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

The quantitative nature of reduced Floer theory

Sara Venkatesh
Abstract

We study the reduced symplectic cohomology of disk subbundles in negative symplectic line bundles. We show that this cohomology theory “sees” the spectrum of a quantum action on quantum cohomology. Precisely, quantum cohomology decomposes into generalized eigenspaces of the action of the first Chern class by quantum cup product. The reduced symplectic cohomology of a disk bundle of radius RR sees all eigenspaces whose eigenvalues have size less than RR, up to rescaling by a fixed constant. Similarly, we show that the reduced symplectic cohomology of an annulus subbundle between radii R1R_{1} and R2R_{2} captures all eigenspaces whose eigenvalues have size between R1R_{1} and R2R_{2}, up to a rescaling. We show how local closed-string mirror symmetry statements follow from these computations.

1 Introduction

Symplectic cohomology is a tool for delving the geometry of an open symplectic manifold. It was initially studied by Cieliebak, Floer, Hofer, and Wysocki to probe quantitative aspects of domains in 2n\mathbb{R}^{2n}; applications were focused on embedding problems and capacities. The focus on quantitative geometry changed when Viterbo introduced a “qualitative” version of symplectic cohomology [viterbo]. This qualitative definition has proved indispensable to the study of global symplectic properties, from attacking classification problems to analyzing Lagrangian embeddings [seidel-biased]. The open symplectic manifolds considered have been, for the most part, Liouville manifolds: manifolds that contain the entire symplectization ×Σ\mathbb{R}\times\Sigma of a contact manifold Σ\Sigma.

Very little has been done to understand symplectic cohomology away from the Liouville setting. The first attempt is due to Ritter, who computed the symplectic cohomology of symplectic line bundles satisfying a negativity condition [ritter-gromov]. Rather than containing an entire symplectization, these line bundles only contain the positive “piece” of a symplectization: they are compactifications of the space (0,)×Σ(0,\infty)\times\Sigma. Ritter’s result surprisingly tied the symplectic cohomology of a line bundle EE, denoted by SH(E)SH^{*}(E), to its quantum cohomology. The main theorem in [ritter-gromov] shows that symplectic cohomology sees almost all of quantum cohomology; it misses only the zeroth generalized eigenspace, denoted by QH0(E)QH^{*}_{0}(E), of a particular action on quantum cohomology. Precisely, there is a ring isomorphism

(1) SH(E)QH(E)/QH0(E).SH^{*}(E)\simeq{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$QH^{*}_{0}(E)$}\right.}.

The action in question is given by quantum cup product with the pull-back of the first Chern class of the the line bundle EE.

In [venkatesh] we showed how to refine Viterbo’s qualitative symplectic cohomology to produce a quantitative invariant. A chain complex computing symplectic cohomology has a natural family of non-Archimedean metrics; one can complete the chain complex with respect to a fixed metric and produce a reduced cohomology theory, à la L2L^{2}-cohomology [dai]. We call the resulting cohomology theory reduced symplectic cohomology. Each fixed metric encodes information about the size of the Reeb orbits on a fixed contact hypersurface. In line bundles, these fixed contact hypersurfaces are circle subbundles, and they come in an >0\mathbb{R}_{>0} family, indexed by radius, that induces an >0\mathbb{R}_{>0} family of metrics. We write SH^(DR)\widehat{SH^{*}}(D_{R}) for the reduced symplectic cohomology theory associated to the radius RR, where DRD_{R} is the disk subbundle of radius RR.

In [venkatesh-thesis] we studied reduced symplectic cohomology on monotone toric line bundles. These line bundles were previously studied by Ritter-Smith in the context of wrapped Fukaya categories [ritter-s]. They showed that the wrapped Fukaya category of a monotone, toric line bundle is split-generated by a single Lagrangian torus LL lying in a particular circle subbundle. In [venkatesh-thesis], we showed that

(2) SH^(DR)0LDR.\widehat{SH^{*}}(D_{R})\neq 0\iff L\subset D_{R}.

Thus, reduced symplectic cohomology contains local information about the disk bundle DRD_{R}.

It is not known what the Fukaya category looks like away from the monotone toric case. However, we can still hope to say something about reduced symplectic cohomology. Take coefficients in the Novikov field over \mathbb{C}, denoted by Λ\Lambda and defined by

Λ={i=0ciTαi|ci,αi}.\Lambda=\left\{\sum_{i=0}^{\infty}c_{i}T^{\alpha_{i}}\hskip 2.84544pt\big{|}\hskip 2.84544ptc_{i}\in\mathbb{C},\mathbb{R}\ni\alpha_{i}\rightarrow\infty\right\}.

Quantum cohomology splits into generalized eigenspaces

QH(E)=λQHλ(E),QH^{*}(E)=\bigoplus_{\lambda}QH^{*}_{\lambda}(E),

where QHλ(E)QH^{*}_{\lambda}(E) is the λ\lambda-generalized eigenspace of the pullback of c1Ec_{1}^{E} acting by quantum cup product. Each non-zero eigenvalue λ\lambda has a “size”, denoted by ev(λ)ev(\lambda), given by the natural valuation on the Novikov field (see (12)).

The full reduced symplectic cohomology is difficult to compute, and it may have subtle invariance. We work instead with a modified version, denoted by SH¯(DR)\overline{SH^{*}}(D_{R}), which is formed by completing only the subspace spanned by the kernel of the differential. The main result of this paper is the following Theorem.

Theorem 1

Let EE be a weak+, negative, symplectic line bundle with negativity constant kk. The reduced symplectic cohomology of a disk subbundle DRED_{R}\subset E is isomorphic as a Λ\Lambda-algebra to

SH¯(DR)QH(E)/ev(λ)>kπR2QHλ(E).\overline{SH^{*}}(D_{R})\simeq{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}QH^{*}_{\lambda}(E)$}\right.}.

Here, the negativity constant kk determines the line bundle EE and affects the symplectic structure.

  • Remark 1)

    Theorem 1 generalizes Ritter’s result (1), which may be rephrased as a computation of the reduced symplectic cohomology of the disk bundle of infinite radius. Note that, in this non-Archimedean setting, ev(0)=ev(0)=\infty.

If EE additionally lies over a toric base, there is an isomorphism

SH^(DR)SH¯(DR)\widehat{SH^{*}}(D_{R})\simeq\overline{SH^{*}}(D_{R})

(see Proposition 4).

  • Remark 2)

    If EE is monotone over a toric base, all non-zero eigenvalues have the same size. For λ0\lambda\neq 0, the split-generating Lagrangian LL lies in a circle bundle of radius R=1kπev(λ)R=\frac{1}{\sqrt{k\pi\cdot ev(\lambda)}} [ritter-fano]. Thus, Theorem 1 is also a generalization of the equivalence (2).

The full theory SH^(DR)\widehat{SH^{*}}(D_{R}) has an extension to the symplectic cohomology of a trivial cobordism, initially defined and studied in [cieliebak-o] and [c-f-o]. In the case of a negative line bundle, the trivial cobordisms are the annulus subbundles. We denote an annulus subbundle between radii R1<R2R_{1}<R_{2} by AR1,R2A_{R_{1},R_{2}}. Completed symplectic cohomology for a trivial cobordism yields an invariant SH^(AR1,R2)\widehat{SH^{*}}(A_{R_{1},R_{2}}), defined in [venkatesh].

Theorem 2

If EE is furthermore a line bundle over a toric base, there is a vector-space isomorphism

(3) SH^(AR1,R2)QH(E)/ev(λ)>kπR22orev(λ)kπR12QHλ(E).\widehat{SH^{*}}(A_{R_{1},R_{2}})\simeq{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{\begin{subarray}{c}ev(\lambda)>k\pi R_{2}^{2}\\ \text{or}\\ ev(\lambda)\leq k\pi R_{1}^{2}\end{subarray}}QH^{*}_{\lambda}(E)$}\right.}.
  • Remark 3)

    The main result of [c-f-o] shows that, in the Liouville setting, the uncompleted symplectic cohomology of a trivial cobordism (ϵ,ϵ)×Σ(-\epsilon,\epsilon)\times\Sigma is isomorphic to the Rabinowitz Floer homology of Σ\Sigma. Expounding upon this result, the reduced theory SH^(AR1,R2)\widehat{SH^{*}}(A_{R_{1},R_{2}}) is conjecturally related to the Rabinowitz Floer theory developed by Albers-Kang for contact-type hypersurfaces in negative line bundles [albers-k].

