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

Robustly Non-Convex Hypersurfaces In Contact Manifolds

Julian Chaidez Department of Mathematics
University of Southern California
Los Angeles, CA
90007
USA
julian.chaidez@usc.edu
Abstract.

We construct the first examples of hypersurfaces in any contact manifold of dimension 5 and larger that cannot be C2C^{2}-approximated by convex hypersurfaces. This contrasts sharply with the foundational result of Giroux in dimension 33 and the work of Honda-Huang in the C0C^{0} case. The main technical step is the construction of a Bonatti-Diaz type blender in the contact setting.

1. Introduction

A hypersurface in a contact manifold (Y,ξ)(Y,\xi) is convex if there is a contact vector-field VV that is transverse to . Convex surface theory was first introduced by Giroux [35], and has since proven to be a deep and powerful tool for the study of contact 33-manifolds. Applications include classifications of contact structures [60, 40, 41, 45, 34, 24, 57, 49] and Legendrians [31, 44, 29, 19]; the construction of 33-manifolds with no tight contact structures [27] and tight, non-fillable contact structures [28]; and finiteness results for tight contact structures [22, 23]. There have also been many fruitful interactions with Floer homology [21, 2, 3, 30, 4, 46]

Recently, the study of higher dimensional convex surface theory was initiated by Honda-Huang [42, 43] and Breen-Honda-Huang [14], who have systematically generalized many of the foundational results of Giroux to dimensions larger than three. One such result is the following.

Theorem 1 (Giroux).

Any closed surface in a contact manifold (Y,ξ)(Y,\xi) in dimension 33 can be CC^{\infty}-approximated by a convex surface .

In [43] (and in Eliashberg-Pancholi [26]) the following partial generalization was proven.

Theorem 2 (Honda-Huang).

Any closed hypersurface in a contact manifold (Y,ξ)(Y,\xi) can be C0C^{0}-approximated by a Weinstein convex hypersurface.

Any convex hypersurface naturally divides into two ideal Liouville manifolds meeting along their boundary, and is called Weinstein if these Liouville manifolds are Weinstein.

Theorem 2 is both stronger than Theorem 1 due to the Weinstein condition, but also weaker since it only provides C0C^{0}-approximations. Indeed, Honda-Huang noted in [43] that the precise generalization of Giroux’s theorem was left unresolved by their work.

Question 3.

[43, Rmk 1.2.4] or [54, Problem 2.1]. Can any closed hypersurface in a contact manifold (Y,ξ)(Y,\xi) be CC^{\infty}-approximated by a convex hypersurface?

Remark 4.

A counter-example to Question 3 was previously proposed by Mori [51, 50] but was later proven to be CC^{\infty}-approximable by Weinstein convex hypersurfaces by Breen [12, Cor 1.8]. A different candidate counter-example (with boundary) remains unverified [12, Rmk 1.9].

The goal of this paper is to resolve this longstanding question by proving the following theorem.

Theorem 5 (Main Theorem).

For any n2n\geq 2, there is a closed hypersurface in standard contact 2n+1 (and thus in any contact (2n+1)(2n+1)-manifold) that cannot be C2C^{2}-approximated by convex hypersurfaces.

Our proof uses a novel combination of constructions in contact topology and partially hyperbolic dynamics. The main idea is to use a dynamical property, namely topological mixing, as an obstruction to convexity. Specifically, we adapt a blender construction pioneered by Bonatti-Diaz [7, 10] to construct C1C^{1}-robustly topologically mixing contactomorphisms, as perturbations of time 11 maps of Anosov flows. We then use suspension and embedding constructions to produce hypersurfaces in 2n+1 with C2C^{2}-robustly mixing characteristic foliations. The characteristic foliation of a convex surface cannot be topologically mixing, so this will prove Theorem 5.

1.1. Characteristic Foliations

Let us briefly recall some dynamical aspects of the structure of hypersurfaces in contact manifolds, before discussing the key results in our proof.

Let be any hypersurface in a contact manifold (Y,ξ)(Y,\xi). Recall that the characteristic foliation ξ of is the (generally singular) oriented, 1-dimensional foliation given by

=ξ(Tξ)ωT{}_{\xi}=(T\Sigma\cap\xi)^{\omega}\subset T\Sigma

Here VωV^{\omega} denotes the symplectic perpendicular of a subspace VξV\subset\xi with respect to the symplectic structure on ξ\xi. Relatedly, a characteristic vector-field ZZ of is a vector-field that spans ξ everywhere and that generates the given orientation.

Any convex surface can be divided into two regions that act as a source and a sink for the flow of any characteristic vector-field for the characteristic foliation.

Definition 6 (Dividing Set).

Let be a convex surface in a contact manifold (Y,ξ)(Y,\xi) with transverse contact vector-field VV. The dividing set with respect to VV is given by

=H1(0)where H=α(V) for any contact form α for ξ\Gamma=H^{-1}(0)\cap\Sigma\quad\text{where }H=\alpha(V)\text{ for any contact form $\alpha$ for $\xi$}

The dividing set is the intersection of the negative region - and positive region + given by the inverse images H1(,0]H^{-1}(-\infty,0]\cap\Sigma and H1[0,)H^{-1}[0,\infty)\cap\Sigma respectively.

The dividing set is always a transversely cutout hypersurface in with a natural contact structure TξT\Gamma\cap\xi [43]. The two regions + and - are ideal Liouville domains with Liouville forms given by the restriction of α\alpha, and the characteristic foliation is given by the span of the corresponding Liouville vector-field. Thus is equipped with a folded symplectic structure [37, 13]. Moreover, this structure is independent of the choices of VV and α\alpha up to isotopy, and is thus canonical up to deformation.

Refer to caption
Figure 1. A convex surface with dividing set and characteristic foliation.

The characteristic foliation ξ of a convex surface is always transverse to the dividing set , pointing out of + and into -, and this has some significant implications for the dynamics. For example, we recall the following definition.

Definition 7 (Topological Mixing).

A smooth flow :×\Phi:\px@BbbR\times\Sigma\to\Sigma is topologically mixing if, for any two non-empty open subsets U,VU,V\subset\Sigma, there is a time TT such that

(U)tVfor all t>T{}_{t}(U)\cap V\neq\emptyset\qquad\text{for all }t>T

We adopt the analogous definition for a diffeomorphism :YY\Phi:Y\to Y of a manifold YY.

Lemma 8.

Let (Y,ξ)\Sigma\subset(Y,\xi) be a closed convex surface and let ZZ be a vector-field that is oriented tangent to ξ. Then ZZ is not topologically mixing.

Proof.

Choose a dividing set on and a pair of disjoint open subsets AA and BB of . Let be the flow of a characteristic vector-field ZZ and consider the subsets

U=(×A)andV=(×B)U=\Phi(\px@BbbR\times A)\quad\text{and}\quad V=\Phi(\px@BbbR\times B)

Any flowline of the characteristic foliation ξ intersects at most once, so UU and VV are open and disjoint. Moreover, (U)tV=UV={}_{t}(U)\cap V=U\cap V=\emptyset for all tt. Therefore is not topologically mixing. ∎

This is the only dynamical property of convex surfaces that we will need for the rest of the paper. Our goal (in view of Lemma 8) is now to construct examples of hypersurfaces that are robustly topologically mixing in the following sense.

Definition 9.

A smooth flow :×\Phi:\px@BbbR\times\Sigma\to\Sigma is robustly mixing if there is a C1C^{1}-open set

𝒰Flow()with𝒰\mathcal{U}\subset\operatorname{Flow}(\Sigma)\quad\text{with}\quad\Phi\in\mathcal{U}

such that any 𝒰\Psi\in\mathcal{U} is topologically mixing. We adopt the analogous definition for a diffeomorphism :YY\Phi:Y\to Y of a manifold YY.

1.2. Suspension

The first step towards this goal is to reduce the problem to a question about contactomorphisms. Fix a contact manifold (Y,ξ)(Y,\xi) and a contactomorphism

:YY\Phi:Y\to Y

We may take the suspension (or mapping torus) to get an even dimensional space

()=[0,1]s×Y/with a hyperplane field η=span(s)ξ\Sigma(\Phi)=[0,1]_{s}\times Y/\sim\qquad\text{with a hyperplane field }\eta=\operatorname{span}(\partial_{s})\oplus\xi

The hyperplane field η\eta is an example of an even contact structure, and so ()\Sigma(\Phi) has the structure of a contact Hamiltonian manifold (also referred to as an even contact manifold [5]). Any such space has a natural characteristic foliation, which in this case is given by

()ξ=span(s)\Sigma(\Phi)_{\xi}=\operatorname{span}(\partial_{s})

In particular, the flow of the characteristic foliation is simply the suspension flow of the contactomorphism . One may also take the contactization

×s(){}_{s}\times\Sigma(\Phi)

such that the characteristic foliation on 0×()0\times\Sigma(\Phi) (as a hypersurface within the contactization) agrees with the intrinsic characteristic foliation ()ξ\Sigma(\Phi)_{\xi}.

Refer to caption
Figure 2. The suspension of a half rotation of the circle Y=S1Y=S^{1}.

As an example, we have depicted the suspension ()\Sigma(\Phi) of a half rotation :S1S1\Phi:S^{1}\to S^{1} of the circle in Figure 2. This is an example of a contactomorphism that gives rise to a non-convex surface in the contactization of its suspension. It is a nice exercise for the reader to find a perturbation of this foliation that satisfies Giroux’s convexity criterion [35, 12].

In Section 2, we will discuss contact Hamiltonian manifolds and the suspension construction in detail. As an elementary application of this construction, we will prove the following result.

Theorem 10 (Proposition 2.17).

Let :YY\Phi:Y\to Y be a robustly mixing contactomorphism. Then ()\Sigma(\Phi) has a C2C^{2}-open neighborhood 𝒰\mathcal{U}, in the space of hypersurfaces in its contactization, such that every 𝒰{}^{\prime}\in\mathcal{U} has topologically mixing characteristic foliation.

1.3. Robustly Mixing Contactomorphisms And Blenders

The next step is the construction of robustly mixing contactomorphisms. This is the subject of our most difficult theorem.

Theorem 11 (Theorem 6.1).

Let (Y,ξ)(Y,\xi) be a closed contact manifold admitting an Anosov Reeb flow with CC^{\infty} stable and unstable foliations. Then the C1C^{1}-open set of robustly mixing contactomorphisms

ContRM(Y,ξ)Cont(Y,ξ)\operatorname{Cont}_{\operatorname{RM}}(Y,\xi)\subset\operatorname{Cont}(Y,\xi)

is non-empty. More precisely, if TT is a (non-zero) multiple of the period of a closed Reeb orbit of , then T is in the CC^{\infty}-closure of the set of robustly mixing contactomorphisms.

Before discussing the proof, we briefly note that there are examples where Theorem 11 applies.

Example 12.

Let XX be a closed nn-manifold with a hyperbolic metric gg. Then the geodesic flow

:×SXSX\Phi:\px@BbbR\times SX\to SX

on the unit cosphere bundle SXSX is Anosov with smooth stable and unstable foliations. These foliations are precisely the quotients of the foliations by the positive and negative unit conormal bundles of the horospheres in n, respectively. Specific constructions of closed hyperbolic manifolds in any dimension 22 or greater can be found in [11, 48, 36].

Theorem 11 is an adaptation of a seminal theorem of Bonatti-Diaz [7, Thm A] to the contact setting, and our main task is to adapt their construction of a certain dynamical structure called a blender. Roughly speaking, a blender is a robust, horseshoe-type structure that forces certain invariant manifolds to be larger than expected. We will give a precise discussion of blenders, along with some background from partially hyperbolic dynamics, in Section 3.

Since their introduction in [7], blenders have become an essential tool in the study of robust and generic properties of smooth dynamical systems. We refer the reader to [6, 10] for a survey on this topic and [9, 8, 47, 53] for just a few examples of their applications. Although constructions of blenders in the symplectic setting have appeared previously (c.f. Nassiri-Pujals [53]), this work is (to our knowledge) the first application of this fundamental dynamical tool to an open problem in symplectic topology.

Our construction of a contact blender will occupy the majority of this paper. The original construction of Bonatti-Diaz [7] does not work without some modifications, and it is quite delicate in certain places, so we have carefully reworked it. We have also included many details that did not appear in [7]. We hope that these details will make this paper more accessible to non-experts in dynamics, e.g. readers with a background primarily in contact topology.

Remark 1.1.

The hypothesis that the Anosov Reeb flow has CC^{\infty} stable and unstable foliation is not present in [7, Thm A] and is quite restrictive. In our proof, we will use it in an essential way to prove one of the axioms of a blender (see Definition 3.22 and Section 5.3). However, we view it as a purely technical hypothesis that can likely be eliminated with more careful analysis.

Refer to caption
Figure 3. A cartoon of the blender appearing in [7] and in this paper, created by the interaction of trajectories between two fixed points (in black). The blue shapes are time evolving balls starting in the green blender box (see Definition 3.22)

1.4. Non-Convex Hypersurfaces

We are now ready to combine the results discussed thus far to prove the main theorem. Already, we can use Lemma 8, Theorems 10-11 and Example 12 to immediately acquire a specific case of our main theorem.

Theorem 13.

The cosphere bundle SXSX of a closed hyperbolic manifold XX has a contactomorphism :XX\Phi:X\to X such that the suspension

()×()\Sigma(\Phi)\subset\px@BbbR\times\Sigma(\Phi)

cannot be C2C^{2}-approximated by a convex hypersurface.

In order to enhance this result to acquire the more general Theorem 5, we apply two difficult theorems. First, we have the following theorem of Sullivan.

Theorem 14.

[56] Every closed hyperbolic manifold WW has a finite cover XX that is stably parallelizable.

We also have the following existence theorem for Legendrian embeddings. This follows from the h-principle of Murphy [52], although this case follows from the earlier h-principle of Gromov.

Theorem 15.

[52] Any closed, stably parallelizable manifold XX has a Legendrian embedding X2n+1X\to{}^{2n+1}.

Finally, we need the following lemma that will be proven in Section 2. Recall that a contactomorphism is positive if it is the time 11 map of a (possibly time-dependent) contact Hamiltonian that is positive as a smooth function.

Lemma 16 (Lemma 2.20).

Let (Y,ξ)\Lambda\subset(Y,\xi) be a closed Legendrian and let :SS\Phi:S\Lambda\to S\Lambda be a positive contactomorphism of the cosphere bundle SS\Lambda. Then there exists a contact embedding

(ϵ,ϵ)×()(Y,ξ)for small ϵ(-\epsilon,\epsilon)\times\Sigma(\Phi)\to(Y,\xi)\qquad\text{for small $\epsilon$}

With these preliminary results in hand, we can now proceed with the proof of the main result.

Theorem 17.

There is a closed, embedded hypersurface 2n+1\Sigma\subset{}^{2n+1} for any n2n\geq 2 that cannot be C2C^{2}-approximated by convex hypersurfaces.

Proof.

Take a stably parallelizable closed hyperbolic manifold XX (via Theorem 14 and Example 12) with a Legendrian embedding X2n+1X\to{}^{2n+1} (via Theorem 15). By Theorem 6.1, there is a robustly mixing contactomorphism

:SXSX\Phi:SX\to SX

that is CC^{\infty}-close to the time TT Reeb flow, where TT is the period of a closed Reeb orbit. Since the time TT Reeb flow is a positive contactomorphism, is also positive. By Lemma 16, there is a contact embedding U2n+1U\to{}^{2n+1} of a neighborhood U×()U\subset\px@BbbR\times\Sigma(\Phi) of the suspension ()\Sigma(\Phi) in its contactization. By Theorem 10, ()\Sigma(\Phi) cannot be C2C^{2}-approximated by a convex surface. ∎

Every contact manifold contains a contact Darboux ball, so this also resolves the general case.

1.5. Open Problems

This work raises many interesting questions at the interface of contact topology and dynamics. We conclude this introduction by mentioning a few of these problems.

Definition 18 (Robust Non-Convexity).

A hypersurface in a contact manifold is robustly non-convex if there is a C2C^{2}-neighborhood 𝒰\mathcal{U} in the space of embedded hypersurfaces such that

 is not convex for any 𝒰{}^{\prime}\text{ is not convex for any }{}^{\prime}\in\mathcal{U}

Similarly, a contactomorphism :YY\Phi:Y\to Y is called robustly non-convex if the suspension ()\Sigma(\Phi) is.

Theorem 5 states that robustly non-convex hypersurfaces exist in any contact manifold of dimension five and higher. It is thus natural to ask about the diversity of such hypersurfaces.

Question 19.

Is every smoothly embedded hypersurface Y\Sigma\to Y in a contact manifold (Y,ξ)(Y,\xi) isotopic to a robustly non-convex one in dimensions 55 and higher?

Our proof of Theorem 5 relies heavily on techniques from partially hyperbolic dynamics and thus may not be applicable to address the full version of Question 19. Indeed, some spaces are known to have no partially hyperbolic diffeomorphisms. For example, we have the following result of Burago-Ivanov.

Theorem 20.

[15] The 33-sphere S3S^{3} does not admit any partially hyperbolic diffeomorphisms.

A result of Bonnati-Diaz-Pujals [9, Thm 2] states that any robustly mixing diffeomorphism must be partially hyperbolic. Thus S3S^{3} admits no robustly mixing diffeomorphisms and the obstructions developed in this paper cannot be applied. This leads naturally to the following question.

Question 21.

Does (S3,ξstd)(S^{3},\xi_{\operatorname{std}}) admit a robustly non-convex contactomorphism?

Question 22.

More generally, is the suspension (IdS)\Sigma(\operatorname{Id}_{S}) of the identity contactomorphism IdS\operatorname{Id}_{S} on S3S^{3} isotopic to a robustly non-convex hypersurface in its contactization?

A negative answer to Question 21 or Question 22 would reveal an interesting connection between the global topology of hypersurfaces and their convex approximability.

Remark 23.

The result [9, Thm 2] of Bonnati-Diaz-Pujals follows from a C1C^{1}-generic dichotomy [9, Thm 1] between maps that admit a dominated splitting and maps with infinitely many sources and sinks (i.e. exhibiting the Newhouse phenomenon), on each homoclinic class.

On the other hand, the methods of Honda-Huang [42, 43] for convexifying hypersurfaces involve the introduction of many new critical points to the characteristic foliation by many small C0C^{0}-perturbations. It is interesting to ask if a hypersurface whose characteristic foliation satisfies a Newhouse-type property along every homoclinic class can be approximated by convex surfaces at some higher-than-expected regularity.

Outline

This concludes the introduction (Section 1) of this paper. In Section 2, we will discuss contact Hamiltonian manifolds (also known as even contact manifolds) and the suspension construction. In Section 3, we will review the necessary background from the theory of partially hyperbolic diffeomorphisms and blenders. In Sections 4 and 5 we will undertake the construction of our contact blender, following Bonatti-Diaz [7]. Finally, we prove Theorem 11 in Section 6.

Acknowledgements

Question 3 was discussed at the workshop Higher-Dimensional Contact Topology at the American Institute of Mathematics in April 2024. We thank the organizers Roger Casals, Yakov Eliashberg, Ko Honda, and Gordana Matic and the AIM staff for a fruitul week.

We also thank the members of the conformal symplectic structures breakout room (Mélanie Bertelson, Kai Celiebak, Fabio Gironella, Pacôme Van Overschelde, Kevin Sackel and Lisa Traynor) for our discussion of this problem, including a very helpful overview of even contact structures and the suspension construction. Finally, we thank Joseph Breen, Kai Celiebak and Austin Christian for input on earlier drafts.

2. Contact Hamiltonian Manifolds

In this brief section, we discuss the theory of contact Hamiltonian manifolds, which are also called even contact manifolds in the terminology of Bertelson-Meigniez [5].

This theory has satisfying parallels with the theory of (stable) Hamiltonian manifolds [20, 59, 17], which motivates our preferred nomenclature. We freely use these two terms as synonyms.

2.1. Fundamentals

We start with the basic facts, which mirror the stable Hamiltonian case.

Definition 2.1.

A contact Hamiltonian manifold (,η)(\Sigma,\eta) is a 2n2n-manifold equipped with a coorientable, maximally non-integrable, plane distribution of codimension one

ηT\eta\subset T\Sigma

Equivalently, η\eta is the kernel of a contact Hamiltonian form ν\nu with νdνn1\nu\wedge d\nu^{n-1} is nowhere vanishing.

Example 2.2 (Product).

Let (Y,ξ)(Y,\xi) be a contact manifold. Then the manifolds

×Yand/×Y\px@BbbR\times Y\qquad\text{and}\qquad\px@BbbR/\px@BbbZ\times Y

are contact Hamiltonian manifolds with distribution η=tξ\eta=\partial_{t}\oplus\xi, where tt is the -coordinate.

Every contact Hamiltonian manifold has a natural line distribution (or equivalently, foliation).

Definition 2.3.

The characteristic foliation η of a contact Hamiltonian manifold (,η)(\Sigma,\eta) is given by

=ηker(dν|η)Tfor any contact Hamiltonian form ν for η{}_{\eta}=\operatorname{ker}(d\nu|_{\eta})\subset T\Sigma\qquad\text{for any contact Hamiltonian form $\nu$ for $\eta$}

A characteristic vector-field ZZ is a section of ξ and a framing form θ\theta is a 11-form whose restriction to ξ is nowhere vanishing. A framing form determines a characteristic vector-field by

(2.1) ν(Z)=0(ιZdν)|η=0andθ(Z)=1\nu(Z)=0\qquad(\iota_{Z}d\nu)|_{\eta}=0\quad\text{and}\quad\theta(Z)=1
Lemma 2.4.

Any characteristic vector-field ZZ on a contact Hamiltonian manifold (,η)(\Sigma,\eta) preserves η\eta.

Proof.

Let be the flow of ZZ. Note that ιZdν=fν\iota_{Z}d\nu=f\nu for some smooth ff since (ιZdν)|η=0(\iota_{Z}d\nu)|_{\eta}=0. Thus

Zν=d(ιZν)+ιZdν=fνand thereforeνt=gtν\mathcal{L}_{Z}\nu=d(\iota_{Z}\nu)+\iota_{Z}d\nu=f\nu\qquad\text{and therefore}\qquad{}_{t}^{*}\nu=g_{t}\cdot\nu\qed
Definition 2.5 (Hamiltonian Vector-fields).

The Reeb vector-field RR of a contact Hamiltonian manifold (,η)(\Sigma,\eta) with contact Hamiltonian form ν\nu and framing θ\theta is the unique vector-field satisfying

θ(R)=0ν(R)=1and(ιRdν)|ξ=0whereξ=ker(θ)\theta(R)=0\qquad\nu(R)=1\quad\text{and}\quad(\iota_{R}d\nu)|_{\xi}=0\quad\text{where}\quad\xi=\operatorname{ker}(\theta)

The Hamiltonian vector-field VHV_{H} of a function H:H:\Sigma\to\px@BbbR is the unique vector-field satisfying

θ(VH)=0ν(VH)=Hand(ιVHdνdH(R)ν+dH)|ξ=0\theta(V_{H})=0\qquad\nu(V_{H})=H\quad\text{and}\quad(\iota_{V_{H}}d\nu-dH(R)\cdot\nu+dH)|_{\xi}=0

Note that the last condition is equivalent to VHν=dH(R)ν+gθ\mathcal{L}_{V_{H}}\nu=dH(R)\cdot\nu+g\cdot\theta given that H=α(VH)H=\alpha(V_{H}).

2.2. Contact Hamiltonian Hypersurfaces

A natural source of contact Hamiltonian hypersurfaces are (special) hypersurfaces in contact manifolds.

Definition 2.6.

A hypersurface in a contact manifold (Y,ξ)(Y,\xi) is contact Hamiltonian called if

ξ is transverse to T\xi\text{ is transverse to }T\Sigma

A framing vector-field UU is a vector-field in a neighborhood of such that

UU is transverse to       and      UU is tangent to ξ\xi
Lemma 2.7.

Let be a contact Hamiltonian hypersurface in (Y,ξ)(Y,\xi) with framing vector-field UU. Then

(,ξT)is contact Hamiltonian with framing formθ=ιUdα(\Sigma,\xi\cap T\Sigma)\quad\text{is contact Hamiltonian with framing form}\quad\theta=\iota_{U}d\alpha
Proof.

Fix a contact form α\alpha on YY. Then dαd\alpha has a 11-dimensional kernel on ker(α)T\operatorname{ker}(\alpha)\cap T\Sigma by standard symplectic linear algebra. It follows that ν=α|T\nu=\alpha|_{T\Sigma} is a contact Hamiltonian form. Similarly, since dαd\alpha is non-degenerate on ξ\xi, we must have

dα(U,Z)0 for any non-vanishing Zηd\alpha(U,Z)\neq 0\text{ for any non-vanishing }Z\in{}_{\eta}\qed

Every contact Hamiltonian manifold arises as a hypersurface in its own contactization.

Definition 2.8.

The contactization (C,α)(C\Sigma,\alpha) of a closed contact Hamiltonian manifold (,η)(\Sigma,\eta) with contact Hamiltonian form ν\nu and framing form θ\theta is given by

C=(ϵ,ϵ)s×with contact form α=sθ+νC\Sigma=(-\epsilon,\epsilon)_{s}\times\Sigma\qquad\text{with contact form }\alpha=s\theta+\nu
Lemma 2.9.

The contactization (C,α)(C\Sigma,\alpha) is a contact manifold for ϵ\epsilon small, and naturally embeds as a contact Hamiltonian hypersurface

=0×Cwithν=α|\Sigma=0\times\Sigma\subset C\Sigma\qquad\text{with}\qquad\nu=\alpha|
Proof.

Note that νdνn1\nu\wedge d\nu^{n-1} is nowhere vanishing and the characteristic vector-field ZZ of θ\theta is a nowhere vanishing vector-field that satisfies

ιZ(νdνn1)=0\iota_{Z}(\nu\wedge d\nu^{n-1})=0

A 11-form μ\mu on thus satisfies μνdνn1\mu\wedge\nu\wedge d\nu^{n-1} if and only if μ(Z)0\mu(Z)\neq 0 everywhere. Therefore

μνdνn1andνdsθdνn1\mu\wedge\nu\wedge d\nu^{n-1}\qquad\text{and}\qquad\nu\wedge ds\wedge\theta\wedge d\nu^{n-1}

are volume forms on and CC\Sigma respectively. The second volume form above agrees with αdαn\alpha\wedge d\alpha^{n} along 0×0\times\Sigma, so there is a neighborhood of 0×0\times\Sigma where αdαn\alpha\wedge d\alpha^{n} is a volume form. ∎

There is a natural way to deform a contact Hamiltonian manifold as a graph in its own contactization (c.f. [18] for a stable Hamiltonian analogue).

