Abstract:
It is known that a nontrivial attractor coexists with trivial basic sets in the nonwandering set of an Ω-stable 3-diffeomorphism if and only if it is either nonorientable one-dimensional or (orientable or not) expanding and two-dimensional. Examples of such diffeomorphisms were constructed previously, with the exception of the case of a nonorientable two-dimensional attractor. The paper fills this gap. In addition, it is constructively shown that the diffeomorphism obtained has an energy function, which extends thereby the class of cascades with global Lyapunov function whose set of critical points coincides with the nonwandering set of the dynamical system.
Bibliography: 20 titles.
With the exception of § 5, this research was supported by the Russian Science Foundation under grant no. 22-11-00027, https://rscf.ru/en/project/22-11-00027/. The results in § 5 were obtained in the framework of the HSE University Basic Research Program.
In this paper we consider Ω-stable1[x]1A diffeomorphism of a closed n-manifold is said to be Ω-stable if the diffeomorphisms C1-close to it are topologically conjugate on the nonwandering sets. diffeomorphisms defined on smooth closed n-manifolds. A spectral decomposition theorem [1] holds for such diffeomorphisms: the nonwandering set of an Ω-stable diffeomorphism is a disjoint finite union of so-called basis subsets, each of which is compact, invariant and topologically transitive. If a basic set is a periodic orbit, then it is said to be trivial, otherwise it is nontrivial.
Let f:Mn→Mn be an Ω-stable diffeomorphism of a manifold Mn with metric d, and let Λ be a basic set of f. Each point x in Λ has so-called invariant manifolds, a stable one Wsx and an unstable one Wux. They are defined by
Wsx={y∈Mn|limn→+∞d(fn(x),fn(y))=0}
and
Wux={y∈Mn|limn→+∞d(f−n(x),f−n(y))=0}.
By the generalized stable manifold theorem [1] the invariant manifolds Wsx and Wux are the images of Euclidean spaces Rk and Rn−k for some k∈{0,1,…,n} under injective immersions2[x]2Here by an immersion we mean a differentiable map whose differential is nondegenerate at each point.Jsx:Rk→Mn and Jux:Rn−k→Mn, respectively. Hence for each r>0 the sets Wsx,r=Jsx(Dkr) and Wux,r=Jux(Dn−kr), where Dkr={x∈Rk:‖ and D^{n-k}_r=\{x\in\mathbb R^{n-k}\colon \|x\|\leqslant r\}, are smoothly embedded discs.
If at least one of the discs W^{\mathrm s}_{x,r} and W^{\mathrm u}_{x,r} has dimension one, that is, k is equal to 1 or n-1, then we can speak about the intersection index of these submanifolds. For simplicity let W^{\mathrm s}_{x,r} be one-dimensional. Let U be a tubular neighbourhood of W^{\mathrm u}_{x,r}, which is the image of an embedding in M^n of the total space of a one-dimensional vector bundle on W^{\mathrm u}_{x,r} ([2], Ch. 4, § 5). Since the ball D^{n-1}_r is contractible, this bundle is trivial and U\setminus{W^{\mathrm u}_{x,r}} consists of two connected components U_+ and U_-. Hence we can define a function \nu\colon U_+\cup U_-\to\mathbb{Z} such that \nu(x)=1 for x\in U_+, and \nu(x)=0 for x\in U_-. If the manifolds W^{\mathrm s}_{x,r} and W^{\mathrm u}_{x,r} intersect transversally at a point y=J^{\mathrm s}_x(t), t\in D^1_r, then there exists \delta>0 such that J^{\mathrm s}_x(t-2\delta,t+2\delta)\subset U.
is called the intersection index of W^{\mathrm s}_{x,r} and W^{\mathrm u}_{x,r} at y. Note that this index takes values in \{1,-1\}. A basic set is said to be orientable if one of its invariant manifolds is one-dimensional and the index of intersection of W^{\mathrm s}_{x,r} and W^{\mathrm u}_{x,r} is the same at all points for each r > 0. Otherwise the set is said to be nonorientable. Note that this definition of the orientability of a basic set can be used for diffeomorphisms of orientable and nonorientable manifolds alike.
A compact invariant set \mathcal A is called an attractor of the diffeomorphism f if it has a compact neighbourhood U_{\mathcal A} such that f(U_{\mathcal A})\subset \operatorname{Int}U_{\mathcal A} and \bigcap_{n\in\mathbb{N}}f^n(U_{\mathcal A})={\mathcal A}. If an attractor is a basic set and its topological dimension is equal to the dimension of the unstable manifolds of points in it, then this attractor is said to be expanding.
One important example of a two-dimensional diffeomorphism with an expanding attractor is the so-called DA-diffeomorphism (Derived from Anosov), originally constructed by Smale [1]. It is obtained from an Anosov diffeomorphism3[x]3A diffeomorphism of a smooth closed manifold is called an Anosov diffeomorphism if the whole supporting manifold is a hyperbolic set. on a 2-torus by using a Smale surgery4[x]4We present a detailed description of a Smale surgery in § 2. in a neighbourhood of a fixed point. The nonwandering set of a DA-diffeomorphism consists of an expanding orientable one-dimensional attractor and a fixed source. By making Smale surgeries in neighbourhoods of several periodic orbits we can obtain generalized DA-diffeomorphisms with a unique nontrivial attractor and periodic sources. A similar operation can be performed for an Anosov diffeomorphism on an n-torus, provided that the stable manifolds of points on the torus are one-dimensional. Such diffeomorphisms are also called \mathrm{DA}-diffeomorphisms if a single source arises, and they are called generalized \mathrm{DA}-diffeomorphisms if there are several sources.
In 1974 Plykin [3] presented a geometric construction of a structurally stable diffeomorphism of a 2-sphere with nonorientable one-dimensional attractor and four fixed points. Subsequently, it was shown (for instance, see [4], § 17.2) that we can obtain such a diffeomorphism from a generalized DA-diffeomorphism with four sources (one of which is fixed and three have period 3), by taking the quotient of the torus by an involution. We call the attractor obtained by a similar construction on an n-torus a generalized Plykin attractor. By construction it is an expanding nonorientable attractor of dimension n-1. Note that, as follows from [5], a 3-manifold admitting an \Omega-stable diffeomorphism with a generalized Plykin attractor is nonorientable, while the diffeomorphism itself is not structurally stable. In the same paper a diffeomorphism with such an attractor was constructed for which the nontrivial repeller dual to this attractor is a nonwandering set.
It is known from the recent results of Barinova, Pochinka and Yakovlev [6] that if all nontrivial basic sets of an \Omega-stable 3-diffeomorphism are attractors, then they can belong to one of the following three types: a nonorientable one-dimensional attractor, an expanding orientable two-dimensional attractor, or an expanding nonorientable two-dimensional attractor.
