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

Homoclinic orbits for geodesic flows of surfaces

Gonzalo Contreras Gonzalo Contreras
Centro de Investigación en Matemáticas
A.P. 402, 36.000, Guanajuato, GTO, Mexico
 and  Fernando Oliveira Fernando Oliveira
Universidade Federal de Minas Gerais
Av. Antônio Carlos 6627, 31270-901, Belo Horizonte, MG, Brasil.
Abstract.

We prove that the geodesic flow of a Kupka-Smale riemannian metric on a closed surface has homoclinic orbits for all of its hyperbolic closed geodesics.

2020 Mathematics Subject Classification:
37D40, 53D25, 37C29
Gonzalo Contreras is partially supported by CONACYT, Mexico, grant A1-S-10145.

1. Introduction.

Let (M,ρ)(M,\rho) be a closed (i.e. compact, boundaryless) riemannian surface. Let SM={(x,v):ρ(v,v)=1}SM=\{(x,v):\rho(v,v)=1\} be its unit tangent bundle with projection π:SMM\pi:SM\to M, π(x,v)=x\pi(x,v)=x. The geodesic flow ϕt:SMSM\phi_{t}:SM\to SM of (M,ρ)(M,\rho) is defined by ϕt(x,v)=(γ(t),γ˙(t))\phi_{t}(x,v)=(\gamma(t),{\dot{\gamma}}(t)), where γ\gamma is the unit speed geodesic with (γ(0),γ˙(0))=(x,v)(\gamma(0),{\dot{\gamma}}(0))=(x,v).

A closed orbit for ϕt\phi_{t} is hyperbolic if its Floquet multipliers do not have modulus 1. The (strong) stable and unstable manifolds of a point zSMz\in SM are

Ws,u(z):={wSM:limt±d(ϕt(w),ϕt(z))=0},W^{s,u}(z):=\{w\in SM:\lim\nolimits_{t\to\pm\infty}d(\phi_{t}(w),\phi_{t}(z))=0\},

respectively. For a subset ASMA\subset SM define

Ws,u(A)=aAWs,u(a).W^{s,u}(A)=\textstyle\bigcup_{a\in A}W^{s,u}(a).

For a hyperbolic closed geodesic γ\gamma the sets Ws(γ˙)W^{s}({\dot{\gamma}}), Wu(γ˙)W^{u}({\dot{\gamma}}) are immersed submanifolds of SMSM either diffeomorphic to a cylinder with one boundary γ˙{\dot{\gamma}} or to a Möbius band where γ˙{\dot{\gamma}} is its equator, according to wether the Floquet multipliers of γ˙{\dot{\gamma}} are positive or negative respectively. A homoclinic point of a hyperbolic closed geodesic γ\gamma is a point in (Ws(γ˙)Wu(γ˙))γ˙(W^{s}({\dot{\gamma}})\cap W^{u}({\dot{\gamma}}))\setminus{\dot{\gamma}}. A heteroclinic point is a point in (Ws(γ˙)γ˙)(Wu(η˙)η˙)(W^{s}({\dot{\gamma}})\setminus{\dot{\gamma}})\cap(W^{u}({\dot{\eta}})\setminus{\dot{\eta}}), where γ˙{\dot{\gamma}}, η˙{\dot{\eta}} are two hyperbolic closed orbits of ϕt\phi_{t}.

Homoclinic points where first discovered by Henri Poincaré in 1889 (cf. Andersson [1]) and named in Poincaré [34, §395]. It is well known the paragraph of Poincaré [34, vol. III, §397][1, §5] describing his admiration of the complexity of the dynamics implied by the existence of a transversal homoclinic point.

We say that the riemannian metric ρ\rho is Kupka-Smale if

  1. (i)

    The Floquet multipliers of every periodic orbit are not roots of unity.

  2. (ii)

    The heteroclinic intersections of hyperbolic orbits Ws(γ˙)Wu(η˙)W^{s}({\dot{\gamma}})\pitchfork W^{u}({\dot{\eta}}) are transversal.

For any rr\in{\mathbb{N}}, r2r\geq 2, the set of CrC^{r} riemannian metrics whose geodesic flow is Kupka-Smale is residual in the set of CrC^{r} riemannian metrics in MM, see Contreras, Paternain [10, Thm. 2.5]. Clarke [5] proves that Kupka-Smale metrics are also residual in the CωC^{\omega} topology for analytic hypersurfaces of n{\mathbb{R}}^{n}, n3n\geq 3. Here we prove

Theorem A.

For a Kupka-Smale riemannian metric on a closed surface every hyperbolic closed geodesic γ\gamma has homoclinic orbits in all the components of Ws(γ˙)γ˙W^{s}({\dot{\gamma}})\setminus{\dot{\gamma}} and of Wu(γ˙)γ˙W^{u}({\dot{\gamma}})\setminus{\dot{\gamma}} and satisfy W¯s(γ˙)=W¯u(γ˙)\overline{W}^{s}({\dot{\gamma}})=\overline{W}^{u}({\dot{\gamma}}).

The importance of finding homoclinic orbits is that in any neighborhood of the homoclinic orbit one finds a horseshoe with complicated dynamics. This dynamics can be coded using symbolic dynamics and implies positive (local) topological entropy, infinitely many periodic orbits shadowing the homoclinic, infinitely many homoclinics and exponential growth of periodic orbits in a neighborhood of the homoclinic. Homoclinics prevent integrability [28, §III.6], and can also be used to obtain Birkhoff sections [8]. They are also the basic skeleton for Mather acceleration theorems in Arnold diffusion [27], [4], [12], [18].

Also theorem A may help to prove that the closed orbits for the geodesic flow of surfaces are generically dense in the phase space. A conjecture by Poincaré [33, vol. I, p.82 §36] stated for the three body problem. By now it is only known that their projection to the surface is generically dense, Irie [22].

It is well known that C2C^{2}, r2r\geq 2, generic riemannian surfaces of genus g2g\geq 2 have homoclinic orbits, see e.g. [10]. Contreras and Paternain [10] proved that C2C^{2} generic metrics on SS2\SS^{2} or 2{\mathbb{R}}{\mathbb{P}}^{2} have some orbits with homoclinic orbits. Knieper and Weiss [24] extended this result to the CC^{\infty} topology and Clarke [5] proved it for analytic convex surfaces in 3{\mathbb{R}}^{3} and the CωC^{\omega} topology. Xia and Zhang [35] prove that for a CC^{\infty} generic metric of positive curvature in SS2\SS^{2}, every hyperbolic periodic orbit has homoclinics. Contreras [6] proves that C2C^{2} generic metrics on any closed manifold have homoclinics.

The CC^{\infty}, CωC^{\omega} results [24], [5], [35] in the sphere use the annular Birkhoff section [3, §VI.10, p.180] for the spheres with positive curvature. Then they apply the techniques of Pixton [32] and Oliveira [29] for area preserving maps on surfaces of genus 0 to obtain the homoclinics. These techniques extend to genus 1 but not to higher genus. The problem with riemannian surfaces which are not spheres of positive curvature is that they have Birkhoff sections in the Kupka-Smale case [9], [7] but their genera is not known.

Instead we construct what we call a complete system of surfaces of section of genera g1g\leq 1. With this we complete a program initiated by Birkhoff in [2, §28, p. 281] with formal justifications using the curve shortening flow [19], [16]. But now the Poincaré maps to these surfaces of section are not continuous. They are essentially discontinuous111There are arbitrarily small curves whose image under the Poincaré map have infinite length and large diameter.. And the standard (continuity) arguments of Mather [26] and Oliveira [29] for area preserving homeomorphisms can not be applied. We show how to take advantage of the discontinuities of the Poincaré map to obtain homoclinic orbits for certain closed orbits. For the remaining hyperbolic orbits we develop in [30] and [31] the theories of Mather and Oliveira for partially defined area preserving homeomorphisms so that they can be applied to our situation.

We also remark that in [31] we show that the usual hypothesis of Moser stability for elliptic periodic points in Mather [26] is not needed. We use instead Theorem 1.2.(4) from [31] which allows to use only condition (i) from our Kupka-Smale definition. Nevertheless, as observed by Xia and Zhang [35], Fayad and Krikorian [14] prove that elliptic periodic points are Moser stable if their Floquet multipliers are diophantine, which is a generic condition for geodesic flows by the Bumpy Metric Theorem.

The Kupka-Smale condition has been chosen in order to have a unified approach using the results from [9], [30], [31]. But the transversality condition (ii) can be relaxed to asking (ii) only for periodic orbits of small period, in order to obtain a Birkhoff section [9]; and a no heteroclinic connections222A heteroclinic connection is the case in which two components of Ws(γ˙)γ˙W^{s}({\dot{\gamma}})\setminus{\dot{\gamma}} and Wu(η˙)η˙W^{u}({\dot{\eta}})\setminus{\dot{\eta}} are equal. condition instead of the transversality (ii). Moreover, since the theorems that we use from [31] on homoclinic points are about fixed points; in order to get an homoclinic orbit for an orbit γ˙{\dot{\gamma}} we only need to ask for such generic conditions on periodic orbits of smaller period than γ˙{\dot{\gamma}}. We shall not pursue such refinements here.

For an elliptic periodic orbit with Floquet multipliers σ\sigma satisfying σk1\sigma^{k}\neq 1 for 1k41\leq k\leq 4, its Poincaré map on a local transversal section can be written in Birkhoff normal form as

P(z)=zei(ω+β|z|2)+R(z),P(z)=ze^{i(\omega+\beta|z|^{2})}+R(z),

with ω,β\omega,\beta\in{\mathbb{R}} and R(z)R(z) with zero 4-jet at z=0z=0. The condition β0\beta\neq 0 is residual for 4-jets of PP. By theorem 2.5 in [10] this condition on all elliptic orbits is residual for CrC^{r} riemannian metrics with the CrC^{r} topology, r5r\geq 5. The condition β0\beta\neq 0 implies that the Poincaré map PP is locally a twist map. Kupka-Smale twist maps have hyperbolic minimizing orbits with homoclinics for every rational rotation number in an interval [ω,ω+ε[[\omega,\omega+\varepsilon[ or ]ωε,ω]]\omega-\varepsilon,\omega], depending on the sign of β\beta, see [25], [36], [17]. These periodic orbits accumulate on the fixed point z=0z=0. Therefore for CrC^{r}, r5r\geq 5, generic riemannian metrics on closed surfaces every closed geodesic is accumulated by homoclinic orbits, and the closure of the periodic orbits is the same as the closure of the homoclinic orbits.

The proof of theorem A needs results in dynamics of area preserving maps, Reeb flows and geodesic flows.

A contact 3-manifold is a pair (N,λ)(N,\lambda) where NN is a closed 3-manifold and λ\lambda is a 1-form in NN such that λdλ\lambda\wedge d\lambda is a volume form. The Reeb vector field XX of (N,λ)(N,\lambda) is defined by iXdλ0i_{X}d\lambda\equiv 0 and λ(X)1\lambda(X)\equiv 1. The Reeb flow ψt:NN\psi_{t}:N\to N of (N,λ)(N,\lambda) is the flow of XX. The Liouville form of a riemannian surface (M,ρ)(M,\rho), given by

λ(v)(ξ):=ρ(v,dπ(ξ)),vTM,ξTvTM,\lambda(v)(\xi):=\rho(v,d\pi(\xi)),\qquad v\in TM,\quad\xi\in T_{v}TM,

is a contact form on SMSM. The Reeb flow of (SM,λ)(SM,\lambda) is the geodesic flow of (M,ρ)(M,\rho).

A surface of section for the Reeb flow ψt\psi_{t} is a compact immersed surface with boundary ΣN\Sigma\subset N, whose interior is embedded and transversal to the Reeb vector field and whose boundary is a cover of a finite union of closed orbits of ψt\psi_{t}.

A Birkhoff section is a connected inmersed surface ΣN\Sigma\subset N whose interior is embedded and transversal to the vector field. Its boundary is a cover of finitely many closed orbits and there is >0\ell>0 such that for all zNz\in N, ψ]0,[(z)Σ\psi_{]0,\ell[}(z)\cap\Sigma\neq\emptyset and ψ],0[(z)Σ\psi_{]-\ell,0[}(z)\cap\Sigma\neq\emptyset.

Contreras and Mazzucchelli proved in [9, Thm. A] that every Kupka-Smale Reeb flow on a closed contact 3-manifold (N,λ)(N,\lambda) has a Birkhoff section. The first return map of the interior of a Birkhoff section is a diffeomorphism which preserves the area form dλd\lambda.

In order to use the results in [30], [31] to obtain homoclinic orbits we need to have area preserving maps defined on surfaces of genus 0 or 1. In higher genus, the time one map of an area preserving flow without heteroclinic connections is an example of a Kupka-Smale map without homoclinics.

In general we don’t know the genus of the Birkhoff sections obtained in [9] or [7]. Instead we use a complete system of surfaces of section with genus 0 or 1, (definition 2.4). This is a finite collection of surfaces of section which intersect every orbit and such that the points which do not return to the collection of surfaces are in the stable or unstable manifold of a finite set of hyperbolic closed orbits Kfix{K_{fix}}, called non rotating boundary orbits, which are some of the boundaries of the surfaces of section of the system. The other closed orbits in the boundaries of the sections are called rotating boundary orbits, their union is denoted Krot{K_{rot}}. They have the property that there is a neighborhood of Krot{K_{rot}} where the return times to the system of sections is uniformly bounded.

If γ\gamma is a hyperbolic orbit of a Reeb flow in a 3-manifold we call separatrices the connected components of Ws(γ)γW^{s}(\gamma)\setminus\gamma and of Wu(γ)γW^{u}(\gamma)\setminus\gamma. Since the contact manifold (N,λ)(N,\lambda) is orientable, they separate any small tubular neighborhood UU of γ\gamma into 2 or 4 connected components. The germs of these components obtained by shrinking UU are called sectors of γ\gamma. We say that a separatrix accumulates on a sector if it intersects such sector for any tubular neighborhood UU. A separatrix is adjacent to a sector if both the closure of the sector and the separatrix contain a component of a local invariant manifold Wεs,u(γ)W^{s,u}_{\varepsilon}(\gamma).

Wεs,u(γ)=zγWεs,u(z),Wεs,u(z)={wWs,u(z):t>0d(ψ±t(w),ψ±t(z))ε}.W^{s,u}_{\varepsilon}(\gamma)=\textstyle\bigcup_{z\in\gamma}W^{s,u}_{\varepsilon}(z),\qquad W^{s,u}_{\varepsilon}(z)=\{w\in W^{s,u}(z):\forall t>0\;\;d(\psi_{\pm t}(w),\psi_{\pm t}(z))\leq\varepsilon\,\}.

A Kupka-Smale contact manifold means that its Reeb flow satifies (i) and (ii).

Theorem B.

Let (N,λ)(N,\lambda) be a Kupka-Smale closed contact 3-manifold.

  1. (1)

    For any hyperbolic closed orbit γ\gamma of (N,λ)(N,\lambda), all the connected components of Ws(γ)γW^{s}(\gamma)\setminus\gamma and Wu(γ)γW^{u}(\gamma)\setminus\gamma have the same closure equal to W¯s(γ)=W¯u(γ)\overline{W}^{s}(\gamma)=\overline{W}^{u}(\gamma).

    Moreover, each separatrix of γ\gamma accumulates on both of its adjacent sectors.

  2. (2)

    If (N,λ)(N,\lambda) has a Birkhoff section Σ\Sigma of genus 0 or 1, then every hyperbolic orbit intersecting the interior of Σ\Sigma has homoclinics in all its seperatrices.

    A hyperbolic boundary orbit in Σ\partial\Sigma has homoclinics in all its separatrices provided that Σ\Sigma has genus 0 or if Σ\Sigma has genus 1 and the union of its local separatrices intersect Σ\Sigma in at least 4 curves.

  3. (3)

    Suppose that (N,λ)(N,\lambda) admits a complete system of surfaces of section. Then:

    Every non rotating boundary orbit in Kfix{K_{fix}} has homoclinics in all its separatrices.

    If the system contains a component SS of genus 0 or 11, then every periodic orbit which intersects the interior of SS has homoclinics in all its separatrices.

    A hyperbolic rotating boundary orbit in KrotS{K_{rot}}\cap\partial S has homoclinics in all its separatrices provided that SS has genus 0 or if SS has genus 1 and the union of its local separatrices intersect SS in at least 4 curves.

See also proposition 2.10 which has no genus restriction.

Observe that the condition of four intersections is satisfied if the hyperbolic boundary orbit has positive Floquet multipliers. Because in that case the separatrices divide a tubular neighborhood of the orbit into four sectors and the trace of the Birkhoff section must turn around the four sectors.

Recall that Hofer, Wysocki and Zehnder prove in [21] corollary 1.8, that any non degenerate tight contact form on the 3-sphere SS3\SS^{3} admits a finite energy foliation whose leaves have genus 0. The rigid surfaces of the finite energy foliation form a complete system of surfaces of section. We check in §3.8 that the transversality condition in item (iii) of definition 2.4 holds. Therefore we get

1.1 Corollary.

Any Kupka-Smale tight contact form on SS3\SS^{3} has homoclinic orbits in all branches of all of its hyperbolic closed orbits.

Since a homoclinic orbit implies the existence of a horseshoe we also obtain

1.2 Corollary.

If a Kupka-Smale tight contact form on SS3\SS^{3} contains a hyperbolic periodic orbit then it has infinitely many periodic orbits.

The geodesic flow is the Reeb flow of the Liouville form in the unit tangent bundle. By lifting the geodesic flow to a double covering if necessary, in order to obtain homoclinic orbits for geodesic flows it is enough to consider orientable surfaces. Theorem A follows from theorem B and the following theorem 1.3 once the conditions on the rotating boundary orbits in Krot{K_{rot}} in item (3) of theorem B are checked.

1.3 Theorem (Contreras, Knieper, Mazzucchelli, Schulz [7, Thm. E]).

Let (M,ρ)(M,\rho) be a closed connected orientable surface all of whose simple contractible closed orbits without conjugate points are non degenerate. Then there is a complete system of surfaces of section for the geodesic flow of (M,ρ)(M,\rho) whose components have genus 0 or 1.

The ideas in theorem 1.3 date back to Birkhoff [2] section 28, together with the modern version of the curve shortening lemma by Grayson [19]. We also provide a proof theorem 1.3 in section §4, theorem 4.8 with a different construction. And we check the conditions on the rotating boundary orbits in item (3) of theorem B. In our case the system has two embedded surfaces of section of genus 1 and finitely many Birkhoff annuli of disjoint simple closed geodesics. This proves theorem A.

For area preserving maps the auto accumulation of invariant manifolds as in item (1) of theorem B usually requires the Kupka-Smale condition and also the condition that elliptic periodic orbits are Moser stable. This is a fundamental step to obtain homoclinics. Instead, using our results in [31], we only use the non-degeneracy condition (i) from our Kupka-Smale definition. In our application the first return map to the complete system of sections is not globally defined. Special care has been taken in [30], [31] to deal with this case.

In section 2.1 we prove theorem B using our results in area preserving maps from [30], [31]. In section 3 we show that the return map in a neighborhood of Krot{K_{rot}} extends to the boundary and that the extension of hyperbolic rotating boundary orbits give rise to saddle periodic orbits for the return map. In section 4 we give a proof of theorem 1.3 adapted to our application.

2. Proof of Theorem B.

2.1. Auto-accumulation of invariant manifolds.

Proof of item (1).

2.1 Theorem (Contreras, Mazzuchelli [9] Thm. A).

Any closed contact 3-manifold satisfying the Kupka-Smale condition admits a Birkhoff section for its Reeb flow.

2.2 Definition.

Let SS be a compact orientable surface with boundary. Suppose that f:SSf:S\to S is a orientation preserving homeomorphism. We say that a periodic point xFix(fn)x\in\operatorname*{Fix}(f^{n}) is hyperbolic or of saddle type if there is an open neighborhood VV and a local chart h:VWh:V\to W such that W=]1,1[2W=]-1,1[^{2}, if pint(S)p\in\operatorname{int}(S) or W=]1,1[×[0,1[W=]-1,1[\times[0,1[, if pSp\in\partial S, h(p)=(0,0)h(p)=(0,0) and hfh1=gh\circ f\circ h^{-1}=g, where g(x,y)=(λx,λ1y)g(x,y)=(\lambda x,\lambda^{-1}y) with λ\lambda\in{\mathbb{R}}, λ{1,0,1}\lambda\notin\{-1,0,1\}.

In such coordinates the set {(x,y)V|x0 and y0}\{(x,y)\in V\,|\,x\neq 0\text{ and }y\neq 0\,\} has two or four connected components that contain pp in their closures. We call them sectors of pp. If Σ\Sigma is one of these sectors and Σ\Sigma^{\prime} is a sector of pp defined by means of another coordinate neighborhood VV^{\prime} of pp then either ΣΣ=\Sigma\cap\Sigma^{\prime}=\emptyset or Σ\Sigma and Σ\Sigma^{\prime} define the same germ at pp. We say that the set AA contains a sector Σ\Sigma if AA contains a set Σ\Sigma^{\prime} germ equivalent to Σ\Sigma at pp. We say that a set BB accumulates on a sector Σ\Sigma of pp if the closure of BΣB\cap\Sigma contains pp. These definitions do not depend on the choice of VV neither on the choice of the linear map (x,y)(λx,λ1y)(x,y)\mapsto(\lambda x,\lambda^{-1}y).

The stable and unstable manifolds of pp are

Ws(p,f)\displaystyle W^{s}(p,f) ={qS:limm+d(fm(p),fm(q))=0},\displaystyle=\{q\in S:\lim_{m\to+\infty}d(f^{m}(p),f^{m}(q))=0\,\},
Wu(p,f)\displaystyle W^{u}(p,f) ={qS:limmd(fm(p),fm(q))=0}.\displaystyle=\{q\in S:\lim_{m\to-\infty}d(f^{m}(p),f^{m}(q))=0\,\}.

The branches of pp are the connected components of Ws(p,f){p}W^{s}(p,f)\setminus\{p\} or of Wu(p,f){p}W^{u}(p,f)\setminus\{p\}. A connection between two periodic points p,qSp,\,q\in S is a branch of pp which is also a branch of qq, i.e. a whole branch which is contained in Ws(p,f)Wu(q,f)W^{s}(p,f)\cap W^{u}(q,f) or in Wu(p,f)Ws(p,f)W^{u}(p,f)\cap W^{s}(p,f). We say that a branch LL and a sector Σ\Sigma are adjacent if a local branch of LL is contained in the closure of Σ\Sigma in SS. Two branches are adjacent if they are adjacent to a single sector.

A periodic point pFix(fn)p\in\operatorname*{Fix}(f^{n}) is irrationally elliptic if ff is C1C^{1} in a neighborhood of pp and no eigenvalue of dfn(p)df^{n}(p) is a root of unity.

Let (N,λ)(N,\lambda) be a Kupka-Smale closed contact 3-manifold and Σ\Sigma a Birkhoff section for its Reeb flow ψ\psi. The first return times τ±:Σ+\tau_{\pm}:\Sigma\to{\mathbb{R}}^{+} and the first return maps f±1:ΣΣf^{\pm 1}:\Sigma\to\Sigma to Σ\Sigma are defined by

τ±(x):=±min{t>0:ψ±t(x)Σ},f±1(x):=ψτ±(x)(x).\displaystyle\tau_{\pm}(x):=\pm\min\{t>0:\psi_{\pm t}(x)\in\Sigma\},\qquad f^{\pm 1}(x):=\psi_{\tau_{\pm}(x)}(x).

We have that f±1f^{\pm 1} are smooth diffeomorphisms of Σ\Sigma preserving the area form dλd\lambda on Σ\Sigma. We are going to apply to ff the following Theorem:

2.3 Theorem (Oliveira, Contreras [31] corollary 4.9).

Let SS be a compact connected orientable surface with boundary provided with a finite measure μ\mu which is positive on open sets and f:SSf:S\to S be an orientation preserving and area preserving homeomorphism of SS.

  1. (1)

    Suppose that LL is a (periodic) branch of ff and that all periodic points of ff contained in clSLcl_{S}L are of saddle type or irrationally elliptic. Then either LL is a connection or LL accumulates on both adjacent sectors. In the later alternative Lω(L)L\subset\omega(L).

  2. (2)

    Let pSSp\in S-\partial S be a periodic point of ff of saddle type and let L1L_{1} and L2L_{2} be adjacent branches of pp that are not connections. If all the periodic points of ff contained in clS(L1L2)cl_{S}(L_{1}\cup L_{2}) are of saddle type or irrationally elliptic, then clSL1=clSL2cl_{S}L_{1}=cl_{S}L_{2}.

