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

Smale regular and chaotic A-homeomorphisms and A-diffeomorphisms

Medvedev V.    Zhuzhoma E.111National Research University Higher School of Economics, 25/12 Bolshaya Pecherskaya, 603005, Nizhni Novgorod, Russia  222Corresponding author: [email protected]
(Dedicated to the memory of Stepin A.M.)
Abstract

We introduce Smale A-homeomorphisms that includes regular, semi-chaotic, chaotic, and super chaotic homeomorphisms of topological nn-manifold MnM^{n}, n2n\geq 2. Smale A-homeomorphisms contain A-diffeomorphisms (in particular, structurally stable diffeomorphisms) provided MnM^{n} admits a smooth structure. Regular A-homeomorphisms contain all Morse-Smale diffeomorphisms, while semi-chaotic and chaotic A-homeomorphisms contain A-diffeomorphisms with trivial and nontrivial basic sets. Super chaotic A-homeomorphisms contain A-diffeomorphisms whose basic sets are nontrivial. We describe invariant sets that determine completely dynamics of regular, semi-chaotic, and chaotic Smale A-homeomorphisms. This allows us to get necessary and sufficient conditions of conjugacy for these Smale A-homeomorphisms. We apply this necessary and sufficient conditions for structurally stable surface diffeomorphisms with arbitrary number of one-dimensional expanding attractors. We also use this conditions to get the complete classification of Morse-Smale diffeomorphisms on projective-like nn-manifolds for n=2,8,16n=2,8,16.

Key words and phrases: conjugacy, topological classification, Smale homeomorphism

Mathematics Subject Classification. Primary 37D05; Secondary 37B35

Introduction

Diffeomorphisms satisfying Smale’s axiom A (in short, A-diffeomorphisms) were introduced by Smale [44] as a magnificent and natural generalization of structurally stable diffeomorphisms. By definition, a non-wandering set of A-diffeomorphism has a uniform hyperbolic structure and is the topological closure of periodic orbits. Smale proved that the non-wandering set splits into closed, transitive, and invariant pieces called basic sets. A basic set is trivial, if it is an isolated periodic orbit. A good example of A-diffeomorphism with trivial basic sets is a Morse-Smale diffeomorphism [36, 43]. Such diffeomorphisms demonstrate regular dynamics. Due to Bowen [9], A-diffeomorphisms with nontrivial basic sets demonstrate chaotic dynamics since any such diffeomorphism has a positive entropy. The most familiar nontrivial basic sets are Plykin’s attractor [37] and codimension one expanding attractors introduced by Williams [46, 47]. Such basic sets appeared in various applications, see for example [15, 25, 45].

Taking in mind that there are manifolds that do not admit smooth structures [33], we introduce Smale A-homeomorphisms with non-wandering sets having a hyperbolic type (see a precise definition below). Such homeomorphisms naturally appear in topological dynamical systems. For example, in [11], it was proved the existence of topological Morse functions with three critical points on topological (including nonsmoothable) closed manifolds. Starting with these examples, one can construct topological (maybe, only topological) Morse-Smale flows and Morse-Smale homeomorphisms with the non-wandering set consisting of three fixed points of hyperbolic type. Deep theory of topological dynamical systems was developed in [1, 2].

The challenging problem in the Theory of Dynamical Systems is the classification up to conjugacy dynamical systems with regular and chaotic dynamics. Recall that homeomorphisms f1f_{1}, f2:MnMnf_{2}:M^{n}\to M^{n} are called conjugate, if there is a homeomorphism h:MnMnh:M^{n}\to M^{n} such that hf1=f2hh\circ f_{1}=f_{2}\circ h. To check whether given f1f_{1} and f2f_{2} are conjugate, one constructs usually an invariant of conjugacy which is a dynamical characteristic keeping under a conjugacy homeomorphism. Normally, such invariant is constructed in the frame of special class of dynamical systems. The famous invariant is Poincare’s rotation number for the class of transitive circle homeomorphisms [38]. This invariant is effective i.e. two transitive circle homeomorphisms are conjugate if and only if they have the same Poincare’s rotation number (see [35] and [5], ch. 7, concerning invariants of low dimensional dynamical systems). Anosov [3] and Smale [44] were first who realize the fundamental role of hyperbolicity for dynamical systems. Numerous topological invariants were constructed for special classes of A-diffeomorphisms including Anosov systems [12, 29, 34] and Morse-Smale systems, see the books [6, 13] and the surveys [14, 30].

In the frame of Smale A-homeomorphisms, we introduce regular, semi-chaotic, chaotic, and super chaotic homeomorphisms. We get necessary and sufficient conditions of conjugacy for regular, semi-chaotic, and chaotic Smale A-homeomorphisms on a closed topological nn-manifold MnM^{n}, n2n\geq 2. Automatically, this gives necessary and sufficient conditions of conjugacy for Morse-Smale diffeomorphisms and a wide class of A-diffeomorphisms with nontrivial basic sets provided MnM^{n} admits a smooth structure. We apply our conditions for structurally stable surface diffeomorphisms with arbitrary number of one-dimensional expanding attractors. We classify Morse-Smale diffeomorphisms with three periodic points on high-dimensional projective-like manifolds. Note that the projective-like manifolds were introduced by the authors in [32] (see also [31]).

Let us give the main definitions and formulate the main results. Later on, closNclos\,N means the topological closure of NN. In [32], the authors introduced the notation of equivalent embedding as follows. Let M1kM^{k}_{1}, M2kMnM^{k}_{2}\subset M^{n} be topologically embedded kk-manifolds, 1kn11\leq k\leq n-1. We say they have the equivalent embedding if there are neighborhoods U(closM1k)U(clos\,M^{k}_{1}), U(closM2k)U(clos\,M^{k}_{2}) of closM1kclos\,M^{k}_{1}, closM2kclos\,M^{k}_{2} respectively and a homeomorphism h:U(closM1k)U(closM2k)h:U(clos\,M^{k}_{1})\to U(clos\,M^{k}_{2}) such that h(M1k)=M2kh(M^{k}_{1})=M^{k}_{2}. This notation allows to classify Morse-Smale topological flows with non-wandering sets consisting of three equilibriums [32]. To be precise, it was proved that two such flows f1tf^{t}_{1}, f2tf^{t}_{2} are topologically equivalent if and only if the stable (or unstable) separatrices of saddles of f1tf^{t}_{1}, f2tf^{t}_{2} have the equivalent embedding. Remark that the notation of equivalent embedding goes back to a scheme introduced by Leontovich and Maier [26, 27] to attack the classification problem for flows on 2-sphere.

Solving the conjugacy problem for homeomorphisms, we have to add conjugacy relations to the equivalent embedding. The modification of (global) conjugacy is a local conjugacy when the conjugacy holds in some neighborhoods of compact invariant sets. We introduce the intermediate notion, so-called a locally equivalent dynamical embedding (in short, dynamical embedding), as follows. Let f1f_{1}, f2:MnMnf_{2}:M^{n}\to M^{n} be homeomorphisms of closed topological nn-manifold MnM^{n}, n2n\geq 2, and N1N_{1}, N2N_{2} invariant sets of f1f_{1}, f2f_{2} respectively i.e. fi(Ni)=Nif_{i}(N_{i})=N_{i}, i=1,2i=1,2. We say that the sets N1N_{1}, N2N_{2} have the same dynamical embedding if there are neighborhoods δ1\delta_{1}, δ2\delta_{2} of closN1clos\leavevmode\nobreak\ N_{1}, closN2clos\leavevmode\nobreak\ N_{2} respectively and a homeomorphism h0:δ1f1(δ1)Mnh_{0}:\delta_{1}\cup f_{1}(\delta_{1})\to M^{n} such that

h0(δ1)=δ2,h0(closN1)=closN2,h0f1|δ1=f2h0|δ1h_{0}(\delta_{1})=\delta_{2},\qquad h_{0}(clos\leavevmode\nobreak\ N_{1})=clos\leavevmode\nobreak\ N_{2},\qquad h_{0}\circ f_{1}|_{\delta_{1}}=f_{2}\circ h_{0}|_{\delta_{1}} (1)

Recall that F:LnLnF:L^{n}\to L^{n} is an A-diffeomorphism of smooth manifold LnL^{n} provided the non-wandering set NW(F)NW(F) is hyperbolic, and the periodic orbits of FF are dense in NW(F)NW(F) [44]. The hyperbolicity implies that every point z0NW(F)z_{0}\in NW(F) has the stable Ws(z0)W^{s}(z_{0}) and unstable Wu(z0)W^{u}(z_{0}) manifolds formed by points yLny\in L^{n} such that ϱL(Fpkz0,Fpky)0\varrho_{L}(F^{pk}z_{0},F^{pk}y)\to 0 as k+k\to+\infty and kk\to-\infty respectively, where ϱL\varrho_{L} is a metric on LnL^{n} [18, 19, 21, 23, 39, 44]. Moreover, Ws(z0)W^{s}(z_{0}) and Wu(z0)W^{u}(z_{0}) are homeomorphic (in the interior topology) to Euclidean spaces dimWs(z0)\mathbb{R}^{\dim W^{s}(z_{0})}, dimWu(z0)\mathbb{R}^{\dim W^{u}(z_{0})} respectively. Note that dimWs(z0)+dimWu(z0)=n\dim W^{s}(z_{0})+\dim W^{u}(z_{0})=n. The non-wandering set NW(F)NW(F) is a finite union of pairwise disjoint FF-invariant closed sets Ω1\Omega_{1}, ,Ωk\ldots,\Omega_{k} such that every restriction F|ΩiF|_{\Omega_{i}} is topologically transitive. These Ωi\Omega_{i} are called basic sets of FF. A basic set is nontrivial if it is not a periodic isolated orbit. Set Ws(u)(Ωi)=xΩiWs(u)(x)W^{s(u)}(\Omega_{i})=\cup_{x\in\Omega_{i}}W^{s(u)}(x). One says that Ωi\Omega_{i} is a sink (source) basic set provided Wu(Ωi)=ΩiW^{u}(\Omega_{i})=\Omega_{i} (Ws(Ωi)=ΩiW^{s}(\Omega_{i})=\Omega_{i}). A basic set Ωi\Omega_{i} is a saddle basic set if it is neither a sink nor a source basic set.

A homeomorphism f:MnMnf:M^{n}\to M^{n} is called a Smale A-homeomorphism if there is an A-diffeomorphism F:LnLnF:L^{n}\to L^{n} such that the non-wandering sets NW(f)NW(f), NW(F)NW(F) have the same dynamical embedding. As a consequence, NW(f)NW(f) is a finite union of pairwise disjoint ff-invariant closed sets Λ1\Lambda_{1}, ,Λk\ldots,\Lambda_{k} called basic sets of ff such that every restriction f|Λif|_{\Lambda_{i}} is topologically transitive. Each basic set Λ\Lambda has the stable manifold Ws(Λ)W^{s}(\Lambda), and the unstable manifold Wu(Λ)W^{u}(\Lambda). Similarly, one introduces the families of sink basic sets ω(f)\omega(f), and source basic sets α(f)\alpha(f), and saddle basic sets σ(f)\sigma(f).

A Smale A-homeomorphism ff is called regular if all basic sets ω(f)\omega(f), σ(f)\sigma(f), α(f)\alpha(f) are trivial.

A Smale A-homeomorphism ff is called semi-chaotic if exactly one family from the families ω(f)\omega(f), σ(f)\sigma(f), α(f)\alpha(f) consists of non-trivial basic sets.

A Smale A-homeomorphism ff is called chaotic if exactly two families from the families ω(f)\omega(f), σ(f)\sigma(f), α(f)\alpha(f) consists of non-trivial basic sets.

A Smale A-homeomorphism ff is called super chaotic if the families ω(f)\omega(f), σ(f)\sigma(f), α(f)\alpha(f) consists of non-trivial basic sets.

In Section 1, we represent examples of all types above of Smale A-homeomorphisms. Actually, all examples are A-diffeomorphisms.

