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

An example derived from Lorenz attractor

Ming Li, Fan Yang, Jiagang Yang, Rusong Zheng M. Li is supported by NSF of China (11971246, 12071231) and National Key R&\&D Program of China (2020YFA0713300); J. Yang is supported by NSF of China (12271538, 11871487, 12071202), CNPq of Brazil, FAPERJ of Brazil, and PRONEX of Brazil; R. Zheng is supported by NSF of China (12071007, 12101293).
Abstract

We consider a DA-type surgery of the famous Lorenz attractor in dimension 4. This kind of surgeries have been firstly used by Smale [S] and Mañé [M1] to give important examples in the study of partially hyperbolic systems. Our construction gives the first example of a singular chain recurrence class which is Lyapunov stable, away from homoclinic tangencies and exhibits robustly heterodimensional cycles. Moreover, the chain recurrence class has the following interesting property: there exists robustly a 2-dimensional sectionally expanding subbundle (containing the flow direction) of the tangent bundle such that it is properly included in a subbundle of the finest dominated splitting for the tangent flow.

1   
Introduction

1.1 Motivation

In the study of differentiable dynamical systems, a central problem is to understand most of the systems. From a topological viewpoint, this could mean to characterize an open and dense subset, or at least a residual subset, of all systems. It was once conjectured by Smale in the early 1960’s that an open and dense subset of all systems should be comprised of uniformly hyperbolic ones. Recall that a dynamical system is uniformly hyperbolic if its limit set consists of finitely many transitive hyperblic sets and without cycles in between. The first counterexamples to this conjecture were given by Abraham and Smale [AS] in the C1C^{1} topology, and by Newhouse [N1, N2] in the C2C^{2} topology. These examples give the following two obstructions to hyperbolicity:

  1. 1.

    heterodimensional cycle : there exist two hyperbolic periodic orbits of different indices such that the stable manifold of one periodic orbit intersects the unstable manifold of the other and vice versa;

  2. 2.

    homoclinic tangency : there exists a hyperbolic periodic orbit whose stable and unstable manifolds have a nontransverse intersection.

Bifurcations of these two mechanisms can exhibit very rich dynamical phenomena. For example, a heterodimensional cycle associated to a pair of periodic saddles with index ii and j=i+1j=i+1 can be perturbed to obtain a robust heterodimensional cycle [BD]: there exist transitive hyperbolic sets Λ1\Lambda_{1} and Λ2\Lambda_{2} with different indices such that the stable manifold of Λ1\Lambda_{1} meets the unstable manifold of Λ2\Lambda_{2} in a C1C^{1} robust way, and vice versa. For more results, see e.g. [PT, D, BDPR]. Trying to give a global view of all dynamical systems, Palis [Pa1, Pa2, Pa3, Pa4] conjectured in the 1990’s that the uniformly hyperbolic systems and the systems with heterodimensional cycles or homoclinic tangencies form a dense subset of all diffeomorphisms. The Palis conjecture has been proved by Pujals and Sambarino [PS] for C1C^{1} diffeomorphisms on surfaces. Moreover, a weaker version, known as the weak Palis conjecture, was shown to hold for C1C^{1} diffeomorphisms on any closed manifold [BGW, C1]. When it comes to flows, or vector fields, the Palis conjecture has to be recast so as to take into consideration the singularities. There have been many progresses on this topic, see e.g. [ARH, GY, CY1]. Also, related conjectures have been given by Bonatti [Bo] in his program for a global view of C1C^{1} systems.

The conjectures by Palis and Bonatti have lead to the study of systems away from homoclinic tangencies and/or heterodimensional cycles. In the C1C^{1} topology, a diffeomorphism (or a vector field) is said to be away from homoclinic tangencies if every system in a C1C^{1} neighborhood exhibits no homoclinic tangency. It is now well understood [W, Go] that for a diffeomorphism away from homoclinic tangencies there exist dominated splittings of the tangent bundle over preperiodic sets. More delicate results can be obtained in the C1C^{1} generic setting, see e.g. [CSY]. Also, it is known [LVY] that every C1C^{1} diffeomorphism away from homoclinic tangencies is entropy expansive, and it follows that there exists a measure of maximum entropy. When considering diffeomorphisms away from homoclinic tangencies and heterodimensional cycles, it is shown in [CP] that a C1C^{1} generic system is essentially hyperbolic. See also [C2].

While there have been plenty of results on diffeomorphisms away from homoclinic tangencies (and heterodimensioinal cycles), parallel studies on vector fields are rare. Regarding the Palis conjecture, it is shown [CY1] that on every 3-manifold the singular hyperbolic vector fields form a C1C^{1} open and dense subset of all vector fields away from homoclinic tangencies. Also in dimension 3, the weak Palis conjecture has been proved for C1C^{1} vector fields [GY]. For higher dimensional (dim 3\geq 3) vector fields, studies are mostly limited to either the singular hyperbolic vector fields [MPP] or the star vector fields, see e.g. [SGW, PYY, CY2, dL, BdL]. General vector fields away from homoclinic tangencies (not star) on a higher dimensional manifold are rarely studied. In [GYZ], for general higher dimensional flows away from homoclinic tangencies, it is shown that C1C^{1} generically a nontrivial Lyapunov stable chain recurrence class contains periodic orbits and hence is a homoclinic class. Their result leaves the possibility that the periodic orbits in the chain recurrence class can have different indices. When this happens robustly, such a chain recurrence class shall exhibit a robust heterodimensional cycle [BD] and the flow can not be star. Apart from this result and possibly a handful others, the world of general higher dimensional vector fields away from homoclinic tangencies is far from being understood.

Inspired by the result of [GYZ], we notice that there is no known example of a singular flow which is away from homoclinic tangencies and is not singular hyperbolic or star. The aim of this paper is to provide such an example, and to open a door to the world of general higher dimensional vector fields away from homoclinic tangencies. Hopefully it will also attract interests to the study of higher dimensional vector fields.

As a by-product, our example exhibits an interesting property which is not observed before: the tangent bundle of the obtained chain recurrence class is shown to admit robustly an invariant subbundle (containing the flow direction), which is properly included in a subbundle of the finest dominated spitting of the tangent bundle. We shall call such an invariant subbundle exceptional (see Definition 1.1). Note that there can not be any exceptional subbundle for diffeomorphisms away from homoclinic tangencies: robustness of an invariant subbundle implies a corresponding dominated spitting of the tangent bundle. Thus the existence of exceptional subbundles may be counted as an essential difference between diffeomorphisms and (singular) vector fields, and is worth studying.

1.2 Precise statement of our result

Let MM be a compact Riemannian manifold without boundary. Given a C1C^{1} vector field on MM, it generates a C1C^{1} flow ϕtX\phi^{X}_{t}, or simply ϕt\phi_{t}, on the manifold. A point xMx\in M with X(x)=0X(x)=0 is called a singularity of XX. Any point other than singularities is called a regular point. A regular point xx satisfying ϕT(x)=x\phi_{T}(x)=x for some T>0T>0 is called a periodic point, and its orbit Orb(x)={ϕt(x):t}\operatorname{Orb}(x)=\{\phi_{t}(x):t\in{\mathbb{R}}\} will be called a periodic orbit. Singularities and periodic orbits are called critical elements.

Let Λ\Lambda be a compact invariant set of the flow ϕt\phi_{t}. It is called chain transitive if for any two points x,yΛx,y\in\Lambda and any constant ε>0{\varepsilon}>0, there exist a sequence of points x0=x,x1,,xn=yx_{0}=x,x_{1},\ldots,x_{n}=y in Λ\Lambda and a sequence of positive numbers t0,t1,,tn11t_{0},t_{1},\ldots,t_{n-1}\geq 1, such that d(ϕti(xi),xi+1)<εd(\phi_{t_{i}}(x_{i}),x_{i+1})<{\varepsilon} for all i=0,1,,n1i=0,1,\ldots,n-1. If Λ\Lambda is maximally chain transitive, i.e. it is chain transitive and is not a proper subset of any other chain transitive set, then Λ\Lambda is called a chain recurrence class. The union of all chain recurrence classes is called the chain recurrent set, denoted by CR(X)\mathrm{CR}(X). For any point xx in CR(X)\mathrm{CR}(X), we denote by C(x,X)C(x,X) the chain recurrence class containing xx.

Denote by Φt\Phi_{t} the tangent flow of XX, i.e. Φt=Dϕt\Phi_{t}=D\phi_{t}. A Φt\Phi_{t}-invariant subbundle EE of TΛMT_{\Lambda}M with dimE2\operatorname{dim}E\geq 2 is called sectionally expanded if there exists constants C1C\geq 1, λ>0\lambda>0 such that for any xΛx\in\Lambda and any 2-dimensional subspace SE(x)S\subset E(x), it holds

|det(Φt|S)|Ceλt,t0.|\det(\Phi_{-t}|_{S})|\leq C\mathrm{e}^{-\lambda t},\quad\forall t\geq 0.

A Φt\Phi_{t}-invariant splitting EFTΛME\oplus F\subset T_{\Lambda}M is called dominated if there exists C1C\geq 1, λ>0\lambda>0 such that for any xΛx\in\Lambda, it holds

Φt|E(x)Φt|F(ϕt(x))Ceλt,t0.\|\Phi_{t}|_{E(x)}\|\cdot\|\Phi_{-t}|_{F(\phi_{t}(x))}\|\leq C\mathrm{e}^{-\lambda t},\quad\forall t\geq 0.

We also say that a Φt\Phi_{t}-invariant splitting E1E2EkTΛME_{1}\oplus E_{2}\oplus\cdots\oplus E_{k}\subset T_{\Lambda}M is dominated if (E1Ei)(Ei+1Ek)(E_{1}\oplus\cdots\oplus E_{i})\oplus(E_{i+1}\oplus\cdots\oplus E_{k}) is dominated for every i=1,,k1i=1,\ldots,k-1. The finest dominated splitting of TΛMT_{\Lambda}M is a Φt\Phi_{t}-invariant splitting TΛM=E1E2EkT_{\Lambda}M=E_{1}\oplus E_{2}\oplus\cdots\oplus E_{k} such that it is dominated and no subbundle EiE_{i} (i=1,,ki=1,\ldots,k) can be decomposed further into a dominated splitting.

Definition 1.1.

Let Λ\Lambda be a chain transitive compact invariant set of a C1C^{1} vector field XX, a subbundle ETΛME\subset T_{\Lambda}M is called exceptional if it satisfies all the following properties:

  • dimE2\operatorname{dim}E\geq 2, and X(x)E(x)X(x)\in E(x) for any xΛx\in\Lambda;

  • EE is invariant for the tangent flow;

  • EE is properly included in a subbundle of the finest dominated splitting of TΛMT_{\Lambda}M.

In particular, if ETΛME\subset T_{\Lambda}M is exceptional, it is not a subbundle of any dominated splitting of TΛMT_{\Lambda}M. Recall that a compact invariant set Λ\Lambda is called Lyapunov stable if for any neighborhood UU of Λ\Lambda there exists a neighborhood VV of Λ\Lambda such that ϕt(V)¯U\overline{\phi_{t}(V)}\subset U for all t0t\geq 0. Let 𝒳1(M)\mathscr{X}^{1}(M) be the set of all C1C^{1} vector fields on MM, endowed with the C1C^{1} topology. We denote by 𝒯\mathcal{H}\mathcal{T} the set of all C1C^{1} vector fields on MM exhibiting a homoclinic tangency. Then 𝒳1(M)𝒯¯\mathscr{X}^{1}(M)\setminus\overline{\mathcal{H}\mathcal{T}} consists of all C1C^{1} vector fields away from homoclinic tangencies.

Theorem A.

On every 4-dimensional closed Riemannian manifold MM there exists a non-empty open set 𝒱𝒳1(M)𝒯¯\mathcal{V}\subset\mathscr{X}^{1}(M)\setminus\overline{\mathcal{H}\mathcal{T}} and an open set UMU\subset M such that for any X𝒱X\in\mathcal{V}, one has ϕtX(U¯)U\phi^{X}_{t}(\overline{U})\subset U for any t>0t>0, and

  • there is a unique singularity σXU\sigma_{X}\in U and it is hyperbolic;

  • C(σX,X)C(\sigma_{X},X) is the unique Lyapunov stable chain recurrence class contained in UU;

  • C(σX,X)C(\sigma_{X},X) contains a robust heterodimensional cycle;

  • the tangent bundle TC(σX,X)MT_{C(\sigma_{X},X)}M contains a 2-dimensional subbundle EXE^{X} which is exceptional and sectionally expanded, and it varies upper semi-continuously with respect to XX.

Remark 1.1.

Since a robust heterodimensional cycle yields non-hyperbolic periodic orbits by arbitrarily small C1C^{1} perturbations, the chain recurrence class C(σX,X)C(\sigma_{X},X) for any X𝒱X\in\mathcal{V} is not singular hyperbolic or star.

Remark 1.2.

In [BLY], a non-empty C1C^{1} open set of vector fields having an attractor containing singularities of different indices is constructed on a given 4-manifold. Their example exhibits also a robust heterodimensional cycle, but is accumulated by vector fields with homoclinic tangencies.

1.3 Idea of the construction

The example (in Theorem A) is derived from Lorenz attractor in the sense that the construction begins with the Lorenz attractor and uses a DA-type surgery. The Lorenz attractor, discovered by the meteorologist E. Lorenz in the early 1960’s [Lo], is among the most important examples of singular flows. It is a chaotic attractor given by a simple group of three ordinary differential equations. We will consider the geometric model of the Lorenz attractor [ABS, GW], which is equivalent to the original attractor for classical parameters [T]. The DA surgery was introduced by Smale [S] as a way of constructing nontrivial basic sets, see also [Wi]. It has been used to give many inspiring and important examples, especially in the study of partially hyperbolic systems, see e.g. [M1, BV].

In the following we give some ideas of the construction.

We begin with a vector field X0X^{0} on the 3-ball B3B^{3} having a Lorenz attractor Λ0\Lambda_{0}. Then we consider a vector field X(x,s)=(X0(x),θs)X(x,s)=(X^{0}(x),-\theta s) on the product B3×[1,1]B^{3}\times[-1,1], where (x,s)(x,s) are coordinates on B3×[1,1]B^{3}\times[-1,1] and θ>0\theta>0 is a constant. In particular, the dynamics is contracting along fibers. Note that Λ=Λ0×{0}\Lambda=\Lambda_{0}\times\{0\} is an attractor and it contains a periodic orbit PP.

Now the idea is to modify the vector field XX in a neighborhood of the periodic orbit PP so that one obtains a robust heterodimensional cycle. For this purpose we combine a DA-type surgery with the construction of a blender [BDV]. Roughly speaking, a blender is a hyperbolic set which verifies some specific geometric properties so that it unstable manifold looks like a manifold of higher dimension. It naturally appears in the unfolding of a heterodimensional cycle of two periodic orbits and can be used to obtain robust heterodimensional cycles.

For this construction, we need to show that the obtained vector field is away from homoclinic tangencies. This requires a careful choice of the contraction rate θ\theta along the fibers. Moreover, we need to show that the singularity and the heterodiemnsional cycle are contained in the same chain recurrence class. This constitutes the most difficult part of our proof.

1.4 Further discussions

For the example in Theorem A, we would like to raise the following questions.

Question 1.

Is the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) (robustly) transitive? Or even isolated?

Apart from topological properties, one can also studies ergodic properties of the example. In [SYY] it is proved that the entropy function for singular flows away from homoclinic tangencies is upper semi-continuous with respect to both invariant measures and the flows, provided that the limiting measure is not supported on singularities.

Question 2.

Is the entropy function upper semi-continuous for the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) in Theorem A?

One may even consider existence of SRB measures or physical measures. However, let us mention one difficulty in the study of this example. In contrast to the Lorenz attractor or a singular hyperbolic attractor, the usual analysis of the return map to a global cross-section may not be applicable in our example. This is because the chain recurrence class in our example admits a partially hyperbolic splitting EsFE^{s}\oplus F, where the 3-dimensional center-unstable subbundle FF contains an exceptional 2-dimensional subbundle and is not sectionally expanded.

Apart from the present example, it may also be interesting to consider a DA-type surgery on the two-sided Lorenz attractor constructed by Barros, Bonatti and Pacifico [BBP].

The rest of this paper is organized as follows. We present some preliminaries of flows in Section 2 and give the construction of the example in Section 3. The proof of Theorem A is also given in Section 3, assuming some robust properties that will be proved in subsequent sections: in Section 4 we show that the example is away from homoclinic tangencies; then in Section 5 we show existence of an exceptional subbundle; finally in Section 6 we show that the singularity and the heterdimensional cycle are contained in the same chain recurrence class.

2   
Preliminaries

In this section we give some preliminaries on vector fields, including hyperboicity, linear Poincaré flow, fundamental limit, and so on.

Let MM be a compact Riemannian manifold without boundary. Let XX be a C1C^{1} vector field on MM. Denoted by Sing(X)\operatorname{Sing}(X) the set of singularities, i.e. Sing(X)={xM:X(x)=0}\operatorname{Sing}(X)=\{x\in M:X(x)=0\}. For any point xMx\in M, we denote by X(x)\langle X(x)\rangle the subspace of TxMT_{x}M that is generated by X(x)X(x). Note that X(x)\langle X(x)\rangle is a one-dimensional subspace of TxMT_{x}M when xSing(X)x\not\in\operatorname{Sing}(X). To simplify notations, we also denote lxX=X(x)l^{X}_{x}=\langle X(x)\rangle for xSing(X)x\notin\operatorname{Sing}(X).

