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

Explicit Dynamical Systems on the Sierpiński curve

Worapan Homsomboon Oregon State University [email protected], [email protected]
(Date: Last update on 7 September 2022)
Abstract.

We apply Boroński and Oprocha’s inverse limit construction of dynamical systems on the Sierpiński carpet by using the initial systems of nn-Chamanara surfaces and their nn-baker transformations, n2n\geq 2. We show that all positive real numbers are realized as metric entropy values of dynamical systems on the carpet. We also produce a simplification of Boroński and Oprocha’s proof showing that dynamical systems on the carpet do not have the Bowen specification property.

1. Introduction

A Sierpiński carpet is a plane fractal created by Wacław Sierpiński in 1916. It is a continuum (i.e. a nonempty, compact, connected, metrizable topological space). Furthermore, it is a universal plane curve (i.e. it contains a homeomorphic copy of any subspace of 2{\mathbb{R}}^{2} with topological dimension 11; see say [11]). Its fractal structure also allows both theoretical exploration of the structures such as Dirichlet forms (see [2]), and practical application in the field of communication (WiFi antenna, see for example [22]).

Nevertheless, the carpet does not only attract interests from topologists, but also from dynamicists. Various studies on homeomorphisms on the carpet have been done over the years. In 1991 (see [17]), Kato showed that the Sierpiński curve does not admit an expansive homeomorphism. In addition, Aarts and Oversteegen were able to show that it is possible for carpet homeomorphisms to be transitive (see [1]). A construction of a homeomorphism given by Biś, Nakayama and Walczak in [3] verified that the carpet admits a homeomorphism with positive topological entropy.

Recently, in 2018, Jan P. Boroński and Piotr Oprocha introduced in their paper [5] a new way to construct a Sierpiński curve. In particular, they built an inverse limit system induced from Arnold’s cat map on the 2-torus 𝕋2{\mathbb{T}}^{2}. The inverse limit was shown to be homeomorphic to the Sierpiński carpet. In addition, the construction produced a function on the inverse limit which was proved to be a homeomorphism. One can say that the strength of Boroński and Oprocha’s inverse limit construction is that it produces both the Sierpiński carpet and its homeomorphism. Another important point is that their construction is based on Whyburn’s topological characterization of the Sierpiński carpet (see [25]). Precisely, one can obtain a Sierpiński curve by deleting an infinite sequence of open discs (Di)i1(D_{i})_{i\geq 1} from the 22-sphere SS2\SS^{2} where

  • Di¯Dj¯=\overline{D_{i}}\cap\overline{D_{j}}=\emptyset for all iji\neq j (i.e. their closures are pairwise disjoint),

  • diam(Dj)0\text{diam}(D_{j})\rightarrow 0 as jj\rightarrow\infty, and

  • i=1Dj¯\cup_{i=1}^{\infty}\overline{D_{j}} is dense on SS2\SS^{2}.

This characterization allows a possibility to extend the inverse limit construction: the initial system Arnold’s cat map on 𝕋2{\mathbb{T}}^{2} can possibly be replaced by a dynamical system (X,T)(X,T) which is a branched covering system of the 22-sphere SS2\SS^{2}.

Arnold’s cat map is an example of a homeomorphism belonging to the class of hyperbolic toral automorphisms. An analogue of hyperbolic toral automorphisms on the surface of genus one 𝕋2{\mathbb{T}}^{2} is the (linear) pseudo-Anosov diffeomorphisms on hyperelliptic translation surfaces of genus g2g\geq 2. These two classes of dynamical systems are branched coverings of SS2\SS^{2}, and in fact, they can replace Arnold’s cat map in the construction. Another family of translation surfaces nn-Chamanara surface and its homeomorphism nn-baker map turns out to be a potential candidate worth investigation.

An α\alpha-Chamanara surface CαC_{\alpha}, α(1,)\alpha\in(1,\infty), is an infinite genus translation surface introduced by Reza Chamanara in [8]. Roughly, one can obtain the surface by assigning side identifications to [0,1]2[0,1]^{2} as shown in Figure 1 below.

Refer to caption
Figure 1. Side identifications impose on the square [0,1]2[0,1]^{2}. The quotient metric space with the single singular point (0,0)(0,0) is the α\alpha-Chamanara surface, α(1,)\alpha\in(1,\infty).

In the special case where α=n\alpha=n\in{\mathbb{N}} with n2n\geq 2, the surface nn-Chamanara CnC_{n} admits a homeomorphism, the nn-baker map BnB_{n}. This is a generalization of the well-known (2-)baker transformation B2B_{2} defined on [0,1)2[0,1)^{2} as

B2(x,y)={(2x,y/2),if 0x<1/2,0y<1(2x1,12(y+1)),if 1/2x<1,0y<1.B_{2}(x,y)=\begin{cases}(2x,y/2),&\mbox{if}\ 0\leq x<1/2,0\leq y<1\\ (2x-1,\frac{1}{2}(y+1)),&\mbox{if}\ 1/2\leq x<1,0\leq y<1.\end{cases}

A rotation by π\pi radians RR centered at (0.5,0.5)(0.5,0.5) induces an equivalence relation R\sim_{R} on CnC_{n}. For the case n=2n=2, Chamanara, Gardiner and Lakic verified in [9] that the quotient space C2/RC_{2}/\sim_{R} is a 22-sphere SS2\SS^{2} (they point to an argument of de Carvalho and Hall in [10] where the result is an application of Moore’s theorem) and there exists a homeomorphism on C2/RC_{2}/\sim_{R} which commutes with B2B_{2} via the projection P2:C2C2/R.P_{2}:C_{2}\rightarrow C_{2}/\sim_{R}. We reprove this fact for (C2,B2)(C_{2},B_{2}) and also extend the result to the pair (Cn,Bn)(C_{n},B_{n}) for n3n\geq 3. Note that our approach is distinct from the one used in [9].

Theorem A.

For each n2n\geq 2, the dynamical system (Cn,Bn)(C_{n},B_{n}) admits an equivalence relation R\sim_{R} given by the rotation by π\pi radians. The quotient space Qn:=Cn/RQ_{n}:=C_{n}/\sim_{R} is homeomorphic to SS2\SS^{2}. The nn-baker map descends down to QnQ_{n} and induces a homeomorphism TnT_{n} on QnQ_{n} such that PnTn=BnPnP_{n}\circ T_{n}=B_{n}\circ P_{n} where Pn:CnQnP_{n}:C_{n}\rightarrow Q_{n} is the natural projection.

As one sees in the discussion of Boroński and Oprocha’s construction given in Section 22, there are additional required conditions on the initial system (X,T)(X,T) besides factoring to give a system on SS2\SS^{2}:

  • The set of periodic points of the map TT, Per(T),\text{Per}(T), is dense in XX.

  • When the equivalence relation inducing the quotient sphere is a local isometry, the homeomorphism TT is an affine diffeomorphism.

It is a well-known result that any hyperbolic toral automorphism FAF_{A} on 𝕋2{\mathbb{T}}^{2} and any pseudo-Anosov homeomorphism TλT_{\lambda} on a hyperelliptic translation surface XgX_{g} satisfy these conditions. For each n2n\geq 2, the pair (Cn,Bn)(C_{n},B_{n}) is a factor of the system of bi-infinite sequence on nn symbols and its left-shift map (Σn,σn)(\Sigma_{n},\sigma_{n}). This gives that Per(Bn)¯=Cn\overline{\text{Per}(B_{n})}=C_{n} which allows one to get the following conclusion.

Theorem B.

Let Γ:={(𝕋2,FA)FA:𝕋2𝕋2is a hyperbolic toral automorphism},\Gamma:=\{({\mathbb{T}}^{2},F_{A})\mid F_{A}:{\mathbb{T}}^{2}\rightarrow{\mathbb{T}}^{2}\ \text{is a hyperbolic toral automorphism}\}, Ω:={(Xg,Tλ)Tλ:XgXgis a pseudo-Anosov homeomorphism},Λ:={(Cn,Bn)n2}.\Omega:=\{(X_{g},T_{\lambda})\mid T_{\lambda}:X_{g}\rightarrow X_{g}\ \text{is a pseudo-Anosov homeomorphism}\},\Lambda:=\{(C_{n},B_{n})\mid n\geq 2\}. Any (Y,S)ΓΩΛ(Y,S)\in\Gamma\cup\Omega\cup\Lambda is a valid initial system to build an inverse limit system of the Sierpiński carpet and its homeomorphism.

Furthermore, each of this topological dynamical system (X,T)(X,T) admits invariant measures for which we can calculate the metric entropy.

Theorem C.

The entropy values of the dynamical systems on the carpet built in Theorem B are

log(|λ|),log(λ)ori=0n1pilog(pi)\log(|\lambda|),\log(\lambda^{\prime})\ \text{or}\ -\sum_{i=0}^{n-1}p_{i}\log(p_{i})

where λ\lambda is the leading eigenvalue of a hyperbolic toral automorphism, λ\lambda^{\prime} is the expansion factor of a pseudo-Anosov diffeomorphism and P=(p0,p1,,pn1)P=(p_{0},p_{1},\ldots,p_{n-1}) is a probability vector (i.e. 0pi<10\leq p_{i}<1 for all ii, and i=0n1pi=1\sum_{i=0}^{n-1}p_{i}=1). In particular, all positive real numbers are realized as entropy values of dynamical systems on the carpet.

Another advantage of Boroński and Oprocha’s construction is that each dynamical system on the carpet inherits various (topological) dynamical properties from the initial system. However, Boroński and Oprocha showed that the system induced from Arnold’s cat map loses the property called the Bowen specification property. Their proof relies on two key facts:

  • A local behavior of orbits of points near a hyperbolic fixed point.

  • A nontrivial result of Pfister-Sullivan on the density of ergodic measures in the space of invariant probability measures (see [21]).

A simplified version of the proof of the failure of Bowen specification in [5], without using Pfister-Sullivan’s result [21], is given as our final result. In fact, a weaker specification-like property called the approximate product property is shown to fail on the systems on the carpet.

Theorem D.

Any dynamical system on the carpet constructed in this paper, regardless of its initial system, does not have the approximate product property.

Acknowledgements. This paper is based on the Ph.D. dissertation of the author. The author expresses the deepest gratitude to Prof. Thomas Andrew Schmidt (Oregon State University) for his guidance and valuable suggestions given throughout the author’s Ph.D. study. A part of results related to the pair (C2,B2)(C_{2},B_{2}) is inspired by the talk given by Dr. Meyer on 27th27^{\text{th}} April 2021 in the Quasiworld seminar entitled: The Solenoid, the Chamanara Space, and Symbolic Dynamics. For this particular matter, the author would like to specially acknowledge Dr. Daniel Meyer (University of Liverpool).

2. Preliminaries

The following notions will be used in this paper:

  • 𝕋2{\mathbb{T}}^{2}

    The surface of genus 11 (the 22-torus).

  • FAF_{A}

    The hyperbolic toral automorphism given by FA(z)=Az(mod1)F_{A}(z)=Az\pmod{1}.

  • XgX_{g}

    A hyperelliptic translation surface of genus g2.g\geq 2.

  • TλT_{\lambda}

    A pseudo-Anosov diffeomorphism with the expansion factor λ>1.\lambda>1.

  • CnC_{n}

    The nn-Chamanara surface, n2n\geq 2.

  • BnB_{n}

    The nn-baker transformation, n2n\geq 2.

  • Σn\Sigma_{n}

    The space of bi-infinite sequences on nn symbols {0,1,,n1}.\{0,1,\ldots,n-1\}.

  • σn\sigma_{n}

    The left-shift homeomorphism on Σn\Sigma_{n}.

  • Per(f)\text{Per}(f)

    The set of periodic points of the function ff.

  • (X){\mathcal{B}}(X)

    The Borel σ\sigma-algebra on the topological space XX.

  • hμ(f)h_{\mu}(f)

    The metric entropy of the function ff with respect to the measure μ\mu.

  • (S(T),H(T))(S_{\infty}(T),H_{\infty}(T))

    The dynamical system on the Sierpiński carpet constructed from the initial system (X,T)(X,T).

2.1. Dynamical systems

We briefly introduce two notions of dynamical systems used in this paper. Refer to references as [18] and [23] for a more thorough discussion.

  • (Topological dynamical systems) A pair (X,T)(X,T) consisting of a topological space XX and a continuous map T:XXT:X\rightarrow X is called a topological dynamical system.

  • (Measure-theoretic dynamical systems) A quadruple (X,T,,μ)(X,T,{\mathcal{F}},\mu) consisting of a topological space XX, a continuous function T:XXT:X\rightarrow X, a σ\sigma-algebra {\mathcal{F}} on XX and a measure μ:[0,)\mu:{\mathcal{F}}\rightarrow[0,\infty) satisfying that μ(A)=μ(T1(A))for allA(this condition is calledμisTinvariant)\mu(A)=\mu(T^{-1}(A))\ \text{for all}\ A\in{\mathcal{F}}\ \text{(this condition is called}\ \mu\ \textit{is}\ T-\text{invariant}) is called a measure-theoretic dynamical system.

In the latter case, without specifying otherwise, we assume that =(X){\mathcal{F}}={\mathcal{B}}(X). Also, we suppresses the words topological or measure-theoretic whenever possible.

There is one particular technique of achieving a measure-theoretic structure for a given pair (X,T)(X,T) called a push-forward of a measure-theoretic structure.

Lemma 2.1.1 (A push-forward of a measure-theoretic structure).

Let (X,T)(X,T) and (Y,S)(Y,S) be dynamical systems. Assume that (Y,S)(Y,S) is a factor of (X,T)(X,T) via π:XY.\pi:X\rightarrow Y. If in addition (X,T,(X),μX)(X,T,{\mathcal{B}}(X),\mu_{X}) is a dynamical system, then there is an extension system (Y,S,(Y),μY)(Y,S,{\mathcal{B}}(Y),\mu_{Y}), which is a factor of (X,T,(X),μX)(X,T,{\mathcal{B}}(X),\mu_{X}), given by μY:=μXπ1\mu_{Y}:=\mu_{X}\circ\pi^{-1}.

Lemma 2.1.1 plays a crucial role in extending the pair (Cn,Bn)(C_{n},B_{n}) for each n2n\geq 2. One shall also see that it is the first step to develop the result all positive real numbers can be achieved as entropy values of dynamical systems on the carpet.

The main dynamical systems in this paper are (Cn,Bn)(C_{n},B_{n}) for all n2n\geq 2. We also discuss briefly dynamical systems of (𝕋2,FA),(Xg,Tλ)({\mathbb{T}}^{2},F_{A}),(X_{g},T_{\lambda}) and (Σn,σn)(\Sigma_{n},\sigma_{n}) for g,n2.g,n\geq 2. One can study definitions and facts regarding the pair (𝕋2,FA)({\mathbb{T}}^{2},F_{A}) in [23]. Standard references for (Xg,Tλ)(X_{g},T_{\lambda}) include [26] and [12], with an exception that we prefer an equivalent definition of pseudo-Anosov diffeomorphisms stated in [14]. For basic notions and results related to symbolic dynamics, consult [18].

2.2. Inverse limit systems

An important result used both in [5] and here is Brown’s inverse limit theorem (see [7]). It concerns a recognition of inverse limit systems via a notion of near homeomorphisms.

Definition 2.2.1.

Let (Xi)i(X_{i})_{i\in{\mathbb{N}}} be a sequence of compact metric spaces. Let fi:XiXi1f_{i}:X_{i}\rightarrow X_{i-1} be a continuous map for all i2i\geq 2. The subspace Lim(Xi,fi)\text{Lim}(X_{i},f_{i}) of the product space i=1Xi\prod_{i=1}^{\infty}X_{i} defined as

Lim(Xi,fi):={(ui)i=1Xi:fij(uj):=fi+1fi+2fj(uj)=uifor allijwithfii=idXi}\text{Lim}(X_{i},f_{i}):=\Big{\{}(u_{i})\in\prod_{i=1}^{\infty}X_{i}:f_{ij}(u_{j}):=f_{i+1}\circ f_{i+2}\circ\cdots\circ f_{j}(u_{j})=u_{i}\ \text{for all}\ i\leq j\ \text{with}\ f_{ii}=\text{id}_{X_{i}}\Big{\}}

is called the limit space of the inverse system (Xi,fi).(X_{i},f_{i}).

