foliations for Reeb flows on the tight 3-sphere
Abstract.
We study the existence of foliations adapted to Reeb flows on the tight -sphere. These foliations admit precisely three binding orbits whose Conley-Zehnder indices are , , and , respectively. All regular leaves are disks and annuli asymptotic to the binding orbits. Our main results provide sufficient conditions for the existence of foliations with prescribed binding orbits. We also exhibit a concrete Hamiltonian on admitting foliations when restricted to suitable energy levels.
Key words and phrases:
Hamiltonian dynamics, pseudo-holomorphic curves, Reeb flows2020 Mathematics Subject Classification:
Primary 53D35; Secondary 37J46, 37J551. Introduction
In this paper, we use the theory of pseudo-holomorphic curves in symplectizations to study the existence of transverse foliations for Reeb flows on the tight -sphere. These flows are equivalent to Hamiltonian flows on star-shaped energy surfaces in .
Some of the relevant results on the existence of global surfaces of section for Reeb flows on the tight -sphere follow from the theory of pseudo-holomorphic curves in symplectizations, initiated by Hofer in [14]. A global surface of section for a -dimensional Reeb flow is an embedded surface with boundary whose interior is transverse to the flow and whose boundary consists of periodic orbits. The surface intersects every orbit, and a first return map describes the qualitative properties of the flow. Hofer, Wysocki, and Zehnder showed in [19] that a dynamically convex Reeb flow on the -sphere admits disk-like global surface of section. The disk is part of an -family of global surfaces of section, forming an open book decomposition. This result applies to Hamiltonian flows on strictly convex energy surfaces in .
A generalization of global surfaces of section and open books adapted to Reeb flows are the transverse foliations. They consist of a singular foliation whose singular set is a finite set of periodic orbits, called binding orbits. The regular leaves are surfaces asymptotic to the binding orbits and transverse to the flow. Such foliations may imply the existence of other periodic orbits and homoclinics to a hyperbolic binding orbit. One may obtain transverse foliations as projections of finite energy foliations in the symplectization. In [21], the authors generalize the results in [19] for generic star-shaped energy surfaces in . They prove that a Hamiltonian flow restricted to a generic star-shaped energy surface admits a finite energy foliation on its symplectization. There are several possible configurations of transverse foliations in [21]. In any case, the binding orbits have Conley-Zehnder indices , , or , and the regular leaves are punctured spheres. If a single periodic orbit forms the binding, then it has Conley-Zehnder index , and the transverse foliation determines an open book decomposition with disk-like pages, which are global surfaces of section for the flow.
In [7, 8], de Paulo and Salomão provide concrete examples of Reeb flows on the tight -sphere that admit transverse foliations other than open books. They study Hamiltonian flows close to a critical energy surface containing two strictly convex sphere-like subsets that touch each other at a saddle-center equilibrium point. For energies slightly above the critical value, they show that the flow admits a transverse foliation, called a foliation, with three binding orbits. The Lyapunoff orbit near the equilibrium is one of the binding orbits, and the other binding orbits have Conley-Zehnder index . See also [3, 4, 11, 33] for interesting results on the existence of transverse foliations.
A natural question that arises from the results in [19] and [21] is to find necessary and sufficient conditions for a finite set of closed orbits to be the binding of a transverse foliation. Some recent studies answer this question in the case of disk-like global surfaces of section. Characterization of the closed orbits that bound a disk-like global surface of section, assuming that the Reeb flow on is dynamically convex, is given in [22] and [23]. A simple orbit bounds a disk-like global surface of section if and only if it is unknotted and has self-linking number . In [25], the authors characterize the closed orbits that bound a disk-like global surface of section for a nondegenerate Reeb flow on the tight -sphere. More precisely, a simple closed orbit bounds a disk-like global surface of section if and only if it is unknotted, has Conley-Zehnder index , has self-linking number , and all index- orbits link with . In both cases, the global surface of section is a page of an open book decomposition.
In the present work, we study the existence of transverse foliations adapted to Reeb flows on the tight -sphere that admit a binding orbit with Conley-Zehnder index . More precisely, we study a particular type of transverse foliation, called foliation, which has three binding orbits , , and , whose Conley-Zehnder indices are , , and , respectively. The rigid leaves are formed by a plane asymptotic to , a pair of cylinders connecting to forming a -torus, and a cylinder connecting to . There is a family of planes asymptotic to , breaking at each end into a cylinder connecting and and the plane asymptotic to . There is also a family of cylinders connecting to , which breaks at each end into a cylinder connecting and and the cylinder connecting and . Our main result provides sufficient conditions for a set of three closed orbits to be the binding orbits of a foliation. We also exhibit a Hamiltonian on whose flow restricted to suitable energy surfaces admits a foliation.
The foliations are one of the possible transverse foliations established in [21]. We expect to use our methods to find concrete Hamiltonians admitting more general transverse foliations.
1.1. Main results
Our objective is to study the existence of transverse foliations for Reeb flows associated to tight contact forms on . A transverse foliation for a flow on a closed oriented -manifold consists of
-
•
A finite set of simple periodic orbits of , called binding orbits;
-
•
A smooth foliation of by properly embedded surfaces. Every leaf is transverse to and has an orientation induced by and . For every leaf there exists a compact embedded surface so that and is a union of connected components of . An end of is called a puncture. To each puncture there is an associated component of , called the asymptotic limit of at . A puncture of is called positive if the orientation on induced by coincides with the orientation induced by . Otherwise, is called negative.
This definition follows [27] and is based on the finite energy foliations from [21].
Let be a contact form on , that is, is a volume form. The Reeb vector field associated to is uniquely determined by
(1) |
The flow of is called the Reeb flow of . The contact structure associated to is the -plane distribution We denote by the projection onto uniquely determined by .
An embedded disk satisfying is called an overtwisted disk. The contact form is tight if the contact structure does not admit an overtwisted disk. Consider with coordinates . The Liouville -form
(2) |
restricts to a contact form on . By results of Bennequim [1] and Eliashberg [10], we know that, up to diffeomorphism, any tight contact form on is of the form for some smooth function . If is a regular energy level of a Hamiltonian function on , then the associated Hamiltonian flow restricted to is equivalent to the Reeb flow of .
We call a pair , where is a periodic trajectory of and is a period of , a Reeb orbit. We identify with the element of induced by the loop where the quotient is relative to the translations . By abuse of notation we sometimes write . If is the minimal positive period of , we call simple. If is an integer, the iterate of will be denoted by . We denote the set of Reeb orbits by .
The Reeb flow preserves the contact structure and the maps are -symplectic. We call the orbit nondegenerate if is not an eigenvalue of . If every orbit is nondegenerate, then the contact form is called nondegenerate.
A simple orbit is called unknotted if is an unknot. We say that a set of simple orbits is an unlink if is an unlink. We say that two orbits and are linked if the linking number is nonzero.
Before stating our main results, we give some necessary definitions.
Definition 1.1 (Strong transverse section).
Let be a contact form on . Let be a compact embedded surface such that is transverse to the Reeb vector field and consists of a finite number of simple orbits in . is called a strong transverse section if every connected component of associated to an orbit has a neighborhood on parametrized by such that , , and the section of defined by
(3) |
satisfies
where is the Lie derivative with respect to . See Remark 2.8.
Definition 1.2 ( foliation).
Let be a contact form on . A foliation adapted to is a transverse foliation for the associated Reeb flow satisfying the following properties. The set of binding orbits consists of three simple orbits , , and with Conley-Zehnder indices , , and respectively, self-linking number , and such that is an unlink. The foliation of is as follows:
-
•
contains a pair of cylinders and , both asymptotic to at their positive punctures and at their negative punctures. is homeomorphic to a torus and is -embedded. divides into two closed regions and .
-
•
contains a disk asymptotic to at its positive puncture and a cylinder asymptotic to at its positive puncture and at its negative puncture. is a -embedded disk with boundary and transverse to .
-
•
contains a one-parameter family of disks , , all of them asymptotic to at their positive punctures, such that foliate .
-
•
contains a one-parameter family of cylinders , , all of them asymptotic to at their positive punctures and at their negative punctures, such that foliate .
-
•
The closure of every leaf of is a strong transverse section.
See figure 1. The Conley-Zehnder index and the self-linking number are discussed in Section 2.

A complex structure on is -compatible if the bilinear form is a positive inner product on , . The space of -compatible structures is nonempty and will be denoted by . For each , the pair determines a natural almost complex structure in the symplectization of , given by (23). We consider -holomorphic curves , where the domain is the Riemann sphere with a finite set of punctures removed. Due to results of [14] and [18], if satisfies an energy condition, it approaches closed orbits of near the punctures. We postpone the precise definitions and statements to Section 2. In what follows we use the notation .
Now we state the main results of this paper.
Theorem 1.3.
Let be a tight contact form on . Assume that there exist Reeb orbits that are nondegenerate, simple, and have Conley-Zehnder indices respectively , , and . Assume further that the orbits , , and are unknotted, and are not linked for , , and the following conditions hold:
-
(i)
Every orbit in having period is nondegenerate;
-
(ii)
is the unique orbit in with Conley-Zehnder index and period less than that is not linked to ;
-
(iii)
is the unique orbit in with Conley-Zehnder index and period less than that is not linked to ;
-
(iv)
There exists such that the almost complex structure admits a finite energy plane asymptotic to at it positive puncture ;
-
(v)
There is no -embedding such that and each hemisphere is a strong transverse section.
Then there exists a foliation adapted to with binding orbits , , and . Consequently, there exists at least one homoclinic orbit to .
Remark 1.4.
We expect to weaken hypothesis (iv) in Theorem 1.3 by assuming the existence of a disk with boundary and interior transverse to the Reeb vector field. It is expected that an analysis similar to that of [6, 8] will show that real-analytic flows admitting a foliation have positive topological entropy if no two branches of the stable and unstable manifold of coincide.
Condition (v) in Theorem 1.3 is necessary to the existence of a foliation. This is the content of Proposition 1.5 below.
Proposition 1.5.
Assume that there exists a foliation adapted to the contact form on and let be the binding orbit with Conley-Zehnder index , as in Definition 1.2. Then there is no -embedding such that and each hemisphere is a strong transverse section.
The following theorem gives another set of sufficient conditions for the existence of a foliation. Some hypotheses are more restrictive than the hypotheses of Theorem 1.3, but we do not assume any non-degeneracy condition for the contact form.
Theorem 1.6.
Let be a tight contact form on . Assume that there exist Reeb orbits that are nondegenerate, simple, and have Conley-Zehnder indices respectively , , and . Assume further that the orbits , , and are unknotted, and are not linked for , , and the following conditions hold:
-
(i)
;
-
(ii)
If satisfies and , then .
-
(iii)
There exists such that the almost complex structure admits a finite energy plane asymptotic to at its positive puncture and a finite energy cylinder asymptotic to at its positive puncture and at its negative puncture ;
-
(iv)
There exists no -embedding such that and each hemisphere is a strong transverse section.
Then there exists a foliation adapted to with binding orbits , , and . Consequently, there exists at least one homoclinic orbit to .
An interesting application of Theorem 1.6 is in the study of bifurcations of finite energy foliations. Consider a one-parameter family of Reeb flows on the tight -sphere admitting adapted open books with disk-like pages. Hypotheses (i) and (ii) can be checked if the orbits , , and bifurcate from the binding orbit at some parameter value. See Remark 1.8 for an example of this phenomenon. The study of more general bifurcations of finite energy foliations is a work in progress of the author, P. Salomão, and A. Schneider. When one of the binding orbits of a finite energy foliation bifurcates into a finite set of binding orbits, it is expected that the transverse foliation bifurcates accordingly and the new Reeb orbits become part of the binding set.
1.2. An example of Reeb flow admitting a foliation
Consider with coordinates and equipped with the canonical symplectic form Consider the Hamiltonian function defined by
(4) |
(5) | ||||
If , then has exactly three critical points: a saddle , a local maximum and a minimum , where . See Figure 2.

The energy level is star-shaped: If and is the radial vector field , then, for sufficiently small , we have
It follows that is diffeomorphic to and is a contact form on . The Hamiltonian vector field associated to is given by
(6) |
where
(7) | ||||
The Reeb vector field is given by where
(8) |
Consider the Reeb orbits , and where
(9) |
(10) |
Note that , which implies that
(11) |
Since , for each , we have
(12) |
for sufficiently small . It is easy to see that is a trivial link.
Proposition 1.7.

Remark 1.8.
If and is sufficiently small, is strictly convex, which implies (see [19]) that is dynamically convex. The simple Reeb orbit with image has Conley-Zehnder index , self-linking number , and is unknotted. It follows from the main statement of [23] that is binding of an open book decomposition with disk-like pages adapted to , where each page is a global surface of section for the associated Reeb flow. The orbits , and bifurcate from when .
1.3. Outline of the main arguments
The paper is organized as follows. In Section 2, we review some facts about contact geometry, Conley-Zehnder indices, and pseudo-holomorphic curves in symplectizations, which will be used throughout the paper. The proof of Theorem 1.3 is split into Propositions 3.1, 4.1, 5.1, and 5.15, proved in Sections 3, 4, and 5, respectively. Proposition 1.5 is proved in Section 6, Theorem 1.6 in Section 7, and Proposition 1.7 in Section 8.
