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

3213-2-1 foliations for Reeb flows on the tight 3-sphere

Carolina Lemos de Oliveira Instituto de Matemática e Estatística - Universidade do Estado do Rio de Janeiro, Rio de Janeiro - RJ, Brasil carolina.lemos@ime.uerj.br
Abstract.

We study the existence of 3213-2-1 foliations adapted to Reeb flows on the tight 33-sphere. These foliations admit precisely three binding orbits whose Conley-Zehnder indices are 33, 22, and 11, respectively. All regular leaves are disks and annuli asymptotic to the binding orbits. Our main results provide sufficient conditions for the existence of 3213-2-1 foliations with prescribed binding orbits. We also exhibit a concrete Hamiltonian on 4\mathbb{R}^{4} admitting 3213-2-1 foliations when restricted to suitable energy levels.

Key words and phrases:
Hamiltonian dynamics, pseudo-holomorphic curves, Reeb flows
2020 Mathematics Subject Classification:
Primary 53D35; Secondary 37J46, 37J55
This study was financed by grant #2016/10466-5, São Paulo Research Foundation (FAPESP), by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance Code 001 and by Serrapilheira Institute through a grant awarded to Prof. Vinicius Ramos.

1. 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 33-sphere. These flows are equivalent to Hamiltonian flows on star-shaped energy surfaces in 4\mathbb{R}^{4}.

Some of the relevant results on the existence of global surfaces of section for Reeb flows on the tight 33-sphere follow from the theory of pseudo-holomorphic curves in symplectizations, initiated by Hofer in [14]. A global surface of section for a 33-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 33-sphere admits disk-like global surface of section. The disk is part of an S1S^{1}-family of global surfaces of section, forming an open book decomposition. This result applies to Hamiltonian flows on strictly convex energy surfaces in 4\mathbb{R}^{4}.

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 4\mathbb{R}^{4}. 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 33, 22, or 11, and the regular leaves are punctured spheres. If a single periodic orbit forms the binding, then it has Conley-Zehnder index 33, 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 33-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 3233-2-3 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 33. 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 S3S^{3} 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 1-1. 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 33-sphere. More precisely, a simple closed orbit PP bounds a disk-like global surface of section if and only if it is unknotted, has Conley-Zehnder index 3\geq 3, has self-linking number 1-1, and all index-22 orbits link with PP. 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 33-sphere that admit a binding orbit with Conley-Zehnder index 11. More precisely, we study a particular type of transverse foliation, called 3213-2-1 foliation, which has three binding orbits P3P_{3}, P2P_{2}, and P1P_{1}, whose Conley-Zehnder indices are 33, 22, and 11, respectively. The rigid leaves are formed by a plane asymptotic to P2P_{2}, a pair of cylinders connecting P3P_{3} to P2P_{2} forming a 22-torus, and a cylinder connecting P2P_{2} to P1P_{1}. There is a family of planes asymptotic to P3P_{3}, breaking at each end into a cylinder connecting P3P_{3} and P2P_{2} and the plane asymptotic to P2P_{2}. There is also a family of cylinders connecting P3P_{3} to P1P_{1}, which breaks at each end into a cylinder connecting P3P_{3} and P2P_{2} and the cylinder connecting P2P_{2} and P1P_{1}. Our main result provides sufficient conditions for a set of three closed orbits to be the binding orbits of a 3213-2-1 foliation. We also exhibit a Hamiltonian on 4\mathbb{R}^{4} whose flow restricted to suitable energy surfaces admits a 3213-2-1 foliation.

The 3213-2-1 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 S3S^{3}. A transverse foliation for a flow φt\varphi^{t} on a closed oriented 33-manifold MM consists of

  • A finite set 𝒫\mathcal{P} of simple periodic orbits of φt\varphi^{t}, called binding orbits;

  • A smooth foliation of MP𝒫PM\setminus\cup_{P\in\mathcal{P}}P by properly embedded surfaces. Every leaf is transverse to φt\varphi^{t} and has an orientation induced by φt\varphi^{t} and MM. For every leaf Σ˙\dot{\Sigma} there exists a compact embedded surface ΣM\Sigma\ \hookrightarrow M so that Σ˙=ΣΣ\dot{\Sigma}=\Sigma\setminus\partial\Sigma and Σ\partial\Sigma is a union of connected components of P𝒫P\cup_{P\in\mathcal{P}}P. An end zz of Σ˙\dot{\Sigma} is called a puncture. To each puncture zz there is an associated component Pz𝒫P_{z}\in\mathcal{P} of Σ\partial\Sigma, called the asymptotic limit of Σ˙\dot{\Sigma} at zz. A puncture zz of Σ˙\dot{\Sigma} is called positive if the orientation on PzP_{z} induced by Σ\Sigma coincides with the orientation induced by φt\varphi^{t}. Otherwise, zz is called negative.

This definition follows [27] and is based on the finite energy foliations from [21].

Let λ\lambda be a contact form on S3S^{3}, that is, λdλ\lambda\wedge d\lambda is a volume form. The Reeb vector field RλR_{\lambda} associated to λ\lambda is uniquely determined by

(1) iRλdλ0,iRλλ1.i_{R_{\lambda}}d\lambda\equiv 0,~{}~{}~{}~{}~{}~{}i_{R_{\lambda}}\lambda\equiv 1.

The flow {φt}\{\varphi^{t}\} of RλR_{\lambda} is called the Reeb flow of λ\lambda. The contact structure associated to λ\lambda is the 22-plane distribution ξ=kerλ.\xi=\ker\lambda. We denote by π:TS3ξ\pi:TS^{3}\to\xi the projection onto ξ\xi uniquely determined by kerπ=Rλ\ker\pi=\mathbb{R}R_{\lambda}.

An embedded disk DS3D\subset S^{3} satisfying TDξ and TpDξp,pDT\partial D\subset\xi\text{ and }T_{p}D\neq\xi_{p},~{}\forall p\in\partial D is called an overtwisted disk. The contact form λ\lambda is tight if the contact structure ξ=kerλ\xi=\ker\lambda does not admit an overtwisted disk. Consider 4\mathbb{R}^{4} with coordinates (x1,x2,y1,y2)(x_{1},x_{2},y_{1},y_{2}). The Liouville 11-form

(2) λ0=12i=12xidyiyidxi\lambda_{0}=\frac{1}{2}\sum_{i=1}^{2}x_{i}dy_{i}-y_{i}dx_{i}

restricts to a contact form on S3S^{3}. By results of Bennequim [1] and Eliashberg [10], we know that, up to diffeomorphism, any tight contact form on S3S^{3} is of the form fλ0|S3f\lambda_{0}|_{S^{3}} for some smooth function f:S3+f:S^{3}\to\mathbb{R}^{+}. If E:={zf(z)|zS3}E:=\{z\sqrt{f(z)}|z\in S^{3}\} is a regular energy level of a Hamiltonian function on (4,ω0:=dλ0)(\mathbb{R}^{4},\omega_{0}:=d\lambda_{0}), then the associated Hamiltonian flow restricted to EE is equivalent to the Reeb flow of fλ0|S3f\lambda_{0}|_{S^{3}}.

We call a pair P=(x,T)P=(x,T), where x:S3x:\mathbb{R}\to S^{3} is a periodic trajectory of x˙(t)=Rλ(x(t))\dot{x}(t)=R_{\lambda}(x(t)) and T>0T>0 is a period of xx, a Reeb orbit. We identify P=(x,T)P=(x,T) with the element of C(/,S3)/\dfrac{C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3})}{\mathbb{R}/\mathbb{Z}} induced by the loop xT:/S3,xT(t)=x(Tt),x_{T}:{\mathbb{R}}/{\mathbb{Z}}\to S^{3},~{}~{}~{}x_{T}(t)=x(Tt), where the quotient is relative to the translations txT(t+c)t\mapsto x_{T}(t+c). By abuse of notation we sometimes write x()=Px(\mathbb{R})=P. If TT is the minimal positive period of xx, we call PP simple. If m1m\geq 1 is an integer, the mthm^{th} iterate of PP will be denoted by Pm:=(x,mT)P^{m}:=(x,mT). We denote the set of Reeb orbits by 𝒫(λ)\mathcal{P}(\lambda).

The Reeb flow φt\varphi^{t} preserves the contact structure ξ\xi and the maps dφt:ξx(0)ξx(t)d\varphi^{t}:\xi_{x(0)}\to\xi_{x(t)} are dλd\lambda-symplectic. We call the orbit P=(x,T)P=(x,T) nondegenerate if 11 is not an eigenvalue of dφT|ξx(0)d\varphi^{T}|_{\xi_{x(0)}}. If every orbit P𝒫(λ)P\in\mathcal{P}(\lambda) is nondegenerate, then the contact form λ\lambda is called nondegenerate.

A simple orbit P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) is called unknotted if x()x(\mathbb{R}) is an unknot. We say that a set of simple orbits i=1nPi=(xi,Ti)\cup_{i=1}^{n}P_{i}=(x_{i},T_{i}) is an unlink if i=1xi()\cup_{i=1}{x_{i}}(\mathbb{R}) is an unlink. We say that two orbits P=(x,T)P=(x,T) and P¯=(x¯,T¯)\bar{P}=(\bar{x},\bar{T}) are linked if the linking number lk(x(),x¯()){\rm lk}(x(\mathbb{R}),\bar{x}(\mathbb{R})) is nonzero.

Before stating our main results, we give some necessary definitions.

Definition 1.1 (Strong transverse section).

Let λ\lambda be a contact form on S3S^{3}. Let ΣS3\Sigma\hookrightarrow S^{3} be a compact embedded surface such that Σ˙=ΣΣ\dot{\Sigma}=\Sigma\setminus\partial\Sigma is transverse to the Reeb vector field RλR_{\lambda} and Σ\partial\Sigma consists of a finite number of simple orbits in 𝒫(λ)\mathcal{P}(\lambda). Σ\Sigma is called a strong transverse section if every connected component of Σ\partial\Sigma associated to an orbit P=(x,T)P=(x,T) has a neighborhood on Σ\Sigma parametrized by ψ:(r0,1]×/Σ\psi:(r_{0},1]\times\mathbb{R}/\mathbb{Z}\to\Sigma such that ψ(1,t)=xT(t)\psi(1,t)=x_{T}(t), t/\forall t\in\mathbb{R}/\mathbb{Z}, and the section of xTξx_{T}^{*}\xi defined by

(3) η(t)=πrψ(r,t)|r=1\eta(t)=\pi\frac{\partial}{\partial r}\psi(r,t)\big{|}_{r=1}

satisfies

dλ(η(t),Rλη(t))0,t/,d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))\neq 0,~{}\forall t\in\mathbb{R}/\mathbb{Z},

where Rλ\mathcal{L}_{R_{\lambda}} is the Lie derivative with respect to RλR_{\lambda}. See Remark 2.8.

Definition 1.2 (3213-2-1 foliation).

Let λ\lambda be a contact form on S3S^{3}. A 3213-2-1 foliation adapted to λ\lambda is a transverse foliation for the associated Reeb flow satisfying the following properties. The set 𝒫\mathcal{P} of binding orbits consists of three simple orbits P1P_{1}, P2P_{2}, and P3P_{3} with Conley-Zehnder indices 11, 22, and 33 respectively, self-linking number 1-1, and such that P1P2P3P_{1}\cup P_{2}\cup P_{3} is an unlink. The foliation \mathcal{F} of S3P𝒫PS^{3}\setminus\cup_{P\in\mathcal{P}}P is as follows:

  • \mathcal{F} contains a pair of cylinders U1U_{1} and U2U_{2}, both asymptotic to P3P_{3} at their positive punctures and P2P_{2} at their negative punctures. T:=U1U2P2P3T:=U_{1}\cup U_{2}\cup P_{2}\cup P_{3} is homeomorphic to a torus and TP3T\setminus P_{3} is C1C^{1}-embedded. TT divides S3S^{3} into two closed regions 1\mathcal{R}_{1} and 2\mathcal{R}_{2}.

  • \mathcal{F} contains a disk D1D\subset\mathcal{R}_{1} asymptotic to P2P_{2} at its positive puncture and a cylinder V2V\subset\mathcal{R}_{2} asymptotic to P2P_{2} at its positive puncture and P1P_{1} at its negative puncture. DP2VD\cup P_{2}\cup V is a C1C^{1}-embedded disk with boundary P1P_{1} and transverse to TT.

  • \mathcal{F} contains a one-parameter family of disks Fτ1F_{\tau}\subset\mathcal{R}_{1}, τ(0,1)\tau\in(0,1), all of them asymptotic to P3P_{3} at their positive punctures, such that {Fτ}τ(0,1)U1U2D\{F_{\tau}\}_{\tau\in(0,1)}\cup U_{1}\cup U_{2}\cup D foliate 1(P2P3)\mathcal{R}_{1}\setminus(P_{2}\cup P_{3}).

  • \mathcal{F} contains a one-parameter family of cylinders Cτ2C_{\tau}\subset\mathcal{R}_{2}, τ(0,1)\tau\in(0,1), all of them asymptotic to P3P_{3} at their positive punctures and P1P_{1} at their negative punctures, such that {Cτ}τ(0,1)U1U2V\{C_{\tau}\}_{\tau\in(0,1)}\cup U_{1}\cup U_{2}\cup V foliate 2(P1P2P3)\mathcal{R}_{2}\setminus(P_{1}\cup P_{2}\cup P_{3}).

  • The closure of every leaf of \mathcal{F} is a strong transverse section.

See figure 1. The Conley-Zehnder index and the self-linking number are discussed in Section 2.

Refer to caption
Figure 1. A section of a 3213-2-1 foliation. The black dots represent the orbit P3P_{3} with Conley-Zehnder index 33, the white dots represent the hyperbolic orbit P2P_{2} with Conley-Zehnder index 22, and the gray dots represent the orbit P1P_{1} with Conley-Zehnder index 11. The bold curves represent the two rigid cylinders connecting P2P_{2} and P3P_{3}, the rigid cylinder connecting P1P_{1} and P2P_{2}, and the rigid disk with boundary P2P_{2}. The dashed curves represent the family of disks with boundary P3P_{3}. The dotted curves represent the family of cylinders connecting P1P_{1} and P3P_{3}. The arrows indicate the direction of the Reeb vector field. The 33-sphere is viewed as 3{}\mathbb{R}^{3}\cup\{\infty\}.

A complex structure JJ on ξ\xi is dλd\lambda-compatible if the bilinear form dλ(,J)d\lambda(\cdot,J\cdot) is a positive inner product on ξp\xi_{p}, pS3\forall p\in S^{3}. The space of dλd\lambda-compatible structures is nonempty and will be denoted by 𝒥(λ)\mathcal{J}(\lambda). For each J𝒥(λ)J\in\mathcal{J}(\lambda), the pair (λ,J)(\lambda,J) determines a natural almost complex structure J~\tilde{J} in the symplectization ×S3\mathbb{R}\times S^{3} of S3S^{3}, given by (23). We consider J~\tilde{J}-holomorphic curves u~:S2Γ×S3\tilde{u}:S^{2}\setminus\Gamma\to\mathbb{R}\times S^{3}, where the domain is the Riemann sphere with a finite set Γ\Gamma of punctures removed. Due to results of [14] and [18], if u~\tilde{u} satisfies an energy condition, it approaches closed orbits of RλR_{\lambda} near the punctures. We postpone the precise definitions and statements to Section 2. In what follows we use the notation J~=(λ,J)\tilde{J}=(\lambda,J).

Now we state the main results of this paper.

Theorem 1.3.

Let λ\lambda be a tight contact form on S3S^{3}. Assume that there exist Reeb orbits P1=(x1,T1),P2=(x2,T2),P3=(x3,T3)𝒫(λ)P_{1}=(x_{1},T_{1}),~{}P_{2}=(x_{2},T_{2}),~{}P_{3}=(x_{3},T_{3})\in\mathcal{P}(\lambda) that are nondegenerate, simple, and have Conley-Zehnder indices respectively 11, 22, and 33. Assume further that the orbits P1P_{1}, P2P_{2}, and P3P_{3} are unknotted, PiP_{i} and PjP_{j} are not linked for iji\neq j, i,j{1,2,3}i,j\in\{1,2,3\}, and the following conditions hold:

  1. (i)

    Every orbit in 𝒫(λ)\mathcal{P}(\lambda) having period T3\leq T_{3} is nondegenerate;

  2. (ii)

    P2P_{2} is the unique orbit in 𝒫(λ)\mathcal{P}(\lambda) with Conley-Zehnder index 22 and period less than T3T_{3} that is not linked to P3P_{3};

  3. (iii)

    P1P_{1} is the unique orbit in 𝒫(λ)\mathcal{P}(\lambda) with Conley-Zehnder index 11 and period less than T2T_{2} that is not linked to P2P_{2};

  4. (iv)

    There exists J𝒥(λ)J\in\mathcal{J}(\lambda) such that the almost complex structure J~=(λ,J)\tilde{J}=(\lambda,J) admits a finite energy plane u~:×S3\tilde{u}:\mathbb{C}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at it positive puncture z=z=\infty;

  5. (v)

    There is no C1C^{1}-embedding Ψ:S23S3\Psi:S^{2}\subset\mathbb{R}^{3}\to S^{3} such that Ψ(S1×{0})=x2()\Psi({S^{1}\times\{0\}})=x_{2}(\mathbb{R}) and each hemisphere is a strong transverse section.

Then there exists a 3213-2-1 foliation adapted to λ\lambda with binding orbits P1P_{1}, P2P_{2}, and P3P_{3}. Consequently, there exists at least one homoclinic orbit to P2P_{2}.

Remark 1.4.

We expect to weaken hypothesis (iv) in Theorem 1.3 by assuming the existence of a disk with boundary P3P_{3} 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 3213-2-1 foliation have positive topological entropy if no two branches of the stable and unstable manifold of P2P_{2} coincide.

Condition (v) in Theorem 1.3 is necessary to the existence of a 3213-2-1 foliation. This is the content of Proposition 1.5 below.

Proposition 1.5.

Assume that there exists a 3213-2-1 foliation adapted to the contact form λ\lambda on S3S^{3} and let P2=(x2,T2)P_{2}=(x_{2},T_{2}) be the binding orbit with Conley-Zehnder index 22, as in Definition 1.2. Then there is no C1C^{1}-embedding ψ:S2S3\psi:S^{2}\to S^{3} such that ψ(S1×{0})=x2()\psi({S^{1}\times\{0\}})=x_{2}(\mathbb{R}) and each hemisphere is a strong transverse section.

The following theorem gives another set of sufficient conditions for the existence of a 3213-2-1 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 λ\lambda be a tight contact form on S3S^{3}. Assume that there exist Reeb orbits P1=(x1,T1),P2=(x2,T2),P3=(x3,T3)𝒫(λ)P_{1}=(x_{1},T_{1}),~{}P_{2}=(x_{2},T_{2}),~{}P_{3}=(x_{3},T_{3})\in\mathcal{P}(\lambda) that are nondegenerate, simple, and have Conley-Zehnder indices respectively 11, 22, and 33. Assume further that the orbits P1P_{1}, P2P_{2}, and P3P_{3} are unknotted, PiP_{i} and PjP_{j} are not linked for iji\neq j, i,j{1,2,3}i,j\in\{1,2,3\}, and the following conditions hold:

  1. (i)

    T1<T2<T3<2T1T_{1}<T_{2}<T_{3}<2T_{1};

  2. (ii)

    If P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) satisfies PP3,TT3P\neq P_{3},~{}T\leq T_{3} and lk(P,P3)=0{\rm lk}(P,P_{3})=0, then P{P1,P2}P\in\{P_{1},P_{2}\}.

  3. (iii)

    There exists J𝒥(λ)J\in\mathcal{J}(\lambda) such that the almost complex structure J~=(λ,J)\tilde{J}=(\lambda,J) admits a finite energy plane u~:×S3\tilde{u}:\mathbb{C}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at its positive puncture z=z=\infty and a finite energy cylinder w~:{0}×S3\tilde{w}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at its positive puncture z=z=\infty and P1P_{1} at its negative puncture z=0z=0;

  4. (iv)

    There exists no C1C^{1}-embedding Ψ:S23S3\Psi:S^{2}\subset\mathbb{R}^{3}\to S^{3} such that Ψ(S1×{0})=x2()\Psi({S^{1}\times\{0\}})=x_{2}(\mathbb{R}) and each hemisphere is a strong transverse section.

Then there exists a 3213-2-1 foliation adapted to λ\lambda with binding orbits P1P_{1}, P2P_{2}, and P3P_{3}. Consequently, there exists at least one homoclinic orbit to P2P_{2}.

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 33-sphere admitting adapted open books with disk-like pages. Hypotheses (i) and (ii) can be checked if the orbits P1P_{1}, P2P_{2}, and P3P_{3} 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 3213-2-1 foliation

Consider 4\mathbb{R}^{4} with coordinates (x1,y1,x2,y2)(x_{1},y_{1},x_{2},y_{2}) and equipped with the canonical symplectic form ω0=i=12dxidyi.\omega_{0}=\sum_{i=1}^{2}dx_{i}\wedge dy_{i}. Consider the Hamiltonian function H=Hϵ:4H=H_{\epsilon}:\mathbb{R}^{4}\to\mathbb{R} defined by

(4) H(x1,y1,x2,y2)=H1(x1,y1)+H2(x2,y2),H(x_{1},y_{1},x_{2},y_{2})=H_{1}(x_{1},y_{1})+H_{2}(x_{2},y_{2}),
(5) H1(x1,y1)\displaystyle H_{1}(x_{1},y_{1}) =x12+y122\displaystyle=\frac{x_{1}^{2}+y_{1}^{2}}{2}
H2(x2,y2)\displaystyle H_{2}(x_{2},y_{2}) =(x22+y22)2ϵ(x22+y22)x2ϵ2(x22+y22)\displaystyle=(x_{2}^{2}+y_{2}^{2})^{2}-\epsilon(x^{2}_{2}+y_{2}^{2})x_{2}-\frac{\epsilon}{2}(x_{2}^{2}+y_{2}^{2})

If ϵ>0\epsilon>0, then H2H_{2} has exactly three critical points: a saddle (p2,0)(p_{2},0), a local maximum (p1,0)(p_{1},0) and a minimum (p3,0)(p_{3},0), where p2<p1=0<p3p_{2}<p_{1}=0<p_{3}. See Figure 2.

Refer to caption
Figure 2. Energy levels of H2H_{2} for ϵ=1\epsilon=1.

The energy level S:=H1(12)S:=H^{-1}\left(\frac{1}{2}\right) is star-shaped: If z=(x1,y1,x2,y2)Sz=(x_{1},y_{1},x_{2},y_{2})\in S and YY is the radial vector field Y(z)=12zY(z)=\frac{1}{2}z, then, for sufficiently small ϵ\epsilon, we have

dHz(Y)\displaystyle dH_{z}(Y) =x12+y122+2(x22+y22)2ϵ(x22+y22)xϵ2(x22+y22)ϵ(x22+y22)x2\displaystyle=\frac{x_{1}^{2}+y_{1}^{2}}{2}+2(x_{2}^{2}+y_{2}^{2})^{2}-\epsilon(x_{2}^{2}+y_{2}^{2})x-\frac{\epsilon}{2}(x_{2}^{2}+y_{2}^{2})-\epsilon(x_{2}^{2}+y_{2}^{2})x_{2}
=H(z)+(x22+y22)2ϵ(x22+y22)x2\displaystyle=H(z)+(x_{2}^{2}+y_{2}^{2})^{2}-\epsilon(x_{2}^{2}+y_{2}^{2})x_{2}
>0\displaystyle>0

It follows that SS is diffeomorphic to S3S^{3} and λ:=λ0|S\lambda:=\lambda_{0}|_{S} is a contact form on SS. The Hamiltonian vector field associated to H:4H:\mathbb{R}^{4}\to\mathbb{R} is given by

(6) XH(x1,y1,x2,y2)=(y1,x1,P(x2,y2),Q(x2,y2)),X_{H}(x_{1},y_{1},x_{2},y_{2})=(-y_{1},x_{1},-P(x_{2},y_{2}),Q(x_{2},y_{2})),

where

(7) Q(x2,y2)\displaystyle Q(x_{2},y_{2}) =H2(x2,y2)x2=4x2(x22+y22)3ϵx22ϵy22ϵx2\displaystyle=\dfrac{\partial H_{2}(x_{2},y_{2})}{\partial x_{2}}=4x_{2}(x_{2}^{2}+y_{2}^{2})-3\epsilon x_{2}^{2}-\epsilon y_{2}^{2}-\epsilon x_{2}
P(x2,y2)\displaystyle P(x_{2},y_{2}) =H2(x2,y2)y2=4y2(x22+y22)2ϵx2y2ϵy2.\displaystyle=\dfrac{\partial H_{2}(x_{2},y_{2})}{\partial y_{2}}=4y_{2}(x_{2}^{2}+y_{2}^{2})-2\epsilon x_{2}y_{2}-\epsilon y_{2}.

The Reeb vector field RλR_{\lambda} is given by Rλ(z)=h(z)XH(z),for z=(x1,y1,x2,y2)S,R_{\lambda}(z)=h(z)X_{H}(z),~{}~{}\text{for }z=(x_{1},y_{1},x_{2},y_{2})\in S, where

(8) h(z)=λz(XH(z))1=2(x12+y12+x2Q(x2,y2)+y2P(x2,y2))1.h(z)=\lambda_{z}(X_{H}(z))^{-1}=2(x_{1}^{2}+y_{1}^{2}+x_{2}Q(x_{2},y_{2})+y_{2}P(x_{2},y_{2}))^{-1}.

Consider the Reeb orbits P1=(γ1,T1)P_{1}=(\gamma_{1},T_{1}), P2=(γ2,T2)P_{2}=(\gamma_{2},T_{2}) and P3=(γ3,T3)P_{3}=(\gamma_{3},T_{3}) where

(9) γi(t)=(ricos(2ri2t),risin(2ri2t),pi,0),for ri=12H2(pi,0),\gamma_{i}(t)=\left(r_{i}\cos\left(\frac{2}{r_{i}^{2}}t\right),r_{i}\sin\left(\frac{2}{r_{i}^{2}}t\right),p_{i},0\right),~{}~{}\text{for }r_{i}=\sqrt{1-2H_{2}(p_{i},0)},
(10) Ti=πri2.T_{i}=\pi r_{i}^{2}.

Note that H2(p3,0)<H2(p2,0)<H2(p1,0)H_{2}(p_{3},0)<H_{2}(p_{2},0)<H_{2}(p_{1},0), which implies that

(11) T1<T2<T3.T_{1}<T_{2}<T_{3}.

Since H2(pi,0)=O(ϵ2)H_{2}(p_{i},0)={O}(\epsilon^{2}), for each i=1,2,3i=1,2,3, we have

(12) T3<2T1,T_{3}<2T_{1},

for sufficiently small ϵ\epsilon. It is easy to see that P1P2P3P_{1}\cup P_{2}\cup P_{3} is a trivial link.

Proposition 1.7.

For sufficiently small ϵ\epsilon, the contact form λ=λ0|S\lambda=\lambda_{0}|_{S}, where S=H1(12)S=H^{-1}(\frac{1}{2}), and the Reeb orbits P1P_{1}, P2P_{2}, and P3P_{3}, defined by (9)-(10), satisfy the hypotheses of Theorem 1.6. Therefore, there exists a 3213-2-1 foliation adapted to λ\lambda with binding orbits P1P_{1}, P2P_{2}, and P3P_{3}.

Refer to caption
Figure 3. Projection of a 3213-2-1 foliation onto the plane (x2,y2)(x_{2},y_{2}).
Remark 1.8.

If ϵ<0\epsilon<0 and |ϵ||\epsilon| is sufficiently small, S=H1(12)S=H^{-1}\left(\frac{1}{2}\right) is strictly convex, which implies (see [19]) that λ=λ0|S\lambda=\lambda_{0}|_{S} is dynamically convex. The simple Reeb orbit PP with image H11(12)×{0}H_{1}^{-1}\left(\frac{1}{2}\right)\times\{0\} has Conley-Zehnder index 33, self-linking number 1-1, and is unknotted. It follows from the main statement of [23] that PP is binding of an open book decomposition with disk-like pages adapted to λ\lambda, where each page is a global surface of section for the associated Reeb flow. The orbits P1P_{1}, P2P_{2} and P3P_{3} bifurcate from PP when ϵ=0\epsilon=0.

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 \mathbb{R}-invariant almost complex structure J~=(λ,J)\tilde{J}=(\lambda,J) admitting a finite energy plane asymptotic to P3P_{3}. By the results of [17] and the Fredholm theory developed in [20], after a quotient by the natural \mathbb{R}-action, the plane lives in a one-dimensional family u~τ=(aτ,uτ):×S3,τ(0,1)\tilde{u}_{\tau}=(a_{\tau},u_{\tau}):\mathbb{C}\to\mathbb{R}\times S^{3},~{}~{}\tau\in(0,1), where each projection uτ:S3u_{\tau}:\mathbb{C}\to S^{3} 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 P3P_{3} to P2P_{2} and a rigid plane asymptotic to P2P_{2}. We use hypothesis (v) and the intersection theory developed in [31] to show that the rigid plane to P2P_{2} is common to the two ends.

Refer to caption
Figure 4. The pair of rigid cylinders U1=ur({0})U_{1}=u_{r}(\mathbb{C}\setminus\{0\}), U2=ur({0})U_{2}=u_{r}^{\prime}(\mathbb{C}\setminus\{0\}), the rigid disk D=uq()D=u_{q}(\mathbb{C}) and the family of disks Fτ=uτ()F_{\tau}=u_{\tau}(\mathbb{C}), τ(0,1)\tau\in(0,1), foliating the region 1\mathcal{R}_{1}.

The projection of the one-parameter family of planes u~τ\tilde{u}_{\tau}, the two rigid cylinders u~r=(ar,ur),u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}),\tilde{u}^{\prime}_{r}=(a^{\prime}_{r},u^{\prime}_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}, and the rigid plane u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3} to S3S^{3} determine a foliation of a closed region 1\mathcal{R}_{1} homeomorphic to a solid torus. See figure 4.

The next step of the proof is Proposition 4.1, where we obtain a rigid cylinder v~r=(br,vr):{0}×S3\tilde{v}_{r}=(b_{r},v_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} from P2P_{2} to P1P_{1}. We consider a non-cylindrical symplectic cobordism between (S3,λ)(S^{3},\lambda) and (S3,λE)(S^{3},\lambda_{E}), where λE\lambda_{E} is a dynamically convex contact form, adapting the ideas of [19]. We define an non \mathbb{R}-invariant almost complex structure J¯\bar{J} on the cobordism such that J¯=J~\bar{J}=\tilde{J} on [2,+)×S3[2,+\infty)\times S^{3}. After an \mathbb{R}-translation, the plane u~q\tilde{u}_{q} is J¯\bar{J}-holomorphic. By the Fredholm theory of [20], the plane lives in a one-dimensional family of J¯\bar{J}-holomorphic planes asymptotic to P2P_{2}. 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 \mathbb{R}-invariant almost complex structure. Using the linking hypotheses and intersection theory, we show that the first level is a cylinder from P2P_{2} to P1P_{1} and projects onto an embedded cylinder in the complement of the region 1\mathcal{R}_{1}.

Refer to caption
Figure 5. The foliation of 1\mathcal{R}_{1}, the cylinder V=vr({0})V=v_{r}(\mathbb{C}\setminus\{0\}) connecting the orbits P2P_{2} and P1P_{1}, and a cylinder Cτ=wτ({0})C_{\tau}=w_{\tau}(\mathbb{C}\setminus\{0\}) connecting P3P_{3} and P1P_{1}.

Next, we glue the cylinders u~r\tilde{u}_{r} and v~r\tilde{v}_{r} to obtain an one-dimensional family w~τ=(cτ,wτ):{0}×S3\tilde{w}_{\tau}=(c_{\tau},w_{\tau}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} of J~\tilde{J}-holomorphic cylinders from P3P_{3} to P1P_{1}. Using hypotheses (ii) we prove that this family breaks into a two-level building formed by the cylinders u~r\tilde{u}^{\prime}_{r} and v~r\tilde{v}_{r}. In Proposition 5.1 we prove that the cylinders Cτ=wτ({0})C_{\tau}=w_{\tau}(\mathbb{C}\setminus\{0\}) complete the foliation.

We show that the orbits P1P_{1}, P2P_{2} and P3P_{3} have self-linking number 1-1 in Proposition 5.15. This follows for the orbits P2P_{2} and P3P_{3} since these orbits bound disks transverse to the Reeb flow. For the orbit P1P_{1} we produce a disk gluing the cylinder V¯\bar{V} and the disk D¯\bar{D}.

The existence of the 3213-2-1 foliation and the arguments in [21, Proposition 7.5] imply the existence of at least one homoclinic orbit to P2P_{2}. 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 {Fτ}\{F_{\tau}\}, τ(0,1)\tau\in(0,1), the one-parameter family of cylinders {Cτ}\{C_{\tau}\}, τ(0,1)\tau\in(0,1), and the cylinders U1,U2U_{1},U_{2} as in Definition 1.2. The local unstable manifold Wloc(P2)W_{loc}^{-}(P_{2}) intersects FτF_{\tau} transversally, for τ\tau close to 0, in an embedded circle bounding an embedded closed disk BF,τ,0FτB_{F,\tau,0}\subset F_{\tau} with dλd\lambda-area T2T_{2}. The local stable manifold Wloc+(P2)W^{+}_{loc}(P_{2}) intersects FτF_{\tau^{\prime}} transversally, for τ\tau^{\prime} close to 11, in an embedded circle bounding an embedded closed disk Bτ+FτB^{+}_{\tau^{\prime}}\subset F_{\tau^{\prime}} with dλd\lambda-area T2T_{2}. All points in FτBτ+F_{\tau^{\prime}}\setminus B^{+}_{\tau^{\prime}} correspond to trajectories that exit 1\mathcal{R}_{1} through the cylinder U2U_{2}. In the same way, Wloc+(P2)W^{+}_{loc}(P_{2}) intersects CτC_{\tau^{\prime}}, for τ\tau^{\prime} close to 11, in an embedded circle Sτ+S^{+}_{\tau^{\prime}} such that P1Sτ+P_{1}\cup S^{+}_{\tau^{\prime}} is the boundary of a closed region Rτ+R^{+}_{\tau^{\prime}} with dλd\lambda-area T2T1>0T_{2}-T_{1}>0. All points in CτRτ+C_{\tau^{\prime}}\setminus R^{+}_{\tau^{\prime}} correspond to trajectories entering 1\mathcal{R}_{1} through the cylinder U1U_{1}.

The existence of a 3213-2-1 adapted to λ\lambda implies that the forward flow sends BF,τ,0B_{F,\tau,0}, into a disk BF,1B_{F,1} inside FτF_{\tau^{\prime}}, for τ\tau^{\prime} close to 11. If BF,1B_{F,1} intersects Bτ+B^{+}_{\tau^{\prime}}, then, since both disks have the same area, their boundaries also intersect, and a homoclinic to P2P_{2} exists. Otherwise, BF,1(FτBτ+)B_{F,1}\subset\left(F_{\tau^{\prime}}\setminus B^{+}_{\tau^{\prime}}\right) and the forward flow sends BF,1B_{F,1} into a disk BC,τ,1CτB_{C,\tau,1}\subset C_{\tau}, for τ\tau close to 0, with dλd\lambda-area T2T_{2}. The forward flow sends BC,τ,1CτB_{C,\tau,1}\subset C_{\tau} into a disk BC,2B_{C,2} inside CτC_{\tau^{\prime}}. If BC,2B_{C,2} intersects Rτ+R_{\tau^{\prime}}^{+}, then, since the area of Rτ+R_{\tau^{\prime}}^{+} is T2T1<T2T_{2}-T_{1}<T_{2}, their boundaries also intersect and there exists a homoclinic to P2P_{2}. Otherwise, BC,2B_{C,2} is contained in CτRτ+C_{\tau^{\prime}}\setminus R_{\tau^{\prime}}^{+} and the forward flow sends BC,2B_{C,2} into a disk BF,τ,2FτB_{F,\tau,2}\subset F_{\tau}, which has area T2T_{2} and is disjoint from BF,τ,0B_{F,\tau,0}. Proceeding this way, we construct disks BF,τ,2nFτB_{F,\tau,2n}\subset F_{\tau}, BC,τ,2n+1CτB_{C,\tau,2n+1}\subset C_{\tau}, for τ\tau close to 0, and BF,2n+1FτB_{F,2n+1}\subset F_{\tau^{\prime}}, BC,2n+2CτB_{C,2n+2}\subset C_{\tau^{\prime}}, for τ\tau^{\prime} close to 11. This procedure must end at some point since the disks BF,τ,2nB_{F,\tau,2n} and BC,τ,2n+1B_{C,\tau,2n+1} are disjoint and have the same area T2T_{2}, while the areas of FτF_{\tau} and CτC_{\tau} are T3T_{3} and T3T1T_{3}-T_{1}, respectively. This implies that either BF,2n+1B_{F,2n+1} intersects Bτ+B^{+}_{\tau^{\prime}} or BC,2n+2B_{C,2n+2} intersects Rτ+R_{\tau^{\prime}}^{+}, for some integer n0n\geq 0. In any case, a homoclinic to P2P_{2} must exist.

To prove Theorem 1.6, we use hypothesis (iii) to obtain a one-dimensional family of pseudo-holomorphic cylinders from P3P_{3} to P1P_{1}. 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 P3P_{3} to P2P_{2} and a rigid cylinder from P2P_{2} to P1P_{1}.

2. Preliminaries

In this section, λ\lambda is a contact form on a 33-manifold MM, and ξ=kerλ\xi=\ker\lambda is the associated contact structure.

2.1. The Conley-Zehnder index

Let φt\varphi^{t} be the Reeb flow associated to the contact form λ\lambda. The bilinear form dλd\lambda turns ξ=kerλ\xi=\ker\lambda into a symplectic vector bundle and the linearized flow dφxt:ξxξφt(x)d\varphi^{t}_{x}:\xi_{x}\to\xi_{\varphi^{t}(x)} is symplectic with respect to dλd\lambda.

Let P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) be a nondegenerate Reeb orbit. Let Ψ:xTξ/×2\Psi:x_{T}^{*}\xi\to\mathbb{R}/\mathbb{Z}\times\mathbb{R}^{2} be a trivialization of the symplectic vector bundle (xTξ,dλ)(x_{T}^{*}\xi,d\lambda) and consider the arc of symplectic matrices Φ([0,1],Sp(1))\Phi\in\mathbb{C}^{\infty}([0,1],Sp(1)) defined by Φ(t)=ΨtdφTt|ξx(0)Ψ01.\Phi(t)=\Psi_{t}\circ d\varphi^{Tt}|_{\xi_{x(0)}}\circ\Psi_{0}^{-1}~{}. Let z{0}z\in\mathbb{C}\setminus\{0\} and let θ:[0,1]\theta:[0,1]\to\mathbb{R} be a continuous argument for z(t):=Φ(t)zz(t):=\Phi(t)z, that is, e2πθ(t)=z(t)|z(t)|,0t1e^{2\pi\theta(t)}=\frac{z(t)}{|z(t)|},~{}\forall~{}0\leq t\leq 1. Define the winding number of z(t)=Φ(t)zz(t)=\Phi(t)z by

(13) Δ(z)=θ(1)θ(0)\Delta(z)=\theta(1)-\theta(0)\in\mathbb{R}

and the winding interval of the arc Φ\Phi by I(Φ)={Δ(z)|z{0}}.I(\Phi)=\{\Delta(z)|z\in\mathbb{C}\setminus\{0\}\}~{}. The interval I(Φ)I(\Phi) is compact and its length is strictly smaller than 12\frac{1}{2}. Since PP is nondegenerate, we have I(Φ)=\partial I(\Phi)\cap\mathbb{Z}=\emptyset, 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 PP relative to the trivialization Ψ\Psi is defined by

(14) μ(P,Ψ)={2k+1, if I(Φ)(k,k+1)2k, if kI(Φ)\mu(P,\Psi)=\left\{\begin{array}[]{lr}2k+1,\text{ if }I(\Phi)\subset(k,k+1)\\ 2k,\text{ if }k\in I(\Phi)\end{array}\right.

This index only depends on the homotopy class of the trivialization Ψ\Psi.

Throughout the paper, we only deal with the tight contact structure on S3S^{3}, which is a trivial symplectic bundle. In this case, the Conley-Zehnder index is independent of the choice of global symplectic trivialization of ξ\xi. We define the Conley-Zehnder index of the Reeb orbit PP by

(15) μ(P):=μ(P,Ψ),\mu(P):=\mu(P,\Psi),

for any global symplectic trivialization Ψ:ξS3×2\Psi:\xi\to S^{3}\times\mathbb{R}^{2}.

We say that the Reeb orbit P=(x,T)P=(x,T) is positive hyperbolic if σ(Φ(1))(0,){1}\sigma(\Phi(1))\subset(0,\infty)\setminus\{1\}, negative hyperbolic if σ(Φ(1))(,0){1}\sigma(\Phi(1))\subset(-\infty,0)\setminus\{-1\}, or elliptic if the eigenvalues are in S1{1}S^{1}\setminus\{1\}. 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 λ\lambda be a tight contact form on S3S^{3} and let P=(x,T)P=(x,T) be a nondegenerate Reeb orbit. Let k1k\geq 1 be an integer such that for every l{1,,k}l\in\{1,\cdots,k\}, the orbit Pl=(x,lT)P^{l}=(x,lT) is nondegenerate. The following assertions hold.

  • (1)

    If μ(Pk)=1\mu(P^{k})=1, then μ(Pl)=1\mu(P^{l})=1, l{1,,k}\forall l\in\{1,\cdots,k\};

  • (2)

    μ(Pk)0\mu(P^{k})\leq 0 \iff μ(Pl)0\mu(P^{l})\leq 0, l{1,,k}\forall l\in\{1,\cdots,k\};

  • (3)

    If PP is a hyperbolic orbit, then μ(Pl)=lμ(P)\mu(P^{l})=l\mu(P), l{1,,k}\forall l\in\{1,\cdots,k\};

  • (4)

    If μ(Pk)=2\mu(P^{k})=2, then k{1,2}k\in\{1,2\} and PP is hyperbolic. If k=2k=2, then μ(P)=1\mu(P)=1.

2.2. The asymptotic operator

Fix P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) and let h:S1TMh:S^{1}\to TM be a vector field along xT:S1=/Mx_{T}:S^{1}=\mathbb{R}/\mathbb{Z}\to M. The Lie derivative Rλh\mathcal{L}_{R_{\lambda}}h of hh with respect to RλR_{\lambda} is defined by

(16) Rλh(t)=dds|s=0dφs(x(Tt+s))h(t+sT),\mathcal{L}_{R_{\lambda}}h(t)=\dfrac{d}{ds}\bigg{|}_{s=0}d\varphi^{-s}(x(Tt+s))h\left(t+\frac{s}{T}\right)~{},

where φt\varphi^{t} is the flow of RλR_{\lambda}. Let \nabla be a symmetric connection on TMTM. We can use dxT(t)t=TRλ(xT(t))dx_{T}(t)\partial_{t}=TR_{\lambda}(x_{T}(t)) to write

(17) TRλh=TRλh=TRλhhTRλ=thThRλ,T\mathcal{L}_{R_{\lambda}}h=\mathcal{L}_{TR_{\lambda}}h=\nabla_{TR_{\lambda}}h-\nabla_{h}TR_{\lambda}=\nabla_{t}h-T\nabla_{h}R_{\lambda},

where t\nabla_{t} is the covariant derivative along xTx_{T}. We conclude that the differential operator tTRλ\nabla_{t}\cdot-T\nabla_{\cdot}R_{\lambda} maps sections of xTξx_{T}^{*}\xi to sections of xTξx_{T}^{*}\xi and is independent of the choice of symmetric connection.

Choosing a dλd\lambda-compatible complex structure J𝒥(λ)J\in\mathcal{J}(\lambda), we associate to the orbit P=(x,T)P=(x,T) the unbounded differential operator

(18) AP,J:𝒟(AP,J)=W1,2(S1,xTξ)L2(S1,xTξ)\displaystyle A_{P,J}:\mathcal{D}(A_{P,J})=W^{1,2}(S^{1},x^{*}_{T}\xi)\subset L^{2}(S^{1},x_{T}^{*}\xi) L2(S1,xTξ)\displaystyle\to L^{2}(S^{1},x_{T}^{*}\xi)
η\displaystyle\eta J(tηTηRλ).\displaystyle\mapsto-J(\nabla_{t}\eta-T\nabla_{\eta}R_{\lambda})~{}.

The operator AP,JA_{P,J} defined by (18) is called the asymptotic operator associated to the orbit PP and the complex structure JJ. In any unitary trivialization Ψ\Psi of (xTξ,dλ,J)(x_{T}^{*}\xi,d\lambda,J), the operator AP,JA_{P,J} takes the form LS:=J0ddtS(t),L_{S}:=-J_{0}\frac{d}{dt}-S(t), where S(t)S(t) is a path of symmetric matrices given by S(t)=J0Φ˙(t)Φ(t)1S(t)=-J_{0}\dot{\Phi}(t)\Phi(t)^{-1} and Φ(t)\Phi(t) is the linearized flow restricted to ξ\xi in the trivialization Ψ\Psi. If η(t)\eta(t) is an eigensection of AP,JA_{P,J} with corresponding eigenvalue λ\lambda\in\mathbb{R}, then n(t):=Ψη(t)n(t):=\Psi\circ\eta(t) satisfies n(t)0n(t)\neq 0 for all tS1t\in S^{1}. It follows that η(t)\eta(t) has a well defined winding number given by

(19) wind(η,Ψ):=deg(tn(t)n(t)).\operatorname{wind}(\eta,\Psi):=\deg\left(t\mapsto\frac{n(t)}{\|n(t)\|}\right).

This definition just depends on the homotopy class of the trivialization Ψ\Psi. The following properties about this winding number are proved in [17].

Proposition 2.2.

[17] The unbounded operator AP,JA_{P,J} has discrete real spectrum accumulating only at ±\pm\infty. Let Ψ\Psi be a unitary trivialization of xTξx_{T}^{*}\xi. Then

  • Given nonzero eigensections η1(t)\eta_{1}(t) and η2(t)\eta_{2}(t) associated to the same eigenvalue λσ(AP,J)\lambda\in\sigma(A_{P,J}), we have wind(η1,Ψ)=wind(η2,Ψ)\operatorname{wind}(\eta_{1},\Psi)=\operatorname{wind}(\eta_{2},\Psi), so that we can define wind(λ,Ψ)=wind(η,Ψ)\operatorname{wind}(\lambda,\Psi)=\operatorname{wind}(\eta,\Psi), for any eigensection η\eta associated to λ\lambda.

