<<

JOURNAL OF MODERN DYNAMICS WEB SITE: http://www.math.psu.edu/jmd VOLUME 1, NO. 4, 2007, 597–613

A NOTE ON REEB DYNAMICS ON THE TIGHT 3-SPHERE

FRÉDÉRIC BOURGEOIS, KAI CIELIEBAK AND TOBIAS EKHOLM (Communicated by Leonid Polterovich)

ABSTRACT. We show that a nondegenerate tight contact form on the 3-sphere has exactly two simple closed Reeb orbits if and only if the differential in lin- earized contact homology vanishes. Moreover, in this case the Floquet multi- pliers and Conley-Zehnder indices of the two Reeb orbits agree withthoseofa suitable irrational ellipsoid in 4-space.

1. INTRODUCTION

A contact form on a closed 3-manifold Y is a 1-form λ such that λ ∧ dλ is a vol- ume form on Y . The contact structure determined by a contact form λ is the tangent hyperplane field ker(λ) ⊂ T Y . The condition on λ ∧ dλ guarantees that the contact structure is a completely non-integrable plane field. The Reeb vector field determined by the contact form λ is the vector field Rλ on Y uniquely de- = = termined by the conditions λ(Rλ) 1 and ιRλ dλ 0. A periodic solution γ of the differential equation determined by Rλ is called a (closed) Reeb orbit. An overtwisted disk in a contact 3-manifold Y is an embedded 2-disk D ⊂ Y such that the foliation of D induced by the contact structure has exactly one singular point and such that the boundary ∂D of D is a closed leaf in this foliation. A contact 3-manifold which does not contain any overtwisted disk is called tight. For a star-shaped (with respect to the origin) hypersurface Y ⊂ R4 =∼ C2 the = 1 2 − = + + canonical 1-form λ 2 j=1(x j d y j y j d x j ), where (z1, z2) (x1 i y1, x2 i y2) 2 are linear coordinates onPC , restricts to a contact form on Y . This contact form is tight, and by Eliashberg’s uniqueness theorem [8] each tight contact form on the three-sphere S3 arises in this way. Particularly simple hypersurfaces of this form are the irrational ellipsoids 2 2 2 |z1| |z2| E(a1, a2) = (z1, z2) ∈ C + = 1 a a ½ ¯ 1 2 ¾ ¯ ¯ Received March 5, 2007, revised July 6, 2007. ¯ 2000 Subject Classification: Primary: 57R58; Secondary: 57R17 35B10. Key words and phrases: Reeb dynamics, contact homology, periodic orbit. FB: Partially supported by the Fonds National de la Recherche Scientifique, Belgium. KC: Partially supported by DFG grant CI 45/2-2. TE: Supported by the Royal Swedish Academy of Sciences, Research Fellow sponsored by the Knut and Alice Wallenberg foundation, by the Alfred P.Sloan Foundation, Research Fellow, and by NSF grant DMS-0505076.

597 ©2007 AIMSCIENCES 598 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM for a1, a2 > 0 and a1/a2 ∉ Q. In [9] Eliashberg, Givental, and Hofer introduced Symplectic Field Theory (SFT). It is a framework for extracting invariants of contact and symplectic manifolds via holomorphic curve counts. For the tight 3-sphere there is a particularly simple SFT type invariant called “lin- earized contact homology”. For the sake of completeness, we will recall its def- inition and a proof of its invariance in Section 2. Roughly speaking, linearized contact homology is defined as follows. Let λ be a contact form on S3 such that all closed Reeb orbits are nondegenerate, i.e. no Floquet multiplier (eigenvalue of the linearized return map on a transverse section) equals 1. Then each closed Reeb orbit has a well-defined Conley-Zehnder index CZ(γ) ∈ Z, see [13]. Follow- ing [13], we call a contact form λ dynamically convex if all closed Reeb orbits are nondegenerate and have Conley-Zehnder index at least 3. E.g. this is the case for the induced contact forms on irrational ellipsoids, and more generally for the contact forms on hypersurfaces in R4 bounding strictly convex domains [13]. For 3 a dynamically convex contact form λ let CC∗(S ,λ) be the Q- gen- erated by the “good” (see Section 2) closed Reeb orbits, graded by their degree |γ| = CZ(γ) − 1. Fixing an R-invariant almost complex structure J on R × S3 com- 3 3 patible with λ, we define a differential ∂ : CC∗(S ,λ) → CC∗−1(S ,λ) by counting rigid J-holomorphic cylinders connecting Reeb orbits. We show in Section 2 that ∂2 = 0, that the linearized contact homology HC lin(S3) = ker∂/im∂ is indepen- dent of the choice of J and of λ, and furthermore that Q for k ≥ 2 even, lin 3 = HCk (S ) (0 otherwise. In this note we prove the following result.

THEOREM 1. Let λ be a dynamically convex tight contact form on the 3-sphere. Then the following are equivalent: (i) The differential in linearized contact homology vanishes. (ii) There are precisely two simple Reeb orbits γ1,γ2.

Moreover, in this case γ1 and γ2 are unknotted, elliptic, have linking number 1, and their Conley-Zehnder indices and Floquet multipliers agree with those of a suitable irrational ellipsoid. Remark 2. The abstract perturbation theory developed in [14] will eventually lead to a definition and an invariance proof of linearized contact homology for arbitrary (not necessarily dynamically convex) tight contact forms on S3 with nondegenerate closed Reeb orbits, cf. Remark 7 below. Using this in combina- tion with formulas for Conley-Zehnder indices, see Section 3.1, it is immediate that condition (i) implies dynamical convexity and an additional argument, see Remark 14, shows that (ii) does as well. Thus, the restriction in Theorem 1 to the technically simpler dynamically convex case turns out not to be any restriction at all.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 599