  • Remark 4)

    The chain complex computing SH^(AR1,R2)\widehat{SH^{*}}(A_{R_{1},R_{2}}) considers positively-traversed Reeb orbits at radius R2R_{2} and negatively-traversed Reeb orbits at radius R1R_{1}. One could just as easily switch the roles of R1R_{1} and R2R_{2}, instead considering the negatively-traversed Reeb orbits at radius R2R_{2} and the positively-traversed Reeb orbits at radius R1R_{1}. Abusing notation, we write SH^(AR2,R1)\widehat{SH^{*}}(A_{R_{2},R_{1}}) for the completed cohomology theory that switches these roles. This theory is dual to SH^(AR1,R2)\widehat{SH^{*}}(A_{R_{1},R_{2}}) and satisfies

    SH^(AR2,R1)kπR12<ev(λ)kπR22QHλ(E),\widehat{SH^{*}}(A_{R_{2},R_{1}})\simeq\bigoplus_{k\pi R_{1}^{2}<ev(\lambda)\leq k\pi R_{2}^{2}}QH_{*}^{\lambda}(E),

    where QHλ(E)QH^{\lambda}_{*}(E) is the λ\lambda-generalized eigenspace of P.D.(c1E)P.D.(c_{1}^{E}) acting by quantum intersection product on the quantum homology QH(E)QH_{*}(E).

Toric line bundles EE have a Landau-Ginzburg mirror (E,W)(E^{\vee},W), where EE^{\vee} is a rigid analytic space, and WW is an analytic function on EE^{\vee}. A closed-string mirror symmetry statement is proved for some cases in [ritter-s], and in the full monotone case in [ritter-fano]. Namely, if EE is monotone, then there is an isomorphism

(4) SH(E)Jac(W).SH^{*}(E)\simeq Jac(W).

A local statement of mirror symmetry for annulus subbundles follows as a corollary of Theorem 1 and the isomorphism (4).

Corollary 1

Let R1<R2R_{1}<R_{2}. An annulus subbundle AR1,R2A_{R_{1},R_{2}} in a monotone, negative line bundle over a toric base has a Landau-Ginzburg mirror (AR1,R2,W)(A_{R_{1},R_{2}}^{\vee},W) such that

(5) SH^(AR1,R2)Jac(W).\widehat{SH^{*}}(A_{R_{1},R_{2}})\simeq Jac(W).

This proves a conjecture of the author made in [venkatesh]. We believe that the correspondence (5) holds more generally, but we do not know what the superpotential WW looks like away from the monotone case.

1.1 Outline of paper

In Section 2 we recall the definition of Hamiltonian Floer theory and symplectic cohomology. We define and discuss reduced symplectic cohomology, and we discuss extensions to the symplectic cohomology of a cobordism. In Section 3 we introduce negative line bundles and the specific Floer data that we will use to compute symplectic cohomology. The entirety of Section 4 is devoted to proving Theorem 1. In Section 5 we prove that, in some cases, completed and reduced symplectic cohomology coincide. We then prove Theorem 2, restated as Theorem 8, and discuss closed string mirror symmetry.

1.2 Acknowledgements

We thank Paul Seidel for motivating this project and Mohammed Abouzaid for helpful discussions. We thank the Institute for Advanced Study for the productive environment under which the bulk of this paper was completed. This work is based upon work supported by the National Science Foundation under Award No. 1902679.

2 Floer theory

We begin this section by recalling the data of Hamiltonian Floer theory and symplectic cohomology, fixing conventions along the way. We refer the reader to [audin-d] and [mcduff-s] for an introduction to Floer theory, and we refer to [seidel-biased] for details on symplectic cohomology. We then introduce the notion of reduced symplectic cohomology, recently defined and studied by the author in [venkatesh], as well as in the independent works of Groman [groman], McLean [mclean], and Varolgunes [varolgunes].

Throughout this paper, we will consider a symplectic manifold (E,Ω)(E,\Omega) of dimension 2m2m that is weak+ monotone. Abusing notation, denote by π2(E)\pi_{2}(E) the image of π2(E)\pi_{2}(E) in H2(E)H_{2}(E) under the Hurewicz homomorphism. Practically, (E,Ω)(E,\Omega) is weak+ monotone if it satisfies at least one of four conditions on π2(E)\pi_{2}(E):

  1. 1.

    there is a positive constant κ>0\kappa>0 such that c1TE=κ[Ω]c_{1}^{TE}=\kappa[\Omega] on π2(E)\pi_{2}(E),

  2. 2.

    c1TE|π2(E)=0c_{1}^{TE}\big{|}_{\pi_{2}(E)}=0,

  3. 3.

    [Ω]|π2(E)=0[\Omega]\big{|}_{\pi_{2}(E)}=0,

  4. 4.

    or the minimal Chern number, defined to be

    minAπ2(E)c1TE(A)0|c1TE(A)|,\min\limits_{\begin{subarray}{c}A\in\pi_{2}(E)\\ c_{1}^{TE}(A)\neq 0\end{subarray}}|c_{1}^{TE}(A)|,

    is at least m1m-1.

This rather ad hoc collection of conditions ensures that moduli spaces are well-defined in both the definition of Floer theory and in the definition of the P.S.S. maps that relate Floer theory to quantum cohomology [hofer-s].

2.1 Hamiltonian Floer theory

Let H:E×S1H:E\times S^{1}\longrightarrow\mathbb{R} be a Hamiltonian. HH induces a Hamiltonian vector field XHX_{H}, which is the unique vector field satisfying

dH()=Ω(,XH).dH(-)=\Omega(-,X_{H}).

The smooth maps x:S1Ex:S^{1}\longrightarrow E satisfying

x˙=XH(x(t))\dot{x}=X_{H}(x(t))

are called the periodic orbits of HH. Assume that the periodic orbits of HH satisfy a generic non-degeneracy condition. Denote the set of orbits by 𝒫(H){\cal P}(H). Define the universal Novikov ring over \mathbb{C} to be

(6) Λ={i=0ciTαi|ci,αi}.\Lambda=\left\{\sum_{i=0}^{\infty}c_{i}T^{\alpha_{i}}\hskip 2.84544pt\big{|}\hskip 2.84544ptc_{i}\in\mathbb{C},\mathbb{R}\ni\alpha_{i}\rightarrow\infty\right\}.

Define the Hamiltonian cochain complex CF(H;Λ)CF^{*}(H;\Lambda) to be a cochain complex with underlying vector space over Λ\Lambda generated by the periodic orbits of XHX_{H}:

CF(H;Λ):=x𝒫(H)Λx.CF^{*}(H;\Lambda):=\bigoplus_{x\in{\cal P}(H)}\Lambda\langle x\rangle.

Let τ\tau be the minimal Chern number. Grading is given by the cohomological Conley-Zehnder index and takes values in /2τ\mathbb{Z}/2\tau\mathbb{Z} [seidel-biased].

To define the differential on CF(H;Λ)CF^{*}(H;\Lambda), fix a capping x~\tilde{x} of each periodic orbit x𝒫(H)x\in{\cal P}(H). The action of xx is defined to be

(7) 𝒜H(x)=Dx~Ω+01H(x(t))𝑑t.{\cal A}_{H}(x)=-\int_{D}\tilde{x}^{*}\Omega+\int_{0}^{1}H(x(t))dt.

Let JJ be an Ω\Omega-tame almost-complex structure that is Ω\Omega-compatible in neighborhoods of the periodic orbits of XHX_{H}. Consider a solution u:s×St1Eu:\mathbb{R}_{s}\times S^{1}_{t}\longrightarrow E of Floer’s equation

(8) us+J(utXH)=0.\frac{\partial u}{\partial s}+J\left(\frac{\partial u}{\partial t}-X_{H}\right)=0.

The energy of uu is defined to be

(9) E(u)=×S1su2𝑑sdt.E(u)=\int_{\mathbb{R}\times S^{1}}||\partial_{s}u||^{2}ds\wedge dt.

A finite-energy Floer solution converges asymptotically in s±s\rightarrow\pm\infty to orbits in 𝒫(H){\cal P}(H). Fix x,x+𝒫(H)x_{-},x_{+}\in{\cal P}(H) and denote by

^(x,x+)\widehat{{\cal M}}(x_{-},x_{+})

the set of finite-energy Floer solutions with

lims±u(s,t)=x±(t).\lim\limits_{s\rightarrow\pm\infty}u(s,t)=x_{\pm}(t).

The moduli space ^(x,x+)\widehat{{\cal M}}(x_{-},x_{+}) is equipped with a free \mathbb{R}-action whenever xx+x_{-}\neq x_{+} that translates the ss-coordinate of a Floer solution. Modding out by this \mathbb{R} action yields a moduli space

(x,x+),{\cal M}(x_{-},x_{+}),

whose components can be indexed by their dimension. The zero-dimensional component, denoted by 0(x,x+){\cal M}^{0}(x_{-},x_{+}), is compact if EE is closed. In this case, the Floer differential fl\partial^{fl} is defined on CF(H;Λ)CF^{*}(H;\Lambda) by

fl(y)=x𝒫(H)u0(x,y)±T[Ω]([y~#(u)#(x~)])x,\partial^{fl}(y)=\sum_{x\in{\cal P}(H)}\sum_{u\in{\cal M}^{0}(x,y)}\pm T^{-[\Omega]([\tilde{y}\#(-u)\#(-\tilde{x})])}\langle x\rangle,

where the sign is determined by fixing choices of orientations on each moduli space and the expression [y~#(u)#(x~)][\tilde{y}\#(-u)\#(-\tilde{x})] refers to the homology class of the sphere constructed by gluing y~\tilde{y}, u-u, and x~-\tilde{x} along their boundaries. We refer to [venkatesh-thesis] for a discussion of orientations in the context of this paper.

