1 Introduction
Throughout, let X denote a compact metric space endowed with a metric d. We denote by $\mathcal {C}(X)$ (respectively, $\mathcal {H}(X)$ ) the set of continuous self-maps (respectively, homeomorphisms) of X. Let $d_{C^0}\colon \mathcal {C}(X)\times \mathcal {C}(X)\to [0,\infty )$ be the metric defined by
for $f,g\in \mathcal {C}(X)$ . A metric $\hat {d}_{C^0}\colon \mathcal {H}(X)\times \mathcal {H}(X)\to [0,\infty )$ is given by
for $f,g\in \mathcal {H}(X)$ . With respect to these metrics, $\mathcal {C}(X)$ and $\mathcal {H}(X)$ are complete metric spaces.
A subset S of X is called a Mycielski set if it is a countable union of Cantor sets. Define $\mathcal {C}_{\mathrm {DC1}_{\ast }}(X)$ to be the set of $f\in \mathcal {C}(X)$ such that there is a Mycielski subset S of X, which is distributionally n- $\delta _n$ -scrambled for all $n\ge 2$ for some $\delta _n>0$ . Let
We say that a subset F of a complete metric space Z is residual if it contains a countable intersection of open and dense subsets of Z. The aim of this paper is to prove the following theorem.
Theorem 1.1. Given any compact smooth manifold M without boundary, $\mathcal {C}_{\mathrm {DC1}_{\ast }}(M)$ is a residual subset of $\mathcal {C}(M)$ , and if $\dim {M}>1$ , then $\mathcal {H}_{\mathrm {DC1}_{\ast }}(M)$ is also a residual subset of $\mathcal {H}(M)$ .
We recall the definition of distributional n-chaos [Reference Li and Oprocha22, Reference Tan and Fu37].
Definition 1.1. For $f\in \mathcal {C}(X)$ , an n-tuple $(x_1,x_2,\ldots ,x_n)\in X^n$ , $n\ge 2$ , is said to be distributionally n- $\delta $ -scrambled for $\delta>0$ if
for all $\epsilon>0$ , and
Let $\mathrm {DC1}_n^{\delta }(X,f)$ denote the set of distributionally n- $\delta $ -scrambled n-tuples and let $\mathrm {DC1}_n(X,f)=\bigcup _{\delta>0}\mathrm {DC1}_n^{\delta }(X,f)$ . A subset S of X is said to be distributionally n-scrambled (respectively, n- $\delta $ -scrambled) if
for any distinct $x_1,x_2,\ldots ,x_n\in S$ . We say that f exhibits the distributional n-chaos of type 1 ( $\mathrm {DC1}_n$ ) if there is an uncountable distributionally n-scrambled subset of X.
For any map $f\in \mathcal {C}(X)$ , let $h_{\mathrm {top}}(f)$ denote the topological entropy of f (see, for example, [Reference Walters38] for its definition). Then, for every compact topological manifold M, possibly with boundary, Yano showed that generic $f\in \mathcal {C}(M)$ (respectively, $f\in \mathcal {H}(M)$ , if $\dim {M}>1$ ) satisfies $h_{\mathrm {top}}(f)=\infty $ [Reference Yano40]. Since the positive topological entropy is a characteristic feature of chaos, generic dynamics on M is, roughly speaking, chaotic. Another definition of chaos is the so-called Li–Yorke chaos derived from [Reference Li and Yorke23]; and in [Reference Blanchard, Glasner, Kolyada and Maass6], any map $f\in \mathcal {C}(X)$ is proved to be Li–Yorke chaotic whenever $h_{\mathrm {top}}(f)>0$ (see [Reference Blanchard, Glasner, Kolyada and Maass6, Corollary 2.4]), so Li–Yorke chaos is topologically generic on manifolds. From a statistical viewpoint, the notion of distributional chaos was introduced by Schweizer and Smítal in [Reference Schweizer and Smítal35] as three variants of Li–Yorke chaos for interval maps. They are numbered in order of decreasing strength ( $\mathrm {DC}\beta _2$ , $\beta \in \{1,2,3\}$ ), therefore $\mathrm {DC1}_2$ is the strongest by definition, and $\mathrm {DC2}_2$ (also called mean Li–Yorke chaos) is still stronger than Li–Yorke chaos. Then it is natural to ask if $\mathrm {DC1}_n$ , $n\ge 2$ , is generic or not. Note that $\mathrm {DC1}_n$ , $n\ge 2$ , does not necessarily imply $\mathrm {DC1}_{n+1}$ [Reference Li and Oprocha22, Reference Tan and Fu37].
For an interval map $f\in \mathcal {C}([0,1])$ , all $\mathrm {DC}\beta _2$ , $\beta \in \{1,2,3\}$ , are equivalent to $h_{\mathrm {top}}(f)>0$ [Reference Schweizer and Smítal35] (see also [Reference Ruette34]). Since there is a Li–Yorke chaotic map $f\in \mathcal {C}([0,1])$ with $h_{\mathrm {top}}(f)=0$ , $\mathrm {DC2}_2$ is strictly stronger than Li–Yorke chaos in general [Reference Smítal36, Reference Xiong39] (see also [Reference Ruette34]). In general, improving the result of [Reference Blanchard, Glasner, Kolyada and Maass6], Downarowicz showed that any map $f\in \mathcal {C}(X)$ with $h_{\mathrm {top}}(f)>0$ exhibits $\mathrm {DC2}_2$ [Reference Downarowicz, Kolyada, Manin and Ward8] (see also [Reference Huang, Li and Ye13]). On the other hand, Pikuła constructed a subshift $(X,f)$ such that $h_{\mathrm {top}}(f)>0$ and $\mathrm {DC1}_2(X,f)=\emptyset $ [Reference Pikuła31]. Since $\mathrm {DC1}_2(X,f)\ne \emptyset $ implies the existence of a distal pair for f (see [Reference Oprocha29, Corollary 15]), any proximal map $f\in \mathcal {C}(X)$ with $h_{\mathrm {top}}(f)>0$ , given in, for example, [Reference Huang, Li and Ye12, Reference Kwietniak19, Reference Oprocha30], does not exhibit $\mathrm {DC1}_2$ . By [Reference Balibrea and Smítal4, Theorem 2], we also know that a minimal map $f\in \mathcal {C}(X)$ with a regularly recurrent point satisfies $\mathrm {DC1}_2(X,f)=\emptyset $ , so every Toeplitz subshift with arbitrary topological entropy does not exhibit $\mathrm {DC1}_2$ (see also [Reference Downarowicz7] and [Reference Kawaguchi15, Remark 2.2]). Thus, some additional assumptions besides $h_{\mathrm {top}}(f)>0$ are needed to ensure $\mathrm {DC1}_n$ , $n\ge 2$ , for a general map $f\in \mathcal {C}(X)$ .
Shadowing is a natural candidate for such an assumption. In [Reference Li, Li and Tu21], Li, Li and Tu proved that for any transitive map $f\in \mathcal {C}(X)$ with the shadowing property, $f\in \mathcal {C}_{\mathrm {DC1}_{\ast }}(X)$ if one of the following properties holds: $(1) \ f$ is non-periodic and has a periodic point; or $(2) \ f$ is non-trivial weakly mixing. Here, we have $h_{\mathrm {top}}(f)>0$ in both cases. This result is extended in [Reference Kawaguchi15] by using a relation defined by Richeson and Wiseman [Reference Richeson and Wiseman33] (see also [Reference Kawaguchi14]). Note that it was previously known that shadowing with (chain) mixing implies the specification and so $\mathrm {DC1}_2$ , except for the degenerate case (see [Reference Kwietniak, Łącka, Oprocha, Kolyada, Möller, Moree and Ward20, Reference Mazur and Oprocha25, Reference Oprocha28]). In [Reference Guihéneuf and Lefeuvre11], for every compact topological manifold M with $\dim {M}>1$ , Guihéneuf and Lefeuvre proved that generic $f\in \mathcal {H}_\mu (M)$ satisfies the shadowing property, where $\mathcal {H}_\mu (M)$ is the set of $f\in \mathcal {H}(M)$ preserving a non-atomic Borel probability measure $\mu $ on M with the full support and $\mu (\partial M)=0$ . Since such $f\in \mathcal {H}_{\mu }(M)$ is also (chain) mixing, $\mathrm {DC1}_n$ , $n\ge 2$ , is generic for the conservative homeomorphisms. As for $\mathcal {H}(M)$ , the situation is very different because, in particular, generic $f\in \mathcal {H}(M)$ has no isolated chain component, at least for any smooth closed M [Reference Akin, Hurley and Kennedy2]. Nevertheless, it is still useful to focus on the chain components. Since shadowing is shown to be generic in $\mathcal {C}(M)$ and $\mathcal {H}(M)$ [Reference Mazur and Oprocha24, Reference Pilyugin and Plamenevskaya32], any consequence of shadowing is a generic property in $\mathcal {C}(M)$ and $\mathcal {H}(M)$ . When we consider the result of Li, Li and Tu, one of the obvious difficulties in proving $\mathrm {DC1}_n$ , $n\ge 2$ , for generic $f\in \mathcal {C}(M)$ (or $f\in \mathcal {H}(M)$ , if $\dim {M}>1$ ) is that even if f has shadowing, its restriction $f|_{C}$ to a chain component C for f does not necessarily have the shadowing property. Another difficulty arises from the additional assumption such as $(1)$ or $(2)$ above. The main results of this paper resolve these difficulties and establish the genericity of $\mathrm {DC1}_n$ , $n\ge 2$ . Note that for generic $f\in \mathcal {C}(M)$ (respectively, $f\in \mathcal {H}(M)$ ), the chain recurrent set $\operatorname {CR}(f)$ is known to be zero-dimensional, or equivalently, totally disconnected [Reference Akin, Hurley and Kennedy2, Reference Krupski, Omiljanowski and Ungeheuer18].
In outline, the proof of Theorem 1.1 goes as follows. In [Reference Good and Meddaugh10], Good and Meddaugh found and investigated a fundamental relationship between subshifts of finite type (SFTs) and shadowing. The following two lemmas are from [Reference Good and Meddaugh10].
Lemma 1.1. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps and let $(X,f)=\lim _\pi (X_n,f_n)$ . If $f_n\colon X_n\to X_n$ has the shadowing property for each $n\ge 1$ , and $\pi $ satisfies the Mittag-Leffler condition $\mathrm{(MLC)}$ , then f has the shadowing property.
Lemma 1.2. Let $f\colon X\to X$ be a continuous map with the shadowing property. If $\dim {X}=0$ , then there is an inverse sequence of equivariant maps
such that the following properties hold:
-
(1) $\pi $ satisfies the $\mathrm{MLC}$ ;
-
(2) $(X_n,f_n)$ is an SFT for each $n\ge 1$ ;
-
(3) $(X,f)$ is topologically conjugate to $\lim _{\pi }(X_n,f_n)$ .
Note that these results concern the so-called Mittag-Leffler condition of an inverse sequence of equivariant maps. Most of this paper is devoted to a study of the MLC focusing on the structure of chain components. By using the above lemmas and a method in [Reference Kawaguchi15] with Mycielski’s theorem, we prove the following lemma. Here, $\mathcal {D}(f)$ is the partition of X with respect to the equivalence relation $\sim _f$ defined by Richeson and Wiseman (see §2.2 for details).
Lemma 1.3. Let $f\colon X\to X$ be a transitive continuous map with the shadowing property such that $h_{\mathrm {top}}(f)>0$ . If there are a compact metric space Y such that $\dim {Y}=0$ , a transitive continuous map $g\colon Y\to Y$ with the shadowing property, and a factor map
then there exists a sequence of positive numbers $(\delta _n)_{n\ge 2}$ such that every $D\in \mathcal {D}(f)$ contains a dense Mycielski subset S which consists of transitive points for f and is distributionally n- $\delta _n$ -scrambled for all $n\ge 2$ .
A question in [Reference Li, Li and Tu21] asks whether or not, for any continuous map $f\colon X\to X$ , the shadowing property, transitivity, and positive topological entropy imply $\mathrm {DC1}_n$ , $n\ge 2$ . Lemma 1.3 gives a partial answer to this question. As a direct consequence, we obtain the following corollary, which answers the question in the zero-dimensional case.