  3. (3)

    Suppose that pSSp\in S-\partial S is a periodic point of ff of saddle type. Assume that all the periodic points contained in clS(WpuWps)cl_{S}(W^{u}_{p}\cup W^{s}_{p}) are of saddle type or irrationally elliptic and pp has no connections. Then the branches of pp have the same closure and each branch of pp accumulates on all the sectors of pp.

    If in addition SS has genus 0 or 1, then the four branches of pp have homoclinic points.

  4. (4)

    Let CC be a connected component of S\partial S and suppose that all the periodic points p1,,p2np_{1},\ldots,p_{2n} of ff in CC are of saddle type. Let LiL_{i} be the branch of pip_{i} contained in SSS-\partial S. Assume that for every ii all the periodic points of ff contained in clSLicl_{S}L_{i} are of saddle type or irrationally elliptic and that LiL_{i} is not a connection. Then for every pair (i,j)(i,j) the branch LiL_{i} accumulates on all the sectors of pjp_{j} and clSLi=clSLjcl_{S}L_{i}=cl_{S}L_{j}.

    If in addition SS has genus 0 then any pair (Li,Lj)(L_{i},L_{j}) of stable and unstable branches intersect. The same happens if the genus of SS is 1 provided that there are at least 4 periodic points in CC.

Item (2) of theorem 2.3 for closed manifolds without boundary and under the further hypothesis that the elliptic periodic points are Moser stable appears in Mather [26] theorem 5.2. It also appears in Franks, Le Calvez [15] theorem 6.2 for S=SS2S=\SS^{2}, the 2-sphere, when the elliptic points are Moser stable. The proof of items (3) and (4) on the existence of homoclinic orbits using item (2), appears in the proof of theorem 4.4 of [31] and can be read independently of the rest of the paper.

In section §3 we prove that if SS is a Birkhoff section for the Reeb flow ψt\psi_{t} of (N,λ)(N,\lambda), then there is a continuous extension of the return map f:SSf:S\to S to the boundary S\partial S which preserves its boundary components as in figure 4. If γ\gamma is a boundary component of S\partial S which is an irrationally elliptic closed orbit then the restriction f|γf|_{\gamma} has no periodic points. If γ\gamma is a hyperbolic closed orbit then the extension f|γf|_{\gamma} has periodic points which are the limits in γ\gamma of the intersections Ws(γ)SW^{s}(\gamma)\cap S and Wu(γ)SW^{u}(\gamma)\cap S. The extension f|γf|_{\gamma} corresponds to the action of the derivative of the flow dψtd\psi_{t} on the projective space of the contact structure ξ\xi, transversal to the vector field. Therefore the limits of the intersections Ws(γ)SW^{s}(\gamma)\cap S are sources in f|γf|_{\gamma} and the limits of the intersections Wu(γ)SW^{u}(\gamma)\cap S are sinks in f|γf|_{\gamma}. The other points in γ\gamma are connections among these sources and sinks, i.e. stable manifolds of sinks which coincide with unstable manifolds of sources inside γ\gamma. These periodic points in γ\gamma are saddles for ff in S¯\overline{S}. The sinks in γ\gamma have an unstable manifold in SS which is a connected component of Wu(γ)SW^{u}(\gamma)\cap S. Similarly, the sources for f|γf|_{\gamma} are saddles in SS with stable manifold a connected component of Ws(γ)SW^{s}(\gamma)\cap S. The Kupka-Smale condition for the Reeb flow ψt\psi_{t} implies that the branches in SSS-\partial S of periodic points in S\partial S are not connections. In fact their intersections with other branches of periodic points of ff are transversal.

Therefore we can apply the first part of items (3) and (4) of theorem 2.3 to the return map ff of a Birkhoff section for the Reeb flow ψt\psi_{t}. This implies item (1) of theorem B. For periodic points which intersect the interior of SS we use item (3) of theorem 2.3 and for hyperbolic periodic orbits in S\partial S we use item (4) of theorem 2.3.

2.2. Homoclinics for Birkhoff sections.

Proof of item (2).

We saw in the proof of item (1) of theorem B in subsection §2.1 that we can apply theorem 2.3 to the first return map of a Birkhoff section for the Kupka-Smale Reeb flow. In the case that the Reeb flow admits a Birkhoff section with genus zero or one, we can also apply the second part of items (3) and (4) of theorem 2.3. This gives homoclinic orbits in every separatrix of all the hyperbolic closed orbits for the Reeb flows ψt\psi_{t} selected in item (2) of theorem B.

2.3. Complete system of surfaces of section.

Let (N,λ)(N,\lambda) be a compact contact 3-manifold and ψt\psi_{t} its Reeb flow. For ZNZ\subset N define the forward trapped set trap+(Z)\operatorname{trap}_{+}(Z) and the backward trapped set trap(Z)\operatorname{trap}_{-}(Z) as

trap±(Z)={zN:τt>τϕ±t(z)Z}.\operatorname{trap}_{\pm}(Z)=\{\,z\in N\,:\;\exists\tau\quad\forall t>\tau\quad\phi_{\pm t}(z)\in Z\,\}.
Refer to caption
Figure 1. This figure shows in the left how a neighborhood NN of a boundary closed orbit γKfix\gamma\in{K_{fix}} arrives to γ\gamma in a local transversal section to the flow. The figure in the right shows a neighborhood NN of a rotating boundary orbit γKrot\gamma\in{K_{rot}} and some of the interates of NN under the Reeb flow.
2.4 Definition.

We say that (Σ1,,Σn)(\Sigma_{1},\ldots,\Sigma_{n}) is a complete system of surfaces of section for (N,λ)(N,\lambda) if

  1. (i)

    Each Σi\Sigma_{i} is a connected surface of section for (N,λ)(N,\lambda), i.e. ΣiN\Sigma_{i}\subset N is a connected immersed compact surface whose interior int(Σi)\operatorname{int}(\Sigma_{i}) is embedded and transversal to the Reeb vector field XX and its boundary is a cover of a finite collection of closed orbits of ψ\psi.

  2. (ii)

    Separate the boundary orbits in two sets333This classification is the same as radial and broken binding orbits for broken book decompositions. iΣi=KrotKfix\cup_{i}\partial\Sigma_{i}={K_{rot}}\cup{K_{fix}}. The non rotating periodic orbits in Kfix{K_{fix}} are hyperbolic and have a neighborhood NN in Σi\Sigma_{i} which arrives to the boundary inside a sector as in figure 1. For the rotating boundary orbits444Rotating boundary orbits can be hyperbolic or elliptic. in Krot{K_{rot}} there is >0\ell>0 such that each γKrot\gamma\in{K_{rot}} has a neighborhood NγN_{\gamma} in Σi\Sigma_{i} such that

    zNγψ]0,[(z)Σi&ψ],0[(z)Σi.\forall z\in N_{\gamma}\quad\psi_{]0,\ell[}(z)\cap\Sigma_{i}\neq\emptyset\quad\&\quad\psi_{]-\ell,0[}(z)\cap\Sigma_{i}\neq\emptyset.
  3. (iii)

    At each555This condition says that the flow rotates more than the surface of section when it approaches its boundary orbit γ\gamma. rotating boundary orbit γKrotΣi\gamma\in{K_{rot}}\cap\Sigma_{i} the extension of Σi\Sigma_{i} to the unit normal bundle 𝒩(γ){\mathcal{N}}(\gamma) of γ\gamma by blowing up a neighborhood of γ\gamma using polar coordinates, is an embedded collection of closed curves transversal to the extension of the Reeb vector field to BγB_{\gamma}.

  4. (iv)

    Every orbit intersects 𝚺=iΣi{\mathbf{\Sigma}}=\cup_{i}\Sigma_{i}.

  5. (v)

    trap±(N𝚺)Ws,u(Kfix)\operatorname{trap}_{\pm}(N\setminus{\mathbf{\Sigma}})\subset W^{s,u}({K_{fix}}).

Recall that a Birkhoff section is a connected embedded surface ΣN\Sigma\subset N whose interior is transversal to the vector field. Its boundary is a cover of finitely many closed orbits and there is >0\ell>0 such that for all zNz\in N, ψ]0,[(z)Σ\psi_{]0,\ell[}(z)\cap\Sigma\neq\emptyset and ψ],0[(z)Σ\psi_{]-\ell,0[}(z)\cap\Sigma\neq\emptyset. We use the same notation Kfix{K_{fix}}, Krot{K_{rot}}, Ki\partial K_{i}, 𝚺\partial{\mathbf{\Sigma}} for a collection of periodic orbits or their union. Here 𝚺:=iΣi\partial{\mathbf{\Sigma}}:=\cup_{i}\;\partial\Sigma_{i} and also int(𝚺):=iintΣi\operatorname{int}({\mathbf{\Sigma}}):=\cup_{i}\operatorname{int}\Sigma_{i}.

2.5 Lemma.

Let (Σi)i=1n(\Sigma_{i})_{i=1}^{n} be a complete system of surfaces of section for (N,λ)(N,\lambda).

Let γKfix\gamma\in{K_{fix}} and a connected component LWs,u(γ)γL\subset W^{s,u}(\gamma)\setminus\gamma, then

ξKfixLWu,s(ξ).\exists\xi\in{K_{fix}}\qquad L\cap W^{u,s}(\xi)\neq\emptyset.
Proof:.

We prove it only for LWu(γ)L\subset W^{u}(\gamma), the other case is similar. Suppose by contradiction that

(1) LWs(Kfix)=.L\cap W^{s}({K_{fix}})=\emptyset.

Let zLz\in L. Let SS be an essential smooth embedded circle in LL. By (1) and (v), the first return time τ:S\tau:S\to{\mathbb{R}}

τ(x):=inf{t>0:ψt(x)𝚺}\tau(x):=\inf\{\,t>0:\psi_{t}(x)\in{\mathbf{\Sigma}}\,\}

is well defined and finite on SS. The return map f:Sint(𝚺)f:S\to\operatorname{int}({\mathbf{\Sigma}}), f(x)=ψτ(x)(x)f(x)=\psi_{\tau(x)}(x) is an immersion. Since SLS\subset L is connected, compact and disjoint from periodic orbits, there is a component Σi0\Sigma_{i_{0}} of 𝚺{\mathbf{\Sigma}} such that f:Sf(S)Σi0f:S\to f(S)\subset\Sigma_{i_{0}} is a diffeomorphism. By the intrinsic dynamics of ψt\psi_{t} on LL, we have that f(S)f(S) is an essential smooth embedded circle in LL. Repeating this argument there is a component Σi1\Sigma_{i_{1}} of 𝚺{\mathbf{\Sigma}} and an infinite collection {Sk}k\{S_{k}\}_{k\in{\mathbb{N}}} of disjoint essential smooth embedded circles in LL given by Sk=fk(S)S_{k}=f_{k}(S), where fkf_{k} is the kk-th return of SS to Σi1\Sigma_{i_{1}}. Observe that for i<ji<j, the circles SiS_{i}, SjS_{j} bound an embedded annulus AijA_{ij} in LL.

The circles SkS_{k} are disjoint and embedded in Σi1\Sigma_{i_{1}}. By Lemma 3.2 or Theorem 3.3 in [23], there is a free homotopy class in Σi1\Sigma_{i_{1}} which contains infinitely many of them {Skn}\{S_{k_{n}}\}_{\in{\mathbb{N}}}.

If the circles SnkS_{n_{k}} are contractible in Σi1\Sigma_{i_{1}}, they bound disjoint disks DknD_{k_{n}} with area

Dkndλ=Sknλ=A(kn)dλ+mγλ=mperiod(γ)>0,\int_{D_{k_{n}}}d\lambda=\int_{S_{k_{n}}}\lambda=\int_{A(k_{n})}d\lambda+\int_{m\cdot\gamma}\lambda=m\cdot\text{period(}\gamma)>0,

where A(kn)A(k_{n}) is the annulus on LL with boundaries {Skn,γ}\{S_{k_{n}},\gamma\} and m{1,2}m\in\{1,2\} wether L\partial L covers γ\gamma mm-times. We have used that λ(X)1\lambda(X)\equiv 1 and that dλ|L0d\lambda|_{L}\equiv 0 because the Reeb vector field is tangent to LL. This contradicts the fact that the area of Σi1\Sigma_{i_{1}} is finite, because

area(Σi1)=Σi1𝑑λ=γΣi1mγγλ<+,\text{area}(\Sigma_{i_{1}})=\int_{\Sigma_{i_{1}}}d\lambda=\operatorname*{{\textstyle\sum}}_{\gamma\in\partial\Sigma_{i_{1}}}m_{\gamma}\int_{\gamma}\lambda<+\infty,

where Σi1\partial\Sigma_{i_{1}} is finite and mγm_{\gamma} is the covering number of Σi1\partial\Sigma_{i_{1}} over γ\gamma.

If the homotopy class of the SknS_{k_{n}} in Σi1\Sigma_{i_{1}} is non trivial then Sk1S_{k_{1}} and Sk2S_{k_{2}} bound an annulus B12B_{12} in Σi1\Sigma_{i_{1}}. The annulus B12B_{12} has positive dλd\lambda-area because the transversality of Σi1\Sigma_{i_{1}} to the Reeb vector field implies that dλd\lambda is non-degenerate on Σi1\Sigma_{i_{1}}. They also bound the annulus Ak1k2A_{k_{1}k_{2}} in LL with zero dλd\lambda-area, because dλ|L0d\lambda|_{L}\equiv 0. Therefore

0=Ak1k2𝑑λ=Sk1λSk2λ=B12𝑑λ>0.0=\int_{A_{k_{1}k_{2}}}d\lambda=\int_{S_{k_{1}}}\lambda-\int_{S_{k_{2}}}\lambda=\int_{B_{12}}d\lambda>0.

A contradiction.

2.6 Remark.

Using proposition 2.12 instead of proposition 2.1 in [8] it is possible to reproduce the proofs of lemma 5.2 and theorem B in [9] to obtain

(2) γKfixWs(γ)¯=Wu(γ)¯,\forall\gamma\in{K_{fix}}\qquad\overline{W^{s}(\gamma)}=\overline{W^{u}(\gamma)},

whenever (N,λ)(N,\lambda) is Kupka-Smale and has a complete systems of surfaces of section. Then proposition 2.8 and section 3 in [9] give a Birkhoff section for (N,λ)(N,\lambda) starting from a complete system instead of a broken book decomposition.

Here we will use theorem B.(1), proved in subsection §2.1, to get (2).

2.7 Lemma.

Let α\alpha, β\beta be hyperbolic periodic orbits of a Kupka-Smale Reeb flow of a closed contact 3-manifold (N,λ)(N,\lambda). Suppose that β\beta has homoclinic orbits. Let QQ be a separatrix of α\alpha. Suppose that

(3) cl(Ws(β)Wu(β))Q¯.cl(W^{s}(\beta)\cup W^{u}(\beta))\subset\overline{Q}.

Then

(4) QWu(α)QWs(β),\displaystyle Q\subset W^{u}(\alpha)\quad\Longrightarrow\quad Q\cap W^{s}(\beta)\neq\emptyset,
(5) QWs(α)QWu(β).\displaystyle Q\subset W^{s}(\alpha)\quad\Longrightarrow\quad Q\cap W^{u}(\beta)\neq\emptyset.

Moreover, all the separatrices of α\alpha have homoclinics.

Proof:.

Let DD be a small disk transversal to the Reeb flow containing a point pβDp\in\beta\cap D and such that

(6) Dα= and D𝑑λ<αλ.D\cap\alpha=\emptyset\quad\text{ and }\quad\int_{D}d\lambda<\int_{\alpha}\lambda.

The Kupka-Smale condition implies that the homoclinic intersections in Ws(β)Wu(β)W^{s}(\beta)\cap W^{u}(\beta) are transversal. By the λ\lambda-lemma there are segments of Wu(β)DW^{u}(\beta)\cap D (resp. Ws(β)DW^{s}(\beta)\cap D) accumulating in the C1C^{1} topology on the whole local component Wεu(β)DW^{u}_{\varepsilon}(\beta)\cap D (resp. Wεs(β)DW^{s}_{\varepsilon}(\beta)\cap D). These segments form a grid in DD nearby pp which contains rectangles of arbitrarily small diameter. Choose small rectangles AA, BB with boundaries in Ws(β)Wu(β)W^{s}(\beta)\cup W^{u}(\beta) such that

cl(A)int(B)cl(B)int(D).cl(A)\subset int(B)\subset cl(B)\subset int(D).

Since cl(Ws(β)Wu(β))Q¯cl(W^{s}(\beta)\cup W^{u}(\beta))\subset\overline{Q}, we have that QDQ\cap D accumulates on the boundary A\partial A. Then there is a point qQint(B)q\in Q\cap int(B). Let JJ be the connected component of QDQ\cap D containing qq. Since by (6) Dα=D\cap\alpha=\emptyset, JJ is either a circle or a curve with endpoints in D\partial D. Suppose first that JJ is a circle. Since JJ is transversal to the flow inside QQ and there are no periodic orbits in QQ, by Poincaré-Bendixon theorem, JJ must be an essential embedded circle in QQ. Let ADA\subset D be a disk with A=J\partial A=J. Since QQ is tangent to the Reeb vector field, dλ|Q0d\lambda|_{Q}\equiv 0. By Stokes theorem

(7) A𝑑λ=Jλ=kαλ,k{1,2},\int_{A}d\lambda=\int_{J}\lambda=k\cdot\int_{\alpha}\lambda,\quad k\in\{1,2\},

with k=2k=2 if α\alpha is negative hyperbolic. But (7) contradicts (6) because ADA\subset D. Therefore JJ is a curve with endpoints in D\partial D. Since qJint(B)q\in J\cap int(B)\neq\emptyset and BD=B\cap\partial D=\emptyset, we have that rJB\exists r\in J\cap\partial B\neq\emptyset. Then rQWu,s(β)r\in Q\cap W^{u,s}(\beta) if QWs,u(α)Q\subset W^{s,u}(\alpha). This proves (4) and (5).

Suppose now that QWu(α)Q\subset W^{u}(\alpha), the case QWs(α)Q\subset W^{s}(\alpha) is similar. By (4)

(8) QWs(β).Q\cap W^{s}(\beta)\neq\emptyset.

But by B.(1), W¯s(α)=W¯u(α)=Q¯\overline{W}^{s}(\alpha)=\overline{W}^{u}(\alpha)=\overline{Q}. Hence cl(Ws(β)Wu(β))Q¯=W¯s(α)cl(W^{s}(\beta)\cup W^{u}(\beta))\subset\overline{Q}=\overline{W}^{s}(\alpha). By (5) applied to a separatrix in Ws(α)W^{s}(\alpha) we have that

(9) Wu(β)Ws(α).W^{u}(\beta)\cap W^{s}(\alpha)\neq\emptyset.

Since by the Kupka-Smale condition the heteroclinic intersections are transversal, equations (8), (9) and the λ\lambda-lemma imply that QWs(α)Q\cap W^{s}(\alpha)\neq\emptyset. Thus the separatrix QQ has homoclinics. Now observe that by B.(1) the condition (3) is satisfied by all the separatrices of α\alpha.

2.8 Proposition.

Let (N,λ)(N,\lambda) be a closed contact 3-manifold satisfying the Kupka-Smale condition. Let (Σ1,,Σm)(\Sigma_{1},\ldots,\Sigma_{m}) be a complete system of surfaces of section with boundary components K=KrotKfixK=K_{\text{rot}}\cup{K_{fix}}. Then every component of Ws,u(γ)γW^{s,u}(\gamma)\setminus\gamma of every non rotating boundary orbit γKfix\gamma\in{K_{fix}} has homoclinics and Ws(γ)¯=Wu(γ)¯\overline{W^{s}(\gamma)}=\overline{W^{u}(\gamma)}.

Proof:.

Write (κ1,,κn)Γ(\kappa_{1},\ldots,\kappa_{n})\in\Gamma if i\forall i κiKfix\kappa_{i}\in{K_{fix}} and Wu(κi)Ws(κi+1)W^{u}(\kappa_{i})\cap W^{s}(\kappa_{i+1})\neq\emptyset for 1i<n1\leq i<n. The definition of Γ\Gamma implies that

(10) (κ1,κ2)Γ,(κ2,κ3)Γ(κ1,κ2,κ3)Γ.(\kappa_{1},\kappa_{2})\in\Gamma,\quad(\kappa_{2},\kappa_{3})\in\Gamma\quad\Longrightarrow\quad(\kappa_{1},\kappa_{2},\kappa_{3})\in\Gamma.

The λ\lambda-lemma implies that

(11) (κ1,,κn)Γ(κ1,κn)Γ.(\kappa_{1},\ldots,\kappa_{n})\in\Gamma\quad\Longrightarrow\quad(\kappa_{1},\kappa_{n})\in\Gamma.

By lemma 2.5

(12) βKfixα,γKfix{(α,β),(β,γ)}Γ.\forall\beta\in{K_{fix}}\quad\exists\alpha,\gamma\in{K_{fix}}\qquad\{(\alpha,\beta),(\beta,\gamma)\}\subset\Gamma.

By (12) for any αKfix\alpha\in{K_{fix}} there is an infinite sequence (α,κ1,κ2,)Γ(\alpha,\kappa_{1},\kappa_{2},\ldots)\in\Gamma. Since Kfix{K_{fix}} is finite, there are nmn\neq m such that kn=km=:βk_{n}=k_{m}=:\beta. By properties (10) and (11), (α,β,β)Γ(\alpha,\beta,\beta)\in\Gamma. Thus

(13) αKfixβKfix(α,β,β)Γ.\forall\alpha\in{K_{fix}}\quad\exists\beta\in{K_{fix}}\qquad(\alpha,\beta,\beta)\in\Gamma.

Let αKfix\alpha\in{K_{fix}} and let QQ be a component of Ws,u(α)αW^{s,u}(\alpha)\setminus\alpha. By theorem B.(1), proved in §2.1, Q¯=Ws(α)¯=Wu(α)¯\overline{Q}=\overline{W^{s}(\alpha)}=\overline{W^{u}(\alpha)}.

Let βKfix\beta\in{K_{fix}} be given by (13). Since the intersection Wu(α)Ws(β)W^{u}(\alpha)\cap W^{s}(\beta) is transversal, by the λ\lambda-lemma Wu(β)Wu(α)¯W^{u}(\beta)\subset\overline{W^{u}(\alpha)}. Then by theorem B.(1),

cl(Ws(β)Wu(β))=Wu(β)¯Wu(α)¯=Q¯.cl(W^{s}(\beta)\cup W^{u}(\beta))=\overline{W^{u}(\beta)}\subset\overline{W^{u}(\alpha)}=\overline{Q}.

Then lemma 2.7 implies that QQ has homoclinics.

2.4. Homoclinics for complete systems.

Proof of item B.(3).

We shall use the following

2.9 Theorem (The accumulation lemma).

Let SS be a connected surface with compact boundary provided with a Borel measure μ\mu such that open non-empty subsets have positive measure and compact subsets have finite measure. Let S0SS_{0}\subset S be an open subset with frSS0fr_{S}S_{0} compact.

Let f,f1:S0Sf,f^{-1}:S_{0}\to S be an area preserving homeomorphism of S0S_{0} onto open subsets f(S0)f(S_{0}), f1(S0)f^{-1}(S_{0}) of SS. Let KS0K\subset S_{0} be a compact connected invariant subset of S0S_{0}.

If LS0L\subset S_{0} is a branch of ff and LKL\cap K\neq\emptyset, then LKL\subset K.

This version of theorem 2.9 is proved in [30, Thm. 4.3], its proof also applies to branches LL of saddle points in the boundary S0\partial S_{0}. Theorem 2.9 was originally proved in Mather [26, corollary 8.3] for surfaces without boundary and global maps (S0=SS_{0}=S). It is also proved in Franks, Le Calvez [15, lemma 6.1] for S=S0=SS2S=S_{0}=\SS^{2}, the 2-sphere. This version is needed to prove theorem 2.13 in  [31]. In proposition 2.10 we only use its global version S0=SS_{0}=S, but in corollary 2.15 we use this version for partially defined maps.

2.10 Proposition.

Let (N,λ)(N,\lambda) be a closed contact 3-manifold satisfying the Kupka-Smale condition with a given complete system of surfaces of section. Let Kfix{K_{fix}} be the set of non rotating boundary orbits let

𝕎=(Ws(Kfix)Wu(Kfix))Kfix.{\mathbb{W}}=(W^{s}({K_{fix}})\cup W^{u}({K_{fix}}))\setminus{K_{fix}}.

Let γ\gamma be a hyperbolic closed orbit of the Reeb flow of (N,λ)(N,\lambda). Let QQ be a separatrix of γ\gamma. If Q¯𝕎\overline{Q}\cap{\mathbb{W}}\neq\emptyset, then all the separatrices of γ\gamma have homoclinics.

