1 Introduction
Entropy is one of the most widely used notions in the characterization of the complexity of topological dynamical systems. In 1965, Adler et al. [Reference Adler, Konheim and McAndrew1] defined topological entropy. In 1973, Bowen [Reference Bowen4] introduced the topological entropy of any subsets resembling Hausdorff dimension. Later in 1984, inspired by Bowen’s approach, Pesin and Pitskel [Reference Pesin25, Reference Pesin and Pitskel26] extended the notion of topological pressure to arbitrary subsets and established a variational principle which is a generalization of classical variational principle established by Goodwyn, Dinaburg, and Goodman [Reference Dinaburg7, Reference Goodman13, Reference Goodwyn14].
In 1983, Brin and Katok [Reference Brin and Katok2] gave the topological version of the Shannon–McMillan–Breiman theorem with a local decomposition of the metric entropy. Further, Feng and Huang [Reference Feng and Huang11] introduced the notion of measure-theoretical entropy for any (not necessarily invariant) Borel measure which is a modification of Brin and Katok’s local metric entropy and established the variational principle between Bowen’s topological entropy and measure-theoretical entropy for any non-empty compact subset.
In 1975, Kieffer [Reference Kieffer22] firstly introduced entropy for amenable group actions. In 2000, Rudolph and Weiss [Reference Rudolph and Weiss28], the properties of entropy for amenable group actions was elaborately explored. In particular, the classical variational principle for sofic group actions was established by Kerr and Li [Reference Kerr and Li19–Reference Kerr and Li21]. The variational principle of topological entropy and measure-theoretical entropy for amenable group actions was obtained by Huang et al. [Reference Huang, Li and Zhou17]. The variational principle related to the Bowen entropy and the Brin-Katok local entropy with respect to invariant measures for amenable group actions was established by Zheng and Chen [Reference Zheng and Chen31]. The variational principle of packing entropy and measure-theoretical entropy for amenable group actions was obtained by Dou et al. [Reference Dou, Zheng and Zhou9]. For other recent related work, we refer to [Reference Bowen4–Reference Deninger6, Reference Dou and Zhang8, Reference Huang, Ye and Zhang15, Reference Huang, Liu and Zhu16–Reference Kong and Chen18, Reference Kerr and Li20, Reference Ren and Sun27, Reference Zhao30, Reference Zheng and Chen31, Reference Zheng and Chen32]. One may refer to Ornstein and Weiss [Reference Ornstein and Weiss24], or Kerr and Li [Reference Kerr and Li21] for more details on dynamics for amenable group actions.
In 2015, Zhao and Pesin [Reference Zhao and Pesin33] defined the scaled topological entropy and scaled measure entropy. In this paper, we will work in the frame of countable discrete amenable group actions and introduce scaled measure entropy and scaled local entropy for any Borel measure. In particular, we prove a variational principle between scaled topological entropy and scaled local entropy:
where K is any non-empty compact subset of X. It is a scaled version of the variational principles obtained by Feng and Huang for continuous maps [Reference Feng and Huang11] and by Huang et al. [Reference Huang, Li and Zhou17] for amenable group actions. It worth to point out that the variational principle in [Reference Pesin25, Reference Pesin and Pitskel26] is not true for scaled entropies with respect to nontrivial sequences, we refer to [Reference Zhao and Pesin33, Example 4.3] for the counter examples. Meanwhile, our result also holds for continuous maps, we would like to directly present this result in a more general setting.
The paper is organized as follows. In Sections 2 and 3, we introduce the notions of scaled topological entropy, scaled weighted topological entropy, scaled measure entropy $E_{\mu }$ and scaled local entropy $h_{\mu }$ for amenable group actions and investigate their properties. We also give the definitions of equivalent scaled sequences and equivalent F $\unicode{xf8} $ lner sequences. In Section 4, a variational principle is established, i.e., the scaled topological entropy is the supremum over all Borel probability measure of the scaled local entropy.
2 Scaled topological entropy
In this section, we give the definitions of scaled topological entropy, lower and upper scaled topological entropies on an arbitrary subset and some related properties. Let $(X,G)$ be a topological dynamical system, where X is a compact metric space and G is a discrete countable amenable group. A group G is amenable if it admits a left invariant mean. This is equivalent to the existence of a sequence of finite subsets $\{F_{n}\}$ of G which are asymptotically invariant, i.e.,
Such sequences are called F $\unicode{xf8} $ lner sequence. One may refer to Ornstein and Weiss [Reference Ornstein and Weiss24], or Kerr and Li [Reference Kerr and Li21] for more details on dynamics for amenable group actions.
Let $\{F_{n}\}$ be a F $\unicode{xf8} $ lner sequence in G. Throughout this paper, we always let the F $\unicode{xf8} $ lner sequence $\{F_{n}\}$ be fixed.
2.1 Scaled topological entropy
We follow the approach described in [Reference Pesin25]. Let $\mathcal {U}$ be an open cover of X. Denote by $\mathcal {W}_{F_{n}}(\mathcal {U})$ the collection of families $\mathbf {U}=\{U_{g}\}_{g\in F_{n}}$ of length $m(\mathbf {U})=|F_{n}|$ with $U_{g}\in \mathcal {U}$ and by $\mathcal {W}(\mathcal {U})=\bigcup _{n\geq 1}\mathcal {W}_{F_{n}}(\mathcal {U})$ . For $\mathbf {U}\in \mathcal {W}_{F_{n}}(\mathcal {U})$ define
Let $Z\subset X$ be a subset of X. We say that a collection of strings $\Gamma \subset \mathcal {W}(\mathcal {U})$ covers Z if $\bigcup _{\mathbf {U\in \Gamma }}X(\mathbf {U})\supset Z$ .
We call a sequence of positive numbers $\mathbf {a}=\{a(n)\}_{n\geq 1}$ a scaled sequence if it is positive and monotonically increasing to infinity. We denote by $\mathcal {SS}$ the set of all scaled sequences.
Let $\mathbb {N}=\{1,2,3,\cdots \}$ . Given a subset $Z\subset X$ , $s\geq 0$ , $N\in \mathbb {N}$ and a scaled sequence $\mathbf {a}\in \mathcal {SS}$ , let
where the infimum is taken over all covers $\Gamma \subset \bigcup _{n\geq N}\mathcal {W}_{F_{n}}(\mathcal {U})$ of Z. It is easy to see that $M(Z,s,N,\{F_{n}\},\mathcal {U},\mathbf {a})$ is monotone in N. Define
By the construction of Carath $\acute {e}$ odory dimension characteristics one can show that when s goes from $-\infty $ to $+\infty $ , $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {a})$ jump from $+\infty $ to 0 at a unique critical value. Hence, let
Definition 2.1 Let $(X,G)$ be a topological dynamical system. For $\mathbf {a}\in \mathcal {SS}$ , $Z\subset X$ ,
is called the scaled topological entropy of $(X,G)$ on the set Z (with respect to the sequence $\mathbf {a}\in \mathcal {SS}$ and the F $\unicode{xf8} $ lner sequence $\{F_{n}\}$ ).
Given a subset $Z\subset X$ , $s\geq 0$ , $N\in \mathbb {N}$ and a scaled sequence $\mathbf {a}\in \mathcal {SS}$ , define
where the infimum is taken over all covers $\Gamma \subset \mathcal {W}_{F_{N}}(\mathcal {U})$ of Z. We set
and define the critical values of $\underline {r}(Z,s,\mathcal {U},\{F_{n}\},\mathbf {a})$ and as
respectively. For existences of the above critical values, we refer to [Reference Pesin25, page 16].
Definition 2.2
and
are called the lower and upper scaled topological entropy of $(X,G)$ on the set Z.
Let $\mathcal {U}$ be an open cover of X and $|\mathcal {U}|=\text {max}\{\text {diam}(U):U\in \mathcal {U}\}$ denote the diameter of the cover $\mathcal {U}$ . In what follows we use the notation $\mathcal {E}$ for either E, $\underline {E,}$ or .
Proposition 2.1 Let $\mathbf {a}\in \mathcal {SS}$ and $\mathcal {U}$ be an open cover of X. Then $\lim \limits _{|\mathcal {U}|\rightarrow 0}\mathcal {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})$ exists and is equal to $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})$ .
Proof Let $\mathcal {V}$ be a finite open cover of X with diameter smaller than the Lebesgue number of $\mathcal {U}$ . Each element $V\in \mathcal {V}$ is contained in some element $U(V)\in \mathcal {U}$ . For any $n\in \mathbb {N}$ and any string $\mathbf {V}=\{V_{g}\}_{g\in F_{n}}\in \mathcal {W}_{F_{n}}(\mathcal {V}),$ there exists a corresponding string $\mathbf {U}(\mathbf {V})=\{U(V_{g})\}_{g\in F_{n}}\in {W}_{F_{n}}(\mathcal {U})$ . If $\Gamma \subset \mathcal {W}_{F_{n}}(\mathcal {V})$ covers a set $Z\subset X$ , then $\mathbf {U}(\Gamma )=\{\mathbf {U}(\mathbf {V}):\mathbf {V}\in \Gamma \}$ also covers Z. By definition of the scaled topological entropy, $M(Z,s,N,\{F_{n}\},\mathcal {U},\mathbf {a})\leq M(Z,s,N,\{F_{n}\},\mathcal {V},\mathbf {a})$ . Then, $E_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})\leq E_{Z}(\{F_{n}\},\mathcal {V},\mathbf {a})$ . Therefore, $E_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})\leq \liminf \limits _{|\mathcal {V}|\rightarrow 0}E_{Z}(\{F_{n}\},\mathcal {V},\mathbf {a})$ , hence
This implies that
$\underline {E}_{Z}(\{F_{n}\},\mathbf {a})=\lim \limits _{|\mathcal {U}|\rightarrow 0}\underline {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})$ and can be proved in a similar manner.