Corollary 1.1. Let $f\colon X\to X$ be a transitive continuous map with the shadowing property. If $\dim {X}=0$ and $h_{\mathrm {top}}(f)>0$ , then there exists a sequence of positive numbers $(\delta _n)_{n\ge 2}$ such that every $D\in \mathcal {D}(f)$ contains a dense Mycielski subset S which consists of transitive points for f and is distributionally n- $\delta _n$ -scrambled for all $n\ge 2$ .
Then, by dropping the transitivity assumption through Lemma 4.1, we obtain the following theorem. Here, $\mathcal {C}(f)$ is the set of chain components for f (see §2.2 for the definition).
Theorem 1.2. Given any continuous map $f\colon X\to X$ with the shadowing property, if $\dim {X}=0$ and $h_{\mathrm {top}}(f)>0$ , then there exist $C\in \mathcal {C}(f)$ and a sequence of positive numbers $(\delta _n)_{n\ge 2}$ such that every $D\in \mathcal {D}(f|_C)$ contains a dense Mycielski subset S which consists of transitive points for $f|_C$ and is distributionally n- $\delta _n$ -scrambled for all $n\ge 2$ .
Let
-
• $\mathcal {C}_{\mathrm {sh}}(X)=\{f\in \mathcal {C}(X)\colon f \text { has the shadowing property}\}$ ,
-
• $\mathcal {C}_{\mathrm{cr}^0}(X)=\{f\in \mathcal {C}(X)\colon \dim \operatorname {CR}(f) = 0 \}$ ,
-
• $\mathcal {C}_{h>0}(X)=\{f\in \mathcal {C}(X)\colon h_{\mathrm {top}}(f)>0 \}$ ,
and let $\mathcal {H}_\sigma (X)=\mathcal {H}(X)\cap \mathcal {C}_\sigma (X)$ for $\sigma \in \{\mathrm{sh},\mathrm{cr}^0, {h>0}\}$ . Note that for any
the restriction $f|_{\operatorname {CR}(f)}\colon \operatorname {CR}(f)\to \operatorname {CR}(f)$ has the following properties:
-
• the shadowing property;
-
• $\dim {\operatorname {CR}(f)}=0$ ;
-
• $h_{\mathrm {top}}(f|_{\operatorname {CR}(f)})=h_{\mathrm {top}}(f)>0$ .
By applying Theorem 1.2 to $f|_{\operatorname {CR}(f)}$ , we obtain $f\in \mathcal {C}_{\mathrm {DC1}_{\ast }}(X)$ ; therefore,
and so
Let M be a compact smooth manifold without boundary. Then Theorem 1.1 follows from the previous claims and the following results in the literature:
-
• shadowing
-
– $\mathcal {C}_{\mathrm {sh}}(M)$ is a residual subset of $\mathcal {C}(M)$ [Reference Mazur and Oprocha24],
-
– $\mathcal {H}_{\mathrm {sh}}(M)$ is a residual subset of $\mathcal {H}(M)$ [Reference Pilyugin and Plamenevskaya32];
-
-
• chain recurrence
-
– $\mathcal {C}_{\mathrm{cr}^0}(M)$ is a residual subset of $\mathcal {C}(M)$ [Reference Akin, Hurley and Kennedy2, Reference Krupski, Omiljanowski and Ungeheuer18],
-
– $\mathcal {H}_{\mathrm{cr}^0}(M)$ is a residual subset of $\mathcal {H}(M)$ [Reference Akin, Hurley and Kennedy2];
-
-
• topological entropy
-
– $\mathcal {C}_{h>0}(M)$ is a residual subset of $\mathcal {C}(M)$ [Reference Yano40],
-
– if $\dim {M}>1$ , then $\mathcal {H}_{h>0}(M)$ is a residual subset of $\mathcal {H}(M)$ [Reference Yano40].
-
The proof also gives an insight into how the distributionally scrambled sets exist in the chain recurrent set. As shown in the proof of Lemma 4.1 in §4, any chain component with positive topological entropy is approximated by one also with the shadowing property. The latter component is partitioned into the equivalence classes of the relation by Richeson and Wiseman. Then Corollary 1.1 implies that all equivalence classes densely contain distributionally scrambled Mycielski sets within them. It deepens the understanding about the chaotic aspect of $C^0$ -generic dynamics on manifolds.
We remark that for generic $f\in \mathcal {H}(M)$ with $\dim {M}>1$ , the set of $x\in \operatorname {CR}(f)$ which is contained in some $C\in \mathcal {C}(f)$ , such that $(C,f|_C)$ has a non-trivial subshift of finite type as a factor, is dense in $\operatorname {CR}(f)$ [Reference Akin, Hurley and Kennedy2]; therefore,
is a dense subset of $\operatorname {CR}(f)$ . We also recall from [Reference Akin, Hurley and Kennedy2] that for generic $f\in \mathcal {H}(M)$ with $\dim {M}>1$ , the set of $x\in \operatorname {CR}(f)$ which lies in some $C\in \mathcal {C}(f)$ such that C is initial or terminal, implying, if $\dim {\operatorname {CR}(f)}=0$ , that C is a periodic orbit or $(C,f|_C)$ is topologically conjugate to an odometer, is a residual subset of $\operatorname {CR}(f)$ . Thus, for such $f\in \mathcal {H}(M)$ with shadowing, the distributionally scrambled Mycielski sets should be contained in intermediate chain components, and the distributional chaos occurs in a dense but meager subset of $\operatorname {CR}(f)$ .
Lastly, Theorem 1.2 also provides a method to prove the genericity of $\mathrm {DC1}_n$ , $n\ge 2$ , for continuous self-maps or homeomorphisms of various underlying spaces which are not necessarily manifolds. We can find many results on the genericity of shadowing, zero-dimensionality of the chain recurrent set, and positive topological entropy in the context of topological dynamics (see, for example, [Reference Glasner and Weiss9, Reference Kościelniak, Mazur, Oprocha and Kubica17, Reference Krupski, Omiljanowski and Ungeheuer18]). Here, let us mention only the case where X is the Cantor set. In this case, it is shown that $\mathcal {H}(X)$ has a residual conjugacy class [Reference Akin, Glasner and Weiss1, Reference Kechris and Rosendal16]. Then generic $f\in \mathcal {H}(X)$ has the shadowing property but satisfies $h_{\mathrm {top}}(f)=0$ and has no Li–Yorke pair [Reference Bernardes and Darji5, Reference Glasner and Weiss9]; therefore, the generic homeomorphisms of X are not chaotic.
This paper consists of six sections. The basic notation, definitions, and facts are briefly collected in §2. In §3 we prove some preparatory lemmas. In §4 we prove Lemma 4.1 to reduce Theorem 1.2 to Lemma 1.3. In §5 we prove Lemma 1.3. In §6, as a bi-product of the proof of Lemma 4.1, we answer a question by Moothathu [Reference Moothathu26] in the zero-dimensional case, and give a related counter-example showing that the chain components with the shadowing property can be relatively few.
2 Preliminaries
In this section we collect some basic definitions, notations, facts, and prove some lemmas which will be used in what follows.
2.1 Chains, cycles, pseudo-orbits and the shadowing property
Given a continuous map $f\colon X\to X$ , a finite sequence $(x_i)_{i=0}^{k}$ of points in X, where $k>0$ is a positive integer, is called a $\delta $ -chain of f if $d(f(x_i),x_{i+1})\le \delta $ for every $0\le i\le k-1$ . A $\delta $ -chain $(x_i)_{i=0}^{k}$ of f is said to be a $\delta $ -cycle of f if $x_0=x_k$ . Let $\xi =(x_i)_{i\ge 0}$ be a sequence of points in X. For $\delta>0$ , $\xi $ is called a $\delta $ -pseudo-orbit of f if $d(f(x_i),x_{i+1})\le \delta $ for all $i\ge 0$ . For $\epsilon>0$ , $\xi $ is said to be $\epsilon $ -shadowed by $x\in X$ if $d(f^i(x),x_i)\leq \epsilon $ for all $i\ge 0$ . We say that f has the shadowing property if, for any $\epsilon>0$ , there is $\delta>0$ such that every $\delta $ -pseudo-orbit of f is $\epsilon $ -shadowed by some point of X.
2.2 Chain components and a relation
2.2.1 Chain recurrence and chain transitivity
Given a continuous map $f\colon X\to X$ , a point $x\in X$ is called a chain recurrent point for f if, for any $\delta>0$ , there is a $\delta $ -cycle $(x_i)_{i=0}^{k}$ of f with $x_0=x_k=x$ . We denote by $\operatorname {CR}(f)$ the set of chain recurrent points for f. It is a closed f-invariant subset of X, and the restriction $f|_{\operatorname {CR}(f)}\colon \operatorname {CR}(f)\to \operatorname {CR}(f)$ satisfies $\operatorname {CR}(f|_{\operatorname {CR}(f)})=\operatorname {CR}(f)$ . It is known that if f has the shadowing property, then so does $f|_{\operatorname {CR}(f)}$ [Reference Moothathu26]. We call f chain recurrent if $X=\operatorname {CR}(f)$ . For any $x,y\in X$ and $\delta>0$ , the notation $x\rightarrow _{f,\delta } y$ means that there is a $\delta $ -chain $(x_i)_{i=0}^k$ of f with $x_0=x$ and $x_k=y$ . Then f is said to be chain transitive if $x\rightarrow _{f,\delta } y$ for any $x,y\in X$ and $\delta>0$ . We say that f is transitive if, for any two non-empty open subsets U, Vof X, there is $n>0$ such that $f^n(U)\cap V\ne \emptyset $ . If f is transitive, then f is chain transitive, and the converse holds when f has the shadowing property.
2.2.2 Chain components
For any continuous map $f\colon X\to X$ , $\operatorname {CR}(f)$ admits a decomposition with respect to a relation $\leftrightarrow _f$ in $\operatorname {CR}(f)^2=\operatorname {CR}(f)\times \operatorname {CR}(f)$ defined as follows: for any $x,y\in \operatorname {CR}(f)$ , $x\leftrightarrow _f y$ if and only if $x\rightarrow _{f,\delta } y$ and $y\rightarrow _{f,\delta } x$ for every $\delta>0$ . Note that $\leftrightarrow _f$ is a closed $(f\times f)$ -invariant equivalence relation in $\operatorname {CR}(f)^2$ . An equivalence class C of $\leftrightarrow _f$ is called a chain component for f. We denote by $\mathcal {C}(f)$ the set of chain components for f. Then the following properties hold.
-
(1) $\operatorname {CR}(f)=\bigsqcup _{C\in \mathcal {C}(f)}C$ , here $\bigsqcup $ denotes the disjoint union.
-
(2) Every $C\in \mathcal {C}(f)$ is a closed f-invariant subset of $\operatorname {CR}(f)$ .
-
(3) $f|_C\colon C\to C$ is chain transitive for all $C\in \mathcal {C}(f)$ .
Note that f is chain transitive if and only if f is chain recurrent and satisfies $\mathcal {C}(f)=\{X\}$ .
2.2.3 A relation
Let $f\colon X\to X$ be a chain transitive map. For $\delta>0$ and a $\delta $ -cycle $\gamma =(x_i)_{i=0}^k$ of f, k is called the length of $\gamma $ . Let $m=m(f,\delta )>0$ be the greatest common divisor of all the lengths of $\delta $ -cycles of f. We define a relation $\sim _{f,\delta }$ in $X^2$ as follows: for any $x,y\in X$ , $x\sim _{f,\delta }y$ if and only if there is a $\delta $ -chain $(x_i)_{i=0}^k$ of f with $x_0=x$ , $x_k=y$ , and $m|k$ . Then the following properties hold.
-
(1) $\sim _{f,\delta }$ is an open and closed $(f\times f)$ -invariant equivalence relation in $X^2$ .
-
(2) For any $x\in X$ and $n\ge 0$ , $x\sim _{f,\delta }f^{mn}(x)$ .
