<<

Exotic and contact circle actions

Dušan Drobnjak∗ and Igor Uljarević†

March 5, 2021

Abstract Using Floer-theoretic methods, we prove that the non-existence of an exotic on the standard symplectic ball, 2n B , implies a rather strict topological condition on the free contact 2n−1 circle actions on the standard contact sphere, S . We also prove an analogue for a Liouville domain and contact circle actions on its boundary. Applications include results concerning the symplectic mapping class group and the fundamental group of the group of contactomorphisms.

1 Introduction

A well-known open problem in symplectic topology is that of the exis- tence of the so-called exotic symplectomorphisms on the standard 2n- dimensional symplectic ball, B2n. A compactly supported symplectomor- phism B2n → B2n is called exotic if it is not isotopic to the identity relative to the boundary through symplectomorphisms. Apart from the cases where n = 1 and n = 2 (where the topology of the group of compactly supported symplectomorphisms B2n → B2n is well understood due to Gromov’s arXiv:2004.10828v2 [math.SG] 3 Mar 2021 theory of J-holomorphic curves [13]) very little is known about exotic sym- plectomorphisms of B2n. For instance, it is an open problem whether an exotic diffeomorphism of B2n can be realized as a symplectomorphism with respect to the standard symplectic structure on B2n (compare to [8]

∗Faculty of mathematics, University of Belgrade, Studentski trg 16, 11158 Belgrade, Serbia; e-mail: [email protected] †Faculty of mathematics, University of Belgrade, Studentski trg 16, 11158 Belgrade, Serbia; e-mail: [email protected]

1 where such a realization is constructed in the case of a non-standard sym- plectic structure on a ball). In other words, it is not known whether the problem of the existence of the exotic symplectomorphisms on the standard symplectic ball can be solved in the framework of differential topology. On the other hand, there are numerous results regarding exotic sym- plectomorphisms on other symplectic [27, 28, 10, 26, 9, 30, 29, 31, 3, 32]. The most notable is work of Seidel [27] who constructed the first known example of a symplectomorphism that is smoothly isotopic to the identity but not symplectically, thus proving that exotic symplectomor- phisms are genuinely a symplectic phenomenon. In the present paper, we prove that non-existence of an exotic sym- plectomorphism on the standard B2n implies a rather strict topological condition on the free contact circle actions on its boundary (i.e. the stan- dard contact sphere S2n−1). We call this condition topological symmetry. It is expressed in terms of the reduced homology (denoted by H˜ ∗ below) of the part of S2n−1 on which the circle action is positively transverse to the contact distribution.

Definition 1.1. Let ξ be a cooriented contact distribution on S2n−1, and let 2n−1 2n−1 ϕt :(S , ξ) → (S , ξ) be a free contact circle action. Denote by Y 2n−1 the vector field that generates ϕt, and by P ⊂ S the set of points in S2n−1 at which the vector field Y represents the positive coorientation of ξ. The contact circle action ϕ is said to be topologically symmetric if there exists m ∈ Z such that

dim H˜ m−k(P; Z2) = dim H˜ k(P; Z2) for all k ∈ Z.

Theorem 1.2. For all n ∈ N, at least one of the following statements is true.

A There exists an exotic symplectomorphism of B2n .

B Every free contact circle action on S2n−1 is topologically symmetric. In fact, we prove a more general statement about exotic symplectomor- phisms of Liouville domains and contact circle actions on their boundaries (see Theorem 4.3 on page 14). A consequence of this more general result is that topologically asymmetric free contact circle actions represent nontriv- ial elements in the fundamental group of the group of contactomorphisms (for a precise statement see Corollary 4.4 on page 15). Theorem 1.2 together with the result of Gromov that there are no exotic symplectomorphisms on B4 implies the following.

2 Corollary 1.3. Every free contact circle action on the standard contact S3 is topologically symmetric.

The paper is organized as follows. Section 2 recalls the preliminaries. Section 3 introduces the notion of topological symmetry and proves some of the basic properties of the free contact circle actions. Section 4 proves Theorem 1.2. Section 5 contains the main technical result, that relies on the on a with boundary.

Acknowledgements We would like to thank Paul Biran, Aleksandra Marinković, Darko Milinković, Vukašin Stojisavljević, and Filip Živanović for useful feedback. This work was partially supported by the Ministry of Education, Science, and Technological development, grant number 174034.

2 Preliminaries

2.1 Notation and conventions

Let (W, ω) be a . The Hamiltonian vector field, XHt , of the Hamiltonian Ht : W → R is the vector field on W defined by H ω(XHt , ·) = dHt. We denote by φt : W → W the Hamiltonian isotopy of H H H the Hamiltonian Ht : W → R, i.e. ∂tφt = XHt ◦ φt and φ0 = id . Let Σ be a contact manifold with a contact form α. The Reeb vector field Rα is the unique vector filed on Σ such that α(Rα) = 1 and such that α h dα(R , ·) = 0. We denote by ϕt : Σ → Σ the contact isotopy furnished by a contact Hamiltonian ht : Σ → R . In our conventions, the vector field Y of h the isotopy ϕt and the contact Hamiltonian h are related by h = −α(Y). A contact Hamiltonian ht : Σ → R is called strict (with respect to the contact α h∗ form α) if dht(R ) = 0. This is equivalent to ϕt α = α, for all t ∈ R.

2.2 Liouville domains Definition 2.1. A Liouville domain is a compact manifold W together with a 1-form λ such that the following holds.

• The 2-form dλ is a symplectic form on W.

• The restriction of λ to the boundary ∂W is a contact form that induces the boundary orientation on ∂W.

3 A part of the symplectization of ∂W naturally embeds into W. More precisely, there exists a unique embedding ι : ∂W × (0, 1] → W such that ι(x, 1) = x, and such that ι∗λ = r·α. Here, r is the (0, 1] coordinate function and α := λ|∂W is the contact form on ∂W. The completion, W,c of the Liouville domain W is obtained by gluing W and ∂W × (0, ) via ι. The sets W and ∂W × (0, ) can be seen as subsets of the completion, W.c The manifold Wc is an exact symplectic manifold with a Liouville∞ form given by ∞ λ on W, bλ := r · α on ∂W × (0, ).

With a slight abuse of notation, we will write λ instead of bλ. ∞ 2.3 A homotopy long exact sequence In this section, we describe a construction that is due to Biran and Giroux [4]. Given a Liouville domain (W, λ), denote by Σ the manifold ∂W, and by α the contact form λ|∂W on Σ. Denote by Cont Σ, Sympc W, Symp(W, λ), and Sympc(W, λ) the groups of diffeomorphisms defined in the list below.

Cont Σ is the group of the contactomorphisms of (Σ, ker α).

Sympc W is the group of the symplectomorphisms of (W, dλ) that are equal to the identity in a neighbourhood of the boundary.

Symp(W, λ) is the group of the so-called exact symplectomorphisms of (W,c dλ) that preserve the Liouville form λ in Σ × [1, ). More precisely, a symplectomorphism φ : Wc → Wc is an element of Symp(W, λ) if, and only if, there exists a smooth function F∞: Wc → R ∗ such that F|Σ×[1−ε, ) ≡ 0, for some ε > 0, and such that φ λ−λ = dF.

Sympc(W, λ) is the group∞ of the symplectomorphisms in Symp(W, λ) that are equal to the identity on Σ × (1 − ε, ) for some ε > 0.

