This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Nonwandering sets and Special α\alpha-limit sets
of Monotone maps on Regular Curves

Aymen Daghar and Habib Marzougui Aymen Daghar,  University of Carthage, Faculty of Science of Bizerte, (UR17ES21), “Dynamical systems and their applications”, Jarzouna, 7021, Bizerte, Tunisia. [email protected] Habib Marzougui, University of Carthage, Faculty of Science of Bizerte, (UR17ES21), “Dynamical systems and their applications”, 7021, Jarzouna, Bizerte, Tunisia [email protected]
Abstract.

Let XX be a regular curve and let f:XXf:X\to X be a monotone map. In this paper, nonwandering set of ff and the structure of special α\alpha-limit sets for ff are investigated. We show that AP(f)=R(f)=Ω(f)(f)=\textrm{R}(f)=\Omega(f), where AP(f)(f), R(f)\textrm{R}(f) and Ω(f)\Omega(f) are the sets of almost periodic points, recurrent points and nonwandering of ff, respectively. This result extends that of Naghmouchi established, whenever ff is a homeomorphism on a regular curve [J. Difference Equ. Appl., 23 (2017), 1485–1490] and [Colloquium Math., 162 (2020), 263–277], and that of Abdelli and Abdelli, Abouda and Marzougui, whenever ff is a monotone map on a local dendrite [Chaos, Solitons Fractals, 71 (2015), 66–72] and [Topology Appl., 250 (2018), 61–73], respectively. On the other hand, we show that for every XP(f)X\setminus\textrm{P}(f), the special α\alpha-limit set sαf(x)s\alpha_{f}(x) is a minimal set, where P(f)(f) is the set of periodic points of ff and that sαf(x)s\alpha_{f}(x) is always closed, for every xXx\in X. In addition, we prove that SA(f)=R(f)\textrm{SA}(f)=\textrm{R}(f), where SA(f)\textrm{SA}(f) denotes the union of all special α\alpha-limit sets of ff; these results extend, for monotone case, recent results on interval and graph maps obtained respectively by Hantáková and Roth in [Preprint: arXiv 2007.10883.] and Foryś-Krawiec, Hantáková and Oprocha in [Preprint: arXiv:2106.05539.]. Further results related to the continuity of the limit maps are also obtained, we prove that the map ωf\omega_{f} (resp. αf\alpha_{f}, resp. sαf\alpha_{f}) is continuous on XP(f)X\setminus\textrm{P}(f) (resp. XP(f)X_{\infty}\setminus\textrm{P}(f)).

Key words and phrases:
Minimal set, regular curve, ω\omega-limit set, local dendrite, monotone map
2010 Mathematics Subject Classification:
37B20, 37B45, 54H20

1. Introduction

In the last two decades, a wide literature on the dynamical properties of maps on some one-dimensional continua has developed. Examples of continua become increasingly studied include, graphs, dendrites and local dendrites (see for instance [1], [2], [3], [30]).

Recently a large class of continua called regular curves has given a special attention (see for example, [5], [12], [13], [14], [20], [21], [28], [29], [31]). These form a large class of continua which includes local dendrites. The Sierpiński triangle is a well known example of a regular curve which is not a local dendrite. Regular curves appear in continuum theory and also in other branches of Mathematics such as complex dynamics; for instance, the Sierpiński triangle can be realized as the Julia set of the complex polynomial p(z)=z2+2p(z)=z^{2}+2 (see [9]). Recall that the Julia set of pp is J(p)={z:(pn(z))n1 is bounded}J(p)=\{z\in\mathbb{C}:(p^{n}(z))_{n\geq 1}\textrm{ is bounded}\}. Seidler [31] proved that every homeomorphism of a regular curve has zero topological entropy (later, this result was extended by Kato in [20] to monotone maps). In [28], [29], Naghmouchi proved that any ω\omega-limit set (resp. α\alpha-limit set) of a homeomorphism ff on a regular curve is a minimal set. Moreover he established the equality between the set of nonwandering points and the set of almost periodic points.  In [12], the first author gave a full characterization of minimal sets for homeomorphisms without periodic points on regular curves. In [15] it was shown that the set of periodic points is either empty or dense in the set of non-wandering points, for homeomorphisms on regular curves. In the present paper, we deal with several questions/problems.

First, we address the question of the equality between the set of nonwandering points and the set of almost periodic points. This was proved in two cases:

- for homeomorphisms on regular curves (Naghmouchi [29])

- for monotone maps on local dendrites (Abdelli et al. [3])

In Theorem 3.1, we prove a more general result by showing that this is true for monotone maps on regular curves. Nevertheless, the later result is false in general for monotone maps on dendroids as it is shown in Example 3.5. Moreover, we show in Theorem 3.3 that the set of nonwandering points coincides with the set of points belonging to their special α\alpha-limit sets (Proposition 6.8).

Second, beside the usual limits sets ω\omega-limit and α\alpha-limit, we are interested in the study of another kind of limit sets, called special α\alpha-limit set (see Section 6). We ask the question whether every special α\alpha-limit set is a minimal set? We show that for a monotone map ff on a regular curve, every special α\alpha-limit set sαf(x)s\alpha_{f}(x) of a non periodic point xx is a minimal set. However, we built an example of a monotone map ff on an infinite star for which sαf(x)s\alpha_{f}(x) is not a minimal set for some periodic point xx. Moreover we show that the inclusion sαf(x)αf(x)s\alpha_{f}(x)\subset\alpha_{f}(x) is strict. In addition, we prove that SA(f)=R(f)\textrm{SA}(f)=\textrm{R}(f), where SA(f)\textrm{SA}(f) respectively R(f)\textrm{R}(f) denotes the union of all special α\alpha-limit sets and the set of recurrent points of ff. Notice that it is shown recently that, for mixing graph maps f:GGf:G\to G, every special sαs\alpha-limit set is the ω\omega-limit set of some point from GG and moreover, for graph maps f:GGf:G\to G with zero topological entropy, every special sαs\alpha-limit set is a minimal set (cf. [24, Theorem 3.8]). On the other hand, Kolyada et al. showed in ([22, Theorem 3.3 and Corollary 3.11]), that sαf(x)s\alpha_{f}(x) is closed whenever ff is either an interval map for which the set of all periodic points is closed or ff is transitive. Recently, Hantáková and Roth [16] showed that a special α\alpha-limit set is closed for a piecewise monotone interval map. It is previously known that sαf(x)s\alpha_{f}(x) is closed for homeomorphisms on any compact metric space. In Corollary 6.7, we extend the later result to monotone maps on regular curves by showing that for any xXx\in X, sαf(x)s\alpha_{f}(x) is closed. In fact we show more precisely that for every xXx\in X_{\infty}, αf(x)Ω(f)=sαf(x)\alpha_{f}(x)\cap\Omega(f)=s\alpha_{f}(x) (cf. Theorem 6.2).

2. Preliminaries

Let XX be a compact metric space with metric dd and let f:XXf:X\to X be a continuous map. The pair (X,f)(X,f) is called a dynamical system. Let ,+\mathbb{Z},\ \mathbb{Z}_{+} and \mathbb{N} be the sets of integers, non-negative integers and positive integers, respectively. For n+n\in\mathbb{Z_{+}}, denote by fnf^{n} the nn-th iterate of ff; that is, f0=identityf^{0}=\textrm{identity} and fn=ffn1f^{n}=f\circ f^{n-1} if nn\in\mathbb{N}. For any xXx\in X, the subset Orbf(x)={fn(x):n+}\textrm{Orb}_{f}(x)=\{f^{n}(x):n\in\mathbb{Z}_{+}\} is called the orbit of xx (under ff). A subset AXA\subset X is called ff-invariant (resp. strongly ff-invariant) if f(A)Af(A)\subset A (resp., f(A)=Af(A)=A); it is further called a minimal set (under ff) if it is closed, non-empty and does not contain any ff-invariant, closed proper non-empty subset of XX. We define the ω\omega-limit set of a point xXx\in X to be the set:

ωf(x)\displaystyle\omega_{f}(x) ={yX:lim infn+d(fn(x),y)=0}\displaystyle=\{y\in X:\liminf_{n\to+\infty}d(f^{n}(x),y)=0\}
=n{fk(x):kn}¯.\displaystyle=\underset{n\in\mathbb{N}}{\cap}\overline{\{f^{k}(x):k\geq n\}}.

The set αf(x)=k0nkfn(x)¯\alpha_{f}(x)=\displaystyle\bigcap_{k\geq 0}\displaystyle\overline{\bigcup_{n\geq k}f^{-n}(x)} is called the α\alpha-limit set of xx. Equivalently a point yαf(x)y\in\alpha_{f}(x) if and only if there exist an increasing sequence of positive integers (nk)k(n_{k})_{k\in\mathbb{N}} and a sequence of points (xk)k0(x_{k})_{k\geq 0} such that fnk(xk)=xf^{n_{k}}(x_{k})=x and limk+xk=y.\displaystyle\lim_{k\rightarrow+\infty}x_{k}=y. It is much harder to deal with α\alpha-limit sets since there are many choices for points in a backward orbit. When ff is a homeomorphism, αf(x)=ωf1(x)\alpha_{f}(x)=\omega_{f^{-1}}(x).

Balibrea et al. [6] considered exactly one branch of the backward orbit as follows.

Definition 2.1.

Let xXx\in X. A negative orbit of xx is a sequence (xn)n(x_{n})_{n} of points in XX such that x0=xx_{0}=x and f(xn+1)=xnf(x_{n+1})=x_{n}, for every n0n\geq 0. The α\alpha-limit set of (xn)n0(x_{n})_{n\geq 0} denoted by αf((xn)n)\alpha_{f}((x_{n})_{n}) is the set of all limit points of (xn)n0(x_{n})_{n\geq 0}.

It is clear that αf((xn)n)αf(x)\alpha_{f}((x_{n})_{n})\subset\alpha_{f}(x). The inclusion can be strict; even for onto monotone maps on regular curves (see Example 6.10). Notice that we have the following equivalence:

(i) αf(x)\alpha_{f}(x)\neq\emptyset, (ii) xXx\in X_{\infty}. In particular, if ff is onto, then αf(x)\alpha_{f}(x)\neq\emptyset for every xXx\in X.

Proposition 2.2.

(([14, Corollary 2.2])) Let f:XXf:X\longrightarrow X be a continuous self mapping of a compact space XX. Let xXx\in X_{\infty}. Then:

  • (i)

    αf(x)\alpha_{f}(x) is non-empty, closed and ff-invariant. In addition, it is strongly ff-invariant whenever limn+diam(fn(x))=0\displaystyle\lim_{n\to+\infty}\mathrm{diam}(f^{-n}(x))=0.

  • (ii)

    αf((xn)n0)\alpha_{f}((x_{n})_{n\geq 0}) is non-empty, closed and strongly ff-invariant, for any negative orbit (xn)n0(x_{n})_{n\geq 0} of xx.

A point xXx\in X is called:

- Periodic of period nn\in\mathbb{N} if fn(x)=xf^{n}(x)=x and fi(x)xf^{i}(x)\neq x for 1in11\leq i\leq n-1; if n=1n=1, xx is called a fixed point of ff i.e. f(x)=xf(x)=x;

- Almost periodic if for any neighborhood UU of xx there is NN\in\mathbb{N} such that {fi+k(x):0iN}U\{f^{i+k}(x):0\leq i\leq N\}\cap U\neq\emptyset, for all kk\in\mathbb{N}.

-Recurrent if xωf(x)x\in\omega_{f}(x).

- Nonwandering if for any neighborhood UU of xx there is nn\in\mathbb{N} such that fn(U)Uf^{n}(U)\cap U\neq\emptyset.

We denote by P(f)(f), AP(f)(f), R(f)(f), Ω(f)\Omega(f) and Λ(f)\Lambda(f) the sets of periodic points, almost periodic points, recurrent points, nonwandering points and the union of all ω\omega-limit sets of ff, respectively. Define the space

X=nfn(X)X_{\infty}=\displaystyle\bigcap_{n\in\mathbb{N}}f^{n}(X). From the definition, we have the following inclusions:

P(f)AP(f)R(f)Λ(f)Ω(f)X.\textrm{P}(f)\subseteq\textrm{AP}(f)\subseteq\textrm{R}(f)\subseteq\Lambda(f)\subseteq\Omega(f)\subseteq X_{\infty}.

In the definitions below, we use the terminology from Nadler [27] and Kuratowski [23].

Definition 2.3.

(([23, page 131])) Let X,YX,~{}Y be two topological spaces. A continuous map f:XYf:X\longrightarrow Y is said to be monotone if for any connected subset CC of YY, f1(C)f^{-1}(C) is connected.

When ff is closed and onto, Definition 2.3 is equivalent to that the preimage of any point by ff is connected (cf. [23, page 131]). In particular, this holds if XX is compact and ff is onto. Notice that fnf^{n} is monotone for every nn\in\mathbb{N} when ff itself is monotone.

