Equicontinuous mappings on finite trees
Abstract.
If is a finite tree and 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 is equicontinuous. To name just a few of the results obtained: the equicontinuity of is equivalent to the fact that there is no arc satisfying for some . It is also equivalent to the fact that for some nonprincial ultrafilter , the function is continuous (in other words, failure of equicontinuity of is equivalent to the failure of continuity of every element of the Ellis remainder ). 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 this paper deals with maps 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)
A discrete dynamical system is a pair such that is a metric space and is a map, i.e. a continuous function.
-
(2)
If is a discrete dynamical system, we define as the identity map on , and, for each , .
-
(3)
If are metric spaces and is a family of functions from to , we say that is equicontinuous at if for every there exists a such that implies for all and every and if is equicontinuous at every , we say that is equicontinuous.
-
(4)
If is a metric space, the function is equicontinuous at if its family of iterates, , is equicontinuous at ; and if is equicontinuous at every 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 , equicontinuity of a family of functions is a pointwise closed condition; consequently if is equicontinuous at then so is , where is the closure of in with the product topology. Note also that, if is compact, then by the usual argument, equicontinuity implies uniform equicontinuity (i.e., given , a can be chosen to work for all ).
Definition 1.2.
Let be a discrete dynamical system, where is compact.
-
(1)
The Ellis semigroup (also called the enveloping semigroup) of is defined as , the closure in (with the product topology) of the family . Note that, as is compact (by Tychonoff’s theorem), so is .
-
(2)
The Ellis remainder of the discrete dynamical system is
Note that .
Composition of functions is what makes a semigroup. In fact, is a compact right-topological semigroup (a semigroup equipped with a topology making all right translations continuous). Since is a metric space, by the observation immediately after Definition 1.1, equicontinuity of is equivalent to equicontinuity of the family , 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 can be made more concrete by means of ultrafilters: for every ultrafilter on , define the ultrafilter-limit function by letting . Then by [8, Theorem 2.2] we have
and consequently
Full definitions of ultrafilters, both principal and nonprincipal, as well as of -limits will be provided in Section 3.
Definition 1.3.
Let be a discrete dynamical system.
-
(1)
is a fixed point if ; the set of fixed points of is denoted by
-
(2)
is a periodic point if , for some , in which case the least such is called the period of ; the set of all periodic points of is denoted by .
-
(3)
is periodic if there exists such that is the identity map on , and is pointwise-periodic if
It is immediate from Definition 1.3 that .
We primarily deal with continua (compact, connected and metric spaces). A simple closed curve is a continuum homeomorphic to the unit circle and an arc is a continuum homeomorphic to the unit interval Other examples of continua are finite graphs (compact, connected one-dimensional polyhedra), dendrites, finite trees and -ods, for each with We give the proper definitions of the last three in Section 2. For the moment it is convenient to note that a -od is an arc, -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 is an arc and is a map, then the following are equivalent:
-
(1)
is equicontinuous;
-
(2)
the restriction is the identity map;
-
(3)
-
(4)
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 and a map , equicontinuity of is equivalent to each of the following conditions: that the restriction is periodic, that for some , and that is connected for all ; furthermore, any of these is also equivalent to the set being connected.
Further interesting results regarding equicontinuity of a map have been obtained by Mai ([11]) in the case where is a finite graph, and by Camargo, Rincón and Uzcátegui ([6]) in the case where is a dendrite. The former shows in [11, Theorem 5.2] that, if is a finite graph, then is equicontinuous if and only if (here denotes the set of recurrent points of , to be defined later in Definition 2.3; for the moment just note that ); the latter proves in [6, Theorem 4.12] that, if is a dendrite, then is equicontinuous if and only if plus an extra condition having to do with the -limit sets of . 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 is a finite tree, is equicontinuous if and only if .
Another concept that will play a central role in this paper is that of an expanding arc. To motivate this concept consider a nonnegative and the map defined by
(1.1) |
It is readily checked that is equicontinuous if and only if , whereas if then fails to be equicontinuous at every . Intuitively speaking, maps that expand the real line fail to be equicontinuous. Note that for the map defined in (1.1), with , we have for all , where is the unit interval. This leads to the following definition.
Definition 1.5.
Let be a discrete dynamical system, and let be an arc. We say that is an -expanding arc if there exists such that .