-
(3) There exists $N>0$ such that, for any $x,y\in X$ with $x\sim _{f,\delta }y$ and $n\ge N$ , there is a $\delta $ -chain $(x_i)_{i=0}^k$ of f with $x_0=x$ , $x_k=y$ and $k=mn$ .
Following [Reference Richeson and Wiseman33], define a relation $\sim _f$ in $X^2$ as follows: for any $x,y\in X$ , $x\sim _f y$ if and only if $x\sim _{f,\delta }y$ for every $\delta>0$ . This is a closed $(f\times f)$ -invariant equivalence relation in $X^2$ . We denote by $\mathcal {D}(f)$ the set of equivalence classes of $\sim _f$ . This gives a closed partition of X. A pair $(x,y)\in X^2$ is said to be chain proximal if, for any $\delta>0$ , there is a pair $((x_i)_{i=0}^k,(y_i)_{i=0}^k)$ of $\delta $ -chains of f such that $(x_0,y_0)=(x,y)$ and $x_k=y_k$ . As claimed in [Reference Richeson and Wiseman33, Remark 8], for any $(x,y)\in X^2$ , $(x,y)$ is chain proximal if and only if $x\sim _f y$ .
2.3 Inverse limit
2.3.1 Inverse limit spaces
Given an inverse sequence of continuous maps
where $(X_n)_{n\ge 1}$ is a sequence of compact metric spaces, define $\pi _n^m\colon X_m\to X_n$ by
for all $m\ge n\ge 1$ . Note that $\pi _n^l=\pi _n^m\circ \pi _m^l$ for any $l\ge m\ge n\ge 1$ . The inverse limit space $X=\lim _\pi X_n$ is defined by
which is a compact metric space.
For any $n\ge 1$ , note that $\pi _n^m(X_m)\supset \pi _n^{m+1}(X_{m+1})$ for every $m\ge n$ , and let
By compactness, we easily see that, for any $n\ge 1$ and $x\in X_n$ , $x\in \hat {X}_n$ if and only if there is a sequence
with $\pi _m^{m+1}(x_{m+1})=x_m$ for every $m\ge n$ . For each $n\ge 1$ , $\pi _n^{n+1}(\hat {X}_{n+1})=\hat {X}_n$ , that is, $\hat {\pi }_n^{n+1}=(\pi _n^{n+1})|_{\hat {X}_{n+1}}\colon \hat {X}_{n+1}\to \hat {X}_n$ is surjective. Let
and $\hat {X}=\lim _{\hat {\pi }}\hat {X}_n$ . Since any $x=(x_n)_{n\ge 1}\in X$ satisfies $x_n\in \hat {X}_n$ for every $n\ge 1$ , we see that the inclusion $i\colon \hat {X}\to X$ is a homeomorphism.
2.3.2 The Mittag-Leffler condition
Given $\pi =(\pi _n^{n+1}\colon X_{n+1}\to X_n)_{n\ge 1}$ , an inverse sequence of continuous maps, $\pi $ is said to satisfy the Mittag-Leffler Condition if, for any $n\ge 1$ , there is $N\ge n$ such that $\pi _n^N(X_N)=\pi _n^m(X_m)$ for all $m\ge N$ . We say that $\pi $ satisfies $\operatorname {MLC}(1)$ if
for any $n\ge 1$ .
Lemma 2.1. Let $\pi =(\pi _n^{n+1}\colon X_{n+1}\to X_n)_{n\ge 1}$ be an inverse sequence of continuous maps. Then the following properties are equivalent.
-
(1) $\pi $ satisfies $\operatorname {MLC}(1)$ .
-
(2) For any $n\ge 1$ and $m\ge n+1$ , $\pi _n^{n+1}(X_{n+1})=\pi _n^m(X_m)$ .
-
(3) For every $n\ge 1$ , $\hat {X}_n=\pi _n^{n+1}(X_{n+1})$ .
Proof. $(1)\Rightarrow (2)$ : We use an induction on m. For $m=n+1$ , $\pi _n^{n+1}(X_{n+1})=\pi _n^m(X_m)$ is trivially true. Assume $\pi _n^{n+1}(X_{n+1})=\pi _n^m(X_m)$ for some $m\ge n+1$ . Then we have
completing the induction.
$(2)\Rightarrow (1)$ : Put $m=n+2$ in (2).
$(2)\Rightarrow (3)$ : Property (2) implies
for every $n\ge 1$ .
$(3)\Rightarrow (2)$ : Since $\pi _n^{n+1}(X_{n+1})\supset \pi _n^{n+2}(X_{n+2})\supset \cdots \supset \hat {X}_n$ , $\hat {X}_n=\pi _n^{n+1}(X_{n+1})$ implies $\pi _n^m(X_m)=\hat {X}_n=\pi _n^{n+1}(X_{n+1})$ for any $m\ge n+1$ , completing the proof.
Remark 2.1. Property (2) in the above lemma implies that $\pi $ satisfies the MLC.
Lemma 2.2. Let $\pi =(\pi _n^{n+1}\colon X_{n+1}\to X_n)_{n\ge 1}$ be an inverse sequence of continuous maps. If $\pi $ satisfies the $\operatorname {MLC}$ , then there is a sequence $1\le n(1)<n(2)<\cdots $ such that, letting $\pi '=(\pi _{n(j)}^{n(j+1)}\colon X_{n(j+1)}\to X_{n(j)})_{j\ge 1}$ , $\pi '$ satisfies $\operatorname {MLC}(1)$ .
Proof. Put $n(0)=1$ . Inductively, define a sequence $1=n(0)<n(1)<n(2)<\cdots $ as follows: given $j\ge 0$ and $n(j)$ , take $n(j+1)>n(j)$ such that $\pi _{n(j)}^{n(j+1)}(X_{n(j+1)})=\pi _{n(j)}^m(X_m)$ for every $m\ge n(j+1)$ . Then, for each $j\ge 1$ , $\pi _{n(j)}^{n(j+1)}(X_{n(j+1)})=\pi _{n(j)}^{n(j+2)}(X_{n(j+2)})$ since $n(j+2)>n(j+1)$ , implying that $\pi '$ satisfies $\operatorname {MLC}(1)$ .
2.3.3 Equivariance, factor and the topological conjugacy
Given two continuous maps $f\colon X\to X$ and $g\colon Y\to Y$ , where X and Y are compact metric spaces, a continuous map $\pi \colon X\to Y$ is said to be equivariant if $g\circ \pi =\pi \circ f$ , and such $\pi $ is also denoted by $\pi \colon (X,f)\to (Y,g)$ . An equivariant map $\pi \colon (X,f)\to (Y,g)$ is called a factor map (respectively, topological conjugacy) if it is surjective (respectively, a homeomorphism). Two systems $(X,f)$ and $(Y,g)$ are said to be topologically conjugate if there is a topological conjugacy $h\colon (X,f)\to (Y,g)$ .
2.3.4 Inverse limit systems
For an inverse sequence of equivariant maps
the inverse limit system $(X,f)=\lim _\pi (X_n,f_n)$ is well defined by $X=\lim _\pi X_n$ , and $f(x)=(f_n(x_n))_{n\ge 1}$ for all $x=(x_n)_{n\ge 1}\in X$ .
For every $n\ge 1$ , note that $\hat {X}_n$ is a closed $f_n$ -invariant subset of $X_n$ , and let $\hat {f}_n=(f_n)|_{\hat {X}_n}\colon \hat {X}_n\to \hat {X}_n$ . For all $n\ge 1$ , $\hat {\pi }_n^{n+1}=(\pi _n^{n+1})|_{\hat {X}_{n+1}}\colon \hat {X}_{n+1}\to \hat {X}_n$ gives a factor map
Let
and $(\hat {X},\hat {f})=\lim _{\hat {\pi }}(\hat {X}_n,\hat {f}_n)$ . Then the inclusion $i\colon \hat {X}\to X$ is a topological conjugacy $i\colon (\hat {X},\hat {f})\to (X,f)$ .
Lemma 2.3. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps and let $(X,f)=\lim _\pi (X_n,f_n)$ . If $f_n\colon X_n\to X_n$ is chain recurrent (respectively, chain transitive) for each $n\ge 1$ , then $f\colon X\to X$ is chain recurrent (respectively, chain transitive).
Proof. Let $n\ge 1$ . Then, for any $m\ge n$ , since $f_m\colon X_m\to X_m$ is chain recurrent (respectively, chain transitive), and $\pi _n^m\colon (X_m,f_m)\to (X_n,f_n)$ is an equivariant map,
is chain recurrent (respectively, chain transitive). Because $\hat {X}_n=\bigcap _{m\ge n}\pi _n^m(X_m)$ ,
is chain recurrent (respectively, chain transitive). Since
is a sequence of factor maps, we see that $\hat {f}\colon \hat {X}\to \hat {X}$ is chain recurrent (respectively, chain transitive). Thus, $f\colon X\to X$ is chain recurrent (respectively, chain transitive), because $(X,f)$ and $(\hat {X},\hat {f})$ are topologically conjugate, completing the proof.
For an inverse sequence of equivariant maps
and a sequence $1\le n(1)<n(2)<\cdots $ , $\pi _{n(j)}^{n(j+1)}\colon X_{n(j+1)}\to X_{n(j)}$ is an equivariant map for each $j\ge 1$ . Letting
and $(Y,g)=\lim _{\pi '}(X_{n(j)},f_{n(j)})$ , we have a topological conjugacy $h\colon (X,f)\to (Y,g)$ given by $h(x)=(x_{n(j)})_{j\ge 1}$ for all $x=(x_n)_{n\ge 1}\in X$ . By this and Lemma 2.2, we obtain the following lemma.
Lemma 2.4. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps. If $\pi $ satisfies the $\operatorname {MLC}$ , then there is a sequence $1\le n(1)<n(2)<\cdots $ such that, letting $\pi '=(\pi _{n(j)}^{n(j+1)}\colon (X_{n(j+1)},f_{n(j+1)})\to (X_{n(j)},f_{n(j)}))_{j\ge 1}$ , $\pi '$ satisfies $\operatorname {MLC}(1)$ , and $\lim _\pi (X_n,f_n)$ is topologically conjugate to $\lim _{\pi '}(X_{n(j)},f_{n(j)})$ .
2.4 Subshifts
2.4.1 Subshifts of finite type
Let S be a finite set with the discrete topology. The shift map $\sigma \colon S^{\mathbb {N}}\to S^{\mathbb {N}}$ is defined by $\sigma (x)=(x_{n+1})_{n\ge 1}$ for all $x=(x_n)_{n\ge 1}\in S^{\mathbb {N}}$ . Note that $\sigma $ is continuous with respect to the product topology of $S^{\mathbb {N}}$ . The product space $S^{\mathbb {N}}$ (and also $(S^{\mathbb {N}},\sigma )$ ) is called the (one-sided) full-shift over S. A closed $\sigma $ -invariant subset X of $S^{\mathbb {N}}$ (and also the subsystem $(X,\sigma |_X)$ of $(S^{\mathbb {N}},\sigma )$ ) is called a subshift. A subshift X of $S^{\mathbb {N}}$ (and also $(X,\sigma |_X)$ of $(S^{\mathbb {N}},\sigma )$ ) is called a subshift of finite type if there are $N>0$ and $F\subset S^{N+1}$ such that, for any $x=(x_n)_{n\ge 1}\in S^{\mathbb {N}}$ , $x\in X$ if and only if $(x_i,x_{i+1},\ldots ,x_{i+N})\in F$ for all $i\ge 1$ . The shift map $\sigma \colon S^{\mathbb {N}}\to S^{\mathbb {N}}$ is positively expansive and has the shadowing property. We know that a subshift X of $S^{\mathbb {N}}$ is of finite type if and only if $\sigma |_X\colon X\to X$ has the shadowing property [Reference Aoki and Hiraide3].
2.4.2 Some properties of SFTs
Let $(X,\sigma |_X)$ be an SFT (of some full-shift over S) and put $f=\sigma |_X$ . Then f has the following properties.
-
(1) $\operatorname {CR}(f)=\overline {\operatorname {Per}(f)}$ , where $\operatorname {Per}(f)$ denotes the set of periodic points for f.
