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

Equicontinuous mappings on finite trees

Gerardo Acosta Instituto de Matemáticas
Universidad Nacional Autónoma de México
Área de la Investigación Científica, Circuito Exterior, Ciudad Universitaria
Coyoacán, 04510, CDMX, Mexico.
gacosta@matem.unam.mx
 and  David Fernández-Bretón Instituto de Matemáticas
Universidad Nacional Autónoma de México
Área de la Investigación Científica, Circuito Exterior, Ciudad Universitaria
Coyoacán, 04510, CDMX, Mexico.
djfernandez@im.unam.mx https://homepage.univie.ac.at/david.fernandez-breton/
Abstract.

If XX is a finite tree and f:XXf\colon X\longrightarrow X is a map, as the Main Theorem of this paper (Theorem 1.8), we find eight conditions, each of which is equivalent to the fact that ff is equicontinuous. To name just a few of the results obtained: the equicontinuity of ff is equivalent to the fact that there is no arc AXA\subseteq X satisfying Afn[A]A\subsetneq f^{n}[A] for some nn\in\mathbb{N}. It is also equivalent to the fact that for some nonprincial ultrafilter uu, the function fu:XXf^{u}\colon X\longrightarrow X is continuous (in other words, failure of equicontinuity of ff is equivalent to the failure of continuity of every element of the Ellis remainder gE(X,f)g\in E(X,f)^{*}). One of the tools used in the proofs is the Ramsey-theoretic result known as Hindman’s theorem. Our results generalize the ones shown by Vidal-Escobar and García-Ferreira in [18], and complement those of Bruckner and Ceder ([4]), Mai ([11]) and Camargo, Rincón and Uzcátegui ([6]).

Key words and phrases:
Dendrites, Discrete Dynamical Systems, Ellis Semigroup, Equicontinuous Functions, Finite Graphs, Finite Trees, Ramsey Theory.
2010 Mathematics Subject Classification:
Primary 37B40, 37E25, 54D80, 54F50; Secondary 54A20, 54D05.

1. Introduction

For a metric space X,X, this paper deals with maps f:XX,f\colon X\longrightarrow X, whose family of iterates is equicontinuous. Such functions represent well-behaved, non-chaotic, dynamical systems (equicontinuity is diametrically opposite to what is known as sensitivity to initial conditions, see [2, Theorem 2.4]).

Definition 1.1.
  1. (1)

    A discrete dynamical system is a pair (X,f)(X,f) such that XX is a metric space and f:XXf\colon X\longrightarrow X is a map, i.e. a continuous function.

  2. (2)

    If (X,f)(X,f) is a discrete dynamical system, we define f0f^{0} as the identity map on XX, and, for each nn\in\mathbb{N}, fn=fn1ff^{n}=f^{n-1}\circ f.

  3. (3)

    If X,YX,Y are metric spaces and \mathcal{F} is a family of functions from XX to YY, we say that \mathcal{F} is equicontinuous at xXx\in X if for every ε>0\varepsilon>0 there exists a δ>0\delta>0 such that d(x,y)<δd(x,y)<\delta implies d(f(x),f(y))εd(f(x),f(y))\leq\varepsilon for all yXy\in X and every f;f\in\mathcal{F}; and if \mathcal{F} is equicontinuous at every xXx\in X, we say that \mathcal{F} is equicontinuous.

  4. (4)

    If XX is a metric space, the function f:XXf\colon X\longrightarrow X is equicontinuous at xXx\in X if its family of iterates, {fn|n}\{f^{n}\big{|}n\in\mathbb{N}\}, is equicontinuous at xx; and if ff is equicontinuous at every xXx\in X we say that it is equicontinuous.

The definition of equicontinuity makes sense for every uniform space, but in this paper we will only consider metric spaces. Note that, upon fixing x,ε,δx,\varepsilon,\delta, equicontinuity of a family of functions \mathcal{F} is a pointwise closed condition; consequently if \mathcal{F} is equicontinuous at xx then so is ¯\overline{\mathcal{F}}, where ¯\overline{\mathcal{F}} is the closure of \mathcal{F} in YXY^{X} with the product topology. Note also that, if XX is compact, then by the usual argument, equicontinuity implies uniform equicontinuity (i.e., given ε>0\varepsilon>0, a δ>0\delta>0 can be chosen to work for all xXx\in X).

Definition 1.2.

Let (X,f)(X,f) be a discrete dynamical system, where XX is compact.

  1. (1)

    The Ellis semigroup (also called the enveloping semigroup) of (X,f)(X,f) is defined as E(X,f)={fn|n}¯E(X,f)=\overline{\{f^{n}\big{|}n\in\mathbb{N}\}}, the closure in XXX^{X} (with the product topology) of the family {fn|n}\{f^{n}\big{|}n\in\mathbb{N}\}. Note that, as XXX^{X} is compact (by Tychonoff’s theorem), so is E(X,f)E(X,f).

  2. (2)

    The Ellis remainder of the discrete dynamical system (X,f)(X,f) is

    E(X,f)=n=1{fk|kn}¯.E(X,f)^{*}=\bigcap_{n=1}^{\infty}\overline{\{f^{k}\big{|}k\geq n\}}.

    Note that E(X,f)=E(X,f){fn|n}E(X,f)=E(X,f)^{*}\cup\{f^{n}\big{|}n\in\mathbb{N}\}.

Composition of functions is what makes E(X,f)E(X,f) a semigroup. In fact, E(X,f)E(X,f) is a compact right-topological semigroup (a semigroup equipped with a topology making all right translations continuous). Since XX is a metric space, by the observation immediately after Definition 1.1, equicontinuity of ff is equivalent to equicontinuity of the family E(X,f)E(X,f), and either of these is equivalent to the same statement with uniform equicontinuity instead of equicontinuity (cf. [8, Theorem 3.3]).

The seemingly abstract object E(X,f)E(X,f) can be made more concrete by means of ultrafilters: for every ultrafilter uu on \mathbb{N}, define the ultrafilter-limit function fuf^{u} by letting fu(x)=u-limnfn(x)f^{u}(x)=u\text{-}\lim_{n\in\mathbb{N}}f^{n}(x). Then by [8, Theorem 2.2] we have

E(X,f)={fu|u is an ultrafilter on },E(X,f)=\{f^{u}\big{|}u\text{ is an ultrafilter on }\mathbb{N}\},

and consequently

E(X,f)={fu|u is a nonprincipal ultrafilter on }.E(X,f)^{*}=\{f^{u}\big{|}u\text{ is a nonprincipal ultrafilter on }\mathbb{N}\}.

Full definitions of ultrafilters, both principal and nonprincipal, as well as of uu-limits will be provided in Section 3.

Definition 1.3.

Let (X,f)(X,f) be a discrete dynamical system.

  1. (1)

    xXx\in X is a fixed point if f(x)=xf(x)=x; the set of fixed points of ff is denoted by Fix(f).\operatorname{Fix}(f).

  2. (2)

    xXx\in X is a periodic point if fn(x)=xf^{n}(x)=x, for some nn\in\mathbb{N}, in which case the least such nn is called the period of xx; the set of all periodic points of ff is denoted by Per(f)\operatorname{Per}(f).

  3. (3)

    ff is periodic if there exists nn\in\mathbb{N} such that fnf^{n} is the identity map on XX, and ff is pointwise-periodic if Per(f)=X.\operatorname{Per}(f)=X.

It is immediate from Definition 1.3 that Per(f)=n=1Fix(fn)\operatorname{Per}(f)=\bigcup_{n=1}^{\infty}\operatorname{Fix}(f^{n}).

We primarily deal with continua (compact, connected and metric spaces). A simple closed curve is a continuum homeomorphic to the unit circle 𝕊1,\mathbb{S}^{1}, and an arc is a continuum homeomorphic to the unit interval [0,1].[0,1]. Other examples of continua are finite graphs (compact, connected one-dimensional polyhedra), dendrites, finite trees and kk-ods, for each kk\in\mathbb{N} with k2.k\geq 2. We give the proper definitions of the last three in Section 2. For the moment it is convenient to note that a 22-od is an arc, kk-ods are finite trees and finite trees are dendrites.

In the early nineties, Bruckner and Hu ([5]) and Bruckner and Ceder ([4]) carried out a very deep and complete study of equicontinuity of maps defined on arcs, obtaining the following result.

Theorem 1.4 (Subset of [4, Theorem 1.2]).

If XX is an arc and f:XXf\colon X\longrightarrow X is a map, then the following are equivalent:

  1. (1)

    ff is equicontinuous;

  2. (2)

    the restriction f2m=1fm[X]f^{2}\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X] is the identity map;

  3. (3)

    Fix(f2)=m=1fm[X];\operatorname{Fix}(f^{2})=\bigcap_{m=1}^{\infty}f^{m}[X];

  4. (4)

    Fix(f2)\operatorname{Fix}(f^{2}) is connected.

Attempting to generalize this result from arcs to finite trees is futile, if taken too literally. Allowing, however, exponents other than 2 in the theorem above yields valid characterizations: we prove that, for an arbitrary finite tree XX and a map f:XXf\colon X\longrightarrow X, equicontinuity of ff is equivalent to each of the following conditions: that the restriction fm=1fm[X]f\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X] is periodic, that Fix(fn)=m=1fm[X]\operatorname{Fix}(f^{n})=\bigcap_{m=1}^{\infty}f^{m}[X] for some nn, and that Fix(fn)\operatorname{Fix}(f^{n}) is connected for all nn; furthermore, any of these is also equivalent to the set Per(f)\operatorname{Per}(f) being connected.

Further interesting results regarding equicontinuity of a map f:XXf\colon X\longrightarrow X have been obtained by Mai ([11]) in the case where XX is a finite graph, and by Camargo, Rincón and Uzcátegui ([6]) in the case where XX is a dendrite. The former shows in [11, Theorem 5.2] that, if XX is a finite graph, then ff is equicontinuous if and only if m=1fm[X]=Rec(f)\bigcap_{m=1}^{\infty}f^{m}[X]=\operatorname{Rec}(f) (here Rec(f)\operatorname{Rec}(f) denotes the set of recurrent points of ff, to be defined later in Definition 2.3; for the moment just note that Per(f)Rec(f)\operatorname{Per}(f)\subseteq\operatorname{Rec}(f)); the latter proves in [6, Theorem 4.12] that, if XX is a dendrite, then ff is equicontinuous if and only if clX(Per(f))=m=1fm[X]\operatorname{cl}_{X}(\operatorname{Per}(f))=\bigcap_{m=1}^{\infty}f^{m}[X] plus an extra condition having to do with the ω\omega-limit sets of ff. Obtaining a simultaneous strengthening of these two results at the expense of considering a less general class of spaces, we prove that, in the case where XX is a finite tree, ff is equicontinuous if and only if Per(f)=m=1fm[X]\operatorname{Per}(f)=\bigcap_{m=1}^{\infty}f^{m}[X].

Another concept that will play a central role in this paper is that of an expanding arc. To motivate this concept consider a nonnegative α\alpha\in\mathbb{R} and the map fα:f_{\alpha}\colon\mathbb{R}\longrightarrow\mathbb{R} defined by

(1.1) fα(x)=αx,for each xX.f_{\alpha}(x)=\alpha x,\hskip 14.22636pt\mbox{for each }x\in X.

It is readily checked that fαf_{\alpha} is equicontinuous if and only if 0α10\leq\alpha\leq 1, whereas if α>1\alpha>1 then fαf_{\alpha} fails to be equicontinuous at every xx\in\mathbb{R}. Intuitively speaking, maps that expand the real line fail to be equicontinuous. Note that for the map fαf_{\alpha} defined in (1.1), with α>1\alpha>1, we have Ifαn[I]I\subsetneq f_{\alpha}^{n}[I] for all nn\in\mathbb{N}, where I=[0,1]I=[0,1] is the unit interval. This leads to the following definition.

Definition 1.5.

Let (X,f)(X,f) be a discrete dynamical system, and let AXA\subseteq X be an arc. We say that AA is an ff-expanding arc if there exists nn\in\mathbb{N} such that Afn[A]A\subsetneq f^{n}[A].

Hence the map fαf_{\alpha} defined in (1.1) fails to be equicontinuous if and only if [0,1][0,1] is fαf_{\alpha}-expanding. Surprisingly, something like this very simple characterization still holds in more general situations. Namely, in [18, Theorems 3.1 and 3.7], Vidal-Escobar and García-Ferreira established the following result (in the case k=2k=2 they assume ff is surjective, but the general case follows from the proof of [4, Theorem 1.2]):

Proposition 1.6.

Let XX be a kk-od for some k2k\geq 2 and f:XXf:X\longrightarrow X be a map where f[X]f[X] is not a singleton. Then ff is not equicontinuous if and only if XX contains an ff-expanding arc.

In this paper, we generalize Proposition 1.6 from kk-ods to arbitrary finite graphs. Our proof of this generalization uses at a crucial point a highly nontrivial Ramsey-theoretic result (Hindman’s theorem). Hence, our result is not a direct use of the proofs presented in [4, Theorem 1.2] and [18, Theorems 3.1 and 3.7].