Now, we describe the second equivalent definition of scaled topological entropy. Let X be a compact metric space. Given $\epsilon>0,n\in \mathbb {N}$ and $x,y\in X$ , denote by $d_{F_{n}}(x,y)=\max _{g\in F_n}d(gx,gy)$ and $B_{F_{n}}(x,\epsilon )$ the open Bowen ball of radius $\epsilon>0$ in the metric $d_{F_{n}}$ around x, i.e., $B_{F_{n}}(x,\epsilon )=\{y \in X:d_{F_{n}}(x,y)< \epsilon \}$ . We follow the approach described in [Reference Pesin25], for each subset $Z\subset X$ , $\mathbf {a}\in \mathcal {SS}$ , $N\in {\mathbb N}$ and $\epsilon ,s>0$ , set
We note that $M(\cdot ,s,N,\{F_{n}\},\epsilon ,\mathbf {a})$ is an outer measure on X.
Since $M(Z,s,N,\{F_{n}\},\epsilon ,\mathbf {a})$ is monotonically increasing with respect to N,
It is easy to show that there is a jump-up value
Let
For any subset $Z\subset X$ and $\mathbf {a}\in \mathcal {SS}$ , let $\aleph (Z,n,\epsilon )$ denote the smallest number of Bowen’s balls $\{B_{F_{n}}(x,\epsilon )\}$ whose union covers the set Z,
and
Set
and
We have the following result.
Proposition 2.2 For any subset $Z\subset X$ and $\mathbf {a}\in \mathcal {SS}$ , we have
(a) $E_{Z}(\{F_{n}\},\mathbf {a})=E_{Z}^{B_1}(\{F_{n}\},\mathbf {a})=E_{Z}^{B_2}(\{F_{n}\},\mathbf {a});$
(b) $\underline {E}_{Z}(\{F_{n}\},\mathbf {a})=\underline {E}_{Z}^{{B_{2}}}(\{F_{n}\},\mathbf {a})$ ; .
Proof We only give the proof of (a). (b) can be proved in a similar manner as Proposition 2.3.
Firstly, we prove $E_{Z}^{B^{1}}(\{F_{n}\},\mathbf {a})\leq E_{Z}(\{F_{n}\},\mathbf {a})$ . Let $\mathcal {U}$ be a finite open cover of X. If s satisfies $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {a})=0$ , then by definition for every $\epsilon>0$ , there exists $N^{'}\in {\mathbb N}$ such that for each $N\in {\mathbb N}$ with $N>N^{'}$ ,
where the infimum is taken over all covers $\Gamma \subset \bigcup _{n\geq N}\mathcal {W}_{F_{n}}(\mathcal {U})$ of Z. So there exists $\Gamma _{N}$ that satisfies $\Gamma _{N}\subset \bigcup _{n\geq N}\mathcal {W}_{F_{n}}(\mathcal {U})$ and $\Gamma _{N}$ covers Z, such that
For any $n\in {\mathbb N}$ and $\mathbf {U}\in \mathcal {W}_{F_{n}}(\mathcal {U})$ , $n^{\{F_{n}\}}_{\mathcal {U}}(X(\mathbf {U}))\geq |F_{n}|$ . Then
and
This implies that
and hence
Next, we prove $ E_{Z}(\{F_{n}\},\mathbf {a})\leq E_{Z}^{B_{1}}(\{F_{n}\},\mathbf {a})$ . Let $\mathcal {U}$ be a finite open cover of X. If s satisfies $M_{\{F_{n}\},\mathcal {U}}(Z,s,\mathbf {a})=0$ , then by definition for every $\epsilon>0$ , there exists $k^{'}\in {\mathbb N}$ such that for each $k\in {\mathbb N}$ with $k>k^{'}$ ,
Thus there exists $\mathcal {D}=\{D_{i}\}_{i=1}^{\infty }\in \mathcal {G}(G,\mathcal {U},Z,k)$ such that
Without loss of generality, we can assume that each $D_i\in \mathcal {D}$ is open with
Indeed, if $n^{\{F_{n}\}}_{\mathcal {U}}(D_i)<\infty $ , we can take $\widehat {D}_i=X(\mathbf {U})$ , where $X(\mathbf {U})= \bigcap _{g\in F_{n}}g^{-1}U_{g}$ and for each $g\in F_n$ , $gD_i\subset U_g$ . It is easy to see that $n^{\{F_{n}\}}_{\mathcal {U}}(D_i)=n^{\{F_{n}\}}_{\mathcal {U}}(\widehat {D}_i)$ and $D_i\subset \widehat {D}_i$ . If $n^{\{F_{n}\}}_{\mathcal {U}}(D_i)=\infty $ , we can take $n\in {\mathbb N}$ sufficiently large and $\widehat {D}_i=X(\mathbf {U})$ , where $X(\mathbf {U})= \bigcap _{g\in F_{n}}g^{-1}U_{g}$ and for each $g\in F_n$ , $gD_i\subset U_g$ , so that $\exp (-sa(n^{\{F_{n}\}}_{\mathcal {U}}(\widehat {D}_i)))<\frac {\epsilon }{2^{i+1}}$ .
Then we can assume that there exists an open cover $\mathcal {D}=\{D_{i}\}_{i=1}^{\infty }$ of Y, such that for every $D_{i}\in \mathcal {D}$ , there exists $\mathbf {U}_i\in \mathcal {W}_{F_{n}}(\mathcal {U})$ with $D_{i}=X(\mathbf {U}_i)$ and
Then $Z\subset \bigcup _{i=1}^{\infty } D_{i}=\bigcup _{i=1}^{\infty } X(\mathbf {U}_i)$ and $\sum _{i=1}^{\infty }\exp (-sa(m(\mathbf {U})))<\epsilon .$ By the arbitrariness of $\epsilon $ , we get $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {a})=0$ and $E_{Z}(\{F_{n}\},\mathbf {a})\leq E_{Z}^{B_{1}}(\{F_{n}\},\mathbf {a}).$ So
Let $\delta (\mathcal {U})$ be the Lebesgue number of $\mathcal {U}$ . Clearly, for every $x\in X$ and $n\in {\mathbb N}$ , if $x\in X(\mathbf {U})$ for some string $\mathbf {U}\in \mathcal {W}_{F_{n}}(\mathcal {U})$ , then $B_{F_{n}}(x,\delta (\mathcal {U}))\subset X(\mathbf {U})\subset B_{F_{n}}(x,|\mathcal {U}|)$ . By Proposition 2.1, this implies that
Therefore, $E_{Z}(\{F_{n}\},\mathbf {a})=E_{Z}^{B_1}(\{F_{n}\},\mathbf {a})=E_{Z}^{B_2}(\{F_{n}\},\mathbf {a})$ .
2.2 Properties of scaled topological entropy
For any subset $Z\subset X$ , any open cover $\mathcal {U}$ of X and $\mathbf {a}\in \mathcal {SS}$ , let $\aleph (\mathcal {U},Z)$ denote the number of sets in a finite subcover of $\mathcal {U}$ with the smallest cardinality. We have the following equivalent definition of the lower and upper scaled topological entropy.
Proposition 2.3 Let $Z\subset X$ , $\mathbf {a}\in \mathcal {SS}$ and $\mathcal {U}$ be an open cover of X, then
Proof We will prove the first equality, the second one can be proved in a similar fashion. Let us put
Given $\eta>0$ , one can choose a subsequence $\{n_i\}$ such that
When $i\in {\mathbb N}$ is sufficiently large, we have
and hence
Let $i\rightarrow \infty $ , we have $\gamma \leq \underline {\beta }+\eta $ , therefore, $\gamma \leq \underline {\beta }.$
Let us choose a subsequence $\{n_i\}$ such that
When $i\in {\mathbb N}$ is sufficiently large, we have
and hence
Let $i\rightarrow \infty $ , we have $\gamma \geq \underline {\beta }-\eta $ , therefore, $\gamma \geq \underline {\beta }.$
So $\gamma =\underline {\beta }$ , i.e.,
Given two open covers $\mathcal {U}$ and $\mathcal {V}$ of X, we say that $\mathcal {U}$ is finer than $\mathcal {V}$ if for every $U\in \mathcal {U}$ there is an element $V\in \mathcal {V}$ such that $U\subset V$ . We denote by $\mathcal {U}\succeq \mathcal {V}$ . Set
In what follows, we use the notation ${\mathcal {E}}$ for either E or ${\underline {E}}$ or . The following Propositions describe some basic properties of scaled topological entropy and lower (upper) scaled topological entropies.