Hence the map defined in (1.1) fails to be equicontinuous if and only if is -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 they assume is surjective, but the general case follows from the proof of [4, Theorem 1.2]):
Proposition 1.6.
Let be a -od for some and be a map where is not a singleton. Then is not equicontinuous if and only if contains an -expanding arc.
In this paper, we generalize Proposition 1.6 from -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 , where is a -od with , if is continuous for every nonprincipal ultrafilter then is equicontinous (note that the converse implication is trivially true as a consequence of the observation right after Definition 1.2). Hence if is not equicontinuous then for some nonprincipal ultrafilter is not continuous. In [18] the authors consider the possibility that, for some nonprincipal ultrafilter distinct from 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 is a finite tree rather than just a -od. As a consequence of this, if fails to be equicontinuous with a finite tree, then every element fails to be continuous. Thus, for maps on a finite tree , we have a strong dichotomy by means of which either every element is continuous, or every element is discontinuous, according to whether or not 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 be a discrete dynamical system, where is a finite tree. Then, either every element of is continuous, or every element of is discontinuous.
We now state the Main Theorem of this paper.
Main Theorem 1.8.
Let be a finite tree and be a map. Then, the following are equivalent:
-
(a)
is equicontinuous;
-
(b)
there is an such that the restriction of to is the identity map;
-
(c)
there exists an such that
-
(d)
-
(e)
there is no -expanding arc in
-
(f)
for every , the set is connected;
-
(g)
the set is connected;
-
(h)
for every nonprincipal ultrafilter , the function is continuous (i.e., every element of is continuous);
-
(i)
for some nonprincipal ultrafilter , the function is continuous (i.e., some element of is continuous).
Remark 1.9.
Some remarks about the equivalences from Theorem 1.8:
-
(1)
The equivalence between (e), (f) and (g) will be established not only for finite trees, but for arbitrary dendrites.
- (2)
-
(3)
The equivalences , where , fail if we allow to be an arbitrary dendrite, and even if is merely a dendrite all of whose branching points are of finite order (see §4.2), or (with the possible exception of , see Remark 4.1) if is merely a dendrite with finitely many branching points (see §4.1).
- (4)
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 of a space we denote by either or the closure of in The interior of in is denoted by We begin by stating some standard results that will be used.
Proposition 2.1.
Let be a decreasing sequence of closed subsets of a compact space , and let . Then,
-
(1)
if is an open set containing , then for all sufficiently large ;
-
(2)
if each is nonempty, then so is ;
-
(3)
if every is connected, then so is ;
-
(4)
if is a map, then .
Proof.
Parts (1) and (3) follow from [7, Corollary 3.1.5 and Corollary 6.1.19]. Part (2) follows from (1) with To show part (4), it is enough to verify that . Let Since is a decreasing sequence of closed, nonempty subsets of ; by (2), and then ∎
If is a discrete dynamical system with compact, then by Proposition 2.1, is a nonempty compact subspace of satisfying
This means that the restricted map is onto . In the case where is a connected space, so is , 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 fails to be equicontinuous at the point if and only if there exists an , a sequence of points converging to , and an increasing sequence of indices such that for all . In this case we will say that , , and witness the failure of equicontinuity of at .
Lemma 2.2.
Let be a metric space, let be a map, and suppose that fails to be equicontinuous at . Then, for every ,
-
(1)
fails to be equicontinuous at , and
-
(2)
there exists an such that fails to be equicontinuous at .
Proof.
Suppose that , the sequence of points , and the sequence of indices witness the failure of equicontinuity of at , and let . To prove (1), assume without loss of generality that ; then, the sequence (which converges to by continuity of the function ), and the increasing sequence of natural numbers, along with , witness the failure of equicontinuity of at . This shows (1). For (2), apply the pigeonhole principle to assume, without loss of generality, that there is a fixed such that for all . Let be such that ; then, the sequence , which converges to , along with the increasing sequence of natural numbers and , witness the failure of equicontinuity of at . ∎
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 be a metric space, and let be a map.
-
(1)
The -limit set of at is the set of all points for which there is an increasing sequence with ; this set is denoted by .
-
(2)
A point is a recurrent point if ; the set of recurrent points of is denoted by .
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 is a finite tree and is an equicontinuous surjective map, then is a homeomorphism which furthermore is periodic.
