<<

The Pennsylvania State University

The Graduate School

Department of Mathematics

FLOER FOR ALMOST HAMILTONIAN

ISOTOPIES

A Thesis in Mathematics by Christopher L. Saunders c 2005 Christopher L. Saunders

Submitted in Partial Fulfillment of the Requirements for the Degree of

Doctor of Philosophy

May 2005 The thesis of Christopher L. Saunders has been reviewed and approved* by the following:

Augustin Banyaga Professor of Mathematics Thesis Advisor Chair of the Committe

Ping Xu Professor of Mathematics

Sergei Tabachnikov Professor of Mathematics

Abhay Ashtekar Professor of Physics

Nigel Higson Professor of Mathematics Department Chair

* Signatures are on file at the Graduate School. Abstract

Floer homology is defined for a closed symplectic which satis-

fies certain technical conditions, and is denoted HF∗(M, ω). Under these conditions, Seidel has introduced a homomorphism from π1(Ham(M)) to a quotient of the of automorphisms of HF∗(M, ω).

The main goal of this thesis is to prove that if two loops of Hamiltonian diffeomorphisms are homotopic through arbitrary loops of diffeomorphisms, then the image of (the classes of) these loops agree under this homomorphism.

This can be interpreted as further evidence for what has been called the “topological rigidity of Hamiltonian loops”. This phenomonon is a collec- tion of results which indicate that the properties of a loop of Hamiltonian diffeomorphisms are more tied to the class of the loop inside of π1(Diff(M)) than one might guess. For example, on a compact manifold M, the path that a point follows under a loop of Hamiltonian diffeomorphisms of M is always contractible. In [14], further results of this type were proved using Seidel’s homomorphism.

To prove this result, we extend the domain of Seidel’s homomorphism to the group of all loops of diffeomorphism of M based at id which are ho- motopic to a Hamiltonian loop. Such a loop will be called almost Hamil- tonian. This requires that we extend the space from which we choose to

HF∗(M, ω).

The groups are defined by choosing a (generic) pair (H,J) in C∞(M × S1) × J (M, ω, S1). Here, J (Mω, S1) means the set of all smooth loops of ω-compatible almost complex structures. Such pairs are

iii called regular pairs. The Floer homology constructed using (H,J) is de- 0 0 noted HF∗(M, ω, H, J). If (H ,J ) is another such regular pair, then there 0 0 is a natural isomorphism between HF∗(M, ω, H, J) and HF∗(M, ω, H ,J ). These isomorphisms are funtorial with respect to composition, so we can speak of HF∗(M, ω) without mention of the pair (H,J) used to define it. These isomorphisms are called the continuation isomorphisms, and are gen- erally denoted by Φ.

Denote by G the group of all smooth loops in Ham(M) at id. Seidel introduces an extension of G, denoted Ge. That is, there is an exact sequence of topological groups 1 → Γ → Ge → G, with Γ discrete. Each g in G defines a new pair (Hg,J g), such that ifg ˜ g g is a lift of g to Ge, theng ˜ induces an isomorphism from HF∗(M, ω, H ,J ) to HF∗(M, ω, H, J). By an abuse of notation, we will still call this iso- morphismg ˜. By precomposing with the continuation isomorphisms, we obtain an automorphism of HF∗(M, ω, H, J). That is,g ˜ ∈ Ge defines an automorphism of HF∗(M, ω, H, J) by

Φ g g g˜ HF∗(M, ω, H, J) / HF∗(M, ω, H ,J ) / HF∗(M, ω, H, J).

Seidel then shows that this automorphism of HF∗(M, ω, H, J) commutes with the continuation isomorphism to any other regular pair, so thatg ˜ ∈ Ge induces an automorphism of HF∗(M, ω) independent of the regular pair.

Finally, Seidel proves that ifg ˜0 andg ˜1 define the same element in π0(Ge), then they induce the same automorphism of HF∗(M, ω). We outline the intuitive argument. Suppose g0, g1 ∈ G have liftsg ˜0, g˜1 ∈ Ge. This means that there is a path gs in G connecting g0 to g1, and a pathg ˜s in Ge connecting

iv g˜0 tog ˜1. Choose a regular pair (H,J), and construct HF∗(M, ω, H, J). Consider the following diagram.

g0 g0 HF∗(M, ω, H ,J ) : I uu II uu II uu II uu II Φ uu II g˜0 uu HF (M, ω, Hgs ,J gs ) II uu ∗ II uu j4 TTT II uu Φ jjjj TTT g˜s II uu jjj TTTT II uu jjjj TTT II uu jjjj TTT* I$ HF∗(M, ω, H, J) HF∗(M, ω, H, J) TT j4 TTTT jjjj TTTT jjjj Φ TTT jjj g˜1 TTT* jjjj g1 g1 HF∗(M, ω, H ,J )

Intuitively, by usingg ˜s to connectg ˜0 andg ˜1, the middle of the di-

g0 g0 agram can be filled in, and we can transform HF∗(M, ω, H ,J ) into

g1 g1 HF∗(M, ω, H ,J ).

We are interested in loops in Diff(M) at id which are not Hamiltonian, but which are homotopic to a Hamiltonian loop. As given, the diagram above cannot be extended to such loops: if g is not a Hamiltonian loop, Hg does not make sense.

To overcome this difficulty, notice that the construction of the Floer H homology groups HF∗(M, ω, H, J) only depends on θ , the Hamiltonian isotopy generated by H, and the almost complex structure J, so we are H justified in writing HF∗(M, ω, θ ,J).

We want to rephrase the above diagram in terms of isotopies, so in section 2.1, we calculate the isotopy generated by the function Hg. We find Hg −1 H that θt = gt θt . So, for θ any isotopy of M and g any loop at id in −1 Diff(M), let g ∗ θ be the isotopy of M given by (g ∗ θ)t := gt θt. In this

v notation, we find that θHg = g ∗ θH .

Then we can rewrite the above diagram as the following.

H g0 HF∗(M, ω, g0 ∗ θ ,J ) s9 KK ss KK ss KK ss KK ss KK Φ ss H g KK g˜0 ssHF (M, ω, g ∗ θ ,J s ) KK ss ∗ s KK ss ii4 UUU KK ss Φ iii UUU g˜s KK ss iiii UUUU KK ss iii UUU KK ss iiii UUU* K% H H HF∗(M, ω, θ ,J) HF∗(M, ω, θ ,J) UU i4 UUUU iiii UUUU iiii Φ UUU iii g˜1 UUU* iiii H g1 HF∗(M, ω, g1 ∗ θ ,J )

While Hg does not make sense if g is not Hamiltonian, g ∗θH does make sense. If we can construct a Floer homology using isotopies of the form g∗θ, where g is an almost Hamiltonian loop in Diff(M) and θ is a Hamiltonian isotopy, then we can fill in the middle of this diagram, and use it to prove the main theorem.

Notice that if g is an almost Hamiltonian loop, and θ is a Hamiltonian isotopy, then g∗θ is homotopic, relative endpoints, to a Hamiltonian isotopy. The first step in proving the main result is to extend the space used to define HF∗(M, ω) to include generic pairs (ψ, J), where ψ is an isotopy of M which is homotopic, relative endpoints, to a Hamiltonian isotopy, and J is a smooth loop of almost complex structures with a suitably adapted compatibility condition. (The details associated with the almost complex structures become rather tech- nical.) This construction culminates in section 4, but the fact that it is

vi well defined uses all of the material developed in the previous sections. We prove that these groups are naturally independent of the choice of (ψ, J) used to define them (theorem 4.16). This effectively extends our space of choices used in defining HF∗(M, ω).

Using the expanded domain of definition for HF∗(M, ω), we can extend the domain of Seidel’s homomorphism to the group of almost Hamiltonian loops. By examining what happens when we homotope these loops, we prove the main result. This is given by theorem 5.3.

This thesis is organized as follows:

The first section gives a quick review of the construction of the Floer homology groups for a weakly monotone , emphasizing the points that will be of concern later. We discuss the monotonicity con- dition, and show why it is helpful. (The condition of being monotone is a technical condition that is no longer needed to define Floer homology. It is the assumption that the founder of the theory, Andreas Floer, used in his expositions. The assumption guarantees that J-holomorphic spheres are well behaved, and simplifies many things. For weakly monotone symplectic , J-holomorphic curves are well behaved for generic J. We suspect the results hold without the assumption.) We then outline the definition of Seidel’s homomorphism.

The next section rephrases Seidel’s homomorphism in terms of isotopies, as is described above. Here, we calculate θHg , the Hamiltonian isotopy gen- erated by Hg. This section includes the proof of the fact that the Hamilto- nian function H is not so important as θH , the isotopy it generates. More

vii precisely, we show that if H and H0 generate the same isotopy, then the 0 groups HF∗(M, ω, H, J) and HF∗(M, ω, H ,J) are actually equal by con- struction. Notice that this truly means equals, and not just isomorphic.

The third section contains the technical tools needed throughout the re- mainder of the thesis. In it, we outline the basic definitions and notions of Hamiltonian and almost Hamiltonian loops as described above. We also in- troduce the space that will be used as the expanded domain of definition for

HF∗(M, ω). Because of the difficulties with ω-compatibility requirements on the almost complex structures, this domain will no longer be a cross product, but should be thought of as a bundle, with base given by the al- most Hamiltonian isotopies, and the fiber over such an isotopy consisting of 1-periodic time dependent almost complex structures with a compatibility requirement that depends on the particular isotopy.

The next section gives the definition of HF∗(M, ω, ψ, J) for a generic almost Hamiltonian isotopy. As in the original construction, the definitions are not as hard as proving that the objects are well-defined. In particular, a boundary operator on a is defined by counting certain objects. The bulk of the difficulty in the construction of the Floer homology groups is in showing that the spaces in question are actually finite. (In fact, not all choices (H,J) are allowed, and we are only allowed to choose from a dense subset of all choices.) To show that our boundary operator is well- defined, we reduce to the Hamiltonian case.

The next section contains the main result. We extend the domain of Seidel’s homomorphism to the group of almost Hamiltonian loops using the expanded domain of definition of HF∗(M, ω). Finally, we show that if two loops are in the same homotopy class, they have the same image under this

viii homomorphism.

The final section consists of remarks about what has been shown, and what more might be said.

ix Contents

Acknowledgements xii

1 Review of Floer Homology 1 1.1 Symplectic Preliminaries ...... 2 1.2 Floer Homology ...... 8 1.2.1 Defining LgM ...... 9 1.2.2 The Conley-Zehnder Index and the Chain Complex . 11 1.2.3 on LgM and the Boundary Operator . 13 1.2.4 Continuation Isomorphisms ...... 20 1.3 Seidel’s Action ...... 23 1.3.1 The Extension Ge ...... 23

1.3.2 The Mapσ ˜ : Ge → Aut(HF∗(M, ω)) ...... 25

2 Rephrasing σ˜ in terms of Isotopies 32 2.1 Calculating θHg ...... 33 2.2 Emphasizing Isotopies ...... 35 2.3 Rephrasingσ ˜ ...... 37

3 Almost Hamiltonian Loops and Isotopies 39 3.1 Almost Hamiltonian Loops of Diffeomorphisms ...... 39 3.2 Almost Hamiltonian Isotopies ...... 42 3.3 Difficulties with Almost Complex Structures ...... 44 3.4 The Maslov Index ...... 49 3.5 LgM Revisited ...... 55

4 Defining HF∗(M, ω, ψ, J) 59 4.1 Defining Pg(ψ)...... 59 4.2 The Chain Complex and Boundary Operator ...... 66

x 4.3 The New HF∗(M, ω)...... 74

5 Extending the Action 81

6 Concluding Remarks 86

References 88

xi Acknowledgements

It’s been quite an adventure, getting to this point. There have been highs and lows, and I am quite sure that the help of many people was in- strumental in my completion of this thesis. My advisor, Professor Augusting Banyaga, inspired me to work in this particular area, and his guidance was invaluable. Many thanks also to my colleague Shantanu Dave, for listening to me ramble on about Floer homology. He understood that explaining mathematics helps tremendously in understanding mathematics. I’d like to thank my wife, Michelle, whose sacrifices don’t go unnoticed. Finally, I thank my children, Brooke and Bailey; with them, I always begin my day smiling.

xii 1 Review of Floer Homology

Floer homology was originally introduced in an attempt to prove a version of the Arnold conjecture. This conjecture states that, in the non-degenerate case, the number of fixed points of a Hamiltonian symplectomorphism of a symplectic manifold is bounded from below by the sum of the Betti numbers of the manifold. (Non-degeneracy means that the graph of the diffeomor- phism inside M × M intersects the diagonal transversally.) The idea of his proof is analogous to proving the Morse inequalities. First, the fixed points are indexed in some manner. The chain complex

CFk(M, ω, H) is defined as all formal sums of elements of the fixed point set of index k with coefficients in Z. A boundary operator is defined on this complex. This boundary operator is defined by counting the number of certain cylinders in M. That we can count these cylinders follows from that fact that the cylin- ders satisfy the same basic analytic properties as Gromov’s J-holomorphic curves, introduced in [12]. The main point is that the differential equation equation defining these curves gives an elliptic (and therefore Fredholm) operator on suitable Banach spaces. Floer proves that ∂2 = 0, and that the homology of this complex recovers the homology of the underlying manifold. Arnold’s conjecture follows. Floer’s construction is for monotone symplectic manifolds. This is a stronger assumption than the assumption of an integral class symplectic form. In [13] and [21], this proof was extended to a larger class of manifolds. Because we intend to extend Seidel’s action, we use the his version of Floer homology, which is only a slightly modified form of that introduced in [13]. For simplicity, we will use coefficients in Z2, to avoid difficulties with orientation.

1 1.1 Symplectic Preliminaries

A symplectic manifold is a manifold M 2n equipped with a 2-form ω ∈ Ω2(M) such that ω is closed and non-degenerate. Any ω ∈ Ω2(M) induces ∗ a bundle map from TM to T M by X 7→ iX ω. Non-degeneracy means that this bundle map is an isomorphism. This is equivalent to the requirement that ωn is a volume form on M. By Diff(M), we mean the set of all diffeomorphisms of M. This set forms a group with the product given by composition of maps. Each element of Diff(M) is a map ψ : M → M, and so Diff(M) ⊂ C∞(M,M). We topologize Diff(M) as a subspace of C∞(M,M) equipped with the C∞ topology. For (M, ω) a symplectic manifold, let

Symp(M, ω) = {ψ ∈ Diff(M) | ψ∗ω = ω}.

Such diffeomorphisms will be called symplectic diffeomorphisms or sym- plectomorphisms. It is clear that Symp(M, ω) is a subgroup of Diff(M), and we topologize Symp(M, ω) as a subspace of Diff(M).

Now, consider Symp0(M, ω), the component of the identity inside the group Symp(M, ω). If we letω ˜ : TM → T ∗M denote the bundle iso- morphism induced by ω, then each ψ ∈ Symp0(M, ω) corresponds to the endpoint of the flow of a time dependent “closed” vector field, i.e., a vector −1 field of the formω ˜ (αt), where each αt is a closed 1-form. By considering “exact” vector fields, we obtain an important normal subgroup, called Ham(M). More precisely, let H ∈ C∞(M × R), and let −1 XHt =ω ˜ (dHt). (This means that ω(XHt , ·) = dHt(·).) Integration of this H H vector field gives a path in Diff(M), denoted θ = (θt ), defined by the equations

2 dθH t (p) = X (θH (p)), θH = id. dt Ht t 0 The group of Hamiltonian diffeomorphisms of M consists of the set of all time-1 diffeomorphisms arising as in the above construction, and is denoted

Ham(M). In fact, Ham(M) is a normal subgroup of Symp0(M, ω) (see [2]).

We topologize Ham(M) as a subspace of Symp0(M, ω).

Conjecture (Arnold) In the non-degenerate case, the number of fixed points for any θ ∈ Ham(M) is bounded from below by the sum of the Betti numbers of M.

This was first proved for Tn by Conley and Zehnder in [6]. Later, Floer introduced the Floer homology groups HF∗(M, ω) to prove the conjecture for certain types of symplectic manifolds, the so called monotone symplec- tic manifolds. This is essentially given in [11], using ideas formulated in [8], [9] and [10]. In [13], the definition of Floer homology was extended to weakly mono- tone symplectic manifolds. To describe these sub-classes, we need the no- tions of an almost complex structure on M, and Chern classes.

Definition 1.1 An almost complex structure on a manifold M is a smooth, fiber preserving bundle map J : TM → TM such that J 2 = −1.

Thus, an almost complex structure is simply a way of assigning a com- plex structure to each tangent space smoothly. Usually, we restrict our attention to those almost complex structures which behave nicely with respect to the symplectic structure. Consider

3 2n R equipped with its standard symplectic structure, ω0, Riemannian met- ric, h, i, and complex structure i. These three structures fit together in the formula ω0(·, i ·) = h·, ·i. Since a manifold has no preferred Rieman- nian structure, we abstract this by simply requiring that ω(·,J·) gives a Riemannian metric on M.

Definition 1.2 An almost complex structure J on a symplectic manifold (M, ω) is called ω-compatible if ω(·,J·) defines a Riemannian metric on M.

Let J (M, ω) denote the space of all ω-compatible almost complex struc- tures on M. This space has a topology induced as a subspace of End(TM).

Proposition 1.3 J (M, ω) is non-empty and contractible.

Proof: This is a standard fact. Given r(·, ·) any Riemannian metric on M, there is a unique automorphism A : TM → TM which satisfies ω(·, ·) = 2 1 r(A·, ·). It is not hard to calculate that J := (−A ) 2 A is an almost complex structure compatible with ω. Thus, J (M, ω) is a deformation retract of the space of all Riemannin metrics.



A choice of an almost complex structure makes (TM,J) into a complex vector bundle over M, and as such, it has Chern classes, ci(TM,J) ∈ H2i(M, Z) (for details about Chern classes, consult [17]). Chern classes are defined for any complex vector bundle over a manifold M, and they are natural in the sense that for a complex bundle V → M ∗ ∗ and a smooth map f : N → M, the Chern classes satisfy ci(f V ) = f (ci), where f ∗V is the pullback bundle over N. The Chern classes are defined for each choice of almost complex struc- ture, and a priori, they may depend on the choice. By restricting to ω-

4 compatible almost complex structures, we avoid this difficulty, as the next proposition shows.

Proposition 1.4 Let (M, ω) be a symplectic manifold, and let J 0 and J 1 be elements of J (M, ω). Then the Chern classes of (TM,J 0) and (TM,J 1) agree. That is, the Chern classes are independent of the choice of almost complex structure J ∈ J (M, ω).

This follows from the fact that J (M, ω) is contractible, and the natu- rality of the Chern classes.

According to this proposition, we can write ci(ω) for ci(TM,J), where J is any ω-compatible almost complex structure. We will only have use for the first Chern class, and we will often drop the mention of ω, simply writing c1.

