Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

DISCRETE AND CONTINUOUS doi:10.3934/dcdss.

2020074
DYNAMICAL SYSTEMS SERIES S
Volume 13, Number 4, April 2020 pp. 1291–1317

A SURVEY OF SOME ASPECTS OF DYNAMICAL TOPOLOGY:


DYNAMICAL COMPACTNESS AND SLOVAK SPACES

Sergiı̆ Kolyada†
Institute of Mathematics, NASU
Tereshchenkivs’ka 3
01601 Kyiv, Ukraine

Dedicated to Professor Jürgen Scheurle on the occasion of his 65th birthday.

Abstract. The area of dynamical systems where one investigates dynamical


properties that can be described in topological terms is “Topological Dyna-
mics”. Investigating the topological properties of spaces and maps that can
be described in dynamical terms is in a sense the opposite idea. This area has
been recently called “Dynamical Topology”. As an illustration, some topolog-
ical properties of the space of all transitive interval maps are described. For
(discrete) dynamical systems given by compact metric spaces and continuous
(surjective) self-maps we survey some results on two new notions: “Slovak
Space” and “Dynamical Compactness”. A Slovak space, as a dynamical ana-
logue of a rigid space, is a nontrivial compact metric space whose homeomor-
phism group is cyclic and generated by a minimal homeomorphism. Dynamical
compactness is a new concept of chaotic dynamics. The omega-limit set of a
point is a basic notion in the theory of dynamical systems and means the col-
lection of states which “attract” this point while going forward in time. It
is always nonempty when the phase space is compact. By changing the time
we introduced the notion of the omega-limit set of a point with respect to a
Furstenberg family. A dynamical system is called dynamically compact (with
respect to a Furstenberg family) if for any point of the phase space this omega-
limit set is nonempty. A nice property of dynamical compactness is that all
dynamical systems are dynamically compact with respect to a Furstenberg
family if and only if this family has the finite intersection property.

1. Introduction. By a (topological ) dynamical system (X, T ) we mean a compact


metric space X with a metric d and a continuous map T : X → X. We call it

2010 Mathematics Subject Classification. Primary: 37B05; Secondary: 54H20, 37B40.


Key words and phrases. Transitive maps, loops, dynamical topology, ω-limit set, dynami-
cal compactness, transitive compactness, sensitive compactness, topological weak mixing, multi-
sensitivity, finite intersection property, Slovak(ian) spaces, rigid spaces, topological entropy, map-
connectedness, chain transitivity.
This survey is based on lectures given by the author at the Max Planck Institute for Mathe-
matics, Technical University of Munich, Paris-Sud University, Luminy Institute of Mathematics,
Institute of Mathematics of Jussieu and several other mathematical departments in 2017.
† Editors’ note: Professor Sergiı̆ Kolyada passed away on May 16, 2018. He will be missed by the

mathematical community, as a mathematician and as a person. Due to his untimely death, Profes-
sor Kolyada could not implement the changes to the first version of the manuscript as suggested in
the (positive) reviews. With the consent of Professor Kolyada’s family, Professor L’ubomı́r Snoha
(Matej Bel University, Banská Bystrica, Slovakia), a friend and colleague of Professor Kolyada,
assumed the responsibility of carrying out the revision. The editors thank Professor Snoha for this
invaluable contribution.

1291
1292 SERGIĬ KOLYADA

trivial if the space is a singleton. Throughout this paper we are only interested in
nontrivial dynamical systems. In Section 3 we additionally assume that X has no
isolated point and T is surjective.
When mathematicians are considering various classes of functions with a natural
topology, often one of the first questions that comes to their mind is about the
topological properties of those spaces. While those questions have been answered
long ago in Mathematical Analysis and Topology, and also in Ergodic Theory there
is a series of papers where topological properties of the group of measure preserving
bijections or homeomorphisms are investigated, see e.g. [25], [31], [15], [10], [43], [51],
it seems that they even have not been asked before in Dynamical Systems. There
are very few papers on topological properties of spaces of maps that have some
specific dynamical properties. Perhaps, the first of them were the paper of Farrell
and Gogolev [14] about the spaces of Anosov diffeomorphisms, the paper [32] that
was being written at the same time as [14] (and completely independently of it),
and the paper [33].
The area of Dynamical Systems where one investigates dynamical properties
that can be described in topological terms is Topological Dynamics. Investigating
topological properties of spaces of maps that can be described in dynamical terms
is in a sense the opposite idea. Therefore in [32] it was proposed to call this area
Dynamical Topology.
In the present paper we survey some recent results in the following areas of
Dynamical Topology:
(1) Topological properties of the space of all transitive maps of a compact interval
to itself and its subspaces.
(2) Dynamical compactness, especially transitive and sensitive compactnesses as
new concepts of chaoticity of a dynamical system.
(3) Slovak spaces, i.e. compact metric spaces whose homeomorphism group is
cyclic and generated by a minimal homeomorphism.
When some notion is not defined where it appears for the first time, the reader
should consult Appendix.

2. Spaces of topologically transitive interval maps. This section is based on


the papers [32], [33].
Let X be a compact metric space and let T : X → X be continuous. The
dynamical system (X, T ) is called topologically transitive (or just transitive) if for
every pair of nonempty open sets U and V in X there is a nonnegative integer n
such that T n (U ) ∩ V 6= ∅. If the space X has no isolated points, this is equivalent
to the existence of a point x ∈ X whose orbit orbT (x) = {x, T (x), . . . , T n (x), . . . }
is dense in X. Consequently, a topologically transitive dynamical system cannot be
decomposed into two disjoint sets with nonempty interiors which do not interact
under the transformation. In particular, transitivity is an ingredient of several
definitions of chaos. For more information on topological transitivity see, e.g., [1],
[37], [40] and references there.
By a map we mean in this section a continuous map on the interval. A lap of an
interval map is a maximal interval on which this map is monotone. The modality of
a piecewise monotone map is the number of laps minus 1. A turning point is a point
that belongs to two distinct laps. When we say “piecewise”, we mean that there are
finitely many pieces. By “slope” we mean the absolute value of the derivative. A
DYNAMICAL TOPOLOGY 1293

full n-horseshoe is a piecewise monotone map with constant slope and n laps, each
of which is mapped to the whole domain of the map.
We will use the following notation.
(1) I = [0, 1];
(2) T – (continuous !) transitive maps I → I
(3) TPM – piecewise monotone transitive maps I → I;
(4) TPL – piecewise linear transitive maps I → I;
(5) Tn – elements of TPM of modality n;
(6) Tn+ – elements of Tn increasing on the first lap;
(7) Tn− – elements of Tn decreasing on the first lap;
(8) CS n – piecewise linear maps I → I, with constant slope and of modality n;
(9) T CS n – transitive maps from CS n .
All those spaces are considered with the C 0 -metric d:
d(f, g) = sup |f (x) − g(x)|.
x∈I

By an interval we mean a nondegenerate interval. If not stated otherwise, it is


assumed to be closed. For an interval J we will denote its length by |J|. When
we speak about symmetry, we mean conjugacy via the symmetry map of I, that is
x 7→ 1 − x.
The space T of all transitive maps on the interval I → I, with uniform metric has
for instance the following properties: is separable; is locally infinite dimensional; is
not complete; is not locally compact; is nowhere dense in the space of all continuous
maps I → I; is a Baire space. Our main aim is to investigate connectedness and
loops in subspaces of T . In such an investigation, we need to be sure that all
considered maps really belong to T . Since we will mostly use piecewise linear maps
with constant slope, before going further we address the problem when such maps
are transitive.

2.1. How to recognize that a piecewise linear map with constant slope is
transitive. We use a coding of maps f ∈ CS n as follows.
(1) Code of f = (f (0), f (1st turning pt), . . . , f (nth turning pt), f (1)).
(2) When we consider a union of spaces CS n with different n0 s, we use the common
(largest) length of codes, say
code of tent map = (0, 1, 0) = (0, 0, 1, 0) = (0, 1, 0, 0) = . . . ,
One can show the following two properties. Pn+1
(3) If code of f = (a0 , . . . , an+1 ) then slope of f = j=1 |aj − aj−1 |.
Sn
(4) f ∈ i=1 CS i depends continuously on the parameters a0 , a1 , . . . , an+1
(jointly).
It is easy to describe the space T CS 1 .
Lemma 2.1 ([33]). A map f ∈ CS 1 is transitive if and only if it has the code

(1) (a, 1, 0) where a ∈ [0,
√ 2 − 2], or
(2) (1, 0, c) where c ∈ [ 2 − 1, 1].
If a map is in T CS 2 then one can show that it has one of the four codes (a, 1, 0, d),
(a, 0, 1, d), (1, 0, c, d), (a, b, 1, 0). To save the space, the results proved in several lem-
mas in [33] are presented in Table 1. Alternatively, one could use [45, Theorem 1].
Using these results, the space T CS 1 ∪ T CS 2 has been drawn in [33], see Figure 1.
1294 SERGIĬ KOLYADA

CS 2
code picture condition equivalent to transitivity

2 2
(a, 1, 0, d) d≤a−4+ a or 1 − a ≤ (1 − d) − 4 + 1−d

(a, 0, 1, d) a>d

1
(1, 0, c, d) d ≤ 2 + 2c − c

1
(a, b, 1, 0) 1 − a ≤ 2 + 2(1 − b) − 1−b
Table 1. Transitive maps in CS 2

Figure 1. The space T CS 1 ∪ T CS 2 .