A continuum is a compact connected metric space. An arc II (resp. a circle) is any space homeomorphic to the compact interval [0,1][0,1] (resp. to the unit circle 𝕊1={z:|z|=1}\mathbb{S}^{1}=\{z\in\mathbb{C}:\ |z|=1\}). A space is called degenerate if it is a single point, otherwise; it is non-degenerate. By a graph GG, we mean a continuum which can be written as the union of finitely many arcs such that any two of them are either disjoint or intersect only in one or both of their endpoints.

A dendrite is a locally connected continuum which contains no circle. Every subcontinuum of a dendrite is a dendrite ([27, Theorem 10.10]) and every connected subset of DD is arcwise connected. A local dendrite is a continuum every point of which has a dendrite neighborhood. By ([23, Theorem 4, page 303]), a local dendrite is a locally connected continuum containing only a finite number of circles. As a consequence every subcontinuum of a local dendrite is a local dendrite ([23, Theorems 1 and 4, page 303]). Every graph and every dendrite is a local dendrite.

A regular curve is a continuum XX with the property that for each point xXx\in X and each open neighborhood VV of xx in XX, there exists an open neighborhood UU of xx in VV such that the boundary set U\partial U of UU is finite. Each regular curve is a 11-dimensional locally connected continuum. It follows that each regular curve is locally arcwise connected (see [23] and [27], for more details). In particular every local dendrite is a regular curve (cf. [23, page 303]).

A continuum XX is said to be finitely suslinean continuum provided that each infinite family of pairwise disjoint continua is null (i.e. for any disjoint open sets U,VU,V, only a finite number of elements of the family meet both UU and VV). Note that each regular curve is finitely suslinean (see [25]). A continuum is called hereditarily locally connected continuum, written hlc, provided that every subcontinuum of XX is locally connected. In particular, finitely suslinean continua and (hence regular curves) are hlc (see [23]). A continuum XX is said to be rational curve provided that each point xx of XX and each open neighborhood VV of xx in XX, there exists an open neighborhood UU of xx in VV such that the boundary set U\partial U of UU is at most countable. Clearly, regular curve are rational.

Let XX be a compact metric space. We denote by 2X2^{X} (resp. C(X)C(X)) the set of all non-empty compact subsets (resp. compact connected subsets) of XX. The Hausdorff metric dHd_{H} on 2X2^{X} (respectively C(X)C(X)) is defined as follows: dH(A,B)=max(supaAd(a,B),supbBd(b,A))d_{H}(A,B)=\max\Big{(}\sup_{a\in A}d(a,B),~{}\sup_{b\in B}d(b,A)\Big{)}, where A,B2XA,B\in 2^{X} (resp. C(X)C(X)). For xXx\in X and M2XM\in 2^{X}, d(x,M)=infyMd(x,y)d(x,M)=\inf_{y\in M}d(x,y). With this distance, (C(X),dH)(C(X),d_{H}) and (2X,dH)(2^{X},d_{H}) are compact metric spaces. Moreover if XX is a continuum, then so are 2X2^{X} and C(X)C(X) (see [27], for more details). Let f:XXf:X\to X be a continuous map of XX. We denote by 2f:2X2X,Af(A)2^{f}:2^{X}\to 2^{X},A\to f(A), called the induced map. Then 2f2^{f} is also a continuous self mapping of (2X,dH)(2^{X},d_{H}) (cf. [27]). For a subset AA of XX, we denote by diam(A)=supx,yAd(x,y)(A)=\sup_{x,y\in A}d(x,y) and card(A)\textrm{card}(A) the cardinality of AA. For any AC(X)A\in C(X), we denote by

Mesh(A)=sup{diam(C):C is a connected component of A}.\textrm{Mesh}(A)=\displaystyle\sup\{\textrm{diam}(C):\ C\textrm{ is a connected component of }A\}.

A family (Ai)iI(A_{i})_{i\in I} of subsets of XX is called a null family if for any infinite sequence (in)n0(i_{n})_{n\geq 0} of I,limn+diam(Ain)=0I,\;\displaystyle\lim_{n\to+\infty}\textrm{diam}(A_{i_{n}})=0. It is well known that each pairwise disjoint family of subcontinua of a regular curve is null (see [25]).

We recall some results which are needed for the sequel.

Proposition 2.4.

Let XX be a regular curve. Then for any ε>0\varepsilon>0 and for any family of pairwise disjoint subcontinua (Ai)iI(A_{i})_{i\in I} of XX, the set
{iI:diam(Ai)ε}\{i\in I:\mathrm{diam}(A_{i})\geq\varepsilon\} is finite. In particular if (An)n0(A_{n})_{n\geq 0} is a sequence of pairwise disjoint continua, then (An)n0(A_{n})_{n\geq 0} is a null family.

Definition 2.5.

((Weak incompressibility)) A set AXA\subset X is said to have the weak incompressibility property if for any proper closed subset FAF\subsetneqq A (i.e. FF is nonempty and distinct from AA), we have that Ff(AF)¯F\cap\overline{f(A\setminus F)}\neq\emptyset.

Notice that the term weak incompressibility seems to have appeared first in [7].

Lemma 2.6.

(([10, Lemma 3, page 71])) For any xXx\in X, ωf(x)\omega_{f}(x) has the weak incompressibility property.

Lemma 2.7.

Let AP(f)A\subset\mathrm{P}(f) be a closed invariant subset of XX with the weak incompressibility property. If some aAa\in A is an isolated point of AA, then A=Of(a)A=O_{f}(a).

Proof.

Assume that for some aA,{a}a\in A,\{a\} is an open subset of AA and Of(a)AO_{f}(a)\subsetneq A. Since fAf_{\mid A} is an homeomorphism, Of(a)O_{f}(a) is a finite open subset of AA, thus also a proper closed subset of AA. AA has the weak incompressibility property and Of(a)O_{f}(a) is a closed subset of AA, we get Of(a)f(AOf(a))¯O_{f}(a)\cap\overline{f(A\setminus O_{f}(a))}\neq\emptyset. Therefore Of(a)(AOf(a))O_{f}(a)\cap(A\setminus O_{f}(a))\neq\emptyset, which contradict the fact that ff is one to one on AA. ∎

Let ff be a monotone map on a regular curve XX. If MM is an infinite minimal set of ff, we call (M)={xX:ωf(x)=M}\mathcal{B}(M)=\{x\in X:\omega_{f}(x)=M\} the basin of attraction of MM.

Theorem 2.8.

[14] Let f:XXf:X\longrightarrow X be a regular curve monotone map. Then the following assertions hold:

  • (1)

    ωf(x)\omega_{f}(x) is a minimal set, for all xXx\in X.

  • (2)

    αf((xn)n0)\alpha_{f}((x_{n})_{n\geq 0}) is a minimal set any negative orbit (xn)n0(x_{n})_{n\geq 0} of xXx\in X_{\infty}. Moreover if xXP(f)x\in X_{\infty}\setminus\mathrm{P}(f), then αf(x)\alpha_{f}(x) is a minimal set and αf(x)=αf((xn)n0)\alpha_{f}(x)=\alpha_{f}((x_{n})_{n\geq 0}), for any negative orbit (xn)n0(x_{n})_{n\geq 0} of xx.

  • (3)

    For every infinite minimal set MM, (M)\mathcal{B}(M) is a closed subset of XX.

  • (4)

    For every xXx\in X_{\infty}, if ωf(x)\omega_{f}(x) is infinite, then αf(x)=ωf(x)\alpha_{f}(x)=\omega_{f}(x).

  • (5)

    AP(f)=Λ(f)=R(f)\mathrm{AP}(f)=\Lambda(f)=\mathrm{R}(f).


3. Nonwandering sets of monotone maps on regular curves

The aim of this section is to prove the following theorem which extends [3, Theorem 1.1] and [29, Theorem 2.1].

Theorem 3.1.

Let XX be a regular curve and ff a monotone self mapping of XX. Then Ω(f)=R(f)=Λ(f)=AP(f)\Omega(f)=\mathrm{R}(f)=\Lambda(f)=\mathrm{AP}(f).

Proof.

Following Theorem 2.8, (5), it suffices to prove that Ω(f)R(f)\Omega(f)\subset\textrm{R}(f). Assume that there exists xΩ(f)R(f)x\in\Omega(f)\setminus\textrm{R}(f). We distinguish two cases:

Case 1: xP(f)¯x\notin\overline{P(f)}. In this case, we follow similarly the proof given in ([29, Theorem 2.1]). Let VV be an open neighborhood of xx such that V¯P(f)=\overline{V}\cap\textrm{P}(f)=\emptyset. As xΩ(f)x\in\Omega(f), there is a sequence (xk)k0(x_{k})_{k\geq 0} in XX converging to xx and a sequence of positive integers (nk)k0(n_{k})_{k\geq 0} such that fnk(xk)f^{n_{k}}(x_{k}) converges to xx. Since xR(f)x\notin\textrm{R}(f), xωf(x)x\notin\omega_{f}(x). Then we can find an open neighborhood U1VU_{1}\subset V of xx with finite boundary and an open neighborhood U2U_{2} of ωf(x)\omega_{f}(x) such that U1U2=U_{1}\cap U_{2}=\emptyset. Since XX is locally connected, so by ([23, Theorem 4, page 257]), for kk large enough, we can find a sequence of arcs (Ik)k0U1(I_{k})_{k\geq 0}\subset U_{1} joining xkx_{k} and xx such that limk+Ik={x}\displaystyle\lim_{k\to+\infty}I_{k}=\{x\} (with respect to the Hausdorff metric). Thus fnk(Ik)f^{n_{k}}(I_{k}) will meets U1U_{1} and U2U_{2}, for kk large enough and so it meets the boundary U1\partial U_{1} in a point bkb_{k}, for kk large enough. As U1\partial U_{1} is finite, one can assume that bk=bb_{k}=b for infinitely many kk. Therefore fnk(b)Ikf^{-n_{k}}(b)\cap I_{k}\neq\emptyset, for infinitely many kk. It follows that xαf(b)x\in\alpha_{f}(b). As bP(f)b\notin\textrm{P}(f), then αf(b)\alpha_{f}(b) is a minimal set of ff (Theorem 2.8). So ωf(x)=αf(b)\omega_{f}(x)=\alpha_{f}(b) and hence xωf(x)x\in\omega_{f}(x). A contradiction.

Case 2: xP(f)¯x\in\overline{P(f)}. Let (xn)n0P(f)(x_{n})_{n\geq 0}\subset\textrm{P}(f) be a sequence converging to xx and set pn=Per(xn)p_{n}=\textrm{Per}(x_{n}) the period of xnx_{n}, n0n\geq 0. Then (pn)n0(p_{n})_{n\geq 0} in unbounded: otherwise, xP(f)R(f)x\in\textrm{P}(f)\subset\textrm{R}(f), a contradiction. So we can assume that (pn)n0(p_{n})_{n\geq 0} goes to infinity. Since αf(x)\alpha_{f}(x) is a minimal set, xαf(x)x\notin\alpha_{f}(x) (because otherwise, xωf(x)x\in\omega_{f}(x), a contradiction). Now as in Case 1, let U2U_{2} be an open neighborhood of αf(x)\alpha_{f}(x) and U1U_{1} an open neighborhood of xx with finite boundary and disjoint from U2U_{2}. Let (Ik)k0(I_{k})_{k\geq 0} be a sequence of arcs joining xx and xkx_{k} such that limk+Ik={x}\displaystyle\lim_{k\to+\infty}I_{k}=\{x\} (with respect to the Hausdorff metric). Thus fpk(Ik)f^{-p_{k}}(I_{k}) meets U1\partial U_{1} infinitely many times and thus there exists bU1b\in\partial U_{1} such that fpk(b)Ikf^{p_{k}}(b)\in I_{k}. This implies that xωf(b)x\in\omega_{f}(b). As ωf(b)\omega_{f}(b) is a minimal, so xR(f)x\in\textrm{R}(f). A contradiction. ∎

Corollary 3.2.

If Ω(f)\Omega(f) is finite or countable, then Ω(f)=P(f)\Omega(f)=\mathrm{P}(f).

Proof.

Since Ω(f)=AP(f)\Omega(f)=\textrm{AP}(f) (Theorem 3.1), so AP(f)\textrm{AP}(f) is at most countable and hence every minimal set for (X,f)(X,f) is a periodic orbit. Therefore AP(f)=P(f)\textrm{AP}(f)=\textrm{P}(f) and then Corollary 3.2 follows. ∎

In [11], Coven and Nitecki have shown that for continuous maps ff of the closed interval [0,1][0,1], Ω(f)={x[0,1]:xαf(x)}.\Omega(f)=\{x\in[0,1]:x\in\alpha_{f}(x)\}. Later, it is extended to graph maps in [26, Corollary 1]. However, for dendrite maps, it does not holds (see [32]). The following theorem extend the above result to monotone maps on regular curves.