Now let us introduce invariant sets that determine dynamics of Smale homeomorphisms. Given any Smale A-homeomorphism f:MnMnf:M^{n}\to M^{n}, denote by A(f)A(f) (resp., R(f)R(f)) the union of ω(f)\omega(f) (resp., α(f)\alpha(f)) and unstable (resp., stable) manifolds of saddle basic sets σ(f)\sigma(f) :

A(f)=ω(f)νσ(f)Wu(ν),R(f)=α(f)νσ(f)Ws(ν).A(f)=\omega(f)\bigcup_{\nu\in\sigma(f)}W^{u}(\nu),\quad R(f)=\alpha(f)\bigcup_{\nu\in\sigma(f)}W^{s}(\nu).

The following statement gives the necessary and sufficient conditions of conjugacy for three types of Smale A-homeomorphisms. Note that if ff is a regular or semi-chaotic Smale A-homeomorphism, then at least one of the families ω(f)\omega(f), α(f)\alpha(f) consists of trivial basic sets. However, if ff is a chaotic Smale A-homeomorphism, then it is possible a priori that the both ω(f)\omega(f) and α(f)\alpha(f) consist of nontrivial basic sets but σ(f)\sigma(f) consists of trivial basic sets.

Theorem 1

Let MnM^{n} be a closed topological nn-manifold MnM^{n}, n2n\geq 2 and fi:MnMnf_{i}:M^{n}\to M^{n} is either regular, or semi-chaotic, or chaotic Smale A-homeomorphism, i=1,2i=1,2. For chaotic fif_{i}, we suppose that either ω(fi)\omega(f_{i}) or α(fi)\alpha(f_{i}) consists of trivial basic sets, i=1,2i=1,2. Then the homeomorphisms f1f_{1}, f2f_{2} are conjugate if and only if one of the following conditions holds:

  • the basic sets α(f1)\alpha(f_{1}), α(f2)\alpha(f_{2}) are trivial while the sets A(f1)A(f_{1}), A(f2)A(f_{2}) have the same dynamical embedding;

  • the basic sets ω(f1)\omega(f_{1}), ω(f2)\omega(f_{2}) are trivial while the sets R(f1)R(f_{1}), R(f2)R(f_{2}) have the same dynamical embedding.

In Section 4, we apply Theorem 1 to consider the conjugacy for structurally stable surface diffeomorphisms M2M2M^{2}\to M^{2} with one-dimensional (orientable and non-orientable) attractors Λ1\Lambda_{1}, \ldots, Λk\Lambda_{k}, k2k\geq 2 (remark that a structurally stable diffeomorphism is an A-diffeomorphism [28]). We also use Theorem 1 to classify Morse-Smale diffeomorphisms with three periodic points on projective-like nn-manifolds for n{2,8,16}n\in\{2,8,16\}.

Refer to caption

Figure 1: (a) one isolated saddle and one expanding attractor on some non-oriented surface; (b) one isolated saddle and two Plykin attractors.

First, we prove the following statement interesting itself (note that a one dimensional expanding attractor is a trivial basic set).

Proposition 1

Let f:M2M2f:M^{2}\to M^{2} be an A-diffeomorphism with the non-wandering set NW(f)NW(f) consisting of one-dimensional expanding attractors Λ1\Lambda_{1}, \ldots, Λk\Lambda_{k}, and s00s_{0}\geq 0 isolated saddle periodic points, and arbitrary number of isolated nodal periodic orbits. Then ks0+1k\leq s_{0}+1.

The case k=s0=1k=s_{0}=1 is represented in Fig. 1, (a), while the case k=2k=2 and s0=1s_{0}=1 is represented in Fig. 1, (b) with two Plykin attractors. See also [7] where one got the estimate for the number of one-dimensional basic sets of surface A-diffeomorphisms depending on a genus of supporting surface.

The following statement shows that the dynamical embedding of unstable manifolds of isolated saddles (trivial basic sets) determine completely global dynamics of structurally stable surface diffeomorphisms with one-dimensional expanding attractors.

Theorem 2

Let fi:M2M2f_{i}:M^{2}\to M^{2}, i=1,2i=1,2, be a structurally stable diffeomorphism of closed 2-manifold M2M^{2} such that the spectral decomposition of fif_{i} consists of k2k\geq 2 one-dimensional expanding attractors Λ1(i)\Lambda_{1}^{(i)}, \ldots, Λk(i)\Lambda_{k}^{(i)}, and isolated source periodic orbits, and k1k-1 isolated saddle periodic points denoted by σ1(i)\sigma_{1}^{(i)}, \ldots, σk1(i)\sigma_{k-1}^{(i)}. Then f1f_{1}, f2f_{2} are conjugate if and only if the sets j=1j=k1Wu(σj(1))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(1)}), j=1j=k1Wu(σj(2))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(2)}) have the same dynamical embedding.

Denote by SRH(Mn)SRH(M^{n}) the class of Smale regular homeomorphisms MnMnM^{n}\to M^{n}. Note that it is possible that fSRH(Mn)f\in SRH(M^{n}) has the empty set σ(f)\sigma(f) of saddle periodic points. In this case the set α(f)\alpha(f) consists of a unique source and the set ω(f)\omega(f) consists of a unique sink, and Mn=SnM^{n}=S^{n} is an nn-sphere. Later on, we’ll assume that fSRH(Mn)f\in SRH(M^{n}) has a non-empty set σ(f)\sigma(f) of saddle periodic points.

Clearly, SRH(Mn)SRH(M^{n}) contains all Morse-Smale diffeomorphisms provided MnM^{n} admits a smooth structure. Note that the class SRH(Mn)SRH(M^{n}) is an essential extension of Morse-Smale diffeomorphisms for a Smale regular diffeomorphism can contain nonhyperbolic periodic points, tangencies, and separatrix connections.

As a consequence of Theorem 1, one gets the following statement (in particular, one gets the necessary and sufficient conditions of conjugacy for any Morse-Smale diffeomorphisms on smooth closed manifolds).

Corollary 1

Let MnM^{n} be a closed topological nn-manifold MnM^{n}, n2n\geq 2. Homeomorphisms f1f_{1}, f2SRH(Mn)f_{2}\in SRH(M^{n}) are conjugate if and only if one of the following conditions holds:

  • the sets A(f1)A(f_{1}), A(f2)A(f_{2}) have the same dynamical embedding;

  • the sets R(f1)R(f_{1}), R(f2)R(f_{2}) have the same dynamical embedding.

Denote by MS(Mn;a,b,c)MS(M^{n};a,b,c) the set of Morse-Smale diffeomorphisms f:MnMnf:M^{n}\to M^{n} whose non-wandering set consists of aa sinks, bb sources, and cc saddles. In [31], the authors proved that for MS(Mn;1,1,1)MS(M^{n};1,1,1) the only values of nn possible are n{2,4,8,16}n\in\{2,4,8,16\}. Moreover, the supporting manifolds for MS(Mn;1,1,1)MS(M^{n};1,1,1) are projective-like provided n{2,8,16}n\in\{2,8,16\} [31, 32].

First, to illustrate the applicability of Corollary 1, we consider very simple class MS(M2;1,1,1)MS(M^{2};1,1,1). In this case, the supporting manifold M2M^{2} is the projective plane M2=2M^{2}=\mathbb{P}^{2} [31]. Below, we define a type for a unique saddle of fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1). Using Corollary 1 we’ll show how to get the following complete classification of Morse-Smale diffeomorphisms MS(2,1,1,1)MS(\mathbb{P}^{2},1,1,1).

Proposition 2

Two diffeomorphisms f1f_{1}, f2MS(2,1,1,1)f_{2}\in MS(\mathbb{P}^{2},1,1,1) are conjugate if and only if the types of their saddles coincide. There are four types TiT_{i}, i=1,2,3,4i=1,2,3,4 of a saddle. Given any type TiT_{i}, i=1,2,3,4i=1,2,3,4, there is a diffeomorphism fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1) with a saddle σ(f)\sigma(f) of the type TiT_{i}.

Thus, up to conjugacy, there are four classes of Morse-Smale diffeomorphisms MS(2,1,1,1)MS(\mathbb{P}^{2},1,1,1).

The most essential application is a complete classification of Morse-Smale diffeomorphisms MS(M8;1,1,1)MS(M^{8};1,1,1) and MS(M16;1,1,1)MS(M^{16};1,1,1). The supporting 2k2k-manifolds for diffeomorphisms from the set MS2k(1,1,1)MS^{2k}(1,1,1) will be denoted by M2k(1,1,1)M^{2k}(1,1,1). Recall that SkS^{k} is a kk-sphere. Below, αf\alpha_{f}, σf\sigma_{f}, and ωf\omega_{f} mean the source, the saddle, and the sink of fMS2k(1,1,1)f\in MS^{2k}(1,1,1) respectively.

An embedding φ:SkM2k(1,1,1)\varphi:S^{k}\to M^{2k}(1,1,1) is called basic if

  • φ(Sk)\varphi(S^{k}) is a locally flat kk-sphere;

  • M2k(1,1,1)φ(Sk)M^{2k}(1,1,1)\setminus\varphi(S^{k}) is an open 2k2k-ball, M2k(1,1,1)=B2kφ(Sk)M^{2k}(1,1,1)=B^{2k}\sqcup\varphi(S^{k}).

It was proved in [31] that every supporting manifold M2k(1,1,1)M^{2k}(1,1,1), k=4,8k=4,8, admits a basic embedding.

Theorem 3

Let f:M2k(1,1,1)M2k(1,1,1)f:M^{2k}(1,1,1)\to M^{2k}(1,1,1) be a diffeomorphism from the set MS2k(1,1,1)MS^{2k}(1,1,1), k=4,8k=4,8. Then the following claims hold :

1) for any fMS2k(1,1,1)f\in MS^{2k}(1,1,1), there are basic embedding

φu(f):SkM2k(1,1,1),φs(f):SkM2k(1,1,1)\varphi_{u}(f):S^{k}\to M^{2k}(1,1,1),\qquad\varphi_{s}(f):S^{k}\to M^{2k}(1,1,1)

such that φu(f)(Sk)=Wσfu{ωf}\varphi_{u}(f)(S^{k})=W^{u}_{\sigma_{f}}\cup\{\omega_{f}\} and φs(f)(Sk)=Wσfs{αf}\varphi_{s}(f)(S^{k})=W^{s}_{\sigma_{f}}\cup\{\alpha_{f}\}.

2) given any basic embedding φ:SkM2k(1,1,1)\varphi:S^{k}\to M^{2k}(1,1,1), there is fMS2k(1,1,1)f\in MS^{2k}(1,1,1) such that one of the following equality holds:

φ(Sk)=Wσfu{ωf} or φ(Sk)=Wσfs{αf}\varphi(S^{k})=W^{u}_{\sigma_{f}}\cup\{\omega_{f}\}\,\,\mbox{ or }\,\,\varphi(S^{k})=W^{s}_{\sigma_{f}}\cup\{\alpha_{f}\}

3) two Morse-Smale diffeomorphisms f1,f2MS2k(1,1,1)f_{1},f_{2}\in MS^{2k}(1,1,1) are conjugate if and only if one of the following conditions holds:

  • the basic embedding φu(f1)(Sk)=Wσf1u{ωf1}\varphi_{u}(f_{1})(S^{k})=W^{u}_{\sigma_{f_{1}}}\cup\{\omega_{f_{1}}\}, φu(f2)(Sk)=Wσf2u{ωf2}\varphi_{u}(f_{2})(S^{k})=W^{u}_{\sigma_{f_{2}}}\cup\{\omega_{f_{2}}\} have the same dynamical embedding;

  • the basic embedding φs(f1)(Sk)=Wσf1s{αf1}\varphi_{s}(f_{1})(S^{k})=W^{s}_{\sigma_{f_{1}}}\cup\{\alpha_{f_{1}}\}, φs(f2)(Sk)=Wσf2s{αf2}\varphi_{s}(f_{2})(S^{k})=W^{s}_{\sigma_{f_{2}}}\cup\{\alpha_{f_{2}}\} have the same dynamical embedding.