Proof.
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 is a dendrite, and then there is a unique (closed) arc in joining and ; such an arc will always be denoted by . Since continuous images of connected sets must be connected, for any map and every we have that .
Whenever is a dendrite and is a subcontinuum of , then there exists a retraction called the first point function, such that for and , is the first point in the arc (equipping such an arc with a linear order where ) that belongs to The mapping does not depend on the specific (see [13, Lemmas 10.24, 10.25 and Terminology 10.26]).
Finally, every dendrite is uniformly locally arcwise connected, that is, for every there exists a such that, whenever and , the arc must have diameter (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 there exists a such that whenever and then there is an arc joining and with diameter ; see [19, Theorem 31.4]).
Proposition 2.6.
Let be a dendrite, let be a map, and let . If is a connected component of such that , then .
Proof.
Notice that is a subcontinuum of –hence is itself a dendrite. We consider the first point function and note that, for , it must be the case that . Since is a dendrite, it has the fixed point property; therefore the map has a fixed point . It is now easy to check that we must have . ∎
The following is another definition that will be crucial throughout the paper.
Definition 2.7.
Let be a dendrite and
-
(1)
The order of a point is the number of connected components of ;
-
(2)
a point is an endpoint if its order is , and a branching point if its order is ;
-
(3)
is a free arc in if no element of is a branching point;
-
(4)
for , is a -od if it contains exactly one branching point (called the vertex of ), which has order ; a -od is simply defined to be an arc (we do not specify a vertex in this case);
-
(5)
is a finite tree if it has only finitely many branching points and each of these branching points has a finite order.
Note that is a free arc in if and only if is open in In a general topological space , the order of a point is defined as the least cardinal number such that, for every open neighbourhood of there exists another open neighbourhood with and (where denotes the boundary of in ), cf. [13, Definition 9.3]; this will be important towards the end of Section 4. If, however, the topological space 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 be a finite tree and let be a map. Then the following conditions are equivalent:
-
(a)
is equicontinuous;
-
(b)
for some , the restriction is the identity map;
-
(c)
for some , ;
-
(d)
.
Proof.
We consider first the case where is surjective. Note that in such situation, and for each . Moreover (b) asserts that is periodic, (c) that for some , and (d) that is pointwise-periodic. By Proposition 2.4, . The implications are obvious and, by [6, Theorem 4.14], the implication holds not only on finite trees, but on every dendrite and with being any (surjective) map.
We now consider the case of an arbitrary (not necessarily surjective) map Since every finite tree is, in particular, a finite graph, we may use [11, Theorem 5.2] to see that is equicontinuous if and only if so is , and since the latter map is onto (and since and ), 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 be a dendrite and let be a map. Then the following are equivalent:
-
()
contains an -expanding arc;
-
()
for some , the set is disconnected;
-
()
there exist points and such that , , and .
Proof.
Let be an -expanding arc and fix an such that . We consider three cases according to whether both, exactly one, or none of are fixed points for .
- Case 1:
-
If and , use the fact that to get a such that . In particular, with and , showing that is not arcwise connected, so is disconnected.
- Case 2:
-
If but , use the fact that to get a such that . Let be the connected component of containing If then, since is arcwise connected, we have a contradiction. Hence, . Letting be the connected component of containing , Proposition 2.6 guarantees the existence of a . Then we have , , and , showing that is disconnected.
- Case 3:
-
If and . Then (since ) we can find with and . Note that . Equip with a linear order via a homeomorphism mapping to and to . If , then a couple of applications of Proposition 2.6 yield fixed points such that ; since , this shows that is disconnected. If, on the other hand, we have , then notice that , so there is a with ; also, and so there is an with . This way we have obtained with and , thus, a couple of applications of Proposition 2.6 yield two fixed points with ; the fact that implies then that is disconnected, and we are done.
Let and be such that . Then there is a with . Considering the first point function and the point , we must have that either or . Since the situation is entirely symmetric, assume without loss of generality that and let . Then we have and .
Under the assumptions we have and so is an -expanding arc. ∎
We now proceed to prove that the failure of equicontinuity of a map on a finite tree implies the existence of an -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 be a finite tree, let be a map and let If has no -expanding arcs, then there is an -invariant subspace (that is, ), where is a -od for some , such that .