Definition 2.2.2.

Let (X,d)(X,d) be a metric space. Then a map f:XXf:X\rightarrow X is call a near homeomorphism if for any ϵ>0\epsilon>0, there exists a homeomorphism hϵ:XXh_{\epsilon}:X\rightarrow X such that

hϵf:=supxXd(hϵ(x),f(x))<ϵ.||h_{\epsilon}-f||:=\mathrm{sup}_{x\in X}d(h_{\epsilon}(x),f(x))<\epsilon.
Proposition 2.2.1.

(Theorem 44 in [7]) Let (Xi)(X_{i}) be a sequence of topological spaces. Assume that there exists a compact metric space XX such that XiX_{i} is homeomorphic to XX for all i.i\in{\mathbb{N}}. If for all i2,fi:XiXi1i\geq 2,f_{i}:X_{i}\rightarrow X_{i-1} is a near homeomorphism, then Lim(Xi,fi)\text{Lim}(X_{i},f_{i}) is homeomorphic to X.X.

2.3. Boroński and Oprocha’s inverse limit construction

Core steps of Boroński and Oprocha’s inverse limit construction found in [5] are listed here.
Step 1: We start with a system of Arnold’s cat map on 𝕋2{\mathbb{T}}^{2}, FC(x,y)t=(2111)(x,y)t(mod1)F_{C}(x,y)^{t}=\begin{pmatrix}2&1\\ 1&1\end{pmatrix}(x,y)^{t}\pmod{1}. This is a well-known dynamical system with various intriguing properties: one of them is that it is a branched covering system of the 22-sphere SS2\SS^{2}. In particular, one defines an equivalence relation (x,y)𝕋2(x,y)(x,y)\sim_{{\mathbb{T}}^{2}}-(x,y) and forms a quotient space 𝒮0=𝕋2/𝕋2.{\mathcal{S}}_{0}={\mathbb{T}}^{2}/\sim_{{\mathbb{T}}^{2}}. It turns out that the quotient 𝒮0{\mathcal{S}}_{0} is homeomorphic to SS2\SS^{2}, and there exists a homeomorphism H0:𝒮0𝒮0H_{0}:{\mathcal{S}}_{0}\rightarrow{\mathcal{S}}_{0} such that πFC=H0π\pi\circ F_{C}=H_{0}\circ\pi where π:𝕋2𝒮0\pi:{\mathbb{T}}^{2}\rightarrow{\mathcal{S}}_{0} is the quotient map. Observe that there are four branch points 𝒞={(0,0),(1/2,0),(0,1/2),(1/2,1/2)}{\mathcal{C}}=\{(0,0),(1/2,0),(0,1/2),(1/2,1/2)\} corresponding to 𝕋2.\sim_{{\mathbb{T}}^{2}}.
Step 2: Since Per(FC)¯=𝕋2\overline{\text{Per}(F_{C})}={\mathbb{T}}^{2}, Per(H0)¯=𝒮0\overline{\text{Per}(H_{0})}={\mathcal{S}}_{0} via the semiconjugacy π\pi. Moreover, using the facts that 𝒮0{\mathcal{S}}_{0} is a connected second-countable Hausdorff topological space, one can decompose Per(H0)\text{Per}(H_{0}) as Per(H0)=𝒪𝒪\text{Per}(H_{0})={\mathcal{O}}\sqcup{\mathcal{O}}^{\prime} where H0(𝒪)=𝒪,H0(𝒪)=𝒪,𝒪π(𝒞)=H_{0}({\mathcal{O}})={\mathcal{O}},H_{0}({\mathcal{O}}^{\prime})={\mathcal{O}}^{\prime},{\mathcal{O}}\cap\pi({\mathcal{C}})=\emptyset and 𝒪¯=𝒪¯=𝒮0.\overline{{\mathcal{O}}}=\overline{{\mathcal{O}}^{\prime}}={\mathcal{S}}_{0}.
Step 3: We further decompose 𝒪=i=1Oi{\mathcal{O}}=\sqcup_{i=1}^{\infty}O_{i} where each OiO_{i} is a full orbit of the periodic point ziz_{i} with the period length nin_{i}. Then a technique called a blow up of a point on a (translation) surface is applied. Roughly, given a point c𝒮0c\in{\mathcal{S}}_{0}, a blow up of a point cc is a pair (Bloc(𝒮0),πc)(\text{Blo}_{c}({\mathcal{S}}_{0}),\pi_{c}) consisting of a topological space Bloc(𝒮0)\text{Blo}_{c}({\mathcal{S}}_{0}) and a continuous function πc:Bloc(𝒮0)𝒮0\pi_{c}:\text{Blo}_{c}({\mathcal{S}}_{0})\rightarrow{\mathcal{S}}_{0} (called a collapsing map at cc) satisfying that

  • πc1({c})\pi_{c}^{-1}(\{c\}) is a circle of directions centered at cc (i.e. a topological copy of the 11-sphere SS1\SS^{1}) denoted by SS1(c)\SS^{1}(c),

  • πc|(Bloc(𝒮0)SS1(c))\pi_{c}|_{(\text{Blo}_{c}({\mathcal{S}}_{0})\setminus\SS^{1}(c))} is a homeomorphism.

Basically, blowing up a point c𝒮0c\in{\mathcal{S}}_{0} is a way to create a topological space from which an interior of a closed disc centered at cc is removed. The collapsing map πc\pi_{c} gives suitable identifications of points between 𝒮0{\mathcal{S}}_{0} and Bloc(𝒮0)\text{Blo}_{c}({\mathcal{S}}_{0}).
Step 4: We first blow up points in O1O_{1} to create a space 𝒮1{\mathcal{S}}_{1} and a continuous map π1:𝒮1𝒮0\pi_{1}:{\mathcal{S}}_{1}\rightarrow{\mathcal{S}}_{0}. Via the collapsing map π1\pi_{1} and the fact that FCF_{C} preserves local radial lines, a homeomorphism H1:𝒮1𝒮1H_{1}:{\mathcal{S}}_{1}\rightarrow{\mathcal{S}}_{1} satisfying that π1H1=H0π1\pi_{1}\circ H_{1}=H_{0}\circ\pi_{1} is induced.
Step 5: An induction based on Step 4 is performed. In fact, if a triple (𝒮k,Hk,πk)({\mathcal{S}}_{k},H_{k},\pi_{k}) consisting of a space 𝒮k{\mathcal{S}}_{k}, a homeomorphism Hk:𝒮k𝒮kH_{k}:{\mathcal{S}}_{k}\rightarrow{\mathcal{S}}_{k} satisfying πkHk=Hk1πk\pi_{k}\circ H_{k}=H_{k-1}\circ\pi_{k} and a collapsing map πk:𝒮k𝒮k1\pi_{k}:{\mathcal{S}}_{k}\rightarrow{\mathcal{S}}_{k-1} is already created by blowing up points belonging to OkO_{k}, then we blow up points in Ok+1O_{k+1} to create a triple (𝒮k+1,Hk+1,πk+1)({\mathcal{S}}_{k+1},H_{k+1},\pi_{k+1}).
Step 6: As a consequence, a limit set

S(FC)={(ui)i0i=0𝒮i|πj(uj)=uj1}S_{\infty}(F_{C})=\Big{\{}(u_{i})_{i\geq 0}\in\prod_{i=0}^{\infty}{\mathcal{S}}_{i}\Big{|}\pi_{j}(u_{j})=u_{j-1}\Big{\}}

is created. The result of Brown in Proposition 2.2.1 is applied to show that S(FC)S_{\infty}(F_{C}) is a Sierpiński curve. A function H(FC):S(FC)S(FC)H_{\infty}(F_{C}):S_{\infty}(F_{C})\rightarrow S_{\infty}(F_{C}) defined by

H(FC)(ui)i0=(Hi(ui))i0H_{\infty}(F_{C})(u_{i})_{i\geq 0}=(H_{i}(u_{i}))_{i\geq 0}

is a homeomorphism on S(FC)S_{\infty}(F_{C}).

As we seek to generalize this construction, we give a list of important properties needed in the construction:

  • One chooses an initial system (X,T)(X,T) with nice dynamical properties.

  • In particular, the system (X,T)(X,T) admits an involution {\mathcal{I}} on XX such that (X,T)(X,T) is a branched covering system of the 22-sphere SS2\SS^{2} via {\mathcal{I}}. That is, the quotient 𝒮0=X/{\mathcal{S}}_{0}=X/{\mathcal{I}} is homeomorphic to SS2\SS^{2}, and there exists a homeomorphism H0H_{0} on 𝒮0{\mathcal{S}}_{0} such that πH0=Tπ.\pi\circ H_{0}=T\circ\pi.

  • The set Per(T)\text{Per}(T) is dense in XX (and hence so is Per(H0)\text{Per}(H_{0}) in 𝒮0{\mathcal{S}}_{0}).

  • The homeomorphism TT preserves local radial lines. Note that any affine diffeomorphism of a plane possesses this property.

Since Arnold’s cat map FCF_{C} belongs to the class of hyperbolic toral automorphisms, it is not hard to believe, and indeed it is easily shown, that any FAF_{A} satisfies all the requirements. Though less trivial, any pair (Xg,Tλ)(X_{g},T_{\lambda}) have all required properties as well (we discuss this formally in Subsection 3.1). We check that each pair (Cn,Bn),n2,(C_{n},B_{n}),n\geq 2, has all these properties in Subsection 3.2.

2.4. Metric entropy of dynamical systems

This paper deals with a metric entropy of a given system (X,T,(X),μX)(X,T,{\mathcal{B}}(X),\mu_{X}). In particular, the two following propositions below are sufficient to obtain entropy values of dynamical systems on the carpet. The first one is Proposition 4.3.16 in [18].

Proposition 2.4.1.

Given dynamical systems (X,T,(X),μX)(X,T,{\mathcal{B}}(X),\mu_{X}) and (Y,S,(Y),(Y,S,{\mathcal{B}}(Y), μY)\mu_{Y}) such that the latter is a metric factor of the former system, then hμY(S)hμX(T).h_{\mu_{Y}}(S)\leq h_{\mu_{X}}(T). In addition, if the two systems are isomorphic, then hμY(S)=hμX(T).h_{\mu_{Y}}(S)=h_{\mu_{X}}(T).

Given that the entropy hμX(T)h_{\mu_{X}}(T) is known, one obtains an upper bound hμY(S)hμX(T)h_{\mu_{Y}}(S)\leq h_{\mu_{X}}(T). In the special case when XX is a compact metric space, the work of F. Ledrappier and P. Walters in [20] gives a lower bound for hμY(S)h_{\mu_{Y}}(S). In many cases, the lower bound is obtained and is equal to hμX(T)h_{\mu_{X}}(T) itself implying that hμY(S)=hμX(T).h_{\mu_{Y}}(S)=h_{\mu_{X}}(T).

Definition 2.4.1.

Let (X,d)(X,d) be a metric space and T:XXT:X\rightarrow X be uniformly continuous.

  • A set FF is said to (n,ϵ)(n,\epsilon)-span a set KK if for each xKx\in K, there is yFy\in F satisfying d(Tj(x),Tj(y))ϵd(T^{j}(x),T^{j}(y))\leq\epsilon for all 0j<n.0\leq j<n.

  • For a compact set KXK\subseteq X, let rn(ϵ,K)r_{n}(\epsilon,K) be the smallest cardinality of any set FF which (n,ϵ)(n,\epsilon)-spans KK. Then define r¯T(ϵ,K):=lim supnlogrn(ϵ,K)n.\overline{r}_{T}(\epsilon,K):=\limsup_{n\rightarrow\infty}\frac{\log r_{n}(\epsilon,K)}{n}.

  • For KXK\subseteq X compact, define h(T,K)=limϵ0r¯T(ϵ,K).h(T,K)=\lim_{\epsilon\rightarrow 0}\overline{r}_{T}(\epsilon,K).

Proposition 2.4.2.

(Ledrappier-Walters [20]) Let XX and YY be compact metric spaces. Let T:XXT:X\rightarrow X and S:YYS:Y\rightarrow Y be continuous. Assume that there is a continuous map π:XY\pi:X\rightarrow Y such that πT=Sπ\pi\circ T=S\circ\pi. Then

supμ:μπ1=νhμ(T)=hν(S)+Yh(T,π1(y))𝑑ν(y).\mathrm{sup}_{\mu:\mu\circ\pi^{-1}=\nu}h_{\mu}(T)=h_{\nu}(S)+\int_{Y}h(T,\pi^{-1}(y))\ d\nu(y).

Additionally, if |π1(y)|<|\pi^{-1}(y)|<\infty for all yYy\in Y, then

supμ:μπ1=νhμ(T)=hν(S).\mathrm{sup}_{\mu:\mu\circ\pi^{-1}=\nu}h_{\mu}(T)=h_{\nu}(S).

2.5. The Bowen specification property

Bowen specification property is a strong shadowing property introduced by Rufus Bowen in 1971 (see [6]). Its usefulness has been evident among dynamicists, especially in the field of hyperbolic dynamics. See [19] for a great overview on this topic. We define formally here only two specification-like properties: the Bowen specification property and the approximate product property.

Definition 2.5.1.

Let (X,d)(X,d) be a compact metric space and let T:XXT:X\rightarrow X be a continuous surjective map.

  • The pair (X,T)(X,T) has the (Bowen) specification property if for any ϵ>0\epsilon>0, there exists a positive integer N(ϵ)N(\epsilon) such that for any integer s2s\geq 2, given any ss points y1,y2,,ysXy_{1},y_{2},...,y_{s}\in X and any sequence of 2s2s integers 0=j1k1<j2k2<<jsks0=j_{1}\leq k_{1}<j_{2}\leq k_{2}<...<j_{s}\leq k_{s} satisfying jm+1kmNj_{m+1}-k_{m}\geq N for all m=1,2,,s1m=1,2,...,s-1, there exists a point xXx\in X (called a shadowing point) with the property for each positive integer msm\leq s,

    d(Ti(x),Ti(ym))<ϵd(T^{i}(x),T^{i}(y_{m}))<\epsilon

    for all i=jm,jm+1,,km.i=j_{m},j_{m}+1,\ldots,k_{m}.

  • The pair (X,T)(X,T) has the approximate product property if for all ϵ>0,δ1>0,δ2>0\epsilon>0,\delta_{1}>0,\delta_{2}>0, there exists N=N(ϵ,δ1,δ2)N=N(\epsilon,\delta_{1},\delta_{2})\in{\mathbb{N}} such that for any nNn\geq N and for any sequence of points (xi)i1X(x_{i})_{i\geq 1}\subseteq X, there are (a sequence of gaps) (hi)i10(h_{i})_{i\geq 1}\subseteq{\mathbb{N}}_{0} satisfying that nhi+1hin+nδ2n\leq h_{i+1}-h_{i}\leq n+n\delta_{2} and a point xXx\in X, for each ii\in{\mathbb{N}}

    |{0jn1:d(Thi+j(x),Tj(xi))ϵ}|<δ1n.|\{0\leq j\leq n-1:d(T^{h_{i}+j}(x),T^{j}(x_{i}))\geq\epsilon\}|<\delta_{1}n.

Observe that compactness together with the Bowen specification imply the approximate product property (see [21]). The converse is false as discussed in [19]. Another important fact is that the Bowen specification is invariant under a topological semiconjugacy. Since (Σn,σn)(\Sigma_{n},\sigma_{n}) and any subshift of finite type have the Bowen specification (see say [24]), hyperbolic toral automorphisms FAF_{A}, pseudo-Anosov diffeomorphisms TλT_{\lambda} and the nn-baker map BnB_{n} all have the Bowen specification (the first two are factors of some shifts of finite type, and the latter, as we soon prove, is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n})). So it is a reasonable question to ask if a dynamical system on the carpet built from an initial system with the specification still has the specification. The answer is negative for the systems induced from (𝕋2,FA),(Xg,Tλ)({\mathbb{T}}^{2},F_{A}),(X_{g},T_{\lambda}) or (Cn,Bn)(C_{n},B_{n}) (see Section 5).