Remark 3. The implication that if there are precisely two simple closed Reeb orbits γ1,γ2 then their Floquet multipliers lie on the unit circle with irrational angles has been proved independently in [11]. Remark 4. It follows from [13] that in the situation of Theorem 1 the closed Reeb orbit of degree 2 is the binding of an open book decomposition with pages whose interiors are transverse to the Reeb vector field. The return map of a page is an area preserving diffeomorphism of the open disk with one fixed point and no other periodic points. However, the return map need not be conjugate to an irrational rotation (see [10]). Remark 5. In view of the result in [13], Theorem 1 implies the following di- chotomy for the Reeb dynamics on a strictly convex hypersurface in R4. Either the differential in linearized contact homology vanishes and there are precisely two simple closed orbits; or the differential does not vanish and there are infin- itely many simple closed orbits. Moreover, generically the second case occurs (even in the class of star-shaped hypersurfaces). This discussion motivates the following conjecture. If, for a star-shaped hyper- surface in R4, the differential in linearized contact homology does not vanish,then it carries infinitely many closed characteristics. More optimistically, one could even conjecture that a star-shaped hypersurface in R2n carries either precisely n (if the differential in linearized contact homology vanishes) or infinitely many (if it does not vanish) simple closed Reeb orbits. See [12] for an exposition of known multiplicity results.

2. LINEARIZEDCONTACTHOMOLOGY

In this section we define the linearized contact homology for a dynamically con- vex contact form on S3 and prove its invariance. See [9] for details of the setup. Fix a dynamically convex contact form λ on S3. If γ is a simple closed orbit of k k the Reeb field Rλ, we denote by γ its k-th iterate. We call γ good if the Conley- Zehnder indices (see Section 3) of γ and γk have the same parity. Otherwise, we k 3 call γ bad. Let CC∗(S ,λ) be the Q-vector space generated by the good closed Reeb orbits, graded by their degree |γ| = CZ(γ) − 1. An R-invariant almost complex structure J on R × S3 is called compatible with λ = · · ∂ = if it preserves ξ kerλ, if dλ( , J ) defines a metric on ξ, and if J ∂t Rλ, where t denotes the coordinate on R. Fix such a J. For good closed orbits γ,γ1,...,γr of Rλ of periods T,T1,...,Tr , let M (γ;γ1,...,γr ) be the moduli space consisting of equivalence classes of tuples (x, p; y1, p1 ..., yr , pr ; f ), where x, y1,..., yr are dis- 2 2 tinct points on S with directions p, p1,..., pr and f = (a,u) : S à{x, y1,..., yr } → R × S3 is a map with the following properties: • d f + J ◦ d f ◦ j = 0, • = +∞ = −∞ limz→x a(z) and limz→yi a(z) , • in polar coordinates (ρ,θ) ∈]0,1]×R/Z around x in which p corresponds to θ = 0 we have limρ→0 u(ρ,θ) = γ(−T θ), and

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 600 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

• in polar coordinates (ρi ,θi ) ∈]0,1]×R/Z around yi in which pi corresponds = = to θi 0 we have limρi →0 u(ρi ,θi ) γi (Ti θi ). ′ ′ ′ ′ ′ ′ ′ Two such tuples (x, p; y1, p1 ..., yr , pr ; f )and(x , p ; y1, p1 ..., yr , pr ; f ) are equiv- 2 = ′ = ′ alent if there exists a biholomorphism h of S such that h(x) x , h(yi ) yi , · = ′ · = ′ = ′ ◦ R R dxh p p , dyi h pi pi and f f h. Since J is -invariant, acts on these moduli spaces by translation and we denote the quotient by M (γ;γ1,...,γr )/R. Using an appropriate functional analytic setup, the moduli spaces can be de- scribed as the zero locus of a Fredholm section of a certain and have expected dimension (determined by the Fredholm index) given by r dim M (γ;γ1,...,γr )/R = |γ|− |γj | − 1. j=1 ¡ ¢ X If this dimension is zero and if the moduli space is compact and cut out transver- sally, then it consists of finitely many points. One can associate a sign to each of these points via coherent orientations on the moduli spaces [3] and we denote by n(γ;γ1,...,γr ) the algebraic count of the elements in M (γ;γ1,...,γr )/R. Fur- thermore, we denote by κγ the multiplicity of the Reeb orbit γ. Define the linear map ′ 3 3 n(γ;γ ) ′ ∂ : CC∗(S ,λ) → CC∗−1(S ,λ), γ 7→ γ . ′ |γ′|=|γ|−1 κγ X Thus ∂ counts rigid J-holomorphic cylinders interpolating between closed Reeb orbits γ and γ′.

THEOREM 6. Let λ be a dynamicallyconvex contact form on S3. Thenfor ageneric S1-dependent compatible almost complex structure J the map ∂ is well-defined and satisfies ∂2 = 0. Moreover, the linearized contact homology HC lin(S3) = ker∂/im∂ is independent of the choice of J and λ, and is given by

Q for k ≥ 2 even, lin 3 = HCk (S ) (0 otherwise.

Proof. The proof follows the familiar scheme from Floer homology (see e.g. [15, 9]), provided we can prove that 1-dimensional moduli spaces of holomorphic cylinders are regular and compact up to breaking into pairs of cylinders. As usual, one needs to prove this in three cases:

• in a symplectization (to establish [9, Proposition 1.9.1]), • in a cobordism (to establish [9, Proposition 1.9.3]), and • for a homotopy of almost complex structures on a cobordism (to establish [9, Proposition 1.9.4]).

We will explain the argument in the case of a homotopy, the other two cases lin 3 being analogous but easier. Once invariance of HCk (S ) is established it can be easily computed for an irrational ellipsoid using the formulae for the Conley- Zehnder indices in Section 3 (∂ = 0 in this case).

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 601

Step 1: Let Y = S3. Consider X = R × Y with an exact symplectic form dλ which t coincides near {±∞} × Y with d(e λ±) for dynamically convex contact forms λ± on Y . Let(Jτ)τ∈[0,1] be a homotopy of almost complex structures on R × Y such that for every τ, Jτ is compatible with dλ and coincides near ±∞ with fixed R- invariant almost complex structures J± compatible with λ±. Fix closed Reeb or- bits γ± for λ± of equal degrees |γ+| = |γ−| and consider the 1-dimensional mod- uli space

M[0,1] = ∪τ∈[0,1]M (γ+;γ−; Jτ).

