that $\Lambda _\varepsilon =\bigcup _{i=1}^\ell P_i$ , we have that

$$ \begin{align*} \lim_{\varepsilon\to0}\dim_H\Lambda_\varepsilon = m_0 = \dim_L\mu. \end{align*} $$

This completes the proof of Theorem A.

3.2 Proof of Theorem B

Let $\Gamma $ be the set of points which are regular in the sense of Oseledets [Reference Oseledets26] with respect to the measure $\mu $ . For every $x\in \Gamma $ , denote its Lyapunov exponents by

To prove Theorem B, we need a coarse upper bound for the unstable and stable pointwise dimension $d_\mu ^u(x)$ , $d_\mu ^s(x)$ of an ergodic f-invariant hyperbolic probability measure $\mu $ for almost every x. We now provide the following useful lemma, which estimates the Hausdorff measure of the image of a small ball along unstable/stable direction under f.

Lemma 3.1. Fix $t\in [0,u]$ , then for any $b_0>2\sqrt {u}$ and $C_0>2^tu^{{t}/{2}}$ , there is $\rho _0>0$ such that for all $x\in \Gamma $ , if $B^u(x,\rho )\subset B(x,\rho _0)\cap W^u(x,f)$ for some $0<\rho <\rho _0$ , then we have

$$ \begin{align*}\mathcal{H}_{b\rho}^t(B^u(x,\rho)) \leq C \mathcal{H}_\rho^t(f(B^u(x,\rho))),\end{align*} $$

where $b=b_0\exp \{-\log \alpha _{u-[t]}(x,f)\} \text { and } C=C_0\exp \{-\psi ^t(x,f)\}$ .

Proof. For simplicity, we just prove the lemma on the assumption that M is the Euclid space $\mathbb {R}^{m_0}$ . For the general case, one can use local charts to prove it.

Given a small positive number $\varepsilon $ with ${e^\varepsilon }/({1-\varepsilon })<2$ , since $f: M\to M$ is a $C^{1+\alpha }$ diffeomorphism on M, there exists $\rho _0>0$ such that for every $y,z\in B(x,\rho _0)\cap W^u(x,f)$ , the following properties hold:

  1. (a) $\|y-z- (D_yf)^{-1}(f(y)-f(z)) \| \leq \varepsilon \|y-z\|$ ;

  2. (b) $|\log \alpha _i(y,f) - \log \alpha _i(z,f)|\leq \varepsilon $ for $i=1,2,\ldots ,u$ .

See [Reference Jordan and Pollicott17, Lemma 4] for the detailed proof of the above properties. Fix $0<\rho <\rho _0$ . Let $A:= B^u(x,\rho )$ and $a=\mathcal {H}^t_\rho (f(A))$ . Assume that a is finite, otherwise the conclusion is clear. For every $\eta>0$ , there are points $\{z_j\}\subset f(B(x,\rho _0)\cap W^u(x,f))$ such that

$$ \begin{align*} f(A) \subset \bigcup_{j}B^u(z_j, r_j) \end{align*} $$

with $r_j\leq \rho $ for each j and

$$ \begin{align*} \sum_{j}r_j^t<a+\eta. \end{align*} $$

Let $B_j'=\{y\in A: f(y)\in B^u(z_j,r_j)\}$ , then $A\subset \bigcup _j B_j'$ . By property (a), we conclude that $B_j'$ is contained in an ellipse with principal axes

$$ \begin{align*} \frac{1}{1-\varepsilon} r_j \cdot \alpha_1(y_j, f)^{-1}, \frac{1}{1-\varepsilon} r_j \cdot \alpha_2(y_j, f)^{-1}, \ldots, \frac{1}{1-\varepsilon} r_j \cdot \alpha_u(y_j, f)^{-1}, \end{align*} $$

where $y_j\in B^u(x,\rho )$ and $f(y_j)=z_j$ . This together with property (b) yield that $B_j'$ is contained in an ellipse with principal axes

$$ \begin{align*}\frac{e^\varepsilon}{1-\varepsilon} r_j \cdot \alpha_1(x, f)^{-1}, \frac{e^\varepsilon}{1-\varepsilon} r_j \cdot \alpha_2(x, f)^{-1}, \ldots, \frac{e^\varepsilon}{1-\varepsilon} r_j \cdot \alpha_u(x, f)^{-1}. \end{align*} $$

Hence, $B_j'$ is covered by

$$ \begin{align*}\frac{\exp\{-\sum_{j=u-[t]+1}^u \log\alpha_j(x,f) \} }{ \exp\{-[t]\log\alpha_{u-[t]}(x,f) \}}\end{align*} $$

balls with radius $({e^\varepsilon }/({1-\varepsilon })) \sqrt {u} r_j\cdot \exp \{-\log \alpha _{u-[t]}(x,f)\}$ . In fact, the radius

$$ \begin{align*} &\hspace{-6pt}\frac{e^\varepsilon}{1-\varepsilon} \sqrt{u} r_j \cdot \exp\{-\log\alpha_{u-[t]}(x,f)\}\\ &\leq \ 2\sqrt{u}\exp\{-\log\alpha_{u-[t]}(x,f)\} \cdot \rho\\ &\leq \ b\rho. \end{align*} $$

Therefore,

$$ \begin{align*} \mathcal{H}^t_{b\rho}(B_j')&\leq \exp\bigg\{-\sum_{j=u-[t]+1}^u \log\alpha_j(x,f) + [t]\log\alpha_{u-[t]}(x,f) \bigg\}\\ &\quad \cdot \bigg(\frac{e^\varepsilon}{1-\varepsilon} \sqrt{u}\bigg)^t r_j^t \cdot \exp\{-t\log\alpha_{u-[t]}(x,f)\}\\ & \leq (2\sqrt{u})^t \cdot \exp\{-\psi^t(x,f)\} \cdot r_j^t. \end{align*} $$

Summing up over all j, we have that

$$ \begin{align*} \mathcal{H}^t_{b\rho}(A)&\leq \sum_{j}\mathcal{H}_{b\rho}^t(B_j')\\ &\leq 2^t (\sqrt{u})^t \exp\{-\psi^t(x,f)\} \cdot \sum_j r_j^t\\ &\leq 2^t (\sqrt{u})^t \exp\{-\psi^t(x,f)\} \cdot (a+\eta). \end{align*} $$

The choice of $C_0$ and the arbitrariness of $\eta>0$ implies the desired result.