One example of such a diffeomorphism with an expanding orientable two-dimensional attractor is a DA-diffeomorphism of a 3-torus. An example of a diffeomorphism with a nonorientable one-dimensional attractor is the cascade on a 3-sphere obtained by compactifying the Cartesian product of a Plykin diffeomorphism of a 2-sphere and a hyperbolic contraction on a line (Figure 1).
In the paper [6] mentioned above no obstructions were found to the simultaneous existence of a generalized Plykin attractor and trivial basic sets in the nonwandering set of an \Omega-stable 3-diffeomorphism, but no examples of such diffeomorphisms have been known so far. In the two-dimensional case, to obtain a cascade on a closed manifold with a single basic set which is a Plykin attractor, it is sufficient to add four fixed hyperbolic sources. We cannot complement similarly the construction of the system in three dimensions because of the structure of the basin of a generalized Plykin attractor.
Our main result in this paper is a constructive proof of the following theorem.
Theorem 1. There exists an \Omega-stable 3-diffeomorphism whose nonwandering set consists of a generalized Plykin attractor and 12 isolated periodic points.5[x]5It was shown in [7] then the number of isolated periodic points cannot be less than 12.
For the proof we present a construction of such diffeomorphisms in §§ 2 and 3. Let \mathcal P denote the class of diffeomorphisms constructed in this way. Their supporting manifolds are obtained by gluing four cylinders over the real projective plane to the boundary of a trapping neighbourhood of a nontrivial attractor. In § 4 we calculate the homology of the supporting manifolds in our examples. It turns out that in this way we can obtain diffeomorphisms which are not topologically conjugate to one another because the fundamental group of the supporting manifold depends essentially on the way we glue the cylinders.
Theorem 2. Let f\colon M^3\to M^3 be a diffeomorphism in the class \mathcal P. Then the homology groups of the manifold M^3 are as follows:
where m\in\{1,2,3\}, and each of these values of m can be realized.
Another result in our paper is the proof that diffeomorphism in the class \mathcal P have energy functions.
Recall that a continuous function \varphi\colon M^n\to\mathbb{R} is called a Lyapunov function for an \Omega-stable diffeomorphism f if \varphi(x)>\varphi(f(x)) when x is a wandering point and \varphi(x)=\varphi(y) when x and y lie in the same basic set. If the Lyapunov function is smooth and its critical point set coincides with the nonwandering set of the diffeomorphism, then it is called an energy function.
In contrast to flows, a cascade does not necessarily have an energy function (for instance, see [8]). The first example of such a cascade on a 3-sphere is due to Pixton [9] (1977): it is a Morse-Smale diffeomorphism6[x]6A diffeomorphism f\colon M^n\to M^n of a closed n-manifold M^n is a Morse–Smale diffeomorphism if it is structurally stable and its nonwandering set is finite. with four fixed points. In the same paper Pixton proved that Morse–Smale diffeomorphisms of surfaces have Morse energy functions. In [10]–[13] Grines, Pochinka and Barinova distinguished several classes of 2- and 3-diffeomorphisms with energy functions. In 2022 a class of dynamical systems on surfaces without energy functions was discovered [14].
In this paper we present a constructive proof of the following result.
Theorem 3. Diffeomorphisms in the class \mathcal P have energy functions.
§ 2. Generalized DA-diffeomorphism
In this section we present the construction of a 3-diffeomorphism obtained from an Anosov diffeomorphism by a Smale surgery. We prove that this diffeomorphism is \Omega-stable and its nonwandering set consists of an expanding orientable two-dimensional attractor and eight fixed sources. Our construction is similar to the two-dimensional case described by Katok and Hasselblatt in their book (see [4], §§ 1.8 and 17.2).
2.1. The construction
Let \mathbb{T}^3=\mathbb{R}^3/\mathbb{Z}^3 be a 3-torus and p\colon \mathbb{R}^3\to\mathbb{T}^3 be the natural projection. We begin with a hyperbolic automorphism \widehat{L}_A\colon \mathbb{T}^3\to\mathbb{T}^3 which is induced by a linear map L_A of \mathbb{R}^3 with hyperbolic matrix A\in \operatorname{GL}(3,\mathbb{Z}) whose eigenvalues \lambda_1, \lambda_2 and \lambda_3 satisfy 0<|\lambda_1|<1<|\lambda_2|\leqslant|\lambda_3|. For instance, we can define L_A by
Then \widehat{L}_A=p L_A p^{-1}\colon \mathbb T^3\to\mathbb T^3 is an Anosov diffeomorphism. The one-dimensional stable two-dimensional unstable manifolds of points on the torus form two transversal invariant foliations, each leaf in which is dense on the torus. Each point with rational coordinates is periodic for \widehat{L}_A, so the set of periodic points is also dense in \mathbb{T}^3.
We describe the action of \Gamma=\{1,-1\} on the torus \mathbb{T}^3=\mathbb{R}^3/\mathbb{Z}^3 by setting \gamma\cdot((x,y,z)+\mathbb{Z}^3)=\gamma (x,y,z)+\mathbb{Z}^3 for all (x,y,z)\in\mathbb{R}^3 and \gamma\in \Gamma. Then \gamma=1 corresponding to the identity map and \gamma=-1 to an involution J with the eight fixed points
These are periodic points of \widehat{L}_A, and therefore there exists k\in\mathbb{N} such that all these points are fixed by \widehat{L}_A^k. For \widehat{L}_A as above k=3, and the matrix A^k has the form
define two invariant foliations on the torus: the one-dimensional stable one W^{\mathrm s} and the two-dimensional unstable one W^{\mathrm u}.
We take r > 0 to be sufficiently small so that the \lambda^{\mathrm u}_2 r-balls with centres \overline v_i, i=1,\dots,8, are disjoint. Setting r_1= \frac{\lambda^{\mathrm s}}{16}r, r_2=\frac{r}{2} and t=\sqrt{y^2+z^2} we fix \varepsilon\ll r_1.
Lemma 1. There exists a diffeomorphism \Delta\colon \mathbb R^3\to\mathbb R^3 of the form
in addition, \delta'_x\leqslant 1 for |x|\geqslant r_1+\varepsilon or t\geqslant r_2.