By the SFT compactness theorem [2, 6], every sequence fk , k = 1,2,... in M[0,1] has a subsequence converging as k → ∞ to a broken holomorphic sphere F = {Fα}α∈T . Here T is a directed tree (i.e. each edge is directed) with the following properties. For each vertex α ∈ T , Fα is a punctured holomorphic sphere with ex- actly one positive puncture and any number of negative punctures in (R×Y , J+), in (R × Y , J−), or in (X , Jτ) for some τ ∈ [0,1]. Each edge e of T is labeled by a closed Reeb orbit γe of λ±. If e is directed from a vertex α to a vertex β, then γe is the asymptotic Reeb orbit at the unique positive puncture of Fα and at one negative puncture of Fβ. Conversely, each puncture on any Fα corresponds to a unique edge in this way except for two free punctures, a positive one asymptotic to γ+ and a negative one asymptotic to γ−. Since the Fredholm index is additive under joining spheres at Reeb orbits and since the only two free (not paired across edges) punctures of F are asymptotic to γ±, the expected dimensions (i.e. the Fredholm indices) ind(Fα) of the moduli spaces of J±-, or Jτ-holomorphic spheres (τ fixed) which contain the Fα satisfy

ind(Fα) = |γ+|−|γ−| = 0. α∈T X The structure of the tree T can be described as follows. Let α± ∈ T be the ver- tices such that Fα± contains the free punctures asymptotic to γ±. Define the stem S of the tree T to be the unique linear (i.e. at most two edges meet at each vertex) subtree S ⊂ T connecting α+ and α−. Define the branches B1,...,Bk to be the connected components of T − S. Since each branch Bi has precisely one free puncture, which is positive and asymptotic to a Reeb orbit γi , its total index = ind(Bi ) α∈Bi ind(Fα) satisfies

P ind(Bi ) = |γi | ≥ 2 by dynamical convexity.

The orbit cylinder over a closed Reeb orbit γ for λ± is the J±-holomorphic cylin- der R × γ ⊂ R × Y . We call a punctured holomorphic sphere good if it is not a branched cover of an orbit cylinder. Now let us assume that the following regu- larity condition holds.

(R) All good components Fα with α ∈ S in the stem are regular, i.e. transversely cut out by the 1-parameter family of Cauchy-Riemann operators corre- sponding to the homotopy (Jτ)τ∈[0,1].

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 602 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

We conclude that each good Fα in (R×Y , J±) is transversely cut out and thus sat- isfies ind(Fα) ≥ 1 because of translation invariance (see Step 2 below for a full explanation of how to reach this conclusion). Moreover, there is a finite collec- tion of exceptional τ-values 0 < τ1 <···< τm < 1 such that the following hold. For every τ 6= τj , j = 1,...,m, each good Fα in (X , Jτ) and in the stem belongs ¯ to a moduli space which is transversely cut out by the ∂Jτ -equation and hence ind(Fα) ≥ 0. For τ = τj some j ∈ {1,...,m}, there exists a unique exceptional good punctured holomorphic sphere Cτ ∈ M (β0;β1,...,βr ; Jτ), which may be- ¯ long to the stem, with the property that the linearized ∂Jτ -operator at Cτ has 1-dimensional cokernel; at all other good spheres Fα 6= Cτ in the stem, the lin- ¯ earized ∂Jτ -operator is surjective. Since the cokernel has dimension 1, it follows that ind(Cτ) ≥ −1.

By Lemma 8 below, a branched cover Fα over an orbit cylinder has index ind(Fα) ≥ 0. So we see that ind(Fα) ≥ 0 for each component Fα in the stem with Fα 6= Cτ.

Recall that each edge e ∈ T was labeled with a Reeb orbit γe. We order the edges by the actions A (γe). For area reasons, this order strictly increases in the direc- tion of the tree and therefore the special component Cτ can occur at most once among the Fα with α ∈ S. In particular, in view of the preceding discussion the total index of the stem S satisfies

ind(S) ≥ −1.

Thus, since ind(T ) = 0 and since ind(Bi ) ≥ 2 for any branch Bi , we conclude that there are no branches Bi , and hence T = S is a linear tree. In particular, this excludes branched covers of orbit cylinders and hence all components Fα are good. Moreover, there are only the following two possibilities for the dimensions of the Fα, α ∈ T .

Case 1: ind(Fα) ≥ 0 for all α ∈ T , in which case T has only one vertex α, no break- ing occurs, and the sequence (fk) converges in M[0,1] to Fα ∈ M[0,1]. ± Case 2: ind(Fβ± ) = ±1 for unique vertices β ∈ T and ind(Fα) = 0 for all vertices + − α ∈ T − {β ,β }. Since the linear tree T contains only one component in (X , Jτ) ± and components in (R × Y , J±) have index at least 1, we conclude that β are the only vertices in T . Hence the limit curve F is a pair of holomorphic cylin- ders, which is precisely what is needed for the chain homotopy property. This concludes the proof modulo the assumption that regularity condition (R) above holds.