Proposition 2.4 Let $\mathcal {U}$ and $\mathcal {V}$ be two open covers of X, $Z\subset X$ and $\mathbf {a}\in \mathcal {SS}$ , the following properties hold:
-
(1) If $\mathcal {U}\preceq {\mathcal {V}}$ , then $\mathcal {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})\leq \mathcal {E}_{Z}(\{F_{n}\},\mathcal {V},\mathbf {a})$ ;
-
(2)
and
Proof $(1)$ Since $\mathcal {U}\preceq \mathcal {V}$ , each element $V\in \mathcal {V}$ is contained in some element in $\mathcal {U}$ which we denote by $U(V)$ . Therefore, for each string $\mathbf {V}\in \mathcal {W}_{F_{n}}(\mathcal {V})$ there exists a corresponding string $\mathbf {U}(\mathbf {V})\in {W}_{F_{n}}(\mathcal {U})$ . This yields that
and hence
and
Let $\Gamma \subset \mathcal {W}(\mathcal {V})$ be a collection of strings that covers Z. The corresponding collection of strings $\{\mathbf {U}(\mathbf {V}):\mathbf {V}\in \Gamma \}\subset \mathcal {W}(\mathcal {U})$ also covers Z. This implies that $M(Z,s,N,\{F_{n}\},\mathcal {U},\mathbf {a})\leq M(Z,s,N,\{F_{n}\},\mathcal {V},\mathbf {a})$ for each $s\geq 0$ and $N>0$ . Thus, $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {a})\leq M(Z,s,\{F_{n}\},\mathcal {V},\mathbf {a})$ . The first statement follows.
$(2)$ The last statement follows immediately from the definitions.
The following proposition shows that the scaled topological entropy as well as lower and upper scaled topological entropies for amenable group actions are invariant under a topological conjugacy. Its proof is similar to the proofs of [Reference Pesin25, Theorems 1.3 and 2.5].
Definition 2.3 [Reference Kerr and Li21, Definition 1.3]
Two continuous actions $G \curvearrowright X_1$ and $G \curvearrowright X_2$ of the same group on compact metric spaces are said to be topologically conjugate if there is a homeomorphism $\phi : X_1 \to X_2$ such that $\phi (gx)=g\phi (x)$ for all $x\in X_1$ and $g\in G$ .
Proposition 2.5 Given two topologically conjugate actions $G \curvearrowright X_1$ and $G \curvearrowright X_2$ and $\mathbf {a}\in \mathcal {SS}$ , then for each $Z\subset X_{1}$ and each open cover $\mathcal {U}$ of $X_{2}$ we have
In particular, $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})=\mathcal {E}_{\phi (Z)}(\{F_{n}\},\mathbf {a})$ .
Proposition 2.6 The following statements hold:
-
(1) If $Z_{1}\subset Z_{2}$ , then $\mathcal {E}_{Z_{1}}(\{F_{n}\},\mathcal {U},\mathbf {a})\leq \mathcal {E}_{Z_{2}}(\{F_{n}\},\mathcal {U},\mathbf {a})$ , hence $\mathcal {E}_{Z_{1}}(\{F_{n}\},\mathbf {a})\leq \mathcal {E}_{Z_{2}}(\{F_{n}\},\mathbf {a})$ ;
-
(2) If $Z_{i}\subset X$ , $i\geq 1$ and $Z=\bigcup _{i\geq 1}Z_{i}$ , then $E_{Z}(\{F_{n}\},\mathbf {a})=\sup _{i\geq 1}E_{Z_{i}}(\{F_{n}\},\mathbf {a})$ , $\underline {E}_{Z}(\{F_{n}\},\mathbf {a})\geq \sup _{i\geq 1}\underline {E}_{Z_{i}}(\{F_{n}\},\mathbf {a})$ and .
Proof $(1)$ The statements follow directly from the definitions.
$(2)$ By $(1)$ , $\mathcal {E}_{Z_{i}}(\{F_{n}\},\mathbf {a})\leq \mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})$ and hence $\sup _{i\geq 1}\mathcal {E}_{Z_{i}}(\{F_{n}\},\mathbf {a})\leq \mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})$ .
Now we are left to show that $E_{Z}(\{F_{n}\},\mathbf {a})\geq \sup _{i\geq 1}E_{Z_{i}}(\{F_{n}\},\mathbf {a})$ . In fact, suppose $E_{Z_{i}}(\{F_{n}\},\mathbf {a})<s(i=1,2,\cdots )$ , it follows that $M(Z_{i},s,\{F_{n}\},\mathcal {U},\mathbf {a})=0$ , and hence $M(\bigcup _{i\geq 1}Z_{i},s,\{F_{n}\},\mathcal {U},\mathbf {a})=0$ . Then $E_{Z=\bigcup _{i\geq 1}Z_{i}}(\{F_{n}\},\mathbf {a})<s$ . This implies that, $E_{Z}(\{F_{n}\},\mathbf {a})\leq \sup _{i\geq 1}E_{Z_{i}}(\{F_{n}\},\mathbf {a})$ .
2.3 Equivalent scaled sequences and equivalent F $\unicode{xf8} $ lner sequences
We call two scaled sequences $\mathbf {a}, \mathbf {b}\in \mathcal {SS}$ equivalent and we write $\mathbf {a}\sim \mathbf {b}$ if the following condition holds
Obviously, $\thicksim $ defines an equivalence relation on $\mathcal {SS}$ . Let $\mathbf {a}\in \mathcal {SS}$ , we denote its equivalence class by $[\mathbf {a}]:=\{\mathbf {b}\in \mathcal {SS}:\mathbf {b}\sim \mathbf {a}\}$ and we let $\mathcal {A}:=\mathcal {SS}/\thicksim $ . Given two equivalence classes $[\mathbf {a}], [\mathbf {b}]\in \mathcal {A}$ , we say that $[\mathbf {a}]\preceq [\mathbf {b}]$ if for each $\mathbf {a}\in [\mathbf {a}]$ and $\mathbf {b}\in [\mathbf {b}]$ the following holds
The following result is immediate.
Proposition 2.7 For $\mathbf {a},\mathbf {b}\in \mathcal {SS}$ , $Z\subset X$ and each open cover $\mathcal {U}$ of X, the following properties hold:
-
(1) If $a(|F_{n}|)\leq b(|F_{n}|)$ for all sufficiently large $n\in {\mathbb N}$ , then $\mathcal {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})\geq \mathcal {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {b})$ and $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{Z}(\{F_{n}\},\mathbf {b})$ ;
-
(2) For each $K> 0$ we have that
$$ \begin{align*} K\cdot\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},K\mathbf{a})=\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},\mathbf{a}),K\cdot\mathcal{E}_{Z}(\{F_{n}\},K\mathbf{a})=\mathcal{E}_{Z}(\{F_{n}\},\mathbf{a}), \end{align*} $$
where $K\mathbf {a}=\{K\cdot a(|F_{n}|)\}$ ;
-
(3) If there exists a constant $C>0$ such that $\frac {1}{C}\cdot b(|F_{n}|)\leq a(|F_{n}|)\leq C\cdot b(|F_{n}|)$ for all sufficiently large $n\in {\mathbb N}$ , then
$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},\mathbf{b})\leq\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},\mathbf{a})\leq C\cdot\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},\mathbf{b}) \end{align*} $$and$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{Z}(\{F_{n}\},\mathbf{b})\leq\mathcal{E}_{Z}(\{F_{n}\},\mathbf{a})\leq C\cdot\mathcal{E}_{Z}(\{F_{n}\},\mathbf{b}). \end{align*} $$
Remark 2.1 By Statement $(3)$ of Proposition 2.7, for each equivalence class $[\mathbf {a}]\in \mathcal {A}$ and for each $\mathbf {a_{1}},\mathbf {a_{2}}\in [\mathbf {a}]$ we have that $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{1}})=\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{2}})=0$ , or $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{1}})=\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{2}})=\infty $ or both $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{1}})$ and $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a_{2}})$ are positive and finite. In the first two cases, we write $\mathcal {E}_{Z}(\{F_{n}\},[\mathbf {a}])=0$ and $\mathcal {E}_{Z}(\{F_{n}\},[\mathbf {a}])=\infty $ respectively and in the third case, we say that $\mathcal {E}_{Z}(\{F_{n}\},[\mathbf {a}])$ is positive and finite. In this sense, entropy depends not on the scaled sequence but on its class of equivalence.
By Statement $(1)$ of Proposition 2.7, we have $\mathcal {E}_{Z}(\{F_{n}\},[\mathbf {a}])\leq \mathcal {E}_{Z}(\{F_{n}\},[\mathbf {b}])$ whenever $[\mathbf {a}]\succeq [\mathbf {b}]$ .
Theorem 2.1 If there is $[\mathbf {a}]\in \mathcal {A}$ such that $E_{Z}(\{F_{n}\},[\mathbf {a}])$ is positive and finite, then
In particular, there may exist at most one element in $(\mathcal {A,\preceq })$ such that the corresponding scaled topological entropy is positive and finite. Similar results hold for lower and upper scaled topological entropy.
Proof We shall prove the result for the scaled topological entropy $E_{Z}(\{F_{n}\},[\mathbf {a}])$ , the arguments for the lower and upper scaled topological entropies are similar.