Proof. Consider a smooth odd function \delta_0(x)\colon \mathbb{R}\to\mathbb{R} obtained from a piecewise linear function by smoothing it by means of arcs of circles.7[x]7Note that \delta_0 can be made arbitrarily smooth; we smoothen it by arcs of circles for geometric transparency. For x\geqslant 0 we show \delta_0(x) in Figure 2; it is expressed by the formula
\begin{equation*}
\delta_0(x)=\begin{cases} \dfrac{2x}{\lambda^{\mathrm s}} & \text{for } 0\leqslant x < r_1-\varepsilon_1, \\ \sqrt{\rho^2_{1}-(x-x_{O_1})^2}+y_{O_1} & \text{for } r_1-\varepsilon_1\leqslant x < r_1+\varepsilon, \\ \dfrac{2}{4-\lambda^{\mathrm s}}x+\dfrac{2 -\lambda^{\mathrm s}}{4-\lambda^{\mathrm s}}r_2 & \text{for } r_1+\varepsilon \leqslant x \leqslant r_2-\varepsilon, \\ - \sqrt{\rho^2_2-(x-x_{O_2})^2}+y_{O_2} & \text{for } r_2-\varepsilon < x \leqslant r_2+\varepsilon_2, \\ x & \text{for } x > r_2+\varepsilon_2, \end{cases}
\end{equation*}
\notag
where (r_1-\varepsilon_1,r_1+\varepsilon) and (r_2-\varepsilon,r_2+\varepsilon_2) are the smoothing intervals and we have \varepsilon_1<\varepsilon and \varepsilon_2<\varepsilon, and where (x_{O_1},y_{O_1}) and (x_{O_2},y_{O_2}) are the coordinates of the centres of the circles used for smoothing and \rho_1 and \rho_2 are their radii.
We extend \delta_t(x) to the negative values of x to have an odd function, that is, \delta_t(x)=-\delta_t(-x) for x<0. We show the graphs of \delta_t(x) for various t in Figure 4. Then \delta(x,y,z)=\delta_{\sqrt{y^2+z^2}}(x) is the required function.
The function \delta(x,y,z) is C^1-smooth in the whole domain of definition because \delta_0(x) is C^1-smooth and \sigma(t) is C^\infty-smooth. Now we show that the diffeomorphism \Delta=(\delta(x,y,z),y,z) we have constructed satisfies all conditions in the lemma. In fact,
Since \delta_0'(x)>0 and 0\leqslant\sigma\leqslant 1, it follows that \delta'_x>0.
Note that the odd function (\delta_0(x)-x) is nonnegative for x>0 and attains its maximum on the interval (r_1-\varepsilon,r_1+\varepsilon). For estimates of the derivatives with respect to y and z we look at the following chains of inequalities bearing in mind that r_1=\frac{\lambda^{\mathrm s}}{8}r_2, r_1+\varepsilon<2r_1 and \delta_0\leqslant 2x/\lambda^{\mathrm s} for x>0:
If |x|\geqslant r_1+\varepsilon, then \delta'_0 is nondecreasing and \delta'_0(x)\leqslant 1. The range of \sigma(t) is the interval [0,1]. Hence \delta'_x\leqslant 1 if |x|\geqslant r_1+\varepsilon or t\geqslant r_2.
Let \overline D_i\subset \mathbb R^3 be the ball neighbourhood of \overline v_i of radius r, and let D_i=p(\overline D_i). In the neighbourhoods \overline D_i of the points \overline v_i we introduce new normalized coordinates so that the new x-axis corresponds to the vector \mathbf e^{\mathrm s}, while the y- and z-axes correspond to \mathbf e^{\mathrm u}_1 and \mathbf e^{\mathrm u}_2. Then T=\begin{pmatrix} \mathbf e^{\mathrm s} & \mathbf e^{\mathrm u}_1 & \mathbf e^{\mathrm u}_2 \end{pmatrix} is the transition matrix from the basis \begin{Bmatrix}\mathbf e^{\mathrm s}, \mathbf e^{\mathrm u}_1, \mathbf e^{\mathrm u}_2\end{Bmatrix} to
to \begin{Bmatrix}\mathbf e^{\mathrm s}, \mathbf e^{\mathrm u}_1, \mathbf e^{\mathrm u}_2\end{Bmatrix}. We denote the linear maps with matrices T and T^{-1} by L_T and L_{T^{-1}}. Then in the neighbourhood D_i of the point v_i, i \in \{1, 2,\dots,8\}, the transition map \psi_i\colon D_i\to\mathbb R^3 looks as follows:8[x]8Below we need the inverse map of \psi in a domain wider than \psi(D_i). We set \psi^{-1}(x,y,z)=p(L_{T}(x,y,z)+\overline v_i) for all (x,y,z)\in\mathbb R^3.
Defined in this way, f_{\mathrm{DA}} commutes with the involution J, there are fixed hyperbolic sources at the fixed points of the involution, and the one-dimensional foliation W^{\mathrm s} for the original Anosov diffeomorphism remains invariant under f_{\mathrm{DA}}.
2.2. \Omega-stability
Theorem 4. The diffeomorphism f_{\mathrm{DA}} is \Omega-stable.
Proof. It follows from [15], Theorems 4.1, 5.7 and 9.1, that when the chain-recurrent set \mathrm{CR}(f_{\mathrm{DA}}) of the diffeomorphism f_{\mathrm{DA}} is hyperbolic, then this diffeomorphism is \Omega-stable.
Let \mathcal A denote the set \mathrm{CR}(f_{\mathrm{DA}})\setminus\bigcup_{i=1}^8\{v_i\}. First of all, we prove a lemma on the position of \mathcal A. Let K_i be subsets of the D_i such that their images \overline K_i in the local charts (D_i,\psi_i) are cylinders
Lemma 2. The domain K_i is a subset of the basin W^{\mathrm u}_{v_i} of the source v_i.
Proof. Let \overline A_i=\{(x,y,z)\in\mathbb R^3\mid |x|<r_1-\varepsilon,\,y^2+z^2<(r_2/2)^2\}. In the domain A_i=\psi^{-1}(\overline A_i) the map f_{\mathrm{DA}} has the form (2x,\lambda^{\mathrm u}_1 y,\lambda^{\mathrm u}_2 z) in the local charts under consideration, so that A_i is a subset of the basin of v_i. In these charts the image of A_i has the form
because \lambda^{\mathrm u}_2>\lambda^{\mathrm u}_1>2. Since \varepsilon is small, we have K_i\subset f(A_i), so that K_i\subset W^{\mathrm u}_{v_i}.
Thus, the set \mathcal A lies outside the domains K_i, i=1,\dots,8. Now we show that \mathcal A is a hyperbolic set, that is, the tangent subbundle T_{\mathcal A}(\mathbb T^3) has a continuous decomposition into a direct sum E^{\mathrm s}_{\mathcal A}\oplus E^{\mathrm u}_{\mathcal A}=\bigcup_{x\in\mathcal A}E^{\mathrm s}_x\oplus E^{\mathrm u}_x, which is invariant under the differential Df_{\mathrm{DA}}, and there exists constants c>0 and \mu\in (0,1) such that
Now we find the differential Df_{\mathrm{DA}} of f_{\mathrm{DA}}. If we choose local charts so that the eigenvectors of A are the unit coordinate vectors, similarly to the charts (D_i,\psi_i), then this differential has the form
and we have \delta'_x=1 and \delta'_y=\delta'_z=0 if the point lies outside the domains D_i.