The following result relates the zero of measure-theoretic pressure with the upper bound of the unstable pointwise dimension of $\mu $ .

Lemma 3.2. For $\mu $ -a.e. x, $d_\mu ^u(x)\leq t_{u,1}^*$ , where $t_{u,1}^*$ is the unique solution of the equation $P_{\mu }(f, -\psi ^t(\cdot ,f))=0$ .

Proof. Fix a small number $\varepsilon>0$ such that $-\unicode{x3bb} _u(\mu )+2\varepsilon <0$ and choose $t>t_{u,1}^*$ such that

$$ \begin{align*}h_\mu(f)-\int\psi^t(x,f)\,d\mu = -3\varepsilon. \end{align*} $$

Claim. There exists an integer $N_1$ (depending only on $\varepsilon $ ) such that, for $\mu $ -a.e. x and every $N\ge N_1$ , the Birkhoff averages

$$ \begin{align*} \frac{1}{kN}\sum_{j=0}^{k-1}\log \alpha_{u}(f^{jN}x, f^N) \end{align*} $$

converge towards a number bigger than $\unicode{x3bb} _u(\mu )-\varepsilon $ , as k goes to $+\infty $ .

Proof of the Claim

We give the proof of the Claim by modifying slightly the arguments in the proof of [Reference Abdenur, Bonatti and Crovisier1, Lemma 8.4].

Since ${\lim _{n\to \infty } ({1}/{n}) \log \alpha _u(x,f^n) = \lim _{n\to \infty } (1/n) \int \log \alpha _u(x,f^n)\,d\mu = \unicode{x3bb} _u(\mu )}$ for $\mu $ -a.e. x, there exists a positive integer L such that

(11)

The measure $\mu $ may be not ergodic for $f^L$ , one can decompose it as

$$ \begin{align*}\mu=\frac{1}{m}(\mu_1+\mu_2+\cdots+\mu_m),\end{align*} $$

where $m\in \mathbb {N}^+$ divides L and each $\mu _i$ is an ergodic $f^L$ -invariant measure such that $f_*\mu _i=\mu _{i+1}$ for each $i(\text {mod}\,\, m)$ . Let $A_1\cup A_2\cup \cdots \cup A_m$ be a measurable partition of $(M,\mu )$ such that $f(A_i)=A_{i+1}$ for each $i(\text {mod}\,\, m)$ and $\mu _i(A_i)=1$ . By equation (11), there exists $j_0\in \{1,2,\ldots , m\}$ such that

For every $N\ge 1$ and $\mu $ -a.e. x, one decomposes the orbit $\{f^i(x)\}_{i=0}^{N-1}$ as $(x,\ldots, f^{j-1}(x))$ , $(f^j(x),\ldots ,f^{j+(r-1)L-1}(x))$ and $(f^{j+(r-1)L}(x),\ldots , f^{N-1}(x))$ , where $j{\kern-1.2pt}<{\kern-1.2pt}L$ , $j+rL{\kern-1.2pt}\ge{\kern-1.2pt} N$ and the points $\{f^{j+sL}(x)\}_{s=0}^{r}$ belong to $A_{j_0}$ . Using the super-additivity of $\{\log \alpha _u(x,f^n) \}_{n\ge 1}$ , we have that

$$ \begin{align*} \log \alpha_u(x,f^N)&\ge \log \alpha_u(x,f^j) + \sum_{s=0}^{r-2} \log \alpha_u(f^{j+sL}x,f^L)\\ &\quad +\log \alpha_u(f^{j+(r-1)L}x,f^{N-j-L(r-1)}). \end{align*} $$

Hence, one has

$$ \begin{align*} \log \alpha_u(x,f^N)\ge 2C_f + \sum_{s=0}^{r-2} \log \alpha_u(f^{j+sL}x,f^L), \end{align*} $$

where ${C_f=\max _{0\le i<L}\max _{x\in M}|{\log}\, \alpha _u(x,f^i)|}$ with the convention that $|{\log}\, \alpha _u (x,f^0)|=0$ . Since

and

$$ \begin{align*} \lim_{k\to +\infty}\frac{1}{kN}\sum_{j=0}^{k-1}\log \alpha_{u}(f^{jN}x, f^N)\ge \frac{2C_f}{N}+\lim_{k\to +\infty} \frac{1}{kL}\sum_{\ell=0}^{k-1}\log \alpha_u(f^{j+\ell L}x,f^L), \end{align*} $$

and there exists an integer $N_1$ (depending on $\varepsilon $ ) so that $|{2C_f}/{N}|< {\varepsilon }/{2}$ for every $N>N_1$ , for $\mu $ -a.e. x and every $N>N_1$ , we have that

Take $b_0>2\sqrt {u}$ and $C_0>2^tu^{{t}/{2}}$ , choose $N>N_1$ large enough such that

(12)

By the above Claim and Birkhoff ergodic theorem, for $\mu $ -a.e. $x\in M$ , we have that

and

$$ \begin{align*} \lim_{n\to\infty}\frac{1}{nN}\sum_{j=0}^{nN-1} \psi^t(f^jx,f)=\int\psi^t(x,f)\,d\mu. \end{align*} $$

Let $\rho _0$ be as in Lemma 3.1. Fix $\delta \in (0,\rho _0)$ . Ledrappier and Young [Reference Ledrappier and Young22] proved that

$$ \begin{align*} \limsup_{n\to\infty} \frac{-\log\mu_x^u(B^u(x,n,\delta/2))}{n} &\leq \lim_{\delta\to0}\limsup_{n\to\infty} \frac{-\log\mu_x^u(B^u(x,n,{\delta}/{2}))}{n}\\ &=h_\mu(f)\, \mu\text{-a.e.}\, x, \end{align*} $$

where $B^u(x,n, \delta /2):=\{y\in W^u(x,f): d^u(f^jx,f^jy)< \delta /2 \text { for } 0\leq j<n\}$ . Hence, one can find sets $A_n\subset \Gamma $ with $\mu (A_n)\to 1\, (n\to \infty )$ , for every $x\in A_n$ where the following properties hold:

  1. (a) ${\exp [-nN(h_\mu (f)+\varepsilon )] \leq \mu _x^u(B^u(x,nN, {\delta }/{2}))}$ ;

  2. (b) ${nN(-\int \psi ^t(x,f)-\varepsilon ) \leq -\sum _{j=0}^{nN-1}\psi ^t(f^{j}x,f) \leq nN(-\int \psi ^t(x,f)+\varepsilon )}$ ;

  3. (c) ${\sum _{j=0}^{n-1}\log \alpha _{u-[t]}(f^{jN}x, f^N) \geq nN(\unicode{x3bb} _u(\mu )-2\varepsilon )}$ .