Proof.
Let be the order of Since is a finite tree, for some the closed ball centred at is a union of finitely many arcs, say , which pairwise intersect at only. By continuity of , we can pick a with such that, if , then for every . Let be the closed ball of radius centred at , and define
Due to our choice of and , we have . Since is a fixed point for , whenever we must have . Hence is a -od for some ( and is the vertex of if , if is an arc, i.e., moreover is an interior point of the arc if , and it is an endpoint of the arc if ). It remains to show that is an -invariant subspace, so let and let us argue that . We can write for some (this includes the case , interpreted as ); there is nothing to do if , so assume that for . Looking at the finite sequence of points , the pigeonhole principle guarantees the existence of and such that . Let us linearly order the arc by copying the order of along a homeomorphism mapping to . If , this would mean that
is an -expanding arc, a contradiction. Therefore we must have , implying that
and so there exists a point with . Hence and so , and we are done. ∎
Remark 2.11.
Suppose that is a metric space, is a map, and is an -invariant subspace with . If is not equicontinuous at , then the restriction also fails to be equicontinuous at : given , if witnesses the equicontinuity of at , then witnesses the equicontinuity of at .
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 of elements of , its set of finite sums is defined as
the set of all numbers that can be obtained by adding a finite amount of terms of the sequence without repetitions. The result known as Hindman’s theorem ([9, Theorem 3.1]) states that for any finite partition of there exists an infinite (strictly increasing) sequence such that the set 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 and , we say that the sequence is a sum subsystem of the sequence if there are finite subsets of such that, for each , we have and . Note that this implies in particular that . 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 and for every finite partition of the set , there exists an and a sum subsystem of the sequence such that .
Since , the original version of Hindman’s theorem follows immediately from Theorem 2.12 above.
Theorem 2.13.
Let be a finite tree and let be a map. If is not equicontinuous, then has an -expanding arc.
Proof.
Take an such that is not equicontinuous at . We have two cases.
Case 1: The point is eventually periodic (i.e., there is such that is a periodic point; equivalently, the set is finite). This means that, replacing by some if necessary (and using clause (1) of Lemma 2.2), we may assume that is a periodic point, say with period . Now use clause (2) of Lemma 2.2 to find an such that fails to be equicontinuous at , and notice that is a fixed point for . If has an -expanding arc, then this is also an -expanding arc and we are done. If, on the contrary, there are no -expanding arcs, then we can use Lemma 2.10 to obtain an -invariant subspace such that is a -od for some , and . By Remark 2.11, the restricted map fails to be equicontinuous at , and so by Proposition 1.6, must contain an -expanding arc . Then is also an -expanding arc, and we are done.
Case 2: The point is not eventually periodic. Then the set is infinite. The space is a finite tree and so it can be written as a union of finitely many maximal free arcs in such that any two distinct have at most one (branching) point in common. Now we define subsets of as follows: iff is a branching point of and, for each , iff and is not a branching point of Clearly
Given the set can be empty. Since the set is infinite, there exist distinct such that for every and for each . Hence is a finite partition of . Theorem 2.12 provides us with a and an infinite strictly increasing sequence of natural numbers, such that the set . Since the points as varies are pairwise distinct and only has finitely many branching points, is finite. Therefore and so
Use a homeomorphism of onto to equip with a linear order . We now partition the set according to whether or ; a further application of Theorem 2.12 allows us to obtain a sum subsystem of such that is contained in one piece of this partition. This means that there is an such that,
Continuing this process by induction, we obtain, for each , a sum subsystem of and an such that
Now, an application of the pigeonhole principle allows us to obtain an infinite increasing sequence such that all the are equal, say, without loss of generality, to . What this means is that, if we define the sequence by , then for every and each we have .
Now, for each we define a point by
Since for every , we must have for every . We may now break the proof into two further subcases (recall that is the metric on ).