Since the surgery described above preserved the foliation W^{\mathrm s}, as the Df_{\mathrm{DA}}-invariant summand E^{\mathrm s}_{x}, x\in\mathcal A, we take the linear span of \begin{pmatrix} 1\\ 0\\ 0 \end{pmatrix}. Let \mathbf{v}=\begin{pmatrix} x\\ 0\\ 0 \end{pmatrix}\in E^{\mathrm s}_{\mathcal A}; then
so that the required conditions on E^{\mathrm s}_{\mathcal A} are satisfied.
To prove that there exists a two-dimensional subspace E^{\mathrm u}_{\mathcal A} we use the cone criterion ([4], Corollary 6.4.8). We show that for the family H of cones of the form y^2+z^2 \geqslant \gamma^2 x^2 there exists \mu>1 such that
\begin{equation*}
D f_{\mathrm{DA}} (H) \subset \operatorname{Int}H, \qquad \|D f_{\mathrm{DA}} (\mathbf{v})\| \geqslant \mu \|\mathbf{v}\|, \quad \mathbf{v} \in H.
\end{equation*}
\notag
We verify the first condition. Let
\begin{equation*}
\begin{pmatrix} \overline x\\ \overline y\\ \overline z \end{pmatrix} =Df_{\mathrm{DA}}\begin{pmatrix} x\\ y\\ z \end{pmatrix}.
\end{equation*}
\notag
We show that if t\geqslant\gamma |x|, \gamma>0, then \sqrt{\overline{y}^2+\overline{z}^2}=\overline{t} \geqslant \gamma |\overline{x}|. To do this we find estimates for the left- and right-hand sides taking the bounds in Lemma 1 into account and bearing in mind that we only need to consider t\neq 0:
We verify that there exists \mu>1 such that \|D f_{\mathrm{DA}} (v)\| \geqslant \mu \|\mathbf{v}\| for \mathbf{v} \in H, that is, we verify the inequality \overline x^2+\overline y^2+\overline z^2\geqslant \mu(x^2+y^2+z^2):
We see that the family of cones y^2+z^2\geqslant (0.5x)^2 and \mu=1.5 satisfy all the assumptions of the cone criterion. In combination with the calculations for E^{\mathrm s}_{\mathcal A}, this shows that the set \mathcal A is hyperbolic, and therefore f_{\mathrm{DA}} is \Omega-stable.
2.3. The existence and uniqueness of a two-dimensional expanding attractor
The diffeomorphism f_{\mathrm{DA}} is \Omega-stable, and \mathrm{NW}(f_{\mathrm{DA}})=\mathcal A\cup\bigcup_{i=1}^8 \{v_i\}. Hence \mathcal A is the union of a nonempty finite system of basic sets. In this subsection we find saddle points whose two-dimensional unstable manifolds consist of nonwandering points and are dense in \mathcal A. Then we can prove that \mathcal A contains a two-dimensional expanding attractor and no other basic sets.
Lemma 3. Each domain D_i, i=1,\dots,8, contains two fixed saddle points \beta^+_i and \beta^-_i.
Proof. In the local charts (D_i,\psi_i) the diffeomorphism f_{\mathrm{DA}} has the form
so that fixed points can only lie on the Ox-axis. We consider the restriction {f_x=\lambda^{\mathrm s}\delta_0(x)} of f_{\mathrm{DA}} to the Ox-axis and find its fixed points from the equation \lambda^{\mathrm s}\delta_0(x)=x (Figure 5). Since \varepsilon is small, for x>0 the solution lies in the interval (r_1-\varepsilon;r_1+\varepsilon), so it is sufficient to solve the equation
Then the points \beta^+_i=\phi_i^{-1}(x^*,0,0) and \beta^-_i=\phi_i^{-1}(-x^*,0,0) are fixed and thus lie in \mathcal A. As the set \mathcal A has a hyperbolic structure, the dimensions of the sets E^{\mathrm s}_x and E^{\mathrm u}_x, x\in\mathcal A, show that \beta^+_i and \beta^-_i are hyperbolic saddle points with one-dimensional stable separatrices and two-dimensional unstable ones.
Lemma 4. The unstable manifolds of the saddle points \beta^+_i and \beta^-_i, i=1,\dots,8, consist of nonwandering points.
Proof. The diffeomorphism f_{\mathrm{DA}} has a one-dimensional invariant foliation W^{\mathrm s}, the same as the one for the Anosov diffeomorphism \widehat L_A. Each leaf l^{\mathrm s} of W^{\mathrm s} is dense on the torus; moreover, if w\in l^{\mathrm s}, then each connected component of the set l^{\mathrm s}\setminus\{w\} is dense. Consider the one-dimensional leaf l^{\mathrm s} through \beta^+_i. The lift of l^{\mathrm s} to \mathbb R^3 that passes through \overline v_i contains no other points all of whose coordinates are rational. Hence the connected component l^+ of the set l^{\mathrm s}\setminus\{\beta^+_i\} that does not contain the fixed source does not pass through any source v_j, j\neq i, and therefore it coincides with the separatrix of \beta^+_i.
It follows from the construction of f_{\mathrm{DA}} that the disc
is a subset of the image \psi_i(W^{\mathrm u}_{\beta^+_i}) of the unstable separatrix of \beta^+_i. Since l^+ is dense on the torus and in the local charts under consideration the image of l^+\cap D_i is a union of line intervals parallel to the Ox-axis which is dense in \overline D_i, the homoclinic points are dense in the set \psi^{-1}_i(\overline D^{\mathrm u}) and therefore also in W^{\mathrm u}_{\beta^+_i} (Figure 6). Then the fact that points in the nonstable separatrix of \beta^+_i are nonwandering is obvious because so are homoclinic points. For the unstable separatrices of the points \beta^-_i the proof is similar.
Theorem 5. The nonwandering set of f_{\mathrm{DA}} contains a two-dimensional expanding attractor \Lambda_{\mathrm{DA}}.