2.1 Hyperbolicity

Recall that the flow and the tangent flow generated by XX are denoted as ϕt\phi_{t} and Φt\Phi_{t}, respectively. Let Λ\Lambda be a compact ϕt\phi_{t}-invariant set and ETΛME\subset T_{\Lambda}M a Φt\Phi_{t}-invariant subbundle. The dynamics Φt|E\Phi_{t}|_{E} is contracting, or EE is contracted, if there exists constants C1C\geq 1, λ>0\lambda>0, such that

Φt|E(x)Ceλt,for anyxΛ,t0.\|\Phi_{t}|_{E(x)}\|\leq C\mathrm{e}^{-\lambda t},\quad\text{for any}\ x\in\Lambda,\ t\geq 0.

One says that Φt|E\Phi_{t}|_{E} is expanding if Φt|E\Phi_{-t}|_{E} is contracting. The set Λ\Lambda is called hyperbolic if its tangent bundle admits a continuous splitting TΛM=EsXEuT_{\Lambda}M=E^{s}\oplus\langle X\rangle\oplus E^{u}, where X\langle X\rangle is the subbundle generated by the vector field XX, and EsE^{s} (resp. EuE^{u}) is contracted (resp. expanded). When Λ\Lambda is hyperbolic, for every xΛx\in\Lambda, the sets

Wss(x)={yM:dist(ϕt(x),ϕt(y))0ast+}W^{ss}(x)=\{y\in M:\mathrm{dist}(\phi_{t}(x),\phi_{t}(y))\to 0\ \text{as}\ t\to+\infty\}

and

Wuu(x)={yM:dist(ϕt(x),ϕt(y))0ast}W^{uu}(x)=\{y\in M:\mathrm{dist}(\phi_{t}(x),\phi_{t}(y))\to 0\ \text{as}\ t\to-\infty\}

are invariant C1C^{1} submanifolds tangent to Es(x)E^{s}(x) and Eu(y)E^{u}(y) respectively at xx, see [HPS]. We denote Ws(Orb(x))=zOrb(x)Wss(z)W^{s}(\operatorname{Orb}(x))=\bigcup_{z\in\operatorname{Orb}(x)}W^{ss}(z) and Wu(Orb(x))=zOrb(x)Wuu(z)W^{u}(\operatorname{Orb}(x))=\bigcup_{z\in\operatorname{Orb}(x)}W^{uu}(z), which are the stable and unstable manifolds of Orb(x)\operatorname{Orb}(x) respectively. For a chain transitive hyperbolic set, all stable manifolds Wss(x)W^{ss}(x) have the same dimension dimEs\operatorname{dim}E^{s}, which will be called the stable index of the hyperbolic set. Sometimes we simply say index with the same meaning of the stable index.

Flows with singularities, or singular flows, can exhibit complex dynamics with regular orbits accumulating on singularities, such as the famous Lorenz attractor [Lo, GW]. In such cases, the set Λ\Lambda can not be hyperbolic. Nevertheless, a weaker version of hyperbolicity can be defined. One says that Λ\Lambda is partially hyperbolic if there exists a Φt\Phi_{t}-invariant splitting of the tangent bundle TΛM=EsEcuT_{\Lambda}M=E^{s}\oplus E^{cu} such that EsE^{s} is contracted and it is dominated by EcuE^{cu}. If, moreover, the bundle EcuE^{cu} is sectionally expanded, then Λ\Lambda is said to be singular hyperbolic [MPP]. Note that the partial hyperbolicity or singular hyperbolicity ensures also the existence of stable manifolds.

We will study mainly chain recurrence classes. A chain recurrence class C(x,X)C(x,X) is nontrivial if it is not reduced to a critical element. In [PYY], it is shown that a nontrivial, Lyapunov stable, and singular hyperbolic chain recurrence class contains a periodic orbit, and C1C^{1} generically, such a chain recurrence class is indeed an attractor. This result has been generalized to C1C^{1} open and dense vector fields in [CY2].

Lemma 2.1 ([CY2]).

There is an open and dense set 𝒰𝒳1(M)\mathcal{U}\in\mathscr{X}^{1}(M) such that for any X𝒰X\in\mathcal{U}, any nontrivial, Lyapunov stable, and singular hyperbolic chain recurrence class C(σ)C(\sigma) of XX is a homoclinic class. Moreover, for any hyperbolic periodic orbit γC(σ)\gamma\subset C(\sigma), there exists a neighborhood UU of C(σ)C(\sigma) and a neighborhood 𝒰X\mathcal{U}_{X} of XX such that for any Y𝒰XY\in\mathcal{U}_{X}, the stable manifold Ws(γY,Y)W^{s}(\gamma_{Y},Y) is dense in UU.

2.2 Linear Poincaré flow and its extension

Given a ϕt\phi_{t}-invariant set ΛM\Lambda\subset M. Define the normal bundle over Λ\Lambda as

𝒩Λ=xΛSing(X)𝒩x,\mathcal{N}_{\Lambda}=\bigcup_{x\in\Lambda\setminus\operatorname{Sing}(X)}\mathcal{N}_{x},

where 𝒩x\mathcal{N}_{x} is the orthogonal complement of the one-dimensional subspace lxXTxMl^{X}_{x}\subset T_{x}M generated by X(x)X(x). In particular, if Λ=M\Lambda=M, then 𝒩M=MSing(X)𝒩x\mathcal{N}_{M}=\bigcup\limits_{M\setminus\operatorname{Sing}(X)}\mathcal{N}_{x}. For any xMSing(X)x\in M\setminus\operatorname{Sing}(X), v𝒩xv\in\mathcal{N}_{x} and tt\in{\mathbb{R}}, let ψt(v)\psi_{t}(v) to be the orthogonal projection of Φt(v)\Phi_{t}(v) to 𝒩ϕt(x)\mathcal{N}_{\phi_{t}(x)}. In this way one defines a flow ψt=ψtX\psi_{t}=\psi^{X}_{t} on the normal bundle 𝒩M\mathcal{N}_{M}, which is called the linear Poincaré flow. Note that the normal bundle is not defined at singularities. Hence the base of the normal bundle is non-compact for singular flows. We present below a compactification given by [LGW].

Define the fiber bundle G1=xMGr(1,TxM)G^{1}=\bigcup\limits_{x\in M}Gr(1,T_{x}M), where Gr(1,TxM)Gr(1,T_{x}M) is the Grassmannian of lines in the tangent space TxMT_{x}M through the origin. Let β:G1M\beta:G^{1}\to M be the corresponding bundle projection that sends a line lGr(1,TxM)l\in Gr(1,T_{x}M) to xx. The tangent flow Φt\Phi_{t} induces a flow Φ^t\hat{\Phi}_{t} on G1G^{1} defined by Φ^t(v)=Φt(v)\hat{\Phi}_{t}(\langle v\rangle)=\langle\Phi_{t}(v)\rangle, where vTMv\in TM is a nonzero vector and v\langle v\rangle is the linear subspace spanned by vv.

Let ξ:TMM\xi:TM\to M be the bundle projection of the tangent bundle. One then defines a vector bundle with base G1G^{1} as

β(TM)={(l,v)G1×TM:β(l)=ξ(v)}.\beta^{*}(TM)=\{(l,v)\in G^{1}\times TM:\beta(l)=\xi(v)\}.

The corresponding bundle projection ι\iota sends every vector (l,v)β(TM)(l,v)\in\beta^{*}(TM) to ll. The tangent flow induces also a flow Φ~t\tilde{\Phi}_{t} on β(TM)\beta^{*}(TM):

Φ~t(l,v)=(Φ^t(l),Φt(v)),(l,v)β(TM).\tilde{\Phi}_{t}(l,v)=(\hat{\Phi}_{t}(l),\Phi_{t}(v)),\quad\forall(l,v)\in\beta^{*}(TM).

The flow Φ~t\tilde{\Phi}_{t} will be called the extended tangent flow.

For any lG1l\in G^{1} such that β(l)=x\beta(l)=x, there is a natural identification between TxMT_{x}M and {l}×TxM\{l\}\times T_{x}M. Thus we define the extended normal bundle as

𝒩~={(l,v)β(TM):vl}.\widetilde{\mathcal{N}}=\{(l,v)\in\beta^{*}(TM):v\perp l\}.

We also define the extended normal bundle over any nonempty subset ΔG1\Delta\subset G^{1}:

𝒩~Δ={(l,v)ι1(Δ):vl}.\widetilde{\mathcal{N}}_{\Delta}=\{(l,v)\in\iota^{-1}(\Delta):v\perp l\}.

Finally, we define the extended linear Poincaré flow ψ~t\tilde{\psi}_{t} on 𝒩~\widetilde{\mathcal{N}} as follows:

ψ~t(l,v)=Φ~t(l,v)Φt(u),Φt(v)Φ~t(l,u),\tilde{\psi}_{t}(l,v)=\tilde{\Phi}_{t}(l,v)-\langle\Phi_{t}(u),\Phi_{t}(v)\rangle\cdot\tilde{\Phi}_{t}(l,u),

where ulu\in l is a unit vector, and ,\langle\cdot,\cdot\rangle stands for the inner product on the tangent bundle.

Remark 2.1.

For any xSing(X)x\notin\operatorname{Sing}(X), let l=lxXl=l^{X}_{x}, then 𝒩~l\widetilde{\mathcal{N}}_{l} can be naturally identified with 𝒩x\mathcal{N}_{x} and ψ~t|𝒩~l\tilde{\psi}_{t}|_{\widetilde{\mathcal{N}}_{l}} can be naturally identified with ψt|𝒩x\psi_{t}|_{\mathcal{N}_{x}}.

We will also consider invariant measures on G1G^{1}. Let μ\mu be any Borel measure on G1G^{1}, the bundle projection β:G1M\beta:G^{1}\to M induces a measure βμ\beta_{*}\mu on MM such that (βμ)(A)=μ(β1(A))(\beta_{*}\mu)(A)=\mu(\beta^{-1}(A)) for any Borel set AMA\subset M. When μ\mu is Φ^t\hat{\Phi}_{t}-invariant, the induced measure βμ\beta_{*}\mu is ϕt\phi_{t}-invariant.

2.3 Fundamental limit and dominated splitting

Given a chain recurrent set Γ\Gamma of the vector field XX. Let XnX_{n} be a sequence of C1C^{1} vector fields such that XnXX_{n}\to X in the C1C^{1} topology. Suppose there exists a periodic orbit γn\gamma_{n} of XnX_{n} such that γn\gamma_{n} converges to a compact subset of Γ\Gamma in the Hausdorff topology, then the sequence of pairs (γn,Xn)(\gamma_{n},X_{n}) is called a fundamental sequence of Γ\Gamma, and will be denoted as (γn,Xn)(Γ,X)(\gamma_{n},X_{n})\hookrightarrow(\Gamma,X). The sequence (γn,Xn)(\gamma_{n},X_{n}) is called an ii-fundamental sequence if Ind(γn)=i\operatorname{Ind}(\gamma_{n})=i for all nn large enough. When there is no possible ambiguity, we will simply say that γn\gamma_{n} is a fundamental sequence of Γ\Gamma. We denote by (Γ)\mathcal{F}(\Gamma) the limit of directions of all fundamental sequences of Γ\Gamma, i.e.

(Γ)={lG1:(γn,Xn)(Γ,X),pnγn,such thatlpnXnl}.\mathcal{F}(\Gamma)=\{l\in G^{1}:\exists(\gamma_{n},X_{n})\hookrightarrow(\Gamma,X),\ p_{n}\in\gamma_{n},\ \text{such that}\ l^{X_{n}}_{p_{n}}\to l\}.

As an immediate consequence of the definition, the map ()\mathcal{F}(\cdot) is upper semi-continuous in the following sense.

Lemma 2.2.

Suppose Γ\Gamma is a chain transitive set of XX. Then for any neighborhood \mathcal{B} of (Γ)\mathcal{F}(\Gamma) in G1G^{1}, there exist a neighborhood UU of Γ\Gamma and C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U} and any chain transitive set ΓY\Gamma_{Y} of YY contained in UU, it holds (ΓY)\mathcal{F}(\Gamma_{Y})\subset\mathcal{B}.

We also denote by B(Γ)B(\Gamma) the limit set of directions of regular points contained in Γ\Gamma, i.e.

B(Γ)={lG1:xnΓSing(X),such thatlxnXl}.B(\Gamma)=\{l\in G^{1}:\exists x_{n}\in\Gamma\setminus\operatorname{Sing}(X),\ \text{such that}\ l^{X}_{x_{n}}\to l\}.

Note that for any l(Γ)l\in\mathcal{F}(\Gamma), if x=β(l)x=\beta(l) is a regular point, then l=lxXl=l^{X}_{x}. This also holds for B(Γ)B(\Gamma). Thus (Γ)B(Γ)\mathcal{F}(\Gamma)\setminus B(\Gamma) is contained in the tangent spaces of singularities.

A hyperbolic singularity ρ\rho is called Lorenz-like if there is a partially hyperbolic splitting of the tangent bundle TρM=EρssEρcuT_{\rho}M=E^{ss}_{\rho}\oplus E^{cu}_{\rho} such that EρcuE^{cu}_{\rho} is sectionally expanded and it decomposes further into a dominated spitting EρcEρuE^{c}_{\rho}\oplus E^{u}_{\rho} with dimEρc=1\operatorname{dim}E^{c}_{\rho}=1 and Φt|Eρc\Phi_{t}|_{E^{c}_{\rho}} is contracting. For a Lorenz-like singularity ρ\rho, denote by Wss(ρ)W^{ss}(\rho) its strong stable manifold tangent to EρssE^{ss}_{\rho} at the singularity.

Lemma 2.3.

Let C(ρ,X)C(\rho,X) be a chain recurrence class of XX containing a Lorenz-like singularity ρ\rho such that C(ρ,X)(Wss(ρ){ρ})=C(\rho,X)\cap(W^{ss}(\rho)\setminus\{\rho\})=\emptyset. Then there exists a C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U}, it holds C(ρY,Y)(Wss(ρY,Y){ρY})=C(\rho_{Y},Y)\cap(W^{ss}(\rho_{Y},Y)\setminus\{\rho_{Y}\})=\emptyset and (C(ρY,Y))TρYMEρYcu\mathcal{F}(C(\rho_{Y},Y))\cap T_{\rho_{Y}}M\subset E^{cu}_{\rho_{Y}}.111Here, to simplify notations, we have identified TρYMT_{\rho_{Y}}M with Gr(1,TρYM)Gr(1,T_{\rho_{Y}}M), and EρYcuE^{cu}_{\rho_{Y}} with Gr(1,EρYcu)Gr(1,E^{cu}_{\rho_{Y}}). We hope that such abuses of notations shall not cause much confusion.

Proof.

Note that the properties defining a Lorenz-like singularity is C1C^{1} robust. Since the singularity ρ\rho is Lorenz-like, there is a C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U}, the continuation ρY\rho_{Y} is also Lorenz-like. Moreover, by the upper semi-continuity of chain recurrence class and the continuity of local strong stable manifolds, one deduces that C(ρY,Y)(Wss(ρY,Y){ρY})=C(\rho_{Y},Y)\cap(W^{ss}(\rho_{Y},Y)\setminus\{\rho_{Y}\})=\emptyset for any Y𝒰Y\in\mathcal{U} (shrinking 𝒰\mathcal{U} if necessary). Then, a similar argument as in the proof of [LGW, Lemma 4.4] gives (C(ρY,Y))TρYMEρYcu\mathcal{F}(C(\rho_{Y},Y))\cap T_{\rho_{Y}}M\subset E^{cu}_{\rho_{Y}}. ∎

A fundamental sequence (γn,Xn)(\gamma_{n},X_{n}) of Γ\Gamma is said to admit an index ii dominated splitting if there exist a ψtXn\psi^{X_{n}}_{t}-invariant splitting 𝒩γn=𝒩ncs𝒩ncu\mathcal{N}_{\gamma_{n}}=\mathcal{N}_{n}^{cs}\oplus\mathcal{N}_{n}^{cu} (nn\in\mathbb{N}) of the normal bundle and constants T>0T>0, n0>0n_{0}>0 such that for any t>Tt>T and n>n0n>n_{0}, it holds dim𝒩ncs=i\operatorname{dim}\mathcal{N}_{n}^{cs}=i and

ψtXn|𝒩ncs(x)ψtXn|𝒩ncu(ϕtXn(x))<1/2.\|\psi^{X_{n}}_{t}|_{\mathcal{N}_{n}^{cs}(x)}\|\cdot\|\psi^{X_{n}}_{-t}|_{\mathcal{N}_{n}^{cu}(\phi^{X_{n}}_{t}(x))}\|<1/2.

Identifying the extended normal bundle 𝒩~B(γn)\widetilde{\mathcal{N}}_{B(\gamma_{n})} with 𝒩γn\mathcal{N}_{\gamma_{n}} (see Remark 2.1), we see that 𝒩~B(γn)\widetilde{\mathcal{N}}_{B(\gamma_{n})} also admits a dominated splitting with respect to ψ~tXn\tilde{\psi}_{t}^{X_{n}} and of the same index. Thus, the following lemma is a consequence of continuity of domination.

Lemma 2.4.