Another result of Vidal-Escobar and García-Ferreira ([18, Theorem 3.7]) is that for each map f:XXf:X\longrightarrow X, where XX is a kk-od with k3k\geq 3, if fuf^{u} is continuous for every nonprincipal ultrafilter uu then ff is equicontinous (note that the converse implication is trivially true as a consequence of the observation right after Definition 1.2). Hence if ff is not equicontinuous then for some nonprincipal ultrafilter u,u, fuf^{u} is not continuous. In [18] the authors consider the possibility that, for some nonprincipal ultrafilter v,v, distinct from u,u, fvf^{v} might be continuous.

In this paper we show that the possibility mentioned in the previous line cannot occur by proving that the statement from the preceding paragraph is true with every replaced by some, even if XX is a finite tree rather than just a kk-od. As a consequence of this, if ff fails to be equicontinuous with XX a finite tree, then every element gE(X,f)g\in E(X,f)^{*} fails to be continuous. Thus, for maps f:XXf\colon X\longrightarrow X on a finite tree XX, we have a strong dichotomy by means of which either every element gE(X,f)g\in E(X,f)^{*} is continuous, or every element gE(X,f)g\in E(X,f)^{*} is discontinuous, according to whether or not ff is equicontinuous. This is a direct generalization of a result of Szuca ([17, Theorem 2]), who obtains the same dichotomy for maps in an arc. This result is therefore worth stating explicitly.

Theorem 1.7.

Let (X,f)(X,f) be a discrete dynamical system, where XX is a finite tree. Then, either every element of E(X,f)E(X,f)^{*} is continuous, or every element of E(X,f)E(X,f)^{*} is discontinuous.

We now state the Main Theorem of this paper.

Main Theorem 1.8.

Let XX be a finite tree and f:XXf:X\longrightarrow X be a map. Then, the following are equivalent:

  1. (a)

    ff is equicontinuous;

  2. (b)

    there is an nn\in\mathbb{N} such that the restriction of fnf^{n} to m=1fm[X]\bigcap_{m=1}^{\infty}f^{m}[X] is the identity map;

  3. (c)

    there exists an nn\in\mathbb{N} such that Fix(fn)=m=1fm[X];\operatorname{Fix}(f^{n})=\bigcap_{m=1}^{\infty}f^{m}[X];

  4. (d)

    Per(f)=m=1fm[X];\operatorname{Per}(f)=\bigcap_{m=1}^{\infty}f^{m}[X];

  5. (e)

    there is no ff-expanding arc in X;X;

  6. (f)

    for every nn\in\mathbb{N}, the set Fix(fn)\operatorname{Fix}(f^{n}) is connected;

  7. (g)

    the set Per(f)\operatorname{Per}(f) is connected;

  8. (h)

    for every nonprincipal ultrafilter uu, the function fuf^{u} is continuous (i.e., every element of E(X,f)E(X,f)^{*} is continuous);

  9. (i)

    for some nonprincipal ultrafilter uu, the function fuf^{u} is continuous (i.e., some element of E(X,f)E(X,f)^{*} is continuous).

Remark 1.9.

Some remarks about the equivalences from Theorem 1.8:

  1. (1)

    The equivalence between (e), (f) and (g) will be established not only for finite trees, but for arbitrary dendrites.

  2. (2)

    The equivalence between (a), (e) and (h) shows that [18, Theorem 3.7] can be extended from kk-ods to finite trees. This is a partial answer to [18, Question 3.10].

  3. (3)

    The equivalences (a)(x)(\mathrm{a})\iff(\mathrm{x}), where x{b,c,,i}\mathrm{x}\in\{\mathrm{b},\mathrm{c},\ldots,\mathrm{i}\}, fail if we allow XX to be an arbitrary dendrite, and even if XX is merely a dendrite all of whose branching points are of finite order (see §4.2), or (with the possible exception of (a)(d)(\mathrm{a})\iff(\mathrm{d}), see Remark 4.1) if XX is merely a dendrite with finitely many branching points (see §4.1).

  4. (4)

    Further conditions equivalent to equicontinuity of a map ff on a space XX have been established in [14, Theorem 2, p. 62] for XX a finite tree, in [11, Theorem 5.2] when XX is a finite graph, and in [6, Theorem 4.12] in the case of XX an arbitrary dendrite.

The paper is structured around the equivalence that constitutes its main result (Theorem 1.8). In Section 2 we begin by proving the equivalence of items (a), (b), (c) and (d), which is a fairly elementary result, and the rest of the section is devoted to the study of expanding arcs, starting with the equivalence of (e) and (f), and concluding with the implication from (e) to (a). Then, in Section 3, we establish the equivalence between (e) and (g), in order to later on focus on ultrafilter-limit functions to establish that (i) implies (e) (this finishes the main theorem, since the implication from (h) to (i) is obvious and that from (a) to (h) is well-known). Finally, in Section 4 we describe the examples that exhibit the failure of all these characterizations in the context of arbitrary dendrites, and state some questions that remain open.

2. Equicontinuity and expanding arcs

Given a subset AA of a space X,X, we denote by either A¯\overline{A} or clX(A),\operatorname{cl}_{X}(A), the closure of AA in X.X. The interior of AA in XX is denoted by intX(A).\mathop{\mathrm{int}}_{X}(A). We begin by stating some standard results that will be used.

Proposition 2.1.

Let (An)n(A_{n})_{n\in\mathbb{N}} be a decreasing sequence of closed subsets of a compact space XX, and let A=n=1AnA=\bigcap_{n=1}^{\infty}A_{n}. Then,

  1. (1)

    if UU is an open set containing AA, then AnUA_{n}\subseteq U for all sufficiently large nn;

  2. (2)

    if each AnA_{n} is nonempty, then so is AA;

  3. (3)

    if every AnA_{n} is connected, then so is AA;

  4. (4)

    if f:XYf:X\longrightarrow Y is a map, then n=1f[An]=f[A]\bigcap_{n=1}^{\infty}f[A_{n}]=f[A].

Proof.

Parts (1) and (3) follow from [7, Corollary 3.1.5 and Corollary 6.1.19]. Part (2) follows from (1) with U=.U=\varnothing. To show part (4), it is enough to verify that n=1f[An]f[A]\bigcap_{n=1}^{\infty}f[A_{n}]\subseteq f[A]. Let yn=1f[An].y\in\bigcap_{n=1}^{\infty}f[A_{n}]. Since (f1[y]An)n(f^{-1}[y]\cap A_{n})_{n\in\mathbb{N}} is a decreasing sequence of closed, nonempty subsets of XX; by (2), f1[y]Af^{-1}[y]\cap A\neq\varnothing and then yf[A].y\in f[A].

If (X,f)(X,f) is a discrete dynamical system with XX compact, then by Proposition 2.1, m=1fm[X]\bigcap_{m=1}^{\infty}f^{m}[X] is a nonempty compact subspace of XX satisfying

f[m=1fm[X]]=m=1fm[X].f\left[\bigcap_{m=1}^{\infty}f^{m}[X]\right]=\bigcap_{m=1}^{\infty}f^{m}[X].

This means that the restricted map fm=1fm[X]f\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X] is onto m=1fm[X]\bigcap_{m=1}^{\infty}f^{m}[X]. In the case where XX is a connected space, so is m=1fm[X]\bigcap_{m=1}^{\infty}f^{m}[X], again by Proposition 2.1.

2.1. Basic lemmas, definitions, and the first equivalences

Before delving deep into the study of dendrites and finite trees, we state two general lemmas (on arbitrary metric spaces) containing some useful consequences of the failure of equicontinuity of a map. First note that a map f:XXf\colon X\longrightarrow X fails to be equicontinuous at the point xXx\in X if and only if there exists an ε>0\varepsilon>0, a sequence of points (xk)k(x_{k})_{k\in\mathbb{N}} converging to xx, and an increasing sequence of indices (nk)k(n_{k})_{k\in\mathbb{N}} such that d(fnk(xk),fnk(x))>εd(f^{n_{k}}(x_{k}),f^{n_{k}}(x))>\varepsilon for all kk\in\mathbb{N}. In this case we will say that ε\varepsilon, (xk)k(x_{k})_{k\in\mathbb{N}}, and (nk)k(n_{k})_{k\in\mathbb{N}} witness the failure of equicontinuity of ff at xx.

Lemma 2.2.

Let XX be a metric space, let f:XXf\colon X\longrightarrow X be a map, and suppose that ff fails to be equicontinuous at xXx\in X. Then, for every nn\in\mathbb{N},

  1. (1)

    ff fails to be equicontinuous at fn(x)f^{n}(x), and

  2. (2)

    there exists an 0i<n0\leq i<n such that fnf^{n} fails to be equicontinuous at fi(x)f^{i}(x).

Proof.

Suppose that ε>0\varepsilon>0, the sequence of points (xk)k(x_{k})_{k\in\mathbb{N}}, and the sequence of indices (nk)k(n_{k})_{k\in\mathbb{N}} witness the failure of equicontinuity of ff at xx, and let nn\in\mathbb{N}. To prove (1), assume without loss of generality that n1>nn_{1}>n; then, the sequence (fn(xk))k(f^{n}(x_{k}))_{k\in\mathbb{N}} (which converges to fn(x)f^{n}(x) by continuity of the function fnf^{n}), and the increasing sequence (nkn)k(n_{k}-n)_{k\in\mathbb{N}} of natural numbers, along with ε\varepsilon, witness the failure of equicontinuity of ff at fn(x)f^{n}(x). This shows (1). For (2), apply the pigeonhole principle to assume, without loss of generality, that there is a fixed 0i<n0\leq i<n such that nkimodnn_{k}\equiv i\mod n for all kk\in\mathbb{N}. Let mkm_{k} be such that nk=nmk+in_{k}=nm_{k}+i; then, the sequence (fi(xk))k(f^{i}(x_{k}))_{k\in\mathbb{N}}, which converges to fi(x)f^{i}(x), along with the increasing sequence (mk)k(m_{k})_{k\in\mathbb{N}} of natural numbers and ε\varepsilon, witness the failure of equicontinuity of fnf^{n} at fi(x)f^{i}(x). ∎

Before considering the specific case of dendrites, we introduce some more definitions and a general result that shall be used later.

Definition 2.3.

Let XX be a metric space, and let f:XXf\colon X\longrightarrow X be a map.

  1. (1)

    The ω\omega-limit set of ff at xX,x\in X, is the set of all points yXy\in X for which there is an increasing sequence (ni)i(n_{i})_{i\in\mathbb{N}} with limifni(x)=y\lim_{i\to\infty}f^{n_{i}}(x)=y; this set is denoted by ω(x,f)\omega(x,f).

  2. (2)

    A point xXx\in X is a recurrent point if xω(x,f)x\in\omega(x,f); the set of recurrent points of ff is denoted by Rec(f)\operatorname{Rec}(f).

It is immediate that every periodic point is recurrent; the converse is not necessarily true. The next proposition follows from known results and will be used for our Main Theorem.

Proposition 2.4.

If XX is a finite tree and f:XXf\colon X\longrightarrow X is an equicontinuous surjective map, then ff is a homeomorphism which furthermore is periodic.

Proof.

By [11, Proposition 2.4] and [5, Corollary 8], cf. [11, Corollary 3.2], ff is a homeomorphism that is pointwise-recurrent, i.e., such that xω(x,f)x\in\omega(x,f) for each xXx\in X. Hence, by [12, Theorem 4.4], ff is periodic. ∎

We now mention some standard facts about dendrites that will be used throughout the paper.

Definition 2.5.

A dendrite is a locally connected continuum without simple closed curves.

A map of a dendrite into itself has a fixed point ([13, Theorem 10.31]). Every subcontinuum of a dendrite is again a dendrite ([13, Corollary 10.6]), and every connected subset of a dendrite is arcwise connected ([13, Proposition 10.9]). If XX is a dendrite, x,yXx,y\in X and xy,x\neq y, then there is a unique (closed) arc in XX joining xx and yy; such an arc will always be denoted by xyxy. Since continuous images of connected sets must be connected, for any map f:XXf\colon X\longrightarrow X and every x,yXx,y\in X we have that f(x)f(y)f[xy]f(x)f(y)\subseteq f[xy].

Whenever XX is a dendrite and YY is a subcontinuum of XX, then there exists a retraction rY:XY,r_{Y}\colon X\longrightarrow Y, called the first point function, such that for xXx\in X and yYy\in Y, rY(x)r_{Y}(x) is the first point in the arc xyxy (equipping such an arc with a linear order where xyx\leq y) that belongs to Y.Y. The mapping rYr_{Y} does not depend on the specific yYy\in Y (see [13, Lemmas 10.24, 10.25 and Terminology 10.26]).

Finally, every dendrite is uniformly locally arcwise connected, that is, for every ε>0\varepsilon>0 there exists a δ>0\delta>0 such that, whenever d(x,y)<δd(x,y)<\delta and xyx\neq y, the arc xyxy must have diameter <ε<\varepsilon (as a matter of fact, every compact, connected and locally connected metric space has this property, which in this more general context must be phrased as: for every ε>0\varepsilon>0 there exists a δ>0\delta>0 such that whenever d(x,y)<δd(x,y)<\delta and xy,x\neq y, then there is an arc joining xx and yy with diameter <ε<\varepsilon; see [19, Theorem 31.4]).

Proposition 2.6.

Let XX be a dendrite, let f:XXf\colon X\longrightarrow X be a map, and let xXx\in X. If YX{x}Y\subseteq X\setminus\{x\} is a connected component of X{x}X\setminus\{x\} such that f(x)Yf(x)\in Y, then YFix(f)Y\cap\operatorname{Fix}(f)\neq\varnothing.