Take a point $x\in A_n$ . Let E be a maximal $(nN, \delta )$ -separated subset of $A_n\cap \xi ^u(x)$ , then

$$ \begin{align*} A_n\cap\xi^u(x) \subset \bigcup_{x_j\in E} B^u(x_j, nN, \delta). \end{align*} $$

Furthermore, by property (a), the number of balls $B^u(x_j,nN, \delta /2)$ is less than or equal to $\exp \{nN[h_\mu (f)+\varepsilon ]\}$ . Let

$$ \begin{align*}b_k(x)=(b_0)^k \exp \bigg[-\sum_{j=n-k}^{n-1}\log\alpha_{u-[t]}(f^{jN}x, f^N)\bigg]\end{align*} $$

for $k=1,2,\ldots , n$ and $\beta _n=\{b_0\exp [(-\unicode{x3bb} _u(\mu )+2\varepsilon )N]\}^n \cdot \rho $ , where $0<\rho <\rho _0$ . By property (c), we have

For each $x_j\in E$ , using Lemma 3.1 n times, we conclude that

$$ \begin{align*} \mathcal{H}^t_{\beta_n}(B^u(x_j, nN, \delta)) &\leq \mathcal{H}^t_{b_n(x_j)\rho}(B^u(x_j,nN,\delta))\\ &\leq C_0 \exp\{-\psi^t(x_j, f^N)\} \cdot \mathcal{H}^t_{b_{n-1}(x_j)\rho} (f^N(B^u(x_j,nN, \delta)))\\ &\leq C_0 \exp\{-\psi^t(x_j, f^N)\} \cdot \mathcal{H}^t_{b_{n-1}(x_j)\rho} (B^u(f^N(x_j),(n-1)N, \delta))\\ &\leq (C_0)^2 \exp\{-\psi^t(x_j, f^N)\} \cdot \exp\{-\psi^t(f^N(x_j), f^N)\} \\ &\quad\cdot \mathcal{H}^t_{b_{n-2}(x_j)\rho} (B^u(f^{2N}(x_j),(n-2)N, \delta))\\ &\leq \cdots\\&\leq (C_0)^n \exp\bigg\{ -\sum_{j=0}^{n-1}\psi^t(f^{jN}x_j, f^N) \bigg\} \cdot \mathcal{H}^t_\rho(B^u(f^{nN}x_j, \delta))\\ &\leq (C_0)^n C_1 \cdot \exp\bigg\{ -\sum_{j=0}^{n-1}\psi^t(f^{jN}x_j, f^N) \bigg\}, \end{align*} $$

where $C_1=\sup _{y\in M}\mathcal {H}^t_\rho (B(y,\delta ))$ . By property (b) and the sub-additivity of $\{-\psi ^t(\cdot ,f^n)\}_{n\geq 1}$ , we have that

$$ \begin{align*} \mathcal{H}_{\beta_n}^t(A_n\cap\xi^u(x)) &\leq \sum_{x_j\in E}\mathcal{H}^t_{\beta_n}(B^u(x_j, nN, \delta))\\ &\leq \sum_{x_j\in E} (C_0)^n C_1 \cdot \exp\bigg\{ -\sum_{i=0}^{n-1}\psi^t(f^{jN}x_j, f^N) \bigg\}\\ &\leq \sum_{x_j\in E} (C_0)^n C_1 \cdot \exp\bigg\{ -\sum_{i=0}^{nN-1}\psi^t(f^{i}x_j, f) \bigg\}\\ &\leq (C_0)^n C_1 \cdot \exp[nN(h_\mu(f)+\varepsilon)] \cdot \exp\bigg[nN\bigg({-}\int\psi^t(x,f)\,d\mu + \varepsilon\bigg)\kern-1pt\bigg] \\ &= (C_0)^n C_1 \cdot \exp\bigg[nN\bigg(h_\mu(f)- \int\psi^t(x,f)\,d\mu +2\varepsilon\bigg) \bigg]\\ &= (C_0)^n C_1\cdot e^{-nN\varepsilon}\\ &=(C_0 e^{-N\varepsilon})^n C_1. \end{align*} $$

Since N satisfies $C_0e^{-N\varepsilon }<1$ , we have that

$$ \begin{align*} \lim_{n\to\infty} \mathcal{H}^t_{\beta_n}(A_n\cap\xi^u(x))=0. \end{align*} $$

Since ${\lim _{n\to \infty }\beta _n=0}$ and ${\lim _{n\to \infty }\mu _x^u(A_n\cap \xi ^u(x))= 1}$ for $\mu $ -a.e. x, by [Reference Jordan and Pollicott17, Lemma 6], we obtain that

$$ \begin{align*}\dim_H\mu_x^u\leq t\end{align*} $$

for $\mu $ -a.e. x. Combining with equation (5) and the choice of t yield that $d_\mu ^u(x)\leq t_{u,1}^*$ for $\mu $ -a.e. x.

Now we are ready to present the proof of Theorem B.

Proof of Theorem B

For each $n>1$ , the measure $\mu $ is f-invariant ergodic, but it may be not ergodic for $f^n$ although $\mu $ is still $f^n$ -invariant. In either case, one can find an $f^n$ -invariant ergodic probability measure $\nu $ such that

$$ \begin{align*}\mu=\frac1m[\nu + f_*\nu + \cdots + f^{m-1}_*\nu], \end{align*} $$

where $m\in \mathbb {N}\setminus \{0\}$ divides n. Let

$$ \begin{align*} \widetilde{P}_\mu(f^n, -\psi^n(\cdot, f^n)):=h_\mu(f^n)-\int \psi^n(x, f^n)\,d\mu, \end{align*} $$

then one can show that

$$ \begin{align*} \widetilde{P}_\mu(f^n, -\psi^n(\cdot, f^n))&= \frac1m \sum_{i=0}^{m-1}\bigg( h_{f_*^i\nu}(f^n)-\int \psi^n(x, f^n) \,df_*^i\nu \bigg)\\ &=\frac1m \sum_{i=0}^{m-1} P_{f_*^i\nu}(f^n, -\psi^n(\cdot, f^n)). \end{align*} $$