Definition 2.10 (Deformation).

The deformation (,ηH)(\Sigma,\eta_{H}) of the contact Hamiltonian manifold (,η)(\Sigma,\eta) by the Hamiltonian H:H:\Sigma\to\px@BbbR is given by

ηH=ker(νH)whereνH=Hθ+ν\eta_{H}=\operatorname{ker}(\nu_{H})\qquad\text{where}\qquad\nu_{H}=H\cdot\theta+\nu

This is precisely the pullback of the induced contact Hamiltonian structure on the graph

GrHCgiven byGrH={(H(x),x):x}\operatorname{Gr}H\subset C\Sigma\qquad\text{given by}\qquad\operatorname{Gr}H=\big{\{}(H(x),x)\;:\;x\in\Sigma\big{\}}

Finally, we note that the contactization provides a local model for the neighborhood of any contact Hamiltonian hypersurface. Specifically, we have the following (strict) standard neighborhood lemma.

Lemma 2.11 (Collar Neighborhood).

Let be a contact Hamiltonian hypersurface in a contact manifold (Y,ξ)(Y,\xi). Fix a contact form α\alpha on YY and a framing vector-field UU of such that

ιUdα is closed\iota_{U}d\alpha\text{ is closed}

Then the flow by UU yields a strict contact embedding

ι:(ϵ,ϵ)×Ywithια=sθ+νwhereν=α| and θ=ιUdα|\iota:(-\epsilon,\epsilon)\times\Sigma\to Y\qquad\text{with}\qquad\iota^{*}\alpha=s\cdot\theta+\nu\quad\text{where}\quad\text{$\nu=\alpha|$ and $\theta=\iota_{U}d\alpha|$}
Proof.

First note that we have the following calculation.

U(ιUdα)=d(dα(U,U))+d(ιUdα)=0\mathcal{L}_{U}(\iota_{U}d\alpha)=d(d\alpha(U,U))+d(\iota_{U}d\alpha)=0

Now let ι:(ϵ,ϵ)s×Y\iota:(-\epsilon,\epsilon)_{s}\times\Sigma\to Y be the tubular neighborhood coordinates of induced by UU. Then the previous calculation and the fact that ιUθ=0\iota_{U}\theta=0 shows that the 11-form

θ=ι(ιUdα)=ιsd(ια)satisfiessθ=0 and θ(s)=0\theta=\iota^{*}(\iota_{U}d\alpha)=\iota_{\partial_{s}}d(\iota^{*}\alpha)\qquad\text{satisfies}\qquad\mathcal{L}_{\partial_{s}}\theta=0\text{ and }\theta(\partial_{s})=0

Thus θ\theta is the pullback of a differential form on to (ϵ,ϵ)×(-\epsilon,\epsilon)\times\Sigma. Moreover, we see that

Uα=d(ιUα)+ιUdα=ιUdαand thussια=θ\mathcal{L}_{U}\alpha=d(\iota_{U}\alpha)+\iota_{U}d\alpha=\iota_{U}d\alpha\qquad\text{and thus}\qquad\mathcal{L}_{\partial_{s}}\iota^{*}\alpha=\theta

It follows that ια\iota^{*}\alpha and sθ+νs\theta+\nu satisfy the same ODE and have the same restriction to 0×0\times\Sigma. Therefore they are equal on the given tubular neighborhood. ∎

2.3. Suspensions

The key examples of contact Hamiltonian manifolds for the purposes of this paper are suspensions of contactomorphisms (or synonymously, mapping tori). This is analogous to the mapping torus construction of stable Hamiltonian manifolds [20, §2.1].

Fix a contact manifold (Y,ξ)(Y,\xi) with a contactomorphism of YY. Recall that the suspension ()\Sigma(\Phi) of is the quotient of ×Y\px@BbbR\times Y by the map

bar:×Y×Ygiven bybar(t,y)=(t1,(y))\bar{\Phi}:\px@BbbR\times Y\to\px@BbbR\times Y\qquad\text{given by}\qquad\bar{\Phi}(t,y)=(t-1,\Phi(y))

Since ξ\xi is preserved by , the contact Hamiltonian structure span(t)ξ\operatorname{span}(\partial_{t})\oplus\xi on the product ×Y\px@BbbR\times Y (see Example 2.2) is bar\bar{\Phi}-invariant. It descends to a contact Hamiltonian structure η\eta on the suspension.

Definition 2.12 (Contact Suspension).

The contact suspension of a contactomorphism :(Y,ξ)(Y,ξ)\Phi:(Y,\xi)\to(Y,\xi) is the contact Hamiltonian manifold given by

((),η)with the framing formθ=dt(\Sigma(\Phi),\eta)\qquad\text{with the framing form}\qquad\theta=dt

The characteristic foliation and a natural framing form are given by the coordinate vector-field and covector-field in the tt-direction.

()η=span(t)andθ=dt\Sigma(\Phi)_{\eta}=\operatorname{span}(\partial_{t})\qquad\text{and}\qquad\theta=dt
Remark 2.13.

In this case, the contact structure on the contactization extends to all of ×()\px@BbbR\times\Sigma(\Phi), and we will refer to this latter space as the contactization.

The most important result of this section is the following lemma, which relates graph-like perturbations of the suspension hypersurface to perturbations of the underlying contactomorphism.

Lemma 2.14 (Hamiltonian Perturbation).

Let H:()H:\Sigma(\Phi)\to\px@BbbR be a smooth function on the suspension of :(Y,ξ)(Y,ξ)\Phi:(Y,\xi)\to(Y,\xi) such that θ=dt\theta=dt frames ηH\eta_{H}. Then there exists a contactomorphism

:H(Y,ξ)(Y,ξ)with an isomorphism:H(()H,η)((),ηH){}^{H}:(Y,\xi)\to(Y,\xi)\qquad\text{with an isomorphism}\qquad{}^{H}:(\Sigma({}^{H}),\eta)\to(\Sigma(\Phi),\eta_{H})

such that distC1(,H)C\|H\|C2\operatorname{dist}_{C^{1}}({}^{H},\Phi)\leq C\cdot\|H\|_{C^{2}} for any Riemannian metric gg and a constant C=C(g)C=C(g).

Proof.

Let ZHZ_{H} denote the characteristic vector-field of ηH\eta_{H} with respect to the framing form dtdt. The characteristic flow H of ZHZ_{H} satisfies (0×Y)tH=t×Y{}^{H}_{t}(0\times Y)=t\times Y since ZH(t)=1Z_{H}(t)=1 and H preserves ηH\eta_{H} by Lemma 2.4. Finally, note that νH\nu_{H} restricts to α\alpha on 0×Y0\times Y and thus ηH(0×Y)\eta_{H}\cap(0\times Y) is ξ\xi. By restriction to ×Y\px@BbbR\times Y where YY is identified with 0×Y0\times Y in ()\Sigma(\Phi), we get a map

:H×Y()with()Hdt=dt and ()HηH=span(s)ξ{}^{H}:\px@BbbR\times Y\to\Sigma(\Phi)\qquad\text{with}\qquad({}^{H})^{*}dt=dt\text{ and }({}^{H})^{*}\eta_{H}=\operatorname{span}(\partial_{s})\oplus\xi

We now define H to be the time 1 map 1H{}^{H}_{1} of the flow. Then the map H satisfies

(s,x)H=(s1,(x)H)H{}^{H}(s,x)={}^{H}(s-1,{}^{H}(x))

In particular, H descends to a map :H()H(){}^{H}:\Sigma({}^{H})\to\Sigma(\Phi) with ()HηH=η({}^{H})^{*}\eta_{H}=\eta. Finally, note that by Definition 2.3, ZHZ_{H} is defined by the formulas

θ(ZH)=1ιZH(Hθ+ν)=0ιZH(dHθ+dν)=0\theta(Z_{H})=1\qquad\iota_{Z_{H}}(H\cdot\theta+\nu)=0\qquad\iota_{Z_{H}}(dH\wedge\theta+d\nu)=0

It follows that there is a smooth linear bundle map

T:(())0(())1(())2T()such thatZH=T(H,ν,dν)T:{}^{0}(\Sigma(\Phi))\oplus{}^{1}(\Sigma(\Phi))\oplus{}^{2}(\Sigma(\Phi))\to T\Sigma(\Phi)\qquad\text{such that}\qquad Z_{H}=T(H,\nu,d\nu)

In particular, for any choice of metric on ()\Sigma(\Phi), there is a constant C>0C>0 and an estimate

\|ZHZ\|C1C\|H\|C1\|Z_{H}-Z\|_{C^{1}}\leq C\cdot\|H\|_{C^{1}}

The same estimate holds for the flow and the time-1 maps. ∎

Example 2.15 (Mapping Torus Of Identity).

Let (Y,ξ)(Y,\xi) be a contact manifold with contact form α\alpha and consider the suspension of the identity

(IdY)=(/)t×Ywith contact Hamiltonian form ν=β\Sigma(\operatorname{Id}_{Y})=(\px@BbbR/\px@BbbZ)_{t}\times Y\qquad\text{with contact Hamiltonian form }\nu=\beta

Fix a Hamiltonian H:/×YH:\px@BbbR/\px@BbbZ\times Y\to\px@BbbR and let VH:/×YTYV_{H}:\px@BbbR/\px@BbbZ\times Y\to TY be the contact vector-field of HH. Consider the deformation

(/×Y,νH)withνH=Hdt+α(\px@BbbR/\px@BbbZ\times Y,\nu_{-H})\qquad\text{with}\qquad\nu_{-H}=-Hdt+\alpha

It is simple to check that in this case the characteristic vector-field for framing form dtdt is given by

ZH=t+VHZ_{H}=\partial_{t}+V_{H}

It follows that the contactomorphism H constructed in Lemma 2.14 is precisely the time 1 map of the contactomorphism generated by H-H and map defines an isomorphism

(()H,η)(/×Y,ηH)(\Sigma({}^{H}),\eta)\simeq(\px@BbbR/\px@BbbZ\times Y,\eta_{-H})

We easily derive the following analogue of Lemma 2.14 for perturbations of hypersurfaces.

Lemma 2.16 (Surface Perturbation).

Let be a contactomorphism. Then there is a C2C^{2}-neighborhood

𝒰Emb(C())of the sub-manifoldι:()C()\mathcal{U}\subset\operatorname{Emb}(C\Sigma(\Phi))\qquad\text{of the sub-manifold}\qquad\iota:\Sigma(\Phi)\to C\Sigma(\Phi)

in the space of embedded smooth sub-manifolds such that any element 𝒰\Sigma\in\mathcal{U} has an isomorphism

:(,η)((),η)wheredistC1(,)CdistC2(,()):(\Sigma,\eta\cap\Sigma)\to(\Sigma(),\eta)\qquad\text{where}\qquad\operatorname{dist}_{C^{1}}(,\Phi)\leq C\cdot\operatorname{dist}_{C^{2}}(\Sigma,\Sigma(\Phi))
Proof.

Any surface in the contactization (ϵ,ϵ)×()(-\epsilon,\epsilon)\times\Sigma(\Phi) that is C2C^{2}-close to 0×()0\times\Sigma(\Phi) is the graph of a function HH on ()\Sigma(\Phi) with C2C^{2}-norm controlled by the C2C^{2}-distance of to 0×0\times\Sigma. Thus this lemma is immediate from Lemma 2.14 and Definition 2.10. ∎

We are now ready to prove Theorem 10 from the introduction, which is very simple.

Proposition 2.17 (Theorem 10).

Let :YY\Phi:Y\to Y be a C1C^{1}-robustly mixing contactomorphism. Then the suspension ()\Sigma(\Phi) has a C2C^{2}-neighborhood

𝒰Emb(C())\mathcal{U}\subset\operatorname{Emb}(C\Sigma(\Phi))

consisting of hypersurfaces with topologically mixing characteristic foliation.

Proof.

By Lemma 2.16, there is a neighborhood 𝒰\mathcal{U} of ()\Sigma(\Phi) such that every 𝒰\Sigma\in\mathcal{U} is the suspension of a contactomorphism that is C1C^{1}-close to . Since is robustly mixing (Definition 9), we can assume after shrinking 𝒰\mathcal{U} that is topologically mixing for any in 𝒰\mathcal{U}. ∎

2.4. Constructions Of Contact-Hamiltonian Hypersurfaces

We conclude this section with constructions of contact Hamiltonian hypersurfaces. The following elementary embedding lemma will be our main tool.

Lemma 2.18 (Disk Neighborhood).

Let be a closed contact manifold and let :\Phi:\Gamma\to\Gamma be a positive contactomorphism. Fix a contact form β\beta and let

H:/×(0,)H:\px@BbbR/\px@BbbZ\times\Gamma\to(0,\infty)

be the corresponding generating Hamiltonian such that =1H\Phi={}^{H}_{1}. Then there is a contact embedding

(2.2) (ϵ,ϵ)s×()(D2×,ar2dθ+β) for any a>12πmaxH(-\epsilon,\epsilon)_{s}\times\Sigma(\Phi)\to(D^{2}\times\Gamma,-a\cdot r^{2}d\theta+\beta)\qquad\text{ for any }a>\frac{1}{2\pi}\cdot\operatorname{max}H
Proof.

We use the following smooth map in radial coordinates.

ι:(2πa,0)s×/×t(D20)×given byι(s,t,x)=((s/2πa)1/2,2πt,x)\iota:(-2\pi a,0)_{s}\times\px@BbbR/{}_{t}\times\Gamma\to(D^{2}-0)\times\Gamma\quad\text{given by}\quad\iota(s,t,x)=((-s/2\pi a)^{1/2},2\pi t,x)

This map satisfies ι(ar2dθ+β)=sdt+β\iota^{*}(-ar^{2}d\theta+\beta)=sdt+\beta. Fix aa satisfying 2πa>maxH2\pi a>\operatorname{max}H. By Example 2.15, the graph of H-H defines an embedding

((),η)(/×,Hdt+β)((2πa,0)s×/×t,sdt+β)(\Sigma(\Phi),\eta)\simeq(\px@BbbR/\px@BbbZ\times\Gamma,-Hdt+\beta)\to((-2\pi a,0)_{s}\times\px@BbbR/{}_{t}\times\Gamma,sdt+\beta)

This embedding extends to a contactomorphism (2.2) by the flow of s\partial_{s} (see Lemma. 2.11). ∎

The next lemma shows that small regions of the jet bundle contains arbitrarily large tubular neighborhoods of the cosphere bundle.

Lemma 2.19.

Let XX be a closed smooth manifold with cosphere bundle SXSX and jet bundle JXJX. Fix a neighborhood UU of XJXX\subset JX and a contact form β\beta on SXSX. Then for any a>0a>0, there is a contact embedding

(D2×SX,ξa,β)Uwhereξa,β=ker(ar2dθ+β)(D^{2}\times SX,\xi_{a,\beta})\to U\qquad\text{where}\qquad\xi_{a,\beta}=\operatorname{ker}(-a\cdot r^{2}d\theta+\beta)
Proof.

Recall that the jet bundle JXJX is given by ×tTX{}_{t}\times T^{*}X with the standard contact form αstd=dt+λstd\alpha_{\operatorname{std}}=dt+\lambda_{\operatorname{std}}. We break the proof into two steps.

Step 1. First assume that U=JXU=JX (so that we may ignore the neighborhood). There is a standard embedding of the symplectization of ×SX\px@BbbR\times SX into TXT^{*}X via the Liouville flow of TXT^{*}X. By using tt-translation, we can extend this to an embedding

κ:×ρ×tSXJXwithκαstd=dt+eρβ\kappa:{}_{\rho}\times{}_{t}\times SX\to JX\qquad\text{with}\qquad\kappa^{*}\alpha_{\operatorname{std}}=dt+e^{\rho}\beta

By applying a further change coordinates by taking s=eρs=-e^{-\rho}, we get a map

ȷ:(,0]s××tSXJXwithȷ(dt+eρβ)=dts1β=s1(sdt+β)\jmath:(-\infty,0]_{s}\times{}_{t}\times SX\to JX\qquad\text{with}\qquad\jmath^{*}(dt+e^{\rho}\beta)=dt-s^{-1}\beta=-s^{-1}\cdot(-s\cdot dt+\beta)

Now we construct an embedding to (,0]s××tSX(-\infty,0]_{s}\times{}_{t}\times SX. Take a disk D(,0]s×tD\subset(-\infty,0]_{s}\times{}_{t} of radius (2a)1/2(2a)^{1/2} centered at a point (s0,t0)(s_{0},t_{0}). We consider the Liouville form

λ=12((tt0)ds(ss0)dt))satisfyingdλ=dsdt\lambda=\frac{1}{2}((t-t_{0})ds-(s-s_{0})dt))\qquad\text{satisfying}\qquad d\lambda=-ds\wedge dt

Next, let τ:2\tau:{}^{2}\to\px@BbbR be a primitive such that λ=sdt+dτ\lambda=-sdt+d\tau and consider the diffeomorphism

:×s×tSX×s×tSXgiven by(s,t,x)=(s,t,(τ(s,t),x)R\Psi:{}_{s}\times{}_{t}\times SX\to{}_{s}\times{}_{t}\times SX\quad\text{given by}\quad\Psi(s,t,x)=(s,t,{}^{R}(\tau(s,t),x)

Here :R×τSXSX{}^{R}:{}_{\tau}\times SX\to SX denote the Reeb flow of β\beta. We compute that

(sdt+β)=sdt+β(R)dτ+()Rβ=sdt+dτ+β=λ+β{}^{*}(-sdt+\beta)=-sdt+\beta(R)\cdot d\tau+({}^{R})^{*}\beta=-sdt+d\tau+\beta=\lambda+\beta

Finally, we compose with the map =ϕ×IdSX\Phi=\phi\times\operatorname{Id}_{SX} where

ϕ:D2×stgiven by(r,θ)(arcos(θ)+s0,arsin(θ)+t0)\phi:D^{2}\to{}_{s}\times{}_{t}\qquad\text{given by}\qquad(r,\theta)\mapsto(ar\cos(\theta)+s_{0},ar\sin(\theta)+t_{0})

The composition \Psi\circ\Phi now restricts to a contact embedding

D2×SX(,0]s××SXwith()(sdt+β)=a2r2dθ+βD^{2}\times SX\to(-\infty,0]_{s}\times\times SX\qquad\text{with}\qquad(\Psi\circ\Phi)^{*}(-s\cdot dt+\beta)=-\frac{a}{2}\cdot r^{2}d\theta+\beta

The composition ȷ:D2×SXJX\jmath\circ\Psi\circ\Phi:D^{2}\times SX\to JX is the desired embedding in the lemma.

Step 2. Now consider the general case where UJXU\subset JX is a proper open set. There is a natural flow of contactomorphisms :×rJXJX{}^{\prime}:{}_{r}\times JX\to JX given by (t,z)rZ=(ert,(z)rZ){}_{r}^{Z}(t,z)=(e^{r}t,{}^{Z}_{r}(z)) where Z is the Liouville flow on TXT^{*}X. This flow is generated by a vector-field U=tt+ZU=t\partial_{t}+Z satisfying

Uαstd=αstd\mathcal{L}_{U}\alpha_{\operatorname{std}}=\alpha_{\operatorname{std}}

In particular, any compact set in JXJX can be pushed into UU by r{}^{\prime}_{r} for rr sufficiently negative. We may then compose the embedding from Step 1 with r{}^{\prime}_{r} to acquire the desired embedding. ∎

By using the Weinstein neighborhood theorem for Legendrians [33] to convert a Legendrian into an embedding of the cosphere bundle of the Legendrian, we acquire the following corollary.

Lemma 2.20 (Lemma 16).

Let (Y,ξ)\Lambda\subset(Y,\xi) be a closed Legendrian sub-manifold and let :SS\Phi:S\Lambda\to S\Lambda be a positive contactomorphism of the cosphere bundle SS\Lambda. Then there is a contact embedding

(ϵ,ϵ)×()(Y,ξ)(-\epsilon,\epsilon)\times\Sigma(\Phi)\to(Y,\xi)
Proof.

By the Weinstein neighborhood theorem, we have a contact embedding

(U,ξstd)(Y,ξ)for a neighborhood UJX of X(U,\xi_{\operatorname{std}})\to(Y,\xi)\qquad\text{for a neighborhood }U\subset JX\text{ of }X

By Lemma 2.18 and Lemma 2.19, for any contact form β\beta on SS\Lambda, there are constants ϵ,a>0\epsilon,a>0 and a contact embedding of the form

((ϵ,ϵ)×(),η)(D2×SX,ξa,β)(U,ξstd)((-\epsilon,\epsilon)\times\Sigma(\Phi),\eta)\to(D^{2}\times SX,\xi_{a,\beta})\to(U,\xi_{\operatorname{std}})\qed

3. Partial Hyperbolicity

In this section, we review the theory of partially hyperbolic maps and blenders.

Remark 3.1.

This section contains extensive background aimed at non-experts in dynamics. We recommend that the reader look to Crovisier-Potrie [25], Hertz-Hertz-Ures [38], Hirsch-Pugh-Shub [39] or Bonatti-Diaz-Viana [10] for a more comprehensive treatment. We also recommend the excellent book of Fisher-Hasselblatt [32] for an accessible treatment of hyperbolic dynamics.

3.1. Fundamentals

Fix a compact smooth manifold YY and a diffeomorphism

:YY\Phi:Y\to Y
Definition 3.2 (Expansion/Contraction).

A sub-bundle ETYE\subset TY is uniformly expanding with respect to and a Riemannian metric gg on YY if there are constants C>0C>0 and λ>1\lambda>1 such that

|T(u)n|Cλn|u|for all n0|T{}^{n}(u)|\geq C\cdot\lambda^{n}\cdot|u|\qquad\text{for all }n\geq 0

Similarly, EE is uniformly contracting for and gg if it is uniformly expanding for -1 and gg.

Definition 3.3 (Domination).

A continuous splitting of TYTY into continuous sub-bundles

TY=E1EkTY=E_{1}\oplus\dots\oplus E_{k}

is dominated with respect to and a metric gg if there are constants C>0C>0 and λ>1\lambda>1 such that

|T(u)n|Cλn|T(v)n|for all n0 and any unit vectors uEi+1 and vEi|T{}^{n}(u)|\geq C\cdot\lambda^{n}\cdot|T{}^{n}(v)|\qquad\text{for all }n\geq 0\text{ and any unit vectors }u\in E_{i+1}\text{ and }v\in E_{i}

The constant λ\lambda is the constant of dilation and the splitting may also be called λ\lambda-dominated.

Dominated splittings obey several fundamental properties (cf. [10, Appendix B] or [25]) that we now discuss briefly. First, dominated splittings are persistent with respect to C1C^{1}-perturbation.

Theorem 3.4 (C1C^{1}-Persistence).

Let :YY\Phi:Y\to Y be a diffeomorphism with a λ\lambda-dominated splitting

TY=E1EkTY=E_{1}\oplus\dots\oplus E_{k}

Then for μ<λ\mu<\lambda, there is a C1C^{1}-neighborhood 𝒰\mathcal{U} of such that every 𝒰\Psi\in\mathcal{U} has a μ\mu-dominated splitting

TY=E1()Ek()such thatdimEi()=dimEi()TY=E_{1}(\Psi)\oplus\dots\oplus E_{k}(\Psi)\qquad\text{such that}\qquad\operatorname{dim}E_{i}(\Psi)=\operatorname{dim}E_{i}(\Phi)

Next, dominated splittings with an expanding (or contracting) factor possess a unique, expanding (or contracting) invariant foliation (cf. [39] and [10, Thm B.7]).

Theorem 3.5 (Foliations).

[39] Let :YY\Phi:Y\to Y be a diffeomorphism with a dominated splitting

TY=DEwith E uniformly expandingTY=D\oplus E\quad\text{with $E$ uniformly expanding}

Then there is a unique Hölder foliation FF with smooth leaves such that

F is tangent to EF=FandF(P)rk(E)\text{$F$ is tangent to $E$}\quad\qquad{}_{*}F=F\qquad\text{and}\qquad F(P)\simeq{}^{\operatorname{rk}(E)}

Finally, the stable manifold theorem (c.f. Hirsch-Pugh-Shub [39, Thm 4.1]) asserts the existence of stable and unstable manifolds for hyperbolic invariant sets.

Theorem 3.6 (Invariant Manifolds).

Let be an invariant sub-manifold of a diffeomorphism :YY\Phi:Y\to Y with normal hyperbolic splitting EuTEsE^{u}\oplus T\Gamma\oplus E^{s}. Then there are unique invariant sub-manifolds

Ws(,)andWu(,)containing W^{s}(\Gamma,\Phi)\qquad\text{and}\qquad W^{u}(\Gamma,\Phi)\qquad\text{containing }\Gamma

that are locally invariant near , C1C^{1}-robust and that satisfy

TWs(,)=EsTandTWu(,)=EuTalong TW^{s}(\Phi,\Gamma)=E^{s}\oplus T\Gamma\qquad\text{and}\qquad TW^{u}(\Phi,\Gamma)=E^{u}\oplus T\Gamma\qquad\text{along }\Gamma
Notation 3.7 (Local Invariant Manifolds).

Given a normally hyperbolic invariant set and an open neighborhood UU of we will use the notation

Wlocs(,;U)andWlocu(,;U)W^{s}_{\operatorname{loc}}(\Gamma,\Phi;U)\qquad\text{and}\qquad W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U)

to denote the respective components of Ws(,)UW^{s}(\Gamma,\Phi)\cap U and W(u,)UW^{u}_{(}\Gamma,\Phi)\cap U that contain . We adopt analogous notation for the leaves of the foliations Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) when defined.

This paper will be entirely oriented towards the following class of diffeomorphisms.

Definition 3.8 (Partially Hyperbolic).

A diffeomorphism :YY\Phi:Y\to Y is partially hyperbolic if there is a splitting the tangent bundle into -invariant, continuous sub-bundles

TY=Es()Ec()Eu()TY=E^{s}(\Phi)\oplus E^{c}(\Phi)\oplus E^{u}(\Phi)

such that Es()E^{s}(\Phi) is uniformly contracting, Eu()E^{u}(\Phi) is uniformly expanding and the splitting is dominated with respect to and some (or equivalently any) Riemannian metric gg.

Example 3.9.

Let :×YY\Psi:\px@BbbR\times Y\to Y be an Anosov flow generated by a vector-field VV. Then for any TT, the time TT map

:TYY{}_{T}:Y\to Y

is partially hyperbolic. The stable and unstable bundles of T are those of , while the central bundle is the span of VV.