Proof:.

Assume that QWu(γ)Q\subset W^{u}(\gamma), the case QWs(γ)Q\subset W^{s}(\gamma) is similar.

By theorem 2.1 there is a Birkhoff section {\mathcal{B}} for the Kupka-Smale Reeb flow of (N,λ)(N,\lambda). Let ff be the first return map of the Reeb flow to {\mathcal{B}}.

In section 3 we show that in case γ\gamma\subset\partial{\mathcal{B}} is a hyperbolic periodic orbit, then the connected components of QQ\cap{\mathcal{B}} are interior branches of saddle points in γ\gamma for the return map ff to {\mathcal{B}}. Choose a connected component LpL_{p} of QQ\cap{\mathcal{B}}. Then LpL_{p} is a branch of a periodic point pp\in{\mathcal{B}} of ff, possibly at the boundary pp\in\partial{\mathcal{B}} if γ\gamma\subset\partial{\mathcal{B}}. Let nn be the minimal period of LpL_{p}, fn(Lp)=Lpf^{n}(L_{p})=L_{p}. In particular fn(p)=pf^{n}(p)=p. Then LpL_{p} is a branch of Wu(p,fn)W^{u}(p,f^{n}). And K=L¯pK=\overline{L}_{p} is a compact fnf^{n}-invariant subset of {\mathcal{B}}.

Let κKfix\kappa\in{K_{fix}} be a non rotating boundary orbit666In lemma 3.3 in [9] an argument of Fried is used to show that, since by proposition 2.8 an orbit κKfix\kappa\in{K_{fix}} has homoclinics in all its branches, one can obtain a Birkhoff section {\mathcal{B}} which intersects κ\kappa in its interior. and let qκq\in\kappa\cap{\mathcal{B}} be a (saddle) periodic point for ff. Choose a multiple mm of nn such that fm(q)=qf^{m}(q)=q, then {p,q}Fix(fm)\{p,q\}\subset\operatorname*{Fix}(f^{m}) and fm(K)=Kf^{m}(K)=K. We will apply the accumulation lemma 2.9 to fmf^{m} and the compact fmf^{m}-invariant set KK. Let Lqint()L_{q}\subset\operatorname{int}({\mathcal{B}}) be an interior branch of qq, i.e. a connected component of Wτ(q,fm){q}W^{\tau}(q,f^{m})\setminus\{q\}, τ{s,u}\tau\in\{s,u\}, which is also a connected component of Wτ(κ)W^{\tau}(\kappa)\cap{\mathcal{B}}.

By the accumulation lemma 2.9,

LqKLqK.L_{q}\cap K\neq\emptyset\qquad\Longrightarrow\qquad L_{q}\subset K.

This implies in the Reeb flow that

(Wτ(κ)κ)Q¯ (a component of Wτ(κ)κ)Q¯.(W^{\tau}(\kappa)\setminus\kappa)\cap\overline{Q}\neq\emptyset\qquad\Longrightarrow \qquad(\text{a component of }W^{\tau}(\kappa)\setminus\kappa)\subset\overline{Q}.

By item (1) of theorem B we have that W¯s(κ)=W¯u(κ)=W¯τ(κ)\overline{W}^{s}(\kappa)=\overline{W}^{u}(\kappa)=\overline{W}^{\tau}(\kappa) is also the closure of any component of Wτ(κ)κW^{\tau}(\kappa)\setminus\kappa, therefore we get

(14) τ{s,u}Q¯(Wτ(κ)κ)cl(Ws(κ)Wu(κ))Q¯.\exists\tau\in\{s,u\}\qquad\overline{Q}\cap(W^{\tau}(\kappa)\setminus\kappa)\neq\emptyset\qquad\Longrightarrow\qquad cl(W^{s}(\kappa)\cup W^{u}(\kappa))\subset\overline{Q}.

Suppose that Q¯𝕎\overline{Q}\cap{\mathbb{W}}\neq\emptyset. Then there is κKfix\kappa\in{K_{fix}} and τ{s,u}\tau\in\{s,u\} such that Q¯(Wτ(κ)κ)\overline{Q}\cap(W^{\tau}(\kappa)\setminus\kappa)\neq\emptyset. By (14),

cl(Ws(κ)Wu(κ))Q¯.cl(W^{s}(\kappa)\cup W^{u}(\kappa))\subset\overline{Q}.

By proposition 2.8 we have that κ\kappa has transversal homoclinics. Then by lemma 2.7 all the separatrices of γ\gamma have homoclinics.

Let SS be a component of a complete system of surfaces of section for a closed contact 3-manifold (N,λ)(N,\lambda). Define the first return times τ±\tau_{\pm} to SS and the first return maps f±1f^{\pm 1} as

τ+:int(S)]0,+[{+} and τ:int(S)],0[{} by\displaystyle\tau_{+}:\operatorname{int}(S)\to]0,+\infty[\cup\{+\infty\}\quad\text{ and }\quad\tau_{-}:\operatorname{int}(S)\to]-\!\infty,0[\cup\{-\infty\}\quad\text{ by}
(15) τ±(z):=±inf{t>0:ϕ±t(x)S}.\displaystyle\tau_{\pm}(z):=\pm\inf\{t>0:\phi_{\pm t}(x)\in S\,\}.
(16) f(x):=ϕτ+(x)(x),f1(x):=ϕτ(x)(x).\displaystyle f(x):=\phi_{\tau_{+}(x)}(x),\qquad f^{-1}(x):=\phi_{\tau_{-}(x)}(x).

By the implicit function theorem ff and f1f^{-1} are defined in the open subsets [τ+<+][\tau_{+}<+\infty] and [τ>][\tau_{-}>-\infty] of int(S)\operatorname{int}(S) respectively.

In section 4 we show that ff and f1f^{-1} extend to a neighborhood of KrotS{K_{rot}}\cap\partial S as in figure 4. All the periodic points for f±1f^{\pm 1} in any γKrotS\gamma\in{K_{rot}}\cap\partial S are of saddle type, their invariant manifolds for f±1f^{\pm 1} are either the intersections Ws,u(γ)SW^{s,u}(\gamma)\cap S or heteroclinic connections in γ\gamma. Irrationally elliptic orbits in S\partial S are in KrotS{K_{rot}}\cap\partial S, but they have no periodic orbits for f±1f^{\pm 1}. And f±1f^{\pm 1} are not defined777In fact the natural extension of ff to a point xKfixx\in{K_{fix}} would be the whole circle of a first intersection of a component of Wu(γ)γW^{u}(\gamma)\setminus\gamma, γ=ψ(x)\gamma=\psi_{\mathbb{R}}(x), with SS. See figure 1. on KfixS{K_{fix}}\cap\partial S. Let

(17) S0=([τ+<+][τ>])(KrotS).S_{0}=([\tau_{+}<+\infty]\cap[\tau_{-}>-\infty])\cup({K_{rot}}\cap\partial S).

By condition 2.4.(ii) the maps τ±\tau_{\pm} are finite in a neighborhood in int(S)\operatorname{int}(S) of KrotS{K_{rot}}\cap\partial S. Then S0S_{0} is an open submanifold of SS with compact boundary S0S\partial S_{0}\subset\partial S and f±1:S0Sf^{\pm 1}:S_{0}\to S are differentiable, area preserving and f(S0)S0f(\partial S_{0})\subset\partial S_{0}.

Observe that condition  2.4.(ii) implies that Kfixn[|τ±|>n]¯{K_{fix}}\subset\cap_{n\in{\mathbb{N}}}\overline{[|\tau_{\pm}|>n]}. In section 3 we see that the functions τ±\tau_{\pm} can be extended to Krot{K_{rot}}. So we use the notation

(18) Kfix[τ±=±],Krot[|τ±|<],\displaystyle{K_{fix}}\subset[\tau_{\pm}=\pm\infty],\qquad{K_{rot}}\subset[|\tau_{\pm}|<\infty],
(19) S0=[τ+<+][τ>].\displaystyle S_{0}=[\tau_{+}<+\infty]\cap[\tau_{-}>-\infty].
2.11 Lemma.
(20) ε>0Nxint(S)&N|τ±(x)|<d(x,[|τ±|=])<ε.\displaystyle\forall\varepsilon>0\quad\exists N\in{\mathbb{N}}\quad x\in\operatorname{int}(S)\;\;\&\;\;N\leq|\tau_{\pm}(x)|<\infty\quad\Longrightarrow\quad d(x,[|\tau_{\pm}|=\infty])<\varepsilon.
Proof:.

We only prove it for τ+\tau_{+}. For n{+}n\in{\mathbb{N}}\cup\{+\infty\} let An:=[τ+<n]A_{n}:=[\tau_{+}<n]. Then AnA_{n} is an increasing family of open sets in the closure cl(S)cl(S) with A=nAnA_{\infty}=\cup_{n\in{\mathbb{N}}}A_{n}. For δ>0\delta>0 let B(A,δ):={xS:d(x,A)<δ}B(\partial A_{\infty},\delta):=\{\,x\in S:d(x,\partial A_{\infty})<\delta\,\}. Observe that A=[τ+=]\partial A_{\infty}=[\tau_{+}=\infty] in cl(S)cl(S). Indeed, by Poincaré recurrence theorem, AA_{\infty} has total measure in SS, then KfixA{K_{fix}}\subset\partial A_{\infty}. It is enough to prove that

(21) ε>0NAANB(A,ε).\forall\varepsilon>0\quad\exists N\in{\mathbb{N}}\qquad A_{\infty}\subset A_{N}\cup B(\partial A_{\infty},\varepsilon).

Let Kε:=AB(A,ε2)K_{\varepsilon}:=A_{\infty}\setminus B(\partial A_{\infty},\frac{\varepsilon}{2}). Then KεK_{\varepsilon} is compact and {An}n\{A_{n}\}_{n\in{\mathbb{N}}} is an open cover of KεK_{\varepsilon}. Since the family {An}\{A_{n}\} is increasing, there is NN\in{\mathbb{N}} such that KεANK_{\varepsilon}\subset A_{N}. Then AKεB(A,ε)ANB(A,ε)A_{\infty}\subset K_{\varepsilon}\cup B(\partial A_{\infty},\varepsilon)\subset A_{N}\cup B(\partial A_{\infty},\varepsilon).

2.12 Proposition (M. Mazzucchelli).

Let NN be a compact 3-manifold with a flow ψt\psi_{t}. Let 𝚺{\mathbf{\Sigma}} be a finite union of connected surfaces of section and KK a finite collection of hyperbolic periodic orbits in 𝚺\partial{\mathbf{\Sigma}}. Suppose that

  1.  (a)

    Every orbit of ψ\psi intersects 𝚺{\mathbf{\Sigma}}.

  2.  (b)

    zN&ψ[0,+[(z)𝚺=zWs(K)z\in N\quad\&\quad\psi_{[0,+\infty[}(z)\cap{\mathbf{\Sigma}}=\emptyset\qquad\Longrightarrow\qquad z\in W^{s}(K).

Let Σ1\Sigma_{1} be a connected component of int(𝚺)\operatorname{int}({\mathbf{\Sigma}}). Let τ:Σ1]0,+[{+}\tau:\Sigma_{1}\to]0,+\infty[\cup\{+\infty\} be the first return time to the component Σ1\Sigma_{1}, i.e.

τ(z):=inf{t>0|ψt(z)Σ1}.\tau(z):=\inf\{\,t>0\;|\;\psi_{t}(z)\in\Sigma_{1}\}.

Let α:[0,1[Σ1\alpha:[0,1[\to\Sigma_{1} be continuous and suppose that

  1. (i) 

    s[0,1[τ(α(s))<+\forall s\in[0,1[\quad\tau(\alpha(s))<+\infty.

  2. (ii) 

    {sn}n[0,1[limα(sn)=wint(Σ1),τ(w)=+.\exists\{s_{n}\}_{n\in{\mathbb{N}}}\subset[0,1[\quad\lim\alpha(s_{n})=w\in\operatorname{int}(\Sigma_{1}),\quad\tau(w)=+\infty.

Then wWs(K)w\in W^{s}(K).

Proof:.

Let

k(s)=#{t[0,τ(α(s))]:ψt(α(s))𝚺}.k(s)=\#\{\,t\in[0,\tau(\alpha(s))]:\psi_{t}(\alpha(s))\in{\mathbf{\Sigma}}\,\}.

By the implicit function theorem k(s)=k0k(s)=k_{0} is constant in s[0,1[s\in[0,1[.

Suppose by contradiction that wWs(K)w\notin W^{s}(K). Then ψt(w)Ws(K)\psi_{t}(w)\notin W^{s}(K) for all t>0t>0. Hypothesis (b) then implies that ψ[0,+[(w)𝚺\psi_{[0,+\infty[}(w)\cap{\mathbf{\Sigma}} is infinite. Let T>0T>0 be such that ψ[0,T[(w)\psi_{[0,T[}(w) intersects (k0+1)(k_{0}+1) times the surface 𝚺{\mathbf{\Sigma}}. By the implicit function theorem there is a neighborhood UU of ww in int(Σ1)\operatorname{int}(\Sigma_{1}) such that for each xUx\in U, the curve ψ[0,T[(x)\psi_{[0,T[}(x) intersects (k0+1)(k_{0}+1)-times 𝚺{\mathbf{\Sigma}}. Therefore τ(α(s))T\tau(\alpha(s))\leq T whenever α(s)U\alpha(s)\in U.

Thus τ(α(sn))T\tau(\alpha(s_{n}))\leq T for nn large enough. There is a subsequence snks_{n_{k}} such that 0<τ1:=limkτ(α(snk))T0<\tau_{1}:=\lim_{k}\tau(\alpha(s_{n_{k}}))\leq T exists. Then

ψτ1(w)=limkψτ(α(snk))(α(snk))Σ1.\psi_{\tau_{1}}(w)=\lim_{k}\psi_{\tau(\alpha(s_{n_{k}}))}(\alpha(s_{n_{k}}))\in\Sigma_{1}.

Therefore τ(w)τ1<+\tau(w)\leq\tau_{1}<+\infty. A contradiction.

The previous results will allow us to obtain homoclinics for branches of periodic points whose closure is not included in S0S_{0}. For the remaining case we will use theorem 2.13.

We remark that the proof of existence of homoclinic orbits in [31], once the auto accumulation of invariant manifolds is known, only uses the dynamics of the map in a neighborhood of the invariant manifolds. We will use the following

2.13 Theorem (Oliveira, Contreras [31], corollary 4.10).

Let SS be a compact connected orientable surface with boundary. Let S0SS_{0}\subset S be a submanifold with compact boundary S0S\partial S_{0}\subset\partial S and let f,f1:S0Sf,f^{-1}:S_{0}\to S be an orientation preserving and area preserving homeomorphism of S0S_{0} onto open subsets fS0fS_{0}, f1S0f^{-1}S_{0} of SS with f(S0)S0f(\partial S_{0})\subset\partial S_{0}.

  1. (1)

    Let pS0Sp\in S_{0}-\partial S be a periodic point of ff of saddle type. Assume that the branches of pp have closure included in S0S_{0}. Assume also that each branch of pp accumulates on both of its adjacent sectors and that all the branches of pp have the same closure in SS. If in addition SS has genus 0 or 1, then the four branches of pp have homoclinic points.

  2. (2)

    Let CC be a connected component of S0\partial S_{0} and suppose that all the periodic points p1,,p2np_{1},\ldots,p_{2n} of ff in CC are of saddle type. Let LiL_{i} be the branch of pip_{i} contained in SSS-\partial S. Assume that for every ii, LiL_{i} is not a connection and clSLi=clSLjS0cl_{S}L_{i}=cl_{S}L_{j}\subset S_{0} for every pair (i,j)(i,j).

    If in addition SS has genus 0, then every pair LiL_{i}, LjL_{j} of stable and unstable branches intersect. The same happens if the genus of SS is 1 provided that there are at least 44 periodic points in CC.

Theorem 2.13 is the version for periodic points of theorem 4.4 in [31]. The proof of theorem 4.4 in [31] can be read independently of the rest of the paper.


Proof of item (3) of theorem B:

By proposition 2.8 every orbit γKfix\gamma\in{K_{fix}} has homoclinics in all its separatrices.

By proposition 2.10 the same happens for a periodic orbit γ\gamma if W¯s(γ)W¯u(γ)\overline{W}^{s}(\gamma)\cup{\overline{W}^{u}(\gamma)} intersects

𝕎:=Ws(Kfix)Wu(Kfix)Kfix.{\mathbb{W}}:=W^{s}({K_{fix}})\cup W^{u}({K_{fix}})\setminus{K_{fix}}.

So assume that γ\gamma is a hyperbolic periodic orbit with γS¯\gamma\cap\overline{S}\neq\emptyset and

(22) (W¯s(γ)W¯u(γ))𝕎=.\big{(}\overline{W}^{s}(\gamma)\cup\overline{W}^{u}(\gamma)\big{)}\cap{\mathbb{W}}=\emptyset.

Let SS be a component of the complete system. Let τ±:S{,+}\tau_{\pm}:S\to{\mathbb{R}}\cup\{-\infty,+\infty\} be the first return times to SS, defined in (15), (18), let S0=[τ+<+][τ>]S_{0}=[\tau_{+}<+\infty]\cap[\tau_{-}>-\infty] be as in (17), (19), and let f,f1:S0Sf,\,f^{-1}:S_{0}\to S be the extensions of the first return maps as in (16) and §3.

Since γ\gamma is a periodic orbit, |τ±||\tau_{\pm}| are finite on γS\gamma\cap S, bounded by the period of γ\gamma. Thus γSS0\gamma\cap S\subset S_{0}. Let pS0=int(S0)S0p\in S_{0}=\operatorname{int}(S_{0})\cup\partial S_{0}, pγS¯p\in\gamma\cap\overline{S}, be a saddle point for ff and let Lint(S)L\subset\operatorname{int}(S) be an interior branch of pp. Let QQ be the separatrix of γ\gamma which contains LL.

Suppose that τ+\tau_{+} is unbounded on LL. Let L1LL_{1}\subset L be the connected component of L[τ+<]L\cap[\tau_{+}<\infty] with pL1p\in L_{1}. Then τ+\tau_{+} is unbounded on L1L_{1}. Let α:[0,1[L1\alpha:[0,1[\to L_{1} be a parametrization of L1L_{1}. Then there is a sequence sn[0,1[s_{n}\in[0,1[ with limnτ+(α(sn))=+\lim_{n}\tau_{+}(\alpha(s_{n}))=+\infty. Extracting a subsequence we can assume that z0=limnα(sn)z_{0}=\lim_{n}\alpha(s_{n}) exists. Since [τ+=][\tau_{+}=\infty] is compact, lemma 2.11 implies that τ+(z0)=\tau_{+}(z_{0})=\infty. But condition 2.4.(ii) implies that τ+\tau_{+} is bounded in a neighborhood of KrotS{K_{rot}}\cap\partial S. Thus z0int(S)Kfixz_{0}\in\operatorname{int}(S)\cup{K_{fix}}. If z0int(S)z_{0}\in\operatorname{int}(S) then proposition 2.12 implies that z0Ws(Kfix)z_{0}\in W^{s}({K_{fix}}). Therefore z0L¯(Ws(Kfix)Kfix)z_{0}\in\overline{L}\cap(W^{s}({K_{fix}})\setminus{K_{fix}})\neq\emptyset. This contradicts (22).

Refer to caption
Figure 2.
When limnα(sn)ηKfix\lim_{n}\alpha(s_{n})\in\eta\subset{K_{fix}}, the forward orbits of the α(sn)\alpha(s_{n}) approach Wu(η)ηW^{u}(\eta)\setminus\eta.

Then z0=limnα(sn)KfixSz_{0}=\lim_{n}\alpha(s_{n})\in{K_{fix}}\cap\partial S. Let η=ψ(z0)Kfix\eta=\psi_{\mathbb{R}}(z_{0})\in{K_{fix}}. The surface SS approaches the non rotating boundary orbit ηSKfix\eta\subset\partial S\cap{K_{fix}} through a quadrant of η\eta as in figure 2. There are tn0t_{n}\geq 0 such that limn(ψtn(α(sn)))Wu(η)η\lim_{n}(\psi_{t_{n}}(\alpha(s_{n})))\in W^{u}(\eta)\setminus\eta. Since Wu(η)ηW^{u}(\eta)\setminus\eta does not contain periodic orbits, this limit is in Wu(η)KfixW^{u}(\eta)\setminus{K_{fix}}. Since L1QL_{1}\subset Q and QQ is invariant, ψtn(α(sn))Q\psi_{t_{n}}(\alpha(s_{n}))\in Q. Therefore Q¯(Wu(η)Kfix)\overline{Q}\cap(W^{u}(\eta)\setminus{K_{fix}})\neq\emptyset. This contradicts (22).

This proves that τ+\tau_{+} is bounded on LL. A similar888For the boundedness of τ\tau_{-} we apply proposition 2.12 to the inverse flow ψt\psi_{-t}. proof shows that τ\tau_{-} is bounded on LL.

Now assume that τ±\tau_{\pm} are bounded on LL and genus(S)1\text{genus}(S)\leq 1. Then there is N>0N>0 such that

L¯[τ+N][τN]S0,\overline{L}\subset[\tau_{+}\leq N]\cap[\tau_{-}\leq N]\subset S_{0},

as required in theorem 2.13. In order to apply theorem 2.13 we need to show that

  1. (a)

    If pint(S)p\in\operatorname{int}(S) then each branch of pp accumulates on both of its adjacent sectors and all branches of pp in int(S)\operatorname{int}(S) have the same closure.

  2. (b)

    If pSp\in\partial S (and hence γ=ψ(p)KrotS\gamma=\psi_{\mathbb{R}}(p)\subset{K_{rot}}\cap\partial S), then all the components of (Ws(γ)γ)S(W^{s}(\gamma)\setminus\gamma)\cap S and of (Wu(γ)γ)S(W^{u}(\gamma)\setminus\gamma)\cap S have the same closure.

Then corollary 2.15 finishes the proof of item 3 of theorem B.

2.14 Lemma.

Let (N,λ)(N,\lambda) be a Kupka-Smale closed contact 3-manifold. Let SS be a component of a complete systems of surfaces of section for (N,λ)(N,\lambda). Let S0S_{0} be as in (19) and let f:S0S0Sf:S_{0}\cup\partial S_{0}\to S be the extension of the first return map to SS made in section 3. Let γ\gamma be a hyperbolic closed orbit for (N,λ)(N,\lambda), γKfix\gamma\notin{K_{fix}}, such that

(23) S(W¯s(γ)W¯u(γ))S0.S\cap(\overline{W}^{s}(\gamma)\cup\overline{W}^{u}(\gamma))\subset S_{0}.

Let pγ(SS)p\in\gamma\cap(S\cup\partial S) be a periodic point for ff and let LpSSL_{p}\subset S\setminus\partial S be an interior branch of pp.

Then LpL_{p} accumulates on both of its adjacent sectors.

Proof:.

Let τ+\tau_{+} and ff be from (15) and (16). The branches in SSS\setminus\partial S of pp for ff are the connected components of the intersection of the separatrices of γ\gamma with SS that contain pp as an endpoint. Let QQ be the separatrix of γ\gamma containing the branch LpL_{p}. By item (1) of theorem B we know that QQ accumulates on both of its adjacent sectors in (N,λ)(N,\lambda).

By hypothesis (23), all the branches of the ff-orbit of pp in SS have closure included in S0S_{0}. The map f:S0Sf:S_{0}\to S is well defined in S0S_{0} and is an injective immersion. In particular ff is continuous in a neighborhood in SS of the branches of the orbit of pp. And every connected component of S(Ws(γ)Wu(γ))S\cap(W^{s}(\gamma)\cup W^{u}(\gamma)) has an endpoint in an element of the ff-orbit of pp.

Let AA be a sector for (f,p)(f,p) in SS adjacent to LpL_{p}. Let nn be the minimal period of pp, fn(p)=pf^{n}(p)=p. There are at most 2n2n connected components of QSQ\cap S and they are branches of the iterates fi(p)f^{i}(p). At least one of these components accumulates on the sector AA.

Suppose first that γ\gamma is a positive hyperbolic orbit. Then for every ii\in{\mathbb{Z}}, fi(Lp)f^{i}(L_{p}) is the unique connected component of QSQ\cap S with endpoint fi(p)f^{i}(p). Let LkL_{k} be a connected component of QSQ\cap S which accumulates on the sector AA. Then LkL_{k} is a branch of an ff-periodic point pkγSp_{k}\in\gamma\cap S. There is 0k<n0\leq k<n such that fk(p)=pkf^{k}(p)=p_{k}. If k=0k=0 then the lemma holds. Assume k1k\geq 1. Since LkL_{k} accumulates on the sector AA adjacent to LpL_{p}, we have that LpLk¯L_{p}\cap\overline{L_{k}}\neq\emptyset. The compact set Lk¯\overline{L_{k}} is invariant under fnf^{n} and Lk¯S0\overline{L_{k}}\subset S_{0}. By the accumulation lemma 2.9 applied to fnf^{n}, LpLk¯L_{p}\subset\overline{L_{k}}.