Thus, every fMS2k(1,1,1)f\in MS^{2k}(1,1,1) corresponds the basic embedding φ(f):SkM2k(1,1,1)\varphi(f):S^{k}\to M^{2k}(1,1,1). Given any basic embedding φ\varphi, there is fMS2k(1,1,1)f\in MS^{2k}(1,1,1) such that φ(f)=φ\varphi(f)=\varphi. At last, a dynamical embedding of basic embedding defines completely a conjugacy class in MS2k(1,1,1)MS^{2k}(1,1,1). We see, that the set of basic embedding (up to isotopy) form the admissible set of conjugacy invariants for the Morse-Smale diffeomorphisms MS2k(1,1,1)MS^{2k}(1,1,1), k=4,8k=4,8. As to the class MS(M4;1,1,1)MS(M^{4};1,1,1), the existence of realizable and effective conjugacy invariant is still the open problem. The main reason is the possibility of wild embedding for topological closure of separatrix of a saddle [31].

The next statement shows that an equivalent embedding gives an invariant of conjugacy for iterations of Morse-Smale diffeomorphisms fMS2k(1,1,1)f\in MS^{2k}(1,1,1), k=4,8k=4,8.

Theorem 4

Let σfi\sigma_{f_{i}} be a unique saddle of fiMS2k(1,1,1)f_{i}\in MS^{2k}(1,1,1), i=1,2i=1,2. If the stable (unstable) manifolds Ws(u)(σf1)W^{s(u)}(\sigma_{f_{1}}), Ws(u)(σf2)W^{s(u)}(\sigma_{f_{2}}) have equivalent embedding, then there are k1k_{1}, k2k_{2}\in\mathbb{N} such that the diffeomorphisms f1k1f_{1}^{k_{1}}, f2k2f_{2}^{k_{2}} are conjugate. Vice versa, if f1k1f_{1}^{k_{1}}, f2k2f_{2}^{k_{2}} are conjugate for some k1k_{1}, k2k_{2}\in\mathbb{N}, then the stable (unstable) manifolds Ws(u)(σf1)W^{s(u)}(\sigma_{f_{1}}), Ws(u)(σf2)W^{s(u)}(\sigma_{f_{2}}) have equivalent embedding.

The structure of the paper is the following. In Section 2, we give some previous results. In Section 3, we prove Theorem 1. In Section 4, we prove Propositions 1, 2, and Theorems 2, 3, 4.

Acknowledgments. This work is supported by Laboratory of Dynamical Systems and Applications of National Research University Higher School of Economics, of the Ministry of science and higher education of the RF, grant ag. № 075-15-2019-1931.

1 Examples of A-diffeomorphisms

1) Regular A-diffeomorphisms. Obvious example of regular A-diffeomorphism is a Morse-Smale diffeomorphism. Note that there are regular A-diffeomorphisms that do not belong to the set of Morse-Smale diffeomorphisms. For example, they can belong to the boundary of the set of Morse-Smale diffeomorphisms in the space of diffeomorphisms. There are regular A-diffeomorphisms which can not be approximated by Morse-Smale diffeomorphisms [40].

2) Semi-chaotic A-diffeomorphisms. A good example of semi-chaotic diffeomorphism is a so-called DA-diffeomorphism obtained from Anosov automorphism after Smale surgery [44], see Fig. 2.

Refer to caption

Figure 2: Examples of semi-chaotic Smale diffeomorphisms.

A classical DA-diffeomorphism f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} contains a nontrivial attractor ω(f)\omega(f), trivial repeller α(f)\alpha(f), and empty set σ(f)\sigma(f). A generalized DA-diffeomorphism contains nonempty set σ(f)\sigma(f) [17]. Another example of semi-chaotic A-diffeomorphism is a classical Smale horseshoe gs:𝕊2𝕊2g_{s}:\mathbb{S}^{2}\to\mathbb{S}^{2}. Well-known that there is gsg_{s} with trivial attractor ω(gs)\omega(g_{s}) and repeller α(gs)\alpha(g_{s}), and nontrivial σ(gs)\sigma(g_{s}).

Starting with DA-diffeomorphisms, Williams [46] constructed an open domain 𝒩Diff1(𝕋2)\mathcal{N}\subset Diff^{1}(\mathbb{T}^{2}) consisting of structurally unstable diffeomorphisms. It is easy to see that 𝒩\mathcal{N} contains semi-chaotic A-diffeomorphisms.

One more example of semi-chaotic A-diffeomorphism is shown in Fig. 3 with a DA-attractor and Plykin attractor on a torus.

Refer to caption

Figure 3: One isolated saddle and two expanding attractors on a torus.

3) Chaotic A-diffeomorphisms. Take the classical DA-diffeomorphism f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} with the non-wandering set consisting of a source α\alpha and one-dimensional expanding attractor Λa\Lambda_{a}. The diffeomorphism f1f^{-1} defined on a copy 𝕋2\mathbb{T}^{2} has the non-wandering set consisting of a sink ω\omega and one-dimensional contracting repeller Λr\Lambda_{r}. Let us delete a small neighborhood UaU_{a} (resp., UsU_{s}) of α\alpha (resp., ω\omega) homeomorphic to a disk. Take an orientation reversing diffeomorphism h:UaUrh:\partial U_{a}\to\partial U_{r}. Then the surface M2=(𝕋2Ua)h(𝕋2Ur)M^{2}=\left(\mathbb{T}^{2}\setminus U_{a}\right)\cup_{h}\left(\mathbb{T}^{2}\setminus U_{r}\right) is a pretzel (closed orientable surface of genus 2). Following [42], one can construct A-diffeomorphism g:M2M2g:M^{2}\to M^{2} with the non-wandering set consisting of ΛaΛr\Lambda_{a}\cup\Lambda_{r} such that g|Λa=fg|_{\Lambda_{a}}=f and g|Λr=f1g|_{\Lambda_{r}}=f^{-1}. Thus, α(g)=Λr\alpha(g)=\Lambda_{r} and ω(g)=Λa\omega(g)=\Lambda_{a}. Clearly, gg is a chaotic A-diffeomorphism. Due to [42], there is a construction such that gg has a closed simple curve consisting of the tangencies of the invariant stable manifolds of Λa\Lambda_{a} and the invariant unstable manifolds of Λr\Lambda_{r}.

Another example one gets starting with a Smale solenoid [44], see Fig. 2. This mapping can be extended to Ω\Omega-stable diffeomorphism fs:M3M3f_{s}:M^{3}\to M^{3} with a one-dimensional expanding attractor, say Ω1\Omega_{1}, and one-dimensional contracting repeller, say Ω2\Omega_{2}, where M3M^{3} is a 3-sphere or lens space [8, 24]. This chaotic diffeomorphism is similar to the Robinson-Williams diffeomorphism gg considered above. There is a bifurcation of Ω1\Omega_{1} into a zero-dimensional saddle type basic set and isolated attracting periodic orbits [48]. As a result, one gets a chaotic Smale diffeomorphism f0:M3M3f_{0}:M^{3}\to M^{3} with trivial basic sets ω(f0)\omega(f_{0}), and the nontrivial source basic set α(f0)=Ω2\alpha(f_{0})=\Omega_{2}, and the nontrivial zero-dimensional saddle basic set σ(f0)\sigma(f_{0}).

4) Super chaotic A-diffeomorphisms. Let gs:𝕊2𝕊2g_{s}:\mathbb{S}^{2}\to\mathbb{S}^{2} be the classical Smale horseshoe and f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} the classical DA-diffeomorphism considered above. Delete small neighborhoods U1U_{1}, U2U_{2} of the sink ω(gs)\omega(g_{s}) and the source α(gs)\alpha(g_{s}) respectively each homeomorphic to a disk. There are reversing orientation diffeomorphisms h1:U1Uah_{1}:\partial U_{1}\to\partial U_{a} and h2:U2Urh_{2}:\partial U_{2}\to\partial U_{r}. Then the surface M2=(𝕋2Ua)h1(S2U1U2)h2(𝕋2Ur)M^{2}=\left(\mathbb{T}^{2}\setminus U_{a}\right)\bigcup_{h_{1}}\left(S^{2}\setminus U_{1}\cup U_{2}\right)\bigcup_{h_{2}}\left(\mathbb{T}^{2}\setminus U_{r}\right) is a pretzel. Similarly to Robinson-Williams’s method developed in [42], one can construct a diffeomorphism g0:M2M2g_{0}:M^{2}\to M^{2} with α(g0)=Λr\alpha(g_{0})=\Lambda_{r}, ω(g0)=Λa\omega(g_{0})=\Lambda_{a}, and σ(g0)\sigma(g_{0}) homeomorphic to the Smale horseshoe σ(gs)\sigma(g_{s}). Thus, g0g_{0} is a super chaotic A-diffeomorphism. By similar way, one can get another examples starting with semi-chaotic A-diffeomorphisms.

2 Properties of Smale homeomorphisms

We begin by recalling several definitions. Further details may be found in [5, 6, 44]. Denote by Orb(x)Orb(x) the orbit of point xMnx\in M^{n} under a homeomorphism f:MnMnf:M^{n}\to M^{n}. The ω\omega-limit set ω(x)\omega(x) of the point xx consists of the points yMny\in M^{n} such that fki(x)yf^{k_{i}}(x)\to y for some sequence kik_{i}\to\infty. Clearly that any points of Orb(x)Orb(x) have the same ω\omega-limit. Replacing ff with f1f^{-1}, one gets an α\alpha-limit set. Obviously, ω(x)α(x)NW(f)\omega(x)\cup\alpha(x)\subset NW(f) for every xMnx\in M^{n}.

Later on, fSsH(Mn)f\in SsH(M^{n}). Given a family C={c1,,cl}C=\{c_{1},\ldots,c_{l}\} of sets ciMnc_{i}\subset M^{n}, denote by C~\widetilde{C} the union c1clc_{1}\cup\ldots\cup c_{l}. It follows immediately from definitions that

NW(f)=α(f)~ω(f)~σ(f)~,fSsH(Mn)NW(f)=\widetilde{\alpha(f)}\cup\widetilde{\omega(f)}\cup\widetilde{\sigma(f)},\quad f\in SsH(M^{n}) (2)
Lemma 1

Let fSsH(Mn)f\in SsH(M^{n}) and xMnx\in M^{n}. Then

  1. 1.

    if ω(x)σ(f)~\omega(x)\subset\widetilde{\sigma(f)}, then xWs(σ)x\in W^{s}(\sigma_{*}) for some saddle basic set σσ(f)\sigma_{*}\in\sigma(f).

  2. 2.

    if α(x)σ(f)~\alpha(x)\subset\widetilde{\sigma(f)}, then xWu(σ)x\in W^{u}(\sigma_{*}) for some saddle basic set σσ(f)\sigma_{*}\in\sigma(f).

Proof. Suppose that ω(x)σ(f)~\omega(x)\subset\widetilde{\sigma(f)}. Since α(f)~\widetilde{\alpha(f)} and ω(f)~\widetilde{\omega(f)} are invariant sets, xα(f)~ω(f)~x\notin\widetilde{\alpha(f)}\cup\widetilde{\omega(f)}. Therefore, there are exist a neighborhood U(α)U(\alpha) of α(f)\alpha(f) and neighborhood U(ω)U(\omega) of ω(f)\omega(f) such that the positive semi-orbit Orb+(x)Orb^{+}(x) belongs to the compact set N=Mn(U(ω)U(α))N=M^{n}\setminus\left(U(\omega)\cup U(\alpha)\right). Let V(σ1)V(\sigma_{1}), \ldots, V(σm)V(\sigma_{m}) be pairwise disjoint neighborhoods of saddle basic sets σ1\sigma_{1}, \ldots, σm\sigma_{m} respectively such that i=1mV(σi)N\cup_{i=1}^{m}V(\sigma_{i})\subset N. Since every V(σi)V(\sigma_{i}) does not intersect jiV(σj)\cup_{j\neq i}V(\sigma_{j}) and all saddle basic sets are invariant, one can take the neighborhoods V(σ1)V(\sigma_{1}), \ldots, V(σm)V(\sigma_{m}) so small that every f(V(σi))f(V(\sigma_{i})) does not intersect jiV(σj)\cup_{j\neq i}V(\sigma_{j}). Suppose the contrary, i.e. there is no a unique saddle basic set σσ(f)\sigma_{*}\in\sigma(f) with xWs(σ)x\in W^{s}(\sigma_{*}). Thus, there are at least two different saddle basic sets σ1\sigma_{1}, σ2\sigma_{2} such that xWs(σ1)x\in W^{s}(\sigma_{1}) and xWs(σ2)x\in W^{s}(\sigma_{2}). Hence, ω(x)\omega(x) have to intersect σ1\sigma_{1}, σ2\sigma_{2}. It follows that the compact set N0=N(i=1mV(σi))N_{0}=N\setminus\left(\cup_{i=1}^{m}V(\sigma_{i})\right) contains infinitely many points of the semi-orbit Orb+(x)Orb^{+}(x). This implies ω(x)N0\omega(x)\cap N_{0}\neq\emptyset that contradicts (2). The second assertion is proved similarly. \Box

A set UU is a trapping region for ff if f(closU)intU.f\left(clos\leavevmode\nobreak\ U\right)\subset int\leavevmode\nobreak\ U. A set AA is an attracting set for ff if there exists a trapping set UU such that

A=k0fk(U).A=\bigcap_{k\geq 0}f^{k}(U).