- Subcase 2.A:
-
There is a such that . Let and note that, for every , there is an with
We have ; in particular, we also have
It follows that
By continuity of the function , we may conclude that . Thus, all the points
are fixed points of the map . If contains an -expanding arc, then this arc is also -expanding and we are done, so assume otherwise. Then we may apply Lemma 2.10 to each of the points to get -invariant subcontinua
such that, for every we have and each is a -od for some . Let be such that for every , the ball centered at with radius is contained in By the continuity of the functions we get a such that, if , then Hence, for each , if then . Note that contains points of the form arbitrarily close, and all the points of the form are points where the map is not equicontinuous (by clause (1) of Lemma 2.2). Hence we can find a with such that is not equicontinuous at ; now use clause (2) of Lemma 2.2 to get such that is not equicontinuous at . Since is -invariant, we may conclude that is not an equicontinuous map (see Remark 2.11). Since is a -od, by Proposition 1.6, the subcontinuum of must have an -expanding arc, and we are done.
- Subcase 2.B:
-
for every Then let . For each fix an such that . If for some , it is not the case that , then we must have
and therefore there is an -expanding arc and we are done, so assume that for all we have . The points for are arbitrarily close to and they all satisfy , so by continuity of we have .
We define connected subspaces as follows. is the connected component of that does not contain , and is the connected component of containing . Since
and all such points belong to the maximal free arc in , we can write
and the union is disjoint. Since , by Proposition 2.6 there is a ; now if we let be sufficiently large that then we will have and so again by Proposition 2.6 there exists a . Letting , we get that , and . Hence is a disconnected set, and so by Lemma 2.9, must have an -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)
An ultrafilter on is a family of subsets of such that
-
(a)
is nonempty and ;
-
(b)
if then ;
-
(c)
if and then ;
-
(d)
whenever , then either or .
-
(a)
-
(2)
An ultrafilter on is principal if there exists an such that ; otherwise we say that is nonprincipal.
-
(3)
We use the symbol to denote the set of all ultrafilters on , and we denote with the set of all nonprincipal ultrafilters on .
-
(4)
Given a metric space , a sequence of points on , and an ultrafilter , we say that is the ultrafilter-limit of in symbols if for every the set .
-
(5)
Given a metric space , a function , and an ultrafilter , we define the ultrafilter-limit function (also called the -th iterate of ) by .
Given a compact metric space a map and , it can be shown that
A few comments about the above definitions are in order. For each , it is common to identify the natural number with the principal ultrafilter ; this way we can think of as a subset of , and we have . Furthermore, one can topologize by declaring the sets to be open, for each ; this endows with a compact Hausdorff topology containing as a discrete dense subspace ([10, Lemma 3.17 and Theorems 3.18 and 3.28]). Regarding the concept of a -limit, it is worth pointing out that, in a compact metric space , every sequence of points will have a unique -limit (for every )([10, Theorem 3.48]). Moreover, if is the principal ultrafilter , then ; similarly (and as a consequence of the above), for a function we will have that . Thus, no confusion should arise if we sometimes abuse notation and write instead of .
Furthermore, it is possible to equip with a right-topological semigroup operation, denoted by . That is, is an associative binary operation on such that, for each fixed , the function is continuous. The operation is given by the formula
This operation extends the usual sum on , in the sense that, if and are the corresponding principal ultrafilters, then , although is not commutative on all of . It is possible to verify that, for any , we have (see [3, p. 38]).
As we mentioned in the Introduction, the equation
shown in [8, Theorem 2.2] and which holds for every map on a compact metric space , is the main reason why obtaining information about the ultrafilter-limit functions has a great deal of importance within the study of the discrete dynamical system . 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 be a metric space.
-
(1)
Let be an arc, and let be a sequence of elements of . We say that the sequence is -monotone if it is monotone (i.e., either increasing or decreasing) when viewed as a sequence on the unit interval via a homeomorphism . Equivalently, the sequence is monotone if for each (noting that ).
-
(2)
If is a map, a sequence of elements of some arc is said to be -backwards if it is -monotone and for each we have .
Remark 3.3.
Note that, by compactness of an arc and monotonicity of backward sequences, any -backward sequence on a dendrite is always convergent. Furthermore, the limit of the sequence is a fixed point of .
Lemma 3.4.
Let be a dendrite and let be a map, and suppose that there is an -expanding arc . Then the following two conditions hold:
-
(1)
for some there exists an -backward sequence in ;
-
(2)
the set is disconnected.
Proof.
If there is an -expanding arc in then, by Lemma 2.9, there exist points and such that , , and , so we can find a such that . Now , so there exists a such that . Continuing by induction, if we already know with and , then
and so there exists a such that . This way we obtain a sequence which is -backward. So (1) holds.