Partially hyperbolic contactomorphisms have some special compatibility properties with the underlying contact structure. For instance, we have the following lemma (for use later).

Lemma 3.10.

Fix a contact (2n+1)(2n+1)-manifold (Y,ξ)(Y,\xi) and a partially hyperbolic contactomorphism

:(Y,ξ)(Y,ξ)\Phi:(Y,\xi)\to(Y,\xi)

such that ξ=Es()Eu()\xi=E^{s}(\Phi)\oplus E^{u}(\Phi) and Ec()E^{c}(\Phi) is transverse to ξ\xi. Then the stable and unstable foliations Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) have Legendrian leaves.

Proof.

The leaves of Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) are tangent to ξ\xi, and the dimensions of the leaves of Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) add to the rank of ξ\xi. It follows the dimension of the leaves Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) must be nn, so that the leaves are Legendrian. ∎

3.2. Holonomy

The holonomy of the stable and unstable foliations are a key tool in the analysis of partially hyperbolic maps, which we will need in Section 4. We next briefly review this concept.

Let FF be a transversely continuous foliation with smooth leaves on a manifold YY. Let be a leaf and let PP be a point in . Recall that a a transversal to at PP is a sub-manifold SYS\subset Y of dimension codimF\operatorname{codim}F tranverse to the leaves of FF and intersecting at PP.

Definition 3.11 (Holonomy).

[16, Ch 2] Let be a leaf of FF equipped with transversals SS and TT at points PP and QQ in . Fix a continuous path

:[0,1]with(0)=P and (1)=Q\Gamma:[0,1]\to\Lambda\qquad\text{with}\qquad\Gamma(0)=P\text{ and }\Gamma(1)=Q

The holonomy is the unique correspondence assigning to (,S,T)(\Gamma,S,T) the germ of a smooth map

HolF,:SNbhd(P)TNbhd(Q) with HolF,(P)=Q\operatorname{Hol}_{F,\Gamma}:S\cap\operatorname{Nbhd}(P)\to T\cap\operatorname{Nbhd}(Q)\qquad\text{ with }\qquad\operatorname{Hol}_{F,\Gamma}(P)=Q

that satisfies the following properties.

  • The correspondence Hol\operatorname{Hol} respects path composition.

  • If ,S,T\Gamma,S,T are in a foliation chart, then Hol(s)\operatorname{Hol}(s) is the point in TT in the same plaque as ss.

The holonomy only depends on up to homotopy relative to PP and QQ in [16, Prop 2.3.2]. Moreover, given a foliation GG such that TFTF and TGTG are C0C^{0}-close, there is a continuation point

QGNbhd(Q)TQ_{G}\in\operatorname{Nbhd}(Q)\cap T

on the leaf of GG through PP, converging to QQ as GG converges to FF. There is also a unique homotopy class of path PP to QGQ_{G} corresponding to . Thus there is a holonomy map

HolG,:SNbhd(P)TNbhd(Q)\operatorname{Hol}_{G,\Gamma}:S\cap\operatorname{Nbhd}(P)\to T\cap\operatorname{Nbhd}(Q)

The holonomy HolF,\operatorname{Hol}_{F,\Gamma} varies continuously with respect to the foliation FF.

By Theorem 3.5, the leaves of the stable foliation FsF^{s} and unstable foliation FuF^{u} of a partially hyperbolic diffeomorphism are contractible and so the holonomy is independent of the path . In this case, we denote the corresponding holonomies

Hols=HolFs,andHolu=HolFu,\operatorname{Hol}^{s}=\operatorname{Hol}_{F^{s},\Gamma}\qquad\text{and}\qquad\operatorname{Hol}^{u}=\operatorname{Hol}_{F^{u},\Gamma}

The main property of holonomy that we will need is the following regularity result, which follows from (for instance) the uniform Hölder regularity results of Pugh-Shub-Wilkenson [55, Thm A].

Lemma 3.12 (Hölder Holonomy).

Let :YY\Phi:Y\to Y be a partially hyperbolic diffeomorphism. Fix points PP and QQ in an unstable leaf with transversals SS and TT. Then is exists

a C1C^{1}-neighborhood 𝒰\mathcal{U} of       a Hölder constant κ(0,1)\kappa\in(0,1)      and      a neighborhod Nbhd(P)\operatorname{Nbhd}(P)

so that the holonomy Holu:SNbhd(P)TNbhd(Q)\operatorname{Hol}^{u}:S\cap\operatorname{Nbhd}(P)\to T\cap\operatorname{Nbhd}(Q) of any diffeomorphism in 𝒰\mathcal{U} satisfies

(3.1) dist(Holu(x),Holu(y))<dist(x,y)κ\operatorname{dist}\big{(}\operatorname{Hol}^{u}(x),\operatorname{Hol}^{u}(y)\big{)}<\operatorname{dist}(x,y)^{\kappa}

3.3. Cone-Fields

There is an alternative formulation of dominated splittings in terms of cone-fields. Cone-fields will also be used extensively in the definition and construction of blenders.

Definition 3.13 (Cone Fields).

A continuous cone-field KK on a manifold XX is a bundle of the form

K={v:Q(v)0}for a continuous, non-degenerate quadratic form Q:TXK=\big{\{}v\;:\;Q(v)\leq 0\big{\}}\qquad\text{for a continuous, non-degenerate quadratic form }Q:TX\to\px@BbbR

A diffeomorphism :XY\Phi:X\to Y induces a natural pushforward cone-field

Kwith fiberKx=T(K(x)1)x{}_{*}K\qquad\text{with fiber}\qquad{}_{*}K_{x}=T{}_{x}(K_{{}^{-1}(x)})

The interior intK\operatorname{int}K of a cone-field KK is the (fiberwise) cone over the interior of KK.

Example 3.14 (Metric Cones).

The cone-field KϵEK_{\epsilon}E of width ϵ\epsilon around a sub-bundle ETXE\subset TX of the tangent bundle of a Riemannian manifold (X,g)(X,g) is defined by

KϵE:={vTX:|vπE(v)|gϵ|πE(v)|g}where πE is orthogonal projection to EK_{\epsilon}E:=\big{\{}v\in TX\;:\;|v-\pi_{E}(v)|_{g}\leq\epsilon\cdot|\pi_{E}(v)|_{g}\big{\}}\qquad\text{where $\pi_{E}$ is orthogonal projection to $E$}

A cone-field KK has width less than ϵ\epsilon with respect to gg if KKϵEK\subset K_{\epsilon}E for some linear sub-bundle ETXE\subset TX.

Definition 3.15 (Contraction/Dilation).

Fix a subset AXA\subset X and a cone-field KK over XX. A diffeomorphism contracts KK over AA if

(K|A)intK{}_{*}(K|_{A})\subset\operatorname{int}K

Given a Riemannian metric gg, we say that dilates KK over AA with constant of dilation λ>0\lambda>0 if

λ|v|g|v|gfor every vTX|A\lambda\cdot|v|_{g}\leq|{}_{*}v|_{g}\qquad\text{for every }v\in TX|_{A}
Theorem 3.16 (Invariant Cone-Fields).

(c.f. [25, §2.2]) Let :YY\Phi:Y\to Y be a diffeomorphism with a dominated splitting TY=DETY=D\oplus E. Then for any Riemannian metric gg and any ϵ>0\epsilon>0, there is an N>0N>0 such that

 contracts NKϵE for all nN{}^{N}\text{ contracts }K_{\epsilon}E\text{ for all }n\geq N

Moreover, if EE is uniformly expanding, then for any λ>1\lambda>1, we may choose NN so that

 dilates NKϵE with dilation factor λ for all nN{}^{N}\text{ dilates }K_{\epsilon}E\text{ with dilation factor $\lambda$ for all }n\geq N

3.4. Blenders

A blender is a type of robust hyperbolic set within a dynamical system, introduced by Bonatti-Diaz [7]. Blenders play a central role in the construction of robustly mixing maps in the partially hyperbolic setting.

In this section, we precisely define blenders (Definition 3.22) and associated structures. We start with the notion of a blender box, which is a chart where the blender structure will reside.

Definition 3.17 (Blender Box).

A blender box BB of type (k,l)(k,l) in an nn-manifold XX is a C1C^{1}-embedded nn-manifold with corners BB with coordinates

(s,t,u):BDsk×Dt1×Duln(s,t,u):B\xrightarrow{\sim}D^{k}_{s}\times D^{1}_{t}\times D^{l}_{u}\subset{}^{n}

Here DxmD^{m}_{x} denotes a closed metric mm-ball of some (unspecified) radius in xm{}^{m}_{x}, with coordinates xix_{i}. The boundary of BB has distinguished subsets that we denote as follows.

sB=Dk×D1×DlcB=Dk×D1×DluB=Dk×D1×Dl\partial^{s}B=\partial D^{k}\times D^{1}\times D^{l}\qquad\partial^{c}B=D^{k}\times\partial D^{1}\times D^{l}\qquad\partial^{u}B=D^{k}\times D^{1}\times\partial D^{l}

These are the stable, central and unstable boundaries, respectively. The boundary of D1D^{1} is a union of two points, a negative point D1\partial^{-}D^{1} and a positive point +D1\partial^{+}D^{1}. We also fix the notation

lB=Dk×D1×Dland+B=Dk×+D1×Dl\partial^{l}B=D^{k}\times\partial^{-}D^{1}\times D^{l}\qquad\text{and}\qquad\partial^{+}B=D^{k}\times\partial^{+}D^{1}\times D^{l}

These subsets are the left-side and right-side of BB, respectively.

Next, we introduce the notion of compatible cone-fields and vertical/horizontal disks. These structures can be viewed as local (and more flexible) versions of the invariant bundles and foliations associated to a partially hyperbolic map.

Definition 3.18 (Compatible Cones).

A triple of smooth cone-fields Ks,KcuK^{s},K^{cu} and KuK^{u} on XX are compatible with a blender box BB if there are inclusions

TDk×D1×DlKsDk×TD1×TDlKcuDk×D1×TDlKuTD^{k}\times D^{1}\times D^{l}\subset K^{s}\qquad D^{k}\times TD^{1}\times TD^{l}\subset K^{cu}\qquad D^{k}\times D^{1}\times TD^{l}\subset K^{u}

We refer to these cone-fields as stable, central-unstable and unstable cones for BB, respectively.

Definition 3.19 (Vertical/Horizontal Disks).

An embedded kk-disk DBD\subset B is called vertical with respect to an unstable cone-field KuK^{u} if

TDKuandDuBTD\subset K^{u}\qquad\text{and}\qquad\partial D\subset\partial^{u}B

Similarly, an embedded ll-disk DBD\subset B is called horizontal with respect to a stable cone-field KsK^{s} if

TDKsandDsBTD\subset K^{s}\qquad\text{and}\qquad\partial D\subset\partial^{s}B

Given a horizontal disk HBH\subset B that is disjoint from uB\partial^{u}B, there are precisely two homotopy classes of vertical disk disjoint from HH, corresponding to the disks

DRight=0s×+1×DlandDLeft=0s×1×DlD_{\operatorname{Right}}=0_{s}\times+1\times D^{l}\qquad\text{and}\qquad D_{\operatorname{Left}}=0_{s}\times-1\times D^{l}

A vertical disk DD is right of HH if it is homotopic to DRightD_{\operatorname{Right}} and left of HH if it is homotopic to DLeftD_{\operatorname{Left}}.

It will be useful to record some basic properties of vertical disks in the following lemmas. These are entirely elementary and left to the reader (also see [7, §1] or [10]).

Lemma 3.20 (Vertical Manifolds).

Let B\Sigma\subset B be any smooth sub-manifold in a blender box BB with

TKuanduBT\Sigma\subset K^{u}\qquad\text{and}\qquad\partial\Sigma\subset\partial^{u}B

Then is diffeomorphic to an ll-disk, and therefore is a vertical disk.

Lemma 3.21 (Graphs).

Let DD be a vertical disk in a blender box BB with respect to an unstable cone-field KuK^{u} of width ϵ\epsilon. Then DD is the graph

DDsk×Dt1×Dulof an 2ϵ-Lipschitz map f:DulDsk×Dt1D\subset D^{k}_{s}\times D^{1}_{t}\times D^{l}_{u}\qquad\text{of an $2\epsilon$-Lipschitz map }f:D^{l}_{u}\to D^{k}_{s}\times D^{1}_{t}

We are now prepared to introduce our operating definition of a blender (Definition 3.22). We warn the reader that this definition is quite cumbersome, since it is entirely functional and tailored to the specific application. We will discuss more motivation in Remark 3.23 below.

Definition 3.22 (Blender).

Let :XX\Phi:X\to X be a C1C^{1}-diffeomorphism and let BXB\subset X be a blender box (see Definition 3.17). The pair

(B,)(B,\Phi)

is a simple stable blender (or just a stable blender) if it satisfies the following properties.

  1. (a)

    There is a connected component AA of B(B)B\cap\Phi(B) that is disjoint from

    sB(cB)and(uB)\partial^{s}B\qquad\Phi(\partial^{c}B)\quad\text{and}\quad\Phi(\partial^{u}B)
  2. (b)

    There is an mm\in\px@BbbN and a connected component AA^{\prime} of B(B)mB\cap{}^{m}(B) that is disjoint from

    rBsBand(uB)\partial^{r}B\qquad\partial^{s}B\quad\text{and}\quad\Phi(\partial^{u}B)

Moreover, there are constants μ>1>ϵ\mu>1>\epsilon and a compatible set of cone-fields Ks,KcuK^{s},K^{cu} and KuK^{u} on YY of width less than ϵ\epsilon (with respect to the standard metric on BB) such that

  1. (c)

    The cone-fields KsK^{s} and KuK^{u} are contracted and dilated (with constant μ\mu) by -1 and .

    Ks1intKsKuintKu dilates 1Ks dilates Ku{}^{-1}_{*}K^{s}\subset\operatorname{int}K^{s}\qquad{}_{*}K^{u}\subset\operatorname{int}K^{u}\qquad{}^{-1}\text{ dilates }K^{s}\qquad\Phi\text{ dilates }K^{u}
  2. (d)

    The cone-field KcuK^{cu} is contracted and dilated (with constant μ\mu) by as follows.

    KcuintKcuand dilates Kcu over (A)1{}_{*}K^{cu}\subset\operatorname{int}K^{cu}\quad\text{and}\quad\Phi\text{ dilates }K^{cu}\qquad\text{ over }{}^{-1}(A)
    KcumintKcuand dilates mKcu over (A)m{}_{*}^{m}K^{cu}\subset\operatorname{int}K^{cu}\quad\text{and}\quad{}^{m}\text{ dilates }K^{cu}\qquad\text{ over }{}^{-m}(A^{\prime})

Finally, in any box BB equipped with such cone-fields KK^{\bullet}, has a unique hyperbolic fixed point QQ in the component AA of index ll (via [7, Lemma 1.6]), and the local stable manifold

W:=Wlocs(Q,;B)Ws(Q,)BW:=W^{s}_{\operatorname{loc}}(Q,\Phi;B)\subset W^{s}(Q,\Phi)\cap B

is a horizontal kk-disk in the blender box BB. We further assume that there are neighborhoods

U of lBU+ of rBU of WU_{-}\text{ of }\partial^{l}B\qquad U_{+}\text{ of }\partial^{r}B\qquad U\text{ of }W

satisfying the following assumptions.

  1. (e)

    Every vertical disk DD through BB to the right of WW is disjoint from UU_{-}.

  2. (f)

    Every vertical disk DD through BB to the right of WW satisfies one of two possibilities.

    1. (i)

      The intersection (D)A\Phi(D)\cap A contains a vertical disk DD^{\prime} through BB to the right of WW and disjoint from U+U_{+}.

    2. (ii)

      The intersection (D)mA{}^{m}(D)\cap A^{\prime} contains a vertical disk DD^{\prime} through BB to the right of WW and disjoint from UU.

The pair (B,)(B,\Phi) is an simple unstable blender if the pair (B,)1(B,{}^{-1}) is a simple stable blender.

Remark 3.23 (Blender Property).

Definition 3.22 is best understood as being specifically tailored to enforce the following distinctive blender property (c.f. [7, Lem 1.8] or [10, Lem 6.6 and §6.2.2]):

(3.2) Every vertical disk to the right of WW must intersect the (global) stable manifold of QQ

This property has the following easy consequence ([7, Lem 1.8] or [10, Lem 6.8]): if PP is a hyperbolic periodic point with unstable manifold Wu(P,)W^{u}(P,\Phi) intersecting the blender box along a vertical disk to the right of WW, then

Ws(P,)close(Ws(Q,))W^{s}(P,\Phi)\subset\operatorname{close}(W^{s}(Q,\Phi))

If Ws(P,)W^{s}(P,\Phi) has larger dimension than Ws(Q,)W^{s}(Q,\Phi), then this says that Ws(Q,)W^{s}(Q,\Phi) is bigger than expecte. Since the distinctive property is robust, this can lead to robust transitivity properties of the global dynamics of [10, §7.1.3].

More contemporary accounts of blenders (c.f. [6] and [10, §6-7]) take the distinctive property (or a related property) as the definition of a blender. With this in mind, we highly encourage the reader to view Definition 3.22 as a concrete list of criteria leading to the distinctive property (3.2).

Remark 3.24 (Blender Variants).

Several alternative definitions have been introduced since [7] (cf. [10, §6.2.2]). Definition 3.22 matches the original definition of Bonatti-Diaz [7, p. 365] except that we replace the hypothesis [7, p. 365 H3] with the simplified hypothesis (c-d).

4. Blender Construction

In this section, we construct a contact version of the blenders used by Bonatti-Diaz [7] to demonstrate robust transitivity of certain perturbed partially hyperbolic maps.

Setup 4.1 (Blender Setup).

Let (Y,ξ)(Y,\xi) be a contact (2n+1)(2n+1)-manifold with contact form α\alpha and Reeb vector-field RR. Fix a strict contactomorphism

:YY\Phi:Y\to Y

that satisfies the following properties.

  1. (a)

    The contactomorphism is partially hyperbolic with stable, central and unstable splitting

    Es()Ec()Eu()withξ=Es()Eu()andEc()=span(R)E^{s}(\Phi)\oplus E^{c}(\Phi)\oplus E^{u}(\Phi)\qquad\text{with}\qquad\xi=E^{s}(\Phi)\oplus E^{u}(\Phi)\quad\text{and}\quad E^{c}(\Phi)=\operatorname{span}(R)
  2. (b)

    There is a closed Reeb orbit Y\Gamma\subset Y that is a normally hyperbolic set of fixed points of .

  3. (c)

    There is a neighborhood VV of and a smooth, integrable sub-bundle

    Esms()ξwith foliationFsms() over VE_{\operatorname{sm}}^{s}(\Phi)\subset\xi\quad\text{with foliation}\quad F_{\operatorname{sm}}^{s}(\Phi)\qquad\text{ over }V

    that is uniformly contracted by and that agrees with Es()E^{s}(\Phi) on Ws(,)W^{s}(\Gamma,\Phi).

  4. (d)

    There are two points PP and QQ in such that the stable and unstable manifolds

    Ws(Q,)andWu(P,)W^{s}(Q,\Phi)\qquad\text{and}\qquad W^{u}(P,\Phi)

    intersect cleanly along a heteroclinic orbit χ\chi of orbit from PP to QQ.

Our goal over the following two sections (Sections 4 and 5) is to prove the following result. We will apply this result to prove the existence of robustly mixing diffeomorphisms in Section 6.

Theorem 4.2 (Heteroclinic Contact Blender).

For any contactomorphism :YY\Phi:Y\to Y as in Setup 4.1, there is an integer N>0N>0 and a smooth family of contactomorphisms

:[0,1]×YYwith=0\Psi:[0,1]\times Y\to Y\qquad\text{with}\qquad{}_{0}=\Phi

satisfying the following properties for all sufficiently small r>0r>0.

  1. (a)

    The points PP and QQ are hyperbolic fixed points of r of index n+1n+1 and nn, respectively.

  2. (b)

    There is a neighborhood Br(Q)B_{r}(Q) of QQ such that (Br(Q),)rN(B_{r}(Q),{}_{r}^{N}) is a stable blender.

  3. (c)

    The intersections Wu(P,)rBr(Q)W^{u}(P,{}_{r})\cap B_{r}(Q) contains a vertical disk DQD_{Q} to the right of Ws(Q,)rW^{s}(Q,{}_{r}).

In this section (Section 4), we will construct the objects appearing in Theorem 4.2, and in the next section (Section 5), we will prove the blender properties (Definition 3.22). We will use the notation of Setup 4.1 and Theorem 4.2 throughout. We also use the following standing notation.

Notation 4.3 (Reeb Intervals).

Let γ:[0,T]Y\gamma:[0,T]\to Y be an embedded segment of a Reeb trajectory with end-points x=γ(0)x=\gamma(0) and y=γ(Ty=\gamma(T. Then we use the shorthand

[x,y]=γ[0,T]Y[x,y]=\gamma[0,T]\subset Y

4.1. Standard Chart

We will require a particular standard rectangular neighborhood of the segment of from QQ to PP. Our first task is to construct this chart carefully.

Lemma 4.4 (Standard Chart).

Let VYV\subset Y be an open neighborhood of . Then there is a smoothly embedded cube UVU\subset V, constants L,δ,ϵ>0L,\delta,\epsilon>0 and local coordinates

(s,t,u):U[3,3]sn×[δ,L+δ]t×[ϵ,ϵ]un(s,t,u):U\simeq[-3,3]^{n}_{s}\times[-\delta,L+\delta]_{t}\times[-\epsilon,\epsilon]^{n}_{u}

satisfying the following properties.

  1. (a)

    The contact form is given by α|U=dt(s1du1++sndun)\alpha|_{U}=dt-(s_{1}du_{1}+\dots+s_{n}du_{n}).

  2. (b)

    The Reeb vector-field RR of α\alpha is given by R|U=tR|_{U}=\partial_{t}.

  3. (c)

    The points PP and QQ are given by (0s,Lt,0u)(0_{s},L_{t},0_{u}) and (0s,0t,0u)(0_{s},0_{t},0_{u}) respectively.

  4. (d)

    The local stable and unstable manifolds Wlocs(,;U)W_{\operatorname{loc}}^{s}(\Gamma,\Phi;U) and Wlocu(,;U)W_{\operatorname{loc}}^{u}(\Gamma,\Phi;U) of U\Gamma\cap U are given by

    [3,3]sn×[δ,Lδ]t×0uand0s×[δ,Lδ]t×[ϵ,ϵ]u[-3,3]^{n}_{s}\times[-\delta,L-\delta]_{t}\times 0_{u}\qquad\text{and}\qquad 0_{s}\times[-\delta,L-\delta]_{t}\times[-\epsilon,\epsilon]_{u}
  5. (e)

    The stable and unstable bundles Es()E^{s}(\Phi) and Fu()F^{u}(\Phi) satisfy

    TFs()=T on snWlocs(,;U)andTFu()=T on unWlocu(,;U)TF^{s}(\Phi)=T{}^{n}_{s}\text{ on }W_{\operatorname{loc}}^{s}(\Gamma,\Phi;U)\qquad\text{and}\qquad TF^{u}(\Phi)=T{}^{n}_{u}\text{ on }W_{\operatorname{loc}}^{u}(\Gamma,\Phi;U)
  6. (f)

    The bundle T=snEsms()T{}^{n}_{s}=E^{s}_{\operatorname{sm}}(\Phi) is invariant under and uniformly contracted by on UU.

  7. (g)

    The heteroclinic orbit χ\chi contains the point

    a=(1s,0t,0u)where 1s=(1,0,,0)[3,3]sna=(1_{s},0_{t},0_{u})\qquad\text{where }1_{s}=(1,0,\dots,0)\in[-3,3]^{n}_{s}
  8. (h)

    There is an integer k>0k>0 such that

    (a)kU{}^{-k}(a)\notin U
Proof.

We construct this chart in two steps: the construction of a nice transverse hypersurface and the construction of a good chart using that hypersurface.

Step 1: Transverse Hypersurface. In this step, we construct a nice embedded hypersurface. We shrink VV so that the smooth unstable bundle Esms()E^{s}_{\operatorname{sm}}(\Phi) is defined on VV (see Setup 4.1). Consider the foliations Fu()F^{u}(\Phi) and Fsms()F^{s}_{\operatorname{sm}}(\Phi) corresponding to the unstable bundle Fu()F^{u}(\Phi) and the smooth stable bundle Esms()E^{s}_{\operatorname{sm}}(\Phi). These foliations are Legendrian (Lemma 3.10) and Fu()F^{u}(\Phi) is Hölder with smooth leaves (Theorem 3.5).

Next, choose a ball u in the leaf of Fu(;Q)F^{u}(\Phi;Q) containing QQ. Since Fsmu()F^{u}_{\operatorname{sm}}(\Phi) is smooth, we may choose an embedded codimension 11 surface that is contained in the union of the leaves of Fsms()F^{s}_{\operatorname{sm}}(\Phi) intersecting u. Denote this surface by

Ywithu\Sigma\subset Y\qquad\text{with}\qquad{}_{u}\subset\Sigma

This surface is transverse to the Reeb vector-field, and is thus symplectic with symplectic form

dα|with primitive α| vanishing on the leaves of Fs() and Fu()d\alpha|\qquad\text{with primitive $\alpha|$ vanishing on the leaves of $F^{s}(\Phi)$ and $F^{u}(\Phi)$}

By flowing slightly by the Reeb flow, we acquire an embedding as follows (for small δ>0\delta>0).