Since a symplectomorphism φ ∈ Symp(∞W, λ) preserves the Liouville form λ on the cylindrical end Σ × [1, ), there exists a contactomorphism ϕ : Σ → Σ such that φ has the following form on the cylindrical end ∞ φ(x, r) = (ϕ(x), ?) ∈ Σ × (0, )

(x, r) ∈ Σ × [a, ), a ∈ + for all where R is∞ big enough. The map Symp(W, λ) → Cont Σ defined by φ 7→ ϕ is called the ideal restriction ∞ 4 map. It turns out that the ideal restriction map is a Serre fibration with the

fibre above the identity equal to Sympc(W, λ). Hence, there is a homotopy long exact sequence

··· πk Sympc(W, λ) πk Symp(W, λ) πk Cont Σ

πk−1 Sympc(W, λ) ··· .

The groups Sympc(W, λ) and Sympc W are homotopy equivalent [4] (see also [31, Lemma 3.3]). Therefore, there is a homotopy long exact sequence

··· πk Sympc W πk Symp(W, λ) πk Cont Σ

πk−1 Sympc W ··· .

Particularly important in this paper is the boundary map

Θ : π1 Cont Σ → π0 Sympc W from the exact sequence above. The map Θ can be described as follows. Given a loop 1 ϕ : S → Cont Σ : t 7→ ϕt.

Let ht : Σ → R be a contact Hamiltonian that generates it. Choose a Hamiltonian Ht : Wc → R such that

Ht(x, r) = r · ht(x) for (x, r) ∈ Σ × (1 − ε, ) for some ε > 0. Then, Θ([ϕ]) = φH . 1 ∞ 2.4 Floer theory The that is utilized in this paper is the Floer homology for contact Hamiltonians. It has been introduced by Merry and the second author in [22] as a consequence of a generalized no-escape lemma. Our re- sults, however, use only the case of strict contact Hamiltonians (i.e. contact Hamiltonians whose isotopies preserve not only the contact distribution but also the contact form), the case considered already in [25], [12] and [31]. The Floer homology for a contact Hamiltonian, HF∗(h), is associated to a contact Hamiltonian h : ∂W × S1 → R defined on the boundary of a

5 Liouville domain W such that the time-one map of h has no fixed points. By definition, HF∗(h) is equal to the Hamiltonian loop Floer homology for 1 a Hamiltonian H : Wc × S → R such that Ht(x, r) = r · ht(x) whenever (x, r) ∈ ∂W × [1, ).

2.4.1 Floer data∞

To define HF∗(h), one has to choose auxiliary data consisting of a Hamilto- 1 1 H : W × → {J } 1 nian c S R and an S family t t∈S of almost complex structures on Wc that satisfy the following conditions.

1. Conditions on the cylindrical end.

• Ht(x, r) = r · ht(x), for (x, r) ∈ ∂W × [1, + ), • dr ◦ Jt = −λ, in W × [1, + ). ∞ 2. Non-degeneracy. For each fixed point x of φH : W → W, ∞ 1 c c H  det dφ1 (x) − id 6= 0.

1 3. dλ-compatibility. dλ(·,Jt·) is an S family of Riemannian metrics on W.c 4. Regularity. The linearized operator of the Floer equation

1 u : R ×S → W,c

∂su + Jt(u)(∂tu − XHt (u)) = 0

is surjective.

2.4.2 Floer complex

The Floer complex, CF∗(H, J), is generated by the contractible 1-periodic orbits of the Hamiltonian H. I.e. M CFk(H, J) := Z2 hγi , deg γ=k where the sum goes through the set of the contractible 1-periodic orbits of H that have the degree equal to k. The degree, deg γ, is defined as the negative Conley-Zehnder index of the path of symplectic matrices obtained from H dφt (γ(0)) by trivializing TWc along a disk that bounds γ. Due to different

6 choices of the disk that bounds γ, deg γ is only defined up to a multiple of 2N, where 2 N := min c1(u) > 0 | u : S → Wc is the minimal Chern number. Hence we see deg γ as an element of Z2N, and the Floer complex CF∗(H, J) is Z2N graded. The differential ∂ : CFk+1(H, J) → CFk(H, J) is defined by counting the isolated un- parametrized solutions of the Floer equation. More precisely,

∂ hγi = n(γ, γ˜ ) hγ˜ i , X where n(γ, γ˜ ) is the number modulo 2 of the isolated unparametrized solutions u : R × S1 → Wc of the following problem

∂su + Jt(u)(∂tu − XHt (u)) = 0, lim u(s, t) = γ(t), s→− lim u(s, t) = γ˜ (t). s→+∞

Note that, if deg γ˜ 6= deg γ∞− 1, then there are no isolated unparametrized solutions of the Floer equation from γ to γ.˜ Hence, in this case, n(γ, γ˜ ) = 0, and ∂ is well defined. The Floer homology, HF∗(H, J), is the homology of the Floer complex CF∗(H, J).

2.4.3 Continuation maps Given Floer data (H−,J−) and (H+,J+). Continuation data from (H−,J−) to (H+,J+) consists of a (s-dependent) Hamiltonian H : Wc × R ×S1 → R {J } 1 W and a family s,t (s,t)∈R ×S of almost complex structures on c such that 1. there exists a smooth function h : ∂W × R ×S1 → R that is non- decreasing in R-variable and such that

Hs,t(x, r) := H((x, r), s, t) = r · h(x, s, t) =: r · hs,t(x)

for all (x, r) ∈ ∂W × [1, ),

1 2. dr ◦ Js,t = −λ in ∂W × [1, ), for all (s, t) ∈ × , ∞ R S dλ(·,J ·) W (s, t) ∈ × 1, 3. s,t is a Riemannian∞ metric on c for all R S ± ± 4. (Hs,t,Js,t) = (Ht ,Jt ) for ±s >> 0.

7 Let γ− be a 1-periodic orbit of H−, and let γ+ be a 1-periodic orbit of + − − + + H . For generic continuation data (Hs,t,Js,t) from (H ,J ) to (H ,J ), the set of the solutions u : R ×S1 → Wc of the problem

∂su + Js,t(u)(∂tu − XHs,t (u)) = 0, lim u(s, t) = γ±(t) s→± is a finite union of compact∞ manifolds (possibly of different dimensions) cut out transversely by the Floer equation. Denote by n(γ−, γ+) the number of its 0-dimensional components. The continuation map − − + + Φ = Φ({Hs,t}, {Js,t}): CF∗(H ,J ) → CF∗(H ,J ) is the linear map defined on the generators by

Φ(γ−) := n(γ−, γ+) hγ+i . γ+ X Since there are no 0-dimensional components of the above mentioned man- ifold if deg γ− 6= deg γ+, continuation maps preserve the grading. By the condition 1 for continuation data, continuation maps − − + + CF∗(H ,J ) → CF∗(H ,J )

− + are defined only if H 6 H on the cylindrical end ∂W×[1, ). The contin- − − + + uation maps CF∗(H ,J ) → CF∗(H ,J ) defined with respect to different − − + + continuation data form (H ,J ) to (H ,J ) are chain homotopic.∞ Hence, − − + + they induce the same map HF∗(H ,J ) → HF∗(H ,J ) on the homology level. The induced map is also called the continuation map. As opposed to the compact case, the continuation maps need not be isomorphisms. They do, however, satisfy the following relations. α α α α α 1. The continuation map Φα : HF∗(H ,J ) → HF∗(H ,J ) is equal to the identity. 2. The composition of the continuation maps β α α β β Φα : HF∗(H ,J ) → HF∗(H ,J ) and γ β β γ γ Φβ : HF∗(H ,J ) → HF∗(H ,J ) is equal to the continuation map γ α α γ γ Φα : HF∗(H ,J ) → HF∗(H ,J ), γ β γ i.e. Φβ ◦ Φα = Φα.