Suppose there is $[\mathbf {a}]\in \mathcal {A}$ such that $E_{Z}(\{F_{n}\},[\mathbf {a}])$ is positive and finite. Then for each $[\mathbf {b}]\succeq [\mathbf {a}]$ ,
for arbitrary $\mathbf {a_{1}}=\{a_{1}(|F_{n}|)\}\in [\mathbf {a}]$ and $\mathbf {b_{1}}=\{b_{1}(|F_{n}|)\}\in [\mathbf {b}]$ . Let us fix such two scaled sequences $\mathbf {a_{1}}$ and $\mathbf {b_{1}}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $a_{1}(|F_{n}|)<\beta b_{1}(|F_{n}|)$ and hence, $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {a_{1}})\geq M(Z,s,\{F_{n}\},\mathcal {U},\beta \mathbf {b_{1}})$ . This implies that
i.e., $\beta E_{Z}(\{F_{n}\},\mathbf {a_{1}})\geq E_{Z}(\{F_{n}\},\mathbf {b_{1}})$ . Since $\beta $ is arbitrary, we conclude that
and hence
On the other hand, if $[\mathbf {b}]\preceq [\mathbf {a}]$ , then
for arbitrary $\mathbf {a_{2}}=\{a_{2}(|F_{n}|)\}\in [\mathbf {a}]$ and $\mathbf {b_{2}}=\{b_{2}(|F_{n}|)\}\in [\mathbf {b}]$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $b_{2}(|F_{n}|)<\beta a_{2}(|F_{n}|)$ and hence, $M(Z,s,\{F_{n}\},\mathcal {U},\mathbf {b_{2}})\geq M(Z,s,\{F_{n}\},\mathcal {U},\beta \mathbf {a_{2}})$ . It follows that $E_{Z}(\{F_{n}\},\mathbf {b_{2}})>\frac {1}{\beta }E_{Z}(\{F_{n}\},\mathbf {a_{2}})$ . Again since $\beta $ is arbitrary,
then
Next we discuss the equivalence of F $\unicode{xf8} $ lner sequence. Let $(X,G)$ be a topological dynamical system, we denote by $\mathcal {SF}(G)$ the set of all F $\unicode{xf8} $ lner sequences in G. We call two F $\unicode{xf8} $ lner sequences $\{F_n\}, \{Q_n\}\in \mathcal {SF}(G)$ equivalent with respect to $\mathbf {a}\in \mathcal {SS}$ and we write $\{F_n\}\stackrel {\mathbf {a}}{\sim }\{Q_n\}$ if the following condition holds:
Obviously, $\{F_n\}\stackrel {\mathbf {a}}{\sim }\{Q_n\}$ defines an equivalence relation on $\mathcal {SF}(G)$ . Let $\{F_n\}\in \mathcal {SF}(G)$ , we denote its equivalence class by $\Big [\{F_n\}\Big ]_{\mathbf {a}}:=\Big \{\{Q_n\}\in \mathcal {SF}(G):\{Q_n\}\{F_n\}\stackrel {\mathbf {a}}{\sim }\{Q_n\}\{F_n\}\Big \}$ and we let $\mathcal {F}(G)_{\mathbf {a}}:=\mathcal {SF}(G)/\stackrel {\mathbf {a}}{\sim }$ . Given two equivalence classes $\Big [\{F_n\}\Big ]_{\mathbf {a}}, \Big [\{Q_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ , we say that $\Big [\{F_n\}\Big ]_{\mathbf {a}}{\preceq }\Big [\{Q_n\}\Big ]_{\mathbf {a}}$ if for each $\{F_n\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ and $\{Q_n\}\in \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ the following holds:
The following result is immediate.
Proposition 2.8 For $\{F_n\},\{Q_n\}\in \mathcal {SF}(G)$ , $\mathbf {a}\in \mathcal {SS}$ , $Z\subset X$ , and each open cover $\mathcal {U}$ of X, the following properties hold:
-
(1) If $a(\vert F_{n}\vert )\leq a(\vert Q_{n}\vert )$ for all sufficiently large $n\in {\mathbb N}$ , then $\mathcal {E}_{Z}(\{F_{n}\},\mathcal {U},\mathbf {a})\geq \mathcal {E}_{Z}(\{Q_{n}\},\mathcal {U},\mathbf {a})$ and $\mathcal {E}_{Z}(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{Z}(\{Q_{n}\},\mathbf {a})$ ;
-
(2) If there exists a constant $C>0$ such that $\frac {1}{C}\cdot a(|Q_{n}|)\leq a(|F_{n}|)\leq C\cdot a(|Q_{n}|)$ for all sufficiently large $n\in {\mathbb N}$ , then
$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{Z}(\{Q_{n}\},\mathcal{U},\mathbf{a})\leq\mathcal{E}_{Z}(\{F_{n}\},\mathcal{U},\mathbf{a})\leq C\cdot\mathcal{E}_{Z}(\{Q_{n}\},\mathcal{U},\mathbf{a}) \end{align*} $$and$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{Z}(\{Q_{n}\},\mathbf{a})\leq\mathcal{E}_{Z}(\{F_{n}\},\mathbf{a})\leq C\cdot\mathcal{E}_{Z}(\{Q_{n}\},\mathbf{a}). \end{align*} $$
Remark 2.2 By Statement $(2)$ of Proposition 2.8, for each $\mathbf {a}\in \mathcal {SS}$ , equivalence class $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ and $\{F_n^*\},\{F_n^{**}\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ , we have that $\mathcal {E}_{Z}(\{F_n^{*}\},\mathbf {a})=\mathcal {E}_{Z}(\{F_n^{**}\},\mathbf {a})=0$ , or $\mathcal {E}_{Z}(\{F_n^{*}\},\mathbf {a})=\mathcal {E}_{Z}(\{F_n^{**}\},\mathbf {a})=\infty $ or both $\mathcal {E}_{Z}(\{F_n^{*}\},\mathbf {a})$ and $\mathcal {E}_{Z}(\{F_n^{**}\},\mathbf {a})$ are positive and finite. In the first two cases, we write $\mathcal {E}_{Z}(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a})=0$ and $\mathcal {E}_{Z}(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a})=\infty ,$ respectively, and in the third case, we say that $\mathcal {E}_{Z}(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a})$ is positive and finite. In this sense, entropy depends not on the F $\unicode{xf8} $ lner sequence but on its class of equivalence.
By Statement $(1)$ of Proposition 2.8, we have $\mathcal {E}_{Z}(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a})\leq \mathcal {E}_{Z}(\Big [\{Q_n\}\Big ]_{\mathbf {a}},\mathbf {a})$ whenever $\Big [\{F_n\}\Big ]_{\mathbf {a}}\succeq \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ .
Theorem 2.2 Let $\mathbf {a}\in \mathcal {SS}$ . If there is $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ such that $E_{Z}\Big(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a}\Big)$ is positive and finite, then
In particular, there may exist at most one element in $(\mathcal {F}(G)_{\mathbf {a}},\preceq )$ such that the corresponding scaled topological entropy is positive and finite. Similar results hold for lower and upper scaled topological entropy.
Proof We shall prove the result for the scaled topological entropy $E_{Z}\Big(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a}\Big)$ , the arguments for the lower and upper scaled topological entropies are similar.
Suppose there is $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ such that $E_{Z}\Big(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a}\Big)$ is positive and finite. Then for each $\Big [\{Q_n\}\Big ]_{\mathbf {a}}\succeq \Big [\{F_n\}\Big ]_{\mathbf {a}}$ ,
for arbitrary $\{F_{n}^*\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ and $\{Q_{n}^*\}\in \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ . Let us fix such two F $\unicode{xf8} $ lner sequences $\{F_{n}^*\}$ and $\{Q_{n}^*\}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $a(|F_{n}^*|)<\beta a(|Q_{n}^*|)$ and hence, $M(Z,s,\{F_{n}^*\},\mathcal {U},\mathbf {a})\geq M(Z,s,\\ \{Q_{n}^*\},\mathcal {U},\beta \mathbf {a})$ . By $(2)$ of Proposition 2.7,
i.e., $\beta E_{Z}(\{F_{n}^*\},\mathbf {a})\geq E_{Z}(\{Q_{n}^*\},\mathbf {a})$ . Since $\beta $ is arbitrary, we conclude that
and hence
On the other hand, if $\Big [\{F_n\}\Big ]_{\mathbf {a}}\succeq \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ , then
for arbitrary $\{F_{n}^{**}\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ and $\{Q_{n}^{**}\}\in \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $a(|Q_{n}^{**}|)<\beta a(|F_{n}^{**}|)$ and hence, $M(Z,s,\{Q_{n}^{**}\},\mathcal {U},\mathbf {a})\geq M(Z,s,\{F_{n}^{**}\},\mathcal {U},\beta \mathbf {a})$ . It follows that $E_{Z}(\{Q_{n}^{**}\},\mathbf {a})>\frac {1}{\beta }E_{Z}(\{F_{n}^{**}\},\mathbf {a})$ . Again since $\beta $ is arbitrary,
then
Remark 2.3 As is shown in Theorem 2.2, the scaled entropy depends on the choice of Følner sequence.
2.4 Scaled weighted topological entropy
For any positive function $f:X\rightarrow [0,\infty ), N\in \mathbb {N},\epsilon>0$ and $\mathbf {a}\in \mathcal {SS}$ , we define
where the infimum is taken over all finite or countable families $\{(B_{F_{n_{i}}}(x_i,\epsilon ),c_i)\}$ such that $x_i\in X, n_i\geq N,0<c_i<\infty $ and $\sum _ic_i\chi _{B_i}\geq f$ , where $B_{i}=B_{F_{n_{i}}}(x_i,\epsilon )$ . We note that $W(\cdot ,s,N,\epsilon ,\{F_{n}\},\mathbf {a})$ is an outer measure on X.