(Γ)\mathcal{F}(\Gamma) admits a dominated splitting of index ii with respect to ψ~tX\tilde{\psi}^{X}_{t} if and only if every fundamental sequence of Γ\Gamma admits an index ii dominated splitting (with the same domination constant TT).

3   
The example: construction and robust properties

In this section we construct the example and claim some key properties of the example. With these properties, a proof of Theorem A is given at the end of this section.

3.1 The Lorenz attractor

Let us begin with the geometric model of the Lorenz attractor, see e.g. [GH, GW, BDV]. Let (x1,x2,x3)(x_{1},x_{2},x_{3}) be a Cartesian coordinate system in 3{\mathbb{R}}^{3}. Let B3B^{3} be a ball in 3\mathbb{R}^{3} centered at the origin O=(0,0,0)O=(0,0,0). We shall consider a C1C^{1} vector field X0X^{0} on 3{\mathbb{R}}^{3} such that it is transverse to the boundary of B3B^{3} and that the ball B3B^{3} is an attracting region, i.e. ϕtX0(x)B3\phi_{t}^{X^{0}}(x)\in B^{3} for any xB3x\in B^{3} and t>0t>0. Moreover, the following properties are assumed (see [AP, Section 3]):

  1. (P1)

    (Lorenz-like singularity) The origin OO is a hyperbolic singularity of stable index 2 such that its local stable manifold Wlocs(O)W^{s}_{loc}(O) coincides with the x1x3x_{1}x_{3}-plane, and its local unstable manifold Wlocu(O)W^{u}_{loc}(O) coincides with the x2x_{2}-axis. Moreover, the two-dimensional stable subspace EOsE^{s}_{O} decomposes into a dominated splitting EOs=EOssEOcE^{s}_{O}=E^{ss}_{O}\oplus E^{c}_{O}. Assume that the local strong stable manifold (tangent to EOssE^{ss}_{O}) coincides with the x1x_{1}-axis. The three Lyapunov exponents at the singularity are

    λs=logDΦ1X0|EOss,λc=logDΦ1X0|EOc,andλu=logDΦ1X0|EOu.\lambda^{s}=\log\|D\Phi^{X^{0}}_{1}|_{E^{ss}_{O}}\|,\quad\lambda^{c}=\log\|D\Phi^{X^{0}}_{1}|_{E^{c}_{O}}\|,\quad\text{and}\ \lambda^{u}=\log\|D\Phi^{X^{0}}_{1}|_{E^{u}_{O}}\|.

    Here, EOuE^{u}_{O} is the unstable subspace of OO. One has λs<λc<0<λu\lambda^{s}<\lambda^{c}<0<\lambda^{u}. We assume that λc+λu>0\lambda^{c}+\lambda^{u}>0, which implies sectional expanding property of the subspace EOcu=EOcEOuE^{cu}_{O}=E^{c}_{O}\oplus E^{u}_{O}.

  2. (P2)

    (Cross-section and the first return map) The square Σ0={(x1,x2,1):1x1,x21}\Sigma_{0}=\{(x_{1},x_{2},1):-1\leq x_{1},x_{2}\leq 1\} is a cross-section of the flow ϕtX0\phi^{X^{0}}_{t}, meaning that the vector field at every point of Σ0\Sigma_{0} is transverse to Σ0\Sigma_{0}. For simplicity, we assume that the vector field is orthogonal to Σ0\Sigma_{0} and X0(x)=1\|X^{0}(x)\|=1 for any xΣ0x\in\Sigma_{0}. The cross-section Σ0\Sigma_{0} intersects Wlocs(O)W^{s}_{loc}(O) at a line segment L0={(x1,0,1):1x11}L_{0}=\{(x_{1},0,1):-1\leq x_{1}\leq 1\}, which cuts Σ0\Sigma_{0} into a left part Σ0\Sigma_{0}^{-} and a right part Σ0+\Sigma_{0}^{+}, Σ0L0=Σ0Σ0+\Sigma_{0}\setminus L_{0}=\Sigma_{0}^{-}\cup\Sigma_{0}^{+}. There is a first return map R0:Σ0L0Σ0R_{0}:\Sigma_{0}\setminus L_{0}\to\Sigma_{0} such that the images R0(Σ0)R_{0}(\Sigma_{0}^{-}) and R0(Σ0+)R_{0}(\Sigma_{0}^{+}) are each a square pinched at one end and contained in the interior of Σ0\Sigma_{0}, as shown in Figure 1. Let us assume that for any xΣ0L0x\in\Sigma_{0}\setminus L_{0}, the time tx>0t_{x}>0 satisfying ϕtxX(x)=R0(x)\phi^{X}_{t_{x}}(x)=R_{0}(x) for its first return is at least 2, i.e. tx>2t_{x}>2.

    Refer to caption
    Figure 1: The Lorenz attractor
  3. (P3)

    (Cone field on the cross-section) For each α>0\alpha>0, there is a cone Cα(x)C_{\alpha}(x) at the point x=(0,0,1)x=(0,0,1):

    Cα(x)={(x1,x2)TxΣ0:x1αx2}.C_{\alpha}(x)=\{(x_{1},x_{2})\in T_{x}\Sigma_{0}:\|x_{1}\|\leq\alpha\|x_{2}\|\}.

    Here, we have identified TxΣ0T_{x}\Sigma_{0} with Σ0\Sigma_{0} and use the same coordinates x1,x2x_{1},x_{2}. By translating Cα(x)C_{\alpha}(x) to every other point on Σ0\Sigma_{0}, one defines a cone field CαC_{\alpha} on Σ0\Sigma_{0}. We assume that R0R_{0} preserves the cone CαC_{\alpha} with α=1\alpha=1. More precisely, for any xΣ0L0x\in\Sigma_{0}\setminus L_{0} and any vector uC1(x)u\in C_{1}(x), we assume DR0(u)C1/2(R0(x))DR_{0}(u)\in C_{1/2}(R_{0}(x)).

  4. (P4)

    (The attractor) Let Λ0=Cl(ϕtX0(ΛΣ0))\Lambda_{0}={\rm Cl}(\phi^{X^{0}}_{t\in\mathbb{R}}(\Lambda_{\Sigma_{0}})), where ΛΣ0=n0R0¯n(Σ0)\Lambda_{\Sigma_{0}}=\cap_{n\geq 0}\overline{R_{0}}^{n}(\Sigma_{0}) and R0¯n(Σ0)\overline{R_{0}}^{n}(\Sigma_{0}) is the closure of nn-th return of Σ0\Sigma_{0}, ignoring L0L_{0} for each return. Then Λ0\Lambda_{0} contains OO and is the unique attractor in B3B^{3}, with an attracting neighborhood UΛ0B3U_{\Lambda_{0}}\subset B^{3} that contains Σ0\Sigma_{0}.

  5. (P5)

    (Singular hyperbolicity) The attractor Λ0\Lambda_{0} admits a singular hyperbolic splitting TΛ03=EssEcuT_{\Lambda_{0}}{\mathbb{R}}^{3}=E^{ss}\oplus E^{cu}, where EssE^{ss} is uniform contracted and EcuE^{cu} is sectionally expanded. Precisely, we assume that for any two dimensional subspace SExcuS\subset E^{cu}_{x}, xΛ0x\in\Lambda_{0}, it holds

    |det(ΦtX0|S)|>eγt,t1,\left|\det(\Phi^{X^{0}}_{t}|_{S})\right|>e^{\gamma t},\quad\forall t\geq 1, (3.1)

    where γ>0\gamma>0 is a constant. The constant γ\gamma will be assumed to be large so that the inequality (3.1) implies expanding property for the first return map R0R_{0}: there exists ρ>1\rho>1 such that for any xΛ0(Σ0L0)x\in\Lambda_{0}\cap(\Sigma_{0}\setminus L_{0}), vExcu(TxΣ0{0})v\in E^{cu}_{x}\cap(T_{x}\Sigma_{0}\setminus\{0\}), it holds

    DR0(v)>ρv.\|DR_{0}(v)\|>\rho\|v\|.

    As the flow direction X0X^{0} is invariant and can not be uniformly contracted, it is contained in the subbundle EcuE^{cu}, i.e. X0(x)Excu{X^{0}}(x)\in E^{cu}_{x} for every xΛ0Sing(X)x\in\Lambda_{0}\setminus\operatorname{Sing}(X) (see [BGY, Lemma 3.4]). At the singularity OO, we have EOcu=EOcEOuE^{cu}_{O}=E^{c}_{O}\oplus E^{u}_{O}. Define

    λ0s=logDΦ1X0|EΛ0ss.\lambda^{s}_{0}=\log\|D\Phi^{X^{0}}_{1}|_{E^{ss}_{\Lambda_{0}}}\|.

    Note that λsλ0s\lambda^{s}\leq\lambda^{s}_{0}. We assume λ0s<λc\lambda^{s}_{0}<\lambda^{c}.

  6. (P6)

    (Stable foliation) The stable foliation 𝒲s(Λ0)={Ws(Orb(x)):xΛ0}\mathcal{W}^{s}(\Lambda_{0})=\{W^{s}(\operatorname{Orb}(x)):x\in\Lambda_{0}\} is C1C^{1} (see [AM]), which induces a C1C^{1} foliation s\mathcal{F}^{s} on Σ0\Sigma_{0} such that for any xΣ0x\in\Sigma_{0}, s(x)\mathcal{F}^{s}(x) is transverse to C1(x)C_{1}(x). Since L0L_{0} is the intersection of Wlocs(O)W^{s}_{loc}(O) with Σ0\Sigma_{0}, it is a leaf of s\mathcal{F}^{s}.

  7. (P7)

    (Transitivity) The attractor Λ0\Lambda_{0} is a homoclinic class (see [Ba]), so that Λ0\Lambda_{0} is transitive and there is a periodic orbit Q0Λ0Q_{0}\subset\Lambda_{0} whose stable manifold is dense in UΛ0U_{\Lambda_{0}}.

3.2 The skew product construction

Let Ω=B3×I\Omega=B^{3}\times I, where I=[1,1]I=[-1,1]. Let xx, ss be the coordinates on B3B^{3} and II, respectively. Define on Ω\Omega the vector field X^(x,s)=(X0(x),θs)\hat{X}(x,s)=(X^{0}(x),-\theta s), where θ>0\theta>0 is a constant. As the fiber direction (along II) is uniformly contracting, there exists a unique attractor Λ=Λ0×{0}\Lambda=\Lambda_{0}\times\{0\}, which is a trivial construction of the Lorenz attractor in dimension 4. All properties of the 3-dimensional attractor Λ0\Lambda_{0} can be easily generalized to Λ\Lambda. We now modify the vector field X^\hat{X} along fibers to obtain a vector field XX so that Λ\Lambda remains an attractor but not singular hyperbolic. This is why we say the example is derived from the Lorenz attractor. The modification will be done in a small neighborhood of a periodic orbit in Λ\Lambda.

Let P0Λ0P_{0}\subset\Lambda_{0} be a periodic orbit of X0X^{0} other than Q0Q_{0} such that it is homoclinically related to Q0Q_{0}. Let P=P0×{0}P=P_{0}\times\{0\} and Q=Q0×{0}Q=Q_{0}\times\{0\}. Consider a small neighborhood VPV_{P} of PP contained in UΛ0×(1,1)U_{\Lambda_{0}}\times(-1,1) such that VP¯(Q{σ})=\overline{V_{P}}\cap(Q\cup\{\sigma\})=\emptyset, where σ=(O,0)\sigma=(O,0). Let η\eta be a CC^{\infty} function on M0M_{0} that satisfies the following conditions:

  • 0η(x,s)10\leq\eta(x,s)\leq 1, for all (x,s)Ω(x,s)\in\Omega;

  • η\eta is supported on VPV_{P}, i.e. η(x,s)=0\eta(x,s)=0 for any (x,s)VP(x,s)\not\in V_{P};

  • η(x,s)=1\eta(x,s)=1 if and only if (x,s)P(x,s)\in P.

The vector field XX on Ω\Omega is defined as the following:

X(x,s)=(X0(x),θs(1η(x,s))).X(x,s)=(X^{0}(x),-\theta s(1-\eta(x,s))). (3.2)

Observe that the dynamics of the vector field is contracting along the fibers, except only for PP where it is neutral. Also, UΛ=UΛ0×(1,1)U_{\Lambda}=U_{\Lambda_{0}}\times(-1,1) is an attracting region of XX and the maximal invariant set Λ=Λ0×{0}\Lambda=\Lambda_{0}\times\{0\} in UΛU_{\Lambda} is an attractor. The attractor contains a unique singularity σ=(O,0)\sigma=(O,0). The tangent space at σ\sigma admits a ΦtX\Phi^{X}_{t}-invariant splitting TσΩ=TσB3IT_{\sigma}\Omega=T_{\sigma}B^{3}\oplus{\mathbb{R}}^{I}, in which we identify B3B^{3} with B3×{0}B^{3}\times\{0\} and I{\mathbb{R}}^{I} is the subspace corresponding to fiber. Thus, the Lyapunov exponents of σ\sigma are λs,λc,λu\lambda^{s},\lambda^{c},\lambda^{u} and θ-\theta. Note that θ-\theta is the Lyapunov exponent of the singularity along the fiber direction. We assume further that

λ0s<θ<λc.\lambda_{0}^{s}<-\theta<\lambda^{c}. (3.3)

This implies in particular that there is a dominated splitting of the stable subspace Eσs=EssIEcsE^{s}_{\sigma}=E^{ss}\oplus{\mathbb{R}}^{I}\oplus E^{cs}. The following properties can be easily verified:

  1. (C1)

    The cube Σ=Σ0×I\Sigma=\Sigma_{0}\times I is a cross section of the flow ϕtX\phi^{X}_{t}. The local stable manifold of σ\sigma cuts Σ\Sigma along a two-dimensional disk L=L0×IL=L_{0}\times I. So ΣL\Sigma\setminus L consists of a left part Σ=Σ0×I\Sigma^{-}=\Sigma_{0}^{-}\times I and a right part Σ+=Σ0+×I\Sigma^{+}=\Sigma_{0}^{+}\times I. Let R:ΣLΣR:\Sigma\setminus L\to\Sigma be the first return map. Then R(Σ)R(\Sigma^{-}) is a cube pinched at one end and contained in the interior of Σ\Sigma, and similarly for R(Σ+)R(\Sigma^{+}). For any pΣLp\in\Sigma\setminus L, let tp>0t_{p}>0 be the smallest time that ϕtpX(p)=R(p)\phi^{X}_{t_{p}}(p)=R(p). By assumption on R0R_{0}, we have tp>2t_{p}>2 for any pΣLp\in\Sigma\setminus L.

  2. (C2)

    QQ is a hyperbolic periodic orbit contained in Λ\Lambda with stable index 2. The stable manifold Ws(Q,X)W^{s}(Q,X) is dense in UΛU_{\Lambda}. This follows from the fact that the stable manifold of the periodic orbit Q0Q_{0} is dense in UΛ0U_{\Lambda_{0}} and that the tangent flow along the fibers is topologically contracting. Moreover, Ws(Q,X)W^{s}(Q,X) intersects Σ\Sigma along a dense family of 2-dimensional C1C^{1} disks, which are leaves of the foliation s×I\mathcal{F}^{s}\times I. Here, s\mathcal{F}^{s} is the foliation on Σ0\Sigma_{0} as in (P6).

  3. (C3)

    PΛP\subset\Lambda is a non-hyperbolic periodic orbit of XX: it has a zero exponent along the fiber direction. Nonetheless, PP has a 3-dimensional topologically stable manifold and its strong stable manifold Wss(P,X)W^{ss}(P,X) (2-dimensional) has nonempty intersection with the unstable manifold of QQ. Also, its unstable manifold has a transverse intersection with the stable manifold of QQ.

3.3 Robust properties of the example

As a first result, we show that Λ\Lambda admits a partially hyperbolic splitting.

Lemma 3.1.

There exists a partially hyperbolic splitting TΛΩ=EssFT_{\Lambda}\Omega=E^{ss}\oplus F for the tangent flow with dimEss=1\operatorname{dim}E^{ss}=1.

Proof.

Since TΛΩ=TΛ0B3IT_{\Lambda}\Omega=T_{\Lambda_{0}}B^{3}\oplus{\mathbb{R}}^{I} and TΛ0B3T_{\Lambda_{0}}B^{3} admits a partially hyperbolic splitting EssEcuE^{ss}\oplus E^{cu}, one obtains an invariant splitting TΛΩ=EssFT_{\Lambda}\Omega=E^{ss}\oplus F, where F=EcuIF=E^{cu}\oplus{\mathbb{R}}^{I}. Since EssE^{ss} is already dominated by EcuE^{cu}, equation (3.3) implies that EssE^{ss} is dominated by FF. Moreover, ΦtX|Ess\Phi^{X}_{t}|_{E^{ss}} is contracting as it can be identified with ΦtX0|Ess\Phi^{X_{0}}_{t}|_{E^{ss}}. Hence TΛΩ=EssFT_{\Lambda}\Omega=E^{ss}\oplus F is a partially hyperbolic splitting. ∎

Note that the splitting F=EcuIF=E^{cu}\oplus{\mathbb{R}}^{I} is invariant but not dominated. This is because the bundle I{\mathbb{R}}^{I} is neutral on the periodic orbit P=P0×{0}P=P_{0}\times\{0\} and cannot be dominated by the bundle EcuE^{cu} which contains the flow direction, and vice versa. Also, FF is not sectionally expanded. Hence Λ\Lambda is not singular hyperbolic. Nonetheless, the invariance of the splitting TΛΩ=EssEcuIT_{\Lambda}\Omega=E^{ss}\oplus E^{cu}\oplus{\mathbb{R}}^{I} suggests also an invariant splitting of the extended normal bundle with one-dimensional subbundles.