Proof.

Notice that clX(Y)=Y{x}\operatorname{cl}_{X}(Y)=Y\cup\{x\} is a subcontinuum of XX –hence clX(Y)\operatorname{cl}_{X}(Y) is itself a dendrite. We consider the first point function rclX(Y):XclX(Y)r_{\operatorname{cl}_{X}(Y)}:X\longrightarrow\operatorname{cl}_{X}(Y) and note that, for yclX(Y)y\notin\operatorname{cl}_{X}(Y), it must be the case that rclX(Y)(y)=xr_{\operatorname{cl}_{X}(Y)}(y)=x. Since clX(Y)\operatorname{cl}_{X}(Y) is a dendrite, it has the fixed point property; therefore the map rclX(Y)(fclX(Y)):clX(Y)clX(Y)r_{\operatorname{cl}_{X}(Y)}\circ(f\upharpoonright\operatorname{cl}_{X}(Y))\colon\operatorname{cl}_{X}(Y)\longrightarrow\operatorname{cl}_{X}(Y) has a fixed point yy. It is now easy to check that we must have f(y)=yf(y)=y. ∎

The following is another definition that will be crucial throughout the paper.

Definition 2.7.

Let XX be a dendrite and k.k\in\mathbb{N}.

  1. (1)

    The order of a point xXx\in X is the number of connected components of X{x}X\setminus\{x\};

  2. (2)

    a point xXx\in X is an endpoint if its order is 11, and a branching point if its order is 3\geq 3;

  3. (3)

    xyXxy\subseteq X is a free arc in X,X, if no element of xy{x,y}xy\setminus\{x,y\} is a branching point;

  4. (4)

    for k3k\geq 3, XX is a kk-od if it contains exactly one branching point (called the vertex of XX), which has order kk; a 22-od is simply defined to be an arc (we do not specify a vertex in this case);

  5. (5)

    XX is a finite tree if it has only finitely many branching points and each of these branching points has a finite order.

Note that xyXxy\subseteq X is a free arc in XX if and only if xy{x,y}xy\setminus\{x,y\} is open in X.X. In a general topological space XX, the order of a point xXx\in X is defined as the least cardinal number κ\kappa such that, for every open neighbourhood UU of x,x, there exists another open neighbourhood VV with xVUx\in V\subseteq U and |(V)|κ|\partial(V)|\leq\kappa (where (V)\partial(V) denotes the boundary of VV in XX), cf. [13, Definition 9.3]; this will be important towards the end of Section 4. If, however, the topological space XX under consideration is a dendrite, then Definition 2.7 agrees with the general definition just mentioned (see [13, Lemma 10.12, Theorem 10.13 and Corollary 10.20.1]).

We now show the equivalence of the first four conditions in Theorem 1.8.

Proposition 2.8.

Let XX be a finite tree and let f:XXf\colon X\longrightarrow X be a map. Then the following conditions are equivalent:

  1. (a)

    ff is equicontinuous;

  2. (b)

    for some nn\in\mathbb{N}, the restriction fnm=1fm[X]f^{n}\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X] is the identity map;

  3. (c)

    for some nn\in\mathbb{N}, Fix(fn)=m=1fm[X]\operatorname{Fix}(f^{n})=\bigcap_{m=1}^{\infty}f^{m}[X];

  4. (d)

    Per(f)=m=1fm[X]\operatorname{Per}(f)=\bigcap_{m=1}^{\infty}f^{m}[X].

Proof.

We consider first the case where ff is surjective. Note that in such situation, X=m=1fm[X]X=\bigcap_{m=1}^{\infty}f^{m}[X] and fnm=1fm[X]=fnf^{n}\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X]=f^{n} for each nn\in\mathbb{N}. Moreover (b) asserts that ff is periodic, (c) that Fix(fn)=X\operatorname{Fix}(f^{n})=X for some nn\in\mathbb{N}, and (d) that ff is pointwise-periodic. By Proposition 2.4, (a)(b)(a)\Rightarrow(b). The implications (b)(c)(d)(b)\Rightarrow(c)\Rightarrow(d) are obvious and, by  [6, Theorem 4.14], the implication (d)(a)(d)\Rightarrow(a) holds not only on finite trees, but on every dendrite and with ff being any (surjective) map.

We now consider the case of an arbitrary (not necessarily surjective) map f:XX.f\colon X\longrightarrow X. Since every finite tree is, in particular, a finite graph, we may use [11, Theorem 5.2] to see that ff is equicontinuous if and only if so is fm=1fm[X]f\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X], and since the latter map is onto m=1fm[X]\bigcap_{m=1}^{\infty}f^{m}[X] (and since Fix(fn)=Fix(fnm=1fm[X])\operatorname{Fix}(f^{n})=\operatorname{Fix}(f^{n}\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X]) and Per(f)=Per(fm=1fm[X])\operatorname{Per}(f)=\operatorname{Per}(f\upharpoonright\bigcap_{m=1}^{\infty}f^{m}[X])), then the theorem follows from the surjective case. ∎

2.2. Expanding arcs

We now analyze some implications of the existence of expanding arcs.

Lemma 2.9.

Let XX be a dendrite and let f:XXf\colon X\longrightarrow X be a map. Then the following are equivalent:

  1. (ee^{\prime})

    XX contains an ff-expanding arc;

  2. (ff^{\prime})

    for some nn\in\mathbb{N}, the set Fix(fn)\operatorname{Fix}(f^{n}) is disconnected;

  3. (jj^{\prime})

    there exist points x,yXx,y\in X and nn\in\mathbb{N} such that x=fn(x)x=f^{n}(x), yfn(y)y\neq f^{n}(y), and yxfn(y)y\in xf^{n}(y).

Proof.

(e)(f)(e^{\prime})\Rightarrow(f^{\prime}) Let abab be an ff-expanding arc and fix an nn\in\mathbb{N} such that abfn[ab]ab\subsetneq f^{n}[ab]. We consider three cases according to whether both, exactly one, or none of a,ba,b are fixed points for fnf^{n}.

Case 1:

If fn(a)=af^{n}(a)=a and fn(b)=bf^{n}(b)=b, use the fact that abfn[ab]ab\subsetneq f^{n}[ab] to get a cab{a,b}c\in ab\setminus\{a,b\} such that fn(c)abf^{n}(c)\notin ab. In particular, cFix(fn)c\notin\operatorname{Fix}(f^{n}) with cabc\in ab and a,bFix(fn)a,b\in\operatorname{Fix}(f^{n}), showing that Fix(fn)\operatorname{Fix}(f^{n}) is not arcwise connected, so Fix(fn)\operatorname{Fix}(f^{n}) is disconnected.

Case 2:

If fn(a)=af^{n}(a)=a but fn(b)bf^{n}(b)\neq b, use the fact that abfn[ab]ab\subseteq f^{n}[ab] to get a cab{a,b}c\in ab\setminus\{a,b\} such that fn(c)=bf^{n}(c)=b. Let ZZ be the connected component of X{c}X\setminus\{c\} containing a.a. If bZb\in Z then, since ZZ is arcwise connected, we have cabZX{c},c\in ab\subseteq Z\subseteq X\setminus\{c\}, a contradiction. Hence, bZb\notin Z. Letting YZY\neq Z be the connected component of X{c}X\setminus\{c\} containing bb, Proposition 2.6 guarantees the existence of a dFix(fn)Yd\in\operatorname{Fix}(f^{n})\cap Y. Then we have cFix(fn)c\notin\operatorname{Fix}(f^{n}), a,dFix(fn)a,d\in\operatorname{Fix}(f^{n}), and cadc\in ad, showing that Fix(fn)\operatorname{Fix}(f^{n}) is disconnected.

Case 3:

If fn(a)af^{n}(a)\neq a and fn(b)bf^{n}(b)\neq b. Then (since abfn[ab]ab\subseteq f^{n}[ab]) we can find x,yab{a,b}x,y\in ab\setminus\{a,b\} with fn(x)=af^{n}(x)=a and fn(y)=bf^{n}(y)=b. Note that xyx\neq y. Equip abab with a linear order via a homeomorphism :[0,1]ab:[0,1]\longrightarrow ab mapping 0 to aa and 11 to bb. If x<yx<y, then a couple of applications of Proposition 2.6 yield fixed points c,dFix(fn)c,d\in\operatorname{Fix}(f^{n}) such that x,ycdx,y\in cd; since x,yFix(fn)x,y\notin\operatorname{Fix}(f^{n}), this shows that Fix(fn)\operatorname{Fix}(f^{n}) is disconnected. If, on the other hand, we have y<xy<x, then notice that yab=fn(x)fn(y)fn[xy]y\in ab=f^{n}(x)f^{n}(y)\subseteq f^{n}[xy], so there is a yxyy^{\prime}\in xy with fn(y)=yf^{n}(y^{\prime})=y; also, xyb=fn(y)fn(y)fn[yy]x\in yb=f^{n}(y^{\prime})f^{n}(y)\subseteq f^{n}[yy^{\prime}] and so there is an xyyx^{\prime}\in yy^{\prime} with fn(x)=xf^{n}(x^{\prime})=x. This way we have obtained x<yx^{\prime}<y^{\prime} with f2n(x)=fn(x)=af^{2n}(x^{\prime})=f^{n}(x)=a and f2n(y)=fn(y)=bf^{2n}(y^{\prime})=f^{n}(y)=b, thus, a couple of applications of Proposition 2.6 yield two fixed points c,dFix(f2n)c,d\in\operatorname{Fix}(f^{2n}) with x,ycdx^{\prime},y^{\prime}\in cd; the fact that x,yFix(f2n)x^{\prime},y^{\prime}\notin\operatorname{Fix}(f^{2n}) implies then that Fix(f2n)\operatorname{Fix}(f^{2n}) is disconnected, and we are done.

(f)(j)(f^{\prime})\Rightarrow(j^{\prime}) Let nn\in\mathbb{N} and a,bFix(fn)a,b\in\operatorname{Fix}(f^{n}) be such that abFix(fn)ab\nsubseteq\operatorname{Fix}(f^{n}). Then there is a yaby\in ab with fn(y)yf^{n}(y)\neq y. Considering the first point function rab:Xabr_{ab}\colon X\longrightarrow ab and the point z=rab(fn(y))z=r_{ab}(f^{n}(y)), we must have that either yazy\in az or ybzy\in bz. Since the situation is entirely symmetric, assume without loss of generality that yazy\in az and let x=ax=a. Then we have fn(x)=xf^{n}(x)=x and yxzxzzfn(y)=xfn(y)y\in xz\subseteq xz\cup zf^{n}(y)=xf^{n}(y).

(j)(e)(j^{\prime})\Rightarrow(e^{\prime}) Under the assumptions we have xyxfn(y)=fn(x)fn(y)fn[xy]xy\subsetneq xf^{n}(y)=f^{n}(x)f^{n}(y)\subseteq f^{n}[xy] and so xyxy is an ff-expanding arc. ∎

We now proceed to prove that the failure of equicontinuity of a map on a finite tree implies the existence of an ff-expanding arc. The next result allows us to restrict any map without expanding arcs from a finite tree to a simpler subcontinuum.

Lemma 2.10.

Let XX be a finite tree, let f:XXf\colon X\longrightarrow X be a map and let xFix(f).x\in\operatorname{Fix}(f). If XX has no ff-expanding arcs, then there is an ff-invariant subspace YXY\subseteq X (that is, f[Y]Yf[Y]\subseteq Y), where YY is a kk-od for some k2k\geq 2, such that xintX(Y)x\in\mathop{\mathrm{int}}_{X}(Y).

Proof.

Let nn be the order of xX.x\in X. Since XX is a finite tree, for some ε>0\varepsilon>0 the closed ball centred at xx is a union of finitely many arcs, say I1,,InI_{1},\ldots,I_{n}, which pairwise intersect at xx only. By continuity of f,f2,,fnf,f^{2},\ldots,f^{n}, we can pick a δ\delta with 0<δε0<\delta\leq\varepsilon such that, if d(y,z)δd(y,z)\leq\delta, then d(fi(y),fi(z))<εd(f^{i}(y),f^{i}(z))<\varepsilon for every 0in0\leq i\leq n. Let ZZ be the closed ball of radius δ\delta centred at xx, and define

Y=Zf[Z]fn[Z].Y=Z\cup f[Z]\cup\cdots\cup f^{n}[Z].