For $Z\subset X,f=\chi _{Z}$ , set $W(Z,s,N,\epsilon ,\{F_{n}\},\mathbf {a})=W(\chi _Z,s,N,\epsilon ,\{F_{n}\},\mathbf {a})$ . Clearly, the function $W(Z,s,N,\epsilon ,\{F_{n}\},\mathbf {a})$ does not decrease as N increases and $\epsilon $ decreases. So the following limits exist:
It’s not difficult to prove that there exists a critical value of parameter s, which we will denote by $h_{top}^W(Z,\{F_{n}\},\mathbf {a})$ , such that
We call $h_{top}^W(Z,\{F_{n}\},\mathbf {a})$ scaled weighted topological entropy of $(X,G)$ on the set Z (with respect to the sequence $\mathbf {a}\in \mathcal {SS}$ and the F $\unicode{xf8} $ lner sequence $\{F_{n}\}$ ).
2.5 Examples
Example 2.1 [Reference Zhao and Pesin33, Example 4.1]
Suppose X is a compact metric space and $G=\mathbb {Z}$ . Let $(X,G)$ is a topological dynamical system induced by an expansive homeomorphism $f:X\to X$ . Let $F_{n}=[0,n-1]\cap \mathbb {Z}$ for each $n\in \mathbb {N}$ and $\mathcal {V}$ be a generating open cover of X. Then $\{F_{n}\}$ is a F $\unicode{xf8} $ lner sequence. For $\mathbf {a}\in \mathcal {SS}$ , let $Z\subset X$ with
-
(1) ;
-
(2) There is an open cover $\mathcal {U}$ of X such that $\mathcal {U}$ is finer than $\mathcal {V}$ and $\aleph (\bigvee _{g\in F_{n}}g^{-1}\mathcal {U},Z)\to \infty $ as $n\to \infty $ . By [Reference Zhao and Pesin33, Proposition 2.2], if $\lim \limits _{n\to \infty }\frac {a(n)}{a(n+1)}=1$ , then
$$ \begin{align*} \underline{E}_{Z}(\{F_{n}\},\mathbf{a})=\liminf_{n\rightarrow\infty}\frac{1}{a(|F_{n}|)}\log\aleph\left(\bigvee_{g\in F_{n}}g^{-1}\mathcal{U},Z\right) \end{align*} $$andIf $a(|F_n|)=\aleph (\bigvee _{g\in F_{n}}g^{-1}\mathcal {U},Z)$ increases monotonically, then
Example 2.2 For $k,d,n\in \mathbb {N}$ , let $X=\{0,1,\ldots ,k\}^G$ , $G=\mathbb {Z}^d$ , $F_n=[-n+1,n-1]^d\cap \mathbb {Z}^d$ and $\sigma _g:X\to X$ be the natural shift action for $g\in G$ , then $\{F_n\}$ is a Følner sequence. Suppose $x=(x_g)_{g\in G}, y=(y_g)_{g\in G}\in X$ and $\mathbf {a}\in \mathcal {SS}$ with $a(n)=\log n$ , set $n(x,y)=\min \{n\in \mathbb {N}:~x_g=y_g \text {~for~all~} g\in F_n \text {~and~}x_g\neq y_g\text {~for~some~}g\in F_{n+1}\backslash F_{n}\},$ then $d(x,y)=\exp (-a(|F_{n(x,y)}|))$ is a compatible metric and $(X,G)$ is a topological dynamical system.
We claim that for any $Z\subset X$ , $E_{Z}(\{F_{n}\},\mathbf {a})=dim_H (Z)$ , where $dim_H (Z)$ is the Hausdorff dimension in $(X,d)$ . Moreover, by [Reference Mattila23, Theorem 8.19], for any $0\leq t\leq dim_H(X)$ , there exists a compact subset $Z_t\subset X$ such that $E_{Z_t}(\{F_{n}\},\mathbf {a})=t$ .
In fact, for any $n\in \mathbb {N}$ and $x\in X$ , let $C_n(x)=\{y\in X:~x_g=y_g,~g\in F_n\}$ be the cylinder set. For any $s\geq 0$ , one can show that
where $\mathcal {H}^s(Z)$ is the s-Hausdorff outer measure of Z and the infimum is taken over all finite or countable family $\mathcal {C}:=\{C_{k_i}(x_i)\}$ which covers Z with ${\sup _{i}\text {diam}(C_{k_i}(x_i))<\varepsilon .}$ For any sufficiently small $\epsilon>0$ , there exists $n\in \mathbb {N}$ such that $\exp (-a(|F_{n+1}|))\leq \epsilon <\exp (-a(|F_{n}|)).$ By the choice of the metric d, we have $B_{F_{k}}(x,\epsilon )=C_{k+n-1}(x)$ and $\text {diam}(C_{k}(x))=\exp (-a(|F_{k+1}|))$ for all $k\in \mathbb {N}$ and ${x\in X.}$ The desired conclusion follows exactly by the definitions of $E_{Z}(\{F_{n}\},\mathbf {a})$ and $dim_H (Z)$ , we refer to [Reference Simpson29, Theorem 4.2] for a similar and detailed proof.
3 Scaled measure entropy
In this Section, we introduce different types of scaled measure entropy and study their properties .
3.1 Scaled measure entropy
Let X be a compact Hausdorff space, $(X,G)$ be a measurable dynamical system. Denote by $\mathcal {M}(X)$ the set of all Borel probability measures on X. Denote by $\mathcal {M}(X,G)$ (respectively, $\mathcal {M}^{e}(X,G)$ ) the set of all G-invariant (respectively, ergodic G-invariant) Borel probability measure on X. We follow the approach described in [Reference Pesin25] and introduce the notion of scaled measure entropy using the inverse variational principle. Given $\mu \in \mathcal {M}(X,G)$ and $\mathbf {a}\in \mathcal {SS}$ , let
The fact that the second equality holds can be proven in the same way as [Reference Pesin25, p. 22], and for that reason we shall omit its proof.
Let
We call the quantity $E_{\mu }(\{F_{n}\},\mathbf {a})$ the scaled measure entropy of $(X,G)$ with respect to $\mu $ and $\mathbf {a}\in \mathcal {SS}$ . Let further
We call the quantities
respectively the lower and upper scaled measure entropy of $(X,G)$ with respect to $\mu $ and $\mathbf {a}\in \mathcal {SS}$ .
We describe another equivalent definition of scaled measure entropy. Given $\mu \in \mathcal {M}(X,G)$ and $\mathbf {a}\in \mathcal {SS}$ , let
and then let
Let further
Set
and
It is easy to see that
and
We have the following result.
Proposition 3.1 For any $\mu \in \mathcal {M}(X,G)$ and $\mathbf {a}\in \mathcal {SS}$ , we have
(a) $E_{\mu }(\{F_{n}\},\mathbf {a})=E_{\mu }^{B_1}(\{F_{n}\},\mathbf {a})=E_{\mu }^{B_2}(\{F_{n}\},\mathbf {a});$
(b) $\underline {E}_{\mu }(\{F_{n}\},\mathbf {a})=\underline {E}_{\mu }^{{B_{2}}}(\{F_{n}\},\mathbf {a})$ ; .
This proof is similar to the proof of Proposition 2.2, so we omit it.
3.2 Properties of the scaled measure entropy
In what follows, we use the notation $\mathcal {E}_{\mu }$ for either $E_{\mu }$ , $\underline {E}_{\mu }$ or . The following Propositions describe some basic properties of scaled measure entropy and lower and upper scaled measure entropies.
The following Proposition is a direct consequence of the definition of scaled measure entropy and Proposition 2.4.
Proposition 3.2 Let $\mathcal {U}$ and $\mathcal {V}$ be two open covers of X, $\mu \in \mathcal {M}(X,G)$ and $\mathbf {a}\in \mathcal {SS}$ , the following properties hold:
-
(1) If $\mathcal {U}\preceq \mathcal {V}$ , then $\mathcal {E}_{\mu }(\{F_{n}\},\mathcal {U},\mathbf {a})\leq \mathcal {E}_{\mu }(\{F_{n}\},\mathcal {V},\mathbf {a})$ ;
-
(2)
The following Proposition shows that the scaled measure entropy as well as lower and supper scaled measure entropies for amenable group actions are invariant under a measure conjugacy.
Definition 3.1 [Reference Kerr and Li21, Definition 1.4]
Two probability measure preserving actions $G\curvearrowright X_{1}$ , $G\curvearrowright X_{2}$ of the same group are said to be measure conjugate if there are conull sets $X_{1}^{'}\subset X_{1}$ and $X_{2}^{'}\subset X_{2}$ with $GX_{1}^{'}\subset X_{1}$ and $GX_{2}^{'}\subset X_{2}$ and an equivariant measure isomorphism $\varphi :X_{1}^{'}\rightarrow X_{2}^{'}$ .
Proposition 3.3 Given two measure conjugate actions $G \curvearrowright X_1$ and $G \curvearrowright X_2$ . For any $\mathbf {a}\in \mathcal {SS}$ and $\mu \in \mathcal {M}(X,G)$ , we have
where $\varphi $ is the equivariant measure isomorphism and $\varphi _{\ast }\mu =\mu \circ \varphi ^{-1}$ .
3.3 Scaled measure entropy for equivalent sequences and equivalent F $\unicode{xf8} $ lner sequences
Following the discussion on scaled topological entropy for equivalent scaled sequence and equivalent F $\unicode{xf8} $ lner sequencesin Section 2.3, we introduce a similar notion of equivalence for the scaled measure entropy.
The following Proposition is a direct consequence of Proposition 2.7.