  • If λμσ(AP,J)\lambda\neq\mu\in\sigma(A_{P,J}) satisfy wind(λ,Ψ)=wind(μ,Ψ)\operatorname{wind}(\lambda,\Psi)=\operatorname{wind}(\mu,\Psi) and ηλ(t),ημ(t)\eta_{\lambda}(t),\eta_{\mu}(t) are non-vanishing λ,μ\lambda,\mu-eigensections, respectively, then ηλ(t),ημ(t)\eta_{\lambda}(t),\eta_{\mu}(t) are pointwise linearly independent.

  • Given kk\in\mathbb{Z}, there exists precisely two eigenvalues λ,μσ(AP,J)\lambda,\mu\in\sigma(A_{P,J}), counting multiplicities, such that wind(λ,Ψ)=wind(μ,Ψ)=k\operatorname{wind}(\lambda,\Psi)=\operatorname{wind}(\mu,\Psi)=k

  • If λ,μσ(AP,J)\lambda,\mu\in\sigma(A_{P,J}) and λμ\lambda\leq\mu, then wind(λ,Ψ)wind(μ,Ψ)\operatorname{wind}(\lambda,\Psi)\leq\operatorname{wind}(\mu,\Psi).

  • 0σ(AP,J)0\notin\sigma(A_{P,J}) if and only if the orbit P=(x,T)P=(x,T) is nondegenerate.

Define

(20) νPneg=max{ν<0|ν is an eigenvalue of AP}\nu^{neg}_{P}=\max\{\nu<0|\nu\text{ is an eigenvalue of }A_{P}\}
(21) νPpos=min{ν0|ν is an eigenvalue of AP}\nu^{pos}_{P}=\min\{\nu\geq 0|\nu\text{ is an eigenvalue of }A_{P}\}

and given a trivialization Ψ\Psi of xTξx_{T}^{*}\xi, define p=wind(νPpos,Ψ)wind(νPneg,Ψ){0,1}.p=\operatorname{wind}(\nu^{pos}_{P},\Psi)-\operatorname{wind}(\nu^{neg}_{P},\Psi)\in\{0,1\}.

Definition 2.3.

The (generalized) Conley-Zehnder index of the orbit PP relative to the unitary trivialization Ψ\Psi is defined by

(22) μ~(P,Ψ)=2wind(νPneg,Ψ)+p.\tilde{\mu}(P,\Psi)=2\operatorname{wind}(\nu^{neg}_{P},\Psi)+p.

It is proved in [17, Theorem 3.10] that for any nondegenerate orbit P𝒫(λ)P\in\mathcal{P}(\lambda), μ~(P,Ψ)=μ(P,Ψ),\tilde{\mu}(P,\Psi)=\mu(P,\Psi)~{}, where μ(P,Ψ)\mu(P,\Psi) is the Conley-Zehnder index defined by (14).

2.3. Self-linking number

The self-linking number sl(L)\operatorname{sl}(L) of a trivial knot LML\subset M transverse to ξ\xi is defined as follows. Consider MM oriented by λdλ\lambda\wedge d\lambda. Let DMD\subset M be an embedded disk satisfying D=L\partial D=L and let ZZ be a smooth nonvanishing section of ξ|D\xi|_{D}. The section ZZ is used to slightly perturb LL to another knot Lϵ={expx(ϵZx)|xL}L_{\epsilon}=\{\operatorname{exp}_{x}(\epsilon Z_{x})|x\in L\} transverse to ξ\xi and DD, where exp\operatorname{exp} is any exponential map. A choice of orientation for LL induces orientations of DD and LϵL_{\epsilon}. The self-linking number of LL is defined by LϵD,L_{\epsilon}\cdot D\in\mathbb{Z}, where LϵDL_{\epsilon}\cdot D is the oriented intersection number of LϵL_{\epsilon} and DD. If, for instance, ξ\xi is trivial, this definition is independent of the choices of ZZ and DD. If P=(x,T)P=(x,T) is an unknotted Reeb orbit, we define its self-linking number by sl(P)=sl(x())\operatorname{sl}(P)=\operatorname{sl}(x(\mathbb{R})).

2.4. Finite energy surfaces

The symplectization of the contact manifold (M,λ)(M,\lambda) is the symplectic manifold (×S3,d(eaλ))(\mathbb{R}\times S^{3},d(e^{a}\lambda)), where aa is the coordinate on \mathbb{R}. Given a complex structure J𝒥(λ)J\in\mathcal{J}(\lambda), we consider the almost-complex structure J~=(λ,J)\tilde{J}=(\lambda,J) on ×M\mathbb{R}\times M defined by

(23) J~a=Rλ,J~|ξ=J,\tilde{J}\partial_{a}=R_{\lambda},~{}~{}~{}\tilde{J}|_{\xi}=J,

where we see RλR_{\lambda} and ξ\xi as \mathbb{R}-invariant objects on ×M\mathbb{R}\times M. It is easy to check that the almost complex structure J~\tilde{J} defined by (23) is d(eaλ)d(e^{a}\lambda)-compatible.

Let (S,j)(S,j) be a closed Riemann surface and let ΓS\Gamma\subset S be a finite set. Let u~:SΓ×M\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times M be a J~\tilde{J}-holomorphic map, that is, u~\tilde{u} is smooth and satisfies the Cauchy-Riemann equation

¯J~(u~)=12(u~+J~(u~)du~j)=0.\bar{\partial}_{\tilde{J}}(\tilde{u})=\frac{1}{2}\left(\tilde{u}+\tilde{J}(\tilde{u})\circ d\tilde{u}\circ j\right)=0.

The Hofer energy E(u~)E(\tilde{u}) of u~\tilde{u} is defined by

E(u~)=supϕΣSΓu~d(ϕλ),E(\tilde{u})=\sup_{\phi\in\Sigma}\int_{S\setminus\Gamma}\tilde{u}^{*}d(\phi\lambda)~{},

where Σ={ϕC(,[0,1])|ϕ0)}\Sigma=\{\phi\in C^{\infty}(\mathbb{R},[0,1])|\phi^{\prime}\geq 0)\}.

Definition 2.4.

The J~\tilde{J}-holomorphic map u~:SΓ×M\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times M is called a finite energy surface if it satisfies 0<E(u~)<+0<E(\tilde{u})<+\infty.

The elements of Γ\Gamma are called punctures. Let zΓz\in\Gamma be a puncture and take a holomorphic chart φ:(U,0)(φ(U),z)\varphi:(U,0)\to(\varphi(U),z) centered at zz. We call (s,t)φ(e2π(s+it))(s,t)\simeq\varphi(e^{-2\pi(s+it)}) positive exponential coordinates and (s,t)φ(e2π(s+it))(s,t)\simeq\varphi(e^{2\pi(s+it)}) negative exponential coordinates around zz. Set u~(s,t)=u~φ(e2π(s+it))\tilde{u}(s,t)=\tilde{u}\circ\varphi(e^{-2\pi(s+it)}), for s>>1s>>1. Write u~=(a,u)\tilde{u}=(a,u). Using Stokes Theorem, one can prove that the limit

(24) m(z)=lims+{s}×S1uλm(z)=\lim_{s\to+\infty}\int_{\{s\}\times S^{1}}u^{*}\lambda

exists. The puncture zz is called removable if m=0m=0, positive if m>0m>0 and negative if m<0m<0. By an application of Gromov’s removable singularity theorem [12], one can prove that u~\tilde{u} 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 Γ=Γ+Γ\Gamma=\Gamma^{+}\cup\Gamma^{-} to distinguish positive and negative punctures. If u~:SΓ×M\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times M is a finite energy surface, then Γ+\Gamma^{+}\neq\emptyset.

Theorem 2.5.

[14, 18] Let u~=(a,u):SΓ×M\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times M be a finite energy surface. Assume zz is non removable and let ϵ=±1\epsilon=\pm 1 be the sign zz. Fix a sequence sn+s_{n}\to+\infty. Then there exists a nonconstant trajectory of the Reeb flow x:Mx:\mathbb{R}\to M with period T>0T>0 and a subsequence snks_{n_{k}} such that

limk+{tu(snk,t)}={tx(ϵTt)}\lim_{k\to+\infty}\{t\mapsto u(s_{n_{k}},t)\}=\{t\mapsto x(\epsilon Tt)\}

in the C(S1,M)C^{\infty}(S^{1},M) topology. If the orbit (x,T)(x,T) is nondegenerate, then

lims+{tu(s,t)}={tx(ϵTt)}.\lim_{s\to+\infty}\{t\mapsto u(s,t)\}=\{t\mapsto x(\epsilon Tt)\}~{}.

We say that the periodic orbit P=(x,T)P=(x,T) given by Theorem 2.5 is an asymptotic limit of u~\tilde{u} at the puncture zz. If P=(x,T)P=(x,T) is nondegenerate, it is called the asymptotic limit of u~\tilde{u} at zz.

The dλd\lambda-area of a J~\tilde{J}-holomorphic curve u~:SΓ×M\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times M is given by the formula

A(u~)=SΓu~𝑑λ=SΓu𝑑λ.A(\tilde{u})=\int_{S\setminus\Gamma}\tilde{u}^{*}d\lambda=\int_{S\setminus\Gamma}u^{*}d\lambda~{}.

One can check that A(u~)0A(\tilde{u})\geq 0 and A(u~)=0A(\tilde{u})=0 if and only if πdu0\pi\cdot du\equiv 0. The following theorem concerning J~\tilde{J}-holomorphic curves with vanishing dλd\lambda-area will be useful throughout the paper.

Theorem 2.6.

[17, Theorem 6.11] Let u~=(a,u):Γ×M\tilde{u}=(a,u):\mathbb{C}\setminus\Gamma\to\mathbb{R}\times M be a finite energy sphere, where Γ\Gamma\subset\mathbb{C} is the finite set of negative punctures and \infty is the unique positive puncture. If πdu0\pi\cdot du\equiv 0, then there exists a nonconstant polynomial p:p:\mathbb{C}\to\mathbb{C} and a Reeb orbit P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) such that p1(0)=Γp^{-1}(0)=\Gamma and u~=FPp\tilde{u}=F_{P}\circ p , where FP:{0}×MF_{P}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times M is defined by FP(z=e2π(s+it))=(Ts,x(Tt))F_{P}(z=e^{2\pi(s+it)})=(Ts,x(Tt)).

2.5. Asymptotic behavior

A Martinet’s tube for a simple orbit P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) is a pair (U,ψ)(U,\psi), where UU is a neighborhood of x()x(\mathbb{R}) in MM and ψ:US1×B\psi:U\to S^{1}\times B is a diffeomorphism (here B2B\subset\mathbb{R}^{2} is an open ball centered at the origin) satisfying

  • There exists f:S1×B+f:S^{1}\times B\to\mathbb{R}^{+} such that f|S1×{0}Tf|_{S^{1}\times\{0\}}\equiv T, df|S1×{0}0df|_{S^{1}\times\{0\}}\equiv 0 and ψ(f(dθ+x1dx2))=λ\psi^{*}(f(d\theta+x_{1}dx_{2}))=\lambda, where θ\theta is the coordinate on S1S^{1} and (x1,x2)(x_{1},x_{2}) are coordinates on 2\mathbb{R}^{2};

  • ψ(xT(t))=(t,0,0)\psi(x_{T}(t))=(t,0,0).

The coordinates (θ,x1,x2)(\theta,x_{1},x_{2}) are called Martinet’s coordinates. The existence of such Martinet’s tubes is proved in [18] for any simple orbit P𝒫(λ)P\in\mathcal{P}(\lambda).

A more precise description of the asymptotic behavior of a finite energy surface is given in [18]. Let u~=(a,u):SΓ×M\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times M be a finite energy surface asymptotic to a nondegenerate orbit P=(x,T)P=(x,T) at the positive puncture z0Γz_{0}\in\Gamma and let (s,t)(s,t) be positive exponential coordinates near z0z_{0}. Let kk be a positive integer such that T=kTminT=kT_{min}, where TminT_{min} is the least positive period of xx and let (θ,z)=(θ,x1,x2)(\theta,z)=(\theta,x_{1},x_{2}) be Martinet’s coordinates in a neighborhood UMU\subset M of Pmin=(x,Tmin)P_{min}=(x,T_{min}). Then there exists s0+s_{0}\in\mathbb{R}^{+} such that, for (s,t)[s0,+)×(s,t)\in[s_{0},+\infty)\times\mathbb{R}, the map u(s,t)u(s,t) can be represented in Martinet’s coordinates by

u(s,t)=(θ(s,t),z(s,t))×2,u(s,t)=(\theta(s,t),z(s,t))\in\mathbb{R}\times\mathbb{R}^{2},

where θ\theta is seen as a map on the universal cover \mathbb{R} of S1=/S^{1}=\mathbb{R}/\mathbb{Z} satisfying θ(s,t+1)=θ(s,t)+k\theta(s,t+1)=\theta(s,t)+k and zz is 11-periodic in tt.

Theorem 2.7 ([18]).

If A(u~)>0A(\tilde{u})>0, there are constants Aij+A_{ij}\in\mathbb{R}^{+}, a0,θ0a_{0},\theta_{0}\in\mathbb{R}, a function R:×S12R:\mathbb{R}\times S^{1}\to\mathbb{R}^{2} and an eigensection η(t)\eta(t) of the asymptotic operator AP,JA_{P,J} (18), associated to a negative eigenvalue ασ(AP,J)\alpha\in\sigma(A_{P,J}), such that

(25) |sitj(a(s,t)(Ts+a0))|Aijer0s\displaystyle|\partial_{s}^{i}\partial_{t}^{j}(a(s,t)-(Ts+a_{0}))|\leq A_{ij}e^{-r_{0}s}
|sitj(θ(s,t)(kt+θ0)|Aijer0s\displaystyle|\partial_{s}^{i}\partial_{t}^{j}(\theta(s,t)-(kt+\theta_{0})|\leq A_{ij}e^{-r_{0}s}
z(s,t)=es0sα(r)𝑑r(e(t)+R(s,t))\displaystyle z(s,t)=e^{\int_{s_{0}}^{s}\alpha(r)dr}(e(t)+R(s,t))
|sitjR(s,t)|,|sitj(α(s)α)|Aijer0s\displaystyle|\partial_{s}^{i}\partial_{t}^{j}R(s,t)|,|\partial_{s}^{i}\partial_{t}^{j}(\alpha(s)-\alpha)|\leq A_{ij}e^{-r_{0}s}

for all large ss and i,ji,j\in\mathbb{N}. Here e:S12e:S^{1}\to\mathbb{R}^{2} represents the eigensection η(t)\eta(t) in the coordinates induced by ψ\psi, and α:[s0,)\alpha:[s_{0},\infty)\to\mathbb{R} is a smooth function such that α(s)α,as s\alpha(s)\to\alpha,\text{as }s\to\infty.

A similar statement holds if z0z_{0} is a negative puncture. In this case, we use negative exponential coordinates near z0z_{0}, er0se^{-r_{0}s} is replaced by er0se^{r_{0}s} and the eigenvalue α\alpha of AP,JA_{P,J} is positive.

The eigenvalue α\alpha and the eigensection η(t)\eta(t), as in Theorem 2.7, will be referred to as the asymptotic eigenvalue and asymptotic eigensection of u~\tilde{u} at the puncture z0z_{0}.

Remark 2.8.

Let u~=(a,u):SΓ×S3\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times S^{3} be a finite energy surface that is asymptotic to nondegenerate simple orbits at all of its punctures and such that πdu\pi\cdot du is not identically zero. By Theorem 2.7, the surface Σ=u(SΓ)¯\Sigma=\overline{u(S\setminus\Gamma)} is a strong transverse section according to Definition 1.1. Indeed, for every puncture zΓz\in\Gamma, the asymptotic eigensection η(t)\eta(t) of u~\tilde{u} at zz satisfies

dλ(η(t),Rλη(t))=1Tdλ(η(t),JAP,Jη(t))=1Tαdλ(η(t),Jη(t))0,t/.d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))=\frac{1}{T}d\lambda(\eta(t),JA_{P,J}\eta(t))=\frac{1}{T}\alpha d\lambda(\eta(t),J\eta(t))\neq 0,~{}\forall t\in\mathbb{R}/\mathbb{Z}.

2.6. Algebraic invariants

Let u~=(a,u):SΓ×M\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times M be a J~\tilde{J}-holomorphic finite energy surface. Assume that πdu\pi\cdot du is not identically zero and that u~\tilde{u} has nondegenerate asymptotic limits at all of its punctures. In [17], it is proved that the set where πdu\pi\cdot du vanishes is finite, and it is defined a local degree associated to each zero of πdu\pi\cdot du, which is always positive. The integer

(26) windπ(u~)0\operatorname{wind}_{\pi}(\tilde{u})\geq 0

is defined as the sum of such local degrees over all zeros of πdu\pi\cdot du.

Consider a unitary trivialization Ψ:(uξ,dλ,J)(SΓ)×2\Psi:(u^{*}\xi,d\lambda,J)\to(S\setminus\Gamma)\times\mathbb{R}^{2}. For zΓz\in\Gamma, fix positive cylindrical coordinates (s,t)(s,t) at zz and define

(27) wind(u~,z,Ψ)=lims+wind(tπsu(s,ϵzt),Ψ|u(s,ϵz)ξ)\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)=\lim_{s\to+\infty}\operatorname{wind}\left(t\mapsto\pi\cdot\partial_{s}u(s,\epsilon_{z}t),\Psi|_{u(s,\epsilon_{z}\cdot)^{*}\xi}\right)\in\mathbb{Z}

where ϵz\epsilon_{z} is the sign of the puncture zz. The winding number on the right is defined as in (19), and is independent of the choice of JJ and of the holomorphic chart. This limit is well defined since πsu(s,t)\pi\cdot\partial_{s}u(s,t) does not vanish for ss sufficiently large. The asymptotic winding number of u~\tilde{u} is defined by

(28) wind(u~)=zΓ+wind(u~,z,Ψ)zΓwind(u~,z,Ψ).\operatorname{wind}_{\infty}(\tilde{u})=\sum_{z\in\Gamma^{+}}\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)-\sum_{z\in\Gamma^{-}}\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)~{}.

It is proved in [17] that this sum does not depend on the chosen trivialization Ψ\Psi.

Remark 2.9.

Throughout the paper we will only deal with the tight contact structure on S3S^{3}, which is a trivial symplectic bundle. Consider a global symplectic trivialization Ψ:ξS3×2.\Psi:\xi\to S^{3}\times\mathbb{R}^{2}~{}. Then wind(u~,z,Ψ)\operatorname{wind}_{\infty}(\tilde{u},z,\Psi) is the winding number of the asymptotic eigensection given by Theorem 2.7, with respect to the trivialization Ψ\Psi. Moreover, wind(u~,z,Ψ)\operatorname{wind}_{\infty}(\tilde{u},z,\Psi) does not depend on the chosen global symplectic trivialization and we denote wind(u~,z,Ψ)\operatorname{wind}_{\infty}(\tilde{u},z,\Psi) by wind(u~,z)\operatorname{wind}_{\infty}(\tilde{u},z).

It is also proved in [17] that the invariants windπ(u~)\operatorname{wind}_{\pi}(\tilde{u}) and wind(u~)\operatorname{wind}_{\infty}(\tilde{u}) satisfy

(29) windπ(u~)=wind(u~)χ(S)+#Γ,\operatorname{wind}_{\pi}(\tilde{u})=\operatorname{wind}_{\infty}(\tilde{u})-\chi(S)+\#\Gamma,

where χ(S)\chi(S) is the Euler characteristic of SS.

The following lemma is a consequence of Proposition 2.2 and Remark 2.9.

Lemma 2.10.

Let u~:SΓ×M\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times M be a finite energy surface and let zΓz\in\Gamma be a puncture with asymptotic limit P=(x,T)P=(x,T). Assume that PP is nondegenerate and define νPpos\nu^{pos}_{P} and νPneg\nu^{neg}_{P} by (20) and (21) respectively. If πdu\pi\cdot du does not vanish identically, then

  • (1)

    wind(u~,z)wind(νPneg)\operatorname{wind}_{\infty}(\tilde{u},z)\leq\operatorname{wind}(\nu^{neg}_{P}) if zz is a positive puncture.

  • (2)

    wind(u~,z)wind(νPpos)\operatorname{wind}_{\infty}(\tilde{u},z)\geq\operatorname{wind}(\nu^{pos}_{P}) if zz is a negative puncture.

2.7. Fredholm theory

Let u~=(a,u):SΓ×M\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times M be a J~\tilde{J}-holomorphic finite energy surface and assume that u~\tilde{u} has nondegenerate asymptotic limits at all of its punctures. Let Ψ:(uξ,dλ)(SΓ)×2\Psi:(u^{*}\xi,d\lambda)\to(S\setminus\Gamma)\times\mathbb{R}^{2} be a symplectic trivialization. Fix a puncture zΓz\in\Gamma and let Pz=(x,T)P_{z}=(x,T) be the asymptotic limit of u~\tilde{u} at zz. The trivialization Ψ\Psi induces a homotopy class of oriented trivializations [Ψz][\Psi_{z}] of xTξx_{T}^{*}\xi. The Conley-Zehnder index of u~\tilde{u} is defined by

(30) μ(u~)=zΓ+μ(Pz,Ψz)zΓμ(Pz,Ψz).\mu(\tilde{u})=\sum_{z\in\Gamma^{+}}\mu(P_{z},\Psi_{z})-\sum_{z\in\Gamma^{-}}\mu(P_{z},\Psi_{z})~{}.

It is proved in [17] that this sum does not depend on the chosen trivialization Ψ\Psi.

An unparametrized finite energy surface is an equivalence class [(u~,(S,j),Γ)][(\tilde{u},(S,j),\Gamma)], where u~:(SΓ,j)(×M,J~)\tilde{u}:(S\setminus\Gamma,j)\to(\mathbb{R}\times M,\tilde{J}) is a J~\tilde{J}-holomorphic finite energy surface and Γ\Gamma is an ordered set. The equivalence class is defined as follows: (u~,(S,j),Γ)(\tilde{u}^{\prime},(S^{\prime},j^{\prime}),\Gamma^{\prime}) is equivalent to (u~,(S,j),Γ)(\tilde{u},(S,j),\Gamma) if there exists a biholomorphism ϕ:(S,j)(S,j)\phi:(S,j)\to(S^{\prime},j^{\prime}) such that u~=u~ϕ\tilde{u}^{\prime}=\tilde{u}\circ\phi, ϕ(Γ)=Γ\phi(\Gamma)=\Gamma^{\prime} and ϕ\phi preserves the ordering of Γ\Gamma and Γ\Gamma^{\prime}. To shorten notation we usually denote an unparametrized surface by [u~][\tilde{u}].

In [9] it is proved that the set of unparametrized finite energy surfaces in the neighborhood of [(u~,(S,j),Γ)][(\tilde{u},(S,j),\Gamma)], where u~\tilde{u} is a somewhere injective finite energy surface, is described by a nonlinear Fredholm equation having Fredholm index equal to

(31) ind(u~):=μ(u~)χ(S)+#Γ.\operatorname{ind}(\tilde{u}):=\mu(\tilde{u})-\chi(S)+\#\Gamma.

This generalizes the result for embedded finite energy surfaces proved in [20].

Theorem 2.11.

[9] There exists a dense subset 𝒥reg{J~|J𝒥(λ)}\mathcal{J}_{reg}\subset\{\tilde{J}|J\in\mathcal{J}(\lambda)\} such that, if u~=(a,u):SΓ×M\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times M is a somewhere injective finite energy surface which is pseudo-holomorphic with respect to J~𝒥reg{\tilde{J}}\in\mathcal{J}_{reg} and has nondegenerate asymptotic limits at all of its punctures, then

0ind(u~)=μ(u~)χ(S)+#Γ.0\leq\operatorname{ind}(\tilde{u})=\mu(\tilde{u})-\chi(S)+\#\Gamma~{}.

If πdu\pi\circ du is not identically zero, then

1ind(u~)=μ(u~)χ(S)+#Γ.1\leq\operatorname{ind}(\tilde{u})=\mu(\tilde{u})-\chi(S)+\#\Gamma~{}.

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 wind\operatorname{wind}_{\infty}, which differs from that of [31]. We assume that all the finite energy curves considered have nondegenerate asymptotic limits.

In the following, u~=(a,u):(SΓ,j)×M\tilde{u}=(a,u):(S\setminus\Gamma,j)\to\mathbb{R}\times M and v~=(b,v):(SΓ,j)×M\tilde{v}=(b,v):(S^{\prime}\setminus\Gamma^{\prime},j^{\prime})\to\mathbb{R}\times M are finite energy J~\tilde{J}-holomorphic curves and Ψ\Psi denotes a choice of trivialization of the contact structure along all simple periodic orbits with covers appearing as asymptotic limits of u~\tilde{u} or v~\tilde{v}.

Theorem 2.12.

[31, Theorem 2.4] Assume that no component of u~\tilde{u} or v~\tilde{v} lies in an orbit cylinder and that the projected curves uu and vv do not have identical image on any component of their domains. Then the following are equivalent:

  • (1)

    The projected curves uu and vv do not intersect;

  • (2)

    All of the following hold:

    • The map uu does not intersect any of the positive asymptotic limits of v~\tilde{v};

    • The map vv does not intersect any of the negative asymptotic limits of u~\tilde{u};

    • If PP is a periodic orbit so that, at zΓz\in\Gamma, u~\tilde{u} is asymptotic PmzP^{m_{z}} and, at wΓw\in\Gamma^{\prime}, v~\tilde{v} is asymptotic to PmwP^{m_{w}}, then: If zz and ww are both positive or both negative punctures, we have wind(u~,z,Ψ)mzwind(v~,w,Ψ)mw.\frac{\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)}{m_{z}}\geq\frac{\operatorname{wind}_{\infty}(\tilde{v},w,\Psi)}{m_{w}}~{}. If zz is a negative puncture and ww is a positive puncture, we have

      wind(u~,z,Ψ)mz=μ(Pmz,Ψ)/2mz=μ(Pmw,Ψ)/2mw=wind(v~,w,Ψ)mw\frac{\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)}{m_{z}}=\frac{\lfloor\mu(P^{m_{z}},\Psi)/2\rfloor}{m_{z}}=\frac{-\lfloor-\mu(P^{m_{w}},\Psi)/2\rfloor}{m_{w}}=\frac{\operatorname{wind}_{\infty}(\tilde{v},w,\Psi)}{m_{w}}
  • (3)

    All of the following hold:

    • The map uu does not intersect any of the asymptotic limits of v~\tilde{v}.

    • The map vv does not intersect any of the asymptotic limits of u~\tilde{u}.

    • If PP is a periodic orbit so that, at zΓz\in\Gamma, u~\tilde{u} is asymptotic to PmzP^{m_{z}} and, at wΓw\in\Gamma^{\prime}, v~\tilde{v} is asymptotic to PmwP^{m_{w}}, then wind(u~,z,Ψ)mz=wind(v~,w,Ψ)mw\frac{\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)}{m_{z}}=\frac{\operatorname{wind}_{\infty}(\tilde{v},w,\Psi)}{m_{w}}.

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 u~\tilde{u} is somewhere injective, connected and does not have image contained in an orbit cylinder. Assume further that u~\tilde{u} is not asymptotic to a covering of the same simple orbit at two distinct punctures. Then the following are equivalent:

  1. (1)

    The projected map u:SΓMu:S\setminus\Gamma\to M is an embedding.

  2. (2)

    uu does not intersect any of its asymptotic limits.

  3. (3)

    All of the following hold:

    • The map u~\tilde{u} is an embedding.

    • The projected map uu is an immersion which is everywhere transverse to the Reeb vector field.

    • For each zΓz\in\Gamma, we have gcd(mz,wind(u~,z,Ψ))=1\gcd(m_{z},\operatorname{wind}_{\infty}(\tilde{u},z,\Psi))=1.

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 u~\tilde{u} and v~\tilde{v} are asymptotic to the same Reeb orbit PP at certain punctures zΓz\in\Gamma and wΓw\in\Gamma^{\prime} with the same sign. We say that u~\tilde{u} and v~\tilde{v} approach PP in the same direction at these punctures if ηu=cηv\eta_{u}=c\eta_{v} for some c>0c>0 , where ηu\eta_{u} and ηv\eta_{v} are the asymptotic eigensections of u~\tilde{u} at zz and of v~\tilde{v} at ww respectively. In case ηu=cηv\eta_{u}=c\eta_{v} for c<0c<0, we say that u~\tilde{u} and v~\tilde{v} approach PP 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 u~\tilde{u} and v~\tilde{v} are asymptotic to an even Reeb orbit PP at punctures zΓz\in\Gamma and wΓw\in\Gamma^{\prime} with the same sign. Assume that u~\tilde{u} and v~\tilde{v} have extremal winding number, that is,

wind(u~,z,Ψ)=wind(v~,w,Ψ)=μ(P,Ψ)2\operatorname{wind}_{\infty}(\tilde{u},z,\Psi)=\operatorname{wind}_{\infty}(\tilde{v},w,\Psi)=\dfrac{\mu(P,\Psi)}{2}

and that u~\tilde{u} and v~\tilde{v} approach PP in the same direction. Then the projections uu and vv 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 λ\lambda be a tight contact form on S3S^{3} satisfying the hypotheses of Theorem 1.3. Then there exists a family of J~\tilde{J}-holomorphic planes {u~τ=(aτ,uτ):×S3}τ(0,1)\{\tilde{u}_{\tau}=(a_{\tau},u_{\tau}):\mathbb{C}\to\mathbb{R}\times S^{3}\}_{\tau\in(0,1)}, all of them asymptotic to P3P_{3}, a pair of J~\tilde{J}-holomorphic cylinders u~r=(ar,ur),u~r=(ar,vr):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}),\tilde{u}_{r}^{\prime}=(a_{r}^{\prime},v_{r}^{\prime}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}, both asymptotic to P3P_{3} at their positive punctures z=+z=+\infty and P2P_{2} at their negative punctures z=0z=0, and a finite energy plane u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3} asymptotic to P2P_{2}. The projections uτ,ur,ur,uqu_{\tau},u_{r},u_{r}^{\prime},u_{q} are embeddings transverse to the Reeb vector field. Define Fτ=uτ()F_{\tau}=u_{\tau}(\mathbb{C}), U1=ur({0})U_{1}=u_{r}(\mathbb{C}\setminus\{0\}), U2=ur({0})U_{2}=u_{r}^{\prime}(\mathbb{C}\setminus\{0\}) and D=uq()D=u_{q}(\mathbb{C}). Then the surface T:=P2P3U1U2T:=P_{2}\cup P_{3}\cup U_{1}\cup U_{2} is homeomorphic to a torus and TP3T\setminus P_{3} is C1C^{1}-embedded. The union of the family {Fτ}τ(0,1)\{F_{\tau}\}_{\tau\in(0,1)} with U1U_{1}, U2U_{2} and DD determine a smooth foliation of 1(P2P3)\mathcal{R}_{1}\setminus(P_{2}\cup P_{3}), where 1S3\mathcal{R}_{1}\subset S^{3} is a closed region with boundary TT and homeomorphic to a solid torus.

By hypothesis, there exists a finite energy J~\tilde{J}-holomorphic plane

(32) u~=(a,u):×S3\tilde{u}=(a,u):\mathbb{C}\to\mathbb{R}\times S^{3}

asymptotic to the orbit P3P_{3}. Both the J~\tilde{J}-holomorphic plane u~\tilde{u} and the projection u:S3u:\mathbb{C}\to S^{3} 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 S3S^{3} equipped with a tight contact form λ\lambda. Assume u~=(a,u):×S3\tilde{u}=(a,u):\mathbb{C}\to\mathbb{R}\times S^{3} is a finite energy plane asymptotic to an unknotted orbit PP. If μ(P)3\mu(P)\leq 3, then u()P=u(\mathbb{C})\cap P=\emptyset and u:S3Pu:\mathbb{C}\to S^{3}\setminus P is an embedding.

The finite energy plane u~\tilde{u} is automatically Fredholm regular. If J𝒥regJ^{\prime}\in\mathcal{J}_{reg} is a CC^{\infty}-small perturbation of JJ, where 𝒥reg\mathcal{J}_{reg} is the dense set obtained by Theorem 2.11, then we can find a J~\tilde{J}^{\prime}-holomorphic plane u~\tilde{u}^{\prime} asymptotic to P3P_{3} as a CC^{\infty} small perturbation of u~\tilde{u}. It follows that we can revert the notation back to u~\tilde{u} and J~\tilde{J} and assume that J~𝒥reg\tilde{J}\in\mathcal{J}_{reg}. 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 𝑷𝟑\boldsymbol{P_{3}}

The following is a particular case of Theorem 4.5.44 of [32]. It generalizes Theorem 1.5 of [20].

Theorem 3.3 ([32]).

Let λ\lambda be a contact form on a closed 33-manifold MM and let J𝒥(ξ=kerλ,dλ)J\in\mathcal{J}(\xi=\ker\lambda,d\lambda). Let u~=(a,u):S2Γ×M\tilde{u}=(a,u):S^{2}\setminus\Gamma\to\mathbb{R}\times M is an embedded J~\tilde{J}-holomorphic finite energy sphere such that every asymptotic limit is nondegenerate, simple and has odd Conley-Zehnder index. Suppose that ind(u~)=2\operatorname{ind}(\tilde{u})=2. Then there exists a number δ>0\delta>0 and an embedding

F~:×(δ,δ)×S2Γ\displaystyle\tilde{F}:\mathbb{R}\times(-\delta,\delta)\times S^{2}\setminus\Gamma ×M\displaystyle\to\mathbb{R}\times M
(σ,τ,z)\displaystyle(\sigma,\tau,z) (aτ(z)+σ,uτ(z))\displaystyle\mapsto(a_{\tau}(z)+\sigma,u_{\tau}(z))

such that

  • For σ\sigma\in\mathbb{R} and τ(δ,δ)\tau\in(-\delta,\delta), the maps u~(σ,τ):=F~(σ,τ,)\tilde{u}_{(\sigma,\tau)}:=\tilde{F}(\sigma,\tau,\cdot) are (up to parametrization) embedded J~\tilde{J}-holomorphic finite energy spheres and u~(0,0)=u~\tilde{u}_{(0,0)}=\tilde{u}.

  • The map F(τ,z)=uτ(z)F(\tau,z)=u_{\tau}(z) is an embedding (δ,δ)×S2ΓM(-\delta,\delta)\times S^{2}\setminus\Gamma\to M and its image never intersects the asymptotic limits. In particular, the maps uτ:S2ΓMu_{\tau}:S^{2}\setminus\Gamma\to M are embeddings for each τ(δ,δ)\tau\in(-\delta,\delta), with mutually disjoint images which do not intersect their asymptotic limits.

  • For any sequence v~k:S2Γ×M\tilde{v}_{k}:S^{2}\setminus\Gamma\to\mathbb{R}\times M such that for each puncture in Γ\Gamma, v~k\tilde{v}_{k} has the same asymptotic limit as u~\tilde{u}, with the same sign, and v~ku~\tilde{v}_{k}\to\tilde{u} in Cloc(S2Γ)C^{\infty}_{loc}(S^{2}\setminus\Gamma), there is a sequence (σk,τk)(0,0)×(δ,δ)(\sigma_{k},\tau_{k})\to(0,0)\in\mathbb{R}\times(-\delta,\delta) such that v~k=u~(σk,τk)φk\tilde{v}_{k}=\tilde{u}_{(\sigma_{k},\tau_{k})}\circ\varphi_{k} for some sequence of biholomorphisms φk:S2S2\varphi_{k}:S^{2}\to S^{2} and kk sufficiently large.

Remark 3.4.

In Theorem 3.3, we do not require J~𝒥reg\tilde{J}\in\mathcal{J}_{reg}.

Applying Theorem 3.3 to the finite energy plane (32) we obtain a maximal one-parameter family of finite energy planes

(33) u~τ=(aτ,uτ):×S3,τ(τ,τ+)\tilde{u}_{\tau}=(a_{\tau},u_{\tau}):\mathbb{C}\to\mathbb{R}\times S^{3},~{}~{}\tau\in(\tau_{-},\tau_{+})

asymptotic to the orbit P3P_{3}. The family (33) satisfies uτ1()uτ2()=,τ1τ2.u_{\tau_{1}}(\mathbb{C})\cap u_{\tau_{2}}(\mathbb{C})=\emptyset,~{}\forall\tau_{1}\neq\tau_{2}. Indeed, suppose that there exist τ1τ2\tau_{1}\neq\tau_{2} such that uτ1()uτ2()u_{\tau_{1}}(\mathbb{C})\cap u_{\tau_{2}}(\mathbb{C})\neq\emptyset. Then uτ1()=uτ2()u_{\tau_{1}}(\mathbb{C})=u_{\tau_{2}}(\mathbb{C}) (see Theorem 2.12 or [17, Theorem 1.4]) and we would obtain an S1S^{1}-family of embedded planes that provides an open book decomposition of S3S^{3}, with binding P3P_{3} and disk-like pages. It follows from equation (29), Lemma 2.10 and formula (22) that windπ(u~τ)=0,τ(τ,τ+).\operatorname{wind}_{\pi}(\tilde{u}_{\tau})=0,~{}\forall\tau\in(\tau_{-},\tau_{+}). As a consequence of the definition of windπ\operatorname{wind}_{\pi}, we conclude that uτu_{\tau} is transverse to RλR_{\lambda}, τ(τ,τ+)\forall\tau\in(\tau_{-},\tau_{+}). This implies that every Reeb orbit in S3P3S^{3}\setminus P_{3} is linked to P3P_{3}, which contradicts the existence of P1P_{1} and P2P_{2}.

Now we describe how the family {u~τ}\{\tilde{u}_{\tau}\} breaks as ττ±\tau\to\tau_{\pm}. We assume that τ=0\tau_{-}=0 and τ+=1\tau_{+}=1 and that τ\tau strictly increases in the direction of RλR_{\lambda}.

Proposition 3.5.

Consider a sequence u~n:=u~τn\tilde{u}_{n}:=\tilde{u}_{\tau_{n}} in the family (33) satisfying τn0+\tau_{n}\to 0^{+}. Then there exists a J~\tilde{J}-holomorphic finite energy cylinder u~r:{0}×S3\tilde{u}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}, which is asymptotic to P3P_{3} at the positive puncture z=z=\infty and to P2P_{2} at the negative puncture z=0z=0, and a finite energy J~\tilde{J}-holomorphic plane u~q:×S3\tilde{u}_{q}:\mathbb{C}\to\mathbb{R}\times S^{3} asymptotic to P2P_{2} at z=z=\infty, such that, after suitable reparametrizations and \mathbb{R}-translations of u~n\tilde{u}_{n}, the following hold

  1. (i)

    up to a subsequence, u~nu~r\tilde{u}_{n}\to\tilde{u}_{r} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  2. (ii)

    There exist sequences δn0+\delta_{n}\to 0^{+}, znz_{n}\in\mathbb{C} and cnc_{n}\in\mathbb{R} such that, up to a subsequence, u~n(zn+δn)+cnu~q\tilde{u}_{n}(z_{n}+\delta_{n}\cdot)+c_{n}\to\tilde{u}_{q} in Cloc()C^{\infty}_{loc}(\mathbb{C}) as nn\to\infty.

Here (a,x)+c:=(a+c,x),(a,x)×S3,c(a,x)+c:=(a+c,x),~{}\forall(a,x)\in\mathbb{R}\times S^{3},~{}c\in\mathbb{R}. A similar statement holds for any sequence u~τn\tilde{u}_{\tau_{n}} satisfying τn1\tau_{n}\to 1^{-}. In this case we change the notation from u~r\tilde{u}_{r} and u~q\tilde{u}_{q} to u~r\tilde{u}_{r}^{\prime} and u~q\tilde{u}_{q}^{\prime} respectively.

The proof of proposition 3.5 is left to Subsection 3.3. Subsection 3.2 below consists of preliminary results.

3.2. Bubbling-off analysis

By hypotheses, every orbit in 𝒫(λ)\mathcal{P}(\lambda) having period T3\leq T_{3} is nondegenerate. Consequently, there exists only a finite number of such orbits. Define σ(T3)\sigma(T_{3}) as any real number satisfying

(34) 0<σ(T3)<min{T,|TT|:TT periods ,T,TT3}.0<\sigma(T_{3})<\min\{T,|T-T^{\prime}|:T\neq T^{\prime}\text{ periods },T,T^{\prime}\leq T_{3}\}~{}.

Most of the material in §3.2.1 and §3.2.2 is adapted from [24].

3.2.1. Germinating sequences

Now we fix J𝒥(λ)J\in\mathcal{J}(\lambda) and consider a sequence of J~\tilde{J}-holomorphic curves v~n=(bn,vn):BRn(0)×S3\tilde{v}_{n}=(b_{n},v_{n}):B_{R_{n}}(0)\subset\mathbb{C}\to\mathbb{R}\times S^{3} satisfying

(35) Rn,Rn(0,+]\displaystyle R_{n}\to\infty,~{}~{}R_{n}\in(0,+\infty]
(36) E(v~n)T3,n\displaystyle E(\tilde{v}_{n})\leq T_{3},~{}\forall n
(37) BRn(0)𝔻vn𝑑λσ(T3),n\displaystyle\int_{B_{R_{n}}(0)\setminus\mathbb{D}}v_{n}^{*}d\lambda\leq\sigma(T_{3}),~{}\forall n
(38) {bn(2)} is uniformily bounded.\displaystyle\{b_{n}(2)\}\text{ is uniformily bounded}.
Definition 3.6.

A sequence v~n\tilde{v}_{n} of J~\tilde{J}-holomorphic curves satisfying (35)-(38) is called a germinating sequence.

Proposition 3.7.

There exists a finite set Γ𝔻\Gamma\subset\mathbb{D}, a J~\tilde{J}-holomorphic map v~=(b,v):Γ×S3\tilde{v}=(b,v):\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} and a subsequence of v~n\tilde{v}_{n}, still denoted by v~n\tilde{v}_{n}, such that v~nv~in Cloc(Γ,×S3)\tilde{v}_{n}\to\tilde{v}~{}~{}\text{in }C^{\infty}_{loc}(\mathbb{C}\setminus\Gamma,\mathbb{R}\times S^{3}) and E(v~)T3E(\tilde{v})\leq T_{3}.

Proof.

Let Γ0\Gamma_{0}\subset\mathbb{C} be the set of points zz\in\mathbb{C} such that there exists a subsequence v~nj\tilde{v}_{n_{j}} and a sequence zjBRnj(0)z_{j}\in B_{R_{n_{j}}}(0) with zjzz_{j}\to z and |dv~nj(zj)|,j.|d\tilde{v}_{n_{j}}(z_{j})|\to\infty,~{}j\to\infty~{}. If Γ0=\Gamma_{0}=\emptyset, then by (38) and usual elliptic estimates, see [28, Chapter 4], we find a J~\tilde{J}-holomorphic map v~:×S3\tilde{v}:\mathbb{C}\to\mathbb{R}\times S^{3} such that, up to a subsequence, v~nv~\tilde{v}_{n}\to\tilde{v} in Cloc(,×S3)C^{\infty}_{loc}(\mathbb{C},\mathbb{R}\times S^{3}). In this case, Γ=\Gamma=\emptyset.

Now assume Γ0\Gamma_{0}\neq\emptyset and let z0Γ0z_{0}\in\Gamma_{0}. It follows from results in [14, §3.2] that there exists a period 0<T0T30<T_{0}\leq T_{3} and sequences rj00+r_{j}^{0}\to 0^{+}, nj0n_{j}^{0}\to\infty and zj0z0z_{j}^{0}\to z_{0} such that

limjBrj(zj0)vnj0𝑑λT0.\lim_{j\to\infty}\int_{B_{r_{j}}(z_{j}^{0})}v_{n_{j}^{0}}^{*}d\lambda\geq T_{0}~{}.

Consider v~nj0\tilde{v}_{n^{0}_{j}} as the new sequence v~n\tilde{v}_{n}. Now let Γ1{z0}\Gamma_{1}\subset\mathbb{C}\setminus\{z_{0}\} be the set of points z1z0z_{1}\neq z_{0} such that there exists a subsequence v~nj\tilde{v}_{n_{j}} and sequence zjBRnj(0)z_{j}\in B_{R_{n_{j}}}(0) with zz1z\to z_{1} and |dv~nj(zj)||d\tilde{v}_{n_{j}}(z_{j})|\to\infty. As before, if Γ1=\Gamma_{1}=\emptyset, we have a J~\tilde{J}-holomorphic map v~:{z0}×S3\tilde{v}:\mathbb{C}\setminus\{z_{0}\}\to\mathbb{R}\times S^{3} such that, up to a subsequence, v~nv~\tilde{v}_{n}\to\tilde{v} in Cloc({z0},×S3)C^{\infty}_{loc}(\mathbb{C}\setminus\{z_{0}\},\mathbb{R}\times S^{3}). In this case, we define Γ=Γ0={z0}\Gamma=\Gamma_{0}=\{z_{0}\}. If Γ1\Gamma_{1}\neq\emptyset and z1Γ1z_{1}\in\Gamma_{1}, there exist a period 0<T1T30<T_{1}\leq T_{3} and sequences rj10r_{j}^{1}\to 0, nj1n_{j}^{1}\to\infty and zj1z1z_{j}^{1}\to z_{1} such that

limjBrj(zj1)vnj1𝑑λT1.\lim_{j\to\infty}\int_{B_{r_{j}}(z_{j}^{1})}v_{n^{1}_{j}}^{*}d\lambda\geq T_{1}~{}.

Considering v~nj1\tilde{v}_{n_{j}^{1}} as the new sequence v~n\tilde{v}_{n}, define Γ2{z0,z1}\Gamma_{2}\subset\mathbb{C}\setminus\{z_{0},z_{1}\} as before. Repeating this argument, let ziΓi{z0,,zi1}z_{i}\in\Gamma_{i}\subset\mathbb{C}\setminus\{z_{0},\dots,z_{i-1}\}. Note that

C\displaystyle C limnE(v~n)limnvn𝑑λ\displaystyle\geq\lim_{n\to\infty}E(\tilde{v}_{n})\geq\lim_{n\to\infty}\int_{\mathbb{C}}v_{n}^{*}d\lambda l=0ilimjBrjl(zjl)vnjl𝑑λT0++Ti.\displaystyle\geq\sum_{l=0}^{i}\lim_{j\to\infty}\int_{B_{r_{j}^{l}}(z_{j}^{l})}v_{n_{j}^{l}}^{*}d\lambda\geq T_{0}+\dots+T_{i}.

It follows that there exists i0i_{0} such that Γi00\Gamma_{i_{0}}\neq 0 and Γi=\Gamma_{i}=\emptyset for i>i0i>i_{0}. We end up with a finite set Γ={z0,,zi0}\Gamma=\{z_{0},\dots,z_{i_{0}}\} and a J~\tilde{J}-holomorphic map v~:Γ×S3\tilde{v}:\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} such that, up to a subsequence, v~nv~ in Cloc.\tilde{v}_{n}\to\tilde{v}\text{ in }C^{\infty}_{loc}~{}.