8 In other words, the family of groups {HF∗(H, J)} together with the continua- tion maps form a directed system of groups. As a consequence, the groups α α β β α β HF∗(H ,J ) and HF∗(H ,J ) are canonically isomorphic if H = H on 1 ∂W × [1, ). Therefore, the group HF∗(h), where h : ∂W × S → R is a 1-periodic contact Hamiltonian whose time-1 map has no fixed α β 1 points, is∞ well defined. Moreover, if h , h : ∂W × S → R are two 1- periodic contact Hamiltonians whose time-1 maps have no fixed points α β and such that h 6 h , then there is a well defined continuation map α β HF∗(h ) → HF∗(h ).

2.4.4 Naturality isomorphisms Let H, F be Hamiltonians, denote by H and H#F the Hamiltonians that H−1 H F generate Hamiltonian isotopies φt and φt ◦ φt , respectively. The naturality isomorphism

N(F): HF(H) → HF(F#H) is associated to a 1-periodic Hamiltonian F : Wc × S1 → R whose Hamil- tonian isotopy is 1-periodic (i.e. F generates a loop of Hamiltonian diffeo- morphisms). On the chain level, on generators, the map N(F) is defined by hγi 7→ (φF)∗γ ,

F ∗ 1 F −1 where (φ ) γ : S → Wc is the loop t 7→ (φt ) (γ(t)). The naturality map is an isomorphism already on the chain level. The naturality isomorphisms do not preserve the grading in general. They do respect the grading up to a constant shift though.

3 Topologically symmetric contact circle actions

A contact circle action is a action of S1 on a contact manifold by contactomorphisms. A contact circle action on Σ can be seen as a 1-periodic family ϕt : Σ → Σ, t ∈ R of contactomorphisms such that

ϕs ◦ ϕt = ϕs+t, for all s, t ∈ R . In particular, every contact circle action on Σ can be seen as a flow of an autonomous vector field on Σ.

Definition 3.1. Let ϕt : Σ → Σ be a contact circle action on a (cooriented) contact manifold (Σ, ξ) that is the boundary of a Liouville domain W with

9 c1(W) = 0. Let h : Σ → R be the contact Hamiltonian that generates ϕt defined with respect to some contact form on Σ. Denote Σ+ := {p ∈ Σ | h(p) > 0} .

The contact circle action ϕt is called topologically symmetric (with respect to the filling W) if there exists an integer m ∈ Z such that + + (∀k ∈ Z) dim Hm−k(W, Σ ; Z2) = dim Hk(W, Σ ; Z2).

+ + Here, Hk(W, Σ ; Z2) stands for the singular homology of the pair (W, Σ ) + in Z2 coefficients. The set Σ is referred to as the positive region of the contact circle action ϕ. Similarly, the negative region of the contact circle action ϕ is the set Σ− := {p ∈ Σ | h(p) < 0} . The set Σ+ in the definition above (and consequently the notion of the topologically symmetric contact circle action) does not depend on the choice of the contact form on Σ, although h does. Namely, Σ+ can be defined as the set of the points p ∈ Σ such that the vector field of ϕt at the point p represents the negative coorientation of the contact distribution ξ at the point p.

Example 3.2. The Reeb flow on the standard contact sphere Σ := S2n−1 is an example of a topologically symmetric contact circle action with respect to the standard symplectic ball W := B2n. Indeed, the Reeb flow is generated by the constant contact Hamiltonian

Σ → R : x 7→ −1. Hence, the corresponding positive region Σ+ is equal to the empty set. + 2n Consequently, H∗(W, Σ ; Z2) is isomorphic to H∗(B ; Z2). This implies the topological symmetry. Example 3.3. Let ε > 0 be a sufficiently small number and let W be the Brieskorn variety

n+1 3 3 W := (z0, . . . , zn) ∈ C | z0 + ··· + zn = ε & |z| 6 1 . The boundary∂W is contactomorphic to the Brieskorn manifold

n+1 3 3 Σ := (z0, . . . , zn) ∈ C | z0 + ··· + zn = 0 & |z| = 1 . Consider the free contact circle action on Σ given by

2πit Σ × R 3 (x, t) 7→ e 3 · z ∈ Σ.

10 + + The positive region, Σ , is equal to the empty set. Therefore, H∗(W, Σ ; Z2) is isomorphic to H∗(W; Z2). The Brieskorn variety W is homotopy equiv- alent to the wedge of 2n copies of Sn. Hence, the contact circle action above is not topologically symmetric. For a detailed account on Brieskorn manifolds, see [19].

Lemma 3.4. Let ϕt : Σ → Σ be a contact circle action on a cooriented contact ∗ manifold Σ. Then, there exists a contact form α on Σ such that ϕtα = α for all t ∈ R . Proof. The lemma is a special case of Proposition 2.8 in [21].

Lemma 3.5. Let ϕt : Σ → Σ be a free contact circle action on a (cooriented) contact manifold Σ, and let α be a contact form on Σ that is invariant under ϕt. Denote by h : Σ → R the contact Hamiltonian of ϕt defined with respect to α. Then, 0 is a regular value of the function h.

Proof. Let Y be the vector field on Σ that generates ϕt, i.e. ∂tϕt = Y ◦ ϕt. Then, by definition, the generating contact Hamiltonian is h = −α(Y). The Cartan formula (together with the invariance of α under ϕt) implies d 0 = (ϕ∗α) = ϕ∗ (dα(Y, ·) + d(α(Y))) . dt t t

−1 Hence, dh = dα(Y, ·). If p ∈ h (0), then Y(p) ∈ ker αp is a non-zero vector that belongs to the contact distribution (it is non-zero because the circle action ϕt is a free action). Since dα is non-degenerate when restricted to the contact distribution, the 1-form dh(p) = dα(Y(p), ·) is non-degenerate. Therefore, p ∈ h−1(0) cannot be a critical point of h, i.e. 0 is a regular value of h.

Remark 3.6. In the situation of Definition 3.1, if 0 is a regular value of h, [15, Theorem 3.43] implies that by replacing Σ+ by Σ− in the definition one obtains an equivalent definition. More precisely, the contact circle action ϕ is topologically symmetric with respect to W if, and only if, there exists m ∈ Z such that

− − dim Hm−k(W, Σ ; Z2) = dim Hk(W, Σ ; Z2), for all k ∈ Z.

11 4 Topologically asymmetric contact circle actions and the topology of transformation groups

Section 2.3 discussed a method of constructing a symplectomorphism φ : W → W of a Liouville domain from a loop of contactomorphisms ϕt : ∂W → ∂W of its boundary. If ht : ∂W → R is the contact Hamiltonian generating ϕt, then the symplectomorphism φ is obtained as the time-1 map of a Hamiltonian Ht : W → R that is equal to r · ht on the cylindrical end. The method gives rise to a homomorphism

h H Θ : π0 Cont(∂W) → π0 Sumpc W :[ϕ] = [ϕ ] 7→ [φ1 ] = [φ]. This section proves the main result of the paper: topologically asym- metric free contact circle actions furnish (via Θ) non-trivial elements of

π0 Sympc W. The next lemma will be used to reduce to the case where the free contact circle action preserves not only the contact distribution but also the contact form on ∂W. Lemma 4.1. Let (W, λ) be a Liouville domain, and let β be a contact form on Σ := ∂W. Then, there exists a Liouville form µ on W such that β = µ|Σ , and such that the following holds. If

λ Θ : π1 Cont Σ → π0 Sympc(W, dλ), µ Θ : π1 Cont Σ → π0 Sympc(W, dµ)

λ are the homomorphisms from Section 2.3, and if η ∈ π1 Cont Σ, then Θ (η) = 0 if and only if Θµ(η) = 0.