DYNAMICAL TOPOLOGY 1295

If n ≥ 3, the situation is more complicated. However, also in this case there


are useful sufficient conditions for transitivity. We mention at least some of them
(alternatively, one can use [45, Theorem 1]).
Lemma 2.2 ([33]). Let f ∈ CS n with slope λ > 2. Assume that the image of every
lap (except perhaps the leftmost and the rightmost ones) is the whole I. Then f is
transitive.
Lemma 2.3 ([33]). Let f ∈ CS n with slope λ > 3. Assume that the image of
every lap (except perhaps one or two leftmost or one or two rightmost ones) is the
whole I. Then f is transitive.
Put E(f ) := {turning points of f } ∪ {endpoints of the interval I}.
Lemma 2.4 ([33]). Let f ∈ CS n with slope λ > 3. Assume that f ({0, 1}) ⊂ {0, 1}
and that out of any four consecutive points of E(f ) at least one is mapped to 0 and
at least one is mapped to 1. Then f is transitive.
2.2. Connectedness properties of spaces of transitive interval maps. The
following theorem shows what kind of results belong to Dynamical Topology.
Theorem 2.5 ([32]). The spaces T , TPM and TPL are contractible (hence arcwise
connected) and locally arcwise connected.
We recall here the idea of the proof of contractibility of T .
Given a closed interval K ⊂ I, we consider box map as in Figure 2.

Figure 2. Box map

It depends continuously (jointly) on 5 parameters:


• al and ar – the values at the left and right endpoint of K,
• ab and at – the bottom and the top of the box,
−ab
• as – the slope multiplier (the slope is as · at|K| , here as = 20)
Given f ∈ T , we put
1. gf,0 = f ;
2. to define gf,t for 0 < t ≤ 1, we partition [0, 1] into intervals I0 , I1 , . . . , Is of
lengths t (the rightmost one can be shorter). Over each Ii we construct a box
as in Figure 3. In each box we choose a box map which coincides with f at
the endpoints of Ii and has as = 20. The obtained map I → I is gf,t .
Vertical sides of the boxes can be chosen such that:
• gf,t is transitive;
• gf,t depends continuously on f and t (jointly).
If t = 1 then we have just one box and so the maps gf,1 are box maps with K =
[0, 1], ab = 0, at = 1 and as = 20. They depend only on al = f (0) ∈ [0, 1] and
1296 SERGIĬ KOLYADA

Figure 3. Boxes Ii × Ij

ar = f (1) ∈ [0, 1]. Hence the set Z = {gf,1 : f ∈ T }, being homeomorphic to the
square, is contractible.
Our family gf,t can be treated as a homotopy joining
1. the identity idT : T → T , f 7→ f = gf,0 and
2. the map T → Z ⊂ T , f 7→ gf,1 .
Since Z is contractible, also T is contractible.
Theorem 2.6 ([32]). Each space Tn has two connected components, Tn+ and Tn− ,
and they are arcwise connected. The distance between T1+ and T1− is positive, while
the distance between Tn+ and Tn− is zero for all n ≥ 2.
2.3. Loops of transitive interval maps. We are going to consider loops of tran-
sitive maps with constant slopes (the reader may wish to check that transitivity
of maps considered below in this subsection always follows from results in Subsec-
tion 2.1; we will not emphasize it any more).
Theorem 2.7 ([32]). For every n ≥ 1 there is a loop in Tn ∪ Tn+1 , which is not
contractible in Tn ∪ Tn+1 .
In fact, we have found such a loop in T CS n ∪ T CS n+1 . We call it the basic loop
of order n and we denote it by Ln . It can be described as follows. We start with
the full (n + 1)-horseshoe which is increasing on the first lap, then an additional
rightmost lap starts to grow until we get a full (n + 2)-horseshoe. Next the leftmost
lap starts to disappear, until we get the full (n + 1)-horseshoe which is decreasing
on the first lap. Then we repeat those operations and we end up with the initial
(n + 1)-horseshoe. For n = 2 see Figure 4. A rigorous description of the basic loops
can be obtained by using codes, see [33].

Figure 4. Basic loop L2 . It consists of four arcs represented by


the four rows in this picture (instead of all elements of such an arc
only five of them are shown).
DYNAMICAL TOPOLOGY 1297

In particular, L2 ∈ T CS 2 ∪ T CS 3 is not contractible in T CS 2 ∪ T CS 3 (even not


contractible in T2 ∪ T3 , as claimed in Theorem 2.7).
We show that L2 is contractible in T CS 2 ∪ T CS 3 ∪ T CS 4 . First we find the
auxiliary loop of order 2 homotopic to L2 in the space T CS 2 ∪ T CS 3 ∪ T CS 4 . Then
one can show that the auxiliary loop is contractible in T CS 2 ∪ T CS 4 .
The basic loop L2 consists of four arcs, so we need to show how to deform these
four arcs, to obtain the auxiliary loop. This is shown in Figures 5–8. Of course,
those deformations can be rigorously described in codes. Say, for the deformation
of the 1st arc, Figure 5, the codes are (0, 1, 0, 1, 1 − s, 1 − s + st), s varies from 0
(left end) to 1 (right end), t varies from 0 (for basic loop) to 1 (for auxiliary loop).

Figure 5. Deformation Figure 6. Deformation


of the 1st arc of L2 . of the 2nd arc of L2 .

Figure 7. Deformation Figure 8. Deformation


of the 3rd arc of L2 . of the 4th arc of L2 .

The basic loop L2 consisting of four arcs has been deformed to the auxiliary loop
consisting of two arcs (formally four arcs, but the second and the third are constant,
so we can ignore them), see Figure 9.
The auxiliary loop consisting of two arcs can be seen in Figures 10 and 11.
Still we need to show that the auxiliary loop consisting of two arcs is contractible
in T CS 2 ∪ T CS 4 . The techniques for showing that are similar to those already used
1298 SERGIĬ KOLYADA

Figure 9. From L2 to auxiliary loop consisting of two arcs.

Figure 10. The 1st arc Figure 11. The 2nd arc
of the auxiliary loop ob- of the auxiliary loop ob-
tained from L2 . tained from L2 .

above and therefore we refer the reader to [33] for details. Here we only reveal the
idea: One can show that the first of these two arcs can be deformed, in T CS 2 ∪T CS 4 ,
to the second arc run backward. Moreover, during that homotopy the endpoints
of the arc are not modified and so we get contractibility of the auxiliary loop in
T CS 2 ∪ T CS 4 . Thus, L2 is really contractible in T CS 2 ∪ T CS 3 ∪ T CS 4 .
We worked with the basic loop L2 . More generally, in [33] the following is shown
(recall that Ln was a loop in T CS n ∪ T CS n+1 and not contractible there, even not
contractible in Tn ∪ Tn+1 ).

Theorem 2.8 ([33]). Let Ln be the basic loop constructed in the proof of Theo-
rem 2.7.
(1) If n ≥ 2 then Ln can be contracted in T CS n ∪ T CS n+1 ∪ T CS n+2 .
(2) If n ≥ 3 then Ln can be contracted also in T CS n−1 ∪ T CS n ∪ T CS n+1 .
(3) L1 can be contracted in T CS 1 ∪ T CS 2 ∪ T CS 4 .

The situation seems to be similar as for the following model, although we do not
know how far we can go with this analogy. Think about the sequence of spaces Rn ,
n = 0, 1, 2, . . . , where each space is a subset of the next one. Set Rn = Rn \ Rn−1
for n = 1, 2, 3, . . . . Then the fundamental group of the space
Rn ∪ Rn+1 = Rn+1 \ Rn−1 = (R2 \ {0}) × Rn−1
is nontrivial, while the fundamental group of the space
Rn ∪ Rn+1 ∪ Rn+2 = Rn+2 \ Rn−1 = (R3 \ {0}) × Rn−1
is trivial.

3. Dynamical compactness. This section is based on the papers [28] and [29].
Throughout this section, when considering a dynamical system (X, T ), we addition-
ally assume that the compact metric space X has no isolated point and that the
continuous map T is surjective.
DYNAMICAL TOPOLOGY 1299

3.1. Furstenberg families and dynamical compactness. Let Z+ be the set of


all nonnegative integers and N the set of all positive integers. Before going on, let
us recall the notion of a Furstenberg family from [1]. Denote by P = P(Z+ ) the set
of all subsets of Z+ . A subset F ⊂ P is a (Furstenberg) family, if it is hereditary
upward, that is, F1 ⊂ F2 and F1 ∈ F imply F2 ∈ F. Any subset A of P clearly
generates a family {F ∈ P : F ⊃ A for some A ∈ A}. Denote by B the family of all
infinite subsets of Z+ , and by P+ the family of all nonempty subsets of Z+ . For a
family F, the dual family of F, denoted by kF, is defined as
kF = {F ∈ P : F ∩ F 0 6= ∅ for any F 0 ∈ F}.
A family F is proper if it is a proper subset of P, that is, Z+ ∈ F and ∅ ∈ / F. By
a filter F we mean a proper family closed under intersection, that is, F1 , F2 ∈ F
implies F1 ∩ F2 ∈ F. A filter is free if the intersection of all its elements is empty.
We extend this concept, a family F is called free if the intersection of all elements
of F is empty.
Let (X, T ) be a dynamical system. For any F ∈ P, every point x ∈ X and each
subset G ⊂ X, we define the orbit T F x = {T i x : i ∈ F }, and the visiting times
nT (x, G) = {n ∈ Z+ : T n x ∈ G}. The ω-limit set of x with respect to a family F
(see [1]), or shortly the ωF -limit set of x, denoted by ωF (x), is defined as
\
T F x = {z ∈ X : nT (x, G) ∈ kF for every neighborhood G of z}.
F ∈F

Let us note that not always ωF (x) is a subset of the ω-limit set ωT (x), which is
defined as
\∞
{T k x : k ≥ n} = {z ∈ X : nT (x, G) ∈ B for every neighborhood G of z}.
n=1