Observe that fk(Lk)=:L2kf^{k}(L_{k})=:L_{2k} is a connected component of QSQ\cap S with endpoint p2k=fk(pk)p_{2k}=f^{k}(p_{k}) and it accumulates on the sector fk(A)f^{k}(A) of pkp_{k}. Similarly LkL2k¯L_{k}\subset\overline{L_{2k}}. And then LpLk¯L2k¯L_{p}\subset\overline{L_{k}}\subset\overline{L_{2k}}. Inductively Lnk=fnk(Lp)L_{nk}=f^{nk}(L_{p}) is a component of QSQ\cap S with endpoint pnk=fnk(p)=pp_{nk}=f^{nk}(p)=p. Thus Lnk=LpL_{nk}=L_{p}. Moreover Lk¯L2k¯Lnk¯=Lp¯\overline{L_{k}}\subset\overline{L_{2k}}\subset\cdots\subset\overline{L_{nk}}=\overline{L_{p}}. Therefore LpL_{p} accumulates on the sector AA.

Suppose now that pγSp\in\gamma\subset\partial S is a boundary ff-periodic point. The return map ff preserves the area form of SS and hence it preserves orientation. This implies that pp is a positive hyperbolic orbit for ff. The orbit γ\gamma for the flow may be negative hyperbolic but the return map ff permutes the interior components of QSQ\cap S. The previous proof of the positive hyperbolic case applies here.

Now suppose that pSSp\in S\setminus\partial S is a negative hyperbolic periodic point for ff. Let n>0n>0 be its minimal period, fn(p)=pf^{n}(p)=p. Let JkJ_{k} be a connected component of QSQ\cap S which accumulates on the sector AA. Let pkp_{k} be the endpoint of JkJ_{k} and let kk\in{\mathbb{N}} be such that fk(p)=pkf^{k}(p)=p_{k}. In the case Lk:=fk(Lp)=JkL_{k}:=f^{k}(L_{p})=J_{k} the same proof as in the positive hyperbolic case follows, with Lj:=fj(Lp)L_{j}:=f^{j}(L_{p}) and Lk¯L2k¯L2nk¯=Lp¯\overline{L_{k}}\subset\overline{L_{2k}}\subset\cdots\subset\overline{L_{2nk}}=\overline{L_{p}}. The accumulation lemma 2.9 is applied to f2nf^{2n} which leaves the branches LikL_{ik} invariant. We iterate 2n2n times fkf^{k}, because the map f2nkf^{2nk} fixes the branch LpL_{p}.

Suppose then that fk(Lp)Jkf^{k}(L_{p})\neq J_{k}. For jj\in{\mathbb{Z}}, let Lj:=fj(Lp)L_{j}:=f^{j}(L_{p}) and Aj:=fj(A)A_{j}:=f^{j}(A). Let KjK_{j} be the other component of QSQ\cap S with endpoint fj(p)f^{j}(p). In local coordinates KjK_{j} is the branch Lj-L_{j} adjacent to the sector Aj-A_{j}. Then Jk=KkJ_{k}=K_{k} and Kj=fj(K0)K_{j}=f^{j}(K_{0}). The branch KkK_{k} accumulates on the sector AA. Then the branch Lk=fn(Kk)L_{k}=f^{n}(K_{k}) accumulates on the sector fn(A)=Af^{n}(A)=-A, adjacent to the branch K0K_{0}. By the accumulation lemma 2.9 applied to f2nf^{2n}, for which the branches are invariant, LpKk¯L_{p}\subset\overline{K_{k}} and Kp:=K0Lk¯K_{p}:=K_{0}\subset\overline{L_{k}}.

Since the branch KkK_{k} accumulates on the sector AA, we have that the branch K2k=fk(Kk)K_{2k}=f^{k}(K_{k}) accumulates on the sector Ak=fk(A)A_{k}=f^{k}(A), adjacent to LkL_{k}. And using fnf^{n}, the branch L2k:=f2k(Lp)=fn(K2k)L_{2k}:=f^{2k}(L_{p})=f^{n}(K_{2k}) accumulates on the sector Ak=fn(Ak)-A_{k}=f^{n}(A_{k}), adjacent to KkK_{k}. Using the accumulation lemma 2.9 we get that

(24) LkK2k¯ and KkL2k¯.L_{k}\subset\overline{K_{2k}}\qquad\text{ and }\qquad K_{k}\subset\overline{L_{2k}}.

Applying fjf^{j} to the inclusions in (24) and using that Lj=fj(L0)L_{j}=f^{j}(L_{0}), Kj=fj(K0)K_{j}=f^{j}(K_{0}) we have that

LpKk¯L2k¯K3k¯L4k¯L_{p}\subset\overline{K_{k}}\subset\overline{L_{2k}}\subset\overline{K_{3k}}\subset\overline{L_{4k}}\subset\cdots

In the iterate 2nk2nk we have that

KkL2nk¯=f2nk(Lp¯)=Lp¯,K_{k}\subset\overline{L_{2nk}}=f^{2nk}(\overline{L_{p}})=\overline{L_{p}},

then LpL_{p} accumulates on the sector AA.

The hypothesis in theorem 2.13 asks for more than lemma 2.14, namely

2.15 Corollary.

Let pS0S0p\in S_{0}\cup\partial S_{0} be a saddle point for the extension f:S0Sf:S_{0}\to S of the return map. Assume that

(25) γ:=ψ(p),SWs(γ)¯S0&SWu(γ)¯S0.\gamma:=\psi_{\mathbb{R}}(p),\qquad\overline{S\cap W^{s}(\gamma)}\subset S_{0}\qquad\&\qquad\overline{S\cap W^{u}(\gamma)}\subset S_{0}.

Then

  1. (1)

    If pSSp\in S\setminus\partial S, all the branches of pp have the same closure and accumulate on all the sectors of pp.

  2. (2)

    If pSp\in\partial S, γ=ψ(p)\gamma=\psi_{\mathbb{R}}(p), p1,,p2mp_{1},\ldots,p_{2m} are the periodic points of ff in γS0\gamma\subset\partial S_{0} and LiSSL_{i}\subset S\setminus\partial S is the interior branch of pip_{i}, then Li¯=Lj¯\overline{L_{i}}=\overline{L_{j}} for every (i,j)(i,j).

Proof:.

(1). By lemma 2.14, it is enough to prove that the branches of pp in SSS\setminus\partial S have the same closure. Let nn be such that fn(p)=pf^{n}(p)=p. Let L0L_{0}, L1L_{1} be two branches of pp adjacent to the same sector AA. Since by lemma 2.14, L0L_{0} accumulates on the sector AA; we have that L1L0¯L_{1}\cap\overline{L_{0}}\neq\emptyset. By hypothesis L0¯S0\overline{L_{0}}\subset S_{0}. Also f2n(L0)=L0f^{2n}(L_{0})=L_{0}. By the accumulation lemma 2.9 applied to f2nf^{2n}, we have that L1L0¯L_{1}\subset\overline{L_{0}}. Iterating this argument we have that all branches of pp have the same closure.

Refer to caption
Figure 3. Saddle periodic points and sectors at a hyperbolic rotating boundary orbit.

(2). Let p=p0,p1,,p2m=p0p=p_{0},p_{1},\ldots,p_{2m}=p_{0} be the ordered periodic points of ff in γS\gamma\subset\partial S, as in figure 3. Let Lp=L0=L2mL_{p}=L_{0}=L_{2m} and let LiL_{i} be the branch of pip_{i} in SSS\setminus\partial S. Then the LiL_{i}’s are connected components of SWs,u(γ)S\cap W^{s,u}(\gamma). By (25), Li¯S0\overline{L_{i}}\subset S_{0} and the LiL_{i}’s are all the components of SWs,u(γ)S\cap W^{s,u}(\gamma). Let A2i,A2i+1A_{2i},A_{2i+1} be the sectors adjacent to pip_{i} chosen so that A2i1A_{2i-1} and A2iA_{2i} are adjacent to a connection between pi1p_{i-1} and pip_{i}.

By lemma 2.14, LiL_{i} accumulates in its adjacent sector A2i+1A_{2i+1}. Due to the connection, LiL_{i} also accumulates on the sector A2i+2A_{2i+2}, adjacent to Li+1L_{i+1}. Then Li¯Li+1\overline{L_{i}}\cap L_{i+1}\neq\emptyset. By the accumulation lemma 2.9 applied to fNf^{N} where NN is a multiple of the periods of LiL_{i} and Li+1L_{i+1}, we have that Li+1Li¯L_{i+1}\subset\overline{L_{i}}. Thus

L0¯L1¯L2¯L2m¯=L0¯.\overline{L_{0}}\supset\overline{L_{1}}\supset\overline{L_{2}}\supset\cdots\supset\overline{L_{2m}}=\overline{L_{0}}.

Therefore Li¯=Lj¯\overline{L_{i}}=\overline{L_{j}} for every (i,j)(i,j).


3. The extension to the boundary of the return map.

In this section we study the extension to the boundary at a rotating boundary orbit of the return map to a component of a complete system of surfaces of section a Reeb flow. On non rotating boundary orbits the return time is infinite and the return map is not defined.

We first consider the case in which rotating boundary orbit Γ\Gamma at the boundary of a component Σ\Sigma is irrationally elliptic. In that case we prove that if the Floquet multipliers of the elliptic orbit are not roots of unity then the extension to the boundary has no periodic point.

Afterwards we deal with the case in which the periodic orbit Γ\Gamma at the boundary of the component Σ\Sigma is hyperbolic. We show that there exist a continuous extension of the Poincaré map to the boundary Γ\Gamma of Σ\Sigma. This extension is a Morse-Smale map in Γ\Gamma with periodic points on Γ\Gamma that, seen in Σ\Sigma, are saddle points. These periodic points do not correspond to other periodic orbits of the Reeb flow but their interior invariant manifolds WsW^{s}, WuW^{u} are the intersections of the invariant manifolds Wu(Γ)W^{u}(\Gamma), Ws(Γ)W^{s}(\Gamma) of the closed orbit Γ\Gamma with the component Σ\Sigma and hence their intersections with other invariant manifolds of the return map to Σ\Sigma are transversal if the Reeb flow is Kupka-Smale.

Refer to caption
Figure 4. The figure shows the dynamics of the extension of the return map PP to the surface of section Σ\Sigma with hyperbolic rotating boundary orbits.

3.1. The elliptic case.

3.1 Proposition.

If the binding periodic orbit Γ\Gamma is elliptic and its Floquet multipliers are not roots of unity, then the extension of the return map to the boundary Γ=Σ\Gamma=\partial\Sigma has no periodic points.

Proof:.

As we shall see in §3.6 the component Σ\Sigma has an asymptotic direction e(t)ξ(Γ(t))e(t)\in\xi(\Gamma(t)), where ξ\xi is the contact structure. The direction of e(t)e(t) turns more slowly than its movement sdφs(e(t0))s\mapsto d\varphi_{s}(e(t_{0})) under the linearized Reeb flow. Then the extension PP of the return map to Γ=Σ\Gamma=\partial\Sigma is given by Γ(t)Γ(b(t))\Gamma(t)\mapsto\Gamma(b(t)), where

b(t)=t+min{s>0|λ>0:dφs(e(t))=λe(t+s)}.b(t)=t+\min\{\,s>0\;|\;\exists\lambda>0:\;d\varphi_{s}(e(t))=\lambda\,e(t+s)\,\}.

If the extension PP has a periodic point at Γ(t)\Gamma(t) then the subspace generated by e(t)e(t) is invariant under dφTd\varphi_{T}, where TT is a multiple of the period of Γ\Gamma. Then some iterate of the Poincaré map of the periodic orbit Γ\Gamma has an invariant 1-dimensional subspace. Since Γ\Gamma is elliptic, this implies that the Poincaré map of Γ\Gamma has an eigenvalue which is a root of unity. This is a contradiction. ∎


3.2. The hyperbolic case. Sketch of the proof.


Let Γ\Gamma be a periodic orbit for the Reeb flow of (N,λ)(N,\lambda). The contact structure ξ=kerλ\xi=\ker\lambda is a subspace transversal to the Reeb vector field which is non integrable. The image of the exponential map of a small ball Bδ(0)ξ(Γ(t))B_{\delta}(0)\cap\xi(\Gamma(t)), Ξ(Γ(t))=expΓ(t)(ξ(Γ(t))Bδ(0))\Xi(\Gamma(t))=\exp_{\Gamma(t)}(\xi(\Gamma(t))\cap B_{\delta}(0)) is a system of transversal sections to the Reeb flow in a neighborhood of Γ\Gamma which is tangent to the contact structure at Γ\Gamma.

The picture of the return map to the surface of section nearby the boundary orbit and its extension to the boundary is clear when the flow is linear in a neighborhood of the biding orbit Γ\Gamma, the contact structure ξ\xi is orthogonal to the vector field X(Γ)=Γ˙X(\Gamma)=\dot{\Gamma} and the intersections, near the boundary Σ\partial\Sigma, of the surface of section Σ\Sigma with the transversals Ξ\Xi, are images under the exponential map of straight lines. In this case the set of images of straight lines in ξ\xi passing through Γ\Gamma is a singular foliation invariant999 This does not happen on a contact flow but may happen for a reparametrization of the flow. by the flow.

In the following paragraphs we show that the return map to Σ\Sigma is conjugate to the situation described above. We choose a coordinate system in which the strong invariant manifolds Ws(ψtz)W^{s}(\psi_{t}z), Wu(ψtz)W^{u}(\psi_{t}z) coincide with coordinate axes, and the flow is linear. In these coordinates the surface of section turns around the axis at least an angle π\pi. We construct an open book decomposition {\mathcal{F}} of a neighborhood of Γ\Gamma with spine Γ\Gamma, which is invariant under the Reeb flow, whose intersections with the transversals Ξ\Xi are straight lines in these coordinates.

Then we show that nearby Γ\Gamma there is an isotopy Σs\Sigma_{s} between Σ\Sigma and a surface Σ1\Sigma_{1}, with the properties that for all ss, Σs\Sigma_{s} is transversal to the Reeb vector field XX and that, for the final surface, Σ1Ξ(Γ(t))\Sigma_{1}\cap\Xi(\Gamma(t)) is a leaf of Ξ(Γ(t)){\mathcal{F}}\cap\Xi(\Gamma(t)) for all tt. Then the return map is conjugated to the map that arises in the simpler situation described above.

For this, in §3.4 we compute the condition for a small cylinder, with Γ\Gamma as one boundary component, to be tangent to the Reeb vector field XX. In §3.5 we give sufficient conditions for such isotopy to give surfaces transversal to XX. Finally, in §3.6, we use the transversal approach of Σ\Sigma to Γ\Gamma to prove that the isotopy is made by surfaces transverse to XX.


3.3. Coordinates and preliminary equations.

Let (N,λ)(N,\lambda) be a 3-dimensional contact manifold, let XX be its Reeb vector field, ψt\psi_{t} its Reeb flow and ξ=kerλ\xi=\ker\lambda its contact structure. Let Σ\Sigma be a surface of section for XX and ΓΣ\Gamma\subset\partial\Sigma a rotating boundary periodic orbit. This means that there is a neighborhood UU of Γ\Gamma in Σ\Sigma such that the first arrival times of ψt\psi_{t} and ψt\psi_{-t} from UU to Σ\Sigma are bounded.

From now on we assume that the boundary periodic orbit Γ\Gamma is hyperbolic. For simplicity assume that the periodic orbit Γ\Gamma has period 11. To simplify the notation we shall also assume that Γ\Gamma has negative eigenvalues and that each local invariant manifold Wεs(Γ)W^{s}_{\varepsilon}(\Gamma), Wεu(Γ)W^{u}_{\varepsilon}(\Gamma) intersects Σ\Sigma in a neighborhood of Γ\Gamma in 3 connected components. The other cases are similar.

Since Γ\Gamma has negative eigenvalues we can choose a smooth coordinate system (x,y,z)(x,y,z) near the periodic orbit Γ\Gamma such that

Γ(t)=(0,0,t+)2×SS1/,SS1=/,(x,y,t+1)(x,y,t),\Gamma(t)=(0,0,t+{\mathbb{Z}})\in{\mathbb{R}}^{2}\times\SS^{1}/\equiv,\qquad\SS^{1}={\mathbb{R}}/{\mathbb{Z}},\quad(x,y,t+1)\equiv(-x,-y,t),
Wεu(Γ(t))\displaystyle W^{u}_{\varepsilon}(\Gamma(t)) ={(x,0,tmod 1):|x|<ε},\displaystyle=\{(x,0,t\,\text{mod }1)\,:\,|x|<\varepsilon\;\},
Wεs(Γ(t))\displaystyle W^{s}_{\varepsilon}(\Gamma(t)) ={(0,y,tmod 1):|y|<ε},\displaystyle=\{(0,y,t\,\text{mod }1)\,:\,|y|<\varepsilon\;\},

and along the periodic orbit one has

(26) eu:=x𝔼u,es:=y𝔼s,dη(eu,es)=1 and z=X,e_{u}:=\tfrac{\partial}{\partial x}\in{\mathbb{E}}^{u},\quad e_{s}:=\tfrac{\partial}{\partial y}\in{\mathbb{E}}^{s},\quad d\eta(e_{u},e_{s})=1\quad\text{ and }\quad\tfrac{\partial}{\partial z}=X,

where 𝔼s{\mathbb{E}}^{s}, 𝔼u{\mathbb{E}}^{u} are the stable and unstable subspaces for Γ\Gamma:

dψs(𝔼s,u(Γ(t)))=𝔼s,u(Γ(t+s)),dψ1|𝔼s<1,dψ1|𝔼u>1.\displaystyle d\psi_{s}({\mathbb{E}}^{s,u}(\Gamma(t)))={\mathbb{E}}^{s,u}(\Gamma(t+s)),\quad\left\|d\psi_{1}|{{\mathbb{E}}^{s}}\right\|<1,\quad\left\|d\psi_{1}|{\mathbb{E}}^{u}\right\|>1.

Consider the derivative DX(0,0,t)DX(0,0,t) of the vector field along the orbit Γ(t)\Gamma(t). Since the subspaces 𝔼s{\mathbb{E}}^{s}, 𝔼u{\mathbb{E}}^{u} are invariant under the linearized flow, we have that

(27) DX(0,0,t)=[λt000μt0000].DX(0,0,t)=\left[\begin{matrix}\lambda_{t}&0&0\\ 0&\mu_{t}&0\\ 0&0&0\end{matrix}\right].

The derivative of the flow at the periodic orbit F(t):=dφt(0,0,0)F(t):=d\varphi_{t}(0,0,0) satisfies the differential equation F˙=DX(0,0,t)F\dot{F}=DX(0,0,t)\;F. Its solution is

dφt[x0eu(τ)\displaystyle d\varphi_{t}\big{[}x_{0}\;e_{u}(\tau) +y0es(τ)+z0X(0,0,τ)]=\displaystyle+y_{0}\;e_{s}(\tau)+z_{0}\;X(0,0,\tau)\big{]}=
=x0eττ+tλs𝑑seu(t+τ)+y0eττ+tμs𝑑ses(t+τ)+z0X(0,0,t+τ).\displaystyle=x_{0}\,e^{\int_{\tau}^{\tau+t}\lambda_{s}\,ds}\;e_{u}(t+\tau)+y_{0}\,e^{\int_{\tau}^{\tau+t}\mu_{s}\,ds}\;e_{s}(t+\tau)+z_{0}\,X(0,0,t+\tau).

Since φt\varphi_{t} preserves dηd\eta, dη[dφt(eu),dφt(es)]1d\eta[d\varphi_{t}(e_{u}),d\varphi_{t}(e_{s})]\equiv 1. This implies that μt=λt\mu_{t}=-\lambda_{t} for all tt.

Since dφtd\varphi_{t} is 1-periodic in tt, the unique invariant (i.e. 1-periodic) subspace transversal to XX is given by ξ=spaneu,es\xi=\text{span}\langle e_{u},e_{s}\rangle, which necessarily coincides with the contact structure ξ=kerη\xi=\ker\eta along Γ\Gamma. Indeed, if in our coordinates the invariant transversal subspace EtE_{t} at (0,0,t)(0,0,t) is given by EtE_{t}atx+bty+ctz=0a_{t}\,x+b_{t}\,y+c_{t}\,z=0, with 1-periodic functions ata_{t}, btb_{t}, ctc_{t}. Then for any (x,y,z)(x,y,z) such that

aτx+bτy+cτz=0,a_{\tau}\,x+b_{\tau}\,y+c_{\tau}\,z=0,

we must have that d(0,0,τ)φn(x,y,z)Eτ+nd_{(0,0,\tau)}\varphi_{n}(x,y,z)\in E_{\tau+n}, i.e.

aτ+nxe0nλs𝑑s+bτ+nye0nλs𝑑s+cτ+nz=0,a_{\tau+n}\,x\,e^{\int_{0}^{n}\lambda_{s}\;ds}+b_{\tau+n}\,y\,e^{-\int_{0}^{n}\lambda_{s}\;ds}+c_{\tau+n}\,z=0,

for all nn\in{\mathbb{Z}}. Since ata_{t}, btb_{t}, ctc_{t} are 1-periodic, this implies that aτ=bτ=0a_{\tau}=b_{\tau}=0 for all τ\tau.

Thus the equation for the derivative of the flow dψt|ξd\psi_{t}|_{\xi} restricted to the contact structure ξ\xi over Γ\Gamma is

(28) w˙=A(t)w,A(t)=[λt00λt].{\dot{w}}=A(t)\cdot w,\qquad A(t)=\begin{bmatrix}\lambda_{t}&0\\ 0&-\lambda_{t}\end{bmatrix}.

We write now this differential equation in polar coordinates.

Let w(t)=(x(t),y(t))=ζ(t)uβ(t)w(t)=(x(t),y(t))=\zeta(t)\,u_{\beta(t)} be a solution of (28), where uβ=(cosβ,sinβ)u_{\beta}=(\cos\beta,\,\sin\beta). Then w(t)=(x0eΛt,y0eΛt)w(t)=(x_{0}\,e^{\Lambda_{t}},y_{0}\,e^{-\Lambda_{t}}), Λt=0tλs𝑑s\Lambda_{t}=\int_{0}^{t}\lambda_{s}\;ds, w(0)=(x0,y0)w(0)=(x_{0},y_{0}). Hence

tanβ(t)=e2Λttanβ(0).\tan\beta(t)=e^{-2\Lambda_{t}}\tan\beta(0).

Differentiating this equation we obtain

(sec2β)β˙=2λttanβ,(\sec^{2}\beta)\,{\dot{\beta}}=-2\lambda_{t}\,\tan\beta,
β˙\displaystyle{\dot{\beta}} =2λtcos2βtanβ,\displaystyle=-2\lambda_{t}\,\cos^{2}\beta\;\tan\beta,
(29) β˙\displaystyle{\dot{\beta}} =λtsin(2β).\displaystyle=-\lambda_{t}\,\sin(2\beta).

Also,

ζ(t)2\displaystyle\zeta(t)^{2} =e2Λtx02+e2Λty02\displaystyle=e^{2\Lambda_{t}}x_{0}^{2}+e^{-2\Lambda_{t}}y_{0}^{2}
2ζζ˙\displaystyle 2\,\zeta\,{\dot{\zeta}} =2λt(e2Λtx02e2Λty02)\displaystyle=2\lambda_{t}\,\big{(}e^{2\Lambda_{t}}x^{2}_{0}-e^{-2\Lambda_{t}}y^{2}_{0}\big{)}
ζζ˙\displaystyle\zeta\,{\dot{\zeta}} =λtζ2(cos2βsin2β)\displaystyle=\lambda_{t}\,\zeta^{2}\,(\cos^{2}\beta-\sin^{2}\beta)
(30) ζ˙\displaystyle{\dot{\zeta}} =λtζcos(2β).\displaystyle=\lambda_{t}\,\zeta\,\cos(2\beta).