Proof. Denote α := λ|Σ . Since α and β determine the same (cooriented) + contact structure on Σ, there exists a positive function f : Σ → R such that β = fα. Let V ⊂ Wc be the complement of the set

+ (x, r) ∈ Σ × R ⊂ Wc | r > f(x) .

Since the Liouville vector field Xλ (defined by λ = Xλydλ) is nowhere + vanishing in Σ × R , and since it is transverse to both ∂V and ∂W, the manifolds V and W are diffeomorphic. Denote by Ψ : W → V the dif- feomorphism furnished by Xλ. For x ∈ ∂W, Ψ(x) = (x, f(x)). Hence, the ∗ one-form µ := Ψ λ satisfies µ|Σ = β. Let ϕt : Σ → Σ be a loop of contactomorphisms that represents the class η, and let ht : Σ → R be the contact Hamiltonian with respect to the contact form α associated to ϕ. Then, the contact Hamiltonian of ϕ with respect

12 to β is equal to f · ht : Σ → R . Let Ht : Wc → R be a Hamiltonian that is equal to r · ht on the complement of int V ∩ int W (this condition makes sure that the time-1 maps of both H and H ◦ Ψ are compactly supported in int W). Then, the Hamiltonian Ht ◦ Ψ : W → R is equal to ρ · (f · ht) near the boundary, where ρ stands for the cylindrical coordinate of the Liouville domain (W, µ). This implies λ  H Θ (η) = φ1 ∈ π0 Sympc(W, dλ), µ  H◦Ψ  −1 H  Θ (η) = φ1 = Ψ ◦ φ1 ◦ Ψ ∈ π0 Sympc(W, dµ). λ If Θ (η) = 0, then there exists a symplectic isotopy φt :(W, dλ) → (W, dλ) H λ relative to the boundary from the identity to φ1 . Denote by φt : Wc → Wc + the flow of the Liouville vector field Xλ. For c ∈ R large enough, the symplectomorphism λ−1 λ φ˜ t := φc ◦ φt ◦ φc is compactly supported in int V ∩ int W for all t ∈ [0, 1]. Additionally,

λ −1 λ s 7→ φc·s ◦ φ1 ◦ φc·s, s ∈ [0, 1],

H is a symplectic isotopy, compactly supported in int V ∩int W, from φ1 = φ1 to φ˜ 1. Denote by φt, t ∈ [0, 1], the isotopy that is obtained by concate- −1 nating φ˜ and the inverse of the isotopy above. Then, Ψ ◦ φt ◦ Ψ is a symplectic isotopy in (W, dµ) relative to the boundary from the identity −1 H H◦Ψ µ to Ψ ◦ φ1 ◦ Ψ = φ1 . Hence, Θ (η) = 0. The other direction can be proven similarly. Theorem 4.3 below is a generalisation of Theorem 1.2 that was stated in the introduction. It uses Proposition 4.2 which is stated here and proved in Section 5. Proposition 4.2. Let W be a Liouville domain with the boundary Σ := ∂W, and let h : Σ → R be a contact Hamiltonian such that 0 is a regular value of h, and such that h has no periodic orbits of period less than or equal to ε, for some ε > 0. Denote

Σ− := {x ∈ Σ | h(x) < 0} , Σ+ := {x ∈ Σ | h(x) > 0} .

Then, ∼ + HFk(εh) = Hk+n(W, Σ ; Z2), ∼ − HFk(−εh) = Hk+n(W, Σ ; Z2).

13 Theorem 4.3. Let W be a Liouville domain such that c1(W) = 0, and let ϕt : ∂W → ∂W be a free contact circle action that is not topologically symmetric with respect to W. Then, Θ([ϕt]) is a non-trivial symplectic mapping class in

π0 Sympc(W). Proof. Denote by λ the Liouville form on W, and by h : ∂W → R the contact Hamiltonian of ϕt with respect to the contact form α := λ|∂W . Lemma 3.4 on page 11 above implies that there exists a contact form α0 on ∂W such ∗ 0 0 that ϕtα = α for all t ∈ R. By this lemma and Lemma 4.1, without loss of generality, we may assume that ϕt preserves the contact form α for all t. Assume, by contradiction, that Θ([ϕt]) is a trivial symplectic mapping class in π0 Sympc(W). As explained in Section 2.3, Θ([ϕt]) is represented by H the time-one map, φ1 , of a Hamiltonian H : Wc → R that is equal to r · h on the cylindrical end (the Hamiltonian H can be chosen to be autonomous). The inclusion

Sympc(W, λ) ,→ Sympc(W) H is a homotopy equivalence [31, Lemma 3.3], and φ1 is an exact symplec- tomorphism that is isotopic to the identity through symplectomorphisms H relative to the cylindrical end. Therefore, φ1 is isotopic to the identity through exact symplectomorphisms relative to the cylindrical end. Since every isotopy through exact symplectomorphisms is actually a Hamilto- nian isotopy, there exists a Hamiltonian Gt : Wc → R that is equal to 0 on G H the cylindrical end, and such that φ1 = φ1 . Denote F := H#G. Let ε ∈ (0, 1). The Hamiltonian F is equal to r·h on the cylindrical end. This, together with h being a strict contact Hamiltonian, implies F h  φt (x, r) = ϕt (x), r , for r big enough. In particular,

F r ◦ φt = r, where r is seen as the coordinate function r :(x, r) 7→ r. Since h is au- h tonomous and strict, h ◦ ϕt = h for all t ∈ R. Hence, the Hamiltonian

−1   F F#(ε · H) t = Ft + (εHt) ◦ φt F F = −Ft ◦ φt + (εHt) ◦ φt F = (εHt − Ft) ◦ φt is equal to (εh − h) · r = (ε − 1) · h · r

14 on the cylindrical end. The naturality isomorphism

N(F): HF(ε · h) → HF((ε − 1) · h) is well defined because F generates a loop of Hamiltonian diffeomorphisms. Hence, if c ∈ Z denotes the shift in grading,

dim HFk(ε · h) = dim HFk+c((ε − 1) · h), for all k ∈ Z. Since the contact Hamiltonian h has no orbits of the period in (0, ε) ∪ (0, 1 − ε), Proposition 4.2 on page 13 below implies

+ dim HFk(ε · h) = dim Hk+n(W, Σ ; Z2), − dim HFk+c((ε − 1) · h) = dim Hk+c+n(W, Σ ; Z2), for all k ∈ Z. Here, Σ+ := {p ∈ Σ | h(p) > 0} , Σ− := {p ∈ Σ | h(p) < 0} .

Therefore, + − dim Hk(W, Σ ; Z2) = dim Hk+c(W, Σ ; Z2), for all k ∈ Z. A generalization of the Lefschetz duality (Theorem 3.43. in [15]) implies − ∼ 2n−k + Hk(W, Σ ; Z2) = H (W, Σ ; Z2), for all k ∈ Z. Consequently,

+ − dim Hk(W, Σ ; Z2) = dim Hk+c(W, Σ ; Z2) 2n−c−k + = dim H (W, Σ ; Z2) + = dim Hom (H2n−c−k(W, Σ ; Z2), Z2) + = dim H2n−c−k(W, Σ ; Z2), for all k ∈ Z. This contradicts the assumption that ϕ is not topologically symmetric.