Definition 1.5 A symplectic manifold (M, ω) is called monotone if [ω] =

τc1(ω) for some τ > 0.

It follows from the preceding proposition that the condition of being monotone is independent of the choice of almost complex structure. We will see that this assumption guarantees the non-existence of J-holomorphic spheres of negative Chern number, and simplifies many things. Hopefully, this assumption can be weakened with further study.

Definition 1.6 A map f :(N, j) → (M,J) between almost complex mani- folds is called (j, J)-holomorphic, or pseudo-holomorphic, if Df ◦j = J◦Df.

In the case (N, j) = (M,J) = (C, i), this condition reduces to the Cauchy-Riemann equations, and so generalizes the notion of holomorphic maps. We will mostly consider pseudo-holomorphic maps with domain S2.

5 Definition 1.7 Given A : S2 → (M, ω), the Chern number of A is R ∗ c1(A) := S2 A c1 ∈ Z.

It is clear that the set of all Chern numbers of classes A ∈ π2(M) is a subgroup of Z. We call the positive generator of this subgroup the minimal Chern number of (M, ω). That is, the minimal Chern number of (M, ω) is the smallest positive value that c1 takes on π2(M). Throughout the thesis, we will write N for the minimal Chern number of (M, ω). Endow S2 with it’s standard complex structure, j, and standard sym- plectic form, and choose J ∈ J (M, ω). We will say that a map f :(S2, j) → (M,J) is J-holomorphic if f is (j, J)-holomorphic, and we will call f a J- 2 holomorphic curve in M. Then, given f :(S , j) → (M, ω, J), Dfz is a linear map between normed linear vector spaces for each z ∈ S2, and as such, it has a norm, |Dfz|.

Definition 1.8 For f :(S2, j) → (M, ω, J), the energy of f is given by R 2 2 e(f) = S2 |Dfz| dν, where ν is the standard area form on S .

For a proof of the following proposition, see [1].

Proposition 1.9 Given a symplectic manifold (M, ω), a choice of J ∈ J (M, ω), and a map A :(S2, j) → (M, ω, J), if A is J-holomorphic, then 1 ω(A) = 2 e(A).

This proposition shows that whenever A : S2 → (M, ω, J) is a J- holomorphic map, then ω(A) is positive. If (M, ω) is monotone, this means that c1(A) is also positive, so monotonicity guarantees the non-existence of J-holomorphic spheres of negative Chern number, as claimed. The ex- istence of such spheres obstruct the compactness of the of connecting orbits, and thus must be avoided. We can now introduce the weakly monotone condition.

6 Definition 1.10 A symplectic manifold (M 2n, ω) is called weakly mono- tone if, for all A ∈ π2(M), 3 − n ≤ c1(A) < 0 =⇒ ω(A) ≤ 0.

All monotone manifolds are weakly monotone. As described above, the monotonicity condition guarantees that there are no J-holomorphic spheres of negative Chern number. In the weakly monotone case, this is only generically true.

Lemma 1.11 (Hofer, Salamon) If (M, ω) is a weakly monotone sym- plectic manifold, then there is a generic subset of J (M, ω), such that for all J in this subset, there are no J-holomorphic spheres with negative Chern number.

Since Floer homology is defined using generic (H,J), this will cause no difficulty. (This is not the only additional difficulty in extending Floer homology to the weakly monotone case - others will be mentioned later.) To define his action, Seidel uses time dependent almost complex struc- tures: we let J (M, ω, S1) denote all ω-compatible almost complex struc- 1 1 tures paramaterized by S . That is, J ∈ J (M, ω, S ) is a family (Jt) such that each Jt is an element of J (M, ω), and Jt varies smoothly with t. Be- cause of this use of time dependent almost complex structures, Seidel has to use a slightly stronger assumption than the weakly monotone assumption.

Definition 1.12 A symplectic manifold (M 2n, ω) will be said to satisfy con- + dition W if, for all A ∈ π2(M), 2 − n ≤ c1(A) < 0 =⇒ ω(A) ≤ 0.

Notice that the only difference between condition W + and weakly mono- tone is that 3 − n is replaced by 2 − n. Seidel introduces this slightly stronger statement because he uses time dependent almost complex struc- tures. We need to assume condition W + in order to guarantee that we

7 can choose generic time dependent almost complex structures for which no 1 Jt-holomorphic sphere has negative Chern number for any fixed t ∈ S .

1.2 Floer Homology

Floer homology for a symplectic manifold satisfying condition W + is de- fined by choosing a 1-periodic Hamiltonian function H ∈ C∞(M × S1, R) and a 1-periodic family of almost complex structures J ∈ J (M, ω, S1).

The Hamiltonian defines XHt , a time dependent vector field on M, by the formula i ω = dH . As described above, integration of this family of XHt t H H H vector fields yields θ = (θt ), a path in Diff(M). For any fixed t, θt is a Hamiltonian diffeomorphism. Recall that Floer homology was introduced in an attempt to prove the Arnold Conjecture, which is a statement concerning the fixed points of Hamiltonian diffeomorphisms. Let θ be a Hamiltonian diffeomorphism, and suppose there is a Hamiltonian function H ∈ C∞(M × S1) for which the H H generated path in Diff(M) satisfies θ1 = θ. Then the fixed points of θ = θ1 are in one-to-one correspondence with the set of maps x : S1 → M which are integral curves of the time dependent vector field XHt . We will define a chain complex using this set of loops in M rather than the set of fixed points of θ. Accordingly, let ΛM consist of all smooth, unbased, parameterized maps x : S1 → M. Again, we topologize ΛM as the mapping space C∞(S1,M) with the C∞ topology. Let LM be the subspace of contractible loops. Then LM is topologized as a subspace of ΛM. Notice that an element of LM includes the parameterization. In Floer homology, we use H ∈ C∞(M × S1) to define a 1-form on LM

8 by the formula

Z 1  αH (x)(ξ) = ω x˙(t) − XHt (x(t)), ξ(t) dt. 0

We define P(H) as the set of zeroes of αH . That is,

P(H) = Zeroes(αH ) = {x ∈ LM | x˙(t) = XHt (x(t))}. (1)

H Notice that each x ∈ P(H) corresponds to a fixed point of θ1 . In fact, H each x ∈ P(H) is defined by x(0), because by definition, x(t) = θt (x(0)). We will actually create a chain complex using the set P(H) instead of the fixed point set. To construct the Floer homology groups, we also require that the fixed point set be non-degenerate in the following sense:

Definition 1.13 A loop x ∈ P(H) is called non-degenerate if

1 H  det − Dθ1 (x(0)) 6= 0,

H H where θ = θt is the Hamiltonian isotopy generated by H.

The non-degeneracy condition appears to be necessary to the Fredholm theory needed in the construction of HF∗(M, ω, H, J).

The 1-form αH is not exact on LM, but becomes exact when pulled back to an appropriate covering space, LgM, which we introduce now.

1.2.1 Defining LgM

The space LgM is constructed as first introduced in [13] : an element of LgM is an equivalence class of pairs (v, x) with x ∈ LM and v : D2 → M satisfying x = v|∂D2 . To define the equivalence relation, notice that for

9 ∞ 2 (v0, x) and (v1, x) in C (D ,M) × LM which satisfy v0|∂D2 = v1|∂D2 = x, 2 2 2 we can define v0#v1 : S → M by identifying S with two copies of D glued together along their boundary circle. (The map from S2 to M constructed in this way may not be smooth, but it nevertheless defines an element of

π2(M).)

The equivalence is defined by (v0, x0) ∼ (v1, x1) ⇐⇒ x0 = x1 and

ω(v0#v1) = c1(v0#v1) = 0. Let p : LgM → LM be the canonical projection that takes [v, x] to x. Then p is a covering map.

Let Γ = π2(M)/π2(M)0 where π2(M)0 is the subgroup of classes A for which ω(A) = c1(A) = 0. Then Γ is the group of deck transformations for the covering p : LgM → LM. Geometrically, Γ acts on LgM by “gluing in spheres”. That is, for γ ∈ Γ and v : D2 → M, we define γ#v : D2 → M by cutting out a small disc from both D2 and S2, and identifying the bound- aries of the missing pieces. We will sometimes think of an element of Γ as a homeomorphism of LgM, and sometimes think of it (via a representative) as an element of π2(M). Using this covering, we define

P^(H) = p−1(P(H)).

An element of P^(H) is a 1-periodic integral curve of XHt , with a choice of disc filling it out (up to an equivalence). ∗ When we pull back to LgM, p αH = daH , where aH is defined by

Z Z 1 ∗ aH (c) = v ω − H(t, x(t))dt, D2 0 with c = [v, x]. By Stoke’s theorem, this is independent of the choice of

10 representative for c. This function is called the action functional. Then

−1 ∗ P^(H) = p (Zeroes(αH )) = Zeroes(p αH ) = Crit(aH ).

The idea is to define a chain complex using P^(H), and a boundary operator on this complex for which ∂2 = 0. First, we index P^(H), which requires the non-degeneracy condition.

1.2.2 The Conley-Zehnder Index and the Chain Complex

When the set P^(H) contains only non-degenerate solutions, it can be in- dexed by the Conley-Zehnder index. We will have to use this later, so we give a quick review of how this index is defined. For details, see [20]. Choose any c = [v, x] ∈ P^(H). Then (v∗T M, ω) is a symplectic bundle over D2. Since D2 is contractible, there is a symplectic trivialization τ : ∗ 2 2n (v T M, ω) → D × (R , ω0). This choice of symplectic trivialization is independent of the representative for c up to homotopy. By restricting this 2n trivialization to the boundary circle, we obtain maps τ(t): Tx(t)M → R ∗ 2n such that τ(t) ω0 = ωx(t) where ω0 is the canonical symplectic form on R . 1 H −1 For a fixed t ∈ S , l(t) := τ(t)Dθt τ(0) is a symplectic, linear map from R2n → R2n. By varying t across S1, we obtain a path of symplectic matrices. Let Sp∗(2n) denote the open dense subset of Sp(2n) consisting of all symplectic matrices which do not have 1 as an eigenvalue. In [20], Salamon and Zehnder showed how to assign an integer to each path Ψ : [0, 1] → Sp(2n) which satisfies Ψ(0) = 1 and Ψ(1) ∈ Sp∗(2n). Denote this map by

µCZ . Now, for any c = [v, x] ∈ P^(H), if x ∈ P(H) is non-degenerate, then the path in Sp(2n) given by l(t) defined above satisfies the conditions necessary

11 to define µCZ (l(t)). Thus, we define µH : P^(H) → Z by µH (c) = µCZ (l(t)). Using the Conley-Zehnder index, we set

P^(H)k = {c ∈ P^(H) | µH (c) = k}.

In the monotone case, Floer defined the k-chains as formal sums of elements of P^(H)k with coefficients in Z. For simplicity, we will use co- efficients in Z2. In extending the theory to the weakly monotone and W + cases, Floer’s boundary operator would not be well-defined (this will be explained later). To overcome this difficulty, we have to sacrifice the some simplicity in the chains: define CFk(M, ω, H) as the set of all formal P sums, mchci, with c ∈ P^(H)k and mc ∈ Z2, which satisfy the condition

{c ∈ P^(H)k | mc 6= 0, aH (c) > C} is finite for all C ∈ R. This finiteness condition will be necessary to guarantee that the boundary operator is well defined.

This set is clearly a group. We define CF∗(M, ω, H) as the graded group L CFk(M, ω, H). This graded group clearly has the structure of a module over Z2. It also has a natural structure of a module over the so-called Novikov ring of (M, ω), which we describe now. Notice that Γ, (the group of deck transformations of the covering map p : LgM → LM), acts on CF∗(M, ω, H) by γ · [v, x] = [γ(v), x]. (Since γ ∈ Γ, it defines a homeomorphism of LgM, and thus γ(v) makes sense.) This action does not preserve the grading, and changes according to the formula

µH (γ · c) = µH (c) − 2c1(γ).

This formula is proved in theorem 4 of [8].

To preserve the grading, we index Γ by µ(γ) = −2c1(γ), and we define

12 Λk as all formal sums X mγhγi

γ∈Γk such that mγ ∈ Z2 and {γ ∈ Γk | mγ 6= 0, ω(γ) ≤ C} is finite for all C ∈ R. L Let Λ = k Λk. Then the multiplication in Λ induced by the multiplication in π2(M) makes Λ into a commutative graded ring which we call the Novikov ring of (M, ω). The Γ action on CF∗(M, ω, H) extends naturally to give

CF∗(M, ω, H) a graded Λ-module structure.

Remark 1 The point is that CFk(M, ω, H) is not finitely generated as a module over Z2, but it is finitely generated over Λ, and its dimension is equal to the number of elements of P(H) with a lift to P^(H)(k mod(2N)), where N is the minimal Chern number of ω.

1.2.3 Morse Theory on LgM and the Boundary Operator

The plan is to apply Morse theory to LgM using the function aH . In order to follow Morse theory, we want to give LgM a Riemannian metric, and consider the gradient flow of aH . We will supply LgM with a metric by using a choice of J ∈ J (M, ω, S1). Notice that LM is an infinite dimensional manifold, with local model space near x given by a neighborhood of the zero section inside C∞(x∗TM), where x∗TM is the pullback bundle. The charts are given by exponentia- tion, using an arbitrary metric. (The size of the neighborhood depends on the choice of metric.) A tangent vector to x ∈ LM is a smooth vector field along x. For two vectors in TxLM, we define

Z 1  hξ, ηi = ω ξ(t),Jtη(t) dt. 0

13 This is a metric whenever Jt is ω-compatible. We simply pull back this metric to LgM using the projection map p.

As in the finite dimensional case, the metric and the function aH com- bine to give a vector field on LM, called ∇aH , by h∇aH , ·i = daH (·). In Morse theory, one uses this vector field to define the stable and unstable manifolds of a critical point by flowing along the vector field. However, in the infinite dimensional setting, this vector field cannot be integrated to a flow, because we no longer have existence and uniqueness of solutions to partial differential equations. In any case, it makes sense to speak of integral curves of the vector field. These integral curves can be describe using the following operator. A curveu ˜ : R → LgM projects to a map u : R × S1 → M. For smooth u ∈ C∞(R × S1,M) and (H,J) ∈ C∞(M × S1) × J (M, ω, S1), define ∞ ∗ ∂H,J (u) ∈ C (u TM) by

∂u ∂u ∂ (u)(s, t) = (s, t) + J(u(s, t)) (s, t) − ∇H (u(s, t)). H,J ∂s ∂t t

Here, ∇Ht is the gradient using the metric rt(·, ·) = ω(·,Jt·).

Proposition 1.14 A smooth map u˜ : R → LgM is an integral curve of 1 −∇aH if and only if it’s projection to LM is given by a map u : R×S → M which satisfies ∂H,J (u) = 0.

Proof: ∗ Since the differential of aH is the 1-form p αH , a simple calculation shows that ∇aH (x)(t) = Jt(x(t))x ˙(t) − ∇Ht(x(t)). The integral curves ∂u satisfy ∂s + ∇aH (u(s, ·)) = 0, so the proposition follows.



14 Given c−, c+ ∈ P^(H), the space M(c−, c+,H,J) is defined as the set 1 of all maps u : R × S → M satisfying ∂H,J (u) = 0, which lift to a map u˜ : R → LgM such that lims→±∞ u˜(s) = c±. These are the “flow lines” of the (negative) gradient vector field on LgM defined by (H,J). In Morse Theory, a generic choice of metric guarantees that we can count these connecting orbits between two appropriately in- dexed critical points, and use this to define a boundary operator. Floer showed how to do the same in the infinite dimensional case.

Notice that in the case H = 0, an element of M(c−, c+,H,J) is just a J-holomorphic curve, introduced by Gromov in [12]. Floer realized that many of the ideas that Gromov introduced to deal with J-holomorphic curves apply to these perturbed curves. In particular, for a generic choice of (H,J), the spaces of connecting orbits are manifolds.

Theorem 1.15 (Floer, Hofer, Salamon) Suppose (M, ω) satisfies con- dition W +. There is a generic subset of C∞(M ×S1)×J (M, ω, S1), denoted ∞ 1  ∞ 1 by C (M × S ) × J reg, such that for any pair (H,J) in C (M × S ) ×  ^ J reg, each x ∈ P(H) is non-degenerate, and for c− and c+ in P(H),

M(c−, c+,H,J) is a finite dimensional manifold of dimension µH (c−) −

µH (c+).

Sketch of proof:

Fix c− = [v−, x−] and c+ = [v+, x+] in P^(H). Choose an arbitrary Riemannian metric on M. Let B be the Banach manifold of all W 1,p maps 1 1,p u : R × S → M with limits (in the W sense) given by x− and x+. Define a bundle E → B by letting the fiber over u ∈ B be Lp(u∗TM), the space of p ∗ L sections of u TM. Then ∂H,J defines a section of this Banach bundle.

The parameters (H,J) can be chosen generically so that ∂H,J is transverse −1 to the zero section, and thus ∂H,J (0) is a (Banach) manifold.

15 1 Let u : R × S → M be in M(c−, c+,H,J). The bundle E → B trivializes near u ∈ B as W 1,p(u∗TM) × Lp(u∗TM). The key point is that any u0 ∈ B which is close enough to u is given by exponentiation 1,p ∗ of some ξ ∈ W (u TM). Under this trivialization, the section ∂H,J becomes an elliptic (and therefore Fredholm) operator, denoted Du. By elliptic regularity, all solutions to ∂H,J (u) = 0 are actually smooth, so −1 ∂H,J (0) = M(c−, c+,H,J). There is a homeomorphism from a neigh- −1 borhood of 0 in ker Du to a neighborhood of 0 in ∂H,J (0). Since Du is Fredholm, this kernel is finite dimensional.

The section ∂H,J is transverse to the zero section if and only if each

Du is onto, and in this case, M(c−, c+,H,J) is actually finite dimensional, and the dimension near u is given by index(Du). Finally, it can be shown that index(Du) is exactly given by µH (c−) − µH (c+), if u lifts to a map u˜ : R → LgM with limits c− and c+.



There is a natural R-action on M(c−, c+,H,J) via translation in the infinite direction. We let M(c−, c+,H,J)/R denote the quotient manifold.

If (H,J) is as in the above theorem, and if µH (c−) − µH (c+) = 1, then

M(c−, c+,H,J)/R is a 0-dimensional manifold. We want to count these spaces, so we need compactness.

∞ 1 Theorem 1.16 (Floer, Hofer, Salamon) The set C (M ×S )×J )reg of theorem 1.15 can be chosen such that for any c− and c+ in P^(H) with

µH (c−) − µH (c+) = 1, M(c−, c+,H,J)/R is compact, and thus finite.

Sketch of proof: This again essentially follows from the ellipticity of the linearized oper- ators Du, and the ability to generically avoid bubbling.