Proposition 3.2.

The extended normal bundle 𝒩~(Λ)\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)} admits a dominated splitting 𝒩ss𝒩I𝒩2\mathcal{N}^{ss}\oplus\mathcal{N}^{I}\oplus\mathcal{N}_{2} with one-dimensional subbundles. Consequently, X|ΛX|_{\Lambda} is C1C^{1} locally away from homoclinic tangencies.

Here, a chain recurrence class is called locally away from homoclinic tangencies if there exists a neighborhood UU of the class and a C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that any Y𝒰Y\in\mathcal{U} admits no homoclinic tangency in UU.

Also, the sectionally expanding subbundle EcuE^{cu} in the splitting TΛΩ=EssEcuIT_{\Lambda}\Omega=E^{ss}\oplus E^{cu}\oplus{\mathbb{R}}^{I} will be shown to exist robustly in the following sense.

Proposition 3.3.

There exist a neighborhood UUΛU\subset U_{\Lambda} of Λ\Lambda and a neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U} and any chain recurrence class CYUC_{Y}\subset U, there is a continuous 2-dimensional subbundle EYE^{Y} of TCYΩT_{C_{Y}}\Omega containing the flow direction such that it is invariant for ΦtY\Phi^{Y}_{t} and sectionally expanded. Moreover, if CY=C(σY,Y)C_{Y}=C(\sigma_{Y},Y) is the chain recurrence class containing σY\sigma_{Y}, then EYE^{Y} varies upper semi-continuously with respect to Y𝒰Y\in\mathcal{U}.

The proof of Proposition 3.2 and Proposition 3.3 will be given in Section 4 and Section 5, respectively.

By construction, the chain recurrence class Λ\Lambda is an attractor containing the unique singularity σ=(O,0)\sigma=(O,0). We will show that for any C1C^{1} vector field YY close enough to XX, the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) is nontrivial and Lyapunov stable. Moreover, it is the only Lyapunov stable chain recurrence class in UΛU_{\Lambda}, where σY\sigma_{Y} is the continuation of the singularity. Precisely, we have

Proposition 3.4.

There exists a C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U}, the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) is the unique Lyapunov stable chain recurrence contained in UΛU_{\Lambda}. Moreover, C(σY,Y)C(\sigma_{Y},Y) contains the periodic orbit QYQ_{Y}, which is the continuation of QQ.

The proof of Proposition 3.4 will be given in Section 6.

3.4 Robust heterodimensional cycles

Let us continue the construction of the example. We will prove the following result.

Lemma 3.5.

In every C1C^{1} neighborhood of XX there is an open subset 𝒱\mathcal{V} of vector fields such that for each Y𝒱Y\in\mathcal{V} the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) contains a robust heterodimensional cycle.

Remark 3.1.

By the lemma, the vector field XX can be C1C^{1} approximated by vector fields with robust heterodimensional cycles. Also, it follows from the construction that XX can be C1C^{1} approximated by vector fields with singular hyperbolic attractors. Therefore, XX is on the boundary between singular hyperbolicity and robust heterodimensional cycles.

The proof of Lemma 3.5 is essentially contained in [BD, Section 4]. We say that a periodic orbit is a saddle-node if its normal bundle has an (orientable) one-dimensional center along which the Lyapunov exponent is zero, and all other Lyapunov exponents are nonzero. For the vector field XX, the periodic orbit PP is a saddle-node with a one-dimensional center in the fiber direction, along which the Lyapunov exponent is zero. The saddle-node PP has a strong stable manifold Wss(P,X)W^{ss}(P,X) contained in B3×{0}B^{3}\times\{0\} (as in the 3-dimensional Lorenz attractor). Also its unstable manifold Wu(P,X)W^{u}(P,X) is contained in B3×{0}B^{3}\times\{0\}. The saddle-node PP has a strong homoclinic intersection: there exists rPr\notin P which belongs to the intersection Wss(P,X)Wu(P,X)W^{ss}(P,X)\cap W^{u}(P,X). Moreover, the intersection is quasi-transverse: dim(TrWss(P,X)TrWu(P,X))=3<dimΩ=4\operatorname{dim}(T_{r}W^{ss}(P,X)\oplus T_{r}W^{u}(P,X))=3<\operatorname{dim}\Omega=4.

Lemma 3.6 ([BD, Theorem 4.1]).

Let XX be a C1C^{1} vector field with a quasi-transverse strong homoclinic intersection associated to a saddle-node. Then in every C1C^{1} neighborhood of XX there is an open set of vector fields exhibiting robust heterodimensional cycles.

It follows from the discussions above and Lemma 3.6 that there exists vector field YY arbitrarily close to XX which exhibits a robust heterodimensional cycle. Thus, to prove Lemma 3.5, it remains to show that the robust heterodimensional cycle is contained in C(σY,Y)C(\sigma_{Y},Y). For this, we need to slightly change the proof of Lemma 3.6 ([BD, Section 4]), in which a key ingredient is the creation of robust heterodimensional cycles through blenders. Roughly speaking, a blender is a hyperbolic set whose stable (or unstable) set behaves as if it is one dimensional higher than it actually is. One can refer to [BDV, Chapter 6] for more discussions on blenders.

Proof of Lemma 3.5.

We have seen that the periodic orbit PP is a saddle-node associated to which there is a quasi-transverse strong homoclinic intersection. In fact, we can take two quasi-transverse strong homoclinic intersections zz and ww associated to PP such that their orbits are disjoint. Following the proof of Lemma 3.6 ([BD, Section 4]), one can make arbitrarily small C1C^{1} perturbations in a neighborhood of Cl(Orb(z,X)Orb(w,X)){\rm Cl}(\operatorname{Orb}(z,X)\cup\operatorname{Orb}(w,X)) (containing PP) to obtain a vector field YY having a blender Λ\Lambda^{\prime}, together with a hyperbolic periodic orbit PδP_{\delta} close to PY=PP_{Y}=P such that:

  • the blender Λ\Lambda^{\prime} is a hyperbolic set of stable index 11 containing the periodic orbit PYP_{Y};

  • the periodic orbit PδP_{\delta} has stable index 22, and its stable manifold intersect transversely the unstable manifold of PYP_{Y};

  • the unstable manifold of PδP_{\delta} meets robustly the stable set of the blender Λ\Lambda^{\prime} (by the property of the blender);

  • thus, the vector field YY has a robust heterodimensional cycle associated with the hyperbolic set ΛPY\Lambda^{\prime}\supset P_{Y} and the hyperbolic saddle PδP_{\delta}.

We can assume that YY is close enough to XX so that by Proposition 3.4, C(σY,Y)C(\sigma_{Y},Y) robustly contains the periodic orbit QYQ_{Y}. Thus, to guarantee that C(σY,Y)C(\sigma_{Y},Y) contains the robust heterodimensional cycle, we make sure that QYQ_{Y} and PYP_{Y} are robustly contained in the same chain recurrence class. A slightly different procedure should be followed when making the perturbations to obtain the robust heterodimensional cycle.

Note that Ws(Q,X)W^{s}(Q,X) intersects Wu(P,X)W^{u}(P,X) transversely. As YY is C1C^{1} close to XX, we can assume that Ws(QY,Y)W^{s}(Q_{Y},Y) and Wu(PY,Y)W^{u}(P_{Y},Y) intersects robustly. Now if Wu(QY,Y)Ws(Pδ,Y)W^{u}(Q_{Y},Y)\pitchfork W^{s}(P_{\delta},Y)\neq\emptyset, the Inclination Lemma would imply that Wu(QY,Y)W^{u}(Q_{Y},Y) accumulates on Wu(Pδ,Y)W^{u}(P_{\delta},Y) and hence meets transversely the characteristic region of the blender Λ\Lambda^{\prime}. By the property of blender and since Ws(QY,Y)W^{s}(Q_{Y},Y) intersects Wu(PY,Y)W^{u}(P_{Y},Y) robustly, we would obtain that QYQ_{Y} and PYP_{Y} are robustly contained in the same chain recurrence class.

Refer to caption
(a) Blow up to twins
Refer to caption
(b) Blow up to triplets
Figure 2: DA-type surgery at the saddle-node PP

Therefore, when perturbing to obtained the periodic orbit PδP_{\delta} we need to make sure that Wu(QY,Y)Ws(Pδ,Y)W^{u}(Q_{Y},Y)\pitchfork W^{s}(P_{\delta},Y)\neq\emptyset. Recall that in [BD, Section 4], PδP_{\delta} is obtained as a twin of PP by blowing up the saddle-node PP along the neutral center, as shown in Figure 2(a). In stead of obtaining only a twin PδP_{\delta}, we blow up the saddle-node to obtain a set of triplet saddles, PP, PδP_{\delta} and PδP_{-\delta}, as shown in Figure 2(b). See also [M1] for a similar construction. The saddles PδP_{\delta} and PδP_{-\delta} are both of index 2, while PP has index 1. Note that for the vector XX, there is a quasi-transverse intersection between Wu(Q,X)W^{u}(Q,X) and Wss(P,X)W^{ss}(P,X). After the blow-up, we can assume without loss of generality that Wu(QY,Y)Ws(Pδ,Y)W^{u}(Q_{Y},Y)\pitchfork W^{s}(P_{\delta},Y)\neq\emptyset. Then the construction continues as in [BD, Section 4]. This ends the proof of Lemma 3.5. ∎

3.5 Proof of Theorem A

Let MM be any 4-dimensional closed Riemannian manifold. One considers a gradient-like vector field Z0Z_{0} on MM, which has at least a sink. Consider a local chart around a sink and a ball B4B^{4} in the chart, centered at the sink. By shrinking the local chart, we assume that the ball B4B^{4} is an attracting region of the flow. In particular, the vector field Z0Z_{0} is transverse to B4\partial B^{4} and inwardly pointing. By changing the metric, we assume that 12B4\frac{1}{2}B^{4} contains Ω=B3×[1,1]\Omega=B^{3}\times[-1,1]. Let XX be a vector field on Ω\Omega as constructed in Section 3.1 and Section 3.2. One then modifies the vector field Z0Z_{0} in a neighborhood of 12B4\frac{1}{2}B^{4} to obtain a C1C^{1} vector field ZZ such that it coincides with XX on Ω12B4\Omega\subset\frac{1}{2}B^{4}. The modification shall be taken outside a neighborhood of MB4M\setminus B^{4}, thus ZZ coincides with Z0Z_{0} in a neighborhood of MB4M\setminus B^{4}. Moreover, since the vector field on Ω\Omega is transverse to each smooth component of Ω\partial\Omega and inwardly pointing, one can require that the obtained vector field ZZ has no recurrence in B4ΩB^{4}\setminus\Omega. Then the chain recurrence set of ZZ is composed of finitely many hyperbolic critical elements and the maximal invariant set in Ω\Omega.

The following properties can be verified.

  1. (1)

    There is nontrivial chain recurrence class C(σ,Z)C(\sigma,Z) associated to a hyperbolic singularity σΩ\sigma\in\Omega.

  2. (2)

    There exist a neighborhood UΩU\subset\Omega of C(σ,Z)C(\sigma,Z) and a neighborhood 𝒰\mathcal{U} of ZZ such that for any Y𝒰Y\in\mathcal{U}, the continuation C(σY,Y)C(\sigma_{Y},Y) is well-defined and contained in UU. In particular, σY\sigma_{Y} is the only singularity in UU. Moreover, since C(σ,Z)C(\sigma,Z) is an attractor, UU can be chosen to be an attracting neighborhood of C(σ,Z)C(\sigma,Z). By reducing 𝒰\mathcal{U}, we assume that UU remains an attracting region for any Y𝒰Y\in\mathcal{U} and the chain recurrent set of YY is composed of finitely many hyperbolic critical elements plus the maximal invariant set in UU.

  3. (3)

    For any Y𝒰Y\in\mathcal{U}, there exists no homoclinic tangency in UU (Proposition 3.2). As the chain recurrent set of YY outside UU is composed of finitely many hyperbolic critical elements, one concludes that YY is away from homoclinic tangencies.

  4. (4)

    There is a partially hyperbolic splitting TC(σY,Y)M=EssFT_{C(\sigma_{Y},Y)}M=E^{ss}\oplus F with dimEss=1\operatorname{dim}E^{ss}=1 (Lemma 3.1). The subbundle FF contains a continuous 2-dimensional subbundle EYTC(σY,Y)ME^{Y}\subset T_{C(\sigma_{Y},Y)}M containing the flow direction such that it is invariant for ΦtY\Phi^{Y}_{t} and sectionally expanded. Moreover, EYE^{Y} varies upper semi-continuously with respect to Y𝒰Y\in\mathcal{U}. (Proposition 3.3)

  5. (5)

    C(σY,Y)C(\sigma_{Y},Y) is the unique Lyapunov stable chain recurrence class contained in UU. (Proposition 3.4)

By Lemma 3.5, there exists an open set 𝒱𝒰\mathcal{V}\subset\mathcal{U} such that for any Y𝒱Y\in\mathcal{V}, the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) exhibits a robust heterodimensional cycle. To finish the proof of Theorem A, it remains to show that for any Y𝒱Y\in\mathcal{V}, the subbundle EYE^{Y} is exceptional. Arguing by contradiction, if EYE^{Y} is not exceptional, then there exists a dominated splitting F=EYEF=E^{Y}\oplus E^{\prime} or F=EEYF=E^{\prime}\oplus E^{Y}. In the first case, since EYE^{Y} is sectionally expanded and is dominated by EE^{\prime}, one deduces that EE^{\prime} is expanded. It follows that C(σY,Y)C(\sigma_{Y},Y) is singular hyperbolic, contradicting to the fact that C(σY,Y)C(\sigma_{Y},Y) contains a robust heterodimensional cycle. In the second case, since EE^{\prime} is dominated by EYE^{Y} which contains the flow direction, one deduces that EE^{\prime} should be contracted. Then C(σY,Y)C(\sigma_{Y},Y) would also be singular hyperbolic, a contradiction again.

This completes the proof of Theorem A.

4   
Away from tangencies: proof of Proposition 3.2

This section is devoted to the proof of Proposition 3.2. Recall that the statement of Proposition 3.2 contains two parts:

  1. (a)

    𝒩~(Λ)\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)} admits a dominated splitting 𝒩ss𝒩I𝒩2\mathcal{N}^{ss}\oplus\mathcal{N}^{I}\oplus\mathcal{N}_{2} with one-dimensional subbundles;

  2. (b)

    the dynamics X|ΛX|_{\Lambda} is locally away from homoclinic tangencies.

Note first that part (b) follows from part (a) and the next lemma.

Lemma 4.1 ([W, Go]).

The dynamics X|ΛX|_{\Lambda} is locally away from homoclinic tangencies if and only if every ii-fundamental sequence of Λ\Lambda admits an index ii dominated splitting.

Part (a) holds if we show that (Λ)=B(Λ)\mathcal{F}(\Lambda)=B(\Lambda) and 𝒩~B(Λ)\widetilde{\mathcal{N}}_{B(\Lambda)} admits a dominated splitting with one dimensional subbundles. These two steps are given below in Lemma 4.2 and Lemma 4.3, respectively. Note that these results are obtained for the unperturbed vector field XX whose dynamics in Ω=B3×I\Omega=B^{3}\times I is a skew-product.

Lemma 4.2.

(Λ)=B(Λ)\mathcal{F}(\Lambda)=B(\Lambda).

Proof.

Since Λ\Lambda is a homoclinic class, periodic orbits are dense in Λ\Lambda. This implies that B(Λ)(Λ)B(\Lambda)\subset\mathcal{F}(\Lambda). We need to show that the converse also holds. By definition of the function ()\mathcal{F}(\cdot), for any l(Λ)l\in\mathcal{F}(\Lambda), we have β(l)Λ\beta(l)\in\Lambda, and if β(l)=x\beta(l)=x is a regular point, then l=lxXl=l^{X}_{x}. Therefore, as σ\sigma is the only singularity in Λ\Lambda, the set (Λ)B(Λ)\mathcal{F}(\Lambda)\setminus B(\Lambda) is contained in TσΩT_{\sigma}\Omega. At the singularity, there is a dominated splitting TσΩ=(EσssI)EσcEσuT_{\sigma}\Omega=(E^{ss}_{\sigma}\oplus{\mathbb{R}}^{I})\oplus E^{c}_{\sigma}\oplus E^{u}_{\sigma}, see (3.3). Let Wss(σ)W^{ss}(\sigma) be the strong stable manifold of σ\sigma that is tangent to EσssIE^{ss}_{\sigma}\oplus{\mathbb{R}}^{I} at σ\sigma. By construction, ΛWss(σ){σ}=\Lambda\cap W^{ss}(\sigma)\setminus\{\sigma\}=\emptyset. Thus it follows from Lemma 2.3 that (Λ)TσΩEσcEσu\mathcal{F}(\Lambda)\cap T_{\sigma}\Omega\subset E^{c}_{\sigma}\oplus E^{u}_{\sigma}. Note that for the 3-dimensional Lorenz attractor Λ0\Lambda_{0}, the intersection of B(Λ0)B(\Lambda_{0}) with TOB3T_{O}B^{3} coincides with EOcu=EOcEOuE^{cu}_{O}=E^{c}_{O}\oplus E^{u}_{O}. Identifying OO with σ=(O,0)\sigma=(O,0) and Λ0\Lambda_{0} with Λ=Λ×{0}\Lambda=\Lambda\times\{0\}, we obtain that B(Λ)=B(Λ0)=EσcEσuB(\Lambda)=B(\Lambda_{0})=E^{c}_{\sigma}\oplus E^{u}_{\sigma}. Thus, (Λ)TσΩB(Λ)TσΩ\mathcal{F}(\Lambda)\cap T_{\sigma}\Omega\subset B(\Lambda)\cap T_{\sigma}\Omega. Consequently, (Λ)B(Λ)\mathcal{F}(\Lambda)\subset B(\Lambda). Therefore, (Λ)=B(Λ)\mathcal{F}(\Lambda)=B(\Lambda). ∎

Lemma 4.3.