Corollary 4.4. Let W be a Liouville domain such that c1(W) = 0, and let ϕt : ∂W → ∂W be a free contact circle action that is not topologically symmetric with respect to W. Then, ϕt determines a non-trivial element [ϕt] ∈ π1 Cont(∂W). In other words, the loop of contactomorphisms {ϕt} is not contractible.

15 Proof. Since Θ is a group homomorphism, the triviality of

[ϕt] ∈ π1 Cont(∂W) implies the triviality of Θ ([ϕt]) ∈ π0 Sympc(W), which contradicts Theo- rem 4.3. Remark 4.5. Theorem 4.3 and Corollary 4.4 hold also in the case where c1(W) 6= 0. However, one should then understand the notion of topological symmetry in the following way. Let N be the minimal Chern number of W, and let ϕt : ∂W → ∂W be a free contact circle action with the positive + region Σ ⊂ ∂W. Denote by aj, j ∈ Z2N the number

+ aj := dim Hk(W, Σ ; Z2). k≡j (Xmod 2N) The contact circle action ϕ is topologically symmetric if there exists m ∈ Z2N such that ak = am−k, for all k ∈ Z2N. Proofs of Theorem 4.3 and Corollary 4.4 in the case where c1(W) 6= 0 are the same as in the case where c1(W) = 0 except at one point, which is discussed next. Let (H, J) be a regular Floer data such that H : W → R is a C2 small Morse function. If c1(W) 6= 0, the chain complexes CF∗(H, J) and CM∗+n(H, JXH) are not identical. Namely, the chain complex HF∗(H, J) is Z2N-graded whereas CM∗+n(H, JXH) is Z-graded. Instead, HF∗(H, J) coincides with the Z2N- graded chain complex obtained by “rolling up” CM∗+n(H, JXH) modulo 2N. More precisely, M CFj(H, J) = CMk+n(H, JXH). k≡j (mod 2N)

Hence, the number aj above is actually the dimension of the group HFj−n(ε · h), where h is the contact Hamiltonian of ϕt and where ε > 0 is a sufficiently small positive number. Let us now explain how to derive Theorem 1.2 from Theorem 4.3. 2n−1 2n−1 Lemma 4.6. Let ϕt : S → S be a free contact circle action on the standard 2n−1 sphere, let α be a contact form on S . Denote by Y the vector field of ϕt, and by P the set 2n−1 P := p ∈ S | α(Y(p)) > 0 . Then, ϕt is topologically symmetric with respect to the standard symplectic ball, B2n, if and only if, there exists m ∈ Z such that

dim H˜ m−k(P; Z2) = dim H˜ k(P; Z2) for all k ∈ Z. Here, H˜ stands for the reduced singular homology.

16 Proof. Since B2n is contractible, from the long exact sequence for the re- duced singular homology of the pair (B2n,P) we deduce ∼ 2n H˜ k(P; Z2) = Hk+1(B ,P; Z2), for all k ∈ Z. The contact Hamiltonian that generates ϕt is given by h = −α(Y), and therefore, the set P is the negative region of the contact circle action ϕ in the sense of Definition 3.1. The topological symmetry of 2n ϕt with respect to B is equivalent to the existence of m ∈ Z such that

2n 2n dim Hm−k(B ,P; Z2) = dim Hk(B ,P; Z2), for all k ∈ Z. Using the above mentioned isomorphism, this is further equivalent to

dim H˜ m−2−(k−1)(P; Z2) = dim H˜ k−1(P; Z2), for all k ∈ Z, which finishes the proof. Theorem 1.2. For all n ∈ N, at least one of the following statements is true. A There exists an exotic symplectomorphism of B2n (i.e. there exists a non-trivial 2n element of π0 Sympc(B )).

B Every free contact circle action on S2n−1 is topologically symmetric. Proof. This is a direct consequence of Theorem 4.3 and Lemma 4.6.

5 Morse theory

In this section, we prove the main technical result that computes the Floer homology for a contact Hamiltonian with sufficiently small absolute value. Using the standard argument, one can reduce the Floer homology to the Morse homology of a function on a manifold with non-empty boundary. The Morse theory for manifolds with boundary has been intensively stud- ied [23, 17, 7, 18, 20]. A Morse function whose restriction to the boundary is also Morse is called an m-function. Given an m-function f : W → R, it is known that the critical points of f together with some of the critical points of f|∂W recover the singular homology of W [18, 20]. Whether a critical point of f|∂W will be taken into account or not is determined by the direction in which the gradient ∇f points at that point. Namely, a critical point p of f|∂W will be ignored if, and only if, the gradient ∇f(p) points outwards.

17 The results in the literature do not cover entirely the Morse theory re- quired in the proof of Proposition 4.2. For instance, the proof will deal with Morse functions whose restrictions to the boundary are not necessar- ily Morse. For the convenience of the reader, we recall the statement of Proposition 4.2.

Proposition 4.2. Let W be a Liouville domain with the boundary Σ := ∂W, and let h : Σ → R be a contact Hamiltonian such that 0 is a regular value of h, and such that h has no periodic orbits of period less than or equal to ε, for some ε > 0. Denote

Σ− := {x ∈ Σ | h(x) < 0} , Σ+ := {x ∈ Σ | h(x) > 0} .

Then,

∼ + HFk(εh) = Hk+n(W, Σ ; Z2), ∼ − HFk(−εh) = Hk+n(W, Σ ; Z2).

Proof. The proof is divided into several steps. In the first step, we pass from Floer to Morse homology. Steps 2-7 reduce the proof (by pasting and cutting) to the known case where the Morse function is constant on the boundary components. Figure 2 on page 26 illustrates the proof. Lemma 5.1 below states that the groups HF∗(a · h) and HF∗(b · h) are isomorphic provided that the contact Hamiltonian h has no closed orbits of period in [a, b]. Hence, we may assume, without loss of generality, that ε > 0 is arbitrary small. Step 1 (Passing to the Morse homology). Let H : Wc → R be a Morse function such that H(x, r) = h(x) · r for (x, r) ∈ Σ × [1, ). For ε > 0 small enough, there exists an almost complex structure J on Wc such that (ε · H, J) is Floer data for ε·h and such that the pair (ε·H, ε·JXH)∞is Morse-Smale [1, Chapter 10] (note that we are using different conventions from [1], namely we define XH as the vector field that satisfies dH = dλ(XH, ·)). Moreover,

CF∗(ε · H, J) = CM∗+n(ε · H, ε · JXH), where CM∗ stands for the Morse complex. Consequently,

dim HFk(εh) = dim HMk+n(ε · H, ε · JXH), where HM∗ is the Morse homology.

18 Step 2 (Extending the Liouville domain). Let

L := max{ε · H(p) | p ∈ Crit H}, ` := min{ε · H(p) | p ∈ Crit H}.