Hence, there exists $j_0\in \{0,1,\ldots , m-1\}$ such that

$$ \begin{align*} \widetilde{P}_\mu(f^n, -\psi^t(\cdot, f^n)) \geq P_{f_*^{j_0}\nu}(f^n, -\psi^t(\cdot, f^n)). \end{align*} $$

Since $f_*^{j_0}\nu $ is hyperbolic and $f^n$ -invariant ergodic, by Lemma 3.2, there is a set $\tilde {A}$ with $\nu \circ f^{-j_0}(\tilde {A})=1$ such that for each $x\in \tilde {A}$ ,

$$ \begin{align*}d_{f_*^{j_0}\nu}^u (x) \leq t_{u,n}^*, \end{align*} $$

where $t_{u,n}^*$ is the unique root of the equation $P_\mu (f^n,-\psi ^t(\cdot , f^n))=0$ . Note that $d_\mu ^u(x), d_{f_*^{j_0}\nu }^u(x)$ are constants almost everywhere (see [Reference Ledrappier and Young22]) and $d_\mu ^u(x)\leq d_{f_*^{j_0}\nu }^u(x) \leq t_{u,n}^*$ for each $x\in \tilde {A}$ with $\mu (\tilde {A})\geq 1/m$ . Consequently, we have that

$$ \begin{align*}d_\mu^u(x)\leq t_{u,n}^*\end{align*} $$

for $\mu $ -a.e. x.

By the sub-additive of $\{-\psi ^t(\cdot , f^n)\}_{n\geq 1}$ , we obtain

$$ \begin{align*}\frac{1}{2^{k+1}} \bigg[h_\mu(f^{2^{k+1}}) - \int \psi^t(x,f^{2^{k+1}})\, d\mu\bigg] \leq \frac{1}{2^k} \bigg[ h_\mu(f^{2^k}) -\int \psi^t(x,f^{2^k})\, d\mu \bigg]. \end{align*} $$

Hence,

$$ \begin{align*}\frac{1}{2^{k+1}} \widetilde{P}_\mu(f^{2^{k+1}}, -\psi^t(\cdot, f^{2^{k+1}})) \leq \frac{1}{2^k} \widetilde{P}_\mu(f^{2^{k}}, -\psi^t(\cdot, f^{2^{k}})). \end{align*} $$

This yields that $t_{u,2^{k+1}}^*\leq t_{u,2^k}^*$ for every $k\ge 1$ . Let ${t_u^*:=\lim _{k\to \infty } t_{u, 2^k}^*}$ , then one has that

$$ \begin{align*}d_\mu^u(x)\leq t_u^*\quad \mu\text{-a.e.}\, x. \end{align*} $$

Since $P_\mu (f, \{-\psi ^t(\cdot , f^n)\})$ is continuous and strictly decreasing with respect to t, there exists at most one solution of the equation. To complete the proof of Theorem B, it suffices to show that $P_\mu (f, \{-\psi ^{t^*}(\cdot , f^n)\})=0$ .

Since $t_{u,2^k}^*\ge t_u^{*}$ for every $k\ge 1$ , by Theorem 2.2, one has that

$$ \begin{align*} 0\le \lim_{k\to\infty}\frac{1}{2^k} \widetilde{P}_\mu(f^{2^k}, -\psi^{t_u^{*}}(\cdot, f^{2^k}))= P_\mu(f, \{-\psi^{t_u^{*}}(\cdot, f^n)\}). \end{align*} $$

However, for each small number $\varepsilon>0$ , there exists K so that $t_{u,2^k}^*\le t_u^{*}+\varepsilon $ for every ${k\ge K}$ . Hence, we have that

$$ \begin{align*} P_\mu(f, \{-\psi^{t_u^*+\varepsilon}(\cdot, f^n)\}) &= h_\mu(f) - \lim_{n\to\infty} \frac1n\int\psi^{t_u^*+\varepsilon}(x,f^n)\, d\mu\\ &= \lim_{k\to\infty} \frac{1}{2^k}\widetilde{P}_\mu(f^{2^k}, -\psi^{t_u^*+\varepsilon}(\cdot, f^{2^k}))\le 0. \end{align*} $$

The previous arguments imply that $P_\mu (f, \{-\psi ^{t^*}(\cdot , f^n)\})=0$ . One can prove in a similar fashion that $d_\mu ^s(x)\leq t_s^*$ for $\mu $ -a.e. x. This completes the proof of Theorem B.

3.3 Proof of Theorem C

For each $\varepsilon>0$ , there exists a hyperbolic set $\Lambda _\varepsilon $ satisfying properties (a)–(d) in Theorem 2.4. The following lemma shows that the zero of the super-additive topological pressure of $\Phi _f(t)$ provides a lower bound of the Carathéodory singular dimension of the hyperbolic set on the local unstable leaf with respect to the super-additive singular valued potential $\Phi _f(t)$ .

Lemma 3.3. Let $f: M\to M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact smooth Riemannian manifold M and let $\Lambda \subset M$ be a hyperbolic set. Assume that $f|_{\Lambda }$ is topologically transitive, then for every $x\in \Lambda $ ,

$$ \begin{align*}\dim_C^{\Phi_f} (\Lambda\cap W^u_{\mathrm{loc}}(x,f)) \geq t_*, \end{align*} $$

where $t_*$ is the unique root of the equation $P(f|_\Lambda , \Phi _f(t))=0$ .

Proof. For every $x\in \Lambda $ , we denote $Z=\Lambda \cap W^u_{\mathrm {loc}}(x, f)$ and $P(t)=P(f|_\Lambda , \Phi _f(t))$ . Since the function $P(t)$ is strictly decreasing in t, then for each $t<t_*$ , we have that $P(t)>0$ . Fix such a number t and take $\varepsilon>0$ with $P(t)-\varepsilon>0$ . By Proposition 2.2, one has that

$$ \begin{align*} P(t)=\lim_{n\to\infty}\frac1nP(f^n|_\Lambda, -\phi^t(\cdot, f^n)), \end{align*} $$

then there exists $N_1\in \mathbb {N}$ such that for every $n\geq N_1$ , we obtain

$$ \begin{align*} P(f^n|_\Lambda, -\phi^t(\cdot, f^n))> n( P(t) - \varepsilon ) >0. \end{align*} $$