-
(2) For any $x\in X$ , there is $y\in \operatorname {CR}(f)$ such that $\lim _{n\to \infty }d(f^n(x),f^n(y))=0$ .
In fact, these two properties are consequences of the positive expansiveness and the shadowing property of $f\colon X\to X$ . Since the restriction $f|_{\operatorname {CR}(f)}\colon \operatorname {CR}(f)\to \operatorname {CR}(f)$ is surjective and positively expansive, it is c-expansive. Also, it has the shadowing property. Applying [Reference Aoki and Hiraide3, Theorem 3.4.4] to $f|_{\operatorname {CR}(f)}$ , we obtain the following property.
-
(3) There is a finite set $\mathcal {C}$ of clopen f-invariant subsets of $\operatorname {CR}(f)$ such that
$$ \begin{align*} \operatorname{CR}(f)=\bigsqcup_{C\in\mathcal{C}}C, \end{align*} $$and $f|_{C}\colon C\to C$ is transitive for every $C\in \mathcal {C}$ .
An element of $\mathcal {C}$ is called a basic set. We easily see that $\mathcal {C}=\mathcal {C}(f)$ , that is, the basic sets coincide with the chain components for f. For every $C\in \mathcal {C}$ , $f|_C$ has the shadowing property, so C (or $(C,f|_C)$ ) is a transitive SFT.
Consider the case where f is transitive (or $(X,f)$ is a transitive SFT). Then we have $X=\operatorname {CR}(f)$ and $\mathcal {C}=\mathcal {C}(f)=\{X\}$ . Again by [Reference Aoki and Hiraide3, Theorem 3.4.4], X admits a decomposition
where $m>0$ is a positive integer, such that $f^i(D)$ , $0\le i\le m-1$ , are clopen $f^m$ -invariant subsets of X, and
is mixing for every $0\le i\le m-1$ . Here, a continuous map $g\colon Y\to Y$ is said to be mixing if, for any two non-empty open subsets U, V of Y, there is $N>0$ such that $g^n(U)\cap V\ne \emptyset $ for all $n\ge N$ . In this case, we easily see that $\mathcal {D}(f)=\{f^i(D)\colon 0\le i\le m-1\}$ .
3 Preparatory lemmas
In this section we prove some preparatory lemmas needed for the proofs of the main results. The first two lemmas give an expression of the chain recurrent set (respectively, chain components) for the inverse limit system.
Lemma 3.1. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps and let $(X,f)=\lim _\pi (X_n,f_n)$ . Then
Proof. Let R denote the right-hand side of the equation. $\operatorname {CR}(f)\subset R$ is clearly true. Let us prove $R\subset \operatorname {CR}(f)$ . Note that $\pi _n^{n+1}(\operatorname {CR}(f_{n+1}))\subset \operatorname {CR}(f_n)$ for every $n\ge 1$ . Let $Y_n=\operatorname {CR}(f_n)$ , $g_n=(f_n)|_{Y_n}\colon Y_n\to Y_n$ , and $\tilde {\pi }_n^{n+1}=(\pi _n^{n+1})|_{Y_{n+1}}\colon Y_{n+1}\to Y_n$ for each $n\ge 1$ . Consider the inverse sequence of equivariant maps
and let $(Y,g)=\lim _{\tilde {\pi }}(Y_n,g_n)$ . Since $g_n$ is chain recurrent for all $n\ge 1$ , by Lemma 2.3, g is chain recurrent. On the other hand, R is a closed f-invariant subset of X and satisfies $Y=R$ . The inclusion $i\colon Y\to R$ gives a topological conjugacy $i\colon (Y,g)\to (R,f|_R)$ , and so $f|_R\colon R\to R$ is chain recurrent, which clearly implies $R\subset \operatorname {CR}(f)$ ; therefore, the lemma has been proved.
Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps and let $(X,f)=\lim _\pi (X_n,f_n)$ . Note that for any $n\ge 1$ and $C_{n+1}\in \mathcal {C}(f_{n+1})$ , there is $C_n\in \mathcal {C}(f_n)$ such that $\pi _n^{n+1}(C_{n+1})\subset C_n$ . Let
Also, for any $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ , let
The next lemma gives an expression of $\mathcal{C}(f)$ .
Lemma 3.2. It holds that $\mathcal {C}(f)=\{[C_{\ast }]\colon C_{\ast }\in \mathcal {C}_\pi \}$ , where $\mathcal{C}(f)$ is the set of chain components for f.
Proof. For any $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ , $[C_{\ast }]$ is a closed f-invariant subset of X. We prove that $f|_{[C_{\ast }]}\colon [C_{\ast }]\to [C_{\ast }]$ is chain transitive. For each $n\ge 1$ , let $g_n=f|_{C_n}\colon C_n\to C_n$ and let
Consider the inverse sequence of equivariant maps
and let $(Y,g)=\lim _{\tilde {\pi }}(C_n,g_n)$ . Since $g_n$ is chain transitive for all $n\ge 1$ , by Lemma 2.3, g is chain transitive. On the other hand, we have $Y=[C_{\ast }]$ . The inclusion $i\colon Y\to [C_{\ast }]$ gives a topological conjugacy $i\colon (Y,g)\to ([C_{\ast }],f|_{[C_{\ast }]})$ , which implies that $f|_{[C_{\ast }]}$ is chain transitive.
Given any $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ , from what is shown above, there is $C\in \mathcal {C}(f)$ such that $[C_{\ast }]\subset C$ . Fix $x=(x_n)_{n\ge 1}\in [C_{\ast }]$ . Then, for every $y=(y_n)_{n\ge 1}\in C$ , we easily see that $\{x_n,y_n\}\in C_n$ for all $n\ge 1$ ; therefore, $y\in [C_{\ast }]$ . This implies $C\subset [C_{\ast }]$ and so $[C_{\ast }]=C$ , proving
To prove
for any $C\in \mathcal {C}(f)$ , fix $x=(x_n)_{n\ge 1}\in C$ , and take $C_n\in \mathcal {C}(f_n)$ with $x_n\in C_n$ for each $n\ge 1$ . Then $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ and $x\in [C_{\ast }]\subset C$ . Similarly to the argument above, we obtain $C=[C_{\ast }]$ , completing the proof.
The next lemma gives an expression for $\mathcal {D}(f)$ , which was introduced in §2.2, for the inverse limit system under $\operatorname {MLC}(1)$ . Given
an inverse sequence of equivariant maps, let $(X,f)=\lim _\pi (X_n,f_n)$ and suppose that $f_n\colon X_n\to X_n$ is chain transitive for all $n\ge 1$ . Then, by Lemma 2.3, $f\colon X\to X$ is chain transitive. Note that, for any $n\ge 1$ and $D_{n+1}\in \mathcal {D}(f_{n+1})$ , there is $D_n\in \mathcal {D}(f_n)$ such that $\pi _n^{n+1}(D_{n+1})\subset D_n$ . Let
Also, for any $D_{\ast }=(D_n)_{n\ge 1}\in \mathcal {D}_\pi $ , let
Lemma 3.3. If $\pi $ satisfies $\operatorname {MLC}(1)$ , then $\mathcal {D}(f)=\{[D_{\ast }]\colon D_{\ast }\in \mathcal {D}_\pi \}$ .
Proof. Let $D_{\ast }=(D_n)_{n\ge 1}\in \mathcal {D}_\pi $ and let $x=(x_n)_{n\ge 1},y=(y_n)_{n\ge 1}\in [D_{\ast }]$ . We prove that $(x,y)\in X^2$ is chain proximal for f. Fix any $N>0$ and $\delta>0$ . Since $\{x_{N+1},y_{N+1}\}\subset D_{N+1}\in \mathcal {D}(f_{N+1})$ , $(x_{N+1},y_{N+1})\in X_{N+1}^2$ is chain proximal for $f_{N+1}$ , implying that there is a pair
of $\delta $ -chains of $f_{N+1}$ such that $(x_{N+1}^{(0)},y_{N+1}^{(0)})=(x_{N+1},y_{N+1})$ and $x_{N+1}^{(k)}=y_{N+1}^{(k)}$ . Let $(z^{(0)},w^{(0)})=(x,y)\in X^2$ and note that
for every $1\le n\le N$ . For each $0<i\le k$ , since
by $\operatorname {MLC}(1)$ (see Lemma 2.1), there are $(z^{(i)},w^{(i)})\in X^2$ , $0<i \le k-1$ , and $z^{(k)}=w^{(k)}\in X$ such that
and also
for every $1\le n\le N$ . Let $d_n$ , $n\ge 1$ , be the metric on $X_n$ . For any $0\le i\le k-1$ and $1\le n\le N$ , we have
with $d_{N+1}(f_{N+1}(x_{N+1}^{(i)}),x_{N+1}^{(i+1)})\le \delta $ , and similarly,
with $d_{N+1}(f_{N+1}(y_{N+1}^{(i)}),y_{N+1}^{(i+1)})\le \delta $ . Therefore, for every $\epsilon>0$ , if N is large enough, and then $\delta $ is sufficiently small,
is a pair of $\epsilon $ -chains of f with $(z^{(0)},w^{(0)})=(x,y)$ and $z^{(k)}=w^{(k)}$ , proving that $(x,y)\in X^2$ is chain proximal for f.
Given any $D_{\ast }=(D_n)_{n\ge 1}\in \mathcal {D}_\pi $ , from what is shown above, we have $[D_{\ast }]\subset D$ for some $D\in \mathcal {D}(f)$ . The rest of the proof is identical to that of Lemma 3.2.
The final lemma gives a sufficient condition for an inverse sequence of subsystems to continue satisfying condition $\operatorname {MLC}(1)$ .
Lemma 3.4. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ . Let $(X,f)=\lim _\pi (X_n,f_n)$ and suppose that a sequence of closed $f_n$ -invariant subsets $Y_n$ of $X_n$ , $n\ge 1$ , has the following properties:
-
(1) $\pi _n^{n+1}(Y_{n+1})\subset Y_n$ for every $n\ge 1$ ;
-
(2) any $x=(x_n)_{n\ge 1}\in X$ satisfies $x_n\in Y_n$ for all $n\ge 1$ .
For each $n\ge 1$ , let $g_n=(f_n)|_{Y_n}\colon Y_n\to Y_n$ and let $\tilde {\pi }_n^{n+1}=(\pi _n^{n+1})|_{Y_{n+1}}\colon Y_{n+1}\to Y_n$ . Then the inverse sequence of equivariant maps
satisfies $\operatorname {MLC}(1)$ .
Proof. For any $n\ge 1$ and $q\in X_{n+1}$ , since $\pi _n^{n+1}(q)\in \pi _n^{n+1}(X_{n+1})=\hat {X}_n$ by $\operatorname {MLC}(1)$ of $\pi $ (see Lemma 2.1), we have $x_n=\pi _n^{n+1}(q)$ for some $x=(x_j)_{j\ge 1}\in X$ . Since $\pi _n^{n+1}(q)=x_n=\pi _n^{n+2}(x_{n+2})$ , by property (2), we obtain $\pi _n^{n+1}(q)\in \pi _n^{n+2}(Y_{n+2})$ , implying $\pi _n^{n+1}(X_{n+1})\subset \pi _n^{n+2}(Y_{n+2})$ . Then
therefore $\pi _n^{n+1}(Y_{n+1})=\pi _n^{n+2}(Y_{n+2})$ . Since $n\ge 1$ is arbitrary, $\tilde {\pi }$ satisfies $\operatorname {MLC}(1)$ .
4 Reduction of Theorem 1.2 to Lemma 1.3
The aim of this section is to prove the following lemma to reduce Theorem 1.2 to Lemma 1.3.
Lemma 4.1. Let $f\colon X\to X$ be a continuous map with the shadowing property. If $\dim {X}=0$ and $h_{\mathrm {top}}(f)>0$ , then there is $C\in \mathcal {C}(f)$ such that $f|_{C}\colon C\to C$ has the shadowing property and satisfies $h_{\mathrm {top}}(f|_C)>0$ .
A lemma is needed for the proof. It states that, for an inverse sequence of SFTs, we can consider the inverse sequence of chain recurrent sets without losing $\operatorname {MLC}(1)$ .