In this step, we prove that there exists R > 0 such that the extension

W(R) := Wc \ (Σ × (R, )) of the Liouville domain W has the following property: the critical values ∞ of ε · H|∂W(R) do not lie in the interval [`, L]. Since 0 is a regular value of ε·h and since the domain of ε·h, Σ, is compact, there exists δ > 0 such that ε · h has no critical values in (−δ, δ). A point p ∈ Σ is a critical point of ε · h if, and only if, (p, R) ∈ ∂W(R) = Σ × {R} is a critical point of ε · H|∂W(R) . Hence ε · H|∂W(R) has no critical values in (−δR, δR). If R is big enough, then [`, L]⊂ (−δR, δR). See Figure 2, Step 2 on page 26. Step 3 (Constructing the double manifold). In this step, we construct a closed manifold (not necessarily symplectic), M, by gluing two copies of W(R) along the boundary. See Figure 2, Step 3 on page 26. Denote by WA and WB those two copies. Explicitly,

M := (WA t (Σ × R) t WB)/ ∼, where ∼ stands for the following identifications   r  WA ⊃ Σ × (0, R] → Σ × :(x, r) 7→ x, log , R R   r  WB ⊃ Σ × (0, R] → Σ × :(x, r) 7→ x, − log . R R Let F : M → R be a function that is obtained by smoothing out the function on M equal to ε · H on both WA and WB. A formal definition of F follows. Let χ : R → R<0 be a smooth concave function such that χ(s) = s for R−1  R−1  s 6 log R , χ(s) = −s for s > − log R , and such that χ has a unique maximum at s = 0. The function χ can be seen as a smoothening of the function s 7→ − |s| by a compactly supported perturbation. Denote by F : M → R the function defined by

ε · H(p) for p ∈ W(R − 1) ⊂ WA, F(p) := R · eχ(s) · ε · h(y) for p = (y, s) ∈ Σ × ,  R  ε · H(p) for p ∈ W(R − 1) ⊂ WB. Step 4 (Truncated double manifold). There are three types of critical points of the function F : the critical points of H in WA, the critical points of

19 R−1  R−1  log R − log R R − 1 R R − 1

WA WB

Figure 1: An illustration of the functions χ (left) and F (right). The function F is graphed along M in the case of a constant contact Hamiltonian h.

H in WB, and the critical points on Σ × {0} ⊂ Σ × R that correspond to the critical points of h : Σ → R . As explained in Step 2, for R > 0 big enough, the critical points of the third type have values outside the interval [`, L]. In fact, one can choose R big enough so that the critical points of the third type have values outside an interval [−K, K] ⊃ [`, L] where K satisfies

−1 −1 F (−K),F (K) ⊂ Σ × R ⊂ M. Then, −K and K are regular values of F and the only critical points of F in F−1([−K, K]) are the ones of type 1 and type 2. Note that these critical points are nondegenerate. Hence F−1([−K, K]) is a manifold with boundary whose boundary components are regular level sets of F, and F is a Morse −1 −1 function on F ([−K, K]). Denote MT := F ([−K, K]) (see Figure 2, Step 4 on page 26). Step 5 (No crossing). Let g be a Riemannian metric on M such that g = dλ(·,J·) on both W(R/2) ⊂ WA and W(R/2) ⊂ WB. Let X = ∇F, where ∇ is the gradient with respect to g. In particular, if R > 2, X is equal to ε·JXH on both W(R/2) ⊂ WA and W(R/2) ⊂ WB. In this step, we prove that for R large enough, there are no integral curves of X that connect two critical points of F in MT and intersect the submanifold Σ × {0} ⊂ Σ × R ⊂ M. This eliminates the integral curves of X that connect a critical point of F in WA with a critical point of F in WB, and also the integral curves of X that connect two critical points of F in WA (or WB) but at some point leave WA (WB, respectively).

20 Assume there exists an integral curve γ : R → M of X such that γ connects two critical points of F in MT and such that it is not contained in one of the regions WA and WB. Without loss of generality assume lim γ(t) ∈ WA. Then, for R > 2, there exists an interval [a, b] ⊂ such t→− R that γ(a) ∈ Σ × {1} ⊂ WA and γ(b) ∈ Σ × {R/2} ⊂ WA. Since ∞ L − ` F( lim γ(t)) − F( lim γ(t)) > t→+ t→− + = dF∞(γ˙ (t))dt ∞ − ∞ Z + = ∞ g (∇F(γ(t)), γ˙ (t)) dt − ∞ Z + = ∞ g (X(γ(t)),X(γ(t))) dt − ∞ Z b > ∞g(ε · JXH(γ(t)), ε · JXH(γ(t)))dt a Z b 2 = ε · g(XH(γ(t)),XH(γ(t)))dt a Z   2 2 R ε · (b − a) · min kXH(p)k | p ∈ Σ × 1, , > g 2 we have L − ` b − a 6 . 2 2  R  ε · min kXH(p)kg | p ∈ Σ × 1, 2 Since h is a strict contact Hamiltonian (i.e. it generates an isotopy that preserves the contact form), the vector XH(x, r) is independent of r. Hence   2 R 2 min kXH(p)kg | p ∈ Σ × 1, = min kXH(p)kg , 2 p∈Σ×{1}

+ and, consequently, b − a < C, where C ∈ R is a constant that does not  R  depend on R. For r ∈ 1, 2 , the vector field ∂r is orthogonal (with respect

21 √1 to g) to Σ × {r} and k∂rkg = r . The latter follows because

2 k∂rkg = dλ(∂r, J∂r)

= dr ∧ α(∂r, J∂r) + r · dα(∂r, J∂r)

= dr(∂r) · α(J∂r) − α(∂r) · dr(J∂r)

= α(J∂r) 1 = · (λ ◦ J)(∂ ) r r 1 = · (−dr ◦ J ◦ J)(∂ ) r r 1 = . r The orthogonal projection (with respect to g) of the vector γ˙ (t) to the 1-dimensional vector space

{s · ∂r | s ∈ R}

d   is equal to dt r(γ(t)) · ∂r. Hence, the Pythagorean theorem implies

  2 2 d 2 kγ˙ (t)k r(γ(t)) · k∂rk g > dt g   2 d 1 = r(γ(t)) · dt r(γ(t)) 2 d  p  = 2 r(γ(t)) . dt

Therefore,

b 2 L − ` > kγ˙ (t)kg dt Za b   2 d p > 2 r(γ(t)) dt a dt Z 2 1  b d    · 2pr(γ(t)) dt > b − a dt Za 1  2 · 2pr(γ(b)) − 2pr(γ(a)) > b − a 1 √ 2 · 2R − 2 . > C

22 This, however, cannot be possible for R big enough. In other words, for R big enough, there are no integral curves of X that cross Σ×{0} ⊂ Σ×R ⊂ M and connect critical points of F in MT . In fact, we proved that such integral curves are all contained in either W(R/2) ⊂ WA or W(R/2) ⊂ WB. As a consequence, F and X satisfy the Smale condition in MT , and

CM∗(F, X) = CM∗(ε · H, ε · JXH) ⊕ CM∗(ε · H, ε · JXH). Hence 2 dim HFk(ε · h) = dim HMk+n(F, X). + Denote by ∂ MT the part of the boundary ∂MT on which the vector field − + − X points inwards. Similarly, denote ∂ MT := ∂MT \ ∂ MT , i.e. ∂ MT is the part of the boundary ∂MT on which the vector field X points outwards. + − Alternatively, ∂ MT and ∂ MT can be defined by + −1 − −1 ∂ MT := F (−K), ∂ MT := F (K).

Since the boundary components of the manifold MT are regular level sets of F, the Morse homology of F : MT → R can be expressed as the singular homology of a pair (see, for instance, [16, Theorem 3.9]) ∼ − HM∗(F, X) = H∗(MT , ∂ MT ).

Step 6 (Mayer-Vietoris long exact sequence). Denote VA := WA ∩ MT , − − VB := WB ∩ MT ,UA := ∂ MT ∩ WA, and UB := ∂ MT ∩ WB. The relative form of the Mayer-Vietoris long exact sequence implies that there exists a long exact sequence . .