Theorem 3.3.

Let XX be a regular curve and ff a self monotone mapping of XX. Then Ω(f)={xX:xαf(x)}\Omega(f)=\{x\in X:x\in\alpha_{f}(x)\}.

Proof.

Let xΩ(f)x\in\Omega(f). By Theorem 3.1, xωf(x)x\in\omega_{f}(x) and hence there exists a negative orbit (xn)n0(x_{n})_{n\geq 0} of xx such that (xn)n0ωf(x)(x_{n})_{n\geq 0}\subset\omega_{f}(x). By minimality of ωf(x)\omega_{f}(x), we have xωf(x)=αf((xn)n0)αf(x)x\in\omega_{f}(x)=\alpha_{f}((x_{n})_{n\geq 0})\subset\alpha_{f}(x). Conversely, let xXx\in X such that xαf(x)x\in\alpha_{f}(x). We can assume that xP(f)x\notin\textrm{P}(f) (if xP(f)x\in\textrm{P}(f) then xΩ(f)x\in\Omega(f) and we are done). So by Theorem 2.8, αf(x)\alpha_{f}(x) is a minimal set and xαf(x)x\in\alpha_{f}(x). Therefore ωf(x)=αf(x)\omega_{f}(x)=\alpha_{f}(x). Hence xωf(x)x\in\omega_{f}(x) and so xR(f)=Ω(f)x\in\textrm{R}(f)=\Omega(f). ∎

Corollary 3.4.

Let XX be a regular curve and ff a self monotone mapping of XX. If xΩ(f)x\in\Omega(f), then :
(i) ωf(x)αf(x)\omega_{f}(x)\subset\alpha_{f}(x).
(ii) If xP(f)x\notin\mathrm{P}(f), then ωf(x)=αf((xn)n0)=αf(x)\omega_{f}(x)=\alpha_{f}((x_{n})_{n\geq 0})=\alpha_{f}(x), for any negative orbit (xn)n0(x_{n})_{n\geq 0} of xx.

Proof.

(i) Since xΩ(f)x\in\Omega(f), so xαf(x)x\in\alpha_{f}(x) (Theorem 3.3) and hence ωf(x)αf(x)\omega_{f}(x)\subset\alpha_{f}(x) (since by Proposition 2.2, αf(x)\alpha_{f}(x) is closed and ff-invariant).
(ii) If xP(f)x\notin\textrm{P}(f), then αf((xn)n)=αf(x)\alpha_{f}((x_{n})_{n})=\alpha_{f}(x) (Theorem 2.8). Again by Theorem 2.8 and from (i), we get αf(x)=ωf(x)\alpha_{f}(x)=\omega_{f}(x). ∎

In the following example, we show that Theorem 3.1 cannot be extended to rational curves; we construct a self monotone map ff on a rational curve DD (it is in fact a dendroid) such that the inclusion is strict: R(f)Ω(f)\textrm{R}(f)\subsetneq\Omega(f).

Notice also that Theorem 3.1 does not hold even for continuous maps on intervals; for instance; we may extend the map g:LLg:L\to L (defined in Example 3.5) into a continuous map hh of [0,1][0,1] so that T1Λ(h)Ω(h)T_{1}\in\Lambda(h)\subset\Omega(h) but T1R(h)T_{1}\notin\textrm{R}(h).

Example 3.5.

The idea of the construction consists to define a continuous map gg on a countable compact set LL satisfying R(g)Ω(g)\mathrm{R}(g)\subsetneq\Omega(g) and then extends it to a monotone map ff defined on a dendroid DD.

\bullet Let σ\sigma be the shift map on the Cantor set {0,1}\{0,1\}^{\mathbb{N}} endowed with the metric dd defined as follows: For x,y{0,1}x,y\in\{0,1\}^{\mathbb{N}}, d(x,y)=2N(x,y)d(x,y)=2^{-N(x,y)}, where N(x,y)=min{n:xnyn}N(x,y)=\min\{n\in\mathbb{N}:x_{n}\neq y_{n}\} with N(x,x)=+,x{0,1}N(x,x)=+\infty,\forall x\in\{0,1\}^{\mathbb{N}}. We denote by:

We let T0=0¯T_{0}=\overline{0} and Ti=(0000001ithposition0), for i>0T_{i}=(000000\dots\underbrace{1}_{i^{th}-\textrm{position}}0\dots),\textrm{ for }~{}i>0. Set

𝒵=(101001000100nzero1)\mathcal{Z}=(\underbrace{10}\underbrace{100}\underbrace{1000}\dots\underbrace{10\dots 0}_{n-zero}1\dots),

O(𝒵)={Ti=(000𝒵ithposition):i1}O_{-}(\mathcal{Z})=\{T_{-i}=(000\dots\underbrace{\mathcal{Z}}_{i^{th}-\textrm{position}}):i\geq 1\},

L=Oσ(𝒵){Ti:i0}O(𝒵)L=O_{\sigma}(\mathcal{Z})\cup\{T_{i}:i\geq 0\}\cup O_{-}(\mathcal{Z}).

Observe that (Ti)i1(T_{-i})_{i\geq 1} converges to T0T_{0}. Let us show that ωσ(Z)={Ti:i0}\omega_{\sigma}(Z)=\{T_{i}:i\geq 0\}.

For each nn\in\mathbb{N}, we let kn=k=1n1k+n=n(n1)2+nk_{n}=\sum_{k=1}^{n-1}k+n=\frac{n(n-1)}{2}+n. It is clear that (kn)n(k_{n})_{n\in\mathbb{N}} is an increasing sequence. Observe that σkn(𝒵)\sigma^{k_{n}}(\mathcal{Z}) gives the first apparition of the number one after nn zeros. Then for each nn\in\mathbb{N}, d(σkn(𝒵),0¯)<2nd(\sigma^{k_{n}}(\mathcal{Z}),\bar{0})<2^{-n} and so 0¯ωσ(𝒵)\bar{0}\in\omega_{\sigma}(\mathcal{Z}). Let i2i\geq 2 and suppose that nin\geq i. Then σkn+(ni+1)(𝒵)\sigma_{k_{n}+(n-i+1)}(\mathcal{Z}) gives i1i-1 zero before the first apparition of one. Then d(σkn+(ni+1)(𝒵),Ti)<2(n+1)d(\sigma^{k_{n}+(n-i+1)}(\mathcal{Z}),T_{i})<2^{-(n+1)}, so Tiωσ(Z),i2T_{i}\in\omega_{\sigma}(Z),\forall i\geq 2. For i=1i=1, we have T1=(100)=σ(T2)ωσ(𝒵)T_{1}=(100\dots)=\sigma(T_{2})\in\omega_{\sigma}(\mathcal{Z}). In result, {0¯}{Ti,i}={Ti:i0}ωσ(𝒵)\{\bar{0}\}\cup\{T_{i},\;i\in\mathbb{N}\}=\{T_{i}:i\geq 0\}\subset\omega_{\sigma}(\mathcal{Z}). Conversely, let y=(yn)nωσ(𝒵)y=(y_{n})_{n\in\mathbb{N}}\in\omega_{\sigma}(\mathcal{Z}) and suppose that there exist i,ji,j\in\mathbb{N} such that i<ji<j and yi=yj=1y_{i}=y_{j}=1, let (ns)s0(n_{s})_{s\geq 0} be an increasing sequence of integer be such that lims+σns(𝒵)=y\lim_{s\to+\infty}\sigma^{n_{s}}(\mathcal{Z})=y. Then there exists s1s_{1}\in\mathbb{N} such that ss1\forall s\geq s_{1}, we have ns>kj+1n_{s}>k_{j+1}, so d(σns(𝒵),y)2j,ss1d(\sigma^{n_{s}}(\mathcal{Z}),y)\geq 2^{-j},\forall s\geq s_{1} ( we have only a single one in the first block with length jj in σns(𝒵)\sigma^{n_{s}}(\mathcal{Z}) but, we have two ones in the first block with length jj in yy). So every point of ωσ(𝒵)\omega_{\sigma}(\mathcal{Z}) contains at most only one. Consequently, ωσ(𝒵)={Ti,i}{0¯}={Ti:i0}\omega_{\sigma}(\mathcal{Z})=\{T_{i},\;i\in\mathbb{N}\}\cup\{\bar{0}\}=\{T_{i}:i\geq 0\}.

We conclude that LL is a countable compact set. We let g=σLg=\sigma_{\mid L}. Then g(Ti)=Ti1g(T_{i})=T_{i-1}, for any i1i\geq 1. Pick a homeomorphic copy of LL in [0,1]×{1}[0,1]\times\{1\}. For ii\in\mathbb{Z}, let IiI_{i} be an arc joining TiT_{i} and T0T_{0} so that S:=iIiS:=\displaystyle\bigcup_{i\in\mathbb{Z}}I_{i} is an infinite star centered at T0T_{0}. in particular I0={T0}I_{0}=\{T_{0}\}.

For each kk\in\mathbb{Z}, let iki_{k}\in\mathbb{Z} such that d(gk(𝒵),S)=d(gk(𝒵),Iik)d(g^{k}(\mathcal{Z}),S)=d(g^{k}(\mathcal{Z}),I_{i_{k}}) and let JkJ_{k} be an arc joining T0T_{0} and gk(𝒵)g^{k}(\mathcal{Z}) which is included in the closed ball BF(Iik,d(gk(𝒵),Iik))B_{F}(I_{i_{k}},d(g^{k}(\mathcal{Z}),I_{i_{k}})) of the plan. We may assume that for any kk\in\mathbb{Z}, SJk={T0}S\cap J_{k}=\{T_{0}\} and for any kl,JkJl={T0}k\neq l,\;J_{k}\cap J_{l}=\{T_{0}\}. Let D=S(k0Jk)D=S\bigcup\big{(}\displaystyle\bigcup_{k\geq 0}J_{k}\big{)}. Clearly DD is a countable union of arcs, so by ([23, Theorem 6, page 286]) DD is a rational curve which is a dendroid.

\bullet We extend gg into a map ff on DD so that f(Jk)=Jk+1f(J_{k})=J_{k+1} and f|Jkf_{|J_{k}} is affine, for each kk\in\mathbb{Z} and f(Ii)=Ii1f(I_{i})=I_{i-1} for i1i\geq 1. Then f(I1)={T0}f(I_{1})=\{T_{0}\} and fL=gf_{\mid L}=g. It is clear that ff is a pointwise monotone map on DD and onto, thus monotone. Moreover R(f)Ω(f)\textrm{R}(f)\subsetneq\Omega(f), since for instance T1Λ(f)Ω(f)T_{1}\in\Lambda(f)\subset\Omega(f) but T1R(f)T_{1}\notin\textrm{R}(f).

Refer to caption
Figure 1. The map ff on the dendroid DD

4. The space of minimal sets with respect to the Hausdorff metric

The main result of this section is to prove the following theorem.

Theorem 4.1.

Let XX be a regular curve and ff a monotone self mapping of XX. Then any limit (with respect to the Hausdorff metric) of a sequence of minimal sets (Mn)n0(M_{n})_{n\geq 0} is a minimal set.

Proof.

Let (Mn)n0(M_{n})_{n\geq 0} be a sequence a sequence of minimal sets converging to some MXM\subset X. Then f(M)=Mf(M)=M. Since each MnR(f)M_{n}\subset\textrm{R}(f), so by Theorem 3.1, MR(f)=AP(f)M\subset\textrm{R}(f)=\textrm{AP}(f). Therefore M=iIMiM=\displaystyle\bigcup_{i\in I}M_{i}, where MiM_{i} is a minimal set for any iIi\in I. Assume that MM is not a minimal set. Then card(I)>1\textrm{card}(I)>1.