2.2 Symplectic cohomology

Assume now that EE is an open symplectic manifold, which, outside of a compact set, is modeled on the symplectization

(10) (0,)r×Σ,Ω=d(πr2α)(0,\infty)_{r}\times\Sigma,\Omega=d(\pi r^{2}\alpha)

of a contact manifold Σ\Sigma with contact form α\alpha. Fix a radius R>0R>0 and a monotone-increasing sequence τn>0\tau_{n}\in\mathbb{R}_{>0} with τn\tau_{n}\rightarrow\infty. Let {Hn}n\{H_{n}\}_{n\in\mathbb{N}} be a family of Hamiltonians satisfying the following constraints.

  1. 1.

    |Hn||H_{n}| is bounded uniformly by a fixed constant C>0C>0 on E(R,)×ΣE\setminus(R,\infty)\times\Sigma,

  2. 2.

    HnH_{n} is linear in πr2\pi r^{2} with slope τn\tau_{n} on (R,)×Σ(R,\infty)\times\Sigma, and

  3. 3.

    the non-constant orbits of HnH_{n} lie in some small neighborhood (R1n,R)×Σ(R-\frac{1}{n},R)\times\Sigma.

The first condition ensures that the action of a periodic orbit is not heavily influenced by HnH_{n}, the second condition ensures that Floer solutions are well-behaved, and the final condition ensures that the set of periodic orbits n𝒫(Hn)\cup\limits_{n}{\cal P}(H_{n}) cluster near the contact hypersurface {R}×Σ\{R\}\times\Sigma. Assume that each HnH_{n} is chosen so that all one-periodic orbits of HnH_{n} are non-degenerate. Choose an almost-complex structure that is cylindrical on (R,)×Σ(R,\infty)\times\Sigma:

Jdr=πr2α.J^{*}dr=-\pi r^{2}\alpha.

Under the assumptions on EE, {Hn}\{H_{n}\}, and JJ, the Floer cochain complex defined in Section 2.1 is well-defined.

By assumption, HnHn+1H_{n}\leq H_{n+1} outside a compact set of EE. Thus, there are chain maps, called continuation maps,

cn:CF(Hn;Λ)CF(Hn+1;Λ)c_{n}:CF^{*}(H_{n};\Lambda)\rightarrow CF^{*}(H_{n+1};\Lambda)

for each nn. Let 𝐪{\bf q} be a formal variable of degree 1-1, satisfying 𝐪2=0{\bf q}^{2}=0. Define the symplectic cochain complex of EE to be

(11) SC(HR)=n=0CF(Hn;Λ)[𝐪],SC^{*}(H_{R})=\bigoplus_{n=0}^{\infty}CF^{*}(H_{n};\Lambda)[{\bf q}],

with differential (x+y𝐪)=fl(x)+(cid)(y)+fl(y)𝐪\partial(x+y{\bf q})=\partial^{fl}(x)+(c-id)(y)+\partial^{fl}(y){\bf q}, where

c=n=0cn.c=\bigoplus_{n=0}^{\infty}c_{n}.

The cohomology of SC(HR)SC^{*}(H_{R}), denoted by SH(HR)SH^{*}(H_{R}), is called the symplectic cohomology of EE.

A standard result in Floer theory is

Theorem 3

The symplectic cohomology SH(HR)SH^{*}(H_{R}) is independent of choice of radius RR, family {Hn}\{H_{n}\}, and cylindrical almost-complex structure JJ.

We will often write SH(E)SH^{*}(E) instead of SH(HR)SH^{*}(H_{R}).

2.3 Reduced symplectic cohomology

Having fixed a definition of action in (7), the Floer cochain complex has a natural non-Archimedean metric. To see this, first recall that the Novikov ring has a valuation

(12) ev:Λ\displaystyle ev:\Lambda {}\displaystyle\rightarrow\mathbb{R}\cup\{\infty\}
(13) i=0ciTαi\displaystyle\sum_{i=0}^{\infty}c_{i}T^{\alpha_{i}} minci0αiif ci0\displaystyle\mapsto\min_{c_{i}\neq 0}\alpha_{i}\hskip 14.22636pt\text{if }\exists\hskip 2.84544ptc_{i}\neq 0
(14) 0\displaystyle 0 .\displaystyle\mapsto\infty.

Define a valuation 𝒜{\cal A} on CF(Hn)CF^{*}(H_{n}) by

𝒜(xi𝒫(Hn)jCixi)=mini(ev(Ci)+𝒜Hn(xi)),{\cal A}\left(\sum_{x_{i}\in{\cal P}(H_{n})}^{j}C_{i}x_{i}\right)=\min_{i}\left(ev(C_{i})+{\cal A}_{H_{n}}(x_{i})\right),

where CiΛC_{i}\in\Lambda are Λ\Lambda-valued coefficients. Extend this to a valuation on SC(H)SC^{*}(H) by

𝒜(i=0jXi𝐪i)=mini𝒜Hni(Xi),{\cal A}\left(\sum_{i=0}^{j}X_{i}{\bf q}^{\ell_{i}}\right)=\min_{i}{\cal A}_{H_{n_{i}}}(X_{i}),

where each XiX_{i} lies in a fixed CF(Hni)CF^{*}(H_{n_{i}}) and i{0,1}\ell_{i}\in\{0,1\}. The non-Archimedean metric ||||||\cdot|| on SC(HR)SC^{*}(H_{R}) is given by

X=e𝒜(X).||X||=e^{-{\cal A}(X)}.

Following [groman], denote by ker()^\widehat{\ker(\partial)} the completion of ker()SC(H)\ker(\partial)\subset SC^{*}(H) with respect to ||||||\cdot||. Denote by im()¯\overline{\textnormal{im}(\partial)} the closure of im()\textnormal{im}(\partial) in ker()^\widehat{\ker(\partial)}. The quotient

SC¯(HR):=ker()^/im()¯\overline{SC^{*}}(H_{R}):={\raisebox{1.99997pt}{$\widehat{\ker(\partial)}$}\left/\raisebox{-1.99997pt}{$\overline{\textnormal{im}(\partial)}$}\right.}

is the reduced symplectic cohomology of the Floer data {Hn}\{H_{n}\}.

An alternative definition of reduced symplectic cohomology, and one which is prevalent in the literature, is to first complete SC(H)SC^{*}(H) with respect to ||||||\cdot||, and then define ^\widehat{\partial} on SC^(H)\widehat{SC^{*}}(H) to be the natural extension of \partial. This defines a symplectic cohomology theory

(15) SH^(HR)=ker(^)/im(^)¯.\widehat{SH^{*}}(H_{R})={\raisebox{1.99997pt}{$\ker(\widehat{\partial})$}\left/\raisebox{-1.99997pt}{$\overline{\textnormal{im}(\widehat{\partial})}$}\right.}.

To avoid confusion, we call SH^(HR)\widehat{SH^{*}}(H_{R}) the completed symplectic cohomology.

Completed symplectic cohomology has an equivalent definition as the algebraic completion induced by a filtration. By choosing HnHn+1H_{n}\leq H_{n+1} everywhere, the action is increased by both the Floer differential and continuation maps. It therefore defines a filtration on SC(H)SC^{*}(H). Define the subcomplex

(16) SC(a,)(HR)=XSC(HR)|𝒜(X)>a,(a,)SC^{*}_{(a,\infty)}(H_{R})=\big{\langle}X\in SC^{*}(H_{R})\hskip 5.69046pt\bigg{|}\hskip 5.69046pt{\cal A}(X)>a\big{\rangle},\partial_{(a,\infty)}

and quotient complex

(17) SCa(HR)=SC(HR)/(SC(a,)(HR),aSC^{*}_{a}(H_{R})={\raisebox{1.99997pt}{$SC^{*}(H_{R})$}\left/\raisebox{-1.99997pt}{$(SC^{*}_{(a,\infty)}(H_{R})$}\right.},\partial_{a}

These complexes are modules over the positive Novikov ring

Λ0:=ev1([0,]]\Lambda_{0}:=ev^{-1}([0,\infty]]

and define cohomology theories denoted, respectively, by SH(a,)(HR)SH^{*}_{(a,\infty)}(H_{R}) and SHa(HR)SH^{*}_{a}(H_{R}). Running over all aa\in\mathbb{R}, these theories form a directed, respectively inversely directed, system through the canonical inclusion, respectively projection at the level of cochains. This gives an alternative description of completed symplectic cohomology, through a Theorem of Groman.

Theorem 4 (Groman: Theorem 8.4 in [groman])
SH^(HR)=limaSHa(HR).\widehat{SH^{*}}(H_{R})=\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}SH^{*}_{a}(H_{R}).
  • Remark 5)

    In practice, we will take the inverse limit over a countable set a1,a2,a_{1},a_{2},...\rightarrow\infty. This simplifies computations.