Due to our choice of δ\delta and ZZ, we have xintX(Y)Yi=1nIix\in\mathop{\mathrm{int}}_{X}(Y)\subseteq Y\subseteq\bigcup_{i=1}^{n}I_{i}. Since xx is a fixed point for ff, whenever yYy\in Y we must have xyYxy\subseteq Y. Hence YY is a kk-od for some k2k\geq 2 (k=nk=n and xx is the vertex of YY if n3n\geq 3, k=2k=2 if YY is an arc, i.e., n{1,2};n\in\{1,2\}; moreover xx is an interior point of the arc YY if n=2n=2, and it is an endpoint of the arc YY if n=1n=1). It remains to show that YY is an ff-invariant subspace, so let yYy\in Y and let us argue that f(y)Yf(y)\in Y. We can write y=fm(z)y=f^{m}(z) for some zZz\in Z (this includes the case m=0m=0, interpreted as y=zZy=z\in Z); there is nothing to do if m<nm<n, so assume that y=fn(z)y=f^{n}(z) for zZz\in Z. Looking at the finite sequence of points z,f(z),,fn(z)z,f(z),\ldots,f^{n}(z), the pigeonhole principle guarantees the existence of 0i<jn0\leq i<j\leq n and l{1,,n}l\in\{1,\ldots,n\} such that fi(z),fj(z)Ilf^{i}(z),f^{j}(z)\in I_{l}. Let us linearly order the arc IlI_{l} by copying the order of [0,1][0,1] along a homeomorphism mapping xx to 0. If fi(z)<fj(z)f^{i}(z)<f^{j}(z), this would mean that

xfi(z)xfj(z)=fji(x)fji(fi(z))fji[xfi(z)]xf^{i}(z)\subsetneq xf^{j}(z)=f^{j-i}(x)f^{j-i}(f^{i}(z))\subseteq f^{j-i}[xf^{i}(z)]

is an ff-expanding arc, a contradiction. Therefore we must have fj(z)fi(z)f^{j}(z)\leq f^{i}(z), implying that

fj(z)xfi(z)=fi(x)fi(z)fi[xz]f^{j}(z)\in xf^{i}(z)=f^{i}(x)f^{i}(z)\subseteq f^{i}[xz]

and so there exists a point zxzZz^{\prime}\in xz\subseteq Z with fj(z)=fi(z)f^{j}(z)=f^{i}(z^{\prime}). Hence y=fn(z)=fnj+i(z)y=f^{n}(z)=f^{n-j+i}(z^{\prime}) and so f(y)=fn(ji)+1(z)Yf(y)=f^{n-(j-i)+1}(z^{\prime})\in Y, and we are done. ∎

Remark 2.11.

Suppose that XX is a metric space, f:XXf\colon X\longrightarrow X is a map, and YXY\subseteq X is an ff-invariant subspace with xintX(Y)x\in\mathop{\mathrm{int}}_{X}(Y). If ff is not equicontinuous at xx, then the restriction fY:YYf\upharpoonright Y\colon Y\longrightarrow Y also fails to be equicontinuous at xx: given ε>0\varepsilon>0, if δ>0\delta>0 witnesses the equicontinuity of fYf\upharpoonright Y at xx, then min{δ,d(x,XintX(Y))}\min\{\delta,d(x,X\setminus\mathop{\mathrm{int}}_{X}(Y))\} witnesses the equicontinuity of f:XXf\colon X\longrightarrow X at xx.

For our next proof we will use a Ramsey-theoretic result known as Hindman’s theorem, so we proceed to explain the relevant concepts and notations. Given a sequence (nk)k(n_{k})_{k\in\mathbb{N}} of elements of \mathbb{N}, its set of finite sums is defined as

FS(nk)k\displaystyle\operatorname{FS}(n_{k})_{k\in\mathbb{N}} =\displaystyle= {kFnk|F is finite and nonempty}\displaystyle\left\{\sum_{k\in F}n_{k}\bigg{|}F\subseteq\mathbb{N}\text{ is finite and nonempty}\right\}
=\displaystyle= {nk1++nkm|m and k1<<km},\displaystyle\left\{n_{k_{1}}+\cdots+n_{k_{m}}\big{|}m\in\mathbb{N}\text{ and }k_{1}<\cdots<k_{m}\right\},

the set of all numbers that can be obtained by adding a finite amount of terms of the sequence (nk)k(n_{k})_{k\in\mathbb{N}} without repetitions. The result known as Hindman’s theorem ([9, Theorem 3.1]) states that for any finite partition of ,\mathbb{N}, there exists an infinite (strictly increasing) sequence (nk)k(n_{k})_{k\in\mathbb{N}} such that the set FS(nk)k\operatorname{FS}(n_{k})_{k\in\mathbb{N}} is completely contained in a single cell of the partition.

We will use a slightly stronger form of the aforementioned theorem. Given two (strictly increasing) sequences of natural numbers (nk)k(n_{k})_{k\in\mathbb{N}} and (mk)k(m_{k})_{k\in\mathbb{N}}, we say that the sequence (mk)k(m_{k})_{k\in\mathbb{N}} is a sum subsystem of the sequence (nk)k(n_{k})_{k\in\mathbb{N}} if there are finite subsets F1,F2,,Fk,F_{1},F_{2},\ldots,F_{k},\ldots of \mathbb{N} such that, for each kk\in\mathbb{N}, we have max(Fk)<min(Fk+1)\max(F_{k})<\min(F_{k+1}) and mk=jFknjm_{k}=\sum_{j\in F_{k}}n_{j}. Note that this implies in particular that FS(mk)kFS(nk)k\operatorname{FS}(m_{k})_{k\in\mathbb{N}}\subseteq\operatorname{FS}(n_{k})_{k\in\mathbb{N}}. With this terminology, we record Hindman’s theorem in the form that will be used later:

Theorem 2.12 ([10, Corollary 5.15]).

For every infinite (strictly increasing) sequence (nk)k(n_{k})_{k\in\mathbb{N}} and for every finite partition {A1,,Am}\{A_{1},\ldots,A_{m}\} of the set FS(nk)k\operatorname{FS}(n_{k})_{k\in\mathbb{N}}, there exists an i{1,,m}i\in\{1,\ldots,m\} and a sum subsystem (mk)k(m_{k})_{k\in\mathbb{N}} of the sequence (nk)k(n_{k})_{k\in\mathbb{N}} such that FS(mk)kAi\operatorname{FS}(m_{k})_{k\in\mathbb{N}}\subseteq A_{i}.

Since =FS(2k1)k\mathbb{N}=\operatorname{FS}(2^{k-1})_{k\in\mathbb{N}}, the original version of Hindman’s theorem follows immediately from Theorem 2.12 above.

Theorem 2.13.

Let XX be a finite tree and let f:XXf\colon X\longrightarrow X be a map. If ff is not equicontinuous, then XX has an ff-expanding arc.

Proof.

Take an xXx\in X such that ff is not equicontinuous at xx. We have two cases.

Case 1: The point xx is eventually periodic (i.e., there is kk\in\mathbb{N} such that fk(x)f^{k}(x) is a periodic point; equivalently, the set {fn(x)|n}\{f^{n}(x)\big{|}n\in\mathbb{N}\} is finite). This means that, replacing xx by some fk(x)f^{k}(x) if necessary (and using clause (1) of Lemma 2.2), we may assume that xx is a periodic point, say with period nn. Now use clause (2) of Lemma 2.2 to find an i<ni<n such that fnf^{n} fails to be equicontinuous at y=fi(x)y=f^{i}(x), and notice that yy is a fixed point for fnf^{n}. If XX has an fnf^{n}-expanding arc, then this is also an ff-expanding arc and we are done. If, on the contrary, there are no fnf^{n}-expanding arcs, then we can use Lemma 2.10 to obtain an fnf^{n}-invariant subspace YXY\subseteq X such that YY is a kk-od for some k2k\geq 2, and yintX(Y)y\in\mathop{\mathrm{int}}_{X}(Y). By Remark 2.11, the restricted map fnY:YYf^{n}\upharpoonright Y\colon Y\longrightarrow Y fails to be equicontinuous at yy, and so by Proposition 1.6, YY must contain an fnf^{n}-expanding arc II. Then IXI\subseteq X is also an ff-expanding arc, and we are done.

Case 2: The point xx is not eventually periodic. Then the set {fn(x)|n}\{f^{n}(x)\big{|}n\in\mathbb{N}\} is infinite. The space XX is a finite tree and so it can be written as a union of finitely many maximal free arcs I1,,ItI_{1},\ldots,I_{t} in XX such that any two distinct Ii,IjI_{i},I_{j} have at most one (branching) point in common. Now we define subsets A0,A1,,AtA_{0},A_{1},\ldots,A_{t} of \mathbb{N} as follows: nA0n\in A_{0} iff fn(x)f^{n}(x) is a branching point of XX and, for each i{1,,t}i\in\{1,\ldots,t\}, nAin\in A_{i} iff fn(x)Iif^{n}(x)\in I_{i} and fn(x)f^{n}(x) is not a branching point of X.X. Clearly

AiAj=for every i,j{0,1,,t} with ij.A_{i}\cap A_{j}=\varnothing\hskip 14.22636pt\mbox{for every }i,j\in\{0,1,\ldots,t\}\mbox{ with }i\neq j.

Given i{0,1,,t}i\in\{0,1,\ldots,t\} the set AiA_{i} can be empty. Since the set {fn(x)|n}\{f^{n}(x)\big{|}n\in\mathbb{N}\} is infinite, there exist distinct m1,m2,,mr{0,1,t}m_{1},m_{2},\ldots,m_{r}\in\{0,1,\ldots t\} such that AmjA_{m_{j}}\neq\varnothing for every j{1,2,,r}j\in\{1,2,\ldots,r\} and Ai=A_{i}=\varnothing for each i{0,1,,t}{m1,m2,,mr}i\in\{0,1,\ldots,t\}\setminus\{m_{1},m_{2},\ldots,m_{r}\}. Hence {Am1,Am2,,Amr}\{A_{m_{1}},A_{m_{2}},\ldots,A_{m_{r}}\} is a finite partition of \mathbb{N}. Theorem 2.12 provides us with a j{1,2,,r}j\in\{1,2,\ldots,r\} and an infinite strictly increasing sequence n11<<nk1<nk+11<n_{1}^{1}<\cdots<n_{k}^{1}<n_{k+1}^{1}<\cdots of natural numbers, such that the set FS(nk1)kAmj\operatorname{FS}(n_{k}^{1})_{k\in\mathbb{N}}\subseteq A_{m_{j}}. Since the points fn(x)f^{n}(x) as nn\in\mathbb{N} varies are pairwise distinct and XX only has finitely many branching points, A0A_{0} is finite. Therefore mj0m_{j}\neq 0 and so

{fn(x)|nFS(nk1)k}Imj.\{f^{n}(x)\big{|}n\in\operatorname{FS}(n_{k}^{1})_{k\in\mathbb{N}}\}\subseteq I_{m_{j}}.

Use a homeomorphism of [0,1][0,1] onto ImjI_{m_{j}} to equip ImjI_{m_{j}} with a linear order \leq. We now partition the set FS(nk1)k2\operatorname{FS}(n_{k}^{1})_{k\geq 2} according to whether fn(x)<fn11+n(x)f^{n}(x)<f^{n_{1}^{1}+n}(x) or fn11+n(x)<fn(x)f^{n_{1}^{1}+n}(x)<f^{n}(x); a further application of Theorem 2.12 allows us to obtain a sum subsystem (nk2)k2(n_{k}^{2})_{k\geq 2} of (nk1)k2(n_{k}^{1})_{k\geq 2} such that FS(nk2)k2\operatorname{FS}(n_{k}^{2})_{k\geq 2} is contained in one piece of this partition. This means that there is an R1{>,<}R_{1}\in\{>,<\} such that,

fn11+n(x)R1fn(x),for every nFS(nk2)k2.f^{n_{1}^{1}+n}(x)\ R_{1}\ f^{n}(x),\hskip 14.22636pt\mbox{for every }n\in\operatorname{FS}(n_{k}^{2})_{k\geq 2}.

Continuing this process by induction, we obtain, for each KK\in\mathbb{N}, a sum subsystem (nkK+1)kK+1(n_{k}^{K+1})_{k\geq K+1} of (nkK)kK+1(n_{k}^{K})_{k\geq K+1} and an RK{>,<}R_{K}\in\{>,<\} such that

fnKK+n(x)RKfn(x),for each nFS(nkK+1)kK+1.f^{n_{K}^{K}+n}(x)\ R_{K}\ f^{n}(x),\hskip 14.22636pt\mbox{for each }n\in\operatorname{FS}(n_{k}^{K+1})_{k\geq K+1}.

Now, an application of the pigeonhole principle allows us to obtain an infinite increasing sequence (Kk)k(K_{k})_{k\in\mathbb{N}} such that all the RKkR_{K_{k}} are equal, say, without loss of generality, to >>. What this means is that, if we define the sequence (nk)k(n_{k})_{k\in\mathbb{N}} by nk=nKkKkn_{k}=n_{K_{k}}^{K_{k}}, then for every KK\in\mathbb{N} and each nFS(nk)kK+1n\in\operatorname{FS}(n_{k})_{k\geq K+1} we have fn(x)<fnK+n(x)f^{n}(x)<f^{n_{K}+n}(x).

Now, for each KK\in\mathbb{N} we define a point yKImjy_{K}\in I_{m_{j}} by

yK=sup{fn(x)|nFS(nk)kK}.y_{K}=\sup\{f^{n}(x)\big{|}n\in\operatorname{FS}(n_{k})_{k\geq K}\}.

Since fn(x)<fnK+n(x)f^{n}(x)<f^{n_{K}+n}(x) for every nFS(nk)kK+1n\in\operatorname{FS}(n_{k})_{k\geq K+1}, we must have yK+1yKy_{K+1}\leq y_{K} for every KK\in\mathbb{N}. We may now break the proof into two further subcases (recall that dd is the metric on XX).