The vector field XX satisfies X(x,y,t)=(λtx,λty, 1)+𝒪(x2+y2)X(x,y,t)=(\lambda_{t}\,x,\;-\lambda_{t}\,y,\;1)+{\mathcal{O}}(x^{2}+y^{2}). Let

γ(t)=(ρ(t)uα(t),z(t))\gamma(t)=(\rho(t)\,u_{\alpha(t)},z(t))

be a solution of γ˙=X(γ){\dot{\gamma}}=X(\gamma). Then γ˙=(ρ˙uα+ρα˙uα,z˙){\dot{\gamma}}=({\dot{\rho}}\,u_{\alpha}+\rho\,{\dot{\alpha}}\,u_{\alpha}^{\perp},\;{\dot{z}}), uα:=(sinα,cosα)u_{\alpha}^{\perp}:=(-\sin\alpha,\cos\alpha), and

ρ˙\displaystyle{\dot{\rho}} =γ˙,uα=X(γ),uα=A(t)γ,uα+𝒪(ρ2),\displaystyle=\langle{\dot{\gamma}},u_{\alpha}\rangle=\langle X(\gamma),u_{\alpha}\rangle=\langle A(t)\,\gamma,u_{\alpha}\rangle+{\mathcal{O}}(\rho^{2}),
ρα˙\displaystyle\rho\,{\dot{\alpha}} =γ˙,uα=X(γ),uα=A(t)γ,uα+𝒪(ρ2).\displaystyle=\langle{\dot{\gamma}},u_{\alpha}^{\perp}\rangle=\langle X(\gamma),u_{\alpha}^{\perp}\rangle=\langle A(t)\,\gamma,u_{\alpha}^{\perp}\rangle+{\mathcal{O}}(\rho^{2}).

Therefore, using (30) and (29),

(31) ρ˙\displaystyle{\dot{\rho}} =λtρcos(2α)+𝒪(ρ2),\displaystyle=\lambda_{t}\,\rho\,\cos(2\alpha)+{\mathcal{O}}(\rho^{2}),
ρα˙\displaystyle\rho\,{\dot{\alpha}} =λtρsin(2α)+𝒪(ρ2),\displaystyle=-\lambda_{t}\,\rho\,\sin(2\alpha)+{\mathcal{O}}(\rho^{2}),
(32) α˙\displaystyle{\dot{\alpha}} =λtsin(2α)+𝒪(ρ).\displaystyle=-\lambda_{t}\,\sin(2\alpha)+{\mathcal{O}}(\rho).

Writing X=(f1,f2,f3)X=(f_{1},f_{2},f_{3}) let Y=1f3XY=\tfrac{1}{f_{3}}\,X and let ϕt\phi_{t} be the flow of YY. Then ϕt\phi_{t} is the reparametrization of XX which preserves the solution Γ(t)\Gamma(t) and for which the foliation “z=z= constant” is invariant. Observe that from (27), f3x|(0,0,z)=f3y|(0,0,z)0\tfrac{\partial f_{3}}{\partial x}|_{(0,0,z)}=\tfrac{\partial f_{3}}{\partial y}|_{(0,0,z)}\equiv 0. The vector field YY is not a Reeb vector field of a contact form but it is smooth and along Γ\Gamma, Y=XY=X and DY=DXDY=DX. In particular, the arguments above remain valid for YY.

Consider the ϕ\phi\,-invariant foliation {\mathcal{F}} of Uε({𝟎}×SS1)U_{\varepsilon}\setminus(\{{\mathbf{0}}\}\times\SS^{1}), where

Uε:=𝔻ε×SS1,𝔻ε={z2:|z|<ε},SS1=/,U_{\varepsilon}:={\mathbb{D}}_{\varepsilon}\times\SS^{1},\quad{\mathbb{D}}_{\varepsilon}=\{z\in{\mathbb{R}}^{2}:|z|<\varepsilon\},\quad\SS^{1}={\mathbb{R}}/{\mathbb{Z}},

whose leaves are

(33) t(uαi):={ϕt(ruαi,s):r]0,δ[,sSS1}Uε,αi=π4i,i=0,1,,7,{\mathcal{F}}_{t}(u_{\alpha_{i}}):=\{\phi_{t}(r\,u_{\alpha_{i}},s):r\in]0,\delta[,\;s\in\SS^{1}\,\}\cap U_{\varepsilon},\quad\alpha_{i}=\tfrac{\pi}{4}i,\quad i=0,1,\ldots,7,

with 0<εδ10<\varepsilon\ll\delta\ll 1. This is a “radial” foliation which satisfies ϕs(t(αi))=s+t(αi)\phi_{s}({\mathcal{F}}_{t}(\alpha_{i}))={\mathcal{F}}_{s+t}(\alpha_{i}). Observe that s(u0)=t(u0)Wu(Γ){\mathcal{F}}_{s}(u_{0})={\mathcal{F}}_{t}(u_{0})\subset W^{u}(\Gamma) for all s,ts,t and also t(uπ2){\mathcal{F}}_{t}(u_{\frac{\pi}{2}})t(u3π2)Ws(Γ){\mathcal{F}}_{t}(u_{\frac{3\pi}{2}})\subset W^{s}(\Gamma) and t(uπ)Wu(Γ){\mathcal{F}}_{t}(u_{\pi})\subset W^{u}(\Gamma).

Let Σ\Sigma be a surface of section having Γ\Gamma as a rotating boundary orbit. We will construct an isotopy of ΣUε\Sigma\cap U_{\varepsilon} along surfaces Σσ\Sigma_{\sigma}, σ[0,1]\sigma\in[0,1] such that Σσint(Uε)=Γ\partial\Sigma_{\sigma}\cap\operatorname{int}(U_{\varepsilon})=\Gamma, Σ0=Σ\Sigma_{0}=\Sigma, Σσ\Sigma_{\sigma} is transversal to YY and such that for all τSS1\tau\in\SS^{1}, Σ1[z=τ]\Sigma_{1}\cap[z=\tau] is included in one leaf of {\mathcal{F}}.

3.4. The tangency condition.

Consider an annular smooth surface SS in UεU_{\varepsilon} with boundary Sint(Uε)=Γ\partial S\cap\operatorname{int}(U_{\varepsilon})=\Gamma which has a well defined limit tangent space at the points in S=Γ\partial S=\Gamma. Then, for ε>0\varepsilon>0 small enough, SUεS\cap U_{\varepsilon} is the image of a map F:[0,ε[×SS1UεF:[0,\varepsilon[\times\SS^{1}\to U_{\varepsilon}, F(r,t)=(ruθ(r,t),t)F(r,t)=(r\,u_{\theta(r,t)},t), where θ\theta is a C1C^{1} map, with continuous derivatives at r=0r=0. We obtain now the conditions for SS to be tangent to the vector field YY on r>0r>0:

3.2 Lemma.

If the surface F(r,t)=(ruθ(r,t),t)F(r,t)=(r\,u_{\theta(r,t)},t) is tangent to the vector field YY at a point (r,t)(r,t) then θ(r,t)\theta(r,t) satisfies

(34) θt=λt[sin(2θ)+rθrcos(2θ)]+𝒪(r) at (r,t).\theta_{t}=-\lambda_{t}\,\big{[}\,\sin(2\theta)+r\,\theta_{r}\,\cos(2\theta)\,\big{]}+{\mathcal{O}}(r)\qquad\text{ at }(r,t).
Proof:.

The tangent plane to SS is generated by

rFr\displaystyle r\cdot\tfrac{\partial F}{\partial r} =(ruθ+r2θruθ, 0),\displaystyle=\big{(}r\,u_{\theta}+r^{2}\,\theta_{r}\,u_{\theta}^{\perp},\;0\big{)},
Ft\displaystyle\tfrac{\partial F}{\partial t} =(rθtuθ, 1).\displaystyle=\big{(}r\,\theta_{t}\,u_{\theta}^{\perp},\;1\big{)}.

Let γ(t)=(ρ(t)uα(t),t)\gamma(t)=\big{(}\rho(t)\,u_{\alpha(t)},t\big{)} be an orbit of the flow ϕ\phi of YY. Then ρ(t)\rho(t) and α(t)\alpha(t) also satisfy equations (31) and (32). The surface SS is tangent to the vector field YY at ruθ=ρuαr\,u_{\theta}=\rho\,u_{\alpha} if and only if there exists a,ba,\,b\in{\mathbb{R}} such that

(ρ˙uα+ρα˙uα, 1)\displaystyle\big{(}{\dot{\rho}}\,u_{\alpha}+\rho\,{\dot{\alpha}}\,u_{\alpha}^{\perp}\,,\,1) =arFr+bFt\displaystyle=a\;r\,\tfrac{\partial F}{\partial r}+b\,\tfrac{\partial F}{\partial t}
=a(ruθ+r2θruθ, 0)+b(rθtuθ, 1).\displaystyle=a\,\big{(}r\,u_{\theta}+r^{2}\,\theta_{r}\,u_{\theta}^{\perp}\,,\,0)+b\,\big{(}r\,\theta_{t}\,u_{\theta}^{\perp}\,,\,1).

In this case b=1b=1 and, using (31) and (32),

(35) ρ˙=ar=λtρcos(2α)+𝒪(ρ2),\displaystyle{\dot{\rho}}=a\,r=\lambda_{t}\,\rho\,\cos(2\alpha)+{\mathcal{O}}(\rho^{2}),
(36) ρα˙=ar2θr+rθt=ρλtsin(2α)+𝒪(ρ2).\displaystyle\rho\,{\dot{\alpha}}=a\,r^{2}\,\theta_{r}+r\,\theta_{t}=-\rho\,\lambda_{t}\,\sin(2\alpha)+{\mathcal{O}}(\rho^{2}).

Since ρ=r\rho=r and α=θ\alpha=\theta, from (35) we get that a=λtcos(2θ)+𝒪(r)a=\lambda_{t}\,\cos(2\theta)+{\mathcal{O}}(r). Substituting aa in (36) we get  (34).

3.5. The isotopy.

Let F:[0,ε[×SS1UεF:[0,\varepsilon[\times\SS^{1}\to U_{\varepsilon}, F(r,t)=(ruθ(r,t),t)F(r,t)=(r\,u_{\theta(r,t)},\,t) be a local parametrization of the surface of section Σ=:Σ0\Sigma=:\Sigma_{0}. Let θ¯(t):=θ(0,t){\overline{\theta}}(t):=\theta(0,t). Let G:[0,ε[×SS1UεG:[0,\varepsilon[\times\SS^{1}\to U_{\varepsilon}, G(r,t)=(ruω(r,t),t)G(r,t)=(r\,u_{\omega(r,t)},\,t) be defined by G([0,ε[,t)𝔽(2×{t})G([0,\varepsilon[,t)\in{\mathbb{F}}\cap({\mathbb{R}}^{2}\times\{t\}), where 𝔽{\mathbb{F}} is the leaf of the foliation {\mathcal{F}} in (33) such that its tangent space at (0,0,t)(0,0,t) is E=span(0,0,1),(uθ¯(t),0)E=\text{span}\langle\,(0,0,1),\,(u_{{\overline{\theta}}(t)},0)\,\rangle.

3.3 Lemma.

Let θ,ω:[0,ε[×SS1\theta,\,\omega:[0,\varepsilon[\times\SS^{1}\to{\mathbb{R}} be of class C1C^{1}. Write θ¯(t):=θ(0,t){\overline{\theta}}(t):=\theta(0,t) and ω¯(t):=ω(0,t){\overline{\omega}}(t):=\omega(0,t).
For μ[0,1]\mu\in[0,1] write φμ(r,t):=μω(r,t)+(1μ)θ(r,t)\varphi^{\mu}(r,t):=\mu\,\omega(r,t)+(1-\mu)\,\theta(r,t). Suppose that

(37) θ¯t<λtsin(2θ¯) and ω¯(t)=θ¯(t) for all tSS1.{\overline{\theta}}_{t}<-\lambda_{t}\,\sin(2{\overline{\theta}})\quad\text{ and }\quad{\overline{\omega}}(t)={\overline{\theta}}(t)\quad\text{ for all }\;t\in\SS^{1}.

Then there is ρ0>0\rho_{0}>0 such that for all μ[0,1]\mu\in[0,1], the surface Hμ(r,t)=(ruφμ(r,t),t)H^{\mu}(r,t)=\big{(}r\,u_{\varphi^{\mu}(r,t)},t\big{)} is transversal to the vector field YY at all points (r,t)(r,t) with 0<r<ρ00<r<\rho_{0}, tSS1t\in\SS^{1}.

Proof:.

Since SS1\SS^{1} is compact, there exists ε>0\varepsilon>0 such that

θ¯t<λtsin(2θ¯)3ε for all tSS1.{\overline{\theta}}_{t}<-\lambda_{t}\sin(2{\overline{\theta}})-3\varepsilon\quad\text{ for all }\;t\in\SS^{1}.

Choose ρ1>0\rho_{1}>0 such that for all 0r<ρ10\leq r<\rho_{1} and tSS1t\in\SS^{1},

|λt|r|θr|<ε4,|λt|r|ωr|<ε4,\displaystyle|\lambda_{t}|\,r\,|\theta_{r}|<\tfrac{\varepsilon}{4}\,,\quad|\lambda_{t}|\,r\,|\omega_{r}|<\tfrac{\varepsilon}{4}\,,
|λt||θ¯(t)θ(r,t)|<ε16,|λt||θ¯(t)ω(r,t)|<ε16 and \displaystyle|\lambda_{t}|\,|{\overline{\theta}}(t)-\theta(r,t)|<\tfrac{\varepsilon}{16}\,,\quad|\lambda_{t}|\,|{\overline{\theta}}(t)-\omega(r,t)|<\tfrac{\varepsilon}{16}\quad\text{ and }
θt<λt[sin(2θ)+rθrcos(2θ)]2ε,\displaystyle\theta_{t}<-\lambda_{t}\,\big{[}\sin(2\theta)+r\,\theta_{r}\cos(2\theta)\big{]}-2\varepsilon,
ωt<λt[sin(2ω)+rωrcos(2ω)]2ε.\displaystyle\omega_{t}<-\lambda_{t}\,\big{[}\sin(2\omega)+r\,\omega_{r}\cos(2\omega)\big{]}-2\varepsilon.

Then

|λt||φμθ¯|<ε16,|λt||φμθ|<ε8,|λt||φμω|<ε8,\displaystyle|\lambda_{t}|\,|\varphi^{\mu}-{\overline{\theta}}|<\tfrac{\varepsilon}{16},\quad|\lambda_{t}|\,|\varphi^{\mu}-\theta|<\tfrac{\varepsilon}{8},\quad|\lambda_{t}|\,|\varphi^{\mu}-\omega|<\tfrac{\varepsilon}{8},
|λt||sin(2φμ)sin(2θ¯)|<ε8 and\displaystyle|\lambda_{t}|\,|\sin(2\varphi^{\mu})-\sin(2{\overline{\theta}})|<\tfrac{\varepsilon}{8}\quad\text{ and}
|λt||sin(2φμ)sin(2θ)|<ε4,|λt||sin(2φμ)sin(2ω)|<ε4.\displaystyle|\lambda_{t}|\,|\sin(2\varphi^{\mu})-\sin(2\theta)|<\tfrac{\varepsilon}{4},\quad|\lambda_{t}|\,|\sin(2\varphi^{\mu})-\sin(2\omega)|<\tfrac{\varepsilon}{4}.

Hence

|λt||sin(2φμ)[μsin(2ω)+(1μ)sin(2θ)]|<ε4.|\lambda_{t}|\,\big{|}\sin(2\varphi^{\mu})-[\mu\,\sin(2\omega)+(1-\mu)\,\sin(2\theta)]\big{|}<\tfrac{\varepsilon}{4}.

Also, for all (r,t)[0,ρ1[×SS1(r,t)\in[0,\rho_{1}[\times\SS^{1} and μ[0,1]\mu\in[0,1], since φrμ=μωr+(1μ)θr\varphi^{\mu}_{r}=\mu\,\omega_{r}+(1-\mu)\,\theta_{r}, we have that

|λtrθrcos(2θ)|<ε4,|λtrωrcos(2ω)|<ε4 and |λtrφrμcos(2φμ)|<ε4.|\lambda_{t}\,r\,\theta_{r}\,\cos(2\theta)|<\tfrac{\varepsilon}{4},\quad|\lambda_{t}\,r\,\omega_{r}\,\cos(2\omega)|<\tfrac{\varepsilon}{4}\quad\text{ and }\quad|\lambda_{t}\,r\,\varphi^{\mu}_{r}\cos(2\varphi^{\mu})|<\tfrac{\varepsilon}{4}.

Thus, we have that

φtμ\displaystyle\varphi^{\mu}_{t} =μωt+(1μ)θt\displaystyle=\mu\,\omega_{t}+(1-\mu)\,\theta_{t}
<2ελt[μsin(2ω)+(1μ)sin(2θ)]μλtrωrcos(2ω)(1μ)λtrθrcos(2θ)\displaystyle<-2\varepsilon-\lambda_{t}\,\big{[}\mu\,\sin(2\omega)+(1-\mu)\,\sin(2\theta)\big{]}-\mu\,\lambda_{t}\,r\,\omega_{r}\cos(2\omega)-(1-\mu)\,\lambda_{t}\,r\,\theta_{r}\cos(2\theta)
<2ε+ε4λtsin(2φμ)+ε4+ε4\displaystyle<-2\varepsilon+\tfrac{\varepsilon}{4}-\lambda_{t}\,\sin(2\varphi^{\mu})+\tfrac{\varepsilon}{4}+\tfrac{\varepsilon}{4}
<2ε+ε4λtsin(2φμ)+ε2λtrφrμcos(2φμ)+ε4\displaystyle<-2\varepsilon+\tfrac{\varepsilon}{4}-\lambda_{t}\,\sin(2\varphi^{\mu})+\tfrac{\varepsilon}{2}-\lambda_{t}\,r\,\varphi^{\mu}_{r}\,\cos(2\varphi^{\mu})+\tfrac{\varepsilon}{4}
<λt[sin(2φμ)+rφrμcos(2φμ)]ε.\displaystyle<-\lambda_{t}\big{[}\sin(2\varphi^{\mu})+r\,\varphi^{\mu}_{r}\,\cos(2\varphi^{\mu})\big{]}-\varepsilon.

Then there is 0<ρ0<ρ10<\rho_{0}<\rho_{1} such that if 0<r<ρ00<r<\rho_{0},

φtμ<λt[sin(2φμ)+rφrμcos(2φμ)]+𝒪(r),\varphi^{\mu}_{t}<-\lambda_{t}\big{[}\sin(2\varphi^{\mu})+r\,\varphi^{\mu}_{r}\,\cos(2\varphi^{\mu})\big{]}+{\mathcal{O}}(r),

where 𝒪(r){\mathcal{O}}(r) is from lemma 3.2. Then lemma 3.2 finishes the proof. ∎

3.6. The transversality condition.

The equation for the dynamics under dψtd\psi_{t} of subspaces along the periodic orbit Γ(t)\Gamma(t) is (29). Then equation (37) just says that the limit direction θ¯{\overline{\theta}} of the surface of section at Γ\Gamma turns more slowly than its iteration under the linearized Reeb flow. Condition 2.4.(iii) in the definition of complete systems implies (37).

We check in §3.8 equation (37) for surfaces of section which are projections of pseudo holomorphic curves in a symplectization, and in §3.7 for Birkhoff annular surfaces of section. The surfaces of section that we use in theorem 1.3 are obtained by topological surgery from Birkhoff annuli. These surgeries mantain inequality (37).

Since a surface of section is transversal to the Reeb flow in its interior, a weak inequality equation (37) must hold at a rotating boundary orbit. If needed one can modify the surface nearby its boundary binding orbit so that the asymptotic rotation of the surface in the boundary is uniform with respect to the rotation of the flow, satisfying (37).

3.7. The transversality condition for a Birkhoff annulus.

Let MM be a closed oriented riemannian surface, SMSM its unit tangent bundle and let ϕt:SM\phi_{t}:SM\hookleftarrow be the geodesic flow of MM. Let λ\lambda be the Liouville form on TMTM:

λ(x,v)(ξ)=v,dπ(ξ)x,(x,v)TM,ξT(x,v)TM.\lambda_{(x,v)}(\xi)=\langle v,d\pi(\xi)\rangle_{x},\qquad(x,v)\in TM,\quad\xi\in T_{(x,v)}TM.

Let 𝕍=kerdπ{\mathbb{V}}=\ker d\pi, and =kerK{\mathbb{H}}=\ker K be the vertical and horizontal subspaces, where K:TTMTMK:TTM\to TM is the connection. The subbundle 𝒩=kerλ{\mathcal{N}}=\ker\lambda of of T(SM)T(SM),

(38) 𝒩=kerλ={(h,w)TθSM𝕍|h,θπ(θ)=0},{\mathcal{N}}=\ker\lambda=\{\,(h,w)\in T_{\theta}SM\subset{\mathbb{H}}\oplus{\mathbb{V}}\;|\;\langle h,\theta\rangle_{\pi(\theta)}=0\;\},

is invariant under the linearized geodesic flow dϕtd\phi_{t}, which is given by

dϕt(J(0),J˙(0))=(J(t),J˙(t))𝒩𝕍,d\phi_{t}(J(0),\dot{J}(0))=(J(t),\dot{J}(t))\in{\mathcal{N}}\subset{\mathbb{H}}\oplus{\mathbb{V}},

where tJ(t)t\mapsto J(t) is a Jacobi field along a geodesic c(t)=πJ(t)c(t)=\pi J(t) which is orthogonal to c˙(t)\dot{c}(t). The tangent space to the unit tangent bundle SMSM is

TθSM=𝕏(θ)𝒩(θ),T_{\theta}SM=\langle{\mathbb{X}}(\theta)\rangle\oplus{\mathcal{N}}(\theta),

where 𝕏{\mathbb{X}} is the geodesic vector field.

Let 𝒥:TxM{\mathcal{J}}:T_{x}M\hookleftarrow be the rotation of angle +π2+\frac{\pi}{2}. Given a simple closed geodesic γ(t)\gamma(t) parametrized with unit speed, define its Birkhoff annulus by

A(γ˙):={(x,v):t,x=γ(t),v,𝒥γ˙(t)γ(t)0}A({\dot{\gamma}}):=\{\,(x,v):\exists t,\;x=\gamma(t),\;\langle v,{\mathcal{J}}{\dot{\gamma}}(t)\rangle_{\gamma(t)}\geq 0\,\}

The interior of A(γ˙)A({\dot{\gamma}}) is tranversal to the geodesic flow. The tangent space of the Birkhoff annulus at a boundary point ±γ˙(t)\pm{\dot{\gamma}}(t), is generated by the geodesic vector field 𝕏(γ˙(t)){\mathbb{X}}({\dot{\gamma}}(t)) and the vertical direction 𝕍𝒩{\mathbb{V}}\cap{\mathcal{N}}.

In order to obtain the transversality condition θ¯t<λtsin(2θ¯){\overline{\theta}}_{t}<-\lambda_{t}\sin(2{\overline{\theta}}) it is enough to show that the (vertical) limit tangent space of the Birkhoff annulus moves slower than the movement of the vertical subspace under the derivative of the flow. This is done as follows:

Let J(t)Tγ(t)MJ(t)\in T_{\gamma(t)}M be an orthogonal Jacobi field. Since both JJ and J˙\dot{J} are multiples of the orthogonal vector γ˙(t){\dot{\gamma}}(t)^{\perp} they can be regarded as scalar quantities. When (J,J˙)(J,\dot{J}) is not horizontal, i.e. when J˙(t)0\dot{J}(t)\neq 0, define W(t)=J(t)/J˙(t)W(t)=J(t)/\dot{J}(t). From the Jacobi equation

J¨+KJ=0andJ=WJ˙,\ddot{J}+K\,J=0\qquad\text{and}\qquad J=W\,\dot{J},

we get

J˙=W˙J˙WKJ.\dot{J}=\dot{W}\dot{J}-WKJ.

Replacing J=WJ˙J=W\dot{J} when J˙0\dot{J}\neq 0 one obtains the Riccati equation

(39) W˙=KW2+1.\dot{W}=KW^{2}+1.

A solution W(t)W(t) is the slope of the iteration under dϕtd\phi_{t} of a linear subspace, i.e. if

𝒲0=graphW(0)={(W(0)v,v)𝕍|v}{\mathcal{W}}_{0}=\operatorname{graph}W(0)=\{(W(0)v,v)\in{\mathbb{H}}\oplus{\mathbb{V}}\,|\,v\in{\mathbb{H}}\}

then dϕt(𝒲0)=graph(W(t))d\phi_{t}({\mathcal{W}}_{0})=\operatorname{graph}(W(t)). The subspace 𝒲0{\mathcal{W}}_{0} is the vertical subspace 𝕍{\mathbb{V}} precisely when W(0)=0W(0)=0. In this case, from (39) we have that W˙(0)=1\dot{W}(0)=1. If V(t)V(t) is the slope of the vertical subspace 𝒩𝕍{\mathcal{N}}\cap{\mathbb{V}}, then V(t)0V(t)\equiv 0 and V˙(0)=0\dot{V}(0)=0. This means that the iteration W(t)W(t) of the vertical subspace under the linearized geodesic flow moves faster than the vertical subspace V(t)V(t) (tangent to the Birkhoff annulus).