A set AA^{*} is a repelling set for ff if there exists a trapping region UU for ff such that

A=k0fk(MnU).A^{*}=\bigcap_{k\leq 0}f^{k}(M^{n}\setminus U).

Another words, AA^{*} is an attracting set for f1f^{-1} with the trapping region MnUM^{n}\setminus U for f1f^{-1}. When we wish to emphasize the dependence of an attracting set AA or a repelling set AA^{*} on the trapping region UU from which it arises, we denote it by AUA_{U} or AUA^{*}_{U} respectively.

Let AA be an attracting set for ff. The basin B(A)B(A) of AA is the union of all open trapping regions UU for ff such that AU=AA_{U}=A. One can similarly define the notion of basin for a repelling set.

Let NN be an attracting or repelling set and B(N)B(N) the basin of NN. A closed set G(N)B(N)NG(N)\subset B(N)\setminus N is called a generating set for the domain B(N)NB(N)\setminus N if

B(N)N=kfk(G(N)).B(N)\setminus N=\cup_{k\in\mathbb{Z}}f^{k}\left(G(N)\right).

Moreover,

1) every orbit from B(N)NB(N)\setminus N intersects G(N)G(N); 2) if an orbit from B(N)NB(N)\setminus N intersects the interior of G(N)G(N), then this orbit intersects G(N)G(N) at a unique point; 3) if an orbit from B(N)NB(N)\setminus N intersects the boundary of G(N)G(N), then the intersection of this orbit with G(N)G(N) consists of two points; 4) the boundary of G(N)G(N) is the union of finitely many compact codimension one topological submanifolds.

Lemma 2

Let fSsH(Mn)f\in SsH(M^{n}).

1) Suppose that all basic sets α(f)\alpha(f) are trivial. Then α(f)~\widetilde{\alpha(f)} is a repelling set while A(f)A(f) is an attracting set with

B(α(f)~)α(f)~=B(A(f))A(f).B\left(\widetilde{\alpha(f)}\right)\setminus\widetilde{\alpha(f)}=B\left(A(f)\right)\setminus A(f).

Moreover,

  • there is a trapping region T(α)T(\alpha) for f1f^{-1} of the set α(f)~\widetilde{\alpha(f)} consisting of pairwise disjoint open nn-balls b1b_{1}, \ldots, brb_{r} such that each bib_{i} contains a unique periodic point from α(f)\alpha(f);

  • the regions B(α(f)~)α(f)~B(\widetilde{\alpha(f)})\setminus\widetilde{\alpha(f)}, B(A(f))A(f)B(A(f))\setminus A(f) have the same generating set G(α)G(\alpha) consisting of pairwise disjoint closed nn-annuluses a1a_{1}, \ldots, ara_{r} such that ai=closfpi(bi)bia_{i}=clos\leavevmode\nobreak\ f^{p_{i}}(b_{i})\setminus b_{i} where pip_{i}\in\mathbb{N} is a minimal period of a periodic point belonging to bib_{i}, i=1,,ri=1,\ldots,r :

    G(α)=i=1rai=i=1r(closfpi(bi)bi);G(\alpha)=\cup_{i=1}^{r}a_{i}=\cup_{i=1}^{r}\left(clos\leavevmode\nobreak\ f^{p_{i}}(b_{i})\setminus b_{i}\right);
  • B(A(f))A(f)=kfk(G(α))B(A(f))\setminus A(f)=\cup_{k\in\mathbb{Z}}f^{k}(G(\alpha)).

2) Suppose that all basic sets ω(f)\omega(f) are trivial. Then ω(f)~\widetilde{\omega(f)} is an attracting set while R(f)R(f) is a repelling set with

B(ω(f)~)ω(f)~=B(R(f))R(f).B(\widetilde{\omega(f)})\setminus\widetilde{\omega(f)}=B(R(f))\setminus R(f).

Moreover,

  • there is a trapping region T(ω)T(\omega) for ff of the set ω(f)~\widetilde{\omega(f)} consisting of pairwise disjoint an open nn-balls b1b_{1}, \ldots, blb_{l} such that each bib_{i} contains a unique periodic point from ω(f)\omega(f);

  • the regions B(ω(f)~)ω(f)~B(\widetilde{\omega(f)})\setminus\widetilde{\omega(f)}, B(R(f))R(f)B(R(f))\setminus R(f) have the same generating set G(ω)G(\omega) consisting of pairwise disjoint closed nn-annuluses a1a_{1}, \ldots, ala_{l} such that ai=biintfpi(bi)a_{i}=b_{i}\setminus int\leavevmode\nobreak\ f^{p_{i}}(b_{i}) where pip_{i}\in\mathbb{N} is a minimal period of a periodic point belonging to bib_{i}, i=1,,li=1,\ldots,l :

    G(ω)=i=1rai=i=1r(biintfpi(bi));G(\omega)=\cup_{i=1}^{r}a_{i}=\cup_{i=1}^{r}\left(b_{i}\setminus int\leavevmode\nobreak\ f^{p_{i}}(b_{i})\right);
  • B(R(f))R(f)=kfk(G(ω))B(R(f))\setminus R(f)=\cup_{k\in\mathbb{Z}}f^{k}(G(\omega)).

Proof. It is enough to prove the first statement only. Since all basic sets α(f)\alpha(f) are trivial and consists of locally hyperbolic source periodic points, there is a trapping region T(α)T(\alpha) for f1f^{-1} of the set α(f)~\widetilde{\alpha(f)} consisting of pairwise disjoint open nn-balls b1b_{1}, \ldots, brb_{r} such that each bib_{i} contains a unique periodic point qiq_{i} from α(f)\alpha(f) [36, 43]. Thus,

T(α)=i=1rbi,k0fkpi(bi)=qi,i=1,,r.T(\alpha)=\cup_{i=1}^{r}b_{i},\quad\cap_{k\leq 0}f^{kp_{i}}(b_{i})=q_{i},\quad i=1,\ldots,r.

As a consequence, there is the generating set G(α)=i=1r(closfpi(bi)bi)G(\alpha)=\cup_{i=1}^{r}\left(clos\leavevmode\nobreak\ f^{p_{i}}(b_{i})\setminus b_{i}\right) consisting of pairwise disjoint closed nn-annuluses ai=closfpi(bi)bia_{i}=clos\leavevmode\nobreak\ f^{p_{i}}(b_{i})\setminus b_{i}, i=1,,ri=1,\ldots,r.

Since the balls b1b_{1}, \ldots, brb_{r} are pairwise disjoint and closbifpi(bi)clos\leavevmode\nobreak\ b_{i}\subset f^{p_{i}}(b_{i}), the balls fp1(b1)f^{p_{1}}(b_{1}), \ldots, fpr(br)f^{p_{r}}(b_{r}) are pairwise disjoint also. For simplicity of exposition, we’ll assume that α(f)\alpha(f) consists of fixed points (otherwise, α(f)\alpha(f) is divided into periodic orbits each considered like a point). Therefore,

f(Mni=1rbi)=Mni=1rf(bi)Mni=1rclosbiint(Mni=1rbi).f\left(M^{n}\setminus\cup_{i=1}^{r}b_{i}\right)=M^{n}\setminus\cup_{i=1}^{r}f(b_{i})\subset M^{n}\setminus\cup_{i=1}^{r}clos\leavevmode\nobreak\ b_{i}\subset int\leavevmode\nobreak\ \left(M^{n}\setminus\cup_{i=1}^{r}b_{i}\right).

Hence, Mni=1rbiM^{n}\setminus\cup_{i=1}^{r}b_{i} is a trapping region for ff. Clearly, A(f)Mni=1rbiA(f)\subset M^{n}\setminus\cup_{i=1}^{r}b_{i}.

Take a point xMni=1rbix\in M^{n}\setminus\cup_{i=1}^{r}b_{i}. Obviously, ω(x)α(f)~\omega(x)\notin\widetilde{\alpha(f)}. It follows from (2) that ω(x)ω(f)~σ(f)~\omega(x)\in\widetilde{\omega(f)}\cup\widetilde{\sigma(f)}. By Lemma 1, ω(x)A(f)\omega(x)\in A(f). Therefore, A(f)A(f) is an attracting set with the trapping region Mni=1rbiM^{n}\setminus\cup_{i=1}^{r}b_{i} for ff :

A(f)=AMni=1rbi.A(f)=A_{M^{n}\setminus\cup_{i=1}^{r}b_{i}}.

Moreover,

Mn=α(f)~B(A(f))M^{n}=\widetilde{\alpha(f)}\cup B(A(f))

because of k0fk(bi)=qi\cap_{k\leq 0}f^{k}(b_{i})=q_{i}, i=1,,ri=1,\ldots,r.

Let us prove the quality B(α(f)~)α(f)~=B(A(f))A(f)B\left(\widetilde{\alpha(f)}\right)\setminus\widetilde{\alpha(f)}=B\left(A(f)\right)\setminus A(f). Take xB(α(f)~)α(f)~x\in B\left(\widetilde{\alpha(f)}\right)\setminus\widetilde{\alpha(f)}. Since xα(f)~x\notin\widetilde{\alpha(f)} and Mn=α(f)~B(A(f))M^{n}=\widetilde{\alpha(f)}\cup B(A(f)), xB(A(f))x\in B(A(f)). Since xB(α(f)~)x\in B\left(\widetilde{\alpha(f)}\right), α(x)α(f)\alpha(x)\subset\alpha(f). Hence, xA(f)x\notin A(f) and xB(A(f))A(f)x\in B(A(f))\setminus A(f). Now, set xB(A(f))A(f)x\in B(A(f))\setminus A(f). Then xα(f)x\notin\alpha(f). Since xA(f)x\notin A(f), α(x)σ(f)~α(f)~\alpha(x)\subset\widetilde{\sigma(f)}\cup\widetilde{\alpha(f)}. If one assumes that α(x)σ(f)~\alpha(x)\subset\widetilde{\sigma(f)}, then according to Lemma 1, xWu(ν)x\in W^{u}(\nu) for some saddle basic set ν\nu. Thus, xA(f)x\in A(f) which contradicts to xA(f)x\notin A(f). Therefore, α(x)α(f)~\alpha(x)\subset\widetilde{\alpha(f)}. Hence xB(α(f)~)x\in B\left(\widetilde{\alpha(f)}\right). As a consequence, xB(α(f)~)α(f)~x\in B\left(\widetilde{\alpha(f)}\right)\setminus\widetilde{\alpha(f)}.

The last assertion of the first statement follows from the previous ones. This completes the proof. \Box

In the next statement, we keep the notation of Lemma 2.

Lemma 3

Let fSsH(Mn)f\in SsH(M^{n}).

1) Suppose that all basic sets α(f)\alpha(f) are trivial. Then given any neighborhood V0(A)V_{0}(A) of A(f)A(f), there is n0n_{0}\in\mathbb{N} such that

kn0fk(G(α))V0(A)\cup_{k\geq n_{0}}f^{k}\left(G(\alpha)\right)\subset V_{0}(A)

where G(α)G(\alpha) is the generating set of the region B(α(f)~)α(f)~B(\widetilde{\alpha(f)})\setminus\widetilde{\alpha(f)}.