(4.1) ι:(δ,δ)t×Ywithια=dt+α|\iota:(-\delta,\delta)_{t}\times\Sigma\to Y\qquad\text{with}\qquad\iota^{*}\alpha=dt+\alpha|

Next, apply the neighborhood theorem for Lagrangian foliations [58, Thm 7.1] to the smooth Legendrian foliation Fsms()F^{s}_{\operatorname{sm}}(\Phi) near the transverse smooth Lagrangian u{}_{u}\subset\Sigma. After shrinking , this yields a symplectic embedding

(4.2) ȷ:TFu(;Q)withFsms()=ȷFstd\jmath:\Sigma\to T^{*}F^{u}(\Phi;Q)\qquad\text{with}\qquad F^{s}_{\operatorname{sm}}(\Phi)\cap\Sigma=\jmath^{*}F_{\operatorname{std}}

Here FstdF_{\operatorname{std}} is the standard Lagrangian foliation of TFu(;Q)T^{*}F^{u}(\Phi;Q) by cotangent fibers. Finally, consider the pullback λstd\lambda_{\operatorname{std}} of the standard Liouville form on TuT^{*}{}_{u} by (4.2). Since λstd\lambda_{\operatorname{std}} and α|\alpha| vanish on s and u, we can find a primitive ff such that

λstdα|=df\lambda_{\operatorname{std}}-\alpha|=df

The primitives α|\alpha| and λstd\lambda_{\operatorname{std}} vanish on LuL_{u} and on the foliation Fs()F^{s}(\Phi)\cap\Sigma, so ff is constant these manifolds. It follows that ff is constant on and so

(4.3) λstd=α|\lambda_{\operatorname{std}}=\alpha|

Step 2: Cube Coordinates. In this step, we use to construct a rectangular chart and verify the requirements. Since χ\chi is a homoclinic orbit asymptotic to QQ in the forward direction, we may choose a point aχa\in\chi such that

aFsms(;Q)Fs(;Q)a\in\Sigma\cap F^{s}_{\operatorname{sm}}(\Phi;Q)\subset F^{s}(\Phi;Q)

Here we use the fact that QQ\in\Gamma so that Fsms(;Q)=Fs(;Q)Ws(,)F^{s}_{\operatorname{sm}}(\Phi;Q)=F^{s}(\Phi;Q)\subset W^{s}(\Gamma,\Phi) by Setup 4.1(c). Choose a smooth embedding of a cube of the form

(4.4) [ϵ,ϵ]unuFu(Q)with0uQ[-\epsilon,\epsilon]_{u}^{n}\to{}_{u}\subset F^{u}(Q)\qquad\text{with}\qquad 0_{u}\mapsto Q

This extends naturally to a map of cotangent bundles, giving a Liouville embedding

κ:([3,3]sn×[ϵ,ϵ]un,λstd)(,α|)whereλstd=\slimits@isidui\kappa:([-3,3]_{s}^{n}\times[-\epsilon,\epsilon]_{u}^{n},\lambda_{\operatorname{std}})\to(\Sigma,\alpha|)\qquad\text{where}\qquad\lambda_{\operatorname{std}}=-\sumop\slimits@_{i}s_{i}\cdot du_{i}

By shrinking around s, we may assume that this map is a symplectomorphism. Using the Reeb flow as in (4.1), we may extend κ\kappa to a local contactomorphism

κ:([3,3]sn×[δ,L+δ]t×[ϵ,ϵ]u,dt+λstd)(Y,α)\kappa:\big{(}[-3,3]_{s}^{n}\times[-\delta,L+\delta]_{t}\times[-\epsilon,\epsilon]_{u}\;,\;dt+\lambda_{\operatorname{std}}\big{)}\to(Y\;,\;\alpha)

The restriction of κ\kappa to 0s×[ϵ,L+ϵ]t×0u0_{s}\times[-\epsilon,L+\epsilon]_{t}\times 0_{u} is a parametrization of a sub-arc of containing QQ and PP sending 0 to QQ and LL to PP. Thus by choosing a sufficiently small surface and choosing δ,ϵ\delta,\epsilon small, we may guarantee that κ\kappa is an embedding and the image of κ\kappa lies in VV.

Now take UU to be the image of κ\kappa and take (x,s,u)(x,s,u) as the coordinates induced by κ\kappa. By construction UVU\subset V. We now check that UU and (s,t,u)(s,t,u) satisfy (a-g). For (a), we note that

κα=dt+λstd=dt(s1du1++sndun)\kappa^{*}\alpha=dt+\lambda_{\operatorname{std}}=dt-(s_{1}du_{1}+\dots+s_{n}du_{n})

The requirements (b) follows immediately from (a). Requirements (c) and (f,g) follow trivially from the construction. To see property (d), note that the local stable manifold Wlocs(,;U)W^{s}_{\operatorname{loc}}(\Gamma,\Phi;U) of is precisely the RR-orbit of the component of the local stable leaf Flocs(,Q;U)F^{s}_{\operatorname{loc}}(\Phi,Q;U) containing QQ. By construction, κ\kappa maps [3,3]sn×0t×0u[-3,3]_{s}^{n}\times 0_{t}\times 0_{u} to Flocs(,Q;U)F^{s}_{\operatorname{loc}}(\Phi,Q;U). It follows that Ws(,;U)W^{s}(\Gamma,\Phi;U) is identified with the zero set

{u=0}=[3,3]sn×[ϵ,1+ϵ]t×0u\{u=0\}=[-3,3]_{s}^{n}\times[-\epsilon,1+\epsilon]_{t}\times 0_{u}

An analogous discussion applies to Wlocu(,;U)W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U), verifying (d). Requirement (e) follows from the same discussion, and the fact that Fs()F^{s}(\Phi) and Fu()F^{u}(\Phi) are RR-invariant. Finally, by shrinking the neighborhood VV in the construction, we may guarantee that the orbit χ\chi contains a point that is not in VV or UU. It follows that (a)kU{}^{-k}(a)\notin U for some kk. This verifies (h).∎

We will require an enhancement of Lemma 4.4 that incorporates a set of invariant cone-fields.

Lemma 4.5 (Standard Chart With Cone-Fields).

For any μ>1\mu>1 and ϵ(0,1)\epsilon\in(0,1), there exists

  • An integer N1N\geq 1.

  • A cube UYU\subset Y with coordinates (s,t,u)(s,t,u) as in Lemma 4.4.

  • A Riemannian metric gg on YY that is compatible with the splitting TY=EsEcEuTY=E^{s}\oplus E^{c}\oplus E^{u}

  • Continuous cone-fields KsK^{s} and KuK^{u} on YY.

that satisfy the following properties (after possibly rescaling the contact form α\alpha).

  1. (a)

    The cone-fields KsK^{s} and KuK^{u} are compatible with the blender box UU (see Definition 3.18).

    TsnKsandTunKuT{}^{n}_{s}\subset K^{s}\quad\text{and}\quad T{}^{n}_{u}\subset K^{u}
  2. (b)

    The cone-fields KsK^{s} is contracted by -N, and KcuK^{cu} and KuK^{u} are contracted by N.

    (Ks)NintKsand(Ku)NintKu{}_{*}^{-N}(K^{s})\subset\operatorname{int}K^{s}\quad\text{and}\quad{}_{*}^{N}(K^{u})\subset\operatorname{int}K^{u}
  3. (c)

    The cone-fields KsK^{s} and KuK^{u} are dilated by -N and N with contant of dilation μ\mu, respectively.

    μ|T(v)N|<|v| if vKuandμ|T(v)N|<|v| if vKs\mu\cdot|T{}^{N}(v)|<|v|\text{ if }v\in K^{u}\qquad\text{and}\qquad\mu\cdot|T{}^{-N}(v)|<|v|\text{ if }v\in K^{s}
  4. (d)

    The cone-fields KsK^{s} and KuK^{u} are width less than ϵ\epsilon for both gg and the standard metric on UU.

Proof.

Choose an auxilliary chart UU^{\prime} and coordinates (s,t,u)(s^{\prime},t^{\prime},u^{\prime}) as in Lemma 4.4. Also choose a continuous metric gg on YY that is compatible with the splitting EsEcEuE^{s}\oplus E^{c}\oplus E^{u}. Finally, choose a δ<ϵ\delta<\epsilon sufficiently small such that the cone-fields

Ks=Kδ(Es())Kcu=Kδ(Ec()Eu())Ku=Kδ(Eu())K^{s}=K_{\delta}(E^{s}(\Phi))\qquad K^{cu}=K_{\delta}(E^{c}(\Phi)\oplus E^{u}(\Phi))\qquad K^{u}=K_{\delta}(E^{u}(\Phi))

are width less than ϵ\epsilon with respect to gg. By construction, these cone-fields satisfy (d). By Theorem 3.16 and Setup 4.1, we may choose an N1N\geq 1 such that satisfies (b) and (c).

To achieve (a), note that Es(),Ec()Eu()E^{s}(\Phi),E^{c}(\Phi)\oplus E^{u}(\Phi) and Eu()E^{u}(\Phi) agree with the tangent spaces T,snT(×t1)unT{}^{n}_{s},T({}^{1}_{t}\times{}^{n}_{u}) and TunT{}^{n}_{u} along , respectively. Thus in a small neighborhood VV of , the chosen cone-fields satisfy (a). We may then rescale the coordinates (s,t,u)(s^{\prime},t^{\prime},u^{\prime}) by taking

(s,t,u)=(cs,ct,cu)for c large(s,t,u)=(c\cdot s^{\prime},c\cdot t^{\prime},c\cdot u^{\prime})\qquad\text{for $c$ large}

For cc sufficiently large, the cube with (s,t,u)(s,t,u)-coordinates [2,2]s×[δ,L+δ]t×[ϵ,ϵ]un[2,-2]_{s}\times[-\delta,L+\delta]_{t}\times[-\epsilon,\epsilon]_{u}^{n} will lie within VV. Properties (a-b) are preserved after we scale the contact form by c1c^{-1}. Properties (c-f,h) in Lemma 4.4 are preserved by this coordinate change. Property (g) in in Lemma 4.4 can be preserved by replacing aa with (a)j{}^{j}(a) for large jj.∎

4.2. Family of Contactomorphisms

Our next task is to construct the family of contactomorphisms in Theorem 4.2 and discuss its basic properties. For the rest of the section, we fix

constants LNmUY with coordinates (s,t,u)metric gandcone-fields K\text{constants $L$, $N$, $m$}\qquad U\subset Y\text{ with coordinates }(s,t,u)\qquad\text{metric $g$}\quad\text{and}\quad\text{cone-fields }K^{\bullet}

as in Lemmas 4.4 and 4.5. We can now begin the main construction of the family of maps. We will require two auxilliary contact Hamiltonians for the construction.

Construction 4.6 (Hamiltonian HH).

Let H:YH:Y\to\px@BbbR be a contact Hamiltonian such that

H(s,t,u)=h(t)H(s,t,u)=h(t)

Here h:h:\px@BbbR\to\px@BbbR is a smooth function of the tt-variable satisfying the following constraints.

h(t)=tiftL/3andh(t)=Ltift2L/3h(t)=t\quad\text{if}\quad t\leq L/3\qquad\text{and}\qquad h(t)=L-t\quad\text{if}\quad t\geq 2L/3
h(t)>0and|h(t)|<1/Lif0<t<Lh(t)>0\quad\text{and}\quad|h^{\prime}(t)|<1/L\qquad\text{if}\quad 0<t<L

The contact Hamiltonian vector-field VHV_{H} generating the flow of contactomorphisms H is

(4.5) VH=h(t)th(t)\slimits@isisi in the chart UV_{H}=h(t)\cdot\partial_{t}-h^{\prime}(t)\cdot\sumop\slimits@_{i}s_{i}\cdot\partial_{s_{i}}\qquad\text{ in the chart }U

In particular, we have the following formulas for special ranges of tt.

(4.6) VH=tt\slimits@isisiif tL/3VH=(Lt)t+\slimits@isisiif t2L/3V_{H}=t\partial_{t}-\sumop\slimits@_{i}s_{i}\partial_{s_{i}}\;\;\text{if $t\leq L/3$}\qquad\qquad V_{H}=(L-t)\partial_{t}+\sumop\slimits@_{i}s_{i}\partial_{s_{i}}\;\;\text{if $t\geq 2L/3$}

Using (4.5), it is a simple calculation to show that H takes the following general form on UU.

(4.7) (s,t,u)rH=(fr(t)s,ψr(t),u)whererψ=hψandrf=hf{}^{H}_{r}(s,t,u)=(f_{r}(t)\cdot s,\psi_{r}(t),u)\quad\text{where}\quad\partial_{r}\psi=h\circ\psi\quad\text{and}\quad\partial_{r}f=-h^{\prime}\cdot f

In particular, we have the following formulas for H on the ranges of tt appearing in (4.6).

(4.8) (s,t,u)rH=(ers,ert,u)if t1/3{}^{H}_{r}(s,t,u)=(e^{-r}s,e^{r}t,u)\;\;\text{if $t\leq 1/3$}
(4.9) (s,t,u)rH=(ers,L+er(tL),u)if t2L/3{}^{H}_{r}(s,t,u)=(e^{r}s,L+e^{-r}(t-L),u)\;\;\text{if $t\geq 2L/3$}
Construction 4.7 (Hamiltonian GG).

We define the (time-dependent) contact Hamiltonian

G:[0,1]r×YG:[0,1]_{r}\times Y\to\px@BbbR

as follows. Recall that the heteroclinic point aa in Lemma 4.4 is in the unstable manifold of PP. Moreover, the local unstable manifold of PP in UU is the set 0s×Lt×[δ,δ]un0_{s}\times L_{t}\times[-\delta,\delta]^{n}_{u}. Thus we choose

(4.10) m2such that(a)Nm=(0s,Lt,xu) for some xu[δ,δ]unm\geq 2\qquad\text{such that}\qquad{}^{-Nm}(a)=(0_{s},L_{t},x_{u})\text{ for some }x_{u}\in[-\delta,\delta]^{n}_{u}

Choose a neighborhood WW of (a)Nm{}^{-Nm}(a) that is small enough so that the sets

(rH)j(W)for j=0m({}^{H}_{r}\circ\Phi)^{j}(W)\qquad\text{for }j=0\dots m

are all disjoint for sufficiently small rr. Since (a)Nm{}^{-Nm}(a) and aa are in UU, we may also assume that

WUand(rH)Nm(W)UW\subset U\qquad\text{and}\qquad({}^{H}_{r}\circ\Phi)^{Nm}(W)\subset U

for sufficiently small rr. Moreover, since (a)k{}^{-k}(a) is not in (U)1{}^{-1}(U) for some kk, we have that

(rH)Nmk(W)(U)1=({}^{H}_{r}\circ\Phi)^{Nm-k}(W)\cap{}^{-1}(U)=\emptyset

for sufficiently small rr and neighborhood WW. We may thus fix a pair of open sets WW′′W^{\prime}\subset W^{\prime\prime} that have the following properties for sufficiently small rr.

(rH)Nmk(W)WW′′(U)1=({}^{H}_{r}\circ\Phi)^{Nm-k}(W)\subset W^{\prime}\qquad W^{\prime\prime}\cap{}^{-1}(U)=\emptyset
W′′(rH)j(W)=if0jNmandjNmkW^{\prime\prime}\cap({}^{H}_{r}\circ\Phi)^{j}(W)=\emptyset\quad\text{if}\quad 0\leq j\leq Nm\quad\text{and}\quad j\neq Nm-k

Now we define a smooth family of contactomorphism :rKYY{}^{K}_{r}:Y\to Y (generated by a parameter-dependent contact Hamiltonian KK) implicitly by the following equation.

(4.11) rRNm=NmrH(rH)kGr(rH)Nmk{}^{R}_{r}\circ{}^{Nm}\circ{}^{H}_{Nmr}=({}^{H}_{r}\circ\Phi)^{k}\circ G_{r}\circ({}^{H}_{r}\circ\Phi)^{Nm-k}

Here R is the Reeb flow of YY. Note that =0KId{}^{K}_{0}=\operatorname{Id}. We now define GG so that it satisfies

Gr=Kr on WandGr=0 on YW′′G_{r}=K_{r}\;\text{ on }\;W^{\prime}\qquad\text{and}\qquad G_{r}=0\;\text{ on }\;Y\setminus W^{\prime\prime}
Construction 4.8 (Family Of Maps).

We define the family of contactomorphisms

:[0,1]r×YYas the composition=rrGrH\Psi:[0,1]_{r}\times Y\to Y\qquad\text{as the composition}\qquad{}_{r}={}^{G}_{r}\circ{}^{H}_{r}\circ\Phi

Note that r satisfies the following elementary properties for WW and mm as in Construction 4.7.

(4.12) =rrHon(U)1and(x)rNm=rRNmonNmrHW{}_{r}={}^{H}_{r}\circ\Phi\;\;\text{on}\;\;{}^{-1}(U)\qquad\text{and}\qquad{}_{r}^{Nm}(x)={}^{R}_{r}\circ{}^{Nm}\circ{}^{H}_{Nmr}\;\;\text{on}\;\;W

4.3. Properties Of Family

In this section, we prove several properties of the family in Construction 4.8. We will use these properties extensively in the proof of the blender properties.

We start by recording the effect of the maps on the coordinates in the standard chart UU.

Lemma 4.9 (Coordinate Projections).

Let πW\pi_{W} and π\pi be the coordinate projections

πW:UWlocu(,;U)=U{s=0}given byπW(s,t,u)=(0s,t,u)\pi_{W}:U\to W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U)=U\cap\{s=0\}\qquad\text{given by}\qquad\pi_{W}(s,t,u)=(0_{s},t,u)
π:UU=U{s,u=0}given byπ(s,t,u)=(0s,t,0u)\pi:U\to\Gamma\cap U=U\cap\{s,u=0\}\qquad\text{given by}\qquad\pi(s,t,u)=(0_{s},t,0_{u})

Then πW\pi_{W} and π\pi commute with r over U(U)1U\cap{}^{-1}(U). In particular, if ψ\psi is the flow of hth\cdot\partial_{t} on t then

t((x)r)=t(π(x)r)=t(rπ(x))=ψr(t(x))t({}_{r}(x))=t(\pi\circ{}_{r}(x))=t({}_{r}\circ\pi(x))=\psi_{r}(t(x))
Proof.

On the set (U)1{}^{-1}(U), we have =rrH{}_{r}={}^{H}_{r}\circ\Phi by the first identity in (4.12). By Lemma 4.4(d-f) and by examination of (4.7), both and rH{}^{H}_{r} preserves the foliation

(4.13) Esms()=T on snUandT on unWlocu(,;U)E^{s}_{\operatorname{sm}}(\Phi)=T{}^{n}_{s}\text{ on }U\qquad\text{and}\qquad T{}^{n}_{u}\text{ on }W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U)

Moreover, both and rH{}^{H}_{r} (and therefore r) preserve the sets

Wlocu(,;U)=U{s=0}andU=U{s,t=0}W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U)=U\cap\{s=0\}\qquad\text{and}\qquad\Gamma\cap U=U\cap\{s,t=0\}

This r sends the leaf of TsnT{}^{n}_{s} through xU{s=0}x\in U\cap\{s=0\} to the leaf through (x)r{}_{r}(x). In other words, πW\pi_{W} commutes with r. Similarly, r preserves the leaves of TunT{}^{n}_{u} on U{s=0}U\cap\{s=0\} and so

π(x)r=rπ(x)if xU{s=0}\pi\circ{}_{r}(x)={}_{r}\circ\pi(x)\qquad\text{if }x\in U\cap\{s=0\}

Since ππW=π\pi\circ\pi_{W}=\pi, this implies that π\pi commutes with r in general. The final claim follows since t=rψrtt\circ{}_{r}=\psi_{r}\circ t on U\Gamma\cap U, since fixes pointwise and t=rHψrtt\circ{}^{H}_{r}=\psi_{r}\circ t on U\Gamma\cap U. ∎

Lemma 4.10 (Coordinate Contraction).

Let U=U(U)NU^{\prime}=U\cap\dots\cap{}^{-N}(U). Then for small rr, we have

μ|s((x)rN)|<|s(x)|andμ|u(x)|<|u((x)rN)|\mu\cdot|s({}_{r}^{N}(x))|<|s(x)|\qquad\text{and}\qquad\mu\cdot|u(x)|<|u({}_{r}^{N}(x))|
Proof.

Note that TsnT{}^{n}_{s} is uniformly contracted by N by a factor of μ\mu on UU and TunT{}^{n}_{u} is uniformly expanded by N by a factor of μ\mu on Wlocu(,;U)W^{u}_{\operatorname{loc}}(\Gamma,\Phi;U) due to Lemma 4.5.

These properties are robust in the C1C^{1}-topology, so the map r also satisfies these properties for small rr. Since xx and π(x)\pi(x) lie in the same fiber of TsnT{}^{n}_{s}, we can apply Lemma 4.10 to see that

|s((x)rN)|=|s((x)rN)s((πW(x))rN|μ1|s(x)s(πW(x))|=μ1|s(x)||s({}^{N}_{r}(x))|=|s({}^{N}_{r}(x))-s({}^{N}_{r}(\pi_{W}(x))|\leq\mu^{-1}\cdot|s(x)-s(\pi_{W}(x))|=\mu^{-1}\cdot|s(x)|

Also, the maps u,πu,\pi and πW\pi_{W} satisfy uπW=uu\circ\pi_{W}=u and uπ=0u\circ\pi=0. Thus by Lemma 4.10, we have

|u((x)rN)|=|uπW((x)rN)uπ((x)rN)|μ|u(πW(x))u(π(x))||u(x)||u({}^{N}_{r}(x))|=|u\circ\pi_{W}({}^{N}_{r}(x))-u\circ\pi({}^{N}_{r}(x))|\geq\mu\cdot|u(\pi_{W}(x))-u(\pi(x))|\geq|u(x)|\qed

Next, note that the maps in the family is partially hyperbolic near . This follows from Theorems 3.4-3.5 and the partial hyperbolicity of .

Lemma 4.11 (Partial Hyperbolicity).

The maps r (for sufficiently small rr) has a dominated splitting

TY=Es()rEc()rEu()rwith stable/unstable foliationsFs()r and Fu()rTY=E^{s}({}_{r})\oplus E^{c}({}_{r})\oplus E^{u}({}_{r})\quad\text{with stable/unstable foliations}\quad F^{s}({}_{r})\text{ and }F^{u}({}_{r})

The points PP and QQ become non-degenerate and hyperbolic fixed points of r.

Lemma 4.12 (Fixed Points).

The points QQ and PP are hyperbolic fixed points of r of index

ind(;rQ)=nandind(;rP)=n+1\operatorname{ind}({}_{r};Q)=n\qquad\text{and}\qquad\operatorname{ind}({}_{r};P)=n+1

The local stable and unstable manifolds of PP and QQ in UU with respect to r are given by

Wlocs(P,;rU)=[3,3]sn×(0,L+ϵ]t×0uWlocu(P,;rU)=0s×Lt×[δ,δ]uW^{s}_{\operatorname{loc}}(P,{}_{r};U)=[-3,3]^{n}_{s}\times(0,L+\epsilon]_{t}\times 0_{u}\qquad\quad W^{u}_{\operatorname{loc}}(P,{}_{r};U)=0_{s}\times L_{t}\times[-\delta,\delta]_{u}
Wlocs(Q,;rU)=[3,3]sn×0t×0uWlocu(Q,;rU)=0s×[ϵ,L)t×[δ,δ]u\hskip 28.0ptW^{s}_{\operatorname{loc}}(Q,{}_{r};U)=[-3,3]^{n}_{s}\times 0_{t}\times 0_{u}\hskip 71.0ptW^{u}_{\operatorname{loc}}(Q,{}_{r};U)=0_{s}\times[-\epsilon,L)_{t}\times[-\delta,\delta]_{u}
Proof.

The points PP and QQ are fixed by due to Setup 4.1(d) and fixed by rH{}^{H}_{r} due to (4.8-4.9). Thus (4.12) implies that PP and QQ are fixed by r. To compute the index of PP, note that TT\Phi and TrHT{}^{H}_{r} at PP decomposes via the splitting TY=Es()Ec()Eu()TY=E^{s}(\Phi)\oplus E^{c}(\Phi)\oplus E^{u}(\Phi) as follows.

TP=TPsIdcTPuandTP=rHerIdserIdcIduT_{P}\Phi=T_{P}{}^{s}\oplus\operatorname{Id}_{c}\oplus T_{P}{}^{u}\qquad\text{and}\qquad T_{P}{}^{H}_{r}=e^{r}\operatorname{Id}_{s}\oplus\;e^{-r}\operatorname{Id}_{c}\oplus\operatorname{Id}_{u}

The linear maps TPsT_{P}{}^{s} and TPuT_{P}{}^{u} are the restrictions to Es()E^{s}(\Phi) and Eu()E^{u}(\Phi) respectively. Since PP is in UU, the first formula in (4.12) implies that the differential of r at PP is given by

TP=rerTPrserIdcTPruT_{P}{}_{r}=e^{r}T_{P}{}^{s}_{r}\;\oplus\;e^{-r}\operatorname{Id}_{c}\;\oplus\;T_{P}{}^{u}_{r}

Since is a normally hyperbolic fixed set of by Setup 4.1, the maps erTPrse^{r}T_{P}{}^{s}_{r} (for small rr) and TPruT_{P}{}^{u}_{r} have real eigenvalues of norm bounded above by 11 and below by 11, respectively. It follows that TPrT_{P}{}_{r} will be hyperbolic of index n+1n+1. The same analysis applies to the index of QQ.

To find the stable and unstable manifolds, fix xUx\in U. By Lemma 4.9, we know that t((x)rk)=ψrk(t(x))t({}_{r}^{k}(x))=\psi^{k}_{r}(t(x)) where ψr\psi_{r} of the vector-field hth\cdot\partial_{t} and HH is as in Construction 4.6. Moreover h>0h>0 on (0,L)(0,L) and h<0h<0 on [ϵ,0)[-\epsilon,0) and (L,L+ϵ](L,L+\epsilon]. This implies that t((x)rk)t({}_{r}^{k}(x)) converges to

L if t(x)(0,L+ϵ]0 if t(x)=0L\text{ if }t(x)\in(0,L+\epsilon]\qquad 0\text{ if }t(x)=0

and that (x)rk{}_{r}^{k}(x) leaves UU if t(x)[ϵ,0)t(x)\in[-\epsilon,0). Now Lemma 4.10 implies that |u((x)rk)||u({}^{k}_{r}(x))| diverges if u(x)0u(x)\neq 0, and so (x)rk{}^{k}_{r}(x) leaves UU. On the other hand, if u(x)=0u(x)=0 then |s((x)rk)|0|s({}^{k}_{r}(x))|\to 0 as kk\to\infty by Lemma 4.10. This shows that

Wlocs(P,;rU)=[3,3]sn×(0,L+ϵ]t×0uWlocs(Q,;rU)=[3,3]sn×0t×0uW^{s}_{\operatorname{loc}}(P,{}_{r};U)=[-3,3]^{n}_{s}\times(0,L+\epsilon]_{t}\times 0_{u}\qquad W^{s}_{\operatorname{loc}}(Q,{}_{r};U)=[-3,3]^{n}_{s}\times 0_{t}\times 0_{u}

An identical analysis works for the unstable manifolds. ∎

Next, we have the following description of the local stable and unstable foliations, and the action of r on the leaves in the chart UU.

Lemma 4.13 (Stable/Unstable Foliations).