Step 2: We adapt the technique used in [7] to achieve the regularity condition (R). 2 To three distinct points z0, z1, z2 on the Riemann sphere S = C ∪ {∞} and a tan- ∈ 2 gent direction p0 Tz0 S /R+ at z0, where R+ acts on tangent vectors by scalar 1 multiplication, we associate an angle w(z0, p0, z1, z2) ∈ S as follows. Let φ ∈ 2 Aut(S ) be the unique Möbius transformation with φ(z0) = 0, φ(z1) = 1, and

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 603

φ(z2) = ∞, and define = · ∈ ∼ 1 w(z0, p0, z1, z2) dz0 φ p0 T0C/R+ = S . The map w is clearly invariant under simultaneous action of Möbius transfor- mations on (z0, p0, z1, z2) and thus it induces a diffeomorphism $ → 1 w : M3 S , $ where M3 is the decorated Deligne-Mumford space of 3 ordered distinct points (z0, z1, z2) on the Riemann sphere with a specified direction p0 at z0 (see [2]). One easily sees that this map extends to arbitrary trees of spheres for which z2 lies between z0 and z1 (i.e. z2 lies on the unique embedded path connecting z0 and z1). 1 Let JS1 be the space of S -dependent almost complex structures on R×Y of the type considered above, i.e., which are compatible with dλ and which coincides near ±∞ with fixed R-invariant almost complex structures J±. Each J ∈ JS1 in- duces a Cauchy-Riemann operator acting on tuples (z0, p0, z1, f ) consisting of 2 2 distinct points z0, z1 ∈ S , a direction p0 at z0, andamap f : S − {z0, z1} → R × Y by ¯ 1 ∂J (z0, p0, z1, f )(z) = d f (z) + J w(z0, p0, z1, z) ◦ d f (z) ◦ i , 2 2 ³ ¡ ¢ ´ where z ∈ S − {z0, z1}. This operator is clearly invariant under the simultaneous 2 action of Aut(S ) on (z0, p0, z1, f ). For fixed (z0, p0, z1), after applying a Möbius transformation that sends z0 to 0, z1 to ∞ and p0 to R+, we obtain a Cauchy- Riemann operator on maps f : R × (R/2π) → R × Y where we think of the source as an infinite cylinder (using C − {0} =∼ R × (R/2π)). In polar coordinates (s, t) ∈ R × (R/2πZ) this operator is given by

1 −it ∂¯ J (f )(s, t) = d f (s, t) + J(e ) ◦ d f (s, t) ◦ i , 2 ³ ´ s+it −it since w(0,R+,∞,e ) = e . A standard argument (see e.g. [15]) shows that regularity for 1-parameter families of such Cauchy-Riemann operators can be 1 achieved by choosing a generic path of S -dependent J in JS1 . More precisely, for a Baire set of paths (Jτ)τ∈[0,1] ∈ JS1 the following holds: at any pair (f ,τ) such ¯ = that ∂Jτ f 0 and such that f is not a branched cover of an orbit cylinder, the ¯ linearization of ∂Jτ at (f ,τ) is surjective. We use the regularity result for S1-dependent almost complex structures to es- tablish condition (R) as follows. Fix a point pγ¯ on each simple closed Reeb or- bit γ¯. For closed Reeb orbits γ0,γ1,...,γk with underlying simple orbits γ¯i and J ∈ JS1 denote by M (γ0;γ1,...,γk; J) the moduli space of equivalence classes of tuples (z0, p0, z1,..., zk, f ) of the following form: 2 • z0,..., zk are distinct points in S ; • p0 is a tangent direction at z0; 2 ¯ • f : S − {z0,..., zk} → R × Y satisfies ∂J (z0, p0, z1, f ) = 0;

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 604 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

• at the positive puncture at z0, f is asymptotic to γ0 and takes the tangent

direction p0 to the point pγ¯0 ; • at the negative punctures at z1,..., zk, f is asymptotic to γ1,...,γk, respec- tively; • f is not a branched cover of an orbit cylinder.

′ ′ ′ ′ ′ Two tuples (z0, p0, z1,..., zk, f )and(z0, p0, z1,..., zk, f ) are equivalent if they are related under the natural action of Aut(S2). The regularity result for the Cauchy-Riemann operator above implies that, for generic paths (Jτ)τ∈[0,1] in JS1 , any moduli space

M[0,1] = ∪τ∈[0,1]M (γ0;γ1,...,γk; Jτ) is cut out transversally and is consequently a manifold of dimension

k dim(M[0,1]) = |γ0|− |γj |+ 1. j=1 X Here a manifold of negative dimension is understood to be empty. In order to establish (R) it remains to study the boundary of a 1-dimensional moduli space M[0,1] as above. Let fk ∈ M[0,1] be a sequence as in Step 1 which converges to a broken holomorphic sphere F modeled on a tree T with stem S. Then each α ∈ S lies between the special vertices α± corresponding to the free asymptotic orbits γ± (the asymptotic orbits of fk), so any good holomorphic sphere Fα, α ∈ S, in (X , Jτ) belong to some moduli space M (γ0;γ1,...,γk; Jτ) of the type above. Thus the preceding discussion implies the following regu- larity properties for Fα. If Fα lies in (R × Y , J±) it belongs to a moduli space M (γ0;γ1,...,γk ; J±) which is transversely cut out, so ind(Fα) ≥ 1 by R-invariance. If Fα lies in (X , Jτ) for some τ it belongs to a moduli space M[0,1] which is trans- versely cut out, so ind(Fα) ≥ −1. Moreover, ind(Fα) = −1 occurs only for finitely many components Cτj at parameter values τ1,...,τm and all other components have ind(Fα) ≥ 0. This proves that the regularity condition (R) holds and hence demonstrates Theorem 6.

Remark 7. Theorem 6 defines linearized contact homology for dynamically con- vex contact forms on S3, which turn out to be sufficient for the purposes of this paper, see Remark 2. Linearized contact homology is expected to exist more gen- erally for any contact manifold (Y ,λ) with an exact symplectic filling (X ,dλ). Here the boundary map between closed Reeb orbits γ± will count punctured holomorphic spheres in R × Y with one positive puncture asymptotic to γ+ and any number of negative punctures asymptotic to Reeb orbits γ−,γ1,...,γk, to- gether with rigid holomorphic planes in X asymptotic to the γi , i = 1,...,k. See [5, 1] for details of this construction. However, it seems that in this more general situation transversality cannot be achieved by the method in the proof of Theorem 6, but requires the use of abstract perturbations. Hofer, Wysocki and

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 605

Zehnder are developing a theory of such perturbations called “polyfold Fred- holm theory” [14]. However, at the time of this writing their theory is not yet completed.