Hk(VA ∩ VB,UA ∩ UB) Hk−1(VA ∩ VB,UA ∩ UB)

. Hk(VA,UA) ⊕ Hk(VB,UB) .

− Hk(MT , ∂ MT )

F| The function VA∩VB is a Morse function that has no critical points. Hence VA ∩ VB is a trivial cobordism, i.e. VA ∩ VB ≈ (UA ∩ UB) × [0, 1]. In particular, Hk(VA ∩ VB,UA ∩ UB) = 0

23 for all k ∈ Z. Consequently, − ∼ Hk(MT , ∂ MT ) = Hk(VA,UA) ⊕ Hk(VB,UB) for all k ∈ Z. Since the pairs (VA,UA) and (VB,UB) are homeomorphic, the following holds

− (∀k ∈ Z) dim Hk(MT , ∂ MT ) = 2 dim Hk(VA,UA).

Therefore,

(∀k ∈ Z) dim HFk(εh) = dim Hk+n(VA,UA).

See Figure 2, Step 6 on page 26. Step 7. In this step, we prove that the pair (VA,UA) is homeomorphic to + the pair (W, Σ ). The boundary of VA consists of parts of the submanifolds Σ×{0} ⊂ Σ×R ⊂ M, F−1(−K), and F−1(K). These parts belong to Σ×R ⊂ M. Recall that, on Σ × R ⊂ M, the function F is given by (y, s) 7→ R · eχ(s) · ε · h(y).

The function F does not change the sign along the curve {x} × R, for x ∈ Σ. Therefore, there does not exist x ∈ Σ such that {x} × R intersects both of the sets −1 −1 (∂VA) ∩ F (−K), (∂VA) ∩ F (K). The derivative d Reχ(s)εh(y) = εReχ(s)χ0(s)h(y) ds is positive for positive h(y) and negative for negative h(y) on Σ × (− , 0]. Therefore, the value of F on the curve {x} × (− , 0] belongs to (−K, K) if   ∞ −1 (x, 0) ∈ int VA ∩ (Σ × {0}) = (Σ ×∞{0}) ∩ F (−K, K).

Consequently, the curve {x} × R, x ∈ Σ, cannot intersect two of the sets

−1 −1 (∂VA) ∩ F (−K), (∂VA) ∩ (Σ × {0}), (∂VA) ∩ F (K) except at the points where those two sets intersect themselves. Since d Reχ(s)εh(y) = εReχ(s)χ0(s)h(y) 6= 0 ds

24 for s 6= 0, and since the curve {x} × R cannot intersect different parts of ∂VA (except at the points where they intersect) for all x ∈ Σ, the boundary of VA can be seen as the graph in Σ×(− , 0] of a continuous (in fact, a piecewise- smooth) function. Lemma 5.2 on page 28 below states that in this case −1 VA = WA ∩ F ([−K, K]) is homeomorphic∞ to WA via a homeomorphism −1 −1 that sends UA = F (K) ∩ WA to F ([K, + )) ∩ ∂WA. The continuous function f to which Lemma 5.2 is applied is given by

−1 ∞ f(y) := min Fy (K), 0 ,  where Fy : R → R : s 7→ F(y, s). The homeomorphism furnished by −1 Lemma 5.2 maps indeed UA to F ([L, + )) ∩ ∂WA for the following −1 reason. For all (y, f(y)) ∈ UA, either f(y) = 0 or 0 > f(y) = Fy (K) (which implies F(y, 0) > K). ∞ F| F−1([−K, K]), Since there are no critical points of ∂WA in the pair

−1  WA, ∂WA ∩ F ([K, + )) is homeomorphic (in fact, diffeomorphic) to the pair (WA, ∂WA ∩ {F 0}). ∞ > Consequently, the pair (VA,UA) is homeomorphic to the pair (W, {h > 0}). Hence ∼ H∗(VA,UA) = H∗(W, {h > 0}) ∼ = H∗(W, {h > 0}) + = H∗(W, Σ ).

See Figure 2, Step 7 on page 26. The following lemma was used at the beginning of the proof of Propo- sition 4.2. It allowed passing from Floer to Morse homology. Lemma 5.1. Let W be a Liouville domain with the boundary Σ := ∂W, and let h : Σ → R be a strict contact Hamiltonian (i.e. dh(Rα) = 0 where Rα is the Reeb vector field). Assume that the contact Hamiltonian c · h has no 1-periodic orbits for c ∈ [a, b]. Then, the groups HF∗(a · h) and HF∗(b · h) are isomorphic. Proof. We prove here that there exists a smooth family Hs : Σ × [1, + ) → R, s ∈ [0, 1] of Hamiltonians without 1-periodic orbits such that 0 ∞ • H (x, r) = a · h(x) · r, for (x, r) ∈ Σ × [1, + ),

• Hs(x, r) = a · h(x) · r, for s ∈ [0, 1] and for r in a neighbourhood of 1, ∞ • H1(x, r) = b · h(x) · r, for r big enough.

25 Figure 2: An illustration of the proof of Proposition 4.2

26 Informally, this means that one can modify the slope smoothly on the cylindrical end from a·h to b·h without creating any additional 1-periodic orbits. Hence, it is possible to find Floer data (Ha,Ja) and (Hb,Jb) that compute HF∗(a · h) and HF∗(b · h), respectively, and such that the chain a a b b complexes CF∗(H ,J ) and CF∗(H ,J ) are identical. As opposed to the case where h is constant, the proof is not direct. Denote by X(Σ) the space of smooth vector fields on Σ endowed with 1 X the C topology, and denote by φt : Σ → Σ the flow of a vector field X ∈ X(Σ). Denote by Yh the vector field of the contact isotopy furnished by the contact Hamiltonian h. The map

X X(Σ) × Σ → Σ :(X, p) 7→ φt (p) is continuous for all t ∈ R . This (together with the time-1 map of the flow of c · Yh not having any fixed points for c ∈ [a, b]) implies that there exists an open neighbourhood U ⊂ X(Σ) of

{c · Yh | c ∈ [a, b]}

Z such that φ1 has no fixed points for all Z ∈ U. + Let δ > 0, and let µ : R → R be a smooth function such that

• µ(r) = a · r, for r < 2,

• µ(r) = b · r, for r sufficiently large,

µ(r) + • r ∈ [a, b], for all r ∈ R ,

µ0(r) − µ(r) < δ, r ∈ + . • r for all R

+ (Such a function µ can be constructed in the following way. Let κ : R → R be a compactly supported smooth function such that supp κ ⊂ (2, + ), δ + such that κ(r) 6 r for all r ∈ R , and such that κ(r)dr = b − a. The R + conditions on κ are not contradicting each other because δ dr = + , R 2 r ∞ and therefore, κ(r)dr can be chosen arbitrary large without∞ violating the R r R condition κ(r) δ . The function µ(r) := a · r + r · κ(s)ds satisfies the R6 r 0 ∞ conditions above. See Figure 3 for graphs of κ and µ.) R We will show that when δ > 0 is sufficiently small, the homotopy

s H : Σ × [1, + ) → R :(x, r) 7→ ((1 − s) · a · r + s · µ(r)) · h(x) satisfies the conditions from the beginning of this proof. The only non- ∞ trivial condition to check is that Hs has no 1-periodic orbits for all s ∈ [0, 1].

27 2 2

0 2 4 0 2 4

Figure 3: An illustration of the functions κ (left) and µ (right).