In the following, we sketch the main steps of the proof of Theorems 1.3 and 1.6. By hypothesis (iv) of Theorem 1.3, there exists an -invariant almost complex structure admitting a finite energy plane asymptotic to . By the results of [17] and the Fredholm theory developed in [20], after a quotient by the natural -action, the plane lives in a one-dimensional family , where each projection is an embedding transverse to the Reeb vector field. In Proposition 3.1, using the linking hypotheses and bubbling-off analysis, we show that the family breaks in both ends into two-level holomorphic buildings, each consisting of a rigid cylinder from to and a rigid plane asymptotic to . We use hypothesis (v) and the intersection theory developed in [31] to show that the rigid plane to is common to the two ends.

The projection of the one-parameter family of planes , the two rigid cylinders , and the rigid plane to determine a foliation of a closed region homeomorphic to a solid torus. See figure 4.
The next step of the proof is Proposition 4.1, where we obtain a rigid cylinder from to . We consider a non-cylindrical symplectic cobordism between and , where is a dynamically convex contact form, adapting the ideas of [19]. We define an non -invariant almost complex structure on the cobordism such that on . After an -translation, the plane is -holomorphic. By the Fredholm theory of [20], the plane lives in a one-dimensional family of -holomorphic planes asymptotic to . Using hypothesis (v) and the SFT compactness theorem, we show that the family breaks into a two-level holomorphic building, where the first level is pseudo-holomorphic for the original -invariant almost complex structure. Using the linking hypotheses and intersection theory, we show that the first level is a cylinder from to and projects onto an embedded cylinder in the complement of the region .

Next, we glue the cylinders and to obtain an one-dimensional family of -holomorphic cylinders from to . Using hypotheses (ii) we prove that this family breaks into a two-level building formed by the cylinders and . In Proposition 5.1 we prove that the cylinders complete the foliation.
We show that the orbits , and have self-linking number in Proposition 5.15. This follows for the orbits and since these orbits bound disks transverse to the Reeb flow. For the orbit we produce a disk gluing the cylinder and the disk .
The existence of the foliation and the arguments in [21, Proposition 7.5] imply the existence of at least one homoclinic orbit to . For completeness, we sketch a proof of the existence of a homoclinic here. Our argument follows [7, §2] and [8, §4].
Consider the one-parameter family of disks , , the one-parameter family of cylinders , , and the cylinders as in Definition 1.2. The local unstable manifold intersects transversally, for close to , in an embedded circle bounding an embedded closed disk with -area . The local stable manifold intersects transversally, for close to , in an embedded circle bounding an embedded closed disk with -area . All points in correspond to trajectories that exit through the cylinder . In the same way, intersects , for close to , in an embedded circle such that is the boundary of a closed region with -area . All points in correspond to trajectories entering through the cylinder .
The existence of a adapted to implies that the forward flow sends , into a disk inside , for close to . If intersects , then, since both disks have the same area, their boundaries also intersect, and a homoclinic to exists. Otherwise, and the forward flow sends into a disk , for close to , with -area . The forward flow sends into a disk inside . If intersects , then, since the area of is , their boundaries also intersect and there exists a homoclinic to . Otherwise, is contained in and the forward flow sends into a disk , which has area and is disjoint from . Proceeding this way, we construct disks , , for close to , and , , for close to . This procedure must end at some point since the disks and are disjoint and have the same area , while the areas of and are and , respectively. This implies that either intersects or intersects , for some integer . In any case, a homoclinic to must exist.
To prove Theorem 1.6, we use hypothesis (iii) to obtain a one-dimensional family of pseudo-holomorphic cylinders from to . Using bubbling-off analysis and hypotheses (i) and (ii), we show that this family breaks in both ends into a two-level holomorphic building consisting of a rigid cylinder from to and a rigid cylinder from to .
2. Preliminaries
In this section, is a contact form on a -manifold , and is the associated contact structure.
2.1. The Conley-Zehnder index
Let be the Reeb flow associated to the contact form . The bilinear form turns into a symplectic vector bundle and the linearized flow is symplectic with respect to .
Let be a nondegenerate Reeb orbit. Let be a trivialization of the symplectic vector bundle and consider the arc of symplectic matrices defined by Let and let be a continuous argument for , that is, . Define the winding number of by
(13) |
and the winding interval of the arc by The interval is compact and its length is strictly smaller than . Since is nondegenerate, we have , see [23, §2]. Thus, the winding interval either lies between two consecutive integers or contains precisely one integer. The Conley-Zehnder index of the orbit relative to the trivialization is defined by
(14) |
This index only depends on the homotopy class of the trivialization .
Throughout the paper, we only deal with the tight contact structure on , which is a trivial symplectic bundle. In this case, the Conley-Zehnder index is independent of the choice of global symplectic trivialization of . We define the Conley-Zehnder index of the Reeb orbit by
(15) |
for any global symplectic trivialization .
We say that the Reeb orbit is positive hyperbolic if , negative hyperbolic if , or elliptic if the eigenvalues are in . An orbit is positive hyperbolic if and only if it has even Conley-Zehnder index.
The following lemma is a consequence of the properties of the Conley-Zehnder index proved in [21].
Lemma 2.1.
Let be a tight contact form on and let be a nondegenerate Reeb orbit. Let be an integer such that for every , the orbit is nondegenerate. The following assertions hold.
-
(1)
If , then , ;
-
(2)
, ;
-
(3)
If is a hyperbolic orbit, then , ;
-
(4)
If , then and is hyperbolic. If , then .
2.2. The asymptotic operator
Fix and let be a vector field along . The Lie derivative of with respect to is defined by
(16) |
where is the flow of . Let be a symmetric connection on . We can use to write
(17) |
where is the covariant derivative along . We conclude that the differential operator maps sections of to sections of and is independent of the choice of symmetric connection.
Choosing a -compatible complex structure , we associate to the orbit the unbounded differential operator
(18) | ||||
The operator defined by (18) is called the asymptotic operator associated to the orbit and the complex structure . In any unitary trivialization of , the operator takes the form where is a path of symmetric matrices given by and is the linearized flow restricted to in the trivialization . If is an eigensection of with corresponding eigenvalue , then satisfies for all . It follows that has a well defined winding number given by
(19) |
This definition just depends on the homotopy class of the trivialization . The following properties about this winding number are proved in [17].
Proposition 2.2.
[17] The unbounded operator has discrete real spectrum accumulating only at . Let be a unitary trivialization of . Then
-
•
Given nonzero eigensections and associated to the same eigenvalue , we have , so that we can define , for any eigensection associated to .
-
•
If satisfy and are non-vanishing -eigensections, respectively, then are pointwise linearly independent.
-
•
Given , there exists precisely two eigenvalues , counting multiplicities, such that
-
•
If and , then .
-
•
if and only if the orbit is nondegenerate.
Define
(20) |
(21) |
and given a trivialization of , define
Definition 2.3.
The (generalized) Conley-Zehnder index of the orbit relative to the unitary trivialization is defined by
(22) |
2.3. Self-linking number
The self-linking number of a trivial knot transverse to is defined as follows. Consider oriented by . Let be an embedded disk satisfying and let be a smooth nonvanishing section of . The section is used to slightly perturb to another knot transverse to and , where is any exponential map. A choice of orientation for induces orientations of and . The self-linking number of is defined by where is the oriented intersection number of and . If, for instance, is trivial, this definition is independent of the choices of and . If is an unknotted Reeb orbit, we define its self-linking number by .
2.4. Finite energy surfaces
The symplectization of the contact manifold is the symplectic manifold , where is the coordinate on . Given a complex structure , we consider the almost-complex structure on defined by
(23) |
where we see and as -invariant objects on . It is easy to check that the almost complex structure defined by (23) is -compatible.
Let be a closed Riemann surface and let be a finite set. Let be a -holomorphic map, that is, is smooth and satisfies the Cauchy-Riemann equation
The Hofer energy of is defined by
where .
Definition 2.4.
The -holomorphic map is called a finite energy surface if it satisfies .
The elements of are called punctures. Let be a puncture and take a holomorphic chart centered at . We call positive exponential coordinates and negative exponential coordinates around . Set , for . Write . Using Stokes Theorem, one can prove that the limit
(24) |
exists. The puncture is called removable if , positive if and negative if . By an application of Gromov’s removable singularity theorem [12], one can prove that can be smoothly extended to a removable puncture. Thus, in the following we assume that all punctures are positive or negative and use the notation to distinguish positive and negative punctures. If is a finite energy surface, then .
Theorem 2.5.
We say that the periodic orbit given by Theorem 2.5 is an asymptotic limit of at the puncture . If is nondegenerate, it is called the asymptotic limit of at .
The -area of a -holomorphic curve is given by the formula
One can check that and if and only if . The following theorem concerning -holomorphic curves with vanishing -area will be useful throughout the paper.
Theorem 2.6.
[17, Theorem 6.11] Let be a finite energy sphere, where is the finite set of negative punctures and is the unique positive puncture. If , then there exists a nonconstant polynomial and a Reeb orbit such that and , where is defined by .
2.5. Asymptotic behavior
A Martinet’s tube for a simple orbit is a pair , where is a neighborhood of in and is a diffeomorphism (here is an open ball centered at the origin) satisfying
-
•
There exists such that , and , where is the coordinate on and are coordinates on ;
-
•
.
The coordinates are called Martinet’s coordinates. The existence of such Martinet’s tubes is proved in [18] for any simple orbit .
A more precise description of the asymptotic behavior of a finite energy surface is given in [18]. Let be a finite energy surface asymptotic to a nondegenerate orbit at the positive puncture and let be positive exponential coordinates near . Let be a positive integer such that , where is the least positive period of and let be Martinet’s coordinates in a neighborhood of . Then there exists such that, for , the map can be represented in Martinet’s coordinates by
where is seen as a map on the universal cover of satisfying and is -periodic in .
Theorem 2.7 ([18]).
If , there are constants , , a function and an eigensection of the asymptotic operator (18), associated to a negative eigenvalue , such that
(25) | |||
for all large and . Here represents the eigensection in the coordinates induced by , and is a smooth function such that .
A similar statement holds if is a negative puncture. In this case, we use negative exponential coordinates near , is replaced by and the eigenvalue of is positive.
The eigenvalue and the eigensection , as in Theorem 2.7, will be referred to as the asymptotic eigenvalue and asymptotic eigensection of at the puncture .
Remark 2.8.
Let be a finite energy surface that is asymptotic to nondegenerate simple orbits at all of its punctures and such that is not identically zero. By Theorem 2.7, the surface is a strong transverse section according to Definition 1.1. Indeed, for every puncture , the asymptotic eigensection of at satisfies
2.6. Algebraic invariants
Let be a -holomorphic finite energy surface. Assume that is not identically zero and that has nondegenerate asymptotic limits at all of its punctures. In [17], it is proved that the set where vanishes is finite, and it is defined a local degree associated to each zero of , which is always positive. The integer
(26) |
is defined as the sum of such local degrees over all zeros of .
Consider a unitary trivialization . For , fix positive cylindrical coordinates at and define
(27) |
where is the sign of the puncture . The winding number on the right is defined as in (19), and is independent of the choice of and of the holomorphic chart. This limit is well defined since does not vanish for sufficiently large. The asymptotic winding number of is defined by
(28) |
It is proved in [17] that this sum does not depend on the chosen trivialization .
Remark 2.9.
Throughout the paper we will only deal with the tight contact structure on , which is a trivial symplectic bundle. Consider a global symplectic trivialization Then is the winding number of the asymptotic eigensection given by Theorem 2.7, with respect to the trivialization . Moreover, does not depend on the chosen global symplectic trivialization and we denote by .
It is also proved in [17] that the invariants and satisfy
(29) |
where is the Euler characteristic of .
2.7. Fredholm theory
Let be a -holomorphic finite energy surface and assume that has nondegenerate asymptotic limits at all of its punctures. Let be a symplectic trivialization. Fix a puncture and let be the asymptotic limit of at . The trivialization induces a homotopy class of oriented trivializations of . The Conley-Zehnder index of is defined by
(30) |
It is proved in [17] that this sum does not depend on the chosen trivialization .
An unparametrized finite energy surface is an equivalence class , where is a -holomorphic finite energy surface and is an ordered set. The equivalence class is defined as follows: is equivalent to if there exists a biholomorphism such that , and preserves the ordering of and . To shorten notation we usually denote an unparametrized surface by .
In [9] it is proved that the set of unparametrized finite energy surfaces in the neighborhood of , where is a somewhere injective finite energy surface, is described by a nonlinear Fredholm equation having Fredholm index equal to
(31) |
This generalizes the result for embedded finite energy surfaces proved in [20].
Theorem 2.11.
[9] There exists a dense subset such that, if is a somewhere injective finite energy surface which is pseudo-holomorphic with respect to and has nondegenerate asymptotic limits at all of its punctures, then
If is not identically zero, then
2.8. Intersection theory of punctured pseudo-holomorphic curves
In this section we state some results concerning the intersection theory of punctured pseudo-holomorphic curves from [31]. A nice exposition of these results is given in [11]. Here we use the same convention as [11] for computing Conley-Zehnder indices and the asymptotic winding number , which differs from that of [31]. We assume that all the finite energy curves considered have nondegenerate asymptotic limits.