3. PROOF OF THEOREM 1

In this section we prove Theorem 1. To that end we study properties of Reeb orbits in dimension 3 and dynamical properties of translations on a flat torus.

3.1. Conley-Zehnder indices. The grading in the contact homology algebra is induced by Conley-Zehnder indices, see [13],[12]. Here we recall some proper- ties of the Conley-Zehnder index in dimension 3 in the case of non-degenerate Reeb orbits. Let γ be a simple closed Reeb orbit and denote by γk its k-th it- erate. Then the Floquet multipliers (eigenvalues of the linearized return map on a transverse section) occur in a pair µ,1/µ ∈ R à {0,±1} if they are real, or µ,µ¯ ∈ S1 à {±1} if not. If γ is a Reeb orbit we write CZ(γ) for its Conley-Zehnder index. According to Section 8.1 in [12], we need to distinguish three cases (the first case is covered by Theorem 7 and the other two by Theorem 6). Elliptic case: γ has a non-real Floquet multiplier µ = e2πiα with α ∈ (0,1) irra- tional. Then CZ(γk) = 2kr + 2[kα] + 1 = 2[k(r + α)] + 1, for some integer r ∈ Z,where[x] denotes the largest integer smaller than or equal to x. Even hyperbolic case: γ has a positive real Floquet multiplier µ ∈ (0,1). Then

CZ(γk ) = 2kr, for some integer r ∈ Z. Odd hyperbolic case: γ has a negative real Floquet multiplier µ ∈ (−1,0). Then

CZ(γk) = (2r + 1)k, for some integer r ∈ Z. Note that the even multiplies of γ are “bad” in the sense of [9] and do not contribute to contact homology. As mentioned in Section 1, the degree |γ| of a closed Reeb orbit γ as a generator of the contact homology algebra as well as a generator of the linearized contact homology chain complex is given by |γ| = CZ(γ) − 1. The following lemma was used in the proof of Theorem 6 above.

LEMMA 8. Let γ be a simple closed Reeb orbit all of whose iterates are nondegen- erate. Then for any positive integers k1,...,ks |γk1+···+ks | ≥ |γk1 |+···+|γks |.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 606 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

Proof. Set k = k1 +···+ ks. We treat each type of γ separately. If γ is elliptic, then s s s k ki |γ |− |γ | = 2r (k − ki ) + 2([kα] − [ki α]) ≥ 0 i=1 i=1 i=1 X X X because [a + b] ≥ [a] + [b] for any real numbers a,b. If γ is even hyperbolic, then s s k ki |γ |− |γ | = (2kr − 1) − (2ki r − 1) = s − 1 ≥ 0. i=1 i=1 X X Finally, if γ is odd hyperbolic, then s s k ki |γ |− |γ | = (2r + 1)k − 1 − (2r + 1)ki − 1 = s − 1 ≥ 0. i=1 i=1 X ¡ ¢ X¡ ¢

3.2. Torus dynamics. Consider the n-dimensional torus Tn = Rn/Zn. Let n n n (ξ1,...,ξn) ∈ [0,1] . Let τ: T → T denote translation by (ξ1,...,ξn). m n LEMMA 9. The orbit {τ (0)}m∈Z of 0 ∈ T under τ is dense in a finite collection of translates of an l-torus Tl which is a subgroup of Tn, where

l + 1 = dimQ SpanQ(ξ1,...,ξn,1) . ³ ´ Here SpanQ(ξ1,...,ξn,1) denotes the vector subspace of R spanned by ξ1,...,ξn,1, where R is viewed as a vector space over Q. Remark 10. This lemma is a discrete version of the well-known fact that the ge- m k m odesic through 0 ∈ T in direction (ξ1,...,ξm) is dense in a subtorus T ⊂ T , where k = dimQ SpanQ(ξ1,...,ξm) . ³ ´ Proof. Assume first that l = n. Then by the preceding remark the geodesic γ n+1 n+1 through 0 ∈ T in direction (ξ1,...,ξn,1) is dense in the torus T . Since γ is transverse to the subtorus Tn where the last coordinate equals zero, the inter- n n m section γ ∩ T is dense in T . But this intersection is just the orbit {τ (0)}m∈Z of 0 in Tn. Consider next the case when l < n. In this case there is an equation of the form

m1ξ1 + m2ξ2 +···+ mnξn + mn+1 = 0, where all m j are integers. If d is the greatest common divisor of m1,...,mn we may rewrite this as ′ ′ m1(dξ1) +···+ mn(dξn) + mn+1 = 0. ′ ′ Since m1,...,mn do not have any common divisor, it follows that the hyperplane H in Rn given by the equation ′ ′ m1(d x1) +···+ mn(d xn) + mn+1 = 0 contains a point with integer coordinates. It follows that any iterate τkd (0) which is a multiple of d lies in the torus Tn−1 ⊂ Tn which is the subgroup with preimage

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 607 in Rn given by the integer translates of the hyperplane H. Any other iterate lies in a translate of this subgroup by τj (0), j = 1,...,d − 1. To finish the proof we use induction. The intersection of the lattice Zn and the hyperplane H is again a lattice generated by vectors with integer coordinates. Let v1,..., vn−1 be a basis. Writing

(dξ1,...,dξn) = η1v1 +···+ ηn−1vn−1, we find that the vector spaces over Q spanned by η1,...,ηn−1 and the vector space spanned by ξ1,...,ξn are equal. Hence

dimQ SpanQ(η1,...,ηn−1,1) = l. Using the argument above we³ can either confine the´ orbits to translates of lower dimensional tori (if l < n−1) or the orbit is dense (if l = n−1). The lemma follows after n − l steps.

3.3. Jump sequences. To each real number ξ ∈ (0,1) we associate its jump se- quence j (ξ) = {jn(ξ)}n∈N via

jn(ξ) = [n/ξ], n ∈ N.