3. The pair (Cn,Bn)(C_{n},B_{n}) as a branched covering system of SS2\SS^{2}

3.1. A brief remark on (𝕋2,FA)({\mathbb{T}}^{2},F_{A}) and (Xg,Tλ),g2(X_{g},T_{\lambda}),g\geq 2

The result for a pair (𝕋2,FA)({\mathbb{T}}^{2},F_{A}) is immediate as all arguments and proofs in [5] using Arnold’s cat map FCF_{C} remain valid when one replaces FCF_{C} by any FAF_{A}.

For the pair (Xg,Tλ)(X_{g},T_{\lambda}), we collect here all necessary facts:

  • The set Per(Tλ)\text{Per}(T_{\lambda}) is dense in XgX_{g} (see Proposition 9.20 in [13]).

  • It is from the definition of XgX_{g} that it admits an involution :XgXg{\mathcal{I}}:X_{g}\rightarrow X_{g} such that 𝒮0=Xg/{\mathcal{S}}_{0}=X_{g}/{\mathcal{I}} is a 22-sphere. Moreover, Lemma 2.3 in [4] gives that there is a homeomorphism H0H_{0} on 𝒮0{\mathcal{S}}_{0} which properly commutes with TλT_{\lambda}.

  • The fact that TλT_{\lambda} being an affine diffeomorphism gives that it preserves radial lines locally.

As a result, one concludes that both families of (𝕋2,FA)({\mathbb{T}}^{2},F_{A}) and (Xg,Tλ)(X_{g},T_{\lambda}) are valid as initial systems for Boroński and Oprocha’s inverse limit construction.

3.2. The proof of the first main result

We now verify the first main result corresponding to Theorem A and B above. We present the proof in three parts. The first part is to show that each pair (Cn,Bn)(C_{n},B_{n}) is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n}). This establishes that Per(Bn)¯=Cn\overline{\text{Per}(B_{n})}=C_{n} and a way to obtain a measure-theoretic structure for the pair (Cn,Bn)(C_{n},B_{n}) (which we discuss later in Section 4). After this we form a quotient space QnQ_{n} of CnC_{n} by an involution of a rotation by π\pi radians about (0.5,0.5)(0.5,0.5). The second part is devoted to showing that QnQ_{n} is a 22-sphere. The last part verifies an existence of a homeomorphism TnT_{n} on QnQ_{n} commuting with BnB_{n}.

3.2.1. The pair (Cn,Bn)(C_{n},B_{n}) as a factor of (Σn,σn)(\Sigma_{n},\sigma_{n})

Let n2n\geq 2 be fixed. We begin by listing each element ω=(x,y)¯=(0.x1x2n,0.y1y2n)¯Cn\omega=\overline{(x,y)}=\overline{(0.x_{1}x_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n})}\in C_{n} explicitly

ω={{ω}ifω(0,1)2,{(0,0.y1y2n),(1,0.y1y2n)}if(x,y)I1I1,{(1,0.nnnykyk+1n),(0,0.000yk+yk+1n)}if(x,y)IkIk,k2{(0.x1x2n,0),(0.x1x2n,1)}if(x,y)J1J1,{(0.nnnnxkxk+1n,1),(0.0000xk+xk+1n,0)}if(x,y)JkJk,k2\omega=\begin{cases}\{\omega\}&\mbox{if}\ \omega\in(0,1)^{2},\\ \{(0,0.y_{1}y_{2}\ldots_{n}),(1,0.y_{1}^{-}y_{2}\ldots_{n})\}&\mbox{if}\ (x,y)\in I_{1}\cup I_{1}^{\prime},\\ \{(1,0.n^{-}n^{-}\ldots n^{-}y_{k}y_{k+1}\ldots_{n}),(0,0.00\ldots 0y_{k}^{+}y_{k+1}\ldots_{n})\}&\mbox{if}\ (x,y)\in I_{k}\cup I_{k}^{\prime},k\geq 2\\ \{(0.x_{1}x_{2}\ldots_{n},0),(0.x_{1}^{-}x_{2}\ldots_{n},1)\}&\mbox{if}\ (x,y)\in J_{1}\cup J_{1}^{\prime},\\ \{(0.n^{-}n^{-}n^{-}\ldots n^{-}x_{k}x_{k+1}\ldots_{n},1),(0.000\ldots 0x_{k}^{+}x_{k+1}\ldots_{n},0)\}&\mbox{if}\ (x,y)\in J_{k}\cup J_{k}^{\prime},k\geq 2\end{cases}

where a:=a1,a+:=a+1,0.a1a2n:=i=1ainia^{-}:=a-1,a^{+}:=a+1,0.a_{1}a_{2}\ldots_{n}:=\sum_{i=1}^{\infty}\frac{a_{i}}{n^{i}},

(0,0)¯={(0,0),(1,1),(1,0),(0,1)}{(0,1nk),(1nk,0),(1,11nk),(11nk,1):k}and fori1,\overline{(0,0)}=\{(0,0),(1,1),(1,0),(0,1)\}\cup\Big{\{}\Big{(}0,\frac{1}{n^{k}}\Big{)},\Big{(}\frac{1}{n^{k}},0\Big{)},\Big{(}1,1-\frac{1}{n^{k}}\Big{)},\Big{(}1-\frac{1}{n^{k}},1\Big{)}:k\in{\mathbb{N}}\Big{\}}\ \text{and for}\ i\geq 1,
Ii={0}×(1ni,1ni1),Ii={1}×(11ni1,11ni),Ji=(1ni,1ni1)×{0},Ji=(11ni1,11ni)×{1}.I_{i}=\{0\}\times\Big{(}\frac{1}{n^{i}},\frac{1}{n^{i-1}}\Big{)},I_{i}^{\prime}=\{1\}\times\Big{(}1-\frac{1}{n^{i-1}},1-\frac{1}{n^{i}}\Big{)},J_{i}=\Big{(}\frac{1}{n^{i}},\frac{1}{n^{i-1}}\Big{)}\times\{0\},J_{i}^{\prime}=\Big{(}1-\frac{1}{n^{i-1}},1-\frac{1}{n^{i}}\Big{)}\times\{1\}.
Refer to caption
Figure 2. The surface CnC_{n} described using identifications on Ii,Ii,JiI_{i},I_{i}^{\prime},J_{i} and JjJ_{j}^{\prime}.

One observes that for (0,y)I1(0,y)\in I_{1}, y=0.y1y2ny=0.y_{1}y_{2}\ldots_{n} where y1{1,2,,n1}.y_{1}\in\{1,2,\ldots,n-1\}. The identification to I1I_{1}^{\prime} translates the yy-coordinate down by 1n=0.1n\frac{1}{n}=0.1_{n}. So (0,y)(0,y) is identified to (1,y0.1n)=(1,y1y2n)I1.(1,y-0.1_{n})=(1,y_{1}^{-}y_{2}\ldots_{n})\in I_{1}^{\prime}. As for k2,k\geq 2, let z=(0,y)Ikz=(0,y)\in I_{k}. Then y=0.y1y2yk1ykny=0.y_{1}y_{2}\ldots y_{k-1}y_{k}\ldots_{n} where y1=y2==yk1=0y_{1}=y_{2}=\ldots=y_{k-1}=0 and yk{1,2,,n1}.y_{k}\in\{1,2,\ldots,n-1\}. One sees that nk1zn^{k-1}z shifts zIkz\in I_{k} to nk1zI1n^{k-1}z\in I_{1}. Then nk1z=(0,0.ykn)n^{k-1}z=(0,0.y_{k}\ldots_{n}) is identified with w=(1,yk1ykn)I1.w=(1,y_{k-1}^{-}y_{k}\ldots_{n})\in I_{1}^{\prime}. After that the yy-coordinate of ww is retracted by n1kn^{1-k}. It is lastly translated up to Ik.I_{k}^{\prime}. This gives that z=(0.000ykyk+1n)is identified with(1,0.nnnykyk+1n)Ik.z=(0.00\ldots 0y_{k}y_{k+1}\ldots_{n})\ \text{is identified with}\ (1,0.n^{-}n^{-}\ldots n^{-}y_{k}^{-}y_{k+1}\ldots_{n})\in I_{k}^{\prime}.

An n-baker map is a function Bn:[0,1)2[0,1)2B_{n}:[0,1)^{2}\rightarrow[0,1)^{2} defined by Bn(x,y)=(nx(k1),y+(k1)n)B_{n}(x,y)=\Big{(}nx-(k-1),\frac{y+(k-1)}{n}\Big{)} for k1nx<kn\frac{k-1}{n}\leq x<\frac{k}{n} for each k=1,2,,n.k=1,2,\ldots,n. The definition of BnB_{n} naturally extends to CnC_{n} since [0,1)2[0,1)^{2} is a fundamental domain of Cn.C_{n}. We now verify that (Cn,Bn)(C_{n},B_{n}) is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n}). Then, as an immediate consequence, Per(Bn)\text{Per}(B_{n}) is dense in CnC_{n}.

Lemma 3.2.1.

The function BnB_{n} is a homeomorphism on CnC_{n}. The topological dynamical system (Cn,Bn)(C_{n},B_{n}) is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n}) via the map Pn:ΣnCnP_{n}:\Sigma_{n}\rightarrow C_{n} defined by

Pn(bm;an)=(0.a1a2n,0.b1b2n)¯.P_{n}(b_{m};a_{n})=\overline{(0.a_{1}a_{2}\ldots_{n},0.b_{1}b_{2}\ldots_{n})}.

Consequently, Per(Bn)\text{Per}(B_{n}) is dense in Cn.C_{n}.

Proof.

Define vertical strips i{\mathcal{E}}_{i} and vertical right lines 𝒱j{\mathcal{V}}_{j} as, for each 1in,1jn11\leq i\leq n,1\leq j\leq n-1,

i=(i1n,in)×(0,1)and𝒱j={jn}×(0,1).{\mathcal{E}}_{i}=\Big{(}\frac{i-1}{n},\frac{i}{n}\Big{)}\times(0,1)\ \text{and}\ {\mathcal{V}}_{j}=\Big{\{}\frac{j}{n}\Big{\}}\times(0,1).

Then for each 1in1\leq i\leq n, define fif_{i} on i¯\overline{{\mathcal{E}}_{i}} by fi(x,y)=(nx(i1),y+(i1)n).f_{i}(x,y)=\big{(}nx-(i-1),\frac{y+(i-1)}{n}\big{)}. Then fif_{i} is continuous on i¯\overline{{\mathcal{E}}_{i}} for all 1in.1\leq i\leq n. Since j¯j+1¯=𝒱i\overline{{\mathcal{E}}_{j}}\cap\overline{{\mathcal{E}}_{j+1}}={\mathcal{V}}_{i} for 1jn11\leq j\leq n-1, and fj(𝒱j)=fj+1(𝒱j)f_{j}({\mathcal{V}}_{j})=f_{j+1}({\mathcal{V}}_{j}) by the identifications on I1I_{1} and I1I_{1}^{\prime}, BnB_{n} is well-defined and continuous on CnC_{n}. Analogously, one uses the identifications on J1J_{1} and J1J_{1}^{\prime} together with an introduction of suitable horizontal strips and horizontal upper lines to show that Bn1B_{n}^{-1} is well-defined and continuous on Cn.C_{n}.

To show the continuity of PnP_{n}, note that Pn=qnpnP_{n}=q_{n}\circ p_{n} where pn:Σn[0,1]2,qn:[0,1]2Cnp_{n}:\Sigma_{n}\rightarrow[0,1]^{2},q_{n}:[0,1]^{2}\rightarrow C_{n}. It then suffices to show that pn(am;bn)=(0.a1a2n,0.b1b2n)p_{n}(a_{m};b_{n})=(0.a_{1}a_{2}\ldots_{n},0.b_{1}b_{2}\ldots_{n}) is continuous, which is immediate from the fact that if dΣn((am;bn),(cm;dn))2kd_{\Sigma_{n}}((a_{m};b_{n}),(c_{m};d_{n}))\leq 2^{-k}, then

(i=1aicini)2+(i=1bidini)2=(i>kaicini)2+(i>kbidini)22(ik+1n1ni)2=2n2k.\displaystyle\Big{(}\sum_{i=1}^{\infty}\frac{a_{i}-c_{i}}{n^{i}}\Big{)}^{2}+\Big{(}\sum_{i=1}^{\infty}\frac{b_{i}-d_{i}}{n^{i}}\Big{)}^{2}=\Big{(}\sum_{i>k}\frac{a_{i}-c_{i}}{n^{i}}\Big{)}^{2}+\Big{(}\sum_{i>k}\frac{b_{i}-d_{i}}{n^{i}}\Big{)}^{2}\leq 2\Big{(}\sum_{i\geq k+1}\frac{n-1}{n^{i}}\Big{)}^{2}=\frac{2}{n^{2k}}.

Lastly, we verify the commutativity of the diagram. For 0in10\leq i\leq n-1, let

𝕍i={(ym;xn):x1=i,xj=n1for allj2}{(ym;xn):x1=i+1,xj=0for allj2}.{\mathbb{V}}_{i}=\{(y_{m};x_{n}):x_{1}=i,x_{j}=n-1\ \mbox{for all}\ j\geq 2\}\cup\{(y_{m};x_{n}):x_{1}=i+1,x_{j}=0\ \mbox{for all}\ j\geq 2\}.