Lemma 4.2. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ . Let $(X,f)=\lim _{\pi }(X_n,f_n)$ and suppose that $(X_n,f_n)$ is an SFT for each $n\ge 1$ . Let $Y_n=\operatorname {CR}(f_n)$ , $g_n=(f_n)|_{Y_n}\colon Y_n\to Y_n$ , and $\tilde {\pi }_n^{n+1}=(\pi _n^{n+1})|_{Y_{n+1}}\colon Y_{n+1}\to Y_n$ for every $n\ge 1$ . Then the inverse sequence of equivariant maps
satisfies $\operatorname {MLC}(1)$ .
Proof. Let $n\ge 1$ . Since $Y_{n+1}=\overline {\operatorname {Per}(f_{n+1})}$ and $\pi _n^{n+1}(\operatorname {Per}(f_{n+1}))\subset \operatorname {Per}(f_n)$ ,
By $\operatorname {MLC}(1)$ of $\pi $ , $\pi _n^{n+1}(Y_{n+1})\subset \pi _n^{n+1}(X_{n+1})=\pi _n^{n+2}(X_{n+2})$ ; therefore, for any $p\in \pi _n^{n+1}(Y_{n+1})\cap \operatorname {Per}(f_n)$ , there is $q\in X_{n+2}$ such that $p=\pi _n^{n+2}(q)$ . Then there is $r\in Y_{n+2}$ such that
implying
where $d_n$ , $d_{n+2}$ are the metrics on $X_n$ , $X_{n+2}$ . Note that $\pi _{n}^{n+2}(r)\in \pi _n^{n+2}(Y_{n+2})$ . From $p\in \operatorname {Per}(f_n)$ and the $f_n$ -invariance of $\pi _n^{n+2}(Y_{n+2})$ , it follows that $p\in \overline {\pi _n^{n+2}(Y_{n+2})}=\pi _n^{n+2}(Y_{n+2})$ . Since $p\in \pi _n^{n+1}(Y_{n+1})\cap \operatorname {Per}(f_n)$ is arbitrary, we obtain
and so
Thus, $\pi _n^{n+1}(Y_{n+1})=\pi _n^{n+2}(Y_{n+2})$ , proving the lemma.
We now prove Lemma 4.1. The proof is based on Lemma 1.2 and by carefully choosing an inverse sequence of chain components with $\operatorname {MLC}(1)$ .
Proof of Lemma 4.1
By Lemmas 1.2 and 2.4, we may assume $(X,f)=\lim _{\pi }(X_n,f_n)$ , where $(X_n,f_n)$ , $n\ge 1$ , are SFTs, and
is an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ . Let $Y_n=\operatorname {CR}(f_n)$ , $g_n=(f_n)|_{Y_n}\colon Y_n\to Y_n$ , and $\tilde {\pi }_n^{n+1}=(\pi _n^{n+1})|_{Y_{n+1}}\colon Y_{n+1}\to Y_n$ for every $n\ge 1$ . Then $(Y_n,g_n)$ , $n\ge 1$ , are chain recurrent SFTs, and by Lemma 4.2,
satisfies $\operatorname {MLC}(1)$ . Letting $(Y,g)=\lim _{\tilde {\pi }}(Y_n,g_n)$ and
we have $R=\operatorname {CR}(f)$ by Lemma 3.1, and as in the proof of Lemma 3.1, the inclusion $i\colon Y\to \operatorname {CR}(f)$ is a topological conjugacy $i\colon (Y,g)\to (\operatorname {CR}(f),f|_{\operatorname {CR}(f)})$ . By this, again without loss of generality, we may assume that f and $f_n$ , $n\ge 1$ , are chain recurrent.
Since $h_{\mathrm {top}}(f)>0$ , there is $C^\dagger \in \mathcal {C}(f)$ such that $h_{\mathrm {top}}(f|_{C^\dagger })>0$ . A proof of this fact is as follows. By the variational principle, there is an ergodic f-invariant Borel probability measure $\mu $ on X such that the measure-theoretical entropy $h_\mu (f)$ is positive. Since $f|_{\mathrm {supp}(\mu )}\colon \mathrm {supp}(\mu )\to \mathrm {supp}(\mu )$ , the restriction of f to the support of $\mu $ , is transitive, there is $C^\dagger \in \mathcal {C}(f)$ such that $\mathrm {supp}(\mu )\subset C^\dagger $ . By the variational principle again, we obtain
Then, by Lemma 3.2, there is $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ such that $C^\dagger =[C_{\ast }]$ . Letting $\Gamma =\prod _{n\ge 1}\pi _n^{n+1}(C_{n+1})$ , a closed F-invariant subset of $\prod _{n\ge 1}X_n$ where $F=\prod _{n\ge 1}f_n$ is the product map, since $C^\dagger \subset \Gamma $ , we have
implying $h_{\mathrm {top}}((f_n)|_{\pi _n^{n+1}(C_{n+1})})>0$ for some $n\ge 1$ .
Note that $\mathcal {C}(f_m)$ , $m\ge 1$ , are finite sets, and for any $m\ge 1$ and $D\in \mathcal {C}(f_m)$ , $(f_m)|_{D}\colon D\to D$ is transitive (see §2.4). Let us prove the following claim.
Claim. There is $C^{\prime }_{\ast }=(C^{\prime }_m)_{m\ge n}\in \prod _{m\ge n}\mathcal {C}(f_m)$ with the following properties:
-
(1) $C^{\prime }_n=C_n$ ;
-
(2) $\pi _n^{n+1}(C_{n+1})\subset \pi _n^{n+1}(C^{\prime }_{n+1})$ ;
-
(3) $\pi _m^{m+1}(C^{\prime }_{m+1})\subset C^{\prime }_m$ for every $m\ge n$ ;
-
(4) $\pi _m^{m+1}(C^{\prime }_{m+1})=\pi _m^{m+2}(C^{\prime }_{m+2})$ for all $m\ge n$ .
Proof of the claim
Step 1. Let $C^{\prime }_n=C_n$ . Take $D_{n+1}\in \mathcal {C}(f_{n+1})$ such that
and $\pi _n^{n+1}(D_{n+1})$ is maximal among
with respect to the inclusion relation.
Step 2. Note that $(f_n)|_{\pi _n^{n+1}(D_{n+1})}$ is transitive, and take a transitive point $p_1\in \pi _n^{n+1}(D_{n+1})$ , that is, $\pi _n^{n+1}(D_{n+1})=\omega (p_1,f_n)$ , the $\omega $ -limit set. Since
we have $p_1=\pi _n^{n+2}(q_1)$ for some $q_1\in X_{n+2}$ . Take $C^{\prime }_{n+1}\in \mathcal {C}(f_{n+1})$ with $\pi _{n+1}^{n+2}(q_1) \in C^{\prime }_{n+1}$ . Then choose $D_{n+2}\in \mathcal {C}(f_{n+2})$ such that
and $\pi _{n+1}^{n+2}(D_{n+2})$ is maximal among
with respect to the inclusion relation. By (P2), we have
implying
since $\pi _n^{n+1}(D_{n+1})=\omega (p_1,f_n)$ , and $\pi _n^{n+2}(D_{n+2})$ is $f_n$ -invariant. By (P1) and $p_1\in \pi _n^{n+1}(D_{n+1})$ , we see that $\pi _n^{n+1}(C_{n+1})\subset \pi _n^{n+1}(D_{n+1})$ and $p_1\in C^{\prime }_n\cap \pi _n^{n+1}(C^{\prime }_{n+1})$ ; therefore,
By the maximality of $\pi _n^{n+1}(D_{n+1})$ in Step 1, we obtain
Step 3. Note that $(f_{n+1})|_{\pi _{n+1}^{n+2}(D_{n+2})}$ is transitive, and take a transitive point $p_2\in \pi _{n+1}^{n+2}(D_{n+2})$ , that is, $\pi _{n+1}^{n+2}(D_{n+2})=\omega (p_2,f_{n+1})$ . Since
we have $p_2=\pi _{n+1}^{n+3}(q_2)$ for some $q_2\in X_{n+3}$ . Take $C^{\prime }_{n+2}\in \mathcal {C}(f_{n+2})$ with $\pi _{n+2}^{n+3}(q_2)\in C^{\prime }_{n+2}$ . Then choose $D_{n+3}\in \mathcal {C}(f_{n+3})$ such that
and $\pi _{n+2}^{n+3}(D_{n+3})$ is maximal among
with respect to the inclusion relation. By (P3), we have
implying
since $\pi _{n+1}^{n+2}(D_{n+2})=\omega (p_2,f_{n+1})$ , and $\pi _{n+1}^{n+3}(D_{n+3})$ is $f_{n+1}$ -invariant. By (P2) and $p_2\in \pi _{n+1}^{n+2}(D_{n+2})$ , we see that $\pi _{n+1}^{n+2}(q_1)\in \pi _{n+1}^{n+2}(D_{n+2})$ and $p_2\in C^{\prime }_{n+1}\cap \pi _{n+1}^{n+2}(C^{\prime }_{n+2})$ ; therefore,
By the maximality of $\pi _{n+1}^{n+2}(D_{n+2})$ in Step 2, we obtain
Assertions (Q1) and (Q2) yield $\pi _n^{n+1}(C^{\prime }_{n+1})=\pi _n^{n+2}(C^{\prime }_{n+2})$ .
Step 4. Note that $(f_{n+2})|_{\pi _{n+2}^{n+3}(D_{n+3})}$ is transitive, and take a transitive point $p_3\in \pi _{n+2}^{n+3}(D_{n+3})$ , that is, $\pi _{n+2}^{n+3}(D_{n+3})=\omega (p_3,f_{n+2})$ . Since
we have $p_3=\pi _{n+2}^{n+4}(q_3)$ for some $q_3\in X_{n+4}$ . Take $C^{\prime }_{n+3}\in \mathcal {C}(f_{n+3})$ with $\pi _{n+3}^{n+4}(q_3)\in C^{\prime }_{n+3}$ . Then choose $D_{n+4}\in \mathcal {C}(f_{n+4})$ such that
and $\pi _{n+3}^{n+4}(D_{n+4})$ is maximal among
with respect to the inclusion relation. By (P4), we have
implying
since $\pi _{n+2}^{n+3}(D_{n+3})=\omega (p_3,f_{n+2})$ , and $\pi _{n+2}^{n+4}(D_{n+4})$ is $f_{n+2}$ -invariant. By (P3) and $p_3\in \pi _{n+2}^{n+3}(D_{n+3})$ , we see that $\pi _{n+2}^{n+3}(q_2)\in \pi _{n+2}^{n+3}(D_{n+3})$ and $p_3\in C^{\prime }_{n+2}\cap \pi _{n+2}^{n+3}(C^{\prime }_{n+3})$ ; therefore,
By the maximality of $\pi _{n+2}^{n+3}(D_{n+3})$ in Step 3, we obtain
Assertions (Q2) and (Q3) yield $\pi _{n+1}^{n+2}(C^{\prime }_{n+2})=\pi _{n+1}^{n+3}(C^{\prime }_{n+3})$ .
Continuing inductively, we obtain a sequence $C^{\prime }_{\ast }=(C^{\prime }_m)_{m\ge n}\in \prod _{m\ge n}\mathcal {C}(f_m)$ . Then properties (1) and (2) are ensured in Steps 1 and 2. For any $k\ge 0$ , $\pi _{n+k}^{n+k+1}(C^{\prime }_{n+k+1})\subset C^{\prime }_{n+k}$ and $\pi _{n+k}^{n+k+1}(C^{\prime }_{n+k+1})=\pi _{n+k}^{n+k+2}(C^{\prime }_{n+k+2})$ are established in Steps $k+2$ and $k+3$ , respectively. Thus, $C^{\prime }_{\ast }$ satisfies the required properties, and so the claim has been proved.