Thus jn = jn(ξ) is the unique integer satisfying

jnξ ≤ n < (jn + 1)ξ, so the n-th jump in the sequence {[kξ]}k∈N occurs at k = jn(ξ). The jump se- quence j (ξ) determines ξ via n lim = ξ. →∞ n jn(ξ)

LEMMA 11. For i = 1,2,3 let ξi ∈ (0,1) be irrational with jump sequences j (ξi ).

(a) If j (ξ2) is a subsequenceof j (ξ1), then there exist a linear relation

ξ2 = pξ1 + q, p, q ∈ Q, p > 0.

(b) If j (ξ2) and j (ξ3) are both subsequences of j (ξ1), then ξ2 and ξ3 have a com- ∈ = mon jump, i.e. there exist m2,m3 N such that jm2 (ξ2) jm3 (ξ3).

Example 12. The situation in (a) occurs e.g. for ξ2 = ξ1/k, k ∈ N, in which case = 6= = − 1 = jn(ξ2) jnk (ξ1). An example with q 0 is given by ξ2 ξ1 2 : If k jn(ξ2) is in the jump sequence of ξ2 then k 1 kξ1 < n + < (k + 1)ξ1 − . 2 2 It follows that for k even, k kξ1 < n + < (k + 1)ξ1, 2 and for k odd, k 1 kξ1 < n + + < (k + 1)ξ1. 2 2 Hence in either case k is in the jump sequence of ξ1.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 608 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

Remark 13. D. Kotschick has pointed out that if j (ξ2) is a subsequence of j (ξ1) and ξ1 ≤ 1/2, then ξ1 = kξ2 for some k ∈ N. This can be seen as follows. Suppose jn(ξ2) = jm(n)(ξ1) for a sequence m : N → N. It easily follows that m is a quasi- morphism with defect (deviation from being a semi-group homomorphism) at most [3ξ1] ≤ 1. A more careful estimate shows that for ξ1 ≤ 1/2 the defect equals zero, so m is a group homomorphism, i.e. m(n) = kn for some k ∈ N.

As the preceding example shows, this result fails as soon as ξ1 > 1/2. It might be interesting to characterize all pairs (ξ1,ξ2) for which j (ξ2) is a subsequence of j (ξ1).

Proof of Lemma 11. Consider (a). Let T2 = R2/Z2 and let τ: T2 → T2 be trans- lation by the vector (ξ1,ξ2). Lemma 9 implies that if ξ1, ξ2 and 1 are linearly independent over Q, then the orbit of 0 under τ is dense in T2. Thus there exists k 2 an iterate τ (0) = (x1, x2) ∈ (0,1) such that x1 + ξ1 < 1 and x2 + ξ2 > 1. So k is a jump of ξ2 but not of ξ1, contradicting the hypothesis. We conclude that ξ1, ξ2 and 1 satisfy a linear relation

ξ2 = pξ1 + q, p, q ∈ Q. The proof of Lemma 9 shows that the orbit of 0 is dense in the straight line through 0 with slope p. If p < 0, this implies that there is an iterate τk (0) with ∈ − 1 1 2 > < representative (x1, x2) ( 2 , 2 ) arbitrarily close to 0 and such that x1 0, x2 0. So k is a jump of ξ2 but not of ξ1, contradicting the hypothesis. Hence the slope p is positive. Consider (b). Part (a) implies that there are linear relations

ξj = p j ξ1 + q j , p j , q j ∈ Q, p j > 0, j = 2,3. It follows that there is a linear relation

ξ3 = pξ2 + q, p, q ∈ Q, p > 0. Again by the proof of Lemma 9, the orbit of 0 under the translation τ : T2 → T2 by (ξ2,ξ3) is dense in the straight line through 0 with slope p. This implies that k ∈ − 1 1 2 there is an iterate τ (0) with representative (x1, x2) ( 2 , 2 ) arbitrarily close to 0 and such that x1 < 0, x2 < 0. So k is a jump of both ξ2 and ξ3.

3.4. Proof of Theorem 1. Assume that the differential in linearized contact ho- mology vanishes. Then in view of the discussion preceding Theorem 1, thereis a unique (not necessarily simple) closed Reeb orbit of every positive even degree and no other Reeb orbits. In particular, there are no even hyperbolic orbits. We distinguish three cases. Case 1: There are at least two simple Reeb orbits which are odd hyperbolic. In this case, the formula for the Conley-Zehnder index shows that suitable odd multiplies of the simple orbits have the same degree. This contradicts vanishing differential. Hence, Case 1 is ruled out. Case 2: There is exactly one simple Reeb orbit which is odd hyperbolic.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 609

Denote this simple orbit by γ1. Write CZ(γ1) = 2r1 + 1. Dynamical convexity ≥ 2ℓ−1 | 2ℓ−1| = + − − implies r1 1. Odd multiples γ1 of γ1 have degrees γ1 (2r1 1)(2ℓ 1) + − − ∞ 1. Since {(2r1 1)(2ℓ 1) 1}ℓ=1 does not contain all positive even integers, there exists some other orbit of even degree. Since γ1 is assumed to be the only odd hyperbolic orbit we conclude that there exists an elliptic orbit. Let γ2 be a simple elliptic orbit. Write CZ(γ2) = 2[r2 +α2]+1 = 2r2 +1, where r2 ∈ Z and α2 ∈ (0,1) is ≥ k irrational. Dynamical convexity implies r2 1. Multiples γ2 of γ2 have degrees 2[k(r2 + α2)]. We claim that there exists multiples k and ℓ such that

(1) (2r1 + 1)(2ℓ − 1) − 1 = 2[k(r2 + α2)]. | 2ℓ−1| = | k| This claim implies that γ1 γ2 which contradicts vanishing differential and allows us to rule out Case 2.