Claim: MP(f)M\subset\textrm{P}(f).  Suppose that MP(f)M\nsubseteq\textrm{P}(f), so there exists i0Ii_{0}\in I such that Mi0M_{i_{0}} is infinite. Set P=Mi0P=M_{i_{0}} and N=Mj0N=M_{j_{0}}, where j0i0j_{0}\neq i_{0}. Then PP and NN are disjoint minimal sets and so does (P)\mathcal{B}(P) and NN. By Theorem 2.8, (P)\mathcal{B}(P) is a closed set. So let UU an open set with finite boundary U\partial U of cardinality kk such that (P)U\mathcal{B}(P)\subset U and NU¯=N\cap\overline{U}=\emptyset.
Recall that PNM=limn+MnP\cup N\subset M=\displaystyle\lim_{n\to+\infty}M_{n}, so fix some yP,zNy\in P,\;z\in N and let yn,znMny_{n},\ z_{n}\in M_{n} such that (yn)n0(y_{n})_{n\geq 0} (resp. (zn)n0(z_{n})_{n\geq 0}) converges to yy (resp. zz). Let (In)n0(I_{n})_{n\geq 0} be a sequence of arcs joining yy and yny_{n} so that {y}=limn+In\{y\}=\displaystyle\lim_{n\to+\infty}I_{n}.
Since MnM_{n} is a minimal set, we can find sn0s_{n}\geq 0 such that d(fsn(yn),zn)1nd(f^{-s_{n}}(y_{n}),z_{n})\leq\frac{1}{n}, for every n1n\geq 1. Hence for nn large enough, fsn(yn)Uf^{-s_{n}}(y_{n})\nsubseteq U and fsn(In)Uf^{-s_{n}}(I_{n})\nsubseteq U. On the other hand, fsn(In)f^{-s_{n}}(I_{n}) meets PP at some point of fsn(y)f^{-s_{n}}(y). As ff is monotone, thus for nn large enough, fsn(In)Uf^{-s_{n}}(I_{n})\cap\partial U\neq\emptyset. As U\partial U is finite, then there exists zUz\in\partial U such that zfsn(In)z\in f^{-s_{n}}(I_{n}), for infinitely many nn. Hence yωf(z)y\in\omega_{f}(z) (since d(fsn(z),y)diam(In)+d(yn,y)d(f^{s_{n}}(z),y)\leq\textrm{diam}(I_{n})+d(y_{n},y)). By minimality of ωf(z)\omega_{f}(z) (see Theorem 2.8), we get ωf(z)=P\omega_{f}(z)=P. Hence z(P)Uz\in\mathcal{B}(P)\cap\partial U. A contradiction with (P)U\mathcal{B}(P)\subset U. The proof is complete.

Proof of Theorem 4.1. As proven by the claim above, we have MP(f)M\subset\textrm{P}(f). Since any minimal set MnM_{n} has the weakly incompressibility property (Lemma 2.6), so does its limit MM by [4, Proposition 3.1]. Since MP(f)M\subset\textrm{P}(f), it is uncountable, otherwise, MM will have an isolated point aa, so by Lemma 2.7, M=Of(a)M=O_{f}(a) is a periodic orbit. A contradiction. From M=nFix(fn)MM=\displaystyle\cup_{n\in\mathbb{N}}\textrm{Fix}(f^{n})\cap M, there exist N>0N>0 and an infinite sequence (zn)n0(z_{n})_{n\geq 0} in Fix(fN)M(f^{N})\cap M with disjoint orbits converging to some zFix(fN)Mz\in\textrm{Fix}(f^{N})\cap M. As Of(z0)Of(z)=O_{f}(z_{0})\cap O_{f}(z)=\emptyset, there exists an open set VzV_{z} with finite boundary such that Of(z)VzO_{f}(z)\subset V_{z} and Of(z0)Vz¯=O_{f}(z_{0})\cap\overline{V_{z}}=\emptyset. Since (zn)n0(z_{n})_{n\geq 0} converges to zz, there is some pp\in\mathbb{N} such that for any np,Of(zn)Vzn\geq p,\;O_{f}(z_{n})\subset V_{z}. For each npn\geq p, let (In,q)q0(I_{n,q})_{q\geq 0} be a sequence of arcs joining znz_{n} and zn,qMqz_{n,q}\in M_{q} so that (In,q)q0(I_{n,q})_{q\geq 0} converges to {zn}\{z_{n}\} as done in the proof of the claim above. As MqM_{q} is minimal and MqUM_{q}\nsubseteq U, we can find sq>0s_{q}>0 such that fsq(zn,q)Uf^{-s_{q}}(z_{n,q})\nsubseteq U. Recall that Of(zn)UO_{f}(z_{n})\subset U, so fsq(In,q)Uf^{-s_{q}}(I_{n,q})\cap\partial U\neq\emptyset, for any q0q\geq 0. By the same argument as in the proof of the claim above, we can find, for each npn\geq p, bnUb_{n}\in\partial U such that ωf(bn)=Of(zn)\omega_{f}(b_{n})=O_{f}(z_{n}). This leads to a contradiction since {Of(zn):np}\{O_{f}(z_{n}):n\geq p\} is an infinite family of disjoint periodic orbits. ∎

5. On the continuity of limit maps ωf\omega_{f} and αf\alpha_{f}

In this section, we shall investigate the continuity of the limit maps:
ωf:X2X;xωf(x)\omega_{f}:X\to 2^{X};~{}x\to\omega_{f}(x) and αf:X2X;xαf(x)\alpha_{f}:X_{\infty}\to 2^{X};~{}x\to\alpha_{f}(x).

Theorem 5.1.

The maps ωf\omega_{f} and αf\alpha_{f} are continuous everywhere except may be at the periodic points of ff.

We use the following lemmas.

Lemma 5.2.

(([13, Lemma 4.2])) Let XX be a hereditarily locally connected continuum, FXF\subsetneq X a closed subset and (On)n0(O_{n})_{n\geq 0} a sequence of open subsets of XX such that F=n0On¯F=\displaystyle\bigcap_{n\geq 0}\overline{O_{n}}. Then limn+Mesh(OnF)=0\displaystyle\lim_{n\rightarrow+\infty}\mathrm{Mesh}(O_{n}\setminus F)=0.

Lemma 5.3.

The restriction map (ωf)|Ω(f)(\omega_{f})_{|\Omega(f)} is continuous.

Proof.

Let (xn)n0(x_{n})_{n\geq 0} a sequence of Ω(f)\Omega(f) converging to some xΩ(f)x\in\Omega(f). By Theorem 3.1, xnωf(xn)x_{n}\in\omega_{f}(x_{n}) and by Theorem 2.8, (ωf(xn))n0(\omega_{f}(x_{n}))_{n\geq 0} is a sequence of minimal set, thus by Theorem 4.1, any limit point of that sequence should be a minimal set. Observe that xx belongs to any limit point of the sequence (ωf(xn))n0(\omega_{f}(x_{n}))_{n\geq 0}, therefore any limit point of the sequence (ωf(xn))n0(\omega_{f}(x_{n}))_{n\geq 0} is a minimal set that contains xx, therefore it has to be ωf(x)\omega_{f}(x).

Let (X,f)(X,f) be a dynamical system. For xXx\in X, we denote by
Sx=i0fi(x)S_{x}=\underset{i\geq 0}{\cup}f^{-i}(x).

Lemma 5.4.

Let x,yXx,y\in X such that Of(x)O_{f}(x) and Of(y)O_{f}(y) are two disjoint orbits. Then SxS_{x} and SyS_{y} are disjoint.

Proof.

Assume that SxSyS_{x}\cap S_{y}\neq\emptyset. Then we can find two positive integers mnm\leq n and some zfm(x)fn(y)z\in f^{-m}(x)\cap f^{-n}(y). Therefore fn(z)Of(x)Of(y)f^{n}(z)\in O_{f}(x)\cap O_{f}(y), which will leads to a contradiction. ∎

Lemma 5.5.

Let XX be a regular curve and ff a monotone self mapping of XX. The limit maps ωf\omega_{f} and αf\alpha_{f} are continuous at any point of XΩ(f)X\setminus\Omega(f).

Proof.

First note that X\R(f)=X\Ω(f)X\backslash\textrm{R}(f)=X\backslash\Omega(f) is a open set disjoint from P(f)¯\overline{P(f)}. Thus for any point aX\Ω(f)a\in X\backslash\Omega(f), αf(a)\alpha_{f}(a) and ωf(a)\omega_{f}(a) are minimal sets. Let xX\R(f)x\in X\backslash\textrm{R}(f):
\bullet Continuity of ωf\omega_{f}: Assume that ωf\omega_{f} is not continuous at xx and set M=ωf(x)M=\omega_{f}(x), then we can find a sequence (xn)n0(x_{n})_{n\geq 0} that converges to xx such that ωf(xn)\omega_{f}(x_{n}) converges to AMA\neq M. By Theorem 4.1, AA is a minimal set, thus AM=A\cap M=\emptyset. Let m0m\geq 0 and VA,mB(A,1m)V_{A,m}\subset B(A,\frac{1}{m}) be an open set with finite boundary such that MVA,m¯=M\cap\overline{V_{A,m}}=\emptyset. Let (In)n0(I_{n})_{n\geq 0} be a sequence of arcs joining xx and xnx_{n}, where (In)n0(I_{n})_{n\geq 0} converge to {x}\{x\} with respect to the Hausdorff metric. Recall that A=limn+ωf(xn)A=\displaystyle\lim_{n\to+\infty}\omega_{f}(x_{n}), so for nn large enough we can find some pnp_{n} such that fpn(xn)VA,mf^{p_{n}}(x_{n})\in V_{A,m} and fpn(x)VA,mf^{p_{n}}(x)\notin V_{A,m}. Then fpn(In)f^{p_{n}}(I_{n}) will meets VA,m\partial V_{A,m} and so xαf(bm)x\in\alpha_{f}(b_{m}) for some bmVA,mb_{m}\in\partial V_{A,m}. Since xΩ(f)x\notin\Omega(f), we conclude that αf(bm)\alpha_{f}(b_{m}) is not a minimal set, thus By Theorem 2.8, bmP(f)b_{m}\in\textrm{P}(f).

Claim: The sequence (bm)m0(b_{m})_{m\geq 0} satisfies the follow properties:
(i) For any m0,bmP(f)m\geq 0,\;b_{m}\in\textrm{P}(f) and xαf(bm)x\in\alpha_{f}(b_{m}).
(ii) The sequence (Of(bm))m0(O_{f}(b_{m}))_{m\geq 0} converges to AA.
(iii) By passing to a subsequence if needed, the family (Of(bm))m0(O_{f}(b_{m}))_{m\geq 0} is pairwise disjoint.

Proof of the claim: (i) is already proven above, for the prove of (ii) observe that any limit point of (bm)m0(b_{m})_{m\geq 0} is a point of AA and since {bm:m0}AΩ(f)\{b_{m}:\;m\geq 0\}\cup A\subset\Omega(f), so by Lemma 5.3 ends the proof of (ii).
(iii) Clearly for any m0m\geq 0, Of(bm)A=O_{f}(b_{m})\cap A=\emptyset. Let bm0=b0b_{m_{0}}=b_{0} and O1O_{1} be an open neighborhood of AA such that Of(bm0)O1=O_{f}(b_{m_{0}})\cap O_{1}=\emptyset. As (bm)m0(b_{m})_{m\geq 0} converges to AA we can find m1>0m_{1}>0 such that bm1O1b_{m_{1}}\in O_{1}. Thus Of(b0)O_{f}(b_{0}) and Of(b1)O_{f}(b_{1}) are disjoint.
For N0,(Of(bmi))0iNN\geq 0,\;(O_{f}(b_{m_{i}}))_{0\leq i\leq N} is defined and pairwise disjoint. Let ON+1O_{N+1} an open neighborhood of AA such that
0iNOf(bmi)ON+1=\displaystyle\bigcup_{0\leq i\leq N}O_{f}(b_{m_{i}})\cap O_{N+1}=\emptyset. We can find mN+1>mNm_{N+1}>m_{N} such that bmN+1ON+1b_{m_{N+1}}\in O_{N+1} and then (Of(bni))0iN+1(O_{f}(b_{n_{i}}))_{0\leq i\leq N+1} is pairwise disjoint. We have defined by indication the subsequence (bmi)i0(b_{m_{i}})_{i\geq 0} so that (Of(bmi))i0(O_{f}(b_{m_{i}}))_{i\geq 0} are pairwise disjoint. This ends the proof of the claim.

Now xR(f)x\notin\textrm{R}(f), so xAx\notin A. Let VV be an open neighborhood of AA such that xV¯x\notin\overline{V} and set k=card(V)k=\textrm{card}(\partial V). By (iii) there exist N0N\geq 0 such that for any mN,ωf(bm)=Of(bm)Vm\geq N,\;\omega_{f}(b_{m})=O_{f}(b_{m})\subset V. By (i), we may find for each m0m\geq 0 some sm0s_{m}\geq 0 such that fsm(bm)Vf^{-s_{m}}(b_{m})\nsubseteq V. Fix Nm0<m2<mkN\leq m_{0}<m_{2}<\dots m_{k}. By Lemma 5.4 and monotonicity of ff, the family (fsmi(bmi))0ik(f^{-s_{m_{i}}}(b_{m_{i}}))_{0\leq i\leq k} is a pairwise disjoint family of connected subset of XX. Observe that for each 0iN,fsmi(bmi)V0\leq i\leq N,\;f^{-s_{m_{i}}}(b_{m_{i}})\nsubseteq V and fsmi(bmi)f^{-s_{m_{i}}}(b_{m_{i}}) meets VV at some point of Of(bmi)O_{f}(b_{m_{i}}). Therefore for each 0iN,fsmi(bmi)0\leq i\leq N,\;f^{-s_{m_{i}}}(b_{m_{i}}) meets V\partial V. This leads to a contradiction.

\bullet Continuity of αf\alpha_{f}.