Fix an integer $L\geq N_1$ , by [Reference Climenhaga, Pesin and Zelerowicz11, Proposition 5.4], one has that

$$ \begin{align*} P(f^L|_\Lambda, -\phi^t(\cdot, f^L)) = P_{Z}(f^L|_\Lambda, -\phi^t(\cdot, f^L)). \end{align*} $$

Hence, there is $\delta _1>0$ such that

$$ \begin{align*}P_{Z} (f^L|_\Lambda, -\phi^t(\cdot, f^L), \delta)> (P(t)-\varepsilon)L \end{align*} $$

for every $0<\delta <\delta _1$ . Consequently, fixing such a $\delta>0$ , one has that

$$ \begin{align*}m(Z, -\phi^t(\cdot, f^L), (P(t)-\varepsilon)L, \delta)=+\infty. \end{align*} $$

Hence, for each $K>0$ , there exists $S\in \mathbb {N}$ such that for each $N\ge S$ , we have that

(13) $$ \begin{align} \nonumber K & \leq \inf \sum_i\exp [-(P(t)-\varepsilon)Lm_i -S_{m_i}\phi^t(x_i, f^L)] \\ &\le e^{-NL(P(t)-\varepsilon)} \inf \sum_i\exp[ -S_{m_i}\phi^t(x_i, f^L)], \end{align} $$

where the infimum is taken over all collections $\{B^u_{m_i}(x_i, \delta , f^L)\}$ with $x_i\in \Lambda $ , $m_i\geq N$ which cover Z, $-S_{m_i}\phi ^t(x_i, f^L)\kern1.2pt{=}\kern1.2pt{-}\phi ^t(x_i, f^L)\kern1.2pt{-}\kern1.2pt\phi ^t(f^Lx_i,f^L)-\cdots -\phi ^t(f^{(m_i-1)L}x_i, f^L)$ and

$$ \begin{align*}\displaystyle{B^u_{m_i}(x_i, \delta, f^L):= \left\{y\in W^u(x_i,f): \max_{0\le j <m_i}d^u(f^{jL}(y), f^{jL}(x_i))<\delta \right\}}.\end{align*} $$

Fixing such an N and taking an integer $R\ge NL$ , let the collection of balls $\{B_{n_i}^u(x_i,\delta )\}$ with $x_i\in \Lambda $ , $n_i\ge R$ be a cover of Z. One can write $n_i=m_iL+s_i$ with $0\le s_i<L$ and $m_i\geq N$ for each i. Since $B_{n_i}^u(x_i,\delta )\subset B_{m_i}^u(x_i,\delta , f^L)$ for each i, the collection of balls $B_{m_i}^u(x_i,\delta , f^L)$ is also a cover of Z with $x_i\in \Lambda $ , $m_i\ge N$ . By the super-additivity of $\{-\phi ^t(\cdot , f^n)\}_{n\ge 1}$ , one has

$$ \begin{align*} \sum_{i}\exp [-\phi^t(x_i,f^{n_i})]&\ge \sum_{i}\exp [-S_{m_i}\phi^t(x_i,f^{L})-\phi^t(f^{m_iL}y, f^{s_i})]\\ &\ge C \sum_{i}\exp [-S_{m_i}\phi^t(x_i,f^{L})], \end{align*} $$

where ${C=\min _{0\le s<L}\min _{x\in M}\exp [-\phi ^t(x, f^{s})]}$ . This together with equation (13) yield that

$$ \begin{align*} \sum_{i}\exp [-\phi^t(x_i,f^{n_i})]\ge CK e^{NL(P(t)-\varepsilon)}. \end{align*} $$

Since the cover of Z is taken arbitrarily, one can conclude that

$$ \begin{align*} \inf\sum_{i}\exp [-\phi^t(x_i,f^{n_i})] \geq CK e^{NL(P(t)-\varepsilon)}, \end{align*} $$

where the infimum is taken over all collections $\{B^u_{n_i}(x_i, \delta )\}$ with $x_i\in \Lambda $ , $n_i\geq NL$ which cover Z. Letting $N\to \infty $ , we obtain

$$ \begin{align*}m(Z, \Phi_f(t), \delta)=+\infty\end{align*} $$

for every $t<t_*$ . This implies that

$$ \begin{align*}\dim_{C}^{\Phi_f}Z\geq t_*.\\[-37pt] \end{align*} $$

Proof of Theorem C(i)

By Lemma 3.3, for every $x\in \Lambda _\varepsilon $ , we obtain

$$ \begin{align*}\dim_C^{\Phi_f}(\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f)) \geq t_{\varepsilon*}, \end{align*} $$

where $t_{\varepsilon *}$ is the unique root of the equation $P(f|_{\Lambda _\varepsilon }, \Phi _f(t))=0$ . By the variational principle of topological entropy, take $\nu \in \mathcal {M}_{\mathrm {inv}}(f|_{\Lambda _\varepsilon })$ such that $h_{\text {top}}(f|_{\Lambda _\varepsilon })=h_\nu (f|_{\Lambda _\varepsilon })$ . By properties (b) and (d) in Theorem 2.4, it holds that

(14)

where $\unicode{x3bb} _1(\nu )\geq \unicode{x3bb} _2(\nu ) \geq \cdots \geq \unicode{x3bb} _{m_0}(\nu )$ are the Lyapunov exponents of $\nu $ . However, let $\tau \in \mathcal {M}_{\mathrm {inv}}(f|_{\Lambda _\varepsilon })$ be an equilibrium state of $P(f|_{\Lambda _\varepsilon }, \Phi _f(t_{\varepsilon *}))$ , then one has that

This together with equation (14) yield that

$$ \begin{align*} -(u+1)\varepsilon\le P_{\mu}(f, \Phi_f(t_{\varepsilon*}))\le (u+1)\varepsilon. \end{align*} $$

Hence, we have that

$$ \begin{align*} \lim_{\varepsilon\to 0}P_{\mu}(f, \Phi_f(t_{\varepsilon*}))=0. \end{align*} $$

This implies that ${\lim _{\varepsilon \to 0}t_{\varepsilon *}=t_{u*}}$ , where $t_{u*}$ is the unique root of $P_\mu (f, \Phi _f(t))=0$ . Consequently, we have that

(15) $$ \begin{align} \liminf_{\varepsilon\to0}\dim_C^{\Phi_f}(\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f)) \geq t_{u*}. \end{align} $$

As a counterpart of Lemma 3.3, we have the following result.