To verify the claim we argue as follows. Let v = 2r1 + 1 and note that the num- bers in the left hand side of (1) can be written as (ℓ − 1)(2v) + (v − 1), ℓ = 1,2,.... Consider the circle R/(v · Z) and the irrational rotation τ: R/(v · Z) → R/(v · Z), τ(x) = x + (r2 + α2). Lemma 9 implies that there exists k such that v − 1 1 dist τk(0), + < 10−80. 2 4 µ µ ¶¶ This means that there exists ℓ ≥ 1 such that

v − 1 1 −80 (ℓ − 1)v + + − k(r2 + α2) < 10 . ¯ 2 4 ¯ ¯ ¯ It follows that ((2r ¯ + 1)(2ℓ − 1) − 1 + 1 ) − 2k(r + α ¯) < 10−79. This implies the 1¯ 2 2 2¯ claim. ¯ ¯ ¯ ¯ Case 3: All simple Reeb orbits are elliptic.

Let γj , j = 1,2,... denote the simple orbits. Write CZ(γj ) = 2[r j + αj ] + 1 = 2r j + 1 where r j ∈ Z and αj ∈ (0,1) is irrational and note that dynamical convexity ≥ k + implies r j 1. Multiples γj of γj have degrees 2[k(r j αj )]. For fixed j the set + ∞ 2[k(r j αj )] k=1 does not contain all even multiples of r j : For every multiplicity k such that there exists an integer m with © ª kαj < m < (k + 1)αj , the equation k+1 k |γj |−|γj | = 2[(k + 1)(r j + αj )] − 2[k(r j + αj )] = 2(r j + 1) ≥ 4 holds. We conclude that there are at least two simple orbits. Now we invoke a result proved in [4]: For all k ∈ N there exist isomorphisms lin 3 =∼ lin 3 HC2k+2(S ) HC2k (S ) between linearized contact homology induced by chain maps counting holomorphic curves in the symplectization. Since the dλ-area of a holomorphic curve is positive, this implies the following, in the situation under consideration: k = (O) The two orderings of the set {γj }, j,k 1,2,... of all closed Reeb orbits by increasing degree and by increasing action, respectively, coincide.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 610 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

k If Aj denotes the action of the simple Reeb orbit γj , then the action of γj equals Ai kAj . It follows from (O) that each action ratio , i 6= j , must be irrational: if it Aj were rational some multiples of γi and γj would have the same action and hence the same degree, which is impossible since the differential vanishes.

Let γ1 denote the simple orbit of smallest action. It follows from (O) and the fact that the minimal degree is 2 that |γ1| = 2[r1 + α1] = 2r1 = 2. Hence r1 = 1 and the degrees of adjacent multiples of γ1 differ either by 2 or by 4. Consider a simple orbit γ , j 6= 1. Since A1 ∈ (0,1) is irrational, for any integer m there is a unique j Aj integer k such that kA1 < mAj < (k + 1)A1, or equivalently A A (2) k 1 < m < (k + 1) 1 . Aj Aj | k+1|−| k | = In particular, (O) implies that γ1 γ1 4, or equivalently ′ (3) kα1 < m < (k + 1)α1, for some integer m′. In the language of Section 3.3, this means:

A1 (J) For each j > 1 the jump sequence of ξj = is a subsequence of the jump Aj sequence of ξ1 = α1. Now we prove that there are at most two simple orbits. To see this, assume that there are three distinct orbits γj , j = 1,2,3. Condition (J) and Lemma 11 (b) A1 imply that ξj = , j = 2,3, have a common jump, i.e. that there exist integers k, Aj m2, and m3 such that

A1 A1 A1 A1 k < m2 < (k + 1) , k < m3 < (k + 1) . A2 A2 A3 A3

This implies that both m2A2 and m3A3 lie between kA1 and (k + 1)A1. Since | k+1| − | k| ≤ | m2 | = | m3 | γ1 γ1 4, it follows from (O) that γ2 γ3 , which contradicts vanish- ing differential. Hence there are at most two simple Reeb orbits. Since we already know that there are at least two orbits, this proves that (i) implies (ii). For the converse implication suppose that the differential in linearized contact homology does not vanish. This can only happen if there is at least one even hyperbolic orbit γ1 (whose iterates have odd degree) and one elliptic or odd hy- perbolic orbit γ2 (whose good iterates have even degree). Dynamical convexity implies r2 > 0 and the iteration formulae for Conley-Zehnder indices show that multiples of γ2 cannot attain all even degrees, so there must be a third orbit. Thus (ii) implies (i). Next suppose that (i) and (ii) hold, so there are precisely two simple closed orbits γ1,γ2 with actions Ai , Floquet multipliers αi and rotation numbers ri . We first show that these data are realized by a suitable irrational ellipsoid. Since there are only two simple closed orbits, condition (O) implies that the jump sequences of ξ = α and ξ = A1 agree. Since the ξ are determined by their 1 1 2 A2 i

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 611 jump sequences, it follows that

A1 (4) α1 = . A2 Next note that (O) implies that for each m, if k is such that

kA1 < mA2 < (k + 1)A1, then, since r1 = 1,

[k(1 + α1)] < [m(r2 + α2)] < [(k + 1)(1 + α1)]. These inequalities combine to

[k(1 + α1)] [m(r2 + α2)]A1 [(k + 1)(1 + α1)] < < , k + 1 mA2 k which in the limit m → ∞ (hence k → ∞) yields A r + α (5) 2 = 2 2 . A1 1 + α1

Next note that the first gap in the degrees of multiples of γ1 occurs at the degree 2([1/α1] + 1). So this must be the degree of γ2 and we conclude

(6) r2 = [1/α1] + 1.