It follows from (37) that Γ𝔻\Gamma\subset\mathbb{D}. The inequality E(v~)T3E(\tilde{v})\leq T_{3} follows from (36) and Fatou’s Lemma. ∎

Definition 3.8.

A J~\tilde{J}-holomorphic map v~:Γ×S3\tilde{v}:\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} as in Proposition 3.7 is called a limit of the germinating sequence v~n\tilde{v}_{n}.

If Γ\Gamma\neq\emptyset, then v~\tilde{v} is nonconstant. In this case, all the punctures z=ziΓz=z_{i}\in\Gamma are negative and \infty is a positive puncture. To prove this, define, for any ϵ>0\epsilon>0,

(39) mϵ(z):=Bϵ(z)vλ,m_{\epsilon}(z):=\int_{\partial B_{\epsilon}(z)}v^{*}\lambda~{},

where Bϵ(z)\partial B_{\epsilon}(z) is oriented counterclockwise. Then

mϵ(z)=Bϵ(z)vλ=limnBϵ(z)vnλ=limnBϵ(z)vn𝑑λ.m_{\epsilon}(z)=\int_{\partial B_{\epsilon}(z)}v^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(z)}v_{n}^{*}\lambda=\lim_{n\to\infty}\int_{B_{\epsilon}(z)}v_{n}^{*}d\lambda~{}.

For jj large, Brji(zji)B_{r_{j}^{i}}(z_{j}^{i}), defined as in the proof of Proposition 3.7, is contained in Bϵ(z)B_{\epsilon}(z). It follows that

mϵ(z)=limnBϵ(z)vn𝑑λlimjBrj(zji)vnji𝑑λTi>0.m_{\epsilon}(z)=\lim_{n\to\infty}\int_{B_{\epsilon}(z)}v_{n}^{*}d\lambda\geq\lim_{j\to\infty}\int_{B_{r_{j}}(z_{j}^{i})}v_{n_{j}^{i}}^{*}d\lambda\geq T_{i}>0~{}.

This implies that v~\tilde{v} is nonconstant and the puncture zz is negative. Moreover, as a consequence of 0<E(v~)<0<E(\tilde{v})<\infty, we know that v~\tilde{v} has at least one positive puncture. Thus, \infty is a positive puncture.

3.2.2. Soft-rescaling near a negative puncture

Assume Γ\Gamma\neq\emptyset and let v~=(b,v):Γ×M\tilde{v}=(b,v):\mathbb{C}\setminus\Gamma\to\mathbb{R}\times M be a limit of the germinating sequence v~n=(bn,vn)\tilde{v}_{n}=(b_{n},v_{n}). Let zΓz\in\Gamma. We define the mass m(z)m(z) of zz by

(40) m(z)=limϵ0+mϵ(z)=limϵ0+Bϵ(z)vλ=Tz>σ(T3)>0,m(z)=\lim_{\epsilon\to 0^{+}}m_{\epsilon}(z)=\lim_{\epsilon\to 0^{+}}\int_{\partial B_{\epsilon}(z)}v^{*}\lambda=T_{z}>\sigma(T_{3})>0~{},

where TzT_{z} is the period of the asymptotic limit of v~\tilde{v} at zz. Since mϵ(z)m_{\epsilon}(z) is a nondecreasing function of ϵ\epsilon, we can fix ϵ\epsilon small enough so that

(41) 0mϵ(z)m(z)σ(T3)2.0\leq m_{\epsilon}(z)-m(z)\leq\frac{\sigma(T_{3})}{2}.

Choose sequences znBϵ(z)¯z_{n}\in\overline{B_{\epsilon}(z)} and 0<δn<ϵ,n0<\delta_{n}<\epsilon,~{}\forall n, so that

(42) bn(zn)bn(ζ),ζBϵ(z),\displaystyle b_{n}(z_{n})\leq b_{n}(\zeta),~{}\forall\zeta\in B_{\epsilon}(z),
(43) Bϵ(z)Bδn(zn)vn𝑑λ=σ(T3).\displaystyle\int_{B_{\epsilon}(z)\setminus B_{\delta_{n}}(z_{n})}v_{n}^{*}d\lambda=\sigma(T_{3}).

Since zz is a negative puncture, (42) implies that znzz_{n}\to z. Hence the existence of δn\delta_{n} as in (43) follows from (40). We claim that lim infδn=0\liminf\delta_{n}=0. Otherwise, we choose 0<ϵ<lim infδnϵ0<\epsilon^{\prime}<\liminf\delta_{n}\leq\epsilon. From (41), we get the contradiction

σ(T3)2\displaystyle\frac{\sigma(T_{3})}{2} mϵ(z)m(z)mϵ(z)mϵ(z)\displaystyle\geq m_{\epsilon}(z)-m(z)\geq m_{\epsilon}(z)-m_{\epsilon^{\prime}}(z)
=limnBϵ(z)Bϵ(z)vn𝑑λ\displaystyle=\lim_{n\to\infty}\int_{B_{\epsilon}(z)\setminus B_{\epsilon^{\prime}}(z)}v_{n}^{*}d\lambda
limnBϵ(z)Bδn(zn)vn𝑑λ=σ(T3).\displaystyle\geq\lim_{n\to\infty}\int_{B_{\epsilon}(z)\setminus B_{\delta_{n}}(z_{n})}v_{n}^{*}d\lambda=\sigma(T_{3})~{}.

Thus, we can assume that δn0\delta_{n}\to 0.

Now take any sequence Rn+R_{n}\to+\infty satisfying δnRn<ϵ2\delta_{n}R_{n}<\frac{\epsilon}{2}~{} and define the sequence of J~\tilde{J}-holomorphic maps w~n=(cn,wn):BRn(0)×S3\tilde{w}_{n}=(c_{n},w_{n}):B_{R_{n}}(0)\to\mathbb{R}\times S^{3} by

(44) w~n(ζ)=(bn(zn+δnζ)bn(zn+2δn),vn(zn+δnζ)).\tilde{w}_{n}(\zeta)=(b_{n}(z_{n}+\delta_{n}\zeta)-b_{n}(z_{n}+2\delta_{n}),v_{n}(z_{n}+\delta_{n}\zeta))~{}.

It follows from (43) that

BRn(0)𝔻wn𝑑λσ(T3),n.\int_{B_{R_{n}}(0)\setminus\mathbb{D}}w_{n}^{*}d\lambda\leq\sigma(T_{3}),~{}\forall n~{}.

Moreover, by the definition of w~n\tilde{w}_{n}, E(w~n)E(v~n)T3E(\tilde{w}_{n})\leq E(\tilde{v}_{n})\leq T_{3} and w~n(2){0}×M\tilde{w}_{n}(2)\in\{0\}\times M. Thus, w~n\tilde{w}_{n} is a germinating sequence.

Let w~=(c,w):Γ×S3\tilde{w}=(c,w):\mathbb{C}\setminus\Gamma^{\prime}\to\mathbb{R}\times S^{3} be a limit of w~n\tilde{w}_{n}, as in Proposition 3.7. If Γ\Gamma^{\prime}\neq\emptyset, then w~\tilde{w} is not constant. If Γ=\Gamma^{\prime}=\emptyset, then

(45) 𝔻w𝑑λ\displaystyle\int_{\mathbb{D}}w^{*}d\lambda =limn𝔻wn𝑑λ=limnBδn(zn)vn𝑑λ\displaystyle=\lim_{n\to\infty}\int_{\mathbb{D}}w_{n}^{*}d\lambda=\lim_{n\to\infty}\int_{B_{\delta_{n}}(z_{n})}v_{n}^{*}d\lambda
=limn(Bϵ(z)vn𝑑λBϵ(z)Bδn(zn)vn𝑑λ)\displaystyle=\lim_{n\to\infty}\left(\int_{B_{\epsilon}(z)}v_{n}^{*}d\lambda-\int_{B_{\epsilon}(z)\setminus B_{\delta_{n}}(z_{n})}v_{n}^{*}d\lambda\right)
=mϵ(z)σ(T3)\displaystyle=m_{\epsilon}(z)-\sigma(T_{3})
Tzσ(T3)>0.\displaystyle\geq T_{z}-\sigma(T_{3})>0~{}.

Thus w~\tilde{w} is nonconstant as well. From Fatou’s Lemma we get 0<E(w~)T30<E(\tilde{w})\leq T_{3}. This also implies that the periods of the asymptotic limits of w~\tilde{w} are bounded by T3T_{3}.

Proposition 3.9.

The asymptotic limit PP_{\infty} of w~\tilde{w} at \infty coincides with the asymptotic limit PzP_{z} of v~\tilde{v} at the negative puncture zΓz\in\Gamma.

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 e>0e>0 and let σ(T3)\sigma(T_{3}) be defined by (34). Identifying S1=/S^{1}=\mathbb{R}/\mathbb{Z}, let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be an open neighborhood of the set of periodic orbits P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) with TT3T\leq T_{3}, viewed as maps xT:S1S3x_{T}:S^{1}\to S^{3}, xT(t)=x(Tt)x_{T}(t)=x(Tt). We assume that 𝒲\mathcal{W} is S1S^{1}-invariant, meaning that y(+c)𝒲y𝒲,cS1y(\cdot+c)\in\mathcal{W}\Leftrightarrow y\in\mathcal{W},\forall c\in S^{1}, and that each of the connected components of 𝒲\mathcal{W} contains at most one periodic orbit modulo S1S^{1}-reparametrizations. Then there exists a constant h>0h>0 such that the following holds. If u~=(a,u):[r,R]×S1×S3\tilde{u}=(a,u):[r,R]\times S^{1}\to\mathbb{R}\times S^{3} is a J~\tilde{J}-holomorphic cylinder satisfying

E(u~)T3,[r,R]×S1u𝑑λσ(T3),{r}×S1uλeandr+hRh,E(\tilde{u})\leq T_{3},~{}~{}~{}\int_{[r,R]\times S^{1}}u^{*}d\lambda\leq\sigma(T_{3}),~{}~{}~{}~{}\int_{\{r\}\times S^{1}}u^{*}\lambda\geq e~{}~{}~{}~{}\text{and}~{}~{}r+h\leq R-h,

then each loop tS1u(s,t)t\in S^{1}\to u(s,t) is contained in 𝒲\mathcal{W} for all s[r+h,Rh]s\in[r+h,R-h].

Proof of Proposition 3.9.

Let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be as in the statement of Lemma 3.10. Let 𝒲\mathcal{W}_{\infty} and 𝒲z\mathcal{W}_{z} be connected components of 𝒲\mathcal{W} containing PP_{\infty} and PzP_{z} respectively. Since v~nv~\tilde{v}_{n}\to\tilde{v}, we can choose 0<ϵ0<ϵ0<\epsilon_{0}<\epsilon small enough so that, if 0<ρϵ00<\rho\leq\epsilon_{0} is fixed, then the loop

tS1vn(zn+ρei2πt)t\in S^{1}\mapsto v_{n}(z_{n}+\rho e^{i2\pi t})

belongs to 𝒲z\mathcal{W}_{z} for large nn. Since w~nw~\tilde{w}_{n}\to\tilde{w}, we can choose R0>1R_{0}>1 large enough so that, if RR0R\geq R_{0} is fixed, then the loop

tS1wn(Rei2πt)=vn(zn+δnRei2πt)t\in S^{1}\mapsto w_{n}(Re^{i2\pi t})=v_{n}(z_{n}+\delta_{n}Re^{i2\pi t})

belongs to 𝒲\mathcal{W}_{\infty} for large nn. By (40) and (43), we can show that

(46) e:=lim infBδnR0(zn)vnλ>0.e:=\liminf\int_{\partial B_{\delta_{n}R_{0}(z_{n})}}v_{n}^{*}\lambda>0.

Consider, for each nn, the J~\tilde{J}-holomorphic cylinder C~n:[lnRoδn2π,lnϵ02π]×S1×S3\tilde{C}_{n}:\left[\frac{\ln R_{o}\delta_{n}}{2\pi},\frac{\ln\epsilon_{0}}{2\pi}\right]\times S^{1}\to\mathbb{R}\times S^{3}, defined by C~n(s,t)=v~n(zn+e2π(s+it))\tilde{C}_{n}(s,t)=\tilde{v}_{n}(z_{n}+e^{2\pi(s+it)}). It follows from (43) that

(47) [lnRoδn2π,lnϵ02π]×S1Cn𝑑λσ(T3)\int_{\left[\frac{\ln R_{o}\delta_{n}}{2\pi},\frac{\ln\epsilon_{0}}{2\pi}\right]\times S^{1}}C_{n}^{*}d\lambda\leq\sigma(T_{3})

for large nn. Using (46) and (47) and applying Lemma 3.10, we find h>0h>0 so that the loop

tCn(s,t)t\mapsto C_{n}(s,t)

is contained in 𝒲\mathcal{W} for all s[lnR0δn2π+h,lnϵ02πh]s\in\left[\frac{\ln R_{0}\delta_{n}}{2\pi}+h,\frac{\ln\epsilon_{0}}{2\pi}-h\right] and large nn. But

Cn(lnϵ02πh,t)=vn(zn+ϵ0e2πhe2πit)𝒲z, for all n large, andC_{n}\left(\frac{\ln\epsilon_{0}}{2\pi}-h,t\right)=v_{n}(z_{n}+\epsilon_{0}e^{-2\pi h}e^{2\pi it})\in\mathcal{W}_{z},~{}\text{ for all }n\text{ large, and}
Cn(lnR0δn2π+h,t)=vn(zn+R0δne2πhe2πt)𝒲,for all n large.C_{n}\left(\frac{\ln R_{0}\delta_{n}}{2\pi}+h,t\right)=v_{n}(z_{n}+R_{0}\delta_{n}e^{2\pi h}e^{2\pi t})\in\mathcal{W}_{\infty},~{}\text{for all }n\text{ large}.

Thus, 𝒲=𝒲z\mathcal{W}_{\infty}=\mathcal{W}_{z} and P=PzP_{\infty}=P_{z}. ∎

Proposition 3.11.

Either

  • Γw𝑑λ>0\int_{\mathbb{C}\setminus\Gamma^{\prime}}w^{*}d\lambda>0 or

  • Γw𝑑λ=0\int_{\mathbb{C}\setminus\Gamma^{\prime}}w^{*}d\lambda=0 and #Γ2\#\Gamma^{\prime}\geq 2.

Proof.

If Γ\Gamma^{\prime}\neq\emptyset, then 0Γ0\in\Gamma^{\prime}. This fact follows from cn(0)=infcn(BRn)c_{n}(0)=\inf c_{n}(B_{R_{n}}) and the fact that the punctures in Γ\Gamma^{\prime} are negative. Arguing by contradiction, assume

Γw𝑑λ=0 and #Γ=1.\int_{\mathbb{C}\setminus\Gamma^{\prime}}w^{*}d\lambda=0\text{ and }\#\Gamma^{\prime}=1~{}.

Thus, Γ={0}\Gamma^{\prime}=\{0\}. Using Theorem 2.6, we conclude that w(ζ=e2π(s+it))=xz(Tzt)w(\zeta=e^{2\pi(s+it)})=x_{z}(T_{z}t), where Pz=(xz,Tz)P_{z}=(x_{z},T_{z}) is the asymptotic limit of v~\tilde{v} at zz. Thus, we have the contradiction

m(z)\displaystyle m(z) =Tz=𝔻wλ=limn𝔻wnλ=limnBϵ(z)vnλσ(T3)\displaystyle=T_{z}=\int_{\partial\mathbb{D}}w^{*}\lambda=\lim_{n\to\infty}\int_{\partial\mathbb{D}}w_{n}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(z)}v_{n}^{*}\lambda-\sigma(T_{3})
=Bϵ(z)vλσ(T3)m(z)σ(T3)2.\displaystyle=\int_{\partial B_{\epsilon}(z)}v^{*}\lambda-\sigma(T_{3})\leq m(z)-\frac{\sigma(T_{3})}{2}.

Here we have used (41), (43) and (45). ∎

3.2.3. Some index estimates

Lemma 3.12.

Let u~=(a,u):Γ×S3\tilde{u}=(a,u):\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} be a finite energy surface such that every puncture in Γ\Gamma is negative, πdu\pi\circ du does not vanish identically and for every asymptotic limit PP of u~\tilde{u}, PP is nondegenerate and μ(P){1,2,3}\mu(P)\in\{1,2,3\}. Then windπ(u~)=0\operatorname{wind}_{\pi}(\tilde{u})=0 and for all zΓ{}z\in\Gamma\cup\{\infty\}, wind(u~,z)=1\operatorname{wind}_{\infty}(\tilde{u},z)=1.

Proof.

We follow [21, Prop 5.9]. By equation (22), for every asymptotic limit PP of u~\tilde{u}, one of the following options hold

(48) μ(P)\displaystyle\mu(P) =1wind(νPneg)=0,wind(νPpos)=1\displaystyle=1~{}\Rightarrow\operatorname{wind}(\nu^{neg}_{P})=0,~{}\operatorname{wind}(\nu_{P}^{pos})=1
μ(P)\displaystyle\mu(P) =2wind(νPneg)=1,wind(νPpos)=1\displaystyle=2~{}\Rightarrow\operatorname{wind}(\nu^{neg}_{P})=1,~{}\operatorname{wind}(\nu_{P}^{pos})=1
μ(P)\displaystyle\mu(P) =3wind(νPneg)=1,wind(νPpos)=2.\displaystyle=3~{}\Rightarrow\operatorname{wind}(\nu^{neg}_{P})=1,~{}\operatorname{wind}(\nu_{P}^{pos})=2.

Then, by Lemma 2.10, if PP is the asymptotic limit of u~\tilde{u} at zz we have

(49) wind(u~,z)\displaystyle\operatorname{wind}_{\infty}(\tilde{u},z) wind(νPneg)1,if z=,\displaystyle\leq\operatorname{wind}(\nu^{neg}_{P})\leq 1,~{}\text{if }z=\infty,
wind(u~,z)\displaystyle\operatorname{wind}_{\infty}(\tilde{u},z) wind(νPpos))1,if zΓ.\displaystyle\geq\operatorname{wind}(\nu^{pos}_{P}))\geq 1,~{}\text{if }z\in\Gamma.

It follows that wind(u~)1#Γ\operatorname{wind}_{\infty}(\tilde{u})\leq 1-\#\Gamma. Using equation (29), we obtain

0windπ(u~)=wind(u~)1+#Γ0.0\leq\operatorname{wind}_{\pi}(\tilde{u})=\operatorname{wind}_{\infty}(\tilde{u})-1+\#\Gamma\leq 0~{}.

Thus windπ=0\operatorname{wind}_{\pi}=0 and wind(u~)=1#Γ\operatorname{wind}_{\infty}(\tilde{u})=1-\#\Gamma. Using (49), we conclude

0wind(u~,)1=zΓwind(u~,z)#Γ=zΓwind(u~,z)10.0\geq\operatorname{wind}_{\infty}(\tilde{u},\infty)-1=\sum_{z\in\Gamma}\operatorname{wind}_{\infty}(\tilde{u},z)-\#\Gamma=\sum_{z\in\Gamma}\operatorname{wind}_{\infty}(\tilde{u},z)-1\geq 0~{}.

We conclude that wind(u~,z)=1\operatorname{wind}_{\infty}(\tilde{u},z)=1, for all zΓ{}z\in\Gamma\cup\{\infty\}. ∎

The following estimate is proved in [25].

Lemma 3.13.

[25, Lemma 3.9] Let v~=(b,v)\tilde{v}=(b,v) be a nonconstant limit of a sequence {v~n=(bn,vn)}\{\tilde{v}_{n}=(b_{n},v_{n})\} satisfying (35)-(38). Then the asymptotic limit PP_{\infty} at the (unique) positive puncture of v~\tilde{v} satisfies μ(P)2\mu(P_{\infty})\geq 2.

3.3. Proof of Proposition 3.5

After a reparametrization of u~n\tilde{u}_{n} and an translation in the \mathbb{R}-direction, we can assume that {u~n}\{\tilde{u}_{n}\} is a germinating sequence as defined in (35)-(38). Indeed, we can take sequences znz_{n}\in\mathbb{C} and 0<δn0<\delta_{n}, such that an(zn)=infan()a_{n}(z_{n})=\inf a_{n}(\mathbb{C}) and Bδn(zn)un𝑑λ=σ(T3)\int_{\mathbb{C}\setminus B_{\delta_{n}(z_{n})}}u_{n}^{*}d\lambda=\sigma(T_{3}), and define

v~n(z)=(bn(z),vn(z))=(an(zn+δnz)an(zn+2δn),un(zn+δnz)).\tilde{v}_{n}(z)=(b_{n}(z),v_{n}(z))=(a_{n}(z_{n}+\delta_{n}z)-a_{n}(z_{n}+2\delta_{n}),u_{n}(z_{n}+\delta_{n}z))~{}.

Then bn(2)=0,nb_{n}(2)=0,\forall n, and

(50) 𝔻vn𝑑λ=Bδn(zn)un𝑑λ=σ(T3).\int_{\mathbb{C}\setminus\mathbb{D}}v_{n}^{*}d\lambda=\int_{\mathbb{C}\setminus B_{\delta_{n}}(z_{n})}u_{n}^{*}d\lambda=\sigma(T_{3}).

Let u~r=(ar,ur):Γr×S3\tilde{u}_{r}=(a_{r},u_{r}):\mathbb{C}\setminus\Gamma_{r}\to\mathbb{R}\times S^{3} be a limit of u~n\tilde{u}_{n} as defined in 3.8. We claim that P3P_{3} is the asymptotic limit of u~r\tilde{u}_{r} at \infty. Indeed, let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be as in the statement of Lemma 3.10. Using the normalization condition (37), we can apply Lemma 3.10 and find R0>1R_{0}>1 such that for RR0R\geq R_{0}, the loops tur(Rei2πt)t\mapsto u_{r}(Re^{i2\pi t}) and {tun(Rei2πt)},n\{t\mapsto u_{n}(Re^{i2\pi t})\},n\in\mathbb{N} belong to 𝒲\mathcal{W}. For fixed RR, the sequence of loops tun(Rei2πt)t\mapsto u_{n}(Re^{i2\pi t}) converges to tur(Rei2πt)t\mapsto u_{r}(Re^{i2\pi t}) in C(S1,S3)C^{\infty}(S^{1},S^{3}), so that for large nn and RR0R\geq R_{0}, tur(Rei2πt)t\mapsto u_{r}(Re^{i2\pi t}) and tun(Rei2πt)t\mapsto u_{n}(Re^{i2\pi t}) belong to the same connected component of 𝒲\mathcal{W}. This implies that P3P_{3} is the asymptotic limit of u~r\tilde{u}_{r} at \infty.

Note that #Γr0\#\Gamma_{r}\neq 0 and 0Γr0\in\Gamma_{r}. Indeed, if Γr=\Gamma_{r}=\emptyset, then u~r:×S3\tilde{u}_{r}:\mathbb{C}\to\mathbb{R}\times S^{3} would satisfy the hypotheses of Theorem 3.3, which contradicts the fact that the family (33) is maximal. Since an(0)=infan()a_{n}(0)=\inf a_{n}(\mathbb{C}) and the punctures in Γr\Gamma_{r} are negative, we have 0Γr0\in\Gamma_{r}.

We claim that {Γr}ur𝑑λ>0\int_{\mathbb{C}\setminus\{\Gamma_{r}\}}u_{r}^{*}d\lambda>0. Suppose, by contradiction, that πdur0\pi\cdot du_{r}\equiv 0. Using Theorem 2.6 and the fact that u~r\tilde{u}_{r} is asymptotic to the simple orbit P3P_{3} at its positive puncture P3P_{3}, we conclude that Γr={0}\Gamma_{r}=\{0\} and ur(ζ=e2π(s+it))=x3(T3t)u_{r}(\zeta=e^{2\pi(s+it)})=x_{3}(T_{3}t). This leads to the contradiction

T3=𝔻urλ=limn𝔻un𝑑λ=T3σ(T3).T_{3}=\int_{\partial\mathbb{D}}u_{r}^{*}\lambda=\lim_{n\to\infty}\int_{\mathbb{D}}u_{n}^{*}d\lambda=T_{3}-\sigma(T_{3}).

Here we have used (50).

Now we prove that Γr={0}\Gamma_{r}=\{0\}. Since P3P_{3} is simple, we know that u~r\tilde{u}_{r} is somewhere injective. By Theorem 2.11, we have the estimate

1ind(u~r)=3zΓrμ(Pz)2+#Γr+1,1\leq\operatorname{ind}(\tilde{u}_{r})=3-\sum_{z\in\Gamma_{r}}\mu(P_{z})-2+\#\Gamma_{r}+1,

where PzP_{z} is the asymptotic limit of u~r\tilde{u}_{r} at zΓrz\in\Gamma_{r}. By Lemma 3.13, we have μ(Pz)2\mu(P_{z})\geq 2, zΓr\forall z\in\Gamma_{r}. Thus, the only possibility is Γ={0}\Gamma=\{0\} and μ(P0)=2\mu(P_{0})=2, where P0=(x0,T0)P_{0}=(x_{0},T_{0}) be the asymptotic limit of u~r\tilde{u}_{r} at 0.

Our next claim is that u~r\tilde{u}_{r} is asymptotic to P2P_{2} at 0. Since, for each nn, unu_{n} is an embedding whose image does not intersect P3P_{3}, it follows that the image of any loop under unu_{n} is not linked to P3P_{3}. This implies that the image of any loop under uru_{r} is not linked to P3P_{3} and consequently that P0P_{0} is not linked to P3P_{3}. Moreover, T3T0={0}ur𝑑λ>0T_{3}-T_{0}=\int_{\mathbb{C}\setminus\{0\}}u_{r}^{*}d\lambda>0. We conclude that P0=P2P_{0}=P_{2}.

As proved in §3.2.2, there exist sequences δn0+\delta_{n}\to 0^{+}, zn0z_{n}\to 0 and RnR_{n}\to\infty such that the sequence of J~\tilde{J}-holomorphic maps w~n=(cn,wn):BRn(0)×S3\tilde{w}_{n}=(c_{n},w_{n}):B_{R_{n}}(0)\to\mathbb{R}\times S^{3} defined by

w~n:=(an(zn+δn)an(zn+2δn),un(zn+δn)).\tilde{w}_{n}:=(a_{n}(z_{n}+\delta_{n}\cdot)-a_{n}(z_{n}+2\delta_{n}),u_{n}(z_{n}+\delta_{n}\cdot)).

is a germinating sequence. Let u~q=(aq,uq):Γq×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3} be a limit of the sequence w~n\tilde{w}_{n}. Then u~q\tilde{u}_{q} is asymptotic to P2P_{2} at z=z=\infty. By Lemma 3.13, we conclude that μ(Pz)2,zΓq,\mu(P_{z})\geq 2,~{}\forall z\in\Gamma_{q}, where PzP_{z} is the asymptotic limit of u~q\tilde{u}_{q} at zz. Since P2P_{2} is simple, we know that u~q\tilde{u}_{q} is somewhere injective. We can apply Theorem 2.11 to u~q\tilde{u}_{q} and obtain

0ind(u~q)=22#Γq2+#Γq+1,0\leq\operatorname{ind}(\tilde{u}_{q})=2-2\#\Gamma_{q}-2+\#\Gamma_{q}+1~{},

which implies #Γq1\#\Gamma_{q}\leq 1. By Proposition 3.11, we have Γquq𝑑λ>0\int_{\mathbb{C}\setminus\Gamma_{q}}u^{*}_{q}d\lambda>0. Again by Theorem 2.11, we have ind(u~q)1\operatorname{ind}(\tilde{u}_{q})\geq 1 and consequently #Γq=0\#\Gamma_{q}=0. We conclude that u~q:×S3\tilde{u}_{q}:\mathbb{C}\to\mathbb{R}\times S^{3} is a finite energy plane asymptotic to the orbit P2P_{2}. This finishes the proof of Proposition 3.5.

3.4. The foliation

Proposition 3.14.

Consider a sequence u~n:=u~τn\tilde{u}_{n}:=\tilde{u}_{\tau_{n}} in the family (33) satisfying τn0+\tau_{n}\to 0^{+}. Let u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3} be the finite energy spheres obtained in Proposition 3.5.

  1. (i)

    Given an S1S^{1}-invariant neighborhood 𝒲3C(/,S3)\mathcal{W}_{3}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx3(T3t)t\mapsto x_{3}(T_{3}t), there exists R0>>1R_{0}>>1 such that, for RR0R\geq R_{0} and large nn, the loop tun(Re2πit)t\mapsto u_{n}(Re^{2\pi it}) belongs to 𝒲3\mathcal{W}_{3}.

  2. (ii)

    Given an S1S^{1}-invariant neighborhood 𝒲2C(/,S3)\mathcal{W}_{2}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx2(T2t)t\mapsto x_{2}(T_{2}t), there exist ϵ1>0\epsilon_{1}>0 and R1>>0R_{1}>>0 such that, for R1δnRϵ1R_{1}\delta_{n}\leq R\leq\epsilon_{1} and large nn, the loop tun(zn+Re2πit)t\mapsto u_{n}(z_{n}+Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}.

  3. (iii)

    Given any neighborhood 𝒱\mathcal{V} of ur({0})uq()P2P3u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}\cup P_{3}, we have un()𝒱u_{n}(\mathbb{C})\subset\mathcal{V} for large nn.

A similar statement works for any sequence τn1\tau_{n}\to 1^{-} with uqu_{q} and uru_{r} replaced by uqu^{\prime}_{q} and uru^{\prime}_{r} respectively.

Proof.

We can assume that 𝒲i,i=2,3\mathcal{W}_{i},i=2,3, contains only the periodic orbit txi(Ti)t\mapsto x_{i}(T_{i}\cdot) modulo S1S^{1}-reparametrizations. Let 𝒲\mathcal{W} be an S1S^{1}-invariant neighborhood of the set of periodic orbits P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) with TT3T\leq T_{3}, viewed as maps xT:S1S3x_{T}:S^{1}\to S^{3}, xT(t)=x(Tt)x_{T}(t)=x(Tt), such that 𝒲2,𝒲3𝒲\mathcal{W}_{2},\mathcal{W}_{3}\subset\mathcal{W}. Using the normalization condition (50) and Lemma 3.10, we find R0>>1R_{0}>>1 such that, for RR0R\geq R_{0}, the loops {tun(Rei2πt)},n\{t\mapsto u_{n}(Re^{i2\pi t})\},n\in\mathbb{N} belong to 𝒲\mathcal{W}. By the asymptotic behavior of the planes u~n\tilde{u}_{n}, we conclude that {tun(Rei2πt)},n\{t\mapsto u_{n}(Re^{i2\pi t})\},n\in\mathbb{N} belong to 𝒲3\mathcal{W}_{3} for RR0R\geq R_{0}. This proves (i).

Applying Lemma 3.10 as in the proof of Proposition 3.9, we find ϵ1>0\epsilon_{1}>0 small and R1>>1R_{1}>>1 such that for every RR satisfying δnR1Rϵ1\delta_{n}R_{1}\leq R\leq\epsilon_{1} and large nn, the loop tun(zn+Re2πit)t\mapsto u_{n}(z_{n}+Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}. This proves (ii).

The proof of (iii) follows from (i), (ii) and Proposition 3.5.∎

Proposition 3.15.

Let u~r=(ar,ur),u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}),\tilde{u}_{r}=(a_{r}^{\prime},u_{r}^{\prime}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and u~q=(aq,uq),u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}),\tilde{u}_{q}^{\prime}=(a_{q}^{\prime},u_{q}^{\prime}):\mathbb{C}\to\mathbb{R}\times S^{3} be the finite energy spheres obtained in Proposition 3.5. Then the projected curves ur:{0}S3u_{r}:\mathbb{C}\setminus\{0\}\to S^{3}, ur:{0}S3u_{r}^{\prime}:\mathbb{C}\setminus\{0\}\to S^{3}, uq:S3u_{q}:\mathbb{C}\to S^{3} and uq:S3u_{q}^{\prime}:\mathbb{C}\to S^{3} are embeddings which are transverse to the Reeb vector field and do not intersect P2P3P_{2}\cup P_{3}.

Proof.

We first prove that ur({0})P3=u_{r}(\mathbb{C}\setminus\{0\})\cap P_{3}=\emptyset. Consider F:{0}×S3F:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} defined by F(e2π(s+it))=(T3s,x3(T3t))F(e^{2\pi(s+it)})=(T_{3}s,x_{3}(T_{3}t)). Note that FF is a finite energy J~\tilde{J}-holomorphic immersion. By Carleman’s similarity principle, the intersections of u~r\tilde{u}_{r} with FF are isolated. By positivity and stability of intersections of pseudo-holomorphic curves, any such intersection implies intersection of u~n\tilde{u}_{n} with FF for large nn, contradicting Theorem 3.3. To prove that ur({0})P2=u_{r}(\mathbb{C}\setminus\{0\})\cap P_{2}=\emptyset, we proceed in the same way, noting that un()P2=,nu_{n}(\mathbb{C})\cap P_{2}=\emptyset,~{}\forall n. Indeed, P2P_{2} and P3P_{3} are not linked and unu_{n} is transverse to the Reeb vector field.

In the same way, we prove that ur({0})u_{r}^{\prime}(\mathbb{C}\setminus\{0\}), uq()u_{q}(\mathbb{C}) and uq()u_{q}^{\prime}(\mathbb{C}) do not intersect P2P3P_{2}\cup P_{3}. Theorem 2.13 shows that uru_{r}, uru_{r}^{\prime}, uqu_{q} and uqu_{q}^{\prime} are embeddings which are transverse to the Reeb vector field. ∎

Proposition 3.16.

Let u~r=(ar,ur),u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}),\tilde{u}_{r}^{\prime}=(a_{r}^{\prime},u_{r}^{\prime}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and u~q=(aq,uq),u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}),\tilde{u}_{q}^{\prime}=(a_{q}^{\prime},u_{q}^{\prime}):\mathbb{C}\to\mathbb{R}\times S^{3} be the finite energy spheres obtained in Proposition 3.5. Then, up to reparametrization and \mathbb{R}-translation, u~q=u~q\tilde{u}_{q}=\tilde{u}^{\prime}_{q}, and u~q\tilde{u}_{q} is the unique finite energy J~\tilde{J}-holomorphic plane asymptotic to P2P_{2}. If ur({0})ur({0})u_{r}(\mathbb{C}\setminus\{0\})\neq u_{r}^{\prime}(\mathbb{C}\setminus\{0\}), then (up to reparametrization and \mathbb{R}-translation) u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} are the unique finite energy J~\tilde{J}-holomorphic cylinders asymptotic to P3P_{3} and P2P_{2} at z=z=\infty and z=0z=0 respectively that do not intersect ×(P2P3)\mathbb{R}\times\left(P_{2}\cup P_{3}\right). Moreover, u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} approach P2P_{2} in opposite directions according to Definition 2.14.

Proof.

Our proof follows [7, Proposition C.1]. First we prove that u~q\tilde{u}_{q} and u~q\tilde{u}^{\prime}_{q} coincide up to reparametrization and \mathbb{R}-translation. Following Theorem 2.7, let ηq\eta_{q} be the asymptotic eigensection of u~q\tilde{u}_{q} at \infty and ηq\eta_{q}^{\prime} the asymptotic eigensection of u~q\tilde{u}^{\prime}_{q} at \infty. It follows from Lemma 3.12 that

(51) wind(ηq)=wind(u~q,)=1 and wind(ηq)=wind(u~q,)=1.\operatorname{wind}(\eta_{q})=\operatorname{wind}_{\infty}(\tilde{u}_{q},\infty)=1\text{ and }\operatorname{wind}(\eta_{q}^{\prime})=\operatorname{wind}_{\infty}(\tilde{u}_{q}^{\prime},\infty)=1.

Since, by Proposition 2.2 and formula (22), νP2neg\nu_{P_{2}}^{neg} is the unique negative eigenvalue of AP2,JA_{P_{2},J} with winding number equal to 11, it follows that ηq\eta_{q} and ηq\eta_{q}^{\prime} are νP2neg\nu^{neg}_{P_{2}}-eigensections. Since the eigenspace of νP2neg\nu_{P_{2}}^{neg} is one dimensional, we find a constant c0c\neq 0 such that ηq=cηq.\eta_{q}=c\eta_{q}^{\prime}~{}. Suppose, contrary to our claim, that uq()uq()u_{q}(\mathbb{C})\neq u^{\prime}_{q}(\mathbb{C}). Then, by (51), Proposition 3.15 and Theorem 2.12, we have

(52) uq()uq()=.u_{q}(\mathbb{C})\cap u^{\prime}_{q}(\mathbb{C})=\emptyset~{}.

By (51), (52) and Theorem 2.15, we conclude that c<0c<0, that is, u~q\tilde{u}_{q} and u~q\tilde{u}_{q}^{\prime} approach P2P_{2} in opposite directions. This implies that uq()uq()P2u_{q}(\mathbb{C})\cup u^{\prime}_{q}(\mathbb{C})\cup P_{2} is a C1C^{1}-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 u~q\tilde{u}_{q} and u~q\tilde{u}^{\prime}_{q} coincide up to reparametrization and \mathbb{R}-translation.

Using the same arguments above and noting that, by Theorem 3.2, any J~\tilde{J}-holomorphic plane asymptotic to P2P_{2} is embedded and does not intersect P2P_{2}, we prove that any finite energy J~\tilde{J}-holomorphic plane asymptotic to P2P_{2} coincides with u~q\tilde{u}_{q} up to reparametrization and \mathbb{R}-translation.

Now we prove the assertions about the cylinders u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime}. Let ηr\eta_{r} and ηr\eta_{r}^{\prime} be the asymptotic eigensections of u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} at 0, respectively. By Lemma 3.12, we have

(53) wind(u~r,0)=wind(u~r,0)=1.\operatorname{wind}_{\infty}(\tilde{u}_{r},0)=\operatorname{wind}_{\infty}(\tilde{u}^{\prime}_{r},0)=1.

Using Proposition 2.2 and formula (22), we conclude that ηr\eta_{r} and ηr\eta_{r}^{\prime} are νP2pos\nu^{pos}_{P_{2}}-eigensections. Since the eigenspace of νP2pos\nu_{P_{2}}^{pos} is one dimensional, we conclude that there exists a nonzero constant cc such that ηr=cηr\eta_{r}^{\prime}=c\eta_{r}. Assume that ur({0})ur({0})u_{r}(\mathbb{C}\setminus\{0\})\neq u_{r}^{\prime}(\mathbb{C}\setminus\{0\}). Then, by (53), Proposition 3.15 and Theorem 2.12, we have

(54) ur({0})ur({0})=.u_{r}(\mathbb{C}\setminus\{0\})\cap u_{r}^{\prime}(\mathbb{C}\setminus\{0\})=\emptyset.

By (53), (54) and Theorem 2.15, we conclude that c<0c<0, that is, u~q\tilde{u}_{q} and u~q\tilde{u}_{q}^{\prime} approach P2P_{2} in opposite directions.

By the same arguments above, we conclude that u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} are the unique cylinders with the properties given in the statement. ∎

Proposition 3.17.

Let u~r=(ar,ur),u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}),\tilde{u}_{r}^{\prime}=(a_{r}^{\prime},u_{r}^{\prime}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3} be the finite energy spheres obtained in Proposition 3.5. Then the cylinders uru_{r} and uru_{r}^{\prime} satisfy ur({0})ur({0})=u_{r}(\mathbb{C}\setminus\{0\})\cap u_{r}^{\prime}(\mathbb{C}\setminus\{0\})=\emptyset, the surface T:=P2P3ur({0})ur({0})T:=P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\}) is homeomorphic to a torus and TP3T\setminus P_{3} is C1C^{1}-embedded. The union of the image of the family {uτ:S3},τ(0,1)\{u_{\tau}:\mathbb{C}\to S^{3}\},~{}\tau\in(0,1), given by (33) with the images of uqu_{q}, uru_{r} and uru^{\prime}_{r} determine a smooth foliation of 1(P2P3)\mathcal{R}_{1}\setminus(P_{2}\cup P_{3}), where 1S3\mathcal{R}_{1}\subset S^{3} is a closed region with boundary TT. Moreover, 1\mathcal{R}_{1} is homeomorphic to a solid torus.

Proof.

By Proposition 3.15 and Theorem 2.12 we conclude that the images of the projected curves uru_{r}, uqu_{q} and {uτ},τ(0,1)\{u_{\tau}\},\tau\in(0,1), are mutually disjoint. Moreover, if ur({0})ur({0})u_{r}(\mathbb{C}\setminus\{0\})\neq u_{r}^{\prime}(\mathbb{C}\setminus\{0\}), then the images of uru_{r}, uru_{r}^{\prime}, uqu_{q} and {uτ},τ(0,1)\{u_{\tau}\},\tau\in(0,1) are mutually disjoint.

Now we prove that ur({0})ur({0})=u_{r}(\mathbb{C}\setminus\{0\})\cap u_{r}^{\prime}(\mathbb{C}\setminus\{0\})=\emptyset. Suppose, by contradiction, that the images of uru_{r} and uru_{r}^{\prime} have a nonempty intersection. Then we have ur({0})=ur({0})u_{r}(\mathbb{C}\setminus\{0\})=u_{r}^{\prime}(\mathbb{C}\setminus\{0\}). Let pS3(ur({0})uq()P2P3)p\in S^{3}\setminus\left(u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}\cup P_{3}\right) and consider a neighborhood 𝒱\mathcal{V} of ur({0})uq()P2P3u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}\cup P_{3} such that p𝒱p\notin\mathcal{V}. By Proposition 3.14, we have uτ0(),uσ0()𝒱u_{\tau_{0}}(\mathbb{C}),u_{\sigma_{0}}(\mathbb{C})\subset\mathcal{V} for some τ0\tau_{0} sufficiently close to 0 and σ0\sigma_{0} sufficiently close to 11. The surface S=uτ0()uσ0()P3S=u_{\tau_{0}}(\mathbb{C})\cup u_{\sigma_{0}}(\mathbb{C})\cup P_{3} is a piecewise smooth embedded sphere. By Jordan-Brouwer separation theorem, SS divides S3SS^{3}\setminus S into two disjoint regions A1A_{1} and A2A_{2} with boundary SS. One of these regions, say A1A_{1}, contains pp and the other contains ur({0})uq()P2u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}. The intersection of the image of the family {uτ}τ(0,1)\{u_{\tau}\}_{\tau\in(0,1)} with A1{A_{1}} is nonempty, open and closed in A1A_{1}. Thus pp is in the image of the family {uτ}τ(0,1)\{u_{\tau}\}_{\tau\in(0,1)}. We conclude that

S3=ur({0})uq()P2P3{uτ()}τ(0,1).S^{3}=u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}\cup P_{3}\cup\{u_{\tau}(\mathbb{C})\}_{\tau\in(0,1)}.

This contradicts the fact that the orbit P1P_{1} is not linked to P3P_{3} and the curves ur,uqu_{r},u_{q} and uτu_{\tau} are transverse to the Reeb vector field. We have proved that ur({0})ur({0})=.u_{r}(\mathbb{C}\setminus\{0\})\cap u_{r}^{\prime}(\mathbb{C}\setminus\{0\})=\emptyset. Consequently, the surface T=P2P3ur({0})ur({0})T=P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\}) is a piecewise smooth embedded torus, and it follows from Proposition 3.16 that TP3T\setminus P_{3} is C1C^{1}-embedded.

By Jordan-Brouwer separation theorem, TT divides S3S^{3} into two closed regions 1\mathcal{R}_{1} and 2\mathcal{R}_{2} with disjoint interiors and boundary TT. One of the regions, say 1\mathcal{R}_{1}, contains the image of the family {uτ}\{u_{\tau}\} and the plane uq()u_{q}(\mathbb{C}). Now we show that

1=P2P3ur({0})ur({0})uq(){uτ()}τ(0,1).\mathcal{R}_{1}=P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup\{u_{\tau}(\mathbb{C})\}_{\tau\in(0,1)}.

Let p1(P2P3ur({0})ur({0})uq())p\in\mathcal{R}_{1}\setminus\left(P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\right) and let 𝒱\mathcal{V} be a neighborhood of P2P3ur({0})ur({0})uq()P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C}) such that p𝒱p\notin\mathcal{V}. By Proposition 3.14, we have uτ0(),uσ0()𝒱u_{\tau_{0}}(\mathbb{C}),u_{\sigma_{0}}(\mathbb{C})\subset\mathcal{V} for some τ0\tau_{0} sufficiently close to 0 and σ0\sigma_{0} sufficiently close to 11. The surface S=uτ0()uσ0()P3S=u_{\tau_{0}}(\mathbb{C})\cup u_{\sigma_{0}}(\mathbb{C})\cup P_{3} is a piecewise smooth embedded sphere. By Jordan-Brouwer separation theorem, SS divides S3SS^{3}\setminus S into two disjoint regions A1A_{1} and A2A_{2} with boundary SS. One of these regions, say A1A_{1} contains pp and is contained in 1\mathcal{R}_{1}. Thus the intersection of the image of the family {uτ}\{u_{\tau}\} with A1A_{1} is open, closed and nonempty in A1A_{1}. This implies that pp is in the image of uτu_{\tau}, for some τ(0,1)\tau\in(0,1).

To prove that 1\mathcal{R}_{1} is homeomorphic to a solid torus, we follow the proof of the Solid torus theorem in [30]. The curve P2P_{2} is the boundary of an embedded 2-disk 𝒟\mathcal{D} contained in 1\mathcal{R}_{1} such that 𝒟̊1̊\mathring{\mathcal{D}}\subset\mathring{\mathcal{R}_{1}}. Let NN be a bicollar neighborhood of 𝒟\mathcal{D} in 1\mathcal{R}_{1}, so that NTN\cap T is an annular neighborhood of P2P_{2} in TT. The boundary of 1N\mathcal{R}_{1}\setminus N is the union of two disks on N\partial N and the set TNT\setminus N, which is an annulus. Thus 1N\mathcal{R}_{1}\setminus N is bounded by a piecewise smooth 22-sphere in S3S^{3}. By the generalized Schönflies theorem, 1N\mathcal{R}_{1}\setminus N is homeomorphic to a closed 33-ball. It follows that 1\mathcal{R}_{1} is homeomorphic to a 33-ball with a D×[0,1]D\times[0,1] attached, where DD is a closed 2-disk (by mapping D×{0}D\times\{0\} and D×{1}D\times\{1\} onto disjoint disks on the boundary of the 33-ball). Since 1\mathcal{R}_{1} is orientable, 1\mathcal{R}_{1} is homeomorphic to D×S1D\times S^{1}. ∎

The proof of Proposition 3.1 is complete.

4. A cylinder asymptotic to P2P_{2} and P1P_{1}

In this section, we continue the proof of Theorem 1.3. Recall that we have obtained a foliation of a closed region 1S3\mathcal{R}_{1}\subset S^{3}. Now we find a finite energy cylinder asymptotic to P2P_{2} at its positive puncture and P1P_{1} at its negative puncture, whose projection to S3S^{3} is contained in the complement of 1\mathcal{R}_{1}.