Set M=αf(x)M=\alpha_{f}(x). By Theorem 2.8, MM is a minimal set, assume that (αf(xn))n0(\alpha_{f}(x_{n}))_{n\geq 0} converges to LML\neq M, then LML\nsubseteq M, so let tLMt\in L\setminus M.
Let VtV_{t} be an open neighborhood of tt with finite boundary such that MVt¯=M\cap\overline{V_{t}}=\emptyset and (In)n0(I_{n})_{n\geq 0} a sequence of arcs joining xx and xnx_{n} converging to {x}\{x\} with respect to the Hausdorff metric. Since tL=limn+αf(xn)t\in L=\displaystyle\lim_{n\to+\infty}\alpha_{f}(x_{n}), for nn large enough, we can find some pn0p_{n}\geq 0 such that fpn(xn)Vtf^{-p_{n}}(x_{n})\subset V_{t} and fpn(x)Vtf^{-p_{n}}(x)\nsubseteq V_{t}. Then by monotonicity of ff fpn(In)f^{-p_{n}}(I_{n}) will meets Vt\partial V_{t} and thus xωf(b)x\in\omega_{f}(b) for some bVtb\in\partial V_{t}. By Theorem 3.1, ωf(b)R(f)\omega_{f}(b)\subset\textrm{R}(f), hence xR(f)x\in\textrm{R}(f). A contradiction. ∎

Proof of Theorem 5.1. Let xXP(f)x\in X\setminus\textrm{P}(f). If xΩ(f)x\notin\Omega(f), then by Lemma 5.5, ωf\omega_{f} is continuous at xx. So assume that xΩ(f)\P(f)x\in\Omega(f)\backslash\textrm{P}(f) and set M=ωf(x)M=\omega_{f}(x). Then xMx\in M and MM is an infinite minimal set.

\bullet Continuity of αf\alpha_{f} on XP(f)X\setminus\textrm{P}(f): By Theorem 2.8, αf(x)=M\alpha_{f}(x)=M. Assume that there is some sequence (xn)n0(x_{n})_{n\geq 0} of XX converging to xx such that αf(xn)\alpha_{f}(x_{n}) converges to LL and LML\neq M. By minimality of MM, LML\nsubseteq M. Assume that there exists yLMy\in L\setminus M. We distinguish two cases:

Case 1: For infinitely many n0n\geq 0, xnP(f)x_{n}\in\textrm{P}(f). Since xx in non periodic point, so per((xn))n(x_{n}))_{n} is unbounded. By choosing (xn))n0(x_{n}))_{n\geq 0} with pairwise distinct periods per((xn))n(x_{n}))_{n}, one can assume that (Of(xn))n0(O_{f}(x_{n}))_{n\geq 0} is pairwise disjoint. Therefore by Lemma 5.4, the family (Sxn)n0(S_{x_{n}})_{n\geq 0} is pairwise disjoint. Let now Vy,VMV_{y},V_{M} be two disjoint open neighborhoods of y,My,M with finite boundary. By Lemma 5.3, the restriction map (ωf)|Ω(f)(\omega_{f})_{|\Omega(f)} is continuous. Therefore as xR(f)\P(f)x\in\textrm{R}(f)\backslash\textrm{P}(f), we conclude that the sequence Of(xn))n0O_{f}(x_{n}))_{n\geq 0} converges to ωf(x)=M\omega_{f}(x)=M. So one can assume that for any n0n\geq 0, we have Of(xn)VMO_{f}(x_{n})\subset V_{M}. As ylimn+αf(xn)y\in\displaystyle\lim_{n\to+\infty}\alpha_{f}(x_{n}), then for each n0n\geq 0, one can find mn>0m_{n}>0 such that fmn(xn)Vyf^{-m_{n}}(x_{n})\cap V_{y}\neq\emptyset. So let Smn,nS_{m_{n},n} be the connected component of SnS_{n} containing fmn(xn)f^{-m_{n}}(x_{n}). Hence (Smn,n)n0(S_{m_{n},n})_{n\geq 0} is a family of pairwise disjoint connected sets each of which meets VMV_{M} and VyV_{y} (since xnx_{n} is a periodic point such that Of(xn)VMO_{f}(x_{n})\subset V_{M}). This leads to a contradiction since (VM)\partial(V_{M}) is finite.

Case 2: For nn large enough, xnP(f)x_{n}\notin\textrm{P}(f). In this case, by Theorem 2.8, αf(xn)\alpha_{f}(x_{n}) is a minimal set. Thus by Theorem 4.1, LL is also a minimal set and then ML=M\cap L=\emptyset. So let VMV_{M} and VLV_{L} be two disjoint neighborhoods of M,LM,L with finite boundary and set k=card((VM))k=\textrm{card}(\partial(V_{M})). As (xn)n0(x_{n})_{n\geq 0} converges to xx, then we can find k+1k+1 pairwise disjoint arcs I0,I1,,IkI_{0},I_{1},\dots,I_{k}, each IjI_{j} joins fj(x)f^{j}(x) and fj(xN)f^{j}(x_{N}), for some NN large enough and satisfying αf(xN)VL\alpha_{f}(x_{N})\subset V_{L}. Then we can find s0s\geq 0 such that for any isi\geq s, fi(xN)VLf^{-i}(x_{N})\subset V_{L}. Consider the family (f(s+k)(Ij))0jk(f^{-(s+k)}(I_{j}))_{0\leq j\leq k}, which is a family of pairwise disjoint connected sets, each of which meets VLV_{L} at f(s+k)(xN)f^{-(s+k)}(x_{N}) and meets VMV_{M} at some point of f(s+k)(fj(x))f^{-(s+k)}(f^{j}(x)), this will lead to a nonempty intersection with (VM)\partial(V_{M}) for each 0jk0\leq j\leq k. A contradiction.
This ends the proof of the continuity of αf\alpha_{f}.

\bullet Continuity of ωf\omega_{f}. Assume that there is some sequence (xn)n0(x_{n})_{n\geq 0} of XX converging to xx such that ωf(xn)\omega_{f}(x_{n}) converges to LL and LML\neq M. As ωf(xn)\omega_{f}(x_{n}) is a minimal set, for any n0n\geq 0, thus by Theorem 4.1, LL is also a minimal set and then ML=M\cap L=\emptyset.

Claim 1. For any neighborhood VLV_{L} of LL, there exists zVLP(f)z\in V_{L}\cap\textrm{P}(f) such that Mαf(z)M\subset\alpha_{f}(z).

Proof. Fix VLV_{L} some neighborhood of LL. Let UMU_{M} and ULU_{L} be two disjoint neighborhoods of M,LM,L with finite boundary such that ULVLU_{L}\subset V_{L}. Set k=card((UL))k=\textrm{card}(\partial(U_{L})). As (xn)n0(x_{n})_{n\geq 0} converges to xx and Of(x)O_{f}(x) is infinite, then for each 0jk0\leq j\leq k, we can find a sequence of arcs (In,j)n0(I_{n,j})_{n\geq 0} joining fj(xn)f^{j}(x_{n}) and fj(x)f^{j}(x) converging to {fj(x)}\{f^{j}(x)\} such that for any n,m0n,m\geq 0 and for any 0i<jk0\leq i<j\leq k, In,iIm,j=I_{n,i}\cap I_{m,j}=\emptyset. Let η=min{d(fp(x),fq(x)), 0p<qk}\eta=\min\{d(f^{p}(x),f^{q}(x)),\;0\leq p<q\leq k\}. Clearly η>0\eta>0 and we may assume that ε=infn0{d(In,p,In,q):0p<qk}>0\varepsilon=\inf_{n\geq 0}\{d(I_{n,p},I_{n,q}):0\leq p<q\leq k\}>0.
We can assume that for any n0,ωf(xn)ULn\geq 0,\;\omega_{f}(x_{n})\subset U_{L}. Thus for each n0n\geq 0, we can find mn>0m_{n}>0 such that {fmn(fj(xn)):0jk}UL\{f^{m_{n}}(f^{j}(x_{n})):0\leq j\leq k\}\subset U_{L}. Therefore for each n0, 0jk,fmn(In,j)(UL)n\geq 0,\;0\leq j\leq k,\;f^{m_{n}}(I_{n,j})\cap\partial(U_{L})\neq\emptyset. Then for each n0n\geq 0, there exists zn(UL)z_{n}\in\partial(U_{L}) and 0j1,n<j2,nk0\leq j_{1,n}<j_{2,n}\leq k such that znfmn(In,j1,n)fmn(In,j2,n)z_{n}\in f^{m_{n}}(I_{n,j_{1,n}})\cap f^{m_{n}}(I_{n,j_{2,n}}). Hence fmn(zn)In,jp,nf^{-m_{n}}(z_{n})\cap I_{n,j_{p,n}}\neq\emptyset, for p{0,1}p\in\{0,1\}. Recall that zn(OL)z_{n}\in\partial(O_{L}) which is finite and 0jn,1<jn,2k0\leq j_{n,1}<j_{n,2}\leq k. Hence there exists z(UL)z\in\partial(U_{L}) and 0j1<j2k0\leq j_{1}<j_{2}\leq k such that fmn(z)In,jpf^{-m_{n}}(z)\cap I_{n,j_{p}}\neq\emptyset, for infinitely many nn, p{0,1}p\in\{0,1\}. Therefore for infinitely many nn, we have that diam(fmn(z))ε\textrm{diam}(f^{-m_{n}}(z))\geq\varepsilon. Hence zP(f)z\in\textrm{P}(f). Moreover as fmn(z)In,jpf^{-m_{n}}(z)\cap I_{n,j_{p}}\neq\emptyset, then Of(x)αf(z)O_{f}(x)\cap\alpha_{f}(z)\neq\emptyset and hence Mαf(z)M\subset\alpha_{f}(z). This ends the proof of Claim 1.

Now by Claim 1, we may find a sequence (zn)n0(z_{n})_{n\geq 0} of periodic points with disjoint orbits converging to some point lLl\in L such that Mαf(zn)M\subset\alpha_{f}(z_{n}). By Lemma 5.3, (Of(zn))n1(O_{f}(z_{n}))_{n\geq 1} converges to LL. Moreover for each n1n\geq 1, we can find sn>0s_{n}>0 such that d(fsn(zn),M)1nd(f^{-s_{n}}(z_{n}),M)\leq\frac{1}{n}. As Of(zn)P(f)O_{f}(z_{n})\subset\textrm{P}(f), then Of(zn)fsn(zn)O_{f}(z_{n})\cap f^{-s_{n}}(z_{n})\neq\emptyset. Therefore we can find N0N\geq 0 such that for any nNn\geq N, we have diam(fsn(zn))d(M,L)2\textrm{diam}(f^{-s_{n}}(z_{n}))\geq\frac{d(M,L)}{2}. Recall that (zn)n0(z_{n})_{n\geq 0} is a sequence of periodic points with disjoint orbits, hence (fsn(zn))nN(f^{-s_{n}}(z_{n}))_{n\geq N} is a non null family of pairwise disjoint subcontinua. A contradiction with the fact that XX is a regular curve and hence finitely Suslinean. This ends the proof of the continuity of ωf\omega_{f}. ∎

6. On special α\alpha-limit sets

In [18], Hero introduced another kind of limit sets, called the special α\alpha-limit sets. He considered the union of the α\alpha-limit sets over all backward orbits of xx.

Definition 6.1.

Let XX be a metric compact space, f:XXf:X\to X a continuous map and xXx\in X. The special α\alpha-limit set of xx, denoted sαf(x)s\alpha_{f}(x), is the union sαf(x)=αf((xn)n0)s\alpha_{f}(x)=\displaystyle\bigcup\alpha_{f}((x_{n})_{n\geq 0}) taken over all negative orbits (xn)n0(x_{n})_{n\geq 0} of xx.

Notice that we have the following equivalence:

(i) αf(x)\alpha_{f}(x)\neq\emptyset, (ii) sαf(x)s\alpha_{f}(x)\neq\emptyset, (iii) xXx\in X_{\infty}.

In particular, if ff is onto, then sαf(x)s\alpha_{f}(x)\neq\emptyset for every xXx\in X.

It is clear that sαf(x)αf(x)s\alpha_{f}(x)\subset\alpha_{f}(x). The inclusion can be strict. Hero [18] provided an example of a continuous map on the interval for which the inclusion sαf(x)αf(x)s\alpha_{f}(x)\subset\alpha_{f}(x) is strict. Even, one can provide an onto monotone interval map. Indeed, consider the monotone map g:[0,1][0,1]:xmax{0,2x1}g:[0,1]\longrightarrow[0,1]:x\longmapsto\max\{0,2x-1\}. We see that sαg(0)={0,1}αg(0)=[0,1]s\alpha_{g}(0)=\{0,1\}\subsetneq\alpha_{g}(0)=[0,1].