Lemma 3.4. Let $f: M\to M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact smooth Riemannian manifold M and let $\Lambda \subset M$ be a hyperbolic set. Assume that $f|_{\Lambda }$ is topologically transitive. Then for every $x\in \Lambda $ ,

$$ \begin{align*}\dim_C^{\Psi_f} (\Lambda\cap W^u_{\mathrm{loc}}(x,f)) \leq t^*, \end{align*} $$

where $t^*$ is the unique root of the equation $P(f|_\Lambda , \Psi _f(t))=0$ .

Proof. Denote $P(t)=P(f|_\Lambda , \Psi _f(t))$ . For each $t>t_*$ ,

$$ \begin{align*}0>P(t)=\lim_{n\to\infty}\frac1nP(f^n, -\psi^t(\cdot, f^n)). \end{align*} $$

Fix such a number t and take $\varepsilon>0$ with $P(t)+\varepsilon <0$ . Then there exists $N_1\in \mathbb {N}$ such that for every $n\geq N_1$ , we obtain

$$ \begin{align*}P(f^n, -\psi^t(\cdot, f^n)) < n( P(t) + \varepsilon ) <0. \end{align*} $$

Fix an integer $L\geq N_1$ such that

$$ \begin{align*}P(f^L, -\psi^t(\cdot, f^L)) < L ( P(t) + \varepsilon ) <0. \end{align*} $$

For each $x\in \Lambda $ , set $Z=\Lambda \cap W^u_{\mathrm {loc}}(x,f)$ . By [Reference Climenhaga, Pesin and Zelerowicz11, Proposition 5.4], one has that

$$ \begin{align*} P(f^L, -\psi^t(\cdot, f^L)) = P_{Z}(f^L, -\psi^t(\cdot, f^L)). \end{align*} $$

Thus, there is $\delta _1>0$ such that for every $0<\delta <\delta _1$ , one has

$$ \begin{align*}P_{Z} (f^L, -\psi^t(\cdot, f^L), \delta) < (P(t)+\varepsilon)L. \end{align*} $$

Hence, one has that

$$ \begin{align*} m(Z, -\psi^t(\cdot, f^L), (P(t)+\varepsilon)L, \delta)=0. \end{align*} $$

For each $\xi>0$ , there exists $N\in \mathbb {N}$ and a cover $\{B^u_{n_i}(x_i, \delta , f^L)\}$ of Z with $x_i\in \Lambda $ , $n_i\geq N$ such that

$$ \begin{align*} \xi &\geq \sum_i\exp\Big[-(P(t)+\varepsilon)Ln_i + \sup_{y\in B^u_{n_i}(x_i, \delta, f^L)}-S_{n_i}\psi^t(y, f^L)\Big]. \\ &\geq e^{-NL(P(t)+\varepsilon)} \sum_i \exp \Big[\sup_{y\in B^u_{n_i}(x_i, \delta, f^L)}-S_{n_i}\psi^t(y, f^L)\Big]. \end{align*} $$

Note that $d^u(f^Lx, f^Ly)<\delta $ implies $d^u(f^ix, f^iy)<\delta $ for $i=0, 1,\ldots , L-1$ , since f is expanding along the unstable manifold. This implies that $ B^u_{(n_i-1)L+1}(x_i,\delta ) = B^u_{n_i}(x_i, \delta , f^L) $ for every i. Since

$$ \begin{align*} -S_{n_i}\psi^t(y, f^L)&=-\psi^t(y, f^L)-\psi^t(f^Ly,f^L)-\cdots -\psi^t(f^{(n_i-1)L}y, f^L)\\ &\geq -\psi^t(y,f^{(n_i-1)L}) +C_1\\ &= -\psi^t(y,f^{(n_i-1)L}) - \psi^t(f^{(n_i-1)L}y,f)+\psi^t(f^{(n_i-1)L}y,f)+C_1\\ &\geq -\psi^t(y,f^{(n_i-1)L+1})+ C_1+C_2, \end{align*} $$

where $C_1=\min _{x\in M}\{-\psi ^t(x,f^L)\}$ and $C_2=\min _{x\in M}\psi ^t(x,f)$ , we have that

$$ \begin{align*} \xi &\geq e^{-NL(P(t)+\varepsilon)}e^{C_1+C_2} \sum_i \exp\Big[\sup_{y\in B^u_{(n_i-1)L+1}(x_i,\delta) }-\psi^t(y,f^{(n_i-1)L+1}) \Big] \\ &\geq e^{-NL(P(t)+\varepsilon)}e^{C_1+C_2} \inf \sum_i\exp \Big[\sup_{y\in B^u_{m_i}(x_i,\delta)} -\psi^t(y, f^{m_i})\Big] \end{align*} $$

and

$$ \begin{align*} \inf \sum_i\exp \Big[\sup_{y\in B^u_{m_i}(x_i,\delta)} -\psi^t(y, f^{m_i})\Big]\leq \xi e^{NL(P(t)+\varepsilon)}e^{-C_1-C_2}, \end{align*} $$

where the infimum is taken over all collections $\{B^u_{m_i}(x_i, \delta )\}$ with $x_i\in \Lambda $ , $m_i\geq (N-1)L$ which cover Z. Letting $N\to \infty $ , we obtain

$$ \begin{align*}m(Z, \Psi_f(t), \delta)=0 \end{align*} $$

for every $t>t_*$ . This yields that

$$ \begin{align*} \dim_{C}^{\Psi_f}Z\leq t_*.\\[-42pt] \end{align*} $$

Remark 3.1. Let $f: M\to M$ be a $C^{1+\alpha }$ diffeomorphism on an $m_0$ -dimensional compact Riemannian manifold M and $\Lambda \subset M$ be a hyperbolic set. Assume that $f|_{\Lambda }$ is topologically transitive. Then for every $x\in \Lambda $ ,

$$ \begin{align*}\dim_C^{\Phi_f} (\Lambda\cap W^u_{\mathrm{loc}}(x,f)) = t^{\Phi_f}_u,\quad \dim_C^{\Psi_f} (\Lambda\cap W^u_{\mathrm{loc}}(x,f)) = t^{\Psi_f}_u, \end{align*} $$

where $t^{\Phi _f}_u$ , $t^{\Psi _f}_u$ are the unique roots of the equations

$$ \begin{align*}P_{\Lambda\cap W^u(x,f)}(f, \Phi_f(t))=0, \quad P_{\Lambda\cap W^u(x,f)}(f, \Psi_f(t))=0,\end{align*} $$

respectively. The proof is a slight modification of Lemmas 3.3 and 3.4. See [Reference Cao, Wang and Zhao9] for more details about the Carathéodory singular dimension of each subset of a repeller. However, we do not know whether $P_{\Lambda \cap W^u(x,f)}(f, \Phi _f(t))=P_\Lambda (f, \Phi _f(t))$ and $P_{\Lambda \cap W^u(x,f)}(f, \Psi _f(t))=P_\Lambda (f, \Psi _f(t))$ hold.