Proposition 4.1.

Let λ\lambda be a tight contact form on S3S^{3} satisfying the hypotheses of Theorem 1.3. Then there exists a finite energy cylinder v~r=(br,vr):{0}×S3\tilde{v}_{r}=(b_{r},v_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P2P_{2} at its positive puncture z=z=\infty and P1P_{1} at its negative puncture z=0z=0. The projection vrv_{r} is an embedding transverse to the Reeb vector field. Define V=vr({0})V=v_{r}(\mathbb{C}\setminus\{0\}). Then V1=V\cap\mathcal{R}_{1}=\emptyset and DVP2D\cup V\cup P_{2} is a C1C^{1}-embedded disk, where D=uq()D=u_{q}(\mathbb{C}) is given by Proposition 3.1.

The contact structure ξ=kerλ\xi=\ker\lambda coincides, up to a diffeomorphism, with ξ0:=kerfλ0\xi_{0}:=\ker f\lambda_{0}, where λ0\lambda_{0} is defined by (2) and f:S3(0,+)f:S^{3}\to(0,+\infty) is a smooth function. Thus, we can assume that λ=fλ0\lambda=f\lambda_{0} for some smooth function f:S3(0,+)f:S^{3}\to(0,+\infty) without loss of generality. Following [19] we define a symplectic cobordism between (S3,λ=fλ0)(S^{3},\lambda=f\lambda_{0}) and (S3,λE)(S^{3},\lambda_{E}), where λE\lambda_{E} is a dynamically convex contact form on S3S^{3}. Given 0<r1<r20<r_{1}<r_{2}, with r12r22\frac{r_{1}^{2}}{r_{2}^{2}} irrational, let λE=fEλ0\lambda_{E}=f_{E}\lambda_{0} be the contact form associated to the ellipsoid

E={(x1,y1,x2,y2)4|x12+y12r12+x22+y22r22=1},E=\left\{(x_{1},y_{1},x_{2},y_{2})\in\mathbb{R}^{4}|\frac{x_{1}^{2}+y_{1}^{2}}{r_{1}^{2}}+\frac{x_{2}^{2}+y_{2}^{2}}{r_{2}^{2}}=1\right\}~{},

that is, fE(x,y)=(x12+y12r12+x22+y22r22)1f_{E}(x,y)=\left(\frac{x_{1}^{2}+y_{1}^{2}}{r_{1}^{2}}+\frac{x_{2}^{2}+y_{2}^{2}}{r_{2}^{2}}\right)^{-1}. The Reeb vector field RER_{E} defined by λE\lambda_{E} has precisely two simple periodic orbits P¯0\bar{P}_{0} and P¯1\bar{P}_{1}. Both periodic orbits and its iterates are nondegenerate. Their Conley-Zehnder indices are μ(P¯0)=3\mu(\bar{P}_{0})=3 and μ(P¯1)=2k+15\mu(\bar{P}_{1})=2k+1\geq 5 where k2k\geq 2 is determined by k<1+(r12r22)<k+1k<1+\left(\frac{r_{1}^{2}}{r_{2}^{2}}\right)<k+1. See [16, Lemma 1.6] for a proof of these facts.

We choose 0<r1<r20<r_{1}<r_{2} small enough so that fE<f pointwise on S3f_{E}<f\text{ pointwise on }S^{3} and a smooth function h:×S3+h:\mathbb{R}\times S^{3}\to\mathbb{R}^{+} satisfying

(55) h(a,)\displaystyle h(a,\cdot) =fE, if a2,\displaystyle=f_{E},\text{ if }a\leq-2,
h(a,)\displaystyle h(a,\cdot) =f, if a2,\displaystyle=f,\text{ if }a\geq 2,
(56) ha0 on ×S3 and ha>σ>0 on [1,1]×S3.\dfrac{\partial h}{\partial a}\geq 0\text{ on }\mathbb{R}\times S^{3}\text{ and }\dfrac{\partial h}{\partial a}>\sigma>0\text{ on }[-1,1]\times S^{3}.

In view of (56), the 22-form d(hλ0)d(h\lambda_{0}) restricted to [1,1]×S3[-1,1]\times S^{3} is a symplectic form.

We consider the family of contact forms {λa=h(a,)λ0,a}\{\lambda_{a}=h(a,\cdot)\lambda_{0},a\in\mathbb{R}\}. The contact structure ξ=kerλa\xi=\ker\lambda_{a} does not depend on aa. Choose JE𝒥(ξ,dλE)J_{E}\in\mathcal{J}(\xi,d\lambda_{E}) and let {Ja𝒥(ξ,dλa),a}\{J_{a}\in\mathcal{J}(\xi,d\lambda_{a}),a\in\mathbb{R}\} be a smooth family of dλad\lambda_{a}-compatible complex structures on ξ\xi so that Ja=JJ_{a}=J if a2a\geq 2 and Ja=JEJ_{a}=J_{E} if a2a\leq-2. We consider smooth almost complex structures J¯\bar{J} on the symplectization ×S3\mathbb{R}\times S^{3} with the following properties. On ([1,1])×S3(\mathbb{R}\setminus[-1,1])\times S^{3}, we consider

J¯|ξ=Ja and J¯a=Rλa.\bar{J}|_{\xi}=J_{a}\text{ and }\bar{J}\partial_{a}=R_{\lambda_{a}}~{}.

On [1,1]×S3[-1,1]\times S^{3} we only require J¯\bar{J} to be compatible with the symplectic form d(hλ0)d(h\lambda_{0}). The space of such almost complex structures on ×S3\mathbb{R}\times S^{3} is nonempty and contractible in the CC^{\infty}-topology and will be denoted by 𝒥(λ,J,λE,JE)\mathcal{J}(\lambda,J,\lambda_{E},J_{E}).

4.1. Generalized finite energy surfaces

Definition 4.2.

Let (S,j)(S,j) be a closed Riemann surface and let ΓS\Gamma\subset S be a nonempty finite set. A nonconstant smooth map u~:SΓ×S3\tilde{u}:S\setminus\Gamma\to\mathbb{R}\times S^{3} is called a generalized finite energy surface if it is J¯\bar{J}-holomorphic, that is, satisfies du~j=J¯(u~)du~d\tilde{u}\circ j=\bar{J}(\tilde{u})\circ d\tilde{u}, for some J¯𝒥(λ,J,λE,JE)\bar{J}\in\mathcal{J}(\lambda,J,\lambda_{E},J_{E}), as well as the energy condition

E(u~)<+,E(\tilde{u})<+\infty~{},

where the energy E(u~)E(\tilde{u}) is defined as follows. Let Σ\Sigma be the collection of smooth functions ϕ:[0,1]\phi:\mathbb{R}\to[0,1] satisfying ϕ0\phi^{\prime}\geq 0 and ϕ=12\phi=\frac{1}{2} on [1,1][-1,1]. Given ϕΣ\phi\in\Sigma we define the 11-form τϕ\tau_{\phi} on ×S3\mathbb{R}\times S^{3} by

τϕ(a,x)(h,k)=ϕ(a)λa(x)(k),\tau_{\phi}(a,x)(h,k)=\phi(a)\lambda_{a}(x)(k),

for any (a,x)×S3(a,x)\in\mathbb{R}\times S^{3} and (h,k)T(a,x)(×S3)(h,k)\in T_{(a,x)}(\mathbb{R}\times S^{3}). Then

(57) E(u~)=supϕΣSΓu~𝑑τϕ.E(\tilde{u})=\sup_{\phi\in\Sigma}\int_{S\setminus\Gamma}\tilde{u}^{*}d\tau_{\phi}.

Theorem 2.11 is still valid for almost complex structures in 𝒥(λ,J,λE,JE)\mathcal{J}(\lambda,J,\lambda_{E},J_{E}).

Theorem 4.3 ([9]).

There exists a dense subset 𝒥regE𝒥(λ,J,λE,JE)\mathcal{J}^{E}_{reg}\subset\mathcal{J}(\lambda,J,\lambda_{E},J_{E}) such that if u~=(a,u):SΓ×S3\tilde{u}=(a,u):S\setminus\Gamma\to\mathbb{R}\times S^{3} is a somewhere injective generalized finite energy surface for J¯𝒥regE\bar{J}\in\mathcal{J}^{E}_{reg}, and has nondegenerate asymptotic limits at all of its punctures, then

0ind(u~)=μ(u~)χ(S)+#Γ.0\leq\operatorname{ind}(\tilde{u})=\mu(\tilde{u})-\chi({S})+\#\Gamma~{}.

4.2. A family of J¯\bar{J}-holomorphic planes asymptotic to P2P_{2}

We are interested in the space of finite energy generalized J¯\bar{J}-holomorphic planes asymptotic to the orbit P2P_{2}, for fixed J¯𝒥regE\bar{J}\in\mathcal{J}^{E}_{reg}. The following theorem is a consequence of results from [20].

Theorem 4.4 ([20]).

Let u~0:×S3\tilde{u}_{0}:\mathbb{C}\to\mathbb{R}\times S^{3} be an embedded finite energy J¯\bar{J}-holomorphic plane, asymptotic to a nondegenerate, simple Reeb orbit P=(x,T)P=(x,T) satisfying μ(P)=2\mu(P)=2. Then there exists a smooth embedding

Φ~:×(ϵ,ϵ)×S3\tilde{\Phi}:\mathbb{C}\times(-\epsilon,\epsilon)\to\mathbb{R}\times S^{3}

with the following properties:

  • Φ~(,0)=u~0\tilde{\Phi}(\cdot,0)=\tilde{u}_{0};

  • For every τ(ϵ,ϵ)\tau\in(-\epsilon,\epsilon), the map zΦ~(z,τ)z\mapsto\tilde{\Phi}(z,\tau) is a generalized finite energy J¯\bar{J}-holomorphic plane asymptotic to PP;

  • If u~n\tilde{u}_{n} is a sequence of finite energy J¯\bar{J}-holomorphic planes asymptotic to PP satisfying u~nu~0\tilde{u}_{n}\to\tilde{u}_{0} in Cloc()C^{\infty}_{loc}(\mathbb{C}) as n+n\to+\infty, then there exist sequences An,BnA_{n},B_{n} in \mathbb{C} with An1A_{n}\to 1, Bn0B_{n}\to 0 and τn\tau_{n} in (ϵ,ϵ)(-\epsilon,\epsilon) with τn0\tau_{n}\to 0 such that

    u~n(z)=Φ~(Anz+Bn,τn)\tilde{u}_{n}(z)=\tilde{\Phi}(A_{n}z+B_{n},\tau_{n})

    for sufficiently large nn.

From now on we fix J¯𝒥regE\bar{J}\in\mathcal{J}^{E}_{reg}, where 𝒥regE\mathcal{J}^{E}_{reg} is given by Theorem 4.3. Let Θ\Theta be the space of generalized finite energy J¯\bar{J}-holomorphic planes asymptotic to P2P_{2}, modulo holomorphic reparametrizations. By Theorem 4.4, Θ\Theta is a smooth 11-dimensional manifold.

Lemma 4.5.

The space Θ\Theta is nonempty.

Proof.

Let u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3} be the J~\tilde{J}-holomorphic plane asymptotic to P2P_{2} given by Theorem 3.5. After an \mathbb{R}-translation we can assume that minzaq(z)>2\min_{z\in\mathbb{C}}a_{q}(z)>2, so that u~q=(aq,uq)\tilde{u}_{q}=(a_{q},u_{q}) can be viewed as a generalized finite energy J¯\bar{J}-holomorphic plane asymptotic to P2P_{2}. ∎

By the proof of Lemma 4.5, we have u~qΘ\tilde{u}_{q}\in\Theta. Let Θ\Theta^{\prime} be the connected component of Θ\Theta containing u~q\tilde{u}_{q}.

4.3. Existence of a 𝑱~\boldsymbol{\tilde{J}}-holomorphic cylinder

Consider a sequence v~n=(bn,vn)\tilde{v}_{n}=(b_{n},v_{n}) of generalized finite energy planes representing elements of Θ\Theta^{\prime} . The energy E(v~n)E(\tilde{v}_{n}) is uniformly bounded by T2T_{2}. 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 v~n\tilde{v}_{n}, still denoted by v~n\tilde{v}_{n}, there exists a bubbling-off tree =(𝒯,𝒰)\mathcal{B}=(\mathcal{T},\mathcal{U}) with the following properties

  • For every vertex qq of 𝒯\mathcal{T} there exist sequences znq,δnqz_{n}^{q},~{}\delta_{n}^{q}\in\mathbb{C} and cnqc_{n}^{q}\in\mathbb{R} such that

    (58) v~n(znq+δnq)+cnqv~qin Cloc(Γq) as n\tilde{v}_{n}(z^{q}_{n}+\delta_{n}^{q}\cdot)+c_{n}^{q}\to\tilde{v}_{q}~{}\text{in }C^{\infty}_{loc}(\mathbb{C}\setminus\Gamma_{q})\text{ as }n\to\infty

    Here v~+c:=(b+c,v)\tilde{v}+c:=(b+c,v), where v~=(b,v)\tilde{v}=(b,v) and cc\in\mathbb{R}

  • The curve v~r\tilde{v}_{r} is asymptotic to P2P_{2} at \infty and the asymptotic limits of all curves v~q\tilde{v}_{q} are closed orbits with periods T2\leq T_{2} of the Reeb flow of either λ\lambda or λE\lambda_{E}.

Here, a bubbling off tree =(𝒯,𝒰)\mathcal{B}=(\mathcal{T},\mathcal{U}) consists of a finite, rooted tree 𝒯=(E,{r},V)\mathcal{T}=(E,\{r\},V), with edges oriented away from the root, and a finite set of pseudo-holomorphic spheres 𝒰\mathcal{U}, satisfying the following properties.

  • There is a bijective correspondence between vertices q𝒯q\in\mathcal{T} and finite-energy punctured spheres v~q:Γq×S3𝒰\tilde{v}_{q}:\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3}\in\mathcal{U}. Each v~q:Γq×S3\tilde{v}_{q}:\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3} is pseudo-holomorphic with respect to either J~\tilde{J}, J~E\tilde{J}_{E} or J¯\bar{J} and has finite energy.

  • Each sphere v~q\tilde{v}_{q} has exactly one positive puncture at \infty and a finite set Γq\Gamma_{q} of negative punctures. If v~q\tilde{v}_{q} is J~\tilde{J} or J~E\tilde{J}_{E}-holomorphic and its contact area vanishes, then #Γq2\#\Gamma_{q}\geq 2.

  • Each ordered path (q1,,qN)(q_{1},\dotsc,q_{N}) from the root q1=rq_{1}=r to a leaf qNq_{N}, where qk+1q_{k+1} is a direct descendant of qkq_{k}, contains at most one vertex qiq_{i} such that v~qi\tilde{v}_{q_{i}} is J¯\bar{J}-holomorphic, in which case v~qj\tilde{v}_{q_{j}} is J~\tilde{J}-holomorphic 1j<i\forall 1\leq j<i, and v~qj\tilde{v}_{q_{j}} is J~E\tilde{J}_{E}-holomorphic i<jN\forall i<j\leq N.

  • If the vertex qq is not the root then qq has an incoming edge ee from a vertex qq^{\prime}, and #Γq\#\Gamma_{q} outgoing edges f1,,f#Γqf_{1},\dots,f_{\#\Gamma_{q}} to vertices p1,,p#Γqp_{1},\dots,p_{\#\Gamma_{q}} of 𝒯\mathcal{T}, respectively. The edge ee is associated to the positive puncture of v~q\tilde{v}_{q} and the edges f1,,f#Γqf_{1},\dots,f_{\#\Gamma_{q}} are associated to the negative punctures of v~q\tilde{v}_{q}. The asymptotic limit of v~q\tilde{v}_{q} at its positive puncture coincides with the asymptotic limit of v~q\tilde{v}_{q^{\prime}} at its negative puncture associated to ee. In the same way, the asymptotic limit of v~q\tilde{v}_{q} at a negative puncture corresponding to fif_{i} coincides with the asymptotic limit of v~pi\tilde{v}_{p_{i}} 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 znq,δnq,cnqz_{n}^{q},\delta_{n}^{q},c_{n}^{q} be sequences such that (58) holds for all vertices qq of 𝒯\mathcal{T}. Then we can assume, up to a selection of a subsequence still denoted by v~n\tilde{v}_{n}, that one of the three mutually excluding possibilities holds for each vertex qq.

  • (I)

    cnqc_{n}^{q} is bounded, bn(zn+δnq)b_{n}(z_{n}+\delta_{n}^{q}\cdot) is Cloc0(Γq)C^{0}_{loc}(\mathbb{C}\setminus\Gamma_{q})-bounded and v~q\tilde{v}_{q} is a J¯\bar{J}-holomorphic curve;

  • (II)

    cnqc_{n}^{q}\to-\infty, bn(znq+δnq)+b_{n}(z_{n}^{q}+\delta_{n}^{q}\cdot)\to+\infty in Cloc0(Γq)C^{0}_{loc}(\mathbb{C}\setminus\Gamma_{q}) as nn\to\infty and v~q\tilde{v}_{q} is a J~\tilde{J}-holomorphic curve;

  • (III)

    cnq+c_{n}^{q}\to+\infty, bn(znq+δnq)b_{n}(z_{n}^{q}+\delta_{n}^{q}\cdot)\to-\infty in Cloc0(Γq)C^{0}_{loc}(\mathbb{C}\setminus\Gamma_{q}) as nn\to\infty and v~q\tilde{v}_{q} is a J~E\tilde{J}_{E}-holomorphic curve.

Moreover, if qq is a vertex for which (III) holds, then v~q\tilde{v}_{q} is asymptotic at its positive puncture to a Reeb orbit having period strictly less than T2T_{2}. In particular, (III) does not hold for the root rr.

The following lemma is proved using the maximum principle combined with estimates for cylinders with small area (Lemma 3.10).

Lemma 4.8.

Let znr,δnr,cnrz_{n}^{r},\delta_{n}^{r},c_{n}^{r} be sequences such that (58) holds for the root rr. Then, by Theorem 4.6, P2=(x2,T2)P_{2}=(x_{2},T_{2}) is the asymptotic limit of v~r\tilde{v}_{r} at the positive puncture \infty. For every /\mathbb{R}/\mathbb{Z}-invariant neighborhood 𝒲\mathcal{W} of tx2(T2t)t\mapsto x_{2}(T_{2}t) in C(/,S3)C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) and for every number M>0M>0, there exist R0>0R_{0}>0 and n0n_{0} such that if R>R0R>R_{0} and n>n0n>n_{0}, then the loop tvn(znr+δnrRei2πt)t\mapsto v_{n}(z_{n}^{r}+\delta_{n}^{r}Re^{i2\pi t}) belongs to 𝒲\mathcal{W} and bnr(znr+δnrei2πt)+cnr>Mb_{n}^{r}(z_{n}^{r}+\delta_{n}^{r}e^{i2\pi t})+c_{n}^{r}>M.

Since Θ\Theta^{\prime} is a connected 11-dimensional manifold without boundary, it is diffeomorphic either to S1S^{1} or to an open interval. It can not be diffeomorphic to S1S^{1}, since it contains the family {[(aq+t,uq)]}t>0\{[(a_{q}+t,u_{q})]\}_{t>0} of equivalence classes of translations of u~q\tilde{u}_{q}, where u~q=(aq,uq)\tilde{u}_{q}=(a_{q},u_{q}) is the plane obtained in Theorem 3.5. Thus, the family is diffeomorphic to an interval and we can assume that Θ={[v~τ=(bτ,vτ)]}τ(τ,+)\Theta^{\prime}=\{[\tilde{v}_{\tau}=(b_{\tau},v_{\tau})]\}_{\tau\in(\tau_{-},+\infty)}, where for τ0\tau\geq 0, v~τ=(aq+τ,uq)\tilde{v}_{\tau}=(a_{q}+\tau,u_{q}). Consider a sequence v~n:=v~τn\tilde{v}_{n}:=\tilde{v}_{\tau_{n}} satisfying τnτ\tau_{n}\to\tau_{-}.

Proposition 4.9.

The bubbling-off tree obtained as an SFT-limit of the sequence [v~n][\tilde{v}_{n}], as in Theorem 4.6, is as follows. The tree has vertices r,qr,q, where rr is the root and qq is a leaf and direct descendant of rr. The root rr corresponds to a J~\tilde{J}-holomorphic cylinder v~r=(br,vr):{0}×S3\tilde{v}_{r}=(b_{r},v_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P2P_{2} at its positive puncture z=z=\infty and to P1P_{1} at its negative puncture z=0z=0. The leaf qq corresponds to a J¯\bar{J}-holomorphic plane asymptotic to P1P_{1}.

Proof.

Let =(𝒯,𝒰)\mathcal{B}=(\mathcal{T},\mathcal{U}) be the bubbling-off tree given by Theorem 4.6 and let v~r:Γr×S3\tilde{v}_{r}:\mathbb{C}\setminus\Gamma_{r}\to\mathbb{R}\times S^{3} be the finite energy sphere associated to the root of 𝒯\mathcal{T}. By Lemma 4.7, v~r\tilde{v}_{r} is not J~E\tilde{J}_{E}-holomorphic. Now we show that v~r\tilde{v}_{r} is J~\tilde{J}-holomorphic.

Suppose, by contradiction, that v~r\tilde{v}_{r} is J¯\bar{J}-holomorphic. By Theorem 4.6, v~r\tilde{v}_{r} is asymptotic to P2P_{2} at \infty. Since P2P_{2} is simple, v~r\tilde{v}_{r} is somewhere injective. If Γr=\Gamma_{r}=\emptyset, then v~rΘ\tilde{v}_{r}\in\Theta, contradicting the fact that the interval Θ\Theta^{\prime} is maximal. Assume Γr\Gamma_{r}\neq\emptyset. By Theorem 4.3, it follows that

(59) 0ind(v~r)=2zΓrμ(Pz)2+#Γr+1.0\leq\operatorname{ind}(\tilde{v}_{r})=2-\sum_{z\in\Gamma_{r}}\mu(P_{z})-2+\#\Gamma_{r}+1~{}.

Then

(60) zΓrμ(Pz)#Γr1.\sum_{z\in\Gamma_{r}}\mu(P_{z})-\#\Gamma_{r}\leq 1~{}.

Contradicting the fact that for all zΓrz\in\Gamma_{r}, the asymptotic limit PzP_{z} of v~r\tilde{v}_{r} at zz is a Reeb orbit of λE\lambda_{E}, which is a dynamically convex contact from, that is, μ(Pz)3\mu(P_{z})\geq 3, for all zΓrz\in\Gamma_{r}. Thus, v~r\tilde{v}_{r} is J~\tilde{J}-holomorphic.

Let m:=minaq()m:=\min a_{q}(\mathbb{C}). We claim that

lim supn(minbn())m.\limsup_{n}\left(\min b_{n}(\mathbb{C})\right)\leq m~{}.

To prove the claim, suppose by contradiction that lim supn(minbn())>m>2.\limsup_{n}\left(\min b_{n}(\mathbb{C})\right)>m>2. Then there exists a subsequence v~nk=(bnk,vnk)\tilde{v}_{n_{k}}=(b_{n_{k}},v_{n_{k}}) satisfying minbnk()>m\min b_{n_{k}}(\mathbb{C})>m. By the definiton of J¯\bar{J}, the planes v~nk\tilde{v}_{n_{k}} are J~\tilde{J}-holomorphic. By Proposition 3.16, we know that u~q\tilde{u}_{q} is the unique finite energy J~\tilde{J}-holomorphic plane asymptotic to P2P_{2}, up to reparametrization and \mathbb{R}-translation. This implies [v~nk]=[aq+τk,uq][\tilde{v}_{n_{k}}]=[a_{q}+\tau_{k},u_{q}] for a sequence τk>0\tau_{k}>0. This contradicts the fact that Θ\Theta^{\prime} is an interval.

Now we show that Γr\Gamma_{r}\neq\emptyset. Suppose, by contradiction, that v~r\tilde{v}_{r} is a J~\tilde{J}-holomorphic plane. Let znz_{n}, δn\delta_{n} and cnc_{n} be the sequences given by Theorem 4.6, such that

v~n(zn+δn)+cnv~rin Cloc() as n.\tilde{v}_{n}(z_{n}+\delta_{n}\cdot)+c_{n}\to\tilde{v}_{r}~{}\text{in }C^{\infty}_{loc}(\mathbb{C})\text{ as }n\to\infty.

The limit v~r\tilde{v}_{r} satisfies 4.7(II), so that cnc_{n}\to-\infty and bn(zn+δn)+b_{n}(z_{n}+\delta_{n}\cdot)\to+\infty in Cloc0()C^{0}_{loc}(\mathbb{C}) as nn\to\infty. By Lemma 4.8, there exists R0>0R_{0}>0 such that bn(zn+δnz)>mcn>mb_{n}(z_{n}+\delta_{n}z)>m-c_{n}>m for |z|>R0|z|>R_{0} and nn large enough. For |z|R0|z|\leq R_{0}, by Lemma 4.7(II), we have bn(zn+δnz)>mb_{n}(z_{n}+\delta_{n}z)>m, for nn large enough. Thus, infbn()>m\inf b_{n}(\mathbb{C})>m for nn large enough. This contradicts lim supninfbn()m\limsup_{n}\inf b_{n}(\mathbb{C})\leq m, and concludes the proof of Γr\Gamma_{r}\neq\emptyset.

So far, we know that v~r:Γr×M\tilde{v}_{r}:\mathbb{C}\setminus\Gamma_{r}\to\mathbb{R}\times M is a J~\tilde{J}-holomorphic sphere and Γr\Gamma_{r}\neq\emptyset. The next step is to prove that every negative asymptotic limit of v~r\tilde{v}_{r} has Conley-Zehnder index equal to 1.

Claim I

If qq is not the root and PP_{\infty} is the asymptotic limit of v~q\tilde{v}_{q} at \infty, then μ(P)1\mu(P_{\infty})\geq 1. To prove Claim I, we argue indirectly assuming that μ(P)0\mu(P_{\infty})\leq 0. The curve v~q:Γq×S3\tilde{v}_{q}:\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3} factors as v~q=u~p,\tilde{v}_{q}=\tilde{u}\circ p~{}, where u~:Γ×S3\tilde{u}:\mathbb{C}\setminus\Gamma^{\prime}\to\mathbb{R}\times S^{3} is a somewhere injective finite energy sphere and pp is a polynomial. If PP is the asymptotic limit of u~\tilde{u} at \infty, then Pdegp=PP^{\deg p}=P_{\infty}. By Lemma 2.1, μ(P)0\mu(P_{\infty})\leq 0 implies μ(P)0\mu(P)\leq 0. By Theorem 4.3, we have

0indu~:=μ(P)zΓμ(Pz)+#Γq1,0\leq\operatorname{ind}\tilde{u}:=\mu(P)-\sum_{z^{\prime}\in\Gamma^{\prime}}\mu(P_{z^{\prime}})+\#\Gamma_{q}-1,

where PzP_{z^{\prime}} is the asymptotic limit of u~\tilde{u} at zz^{\prime}. If Γ=\Gamma^{\prime}=\emptyset, we already have a contradiction. Otherwise, there exists z0Γz_{0}^{\prime}\in\Gamma^{\prime} such that μ(Pz0)0\mu(P_{z_{0}^{\prime}})\leq 0. Let z0Γqz_{0}\in\Gamma_{q} be such that p(z0)=z0p(z_{0})=z_{0}^{\prime}. Then Pz0=Pz0kP_{z_{0}}=P_{z_{0}^{\prime}}^{k}, for some kdegpk\leq\deg p, where Pz0P_{z_{0}} is the asymptotic limit of v~q\tilde{v}_{q} at z0z_{0}. By Lemma 2.1, we have μ(Pz0)0\mu(P_{z_{0}})\leq 0. Since the tree has a finite number of vertices, by induction we find a leaf ll of the tree such that the finite energy plane u~l:×S3\tilde{u}_{l}:\mathbb{C}\to\mathbb{R}\times S^{3} is asymptotic to an orbit PP with μ(P)0\mu(P)\leq 0, a contradiction. This proves Claim I.

Claim II

For every zΓrz\in\Gamma_{r}, we have μ(Pz)=1\mu(P_{z})=1, where PzP_{z} is the asymptotic limit of v~r\tilde{v}_{r} at zz. To prove Claim II, first note that, by Theorem 2.11, we have

0indv~r=2zΓrμ(Pz)+#Γr1.0\leq\operatorname{ind}\tilde{v}_{r}=2-\sum_{z\in\Gamma_{r}}\mu(P_{z})+\#\Gamma_{r}-1~{}.

It follows that

zΓrμ(Pz)#Γr+1.\sum_{z\in\Gamma_{r}}\mu(P_{z})\leq\#\Gamma_{r}+1~{}.

Since, by Claim I, we have 1μ(Pz)1\leq\mu(P_{z}), for every zΓrz\in\Gamma_{r}, there exists at most one puncture z0Γrz_{0}\in\Gamma_{r} such that μ(Pz0)2\mu(P_{z_{0}})\geq 2. If there exists such puncture, we have ind(v~r)=0\operatorname{ind}(\tilde{v}_{r})=0. By Theorem 2.11, this implies πdvr0\pi\circ dv_{r}\equiv 0. By the definition of bubbling-off tree, this implies #Γr2\#\Gamma_{r}\geq 2. By Theorem 2.6, there exists a periodic orbit PP and a polynomial p:p:\mathbb{C}\to\mathbb{C} such that p1(0)=Γrp^{-1}(0)=\Gamma_{r} and v~r=FPp\tilde{v}_{r}=F_{P}\circ p, where FPF_{P} is the cylinder over the orbit PP, which contradicts the fact that P2P_{2} is simple. This proves Claim II. We have also proved that Γrvr𝑑λ>0\int_{\mathbb{C}\setminus\Gamma_{r}}v_{r}^{*}d\lambda>0.

We are now in a position to show that Γr={0}\Gamma_{r}=\{0\} and v~r:{0}×S3\tilde{v}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} is a cylinder asymptotic to P1P_{1} at z=0z=0. By Lemma 3.12, we conclude that wind(v~r,z)=1,zΓr{}\operatorname{wind}_{\infty}(\tilde{v}_{r},z)=1,~{}\forall z\in\Gamma_{r}\cup\{\infty\}. Recall that uq()P2=.u_{q}(\mathbb{C})\cap P_{2}=\emptyset~{}. Thus, the curves u~q\tilde{u}_{q} and v~r\tilde{v}_{r} satisfy condition (2) of Theorem 2.12. Here u~q\tilde{u}_{q} is the plane asymptotic to P2P_{2} obtained by Theorem 3.5. It follows from Theorem 2.12 that the projected curve uqu_{q} does not intersect any of the negative asymptotic limits of v~r\tilde{v}_{r}. This implies that P2P_{2} is contractible in S3PzS^{3}\setminus P_{z}, for every PzP_{z} asymptotic limit of v~r\tilde{v}_{r} at zΓrz\in\Gamma_{r}. Consequently,

lk(P2,Pz)=lk(Pz,P2)=0,zΓr.{\rm lk}(P_{2},P_{z})={\rm lk}(P_{z},P_{2})=0,~{}\forall z\in\Gamma_{r}.

Since, by hypothesis, the orbit P1P_{1} is the only orbit with Conley-Zehnder index 11 and period less than T2T_{2} that is not linked to P2P_{2}, it follows that

Pz=P1,zΓr.P_{z}=P_{1},~{}~{}\forall z\in\Gamma_{r}~{}.

Applying Theorem 2.12 to the finite energy curves v~r:Γr×S3\tilde{v}_{r}:\mathbb{C}\setminus\Gamma_{r}\to\mathbb{R}\times S^{3} and u~τ:×S3\tilde{u}_{\tau}:\mathbb{C}\to\mathbb{R}\times S^{3}, where u~τ\tilde{u}_{\tau} is any of the planes in the family (33), we prove that vrv_{r} does not intersect uτu_{\tau}. We conclude that vr(Γr)2v_{r}(\mathbb{C}\setminus\Gamma_{r})\subset\mathcal{R}_{2}, where 2:=S31̊\mathcal{R}_{2}:=S^{3}\setminus\mathring{\mathcal{R}_{1}} and 1\mathcal{R}_{1} is the closed region given by Proposition 3.17. The region 1\mathcal{R}_{1} contains an embedded disk with boundary P2P_{2}, so that P2P_{2} is contractible in 1\mathcal{R}_{1}. One can show, using Mayer-Vietoris sequence, that the holomology class of /tx2(T2t)\mathbb{R}/\mathbb{Z}\ni t\mapsto x_{2}(T_{2}t) generates H1(2,)H_{1}(\mathcal{R}_{2},\mathbb{Z}). The projected curve vrv_{r} defines a singular 2-chain in C2(2,)C_{2}(\mathcal{R}_{2},\mathbb{Z}), so that [x2(T2)]=n[x1(T1)][x_{2}(T_{2}\cdot)]=n[x_{1}(T_{1}\cdot)] in H1(2,)H_{1}(\mathcal{R}_{2},\mathbb{Z}), where n=#Γrn=\#\Gamma_{r}. It follows that #Γr=1\#\Gamma_{r}=1 and we can assume that Γr={0}\Gamma_{r}=\{0\}.

Now we prove that the next (and last) level of the bubbling-off tree consists of a J¯\bar{J}-holomorphic plane. Let v~q:Γq×S3\tilde{v}_{q}:\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3} be the finite energy sphere associated to the unique vertex that is a direct descendant of the root rr. The asymptotic limit of v~q\tilde{v}_{q} at \infty is P1P_{1}, that is a simple orbit. It follows that v~q\tilde{v}_{q} is somewhere injective. By the definition of bubbling-off tree, v~q\tilde{v}_{q} is either J~\tilde{J}-holomorphic or J¯\bar{J}-holomorphic. By Theorem 4.3, we have

0indv~q=1zΓqμ(Pz)+#Γq1=#ΓqzΓqμ(Pz),0\leq\operatorname{ind}\tilde{v}_{q}=1-\sum_{z\in\Gamma_{q}}\mu(P_{z})+\#\Gamma_{q}-1=\#\Gamma_{q}-\sum_{z\in\Gamma_{q}}\mu(P_{z})~{},

where PzP_{z} is the asymptotic limit of v~q\tilde{v}_{q} at zΓqz\in\Gamma_{q}. Since, by Claim I, μ(Pz)1\mu(P_{z})\geq 1 for all zΓqz\in\Gamma_{q}, it follows that indv~q=0\operatorname{ind}\tilde{v}_{q}=0 and μ(Pz)=1\mu(P_{z})=1, for all zΓqz\in\Gamma_{q}.

Suppose, by contradiction, that v~q\tilde{v}_{q} is J~\tilde{J}-holomorphic. By Theorem 2.11, we have πdvq0\pi\circ dv_{q}\equiv 0. By the definition of bubbling-off tree, this implies #Γq2\#\Gamma_{q}\geq 2. Theorem 2.6 and the fact that P1P_{1} is simple lead to a contradiction. We have proved that v~q\tilde{v}_{q} is J¯\bar{J}-holomorphic.

Suppose that Γq\Gamma_{q}\neq\emptyset. By the definition of bubbling-off tree, if ll is a vertex of the tree that is a direct descendant of qq, then v~l\tilde{v}_{l} is necessarily J~E\tilde{J}_{E}-holomorphic. The asymptotic limit of v~l\tilde{v}_{l} at \infty is equal to PzP_{z} for some zΓqz\in\Gamma_{q}. But μ(Pz)=1\mu(P_{z})=1 for all zΓqz\in\Gamma_{q}, contradicting the fact that all closed orbits of RER_{E} have Conley-Zehnder index 3\geq 3. We have proved that Γq=\Gamma_{q}=\emptyset. This finishes the proof of Proposition 4.9.

Proposition 4.10.

The curve v~r=(br,vr):{0}×S3\tilde{v}_{r}=(b_{r},v_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} obtained in Proposition 4.9 is an embedding. The projected curve vr:{0}S3v_{r}:\mathbb{C}\setminus\{0\}\to S^{3} is an embedding which is transverse to the Reeb vector field and does not intersect any of its asymptotic limits. Moreover, vr({0})1=v_{r}(\mathbb{C}\setminus\{0\})\cap\mathcal{R}_{1}=\emptyset, where 1\mathcal{R}_{1} is the closed region obtained in Proposition 3.17.

To prove Proposition 4.10, we need the following.

Lemma 4.11.

Let λ\lambda be a tight contact form on S3S^{3}. Let u~:×S1×S3\tilde{u}:\mathbb{R}\times S^{1}\to\mathbb{R}\times S^{3} be a finite energy cylinder asymptotic to nondegenerate simple Reeb orbits P=(x,T)P=(x,T) at ++\infty and P¯=(x¯,T¯)\bar{P}=(\bar{x},\bar{T}) at -\infty. Assume that PP¯P\neq\bar{P}, PP¯P\cup\bar{P} is an unlink and μ(P),μ(P¯){1,2,3}\mu(P),~{}\mu(\bar{P})\in\{1,2,3\}. Then u(×S1)P=u(\mathbb{R}\times S^{1})\cap P=\emptyset and u(×S1)P¯=u(\mathbb{R}\times S^{1})\cap\bar{P}=\emptyset.

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 u~:Γ×S3\tilde{u}:\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} such that πdu\pi\cdot du 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 22-disk 𝒟\mathcal{D} transversal to the periodic orbit at a point x(t)x(t) of PP and tangent to ξ=kerλ\xi=\ker\lambda at x(t)x(t), that is, Tx(t)𝒟=ξx(t)T_{x(t)\mathcal{D}}=\xi_{x(t)}. We orient the disk in such a way that Tx(t)𝒟T_{x(t)\mathcal{D}} and ξx(t)\xi_{x(t)} have the same orientation. Let MM\in\mathbb{R} be such that all the intersection points of uu with PP are contained in u((,M)×S1)u((-\infty,M)\times S^{1}). Let φ:𝔻S3\varphi:\mathbb{D}\to S^{3} be a disk map to P¯\bar{P} such that φ(𝔻)P=\varphi(\mathbb{D})\cap P=\emptyset. Such disk exists since PP¯P\cup\bar{P} is a trivial link. Glue the disk 𝔻\mathbb{D} to [,M]×S1[-\infty,M]\times S^{1} along ×S1-\infty\times S^{1} to form a new disk DD. Let u¯:¯×S1S3\bar{u}:\bar{\mathbb{R}}\times S^{1}\to S^{3} be the map obtained by defining u¯(,t)=x¯(T¯t)\bar{u}(-\infty,t)=\bar{x}(\bar{T}t) and u¯(+,t)=x(Tt)\bar{u}(+\infty,t)=x(Tt) and define U:DS3U:D\to S^{3} by

U|[,M]×S1=u¯,U|𝔻=φ.U|_{[-\infty,M]\times S^{1}}=\bar{u},~{}~{}~{}U|_{\mathbb{D}}=\varphi~{}.

Consider U:H1(D,)H1(S3P,)U_{*}:H_{1}(\partial D,\mathbb{Z})\to H_{1}(S^{3}\setminus P,\mathbb{Z}). If α\alpha is the generator of H1(D,)H_{1}(\partial D,\mathbb{Z}), there exists an integer, called int(u)int(u), such that

U(α)=int(u)[𝒟]H1(S3P,).U_{*}(\alpha)=int(u)[\partial\mathcal{D}]\in H_{1}(S^{3}\setminus P,\mathbb{Z})~{}.

Here we have used the fact that H1(S3P,)H_{1}(S^{3}\setminus P,\mathbb{Z}) is generated by [𝒟][\partial\mathcal{D}].

It follows from the proof of Theorem 4.6 in [17] that int(u)int(u) is the oriented intersection number of UU and PP. Moreover, all the intersections are in the image of the map uu, since there is no intersections of PP with φ(𝔻)\varphi(\mathbb{D}). Following the proof of Theorem 4.6 in [17], one can show that int(u)0int(u)\geq 0 and int(u)=0int(u)=0 if and only if u(×S1)P=u(\mathbb{R}\times S^{1})\cap P=\emptyset. Now we show that int(u)=0int(u)=0.

Let Φ:𝒰S3\Phi:\mathcal{U}\to S^{3} be an embedding of an open neighborhood 𝒰\mathcal{U} of the zero section of ξ|P\xi|_{P}. We require that Φ(0p)=p\Phi(0_{p})=p and the fiberwise derivative of Φ\Phi at 0p0_{p} is the inclusion of ξ\xi into TpS3T_{p}S^{3}. Consider a nonvanishing section v(t)v(t) of ξ\xi along PP which is contained in 𝒰\mathcal{U}. Define the loop β(v)\beta(v) by β(v)(t)=Φv(t)\beta(v)(t)=\Phi\circ v(t) for 0tT0\leq t\leq T. It is contained in S3PS^{3}\setminus P and we denote by [β(v)]H1(S3P)[\beta(v)]\in H_{1}(S^{3}\setminus P) the homology class generated by this loop. Fix the global trivialization Ψ:ξS3×2\Psi:\xi\to S^{3}\times\mathbb{R}^{2}. Let wind(v,Ψ)\operatorname{wind}(v,\Psi) be the winding number of the small section v(t)v(t) with respect to the trivialization Ψ\Psi. It is proved in [17] that

wind(v,Ψ)[𝒟][β(v)]H1(S3P)\operatorname{wind}(v,\Psi)[\partial\mathcal{D}]-[\beta(v)]\in H_{1}(S^{3}\setminus P)

is independent of the section vv as described above. We define a constant c(u)c(u) by

(wind(v,Ψ)c(u))[𝒟]=[β(v)].(\operatorname{wind}(v,\Psi)-c(u))[\partial\mathcal{D}]=[\beta(v)]~{}.

If we choose, for example, the special section v(t)v(t) such that Φv(t)=u(s,t)\Phi\circ v(t)=u(s^{*},t) for some large ss^{*}\in\mathbb{R}, then by definition [β(v)]=int(u)[𝒟][\beta(v)]=int(u)[\partial\mathcal{D}] and taking the limit as ss^{*}\to\infty, we have

wind(u~,)c(u)=int(u).\operatorname{wind}_{\infty}(\tilde{u},\infty)-c(u)=int(u)~{}.

It is proved in [17] that there exists an embedded disk F=φ(𝔻)F=\varphi(\mathbb{D}) with φ(𝔻)=P\varphi(\partial\mathbb{D})=P whose characteristic distribution has e+1e^{+}\geq 1 positive elliptic points, and that

c(u)=2e+1.c(u)=2e^{+}-1~{}~{}.

By Lemma 3.12, we have wind(u~,)=1\operatorname{wind}_{\infty}(\tilde{u},\infty)=1, so that

int(u)=22e+.int(u)=2-2e^{+}~{}.

Since int(u)0int(u)\geq 0 and e+1e^{+}\geq 1, we have int(u)=0int(u)=0 and e+=1e^{+}=1. This shows that u(×S1)P=u(\mathbb{R}\times S^{1})\cap P=\emptyset. We can repeat the arguments replacing PP by P¯\bar{P} to show that u(×S1)P¯=u(\mathbb{R}\times S^{1})\cap\bar{P}=\emptyset and conclude the proof. ∎

Proof of Proposition 4.10.

From Lemma 4.11, we conclude that vrv_{r} does not intersect its asymptotic limits. Thus, v~r\tilde{v}_{r} satisfies condition (2) of Theorem 2.13. We conclude that v~r\tilde{v}_{r} is an embedding and vrv_{r} is an embedding transverse to the Reeb vector field. Applying Theorem 2.12 to the finite energy cylinders v~r:{0}×S3\tilde{v}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and u~r:{0}×S3\tilde{u}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}, where u~r\tilde{u}_{r} is the cylinder obtained in Proposition 3.5, we prove that vrv_{r} does not intersect uru_{r}. Similarly, we prove that vrv_{r} does not intersect uru_{r}^{\prime}. Since P1(S31)P_{1}\subset(S^{3}\setminus\mathcal{R}_{1}), we conclude that vr({0})(S31)v_{r}(\mathbb{C}\setminus\{0\})\subset(S^{3}\setminus\mathcal{R}_{1}). ∎

Proposition 4.12.

Let v~r=(br,vr):{0}×S3\tilde{v}_{r}=(b_{r},v_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} be the J~\tilde{J}-holomorphic cylinder obtained in Proposition 4.9. Then, up to reparametrization and \mathbb{R}-translation, v~r\tilde{v}_{r} is the unique J~\tilde{J}-holomorphic cylinder asymptotic to P2P_{2} at \infty and to P1P_{1} at 0 that does not intersect ×(P1P2)\mathbb{R}\times\left(P_{1}\cup P_{2}\right). Moreover, the cylinder v~r\tilde{v}_{r} and the plane u~q\tilde{u}_{q} obtained in Proposition 3.5 approach P2P_{2} in opposite directions, according to Definition 2.14. Consequently, uq()vr({0})P2u_{q}(\mathbb{C})\cup v_{r}(\mathbb{C}\setminus\{0\})\cup P_{2} is a C1C^{1}-embedded disk.

Proof.

The proof follows [7, Proposition C.1]. Following Theorem 2.7, let ηq\eta_{q} be the asymptotic eigensection of u~q\tilde{u}_{q} at \infty and let ηr\eta_{r} be the asymptotic eigensection of v~r\tilde{v}_{r} at \infty. By Lemma 3.12, we have wind(u~q,)=wind(v~r,)=1\operatorname{wind}_{\infty}(\tilde{u}_{q},\infty)=\operatorname{wind}(\tilde{v}_{r},\infty)=1. Using formula (22) and Proposition 2.2, we conclude that νP2neg\nu_{P_{2}}^{neg} is the unique negative eigenvalue of AP2,JA_{P_{2},J} with winding number 11. By Proposition 2.2, we know that the eigenspace of νP2neg\nu_{P_{2}}^{neg} is one dimensional. Thus, there exists c0c\neq 0 such that ηr=cηq\eta_{r}=c\eta_{q}. By Proposition 4.10 and Theorem 2.15, we conclude that c<0c<0, that is, v~r\tilde{v}_{r} and u~q\tilde{u}_{q} approach P2P_{2} 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 v~r\tilde{v}_{r}. ∎

The proof of Proposition 4.1 is complete.

5. A family of cylinders asymptotic to P3P_{3} and P1P_{1}

Recall that we have obtained a foliation of a closed region 1S3\mathcal{R}_{1}\subset S^{3} and a finite energy cylinder asymptotic to P2P_{2} at its positive puncture and P1P_{1} at its negative puncture, whose projection to S3S^{3} is contained in 2:=S31̊\mathcal{R}_{2}:=S^{3}\setminus\mathring{\mathcal{R}_{1}}. Now we construct a foliation of 2\mathcal{R}_{2}. More precisely, we prove the following statement.

Proposition 5.1.