3.8. The transversality condition for finite energy surfaces.

In this section we prove that condition (37) holds for projections on SS3\SS^{3} of pseudo holomorphic curves in the simplectization of a tight contact form on SS3\SS^{3}. Then we can apply item (3) of theorem B to the complete system of surfaces of section of genus 0 obtained by Hofer, Wysocki, Zehnder in [21, Cor. 1.8], in order to obtain Corollary 1.1.

In this case the complete system is given by the rigid surfaces of the finite energy foliation. Let Σ\Sigma be a rigid surface and let Γ\Gamma be a boundary periodic orbit of Σ\Sigma where the foliation is radial. The equation for the dynamics under dψtd\psi_{t} of subspaces along a periodic orbit Γ(t)\Gamma(t) is (29). So we want to prove that the limit direction θ¯{\overline{\theta}} of the surface of section Σ\Sigma at Γ\Gamma turns slower than its iteration under the linearized Reeb flow.

Recall that the contact structure ξ\xi is invariant under the Reeb flow ψt\psi_{t}. The linearized Reeb flow on ξ\xi satisfies v(t)=dψt(v(0))v(t)=d\psi_{t}(v(0)) where v(0)ξv(0)\in\xi and

(40) v˙=DX(ψt(Γ(0)))v=S(t)v.\dot{v}=DX(\psi_{t}(\Gamma(0)))\cdot v=S(t)\,v.

Here the matrix S(t)=DX(Γ(t))S(t)=DX(\Gamma(t)) is symmetric on symplectic linear coordinates in ξ(Γ(t))\xi(\Gamma(t)) and v(t)=ζ(t)uβ(t)v(t)=\zeta(t)\,u_{\beta(t)} satisfies (29) and (30).

From theorem 1.4 in [20] (where S=J0S(t)S_{\infty}=-J_{0}\,S(t) and J0=J|ξJ_{0}=J|_{\xi}), there is a periodic vector e(t)=ε(t)uθ¯(t)ξ(Γ(t))e(t)=\varepsilon(t)\,u_{{\overline{\theta}}(t)}\in\xi(\Gamma(t)) in the asymptotic direction of the rigid surface Σ\Sigma which satisfies the (eigenvalue) equation

(41) e˙=S(t)e(t)+μJe\dot{e}=S(t)\,e(t)+\mu\,Je

with μ<0\mu<0 and J:ξJ:\xi\hookleftarrow an almost complex structure on ξ\xi.

Our choice of coordinates (26) about Γ(t)\Gamma(t) is symplectic and the almost complex structure can be taken J(x,y)=(y,x)J(x,y)=(-y,x) in these coordinates. Comparing equations (40) and (41) at an initial condition for v(t)v(t) such that v(t0)=e(t0)v(t_{0})=e(t_{0}) we get that the rigid surface Σ\Sigma turns slower than the linearized flow.

Indeed, in polar coordinates ruβ=v(t0)=e(t0)=εuθ¯r\,u_{\beta}=v(t_{0})=e(t_{0})=\varepsilon\,u_{\overline{\theta}}, from (40), at t=t0t=t_{0} we have that

S(t0)e\displaystyle S(t_{0})\,e =v˙=r˙uβ+rβ˙uβ.\displaystyle=\dot{v}=\dot{r}\,u_{\beta}+r\,\dot{\beta}\,u_{\beta}^{\perp}.
e˙\displaystyle\dot{e} =ε˙uθ¯+εθ¯˙uθ¯,\displaystyle=\dot{\varepsilon}\,u_{\overline{\theta}}+\varepsilon\,\dot{\overline{\theta}}\,u_{\overline{\theta}}^{\perp},
=S(t0)e+μJe,\displaystyle=S(t_{0})\,e+\mu\,Je,

where Je=εuθ¯Je=\varepsilon\,u_{\overline{\theta}}^{\perp}. In the component uθ¯u_{\overline{\theta}}^{\perp} these equations are

εθ¯˙=εβ˙+με.\varepsilon\,\dot{\overline{\theta}}=\varepsilon\,\dot{\beta}+\mu\,\varepsilon.

From (29), β˙=λtsin2β=λtsin2θ¯\dot{\beta}=-\lambda_{t}\,\sin 2\beta=-\lambda_{t}\,\sin 2{\overline{\theta}}. Therefore

θ¯˙=λtsin2θ¯+μ,\dot{\overline{\theta}}=-\lambda_{t}\,\sin 2{\overline{\theta}}+\mu,

with the (constant) eigenvalue μ<0\mu<0. This implies (37).

3.9. The return map.

Lemma 3.3 gives an isotopy of the local surface of section Σ0\Sigma_{0} by surfaces Σμ\Sigma_{\mu}, μ[0,1]\mu\in[0,1], which are transversal to the vector field. Then the return map of the Reeb flow to the final surface Σ1\Sigma_{1} is topologically conjugate to the return map to the surface Σ0\Sigma_{0} in a neighbourhood of its boundary. Moreover, the intersections Σ1(2×{t})\Sigma_{1}\cap({\mathbb{R}}^{2}\times\{t\}) are included in a leaf of the radial invariant foliation {\mathcal{F}}. This implies that the return map to Σ1\Sigma_{1} near the boundary Σ1\partial\Sigma_{1} preserves the foliation of Σ1\Sigma_{1} given by the sections Σ1(2×{t})tSS1\langle\Sigma_{1}\cap({\mathbb{R}}^{2}\times\{t\})\rangle_{t\in\SS^{1}}. The surface Σ1\Sigma_{1} is parametrized by

G(r,t)=(ruω(r,t),t).G(r,t)=(r\,u_{\omega(r,t)},t).

At its boundary points (0,t)(0,t), the surface Σ1\Sigma_{1} has a well defined tangent plane generated by (uω(0,t),0)(u_{\omega(0,t)},0) and the Reeb vector field X=(0,0,1)X=(0,0,1). Here ω(0,t)=ω¯(t)=θ¯(t)=θ(0,t)\omega(0,t)={\overline{\omega}}(t)={\overline{\theta}}(t)=\theta(0,t) is the same angular approach of the surface Σ0\Sigma_{0}, which by §3.6 satisfies

θ¯t<λtsin2θ¯.{\overline{\theta}}_{t}<-\lambda_{t}\,\sin 2{\overline{\theta}}.

Let P:Σ1Σ1P:\Sigma_{1}\to\Sigma_{1} be the first return map to Σ1\Sigma_{1}. Then in coordinates (r,t)(r,t) given by the parametrization G(r,t)G(r,t) we have that

(42) P(r,t)=(a(r,t),b(t)).P(r,t)=\big{(}a(r,t),\,b(t)\big{)}.

Here b(t)b(t) is given by the time in which the leaf G([0,ε[,t)G([0,\varepsilon[,t) returns to Σ1\Sigma_{1}. This is the same as the time in which the derivative of the flow sends the tangent subspace T(0,t)Σ1T_{(0,t)}\Sigma_{1} to T(0,b(t))Σ1T_{(0,b(t))}\Sigma_{1}. The equations for the derivative of the flow in polar coordinates are (29) and  (30). Then b(t)b(t) is determined by the minimal b(t)>tb(t)>t satisfying

(43) β(t)=θ¯(t),β(b(t))=θ¯(b(t)),β˙=λtsin(2β).\displaystyle\beta(t)={\overline{\theta}}(t),\qquad\beta(b(t))={\overline{\theta}}(b(t)),\qquad{\dot{\beta}}=-\lambda_{t}\,\sin(2\beta).

This is the first return map QQ of the flow t(t,β(t))t\mapsto(t,\beta(t)) of the differential equation (29) to the graph of θ¯{\overline{\theta}} (see figures 6, 6).

Refer to caption
Figure 5. This figure shows the flow of the differential equation (29) for β(t)\beta(t) describing the action of the derivative of the Reeb flow on 2-planes tangent to the periodic orbit γ\gamma. It also shows the curve (t,θ¯(t))(t,{\overline{\theta}}(t)), which corresponds to the movement of the limit tangent plane of the surface of section Σ\Sigma along its rotating boundary periodic orbit γ\gamma. The graph of θ¯{\overline{\theta}} is transversal to the flow of β\beta and θ¯(t+1)=θ¯(t)+3π{\overline{\theta}}(t+1)={\overline{\theta}}(t)+{3\pi}, but θ(t)¯\overline{\theta(t)} may not be monotonous.
Observe that the first return map of the flow of (29) to the graph of θ¯{\overline{\theta}} has repelling periodic points at θ¯=0,π,2π{\overline{\theta}}=0,\pi,2\pi and attracting periodic points at θ¯=π2,π2,3π2{\overline{\theta}}=-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2}.
Refer to caption
Refer to caption
Figure 6. The figure at the left shows that the points (t,θ¯)(t,{\overline{\theta}}) in the graph 𝒢(θ¯){\mathcal{G}}({\overline{\theta}}) of θ¯{\overline{\theta}} with θ¯π2(mod 3π){\overline{\theta}}\notin\frac{\pi}{2}{\mathbb{Z}}~{}(\text{mod }3\pi) are wandering under the return map. The figure at the right represents the dynamics of the return map tb(t)(mod 1)t\mapsto b(t)~{}(\text{mod }1).
Refer to caption
Figure 7. The dynamics at the boundary.

The figure shows the dynamics of the extension to Σ\partial\Sigma of the return map PP to the surface of section Σ\Sigma in a neighborhood of a rotating boundary component ΓΣ\Gamma\subset\partial\Sigma, when the periodic orbit Γ\Gamma is hyperbolic. Here the map PP has two periodic points in Γ=Σ\Gamma=\partial\Sigma of period 3 which are saddles on Σ\Sigma. The periodic points correspond to the times in which the stable and unstable subspaces intersect the tangent space of the section Σ\Sigma at its boundary.

The vertical axis is the time parameter and the periodic orbit Γ\Gamma which is supposed to have period 1. The three shadowed rectangles are copies of the 2-torus formed by the periodic orbit (the time circle) and the one dimensional subspaces orthogonal to the periodic orbit, parametrized by their angle with one branch of the stable subspace. The movement of the stable subspace 𝔼s{\mathbb{E}}^{s} is represented by the angles 0 and π\pi and the unstable subspace 𝔼u{\mathbb{E}}^{u} by the angles π2\frac{\pi}{2} and 3π2\frac{3\pi}{2}.

The periodic orbit has negative eigenvalues, then after one period the normal subspaces are identified by a rotation of angle π\pi. For example, the 0 branch of the stable subspace 𝔼s{\mathbb{E}}^{s} is identified with the π\pi branch of 𝔼s{\mathbb{E}}^{s}. This can be seen in the picture as a shift of length π\pi in the second shadowed square. The black lines are the dynamics of linear subspaces orthogonal to Γ\Gamma under the derivative of the flow, we will call it the projective flow. The subspaces converge to the unstable subspace 𝔼u{\mathbb{E}}^{u} in the future and to 𝔼s{\mathbb{E}}^{s} in the past.

The transversal lines are the graph of the asymptotic limit θ¯(t)\overline{\theta}(t) of the surface of section Σ\Sigma. We have assumed that this graph intersects three times 𝔼s{\mathbb{E}}^{s}, 𝔼u{\mathbb{E}}^{u} in one period. The dynamics of the extension to the boundary in this figure is given by the return map of the projective flow in the figure, to the graph of the asymptotic direction θ¯(t)\overline{\theta}(t). The periodic orbit corresponding to the unstable subspace 𝔼u{\mathbb{E}}^{u} is shown in the figure with the numbers 1,2,31,~{}2,~{}3, in the order of the orbit. It is an attracting periodic orbit, and the stable subspace in Σ\partial\Sigma gives a repelling periodic orbit. The extension of the return map to Σ\partial\Sigma is Morse Smale.

The graph of θ¯{\overline{\theta}}, 𝒢(θ¯)={(t,θ¯(t))|tSS1}{\mathcal{G}}({\overline{\theta}})=\{\,(t,{\overline{\theta}}(t))\,|\,t\in\SS^{1}\,\}, is transversal to the flow lines of  (29). We are assuming that the angle θ¯(t)\overline{\theta}(t) of Σ\Sigma turns 3π3\pi in one period t[0,1]t\in[0,1]. The return map Q:𝒢(θ¯)𝒢(θ¯)Q:{\mathcal{G}}({\overline{\theta}})\to{\mathcal{G}}({\overline{\theta}}) under the flow of the differential equation (43) for β\beta is the continuous extension of the return map of ψt\psi_{t} to Σ1\Sigma_{1} to the boundary [r=0]𝒢(θ¯)/3πΣ1[r=0]\approx{\mathcal{G}}({\overline{\theta}})\approx{\mathbb{R}}/3\pi{\mathbb{Z}}\subset\partial\Sigma_{1}. The return map to 𝒢(θ¯){\mathcal{G}}({\overline{\theta}}) has periodic points at θ¯=0,π,2π{\overline{\theta}}=0,\pi,2\pi and at θ¯=π2,π2,3π2{\overline{\theta}}=-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2}. The periodic orbit at θ¯=0,π,2π{\overline{\theta}}=0,\pi,2\pi is a repellor and the periodic orbit at at θ¯=π2,π2,3π2{\overline{\theta}}=-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2} is an attractor. Lemma 3.4 shows that there are no other periodic points for the return map to 𝒢(θ¯){\mathcal{G}}({\overline{\theta}}).

3.4 Lemma.

There are no periodic points (τ,θ¯(τ))(\tau,{\overline{\theta}}(\tau)) for the return map Q:𝒢(θ¯)𝒢(θ¯)Q:{\mathcal{G}}({\overline{\theta}})\to{\mathcal{G}}({\overline{\theta}})

Q(t,θ¯(t))=(b(t),θ¯(b(t))),t/,θ¯/3πQ(t,{\overline{\theta}}(t))=\big{(}b(t),{\overline{\theta}}(b(t))\big{)},\quad t\in{\mathbb{R}}/{\mathbb{Z}},\quad\overline{\theta}\in{\mathbb{R}}/3\pi{\mathbb{Z}}

with θ¯(τ)π2{\overline{\theta}}(\tau)\notin\frac{\pi}{2}\,{\mathbb{Z}}. The periodic orbit θ¯=0,π,2π{\overline{\theta}}=0,\pi,2\pi is a repellor and the periodic orbit θ¯=π2,π2,3π2{\overline{\theta}}=-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2} is an attractor.

Proof:.

Observe that since tλtt\mapsto\lambda_{t} is a 1-periodic function, the equation (29) defines a 1-periodic flow ϕt(s,β(s))=(s+t,β(s+t))\phi_{t}(s,\beta(s))=(s+t,\beta(s+t)) on ×/2π{\mathbb{R}}\times{\mathbb{R}}/2\pi{\mathbb{Z}}.

Since the line θ¯=π2{\overline{\theta}}=\frac{\pi}{2} corresponds to the unstable subspace 𝔼u{\mathbb{E}}^{u} of Γ\Gamma and tβ(t)t\mapsto\beta(t) describes the dynamics of the linearized flow on 1-dimensional subspaces along Γ\Gamma, we have that

(44) 0<β(0)<π\displaystyle 0<\beta(0)<\pi limt+β(t)=π2,\displaystyle\Longrightarrow\quad\lim_{t\to+\infty}\beta(t)=\tfrac{\pi}{2},
π<β(0)<2π\displaystyle\pi<\beta(0)<2\pi limt+β(t)=3π2.\displaystyle\Longrightarrow\quad\lim_{t\to+\infty}\beta(t)=\tfrac{3\pi}{2}.

This implies that the periodic orbit θ¯=π2,π2,3π2{\overline{\theta}}=-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2} for QQ is an attractor. Similarly, the periodic orbit θ¯=0,π,2π{\overline{\theta}}=0,\pi,2\pi is a repellor because it corresponds to the stable subspace 𝔼s{\mathbb{E}}^{s}.

Suppose that there is a periodic point (τ,θ¯(τ))(\tau,{\overline{\theta}}(\tau)) of the return map QQ with θ¯(τ)π2{\overline{\theta}}(\tau)\notin-\frac{\pi}{2}{\mathbb{Z}}. Then there is n+n\in{\mathbb{Z}}^{+} such that Q(τ,θ¯(τ))=(τ+n,θ¯(τ))Q(\tau,{\overline{\theta}}(\tau))=(\tau+n,{\overline{\theta}}(\tau)). The solution β\beta of (29) with β(τ)=θ¯(τ)\beta(\tau)={\overline{\theta}}(\tau) satisfies β(τ+n)=θ¯(τ)\beta(\tau+n)={\overline{\theta}}(\tau) and hence it is nn-periodic. This contradicts (44). ∎

Refer to caption
Figure 8. The figure at the left shows the dynamics of the return map PP to the surface of section Σ1\Sigma_{1}. The figure at the right illustrates the construction of the conjugacy to a hyperbolic periodic point.
3.5 Proposition.

The periodic points at a hyperbolic rotating boundary of a surface of section are saddles for the return map.

Proof:.

In a neighbourhood of the periodic orbit Γ\Gamma, the foliation of the surface Σ1\Sigma_{1} whose leaves are G(]0,ε[,t)G(]0,\varepsilon[,t), t[0,1]t\in[0,1] is invariant under the return map P:Σ1Σ1P:\Sigma_{1}\to\Sigma_{1}. Let u,s[0,1]u,s\in[0,1] be such that θ¯(u){0,π}{\overline{\theta}}(u)\in\{0,\pi\} and θ¯(s){π2,3π2}{\overline{\theta}}(s)\in\{\tfrac{\pi}{2},\frac{3\pi}{2}\}. Then G(]0,ε[,u)G(]0,\varepsilon[,u) and G(]0,ε[,s)G(]0,\varepsilon[,s) are components of Wu(γ)Σ1W^{u}(\gamma)\cap\Sigma_{1} and Ws(γ)Σ1W^{s}(\gamma)\cap\Sigma_{1} respectively. For such uu’s, using formula (42), the third iterate P3P^{3} of return map ra(r,u)[0,ε[r\mapsto a(r,u)\in[0,\varepsilon[, which is the dynamics in Wu(γ)Σ1W^{u}(\gamma)\cap\Sigma_{1}, is expanding with fixed point r=0r=0 and on the components of Ws(γ)Σ1W^{s}(\gamma)\cap\Sigma_{1} it is a contraction with fixed point at the boundary of Σ1\Sigma_{1}.

Refer to caption
Figure 9. The figure shows that the iteration P3n(α)P^{-3n}(\alpha) by the return map to the section Σ1\Sigma_{1} of a curve αΣ1\alpha\subset\Sigma_{1} which is transversal to the intersection Σ1Wu(γ)\Sigma_{1}\cap W^{u}(\gamma). The backward iteration of α\alpha under a long time φ]T,T+ε[(α)\varphi_{]-T,-T+\varepsilon[}(\alpha) is a surface which approaches the stable manifold Ws(γ)W^{s}(\gamma). Its intersection with Σ1\Sigma_{1} is the return P3n(α)P^{-3n}(\alpha), which in a small neighborhood of Σ1Wu(γ)\Sigma_{1}\cap W^{u}(\gamma) converges to the boundary Σ1=γ\partial\Sigma_{1}=\gamma.

Consider a small curve α\alpha transversal to Wu(γ)Σ1W^{u}(\gamma)\cap\Sigma_{1} as in figure 9. The inverse image P3(α)P^{-3}(\alpha) intersects a larger set of leaves of the PP-invariant foliation 𝔽=Σ1{\mathbb{F}}={\mathcal{F}}\cap\Sigma_{1}, this depends only on the dynamics of b(t)b(t). Extend {α,P3(α)}\{\alpha,\,P^{-3}(\alpha)\} to a 1-dimensional foliation 𝔸{\mathbb{A}} on Σ1\Sigma_{1} between α\alpha and P3(α)P^{-3}(\alpha). By the λ\lambda-lemma, the backward flow ϕT(ϕ[0,ε](α))\phi_{-T}(\phi_{[0,\varepsilon]}(\alpha)) of ϕ[0,ε](α)\phi_{[0,\varepsilon]}(\alpha) approaches in the C1C^{1} topology to the stable manifold Ws(γ)W^{s}(\gamma). The intersection of ϕT(ϕ[0,ε](α))\phi_{-T}(\phi_{[0,\varepsilon]}(\alpha)) with Σ1\Sigma_{1} are leaves of P3n(𝔸)P^{-3n}({\mathbb{A}}), which approach the boundary of Σ\Sigma. Extend the foliation by iteration to a neighbourhood nP3n(𝔸)\cup_{n\in{\mathbb{N}}}P^{-3n}({\mathbb{A}}) of the fixed point at the boundary r=0r=0, t=ut=u. Use the foliations 𝔽{\mathbb{F}} and 𝔸{\mathbb{A}} as in figure 8, to construct a coordinate system in a neighbourhood of the fixed point r=0r=0, t=ut=u which conjugates the dynamics to two sectors of a saddle fixed point. A similar construction can be made in a neighbourhood of the periodic points r=0r=0, t=st=s.

4. The complete system for geodesic flows.

Refer to caption
Figure 10. The minimizing geodesics in the homotopy classes of γi\gamma_{i} separate MM into four simply connected regions RjR_{j} and the smoothing of their Birkhoff annuli 𝒯=R1R3,𝒯=R2R4,{\mathcal{T}}=\partial R_{1}\cup\partial R_{3},\qquad-{\mathcal{T}}=\partial R_{2}\cup\partial R_{4}, are two embedded surfaces of section of genus 1. The orbits γ˙i{\dot{\gamma}}_{i}, γ˙i-{\dot{\gamma}}_{i} are simply covered rotating boundary orbits for 𝒯{\mathcal{T}} and 𝒯-{\mathcal{T}}.

The set of ideas in this section descent from G. Birkhoff, notably [2] section §28. By using an orientable double cover of MM if necessary, for theorem A it is enough to assume that the surface MM is orientable.

We denote SM={(x,v)TM:ρ(v,v)=1}SM=\{(x,v)\in TM:\rho(v,v)=1\} the unit tangent bundle, π:SMM\pi:SM\to M the projection, ϕt\phi_{t} the geodesic flow on SMSM and SA=π1(A)SMSA=\pi^{-1}(A)\cap SM for every AMA\subset M.

4.1. Two surfaces of section of genus 1.

Let γ1,,γ2g+2\gamma_{1},\ldots,\gamma_{2g+2} be minimizing geodesics in the homotopy classes of the curves shown in figure 10. We show now that they divide the surface MM into four regions R1,,R4R_{1},\ldots,R_{4} which are simply connected.

A bigon is a simply connected open subset of MM whose boundary is two geodesic segments. Two minimizing geodesics in their homotopy classes can not form a bigon. Then they must have minimal intersection number in their homotopy classes c.f. [13, Prop. 1.7]. Therefore

|#(γiγj)|=δi,j1+δi,j+1 if ij.|\#(\gamma_{i}\cap\gamma_{j})|=\delta_{i,j-1}+\delta_{i,j+1}\quad\text{ if }\quad i\neq j.

Now M(γ1γ3γ2g+1)M\setminus(\gamma_{1}\cup\gamma_{3}\cup\cdots\cup\gamma_{2g+1}) is the union of two surfaces N1N_{1}, N2N_{2} of genus zero with 2g+12g+1 boundary components. The segments γ2iNj\gamma_{2i}\cap N_{j} are curves connecting the boundary components γ2i1\gamma_{2i-1} and γ2i+1\gamma_{2i+1}. They form two simple closed curves bounding two simply connected regions R2j1R_{2j-1}, R2jR_{2j}.

Let 𝒥:TMTM{\mathcal{J}}:TM\to TM be a linear map such that (v,𝒥v)(v,{\mathcal{J}}v) is an oriented orthonormal basis for every unit vector vv. Given an oriented simple closed geodesic γ\gamma, define the Brikhoff annulus of γ˙{\dot{\gamma}} as

A(γ˙):={(x,v)SM|t,x=γ(t),v,𝒥γ˙(t)0}.A({\dot{\gamma}}):=\{(x,v)\in SM\;|\;\exists t,\;x=\gamma(t),\;\langle v,{\mathcal{J}}{\dot{\gamma}}(t)\rangle\geq 0\,\}.

Then A(γ˙)A({\dot{\gamma}}) is an annulus in SMSM with boundaries γ˙{\dot{\gamma}}, γ˙-{\dot{\gamma}} whose interior is transversal to the geodesic flow. Because other geodesics intersecting γ\gamma must be tranversal to γ\gamma.