In ([22, Theorem 3.3 and Corollary 3.11]), Kolyada et al. provided an example of a map on a subset of 2\mathbb{R}^{2} where a special α\alpha-limit set is not closed. Recently, Hantáková and Roth proved in ([16], Theorem 37) that a special α\alpha-limit set for interval map is always Borel, and in fact both FσF_{\sigma} and GδG_{\delta}. Furthermore, they provided a counterexample of an interval map with a special α\alpha-limit set which is not closed, this disproves the conjecture 1 in [22]. However they showed that a special α\alpha-limit set is closed for a piecewise monotone interval map. Jackson et al. ([19]) proved that a special α\alpha-limit set is always analytic (i.e. a continuous image of a Polish space) and provide an example of a map of the unit square with special α\alpha-limit set not a Borel set. Here we show that, for a monotone map on a regular curve, the special α\alpha-limit set is always closed.

6.1. Relation between Nonwandering sets, α\alpha-limit and Special α\alpha-limit sets

The aim of this paragraph is to prove the following theorem.

Theorem 6.2.

Let XX be a regular curve and ff a monotone self mapping of XX. Then for every xXx\in X_{\infty}, we have that  sαf(x)=αf(x)Ω(f)s\alpha_{f}(x)=\alpha_{f}(x)\cap\Omega(f).

First, we derive from Theorem 2.8 the following corollary.

Corollary 6.3.

Let XX be a regular curve and ff a monotone self mapping of XX. Then or any xXx\in X_{\infty}:
(i) sαf(x)Ω(f)s\alpha_{f}(x)\subset\Omega(f),
(ii) sαf(x)s\alpha_{f}(x) is a union of minimal sets.
(iii) if xXP(f)x\in X_{\infty}\setminus\mathrm{P}(f), then sαf(x)=αf(x)s\alpha_{f}(x)=\alpha_{f}(x) is a minimal set.

Proof.

Let xXx\in X_{\infty}. By Theorem 2.8, (2), αf((xn)n0)\alpha_{f}((x_{n})_{n\geq 0}) is a minimal set for every negative orbit (xn)n0(x_{n})_{n\geq 0} of xx, thus (i) and (ii) follow. Assume now that xXP(f)x\in X_{\infty}\setminus\textrm{P}(f), again by Theorem 2.8, (2), we have αf((xn)n0)=αf(x)\alpha_{f}((x_{n})_{n\geq 0})=\alpha_{f}(x) is a minimal set for any negative orbit (xn)n0(x_{n})_{n\geq 0} of xx, thus (iii) follows. ∎

The proof of Theorem 6.2 needs the following lemma.

Lemma 6.4.

Let XX be a regular curve and ff a monotone self mapping of XX. If xP(f)x\in\mathrm{P}(f), then every minimal set Mαf(x)M\subset\alpha_{f}(x) is a periodic orbit.

Proof.

Let xP(f)x\in\textrm{P}(f) and MM an infinite minimal set such that Mαf(x)M\subset\alpha_{f}(x). Fix some yMαf(x)y\in M\subset\alpha_{f}(x), we can find an increasing sequence of integers (mn)n0(m_{n})_{n\geq 0} and xnfmn(x)x_{n}\in f^{-m_{n}}(x) such that (xn)n0(x_{n})_{n\geq 0} converges to yy. Clearly for any n0n\geq 0, we have ωf(xn)=Of(x)\omega_{f}(x_{n})=O_{f}(x). By Theorem 5.1, ωf\omega_{f} is continuous at yy, therefore we get ωf(y)=M=Of(x)\omega_{f}(y)=M=O_{f}(x), which is finite. A contradiction. ∎

Recall that given a subset AA of a topological space XX, the arc connected component CC of aAa\in A is defined as

C={yA:there exists an arc IA joining a and y}.C=\{y\in A:\textrm{there exists an arc }I\subset A\textrm{ joining }a\textrm{ and }y\}.
Lemma 6.5.

Let XX be a regular curve and AA a subset of XX. Then the following hold:
(i) Every arc connected component of AA is closed in AA.
(ii) If CC is arcwise connected, then it is locally arcwise connected.

Proof.

Since a regular curve is finitely suslinean continuum, so the proof of (i) results from [35, Corollary 2.2] and the proof of (ii) results from [17, Corollary 5.5]. ∎

Lemma 6.6.

Let XX be a regular curve and ff a monotone self mapping of XX. If xP(f)x\in\mathrm{P}(f) and Of(x)αf(x)O_{f}(x)\subsetneq\alpha_{f}(x), then Ω(f)X\Omega(f)\neq X.

Proof.

Assume that Ω(f)=X\Omega(f)=X, by Theorem 3.1, Ω(f)=R(f)\Omega(f)=\textrm{R}(f), thus R(f)=X\textrm{R}(f)=X. Observe that for some n0,fn(x)Of(x)n\geq 0,f^{-n}(x)\nsubseteq O_{f}(x), if not Of(x)=αf(x)O_{f}(x)=\alpha_{f}(x). So let tXOf(x)t\in X\setminus O_{f}(x) and n>0n>0 such that fn(t)=xf^{n}(t)=x, then ωf(t)=Of(x)\omega_{f}(t)=O_{f}(x) and tOf(x)t\notin O_{f}(x). This will lead to a contradiction since tR(f)t\in\textrm{R}(f). ∎

Proof of Theorem 6.2.

If xXP(f)x\in X_{\infty}\setminus\textrm{P}(f), then by Corollary 6.3, sαf(x)=αf(x)Ω(f)s\alpha_{f}(x)=\alpha_{f}(x)\subset\Omega(f) and so sαf(x)=αf(x)Ω(f)s\alpha_{f}(x)=\alpha_{f}(x)\cap\Omega(f). Now assume that xP(f)x\in\textrm{P}(f). The inclusion sαf(x)αf(x)Ω(f)s\alpha_{f}(x)\subset\alpha_{f}(x)\cap\Omega(f) follows from Corollary 6.3. Now let yαf(x)Ω(f)y\in\alpha_{f}(x)\cap\Omega(f). By Lemma 6.4, ωf(y)\omega_{f}(y) is a periodic orbit and so yP(f)y\in\textrm{P}(f). If Of(y)Of(x)O_{f}(y)\cap O_{f}(x)\neq\emptyset, then yOf(x)sαf(x)y\in O_{f}(x)\subset s\alpha_{f}(x). Now assume that Of(y)Of(x)=O_{f}(y)\cap O_{f}(x)=\emptyset, then η=d(Of(x),Of(y))>0\eta=d(O_{f}(x),O_{f}(y))>0 and Of(x)αf(x)O_{f}(x)\subsetneq\alpha_{f}(x). Therefore by Lemma 6.6, Ω(f)X\Omega(f)\subsetneq X. By Lemma 5.2, we can find an open set UU such that Ω(f)U\Omega(f)\subset U and Mesh(UΩ(f))<η2(U\setminus\Omega(f))<\frac{\eta}{2}. Denote by A=n0fn(x)Of(x)Of(y)A=\displaystyle\bigcup_{n\geq 0}f^{-n}(x)\cup O_{f}(x)\cup O_{f}(y), observe that f(A)Af(A)\subset A.

Claim: AA has a finite number of arc connected components.
Denote by qq the period of xx. We have

A=i=0q1(n0fi(fqn(x)))Of(x)Of(y).A=\displaystyle\bigcup\limits_{i=0}^{q-1}\big{(}\displaystyle\bigcup_{n\geq 0}f^{-i}(f^{-qn}(x))\big{)}\cup O_{f}(x)\cup O_{f}(y).

Since ff is monotone then for each n0,fqn(x)n\geq 0,\ f^{-qn}(x) is a subcontinuum of the regular curve XX, thus arcwise connected, moreover it contains xx. We conclude that n0(fqn(x))\displaystyle\bigcup_{n\geq 0}(f^{-qn}(x)) is an arcwise connected subset of XX. Again since ff is monotone and n0(fqn(x))\displaystyle\bigcup_{n\geq 0}(f^{-qn}(x)) is arcwise connected then for each 0iq1,n0fi(fqn(x))0\leq i\leq q-1,\;\displaystyle\bigcup_{n\geq 0}f^{-i}(f^{-qn}(x)) is also arcwise connected. Indeed for a,bn0fi(fqn(x))a,b\in\displaystyle\bigcup_{n\geq 0}f^{-i}(f^{-qn}(x)), we can find an arc II joining fi(a)f^{i}(a) and fi(b)f^{i}(b) in n0(fqn(x))\displaystyle\bigcup_{n\geq 0}(f^{-qn}(x)), thus we can find some arc JJ in fi(I)n0fi(fqn(x))f^{-i}(I)\subset\displaystyle\bigcup_{n\geq 0}f^{-i}(f^{-qn}(x)) joining aa and bb. We conclude that AA can be written as a finite union of arcwise connected subset of XX, this end the proof of the claim. Recall that yαf(x)y\in\alpha_{f}(x) then we can find an increasing sequence of integer (mn)n0(m_{n})_{n\geq 0} and xnfmn(x)x_{n}\in f^{-m_{n}}(x) so that (xn)n0(x_{n})_{n\geq 0} converges to yy. For each n0,xnAn\geq 0,\;x_{n}\in A which has a finite number of arc connected component by the claim above, namely C1,CmC_{1},\dots C_{m}, so we may find 1jm1\leq j\leq m such that for each n0,xnCjn\geq 0,\;x_{n}\in C_{j} so that yCj¯y\in\overline{C_{j}}. By Lemma 6.5, (i), Cj{y}C_{j}\cup\{y\} is arcwise connected. Therefore by Lemma 6.5, (ii), locally arcwise connected. Let then In(Cj{y})B(y,1n)I_{n}\subset\big{(}C_{j}\cup\{y\}\big{)}\cap B(y,\frac{1}{n}) be an arc in Cj{y}C_{j}\cup\{y\} joining yy and xnfmn(x)x_{n}\in f^{-m_{n}}(x) so that (In)n0(I_{n})_{n\geq 0} converges to {y}\{y\}. We have for each n0{x,fmn(y)}fmn(In)A(XΩ(f)){Of(x),Of(y)}n\geq 0\;\{x,f^{m_{n}}(y)\}\subset f^{m_{n}}(I_{n})\subset A\subset\big{(}X\setminus\Omega(f)\big{)}\cup\{O_{f}(x),O_{f}(y)\}. Therefore for each n0,fmn(In)n\geq 0,\;f^{m_{n}}(I_{n}) is an arcwise connected set of XX meeting Ω(f)\Omega(f) only at Of(x),Of(y)O_{f}(x),\;O_{f}(y). Therefore fmn(In)Uf^{m_{n}}(I_{n})\nsubseteq U. Thus for each n0n\geq 0, there exists znfmn(In)UXΩ(f)z_{n}\in f^{m_{n}}(I_{n})\cap\partial U\subset X\setminus\Omega(f). As U\partial U is finite there exists zUz\in\partial U such that zn=zz_{n}=z, for infinitely many nn. Hence fmn(z)Inf^{-m_{n}}(z)\cap I_{n}\neq\emptyset for infinitely many n0n\geq 0. Since zUXP(f)z\in\partial U\subset X\setminus\textrm{P}(f), so by Theorem 2.8, αf(z)\alpha_{f}(z) is a minimal set containing yy. Thus zXz\in X_{\infty} and αf(z)=Of(y)\alpha_{f}(z)=O_{f}(y). Recall that for each n0n\geq 0, znAz_{n}\in A and hence zAz\in A. Then for some p>0p>0, we have fp(z)=xf^{p}(z)=x, therefore any negative orbit of zz can be completed into a negative orbit of xx hence sαf(z)sαf(x)s\alpha_{f}(z)\subset s\alpha_{f}(x). By Theorem 2.8, we have sαf(z)=αf(z)=Of(y)s\alpha_{f}(z)=\alpha_{f}(z)=O_{f}(y) and so ysαf(x)y\in s\alpha_{f}(x). ∎

Corollary 6.7.

Let XX be a regular curve and ff a monotone self mapping of XX. Then for every xXx\in X, sαf(x)s\alpha_{f}(x) is a closed set.

Proof.

The proof follows from Theorem 6.2  (since Ω(f)\Omega(f) and αf(x)\alpha_{f}(x) are closed). ∎

The following proposition improves Theorem 3.3.

Proposition 6.8.

Let XX be a regular curve and ff a monotone self mapping of XX. Then Ω(f)={xX:xsαf(x)}\Omega(f)=\{x\in X:x\in s\alpha_{f}(x)\}.

Proof.

Let xXx\in X such that xsαf(x)x\in s\alpha_{f}(x). By Corollary 6.3, sαf(x)s\alpha_{f}(x) is a union of minimal sets, thus xΩ(f)x\in\Omega(f). Conversely, let xΩ(f)x\in\Omega(f). Then by Theorem 3.1, xωf(x)x\in\omega_{f}(x). Thus xx has a negative orbit (xn)n0ωf(x)(x_{n})_{n\geq 0}\subset\omega_{f}(x). Therefore by minimality of ωf(x)\omega_{f}(x) (cf. Theorem 2.8), we have αf((xn)n0)=ωf(x)\alpha_{f}((x_{n})_{n\geq 0})=\omega_{f}(x). Hence xαf((xn)n0)sαf(x)x\in\alpha_{f}((x_{n})_{n\geq 0})\subset s\alpha_{f}(x). The proof is complete. ∎

6.2. Further results on special α\alpha-limit sets

Theorem 6.9 (Continuity of special α\alpha-limit maps).