Let λ\lambda be a tight contact form on S3S^{3} satisfying the hypotheses of Theorem 1.3. Then there exists a family of finite energy cylinders {w~τ=(cτ,wτ):{0}×S3}τ(0,1)\{\tilde{w}_{\tau}=(c_{\tau},w_{\tau}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}\}_{\tau\in(0,1)}, all of them asymptotic to P3P_{3} at their positive punctures z=z=\infty and P1P_{1} at their negative punctures z=0z=0. The projections wτw_{\tau} are embeddings transverse to the Reeb vector field and {Cτ:=wτ({0})}τ(0,1)2(P1P2P3)\{C_{\tau}:=w_{\tau}(\mathbb{C}\setminus\{0\})\}_{\tau\in(0,1)}\subset\mathcal{R}_{2}\setminus(P_{1}\cup P_{2}\cup P_{3}). The union of {Cτ}τ(0,1)\{C_{\tau}\}_{\tau\in(0,1)}, U1U_{1}, U2U_{2} and VV determine a smooth foliation of 2(P1P2P3)\mathcal{R}_{2}\setminus(P_{1}\cup P_{2}\cup P_{3}). Here the surfaces U1U_{1}, U2U_{2} and VV are given by Propositions 3.1 and 4.1.

In Subsection 5.5, we also prove that sl(Pi)=1{\rm{sl}}(P_{i})=-1, i=1,2,3i=1,2,3, completing the proof of Theorem 1.3.

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 P+,P,P𝒫(λ)P_{+},P,P_{-}\in\mathcal{P}(\lambda) be simple Reeb orbits such that μ(P+)=μ(P)+1=μ(P)+2\mu(P_{+})=\mu(P)+1=\mu(P_{-})+2. Let u~+,u~:{0}×S3\tilde{u}_{+},\tilde{u}_{-}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} be finite energy cylinders, which are pseudo-holomorphic with respect to J~𝒥reg\tilde{J}\in\mathcal{J}_{reg} and such that u~+\tilde{u}_{+} is asymptotic to P+P_{+} at the positive puncture z=z=\infty and to PP at the negative puncture z=0z=0 and u~\tilde{u}_{-} is asymptotic to PP at the positive puncture z=z=\infty and to PP_{-} at the negative puncture z=0z=0. Then there exists R0R_{0}\in\mathbb{R} and a family of finite energy cylinders

{u~R:{0}×S3},R[R0,+)\{\tilde{u}_{R}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}\},~{}R\in[R_{0},+\infty)

asymptotic to P+P_{+} at the positive puncture z=z=\infty and to PP_{-} at the negative puncture z=0z=0. For every sequence Rn[R0,+)R_{n}\in[R_{0},+\infty) satisfying RnR_{n}\to\infty, there exist sequences δn±\delta_{n}^{\pm}\in\mathbb{C} and cn±c_{n}^{\pm}\in\mathbb{R} such that

u~R(δn+)+cn+\displaystyle\tilde{u}_{R}(\delta_{n}^{+}\cdot)+c_{n}^{+} u~+in Cloc({0})\displaystyle\to\tilde{u}_{+}~{}~{}\text{in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\})
u~R(δn)+cn\displaystyle\tilde{u}_{R}(\delta_{n}^{-}\cdot)+c_{n}^{-} u~in Cloc({0})\displaystyle\to\tilde{u}_{-}~{}~{}\text{in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\})

5.2. A family of J~\tilde{J}-holomorphic cylinders

Let u~r:{0}×S3\tilde{u}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and v~r:{0}×S3\tilde{v}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} be the finite energy cylinders obtained by Proposition 3.5 and Proposition 4.9 respectively. Applying Theorem 5.2 to u~r\tilde{u}_{r} and v~r\tilde{v}_{r} we obtain a family of finite energy J~\tilde{J}-holomorphic cylinders

(61) {w~τ=(cτ,wτ):{0}×S3},τ[R0,+),\{\tilde{w}_{\tau}=(c_{\tau},w_{\tau}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}\},\tau\in[R_{0},+\infty),

all of them asymptotic to the orbit P3P_{3} at the positive puncture z=z=\infty and P1P_{1} at the negative puncture z=0z=0.

Proposition 5.3.

For every τ[R0,+)\tau\in[R_{0},+\infty), w~τ\tilde{w}_{\tau} is an embedding. The projection wτ:{0}S3w_{\tau}:\mathbb{C}\setminus\{0\}\to S^{3} is an embedding which does not intersect its asymptotic limits and is transverse to the Reeb vector field. Moreover, wτ({0})1=w_{\tau}(\mathbb{C}\setminus\{0\})\cap\mathcal{R}_{1}=\emptyset, where 1\mathcal{R}_{1} is the closed region defined in Proposition 3.17.

Proof.

By Lemma 4.11 and Theorem 2.13, we conclude that w~τ\tilde{w}_{\tau} is an embedding and wτw_{\tau} 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 w~τ\tilde{w}_{\tau} and u~r\tilde{u}_{r}, we conclude that wτw_{\tau} and uru_{r} do not intersect and that wτw_{\tau} does not intersect the orbit P2P_{2}. Similarly, we conclude that wτw_{\tau} and uru_{r}^{\prime} do not intersect, where the J~\tilde{J}-holomorphic cylinder u~r=(ar,wr):{0}×S3\tilde{u}_{r}^{\prime}=(a^{\prime}_{r},w_{r}^{\prime}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} is obtained in Proposition 3.5. Consequently, wτ({0})1=w_{\tau}(\mathbb{C}\setminus\{0\})\cap\mathcal{R}_{1}=\emptyset. ∎

Applying Theorem 3.3 to the maps w~τ\tilde{w}_{\tau} in (61), we obtain a maximal smooth one-parameter family of finite energy cylinders, containing the family (61). Assuming that τ\tau strictly increases in the direction of RλR_{\lambda} and the normalization τ(0,1)\tau\in(0,1), we denote this maximal family by

(62) {w~τ=(cτ,wτ):{0}×S3},τ(0,1).\{\tilde{w}_{\tau}=(c_{\tau},w_{\tau}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}\},~{}\tau\in(0,1).
Proposition 5.4.

Consider a sequence w~n=(cn,wn):{0}×S3\tilde{w}_{n}=(c_{n},w_{n}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} in the family (62), where w~n=w~τn\tilde{w}_{n}=\tilde{w}_{\tau_{n}} and τn1\tau_{n}\to 1^{-}. Let u~r,u~r:{0}×S3\tilde{u}_{r},\tilde{u}_{r}^{\prime}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} and v~r:{0}×S3\tilde{v}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} be the finite energy cylinders obtained by Proposition 3.5 and Proposition 4.9 respectively. Then after suitable reparametrizations and \mathbb{R}-translations of w~n\tilde{w}_{n}, u~r,u~r\tilde{u}_{r},~{}\tilde{u}_{r}^{\prime} and v~r\tilde{v}_{r}, we have

  1. (i)

    up to a subsequence, w~nu~r\tilde{w}_{n}\to\tilde{u}_{r} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  2. (ii)

    There exist sequences δn+0+\delta_{n}^{+}\to 0^{+} and dnd_{n}\in\mathbb{R} such that, up to a subsequence, w~n(δn)+dnv~r\tilde{w}_{n}(\delta_{n}\cdot)+d_{n}\to\tilde{v}_{r} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

A similar statement holds for any sequence τn0+\tau_{n}\to 0^{+}, with u~r\tilde{u}_{r} replaced by u~r\tilde{u}_{r}^{\prime}.

The Proof of Proposition 5.4 is given in Subsections 5.3 and 5.4.

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 w~n=(cn,wn):{0}×S3\tilde{w}_{n}=(c_{n},w_{n}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} in the family (62), where w~n=w~τn\tilde{w}_{n}=\tilde{w}_{\tau_{n}} and τn1\tau_{n}\to 1^{-}. Note that since all cylinders w~n\tilde{w}_{n} are asymptotic to P3P_{3} at z=0z=0, we have 0<E(w~n)=T30<E(\tilde{w}_{n})=T_{3}. We reparametrize the sequence so that

(63) 𝔻wn𝑑λ=σ(T3)2.\int_{\mathbb{C}\setminus\mathbb{D}}w_{n}^{*}d\lambda=\frac{\sigma(T_{3})}{2}~{}.

Define Θ={z{0}| subsequence w~nj and zjz s.t. |dw~nj(zj)|}.\Theta=\{z\in\mathbb{C}\setminus\{0\}|\exists\text{ subsequence }\tilde{w}_{n_{j}}\text{ and }z_{j}\to z\text{ s.t. }|d\tilde{w}_{n_{j}}(z_{j})|\to\infty\}. By the same arguments used in the proof of Proposition 3.7, we can assume that Θ\Theta is finite and Θ𝔻{0}\Theta\subset\mathbb{D}\setminus\{0\}. Moreover, there exists a J~\tilde{J}-holomorphic map

(64) w~=(c,w):({0}Θ)×S3\tilde{w}=(c,w):\mathbb{C}\setminus\left(\{0\}\cup\Theta\right)\to\mathbb{R}\times S^{3}

such that, up to a subsequence, still denoted by w~n\tilde{w}_{n},

w~nw~ in Cloc(({0}Θ),×S3)\tilde{w}_{n}\to\tilde{w}\text{ in }C^{\infty}_{loc}\left(\mathbb{C}\setminus(\{0\}\cup\Theta),\mathbb{R}\times S^{3}\right)

and E(w~)T3E(\tilde{w})\leq T_{3}. The punctures in {0}Θ\{0\}\cup\Theta are non-removable and negative, and the puncture z=z=\infty is positive. Indeed, for any ϵ\epsilon sufficiently large or small, we have

Bϵ(0)wλ=limnBϵ(0)wnλ[T1,T3],\int_{\partial B_{\epsilon}(0)}w^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(0)}w_{n}^{*}\lambda\in[T_{1},T_{3}],

where Bϵ(z)\partial B_{\epsilon}(z) is oriented counterclockwise. It follows that \infty is a positive puncture and 0 is a negative puncture. If zΘz\in\Theta, then for any sufficiently small ϵ\epsilon, we have

Bϵ(z)wλ=limnBϵ(z)wnλ=limnBϵ(z)wn𝑑λT>0,\int_{\partial B_{\epsilon}(z)}w^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(z)}w_{n}^{*}\lambda=\lim_{n\to\infty}\int_{B_{\epsilon}(z)}w_{n}^{*}d\lambda\geq T>0,

where TT3T\leq T_{3} is a period. This follows from the same arguments used in §3.2.1. We conclude that zz is a negative puncture. By the same arguments used in the proof of Proposition 3.5 we conclude that the asymptotic limit of w~\tilde{w} at \infty is P3P_{3}.

Lemma 5.5.

({0}Θ)w𝑑λ>0.\int_{\mathbb{C}\setminus(\{0\}\cup\Theta)}w^{*}d\lambda>0~{}.

Proof.

If Θ=\Theta=\emptyset, then it follows from (63) that {0}w𝑑λσ(T3)2\int_{\mathbb{C}\setminus\{0\}}w^{*}d\lambda\geq\frac{\sigma(T_{3})}{2}. Now assume Θ\Theta\neq\emptyset and suppose, contrary to our claim, that ({0}Θ)w𝑑λ=0\int_{\mathbb{C}\setminus(\{0\}\cup\Theta)}w^{*}d\lambda=0. By Theorem 2.6, there exists a polynomial p:p:\mathbb{C}\to\mathbb{C} and a periodic orbit P𝒫(λ)P\in\mathcal{P}(\lambda) such that p1(0)={0}Θp^{-1}(0)=\{0\}\cup\Theta and w~=FPp\tilde{w}=F_{P}\circ p, where FPF_{P} is the cylinder over the orbit PP. But this implies degp2\deg p\geq 2, contradicting the fact that the asymptotic limit of w~\tilde{w} at \infty is P3P_{3}, that is a simple orbit. ∎

5.3.1. Soft rescaling near 𝒛𝚯\boldsymbol{z\in\Theta}

Assume Θ\Theta\neq\emptyset and take a puncture zΘz\in\Theta. Now we proceed as in the soft rescaling done in §3.2.2. Define the mass m(z)m(z) of zz as in (40). Fix ϵ>0\epsilon>0 satisfying (41) and choose sequences znBϵ(z)¯z_{n}\in\overline{B_{\epsilon}(z)} and 0<δn<ϵ0<\delta_{n}<\epsilon satisfying (42)-(43). It follows that znzz_{n}\to z and, passing to a subsequence, we have δn0\delta_{n}\to 0. Take RnR_{n}\to\infty such that δnRn<ϵ2\delta_{n}R_{n}<\frac{\epsilon}{2} and define

(65) v~n=(bn,vn):BRn(0)\displaystyle\tilde{v}_{n}=(b_{n},v_{n}):B_{R_{n}}(0) ×S3\displaystyle\to\mathbb{R}\times S^{3}
ζ\displaystyle\zeta (cn(zn+δnζ)cn(zn+2δn),wn(zn+δnζ)).\displaystyle\mapsto(c_{n}(z_{n}+\delta_{n}\zeta)-c_{n}(z_{n}+2\delta_{n}),w_{n}(z_{n}+\delta_{n}\zeta)).

The sequence v~n\tilde{v}_{n} is a germinating sequence according to Definition 3.6. Let

(66) Θ1={ζ|ζjζ and subsequence v~nj s.t. |dv~nj(ζj)|}.\Theta_{1}=\{\zeta\in\mathbb{C}|\exists\zeta_{j}\to\zeta\text{ and subsequence }\tilde{v}_{n_{j}}\text{ s.t. }|d\tilde{v}_{n_{j}}(\zeta_{j})|\to\infty\}~{}.

Passing to a subsequence, we can assume that Θ1\Theta_{1} is finite. Let

v~z:Θ1×S3\tilde{v}_{z}:\mathbb{C}\setminus\Theta_{1}\to\mathbb{R}\times S^{3}

be a limit of v~n\tilde{v}_{n} as defined in 3.8. Let PzP_{z} be the asymptotic limit of w~\tilde{w} at zz. Then v~z\tilde{v}_{z} is asymptotic to PzP_{z} at its unique positive puncture \infty. Using Lemma 3.13, we conclude the following.

Lemma 5.6.

If zΘz\in\Theta and w~\tilde{w} is asymptotic to PzP_{z} at zz, then μ(Pz)2\mu(P_{z})\geq 2.

Since P3P_{3} is simple, it follows that w~\tilde{w} is somewhere injective. By Theorem 2.11, we have

1ind(w~)=3z{0}Θμ(Pz)+#Θ1\leq\operatorname{ind}(\tilde{w})=3-\sum_{z\in\{0\}\cup\Theta}\mu(P_{z})+\#\Theta~{}

and consequently

(67) z{0}Θμ(Pz)2+#Θ.\sum_{z\in\{0\}\cup\Theta}\mu(P_{z})\leq 2+\#\Theta~{}.

This proves the following lemma.

Lemma 5.7.

Assume Θ\Theta\neq\emptyset. Then μ(P0)1\mu(P_{0})\leq 1. If μ(P0)=1\mu(P_{0})=1, then Θ={z}\Theta=\{z\} and μ(Pz)=2\mu(P_{z})=2. Here PzP_{z} is the asymptotic limit of w~\tilde{w} at the puncture z{0}Θz\in\{0\}\cup\Theta.

5.3.2. Soft-rescaling near 𝒛=𝟎\boldsymbol{z=0}

For any ϵ>0\epsilon>0, define

(68) mϵ(0)=limnBϵ(0){0}wn𝑑λ,m_{\epsilon}(0)=\lim_{n\to\infty}\int_{B_{\epsilon}(0)\setminus\{0\}}w_{n}^{*}d\lambda,

and define the mass of the puncture z=0z=0 by

(69) m(0)=limϵ0mϵ(0).m(0)=\lim_{\epsilon\searrow 0}m_{\epsilon}(0)~{}.

Note that, for large nn and small ϵ\epsilon, we have

(70) Bϵ(0){0}wn𝑑λ=Bϵ(0)wnλlimδ0Bδ(0)wnλ=Bϵ(0)wnλT1.\int_{B_{\epsilon}(0)\setminus\{0\}}w^{*}_{n}d\lambda=\int_{\partial B_{\epsilon}(0)}w^{*}_{n}\lambda-\lim_{\delta\to 0}\int_{\partial B_{\delta}(0)}w^{*}_{n}\lambda=\int_{\partial B_{\epsilon}(0)}w_{n}^{*}\lambda-T_{1}.

It follows that

(71) m(0)=limϵ0Bϵ(0)wλT1=T0T1,m(0)=\lim_{\epsilon\searrow 0}\int_{\partial B_{\epsilon}(0)}w^{*}\lambda-T_{1}=T_{0}-T_{1},

where T0T_{0} is the period of the asymptotic limit P0=(x0,T0)P_{0}=(x_{0},T_{0}) of w~\tilde{w} at the puncture z=0z=0. We have two cases:

  • either m(0)=0m(0)=0 or

  • m(0)>0m(0)>σ(T3)m(0)>0\Rightarrow m(0)>\sigma(T_{3}).

I

First assume that m(0)>σ(T3)>0m(0)>\sigma(T_{3})>0. We claim that there is a sequence δn0\delta_{n}\to 0 satisfying

(72) Bδn(0){0}wn𝑑λ=m(0)σ(T3)2.\int_{B_{\delta_{n}(0)}\setminus\{0\}}w_{n}^{*}d\lambda=m(0)-\frac{\sigma(T_{3})}{2}.

Indeed, there exists a sequence δn\delta_{n} satisfying the equation above, since using (63) and {0}wn𝑑λ=T3T1m(0)>σ(T3)\int_{\mathbb{C}\setminus\{0\}}w_{n}^{*}d\lambda=T_{3}-T_{1}\geq m(0)>\sigma(T_{3}), we conclude

𝔻{0}wn𝑑λm(0)σ(T3)2>0.\int_{\mathbb{D}\setminus\{0\}}w_{n}^{*}d\lambda\geq m(0)-\frac{\sigma(T_{3})}{2}>0~{}.

Now we show that lim infδn=0\liminf\delta_{n}=0, so that, passing to a subsequence, still denoted by δn\delta_{n}, the claim is true. Suppose that there exists 0<ϵ<lim infδn0<\epsilon^{\prime}<\liminf\delta_{n}. Then we have the contradiction

m(0)σ(T3)2=limjBδn(0){0}wn𝑑λlimjBϵ(0){0}wn𝑑λm(0).m(0)-\frac{\sigma(T_{3})}{2}=\lim_{j\to\infty}\int_{B_{\delta_{n}(0)}\setminus\{0\}}w_{n}^{*}d\lambda\geq\lim_{j\to\infty}\int_{B_{\epsilon}^{\prime}(0)\setminus\{0\}}w^{*}_{n}d\lambda\geq m(0)~{}.

This proves our claim.

Let ϵ0>0\epsilon_{0}>0 be small enough so that the disks Bϵ0(z),zΘ{0}B_{\epsilon_{0}}(z),~{}z\in\Theta\cup\{0\} are disjoint. Define

(73) v~n(z)=(bn(z),vn(z))=(cn(δnz)cn(2δn),wn(δnz))\tilde{v}_{n}(z)=(b_{n}(z),v_{n}(z))=(c_{n}(\delta_{n}z)-c_{n}(2\delta_{n}),w_{n}(\delta_{n}z))

for zBϵ0δn(0){0}z\in B_{\frac{\epsilon_{0}}{\delta_{n}}}(0)\setminus\{0\}. It follows from (68), (69), (72) and (73) that, for large nn and small ϵ0\epsilon_{0}, we have the estimate

(74) Bϵ0δn(0)𝔻vn𝑑λ\displaystyle\int_{B_{\frac{\epsilon_{0}}{\delta_{n}}(0)\setminus\mathbb{D}}}v_{n}^{*}d\lambda =Bϵ0δn(0){0}vn𝑑λ𝔻{0}vn𝑑λ\displaystyle=\int_{B_{\frac{\epsilon_{0}}{\delta_{n}}(0)\setminus\{0\}}}v_{n}^{*}d\lambda-\int_{\mathbb{D}\setminus\{0\}}v_{n}^{*}d\lambda
=Bϵ0(0){0}wn𝑑λ(m(0)σ(T3)2)\displaystyle=\int_{B_{\epsilon_{0}}(0)\setminus\{0\}}w_{n}^{*}d\lambda-\left(m(0)-\frac{\sigma(T_{3})}{2}\right)
m(0)+σ(T3)2(m(0)σ(T3)2)=σ(T3).\displaystyle\leq m(0)+\frac{\sigma(T_{3})}{2}-\left(m(0)-\frac{\sigma(T_{3})}{2}\right)=\sigma(T_{3}).

Define

(75) Θ0={z|zjz and subsequence v~nj s.t. |dv~nj(zj)|}.\Theta_{0}=\{z\in\mathbb{C}|\exists z_{j}\to z\text{ and subsequence }\tilde{v}_{n_{j}}\text{ s.t. }|d\tilde{v}_{n_{j}}(z_{j})|\to\infty\}~{}.

Using the proof of Proposition 3.7 we conclude, passing to a subsequence, that Θ0\Theta_{0} is finite and Θ0𝔻{0}\Theta_{0}\subset\mathbb{D}\setminus\{0\}. Moreover, there exists a J~\tilde{J}-holomorphic map v~0=(b0,v0):{0}Θ0×S3\tilde{v}_{0}=(b_{0},v_{0}):\mathbb{C}\setminus\{0\}\cup\Theta_{0}\to\mathbb{R}\times S^{3} such that, passing to a subsequence

v~nv~0 in Cloc({0}Θ0).\tilde{v}_{n}\to\tilde{v}_{0}\text{ in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}\cup\Theta_{0})~{}.

The map v~0\tilde{v}_{0} is nonconstant, the punctures in {0}Θ0\{0\}\cup\Theta_{0} are non-removable and negative, and the puncture z=z=\infty is positive.

Lemma 5.8.

The asymptotic limit of v~0\tilde{v}_{0} at its unique positive puncture z=z=\infty is equal to P0P_{0}, the asymptotic limit of w~\tilde{w} at {0}\{0\}.

Proof.

Let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be as in the statement of Lemma 3.10. Let PP_{\infty} be the asymptotic limit of v~\tilde{v} at \infty and let 𝒲\mathcal{W}_{\infty} and 𝒲0\mathcal{W}_{0} be connected components of 𝒲\mathcal{W} containing PP_{\infty} and P0P_{0} respectively. Since w~nw~\tilde{w}_{n}\to\tilde{w} in ClocC^{\infty}_{loc}, we can choose 0<ϵ0<ϵ00<\epsilon_{0}^{\prime}<\epsilon_{0} small enough so that, if 0<ρϵ00<\rho\leq\epsilon_{0}^{\prime} is fixed, then the loop tS1wn(ρei2πt)t\in S^{1}\mapsto w_{n}(\rho e^{i2\pi t}) belongs to 𝒲0\mathcal{W}_{0} for large nn. Since v~nv~\tilde{v}_{n}\to\tilde{v} in ClocC^{\infty}_{loc}, we can choose R0>1R_{0}>1 large enough so that, if RR0R\geq R_{0} is fixed, then the loop tS1vn(Rei2πt)=wn(δnRei2πt)t\in S^{1}\mapsto v_{n}(Re^{i2\pi t})=w_{n}(\delta_{n}Re^{i2\pi t}) belongs to 𝒲\mathcal{W}_{\infty} for large nn. By (72), we can show that

(76) e:=lim infBδnR0(0)wnλ>0.e:=\liminf\int_{\partial B_{\delta_{n}R_{0}(0)}}w_{n}^{*}\lambda>0.

Consider, for each nn, the J~\tilde{J}-holomorphic cylinder C~n:[lnRoδn2π,lnϵ02π]×S1×S3\tilde{C}_{n}:\left[\frac{\ln R_{o}\delta_{n}}{2\pi},\frac{\ln\epsilon_{0}^{\prime}}{2\pi}\right]\times S^{1}\to\mathbb{R}\times S^{3}, defined by C~n(s,t)=w~n(e2π(s+it))\tilde{C}_{n}(s,t)=\tilde{w}_{n}(e^{2\pi(s+it)}). It follows from (74) that

(77) [lnRoδn2π,lnϵ02π]×S1Cn𝑑λσ(T3)\int_{\left[\frac{\ln R_{o}\delta_{n}}{2\pi},\frac{\ln\epsilon_{0}^{\prime}}{2\pi}\right]\times S^{1}}C_{n}^{*}d\lambda\leq\sigma(T_{3})

for large nn. Using (76) and (77) and applying Lemma 3.10 as in the proof of Proposition 3.9, we conclude that 𝒲=𝒲0\mathcal{W}_{\infty}=\mathcal{W}_{0} and consequently that P=P0P_{\infty}=P_{0}. ∎

Lemma 5.9.

Either

  • {0}Θ0v0𝑑λ>0\int_{\mathbb{C}\setminus\{0\}\cup\Theta_{0}}v_{0}^{*}d\lambda>0 or

  • {0}Θ0v0𝑑λ=0 and #Θ01\int_{\mathbb{C}\setminus\{0\}\cup\Theta_{0}}v_{0}^{*}d\lambda=0\text{ and }\#\Theta_{0}\geq 1.

Proof.

On the contrary, suppose that {0}Θ0v0𝑑λ=0 and Θ0=\int_{\mathbb{C}\setminus\{0\}\cup\Theta_{0}}v_{0}^{*}d\lambda=0\text{ and }\Theta_{0}=\emptyset. Then

T0=𝔻v0λ\displaystyle T_{0}=\int_{\partial\mathbb{D}}v_{0}^{*}\lambda =limnBδn(0)wnλ\displaystyle=\lim_{n\to\infty}\int_{\partial B_{\delta_{n}}(0)}w_{n}^{*}\lambda
=Bδn(0){0}wnλ+limϵ0Bϵ(0)wnλ\displaystyle=\int_{B_{\delta_{n}}(0)\setminus\{0\}}w_{n}^{*}\lambda+\lim_{\epsilon\to 0}\int_{\partial B_{\epsilon}(0)}w_{n}^{*}\lambda
=m(0)σ(T3)2+T1\displaystyle=m(0)-\frac{\sigma(T_{3})}{2}+T_{1}
=T0σ(T3)2,\displaystyle=T_{0}-\frac{\sigma(T_{3})}{2},

a contradiction. ∎

II

Now assume that m(0)=0m(0)=0. Let ϵ>0\epsilon>0 be small enough so that mϵ(0)σ(T3)2m_{\epsilon}(0)\leq\frac{\sigma(T_{3})}{2}. Define, for any sequence δn0\delta_{n}\to 0,

v~n(z)=(bn(z),vn(z))=(cn(δnz)cn(2δn),wn(δnz)),zBϵδn(0){0}.\tilde{v}_{n}(z)=(b_{n}(z),v_{n}(z))=(c_{n}(\delta_{n}z)-c_{n}(2\delta_{n}),w_{n}(\delta_{n}z)),~{}~{}~{}z\in B_{\frac{\epsilon}{\delta_{n}}}(0)\setminus\{0\}.

Using (68), we conclude that

(78) limnBϵδn(0){0}vn𝑑λ=Bϵ(0){0}wn𝑑λ=mϵ(0)σ(T3)2.\lim_{n\to\infty}\int_{B_{\frac{\epsilon}{\delta_{n}}}(0)\setminus\{0\}}v_{n}^{*}d\lambda=\int_{B_{\epsilon}(0)\setminus\{0\}}w_{n}^{*}d\lambda=m_{\epsilon}(0)\leq\frac{\sigma(T_{3})}{2}~{}.

It follows that Θ0=\Theta_{0}=\emptyset, where Θ0\Theta_{0} is defined as in (75). Indeed, if zΘ0z\in\Theta_{0}, arguing as in the proof of Proposition 3.7, we find sequences rj0+r_{j}\to 0^{+}, zjzz_{j}\to z and nj+n_{j}\to+\infty such that limnBrj(zj)vn𝑑λT,\lim_{n\to\infty}\int_{B_{r_{j}}(z_{j})}v_{n}^{*}d\lambda\geq T, for some period TT, contradicting (78). Thus, passing to a subsequence, still denoted v~n\tilde{v}_{n}, there exists a J~\tilde{J}-holomorphic map v~0=(b0,v0):{0}×S3\tilde{v}_{0}=(b_{0},v_{0}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} such that v~nv~0 in Cloc({0}).\tilde{v}_{n}\to\tilde{v}_{0}\text{ in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\})~{}.

Lemma 5.10.

If m(0)=0m(0)=0, the curve v~0\tilde{v}_{0} is a trivial cylinder over the orbit P1P_{1}. Moreover, w~\tilde{w} is asymptotic to P1P_{1} at its negative puncture z=0z=0.

Proof.

It follows from (78) that {0}v0𝑑λ=0.\int_{\mathbb{C}\setminus\{0\}}v_{0}^{*}d\lambda=0~{}. Indeed, if {0}v0𝑑λ>0\int_{\mathbb{C}\setminus\{0\}}v_{0}^{*}d\lambda>0, then {0}v0𝑑λ>σ(T3)\int_{\mathbb{C}\setminus\{0\}}v_{0}^{*}d\lambda>\sigma(T_{3}), which contradicts (78). Note that

(79) 𝔻v0λ\displaystyle\int_{\partial\mathbb{D}}v_{0}^{*}\lambda =limn𝔻vnλ\displaystyle=\lim_{n\to\infty}\int_{\partial\mathbb{D}}v_{n}^{*}\lambda
=limnBδn(0)wnλ\displaystyle=\lim_{n\to\infty}\int_{\partial B_{\delta_{n}}(0)}w^{*}_{n}\lambda
=limn(Bϵ(0)wnλBϵ(0)Bδn(0)wnλ).\displaystyle=\lim_{n\to\infty}\left(\int_{\partial B_{\epsilon}(0)}w^{*}_{n}\lambda-\int_{B_{\epsilon}(0)\setminus B_{\delta_{n}}(0)}w_{n}^{*}\lambda\right).

Using (70) we conclude that

(80) 0limnBϵ(0)Bδn(0)wnλlimnBϵ(0){0}wnλ=Bϵ(0)wλT1.0\leq\lim_{n\to\infty}\int_{B_{\epsilon}(0)\setminus B_{\delta_{n}}(0)}w_{n}^{*}\lambda\leq\lim_{n\to\infty}\int_{B_{\epsilon}(0)\setminus\{0\}}w_{n}^{*}\lambda=\int_{\partial B_{\epsilon}(0)}w^{*}\lambda-T_{1}.

It follows from (79), (80) and (71) that

(81) 𝔻v0λ=limn(BϵwnλBϵBδnwnλ)=T1.\int_{\partial\mathbb{D}}v_{0}^{*}\lambda=\lim_{n\to\infty}\left(\int_{\partial B_{\epsilon}}w^{*}_{n}\lambda-\int_{B_{\epsilon}\setminus B_{\delta_{n}}}w_{n}^{*}\lambda\right)=T_{1}.

We conclude that v~0\tilde{v}_{0} is a cylinder over a T1T_{1}-periodic orbit P=(x,T1)P=(x,T_{1}).

Now we prove that P=P1=P0P=P_{1}=P_{0}, where P0=(x0,T0)P_{0}=(x_{0},T_{0}) is the asymptotic limit of w~\tilde{w} at the puncture z=0z=0. Let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be as in the statement of Lemma 3.10. For fixed nn, we know that (twn(ϵe2πit))x1(T1t) as ϵ0.(t\mapsto w_{n}(\epsilon e^{2\pi it}))\to x_{1}(T_{1}t)\text{ as }\epsilon\to 0~{}. Thus, we can choose a sequence δn0\delta_{n}\to 0 such that

(twn(δne2πit)=vn(e2πit))𝒲1,n,(t\mapsto w_{n}(\delta_{n}e^{2\pi it})=v_{n}(e^{2\pi it}))\in\mathcal{W}_{1},~{}~{}\forall n~{},

where 𝒲1\mathcal{W}_{1} is the connected component of 𝒲\mathcal{W} containing P1P_{1}. We conclude that P=P1P=P_{1}. Since (tw(ϵe2πit))x0(T0t) as ϵ0,(t\mapsto w(\epsilon e^{2\pi it}))\to x_{0}(T_{0}t)~{}\text{ as }\epsilon\to 0, we conclude, from estimate (78) and Lemma 3.10, arguing as in the proof of Proposition 3.9, that P=P0P=P_{0}. ∎

This completes the analysis of the case m(0)=0m(0)=0.

Going back to the case m(0)>σ(T3)>0m(0)>\sigma(T_{3})>0, if the mass of the puncture z=0z=0 of v~0\tilde{v}_{0} is positive or Θ00\Theta_{0}\neq 0, 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 P1P_{1}, originating from the puncture z=0z=0.

5.4. Proof of Proposition 5.4

The finite energy curve w~:(Θ{0})×S3\tilde{w}:\mathbb{C}\setminus(\Theta\cup\{0\})\to\mathbb{R}\times S^{3} defined by (64) is asymptotic to P3P_{3} at its positive puncture z=z=\infty and to an orbit P0=(x0,T0)P_{0}=(x_{0},T_{0}) at the negative puncture z=0z=0. If Θ\Theta\neq\emptyset, the punctures in Θ\Theta are negative.

One of the leaves of the bubbling-off tree obtained from the sequence w~n\tilde{w}_{n} is a cylinder over the orbit P1P_{1} originated from the puncture z=0z=0. Using the fact that μ(P1)=1\mu(P_{1})=1 and an argument similar to Claim I in the proof of Theorem 4.9, we conclude that

(82) μ(P0)1.\mu(P_{0})\geq 1.

We first show that Θ=\Theta=\emptyset. On the contrary, suppose that Θ\Theta\neq\emptyset. Then by Lemma 5.7, we have μ(P0)=1\mu(P_{0})=1, #Θ=1\#\Theta=1 and μ(Pz)=2\mu(P_{z})=2, where Pz=(xz,Tz)P_{z}=(x_{z},T_{z}) is the asymptotic limit at the unique puncture zΘz\in\Theta. The orbit PzP_{z} is not linked to P3P_{3}. Indeed, for each nn, wnw_{n} is an embedding whose image does not intersect P3P_{3}, which implies that any contractible loop in wn({0})w_{n}(\mathbb{C}\setminus\{0\}) is not linked to P3P_{3} as well. It follows that any loop in the image of ww near the end zΘz\in\Theta is not linked to P3P_{3} and consequently that PzP_{z} is not linked to P3P_{3}. Since Tz<T3T_{z}<T_{3}, by hypothesis we have Pz=P2P_{z}=P_{2}. This implies that P2P_{2} is contractible in 2\mathcal{R}_{2}. Since the region 1\mathcal{R}_{1} contains an embedded disk with boundary P2P_{2} and consequently P2P_{2} is contractible in 1\mathcal{R}_{1}, it is easy to prove, using Mayer-Vietoris sequence, that the holomology class of /tx2(T2t)\mathbb{R}/\mathbb{Z}\ni t\to x_{2}(T_{2}t) generates H1(2,)H_{1}(\mathcal{R}_{2},\mathbb{Z}), a contradiction.

Now we show that the mass m(0)m(0) of the puncture z=0z=0, defined by (69), is positive. On the contrary, suppose that m(0)=0m(0)=0. By Lemma 5.10, we know that w~\tilde{w} is asymptotic to P1P_{1} at its negative puncture z=0z=0. This contradicts the fact that the family of cylinders (62) is maximal.

So far, we know that w~:{0}×S3\tilde{w}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} is a J~\tilde{J}-holomorphic cylinder asymptotic to P3P_{3} at \infty and to the orbit P0P_{0} at 0. By (67) and (82), we have 1μ(P0)2.1\leq\mu(P_{0})\leq 2.

The second level of the bubbling-off tree obtained from the sequence w~n\tilde{w}_{n} consists of a unique vertex associated to a finite energy sphere v~0:{0}Θ0×S3\tilde{v}_{0}:\mathbb{C}\setminus\{0\}\cup\Theta_{0}\to\mathbb{R}\times S^{3}, asymptotic to P0P_{0} at its positive puncture \infty. Observe that since the orbit P1P_{1} is not linked to P3P_{3} and, for each nn, P3P_{3} does not intersect the image of wnw_{n}, we know that any loop in the image of wnw_{n} is also not linked to P3P_{3}. It follows that any loop in the image of w~\tilde{w} is not linked to P3P_{3} and consequently, that P0P_{0} is also not linked to P3P_{3}.

Claim I: μ(P0)=2\mu(P_{0})=2.

To prove the claim, suppose by contradiction that μ(P0)=1\mu(P_{0})=1. If v~0\tilde{v}_{0} is somewhere injective, using Theorem 2.11, we have

01μ(P0v)zΘ0μ(Pzv)+#Θ0,0\leq 1-\mu(P_{0}^{v})-\sum_{z\in\Theta_{0}}\mu(P_{z}^{v})+\#\Theta_{0},

where PzvP_{z}^{v} is the asymptotic limit of v~0\tilde{v}_{0} at the puncture zΘ0{0}z\in\Theta_{0}\cup\{0\}. Using the fact that μ(P1)=1\mu(P_{1})=1 and an argument similar to Claim I in the proof of Theorem 4.9, we conclude that μ(P0v)1\mu(P_{0}^{v})\geq 1. If zΘ0z\in\Theta_{0}, by Lemma 3.13, we have μ(Pzv)2\mu(P_{z}^{v})\geq 2. We conclude that #Θ0=0\#\Theta_{0}=0 and πdv00\pi\cdot dv_{0}\equiv 0, which contradicts Lemma 5.9. If v~0\tilde{v}_{0} is not somewhere injective, there exists a somewhere injective J~\tilde{J}-holomorphic curve u~0:Γ×S3\tilde{u}_{0}:\mathbb{C}\setminus\Gamma\to\mathbb{R}\times S^{3} and a polynomial p:p:\mathbb{C}\to\mathbb{C} such that v~0=u~0p\tilde{v}_{0}=\tilde{u}_{0}\circ p, p(Θ0{0})=Γp(\Theta_{0}\cup\{0\})=\Gamma and p1(Γ)=Θ0{0}p^{-1}(\Gamma)=\Theta_{0}\cup\{0\}. This implies that P0=PdegpP_{0}=P_{\infty}^{\deg p}, where PP_{\infty} is the asymptotic limit of u~0\tilde{u}_{0} at \infty. Using Lemma 2.1 and the assumption μ(P0)=1\mu(P_{0})=1, we conclude that μ(P)=1\mu(P_{\infty})=1. For every zΘ0{0}z\in\Theta_{0}\cup\{0\}, we have Pzv=(Pwu)kP_{z}^{v}=(P^{u}_{w})^{k}, where p(z)=wp(z)=w, kdegpk\mid\deg p and PwuP_{w}^{u} is the asymptotic limit of u~0\tilde{u}_{0} at the puncture ww. Since μ(Pzv)1,zΘ0{0}\mu(P_{z}^{v})\geq 1,~{}\forall z\in\Theta_{0}\cup\{0\}, using Lemma 2.1 we conclude that μ(Pzu)1,zΓ\mu(P_{z}^{u})\geq 1,~{}\forall z\in\Gamma. Applying Theorem 2.11 to the curve u~0\tilde{u}_{0} we have

01zΓμ(Pzu)+#Γ1,0\leq 1-\sum_{z\in\Gamma}\mu(P_{z}^{u})+\#\Gamma-1,

where PzuP_{z}^{u} is the asymptotic limit of u~0\tilde{u}_{0} at the puncture zΓz\in\Gamma. It follows that μ(Pzu)=1,zΓ\mu(P_{z}^{u})=1,~{}\forall z\in\Gamma. By Theorem 2.11, we have πdu00\pi\cdot du_{0}\equiv 0, which implies πdv00\pi\cdot dv_{0}\equiv 0. It follows from Theorem 2.6 that v~0=FPp\tilde{v}_{0}=F_{P}\circ p, where FPF_{P} is a cylinder over an orbit P𝒫(λ)P\in\mathcal{P}(\lambda), p:p:\mathbb{C}\to\mathbb{C} is a polynomial and P0=PdegpP_{0}=P^{\deg p}. Since μ(P0)=1\mu(P_{0})=1, we conclude, using Lemma 2.1, that μ(P)=1\mu(P)=1. By Lemma 5.9, we have Θ0\Theta_{0}\neq\emptyset. Let zΘ0z\in\Theta_{0}. Then μ(Pzv)2\mu(P_{z}^{v})\geq 2 and Pzv=PkP_{z}^{v}=P^{k}, where kdegpk\mid\deg p. This implies that 2μ(Pk)μ(Pdegp)=12\leq\mu(P^{k})\leq\mu(P^{\deg p})=1, a contradiction. This proves our claim.

Since T0<T3T_{0}<T_{3} and P0P_{0} is not linked to P3P_{3}, by hypothesis we have P0=P2.P_{0}=P_{2}.

Now we prove that the mass m(0)m(0) of the puncture z=0z=0 of v~0\tilde{v}_{0} is zero, according to definition (69). This implies that v~0\tilde{v}_{0} is asymptotic to P1P_{1} at its unique negative puncture z=0z=0. Suppose, by contradiction, that m(0)>0m(0)>0. Since P2P_{2} is simple, we know that v~0\tilde{v}_{0} is somewhere injective. Applying Theorem 2.11 to v~0\tilde{v}_{0}, we have

0ind(v~0)=2μ(P0v)zΘ0μ(Pzv)+#Θ0.0\leq\operatorname{ind}(\tilde{v}_{0})=2-\mu(P_{0}^{v})-\sum_{z\in\Theta_{0}}\mu(P^{v}_{z})+\#\Theta_{0}.

If ind(v~0)=0\operatorname{ind}(\tilde{v}_{0})=0, then πdv00\pi\cdot dv_{0}\equiv 0. By Theorem 2.6, we conclude that v~0\tilde{v}_{0} is a cylinder over P2P_{2}, contradicting Lemma 5.9. Therefore, ind(v~0)1\operatorname{ind}(\tilde{v}_{0})\geq 1 and we have

1μ(P0v)1zΘ0μ(Pzv)+#Θ0.1\leq\mu(P_{0}^{v})\leq 1-\sum_{z\in\Theta_{0}}\mu(P^{v}_{z})+\#\Theta_{0}.

Thus #Θ0=\#\Theta_{0}=\emptyset and μ(P0v)=1\mu(P_{0}^{v})=1. Following the same arguments used to prove Claim I, we get a contradiction with m(0)>0m(0)>0. We have proved that v~0\tilde{v}_{0} is asymptotic to P1P_{1} at z=0z=0.

We claim that Θ0=\Theta_{0}=\emptyset. On the contrary, suppose that Θ0\Theta_{0}\neq\emptyset. Since v~0\tilde{v}_{0} is somewhere injective and πdv0\pi\circ dv_{0} is not identically zero, by Theorem 2.11 we have

1ind(v~0)=2μ(P1)zΘ0μ(Pzv)+#Θ0=1zΘ0μ(Pzv)+#Θ0.1\leq\operatorname{ind}(\tilde{v}_{0})=2-\mu(P_{1})-\sum_{z\in\Theta_{0}}\mu(P^{v}_{z})+\#\Theta_{0}=1-\sum_{z\in\Theta_{0}}\mu(P^{v}_{z})+\#\Theta_{0}.

This implies that μ(Pzv)1,zΘ0\mu(P^{v}_{z})\leq 1,~{}\forall z\in\Theta_{0}, contradicting 3.13.

So far, we have proved the following statement.

Proposition 5.11.

Consider a sequence w~n=(cn,wn):{0}×S3\tilde{w}_{n}=(c_{n},w_{n}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} in the family (62), where w~n=w~τn\tilde{w}_{n}=\tilde{w}_{\tau_{n}} and τn1\tau_{n}\to 1^{-}. There exists a cylinder w~:{0}×S3\tilde{w}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at its positive puncture \infty and to P2P_{2} at its negative puncture 0 and a cylinder v~0:{0}×S3\tilde{v}_{0}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P2P_{2} at its positive puncture \infty and to P1P_{1} at its negative puncture 0, such that, after suitable reparametrizations and \mathbb{R}-translations of w~n\tilde{w}_{n}, we have

  1. (i)

    up to a subsequence, w~nw~\tilde{w}_{n}\to\tilde{w} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  2. (ii)

    There exist sequences δn+0+\delta_{n}^{+}\to 0^{+} and dnd_{n}\in\mathbb{R} such that, up to a subsequence, w~n(δn)+dnv~0\tilde{w}_{n}(\delta_{n}\cdot)+d_{n}\to\tilde{v}_{0} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

A similar statement holds for any sequence τn0+\tau_{n}\to 0^{+} with w~\tilde{w} replaced with a cylinder w~\tilde{w}^{\prime} with the same asymptotics as w~\tilde{w} and v~0\tilde{v}_{0} replaced with a cylinder v~0\tilde{v}_{0}^{\prime} with the same asymptotics as v~0\tilde{v}_{0}.

Proposition 5.12.

Let w~n\tilde{w}_{n}, w~\tilde{w} and v~0\tilde{v}_{0} be as in Proposition 5.11. Then

  1. (i)

    Given an S1S^{1}-invariant neighborhood 𝒲3C(/,S3)\mathcal{W}_{3}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx3(T3t)t\mapsto x_{3}(T_{3}t), there exists R3>>1R_{3}>>1 such that, for RR3R\geq R_{3} and large nn, the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲3\mathcal{W}_{3}.

  2. (ii)

    Given an S1S^{1}-invariant neighborhood 𝒲2C(/,S3)\mathcal{W}_{2}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx2(T2t)t\mapsto x_{2}(T_{2}t), there exist ϵ2>0\epsilon_{2}>0 and R2>1R_{2}>1 such that the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}, for R2δnRϵ2R_{2}\delta_{n}\leq R\leq\epsilon_{2} and large nn.

  3. (iii)

    Given an S1S^{1}-invariant neighborhood 𝒲1C(/,S3)\mathcal{W}_{1}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx1(T1t)t\mapsto x_{1}(T_{1}t), there exist ϵ1>0\epsilon_{1}>0 such that the loop twn(ρe2πit)t\mapsto w_{n}(\rho e^{2\pi it}) belongs to 𝒲1\mathcal{W}_{1}, for ρϵ1δn\rho\leq\epsilon_{1}\delta_{n} and large nn.

  4. (iv)

    Given any neighborhood 𝒱2\mathcal{V}\subset\mathcal{R}_{2} of w({0})v0(0)P1P2P3w(\mathbb{C}\setminus\{0\})\cup v_{0}(\mathbb{C}\setminus{0})\cup P_{1}\cup P_{2}\cup P_{3}, we have wn({0})𝒱w_{n}(\mathbb{C}\setminus\{0\})\subset\mathcal{V}, for large nn.

A similar statement holds for any sequence w~τn\tilde{w}_{\tau_{n}} such that τn0+\tau_{n}\to 0^{+}, with ww replaced by ww^{\prime} and v0v_{0} replaced by v0v_{0}^{\prime}.

Proof.