Proposition 3.4 Let $\mathbf {a},\mathbf {b}\in \mathcal {SS}$ , for every G-invariant measure $\mu \in \mathcal {M}(X,G)$ , the following properties hold:
-
(1) If $a(|F_{n}|)\leq b(|F_{n}|)$ for all sufficiently large $n\in {\mathbb N}$ , then $\mathcal {E}_{\mu }(\{F_{n}\},\mathcal {U},\mathbf {a})\geq \mathcal {E}_{\mu }(\{F_{n}\},\mathcal {U},\mathbf {b})$ and $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{\mu }(\{F_{n}\},\mathbf {b})$ ;
-
(2) For each $K> 0$ we have that
$$ \begin{align*} K\cdot\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},K\mathbf{a})=\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},\mathbf{a}),~K\cdot\mathcal{E}_{\mu}(\{F_{n}\},K\mathbf{a})=\mathcal{E}_{\mu}(\{F_{n}\},\mathbf{a}), \end{align*} $$
where $K\mathbf {a}=\{K\cdot a(|F_{n}|)\}$ ;
-
(3) If there exists a constant $C>0$ such that $\frac {1}{C}\cdot b(|F_{n}|)\leq a(|F_{n}|)\leq C\cdot b(|F_{n}|)$ for all sufficiently large n, then
$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},\mathbf{b})\leq\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},\mathbf{a})\leq C\cdot\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},\mathbf{b}) \end{align*} $$and$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{\mu}(\{F_{n}\},\mathbf{b})\leq\mathcal{E}_{\mu}(\{F_{n}\},\mathbf{a})\leq C\cdot\mathcal{E}_{\mu}(\{F_{n}\},\mathbf{b}). \end{align*} $$
Remark 3.1 By Statement $(1)$ of Proposition 3.4, we have that $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{\mu }(\{F_{n}\},\mathbf {b})$ whenever $[\mathbf {a}]\preceq [\mathbf {b}]$ and by Statement $(3)$ of Proposition 3.4, for each equivalence class $[\mathbf {a}]\in \mathcal {A}$ and for each $\mathbf {a_{1}},\mathbf {a_{2}}\in [\mathbf {a}]$ we have that $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{1}})=\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{2}})=0$ or $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{1}})=\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{2}})=\infty $ or both $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{1}})$ and $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a_{2}})$ are positive and finite.
Theorem 3.1 For every $\mu \in \mathcal {M}(X,G)$ , if there is $[\mathbf {a}]\in \mathcal {A}$ such that $E_{\mu }(\{F_{n}\},[\mathbf {a}])$ is positive and finite, then
Similar results holds for lower and upper scaled measure entropy.
Proof Suppose there is $[\mathbf {a}]\in \mathcal {A}$ such that $E_{\mu }(\{F_{n}\},[\mathbf {a}])$ is positive and finite. Then for each $[\mathbf {b}]\succeq [\mathbf {a}]$ ,
for arbitrary $\mathbf {a_{1}}=\{a_{1}(|F_{n}|)\}\in [\mathbf {a}]$ and $\mathbf {b_{1}}=\{b_{1}(|F_{n}|)\}\in [\mathbf {b}]$ . Let us fix such two scaled sequences $\mathbf {a_{1}}$ and $\mathbf {b_{1}}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $a_{1}(|F_{n}|)<\beta b_{1}(|F_{n}|)$ . By Proposition 3.4, we have that
i.e., $\beta E_{\mu }(\{F_{n}\},\mathbf {a_{1}})\geq E_{\mu }(\{F_{n}\},\mathbf {b_{1}})$ . Since $\beta $ is arbitrary, we conclude that
and hence
On the other hand, if $[\mathbf {b}]\preceq [\mathbf {a}]$ then
for arbitrary $\mathbf {a_{2}}=\{a_{2}(|F_{n}|)\}\in [\mathbf {a}]$ and $\mathbf {b_{2}}=\{b_{2}(|F_{n}|)\}\in [\mathbf {b}]$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $b_{2}(|F_{n}|)<\beta a_{2}(|F_{n}|)$ . It follows that $E_{\mu }(\{F_{n}\},\mathbf {b_{2}})>\frac {1}{\beta }E_{\mu }(\{F_{n}\},\mathbf {a_{2}})$ . Again since $\beta $ is arbitrary,
and hence
And the following proposition is a direct consequence of Proposition 2.8.
Proposition 3.5 Let $\{F_n\},\{Q_n\}\in \mathcal {SF}(G)$ . For $\mathbf {a}\in \mathcal {SS}$ and every G-invariant measure $\mu \in \mathcal {M}(X,G)$ , the following properties hold:
-
(1) If $a(\vert F_{n}\vert )\leq a(\vert Q_{n}\vert )$ for all sufficiently large $n\in {\mathbb N}$ , then $\mathcal {E}_{\mu }(\{F_{n}\},\mathcal {U},\mathbf {a})\geq \mathcal {E}_{\mu }(\{Q_{n}\},\mathcal {U},\mathbf {a})$ and $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{\mu }(\{Q_{n}\},\mathbf {a})$ ;
-
(2) If there exists a constant $C>0$ such that $\frac {1}{C}\cdot a(|Q_{n}|)\leq a(|F_{n}|)\leq C\cdot a(|Q_{n}|)$ for all sufficiently large $n\in {\mathbb N}$ , then
$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{\mu}(\{Q_{n}\},\mathcal{U},\mathbf{a})\leq\mathcal{E}_{\mu}(\{F_{n}\},\mathcal{U},\mathbf{a})\leq C\cdot\mathcal{E}_{\mu}(\{Q_{n}\},\mathcal{U},\mathbf{a}) \end{align*} $$and$$ \begin{align*} \frac{1}{C}\cdot\mathcal{E}_{\mu}(\{Q_{n}\},\mathbf{a})\leq\mathcal{E}_{\mu}(\{F_{n}\},\mathbf{a})\leq C\cdot\mathcal{E}_{\mu}(\{Q_{n}\},\mathbf{a}). \end{align*} $$
Remark 3.2 By Statement $(1)$ of Proposition 3.5, we have that $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})\geq \mathcal {E}_{\mu }(\{Q_{n}\},\mathbf {a})$ whenever $\Big [\{F_n\}\Big ]_{\mathbf {a}}\preceq \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ and by Statement $(2)$ of Proposition 3.5, for each equivalence class $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ and for each $\{F_n^*\},\{F_n^{**}\}\in \Big [\{F_n\}\Big ]$ we have that $\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})=\mathcal {E}_{\mu }(\{F_{n}\},\mathbf {a})=0$ or $\mathcal {E}_{\mu }(\{F_{n}^*\},\mathbf {a})=\mathcal {E}_{\mu }(\{F_{n}^{**}\},\mathbf {a})=\infty $ or both $\mathcal {E}_{\mu }(\{F_{n}^*\},\mathbf {a})$ and $\mathcal {E}_{\mu }(\{F_{n}^{**}\},\mathbf {a})$ are positive and finite.
Theorem 3.2 Let $\mu \in \mathcal {M}(X,G)$ and $\mathbf {a}\in \mathcal {SS}$ . If there is $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ such that $E_{\mu }\Big(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a}\Big)$ is positive and finite, then
Similar results hold for lower and upper scaled measure entropy.
Proof Suppose there is $\Big [\{F_n\}\Big ]_{\mathbf {a}}\in \mathcal {F}(G)_{\mathbf {a}}$ such that $E_{\mu }\Big(\Big [\{F_n\}\Big ]_{\mathbf {a}},\mathbf {a}\Big)$ is positive and finite. Then for each $\Big [\{Q_n\}\Big ]_{\mathbf {a}}\succeq \Big [\{F_n\}\Big ]_{\mathbf {a}}$ ,
for arbitrary $\{F_{n}^*\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ and $\{Q_{n}^*\}\in \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ . Let us fix such two F $\unicode{xf8} $ lner sequences $\{F_{n}^*\}$ and $\{Q_{n}^*\}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N,}$ we have that $a(|F_{n}^*|)<\beta a(|Q_{n}^*|)$ . By $(2)$ of Proposition 3.4, we have that
i.e., $\beta E_{\mu }(\{F_{n}^*\},\mathbf {a})\geq E_{\mu }(\{Q_{n}^*\},\mathbf {a})$ . Since $\beta $ is arbitrary, we conclude that
and hence
On the other hand, if $\Big [\{F_n\}\Big ]_{\mathbf {a}}\succeq \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ then
for arbitrary $\{F_{n}^{**}\}\in \Big [\{F_n\}\Big ]_{\mathbf {a}}$ and $\{Q_{n}^{**}\}\in \Big [\{Q_n\}\Big ]_{\mathbf {a}}$ . Given a small number $\beta>0$ , for all sufficiently large $n\in {\mathbb N}$ we have that $a(|Q_{n}^{**}|)<\beta a(|F_{n}^{**}|)$ . It follows that $E_{\mu }(\{Q_{n}^{**}\},\mathbf {a})>\frac {1}{\beta }E_{\mu }(\{F_{n}^{**}\},\mathbf {a})$ . Again since $\beta $ is arbitrary,
then
3.4 Scaled local entropy
In this Section, we introduce the scaled local entropy following the approach of Brin and Katok defined for amenable group actions as follows:
Definition 3.2 For any $\mathbf {a}\in \mathcal {SS}$ , $\epsilon>0$ and $\mu \in \mathcal {M}(X)$ , define
where
We call $\underline {h}_\mu (\{F_{n}\},\mathbf {a})$ , the lower and upper scaled local entropy of $(X,G)$ (with respect to the sequence $\mathbf {a}\in \mathcal {SS}$ , the F $\unicode{xf8} $ lner sequence $\{F_{n}\}$ and $\mu \in \mathcal {M}(X)$ ).