𝒩~B(Λ)\widetilde{\mathcal{N}}_{B(\Lambda)} admits a dominated splitting 𝒩ss𝒩I𝒩2\mathcal{N}^{ss}\oplus\mathcal{N}^{I}\oplus\mathcal{N}_{2} with one-dimensional subbundles.

Proof.

For the Lorenz attractor Λ0\Lambda_{0} in B3B^{3}, there exists a singular hyperbolic splitting TΛ0B3=EssEcuT_{\Lambda_{0}}B^{3}=E^{ss}\oplus E^{cu}. For any point xΛ0{O}x\in\Lambda_{0}\setminus\{O\}, one has lxX0Excul^{X^{0}}_{x}\subset E^{cu}_{x}, see (P5). In particular, the domination gives a lower bound of the angle between ExssE^{ss}_{x} and lxX0l^{X^{0}}_{x}, which is uniform for xΛ0{O}x\in\Lambda_{0}\setminus\{O\}. Let 𝒩xss\mathcal{N}^{ss}_{x} be the orthogonal projection of ExssE^{ss}_{x} to 𝒩x\mathcal{N}_{x}, and let 𝒩2(x)=Excu𝒩x\mathcal{N}_{2}(x)=E^{cu}_{x}\cap\mathcal{N}_{x}. The domination property of TΛ0B3=EssEcuT_{\Lambda_{0}}B^{3}=E^{ss}\oplus E^{cu} and the uniform lower bound of the angle between ExssE^{ss}_{x} and lxX0l^{X^{0}}_{x} imply that the bundle 𝒩ss\mathcal{N}^{ss} is also dominated by 𝒩2\mathcal{N}_{2} for the linear Poincaré flow ([BGY, Lemma 2.3]). Moreover, with the contracting property of ΦtX0|Ess\Phi^{X_{0}}_{t}|_{E^{ss}} one can show that the bundle 𝒩ss\mathcal{N}^{ss} is also contracted by the linear Poncaré flow. Taking limits of the splitting 𝒩Λ0=𝒩ss𝒩2\mathcal{N}_{\Lambda_{0}}=\mathcal{N}^{ss}\oplus\mathcal{N}_{2} in the Grassmannian G1G^{1}, one obtains a dominated splitting 𝒩~B(Λ0)=𝒩ss𝒩2\widetilde{\mathcal{N}}_{B(\Lambda_{0})}=\mathcal{N}^{ss}\oplus\mathcal{N}_{2}. Moreover, the subbundle 𝒩ss\mathcal{N}^{ss} is contracting for the extended linear Poincaré flow.

Note that the dynamics in Ω=B3×I\Omega=B^{3}\times I is a skew-product and Λ=Λ0×{0}\Lambda=\Lambda_{0}\times\{0\} is an invariant set for XX. We may assume that the one-dimensional subbundle I{\mathbb{R}}^{I} tangent to the fiber II is orthogonal to T(B3×{0})T(B^{3}\times\{0\}). Identifying B3B^{3} with B3×{0}B^{3}\times\{0\}, one can see that 𝒩~B(Λ)\widetilde{\mathcal{N}}_{B(\Lambda)} admits an invariant splitting 𝒩ss𝒩I𝒩2\mathcal{N}^{ss}\oplus\mathcal{N}^{I}\oplus\mathcal{N}_{2}, where 𝒩ss\mathcal{N}^{ss} and 𝒩2\mathcal{N}_{2} are given by the splitting 𝒩~(B(Λ0))=𝒩ss𝒩2\widetilde{\mathcal{N}}(B(\Lambda_{0}))=\mathcal{N}^{ss}\oplus\mathcal{N}_{2} and 𝒩I=I\mathcal{N}^{I}={\mathbb{R}}^{I}. We need to prove that this splitting is dominated. Following [M2], it suffices to show that for any ergodic invariant measure supported on B(Λ)B(\Lambda), the Lyapunov exponents corresponding to the three subbundles satisfy

ηss<ηI<η2.\eta^{ss}<\eta^{I}<\eta_{2}. (4.1)

Recall from the construction that there is a dominated splitting TσΩ=EssIEcEuT_{\sigma}\Omega=E^{ss}\oplus{\mathbb{R}}^{I}\oplus E^{c}\oplus E^{u}, such that the corresponding Lyapunov exponents of the subbundles are

λs<θ<λc<λu.\lambda^{s}<-\theta<\lambda^{c}<\lambda^{u}.

Moreover, we have defined λ0s=logDΦ1X0|EΛ0ss\lambda^{s}_{0}=\log\|D\Phi_{1}^{X^{0}}|_{E^{ss}_{\Lambda_{0}}}\| which satisfies λsλ0s<θ\lambda^{s}\leq\lambda^{s}_{0}<-\theta. See Section 3.2.

Let μ\mu be any ergodic invariant measure supported on B(Λ)B(\Lambda). Suppose βμ\beta_{*}\mu is the Dirac measure at σ\sigma, then μ\mu is the Dirac measure supported either on Eσc\langle E^{c}_{\sigma}\rangle or on Eσu\langle E^{u}_{\sigma}\rangle. In both cases, we have

ηss=λs<θ=ηI.\eta^{ss}=\lambda^{s}<-\theta=\eta^{I}.

Then we show that ηI<η2\eta^{I}<\eta_{2} as follows: if μ\mu is the Dirac measure on Eσc\langle E^{c}_{\sigma}\rangle, then η2=λu\eta_{2}=\lambda^{u}; otherwise, μ\mu is the Dirac measure on Eσu\langle E^{u}_{\sigma}\rangle, then η2=λc\eta_{2}=\lambda^{c}. In both cases, it follows from equation (3.3) that

ηI=θ<λcη2.\eta^{I}=-\theta<\lambda^{c}\leq\eta_{2}.

Hence (4.1) holds when βμ\beta_{*}\mu is the Dirac measure at σ\sigma.

Now, suppose βμ\beta_{*}\mu is not the Dirac measure at σ\sigma, then it is a nonsingular ergodic measure for the flow ϕtX\phi_{t}^{X}. Since the bundle 𝒩ss\mathcal{N}^{ss} is obtained from EssE^{ss} by orthogonal projection, the Lyapunov exponent along 𝒩ss\mathcal{N}^{ss} is no larger than logDΦ1X0|EΛ0ss=λ0s\log\|D\Phi_{1}^{X^{0}}|_{E^{ss}_{\Lambda_{0}}}\|=\lambda^{s}_{0}, i.e.

ηssλ0s.\eta^{ss}\leq\lambda^{s}_{0}.

By construction of the example, on every point (x,0)Λ(x,0)\in\Lambda, it holds

θζ(x,s)s|s=00,-\theta\leq\frac{\partial\zeta(x,s)}{\partial s}\bigg{|}_{s=0}\leq 0,

where ζ(x,s)=θs(1η(x,s))\zeta(x,s)=-\theta s(1-\eta(x,s)) is the last component of X(x,s)X(x,s). This implies that θηI0-\theta\leq\eta^{I}\leq 0. Then by the inequality (3.3), one has

ηssλ0s<θηI0.\eta^{ss}\leq\lambda^{s}_{0}<-\theta\leq\eta^{I}\leq 0.

Since EcuE^{cu} is sectionally expanding and 𝒩2\mathcal{N}_{2} is obtained from EcuE^{cu} by intersecting with the normal bundle, one deduces that the Lyapunov exponent along 𝒩2\mathcal{N}_{2} is larger than zero, i.e. η2>0\eta_{2}>0. Therefore, (4.1) also holds in this case. This finishes the proof of the lemma. ∎

Remark 4.1.

From the proof, we can see that EΛcu=β((Λ)𝒩2)E^{cu}_{\Lambda}=\beta_{*}(\mathcal{F}(\Lambda)\oplus\mathcal{N}_{2}), where the map β\beta_{*} takes each vector (l,v)β(TM)(l,v)\in\beta^{*}(TM) to vTMv\in TM.

Proof of Proposition 3.2.

By Lemma 4.2 and Lemma 4.3, the extended normal bundle 𝒩~(Λ)\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)} admits a dominated splitting with 1-dimensional subbundles. Then it follows from Lemma 2.4 that any fundamental sequence of Λ\Lambda admits an index ii dominated splitting, for i=1i=1 and i=2i=2. Thus Lemma 4.1 shows that X|ΛX|_{\Lambda} is locally away from homoclinic tangencies. ∎

5   
Existence of sectionally expanding subbundle: proof of Proposition 3.3

By construction, the chain recurrence class Λ\Lambda admits a 2-dimensional invariant bundle EcuTΛΩE^{cu}\subset T_{\Lambda}\Omega which is sectionally expanding for ΦtX\Phi^{X}_{t}. Proposition 3.3 asserts that this property also holds for nearby chain recurrence classes for C1C^{1} perturbations.

This section gives a proof of Proposition 3.3. For this purpose, we turn to the dominated splitting 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} given by Proposition 3.2, where 𝒩1=𝒩ss𝒩I\mathcal{N}_{1}=\mathcal{N}^{ss}\oplus\mathcal{N}^{I}. As in Remark 4.1, we have EΛcu=β((Λ)𝒩2)E^{cu}_{\Lambda}=\beta_{*}(\mathcal{F}(\Lambda)\oplus\mathcal{N}_{2}), which is a two-dimensional sectionally expanding subbundle for the tangent flow.

Definition 5.1.

Let β(TM)\mathcal{E}\subset\beta^{*}(TM) be any invariant bundle for the extended tangent flow Φ~t\tilde{\Phi}_{t}. We say that the bundle \mathcal{E} projects to a continuous bundle ETME\subset TM if EE is continuous and for any lι()l\in\iota(\mathcal{E}), it holds β(l)=Eβ(l)\beta_{*}(\mathcal{E}_{l})=E_{\beta(l)}.

Remark 5.1.

If \mathcal{E} is continuous and it satisfies β(l1)=β(l2)\beta_{*}(\mathcal{E}_{l_{1}})=\beta_{*}(\mathcal{E}_{l_{2}}) for any l1,l2ι()l_{1},l_{2}\in\iota(\mathcal{E}) with β(l1)=β(l2)\beta(l_{1})=\beta(l_{2}), then E=β()E=\beta_{*}(\mathcal{E}) is a well-defined continuous subbundle of TMTM. Moreover, it is easy to see that EE is invariant for the tangent flow Φt\Phi_{t} if \mathcal{E} is invariant for the extended tangent flow Φ~t\tilde{\Phi}_{t}.

Proposition 3.3 will be proved as a corollary of the following result, which is stated in a slightly more general setting.

Proposition 5.2.

Let CXC_{X} be a chain recurrence class of a vector field X𝒳1(M)X\in\mathscr{X}^{1}(M) such that every singularity in the class, say ρ\rho, is Lorenz-like with stable index i0i_{0} and CX(Wss(ρ,X){ρ})=C_{X}\cap(W^{ss}(\rho,X)\setminus\{\rho\})=\emptyset. Suppose 𝒩~(CX)\widetilde{\mathcal{N}}_{\mathcal{F}(C_{X})} admits a dominated splitting 𝒩1𝒩2\mathcal{N}_{1}\oplus\mathcal{N}_{2} such that dim𝒩1=i01\operatorname{dim}\mathcal{N}_{1}=i_{0}-1. Then there is a neighborhood UU of CXC_{X} and a neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U} and any chain recurrence class CYUC_{Y}\subset U of YY, the class admits a dominated splitting 𝒩~(CY)=𝒩1Y𝒩2Y\widetilde{\mathcal{N}}_{\mathcal{F}(C_{Y})}=\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2}, such that (CY)𝒩2Y\mathcal{F}(C_{Y})\oplus\mathcal{N}^{Y}_{2} projects to a continuous bundle EYTCYME^{Y}\subset T_{C_{Y}}M with dimEY=dimMi0+1\operatorname{dim}E^{Y}=\operatorname{dim}M-i_{0}+1.

Proof.

By robustness of dominated splitting, there exist a neighborhood \mathcal{B} of (CX)\mathcal{F}(C_{X}) and a neighborhood 𝒰\mathcal{U} of XX, such that for any Y𝒰Y\in\mathcal{U} and for any invariant set Δ\Delta\subset\mathcal{B} of the extended tangent flow of YY, the normal bundle 𝒩~(Δ)\widetilde{\mathcal{N}}(\Delta) admits a dominated splitting 𝒩1Y𝒩2Y\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2} with respect to the extend linear Poincaré flow ψ~tY\tilde{\psi}_{t}^{Y}, and it satisfies dim𝒩1Y=i01\operatorname{dim}\mathcal{N}^{Y}_{1}=i_{0}-1. By Lemma 2.2, there exists a neighborhood UU of CXC_{X} such that for any Y𝒰Y\in\mathcal{U} (shrinking 𝒰\mathcal{U} if necessary), for any chain recurrence class CYUC_{Y}\subset U, one has (CY)\mathcal{F}(C_{Y})\subset\mathcal{B}. Hence there is a dominated splitting 𝒩~(CY)=𝒩1Y𝒩2Y\widetilde{\mathcal{N}}_{\mathcal{F}(C_{Y})}=\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2} with index i01i_{0}-1, i.e. dim𝒩1Y=i01\operatorname{dim}\mathcal{N}^{Y}_{1}=i_{0}-1.

Since the singularities in CXC_{X} are all Lorenz-like, by reducing UU and 𝒰\mathcal{U}, we can assume that singularities in CYC_{Y} are also Lorenz-like. In particular, for any singularity ρCY\rho\in C_{Y}, there is a dominated splitting TρM=EρssEρcuT_{\rho}M=E^{ss}_{\rho}\oplus E^{cu}_{\rho} such that dimEρss=i01\operatorname{dim}E^{ss}_{\rho}=i_{0}-1. Moreover, by Lemma 2.3 we can assume that for any l(CY)l\in\mathcal{F}(C_{Y}), if β(l)=ρ\beta(l)=\rho is a singularity, then lEρcul\in E^{cu}_{\rho}. As the dominated splitting 𝒩~(CY)=𝒩1Y𝒩2Y\widetilde{\mathcal{N}}_{\mathcal{F}(C_{Y})}=\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2} has index i01i_{0}-1, which equals to dimEρss\operatorname{dim}E^{ss}_{\rho}, the uniqueness of domination implies that 𝒩2Y(l)Eρcu\mathcal{N}^{Y}_{2}(l)\subset E^{cu}_{\rho} and hence β(l𝒩2Y(l))=Eρcu\beta_{*}(l\oplus\mathcal{N}^{Y}_{2}(l))=E^{cu}_{\rho}. Note that the subspace EρcuE^{cu}_{\rho} is unique for all l(CY)β1(ρ)l\in\mathcal{F}(C_{Y})\cap\beta^{-1}(\rho). For any l(CY)l\in\mathcal{F}(C_{Y}) with β(l)=x\beta(l)=x a regular point, we have l=lxYl=l^{Y}_{x}. Thus, we can define Excu=β(l𝒩2(l))E^{cu}_{x}=\beta_{*}(l\oplus\mathcal{N}_{2}(l)). In this way, we define a bundle EcuE^{cu} over CYC_{Y}. By continuity of the map β\beta_{*} and the bundle 𝒩2Y\mathcal{N}^{Y}_{2}, one can see that the bundle EcuE^{cu} is continuous at regular points. For any singularity ρCY\rho\in C_{Y}, let xnx_{n} be a sequence in CYC_{Y} such that xnρx_{n}\to\rho and ln=lxnYll_{n}=l^{Y}_{x_{n}}\to l, we have

Exncu=β(ln𝒩2Y(ln))β(l𝒩2Y(l))=Eρcu.E^{cu}_{x_{n}}=\beta_{*}(l_{n}\oplus\mathcal{N}_{2}^{Y}(l_{n}))\to\beta_{*}(l\oplus\mathcal{N}_{2}^{Y}(l))=E^{cu}_{\rho}.

Hence EcuE^{cu} is also continuous at ρ\rho. Therefore (CY)𝒩2Y\mathcal{F}(C_{Y})\oplus\mathcal{N}^{Y}_{2} projects to the continuous bundle EcuE^{cu} with dimEcu=dimMi0+1\operatorname{dim}E^{cu}=\operatorname{dim}M-i_{0}+1. ∎

Proof of Proposition 3.3.