We can assume that 𝒲i,i=1,2,3\mathcal{W}_{i},i=1,2,3, contains only the periodic orbit txi(Ti)t\mapsto x_{i}(T_{i}\cdot) modulo S1S^{1}-reparametrizations. Let 𝒲C(/,S3)\mathcal{W}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) be as in the statement of Lemma 3.10 and such that 𝒲1𝒲2𝒲3𝒲\mathcal{W}_{1}\cup\mathcal{W}_{2}\cup\mathcal{W}_{3}\subset\mathcal{W}.

Using the normalization condition (63) we can apply Lemma 3.10 and find R3>>1R_{3}>>1 such that for RR3R\geq R_{3}, the loops {twn(Rei2πt)},n\{t\mapsto w_{n}(Re^{i2\pi t})\},n\in\mathbb{N} belong to 𝒲\mathcal{W}. By the asymptotic behavior of the cylinders w~n\tilde{w}_{n}, we conclude that for any RR3R\geq R_{3} and large nn, the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲3\mathcal{W}_{3}. Assertion (i) is proved.

Now we prove (ii). Recall that the asymptotic limit of w~\tilde{w} at z=0z=0 is the orbit P2P_{2}. We can apply 3.10 as in the proof of Proposition 5.8 to find ϵ2>0\epsilon_{2}>0 small and R2>>1R_{2}>>1 such that for every RR satisfying δnR2Rϵ2\delta_{n}R_{2}\leq R\leq\epsilon_{2} and large nn, the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}.

To prove (iii), first recall that the mass of the puncture z=0z=0 of the sequence v~n\tilde{v}_{n} is zero. Applying 3.10 as in the proof of Lemma 5.10, we find ϵ1>0\epsilon_{1}>0 small and a sequence δn0\delta^{\prime}_{n}\to 0 such that twn(δnρe2πit)=vn(ρe2πit)𝒲1t\mapsto w_{n}(\delta_{n}\rho^{\prime}e^{2\pi it})=v_{n}(\rho^{\prime}e^{2\pi it})\in\mathcal{W}_{1} for δnρϵ1\delta^{\prime}_{n}\leq\rho^{\prime}\leq\epsilon_{1}. The sequence vn(δn)v_{n}(\delta_{n}^{\prime}\cdot) converges in ClocC^{\infty}_{loc} to the cylinder over P1P_{1}, so that applying Lemma 3.10 again, we conclude that for large nn, the loop tvn(δnre2πit)t\mapsto v_{n}(\delta_{n}^{\prime}re^{2\pi it}) belongs to 𝒲1\mathcal{W}_{1} for every r1r\leq 1. We conclude that for any ρδnϵ1\rho\leq\delta_{n}\epsilon_{1} and nn large enough, the loop twn(ρe2πit)t\mapsto w_{n}(\rho e^{2\pi it}) belongs to 𝒲1\mathcal{W}_{1}.

The proof of (iv) follows from (i), (ii), (iii) and Proposition 5.11.∎

Proposition 5.13.

Let w~\tilde{w}, v~0\tilde{v}_{0}, w~\tilde{w}^{\prime} and v~0\tilde{v}_{0}^{\prime} be as in Proposition 5.11. Let u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} be the finite energy cylinders obtained in Proposition 3.5 and let v~r\tilde{v}_{r} be the finite energy cylinder obtained in Proposition 4.9. Then, up to reparametrization and \mathbb{R}-translation, v~0=v~0=v~r\tilde{v}_{0}=\tilde{v}_{0}^{\prime}=\tilde{v}_{r}, w~=u~r\tilde{w}=\tilde{u}_{r} and w~=u~r\tilde{w}^{\prime}=\tilde{u}_{r}^{\prime}.

Proof.

By positivity and stability of intersection of pseudo-holomorphic curves and the fact that the cylinders w~τ\tilde{w}_{\tau} in the family (62) do not intersect ×(P1P2P3)\mathbb{R}\times\left(P_{1}\cup P_{2}\cup P_{3}\right), we conclude that the cylinders w~\tilde{w}, v~0\tilde{v}_{0}, w~\tilde{w}^{\prime} and v~0\tilde{v}_{0}^{\prime} do not intersect ×(P1P2P3)\mathbb{R}\times\left(P_{1}\cup P_{2}\cup P_{3}\right). It follows from Proposition 4.12 that v~0=v~0=v~r\tilde{v}_{0}=\tilde{v}_{0}^{\prime}=\tilde{v}_{r}, up to reparametrization and \mathbb{R}-translation. It follows from Proposition 3.16 that w~\tilde{w} is either equal to u~r\tilde{u}_{r} or u~r\tilde{u}_{r}^{\prime}, up to reparametrization and \mathbb{R}-translation. Similarly, w~\tilde{w}^{\prime} is either equal to u~r\tilde{u}_{r} or u~r\tilde{u}_{r}^{\prime}. Since the family of cylinders (61) was obtained by Theorem 5.2 applied to u~r\tilde{u}_{r} and v~r\tilde{v}_{r}, we conclude, using Proposition 5.12, that w~=u~r\tilde{w}=\tilde{u}_{r}.

It remains to prove that w~=u~r\tilde{w}^{\prime}=\tilde{u}_{r}^{\prime}. Fix τ0(0,1)\tau_{0}\in(0,1). By Jordan-Brouwer separation theorem, the surface

S:=wτ0({0})ur({0})vr({0})P1P2P3S:=w_{\tau_{0}}(\mathbb{C}\setminus\{0\})\cup u_{r}(\mathbb{C}\setminus\{0\})\cup v_{r}(\mathbb{C}\setminus\{0\})\cup P_{1}\cup P_{2}\cup P_{3}

divides S3/SS^{3}/S into two disjoint regions with boundary SS. One of these regions, that we call AA, is contained in 2\mathcal{R}_{2}, since SS does not intersect any of the curves foliating the interior of the region 1\mathcal{R}_{1}.

We will show that AA is foliated by cylinders in the family {wτ}\{w_{\tau}\}. Let pAp\in A and let 𝒱2\mathcal{V}\subset\mathcal{R}_{2} be a small neighborhood of SS in 2\mathcal{R}_{2} separating pp and SS. By Proposition 5.12, for τ1>τ0\tau_{1}>\tau_{0} sufficiently close to 11, we have wτ1({0})𝒱w_{\tau_{1}}(\mathbb{C}\setminus\{0\})\subset\mathcal{V}. Moreover, wτ1({0})𝒱Aw_{\tau_{1}}(\mathbb{C}\setminus\{0\})\subset\mathcal{V}\cap A, since there are points in the image of the family wτw_{\tau} converging to points in a compact subset of ur({0})u_{r}(\mathbb{C}\setminus\{0\}), as τ1\tau\to 1^{-}. Let BAB\subset A be the region limited by wτ0({0})P3P1wτ1({0})w_{\tau_{0}}(\mathbb{C}\setminus\{0\})\cup P_{3}\cup P_{1}\cup w_{\tau_{1}}(\mathbb{C}\setminus\{0\}). Thus, the image of the family wτw_{\tau}, τ(τ0,τ1)\tau\in(\tau_{0},\tau_{1}) is open, closed and nonempty in BB. It follows that pp is in the image of wτw_{\tau} for some τ(τ0,τ1)\tau\in(\tau_{0},\tau_{1}). We conclude that AA is foliated by the images of the cylinders {wτ}\{w_{\tau}\}, τ(τ0,1)\tau\in(\tau_{0},1).

Since every neighborhood of a compact set of ur({0})u_{r}(\mathbb{C}\setminus\{0\}) in 2\mathcal{R}_{2} is contained in A¯\bar{A}, the family {wτ}\{w_{\tau}\}, τ(τ0,1)\tau\in(\tau_{0},1), foliates AA and the cylinders in the family {wτ}\{w_{\tau}\} do not intersect each other, we conclude that w~=u~r\tilde{w}^{\prime}=\tilde{u}_{r}^{\prime}. ∎

The proof of Proposition 5.4 is complete.

5.5. Conclusion of the proof of Theorem 1.3

Proposition 5.14.

Let u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} be the finite energy cylinders obtained in Proposition 3.5 and let v~r\tilde{v}_{r} be the finite energy cylinder obtained in Proposition 4.9. The images of the family {wτ},τ(0,1)\{w_{\tau}\},\tau\in(0,1) given by (62), uru_{r}, uru_{r}^{\prime}, and vrv_{r} determine a smooth foliation of 2(P1P2P3)\mathcal{R}_{2}\setminus(P_{1}\cup P_{2}\cup P_{3}). Here 2=S31¯\mathcal{R}_{2}=\overline{S^{3}\setminus\mathcal{R}_{1}}, where 1\mathcal{R}_{1} is the region obtained in Proposition 3.17.

Proof.

Arguing as in the proof of Proposition 5.13, we conclude that for τ0(0,1)\tau_{0}\in(0,1) fixed, {wτ({0})},τ(τ0,1)\{w_{\tau}(\mathbb{C}\setminus\{0\})\},~{}\tau\in(\tau_{0},1) determine a foliation of a region AA in 2\mathcal{R}_{2} bounded by

S:=wτ0({0})ur({0})vr({0})P1P2P3.S:=w_{\tau_{0}}(\mathbb{C}\setminus\{0\})\cup u_{r}(\mathbb{C}\setminus\{0\})\cup v_{r}(\mathbb{C}\setminus\{0\})\cup P_{1}\cup P_{2}\cup P_{3}.

Repeating the arguments in the proof of Proposition 5.13, with uru_{r} replaced by uru_{r}^{\prime}, we find a foliation of the complement of AA in 2\mathcal{R}_{2}. ∎

The proof of Proposition 5.1 is complete. Now we complete the proof of Theorem 1.3.

Proposition 5.15.

sl(Pi)=1{\rm sl}(P_{i})=-1, for i=1,2,3i=1,2,3.

Proof.

First we prove that sl(P3)=1{\rm{sl}}(P_{3})=-1. Our proof is adapted from [25, §6]. Let F=FτF=F_{\tau} be one of the disks with boundary P3P_{3} obtained in Proposition 3.1. Consider the orientation of S3S^{3} induced by λdλ\lambda\wedge d\lambda and let oo be the orientation on F¯\bar{F} induced by RλR_{\lambda}. The characteristic distribution of F¯\bar{F}, defined by (TF¯ξ),(T\bar{F}\cap\xi)^{\bot}, where \bot means the dλd\lambda-symplectic orthogonal, can be parametrized by a smooth vector field XX on F¯\bar{F}, which is transverse to F\partial F pointing outwards. After a CC^{\infty} perturbation of FF away from a neighborhood of F\partial F and keeping the transversality of RλR_{\lambda} and F{F}, we can assume that all singular points of XX are nondegenerate. More details about these facts can be found in [17] and [16]. Let oo^{\prime} be the orientation on ξ\xi given by dλ|ξd\lambda|_{\xi}. A zero pp of XX is called positive if oo coincides with oo^{\prime} at pp, and negative otherwise. Since RλR_{\lambda} is transverse to F{F}, all singularities of XX have the same sign and they must be positive since F𝑑λ=P2λ>0\int_{F}d\lambda=\int_{P_{2}}\lambda>0. Considering XX as a section of the bundle ξF¯\xi\to\bar{F}, we have

(83) wind(X|F,Ψ|F¯)=X(z)=0sign(DX(z):(TzF,oz)(ξz,oz)),\operatorname{wind}(X|_{\partial F},\Psi|_{\bar{F}})=\sum_{X(z)=0}{\rm sign}\left(DX(z):(T_{z}F,o_{z})\to(\xi_{z},o^{\prime}_{z})\right),

where Ψ\Psi is a trivialization of ξS3\xi\to S^{3}. Let ΨF:TF¯F¯×2\Psi_{F}:T\bar{F}\to\bar{F}\times\mathbb{R}^{2} be a trivialization of TF¯F¯T\bar{F}\to\bar{F}. Considering XX as a section of the bundle TF¯F¯T\bar{F}\to\bar{F} and using the fact that XX points outwards at F\partial F, we have

(84) 1=wind(X|F,ΨF)=X(z)=0sign(DX(z):(TzF,oz)(TzF,oz)).1=\operatorname{wind}(X|_{\partial F},\Psi_{F})=\sum_{X(z)=0}{\rm sign}\left(DX(z):(T_{z}F,o_{z})\to(T_{z}F,o_{z})\right).

Since all singularities of XX are positive and DX(z)DX(z) at a singularity zz of XX does not depend on whether we consider XX as a section of ξF¯\xi\to\bar{F} or as a section of TF¯F¯T\bar{F}\to\bar{F}, the right side of equations (83) and (84) coincide.

Consider the nonvanishing section of ξF¯\xi\to\bar{F} defined by Z(z)=Ψ1(z,(1,0))Z(z)=\Psi^{-1}(z,(1,0)). After a small perturbation of P3P_{3} in the direction of ZZ, which we can assume to be transverse to F{F} and ξ\xi, we find a new closed curve P~3\tilde{P}_{3}. Consider a trivialization Ψ~\tilde{\Psi} of (ξ|F,o)(\xi|_{\partial F},o) induced by the nonvanishing section X|FX|_{\partial F}. We have

sl(P3)=P~3F¯=wind(Z,Ψ~)=wind(X|F,Ψ)=1.{\rm sl}(P_{3})=\tilde{P}_{3}\cdot\bar{F}=\operatorname{wind}(Z,\tilde{\Psi})=-\operatorname{wind}(X|_{\partial F},\Psi)=-1.

The proof of sl(P2)=1{\rm sl}(P_{2})=-1 is completely analogous, replacing FF with the disk DD given by Proposition 3.1.

Now we prove that sl(P1)=1{\rm sl}(P_{1})=-1. Let VV be the cylinder with boundary P1P2P_{1}\cup P_{2} given by Proposition 4.1. Let oo be the orientation on V¯\bar{V} induced by RλR_{\lambda}. The characteristic distribution of V¯\bar{V} can be parametrized by a smooth vector field X~\tilde{X} on V¯\bar{V}, which is transverse to V\partial V pointing outwards. Note that, along P2P_{2}, we have X~(z)=X(z)\tilde{X}(z)=-X(z). After a CC^{\infty} perturbation of XX away from a neighborhood of V\partial V and keeping the transversality of RλR_{\lambda} and V{V}, we can assume that all singular points of X~\tilde{X} are nondegenerate. Since RλR_{\lambda} is transverse to V{V}, all singularities of X~\tilde{X} have the same sign and they must be positive since V𝑑λ=P2λP1λ>0\int_{V}d\lambda=\int_{P_{2}}\lambda-\int_{P_{1}}\lambda>0. Let ΨV\Psi_{V} be a trivialization of TV¯V¯T\bar{V}\to\bar{V}. Then we have

wind(X~|P2,Ψ)wind(X~|P1,Ψ)\displaystyle\operatorname{wind}(\tilde{X}|_{P_{2}},\Psi)-\operatorname{wind}(\tilde{X}|_{P_{1}},\Psi) =X~(z)=0sign(DX~(z):(TzV,oz)(ξz,oz))\displaystyle=\sum_{\tilde{X}(z)=0}{\rm sign}\left(D\tilde{X}(z):(T_{z}V,o_{z})\to(\xi_{z},o^{\prime}_{z})\right)
=X~(z)=0sign(DX~(z):(TzV,oz)(TzV,oz))\displaystyle=\sum_{\tilde{X}(z)=0}{\rm sign}\left(D\tilde{X}(z):(T_{z}V,o_{z})\to(T_{z}V,o_{z})\right)
=wind(X~|P2,ΨV)wind(X~|P1,ΨV)\displaystyle=\operatorname{wind}(\tilde{X}|_{P_{2}},\Psi_{V})-\operatorname{wind}(\tilde{X}|_{P_{1}},\Psi_{V})
=0.\displaystyle=0.

Here we have used the fact that X~\tilde{X} points outwards. Consequently,

wind(X~|P1,Ψ)=wind(X~|P2,Ψ)=wind(X|P2,Ψ)=1.\operatorname{wind}(\tilde{X}|_{P_{1}},\Psi)=\operatorname{wind}(\tilde{X}|_{P_{2}},\Psi)=\operatorname{wind}({X}|_{P_{2}},\Psi)=1.

We know that 𝒟:=DVP1P2\mathcal{D}:=D\cup V\cup P_{1}\cup P_{2} is a C1C^{1}-embedded disk with boundary P1P_{1}. Note that 𝒟\mathcal{D} with the orientation induced by P1P_{1} is an Seifert surface for P1P_{1}. Consider the nonvanishing section of ξ𝒟\xi\to\mathcal{D} defined by Z(z)=Ψ1(z,(1,0))Z(z)=\Psi^{-1}(z,(1,0)). After a small perturbation of P1P_{1} in the direction of ZZ, which we can assume to be transverse to the interior of 𝒟\mathcal{D} and ξ\xi, we find a new closed curve P~1\tilde{P}_{1}. Consider a trivialization Ψ~\tilde{\Psi} of (ξ|𝒟,o)(\xi|_{\partial\mathcal{D}},o) induced by the nonvanishing section X~|P1\tilde{X}|_{P_{1}}. Thus,

sl(P1)=P~1(DV)=wind(Z,Ψ~)=wind(X~|P1,Ψ)=1.{\rm sl}(P_{1})=\tilde{P}_{1}\cdot(D\cup V)=\operatorname{wind}(Z,\tilde{\Psi})=-\operatorname{wind}(\tilde{X}|_{P_{1}},\Psi)=-1.

We have proved the existence of a 3213-2-1 foliation adapted to λ\lambda. 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 3213-2-1 foliation adapted to the contact form λ\lambda and let P2=(x2,T2)P_{2}=(x_{2},T_{2}) be the binding orbit with Conley-Zehnder index 22, as in Definition 1.2. Then there is no C1C^{1}-embedding ψ:S2S3\psi:S^{2}\to S^{3} such that ψ(S1×{0})=x2()\psi({S^{1}\times\{0\}})=x_{2}(\mathbb{R}) 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 ψ:S2S3\psi:S^{2}\to S^{3} is a C1C^{1}-embedding such that ψ(S1×{0})=P2\psi({S^{1}\times\{0\}})=P_{2} and ψ(S2)P2\psi(S^{2})\setminus P_{2} is transverse to the Reeb vector field, then we can assume that ψ|S2(S1×{0})\psi|_{S^{2}\setminus(S^{1}\times\{0\})} is also transverse to T(P3P2)=U1U2T\setminus(P_{3}\cup P_{2})=U_{1}\cup U_{2}.

Proof.

Define F:×S2S3F:\mathbb{R}\times S^{2}\to S^{3} by (t,x)φtψ(x)(t,x)\mapsto\varphi^{t}\circ\psi(x), where φt\varphi^{t} is the Reeb flow. Then FF satisfies F(t,)(S1×{0})=ψ(S1×{0})F(t,\cdot)(S^{1}\times\{0\})=\psi(S^{1}\times\{0\}), t\forall t\in\mathbb{R}, and

dF(t,x)(a,v)=aRλ(φtψ(x))+dφψ(x)tdψxv,dF_{(t,x)}(a,v)=aR_{\lambda}(\varphi^{t}\circ\psi(x))+d\varphi^{t}_{\psi(x)}\cdot d\psi_{x}v~{},

for every (t,x)×S2(t,x)\in\mathbb{R}\times S^{2} and (a,v)×TxS2(a,v)\in\mathbb{R}\times T_{x}S^{2}. Since ψ\psi is transverse to the Reeb flow away from S1×{0}S^{1}\times\{0\}, we know that πdψx\pi\circ d\psi_{x} is surjective. Since dφtd\varphi^{t} preserves the splitting TS3=RλξTS^{3}=\mathbb{R}R_{\lambda}\oplus\xi, it follows that πdφψ(x)tdψx\pi\circ d\varphi^{t}_{\psi(x)}\circ d\psi_{x} is surjective. We conclude that dF(t,x)dF(t,x) is surjective for all xS2(S1×{0})x\in S^{2}\setminus(S^{1}\times\{0\}) and tt\in\mathbb{R}. In particular, FF is transverse to T(P2P3)T\setminus(P_{2}\cup P_{3}) on ×(S2(S1×{0}))\mathbb{R}\times(S^{2}\setminus(S^{1}\times\{0\})). This implies 111If XX, SS and YY are smooth manifolds, ZZ is a submanifold of YY and F:S×XYF:S\times X\to Y is a smooth map transverse to ZZ, then for almost every sSs\in S, fs:=F(s,)f_{s}:=F(s,\cdot) is transverse to ZZ. A proof of this fact can be found in [13, §3]. that for almost all tt\in\mathbb{R}, F(t,)|S2S1×{0}F(t,\cdot)|_{S^{2}\setminus S^{1}\times\{0\}} is transverse to T(P2P3)T\setminus(P_{2}\cup P_{3}).

Moreover, F(t,)dλ=(φtψ)dλ=ψdλF(t,\cdot)^{*}d\lambda=(\varphi^{t}\circ\psi)^{*}d\lambda=\psi^{*}d\lambda, and it follows that F(t,)F(t,\cdot) is still transverse to the flow for all tt\in\mathbb{R}. Thus, we can replace ψ\psi with F(t,)F(t,\cdot) for some tt satisfying F(t,)|S2S1×{0}T(P2P3)F(t,\cdot)|_{S^{2}\setminus S^{1}\times\{0\}}\pitchfork T\setminus(P_{2}\cup P_{3}) to get the desired embedding. ∎

Lemma 6.3.

Assume that ψ:S2S3\psi:S^{2}\to S^{3} is a C1C^{1} embedding such that ψ(S1×{0})=P2\psi({S^{1}\times\{0\}})=P_{2} and such that the image of each closed hemisphere is a strong transverse section. Then ψ\psi is transverse to the torus TT along P2P_{2}.

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 ψ:S2S3\psi:S^{2}\to S^{3} is a C1C^{1} embedding such that ψ(S1×{0})=P2\psi({S^{1}\times\{0\}})=P_{2} and such that each hemisphere is a strong transverse section.

Let T=U1U2P2P3T=U_{1}\cup U_{2}\cup P_{2}\cup P_{3} be the torus given by Definition 1.2. TT divides S3S^{3} into two closed regions 1\mathcal{R}_{1} e 2\mathcal{R}_{2}. The region 1\mathcal{R}_{1} contains an embedded disk with boundary P2P_{2}, so that P2P_{2} is contractible in 1\mathcal{R}_{1}. One can also show, using Mayer-Vietoris sequence, that the holomology class of xT2x_{T_{2}} generates H1(2,)H_{1}(\mathcal{R}_{2},\mathbb{Z}). Lemma 6.3 shows that the image of any neighborhood of S1×{0}S2S^{1}\times\{0\}\subset S^{2} by ψ\psi intersects both 1\mathcal{R}_{1} and 2\mathcal{R}_{2} away from ψ(S1×{0})\psi(S^{1}\times\{0\}). This implies that the sphere ψ(S2)\psi(S^{2}) intersects the torus TT away from P2P_{2}. Indeed, one of the hemispheres of ψ(S2)\psi(S^{2}) intersects 2\mathcal{R}_{2} and can not be contained in 2\mathcal{R}_{2}, since this would imply P2P_{2} contractible in 2\mathcal{R}_{2}, a contradiction.

Since ψ|S2S1×{0}\psi|_{S^{2}\setminus S^{1}\times\{0\}} is transverse to P3P_{3} and lk(P2,P3)=0{\rm lk}(P_{2},P_{3})=0, we conclude that ψ(S2)P3=\psi(S^{2})\cap P_{3}=\emptyset. Using Lemmas 6.2 and 6.3, we conclude that ψ\psi intersects TT transversely and the intersection ψ(S2)T\psi(S^{2})\cap T is contained in a closed subset of TP3T\setminus P_{3}. We conclude that the preimage of the intersection ψ(S2)T\psi(S^{2})\cap T by ψ\psi is a 11-dimensional submanifold of S2S^{2} which is a closed subset of S2S^{2}. It follows that each connected component of ψ1(ψ(S2)T)\psi^{-1}(\psi(S^{2})\cap T) is diffeomorphic to S1S^{1}.

The equator S1×{0}S^{1}\times\{0\} is one of the connected components of the boundary of a region RS2R\subset S^{2} such that ψ(R)2\psi(R)\subset\mathcal{R}_{2}. Thus, one of the other connected components of the boundary of RR, that we denote by SS, is such that ψ|S\psi|_{S} is homologous to x2T2{x_{2}}_{T_{2}} in 2\mathcal{R}_{2}. Denoting H1(T,)=[x2T2][m]H^{1}(T,\mathbb{Z})=\mathbb{Z}[{x_{2}}_{T_{2}}]\oplus\mathbb{Z}[m], the homology class of ψ|S\psi|_{S} in H1(T,)H^{1}(T,\mathbb{Z}) is (1,l)(1,l) for some ll\in\mathbb{Z}. Since SS does not intersect P2P_{2}, ll must be zero. This implies that ψ(S)\psi({S}) and P2P_{2} divide TT into two connected regions.

Now fix an orientation on S1×{0}S2S^{1}\times\{0\}\subset S^{2} in such a way that ψ|S1×{0}\psi|_{S^{1}\times\{0\}} preserves orientation. Consider the closed hemispheres of S2S^{2}, that we call H+H_{+} and HH_{-}, with the orientation induced by the orientation of S1×{0}S^{1}\times\{0\}. It follows that

0<T2=/x2T2λ=H±ψ𝑑λ.0<T_{2}=\int_{\mathbb{R}/\mathbb{Z}}{x_{2}}_{T_{2}}^{*}\lambda=\int_{H_{\pm}}\psi^{*}d\lambda~{}.

Since ψ\psi is transverse to the Reeb vector field RλR_{\lambda} in H±S1×{0}H_{\pm}\setminus S^{1}\times\{0\}, we have ψdλ>0\psi^{*}d\lambda>0 in H±S1×{0}H_{\pm}\setminus S^{1}\times\{0\}. Let BB be the connected region of S2S^{2} bounded by S1×{0}S^{1}\times\{0\} and SS. Since BB is contained in one of the hemispheres of S2S^{2}, we have

(85) 0<Bψ𝑑λ=/x2T2λSψλ=T2ψ(S)λ.0<\int_{B}\psi^{*}d\lambda=\int_{\mathbb{R}/\mathbb{Z}}{x_{2}}_{T_{2}}^{*}\lambda-\int_{S}\psi^{*}\lambda=T_{2}-\int_{\psi(S)}\lambda~{}.

Since ψ(S2)\psi(S^{2}) does not intersect P3P_{3}, we know that one of the regions of TT bounded by P2P_{2} and ψ(S)\psi(S), that we denote by AA, satisfies either AU1A\subset U_{1} or AU2A\subset U_{2}. Recall that U1U_{1} and U2U_{2} are oriented in such a way that dλ|U1,2d\lambda|U_{1,2} is an area form, P3P_{3} is a positive asymptotic limit and P2P_{2} is a negative asymptotic limit. Then we have

(86) 0<A𝑑λ=ψ(S)λP2λ=ψ(S)λT2,0<\int_{A}d\lambda=\int_{\psi(S)}\lambda-\int_{P_{2}}\lambda=\int_{\psi(S)}\lambda-T_{2}~{},

contrary to (85). This proves the proposition. ∎

6.1. Relative position of sections along 𝑷𝟐\boldsymbol{P_{2}}

The orbit P2P_{2} is hyperbolic and lies in the intersection of its stable manifold W+(P2)W^{+}(P_{2}) and its unstable manifold W(P2)W^{-}(P_{2}). The tangent spaces of W±(P2)W^{\pm}(P_{2}) along the periodic solution tx2(t)t\mapsto x_{2}(t) are spanned by the Reeb vector field Rλ(x2(t))R_{\lambda}(x_{2}(t)) and vector fields v±(t)ξx2(t)v^{\pm}(t)\in\xi_{x_{2}(t)} defined below.

Consider the path of symplectic matrices Φ(t)=Ψx2(t)dφt|ξ(x2(0))Ψx2(0)1,t,\Phi(t)=\Psi_{x_{2}(t)}\circ d\varphi^{t}|_{\xi(x_{2}(0))}\circ\Psi^{-1}_{x_{2}(0)},~{}t\in\mathbb{R}, where Ψ\Psi is any global symplectic trivialization of ξ\xi. Since μ(P2)=2\mu(P_{2})=2, we know that Φ(T2)\Phi(T_{2}) has two eigenvalues β,β1\beta,\beta^{-1}, where β>1\beta>1. The vector v(0)v^{-}(0) is an eigenvector of dφT2|ξ(x2(0))d\varphi^{T_{2}}|_{\xi(x_{2}(0))} associated to the eigenvalue β\beta and the vector v+(0)v^{+}(0) is an eigenvector of dφT2|ξ(x2(0))d\varphi^{T_{2}}|_{\xi(x_{2}(0))} associated to the eigenvalue β1\beta^{-1}. For each tt, we can define

v±(t)=dφt|ξ(x2(0))v±(0),v^{\pm}(t)=d\varphi^{t}|_{\xi(x_{2}(0))}v^{\pm}(0),

so that v(t)v^{-}(t) is an eigenvector of dφT2|ξ(x(t))d\varphi^{T_{2}}|_{\xi(x(t))} associated to the eigenvalue β>1\beta>1 and v+(t)v^{+}(t) is an eigenvector of dφT2|ξ(x2(t))d\varphi^{T_{2}}|_{\xi(x_{2}(t))} associated to the eigenvalue β1\beta^{-1}.

Replacing v±(t)v^{\pm}(t) with v±(t)-v^{\pm}(t) if necessary, we can assume that {v(t),v+(t)}\{v^{-}(t),v^{+}(t)\} is a positive basis for ξx(t)\xi_{x(t)}, for each tt. The basis {v(t),v+(t)}\{v^{-}(t),v^{+}(t)\} determines four open quadrants in ξx(t)\xi_{x(t)}. Let (I)\rm{(I)} and (III)\rm{(III)} be the open quadrants between v(t)\mathbb{R}v^{-}(t) and v+(t)\mathbb{R}v^{+}(t) following the positive orientation and (II)\rm{(II)} and (IV)\rm{(IV)} the open quadrants between v+(t)\mathbb{R}v^{+}(t) and v(t)\mathbb{R}v^{-}(t).

Proposition 6.4.

Let tv(t)ξx2(t)t\mapsto v(t)\in\xi_{x_{2}(t)} be a T2T_{2}-periodic nonvanishing section such that wind(v,Ψ)=1.\operatorname{wind}(v,\Psi)=1~{}. If dλ(v(t),Rλv(t))>0,t,d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t))>0,~{}\forall t\in\mathbb{R}, then v(t)v(t) belongs to regions (I)(I) or (III)(III) for every tt\in\mathbb{R}. If dλ(v(t),Rλv(t))<0,t,d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t))<0,~{}\forall t\in\mathbb{R}, then v(t)v(t) belongs to regions (II)(II) or (IV)(IV) for every tt\in\mathbb{R}.

Proof.

Let S(t)=J0Φ˙(t)Φ1(t)S(t)=-J_{0}\dot{\Phi}(t)\Phi^{-1}(t). Recall that tS(t)t\mapsto S(t) is T2T_{2}-periodic, and for each tt, S(t)S(t) is real and symmetric matrix. Define A(t)=J0S(t)A(t)=J_{0}S(t). In what follows we will also write v±(t)v^{\pm}(t) for the representations of the sections v±(t)v^{\pm}(t) in the trivialization Ψ\Psi. The sections v±(t)v^{\pm}(t) are solutions of the equation

(87) x˙(t)=A(t)x(t)\dot{x}(t)=A(t)x(t)

satisfying v+(T2)=1βv+(0),v(T2)=βv(0).v^{+}(T_{2})=\frac{1}{\beta}v^{+}(0),~{}~{}~{}v^{-}(T_{2})=\beta v^{-}(0)~{}. In the basis {v(0),v+(0)}\{v^{-}(0),v^{+}(0)\}, Φ(T2)\Phi(T_{2}) has the form

Φ(T2)=[β001β].\Phi(T_{2})=\begin{bmatrix}\beta&0\\ 0&\frac{1}{\beta}\end{bmatrix}.

We want to find a matrix BB satisfying eT2B=Φ(T2)e^{T_{2}B}=\Phi(T_{2}) and a T2T_{2}-periodic map tP(t)t\mapsto P(t) such that Φ(t)=P(t)etB,t\Phi(t)=P(t)e^{tB},~{}\forall t\in\mathbb{R}, so that with the change of variables y=P1(t)xy=P^{-1}(t)x, the equation (87) becomes

(88) y˙(t)=By(t).\dot{y}(t)=By(t).

Define

B=1T2[lnβ00lnβ],B=\frac{1}{T_{2}}\begin{bmatrix}\ln\beta&0\\ 0&-\ln\beta\end{bmatrix},

and tP(t)t\mapsto P(t) by Φ(t)=P(t)etB.\Phi(t)=P(t)e^{tB}. The map tP(t)t\mapsto P(t) is T2T_{2}-periodic. In fact,

Φ(t)eT2B=Φ(t)Φ(T2)=Φ(t+T2)=P(t+T2)etBeT2BΦ(t)=P(t+T2)etB.\Phi(t)e^{T_{2}B}=\Phi(t)\Phi(T_{2})=\Phi(t+T_{2})=P(t+T_{2})e^{tB}e^{T_{2}B}\Rightarrow\Phi(t)=P(t+T_{2})e^{tB}.

If x(t)x(t) is a solution of (87), then y(t)=P(t)1x(t)y(t)=P(t)^{-1}x(t) satisfies

y˙(t)\displaystyle\dot{y}(t) =(P(t)1˙)x(t)+P(t)1x˙(t)\displaystyle=(\dot{P(t)^{-1}})x(t)+P(t)^{-1}\dot{x}(t)
=P(t)1P˙(t)P(t)1(P(t)y(t))+P(t)1A(t)(P(t)y(t))\displaystyle=-P(t)^{-1}\dot{P}(t)P(t)^{-1}(P(t)y(t))+P(t)^{-1}A(t)(P(t)y(t))
=(P(t)1P˙(t)+P(t)1A(t)P(t))y(t).\displaystyle=(-P(t)^{-1}\dot{P}(t)+P(t)^{-1}A(t)P(t))y(t).

Using the identities Φ˙(t)=A(t)Φ(t)\dot{\Phi}(t)=A(t)\Phi(t) and Φ(t)=P(t)etB\Phi(t)=P(t)e^{tB}, we get

A(t)P(t)etB=Φ˙(t)=P(t)etBB+P˙(t)etB\displaystyle A(t)P(t)e^{tB}=\dot{\Phi}(t)=P(t)e^{tB}B+\dot{P}(t)e^{tB}
B=P(t)1P˙(t)+P(t)1A(t)P(t).\displaystyle\Rightarrow B=-P(t)^{-1}\dot{P}(t)+P(t)^{-1}A(t)P(t).

Thus, y(t)y(t) is a solution of (88). In the same way, if y(t)y(t) solves (88), then x(t)=P(t)y(t)x(t)=P(t)y(t) solves (87).

Writing equation (88) in coordinates, we get

{y1˙(t)=λy1(t)y2˙(t)=λy2(t),\left\{\begin{array}[]{l}\dot{y_{1}}(t)=\lambda y_{1}(t)\\ \dot{y_{2}}(t)=-\lambda y_{2}(t),\end{array}\right.

where λ=1T2lnβ\lambda=\frac{1}{T_{2}}\ln\beta. Writing y=(y1,y2)y=(y_{1},y_{2}) in polar coordinates,

{y1(t)=ρ(t)cosη(t)y2(t)=ρ(t)sinη(t),\left\{\begin{array}[]{l}{y_{1}}(t)=\rho(t)\cos\eta(t)\\ {y_{2}}(t)=\rho(t)\sin\eta(t),\end{array}\right.

we get

(89) ρ2(t)η˙(t)=ρ2(t)(2λsinη(t)cosη(t)).\rho^{2}(t)\dot{\eta}(t)=\rho^{2}(t)\left(-2\lambda\sin\eta(t)\cos\eta(t)\right).

Let tv(t)ξx2(t)t\mapsto v(t)\in\xi_{x_{2}(t)} be a T2T_{2}-periodic nonvanishing section on x2ξx_{2}^{*}\xi. We want to compute dλ(v(t),Rλv(t)).d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t)). In the symplectic trivialization Ψ\Psi, Rλv(t)\mathcal{L}_{R_{\lambda}}v(t) takes the form

Rλv(t)=ddtv(t)J0S(t)v(t),\mathcal{L}_{R_{\lambda}}v(t)=\frac{d}{dt}v(t)-J_{0}S(t)v(t),

where we also denote by v(t)v(t) its representation in the trivialization Ψ\Psi. Thus,

dλ(v(t),Rλv(t))\displaystyle d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t)) =dλ0(v(t),v˙(t)J0S(t)v(t))\displaystyle=d\lambda_{0}(v(t),\dot{v}(t)-J_{0}S(t)v(t))
=dλ0(v(t),v˙(t))dλ0(v(t),A(t)v(t)).\displaystyle=d\lambda_{0}(v(t),\dot{v}(t))-d\lambda_{0}(v(t),A(t)v(t)).

Writing v(t)v(t) in the basis {v(0),v+(0)}\{v^{-}(0),v^{+}(0)\} as v(t)=(v1(t),v2(t))v(t)=(v_{1}(t),v_{2}(t)) and in the basis {P(t)v(0),P(t)v+(0)}\{P(t)v^{-}(0),P(t)v^{+}(0)\} as v(t)=(u1(t),u2(t))v(t)=(u_{1}(t),u_{2}(t)), we have

(r(t)cosθ(t),r(t)sinθ(t)):=(u1(t),u2(t))=P(t)1(v1(t),v2(t)).(r(t)\cos\theta(t),r(t)\sin\theta(t)):=(u_{1}(t),u_{2}(t))=P(t)^{-1}(v_{1}(t),v_{2}(t)).

Thus, we have

dλ0(v(t),v˙(t))\displaystyle d\lambda_{0}(v(t),\dot{v}(t)) =CdetP(t)dλ0((u1(t),u2(t)),P(t)1(v1˙(t),v2˙(t)))\displaystyle=C\cdot\det P(t)\cdot d\lambda_{0}\left((u_{1}(t),u_{2}(t)),P(t)^{-1}(\dot{v_{1}}(t),\dot{v_{2}}(t))\right)
=C(r(t)2θ˙(t)+dλ0((u1(t),u2(t)),ddt(P(t)1)(v1(t),v2(t)))),\displaystyle=C\left(r(t)^{2}\dot{\theta}(t)+d\lambda_{0}\left((u_{1}(t),u_{2}(t)),\frac{d}{dt}{(P(t)^{-1})}(v_{1}(t),v_{2}(t))\right)\right),

where CC is a positive constant. For fixed tt, let sxt(s)s\mapsto x_{t}(s) be the solution of (87) such that xt(t)=v(t)x_{t}(t)=v(t). Then

A(t)v(t)\displaystyle A(t)v(t) =A(t)xt(t)=xt˙(t).\displaystyle=A(t)x_{t}(t)=\dot{x_{t}}(t).

Writing xt(s)x_{t}(s) is the basis {v(0),v+(0)}\{v^{-}(0),v^{+}(0)\} as xt(s)=(x1(s),x2(s))x_{t}(s)=(x_{1}(s),x_{2}(s)) and in the basis {P(t)v(0),P(t)v+(0)}\{P(t)v^{-}(0),P(t)v^{+}(0)\} as (y1(s),y2(s))(y_{1}(s),y_{2}(s)), we have

(ρ(s)cosη(s),ρ(s)sinη(s))=(y1(s),y2(s))=P1(x1(s),x2(s)).(\rho(s)\cos\eta(s),\rho(s)\sin\eta(s))=(y_{1}(s),y_{2}(s))=P^{-1}(x_{1}(s),x_{2}(s)).

Thus, we have

dλ0(v(t),A(t)v(t))\displaystyle d\lambda_{0}(v(t),A(t)v(t)) =dλ0(xt(t),x˙t(t))\displaystyle=d\lambda_{0}(x_{t}(t),\dot{x}_{t}(t))
=C(ρ2(t)η˙(t)+dλ0((u1(t),u2(t)),ddt(P(t)1)(v1(t),v2(t)))).\displaystyle=C\left(\rho^{2}(t)\dot{\eta}(t)+d\lambda_{0}\left((u_{1}(t),u_{2}(t)),\frac{d}{dt}(P(t)^{-1})(v_{1}(t),v_{2}(t))\right)\right).

We conclude that

(90) dλ(v(t),Rλv(t))=Cr2(t)(θ˙(t)η˙(t)).d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t))=Cr^{2}(t)(\dot{\theta}(t)-\dot{\eta}(t)).

Assume that tv(t)ξx2(t)t\mapsto v(t)\in\xi_{x_{2}(t)} satisfies wind(v,Ψ)=1\operatorname{wind}(v,\Psi)=1 and that dλ(v(t),Rλv(t))>0,t.d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t))>0,~{}\forall t\in\mathbb{R}~{}. By (90), we have θ˙(t)>η˙(t),t.\dot{\theta}(t)>\dot{\eta}(t),~{}\forall t\in\mathbb{R}. Note that wind(v±(t),Ψ)=1\operatorname{wind}(v^{\pm}(t),\Psi)=1. This follows from μ(P2)=2\mu(P_{2})=2 and the geometric description of the Conley-Zehnder index given in §2.1. This implies that

(91) wind(v(t),P(t)v±(0))=wind(v(t),v±(t))=0.\operatorname{wind}(v(t),P(t)v^{\pm}(0))=\operatorname{wind}(v(t),v^{\pm}(t))=0~{}.

Now we show that for all tt\in\mathbb{R}, v(t)v(t) lies in regions (I)\rm{(I)} or (III)\rm{(III)}. On the contrary, suppose that there exists t0t_{0}\in\mathbb{R} such that v(t0)v(t_{0}) belongs to regions (II)\rm{(II)}, (IV)\rm{(IV)}, or is the direction of P(t0)v(0)P(t_{0})v^{-}(0). By (89), we have θ˙(t0)>η˙(t0)0\dot{\theta}(t_{0})>\dot{\eta}(t_{0})\geq 0. This implies that θ(t)\theta(t) is increasing near t0t_{0}. Since wind(v(t),P(t)v±(0))=0\operatorname{wind}(v(t),P(t)v^{\pm}(0))=0, this would force the existence of t1>t0t_{1}>t_{0} such that θ(t1)=θ(t0)\theta(t_{1})=\theta(t_{0}) and θ˙(t1)0\dot{\theta}(t_{1})\leq 0, a contradiction with θ˙(t1)>η˙(t1)0\dot{\theta}(t_{1})>\dot{\eta}(t_{1})\geq 0. Now, suppose that, for some t0t_{0}, v(t0)v(t_{0}) is in the direction of P(t0)v+(0)P(t_{0})v^{+}(0), then, using (89), we have θ˙(t0)>η˙(t0)=0\dot{\theta}(t_{0})>\dot{\eta}(t_{0})=0. But this would force v(t)v(t) to go to (II)\rm{(II)} or (IV)\rm{(IV)}, which is impossible. We conclude that v(t)v(t) lies in regions (I)\rm{(I)} or (III)\rm{(III)}, for every tt\in\mathbb{R}.

By a similar argument, we conclude that if tv(t)ξx2(t)t\mapsto v(t)\in\xi_{x_{2}(t)} satisfies wind(v,Φ)=1\operatorname{wind}(v,\Phi)=1 and dλ(v(t),Rλv(t))<0,t,d\lambda(v(t),\mathcal{L}_{R_{\lambda}}v(t))<0,~{}\forall t\in\mathbb{R}~{}, then we have v(t)(II)v(t)\in\rm{(II)} or v(t)(IV)v(t)\in\rm{(IV)}, t\forall t\in\mathbb{R}. ∎

6.2. Proof of Lemma 6.3

Lemma 6.5.

Let /tη(t)ξx2(T2t)\mathbb{R}/\mathbb{Z}\ni t\mapsto\eta(t)\in\xi_{x_{2}(T_{2}t)} be a section on x2T2ξ{x_{2}}_{T_{2}}^{*}\xi such that η()\eta(\cdot) and Rλ(x2(T2)){R_{\lambda}}(x_{2}(T_{2}\cdot)) generate dψ(TS2)d\psi(TS^{2}) along x2(T2)x_{2}(T_{2}\cdot). Let /tη(t)ξx2(T2t)\mathbb{R}/\mathbb{Z}\ni t\mapsto\eta^{\prime}(t)\in\xi_{x_{2}(T_{2}t)} be a section on x2T2ξ{x_{2}}_{T_{2}}^{*}\xi such that η()\eta^{\prime}(\cdot) and Rλ(x2(T2)){R_{\lambda}}(x_{2}(T_{2}\cdot)) generate the tangent space of TT along x2(T2)x_{2}(T_{2}\cdot). Then wind(η,Ψ)=wind(η,Ψ)=1\operatorname{wind}(\eta^{\prime},\Psi)=\operatorname{wind}(\eta,\Psi)=1.

Proof.

Let HS2H\subset S^{2} be one of the closed hemispheres of S2S^{2} and let F:=ψ(H)S3F:=\psi(H)\subset S^{3}. Then FF is an embedded disk satisfying F=P2\partial F=P_{2} and F̊\mathring{F} is transverse to the Reeb vector field RλR_{\lambda}. The characteristic distribution (TFξ)(TF\cap\xi)^{\bot} can be parametrized by a smooth vector field XX on FF, which is transverse to F\partial F pointing outwards. After a CC^{\infty} perturbation of FF keeping the transversality of RλR_{\lambda} and F̊\mathring{F}, we can assume that all singular points of XX are nondegenerate. Let oo be the orientation on FF induced by the orientation of F\partial F in the direction of RλR_{\lambda}. Let oo^{\prime} be the orientation on ξ\xi given by dλ|ξd\lambda|_{\xi}. A zero pp of XX is called positive if oo coincides with oo^{\prime} at pp, and negative otherwise. Since RλR_{\lambda} is transverse to F̊\mathring{F}, all singularities of XX have the same sign and they must be positive since F𝑑λ=P2λ>0\int_{F}d\lambda=\int_{P_{2}}\lambda>0. Considering XX as a section of the bundle ξF\xi\to F, we have

(92) wind(η,Ψ)=wind(X|F,Ψ)=X(z)=0sign(DX(z):(TzF,oz)(ξz,oz)).\operatorname{wind}(\eta,\Psi)=\operatorname{wind}(X|_{\partial F},\Psi)=\sum_{X(z)=0}{\rm sign}\left(DX(z):(T_{z}F,o_{z})\to(\xi_{z},o^{\prime}_{z})\right).

Let ΨF:TFF×2\Psi_{F}:TF\to F\times\mathbb{R}^{2} be a trivialization of TFFTF\to F. Considering XX as a section of the bundle TFFTF\to F and using the fact that XX points outwards at F\partial F, we have

(93) 1=wind(X|F,ΨF)=X(z)=0sign(DX(z):(TzF,oz)(TzF,oz)).1=\operatorname{wind}(X|_{\partial F},\Psi_{F})=\sum_{X(z)=0}{\rm sign}\left(DX(z):(T_{z}F,o_{z})\to(T_{z}F,o_{z})\right).