The vector field of the Hamiltonian Hs is equal to     µ(r) h µ(r) 0 X s (x, r) = (1 − s) · a + s · · Y (x) + s · − µ (r) · h(x) · R(x), H r r where R = Rα is the Reeb vector field on Σ (the computation used dh(R) = 0). In particular, the flow of Hs preserves the submanifolds Σ × {r}, r ∈ [1, + ). Therefore, it is enough to prove that the restriction of the flow of Hs to Σ × {r} has no 1-periodic orbits for all r ∈ [1, + ) and all s ∈ [0, 1]. µ(r) By the∞ assumptions, (1 − s) · a + s · r ∈ [a, b] for all s ∈ [0, 1] and r ∈ [1, + ). Hence, for δ > 0 sufficiently small, XHs (·∞, r) ∈ U for s ∈ [0, 1] and r ∈ [1, + ). Consequently, for δ > 0 sufficiently small, the flow of XHs has no 1-periodic∞ orbits. The next∞ lemma is a topological fact that was used in the final step in the proof of Proposition 4.2 above. + Lemma 5.2. Let X be a compact topological space, let ε ∈ R be a positive real number, and let f : X → [0, + ) be a continuous function. Then, there exists a homeomorphism ∞ ψ : X × [0, + ) → {(x, r) ∈ X × [0, + ) | r > f(x)} such that ψ(x, 0) = (x, f(x)) for all x ∈ X, and such that ψ(x, r) = (x, r) if ∞ ∞ r > f(x) + ε. Proof. Denote Y := {(x, r) ∈ X × [0, + ) | r > f(x)} . A homeomorphism satisfying the conditions of the lemma can be con- structed explicitly as follows. Let ∞   x, f(x) + r · ε for r ∈ [0, f(x) + ε] ψ(x, r) := f(x)+ε  (x, r) for r ∈ [f(x) + ε, + ).

28 ∞ The function ψ : X × [0, + ) → Y is well defined and continuous. The function  ∞  y, (s − f(y)) · f(y)+ε for s ∈ [f(y), f(y) + ε] (y, s) 7→ ε  (y, s) for s ∈ [f(y) + ε, + ).

Y → X × [0, + ) is a well-defined continuous function , and it∞ is inverse to ψ. Hence ψ is a homeomorphism. ∞ References

[1] Michèle Audin and Mihai Damian. Morse theory and Floer homology. Springer, 2014. [2] Serguei Barannikov. “The framed Morse complex and its invariants”. In: American Mathematical Society Translations. 2nd ser. (1994). [3] Kilian Barth, Hansjörg Geiges, and Kai Zehmisch. “The diffeomor- phism type of symplectic fillings”. In: Journal of 17.4 (2019), pp. 929–971. [4] Paul Biran and Emmanuel Giroux. “Symplectic mapping classes and fillings”. unpublished manuscript. 2005. [5] Jonathan M. Bloom. “The combinatorics of Morse theory with boundary”. In: arXiv preprint arXiv:1212.6467 (2012). [6] Maciej Borodzik, András Némethi, and Andrew Ranicki. “Morse the- ory for manifolds with boundary”. In: Algebraic & Geometric Topology 16.2 (2016), pp. 971–1023. [7] Dietrich Braess. “Morse-theorie für berandete Mannigfaltigkeiten”. In: Mathematische Annalen 208.2 (1974), pp. 133–148. [8] Roger Casals et al. “Symplectomorphisms of exotic discs”. In: Journal de l’École polytechnique - Mathématiques 5 (2018), pp. 289–316. [9] River Chiang, Fan Ding, and Otto van Koert. “Non-fillable in- variant contact structures on principal circle bundles and left- handed twists”. In: International Journal of Mathematics 27.03 (2016), p. 1650024. [10] River Chiang, Fan Ding, and Otto van Koert. “Open books for Boothby-Wang bundles, fibered Dehn twists and the mean Euler characteristic”. In: Journal of Symplectic Geometry 12.2 (2014), pp. 379– 426.

29 [11] Octav Cornea and Andrew Ranicki. “Rigidity and gluing for Morse and Novikov complexes”. In: Journal of the European Mathematical Society 5.4 (2003), pp. 343–394. [12] Alexander Fauck. “Rabinowitz-Floer homology on Brieskorn mani- folds”. PhD thesis. Humboldt-Universität zu Berlin, Mathematisch- Naturwissenschaftliche Fakultät, 2016. [13] Mikhael Gromov. “Pseudo holomorphic curves in symplectic mani- folds”. In: Inventiones mathematicae 82.2 (1985), pp. 307–347. [14] Bogusłav Hajduk. “Minimal m-functions”. In: Fundamenta Mathemat- icae 111.3 (1981), pp. 179–200. [15] Allen Hatcher. Algebraic Topology. Cambridge University Press, 2002. [16] Michael Hutchings. “Lecture notes on Morse homology (with an eye towards Floer theory and pseudoholomorphic curves)”. In: Lecture Notes. UC Berkeley (2002). [17] Andrzej W. Jankowski and Ryszard L. Rubinsztejn. “Functions with non-degenerate critical points on manifolds with boundary”. In: Commentationes Mathematicae 16.1 (1972). [18] Peter B. Kronheimer and Tomasz Mrowka. Monopoles and three- manifolds. Vol. 10. Cambridge University Press Cambridge, 2007. [19] Myeonggi Kwon and Otto van Koert. “Brieskorn manifolds in contact topology”. In: Bulletin of the London Mathematical Society 48.2 (2016), pp. 173–241. [20] François Laudenbach. “A Morse complex on manifolds with bound- ary”. In: Geometriae Dedicata 153.1 (2011), pp. 47–57. [21] Eugene Lerman. “Contact cuts”. In: Israel Journal of Mathematics 124.1 (2001), pp. 77–92. [22] Will J. Merry and Igor Uljarević. “Maximum principles in symplectic homology”. In: Israel Journal of Mathematics 229.1 (2019), pp. 39–65. [23] Marston Morse and George B. van Schaack. “The critical point theory under general boundary conditions”. In: Annals of Mathematics (1934), pp. 545–571. [24] Petr E. Pushkar. “Morse theory on manifolds with boundary I. Strong Morse function, cellular structures and algebraic simplification of cellular differential”. In: arXiv preprint arXiv:1912.06437 (2019). [25] Alexander Ritter. “Circle actions, quantum cohomology, and the Fukaya category of Fano toric varieties”. In: Geometry & Topology 20.4 (2016), pp. 1941–2052.

30 [26] Paul Seidel. “Exotic iterated Dehn twists”. In: Algebraic & Geometric Topology 14.6 (2015), pp. 3305–3324. [27] Paul Seidel. “Floer homology and the symplectic isotopy problem”. PhD thesis. University of Oxford, 1997. [28] Paul Seidel. “Lectures on four-dimensional Dehn twists”. In: Sym- plectic 4-manifolds and algebraic surfaces. Springer, 2008, pp. 231–267. [29] Vsevolod Shevchishin and Gleb Smirnov. “Elliptic diffeomorphisms of symplectic 4-manifolds”. In: arXiv preprint arXiv:1708.01518 (2017). [30] Dmitry Tonkonog. “Commuting symplectomorphisms and Dehn twists in divisors”. In: Geometry & Topology 19.6 (2016), pp. 3345– 3403. [31] Igor Uljarević. “Floer homology of automorphisms of Liouville do- mains”. In: Journal of Symplectic Geometry 15.3 (2017), pp. 861–903. [32] Igor Uljarević. “Viterbo’s transfer morphism for symplectomor- phisms”. In: Journal of Topology and Analysis 11.01 (2019), pp. 149– 180.

31