  • Remark 6)

    SH¯(HR)\overline{SH^{*}}(H_{R}) and SH^(HR)\widehat{SH^{*}}(H_{R}) now depend upon the families {Hn}\{H_{n}\} used to define them. As we will see, these cohomology theories are quantitative invariants encoding local information about domains contained in EE.

  • Remark 7)

    In this paper we study SH¯(HR)\overline{SH^{*}}(H_{R}), rather than SH^(HR)\widehat{SH^{*}}(H_{R}), simply because this is the object that we can compute. We give examples in Section 5 of manifolds EE and families {Hn}\{H_{n}\} on EE for which

    (18) SH¯(HR)=SH^(HR).\overline{SH^{*}}(H_{R})=\widehat{SH^{*}}(H_{R}).

    We do not know in how much generality (18) holds. However, we believe SH¯(HR)\overline{SH^{*}}(H_{R}) to be at least as robust an invariant as SH^(HR)\widehat{SH^{*}}(H_{R}). While we have no Floer theory examples to support this belief, it is straight-forward to find examples of chain complexes for which these two theories disagree. Consider the following example from Morse theory.

    Refer to caption
    Refer to caption
    Figure 1: The functions g1g_{1} and g4g_{4}

    Let {gn:I}\{g_{n}:I\subset\mathbb{R}\longrightarrow\mathbb{R}\} be the family of Morse functions pictured in Figure 1. Denote the Morse cochain complex of gng_{n} over a ring RR by CM(gn;R)CM^{*}(g_{n};R). Choose continuation maps

    cn:CM(gn;R)CM(gn+1;R)c_{n}:CM^{*}(g_{n};R)\longrightarrow CM^{*}(g_{n+1};R)

    that correspond to the inclusion of a subcomplex. Analogously to symplectic cohomology, define a new Morse-type complex

    CM(g)=n=0CM(gn;R)[𝐪],CM^{*}(g)=\bigoplus_{n=0}^{\infty}CM^{*}(g_{n};R)[{\bf q}],

    where 𝐪{\bf q} is a formal variable of degree 1-1, satisfying 𝐪2=0{\bf q}^{2}=0. Denoting the Morse differential by M\partial^{M}, the differential on SC(M)SC^{*}(M) is given by

    (x+y𝐪)=M(x)+(cid)(y)+M(y)𝐪.\partial(x+y{\bf q})=\partial^{M}(x)+(c-id)(y)+\partial^{M}(y){\bf q}.

    For xCrit(gn)x\in\textnormal{Crit}(g_{n}), define

    𝒜(x)=gn(x).{\cal A}(x)=g_{n}(x).

    The assignment 𝒜{\cal A} extends to a valuation on the Morse complex CM(g)CM^{*}(g) that, in turn, defines a non-Archimedean metric. We use these metrics to define reduced and completed Morse cohomology theories, denoted by HM¯(g)\overline{HM^{*}}(g) and HM^(g)\widehat{HM^{*}}(g). It is straight-forward to see that

    HM¯(g)=0,\overline{HM^{*}}(g)=0,

    but

    HM^(g)=R.\widehat{HM^{*}}(g)=R.

    For any single gng_{n},

    HM¯(gn)=HM^(gn)=0.\overline{HM^{*}}(g_{n})=\widehat{HM^{*}}(g_{n})=0.

    Thus, in this contrived scenario, the reduced cohomology seems to behave stably, while the completed theory does not.

    In practice, we will assume that all Hamiltonians have bounded sup norm. We do not know if this is sufficient to always achieve an equality

    SH¯(HR)=SH^(HR).\overline{SH^{*}}(H_{R})=\widehat{SH^{*}}(H_{R}).

2.4 Rabinowitz Floer homology

Recall that a Liouville cobordism WW is an exact symplectic manifold with contact-type boundary. A boundary component is positive if the contact orientation agrees with the boundary orientation. Otherwise, it is negative. A filled Liouville cobordism is an exact symplectic manifold MM with positive contact boundary such that

  1. 1.

    WW symplectically embeds into MM and

  2. 2.

    under this embedding, the boundary of MM is the positive boundary of WW.

The completed symplectic cohomology of a domain has a natural extension to a theory for filled cobordisms. This theory was first defined in [c-f-o] for trivial Liouville cobordism with Liouville filling and extended to general Liouville cobordism with Liouville filling in [cieliebak-o]. The definition was extended by the author to a completed theory in [venkatesh]. We briefly recall the construction for a trivial cobordism.

Define a “dual” symplectic homology theory

SC(HR)=n=0CF(Hn)[𝒒].SC_{*}(H_{R})=\prod_{n=0}^{\infty}CF^{*}(-H_{n})[\bm{q}].

As with symplectic cohomology, SC(HR)SC_{*}(H_{R}) admits a filtration by action, given by the chain complexes CF(a,)(Hi)CF^{*}_{(a,\infty)}(H_{i}). Define subcomplexes

SC(a,)(HR)=n=0CF(a,)(Hn)[𝐪]SC_{*}^{(a,\infty)}(H_{R})=\prod_{n=0}^{\infty}CF^{*}_{(a,\infty)}(-H_{n})[{\bf q}]

with homology SH(a,)(HR)SH_{*}^{(a,\infty)}(H_{R}). Action-completed symplectic homology is defined to be

SH^(HR)=H(limaSC(a,)(HR)).\widehat{SH_{*}}(H_{R})=H\left(\lim_{\begin{subarray}{c}\rightarrow\\ a\end{subarray}}SC_{*}^{(a,\infty)}(H_{R})\right).

Let {Hn}\{H_{n}^{\prime}\} be a family of Hamiltonians corresponding to some radius RR^{\prime}, and let {Hn}\{H_{n}\}, as always, be a family of Hamiltonians corresponding to some radius RR. We define a map

𝔠^:SC^(HR)SC^(HR),\mathfrak{\widehat{c}}:\widehat{SC_{*}}(H_{R^{\prime}})\longrightarrow\widehat{SC^{*}}(H_{R}),

as follows. There is a map

SC^(HR)SC(HR)\widehat{SC_{*}}(H_{R^{\prime}})\longrightarrow SC_{*}(H_{R^{\prime}})

that is the direct limit of the inclusion maps

SC(a,)(HR)SC(HR)SC_{*}^{(a,\infty)}(H_{R^{\prime}})\hookrightarrow SC_{*}(H_{R^{\prime}})

and a map

Ψ:SC(HR)CF(H0)\Psi:SC_{*}(H_{R^{\prime}})\longrightarrow CF^{*}(-H_{0}^{\prime})

that is projection onto the first component. Similarly, there is a map

SC(HR)SC^(HR)SC^{*}(H_{R})\longrightarrow\widehat{SC^{*}}(H_{R})

that is the inverse limit of the projection maps

SC(HR)SCa(HR)SC^{*}(H_{R})\twoheadrightarrow SC^{*}_{a}(H_{R})

and a map

Φ:CF(H0)SC(HR)\Phi:CF^{*}(H_{0})\hookrightarrow SC^{*}(H_{R})

that is inclusion into the first component. The map 𝔠^\mathfrak{\widehat{c}} is the composition

SC^(HR){\widehat{SC_{*}}(H_{R^{\prime}})}SC^(HR){\widehat{SC^{*}}(H_{R})}SC(HR){SC_{*}(H_{R^{\prime}})}SC(HR){SC^{*}(H_{R})}CF(H0){CF^{*}(-H_{0}^{\prime})}CF(H0){CF^{*}(H_{0})}Ψ\scriptstyle{\Psi}𝔠\scriptstyle{\mathfrak{c}}c\scriptstyle{c}Φ\scriptstyle{\Phi}

that projects an infinite sum onto the “CF(H0)CF^{*}(-H_{0}^{\prime})” component, maps this component onto CF(H0)CF^{*}(H_{0}) via continuation, and finally includes into the symplectic cochain complex.

Without loss of generality assume that RRR^{\prime}\leq R. Define the symplectic cohomology of the cobordism [R,R]×Σ[R^{\prime},R]\times\Sigma to be the cohomology of the cone of 𝔠\mathfrak{c}, the latter written as

SC([R,R]×Σ)=Cone(𝔠),SC^{*}([R^{\prime},R]\times\Sigma)=Cone(\mathfrak{c}),

and the completed symplectic cohomology to be the cohomology of the cone of 𝔠^\widehat{\mathfrak{c}},

SC^([R,R]×Σ)=Cone(𝔠^).\widehat{SC^{*}}([R^{\prime},R]\times\Sigma)=Cone(\widehat{\mathfrak{c}}).

Denote these cohomology theories by SH([R,R]×Σ)SH^{*}([R^{\prime},R]\times\Sigma), respectively SH^([R,R]×Σ)\widehat{SH^{*}}([R^{\prime},R]\times\Sigma).