Proof. It follows from Lemma 4 that the unstable two-dimensional manifolds of fixed saddle points lie in \mathcal A. By the spectral decomposition theorem (see [1]) there exists a basic set \Lambda_{\mathrm{DA}} of type9[x]9The type of a basic set \Lambda is the pair of numbers (\dim W^{\mathrm u}_x,\dim W^{\mathrm s}_x), x\in\Lambda.(2,1) that contains the two-dimensional unstable manifold of the saddle. Then \Lambda_{\mathrm{DA}} has topological dimension at least 2. Let {\Lambda_{\mathrm{DA}}=3}; however, then by Lemma 8.1 in [16] \Lambda_{\mathrm{DA}} coincides with the supporting manifold \mathbb T^3, which is impossible since the nonwandering set of f_{\mathrm{DA}} contains sources. Then the basic set \Lambda_{\mathrm{DA}} has dimension 2, so it is an attractor (see [17], Theorem 3), which is two-dimensional and expanding.
The proof is complete.
Theorem 6. The set \mathcal A coincides with the attractor \Lambda_{\mathrm{DA}}.
Proof. Since basic sets are compact, it is sufficient to show that the unstable manifold of the saddle point \beta^+_1 is dense in \mathcal A. Let a\in\mathcal A, and let U_a be some neighbourhood of a. Since periodic points are dense in the nonwandering sets of \Omega-stable diffeomorphisms, there exists a point p\in U_a that is periodic with period m_p. At least one of its one-dimensional separatrices, we denote it by l^{\mathrm s}_p, is dense on the torus and intersects the unstable manifold of \beta^+_1. Let q be one of their points of intersection. Since the stable separatrix l^{\mathrm s}_p is f^{m_p}-invariant, it follows that
However, q also belongs to the invariant manifold of \beta^+_1, so there exists an index k such that W^{\mathrm u}_{\beta^+_1}\cap U_a\neq\varnothing. Thus, \mathcal A contains no basic sets other than \Lambda_{\mathrm{DA}}.
The proof is complete.
Summarizing the results of Theorems 4–6 we see that f_{\mathrm{DA}} is an \Omega-stable diffeomorphism and its nonwandering set consists of eight fixed hyperbolic sources and a two-dimensional expanding attractor, that is, it is a generalized DA-diffeomorphism. Note that, as the attractor \Lambda_{\mathrm{DA}} is the complement to the basins of sources, which are homeomorphic to 3-balls, it is connected. One separatrix of each saddle point \beta_i^+ and each point \beta_i^-, i=1,\dots,8, lies fully in the basin of the corresponding source, so these saddle points are boundary periodic points of the expanding attractor \Lambda_{\mathrm{DA}} and their unstable manifolds combine into 2-connectives. Hence the attractor obtained is orientable.
§ 3. Constructing a 3-diffeomorphism with generalized Plykin attractor
In this section we present the construction of a 3-diffeomorphism with generalized Plykin attractor and 12 fixed points, that is, we prove Theorem 1.
Consider the quotient space M_s=\mathbb{T}^3/\Gamma=\mathbb T^3/J. This is a three-dimensional orbifold with eight singular (orbifold) points. Then the quotient space \widetilde M=(\mathbb{T}^3\setminus \bigcup^8_{i=1}\{v_i\})/{\Gamma} is a manifold, and the natural projection \widetilde p\colon \mathbb{T}^3\setminus\bigcup^8_{i=1}\{v_i\}\to \widetilde M is a two-sheeted cover. Since f_{\mathrm{DA}}J=Jf_{\mathrm{DA}}, f_{\mathrm{DA}} induces a diffeomorphism \widetilde f=\widetilde pf_{\mathrm{DA}}\widetilde p^{-1}\colon \widetilde M\to\widetilde M with a generalized Plykin attractor \Lambda=\widetilde p(\Lambda_{\mathrm{DA}}), whose stable manifold coincides with \widetilde M. The set \widetilde M\setminus \Lambda is the wandering set of \widetilde f; it consists of eight connected components \widetilde B_i, i=1,\dots,8, each obtained as a quotient of the basin W^{\mathrm u}_{v_i}\setminus \{v_i\} of a source v_i. It follows from [18] that \widetilde B_i, i=1,\dots,8, is diffeomorphic to \mathbb RP^2\times \mathbb{R}, where \mathbb RP^2 is the real projective plane.
To understand the topological structure of the orbit space \widetilde B_i/\widetilde f of the restriction of \widetilde f to each connected component of the wandering set \widetilde B_i, we find a fundamental domain F_i of the restriction of f_{\mathrm{DA}} to the sets W^{\mathrm u}_{v_i}\setminus \{v_i\}, i=1,\dots,8, and project it onto \widetilde M. To do this we consider the local charts (\psi_i, D_i) introduced above. For x^2+y^2+z^2\leqslant r_1^2 the diffeomorphism f_{\mathrm{DA}} looks like f_{\mathrm{DA}}(x,y,z)=(2x,\lambda^{\mathrm u}_1 y,\lambda^{\mathrm u}_2 z) in these charts. Then
Note that each leaf of this foliation is invariant under the involution J on \mathbb{T}^3, so the fundamental domain \widetilde F_i of the restriction of \widetilde f to \widetilde B_i is the projection of the fundamental domain F_i, that is, \widetilde F_i=\widetilde p(F_i); moreover, \widetilde F_i is diffeomorphic to {\mathbb RP^2\times [0,1]} because after taking the quotient each ellipse becomes a set diffeomorphic to \mathbb RP^2. Then the required orbit space {\widetilde B}_i/{\widetilde f} has the form \mathbb RP^2\times [0,1]|_{(x,0)\sim (\widetilde{f}(x),1)} and is diffeomorphic to \mathbb{R}P^2\times S^1, since any diffeomorphism on \mathbb RP^2 is isotopic to the identity map [19].
Consider a structurally stable diffeomorphism g_{RP^2}\colon \mathbb RP^2\to\mathbb RP^2 with three fixed points: a source \alpha, a sink \omega and a saddle point \sigma (Figure 7).
Also let g_R\colon \mathbb R\to\mathbb R be the diffeomorphism defined by g_{R}(x)=2x, let g_k={g_{RP^2}\times g_R}\colon C_k\to C_k for k=1,\dots,4, and let g\colon C\to C be the diffeomorphism made up of the g_k. It is obvious that the orbit spaces C_k^-/g_k and C_k^+/g_k are diffeomorphic to \mathbb RP^2\times\mathbb S^1.
Thus, for all i=1,\dots,8 and k=1, \dots,4 the orbit spaces C_k^-/g_k and C_k^+/g_k are diffeomorphic to {\widetilde B_i}/\widetilde f, and therefore for each permutation P=\{i_1,j_1,i_2,j_2,i_3,j_3,i_4,j_4\} of the string \{1,\dots,8\} there exists a diffeomorphism h\colon C^{\pm}\to \widetilde M\setminus\Lambda=\bigcup_{i=1}^8 \widetilde B_i that conjugates g|_{C^{\pm}} to \widetilde f|_{\widetilde M\setminus\Lambda} and such that
\begin{equation*}
M_P=C\bigcup_h \widetilde M,
\end{equation*}
\notag
and let q\colon C\sqcup \widetilde M\to M_P be the natural projection. Let f_{P}\colon M_P \to M_P be a diffeomorphism coinciding with q \widetilde f (q| _{\widetilde M})^{- 1} on q (\widetilde M) and with q g (q | _{C})^{- 1} on q(C). Then, as a diffeomorphism required in Theorem 1, we can choose, for example, the diffeomorphism f_P corresponding to the permutation P=\{1,\dots,8\}; all diffeomorphisms corresponding to permutations form a class \mathcal P.