16 Let u : R × S1 → M be smooth. Define the energy of such a map by

Z 1 Z ∞   1 ∂u 2 ∂u θH 2 E(u) = | | + | − Xt (u)| ds dt. 2 0 −∞ ∂s ∂t

The main point is that if ∂H,J (u) = 0, then E(u) < ∞ if and only if it converges to two limit circles x− and x+ in P(H). Now, it follows from elliptic methods that for any sequence of maps 1 uj : R × S → M in M(c−, c+,H,J) and which have a uniform bound on

E(uj), and any sequence sj ∈ R, there is a subsequence of uj, still denoted uj, for which the reparameterized sequence vj(s, t) = uj(s − sj, t) has a subsequence which converges uniformly with all derivatives on compact sets, modulo a finite set, to some limit map u : R × S1 → M. This limit map then also satisfies ∂H,J (u) = 0. The energy of the limit map must be less than or equal to the bound on the energy for the sequence, and therefore u lims→±∞(u) = x± ∈ P(H). By choosing different sequences sj, we obtain

finitely many circles x0, x1, ... , xk ∈ P(H), and finitely many maps 1 ui : R × S → M, i = 1...k such that lims→−∞ ui(s, t) = xi−1(t) and lims→∞ ui(s, t) = xi(t). Moreover, all the maps ui satisfy ∂H,J (ui) = 0. What remains is dealing with the finite set involved in the (weak) con- vergence of uj to u. This is described by the phenomenon of bubbling, which is a fundamental part of Gromov compactness. The process can be sum- marized as follows. Let {z1 = (s1, t1), z2 = (s2, t2), ..., zl = (sl, tl)} be the

finite set which arises in the convergence of uj. (That is, uj → u uniformly 1 with all derivatives on compact sets on the domain R × S − {z1, z2, ...zl}.) ν ν 1 One can can find sequences (sj , tj ) ∈ R × S for which the reparameterized ν 1 ν ν ν maps uj : R × S → M given by uj (s, t) = uj(s − sj , t − tj ) converge uniformly with all derivatives on compact sets to a Jtν holomorphic sphere 2 vν : S → M.

17 Suppose µH (c−) − µH (c+) = 1. Then in fact, u#v1#...#vl satisfies

l X E(u) + E(vν) = lim E(uj), j→∞ ν=1

k l X X µ(ui) + 2c1(vj) = dim M(c−, c+,H,J) = 1. i=1 j=1

Since there are no Jt holomorphic sphere with negative Chern number for regular pairs (H,J), this implies that k = 1 and l = 0. That is, there is a single limit cylinder u, and no bubbling occurs.

Finally, since every u ∈ M(c−, c+,H,J) is an integral curve of −∇aH , one can show that E(u) = aH (c+) − aH (c−). This shows that the energy of each u ∈ M(c−, c+,H,J) is constant, so the theorem follows. Notice that this is the point at which the W + condition is necessary.



This theorem was proved in the monotone case by Floer, then extended to the weakly monotone case by Hofer and Salamon.

This fact allows us to define a boundary operator on CF∗(M, ω, H) by counting the number of connecting orbits between two generators. For a regular pair (H,J) and c− ∈ P^(H)k, set

X  ∂k(H,J)(c−) = # M(c−, c+,H,J)/R hc+i.

c+∈P^(H)k−1

Here, # means counting modulo 2. There are two difficulties with this: it is not yet clear that this formal sum satisfies the finiteness condition on the chains in CF∗(M, ω, H), and we

18 need to extend this to infinite linear combinations. Both of these problems can be solved exactly because of the finiteness condition. P For a k-chain κ = mc− hc−i, ∂k(H,J)(κ) is a formal sum c−∈P^(H)k P mc+ hc+i, where the coefficient of c+ ∈ P](ψ)k−1 is given by c+∈P](ψ)k−1 the number of orbits (modulo time shift) connecting any c− ∈ P](ψ)k with mc− 6= 0 to c+. The above map wouldn’t extend to infinite combinations if the number of such orbits could be infinite. The finiteness condition guarantees a uniform bound on the energy of all the orbits contributing to the coefficient of c+, and thus the space of all such connecting orbits is compact as before, and thus finite.

This shows that ∂∗(H,J) gives a boundary operator on CF∗(M, ω, H).

Theorem 1.17 (Floer) Let (M, ω) be a symplectic manifold which satis- + ∞ 1 fies condition W . The space (C (M × S ) × J )reg can be chosen so that for regular pairs (H,J), the boundary operator ∂(H,J) satisfies and ∂2 = 0.

Sketch of proof:

Let c− and c+ be such that µH (c−) − µH (c+) = 2. Now, the mani- fold M(c−, c+,H,J)/R is not compact, but it is compact up to splitting of orbits. That is, we can think of M(c−, c+,H,J)/R as a compact 1- dimensional manifold with boundary given by

[ ∂(M(c−, c+,H,J)/R) = (M(c−, c0,H,J)/R) × (M(c0, c+,H,J)/R).

c0∈P^(H)k−1

This depends on some general position arguments to avoid bubbling, coupled with Floer’s gluing procedure.

Now, ∂ ◦ ∂(c−) is a formal sum of elements of P^(H)k−1, and the coeffi- cient c+ in the formal sum ∂ ◦ ∂(c−) is exactly the number of elements in

∂(M(c−, c+,H,J)/R). But this must be an even number, by the classifica-

19 tion of compact 1-manifolds.



 The homology of the chain complex CF∗(M, ω, H), ∂∗(H,J) is denoted

HF∗(M, ω, H, J).

The differentials ∂∗(H,J) are linear over the Novikov ring Λ, so we see that HF∗(M, ω, H, J) is also a Λ-module.

1.2.4 Continuation Isomorphisms

Finally, one can prove that these homology groups are actually independent of the choice of (regular) pair (H,J). This is done by use of the so called “continuation maps”. For two regular pairs (H−,J −) and (H+,J +), a homotopy of regular ∞ 1 s 1 pairs is a function H ∈ C (M ×R×S ) and a J = (Jt ) ∈ J (M, ω, R×S ) that satisfy the following condition:

s − − s ≤ −1 =⇒ (Ht(·, s),Jt ) = (Ht (·),Jt ), s + + s ≥ 1 =⇒ (Ht(·, s),Jt ) = (Ht (·),Jt ).

Given a homotopy Φ from (H−,J −) to (H+,J +) as above and a fixed s s ∈ R, we will usually denote by Φ(s) the pair H(·, s, ·),J . − + − − For c− ∈ P^(H ), c+ ∈ P^(H ) and Φ a homotopy of pairs from (H ,J ) + + to (H ,J ), the space M(c−, c+, Φ) is defined as the space of all smooth maps u : R × S1 → M which satisfy

∂u ∂u + J (u)( ) − ∇H(s, t, u) = 0, ∂s s,t ∂t and which can be lifted to a pathu ˜ : R → LgM with limits c−, c+.

20 Theorem 1.18 (Floer) For a generic homotopy of regular pairs, all of the spaces M(c−, c+, Φ) have a finite number of elements whenever µH− (c−) =

µH+ (c+).

This is proved in the same fashion as theorem 1.16. Such a homotopy is called regular. We define the “continuation homomorphisms” for a regular homotopy by − + Φk(H, J): CFk(M, ω, H ) → CFk(M, ω, H )

X Φ − hc−i 7→ #M (c−, c+,H, J)hc+i, c− ∈ P^(H )k.

+ c+∈P^(H )k Again, this formula is to be extended linearly to infinite formal sums, and # means counting modulo 2. Regularity implies that # again makes sense. This is a homomorphism of chain complexes and it induces an isomorphism on the homology level. This isomorphism is denoted Φ∗(H, J).

− − Theorem 1.19 The continuation isomorphism from HF∗(M, ω, H ,J ) + + to HF∗(M, ω, H ,J ) is independent of the choice of regular homotopy from (H−,J −) to (H+,J +), and the isomorphisms are functorial with respect of composition of homotopies.

Sketch of proof: Given two homotopies Φα and Φβ, one chooses a “homotopy of homo- topies” connecting Φα and Φβ. This consists of a function H ∈ C∞(M × [0, 1]×R×S1) and an almost complex structure J ∈ J (M, ω, [0, 1]×R×S1). These must be such that

0,s α H(p, 0, s, t),Jt = Φ (s), 1,s β H(p, 1, s, t),Jt = Φ (s).

21 The idea is that by fixing r ∈ [0, 1], we get a homotopy Φr from (H−,J −) to (H+,J +), and by varying r, we move from Φα to Φβ. ± Given such a homotopy of homotopies Ψ and c± ∈ P^(H ) , consider the parameterized modulii space M(c−, c+, Ψ) which consists of the set of r all pairs (u, r) for which u ∈ M(c−, c+, Φ ), and index Du = −1. For a generic homotopy of homotopies, these spaces are compact, 0-dimensional manifolds whenever µH− (c−) = µH+ (c+) − 1, and this gives a map Ψ : − + CF∗(M, ω, H ) → CF∗+1(M, ω, H ). This map is a chain homotopy be- tween the the isomorphisms corresponding to Φα and Φβ, and thus the continuations induce the same isomorphisms on homology.



This means that HF∗(M, ω) makes sense without mention of the choice of regular pair (H,J). In fact, the continuation homomorphisms are also linear over Λ, so we have an induced Λ-module structure on HF∗(M, ω) independent of the choice of regular pair used to define it.

The final step is in proving that HF∗(M, ω) recovers to homology of the underlying manifold. The quantum homology QH∗(M, ω) is the graded Λ-module given by

M QHk(M, ω) = Hi(M, Z2) ⊗ Λj. i+j=k

Theorem 1.20 (Piunikhin, Salamon, Schwarz) The quantum homol- ogy and Floer homology agree. That is, HF∗(M, ω) w QH∗(M, ω) as graded Λ-modules.

∞ 1 To summarize, H ∈ C (M × S ) is used to create CF∗(M, ω, H). For a generic choice of (H,J) ∈ C∞(M × S1) × J (M, ω, S1), we can define a boundary operator on CF∗(M, ω, H), and we can define HF∗(M, ω, H, J) as

22  the homology of the pair CF∗(M, ω, H), ∂∗(H,J) . There are continuation isomorphisms between the groups defined using any choice of regular pairs, so we may speak of HF∗(M, ω) independently of the choice of regular pair used to define it.

1.3 Seidel’s Action

In [21], P. Seidel introduced an action of a certain extension of π1(Ham(M)) on the Floer homology groups of (M, ω). More precisely, he introduces a topological group Ge which fits into the exact sequence

1 → Γ → Ge → G → 1, where Γ is discreet. To describe this action, we first need to introduce the extension he uses.

1.3.1 The Extension Ge

Let G be the group of smooth loops in Ham(M) based at id. More precisely, 1 1 for continuous g : S → Diff(M), defineg ¯ : M × S → M byg ¯(p, t) = gt(p). We say that g is smooth ifg ¯ is smooth. Then let

1 G = {g : S → Ham(M) | g is smooth, g0 = g1 = id}.

We topologize G as a subspace of C∞(M × S1,M) with the C∞ topology.

Lemma 1.21 π0(G) = π1(Ham(M))

Proof: This is not immediate, because the fundamental group considers contin- uous loops, and G contains only smooth loops, but it follows from the fact

23 that any continuous loop in Ham(M) can be approximated by a smooth loop, coupled with the fact that there is a neighborhood of id in Ham(M) such that each path component of this neighborhood is contractible .



Recall that ΛM consists of all parameterized loops in M, and LM is the subspace of contractible loops. There is a natural action of G on ΛM: for x ∈ ΛM and g ∈ G, define (g · x) by

(g · x)(t) = gt(x(t)). (2)

Proposition 1.22 G ·LM = LM

This is actually a consequence of theorem 1.20. We will reproduce Sei- del’s proof after some preliminary facts are established. Notice that if we consider this action applied to loops which are con- stant in M, we obtain the folkloric fact that the image of a point under a Hamiltonian loop is contractible. It was first noticed in [4] that the image of a point under a Hamiltonian loop is homologous to zero, but contractibility is essentially a consequence of theorem 1.20. Recall that LgM is a covering space over LM. For g ∈ G, we have a map g : LM → LM given by equation 2. (See the following diagram.)

LgM LgM p p  g  LM / LM

Lemma 1.23 (Seidel) The action of any g ∈ G lifts to a homeomorphism of LgM.

24 Proof: By the Lifting Theorem, we need only show that the action of g on LM preserves the set of smooth maps S1 → LM which lift to a map S1 × S1 → LgM. Any such map is given by a map A ∈ C∞(S1 × S1,M), with ω(A) = c1(A) = 0. The image of A under the loop g is given by ∗ ∗ (g · A)(s, t) = gt(A(s, t)). Because g is Hamiltonian, (g · A) ω = A (ω), and thus ω(g · A) = ω(A). To prove that c1(g · A) = 0, notice that g induces an isomorphism of symplectic vector bundles from A∗TM to (g · A)∗TM by

(s, t, ξ) → (s, t, Dgtξ). The statement then follows from naturality of the Chern classes.



We can now introduce the extension of G which is used in Seidel’s action.  Let Ge ⊂ G × Homeo(LgM) consist of all pairs (g, g˜) such thatg ˜ is a lift of the action of g on LM to a homeomorphism of LgM. We topologize Ge by using the C∞ topology on G and the topology of pointwise convergence on Homeo(LgM). The preceding lemma shows that Ge is non-empty. Then Ge is a topological group. The kernel of the projection Ge → G is exactly all pairs (id, γ), with γ ∈ Γ. (Recall that Γ is the group of deck transformations of the covering LgM → LM.) Thus, there is the following exact sequence of topological groups.

1 → Γ → Ge → G → 1 (3)

1.3.2 The Map σ˜ : Ge → Aut(HF∗(M, ω))

The basic idea in Seidel’s action is to choose a regular pair (H,J), and to use g ∈ G to define a new regular pair (Hg,J g). This new pair is defined

25 in such a way that the action of g on LM gives

g ·P(Hg) = P(H).

By lifting to (g, g˜) ∈ Ge, we see that

g˜(P^(Hg)) = P^(H).

The map hci 7→ hg˜(c)i induces a chain map between the complexes g g g   CF∗(M, ω, H ), ∂∗(H ,J ) and CF∗(M, ω, H), ∂∗(H,J) . It follows that the map hci → hg˜(c)i induces a homomorphism

g g g˜ HF∗(M, ω, H ,J ) → HF∗(M, ω, H, J).

This map is an isomorphism, with inverse induced by (g−1, g˜−1). By precomposing with the continuation isomorphism (see section 1.2.4) g g from HF∗(M, ω, H, J) to HF∗(M, ω, H ,J ) we obtain an automorphism of

HF∗(M, ω, H, J). Next, Seidel proves that this automorphism is indepen- dent of the regular pair used to define HF∗(M, ω), in the sense that the appropriate diagrams are commutative. Finally, he shows that the auto- morphism is well defined on π0(Ge). We will have need of some of the details contained inside this quick description, so we will expand the explanation a bit now. Begin by choosing a regular pair (H,J), and let g ∈ G. Choose Kg ∈ C∞(M × S1), a Hamiltonian function generating g.

Definition 1.24 For H, g and Kg as above, define Hg ∈ C∞(M × S1) by g g H (p, t) = H(gt(p), t) − K (gt(p), t).

∞ 1 ∗ Hg Lemma 1.25 For all g ∈ G and H ∈ C (M × S ), g αH = α .

26 Proof: This is a straightforward caclulation.



We can use this to give a proof of proposition 1.22. Proof of proposition 1.22: If g ·LM 6= LM, then g ·LM does not intersect LM. (This is because LM is a connected component of ΛM.) Suppose (H,J) is a regular pair, and

H chosen small enough so that the only zeros of αH are the critical points of H, and thus αH has no zeros in g ·LM. It then follows from 1.25 that Hg g g α has no zeros in LM. But this would mean that HF∗(M, ω, H ,J ) = 0, which contradicts theorem 1.20.



Lemma 1.26 For g ∈ G, g ·P(Hg) = P(H).

Proof:

Recall that the function H is used to define a 1-form on LM, called αH , and P(H) = Zeroes(αH ). The lemma follows from lemma 1.25.



It follows from the previous lemma that for any choice of (g, g˜) ∈ Ge, g˜(P^(Hg)) = P^(H). However, this map does not preserve the Conley- Zehnder index. To describe the change in index, Seidel introduces a “Maslov Index” on Ge. The index is defined as follows: Choose a point c = [v, x] ∈ ∗ 2 2n LgM. Trivialize the symplectic bundle (v T M, ω) as D × (R , ω0). Let 2n At :(Tx(t)M, ω) → (R , ω0) be this trivialization on the boundary circle. 0 0 ∗  2 2n Considerg ˜(c) = [v , g · x] ∈ LgM. Trivialize (v ) T M, ω as D × (R , ω0),

27 2n and let Bt :(Tgt(x(t))M, ω) → (R , ω0) be the trivialization on the bound- −1 ary, as before. Then BtDgt(At) gives a linear, symplectic map from R2n to R2n, and is thus given by a symplectic matrix, denoted l(t). Then {l(t)} is a loop in Sp(2n,R), and this loop is independent, up to homo- topy, of all the choices involved, including the choice of c ∈ LgM. Let deg : π1(Sp(2n, R)) → Z be the canonical isomorphism induced by the de- terminant function on U(n) ⊂ Sp(2n, R). For details on this isomorphism, consult [1].

Definition 1.27 Define an index map I : Ge → Z by I(g, g˜) = deg(l), where l is the loop in Symp(2n) defined above.

The following is a simple lemma of Seidel:

Lemma 1.28 (Seidel) I(g, g˜) depends only on [g, g˜] ∈ π0(Ge),I is a ho- momorphism, and I(id, γ) = c1(γ).

This index is related to lemma 1.26 by the following lemma.

Lemma 1.29 (Seidel) For (H,J) and g as above,

µHg (c) = µH (˜g(c)) − 2I(g, g˜).

g This shows that hci 7→ hg˜(c)i induces a map from CFk(M, ω, H ) to

CFk−2I(g,g˜)(M, ω, H). To guarantee that this is a chain map, we have to adjust the choice of almost complex structure. For J ∈ J (M, ω, S1) and g ∈ G, define J g ∈ J (M, ω, S1) by

g −1 Jt = Dgt JtDgt.

28 Seidel shows that if (H,J) is a regular pair, then so is (Hg,J g).

Proposition 1.30 Suppose (H,J) is a regular pair. Let c− and c+ be g g g in P^(H ). Then there is a bijection between M(c−, c+,H ,J )/R and

M(˜g(c−), g˜(c+),H,J)/R.

Proof: g g Let u : R → LgM be and element of M(c−, c+,H ,J ). Define v : R → LgM by v(s) =g ˜(u(s)). It is not hard to check that v is an element of

M(˜g(c−), g˜(c+),H,J), and that this map is R-equivariant.