  • Remark 8)

    If R>RR^{\prime}>R, we can still follow the above recipe to define a symplectic cohomology theory. We continue to write SH([R,R]×Σ)SH^{*}([R^{\prime},R]\times\Sigma), respectively SH^([R,R]×Σ)\widehat{SH^{*}}([R^{\prime},R]\times\Sigma). These theories are dual to SH([R,R]×Σ)SH^{*}([R,R^{\prime}]\times\Sigma), respectively SH^([R,R]×Σ)\widehat{SH^{*}}([R,R^{\prime}]\times\Sigma).

  • Remark 9)

    Note that, a priori,

    H(SC^(HR))SH^(HR).H\left(\widehat{SC^{*}}(H_{R})\right)\neq\widehat{SH^{*}}(H_{R}).

    The left-hand side is a quotient by im(^)\textnormal{im}(\widehat{\partial}), while the right-hand side is a quotient by im(^)¯\overline{\textnormal{im}(\widehat{\partial})}. In the examples we consider, however, the two will coincide. See Section 5.

  • Remark 10)

    Suppose that EE is exact and ΣE\Sigma\subset E is a convex contact hypersurface as in (10). Σ\Sigma has associated to it a Floer-type invariant called Rabinowitz Floer homology, denoted by RFH(Σ)RFH^{*}(\Sigma). Cieliebak-Frauenfelder-Oancea showed in [c-f-o] that there is an isomorphism

    H(Cone(𝔠))RFH(Σ).H\left(Cone(\mathfrak{c})\right)\simeq RFH(\Sigma).

    In the non-exact case, there is a completed version of Rabinowitz Floer homology associated to a contact hypersurface Σ×{R}\Sigma\times\{R\}, which we denote by RFH^(Σ×{R})\widehat{RFH^{*}}(\Sigma\times\{R\}). This was first studied by Albers-Kang in [albers-k]. In Section 5 we will give examples of scenarios in which, for R<R′′<RR^{\prime}<R^{\prime\prime}<R,

    (19) SH^([R,R]×Σ)RFH^(Σ×{R′′}).\widehat{SH^{*}}([R^{\prime},R]\times\Sigma)\simeq\widehat{RFH^{*}}(\Sigma\times\{R^{\prime\prime}\}).

    There are maps

    SH^([R,R]×Σ)SH^(Σ×[S,S])\widehat{SH^{*}}([R^{\prime},R]\times\Sigma)\rightarrow\widehat{SH^{*}}(\Sigma\times[S^{\prime},S])

    whenever [S,S][R,R][S^{\prime},S]\subset[R^{\prime},R] [cieliebak-o]. We expect the isomorphism (20) to generalize to an isomorphism

    (20) limR<R′′<RSH^([R,R]×Σ)RFH^(Σ×{R′′}).\lim_{\begin{subarray}{c}\rightarrow\\ R^{\prime}<R^{\prime\prime}<R\end{subarray}}\widehat{SH^{*}}([R^{\prime},R]\times\Sigma)\simeq\widehat{RFH^{*}}(\Sigma\times\{R^{\prime\prime}\}).

2.5 Morse-Bott Floer theory and cascades

Thus far we have assumed that all periodic orbits are non-degenerate. In the examples considered in this paper, however, all periodic orbits will be transversely non-degenerate, requiring Morse-Bott techniques. We follow the exposition in [bourgeois-o]. Let HH be a Hamiltonian whose orbits are each either constant and non-degenerate or transversely non-degenerate. For each non-constant orbit x𝒫(H)x\in{\cal P}(H) choose a generic perfect Morse function fx:S1f_{x}:S^{1}\longrightarrow\mathbb{R}. A choice of Ω\Omega-tame almost-complex structure JJ defines cascades: tuples 𝐮=(cm,um,cm1,um1,,u1,c0){\bf u}=(c_{m},u_{m},c_{m-1},u_{m-1},\dots,u_{1},c_{0}) associated to a sequence of orbits xm,xm1,,x0x_{m},x_{m-1},\dots,x_{0} with xm1,,x1x_{m-1},\dots,x_{1} non-constant, such that

  1. 1.

    ciim(xi)c_{i}\in\textnormal{im}(x_{i})

  2. 2.

    uiu_{i} is a finite-energy Floer solution corresponding to the Floer data (H,J)(H,J),

  3. 3.

    limsui(s,0)\lim\limits_{s\rightarrow\infty}u_{i}(s,0) is in the stable manifold of cic_{i} (or ci=limsui(s,t)c_{i}=\lim\limits_{s\rightarrow\infty}u_{i}(s,t) if xix_{i} is constant), and

  4. 4.

    cic_{i} is in the unstable manifold of limsui+1(s,0)\lim\limits_{s\rightarrow-\infty}u_{i+1}(s,0) (or ci=limsui+1(s,t)c_{i}=\lim\limits_{s\rightarrow-\infty}u_{i+1}(s,t) if xix_{i} is constant).

Choose capping discs x~0\tilde{x}_{0} for x0x_{0} and x~m\tilde{x}_{m} for xmx_{m}. The union x~m#um#um1##u1#x~0\tilde{x}_{m}\#-u_{m}\#-u_{m-1}\#\dots\#-u_{1}\#-\tilde{x}_{0} represents a homology class βH2(E)\beta\in H_{2}(E). Let pp and qq be constant orbits of HH or critical points of some functions fxf_{x} and fxf_{x^{\prime}}. The moduli space ^β,m(q,p)\widehat{{\cal M}}_{\beta,m}(q,p) is the space of tuples (cm,um,cm1,um1,,u1,c0)(c_{m},u_{m},c_{m-1},u_{m-1},\dots,u_{1},c_{0}) representing class β\beta such that cmc_{m} is in the stable manifold of pp (or is equal to a constant orbit pp) and c0c_{0} is in the stable manifold of qq (or equal to a constant orbit qq). Each component of ^β,m(q,p)\widehat{{\cal M}}_{\beta,m}(q,p) carries an m\mathbb{R}^{m} action, induced by the \mathbb{R}-actions on each Floer trajectory. By Proposition 3.2 in [bourgeois-o],

β(q,p):=m1^β,m(q,p)/m{\cal M}_{\beta}(q,p):=\bigsqcup_{m\geq 1}{\raisebox{1.99997pt}{$\widehat{{\cal M}}_{\beta,m}(q,p)$}\left/\raisebox{-1.99997pt}{$\mathbb{R}^{m}$}\right.}

is a manifold of the expected dimension. The Floer differential now counts

fl(p)=|q||p|=1βπ2(E)uβ(q,p)±T[Ω](β)q.\partial^{fl}(p)=\sum_{\begin{subarray}{c}|q|-|p|=1\\ \beta\in\pi_{2}(E)\end{subarray}}\sum_{u\in{\cal M}_{\beta}(q,p)}\pm T^{-[\Omega](\beta)}q.

The continuation maps are similarly modified.

Instead of using cascades, one could just generically perturb the Hamiltonian. However, the S1S^{1}-symmetry of the unperturbed Hamiltonian will be useful in the computations in this paper. Our use of cascades is justified by a result by Bourgeois-Oancea, showing the equivalence of the two approaches.

Theorem 1 (Bourgeois-Oancea: Theorem 3.7 in [bourgeois-o])

If HH is a transversally-nondegenerate Hamiltonian, there exists a non-degenerate Hamiltonian HH^{\prime} – a perturbation of HH – such that

SC(H)SC(H)SC^{*}(H)\simeq SC^{*}(H^{\prime})

are chain-isomorphic.

3 Negative line bundles

Let (M,ω)(M,\omega) be a symplectic manifold of dimension 2m22m-2. Denote by E\xlongrightarrowρME\xlongrightarrow{\rho}M the line bundle satisfying c1E=k[ω]c_{1}^{E}=-k[\omega] for some fixed k>0k>0. Such a line bundle is called negative. EE is a symplectic manifold; assume that EE is weak+ monotone. Following [oancea-leray] and [ritter-gromov], we construct a canonical symplectic form on EE through ω\omega.

Let JJ be an ω\omega-compatible complex structure on MM. Let |||\cdot| be a Hermitian metric on EE with induced Chern curvature {\cal F}. Define a radial coordinate rr by r(w)=|w|r(w)=|w| and a fiber-wise angular one-form on the complement of the zero-section by

α=14πkdclog(r2)\alpha=\frac{1}{4\pi k}d^{c}\log(r^{2})

so that

dα=14πkddclog(r2)=i2πk¯log(r2)=i2πkρ1kρc1Eρ[ω].d\alpha=\frac{1}{4\pi k}dd^{c}\log(r^{2})=\frac{i}{2\pi k}\partial\bar{\partial}\log(r^{2})=-\frac{i}{2\pi k}\rho^{*}{\cal F}\equiv-\frac{1}{k}\rho^{*}c_{1}^{E}\equiv\rho^{*}[\omega].