Remark 3.3 It is pointed out in [Reference Foreman and Weiss12] that Dan Rudolph showed that for an amenable group G, the generic measure-preserving action of G on a Lebesgue space has zero entropy. Indeed, this is extended to nonamenable groups by Lewis Bowen in [Reference Bowen3] in which the proof shows that every action is a factor of a zero entropy action. In this sense, for generic measure-preserving actions, if we can find certain sub-exponential scaled sequences, then the scaled measure entropy could be positive. Thus the scaled measure entropy is a possible candidate to classify generic measure-preserving actions. For more examples of scaled measure entropies for $\mathbb {Z}$ or $\mathbb {N}$ actions, we refer to [Reference Zhao and Pesin33, 4 Examples].
4 Variational principle
The notion of scaled weighted topological entropy is introduced, which is important to prove the variational principle.
4.1 Equivalence of $E^{B_{2}}_{Z}(\{F_{n}\},\mathbf {a})$ and $h_{top}^W(Z,\{F_{n}\},\mathbf {a})$ .
Lemma 4.1 [Reference Mattila23, Theorem 2.1]
Let $(X,d)$ be a compact metric space and $\mathcal {B}=\{B(x_i,r_i)\}_{i\in \mathcal {I}}$ be a family open of (or closed) balls in X. Then there exists a finite or countable subfamily $\mathcal {B}^{'}=\{B(x_i,r_i)\}_{i\in \mathcal {I}^{'}}$ of pairwise disjoint balls in $\mathcal {B}$ such that
Theorem 4.1 For any $\mathbf {a}\in \mathcal {SS}$ , Borel set $L\subset X$ , $\mu \in \mathcal {M}(X)$ and $s\geq 0$ , we have
-
(1) If $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x)\leq s$ for all $x\in L$ , then $E_{L}(\{F_{n}\},\mathbf {a})\leq s$ ;
-
(2) If $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x)\geq s$ for all $x\in L$ and $\mu (L)>0$ , then $E_{L}(\{F_{n}\},\mathbf {a})\geq s$ .
Proof $(1)$ For a fixed $r>0$ and $k\in \mathbb {N}$ , let
Then we have $L=\bigcup _{k=1}^{\infty }L_{k}$ , since $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x)\leq s$ for all $x\in L$ .
Now fix $k\geq 1$ and $0<\epsilon <\frac {1}{5k}$ . For each $x\in L_{k}$ , there exists a strictly increasing sequence $\{n_{j}\}_{j=1}^{\infty }$ (depending on the point x) such that
For any $N\in \mathbb {N}$ , the set $L_{k}$ is contained in the union of the sets in the family
By Lemma 4.1, there exists a finite or countable subfamily $\mathcal {B}=\{B_{F_{n_{i}}}(x_i,\epsilon )\}_{i\in \mathcal {I}}\subset \mathcal {F}$ of pairwise disjoint balls such that
The subfamily is at most countable since $\mu $ is a probability measure and the elements in $\mathcal {B}$ are pairwise disjoint and have positive $\mu $ -measure. Note that
The disjointness of $\{B_{F_{n_{i}}}(x_i,\epsilon )\}_{i\in \mathcal {I}}$ yields that
It follows that
Hence,
which implies that
Hence,
Since r can be arbitrary, this implies that $E_{L}(\{F_{n}\},\mathbf {a})\leq s.$
$(2)$ For a fixed $r>0$ and $k\in \mathbb {N}$ , let
Since $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x)\geq s$ for all $x\in L$ , we have that $L_{k}\subset L_{k+1}$ and $L=\bigcup _{k=1}^{\infty }L_{k}$ . Fix a sufficiently large $k\geq 1$ with $\mu (L_{k})>\frac {1}{2}\mu (L)>0$ . For each $N\in \mathbb {N}$ , set
It is easy to see that $L_{k,N}\subset L_{k,N+1}$ and $L_{k}=\bigcup _{N=1}^{\infty }L_{k,N}$ . Thus we can pick $N^{\ast }\geq 1$ such that $\mu (L_{k,N^{\ast }})>\frac {1}{2}\mu (L_{k})>0$ . For simplicity of notation, let $L^{\ast }=L_{k,N^{\ast }}$ and $\epsilon ^{\ast }=\frac {1}{k}$ . By the choice of $L^{\ast }$ , we have that
Fix a sufficiently large $N>N^{\ast }$ . For each cover $\mathcal {F}=\{B_{F_{n_{i}}}(y_{i},\frac {\epsilon }{2})\}_{i\geq 1}$ of $L^{\ast }$ with $0<\epsilon <\epsilon ^{\ast }$ and $n_{i}\geq N\geq N^{\ast }$ for each $i\geq 1$ . Without loss of generality, assume that $L^{\ast }\bigcap B_{F_{n_{i}}}(y_{i},\frac {\epsilon }{2})\neq \emptyset $ for all i. Thus, for each $i\geq 1$ pick a point $x_{i}\in L^{\ast }\bigcap B_{F_{n_{i}}}(y_{i},\frac {\epsilon }{2})$ so that
It follows that
Therefore,
Consequently,
which implies that $E_{L^{\ast }}(\{F_{n}\},\mathbf {a})\geq s-r.$
It follows that
Since r can be arbitrary, this implies that $E_{L}(\{F_{n}\},\mathbf {a})\geq s$ completing the proof of the theorem.
We need to emphasize that the following proof uses the similar idea as [Reference Feng and Huang11, Proposition 3.2].
Proposition 4.1 For any $\mathbf {a}\in \mathcal {SS}$ , $s\geq 0,\epsilon ,\delta>0$ and $Z\subset X$ . If $\kern1pt\liminf \limits _{n\rightarrow +\infty }\frac {a(|F_{n}|)}{n}>0$ , we have
for large enough $N\in \mathbb {N}$ , and then $E^{B_{2}}_{Z}(\{F_{n}\},\mathbf {a})=h_{top}^W(Z,\{F_{n}\},\mathbf {a})$ .
Proof Let $Z\subset X,s\geq 0,\epsilon ,\delta>0$ , set $f=\chi _ Z,c_i\equiv 1$ in the definition of scaled weighted topological entropy, we have
Next, we prove $M(Z,s+\delta ,N,\{F_{n}\},6\epsilon ,\mathbf {a})\leq W(Z,s,N,\epsilon ,\{F_{n}\},\mathbf {a})$ for large enough $N\in \mathbb {N}$ .
There are $\xi>0$ and $N_{1}$ such that $a(|F_{n}|)\geq n\xi $ for all $n\geq N_{1}$ . Let $N>\max \{N_{1},2\}$ such that $n^{2}e^{-n\delta \xi }\leq 1$ for all $n\geq N$ . Let $\{(B_{F_{n_{i}}}(x_i,\epsilon ),c_i)\}_{i\in \mathcal {I}}$ be a family so that $\mathcal {I}\subset \mathbb {N},x_i\in X,0\leq c_i<\infty ,n_i\geq N$ and
where $B_i:=B_{F_{n_{i}}}(x_i,\epsilon )$ . We claim that
and hence $M(Z,s+\delta ,N,\{F_{n}\},6\epsilon ,\mathbf {a})\leq W(Z,s,N,\epsilon ,\{F_{n}\},\mathbf {a})$ .
We denote
and
for $n\geq N,k\in \mathbb {N}.$ We write $B_i:=B_{F_{n_{i}}}(x_i,\epsilon ),5B_i:=B_{F_{n_{i}}}(x_i,5\epsilon )$ for $i\in \mathcal {I}$ . Obviously, we may assume $B_i\neq B_j$ for $i\neq j$ . For $t>0$ , set
and
We divide the proof of (4.1) into the following three steps.
Step 1. For each $n\geq N, k\in \mathbb {N}$ and $t>0$ , there exists a finite set $\mathcal {J}_{n,k,t}\subset \mathcal {I}_{n,k}$ such that the ball $B_i(i\in \mathcal {J}_{n,k,t})$ are pairwise disjoint, $Z_{n,t,k}\subset \cup _{i\in \mathcal {J}_{n,k,t}}5B_i$ , and
We will use the method of Federer [Reference Federer10, 2.10.24] and Mattila [Reference Mattila23, Lemma 8.16] for amenable group actions. Since $\mathcal {I}_{n,k}$ is finite, by approximating the $c_i$ ’s from above, we may assume that each $c_i$ is a positive rational, and then by multiplying with a common denominator we may assume that each $c_i$ is a positive integer. Let m be the least integer with $m\geq t$ . Denote $\mathcal {B}=\{B_i,i\in \mathcal {I}_{n,k}\}$ , and define ${u:\mathcal {B}\rightarrow \mathbb {Z}}$ , by $u(B_i)=c_i$ . Since $B_i\neq B_j$ for $i\neq j$ , so u is well defined. We define by introduction integer-valued functions $v_0,v_1,\cdots ,v_m$ on $\mathcal {B}$ and sub-families $\mathcal {B}_1,\mathcal {B}_{2},\ldots ,\mathcal {B}_m$ of $\mathcal {B}$ starting with $v_0=u$ . Using Lemma 4.1 repeatedly, we define inductively for $j=1,\cdots ,m$ , disjoint subfamilies $\mathcal {B}_{i}$ of $\mathcal {B}$ such that
and the functions $v_j$ such that
This is possible for $j<m$ ,
whence every $x\in Z_{n,k,t}$ belongs to some ball $B\in \mathcal {B}$ with $v_j(B)\geq 1$ . Thus
Choose $j_0\in \{1,\cdots ,m\}$ so that $\aleph (\mathcal {B}_{j_0})$ is the smallest. Then
So $\mathcal {J}_{n,k,t}=\{i\in \mathcal {I}:B_i\in \mathcal {B}_{j_0}\}$ is desired.