In the following, and are finite energy -holomorphic curves and denotes a choice of trivialization of the contact structure along all simple periodic orbits with covers appearing as asymptotic limits of or .
Theorem 2.12.
[31, Theorem 2.4] Assume that no component of or lies in an orbit cylinder and that the projected curves and do not have identical image on any component of their domains. Then the following are equivalent:
-
(1)
The projected curves and do not intersect;
-
(2)
All of the following hold:
-
–
The map does not intersect any of the positive asymptotic limits of ;
-
–
The map does not intersect any of the negative asymptotic limits of ;
-
–
If is a periodic orbit so that, at , is asymptotic and, at , is asymptotic to , then: If and are both positive or both negative punctures, we have If is a negative puncture and is a positive puncture, we have
-
–
-
(3)
All of the following hold:
-
–
The map does not intersect any of the asymptotic limits of .
-
–
The map does not intersect any of the asymptotic limits of .
-
–
If is a periodic orbit so that, at , is asymptotic to and, at , is asymptotic to , then .
-
–
A set of necessary and sufficient conditions for the projection of a curve to be embedded are given in Theorem 2.6 of [31]. The following statement is a direct consequence of [31, Theorem 2.6] that is enough for our purposes.
Theorem 2.13.
[31] Assume that is somewhere injective, connected and does not have image contained in an orbit cylinder. Assume further that is not asymptotic to a covering of the same simple orbit at two distinct punctures. Then the following are equivalent:
-
(1)
The projected map is an embedding.
-
(2)
does not intersect any of its asymptotic limits.
-
(3)
All of the following hold:
-
•
The map is an embedding.
-
•
The projected map is an immersion which is everywhere transverse to the Reeb vector field.
-
•
For each , we have .
-
•
Next we recall the notion of two pseudo-holomorphic curves approaching an orbit in the same (or opposite) direction. The definition given here follows [11]. It applies to any nondegenerate orbit and is stricter than the definition from [31].
Definition 2.14 ([31], [11]).
Assume that and are asymptotic to the same Reeb orbit at certain punctures and with the same sign. We say that and approach in the same direction at these punctures if for some , where and are the asymptotic eigensections of at and of at respectively. In case for , we say that and approach in opposite directions.
The following theorem of [11] is a consequence of results on [31] concerning the intersection properties of pseudo-holomorphic curves approaching an even orbit in the same direction.
Theorem 2.15.
[11, Theorem 3.16] Assume that and are asymptotic to an even Reeb orbit at punctures and with the same sign. Assume that and have extremal winding number, that is,
and that and approach in the same direction. Then the projections and intersect.
3. Foliating a solid torus
In this section, we start the proof of Theorem 1.3 with the following statement.
Proposition 3.1.
Let be a tight contact form on satisfying the hypotheses of Theorem 1.3. Then there exists a family of -holomorphic planes , all of them asymptotic to , a pair of -holomorphic cylinders , both asymptotic to at their positive punctures and at their negative punctures , and a finite energy plane asymptotic to . The projections are embeddings transverse to the Reeb vector field. Define , , and . Then the surface is homeomorphic to a torus and is -embedded. The union of the family with , and determine a smooth foliation of , where is a closed region with boundary and homeomorphic to a solid torus.
By hypothesis, there exists a finite energy -holomorphic plane
(32) |
asymptotic to the orbit . Both the -holomorphic plane and the projection are embeddings. This is a consequence of the following result, which is a particular case of Theorem 1.3 of [17].
Theorem 3.2 ([17]).
Consider equipped with a tight contact form . Assume is a finite energy plane asymptotic to an unknotted orbit . If , then and is an embedding.
The finite energy plane is automatically Fredholm regular. If is a -small perturbation of , where is the dense set obtained by Theorem 2.11, then we can find a -holomorphic plane asymptotic to as a small perturbation of . It follows that we can revert the notation back to and and assume that . This assumption will be necessary since we will use Theorem 2.11 in the proof of Theorem 1.3.
3.1. A family of planes asymptotic to
Theorem 3.3 ([32]).
Let be a contact form on a closed -manifold and let . Let is an embedded -holomorphic finite energy sphere such that every asymptotic limit is nondegenerate, simple and has odd Conley-Zehnder index. Suppose that . Then there exists a number and an embedding
such that
-
•
For and , the maps are (up to parametrization) embedded -holomorphic finite energy spheres and .
-
•
The map is an embedding and its image never intersects the asymptotic limits. In particular, the maps are embeddings for each , with mutually disjoint images which do not intersect their asymptotic limits.
-
•
For any sequence such that for each puncture in , has the same asymptotic limit as , with the same sign, and in , there is a sequence such that for some sequence of biholomorphisms and sufficiently large.
Remark 3.4.
In Theorem 3.3, we do not require .
Applying Theorem 3.3 to the finite energy plane (32) we obtain a maximal one-parameter family of finite energy planes
(33) |
asymptotic to the orbit . The family (33) satisfies Indeed, suppose that there exist such that . Then (see Theorem 2.12 or [17, Theorem 1.4]) and we would obtain an -family of embedded planes that provides an open book decomposition of , with binding and disk-like pages. It follows from equation (29), Lemma 2.10 and formula (22) that As a consequence of the definition of , we conclude that is transverse to , . This implies that every Reeb orbit in is linked to , which contradicts the existence of and .
Now we describe how the family breaks as . We assume that and and that strictly increases in the direction of .
Proposition 3.5.
Consider a sequence in the family (33) satisfying . Then there exists a -holomorphic finite energy cylinder , which is asymptotic to at the positive puncture and to at the negative puncture , and a finite energy -holomorphic plane asymptotic to at , such that, after suitable reparametrizations and -translations of , the following hold
-
(i)
up to a subsequence, in as .
-
(ii)
There exist sequences , and such that, up to a subsequence, in as .
Here . A similar statement holds for any sequence satisfying . In this case we change the notation from and to and respectively.
3.2. Bubbling-off analysis
By hypotheses, every orbit in having period is nondegenerate. Consequently, there exists only a finite number of such orbits. Define as any real number satisfying
(34) |
3.2.1. Germinating sequences
Now we fix and consider a sequence of -holomorphic curves satisfying
(35) | |||
(36) | |||
(37) | |||
(38) |
Definition 3.6.
Proposition 3.7.
There exists a finite set , a -holomorphic map and a subsequence of , still denoted by , such that and .
Proof.
Let be the set of points such that there exists a subsequence and a sequence with and If , then by (38) and usual elliptic estimates, see [28, Chapter 4], we find a -holomorphic map such that, up to a subsequence, in . In this case, .
Now assume and let . It follows from results in [14, §3.2] that there exists a period and sequences , and such that
Consider as the new sequence . Now let be the set of points such that there exists a subsequence and sequence with and . As before, if , we have a -holomorphic map such that, up to a subsequence, in . In this case, we define . If and , there exist a period and sequences , and such that
Considering as the new sequence , define as before. Repeating this argument, let . Note that
It follows that there exists such that and for . We end up with a finite set and a -holomorphic map such that, up to a subsequence,
Definition 3.8.
A -holomorphic map as in Proposition 3.7 is called a limit of the germinating sequence .
If , then is nonconstant. In this case, all the punctures are negative and is a positive puncture. To prove this, define, for any ,
(39) |
where is oriented counterclockwise. Then
For large, , defined as in the proof of Proposition 3.7, is contained in . It follows that
This implies that is nonconstant and the puncture is negative. Moreover, as a consequence of , we know that has at least one positive puncture. Thus, is a positive puncture.
3.2.2. Soft-rescaling near a negative puncture
Assume and let be a limit of the germinating sequence . Let . We define the mass of by
(40) |
where is the period of the asymptotic limit of at . Since is a nondecreasing function of , we can fix small enough so that
(41) |
Choose sequences and , so that
(42) | |||
(43) |
Since is a negative puncture, (42) implies that . Hence the existence of as in (43) follows from (40). We claim that . Otherwise, we choose . From (41), we get the contradiction
Thus, we can assume that .
Now take any sequence satisfying and define the sequence of -holomorphic maps by
(44) |
It follows from (43) that
Moreover, by the definition of , and . Thus, is a germinating sequence.
Let be a limit of , as in Proposition 3.7. If , then is not constant. If , then
(45) | ||||
Thus is nonconstant as well. From Fatou’s Lemma we get . This also implies that the periods of the asymptotic limits of are bounded by .
Proposition 3.9.
The asymptotic limit of at coincides with the asymptotic limit of at the negative puncture .
To prove Proposition 3.9, we need the following lemma, which is a restatement of Lemma 4.9 from [21].
Lemma 3.10.
[21] Consider a constant and let be defined by (34). Identifying , let be an open neighborhood of the set of periodic orbits with , viewed as maps , . We assume that is -invariant, meaning that , and that each of the connected components of contains at most one periodic orbit modulo -reparametrizations. Then there exists a constant such that the following holds. If is a -holomorphic cylinder satisfying
then each loop is contained in for all .
Proof of Proposition 3.9.
Let be as in the statement of Lemma 3.10. Let and be connected components of containing and respectively. Since , we can choose small enough so that, if is fixed, then the loop
belongs to for large . Since , we can choose large enough so that, if is fixed, then the loop
belongs to for large . By (40) and (43), we can show that
(46) |
Consider, for each , the -holomorphic cylinder , defined by . It follows from (43) that
(47) |
for large . Using (46) and (47) and applying Lemma 3.10, we find so that the loop
is contained in for all and large . But
Thus, and . ∎
Proposition 3.11.
Either
-
•
or
-
•
and .
3.2.3. Some index estimates
Lemma 3.12.
Let be a finite energy surface such that every puncture in is negative, does not vanish identically and for every asymptotic limit of , is nondegenerate and . Then and for all , .
Proof.
The following estimate is proved in [25].
3.3. Proof of Proposition 3.5
After a reparametrization of and an translation in the -direction, we can assume that is a germinating sequence as defined in (35)-(38). Indeed, we can take sequences and , such that and , and define
Then , and
(50) |
Let be a limit of as defined in 3.8. We claim that is the asymptotic limit of at . Indeed, let be as in the statement of Lemma 3.10. Using the normalization condition (37), we can apply Lemma 3.10 and find such that for , the loops and belong to . For fixed , the sequence of loops converges to in , so that for large and , and belong to the same connected component of . This implies that is the asymptotic limit of at .
Note that and . Indeed, if , then would satisfy the hypotheses of Theorem 3.3, which contradicts the fact that the family (33) is maximal. Since and the punctures in are negative, we have .
We claim that . Suppose, by contradiction, that . Using Theorem 2.6 and the fact that is asymptotic to the simple orbit at its positive puncture , we conclude that and . This leads to the contradiction
Here we have used (50).
Now we prove that . Since is simple, we know that is somewhere injective. By Theorem 2.11, we have the estimate
where is the asymptotic limit of at . By Lemma 3.13, we have , . Thus, the only possibility is and , where be the asymptotic limit of at .
Our next claim is that is asymptotic to at . Since, for each , is an embedding whose image does not intersect , it follows that the image of any loop under is not linked to . This implies that the image of any loop under is not linked to and consequently that is not linked to . Moreover, . We conclude that .
As proved in §3.2.2, there exist sequences , and such that the sequence of -holomorphic maps defined by
is a germinating sequence. Let be a limit of the sequence . Then is asymptotic to at . By Lemma 3.13, we conclude that where is the asymptotic limit of at . Since is simple, we know that is somewhere injective. We can apply Theorem 2.11 to and obtain
which implies . By Proposition 3.11, we have . Again by Theorem 2.11, we have and consequently . We conclude that is a finite energy plane asymptotic to the orbit . This finishes the proof of Proposition 3.5.
3.4. The foliation
Proposition 3.14.
Consider a sequence in the family (33) satisfying . Let and be the finite energy spheres obtained in Proposition 3.5.
-
(i)
Given an -invariant neighborhood of the loop , there exists such that, for and large , the loop belongs to .
-
(ii)
Given an -invariant neighborhood of the loop , there exist and such that, for and large , the loop belongs to .
-
(iii)
Given any neighborhood of , we have for large .
A similar statement works for any sequence with and replaced by and respectively.
Proof.
We can assume that , contains only the periodic orbit modulo -reparametrizations. Let be an -invariant neighborhood of the set of periodic orbits with , viewed as maps , , such that . Using the normalization condition (50) and Lemma 3.10, we find such that, for , the loops belong to . By the asymptotic behavior of the planes , we conclude that belong to for . This proves (i).
Applying Lemma 3.10 as in the proof of Proposition 3.9, we find small and such that for every satisfying and large , the loop belongs to . This proves (ii).
The proof of (iii) follows from (i), (ii) and Proposition 3.5.∎
Proposition 3.15.
Let and be the finite energy spheres obtained in Proposition 3.5. Then the projected curves , , and are embeddings which are transverse to the Reeb vector field and do not intersect .
Proof.
We first prove that . Consider defined by . Note that is a finite energy -holomorphic immersion. By Carleman’s similarity principle, the intersections of with are isolated. By positivity and stability of intersections of pseudo-holomorphic curves, any such intersection implies intersection of with for large , contradicting Theorem 3.3. To prove that , we proceed in the same way, noting that . Indeed, and are not linked and is transverse to the Reeb vector field.