Note that α\alpha defines a contact one-form on the unit circle bundle. Let

Ω:=(1+kπr2)dα+2kπrdrα=dα+d(kπr2α)\Omega:=(1+k\pi r^{2})d\alpha+2k\pi rdr\wedge\alpha=d\alpha+d(k\pi r^{2}\alpha)

be a symplectic form on the complement of the zero section. Extend Ω\Omega smoothly over the zero-section by

Ω|zero section=i2πkρ+{area form of fiber}.\Omega\big{|}_{\text{zero section}}=-\frac{i}{2\pi k}\rho^{*}{\cal F}+\text{\{area form of fiber\}}.

Then Ω\Omega is a symplectic form on EE and [Ω]=[ρω][\Omega]=[\rho^{*}\omega].

We wish to compute the symplectic cohomology of EE. There is a simple family of Hamiltonians to take, which lead to an elegant computation of SH(E)SH^{*}(E). This was developed by Ritter in [ritter-gromov] and [ritter-fano] using the S1S^{1}-action on EE that rotates the fibers. To fit our framework, we modify this construction slightly, and we appeal to [groman] to assert that the two frameworks yield isomorphic homology theories.

Let {gn:}n\{g_{n}:\mathbb{R}\longrightarrow\mathbb{R}\}_{n\in\mathbb{Z}} be a family of functions defined as

gn(r)=(nk+12k)r,g_{n}(r)=\left(\frac{n}{k}+\frac{1}{2k}\right)r,

as depicted in Figure 2. Choose a 𝒞2{\cal C}^{2}-small Morse function f:Mf:M\longrightarrow\mathbb{R} on MM. Define

Gn=gn(kπr2)+(1+kπr2)ρf.G_{n}=g_{n}(k\pi r^{2})+(1+k\pi r^{2})\rho^{*}f.
\mathbb{R}kπr2k\pi r^{2}g0g_{0}g1g_{1}g2g_{2}
Figure 2: The Hamiltonians g0(kπr2),g1(kπr2),g_{0}(k\pi r^{2}),g_{1}(k\pi r^{2}), and g2(kπr2)g_{2}(k\pi r^{2})

Recall that, away from the zero section,

Ω=(1+kπr2)dα+2kπrdrα\Omega=(1+k\pi r^{2})d\alpha+2k\pi rdr\wedge\alpha

Let XfhX_{f}^{h} be the horizontal lift of the Hamiltonian vector field XfX_{f} on MM, uniquely defined through the connection one-form α\alpha. As

dGn\displaystyle dG_{n} =2kπr(gn(kπr2)+ρf)dr+(1+kπr2)ρdf\displaystyle=2k\pi r(g_{n}^{\prime}(k\pi r^{2})+\rho^{*}f)dr+(1+k\pi r^{2})\rho^{*}df
=2kπr(nk+12k+ρf)dr+(1+kπr2)ρdf,\displaystyle=2k\pi r\left(\frac{n}{k}+\frac{1}{2k}+\rho^{*}f\right)dr+(1+k\pi r^{2})\rho^{*}df,

the Hamiltonian vector field of GnG_{n} away from the zero section is

XGn=(nk+12k+ρf)Rα+Xfh.X_{G_{n}}=\left(\frac{n}{k}+\frac{1}{2k}+\rho^{*}f\right)R_{\alpha}+X_{f}^{h}.

The Reeb orbits of α\alpha have period 1k\frac{1}{k}, and so the periodic orbits of XGnX_{G_{n}} exist only where nk+12k+ρf\frac{n}{k}+\frac{1}{2k}+\rho^{*}f is an integer multiple of 1k\frac{1}{k}. By construction, ρf\rho^{*}f is 𝒞2{\cal C}^{2}-small, and so XGnX_{G_{n}} has no periodic orbits away from the zero section. On the zero section XGn=XfhX_{G_{n}}=X_{f}^{h}, which can be identified with XfX_{f}. Thus, the periodic orbits of XGnX_{G_{n}} correspond to the periodic orbits of XfX_{f}. As ff is 𝒞2{\cal C}^{2} small, these correspond precisely to the critical points of the Morse function ff. It follows that there is a vector space isomorphism

CF(Gn)CM(f)CF^{*}(G_{n})\simeq CM^{*}(f)

for every nn\in\mathbb{Z}.

Suppose that σ:S1Ham(E,id)\sigma:S^{1}\longrightarrow Ham(E,id) is a loop of Hamiltonians based at the identity. Then σ\sigma acts on the loopspace (E){\cal L}(E) by σx(t)=σ(t)(x(t))\sigma^{*}x(t)=\sigma(t)(x(t)). This lifts to an action σ~\tilde{\sigma} on a cover of the loopspace (E)~\widetilde{{\cal L}(E)}. We fix (E)~\widetilde{{\cal L}(E)} to be the cover defined by the deck transformation group

Γ=ker([Ω]|π2(E))ker([c1TE(E)]|π2(E)).\Gamma=\ker([\Omega]\big{|}_{\pi_{2}(E)})\cap\ker([c_{1}^{TE}(E)]\big{|}_{\pi_{2}(E)}).

In other words, (E)~\widetilde{{\cal L}(E)} is the group of cappings of each loop, under the equivalence relation uvu\simeq v if

  • u=v\partial u=\partial v,

  • [Ω]([u#(v)])=0[\Omega]([u\#(-v)])=0, and

  • c1TE([u#(v)])=0c_{1}^{TE}([u\#(-v)])=0.

Let σt\sigma_{t} be the action on EE that rotates each fiber by e2πite^{2\pi it}. This is a Hamiltonian action generated by the Hamiltonian 1k(kπr2)\frac{1}{k}(k\pi r^{2}), preserving the radial coordinate kπr2k\pi r^{2}, and so

σtGn:=Gnσt(1kkπr2)σt=(Gn1k(kπr2))σt=Gn1.\sigma_{t}^{*}G_{n}:=G_{n}\circ\sigma_{t}-\left(\frac{1}{k}k\pi r^{2}\right)\circ\sigma_{t}=\left(G_{n}-\frac{1}{k}(k\pi r^{2})\right)\circ\sigma_{t}=G_{n-1}.

Define the lift σ~\tilde{\sigma} to preserve cappings of constant loops.

Let JtJ_{t} be any one-parameter family of Ω\Omega-tame almost-complex structures. The action of σt\sigma_{t} on JtJ_{t} defined by

σtJt=dσt1Jtdσt,\sigma_{t}^{*}J_{t}=d\sigma_{t}^{-1}\circ J_{t}\circ d\sigma_{t},

produces another one-parameter family of Ω\Omega-tame almost complex structures. Recall that Floer data (H,J)(H,J) is generic if the Floer cochain complex CF(H,J)CF^{*}(H,J) is well-defined. The following two theorems, due to Ritter, yield the promised computation of symplectic cohomology (see Section 7 of [ritter-gromov]).

Theorem 5 (Ritter [ritter-gromov])

If (Gn,J)(G_{n},J) is generic, then so is (σtGn,σtJ)(\sigma_{t}^{*}G_{n},\sigma_{t}^{*}J). The action σ~\tilde{\sigma} induces a chain isomorphism

𝒮:CF(Gn,J)CF+2(σGn,σJ)=CF+2(Gn1,σJ).{\cal S}:CF^{*}(G_{n},J)\longrightarrow CF^{*+2}(\sigma^{*}G_{n},\sigma^{*}J)=CF^{*+2}(G_{n-1},\sigma^{*}J).

Choose continuation maps

cGn:CF(Gn,J)CF(Gn+1,J)c^{G}_{n}:CF^{*}(G_{n},J)\longrightarrow CF^{*}(G_{n+1},J^{\prime})

between Floer data (Gn,J)(G_{n},J) and (Gn+1,J)(G_{n+1},J^{\prime}).

Theorem 6 (Ritter [ritter-gromov])

The following diagram induces a commutative diagram on the level of cohomology.

CF(Gn1,J){CF^{*}(G_{n-1},J)}CF(Gn,J){CF^{*}(G_{n},J^{\prime})}CF+2n(G1,(σ)nJ){CF^{*+2n}(G_{-1},(\sigma^{*})^{n}J)}CF+2n(G0,(σ)nJ){CF^{*+2n}(G_{0},(\sigma^{*})^{n}J^{\prime})}cn1G\scriptstyle{c_{n-1}^{G}}𝒮n\scriptstyle{{\cal S}^{n}}𝒮n\scriptstyle{{\cal S}^{n}}c1G\scriptstyle{c_{-1}^{G}}
Corollary 2 (Ritter [ritter-gromov])

The symplectic cochain complex

SC(G):=n=0CF(Gn)[𝐪]SC^{*}(G):=\bigoplus_{n=0}^{\infty}CF^{*}(G_{n})[{\bf q}]

is quasi-isomorphic to a complex

SC(G0):=n=0CF(G0)[2n][𝐪]SC^{*}(G_{0}):=\bigoplus_{n=0}^{\infty}CF^{*}(G_{0})[-2n][{\bf q}]

with differential

(x+y𝐪)=fl(x)+yc1G𝒮(y)+fl(y)𝐪.\partial(x+y{\bf q})=\partial^{fl}(x)+y-c_{-1}^{G}\circ{\cal S}(y)+\partial^{fl}(y){\bf q}.