The local stable and unstable foliations of r satisfy

Fs()r=TonsnWlocs(P,;rU)andFu()r=TonunWlocu(Q,;rU)F^{s}({}_{r})=T{}^{n}_{s}\;\;\text{on}\;\;W_{\operatorname{loc}}^{s}(P,{}_{r};U)\qquad\text{and}\qquad F^{u}({}_{r})=T{}^{n}_{u}\;\;\text{on}\;W_{\operatorname{loc}}^{u}(Q,{}_{r};U)

Moreover, if s and u are leaves of Fs()rF^{s}({}_{r}) and Fu()rF^{u}({}_{r}) intersecting U\Gamma\cap U, respectively, then

()srU=()srHUand()urU=()urHU{}_{r}({}^{s})\cap U={}^{H}_{r}({}^{s})\cap U\qquad\text{and}\qquad{}_{r}({}^{u})\cap U={}^{H}_{r}({}^{u})\cap U
Proof.

For the first claim, note that TsnT{}^{n}_{s} and TunT{}^{n}_{u} are tangent to Wlocs(P,;rU)W^{s}_{\operatorname{loc}}(P,{}_{r};U) and Wlocu(Q,;rU)W^{u}_{\operatorname{loc}}(Q,{}_{r};U) by Lemma 4.12. Moreover, TsnT{}^{n}_{s} and TunT{}^{n}_{u} are uniformly contracted and expanded, respectively, on those sets via (4.12). The lemma then follows from the uniqueness of the strong stable and unstable foliations on the stable and unstable manifolds of a hyperbolic invariant set [39, §4]. The second claim follows from the first claim, the formula (4.12) and the fact that preserves the stable and unstable foliations in UU. ∎

Finally, we have the following computation of a certain important family of heteroclinics.

Lemma 4.14 (Heteroclinics).

Consider the Reeb segment [br,br][b_{-r},b_{r}] containing the point aa, where

bτ=(1s,τ,0u)and we recall thata=(1s,0,0u)b_{\tau}=(1_{s},\tau,0_{u})\qquad\text{and we recall that}\qquad a=(1_{s},0,0_{u})

Then for r>0r>0 sufficiently small, we have

  1. (a)

    The leaf of Fu()rF^{u}({}_{r}) through (0s,Lδ,0u)(0_{s},L-\delta,0_{u}) meets [br,br][b_{-r},b_{r}] at the point (1s,rδ,0u)(1_{s},r-\delta,0_{u}) for δ[0,2r]\delta\in[0,2r].

  2. (b)

    The interval (a,br)(a,b_{r}) is a connected component of the intersection of Ws(P,)rW^{s}(P,{}_{r}) and Wu(Q,)rW^{u}(Q,{}_{r}).

  3. (c)

    The points aa and brb_{r} are transverse homoclinic points of QQ and PP, respectively.

Proof.

Let mm and WW be the integer and fixed (rr-independent) neighborhood in Construction 4.8. Note that WW is a fixed neighborhood of the point

(a)Nm=(0,Lt,1u){}^{-Nm}(a)=(0,L_{t},1_{u})

We may thus choose rr small enough so that we have the following inclusion for all δ[0,2r]\delta\in[0,2r].

(0s,Lδ,1u)NmrHW{}_{-Nmr}^{H}(0_{s},L-\delta,1_{u})\in W

We first compute the image of (0s,Lδ,1u)(0_{s},L-\delta,1_{u}) under rNmNmrH{}_{r}^{Nm}\circ{}^{H}_{-Nmr}. By (4.12), we see that

rNm(0s,Lδ,1u)NmrH=rRNmNmrH(0s,Lδ,1u)NmrH=rR(0s,Lδ,1u)Nm{}_{r}^{Nm}\circ{}_{-Nmr}^{H}(0_{s},L-\delta,1_{u})={}^{R}_{r}\circ{}^{Nm}\circ{}^{H}_{Nmr}\circ{}^{H}_{-Nmr}(0_{s},L-\delta,1_{u})={}^{R}_{r}\circ{}^{Nm}(0_{s},L-\delta,1_{u})

Since the Reeb vector-field commutes with , the latter expression becomes

rR(0s,Lδ,1u)Nm=rRNm(0s,L,1u)δR=rδR(0s,L,1u)m{}^{R}_{r}\circ{}^{Nm}(0_{s},L-\delta,1_{u})={}^{R}_{r}\circ{}^{Nm}\circ{}^{R}_{-\delta}(0_{s},L,1_{u})={}^{R}_{r-\delta}\circ{}^{m}(0_{s},L,1_{u})

Since (1s,0,0u)=(0s,L,1u)Nm(1_{s},0,0_{u})={}^{Nm}(0_{s},L,1_{u}) by (4.10) in Construction 4.8, we then acquire the equality

rδR(0s,L,1u)Nm=(1s,0,0u)rδR=(1s,rδ,0u){}^{R}_{r-\delta}\circ{}^{Nm}(0_{s},L,1_{u})={}^{R}_{r-\delta}(1_{s},0,0_{u})=(1_{s},r-\delta,0_{u})

We thus acquire the following formula that we will shortly use to prove (a-c).

(4.14) rNm(0s,Lδ,1u)NmrH=(1s,rδ,0u){}^{Nm}_{r}\circ{}^{H}_{-Nmr}(0_{s},L-\delta,1_{u})=(1_{s},r-\delta,0_{u})

Now we prove the claims above. For (a), note that by Lemma 4.13, rNm{}^{Nm}_{r} maps the leaf of Fu()rF^{u}({}_{r}) containing (0s,Lδ,1u)NmrH{}^{H}_{-Nmr}(0_{s},L-\delta,1_{u}) to the leaf ()rNm=()NmrH{}^{Nm}_{r}(\Lambda)={}^{H}_{Nmr}(\Lambda) containing the points

rNm(0s,Lδ,1u)NmrH=(1s,rδ,0u){}^{Nm}_{r}\circ{}^{H}_{-Nmr}(0_{s},L-\delta,1_{u})=(1_{s},r-\delta,0_{u})
NmrH(0s,Lδ,1u)NmrH=(0s,Lδ,1u){}_{Nmr}^{H}\circ{}^{H}_{-Nmr}(0_{s},L-\delta,1_{u})=(0_{s},L-\delta,1_{u})

By Lemma 4.13, the leaf that contains (0s,Lδ,1u)(0_{s},L-\delta,1_{u}) also contains (0s,Lδ,0u)(0_{s},L-\delta,0_{u}), proving (a). For (b), we note that by (a), the point (0s,rδ,0u)(0_{s},r-\delta,0_{u}) is contained in the stable manifold Ws(P,)rW^{s}(P,{}_{r}) by Lemma 4.12. On the other hand, it is also contained in the leaf of in Fu()rF^{u}({}_{r}) passing through (1s,rδ,0u)(1_{s},r-\delta,0_{u}) by (a), and this leaf is contained in the unstable manifold Wu(Q,)rW^{u}(Q,{}_{r}) by Lemma 4.12. This proves (b). For (c), we argue similarly to (b). By (a), the point a=(1s,0t,0u)a=(1_{s},0_{t},0_{u}) is contained in the leaf of Fu()rF^{u}({}_{r}) containing (0s,Lr,0u)(0_{s},L-r,0_{u}), which is contained in the unstable manifold Wu(Q,)rW^{u}(Q,{}_{r}) by Lemmas 4.12. On the other hand, aa is contained in Ws(Q,)rW^{s}(Q,{}_{r}) (also by Lemma 4.12). Thus it is a homoclinic point for QQ. A similar discussion holds for brb_{r} and PP.∎

4.4. Family Of Boxes

In this section, we describe the families of smoothly embedded boxes

Br(Q)B_{r}(Q)

appearing in the blenders in Theorem 4.2. We start by introducing notation for several important points and quantities appearing in the description of the boxes.

Notation 4.15.

For r[0,1]r\in[0,1], let PrP_{r} and QrQ_{r} denote the points in UU given by

Pr=(0s,Lr,0u)andQr=(0s,r(1e8Nr),0u)P_{r}=(0_{s},L-r,0_{u})\qquad\text{and}\qquad Q_{r}=(0_{s},r\cdot(1-e^{-8Nr}),0_{u})

Then we define the constant mrm_{r} to be the unique integer satisfying

(4.15) (Qr)rNmr[(Pr)r5N,(Pr)r6N]{}_{r}^{Nm_{r}}(Q_{r})\in[{}^{5N}_{r}(P_{r}),{}^{6N}_{r}(P_{r})]
Lemma 4.16.

The integers mrm_{r} diverge to \infty as r0r\to 0. Precisely, there are constants C,r0>0C,r_{0}>0 such that

mr>Cr1log(r)for all r(0,r0)m_{r}>-Cr^{-1}\cdot\log(r)\qquad\text{for all }r\in(0,r_{0})
Proof.

Since r restricts to the flow of ψr\psi_{r} on (see Construction 4.6 or Lemma 4.9), we may equivalently characterize mrm_{r} by

ψrNmr(r(1e8Nr))[Lre5Nr,Lre6Nr]\psi_{rNm_{r}}(r(1-e^{-8Nr}))\in[L-re^{-5Nr},L-re^{-6Nr}]

Here ψr:tt\psi_{r}:{}_{t}\to{}_{t} is the flow of the vector-field hth\cdot\partial_{t} and hh is as in Construction 4.6. Briefly switch notation by letting ψr(t)=ψ(r,t)\psi_{r}(t)=\psi(r,t), and let the quantities Tst(r),TmidT_{\operatorname{st}}(r),T_{\operatorname{mid}} and Tend(r)T_{\operatorname{end}}(r) be the unique values such that

ψ(Tst(r),r(1e8Nr))=L/3ψ(Tmid,L/3)=2L/3ψ(Tend(r),2L/3)=Lre5Nr\psi(T_{\operatorname{st}}(r),r(1-e^{-8Nr}))=L/3\qquad\psi(T_{\operatorname{mid}},L/3)=2L/3\qquad\psi(T_{\operatorname{end}}(r),2L/3)=L-re^{-5Nr}

Note that TmidT_{\operatorname{mid}} is independent of rr. Moreover, using the formulas (4.8) and (4.9) on the intervals [0,L/3]t[0,L/3]_{t} and [2L/3,L]t[2L/3,L]_{t} (see Construction 4.6), we can compute that

Tst(r)=log(L/3)log(r)log(1e8Nr)T_{\operatorname{st}}(r)=\log(L/3)-\log(r)-\log(1-e^{-8Nr})
Tend(r)=log(3/2)log(r)+5NrT_{\operatorname{end}}(r)=-\log(3/2)-\log(r)+5Nr

Now we simply note that in the r0r\to 0 limit, we have

rNmr(Tst(r)+Tmid+Tend(r))3log(r)rNm_{r}\sim(T_{\operatorname{st}}(r)+T_{\operatorname{mid}}+T_{\operatorname{end}}(r))\geq-3\log(r)

Here \sim denotes that the limit of the ratio is 11. This proves the result.∎

We can now introduce the definition of the blender box Br(Q)B_{r}(Q).

Construction 4.17 (Blender Box For ).

The blender box Br(Q)UB_{r}(Q)\subset U is defined as the set

Br(Q)=Dsn(2)×Dt1(t(Qr))×Dun(lμmr)B_{r}(Q)=D^{n}_{s}(2)\times D^{1}_{t}(t(Q_{r}))\times D^{n}_{u}(l\cdot\mu^{-m_{r}})

The constants in the formula are defined as follows.

  • The quantity t(Qr)t(Q_{r}) is the tt-coordinate r(1e8Nr)r\cdot(1-e^{-8Nr}) of the point QrQ_{r}.

  • The constant ll is a positive number such that l/3l/3 is larger than the minimum radius of a ball in the unstable leaf Fu(;rP)F^{u}({}_{r};P) containing aa.

  • The constant μ\mu is an rr-independent constant of dilation for r via Lemma 4.11.

  • The constant mrm_{r} is the integer defined by the formula (4.15).

The stable, central, unstable, left and right boundaries are defined as in Definition 3.17.

We will also need an auxilliary Hölder continuous box constructed using unstable leaves. Let

(4.16) Sr(Q)=Dsn(2+μmr/2)×Dt1(t((Qr)rN))×0uS_{r}(Q)=D^{n}_{s}(2+\mu^{-m_{r}/2})\times D^{1}_{t}(t({}^{N}_{r}(Q_{r})))\times 0_{u}

The Hölder box Btilder(Q)\tilde{B}_{r}(Q) is as the union of disks in the leaves of the unstable foliation Fu()rF^{u}({}_{r}) that have boundary on the set {|u|=lμmr}\{|u|=l\cdot\mu^{-m_{r}}\} and that intersect Sr(Q)S_{r}(Q) . That is

(4.17) Btilder(Q)={x:Fu(,rx)Sr(Q)and|u(x)|lμmr}\tilde{B}_{r}(Q)=\big{\{}x\;:\;F^{u}({}_{r},x)\cap S_{r}(Q)\neq\emptyset\quad\text{and}\quad|u(x)|\leq l\cdot\mu^{-m_{r}}\big{\}}

There is a map π:Btilder(Q)Sr(Q)\pi:\tilde{B}_{r}(Q)\to S_{r}(Q) mapping a point xx to the intersection point π(x)Fu(,rx)Sr(Q)\pi(x)\in F^{u}({}_{r},x)\cap S_{r}(Q). The following lemma allows us to replace Btilder(Q)\tilde{B}_{r}(Q) with Br(Q)B_{r}(Q) in some arguments.

Lemma 4.18.

The Hölder box B𝑡𝑖𝑙𝑑𝑒r(Q)\tilde{B}_{r}(Q) contains the blender box Br(Q)B_{r}(Q) for sufficiently small rr.

Proof.

Let xBr(Q)x\in B_{r}(Q). By Construction 4.17 we know that |u(x)|lμmr|u(x)|\leq l\cdot\mu^{-m_{r}}. Let DFu(,rx)D\subset F^{u}({}_{r},x) be the disk given by Fu(,rx){|u(x)|lμmr}F^{u}({}_{r},x)\cap\{|u(x)|\leq l\cdot\mu^{-m_{r}}\}. This disk is tangent to the vertical cone-field KuK^{u} in UU (see Lemma 4.5), and thus is the graph of a Lipschitz map

f:Dun(lμmr)×snt1f:D_{u}^{n}(l\cdot\mu^{-m_{r}})\to{}^{n}_{s}\times{}^{1}_{t}

with a uniform Lipschitz constant CC independent of rr (see Lemma 3.21) and so the image of ff in ×st{}_{s}\times{}_{t} has diameter bounded by ClμmrC\cdot l\cdot\mu^{-m_{r}}. Thus if y=f(0u)y=f(0_{u}) then

|s(y)s(x)|<5Clμmrand|t(y)t(x)|<5Clμmr|s(y)-s(x)|<5C\cdot l\cdot\mu^{-m_{r}}\qquad\text{and}\qquad|t(y)-t(x)|<5C\cdot l\cdot\mu^{-m_{r}}

In particular, this implies that for sufficiently small rr, we have

|s(y)||s(x)|+5Clμmr2+μmr/2|s(y)|\leq|s(x)|+5C\cdot l\cdot\mu^{-m_{r}}\leq 2+\mu^{-m_{r}/2}
|t(y)||t(x)|+5Clμmrt(Qr)+5Clμmrt((Qr)r)|t(y)|\leq|t(x)|+5C\cdot l\cdot\mu^{-m_{r}}\leq t(Q_{r})+5C\cdot l\cdot\mu^{-m_{r}}\leq t({}_{r}(Q_{r}))

The last inequality follows from the fact that, for rr small, we have

t((Qr)r)t(Qr)=r(e9Nre8Nr)12r2t({}_{r}(Q_{r}))-t(Q_{r})=r(e^{9Nr}-e^{8Nr})\geq\frac{1}{2}r^{2}

Thus DD intersects Sr(Q)S_{r}(Q) at y×0uy\times 0_{u} and Fu(,rx)Sr(Q)F^{u}({}_{r},x)\cap S_{r}(Q) is non-empty. In particular, xx is in Sr(Q)S_{r}(Q) and Br(Q)Btilder(Q)B_{r}(Q)\subset\tilde{B}_{r}(Q). ∎

5. Proof Of Blender Axioms

In the previous section, we constructed the family of maps and an accompanying family of blender boxes. Our objective in this section is to prove that the pair

(,rNBr(Q))for sufficiently small r({}_{r}^{N},B_{r}(Q))\qquad\text{for sufficiently small }r

satisfies the axioms of a stable blender given in Definition 3.22.

Construction 5.1 (Useful Holonomy).

The following holonomy map will be useful in the proofs below. Recall that aa and PP lie on a leaf of the unstable foliation Fu()F^{u}(\Phi) of , with transversals

(5.1) S=T=[2,2]sn×[ϵ,L+ϵ]t×0uS=T=[-2,2]^{n}_{s}\times[-\epsilon,L+\epsilon]_{t}\times 0_{u}

By Lemma 3.12, we can choose a neighborhood Nbhd(P)\operatorname{Nbhd}(P) and a Hölder constant κ\kappa so that the corresponding holonomy maps Holr\operatorname{Hol}_{{}_{r}} from Nbhd(P)S\operatorname{Nbhd}(P)\cap S to TT are Hölder with Hölder constant κ\kappa for sufficiently small rr. For small rr, this restricts to a holonomy map

(5.2) Holr:Dsn(2μmr)×[L2r,L]t×0u[2,2]sn×[δ,L+δ]t×0u\operatorname{Hol}_{{}_{r}}:D^{n}_{s}(2\cdot\mu^{-m_{r}})\times[L-2r,L]_{t}\times 0_{u}\to[-2,2]^{n}_{s}\times[-\delta,L+\delta]_{t}\times 0_{u}

Lemma 4.14(a) can be restated as the following formula.

(5.3) Holr(0s,Lδ,0u)=(1s,rδ,0u)for any δ[0,2r]\operatorname{Hol}_{{}_{r}}(0_{s},L-\delta,0_{u})=(1_{s},r-\delta,0_{u})\qquad\text{for any }\delta\in[0,2r]

5.1. Axiom A

We start by defining the subset AA and proving the axiom in Definition 3.22(a).

Definition 5.2 (Blender Set AA).

We let Ar(Q)Br(Q)A_{r}(Q)\subset B_{r}(Q) be the connected component of the intersection Br(Q)(Br(Q))NB_{r}(Q)\cap{}^{N}(B_{r}(Q)) that contains the point Q=(0s,0t,0u)Q=(0_{s},0_{t},0_{u}).

Lemma 5.3 (Blender Axiom A).

The intersection Br(Q)(Br(Q))rNB_{r}(Q)\cap{}_{r}^{N}(B_{r}(Q)) is disjoint from

sBr(Q)(cBr(Q))rNand(uBr(Q))rNfor sufficiently small r\partial^{s}B_{r}(Q)\qquad{}^{N}_{r}(\partial^{c}B_{r}(Q))\quad\text{and}\quad{}_{r}^{N}(\partial^{u}B_{r}(Q))\qquad\text{for sufficiently small }r

In particular, the connected component Ar(Q)A_{r}(Q) of Br(Q)(Br(Q))rNB_{r}(Q)\cap{}_{r}^{N}(B_{r}(Q)) satisfies these properties.

Proof.

We start by noting that for sufficiently small rr, we have

(5.4) Br(Q)U(U)r1(U)rNB_{r}(Q)\subset U\cap{}_{r}^{-1}(U)\cap\dots\cap{}_{r}^{-N}(U)

Now fix points xsBr(Q)x\in\partial^{s}B_{r}(Q) and y(uBr(Q))rNy\in{}^{N}_{r}(\partial^{u}B_{r}(Q)). By Lemma 4.10 and Construction 4.17, we know that the coordinates of xx and yy satisfy

|s(x)|=2and|u(y)|>μlμmr|s(x)|=2\qquad\text{and}\qquad|u(y)|>\mu\cdot l\cdot\mu^{-m_{r}}

Also by Lemma 4.10 and Construction 4.17) we know that any zBr(Q)(Br(Q))rNz\in B_{r}(Q)\cap{}^{N}_{r}(B_{r}(Q)) satisfies

|s(z)|<2μ1and|u(z)|=lμmr|s(z)|<2\cdot\mu^{-1}\qquad\text{and}\qquad|u(z)|=l\cdot\mu^{-m_{r}}

It follows that Br(Q)(Br(Q))rNB_{r}(Q)\cap{}^{N}_{r}(B_{r}(Q)) is disjoint from sBr(Q)\partial^{s}B_{r}(Q) and (uBr(Q))rN{}^{N}_{r}(\partial^{u}B_{r}(Q)). For the central boundary, note that Br(Q)B_{r}(Q) consists of points where tL/3t\leq L/3 for small rr. Thus by Lemma 4.10

|t((w)rN)|=|ψNr(t(w))|=eNrr(1e8Nr) for any wcBr(Q)|t({}^{N}_{r}(w))|=|\psi_{Nr}(t(w))|=e^{Nr}\cdot r\cdot(1-e^{-8Nr})\qquad\text{ for any }w\in\partial^{c}B_{r}(Q)

On the other hand, any zBr(Q)z\in B_{r}(Q) has |t(z)|r(1e8Nr)|t(z)|\leq r\cdot(1-e^{-8Nr}) by Construction 4.17. Thus Br(Q)B_{r}(Q) and (cBr(Q))rN{}^{N}_{r}(\partial^{c}B_{r}(Q)) are disjoint, and the proof is finished. ∎

5.2. Axiom B

Next, we introduce the subset AA^{\prime} appearing in Definition 3.22 and prove the axiom in Definition 3.22(b). We require the following lemma for the definition of AA^{\prime}.

Lemma 5.4.

The heteroclinic point aχa\in\chi is contained in the intersection

Br(Q)(Br(Q))NmrB_{r}(Q)\cap{}^{Nm_{r}}(B_{r}(Q))
Proof.

By the construction of the integer mrm_{r} (see Notation 4.15), we know that

(Qr)rNmr[(Pr)r5,(Pr)r6](using Notation 4.3){}^{Nm_{r}}_{r}(Q_{r})\in\big{[}{}_{r}^{5}(P_{r}),{}_{r}^{6}(P_{r})\big{]}\qquad\text{(using Notation \ref{not:Reeb_intervals})}

Since r and rH{}^{H}_{r} agree on the Reeb segment U\Gamma\cap U, and ()rHNmr({}^{H}_{r})^{Nm_{r}} maps the interval Br(Q)\Gamma\cap B_{r}(Q) to an interval in containing [0,(Pr)r5][0,{}^{5}_{r}(P_{r})], we thus deduce that

Pr([Qr,Qr])Nmrwhere±Qr=0s×±r(1e8Nr)×0u)P_{r}\in{}^{Nm_{r}}([-Q_{r},Q_{r}])\qquad\text{where}\qquad\pm Q_{r}=0_{s}\times\pm r(1-e^{-8Nr})\times 0_{u})

Let x[Qr,Qr]x\in[-Q_{r},Q_{r}] be the point with (x)Nmr=Pr{}^{Nm_{r}}(x)=P_{r} and let DBr(Q)D\subset B_{r}(Q) be the disk

D=0s×x×Dun(lμmr)contained inFu(,rx)Br(Q)D=0_{s}\times x\times D^{n}_{u}(l\cdot\mu^{-m_{r}})\qquad\text{contained in}\qquad F^{u}({}_{r},x)\cap B_{r}(Q)

This disk is a disk of radius lμmrl\cdot\mu^{-m_{r}} in Fu(,rx)F^{u}({}_{r},x). Since rN{}^{N}_{r} uniformly expands distances in leaves of FuF^{u} (see Lemma 4.5), the disk (D)Nmr{}^{Nm_{r}}(D) has radius larger than ll in Fu(,rPr)F^{u}({}_{r},P_{r}). It follows from the definition of ll that

a(D)Nmr(Br(Q))Nmra\in{}^{Nm_{r}}(D)\subset{}^{Nm_{r}}(B_{r}(Q))\qed
Definition 5.5 (Blender Set AA^{\prime}).

The sets Ar(Q)A^{\prime}_{r}(Q) and Atilder(Q)\tilde{A}_{r}(Q) are the components of the intersections

Br(Q)(Br(Q))NmrandBtilder(Q)(Btilder(Q))NmrB_{r}(Q)\cap{}^{Nm_{r}}(B_{r}(Q))\qquad\text{and}\qquad\tilde{B}_{r}(Q)\cap{}^{Nm_{r}}(\tilde{B}_{r}(Q))

that contain the point a=(1s,0t,0u)a=(1_{s},0_{t},0_{u}), respectively. Note that Ar(Q)Atilder(Q)A_{r}(Q)\subset\tilde{A}_{r}(Q).

We next begin working towards the proof of the corresponding axiom, Definition 3.22(b). We need some technical lemmas about the holonomy map (see Construction 5.1). Consider the set

Cr=Dsn(5μmr)×[L2r,L]t×0uC_{r}=D^{n}_{s}(5\cdot\mu^{-m_{r}})\times[L-2r,L]_{t}\times 0_{u}

The holonomy from Construction 5.1 is well-defined on CrC_{r} for small rr, yielding a smooth map

Holr:Cr[3,3]sn×[ϵ,L+ϵ]t×0ufor small r\operatorname{Hol}_{{}_{r}}:C_{r}\to[-3,3]^{n}_{s}\times[-\epsilon,L+\epsilon]_{t}\times 0_{u}\qquad\text{for small }r

Our first goal is to analyze the image of CrC_{r} under the holonomy map.

Lemma 5.6 (Holonomy Estimates).

There is a κ>0\kappa>0 independent of rr with the following property. Fix

xCrandy=Holr(x)x\in C_{r}\qquad\text{and}\qquad y=\operatorname{Hol}_{{}_{r}}(x)

Then for sufficiently small rr, the coordinates of yy satisfy

|s(y)|μκmrand|t(y)(r+t(x)L)|μκmr|s(y)|\leq\mu^{-\kappa m_{r}}\qquad\text{and}\qquad|t(y)-(r+t(x)-L)|\leq\mu^{-\kappa m_{r}}
Proof.

If xx has s(x)=0s(x)=0, then (5.3) says that s(y)=0s(y)=0 and t(y)=r+t(x)Lt(y)=r+t(x)-L. In general, let x=π(x)x^{\prime}=\pi(x) be the projection of xx to the tt-axis and let y=Holr(x)y^{\prime}=\operatorname{Hol}_{{}_{r}}(x^{\prime}). Then by Construction 5.1

dist(y,y)dist(x,x)κ=|s(x)|κ(5lμmr)κ\operatorname{dist}(y^{\prime},y)\leq\operatorname{dist}(x^{\prime},x)^{\kappa}=|s(x)|^{\kappa}\leq(5l\mu^{-m_{r}})^{\kappa}

Here κ\kappa is the Hölder constant in Construction 5.1. The implies that

|s(y)|=|s(y)s(y)|(5lμmr)κand|s(y)(r+t(x)L)|(5lμmr)κ|s(y)|=|s(y)-s(y^{\prime})|\leq(5l\mu^{-m_{r}})^{\kappa}\qquad\text{and}\qquad|s(y)-(r+t(x)-L)|\leq(5l\mu^{-m_{r}})^{\kappa}

Since mrm_{r} grows faster than 1/r1/r, we can eliminate the factor of 5l5l by shrinking κ\kappa. ∎

Next let r be the union of the disks DD in the unstable foliation satisfying the following properties.