Let 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} with dim𝒩2=1\operatorname{dim}\mathcal{N}_{2}=1 be the dominated splitting given by Proposition 3.2. Then (Λ)𝒩2\mathcal{F}(\Lambda)\oplus\mathcal{N}_{2} projects to the two-dimensional continuous bundle EΛcuE^{cu}_{\Lambda}, which is sectionally expanded. The neighborhoods UUΛU\subset U_{\Lambda} of Λ\Lambda and 𝒰\mathcal{U} of XX are given by Proposition 5.2. By reducing these neighborhoods, we can assume that for any Y𝒰Y\in\mathcal{U} and any chain recurrence class CYUC_{Y}\subset U, the splitting 𝒩~(CY)=𝒩1Y𝒩2Y\widetilde{\mathcal{N}}_{\mathcal{F}(C_{Y})}=\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2} is close enough to 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2}. This implies that (CY)𝒩2Y\mathcal{F}(C_{Y})\oplus\mathcal{N}^{Y}_{2} projects to a continuous bundle EYE^{Y} that is close enough to EΛcuE^{cu}_{\Lambda}, hence is also sectionally expanded. Moreover, as EYE^{Y} contains β((CY))\beta_{*}(\mathcal{F}(C_{Y})), it contains Y(x)Y(x) for all xCYx\in C_{Y}.

Finally, if CY=C(σY,Y)C_{Y}=C(\sigma_{Y},Y), then CYC_{Y} varies upper semi-continuously with respect to YY. By Lemma 2.2, (CY)\mathcal{F}(C_{Y}) also varies upper semi-continuously with respect to YY. Note that EY=β((CY)𝒩2Y)E^{Y}=\beta_{*}(\mathcal{F}(C_{Y})\oplus\mathcal{N}_{2}^{Y}). As the dominated splitting 𝒩~F(CY)=𝒩1Y𝒩2Y\widetilde{\mathcal{N}}_{F(C_{Y})}=\mathcal{N}^{Y}_{1}\oplus\mathcal{N}^{Y}_{2} varies continuously with respect to YY and the projection β\beta_{*} is continuous, one conclude that EYE^{Y} varies upper semi-continuously with respect to YY. ∎

In the case CYC_{Y} is non-singular, the sectional expanding property of the EYE^{Y} bundle implies that the normal bundle 𝒩2Y\mathcal{N}^{Y}_{2} is expanded. This also holds for any non-singular compact invariant set in UU.

Corollary 5.3.

Let the neighborhoods UU and 𝒰\mathcal{U} be given by Proposition 3.3. For any Y𝒰Y\in\mathcal{U} and any compact ϕtY\phi^{Y}_{t}-invariant set ΛYU\Lambda_{Y}\subset U such that ΛYSing(Y)=\Lambda_{Y}\cap\operatorname{Sing}(Y)=\emptyset, the normal bundle 𝒩ΛY\mathcal{N}_{\Lambda_{Y}} admits a dominated splitting 𝒩1𝒩2\mathcal{N}_{1}\oplus\mathcal{N}_{2} such that 𝒩2\mathcal{N}_{2} is one-dimensional and expanded.

6   
Uniqueness of Lyapunov stable class: proof of Proposition 3.4

In this section we prove Proposition 3.4. In other words, we show that for C1C^{1} perturbations of the vector field XX, any Lyapunov stable chain recurrence class in a small neighborhood of Λ\Lambda contains the singularity σ\sigma and the periodic orbit QQ. We shall make use of the sectional expanding property of the EcuE^{cu} subbundle in the splitting TΛΩ=EssEcuIT_{\Lambda}\Omega=E^{ss}\oplus E^{cu}\oplus{\mathbb{R}}^{I}. The difficulty lies in the fact that EcuE^{cu} is exceptional: there is no domination between the subbundles EcuE^{cu} and I{\mathbb{R}}^{I}. As a bypass, we will consider the dominated splitting 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} given by Proposition 3.2, where 𝒩1=𝒩ss𝒩I\mathcal{N}_{1}=\mathcal{N}^{ss}\oplus\mathcal{N}^{I}.

The rough idea for the proof of Propostion 3.4 goes as follows. We will first extend the dominated splitting 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} to a neighborhood 0\mathcal{B}_{0} of (Λ)\mathcal{F}(\Lambda) and define a cone around 𝒩2\mathcal{N}_{2} on the normal bundle (Section 6.1), then use the cone on the normal bundle to construct an invariant cone on the cross-section Σ\Sigma for the first return map and show that vectors in the cone on Σ\Sigma are expanded by the first return map (Section 6.2 and Section 6.3), and finally finish the proof of the proposition by an analysis of the first return map (Section 6.4).

6.1 Dominated splitting and cone field on the normal bundle

For simplicity, we assume that for the dominated splitting 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} the domination constant T=1T=1. In other words,

ψ~t|𝒩1(l)ψ~t|𝒩2(lt)<12,t1,\|\tilde{\psi}_{t}|_{\mathcal{N}_{1}(l)}\|\cdot\|\tilde{\psi}_{-t}|_{\mathcal{N}_{2}(l_{t})}\|<\frac{1}{2},\quad\forall t\geq 1,

where lt=Φ^t(l)l_{t}=\hat{\Phi}_{t}(l). We extend the splitting 𝒩~(Λ)=𝒩1𝒩2\widetilde{\mathcal{N}}_{\mathcal{F}(\Lambda)}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} to a neighborhood 0\mathcal{B}_{0} of (Λ)\mathcal{F}(\Lambda) in a continuous way. Define on 0\mathcal{B}_{0} a cone field 𝒞α\mathcal{C}_{\alpha} on the normal bundle around the 𝒩2\mathcal{N}_{2} subbundle for any α>0\alpha>0:

𝒞α={v=v1+v2𝒩0:v1𝒩1,v2𝒩2,v1αv2}.\mathcal{C}_{\alpha}=\{v=v_{1}+v_{2}\in\mathcal{N}_{\mathcal{B}_{0}}:v_{1}\in\mathcal{N}_{1},v_{2}\in\mathcal{N}_{2},\|v_{1}\|\leq\alpha\|v_{2}\|\}.

We define also the cone222Technically, 𝒞~α\widetilde{\mathcal{C}}_{\alpha} is not a cone, but a “wedge”. 𝒞~α=0𝒞α\widetilde{\mathcal{C}}_{\alpha}=\mathcal{B}_{0}\oplus\mathcal{C}_{\alpha}. In other words, v𝒞~αv\in\widetilde{\mathcal{C}}_{\alpha} if and only if there is a decomposition v=v1+v2v=v_{1}+v_{2} such that v10v_{1}\in\mathcal{B}_{0} and v2𝒞αv_{2}\in\mathcal{C}_{\alpha}. By domination of the splitting 𝒩1𝒩2\mathcal{N}_{1}\oplus\mathcal{N}_{2} and continuity, we have the following lemma.

Lemma 6.1.

For any α>0\alpha>0, there exist a neighborhood 𝒰\mathcal{U} of XX and a neighborhood 0\mathcal{B}\subset\mathcal{B}_{0} of (Λ)\mathcal{F}(\Lambda) such that for any any Y𝒰Y\in\mathcal{U} and ll\in\mathcal{B}, if Φ^[0,t]Y(l)\hat{\Phi}^{Y}_{[0,t]}(l)\subset\mathcal{B} with t1t\geq 1, then

ψ~tY(𝒞α(l))𝒞α/2(Φ^tY(l))andΦ~tY(𝒞~α(l))𝒞~α/2(Φ^tY(l)).\tilde{\psi}^{Y}_{t}(\mathcal{C}_{\alpha}(l))\subset\mathcal{C}_{\alpha/2}(\hat{\Phi}^{Y}_{t}(l))\quad\text{and}\quad\tilde{\Phi}^{Y}_{t}(\widetilde{\mathcal{C}}_{\alpha}(l))\subset\widetilde{\mathcal{C}}_{\alpha/2}(\hat{\Phi}^{Y}_{t}(l)).

Note that the vector bundle (Λ)𝒩2\mathcal{F}(\Lambda)\oplus\mathcal{N}_{2} is invariant for the extended tangent flow Φ~t\tilde{\Phi}_{t}. By Remark 4.1, β((Λ)𝒩2)=EΛcu\beta_{*}(\mathcal{F}(\Lambda)\oplus\mathcal{N}_{2})=E^{cu}_{\Lambda}, which is sectionally expanded by the tangent flow. Following from the continuity of the map β\beta_{*} and by reducing 0\mathcal{B}_{0}, we can fix α0>0\alpha_{0}>0 small and a neighborhood 𝒰0\mathcal{U}_{0} of XX such that for any Y𝒰0Y\in\mathcal{U}_{0} and l0l\in\mathcal{B}_{0}, if Φ^[0,t]Y(l)0\hat{\Phi}^{Y}_{[0,t]}(l)\subset\mathcal{B}_{0} with t1t\geq 1, then for any 2-dimensional subspace 𝒫𝒞~α0(l)\mathcal{P}\subset\widetilde{\mathcal{C}}_{\alpha_{0}}(l), the projection β(𝒫)\beta_{*}(\mathcal{P}) is close enough to EΛcuE^{cu}_{\Lambda} and hence

|det(ΦtY|β(𝒫))|>eγt/2,\left|\det\left(\Phi^{Y}_{t}|_{\beta_{*}(\mathcal{P})}\right)\right|>e^{\gamma t/2}, (6.1)

where γ>0\gamma>0 is given as in (3.1).

We assume that the results of Lemma 6.1 hold for the fixed α=α0\alpha=\alpha_{0} with 𝒰=𝒰0\mathcal{U}=\mathcal{U}_{0} and =0\mathcal{B}=\mathcal{B}_{0}. Moreover, by Lemma 2.3 we assume that

Wss(σY,Y)C(σY,Y)={σY},for any Y𝒰0,W^{ss}(\sigma_{Y},Y)\cap C(\sigma_{Y},Y)=\{\sigma_{Y}\},\quad\text{for any }\ Y\in\mathcal{U}_{0}, (6.2)

where Wss(σY,Y)W^{ss}(\sigma_{Y},Y) is the strong stable manifold of σY\sigma_{Y} tangent to EσssIE^{ss}_{\sigma}\oplus{\mathbb{R}}^{I}.

6.2 A forward invariance of 0\mathcal{B}_{0}

Recall that there is a cross-section Σ=Σ0×I\Sigma=\Sigma_{0}\times I of the flow ϕtX\phi^{X}_{t}. One can see that Σ\Sigma remains a cross-section for vector fields close to XX. In this section we prove the following lemma.

Lemma 6.2.

There exist a neighborhood U1U_{1} of Λ\Lambda and a C1C^{1} neighborhood 𝒰1\mathcal{U}_{1} of XX, such that for any pΣU1¯p\in\Sigma\cap\overline{U_{1}}, for any Y𝒰1Y\in\mathcal{U}_{1}, and t0t\geq 0, it holds lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0}, where pt=ϕtY(p)p_{t}=\phi^{Y}_{t}(p).

For the proof of Lemma 6.2 we need consider orbit segments arbitrarily close to the singularity. Note that the tangent space at the singularity σ\sigma admits a dominated splitting TσΩ=FσssEσcuT_{\sigma}\Omega=F^{ss}_{\sigma}\oplus E^{cu}_{\sigma}, where Fσss=EσssIF^{ss}_{\sigma}=E^{ss}_{\sigma}\oplus{\mathbb{R}}^{I}; and by construction, it holds (Λ)TσΩ=Eσcu\mathcal{F}(\Lambda)\cap T_{\sigma}\Omega=E^{cu}_{\sigma}. One defines on TσΩT_{\sigma}\Omega a cucu-cone CαcuC^{cu}_{\alpha} around the EσcuE^{cu}_{\sigma} bundle. By domination, the cone CαcuC^{cu}_{\alpha} is forward contracting. We extend the cone continuously to a ball around σ\sigma, which is denoted by W0W_{0}. Let KK be the intersection of t0ϕtX(L)\bigcup_{t\geq 0}\phi^{X}_{t}(L) with W0\partial W_{0}, which is a compact subset of W0\partial W_{0}. We assume that the ball W0W_{0} is small such that W0Σ=W_{0}\cap\Sigma=\emptyset and since the singularity is Lorenz-like, there exists α>0\alpha>0 such that for any pKp\in K, it holds lpXCαcu(p)l^{X}_{p}\subset C^{cu}_{\alpha}(p). By reducing 𝒰0\mathcal{U}_{0} if necessary, there exists a neighborhood NKW0N_{K}\subset\partial W_{0} of KK such that lpYCαcu(p)l^{Y}_{p}\subset C^{cu}_{\alpha}(p) for any pNKp\in N_{K} and Y𝒰0Y\in\mathcal{U}_{0}. Moreover, we assume that the cone CαcuC^{cu}_{\alpha} is forward contracting under the tangent flow ΦtY\Phi^{Y}_{t} for any Y𝒰0Y\in\mathcal{U}_{0}.

Lemma 6.3.

There exists a neighborhood W1W0W_{1}\subset W_{0} of σ\sigma and a neighborhood 𝒰1𝒰0\mathcal{U}_{1}\subset\mathcal{U}_{0} of XX such that for any Y𝒰1Y\in\mathcal{U}_{1}, any orbit segment from a point pΣp\in\Sigma to a point qW1¯q\in\overline{W_{1}} crosses NKN_{K}, and moreover, lqY0l^{Y}_{q}\in\mathcal{B}_{0}.

Proof.

Since Σ\Sigma can be a cross-section for any vector field C1C^{1} close to XX and its image under the first return map remains in the interior of Σ\Sigma, we need only consider orbit segments from pΣp\in\Sigma to qW1¯q\in\overline{W_{1}} without returns to Σ\Sigma. Since NKN_{K} is a neighborhood of KK, there exists a neighborhood W1W0W_{1}\subset W_{0} of σ\sigma such that any orbit segment SS of XX from a point pΣp\in\Sigma to a point qW1¯q\in\overline{W_{1}} (without returns) crosses NKN_{K}. Moreover, let rr be the intersection of SS with NKN_{K}, then the orbit segment from rr to qq is contained in W0¯\overline{W_{0}}. See Figure 3. Shrinking W1W_{1} if necessary, there exists a neighborhood 𝒰1𝒰0\mathcal{U}_{1}\subset\mathcal{U}_{0} of XX such that the same property holds for every Y𝒰1Y\in\mathcal{U}_{1}. The following argument shows that lqY0l^{Y}_{q}\in\mathcal{B}_{0} if we choose W1W_{1} small enough.

Refer to caption
Figure 3: Local analysis around the singularity. Note
that the strong stable direction is 2-dimensional.

Since Eσcu(Λ)E^{cu}_{\sigma}\subset\mathcal{F}(\Lambda) and 0\mathcal{B}_{0} is a neighborhood of (Λ)\mathcal{F}(\Lambda), we can assume that W1W_{1} is small enough and there exists a constant α>0\alpha^{\prime}>0 such that for any qW1¯q\in\overline{W_{1}} and nonzero vector vTqΩv\in T_{q}\Omega, if vCαcu(q)v\in C^{cu}_{\alpha^{\prime}}(q), then v0\langle v\rangle\in\mathcal{B}_{0}; and if the backward orbit of qq hits NKN_{K} at a point rr before leaving W0W_{0}, then the time tt from rr to qq is large enough so that ΦtY(Cαcu(r))Cαcu(q)\Phi^{Y}_{t}(C^{cu}_{\alpha}(r))\subset C^{cu}_{\alpha^{\prime}}(q), by the contracting property of the cone. In particular, since lrYCαcu(r)l^{Y}_{r}\subset C^{cu}_{\alpha}(r), we have lqY=Φ^tY(lrY)Cαcu(q)l^{Y}_{q}=\hat{\Phi}^{Y}_{t}(l^{Y}_{r})\subset C^{cu}_{\alpha^{\prime}}(q). It follows that lqY0l^{Y}_{q}\in\mathcal{B}_{0}. ∎

The next result considers orbit segments away from the singularity.

Lemma 6.4.

Reducing 𝒰1\mathcal{U}_{1} if necessary, there exists an attracting neighborhood U1U_{1} of Λ\Lambda, such that for any Y𝒰1Y\in\mathcal{U}_{1}, one has ϕ1Y(U1¯)U1\phi^{Y}_{1}(\overline{U_{1}})\subset U_{1}. Moreover, for any pU1¯W1p\in\overline{U_{1}}\setminus W_{1}, it holds that lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for all t[0,1]t\in[0,1], where pt=ϕtY(p)p_{t}=\phi^{Y}_{t}(p), and ψ~1Y(𝒞α0(lpY))𝒞α0/2(lp1Y)\tilde{\psi}^{Y}_{1}(\mathcal{C}_{\alpha_{0}}(l^{Y}_{p}))\subset\mathcal{C}_{\alpha_{0}/2}(l^{Y}_{p_{1}}).

Proof.