Note that the grading shift [2n][-2n] increases the Conley-Zehnder index by 2n2n.

Denote the 0th generalized eigenspace of the map

𝒄𝟏𝑮𝓢:HF(H0)HF(H0).\bm{c_{-1}^{G}}\circ\bm{{\cal S}}:HF^{*}(H_{0})\longrightarrow HF^{(}H_{0}).

by HF0(H0)HF^{*}_{0}(H_{0}). A linear algebra consequence of Corollary 2 is

Corollary 3 (Ritter [ritter-gromov])

The uncompleted symplectic cohomology of EE is

SH(E)=SH(G)SH(G0)HF(H0)/HF0(H0).SH^{*}(E)=SH^{*}(G)\cong SH^{*}(G_{0})\cong{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$HF^{*}_{0}(H_{0})$}\right.}.

Under this isomorphism, the map

HF(H0)SH(E)HF^{*}(H_{0})\longrightarrow SH^{*}(E)

induced by inclusion on cochains is the quotient map.

Finally, symplectic cohomology can be rephrased completely in terms of topological information. There is a map, termed the P.S.S. map after its creators Piukhin-Salamon-Schwartz, that identifies the quantum cohomology QH(E)QH^{*}(E) with HF(H0)HF^{*}(H_{0}).

Theorem 7 (Ritter [ritter-gromov])

The P.S.S. isomorphism

QH(E)HF(H0)QH^{*}(E)\cong HF^{*}(H_{0})

identifies up to non-zero scalar η\eta the map 𝐜𝟏𝐆𝓢\bm{c_{-1}^{G}}\circ\bm{{\cal S}} with the action of quantum cup product \cup_{*} with ρc1E\rho^{*}c_{1}^{E}. If QH0(E)QH^{*}_{0}(E) is the 0th-generalized eigenspace of the map xρc1Exx\mapsto\rho^{*}c_{1}^{E}\cup_{*}x, there is an isomorphism of Λ\Lambda-algebras

SH(E)QH(E)/QH0(E).SH^{*}(E)\cong{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$QH^{*}_{0}(E)$}\right.}.

Indeed, Ritter shows that the scalar η\eta has valuation ev(η)=0ev(\eta)=0 (see Theorem 67 of [ritter-gromov]).

Corollary 4

The P.S.S. isomorphism identifies a λ\lambda-generalized eigenspace of 𝐜𝟏𝐆𝓢𝟏\bm{c_{-1}^{G}}\circ\bm{{\cal S}^{-1}} with a ηλ\eta\lambda-generalized eigenspace of ρc1E\rho^{*}c_{1}^{E}\cup_{*}-, and

ev(λ)=ev(ηλ).ev(\lambda)=ev(\eta\lambda).

3.1 Disk subbundles

We now use reduced symplectic cohomology to define an invariant of a disk subbundle contained in EE. Fix a radius R(0,)R\in(0,\infty), and let DRD_{R} be the disk bundle of radius RR. As in Subsection 2.2, we will construct a chain complex generated by one-periodic orbits that cluster near the boundary of DRD_{R}. We use a very precise family of Hamiltonians, as well as almost-complex structures, to compute SH¯(DR)\overline{SH^{*}}(D_{R}).

3.1.1 The almost-complex structure

Let 𝒥(ωstd){\cal J}(\omega_{std}) be the space of S1S^{1}-families of almost-complex structures on \mathbb{C} compatible with the standard symplectic form, and such that each almost-complex structure is cylindrical in a neighborhood of periodic orbits and at infinity. Let 𝒥(ω){\cal J}(\omega) be the space of S1S^{1}-families of almost-complex structures on MM compatible with ω\omega. Choose it𝒥(ωstd)i_{t}\in{\cal J}(\omega_{std}) and jt𝒥(ω)j_{t}\in{\cal J}(\omega). The one-form α\alpha determines a splitting of TETE into a vertical component VV\cong\mathbb{C} and horizontal component HH. Let Lt(H,V)L_{t}(H,V) be the space of S1S^{1}-families of linear maps from HH to VV. Let 𝒰{\cal U} be an open set comprised of small neighborhoods of the circle bundles on which live non-constant periodic orbits of each HnH_{n}, as well as small neighborhoods of the constant orbits. Let 𝔅(it,jt)\mathfrak{B}(i_{t},j_{t}) be the elements BtLt(H,V)B_{t}\in L_{t}(H,V) with compact support in the complement of 𝒰{\cal U}, and satisfying itBt+Btρjt=0i_{t}B_{t}+B_{t}\rho^{*}j_{t}=0 for all tt. We will use this data to define a set of almost-complex structures. These conditions will ensure that Floer trajectories “flow outward”, that Floer trajectories converge to periodic orbits, and that JtJ_{t} is almost-complex. Define

𝒥(Ω)={Jt=[itBt0ρjt]End(TE)|it𝒥(ωstd),jt𝒥(ω),Bt𝔅(it,jt),Jt is Ω–tame}.{\cal J}(\Omega)=\left\{J_{t}=\left[\begin{array}[]{cc}i_{t}&B_{t}\\ 0&\rho^{*}j_{t}\end{array}\right]\in End(TE)\hskip 2.84544pt\bigg{|}\hskip 2.84544pti_{t}\in{\cal J}(\omega_{std}),j_{t}\in{\cal J}(\omega),B_{t}\in\mathfrak{B}(i_{t},j_{t}),J_{t}\text{ is }\Omega\text{--tame}\right\}.

Transversality and regularity for J𝒥(Ω)J\in{\cal J}(\Omega) was proven by Albers-Kang in [albers-k] for the completely analogous case of Floer solutions in Rabinowitz Flor homology. We therefore omit these proofs, and refer to [albers-k].

Lemma 1 (Albers-Kang: Proposition 2.11 in [albers-k]))

There exists a comeager subset of 𝒥(Ω){\cal J}(\Omega) for which the finite-energy cascades of HnH_{n} and HnsH_{n}^{s}, for any nn, are cut out transversally.

Lemma 2 (Albers-Kang: Lemma 2.15 in [albers-k])

All simple JJ-holomorphic spheres are regular.

3.1.2 The Hamiltonians

Fix a constant C>0C>0 and let {hn:}n\left\{h_{n}\colon\mathbb{R}\longrightarrow\mathbb{R}\right\}_{n\in\mathbb{N}} be a family of functions where each hnh_{n} is

  1. 1.

    convex and monotone increasing on 0\mathbb{R}_{\geq 0},

  2. 2.

    bounded in absolute value by CC on [0,kπR2][0,k\pi R^{2}], and

  3. 3.

    of slope nk+12k\frac{n}{k}+\frac{1}{2k} on (kπRn2,)(k\pi R_{n}^{2},\infty), for some Rn<RR_{n}<R.

Further assume that the sequence {Rn}\{R_{n}\} tends to RR as nn tends to \infty. To simplify later proofs, we also choose hnh_{n} such that

  1. 1.

    hnh_{n} and hnh_{n}^{\prime} are monotone increasing on 0\mathbb{R}_{\geq 0} and

  2. 2.

    hn=hn+1h_{n}=h_{n+1} on [0,kπRn2][0,k\pi R_{n}^{2}].

Choose a Morse function function f:Mf\colon M\longrightarrow\mathbb{R} that is 𝒞2{\cal C}^{2}-small. Define a family of Hamiltonians {Hn:E}n\left\{H_{n}\colon E\longrightarrow\mathbb{R}\right\}_{n\in\mathbb{N}} by

Hn=hn(kπr2)+(1+kπr2)ρf.H_{n}=h_{n}(k\pi r^{2})+(1+k\pi r^{2})\rho^{*}f.

We assume that the one-periodic orbits of HnH_{n} are transversally nondegenerate. For example, we can take each hnh_{n} to be a smoothing of a piecewise linear function, as in Figure 3.

\mathbb{R}kπr2k\pi r^{2}h0h_{0}h1h_{1}h2h_{2}h3h_{3}...
Figure 3: A family of Hamiltonians {hn(kπr2)}\{h^{n}(k\pi r^{2})\}

Let RαR_{\alpha} be the unique vector field on E{zero section}E\setminus\{\text{zero section}\} satisfying

{α(Rα)=1dr(Rα)=0iRαdα=0.\left\{\begin{array}[]{c}\alpha(R_{\alpha})=1\\ dr(R_{\alpha})=0\\ i_{R_{\alpha}}d\alpha=0\end{array}\right..