D{|u|lμmr}andD{u=0}Holr(Cr)\partial D\subset\{|u|\leq l\cdot\mu^{-m_{r}}\}\qquad\text{and}\qquad D\cap\{u=0\}\in\operatorname{Hol}_{{}_{r}}(\partial C_{r})
Lemma 5.7.

The set r is disjoint from B𝑡𝑖𝑙𝑑𝑒r(Q)(B𝑡𝑖𝑙𝑑𝑒r(Q))Nmr\tilde{B}_{r}(Q)\cap{}^{Nm_{r}}(\tilde{B}_{r}(Q)).

Proof.

Fix zrz\in{}_{r} and let DD be the corresponding unstable disk of radius lμmrl\cdot\mu^{-m_{r}}, centered at y=Holr(x)y=\operatorname{Hol}_{{}_{r}}(x) where xCx\in\partial C. Recall that Sr(Q)S_{r}(Q) is the intersection of Btilder(Q)\tilde{B}_{r}(Q) with {u=0}\{u=0\}, given by

Sr(Q)=Dsn(2+μmr/2)×Dt1(t((Qr)rN))×0uS_{r}(Q)=D^{n}_{s}(2+\mu^{-m_{r}/2})\times D^{1}_{t}(t({}_{r}^{N}(Q_{r})))\times 0_{u}

Note that PP is connected to Hol(a)\operatorname{Hol}(a) by a path of length less than l/3l/3 in the unstable leaf Fu(,P)F^{u}(\Phi,P). Moreover, Holr\operatorname{Hol}_{{}_{r}} and CrC_{r} converge to Hol\operatorname{Hol} and PP as r0r\to 0. Therefore Holr(x)\operatorname{Hol}_{{}_{r}}(x) and xx are connected by a path of length less than l/2l/2 in Fu(,rx)F^{u}({}_{r},x) for small rr, and zz is connected to xx by a path in of length less than ll. Since (Btilder(Q))Nmr{}^{Nm_{r}}(\tilde{B}_{r}(Q)) contains the unstable disks of radius ll around (Sr(Q))Nmr{}^{Nm_{r}}(S_{r}(Q)), it follows that

z(Btilder(Q))Nmrif and only ifx(Sr(Q))Nmrz\in{}^{Nm_{r}}(\tilde{B}_{r}(Q))\qquad\text{if and only if}\qquad x\in{}^{Nm_{r}}(S_{r}(Q))

We now claim that the intersection (Sr(Q))NmrCr{}^{Nm_{r}}(S_{r}(Q))\cap\partial C_{r} is contained in the set

(5.5) Dsn(5lμmr)×{L2r}t×0uD_{s}^{n}(5\cdot l\cdot\mu^{-m_{r}})\times\{L-2r\}_{t}\times 0_{u}

Indeed, any point in Sr(Q)S_{r}(Q) satisfies |s(x)|2+μmr/2|s(x)|\leq 2+\mu^{-m-r/2} and t(x)t((Qr)rN)t(x)\leq t({}_{r}^{N}(Q_{r})). Therefore it follows from Lemmas 4.9-4.10 and the construction of mrm_{r} that

|s((x)Nmr)|(2+μmr/2)μmr<5μmr|s({}^{Nm_{r}}(x))|\leq(2+\mu^{-m_{r}/2})\cdot\mu^{-m_{r}}<5\mu^{-m_{r}}
t((x)Nmr)t((Qr)N(mr+1))t((Pr)7N)=Lre7Nr<Lt({}^{Nm_{r}}(x))\leq t({}^{N(m_{r}+1)}(Q_{r}))\leq t({}^{7N}(P_{r}))=L-re^{7Nr}<L

By the definition of CrC_{r}, this implies that xx is in (Sr(Q))NmrCr{}^{Nm_{r}}(S_{r}(Q))\cap\partial C_{r} only if xx is in the set (5.5).

Finally, we claim that if xx is in (5.5), then zz is disjoint from Btilder(Q)\tilde{B}_{r}(Q) for small rr. Indeed, by Lemma 5.6, we know that

t(y)r+t(x)L+μκmr=Lr+μκmr<r(1r9Nr)=t((Qr)rN)t(y)\leq r+t(x)-L+\mu^{-\kappa m_{r}}=L-r+\mu^{-\kappa m_{r}}<-r(1-r^{-9Nr})=-t({}_{r}^{N}(Q_{r}))

On the other hand, if zz is in Btilder(Q)\tilde{B}_{r}(Q), then yy must be contained in Sr(Q)S_{r}(Q) which only consists of points yy with t(y)t((Qr)rN)t(y)\geq-t({}^{N}_{r}(Q_{r})). This shows that zz must be disjoint from Btilder(Q)(Btilder(Q))Nmr\tilde{B}_{r}(Q)\cap{}^{Nm_{r}}(\tilde{B}_{r}(Q)), concluding the proof. ∎

We now apply Lemmas 5.6 and 5.7 to acquire some estimates on Ar(Q)A^{\prime}_{r}(Q) and Atilder(Q)\tilde{A}_{r}(Q).

Lemma 5.8 (Bounds On Atilde\tilde{A}).

There is a κ>0\kappa>0 so that for any point zz in A𝑡𝑖𝑙𝑑𝑒r(Q)\tilde{A}_{r}(Q) and small rr, we have

(5.6) |s(z)1s|μκmrt(Qr)t(z)r(1e6Nr)+μκmr|u(z)|lμmr|s(z)-1_{s}|\leq\mu^{-\kappa m_{r}}\qquad-t(Q_{r})\leq t(z)\leq r(1-e^{-6Nr})+\mu^{-\kappa m_{r}}\qquad|u(z)|\leq l\cdot\mu^{-m_{r}}
Proof.

The bound on u(z)u(z) and the lower bound on tt follow since Atilder(Q)Btilder(Q)\tilde{A}_{r}(Q)\subset\tilde{B}_{r}(Q).

For the remaining bounds, let VrUV_{r}\subset U denote the tube [2,2]sn×[δ,L+δ]×Dun(lμmr)[-2,2]^{n}_{s}\times[-\delta,L+\delta]\times D^{n}_{u}(l\cdot\mu^{-m_{r}}). Note that rVr{}_{r}\subset V_{r} and VrrV_{r}\setminus{}_{r} consists of two components VrinV^{\operatorname{in}}_{r} and VroutV^{\operatorname{out}}_{r} where

Vrin={zD:D is unstable disk with DSr(Q)Holr(int(Cr)) and D{|u|=lμmr}}V^{\operatorname{in}}_{r}=\big{\{}z\in D\;:\;D\text{ is unstable disk with $D\cap S_{r}(Q)\in\operatorname{Hol}_{{}_{r}}(\operatorname{int}(C_{r}))$ and $\partial D\subset\{|u|=l\cdot\mu^{-m_{r}}\}$}\big{\}}

We will not need a description of VroutV^{\operatorname{out}}_{r}. Since Btilder(Q)Vr\tilde{B}_{r}(Q)\subset V_{r} by construction, Atilder(Q)Vr\tilde{A}_{r}(Q)\subset V_{r} as well. By Lemma 5.7, Atilder(Q)\tilde{A}_{r}(Q) must be in one of the components VrinV^{\operatorname{in}}_{r} and VroutV^{\operatorname{out}}_{r}. A direct computation gives aVrina\in V_{r}^{\operatorname{in}}, so it follows from Definition 5.5 that Atilder(Q)Vrin\tilde{A}_{r}(Q)\subset V^{\operatorname{in}}_{r}.

It therefore suffices to prove the remaining bounds for a point zz in VrinV^{\operatorname{in}}_{r}. Let DD be the unstable disk through zz and let y=Holr(x)y=\operatorname{Hol}_{{}_{r}}(x) be the intersection of DD with Dr(Q)D_{r}(Q). By Lemma 5.6

|s(y)|μκmrandt(y)Lr+t(x)Lr+t((Pr)r7)=r(1e7Nr)|s(y)|\leq\mu^{-\kappa m_{r}}\qquad\text{and}\qquad t(y)\leq L-r+t(x)\leq L-r+t({}^{7}_{r}(P_{r}))=r(1-e^{7Nr})

The point zz lies on the vertical disk DD, which is a graph of a 2ϵ2\epsilon-Lipschitz graph from Dun(lμmr)D^{n}_{u}(l\cdot\mu^{-m_{r}}). It follows that dist(z,y)Cμmr\operatorname{dist}(z,y)\leq C\mu^{-m_{r}} for some CC independent of rr, and so

|s(x)|μκmr+Cμmrt(y)r(1=e7Nr)+Cμmr|s(x)|\leq\mu^{-\kappa m_{r}}+C\mu^{-m_{r}}\qquad t(y)\leq r(1=e^{7Nr})+C\mu^{-m_{r}}

The remaining estimates follow by taking rr small and possibly shrinking κ\kappa. ∎

The axiom in Definition 3.22(b) is now an easy consequence of Lemma 5.8.

Lemma 5.9 (Blender Axiom B).

The regions A𝑡𝑖𝑙𝑑𝑒r(Q)\tilde{A}_{r}(Q) and Ar(Q)A^{\prime}_{r}(Q) is disjoint from

rBr(Q)sBr(Q)and(uBr(Q))rN\partial^{r}B_{r}(Q)\qquad\partial^{s}B_{r}(Q)\quad\text{and}\quad{}_{r}^{N}(\partial^{u}B_{r}(Q))
Proof.

Disjointness from (uBr(Q))N{}^{N}(\partial^{u}B_{r}(Q)) follows from the same argument as in Lemma 5.3. For the other two boundary regions, note that by Lemma 5.8 we have

|s(x)1s|μκmrandt(x)r(1e7Nr)+μκmrfor all xAtilder(Q)|s(x)-1_{s}|\mu^{-\kappa m_{r}}\quad\text{and}\quad t(x)\leq r(1-e^{-7Nr})+\mu^{-\kappa m-r}\quad\text{for all }x\in\tilde{A}_{r}(Q)

This implies that for sufficiently small rr, the set Atilder(Q)\tilde{A}_{r}(Q) is disjoint from the sets

sBr(Q)=Br(Q){|s|=2}andrBr(Q)=Br(Q){t=r(1e8Nr)}\partial^{s}B_{r}(Q)=B_{r}(Q)\cap\{|s|=2\}\qquad\text{and}\qquad\partial^{r}B_{r}(Q)=B_{r}(Q)\cap\{t=r\cdot(1-e^{-8Nr})\}\qed

5.3. Axioms C And D

Next, we prove the blender axioms related to cone-fields (see Definition 3.22(c-d)). The first of these axioms is relatively straightforward.

Lemma 5.10 (Blender Axiom C).

There are compatible cone-fields KsK^{s} and KuK^{u} for Br(Q)B_{r}(Q) of width less than ϵ\epsilon (with respect to the standard metric) that are contracted and dilated (with constant μ\mu) as follows.

()rNKsintKs()rNKuintKu dilates rNKs dilates rNKu({}_{r}^{-N})_{*}K^{s}\subset\operatorname{int}K^{s}\qquad({}_{r}^{N})_{*}K^{u}\subset\operatorname{int}K^{u}\qquad{}_{r}^{-N}\text{ dilates }K^{s}\qquad{}_{r}^{N}\text{ dilates }K^{u}
Proof.

Let KsK^{s} and KuK^{u} be the cone-fields in Lemma 4.5. Note that r{}_{r}\to\Phi in the CC^{\infty}-topology. Moreover, contraction and dilation (with constant μ\mu) of a given cone-field are C1C^{1}-robust properties. Thus, Lemma 4.5 implies this axiom for our choice of constants N,μ,ϵN,\mu,\epsilon and small rr. ∎

The second cone-field related axiom (regarding the central-unstable cone-field) is more difficult and will require some preliminary results. To start, choose a Riemannian metric

g on TYsuch that the splitting Eu()rTt1T is orthogonal on Usng\text{ on }TY\qquad\text{such that the splitting }E^{u}({}_{r})\oplus T{}^{1}_{t}\oplus T{}^{n}_{s}\text{ is orthogonal on $U$}

We consider the following cone-fields of width ϵ\epsilon with respect to gg.

Kϵus=KϵEu()r(Eu()rT)sn={u+v:uEu()r and vT with sn|u|ϵv}K^{us}_{\epsilon}=K_{\epsilon}E^{u}({}_{r})\cap(E^{u}({}_{r})\oplus T{}^{n}_{s})=\{u+v\;:\;u\in E^{u}({}_{r})\text{ and }v\in T{}^{n}_{s}\text{ with }|u|\geq\epsilon\cdot v\}
Kδcs=Kδ(T)t1(Tt1T)sn={u+v:uT and t1vT with sn|u|δv}K^{cs}_{\delta}=K_{\delta}(T{}^{1}_{t})\cap(T{}^{1}_{t}\oplus T{}^{n}_{s})=\{u+v\;:\;u\in T{}^{1}_{t}\text{ and }v\in T{}^{n}_{s}\text{ with }|u|\geq\delta\cdot v\}

These cone-fields are well-defined over UU (i.e. wherever (s,t,u)(s,t,u)-coordinates are well-defined). Finally, we define the fiberwise sum of cones

Kδ,ϵcu=Kϵus+KδcsK^{cu}_{\delta,\epsilon}=K^{us}_{\epsilon}+K^{cs}_{\delta}

The following key lemma describes choices of the parameters of the cone KcuK^{cu} that guarantee contraction and dilation.

Lemma 5.11 (Stretching KcuK^{cu}).

Fix a positive integer kk, positive constants μ,ϵ,ν,η\mu,\epsilon,\nu,\eta , and a subset VUV\subset U with the following properties.

  • The constants satisfy μ2>1+ϵ2\mu^{2}>1+\epsilon^{2} and ν>1>η\nu>1>\eta.

  • TrkT{}^{k}_{r} dilates Eu()rE^{u}({}_{r}) with constant μ\mu and TrkT{}^{k}_{r} dilates TsnT{}^{n}_{s} with constant μ1\mu^{-1} over VV.

  • T(R)rk=νR+wT{}^{k}_{r}(R)=\nu\cdot R+w where wEs()=Tsnw\in E^{s}(\Phi)=T{}^{n}_{s} and |w|η|R||w|\leq\eta\cdot|R| over the subset VV, where R=tR=\partial_{t}.

Then Kϵ,δcuK^{cu}_{\epsilon,\delta} is contracted and uniformly dilated by kr{}_{r}^{k} over the susbet VV for

η1μ1<δ<ν21\frac{\eta}{1-\mu^{-1}}<\delta<\sqrt{\nu^{2}-1}
Proof.

To prove that Kϵ,δcuK^{cu}_{\epsilon,\delta} is uniformly dilated with some positive constant of dilation, we write an arbitrary vector vv in Kϵ,δcuK^{cu}_{\epsilon,\delta} as follows.

v=(vu+vs)+(aR+ws)where|vu|ϵ|vs| and |aR|=aδ|ws|v=(v^{u}+v^{s})+(aR+w^{s})\qquad\text{where}\qquad|v^{u}|\geq\epsilon|v^{s}|\text{ and }|aR|=a\geq\delta|w^{s}|

We then compute the norm of the image of vv under Nk.

|(v)rk|2=|(vu)rk|2+|(aR)rk|2+|(vs+ws)rk|2|{}^{k}_{r}(v)|^{2}=|{}^{k}_{r}(v^{u})|^{2}+|{}^{k}_{r}(aR)|^{2}+|{}^{k}_{r}(v^{s}+w^{s})|^{2}
μ2|vu|2+ν2|aR|2+|aw|2+μ2|vs+ws|2μ2|vu|2+ν2|aR|2\geq\mu^{2}\cdot|v^{u}|^{2}+\nu^{2}\cdot|aR|^{2}+|aw|^{2}+\mu^{-2}\cdot|v^{s}+w^{s}|^{2}\geq\mu^{2}\cdot|v^{u}|^{2}+\nu^{2}\cdot|aR|^{2}

Now since |vu|ϵ|vs||v^{u}|\geq\epsilon|v^{s}| and |aR|δ|ws||aR|\geq\delta|w^{s}|, we see that

(1+ϵ2)|vu|2|vu+vs|2and(1+δ2)|aR|2|aR+ws|2(1+\epsilon^{2})\cdot|v^{u}|^{2}\geq|v^{u}+v^{s}|^{2}\qquad\text{and}\qquad(1+\delta^{2})\cdot|aR|^{2}\geq|aR+w^{s}|^{2}

Finally, we calculate that

|(v)rNk|2μ1+ϵ2|vu+vs|2+ν21+δ2|aR|2min(μ1+ϵ2,ν21+δ2)|v|2|{}^{Nk}_{r}(v)|^{2}\geq\frac{\mu}{1+\epsilon^{2}}\cdot|v^{u}+v^{s}|^{2}+\frac{\nu^{2}}{1+\delta^{2}}\cdot|aR|^{2}\geq\operatorname{min}(\frac{\mu}{1+\epsilon^{2}},\frac{\nu^{2}}{1+\delta^{2}})\cdot|v|^{2}

Since μ2>1+ϵ2\mu^{2}>1+\epsilon^{2} and ν2>1+δ2\nu^{2}>1+\delta^{2} by assumption, we thus find that vv is uniformly expanded.

To prove that Kδ,ϵcuK^{cu}_{\delta,\epsilon} is contracted, we argue as follows. First, note that for any v=vu+vsv=v^{u}+v^{s} in KϵusK^{us}_{\epsilon} with |vu|ϵ|vs||v^{u}|\geq\epsilon|v^{s}|, we have

(v)rk=(vu)rk+(vs)rkwhere(vu)rkEu()r and (vs)rkTsn{}^{k}_{r}(v)={}^{k}_{r}(v^{u})+{}^{k}_{r}(v^{s})\qquad\text{where}\qquad{}^{k}_{r}(v^{u})\in E^{u}({}_{r})\text{ and }{}^{k}_{r}(v^{s})\in T{}^{n}_{s}

By our hypothesis on the dilation of TrkT{}_{r}^{k}, we know that

|T(vu)rk|μ2ϵ|T(vs)rk|>ϵ|T(vs)rk||T{}^{k}_{r}(v^{u})|\geq\mu^{2}\cdot\epsilon\cdot|T{}^{k}_{r}(v^{s})|>\epsilon\cdot|T{}^{k}_{r}(v^{s})|

Therefore T(v)rkT{}_{r}^{k}(v) is strictly contracted by TrkT{}_{r}^{k}. Likewise, take any vector v=aR+vsv=aR+v^{s} in KδusK^{us}_{\delta} with |aR|δ|vs||aR|\geq\delta\cdot|v^{s}|. Then

T(v)rk=aνR+(aw+(ws)rkT{}^{k}_{r}(v)=a\nu R+(aw+{}^{k}_{r}(w^{s})

Now we note that we have the following estimate.

|aw+(ws)rk||aw|+|(ws)k|η|aR|+ϵμ1|aR|(η+δμ1)|aR||aw+{}^{k}_{r}(w^{s})|\leq|aw|+|{}^{k}(w^{s})|\leq\eta|aR|+\epsilon\cdot\mu^{-1}\cdot|aR|\leq(\eta+\delta\cdot\mu^{-1})\cdot|aR|

By assumption we have η+δμ1<δ\eta+\delta\cdot\mu^{-1}<\delta and thus TrkT{}^{k}_{r} contracts the cone KδusK^{us}_{\delta}. We have thus proven that

T(Kδ,ϵcu)rkintKδ,ϵcuT{}^{k}_{r}(K^{cu}_{\delta,\epsilon})\subset\operatorname{int}K^{cu}_{\delta,\epsilon}\qed

We next verify the third criterion in Lemma 5.11 in the cases relevant to our axiom. Recall that we have fixed constants N,mN,m (see the beginning of Section 4.2).

Lemma 5.12 (Axiom D, Part 1).

For sufficiently small rr, we have

T(R)rN=eNrRover the subset Ar(Q)T{}^{N}_{r}(R)=e^{Nr}\cdot R\qquad\text{over the subset }A_{r}(Q)
Proof.

For sufficiently small rr, we have

(Br(Q))rj[2,2]s×[δ,L/3]t×[ϵ,ϵ] for all j=0,,N{}^{j}_{r}(B_{r}(Q))\subset[-2,2]_{s}\times[-\delta,L/3]_{t}\times[-\epsilon,\epsilon]\text{ for all }j=0,\dots,N

Here δ,ϵ,L\delta,\epsilon,L are the parameters of the chart in Lemma 4.4. In this region, we know that =rrH{}_{r}={}^{H}_{r}\circ\Psi. Therefore by Construction 4.6 (and more specifically (4.8)) we find that

T(R)rH=erRandT(R)=Rand thereforeT(R)rN=eNrRT{}^{H}_{r}(R)=e^{r}R\quad\text{and}\quad T\Psi(R)=R\qquad\text{and therefore}\quad T{}_{r}^{N}(R)=e^{Nr}\cdot R\qed

For the other part of Axiom D, we require the following lemma tracking the behavior of the set Ar(Q)A^{\prime}_{r}(Q) under r1{}^{-1}_{r}.

Lemma 5.13 (Atilde\tilde{A} Stays In UU).

Let mm and WW be as in Construction 4.7 and (4.10). Then

Atilder(Q)(W)rNmand(Atilder(Q))jU for all j=Nm,,mrfor small r\tilde{A}_{r}(Q)\subset{}^{Nm}_{r}(W)\qquad\text{and}\qquad{}^{-j}(\tilde{A}_{r}(Q))\subset U\text{ for all }j=Nm,\dots,m_{r}\qquad\text{for small $r$}
Proof.

For the first claim, note that NN and mm are independent of rr and r{}_{r}\to\Phi in CC^{\infty} as r0r\to 0. Thus we may choose an rr-independent open neighborhood V(W)rNmV\subset{}^{Nm}_{r}(W) with aVa\in V. By Lemma 5.8, we know that Atilder(Q)\tilde{A}_{r}(Q) is contained in the region

Br(Q){|s1s|βlμκmr}for β,κ>0B_{r}(Q)\cap\{|s-1_{s}|\leq\beta\cdot l\cdot\mu^{-\kappa m_{r}}\}\qquad\text{for $\beta,\kappa>0$}

It follows from Construction 4.17 that this region is contained in a ball of radius bounded by rr around aa. Thus for small rr, this region is contained in VV and the first claim is proven.

For the second claim, we require some preliminary observations. Consider the subsets \Xi\subset\Gamma and Wlocs(;U)\Sigma\subset W_{\operatorname{loc}}^{s}(\Phi;U) given by

=0s×[0,t(Qr)]t×0u=[Q,Qr]and=Dsn(5/2)×[0,t(Qr)]t×0u\Xi=0_{s}\times[0,t(Q_{r})]_{t}\times 0_{u}=[Q,Q_{r}]\qquad\text{and}\qquad\Sigma=D^{n}_{s}(5/2)\times[0,t(Q_{r})]_{t}\times 0_{u}

Note that Btilder(Q){u=0,t0}\tilde{B}_{r}(Q)\cap\{u=0,t\geq 0\} is contained in by Construction 4.17 and (4.16). Let ψr\psi_{r} be the flow of hth\cdot\partial_{t} (see Lemma 4.10). Then ψrj(0)=0\psi_{r}^{j}(0)=0 for all jj and by construction of mrm_{r}, we know that

ψrj(t(Qr))[δ,L+δ]tand(Qr)rUfor all j=0,,Nmr\psi_{r}^{j}(t(Q_{r}))\in[-\delta,L+\delta]_{t}\quad\text{and}\quad{}_{r}(Q_{r})\in\Gamma\cap U\qquad\text{for all }j=0,\dots,Nm_{r}

Moreover, since ()r=0s×[0,ψr(t(Qr))]t×0u{}_{r}(\Xi)=0_{s}\times[0,\psi_{r}(t(Q_{r}))]_{t}\times 0_{u} and r contracts the ss-coordinate (see Lemma 4.10), this implies that

()rjUand()rj{u=0}Ufor all j=0,,Nmr{}_{r}^{j}(\Xi)\subset\Gamma\cap U\quad\text{and}\quad{}_{r}^{j}(\Sigma)\subset\{u=0\}\cap U\qquad\text{for all }j=0,\dots,Nm_{r}

Now we prove the second claim. By the first claim and the definition of Atilder(Q)\tilde{A}_{r}(Q), we know that

(5.7) (Atilder(Q))NmWUand(Atilder(Q))rNm(Btilder(Q))rNmrNm{}^{-Nm}(\tilde{A}_{r}(Q))\subset W\subset U\qquad\text{and}\qquad{}^{-Nm}_{r}(\tilde{A}_{r}(Q))\subset{}^{Nm_{r}-Nm}_{r}(\tilde{B}_{r}(Q))

From these two inclusions, it follows that (Atilder(Q))rNm{}^{-Nm}_{r}(\tilde{A}_{r}(Q)) is included in the set

V={xFu(;ry):y()rNmrNmand|u(x)|2}V^{\prime}=\{x\in F^{u}({}_{r};y)\;:\;y\in{}^{Nm_{r}-Nm}_{r}(\Sigma)\quad\text{and}\quad|u(x)|\leq 2\}

Here we choose WW in Construction 4.6 so that WW is contained in {|u|2}\{|u|\leq 2\}. Finally, we note that

(V)rjUfor all j=0,,NmrNm{}_{r}^{-j}(V^{\prime})\subset U\qquad\text{for all }j=0,\dots,Nm_{r}-Nm

Indeed, the intersection of (V)rj{}_{r}^{-j}(V^{\prime}) with {u=0}\{u=0\} is ()NmrNmj{}^{Nm_{r}-Nm-j}(\Sigma) and the union V′′V^{\prime\prime} of unstable disks intersecting ()NmrNmj{}^{Nm_{r}-Nm-j}(\Sigma) with boundary on {u=3/2}\{u=3/2\} is contained in UU. On the other hand, since r1{}^{-1}_{r} contracts distances in UU, (V)rjV′′{}_{r}^{-j}(V^{\prime})\subset V^{\prime\prime}. This proves the second claim. ∎

Lemma 5.14 (Axiom D, Part 2).