To show (2), we use the points and the sequence obtained in (1). Since , by Proposition 2.6 there is a point such that ; then we have , so it suffices to show that . If we are done, so assume that , say with period . Then ; since the are pairwise distinct we can choose an such that . Then
thus for each we have and so . Since , the proof is finished. ∎
Corollary 3.5.
For a map with a dendrite, the following are equivalent:
-
(e)
there is no -expanding arc in ;
-
(g)
the set is connected.
Proof.
Now, in order to use -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 be a compact metric space, and let be a map. A sequence of elements of will be said to be -divergent if the following three conditions hold:
-
(1)
exists in ;
-
(2)
for each , (this implies that );
-
(3)
there exists an open neighbourhood of such that .
It is not hard to see that -divergent sequences can only exist if fails to be equicontinuous. As a matter of fact, much more is true, as seen in the following theorem.
Theorem 3.7.
Let be an arbitrary compact metric space and let be a map. If there is an such that contains a -divergent sequence, then for every nonprincipal ultrafilter , the function is discontinuous.
Proof.
Let be the hypothesized -divergent sequence, let , and let be an open set containing such that .
Now let be an arbitrary nonprincipal ultrafilter. There exists a unique such that , so that . This means that it makes sense to consider the Rudin–Keisler image of the ultrafilter under the mapping given by . So we have that (where denotes the Rudin–Keisler image of the ultrafilter under the mapping ).
Define a new sequence by letting , and let . Since converges to and is continuous, will converge to . We now proceed to observe that
and, for each , we have
By definition of ultrafilter-limits, must be an accumulation point of the set . However, for we have , so for every , and therefore the sequence does not converge to , showing that the function is discontinuous at , and we are done. ∎
The previous lemma works for every compact metric space. For certain dendrites, there is a relation between -backwards sequences and -divergent sequences.
Lemma 3.8.
Let be a dendrite with only finitely many branching points, and let be a map. If there is an arc such that contains a -backwards sequence, then there exists an such that has a -divergent sequence.
Proof.
Let be a -backwards sequence in the arc , and let . Notice that is a fixed point of , and therefore .
Now let us fix some notation. First of all, since has only finitely many branching points, we may shrink and drop finitely many terms of the sequence (and shift indices afterwards so that our sequence indexing still starts at 0) to ensure that is a free arc in . Now linearly order the arc via a homeomorphism with in such a way that . Then the monotonicity of the -backwards sequence means in this case that the sequence is increasing. Now let be the first point function from onto the subcontinuum of . We will analyze the -orbit of There are two cases to consider.
Case 1: For every , . In this case, for each fixed we have that , and so for every . Since is a dendrite and hence uniformly locally arcwise connected, there must be a such that, whenever and , then the arc must have diameter smaller than that of the arc . In particular, if , then . So if we let be the ball centred at with radius , then for every it is the case that , and consequently the sequence itself is already -divergent.
Case 2: There exists an such that . Fix one such , and notice that the function satisfies
Therefore (by a standard result for maps in the unit interval) this map must have a fixed point in , that is, there is a with . Since is a free arc in , we have that is one of the endpoints of whenever . Since (so is an interior point of ), from it follows that and so is actually a fixed point of the map .
Now , so there must exist a such that . We continue this process by induction: given a
we find a such that . This way we obtain a monotone sequence , with limit , which is -backwards and where . Since , any open set containing and not containing will satisfy , for every . Therefore the sequence is -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 -expanding arc in then there is an -backward sequence for some , by Lemma 3.4; this yields an such that there is an -divergent sequence by Lemma 3.8, and this in turn implies that there is no such that 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 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 and a map such that fails to be equicontinuous (although is equicontinuous) yet . 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 with a unique branching point, which has infinite order, and maps such that satisfies all conditions from (e) through (i) of Theorem 1.8 but fails to be equicontinuous, while is equicontinuous but fails to satisfy conditions (b) and (c) of Theorem 1.8.
For other purposes, the dendrite together with the map , 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 that are not made in [6], and in order to be able to also describe the map . We build by taking infinitely many disjoint arcs indexed by , , with each of length , and identifying in a single point (the vertex) one end of each . The result is a dendrite with a single infinite-order branching point , as in Figure 1.