Subcase 2.A:

There is a KK\in\mathbb{N} such that yK+1=yKy_{K+1}=y_{K}. Let y=yK=yK+1y=y_{K}=y_{K+1} and note that, for every NN\in\mathbb{N}, there is an lNFS(nk)kK+1l_{N}\in\operatorname{FS}(n_{k})_{k\geq K+1} with

flN(x)<yandd(flN(x),y)<1N.f^{l_{N}}(x)<y\hskip 14.22636pt\mbox{and}\hskip 14.22636ptd(f^{l_{N}}(x),y)<\frac{1}{N}.

We have flN(x)<fnK+lN(x)<yf^{l_{N}}(x)<f^{n_{K}+l_{N}}(x)<y; in particular, we also have

d(fnK+lN(x),y)<1N.d(f^{n_{K}+l_{N}}(x),y)<\frac{1}{N}.

It follows that

limNflN(x)=yandlimNfnK(flN(x))=y.\lim_{N\to\infty}f^{l_{N}}(x)=y\hskip 14.22636pt\mbox{and}\hskip 14.22636pt\lim_{N\to\infty}f^{n_{K}}(f^{l_{N}}(x))=y.

By continuity of the function fnKf^{n_{K}}, we may conclude that fnK(y)=yf^{n_{K}}(y)=y. Thus, all the points

y,f(y),,fnK1(y)y,f(y),\ldots,f^{n_{K}-1}(y)

are fixed points of the map fnKf^{n_{K}}. If XX contains an fnKf^{n_{K}}-expanding arc, then this arc is also ff-expanding and we are done, so assume otherwise. Then we may apply Lemma 2.10 to each of the points y,f(y),,fnK1(y)y,f(y),\ldots,f^{n_{K}-1}(y) to get fnKf^{n_{K}}-invariant subcontinua

Y0,Y1,,YnK1XY_{0},Y_{1},\ldots,Y_{n_{K}-1}\subseteq X

such that, for every i{0,1,,nK1},i\in\{0,1,\ldots,n_{K}-1\}, we have fi(y)intX(Yi)f^{i}(y)\in\mathop{\mathrm{int}}_{X}(Y_{i}) and each YiY_{i} is a kik_{i}-od for some ki2k_{i}\geq 2. Let ε>0\varepsilon>0 be such that for every i<nKi<n_{K}, the ball centered at fi(y)f^{i}(y) with radius ε\varepsilon is contained in Yi.Y_{i}. By the continuity of the functions f,f2,,fnK1f,f^{2},\ldots,f^{n_{K}-1} we get a δ>0\delta>0 such that, if d(z,y)<δd(z,y)<\delta, then d(fi(z),fi(y))<ε.d(f^{i}(z),f^{i}(y))<\varepsilon. Hence, for each i<nKi<n_{K}, if d(z,y)<δd(z,y)<\delta then fi(z)intX(Yi)f^{i}(z)\in\mathop{\mathrm{int}}_{X}(Y_{i}). Note that yy contains points of the form fn(x)f^{n}(x) arbitrarily close, and all the points of the form fn(x)f^{n}(x) are points where the map ff is not equicontinuous (by clause (1) of Lemma 2.2). Hence we can find a zz with d(z,y)<δd(z,y)<\delta such that ff is not equicontinuous at zz; now use clause (2) of Lemma 2.2 to get i<nKi<n_{K} such that fnKf^{n_{K}} is not equicontinuous at fi(z)intX(Yi)f^{i}(z)\in\mathop{\mathrm{int}}_{X}(Y_{i}). Since YiY_{i} is fnKf^{n_{K}}-invariant, we may conclude that fnKYif^{n_{K}}\upharpoonright Y_{i} is not an equicontinuous map (see Remark 2.11). Since YiY_{i} is a kik_{i}-od, by Proposition 1.6, the subcontinuum YiY_{i} of XX must have an fnKf^{n_{K}}-expanding arc, and we are done.

Subcase 2.B:

yK+1<yKy_{K+1}<y_{K} for every K.K\in\mathbb{N}. Then let y=inf{yK|K}y=\inf\{y_{K}\big{|}K\in\mathbb{N}\}. For each KK\in\mathbb{N} fix an mKFS(nk)kKm_{K}\in\operatorname{FS}(n_{k})_{k\geq K} such that yK+1<fmK(x)<yKy_{K+1}<f^{m_{K}}(x)<y_{K}. If for some KK\in\mathbb{N}, it is not the case that y<fnK(y)y<f^{n_{K}}(y), then we must have

yfmK+1(y)fnK(y)fnK+mK+1(x)fnK[yfmK+1(x)]yf^{m_{K+1}}(y)\subsetneq f^{n_{K}}(y)f^{n_{K}+m_{K+1}}(x)\subseteq f^{n_{K}}[yf^{m_{K+1}}(x)]

and therefore there is an ff-expanding arc and we are done, so assume that for all KK\in\mathbb{N} we have y<fnK(y)y<f^{n_{K}}(y). The points fmk(x)f^{m_{k}}(x) for k>Kk>K are arbitrarily close to yy and they all satisfy fnK(fmk(x))=fnK+mk(x)yKf^{n_{K}}(f^{m_{k}}(x))=f^{n_{K}+m_{k}}(x)\leq y_{K}, so by continuity of fnKf^{n_{K}} we have fnK(y)yKf^{n_{K}}(y)\leq y_{K}.

We define connected subspaces Y1,Y2XY_{1},Y_{2}\subseteq X as follows. Y1Y_{1} is the connected component of X{fm2(x)}X\setminus\{f^{m_{2}}(x)\} that does not contain yy, and Y2Y_{2} is the connected component of X{fn3(y)}X\setminus\{f^{n_{3}}(y)\} containing yy. Since

y<fn3(y)y3<fm2(x)<fn1+m2(x),y<f^{n_{3}}(y)\leq y_{3}<f^{m_{2}}(x)<f^{n_{1}+m_{2}}(x),

and all such points belong to the maximal free arc ImjI_{m_{j}} in XX, we can write

X=Y1fn3(y)fm2(x)Y2,X=Y_{1}\cup f^{n_{3}}(y)f^{m_{2}}(x)\cup Y_{2},

and the union is disjoint. Since fn1+m2(x)Y1f^{n_{1}+m_{2}}(x)\in Y_{1}, by Proposition 2.6 there is a z1Y1Fix(fn1)z_{1}\in Y_{1}\cap\operatorname{Fix}(f^{n_{1}}); now if we let KK be sufficiently large that yK<fn3(y)y_{K}<f^{n_{3}}(y) then we will have fnK(y)Y2f^{n_{K}}(y)\in Y_{2} and so again by Proposition 2.6 there exists a z2Y2Fix(fnKn3)z_{2}\in Y_{2}\cap\operatorname{Fix}(f^{n_{K}-n_{3}}). Letting N=(nKn3)n1N=(n_{K}-n_{3})n_{1}, we get that z1,z2Fix(fN)z_{1},z_{2}\in\operatorname{Fix}(f^{N}), and fm2(x)z1z2Fix(fN)f^{m_{2}}(x)\in z_{1}z_{2}\setminus\operatorname{Fix}(f^{N}). Hence Fix(fN)\operatorname{Fix}(f^{N}) is a disconnected set, and so by Lemma 2.9, XX must have an ff-expanding arc.

3. The Ellis remainder and ultrafilter-limits

In this section we introduce the notion of ultrafilter-limits and point out the relation of this concept with that of the Ellis remainder, with the objective of establishing the equivalence of items (a), (e), (h) and (i) from Theorem 1.8.

Definition 3.1.
  1. (1)

    An ultrafilter on \mathbb{N} is a family uu of subsets of \mathbb{N} such that

    1. (a)

      uu is nonempty and u\varnothing\notin u;

    2. (b)

      if A,Bu,A,B\in u, then ABuA\cap B\in u;

    3. (c)

      if AuA\in u and AB,A\subseteq B\subseteq\mathbb{N}, then BuB\in u;

    4. (d)

      whenever =AB\mathbb{N}=A\cup B, then either AuA\in u or BuB\in u.

  2. (2)

    An ultrafilter uu on \mathbb{N} is principal if there exists an nn\in\mathbb{N} such that u={A|nA}u=\{A\subseteq\mathbb{N}\big{|}n\in A\}; otherwise we say that uu is nonprincipal.

  3. (3)

    We use the symbol β\beta\mathbb{N} to denote the set of all ultrafilters on \mathbb{N}, and we denote with \mathbb{N}^{*} the set of all nonprincipal ultrafilters on \mathbb{N}.

  4. (4)

    Given a metric space (X,d)(X,d), a sequence (xn)n(x_{n})_{n\in\mathbb{N}} of points on XX, and an ultrafilter uβu\in\beta\mathbb{N}, we say that xx is the uu ultrafilter-limit of (xn)n,(x_{n})_{n\in\mathbb{N}}, in symbols x=u-limnxn,x=u\text{-lim}_{n\to\infty}x_{n}, if for every ε>0\varepsilon>0 the set {n|d(x,xn)<ε}u\{n\in\mathbb{N}\big{|}d(x,x_{n})<\varepsilon\}\in u.

  5. (5)

    Given a metric space XX, a function f:XXf\colon X\longrightarrow X, and an ultrafilter uβu\in\beta\mathbb{N}, we define the uu ultrafilter-limit function fu:XXf^{u}\colon X\longrightarrow X (also called the uu-th iterate of ff) by fu(x)=u-limnfn(x)f^{u}(x)=u\text{-lim}_{n\to\infty}f^{n}(x).

Given a compact metric space X,X, a map f:XXf\colon X\longrightarrow X and xXx\in X, it can be shown that

ω(x,f)={fu(x)|u}.\omega(x,f)=\{f^{u}(x)\big{|}u\in\mathbb{N}^{*}\}.

A few comments about the above definitions are in order. For each nn\in\mathbb{N}, it is common to identify the natural number nn with the principal ultrafilter un={A|nA}u_{n}=\{A\subseteq\mathbb{N}\big{|}n\in A\}; this way we can think of \mathbb{N} as a subset of β\beta\mathbb{N}, and we have =β\mathbb{N}^{*}=\beta\mathbb{N}\setminus\mathbb{N}. Furthermore, one can topologize β\beta\mathbb{N} by declaring the sets A¯={uβ|Au}\bar{A}=\{u\in\beta\mathbb{N}\big{|}A\in u\} to be open, for each AA\subseteq\mathbb{N}; this endows β\beta\mathbb{N} with a compact Hausdorff topology containing \mathbb{N} as a discrete dense subspace ([10, Lemma 3.17 and Theorems 3.18 and 3.28]). Regarding the concept of a uu-limit, it is worth pointing out that, in a compact metric space XX, every sequence (xn)n(x_{n})_{n\in\mathbb{N}} of points will have a unique uu-limit (for every uβu\in\beta\mathbb{N})([10, Theorem 3.48]). Moreover, if unu_{n} is the principal ultrafilter {A|nA}\{A\subseteq\mathbb{N}\big{|}n\in A\}, then un-limmxm=xnu_{n}\text{-lim}_{m\to\infty}x_{m}=x_{n}; similarly (and as a consequence of the above), for a function f:XXf\colon X\longrightarrow X we will have that fun=fnf^{u_{n}}=f^{n}. Thus, no confusion should arise if we sometimes abuse notation and write nn instead of unu_{n}.

Furthermore, it is possible to equip β\beta\mathbb{N} with a right-topological semigroup operation, denoted by ++. That is, ++ is an associative binary operation on β\beta\mathbb{N} such that, for each fixed uβu\in\beta\mathbb{N}, the function vu+vv\longmapsto u+v is continuous. The operation is given by the formula

u+v={A|{n|{m|n+mA}v}u}.u+v=\{A\subseteq\mathbb{N}\big{|}\{n\in\mathbb{N}\big{|}\{m\in\mathbb{N}\big{|}n+m\in A\}\in v\}\in u\}.

This operation extends the usual sum on \mathbb{N}, in the sense that, if n,mn,m\in\mathbb{N} and un,umu_{n},u_{m} are the corresponding principal ultrafilters, then un+um=un+mu_{n}+u_{m}=u_{n+m}, although ++ is not commutative on all of β\beta\mathbb{N}. It is possible to verify that, for any u,vβu,v\in\beta\mathbb{N}, we have fufv=fu+vf^{u}\circ f^{v}=f^{u+v} (see [3, p. 38]).

As we mentioned in the Introduction, the equation

E(X,f)={fu|uβ},E(X,f)=\{f^{u}\big{|}u\in\beta\mathbb{N}\},

shown in [8, Theorem 2.2] and which holds for every map f:XXf\colon X\longrightarrow X on a compact metric space XX, is the main reason why obtaining information about the ultrafilter-limit functions fuf^{u} has a great deal of importance within the study of the discrete dynamical system (X,f)(X,f). At this moment, we aim to prove that the existence of expanding arcs implies the discontinuity of ultrafilter-limit functions. We begin by introducing a definition that will help to expedite the statement of the subsequent lemmas.

Definition 3.2.

Let XX be a metric space.

  1. (1)

    Let IXI\subseteq X be an arc, and let (xn)n(x_{n})_{n\in\mathbb{N}} be a sequence of elements of II. We say that the sequence is II-monotone if it is monotone (i.e., either increasing or decreasing) when viewed as a sequence on the unit interval [0,1][0,1] via a homeomorphism :I[0,1]:I\longrightarrow[0,1]. Equivalently, the sequence (xn)n(x_{n})_{n\in\mathbb{N}} is monotone if xn+1xnxn+2x_{n+1}\in x_{n}x_{n+2} for each nn\in\mathbb{N} (noting that xnxn+2Ix_{n}x_{n+2}\subseteq I).