For sufficiently small rr, we have

T(R)Nmr=λrR+vover(Ar(Q))NmrT{}^{Nm_{r}}(R)=\lambda_{r}\cdot R+v\qquad\text{over}\quad{}^{-Nm_{r}}(A^{\prime}_{r}(Q))

Here λrr1/2\lambda_{r}\geq r^{-1/2} and |v|r2μlog(r)/r|v|\leq r^{2}\cdot\mu^{-\log(r)/r} for small rr.

Proof.

Fix an arbitrary point xx in (Atilder(Q))Nmr{}^{-Nm_{r}}(\tilde{A}_{r}(Q)). We let nrstn^{\operatorname{st}}_{r} and nrmidn^{\operatorname{mid}}_{r} denote the quantities

nrst=min{j:t((x)rNj)L/3}andnrmid=min{jnrst:t((x)rNj)2L/3}n_{r}^{\operatorname{st}}=\operatorname{min}\big{\{}j\>:\;t({}_{r}^{Nj}(x))\geq L/3\big{\}}\qquad\text{and}\qquad n_{r}^{\operatorname{mid}}=\operatorname{min}\big{\{}j-n^{\operatorname{st}}_{r}\>:\;t({}_{r}^{Nj}(x))\geq 2L/3\big{\}}

Finally, let mm and WW be as in Construction 4.7 and (4.10), and let

nrend=mrNmnrstnrmidn^{\operatorname{end}}_{r}=m_{r}-Nm-n^{\operatorname{st}}_{r}-n^{\operatorname{mid}}_{r}

By essentially identical analysis to Lemma 4.16, we have the following asymptotic behavior for these quantities.

(5.8) nrst2r1log(r)nrmidr1andnrendr1log(r)n^{\operatorname{st}}_{r}\sim-2r^{-1}\cdot\log(r)\qquad n^{\operatorname{mid}}_{r}\sim r^{-1}\quad\text{and}\quad n^{\operatorname{end}}_{r}\sim-r^{-1}\cdot\log(r)

Here \sim means the limit of the ratio is 11 as r0r\to 0. We will analyze the sequence of terms T(R)r,xNjT{}^{Nj}_{r,x}(R) starting at the point xx in four regimes, corresponding to the following intervals for the index jj.

[0,nrst][nrst,nrst+nrmid][nrst+nrmid,mrNm][mrNm,mr][0,n^{\operatorname{st}}_{r}]\qquad[n^{\operatorname{st}}_{r},n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r}]\qquad[n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r},m_{r}-Nm]\qquad[m_{r}-Nm,m-r]

We will call these the starting regime, middle regime, ending regime and extra regime. Crucially, by Lemma 5.13, we know that (x)rNj{}^{Nj}_{r}(x) stays in UU for the first three periods.

Step 1: Starting Regime. Start by assuming that jj is in the interval [0,nrst][0,n^{\operatorname{st}}_{r}]. In this regime, t((r)Nj)L/3t({}_{Nj}(r))\leq L/3. It follows from (4.12) and Construction 4.6 (see (4.8)) that

(5.9) T(R)rNj=(TrHT)Nj(R)=(T)rHNj(R)=erNnrstRT{}^{Nj}_{r}(R)=(T{}^{H}_{r}\circ T\Phi)^{Nj}(R)=(T{}^{H}_{r})^{Nj}(R)=e^{rNn^{\operatorname{st}}_{r}}\cdot R

Step 2: Middle Regime. Next, assume that jj is in the middle interval [nrst,nrst+nrmid][n^{\operatorname{st}}_{r},n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r}]. In this case, it follows from the general form for rH{}^{H}_{r} in (4.7) that for any yy in UU with t(y)[L/3,2L/3]t(y)\in[L/3,2L/3], we have

(5.10) T(R)r,yH=tψr(t(y))R+tfr(t(y))(s(y)s)T{}^{H}_{r,y}(R)=\partial_{t}\psi_{r}(t(y))\cdot R+\partial_{t}f_{r}(t(y))(s(y)\cdot\partial_{s})

under the splitting TU=TsnTt1TunTU=T{}^{n}_{s}\oplus T{}^{1}_{t}\oplus T{}^{n}_{u}. Here ψ\psi is given in (4.7) and satisfies

ψ0(t)=trψ=hψ\psi_{0}(t)=t\qquad\partial_{r}\psi=h\circ\psi

By the assumption that |h|1/L|h^{\prime}|\leq 1/L on the interval [0,L][0,L], we know that |h|1|h|\leq 1 and so

(5.11) |tψr(t(y))|erfor small r|\partial_{t}\psi_{r}(t(y))|\geq e^{-r}\qquad\text{for small }r

Next note that |tfr(t(y))||\partial_{t}f_{r}(t(y))| is bounded by some constant C>0C>0 for small rr. Finally, Nr{}_{r}^{N} and N both preserve TsnT{}^{n}_{s} and contract TsnT{}^{n}_{s} by a factor of μ1\mu^{-1} for small rr (see Lemmas 4.5 and 4.11).

Now let xstx_{\operatorname{st}} and RstR_{\operatorname{st}} denote the point and vector that is left after we exit the early regime.

xst=(x)rNnrstandRst=T(R)r,xNnrstspan(R)x_{\operatorname{st}}={}^{Nn^{\operatorname{st}}_{r}}_{r}(x)\qquad\text{and}\qquad R_{\operatorname{st}}=T{}_{r,x}^{Nn^{\operatorname{st}}_{r}}(R)\in\operatorname{span}(R)

It follows from the discussion above and (5.10) that we can write the following expansion.

(5.12) T(Rst)r,xstNnrmid=T()Hr,xstNnrmid(Rst)+\slimits@k=0NnrmidwkT{}^{Nn^{\operatorname{mid}}_{r}}_{r,x_{\operatorname{st}}}(R_{\operatorname{st}})=T({}^{H})^{Nn^{\operatorname{mid}}_{r}}_{r,x_{\operatorname{st}}}(R_{\operatorname{st}})+\sumop\slimits@_{k=0}^{Nn^{\operatorname{mid}}_{r}}w_{k}

Here the vectors wkw_{k} can be written as follows.

wk=T(ck(s((xst)rk)s))rNnrmidkTsnwhereck=rfr(t((xst)rk))w_{k}=T{}_{r}^{Nn^{\operatorname{mid}}_{r}-k}(c_{k}\cdot(s({}^{k}_{r}(x_{\operatorname{st}}))\cdot\partial_{s}))\in T{}^{n}_{s}\qquad\text{where}\qquad c_{k}=\partial_{r}f_{r}(t({}^{k}_{r}(x_{\operatorname{st}})))

Now we estimate the terms appearing in (5.12). First, by (5.11) we know that

|T()Hr,xstNnrmid(Rst)|eNnrmid|Rst|andT()Hr,xstNnrmid(Rst)span(R)|T({}^{H})^{Nn^{\operatorname{mid}}_{r}}_{r,x_{\operatorname{st}}}(R_{\operatorname{st}})|\geq e^{-Nn^{\operatorname{mid}}_{r}}\cdot|R_{\operatorname{st}}|\qquad\text{and}\qquad T({}^{H})^{Nn^{\operatorname{mid}}_{r}}_{r,x_{\operatorname{st}}}(R_{\operatorname{st}})\in\operatorname{span}(R)

Next we estimate the norm of wkw_{k}. As noted previously, |ck|C|c_{k}|\leq C for some CC independent of kk and rr small. Since (xst)rk{}^{k}_{r}(x_{\operatorname{st}}) is in the image of UU under the (Nnrst+k)(Nn_{r}^{\operatorname{st}}+k)-th power of r and r uniformly contracts TsnT{}^{n}_{s} by a factor of μ1\mu^{-1}, we know that the ss-vector s((xst)k)ss({}^{k}(x_{\operatorname{st}}))\cdot\partial_{s} is bounded as follows.

|s((xst)k)s|2μnrstk/N|s({}^{k}(x_{\operatorname{st}}))\cdot\partial_{s}|\leq 2\cdot\mu^{-n^{\operatorname{st}}_{r}-\lfloor k/N\rfloor}

Finally, TrNT{}^{N}_{r} uniformly contracts vectors in TsnT{}^{n}_{s} by a factor of μ1\mu^{-1}. Combining these estimates, we find that for some constant C>0C^{\prime}>0 independent of xx and small rr, we have

|wk|Cμ(nrst+nrmid)and|\slimits@k=0Nnrmidwk|CNnrmidμ(nrst+nrmid)C′′1rμ2log(r)/r|w_{k}|\leq C^{\prime}\cdot\mu^{-(n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r})}\qquad\text{and}\qquad\Big{|}\sumop\slimits@_{k=0}^{Nn^{\operatorname{mid}}_{r}}w_{k}\Big{|}\leq C^{\prime}\cdot Nn^{\operatorname{mid}}_{r}\mu^{-(n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r})}\leq C^{\prime\prime}\cdot\frac{1}{r}\cdot\mu^{-2\log(r)/r}

The outcome of this analysis of the middle regime is the following formula.

(5.13) Rmid=T(R)r,xN(nrst+nrmid)=erNnrstT(R)r,xNnrmid=AmiderNnrstR+wR_{\operatorname{mid}}=T{}_{r,x}^{N(n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r})}(R)=e^{rNn^{\operatorname{st}}_{r}}\cdot T{}_{r,x}^{Nn^{\operatorname{mid}}_{r}}(R)=A_{\operatorname{mid}}\cdot e^{rNn^{\operatorname{st}}_{r}}\cdot R+w

Here AmidA_{\operatorname{mid}} is some constant bounded by eNnrmide^{-Nn^{\operatorname{mid}}_{r}} and ww is a vector in TsnT{}^{n}_{s} with norm bounded by the quantity C′′r1μ2log(r)/rC^{\prime\prime}\cdot r^{-1}\cdot\mu^{-2\log(r)/r}.

Step 3: Ending Regime. We next examine the ending regime, where jj is in the interval [nrst+nrmid,mrNm][n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r},m_{r}-Nm]. In this regime, t((r)Nj)2L/3t({}_{Nj}(r))\geq 2L/3. By using (4.12) and Construction 4.6 (and specifically (4.9)), we see that we have

T(Rmid)rNnrend=AmiderNnrstT(R)rNnrend+T(w)rNnrend=AmiderN(nrstnrend)R+T(w)rNnrendT{}^{Nn^{\operatorname{end}}_{r}}_{r}(R_{\operatorname{mid}})=A_{\operatorname{mid}}\cdot e^{rNn^{\operatorname{st}}_{r}}\cdot T{}^{Nn^{\operatorname{end}}_{r}}_{r}(R)+T{}^{Nn^{\operatorname{end}}_{r}}_{r}(w)=A_{\operatorname{mid}}\cdot e^{rN(n^{\operatorname{st}}_{r}-n^{\operatorname{end}}_{r})}R+T{}^{Nn^{\operatorname{end}}_{r}}_{r}(w)

Focusing on the first term, the lower bound AmiderNnrmidA_{\operatorname{mid}}\geq e^{-rNn^{\operatorname{mid}}_{r}} and the asymptotic formula (5.8) imply that

AmiderN(nrstnrend)erlog(r)/rr1/2for small rA_{\operatorname{mid}}\cdot e^{rN(n^{\operatorname{st}}_{r}-n^{\operatorname{end}}_{r})}\geq e^{-r\cdot\log(r)/r}\geq r^{-1/2}\qquad\text{for small }r

Moreover, N uniformly contracts TsnT{}^{n}_{s} by a factor of μ1\mu^{-1}, and we thus see that

w=T(w)rNnrendsatisfies|w|Nnrmidμ(nrst+nrmid+nrend)Nmrμmr+Nmw^{\prime}=T{}^{Nn^{\operatorname{end}}_{r}}_{r}(w)\qquad\text{satisfies}\qquad|w^{\prime}|\leq N\cdot n^{\operatorname{mid}}_{r}\cdot\mu^{-(n^{\operatorname{st}}_{r}+n^{\operatorname{mid}}_{r}+n^{\operatorname{end}}_{r})}\leq N\cdot m_{r}\cdot\mu^{-m_{r}+Nm}

Combining all of the analysis up to now, we have proven that

(5.14) T(R)rN(mrm)=λrR+vover (Ar(Q))NmrT{}^{N(m_{r}-m)}_{r}(R)=\lambda_{r}\cdot R+v\qquad\text{over }{}^{-Nm_{r}}(A^{\prime}_{r}(Q))

where λrr1/2\lambda_{r}\geq r^{-1/2} and vTsnv\in T{}^{n}_{s} satisfies |v|r2μlog(r)/r|v|\leq r^{2}\cdot\mu^{-\log(r)/r} for small rr.

Step 4: Extra Regime. Finally, we consider the last regime. By Lemma 5.13, we know that

(x)rN(mrm)W{}^{N(m_{r}-m)}_{r}(x)\subset W

By the construction of r, or more precisely (4.12), we know that

T=rNmTrRTNmTNmrHon WT{}^{Nm}_{r}=T{}^{R}_{r}\circ T{}^{Nm}\circ T{}^{H}_{Nmr}\qquad\text{on }W

Note that: TrRT{}^{R}_{r} preserves RR and TsnT{}^{n}_{s}; TNmT{}^{Nm} preserves RR and shrinks TsnT{}^{n}_{s} by a factor of μ1\mu^{-1}; and TNmrHT{}^{H}_{Nmr} preserves RR and expands TsnT{}^{n}_{s} by at most a factor of eNmre^{Nmr}. It follows that

T(R)rNmr=λrR+vT{}^{Nm_{r}}_{r}(R)=\lambda_{r}\cdot R+v

where λr\lambda_{r} and vv satisfy the same estimates as in (5.14). This concludes the proof.∎

We are (finally) ready to prove the second cone axiom. Thanks to the onerous work of Lemmas 5.12 and 5.14, this will be a simple application of Lemma 5.11.

Lemma 5.15 (Blender Axiom D).

There is a compatible cone-field KcuK^{cu} for Br(Q)B_{r}(Q) of width less than ϵ\epsilon (with respect to the standard metric) that is contracted and dilated uniformly as follows.

()rNKcuintKcuand dilates rNKcu over (Ar(Q))N({}_{r}^{N})_{*}K^{cu}\subset\operatorname{int}K^{cu}\quad\text{and}\quad{}^{N}_{r}\text{ dilates }K^{cu}\qquad\text{ over }{}^{-N}(A_{r}(Q))
()rNmrKcuintKcuand dilates rNmrKcu over (Ar(Q))rNmr({}^{Nm_{r}}_{r})_{*}K^{cu}\subset\operatorname{int}K^{cu}\quad\text{and}\quad{}^{Nm_{r}}_{r}\text{ dilates }K^{cu}\qquad\text{ over }{}_{r}^{-Nm_{r}}(A^{\prime}_{r}(Q))
Proof.

Fix constants N,μ,ϵN,\mu,\epsilon as in Lemma 4.5 such that μ>1+ϵ2\mu>1+\epsilon^{2}. By Lemma 5.13, we may take rr small enough so that TrNT{}^{N}_{r} dilates Eu()rE^{u}({}_{r}) by constant μ\mu and dilates TsnT{}^{n}_{s} by μ1\mu^{-1} over UU. We take

Kcu=Kϵ,δcuK^{cu}=K^{cu}_{\epsilon,\delta}

for a judiciously chosen δ\delta. To choose δ\delta appropriately, note that by Lemma 5.12, we have

T(R)rN=eNrR over the subset (Ar(Q))rNUT{}^{N}_{r}(R)=e^{Nr}\cdot R\qquad\text{ over the subset }{}^{-N}_{r}(A_{r}(Q))\subset U

It follows by Lemma 5.11 that KcuK^{cu} is contracted and uniformly dilated by rN{}^{N}_{r} as long as

(5.15) 0δeNr10\leq\delta\leq\sqrt{e^{Nr}-1}

Likewise, by Lemma 5.14, we know that for large rr

T(R)Nmr=λrR+v over the subset (Ar(Q))rNmrUT{}^{Nm_{r}}(R)=\lambda_{r}\cdot R+v\qquad\text{ over the subset }{}^{-Nm_{r}}_{r}(A_{r}(Q))\subset U

where |λr|r1/2|\lambda_{r}|\geq r^{-1/2} and |v|r2μlog(r)/r|v|\leq r^{2}\cdot\mu^{-\log(r)/r}. It follows by Lemma 5.11 that KcuK^{cu} is contracted and uniformly dilated by rNmr{}^{Nm_{r}}_{r} as long as

(5.16) μlog(r)/r1μ1δr11λr21\frac{\mu^{-\log(r)/r}}{1-\mu^{-1}}\leq\delta\leq\sqrt{r^{-1}-1}\leq\sqrt{\lambda_{r}^{2}-1}

On the other hand, for every small rr we know that

μlog(r)/r1μNmrr2eNr1\frac{\mu^{-\log(r)/r}}{1-\mu^{-Nm_{r}}}\leq r^{2}\leq\sqrt{e^{Nr}-1}

Thus we may choose δ\delta to satisfy both (5.15) and (5.16). The result now follows by Lemma 5.11.∎

5.4. Axioms E And F

Finally, we prove the blender axioms regarding vertical disks, Definition 3.22(e-f). The first such axiom is relatively straightforward.

Lemma 5.16 (Blender Axiom E).

Let DD be a vertical disk through Br(Q)B_{r}(Q) to the right of the local unstable manifold of QQ with respect to r. Then

dist(D,lBr(Q))>r3for small r\operatorname{dist}(D,\partial^{l}B_{r}(Q))>r^{3}\qquad\text{for small }r

In particular, there is a neighborhood UU_{-} of lBr(Q)\partial^{l}B_{r}(Q) that is disjoint from all such disks DD.

Proof.

By Lemma 4.12, the local unstable manifold WW of QQ with respect to r in Br(Q)B_{r}(Q) is given by Dsn(2)×0t×0uD^{n}_{s}(2)\times 0_{t}\times 0_{u}. It follows that a vertical disk DD to the right of WW must intersect the manifold

W=Dsn(2)×Dt1(t(Qr))×0uW^{\prime}=D^{n}_{s}(2)\times D^{1}_{t}(t(Q_{r}))\times 0_{u}

at a point xx with t(x)>0t(x)>0 and u(x)=0u(x)=0. By Lemma 3.21, DD is the graph of a 2ϵ2\epsilon-Lipschitz map

f:Dun(lμμr)Dsn(2)×Dt1(t(Qr))f:D^{n}_{u}(l\cdot\mu^{-\mu_{r}})\to D^{n}_{s}(2)\times D^{1}_{t}(t(Q_{r}))

The image of ff, or equivalently the projection of DD to Dsn(2)×Dt1(t(Qr))D^{n}_{s}(2)\times D^{1}_{t}(t(Q_{r})), is contained in a ball of radius 2ϵlμμr2\epsilon\cdot l\cdot\mu^{-\mu_{r}} and tt-coordinate of DD is lower bounded by

t02ϵlμmr>2ϵlμmrt_{0}-2\epsilon\cdot l\cdot\mu^{-m_{r}}>-2\epsilon\cdot l\cdot\mu^{-m_{r}}

By Construction 4.17, the left boundary lBr(Q)\partial^{l}B_{r}(Q) consists of points in Br(Q)B_{r}(Q) with tt-coordinate t(Qr)-t(Q_{r}). Finally, note that mrm_{r} diverges faster than 1/r1/r as r0r\to 0 by Lemma 4.16. Therefore

dist(D,lBr(Q))>r(1e8Nr)2ϵlμmr>r3for small r \operatorname{dist}(D,\partial^{l}B_{r}(Q))>r\cdot(1-e^{-8Nr})-2\epsilon\cdot l\cdot\mu^{-m_{r}}>r^{3}\qquad\text{for small $r$ }\qed

The final blender axiom is more difficult. We prove the following version.

Lemma 5.17 (Blender Axiom F).

Let DD be a vertical disk through Br(Q)B_{r}(Q) to the right of the local unstable manifold WW of QQ with respect to r. Consider the intersection point

x=(s,t,0u)=DWwhereW:=Dsn×Dt1×0ux=(s,t,0_{u})=D\cap W^{\prime}\qquad\text{where}\qquad W^{\prime}:=D^{n}_{s}\times D^{1}_{t}\times 0_{u}

Then the following two alternatives hold for DD.

  1. (a)

    If 0<t(x)t((Qr)3N)0<t(x)\leq t({}^{-3N}(Q_{r})) then the intersection (D)rNAr(Q){}_{r}^{N}(D)\cap A_{r}(Q) contains a disk Dtilde\tilde{D} through Br(Q)B_{r}(Q) to the right of WW, such that

    dist(Dtilde,rBr(Q))>r3\operatorname{dist}(\tilde{D},\partial^{r}B_{r}(Q))>r^{3}
  2. (b)

    Otherwise, if t((Qr)3N)t(x)t({}^{-3N}(Q_{r}))\leq t(x) then the intersection (D)rNmrAr(Q){}_{r}^{Nm_{r}}(D)\cap A^{\prime}_{r}(Q) contains a disk Dtilde\tilde{D} through Br(Q)B_{r}(Q) to the right of WW such that

    dist(Dtilde,W)>r3\operatorname{dist}(\tilde{D},W)>r^{3}

Thus there are neighborhoods U+U_{+} of rBr(Q)\partial^{r}B_{r}(Q) and UU of WW such that D𝑡𝑖𝑙𝑑𝑒\tilde{D} is disjoint from either UU or U+U_{+}.

Proof.

We proceed in several steps. First, we address (a) which is easy. Second, we

Step 1. We prove (a) first. Note that the tt-coordinate of (x)rN{}_{r}^{N}(x) satisfies

0<t((x)rN)=eNrt(x)e2Nrt(Qr)0<t({}^{N}_{r}(x))=e^{Nr}\cdot t(x)\leq e^{-2Nr}\cdot t(Q_{r})

Take the disk Dtilde\tilde{D} to be the connected component of (D)rN{}_{r}^{N}(D) containing (x)rN{}_{r}^{N}(x). Then Dtilde\tilde{D} is a vertical disk to the right of WW since it intersects WW^{\prime} at a point with positive tt-coordinate. By Lemma 3.21, it is the graph of a 2ϵ2\epsilon-Lipschitz map

f:Dun(lμmr)Dsn(2)×Dt1(t(Qr))f:D^{n}_{u}(l\cdot\mu^{-m_{r}})\to D^{n}_{s}(2)\times D^{1}_{t}(t(Q_{r}))

As in Lemma 5.16, this implies that the tt-coordinate is bounded by

e2Nrt(Qr)+2ϵlμmr<eNrt(Qr)for small re^{-2Nr}\cdot t(Q_{r})+2\epsilon\cdot l\cdot\mu^{-m_{r}}<e^{-Nr}\cdot t(Q_{r})\qquad\text{for small }r

The right side rBr(Q)\partial^{r}B_{r}(Q) is the set of points with tt-coordinate t(Qr)t(Q_{r}), by Definition 4.17. Therefore

dist(rBr(Q),Dtilde)>t(Qr)eNrt(Qr)=eNr(1eNr)(1e8Nr)>r3for small r\operatorname{dist}(\partial^{r}B_{r}(Q),\tilde{D})>t(Q_{r})-e^{-Nr}\cdot t(Q_{r})=e^{-Nr}\cdot(1-e^{-Nr})\cdot(1-e^{-8Nr})>r^{3}\qquad\text{for small }r

Step 2. In this step, we consider a useful special case of (b). Let CrWC^{\prime}_{r}\subset W^{\prime} denote the set

Cr=Dsn(2+μmr/2)×[t((Qr)r4N),t((Qr)rN)]t×0uC^{\prime}_{r}=D^{n}_{s}(2+\mu^{-m_{r}/2})\times[t({}^{-4N}_{r}(Q_{r})),t({}^{N}_{r}(Q_{r}))]_{t}\times 0_{u}

Given a point wCrw\in C^{\prime}_{r}, we let DwD_{w} denote the following vertical disk

Dw={(s,t,u)Fu(,rw):|u|lμmr}D_{w}=\big{\{}(s,t,u)\in F^{u}({}_{r},w)\;:\;|u|\leq l\cdot\mu^{-m_{r}}\big{\}}

Note that by Lemmas 4.10 and 4.9, we have

(Cr)rNmrDsn(2μmr)×[t((Qr)rN(mr4)),t((Qr)rN(mr+1))]{}^{Nm_{r}}_{r}(C^{\prime}_{r})\subset D^{n}_{s}(2\cdot\mu^{-m_{r}})\times[t({}^{N(m_{r}-4)}_{r}(Q_{r})),t({}^{N(m_{r}+1)}_{r}(Q_{r}))]

By the construction of mrm_{r} (see Notation 4.15) we know that

t((Qr)rN(mr4))t((Pr)rN)andt((Qr)rN(mr+1))t((Pr)r7N)t({}_{r}^{N(m_{r}-4)}(Q_{r}))\geq t({}_{r}^{N}(P_{r}))\qquad\text{and}\qquad t({}_{r}^{N(m_{r}+1)}(Q_{r}))\geq t({}_{r}^{7N}(P_{r}))

Therefore we have the following inclusion

(Cr)rNmrDsn(2μmr)×[t((Pr)rN),t((Pr)r7)]{}^{Nm_{r}}_{r}(C^{\prime}_{r})\subset D^{n}_{s}(2\cdot\mu^{-m_{r}})\times[t({}^{N}_{r}(P_{r})),t({}^{7}_{r}(P_{r}))]

For sufficiently small rr, Construction 5.1 yields a well-defined holonomy map

Holr:Dsn(2μmr)×[t((Pr)rN),t((Pr)r7N)]Dsn(3)×[δ,L+δ]t×0u\operatorname{Hol}_{{}_{r}}:D^{n}_{s}(2\cdot\mu^{-m_{r}})\times[t({}^{N}_{r}(P_{r})),t({}^{7N}_{r}(P_{r}))]\to D^{n}_{s}(3)\times[-\delta,L+\delta]_{t}\times 0_{u}

Thus the point z=(w)rNmrz={}^{Nm_{r}}_{r}(w) has well-defined holononomy. By Lemma 5.6, we have

|Holr(z)(r+t(z)L)|μκmr|\operatorname{Hol}_{{}_{r}}(z)-(r+t(z)-L)|\leq\mu^{-\kappa m_{r}}

In particular, this implies that for small rr, we have the following inequality.