Since the boundary operator consists of counting the number of elements in these spaces, it follows that hci → hg˜(c)i induces a chain map from g g g   CF∗(M, ω, H ), ∂∗(H ,J ) to CF∗(M, ω, H), ∂∗(H,J) . As described above, this map is an isomorphism still denotedg ˜. This gives an auto- morphism by HF∗(M, ω, H, J) by

Φ g g g˜ HF∗(M, ω, H, J) → HF∗(M, ω, H ,J ) → HF∗(M, ω, H, J), where Φ denotes the continuation isomorphism.

In order to show that this induces an automorphism of HF∗(M, ω) in- dependently of the choice of (H,J), Seidel proves the following theorem.

Theorem 1.31 (Seidel) Let (H−,J −) and (H+,J +) be regular pairs. For (g, g˜) ∈ Ge, the following diagram commutes. (Here, Φ denotes the contin- uation isomorphism between the appropriate homologies.)

29 − − Φ − g − g g˜ − − HF∗(M, ω, H ,J ) / HF∗(M, ω, (H ) , (J ) ) / HF∗(M, ω, H ,J )

Φ Φ Φ  Φ   HF (M, ω, H+,J +) / HF (M, ω, (H+)g, (J +)g) / HF (M, ω, H+,J +) ∗ ∗ g˜ ∗

Next, Seidel shows that this map is linear over Λ. This shows that each (g, g˜) ∈ Ge induces an automorphism of HF∗(M, ω) as a Λ-module independent of the choice of regular pair used to define HF∗(M, ω). This automorphism is denoted HF∗(g, g˜). To complete the construction, Seidel proves that the automorphism is well defined on π0(Ge).

Theorem 1.32 (Seidel) If (g0, g˜0) and (g1, g˜1) define the same element 0 0 1 1 of π0(Ge), then HF∗(g , g˜ ) = HF∗(g , g˜ ).

This theorem means that for any regular pair (H,J), the following dia- 0 0 1 1 gram commutes whenever [g , g˜ ] = [g , g˜ ] inside π0(Ge).

g˜0 0 g0 0 g0 0 0 HF∗(M, ω, (H ) , (J ) ) / HF∗(M, ω, H ,J )

Φ Φ   1 g1 1 g1 1 1 HF (M, ω, (H ) , (J ) ) / HF∗(M, ω, H ,J ) ∗ g˜1

Again, Φ means the continuation isomorphisms between the appropriate homologies. This gives a map

σ˜ : π0(Ge) → Aut(HF∗(M, ω)),

(g, g˜) 7→ HF∗(g, g˜).

We list some basic properties of this map, which are clear from the

30 definition.

Proposition 1.33 The map σ˜ has the following properties:

1. HF∗(g, g˜) is an automorphism of HF∗(M, ω) as a Λ-module.

2. If (g, g˜) = Id , then HF (g, g˜) = Id . Ge ∗ Aut(HF∗(M,ω))

3. If (g, g˜) = (id, γ) for some γ ∈ Γ, then HF∗(g, g˜) is given by multipli- cation by γ.

4. The map σ˜ is a homomorphism.

In order to apply this to , we consider the related map

σ : π1(Ham(M)) → Aut(HF∗(M, ω))/Γ, which is well defined because of the exactness of the sequence (3). (Here, we identify γ ∈ Γ with the automorphism of HF∗(M, ω) induced by hci 7→ hγ(c)i. Also recall that π0(G) = π1(Ham(M)).) This ends the review of Floer homology and Seidel’s homomorphism.

31 2 Rephrasing σ˜ in terms of Isotopies

The main goal in this thesis is to prove that the homomorphism

σ : π1(Ham(M)) → Aut(HF∗(M, ω))/Γ

is well defined on the group i∗(π1(Ham(M))), where i : Ham(M) → Diff(M) is the inclusion. That is, we would like to show that if two loops in Ham(M) are homotopic through loops in Diff(M), then the image of the classes of these loops agree under the homomorphisms σ. Accordingly, we are interested in the loops in Diff(M) which are homotopic to a Hamiltonian loop.

Definition 2.1 A loop at id in Diff(M) will be called almost Hamilto- nian is it is homotopic to a loop in Ham(M).

Let i : Ham(M) → Diff(M) be the inclusion, and

i∗ : π1(Ham(M)) → π1(Diff(M)) be the map on fundamental groups induced by i. It is clear that g is almost

Hamiltonian if and only if [g] ∈ π1(Diff(M)) is in the image of i∗. We will denote the space of all almost Hamiltonian loops by D. Then D is clearly a subgroup of Diff(M). We equip D with the C∞ topology. Consider σ as a map with domain G. We will extend the domain of definition of σ to the (possibly) larger group D, and then prove this extended map is well defined on π0(D).

32 2.1 Calculating θHg

In order to extend to domain of σ, it it instructive to calculate the Hamil- tonian isotopy generated by Hg in terms of g and the isotopy generated by H. Let N denote the space of all smooth isotopies of M. That is, elements of N are maps ψ : [0, 1] → Diff(M) with ψ(0) = id such that the associated map [0, 1] × M → M given by (t, p) 7→ ψ(t)(p) is smooth. This space is topologized as a subspace of C∞([0, 1] × M,M). We will generally write

ψt for ψ(t). By a slight abuse of notation, ψt will sometimes represent the entire isotopy. For a detailed analysis of these spaces of isotopies, consult [2]. There is a natural action of D on N given by

−1 (g ∗ ψ)t := gt ψt.

The reason that we include an inverse here will become clear later.

ψ Definition 2.2 For ψ = (ψt) an isotopy of M, Xt is the family of vector

ψ dψt −1 fields on M defined by Xt (p) = dt (ψt (p)).

ψ The correspondence ψ ↔ Xt is one-to-one, as each smooth 1-parameter family of vector fields determines an isotopy by integration. (Recall that M is compact.) Notice that if θ is a Hamiltonian isotopy generated by H, then by defi-

dθt −1 θ nition, dt (θt (p)) = XHt , so we see that Xt = XHt . We will need the chain rule for isotopies:

Lemma 2.3 (Chain Rule) For any isotopies φ = (φt) and ψ = (ψt) of M, the following identities hold:

33 ψ·φ ψ φ −1 1) Xt = Xt + DψtXt ◦ ψt φ−1 −1 φ 2) Xt = −Dφt Xt ◦ φt



We will also need the following lemma.

Lemma 2.4 For g ∈ G, H ∈ C∞(M × S1), and Kg ∈ C∞(M × S1) generating g, the following identities hold: −1 1) Dgt XHt ◦ gt = XHt◦gt −1 2) Dg X g ◦ g = X g t Kt t Kt ◦gt Proof: To see this, consider the following string of unravelling definitions:

−1 i ω(Yp) = ω(XHt (gt(p)), DgtYp) Dgt XHt ◦gt = g∗(i ω(Y )) t XHt p ∗ = gt (dHt)(Yp)

= d(Ht ◦ gt)(Yp) = i ω(Y ) XHt◦gt p

Since a vector field is uniquely defined by it’s interior multiplication with ω, the conclusion follows. The same proof holds for the other identity.



We can now calculate the isotopy generated by Hg. Recall that for H ∈ C∞(M × S1), θH denotes the Hamiltonian isotopy of M generated by H.

34 Lemma 2.5 Let g ∈ G with generating function Kg ∈ C∞(M × S1), and let H ∈ C∞(M ×S1). Let Hg ∈ C∞(M ×S1) be as in definition 1.24. Then θHg = g ∗ θH .

Proof: g∗θH g g We need only show that X = X g . Recall that H = H ◦g −K ◦g , t Ht t t t t t H −1 H and (g ∗ θ )t = gt θt . Then lemma 2.3 gives:

g∗θH g−1 −1 θH Xt = Xt + Dgt Xt ◦ gt −1 g −1 θH = −Dgt Xt ◦ gt + Dgt Xt ◦ gt

g θH Now, X = X g , and X = X by definition. This, combined with t Kt t Ht lemma 2.4, gives:

g∗θH X = X − X g t Ht◦gt Kt ◦gt

= X g (Ht−Kt )◦gt

= X g Ht

This proves the proposition.



This is why we include an inverse in the definition of the ∗ operation. This will allow us to rephrase Seidel’s homomorphism in terms of iso- topies. This rephrasing suggests a way of extending the domain of σ.

2.2 Emphasizing Isotopies

The goal of this subsection is to prove the following (nearly obvious) propo- sition:

35 Proposition 2.6 If (H,J) and (H0,J) are two regular pairs such that θH = H0 0 θ , then HF∗(M, ω, H, J) = HF∗(M, ω, H ,J).

Proof: The proof is essentially a string of trivial observations, but there are enough that it seems prudent to mention them. As described in the first section, the construction of the Floer homology groups HF∗(M, ω, H, J) is best described using an analogy with Morse The- ory. The chain complex CF∗(M, ω, H) is generated by the critical points of the action functional, indexed appropriately (with a finiteness condition). The boundary operator consists of counting “flow lines” of the the negative gradient vector field on LgM corresponding to the action functional and a metric on LgM defined by the almost complex structure J. These flow lines are solutions of the differential equation ∂H,J (u) = 0. Thus, to prove this proposition, we must only show that if (H,J) and (H0,J) are as in proposi- 0 tion 2.6, (meaning they generate the same isotopy), then P^(H)k = P^(H )k, 0 M(c−, c+,H,J) = M(c−, c+,H ,J), and ∂H,J = ∂H0,J . Given these facts, proposition 2.6 is obvious, since HF∗(M, ω, H, J) depends only on these objects. Let H and H0 generate the same Hamiltonian isotopy. It then follows that P(H) = P(H0), since these sets consist of loops obtained by flowing along the corresponding isotopy. Thus, P^(H) = P^(H0). Since the Conley- Zehnder index is obtained by linearizing according to the isotopy, we see 0 that P^(H)k = P^(H )k. The assumption of the proposition means that H and H0 differ by a constant, so the finiteness conditions in the definitions 0 of CF∗(M, ω, H) and CF∗(M, ω, H ) agree, showing that CF∗(M, ω, H) = 0 CF∗(M, ω, H ). The last step is to see that the boundary operators match.

This is easy to see by noticing that ∂H,J = ∂H0,J , which is true because ∇H = ∇H0.

36 

Notice that this truly means equals. An isomorphism is guaranteed by the continuation maps, but this propositions says that the two groups are equal by contruction. This proposition allows us to shift our attention away from the function, and toward it’s generated isotopy. In fact, in later sections, we will have no function at all, and begin with an isotopy.

At each step in the construction of the Λ-module HF∗(M, ω, H, J), we can convert the dependence on the function, H, to dependence on the iso- topy, θH . Now that proposition 2.6 has been proved, we are justified in making the following definition.

Definition 2.7 For θ = θH a Hamiltonian isotopy and J ∈ J (M, ω, S1), if (H,J) is a regular, we define HF∗(M, ω, θ, J) as HF∗(M, ω, H, J).

2.3 Rephrasing σ˜

According to proposition 2.6, we can define HF∗(M, ω) via a generic choice of Hamiltonian isotopy and almost complex structure. We can now rephrase the mapσ ˜ in terms of isotopies.

Definition 2.8 For a Hamiltonian isotopy θ = θH and an almost complex structure J ∈ J (M, ω, S1), we will say that the pair (θ, J) is a regular pair ∞ 1 if (H,J) ∈ C (M × S ) × J )reg.

Let (θ, J) be a regular pair, and choose (g, g˜) ∈ Ge. Then in terms g of isotopies, (g, g˜) induces an isomorphism from HF∗(M, ω, g ∗ θ, J ) to Hg HF∗(M, ω, θ, J), because θ = g ∗ θ. By precomposing with the continua- g tion isomorphism from HF∗(M, ω, θ, J) to HF∗(M, ω, g ∗ θ, J ), we recover

37 the automorphism of HF∗(M, ω, θ, J) defined by (g, g˜). That is, HF∗(g, g˜) is the automorphism of HF∗(M, ω, θ, J) given by

Φ g g˜ HF∗(M, ω, θ, J) −→ HF∗(M, ω, g ∗ θ, J ) −→ HF∗(M, ω, θ, J). (4)

This suggests a method of solving the problem at hand: if we extend the space of choices available in defining HF∗(M, ω) to include isotopies of the form g ∗ θ, where g is merely almost Hamiltonian, then we can simply use (4) to extend the domain ofσ ˜. (There are many additional details. First, we will need to adjust the compatibility requirements on the almost complex structures. Second, (g, g˜) is an element of Ge. We will need an analogous extension of D.)

The bulk of the thesis consists of constructing HF∗(M, ω, ψ, J), where ψ is an isotopy of M of the form g ∗ θ, for some g ∈ D and Hamiltonian isotopy θ.

38 3 Almost Hamiltonian Loops and Isotopies

In the previous section, we showed that the Hamiltonian function is not as important as it’s generated (Hamiltonian) isotopy. The fundamental idea in the proof of the main result of the thesis is in constructing a Floer Homology for an isotopy which is not Hamiltonian, but is very close to being Hamiltonian. This section describes the notation and basic properties of these “almost Hamiltonian” loops and isotopies. We also introduce the spaces necessary to deal with the almost complex structures. Since our isotopies may not preserve the symplectic structure, we will no longer require that our choice of almost complex structure be compatible with ω, but we will use a restriction that is related to the isotopy. This section contains most of the technical details that are required in the construction given in section 4.

3.1 Almost Hamiltonian Loops of Diffeomorphisms

We begin by letting Ω Diffid(M) denote the group of smooth loops at id in Diff(M), with the multiplication given by time-wise composition. We have already introduced the subgroups G and D, which consist of Hamiltonian loops and almost Hamiltonian loops, respectively.

There is a natural action of Ω Diffid(M) on ΛM (the space of parame- terized loops in M) given by (g ·x)(t) = gt(x(t)) for all loops g ∈ Ω Diffid M and x ∈ ΛM. If we restrict our attention to Hamiltonian loops of diffeomorphisms, Seidel showed that contractibility of loops in M is preserved. (This is lemma 1.22.) This fact is also true for almost Hamiltonian loops:

Lemma 3.1 D ·LM = LM

39 Proof: Let x be in LM, and let g be an almost Hamiltonian loop. Choose h ∈ G such that g ∼ h. Then g · x ∼ h · x, and since h is Hamiltonian, h · x is contractible by lemma 1.22, and thus g · x is contractible as well.



Recall that in the Hamiltonian case, this fact was used to define an extension of G by proving that the action always lifts to a homeomorphism of LgM. We prove the analogous theorem in the almost Hamiltonian case.

Lemma 3.2 The action of any g ∈ D on LM can be lifted to a homeo- morphism of LgM.

Proof: Apply the Lifting Theorem to the following diagram:

LgM  p y p g LgM −−−→ LM −−−→ LM Then if the condition on the fundamental groups is satisfied, there exists a unique lift (for each choice of base points),g ˜ : LgM → LgM. Keep the base points fixed, and apply the Lifting Theorem to the diagram:

LgM  p y LM ←−−− LM ←−−− LgM g−1 p

Then the unique lift from this diagram is exactlyg ˜−1, so the lift is a home- omorphism.

40 So, we must show that the condition on the fundamental groups holds.    That is, we must show that g# p#(π1(LgM)) ⊂ p#(π1(LgM)). This translates to showing that the action of g on LM preserves the set of smooth maps S1 → LM which can be lifted to LgM, because LgM is a connected covering, and p#(π1(LgM)) consists of loops in LM which have a lift to LgM. To this end, let s : S1 → LM be a smooth map ands ˜ : S1 → LgM a lift of s to LgM. We need to show that (g · s): S1 → LM also lifts to LgM. We know that g ∼ h where h is Hamiltonian. We also know that (h · s) lifts to LgM, because h is Hamiltonian (Seidel proves this). Let hg· s be such a lift. Then (hg· s)(t) is an equivalence class of disc in M with boundary equal to (h · s)(t). Because g ∼ h,(g · s) ∼ (h · s). Each (g · s)(t) is a loop in M. To lift to LgM, we need to smoothly specify a choice of disc in M with boundary equal to (g · s)(t). Fixing t, apply the given homotopy to (g · s)(t). This gives us a cylinder in M with boundary loops (g · s)(t) and (h · s)(t). Cap this cylinder off on the (h · s)(t) end using hg· s(t). Do this for each t. This is a smooth loop because the homotopy is smooth and the lift of h was smooth.



 Analogously to the Hamiltonian case, we let De ⊂ D × Homeo(LgM) consist of all pairs (g, g˜) such thatg ˜ is a lift of the g-action to LgM. Give De the topology induced from the C∞ topology on D and the topol- ogy of pointwise convergence on Homeo(LgM). Then De is a topological group. As before, Γ denotes the group of deck transformations of the covering p : LgM → LM. We again have the following short exact sequence of topological groups:

41 1 → Γ → De → D → 1 (5)

3.2 Almost Hamiltonian Isotopies

Since we are shifting our attention to isotopies of M, we need to develop some notation and basic properties of the subgroups in which we are inter- ested. ψ Recall that if ψ = (ψt) is an isotopy of M, then Xt denotes the family of vector fields on M generated by ψ. Also recall that if θ is a Hamiltonian θ isotopy generated by H, then Xt = XHt . In Floer homology, we choose a smooth 1-periodic Hamiltonian function, so that the generated isotopy is smooth, and has a 1-periodic family of vector fields. We will also restrict to the periodic case. ψ We let K denote the space of all smooth isotopies of M for which Xt is ψ ψ 1-periodic. (This means that X0 = X1 .) This space is a subgroup of the group of all isotopies. By H, we denote the subgroup of K consisting of smooth Hamiltonian isotopies of M which are generated by functions in C∞(M × S1). There is a surjection C∞(M × S1) → H which takes H to θH . By theorems 1.15, 1.16 and 1.17, there is a dense subset,

∞ 1 ∞ 1 1  (C (M × S ) × J )reg ⊂ C (M × S ), J (M, ω, S ) ,

for which we can define HF∗(M, ω, H, J).

Definition 3.3 A pair (θ, J) ∈ H × J (M, ω, S1) will be called a regular pair if θ = θH for some H ∈ C∞(M × S1) and (H,J) ∈ (C∞(M × S1) ×

J )reg. (This simply means that HF∗(M, ω, H, J) is well-defined.) The set

42  of all regular pairs will be denoted H × J reg. This definition simply converts the Hamiltonian function to it’s gener- ated isotopy, as we have been doing all along. In other words, the pair H ∞ 1 (θ ,J) ∈ (H × J ) is regular if and only if (H,J) ∈ (C (M × S ) × J )reg.

Proposition 3.4 (H × J )reg is dense inside of (H × J ).

Proof:

(H × J )reg is the image of a dense set under a surjective, continuous map.