  2. (2)

    If g:XXg\colon X\longrightarrow X is a map, a sequence (xn)n(x_{n})_{n\in\mathbb{N}} of elements of some arc IXI\subseteq X is said to be gg-backwards if it is II-monotone and for each nn\in\mathbb{N} we have g(xn+1)=xng(x_{n+1})=x_{n}.

Remark 3.3.

Note that, by compactness of an arc and monotonicity of backward sequences, any gg-backward sequence on a dendrite is always convergent. Furthermore, the limit of the sequence is a fixed point of gg.

Lemma 3.4.

Let XX be a dendrite and let f:XXf\colon X\longrightarrow X be a map, and suppose that there is an ff-expanding arc IXI\subseteq X. Then the following two conditions hold:

  1. (1)

    for some mm\in\mathbb{N} there exists an fmf^{m}-backward sequence (yn)n(y_{n})_{n\in\mathbb{N}} in II;

  2. (2)

    the set Per(f)\operatorname{Per}(f) is disconnected.

Proof.

If there is an ff-expanding arc II in XX then, by Lemma 2.9, there exist points x,yXx,y\in X and nn\in\mathbb{N} such that fn(x)=xf^{n}(x)=x, fn(y)yf^{n}(y)\neq y, and yxfn(y)fn[xy]y\in xf^{n}(y)\subseteq f^{n}[xy], so we can find a y1xy{y}y_{1}\in xy\setminus\{y\} such that fn(y1)=yf^{n}(y_{1})=y. Now y1xy=fn(x)fn(y1)fn[xy1]y_{1}\in xy=f^{n}(x)f^{n}(y_{1})\subseteq f^{n}[xy_{1}], so there exists a y2xy1{y1}y_{2}\in xy_{1}\setminus\{y_{1}\} such that fn(y2)=y1f^{n}(y_{2})=y_{1}. Continuing by induction, if we already know y1,,yky_{1},\ldots,y_{k} with fn(yi)=yi1f^{n}(y_{i})=y_{i-1} and yixyi1{yi1}y_{i}\in xy_{i-1}\setminus\{y_{i-1}\}, then

ykxyk1=fn(x)fn(yk)fn[xyk],y_{k}\in xy_{k-1}=f^{n}(x)f^{n}(y_{k})\subseteq f^{n}[xy_{k}],

and so there exists a yk+1xyk{yk}y_{k+1}\in xy_{k}\setminus\{y_{k}\} such that fn(yk+1)=ykf^{n}(y_{k+1})=y_{k}. This way we obtain a sequence (yn)n(y_{n})_{n\in\mathbb{N}} which is fnf^{n}-backward. So (1) holds.

To show (2), we use the points x,yx,y and the sequence (yn)n(y_{n})_{n\in\mathbb{N}} obtained in (1). Since yxfn(y)y\in xf^{n}(y), by Proposition 2.6 there is a point zFix(fn)z\in\operatorname{Fix}(f^{n}) such that yxzy\in xz; then we have x,zPer(f)x,z\in\operatorname{Per}(f), so it suffices to show that xzPer(f)xz\setminus\operatorname{Per}(f)\neq\varnothing. If yPer(f)y\notin\operatorname{Per}(f) we are done, so assume that y=Per(f)y=\operatorname{Per}(f), say with period kk. Then {fn(y)|n}={y,f(y),,fk1(y)}\{f^{n}(y)\big{|}n\in\mathbb{N}\}=\{y,f(y),\ldots,f^{k-1}(y)\}; since the yny_{n} are pairwise distinct we can choose an nn\in\mathbb{N} such that yn{y,f(y),,fk1(y)}y_{n}\notin\{y,f(y),\ldots,f^{k-1}(y)\}. Then

fm(yn){yn1,,y1,y,f(y),,fk1(y)}for all m,f^{m}(y_{n})\in\{y_{n-1},\ldots,y_{1},y,f(y),\ldots,f^{k-1}(y)\}\hskip 14.22636pt\mbox{for all }m\in\mathbb{N},

thus for each mm\in\mathbb{N} we have fm(yn)ynf^{m}(y_{n})\neq y_{n} and so ynPer(f)y_{n}\notin\operatorname{Per}(f). Since ynxzy_{n}\in xz, the proof is finished. ∎

Corollary 3.5.

For a map f:XXf\colon X\longrightarrow X with XX a dendrite, the following are equivalent:

  1. (e)

    there is no ff-expanding arc in XX;

  2. (g)

    the set Per(f)\operatorname{Per}(f) is connected.

Proof.

Suppose that Per(f)\operatorname{Per}(f) is disconnected, and find x,yPer(f)x,y\in\operatorname{Per}(f) such that there exists a zxyPer(f)z\in xy\setminus\operatorname{Per}(f). If xx has period nn and yy has period mm, then we have x,yFix(fnm)x,y\in\operatorname{Fix}(f^{nm}); as zz is not periodic, we have zxyFix(fnm)z\in xy\setminus\operatorname{Fix}(f^{nm}). Hence the set Fix(fnm)\operatorname{Fix}(f^{nm}) is disconnected and so, by Lemma 2.9, XX contains an ff-expanding arc. Conversely, if XX contains an ff-expanding arc, use Lemma 3.4. ∎

Now, in order to use gg-backward sequences to deduce discontinuity of elements in the Ellis remainder, we will introduce a fairly stronger definition that allows us to work in a slightly more general context. In what follows, it will be convenient that the indexing of our sequences starts at 0 rather than at 1.

Definition 3.6.

Let XX be a compact metric space, and let g:XXg\colon X\longrightarrow X be a map. A sequence (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} of elements of XX will be said to be gg-divergent if the following three conditions hold:

  1. (1)

    x=limnxnx=\lim_{n\to\infty}x_{n} exists in XX;

  2. (2)

    for each nn\in\mathbb{N}, g(xn+1)=xng(x_{n+1})=x_{n} (this implies that g(x)=xg(x)=x);

  3. (3)

    there exists an open neighbourhood UXU\subseteq X of xx such that U{gn(x0)|n}=U\cap\{g^{n}(x_{0})\big{|}n\in\mathbb{N}\}=\varnothing.

It is not hard to see that gg-divergent sequences can only exist if gg fails to be equicontinuous. As a matter of fact, much more is true, as seen in the following theorem.

Theorem 3.7.

Let XX be an arbitrary compact metric space and let g:XXg\colon X\longrightarrow X be a map. If there is an mm\in\mathbb{N} such that XX contains a gmg^{m}-divergent sequence, then for every nonprincipal ultrafilter uu\in\mathbb{N}^{*}, the function gug^{u} is discontinuous.

Proof.

Let (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} be the hypothesized gmg^{m}-divergent sequence, let x=limkxkx=\lim_{k\to\infty}x_{k}, and let UU be an open set containing xx such that U{gmn(x0)|n}=U\cap\{g^{mn}(x_{0})\big{|}n\in\mathbb{N}\}=\varnothing.

Now let uu\in\mathbb{N}^{*} be an arbitrary nonprincipal ultrafilter. There exists a unique 0i<m0\leq i<m such that m+ium\mathbb{N}+i\in u, so that muim\mathbb{N}\in u-i. This means that it makes sense to consider the Rudin–Keisler image vv of the ultrafilter uiu-i under the mapping :m\colon m\mathbb{N}\longrightarrow\mathbb{N} given by mkkmk\longmapsto k. So we have that mv+i=umv+i=u (where mvmv denotes the Rudin–Keisler image of the ultrafilter vv under the mapping kmkk\longmapsto mk).

Define a new sequence (yn)n{0}(y_{n})_{n\in\mathbb{N}\cup\{0\}} by letting yn=gmi(xn)y_{n}=g^{m-i}(x_{n}), and let y=gmi(x)y=g^{m-i}(x). Since (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} converges to xx and gmig^{m-i} is continuous, (yn){0}(y_{n})_{\mathbb{N}\cup\{0\}} will converge to yy. We now proceed to observe that

gu(y)\displaystyle g^{u}(y) =\displaystyle= gmv+i(gmi(x))=(gm)v(gi(gmi(x)))\displaystyle g^{mv+i}(g^{m-i}(x))=(g^{m})^{v}(g^{i}(g^{m-i}(x)))
=\displaystyle= (gm)v(gm(x))=(gm)v(x)=xU,\displaystyle(g^{m})^{v}(g^{m}(x))=(g^{m})^{v}(x)=x\in U,

and, for each kk\in\mathbb{N}, we have

gu(yk)=gmv+i(gmi(xk))=(gm)v(gm(xk))=gmv(xk1).g^{u}(y_{k})=g^{mv+i}(g^{m-i}(x_{k}))=(g^{m})^{v}(g^{m}(x_{k}))=g^{mv}(x_{k-1}).

By definition of ultrafilter-limits, gmv(xk1)g^{mv}(x_{k-1}) must be an accumulation point of the set {gmn(xk1)|n}\{g^{mn}(x_{k-1})\big{|}n\in\mathbb{N}\}. However, for n>k1n>k-1 we have gmn(xk1)=gm(nk+1)(x0)Ug^{mn}(x_{k-1})=g^{m(n-k+1)}(x_{0})\notin U, so gu(yk)Ug^{u}(y_{k})\notin U for every kk\in\mathbb{N}, and therefore the sequence (gu(yk))k(g^{u}(y_{k}))_{k\in\mathbb{N}} does not converge to x=gu(y)x=g^{u}(y), showing that the function gug^{u} is discontinuous at yy, and we are done. ∎

The previous lemma works for every compact metric space. For certain dendrites, there is a relation between gg-backwards sequences and gg-divergent sequences.

Lemma 3.8.

Let XX be a dendrite with only finitely many branching points, and let g:XXg\colon X\longrightarrow X be a map. If there is an arc IXI\subseteq X such that II contains a gg-backwards sequence, then there exists an mm\in\mathbb{N} such that XX has a gmg^{m}-divergent sequence.

Proof.

Let (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} be a gg-backwards sequence in the arc II, and let x=limnxnx=\lim_{n\to\infty}x_{n}. Notice that xx is a fixed point of gg, and therefore limkgk(x)=x\lim_{k\to\infty}g^{k}(x)=x.

Now let us fix some notation. First of all, since XX has only finitely many branching points, we may shrink II and drop finitely many terms of the sequence (and shift indices afterwards so that our sequence indexing still starts at 0) (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} to ensure that II is a free arc in XX. Now linearly order the arc II via a homeomorphism with [0,1][0,1] in such a way that x0<xx_{0}<x. Then the monotonicity of the gg-backwards sequence (xn)n{0}(x_{n})_{n\in\mathbb{N}\cup\{0\}} means in this case that the sequence is increasing. Now let rI:XIr_{I}\colon X\longrightarrow I be the first point function from XX onto the subcontinuum II of XX. We will analyze the gg-orbit of x0.x_{0}. There are two cases to consider.

Case 1: For every mm\in\mathbb{N}, rI(gm(x0))x0r_{I}(g^{m}(x_{0}))\leq x_{0}. In this case, for each fixed n{0}n\in\mathbb{N}\cup\{0\} we have that gm+n(xn)=gm(x0)g^{m+n}(x_{n})=g^{m}(x_{0}), and so rI(gm+n(xn))x0<xr_{I}(g^{m+n}(x_{n}))\leq x_{0}<x for every mm\in\mathbb{N}. Since XX is a dendrite and hence uniformly locally arcwise connected, there must be a δ>0\delta>0 such that, whenever d(x,z)<δd(x,z)<\delta and xzx\neq z, then the arc xzxz must have diameter smaller than that of the arc x0xx_{0}x. In particular, if rI(z)x0r_{I}(z)\leq x_{0}, then d(x,z)δd(x,z)\geq\delta. So if we let UU be the ball centred at xx with radius δ\delta, then for every nn\in\mathbb{N} it is the case that gn(x0)Ug^{n}(x_{0})\notin U, and consequently the sequence (xn)n(x_{n})_{n\in\mathbb{N}} itself is already gg-divergent.

Case 2: There exists an mm\in\mathbb{N} such that x0rI(gm(x0))x_{0}\leq r_{I}(g^{m}(x_{0})). Fix one such mm, and notice that the function rI(gmI):IIr_{I}\circ(g^{m}\upharpoonright I):I\longrightarrow I satisfies

rI(gm(xm))=rI(x0)=x0xmandx0rI(gm(x0)).r_{I}(g^{m}(x_{m}))=r_{I}(x_{0})=x_{0}\leq x_{m}\hskip 14.22636pt\mbox{and}\hskip 14.22636ptx_{0}\leq r_{I}(g^{m}(x_{0})).

Therefore (by a standard result for maps in the unit interval) this map must have a fixed point in x0xmx_{0}x_{m}, that is, there is a z0x0xmz_{0}\in x_{0}x_{m} with z0=rI(gm(z0))z_{0}=r_{I}(g^{m}(z_{0})). Since II is a free arc in XX, we have that rI(w)r_{I}(w) is one of the endpoints of II whenever wIw\notin I. Since z0x0xm{x0,xm}z_{0}\in x_{0}x_{m}\setminus\{x_{0},x_{m}\} (so z0z_{0} is an interior point of II), from z0=rI(gm(z0))z_{0}=r_{I}(g^{m}(z_{0})) it follows that z0=gm(z0)z_{0}=g^{m}(z_{0}) and so z0z_{0} is actually a fixed point of the map gmg^{m}.