We continue the proof of Lemma 4.1. Define $C^{\prime \prime }_{\ast }=(C^{\prime \prime }_j)_{j\ge 1}\in \prod _{j\ge 1}\mathcal {C}(f_j)$ by
By properties (1) and (3) in the claim, we see that $C^{\prime \prime }_{\ast }\in \mathcal {C}_\pi $ . By Lemma 3.2, letting $C=[C^{\prime \prime }_{\ast }]$ , we obtain $C\in \mathcal {C}(f)$ . Let
and
Then $(C,f|_C)$ (respectively, $\lim _{\pi ''}(C^{\prime \prime }_j,(f_j)|_{C^{\prime \prime }_j})$ ) is topologically conjugate to
(respectively, $\lim _{\pi '}(C^{\prime }_m,(f_m)|_{C^{\prime }_m})$ ), so $(C,f|_C)$ is topologically conjugate to
Let $(Y,g)=\lim _{\pi '}(C^{\prime }_m,(f_m)|_{C^{\prime }_m})$ . By property (4) in the claim, $\pi '$ satisfies $\operatorname {MLC}(1)$ . Since $(f_m)|_{C^{\prime }_m}$ has the shadowing property for each $m\ge 1$ , due to Lemma 1.1, g has the shadowing property. Let us prove $h_{\mathrm {top}}(g)>0$ . Again by $\operatorname {MLC}(1)$ of $\pi '$ , we have $\hat {C}^{\prime }_n=\pi _n^{n+1}(C^{\prime }_{n+1})$ (see Lemma 2.1). Then a map $\phi \colon Y\to \pi _n^{n+1}(C^{\prime }_{n+1})$ , defined by $\phi (y)=y_n$ for all $y=(y_m)_{m\ge n}\in Y$ , gives a factor map
Property (2) in the claim ensures that $\pi _n^{n+1}(C_{n+1})$ is a closed $f_n$ -invariant subset of $\pi _n^{n+1}(C^{\prime }_{n+1})$ , therefore,
Thus, $f|_C$ has the shadowing property and satisfies $h_{\mathrm {top}}(f|_C)>0$ , completing the proof of the lemma.
5 Proof of Lemma 1.3
In this section we prove Lemma 1.3. Let
be an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ and suppose that $(X_n,f_n)$ is a transitive SFT for each $n\ge 1$ . For every $n\ge 1$ , note that $\mathcal {D}(f_n)$ is a finite set, and let $m_n=|\mathcal {D}(f_n)|$ . Then $m_n|m_{n+1}$ for all $n\ge 1$ , and for any $n\ge 1$ and $E\in \mathcal {D}(f_n)$ , $(f_n^{m_n})|_E\colon E\to E$ is mixing. The proof of the first lemma is similar to that of Lemma 4.1.
Lemma 5.1. There is $D_{\ast }=(D_n)_{n\ge 1}\in \mathcal {D}_\pi $ such that
satisfies $\operatorname {MLC}(1)$ .
Proof. We argue as in the proof of Lemma 4.1.
Step 1. Fix $D_1\in \mathcal {D}(f_1)$ with $\pi _1^2(F_2)\subset D_1$ for some $F_2\in \mathcal {D}(f_2)$ . Take $E_2\in \mathcal {D}(f_2)$ such that
and $\pi _1^2(E_2)$ is maximal among
with respect to the inclusion relation.
Step 2. Note that $(f_1^{m_2})|_{\pi _1^2(E_2)}$ is mixing, so $(f_1^{m_3})|_{\pi _1^2(E_2)}$ is transitive, and take a transitive point $p_1\in \pi _1^2(E_2)$ , that is, $\pi _1^2(E_2)=\omega (p_1,f_1^{m_3})$ , the $\omega $ -limit set. Since
we have $p_1=\pi _1^3(q_1)$ for some $q_1\in X_3$ . Take $D_2\in \mathcal {D}(f_2)$ with $\pi _2^3(q_1)\in D_2$ . Then choose $E_3\in \mathcal {D}(f_3)$ such that
and $\pi _2^3(E_3)$ is maximal among
with respect to the inclusion relation. By (P2), we have
implying
since $\pi _1^2(E_2)=\omega (p_1,f_1^{m_3})$ , and $\pi _1^3(E_3)$ is $f_1^{m_3}$ -invariant. By (P1) and $p_1\in \pi _1^2(E_2)$ , we see that $p_1\in D_1\cap \pi _1^2(D_2)$ ; therefore,
By the maximality of $\pi _1^2(E_2)$ in Step 1, we obtain
Step 3. Note that $(f_2^{m_3})|_{\pi _2^3(E_3)}$ is mixing, so $(f_2^{m_4})|_{\pi _2^3(E_3)}$ is transitive, and take a transitive point $p_2\in \pi _2^3(E_3)$ , that is, $\pi _2^3(E_3)=\omega (p_2,f_2^{m_4})$ . Since
we have $p_2=\pi _2^4(q_2)$ for some $q_2\in X_4$ . Take $D_3\in \mathcal {D}(f_3)$ with $\pi _3^4(q_2)\in D_3$ . Then choose $E_4\in \mathcal {D}(f_4)$ such that
and $\pi _3^4(E_4)$ is maximal among
with respect to the inclusion relation. By (P3), we have
implying
since $\pi _2^3(E_3)=\omega (p_2,f_2^{m_4})$ , and $\pi _2^4(E_4)$ is $f_2^{m_4}$ -invariant. By (P2) and $p_2\in \pi _2^3(E_3)$ , we see that $\pi _2^3(q_1)\in \pi _2^3(E_3)$ and $p_2\in D_2\cap \pi _2^3(D_3)$ ; therefore,
By the maximality of $\pi _2^3(E_3)$ in Step 2, we obtain
Assertions (Q1) and (Q2) yield $\pi _1^2(D_2)=\pi _1^3(D_3)$ .
Step 4. Note that $(f_3^{m_4})|_{\pi _3^4(E_4)}$ is mixing, so $(f_3^{m_5})|_{\pi _3^4(E_4)}$ is transitive, and take a transitive point $p_3\in \pi _3^4(E_4)$ , that is, $\pi _3^4(E_4)=\omega (p_3,f_3^{m_5})$ . Since
we have $p_3=\pi _3^5(q_3)$ for some $q_3\in X_5$ . Take $D_4\in \mathcal {D}(f_4)$ with $\pi _4^5(q_3)\in D_4$ . Then choose $E_5\in \mathcal {D}(f_5)$ such that
and $\pi _4^5(E_5)$ is maximal among
with respect to the inclusion relation. By (P4), we have
implying
since $\pi _3^4(E_4)=\omega (p_3,f_3^{m_5})$ , and $\pi _3^5(E_5)$ is $f_3^{m_5}$ -invariant. By (P3) and $p_3\in \pi _3^4(E^4)$ , we see that $\pi _3^4(q_2)\in \pi _3^4(E_4)$ and $p_3\in D_3\cap \pi _3^4(D_4)$ ; therefore,
By the maximality of $\pi _3^4(E_4)$ in Step 3, we obtain
Assertions (Q2) and (Q3) yield $\pi _2^3(D_3)=\pi _2^4(D_4)$ .
Continuing inductively, we obtain a sequence $D_{\ast }=(D_n)_{n\ge 1}\in \prod _{n\ge 1}\mathcal {D}(f_n)$ . For any $n\ge 1$ , $\pi _n^{n+1}(D_{n+1})\subset D_n$ and $\pi _n^{n+1}(D_{n+1})=\pi _n^{n+2}(D_{n+2})$ are established in Steps $n+1$ and $n+2$ , respectively. Thus, $D_{\ast }\in \mathcal {D}_\pi $ , and
satisfies $\operatorname {MLC}(1)$ , completing the proof.
The next lemma relates the previous lemma to a method developed in [Reference Kawaguchi15]. Let $\pi =(\pi _n^{n+1}\colon (X_{n+1},f_{n+1})\to (X_n,f_n))_{n\ge 1}$ be an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ and let $(X,f)=\lim _\pi (X_n,f_n)$ . Suppose that $(X_n,f_n)$ , $n\ge 1$ , are transitive SFTs, and for $D_{\ast }=(D_n)_{n\ge 1}\in \mathcal {D}_\pi $ ,
satisfies $\operatorname {MLC}(1)$ . By Lemma 3.3, letting $D=[D_{\ast }]$ , we have $D\in \mathcal {D}(f)$ . Let d, $d_n$ , $n\ge 1$ , be the metrics on X, $X_n$ .
Lemma 5.2. For any $\epsilon>0$ , there is $\delta>0$ such that every $\delta $ -pseudo-orbit $(x^{(i)})_{i\ge 0}$ of f with $x^{(0)}\in D$ is $\epsilon $ -shadowed by some $x\in D$ .
Proof. Fix any $N>0$ and $\epsilon '>0$ . Note that $f_{N+1}\colon X_{N+1}\to X_{N+1}$ has the shadowing property, and $\mathcal {D}(f_{N+1})$ is a clopen partition of $X_{N+1}$ ; therefore, there is $\delta '>0$ such that, for any $E_{N+1}\in \mathcal {D}_{N+1}$ , every $\delta '$ -pseudo-orbit $(y_{N+1}^{(i)})_{i\ge 0}$ of $f_{N+1}$ with $y_{N+1}^{(0)}\in E_{N+1}$ is $\epsilon '$ -shadowed by some $y_{N+1}\in E_{N+1}$ .
If $\delta>0$ is small enough, then, for every $\delta $ -pseudo-orbit $\xi =(x^{(i)})_{i\ge 0}$ of f with $x^{(0)}\in D$ , $\xi _{N+1}=(x_{N+1}^{(i)})_{i\ge 0}$ is a $\delta '$ -pseudo-orbit of $f_{N+1}$ with $x_{N+1}^{(0)}\in D_{N+1}$ , which is $\epsilon '$ -shadowed by some $x_{N+1}\in D_{N+1}$ . Since
by $\operatorname {MLC}(1)$ of $\tilde {\pi }$ (see Lemma 2.1), there is $x=(x_n)_{n\ge 1}\in D=[D_{\ast }]$ such that $x_n=\pi _n^{N+1}(x_{N+1})$ for each $1\le n\le N$ . Then, for any $i\ge 0$ and $1\le n\le N$ , we have
with $d_{N+1}(f_{N+1}^i(x_{N+1}),x_{N+1}^{(i)})\le \epsilon '$ . Therefore, for every $\epsilon>0$ , if N is large enough, and then $\epsilon '$ is sufficiently small, we have $d(f^i(x),x^{(i)})\le \epsilon $ for all $i\ge 0$ , that is, $\xi $ is $\epsilon $ -shadowed by $x\in D$ . Since $\xi $ is arbitrary, the lemma has been proved.
For the proof of Lemma 1.3, we need a sequence of lemmas. Let $f\colon X\to X$ be a chain transitive continuous map. For $\epsilon ,\delta>0$ , we denote by $\mathcal {D}^{\epsilon ,\delta }(f)$ the set of $D\in \mathcal {D}(f)$ for which every $\delta $ -pseudo-orbit $(x_i)_{i\ge 0}$ of f with $x_0\in D$ is $\epsilon $ -shadowed by some $x\in D$ . We set
Lemma 5.3. Let $f\colon X\to X$ be a chain transitive continuous map and let $\epsilon ,\delta>0$ . For any $D\in \mathcal {D}(f)$ , if $D\in \mathcal {D}^{\epsilon ,\delta }(f)$ , then $f(D)\in \mathcal {D}^{\epsilon ,\delta }(f)$ .
Proof. Let $D\in \mathcal {D}^{\epsilon ,\delta }(f)$ . For any $\delta $ -pseudo-orbit $\xi =(x_i)_{i\ge 0}$ of f with $x_0\in f(D)$ , take $y\in D$ with $f(y)=x_0$ and consider
a $\delta $ -pseudo-orbit of f, which is $\epsilon $ -shadowed by $x\in D$ . Then $\xi $ is $\epsilon $ -shadowed by $f(x)\in f(D)$ . Since $\xi $ is arbitrary, we obtain $f(D)\in \mathcal {D}^{\epsilon ,\delta }(f)$ , proving the lemma.
Lemma 5.4. Let $f\colon X\to X$ be a chain transitive continuous map and let $\epsilon ,\gamma>0$ . If $\mathcal {D}_{\mathrm {sh}}(f)\cap \mathcal {D}^{\epsilon ,\gamma }(f)\ne \emptyset $ , then $\mathcal {D}(f)=\mathcal {D}^{\epsilon ,\delta }(f)$ holds for every $0<\delta <\gamma $ .