Reducing 𝒰1\mathcal{U}_{1} if necessary, we can take a neighborhood W1W1W^{\prime}_{1}\subset W_{1} of σ\sigma such that for any pW1p\in W^{\prime}_{1} and any Y𝒰1Y\in\mathcal{U}_{1}, we have ϕtY(p)W1\phi^{Y}_{t}(p)\in W_{1} for all t[1,1]t\in[-1,1]. Since σ\sigma is the only singularity in Λ\Lambda, we have

(Λ)β1(W1)={lpX:pΛW1}.\mathcal{F}(\Lambda)\setminus\beta^{-1}(W^{\prime}_{1})=\{l^{X}_{p}:p\in\Lambda\setminus W^{\prime}_{1}\}. (6.3)

By construction of the vector field XX, there exist arbitrarily small attracting neighborhoods of Λ\Lambda. Following from (6.3), we can take an attracting neighborhood U1U_{1} of Λ\Lambda and reduce the neighborhood 𝒰1\mathcal{U}_{1} of XX such that ϕ1Y(U1¯)U1\phi^{Y}_{1}(\overline{U_{1}})\subset U_{1} and lpY0l^{Y}_{p}\in\mathcal{B}_{0} for any Y𝒰1Y\in\mathcal{U}_{1} and for any pU1¯W1p\in\overline{U_{1}}\setminus W^{\prime}_{1}. This implies that for all pU1¯W1p\in\overline{U_{1}}\setminus W_{1} and t[0,1]t\in[0,1], we have lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0}, and hence by Lemma 6.1, ψ~1Y(𝒞α0(lpY))𝒞α0/2(lp1Y)\tilde{\psi}^{Y}_{1}(\mathcal{C}_{\alpha_{0}}(l^{Y}_{p}))\subset\mathcal{C}_{\alpha_{0}/2}(l^{Y}_{p_{1}}). ∎

We can now finish the proof of Lemma 6.2.

Proof of Lemma 6.2.

Let the neighborhoods 𝒰1\mathcal{U}_{1} of XX and U1U_{1} of Λ\Lambda be given by Lemma 6.3 and Lemma 6.4. Let pp be any point in ΣU1¯\Sigma\cap\overline{U_{1}}. By Lemma 6.4, lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for t[0,1]t\in[0,1]. If p1W1p_{1}\in W_{1}, let t1>1t_{1}>1 be such that the orbit segment ϕ[1,t1]Y(p)W1¯\phi^{Y}_{[1,t_{1}]}(p)\subset\overline{W_{1}} and pt1W1p_{t_{1}}\in\partial W_{1}. Then it follows from Lemma 6.3 that lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for t[1,t1]t\in[1,t_{1}]. If p1W1p_{1}\notin W_{1}, it follows from Lemma 6.4 again that lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for any t[1,2]t\in[1,2]. Inductively, one concludes that lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for all t0t\geq 0. ∎

6.3 Invariant cone on the cross-section and expanding property

Recall that R:ΣLΣR:\Sigma\setminus L\to\Sigma is the first return map, where LL is the intersection of the local stable manifold of σ\sigma with Σ\Sigma (Section 3.2). For any Y𝒰1Y\in\mathcal{U}_{1}, let LY=Wlocs(σY,Y)ΣL_{Y}=W^{s}_{loc}(\sigma_{Y},Y)\cap\Sigma and RY:ΣLYΣR^{Y}:\Sigma\setminus L_{Y}\to\Sigma be the first return map. Let TRYTR^{Y} be the tangent map of RYR^{Y}. For simplicity, let us denote

Σ=ΣU1¯.\Sigma^{*}=\Sigma\cap\overline{U_{1}}.

We define in this section a cone on Σ\Sigma^{*} such that it is invariant under TRYTR^{Y} and moreover, vectors in the cone will be shown to be expanded by TRYTR^{Y}. This allows us to consider curves tangent to the cone on Σ\Sigma^{*} and show that the length of such curves increase under the iteration of the first return map.

In the following, for any 0<αα00<\alpha\leq\alpha_{0} and pU1¯{σY}p\in\overline{U_{1}}\setminus\{\sigma_{Y}\} such that lpY0l^{Y}_{p}\in\mathcal{B}_{0} we denote 𝒞αY(p)=β(𝒞~α(lpY))\mathcal{C}^{Y}_{\alpha}(p)=\beta_{*}(\widetilde{\mathcal{C}}_{\alpha}(l^{Y}_{p})), which is a cone on the tangent space TpΩT_{p}\Omega. We then define a cone field 𝒟αY\mathcal{D}^{Y}_{\alpha} on Σ\Sigma^{*} by letting 𝒟αY(p)=𝒞αY(p)TpΣ\mathcal{D}^{Y}_{\alpha}(p)=\mathcal{C}^{Y}_{\alpha}(p)\cap T_{p}\Sigma. By shrinking U1U_{1} and 𝒰1\mathcal{U}_{1}, the following result holds.

Lemma 6.5.

There exist 0<α1<α00<\alpha_{1}<\alpha_{0} and λ>1\lambda>1 such that for any Y𝒰1Y\in\mathcal{U}_{1}, any pΣLYp\in\Sigma^{*}\setminus L_{Y} and any v𝒟α1X(p)v\in\mathcal{D}^{X}_{\alpha_{1}}(p), one has TRY(v)𝒟α1X(p)TR^{Y}(v)\in\mathcal{D}^{X}_{\alpha_{1}}(p^{\prime}), where p=RY(p)p^{\prime}=R^{Y}(p). Moreover, TRY(v)λv\|TR^{Y}(v)\|\geq\lambda\|v\|.

Proof.

Note that restricted to Σ\Sigma^{*}, the cone 𝒞αY(p)\mathcal{C}^{Y}_{\alpha}(p) and hence 𝒟αY(p)\mathcal{D}^{Y}_{\alpha}(p) are uniformly continuous with respect to the vector field YY in the C1C^{1} topology. Let us fix α1(α0/2,α0)\alpha_{1}\in(\alpha_{0}/2,\alpha_{0}). Then by reducing 𝒰1\mathcal{U}_{1}, we can assume that for any Y𝒰1Y\in\mathcal{U}_{1}, it holds

𝒟α0/2Y(p)𝒟α1X(p)𝒟α0Y(p).\mathcal{D}^{Y}_{\alpha_{0}/2}(p)\subset\mathcal{D}^{X}_{\alpha_{1}}(p)\subset\mathcal{D}^{Y}_{\alpha_{0}}(p).

Also, from (C1) in Section 3.2 we can assume that tp>2t_{p}>2 for any pΣLYp\in\Sigma^{*}\setminus L_{Y}. Now, for any v𝒟α1X(p)𝒟α0Y(p)v\in\mathcal{D}^{X}_{\alpha_{1}}(p)\subset\mathcal{D}^{Y}_{\alpha_{0}}(p), we have (lpY,v)𝒞~α0Y(lpY)(l^{Y}_{p},v)\in\widetilde{\mathcal{C}}^{Y}_{\alpha_{0}}(l^{Y}_{p}). From Lemma 6.2 it follows that lptY0l^{Y}_{p_{t}}\in\mathcal{B}_{0} for all t0t\geq 0. Thus, by Lemma 6.1, we have Φ~tpY(lpY,v)𝒞~α0/2(lpY)\widetilde{\Phi}^{Y}_{t_{p}}(l^{Y}_{p},v)\in\widetilde{\mathcal{C}}_{\alpha_{0}/2}(l^{Y}_{p^{\prime}}), where p=RY(p)p^{\prime}=R^{Y}(p). Equivalently, we have ΦtpY(v)𝒞α0/2Y(p)\Phi^{Y}_{t_{p}}(v)\in\mathcal{C}^{Y}_{\alpha_{0}/2}(p^{\prime}). Projecting ΦtpY(v)\Phi^{Y}_{t_{p}}(v) to TpΣT_{p^{\prime}}\Sigma along lpYl^{Y}_{p^{\prime}}, we obtain that

TRY(v)𝒟α0/2Y(p)𝒟α1X(p).TR^{Y}(v)\in\mathcal{D}^{Y}_{\alpha_{0}/2}(p^{\prime})\subset\mathcal{D}^{X}_{\alpha_{1}}(p^{\prime}).

This establishes the invariance of the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}}.

To see that TRYTR^{Y} is expanding on 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}}, we shall refer to the sectional expanding property (6.1). Recall that in the construction of the example, we have assumed the vector field X0X^{0} to be orthogonal to the cross-section Σ0\Sigma_{0} and X0(x)=1\|X^{0}(x)\|=1 for xΣ0x\in\Sigma_{0}, see (P2) in Section 3.1. Then it follows from (3.2) that the vector field XX is orthogonal to the cross-section Σ\Sigma at points pΣ0×{0}p\in\Sigma_{0}\times\{0\} and X(p)=1\|X(p)\|=1. In particular, this holds for pΛΣp\in\Lambda\cap\Sigma. By shrinking U1U_{1} and 𝒰1\mathcal{U}_{1} if necessary, we can assume that for any Y𝒰1Y\in\mathcal{U}_{1} and any pΣ=ΣU1¯p\in\Sigma^{*}=\Sigma\cap\overline{U_{1}}, Y(p)Y(p) is almost orthogonal to the cross-section and Y(p)\|Y(p)\| is close to 1. Then the sectional expanding property (6.1) allows us to obtain a constant λ>1\lambda>1, independent of YY, such that TRY(v)λv\|TR^{Y}(v)\|\geq\lambda\|v\|. ∎

The previous result shows that the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}} on Σ\Sigma^{*} is invariant for TRYTR^{Y}, for any Y𝒰1Y\in\mathcal{U}_{1}. Moreover, TRYTR^{Y} is expanding on 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}}. Note the first return time tpt_{p} of a point pΣLYp\in\Sigma^{*}\setminus L_{Y} can be arbitrarily large if it is close enough to LYL_{Y}. In this case, the sectional expanding property (6.1) will guarantee a large expansion rate for TRYTR^{Y}.

Lemma 6.6.

There exists a neighborhood NLN_{L} of LL in Σ\Sigma such that for any Y𝒰1Y\in\mathcal{U}_{1} (reducing 𝒰1\mathcal{U}_{1} if necessary), one has LYNLL_{Y}\subset N_{L} and for any p(NLLY)Σp\in(N_{L}\setminus L_{Y})\cap\Sigma^{*},

TRY(v)>3v,v𝒟α1X(p){0}.\|TR^{Y}(v)\|>3\|v\|,\quad\forall v\in\mathcal{D}^{X}_{\alpha_{1}}(p)\setminus\{0\}. (6.4)
Proof.

By Lemma 6.2 and the sectional expanding property (6.1), there exists t0>0t_{0}>0 such that for any Y𝒰1Y\in\mathcal{U}_{1} and any pΣLYp\in\Sigma^{*}\setminus L_{Y}, if the first return time tp>t0t_{p}>t_{0}, then equation (6.4) holds. By continuity of local stable manifold, for any neighborhood NLN_{L} of LL in Σ\Sigma there exists a C1C^{1} neighborhood 𝒰\mathcal{U} of XX such that for any Y𝒰Y\in\mathcal{U}, the intersection LY=Wlocs(σY,Y)ΣL_{Y}=W^{s}_{loc}(\sigma_{Y},Y)\cap\Sigma is contained in NLN_{L}. Therefore, by reducing 𝒰1\mathcal{U}_{1}, we can assume that this property holds for Y𝒰1Y\in\mathcal{U}_{1} and NLN_{L} is small enough such that for any p(NLLY)Σp\in(N_{L}\setminus L_{Y})\cap\Sigma^{*}, the first return time tpt_{p} is larger than t0t_{0}. Hence equation (6.4) holds. ∎

Definition 6.7 (cucu-curve).

A C1C^{1} curve JJ on Σ\Sigma is called a cucu-curve if it is contained in Σ\Sigma^{*} and is tangent to the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}}.

Lemma 6.8.

For any Y𝒰1Y\in\mathcal{U}_{1} and any cucu-curve JJ such that JLY=J\cap L_{Y}=\emptyset, the image RY(J)R^{Y}(J) remains a cucu-curve.

Proof.

Since JLY=J\cap L_{Y}=\emptyset, for any pJp\in J, there is a neighborhood JpJJ_{p}\subset J of pp and a constant δ>0\delta>0 such that the set Γ=|ttp|δϕtY(Jp)\Gamma=\bigcup_{|t-t_{p}|\leq\delta}\phi^{Y}_{t}(J_{p}) is a 2-dimensional submanifold tangent to the cone 𝒞α0Y\mathcal{C}^{Y}_{\alpha_{0}}, transverse to Σ\Sigma. The intersection ΓΣ\Gamma\cap\Sigma is contained in Σ\Sigma^{*} and by Lemma 6.5, is a cucu-curve. This implies that RY(J)R^{Y}(J) is a cucu-curve. ∎

Assuming 𝒰1\mathcal{U}_{1} small enough, the following result holds.

Lemma 6.9.

There exists ε0>0{\varepsilon}_{0}>0 such that for any Y𝒰1Y\in\mathcal{U}_{1} and any cucu-curve JJ, there exists k>0k>0 such that n=0k1(RY)n(J)\bigcup_{n=0}^{k-1}(R^{Y})^{n}(J) contains a cucu-curve with length larger than ε0{\varepsilon}_{0}.

Proof.

Assume that the C1C^{1} neighborhood 𝒰1\mathcal{U}_{1} of XX is small enough such that LYL_{Y} is C1C^{1} close to LL and transverse to 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}} with uniform angle. Moreover, NLN_{L} contains an ε\varepsilon-neighborhood of LYL_{Y} in Σ\Sigma, where ε>0{\varepsilon}>0 is uniform for Y𝒰1Y\in\mathcal{U}_{1}. Then there exists ε0(0,ε){\varepsilon}_{0}\in(0,{\varepsilon}) such that any cucu-curve with length smaller than ε0\varepsilon_{0} intersects LYL_{Y} at most once, and moreover, when it does intersect LYL^{Y} then it is contained in NLN_{L}.

Let JJ be any cucu-curve that does not intersect LYL_{Y}. By Lemma 6.8, RY(J)R^{Y}(J) remains a cucu-curve. Moreover, Lemma 6.5 shows that len(RY(J))>λlen(J)\operatorname{len}(R^{Y}(J))>\lambda\operatorname{len}(J), where len()\operatorname{len}(\cdot) denotes the length of a C1C^{1} curve. By iteration, the length of (RY)n(J)(R^{Y})^{n}(J) keeps growing if it does not intersect LYL_{Y}. When (RY)n(J)(R^{Y})^{n}(J) does intersect LYL_{Y}, then either (RY)n(J)(R^{Y})^{n}(J) has length larger than ε0\varepsilon_{0}, or it is contained in NLN_{L} and is cut by LYL_{Y} into two pieces. In the latter case, let JJ^{\prime} be the longer piece. Then Lemma 6.6 implies that RY(J)R^{Y}(J^{\prime}) has length larger than 3/23/2 times the length of (RY)n(J)(R^{Y})^{n}(J). The conclusion of the lemma holds by induction. ∎

6.4 Proof of Proposition 3.4

We now fix the neighborhoods U1U_{1} and 𝒰1\mathcal{U}_{1} small enough so that all the results on Section 6.2 and Section 6.3 hold.

Recall that there is a periodic orbit QQ in Λ\Lambda with stable index 2 such that its stable manifold is dense in UΛU_{\Lambda}, see (C2) in Section 3.2. In particular, the stable manifold of QQ intersects Σ\Sigma along a dense family of 2-dimensional C1C^{1} disks of the form s(x)×I\mathcal{F}^{s}(x)\times I. As is easy to see that the foliation s×I\mathcal{F}^{s}\times I is transverse to the bundle EcuE^{cu}, we can assume that α0\alpha_{0} is small such that s×I\mathcal{F}^{s}\times I is transverse to the cone 𝒟α0X\mathcal{D}^{X}_{\alpha_{0}}. This implies, in particular, that any cucu-curve is transverse to the foliation s×I\mathcal{F}^{s}\times I and the angle between them is uniformly bounded below. By continuity of local stable manifold, there exists a C1C^{1} neighborhood 𝒰2\mathcal{U}_{2} of XX such that for any Y𝒰2Y\in\mathcal{U}_{2}, any cucu-curve with length ε0{\varepsilon}_{0} intersects the stable manifold Ws(QY,Y)W^{s}(Q_{Y},Y), where ε0>0{\varepsilon}_{0}>0 is given by Lemma 6.9 and QYQ_{Y} is the continuation of QQ.

Lemma 6.10.

For any Y𝒰1𝒰2Y\in\mathcal{U}_{1}\cap\mathcal{U}_{2}, for any ϕtY\phi^{Y}_{t}-invariant compact set ΓU1\Gamma\subset U_{1} such that ΓSing(Y)=\Gamma\cap\operatorname{Sing}(Y)=\emptyset, it holds Wu(Γ,Y)Ws(QY,Y)W^{u}(\Gamma,Y)\cap W^{s}(Q_{Y},Y)\neq\emptyset and Wu(Γ,Y)Ws(σY,Y)W^{u}(\Gamma,Y)\cap W^{s}(\sigma_{Y},Y)\neq\emptyset.

Proof.

By Corollary 5.3, there is a dominated splitting 𝒩Γ=𝒩1𝒩2\mathcal{N}_{\Gamma}=\mathcal{N}_{1}\oplus\mathcal{N}_{2} such that 𝒩2\mathcal{N}_{2} is one-dimensional and expanding. This implies that Γ\Gamma has an unstable manifold tangent to Y𝒩2\langle Y\rangle\oplus\mathcal{N}_{2}. By Lemma 6.1, the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}} is invariant under TRYTR^{Y}, which implies that (lpY𝒩2(p))TpΣ𝒟α1X(p)(l^{Y}_{p}\oplus\mathcal{N}_{2}(p))\cap T_{p}\Sigma\subset\mathcal{D}^{X}_{\alpha_{1}}(p) for any pΓΣp\in\Gamma\cap\Sigma. Then the intersection of the unstable manifold of Orb(p)\operatorname{Orb}(p) with Σ\Sigma contains a cucu-curve JJ. It follows from Lemma 6.9 that the length of JJ can be assumed to be larger than ε0{\varepsilon}_{0}. Since Y𝒰2Y\in\mathcal{U}_{2}, JWs(QY,Y)J\cap W^{s}(Q_{Y},Y)\neq\emptyset. Hence Wu(Γ,Y)Ws(QY,Y)W^{u}(\Gamma,Y)\cap W^{s}(Q_{Y},Y)\neq\emptyset.