Now x0xm=gm(xm)gm(x2m)gm[xmx2m]x_{0}x_{m}=g^{m}(x_{m})g^{m}(x_{2m})\subseteq g^{m}[x_{m}x_{2m}], so there must exist a z1xmx2mz_{1}\in x_{m}x_{2m} such that gm(z1)=z0g^{m}(z_{1})=z_{0}. We continue this process by induction: given a

znxnmx(n+1)m=gm(x(n+1)m)gm(x(n+2)m)gm[x(n+1)mx(n+2)m],z_{n}\in x_{nm}x_{(n+1)m}=g^{m}(x_{(n+1)m})g^{m}(x_{(n+2)m})\subseteq g^{m}[x_{(n+1)m}x_{(n+2)m}],

we find a zn+1x(n+1)mx(n+2)mz_{n+1}\in x_{(n+1)m}x_{(n+2)m} such that gm(zn+1)=zng^{m}(z_{n+1})=z_{n}. This way we obtain a monotone sequence (zn)n{0}(z_{n})_{n\in\mathbb{N}\cup\{0\}}, with limit xx, which is gmg^{m}-backwards and where z0Fix(gm)z_{0}\in\operatorname{Fix}(g^{m}). Since z0xz_{0}\neq x, any open set UU containing xx and not containing z0z_{0} will satisfy (gm)n(z0)=z0U(g^{m})^{n}(z_{0})=z_{0}\notin U, for every nn\in\mathbb{N}. Therefore the sequence (zn)n{0}(z_{n})_{n\in\mathbb{N}\cup\{0\}} is gmg^{m}-divergent. ∎

We are ready to prove the Main Theorem of this paper.

Proof of Main Theorem 1.8 (and of clause (1) of Remark 1.9).

The equivalence of (a), (b), (c) and (d) is established in Proposition 2.8. The equivalence of (e) and (f) is Lemma 2.9, and that of (e) and (g) is Corollary 3.5; in both cases this equivalence works for arbitrary dendrites. Finally, (e) implies (a) by Theorem 2.13; (a) implies (h) easily (by the remark in the Introduction right after Definition 1.2), and it is obvious that (h) implies (i) and that (d) implies (g). We also have that (i) implies (e): by contrapositive, if there exists an ff-expanding arc in XX then there is an fmf^{m}-backward sequence for some mm, by Lemma 3.4; this yields an nn\in\mathbb{N} such that there is an fmnf^{mn}-divergent sequence by Lemma 3.8, and this in turn implies that there is no uu\in\mathbb{N}^{*} such that fuf^{u} is continuous, by Theorem 3.7. The last chain of implications establishes the equivalence of (a) with (e), (h) and (i), which finishes the proof. ∎

4. Examples and open problems

This section contains examples showing that the previous results cannot be extended to other kinds of dendrites. Theorem 1.8 holds for finite trees, and trees are dendrites satisfying two additional conditions: that they have finitely many branching points, and that each branching point has finite order. We show examples of dendrites where one of these two conditions fails. Afterwards, we finish the paper by making a few observations about functions defined on finite graphs.

4.1. Dendrites with finitely many branching points

In this subsection we proceed to exhibit an example of a dendrite, and two maps defined on it, which together show that none of the equivalences between (a) and (b), (c), or any of (e)-(i) from Theorem 1.8 are generalizable to dendrites with finitely many branching points (meaning that the hypothesis that all branching points are of finite order is really necessary in Theorem 1.8).

Remark 4.1.

A few words regarding the equivalence between (a) and (d) are in order. By [6, Theorem 4.12] together with [16, Lemma 2.6], the implication from (a) to (d) still holds if XX is merely a dendrite with finitely many branching points. The reverse implication holds for an arbitrary dendrite and a surjective map by [11, Theorem 5.2]. Surjectivity, however, is necessary: Sun et. al. [15, Example 2.9] exhibit an example of a dendrite XX and a map f:XXf:X\longrightarrow X such that ff fails to be equicontinuous (although fn=1fn[X]f\upharpoonright\bigcap_{n=1}^{\infty}f^{n}[X] is equicontinuous) yet Per(f)=n=1fn[X]\operatorname{Per}(f)=\bigcap_{n=1}^{\infty}f^{n}[X]. The dendrite in this example has infinitely many branching points; we do not know of an example of this phenomenon with a dendrite that has finitely many branching points (cf. Question 4.7).

Example 4.2.

A dendrite XX with a unique branching point, which has infinite order, and maps f,g:XXf,g\colon X\longrightarrow X such that ff satisfies all conditions from (e) through (i) of Theorem 1.8 but fails to be equicontinuous, while gg is equicontinuous but fails to satisfy conditions (b) and (c) of Theorem 1.8.

For other purposes, the dendrite XX together with the map ff, appear in [6, Example 5.1]. We reproduce their description here for three reasons: for the reader’s convenience, to point out a few observations about the map ff that are not made in [6], and in order to be able to also describe the map gg. We build XX by taking infinitely many disjoint arcs indexed by \mathbb{Z}, {In|n}\{I_{n}\big{|}n\in\mathbb{Z}\}, with each InI_{n} of length 12|n|\frac{1}{2^{|n|}}, and identifying in a single point vv (the vertex) one end of each InI_{n}. The result X=nInX=\bigcup_{n\in\mathbb{Z}}I_{n} is a dendrite with a single infinite-order branching point vv, as in Figure 1.

Refer to caption
Figure 1. The dendrite XX that has vv as its only branching point of infinite order.

Now, for each nn\in\mathbb{Z}, let us consider a map hn:InIn+1h_{n}\colon I_{n}\longrightarrow I_{n+1} defined by fixing vv, and, for each e(0,12|n|]e\in(0,\frac{1}{2^{|n|}}], if xx is the unique element of In{v}I_{n}\setminus\{v\} at distance ee from vv, then hn(x)h_{n}(x) is the unique element of In+1{v}I_{n+1}\setminus\{v\} at distance 2n|n|e2^{-\frac{n}{|n|}}e (at distance e2\frac{e}{2} in the case n=0n=0) from vv. Note that hnh_{n} maps InI_{n} homeomorphically onto In+1I_{n+1}.

We let f:XXf\colon X\longrightarrow X be defined by f=nhnf=\bigcup_{n\in\mathbb{Z}}h_{n}, that is, by f(v)=vf(v)=v and f(x)=hn(x)f(x)=h_{n}(x) whenever nn is the unique element in \mathbb{Z} so that xIn{v}x\in I_{n}\setminus\{v\}. The sequence (xk)k(x_{k})_{k\in\mathbb{N}}, where xkx_{k} is the endpoint of IkI_{-k} that is distinct from vv (this sequence converges to vv), together with ε=12\varepsilon=\frac{1}{2} and the sequence of indices (nk)k(n_{k})_{k\in\mathbb{N}} given by nk=kn_{k}=k, witness the failure of the equicontinuity of ff at vv (since fnk(xk)=x0f^{n_{k}}(x_{k})=x_{0}, where x0x_{0} is the endpoint of I0I_{0} that is distinct from vv). For each nn\in\mathbb{N}, we have

(4.1) Fix(fn)=Per(f)={v},\operatorname{Fix}(f^{n})=\operatorname{Per}(f)=\{v\},

which is a connected set. Using (4.1) it is straightforward to see that property (j)(j^{\prime}) of Lemma 2.9 is not satisfied. Hence, by the same lemma, XX has no ff-expanding arcs. Note that

ω(v,f)={v}X=m=1fm[X].\omega(v,f)=\{v\}\subsetneq X=\bigcap_{m=1}^{\infty}f^{m}[X].

Moreover, for every xXx\in X we have limnfn(x)=v\lim_{n\to\infty}f^{n}(x)=v, which implies that, for each nonprincipal ultrafilter uu\in\mathbb{N}^{*}, fu:XXf^{u}\colon X\longrightarrow X is the map with constant value vv, which is continuous.

We now proceed to describe the map gg. We stipulate that gn2Ing\upharpoonright\bigcup_{n\leq 2}I_{n} is the identity map. For each positive m{2n|n{0}}m\in\mathbb{N}\setminus\{2^{n}\big{|}n\in\mathbb{N}\cup\{0\}\}, we let gIm=hmg\upharpoonright I_{m}=h_{m}; finally, we let

gIm=hm(2n11)1hm(2n12)1hm11,whenever m=2n with n2.g\upharpoonright I_{m}=h_{m-(2^{n-1}-1)}^{-1}\circ h_{m-(2^{n-1}-2)}^{-1}\circ\cdots\circ h_{m-1}^{-1},\hskip 5.69046pt\mbox{whenever }m=2^{n}\mbox{ with }n\geq 2.

Hence we have gI2n:I2nI2n1+1g\upharpoonright I_{2^{n}}:I_{2^{n}}\longrightarrow I_{2^{n-1}+1}. Therefore, for every n2n\geq 2, the map gg will cyclically permute the finite sequence of arcs (I2n1+1,I2n1+2,,I2n)(I_{2^{n-1}+1},I_{2^{n-1}+2},\ldots,I_{2^{n}}), in such a way that f2n1i=2n1+12nIif^{2^{n-1}}\upharpoonright\bigcup_{i=2^{n-1}+1}^{2^{n}}I_{i} is the identity map (and gg will fix every point in each of the ImI_{m} for mm\in\mathbb{Z} with m2m\leq 2). As a result of this, we will have that Per(g)=X\operatorname{Per}(g)=X, and so gg will be equicontinuous by [6, Theorem 4.14]; at the same time, although ff is pointwise-periodic, XX contains points of arbitrarily high period (if xImx\in I_{m} for 2n1+1m2n2^{n-1}+1\leq m\leq 2^{n}, n{1}n\in\mathbb{N}\setminus\{1\}, then the period of xx is equal to 2n12^{n-1}) and therefore, for every nn\in\mathbb{N}, we have Fix(gn)m=1gm[X]\operatorname{Fix}(g^{n})\neq\bigcap_{m=1}^{\infty}g^{m}[X] and gnm=1gm[X]g^{n}\upharpoonright\bigcap_{m=1}^{\infty}g^{m}[X] is not the identity map.

4.2. Dendrites with branching points of finite order

We now show that, if we drop the requirement that the dendrite XX has finitely many branching points, then none of the equivalences of equicontinuity from Theorem 1.8 holds. The first few equivalences can be seen to fail by looking at [6, Example 5.4], which is the Gehman dendrite XX (this dendrite is described and pictured in [13, Example 10.39]) with all branching points of finite order (with infinitely many branching points), and a surjective equicontinuous map f:XXf\colon X\longrightarrow X such that Per(f)X\operatorname{Per}(f)\neq X (consequently, X=m=1fm[X]X=\bigcap_{m=1}^{\infty}f^{m}[X] and Per(f)m=1fm[X]\operatorname{Per}(f)\neq\bigcap_{m=1}^{\infty}f^{m}[X]). Therefore ff is equicontinuous but fails to satisfy conditions (b), (c) and (d) of Theorem 1.8.

Now for the remaining equivalences, the following example finishes our analysis.

Example 4.3.

A dendrite XX with infinitely many branching points (each of which has finite order) and a map f:XXf\colon X\longrightarrow X that fails to be equicontinuous, but satisfies (e) through (i) of Theorem 1.8.

We build XX as a subset of 2\mathbb{R}^{2} as follows. We let K=[0,1]×{0}K=[0,1]\times\{0\} and, for each n{0}n\in\mathbb{N}\cup\{0\}, we let In={12n}×[0,12n]I_{n}=\left\{\frac{1}{2^{n}}\right\}\times\left[0,\frac{1}{2^{n}}\right] and Jn={12n}×[12n,0]J_{n}=\left\{\frac{1}{2^{n}}\right\}\times\left[-\frac{1}{2^{n}},0\right]. Define

X=K(n=0In)(n=0Jn).X=K\cup\left(\bigcup_{n=0}^{\infty}I_{n}\right)\cup\left(\bigcup_{n=0}^{\infty}J_{n}\right).