We are interested in isotopies of M which are of the form g ∗ θ, for g ∈ D and θ ∈ H. Any such isotopy is homotopic, relative endpoints, to a Hamiltonian isotopy. (Just homotope g to a Hamiltonian loop.)

Definition 3.5 An isotopy ψ ∈ K will be called almost Hamiltonian if ψ is homotopic, relative endpoints, to a Hamiltonian isotopy θ ∈ H. The group of all almost Hamiltonian isotopies will be denoted I.

Notice that by definition, almost Hamiltonian isotopies are smooth and 1-periodic.

Clearly, an isotopy ψ is almost Hamiltonian if and only if [ψ] ∈ Diff^id(M) has a representative Hamiltonian isotopy. Also, I is obviously a subgroup of Ωid Diff(M). We will work predominantly with the spaces H and I of smooth periodic Hamiltonian isotopies and almost Hamiltonian isotopies respectively.

43 3.3 Difficulties with Almost Complex Structures

We intend to define a Floer homology for an isotopy in I. To do so, we need to choose a J ∈ J (M,S1). In the standard setting, (when the isotopy is Hamiltonian), this choice is limited to J (M, ω, S1). That is, we must choose ω-compatible almost complex structures. When extending this to I (almost Hamiltonian isotopies), we are faced with the difficulty that ψ ∈ I may include diffeomorphisms which do not preserve the symplectic structure. The solution is to consider the loop of symplectic structures defined by the isotopy. In other words, given ψ ∈ I, consider the family of symplectic −1 ∗ structures defined by ωt := (ψt ) ω. (The inverse is an annoying, but necessary technical detail.) Since ψ ∈ I, we have that ψ1 ∈ Ham(M), so −1 ∗ (ψ1 ) ω = ω, thus ω0 = ω1 = ω, so this is actually a loop of symplectic structures. ψ 1 1 For ψ = ψt ∈ I, let J (M, ω, S ) ⊂ J (M,S ) be all time dependent almost complex structures on M such that Jt is compatible with ωt. That is, ψ 1 1 −1 ∗ J (M, ω, S ) = {J ∈ J (M,S ) | Jt ∈ J (M, (ψt ) ω)}.

This space is used to create a bundle over I by setting the fiber over ψ ∈ I to be J ψ(M, ω, S1). That is,

F = {(ψ, J) ∈ I × J (M,S1) | J ∈ J ψ(M, ω, S1)}.

This space has a topology induced as a subspace of I × J (M,S1). So, in this new case, the isotopy we begin with determines a subspace of J (M,S1) from which we must choose. The ∗ operation extends to an action of D on I given by

44 −1 (g ∗ ψ)t := gt ψt, g ∈ D, ψ ∈ I.

There is also an action of D on J (M,S1) given by

g −1 1 (g ∗ J)t = Jt := Dgt JtDgt, g ∈ D,J ∈ J (M,S ).

We will to use the following two technical lemmas throughout.

Lemma 3.6 If J is an ω-compatible almost complex structure, and ρ is any diffeomorphism of M, then ρ∗J := Dρ−1JDρ is compatible with ρ∗ω.

Proof: This is a simple calculation. Since J is ω-compatible, r(·, ·) := ω(·,J·) is a Riemannian metric. So,

ρ∗ω(·, ρ∗J ·) = ω(Dρ ·, JDρ ·) = r(Dρ ·, Dρ ·) = ρ∗r(·, ·)

Since r is a metric, so is ρ∗r. We also need to check that that ρ∗ω is ρ∗J invariant.

ρ∗ω(ρ∗J·, rho∗J·) = ω(JDρ ·, JDρ ·) = ω(Dρ ·, Dρ ·) = ρ∗ω(·, ·)



Lemma 3.7 For a given g ∈ D, θ ∈ H, and J ∈ J g∗θ(M, ω, S1), we have that J g−1 ∈ J (M, ω, S1).

45 Proof: First notice that

−1 −1∗ −1 ∗ (gt θt) ω(·, ·) = θt gt ω(·, ·) ∗ −1 ∗ = gt ((θt ) ω)(·, ·) −1∗ = (θt ω)(Dgt·, Dgt·)

= ω(Dgt·, Dgt·) ∗ = gt ω(·, ·).

The second to last equality follows from the fact that θ ∈ H, and thus each θt is a symplectic diffeomorphism. g∗θ 1 −1 −1∗ We have that J ∈ J (M, ω, S ), so let rt(·, ·) = (gt θt) ω(·,Jt·) be the Riemannian metric guaranteed by this hypothesis. Then,

g−1 −1 ω(·,J ·) = ω(·, DgtJtDgt ·) ∗ −1 −1 = (gt) ω(Dgt ·,JtDgt ·) −1 −1∗ −1 −1 = (gt θt) ω(Dgt ·,JtDgt ·) −1 −1 = rt(Dgt ·, Dgt ·) −1 ∗ = (gt ) rt(·, ·).

−1 −1∗ The third equality follows from the above calculation of gt θt) ω. −1 ∗ But rt is a metric, thus so is (gt ) rt. The only fact remaining to prove is that ω is J g−1 invariant. This is just a direct calculation:

46 −1 −1 gt gt −1 −1 ω(J ·,J ·) = ω(DgtJtDgt ·, DgtJtDgt ·) ∗ −1 −1 = gt ω(JtDgt ·,JtDgt ·) ∗ −1 −1 = gt ω(Dgt ·, Dgt ·) = ω(·, ·).

This shows that J g−1 ∈ J (M, ω, S1) as promised.



The following proposition is fundamental to the construction given in section 4. It says that any element of F can be decomposed into an element of D and a Hamiltonian pair (θ, j) ∈ H × J (M, ω, S1).

Proposition 3.8 F = D ∗ (H × J (M, ω, S1))

Proof: We first prove the simpler assertion that D ∗ H × J (M, ω, S1) ⊂ F. Let g ∈ D, θ ∈ H, and J ∈ J (M, ω, S1). We need to show that g ∗ θ is almost Hamiltonian, and that J g ∈ J g∗θ(M, ω, S1). Smoothly homotope g to some h ∈ G, which exists by definition of D. This defines a smooth homotopy from g ∗ θ to h ∗ θ. But since h ∈ G, h ∗ θ ∈ H, so g ∗ θ ∈ I. g g∗θ 1 To show that J ∈ J (M, ω, S ), let rt(·, ·) be the Riemannian metric given by rt(·, ·) = ω(·,Jt·). Then,

47 −1 ∗ g −1 ∗ −1 ((g ∗ θ)t ) ω(·,Jt ·) = (θt gt) ω(·, Dgt JtDgt·) ∗ −1 = (gt) ω(·, Dgt JtDgt·)

= ω(Dgt·,JtDgt·)

= rt(Dgt·, Dgt·) ∗ = gt rt(·, ·)

∗ g We also need to show that (g ∗ θt) ω is invariant under Jt .

−1 ∗ g g ∗ g g ((g ∗ θ)t ) ω(Jt ·,Jt ·) = (gt) ω(Jt ·,Jt ·)

= ω(JtDgt·,JtDgt·)

= ω(Dgt·, Dgt·) ∗ = (gt) ω(·, ·) −1 ∗ = ((g ∗ θ)t ) ω(·, ·).

This shows that J g ∈ J g∗θ(M, ω, S1). For the other inclusion, let (ψ, J) ∈ F, and choose θ ∈ H such that ψ is homotopic to θ, relative endpoints. Such a θ exists by the definition of I. −1 Define g ∈ Ω Diffid(M) by gt = θtψt . Then g is contractible, so g ∈ D. −1 g−1 g But ψt = gt θt, so ψ = g ∗ θ. Also, since (J ) = J,(ψ, J) = g ∗ −1 −1 (θ, J gt ). By lemma 3.7, J g ∈ J (M, ω, S1), showing that F ⊂ D ∗ (H × J (M, ω, S1)).



Remark 2 Note that a decomposition ψ = g ∗ θ as in proposition 3.8 is not unique.

48 Definition 3.9 A pair (ψ, J) ∈ F will be called an almost Hamiltonian regular pair if (ψ, J) decomposes according to proposition 3.8 as (ψ, J) = g ∗ (θ, j) where (θ, j) is a Hamiltonian regular pair.

We denote the space of all almost Hamiltonian regular pairs by Freg. Thus,

Freg = D ∗ (H × J )reg.

Lemma 3.10 Freg is a dense subset of F.

Proof: Suppose (ψ, J) is an element of F. Decompose (ψ, J) according to ν ν proposition 3.8 as (ψ, J) = g ∗ (θ, j). Choose (θ , j ) ∈ (H × J )reg such that (θν, jν) converges to (θ, j). That this is possible follows from the fact  1 that H × J reg is a dense subset of H × J (M, ω, S ) (see proposition 3.4). Then the sequence g ∗(θν, jν) converges to g ∗(θ, j) = (ψ, J). Since each ν ν ν ν (θ , j ) is a (Hamiltonian) regular pair, g ∗ (θ , j ) ∈ Freg. It follows that

Freg is a dense subset of F.



3.4 The Maslov Index

Recall the Maslov index I : Ge → Z. This index described the grading shift in the automorphism of HF∗(M, ω) induced by (g, g˜) ∈ Ge. We will need to extend this index to De. In dealing with almost Hamiltonian loops, (i.e., the space De), we run into a difficulty using the definition of the index described in the first sec- tion. Since our isotopies go through Diff(M), and not just Symp(M), the path of matrices obtained in the definition of the Maslov index will not be symplectic, just invertible.

49 We get around this difficulty by showing that if two elements of Ge can be connected by a path through De, then their indexes agree. We then show that any element of De can be connected to an element of Ge, and thus the index can be extended to all of De.

Proposition 3.11 If (g0, g˜0), (g1, g˜1) ∈ Ge are such that [g0, g˜0] = [g1, g˜1] 0 0 1 1 inside π0(De), then I(g , g˜ ) = I(g , g˜ ).

Proof: First, recall the definition if the index map I in the Hamiltonian case: Choose a point c = [v, x] ∈ LgM. Trivialize the symplectic bundle ∗ 2 2n 2n (v T M, ω) as D × (R , ω0). Let At :(Tx(t)M, ω) → (R , ω0) be this trivi- alization on the boundary circle. Considerg ˜(c) = [v0, g·x] ∈ LgM. Trivialize 0 ∗  2 2n 2n (v ) T M, ω as D × (R , ω0), and let Bt :(Tgt(x(t))M, ω) → (R , ω0) be −1 the trivialization on the boundary, as before. Then BtDgt(At) gives a linear, symplectic map from R2n to R2n, and is thus given by a symplectic matrix, denoted l(t). Then {l(t)} is a loop in Sp(2n,R), and this loop is in- dependent, up to homotopy, of all the choices involved, including the choice of c ∈ LgM. Let deg : π1(Sp(2n, R)) → Z be the canonical isomorphism. The index map is given by I(g, g˜) = deg(l(t)), and I is well defined on

π0(Ge). For our proof, pick any c = [v, x] ∈ LgM. Since [g0, g˜0] = [g1, g˜1] in s s π0(De), there is a smooth path g , 0 ≤ s ≤ 1, and a smooth liftg ˜ , 0 ≤ s ≤ 1, 0 0 1 1 s 0 1 connecting (g , g˜ ) and (g , g˜ ). Choose this path so thatg ˜ =g ˜ for s ≤ 2 . 2 θ Consider D with coordinates (s, t) = (r, 2π ). (The 2π factor is just so that (s, 0) = (s, 1).) Notice that s and t are use both as parameters for diffeomorphisms and 2 s as coordinates for D . This is because we will assign the diffeomorphism gt to the point (s, t) ∈ D2, and it is convenient to have the same notation.

50 We are given v : D2 → M. Consider the pullback bundle v∗TM. We give v∗TM the structure of a symplectic fiber bundle as follows. Let ξ and η be in the fiber over the point (s, t) ∈ D2. (That is, they are elements of

Tv(s,t)M.) Give Tv(s,t)M the symplectic structure defined by

g s ∗ s s ω (ξ, η) := (g ) ω (ξ, η) = ω s (Dg ξ, Dg η). s,t t v(s,t) gt (v(s,t)) t t

This gives v∗TM the structure of a smooth, symplectic vector bundle, de- ∗ g g noted (v T M, ω ). To emphasize, ωs,t is a symplectic form on the vector space Tv(s,t)M. (The symplectic bundle is smooth at (0, 0) because we choose gs to be constant near s = 0.) Since D2 is contractible, there is a trivialization A :(v∗T M, ωg) → 2 2n g D × (R , ω0). Let As,t be the symplectic isomorphism from (Tv(s,t)M, ωs,t) 2n to (R , ω0) given by A. For clarity:

g 2n As,t :(Tv(s,t)M, ωs,t) → (R , ω0), (6)

−1 ∗ g (As,t ) ωs,t = ω0. (7)

Fix s ∈ [0, 1]. (Recall that s describes the radial coordinate of the disc.) Define xs ∈ LM by xs(t) = v(s, t). Define vs : D2 → M by vs(s0, t) = v(s · s0, t). Then vs is simply the portion of v which is inside xs, suitably reparameterized. Since vs(1, t) = v(s, t) = xs(t), we can define cs ∈ LgM by cs = [vs, xs]. Considerg ˜s(cs) ∈ LgM. Theng ˜s(cs) = [v0, gs · xs], for some v0 : D2 → 0 ∗  2 2n M. Let B : (v ) T M, ω → D × (R , ω0) be a trivialization as before. (Note that in this case, we do not adjust the symplectic structure.) Let s s ∗  1 2n Bs,t : g · x ) T M, ω → S × (R , ω0) be the trivialization restricted to the boundary circle. Then Bs,t gives an isomorphism of symplectic vector

51  2n spaces from T s s M, ω s s to ( , ω ), and in particular, gt (x (t)) gt (x (t)) R 0

2n B : T s s M → , (8) s,t gt (x (t)) R

∗ (B ) ω = ω s s . (9) s,t 0 gt (x (t))

s Finally, notice that Dgt gives an isomorphism of vector spaces between

T s M and T s s M, and x (t) gt (x (t))

s ∗  s ∗ s ∗  g (Dg ) (ω s s ) = ((g ) ω) s = (g ) ω = ω . (10) t gt (x (t)) t x (t) t v(s,t) s,t

1 s 2n 2n For s ∈ [0, 1] and t ∈ S , define lt : R → R by

s s s −1 l (t) = Bs,tDgt (x (t))As,t .

s We claim that lt is a symplectic map. To see this, notice that   s ∗ −1 ∗ s∗ ∗ (lt ) ω0 = (As,t ) Dgt (Bs,tω0) .

s ∗ It follows from equations 9, 10 and 7 that (lt ) ω0 = ω0. This shows that s lt is in fact symplectic. s s It is clear that lt is also a linear map. Thus, lt is a linear, symplectic map from R2n to R2n. Fix s ∈ [0, 1], and allow t to vary across S1. By s considering {lt} for such s and t, we obtain a loop in Sp(2n, R). Notice that the above procedure is very similar to that used to define I(g, g˜) for (g, g˜) ∈ Ge. The main difference is that we have adjusted the sym- plectic form on each fiber of v∗TM. By taking s = 0, and using c0 to define 0 0 0 0 0 0 I(g , g˜ ), we see that I(g , g˜ ) = deg({lt} ), because g is a Hamiltonian 0 loop, and so we don’t have any adjustment on ω when defining {l }t. 1 1 1 1 We claim that deg({lt} ) = I(g , g˜ ). We will use c = c = [v, x] to define

52 I(g1, g˜1). (Recall that any point in LgM will do.) In using c1 = [v, x], we ∗ 2 2n trivialize (v T M, ω) as D ×(R , ω0). In our above process, we trivialize the adjusted bundle (v∗T M, ωg). We must only show that these trivializations can be chosen so that they agree on the boundary circle, as every other step s 1 1 in defining {l }t matches the definition of I(g , g˜ ). A trivialization of (v∗T M, ω) can be given by the following process. Choose an almost complex structure on each fiber of v∗TM which varies smoothly with s and t, and such that the almost complex structure cor- responding to the fiber over (s, t) is compatible with ωv(s,t). Given this, choose an oriented, smooth frame field on v∗TM which is orthonormal with respect to the metric induced by ω and the almost complex structures. The same process can be used to trivialize (v∗T M, ωg), but the almost complex structure on Tv(s,t)M will have to be chosen so that it is compatible g with ωs,t. To accomplish this, choose any J ∈ J (M, ω). We use J to define an almost complex structure on each fiber of v∗TM as follows. On the 2 fiber over (s, t) ∈ D , (which is exactly Tv(s,t)M), supply the structure s −1 0 0 −1 s (Dgt ) Dgt J(Dgt ) Dgt . This defines a complex structure on the vector bundle v∗TM, and we will denote the complex structure on the fiber over ¯s ¯0 ¯s (s, t) by Jt . Then Jt = J for all t. It follows from lemma 3.6 that Jt is g compatible with ωs,t as required. So, let J be a choice of almost complex strucure on each fiber of v∗TM which is compatible with ω, and let ζ be an oriented frame field of v∗TM which is compatible with ω, and which is orthonomal with respect to the metric induced by ω and J. Let J g be a choice of almost complex structure on each fiber such that J g(s, t) is compatible with ωg(s, t). Choose J g such that J g(1, t) = J(1, t). This is possible because g1 is a Hamiltonian loop, g and thus ω1,t = ωv(1,t). (In fact, simply use the same procedure given in the

53 previous paragraph.) We can modify ζ so that it becomes an oriented frame field of v∗TM which is orthonormal with respect to the metric induced by ωg and J g by applying the Graham-Schmidt procedure to each basis simultaneously. Call this modified frame field ζg. This procedure will not change ζ(1, t) g g for any t because ω1,t = ωv(1,t), and J (1, t) = J(1, t), and thus ζ(1, t) is already orthonormal for ωg and J g. Since ζg(1, t) = ζ(1, t) for all t, the trivializations of (v∗T M, ωg) and (v∗T M, ω) agree on the boundary circle, as desired. s By varying s across [0, 1], lt gives a homotopy through loops in Sp(2n, R) 0 1 0 0 between lt and lt . Since deg is homotopy invariant, we see that I(g , g˜ ) = I(g1, g˜1), as desired.



Lemma 3.12 For any (g, g˜) ∈ De, there is some (h, h˜) ∈ Ge such that ˜ [g, g˜] = [h, h] inside π0(De).

Proof: Choose a path in D from g to some h ∈ G which exists by the definition of D. Since [0, 1] is contractible, the Lifting Theorem guarantees that this ˜ path has a lift to De. Then [g, g˜] = [h, h] inside π0(De).



According to these lemmas, we are justified in making the following definition.

Definition 3.13 For (g, g˜) ∈ De, let I(g, g˜) = I(h, h˜) for any (h, h˜) ∈ Ge ˜ with [g, g˜] = [h, h] inside π0(De).