2) Suppose that all basic sets ω(f)\omega(f) are trivial. Then given any neighborhood V0(R)V_{0}(R) of R(f)R(f), there is n0n_{0}\in\mathbb{N} such that

kn0fk(G(ω))V0(R)\cup_{k\leq-n_{0}}f^{k}\left(G(\omega)\right)\subset V_{0}(R)

where G(ω)G(\omega) is the generating set of the region B(ω(f)~)ω(f)~B(\widetilde{\omega(f)})\setminus\widetilde{\omega(f)}.

Proof. It is enough to prove the first statement only. Take a closed tripping neighborhood UU of A(f)A(f) for ff. Since kfk(U)=A(f)V0(A)\cap_{k\in\mathbb{N}}f^{k}(U)=A(f)\subset V_{0}(A), there is k0k_{0}\in\mathbb{N} such that fk0(U)V0(A)f^{k_{0}}(U)\subset V_{0}(A). Clearly, fk0(U)f^{k_{0}}(U) is a tripping region of A(f)A(f) for ff. Hence, fk0+k(U)fk0(U)V0(A)f^{k_{0}+k}(U)\subset f^{k_{0}}(U)\subset V_{0}(A) for every kk\in\mathbb{N}.

Let G(α)G(\alpha) be a generating set of the region B(α(f)~)α(f)~B(\widetilde{\alpha(f)})\setminus\widetilde{\alpha(f)}. By Lemma 2, G(α)G(\alpha) is the generating set of the region B(A(f))A(f)B\left(A(f)\right)\setminus A(f) as well. Since G(α)G(\alpha) is a compact set, there is n0n_{0}\in\mathbb{N} such that fn0(G(α))fk0(U)f^{n_{0}}\left(G(\alpha)\right)\subset f^{k_{0}}(U). It follows that fn0+k(G(α))fk0+k(U)fk0(U)V0(A)f^{n_{0}+k}\left(G(\alpha)\right)\subset f^{k_{0}+k}(U)\subset f^{k_{0}}(U)\subset V_{0}(A) for every kk\in\mathbb{N}. As a consequence, kn0fk(G(α))V0(A)\cup_{k\geq n_{0}}f^{k}\left(G(\alpha)\right)\subset V_{0}(A). \Box

3 Proof of Theorem 1

Suppose that homeomorphisms f1f_{1}, f2SsH(Mn)f_{2}\in SsH(M^{n}) are conjugate. Since a conjugacy mapping MnMnM^{n}\to M^{n} is a homeomorphism, the sets A(f1)A(f_{1}), A(f2)A(f_{2}), as well as the sets R(f1)R(f_{1}), R(f2)R(f_{2}) have the same dynamical embedding.

To prove the inverse assertion, let us suppose for definiteness that the basic sets α(f1)\alpha(f_{1}), α(f2)\alpha(f_{2}) are trivial while the sets A(f1)A(f_{1}), A(f2)A(f_{2}) have the same dynamical embedding. Taking in mind that A(f1)A(f_{1}) and A(f2)A(f_{2}) are attracting sets, we see that there are neighborhoods δ1\delta_{1}, δ2\delta_{2} of A(f1)A(f_{1}), A(f2)A(f_{2}) respectively, and a homeomorphism h0:δ1δ2h_{0}:\delta_{1}\to\delta_{2} such that

h0f1|δ1=f2h0|δ1,f1(δ1)δ1,h0(A(f1))=A(f2).h_{0}\circ f_{1}|_{\delta_{1}}=f_{2}\circ h_{0}|_{\delta_{1}},\quad f_{1}(\delta_{1})\subset\delta_{1},\quad h_{0}(A(f_{1}))=A(f_{2}). (3)

Without loss of generality, one can assume that δ1B(A(f1))\delta_{1}\subset B(A(f_{1})). Moreover, taking δ1\delta_{1} smaller if one needs, we can assume that closδ1clos\leavevmode\nobreak\ \delta_{1} is a trapping region for f1f_{1} of the set A(f1)A(f_{1}). By (3), one gets

f2(closδ2)=f2h0(closδ1)=h0f1(closδ1)h0(δ1)=δ2.f_{2}(clos\leavevmode\nobreak\ \delta_{2})=f_{2}\circ h_{0}(clos\leavevmode\nobreak\ \delta_{1})=h_{0}\circ f_{1}(clos\leavevmode\nobreak\ \delta_{1})\subset h_{0}(\delta_{1})=\delta_{2}.

Thus, closδ2clos\leavevmode\nobreak\ \delta_{2} is a trapping region for f2f_{2} of the set A(f2)A(f_{2}). As a consequence, we get the following generalization of (3)

h0f1k|δ1=f2kh0|δ1,k,f1(closδ1)δ1,h0(A(f1))=A(f2).h_{0}\circ f^{k}_{1}|_{\delta_{1}}=f^{k}_{2}\circ h_{0}|_{\delta_{1}},\quad k\in\mathbb{N},\quad f_{1}(clos\leavevmode\nobreak\ \delta_{1})\subset\delta_{1},\quad h_{0}(A(f_{1}))=A(f_{2}). (4)

By Lemma 2, there is the trapping region T(α1)T(\alpha_{1}) for f11f^{-1}_{1} of the set α(f1)~\widetilde{\alpha(f_{1})} consisting of pairwise disjoint open nn-balls b1b_{1}, \ldots, brb_{r} such that each bib_{i} contains a unique periodic point qiq_{i} from α(f1)\alpha(f_{1}). In addition, the region B(α(f1)~)α(f1)~B(\widetilde{\alpha(f_{1})})\setminus\widetilde{\alpha(f_{1})} has the generating set G(α1)G(\alpha_{1}) consisting of pairwise disjoint closed nn-annuluses a1a_{1}, \ldots, ara_{r} such that ai=closf1pi(bi)bia_{i}=clos\leavevmode\nobreak\ f^{p_{i}}_{1}(b_{i})\setminus b_{i} where pip_{i}\in\mathbb{N} is a minimal period of the periodic point qiq_{i}.

Due to Lemma 3, one can assume without loss of generality that G(α1)=defG1δ1G(\alpha_{1})\stackrel{{\scriptstyle\rm def}}{{=}}G_{1}\subset\delta_{1}. Hence,

A(f1)(k0fk(G1))=A(f1)N+δ1,N+=k0fk(G1).A(f_{1})\bigcup\left(\cup_{k\geq 0}f^{k}(G_{1})\right)=A(f_{1})\bigcup N^{+}\subset\delta_{1},\quad N^{+}=\cup_{k\geq 0}f^{k}(G_{1}).

According to Lemma 2, G1G_{1} is a generating set of the region B(A(f1))A(f1)B(A(f_{1}))\setminus A(f_{1}). Let us show that h0(G1)=defG2h_{0}(G_{1})\stackrel{{\scriptstyle\rm def}}{{=}}G_{2} is a generating set for the region B(A(f2))A(f2)B(A(f_{2}))\setminus A(f_{2}). Take a point z2G2z_{2}\in G_{2}. There is a unique point z1G1z_{1}\in G_{1} such that h0(z1)=z2h_{0}(z_{1})=z_{2}. Note that z2A(f2)z_{2}\notin A(f_{2}) since z1A(f1)z_{1}\notin A(f_{1}). Since G1(B(A(f1))A(f1))G_{1}\subset\left(B(A(f_{1}))\setminus A(f_{1})\right), f1k(z1)A(f1)f_{1}^{k}(z_{1})\to A(f_{1}) as kk\to\infty. It follows from (4) that

f2k(z2)=f2kh0(z1)=h0f1k(z1)h0(A(f1))=A(f2) as k.f_{2}^{k}(z_{2})=f_{2}^{k}\circ h_{0}(z_{1})=h_{0}\circ f_{1}^{k}(z_{1})\to h_{0}(A(f_{1}))=A(f_{2})\quad\mbox{ as }\quad k\to\infty.

Hence, z2B(A(f2))z_{2}\in B(A(f_{2})) and G2B(A(f2))A(f2)G_{2}\in B(A(f_{2}))\setminus A(f_{2}).

Take an orbit Orbf2B(A(f2))A(f2)Orb_{f_{2}}\subset B(A(f_{2}))\setminus A(f_{2}). Since this orbit intersects a trapping region of A(f2)A(f_{2}), Orbf2δ2Orb_{f_{2}}\cap\delta_{2}\neq\emptyset. Therefore there exists a point x2Orbf2δ2x_{2}\in Orb_{f_{2}}\cap\delta_{2}. Since h0(A(f1))=A(f2)h_{0}(A(f_{1}))=A(f_{2}) and x2B(A(f2))A(f2)x_{2}\in B(A(f_{2}))\setminus A(f_{2}), the orbit Orbf1Orb_{f_{1}} of the point x1=h01(x2)δ1x_{1}=h_{0}^{-1}(x_{2})\subset\delta_{1} under f1f_{1} belongs to B(A(f1))A(f1)B(A(f_{1}))\setminus A(f_{1}). Hence, Orbf1Orb_{f_{1}} intersects G1G_{1} at some point w1δ1w_{1}\in\delta_{1}. Since x1x_{1}, w1Orbf1w_{1}\in Orb_{f_{1}}, there is kk\in\mathbb{N} such that either x1=f1k(w1)x_{1}=f_{1}^{k}(w_{1}) or w1=f1k(x1)w_{1}=f_{1}^{k}(x_{1}). Suppose for definiteness that w1=f1k(x1)w_{1}=f_{1}^{k}(x_{1}). Using (3), one gets

w2=h0(w1)=h0f1k(x1)=h0f1kh01(x2)=f2k(x2)G2Orbf2.w_{2}=h_{0}(w_{1})=h_{0}\circ f_{1}^{k}(x_{1})=h_{0}\circ f_{1}^{k}\circ h_{0}^{-1}(x_{2})=f_{2}^{k}(x_{2})\in G_{2}\cap Orb_{f_{2}}.

Similarly one can prove that if Orbf2Orb_{f_{2}} intersects the interior of G2G_{2}, then Orbf2Orb_{f_{2}} intersects G2G_{2} at a unique point, and if Orbf2Orb_{f_{2}} intersects the boundary of G2G_{2} then Orbf2Orb_{f_{2}} intersects G2G_{2} at two points. Thus, G2G_{2} is a generating set for the region B(A(f2))A(f2)B(A(f_{2}))\setminus A(f_{2}).

Set

k0fik(Gi)=defO(Gi),k0fik(Gi)=defO+(Gi),i=1,2.\cup_{k\geq 0}f^{-k}_{i}(G_{i})\stackrel{{\scriptstyle\rm def}}{{=}}O^{-}(G_{i}),\quad\cup_{k\geq 0}f^{k}_{i}(G_{i})\stackrel{{\scriptstyle\rm def}}{{=}}O^{+}(G_{i}),\quad i=1,2.

We see that O(Gi)O+(Gi)O^{-}(G_{i})\cup O^{+}(G_{i}) is invariant under fif_{i}, i=1,2i=1,2. Given any point xO(G1)O+(G1)x\in O^{-}(G_{1})\cup O^{+}(G_{1}), there is mm\in\mathbb{Z} such that xf1m(G1)x\in f^{-m}_{1}(G_{1}). Let us define the mapping

h:O(G1)O+(G1)O(G2)O+(G2)h:O^{-}(G_{1})\cup O^{+}(G_{1})\to O^{-}(G_{2})\cup O^{+}(G_{2})

as follows

h(x)=f2mh0f1m(x),wherexf1m(G1).h(x)=f^{-m}_{2}\circ h_{0}\circ f^{m}_{1}(x),\quad where\quad x\in f^{-m}_{1}(G_{1}).

Since G1G_{1} and G2G_{2} are generating sets, hh is correct. It is easy to check that

hf1|O(G1)O+(G1)=f2h|O(G1)O+(G1).h\circ f_{1}|_{O^{-}(G_{1})\cup O^{+}(G_{1})}=f_{2}\circ h|_{O^{-}(G_{1})\cup O^{+}(G_{1})}.