For instance, if each element of F contains 0 then any point x ∈ ωF (x). Neverthe-
less, if a family F is free, then ωF (x) ⊂ ωT (x) for any point x ∈ X and if (X, T )
has a nonrecurrent point1 , then the converse is true.
A dynamical system (X, T ) is called compact with respect to F, or shortly dy-
namically compact, if the ωF -limit set ωF (x) is nonempty for all x ∈ X.
H. Furstenberg started a systematic study of transitive systems in his paper on
disjointness in topological dynamics and ergodic theory [17], and the theory was
further developed in [19] and [18]. Recall that the system (X, T ) is (topologically)
transitive if NT (U1 , U2 ) = {n ∈ Z+ : U1 ∩ T −n U2 6= ∅} (= {n ∈ Z+ : T n U1 ∩ U2 6=
∅}) ∈ P+ for any opene 2 subsets U1 , U2 ⊂ X, equivalently, NT (U1 , U2 ) ∈ B for any
opene subsets U1 , U2 ⊂ X.
3.2. Transitive compactness, sensitive compactness, FIP. Transitive com-
pactness is one of possible dynamical compactnesses. Let NT be the set of all
subsets of Z+ containing some NT (U, V ), where U, V are opene subsets of X. A
dynamical system (X, T ) is called transitively compact, if for any point x ∈ X the
ωNT -limit set ωNT (x) is nonempty, in other words, for any point x ∈ X there exists
a point z ∈ X such that
nT (x, G) ∩ NT (U, V ) 6= ∅
for any neighborhood G of z and any opene subsets U, V of X.
1A point x ∈ X is called recurrent if x ∈ ωT (x).
2 Because we so often have to refer to open, nonempty subsets, we will call such subsets opene.
1300 SERGIĬ KOLYADA

Let (X, T ) and (Y, S) be two dynamical systems and k ∈ N. The product system
(X × Y, T × S) is defined naturally, and denote by (X k , T (k) ) the product system of
k copies of the system (X, T ). Recall that the system (X, T ) is minimal if it does
not admit a nonempty, closed, proper subset K of X with T K ⊂ K, and is weakly
mixing if the product system (X 2 , T (2) ) is transitive. Any transitively compact
system is obviously topologically transitive, and observe that each weakly mixing
system is transitively compact ([3]).
Sensitive compactness is another possible dynamical compactness. The notion
of sensitivity was first used by Ruelle [46]. It captures the idea that in a chaotic
system a small change in the initial condition can cause a big change in the trajec-
tory. According to the works by Guckenheimer [24] and Auslander and Yorke [7],
a dynamical system (X, T ) is called sensitive if there exists δ > 0 such that for
every x ∈ X and every neighborhood Ux of x, there exist y ∈ Ux and n ∈ N with
d(T n x, T n y) > δ. Such a δ is called a sensitive constant of (X, T ). Recently in
[42] Moothathu initiated a way to measure the sensitivity of a dynamical system,
by checking how large is the set of nonnegative integers for which the sensitivity
occurs. For a positive δ and a subset U ⊂ X define
ST (U, δ) = {n ∈ Z+ : there are x1 , x2 ∈ U such that d(T n x1 , T n x2 ) > δ}.
Let ST (δ) be the set of all subsets of Z+ containing some ST (U, δ), where U is
an opene subset of X. A dynamical system (X, T ) is called sensitively compact, if
there exists δ > 0 such that for any point x ∈ X the ωST (δ) -limit set ωST (δ) (x) is
nonempty, in other words, for any point x ∈ X there exists a point z ∈ X such that
nT (x, G) ∩ ST (U, δ) 6= ∅
for any neighborhood G of z and any opene subset U in X.
A dynamical system (X, T ) is called multi-sensitive if there exists δ > 0 such
Tk
that i=1 ST (Ui , δ) 6= ∅ for any finite collection of opene U1 , . . . , Uk ⊂ X. Such a
δ is called a constant of multi-sensitivity of (X, T ).
Recall that a collection A of subsets of a set Y has the finite intersection property
(FIP) if the intersection of all sets in any finite subcollection of A is nonempty. If
such intersections are infinite, we speak on strong finite intersection property (SFIP).
It is well known that the FIP is useful in formulating an alternative definition of
compactness of a topological space: a topological space is compact if and only if
every collection of closed subsets satisfying the FIP has a nonempty intersection
itself.
We recall that if (X, T ) is weakly mixing then it is well known that the family
NT is a filter (see [17], [1]), and hence has FIP. If (X, T ) is a multi-sensitive system
with a constant of multi-sensitivity δ > 0 then obviously the family ST (δ) has FIP.
Since these families are also free, actually they have SFIP.
In fact we can say more — the FIP is useful in characterizing the dynamical
compactness.
Theorem 3.1 (Theorem FIP, [28]). All dynamical systems are dynamically compact
with respect to F if and only if the family F has the finite intersection property.
Proof. Sufficiency. Suppose that F has FIP. Take arbitrary dynamical system
(X, T ) and let x ∈ X. Obviously the family {T F x : F ∈ F} also has FIP, and
then by compactness of X the family {T F x : F ∈ F} has a nonempty intersection
itself, i.e., ωF (x) 6= ∅. Thus (X, T ) is dynamically compact with respect to F.
DYNAMICAL TOPOLOGY 1301

Necessity. Suppose that the family F does not have FIP. Then there is a collection
Tk
{F1 , . . . , Fk } ⊂ F with i=1 Fi = ∅. Let A = {a1 , . . . , ak } be an alphabet and let
(X, T ) := (Σ, σ) be the full (one-sided) A-shift. We are going to define a point
x ∈ X with ωF (x) = ∅. Let x0 = a1 . For any n ≥ 1 there is i with n ∈ / Fi , else
the intersection of F1 , . . . , Fk would be nonempty. Then define xn := ai . Finally,
let x = x0 x1 x2 x3 . . . and the construction is finished.
Assume the contrary that we can take z ∈ ωF (x), and that z begins with ai ∈ A.
Take Gz = C[ai ]. As z ∈ ωF (x) we have nT (x, Gz ) ∩ Fi 6= ∅. But if n ∈ nT (x, Gz ),
then xn = ai and so n ∈ / Fi by the construction, a contradiction.
3.3. Dynamical compactness with respect to an arbitrary family. As we
have mentioned, any filter has FIP; if (X, T ) is weakly mixing then the family NT
is a filter; if A is a weakly mixing subset of (X, T ) then the family NT (A) has FIP;
and if (X, T ) is a multi-sensitive system with a constant of multi-sensitivity δ > 0
then the family ST (δ) also has FIP.
A collection H ⊂ F will be called a base for F if for any F ∈ F there is H ∈ H
with H ⊂ F . We are interested in those families which have a countable base.
Note that not every Furstenberg family F has a countable base, for example, the
family B. Assume the contrary that B admits a countable base {Fn : n ∈ N}. We
take k1 ∈ F1 , and once km ∈ Fm , m ∈ N is defined we choose km+1 ∈ Fm+1 with
km+1 > km + m + 1. Set E = {kn : n ∈ N} and F = Z+ \ E. Then E ∩ Fn 6= ∅ for
all n ∈ N, and F ⊃ {km + m : m ∈ N} and hence F ∈ B, in particular, there exists
no n ∈ N with Fn ⊂ F , a contradiction.
It is not hard to show even the existence of a family with FIP, but without
a countable base. Nevertheless the families NT and ST (δ) have countable bases.
Indeed, we can consider a countable base U of open sets for the space X. Note that
U1 ⊂ U , V1 ⊂ V implies NT (U1 , V1 ) ⊂ NT (U, V ) and ST (U1 , δ) ⊂ ST (U, δ). Then
{NT (U, V ) : U, V ∈ U } and {ST (U, δ) : U ∈ U } are countable bases for NT and
ST (δ), respectively.
The following is a general result that will be especially useful for families with
countable bases. By TranF (X, T ) we denote the set of all F-transitive points in
(X, T ), see Appendix for the definition.
Proposition 3.2 ([29]). Let (X, T ) be a dynamical system and let F be a family
such that there exists x ∈ TranF (X, T ). Then orbT (x) ⊂ TranF (X, T ).
Proposition 3.3 ([29]). Assume that F admits a countable base H. Then
TrankF (X, T ) is a Gδ subset of X. Moreover, the following are equivalent:
(1) The system (X, T ) is kF-transitive,
(2) TrankF (X, T ) is a dense Gδ subset of X,
(3) TrankF (X, T ) 6= ∅.
3.4. Transitive sensitivity and sensitive compactness. A dynamical system
(X, T ) is transitively sensitive if there exists δ > 0 such that ST (W, δ) ∩ NT (U, V ) 6=
∅ for any opene subsets U, V, W of X. Recall that it is sensitively compact if there
exists δ > 0 such that for any point x ∈ X the set ωST (δ) (x) is nonempty. Sometimes
in such cases we will also say that (X, T ) is transitively sensitive with a sensitive
constant δ and (X, T ) is sensitively compact with a sensitive constant δ. Then
Theorem 3.4 ([29]). Let (X, T ) be a minimal system. Then the following condi-
tions are equivalent:
1. (X, T ) is multi-sensitive.
1302 SERGIĬ KOLYADA

2. (X, T ) is sensitively compact.