54 Lemma 3.14 I : De → Z is well defined on π0(De), and I is a homomor- phism.

Proof:

The fact that I is well defined on π0(De) is clear from it’s definition. This combined with the fact that I| is a homomorphism (lemma 1.28) π0(Ge) proves the second statement.



To summarize up to this point, we have defined the group D of almost Hamiltonian loops, and it’s extension De, which allows a lift of the action to LgM. There is an index map I : π0(De) → Z which reduces to Seidel’s

Maslov index on π0(Ge). We have introduced the space F which is a bundle over the space of al- most Hamiltonian isotopies. The fiber over ψ ∈ I is given by J ψ(M, ω, S1). We have seen that any (ψ, J) ∈ F) decomposes as (ψ, J) = g ∗ (θ, j) for some (θ, j) ∈ (H × J (M, ω, S1) and some g ∈ G, although the decomposi- tion is not unique.

3.5 LgM Revisited

Recall that the space LgM depends on ω. The equivalence relation on LgM 0 0 0 0 0 is [v, x] = [v , x ] ⇐⇒ x = x , ω(v#v ) = c1(v#v ) = 0. In the original construction of the Floer homology, this is well defined for (H,J). But in our case, ω is not fixed, and Jt varies outside of J (M, ω). We need to see that LgM is well defined even for (ψ, J) ∈ F.

∗ Proposition 3.15 For (ψ, J) ∈ F, [ψt ω] = [ω] for all t, and the Chern classes satisfy c1(TM,Jt) = c1(TM,J0) = c1(ω).

55 Before proving the proposition, we prove several lemmas.

Lemma 3.16 Let φ ∈ Diff(M) and J ∈ J (M, ω). Then φ∗J := Dφ−1JDφ is an almost complex structure compatible with φ∗ω, i.e., φ∗J ∈ J (M, φ∗ω).

Proof: Because J is ω -compatible, r(·, ·) := ω(·,J·) is a Riemannian metric and ω(J·,J·) = ω(·, ·). We have:

φ∗ω(·, φ∗J·) = φ∗ω(·, Dφ−1JDφ·) = ω(Dφ·, JDφ·) = r(Dφ·, Dφ·) = φ∗r(·, ·).

Since r is a metric, so is φ∗r. Also,

φ∗ω(φ∗J·, φ∗J·) = φ∗ω(Dφ−1JDφ·, Dφ−1JDφ·) = ω(JDφ·, JDφ·) = ω(Dφ·, Dφ·) = φ∗ω(·, ·).

This shows that φ∗J is a φ∗ω -compatible almost complex structure. 

Recall that c1(ω) is defined for any symplectic structure by choosing any J ∈ J (M, ω), and considering the complex vector bundle (TM,J), and these Chern classes are independent of the choice of J.

56 ∗ ∗ Lemma 3.17 For φ ∈ Diff(M), c1(M, φ ω) = φ c1(M, ω).

Proof: ∗ ∗ ∗ ∗ By lemma 3.16, φ J is φ ω -compatible, so c1(M, φ ω) = c1(T M, φ J). Consider the following diagram:

φ∗(TM,J)(TM,J)   φ∗π π y y M −−−→ M φ Here φ∗(TM,J) denotes the pulled back bundle over M induced by φ. ∗ ∗ By the naturality of the Chern classes, c1(φ (TM,J)) = φ c1(TM,J) (see [17]). ∗ ∗ We need only show that φ (TM,J) w (T M, φ J) as complex vector bundles over M, because given this,

∗ ∗ c1(M, φ ω) = c1(T M, φ J) ∗ = c1(φ (TM,J)) ∗ = φ c1(TM,J) ∗ = φ c1(M,J).

To accomplish this, we exhibit a fiber preserving complex linear map between the bundles that induces the identity on the base space M. ∗ ∗ −1 The fiber over the point p in (φ (TM,J)) is φ π (p) = (Tφ(p)M,J). ∗ −1 ∗ The fiber over the point p in (T M, φ J) is π (p) = (TpM, φ J). Simply ∗ −1 −1 −1 map φ π (p) to π (p) by Dφp : Tφ(p)M → TpM. This is an isomorphism of (real) vector spaces because φ is a diffeomorphism. We must show that

57 the map respects the complex structure on each side:

−1 −1 −1 Dφp (Jx) = Dφp JDφpDφp x ∗ −1 = φ J(Dφp x)

This shows that the map between the bundles is complex linear, and it clearly induces the identity on the base, so it is an isomorphism of com- plex vector bundles. Using the naturality of the Chern class again gives ∗ ∗ that c1(φ (TM,J)) = c1(T M, φ J). Because we discovered above that ∗ ∗ c1(φ (TM,J)) = φ c1(TM,J), the lemma is proved.



Proof of proposition 3.15: Since each ψt is isotopic to the identity, the induced map on cohomology is the identity. The statement about the Chern classes follows from this fact, along with lemmas 3.16 and 3.17.



∗ It follows from proposition 3.15 that for ψ = ψt ∈ I, ω(A) = ψt ω(A) for all A ∈ π2(M), and c1(A) is independent of J, so LgM is well defined.

58 4 Defining HF∗(M, ω, ψ, J)

In section 2.2, we isolated the ingredients needed to construct the Λ-module

HF∗(M, ω, H, J). In fact, we need only P(H) (along with the C-Z index), the action functional aH , and the ∂H,J operator. In this section, we define

HF∗(M, ω, ψ, J) for (ψ, J) ∈ Freg. If (ψ, J) is Hamiltonian (meaning ψ ∈ ∞ H), then HF∗(M, ω, ψ, J) reduces to HF∗(M, ω, H, J), where H ∈ C (M × S1) is a generating function for ψ. Recall that I is the space of 1-periodic almost Hamiltonian isotopies, and it is convenient to consider F as a bundle over I, with fiber over ψ given by J ψ(M, ω, S1) ⊂ J (M,S1).

4.1 Defining Pg(ψ)

For ψ = (ψt) ∈ I, let P(ψ) consist of the contractible loops in M which arise by flowing along the isotopy ψ. That is,

ψ P(ψ) = {x ∈ LM | x˙(t) = Xt (x(t))}.

By considering the definition of P(H) (see section 1.2), it becomes clear that P(ψ) = P(H) if ψ is a Hamiltonian isotopy generated by H ∈ C∞(M × S1).

As before, it is necessary to construct HF∗(M, ω, ψ, J) using the covering space LgM, so we set P](ψ) = p−1(P(ψ)), where p : LgM → LM is the projection. Again, P](ψ) = P^(H) if ψ is Hamiltonian, generated by the Hamiltonian function H. We wish to use the set P](ψ) to create a chain complex, but first we

59 must index the set. Indexing P](ψ) is a bit tricky. For a Hamiltonian isotopy θ, this is done using the Conley-Zehnder index µθ : P](θ) → Z. If ψ is an almost Hamiltonian isotopy, the C-Z index does not make sense. It is obtained by linearizing ψ along the path x(t) for x ∈ P(ψ), and using a symplectic trivialization to create a path of symplectic matrices. The problem in our setup is that this process will not result in a path of symplectic matrices, only invertible matrices, because ψt will not necessarily preserve ω. First recall that the Conley-Zehnder index for Hamiltonian isotopies is only well-defined when the isotopy is non-degenerate. (This means that for all x ∈ P(θ), det(1 − Dθ1) 6= 0.) While the non-degeneracy condition was given for Hamiltonian isotopies, it makes sense for arbitrary isotopies.

Definition 4.1 An isotopy ψ ∈ I will be called non-degenerate if, for all x ∈ P(ψ), det(1 − Dψ1) 6= 0.

We begin by defining the index for non-degenerate Hamiltonian isotopies by using it’s generating function. That is, for θ ∈ H and c ∈ P(H), let

H µθ(c) := µH (c), θ ∈ H, θ = θ .

The next theorem is the first in a series in which we decompose ψ ∈ I as ψ = g ∗ θ according to proposition 3.8. Each time, a decomposition is chosen, and used to define some object. We then show that the construction is independent of the choice of decomposition. We will often need to choose a liftg ˜, so that (g, g˜) ∈ De. Recall that we have extended the Maslov index to a map I : De → Z.

Proposition 4.2 Let ψ = (ψt) ∈ I, and decompose as ψ = g ∗ θ as in proposition 3.8. Suppose that ψ is a non-degenerate isotopy. Choose g˜ such

60 that (g, g˜) ∈ De. Then µψ : P](ψ) → Z is well defined by

µψ(c) = µθ(˜g(c)) − 2I(g, g˜), independent of the choice of decomposition or the choice of lift g˜. Moreover, if ψ is Hamiltonian generated by H, then µψ = µH .

The proof will be given after we establish some initial facts. First notice that we haven’t yet shown thatg ˜(c) ∈ P](θ), as this formula requires. In the Hamiltonian case, this is given by lemma 1.26. The next lemma gives the proof in the almost Hamiltonian case. Notice that lemma 1.26 is proved by considering the action functional. The proof of the next lemma is in purely topological terms. Recall that P(ψ) is contained in LM, and any g ∈ D induces a map (still called g) from LM to LM.

Lemma 4.3 For all (g, g˜) ∈ De and ψ ∈ I, the following equalities hold.

g(P(g ∗ ψ)) = P(ψ) (11)

g˜(P(^g ∗ ψ)) = P](ψ). (12)

Proof: We need only prove equation 11, because then equation 12 follows from the fact thatg ˜ lifts the action of g. We prove that g(P(g ∗ ψ)) ⊂ P(ψ). Given x ∈ P(g ∗ ψ), let xg = g · x. g ψ g  We need to show thatx ˙ (t) = Xt (x )(t) . −1 Because (g ∗ ψ)t = gt ψt, lemma 2.3 gives the following two equations:

61 g∗ψ g−1 −1 ψ −1 g −1 ψ Xt = Xt + Dgt Xt ◦ gt = −Dgt Xt ◦ gt + Dgt Xt ◦ gt, (13)

dg x˙g(t) = Dg (x ˙(t)) + t (x(t)). (14) t dt By definition, dg t (x(t)) = Xg(g (x(t))). (15) dt t t Because x ∈ P(g ∗ ψ),

g∗ψ x˙(t) = Xt (x(t)). (16)

Substituting equations 13, 15, and 16 into equation 14 gives

dg x˙ g(t) = Dg (x ˙(t)) + t (x(t)) t dt g∗ψ g  = DgtXt (x(t)) + Xt gt(x(t))   −1 g −1 ψ g  = Dgt −Dgt Xt (gt(x(t))) + Dgt Xt (gt(x(t))) + Xt gt(x(t))

g  ψ  g  = −Xt gt(x(t)) + Xt gt(x(t)) + Xt gt(x(t)) ψ g = Xt ((x )(t)),

as desired. The same manipulations show the other inclusion.



Now that the formula at least makes sense, we show that the formula given in proposition 4.2 reduces to the Conley-Zehnder index in the Hamil- tonian case.

62 Recall that given a regular pair (H,J), we have defined the pullback pair (Hg,J g), and the C-Z index satisfies

µHg (c) = µH (˜g(c)) − 2I(g, g˜). (17)

Section 2.1 was devoted to calculating the isotopy generated by Hg. We found that θHg = g ∗ θH . (18)

Lemma 4.4 For (g, g˜) ∈ G and θ = θt ∈ H, µg∗θ(c) = µθ(˜g(c)) − 2I(g, g˜).

Proof:

Recall that for θ ∈ H, µθ is well defined by µθ = µH for any H ∈ C∞(M × S1) which generates θ. The lemma then is just the formula given in equation 17, with the functions converted to isotopies, using equation 18.



We also need the following lemma to prove proposition 4.2.

Lemma 4.5 Given θ ∈ H and g ∈ D, g ∗ θ ∈ H ⇐⇒ g ∈ G.

Proof: First, the reverse implication. Let Hθ generate θ, and Hg generate g. θ g Define a Hamiltonian function by Ht = Ht ◦ gt − Ht ◦ gt. Then H generates g ∗ θ by proposition 2.5. For the forward implication, let Hθ generate θ, and Hg∗θ generate g ∗ θ. g θ g∗θ −1 Define a Hamiltonian function by (K )t := Ht − Ht ◦ gt . We claim that Kg generates g, so that g is actually Hamiltonian, and thus g ∈ G.

63 g To see this, we need only show that X g = X . First notice that by Kt t g∗θ θt −1 g −1 g θ g∗θ −1 lemma 2.3, Xt = X ◦ gt − Xt ◦ gt , so Xt = Xt − Xt ◦ gt . It then follows that g θ g∗θ −1 Xt = Xt − Xt ◦ gt (19)

Simply plug these into the definition of X g : Kt

X g = X θ g∗θ −1 Kt (H −H ◦g )t

= X θ − X g∗θ −1 Ht (H ◦g )t

θ (g∗θt) −1 = Xt − X ◦ gt g = Xt .

Thus, g is Hamiltonian generated by Kg.



In subsequent sections, we will use the preceding lemma often in con- junction with proposition 3.8. It allows us to prove that many things are independent of the choice of decomposition.

Proof of proposition 4.2: Let ψ be in I, and choose two decompositions

ψ = g0 ∗ θ0 = g1 ∗ θ1 as in proposition 3.8. Letg ˜0 andg ˜1 be lifts of g0 and g1 respectively, so that (g0, g˜0) and (g1, g˜1) are in De. Since g0 ∗ θ0 = g1 ∗ θ1, we have that 0 −1 0 1 −1 1 0 0 1 −1 1 (gt ) θt = (gt ) θt , and so θt = gt (gt ) θt . Thus, if we define a loop of

64 1 0 −1 diffeomorphisms of M byg ¯t = gt (gt ) , we have

1 −1 1 (¯g ∗ θ )t = (¯gt) θt 1 0 −1 −1 1 = (gt (gt ) ) θt 0 1 −1 1 = gt (gt ) θt 0 = θt .

This means thatg ¯ ∗ θ1 = θ0. But since θ0 and θ1 are Hamiltonian, so isg ¯, by lemma 4.5. Notice also thatg ˜1(˜g0)−1 is a lift of the action ofg ¯ to LgM. Then,

0 0 µθ0 (˜g (c)) = µg¯∗θ1 (˜g (c)) 1 0 −1 0  1 0 −1 = µθ1 (˜g (˜g ) g˜ )(c) − 2I(¯g, g˜ (˜g ) ) 1 1 1 0 0  = µθ1 (˜g (c)) − 2 I(g , g˜ ) − I(g , g˜ ) .

The second equality follows from the fact thatg ¯ is Hamiltonian, and lemma 4.4. The third equality follows from the fact that I is a homomor- phism, proved in lemma 3.14.



This shows that the definition of the index is independent of the choices, and proves proposition 4.2. Now we may index P](ψ) for non-degenerate ψ ∈ I by setting

P](ψ)k = {c ∈ P](ψ)|µψ(c) = k}.

Lemma 4.6 If ψ = g ∗ θ is a decomposition of a non-degenerate isotopy

ψ ∈ I as in proposition 3.8, then g˜(P](ψ)k) = P](θ)k−2I(g,g˜).

Proof:

65 By lemma 4.3, the only thing left to prove is the formula for the index change. This fact follows from the formula given in proposition 4.2.



4.2 The Chain Complex and Boundary Operator

We now consider a fixed (ψ, J) ∈ F, and begin to define HF∗(M, ω, ψ, J). Notice that in the Hamiltonian case, the definition of the chain complex de- pends on the action functional, aH . (It is part of the finiteness condition on the chains.) In the almost Hamiltonian case, there is no action functional.

To overcome this difficulty, decompose ψ = (ψt) using proposition 3.8 as ψ = g ∗θ, and chooseg ˜, a lift of g, so that (g, g˜) is in De. In the Hamiltonian ∗ case, according to lemma 2.3 of [21], aHg =g ˜ aH + (constant). Combined ∗ with proposition 2.5, this suggests that we useg ˜ aHθ as a replacement for the action functional in the definition of the chain complex, where Hθ denotes a Hamiltonian function generating θ.

We define CFk(M, ω, ψ) as all formal sums

X mchci,

c∈P](ψ)k

where mc ∈ Z2, which satisfy the finiteness condition {c ∈ P](ψ)k | mc 6= ∗ 0, g˜ aHθ (c) > C} is finite for all C ∈ R.

Lemma 4.7 CFk(M, ω, ψ) is well defined independent of the choice of de- composition of ψ.

Proof:

66 Choose two decompositions of ψ as in proposition 3.8:

ψ = g0 ∗ θ0 = g1 ∗ θ1.

0 1 0 1 −1 Also choose liftsg ˜ andg ˜ . Defineg ¯ ∈ D byg ¯t = gt (gt ) . Theng ¯ is Hamiltonian by lemma 4.5. Let Hθ0 and Hθ1 generate the isotopies θ0 and θ1 respectively. Then (Hθ0 )g¯ = Hg¯∗θ0 = Hθ1 by proposition 2.5. This implies that

0 ∗ 0 1 −1 1∗ (˜g ) aHθ0 = (˜g ◦ (˜g ) ) ◦ g˜ aHθ0 1 ∗ 0 1 −1 ∗ = (˜g ) (˜g ◦ (˜g ) ) aHθ0 ) 1 ∗  = (˜g ) a(Hθ0 )g¯ + (constant) 1 ∗  = (˜g ) aHg¯∗θ0 + (constant) 1 ∗ = (˜g ) (aHθ1 ) + (constant)

The third equality follows from the fact thatg ¯ is Hamiltonian. It then follows that the finiteness condition is independent of the choice of decomposition.



L H We then set CF∗(M, ω, ψ) = CFk(M, ω, ψ). It is clear that for θ ∈ H H, CF∗(M, ω, θ ) reduces to CF∗(M, ω, H).

As in the Hamiltonian case, CF∗(M, ω, ψ) has the structure of a Λ- module, and the dimension of CFk(M, ω, ψ) as a module over Λ is exactly the number of elements in P(ψ) with index µψ(c) = k(mod 2N).

67 Next, we define a boundary operator on CF∗(M, ω, ψ), using a choice of J ∈ J ψ(M, ω, S1). The boundary operator is given in terms of the ∂ operator. 1 ∞ ∗ Recall that for smooth u : R × S → M, ∂H,J (u) ∈ C (u TM) is given ∂u ∂u by ∂H,J (u)(s, t) = ∂s (s, t) + Jt(u(s, t)) ∂t (s, t) − ∇Ht(u(s, t)). We would like to define an analogous operator in the almost Hamiltonian case, but we have no function H to use. To get around this difficulty, we use the following standard fact.

Lemma 4.8 For J ∈ J (M, ω) and H ∈ C∞(M), let ∇H denote the gradi- ent vector field of H using the metric r(·, ·) = ω(·,J·). Then ∇H = JXH .