Let XX be a regular curve and ff a monotone self mapping of XX. Then the special α\alpha-limit map sαfs\alpha_{f} is continuous everywhere except may be at the periodic points of ff.

Proof.

Let xXP(f)x\in X_{\infty}\setminus\textrm{P}(f) and (xn)n0(x_{n})_{n\geq 0} be a sequence of XX_{\infty} converging to xx. By Theorem 5.1, the sequence αf(xn)n0\alpha_{f}(x_{n})_{n\geq 0} converges to αf(x)\alpha_{f}(x). Let FF be any limit point of sαf(xn)n0s\alpha_{f}(x_{n})_{n\geq 0} with respect to the Hausdorff metric. Clearly Fαf(x)F\subset\alpha_{f}(x), moreover by Corollary 6.7, FF is an invariant closed subset of XX. By Theorem 2.8, (2), αf(x)\alpha_{f}(x) is a minimal set, therefore F=αf(x)F=\alpha_{f}(x). It turn out that αf(x)\alpha_{f}(x) is the unique limit point of the sequence sαf(xn)n0s\alpha_{f}(x_{n})_{n\geq 0}. We conclude that the sequence sαf(xn)n0s\alpha_{f}(x_{n})_{n\geq 0} converges to αf(x)\alpha_{f}(x) (since 2X2^{X} is compact). Again by Theorem 2.8, (2), αf(x)=sαf(x)\alpha_{f}(x)=s\alpha_{f}(x) and the result follows. ∎

Note that it may happened that the maps ωf,αf\omega_{f},\alpha_{f} and sαf(x)s\alpha_{f}(x) are discontinuous at some periodic point: For example consider the map f:XXf:X\to X given in [14, Example 5.10].

Example 6.10.

(([14, Example 5.10])) There exists a monotone map ff on an infinite star XX centered at a point z02z_{0}\in\mathbb{R}^{2} and beam In,n0I_{n},\;n\geq 0 with endpoints z0z_{0} and znz_{n} satisfying the following properties.

  • (i)

    For any xX{zn:n0}x\in X\setminus\{z_{n}:n\geq 0\} we have ωf(x)={z0}\omega_{f}(x)=\{z_{0}\} and αf(x)={znx}\alpha_{f}(x)=\{z_{n_{x}}\}, where nx1n_{x}\geq 1 such that xInxx\in I_{n_{x}}.

  • (ii)

    For any n1n\geq 1 we have αf(zn)=ωf(zn)={zn}\alpha_{f}(z_{n})=\omega_{f}(z_{n})=\{z_{n}\}.

  • (iii)

    ωf(z0)={z0},αf(z0)=X\omega_{f}(z_{0})=\{z_{0}\},\;\alpha_{f}(z_{0})=X.

We have ωf\omega_{f} is discontinuous at znz_{n}, for every n0n\geq 0 and αf\alpha_{f} (resp. sαfs\alpha_{f}) is discontinuous at z0z_{0}.

Denote by SA(f)\textrm{SA}(f) (respectively A(f)\textrm{A}(f)) the union of all special α\alpha-limit sets (respectively, all α\alpha-limit sets) of a map ff. In [24, Corollary 2.2], it was shown that R(f)SA(f)\textrm{R}(f)\subset\textrm{SA}(f) holds for general dynamical system (X,f)(X,f). For continuous map ff on graphs, it was shown that SA(f)R(f)¯Λ(f)\textrm{SA}(f)\subset\overline{\textrm{R}(f)}\subset\Lambda(f) (see [33], cf. [8], [18]) and that there are continuous maps ff on dendrites with SA(f)R(f)¯\textrm{SA}(f)\nsubseteq\overline{\textrm{R}(f)} (cf. [34]). In the following theorem, we extend the above results to regular curves when we restricted to monotone maps.

Theorem 6.11.

Let XX be a regular curve and ff a monotone self mapping of XX, the following hold:
(i) SA(f)=R(f)\mathrm{SA}(f)=\mathrm{R}(f).
(ii) SA(f)\mathrm{SA}(f) and A(f)\mathrm{A}(f) are closed.
(iii) If P(f)=\mathrm{P}(f)=\emptyset, then SA(f)=R(f)=A(f)\mathrm{SA}(f)=\mathrm{R}(f)=\mathrm{A}(f).

Proof.

Recall that by Theorem 3.1, Ω(f)=R(f)\Omega(f)=\textrm{R}(f). Assertion (i) follows immediately from Corollary 6.3 and Proposition 6.8.

(iii): Assume now that P(f)=\textrm{P}(f)=\emptyset. Then by Corollary 6.3, sαf(x)=αf(x)s\alpha_{f}(x)=\alpha_{f}(x) is a minimal set, for any xXx\in X_{\infty}, so assertion (iii) follows.

(ii): By (i), Ω(f)=R(f)=SA(f)\Omega(f)=\textrm{R}(f)=\textrm{SA}(f) and so SA(f)\textrm{SA}(f) is closed. Let us show that A(f)\mathrm{A}(f) is closed. Let (xn)n0(x_{n})_{n\geq 0} be a sequence of A(f)\textrm{A}(f) converging to some xXx\in X. We can assume that xΩ(f)x\notin\Omega(f) since otherwise xΩ(f)A(f)x\in\Omega(f)\subset\textrm{A}(f) and we are done. For each n0n\geq 0, let bnXb_{n}\in X_{\infty} such that xnαf(bn)x_{n}\in\alpha_{f}(b_{n}). Suppose that xA(f)x\notin\textrm{A}(f), then by passing to a subsequence if needed, one can assume that the sequence (bn)n0(b_{n})_{n\geq 0} is infinite (since otherwise xA(f)x\in\textrm{A}(f)).

Since xΩ(f)x\notin\Omega(f), so for nn large enough, xnΩ(f)x_{n}\notin\Omega(f) and therefore αf(bn)\alpha_{f}(b_{n}) is not a minimal set. Therefore by Theorem 2.8, bnP(f)b_{n}\in\textrm{P}(f). So by passing to a subsequence if needed, the family (Of(bn))n0(O_{f}(b_{n}))_{n\geq 0} is pairwise disjoint. Let UU be an open neighborhood of xx with finite boundary such that U¯Ω(f)=\overline{U}\cap\Omega(f)=\emptyset. As xnαf(bn)x_{n}\in\alpha_{f}(b_{n}) and the sequence (xn)n0(x_{n})_{n\geq 0} converges to xx, there exists N0N\geq 0 such that for any nNn\geq N, we can find sn0s_{n}\geq 0 such that fsn(bn)Uf^{-s_{n}}(b_{n})\cap U\neq\emptyset. Since bnP(f)b_{n}\in\textrm{P}(f), we have fsn(bn)Uf^{-s_{n}}(b_{n})\nsubseteq U. Since ff is monotone, it follows that fsn(bn)Uf^{-s_{n}}(b_{n})\cap\partial U\neq\emptyset, for any nNn\geq N. By Lemma 5.4 the family (fsn(bn))nN(f^{-s_{n}}(b_{n}))_{n\geq N} is pairwise disjoint and as proven above fsn(bn)Uf^{-s_{n}}(b_{n})\cap\partial U\neq\emptyset, for each nNn\geq N, this will lead to a contradiction since U\partial U is finite. ∎

Remark 6.12.

If P(f)\mathrm{P}(f)\neq\emptyset, we have the inclusion SA(f)A(f)\textrm{SA}(f)\subset\textrm{A}(f) which can be strict for monotone maps on regular curves. Indeed, consider the map f:SSf_{\infty}:S_{\infty}\to S_{\infty} of Example 6.18. We have that
SA(f)=R(f)=P(f)A(f)=S\textrm{SA}(f_{\infty})=\textrm{R}(f_{\infty})=\textrm{P}(f_{\infty})\subsetneq\textrm{A}(f_{\infty})=S_{\infty}.

Theorem 6.13.

Let XX be a regular curve and ff a monotone self mapping of XX. Then for any xP(f),sαf(x)x\in\mathrm{P}(f),\;s\alpha_{f}(x) is either a finite union of periodic orbits or an infinite sequence of periodic orbits converging to Of(x)O_{f}(x).

Proof.

Let xP(f)x\in\textrm{P}(f). By Corollary 6.3 and Lemma 6.4, sαf(x)s\alpha_{f}(x) is a union of periodic orbits.

Assume that sαf(x)s\alpha_{f}(x) is infinite and that there is an accumulation point aa of sαf(x)s\alpha_{f}(x) such that aOf(x)a\notin O_{f}(x). We can find an infinite sequence (an)n0(a_{n})_{n\geq 0} of sαf(x)s\alpha_{f}(x) converging to aa. By Theorem 4.1, (Of(an))n0(O_{f}(a_{n}))_{n\geq 0} converges to a minimal set LL. As anP(f)a_{n}\in\textrm{P}(f), then aLa\in L and hence L=Of(a)L=O_{f}(a). Then for ii large enough, say ipi\geq p,

d(Of(ai),Of(x))d(Of(x),Of(a))2.d\left(O_{f}(a_{i}),~{}O_{f}(x)\right)\geq\frac{d\Big{(}O_{f}(x),~{}O_{f}(a)\Big{)}}{2}.

Since Ω(f)X\Omega(f)\subsetneq X (Lemma 6.6), so by Lemma 5.2, there is an open set OO such that card(O)=k\textrm{card}(\partial O)=k and

Mesh(OΩ(f))<d(Of(x),Of(a))2.\textrm{Mesh}(O\setminus\Omega(f))<\frac{d\left(O_{f}(x),~{}O_{f}(a)\right)}{2}.

As done in the proof of Theorem 6.2, the set

A=n0fn(x){Of(ai),pik+p}A=\displaystyle\cup_{n\geq 0}f^{-n}(x)\cup\{O_{f}(a_{i}),~{}p\leq i\leq k+p\}

is ff-invariant and has a finite number of arc connected components. For each pik+pp\leq i\leq k+p, let CiC_{i} be the arc connected component of AA containing aia_{i}. Since aiαf(x)\;a_{i}\in\alpha_{f}(x), for each pik+pp\leq i\leq k+p, we can find an increasing sequence of integers (mn,i)n0(m_{n,i})_{n\geq 0} and a sequence (xn,i)n0(x_{n,i})_{n\geq 0} which converges to aia_{i}, where xn,ifmn,i(x)x_{n,i}\in f^{-m_{n,i}}(x). By the same arguments as in the proof of Theorem 6.2, we can assume that xn,iCix_{n,i}\in C_{i} for infinitely many n0n\geq 0. Now as the aia_{i} are pairwise distinct and the CiC_{i} are locally arcwise connected, pik+pp\leq i\leq k+p, so for nn large enough, we may find a sequence of pairwise disjoint arcs (In,i)n0(I_{n,i})_{n\geq 0} in CiC_{i} joining xn,ix_{n,i} and aia_{i} which converges to {ai}\{a_{i}\}. Choose NN large enough and set Ii=IN,iI_{i}=I_{N,i}, pik+pp\leq i\leq k+p. So (Ii)pik+p(I_{i})_{p\leq i\leq k+p} is a family of k+1k+1 pairwise disjoint arcs in AA joining aia_{i} and xN,ifmN,i(x)x_{N,i}\in f^{-m_{N,i}}(x). Set

η:=min{d(Ii,Ij):pi<jk+p}>0\displaystyle\eta:=\min\{d(I_{i},I_{j}):p\leq i<j\leq k+p\}>0

and

m>max{mN,i:pik+p}.m>\max\{m_{N,i}:p\leq i\leq k+p\}.