Proof. Fix $D_0\in \mathcal {D}_{\mathrm {sh}}(f)\cap \mathcal {D}^{\epsilon ,\gamma }(f)$ , $x\in D_0$ , and a sequence $0<\epsilon _1>\epsilon _2>\cdots \to 0$ . Since $D_0\in \mathcal {D}_{\mathrm {sh}}(f)$ , there is a sequence $0<\delta _1>\delta _2>\cdots \to 0$ such that $D_0\in \mathcal {D}^{\epsilon _n,\delta _n}(f)$ for every $n\ge 1$ . Given any $D\in \mathcal {D}(f)$ and any $\delta $ -pseudo-orbit $\xi =(x_i)_{i\ge 0}$ of f with $x_0\in D$ , since f is chain transitive, for every $n\ge 1$ , there is a $\delta _n$ -chain $(x_i^{(n)})_{i=0}^{k_n}$ of f with $x_0^{(n)}=x$ and $x_{k_n}^{(n)}=x_0$ . Then, for each $n\ge 1$ , because $D_0\in \mathcal {D}^{\epsilon _n,\delta _n}(f)$ and $x_0^{(n)}=x\in D_0$ , we have $x_n\in D_0$ with $d(f^i(x_n),x_i^{(n)})\le \epsilon _n$ for all $0\le i\le k_n$ . Since
$n\ge 1$ , $\lim _{n\to \infty }\epsilon _n=0$ , and $\delta <\gamma $ , there exists $N>0$ such that for all $n\ge N$ ,
is a $\gamma $ -pseudo-orbit of f with $f^{k_n}(x_n)\in f^{k_n}(D_0)$ . For any $n\ge N$ , because $D_0\in \mathcal {D}^{\epsilon ,\gamma }(f)$ , by Lemma 5.3, we have $f^{k_n}(D_0)\in \mathcal {D}^{\epsilon ,\gamma }(f)$ , so $\xi _n$ is $\epsilon $ -shadowed by some $y_n\in f^{k_n}(D_0)$ . Taking a sequence $N\le n_1<n_2<\cdots $ such that $\lim _{j\to \infty }y_{n_j}=y$ for some $y\in X$ , we easily see that $\xi $ is $\epsilon $ -shadowed by y. Note that, for each $n\ge N$ , $\{f^{k_n}(x_n),y_n\}\subset f^{k_n}(D_0)$ and so $f^{k_n}(x_n)\sim _f y_n$ . Because $\sim _f$ is closed in $X^2$ , by
we obtain $x_0\sim _f y$ and thus $y\in D$ . In other words, $\xi $ is $\epsilon $ -shadowed by $y\in D$ . Since $\xi $ and then $D\in \mathcal {D}(f)$ are arbitrary, we conclude that $\mathcal {D}(f)=\mathcal {D}^{\epsilon ,\delta }(f)$ .
As a consequence of Lemma 5.4, we obtain the following corollary.
Corollary 5.1. For any chain transitive continuous map $f\colon X\to X$ , if $\mathcal {D}_{\mathrm {sh}}(f)\ne \emptyset $ , then $\mathcal {D}(f)=\mathcal {D}_{\mathrm {sh}}(f)$ .
The next lemma is needed for the proof of Lemma 5.6.
Lemma 5.5. Let $f\colon X\to X$ be a chain transitive continuous map and let $D\in \mathcal {D}(f)$ . If, for any $x\in D$ and $\epsilon>0$ , there is $\delta (x,\epsilon )>0$ such that every $\delta (x,\epsilon )$ -pseudo-orbit $(x_i)_{i\ge 0}$ of f with $x_0=x$ is $\epsilon $ -shadowed by some $y\in D$ , then $D\in \mathcal {D}_{\mathrm {sh}}(f)$ .
Proof. Fix $\epsilon>0$ and, for any $x\in D$ , take $\delta (x,\epsilon /2)>0$ as in the assumption. Then, for each $x\in D$ , there is $0<\delta (x)<\epsilon /2$ such that, for every $\delta (x)$ -pseudo-orbit $\xi =(x_i)_{i\ge 0}$ of f with $d(x,x_0)<\delta (x)$ ,
is a $\delta (x,\epsilon /2)$ -pseudo-orbit of f, $\epsilon /2$ -shadowed by some $y\in D$ . This clearly implies that $\xi $ is $\epsilon $ -shadowed by $y\in D$ . Take a finite subset $F\subset D$ such that
where $B_{\delta (x)}(x)$ denotes the $\delta (x)$ -ball. Let $\delta =\min \{\delta (x)\colon x\in F\}$ . It follows that every $\delta $ -pseudo-orbit $(x_i)_{i\ge 0}$ of f with $x_0\in D$ is a $\delta (x)$ -pseudo-orbit of f with $x_0\in B_{\delta (x)}(x)$ for some $x\in F$ , and so $\epsilon $ -shadowed by some $y\in D$ . Since $\epsilon>0$ is arbitrary, we obtain $D\in \mathcal {D}_{\mathrm {sh}}(f)$ , proving the lemma.
Lemma 5.6. Let $f\colon X\to X$ be a chain transitive continuous map with the shadowing property. Let Y be a compact metric space and let $g\colon Y\to Y$ be a chain transitive continuous map. If there is a factor map $\pi \colon (Y,g)\to (X,f)$ , and if $\mathcal {D}_{\mathrm {sh}}(g)\ne \emptyset $ , then $\mathcal {D}_{\mathrm {sh}}(f)\ne \emptyset $ .
Proof. Fix $D\in \mathcal {D}(f)$ . By Lemma 5.5, it suffices to show that, for any $x\in D$ and $\epsilon>0$ , there exists $\delta>0$ such that every $\delta $ -pseudo-orbit $(x_i)_{i\ge 0}$ of f with $x_0=x$ is $\epsilon $ -shadowed by some $q\in D$ , because this implies $D\in \mathcal {D}_{\mathrm {sh}}(f)$ .
Let $d_Y$ denote the metric on Y. For any $\epsilon '>0$ , Lemma 5.4 with $\mathcal {D}_{\mathrm {sh}}(g)\ne \emptyset $ implies $\mathcal {D}(g)=\mathcal {D}^{\epsilon ',\delta '}(g)$ for some $\delta '>0$ . For any $\gamma>0$ , take $0<\epsilon ''<\epsilon /2$ such that $d(x,y)\le \epsilon ''$ implies
for all $y\in X$ . Since f has the shadowing property, there is $\delta>0$ such that every $\delta $ -pseudo-orbit $\xi =(x_i)_{i\ge 0}$ of f with $x_0=x$ is $\epsilon ''$ -shadowed by some $y\in X$ . Take $z\in \pi ^{-1}(y)$ and note that $d(x,y)=d(x_0,y)\le \epsilon ''$ . By the choice of $\epsilon ''$ , we obtain $w\in \pi ^{-1}(x)$ such that $d_Y(z,w)<\gamma $ . If $\gamma $ is sufficiently small, then
is a $\delta '$ -pseudo-orbit of g. By $\mathcal {D}(g)=\mathcal {D}^{\epsilon ',\delta '}(g)$ , $\xi '$ is $\epsilon '$ -shadowed by some $p\in Y$ with $w\sim _g p$ . Here, $w\sim _g p$ implies $\pi (w)\sim _f \pi (p)$ , so, putting $q=\pi (p)$ , we have $x\sim _f q$ , that is, $q\in D$ . Note that
and
for each $i\ge 1$ . Since $\xi '$ is $\epsilon '$ -shadowed by p, we have $d_Y(g^i(p),y_i)\le \epsilon '$ , $i\ge 0$ , so if $\epsilon '$ is sufficiently small, then $d(f^i(q),x_i)\le \epsilon $ for all $i\ge 0$ , that is, $\xi $ is $\epsilon $ -shadowed by $q\in D$ . Since $\xi $ is arbitrary, this shows the existence of $\delta $ , and thus the lemma has been proved.
For any chain transitive continuous map $f\colon X\to X$ , we denote by $T(f)$ the set of transitive points for f:
where $\omega (\cdot ,f)$ denotes the $\omega $ -limit set.
Lemma 5.7. Let $f\colon X\to X$ be a chain transitive continuous map. For any $D\in \mathcal {D}(f)$ , if $D\in \mathcal {D}_{\mathrm {sh}}(f)$ , then $D\cap T(f)$ is a dense $G_\delta $ -subset of D.
Proof. Let $\{U_n\colon n\ge 1\}$ be a countable basis for the topology of X. Then
a $G_\delta $ -subset of D. For any $n\ge 1$ , take $p_n\in U_n$ and $\epsilon _n>0$ such that
Let $n\ge 1$ and $j\ge 0$ . For any $x\in D$ and $0<\epsilon <\epsilon _n$ , since $D\in \mathcal {D}_{\mathrm {sh}}(f)$ , we have $D\in \mathcal {D}^{\epsilon ,\delta }(f)$ for some $\delta>0$ . Then the chain transitivity of f gives a $\delta $ -chain $(x_i)_{i=0}^k$ of f with $x_0=x$ , $x_k=p_n$ , and also $k\ge j$ . By $x\in D$ and $D\in \mathcal {D}^{\epsilon ,\delta }(f)$ , we obtain $d(f^i(y),x_i)\le \epsilon $ for all $0\le i\le k$ for some $y\in D$ . Note that $d(y,x)=d(y,x_0)\le \epsilon $ , $d(f^k(y),p_n)=d(f^k(y),x_k)\le \epsilon <\epsilon _n$ , and so $f^k(y)\in U_n$ , implying
Since $x\in D$ and $0<\epsilon <\epsilon _n$ are arbitrary, this shows that
is dense in D. Since $n\ge 1$ and $j\ge 0$ are arbitrary, we conclude that $D\cap T(f)$ is a dense $G_\delta $ -subset of D, completing the proof.
To prove Lemma 1.3, we use the method in [Reference Kawaguchi15]. The next lemma is a modification of [Reference Kawaguchi15, Lemma 2.6].
Lemma 5.8. Let $f\colon X\to X$ be a chain transitive continuous map and let $D\in \mathcal {D}(f)$ . If $D\in \mathcal {D}_{\mathrm {sh}}(f)$ , then, for any $y,z\in D$ and $\epsilon>0$ , there is $w\in D$ such that $d(z,w)\le \epsilon $ and $\limsup _{k\to \infty }d(f^k(y),f^k(w))\le \epsilon $ .
Proof. Given any $\epsilon>0$ , take $\delta>0$ so small that $D\in \mathcal {D}^{\epsilon ,\delta }(f)$ . For this $\delta $ , choose $N>0$ as in property (3) of $\sim _{f,\delta }$ (see §2.2.3). Note that $y,z\in D$ implies $y\sim _f z$ and so $y\sim _{f,\delta }z$ . Since $y\sim _{f,\delta }f^{mN}(y)$ , we have $z\sim _{f,\delta }f^{mN}(y)$ . Then the choice of N gives a $\delta $ -chain $\alpha =(y_i)_{i=0}^{mN}$ of f with $y_0=z$ and $y_{mN}=f^{mN}(y)$ . Let
and $\xi =\alpha \beta =(x_i)_{i\ge 0}$ . Then $\xi $ is a $\delta $ -pseudo-orbit of f with $x_0=z\in D$ , so is $\epsilon $ -shadowed by some $w\in D$ . Note that $d(z,w)=d(x_0,w)\le \epsilon $ . Also, we have
for every $i\ge mN$ , so $\limsup _{k\to \infty }d(f^k(y),f^k(w))\le \epsilon $ . This completes the proof.
Let $f\colon X\to X$ be a continuous map. For $n\ge 2$ and $r>0$ , we say that an n-tuple $(x_1,x_2,\ldots ,x_n)\in X^n$ is r-distal if
Then the following lemma is a consequence of [Reference Kawaguchi15, Lemmas 2.4 and 2.5].