Proof of Theorem C(ii)

By Lemma 3.4, we have that

$$ \begin{align*}\dim_C^{\Psi_f}(\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f)) \leq t_{\varepsilon}^*, \end{align*} $$

where $t_{\varepsilon }^*$ is the unique root of the equation $P(f|_{\Lambda _\varepsilon }, \Psi _f(t))=0$ . Take $\nu \in \mathcal {M}_{\mathrm {inv}}(f|_{\Lambda _\varepsilon })$ such that $h_\nu (f|_{\Lambda _\varepsilon })=h_{\text {top}}(f|_{\Lambda _\varepsilon })$ , by properties (b) and (d) in Theorem 2.4, it holds that

(16)

where $\unicode{x3bb} _1(\nu )\geq \unicode{x3bb} _2(\nu ) \geq \cdots \geq \unicode{x3bb} _{m_0}(\nu )$ are the Lyapunov exponents of $\nu $ . Similarly, let ${\tau \in \mathcal {M}_{\mathrm {inv}}(f|_{\Lambda _\varepsilon })}$ be an equilibrium state of $P(f|_{\Lambda _\varepsilon }, \Psi _f(t_{\varepsilon }^*))$ , then one has that

This together with equation (16) yield that

$$ \begin{align*} -(u+1)\varepsilon\le P_{\mu}(f, \Psi_f(t_{\varepsilon}^*))\le (u+1)\varepsilon. \end{align*} $$

Hence, we have that

$$ \begin{align*} \lim_{\varepsilon\to 0}P_{\mu}(f, \Psi_f(t_{\varepsilon}^*))=0. \end{align*} $$

This implies that ${\lim _{\varepsilon \to 0}t_{\varepsilon }^*=t_{u}^*}$ , where $t_{u}^*$ is the unique root of $P_\mu (f, \Psi _f(t))=0$ . Consequently, we have that

$$ \begin{align*}\limsup_{\varepsilon\to0}\dim_C^{\Psi_f}(\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f)) \leq t_{u}^*.\\[-44pt] \end{align*} $$

Finally, to complete the proof of Theorem C, assume that $\mu $ is an SRB measure from now on, then $h_\mu (f)=\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _u(\mu )$ . Thus, $P_\mu (f, \Phi _f(u))=0$ . Since the Carathéodory singular dimension with respect to $\Psi _f$ is always less than u, by property (i) of Theorem C, we have that

$$ \begin{align*} \lim_{\varepsilon\to0}\dim_C^{\Phi_f}(\Lambda_\varepsilon\cap W^u_{\mathrm{loc}}(x,f)) =u. \end{align*} $$

If $\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _{m_0}(\mu )\geq 0$ , then $P_\mu (f, \Xi _f(m_0-u))\ge 0$ since $\mu $ is an SRB measure for f. Consider $f^{-1}$ , by Margulis–Ruelle inequality, we have that

which implies that $P_\mu (f, \Xi _f(m_0-u))\le 0$ . Hence, we have that

$$ \begin{align*} P_\mu(f, \Xi_f(m_0-u))= 0. \end{align*} $$

Thus, we have that $t_s^*=m_0-u$ . By the definition of Lyapunov dimension, we have that $\dim _L\mu =m_0=u+t_s^*$ .

Now, we assume that $\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _{m_0}(\mu )< 0$ and let $\ell $ be the largest integer such that $\unicode{x3bb} _1(\mu )+\unicode{x3bb} _2(\mu )+\cdots +\unicode{x3bb} _{\ell }(\mu )\geq 0$ . By a standard computation, one can show that

Combining with

one has

$$ \begin{align*}\dim_L\mu=u+t_s^*.\end{align*} $$

This completes the proof of Theorem C.

Acknowledgements

This work is partially supported by The National Key Research and Development Program of China (2022YFA1005802). Y.C. is partially supported by NSFC (12371194), Y.Z. is partially supported by NSFC (12271386) and Qinglan project of Jiangsu Province.