In the same way, we prove that , and do not intersect . Theorem 2.13 shows that , , and are embeddings which are transverse to the Reeb vector field. ∎
Proposition 3.16.
Let and be the finite energy spheres obtained in Proposition 3.5. Then, up to reparametrization and -translation, , and is the unique finite energy -holomorphic plane asymptotic to . If , then (up to reparametrization and -translation) and are the unique finite energy -holomorphic cylinders asymptotic to and at and respectively that do not intersect . Moreover, and approach in opposite directions according to Definition 2.14.
Proof.
Our proof follows [7, Proposition C.1]. First we prove that and coincide up to reparametrization and -translation. Following Theorem 2.7, let be the asymptotic eigensection of at and the asymptotic eigensection of at . It follows from Lemma 3.12 that
(51) |
Since, by Proposition 2.2 and formula (22), is the unique negative eigenvalue of with winding number equal to , it follows that and are -eigensections. Since the eigenspace of is one dimensional, we find a constant such that Suppose, contrary to our claim, that . Then, by (51), Proposition 3.15 and Theorem 2.12, we have
(52) |
By (51), (52) and Theorem 2.15, we conclude that , that is, and approach in opposite directions. This implies that is a -embedded sphere, where each hemisphere is a strong transverse section (see Remark 2.8), which is a contradiction with the hypotheses of Theorem 1.3. We have proved that and coincide up to reparametrization and -translation.
Using the same arguments above and noting that, by Theorem 3.2, any -holomorphic plane asymptotic to is embedded and does not intersect , we prove that any finite energy -holomorphic plane asymptotic to coincides with up to reparametrization and -translation.
Now we prove the assertions about the cylinders and . Let and be the asymptotic eigensections of and at , respectively. By Lemma 3.12, we have
(53) |
Using Proposition 2.2 and formula (22), we conclude that and are -eigensections. Since the eigenspace of is one dimensional, we conclude that there exists a nonzero constant such that . Assume that . Then, by (53), Proposition 3.15 and Theorem 2.12, we have
(54) |
By (53), (54) and Theorem 2.15, we conclude that , that is, and approach in opposite directions.
By the same arguments above, we conclude that and are the unique cylinders with the properties given in the statement. ∎
Proposition 3.17.
Let and be the finite energy spheres obtained in Proposition 3.5. Then the cylinders and satisfy , the surface is homeomorphic to a torus and is -embedded. The union of the image of the family , given by (33) with the images of , and determine a smooth foliation of , where is a closed region with boundary . Moreover, is homeomorphic to a solid torus.
Proof.
By Proposition 3.15 and Theorem 2.12 we conclude that the images of the projected curves , and , are mutually disjoint. Moreover, if , then the images of , , and are mutually disjoint.
Now we prove that . Suppose, by contradiction, that the images of and have a nonempty intersection. Then we have . Let and consider a neighborhood of such that . By Proposition 3.14, we have for some sufficiently close to and sufficiently close to . The surface is a piecewise smooth embedded sphere. By Jordan-Brouwer separation theorem, divides into two disjoint regions and with boundary . One of these regions, say , contains and the other contains . The intersection of the image of the family with is nonempty, open and closed in . Thus is in the image of the family . We conclude that
This contradicts the fact that the orbit is not linked to and the curves and are transverse to the Reeb vector field. We have proved that Consequently, the surface is a piecewise smooth embedded torus, and it follows from Proposition 3.16 that is -embedded.
By Jordan-Brouwer separation theorem, divides into two closed regions and with disjoint interiors and boundary . One of the regions, say , contains the image of the family and the plane . Now we show that
Let and let be a neighborhood of such that . By Proposition 3.14, we have for some sufficiently close to and sufficiently close to . The surface is a piecewise smooth embedded sphere. By Jordan-Brouwer separation theorem, divides into two disjoint regions and with boundary . One of these regions, say contains and is contained in . Thus the intersection of the image of the family with is open, closed and nonempty in . This implies that is in the image of , for some .
To prove that is homeomorphic to a solid torus, we follow the proof of the Solid torus theorem in [30]. The curve is the boundary of an embedded 2-disk contained in such that . Let be a bicollar neighborhood of in , so that is an annular neighborhood of in . The boundary of is the union of two disks on and the set , which is an annulus. Thus is bounded by a piecewise smooth -sphere in . By the generalized Schönflies theorem, is homeomorphic to a closed -ball. It follows that is homeomorphic to a -ball with a attached, where is a closed 2-disk (by mapping and onto disjoint disks on the boundary of the -ball). Since is orientable, is homeomorphic to . ∎
The proof of Proposition 3.1 is complete.
4. A cylinder asymptotic to and
In this section, we continue the proof of Theorem 1.3. Recall that we have obtained a foliation of a closed region . Now we find a finite energy cylinder asymptotic to at its positive puncture and at its negative puncture, whose projection to is contained in the complement of .
Proposition 4.1.
Let be a tight contact form on satisfying the hypotheses of Theorem 1.3. Then there exists a finite energy cylinder asymptotic to at its positive puncture and at its negative puncture . The projection is an embedding transverse to the Reeb vector field. Define . Then and is a -embedded disk, where is given by Proposition 3.1.
The contact structure coincides, up to a diffeomorphism, with , where is defined by (2) and is a smooth function. Thus, we can assume that for some smooth function without loss of generality. Following [19] we define a symplectic cobordism between and , where is a dynamically convex contact form on . Given , with irrational, let be the contact form associated to the ellipsoid
that is, . The Reeb vector field defined by has precisely two simple periodic orbits and . Both periodic orbits and its iterates are nondegenerate. Their Conley-Zehnder indices are and where is determined by . See [16, Lemma 1.6] for a proof of these facts.
We choose small enough so that and a smooth function satisfying
(55) | ||||
(56) |
In view of (56), the -form restricted to is a symplectic form.
We consider the family of contact forms . The contact structure does not depend on . Choose and let be a smooth family of -compatible complex structures on so that if and if . We consider smooth almost complex structures on the symplectization with the following properties. On , we consider
On we only require to be compatible with the symplectic form . The space of such almost complex structures on is nonempty and contractible in the -topology and will be denoted by .
4.1. Generalized finite energy surfaces
Definition 4.2.
Let be a closed Riemann surface and let be a nonempty finite set. A nonconstant smooth map is called a generalized finite energy surface if it is -holomorphic, that is, satisfies , for some , as well as the energy condition
where the energy is defined as follows. Let be the collection of smooth functions satisfying and on . Given we define the -form on by
for any and . Then
(57) |
Theorem 2.11 is still valid for almost complex structures in .
Theorem 4.3 ([9]).
There exists a dense subset such that if is a somewhere injective generalized finite energy surface for , and has nondegenerate asymptotic limits at all of its punctures, then
4.2. A family of -holomorphic planes asymptotic to
We are interested in the space of finite energy generalized -holomorphic planes asymptotic to the orbit , for fixed . The following theorem is a consequence of results from [20].
Theorem 4.4 ([20]).
Let be an embedded finite energy -holomorphic plane, asymptotic to a nondegenerate, simple Reeb orbit satisfying . Then there exists a smooth embedding
with the following properties:
-
•
;
-
•
For every , the map is a generalized finite energy -holomorphic plane asymptotic to ;
-
•
If is a sequence of finite energy -holomorphic planes asymptotic to satisfying in as , then there exist sequences in with , and in with such that
for sufficiently large .
From now on we fix , where is given by Theorem 4.3. Let be the space of generalized finite energy -holomorphic planes asymptotic to , modulo holomorphic reparametrizations. By Theorem 4.4, is a smooth -dimensional manifold.
Lemma 4.5.
The space is nonempty.
Proof.
Let be the -holomorphic plane asymptotic to given by Theorem 3.5. After an -translation we can assume that , so that can be viewed as a generalized finite energy -holomorphic plane asymptotic to . ∎
By the proof of Lemma 4.5, we have . Let be the connected component of containing .
4.3. Existence of a -holomorphic cylinder
Consider a sequence of generalized finite energy planes representing elements of . The energy is uniformly bounded by . The following statement adapted from [26] is a corollary of the SFT compactness theorem of [2].
Theorem 4.6 ([26],Theorem 3.11).
Up to a subsequence of , still denoted by , there exists a bubbling-off tree with the following properties
-
•
For every vertex of there exist sequences and such that
(58) Here , where and
-
•
The curve is asymptotic to at and the asymptotic limits of all curves are closed orbits with periods of the Reeb flow of either or .
Here, a bubbling off tree consists of a finite, rooted tree , with edges oriented away from the root, and a finite set of pseudo-holomorphic spheres , satisfying the following properties.
-
•
There is a bijective correspondence between vertices and finite-energy punctured spheres . Each is pseudo-holomorphic with respect to either , or and has finite energy.
-
•
Each sphere has exactly one positive puncture at and a finite set of negative punctures. If is or -holomorphic and its contact area vanishes, then .
-
•
Each ordered path from the root to a leaf , where is a direct descendant of , contains at most one vertex such that is -holomorphic, in which case is -holomorphic , and is -holomorphic .
-
•
If the vertex is not the root then has an incoming edge from a vertex , and outgoing edges to vertices of , respectively. The edge is associated to the positive puncture of and the edges are associated to the negative punctures of . The asymptotic limit of at its positive puncture coincides with the asymptotic limit of at its negative puncture associated to . In the same way, the asymptotic limit of at a negative puncture corresponding to coincides with the asymptotic limit of at its unique positive puncture.
The following statement is a reformulation of Lemma 3.12 from [26], adapted to our set-up.
Lemma 4.7 ([26], Lemma 3.12).
Let be sequences such that (58) holds for all vertices of . Then we can assume, up to a selection of a subsequence still denoted by , that one of the three mutually excluding possibilities holds for each vertex .
-
(I)
is bounded, is -bounded and is a -holomorphic curve;
-
(II)
, in as and is a -holomorphic curve;
-
(III)
, in as and is a -holomorphic curve.
Moreover, if is a vertex for which (III) holds, then is asymptotic at its positive puncture to a Reeb orbit having period strictly less than . In particular, (III) does not hold for the root .
The following lemma is proved using the maximum principle combined with estimates for cylinders with small area (Lemma 3.10).
Lemma 4.8.
Since is a connected -dimensional manifold without boundary, it is diffeomorphic either to or to an open interval. It can not be diffeomorphic to , since it contains the family of equivalence classes of translations of , where is the plane obtained in Theorem 3.5. Thus, the family is diffeomorphic to an interval and we can assume that , where for , . Consider a sequence satisfying .
Proposition 4.9.
The bubbling-off tree obtained as an SFT-limit of the sequence , as in Theorem 4.6, is as follows. The tree has vertices , where is the root and is a leaf and direct descendant of . The root corresponds to a -holomorphic cylinder asymptotic to at its positive puncture and to at its negative puncture . The leaf corresponds to a -holomorphic plane asymptotic to .
Proof.
Let be the bubbling-off tree given by Theorem 4.6 and let be the finite energy sphere associated to the root of . By Lemma 4.7, is not -holomorphic. Now we show that is -holomorphic.
Suppose, by contradiction, that is -holomorphic. By Theorem 4.6, is asymptotic to at . Since is simple, is somewhere injective. If , then , contradicting the fact that the interval is maximal. Assume . By Theorem 4.3, it follows that
(59) |
Then
(60) |
Contradicting the fact that for all , the asymptotic limit of at is a Reeb orbit of , which is a dynamically convex contact from, that is, , for all . Thus, is -holomorphic.
Let . We claim that
To prove the claim, suppose by contradiction that Then there exists a subsequence satisfying . By the definiton of , the planes are -holomorphic. By Proposition 3.16, we know that is the unique finite energy -holomorphic plane asymptotic to , up to reparametrization and -translation. This implies for a sequence . This contradicts the fact that is an interval.
Now we show that . Suppose, by contradiction, that is a -holomorphic plane. Let , and be the sequences given by Theorem 4.6, such that
The limit satisfies 4.7(II), so that and in as . By Lemma 4.8, there exists such that for and large enough. For , by Lemma 4.7(II), we have , for large enough. Thus, for large enough. This contradicts , and concludes the proof of .
So far, we know that is a -holomorphic sphere and . The next step is to prove that every negative asymptotic limit of has Conley-Zehnder index equal to 1.
Claim I
If is not the root and is the asymptotic limit of at , then . To prove Claim I, we argue indirectly assuming that . The curve factors as where is a somewhere injective finite energy sphere and is a polynomial. If is the asymptotic limit of at , then . By Lemma 2.1, implies . By Theorem 4.3, we have
where is the asymptotic limit of at . If , we already have a contradiction. Otherwise, there exists such that . Let be such that . Then , for some , where is the asymptotic limit of at . By Lemma 2.1, we have . Since the tree has a finite number of vertices, by induction we find a leaf of the tree such that the finite energy plane is asymptotic to an orbit with , a contradiction. This proves Claim I.
Claim II
For every , we have , where is the asymptotic limit of at . To prove Claim II, first note that, by Theorem 2.11, we have
It follows that
Since, by Claim I, we have , for every , there exists at most one puncture such that . If there exists such puncture, we have . By Theorem 2.11, this implies . By the definition of bubbling-off tree, this implies . By Theorem 2.6, there exists a periodic orbit and a polynomial such that and , where is the cylinder over the orbit , which contradicts the fact that is simple. This proves Claim II. We have also proved that .