Define 𝔼i={(ym;xn)Σn:x1=i}(𝕍i1𝕍i){\mathbb{E}}_{i}=\{(y_{m};x_{n})\in\Sigma_{n}:x_{1}=i\}\setminus({\mathbb{V}}_{i-1}\cup{\mathbb{V}}_{i}) and 𝔼0={(ym;xn)Σn:x1=0}(𝕍0{(ym;xn):xn=0for alln}){\mathbb{E}}_{0}=\{(y_{m};x_{n})\in\Sigma_{n}:x_{1}=0\}\setminus({\mathbb{V}}_{0}\cup\{(y_{m};x_{n}):x_{n}=0\ \mbox{for all}\ n\in{\mathbb{N}}\}) for each 1in11\leq i\leq n-1. One regards these 𝔼i{\mathbb{E}}_{i} and 𝕍j{\mathbb{V}}_{j} as notions of vertical strips and vertical right lines in the space Σn.\Sigma_{n}. Note that Σn=i=0n1(𝔼i𝕍i){(ym;xn):xn=0for alln}.\Sigma_{n}=\sqcup_{i=0}^{n-1}({\mathbb{E}}_{i}\sqcup{\mathbb{V}}_{i})\sqcup\{(y_{m};x_{n}):x_{n}=0\ \mbox{for all}\ n\in{\mathbb{N}}\}.
Let z=(ym;xn)Σn.z=(y_{m};x_{n})\in\Sigma_{n}.
Case 1: Assume that z𝔼kz\in{\mathbb{E}}_{k} for some 0kn10\leq k\leq n-1.
Then BnPn(z)=Bn(0.kx2n,0.y1y2n)=(0.x2x3n,0.ky1y2n)=Pnσn(z).B_{n}P_{n}(z)=B_{n}(0.kx_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n})=(0.x_{2}x_{3}\ldots_{n},0.ky_{1}y_{2}\ldots_{n})=P_{n}\sigma_{n}(z).
Case 2: Assume z𝕍kz\in{\mathbb{V}}_{k} for some 0kn20\leq k\leq n-2.
Then

BnPn(z)=Bn(0.kn+,0.y1y2n)=(0,0.k+y1y2n)=Pn(y2y1k+;00)=Pnσn(z),andB_{n}P_{n}(z)=B_{n}(0.k^{+}_{n},0.y_{1}y_{2}\ldots_{n})=(0,0.k^{+}y_{1}y_{2}\ldots_{n})=P_{n}(\ldots y_{2}y_{1}k^{+};0\ldots 0\ldots)=P_{n}\sigma_{n}(z),\ \text{and}
BnPn(w)\displaystyle B_{n}P_{n}(w) =Bn(0.knnn,0.y1y2n)=(0.nnn,0.ky1y2n)\displaystyle=B_{n}(0.kn^{-}\ldots n^{-}\ldots_{n},0.y_{1}y_{2}\ldots_{n})=(0.n^{-}\ldots n^{-}\ldots_{n},0.ky_{1}y_{2}\ldots_{n})
=Pn(y2y1k;nn)=Pnσn(w)\displaystyle=P_{n}(\ldots y_{2}y_{1}k;n^{-}\ldots n^{-}\ldots)=P_{n}\sigma_{n}(w)

where z=(y2y1;k+00),w=(y2y2;knn)z=(\ldots y_{2}y_{1};k^{+}0\ldots 0\ldots),w=(\ldots y_{2}y_{2};kn^{-}\ldots n^{-}\ldots).
Case 3: Assume that z=(ym;0)z=(y_{m};0).
Then BnPn(z)=Bn(0,0.y1y2n)=(0,0.0y1y2n)=Pnσn(z).B_{n}P_{n}(z)=B_{n}(0,0.y_{1}y_{2}\ldots_{n})=(0,0.0y_{1}y_{2}\ldots_{n})=P_{n}\sigma_{n}(z).
Case 4: Assume that z=(ym;nn)z=(y_{m};n^{-}\ldots n^{-}\ldots)
Then BnPn(z)=Bn(0.nnn,0.y1y2n)=(0.nnn,0.ny1y2n)=Pnσn(z).B_{n}P_{n}(z)=B_{n}(0.n^{-}\ldots n^{-}\ldots_{n},0.y_{1}y_{2}\ldots_{n})=(0.n^{-}\ldots n^{-}\ldots_{n},0.n^{-}y_{1}y_{2}\ldots_{n})=P_{n}\sigma_{n}(z).
So the system (Cn,Bn)(C_{n},B_{n}) is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n}). ∎

3.2.2. The quotient QnQ_{n} is a 22-sphere SS2\SS^{2}

As mentioned earlier, an involution RR on [0,1]2[0,1]^{2} is a rotation by π\pi radians about the point (0.5,0.5)(0.5,0.5). It is a straight-forward calculation to see that CnC_{n} admits an equivalence relation ωRR(ω).\omega\sim_{R}R(\omega). In fact, for ω1={(0,0.y1y2n),(1,0.y1y2n)}I1I1\omega_{1}=\{(0,0.y_{1}y_{2}\ldots_{n}),(1,0.y_{1}^{-}y_{2}\ldots_{n})\}\subseteq I_{1}\cup I_{1}^{\prime},

R(0,0.y1y2n)=(1,0.y1y2n)Cn(0,0.(y1)y2n)=(0,0.(y1)y2n)=R(1,0.y1y2n)R(0,0.y_{1}y_{2}\ldots_{n})=(1,0.y_{1}^{*}y_{2}^{*}\ldots_{n})\sim_{C_{n}}(0,0.(y_{1}^{*})^{-}y_{2}^{*}\ldots_{n})=(0,0.(y_{1}^{-})^{*}y_{2}^{*}\ldots_{n})=R(1,0.y_{1}^{-}y_{2}\ldots_{n})

and for ω2={(1,0.nnnykyk+1n),(0,0.000yk+yk+1n)}IkIk\omega_{2}=\{(1,0.n^{-}n^{-}\ldots n^{-}y_{k}y_{k+1}\ldots_{n}),(0,0.00\ldots 0y_{k}^{+}y_{k+1}\ldots_{n})\}\subseteq I_{k}\cup I_{k}^{\prime} for some k2,k\geq 2,

R(1,0.nnykn)=(0,00ykn)Cn(1,nn(yk+)n)=R(0,0.00yk+n)R(1,0.n^{-}\ldots n^{-}y_{k}\ldots_{n})=(0,0\ldots 0y_{k}^{*}\ldots_{n})\sim_{C_{n}}(1,n^{-}\ldots n^{-}(y_{k}^{+})^{*}\ldots_{n})=R(0,0.0\ldots 0y_{k}^{+}\ldots_{n})

where a:=na=(n1)a.a^{*}:=n^{-}-a=(n-1)-a.

So we form a quotient space Qn:=Cn/RQ_{n}:=C_{n}/\sim_{R}, which is now shown to be homeomorphic to SS2\SS^{2}. The base case of Q2Q_{2} is first verified. We recall that this result was already known in [9], but the verification here is done via a new approach of building a certain inverse limit system which is proved to be a 22-sphere using Brown’s work in Proposition 2.2.1.

The quotient Q2Q_{2} can be visualized by side identifications on [0,0.5]×[0,1][0,0.5]\times[0,1] as shown below.

Refer to caption
Figure 3. Side identifications on the boundary of [0,0.5]×[0,1][0,0.5]\times[0,1] results in the quotient Q2Q_{2}. One should compare the identifications and notations with the Figure 2.

To show that Q2SS2Q_{2}\cong\SS^{2}, we adopt the idea of zips similar both to that used in Conway’s zip proof of the classification of closed surfaces (see [15]) and to Veech’s notion of zippered rectangles (see say [26]). In particular, a zip represents an equivalence relation imposed on a certain part of a boundary. With the zip notation, one sees that the quotient Q2Q_{2} is homeomorphic to a 22-sphere which is zipped up countably infinitely many times along the single (singular) vertex point, denoted by SS2(Q2)\SS^{2}(Q_{2}) (see Figure 4 below). It suffices to show that SS2(Q2)SS2.\SS^{2}(Q_{2})\cong\SS^{2}.

Refer to caption
Figure 4. The side identification of I0I_{0} is zipping up along the edge of I0I_{0}. A neighborhood N(I0)N(I_{0}) of the zip Z(I0)Z(I_{0}) is shown in the top right figure. Hence, the space Q2Q_{2} is homeomorphic to the 2-sphere with countably infinitely many zips of diminishing sizes, SS2(Q2)\SS^{2}(Q_{2}).

Observe that each zip Z(P)Z(P) has its own neighborhood called N(P)N(P). These neighborhoods are pairwise disjoint N(P)N(Q)=N(P)\cap N(Q)=\emptyset for zips PQP\neq Q. With this observation, we build an inverse limit system as follows.
Step 1: Let M0M_{0} be a 22-sphere obtained by collapsing the boundary of [0,0.5]×[0,1][0,0.5]\times[0,1] to a single point. Next, we create M1M_{1} as a quotient space of [0,0.5]×[0,1][0,0.5]\times[0,1] obtained by imposing the equivalence relation R\sim_{R} on I0I_{0} and collapse ([0,0.5]×[0,1])(I0(I0))\partial([0,0.5]\times[0,1])\setminus(I_{0}\setminus\partial(I_{0})) to a single point. That is, M1M_{1} is a 22-sphere with a zip representing the relation imposed on I0I_{0}.
Step 2: The space M2M_{2} is a 22-sphere with two zips created by imposing the relation R\sim_{R} on I0I_{0} and J1J_{1}^{\prime}, and collapsing ([0,0.5]×[0,1])((I0(I0))(J1(J1)))\partial([0,0.5]\times[0,1])\setminus((I_{0}\setminus\partial(I_{0}))\cup(J_{1}^{\prime}\setminus\partial(J_{1}^{\prime}))) to a single point.
Step 3: In general, we inductively create a new space Mk+1M_{k+1} from the already created space MkM_{k} by imposing the relation R\sim_{R} to an additional part of ([0,0.5]×[0,1])\partial([0,0.5]\times[0,1]), and collapsing the rest of the boundary part to a single point. In particular, Table 11 shows a pattern of imposing the relation R\sim_{R} to an additional part of the boundary for the first five steps.

Step kthk^{th} of the construction 0th0^{th}-step 1st1^{st}-step 2nd2^{nd}-step 3rd3^{rd}-step 4th4^{th}-step 5th5^{th}-step
The R\sim_{R} is imposed on - I0I_{0} J1J_{1}^{\prime} I1I_{1} J2J_{2} I2I_{2}
The created space M0M_{0} M1M_{1} M2M_{2} M3M_{3} M4M_{4} M5M_{5}
Table 1. A pattern of imposing R\sim_{R} to parts of ([0,0.5]×[0,1])\partial([0,0.5]\times[0,1]) for the first five steps.

Step 4: As a result, we have a sequence of topological spaces (Mi)i0(M_{i})_{i\geq 0} such that each pair of consecutive spaces differ by a zip. The last requirement to build an inverse limit system of (Mk)(M_{k}) is to define a suitable continuous map hk:MkMk1h_{k}:M_{k}\rightarrow M_{k-1} for k.k\in{\mathbb{N}}. The facts that each consecutive pair of spaces differ by a zip, called it Z(P)Z(P), and each zip Z(P)Z(P) has it own neighborhood N(P)N(P) allow us to define the map hkh_{k} locally in N(P)N(P). That is, a map is the identity outside of N(P)N(P). We give a precise definition below.

A function h:N(P)¯N(P)¯h:\overline{N(P)}\rightarrow\overline{N(P)} which collapses the zip Z(P)Z(P) is defined as follows:
Let h(Z(P))={[(0,0)]}h(Z(P))=\{[(0,0)]\} and h(z)=zh(z)=z for all zN(P).z\in\partial N(P). Then we define hh on each radial arc by scaling its length corresponding to how close it is to the zip Z(P)Z(P). The resulting function hh satisfies that

  • hh is a homeomorphism on N(P)Z(P)N(P)\setminus Z(P) and hh is continuous on N(P)N(P),

  • hh fixes the boundary N(P),\partial N(P),

  • but, hh is not injective on Z(P)Z(P).

Refer to caption
Figure 5. The continuous function hh collapses the zip Z(P)Z(P) .
Lemma 3.2.2.

The continuous function h:N(P)¯N(P)¯h:\overline{N(P)}\rightarrow\overline{N(P)} is a near homeomorphism.

Proof.

For each small δ>0,\delta>0, we can define a homeomorphism hδ:N(P)¯N(P)¯h_{\delta}:\overline{N(P)}\rightarrow\overline{N(P)} as follows:

  • Let δ\ell_{\delta} be the curve of length δ\delta contains in Z(P)Z(P) and having [(0,0)][(0,0)] as one of the endpoints. The map hδh_{\delta} sends Z(P)Z(P) to the arc δ\ell_{\delta} instead of the point [(0,0)][(0,0)].

  • The map hδh_{\delta} is the identity on N(P)\partial N(P).

  • Let |Z(P)||Z(P)| denote the length of the zip Z(P)Z(P). Each radial arc in N(P)Z(P)N(P)\setminus Z(P) is the union of an arc LL of length |Z(P)||Z(P)| and its continuation Lc.L^{c}. The map hδh_{\delta} sends LLcL\cup L^{c} to itself by linearly contracting LL and expanding Lc.L^{c}. We set the constant of contraction according to the angle between LL and Z(P)Z(P), and the fixed δ>0.\delta>0.

A motivation for definition of hδh_{\delta} is that as δ0\delta\rightarrow 0, hδh_{\delta} tries to be the same as hh. So by the construction, for a fixed ϵ>0\epsilon>0, then there exists δ>0\delta>0 such that hδh<ϵ.||h_{\delta}-h||<\epsilon.

Refer to caption
Figure 6. The homeomorphism hδh_{\delta} decreases the size of the zip Z(P)Z(P) .

Since MiSS2M_{i}\cong\SS^{2} for all i0i\geq 0, Proposition 2.2.1 and Lemma 3.2.2 yield that

Lim(M):={(zi)i=0Mi:hj(zj)=zj1}SS2.\text{Lim}(M):=\Big{\{}(z_{i})\in\prod_{i=0}^{\infty}M_{i}:h_{j}(z_{j})=z_{j-1}\Big{\}}\cong\SS^{2}.
Proposition 3.2.1.

The topological space SS2(Q2)\SS^{2}(Q_{2}) is homeomorphic to SS2.\SS^{2}. In particular, Q2Q_{2} is homeomorphic to SS2.\SS^{2}.

Proof.

For each i0i\geq 0, we then define a function ϕi:SS2(Q2)Mi\phi_{i}:\SS^{2}(Q_{2})\rightarrow M_{i} by

ϕi(z)={hk(z)forzN(Pk),k>izotherwise.\phi_{i}(z)=\begin{cases}h_{k}(z)&\text{for}\ z\in N(P_{k}),k>i\\ z&\text{otherwise}.\end{cases}

Then ϕi\phi_{i} is continuous because each hkh_{k} is identity for all zN(Pk),k>iz\in\partial N(P_{k}),k>i.
We now have a space SS2(Q2)\SS^{2}(Q_{2}) and a family of continuous functions (ϕi:SS2(Q2)Mi)i0.(\phi_{i}:\SS^{2}(Q_{2})\rightarrow M_{i})_{i\geq 0}.
If for ij,i\geq j, let fij=hihi1hj:MiMjf_{ij}=h_{i}\circ h_{i-1}\circ\ldots\circ h_{j}:M_{i}\rightarrow M_{j} with fii=idMi.f_{ii}=\text{id}_{M_{i}}.
Then (ϕi)(\phi_{i}) is compatible with (fij)(f_{ij}) in the sense that for any iji\geq j, fijϕi=ϕj.f_{ij}\phi_{i}=\phi_{j}.
By the universal property of an inverse limit of topological spaces, there exists a unique continuous function F:SS2(Q2)Lim(M)F:\SS^{2}(Q_{2})\rightarrow\text{Lim}(M) such that for each i0,i\geq 0, ψiF=ϕi\psi_{i}\circ F=\phi_{i} where ψi:Lim(M)Mi\psi_{i}:\text{Lim}(M)\rightarrow M_{i} is the ii-coordinate projection. Define a function G:Lim(M)SS2(Q2)G:\text{Lim}(M)\rightarrow\SS^{2}(Q_{2}) as follows: Let zSS2(Q2)z\in\SS^{2}(Q_{2}).
Case 1: zSS2(Q2)(i=1N(Pi))z\in\SS^{2}(Q_{2})\setminus(\cup_{i=1}^{\infty}N(P_{i}))
Then hk(z)=zh_{k}(z)=z for all k1k\geq 1. So define G(zi)=zG(z_{i})=z where zi=zz_{i}=z for all i0.i\geq 0.
Case 2: zN(Pi0)z\in N(P_{i_{0}}) for some i0i_{0}
Then there exists a unique point (αi)Lim(M)(\alpha_{i})\in\text{Lim}(M) such that αj=z\alpha_{j}=z for some jj. In fact, there is a unique pair of a point wSS2(Q2)w\in\SS^{2}(Q_{2}) and j1j\geq 1 such that (αi)Lim(M)(\alpha_{i})\in\text{Lim}(M) where αi=w\alpha_{i}=w for all iji\leq j and αi=z\alpha_{i}=z for all i>j.i>j. Define G(αi)=zG(\alpha_{i})=z. Lastly, define G(zi)=VG(z_{i})=V where zi=Vz_{i}=V for all i0i\geq 0 and V=[(0,0)].V=[(0,0)]. Then GG is a surjective function from Lim(M)\text{Lim}(M) to SS2(Q2).\SS^{2}(Q_{2}).
Let k1.k\geq 1. Let (zi)Lim(M).(z_{i})\in\text{Lim}(M). Notice one important characteristic of an element (zi)(z_{i}) of Lim(M)\text{Lim}(M): it is eventually constant in the sense that either there exists wSS2(Q2)w\in\SS^{2}(Q_{2}) such that zi=wz_{i}=w for all i0i\geq 0, or there exists a unique triple (v,w,j)(v,w,j) of distinct v,wSS2(Q2)v,w\in\SS^{2}(Q_{2}) and j1j\geq 1 such that zi=vz_{i}=v for all i<ji<j and zi=wz_{i}=w for all ij.i\geq j. For the former case, ϕkG(zi)=ϕk(w)=w=ψk(zi).\phi_{k}\circ G(z_{i})=\phi_{k}(w)=w=\psi_{k}(z_{i}). For the latter case,

ϕkG(zi)=ϕk(w)={vfork<j,wforkj.\phi_{k}\circ G(z_{i})=\phi_{k}(w)=\begin{cases}v&\text{for}\ k<j,\\ w&\text{for}\ k\geq j.\end{cases}

So we conclude that ϕkG=ψk\phi_{k}\circ G=\psi_{k} for all k.k\in{\mathbb{N}}. Observe the facts that

H1:SS2(Q2)SS2(Q2)such thatϕkH1=ϕkkH1=idSS2(Q2)andH_{1}:\SS^{2}(Q_{2})\rightarrow\SS^{2}(Q_{2})\ \text{such that}\ \phi_{k}\circ H_{1}=\phi_{k}\ \forall k\in{\mathbb{N}}\rightarrow H_{1}=\text{id}_{\SS^{2}(Q_{2})}\ \text{and}
H2:Lim(M)Lim(M)such thatψkH2=ψkkH2=idLim(M).H_{2}:\text{Lim}(M)\rightarrow\text{Lim}(M)\ \text{such that}\ \psi_{k}\circ H_{2}=\psi_{k}\ \forall k\in{\mathbb{N}}\rightarrow H_{2}=\text{id}_{\text{Lim}(M)}.