It follows from (3) that

h:A(f1)O(G1)O+(G1)A(f2)O(G2)O+(G2)h:A(f_{1})\cup O^{-}(G_{1})\cup O^{+}(G_{1})\to A(f_{2})\cup O^{-}(G_{2})\cup O^{+}(G_{2})

is the homeomorphic extension of h0h_{0} putting h|A(f1)=h0|A(f1)h|_{A(f_{1})}=h_{0}|_{A(f_{1})}. Moreover,

hf1k|A(f1)O(G1)O+(G1)=f2kh|A(f1)O(G1)O+(G1),k.h\circ f^{k}_{1}|_{A(f_{1})\cup O^{-}(G_{1})\cup O^{+}(G_{1})}=f^{k}_{2}\circ h|_{A(f_{1})\cup O^{-}(G_{1})\cup O^{+}(G_{1})},\quad k\in\mathbb{Z}.

By Lemma 2, GiG_{i} is a generating set for the region B(α(fi)~)α(fi)~=B(A(fi))A(fi)B\left(\widetilde{\alpha(f_{i})}\right)\setminus\widetilde{\alpha(f_{i})}=B\left(A(f_{i})\right)\setminus A(f_{i}) and B(A(fi))A(fi)=kfik(Gi)B\left(A(f_{i})\right)\setminus A(f_{i})=\cup_{k\in\mathbb{Z}}f^{k}_{i}(G_{i}), i=1,2i=1,2. Thus, one gets the conjugacy h:Mnα(f1)~Mnα(f2)~h:M^{n}\setminus\widetilde{\alpha(f_{1})}\to M^{n}\setminus\widetilde{\alpha(f_{2})} from f1|Mnα(f1)~f_{1}|_{M^{n}\setminus\widetilde{\alpha(f_{1})}} to f2|Mnα(f2)~f_{2}|_{M^{n}\setminus\widetilde{\alpha(f_{2})}} :

hf1k|Mnα(f1)~=f2kh|Mnα(f1)~,k.h\circ f^{k}_{1}|_{M^{n}\setminus\widetilde{\alpha(f_{1})}}=f^{k}_{2}\circ h|_{M^{n}\setminus\widetilde{\alpha(f_{1})}},\quad k\in\mathbb{Z}. (5)

Recall that the sets α(f1)\alpha(f_{1}), α(f2)\alpha(f_{2}) are periodic sources {αj(f1)}j=1l1\{\alpha_{j}(f_{1})\}_{j=1}^{l_{1}}, {αj(f2)}j=1l2\{\alpha_{j}(f_{2})\}_{j=1}^{l_{2}} respectively. By Lemma 2, the generating set GiG_{i} consists of pairwise disjoint nn-annuluses aj(fi)a_{j}(f_{i}), i=1,2i=1,2. Take an annulus ar(f1)=arG1a_{r}(f_{1})=a_{r}\subset G_{1} surrounding a source periodic point αr(f1))\alpha_{r}(f_{1})) of minimal period prp_{r}, 1rl11\leq r\leq l_{1}. Then the set k0f1kpr(ar){αr(f1))}=Dnr\bigcup_{k\geq 0}f_{1}^{-kp_{r}}(a_{r})\cup\{\alpha_{r}(f_{1}))\}=D^{n}_{r} is a closed nn-ball. Since

MnB(A(f2))=Mn(A(f2)kf2k(G2))M^{n}\setminus B(A(f_{2}))=M^{n}\setminus\left(A(f_{2})\cup_{k\in\mathbb{Z}}f_{2}^{k}(G_{2})\right)

consists of the source periodic points α(f2)\alpha(f_{2}), the annulus

k0f2kprh(ar)=k0hf1kpr(ar)=Dr\bigcup_{k\geq 0}f_{2}^{-kp_{r}}\circ h(a_{r})=\bigcup_{k\geq 0}h\circ f_{1}^{-kp_{r}}(a_{r})=D^{*}_{r}

surrounds a unique source periodic point αj(r)(f2)\alpha_{j(r)}(f_{2}) of the same minimal period prp_{r}. Moreover, Dr{αj(r)(f2)}D^{*}_{r}\cup\{\alpha_{j(r)}(f_{2})\} is a closed n-ball. It implies the one-to-one correspondence rj(r)r\to j(r) inducing the one-to-one correspondence j0:αr(f1))αj(r)(f2))j_{0}:\alpha_{r}(f_{1}))\to\alpha_{j(r)}(f_{2})). Since αr(f1))\alpha_{r}(f_{1})) and αj(r)(f2))\alpha_{j(r)}(f_{2})) have the same period, one gets

j0(f1k(αr(f1))=f2k(j0(αr(f1)))=f2k(αj(r)(f2)),0kpr.j_{0}\left(f_{1}^{k}(\alpha_{r}(f_{1})\right)=f_{2}^{k}\left(j_{0}(\alpha_{r}(f_{1}))\right)=f_{2}^{k}\left(\alpha_{j(r)}(f_{2})\right),\quad 0\leq k\leq p_{r}. (6)

Put by definition, h(αr(f1))=αj(r)(f2))h\left(\alpha_{r}(f_{1})\right)=\alpha_{j(r)}(f_{2})). For sufficiently large mm\in\mathbb{N}, the both f1mpr(Drn)f^{-mp_{r}}_{1}(D^{n}_{r}) and f2mpr(Dr)f^{-mp_{r}}_{2}(D^{*}_{r}) can be embedded in arbitrary small neighborhoods of αr(f1))\alpha_{r}(f_{1})) and αj(r)(f2))\alpha_{j(r)}(f_{2})) respectively, because of α(f1)~\widetilde{\alpha(f_{1})} and α(f2)~\widetilde{\alpha(f_{2})} are repelling sets. Taking in mind (6), it follows that h:MnMnh:M^{n}\to M^{n} is a conjugacy from f1f_{1} to f2f_{2}. This completes the proof. \Box

4 Some applications

Following Smale [43, 44], we write σ1σ2\sigma_{1}\succ\sigma_{2} provided Wu(σ1)Ws(σ2)W^{u}(\sigma_{1})\cap W^{s}(\sigma_{2})\neq\emptyset where σ1\sigma_{1} and σ2\sigma_{2} are saddle periodic points. Later on, we assume a surface M2M^{2} to be closed and connected. Recall that a node is either a sink or a source.

Proof of Proposition 1 is by induction on s0s_{0}. First, we consider the case s0=0s_{0}=0. We have to prove that k=1k=1. Suppose the contrary that is k2k\geq 2. According to [13, 37] (see also [16, 17]), there are disjoint open sets UiU_{i}, i=1,,ki=1,\ldots,k, such that each UiU_{i} is an attracting domain of Λi\Lambda_{i} with no trivial basic sets. Moreover, the boundary Ui\partial U_{i} consists of a finitely many simple closed curves. Therefore, M2i=1kUiM^{2}\setminus\cup_{i=1}^{k}U_{i} is the disjoint union j1Kj=G\cup_{j\geq 1}K_{j}=G of compact connected sets KjK_{j} where f1(G)Gf^{-1}(G)\subset G. Any iteration of ff has at least kk one-dimensional expanding attractors. Thus, without loss of generality, we can assume that f1(Kj)Kjf^{-1}(K_{j})\subset K_{j} for every KjK_{j}. In addition, one can assume that any periodic isolated point is fixed and the restriction of ff on every invariant manifold of saddle isolated point preserves orientation.

Since k2k\geq 2 and M2M^{2} is connected, there is a component of GG, say K1K_{1}, and different sets UlU_{l}, UrU_{r} such that K1Ul\partial K_{1}\cap\partial U_{l}\neq\emptyset and K1Ur\partial K_{1}\cap\partial U_{r}\neq\emptyset where UlUr=U_{l}\cap U_{r}=\emptyset. Any component of the boundary K1\partial K_{1} is a circle. We see that there are at least two components of K1\partial K_{1}. Let us glue a disk to each boundary component of K1\partial K_{1} to get a closed surface K~1\widetilde{K}_{1}. Since f1(K1)K1f^{-1}(K_{1})\subset K_{1}, one can extend f|K1f|_{K_{1}} to an A-diffeomorphism f~:K~1K~1\widetilde{f}:\widetilde{K}_{1}\to\widetilde{K}_{1} with a unique sink in each disk we glued. Note that by construction, the non-wandering set NW(f~)NW(\widetilde{f}) of f~\widetilde{f} consists of isolated nodal fixed points, and NW(f~)NW(\widetilde{f}) contains at least two sinks. According to [43], the surface K~1\widetilde{K}_{1} is the disjoint union of the stable manifolds of sinks and finitely many isolated sources (remark that the stable manifold of a source coincide with this source). This contradicts to the connectedness of K~1\widetilde{K}_{1} because of every stable manifold of a sink is homeomorphic to an open ball, and isolated sources do not separate the stable manifolds of two sinks. This contradiction proves that k=1k=1 provided s0=0s_{0}=0.

Suppose the statement holds for 0,,s0,\ldots,s saddles. We have to prove this statement for s0=s+1s_{0}=s+1 saddles. Recall that due to [43], the isolated saddles endowed with the Smale partial order \succ. Since now the set of isolated saddles is not empty, there is a minimal saddle, say σ\sigma. Then the topological closure of Ws(σ)W^{s}(\sigma) is either a segment II with the endpoints being two sources or a circle SS consisting of one source and Ws(σ)W^{s}(\sigma). In any cases, the both II and SS are repelling sets. Let us consider this cases.

The segment II has a neighborhood U(I)=UU(I)=U homeomorphic to a disk such that closUf(U)clos\,U\subset f(U). Note that σ\sigma is inside of UU. One can change ff inside of UU replacing closWs(σ)clos\,W^{s}(\sigma) by a unique source. One gets a diffeomorphism with kk expanding attractors and ss saddles. By the inductive assumption, ks+1s0<s0+1k\leq s+1\leq s_{0}<s_{0}+1.

Similarly, the circle SS has a neighborhood U(S)U(S) homeomorphic to an annulus such that closU(S)f(U(S))clos\,U(S)\subset f(U(S)). Note that σ\sigma belongs to U(S)U(S). The manifold M12=M2U(S)M^{2}_{1}=M^{2}\setminus U(S) has two boundary components M1M_{1}, M2M_{2} each homeomorphic to a circle. One can attach two disks D12D^{2}_{1}, D22D^{2}_{2} along their boundaries to M1M_{1}, M2M_{2} respectively to get a closed surface M~2\tilde{M}^{2}. This surface either is connected or consists of two connected surfaces. Since SS is a repelling set, one can extend ff on M~2\tilde{M}^{2} to get a diffeomorphism f~:M~2M~2\tilde{f}:\tilde{M}^{2}\to\tilde{M}^{2} with kk expanding attractors and ss saddles. If M~2\tilde{M}^{2} is connected then the inductive assumption implies ks+1s0k\leq s+1\leq s_{0}. Let us consider the case when M~2\tilde{M}^{2} consists of two connected closed surfaces M~12\tilde{M}_{1}^{2}, M~22\tilde{M}_{2}^{2}. Suppose that M~i2\tilde{M}_{i}^{2} contains kik_{i} expanding attractors and sis_{i} isolated saddles, i=1,2i=1,2. Obviously, k=k1+k2k=k_{1}+k_{2} and s1+s2=ss_{1}+s_{2}=s. By the inductive assumption, kisi+1k_{i}\leq s_{i}+1, i=1,2i=1,2. Hence, k(s1+1)+(s2+1)=s1+s2+2=s+2=s0+1k\leq(s_{1}+1)+(s_{2}+1)=s_{1}+s_{2}+2=s+2=s_{0}+1. This concludes the proof. \Box

Proof of Theorem 2. Let us consider a structurally stable diffeomorphism f:M2M2f:M^{2}\to M^{2} with the non-wandering set consisting of k2k\geq 2 one-dimensional expanding attractors Λ1\Lambda_{1}, \ldots, Λk\Lambda_{k}, and isolated source periodic orbits, and k1k-1 saddle periodic points σ1\sigma_{1}, \ldots, σk1\sigma_{k-1}. Each Λi\Lambda_{i} has a neighborhood UiU_{i} that is an attracting region of Λi\Lambda_{i}. Then M2(i=1k1Ui)M^{2}\setminus(\cup_{i=1}^{k-1}U_{i}) is the disjoint union G=j1KjG=\cup_{j\geq 1}K_{j} of compact connected sets where f1(G)Gf^{-1}(G)\subset G. Note that any positive iteration of ff has at least kk one-dimensional expanding attractors. Obviously, any iteration of ff has the same number k1k-1 of saddle periodic points. Due to Proposition 1, any positive iteration of ff has no more than kk one-dimensional expanding attractors. Hence, any positive iteration of ff has exactly the same number kk of expanding attractors. This implies that every attractor Λi\Lambda_{i} is CC-dense [4, 41]. As a consequence, each unstable manifold Wu()ΛiW^{u}(\cdot)\subset\Lambda_{i} is dense in Λi\Lambda_{i} [4, 13].