We are now in a position to show that and is a cylinder asymptotic to at . By Lemma 3.12, we conclude that . Recall that Thus, the curves and satisfy condition (2) of Theorem 2.12. Here is the plane asymptotic to obtained by Theorem 3.5. It follows from Theorem 2.12 that the projected curve does not intersect any of the negative asymptotic limits of . This implies that is contractible in , for every asymptotic limit of at . Consequently,
Since, by hypothesis, the orbit is the only orbit with Conley-Zehnder index and period less than that is not linked to , it follows that
Applying Theorem 2.12 to the finite energy curves and , where is any of the planes in the family (33), we prove that does not intersect . We conclude that , where and is the closed region given by Proposition 3.17. The region contains an embedded disk with boundary , so that is contractible in . One can show, using Mayer-Vietoris sequence, that the holomology class of generates . The projected curve defines a singular 2-chain in , so that in , where . It follows that and we can assume that .
Now we prove that the next (and last) level of the bubbling-off tree consists of a -holomorphic plane. Let be the finite energy sphere associated to the unique vertex that is a direct descendant of the root . The asymptotic limit of at is , that is a simple orbit. It follows that is somewhere injective. By the definition of bubbling-off tree, is either -holomorphic or -holomorphic. By Theorem 4.3, we have
where is the asymptotic limit of at . Since, by Claim I, for all , it follows that and , for all .
Suppose, by contradiction, that is -holomorphic. By Theorem 2.11, we have . By the definition of bubbling-off tree, this implies . Theorem 2.6 and the fact that is simple lead to a contradiction. We have proved that is -holomorphic.
Suppose that . By the definition of bubbling-off tree, if is a vertex of the tree that is a direct descendant of , then is necessarily -holomorphic. The asymptotic limit of at is equal to for some . But for all , contradicting the fact that all closed orbits of have Conley-Zehnder index . We have proved that . This finishes the proof of Proposition 4.9.
∎
Proposition 4.10.
To prove Proposition 4.10, we need the following.
Lemma 4.11.
Let be a tight contact form on . Let be a finite energy cylinder asymptotic to nondegenerate simple Reeb orbits at and at . Assume that , is an unlink and . Then and .
Proof.
Our arguments follow the proof of Theorem 4.4 from [17], so we sketch the proof here and refer to the results of [17] when necessary.
A finite energy surface such that is not identically zero can intersect its asymptotic limits in at most finitely many points (see [18, Theorem 5.2]). This allows the definition of an algebraic intersection index as follows. Take a small embedded -disk transversal to the periodic orbit at a point of and tangent to at , that is, . We orient the disk in such a way that and have the same orientation. Let be such that all the intersection points of with are contained in . Let be a disk map to such that . Such disk exists since is a trivial link. Glue the disk to along to form a new disk . Let be the map obtained by defining and and define by
Consider . If is the generator of , there exists an integer, called , such that
Here we have used the fact that is generated by .
It follows from the proof of Theorem 4.6 in [17] that is the oriented intersection number of and . Moreover, all the intersections are in the image of the map , since there is no intersections of with . Following the proof of Theorem 4.6 in [17], one can show that and if and only if . Now we show that .
Let be an embedding of an open neighborhood of the zero section of . We require that and the fiberwise derivative of at is the inclusion of into . Consider a nonvanishing section of along which is contained in . Define the loop by for . It is contained in and we denote by the homology class generated by this loop. Fix the global trivialization . Let be the winding number of the small section with respect to the trivialization . It is proved in [17] that
is independent of the section as described above. We define a constant by
If we choose, for example, the special section such that for some large , then by definition and taking the limit as , we have
It is proved in [17] that there exists an embedded disk with whose characteristic distribution has positive elliptic points, and that
By Lemma 3.12, we have , so that
Since and , we have and . This shows that . We can repeat the arguments replacing by to show that and conclude the proof. ∎
Proof of Proposition 4.10.
From Lemma 4.11, we conclude that does not intersect its asymptotic limits. Thus, satisfies condition (2) of Theorem 2.13. We conclude that is an embedding and is an embedding transverse to the Reeb vector field. Applying Theorem 2.12 to the finite energy cylinders and , where is the cylinder obtained in Proposition 3.5, we prove that does not intersect . Similarly, we prove that does not intersect . Since , we conclude that . ∎
Proposition 4.12.
Let be the -holomorphic cylinder obtained in Proposition 4.9. Then, up to reparametrization and -translation, is the unique -holomorphic cylinder asymptotic to at and to at that does not intersect . Moreover, the cylinder and the plane obtained in Proposition 3.5 approach in opposite directions, according to Definition 2.14. Consequently, is a -embedded disk.
Proof.
The proof follows [7, Proposition C.1]. Following Theorem 2.7, let be the asymptotic eigensection of at and let be the asymptotic eigensection of at . By Lemma 3.12, we have . Using formula (22) and Proposition 2.2, we conclude that is the unique negative eigenvalue of with winding number . By Proposition 2.2, we know that the eigenspace of is one dimensional. Thus, there exists such that . By Proposition 4.10 and Theorem 2.15, we conclude that , that is, and approach in opposite directions. Using the same arguments above, we conclude that any cylinder with the properties given in the statement must have the same image as . ∎
The proof of Proposition 4.1 is complete.
5. A family of cylinders asymptotic to and
Recall that we have obtained a foliation of a closed region and a finite energy cylinder asymptotic to at its positive puncture and at its negative puncture, whose projection to is contained in . Now we construct a foliation of . More precisely, we prove the following statement.
Proposition 5.1.
Let be a tight contact form on satisfying the hypotheses of Theorem 1.3. Then there exists a family of finite energy cylinders , all of them asymptotic to at their positive punctures and at their negative punctures . The projections are embeddings transverse to the Reeb vector field and . The union of , , and determine a smooth foliation of . Here the surfaces , and are given by Propositions 3.1 and 4.1.
5.1. Gluing
The following statement is a consequence of the usual gluing theorem for pseudo-holomorphic curves in symplectizations, see [29, §7] or [34, §10] for a nice exposition.
Theorem 5.2.
Let be simple Reeb orbits such that . Let be finite energy cylinders, which are pseudo-holomorphic with respect to and such that is asymptotic to at the positive puncture and to at the negative puncture and is asymptotic to at the positive puncture and to at the negative puncture . Then there exists and a family of finite energy cylinders
asymptotic to at the positive puncture and to at the negative puncture . For every sequence satisfying , there exist sequences and such that
5.2. A family of -holomorphic cylinders
Let and be the finite energy cylinders obtained by Proposition 3.5 and Proposition 4.9 respectively. Applying Theorem 5.2 to and we obtain a family of finite energy -holomorphic cylinders
(61) |
all of them asymptotic to the orbit at the positive puncture and at the negative puncture .
Proposition 5.3.
For every , is an embedding. The projection is an embedding which does not intersect its asymptotic limits and is transverse to the Reeb vector field. Moreover, , where is the closed region defined in Proposition 3.17.
Proof.
By Lemma 4.11 and Theorem 2.13, we conclude that is an embedding and is an embedding which does not intersect its asymptotic limits and is transverse to the Reeb vector field. Applying Lemma 3.12 and Theorem 2.12 to and , we conclude that and do not intersect and that does not intersect the orbit . Similarly, we conclude that and do not intersect, where the -holomorphic cylinder is obtained in Proposition 3.5. Consequently, . ∎
Applying Theorem 3.3 to the maps in (61), we obtain a maximal smooth one-parameter family of finite energy cylinders, containing the family (61). Assuming that strictly increases in the direction of and the normalization , we denote this maximal family by
(62) |
Proposition 5.4.
Consider a sequence in the family (62), where and . Let and be the finite energy cylinders obtained by Proposition 3.5 and Proposition 4.9 respectively. Then after suitable reparametrizations and -translations of , and , we have
-
(i)
up to a subsequence, in as .
-
(ii)
There exist sequences and such that, up to a subsequence, in as .
A similar statement holds for any sequence , with replaced by .
5.3. Bubbling-off analysis for the family of cylinders
Most of the material in Subsection 5.3 is adapted from [21, §6.2]. However, we can not directly apply the results of [21] since our hypotheses are slightly different.
Consider a sequence in the family (62), where and . Note that since all cylinders are asymptotic to at , we have . We reparametrize the sequence so that
(63) |
Define By the same arguments used in the proof of Proposition 3.7, we can assume that is finite and . Moreover, there exists a -holomorphic map
(64) |
such that, up to a subsequence, still denoted by ,
and . The punctures in are non-removable and negative, and the puncture is positive. Indeed, for any sufficiently large or small, we have
where is oriented counterclockwise. It follows that is a positive puncture and is a negative puncture. If , then for any sufficiently small , we have
where is a period. This follows from the same arguments used in §3.2.1. We conclude that is a negative puncture. By the same arguments used in the proof of Proposition 3.5 we conclude that the asymptotic limit of at is .
Lemma 5.5.
Proof.
If , then it follows from (63) that . Now assume and suppose, contrary to our claim, that . By Theorem 2.6, there exists a polynomial and a periodic orbit such that and , where is the cylinder over the orbit . But this implies , contradicting the fact that the asymptotic limit of at is , that is a simple orbit. ∎
5.3.1. Soft rescaling near
Assume and take a puncture . Now we proceed as in the soft rescaling done in §3.2.2. Define the mass of as in (40). Fix satisfying (41) and choose sequences and satisfying (42)-(43). It follows that and, passing to a subsequence, we have . Take such that and define
(65) | ||||
The sequence is a germinating sequence according to Definition 3.6. Let
(66) |
Passing to a subsequence, we can assume that is finite. Let
be a limit of as defined in 3.8. Let be the asymptotic limit of at . Then is asymptotic to at its unique positive puncture . Using Lemma 3.13, we conclude the following.
Lemma 5.6.
If and is asymptotic to at , then .
Since is simple, it follows that is somewhere injective. By Theorem 2.11, we have
and consequently
(67) |
This proves the following lemma.
Lemma 5.7.
Assume . Then . If , then and . Here is the asymptotic limit of at the puncture .
5.3.2. Soft-rescaling near
For any , define
(68) |
and define the mass of the puncture by
(69) |
Note that, for large and small , we have
(70) |
It follows that
(71) |
where is the period of the asymptotic limit of at the puncture . We have two cases:
-
•
either or
-
•
.
I
First assume that . We claim that there is a sequence satisfying
(72) |
Indeed, there exists a sequence satisfying the equation above, since using (63) and , we conclude
Now we show that , so that, passing to a subsequence, still denoted by , the claim is true. Suppose that there exists . Then we have the contradiction
This proves our claim.
Let be small enough so that the disks are disjoint. Define
(73) |
for . It follows from (68), (69), (72) and (73) that, for large and small , we have the estimate
(74) | ||||
Define
(75) |
Using the proof of Proposition 3.7 we conclude, passing to a subsequence, that is finite and . Moreover, there exists a -holomorphic map such that, passing to a subsequence
The map is nonconstant, the punctures in are non-removable and negative, and the puncture is positive.
Lemma 5.8.
The asymptotic limit of at its unique positive puncture is equal to , the asymptotic limit of at .
Proof.
Let be as in the statement of Lemma 3.10. Let be the asymptotic limit of at and let and be connected components of containing and respectively. Since in , we can choose small enough so that, if is fixed, then the loop belongs to for large . Since in , we can choose large enough so that, if is fixed, then the loop belongs to for large . By (72), we can show that
(76) |
Consider, for each , the -holomorphic cylinder , defined by . It follows from (74) that
(77) |
for large . Using (76) and (77) and applying Lemma 3.10 as in the proof of Proposition 3.9, we conclude that and consequently that . ∎
Lemma 5.9.
Either
-
•
or
-
•
.
Proof.
On the contrary, suppose that . Then
a contradiction. ∎
II
Now assume that . Let be small enough so that . Define, for any sequence ,
Using (68), we conclude that
(78) |
It follows that , where is defined as in (75). Indeed, if , arguing as in the proof of Proposition 3.7, we find sequences , and such that for some period , contradicting (78). Thus, passing to a subsequence, still denoted , there exists a -holomorphic map such that
Lemma 5.10.
If , the curve is a trivial cylinder over the orbit . Moreover, is asymptotic to at its negative puncture .
Proof.
It follows from (78) that Indeed, if , then , which contradicts (78). Note that
(79) | ||||
Using (70) we conclude that
(80) |
It follows from (79), (80) and (71) that
(81) |
We conclude that is a cylinder over a -periodic orbit .
Now we prove that , where is the asymptotic limit of at the puncture . Let be as in the statement of Lemma 3.10. For fixed , we know that Thus, we can choose a sequence such that
where is the connected component of containing . We conclude that . Since we conclude, from estimate (78) and Lemma 3.10, arguing as in the proof of Proposition 3.9, that . ∎
This completes the analysis of the case .
Going back to the case , if the mass of the puncture of is positive or , we repeat the process. It necessarily stops after finitely many iterations, when we reach punctures with zero mass or run out of bubbling-off points. This follows from Lemmas 3.11 and 5.9. We obtain a bubbling-off tree of finite energy spheres, defined as in Theorem 4.6. The leaves of the tree correspond to finite energy planes originating from the bubbling-off points and a cylinder over the orbit , originating from the puncture .