This gives that GF=idSS2(Q2)G\circ F=\text{id}_{\SS^{2}(Q_{2})} and FG=idLim(M).F\circ G=\text{id}_{\text{Lim}(M)}. So F:SS2(Q2)Lim(M)F:\SS^{2}(Q_{2})\rightarrow\text{Lim}(M) is a bijective continuous function. Since SS2(Q2)\SS^{2}(Q_{2}) is compact, FF is a closed map. Hence, FF is a homeomorphism. ∎

Now, to see that QnSS2Q_{n}\cong\SS^{2} for n3n\geq 3, one first note that the length of J1J_{1} and J1J_{1}^{\prime} exceed 12\frac{1}{2} so there are extra side identifications on QnQ_{n} which cannot be seen on Q2Q_{2}. It can be described via J~1,J^1,Jˇ1\tilde{J}_{1}^{\prime},\hat{J}_{1}^{\prime},\check{J}_{1}^{\prime} and J~1,J^1,Jˇ1:\tilde{J}_{1},\hat{J}_{1},\check{J}_{1}^{*}:

J1~=J1([0,n22n]×{1}),J1~=R(J1~),J1^=J1((n22n,12]×{1}),J1^=R(J^1)and\tilde{J_{1}^{\prime}}=J_{1}^{\prime}\cap\Big{(}\Big{[}0,\frac{n-2}{2n}\Big{]}\times\{1\}\Big{)},\tilde{J_{1}}=R(\tilde{J_{1}^{\prime}}),\hat{J_{1}^{\prime}}=J_{1}^{\prime}\cap\Big{(}\Big{(}\frac{n-2}{2n},\frac{1}{2}\Big{]}\times\{1\}\Big{)},\hat{J_{1}}=R(\hat{J}_{1}^{\prime})\ \text{and}
Jˇ1=J1(J~1J^1),Jˇ1=J1(J~1J^1)which giveJ1=J~1J^1Jˇ1,J1=J~1J^1Jˇ1.\check{J}_{1}^{\prime}=J_{1}^{\prime}\setminus(\tilde{J}_{1}^{\prime}\cup\hat{J}_{1}^{\prime}),\check{J}_{1}=J_{1}\setminus(\tilde{J}_{1}\cup\hat{J}_{1})\ \text{which give}\ J_{1}=\tilde{J}_{1}\cup\hat{J}_{1}\cup\check{J}_{1},J_{1}^{\prime}=\tilde{J}_{1}^{\prime}\cup\hat{J}_{1}^{\prime}\cup\check{J}_{1}^{\prime}.
Refer to caption
Figure 7. Decomposition of J1J_{1} and J1J_{1}^{\prime} as J~1,J^1,Jˇ1\tilde{J}_{1},\hat{J}_{1},\check{J}_{1} and J~1,J^1,Jˇ1\tilde{J}_{1}^{\prime},\hat{J}_{1}^{\prime},\check{J}_{1}^{\prime}, respectively .

One see that QnQ_{n} can be described by identifying each part on the boundary of [0,0.5]×[0,1][0,0.5]\times[0,1] symmetrically to its middle point. There is an identification on J~1\tilde{J}_{1}^{\prime} with Jˇ1\check{J}_{1} which does not appear on Q2Q_{2}. In fact, one can regard QnQ_{n} as Q2Q_{2} with an additional strip between J~1\tilde{J}_{1}^{\prime} and Jˇ1\check{J}_{1} (see the Figure below).

Refer to caption
Figure 8. The quotient QnQ_{n} illustrated as a quotient of [0,0.5]×[0,1][0,0.5]\times[0,1]. It is homeomorphic to SS2(Qn),n3\SS^{2}(Q_{n}),n\geq 3 .

We build a corresponding space SS2(Qn)\SS^{2}(Q_{n}) of QnQ_{n} the same way of SS2(Q2)\SS^{2}(Q_{2}) and Q2.Q_{2}. Via the same method used for Q2Q_{2}, one then builds an inverse limit system which is homeomorphic to SS2\SS^{2}. Then using the universal property of the inverse limit of topological spaces, QnSS2.Q_{n}\cong\SS^{2}. We state this result as a Corollary to Proposition 3.2.1.

Corollary 3.2.2.1.

The space QnQ_{n} is homeomorphic to SS2,n3\SS^{2},n\geq 3

3.2.3. The existence of a dynamical system on QnQ_{n} induced from the system (Cn,Bn)(C_{n},B_{n})

We aim to verify here the existence of a dynamical system on QnQ_{n} induced from (Cn,Bn),n2(C_{n},B_{n}),n\geq 2. However, one realized from the previous subsection that Qn,n3,Q_{n},n\geq 3, has extra side identifications which do not appear in Q2Q_{2}. These extra side identifications complicate the existence proof of systems on QnQ_{n} for n3n\geq 3. Moreover, the proof for the case of Q2Q_{2} was already discussed in the talk given by Dr. Daniel Mayer (see the acknowledgement). We thus focus on these cases, and refer the case of Q2Q_{2} to the author dissertation (see [16]).

Let n3.n\geq 3. Each class ω=[(x,y)]=[(0.x1x2n,0.y1y2n)]\omega=[(x,y)]=[(0.x_{1}x_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n})] under R\sim_{R} can be written explicitly as:

ω={{(x,y)¯,R(x,y)¯}if(x,y)(0,1)2,{(0,0.y1y2y3n),(0,0.y1¯y2y3n)}if(x,y)I1,{(0,0.n¯n¯n¯ykyk+1n),(0,0.00yk^yk+1n)}if(x,y)Ik,k2{(0.x1x2x3n,0),(0.x1^x2x3n,0)}if(x,y)J1~J1~,{(0.x1x2x3n,1),(0.x1¯x2x3n,1)}if(x,y)J1^{(0.00xkxk+1n,0),(0.00xk^xk+1n,0)}if(x,y)Jk,k2\omega=\begin{cases}\{\overline{(x,y)},\overline{R(x,y)}\}&\mbox{if}\ (x,y)\in(0,1)^{2},\\ \{(0,0.y_{1}y_{2}y_{3}\ldots_{n}),(0,0.\bar{y_{1}}^{*}y_{2}^{*}y_{3}^{*}\ldots_{n})\}&\mbox{if}\ (x,y)\in I_{1},\\ \{(0,0.\bar{n}^{*}\bar{n}^{*}\ldots\bar{n}^{*}y_{k}^{*}y_{k+1}^{*}\ldots_{n}),(0,0.0\ldots 0\hat{y_{k}}y_{k+1}\ldots_{n})\}&\mbox{if}\ (x,y)\in I_{k},k\geq 2\\ \{(0.x_{1}x_{2}x_{3}..._{n},0),(0.\hat{x_{1}}^{*}x_{2}^{*}x_{3}^{*}..._{n},0)\}&\mbox{if}\ (x,y)\in\tilde{J_{1}}\cup\tilde{J_{1}^{\prime}},\\ \{(0.x_{1}^{*}x_{2}^{*}x_{3}^{*}\ldots_{n},1),(0.\bar{x_{1}}x_{2}x_{3}\ldots_{n},1)\}&\mbox{if}\ (x,y)\in\hat{J_{1}^{\prime}}\\ \{(0.0\ldots 0x_{k}^{*}x_{k+1}^{*}\ldots_{n},0),(0.0\ldots 0\hat{x_{k}}x_{k+1}\ldots_{n},0)\}&\mbox{if}\ (x,y)\in J_{k},k\geq 2\end{cases}

where a=a¯:=a1,a+=a^:=a+1,a:=(n1)aa^{-}=\bar{a}:=a-1,a^{+}=\hat{a}:=a+1,a^{*}:=(n-1)-a and [(0,0)]={(0,0)¯}.[(0,0)]=\{\overline{(0,0)}\}.

Refer to caption
Figure 9. The identification on the boundary of [0,0.5]×[0,1][0,0.5]\times[0,1] assigned by R\sim_{R} .

Observe that each J^1,Ik1\hat{J}_{1}^{\prime},I_{k-1} and JkJ_{k} is identified under R\sim_{R} by assigning points corresponding to each middle point (m^12,1),(0,lk12)\Big{(}\frac{\hat{m}_{1}}{2},1\Big{)},\Big{(}0,\frac{l_{k-1}}{2}\Big{)} and (mk2,0).\Big{(}\frac{m_{k}}{2},0\Big{)}.

Proposition 3.2.2.

The space CnC_{n} is a 2-to-1 branched-covering space of QnQ_{n} with countably infinite branched points {[(0,0)],[(0.5)],[(m1^/2,1)],[0,lk1/2],[(mk/2,0)]:k2}\{[(0,0)],[(0.5)],[(\hat{m_{1}}/2,1)],[0,l_{k-1}/2],[(m_{k}/2,0)]:k\geq 2\}. In fact, there exists a homeomorphism Tn:QnQnT_{n}:Q_{n}\rightarrow Q_{n} such that P^nBn=TnP^n\hat{P}_{n}\circ B_{n}=T_{n}\circ\hat{P}_{n} where P^n:CnQn\hat{P}_{n}:C_{n}\rightarrow Q_{n} is the quotient map.

Proof.

Let z=[(x,y)]=[(0.x1x2n,0.y1y2n)]Qn.z=[(x,y)]=[(0.x_{1}x_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n})]\in Q_{n}.
Assume first that 0<0.x1x2n<0.5,0<0.y1y2n<1.0<0.x_{1}x_{2}\ldots_{n}<0.5,0<0.y_{1}y_{2}\ldots_{n}<1.
Notice that ()R(i)=n1i=iandR(𝒱i)=𝒱n2i=𝒱i1=𝒱i¯for alli=0,,n1(*)\ \ R({\mathcal{E}}_{i})={\mathcal{E}}_{n-1-i}={\mathcal{E}}_{i^{*}}\ \mbox{and}\ R({\mathcal{V}}_{i})={\mathcal{V}}_{n-2-i}={\mathcal{V}}_{i^{*}-1}={\mathcal{V}}_{\bar{i^{*}}}\ \text{for all}\ i=0,...,n-1 where a¯=a1,a=(n1)a\bar{a}=a-1,a^{*}=(n-1)-a (recall notations i{\mathcal{E}}_{i}, 𝒱j{\mathcal{V}}_{j} used in Lemma 3.2.1).
Also note that P^n1({z})={(0.x1x2n,0.y1y2n),(0.x1x2n,0.y1y2n)}.\hat{P}_{n}^{-1}(\{z\})=\{(0.x_{1}x_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n}),(0.x_{1}^{*}x_{2}^{*}\ldots_{n},0.y_{1}^{*}y_{2}^{*}\ldots_{n})\}.
Since 0<0.x1x2n<0.50<0.x_{1}x_{2}\ldots_{n}<0.5, there exists 0kn120\leq k\leq\lfloor\frac{n-1}{2}\rfloor such that 0.x1x2nk𝒱k0.x_{1}x_{2}\ldots_{n}\in{\mathcal{E}}_{k}\cup{\mathcal{V}}_{k}.
This yields that 0.x1x2nk𝒱k¯.0.x_{1}^{*}x_{2}^{*}\ldots_{n}\in{\mathcal{E}}_{k^{*}}\cup{\mathcal{V}}_{\bar{k^{*}}}.
If (0.x1x2n,y)k,(0.x_{1}x_{2}\ldots_{n},y)\in{\mathcal{E}}_{k}, then

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(0.kx2n,0.y1y2n),(0.kx2n,0.y1y2n)}\displaystyle=\hat{P_{n}}B_{n}\{(0.kx_{2}\ldots_{n},0.y_{1}y_{2}\ldots_{n}),(0.k^{*}x_{2}^{*}\ldots_{n},0.y_{1}^{*}y_{2}^{*}\ldots_{n})\}
=Pn^{(0.x2x3n,0.ky1y2n),(0.x2x3n,0.ky1y2n)}\displaystyle=\hat{P_{n}}\{(0.x_{2}x_{3}\ldots_{n},0.ky_{1}y_{2}\ldots_{n}),(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.k^{*}y_{1}^{*}y_{2}^{*}\ldots_{n})\}
={[(0.x2x3n,0.ky1y2n)],[(0.x2x3n,0.ky1y2n)]}\displaystyle=\{[(0.x_{2}x_{3}\ldots_{n},0.ky_{1}y_{2}\ldots_{n})],[(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.k^{*}y_{1}^{*}y_{2}^{*}\ldots_{n})]\}
={[(0.x2x3n,0.ky1y2n)]}\displaystyle=\{[(0.x_{2}x_{3}\ldots_{n},0.ky_{1}y_{2}\ldots_{n})]\}

because [(0.x2x3n,0.ky1y2n)]=[(0.x2x3n,0.ky1y2n)][(0.x_{2}x_{3}\ldots_{n},0.ky_{1}y_{2}\ldots_{n})]=[(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.k^{*}y_{1}^{*}y_{2}^{*}\ldots_{n})].
If (0.x1x2n,y)=(k+1n,y)𝒱k,(0.x_{1}x_{2}..._{n},y)=(\frac{k+1}{n},y)\in{\mathcal{V}}_{k}, then

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(k+1n,y),(nk1n,1y)}\displaystyle=\hat{P_{n}}B_{n}\Big{\{}\Big{(}\frac{k+1}{n},y\Big{)},\Big{(}\frac{n-k-1}{n},1-y\Big{)}\Big{\}}
=Pn^{(0,y+k+1n),(0,nkyn)}\displaystyle=\hat{P_{n}}\Big{\{}\Big{(}0,\frac{y+k+1}{n}\Big{)},\Big{(}0,\frac{n-k-y}{n}\Big{)}\Big{\}}
={[(0,y+k+1n)],[(0,nkyn)]}.\displaystyle=\Big{\{}\Big{[}\Big{(}0,\frac{y+k+1}{n}\Big{)}\Big{]},\Big{[}\Big{(}0,\frac{n-k-y}{n}\Big{)}\Big{]}\Big{\}}.

Observe that y+k+1n+nkyn=1+1n=l1\frac{y+k+1}{n}+\frac{n-k-y}{n}=1+\frac{1}{n}=l_{1} where (0,l1/2)(0,l_{1}/2) is the middle point of I1I_{1} yielding that

[(0,y+k+1n)]=[(0,nkyn)].\Big{[}\Big{(}0,\frac{y+k+1}{n}\Big{)}\Big{]}=\Big{[}\Big{(}0,\frac{n-k-y}{n}\Big{)}\Big{]}.