Equations (4), (5)and(6) together with r1 = 1 uniquely determine α1, α2, r1 and r2 in terms of the action ratio A2/A1:

A1 A2 A2 A2 (7) α1 = , α2 = − , r1 = 1, r2 = + 1. A2 A1 · A1 ¸ · A1 ¸

It remains to show that any combination of actions Ai , Floquet multipliers αi and rotation numbers ri satisfying equations (7) is realized by an ellipsoid 2 2 2 |z1| |z1| E = (z1, z2) ∈ C + = 1 a a ½ ¯ 1 1 ¾ ¯ with suitable 0 < a < a such that a /a¯ is irrational. 1 2 1 ¯2 2it/ai The Reeb flow on E is given by zi (t) = e , which is periodic of period Ti = πai in the i-th component. Thus for a1/a2 irrational there are precisely two simple closed orbits γ1 = {z2 = 0} and γ2 = {z1 = 0} of period (= action)

Ai = Ti = πai .

Define the ai by this equation. From

2πia1/a2 2πia2/a1 z2(T1) = e , z1(T2) = e we read off the Floquet multipliers

α1 = a1/a2, α2 = a2/a1 − [a2/a1] and the rotation numbers

r1 = 1, r2 = [a2/a1] + 1

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 612 FRÉDÉRIC BOURGEOIS,KAI CIELIEBAK AND TOBIAS EKHOLM

(the additional +1 results from the choice of trivializations that extend over disks). For ai = Ai /π these equations agree with equations (7). It is proved in [13] that for dynamically convex contact forms there exists a sim- ple closed orbit γ1 with CZ(γ1) = 3 (unique in our case) which is the binding of an open book decomposition. The pages are discs whose interiors are transverse to the Reeb vector field. The return map of a page possesses at least one fixed point (see [13]), which in our case must correspond to the second simple closed or- bit γ2. Hence γ1 and γ2 are both unknotted and have linking number 1, which concludes the proof of Theorem 1.

Remark 14. In this remark we assume abstract perturbations exist so that the linearized contact homology is well defined for all tight contact forms on S3 with nondegenerate closed Reeb orbits. As mentioned in Remark 2, it is immediate from the formulas for Conley-Zehnder indices together with the computation of the linearized contact homology of a tight contact form on S3 that condition (i) in Theorem 1 implies dynamical convexity. We show that (ii) implies (i) and hence (ii) implies dynamical convexity as well. Suppose that the differential in linearized contact homology does not vanish. This can only happen if there is at least one even hyperbolic orbit γ1 (whose iterates have odd degree) and one elliptic or odd hyperbolic orbit γ2 (whose good iterates have even degree). If r2 6= 0 then we conclude that (i) holds exactly as in the proof above: the iteration formulas for Conley-Zehnder indices show that multiples of γ2 cannot attain all even degrees, so there must be a third orbit. If = | | = | | = ± | | = − | k | = − r2 0 then γ2 0, so we must have γ1 1. If γ1 1, then γ1 1 for all k ≥ 1, so that infinitely many orbits are required to eliminate these generators in | | = | k| = − ≥ homology. If γ1 1, then γ1 2k 1 for all k 1, so that in order to eliminate these generators in homology and span exactly HC lin(S3) in degrees 2 to 2k − 2, | 2k−2| < − < | 2k| > we must have γ2 2k 1 γ2 for all k 1. This implies that γ2 is elliptic |γ2k | = 2 = and that 4α2 limk→∞ k 2, a contradiction. This proves that (ii) implies (i).

Acknowledgments. We thank E. Volkov for helpful comments, and D. Kotschick for pointing out Remark 13.

REFERENCES

[1] F. Bourgeois, T. Ekholm and Y. Eliashberg, A surgery exact sequence for linearized contact homology, in preparation. [2] F. Bourgeois, Y. Eliashberg, H. Hofer, K. Wysocki and E. Zehnder, Compactness results in Symplectic Field Theory, Geom. and Top., 7 (2003), 799–888. [3] F. Bourgeois and K. Mohnke, Coherent orientations in symplectic field theory, Math. Z., 248 (2004), 123–146. [4] F. Bourgeois and A. Oancea, An exact sequence for contact and symplectic homology,

. [5] K. Cieliebak and J. Latschev, Symplectic field theory and string : the exact case, in preparation.

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613 A NOTEON REEBDYNAMICSONTHETIGHT 3-SPHERE 613

[6] K. Cieliebak and K. Mohnke, Compactness for punctured holomorphic curves, J. Symp. Geom., 3 (2006), 1–65. [7] K. Cieliebak and K. Mohnke, Symplectic hypersurfaces and transversality in Gromov-Witten

theory, . [8] Y. Eliashberg, Contact 3-manifolds twenty years since J. Martinet’s work, Ann. Inst. Fourier (Grenoble) 42 (1992), 165–192. [9] Y. Eliashberg, A. Givental and H. Hofer, Introduction to symplectic field theory, GAFA 2000 Visions in Mathematics special volume, part II, 560–673. [10] B. Fayad and A. Katok, Constructions in elliptic dynamics, Ergodic Theory Dynam. Syst., 24 (2004), 1477–1520. [11] X. Hu, Y. Long and W. Wang, Resonance identity, stability and multiplicity of closed charac-

teristics on compact convex hypersurfaces, . [12] Y. Long, “Index Theory for Symplectic Paths with Applications," Birkhäuser, 2002. [13] H. Hofer, K. Wysocki and E. Zehnder, The dynamics on three-dimensional strictly convex energy surfaces, Ann. of Math., 148 (1998), 197–289.

[14] H. Hofer, K. Wysocki and E. Zehnder, A General Fredholm Theory I-II, . [15] D. Salamon, Lectures on Floer homology, Symplectic and topology (Park City

1997), , IAS/Park City Math. Ser., Amer. Math. Soc., 7 (1999), 143–229.

 FRÉDÉRIC BOURGEOIS : Département de Mathématique, Université Li-

bre de Bruxelles CP 218, Boulevard du Triomphe, 1050 Bruxelles, Belgium

 KAI CIELIEBAK : Mathematisches Institut der LMU München, Theresien-

str. 39, 80333 München, Germany

 TOBIAS EKHOLM : Department of Mathematics Uppsala University, Box 480, 751 06 Uppsala, Sweden

JOURNAL OF MODERN DYNAMICS VOLUME 1, NO. 4 (2007), 597–613