§ 4. Homology of supporting manifolds
In this section we calculate the integer homology of the supporting manifolds M_P for the diffeomorphisms f_{P} constructed in § 3. This will prove Theorem 2.
Since the set q(C^0) is closed in M_P and lies in \operatorname{Int}{N}_P, the inclusion (M_P\setminus q(C^0),N_P\setminus q(C^0))\to(M_P,N_P) induces an isomorphism H_n(M_P\setminus q(C^0),N_P\setminus q(C^0))\cong H_n(M_P,N_P). Since N_P\setminus q(C^0)=q(C^1\setminus C^0)\approx N_1\times[0,1), the inclusion (M_1,N_1)\to(M_P\setminus q(C^0),N_P\setminus q(C^0)) is a homotopy equivalence of pairs, and so induces an isomorphism H_n(M_1,N_1)\cong H_n(M_P\setminus q(C^0),N_P\setminus q(C^0)). As a result, we obtain isomorphisms H_n(M_1,N_1)\cong H_n(M_P,N_P) induced by the inclusion (M_1,N_1)\to(M_P,N_P).
Then q\colon \widetilde{M}_1\to M_1 is a homeomorphism and q(\widetilde{N}_1)=N_1. Hence q induces an isomorphism H_n(\widetilde{M}_1,\widetilde{N}_1)\cong H_n(M_1,N_1).
Let M_s=\mathbb{T}^3/\Gamma, let p\colon \mathbb{R}^3\to\mathbb{T}^3 and p_J\colon \mathbb{T}^3\to M_s be the natural projections, and let p_s=p_J\circ p. Set
\begin{equation*}
\widetilde{N}=(q|_{\widetilde{M}})^{-1}(q(C^1\setminus{C}^0)), \qquad V=\{p_J(v_1),\dots,p_J(v_8)\}\quad\text{and}\quad N_s=\widetilde{N}\cup V.
\end{equation*}
\notag
Then \widetilde{M}=M_s\setminus{V}, and therefore \widetilde{M}\subset M_s and \widetilde{N}\subset M_s; moreover, N_s is the closure of \widetilde{N} in the orbifold M_s. For the same reasons as above we have isomorphisms H_n(\widetilde{M}_1,\widetilde{N}_1)\cong H_n(M_{s}\setminus{V},N_{s}\setminus{V})\cong H_n(M_{s},N_{s}), which are induced by the inclusions of pairs (\widetilde{M}_1,\widetilde{N}_1)\to(M_{s}\setminus{V},N_{s}\setminus{V}) and (M_{s}\setminus{V},N_{s}\setminus{V})\to(M_s,N_s).
Combining these results, we arrive at the following statement.
Lemma 5. Let \imath_s\colon \widetilde{M}\to M_s be the inclusion and q\colon C\sqcup \widetilde M\to M_P be the natural projection. Then the maps of pairs \imath_s\colon (\widetilde{M}_1,\widetilde{N}_1)\to(M_s,N_s) and q\colon (\widetilde{M}_1,\widetilde{N}_1)\to(M_P,N_P) induce isomorphisms
The action of the group \Gamma on \mathbb{T}^3 has a fundamental domain equal to the image Q=p(\overline{Q}) of the parallelepiped \overline{Q}=[0,1]^2\times[0,1/2]. Furthermore, \partial{Q}=p([0,1]^2\times 0)\cup{p}([0,1]^2\times(1/2)). The involution J\in\Gamma acts on \partial{Q} by the formula J(p(x,y,\epsilon))=p(1-x,1-y,\epsilon), where \epsilon\in\{0,1/2\}. This produces a natural cell decomposition of M_s (Figure 8). It consists of the zero-dimensional cells e^0_i=p_J(v_i)=p_s(\overline{v}_i), i=1,\dots,8, the three-dimensional cell e^3=p_s(\operatorname{Int}\overline{Q}), the one-dimensional cells
We denote the closure of the 1-cell e^1_i by a_i, i=1,\dots,7, and the closure of e^2_j by V_j, j=1,\dots,4. Then a_1=[e^0_1e^0_3], a_2=[e^0_2e^0_4], a_3=[e^0_1e^0_5], a_4=[e^0_2e^0_6], a_5=[e^0_5e^0_7], a_6=[e^0_6e^0_8] and a_7=[e^0_1e^0_2]. Each closed 2-cell V_j is homeomorphic to S^2 and is a (‘pillow’) orbifold with four singular points. More precisely,
The above cell decomposition of M_s shows that \partial{e}^2_j=\partial{V}_j=0 for j=1,\dots,4 and, if the orientations of 2-cells are as indicated in Figure 1, \partial{e}^3=2V_3-2V_4. Direct calculations show that Z_1(M_s)=0. Therefore,
By construction \widetilde{N}=(q|_{\widetilde{M}})^{-1}(q(C^1\setminus{C}^0))\subset\bigsqcup_{i=1}^8\widetilde{B}_i. Let \widetilde{N}_i=\widetilde{N}\cap\widetilde{B}_i, and let N_{si}=\widetilde{N}_i\cup\{e^0_i\} be the closure of \widetilde{N}_i in M_s. Then N_{si} is a closed neighbourhood of the singular point e^0_i and N_s=\bigsqcup_{i=1}^8N_{si}.
The sets \widetilde{N}_i are homeomorphic to the cylinder \mathbb{R}P^2\times(0,1] or \mathbb{R}P^2\times[-1,0), and the neighbourhoods N_{si} are homeomorphic to a cone over \mathbb{R}P^2. Hence
Let V_{j1}=V_j\cap M_{s1}, V_{jP}=q(V_{j1}) and \overline{V}_{jP}=V_{jP}+C_2(N_P) for j=1,\dots,4. Then from (4.8), (4.9) and Lemma 5 we deduce the following result.
Comparing these formulae, we conclude that t=1 and H_2(M_P)\cong\mathbb{Z}^3\times\mathbb{Z}_2.