Now assume that x=0x=0 and y=0.y1y2n(0,1)y=0.y_{1}y_{2}..._{n}\in(0,1) such that z=[(0,y)][(0,0)]z=[(0,y)]\neq[(0,0)].
Then (0,y)Ik={0}×(1nk,1nk1)(0,y)\in I_{k}=\{0\}\times\Big{(}\frac{1}{n^{k}},\frac{1}{n^{k-1}}\Big{)} for some k.k\in{\mathbb{N}}.
Note that

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(0,0.n¯n¯n¯ykyk+1n),[(0,0.00yk+yk+1n)}\displaystyle=\hat{P_{n}}B_{n}\{(0,0.\bar{n}^{*}\bar{n}^{*}\ldots\bar{n}^{*}y_{k}^{*}y_{k+1}^{*}\ldots_{n}),[(0,0.0\ldots 0y_{k}^{+}y_{k+1}\ldots_{n})\}
=Pn^{(0,0.0n¯n¯n¯ykyk+1n),(0,0.00yk+yk+1nn)}\displaystyle=\hat{P_{n}}\Big{\{}(0,0.0\bar{n}^{*}\bar{n}^{*}\ldots\bar{n}^{*}y_{k}^{*}y_{k+1}^{*}\ldots_{n}),\Big{(}0,\frac{0.0\ldots 0y_{k}^{+}y_{k+1}\ldots_{n}}{n}\Big{)}\Big{\}}
={[(0,0.0n¯n¯n¯ykyk+1n)]}\displaystyle=\{[(0,0.0\bar{n}^{*}\bar{n}^{*}\ldots\bar{n}^{*}y_{k}^{*}y_{k+1}^{*}\ldots_{n})]\}

because 0.n¯n¯n¯ykyk+1nn+0.00yk^yk+1nn=lkn=lk+1,\frac{0.\bar{n}^{*}\bar{n}^{*}\ldots\bar{n}^{*}y_{k}^{*}y_{k+1}^{*}\ldots_{n}}{n}+\frac{0.0\ldots 0\hat{y_{k}}y_{k+1}\ldots_{n}}{n}=\frac{l_{k}}{n}=l_{k+1}, (0,lk+1/2)(0,l_{k+1}/2) is the middle point of Ik+1I_{k+1}.
Next, let z=[(x,0)]z=[(x,0)] be such that (x,0)J1(x,0)\in J_{1} and x0.5x\leq 0.5 (i.e. (x,0)Jˇ1(x,0)\in\check{J}_{1}).
Observe the following facts

kn<x<k+1nnkn<1+1nx<nk+1n.\frac{k}{n}<x<\frac{k+1}{n}\rightarrow\frac{n-k}{n}<1+\frac{1}{n}-x<\frac{n-k+1}{n}.

This yields that

Pn^BnPn^1({z})=Pn^Bn{(x,0),(1+1nx,0)}=Pn^{(nxk,kn),(n(1+1nx)(nk),nkn)}=Pn^{(nxk,kn),(1+knx,nkn)}\begin{split}\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\})&=\hat{P_{n}}B_{n}\Big{\{}(x,0),\Big{(}1+\frac{1}{n}-x,0\Big{)}\Big{\}}\\ &=\hat{P_{n}}\Big{\{}\Big{(}nx-k,\frac{k}{n}\Big{)},\Big{(}n(1+\frac{1}{n}-x)-(n-k),\frac{n-k}{n}\Big{)}\Big{\}}\\ &=\hat{P_{n}}\Big{\{}\Big{(}nx-k,\frac{k}{n}\Big{)},\Big{(}1+k-nx,\frac{n-k}{n}\Big{)}\Big{\}}\\ \end{split}

because R(nxk,kn)=(1+knx,nkn).R(nx-k,\frac{k}{n})=(1+k-nx,\frac{n-k}{n}).
Next, let z=[(x,y)]z=[(x,y)] with (x,y)J1^(x,y)\in\hat{J_{1}}.
It is necessary to consider two cases.
Case 1: nn is even
Then z=[(x,y)]={(0.x1x2x3n,0),(0.x1¯x2x3n,0)}z=[(x,y)]=\{(0.x_{1}x_{2}x_{3}\ldots_{n},0),(0.\bar{x_{1}}^{*}x_{2}^{*}x_{3}^{*}\ldots_{n},0)\} where x1=n/2.x_{1}=n/2.

Refer to caption
Figure 10. Images of z1,z2z_{1},z_{2} under BnB_{n} where z1=(0.x1x2n,0),z2=(0.x1¯x2n,0)z_{1}=(0.x_{1}x_{2}\ldots_{n},0),z_{2}=(0.\bar{x_{1}}^{*}x_{2}^{*}\ldots_{n},0) .

Hence

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(n2n+j=2xjnj,0),((n1nn/21n)+j=2n1xjnj,0)}\displaystyle=\hat{P_{n}}B_{n}\Big{\{}\Big{(}\frac{n}{2n}+\sum_{j=2}^{\infty}\frac{x_{j}}{n^{j}},0\Big{)},\Big{(}\Big{(}\frac{n-1}{n}-\frac{n/2-1}{n}\Big{)}+\sum_{j=2}^{\infty}\frac{n-1-x_{j}}{n^{j}},0\Big{)}\Big{\}}
=P^n{(0.x2x3n,0.5),(0.x2x3n,0.5)}\displaystyle=\hat{P}_{n}\{(0.x_{2}x_{3}\ldots_{n},0.5),(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.5)\}
={[(0.x2x3n,0.5)],[(0.x2x3n,0.5)]}.\displaystyle=\{[(0.x_{2}x_{3}\ldots_{n},0.5)],[(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.5)]\}.

with the fact that j=1xj+1nj+j=1n1xj+1nj=j=1n1nj=1\sum_{j=1}^{\infty}\frac{x_{j+1}}{n^{j}}+\sum_{j=1}^{\infty}\frac{n-1-x_{j+1}}{n^{j}}=\sum_{j=1}^{\infty}\frac{n-1}{n^{j}}=1 which yields that

R(0.x2x3n,0.5)=(0.x2x3n,0.5).R(0.x_{2}x_{3}\ldots_{n},0.5)=(0.x_{2}^{*}x_{3}^{*}\ldots_{n},0.5).

Case 2: nn is odd
Then z=(x,y)={(0.x1x2x3n,0),(0.x1¯x2x3n,0)}z=(x,y)=\{(0.x_{1}x_{2}x_{3}\ldots_{n},0),(0.\bar{x_{1}}^{*}x_{2}^{*}x_{3}^{*}\ldots_{n},0)\} where x1=n12.x_{1}=\frac{n-1}{2}.

Refer to caption
Figure 11. Images of S1,S2S_{1},S_{2} under BnB_{n} where S1=(0.x1x2n,0),S2=(0.x1¯x2n,0)S_{1}=(0.x_{1}x_{2}\ldots_{n},0),S_{2}=(0.\bar{x_{1}}^{*}x_{2}^{*}\ldots_{n},0) .

Therefore

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(n12n+j=2xjnj,0),((n1nn121n)+j=2n1xjnj,0)}\displaystyle=\hat{P_{n}}B_{n}\Big{\{}\Big{(}\frac{n-1}{2n}+\sum_{j=2}^{\infty}\frac{x_{j}}{n^{j}},0\Big{)},\Big{(}\Big{(}\frac{n-1}{n}-\frac{\frac{n-1}{2}-1}{n}\Big{)}+\sum_{j=2}^{\infty}\frac{n-1-x_{j}}{n^{j}},0\Big{)}\Big{\}}
=Pn^Bn{(n12n+j=2xjnj,0),(n+12n+j=2n1xjnj,0)}\displaystyle=\hat{P_{n}}B_{n}\Big{\{}\Big{(}\frac{n-1}{2n}+\sum_{j=2}^{\infty}\frac{x_{j}}{n^{j}},0\Big{)},\Big{(}\frac{n+1}{2n}+\sum_{j=2}^{\infty}\frac{n-1-x_{j}}{n^{j}},0\Big{)}\Big{\}}
=P^n{(0.x2x3n,n12n),(0.x2x3n,n+12n)}\displaystyle=\hat{P}_{n}\Big{\{}\Big{(}0.x_{2}x_{3}\ldots_{n},\frac{n-1}{2n}\Big{)},\Big{(}0.x_{2}^{*}x_{3}^{*}\ldots_{n},\frac{n+1}{2n}\Big{)}\Big{\}}
={[(0.x2x3n,n12n)]}\displaystyle=\Big{\{}\Big{[}\Big{(}0.x_{2}x_{3}\ldots_{n},\frac{n-1}{2n}\Big{)}\Big{]}\Big{\}}

because [(0.x2x3n,n12n)]=[R(0.x2x3n,n12n)]=[(0.x2x3n,n+12n)].\Big{[}\Big{(}0.x_{2}x_{3}\ldots_{n},\frac{n-1}{2n}\Big{)}\Big{]}=\Big{[}R\Big{(}0.x_{2}x_{3}\ldots_{n},\frac{n-1}{2n}\Big{)}\Big{]}=\Big{[}\Big{(}0.x_{2}^{*}x_{3}^{*}\ldots_{n},\frac{n+1}{2n}\Big{)}\Big{]}.
Lastly, let z=[(x,0)]z=[(x,0)] where (x,0)Jk(x,0)\in J_{k} for some k2.k\geq 2.
In this case, z={(0.00xkxk+1n,0),(0.00xk+xk+1n,0)}z=\{(0.0\ldots 0x_{k}^{*}x_{k+1}^{*}\ldots_{n},0),(0.0\ldots 0x_{k}^{+}x_{k+1}\ldots_{n},0)\}.
Hence,

Pn^BnPn^1({z})\displaystyle\hat{P_{n}}B_{n}\hat{P_{n}}^{-1}(\{z\}) =Pn^Bn{(0.00xkxk+1n,0),(0.00xk+xk+1n,0)}\displaystyle=\hat{P_{n}}B_{n}\{(0.0\ldots 0x_{k}^{*}x_{k+1}^{*}\ldots_{n},0),(0.0\ldots 0x_{k}^{+}x_{k+1}\ldots_{n},0)\}
=Pn^{(0.xkxk+1nnk2,0),(0.xk+xk+1nnk2,0)}\displaystyle=\hat{P_{n}}\Big{\{}\Big{(}\frac{0.x_{k}^{*}x_{k+1}^{*}\ldots_{n}}{n^{k-2}},0\Big{)},\Big{(}\frac{0.x_{k}^{+}x_{k+1\ldots_{n}}}{n^{k-2}},0\Big{)}\Big{\}}
={[(0.xkxk+1nnk2,0)]}\displaystyle=\Big{\{}\Big{[}\Big{(}\frac{0.x_{k}^{*}x_{k+1}^{*}\ldots_{n}}{n^{k-2}},0\Big{)}\Big{]}\Big{\}}

because 0.xkxk+1nnk2+0.xk+xk+1nnk2=1+1nnk2=1nk2+1nk1=mk1\frac{0.x_{k}^{*}x_{k+1}^{*}\ldots_{n}}{n^{k-2}}+\frac{0.x_{k}^{+}x_{k+1}\ldots_{n}}{n^{k-2}}=\frac{1+\frac{1}{n}}{n^{k-2}}=\frac{1}{n^{k-2}}+\frac{1}{n^{k-1}}=m_{k-1} for k2k\geq 2.
Since BnB_{n} is a homeomorphism, we can conclude that there exists a homeomorphism
Tn:QnQnT_{n}:Q_{n}\rightarrow Q_{n} such that Pn^Bn=TnPn^.\hat{P_{n}}\circ B_{n}=T_{n}\circ\hat{P_{n}}.

As a summary, we have that Proposition 3.2.1, Corollary 3.2.2.1 together with Proposition 3.2.2 give Theorem A. Then the fact regarding density of periodic points stated in Lemma 3.2.1 yields Theorem B.

4. Metric entropy values of induced dynamical systems on the carpet

This section starts by listing steps we used to extend a given pair (Y,S)(Y,S) to a quadruple (Y,S,(Y),μY)(Y,S,{\mathcal{B}}(Y),\mu_{Y}). A more detailed discussion can be found in [5]. One notes that this method works regardless of the initial dynamical system (X,T)(X,T).
Step 0: In some cases, there is a pre-initial system (X~,T~,(X~),μ~)(\tilde{X},\tilde{T},{\mathcal{B}}(\tilde{X}),\tilde{\mu}) which provides the initial system (X,T)(X,T) a measure-theoretic structure via Lemma 2.1.1. If an initial system (X,T)(X,T) has a natural measure-theoretic structure, one skips this step.
Step 1: Assume that the initial system (X,T)(X,T) has a measure-theoretic structure (X,T,(X),μX)(X,T,{\mathcal{B}}(X),\mu_{X}). A system on the quotient sphere (𝒮0,H0)({\mathcal{S}}_{0},H_{0}) extends to (𝒮0,H0,(𝒮0),ν)({\mathcal{S}}_{0},H_{0},{\mathcal{B}}({\mathcal{S}}_{0}),\nu) through Lemma 2.1.1.
Step 2: There is the 00-th coordinate projection Π0:S(T)𝒮0,(zi)i0z0\Pi_{0}:S_{\infty}(T)\rightarrow{\mathcal{S}}_{0},(z_{i})_{i\geq 0}\mapsto z_{0}. Then a measure μ(A):=ν(Π0(AM))\mu(A):=\nu(\Pi_{0}(A\cap M_{\infty})) is a probability measure on S(T),S_{\infty}(T), where M={(ui)S:U0Π0(S(T))𝒪}.M_{\infty}=\{(u_{i})\in S_{\infty}:U_{0}\in\Pi_{0}(S_{\infty}(T))\setminus{\mathcal{O}}\}. The quadruple (S(T),H(T),(S(T)),μ)(S_{\infty}(T),H_{\infty}(T),{\mathcal{B}}(S_{\infty}(T)),\mu) is a dynamical system which is isomorphic to (𝒮0,H0,(𝒮0),ν)({\mathcal{S}}_{0},H_{0},{\mathcal{B}}({\mathcal{S}}_{0}),\nu). Since these two families have finite numbers of branch points

We now briefly discuss entropy values of H(FA)H_{\infty}(F_{A}) and H(Tλ)H_{\infty}(T_{\lambda}). Both families (𝕋2,FA)({\mathbb{T}}^{2},F_{A}) and (Xg,Tλ)(X_{g},T_{\lambda}) have their natural measure-theoretic structures induced from 2{\mathbb{R}}^{2}. It is well-known that their entropy with respect to their invariant measures are log(|λ|)\log(|\lambda^{\prime}|) and log(λ)\log(\lambda) where λ\lambda^{\prime} is the leading eigenvalue of FA.F_{A}. Since these two families have a finite numbers of branch points, as pointed out earlier in the discussion preceding Proposition 2.4.1, entropy values of systems on the quotient sphere are log(|λ|)\log(|\lambda^{\prime}|) and log(λ)\log(\lambda), respectively. Since each system on the carpet is isomorphic to its base system of the quotient sphere, log(|λ|)\log(|\lambda^{\prime}|) and log(λ)\log(\lambda) are respectively the entropy values of H(FA)H_{\infty}(F_{A}) and H(Tλ)H_{\infty}(T_{\lambda}). One notes that the crucial step is that Proposition 2.4.2 can be applied immediately to show that entropy values of FAF_{A} and TλT_{\lambda} are lower bounds of entropy values of homeomorphisms on the quotient spheres. This is due to the fact that they have finitely many branch points. This situation is slightly different for cases of (Cn,Bn)(C_{n},B_{n}) for n2n\geq 2, for which we now provide a proof.

Proposition 4.0.1.

Let n2n\geq 2. For a fixed probability vector P=(p0,p1,,pn1)P=(p_{0},p_{1},\ldots,p_{n-1}), there is a measure-theoretic dynamical system on (S,H)(S_{\infty},H_{\infty}) with an entropy value of i=0n1pilog(pi).-\sum_{i=0}^{n-1}p_{i}\log(p_{i}). As a consequence, every positive real number is realized as an entropy value of dynamical systems on S.S_{\infty}.

Proof.