5.4. Proof of Proposition 5.4
The finite energy curve defined by (64) is asymptotic to at its positive puncture and to an orbit at the negative puncture . If , the punctures in are negative.
One of the leaves of the bubbling-off tree obtained from the sequence is a cylinder over the orbit originated from the puncture . Using the fact that and an argument similar to Claim I in the proof of Theorem 4.9, we conclude that
(82) |
We first show that . On the contrary, suppose that . Then by Lemma 5.7, we have , and , where is the asymptotic limit at the unique puncture . The orbit is not linked to . Indeed, for each , is an embedding whose image does not intersect , which implies that any contractible loop in is not linked to as well. It follows that any loop in the image of near the end is not linked to and consequently that is not linked to . Since , by hypothesis we have . This implies that is contractible in . Since the region contains an embedded disk with boundary and consequently is contractible in , it is easy to prove, using Mayer-Vietoris sequence, that the holomology class of generates , a contradiction.
Now we show that the mass of the puncture , defined by (69), is positive. On the contrary, suppose that . By Lemma 5.10, we know that is asymptotic to at its negative puncture . This contradicts the fact that the family of cylinders (62) is maximal.
So far, we know that is a -holomorphic cylinder asymptotic to at and to the orbit at . By (67) and (82), we have
The second level of the bubbling-off tree obtained from the sequence consists of a unique vertex associated to a finite energy sphere , asymptotic to at its positive puncture . Observe that since the orbit is not linked to and, for each , does not intersect the image of , we know that any loop in the image of is also not linked to . It follows that any loop in the image of is not linked to and consequently, that is also not linked to .
Claim I: .
To prove the claim, suppose by contradiction that . If is somewhere injective, using Theorem 2.11, we have
where is the asymptotic limit of at the puncture . Using the fact that and an argument similar to Claim I in the proof of Theorem 4.9, we conclude that . If , by Lemma 3.13, we have . We conclude that and , which contradicts Lemma 5.9. If is not somewhere injective, there exists a somewhere injective -holomorphic curve and a polynomial such that , and . This implies that , where is the asymptotic limit of at . Using Lemma 2.1 and the assumption , we conclude that . For every , we have , where , and is the asymptotic limit of at the puncture . Since , using Lemma 2.1 we conclude that . Applying Theorem 2.11 to the curve we have
where is the asymptotic limit of at the puncture . It follows that . By Theorem 2.11, we have , which implies . It follows from Theorem 2.6 that , where is a cylinder over an orbit , is a polynomial and . Since , we conclude, using Lemma 2.1, that . By Lemma 5.9, we have . Let . Then and , where . This implies that , a contradiction. This proves our claim.
Since and is not linked to , by hypothesis we have
Now we prove that the mass of the puncture of is zero, according to definition (69). This implies that is asymptotic to at its unique negative puncture . Suppose, by contradiction, that . Since is simple, we know that is somewhere injective. Applying Theorem 2.11 to , we have
If , then . By Theorem 2.6, we conclude that is a cylinder over , contradicting Lemma 5.9. Therefore, and we have
Thus and . Following the same arguments used to prove Claim I, we get a contradiction with . We have proved that is asymptotic to at .
We claim that . On the contrary, suppose that . Since is somewhere injective and is not identically zero, by Theorem 2.11 we have
This implies that , contradicting 3.13.
So far, we have proved the following statement.
Proposition 5.11.
Consider a sequence in the family (62), where and . There exists a cylinder asymptotic to at its positive puncture and to at its negative puncture and a cylinder asymptotic to at its positive puncture and to at its negative puncture , such that, after suitable reparametrizations and -translations of , we have
-
(i)
up to a subsequence, in as .
-
(ii)
There exist sequences and such that, up to a subsequence, in as .
A similar statement holds for any sequence with replaced with a cylinder with the same asymptotics as and replaced with a cylinder with the same asymptotics as .
Proposition 5.12.
Let , and be as in Proposition 5.11. Then
-
(i)
Given an -invariant neighborhood of the loop , there exists such that, for and large , the loop belongs to .
-
(ii)
Given an -invariant neighborhood of the loop , there exist and such that the loop belongs to , for and large .
-
(iii)
Given an -invariant neighborhood of the loop , there exist such that the loop belongs to , for and large .
-
(iv)
Given any neighborhood of , we have , for large .
A similar statement holds for any sequence such that , with replaced by and replaced by .
Proof.
We can assume that , contains only the periodic orbit modulo -reparametrizations. Let be as in the statement of Lemma 3.10 and such that .
Using the normalization condition (63) we can apply Lemma 3.10 and find such that for , the loops belong to . By the asymptotic behavior of the cylinders , we conclude that for any and large , the loop belongs to . Assertion (i) is proved.
Now we prove (ii). Recall that the asymptotic limit of at is the orbit . We can apply 3.10 as in the proof of Proposition 5.8 to find small and such that for every satisfying and large , the loop belongs to .
To prove (iii), first recall that the mass of the puncture of the sequence is zero. Applying 3.10 as in the proof of Lemma 5.10, we find small and a sequence such that for . The sequence converges in to the cylinder over , so that applying Lemma 3.10 again, we conclude that for large , the loop belongs to for every . We conclude that for any and large enough, the loop belongs to .
The proof of (iv) follows from (i), (ii), (iii) and Proposition 5.11.∎
Proposition 5.13.
Proof.
By positivity and stability of intersection of pseudo-holomorphic curves and the fact that the cylinders in the family (62) do not intersect , we conclude that the cylinders , , and do not intersect . It follows from Proposition 4.12 that , up to reparametrization and -translation. It follows from Proposition 3.16 that is either equal to or , up to reparametrization and -translation. Similarly, is either equal to or . Since the family of cylinders (61) was obtained by Theorem 5.2 applied to and , we conclude, using Proposition 5.12, that .
It remains to prove that . Fix . By Jordan-Brouwer separation theorem, the surface
divides into two disjoint regions with boundary . One of these regions, that we call , is contained in , since does not intersect any of the curves foliating the interior of the region .
We will show that is foliated by cylinders in the family . Let and let be a small neighborhood of in separating and . By Proposition 5.12, for sufficiently close to , we have . Moreover, , since there are points in the image of the family converging to points in a compact subset of , as . Let be the region limited by . Thus, the image of the family , is open, closed and nonempty in . It follows that is in the image of for some . We conclude that is foliated by the images of the cylinders , .
Since every neighborhood of a compact set of in is contained in , the family , , foliates and the cylinders in the family do not intersect each other, we conclude that . ∎
The proof of Proposition 5.4 is complete.
5.5. Conclusion of the proof of Theorem 1.3
Proposition 5.14.
Proof.
Proposition 5.15.
, for .
Proof.
First we prove that . Our proof is adapted from [25, §6]. Let be one of the disks with boundary obtained in Proposition 3.1. Consider the orientation of induced by and let be the orientation on induced by . The characteristic distribution of , defined by where means the -symplectic orthogonal, can be parametrized by a smooth vector field on , which is transverse to pointing outwards. After a perturbation of away from a neighborhood of and keeping the transversality of and , we can assume that all singular points of are nondegenerate. More details about these facts can be found in [17] and [16]. Let be the orientation on given by . A zero of is called positive if coincides with at , and negative otherwise. Since is transverse to , all singularities of have the same sign and they must be positive since . Considering as a section of the bundle , we have
(83) |
where is a trivialization of . Let be a trivialization of . Considering as a section of the bundle and using the fact that points outwards at , we have
(84) |
Since all singularities of are positive and at a singularity of does not depend on whether we consider as a section of or as a section of , the right side of equations (83) and (84) coincide.
Consider the nonvanishing section of defined by . After a small perturbation of in the direction of , which we can assume to be transverse to and , we find a new closed curve . Consider a trivialization of induced by the nonvanishing section . We have
The proof of is completely analogous, replacing with the disk given by Proposition 3.1.
Now we prove that . Let be the cylinder with boundary given by Proposition 4.1. Let be the orientation on induced by . The characteristic distribution of can be parametrized by a smooth vector field on , which is transverse to pointing outwards. Note that, along , we have . After a perturbation of away from a neighborhood of and keeping the transversality of and , we can assume that all singular points of are nondegenerate. Since is transverse to , all singularities of have the same sign and they must be positive since . Let be a trivialization of . Then we have
Here we have used the fact that points outwards. Consequently,
We know that is a -embedded disk with boundary . Note that with the orientation induced by is an Seifert surface for . Consider the nonvanishing section of defined by . After a small perturbation of in the direction of , which we can assume to be transverse to the interior of and , we find a new closed curve . Consider a trivialization of induced by the nonvanishing section . Thus,
∎
We have proved the existence of a foliation adapted to . The proof of Theorem 1.3 is complete.
6. Proof of Proposition 1.5
In this section we prove Proposition 1.5, which is restated below.
Proposition 6.1.
Assume there exists a foliation adapted to the contact form and let be the binding orbit with Conley-Zehnder index , as in Definition 1.2. Then there is no -embedding such that and each hemisphere is a strong transverse section.
Throughout the section, we follow the notation of Definition 1.2. Before proving Proposition 1.5 we need two lemmas, which are stated below.
Lemma 6.2.
If is a -embedding such that and is transverse to the Reeb vector field, then we can assume that is also transverse to .
Proof.
Define by , where is the Reeb flow. Then satisfies , , and
for every and . Since is transverse to the Reeb flow away from , we know that is surjective. Since preserves the splitting , it follows that is surjective. We conclude that is surjective for all and . In particular, is transverse to on . This implies 111If , and are smooth manifolds, is a submanifold of and is a smooth map transverse to , then for almost every , is transverse to . A proof of this fact can be found in [13, §3]. that for almost all , is transverse to .
Moreover, , and it follows that is still transverse to the flow for all . Thus, we can replace with for some satisfying to get the desired embedding. ∎
Lemma 6.3.
Assume that is a embedding such that and such that the image of each closed hemisphere is a strong transverse section. Then is transverse to the torus along .
The proof of Lemma 6.3 is postponed to Subsection 6.2. Subsection 6.1 below consists of preliminary results. Now we use Lemmas 6.2 and 6.3 to prove Proposition 1.5.
Proof of Proposition 1.5.
Suppose, by contradiction, that is a embedding such that and such that each hemisphere is a strong transverse section.
Let be the torus given by Definition 1.2. divides into two closed regions e . The region contains an embedded disk with boundary , so that is contractible in . One can also show, using Mayer-Vietoris sequence, that the holomology class of generates . Lemma 6.3 shows that the image of any neighborhood of by intersects both and away from . This implies that the sphere intersects the torus away from . Indeed, one of the hemispheres of intersects and can not be contained in , since this would imply contractible in , a contradiction.
Since is transverse to and , we conclude that . Using Lemmas 6.2 and 6.3, we conclude that intersects transversely and the intersection is contained in a closed subset of . We conclude that the preimage of the intersection by is a -dimensional submanifold of which is a closed subset of . It follows that each connected component of is diffeomorphic to .
The equator is one of the connected components of the boundary of a region such that . Thus, one of the other connected components of the boundary of , that we denote by , is such that is homologous to in . Denoting , the homology class of in is for some . Since does not intersect , must be zero. This implies that and divide into two connected regions.
Now fix an orientation on in such a way that preserves orientation. Consider the closed hemispheres of , that we call and , with the orientation induced by the orientation of . It follows that
Since is transverse to the Reeb vector field in , we have in . Let be the connected region of bounded by and . Since is contained in one of the hemispheres of , we have
(85) |
Since does not intersect , we know that one of the regions of bounded by and , that we denote by , satisfies either or . Recall that and are oriented in such a way that is an area form, is a positive asymptotic limit and is a negative asymptotic limit. Then we have
(86) |
contrary to (85). This proves the proposition. ∎
6.1. Relative position of sections along
The orbit is hyperbolic and lies in the intersection of its stable manifold and its unstable manifold . The tangent spaces of along the periodic solution are spanned by the Reeb vector field and vector fields defined below.
Consider the path of symplectic matrices where is any global symplectic trivialization of . Since , we know that has two eigenvalues , where . The vector is an eigenvector of associated to the eigenvalue and the vector is an eigenvector of associated to the eigenvalue . For each , we can define
so that is an eigenvector of associated to the eigenvalue and is an eigenvector of associated to the eigenvalue .
Replacing with if necessary, we can assume that is a positive basis for , for each . The basis determines four open quadrants in . Let and be the open quadrants between and following the positive orientation and and the open quadrants between and .
Proposition 6.4.
Let be a -periodic nonvanishing section such that If then belongs to regions or for every . If then belongs to regions or for every .
Proof.
Let . Recall that is -periodic, and for each , is real and symmetric matrix. Define . In what follows we will also write for the representations of the sections in the trivialization . The sections are solutions of the equation
(87) |
satisfying In the basis , has the form
We want to find a matrix satisfying and a -periodic map such that , so that with the change of variables , the equation (87) becomes
(88) |
Define
and by The map is -periodic. In fact,
If is a solution of (87), then satisfies
Using the identities and , we get
Thus, is a solution of (88). In the same way, if solves (88), then solves (87).