By construction \widetilde{N}_1=\bigsqcup_{i=1}^8\widetilde{N}_{1i}, where \widetilde{N}_{1i}\approx\mathbb{R}P^2 and \widetilde{N}_{1i}\subset\widetilde{B}_i; we also have N_1=q(\widetilde{N}_1). Let \widetilde{u}_i be the 1-cycle in \widetilde{N}_{1i} with homology class equal to the generator of the group H_1(\widetilde{N}_{1i})\cong\mathbb{Z}_2. Then H_1(\widetilde{N}_1)=\langle[\widetilde{u}_1],\dots,[\widetilde{u}_8]\rangle\cong\mathbb{Z}^8. Set u_i=q(\widetilde{u}_i) for i=1,\dots,8. Since the cycles u_{i_k} and u_{j_k} are homologous in N_P, it follows that
For each j=1,\dots,4 the boundary \partial{V}_{j1} is a sum of the four cycles in the set \widetilde{u}_1,\dots,\widetilde{u}_8 that correspond to the singular vertices in V_j. By definition the homomorphism \partial_*^2 in (4.11) satisfies the equalities \partial_*^2([\overline{V}_{jP}])=[\partial{V}_{jP}]=[q(\partial{V}_{j1})]. Hence by (4.1) and Lemma 6 the subgroup \operatorname{im}\partial_*^2 of H_1(N_P) is generated by the elements
Since the permutation P involves all elements of \{1,\dots,8\}, it follows from (4.16) and (4.12) that w_4=0. Therefore, \operatorname{im}\partial_*^2\cong\mathbb{Z}_2^{\mathrm s}, where 0\leqslant{s}\leqslant3.
If in the permutation P=\{i_1,j_1,i_2,j_2,i_3,j_3,i_4,j_4\} we replace a pair \{i_k,j_k\} by \{j_k,i_k\}, then h(C_k^-)=\widetilde{B}_{j_k} and h(C_k^+)=\widetilde{B}_{i_k}. By interchanging two pairs \{i_k,j_k\} and \{i_l,j_l\} in P, k\ne l, we simply renumber the cylinders in the list C_1,\dots,C_4. These transformations do not change the topological type of M_P, so in what follows we assume without loss of generality that
By (4.13), (4.12) and (4.17) the equality w_1\!=\!0 is only possible when \{\mkern-1mu i_1,j_1,i_2,j_2\mkern-1mu\} is a permutation of the elements in the string \{1,2,3,4\} such that i_1=1, and \{i_3,j_3,i_4,j_4\} is a permutation of elements in the string \{5,6,7,8\}. If we have w_2=0 in addition, then j_1=2 by (4.14). However, by (4.15), for i_1=1 and j_1=2 we have w_3\ne0. Thus, \operatorname{im}\partial_*^2\ne0, and therefore s>0.
We see that \operatorname{im}\partial_*^2\cong\mathbb{Z}_2^{\mathrm s}, where 0\!\leqslant\! s\!\leqslant\!3. Moreover, by (4.11) we have {\operatorname{im}\imath_*^1\!\cong\!\mathbb{Z}_2^m}, where m=4-s. But then by the same formula (4.11) we have H_1(M_P)\cong\mathbb{Z}^4\times\mathbb{Z}_2^m and 0\leqslant{m}\leqslant3.
Proof. Let P=\{1,2,3,4,5,6,7,8\}. Then [u_1],[u_3],[u_5],[u_7] is a basis of H_1(N_P), and the relations [u_1]=[u_2], [u_3]=[u_4], [u_5]=[u_6] and [u_7]=[u_8] hold in H_1(N_P). These relations and formulae (4.13) and (4.14) show that w_1=w_2=0. Therefore, \operatorname{im}\partial_*^2=\langle w_3\rangle\cong\mathbb{Z}_2. Thus, in our case s=1 and m=4-s=3.
For P=\{1,2,3,5,4,6,7,8\} the basis of H_1(N_P) is formed by [u_1],[u_3],[u_4] and [u_7], and the relations in (4.12) are as follows:
Furthermore, by (4.13)–(4.15) we have w_1=w_2=[u_3]+[u_4] and w_3=[u_1]+[u_7]. The coordinates of w_1 and w_2 with respect to the basis [u_1],[u_3],[u_4],[u_7] form the matrix
Since \operatorname{rank}{A}=2, it follows that \operatorname{im}\partial_*^2=\langle w_1,w_3\rangle\cong\mathbb{Z}_2^2. Hence in this case s=2 and m=2.
Finally, let P=\{1,3,2,5,4,6,7,8\}. Then [u_1], [u_2], [u_4], [u_7] is a basis of H_1(N) and we have [u_1]=[u_3], [u_2]=[u_5], [u_4]=[u_6] and [u_7]=[u_8]. These relations show that
Since \operatorname{rank}B=3, we have \operatorname{im}\partial_*^2=\langle w_1,w_2,w_3\rangle\cong\mathbb{Z}_2^3. Thus, for P in question we have s=3 and m=1.
Now we show that diffeomorphisms in \mathcal P have energy functions, that is, we prove Theorem 3.
Proof of Theorem 3. Let f_P\colon M_P\to M_P be a diffeomorphism in \mathcal P with generalized Plykin attractor \Lambda. By construction M_P=C\bigcup_h \widetilde M, where the set C consists of four cylinders over the real projective plane, that is, C=\bigsqcup_{k=1}^4 C_k, where C_k=\mathbb{R}P^2\times\mathbb{R}, and the image of \widetilde M under the natural projection q\colon C\sqcup\widetilde M\to M_P is the stable manifold of the attractor \Lambda.
Recall that the restriction of the projection q to C is a diffeomorphism onto its image q(C)=M_P\setminus\Lambda, which conjugates f_P|_{M_P\setminus\Lambda} smoothly to the diffeomorphism g made up of maps g_k\colon C_k\to C_k of the form g_k=g_{RP^2}\times g_R, where g_{RP^2} is a structurally stable diffeomorphism of the real projective plane whose nonwandering set consists of three fixed points: a source \alpha, a sink \omega and a saddle point \sigma (see Figure 7), and g_R(x)=2x for x\in\mathbb{R}.
Pixton [9] proved in 1977 that Morse–Smale diffeomorphisms of closed surfaces have energy functions, so g_{RP^2} has one. We present a geometric version of the construction of such a function in the lemma below.
Lemma 9. Let p_{RP^2}\colon \mathbb S^2\to \mathbb {R}P^2 be a two-sheeted cover and p_z\colon \mathbb R^3\to\mathbb R be the canonical projection onto the Oz-axis. Then there exists an immersion H\colon \mathbb S^2\to\mathbb R^3 such that \varphi_{RP^2}=p_z\circ H\circ p^{-1}_{RP^2}\colon \mathbb RP^2\to\mathbb R is a well-defined function which is an energy function for g_{RP^2}.