One refers to [18] for the following facts: for each (Σn,σn)(\Sigma_{n},\sigma_{n}) and for each probability vector P=(p0,p1,,pn1)P=(p_{0},p_{1},\ldots,p_{n-1}) (i.e. 0pi<10\leq p_{i}<1 and i=0n1pi=1\sum_{i=0}^{n-1}p_{i}=1), there is a probability measure ρP\rho_{P} defined on (Σn){\mathcal{B}}(\Sigma_{n}) such that (Σn,σn,(Σn),ρP)(\Sigma_{n},\sigma_{n},{\mathcal{B}}(\Sigma_{n}),\rho_{P}) is a dynamical system. Moreover, hρP(σn)=i=0n1pilog(pi)h_{\rho_{P}}(\sigma_{n})=-\sum_{i=0}^{n-1}p_{i}\log(p_{i}). Note that Lemma 3.2.1 and Proposition 3.2.2 yield that (Qn,Tn)(Q_{n},T_{n}) is a factor of (Σn,σn)(\Sigma_{n},\sigma_{n}). So we have

hν(Tn)hρP(σn)=i=0n1pilog(pi)whereνis the push-forward measure.h_{\nu}(T_{n})\leq h_{\rho_{P}}(\sigma_{n})=-\sum_{i=0}^{n-1}p_{i}\log(p_{i})\ \text{where}\ \nu\ \text{is the push-forward measure}.

To obtain the reverse inequality, we note that there is a unique point with countably infinite preimage on Σn\Sigma_{n}, and all other points have finite preimage. Since the measure ρP\rho_{P} assigns a measure zero to any countably infinite set, using Proposition 2.4.2, we obtain the reverse inequality and get that hν(Tn)=i=0n1pilog(pi).h_{\nu}(T_{n})=-\sum_{i=0}^{n-1}p_{i}\log(p_{i}). Observe that

limN(N1Nlog(N1N)+log(N)N)=0andi=0N1log(N)N=log(N)N.\lim_{N\rightarrow\infty}\Big{(}-\frac{N-1}{N}\log\Big{(}\frac{N-1}{N}\Big{)}+\frac{\log(N)}{N}\Big{)}=0\ \text{and}\ \sum_{i=0}^{N-1}\frac{\log(N)}{N}=\log(N)\xrightarrow{N\rightarrow\infty}\infty.

By continuity and connectedness, for all N2,N\geq 2, [N1Nlog(N1N)+log(N)N,log(N)]E(N)\Big{[}-\frac{N-1}{N}\log\Big{(}\frac{N-1}{N}\Big{)}+\frac{\log(N)}{N},\log(N)\Big{]}\subseteq E(N) where E(N):={i=0N1pilog(pi)P=(p0,p1,,pN1)is a probability vector}E(N):=\{-\sum_{i=0}^{N-1}p_{i}\log(p_{i})\mid P=(p_{0},p_{1},\ldots,p_{N_{1}})\ \text{is a probability vector}\}\subseteq{\mathbb{R}}. So (0,)N2E(N).(0,\infty)\subseteq\cup_{N\geq 2}E(N).

5. The failure of Bowen specification property

We present in this section a proof that the systems on the carpet do not have the approximate product property (and hence they cannot have the Bowen specification). The proof here is a modification of the proof found in [5]. We discuss two key facts prior to giving the proof: the local behavior of points near a hyperbolic fixed point and the topology of the inverse limit.

5.1. The local behavior of orbits of points near a hyperbolic fixed point

Observe first that any fixed point of FAk,TλkF_{A}^{k},T_{\lambda}^{k} and BnkB_{n}^{k} is hyperbolic for any kk\in{\mathbb{N}}. Hence the behavior of orbits of points under iterations of FA,TλF_{A},T_{\lambda} or BnB_{n} in small neighborhoods of their hyperbolic fixed points are similar. In particular, the local behavior is analogous to the behavior of points under iterations of a hyperbolic linear map LM:22L_{M}:{\mathbb{R}}^{2}\rightarrow{\mathbb{R}}^{2} in a neighborhood of the origin.

Note that our definition of a hyperbolic linear map LML_{M} on 2{\mathbb{R}}^{2} is slightly different from the usual one.

Definition 5.1.1.

A linear map LM:22L_{M}:{\mathbb{R}}^{2}\rightarrow{\mathbb{R}}^{2} is called a hyperbolic linear map if LML_{M} is a linear transformation of the form LM(x,y)t=M(x,y)tL_{M}(x,y)^{t}=M(x,y)^{t} where MM is a 2×22\times 2 matrix with real entries such that both of its eigenvalues do not have modulus 11 and |det(M)|=1.|\text{det}(M)|=1.

Using diagonalization coordinates, let LML_{M} be a hyperbolic linear map of the form LM(x,y)=(λ1x,λy)L_{M}(x,y)=(\lambda^{-1}x,\lambda y) where 0<λ1<λ.0<\lambda^{-1}<\lambda. Then Proposition 5.1.1 describes an iteration pattern of a point in a neighborhood of the origin under LML_{M} in a form of the proportion of visiting a fixed region of interest. In particular, we choose a sufficiently small open square D=(λϵ,λϵ)22D=(-\lambda\epsilon,\lambda\epsilon)^{2}\subseteq{\mathbb{R}}^{2} and a specific region of interest DDD^{*}\subseteq D. Then, roughly, a consequence is that the proportion of the number of iterations of a point falling in DD^{*} over the total number of iterations of the point falling in DD cannot exceed 12.\frac{1}{2}. Observe that Proposition 5.1.1 is a formalization of the discussion stated in [5].

Proposition 5.1.1.

Let E=((λϵ,λϵ)×([ϵ,λϵ)(λϵ,ϵ]))(([ϵ,λϵ)(λϵ,ϵ]))×(λϵ,λϵ)).E=((-\lambda\epsilon,\lambda\epsilon)\times([\epsilon,\lambda\epsilon)\cup(-\lambda\epsilon,-\epsilon]))\cup(([\epsilon,\lambda\epsilon)\cup(-\lambda\epsilon,-\epsilon]))\times(-\lambda\epsilon,\lambda\epsilon)). Let z2z\in{\mathbb{R}}^{2} be such that there exist positive integers N,KN,K satisfying

  • LMN(z)D,LMN+K+1(z)DL_{M}^{N}(z)\notin D,L_{M}^{N+K+1}(z)\notin D and LMN+1(z)D1:=D1(E(,0))L_{M}^{N+1}(z)\in D_{1}^{*}:=D_{1}\setminus(E\cup({\mathbb{R}},0)) (see Figure 1212),

  • for all j=1,2,,K,LMN+j(z)D,j=1,2,\ldots,K,L_{M}^{N+j}(z)\in D,

then

nNK(z):=|{j=1,2,,K:LMN+j(z)D1}|K12.n_{N}^{K}(z):=\frac{|\{j=1,2,\ldots,K:L_{M}^{N+j}(z)\in D_{1}^{*}\}|}{K}\leq\frac{1}{2}.
Refer to caption
Figure 12. Subregions D1,D2,D3,D4D_{1},D_{2},D_{3},D_{4} and D1D_{1}^{*} inside D=(λϵ,λϵ)2D=(-\lambda\epsilon,\lambda\epsilon)^{2} .
Proof.

For simplicity, set β=λ1.\beta=\lambda^{-1}. Assume that z2z\in{\mathbb{R}}^{2} and N,KN,K\in{\mathbb{N}} such that

LMN(z)D,LMN+i(z)D1iK,LMN+K+1(z)D,LMN+1(z)=(p,q)with|p|>|q|>0.L_{M}^{N}(z)\notin D,L_{M}^{N+i}(z)\in D\ \forall 1\leq i\leq K,L_{M}^{N+K+1}(z)\notin D,L_{M}^{N+1}(z)=(p,q)\ \mbox{with}\ |p|>|q|>0.

Assume that (p,q)D1.(p,q)\in D_{1}^{*}. Then there exists an integer mm such that 0mK0\leq m\leq K,

βm|p|>λm|q|,βm+1|p|λm+1|q|<λϵ.\beta^{m}|p|>\lambda^{m}|q|,\beta^{m+1}|p|\leq\lambda^{m+1}|q|<\lambda\epsilon.

We also know that λm+1|q|<λϵ\lambda^{m+1}|q|<\lambda\epsilon since λϵλm+1|q|ϵλm|q|<βm|p||p|<λϵ(p,q)E.\lambda\epsilon\leq\lambda^{m+1}|q|\rightarrow\epsilon\leq\lambda^{m}|q|<\beta^{m}|p|\leq|p|<\lambda\epsilon\ \rightarrow(p,q)\in E. Next, λ2m|q|<λϵ\lambda^{2m}|q|<\lambda\epsilon because λ2m|q|λϵ\lambda^{2m}|q|\geq\lambda\epsilon implies that λϵ(βλ)mβmλm|q|<βmβm|p|=|p|\lambda\epsilon\leq(\beta\lambda)^{m}\beta^{-m}\lambda^{m}|q|<\beta^{-m}\beta^{m}|p|=|p| contradicting (p,q)D.(p,q)\in D. This in fact yields that LMN+i(z)D1,LMN+j(z)D2D4L_{M}^{N+i}(z)\in D_{1}^{*},L_{M}^{N+j}(z)\in D_{2}\cup D_{4} for all i=1,,m,j=m+1,,2m.i=1,...,m,j=m+1,...,2m. Therefore,

|{1jK:LMN+j(z)D1}|Km2m=12.\frac{|\{1\leq j\leq K:L_{M}^{N+j}(z)\in D_{1}^{*}\}|}{K}\leq\frac{m}{2m}=\frac{1}{2}.

Remark.

(1) Proposition 5.1.1 holds for the case the region D1D_{1}^{*} is changed to D3:=D3(E(,0))D_{3}^{*}:=D_{3}\setminus(E\cup({\mathbb{R}},0)) or any other subsets of D1D_{1}^{*} or D3D_{3}^{*}.
(2) We would like to point out a possible error in the discussion found in [5] on page 343343. Using the notation in [5], for each fixed region D=[ϵ,ϵ]2D=[-\epsilon,\epsilon]^{2}, there is a point z=(p,q)Dz=(p,q)\in D with |p||q||p|\geq|q| such that
the minimal mm satisfying am|p|bm|q|a^{m}|p|\geq b^{m}|q| and am+1|p|<bm+1|q|a^{m+1}|p|<b^{m+1}|q| is equal to zero. Our use of the region EE eliminates this exception.

5.2. The topology of an inverse limit

We now realize that all three families of initial systems (𝕋2,FA),({\mathbb{T}}^{2},F_{A}), (Xg,Tλ)(X_{g},T_{\lambda}) and (Cn,Bn)(C_{n},B_{n}) are such that all fixed points are hyperbolic and the induced dynamical systems on the carpet are topologized by the same topology. So the fact concerning the inverse limit topology in this subsection and the proof of the failure of approximate product property in the next subsection are presented on the pair (S,H)(S_{\infty},H_{\infty}) without the necessity to specify the initial system. In fact, we refer to the initial system ambiguously as the pair (X,T)(X,T).

Recall that we blew up the orbit O1O_{1} of the point z1z_{1} with the period length n1n_{1} to create (𝒮1,H1)({\mathcal{S}}_{1},H_{1}) in the first step. Set G=Hn1:SSG=H_{\infty}^{n_{1}}:S_{\infty}\rightarrow S_{\infty} and G0=H0n1:𝒮0𝒮0.G_{0}=H_{0}^{n_{1}}:{\mathcal{S}}_{0}\rightarrow{\mathcal{S}}_{0}. Assume that eigenvalues of the differential of Tn1T^{n_{1}} are 0<λ1<λ0<\lambda^{-1}<\lambda. The contracting direction at the fixed point z1z_{1} intersects the blown up circle S(z1)𝒮1S(z_{1})\subseteq{\mathcal{S}}_{1} in exactly 22 points, denoted them by v1v_{1} and v2v_{2}. Let z(i)z(i) denote the element (z1,vi,vi,)(z_{1},v_{i},v_{i},\ldots) in SS_{\infty}, for i=1,2i=1,2. By normalizing the diagonalization coordinate of Tn1T^{n_{1}} at z1z_{1}, we define a small open square region D=(λϵ,λϵ)2𝒮0D=(-\lambda\epsilon,\lambda\epsilon)^{2}\subseteq{\mathcal{S}}_{0} satisfying that it contains no images of branch points. We also define open subregion D1,D2,D3,D4D_{1},D_{2},D_{3},D_{4} of DD and their corresponding open regions D1,D2,D3,D4D_{1}^{\prime},D_{2}^{\prime},D_{3}^{\prime},D_{4}^{\prime} of 𝒮1{\mathcal{S}}_{1} such that π1(Dj)=Dj\pi_{1}(D_{j}^{\prime})=D_{j}. Lastly, we introduce four arcs C1,C2,C3C_{1},C_{2},C_{3} and C4C_{4} on the boundary circle S(z1)S(z_{1}). All regions D1,D2,D3,D4,D1,D2,D3,D4D_{1},D_{2},D_{3},D_{4},D_{1}^{\prime},D_{2}^{\prime},D_{3}^{\prime},D_{4}^{\prime} and arcs C1,C2,C3,C3C_{1},C_{2},C_{3},C_{3} are shown in Figure 13.

Refer to caption
Figure 13. Regions D1,D2,D3,D4D_{1},D_{2},D_{3},D_{4} in (λϵ,λϵ)2(-\lambda\epsilon,\lambda\epsilon)^{2}, and arcs C1,C2,C3,C4C_{1},C_{2},C_{3},C_{4} on the boundary circle of the fixed point z1z_{1} .

For each i0,i\geq 0, let Πi:j=0𝒮j𝒮i\Pi_{i}:\prod_{j=0}^{\infty}{\mathcal{S}}_{j}\rightarrow{\mathcal{S}}_{i} be the ii-coordinate projection map, (zj)j0zi(z_{j})_{j\geq 0}\mapsto z_{i}. Note that the product topology on i=0𝒮i\prod_{i=0}^{\infty}{\mathcal{S}}_{i} is metrizable with a metric d:=d:i=0𝒮i[0,)d:=d_{\infty}:\prod_{i=0}^{\infty}{\mathcal{S}}_{i}\rightarrow[0,\infty) defined by

d((xi),(yi))=j=0dj(xj,yj)2j(1+dj(xj,yj)).d((x_{i}),(y_{i}))=\sum_{j=0}^{\infty}\frac{d_{j}(x_{j},y_{j})}{2^{j}(1+d_{j}(x_{j},y_{j}))}.

The topology on the carpet SS_{\infty} is the subspace topology of i=0𝒮i\prod_{i=0}^{\infty}{\mathcal{S}}_{i}. An open ball centered at zz with radius r>0r>0 on SS_{\infty} is denoted by BS(z,r):={yS:d(z,y)<r}.B_{S}(z,r):=\{y\in S_{\infty}:d(z,y)<r\}.

Lemma 5.2.1.

For any sufficiently small r>0r>0, Π0(BS(z(1),r)D1.\Pi_{0}(B_{S}(z(1),r)\subseteq D_{1}. That is, the projection of an open ball BS(z(1),r)B_{S}(z(1),r) to 𝒮0{\mathcal{S}}_{0} is contained in the triangular region D1D_{1} for any small r>0r>0.

Refer to caption
Figure 14. For small r>0,r>0, the projection to 𝒮0{\mathcal{S}}_{0} of BS(z(1),r)B_{S}(z(1),r) is contained in D1{z1}D_{1}\cup\{z_{1}\} .
Proof.

Assume that supx,y𝒮1d1(x,y)<1.\mathrm{sup}_{x,y\in{\mathcal{S}}_{1}}d_{1}(x,y)<1. Notice that there is M>0M>0 such that

inf{d1(v1,x):xj=24(DjCj)}>M.\inf\{d_{1}(v_{1},x):x\in\cup_{j=2}^{4}(D_{j}^{\prime}\cup C_{j})\}>M.