Let be a -periodic nonvanishing section on . We want to compute In the symplectic trivialization , takes the form
where we also denote by its representation in the trivialization . Thus,
Writing in the basis as and in the basis as , we have
Thus, we have
where is a positive constant. For fixed , let be the solution of (87) such that . Then
Writing is the basis as and in the basis as , we have
Thus, we have
We conclude that
(90) |
Assume that satisfies and that By (90), we have Note that . This follows from and the geometric description of the Conley-Zehnder index given in §2.1. This implies that
(91) |
Now we show that for all , lies in regions or . On the contrary, suppose that there exists such that belongs to regions , , or is the direction of . By (89), we have . This implies that is increasing near . Since , this would force the existence of such that and , a contradiction with . Now, suppose that, for some , is in the direction of , then, using (89), we have . But this would force to go to or , which is impossible. We conclude that lies in regions or , for every .
By a similar argument, we conclude that if satisfies and then we have or , . ∎
6.2. Proof of Lemma 6.3
Lemma 6.5.
Let be a section on such that and generate along . Let be a section on such that and generate the tangent space of along . Then .
Proof.
Let be one of the closed hemispheres of and let . Then is an embedded disk satisfying and is transverse to the Reeb vector field . The characteristic distribution can be parametrized by a smooth vector field on , which is transverse to pointing outwards. After a perturbation of keeping the transversality of and , we can assume that all singular points of are nondegenerate. Let be the orientation on induced by the orientation of in the direction of . Let be the orientation on given by . A zero of is called positive if coincides with at , and negative otherwise. Since is transverse to , all singularities of have the same sign and they must be positive since . Considering as a section of the bundle , we have
(92) |
Let be a trivialization of . Considering as a section of the bundle and using the fact that points outwards at , we have
(93) |
Since all the singularities of are positive and at a singularity of does not depend on whether we consider as a section of or as a section of , equations (92) and (93) coincide. We conclude that
By the same arguments above, if is a section such that generates the tangent space of the disk along , where is the disk given by the definition of foliation 1.2, we have Since is transverse to along , we obtain ∎
Now we are ready to prove Lemma 6.3.
Proof of Lemma 6.3.
Let be a section along such that generates along and
(94) |
Let ; be a smooth function such that
Fix an orientation on in such a way that preserves orientation. Consider the closed hemispheres of , that we call e , with the orientation induced by the orientation of . It follows that
Since is transverse to the Reeb flow along , we have on . This implies that is a local maximum of the function
where is the action functional defined by . For any , we have
(95) |
where is the asymptotic operator defined by (18). For a proof, see [5, §1.4]. It follows from (17) and (95) that
(96) |
By (94), we have
(97) |
Let be a section along such that generates the tangent space of along and
(98) |
Let , be a smooth function such that
Recall that the cylinders and given by Definition 1.2 are oriented in such a way that is an area form, is a positive asymptotic limit and is a negative asymptotic limit. This implies that is a local minimum of the function . It follows from (17) and (95) that
(99) |
By (98), we have
(100) |
Now we can apply Lemma 6.5 and Proposition 6.4 to the sections and and conclude that is transverse to the torus along . ∎
7. Proof of Theorem 1.6
In this section, we prove Theorem 1.6, which is restated below.
Theorem 7.1.
Let be a tight contact form on . Assume that there exist Reeb orbits that are nondegenerate, simple, and have Conley-Zehnder indices respectively , and . Assume further that the orbits , , and are unknotted, and are not linked for , , and the following conditions hold:
-
(i)
;
-
(ii)
If satisfies and , then .
-
(iii)
There exists such that the almost complex structure admits a finite energy plane asymptotic to at it positive puncture and a finite energy clylinder asymptotic to at its positive puncture and at its negative puncture ;
-
(iv)
There exists no -embedding such that and each hemisphere is a strong transverse section.
Then there exists a foliation adapted to with binding orbits , , and . Consequently, there exists at least one homoclinic orbit to .
Proof of Theorem 1.6.
By the same arguments used in §3.1, we find a maximal one-parameter family of finite energy -holomorphic planes
(101) |
asymptotic to the orbit . For each , is an embedding transverse to the Reeb flow and We assume that strictly increases in the direction of .
Now we describe how the family breaks as and . Consider a sequence satisfying , and define .
Claim I
There exists a -holomorphic finite energy cylinder
(102) |
asymptotic to at its positive puncture and to at its negative puncture , and a finite energy -holomorphic plane
(103) |
asymptotic to at its positive puncture , such that, after a suitable reparametrization and an -translations of , the following hold
-
(i)
up to a subsequence, in as .
-
(ii)
There exist sequences , and such that, up to a subsequence, in as .
-
(iii)
Given an -invariant neighborhood of the loop , there exists such that the loop belongs to , for and large .
-
(iv)
Given an -invariant neighborhood of the loop , there exists and such that the loop belongs to , for and large .
-
(v)
Given any neighborhood of , we have , for large .
Here . A similar claim holds for any sequence satisfying . In this case we change the notation from and to and respectively.
To prove Claim I, define as any real number satisfying
(104) |
After reparametrizing and translating in the direction, we can assume that
(105) | |||
(106) | |||
(107) |
Let be the set of points such that there exist subsequence and sequence satisfying .
If , by elliptic estimates we find a -holomorphic map such that, up to a subsequence, in . Using Fatou’s Lemma we conclude that . By (105) we have
(108) |
Therefore, is nonconstant.
If , let and let be such that, passing to a subsequence still denoted by , we have . Consider a sequence such that . We need the following lemma.
Lemma 7.2 ([15, Lemma 3.3]).
Let be a complete metric space and a continuous function. For any and , there exist and such that
By Lemma 7.2, we have, after perhaps modifying and ,
(109) |
Denoting , define the sequence of -holomorphic maps by
(110) |
By elliptic estimates, there exists a subsequence, still denoted by and a -holomorphic map such that in . Since , we know that is nonconstant. Since , we have . Any asymptotic limit of the plane is not linked to and has action . We conclude that is asymptotic to either , or . Now we show that is asymptotic either to or to . Suppose, by contradiction, that is asymptotic to . From equation (22), Lemma 2.10 and , we have . Using equation (29), we get the contradiction
Using Fatou’s Lemma and passing to a subsequence, still denoted by , we get
(111) |
Since , we conclude from (111) that . From (106) and elliptic estimates, we find a -holomorphic map
such that, up to a subsequence, and . The puncture is negative. Indeed, for any , we have
(112) |
where is oriented counterclockwise. Here we have used (111). The puncture is necessarily positive since . Using (107) we conclude that .
Now we show that is asymptotic to at its positive puncture and to at its negative puncture . We need the following lemma, which is an adaptation of Lemma 4.9 from [21] to our set-up.
Lemma 7.3.
Consider a constant and let be defined by (104). Identifying , let be an open neighborhood of the periodic orbits , and , viewed as maps , . We assume that is -invariant, meaning that , and that each of the connected components of contains at most one periodic orbit modulo -reparametrizations. Then there exists a constant such that the following holds. If is a -holomorphic cylinder such that the image of does not intersect , is not linked to , and
then each loop is contained in for all .
Proof.
Arguing by contradiction, we find a sequence of -holomorphic maps
such that the image of does not intersect , is not linked to ,
(113) |
and
(114) |
Define a sequence of -holomorphic maps by
Note that and by (114) we have
(115) |
We claim that is uniformly bounded in and . Conversely, suppose that there exist a subsequence and a sequence satisfying . Arguing as in the proof of (111), we get , for some sequence . This contradicts
Thus, passing to a subsequence, still denoted by , we have in , where is a finite energy cylinder. Using (113), we conclude that
Consequently, is nonconstant, has a positive puncture at , a negative puncture at and any asymptotic limit of at has period . Since the loops are not linked to , the asymptotic limits of are not linked to as well. Therefore, is asymptotic to either , or at . Since , we conclude that is a cylinder over either , or . However, by (115), we know that , a contradiction. ∎
Any asymptotic limit of at is not linked to and satisfies . Thus, is asymptotic to either , or at . Let be as in the statement of Lemma 7.3. Using (105) and Lemma 7.3, we conclude that for each and large , and . Since the planes are asymptotic to and for each fixed , converges to , as , we conclude that is asymptotic to at . Consequently we have , since would contradict the fact that the family (101) is maximal.
Using (105), we conclude that
(116) |
Therefore, any asymptotic limit of at is not linked to and satisfies . Here we have used (112) and (116). We conclude that .
Now we proceed as in the soft rescaling done in §3.2.2. Let be defined as in (39). Since is a nondecreasing function of , we can fix so that
(117) |
Arguing as in §3.2.2, we choose sequences and satisfying
(118) | |||
(119) |
such that and . Thus, passing to a subsequence, we can assume .
Take any sequence satisfying and define the sequence of -holomorphic maps by
(120) |
Note that and . Let be the set of points such that there exist subsequence and sequence satisfying . Then there exists a -holomorphic map such that, up to a subsequence, in .
We claim that is nonconstant and asymptotic to at its positive puncture . Indeed, using (119) and (120) we conclude that, for any ,
(121) |
and by (105), we have
(122) |
Here we have used and . Therefore, any asymptotic limit of at is not linked to and satisfies . We conclude that .
Now we prove that . Suppose, by contradiction, that and let be such that, passing to a subsequence still denoted by , . Arguing as in the proof of (111), we find a sequence such that, for each sufficiently small , we have
We conclude that and . This implies that is a cylinder over the orbit and we get the contradiction
Here we have used (117) and (119). We have proved (i) and (ii).
An application of Lemma 7.3, similar to the proof of Proposition 3.14, proves (iii) and (iv). Assertion (v) is a consequence of (i)-(iv). Claim I is proved.
Following the same arguments as in the proof of Propositions 3.15-3.17, we conclude that
-
•
Up to reparametrization, ;
-
•
;
-
•
The union of the image of the family (101) with the images of , , , and determine a singular foliation of a closed region , homeomorphic to a solid torus, such that , where .
Now we find the foliation of . Following the same arguments as in Proposition 5.3 we prove that the finite energy cylinder is an embedding. Applying Theorem 3.3 to the finite energy cylinder , we obtain a maximal one-parameter family of finite energy cylinders
(123) |
asymptotic to at its positive puncture and to at its negative puncture . For each , is an embedding, the projection is an embedding which does not intersect its asymptotic limits, and . We assume that strictly increases in the direction of .
Consider a sequence satisfying and define .
Claim II
There exists a finite energy cylinder
(124) |
asymptotic to at its positive puncture and to at its negative puncture , and a finite energy cylinder
(125) |
asymptotic to at its positive puncture and to at its negative puncture , such that, after suitable reparametrizations and -translations of , we have
-
(i)
up to a subsequence, in as .
-
(ii)
There exist sequences and such that, up to a subsequence, in as .
-
(iii)
Given an -invariant neighborhood of the loop , there exists such that the loop belongs to , for and large .
-
(iv)
Given an -invariant neighborhood of the loop , there exists and such that the loop belongs to , for and large .
-
(v)
Given an -invariant neighborhood of the loop , there exists such that the loop belongs to , for and large .
-
(vi)
Given any neighborhood of , we have , for large .
A similar claim holds for any sequence with replaced with a cylinder with the same asymptotics as and replaced with a cylinder with the same asymptotics as .
After a reparametrization and -translation of , we can assume that
(126) | |||
(127) |
where is defined by (104). Observe that is uniformlly bounded in and . Otherwise, arguing as in the proof of (111), we would find sequences , and a subsequence of , still denoted , such that
contradicting the hypothesis . We conclude that there exists a finite energy -holomorphic cylinder such that, up to subsequence, still denoted by ,
The finite energy cylinder is nonconstant, is a negative puncture and is a positive puncture. Indeed, for any , we have where is oriented counterclockwise.
It follows from (126) that . Using Lemma 7.3 as in the proof of Claim I, we conclude that is asymptotic to at the puncture . Any asymptotic limit of at is not linked to and has period . Thus, is asymptotic to either or at . Since the family (123) is maximal, we conclude that is asymptotic to at the negative puncture .
Now we proceed as in the soft rescaling done in §5.3.2. Arguing as in §5.3.2, we find a sequence satisfying
(128) |
and such that . Thus, passing to a subsequence still denoted by , we can assume that .
Fix such that
(129) |
and define the sequence of -holomorphic maps by
(130) |
Using (128) and (129), we conclude that, for large ,
(131) | ||||
Note that and . Moreover, is uniformly bounded on and . Otherwise, arguing as in the proof of (111), we would find sequences , and a subsequence of , still denoted by , such that
a contradiction. Therefore, there exists a finite energy -holomorphic cylinder
such that, up to a subsequence,
For and large , we have
(132) |
Here we have used (126) and (128). Therefore, we have
(133) |
We conclude that is asymptotic to at the positive puncture .
Using (128), we conclude that
Thus, any asymptotic limit of at its negative puncture is an orbit that is not linked to and has period . We conclude that is asymptotic to at . This completes the proof of (i) and (ii).
An application of Lemma 7.3, similar to the proof of Proposition 5.12, proves (iii), (iv) and (v). Assertion (vi) is a consequence of (i)-(v).