3. There exists δ > 0 such that ωST (δ) (x) = X for each x ∈ X.
4. There exist δ > 0 and x ∈ X with ωST (δ) (x) = X.
5. (X, T ) is transitively sensitive.
Before proceeding, we need the following lemma.
Lemma 3.5 ([29]). Let δ > 0 and x ∈ X. If T : X → X is almost open, then the
family ST (δ) is negatively, and the subset ωST (δ) (x) is positively T -invariant.
The following result gives a characterization of transitive sensitivity in terms of
dynamical compactness.
Proposition 3.6 ([29]). Let (X, T ) be a dynamical system. Then the family ST (δ)
is positively invariant for any δ > 0. Furthermore, the following conditions are
equivalent:
1. (X, T ) is transitively sensitive.
2. There exist δ > 0 and a dense Gδ subset X0 ⊂ X such that ωST (δ) (x) = X
for each x ∈ X0 .
3. There exist δ > 0 and a point x ∈ X with ωST (δ) (x) = X.
Recall that by [21, Corollary 1.7] the sensitivity of a dynamical system can be
lifted up from a factor to an extension by an almost open factor map between
transitive systems. The following result shows that the transitive sensitivity can
be lifted up to an extension from a factor by an almost one-to-one factor map and
that the transitive sensitivity is projected from an extension to the sensitivity of a
factor by a weakly almost one-to-one factor map.
Lemma 3.7 ([29]). Let π : (X, T ) → (Z, R) be a factor map between dynamical
systems.
1. Assume that π is almost one-to-one. If (Z, R) is transitively sensitive with a
sensitive constant δ > 0 then (X, T ) is also transitively sensitive.
2. Assume that there exists z ∈ Z whose fiber is a singleton. If (X, T ) is transi-
tively sensitive then (Z, R) is sensitive, in particular, Eq(Z, R) = ∅.
Now let us give the proof of Theorem 3.4 taken from [29].
Proof of Theorem 3.4. (1) ⇒ (2) follows directly from the definitions. As the sys-
tem (X, T ) is minimal, the map T : X → X is almost open. Observing that
ωST (δ) (x) is a closed subset of X for each x ∈ X, the implication (2) ⇒ (3) follows
from Lemma 3.5 and the minimality of (X, T ). The implications (3) ⇒ (4) ⇒ (5)
follow from Proposition 3.6. Since a minimal system is either multi-sensitive or a
weakly almost one-to-one extension of its maximal equicontinuous factor by [30],
then (5) ⇒ (1) follows from Lemma 3.7. This finishes the proof.
Clearly each multi-sensitive system is sensitively compact. Observe that by[28,
Theorem 4.7] we have the following.
Proposition 3.8 ([28]). Each non-proximal, transitively compact system (X, T ) is
multi-sensitive.
In particular, each minimal transitively compact system is multi-sensitive, as each
minimal proximal system is trivial by [3] and all dynamical systems considered are
assumed to be nontrivial. Nevertheless, there are many minimal, non transitively
compact, multi-sensitive systems. For example, consider the classical dynamical
DYNAMICAL TOPOLOGY 1303

system (X, T ) given by X = R2 /Z2 and T : (x, y) 7→ (x + α, x + y) with α ∈ / Q (see


[18, Chapter 1]). As commented in [28, p. 1816], (X, T ) is an invertible minimal
multi-sensitive system; note that (X, T ) is not weakly mixing, since (X, T ) admits
an irrational rotation as its nontrivial equicontinuous factor and any equicontinuous
factor of a weakly mixing system is trivial. Recall that by [28, Corollary 3.10] for a
minimal system it holds that the system is transitively compact if and only if it is
weakly mixing, and so the constructed system (X, T ) is not transitively compact.
Proposition 3.9 ([29]). Each nontrivial weakly mixing system (X, T ) is transitively
sensitive.
We give a sufficient condition for a dynamical system to be transitively sensitive
(by Proposition 3.6) as the last result of this subsection.
Lemma 3.10 ([29]). Assume ωST (ε) (x) = X for some x ∈ X and ε > 0. Then
there is δ > 0 such that for any opene subset U of X and each neighbourhood Ux of
x there are y ∈ Ux and n ∈ nT (x, U ) with d(T n x, T n y) > δ. If in addition, the map
T : X → X is almost one-to-one, then the converse holds.
3.5. Transitively compact (non weakly mixing) systems. Recall that the
system (X, T ) is totally transitive if (X, T k ) is transitive for each k ∈ N; and is
topologically mixing if NT (U, V ) ∈ Fcof for any opene subsets U, V in X. Note that
(X, T ) is weakly mixing if and only if NT (U, V ) ∈ Fthick for any opene sets U, V in
X by [17, 44], and so any weakly mixing system is totally transitive. It is direct
to check that each weakly mixing system is transitively compact. We extend it as
follows.
Theorem 3.11 ([29]). There are non-totally transitive, transitively compact sys-
tems and totally transitive, transitively compact systems which are not weakly mix-
ing.
The following result is proved independently in [11] and [48].
Lemma 3.12 ([29]). No ω-limit set ωT (x) can be decomposed into α disjoint closed,
nonempty, positively T -invariant subsets, where 2 ≤ α ≤ ℵ0 .
Before proceeding we need the following example, as it is crucial in our arguments.
Proposition 3.13 ([29]). For any given compact metric space Z, there exists a
topologically mixing system (X, T ) such that Z can be realized as the set of all of its
minimal points, furthermore, its each minimal point is a fixed point.
The following result shows that in general there is no topological structure similar
to Lemma 3.12 for the ωNT -limit sets.
Theorem 3.14 ([29]). For any given compact metric space Z, there exists a non
totally transitive, transitively compact system (X, T ) such that Z can be realized
as the set of all its minimal points with each minimal point being a fixed point,
furthermore, Z is realized as ωNT (x) for some x ∈ X.
Note that a dynamical system is proximal if and only if it contains the unique
fixed point, which is the only minimal point of the system [3]. Thus, as a direct
corollary of Lemma 3.12 and Theorem 3.14, we have:
Corollary 3.15 ([29]). There exists a non-proximal, non-totally transitive, transi-
tively compact system (X, T ) and a point x0 ∈ X such that ωNT (x0 ) 6= ωT (x) for
all x ∈ X.
1304 SERGIĬ KOLYADA

Nevertheless the following is still open.


Question 3.16 ([29]). Let (X, T ) be a weakly mixing system. Is there a point
x ∈ X and 2 ≤ α ≤ ℵ0 such that ωNT (x) can be decomposed into α disjoint closed,
nonempty, positively T -invariant subsets?
At the end of this subsection let us mention one more chaotic property of tran-
sitively compact systems in addition to already known.
A pair of points x, y ∈ X is proximal (asymptotic) if lim inf n→∞ d(T n x, T n y) =
0 (limn→∞ d(T n x, T n y) = 0, respectively). Denote by P roxT (X) and by AsymT (X)
the set of all proximal pairs and asymptotic pairs of points, respectively. Any pair
(x, y) ∈ ProxT (X) \ AsymT (X) is called a Li-Yorke pair. Recall that a dynamical
system (X, T ) is Li-Yorke chaotic if there exists an uncountable set S ⊂ X with
(S × S) \ ∆2 (X) ⊂ ProxT (X) \ AsymT (X), where ∆2 (X) = {(x, x) : x ∈ X}.
Proposition 3.17 ([29]). Each transitively compact system (X, T ) is Li-Yorke
chaotic.
Observe that in [34] we initiated another way to measure the sensitivity of a
system, that is, we gave quantitative measures of the sensitivity of a dynamical
system by introducing the Lyapunov numbers:
Lr = sup{δ : for every x ∈ X and every open neighborhood Ux of x there
exist y ∈ Ux and a nonnegative integer n with
d(T n x, T n y) > δ};

Lr = sup{δ : for every x ∈ X and every open neighborhood Ux of x there


exists y ∈ Ux with lim sup d(T n x, T n y) > δ};
n→∞
Ld = sup{δ : in any opene U ⊂ X there exist x, y ∈ U and a nonnegative
integer n with d(T n x, T n y) > δ};
Ld = sup{δ : in any opene U ⊂ X there exist x, y ∈ U with
lim sup d(T n x, T n y) > δ}.
n→∞

Here we set sup ∅ = 0 by convention. Various definitions of sensitivity, formally


give us different Lyapunov numbers. Nevertheless, as was shown in [34], for minimal
topologically weakly mixing systems all these Lyapunov numbers are the same.
The motivation of [34] comes from the following proposition.
Proposition 3.18 (cf. [3]). The following conditions are equivalent:
1. (X, T ) is sensitive.
2. There exists δ > 0 such that for every x ∈ X and every neighborhood Ux of x,
there exists y ∈ Ux with lim supn→∞ d(T n x, T n y) > δ.
3. There exists δ > 0 such that in any opene U in X there are x, y ∈ U and a
nonnegative integer n with d(T n x, T n y) > δ.
4. There exists δ > 0 such that in any opene U ⊂ X there are x, y ∈ U with
lim supn→∞ d(T n x, T n y) > δ.
3.6. Concluding remarks. Figure 12 presents a comparison between stronger
forms of sensitivity for transitive systems. It includes also some new elements in
the diagram, which are not in [28] and [29].
DYNAMICAL TOPOLOGY 1305

Recall that (X, T ) is Li-Yorke sensitive if there exists δ > 0 such that for every
x ∈ X and every neighborhood Ux of x, there exist y ∈ Ux such that (x, y) is a
proximal pair while lim supn→+∞ d(T n x, T n y) > δ.
Let (X, T ) be a dynamical system which is compact with respect to a (dynamical)
Furstenberg family F. Say, F is NT or ST . Then it is not so hard to show that in
that case for any point x ∈ X there exist a minimal subset Mx of X and a point
z ∈ Mx such that
nT (x, Gz ) ∩ F 6= ∅
for any neighborhood G of z in X and any F ∈ F. If A is a set in X, by Bε [A]
we will denote the union of all open balls of radius ε > 0 whose centers run over
A. Similarly as in the proof of [28, Lemma 3.12] one can show that for any x ∈ X
and any Bε [Mx ] the set nT (x, Bε [Mx ]) is thickly syndetic. Let nT be the set of all
subsets of Z+ containing some nT (x, Bε [Mx ]), where x is a point of X, ε > 0 and
Mx ⊂ ωT (x) is a minimal subset of X.