References

Abdenur, F., Bonatti, C. and Crovisier, S.. Nonuniform hyperbolicity for ${C}^1$ -generic diffeomorphisms. Israel J. Math. 183 (2011), 160.10.1007/s11856-011-0041-5CrossRefGoogle Scholar
Avila, A., Crovisier, S. and Wilkinson, A.. ${C}^1$ density of stable ergodicity. Adv. Math. 379 (2021), 107496.10.1016/j.aim.2020.107496CrossRefGoogle Scholar
Ban, J., Cao, Y. and Hu, H.. The dimension of a non-conformal repeller and an average conformal repeller. Trans. Amer. Math. Soc. 362 (2010), 727751.10.1090/S0002-9947-09-04922-8CrossRefGoogle Scholar
Barreira, L., Pesin, Y. and Schmeling, J.. Dimension and product structure of hyperbolic measures. Ann. of Math. (2) 149(3) (1999), 755783.10.2307/121072CrossRefGoogle Scholar
Barreira, L. and Wolf, C.. Pointwise dimension and ergodic decompositions. Ergod. Th. & Dynam. Sys. 26 (2006), 653671.10.1017/S0143385705000672CrossRefGoogle Scholar
Cao, Y., Feng, D. and Huang, W.. The thermodynamic formalism for sub-additive potentials. Discrete Contin. Dyn. Syst. 20 (2008), 639657.10.3934/dcds.2008.20.639CrossRefGoogle Scholar
Cao, Y., Hu, H. and Zhao, Y.. Nonadditive measure-theoretic pressure and applications to dimensions of an ergodic measure. Ergod. Th. & Dynam. Sys. 33(3) (2013), 831850.10.1017/S0143385712000090CrossRefGoogle Scholar
Cao, Y., Pesin, Y. and Zhao, Y.. Dimension estimates for non-conformal repellers and continuity of sub-additive topological pressure. Geom. Funct. Anal. 29 (2019), 13251368.10.1007/s00039-019-00510-7CrossRefGoogle Scholar
Cao, Y., Wang, J. and Zhao, Y.. Dimension approximation in smooth dynamical systems. Ergod. Th. & Dynam. Sys. 44 (2024), 383407.10.1017/etds.2023.26CrossRefGoogle Scholar
Chung, Y.. Shadowing properties of non-invertible maps with hyperbolic measures. Tokyo J. Math. 22 (1999), 145166.10.3836/tjm/1270041619CrossRefGoogle Scholar
Climenhaga, V., Pesin, Y. and Zelerowicz, A.. Equilibrium states in dynamical systems via geometric measure theory. Bull. Amer. Math. Soc. (N.S.) 56(4) (2019), 569610.10.1090/bull/1659CrossRefGoogle Scholar
Fang, J., Cao, Y. and Zhao, Y.. Measure theoretic pressure and dimension formula for non-ergodic measures. Discrete Contin. Dyn. Syst. 40(5) (2020), 27672789.10.3934/dcds.2020149CrossRefGoogle Scholar
Frederickson, P., Kaplan, J., Yorke, E. and Yorke, J.. The Lyapunov dimension of strange attractors. J. Differential Equations 49(2) (1983), 185207.10.1016/0022-0396(83)90011-6CrossRefGoogle Scholar
Gelfert, K.. Repellers for non-uniformly expanding maps with singular or critical points. Bull. Braz. Math. Soc. (N.S.) 41 (2010), 237257.10.1007/s00574-010-0012-1CrossRefGoogle Scholar
Gelfert, K.. Horseshoes for diffeomorphisms preserving hyperbolic measures. Math. Z. 283 (2016), 685701.10.1007/s00209-016-1618-9CrossRefGoogle Scholar
He, L., Lv, J. and Zhou, L.. Definition of measure-theoretic pressure using spanning sets. Acta Math. Sin. (Engl. Ser.) 20(4) (2004), 709718.10.1007/s10114-004-0368-5CrossRefGoogle Scholar
Jordan, T. and Pollicott, M.. The Hausdorff dimension of measures for iterated function systems which contract on average. Discrete Contin. Dyn. Syst. 22 (2012), 235246.Google Scholar
Katok, A.. Lyapunov exponents, entropy and periodic orbits for diffeomorphisms. Publ. Math. Inst. Hautes Études Sci. 51 (1980), 137173.10.1007/BF02684777CrossRefGoogle Scholar
Katok, A. and Hasselblatt, B.. Introduction to the Modern Theory of Dynamical Systems (Encyclopedia of Mathematics and Its Applications, 54). Cambridge University Press, Cambridge, 1995.10.1017/CBO9780511809187CrossRefGoogle Scholar
Ledrappier, F.. Dimension of invariant measures. Proceedings of the Conference on Ergodic Theory and Related Topics, II (Georgenthal, 1986) (Mathematics in Stuttgart, 94). Eds. H. Michel, K. Häsler and V. Warstat. Teubner-Tecte, Leipzig, 1987, pp. 116124.Google Scholar
Ledrappier, F. and Young, L. S.. The metric entropy of diffeomorphisms. I. Ann. of Math. (2) 122(3) (1985), 509539.10.2307/1971328CrossRefGoogle Scholar
Ledrappier, F. and Young, L. S.. The metric entropy of diffeomorphisms. II. Relations between entropy, exponents and dimension. Ann. of Math. (2) 122(3) (1985), 540574.10.2307/1971329CrossRefGoogle Scholar
Mañé, R.. A proof of Pesin’s formula. Ergod. Th. & Dynam. Sys. 1 (1981), 95102.10.1017/S0143385700001188CrossRefGoogle Scholar
Mendoza, L.. Ergodic attractors for diffeomorphisms of surfaces. J. Lond. Math. Soc. (2) 37 (1988), 362374.10.1112/jlms/s2-37.2.362CrossRefGoogle Scholar
Misiurewicz, M. and Szlenk, W.. Entropy of piecewise monotone mappings. Studia Math. 67 (1980), 4563.10.4064/sm-67-1-45-63CrossRefGoogle Scholar
Oseledets, V.. A multiplicative ergodic theorem. Trans. Moscow Math. Soc. 19 (1968), 197231.Google Scholar
Persson, T. and Schmeling, J.. Dyadic diophantine approximation and Katok’s horseshoe approximation. Acta Arith. 132 (2008), 205230.10.4064/aa132-3-2CrossRefGoogle Scholar
Pesin, Y.. Characteristic Lyapunov exponents and smooth ergodic theory. Russian Math. Surveys 32 (1977), 55114.10.1070/RM1977v032n04ABEH001639CrossRefGoogle Scholar
Pesin, Y.. Dimension Theory in Dynamical Systems. Contemporary Views and Applications (Chicago Lectures in Mathematics). University of Chicago Press, Chicago, IL, 1997.10.7208/chicago/9780226662237.001.0001CrossRefGoogle Scholar
Przytycki, F. and Urbański, M.. Conformal Fractals: Ergodic Theory Methods (London Mathematical Society Lecture Note Series, 371). Cambridge University Press, Cambridge, 2010.10.1017/CBO9781139193184CrossRefGoogle Scholar
Sánchez-Salas, F. J.. Ergodic attractors as limits of hyperbolic horseshoes. Ergod. Th. & Dynam. Sys. 22(2) (2002), 571589.10.1017/S0143385702000287CrossRefGoogle Scholar
Shu, L.. Dimension theory for invariant measures of endomorphisms. Comm. Math. Phys. 298 (2010), 6599.10.1007/s00220-010-1059-yCrossRefGoogle Scholar
Walters, P.. An Introduction to Ergodic Theory. Springer-Verlag, New York, 1982.10.1007/978-1-4612-5775-2CrossRefGoogle Scholar
Wang, J., Qu, C. and Cao, Y.. Dimension approximation for diffeomorphisms preserving hyperbolic SRB measures. J. Differential Equations 337 (2022), 294322.10.1016/j.jde.2022.07.041CrossRefGoogle Scholar
Yang, Y.. Horseshoes for ${C}^{1+\alpha }$ mappings with hyperbolic measures. Discrete Contin. Dyn. Syst. 35 (2015), 51335152.10.3934/dcds.2015.35.5133CrossRefGoogle Scholar
Young, L. S.. Dimension, entropy and Lyapunov exponents. Ergod. Th. & Dynam. Sys. 2 (1982), 109129.10.1017/S0143385700009615CrossRefGoogle Scholar