Since AA is ff-invariant, fm(Ii)Af^{m}(I_{i})\subset A. Moreover fm(Ii)f^{m}(I_{i}) is a subcontinuum of XX which meets the orbits of xx and aia_{i}. As

AΩ(f)=Of(x){Of(ai):pik+p},A\cap\Omega(f)=O_{f}(x)\cup\{O_{f}(a_{i}):p\leq i\leq k+p\},

then we can find an arc Jfm(Ii)J\subset f^{m}(I_{i}) joining xOf(x)x^{\prime}\in O_{f}(x) and biOf(ai)b_{i}\in O_{f}(a_{i}) such that J{x,bi}AΩ(f)J\setminus\{x^{\prime},b_{i}\}\subset A\setminus\Omega(f). Obviously J{x,bi}J\setminus\{x^{\prime},b_{i}\} is connected and diam(J{x,bi})d(Of(x),Of(a))2\textrm{diam}(J\setminus\{x^{\prime},b_{i}\})\geq\frac{d\left(O_{f}(x),~{}O_{f}(a)\right)}{2}. Therefore JOJ\nsubseteq O and JOJ\cap O\neq\emptyset. Thus JOJ\cap\partial O\neq\emptyset and therefore fm(Ii)Of^{m}(I_{i})\cap\partial O\neq\emptyset, for each pik+pp\leq i\leq k+p. It follows that there exists zmOz_{m}\in\partial O such that zmfm(Ii)fm(Ij)z_{m}\in f^{m}(I_{i})\cap f^{m}(I_{j}), for some pi<jk+pp\leq i<j\leq k+p. Hence diam(fm(zm))>η\textrm{diam}(f^{-m}(z_{m}))>\eta. As O\partial O is finite, there exists zOz\in\partial O such that z=zmz=z_{m} for infinitely many mm, thus lim sups+diam(fs(z))>0\displaystyle\limsup_{s\to+\infty}\textrm{diam}(f^{-s}(z))>0. This implies that the family (fn(z))n0(f^{-n}(z))_{n\geq 0} is not pairwise disjoint (otherwise it will be a null family), and hence for some q1<q2q_{1}<q_{2}, we have fq1(z)fq2(z)f^{-q_{1}}(z)\cap f^{-q_{2}}(z)\neq\emptyset. Thus z=fq2q1(z)z=f^{q_{2}-q_{1}}(z) and so zP(f)z\in\textrm{P}(f). As zOz\in\partial O and P(f)Ω(f)O\textrm{P}(f)\subset\Omega(f)\subset O, then zXP(f)z\in X\setminus\textrm{P}(f). A contradiction.

We conclude that any accumulation point of sαf(x)s\alpha_{f}(x) is a point of Of(x)O_{f}(x). It turns out that sαf(x)s\alpha_{f}(x) is a compact space having a finite set of accumulation points, hence it is countable. ∎

The following corollary is in contrast to the properties of ω\omega-limit sets.

Corollary 6.14.

Let XX be a regular curve and ff a monotone self mapping of XX and xXx\in X_{\infty}. If sαf(x)s\alpha_{f}(x) contains an isolated point, then sαf(x)P(f)s\alpha_{f}(x)\subset\mathrm{P}(f).

Proof.

If xP(f)x\notin\textrm{P}(f), then sαf(x)s\alpha_{f}(x) is a minimal set and hence it is a periodic orbit since it contains zz which is an isolated point in it. If xP(f)x\in\textrm{P}(f), then from Lemma 6.4, we have sαf(x)P(f)s\alpha_{f}(x)\subset\textrm{P}(f). ∎

Corollary 6.15.

A countable sαs\alpha-limit set for a monotone map on a regular curve is a union of periodic orbits.

Proof.

Let xXx\in X_{\infty} such that sαf(x)s\alpha_{f}(x) is countable. We may that sαf(x)s\alpha_{f}(x) is not minimal; otherwise sαf(x)s\alpha_{f}(x) will be a finite minimal set and we are done. Therefore by Theorem 2.8, (2), xP(f)x\in\textrm{P}(f) and by Theorem 6.13, the result follows. ∎

Remark 6.16.

Notice that Corollaries 6.14 and 6.15 extend those of [16, Theorem 20 and Corollary 21] to monotone maps on regular curves.

Remark 6.17.
  • (1)

    A special α\alpha-limit set can be totally periodic and infinite for monotone map on a regular curve; for example, in Example 6.10, we have sαf(z0)s\alpha_{f}(z_{0}) is infinite and composed of fixed points. This is in contrast to some properties of ω\omega-limit sets (cf. [14, Theorem 2.4]).

  • (2)

    In [5], it was constructed a continuous self-mapping (not monotone) of a dendrite having totally periodic ω\omega-limit set with unbounded periods. In Example 6.18 below, we construct analogously, a monotone map on a dendrite having a totally periodic special α\alpha-limit set with unbounded periods.

Example 6.18.

The idea is to slightly change the map ff in Example 6.10 so that sαf(z0)s\alpha_{f}(z_{0}) is infinite and composed of periodic orbits with unbounded period. Let z0z_{0} some point of the plan 2\mathbb{R}^{2}, N1N\geq 1 and denote by SNS_{N}, the NN-star centered at z0z_{0}. For N=1N=1, we define f1:S1S1f_{1}:S_{1}\longrightarrow S_{1} identified as the map f1=g:[0,1][0,1]:xmax{0,2x1}f_{1}=g:[0,1]\longrightarrow[0,1]:x\longmapsto\max\{0,2x-1\}, where z0z_{0} and S1S_{1} play the role of 0 and [0,1][0,1], respectively. Let now N1,SN=k=0N1Ik,NN\geq 1,S_{N}=\bigcup\limits_{k=0}^{N-1}I_{k,N}, each Ik,NI_{k,N} is an arc of 2\mathbb{R}^{2}, where z0z_{0} is one of its endpoints, we denote by zk,Nz_{k,N} the other endpoint of Ik,NI_{k,N} distinct from z0z_{0}. Let hN:SNSNh_{N}:S_{N}\longrightarrow S_{N} be the homeomorphism of SNS_{N} defined as hN(Ik,N)=Ik+1,Nh_{N}(I_{k,N})=I_{k+1,N} and hN(IN1,N)=I0,Nh_{N}(I_{N-1,N})=I_{0,N} in an affine way. We let fN=hNf1f_{N}=h_{N}\circ f_{1}. Clearly fN:SNSNf_{N}:S_{N}\longrightarrow S_{N} is monotone and continuous. Observe that End(SN)={zk,N: 0kN1}\textrm{End}(S_{N})=\{z_{k,N}:\;0\leq k\leq N-1\} is a periodic orbit of period NN, sαfN(z0)={z0,zk,N: 0kN1}s\alpha_{f_{N}}(z_{0})=\{z_{0},z_{k,N}:\;0\leq k\leq N-1\} and αfN(z0)=SN\alpha_{f_{N}}(z_{0})=S_{N}. We may pick (SN)N1(S_{N})_{N\geq 1} in such a way that S=n1SNS_{\infty}=\displaystyle\bigcup_{n\geq 1}S_{N} is an infinite star centered at z0z_{0}. Define the map f:SSf_{\infty}:S_{\infty}\to S_{\infty} given by its restriction (f)|SN=fN(f_{\infty})_{|S_{N}}=f_{N}. Clearly ff_{\infty} is monotone and continuous. Moreover αf(z0)=S\alpha_{f_{\infty}}(z_{0})=S_{\infty} and sαf(z0)={z0,zk,N:N1, 0kN1}s\alpha_{f_{\infty}}(z_{0})=\{z_{0},z_{k,N}:\;N\geq 1,\;0\leq k\leq N-1\} is totally periodic which contains periodic orbits of period NN, for any N1N\geq 1.

Refer to caption
Figure 2. The map gg
Refer to caption
Figure 3. The map ff_{\infty}

Acknowledgements. This work was supported by the research unit: “Dynamical systems and their applications”, (UR17ES21), Ministry of Higher Education and Scientific Research, Faculty of Science of Bizerte, Bizerte, Tunisia.

References

  • [1] H. Abdelli, ω\omega-limit sets for monotone local dendrite maps, Chaos Solitons Fractals, 71 (2015), 66–72.
  • [2] H. Abdelli, H. Marzougui, Invariant sets for monotone local dendrites, Internat. J. Bifur. Chaos Appl. Sci. Engrg., 26 (2016), 1650150.
  • [3] H. Abdelli, H. Abouda and H. Marzougui, Nonwandering points of monotone local dendrite maps revisited, Topology Appl., 250 (2018), 61–73.
  • [4] E.D. Anielloa, T.H. Steele, The persistence of ω\omega-limit sets defined on compact spaces, J. Math. Anal. Appl. 413 (2014), 789–799.
  • [5] G. Askri and I. Naghmouchi, On totally periodic ω\omega-limit sets in regular continua, Chaos Solitons Fractals 75 (2015), 91–95.
  • [6] F. Balibrea, J.L. Guirao and M. Lampart, A note on the definition of α\alpha-limit set, Appl. Math. Inf. Sci. 7 (2013), 1929–1932.
  • [7] F. Balibrea and C. La Paz, A characterization of the ω\omega-limit sets of interval maps. Acta Math. Hungar., 88(4):291–300, 2000.
  • [8] F. Balibrea, G. Dvorníková, M. Lampart, P. Oprocha, On negative limit sets for one-dimensional dynamics, Nonlinear Anal. 75 (2012), 3262–3267.
  • [9] P. Blanchard, R.L. Devaney, D. Look, M. Moreno Rocha, P. Seal, S. Siegmund, and D. Uminsky, Sierpinski carpets and gaskets as Julia sets of rational maps, In Dynamics on the Riemann Sphere, European Mathematical Society, Zürich, 2006, pp. 97–119.
  • [10] L.S. Block and W. A. Coppel, Dynamics in one dimension, Lecture Notes in Mathematics, 1513, Springer-Verlag, Berlin, 1992.
  • [11] E. Coven and Z. Nitecki, Non-wandering sets of the powers of maps of the interval, Ergodic Theory Dyn. Syst. 1 (1981), 9–31.
  • [12] A. Daghar (2021): On regular curve homeomorphisms without periodic points, J. Difference Equ. Appl., DOI: 10.1080/10236198.2021.1912030.
  • [13] A. Daghar, Homeomorphisms of hereditarily locally connected continua. 2021. ⟨hal-03223435v2⟩
  • [14] A. Daghar and H. Marzougui. Limit sets of monotone maps on regular curves. 2021. arXiv: 2106.12418v1
  • [15] A. Daghar, I. Naghmouchi and M. Riahi, Periodic Points of Regular Curve Homeomorphisms, Qual. Theory Dyn. Syst. 20 (2) (2021).
  • [16] J. Hantáková and S. Roth, On backward attractors of interval maps. 2020. arXiv: 2007.10883v2
  • [17] J. Grispolakis, E.D. Tymchatyn, σ\sigma- connectedness in hereditarily locally connected spaces, Trans. Amer. Math. Soc., 253 (1979), 303–315.
  • [18] M.W. Hero, Special α\alpha-limit points for maps of the interval, Proc. Amer. Math. Soc. 116 (1992), 1015–1022.
  • [19] S. Jackson, B. Mance and S. Roth, A non Borel special α\alpha-limit set in the square. 2020. arXiv: 2011.05509v1.
  • [20] H. Kato, Topological entropy of monotone maps and confluent maps on regular curves, Topol. Proc. 28 (2004), 587–593.
  • [21] H. Kato, Topological entropy of piecewise embedding maps on regular curues, Ergodic Theory Dyn. Syst. 26 (2006), 1115–1125.
  • [22] S. Kolyada, M. Misiurewicz and L. Snoha, Special α\alpha-limit sets, In Dynamics: topology and numbers, 157–173, Contemp. Math., 744, Amer. Math. Soc., Providence, RI, (2020).
  • [23] K. Kuratowski, Topology, vol.2, Academic Press, New-York, 1968.
  • [24] M. Foryś-Krawiec, M.J. Hantáková, P. Oprocha, On the structure of α\alpha-limit sets of backward trajectories for graph maps. 2021. arXiv:2106.05539v1.
  • [25] A. Lelek, On the topology of curves. II, Fund. Math. 70 (1971), 131–138.
  • [26] J. Mai, T. Sun, The ω\omega-limit set of a graph map, Topology Appl. 154 (2007) 2306–2311.
  • [27] S.B. Nadler, Continuum Theory: An Introduction, (Monographs and Textbooks in Pure and Applied Mathematics, 158). Marcel Dekker, Inc., New York, 1992.
  • [28] I. Naghmouchi, Homeomorphisms of regular curves, J. Difference Equ. Appl., 23 (2017), 1485–1490.
  • [29] I. Naghmouchi, Dynamics of Homeomorphisms of regular curves, Colloquium Math., 162 (2020), 263–277.
  • [30] I. Naghmouchi. Dynamics of monotone graph, dendrite and dendroid maps. Internat. J. Bifur. Chaos Appl. Sci. Engrg., 21 (2011), 3205–3215.
  • [31] G.T. Seidler, The topological entropy of homeomorphisms on one-dimensional continua, Proc. Amer. Math. Soc., 108 (1990), 1025–1030.
  • [32] T. Sun, Q. He, J. Liu, C. Tao, H. Xi, Non-wandering sets for dendrite maps, Qual. Theory Dyn. Syst. 14 (2015), 101–108.
  • [33] T. Sun, H. Xi, H. Liang, Special α\alpha-limit points and unilateral γ\gamma limit points for graph maps, Sci China Math., 54 (2011), 2013–2018.
  • [34] T. Sun, Y. Tang, G. Su, H. Xi, B. Qin, Special α\alpha-limit points and γ\gamma-limit points of a dendrite map, Qual. Theory Dyn. Syst. 17 (2018), 245–257.
  • [35] E.D. Tymchatyn, Characterizations of continua in which connected subsets are arcwise connected, Trans. Amer. Math. Soc., 222 (1976), 377–388.