Figure 12. Topologically transitive systems.

Notice that Figure 12 presents also an open question. Are all transitively com-
pact systems sensitively compact? More precisely, does there exist a proximal,
transitively compact system which is not sensitively compact? Nevertheless, for
topologically transitive, non-proximal dynamical systems we have more clear situ-
ation, see Figure 13.

4. Slovak spaces. This section is mostly based on the papers [35] and [13].

4.1. Semigroup of continuous maps, group of homeomorphisms, func-


tional envelopes. Given a compact metric space X, let S(X) be the set of all
continuous maps X → X. It is a topological semigroup under composition and
1306 SERGIĬ KOLYADA

Figure 13. Topologically transitive, non-proximal systems.

compact-open topology. Similarly, let H(X) be the topological group of all home-
omorphisms X → X.
A dynamical system (X, T ) gives rise to several “hyper-systems”. For example,
T acts naturally on 2X , the space of all compact subsets of X equipped with the
Hausdorff metric. The induced map here is the continuous map T̃ : 2X → 2X given
by T̃ (A) = T (A) for each A ∈ 2X .
In the same spirit (although now the phase space is not necessarily compact),
jointly with J. Auslander and L’. Snoha in [6], we introduced and started to study
the “hyper-system” (S(X), FT ), where the transformation
FT : S(X) → S(X) is defined by FT ϕ = T ◦ ϕ for any ϕ ∈ S(X).
We called this system the functional envelope of (X, T ). Since the metric space X is
compact, the compact-open topology on S(X) is compatible both with the uniform
metric and with the Hausdorff metric (applied to the graphs of maps). If S(X) is
viewed as a metric space with the uniform or Hausdorff metric, the notation SU (X)
or SH (X) is used. If the properties under consideration do not depend on which of
the two metrics is used in S(X), we will sometimes keep the notation S(X) instead
of SU (X) or SH (X).
Why the name “functional envelope”? Because the system (S(X), FT ) contains
(properly, if card X ≥ 2) an isomorphic copy of (X, T ). The map ι : X → S(X)
sending a point a ∈ X to the constant map x 7→ a, x ∈ X, is an isometry (regardless
of whether the uniform metric or the Hausdorff metric is used in S(X)) and also a
topological conjugacy, i.e., ι ◦ T = FT ◦ ι.
DYNAMICAL TOPOLOGY 1307

The “uniform” functional envelope (SU (X), FT ) and the “Hausdorff” functional
envelope (SH (X), FT ) may differ in some dynamical properties which depend on
the metric (for example the topological entropy in non-compact systems does). On
the other hand, the essential dynamical properties of these functional envelopes do
not depend on the metric.
Now suppose additionally that the map T : X → X is a homeomorphism. Then
the space H(X) is invariant under FT and (H(X), FT ) becomes another kind of a
“functional envelope”. Here again
FT : H(X) → H(X) is defined by FT ϕ = T ◦ ϕ for any ϕ ∈ H(X).
One of the main topics of our research in this section is to study relations between
some properties (structure) of the topological semigroup S(X), group H(X) and
possible values of the topological entropy (and/or some other chaotic properties) of
its elements (continuous maps and homeomorphisms, respectively).
In [35] we have mostly considered the following two questions: 1) when does
a compact metric space admit a continuous map (homeomorphism) with positive
topological entropy? 2) when does the existence of a positive-entropy continuous
map on a compact metric space imply the existence of a +∞-entropy continuous
map? We have proved there the following.
Theorem 4.1 ([35]). Let X be a compact metric space. If S(X) is compact, then
for any f ∈ S(X) topological entropy of (X, f ) and topological entropy of the func-
tional envelope (S(X), Ff ) are zero. If H(X) is compact, then for any f ∈ H(X)
topological entropy of (X, f ) is also zero.
At present, not so much is known about the topological structure of compact
semigroups S(X) and groups H(X). The compactness of S(X) and H(X) is not
a very strict condition and occurs sometimes, because both of them may be “very
small”. Recall that a topological space X is said to be rigid (for homeomorphisms),
if the full topological homeomorphism group H(X) contains only the identity. De
Groot and Wille in [22] showed the existence of rigid spaces even as locally con-
nected one-dimensional continua (Peano curves) of the plane. The main idea of the
construction of such a space is similar to that of the Sierpinski carpet, which is a
square with interiors of a dense family of subsquares removed, see Figure 14. The
Sierpinski carpet, also known as the Sierpinski universal curve, is a one-dimensional
planar Peano continuum.

Figure 14. The first 4 steps in the construction of the Sierpinski carpet.

One of the De Groot – Wille rigid plane locally connected one-dimensional con-
tinua is a disc with interiors of a dense family of propellers (with different numbers
of blades) removed, see Figure 15. To be more precise, by [22], consider a disc D in
the plane. Let {ai } be a countable dense subset of the interior of D. We define a
sequence of “propellers” in D. The first is bounded by a two-bladed curve having
a1 as its centre, which avoids the boundary of D. Suppose the first n − 1 propellers
have already been defined. Let a0n be the first member of the sequence {ai } which
1308 SERGIĬ KOLYADA

is in no previously constructed propeller. Then the n-th propeller is n + 1 bladed,


with centre at a0n , and lies inside a circle which misses all previously constructed
propellers and the boundary of D. Moreover, we take care that the diameters of
the propellers tend to zero. The space P is the disc D with the interiors of all
the propellers removed. Then P is a continuum as the intersection of a countable,
decreasing sequence of continua. Routine procedure shows the local connectedness
of P . P has dimension one, since it does not contain a subset, open in the plane.
Since the total area of the interiors of the propellers can be chosen as small as we
want it to be, P can also have positive measure.

Figure 15. Steps in the construction of the De Groot – Wille rigid


plane continuum.

As a curiosity, notice that the Julia set from Figure 16 (a picture by Volodymyr
Nekrashevych), though not being rigid, reminds the De Groot – Wille rigid conti-
nuum.

Figure 16. The Julia set for the map z 7→ (z 2 + 0.3 + 0.05i)/(z 2 − 1).

Using such kind of ideas, de Groot proved the following.


Theorem 4.2 ([23]). Let G be an arbitrary group. Then there exists a connected,
locally connected, complete metric space X (or alternatively compact, Hausdorff )
for which the group of all autohomeomorphisms H(X) is isomorphic to G.
However, such a space need not exist in the class of compact metric spaces,
because a compact metric space has cardinality at most c, while there are groups
DYNAMICAL TOPOLOGY 1309

of arbitrary cardinalities. Nevertheless, as De Groot and Wille proved, if G is


countable then X can be chosen to be a Peano continuum of any positive dimension.
Cook (in [8]) constructed two metric continua X and Y such that
1. the space S(X) consists only of the constant maps and the identity on X;
2. the space H(Y ) is topologically equivalent to the Cantor set.
But it is still an open problem what can we say about the topological structure
of the compact full topological homeomorphism group H(X) and of the compact
topological semigroup S(X) (see the conjecture below).
Recall that a topological group is called a profinite group if it is Hausdorff, com-
pact, and totally disconnected. Gartside and Glyn ([20]) have established that every
metric profinite group is the full homeomorphism group of a continuum. Recently
Hofmann and Morris in [26] proved that a compact full homeomorphism group of
a Tychonoff space is a profinite topological group. But the following conjecture is
still open.
Conjecture 4.3 ([27]). Let G be a compact group. Then the following conditions
are equivalent:
(1) There is a compact connected space X such that G ∼
= H(X).
(2) There is a compact space X such that G ∼
= H(X).
(3) G is profinite.
Let us gather some basic known facts and open problems concerning the entropy
of the functional envelopes:
1. The entropy of (S(X), FT ) is always not smaller than that of (X, T ). To see
this, it suffices to consider the compact FT -invariant subsemigroup of constant
self-maps, which is conjugate to (X, T ).
2. On the other hand, it was not known whether the same inequality holds for
(H(X), FT ).
3. Our with Semikina conjecture says that the functional envelope (S(X), FT )
has entropy either zero or infinity. In [35] we proved that the conjecture
holds true for all Peano continua and for all compact spaces with continuum
many connected components. Otherwise the conjecture was open. The same
problem can be posed for the functional envelope (H(X), FT ) in case T is a
homeomorphism.
Recently, Downarowicz, Snoha and Tywoniuk [13] gave a positive answer to the
question (2) for homogeneous spaces and extended the knowledge on the question
(3). Namely, they proved the following.
Theorem 4.4 ([13]). Let T : X → X be a homeomorphism of a homogeneous
compact space. Then the entropy of the functional envelope (H(X), FT ) is at least
as large as that of (X, T ).
Theorem 4.5 ([13]). Let T : X → X be a homeomorphism of a compact zero-
dimensional space. Then the entropies of (S(X), FT ) and (H(X), FT ) are either
both zero or both infinite. They are equal to zero if and only if T is equicontinuous.
They also showed that there exists a positive (even infinite) entropy homeomor-
phism T of a compact space X such that (H(X), FT ) has entropy zero (see [13] and
below). This proves that the general question (2) (the inequality between topo-
logical entopies of (X, T ) and its functional envelope (H(X), FT )) has a negative
answer. This was solved by using a new class of spaces, which they have called
1310 SERGIĬ KOLYADA

Slovak spaces, defined by the combination of two properties: the existence of


a (minimal) homeomorphism, say T , and nonexistence of homeomorphisms other
than the powers of T . It is obvious that for such spaces the functional envelope
(H(X), FT ) always has entropy zero.
The readers interested in functional envelopes may consult, besides already men-
tioned papers [6], [35] and [13], also for instance [41] and [9].