Lemma 5.9. Suppose that a continuous map $f\colon X\to X$ is chain transitive and has the shadowing property. If $h_{\mathrm {top}}(f)>0$ , then, for any $n\ge 2$ , there is $r_n>0$ such that, for every $D\in \mathcal {D}(f)$ , there is an $r_n$ -distal n-tuple $(x_1,x_2,\ldots ,x_n)\in X^n$ with $\{x_1,x_2,\ldots ,x_n\}\subset D$ .
We recall a simplified version of Mycielski’s theorem [Reference Mycielski27, Theorem 1]. A topological space is said to be perfect if it has no isolated point.
Lemma 5.10. Let X be a perfect complete metric space. If $R_n$ is a residual subset of $X^n$ for each $n\ge 2$ , then there is a Mycielski set S which is dense in X and satisfies $(x_1,x_2,\ldots ,x_n)\in R_n$ for any $n\ge 2$ and distinct $x_1,x_2,\ldots ,x_n\in S$ .
Finally, we complete the proof of Lemma 1.3.
Proof of Lemma 1.3
Due to Lemmas 1.2 and 2.4, $(Y,g)$ is topologically conjugate to $(Z,h)=\lim _{\pi }(X_j,f_j)$ where $(X_j,f_j)$ , $j\ge 1$ , are SFTs, and
is an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ . Since g is transitive and so is h, we have $Z\in \mathcal {C}(h)$ , so by Lemma 3.2, $Z=[C_{\ast }]$ for some $C_{\ast }=(C_j)_{j\ge 1}\in \mathcal {C}_\pi $ . Then any $z=(x_j)_{j\ge 1}\in Z$ satisfies $x_j\in C_j$ for all $j\ge 1$ , and $(Z,h)$ is topologically conjugate to $\lim _{\pi '}(C_j,(f_j)|_{C_j})$ , where
Note that $(C_j,(f_j)|_{C_j})$ , $j\ge 1$ , are transitive SFTs, and by Lemma 3.4, $\pi '$ satisfies $\operatorname {MLC}(1)$ ; therefore, without loss of generality, we may assume that $f_j$ is transitive for every $j\ge 1$ . Then, by Lemma 5.1, there is $D_{\ast }=(D_j)_{j\ge 1}\in \mathcal {D}_\pi $ such that
satisfies $\operatorname {MLC}(1)$ . Letting $E=[D_{\ast }]\in \mathcal {D}(h)$ , by Lemma 5.2, we see that $E\in \mathcal {D}_{\mathrm {sh}}(h)$ , which implies $\mathcal {D}_{\mathrm {sh}}(h)\ne \emptyset $ and so $\mathcal {D}_{\mathrm {sh}}(g)\ne \emptyset $ . From Lemma 5.6 and Corollary 5.1, it follows that $\mathcal {D}(f)=\mathcal {D}_{\mathrm {sh}}(f)$ . For any $D\in \mathcal {D}(f)$ , since $D\in \mathcal {D}_{\mathrm {sh}}(f)$ , D satisfies the conclusions of Lemmas 5.7 and 5.8. Note that the conclusion of Lemma 5.9 is also satisfied. Similarly to the proof of [Reference Kawaguchi15, Theorem 1.1], it can be shown that there exists a sequence of positive numbers $(\delta _n)_{n\ge 2}$ for which
is a residual subset of $D^n$ for all $D\in \mathcal {D}(f)$ and $n\ge 2$ . By Lemma 5.10, we conclude that every $D\in \mathcal {D}(f)$ contains a dense Mycielski subset S which is included in $T(f)$ and distributionally n- $\delta _n$ -scrambled for all $n\ge 2$ , completing the proof.
6 A remark on the chain components under shadowing
Given any continuous map $f\colon X\to X$ , $\mathcal {C}(f)$ can be seen as a quotient space of $\operatorname {CR}(f)$ with respect to the closed $(f\times f)$ -invariant equivalence relation $\leftrightarrow _f$ in $\operatorname {CR}(f)^2$ . Then $\mathcal {C}(f)=\operatorname {CR}(f)/\leftrightarrow _f$ is a compact metric space.
In the case of $\dim {X}=0$ , if f has the shadowing property, then by Lemmas 1.2 and 2.4, $(X,f)$ is topologically conjugate to $\lim _\pi (X_n,f_n)$ , where $(X_n,f_n)$ , $n\ge 1$ , are SFTs, and
is an inverse sequence of equivariant maps with $\operatorname {MLC}(1)$ . Without loss of generality, we consider the case where $(X,f)=\lim _\pi (X_n,f_n)$ . For any $C\in \mathcal {C}(f)$ , by Lemma 3.2, we have $C=[C_{\ast }]$ for some $C_{\ast }=(C_n)_{n\ge 1}\in \mathcal {C}_\pi $ . As in the proof of Lemma 4.1, it can be shown that for each $N>0$ , there is $C^{\prime }_{\ast }=(C^{\prime }_m)_{m\ge N}\in \prod _{m\ge N}\mathcal {C}(f_m)$ with the following properties:
-
(1) $C^{\prime }_N=C_N$ ;
-
(2) $\pi _m^{m+1}(C^{\prime }_{m+1})\subset C^{\prime }_m$ for every $m\ge N$ ;
-
(3) $\pi _m^{m+1}(C^{\prime }_{m+1})=\pi _m^{m+2}(C^{\prime }_{m+2})$ for all $m\ge N$ .
Define $C^{\prime \prime }_{\ast }=(C^{\prime \prime }_n)_{n\ge 1}\in \prod _{n\ge 1}\mathcal {C}(f_n)$ by
Properties (1) and (2) ensure $C^{\prime \prime }_{\ast }\in \mathcal {C}_\pi $ . By Lemma 3.2, letting $C''=[C^{\prime \prime }_{\ast }]$ , we obtain $C''\in \mathcal {C}(f)$ , and by property (3), similarly to the proof of Lemma 4.1, it can be seen that $f|_{C''}\colon C''\to C''$ has the shadowing property. Note that for any neighborhood U of C in $\operatorname {CR}(f)$ , by property (1) above, if N is sufficiently large, then $C''\subset U$ . Thus, letting
we conclude that $\mathcal {C}_{\mathrm {sh}}(f)$ is dense in $\mathcal {C}(f)$ . In other words, we obtain the following theorem.
Theorem 6.1. Let $f\colon X\to X$ be a continuous map with the shadowing property. If $\dim {X}=0$ , then $\mathcal {C}(f)=\overline {\mathcal {C}_{\mathrm {sh}}(f)}$ .
This theorem gives a positive answer to a question by Moothathu [Reference Moothathu26] in the zero-dimensional case. Note that, for any $C\in \mathcal {C}_{\mathrm {sh}}(f)$ , by the shadowing property of $f|_C$ , we have $C=\overline {M(f|_C)}$ , where $M(f|_C)$ denotes the set of minimal points for $f|_C$ (see [Reference Moothathu26] for details).
As a complement to Theorem 6.1, we give an example of a continuous map $f\colon X\to X$ with the following properties:
-
(1) X is a Cantor space;
-
(2) f has the shadowing property;
-
(3) $\mathcal {C}(f)$ is a Cantor space;
-
(4) $\mathcal {C}_{\mathrm {sh}}(f)$ is a countable set and so is a meager subset of $\mathcal {C}(f)$ .
Example 6.1. For any closed interval $I=[a,b]$ and $c\in (0,1/2)$ , let $\hat {I}=\{a,b\}$ , $I_c^{(0)}=[a,a+c(b-a)]$ , and $I_c^{(1)}=[b-c(b-a),b]$ . Let $(c_j)_{j\ge 1}$ be a sequence of positive numbers with $1/2>c_1>c_2>\cdots $ . For any $s=(s_j)_{j\ge 1}\in \{0,1\}^{\mathbb {N}}$ , let
where $I(s,j)$ is defined by $I(s,0)=[0,1]$ , and $I(s,j+1)=I(s,j)_{c_{j+1}}^{(s_{j+1})}$ for every $j\ge 0$ . Let
and note that $i\colon \{0,1\}^{\mathbb {N}}\to C$ is a homeomorphism, so C is a Cantor space. For any $j\ge 1$ , let
and note that $\hat {I}_1\subset \hat {I}_2\subset \cdots $ . Also, let $A_1=\hat {I}_1$ , $A_{j+1}=\hat {I}_{j+1}\setminus \hat {I}_j$ , $j\ge 1$ , and
Let $\sigma \colon \{0,1\}^{\mathbb {N}}\to \{0,1\}^{\mathbb {N}}$ be the shift map. For each $k\ge 1$ , define
which is a mixing SFT, so $\sigma |_{\Sigma _k}\colon \Sigma _k\to \Sigma _k$ has the shadowing property. Note that $\Sigma _1\supset \Sigma _2\supset \cdots $ and consider
Then it is easily seen that $\sigma |_{\Sigma _\infty }\colon \Sigma _\infty \to \Sigma _\infty $ does not have the shadowing property. Let $\hat {\Sigma }_k=i(\Sigma _k)$ , $k\ge 1$ , and $\hat {\Sigma }_\infty =i(\Sigma _\infty )$ ; here, $i\colon \{0,1\}^{\mathbb {N}}\to C$ is the homeomorphism defined above. Let
which is a perfect compact subset of $C\times C$ and so is a Cantor space.
Let $\hat {\sigma }=i\circ \sigma \circ i^{-1}\colon C\to C$ and
To ensure the shadowing property of f, we define a sequence of positive numbers $(c_j)_{j\ge 1}$ with $1/2>c_1>c_2>\cdots $ as follows. Fix a sequence of positive numbers $(\epsilon _k)_{k\ge 1}$ with $\lim _{k\to \infty }\epsilon _k=0$ . Denote by $\pi \colon X\to C$ the projection onto the first coordinate. Let
$k\ge 1$ . For each $k\ge 1$ , since $\pi ^{-1}(B_k)$ is a finite disjoint union of SFTs,
has the shadowing property, implying the existence of $\delta ^{\prime }_k>0$ such that every $\delta ^{\prime }_k$ -pseudo-orbit $(x_i)_{i\ge 0}$ of $f|_{\pi ^{-1}(B_k)}$ is $\epsilon _k/2$ -shadowed by some $x\in \pi ^{-1}(B_k)$ . Fix any $c_1\in (0,1/2)$ and assume that $c_k$ , $k\ge 1$ , is given. For any $\delta _k\in (0,\epsilon _k/2)$ , if $c_{k+1}\in (0,c_k)$ is small enough, then X is contained in the $\delta _k$ -neighborhood of $\pi ^{-1}(B_k)$ . Then, for every $\delta _k$ -pseudo-orbit $(y_i)_{i\ge 0}$ of f, we have
for all $i\ge 0$ for some $x_i\in \pi ^{-1}(B_k)$ . Since
for every $i\ge 0$ , if $\delta _k$ is small enough, then $(x_i)_{i\ge 0}$ is a $\delta ^{\prime }_k$ -pseudo-orbit of $f|_{\pi ^{-1}(B_k)}$ , $\epsilon _k/2$ -shadowed by some $x\in \pi ^{-1}(B_k)$ . This implies
for all $i\ge 0$ , that is, $(y_i)_{i\ge 0}$ is $\epsilon _k$ -shadowed by x. By defining $(c_j)_{j\ge 1}$ in this way, we conclude that f has the shadowing property.
Note that $X=\operatorname {CR}(f)$ and
which is a countable set. It remains to show that $\mathcal {C}(f)$ is a Cantor space. Let $\pi _{\leftrightarrow _f}\colon X\to \mathcal {C}(f)$ be the quotient map. For any $x,y\in X$ , we easily see that $x\leftrightarrow _f y$ if and only if $\pi (x)=\pi (y)$ . This implies that there is a continuous map $h\colon \mathcal {C}(f)\to C$ with $\pi =h\circ \pi _{\leftrightarrow _f}$ , which is bijective and so is a homeomorphism. Thus, $\mathcal {C}(f)$ is a Cantor space.
Acknowledgements
The author would like to thank the anonymous reviewer for helpful comments and suggestions. The author is a JSPS Research Fellow. This work was supported by JSPS KAKENHI grant number JP20J01143.