Take a connected component KK of the set GG. The boundary K\partial K is the disjoint union of circles c1c_{1}, \ldots. By construction, this circles belong to the boundaries of the attracting regions U1U_{1}, \ldots, UkU_{k}. Therefore, one can glue a disk djd_{j} to each circle cjc_{j} extending ff to djd_{j} with a sink inside of djd_{j}. If KK is without an isolated saddles, then KjdJK\cup_{j}d-J is a 2-sphere with a unique source and a unique sink [14]. Therefore if KK is without an isolated saddles, then KK is a disk with a unique source. Such a set KK we’ll call a disk with no saddles. Now, take a neighborhood UU of some Λi\Lambda_{i}. Suppose that all components of the boundary U\partial U attach to components of GG that are disks with no saddles. Then the union of UU and this disks gives a closed surface with exactly one expanding attractor Λi\Lambda_{i}. This contradicts to either the connectedness of M2M^{2} or the inequality k2k\geq 2. Thus, given any neighborhood UiU_{i} of Λi\Lambda_{i}, the boundary Ui\partial U_{i} has a common part with the boundary Kj\partial K_{j} of some component KGK\subset G which contains at least one isolated saddle.

Let KK be a component of GG containing a saddle σ\sigma and UU a neighborhood of some Λi\Lambda_{i} such that KU\partial K\cap\partial U\neq\emptyset. Let us show that Wu(σ)Ws(Λi)W^{u}(\sigma)\cap W^{s}(\Lambda_{i})\neq\emptyset. Suppose the contrary. We know that Wu(σ){σ}W^{u}(\sigma)\setminus\{\sigma\} belongs to stable manifolds of isolated periodic points lying in KK. Then there is a saddle σ1K\sigma_{1}\in K such that σ1σ\sigma_{1}\prec\sigma, and the topological closure of Ws(σ1)W^{s}(\sigma_{1}) is either a segment II with the endpoints being two sources or a circle SS consisting of one source and Ws(σ1)W^{s}(\sigma_{1}). In any cases, the both II and SS are repelling sets. Therefore, ff can be changed inside of KK so that a diffeomorphism obtained has k2k-2 isolated saddles and kk one-dimensional expanding attractors. This contradicts Proposition 1. Thus, Wu(σ)Ws(Λi)W^{u}(\sigma)\cap W^{s}(\Lambda_{i})\neq\emptyset.

Since ff is a structurally stable diffeomorphism, all intersections Wu(σ)Ws(x)W^{u}(\sigma)\cap W^{s}(x), xΛix\in\Lambda_{i}, are transversal. It follows from Wu(σ)Ws(Λi)W^{u}(\sigma)\cap W^{s}(\Lambda_{i})\neq\emptyset that there is xΛix\in\Lambda_{i} such that Wu(σ)W^{u}(\sigma) intersects transversally the stable manifold Ws(x)W^{s}(x). Recall that the attractor Λi\Lambda_{i} is CC-dense. Since any unstable manifold Wu()ΛiW^{u}(\cdot)\subset\Lambda_{i} is dense in Λi\Lambda_{i}, the topological closure of Wu(σ)W^{u}(\sigma) contains Λi\Lambda_{i}, closWu(σ)Λiclos\,W^{u}(\sigma)\supset\Lambda_{i}.

Clearly that if f1f_{1}, f2f_{2} are conjugate, then j=1j=k1Wu(σj(1))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(1)}), j=1j=k1Wu(σj(2))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(2)}) have the same dynamical embedding. Suppose that the sets j=1j=k1Wu(σj(1))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(1)}) and j=1j=k1Wu(σj(2))\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(2)}) have the same dynamical embedding. It follows from above that

clos(j=1j=k1Wu(σj(i)))j=1j=kΛj(i),i=1,2.clos\,\left(\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(i)})\right)\supset\cup_{j=1}^{j=k}\Lambda_{j}^{(i)},\quad i=1,2.

Since A(fi)=j=1j=k1Wu(σj(i))(j=1j=kΛj(i))A(f_{i})=\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(i)})\bigcup\left(\cup_{j=1}^{j=k}\Lambda_{j}^{(i)}\right), we see that

closA(f1)=clos(j=1j=k1Wu(σj(1))),closA(f2)=clos(j=1j=k1Wu(σj(2))).clos\,A(f_{1})=clos\,\left(\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(1)})\right),\quad clos\,A(f_{2})=clos\,\left(\cup_{j=1}^{j=k-1}W^{u}(\sigma_{j}^{(2)})\right).

Therefore, the sets A(f1)A(f_{1}), A(f2)A(f_{2}) have the same dynamical embedding. As a consequence of Theorem 1, we have that f1f_{1}, f2f_{2} are conjugate. This completes the proof. \Box

Consider fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1) with a unique saddle σ(f)\sigma(f). By definition, ff conjugates in some neighborhood of σ(f)\sigma(f) to a linear diffeomorphism with a saddle hyperbolic fixed point [39]. It easy to check that up to conjugacy there are exactly four such mappings :

T1={x¯=12xy¯=2y,T2={x¯=12xy¯=2y,T3={x¯=12xy¯=2y,T4={x¯=12xy¯=2y.T_{1}=\left\{\begin{array}[]{ccc}\bar{x}&=&\frac{1}{2}x\\ \bar{y}&=&2y,\end{array}\right.\qquad T_{2}=\left\{\begin{array}[]{ccc}\bar{x}&=&-\frac{1}{2}x\\ \bar{y}&=&2y,\end{array}\right.\qquad T_{3}=\left\{\begin{array}[]{ccc}\bar{x}&=&\frac{1}{2}x\\ \bar{y}&=&-2y,\end{array}\right.\qquad T_{4}=\left\{\begin{array}[]{ccc}\bar{x}&=&-\frac{1}{2}x\\ \bar{y}&=&-2y.\end{array}\right.

We’ll say that the saddle σ(f)\sigma(f) is of the type T1T_{1}, T2T_{2}, T3T_{3}, T4T_{4} respectively, see Fig. 4.

Refer to caption

Figure 4: Phase portrait for fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1): the diametrically opposite points are identified.

Proof of Proposition 2. Take fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1) with a unique saddle σ(f)=σ\sigma(f)=\sigma. The attracting set A(f)A(f) is a closed curve consisting of an unstable manifold Wu(σ)W^{u}(\sigma) of a unique saddle σ\sigma and a sink ω\omega. A neighborhood UU of A(f)A(f) is homeomorphic to a Möbius band. Since UU contains only two fixed points, the saddle σ\sigma and the sink ω\omega, the dynamics of f|Uf|_{U} depends completely on a local dynamics of ff at σ\sigma which is defined by one of the types TiT_{i}, i=1,2,3,4i=1,2,3,4. Due to Corollary 1, diffeomorphisms f1f_{1}, f2MS(2,1,1,1)f_{2}\in MS(\mathbb{P}^{2},1,1,1) are conjugate if and only if the types of their saddles coincide.

Choose any type Ti{T1,T2,T3,T4}T_{i}\in\{T_{1},T_{2},T_{3},T_{4}\}. Let BB be a Möbius band with the middle closed curve c0c_{0}. There is a mapping f0:BBf_{0}:B\to B with the attracting set c0c_{0} such that the non-wandering set of f0f_{0} consists of a hyperbolic sink ωc0\omega\in c_{0} and a hyperbolic saddle σc0\sigma\in c_{0} with Wu(σ)=c0{ω}W^{u}(\sigma)=c_{0}\setminus\{\omega\}. Note that the set 2B\mathbb{P}^{2}\setminus B is a 2-disk D2D^{2}. Since c0c_{0} is an attracting set, one can extend f0f_{0} to ff with a hyperbolic source in D2D^{2}. This gives fMS(2,1,1,1)f\in MS(\mathbb{P}^{2},1,1,1) desired. \Box

Proof of Theorem 3. 1) Since ff has a unique saddle, the both Wσfu{ωf}W^{u}_{\sigma_{f}}\cup\{\omega_{f}\} and Wσfs{αf}W^{s}_{\sigma_{f}}\cup\{\alpha_{f}\} are topologically embedded spheres denoted by Sk1S^{k_{1}} and Sk2S^{k_{2}} respectively. Due to [31], k1=k2=kk_{1}=k_{2}=k, and the complements M2k(1,1,1)(Wσfu{ωf})M^{2k}(1,1,1)\setminus\left(W^{u}_{\sigma_{f}}\cup\{\omega_{f}\}\right), M2k(1,1,1)(Wσfs{αf})M^{2k}(1,1,1)\setminus\left(W^{s}_{\sigma_{f}}\cup\{\alpha_{f}\}\right) homeomorphic to an open 2k2k-ball (see also, [32]). Thus, we have the embedding

φu(f):SkWσfu{ωf}M2k(1,1,1),φs(f):SkWσfs{αf}M2k(1,1,1).\varphi_{u}(f):S^{k}\to W^{u}_{\sigma_{f}}\cup\{\omega_{f}\}\subset M^{2k}(1,1,1),\quad\varphi_{s}(f):S^{k}\to W^{s}_{\sigma_{f}}\cup\{\alpha_{f}\}\subset M^{2k}(1,1,1).

Since the codimension of SkS^{k} equals k4k\geq 4, φu(f)(Sk)\varphi_{u}(f)(S^{k}) and φs(f)(Sk)\varphi_{s}(f)(S^{k}) are locally flat spheres [10]. Hence, φu(f)\varphi_{u}(f) and φs(f)\varphi_{s}(f) are basic embedding.

2) According to Théorèm d’approximation by Haefliger [20], we can assume without loss of generality that φ(Sk)\varphi(S^{k}) is a smoothly embedded kk-sphere. Hence, there is a tubular neighborhood T2kT^{2k} of φ(Sk)\varphi(S^{k}) that is the total space of locally trivial fiber bundle p:T2kφ(Sk)p:T^{2k}\to\varphi(S^{k}) with the base S0k=φ(Sk)S^{k}_{0}=\varphi(S^{k}) and a fiber kk-disk DkD^{k} [22]. Let ϑns:S0S0\vartheta_{ns}:S_{0}\to S_{0} be a Morse-Smale diffeomorphism with a unique sink ω0\omega_{0} and a unique source NN, so-called a "north-south"  diffeomorphism. The fiber p1(N)p^{-1}(N) is an open kk-disk. Let ψN:p1(N)p1(N)\psi_{N}:p^{-1}(N)\to p^{-1}(N) be the mapping with a unique hyperbolic sink at NN such that closψN(p1(N))ψN(p1(N))clos\,\psi_{N}(p^{-1}(N))\subset\psi_{N}(p^{-1}(N)) and j0ψNj(p1(N))={N}\cap_{j\geq 0}\psi^{j}_{N}(p^{-1}(N))=\{N\}. Since pp is a locally trivial fiber bundle, one can extend ψN\psi_{N} and ϑns\vartheta_{ns} to get the mapping f0:T2kT2kf_{0}:T^{2k}\to T^{2k} such that a) NN is a hyperbolic saddle with kk-dimensional local stable and unstable manifolds, and ω0\omega_{0} is a hyperbolic sink; b) given any point aT2kp1(N)a\in T^{2k}\setminus p^{-1}(N), f0l(a)f^{l}_{0}(a) tends to ω0\omega_{0} as ll\to\infty; moreover, S0=l0f0l(T2k)S_{0}=\cap_{l\geq 0}f^{l}_{0}(T^{2k}).