4.2. Slovak spaces as uniquely minimal spaces. Recall that a map T : X → X


is called minimal if the orbit {x, T x, T 2 x, ..., T n x, ...} of any point x ∈ X is dense
in X.
For a compact metric space X there are two possibilities:
1. X does not admit any minimal homeomorphism (e.g., interval, disk, any space
with the fixed point property, ...);
2. X admits a minimal homeomorphism (Cantor set, circle, torus, ... ). In this
case, if X is infinite then in known examples usually X admits uncountably
many homeomorphisms and even uncountably many of them are minimal.
Is there a third possibility? That is, does there exist an infinite compact metric
space X such that it admits, but only “a few”, minimal homeomorphisms?
An infinite compact metric space X is called Slovak if it is uniquely minimal in the
following sense: X admits a minimal homeomorphism T and H(X) = {T n : n ∈ Z}.
The assumption that X is infinite eliminates two trivial examples: the one-point
space and the two-point space. If X is Slovak then card X = c, in fact every Slovak
space is a nondegenerate continuum, and all iterates T n , n ∈ Z are different, i.e.
H(X) ≈ Z. Moreover, all iterates T n , except identity, are minimal.
Theorem 4.6 ([13])). There exist Slovak spaces in the class of metric continua.
Moreover, the topological entropies of generating homeomorphisms T exhaust the
interval [0, ∞].
The idea of the construction of a Slovak space in [13] is as follows.
Step 1. Let h : C → C be a minimal homeomorphism on the Cantor set C (with
arbitrary entropy).
Step 2. Define the generalized solenoid induced by (C, h), i.e.
SOL := C × [0, 1]/∼ ,
where (x, 1) ∼ (h(x), 0).
Recall that solenoids, which are special cases of generalized solenoids, are among
the simplest examples of indecomposable homogeneous continua. They are neither
arcwise connected nor locally connected. In Figure 17 there is a solid torus S1 × D
wrapped twice around inside another solid torus in R3 . Each solenoid may be
constructed as the intersection of a nested system of embedded solid tori in R3 .

Figure 17. The first 5 steps in the construction of the solenoid


called the Smale-Williams attractor.
DYNAMICAL TOPOLOGY 1311

Take the suspension flow over h (with ceiling function ≡ 1), Φt , t ∈ R defined on
SOL by the formula
Φt (y, s) := (h[t+s] y, {t + s}),
where [ · ] and {·} denote the integer and fractional parts of a real number, respec-
tively.
Step 3. It is well known that there exists t0 ∈ R such that the map
T := time t0 -map of the flow Φ
is a minimal homeomorphism on SOL (see e.g. [16]).
Step 4. This is the main technical step. The Slovak space constructed in [13] will
be a subset of the cylinder SOL × [0, 1]. The main element of this construction is a
topologist’s sine curve (a subset of the plane that is the union of the graph of the
function f (x) = sin(1/x), 0 < x ≤ 1 with the segment −1 ≤ y ≤ 1 of the y-axis).
1. The continuum SOL has uncountably many composants (orbits of the flow
Φ); choose a composant γ and a point x0 ∈ γ. (Recall that the composant of
a point p in a continuum A is the union of all proper subcontinua of A that
contain p. If a continuum is indecomposable, then its composants are pairwise
disjoint. The composants of a continuum are dense in that continuum.)
2. On a closed arc around x0 (minus the point x0 itself) and lying in γ, we define
a function which looks like a one-sided topologist’s sine curve (values in [0, 1],
wiggles of height 1 in any left neighbourhood of x0 , constant value 0 to the
right of x0 ). It is continuous and not defined at x0 .
3. Extend itPto a continuous function f : SOL \ {x0 } → [0, 1].
n
P F = n∈Z an f ◦ T , where the coefficients an are all strictly positive with
4. Let
n∈Z an = 1 and satisfying some technical assumptions (F is defined on SOL
minus the T -orbit of x0 ).
5. Then one can show that both the mapping (x, F (x)) 7→ (T x, F (T x)) and its
inverse are uniformly continuous homeomorphisms of the graph of F . There-
fore, the map (x, F (x)) 7→ (T x, F (T x)) extends to a homeomorphism T̄ of F̄
(=the closure of the graph of F ).
6. F̄ ⊆ SOL × [0, 1] is our S lovak space. The composant γ̄ of F̄ “above” γ has
basicly the shape as in Figure 18. The other composants of F̄ are continuous

Figure 18. Composant γ̄ of the Slovak space.

bijective images of the real line.


4.3. Some related results. The question when a system (X, f ) can be embedded
as a subsystem of some (Y, g) so that X = ωg (y) for some y ∈ Y was answered by
Dowker and Friedlander [11] for homeomorphisms and by Sharkovsky [47] in general
(see also [49]). They showed that (X, f ) can be embedded as the omega limit set
in some larger system if and only if it is f -connected. Recall that a system (X, f )
is f -connected if, for any proper, nonempty, closed subset U ( X, the intersection
f (U ) ∩ X \ U is nonempty.
1312 SERGIĬ KOLYADA

In [4] Akin and Rautio considered the related problem of when a space X admits
a homeomorphism f so that (X, f ) is the omega limit set in a larger system. As we
will see, their results are somewhat different from the map case. They reinterpreted
the problem by using some other notion. Given ε ≥ 0, a finite or infinite sequence
{xn ∈ X} with at least two terms is an ε-chain for (X, f ) if d(f xk , xk+1 ) ≤ ε for all
terms xk of the sequence (except the last one). The system (X, f ) is called chain
transitive when every pair of points of X can be connected by some finite ε-chain
for every positive ε. A subset A ⊂ X is called a chain transitive subset when it is
closed and f -invariant (i.e., f (A) = A) and the subsystem (A, f ) is chain transitive.
It is well known that chain transitivity and f -connectedness are equivalent con-
cepts (see [4] for details). Akin and Rautio generalized several known results (in
particular, regarding rigid spaces). They proved the existence of the following com-
pact metric spaces:
1. Suppose that G is a finitely generated (and hence countable) group. Then
there exists a space X such that the homeomorphism group H(X) is isomor-
phic to G and every f ∈ H(X) is chain transitive.
2. There exists a space X such that the homeomorphism group H(X) contains a
nontrivial path-connected subgroup and every f ∈ H(X) is chain transitive.
3. There exists a space X such that the homeomorphism group H(X) is isomor-
phic to the homeomorphism group of the Cantor set and every f ∈ H(X) is
chain transitive.
Akin and Rautio in [4] also extended the result of Downarowicz, Snoha and Ty-
woniuk on the existence of Slovak spaces by showing the existence of spaces they
call Slovakian spaces. These are infinite compact metric spaces with H(X) non-
trivial and such that every homeomorphism other than the identity is topologically
transitive. The only Slovakian spaces known to them are variations on the original
construction of [13]; all of these have homeomorphism group isomorphic to Z.

5. Appendix.
5.1. Basic concepts in topological dynamics. Given a dynamical system (X, T ),
recall that x ∈ X is a fixed point if T (x) = x, and an F-transitive point of (X, T )
[39] if NT (x, U ) ∈ F for any opene subset U of X. It is a trivial observation that
if a family F admits an F-transitive dynamical system (X, T ) without isolated
points, then F is free. Since k(kF) = F, it is easy to see that x ∈ X is an F-
transitive point of (X, T ) if and only if ωkF (x) = X. Denote by TranF (X, T ) the
set of all F-transitive points of (X, T ). The system (X, T ) is F-point transitive if
TranF (X, T ) 6= ∅, and is F-transitive if NT (U, V ) ∈ F for any opene subsets U, V of
X. Write Tran(X, T ) = TranP+ (X, T ) for short, and we also call the point x transi-
tive if x ∈ Tran(X, T ) or, equivalently, if its orbit orbT (x) = {T n x : x = 0, 1, 2, . . . }
is dense in X. The system (X, T ) is transitive if and only if Tran(X, T ) is a dense
Gδ subset of X.
A subset A of X is T -invariant if T (A) = A, and positively T -invariant if
T (A) ⊂ A. If A is a closed, nonempty, T -invariant subset then (A, T |A ) is called
the associated subsystem. A minimal subset of X is a closed, nonempty, T -invariant
subset such that the associated subsystem is minimal. Clearly, (X, T ) is minimal if
and only if Tran(X, T ) = X, if and only if it admits no proper, closed, nonempty,
positively T -invariant subset. A point x ∈ X is called minimal if it lies in some
minimal subset. Zorn’s Lemma implies that every closed, nonempty, positively
T -invariant set contains a minimal set.
DYNAMICAL TOPOLOGY 1313

A pair of points x, y ∈ X is called proximal if lim inf n→∞ d(T n x, T n y) = 0. In


this case each of the points from the pair is said to be proximal to the other one.
Denote by ProxT (X) the set of all proximal pairs of points. For each x ∈ X, denote
by ProxT (x) the proximal cell of x, i.e. the set of all points which are proximal to
x. Recall that a dynamical system (X, T ) is called proximal if ProxT (X) = X × X.
The system (X, T ) is proximal if and only if (X, T ) has the unique fixed point,
which is the only minimal point of (X, T ) (see e.g. [3]).
The opposite to the notion of sensitivity is the concept of equicontinuity. Recall
that x ∈ X is an equicontinuity point of (X, T ) if for every ε > 0 there exists a
δ > 0 such that d(x, x0 ) < δ implies d(T n x, T n x0 ) < ε for any n ∈ Z+ . Denote
by Eq(X, T ) the set of all equicontinuity points of (X, T ). The system (X, T ) is
called equicontinuous if Eq(X, T ) = X. Each dynamical system admits a maximal
equicontinuous factor. Recall that by a factor map π : (X, T ) → (Y, S) between
dynamical systems (X, T ) and (Y, S) we mean a continuous surjection π : X → Y
with π ◦ T = S ◦ π. In such a case we say that (X, T ) is an extension of (Y, S) and
(Y, S) is a factor of (X, T ).