Proof. Let \overline{g}_{RP^2}\colon \mathbb S^2\to\mathbb S^2 be a lift of g_{RP^2} such that the nonwandering set of \overline g_{RP^2} consists of six fixed points (Figure 9). Consider a smooth map H\colon \mathbb S^2\to\mathbb{R}^3 of the 2-sphere to three-dimensional Euclidean space such that the height function p_z\circ H commutes with p_{RP^2}, so that for all a\in\mathbb RP^2 the set p_zHp_{RP^2}^{-1}(a) consists of a single point and the image of the fixed-point set of \overline g_{RP^2} is the critical set of the function p_z\circ H.
We define the required image H(\mathbb S^2) of the sphere parametrically. For r\in[0,8] and s\in [0, 2\pi] we consider the map \Theta=(\Theta_x, \Theta_y, \Theta_z)\colon [0,8]\times [0,2\pi]\to\mathbb R^3 such that the first two components have the form
It turn, we can parametrize the sphere by the map \Upsilon=(\Upsilon_x,\Upsilon_y,\Upsilon_z)\colon [0,2\pi]\times[-\pi/2,\pi/2]\to \mathbb S^2\subset\mathbb{R}^3, where
Then H=\Theta\,{\circ}\, H_{sq}\,{\circ}\, \Upsilon^{-1} is the required immersion. The function p_z\circ H is a Lyapunov function for the diffeomorphism10[x]10Note that although we have not formally defined the diffeomorphisms g_{RP^2} and \overline g_{RP^2}, Figure 7 specifies a conjugacy class of diffeomorphisms of the real projective plane, in which we can select a diffeomorphism g_{RP^2} such that p_z\circ H is a Lyapunov function for its lift.\overline g_{RP^2}, which is the lift of g_{RP^2}. The critical points of p_z\circ H coincide with the nonwandering points of \overline g_{RP^2}. Furthermore, we can verify directly that p_z\circ H commutes with the covering map p_{RP^2}, so that \varphi_{RP^2}= p_zHp_{RP^2}^{-1} is an energy function for our system on \mathbb{R}P^2.
We continue the proof of Theorem 3. The cascade g_R has the energy function \varphi_R(x)=-x^2. Then the fact that the direct product of g_{RP^2} and g_R has an energy function \varphi_k\colon C_k\to\mathbb R, k=1,\dots,4, is a consequence of [20]. Moreover, we can take it in the form
By the definition of the \varphi_k, k=1,\dots,4, the function \Phi is continuous.
Although \Phi is an energy function for f_{P} on the set M_P\setminus \Lambda, it is not smooth on the whole of M_P in general. By the smoothing lemma in [14] there exists a function \xi\colon \mathbb R \to \mathbb R such that \xi \circ \Phi is the required energy function for f_{P}.
M. W. Hirsch, Differential topology, Grad. Texts in Math., 33, Corr. reprint of the 1976 original, Springer-Verlag, New York–Heidelberg, 2012, x+221 pp.
3.
R. V. Plykin, “Sources and sinks of A-diffeomorphisms of surfaces”, Math. USSR-Sb., 23:2 (1974), 233–253
4.
A. Katok and B. Hasselblatt, Introduction to the modern theory of dynamical systems, Encyclopedia Math. Appl., 54, Cambridge Univ. Press, Cambridge, 1995, xviii+802 pp.
5.
E. V. Zhuzhoma and V. S. Medvedev, “On non-orientable two-dimensional basic sets on 3-manifolds”, Sb. Math., 193:6 (2002), 869–888
6.
M. Barinova, O. Pochinka and E. Yakovlev, “On a structure of non-wandering set of an \Omega-stable 3-diffeomorphism possessing a hyperbolic attractor”, Discrete Contin. Dyn. Syst., 44:1 (2024), 1–17
7.
M. Barinova, “On isolated periodic points of diffeomorphisms with expanding attractors of codimension 1”, Regul. Chaotic Dyn., 29:5 (2024), 777–802; arXiv: 2404.156992
8.
V. Z. Grines and O. V. Pochinka, “The constructing of energy functions for \Omega-stable diffeomorphisms on 2- and 3-manifolds”, J. Math. Sci. (N.Y.), 250:4 (2020), 537–568
9.
D. Pixton, “Wild unstable manifolds”, Topology, 16:2 (1977), 167–172
10.
V. Z. Grines, F. Laudenbach and O. V. Pochinka, “Dynamically ordered energy function for Morse–Smale diffeomorphisms on 3-manifolds”, Proc. Steklov Inst. Math., 278 (2012), 27–40
11.
V. Z. Grines, M. K. Noskova and O. V. Pochinka, “The construction of an energy function for three-dimensional cascades with a two-dimensional expanding attractor”, Trans. Moscow Math. Soc., 76:2 (2015), 237–249
12.
V. Z. Grines, M. K. Noskova and O. V. Pochinka, “Energy function for A-diffeomorphisms of surfaces with one-dimensional nontrivial basic sets”, Dinamicheskie Sistemy, 5(33):1–2 (2015), 31–37 (Russian)
13.
M. Barinova, V. Grines, O. Pochinka and B. Yu, “Existence of an energy function for three-dimensional
chaotic “sink-source” cascades”, Chaos, 31:6 (2021), 063112, 8 pp.
14.
M. K. Barinova, “On existence of an energy function for \Omega-stable surface diffeomorphisms”, Lobachevskii J. Math., 42:14 (2021), 3317–3323
15.
C. Robinson, Dynamical systems. Stability, symbolic dynamics, and chaos, Stud. Adv. Math., 2nd ed., CRC Press, Boca Raton, FL, 1999, xiv+506 pp.
16.
V. Z. Grines, T. V. Medvedev and O. V. Pochinka, Dynamical systems on 2- and 3-manifolds, Dev. Math., 46, Springer, Cham, 2016, xxvi+295 pp.
17.
R. V. Plykin, “The topology of basis sets for Smale diffeomorphisms”, Math. USSR-Sb., 13:2 (1971), 297–307
18.
M. Barinova, V. Grines and O. Pochinka, “Dynamics of three-dimensional A-diffeomorphisms with two-dimensional attractors and repellers”, J. Difference Equ. Appl., 29:9–12 (2023), 1275–1286
19.
M.-E. Hamstrom, “Homotopy properties of the space of homeomorphisms on P^2 and the Klein bottle”, Trans. Amer. Math. Soc., 120:1 (1965), 37–45
20.
M. K. Barinova and E. K. Shustova, “Dynamical properties of direct products if discrete dynamical systems”, Zh. Srednevolzhsk. Mat. Obshch., 24:1 (2022), 21–30 (Russian)
Citation:
M. K. Barinova, O. A. Kolchurina, E. I. Yakovlev, “On 3-diffeomorphisms with generalized Plykin attractors”, Sb. Math., 215:9 (2024), 1135–1158