Following the same arguments as in the proof of Propositions 5.13-5.15, we conclude that
-
•
Up to reparametrization and -translation, we have and , where and are given by Claim I;
-
•
Up to reparametrization and -translation we have ;
-
•
The images of the family , , , , , and determine a singular foliation of the region .
-
•
, .
It follows that there exists a foliation with binding orbits , and . The existence of a homoclinic to is proved as in §5.5. ∎
8. Proof of Proposition 1.7
In this section we prove Proposition 1.7, which is restated below.
Proposition 8.1.
Lemma 8.2.
Assuming that is sufficiently small, the orbits , and are nondegenerate and their Conley-Zehnder indices are respectively and .
Proof.
First we define a trivialization of . Consider the matrices ,
For each we have an ortonormal frame of given by , where
(134) |
Note that and .
The contact structure is isomorphic to , as a tangent hyperplane distribution, via the projection along . Let be the vector field determined by
(135) |
We can define a symplectic trivialization of by
(136) |
Note that along the orbits , , coincides with the plane . Indeed, along the orbits, we have and
(137) | ||||
We can define another trivialization along the orbits , , by
(138) |
It is easy to check that , , is a symplectic trivialization. A simple calculation shows that, in the trivialization , , is represented by
(139) |
which is a constant linear map. Note that, for sufficiently small , we have .
Let be the flow of . The linearized flow along a trajectory is the solution of the equation
(140) |
For , we have and . Solving equation (140) we find
(141) |
Thus, if is sufficiently small, is nondegenerate and for any , has winding number , where is defined by (13). We conclude that .
For , we have and . Solving equation (140) we find
(142) |
Note that
(143) | ||||
We conclude that is nondegenerate and .
For , we have and . Solving equation (140) we find
(144) |
Thus, if is sufficiently small, is nondegenerate and for any , has winding number . We conclude that .
Now, the Conley-Zehnder index of the orbit , , satisfies the formula
and
(145) | ||||
We conclude that , and . ∎
Lemma 8.3.
There exists such that the almost complex structure admits a finite energy plane asymptotic to at its positive puncture and a finite energy cylinder asymptotic to at its positive puncture and at its negative puncture .
Proof.
Our proof follows [7, §5]. Define a -compatible complex structure by
(146) |
where , is defined by (135). Let be the almost complex structure on defined by (23).
First we look for the -holomorphic plane. Our candidate for the plane will project into the surface
(147) |
We write for . First we assume that satisfies
(148) |
where and are smooth functions to be determined.
Recall that is -holomorphic if and only if it satisfies
(149) | |||
(150) |
where is the projection along the Reeb vector field .
Now we find conditions on and so that defined by (148) satisfies (149). First note that
(151) | ||||
Restricting the frame to the surface (147) we find
(152) | ||||
(153) | ||||
Now using the definition of (146) and equation (149), we find the differential equation
(154) |
Since the image of is the surface (147), we have
(155) |
Moreover, . Thus, we may view (154) as an ODE of the type
(156) |
where is a smooth function. Here is the unique positive solution of .
Note that vanishes in and in , and G is negative on . Thus, for . Consequently is strictly decreasing and satisfies
(157) |
It follows from (155) and (157) that
(158) |
where is defined by (9). Using (157) and (158) we conclude that the loops converge to in as . Moreover, the loops converge to in as .
Now we define the function by
(159) |
Then
(160) | ||||
Consider the map with defined by (148), where is a solution of (154) satisfying , is defined by (155), and defined by (159). Since satisfies (149) and (150), is -holomorphic. Moreover, has finite energy. Indeed, using Stokes theorem and equations (158) and (159) we find
(161) | ||||
where , is smooth and . We conclude that .
The mass of the singularity satisfies
This implies that has a removable singularity at . Removing it, we obtain a finite energy -holomorphic plane asymptotic to at its positive puncture .
The construction of the cylinder is completely analogous, but we start with a solution of (154) satisfying . In this case, we obtain
and the singularity is non-removable, since
∎
Lemma 8.4.
Assuming that is sufficiently small, if satisfies and , then .
Proof.
First we prove the following claim.
Claim I
Let be a nonconstant -periodic solution of , where . Then , where .
To prove the claim, first we note that satisfies
(162) |
This follows from the fact that and from the representations of and as Fourier series. Now we have
(163) |
Consider the function , where
Then is continuous and . Hence there exists such that, for and ,
(164) |
For sufficiently small, we also have
(165) |
Indeed, using the fact that is a proper function, we know that the set is bounded by a uniform constant for every . Thus, for sufficiently small and , we have
(166) | ||||
Claim II
If satisfies
(167) |
then the period of satisfies .
Using Claim I and equation (164) we conclude that the period of satisfies
Note that the orbit is nonconstant, since , and -periodic. This implies that the period of satisfies , for some . The period of satisfies
(168) | ||||
where is the region in the plane limited by the image of . Here we have used (10) and (167). Taking a smaller if necessary and using (12), we have
proving the claim.
Let be such that . We claim that . By Claim II, we know that the solution does not satisfy (167), that is, . The energy levels , for are diffeomorphic to and bound a region containing in its interior. Then intersects the disk
Note that . Moreover is transverse to the Reeb vector field , since . We conclude that , which proves the lemma.
∎
Lemma 8.5.
There is no -embedding such that and each hemisphere of is a strong transverse section.
Proof.
Let and be the stable and unstable manifolds of respectively. First note that
Moreover, consists of two connected components diffeomorphic to cylinders
(169) |
where and are solutions of such that
and is contained in the interior of the bounded region with boundary .
Define Then is a -torus topologically embedded in and divides into two closed regions with boundary . One of these regions, that we call , has holomogy generated by , and is contractible in . Since the disk
has boundary and projects into the complement of in the plane , we know that , , also projects into the complement of in the plane . This implies that projects into , for . Moreover, we have .
Following §6.1, let be the positive basis of where is an eigenvector of associated to the eigenvalue and is an eigenvector of associated to the eigenvalue .Let and be the open quadrants between and following the positive orientation and and the open quadrants between and .
Let be a section of along such that generates along . We claim that
(170) |
By Proposition 6.4 it is enough to show that
(171) | |||
(172) |
where is any global symplectic trivialization of . The proof of (171) is completely analogous to the proof of Lemma 6.5 and the proof of (172) is similar to the proof of (97).
Suppose, by contradiction, that is a embedding such that and each hemisphere is a strong transverse section. Using (170), we conclude that one of the closed hemispheres of , that we denote by , intersects . Since is not contractible in , the image of can not be contained in and we conclude that .
We claim that consists of a disjoint union of embedded circles. To prove this claim, note that by (170), there exists an open neighborhood of in such that . Thus, is a closed subset of such that . Moreover, is transverse to . This implies that is a one dimensional submanifold of that is a closed subset of , proving our claim.
At least one of the connected components of is homologous to both in and in . Let be such that is one of these components. Then and divide into two closed regions. Let be one of these regions, with orientation induced by . Then
(173) |
Consider with the orientation induced by the orientation of . It follows that
(174) |
Since is transverse to the Reeb vector field, we have in . Let be the region bounded by and . Then
(175) |
contradicting (173). This proves the lemma. ∎
Acknowledgments
This work originated from the author’s Ph.D. thesis, written under the supervision of Prof. Pedro Salomão at the University of São Paulo. The author wishes to express her gratitude to Pedro Salomão for suggesting the problem, for many stimulating conversations, and for his support during the preparation of this paper. The author would like to thank Alexsandro Schneider, Ana Kelly de Oliveira, Naiara de Paulo, and Seongchan Kim for pointing out mistakes in previous versions of the paper and the reviewer for the helpful feedback.
References
- [1] D. Bennequin, Entrelacements et equations de pfaff, Astérisque 107-108 (1983), 87–161.
- [2] F. Bourgeois, Y. Eliashberg, H. Hofer, K. Wysocki, and E. Zehnder, Compactness results in symplectic field theory, Geom. Topol. 7 (2003), no. 2, 799–888.
- [3] B. Bramham, Periodic approximations of irrational pseudo-rotations using pseudoholomorphic curves, Ann. of Math. (2) (2015), 1033–1086.
- [4] V. Colin, P. Dehornoy, and A. Rechtman, On the existence of supporting broken book decompositions for contact forms in dimension 3, preprint arXiv:2001.01448 (2020).
- [5] C. L. de Oliveira, 3-2-1 foliations for reeb flows on , Ph.D. thesis, Universidade de São Paulo, 2020.
- [6] N. de Paulo, U. Hryniewicz, S. Kim, and P. Salomão, Genus zero transverse foliations for weakly convex Reeb flows on the tight -sphere, Preprint arxiv 2206.12856v1 (2022).
- [7] N. de Paulo and P. Salomão, Systems of transversal sections near critical energy levels of hamiltonian systems in , Mem. Amer. Math. Soc. 252 (2018), no. 1202.
- [8] by same author, On the multiplicity of periodic orbits and homoclinics near critical energy levels of hamiltonian systems in , Trans. Amer. Math. Soc. 372 (2019), no. 2, 859–887.
- [9] D. Dragnev, Fredholm theory and transversality for noncompact pseudoholomorphic maps in symplectizations, Comm. Pure Appl. Math. 57 (2004), no. 6, 726–763.
- [10] Y. Eliashberg, Contact 3-manifolds twenty years since j. martinet’s work, Ann. Inst. Fourier (Grenoble) 42 (1992), no. 1-2, 165–192.
- [11] J. Fish and R. Siefring, Connected sums and finite energy foliations i: Contact connected sums, J. Symplectic Geom. 16 (2018), no. 6, 1639–1748.
- [12] M. Gromov, Pseudo holomorphic curves in symplectic manifolds, Invent. Math. 82 (1985), 307–347.
- [13] V. Guillemin and A. Pollack, Differential topology, vol. 370, American Mathematical Soc., 2010.
- [14] H. Hofer, Pseudoholomorphic curves in symplectizations with applications to the weinstein conjecture in dimension three, Invent. Math. 114 (1993), no. 1, 515–563.
- [15] H. Hofer and C. Viterbo, The weinstein conjecture in the presence of holomorphic spheres, Comm. Pure Appl. Math. 45 (1992), no. 5, 583–622.
- [16] H. Hofer, K. Wysocki, and E. Zehnder, A characterization of the tight 3-sphere, Duke Math. J. 81 (1995), no. 1, 159–226.
- [17] by same author, Properties of pseudo-holomorphic curves in symplectisations II: Embedding controls and algebriac invariants, Geom. Funct. Anal. 5 (1995), no. 2, 270–328.
- [18] by same author, Properties of pseudoholomorphic curves in symplectisations I: Asymptotics, Ann. Inst. H. Poincaré Anal. Non Linéaire 13 (1996), no. 3, 337–379.
- [19] H. Hofer, K. Wysocki, and E. Zehnder, The dynamics on three-dimensional strictly convex energy surfaces, Ann. of Math. (2) 148 (1998), no. 1, 197–289.
- [20] H. Hofer, K. Wysocki, and E. Zehnder, Properties of pseudoholomorphic curves in symplectizations III: Fredholm theory, Topics in nonlinear analysis, Progr. Nonlinear Differential Equations Appl., vol. 35, Springer, 1999, pp. 381–475.
- [21] H. Hofer, K. Wysocki, and E. Zehnder, Finite energy foliations of tight three-spheres and hamiltonian dynamics, Ann. of Math. (2) 157 (2003), no. 1, 125–255.
- [22] U. Hryniewicz, Fast finite-energy planes in symplectizations and applications, Trans. Amer. Math. Soc. 364 (2012), no. 4, 1859–1931.
- [23] by same author, Systems of global surfaces of section for dynamically convex reeb flows on the 3-sphere, J. Symplectic Geom. 12 (2014), no. 4, 791–862.
- [24] U. Hryniewicz, J. Licata, and P. A. S. Salomão, A dynamical characterization of universally tight lens spaces, Proc. Lond. Math. Soc.(3) 110 (2015), no. 1, 213–269.
- [25] U. Hryniewicz and P. A. S. Salomão, On the existence of disk-like global sections for reeb flows on the tight -sphere, Duke Math. J. 160 (2011), no. 3, 415–465.
- [26] by same author, Elliptic bindings for dynamically convex reeb flows on the real projective three-space, Calc. Var. Partial Differential Equations 55 (2016), no. 2, 43.
- [27] by same author, Global surfaces of section for reeb flows on dimension three and beyond, Proceedings of the International Congress of Mathematicians 2018, vol. 2, 2018, pp. 959–986.
- [28] D. McDuff and D. Salamon, J-holomorphic curves and symplectic topology, Amer. Math. Soc. Colloq. Publ., vol. 52, 2004.
- [29] J. Nelson, Applications of automatic transversality in contact homology, Ph.D. thesis, University of Wisconsin-Madison, 2013.
- [30] D. Rolfsen, Knots and links, American Mathematical Soc., 2003.
- [31] R. Siefring, Intersection theory of punctured pseudoholomorphic curves, Geom. Topol. 15 (2011), no. 4, 2351–2457.
- [32] C. Wendl, Finite energy foliations and surgery on transverse links, Ph.D. thesis, New York University, 2005.
- [33] by same author, Finite energy foliations on overtwisted contact manifolds, Geom. Topol. 12 (2008), no. 1, 531–616.
- [34] by same author, Lectures on symplectic field theory, preprint arXiv:1612.01009 (2016).