5.2. Furstenberg families. We recall basic concepts related to Furstenberg fami-


lies, see [1].
Let F ∈ P. Recall that a subset F is thick if it contains arbitrarily long runs
of positive integers. Denote by Fthick the set of all thick subsets of Z+ , and define
Fsyn = kFthick . Each element of Fsyn is said to be syndetic. Equivalently, F is
syndetic if and only if there is N ∈ N such that {i, i + 1, . . . , i + N } ∩ F 6= ∅ for
every i ∈ Z+ . We say that F is thickly syndetic if for every N ∈ N the positions
where length N runs begin form a syndetic set. Denote by Fcof the set of all cofinite
subsets of Z+ . Note that, by a classical result of Birkhoff, a point x ∈ X is minimal
if and only if nT (x, U ) = {n ∈ Z+ : T n x ∈ U } is syndetic for any neighborhood U
of x. Hence, for any minimal system (X, T ), the set NT (U, V ) is syndetic for any
opene subsets U, V of X.
Recall that a family F is proper if it is a proper subset of P, that is, Z+ ∈ F
and ∅ ∈ / F. By a filter F we mean a proper family closed under intersection, that
is, F1 , F2 ∈ F implies F1 ∩ F2 ∈ F. For families F1 and F2 , we define the family
F1 · F2 := {F1 ∩ F2 : F1 ∈ F1 , F2 ∈ F2 } and call it the interaction of F1 and F2 . We
have F1 ∪ F2 ⊂ F1 · F2 . It is easy to check that F is a filter if and only if F = F · F,
and F1 · F2 is proper if and only if F2 ⊂ kF1 .
For each i ∈ Z+ , we define g i : Z+ → Z+ , j 7→ i + j. Let F be a family.
Recall that F is positively invariant if for every i ∈ Z+ , F ∈ F implies g i (F ) ∈
F; negatively invariant if for every i ∈ Z+ , F ∈ F implies g −i (F ) ∈ F, where
g −i (F ) = (g i )−1 (F ) = {j − i : j ∈ F, j ≥ i}; and translation invariant if it is both
positively and negatively invariant, equivalently, for every i ∈ Z+ , F ∈ F if and
only if g −i (F ) ∈ F.
As g −i (g i A) = A and g i (g −i A) ⊂ A for any i ∈ Z+ , it is easy to obtain that
the family F is positively or negatively or translation invariant if and only if kF
is negatively or positively or translation, respectively, invariant (see for example [1,
Proposition 2.5.b]). Then we have the following.

Proposition 5.1 ([29]). Let x ∈ X. Then T ωF (x) ⊂ ωF (T x). Additionally, if F


is negatively (positively, translation) invariant then ωF (T x) ⊂ (⊃, =, respectively)
ωF (x).
1314 SERGIĬ KOLYADA

Proposition 5.2 ([29]). Let (X, T ) be a dynamical system and let F be a family.
(i) If F is free, then ωF (x) ⊂ ωT (x) for any x ∈ X. Moreover, if (X, T ) has a
nonrecurrent point, then the converse implication is true.
(ii) If F is free and has FIP then it has SFIP.
5.3. Almost one-to-one maps. Let ϕ : X → Y be a continuous surjective map
from a compact metric space X onto a compact Hausdorff space Y . Recall that ϕ
is almost open if ϕ(U ) has a nonempty interior in Y for any opene U ⊂ X. Note
that each factor map between minimal systems is almost open [5, Theorem 1.15].
In particular, for a minimal system (X, T ) the map T : X → X is almost open [38].
Denote by Y0 the set of all points y ∈ Y whose preimage is a singleton. Then Y0
is a Gδ subset of Y , because
\ 1

−1 −1
Y0 = {y ∈ Y : ϕ (y) is a singleton} = y ∈ Y : diam(ϕ (y)) <
n
n∈N
−1
and the map y 7→ diam(ϕ (y)) is upper semi-continuous. Here, we denote by
diam(A) the diameter of a subset A ⊂ X. Recall that the function f : Y → R+ is
upper semi-continuous if lim sup f (y) ≤ f (y0 ) for each y0 ∈ Y .
y→y0
Denote by X0 the set of all points x ∈ X such that the preimage of ϕ(x) is a
singleton. Then X0 = ϕ−1 (Y0 ) is a Gδ subset of X.
We call ϕ weakly almost one-to-one if Y0 is dense in Y , and almost one-to-one 3
if X0 is dense in X. It is not hard to show that: if ϕ is weakly almost one-to-one,
then for any δ > 0 and any opene subset U of Y there exists opene V ⊂ U with
diam(ϕ−1 V ) < δ; and if ϕ is almost one-to-one, then for any opene subset U ∗ of
X there exists an opene subset V ∗ of Y with ϕ−1 V ∗ ⊂ U ∗ . Clearly almost one-
to-one is stronger than weakly almost one-to-one. For instance, define T x = 2x
for x ∈ [0, 12 ] and T x = 1 for x ∈ [ 12 , 1]. Then T : [0, 1] → [0, 1] is weakly almost
one-to-one but not almost one-to-one.
For each minimal system (X, T ), the map T : X → X is weakly almost one-to-one
[38, Theorem 2.7], and in fact almost one-to-one [30, Proposition 2.3]. The following
result characterizes the relationship between weakly almost one-to-one and almost
one-to-one, which extends [30, Proposition 2.3].
Proposition 5.3 ([29]). Let ϕ : X → Y be a continuous surjective map from a
compact metric space X onto a compact Hausdorff space Y . Then ϕ is almost
one-to-one if and only if it is almost open and weakly almost one-to-one.
As a direct corollary, we have:
Corollary 5.4 ([29]). Let ϕ : X → Y and π : Y → Z be continuous surjective
maps between compact metric spaces. Then the composition map π ◦ ϕ : X → Z is
almost one-to-one if and only if both ϕ and π are almost one-to-one.
Let π : (X, T ) → (Y, S) be a factor map between dynamical systems. If the
map π : X → Y is almost one-to-one (weakly almost one-to-one, respectively),
then we also say that (X, T ) is an almost one-to-one extension (a weakly almost
one-to-one extension, respectively) of (Y, S). The main result of [30] states that a
minimal system is either multi-sensitive or a weakly almost one-to-one extension of
its maximal equicontinuous factor. This is an analog of the well-known Auslander-
Yorke dichotomy theorem: a minimal system is either sensitive or equicontinuous.
3 Here we use the concept of almost one-to-one following [2]. The concept of almost one-to-one

used in [12, 30, 38] is in fact our weakly almost one-to-one.


DYNAMICAL TOPOLOGY 1315

5.4. Topological entropy: Bowen and Dinaburg definition. Topological en-


tropy measures the evolution of distinct (distinguishable) orbits over time, thereby
providing an idea of how complex the orbit structure of a system is. Recall the
Bowen-Dinaburg definition of topological entropy (see [50]). Let (Z, %) be a metric
space and f : Z → Z be uniformly continuous. For any integer n ≥ 1 the function
%n (x, y) := max %(f j x, f j y)
0≤j≤n−1

defines a metric on Z equivalent with %.


Fix an integer n ≥ 0 and ε > 0 and let K be a compact set in Z. A subset E ⊂ K
is called (n, f, ε)-separated , if for any two distinct points x, y ∈ E, %n (x, y) > ε. We
say that a subset F ⊂ Z (n, f, ε)-spans the set K, if for every point x ∈ K there
exists a point y ∈ F such that %n (x, y) ≤ ε. Note that since K is compact, E is
finite and F may be infinite, but a finite subset that still spans K exists.
Denote by sep(n, f, ε; K) the maximal cardinality of an (n, f, ε)-separated set in
K, and by span(n, f, ε; K) the minimal cardinality of a set which (n, f, ε)-spans K.
For every ε > 0 and n ≥ 0 it holds
span(n, f, ε; K) ≤ sep(n, f, ε; K) ≤ span(n, f, ε/2; K) .
The topological entropy h(f, K) on a compact set K ⊂ Z is defined by
1 1
h(f, K) := lim lim sup log sep(n, f, ε; K) = lim lim sup log span(n, f, ε; K).
ε→0 n→∞ n ε→0 n→∞ n

Then the topological entropy h(f ) of a map f : Z → Z is defined by


h % (f ) = h(f ) := sup{h(f, K) : K ⊂ Z and K is compact} .
K
If span (n, f, ε; K) denotes the minimal cardinality of subsets of the com-
pact set K which (n, f, ε)-span K then spanK (n, f, ε; K) ≤ sep(n, f, ε; K) ≤
spanK (n, f, ε/2; K) and h(f, K) = limε→0 lim supn→∞ n1 log spanK (n, f, ε; K) (see
[50]).
For uniformly equivalent metrics %1 and %2 we have h%1 (X, f ) = h%2 (X, f ), and
if %1 is stronger than %2 , then h%1 (X, f ) ≥ h%2 (X, f ).
If X is compact, then the map Ff : ϕ 7→ f ◦ ϕ is uniformly continuous on both
the spaces SU (X) and SH (X), and so we can study its topological entropy. Recall
that dU (ϕ1 , ϕ2 ) ≥ dH (ϕ1 , ϕ2 ) for all ϕ1 , ϕ2 ∈ S(X) and that these two metrics are
equivalent on S(X), hence uniformly equivalent on compact subsets of S(X).
5.5. Basic properties of topological entropy. The following are some of the
basic properties of topological entropy for compact dynamical systems (see e.g [50]
and, for 5., [36]).
1. h(T ) ≥ h(S) if (Y, S) is a factor of (X, T ) , i.e., ϕ◦T = S ◦ϕ, where ϕ : X → Y
is a continuous surjection.
2. h(T ) ≥ h(T |Y ) if Y is a nonempty closed invariant subset of X.
3. h(T n ) = n · h(T ) for n ≥ 0.
4. h(T × S) = h(T ) + h(S).
5. h(T ◦ S) = h(S ◦ T ).