Step 2. For each $n\in \mathbb {N}$ and $t>0$ , we have
Assume $Z_{n,t}\neq \emptyset $ , otherwise (4.2) is obvious. Since $Z_{n,k,t}\uparrow Z_{n,t}$ , $Z_{n,k,t}\neq \emptyset $ for large enough $k\in {\mathbb N}$ . Let $\mathcal {J}_{n,k,t}$ be the sets constructed in Step 1. Then $\mathcal {J}_{n,k,t}\neq \emptyset $ for large enough $k\in {\mathbb N}$ . Set $E_{n,k,t}=\{x_i:i\in \mathcal {J}_{n,k,t}\}$ . Note that the family of all non-empty compact subsets of X is compact with respect to Hausdorff distance(Federer [Reference Federer10, 2.10.21]). It follows that there is a subsequence $(k_j)$ of natural numbers and a non-empty compact set $E_{n,t}\subset X$ such that $E_{n,k_j,t}$ converges to $E_{n,t}$ in the Hausdorff distance as $j\rightarrow \infty $ . Since any two points in $E_{n,k,t}$ have a distance (with respect to $d_{F_{n}}$ ) not less than $\epsilon $ , so do the points in $E_{n,t}$ . Thus $E_{n,t}$ is a finite set, moreover, $\aleph (E_{n,k_j,t})=\aleph (E_{n,t})$ when $j\in {\mathbb N}$ is large enough.
Hence
when $j\in {\mathbb N}$ is large enough, and thus $\bigcup _{x\in E_{n,t}}B_{F_{n}}(x,6\epsilon )\supset Z_{n,t}$ . By the way, since $\aleph {(E_{n,k_j,t})}=\aleph (E_{n,t})$ when $j\in {\mathbb N}$ is large enough, we have
Therefore,
Step 3. For any $t\in (0,1)$ , we have
which implies (4.1). In fact, fix $t\in (0,1)$ . Note that $\sum _{n=N}^{\infty }n^{-2}<1$ . Then $Z\subset \bigcup _{n=N}^\infty Z_{n,n^{-2}t}$ . Also note that $M(Z,s,N,\{F_{n}\},\epsilon ,\mathbf {a})$ is an outer measure of X, so we get
We will give a Frostman’s lemma in dynamical system, which is important to our proof.
Lemma 4.2 Suppose K is a non-empty compact subset of X. Let $s\geq 0,N\in \mathbb {N},\epsilon>0$ . If $c:=W(K,s,N,\epsilon ,\{F_{n}\},\mathbf {a})>0$ , then there exists a Borel probability measure $\mu $ on X such that $\mu (K)=1$ and
Proof Clearly $c<\infty $ . We define a function p on the space $C(X)$ of continuous real-valued functions on X by
Let $\mathbf {1}\in C(X)$ denote the constant function $\mathbf {1}(x)\equiv 1$ . It is easy to verify that:
-
(1) $p(tf)=tp(f)$ for $f\in C(X)$ and $t\geq 0$ ,
-
(2) $p(f+g)\leq p(f)+p(g)$ for $f,g\in C(X)$ ,
-
(3) $p(\mathbf {1})=1,0\leq P(f)\leq \parallel f\parallel _{\infty }$ for $f\in C(X)$ , and $p(g)=0$ for $g\in C(X),g\leq 0.$
By the Hahn–Banach theorem, we can extend the linear functional ${t\rightarrow tp(1),t\in \mathbb {R}}$ , from the subspace of constant functions to a linear functional $L:C(X)\rightarrow \mathbb {R}$ satisfying
If $f\in C(X)$ with $f\geq 0$ , then $p(-f)=0$ and so $L(f)\geq 0$ . Hence we can use the Riesz representation theorem to find a Borel probability measure $\mu $ on X such that $L(f)=\int f d\mu $ for $f\in C(X)$ .
Next, we prove $\mu (K)=1$ . For any compact set $E\subset X\backslash K$ , by Urysohn lemma there exists $f\in C(X)$ such that $0\leq f\leq 1,f(x)=1$ for $x\in E$ and $f(x)=0$ for ${x\in K}$ . Then ${f\cdot \chi _K=0}$ and thus $p(f)=0$ . Hence $\mu (E)\leq L(f)\leq p(f)=0$ . This shows ${\mu (X\backslash K)=0,}$ that is $\mu (K)=1$ .
In the end, we prove $\mu (B_{F_n}(x,\epsilon ))\leq \frac {1}{c}\exp (-sa(|F_{n}|))$ for any $x\in X,n\geq N$ . In fact, for any compact set $E\subset B_{F_n}(x,\epsilon )$ , by Urysohn lemma again, there is $f\in C(X)$ , such that $0\leq f\leq 1, f(y)=1$ for $y\in E$ and $f(y)=0$ for $y\in X\backslash B_{F_n}(x,\epsilon )$ . Then $\mu (E)\leq L(f)\leq p(f)$ . Since $\chi _K\cdot f\leq \chi _{B_{F_n}(x,\epsilon )}$ and $n\geq N$ , we get $W(\chi _K\cdot f,s,N,\epsilon ,\{F_{n}\},\mathbf {a})\leq \exp (-sa(|F_{n}|))$ and hence $p(f)\leq \frac {1}{c}\exp (-sa(|F_{n}|))$ . Therefore, we have $\mu (E)\leq \frac {1}{c}\exp (-sa(|F_{n}|))$ . It follows that
Now we are in a position to present our main result: the variational principle between scaled topological entropy and scaled local entropy.
Theorem 4.2 Let $(X,G)$ be a topological dynamical system, $\mathbf {a}\in \mathcal {SS}$ and K be a non-empty compact subset of X. If $\liminf \limits _{n\rightarrow +\infty }\frac {a(|F_{n}|)}{n}>0$ , then
Proof Firstly, we prove $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})\geq \underline {h}_\mu (\{F_{n}\},\mathbf {a})$ , for any $\mu \in \mathcal {M}(X),\mu (K)=1$ . We set
for $x\in X,n\in \mathbb {N},\epsilon>0$ . It’s easy to see that $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x,\epsilon )$ is nonnegative and increases as $\epsilon $ decreases. By the monotone convergence theorem, we get
Thus to show $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})\geq \underline {h}_\mu (\{F_{n}\},\mathbf {a})$ , we only to show
Now we fix $\epsilon>0, l\in \mathbb {N}$ , set $u_{l}=\min \{l,\int \underline {h}_\mu (\{F_{n}\},\mathbf {a},x,\epsilon )d\mu (x)-\frac {1}{l}\}$ , then exist a Borel set $A_{l}\subset X,\mu (A_{l})>0,N\in \mathbb {N}$ such that
Let $\{B_{F_{n_{i}}}(x_i,\epsilon /2)\}$ ba a finite or countable family such that $x_i\in X,n_i\geq N$ and $K\cap A_{l}\subset \bigcup _iB_{F_{n_{i}}}(x_i,\epsilon /2)$ . We may as well assume that for each $i\in {\mathbb N}$ , $B_{F_{n_{i}}}(x_i,\epsilon /2)\bigcap (K\cap A_{l})\neq \emptyset $ , and select $y_i\in B_{F_{n_{i}}}(x_i,\epsilon /2)\bigcap (K\cap A_{l})$ . Then by (4.3), we have
So, we get
Therefore, $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})\geq u_{l}$ . Letting $l\rightarrow \infty $ , we get
Thus $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})\geq \underline {h}_\mu (\{F_{n}\},\mathbf {a})$ .
We next prove $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})\leq \{\underline {h}_\mu (\{F_{n}\},\mathbf {a}):\mu \in \mathcal {M}(X),\mu (K)=1\}$ . We may as well assume $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})>0$ , otherwise the conclusion is obvious. By Proposition 4.1, $E_{K}^{B_{2}}(\{F_{n}\},\mathbf {a})=h_{top}^W(K,\{F_{n}\},\mathbf {a})$ . Suppose $0<s<h_{top}^W(K,\{F_{n}\},\mathbf {a})$ , then there exists $\epsilon>0$ and $N\in \mathbb {N}$ , such that $c=W(K,s,N,\epsilon ,\{F_{n}\},\mathbf {a})>0$ . By Lemma 4.2, there exists $\mu \in \mathcal {M}(X),\mu (K)=1$ , such that
for any $x\in X,n\geq N$ . And then $\underline {h}_\mu (\{F_{n}\},\mathbf {a},x)\geq s$ for each $x\in X$ . Therefore, $\underline {h}_\mu (\{F_{n}\},\mathbf {a})\geq \int \underline {h}_\mu (\{F_{n}\},\mathbf {a},x)d\mu (x)\geq s.$ By Proposition 2.2, the proof is completed.
Funding
This work is supported by National Natural Science Foundation of China (Grant No. 12271432) and Xi'an International Science and Technology Cooperation Base-Ergodic Theory and Dynamical Systems.