Proof: By definition, dH(·) = r(∇H, ·) = ω(∇H,J·) = ω(J∇H, ·). Also by definition, dH(·) = iXH ω(·) = ω(XH , ·). This shows that ω(J∇H, ·) =

ω(XH , ·), and so the lemma follows from the non-degeneracy of ω.



ψ Since XHt = Xt if ψ is a Hamiltonian isotopy generated by H, we may extend the ∂ operator in the following way. For all pairs (ψ, J) ∈ F and smooth maps u : R × S1 → M, define a ∞ ∗ section ∂ψ,J (u) ∈ C (u TM) by

∂u ∂u  ∂ (u)(s, t) = (s, t) + J (u(s, t)) (s, t) − Xψ(u(s, t)) . (20) ψ,J ∂s t ∂t t

It follows from lemma 4.8 that if ψ is Hamiltonian, generated by H, then ∂ψ,J = ∂H,J . This allows us to define the space of “connecting orbits” between two chain complex generators. For (ψ, J) ∈ F and c−, c+ ∈ P](ψ), we define

68 1 M(c−, c+, ψ, J) as the set of all smooth maps u : R × S → M such that

∂ψ,J (u) = 0 and there is a liftu ˜ : R → LgM with limits c−, c+. As in the Hamiltonian case, we would like to use these spaces to define a boundary operator on CF∗(M, ω, H). In particular, if c± ∈ P](ψ) are such that µψ(c−) − µψ(c+) = 1, we would like to be able to count the number of elements of M(c−, c+,H,J)/R, where the R-action is again translation in the s variable. We will only be able to count these spaces for generic pairs (ψ, J) ∈ F.  Recall that Freg = D ∗ (H × J )reg, and that g P(g ∗ θ) = P(θ) for all g ∈ D and θ ∈ H.

Proposition 4.9 Let (θ, j) ∈ (H×J )reg, and let (g, g˜) ∈ De. Given c− and g c+ in P^(g ∗ θ), there is a bijection between the spaces M(c−, c+, g ∗ θ, j )/R and M(˜g(c−), g˜(c+), θ, j)/R.

Proof:

For u ∈ M(˜g(c−), g˜(c+), θ, j) with liftu ˜ : R → LgM, definev ˜ : R → LgM −1 byv ˜(s) =g ˜ (˜u(s)). This map clearly satisfies lims→±∞ v˜(s) = c±. Because u ∈ M(˜g(c−), g˜(c+), θ, j), ∂θ,j(u) = 0. We prove ∂ψ,J (v) = 0. −1 Because v(s, t) = gt (u(s, t)), the chain rule (lemma 2.3) gives the fol- lowing identities:

∂v ∂u (s, t) = Dg−1 (s, t) ∂s t ∂s ∂v ∂g−1 ∂u (s, t) = t (u(s, t)) + Dg−1 (s, t) ∂t ∂t t ∂t gθ g−1 −1 θ  Xt (v(s, t)) = Xt (v(s, t)) + Dgt Xt gt(v(s, t)) .

69 Substitution into the definition of ∂g∗θ,jg (v) gives

  ∂v g ∂v g∗θ ∂ g (v)(s, t) = (s, t) + j (v(s, t)) (s, t) − X (v(s, t)) g∗θ,j ∂s t ∂t t ∂u  ∂u = Dg−1 (s, t) + Dg−1j Dg (v(s, t)) Dg−1 (s, t) t ∂s t t t t ∂t g−1 −1 −1 θ + Xt (gt (u(s, t))) − Dgt Xt (u(s, t))  g−1 −1 − Xt (gt (u(s, t))) ∂u ∂u  = Dg−1 (s, t) + j (s, t) − Xθ(u(s, t)) t ∂s t ∂t t −1 = Dgt ∂θ,j(u)(s, t).

Since ∂θ,j(u) = 0 by assumption, ∂g∗θ,jg (v) = 0 as well, proving the claim. g This map from M(˜g(c−), g˜(c+), θ, j) to M(c−, c+, g ∗ θ, j ) is injective because it is induced by a homeomorphism of LgM. It has an inverse in- duced by the homeomorphismg ˜−1, and so is a bijection. Because the map is equivariant under the R action, it also induces a bijection from g M(˜g(c−), g˜(c+), θ, j)/R to M(c−, c+, g ∗ θ, j )/R.



Corollary 4.10 For (ψ, J) ∈ Freg and c−, c+ ∈ P](ψ), if µψ(c−)−µψ(c+) =

1, then M(c−, c+, ψ, J)/R is finite. Proof: Decompose (ψ, J) as (ψ, J) = g ∗ (θ, j), and choose a liftg ˜ so that

(g, g˜) ∈ De. Then (θ, j) is a Hamiltonian regular pair. Since µψ(c−) −

µψ(c+) = 1, it follows from 4.5 that µθ(˜g(c−)) − µθ(˜g(c+)) = 1 as well.

Since (θ, j) is regular, M(˜g(c−), g˜(c+), θ, j) is finite. Thus, by proposition

4.9, M(c−, c+, ψ, J) is finite as well.

70 

We define a boundary operator on CF∗(M, ω, ψ) by counting these spaces. First, for c− ∈ P](ψ)k, define

X  ∂k(ψ, J)(hc−i) = # M(c−, c+, ψ, J)/R hc+i. (21)

c+∈P](ψ)k−1

Again, # means counting modulo 2. It is perhaps not obvious that this formal sum lies in CFk−1(M, ω, ψ). In addition, to use this as our boundary operator, it must extend to all of CFk(M, ω, ψ). This difficulty is P the same as in the Hamiltonian case - for a k-chain κ = mc− hc−i, c−∈P](ψ)k P ∂k(κ) is a formal sum mc+ hc+i, where the coefficient of c+ ∈ c+∈P](ψ)k−1

P](ψ)k−1 is given by the number of orbits (modulo time shift) connecting any c− ∈ P](ψ)k with mc− 6= 0 to c+. The above map wouldn’t extend to CFk(M, ω, ψ) if the number of such orbits could be infinite. In the Hamiltonian case, this is finite because the finiteness condition on the chains guarantees a uniform bound on the energy of all such connecting orbits.

Proposition 4.11 For (ψ, J) ∈ Freg, the map given by equation 21 ex- tends linearly to define a boundary operator ∂k(H,J): CFk(M, ω, ψ) →

CFk−1(M, ω, ψ).

Proof: P Let κ = mc− hc−i be a formal sum which lies in CFk(M, ω, ψ). c−∈P](ψ)k Decompose (ψ, J) as (ψ, J) = g ∗ (θ, j) as in proposition 3.8, and choose a liftg ˜ so that (g, g˜) ∈ De.  Sinceg ˜ P](ψ)k = P](θ)k−2I(g,g˜), we can applyg ˜ to each term of κ, and obtain a formal sum of elements of P](θ)k−2I(g,g˜). We will denote this formal P sum byg ˜(κ). Theng ˜(κ) = mc+ hc+i. c+∈P](θ)k−2I(g,g˜)

71 In fact, g(κ) lies in CFk−2I(g,g˜)(M, ω, θ). To see this, let C ∈ R. We need to show that {c+ ∈ P](θ)k−2I(g,g˜) | mc+ 6= 0, aHθ (c+) > C} is finite. But this is the same as {g˜(c−) ∈ P](θ)k−2I(g,g˜) | mg˜(c−) 6= 0, aHθ (˜g(c−))}. This set can ∗ be identified with {c− ∈ P](ψ)k | mc− 6= 0, g˜ aH (c−) > C}. This latter set is

finite because κ lies in CFk(M, ω, ψ).

As described above, ∂k(ψ, J)(κ) is a formal sum of elements in P](ψ)k−1, and the coefficient of c+ ∈ P](ψ)k−1 is given by the number of orbits (modulo time shift) connecting any c− ∈ P](ψ)k with mc− 6= 0 to c+. The key point is thatg ˜ again induces a bijection between this set and the corresponding set with ψ repaced by θ. Since (θ, j) is a regular pair, and because g(κ) ∈

CFk−2I(g,g˜)(M, ω, θ), this is in fact finite.

This shows that none of the coeffiecients in ∂k(ψ, J)(κ) is infinity, so it is actually a formal sum of elements in P](ψ)k−1.

To see that ∂k(ψ, J)(κ) actually lies in CFk−1(M, ω, ψ), we will show that −1  ∂k(ψ, J)(κ) =g ˜ ∂k−2I(g,g˜)(θ, j)(˜g(κ)) . This again follows essentially from proposition 4.9, because the boundary operators are defined by counting the spaces of connecting orbits.

We have seen thatg ˜ does in fact induce a map from CFk(M, ω, ψ) to  CFk−2I(g,g˜)(M, ω, θ). We show thatg ˜ ∂k(ψ, J)hc−i = ∂k−2I(g,g˜)(θ, j)hg˜(c−)i for all c− ∈ P](ψ)k. By definition,

 g  g˜ ∂k(ψ, J)hc−i =g ˜ ∂k(g ∗ θ, j )hc−i X g  = # M(c−, c+, g ∗ θ, j )/R hg˜(c+)i.

c+∈P^(g∗θ)k−1

By proposition 4.9,

g   # M(c−, c+, g ∗ θ, j )/R = # M(˜g(c−), g˜(c+), θ, j)/R .

72 It then follows that

 X  g˜ ∂k(ψ, J)hc−i = # M(˜g(c−), g˜(c+), θ, j)/R hg˜(c+)i.

c+∈P^(g∗θ)k−1

Now, by definition,

X 0  0 ∂k−2I(g,g˜)(θ, j)hg˜(c−)i = # M(˜g(c−), c , θ, j)/R hc i. 0 c ∈P](θ)k−2I(g,g˜)−1

By lemma 4.6,g ˜(P](ψ)k) = P](θ)k−2I(g,g˜), so this last sum can be written as

X  # M(˜g(c−), g˜(c+), θ, j)/R hg˜(c+)i,

c+∈P](ψ)k−1 and this agrees withg ˜ ∂k(ψ, J), as claimed.  This shows thatg ˜ ∂k(ψ, J)hc−i = ∂k−2I(g,g˜)(θ, j)hg˜(c−)i for all c− ∈

P](ψ)k. A similar calculation shows that the formula continues to hold for infinite sums in CFk(M, ω, ψ). Notice that since (θ, j) is a Hamiltonian regular pair,

∂∗(θ, j) gives a well defined boundary operator on CF∗(M, ω, θ). (This was used in the above proof.)



The proof of the preceding proposition contains an important fact that we separate for later use.

Proposition 4.12 Let (ψ, J) ∈ Freg, and let (ψ, J) = g ∗ (θ, j) as in  proposition 3.8. Then g˜ ∂k(ψt,Jt)(κ) = ∂k−2I(g,g˜)(θ, j)(˜g(κ)) for all chains

73 κ ∈ CFk(M, ω, ψ). Here, g˜ represents the map from CF∗(M, ω, ψ) to

CF∗(M, ω, θ) induced by hci → hg˜(c)i.



Corollary 4.13 For (ψ, J) ∈ Freg, the boundary operator on CF∗(M, ω, ψ) satisfies ∂2 = 0.

Proof: Decompose (ψ, J) as g ∗ (θ, j), and chooseg ˜ such that (g, g˜) ∈ De. Since 2 (ψ, J) ∈ Freg, we have that (θ, j) ∈ (H × J )reg, and thus ∂(θ, j) = 0. So,

−1  −1  ∂k−1(ψ, J) ◦ ∂k(ψ, J) = g˜ ∂k−1−2I(g,g˜)(θ, j)g ˜ ◦ g˜ ∂k−2I(g,g˜)(θ, j)g ˜ −1 =g ˜ ∂k−1−2I(g,g˜)(θ, j) ◦ ∂k−2I(g,g˜)(θ, j)g ˜ = 0.

The first equality follows from the previous proposition, and the third equality follows from the above remarks.



4.3 The New HF∗(M, ω)

We can now define HF∗(M, ω, ψ, J) for (ψ, J) ∈ Freg as the homology of  the chain complex CF∗(M, ω, ψ), ∂∗(ψ, J) .

Remark 3 Notice that the definition of HF∗(M, ω, ψ, J) doesn’t depend on the decomposition ψ = g∗θ, but showing that the definition makes sense did use such a decomposition.

To complete the new construction of HF∗(M, ω), we will show that

HF∗(M, ω, ψ, J) is naturally independent of the choice of (ψ, J) ∈ Freg. For

74 notational convenience, throughout this section, we will sometimes write

HF∗(ψ, J) for HF∗(M, ω, ψ, J). (This is really just because commutative diagrams don’t fit on the page with the extended notation.) Before we prove the existence of these continuation isomorphisms, we show that HF∗(M, ω, ψ, J) does in fact always recover the standard Floer homology, HF∗(M, ω).

Proposition 4.14 For (ψ, J) ∈ Freg, and (g, g˜) ∈ De, the map hci → hg˜(c)i g induces an isomorphism from HF∗(M, ω, g ∗ ψ, J ) to HF∗(M, ω, ψ, J).

Proof: If follows from proposition 4.12 that multiplication byg ˜ induces a chain   map from CF∗(M, ω, ψ, J), ∂∗(ψ, J) to CF∗(M, ω, θ, j), ∂∗(θ, j) , and as such, it induces a homomorphism on the homology level. This homomor- phism is actually an isomorphism with inverse induced by multiplication by −1 g˜ , so HF∗(M, ω, ψ, j) w HF∗(M, ω, θ, j). But since (θ, j) is a Hamiltonian regular pair, HF∗(M, ω, θ, j) w HF∗(M, ω).



To complete the new construction of HF∗(M, ω), we need to see that

HF∗(M, ω, ψ, J) is naturally independent of the choice of (ψ, J) ∈ Freg. Before we give the main ideas involved, notice that we have been able to bypass the usual analytic difficulties involved in Floer homology. In particular, when ψ is a Hamiltonian isotopy, corollary 4.10 follows from two basic facts about the ∂ψ,J operator: suitably interpreted, it is an elliptic operator, and solutions to ∂ψ,J = 0 with fixed endpoints have uniformly bounded energy (once we lift to LgM). It follows from the first fact that the spaces M(c−, c+,H,J) are manifolds, and it follows from the second fact that in the 0-dimensional case, these manifolds are compact. We have

75 used a topological trick to reduce to the Hamiltonian case in order to prove corollary 4.10 when ψ is merely almost Hamiltonian. In the standard construction, an isomorphism between the Floer homol- α α β β ogy groups HF∗(M, ω, H ,J ) and HF∗(M, ω, H ,J ) is given via a choice of (generic) path of pairs in C∞(M ×S1)×J (M, ω, S1) connecting (Hα,J α) and (Hβ,J β). In our situation, this corresponds to choosing a path in F between two regular pairs (ψα,J α) and (ψβ,J β).

Definition 4.15 A homotopy of regular pairs between (ψα,J α) and β β α α (ψ ,J ) in Freg is a smooth map Φ: R → F such that Φ(s) = (ψ ,J ) for s ≤ −1 and Φ(s) = (ψβ,J β) for s ≥ 1.

Given any two regular pairs (ψα,J α) and (ψβ,J β), there is always a homotopy between them. (This is because we can always homotope ψα and ψβ to become Hamiltonian isotopies.) Given such a homotopy, we will write Φ(s) = (ψs,J s). For a homotopy 1 ∞ ∗ Φ and a smooth map u : R×S → M, define a section ∂Φ(u) ∈ C (u TM) by

∂u ∂u s ∂ (u)(s, t) = (s, t) + J s(u(s, t)) (s, t) − Xψ (u(s, t)). (22) Φ ∂s t ∂t t

We would like to use these spaces to define a chain map between the α β complexes CF∗(M, ω, ψ ) and CF∗(M, ω, ψ ) by counting the solutions to this equation. For cα ∈ P^(ψα) and cβ ∈ P^(ψβ), define M(cα, cβ, Φ) as the 1 set of all smooth maps u : R × S → M which satisfy ∂Φ(u) = 0, and which lift to a mapu ˜ : R → LM with limits cα and cβ. Equation 22 looks very similar to equation 20. The only difference is that we allow the almost complex structure and the vector field to vary with s. We showed above that solutions to equation 20 can be counted (modulo

76 time shift). This suggests that they have the same analytic properties as the solutions to the corresponding Hamiltonian equations (that is, when ψ the vector fields Xt are Hamiltonian).

α α β β Theorem 4.16 For (ψ ,J ) and (ψ ,J ) in Freg, there exists a natural α,β α α β β isomorphism Φ : HF∗(M, ω, ψ ,J ) → HF∗(M, ω, ψ ,J ). (By natural, we mean that the existence of the isomorphism depends on some choices, but the actual isomorphism is independent of the choices involved.) Moreover, these isomorphisms are functorial in the sense that Φα,β ◦ Φγ,α = Φγ,β. In addition, if (ψα,J α) and (ψβ,J β) are Hamiltonian regular pairs, then this isomorphism reduces to the standard continuation isomorphism.

Proof: Choose decompositions (ψα,β,J α,β) = gα,β ∗(θα,β, jα,β) as in thereom 3.8. Make these choices so that gα and gβ are contractible, and choose liftsg ˜α andg ˜β so that (gα, g˜α) and (gβ, g˜β) are in the component of the identity in De. Choose Φ¯ a (Hamiltoniona) homotopy of regular pairs between (θα, jα) and (θβ, jβ). Chooseg ˜s a path in De which satisfiesg ˜s =g ˜α for small enough s, andg ˜s =g ˜β for large enough s. Define Φ a homotopy of regular pairs from (ψα,J α) to (ψβ,J β) by Φ(s) = gs ∗ (Φ(¯ s). Let cα,β ∈ P^(ψα,β). We will show that for carefully chosen Φ, there is a bijection between M(cα, cβ, Φ) and M(˜gα(cα), g˜β(cβ), Φ)).¯ This is essen- tially a “stretching the neck” argument. First notice that if u : R×S1 → M 1 satisfies ∂Φ¯ (u) = 0, and if we define v : R × S → M by v(s, t) = s s −1 s −1 ∂gt  (gt ) (u(s, t)), then ∂Φ(v)(s, t) = (Dgt ) ∂s (u(s, t)) . We want to guar- 0 antee that ∂Φ(v) is small, so that we can assign to it a true solution v with 0 ∂Φ(v ) = 0 using the implicit function theorem. The only factor contribut- s ∂gt ing to the non-vanishing of ∂Φ(v) is the term ∂s . We can guarantee that this term is arbitrarily small by allowing it to vary over an arbitrarily large

77 compact interval during the homotopy. This argument, along with the compactness of M, shows that we can s s −1 ∂gt  find a uniform bound on (Dgt ) ∂s (u(s, t)) , and we can make this bound arbitrary small. By applying the implicit function theorem, we see that there is a bijection between M(cα, cβ, Φ) and M(˜gα(cα), g˜β(cβ), Φ)¯ for such ¯ α β α β homotopies Φ. Thus, if µψα (c ) = µψβ (c ), then M(c , c , Φ) is finite. We have now established that we can always find a homotopy of regular pairs Φ from (ψα,J α) to (ψβ,J β) such that for any pair (cα, cβ) ∈ P^(ψα) × β α β α β P^(ψ ) with µψα (c ) = µψβ (c ), it is the case that M(c , c , Φ) is a finite α,β α β set. This defines a map Φ from CFk(M, ω, ψ ) to CFk(M, ω, ψ ) by

X hcαi 7→ #M(cα, cβ, Φ)hcβi.