Now, for each , let us consider a map defined by fixing , and, for each , if is the unique element of at distance from , then is the unique element of at distance (at distance in the case ) from . Note that maps homeomorphically onto .
We let be defined by , that is, by and whenever is the unique element in so that . The sequence , where is the endpoint of that is distinct from (this sequence converges to ), together with and the sequence of indices given by , witness the failure of the equicontinuity of at (since , where is the endpoint of that is distinct from ). For each , we have
(4.1) |
which is a connected set. Using (4.1) it is straightforward to see that property of Lemma 2.9 is not satisfied. Hence, by the same lemma, has no -expanding arcs. Note that
Moreover, for every we have , which implies that, for each nonprincipal ultrafilter , is the map with constant value , which is continuous.
We now proceed to describe the map . We stipulate that is the identity map. For each positive , we let ; finally, we let
Hence we have . Therefore, for every , the map will cyclically permute the finite sequence of arcs , in such a way that is the identity map (and will fix every point in each of the for with ). As a result of this, we will have that , and so will be equicontinuous by [6, Theorem 4.14]; at the same time, although is pointwise-periodic, contains points of arbitrarily high period (if for , , then the period of is equal to ) and therefore, for every , we have and 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 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 (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 such that (consequently, and ). Therefore 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 with infinitely many branching points (each of which has finite order) and a map that fails to be equicontinuous, but satisfies (e) through (i) of Theorem 1.8.
We build as a subset of as follows. We let and, for each , we let and . Define
For notational convenience, we write where for each . Now we define the map as follows. First make the identity map. Now, for each , is given as follows: for (), we let
so that maps homeomorphically onto . Furthermore, is defined by letting so that maps homeomorphically onto . Finally, for each , we define by letting
whenever (); so that maps homeomorphically onto . The dendrite , as well as the map , are depicted in Figure 2.

We will denote by the point . Notice that the sequence of endpoints of the , (which converges to ), along with the increasing sequence of indices and , witness the failure of equicontinuity of at (since , which is at distance from ). So is not equicontinuous.
For every we have
(4.2) |
Thus the sets , as well as , are all connected. Using (4.2) it is straightforward to see that property of Lemma 2.9 is not satisfied. Hence, by the same lemma, has no -expanding arcs
It remains to show that the function is continuous, whenever is a nonprincipal ultrafilter. To do this, we define an auxiliary (continuous) function as follows. First of all, will be the identity map. For every , we have
Next, if is not an endpoint (that is, if ) then we let be unique such that
and define
and
Finally, if is not an endpoint (i.e., ), we let
The function is continuous and, moreover, for each we have and therefore, for every nonprincipal ultrafilter , it must be the case that . Thus the function is continuous for every nonprincipal ultrafilter .
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 is a finite graph, and there are maps (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 , and corresponding equicontinuous maps so that is disconnected, for . Here has exactly one branching point and only one endpoint, is a simple closed curve in (and thus it has no branching points nor endpoints), and .
The finite graph is given by
and we let . Since is the identity map, is equicontinuous. However, is obviously disconnected. Now
and 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 be a discrete dynamical system. Which of the equivalences from Theorem 1.8 hold if we assume that is an arbitrary finite graph? Which of them hold if we furthermore assume that 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 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 be the harmonic fan and let be a map. Is it true that the existence of an -expanding arc in is equivalent to the fact that is discontinuous for some ? Does the existence of an -expanding arc in implies that there is such that contains a -divergent sequence?
We finish with two more questions related to Remark 4.1.
Problem 4.7.
Is there a dendrite with finitely many branching points and a map such that yet fails to be equicontinuous? An equivalent way of phrasing the same question is: does the implication of Theorem 1.8 hold for any dendrite with finitely many branching points?
A map as requested in Question 4.7 would still need to be equicontinuous when restricted to (even though itself fails to be equicontinuous). Notice that, for a finite tree and a map, equicontinuity of is equivalent to equicontinuity of the restriction by [11, Theorem 5.2]; on the other hand, [15, Example 2.9] shows that this equivalence no longer holds if 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 belongs to the class and is a map, then the equicontinuity of implies the equicontinuity of ?
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 , 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 , 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 -od, Top. Appl. 265 (2019), 106756.
- [19] S. Willard, General Topology, Dover, Mineola, New York, 2004 (unabridged from the original 1970 version).