It was proved in [31] that the boundary T2k\partial T^{2k} of T2kT^{2k} is a (2k1)(2k-1)-sphere, say S2k1S^{2k-1}. Moreover, S2k1S^{2k-1} bounds the ball B2k=M2k(1,1,1)T2kB^{2k}=M^{2k}(1,1,1)\setminus T^{2k}. Take a point a0B2ka_{0}\in B^{2k}. Since B2kB^{2k} is a ball, one can extend f0f_{0} to B2kB^{2k} to get a mapping f:M2k(1,1,1)M2k(1,1,1)f:M^{2k}(1,1,1)\to M^{2k}(1,1,1) with a unique hyperbolic source at a0a_{0}. It follows from (a) and (b) that we get the desired Morse-Smale diffeomorphism fMS2k(1,1,1)f\in MS^{2k}(1,1,1) with the sink ω0=ωf\omega_{0}=\omega_{f}, the saddle N=σfN=\sigma_{f}, and the source a0=αfa_{0}=\alpha_{f}.

3) The last statement immediately follows from Corollary 1. \Box

Proof of Theorem 4. Obviously, if f1k1f_{1}^{k_{1}}, f2k2f_{2}^{k_{2}} are conjugate for some k1k_{1}, k2k_{2}\in\mathbb{N}, then the stable (unstable) manifolds Ws(u)(σf1)W^{s(u)}(\sigma_{f_{1}}), Ws(u)(σf2)W^{s(u)}(\sigma_{f_{2}}) have equivalent embedding. We have to prove the inverse assertion.

Suppose for definiteness that the unstable manifolds Wu(σf1)W^{u}(\sigma_{f_{1}}), Wu(σf2)W^{u}(\sigma_{f_{2}}) have equivalent embedding. Let Ti2kT^{2k}_{i} be a tubular neighborhood of Sik=Ws(σf1){ωi}S^{k}_{i}=W^{s}(\sigma_{f_{1}})\cup\{\omega_{i}\} that is the total space of locally trivial fiber bundle p:Ti2kSikp:T^{2k}_{i}\to S^{k}_{i} with the base SikS^{k}_{i} and a fiber kk-disk DikD^{k}_{i}, i=1,2i=1,2. Here, ωi\omega_{i} is a unique sink of fif_{i}, i=1,2i=1,2. Note that σf1\sigma_{f_{1}}, ωiSik\omega_{i}\subset S^{k}_{i} and SikS^{k}_{i} is an attracting set of fif_{i}, i=1,2i=1,2. It follows that there is kik_{i}\in\mathbb{N} such that fiki(Ti2k)Ti2kf_{i}^{k_{i}}(T^{2k}_{i})\subset T^{2k}_{i}. Moreover, without loss of generality one can assume that the restrictions

fiki|Ws(σfi):Ws(σfi)Ws(σfi),fiki|Ws(σfi):Wu(σfi)Wu(σfi),i=1,2f_{i}^{k_{i}}|_{W^{s}(\sigma_{f_{i}})}:W^{s}(\sigma_{f_{i}})\to W^{s}(\sigma_{f_{i}}),\quad f_{i}^{k_{i}}|_{W^{s}(\sigma_{f_{i}})}:W^{u}(\sigma_{f_{i}})\to W^{u}(\sigma_{f_{i}}),\quad i=1,2

preserve orientation. Taking the neighborhood Ti2kT^{2k}_{i} smaller if necessary, one assume that fikif_{i}^{k_{i}} near the saddle σfi\sigma_{f_{i}} conjugates a linear hyperbolic diffeomorphism due to the Grobman-Hartman Theorem [18, 19, 21] (see also [39]), i=1,2i=1,2. Hence, fikif_{i}^{k_{i}} is embedded into a flow near the saddle σfi\sigma_{f_{i}}, i=1,2i=1,2. Since fiki(Ti2k)Ti2kf_{i}^{k_{i}}(T^{2k}_{i})\subset T^{2k}_{i}, the both f1k1f_{1}^{k_{1}} and f2k2f_{2}^{k_{2}} are embedded into the flows, say f1tf^{t}_{1} and f2tf^{t}_{2}, in the neighborhoods T12kT^{2k}_{1}, T22kT^{2k}_{2} respectively. Clearly, the unstable manifolds Wu(σf1)W^{u}(\sigma_{f_{1}}), Wu(σf2)W^{u}(\sigma_{f_{2}}) are the unstable manifolds of f1tf^{t}_{1} and f2tf^{t}_{2}. Since Wu(σf1)W^{u}(\sigma_{f_{1}}) and Wu(σf2)W^{u}(\sigma_{f_{2}}) have equivalent embedding, it follows from the proof of Theorem 2 [32] that the flows f1tf^{t}_{1} and f2tf^{t}_{2} are conjugate. This implies that f1k1f_{1}^{k_{1}} and f2k2f_{2}^{k_{2}} are conjugate. \Box

Bibliography

  • [1] Akin E. The General Topology of Dynamical Systems. Graduate Studies in Math., v. 1(1993), AMS.
  • [2] Akin E., Hurley M., Kennedy J.A. Dynamics of Topologically Generic Homeomorphisms. Memoirs of AMS, v. 164(2003), no. 783.
  • [3] Anosov D.V. Geodesic flows on closed Riemannian manifolds of negative curvature. Trudy Mat. Inst. Steklov. 90 (1967); English transl., Proc. Steklov Inst. Math. 90 (1969).
  • [4] Anosov D.V. On one class of invariant sets of smooth dynamical systems. Proc. Int. Conf. ‘‘Nonlinear Oscillations Qualitative Methods, Kiev, Vol. 2, 1970, 39-45.
  • [5] Anosov D.V., Zhuzhoma E. Nonlocal asymptotic behavior of curves and leaves of laminations on covering surfaces. Proc. Steklov Inst. of Math., 249(2005), p. 221.
  • [6] Aranson S., Belitsky G., Zhuzhoma E. Introduction to Qualitative Theory of Dynamical Systems on Closed Surfaces. Translations of Math. Monographs, Amer. Math. Soc., 153(1996).
  • [7] Aranson S., Plykin R., Zhirov A., Zhuzhoma E. Exact upper bounds for the number of one-dimensional basic sets of surface A-diffeomorphisms. Journ. Dynamical and Control Systems, 3(1997), no 1, 3-18.
  • [8] Boju Jiang, Yi Ni, Shicheng Wang. 3-manifolds that admit knotted solenoids as attractors. Trans. Amer. Math. Soc., 356(2004), 4371-43-82.
  • [9] Bowen R. Topological entropy and Axiom A. Global Analysis, Proc. Sympos. Pure Math., Amer. Math. Soc., 14(1970), 23-42.
  • [10] Daverman R.J., Venema G.A. Embeddings in Manifolds. GSM, Amer. Math. Soc., Providence, 106(2009).
  • [11] Eells J., Kuiper N. Manifolds which are like projective planes. Publ. Math. IHES, 14(1962), 5-46.
  • [12] Franks J. Anosov diffeomorphisms. Global Analisys. Proc. Symp. in Pure Math., AMS 14(1970), 61-94.
  • [13] Grines V., Medvedev T., Pochinka O. Dynamical systems on 2- and 3-manifolds. Dev. Math., v. 46, Springer, 2016, 295p.
  • [14] Grines V., Gurevich E., Pochinka O., Zhuzhoma E. Classification of Morse-Smale systems and topological structure of the underlying manifolds. Russian Math. Surveys, 74(2019), no 1, 37-110.
  • [15] Grines V., Medvedev T., Pochinka O., Zhuzhoma E. On heteroclinic separators of magnetic fields in electrically conducting fluids. Physica D: Nonlinear Phenomena, 294(2015), 1-5.
  • [16] Grines V., Zhuzhoma E.. The topological classification of orientable attractors on an nn-dimensional torus. Russian Math. Surveys, 34 (1978), no. 4, 163–164.
  • [17] Grines V., Zhuzhoma E. On structurally stable diffeomorphisms with codimension one expanding attractors. Trans. Amer. Math. Soc., 357(2005), 617-667.
  • [18] Grobman D. On a homeomorphism of systems of differential equations. Dokl. Akad. Nauk, USSR, 128(1959), 5, 880-881.
  • [19] Grobman D. Topological classification of neighborhoods of singular point in nn-dimensional space. Matem. sbornik, USSR, 56(1962), 1, 77-94.
  • [20] Haefliger A. Plongements différentiable de variétés dans variétés. Comment. Math. Helv., 36(1961-1962), 47-82.
  • [21] Hartman P. On the local linearization of differential equations. Proc. AMS, 14(1963), no 4, 568-573.
  • [22] Hirch M. Differential Topology. Sringer-Verlag, 1976.
  • [23] Hirsch M., Pugh C., Shub M. Invariant Manifolds. Springer-Verlag (Lect. Nots Math.), 1977.
  • [24] Jiming Ma, Bin Yu. The realization of Smale solenoid type attractors in 3-manifolds. Topology and its Appl., 154(2007), 3021-3031.
  • [25] Kuznetsov S.P. Examples of a physical system with a hyperbolic attractor of the Smale-Williams type. Phys. Rev. Lett., 95(2005), 144101.
  • [26] Leontovich E.A., Maier A.G. On trajectories defining a qualitative structure of decomposition of sphere into trajectories. Doclady Acad. Sci. USSR, 14(1937), 5, 251-257.
  • [27] Leontovich E.A., Maier A.G. On the scheme defining the topological structure of decomposition into trajectories. Doclady Acad. Sci. USSR, 103(1955), 4, 557-560.
  • [28] Mañé R. A proof of C1C^{1} stability conjecture. Publ. Math. IHES 66(1988), 161-210.
  • [29] Manning A. There are no new Anosov diffeomorphisms on tori. Amer. Journ. of Math., 96(1974), 422-429.
  • [30] Medvedev V., Zhuzhoma E. Global dynamics of Morse-Smale systems. Proc. Steklov Inst. of Math., 261(2008), 112-135.
  • [31] Medvedev V., Zhuzhoma E. Morse-Smale systems with few non-wandering points. Topology and its Applications, 160(2013), issue 3, 498-507.
  • [32] Medvedev V., Zhuzhoma E. Continuous Morse-Smale flows with three equilibrium positions. Sbornik: Mathematics, 207(2016), 5, 702–723.
  • [33] Milnor J. On manifolds homeomorphic to the 7-sphere. Annals of Math., 64(1956), no 2, 399-405.
  • [34] Newhouse S. On codimension one Anosov diffeomorphisms. Amer. J. of Math., 92(1970), no 3, 761-770.
  • [35] Nikolaev I., Zhuzhoma E. Flows on 2-dimensional manifolds. Lect. Notes in Math., 1705, Springer 1999.
  • [36] Palis J. On Morse-Smale dynamical systems. Topology, 8(1969), no 4, 385-404.
  • [37] Plykin R.V. On the geometry of hyperbolic attractors of smooth cascades. Russian Math. Surveys, 39(1984), no 6, 85-131.
  • [38] Poincare H. Sur les courbes définies par les equations differentielles. J. Math. Pures Appl., 2(1886), 151-217.
  • [39] Pugh C. On a theorem of P.Hartman. Amer. Journ. Math., 91(1962), no 2, 363-367.
  • [40] Pugh C., Walker R.B., Wilson F.W. On Morse-Smale approximations – a counterexample. Journ. Diff. Equat., 23(1977), 173-182.
  • [41] Robinson C. Dynamical Systems: stability, symbolic dynamics, and chaos. Studies in Adv. Math., Sec. edition, CRC Press, 1999.
  • [42] Robinson C., Williams R. Finite stability is not generic. Dynamical Systems: Proc. Symp. (Brazil, 1971), Academic Press, New York, London, 1973, 451-462.
  • [43] Smale S. Morse inequalities for a dynamical system. Bull. Amer. Math. Soc., 66(1960), 43-49.
  • [44] Smale S. Differentiable dynamical systems. Bull. Amer. Math. Soc., 73(1967), 747-817.
  • [45] Strogatz S. Nonlinear Dynamics and Chaos with Applications to Physixs, Biology, Chemistry and Engineering. Persius Books, Cambridge, Massachusetts.
  • [46] Williams R. The DA maps of Smale and structural stability. Global Anal., Proc. Symp. Pure Math., AMS, 14(1970), 329-334.
  • [47] Williams R. Expanding attractors. Publ. Math. I.H.E.S., 43(1974), 169-203.
  • [48] Zhuzhoma E.V., Isaenkova N.V. Zero-dimensional solenoidal base sets. Sbornik: Mathematics, 202(2011), no 3, 351-372.