For the proof of Wu(Γ,Y)Ws(σY,Y)W^{u}(\Gamma,Y)\cap W^{s}(\sigma_{Y},Y)\neq\emptyset, we need only show that Wu(Γ,Y)LYW^{u}(\Gamma,Y)\cap L_{Y}\neq\emptyset. Assume that JJ does not intersect LYL_{Y}, then by invariance of the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}} and Lemma 6.5, the iterate RY(J)R^{Y}(J) remains a cucu-curve with length larger than λlen(J)\lambda\operatorname{len}(J). By iteration, the length of (RY)n(J)(R^{Y})^{n}(J) keeps growing if the first n1n-1 iterates does not intersects with LYL_{Y}. As the cone 𝒟α1X\mathcal{D}^{X}_{\alpha_{1}} is transverse to s×I\mathcal{F}^{s}\times I, there exists a finite upper bound for the length of cucu-curves. Hence there exists n>0n>0 such that (RY)n(J)LY(R^{Y})^{n}(J)\cap L_{Y}\neq\emptyset. Therefore, Wu(Γ,Y)LYW^{u}(\Gamma,Y)\cap L_{Y}\neq\emptyset. ∎

Corollary 6.11.

For any Y𝒰1𝒰2Y\in\mathcal{U}_{1}\cap\mathcal{U}_{2}, the chain recurrence class C(σY,Y)C(\sigma_{Y},Y) is the only Lyapunov stable one contained in U1U_{1}.

Proof.

Let CYC_{Y} be any Lyapunov stable chain recurrence class of YY contained in U1U_{1}. Suppose that CYC_{Y} does not contain the singularity σY\sigma_{Y}. Then CYC_{Y} is non-singular. By Lemma 6.10, we have Wu(CY,Y)Ws(σY,Y)W^{u}(C_{Y},Y)\cap W^{s}(\sigma_{Y},Y)\neq\emptyset. By Lyapunov stability of the chain recurrence class CYC_{Y}, one would obtain that σYCl(Wu(CY,Y))CY\sigma_{Y}\in{\rm Cl}(W^{u}(C_{Y},Y))\subset C_{Y}, a contradiction. Thus, CYC_{Y} contains σY\sigma_{Y}. It follows that C(σY,Y)C(\sigma_{Y},Y) is the only Lyapunov stable chain recurrence class in U1U_{1}. ∎

Let Y𝒰1𝒰2Y\in\mathcal{U}_{1}\cap\mathcal{U}_{2}, we need to show that QYQ_{Y} is contained in the Lyapunov stable chain recurrence class C(σY,Y)C(\sigma_{Y},Y). Let us recall the following result in [GY].

Lemma 6.12 ([GY, Proposition 4.9]).

Let Δ\Delta be a compact invariant set of Y𝒳1(M)Y\in\mathscr{X}^{1}(M) verifying the following properties:

  • ΔSing(Y)\Delta\setminus{\rm Sing}(Y) admits a dominated splitting 𝒩Δ=GcsGcu\mathcal{N}_{\Delta}=G^{cs}\oplus G^{cu} in the normal bundle w.r.t. the linear Poincaré flow ψt\psi_{t}, with index ii.

  • Every singularity ρΔ\rho\in\Delta is hyperbolic and Ind(ρ)>i\operatorname{Ind}(\rho)>i. Moreover, TρMT_{\rho}M admits a partially hyperbolic splitting TρM=FssFcuT_{\rho}M=F^{ss}\oplus F^{cu} with respect to the tangent flow, where dimFss=i\operatorname{dim}F^{ss}=i and for the corresponding strong stable manifolds Wss(ρ)W^{ss}(\rho), one has Wss(ρ)Δ={ρ}W^{ss}(\rho)\cap\Delta=\{\rho\}.

  • For every xΔx\in\Delta, one has ω(x)Sing(Y)\omega(x)\cap{\rm Sing}(Y)\neq\emptyset.

Then either Δ\Delta admits a partially hyperbolic splitting TΔM=EssFT_{\Delta}M=E^{ss}\oplus F with respect to the tangent flow ΦtY\Phi^{Y}_{t}, where dimEss=i\operatorname{dim}E^{ss}=i, or Δ\Delta intersects the homoclinic class H(γ)H(\gamma) of a hyperbolic periodic orbit γ\gamma of index ii.

Since Y𝒰1Y\in\mathcal{U}_{1}, the singularity σY\sigma_{Y} is Lorenz-like and it follows from Proposition 5.2 that 𝒩C(σY,Y)\mathcal{N}_{C(\sigma_{Y},Y)} admits a dominated splitting 𝒩1𝒩2\mathcal{N}_{1}\oplus\mathcal{N}_{2} with index 22. Moreover, since 𝒰1𝒰0\mathcal{U}_{1}\subset\mathcal{U}_{0} and by (6.2), we have Wss(σY,Y)C(σY,Y)={σY}W^{ss}(\sigma_{Y},Y)\cap C(\sigma_{Y},Y)=\{\sigma_{Y}\}. Let us consider the following two cases:

  1. (1)

    either ω(x)Sing(Y)\omega(x)\cap\operatorname{Sing}(Y)\neq\emptyset, for every xC(σY,Y)x\in C(\sigma_{Y},Y);

  2. (2)

    or there exists xC(σY,Y)x\in C(\sigma_{Y},Y) such that ω(x)Sing(Y)=\omega(x)\cap\operatorname{Sing}(Y)=\emptyset, i.e. σYω(x)\sigma_{Y}\notin\omega(x).

If Case (1) happens, then all the conditions in Lemma 6.12 are satisfied for C(σY,Y)C(\sigma_{Y},Y) and since C(σY,Y)C(\sigma_{Y},Y) can not intersect a homoclinic class, it follows that C(σY,Y)C(\sigma_{Y},Y) admits a partially hyperbolic splitting TC(σY,Y)M=FssFcuT_{C(\sigma_{Y},Y)}M=F^{ss}\oplus F^{cu} with respect to the tangent flow. One can see that the bundle FcuF^{cu} is exactly the sectionally expanding subbundle EYE^{Y} given by Proposition 3.3. Therefore, C(σY,Y)C(\sigma_{Y},Y) is singular hyperbolic, and by Corollary 6.11 it is also Lyapunov stable. By the result [PYY, Corollary D], the class C(σY,Y)C(\sigma_{Y},Y) contains a periodic orbit, which is a contradiction to the assumption that ω(x)Sing(Y)\omega(x)\cap\operatorname{Sing}(Y)\neq\emptyset for all xC(σY,Y)x\in C(\sigma_{Y},Y). So we are left with Case (2), in which C(σY,Y)C(\sigma_{Y},Y) contains a non-singular compact invariant subset Γ\Gamma. Then it follows from Lemma 6.10 that Wu(Γ,Y)Ws(QY,Y)W^{u}(\Gamma,Y)\cap W^{s}(Q_{Y},Y)\neq\emptyset. Hence Wu(C(σY,Y),Y)Ws(QY,Y)W^{u}(C(\sigma_{Y},Y),Y)\cap W^{s}(Q_{Y},Y)\neq\emptyset. Applying Lemma 6.10 to Γ=QY\Gamma=Q_{Y}, we also obtain that Wu(QY,Y)Ws(σY,Y)W^{u}(Q_{Y},Y)\cap W^{s}(\sigma_{Y},Y)\neq\emptyset. Thus QYC(σY,Y)Q_{Y}\subset C(\sigma_{Y},Y).

This ends the proof of Proposition 3.4.

References

  • [ABS] V. Afraimovich, V. Bykov and L. Shilnikov, On the origin and structure of the Lorenz attractor, Dokl. Akad. Nauk SSSR 234 (1977), 336–339; English transl. in Soviet Phys. Dokl. 22 (1977), 253–255.
  • [AM] V. Araújo and I. Melbourne, Existence and smoothness of the stable foliation for sectional hyperbolic attractors, Bulletin of the London Mathematical Society 49 (2017), 351–367.
  • [AP] V. Araújo and M. J. Pacifico, Three-dimensional flows. Ergebnisse der Mathematik und ihrer Grenzgebiete, vol 53. Springer (2010).
  • [ARH] A. Arroyo and F. Rodriguez Hertz, Homoclinic bifurcations and uniform hyperbolicity for three-dimensional flows, Ann. Inst. H. Poincaré - Anal. NonLinéaire, 20 (2003), 805–841.
  • [AS] R. Abraham and S. Smale. Nongenericity of Ω\Omega-stability. Global Analysis I (Proceedings of Symposia in Pure Mathematics, 14). American Mathematical Society, Providence, RI, 1968, pp. 5–8.
  • [Ba] S. Bautista, The geometric Lorenz attractor is a homoclinic class, Bol. Mat. (N.S.) (Dpto. de Matemáticas - Facultad de Ciencias - Universidad Nacional de Colombia) XI(1), 69–78 (2004)
  • [Bo] C. Bonatti, Survey towards a global view of dynamics for the C1C^{1}-topology, Ergodic Theory and Dynamical Systems 31 (2011), 959–993.
  • [BBP] D. Barros, C. Bonatti and M. J. Pacifico, Up, down, two-sided Lorenz attractor, collisions, merging and switching, arXiv:2101.07391v1
  • [BdL] C. Bonatti and A. da Luz, Star flows and multisingular hyperbolicity, arXiv:1705.05799v2
  • [BD] C. Bonatti and L. J. Díaz, Robust heterodimensional cycles and C1C^{1}-generic dynamics, Journal of the Inst. of Math. Jussieu 7 (2008), 469–525.
  • [BDPR] C. Bonatti, L. J. Díaz, E. R. Pujals, and J. Rocha, Robustly transitive sets and heterodimensional cycles, Astérisque 286 (2003), 187–222.
  • [BDV] C. Bonatti, L. J. Díaz, and M. Viana, Dynamics beyond uniform hyperbolicity, Encyclopaedia of Mathematical Sciences (Mathematical Physics), 102, Springer Verlag, (2005).
  • [BGW] C. Bonatti, S. Gan and L. Wen, On the existence of nontrivial homoclinic classes, Ergodic Theory and Dynamical Systems, 27 (2007), 1473–1508.
  • [BGY] C. Bonatti, S. Gan and D. Yang, Dominated chain recurrent class with singularities. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 14 (2015), 83–99.
  • [BLY] C. Bonatti, M. Li and D. Yang, A robustly chain transitive attractor with singularities of different indices. Journal of the Institute of Mathematics of Jussieu, 12 (2013), 449–501.
  • [BV] C. Bonatti and M. Viana, SRB measures for partially hyperbolic systems whose central direction is mostly contracting, Israel Journal of Mathematics 115 (2000), 157–193.
  • [C1] S. Crovisier, Birth of homoclinic intersections: a model for the central dynamics of partially hyperbolic systems, Ann. of Math. 172 (2010), 1641–1677.
  • [C2] S. Crovisier, Partial hyperbolicity far from homoclinic bifurcations, Advances in Math. 226 (2011), 673–726.
  • [CP] S. Crovisier and E. Pujals, Essential hyperbolicity and homoclinic bifurcations: a dichotomy phenomenonmechanism for diffeomorphisms, Inventiones Mathematicae 201 (2015), 385–517.
  • [CSY] S. Crovisier, M. Sambarino and D. Yang, Partial hyperbolicity and homoclinic tangencies, Journal of the European Mathematical Society 17 (2015), 1–49.
  • [CY1] S. Crovisier and D. Yang, Homoclinic tangencies and singular hyperbolicity for three-dimensional vector fields, arXiv:1702.05994v2
  • [CY2] S. Crovisier and D. Yang, Robust transitivity of singular hyperbolic attractors, Mathematische Zeitschrift 298, 469–488 (2021)
  • [dL] A. da Luz, Star flows with singularities of different indices, arXiv:1806.09011v2
  • [D] L. J. Díaz, Persistence of cycles and non-hyperbolic dynamics at heteroclinic bifurcations, Nonlinearity, 8 (1995), 693–715.
  • [Go] N. Gourmelon, Generation of homoclinic tangencies by C1C^{1}-perturbations, Discrete Contin. Dyn. Syst. 26 (2010), 1–42.
  • [GH] J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, Springer, 2002, ISBN 978-1-4612-7020-1
  • [GW] J. Guckenheimer and R. Williams, Structural stability of Lorenz attractors, Inst. Hautes Etudes Sci. Publ. Math. 50 (1979), 59–72.
  • [GY] S. Gan and D. Yang, Morse-Smale systems and horseshoes for three dimensional singular flows, Annales Scientifiques de I École Normale Supérieure, 51(1), 39–112, 2018.
  • [GYZ] S. Gan, J. Yang and R. Zheng, Lyapunov stable chain recurrence classes for singular flows, arXiv:2202.09742
  • [HPS] M. Hirsch, C. Pugh and M. Shub, Invariant manifolds, Lecture Notes in Math., vol 583, Springer Verlar, New York, 1977.
  • [Lo] E. N. Lorenz, Deterministic nonperiodic flow. Journal of the Atmospheric Sciences. 20 (1963), 130–141.
  • [LGW] M. Li, S. Gan and L. Wen, Robustly transitive singular sets via approach of extended linear Poincaré flow. Discrete Contin. Dyn. Syst. 13 (2005), 239–269.
  • [LVY] G. Liao, M. Viana and J. Yang, The entropy conjecture for diffeomorphisms away from tangencies, J. Eur. Math. Soc. 15 (2013), 2043–2060.
  • [M1] R. Mañé, Contributions to the stability conjecture, Topology 17 (1978), 383–396.
  • [M2] R. Mañé, An ergodic closing lemma, Annals of Mathematics, 116 (1982), 503–540.
  • [MPP] C. A. Morales, M. J. Pacifico and E. R. Pujals, Robust transitive singular sets for 3-flows are partially hyperbolic attractors or repellers, Annals of Mathematics, 160 (2004), 375–432.
  • [N1] S. Newhouse, Nondensity of axiom A(a) on S2S^{2}, Global Analysis (Proc. Sympos. Pure Math., Vol. XIV, 1968)(1970), 191–202, Amer. Math. Soc., Providence, R.I.
  • [N2] S. Newhouse, Diffeomorphisms with infinitely many sinks, Topology 13 (1974), 9–18.
  • [Pa1] J. Palis, Homoclinic bifurcations, sensitive-chaotic dynamics and strange attractors. in Dynamical systems and related topics (Nagoya, 1990). Adv. Ser. Dynam. Systems 9 (1991), 466–472.
  • [Pa2] J. Palis, A global view of dynamics and a conjecture of the denseness of finitude of attractors. Astérisque, 261 (2000), 335–347.
  • [Pa3] J. Palis, A global perspective for non-conservative dynamics. Ann. Inst. H. Poincaré Anal. Non Linéaire, 22 (2005), 485–507.
  • [Pa4] J. Palis, Open questions leading to a global perspective in dynamics. Nonliearity, 21 (2008), 37–43.
  • [PS] E. R. Pujals, M. Sambarino, Homoclinic tangencies and hyperbolicity for surface diffeomorphisms. Ann. of math. 151 (2000), 961–1023.
  • [PT] J. Palis and F. Takens, Hyperbolicity and sensitive chaotic dynamics at homoclinic bifurcations, Cambridge studies in advanced mathematics 35, Cambridge University Press (1993).
  • [PYY] M. J. Pacifico, F. Yang and J. Yang, Entropy theory for sectional hyperbolic flows, Annales de l’Institut Henri Poincaré - Analyse non linéaire 38(2021), 1001–1030.
  • [PYY2] M. Pacifico, F. Yang and J. Yang, Equilibrium states for sectional-hyperbolic attractors, arXiv:2209.10784v1
  • [S] S. Smale, Differentiable Dynamical Systems. Bull. Amer. Math. Soc., 73 (1976), 747–817.
  • [SGW] Y. Shi, S. Gan and L. Wen, On the singular-hyperbolicity of star flows, J. of Modern Dynamics, 8(2014), 191–219.
  • [SYY] Y. Shi, F. Yang, and J. Yang, A countable partition for singular flows, and its application on the entropy theory. Isr. J. Math. 249, 375–429 (2022).
  • [T] W. Tucker, The Lorenz attractor exists. C. R. Acad. Sci., Sér I Math., 328 (1999), 1197–1202.
  • [W] L. Wen, Homoclinic tangencies and dominated splittings, Nonlinearity 15 (2002), 1445–1469.
  • [Wi] R. Williams, The “DA” maps of Smale and structural stability. 1970 Global Analysis (Proc. Sympos. Pure Math., Vol. XIV, Berkeley, Calif., 1968) pp. 329–-334

Ming Li, School of Mathematical Sciences, Nankai University, Tianjing, China
E-mail: [email protected]

Fan Yang, Department of Mathematics, Michigan State University, MI, USA
E-mail: [email protected]

Jiagang Yang, Departamento de Geometria, Instituto de Matemática e Estatística, Universidade Federal Fluminense, Niterói, Brazil
E-mail: [email protected]

Rusong Zheng, Joint Research Center on Computational Mathematics and Control, Shenzhen MSU-BIT University, Shenzhen, China
E-mail: [email protected]