Since all the singularities of XX are positive and DX(z)DX(z) at a singularity zz of XX does not depend on whether we consider XX as a section of ξF\xi\to F or as a section of TFFTF\to F, equations (92) and (93) coincide. We conclude that wind(η,Ψ)=1.\operatorname{wind}(\eta,\Psi)=1.

By the same arguments above, if ν\nu is a section such that {ν(),R(x2(T2))}\{\nu(\cdot),R(x_{2}(T_{2}\cdot))\} generates the tangent space of the disk D¯\bar{D} along x2(T2)x_{2}(T_{2}\cdot), where D{D} is the disk given by the definition of 3213-2-1 foliation 1.2, we have wind(ν,Ψ)=1.\operatorname{wind}(\nu,\Psi)=1. Since TT is transverse to D¯\bar{D} along x2x_{2}, we obtain wind(η,Ψ)=1.\operatorname{wind}(\eta^{\prime},\Psi)=1~{}.

Now we are ready to prove Lemma 6.3.

Proof of Lemma 6.3.

Let S1=/tη(t)ξS^{1}=\mathbb{R}/\mathbb{Z}\ni t\mapsto\eta(t)\in\xi be a section along x2(T2)x_{2}(T_{2}\cdot) such that {η(),R(x2(T2))}\{\eta(\cdot),R(x_{2}(T_{2}\cdot))\} generates dψ(TS2)d\psi(TS^{2}) along x2(T2)x_{2}(T_{2}\cdot) and

(94) dλ(η(t),Rλη(t))0,tS1.d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))\neq 0,~{}\forall t\in S^{1}.

Let u:(ϵ,ϵ)×S1Ψ(S2)u:(-\epsilon,\epsilon)\times S^{1}\to\Psi(S^{2}); (s,t)u(s,t)(s,t)\mapsto u(s,t) be a smooth function such that

{u(0,)=x2T2su(s,)|s=0=η\left\{\begin{array}[]{lr}u(0,\cdot)={x_{2}}_{T_{2}}\\ \frac{\partial}{\partial s}u(s,\cdot)\big{|}_{s=0}=\eta\end{array}\right.

Fix an orientation on S1×{0}S2S^{1}\times\{0\}\subset S^{2} in such a way that ψ|S1×{0}\psi|_{S^{1}\times\{0\}} preserves orientation. Consider the closed hemispheres of S2S^{2}, that we call H+H_{+} e HH_{-}, with the orientation induced by the orientation of S1×{0}S^{1}\times\{0\}. It follows that

0<T2=S1x2T2λ=H±ψ𝑑λ.0<T_{2}=\int_{S^{1}}{x_{2}}_{T_{2}}^{*}\lambda=\int_{H_{\pm}}\psi^{*}d\lambda~{}.

Since ψ\psi is transverse to the Reeb flow RλR_{\lambda} along H±(S1×{0})H_{\pm}\setminus(S^{1}\times\{0\}), we have ψdλ>0\psi^{*}d\lambda>0 on H±(S1×{0})H_{\pm}\setminus(S^{1}\times\{0\}). This implies that 0 is a local maximum of the function

(ϵ,ϵ)s𝒜(u(s,)),(-\epsilon,\epsilon)\ni s\mapsto\mathcal{A}(u(s,\cdot))\in\mathbb{R},

where 𝒜:C(/,S3)\mathcal{A}:C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3})\to\mathbb{R} is the action functional defined by 𝒜(γ)=/γλ\mathcal{A}(\gamma)=\int_{\mathbb{R}/\mathbb{Z}}\gamma^{*}\lambda. For any J𝒥(λ)J\in\mathcal{J}(\lambda), we have

(95) d2ds2(𝒜(u(s,)))|s=0\displaystyle\frac{d^{2}}{ds^{2}}(\mathcal{A}(u(s,\cdot)))\bigg{|}_{s=0} =S1𝑑λ(AP,J(η),Jη)𝑑t,\displaystyle=\int_{S^{1}}d\lambda\left(A_{P,J}(\eta),J\eta\right)dt,

where AP,JA_{P,J} is the asymptotic operator defined by (18). For a proof, see [5, §1.4]. It follows from (17) and (95) that

(96) d2ds2(𝒜(u(s,)))|s=0=T2S1𝑑λ(η(t),Rλη(t))𝑑t0.\frac{d^{2}}{ds^{2}}(\mathcal{A}(u(s,\cdot)))\bigg{|}_{s=0}=T_{2}\int_{S^{1}}d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))dt\leq 0~{}.

By (94), we have

(97) dλ(η(t),Rλη(t))<0,tS1.d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))<0,~{}\forall t\in S^{1}.

Let /tη(t)ξ\mathbb{R}/\mathbb{Z}\ni t\mapsto\eta^{\prime}(t)\in\xi be a section along x2(T2)x_{2}(T_{2}\cdot) such that {η(),R(x2(T2))}\{\eta^{\prime}(\cdot),R(x_{2}(T_{2}\cdot))\} generates the tangent space of TT along x2(T2)x_{2}(T_{2}\cdot) and

(98) dλ(η(t),Rλη(t))0,tS1.d\lambda(\eta^{\prime}(t),\mathcal{L}_{R_{\lambda}}\eta^{\prime}(t))\neq 0,~{}\forall t\in S^{1}.

Let v:(ϵ,ϵ)×S1Tv:(-\epsilon,\epsilon)\times S^{1}\to T, (s,t)v(s,t)(s,t)\mapsto v(s,t) be a smooth function such that

{v(0,)=x2T2sv(s,)|s=0=η.\left\{\begin{array}[]{lr}v(0,\cdot)={x_{2}}_{T_{2}}\\ \frac{\partial}{\partial s}v(s,\cdot)\big{|}_{s=0}=\eta^{\prime}.\end{array}\right.

Recall that the cylinders U1U_{1} and U2U_{2} given by Definition 1.2 are oriented in such a way that dλ|U1,2d\lambda|U_{1,2} is an area form, P3P_{3} is a positive asymptotic limit and P2P_{2} is a negative asymptotic limit. This implies that 0 is a local minimum of the function s𝒜(v(s,))s\mapsto\mathcal{A}(v(s,\cdot)). It follows from (17) and (95) that

(99) d2ds2(𝒜(v(s,)))|s=0=T2S1𝑑λ(η(t),Rλη(t))𝑑t0.\frac{d^{2}}{ds^{2}}(\mathcal{A}(v(s,\cdot)))\bigg{|}_{s=0}=T_{2}\int_{S^{1}}d\lambda(\eta^{\prime}(t),\mathcal{L}_{R_{\lambda}}\eta^{\prime}(t))dt\geq 0~{}.

By (98), we have

(100) dλ(η(t),Rλη(t))>0,tS1.d\lambda(\eta^{\prime}(t),\mathcal{L}_{R_{\lambda}}\eta^{\prime}(t))>0,~{}\forall t\in S^{1}.

Now we can apply Lemma 6.5 and Proposition 6.4 to the sections η\eta and η\eta^{\prime} and conclude that ψ\psi is transverse to the torus TT along P2P_{2}. ∎

7. Proof of Theorem 1.6

In this section, we prove Theorem 1.6, which is restated below.

Theorem 7.1.

Let λ\lambda be a tight contact form on S3S^{3}. Assume that there exist Reeb orbits P1=(x1,T1),P2=(x2,T2),P3=(x3,T3)𝒫(λ)P_{1}=(x_{1},T_{1}),~{}P_{2}=(x_{2},T_{2}),~{}P_{3}=(x_{3},T_{3})\in\mathcal{P}(\lambda) that are nondegenerate, simple, and have Conley-Zehnder indices respectively 11, 22 and 33. Assume further that the orbits P1P_{1}, P2P_{2}, and P3P_{3} are unknotted, PiP_{i} and PjP_{j} are not linked for iji\neq j, i,j{1,2,3}i,j\in\{1,2,3\}, and the following conditions hold:

  1. (i)

    T1<T2<T3<2T1T_{1}<T_{2}<T_{3}<2T_{1};

  2. (ii)

    If P=(x,T)𝒫(λ)P=(x,T)\in\mathcal{P}(\lambda) satisfies PP3,TT3P\neq P_{3},~{}T\leq T_{3} and lk(P,P3)=0{\rm lk}(P,P_{3})=0, then P{P1,P2}P\in\{P_{1},P_{2}\}.

  3. (iii)

    There exists J𝒥(λ)J\in\mathcal{J}(\lambda) such that the almost complex structure J~=(λ,J)\tilde{J}=(\lambda,J) admits a finite energy plane u~:×S3\tilde{u}:\mathbb{C}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at it positive puncture z=z=\infty and a finite energy clylinder w~:{0}×S3\tilde{w}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} asymptotic to P3P_{3} at its positive puncture z=z=\infty and P1P_{1} at its negative puncture z=0z=0;

  4. (iv)

    There exists no C1C^{1}-embedding Ψ:S23S3\Psi:S^{2}\subset\mathbb{R}^{3}\to S^{3} such that Ψ(S1×{0})=x2()\Psi({S^{1}\times\{0\}})=x_{2}(\mathbb{R}) and each hemisphere is a strong transverse section.

Then there exists a 3213-2-1 foliation adapted to λ\lambda with binding orbits P1P_{1}, P2P_{2}, and P3P_{3}. Consequently, there exists at least one homoclinic orbit to P2P_{2}.

Proof of Theorem 1.6.

By the same arguments used in §3.1, we find a maximal one-parameter family of finite energy J~\tilde{J}-holomorphic planes

(101) u~τ=(aτ,uτ):×S3,τ(0,1)\tilde{u}_{\tau}=(a_{\tau},u_{\tau}):\mathbb{C}\to\mathbb{R}\times S^{3},~{}~{}\tau\in(0,1)

asymptotic to the orbit P3P_{3}. For each τ(0,1)\tau\in(0,1), uτu_{\tau} is an embedding transverse to the Reeb flow and uτ1()uτ2()=,τ1τ2.u_{\tau_{1}}(\mathbb{C})\cap u_{\tau_{2}}(\mathbb{C})=\emptyset,~{}\forall\tau_{1}\neq\tau_{2}. We assume that τ\tau strictly increases in the direction of RλR_{\lambda}.

Now we describe how the family {u~τ}\{\tilde{u}_{\tau}\} breaks as τ0+\tau\to 0^{+} and τ1\tau\to 1^{-}. Consider a sequence τn(0,1)\tau_{n}\in(0,1) satisfying τn0+\tau_{n}\to 0^{+}, and define u~n:=u~τn\tilde{u}_{n}:=\tilde{u}_{\tau_{n}}.

Claim I

There exists a J~\tilde{J}-holomorphic finite energy cylinder

(102) u~r=(ar,ur):{0}×S3\tilde{u}_{r}=(a_{r},u_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}

asymptotic to P3P_{3} at its positive puncture z=z=\infty and to P2P_{2} at its negative puncture z=0z=0, and a finite energy J~\tilde{J}-holomorphic plane

(103) u~q=(aq,uq):×S3\tilde{u}_{q}=(a_{q},u_{q}):\mathbb{C}\to\mathbb{R}\times S^{3}

asymptotic to P2P_{2} at its positive puncture z=z=\infty, such that, after a suitable reparametrization and an \mathbb{R}-translations of u~n\tilde{u}_{n}, the following hold

  1. (i)

    up to a subsequence, u~nu~r\tilde{u}_{n}\to\tilde{u}_{r} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  2. (ii)

    There exist sequences δn0+\delta_{n}\to 0^{+}, znz_{n}\in\mathbb{C} and cnc_{n}\in\mathbb{R} such that, up to a subsequence, u~n(zn+δn)+cnu~q\tilde{u}_{n}(z_{n}+\delta_{n}\cdot)+c_{n}\to\tilde{u}_{q} in Cloc()C^{\infty}_{loc}(\mathbb{C}) as nn\to\infty.

  3. (iii)

    Given an S1S^{1}-invariant neighborhood 𝒲3C(/,S3)\mathcal{W}_{3}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx3(T3t)t\mapsto x_{3}(T_{3}t), there exists R0>>1R_{0}>>1 such that the loop tun(Re2πit)t\mapsto u_{n}(Re^{2\pi it}) belongs to 𝒲3\mathcal{W}_{3}, for RR0R\geq R_{0} and large nn.

  4. (iv)

    Given an S1S^{1}-invariant neighborhood 𝒲2C(/,S3)\mathcal{W}_{2}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx2(T2t)t\mapsto x_{2}(T_{2}t), there exists ϵ1>0\epsilon_{1}>0 and R1>>0R_{1}>>0 such that the loop tun(zn+Re2πit)t\mapsto u_{n}(z_{n}+Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}, for R1δnRϵ1R_{1}\delta_{n}\leq R\leq\epsilon_{1} and large nn.

  5. (v)

    Given any neighborhood 𝒱\mathcal{V} of ur({0})uq()P2P3u_{r}(\mathbb{C}\setminus\{0\})\cup u_{q}(\mathbb{C})\cup P_{2}\cup P_{3}, we have un()𝒱u_{n}(\mathbb{C})\subset\mathcal{V}, for large nn.

Here (a,x)+c:=(a+c,x),(a,x)×S3,c(a,x)+c:=(a+c,x),~{}\forall(a,x)\in\mathbb{R}\times S^{3},~{}c\in\mathbb{R}. A similar claim holds for any sequence τn(0,1){\tau_{n}}\in(0,1) satisfying τn1\tau_{n}\to 1^{-}. In this case we change the notation from u~r\tilde{u}_{r} and u~q\tilde{u}_{q} to u~r\tilde{u}_{r}^{\prime} and u~q\tilde{u}_{q}^{\prime} respectively.

To prove Claim I, define γ\gamma as any real number satisfying

(104) 0<γ<min{T1,T2T1,T3T2}.0<\gamma<\min\{T_{1},T_{2}-T_{1},T_{3}-T_{2}\}.

After reparametrizing and translating u~n\tilde{u}_{n} in the \mathbb{R} direction, we can assume that

(105) 𝔻un𝑑λ=γ,n\displaystyle\int_{\mathbb{C}\setminus\mathbb{D}}u_{n}^{*}d\lambda=\gamma,~{}\forall n\in\mathbb{N}
(106) an(2)=0,n\displaystyle a_{n}(2)=0,~{}\forall n\in\mathbb{N}
(107) an(0)=infan(),n\displaystyle a_{n}(0)=\inf a_{n}(\mathbb{C}),~{}\forall n\in\mathbb{N}

Let Γ\Gamma\subset\mathbb{C} be the set of points zz\in\mathbb{C} such that there exist subsequence u~nj\tilde{u}_{n_{j}} and sequence zjzz_{j}\to z satisfying |du~nj(zj)||d\tilde{u}_{n_{j}}(z_{j})|\to\infty.

If Γ=\Gamma=\emptyset, by elliptic estimates we find a J~\tilde{J}-holomorphic map u~r:×S3\tilde{u}_{r}:\mathbb{C}\to\mathbb{R}\times S^{3} such that, up to a subsequence, u~nu~r\tilde{u}_{n}\to\tilde{u}_{r} in Cloc(,×S3)C^{\infty}_{loc}(\mathbb{C},\mathbb{R}\times S^{3}). Using Fatou’s Lemma we conclude that E(u~r)T3E(\tilde{u}_{r})\leq T_{3}. By (105) we have

(108) 𝔻ur𝑑λ=limn𝔻unλ=limn𝔻un𝑑λ=T3γ>0.\int_{\partial\mathbb{D}}u_{r}^{*}d\lambda=\lim_{n\to\infty}\int_{\partial\mathbb{D}}u_{n}^{*}\lambda=\lim_{n\to\infty}\int_{\mathbb{D}}u_{n}^{*}d\lambda=T_{3}-\gamma>0.

Therefore, u~r\tilde{u}_{r} is nonconstant.

If Γ\Gamma\neq\emptyset, let zΓz\in\Gamma and let znzz_{n}\to z be such that, passing to a subsequence still denoted by u~n\tilde{u}_{n}, we have |du~n(zn)||d\tilde{u}_{n}(z_{n})|\to\infty. Consider a sequence rn0+r_{n}\to 0^{+} such that rn|du~n(zn)|r_{n}|d\tilde{u}_{n}(z_{n})|\to\infty. We need the following lemma.

Lemma 7.2 ([15, Lemma 3.3]).

Let (X,d)(X,d) be a complete metric space and f:X[0,+)f:X\to[0,+\infty) a continuous function. For any ϵ0>0\epsilon_{0}>0 and x0Xx_{0}\in X, there exist ϵ0(0,ϵ0]\epsilon^{\prime}_{0}\in(0,\epsilon_{0}] and x0B2ϵ0(x0)¯x^{\prime}_{0}\in\overline{B_{2\epsilon_{0}}(x_{0})} such that

{f(x0)ϵ0f(x0)ϵ0d(x,x0)ϵ0f(x)2f(x0).\left\{\begin{array}[]{lr}f(x_{0}^{\prime})\epsilon_{0}^{\prime}\geq f(x_{0})\epsilon_{0}\\ d(x,x_{0}^{\prime})\leq\epsilon_{0}^{\prime}\Rightarrow f(x)\leq 2f(x_{0}^{\prime})\end{array}\right..

By Lemma 7.2, we have, after perhaps modifying rnr_{n} and znz_{n},

(109) |du~n(z)|2|du~n(zn)|, for zBrn(zn).|d\tilde{u}_{n}(z)|\leq 2|d\tilde{u}_{n}(z_{n})|,~{}\text{ for }z\in B_{r_{n}}(z_{n}).

Denoting δn=|du~n(zn)|1\delta_{n}=|d\tilde{u}_{n}(z_{n})|^{-1}, define the sequence of J~\tilde{J}-holomorphic maps v~n=(bn,vn):Brnδn(0)×S3\tilde{v}_{n}=(b_{n},v_{n}):B_{\frac{r_{n}}{\delta_{n}}}(0)\to\mathbb{R}\times S^{3} by

(110) v~n(z)=(an(zn+δnz)an(zn),un(zn+δnz)).\tilde{v}_{n}(z)=(a_{n}(z_{n}+\delta_{n}z)-a_{n}(z_{n}),u_{n}(z_{n}+\delta_{n}z)).

From (109) and (110) we get

v~n(0){0}×S3,n\displaystyle\tilde{v}_{n}(0)\in\{0\}\times S^{3},~{}\forall n\in\mathbb{N}
|du~n(z)|2,zBrnδn(0).\displaystyle|d\tilde{u}_{n}(z)|\leq 2,~{}\forall z\in B_{\frac{r_{n}}{\delta_{n}}}(0).

By elliptic estimates, there exists a subsequence, still denoted by v~n\tilde{v}_{n} and a J~\tilde{J}-holomorphic map v~:×S3\tilde{v}:\mathbb{C}\to\mathbb{R}\times S^{3} such that v~nv~\tilde{v}_{n}\to\tilde{v} in Cloc(,×S3)C^{\infty}_{loc}(\mathbb{C},\mathbb{R}\times S^{3}). Since |dv~(0)|=limn|dv~n(0)|=1|d\tilde{v}(0)|=\lim_{n\to\infty}|d\tilde{v}_{n}(0)|=1, we know that v~\tilde{v} is nonconstant. Since E(v~n)E(u~n)=T3E(\tilde{v}_{n})\leq E(\tilde{u}_{n})=T_{3}, we have E(v~)T3E(\tilde{v})\leq T_{3}. Any asymptotic limit of the plane v~\tilde{v} is not linked to P3P_{3} and has action T3\leq T_{3}. We conclude that v~\tilde{v} is asymptotic to either P1P_{1}, P2P_{2} or P3P_{3}. Now we show that v~\tilde{v} is asymptotic either to P3P_{3} or to P2P_{2}. Suppose, by contradiction, that v~\tilde{v} is asymptotic to P1P_{1}. From equation (22), Lemma 2.10 and μ(P1)=1\mu(P_{1})=1, we have wind(v~,)0\operatorname{wind}_{\infty}(\tilde{v},\infty)\leq 0. Using equation (29), we get the contradiction

0windπ(v~)=wind(v~)11.0\leq\operatorname{wind}_{\pi}(\tilde{v})=\operatorname{wind}_{\infty}(\tilde{v})-1\leq-1.

Using Fatou’s Lemma and passing to a subsequence, still denoted by u~n\tilde{u}_{n}, we get

(111) T2v𝑑λlimnBrn(zn)un𝑑λun𝑑λ=T3.T_{2}\leq\int_{\mathbb{C}}v^{*}d\lambda\leq\lim_{n\to\infty}\int_{B_{r_{n}}(z_{n})}u_{n}^{*}d\lambda\leq\int_{\mathbb{C}}u_{n}^{*}d\lambda=T_{3}.

Since T3<2T2T_{3}<2T_{2}, we conclude from (111) that Γ={z}\Gamma=\{z\}. From (106) and elliptic estimates, we find a J~\tilde{J}-holomorphic map

u~r:{z}×S3\tilde{u}_{r}:\mathbb{C}\setminus\{z\}\to\mathbb{R}\times S^{3}

such that, up to a subsequence, u~nu~r\tilde{u}_{n}\to\tilde{u}_{r} and E(u~r)T3E(\tilde{u}_{r})\leq T_{3}. The puncture zz is negative. Indeed, for any ϵ>0\epsilon>0, we have

(112) mϵ(z):=Bϵ(z)urλ=limnBϵ(z)unλ=limnBϵ(z)un𝑑λT2>0,m_{\epsilon}(z):=\int_{\partial B_{\epsilon}(z)}u_{r}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(z)}u_{n}^{*}\lambda=\lim_{n\to\infty}\int_{B_{\epsilon}(z)}u_{n}^{*}d\lambda\geq T_{2}>0,

where Bϵ(z)B_{\epsilon}(z) is oriented counterclockwise. Here we have used (111). The puncture \infty is necessarily positive since 0<E(u~r)T30<E(\tilde{u}_{r})\leq T_{3}. Using (107) we conclude that z=0z=0.

Now we show that u~r\tilde{u}_{r} is asymptotic to P3P_{3} at its positive puncture z=z=\infty and to P2P_{2} at its negative puncture z=0z=0. 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 e>0e>0 and let γ\gamma be defined by (104). Identifying S1=/S^{1}=\mathbb{R}/\mathbb{Z}, let 𝒲C(S1,S3)\mathcal{W}\subset C^{\infty}(S^{1},S^{3}) be an open neighborhood of the periodic orbits P1P_{1}, P2P_{2} and P3P_{3}, viewed as maps xT:S1S3x_{T}:S^{1}\to S^{3}, xT(t)=x(Tt)x_{T}(t)=x(Tt). We assume that 𝒲\mathcal{W} is S1S^{1}-invariant, meaning that y(+c)𝒲y𝒲,cS1y(\cdot+c)\in\mathcal{W}\Leftrightarrow y\in\mathcal{W},\forall c\in S^{1}, and that each of the connected components of 𝒲\mathcal{W} contains at most one periodic orbit modulo S1S^{1}-reparametrizations. Then there exists a constant h>0h>0 such that the following holds. If u~=(a,u):[r,R]×S1×S3\tilde{u}=(a,u):[r,R]\times S^{1}\to\mathbb{R}\times S^{3} is a J~\tilde{J}-holomorphic cylinder such that the image of u~\tilde{u} does not intersect P3P_{3}, u(r,)u(r,\cdot) is not linked to P3P_{3}, and

E(u~)T3,[r,R]×S1u𝑑λγ,{r}×S1uλeandr+hRh,E(\tilde{u})\leq T_{3},~{}~{}~{}\int_{[r,R]\times S^{1}}u^{*}d\lambda\leq\gamma,~{}~{}~{}~{}\int_{\{r\}\times S^{1}}u^{*}\lambda\geq e~{}~{}~{}~{}\text{and}~{}~{}r+h\leq R-h,

then each loop tS1u(s,t)t\in S^{1}\to u(s,t) is contained in 𝒲\mathcal{W} for all s[r+h,Rh]s\in[r+h,R-h].

Proof.

Arguing by contradiction, we find a sequence of J~\tilde{J}-holomorphic maps

u~n=(an,un):[rn,Rn]×S1×S3\tilde{u}_{n}=(a_{n},u_{n}):[r_{n},R_{n}]\times S^{1}\to\mathbb{R}\times S^{3}

such that the image of u~n\tilde{u}_{n} does not intersect P3P_{3}, un(rn,)u_{n}(r_{n},\cdot) is not linked to P3P_{3},

(113) E(u~n)T3,[rn,Rn]×S1un𝑑λγ,{rn}×S1uλe,rn+nRnnE(\tilde{u}_{n})\leq T_{3},~{}~{}~{}~{}\int_{[r_{n},R_{n}]\times S^{1}}u_{n}^{*}d\lambda\leq\gamma,~{}~{}~{}~{}\int_{\{r_{n}\}\times S^{1}}u^{*}\lambda\geq e,~{}~{}~{}~{}r_{n}+n\leq R_{n}-n

and

(114) un(sn,)𝒲 for some sn[rn+n,Rnn].u_{n}(s_{n},\cdot)\notin\mathcal{W}\text{ for some }s_{n}\in[r_{n}+n,R_{n}-n].

Define a sequence of J~\tilde{J}-holomorphic maps v~n=(bn,vn):[rnsn,Rnsn]×S1×S3\tilde{v}_{n}=(b_{n},v_{n}):[r_{n}-s_{n},R_{n}-s_{n}]\times S^{1}\to\mathbb{R}\times S^{3} by

v~n(s,t)=(an(s+sn,t)an(sn,0),un(s+sn,t)).\tilde{v}_{n}(s,t)=(a_{n}(s+s_{n},t)-a_{n}(s_{n},0),u_{n}(s+s_{n},t)).

Note that bn(0,0)=0,nb_{n}(0,0)=0,\forall n\in\mathbb{N} and by (114) we have

(115) vn(0,)𝒲.v_{n}(0,\cdot)\notin\mathcal{W}.

We claim that |dv~n(s,t)||d\tilde{v}_{n}(s,t)| is uniformly bounded in nn\in\mathbb{N} and (s,t)×S1(s,t)\in\mathbb{R}\times S^{1}. Conversely, suppose that there exist a subsequence v~nj\tilde{v}_{n_{j}} and a sequence zjzz_{j}\to z\in\mathbb{C} satisfying |dv~nj(zj)||d\tilde{v}_{n_{j}}(z_{j})|\to\infty. Arguing as in the proof of (111), we get T2limnBrj(zj)vnj𝑑λT_{2}\leq\lim_{n\to\infty}\int_{B_{r_{j}}(z_{j})}v_{n_{j}}^{*}d\lambda, for some sequence rj0+r_{j}\to 0^{+}. This contradicts

[rnsn,Rnsn]×S1vn𝑑λ=[rn,Rn]×S1un𝑑λγ.\int_{[r_{n}-s_{n},R_{n}-s_{n}]\times S^{1}}v_{n}^{*}d\lambda=\int_{[r_{n},R_{n}]\times S^{1}}u_{n}^{*}d\lambda\leq\gamma.

Thus, passing to a subsequence, still denoted by v~n\tilde{v}_{n}, we have v~nv~\tilde{v}_{n}\to\tilde{v} in Cloc(×S1,×S3)C^{\infty}_{loc}(\mathbb{R}\times S^{1},\mathbb{R}\times S^{3}), where v~:×S1×S3\tilde{v}:\mathbb{R}\times S^{1}\to\mathbb{R}\times S^{3} is a finite energy cylinder. Using (113), we conclude that

E(v~)T3,×S1v𝑑λγ and {s}×S1uλe,s.E(\tilde{v})\leq T_{3},~{}~{}~{}\int_{\mathbb{R}\times S^{1}}v^{*}d\lambda\leq\gamma~{}~{}\text{ and }~{}\int_{\{s\}\times S^{1}}u^{*}\lambda\geq e,\forall s\in\mathbb{R}.

Consequently, v~\tilde{v} is nonconstant, has a positive puncture at s=+s=+\infty, a negative puncture at s=s=-\infty and any asymptotic limit of v~\tilde{v} at s=+s=+\infty has period T3\leq T_{3}. Since the loops v~n(s,)\tilde{v}_{n}(s,\cdot) are not linked to P3P_{3}, the asymptotic limits of v~\tilde{v} are not linked to P3P_{3} as well. Therefore, v~\tilde{v} is asymptotic to either P1P_{1}, P2P_{2} or P3P_{3} at s=+s=+\infty. Since ×S1v𝑑λγ\int_{\mathbb{R}\times S^{1}}v^{*}d\lambda\leq\gamma, we conclude that v~\tilde{v} is a cylinder over either P1P_{1}, P2P_{2} or P3P_{3}. However, by (115), we know that v(0,)𝒲v(0,\cdot)\notin\mathcal{W}, a contradiction. ∎

Any asymptotic limit P+=(x+,T+)P_{+}=(x_{+},T_{+}) of u~r\tilde{u}_{r} at z=z=\infty is not linked to P3P_{3} and satisfies T+T3T_{+}\leq T_{3}. Thus, u~r\tilde{u}_{r} is asymptotic to either P1P_{1}, P2P_{2} or P3P_{3} at z=z=\infty. Let 𝒲\mathcal{W} be as in the statement of Lemma 7.3. Using (105) and Lemma 7.3, we conclude that for each nn\in\mathbb{N} and large ss, {tun(s,t)}𝒲\{t\mapsto u_{n}(s,t)\}\in\mathcal{W} and {tur(s,t)}𝒲\{t\mapsto u_{r}(s,t)\}\in\mathcal{W}. Since the planes u~n\tilde{u}_{n} are asymptotic to P3P_{3} and for each fixed ss, {tun(s,t)}𝒲\{t\mapsto u_{n}(s,t)\}\in\mathcal{W} converges to {tur(s,t)}𝒲\{t\mapsto u_{r}(s,t)\}\in\mathcal{W}, as nn\to\infty, we conclude that u~r\tilde{u}_{r} is asymptotic to P3P_{3} at z=z=\infty. Consequently we have Γ\Gamma\neq\emptyset, since Γ=\Gamma=\emptyset would contradict the fact that the family (101) is maximal.

Using (105), we conclude that

(116) Durλ=limn𝔻unλ=limn𝔻un𝑑λ=T3γ.\int_{\partial D}u_{r}^{*}\lambda=\lim_{n\to\infty}\int_{\partial\mathbb{D}}u_{n}^{*}\lambda=\lim_{n\to\infty}\int_{\mathbb{D}}u_{n}^{*}d\lambda=T_{3}-\gamma.

Therefore, any asymptotic limit P=(x,T)𝒫(λ)P_{-}=(x_{-},T_{-})\in\mathcal{P}(\lambda) of u~r\tilde{u}_{r} at z=0z=0 is not linked to P3P_{3} and satisfies T2T<T3T_{2}\leq T_{-}<T_{3}. Here we have used (112) and (116). We conclude that P=P2P_{-}=P_{2}.

Now we proceed as in the soft rescaling done in §3.2.2. Let mϵ(0)m_{\epsilon}(0) be defined as in (39). Since mϵ(0)m_{\epsilon}(0) is a nondecreasing function of ϵ\epsilon, we can fix 0<ϵ<<10<\epsilon<<1 so that

(117) 0mϵ(0)T2γ2.0\leq m_{\epsilon}(0)-T_{2}\leq\frac{\gamma}{2}.

Arguing as in §3.2.2, we choose sequences znBϵ(0)¯z_{n}\in\overline{B_{\epsilon}(0)} and 0<δn<ϵ0<\delta_{n}<\epsilon satisfying

(118) an(zn)an(ζ),ζBϵ(0),\displaystyle a_{n}(z_{n})\leq a_{n}(\zeta),~{}\forall\zeta\in B_{\epsilon}(0),
(119) Bϵ(0)Bδn(zn)un𝑑λ=γ,\displaystyle\int_{B_{\epsilon}(0)\setminus B_{\delta_{n}}(z_{n})}u_{n}^{*}d\lambda=\gamma,

such that zn0z_{n}\to 0 and lim infδn=0\liminf\delta_{n}=0. Thus, passing to a subsequence, we can assume δn0\delta_{n}\to 0.

Take any sequence Rn+R_{n}\to+\infty satisfying δnRn<ϵ2\delta_{n}R_{n}<\frac{\epsilon}{2}~{} and define the sequence of J~\tilde{J}-holomorphic maps w~n=(cn,wn):BRn(0)×S3\tilde{w}_{n}=(c_{n},w_{n}):B_{R_{n}}(0)\to\mathbb{R}\times S^{3} by

(120) w~n(ζ)=(an(zn+δnζ)an(zn+2δn),un(zn+δnζ)).\tilde{w}_{n}(\zeta)=(a_{n}(z_{n}+\delta_{n}\zeta)-a_{n}(z_{n}+2\delta_{n}),u_{n}(z_{n}+\delta_{n}\zeta))~{}.

Note that E(w~n)E(u~n)=T3E(\tilde{w}_{n})\leq E(\tilde{u}_{n})=T_{3} and w~n(2){0}×S3\tilde{w}_{n}(2)\in\{0\}\times S^{3}. Let Γq\Gamma_{q} be the set of points zz\in\mathbb{C} such that there exist subsequence w~nj\tilde{w}_{n_{j}} and sequence zjzz_{j}\to z satisfying |dw~nj(zj)||d\tilde{w}_{n_{j}}(z_{j})|\to\infty. Then there exists a J~\tilde{J}-holomorphic map u~q:Γq×S3\tilde{u}_{q}:\mathbb{C}\setminus\Gamma_{q}\to\mathbb{R}\times S^{3} such that, up to a subsequence, w~nu~q\tilde{w}_{n}\to\tilde{u}_{q} in Cloc(Γq,×S3)C^{\infty}_{loc}(\mathbb{C}\setminus\Gamma_{q},\mathbb{R}\times S^{3}).

We claim that u~q\tilde{u}_{q} is nonconstant and asymptotic to P2P_{2} at its positive puncture z=z=\infty. Indeed, using (119) and (120) we conclude that, for any r1r\geq 1,

(121) Br(0)uqλ=limnBrδn(zn)unλlimnBδn(zn)un𝑑λT2γ,\int_{\partial B_{r}(0)}u_{q}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{r\delta_{n}}(z_{n})}u_{n}^{*}\lambda\geq\lim_{n\to\infty}\int_{B_{\delta_{n}}(z_{n})}u_{n}^{*}d\lambda\geq T_{2}-\gamma,

and by (105), we have

(122) Br(0)uqλ=limnBrδn(zn)unλlimnBδnRn(zn)un𝑑λT3γ.\int_{\partial B_{r}(0)}u_{q}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{r\delta_{n}}(z_{n})}u_{n}^{*}\lambda\leq\lim_{n\to\infty}\int_{B_{\delta_{n}R_{n}}(z_{n})}u_{n}^{*}d\lambda\leq T_{3}-\gamma.

Here we have used δnRn<ϵ2<12\delta_{n}R_{n}<\frac{\epsilon}{2}<\frac{1}{2} and zn0z_{n}\to 0. Therefore, any asymptotic limit P+=(x+,T+)P_{+}=(x_{+},T_{+}) of u~q\tilde{u}_{q} at \infty is not linked to P3P_{3} and satisfies P2γT+T3γP_{2}-\gamma\geq T_{+}\geq T_{3}-\gamma. We conclude that P+=P2P_{+}=P_{2}.

Now we prove that Γq=\Gamma_{q}=\emptyset. Suppose, by contradiction, that zΓqz\in\Gamma_{q} and let znzz_{n}\to z be such that, passing to a subsequence still denoted by w~n\tilde{w}_{n}, |dw~n(zn)||d\tilde{w}_{n}(z_{n})|\to\infty. Arguing as in the proof of (111), we find a sequence rn0+r_{n}\to 0^{+} such that, for each sufficiently small ϵ>0\epsilon>0, we have

Bϵ(z)uqλ=limnBϵ(z)wnλlimnBrn(zn)wn𝑑λT2.\int_{\partial B_{\epsilon}(z)}u_{q}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\epsilon}(z)}w_{n}^{*}\lambda\geq\lim_{n\to\infty}\int_{B_{r_{n}}(z_{n})}w_{n}^{*}d\lambda\geq T_{2}.

We conclude that Γquq𝑑λ=0\int_{\mathbb{C}\setminus\Gamma_{q}}u_{q}^{*}d\lambda=0 and Γq={z}={0}\Gamma_{q}=\{z\}=\{0\}. This implies that u~q\tilde{u}_{q} is a cylinder over the orbit P2P_{2} and we get the contradiction

T2=𝔻uqλ=limn𝔻wnλ=limnBδn(zn)un𝑑λ=mϵ(z)γT2γ2.T_{2}=\int_{\partial\mathbb{D}}u_{q}^{*}\lambda=\lim_{n\to\infty}\int_{\partial\mathbb{D}}w_{n}^{*}\lambda=\lim_{n\to\infty}\int_{B_{\delta_{n}}(z_{n})}u_{n}^{*}d\lambda=m_{\epsilon}(z)-\gamma\leq T_{2}-\frac{\gamma}{2}.

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, uq=uqu_{q}=u^{\prime}_{q};

  • ur({0})ur({0})=u_{r}(\mathbb{C}\setminus\{0\})\cap u_{r}^{\prime}(\mathbb{C}\setminus\{0\})=\emptyset;

  • The union of the image of the family (101) with the images of uqu_{q}, uru_{r}, uru^{\prime}_{r}, x2x_{2} and x3x_{3} determine a singular foliation of a closed region 1\mathcal{R}_{1}, homeomorphic to a solid torus, such that 1=T\partial\mathcal{R}_{1}=T, where T=P2P3ur({0})ur({0})T=P_{2}\cup P_{3}\cup u_{r}(\mathbb{C}\setminus\{0\})\cup u^{\prime}_{r}(\mathbb{C}\setminus\{0\}).

Now we find the foliation of 2=S31¯\mathcal{R}_{2}=S^{3}\setminus\overline{\mathcal{R}_{1}}. Following the same arguments as in Proposition 5.3 we prove that the finite energy cylinder w~\tilde{w} is an embedding. Applying Theorem 3.3 to the finite energy cylinder w~\tilde{w}, we obtain a maximal one-parameter family of finite energy cylinders

(123) w~τ=(cτ,wτ):{0}×S3,τ(0,1)\tilde{w}_{\tau}=(c_{\tau},w_{\tau}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3},~{}~{}\tau\in(0,1)

asymptotic to P3P_{3} at its positive puncture z=z=\infty and to P1P_{1} at its negative puncture z={0}z=\{0\}. For each τ(0,1)\tau\in(0,1), w~τ\tilde{w}_{\tau} is an embedding, the projection wτ:{0}S3w_{\tau}:\mathbb{C}\setminus\{0\}\to S^{3} is an embedding which does not intersect its asymptotic limits, and wτ({0})1=w_{\tau}(\mathbb{C}\setminus\{0\})\cap\mathcal{R}_{1}=\emptyset. We assume that τ\tau strictly increases in the direction of RλR_{\lambda}.

Consider a sequence τn(0,1)\tau_{n}\in(0,1) satisfying τn1\tau_{n}\to 1^{-} and define w~n:=w~τn\tilde{w}_{n}:=\tilde{w}_{\tau_{n}}.

Claim II

There exists a finite energy cylinder

(124) w~r:{0}×S3\tilde{w}_{r}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}

asymptotic to P3P_{3} at its positive puncture z=z=\infty and to P2P_{2} at its negative puncture z=0z=0, and a finite energy cylinder

(125) w~q:{0}×S3\tilde{w}_{q}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}

asymptotic to P2P_{2} at its positive puncture z=z=\infty and to P1P_{1} at its negative puncture z=0z=0, such that, after suitable reparametrizations and \mathbb{R}-translations of w~n\tilde{w}_{n}, we have

  1. (i)

    up to a subsequence, w~nw~r\tilde{w}_{n}\to\tilde{w}_{r} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  2. (ii)

    There exist sequences δn+0+\delta_{n}^{+}\to 0^{+} and dnd_{n}\in\mathbb{R} such that, up to a subsequence, w~n(δn)+dnw~q\tilde{w}_{n}(\delta_{n}\cdot)+d_{n}\to\tilde{w}_{q} in Cloc({0})C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}) as nn\to\infty.

  3. (iii)

    Given an S1S^{1}-invariant neighborhood 𝒲3C(/,S3)\mathcal{W}_{3}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx3(T3t)t\mapsto x_{3}(T_{3}t), there exists R3>>1R_{3}>>1 such that the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲3\mathcal{W}_{3}, for RR3R\geq R_{3} and large nn.

  4. (iv)

    Given an S1S^{1}-invariant neighborhood 𝒲2C(/,S3)\mathcal{W}_{2}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx2(T2t)t\mapsto x_{2}(T_{2}t), there exists ϵ2>0\epsilon_{2}>0 and R2>1R_{2}>1 such that the loop twn(Re2πit)t\mapsto w_{n}(Re^{2\pi it}) belongs to 𝒲2\mathcal{W}_{2}, for R2δnRϵ2R_{2}\delta_{n}\leq R\leq\epsilon_{2} and large nn.

  5. (v)

    Given an S1S^{1}-invariant neighborhood 𝒲1C(/,S3)\mathcal{W}_{1}\subset C^{\infty}(\mathbb{R}/\mathbb{Z},S^{3}) of the loop tx1(T1t)t\mapsto x_{1}(T_{1}t), there exists ϵ1>0\epsilon_{1}>0 such that the loop twn(ρe2πit)t\mapsto w_{n}(\rho e^{2\pi it}) belongs to 𝒲1\mathcal{W}_{1}, for ρϵ1δn\rho\leq\epsilon_{1}\delta_{n} and large nn.

  6. (vi)

    Given any neighborhood 𝒱2\mathcal{V}\subset\mathcal{R}_{2} of wr({0})wq({0})P1P2P3w_{r}(\mathbb{C}\setminus\{0\})\cup w_{q}(\mathbb{C}\setminus\{0\})\cup P_{1}\cup P_{2}\cup P_{3}, we have wn({0})𝒱w_{n}(\mathbb{C}\setminus\{0\})\subset\mathcal{V}, for large nn.

A similar claim holds for any sequence τn0+\tau_{n}\to 0^{+} with w~r\tilde{w}_{r} replaced with a cylinder w~r\tilde{w}_{r}^{\prime} with the same asymptotics as w~r\tilde{w}_{r} and w~q\tilde{w}_{q} replaced with a cylinder w~q\tilde{w}_{q}^{\prime} with the same asymptotics as w~q\tilde{w}_{q}.

After a reparametrization and \mathbb{R}-translation of w~n\tilde{w}_{n}, we can assume that

(126) 𝔻wn𝑑λ=γ2,n\displaystyle\int_{\mathbb{C}\setminus\mathbb{D}}w_{n}^{*}d\lambda=\frac{\gamma}{2},~{}\forall n\in\mathbb{N}
(127) cn(2)=0,n,\displaystyle c_{n}(2)=0,~{}\forall n\in\mathbb{N},

where γ\gamma is defined by (104). Observe that |dw~n(z)||d\tilde{w}_{n}(z)| is uniformlly bounded in nn\in\mathbb{N} and z{0}z\in\mathbb{C}\setminus\{0\}. Otherwise, arguing as in the proof of (111), we would find sequences znz_{n}\in\mathbb{C}, rn0+r_{n}\to 0^{+} and a subsequence of w~n\tilde{w}_{n}, still denoted w~n\tilde{w}_{n}, such that

T2=v𝑑λlimnBrn(zn)wn𝑑λlimnwn𝑑λ=T3T1,T_{2}=\int_{\mathbb{C}}v^{*}d\lambda\leq\lim_{n\to\infty}\int_{B_{r_{n}}(z_{n})}w_{n}^{*}d\lambda\leq\lim_{n\to\infty}\int_{\mathbb{C}}w_{n}^{*}d\lambda=T_{3}-T_{1},

contradicting the hypothesis T3<2T1T_{3}<2T_{1}. We conclude that there exists a finite energy J~\tilde{J}-holomorphic cylinder w~r=(cr,wr):{0}×S3\tilde{w}_{r}=(c_{r},w_{r}):\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3} such that, up to subsequence, still denoted by w~n\tilde{w}_{n}, w~nw~r in Cloc({0},×S3).\tilde{w}_{n}\to\tilde{w}_{r}~{}\text{ in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\},\mathbb{R}\times S^{3}).

The finite energy cylinder w~r\tilde{w}_{r} is nonconstant, z=0z=0 is a negative puncture and z=z=\infty is a positive puncture. Indeed, for any r>0r>0, we have Br(0)wrλ=limnBr(0)wn(T1,T3),\int_{\partial B_{r}(0)}w_{r}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{r}(0)}w_{n}^{*}\in(T_{1},T_{3}), where Br(0)\partial B_{r}(0) is oriented counterclockwise.

It follows from (126) that {0}wr𝑑λγ2>0\int_{\mathbb{C}\setminus\{0\}}w_{r}^{*}d\lambda\geq\frac{\gamma}{2}>0. Using Lemma 7.3 as in the proof of Claim I, we conclude that w~r\tilde{w}_{r} is asymptotic to P3P_{3} at the puncture z=z=\infty. Any asymptotic limit of w~r\tilde{w}_{r} at z=0z=0 is not linked to P3P_{3} and has period <T3<T_{3}. Thus, w~r\tilde{w}_{r} is asymptotic to either P2P_{2} or P1P_{1} at z=0z=0. Since the family (123) is maximal, we conclude that w~r\tilde{w}_{r} is asymptotic to P2P_{2} at the negative puncture z=0z=0.

Now we proceed as in the soft rescaling done in §5.3.2. Arguing as in §5.3.2, we find a sequence δn>0\delta_{n}>0 satisfying

(128) Bδn(0){0}wn𝑑λ=(T2T1)γ2\int_{{B_{\delta_{n}}(0)}\setminus\{0\}}w_{n}^{*}d\lambda=(T_{2}-T_{1})-\frac{\gamma}{2}

and such that lim infδn=0\liminf\delta_{n}=0. Thus, passing to a subsequence still denoted by w~n\tilde{w}_{n}, we can assume that δn0\delta_{n}\to 0.

Fix ϵ>0\epsilon>0 such that

(129) Bϵ(0)wrλT2+γ4\int_{\partial B_{\epsilon}(0)}w_{r}^{*}\lambda\leq T_{2}+\frac{\gamma}{4}

and define the sequence of J~\tilde{J}-holomorphic maps v~n=(bn,vn):Bϵδn(0){0}×S3\tilde{v}_{n}=(b_{n},v_{n}):B_{\frac{\epsilon}{\delta_{n}}}(0)\setminus\{0\}\to\mathbb{R}\times S^{3} by

(130) v~n(z)=(cn(δnz)cn(2δn),wn(δnz)).\tilde{v}_{n}(z)=(c_{n}(\delta_{n}z)-c_{n}(2\delta_{n}),w_{n}(\delta_{n}z)).