Let 0<r<M40<r<\frac{M}{4}. Then for any y=(yi)Sy=(y_{i})\in S_{\infty} with y1D2D3D4C2C3C4,y_{1}\in D_{2}^{\prime}\cup D_{3}^{\prime}\cup D_{4}^{\prime}\cup C_{2}\cup C_{3}\cup C_{4},

d(z(1),y)d1(v1,y1)2(1+d1(v1,y1))>M4>r.d(z(1),y)\geq\frac{d_{1}(v_{1},y_{1})}{2(1+d_{1}(v_{1},y_{1}))}>\frac{M}{4}>r.

This yields that Π0(BS(z(1),r))D1{z1}\Pi_{0}(B_{S}(z(1),r))\subseteq D_{1}\cup\{z_{1}\} for all sufficiently small r>0.r>0.

Remark.

(1) Lemma 5.2.1 holds if z(1)z(1) is replaced by z(2)z(2). Precisely, for any sufficiently small r>0r>0, Π0(BS(z(1),r))D1andΠ0(BS(z(2)),r)D3.\Pi_{0}(B_{S}(z(1),r))\subseteq D_{1}\ \text{and}\ \Pi_{0}(B_{S}(z(2)),r)\subseteq D_{3}.
(2) Lemma 5.2.1 describes that a projection of an open ball of a small radius on the inverse limit S(FA)S_{\infty}(F_{A}) to the 0th0^{th}-coordinate is not an open ball centered at z1z_{1}. The projection is instead a triangular region contained completely in D1D_{1}.

5.3. The proof of the failure of the approximate product property

Theorem 5.3.1.

The system (S,H)(S_{\infty},H_{\infty}) does not have the approximate product property.

Proof.

Suppose that G=Hn1G=H_{\infty}^{n_{1}} has the approximate product property.
Let ϵ0>0\epsilon_{0}>0 be sufficiently small (Lemma 5.2.1). Let δ1=101\delta_{1}=10^{-1} and δ2=102.\delta_{2}=10^{-2}.
Let N=N(ϵ,δ1,δ2)N=N(\epsilon,\delta_{1},\delta_{2}) be the natural number and fix nmax{N,100}n\geq\max\{N,100\} with 100|n100|n.
Let (x(i))i𝒮(x(i))_{i\in{\mathbb{N}}}\subseteq{\mathcal{S}} where x(2i+1)=z(1)x(2i+1)=z(1) and x(2i+2)=z(2)x(2i+2)=z(2) for all i0i\in{\mathbb{N}}_{0}. Then there exist (hi)0(h_{i})\subseteq{\mathbb{N}}_{0} satisfying h1=0,nhi+1hin+n100h_{1}=0,n\leq h_{i+1}-h_{i}\leq n+\frac{n}{100} and y=(yi)Sy=(y_{i})\in S_{\infty} such that

|{0j<n:d(Ghi+j(y),Gj(x(i)))>ϵ0}|n10|\{0\leq j<n:d(G^{h_{i}+j}(y),G^{j}(x(i)))>\epsilon_{0}\}|\leq\frac{n}{10}

for each i.i\in{\mathbb{N}}. This yields that for each ii\in{\mathbb{N}},

Ghi+j(y)BS(Gj(x(i)),ϵ0)G^{h_{i}+j}(y)\in B_{S}(G^{j}(x(i)),\epsilon_{0})

for at least 9n10\frac{9n}{10} times among hi,hi+1,,hi+n1.h_{i},h_{i}+1,\ldots,h_{i}+n-1. This implies that

G0j(y0)=:Π0(Gj(y))Π0(BS(x(1),ϵ0))D1G_{0}^{j}(y_{0})=:\Pi_{0}(G^{j}(y))\in\Pi_{0}(B_{S}(x(1),\epsilon_{0}))\subseteq D_{1}

for at least 9n10\frac{9n}{10} times out of 0,1,,n1,0,1,\ldots,n-1, where G0=H0n1.G_{0}=H_{0}^{n_{1}}. Then we have that there exist 0m1<m20\leq m_{1}<m_{2} with m2m18n10m_{2}-m_{1}\geq\frac{8n}{10} and 0m3n100\leq m_{3}\leq\frac{n}{10} with m2+m3=n1m_{2}+m_{3}=n-1 such that

G0j(y0)D1for allj=m1,m1+1,,m2.G^{j}_{0}(y_{0})\in D_{1}\ \text{for all}\ j=m_{1},m_{1}+1,\ldots,m_{2}.

According to Proposition 5.1.1, G0j(y0)D2G^{j}_{0}(y_{0})\in D_{2} for all j=n,n+1,,n+7n10.j=n,n+1,\ldots,n+\frac{7n}{10}.
Since h2<n+n100h_{2}<n+\frac{n}{100} and D2D3=D_{2}\cap D_{3}=\emptyset, it must be that

|{0j<n:d(Gh2+j(y),x(2))}|>n10.|\{0\leq j<n:d(G^{h_{2}+j}(y),x(2))\}|>\frac{n}{10}.

Remark.

(1) Theorem 5.3.1 holds for a system (S,H)(S_{\infty},H_{\infty}) with eigenvalues λ<0\lambda<0 or λ1<0\lambda^{-1}<0. In such cases, we consider the system (S,H2k)(S_{\infty},H_{\infty}^{2k}) which satisfies that 0<λ2k<λ2k.0<\lambda^{-2k}<\lambda^{2k}.
(2) As we mentioned earlier, the proof presented in Theorem 5.3.1 is valid on (S,H)(S_{\infty},H_{\infty}) regardless of the initial system being (𝕋2,FA),(Xg,Tλ)({\mathbb{T}}^{2},F_{A}),(X_{g},T_{\lambda}) or (Cn,Bn)(C_{n},B_{n}).

The proof of Theorem 5.3.1 gets a contradiction from an argument on the behavior of points in the neighborhood of z(1)z(1) and z(2)z(2). These two points are representatives of the contracting direction on the boundary circle S(z1)S(z_{1}). Though the notion of the Bowen specification is usually defined on a compact metric space, it is an interesting question if we consider the invariant subset of SS_{\infty} which we exclude all points in 𝒪{\mathcal{O}} (so we exclude both z(1)z(1) and z(2)z(2)), does the specification hold on this subset?

5.4. The invariant subspace (S,H)(S^{*},H^{*}): does it have the specification ?

Let SS^{*} be a subset of SS_{\infty} defined by

S={(zi)i0S:z0𝒮0𝒪}.S^{*}=\{(z_{i})_{i\geq 0}\in S_{\infty}:z_{0}\in{\mathcal{S}}_{0}\setminus{\mathcal{O}}\}.

That is, SS^{*} is a subset of SS_{\infty} where we exclude all blown up boundary circles. Observe that H(𝒪)=𝒪H({\mathcal{O}})={\mathcal{O}} so SS^{*} is HH-invariant. This gives that (S,H)(S^{*},H^{*}) is a dynamical system where H:=H|S:SS.H^{*}:=H_{\infty}|_{S^{*}}:S^{*}\rightarrow S^{*}. We show that even (S,H)(S^{*},H^{*}) cannot have the Bowen specification property.

Before we proceed on, we note two facts. First, SS^{*} is not a compact metric space so it is not true that the Bowen specification implies the approximate product property. Second, the map HH^{*} is a restriction of HH_{\infty} so that Proposition 5.1.1 still can be applied. Hence, one needs to only investigate if there is a variant of Lemma 5.2.1 for HH^{*}.

Lemma 5.4.1.

There exists 0<δ<ϵ0<\delta<\epsilon such that for any Q=(qi)i0SQ=(q_{i})_{i\geq 0}\in S_{\infty} with q0=z1+(xQ,0)q_{0}=z_{1}+(x_{Q},0) where 0<xQ<ϵδ0<x_{Q}<\epsilon-\delta,

Π0(BS(Q,δ))D1.\Pi_{0}(B_{S}(Q,\delta))\subseteq D_{1}.
Proof.

Using the same notation as in the proof of Lemma 5.2.1, there is M>0M>0 such that

inf{d1(x,y):π1(x)=z1+(s,0)with 0<s<ϵ,yj=24(DjCj)}>M.\inf\{d_{1}(x,y):\pi_{1}(x)=z_{1}+(s,0)\ \text{with}\ 0<s<\epsilon,y\in\cup_{j=2}^{4}(D_{j}^{\prime}\cup C_{j})\}>M.

We then have that d(z,Q)>M4d(z,Q)>\frac{M}{4} for all zz with Π2(z)j=24(DjCj)\Pi_{2}(z)\in\cup_{j=2}^{4}(D_{j}^{\prime}\cup C_{j}) and Q=(qi)Q=(q_{i}) with q0=z1+(xQ,0)q_{0}=z_{1}+(x_{Q},0) where 0<xQ<ϵ.0<x_{Q}<\epsilon.

Refer to caption
Figure 15. Π0(BS(Q,δ))D1\Pi_{0}(B_{S}(Q,\delta))\subseteq D_{1} for small δ>0\delta>0 for any points on the contracting lines .

So for any small δ>0\delta>0, Π0(BS(Q,δ))D1\Pi_{0}(B_{S}(Q,\delta))\subseteq D_{1} for all Q=(qi)Q=(q_{i}) with q0=z1+(xQ,0),0<xQ<ϵδ.q_{0}=z_{1}+(x_{Q},0),0<x_{Q}<\epsilon-\delta.

Theorem 5.4.2.

The system (S,H)(S^{*},H^{*}) does not have the specification.

Proof.

Allow us to abuse notation and suppose that G:=(H)n1G:=(H^{*})^{n_{1}} has the specification.
Let δ>0\delta>0 be sufficiently small (according to Lemma 5.4.1).
Let N=N(δ)N=N(\delta) be the parameter for the specification of GG with respect to δ\delta.
Let Q(i)=(qj(i))SQ(i)=(q_{j}(i))\in S^{*} where q0(i)=z1+((1)i+1xQ(i),0)q_{0}(i)=z_{1}+((-1)^{i+1}x_{Q(i)},0) with 0<|xQ(i)|<ϵδ0<|x_{Q(i)}|<\epsilon-\delta for i=1,2.i=1,2.
Let 0=i1<j1=N+10=i_{1}<j_{1}=N+1 and i2=j2=2N+1.i_{2}=j_{2}=2N+1. Then there exists u=(ui)Su=(u_{i})\in S^{*} such that for each 1k21\leq k\leq 2,

Glk(u)BS(Glk(Q(k),δ))G^{l_{k}}(u)\in B_{S}(G^{l_{k}}(Q(k),\delta))

for lk[ik,jk].l_{k}\in[i_{k},j_{k}]. This yields that, H0l(u0)D1H_{0}^{l}(u_{0})\in D_{1} for all l=0,1,,N+1.l=0,1,\ldots,N+1.
So there exists MN+1M\geq N+1 such that

H0l(u0)D1D2D4H_{0}^{l}(u_{0})\in D_{1}\cup D_{2}\cup D_{4}

for all l=0,1,,2Ml=0,1,\ldots,2M. This is a contradiction because 2M>2N+1.2M>2N+1.

Remark.

Lemma 5.4.1 together with a slight modification of the proof of Theorem 5.3.1 yield that (S,H)(S^{*},H^{*}) does not have the approximate product property as well.

References

  • [1] Jan M. Aarts and Lex G. Oversteegen. The dynamics of the Sierpiński curve. Proc. Amer. Math. Soc., 120(3): 965–968, 1994.
  • [2] M. T. Barlow. Analysis on the Sierpinski carpet. In Analysis and geometry of metric measure spaces, volume 56 of CRM Proc. Lecture Notes, pages  27–53. Amer. Math. Soc., Providence, RI, 2013.
  • [3] Andrzej Biś, Hiromichi Nakayama, and Paweł  Walczak. Modelling minimal foliated spaces with positive entropy. Hokkaido Math. J., 36(2): 283–310, 2007.
  • [4] Corentin Boissy and Erwan Lanneau. Pseudo-Anosov homeomorphisms on translation surfaces in hyperelliptic components have large entropy. Geom. Funct. Anal., 22(1): 74–106, 2012.
  • [5] Jan P. Boroński and Piotr Oprocha. On dynamics of the Sierpiński carpet. C. R. Math. Acad. Sci. Paris, 356(3): 340–344, 2018.
  • [6] Rufus Bowen. Periodic points and measures for Axiom AA diffeomorphisms. Trans. Amer. Math. Soc., 154: 377–397, 1971.
  • [7] Morton Brown. Some applications of an approximation theorem for inverse limits. Proc. Amer. Math. Soc., 11: 478–483, 1960.
  • [8] R. Chamanara. Affine automorphism groups of surfaces of infinite type. In In the tradition of Ahlfors and Bers, III, volume 355 of Contemp. Math., pages  123–145. Amer. Math. Soc., Providence, RI, 2004.
  • [9] R. Chamanara, F. P. Gardiner, and N. Lakic. A hyperelliptic realization of the horseshoe and baker maps. Ergodic Theory Dynam. Systems, 26(6): 1749–1768, 2006.
  • [10] André de Carvalho and Toby Hall. Unimodal generalized pseudo-Anosov maps. Geometry & Topology, 8(3): 1127 – 1188, 2004.
  • [11] Robert L. Devaney. Cantor and Sierpinski, Julia and Fatou: complex topology meets complex dynamics. Notices Amer. Math. Soc., 51(1): 9–15, 2004.
  • [12] Benson Farb and Dan Margalit. A primer on mapping class groups, volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 2012.
  • [13] Albert Fathi, Francois Laudenbach, and Valentin Poénaru. Thurston’s work on surfaces, volume 48 of Mathematical Notes. Princeton University Press, Princeton, NJ, 2012. Translated from the 1979 French original by Djun M. Kim and Dan Margalit.
  • [14] Frédéric Faure, Sébastien Gouëzel, and Erwan Lanneau. Ruelle spectrum of linear pseudo-Anosov maps. J. Éc. polytech. Math., 6: 811–877, 2019.
  • [15] George K. Francis and Jeffrey R. Weeks. Conway’s ZIP proof. Amer. Math. Monthly, 106(5): 393–399, 1999.
  • [16] Worapan Homsomboon. Explicit Dynamical Systems on the Sierpiński Curve. PhD thesis, Oregon State University, 2022.
  • [17] Hisao Kato. The nonexistence of expansive homeomorphisms of Peano continua in the plane. Topology Appl., 34(2): 161–165, 1990.
  • [18] Anatole Katok and Boris Hasselblatt. Introduction to the modern theory of dynamical systems, volume 54 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1995. With a supplementary chapter by Katok and Leonardo Mendoza.
  • [19] Dominik Kwietniak, Martha Łacka, and Piotr Oprocha. A panorama of specification-like properties and their consequences. In Dynamics and numbers, volume 669 of Contemp. Math., pages  155–186. Amer. Math. Soc., Providence, RI, 2016.
  • [20] Francois Ledrappier and Peter Walters. A relativised variational principle for continuous transformations. J. London Math. Soc. (2), 16(3): 568–576, 1977.
  • [21] C.-E. Pfister and W. G. Sullivan. Large deviations estimates for dynamical systems without the specification property. Applications to the β\beta-shifts. Nonlinearity, 18(1): 237–261, 2005.
  • [22] Ch. Raghavendra, V. Saritha, and B. Alekhya. Design of modified sierpinski carpet fractal patch antenna for multiband applications. In The Institute of Electrical and Electronics Engineers, Inc. (IEEE) Conference Proceedings, page 868, Piscataway, 2017. The Institute of Electrical and Electronics Engineers, Inc. (IEEE).
  • [23] R. Clark Robinson. An introduction to dynamical systems: continuous and discrete. Pearson Prentice Hall, Upper Saddle River, NJ, 2004.
  • [24] Karl Sigmund. On dynamical systems with the specification property. Trans. Amer. Math. Soc., 190: 285–299, 1974.
  • [25] G. T. Whyburn. Topological characterization of the Sierpiński curve. Fund. Math., 45: 320–324, 1958.
  • [26] Anton Zorich. Flat surfaces. In Frontiers in number theory, physics, and geometry. I, pages  437–583. Springer, Berlin, 2006.