For notational convenience, we write K=n=1KnK=\bigcup_{n=1}^{\infty}K_{n} where Kn=[12n,12n1]×{0}K_{n}=\left[\frac{1}{2^{n}},\frac{1}{2^{n-1}}\right]\times\{0\} for each nn\in\mathbb{N}. Now we define the map f:XXf\colon X\longrightarrow X as follows. First make fKf\upharpoonright K the identity map. Now, for each nn\in\mathbb{N}, fInf\upharpoonright I_{n} is given as follows: for (12n,y)In\left(\frac{1}{2^{n}},y\right)\in I_{n} (0y12n0\leq y\leq\frac{1}{2^{n}}), we let

f(12n,y)={(12n+2y,0),if 0y12n+1;(12n1,4(y12n+1)),if 12n+1y12n,f\left(\frac{1}{2^{n}},y\right)=\left\{\begin{array}[]{ll}\left(\frac{1}{2^{n}}+2y,0\right),&\mbox{if }0\leq y\leq\frac{1}{2^{n+1}};\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \left(\frac{1}{2^{n-1}},4\left(y-\frac{1}{2^{n+1}}\right)\right),&\mbox{if }\frac{1}{2^{n+1}}\leq y\leq\frac{1}{2^{n}},\end{array}\right.

so that ff maps InI_{n} homeomorphically onto KnIn1K_{n}\cup I_{n-1}. Furthermore, fI0f\upharpoonright I_{0} is defined by letting f(0,y)=(0,y)f(0,y)=(0,-y) so that ff maps I0I_{0} homeomorphically onto J0J_{0}. Finally, for each n{0}n\in\mathbb{N}\cup\{0\}, we define fJnf\upharpoonright J_{n} by letting

f(12n,y)={(12n+y,0),if 12n+1y0;(12n+1,y+12n+1),if 12ny12n+1,f\left(\frac{1}{2^{n}},y\right)=\left\{\begin{array}[]{ll}\left(\frac{1}{2^{n}}+y,0\right),&\mbox{if }-\frac{1}{2^{n+1}}\leq y\leq 0;\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \left(\frac{1}{2^{n+1}},y+\frac{1}{2^{n+1}}\right),&\mbox{if }-\frac{1}{2^{n}}\leq y\leq-\frac{1}{2^{n+1}},\end{array}\right.

whenever (12n,y)Jn\left(\frac{1}{2^{n}},y\right)\in J_{n} (12ny0-\frac{1}{2^{n}}\leq y\leq 0); so that ff maps JnJ_{n} homeomorphically onto Kn+1Jn+1K_{n+1}\cup J_{n+1}. The dendrite XX, as well as the map f:XXf:X\longrightarrow X, are depicted in Figure 2.

Refer to caption
Figure 2. The dendrite XX that has infinitely many branching points of finite order. The map f:XXf\colon X\longrightarrow X is not equicontinuous.

We will denote by vv the point (0,0)(0,0). Notice that the sequence of endpoints of the InI_{n}, (12n,12n)n\left(\frac{1}{2^{n}},\frac{1}{2^{n}}\right)_{n\in\mathbb{N}} (which converges to vv), along with the increasing sequence of indices (n)n(n)_{n\in\mathbb{N}} and ε=1\varepsilon=1, witness the failure of equicontinuity of ff at vv (since fn(12n,12n)=(1,1)I0f^{n}\left(\frac{1}{2^{n}},\frac{1}{2^{n}}\right)=(1,1)\in I_{0}, which is at distance >1>1 from fn(v)=vf^{n}(v)=v). So ff is not equicontinuous.

For every nn\in\mathbb{N} we have

(4.2) Fix(fn)=Per(f)=K.\operatorname{Fix}(f^{n})=\operatorname{Per}(f)=K.

Thus the sets Fix(fn)\operatorname{Fix}(f^{n}), as well as Per(f)\operatorname{Per}(f), are all connected. Using (4.2) it is straightforward to see that property (j)(j^{\prime}) of Lemma 2.9 is not satisfied. Hence, by the same lemma, XX has no ff-expanding arcs

It remains to show that the function fuf^{u} is continuous, whenever uu is a nonprincipal ultrafilter. To do this, we define an auxiliary (continuous) function g:XXg\colon X\longrightarrow X as follows. First of all, gKg\upharpoonright K will be the identity map. For every n{0}n\in\mathbb{N}\cup\{0\}, we have

g(12n,12n)=v=g(12n,12n).g\left(\frac{1}{2^{n}},\frac{1}{2^{n}}\right)=v=g\left(\frac{1}{2^{n}},-\frac{1}{2^{n}}\right).

Next, if (12n,y)In\left(\frac{1}{2^{n}},y\right)\in I_{n} is not an endpoint (that is, if 0<y<12n0<y<\frac{1}{2^{n}}) then we let m{0}m\in\mathbb{N}\cup\{0\} be unique such that

12n12n+my<12n12n+m+1,\frac{1}{2^{n}}-\frac{1}{2^{n+m}}\leq y<\frac{1}{2^{n}}-\frac{1}{2^{n+m+1}},

and define

g(12n,y)=(12nm+22m+1(y12n12n+m),0)if m<n,g\left(\frac{1}{2^{n}},y\right)=\left(\frac{1}{2^{n-m}}+2^{2m+1}\left(y-\frac{1}{2^{n}}-\frac{1}{2^{n+m}}\right),0\right)\hskip 14.22636pt\mbox{if }m<n,

and

g(12n,y)=(122n(y(12n122n)),0)if nm.g\left(\frac{1}{2^{n}},y\right)=\left(1-2^{2n}\left(y-\left(\frac{1}{2^{n}}-\frac{1}{2^{2n}}\right)\right),0\right)\hskip 14.22636pt\mbox{if }n\leq m.

Finally, if (12n,y)Jn\left(\frac{1}{2^{n}},y\right)\in J_{n} is not an endpoint (i.e., 12n<y<0-\frac{1}{2^{n}}<y<0), we let

g(12n,y)=(12n+y,0).g\left(\frac{1}{2^{n}},y\right)=\left(\frac{1}{2^{n}}+y,0\right).

The function gg is continuous and, moreover, for each xXx\in X we have limnfn(x)=g(x)\lim_{n\to\infty}f^{n}(x)=g(x) and therefore, for every nonprincipal ultrafilter uu, it must be the case that fu=gf^{u}=g. Thus the function fuf^{u} is continuous for every nonprincipal ultrafilter uu.

4.3. Finite graphs

The case of finite graphs might be harder to analyze than the case of finite trees. The first difficulty that arises is the fact that the unit circle 𝕊1\mathbb{S}^{1} is a finite graph, and there are maps f:𝕊1𝕊1f\colon\mathbb{S}^{1}\longrightarrow\mathbb{S}^{1} (such as, e.g., rotations by an irrational angle), which, though equicontinuous and surjective, lack any periodic points. Thus, items (a), (b), (c) and (d) from Theorem 1.8 are no longer equivalent if one attempts to replace “finite tree” with “finite graph” in its statement. For finite graphs with at least one branching point or at least one endpoint, however, the equivalence between items (a), (b) and (c) can be established by adapting the argument in the proof of Proposition 2.8. The following example shows that the equivalence between statements (a) and (f) from Theorem 1.8 does not hold on finite graphs, even if one demands that the graphs have branching points or endpoints.

Example 4.4.

Two finite graphs X1,X2X_{1},X_{2}, and corresponding equicontinuous maps fi:XiXif_{i}\colon X_{i}\longrightarrow X_{i} so that Fix(fi)\operatorname{Fix}(f_{i}) is disconnected, for i{1,2}i\in\{1,2\}. Here X1X_{1} has exactly one branching point and only one endpoint, X2X_{2} is a simple closed curve in X1X_{1} (and thus it has no branching points nor endpoints), and f2=f1X2f_{2}=f_{1}\upharpoonright X_{2}.

The finite graph X1X_{1} is given by

X1={(x,y)2|x[1,2] and y=0, or x[1,1] and y=±x},X_{1}=\left\{(x,y)\in\mathbb{R}^{2}\big{|}x\in[1,2]\text{ and }y=0,\text{ or }x\in[-1,1]\text{ and }y=\pm x\right\},

and we let f1(x,y)=(x,y)f_{1}(x,y)=(x,-y). Since f12f_{1}^{2} is the identity map, f1f_{1} is equicontinuous. However, Fix(f1)\operatorname{Fix}(f_{1}) is obviously disconnected. Now

X2={(x,y)2|x[1,1] and y=±x}X_{2}=\left\{(x,y)\in\mathbb{R}^{2}\big{|}x\in[-1,1]\text{ and }y=\pm x\right\}

and f2=f1X2:X2X2f_{2}=f_{1}\upharpoonright X_{2}\colon X_{2}\longrightarrow X_{2} satisfy the requirements.

The observations, along with the example from this subsection suggest that the following might be a worthwhile question (a subset of the following question appears as [18, Question 3.10]).

Problem 4.5.

Let (X,f)(X,f) be a discrete dynamical system. Which of the equivalences from Theorem 1.8 hold if we assume that XX is an arbitrary finite graph? Which of them hold if we furthermore assume that XX has at least one branching point or at least one endpoint?

A subset the next question appears as [18, Question 3.9]. First recall that the cone over the harmonic sequence {0,1,12,13,}\{0,1,\frac{1}{2},\frac{1}{3},\ldots\} is called the harmonic fan. Attempting to generalize some clauses of Theorem 1.8 from finite trees to non-locally connected continua, we ask the following question.

Problem 4.6.

Let XX be the harmonic fan and let f:XXf\colon X\longrightarrow X be a map. Is it true that the existence of an ff-expanding arc in XX is equivalent to the fact that fuf^{u} is discontinuous for some uu\in\mathbb{N}^{*}? Does the existence of an ff-expanding arc in XX implies that there is mm\in\mathbb{N} such that XX contains a gmg^{m}-divergent sequence?

We finish with two more questions related to Remark 4.1.

Problem 4.7.

Is there a dendrite XX with finitely many branching points and a map f:XXf:X\longrightarrow X such that Per(f)=n=1fn[X]\operatorname{Per}(f)=\bigcap_{n=1}^{\infty}f^{n}[X] yet ff fails to be equicontinuous? An equivalent way of phrasing the same question is: does the implication (d)(a)(d)\Rightarrow(a) of Theorem 1.8 hold for any dendrite with finitely many branching points?

A map ff as requested in Question 4.7 would still need to be equicontinuous when restricted to n=1fn[X]\bigcap_{n=1}^{\infty}f^{n}[X] (even though ff itself fails to be equicontinuous). Notice that, for XX a finite tree and f:XXf:X\longrightarrow X a map, equicontinuity of ff is equivalent to equicontinuity of the restriction fn=1fn[X]f\upharpoonright\bigcap_{n=1}^{\infty}f^{n}[X] by [11, Theorem 5.2]; on the other hand, [15, Example 2.9] shows that this equivalence no longer holds if XX is an arbitrary dendrite instead. This motivates the following question.

Problem 4.8.

Is there a class of dendrites, broader than the class of finite trees, such that if XX belongs to the class and f:XXf:X\longrightarrow X is a map, then the equicontinuity of fn=1fn[X]f\upharpoonright\bigcap_{n=1}^{\infty}f^{n}[X] implies the equicontinuity of ff?

Acknowledgements

The second author was supported by a postdoctoral fellowship from DGAPA-UNAM under the mentoring of the first author. Both authors are grateful to two anonymous referees and one anonymous editor for numerous suggestions that helped to improve the paper.

References

  • [1]
  • [2] E. Akin, J. Auslander and K. Berg, When is a transitive map chaotic?, in: Convergence in Ergodic Theory and Probability, V. Bergelson, P. March, J. Rosenblatt (eds.), Walter de Gruyter & Co., Berlin, 1996, pp. 25–40.
  • [3] A. Blass, Ultrafilters: where topological dynamics = algebra = combinatorics, Top. Proc. 18 (1993), 33–56.
  • [4] A. M. Bruckner and J. Ceder, Chaos in terms of the map xω(x,f)x\rightarrow\omega(x,f), Pacific J. Math. 156 (1992), 63–96.
  • [5] A. M. Bruckner and T. Hu, Equicontinuity of iterates of an interval map, Tamkang Journal of Mathematics 21 (1990), 287–294.
  • [6] J. Camargo, M. Rincón and C. Uzcátegui, Equicontinuity of maps on dendrites, Chaos, Solitons and Fractals 126 (2019), 1–6.
  • [7] R. Engelking, General Topology, Translated from the Polish by the Author, Sigma Series in Pure Mathematics, vol. 6, 2nd edn. Heldermann Verlag Berlin, 1989.
  • [8] S. García–Ferreira and M. Sanchis, Some remarks on the topology of the Ellis semigroup of a discrete dynamical system, Topology Proceedings 42 (2013), 121–140.
  • [9] N. Hindman, Finite sums from sequences within cells of a partition of NN, J. Combin. Theory Ser. A 17 (1974) 1–11.
  • [10] N. Hindman and D. Strauss, Algebra in the Stone-Čech compactification. Second revised and extended edition, De Gruyter Textbook. Walter de Gruyter & Co., Berlin, 2012.
  • [11] J.-H. Mai, The structure of equicontinuous maps, Trans. Amer. Math. Soc. 355 (2003), 4125–4136.
  • [12] J.-H. Mai, Pointwise-recurrent graph maps, Ergod. Th. & Dynam. Sys 25 (2005), 629–637.
  • [13] S. B. Nadler, Jr., Continuum Theory. An Introduction, Marcel Dekker, New York, Basel, Hong Kong, 1992.
  • [14] T. Sun, Y. Zhang and X. Zhang, Equicontinuity of a graph map, Bull. Austral. Math. Soc. 71 (2005), 61–67.
  • [15] T. Sun, Z. Chen, X. Liu and H. Xi, Equicontinuity of dendrite maps, Chaos, Solitons and Fractals 69 (2014), 10–13.
  • [16] T. Sun, G. W. Su, H. J. Xi and X. Kong, Equicontinuity of maps on a dendrite with finite branch points, Acta Math. Sin. (Engl. Ser.) 33 (2017), 1125–1130.
  • [17] P. Szuca, F-limit points in dynamical systems defined on the interval, Cent. Eur. J. Math. 11 (2013), 170–176.
  • [18] I. Vidal-Escobar and S. García-Ferreira, About the Ellis semigroup of a simple kk-od, Top. Appl. 265 (2019), 106756.
  • [19] S. Willard, General Topology, Dover, Mineola, New York, 2004 (unabridged from the original 1970 version).