(5.17) t(Holr(z))r+t((Pr)rN)Lμκmr=r(1eNr)μκmr>12r2t(\operatorname{Hol}_{{}_{r}}(z))\geq r+t({}^{N}_{r}(P_{r}))-L-\mu^{-\kappa m_{r}}=r(1-e^{-Nr})-\mu^{-\kappa m_{r}}>\frac{1}{2}r^{2}

Moreover, Holr(z)\operatorname{Hol}_{{}_{r}}(z) lies on the unstable disk (Dw)Nmr{}^{Nm_{r}}(D_{w}) through (w)rNmr{}_{r}^{Nm_{r}}(w), since it contains all points in the unstable leaf of distance less than ll from (w)rNmr{}_{r}^{Nm_{r}}(w). We thus acquire a point

xw=(Holr(z))rNmrwith(xw)rNmrAtilder(Q){u=0} and t((xw)rNmr)>12r2x_{w}={}_{r}^{-Nm_{r}}(\operatorname{Hol}_{{}_{r}}(z))\quad\text{with}\quad{}^{Nm_{r}}_{r}(x_{w})\in\tilde{A}_{r}(Q)\cap\{u=0\}\text{ and }t({}^{Nm_{r}}_{r}(x_{w}))>\frac{1}{2}r^{2}

Note that xwx_{w} varies continuously with ww.

Step 3. In this step we discuss the general case of (b). Fix a vertical disk DBr(Q)D\subset B_{r}(Q) as in (b), with t((Qr)3N)<t(x)t({}^{-3N}(Q_{r}))<t(x) where x=DWx=D\cap W^{\prime}. Let

π:U[3,3]sn×[ϵ,L+ϵ]t×0u\pi:U\to[-3,3]^{n}_{s}\times[-\epsilon,L+\epsilon]_{t}\times 0_{u}

be projection to the (s,t)(s,t)-plane and let W\Sigma\subset W^{\prime} be the union of all projections π(Dw)\pi(D_{w}) where DwD_{w} intersects π(D)\pi(D). The disks DD and DwD_{w} are 2ϵ2\epsilon-Lipschitz graphs over Dun(lμmr)D^{n}_{u}(l\cdot\mu^{-m_{r}}) (Lemma 3.21). Therefore these projections are all contained in balls of radius 2ϵlμmr2\epsilon\cdot l\cdot\mu^{-m_{r}} in WW^{\prime} (see also Step 2). It follows that there is a C>0C>0 independent of rr such that

dist(x,y)Cμmrfor all y\operatorname{dist}(x,y)\leq C\cdot\mu^{-m_{r}}\qquad\text{for all }y\in\Sigma

This implies that Cr\Sigma\subset C^{\prime}_{r} for sufficiently small rr, and therefore that

DVwhereV={zDw:wCr}D\subset V\qquad\text{where}\qquad V=\{z\in D_{w}\;:\;w\in C^{\prime}_{r}\}

since contains the (s,t)(s,t)-ball of radius CμmrC\cdot\mu^{-m_{r}} around xx for small rr. We let π:VCr\pi^{\prime}:V\to C^{\prime}_{r} be the obvious projection mapping zDwz\in D_{w} to ww.

By Step 2, the projection π:VCr\pi^{\prime}:V\to C^{\prime}_{r} has a natural continuous section

σ:CrVgiven byσ(w)=xw\sigma:C^{\prime}_{r}\to V\qquad\text{given by}\qquad\sigma(w)=x_{w}

Since DD is vertical, it must necessarily intersect one point xDx\in D in the image of σ\sigma. By Step 2

t((x)Nmr)>12r2and(x)NmrAtilder(Q)Wt({}^{Nm_{r}}(x))>\frac{1}{2}r^{2}\qquad\text{and}\qquad{}^{Nm_{r}}(x)\in\tilde{A}_{r}(Q)\cap W^{\prime}

Note that Atilder(Q)Br(Q)\tilde{A}_{r}(Q)\subset B_{r}(Q) for small rr. We let Dtilde\tilde{D} be the component of (D)rNmrBr(Q){}_{r}^{Nm_{r}}(D)\cap B_{r}(Q) containing xx. This is a vertical disk containing xx, so by the usual considerations we have

dist(x,z)<Cμmrfor all zDtilde\operatorname{dist}(x,z)<C\cdot\mu^{-m_{r}}\qquad\text{for all }z\in\tilde{D}

It follows that for sufficiently small rr, we have the lower bound

dist(Dtilde,W)dist(x,W)Cμmr=t(x)Cμmr>12r2Cμmr>r3\operatorname{dist}(\tilde{D},W)\geq\operatorname{dist}(x,W)-C\mu^{-m_{r}}=t(x)-C\mu^{-m_{r}}>\frac{1}{2}r^{2}-C\mu^{-m_{r}}>r^{3}

This constructs the required disk and concludes the proof.∎

5.5. Proof Of Theorem 4.2

We are finally ready to conclude the proof of Theorem 4.2. We need a final lemma, demonstrating Theorem 4.2(c).

Lemma 5.18.

The intersections Wu(P,)rBr(Q)W^{u}(P,{}_{r})\cap B_{r}(Q) contains a vertical disk DQD_{Q} to the right of Ws(Q,)rW^{s}(Q,{}_{r}).

Proof.

We demonstrate this for Br(Q)B_{r}(Q). By Lemma 4.14, Wu(P,)rW^{u}(P,{}_{r}) contains a disk DrUD_{r}\subset U in Fu()rF^{u}({}_{r}) centered at the homoclinic point

br=(1s,r,0u)ofPb_{r}=(1_{s},r,0_{u})\quad\text{of}\quad P

Let cr=(0s,r,0u)c_{r}=(0_{s},r,0_{u}) be the intersection point Fs(,rbr)F^{s}({}_{r},b_{r})\cap\Gamma and let DrD^{\prime}_{r} be a disk in Fu()rUF^{u}({}_{r})\cap U centered at crc_{r}. Note that we may take the radius of DrD^{\prime}_{r} and DrD_{r} to be lowerbounded by A>0A>0 independent of rr. Let lrl_{r} denote the unique integer such that

(cr)rlr((Pr)r,(Pr)r2]or equivalentlyt((cr)rlr)(Lrer,Lre2r]{}_{r}^{l_{r}}(c_{r})\in({}_{r}(P_{r}),{}_{r}^{2}(P_{r})]\quad\text{or equivalently}\quad t({}_{r}^{l_{r}}(c_{r}))\in(L-re^{-r},L-re^{-2r}]

As in Lemma 4.16, we know that lr>1/rl_{r}>1/r if rr is small. Now note that Nr{}_{r}^{N} uniformly expands Fu()rF^{u}({}_{r}) with constant of dilation (greater than) μ\mu, for small rr. Therefore (Dr)rlr{}_{r}^{l_{r}}(D^{\prime}_{r}) contains the disk in Fu()rF^{u}({}_{r}) of radius greater than 2l2\cdot l around (cr)rlr{}^{l_{r}}_{r}(c_{r}) for small rr. In particular, by Lemma 4.14 and the definition of ll (Construction 4.17), there is a sub-disk

Dr′′(Dr)lrwith a point x=(1s,t(x),0u) with t(x)[r(1er),r(1e2r)]D^{\prime\prime}_{r}\subset{}^{l_{r}}(D^{\prime}_{r})\qquad\text{with a point $x=(1_{s},t(x),0_{u})$ with $t(x)\in[r(1-e^{-r}),r(1-e^{-2r})]$}

Thus the disk is to the right of Wlocs(Q,)rBr(Q)W^{s}_{\operatorname{loc}}(Q,{}_{r})\cap B_{r}(Q). The disk must also be contained in a unstable disk fiber of Btilder(Q)\tilde{B}_{r}(Q) (see Construction 4.17). Therefore every point in Dr′′D^{\prime\prime}_{r} is within distance 2ϵlμmr2\epsilon\cdot l\cdot\mu^{-m_{r}} of xx and so

t(y)t(x)+2ϵlμmrr(1e3r)4r2for every yDr′′ and small rt(y)\leq t(x)+2\epsilon\cdot l\cdot\mu^{-m_{r}}\leq r(1-e^{-3r})\leq 4r^{2}\qquad\text{for every $y\in D^{\prime\prime}_{r}$ and small $r$}

Similarly, t(y)r2t(y)\geq r^{2}. Finally, since brb_{r} and crc_{r} are on the same stable disk in UU, we have

dist((cr)rmr,(br)rmr)μmrμ1/r\operatorname{dist}({}^{m_{r}}_{r}(c_{r}),{}^{m_{r}}_{r}(b_{r}))\leq\mu^{-m_{r}}\leq\mu^{-1/r}

By taking a path γ\gamma from (cr)rmr{}^{m_{r}}_{r}(c_{r}) to xx in Fu()rF^{u}({}_{r}) and using the holonomy map of FuF^{u} (see Construction 5.1) we get a small unstable disk DQD_{Q} in Fu()rF^{u}({}_{r}) centered at a point xQ=Holr(x)x_{Q}=\operatorname{Hol}_{{}_{r}}(x) with

dist(xQ,x)μκ/r\operatorname{dist}(x_{Q},x)\leq\mu^{-\kappa/r}

Here β\beta is the uniform Hölder constants in Construction 5.1. It follows, as with D′′D^{\prime\prime}, that DQD_{Q} is a vertical disk to the right of Wlocs(Q,)rBr(Q)W^{s}_{\operatorname{loc}}(Q,{}_{r})\cap B_{r}(Q).∎

Theorem 4.2 is an immediate consequence of Lemma 4.12 for Theorem 4.2(a), the Lemmas 5.3, 5.9, 5.10, 5.15, 5.16 and 5.17 for Theorem 4.2(b) and the Lemma 5.18 for Theorem 4.2(c).

6. Robustly Mixing Contactomorphisms

In the final section of this paper, we prove Theorem 11. The proof is a small modification of the proof of Theorem A of [7] in [7, Section 4.C].

Theorem 6.1 (Theorem 11).

Let (Y,ξ)(Y,\xi) be a closed contact manifold admitting an Anosov Reeb flow with CC^{\infty} stable and unstable foliations. Then the C1C^{1}-open set of robustly mixing contactomorphisms

ContRM(Y,ξ)Cont(Y,ξ)\operatorname{Cont}_{\operatorname{RM}}(Y,\xi)\subset\operatorname{Cont}(Y,\xi)

is non-empty. Moreover, if TT is the period of a closed Reeb orbit of , then T is in the closure of this set.

Proof.

Since is Reeb Anosov, it must be a transitive by the Plante alternative [32, Thm 8.1.3 and 8.1.4]. A transitive Anosov flow must also have a dense set of closed orbits [32, Thm 6.2.10].

Fix a closed orbit of period TT and an open neighborhood UU of . Also pick an auxilliary orbit with a neighborhood VV. Let α\alpha denote the contact form with Reeb vector-field generating the Reeb flow and choose a 11-parameter family of contact forms αs\alpha_{s} on YY such that

α0=ααs|U=αandαs|V=(1+s)α\alpha_{0}=\alpha\qquad\alpha_{s}|_{U}=\alpha\quad\text{and}\quad\alpha_{s}|_{V}=(1+s)\cdot\alpha

Let s denote the Reeb flow of αs\alpha_{s}. Anosov flows are C1C^{1}-structurally stable ([1] or [32, Thm 5.4.22]), and thus s is a smooth Anosov Reeb flow for sufficiently small ss. Moreover, and are orbits of s for all ss.

Thus, choose an ss such that s is Reeb Anosov and such that has a period that is not a multiple of TT with respect to s. We let be the time TT map of s and note that

  • is a strict, partially hyperbolic contactomorphism with stable and unstable bundle equal to those of s and central bundle given by the span of the Reeb vector-field of αs\alpha_{s}.

  • is a closed Reeb orbit that is a normally hyperbolic fixed set of .

  • The stable bundle Es()E^{s}(\Phi) of T is a smooth, integrable and uniformly contracted by on UU since agrees with T on UU.

Note also the local stable manifolds of and for the set (or for any point PP\in\Gamma) agree.

Wlocs(,;U)=Wlocs(,;TU)andWlocs(P,;U)=Wlocs(P,;TU)W^{s}_{\operatorname{loc}}(\Gamma,\Psi;U)=W^{s}_{\operatorname{loc}}(\Gamma,{}_{T};U)\qquad\text{and}\qquad W^{s}_{\operatorname{loc}}(P,\Psi;U)=W^{s}_{\operatorname{loc}}(P,{}_{T};U)

Since the points PP of are fixed, the stable manifold of PP is equal to the stable leaf of PP with respect to both and T. It follows that

Fs(,P)=Fs(,P) on Wlocs(,;U)F^{s}(\Psi,P)=F^{s}(\Phi,P)\qquad\text{ on }W_{\operatorname{loc}}^{s}(\Gamma,\Psi;U)

This verifies the criteria in Theorem 4.2(a-c). To check Theorem 4.2(d) we apply the following lemma of Bonatti-Diaz.

Lemma 6.2.

[7, Lemma 4.3] Let be a partially hyprbolic map and , be closed orbits of different period, as constructed above. Then there are a pair of points P,QP,Q\in\Gamma and a heteroclinic orbit from PP to QQ.

The argument now proceeds identically to [7, p. 395] and we recall it here. Apply Theorem 4.2 to acquire a 11-parameter family of contactomorphisms r and an N>0N>0 such that

  • r has hyperbolic fixed points PP and QQ on of index n+1n+1 and nn, respectively.

  • There is a neighborhood Br(Q)B_{r}(Q) of QQ such that (Br(Q),)rN(B_{r}(Q),{}_{r}^{N}) is a stable blender.

  • The intersection Wu(P,)rBr(Q)W^{u}(P,{}_{r})\cap B_{r}(Q) contains a vertical disk DQD_{Q} to the right of Ws(Q,)rW^{s}(Q,{}_{r}).

This implies that the tuple ((Br(Q),)rN,P)((B_{r}(Q),{}_{r}^{N}),P) is a chain of blenders in the sense of Bonatti-Diaz (see [7, p. 369] or [7, §7.1]). Lemma 1.12 of [7] now states that there is a C1C^{1}-neighborhood 𝒰\mathcal{U} of r such that, for any 𝒰{}^{\prime}\in\mathcal{U}, there are fixed points PP^{\prime} and QQ^{\prime} (the continuations of PP and QQ, which are non-degenerate hyperbolic fixed points and thus persist in a C1C^{1}-nieghborhood) such that

(6.1) Ws(P,)close(Ws(Q,))W^{s}(P^{\prime},{}^{\prime})\subset\operatorname{close}(W^{s}(Q^{\prime},{}^{\prime}))

Here close()\operatorname{close}(-) denotes the topological closure.

Next, since s is Anosov and transitive, the time TT map is partially hyperbolic with well-defined center-stable and center-unstable foliations

Fcs() tangent to span(R)Es()andFcu() tangent to span(R)Eu()F^{cs}(\Psi)\text{ tangent to }\operatorname{span}(R)\oplus E^{s}(\Psi)\qquad\text{and}\qquad F^{cu}(\Psi)\text{ tangent to }\operatorname{span}(R)\oplus E^{u}(\Psi)

Moreover, the center-stable and center-unstable foliations have dense leaves [32, Thm 6.2.10]. By Hirsch-Pugh-Robinson [39], these properties are C1C^{1}-robust. Thus is partially hyperbolic with invariant foliations

Fcs() tangent to Ec()Es()andFcu() tangent to Ec()Eu()F^{cs}({}^{\prime})\text{ tangent to }E^{c}({}^{\prime})\oplus E^{s}({}^{\prime})\qquad\text{and}\qquad F^{cu}({}^{\prime})\text{ tangent to }E^{c}({}^{\prime})\oplus E^{u}({}^{\prime})

with dense leaves. Now note that we have the following identifications.

Fcs(,P)=Ws(P,)andFcu(,Q)=Wu(Q,)F^{cs}({}^{\prime},P^{\prime})=W^{s}(P^{\prime},{}^{\prime})\qquad\text{and}\qquad F^{cu}({}^{\prime},Q^{\prime})=W^{u}(Q^{\prime},{}^{\prime})

Indeed, the uniqueness of local invariant manifolds near hyperbolic invariant sets [39, Thm 4.1(b)] implies equality near PP^{\prime} and QQ^{\prime}, and then global invariance implies global equality. In particular, Ws(P,)W^{s}(P^{\prime},{}^{\prime}) and Wu(Q,)W^{u}(Q^{\prime},{}^{\prime}) are dense. Moreover, by (6.1) Ws(Q,)W^{s}(Q^{\prime},{}^{\prime}) is also dense.

Now [10, Lem 7.3] states that is robustly transitive (and in fact mixing). We recall the argument. Let UU and VV be neighborhoods in YY. Then

UWs(Q,)andVWu(Q,)U\cap W^{s}(Q^{\prime},{}^{\prime})\neq\emptyset\qquad\text{and}\qquad V\cap W^{u}(Q^{\prime},{}^{\prime})\neq\emptyset

since these invariant manifolds are dense. This implies that ()j(U)()k(V)({}^{\prime})^{j}(U)\cap({}^{\prime})^{-k}(V) is non-empty for all sufficiently large jj and kk. In other words, is topologically mixing. ∎

References

  • [1] Dmitry Victorovich Anosov. Ergodic properties of geodesic flows on closed riemannian manifolds of negative curvature. In Doklady Akademii Nauk, volume 151, pages 1250–1252. Russian Academy of Sciences, 1963.
  • [2] Russell Avdek. An algebraic generalization of giroux’s criterion. arXiv preprint arXiv:2307.09068, 2023.
  • [3] Russell Avdek and Zhengyi Zhou. Bourgeois’ contact manifolds are tight. arXiv preprint arXiv:2404.16311, 2024.
  • [4] John A Baldwin and Steven Sivek. A contact invariant in sutured monopole homology. In Forum of Mathematics, Sigma, volume 4, page e12. Cambridge University Press, 2016.
  • [5] Mélanie Bertelson and Gael Meigniez. Conformal symplectic structures, foliations and contact structures. arXiv preprint arXiv:2107.08839, 2021.
  • [6] Christian Bonatti, Sylvain Crovisier, Lorenzo Diaz, and Amie Wilkinson. What is… a blender? arXiv preprint arXiv:1608.02848, 2016.
  • [7] Christian Bonatti and Lorenzo J. Diaz. Persistent nonhyperbolic transitive diffeomorphisms. Annals of Mathematics, 143(2):357–396, 1996.
  • [8] Christian Bonatti and Lorenzo J Díaz. Robust heterodimensional cycles and-generic dynamics. Journal of the Institute of Mathematics of Jussieu, 7(3):469–525, 2008.
  • [9] Christian Bonatti, Lorenzo J Díaz, and Enrique R Pujals. A c1-generic dichotomy for diffeomorphisms: weak forms of hyperbolicity or infinitely many sinks or sources. Annals of Mathematics, pages 355–418, 2003.
  • [10] Christian Bonatti, Lorenzo J Díaz, and Marcelo Viana. Dynamics beyond uniform hyperbolicity: A global geometric and probabilistic perspective, volume 3. Springer Science & Business Media, 2004.
  • [11] Armand Borel. Compact clifford-klein forms of symmetric spaces. Topology, 2(1-2):111–122, 1963.
  • [12] Joseph Breen. Morse-smale characteristic foliations and convexity in contact manifolds. Proceedings of the American Mathematical Society, 149(9):3977–3989, 2021.
  • [13] Joseph Breen. Folded symplectic forms in contact topology. Journal of Geometry and Physics, 201:105213, 2024.
  • [14] Joseph Breen, Ko Honda, and Yang Huang. The giroux correspondence in arbitrary dimensions. arXiv preprint arXiv:2307.02317, 2023.
  • [15] Dmitri Burago and Sergei Ivanov. Partially hyperbolic diffeomorphisms of 3-manifolds with abelian fundamental groups. J. Mod. Dyn, 2(4):541–580, 2008.
  • [16] A. Candel and L. Conlon. Foliations. Number v. 2 in Foliations. Mathematical Society of Japan, 1985.
  • [17] Robert Cardona. Stability is not open or generic in symplectic four-manifolds. arXiv preprint arXiv:2305.13158, 2023.
  • [18] Julian Chaidez and Shira Tanny. Elementary sft spectral gaps and the strong closing property. arXiv preprint arXiv:2312.17211, 2023.
  • [19] Rima Chatterjee, John B Etnyre, Hyunki Min, and Anubhav Mukherjee. Existence and construction of non-loose knots. arXiv preprint arXiv:2310.04908, 2023.
  • [20] Kai Cieliebak and Evgeny Volkov. First steps in stable hamiltonian topology. Journal of the European Mathematical Society, 17(2):321–404, 2015.
  • [21] Vincent Colin, Paolo Ghiggini, Ko Honda, and Michael Hutchings. Sutures and contact homology i. Geometry & Topology, 15(3):1749–1842, 2011.
  • [22] Vincent Colin, Emmanuel Giroux, and Ko Honda. On the coarse classification of tight contact structures. arXiv preprint math/0305186, 2003.
  • [23] Vincent Colin, Emmanuel Giroux, and Ko Honda. Finitude homotopique et isotopique des structures de contact tendues. Publications mathématiques, 109:245–293, 2009.
  • [24] James Conway and Hyunki Min. Classification of tight contact structures on surgeries on the figure-eight knot. Geometry & Topology, 24(3):1457–1517, 2020.
  • [25] Sylvain Crovisier and Rafael Potrie. Introduction to partially hyperbolic dynamics. School on Dynamical Systems, ICTP, Trieste, 3(1), 2015.
  • [26] Yakov Eliashberg and Dishant Pancholi. Honda-huang’s work on contact convexity revisited. arXiv preprint arXiv:2207.07185, 2022.
  • [27] John B Etnyre and Ko Honda. On the nonexistence of tight contact structures. Annals of Mathematics, pages 749–766, 2001.
  • [28] John B Etnyre and Ko Honda. Tight contact structures with no symplectic fillings. Inventiones Mathematicae, 148(3):609, 2002.
  • [29] John B Etnyre, Hyunki Min, and Anubhav Mukherjee. Non-loose torus knots. arXiv preprint arXiv:2206.14848, 2022.
  • [30] François-Simon Fauteux-Chapleau and Joseph Helfer. Exotic tight contact structures on n. arXiv preprint arXiv:2111.07590, 2021.
  • [31] Eduardo Fernández and Hyunki Min. Spaces of legendrian cables and seifert fibered links. arXiv preprint arXiv:2310.12385, 2023.
  • [32] Todd Fisher and Boris Hasselblatt. Hyperbolic flows. European Mathematical Society, 2019.
  • [33] Hansjörg Geiges. An introduction to contact topology, volume 109. Cambridge University Press, 2008.
  • [34] Paolo Ghiggini, Paolo Lisca, András I Stipsicz, et al. Classification of tight contact structures on small seifert 3-manifolds with e00e_{0}\geq 0. In Proc. Amer. Math. Soc, volume 134, pages 909–916, 2006.
  • [35] Emmanuel Giroux. Convexity in contact topology. Comment. Math. Helv, 66(4):637–677, 1991.
  • [36] Michael Gromov and Ilya Piatetski-Shapiro. Non-arithmetic groups in lobachevsky spaces. Publications Mathématiques de l’IHÉS, 66:93–103, 1987.
  • [37] Victor Guillemin, Ana Cannas da Silva, and Christopher Woodward. On the unfolding of folded symplectic structures. Mathematical Research Letters, 7(1):35–53, 2000.
  • [38] F Rodriguez Hertz, J Rodriguez Hertz, and Raúl Ures. Partially hyperbolic dynamics. Publicaçoes Matemáticas do IMPA, 2011.
  • [39] M.W. Hirsch, C.C. Pugh, and M. Shub. Invariant Manifolds. Lecture Notes in Mathematics. Springer Berlin Heidelberg, 2006.
  • [40] Ko Honda. On the classification of tight contact structures i. arXiv preprint math/9910127, 1999.
  • [41] Ko Honda. On the classification of tight contact structures ii. Journal of Differential Geometry, 55(1):83–143, 2000.
  • [42] Ko Honda and Yang Huang. Bypass attachments in higher-dimensional contact topology. arXiv preprint arXiv:1803.09142, 2018.
  • [43] Ko Honda and Yang Huang. Convex hypersurface theory in contact topology. arXiv preprint arXiv:1907.06025, 2019.
  • [44] Ko Honda, William H Kazez, and Gordana Matić. Convex decomposition theory. International Mathematics Research Notices, 2002(2):55–88, 2002.
  • [45] Ko Honda, William H Kazez, and Gordana Matić. Tight contact structures on fibred. Journal of Differential Geometry, 64(2):305–358, 2003.
  • [46] Ko Honda, William H Kazez, and Gordana Matic. Contact structures, sutured floer homology and tqft. arXiv preprint arXiv:0807.2431, 2008.
  • [47] Dongchen Li. Blender-producing mechanisms and a dichotomy for local dynamics for heterodimensional cycles. arXiv preprint arXiv:2403.15818, 2024.
  • [48] John J Millson. On the first betti number of a constant negatively curved manifold. Annals of Mathematics, 104(2):235–247, 1976.
  • [49] Hyunki Min and Isacco Nonino. Tight contact structures on a family of hyperbolic l-spaces. arXiv preprint arXiv:2311.14103, 2023.
  • [50] Atsuhide Mori. Reeb foliations on s^ 5 and contact 5-manifolds violating the thurston-bennequin inequality. arXiv preprint arXiv:0906.3237, 2009.
  • [51] Atsuhide Mori. On the violation of thurston-bennequin inequality for a certain non-convex hypersurface. arXiv preprint arXiv:1111.0383, 2011.
  • [52] Emmy Murphy. Loose Legendrian embeddings in high dimensional contact manifolds. Stanford University, 2012.
  • [53] Meysam Nassiri and Enrique R Pujals. Robust transitivity in hamiltonian dynamics. In Annales scientifiques de l’École normale supérieure, volume 45, pages 191–239, 2012.
  • [54] Workshop Participants. Aim problem list in higher dimensional contact topology, 2024.
  • [55] Charles Pugh, Michael Shub, and Amie Wilkinson. Hölder foliations. Duke Math. J., 90(1):517–546, 1997.
  • [56] Dennis Sullivan. Hyperbolic geometry and homeomorphisms. In Geometric topology, pages 543–555. Elsevier, 1979.
  • [57] Ichiro Torisu. Convex contact structures and fibered links in 3-manifolds. International Mathematics Research Notices, 2000(9):441–454, 2000.
  • [58] Alan Weinstein. Symplectic manifolds and their lagrangian submanifolds. Advances in Mathematics, 6(3):329–346, 1971.
  • [59] Chris Wendl. Lectures on symplectic field theory. arXiv preprint arXiv:1612.01009, 2016.
  • [60] Kanda Yutaka. The classification of tight contact structures on the 3-torus. Communications in Analysis and Geometry, 5(3):413–438, 1997.