Refer to caption
Figure 11. Fried surgery for a double crossing. The new surface in the center does not self intersect and is transversal to the flow. The right figure shows that the surgery is obtained by cutting the surfaces along two segments and gluing them. The gluing is uniquely determined by the contitions of transversality to the flow and non intersection.
Refer to caption
Figure 12. Fried surgery for an intersection of a boundary orbit. The surgery is obtained by cutting along two segments and gluing. The gluing is uniquely determined by the flow. The resulting surface can be realized in an arbitrarily small neighborhood of the original surfaces. At interior points it is the same surgery as in figure 11.

We perform the Fried surgeries described in figures 11, 13 to the collection of Birkhoff annuli A(γ˙1),,A(γ˙2g+2),A(γ˙1),,A(γ˙2g+2)A({\dot{\gamma}}_{1}),\ldots,A({\dot{\gamma}}_{2g+2}),A(-{\dot{\gamma}}_{1}),\ldots,A(-{\dot{\gamma}}_{2g+2}). Observe that there are not triple intersections of the interior of these annuli because there are no triple intersections of their projected geodesics. We need to use the surgery in figure 13 instead of figure 12 because the annuli A(γ˙i)A({\dot{\gamma}}_{i}), A(γ˙i)A(-{\dot{\gamma}}_{i}) meet at their boundaries.

Refer to caption
Figure 13. The Birkhoff annuli A(γ˙i)A({\dot{\gamma}}_{i}) and A(γ˙i)A(-{\dot{\gamma}}_{i}) intersect at γ˙i{\dot{\gamma}}_{i} and γ˙i-{\dot{\gamma}}_{i}, and the orbit γ˙i{\dot{\gamma}}_{i} intersects transversely the annulus A(γ˙i+1)A({\dot{\gamma}}_{i+1}) so it is necessary to perform the surgery in figure 12 twice.

We prove that the result are two surfaces of section S1S_{1}, S2S_{2}, of genus 1, each of them with the 4g+44g+4 boundary components {γ˙1,,γ˙2g+2,γ˙1,,γ˙2g+2}\{{\dot{\gamma}}_{1},\ldots,{\dot{\gamma}}_{2g+2},-{\dot{\gamma}}_{1},\ldots,-{\dot{\gamma}}_{2g+2}\}. Observe that any orbit Γ\Gamma with

iπΓγiandπΓi=12g+2γi\forall i\;\pi\Gamma\neq\gamma_{i}\quad\text{and}\quad\pi\Gamma\cap\cup_{i=1}^{2g+2}\gamma_{i}\neq\emptyset

intersects S1S_{1} or S2S_{2} transversely.

Refer to caption
Figure 14. The sets A(R1)A(\partial R_{1}) and A(R3)A(\partial R_{3}) are a collage of half of the Birkhoff cylinders A(γ˙i)A({\dot{\gamma}}_{i}) and A(γ˙i)A(-{\dot{\gamma}}_{i}) respectively. Both A(R1)A(\partial R_{1}) and A(R3)A(\partial R_{3}) are cylinders.
Refer to caption
Figure 15. This is the same as figure 15 where the segments of γ˙i{\dot{\gamma}}_{i}, γ˙i-{\dot{\gamma}}_{i} has been drawn as curves and their adjacent vertical dashed segments have been drawn horizontal. Each of the figures is a cylinder, glued at its sides. The two figures are glued at the horizontal dashed lines. They form a torus 𝒯=A(R1)A(R3){\mathcal{T}}=A(\partial R_{1})\cup A(\partial R_{3}) with 4g+44g+4 holes. The complete system contains another torus 𝒯=A(R2)A(R4)-{\mathcal{T}}=A(\partial R_{2})\cup A(\partial R_{4}) which corresponds to a similar construction using the regions R2R_{2} and R4R_{4} with the vectors opposite to those of 𝒯{\mathcal{T}}. Both tori intersect pairwise at their boundaries.

Let A(Ri)A(\partial R_{i}), 1i41\leq i\leq 4 be the closure of the set of unit vectors based at Ri\partial R_{i} pointing outside of RiR_{i}. Then each A(Ri)A(\partial R_{i}) is a cylinder whose boundary projects to Ri\partial R_{i}. Figure 15 shows the cylinders A(R1)A(\partial R_{1}), A(R3)A(\partial R_{3}) and how they are glued after performing the surgeries in figures 1113. Figure 16 shows how the surgeries of figure 13 glue the segments aa and bb in figure 15. Then figure 15 is the same a figure 15 with the boundaries curved and rotated in order to show how the two cylinders A(R1)A(\partial R_{1}), A(R3)A(\partial R_{3}) glue after the surgery to form a torus S1S_{1} with 4g+44g+4 holes. Similarly S2S_{2} is obtained from A(R2)A(\partial R_{2}) and A(R4)A(\partial R_{4}).

Refer to caption
Figure 16. This figure shows in more detail how the annuli A(γ˙i)A({\dot{\gamma}}_{i}), A(γ˙i+1)A({\dot{\gamma}}_{i+1}), A(γ˙i)A(-{\dot{\gamma}}_{i}) are glued in figure 15 after the surgery in figure 13.
Refer to caption
Figure 17. This figure shows how the local invariant manifolds Wlocs(γ˙i)W^{s}_{loc}(\dot{\gamma}_{i}), Wlocu(γ˙i)W^{u}_{loc}(\dot{\gamma}_{i}) intersect the surface SjS_{j} over the point γiγi+1\gamma_{i}\cap\gamma_{i+1}. The surface SjS_{j} stays near the Birkhoff annuli A(γ˙i)A({\dot{\gamma}}_{i}), A(γ˙i+1)A({\dot{\gamma}}_{i+1}), A(γ˙i)A(-{\dot{\gamma}}_{i}). The annulus A(γ˙i)A({\dot{\gamma}}_{i}) is included in the vertical fibre Sγ˙iS{\dot{\gamma}}_{i}. The local manifolds do not intersect Sγ˙iγ˙iS{\dot{\gamma}}_{i}\setminus{\dot{\gamma}}_{i} because γi\gamma_{i} has no conjugate points. There are other intersections over the point γi1γi\gamma_{i-1}\cap\gamma_{i}.

Now we prove that the boundary components γ˙i{\dot{\gamma}}_{i}, γ˙i-{\dot{\gamma}}_{i} are hyperbolic and that their local invariant manifolds WlocsWlocuW^{s}_{loc}\cup W^{u}_{loc} intersect four times each section S1S_{1}, S2S_{2}.

The geodesics γi\gamma_{i} are minimizers in their homotopy class. Since MM is a surface their multiples γin(t):=γi(nt)\gamma_{i}^{n}(t):=\gamma_{i}(nt) are local minimizers, because a curve η\eta homotopic to γin\gamma_{i}^{n} contained in a small tubular neighborhood of γi\gamma_{i} can be separated into nn closed curves homotopic to γi\gamma_{i}. Then the length L(η)nL(γi)=L(γin)L(\eta)\geq n\cdot L(\gamma_{i})=L(\gamma_{i}^{n}). This implies that the whole geodesic γi(t)\gamma_{i}(t), tt\in{\mathbb{R}} has no conjugate points. Since γi\gamma_{i} is non-degenerate, then it must be hyperbolic. Since MM is orientable γi\gamma_{i}, is positive hyperbolic.

By section 3.7 the vertical subspace is not invariant. Then its forward iterates dϕt(V(γ˙i))d\phi_{t}(V({\dot{\gamma}}_{i})) must converge to the unstable subspace Eu(γ˙i)E^{u}({\dot{\gamma}}_{i}). But dϕt(V(γ˙i))d\phi_{t}(V({\dot{\gamma}}_{i})) can not approach the vertical V(γ˙i)V({\dot{\gamma}}_{i}) because by section 3.7 it would intersect the vertical non trivially, producing conjugate points. Thus its limit Eu(γ˙i)=limt+dϕt(V(ϕt(γ˙i)))E^{u}({\dot{\gamma}}_{i})=\lim_{t\to+\infty}d\phi_{t}(V(\phi_{-t}({\dot{\gamma}}_{i}))) satisfies

(45) Eu(γ˙i)V(γ˙i)={0}, and also Es(γ˙i)V(γ˙i)={0}.E^{u}({\dot{\gamma}}_{i})\cap V({\dot{\gamma}}_{i})=\{0\},\quad\text{ and also }\quad E^{s}({\dot{\gamma}}_{i})\cap V({\dot{\gamma}}_{i})=\{0\}.

The tangent space to the Birkhoff annulus A(γ˙i)A({\dot{\gamma}}_{i}) is

Tγ˙iA(γ˙i)=X(γ˙i)V(γ˙i),T_{{\dot{\gamma}}_{i}}A({\dot{\gamma}}_{i})=\langle X({\dot{\gamma}}_{i})\rangle\oplus V({\dot{\gamma}}_{i}),

where XX is the geodesic vector field. Then the invariant subspaces Es(γ˙i)E^{s}({\dot{\gamma}}_{i}), Eu(γ˙i)E^{u}({\dot{\gamma}}_{i}) are bounded away from Tγ˙iA(γ˙i)T_{{\dot{\gamma}}_{i}}A({\dot{\gamma}}_{i}). This implies that the local invariant manifolds Wεs(γ˙i)W^{s}_{\varepsilon}({\dot{\gamma}}_{i}), Wεu(γ˙i)W^{u}_{\varepsilon}({\dot{\gamma}}_{i}) do not intersect the interior of A(γ˙i)A({\dot{\gamma}}_{i}). Figure 17 shows how each of the local invariant manifolds intersect once the surface SjS_{j} over the intersection γi1γi\gamma_{i-1}\cap\gamma_{i} and once more over γiγi+1\gamma_{i}\cap\gamma_{i+1}. By section 3 this gives four saddle periodic points for the return map at each boundary component γ˙i{\dot{\gamma}}_{i} or γ˙i-{\dot{\gamma}}_{i} of each surface of section S1S_{1}, S2S_{2}.

Observe that there is >0\ell>0 and a neighborhood NN of Sj=i(γ˙iγ˙i)\partial S_{j}=\cup_{i}(-{\dot{\gamma}}_{i}\cup{\dot{\gamma}}_{i}) such that

zNϕ]0,[(z)Sj&ϕ],0[(z)Sj.\forall z\in N\quad\phi_{]0,\ell[}(z)\cap S_{j}\neq\emptyset\quad\&\quad\phi_{]-\ell,0[}(z)\cap S_{j}\neq\emptyset.

Therefore the orbits ±γ˙i\pm{\dot{\gamma}}_{i} will be rotating boundary orbits for the sections SjS_{j}.

4.2. Applications of the curve shortening flow.

Here we follow section §2 of [7]. Let (M,ρ)(M,\rho) be an oriented riemannian surface. Let S1=/S^{1}={\mathbb{R}}/{\mathbb{Z}}. For an embedding γ:S1M\gamma:S^{1}\hookrightarrow M, let νγ\nu_{\gamma} be its positively oriented normal vector field and let kγk_{\gamma} be the curvature of γ\gamma. Let Emb(S1,M)\operatorname{Emb}(S^{1},M) be the space of smooth embedded circles in MM endowed with the CC^{\infty} topology. Let

L(γ)=S1γ˙ρL(\gamma)=\int_{S^{1}}\left\|{\dot{\gamma}}\right\|_{\rho}

be the length functional. The curve shortening flow is a continuous map

𝒰Emb(S1,M),(s,γ0)Ψs(γ0)=:γs,{\mathcal{U}}\longrightarrow\operatorname{Emb}(S^{1},M),\qquad(s,\gamma_{0})\mapsto\Psi_{s}(\gamma_{0})=:\gamma_{s},

defined on a maximal open neighborhood 𝒰[0,[×Emb(S1,M){\mathcal{U}}\subset[0,\infty[\times\operatorname{Emb}(S^{1},M) of {0}×Emb(S1,M)\{0\}\times\operatorname{Emb}(S^{1},M) by the following PDE:

sγs=kγsνγs.\partial_{s}\gamma_{s}=k_{\gamma_{s}}\nu_{\gamma_{s}}.

The following properties are proved in [19], [11]:

  1. (i)

    Ψ0=id\Psi_{0}=id  and  ΨsΨt=Ψs+t\Psi_{s}\circ\Psi_{t}=\Psi_{s+t}  for all s,t0s,t\geq 0.

  2. (ii)

    Ψs(γθ)=Ψs(γ)θ\Psi_{s}(\gamma\circ\theta)=\Psi_{s}(\gamma)\circ\theta  for all γEmb(S1,M)\gamma\in\operatorname{Emb}(S^{1},M) and θDiff(S1)\theta\in\operatorname{Diff}(S^{1}).

  3. (iii)

    ddsL(ψs(γ))0\frac{d\,}{ds}L(\psi_{s}(\gamma))\leq 0 for all γEmb(S1,M)\gamma\in\operatorname{Emb}(S^{1},M), with equality if and only if the image of γ\gamma is a geodesic.

  4. (iv)

    Given γEmb(S1,M)\gamma\in\operatorname{Emb}(S^{1},M) let sγ=sup{s0|(s,γ)𝒰}s_{\gamma}=\sup\{\,s\geq 0\,|\,(s,\gamma)\in{\mathcal{U}}\,\}. Then sγs_{\gamma} is finite if and only if Ψs(γ)\Psi_{s}(\gamma) converges to a constant when ssγs\to s_{\gamma}.

A path-connected subset UMU\subset M is weakly convex if for any pair x,yUx,y\in U that can be joined by an absolutely continuous curve in UU of length smaller than the injectivity radius inj(M,g)\operatorname{inj}(M,g), the shortest geodesic joining xx and yy is contained in UU. Another useful property of Ψs\Psi_{s} is that it preserves weakly convex sets, namely

  1. (v)

    If UMU\subset M is weakly convex then

    γEmb(S1,U)s[0,sγ[ψs(γ)Emb(S1,U).\gamma\in\operatorname{Emb}(S^{1},U)\quad\Longrightarrow\quad\forall s\in[0,s_{\gamma}[\quad\psi_{s}(\gamma)\in\operatorname{Emb}(S^{1},U).

This flow is used in [7] to prove the following lemmata.

4.1 Lemma ([7] lemma 2.1).

Let UMU\subseteq M be a weakly convex subset that is not simply connected. Let 𝒞Emb(S1,U){\mathcal{C}}\subset\operatorname{Emb}(S^{1},U) be a connected component containing loops that are non-contractible in UU. Then, there exists a sequence γn𝒞\gamma_{n}\in{\mathcal{C}} converging in the C2C^{2}-topology to a simple closed geodesic γ\gamma contained in U¯\overline{U} of length

L(γ)=infζ𝒞L(ζ)>0.L(\gamma)=\inf_{\zeta\in{\mathcal{C}}}L(\zeta)>0.
4.2 Lemma ([7] lemma 2.2).

If UMU\subset M is weakly convex and KSMK\subset SM is invariant by the geodesic flow (i.e. t\forall t\in{\mathbb{R}} ϕt(K)=K\phi_{t}(K)=K) and such that π(K)U\pi(K)\subset U, then any path-connected component of Uπ(K)U\setminus\pi(K) is weakly convex.

A closed geodesic γ:S1M\gamma:S^{1}\to M is called a waist when any absolutely continuous curve ζ\zeta which is sufficiently C0C^{0}-close to γ\gamma satisfies L(ζ)L(γ)L(\zeta)\geq L(\gamma). By the argument before (45), non degenerate waists are positive hyperbolic and have no conjugate points.


4.3 Lemma.

A simple nondegenerate closed geodesic γ\gamma is a waist if and only if it has no conjugate points.

Proof:.

Suppose that γ\gamma is nondegenerate and has no conjugate points, we prove that it is a waist. The converse is standard. Consider the geodesic lagrangian L:TML:TM\to{\mathbb{R}} and hamiltonian H:TMH:T^{*}M\to{\mathbb{R}}

(46) L(x,v)=12|v|x2,H(x,p)=supvTxM{p(v)L(x,v)},H(x,p)=12|p|x2.L(x,v)=\tfrac{1}{2}|v|_{x}^{2},\quad H(x,p)=\sup_{v\in T_{x}M}\{p(v)-L(x,v)\},\quad H(x,p)=\tfrac{1}{2}\,|p|_{x}^{2}.

The Legendre transform (x,v)=v,x{\mathcal{L}}(x,v)=\langle v,\cdot\rangle_{x} conjugates the geodesic flow to the hamiltonian flow of HH on the energy level H12H\equiv\tfrac{1}{2}. Also (π1{x})=π1{x}{\mathcal{L}}(\pi^{-1}\{x\})=\pi_{*}^{-1}\{x\} identifies the vertical fibers. Observe that γ\gamma must be positive hyperbolic. Since (45) holds in the hamiltonian flow there is a neighborhood UU of γ\gamma where Ws(γ)TMW^{s}(\gamma)\subset T^{*}M is a graph:

TUWs(γ)={(x,ω(x))TxM:xU}.T^{*}U\cap W^{s}(\gamma)=\{\,(x,\omega(x))\in T^{*}_{x}M:x\in U\,\}.

Then ωΛ1(U)\omega\in\Lambda^{1}(U) is a 1-form on UU which is closed because Ws(γ)W^{s}(\gamma) is a lagrangian submanifold. And (dxdp)|Ws(γ)0(dx\wedge dp)|_{W^{s}(\gamma)}\equiv 0 because Ws(γ)W^{s}(\gamma) is tangent to the Reeb vector field of (H1{12},pdx)(H^{-1}\{\tfrac{1}{2}\},p\,dx). Since H(x,ω(x))12H(x,\omega(x))\equiv\tfrac{1}{2}, equation (46) implies that

(x,v)TMω(x)(v)L(x,v)+12.\forall(x,v)\in TM\qquad\omega(x)(v)\leq L(x,v)+\tfrac{1}{2}.

For xγx\in\gamma, we have that ω(x)=(x,γ˙)=γ˙,x\omega(x)={\mathcal{L}}(x,{\dot{\gamma}})=\langle{\dot{\gamma}},\cdot\rangle_{x}. Therefore ω(γ)γ˙1\omega(\gamma)\cdot{\dot{\gamma}}\equiv 1.

Let η\eta be an absolutely continuous curve C0C^{0} close to γ\gamma in UU parametrized by arc length. Then L(η,η˙)12L(\eta,{\dot{\eta}})\equiv\tfrac{1}{2} and

L(γ)=γω=ηωηL+12=L(η),L(\gamma)=\int_{\gamma}\omega=\int_{\eta}\omega\leq\int_{\eta}L+\tfrac{1}{2}=L(\eta),

where the second inequality holds because η\eta is homotopic to γ\gamma inside UU.

We need the following min-max lemma. These geodesic have conjugate points because minimax critical points can not be local minima.

4.4 Lemma ([7] lemma 2.4).

Let (M,ρ)(M,\rho) be an orientable riemannian surface.

  1. (i)

    If AMA\subset M is an annulus bordered by two waists, then int(A)\operatorname{int}(A) contains a non contractible simple closed geodesic with conjugate points.

  2. (ii)

    If DMD\subset M is a compact disk bounded by a waist, then int(D)\operatorname{int}(D) contains a simple closed geodesic with conjugate points.


4.5 Lemma ([11] lemma 5.9).

Let (M,ρ)(M,\rho) be a riemannian surface, and γ:[T,T]M\gamma:[-T,T]\to M a geodesic arc parametrized with unit speed whose interior γ|]T,T[\gamma|_{]-T,T[} contains a pair of conjugate points. Then there exists an open neighborhood USMU\subset SM of (γ(0),γ˙(0))(\gamma(0),{\dot{\gamma}}(0)) such that, for each (x,v)SU(x,v)\in SU, the geodesic ζ(t)=expx(tv)\zeta(t)=\exp_{x}(tv) intersets γ\gamma for some t[T,T]t\in[-T,T].

Lemma 4.5 implies the following corollary:

4.6 Corollary.

Let (M,ρ)(M,\rho) be an orientable riemannian surface and γ\gamma a simple closed geodesic with conjugate points.

  1. (i)

    There exists T>0T>0 and an open neighborhood VSMV\subset SM of the lift γ˙{\dot{\gamma}} such that, for each zVz\in V, the geodesic ζ(t):=πϕt(z)\zeta(t):=\pi\circ\phi_{t}(z) intersects γ\gamma on some positive time t1]0,T]t_{1}\in]0,T] and some negative time t2[T,0[t_{2}\in[-T,0[.

  2. (ii)

    There exists T>0T>0 and an open neighborhood UMU\subset M of γ\gamma such that, for each zSUz\in SU, the geodesic ζ(t):=πϕt(z)\zeta(t):=\pi\circ\phi_{t}(z) intersects γ\gamma on some t[T,T]t\in[-T,T].

A geodesic polygon in a riemannian surface is a simple closed curve which is a union of finitely many distinct geodesic arcs that is not one closed geodesic. Observe that necessarily the geodesic arcs are transversal. Therefore we have


4.7 Remark.
  1. (i)

    If PP is a geodesic polygon then there exist a neighborhood VSMV\subset SM of the lift P˙\dot{P} and >0\ell>0 such that for every zVz\in V and ζ(t):=πϕt(z)\zeta(t):=\pi\phi_{t}(z), both geodesic arcs ζ|]0,]\zeta|_{]0,\ell]}, ζ|[,0[\zeta|_{[-\ell,0[} intersect PP.

  2. (ii)

    If PP is a geodesic polygon there exists a neighborhood UMU\subset M of PP and >0\ell>0 such that for every zSUz\in SU and ζ(t):=πϕt(z)\zeta(t):=\pi\phi_{t}(z), the geodesic arc ζ|[,]\zeta|_{[-\ell,\ell]} intersects PP.


4.3. Complementary Birkhoff annuli.

In this section we obtain a complete system of surfaces of sections for (M,ρ)(M,\rho) provided that all waists are nondegenerate (i.e. hyperbolic). This is done by adding disjoint Birkhoff annuli to the surfaces obtained in section 4.1. The Birkhoff annuli have genus 0, so for them we don’t need to check the condition in theorem B.(3) on the number of intersections of the separatrices.

In fact some are Birkhoff annuli of waists which are in Kfix{K_{fix}} and other are Birkhoff annuli of minimax orbits which have index 1 and are in Krot{K_{rot}}. If these minimax orbits are hyperbolic, then their Floquet multipliers are negative and their invariant subspaces EsE^{s}, EuE^{u} intersect the vertical bundle V=kerdπV=\ker d\pi, π:TMM\pi:TM\to M, twice along one period. So each local invariant manifold Ws(γ)W^{s}(\gamma), Wu(γ)W^{u}(\gamma) intersects each Birkhoff annuli A(γ˙)A({\dot{\gamma}}), A(γ˙)A(-{\dot{\gamma}}) only once.

We prove the following.

4.8 Theorem.

Let (M,ρ)(M,\rho) be an orientable riemannian surface of genus gg with all its waists non degenerate. There are a finite number of surfaces of section Σ1,,Σ2n\Sigma_{1},\ldots,\Sigma_{2n} such that

  1. (a)

    If g=0g=0 then Σ1\Sigma_{1}, Σ2\Sigma_{2} are the Birkhoff annuli of a minimax simple closed geodesic.

  2. (b)

    If g>0g>0, Σ1\Sigma_{1}, Σ2\Sigma_{2} are the surfaces of genus 1 and 4G+44G+4 boundary components described in subsection 4.1.

  3. (c)

    Σ3,,Σ2n\Sigma_{3},\ldots,\Sigma_{2n} are Birkhoff annuli of n1n-1 mutually disjoint simple closed geodesics.

  4. (d)

    Σ3,,Σ2n\Sigma_{3},\ldots,\Sigma_{2n} are disjoint from Σ1\Sigma_{1}, Σ2\Sigma_{2}.

  5. (e)

    Every geodesic orbit intersects Σ1Σ2n\Sigma_{1}\cup\cdots\cup\Sigma_{2n}.

  6. (f)

    Let Kfix{K_{fix}} be the union of the set of closed orbits without conjugate points in i=32nΣi\cup_{i=3}^{2n}\partial\Sigma_{i} and let Krot=i=12nΣiKfix{K_{rot}}=\cup_{i=1}^{2n}\partial\Sigma_{i}\setminus{K_{fix}}. There are 0<<0<\ell<\infty and a neighborhood 𝒰{\mathcal{U}} of Krot{K_{rot}} in SMSM such that

    z𝒰ϕ]0,[(z)𝚺&ϕ],0[(z)𝚺,𝚺:=i=12nΣ.\forall z\in{\mathcal{U}}\quad\phi_{]0,\ell[}(z)\cap{\mathbf{\Sigma}}\neq\emptyset\quad\&\quad\phi_{]-\ell,0[}(z)\cap{\mathbf{\Sigma}}\neq\emptyset,\qquad{\mathbf{\Sigma}}:=\cup_{i=1}^{2n}\Sigma.
  7. (g)

    If γ\gamma is a geodesic with γ˙(]0,+[)𝚺={\dot{\gamma}}(]0,+\infty[)\cap{\mathbf{\Sigma}}=\emptyset then γ˙(t)Ws(zt){\dot{\gamma}}(t)\in W^{s}(z_{t}) for some ztKfixz_{t}\in{K_{fix}}.