Using (128) and (129), we conclude that, for large nn,

(131) Bϵδn(0)𝔻vn𝑑λ\displaystyle\int_{B_{\frac{\epsilon}{\delta_{n}}}(0)\setminus\mathbb{D}}v_{n}^{*}d\lambda =Bϵ(0){0}wn𝑑λBδn(0){0}wn𝑑λ\displaystyle=\int_{B_{\epsilon}(0)\setminus\{0\}}w_{n}^{*}d\lambda-\int_{B_{\delta_{n}}(0)\setminus\{0\}}w_{n}^{*}d\lambda
=Bϵ(0)wnλT1(T2T1γ2)\displaystyle=\int_{\partial B_{\epsilon}(0)}w_{n}^{*}\lambda-T_{1}-\left(T_{2}-T_{1}-\frac{\gamma}{2}\right)
T2+γ2T1(T2T1γ2)=γ\displaystyle\leq T_{2}+\frac{\gamma}{2}-T_{1}-\left(T_{2}-T_{1}-\frac{\gamma}{2}\right)=\gamma

Note that E(v~n)T3E(\tilde{v}_{n})\leq T_{3} and v~n(2){0}×S3\tilde{v}_{n}(2)\in\{0\}\times S^{3}. Moreover, |dv~n(z)||d\tilde{v}_{n}(z)| is uniformly bounded on nn\in\mathbb{N} and z{0}z\in\mathbb{C}\setminus\{0\}. Otherwise, arguing as in the proof of (111), we would find sequences znz{0}z_{n}\to z\in\mathbb{C}\setminus\{0\}, rn0+r_{n}\to 0^{+} and a subsequence of v~n\tilde{v}_{n}, still denoted by v~n\tilde{v}_{n}, such that

T2limnBrn(zn)vn𝑑λlimnBϵ(0){0}wn𝑑λT2+γ2T1,T_{2}\leq\lim_{n\to\infty}\int_{B_{r_{n}}(z_{n})}v_{n}^{*}d\lambda\leq\lim_{n\to\infty}\int_{B_{\epsilon}(0)\setminus\{0\}}w_{n}^{*}d\lambda\leq T_{2}+\frac{\gamma}{2}-T_{1},

a contradiction. Therefore, there exists a finite energy J~\tilde{J}-holomorphic cylinder

w~q:{0}×S3\tilde{w}_{q}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{3}

such that, up to a subsequence, v~nw~q in Cloc({0}).\tilde{v}_{n}\to\tilde{w}_{q}~{}\text{ in }C^{\infty}_{loc}(\mathbb{C}\setminus\{0\}).

For r1r\geq 1 and large nn, we have

(132) T2γ2Bδnr(0)wnλ𝔻wnλ=T3γ2.T_{2}-\frac{\gamma}{2}\leq\int_{\partial B_{\delta_{n}r}(0)}w_{n}^{*}\lambda\leq\int_{\partial\mathbb{D}}w_{n}^{*}\lambda=T_{3}-\frac{\gamma}{2}.

Here we have used (126) and (128). Therefore, we have

(133) Br(0)wqλ=limnBr(0)vnλ=limnBδnr(0)wnλ[T2γ2,T3γ2].\int_{\partial B_{r}(0)}w_{q}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{r}(0)}v_{n}^{*}\lambda=\lim_{n\to\infty}\int_{\partial B_{\delta_{n}r}(0)}w_{n}^{*}\lambda\in\left[T_{2}-\frac{\gamma}{2},T_{3}-\frac{\gamma}{2}\right].

We conclude that w~q\tilde{w}_{q} is asymptotic to P2P_{2} at the positive puncture z=z=\infty.

Using (128), we conclude that

{0}wq𝑑λ𝔻wq𝑑λ=T2limn𝔻vnλ=T2limnBδn(0)wn𝑑λ=γ2>0.\int_{\mathbb{C}\setminus\{0\}}w_{q}^{*}d\lambda\geq\int_{\mathbb{C}\setminus\mathbb{D}}w_{q}^{*}d\lambda=T_{2}-\lim_{n\to\infty}\int_{\partial\mathbb{D}}v_{n}^{*}\lambda=T_{2}-\lim_{n\to\infty}\int_{\partial B_{\delta_{n}}(0)}w_{n}^{*}d\lambda=\frac{\gamma}{2}>0.

Thus, any asymptotic limit of w~q\tilde{w}_{q} at its negative puncture z=0z=0 is an orbit that is not linked to P3P_{3} and has period <T2<T_{2}. We conclude that w~q\tilde{w}_{q} is asymptotic to P1P_{1} at z=0z=0. 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 \mathbb{R}-translation, we have w~r=u~r\tilde{w}_{r}=\tilde{u}_{r} and w~r=u~r\tilde{w}^{\prime}_{r}=\tilde{u}_{r}^{\prime}, where u~r\tilde{u}_{r} and u~r\tilde{u}_{r}^{\prime} are given by Claim I;

  • Up to reparametrization and \mathbb{R}-translation we have w~q=w~q\tilde{w}_{q}=\tilde{w}^{\prime}_{q};

  • The images of the family {wτ},τ(τ,+)\{w_{\tau}\},\tau\in(\tau_{-},+\infty), uru_{r}, uru_{r}^{\prime}, wqw_{q}, x1x_{1}, x2x_{2} and x3x_{3} determine a singular foliation of the region 2=S31¯\mathcal{R}_{2}=\overline{S^{3}\setminus\mathcal{R}_{1}}.

  • sl(Pi)=1{\rm sl}(P_{i})=-1, i=1,2,3i=1,2,3.

It follows that there exists a 3213-2-1 foliation with binding orbits P1P_{1}, P2P_{2} and P3P_{3}. The existence of a homoclinic to P2P_{2} 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.

For sufficiently small ϵ\epsilon, the contact form λ=λ0|S\lambda=\lambda_{0}|_{S}, where S=H1(12)S=H^{-1}(\frac{1}{2}), and the Reeb orbits P1P_{1}, P2P_{2}, and P3P_{3}, defined by (9)-(10), satisfy the hypotheses of Theorem 1.6. Therefore, there exists a 3213-2-1 foliation adapted to λ\lambda with binding orbits P1P_{1}, P2P_{2}, and P3P_{3}.

The proof of Proposition 1.7 follows immediately from (11), (12) and Lemmas 8.2-8.5 below.

Lemma 8.2.

Assuming that ϵ\epsilon is sufficiently small, the orbits P1P_{1}, P2P_{2} and P3P_{3} are nondegenerate and their Conley-Zehnder indices are respectively 1,21,~{}2 and 33.

Proof.

First we define a trivialization of ξ=kerλ\xi=\ker\lambda. Consider the matrices A0=IdA_{0}={\rm{Id}},

A1=[0001001001001000],A2=[0010000110000100],A3=[0100100000010010].A_{1}=\left[\begin{matrix}0&0&0&1\\ 0&0&1&0\\ 0&-1&0&0\\ -1&0&0&0\end{matrix}\right],~{}A_{2}=\left[\begin{matrix}0&0&-1&0\\ 0&0&0&1\\ 1&0&0&0\\ 0&-1&0&0\end{matrix}\right],~{}A_{3}=\left[\begin{matrix}0&-1&0&0\\ 1&0&0&0\\ 0&0&0&-1\\ 0&0&1&0\end{matrix}\right].

For each zSz\in S we have an ortonormal frame of Tz44T_{z}\mathbb{R}^{4}\simeq\mathbb{R}^{4} given by {Xi(z)},i=0,,3\{X_{i}(z)\},~{}i=0,\dotsc,3, where

(134) Xi:=Ai(H(z)H(z)),i=0,1,2,3.X_{i}:=A_{i}\left(\frac{\nabla H(z)}{\|\nabla H(z)\|}\right),~{}i=0,1,2,3.

Note that X3(z)=XH(z)H(z)X_{3}(z)=\frac{X_{H}(z)}{\|\nabla H(z)\|} and TzS=span{X1,X2,X3}T_{z}S={\rm span}\{X_{1},X_{2},X_{3}\}.

The contact structure ξ=kerλ\xi=\ker\lambda is isomorphic to span{X1,X2}{\rm span}\{X_{1},X_{2}\}, as a tangent hyperplane distribution, via the projection π12:TSspan{X1,X2}\pi_{12}:TS\to{\rm span}\{X_{1},X_{2}\} along X3X_{3}. Let X¯iξ\bar{X}_{i}\in\xi be the vector field determined by

(135) π12(X¯i)=Xi,i=1,2.\pi_{12}(\bar{X}_{i})=X_{i},~{}i=1,2.

We can define a symplectic trivialization of ξ\xi by

(136) Ψ:S×2ξ;Ψ(z,α1,α2)=α1X¯1(z)+α2X¯2(z).\Psi:S\times\mathbb{R}^{2}\to\xi;~{}~{}\Psi(z,\alpha_{1},\alpha_{2})=\alpha_{1}\bar{X}_{1}(z)+\alpha_{2}\bar{X}_{2}(z).

Note that along the orbits PiP_{i}, i=1,2,3i=1,2,3, span{X1,X2}{\rm span}\{X_{1},X_{2}\} coincides with the plane (x2,y2)(x_{2},y_{2}). Indeed, along the orbits, we have P=Q=0P=Q=0 and

(137) X1(z)\displaystyle X_{1}(z) =1x12+y12(0,0,y1,x1)\displaystyle=\frac{1}{x_{1}^{2}+y_{1}^{2}}(0,0,-y_{1},-x_{1})
X2(z)\displaystyle X_{2}(z) =1x12+y12(0,0,x1,y1).\displaystyle=\frac{1}{x_{1}^{2}+y_{1}^{2}}(0,0,x_{1},-y_{1}).

We can define another trivialization along the orbits PiP_{i}, i=1,2,3i=1,2,3, by

(138) ρi:×2xiξ;(t,β1,β2)(π121)xi(t)(β1(0,0,1,0)+β2(0,0,0,1)).\rho_{i}:\mathbb{R}\times\mathbb{R}^{2}\to x_{i}^{*}\xi;~{}~{}(t,\beta_{1},\beta_{2})\mapsto\left(\pi_{12}^{-1}\right)_{x_{i}(t)}(\beta_{1}(0,0,1,0)+\beta_{2}(0,0,0,1)).

It is easy to check that ρi\rho_{i}i=1,2,3i=1,2,3, is a symplectic trivialization. A simple calculation shows that, in the trivialization ρi\rho_{i}i=1,2,3i=1,2,3, dRλ(γi(t))dR_{\lambda}(\gamma_{i}(t)) is represented by

(139) dRλ(γi(t))=[0h(γi(t))Py2(pi,0)h(γi(t))Qx2(pi,0)0]=:[0k1ik2i0]dR_{\lambda}(\gamma_{i}(t))=\left[\begin{matrix}0&-h(\gamma_{i}(t))\dfrac{\partial P}{\partial y_{2}}(p_{i},0)\\ h(\gamma_{i}(t))\dfrac{\partial Q}{\partial x_{2}}(p_{i},0)&0\end{matrix}\right]=:\left[\begin{matrix}0&k_{1}^{i}\\ k_{2}^{i}&0\end{matrix}\right]

which is a constant linear map. Note that, for sufficiently small ϵ>0\epsilon>0, we have 0<|k1i|,|k2i|<<1,i=1,2,30<|k^{i}_{1}|,|k_{2}^{i}|<<1,~{}~{}i=1,2,3.

Let φt\varphi^{t} be the flow of RλR_{\lambda}. The linearized flow dφt(x(0))d\varphi^{t}(x(0)) along a trajectory x(t)x(t) is the solution of the equation

(140) ddtdφt(x(0))=dRλ(x(t))dφt(x(0)).\dfrac{d}{dt}d\varphi^{t}(x(0))=dR_{\lambda}(x(t))\cdot d\varphi^{t}(x(0)).

For i=1i=1, we have k11>0k_{1}^{1}>0 and k21<0k_{2}^{1}<0. Solving equation (140) we find

(141) dφx1(0)t=[cos(k11k21t)k11k21sin(k11k21t)k21k11sin(k11k21t)cos(k11k21t)].d\varphi^{t}_{x_{1}(0)}=\left[\begin{matrix}\cos(\sqrt{-k_{1}^{1}k_{2}^{1}}t)&\frac{\sqrt{k_{1}^{1}}}{\sqrt{-k_{2}^{1}}}\sin(\sqrt{-k_{1}^{1}k_{2}^{1}}t)\\ -\frac{\sqrt{-k_{2}^{1}}}{\sqrt{k_{1}^{1}}}\sin(\sqrt{-k_{1}^{1}k_{2}^{1}}t)&\cos(\sqrt{-k_{1}^{1}k_{2}^{1}}t)\end{matrix}\right].

Thus, if ϵ\epsilon is sufficiently small, P1P_{1} is nondegenerate and for any z2z\in\mathbb{R}^{2}, z(t):=dφx1(0)tzz(t):=d\varphi^{t}_{x_{1}(0)}z has winding number 1<Δ(z)<0-1<\Delta(z)<0, where Δ(z)\Delta(z) is defined by (13). We conclude that μ(P1,ρ1)=1\mu(P_{1},\rho_{1})=-1.

For i=2i=2, we have k12>0k_{1}^{2}>0 and k22>0k_{2}^{2}>0. Solving equation (140) we find

(142) dφx2(0)t=[cosh(k12k22t)k12k22sinh(k12k22t)k22k12sinh(k12k22t)cosh(k12k22t)].d\varphi^{t}_{x_{2}(0)}=\left[\begin{matrix}\cosh(\sqrt{k_{1}^{2}k_{2}^{2}}t)&\frac{\sqrt{k_{1}^{2}}}{\sqrt{k_{2}^{2}}}\sinh(\sqrt{k_{1}^{2}k_{2}^{2}}t)\\ \frac{\sqrt{k_{2}^{2}}}{\sqrt{k_{1}^{2}}}\sinh(\sqrt{k_{1}^{2}k_{2}^{2}}t)&\cosh(\sqrt{k_{1}^{2}k_{2}^{2}}t)\end{matrix}\right].

Note that

(143) dφx2(0)t(k12k22,1)\displaystyle d\varphi^{t}_{x_{2}(0)}\left(\frac{\sqrt{k_{1}^{2}}}{\sqrt{k_{2}^{2}}},1\right) =ek12k22t(k12k22,1),t,\displaystyle=e^{\sqrt{k_{1}^{2}k_{2}^{2}}t}\left(\frac{\sqrt{k_{1}^{2}}}{\sqrt{k_{2}^{2}}},1\right),~{}~{}\forall t\in\mathbb{R},
dφx2(0)t(1,k22k12)\displaystyle d\varphi^{t}_{x_{2}(0)}\left(-1,\frac{\sqrt{k_{2}^{2}}}{\sqrt{k_{1}^{2}}}\right) =ek12k22t(1,k22k12),t.\displaystyle=e^{-\sqrt{k_{1}^{2}k_{2}^{2}}t}\left(-1,\frac{\sqrt{k_{2}^{2}}}{\sqrt{k_{1}^{2}}}\right),~{}~{}\forall t\in\mathbb{R}.

We conclude that P2P_{2} is nondegenerate and μ(P2,ρ2)=0\mu(P_{2},\rho_{2})=0.

For i=3i=3, we have k13<0k_{1}^{3}<0 and k23>0k_{2}^{3}>0. Solving equation (140) we find

(144) dφx3(0)t=[cos(k13k23t)k13k23sin(k13k23t)k23k13sin(k13k23t)cos(k13k23t)]d\varphi^{t}_{x_{3}(0)}=\left[\begin{matrix}\cos(\sqrt{-k_{1}^{3}k_{2}^{3}}t)&-\frac{\sqrt{-k_{1}^{3}}}{\sqrt{k_{2}^{3}}}\sin(\sqrt{-k_{1}^{3}k_{2}^{3}}t)\\ \frac{\sqrt{k_{2}^{3}}}{\sqrt{-k_{1}^{3}}}\sin(\sqrt{-k_{1}^{3}k_{2}^{3}}t)&\cos(\sqrt{-k_{1}^{3}k_{2}^{3}}t)\end{matrix}\right]

Thus, if ϵ\epsilon is sufficiently small, P3P_{3} is nondegenerate and for any z2z\in\mathbb{R}^{2}, z(t):=dφx3(0)tzz(t):=d\varphi^{t}_{x_{3}(0)}z has winding number 0<Δ(z)<10<\Delta(z)<1. We conclude that μ(P3,ρ3)=1\mu(P_{3},\rho_{3})=1.

Now, the Conley-Zehnder index of the orbit PiP_{i}, i=1,2,3i=1,2,3, satisfies the formula

μ(Pi)=μ(Pi,Ψ)=μ(Pi,ρi)+2wind((ρi)t(1,0),Ψ)\mu(P_{i})=\mu(P_{i},\Psi)=\mu(P_{i},\rho_{i})+2\operatorname{wind}((\rho_{i})_{t}(1,0),\Psi)

and

(145) wind((ρi)t(1,0),Ψ)\displaystyle\operatorname{wind}((\rho_{i})_{t}(1,0),\Psi) =deg((π121)γi(t)(0,0,1,0),Ψ)\displaystyle=\deg\left((\pi_{12}^{-1})_{\gamma_{i}(t)}(0,0,1,0),\Psi\right)
=deg(/TitΨγi(t)1(π121)γi(t)(0,0,1,0))\displaystyle=\deg\left(\mathbb{R}/T_{i}\mathbb{Z}\ni t\mapsto\Psi^{-1}_{\gamma_{i}(t)}(\pi_{12}^{-1})_{\gamma_{i}(t)}(0,0,1,0)\right)
=deg(/Tit(risin(2ri2t),ricos(2ri2t)))\displaystyle=\deg\left(\mathbb{R}/T_{i}\mathbb{Z}\ni t\mapsto\left(-r_{i}\sin\left(\frac{2}{r_{i}^{2}}t\right),r_{i}\cos\left(\frac{2}{r_{i}^{2}}t\right)\right)\right)
=1.\displaystyle=1.

We conclude that μ(P1)=1\mu(P_{1})=1, μ(P2)=2\mu(P_{2})=2 and μ(P3)=3\mu(P_{3})=3. ∎

Lemma 8.3.

There exists J𝒥(λ)J\in\mathcal{J}(\lambda) such that the almost complex structure J~=(λ,J)\tilde{J}=(\lambda,J) admits a finite energy plane u~:×S\tilde{u}:\mathbb{C}\to\mathbb{R}\times S asymptotic to P3P_{3} at its positive puncture z=z=\infty and a finite energy cylinder v~:{0}×S\tilde{v}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S asymptotic to P3P_{3} at its positive puncture z=z=\infty and P1P_{1} at its negative puncture z=0z=0.

Proof.

Our proof follows [7, §5]. Define a dλd\lambda-compatible complex structure J:ξξJ:\xi\to\xi by

(146) JX¯1=X¯2J\bar{X}_{1}=\bar{X}_{2}

where X¯i\bar{X}_{i}, i=1,2i=1,2 is defined by (135). Let J~=(λ,J)\tilde{J}=(\lambda,J) be the almost complex structure on ×S\mathbb{R}\times S defined by (23).

First we look for the J~\tilde{J}-holomorphic plane. Our candidate for the plane u~\tilde{u} will project into the surface

(147) {(x1,y1,x2,0)4|x2>p3}S.\{(x_{1},y_{1},x_{2},0)\in\mathbb{R}^{4}~{}|~{}x_{2}>p_{3}\}\cap S.

We write u~(s,t)=u~(e2π(s+it))\tilde{u}(s,t)=\tilde{u}(e^{2\pi(s+it)}) for (s,t)×S1(s,t)\in\mathbb{R}\times S^{1}. First we assume that u~=(a(s,t),u(s,t)):×S1×S\tilde{u}=(a(s,t),u(s,t)):\mathbb{R}\times S^{1}\to\mathbb{R}\times S satisfies

(148) u(s,t)=(f(s)cos2πt,f(s)sin2πt,g(s),0),u(s,t)=(f(s)\cos 2\pi t,f(s)\sin 2\pi t,g(s),0),

where f:(0,+)f:\mathbb{R}\to(0,+\infty) and g:{0}g:\mathbb{R}\to\mathbb{R}\setminus\{0\} are smooth functions to be determined.

Recall that u~\tilde{u} is J~\tilde{J}-holomorphic if and only if it satisfies

(149) πus+Jπut=0\displaystyle\pi u_{s}+J\pi u_{t}=0
(150) dai=uλ,\displaystyle da\circ i=-u^{*}\lambda,

where π:TSξ\pi:TS\to\xi is the projection along the Reeb vector field RλR_{\lambda}.

Now we find conditions on f(s)f(s) and g(s)g(s) so that u(s,t)u(s,t) defined by (148) satisfies (149). First note that

(151) πus\displaystyle\pi u_{s} =usλ(us)Rλ\displaystyle=u_{s}-\lambda(u_{s})R_{\lambda}
=(f(s)cos2πt,f(s)sin2πt,g(t),0)\displaystyle=(f^{\prime}(s)\cos 2\pi t,f^{\prime}(s)\sin 2\pi t,g^{\prime}(t),0)
πut\displaystyle\pi u_{t} =utλ(ut)Rλ\displaystyle=u_{t}-\lambda(u_{t})R_{\lambda}
=(2πf(s)sin2πt,2πf(s)cos2πt,0,0)\displaystyle=(-2\pi f(s)\sin 2\pi t,2\pi f(s)\cos 2\pi t,0,0)
πf(s)2h(f(s)sin2πt,f(s)cos2πt,0,Q).\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}-\pi f(s)^{2}h(-f(s)\sin 2\pi t,f(s)\cos 2\pi t,0,Q).

Restricting the frame {X¯1,X¯2}\{\bar{X}_{1},\bar{X}_{2}\} to the surface (147) we find

(152) X¯1\displaystyle\bar{X}_{1} =X1λ(X1)Rλ\displaystyle=X_{1}-\lambda(X_{1})R_{\lambda}
=(0,Q,y1,x1)h(x1Qx2x1)(y1,x1,0,Q)\displaystyle=(0,Q,-y_{1},-x_{1})-h(x_{1}Q-x_{2}x_{1})(-y_{1},x_{1},0,Q)
X¯2\displaystyle\bar{X}_{2} =X2λ(X2)Rλ\displaystyle=X_{2}-\lambda(X_{2})R_{\lambda}
=(Q,0,x1,y1)h(y1Qx2y1)(y1,x1,0,Q).\displaystyle=(-Q,0,x_{1},-y_{1})-h(y_{1}Q-x_{2}y_{1})(-y_{1},x_{1},0,Q).

Using (151) and (152) we find

(153) X¯1(u(s,t))\displaystyle\bar{X}_{1}(u(s,t)) =y1g(s)πus+x1(1+h(Qx2)Q)hπf(s)2Qπut\displaystyle=-\frac{y_{1}}{g^{\prime}(s)}\pi u_{s}+x_{1}\frac{(1+h(Q-x_{2})Q)}{h\pi f(s)^{2}Q}\pi u_{t}
X¯2(u(s,t))\displaystyle\bar{X}_{2}(u(s,t)) =x1g(s)πus+y1(1+h(Qx2)Q)hπf(s)2Qπut.\displaystyle=\frac{x_{1}}{g^{\prime}(s)}\pi u_{s}+y_{1}\frac{(1+h(Q-x_{2})Q)}{h\pi f(s)^{2}Q}\pi u_{t}.

Now using the definition of JJ (146) and equation (149), we find the differential equation

(154) g(s)=hπf(s)2Q1+h(Qx2)Q.g^{\prime}(s)=\frac{-h\pi f(s)^{2}Q}{1+h(Q-x_{2})Q}.

Since the image of uu is the surface (147), we have

(155) 12f(s)2=12(g(s)42+ϵag(s)3+ϵ2cg(s)2).\frac{1}{2}f(s)^{2}=\frac{1}{2}-\left(\frac{g(s)^{4}}{2}+\epsilon ag(s)^{3}+\epsilon^{2}cg(s)^{2}\right).

Moreover, h(u(s,t))=2(f(s)2+g(s)Q(g(s),0))1h(u(s,t))=2(f(s)^{2}+g(s)Q(g(s),0))^{-1}. Thus, we may view (154) as an ODE of the type

(156) g(s)=G(g(s)),g^{\prime}(s)=G(g(s)),

where G:[p3,x¯]G:[p_{3},\bar{x}]\to\mathbb{R} is a smooth function. Here x¯\bar{x} is the unique positive solution of H2(x,0)=12H_{2}(x,0)=\frac{1}{2}.

Note that GG vanishes in g=p3g=p_{3} and in g=x¯g=\bar{x}, and G is negative on (p3,x¯)(p_{3},\bar{x}). Thus, g(x)<0g^{\prime}(x)<0 for g(s)(p3,x¯)g(s)\in(p_{3},\bar{x}). Consequently g(s)g(s) is strictly decreasing and satisfies

(157) limsg(s)=x¯\displaystyle\lim_{s\to-\infty}g(s)=\bar{x} lims+g(s)=p3.\displaystyle\lim_{s\to+\infty}g(s)=p_{3}.

It follows from (155) and (157) that

(158) limsf(s)=0\displaystyle\lim_{s\to-\infty}f(s)=0 lims+f(s)=r3,\displaystyle\lim_{s\to+\infty}f(s)=r_{3},

where r3r_{3} is defined by (9). Using (157) and (158) we conclude that the loops S1tu(s,t)S^{1}\ni t\mapsto u(s,t) converge to S1tx3(T3t)S^{1}\ni t\mapsto x_{3}(T_{3}t) in CC^{\infty} as s+s\to+\infty. Moreover, the loops S1tu(s,t)S^{1}\ni t\mapsto u(s,t) converge to (0,0,x¯,0)(0,0,\bar{x},0) in CC^{\infty} as ss\to\infty.

Now we define the function a:×S1a:\mathbb{R}\times S^{1}\to\mathbb{R} by

(159) a(s,t)=π0sf(τ)𝑑τ.a(s,t)=\pi\int_{0}^{s}f(\tau)d\tau.

Then

(160) as(s,t)\displaystyle a_{s}(s,t) =πf(s)2=λ(ut)\displaystyle=\pi f(s)^{2}=\lambda(u_{t})
at(s,t)\displaystyle a_{t}(s,t) =0=λ(us).\displaystyle=0=-\lambda(u_{s}).

Consider the map u~=(a,u):S1××S\tilde{u}=(a,u):S^{1}\times\mathbb{R}\to\mathbb{R}\times S with uu defined by (148), where gg is a solution of (154) satisfying g(0)(p3,x¯)g(0)\in(p_{3},\bar{x}), ff is defined by (155), and aa defined by (159). Since u~=(a,u)\tilde{u}=(a,u) satisfies (149) and (150), u~\tilde{u} is J~\tilde{J}-holomorphic. Moreover, u~\tilde{u} has finite energy. Indeed, using Stokes theorem and equations (158) and (159) we find

(161) ×S1u~d(ϕλ)\displaystyle\int_{\mathbb{R}\times S^{1}}\tilde{u}^{*}d(\phi\lambda) =limR+(ϕ(a(R))πf(R)2ϕ(a(R))πf(R)2)\displaystyle=\lim_{R\to+\infty}\left(\phi(a(R))\pi f(R)^{2}-\phi(a(-R))\pi f(-R)^{2}\right)
=limR+ϕ(R)πr32\displaystyle=\lim_{R\to+\infty}\phi(R)\pi r_{3}^{2}
=limR+ϕ(R)T3\displaystyle=\lim_{R\to+\infty}\phi(R)T_{3}

where ϕλ(a,u)=ϕ(a)λ(u)\phi\lambda(a,u)=\phi(a)\lambda(u), ϕ:[0,1]\phi:\mathbb{R}\to[0,1] is smooth and ϕ0\phi^{\prime}\geq 0. We conclude that E(u~)=T3E(\tilde{u})=T_{3}.

The mass m()m(-\infty) of the singularity z=z=-\infty satisfies

m()=lims{s}×S1uλ=limsπf(s)2=0.m(-\infty)=\lim_{s\to-\infty}\int_{\{s\}\times S^{1}}u^{*}\lambda=\lim_{s\to-\infty}\pi f(s)^{2}=0.

This implies that u~\tilde{u} has a removable singularity at s=s=-\infty. Removing it, we obtain a finite energy J~\tilde{J}-holomorphic plane u~:×S1\tilde{u}:\mathbb{C}\to\mathbb{R}\times S^{1} asymptotic to P3P_{3} at its positive puncture z=z=\infty.

The construction of the cylinder v~:{0}×S1\tilde{v}:\mathbb{C}\setminus\{0\}\to\mathbb{R}\times S^{1} is completely analogous, but we start with a solution of (154) satisfying g(0)(p1,p3)g(0)\in(p_{1},p_{3}). In this case, we obtain

limsg(s)=p1\displaystyle\lim_{s\to-\infty}g(s)=p_{1} lims+g(s)=p3\displaystyle\lim_{s\to+\infty}g(s)=p_{3}
limsf(s)=r1\displaystyle\lim_{s\to-\infty}f(s)=r_{1} lims+f(s)=r3,\displaystyle\lim_{s\to+\infty}f(s)=r_{3},

and the singularity z=z=-\infty is non-removable, since

m()=lims{s}×S1vλ=limsπf(s)2=πr12=T1.m(-\infty)=\lim_{s\to-\infty}\int_{\{s\}\times S^{1}}v^{*}\lambda=\lim_{s\to-\infty}\pi f(s)^{2}=\pi r_{1}^{2}=T_{1}.

Lemma 8.4.

Assuming that ϵ\epsilon is sufficiently small, if P=(γ,T)𝒫(λ)P=(\gamma,T)\in\mathcal{P}(\lambda) satisfies PP3,TT3P\neq P_{3},~{}T\leq T_{3} and lk(P,P3)=0{\rm lk}(P,P_{3})=0, then P{P1,P2}P\in\{P_{1},P_{2}\}.

Proof.

First we prove the following claim.

Claim I

Let γ:2n\gamma:\mathbb{R}\to\mathbb{R}^{2n} be a nonconstant TT-periodic solution of γ˙(t)=XH(γ(t))\dot{\gamma}(t)=X_{H}(\gamma(t)), where H:2nH:\mathbb{R}^{2n}\to\mathbb{R}. Then hT2πhT\geq 2\pi, where h:=supt[0,T]|d2H(γ(t))|h:=\sup_{t\in[0,T]}|d^{2}H(\gamma(t))|.

To prove the claim, first we note that γ˙\dot{\gamma} satisfies

(162) γ˙L2T2πγ¨L2.\|\dot{\gamma}\|_{L^{2}}\leq\frac{T}{2\pi}\|\ddot{\gamma}\|_{L^{2}}.

This follows from the fact that 0Tγ˙=0\int_{0}^{T}\dot{\gamma}=0 and from the representations of γ˙\dot{\gamma} and γ¨\ddot{\gamma} as Fourier series. Now we have

(163) γ¨L2=(dXHγ)γ˙L2(0T|dXH(γ(t))|2γ˙(t)2𝑑t)12hγ˙L2.\|\ddot{\gamma}\|_{L^{2}}=\|(dX_{H}\circ\gamma)\dot{\gamma}\|_{L^{2}}\leq\left(\int_{0}^{T}|dX_{H}(\gamma(t))|^{2}\|\dot{\gamma}(t)\|^{2}dt\right)^{\frac{1}{2}}\leq h\|\dot{\gamma}\|_{L^{2}}.

We conclude from (162) and (163) that hT2πhT\geq 2\pi.

Consider the function 3(ϵ,x,y)H2ϵ(x,y)\mathbb{R}^{3}\ni(\epsilon,x,y)\mapsto H_{2}^{\epsilon}(x,y), where

H2ϵ(x,y)=(x2+y2)2ϵ(x2+y2)xϵ2(x2+y2).H_{2}^{\epsilon}(x,y)=(x^{2}+y^{2})^{2}-\epsilon(x^{2}+y^{2})x-\frac{\epsilon}{2}(x^{2}+y^{2}).

Then 3(ϵ,x,y)d2(H2ϵ)(x,y)\mathbb{R}^{3}\ni(\epsilon,x,y)\mapsto d^{2}(H_{2}^{\epsilon})(x,y) is continuous and d2(H20(0,0))=0d^{2}(H_{2}^{0}(0,0))=0. Hence there exists r>0r>0 such that, for 0<ϵ<r0<\epsilon<r and (x,y)Br(0)2(x,y)\in B_{r}(0)\subset\mathbb{R}^{2},

(164) |d2H2ϵ(x,y)|<12.|d^{2}H_{2}^{\epsilon}(x,y)|<\frac{1}{2}.

For ϵ>0\epsilon>0 sufficiently small, we also have

(165) Aϵ:={(x,y)2|H2ϵ(x,y)H2ϵ(p1,0)}Br(0).A_{\epsilon}:=\{(x,y)\in\mathbb{R}^{2}~{}|~{}H_{2}^{\epsilon}(x,y)\leq H_{2}^{\epsilon}(p_{1},0)\}\subset B_{r}(0).

Indeed, using the fact that (ϵ,x,y)H2ϵ(x,y)(\epsilon,x,y)\mapsto H_{2}^{\epsilon}(x,y) is a proper function, we know that the set AϵA_{\epsilon} is bounded by a uniform constant for every 0<ϵ<r0<\epsilon<r. Thus, for sufficiently small ϵ\epsilon and (x,y)Aϵ(x,y)\in A_{\epsilon}, we have

(166) (x,y)4\displaystyle\|(x,y)\|^{4} =(x2+y2)2\displaystyle=(x^{2}+y^{2})^{2}
H2ϵ(p1,0)+ϵ(x2+y2)x+ϵ2(x2+y2)\displaystyle\leq H_{2}^{\epsilon}(p_{1},0)+\epsilon(x^{2}+y^{2})x+\frac{\epsilon}{2}(x^{2}+y^{2})
r4.\displaystyle\leq r^{4}.

Fix ϵ>0\epsilon>0 so that (165) is satisfied and consider the Hamiltonian function H=HϵH=H_{\epsilon} defined by (4). Let P=(γ,T)𝒫(λ){P1,P2,P3}P=(\gamma,T)\in\mathcal{P}(\lambda)\setminus\{P_{1},P_{2},P_{3}\} be a simple orbit and let z(t)z(t) be a periodic solution of z˙(t)=XH(z(t))\dot{z}(t)=X_{H}(z(t)) that is a reparametrization of γ\gamma. Then z=z1×z2z=z_{1}\times z_{2}, where z1˙(t)=XH1(z1(t))\dot{z_{1}}(t)=X_{H_{1}}(z_{1}(t)), z2˙(t)=XH2(z2(t))\dot{z_{2}}(t)=X_{H_{2}}(z_{2}(t)) and z2(t)z_{2}(t) is nonconstant. Now we prove the following claim.

Claim II

If z2z_{2} satisfies

(167) H2z2CH2(p1,0),H_{2}\circ z_{2}\equiv C\leq H_{2}(p_{1},0),

then the period TT of PP satisfies T<T3T<T_{3}.

Using Claim I and equation (164) we conclude that the period T~2\tilde{T}_{2} of z2z_{2} satisfies

4π2hT~2<T~2.4\pi\leq 2h\cdot\tilde{T}_{2}<\tilde{T}_{2}.

Note that the orbit z1(t)=(12Ccos(t+t0),12Csin(t+t0))z_{1}(t)=(\sqrt{1-2C}\cos(t+t_{0}),\sqrt{1-2C}\sin(t+t_{0})) is nonconstant, since CH2(p1,0)<12C\leq H_{2}(p_{1},0)<\frac{1}{2}, and 2π2\pi-periodic. This implies that the period T~\tilde{T} of zz satisfies T~=k2π\tilde{T}=k2\pi, for some k>2k>2. The period TT of γ\gamma satisfies

(168) T\displaystyle T =γ()λ\displaystyle=\int_{\gamma(\mathbb{R})}\lambda
=0T~12z1(x1dy1y1dx1)+0T~12z2(x2dy2y2dx2)\displaystyle=\int_{0}^{\tilde{T}}\frac{1}{2}z_{1}^{*}(x_{1}dy_{1}-y_{1}dx_{1})+\int_{0}^{\tilde{T}}\frac{1}{2}z_{2}^{*}(x_{2}dy_{2}-y_{2}dx_{2})
>2π(12C)+area(R2)\displaystyle>2\pi(1-2C)+{\rm area}(R_{2})
2π(12H2(p1,0))\displaystyle\geq 2\pi(1-2H_{2}(p_{1},0))
=2T1\displaystyle=2T_{1}

where R2R_{2} is the region in the plane (x2,y2)(x_{2},y_{2}) limited by the image of z2z_{2}. Here we have used (10) and (167). Taking a smaller ϵ\epsilon if necessary and using (12), we have

T>2T1>T3,T>2T_{1}>T_{3},

proving the claim.

Let P=(γ,T)𝒫(λ){P1,P2,P3}P=(\gamma,T)\in\mathcal{P}(\lambda)\setminus\{P_{1},P_{2},P_{3}\} be such that TT3T\leq T_{3}. We claim that lk(P,P3)0{\rm lk}(P,P_{3})\neq 0. By Claim II, we know that the solution z2z_{2} does not satisfy (167), that is, H2z2C>H2(p1,0)H_{2}\circ z_{2}\equiv C>H_{2}(p_{1},0). The energy levels H2=CH_{2}=C, for C>H2(p1,0)C>H_{2}(p_{1},0) are diffeomorphic to S1S^{1} and bound a region containing (p3,0)(p_{3},0) in its interior. Then PP intersects the disk

D:={(x1,y1,x2,0)4|x2p3}S.D:=\{(x_{1},y_{1},x_{2},0)\in\mathbb{R}^{4}~{}|~{}x_{2}\geq p_{3}\}\cap S.

Note that D=P3\partial D=P_{3}. Moreover DD is transverse to the Reeb vector field RλR_{\lambda}, since XH2(x2,0)=(0,Q)X_{H_{2}}(x_{2},0)=(0,Q). We conclude that lk(P,P3)0{\rm lk}(P,P_{3})\neq 0, which proves the lemma.

Lemma 8.5.

There is no C1C^{1}-embedding ψ:S2S3\psi:S^{2}\to S^{3} such that ψ(S1×{0})=P2\psi({S^{1}\times\{0\}})=P_{2} and each hemisphere of ψ(S2)\psi(S^{2}) is a strong transverse section.

Proof.

Let W+(P2)W^{+}(P_{2}) and W(P2)W^{-}(P_{2}) be the stable and unstable manifolds of P2P_{2} respectively. First note that

W+(P2)=W(P2)=H11(12H2(p2,0))×H21(H2(p2,0)).W^{+}(P_{2})=W^{-}(P_{2})=H_{1}^{-1}\left(\frac{1}{2}-H_{2}(p_{2},0)\right)\times H_{2}^{-1}(H_{2}(p_{2},0)).

Moreover, W±(P2)P2W^{\pm}(P_{2})\setminus P_{2} consists of two connected components diffeomorphic to cylinders

(169) C1=H11(12H2(p2,0))×γ1()\displaystyle C_{1}=H_{1}^{-1}\left(\frac{1}{2}-H_{2}(p_{2},0)\right)\times\gamma^{1}(\mathbb{R}) C2=H11(12H2(p2,0))×γ2(),\displaystyle C_{2}=H_{1}^{-1}\left(\frac{1}{2}-H_{2}(p_{2},0)\right)\times\gamma^{2}(\mathbb{R}),

where γ1:2\gamma^{1}:\mathbb{R}\to\mathbb{R}^{2} and γ2:2\gamma^{2}:\mathbb{R}\to\mathbb{R}^{2} are solutions of γ˙=XH2(γ)\dot{\gamma}=X_{H_{2}}(\gamma) such that

H21(H2(p2,0))={(p2,0)}γ21()γ22(),H^{-1}_{2}(H_{2}(p_{2},0))=\{(p_{2},0)\}\cup\gamma_{2}^{1}(\mathbb{R})\cup\gamma_{2}^{2}(\mathbb{R}),

and γ1()\gamma^{1}(\mathbb{R}) is contained in the interior of the bounded region R22R_{2}\subset\mathbb{R}^{2} with boundary γ2(){(p2,0)}\gamma^{2}(\mathbb{R})\cup\{(p_{2},0)\}.

Define 𝒯i:=CiP2,i=1,2.\mathcal{T}_{i}:=C_{i}\cup P_{2},~{}i=1,2. Then 𝒯i\mathcal{T}_{i} is a 22-torus topologically embedded in SS and divides SS into two closed regions with boundary 𝒯i\mathcal{T}_{i}. One of these regions, that we call i1\mathcal{R}_{i}^{1}, has holomogy generated by P2P_{2}, and P2P_{2} is contractible in i2:=Si1¯\mathcal{R}_{i}^{2}:=\overline{S\setminus\mathcal{R}_{i}^{1}}. Since the disk

D:={(x1,y1,x2,0)4|x2p2}SD:=\{(x_{1},y_{1},x_{2},0)\in\mathbb{R}^{4}~{}|~{}x_{2}\leq p_{2}\}\cap S

has boundary P2P_{2} and projects into the complement of R2R_{2} in the plane (x2,y2)(x_{2},y_{2}), we know that i2\mathcal{R}^{2}_{i}, i=1,2i=1,2, also projects into the complement of R2R_{2} in the plane (x2,y2)(x_{2},y_{2}). This implies that i1\mathcal{R}^{1}_{i} projects into R2R_{2}, for i=1,2i=1,2. Moreover, we have 1121\mathcal{R}_{1}^{1}\subset\mathcal{R}_{2}^{1}.

Following §6.1, let {v(t),v+(t)}\{v^{-}(t),v^{+}(t)\} be the positive basis of ξx2(t)\xi_{x_{2}(t)} where v(t)v^{-}(t) is an eigenvector of dφT|ξ(x2(t))d\varphi^{T}|_{\xi(x_{2}(t))} associated to the eigenvalue β>1\beta>1 and v+(t)v^{+}(t) is an eigenvector of dφT|ξ(x2(t))d\varphi^{T}|_{\xi(x_{2}(t))} associated to the eigenvalue β1\beta^{-1}.Let (I)\rm{(I)} and (III)\rm{(III)} be the open quadrants between v(t)\mathbb{R}v^{-}(t) and v+(t)\mathbb{R}v^{+}(t) following the positive orientation and (II)\rm{(II)} and (IV)\rm{(IV)} the open quadrants between v+(t)\mathbb{R}v^{+}(t) and v(t)\mathbb{R}v^{-}(t).

Let /tη(t)ξx2(T2t)\mathbb{R}/\mathbb{Z}\ni t\mapsto\eta(t)\in\xi_{x_{2}(T_{2}t)} be a section of ξ\xi along x2(T2)x_{2}(T_{2}\cdot) such that {η,Rλ(T2)}\{\eta,R_{\lambda}(T_{2}\cdot)\} generates dψ(TS2)d\psi(TS^{2}) along x2(T2)x_{2}(T_{2}\cdot). We claim that

(170) η(t)(II)(IV),t/.\eta(t)\in(II)\cup(IV),~{}\forall t\in\mathbb{R}/\mathbb{Z}.

By Proposition 6.4 it is enough to show that

(171) wind(η,Ψ)=1\displaystyle\operatorname{wind}(\eta,\Psi)=1
(172) dλ(η(t),Rλη(t))<0,t/.\displaystyle d\lambda(\eta(t),\mathcal{L}_{R_{\lambda}}\eta(t))<0,~{}\forall t\in\mathbb{R}/\mathbb{Z}.

where Ψ\Psi is any global symplectic trivialization of ξ\xi. 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 ψ:S2S3\psi:S^{2}\to S^{3} is a C1C^{1} embedding such that ψ(S1×{0})=P2\psi({S^{1}\times\{0\}})=P_{2} and each hemisphere is a strong transverse section. Using (170), we conclude that one of the closed hemispheres of Ψ(S2)\Psi(S^{2}), that we denote by HS2H\subset S^{2}, intersects 11\mathcal{R}_{1}^{1}. Since P2P_{2} is not contractible in 11\mathcal{R}_{1}^{1}, the image of HH can not be contained in 11\mathcal{R}_{1}^{1} and we conclude that ψ(S2)W±(P2)\psi(S^{2})\cap W^{\pm}(P_{2})\neq\emptyset.

We claim that ψ(S2)W±(P2)\psi(S^{2})\cap W^{\pm}(P_{2}) consists of a disjoint union of embedded circles. To prove this claim, note that by (170), there exists an open neighborhood UU of P2P_{2} in W±(P2)W^{\pm}(P_{2}) such that Uψ(S2)=P2U\cap\psi(S^{2})=P_{2}. Thus, F:=W±(P2)UF:=W^{\pm}(P_{2})\setminus U is a closed subset of SS such that ψ(S2)(W±(P2)P2)F\psi(S^{2})\cap(W^{\pm}(P_{2})\setminus P_{2})\subset F. Moreover, ψ|S2(S1×{0})\psi|_{S^{2}\setminus(S^{1}\times\{0\})} is transverse to W±(P2)P2W^{\pm}(P_{2})\setminus P_{2}. This implies that ψ1(F)=ψ1(ψ(S2)(W±(P2)P2))\psi^{-1}(F)=\psi^{-1}(\psi(S^{2})\cap(W^{\pm}(P_{2})\setminus P_{2})) is a one dimensional submanifold of S2S^{2} that is a closed subset of S2S^{2}, proving our claim.

At least one of the connected components of ψ(S2)W±(P2)\psi(S^{2})\cap W^{\pm}(P_{2}) is homologous to P2P_{2} both in 11\mathcal{R}_{1}^{1} and in 𝒯1\mathcal{T}_{1}. Let LS2L\subset S^{2} be such that ψ(L)\psi(L) is one of these components. Then ψ(L)\psi(L) and P2P_{2} divide T1T_{1} into two closed regions. Let AA be one of these regions, with orientation induced by P2P_{2}. Then

(173) 0=A𝑑λ=P2λψ(L)λ.0=\int_{A}d\lambda=\int_{P_{2}}\lambda-\int_{\psi(L)}\lambda.

Consider HH with the orientation induced by the orientation of S1×{0}S^{1}\times\{0\}. It follows that

(174) 0<T2=/x2(T2)λ=Hψdλ0<T_{2}=\int_{\mathbb{R}/\mathbb{Z}}x_{2}(T_{2}\cdot)^{*}\lambda=\int_{H}\psi^{*}d\lambda

Since ψ|S2(S1×{0})\psi|_{S^{2}\setminus(S^{1}\times\{0\})} is transverse to the Reeb vector field, we have ψdλ>0\psi^{*}d\lambda>0 in H(S1×{0})H\setminus(S^{1}\times\{0\}). Let BHB\subset H be the region bounded by S1×{0}S^{1}\times\{0\} and LL. Then

(175) 0<Bψ𝑑λ=P2λψ(L)λ,0<\int_{B}\psi^{*}d\lambda=\int_{P_{2}}\lambda-\int_{\psi(L)}\lambda,

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 S3{S}^{3}, 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 33-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 4\mathbb{R}^{4}, 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 4\mathbb{R}^{4}, 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 33 -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).