β β c ∈P^(ψ )k

The proof that this is a chain map is exactly the same as in the Hamilto- α α β β α β nian case. Given c ∈ P^(ψ ) and c ∈ P^(ψ ) with µψα (c ) − µψβ (c ) = 1, consider the 1-dimensional manifold M(cα, cβ, Φ). This manifold is not compact, but Floer’s gluing procedure shows that we can consider this space as a manifold with boundary given by

α β [ α α α α  α β ∂M(c , c , Φ) = M(c , b , ψ ,J )/R × M(b , c , Φ) ∪ α α b ∈P^(ψ ) α µψα (c )−1 [ M(cα, bβ, Φ) × M(bβ, cβ, ψβ,J β).

β β b ∈P^(ψ ) α µψα (c )

This means that ∂(ψβ,J β) ◦ Φα,β = Φα,β ◦ ∂(ψα,J α), which means that Φα,β is in fact a chain map, and so it induces a homomorphism on the level of homology.

78 This homomorphism is actually an isomorphism, with inverse induced by the homotopy Φ(¯ s) = Φ(−s). (This is a homotopy from (ψβ,J β) to (ψα,J α).) Homotopies as above will be called regular homotopies. Finally, we show that the continuation isomorphisms are independent of the regular homotopy used to define them. Let Φ0 and Φ1 be two regular homotopies between the regular pairs (ψα,J α) and (ψβ,J β). Most of the ideas needed to extend this to the almost Hamiltonian case were given above. First, we know that Φ0 and Φ1 are of the form Φ0(s) = s ¯ 0 1 s ¯ 1 ¯ 0 ¯ 1 g0 ∗ Φ (s) and Φ (s) = g1 ∗ Φ (s) for some homotopies Φ and Φ and paths ¯ ¯ 0 g1,2 : R → D. Choose Ψ a (Hamiltonian) homotopy of homotopies from Φ to Φ¯ 1, and choose a map g : [0, 1] × R → D, with a liftg ˜ : [0, 1] × R → De, such that eachg ˜(r, s) is in the component of the identity. This defines a path of homotopies connecting Φ0 to Φ1 by Ψ(r)(s) = g(r, s) ∗ Ψ(¯ r)(s). α α For a fixed r ∈ [0, 1], ∂Ψ(r)(u) is well defined. For c ∈ P^(ψ ) and cβ ∈ P^(ψβ), define M(cα, cβ, Ψ) as the set of all pairs (u, r) ∈ C∞(R × 1 S ,M) × [0, 1] such that ∂Ψ(r)(u) = 0, and there exists a liftu ˜ : R → LM α,β with limits lims→−∞, +∞ u˜(s) = c . As above, we use a “stretching the neck” argument to show that for carefully chosen g, there is a bijection between the sets M(cα, cβ, Ψ) and α α β β ¯ α β M(˜g (c ), g˜ (c ), Ψ). Thus, if µψα (c ) − µψβ (c ) = −1, this set is finite. α β Thus, we can define a map from CFk(M, ω, ψ ) to CFk+1(M, ω, ψ ) by α P α β  β hc i 7→ β β # M(c , c , Ψ) hc i. c ∈P^(ψ )k+1 As in the Hamiltonian case, this map defines a chain homotopy between the maps induced by Φ0 and Φ1, proving that the induce they same isomor- phism on the homology level.

79 

According to this theorem, we may speak of HF∗(M, ω) independently of the choice of (ψ, J) ∈ Freg. In effect, we have extended the space from which we may choose to define HF∗(M, ω). Now, rather than just allowing generic pairs (H,J) ∈ C∞(M × S1) × J (M, ω, S1), we allow generic pairs (ψ, J) ∈ F. Next, we use this expanded version of Floer homology to extend

Seidel’s action of π0(Ge) on HF∗(M, ω) to an action of π0(De).

80 5 Extending the Action

We quickly recall Seidel’s homomorphism from Ge to Aut(HF∗(M, ω)). Be- ∞ 1 gin by choosing a regular pair (H,J) ∈ (C (M × S ) × J )reg. For (g, g˜) in Ge, let Φ be the continuation isomorphism from HF∗(M, ω, H, J) to g g g g HF∗(M, ω, H ,J ) (the pair (H ,J ) is given by definition 1.24). This defines an automorphism of HF∗(M, ω, H, J) by

g g HF∗(M, ω, H, J) −−−→ HF∗(M, ω, H ,J ) −−−→ HF∗(M, ω, H, J). Φ g˜

Here,g ˜ means the map on homology induced by the map hci 7→ hg˜(c)i. This induced map is actually an isomorphism. In fact, we can speak of this action independent of the choice of regular pair. Given a different choice of regular pair, (H0,J 0), the following diagram commutes. (As usual, the symbol Φ represents the continuation isomorphism between the appropriate homologies.)

g g HF∗(M, ω, H, J) −−−−→ HF∗(M, ω, H ,J ) −−−−→ HF∗(M, ω, H, J) Φ g˜     Φy yΦ 0 0 0 g 0 g 0 0 HF∗(M, ω, H ,J ) −−−−→ HF∗(M, ω, (H ) , (J ) ) −−−−→ HF∗(M, ω, H ,J ) Φ g˜

This means that (g, g˜) ∈ Ge induces an isomorphism of HF∗(M, ω) indepen- dent of the regular pair used to define HF∗(M, ω). In fact, this action is well defined on π0(Ge).

Theorem 5.1 (Seidel) Let (g0, g˜0) and (g1, g˜1) be elements of Ge such that their classes inside π0(Ge) agree. Then (g0, g˜0) and (g1, g˜1) induce the same automorphisms of HF∗(M, ω). (That is, they induce the same automor-

81 phism of HF∗(M, ω, H, J) for any choice of regular pair (H,J), and this automorphism commutes with the continuation isomorphism to any other regular pair.)



The conditions of the above theorem mean that g0 and g1 give the same element of π1(Ham(M)), and there is a path in Ge from (g0, g˜0) to (g1, g˜1).

This describes the homomorphismσ ˜ : π0(Ge) → Aut(HF∗(M, ω)) intro- duced by Seidel. We want to extend this map to have domain De. Most of the ingredients involved have already been developed. First, we move from the Hamiltonian case to the almost Hamiltonian case, and define HF∗(M, ω) via a choice of

(ψ, J) ∈ Freg. For (g, g˜) ∈ De, let Φ be the continuation isomorphism from g HF∗(M, ω, ψ, J) to HF∗(M, ω, g ∗ ψ, J ) as in theorem 4.16. Define an automorphism of HF∗(M, ω, ψ, J) by

g HF∗(M, ω, ψ, J) −−−→ HF∗(M, ω, g ∗ ψ, J ) −−−→ HF∗(M, ω, ψ, J). Φ g˜

It follows from proposition 2.5 that if (ψ, J) and g were Hamiltonian,

(i.e., (ψ, J) ∈ (H × J )reg and (g, g˜) ∈ Ge), then the above automorphism of

HF∗(M, ω, ψ, J) correspondsσ ˜(g, g˜).

In order to extend this to a homomorphism from De to Aut(HF∗(M, ω)), we need only show that (g, g˜) induces an automorphism of HF∗(M, ω) in- dependent of the (ψ, J) ∈ Freg used to define it.

Proposition 5.2 Let (ψα,J α) and (ψβ,J β) be regular pairs, and let (g, g˜) be in De. The following diagram commutes. (Again, Φ denotes the contin-

82 uation isomorphisms.)

α α α g α α HF∗(ψ ,J ) −−−→ HF∗(g ∗ ψ, (J ) ) −−−→ HF∗(ψ ,J ) Φα g˜      g  yΦ yΦ yΦ β β β β g β β HF∗(ψ ,J ) −−−→ HF∗(g ∗ ψ , (J ) ) −−−→ HF∗(ψ ,J ) Φβ g˜

Proof: The left square of the diagram consists of the continuation isomorphisms, and so it commutes by theorem 4.16. The fact that the right square commutes is proved in the same way as proposition 4.9. Choose a regular homotopy Φ. Define Φg by Φg(s) = g ∗ (Φ(s)). The key point is that for cα,β ∈ P^(ψα,β), there is a bijection between M(cα, cβ, Φ) and M(˜g(cα), g˜(cβ), Φg). Since Φ and Φg consist of counting these spaces, the proposition follows.



Finally, we show that the above action is well defined on π0(De).

Theorem 5.3 If [g0, g˜0] = [g1, g˜1] in π0(De), then they induce the same automorphisms of HF∗(M, ω).

Proof: It is sufficient to consider the case when (g, g˜) is in the component of id in De. Choose (ψ, J) ∈ Freg. Let Φ denote the continuation isomorphism g from HF∗(M, ω, ψ, J) to HF∗(M, ω, g ∗ ψ, J ). We must show thatg ˜ ◦ Φ =

IdHF∗(M,ω,ψ,J). The proof is again analagous to the Hamiltonian case, with the same modifications introduced in order to proof theorem 4.16.

83 Let (gr, g˜r) be a path in D connecting id to g, and let Φ be any regular homotopy from (ψ, J) to (g∗ψ, J g). By a deformation of homotopies compatible with (ψ, J), gr and Φ, we will mean a two parameter family Υ = (ψr,s,J r,s), 0 ≤ r ≤ 1, −∞ < s < ∞, where for a fixed (r, s), (ψr,s,J r,s) is in F. These must satisfy the following conditions.

r,s r,s gr 1) s ≤ −1 =⇒ ψt = gr ∗ ψ, Jt = Jt , r,s r,s 2) s ≥ 1 =⇒ ψt = ψt,Jt = Jt, 0,s 0,s 3) ψt = ψt,Jt = Jt, ∀s ∈ R, 4) (ψ1,s,J 1,s) = Φ(s), ∀s ∈ R.

Choose a deformation of homotopies Υ = (ψr,s,J r,s), and let (r, u) ∈ ∞ 1  [0, 1] × C (R × S ,M) satisfy

  ∂u ∂u s,w (s, t) + J s,w(u(s, t)) (s, t) − Xψ (u(s, t)) = 0. (23) ∂s t ∂t t

We will say that such a pair converges to c−, c+ ∈ P](ψ) if there is a −1 smooth liftu ˜ : R → LM with limits lims→−∞(˜u(s)) = (˜gw) (c−) and lims→∞(˜u(s)) = c+. The set of all such pairs is denoted by M(c−, c+, Υ).

There is a deformation of homotopies such that whenever µψ(c+) =

µψ(c−)+1, M(c−, c+, Υ) is a finite set. This is proved exactly as in theorems 4.16. More precisely, we decompose the homotopy Φ into Φ(s) = hs ∗(θs, js) as in as in proposition 3.8, with each hr contractibile. Denote the homotopy from (θ−∞, j−∞) to (θ∞, j∞) by Φ.¯ This gives a decomposition (ψ, J) = h ∗ (θ, j). We choose a deformation of homotopies compatible with (θ, j), grhr, and Φ.¯ Since this is a Hamiltonian deformation, the finiteness holds (lemma 5.3 of [21]). As in theorem 4.16, there is a bijection from this finite set to the almost Hamiltonian set for carefully chosen decompositions.

84 We now proceed as in the Hamiltonian case, and use this to define a map h : CF (M, ω, ψ) → CF (M, ω, ψ) by hc i 7→ P #M(c , c , Υ)hc i. ∗ ∗+1 − c+ − + +

Finally, it follows from lemma 5.5 in [21] that ∂k+1(ψ, J)hk +hk+1∂k(ψ, J) = −1 Φ ◦ g˜ − IdCFk(ψ). That is, h gives a chain homotopy between Ψ ◦ g˜ and

IdHF∗(ψ,J), proving the theorem.



Notice that this is not an improved proof of the analogous theorem in the Hamiltonian case, because the Hamiltonian theorem is used in proving theorem 4.16, which is used in the above proof.

85 6 Concluding Remarks

The methods of the proof of the main result of the thesis suggest more questions about Floer homology. In the classical case, Floer homology is defined by choosing a generic, periodic, time dependent function on M, and an almost complex structure. These choices define a function and Rieman- nian metric on the infinite dimensional manifold of contractible loops in M, and Floer homology can be considered as Morse theory using this function and metric. To prove the main result, we extended the available choices used in defining HF∗(M, ω). However, it appears that we lose the analogy with Morse theory when we use the extended domain. Thus, one can ask if the extended domain of definition can be shown to arise as Morse theory using some function and metric on the space of contractible loops. Even more fundamentally, exactly how far can this space of choices be extended?

There is a collection of results which seem to suggest that Ham(M) is more related to the topology of M than might be guessed. For example, any homotopy class [ψt] ∈ π1(Diff(M)) defines a map evψt : M → π1(M) by evψt (p)(t) = ψt(p). If ψt is a Hamiltonian loop, then evψt ≡ 0. Since

π1(M) is a topological invariant of M, this shows a deep connection between Ham(M) and the topology of M. The main result of the thesis gives more evidence of this phenomenon. As of now, these rigidity results seem to be symptoms of some deep connection between Ham(M) and the topology of the manifold. A long term goal is to discover the underlying connection that would explain all of the rigidity results.

These questions are intimately related to the homotopy groups of the triple Diff(M), Symp(M, ω), Ham(M). For the latter two, the homotopy properties of the pair (Symp(M), Ham(M)) are known: in dimension n > 1,

86 πn(Symp(M, ω)) = πn(Ham(M)), and in dimension n = 1, the difference is described by the Flux homomorphisms. The rigidy phenomonon seems to be related to the homotopy properties of the pair Diff(M), Ham(M). There is also still very little understanding of the homotopy properties of the pair Diff(M), Symp(M, ω), which may or may not share the same rigidity properties as the pair Diff(M), Ham(M).

In the end, the thesis answers one question, but at the same time, it suggests many more questions in its place, which is as it should be.

87 References

[1] Aebischer, B., Borer, M., K¨alin,M., Leuenberger, Ch. and Reimann, H.M., Symplectic geometry, an introduction based on the seminar in Bern, 1992. Progress in Mathematics, 124. Birkh¨auser Verlag, Basel, 1994. xii+239 pp.

[2] Banyaga, A., The structure of classical diffeomorphism groups, Math- ematics and its Applications, 400, Kluwer Academic Publishers Group, Dordrecht, 1997, xii+197 pp.

[3] Banyaga, A. and Hurtubise, D., Lectures on Morse Homology, Kluwer Texts in the Mathematical Sciences, Vol. 29, 2004, 326 pp.

[4] Bialy, M. and Polterovich, L., Hamiltonian diffeomorphisms and La- grangian distributions, Geom. and Funct. Anal. 2, (1992) 173-210.

[5] Bredon, A., Topology and geometry. Graduate Texts in Mathematics, 139. Springer-Verlag, New York, 1993, xiv+557 pp.

[6] Conley, C. and Zehnder, E., The Birkhoff-Lewis fixed point theorem and a conjecture of V. I. Arnold, Invent. Math. 73, (1983) 33-49.

[7] Dostoglou, S. and Salamon, D., Instanton homology and symplectic fixed points. Symplectic geometry, 57–93, London Math. Soc. Lecture Note Ser., 192, Cambridge Univ. Press, Cambridge, 1993.

[8] Floer, A., A relative Morse index for the symplectic action, it Comm. Pure Appl. Math., XLI, (1988), 393-407.

[9] Floer, A., Morse theory for Lagrangian intersections, J. Diff. Geom., 28 (1988), no.3, 513-547.

88 [10] Floer, A., The unregulized gradient flow of the symplectic action, Com- mun. Pure Appl. Math., 41 (1988), no. 6, 775-813.

[11] Floer, A., Symplectic fixed points and holomorphic spheres, Comm. Math. Phys., 120 (1989), 575-611.

[12] Gromov, M., Pseudoholomorphic curves in symplectic manifolds. In- vent. Math. 82 (1985), no. 2, 307–347.

[13] Hofer, H. and Salamon, D., Floer homology and Novikov rings. The Floer memorial volume, 483–524, Progr. Math., 133, Birkh¨auser, Basel, 1995.

[14] Lalonde, F., McDuff, D. and Polterovich, L., Topological rigidity of Hamiltonian loops and quantum homology, Invent. Math. 135 (1999), no. 2, 369–385.

[15] McDuff, D. and Salamon, D., Introduction to symplectic topology. Oxford Mathematical Monographs. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1995. viii+425 pp.

[16] McDuff, D. and Salamon, D., J-holomorphic curves and quantum coho- mology, University Lecture Series, 6. American Mathematical Society, Providence, RI, 1994. viii+207 pp.

[17] Milnor, J. and Stasheff, J., Characteristic classes. Annals of Mathe- matics Studies, No. 76. Princeton University Press, Princeton, N. J.; University of Tokyo Press, Tokyo, 1974. vii+331 pp.

[18] Salamon, D., Morse theory, the Conley index and Floer homology, Bull. L.M.S. 22 (1990), 113-140.

89 [19] Salamon, D., Lectures on Floer homology, Symplectic geometry and topology (Park City, UT, 1997), 143–229, IAS/Park City Math. Ser., 7, Amer. Math. Soc., Providence, RI, 1999

[20] Salamon, D. and Zehnder, D., Morse theory for periodic solutions of hamiltonian systems and the Maslov index, Comm. Pure Appl. Math., 45 (1992), 1303-1360.

[21] Seidel, P., π1 of symplectic automorphism groups and invertibles in quantum homology rings, Geom. funct. anal., 7 (1997), 1046-1095.

[22] Stelling, L., Fixed points of non-Hamiltonian Symplectomorphisms, Journal of Geom. Anal., 11 (2001), 693-716.

90 Curriculum Vitae

Christopher L. Saunders Department of Mathematics McAllister Building The Pennsylvania State University University Park, PA 16802 (USA) Telephone (814) 692-8160, e-mail: [email protected] Website: http://www.math.psu.edu/saunders

Education

August, 2000 - May, 2005 Doctoral Student at Pennsylvania State University Thesis advisor: Augustin Banyaga

August, 1996 - May, 1999 B.S. in Mathematics, Juniata College, Huntingdon, PA

Awards

Spring, 2005 Pritchard Dissertation Fellowship for Excellence in Thesis Research

Fall, 2004 Teaching Associate Certificate for Successful Teaching