  8. (h)

    If γ\gamma is a geodesic with γ˙(],0[)𝚺={\dot{\gamma}}(]\!-\!\infty,0[)\cap{\mathbf{\Sigma}}=\emptyset then γ˙(t)Wu(zt){\dot{\gamma}}(t)\in W^{u}(z_{t}) for some ztKfixz_{t}\in{K_{fix}}.

The following proposition is proved in lemmas 3.8 and 3.7 in [7], using examples 3.2 and 3.3 in [7].

4.9 Proposition ([7] lemmas 3.8, 3.7).

Let (M,ρ)(M,\rho) be a riemannian surface and let DMD\subset M be a simply connected open set whose boundary D=P\partial D=P is a geodesic polygon or a simple closed geodesic with conjugate points. Suppose that every simple closed geodesic without conjugate points contained in DD is non-degenerate. Then every collection of mutually disjoint simple closed geodesics contained in DD is finite.

A corset (A,w)(A,w) in (M,ρ)(M,\rho) is an annulus AMA\subset M such that int(A)\operatorname{int}(A) contains a simple closed geodesic ww which is a waist and that the boundary components of A\partial A are either a polygon or a simple closed geodesic with conjugate points. A bowl is a disk DMD\subset M whose boundary D\partial D is either a geodesic polygon or a simple closed geodesic with conjugate points. We further require that corsets and bowls are connected components of the complement of finitely many geodesics. Observe that by lemma 4.2, corsets and bowls are weakly convex.

4.10 Lemma.
  1. (1)

    If (A,w1)(A,w_{1}) is a corset, U=intAw1U=\operatorname{int}A\setminus w_{1} and tϕt(SU)\cap_{t\in{\mathbb{R}}}\phi_{-t}(SU)\neq\emptyset; then there are two corsets (A1,w1)(A_{1},w_{1}), (A2,w2)(A_{2},w_{2}) with A=A1A2A=A_{1}\cup A_{2} and A1A2=AiAA_{1}\cap A_{2}=\partial A_{i}\setminus\partial A, i=1,2i=1,2.

  2. (2)

    If DD is a bowl, V=DDV=D\setminus\partial D and tϕt(SV)\cap_{t\in{\mathbb{R}}}\phi_{-t}(SV)\neq\emptyset; then there is a corset (A,w)(A,w) and a bowl BB such that D=ABD=A\cup B and AB=B=ADA\cap B=\partial B=\partial A\setminus\partial D.

Refer to caption
Figure 18. These are examples of a decomposition of a bowl and a corset in lemma 4.10 when there is a new invariant subset projecting in its interior. The star \star marks a point in the projection πΛ\pi\Lambda of the new invariant subset and determines the homotopy class of the new waist.
Proof:.

(1). Write Λ:=tϕt(SU)\Lambda:=\cap_{t\in{\mathbb{R}}}\phi_{-t}(SU). By corollary 4.6.(ii) and remark 4.7.(ii) there is a neighborhood NN of A\partial A such that ΛSN=\Lambda\cap SN=\emptyset. Since Uw1=U\cap w_{1}=\emptyset we have that πΛw1\pi\Lambda\not\subset w_{1}. By lemma 4.2 any path-connected component of A(w1πΛ)A\setminus(w_{1}\cup\pi\Lambda) is weakly convex. Let a2a_{2} be the connected component of A\partial A which is included in a connected component of Aw1A\setminus w_{1} which intersects πΛ\pi\Lambda. Let a1a_{1} be the other component of A\partial A. Let WW be the connected component of A(w1πΛ)A\setminus(w_{1}\cup\pi\Lambda) which contains a2a_{2}. Observe that a2a_{2} is not homotopic to w1w_{1} in WW.

Let xπΛw1x\in\pi\Lambda\setminus w_{1}, and ε>0\varepsilon>0 with d(x,w1)>εd(x,w_{1})>\varepsilon. We claim that if a closed curve γW\gamma\subset W is homotopic to a2a_{2} inside WW then there is yγγy_{\gamma}\in\gamma such that d(yγ,w1)εd(y_{\gamma},w_{1})\geq\varepsilon. For if

γB(w1,ε):={zA:d(z,w1)<ε},\gamma\subset B(w_{1},\varepsilon):=\{z\in A:d(z,w_{1})<\varepsilon\},

then γ\gamma is homotopic to w1w_{1} inside A{x}A\setminus\{x\}. Thus γ\gamma is non homotopic to a2a_{2} inside A{x}A\setminus\{x\}. Then γ\gamma is non homotopic to a2a_{2} inside WA{x}W\subset A\setminus\{x\}. A contradiction. Consequently, if η\eta is a C0C^{0} limit of curves γnW\gamma_{n}\subset W homotopic to a2a_{2} inside WW, then

(47) ηw1.\eta\neq w_{1}.

Let 𝒞{\mathcal{C}} be the connected component of Emb(S1,W)\operatorname{Emb}(S^{1},W) containing a curve homotopic to a2a_{2}. By lemma 4.1 there is a sequence γn𝒞\gamma_{n}\in{\mathcal{C}} converging in the C2C^{2} topology to a simple closed geodesic w2w_{2} in W¯\overline{W} of length L(w1)=infζ𝒞L(ζ)>0L(w_{1})=\inf_{\zeta\in{\mathcal{C}}}L(\zeta)>0. Then w2w_{2} is a waist and by (47), w2w1w_{2}\neq w_{1}. Since w1w_{1}, w2w_{2} are waists in the same homotopy class in AA, we have that w1w2=w_{1}\cap w_{2}=\emptyset (c.f. [13, Prop. 1.7]). Since w2W¯Aw_{2}\subset\overline{W}\subset A, if w2a2w_{2}\cap a_{2}\neq\emptyset then w2w_{2} and a2a_{2} would be tangent geodesics (segments) and hence the same geodesic. But this is not possible because a2a_{2} has conjugate points or is a polygon and w1w_{1} is a waist.

In the annulus AA the curves a1a_{1}, w1w_{1}, w2w_{2}, a2a_{2} are all disjoint and homotopic. There is an annulus A(w1,w2)A(w_{1},w_{2}) in AA with boundaries w1w_{1} and w2w_{2}. By lemma 4.4.(i) there is a non contractibe simple closed geodesic hh with conjugate points in int(A(w1,w2))\operatorname{int}(A(w_{1},w_{2})). In particular hh is disjoint and homotopic to wiw_{i}, aia_{i}, i=1,2i=1,2. Denote the annuli A1:=A(a1,h)A_{1}:=A(a_{1},h), A2=A(h,a2)A_{2}=A(h,a_{2}) with boundaries (a1,h)(a_{1},h), (h,a2)(h,a_{2}) respectively. Then (A1,w1)(A_{1},w_{1}), (A2,w2)(A_{2},w_{2}) are the desired corsets.

(2). Write Λ:=tϕt(SV)\Lambda:=\cap_{t\in{\mathbb{R}}}\phi_{-t}(SV). By corollary 4.6.(ii) and remark 4.7.(ii) there is a neighborhood NN of D\partial D such that ΛSN=\Lambda\cap SN=\emptyset. Let WW be the connected component of DπΛD\setminus\pi\Lambda which contains int(N)\operatorname{int}(N). Observe that D\partial D is non-contractible in W¯\overline{W}. Let 𝒞{\mathcal{C}} be a connected component in Emb(S1,W)\operatorname{Emb}(S^{1},W) containing curves homotopic to D\partial D. By lemma 4.1 apllied to 𝒞{\mathcal{C}}, there is a waist w1w_{1} in W¯\overline{W}. The waist w1w_{1} bounds a disk D1D_{1} in DD. By lemma 4.4.(ii) there is a simple closed geodesic h1h_{1} in int(D1)\operatorname{int}(D_{1}) with conjugate points. Let B1B_{1} be the disk in DD with B1=h1\partial B_{1}=h_{1}. Let A1=A(D,h1)A_{1}=A(\partial D,h_{1}) be the annulus with boundary Dh1\partial D\cup h_{1}. Then B1B_{1} is a bowl and (A1,w1)(A_{1},w_{1}) is a corset with disjoint interiors and D=A1B1D=A_{1}\cup B_{1} as required.

Proof of theorem 4.8:

If M=SS2M=\SS^{2} let Σ1,Σ2\Sigma_{1},\Sigma_{2} be the Birkhoff annuli of a simple closed minimax geodesic mm and let R1,R2R_{1},R_{2} be the two disks bounded by mm. Otherwise let R1,,R4R_{1},\ldots,R_{4} be the disks in MM and Σ1\Sigma_{1}, Σ2\Sigma_{2} be the surfaces of section of genus 1 obtained in subsection 4.1.

Given and open subset VSMV\subset SM define the forward trapped set trap+(V)\operatorname{trap}_{+}(V) and the backward trapped set trap(V)\operatorname{trap}_{-}(V) as

trap±(V)={zSM:τt>τϕ±t(z)V}.\operatorname{trap}_{\pm}(V)=\{\,z\in SM\,:\;\exists\tau\quad\forall t>\tau\quad\phi_{\pm t}(z)\in V\,\}.
4.11 Claim.

For each i=1,,{2,4}i=1,\ldots,\{2,4\} there are finitely many corsets (A1i,w1i),,(A_{1}^{i},w_{1}^{i}),\ldots, (Amii,wmii)(A_{m_{i}}^{i},w_{m_{i}}^{i}) and a bowl BmiiB_{m_{i}}^{i} with disjoint interiors such that Ri=A1iAmiiBmiiR_{i}=A^{i}_{1}\cup\cdots\cup A^{i}_{m_{i}}\cup B^{i}_{m_{i}} and letting

(48) Ki:=Rij=1mi(Ajiwji)Bmii,\displaystyle K_{i}:=\partial R_{i}\cup\textstyle\cup_{j=1}^{m_{i}}(\partial A^{i}_{j}\cup w^{i}_{j})\cup\partial B^{i}_{m_{i}},
(49) γ geodesicγ(0)Riγ()Ki,\displaystyle\gamma\text{ geodesic}\quad\gamma(0)\in R_{i}\quad\Longrightarrow\quad\gamma({\mathbb{R}})\cap K_{i}\neq\emptyset,
(50) trap±SUij=1miWs,u(w˙ji)Ws,u(w˙ji),Ui:=RiKi.\displaystyle\operatorname{trap}_{\pm}SU_{i}\subset\textstyle\bigcup_{j=1}^{m_{i}}W^{s,u}(\dot{w}^{i}_{j})\cup W^{s,u}(-\dot{w}^{i}_{j}),\qquad U_{i}:=R_{i}\setminus K_{i}.
Refer to caption
Figure 19. This is a possible outcome of a decomposition Ri=A1A2B2R_{i}=A_{1}\cup A_{2}\cup B_{2}. The decomposition depends on the order in which invariant remaining subsets are given. The subindices are the order in which the closed orbits are found in this event.

Assume claim 4.11 holds. Let Σ3,,Σ2n\Sigma_{3},\ldots,\Sigma_{2n} be the collection of the two Birkhoff annuli of the geodesics in KiintRiK_{i}\cap\operatorname{int}R_{i} whenever trap±S(RiRi)\operatorname{trap}_{\pm}S(R_{i}\setminus\partial R_{i})\neq\emptyset. Then 4.8.(a)-(d) hold. Also (49) implies 4.8.(e). By (50) we have that

trap±(SMj=12nΣj)i=1{2,4}j=1miWs,u(w˙ji)Ws,u(w˙ji).\operatorname{trap}_{\pm}(SM\setminus\cup_{j=1}^{2n}\Sigma_{j})\subset\textstyle\bigcup_{i=1}^{\{2,4\}}\bigcup_{j=1}^{m_{i}}W^{s,u}(\dot{w}^{i}_{j})\cup W^{s,u}(-\dot{w}^{i}_{j}).

This implies 4.8.(g) and 4.8.(h).

We have that Kfix=i,j{w˙ji,w˙ji}{K_{fix}}=\cup_{i,j}\{\dot{w}^{i}_{j},-\dot{w}^{i}_{j}\} and Krot=i=12nΣiKfix{K_{rot}}=\cup_{i=1}^{2n}\partial\Sigma_{i}\setminus{K_{fix}}. Since the orbits in Krot{K_{rot}} are either in a polygon or have conjugate points, corollary 4.6.(i) and remark 4.7.(i) imply 4.8.(f).

We now prove claim 4.11. Observe that the disks R1,,R{2,4}R_{1},\ldots,R_{\{2,4\}} are bowls. Recall that the surfaces Σ1,Σ2\Sigma_{1},\Sigma_{2} can be constructed inside an arbitrarily small neighborhood of S(i=1{2,4}Ri)S(\cup_{i=1}^{\{2,4\}}\partial R_{i}).

If tϕt(S(intRi))\cap_{t\in{\mathbb{R}}}\phi_{t}(S(\operatorname{int}R_{i}))\neq\emptyset, by lemma 4.10.(2) we can add a corset (A1,w1)(A_{1},w_{1}) and a bowl B1B_{1} with Ri=A1B1R_{i}=A_{1}\cup B_{1}, A1B1=B1=A1Ri=:b1A_{1}\cap B_{1}=\partial B_{1}=\partial A_{1}\setminus\partial R_{i}=:b_{1}. Observe that {w˙1,b˙1}\{\dot{w}_{1},\dot{b}_{1}\} is a set of pairwise disjoint simple closed geodesics in RiR_{i}. Inductively, suppose we have corsets (Aj,wj)(A_{j},w_{j}), j=1,,mj=1,\ldots,m and a bowl BmB_{m} with disjoint interiors and Ri=A1AmBmR_{i}=A_{1}\cup\cdots\cup A_{m}\cup B_{m}. Let U=j=1m(intAiwj)intBmU=\cup_{j=1}^{m}(\operatorname{int}A_{i}-w_{j})\cup\operatorname{int}B_{m}. If Λ:=tϕt(SU)\Lambda:=\cap_{t\in{\mathbb{R}}}\phi_{t}(SU)\neq\emptyset then either πΛintBm\pi\Lambda\cap\operatorname{int}B_{m}\neq\emptyset or πΛ(intAjwj)\pi\Lambda\cap(\operatorname{int}A_{j}-w_{j})\neq\emptyset for some jj. Therefore either tϕt(S(intBm))\cap_{t\in{\mathbb{R}}}\phi_{t}(S(\operatorname{int}B_{m}))\neq\emptyset or tϕt(S(intAjwj))\cap_{t\in{\mathbb{R}}}\phi_{t}(S(\operatorname{int}{A_{j}}-w_{j}))\neq\emptyset. By lemma 4.10 there is a corset (Am+1,wm+1)(A_{m+1},w_{m+1}), and a bowl Bm+1B_{m+1} or a corset (Aj,wj)(A^{\prime}_{j},w_{j}), with either Aj=AjAm+1A_{j}=A^{\prime}_{j}\cup A_{m+1} or Bm=Am+1Bm+1B_{m}=A_{m+1}\cup B_{m+1} where the sets in the unions have disjoint interiors. In any case we obtain a new decomposition

Ri=A1Am+1Bm+1,R_{i}=A_{1}\cup\cdots\cup A_{m+1}\cup B_{m+1},

where (Aj,wj)(A_{j},w_{j}) are corsets and Bm+1B_{m+1} is a bowl, all with disjoint interiors. The process can continue as long as

(51) tϕt(SU),U=j=1m+1(intAiwj)intBm+1.\cap_{t\in{\mathbb{R}}}\phi_{t}(SU)\neq\emptyset,\quad U=\cup_{j=1}^{m+1}(\operatorname{int}A_{i}-w_{j})\cup\operatorname{int}B_{m+1}.

Here the closed simple geodesics {wj}j=1m+1\{w_{j}\}_{j=1}^{m+1} are mutually disjoint waists and contained in RiR_{i}. By proposition 4.9 this process must stop. Then for each RiR_{i}, i=1,,{2,4}i=1,\ldots,\{2,4\} there is m=:mi1m=:m_{i}-1 for which condition (51) does not hold. This implies (49).

Let Ui=RiKiU_{i}=R_{i}\setminus K_{i} be from (50). Suppose that ztrap+(SUi)z\in\operatorname{trap}_{+}(SU_{i}) then its ω\omega-limit

ω(z):=t>Tϕ[T,+[(z)¯\omega(z):=\textstyle\bigcap_{t>T}\overline{\phi_{[T,+\infty[}(z)}

is an invariant set with projection π(ω(z))Ui¯\pi(\omega(z))\subset\overline{U_{i}}. Since condition (51) does not hold for U=UiU=U_{i} we have that π(ω(z))Ui=\pi(\omega(z))\cap U_{i}=\emptyset. Therefore π(ω(z))\pi(\omega(z)) is a connected component of KiK_{i} in (48). By corollary 4.6.(i) and remark 4.7.(i) the forward orbit of zz can not approach the boundary orbits in S(Aj)S(\partial A_{j}) or SBmSB_{m} without intersecting SKiSK_{i}. Thus π(ω(z))=wji\pi(\omega(z))=w^{i}_{j} for some 1jmi1\leq j\leq m_{i}. Since ω(z)\omega(z) is invariant, this implies that ω(z)=±w˙ji\omega(z)=\pm{\dot{w}}^{i}_{j}. This proves (50).

References

  • [1] K. G. Andersson, Poincaré’s discovery of homoclinic points, Arch. Hist. Exact Sci. 48 (1994), no. 2, 133–147.
  • [2] George D. Birkhoff, Dynamical systems with two degrees of freedom, Trans. Amer. Math. Soc. 18 (1917), no. 2, 199–300.
  • [3] by same author, Dynamical systems, American Mathematical Society, Providence, R.I., 1966.
  • [4] Chong-Qing Cheng and Jun Yan, Existence of Diffusion Orbits in a priori Unstable Hamiltonian Systems, Journal of Differential Geometry 67 (2004), 457–518.
  • [5] Andrew Clarke, Generic properties of geodesic flows on analytic hypersurfaces of Euclidean space, Discrete Contin. Dyn. Syst. 42 (2022), no. 12, 5839–5868.
  • [6] Gonzalo Contreras, Geodesic flows with positive topological entropy, twist maps and hyperbolicity, Annals of Math. 172 (2010), no. 2, 761–808.
  • [7] Gonzalo Contreras, Gerhard Knieper, Marco Mazzucchelli, and Benjamin H. Schulz, Surfaces of section for geodesic flows of closed surfaces, Preprint arXiv:2204.11977, 2022.
  • [8] Gonzalo Contreras and Marco Mazzucchelli, Proof of the C2{C}^{2} stability conjecture for geodesic flows of closed surfaces, Preprint arXiv.org 2109.10704, 2021.
  • [9] by same author, Existence of Birkhoff sections for Kupka-Smale Reeb flows of closed contact 3-manifolds, Geom. Funct. Anal. 32 (2022), no. 5, 951–979.
  • [10] Gonzalo Contreras-Barandiarán and Gabriel Paternain, Genericity of geodesic flows with positive topological entropy on S2S^{2}, Jour. Diff. Geom. 61 (2002), 1–49.
  • [11] G. De Philippis, M. Mazzucchelli, M. Marini, and S. Suhr, Closed geodesics on reversible Finsler 2-spheres, to appear in J. Fixed Point Theory Appl., 2020.
  • [12] Amadeu Delshams, Rafael de la Llave, and Tere M. Seara, A geometric approach to the existence of orbits with unbounded energy in generic periodic perturbations by a potential of generic geodesic flows of 𝐓2{\bf T}^{2}, Comm. Math. Phys. 209 (2000), no. 2, 353–392.
  • [13] Benson Farb and Dan Margalit, A primer on mapping class groups, Princeton Mathematical Series, vol. 49, Princeton University Press, Princeton, NJ, 2012.
  • [14] Bassam Fayad and Raphaël Krikorian, Herman’s last geometric theorem, Ann. Sci. Éc. Norm. Supér. (4) 42 (2009), no. 2, 193–219.
  • [15] John Franks and Patrice Le Calvez, Regions of instability for non-twist maps, Ergodic Theory Dynam. Systems 23 (2003), no. 1, 111–141.
  • [16] Michael E. Gage, Curve shortening on surfaces, Ann. Sci. École Norm. Sup. (4) 23 (1990), no. 2, 229–256.
  • [17] Chr. Genecand, Transversal homoclinic orbits near elliptic fixed points of area-preserving diffeomorphisms of the plane, Dynamics reported, Dynam. Report. Expositions Dynam. Systems (N.S.), vol. 2, Springer, Berlin, 1993, pp. 1–30.
  • [18] Marian Gidea, Rafael de la Llave, and Tere M-Seara, A general mechanism of diffusion in Hamiltonian systems: qualitative results, Comm. Pure Appl. Math. 73 (2020), no. 1, 150–209.
  • [19] Matthew A. Grayson, Shortening embedded curves, Ann. of Math. (2) 129 (1989), no. 1, 71–111.
  • [20] H. Hofer, K. Wysocki, and E. Zehnder, Properties of pseudoholomorphic curves in symplectisations. i: Asymptotics, Ann. Inst. Henri Poincaré, Anal. Non Linéaire 13 (1996), no. 3, 337–379, correction ibid. 15, No.4, 535-538 (1998).
  • [21] by same author, Finite energy foliations of tight three-spheres and Hamiltonian dynamics, Ann. of Math. (2) 157 (2003), no. 1, 125–255.
  • [22] Kei Irie, Dense existence of periodic Reeb orbits and ECH spectral invariants, J. Mod. Dyn. 9 (2015), 357–363.
  • [23] M. Juvan, A. Malnič, and B. Mohar, Systems of curves on surfaces, J. Combin. Theory Ser. B 68 (1996), no. 1, 7–22.
  • [24] Gerhard Knieper and Howard Weiss, CC^{\infty} genericity of positive topological entropy for geodesic flows on S2S^{2}, J. Differential Geom. 62 (2002), no. 1, 127–141.
  • [25] Patrice Le Calvez, Étude Topologique des Applications Déviant la Verticale, Ensaios Matemáticos, 2, Sociedade Brasileira de Matemática, Rio de Janeiro, 1990.
  • [26] John N. Mather, Invariant subsets for area preserving homeomorphisms of surfaces, Mathematical analysis and applications, Part B, Academic Press, New York-London, 1981, pp. 531–562.
  • [27] John N. Mather, Arnol’d diffusion. I. Announcement of results, Sovrem. Mat. Fundam. Napravl. 2 (2003), 116–130, translation in J. Math. Sci. (N.Y.) 124 (2004), no. 5, 5275–5289.
  • [28] Jürgen Moser, Stable and random motions in dynamical systems, Princeton University Press, Princeton, N. J., 1973, With special emphasis on celestial mechanics, Hermann Weyl Lectures, the Institute for Advanced Study, Princeton, N. J, Annals of Mathematics Studies, No. 77.
  • [29] Fernando Oliveira, On the generic existence of homoclinic points, Ergodic Theory Dynam. Systems 7 (1987), no. 4, 567–595.
  • [30] Fernando Oliveira and Gonzalo Contreras, The Ideal Boundary and the Accumulation Lemma, Preprint arXiv.2205.14738, 2022.
  • [31] by same author, No elliptic points from fixed prime ends, Preprint arXiv:2205.14768, 2022.
  • [32] Dennis Pixton, Planar homoclinic points, J. Differential Equations 44 (1982), no. 3, 365–382.
  • [33] H. Poincaré, Les méthodes nouvelles de la mécanique céleste. Tome I, Gauthier-Villars, Paris, 1892, Solutions périodiques. Non-existence des intégrales uniformes. Solutions asymptotiques.
  • [34] by same author, Les méthodes nouvelles de la mécanique céleste. Tome III, Gauthier-Villars, Paris, 1899, Invariant intégraux. Solutions périodiques du deuxième genre. Solutions doublement asymptotiques.
  • [35] Zhihong Xia and Pengfei Zhang, Homoclinic intersections for geodesic flows on convex spheres, Dynamical systems, ergodic theory, and probability: in memory of Kolya Chernov, Contemp. Math., vol. 698, Amer. Math. Soc., Providence, RI, 2017, pp. 221–238.
  • [36] E. Zehnder, Homoclinic points near elliptic fixed points, Comm. Pure Appl. Math. 26 (1973), 131–182. MR 345134