Acknowledgments. The author thanks very much Danylo Khilko and L’ubomı́r
Snoha for useful remarks and discussions. The author was supported by Max-
Planck-Institut für Mathematik (MPIM, Bonn) and Institut des Hautes Études
Scientifiques (IHES, Bures-sur-Yvette); he acknowledges the hospitality of the In-
stitutes.
1316 SERGIĬ KOLYADA

REFERENCES

[1] E. Akin, Recurrence in Topological Dynamics. Furstenberg families and Ellis actions, The
University Series in Mathematics, Plenum Press, New York, 1997.
[2] E. Akin and E. Glasner, Residual properties and almost equicontinuity, J. Anal. Math., 84
(2001), 243–286.
[3] E. Akin and S. Kolyada, Li-Yorke sensitivity, Nonlinearity, 16 (2003), 1421–1433.
[4] E. Akin and J. Rautio, Chain transitive homeomorphisms on a space: All or none, Pacific J.
Math., 291 (2017), 1–49.
[5] J. Auslander, Minimal Flows and Their Extensions, North-Holland Mathematics Studies, vol.
153, North-Holland Publishing Co., Amsterdam, 1988, Notas de Matemática [Mathematical
Notes], 122.
[6] J. Auslander, S. Kolyada and L’. Snoha, Functional envelope of a dynamical system, Nonlin-
earity, 20 (2007), 2245–2269.
[7] J. Auslander and J. A. Yorke, Interval maps, factors of maps, and chaos, Tôhoku Math. J.,
(2) 32 (1980), 177–188.
[8] H. Cook, Continua which admit only the identity mapping onto non-degenerate subcontinua,
Fund. Math., 60(1967), 241–249.
[9] T. Das, E. Shah and L’. Snoha, (Non-)expansivity in functional envelopes, J. Math. Anal.
Appl., 410 (2014), 1043–1048.
[10] T. Dobrowolski, Examples of topological groups homeomorphic to l2f , Proc. Amer. Math.
Soc., 98 (1986), 303–311.
[11] Y. N. Dowker and F. G. Friedlander, On limit sets in dynamical systems, Proc. London Math.
Soc., (3) 4 (1954), 168–176.
[12] T. Downarowicz, Survey of odometers and Toeplitz flows, Algebraic and Topological Dynam-
ics, Contemp. Math., vol. 385, Amer. Math. Soc., Providence, RI, 2005, 7–37.
[13] T. Downarowicz, L’. Snoha and D. Tywoniuk, Minimal spaces with cyclic group of homeo-
morphisms, J. Dynam. Differential Equations, 29 (2017), 243–257.
[14] F. T. Farrell and A. Gogolev, The space of Anosov diffeomorphisms, J. Lond. Math. Soc., (2)
89 (2014), 383–396.
[15] A. Fathi, Structure of the group of homeomorphisms preserving a good measure on a compact
manifold, Ann. Sci. École Norm. Sup., 13 (1980), 45–93.
[16] B. R. Fayad, Topologically mixing and minimal but not ergodic, analytic transformation on
T5 , Bol. Soc. Brasil. Mat. (N.S.), 31 (2000), 277–285.
[17] H. Furstenberg, Disjointness in ergodic theory, minimal sets, and a problem in Diophantine
approximation, Math. Systems Theory, 1 (1967), 1–49.
[18] H. Furstenberg, Recurrence in Ergodic Theory and Combinatorial Number Theory, Princeton
University Press, Princeton, N.J., 1981, M. B. Porter Lectures.
[19] H. Furstenberg and B. Weiss, Topological dynamics and combinatorial number theory, J.
Analyse Math., 34 (1978), 61–85 (1979).
[20] P. Gartside and A. Glyn, Autohomeomorphism groups, Topology Appl., 129 (2003), 103–110.
[21] E. Glasner and B. Weiss, Sensitive dependence on initial conditions, Nonlinearity, 6 (1993),
1067–1075.
[22] J. de Groot and R. J. Wille, Rigid continua and topological group-pictures, Arch. Math., 9
(1958), 441–446.
[23] J. de Groot, Groups represented by homeomorphism groups, Math. Ann., 138 (1959), 80–102.
[24] J. Guckenheimer, Sensitive dependence to initial conditions for one-dimensional maps, Comm.
Math. Phys., 70 (1979), 133–160.
[25] S. Harada, Remarks on the topological group of measure preserving transformations, Proc.
Japan Acad., 27 (1951), 523–526.
[26] K. H. Hofmann and S. A. Morris, Compact homeomorphism groups are profinite, Topology
Appl., 159 (2012), 2453–2462.
[27] K. H. Hofmann and S. A. Morris, Representing a profinite group as the homeomorphism
group of a continuum, preprint, arXiv:1108.3876.
[28] W. Huang, D. Khilko, S. Kolyada and G. Zhang, Dynamical compactness and sensitivity, J.
Differential Equations, 260 (2016), 6800–6827.
[29] W. Huang, D. Khilko, S. Kolyada, A. Peris and G. Zhang, Finite intersection property and
dynamical compactness, J. Dynam. Differential Equations, 30 (2018), 1221–1245.
DYNAMICAL TOPOLOGY 1317

[30] W. Huang, S. Kolyada and G. Zhang, Analogues of Auslander-Yorke theorems for multi-
sensitivity, Ergodic Theory Dynam. Systems, 38 (2018), 651–665.
[31] M. Keane, Contractibility of the automorphism group of a nonatomic measure space, Proc.
Amer. Math. Soc., 26 (1970), 420–422.
[32] S. Kolyada, M. Misiurewicz and L’. Snoha, Spaces of transitive interval maps, Ergodic Theory
Dynam. Systems, 35 (2015), 2151–2170.
[33] S. Kolyada, M. Misiurewicz and L’. Snoha, Loops of transitive interval maps, Dynamics and
numbers, Contemp. Math., Amer. Math. Soc., Providence, RI, 669 (2016), 137–154.
[34] S. Kolyada and O. Rybak, On the Lyapunov numbers, Colloq. Math., 131 (2013), 209–218.
[35] S. Kolyada and J. Semikina, On topological entropy: When positivity implies +infinity, Proc.
Amer. Math. Soc., 143 (2015), 1545–1558.
[36] S. Kolyada and L’. Snoha, Topological entropy of nonautonomous dynamical systems, Random
Comput. Dynam., 4 (1996), 205–233.
[37] S. Kolyada and L’. Snoha, Some aspects of topological transitivity – a survey, Iteration Theory
(ECIT 94) (Opava), Grazer Math. Ber., 334 (1997), 3–35.
[38] S. Kolyada, L’. Snoha and S. Trofimchuk, Noninvertible minimal maps, Fund. Math., 168
(2001), 141–163.
[39] J. Li, Transitive points via Furstenberg family, Topology Appl., 158 (2011), 2221–2231.
[40] J. Li and X. Ye, Recent development of chaos theory in topological dynamics, Acta Math.
Sin. (Engl. Ser.), 32 (2016), 83–114.
[41] M. Matviichuk, On the dynamics of subcontinua of a tree, J. Difference Equ. Appl., 19 (2013),
223–233.
[42] T. K. S. Moothathu, Stronger forms of sensitivity for dynamical systems, Nonlinearity, 20
(2007), 2115–2126.
[43] N. T. Nhu, The group of measure preserving transformations of the unit interval is an absolute
retract, Proc. Amer. Math. Soc., 110 (1990), 515–522.
[44] K. E. Petersen, Disjointness and weak mixing of minimal sets, Proc. Amer. Math. Soc., 24
(1970), 278–280.
[45] P. Raith, Topological transitivity for expanding piecewise monotonic maps on the interval,
Aequationes Math., 57 (1999), 303–311.
[46] D. Ruelle, Dynamical systems with turbulent behavior, Mathematical Problems in Theoretical
Physics (Proc. Internat. Conf., Univ. Rome, Rome, 1977), Lecture Notes in Phys., vol. 80,
Springer, Berlin-New York, 1978, 341–360.
[47] A. N. Šarkovskiı̆, On attracting and attracted sets, Dokl. Akad. Nauk SSSR, 160 (1965),
1036–1038. In Russian; translated in Soviet Math. Dokl., 6 (1965), 268–270.
[48] A. N. Šarkovskiı̆, Continuous mapping on the limit points of an iteration sequence, Ukrain.
Mat. Ž., 18 (1966), 127–130.
[49] A. N. Šarkovskiı̆, S. F. Kolyada, A. G. Sivak and V. V. Fedorenko, Dynamics of One-
Dimensional Maps, Kiev, 1989.
[50] P. Walters, An Introduction to Ergodic Theory, Springer-Verlag, New York-Berlin, 1982.
[51] T. Yagasaki, Weak extension theorem for measure-preserving homeomorphisms of noncompact
manifolds, J. Math. Soc. Japan, 61 (2009), 687–721.

Received January 2018; revised September 2018.

You might also like