Note that Rα|r=𝔯R_{\alpha}\big{|}_{r=\mathfrak{r}} is the Reeb vector field of the contact form α|{r=𝔯}\alpha\big{|}_{\{r=\mathfrak{r}\}} and the simply-covered orbits of RαR_{\alpha} have period 1k\frac{1}{k}. As

dHn=2kπr(hn(kπr2)+ρf)dr+(1+kπr2)ρdf,dH_{n}=2k\pi r(h_{n}^{\prime}(k\pi r^{2})+\rho^{*}f)dr+(1+k\pi r^{2})\rho^{*}df,

the Hamiltonian vector field of HnH_{n} is

XHn=(hn(kπr2)+ρf)Rα+Xfh,X_{H_{n}}=(h_{n}^{\prime}(k\pi r^{2})+\rho^{*}f)R_{\alpha}+X_{f}^{h},

Thus, the one-periodic orbits of HnH_{n} correspond bijectively to

  1. 1.

    the S1S^{1}-families of orbits of Rα|DRR_{\alpha}\big{|}_{D_{R}} with period between 1k\frac{1}{k} and 1n\frac{1}{n}, lying in fibers above the critical points of ff, and

  2. 2.

    the critical points of ff itself.

Impose a perfect Morse function on each S1S^{1}-family of orbits, so that each family gives rise to two distinguished orbits: the minimum and the maximum of the perfect Morse function.

Denote by 𝒫(Hn){\cal P}(H_{n}) the union of all minimum and maximum distinguished orbits of HnH_{n}, in addition to the constant orbits, and by 𝒫(H){\cal P}(H) the union n𝒫(Hn)\bigcup\limits_{n}{\cal P}(H_{n}).

Choose generic almost-complex structures in 𝒥(Ω){\cal J}(\Omega) to define the Floer complexes CF(Hn)CF^{*}(H_{n}). For each nn choose a generic \mathbb{R}-family of functions hns:h_{n}^{s}\colon\mathbb{R}\rightarrow\mathbb{R}, monotonely decreasing in ss, with hns=hnh_{n}^{s}=h_{n} when s>>0s>>0 and hns=hn+1h_{n}^{s}=h_{n+1} when s<<0s<<0. Set Hns=hns(kπr2)+(1+kπr2)ρfH_{n}^{s}=h_{n}^{s}(k\pi r^{2})+(1+k\pi r^{2})\rho^{*}f. Define continuation maps

cn:CF(Hn)CF(Hn+1)c_{n}:CF^{*}(H_{n})\rightarrow CF^{*}(H_{n+1})

through HnsH_{n}^{s}. In the notation of Section 2, this collection of data defines cochain complexes SC(H),SC¯(H),SC^{*}(H),\overline{SC^{*}}(H), and SC^(H)\widehat{SC^{*}}(H).

3.1.3 The Floer differential is well-defined

The almost-complex structures we use are in a restricted form; we check that the Floer complexes are still well-defined. Transversality and regularity for J𝒥(Ω)J\in{\cal J}(\Omega) was proven by Albers-Kang in [albers-k] for the completely analogous case of Floer solutions in Rabinowitz Flor homology. We therefore omit these proofs, and refer to [albers-k].

Lemma 3 (Albers-Kang: Proposition 2.11 in [albers-k]))

There exists a comeager subset of 𝒥(Ω){\cal J}(\Omega) for which the finite-energy cascades of HnH_{n} and HnsH_{n}^{s}, for any nn, are cut out transversally.

Lemma 4 (Albers-Kang: Lemma 2.15 in [albers-k])

All simple JJ-holomorphic spheres are regular.

It suffices for us to prove compactness. We must show that Floer solutions do not escape to infinity and that bubbling cannot occur at degenerations of moduli spaces.

Lemma 5

Sequences of cascades of HnH_{n} or HnsH_{n}^{s} between two fixed periodic orbits remain in a compact region of EE, for any nn.

Proof.

This follows from an integrated maximum principle, as in the proof of Lemma 7. Indeed, by assuming that iti_{t} is cylindrical and Bt=0B_{t}=0 outside of DRD_{R}, the integrated maximum principle tells us that all Floer solutions appearing in a cascade remain in DRD_{R}.

Lemma 6

Bubbling does not generically occur in the limit of sequences of index-0, index-1 and index-2 cascades defining the chain complex.

Proof.

Let v:ΣEv\colon\Sigma\longrightarrow E be a non-constant JJ-holomorphic sphere of index 22. As the symplectic form on EE is exact away from the zero-section, Σ\Sigma must intersect the zero-section (else, by Stokes’ Theorem, Σ\Sigma would have zero symplectic area, contradicting that vv is non-constant). Assume for contradiction that Σ\Sigma leaves the zero-section. Recall that we chose iti_{t} cylindrical in a neighborhood of each circle bundle containing a non-constant periodic orbit. Choose a generic disc bundle 𝒟{\cal D} containing no non-constant periodic orbits, but such that iti_{t} is cylindrical on 𝒟\partial{\cal D}. Suppose that vv leaves 𝒟{\cal D}. Apply the integrated maximum principle from [abouzaid] to v1(E𝒟)v^{-1}(E\setminus{\cal D}). This computation shows that the symplectic area of v1(E𝒟)v^{-1}(E\setminus{\cal D}) is negative, contradicting JJ-holomorphicity. Thus, all JJ-holomorphic spheres are constrained to lie on the zero section.

Denote by c(J){\cal M}_{c}(J) the moduli space of simple JJ-holomorphic curves uu with c1TE(u)=cc_{1}^{TE}(u)=c, modded out by the group of biholomorphic automorphisms. By weak+ monotonicity, all JJ-holomorphic spheres have non-negative first Chern class. Thus, the only moduli spaces that can cause bubbling to index-0, -1, or -2 cascades are 0(J){\cal M}_{0}(J) and 2(J){\cal M}_{2}(J). The index formula says that

dim(c(J))=2(m+1)+2c6.dim({\cal M}_{c}(J))=2(m+1)+2c-6.

As we consider S1S^{1} families of almost-complex structures, the spheres in 0(Jt){\cal M}_{0}(J_{t}) sweep out a codimension-44 set in E×S1E\times S^{1}. The relevant moduli spaces of Floer trajectories sweep out a set of dimension at m 33 in E×S1E\times S^{1}. Thus, the space of relevant Floer trajectories is generically disjoint from 0(Jt){\cal M}_{0}(J_{t}). But any bubble must intersect a limiting (broken) Floer trajectory [salamon]. We conclude that any bubble belonging to 0(Jt){\cal M}_{0}(J_{t}) must intersect a periodic orbit. By the first paragraph, the bubble must be contained in the zero-section MM and therefore intersect a constant orbit. 0(Jt){\cal M}_{0}(J_{t}) has codimension 22 in M×S1M\times S^{1}, and constant orbits have dimension zero. A generic perturbation of the Morse function ff therefore displaces the constant orbits from any such bubbles. We conclude that bubbling cannot occur through Chern-0 spheres.

The Chern-22 bubbles will not be seen by moduli spaces of cascades of index less than 2. These moduli spaces include those appearing in the arguments showing that

  • the differential is well-defined as a map on vector spaces (index 1);

  • continuation maps are well-defined as maps on vector spaces (index 0);

  • continuation maps are chain maps (index 0 and index 1);

  • continuation maps are invariant under choice of underlying Hamiltonian and almost-complex families (index 0, or, at non-generic points in an interpolation between two families, perhaps index ±1\pm 1); and

  • choosing h0h_{0} small enough, the Floer differential on CF(H0)CF^{*}(H_{0}) is canonically identified with the Morse differential (index 11).

It remains to consider the index-22 moduli spaces of cascades that appear in showing 2=0\partial^{2}=0. Let 2(x,y){\cal M}_{2}(x,y) be the two-dimensional component of the moduli space of cascades connecting period orbits xx and yy, and associated with generic Floer data (Hn,J)(H_{n},J). Suppose that bubbling occurs within the moduli space 2(x,y){\cal M}_{2}(x,y). By an argument in [salamon], any bubble must intersect a dimension-0 component of (x,y){\cal M}(x,y); in particular, x=yx=y and any bubble passes through xx.

It follows from the first paragraph that xx is a constant orbit. Usually one could argue that an S1S^{1}-family of JtJ_{t}-holomorphic spheres have codimension two in E×S1E\times S^{1}, and therefore do not generically intersect a zero-dimensional constant orbit. However, the requirement Bt=0B_{t}=0 in a neighborhood of the constant orbits means that we cannot necessarily perturb an almost-complex structure in a direction required to “push” it off of a constant orbit, and we cannot perturb HnH_{n} without risking breaking the nice structure coming from the Morse function ff. We will instead show that no sequence of maps in 2(x,x){\cal M}_{2}(x,x) converges to a broken trajectory. This will show that, even if 2(x,x){\cal M}_{2}(x,x) “sees” bubbling, it does not affect the computation 2=0\partial^{2}=0.

Suppose that a sequence of cascades ui2(x,x)u_{i}\in{\cal M}_{2}(x,x) converges to a broken cascade u1(x,z)×1(z,x)u\in{\cal M}_{1}(x,z)\times{\cal M}_{1}(z,x). By Lemma 7, zz is a constant orbit as well, and all curves remain in a region where the Floer dynamics are governed by the 𝒞2{\cal C}^{2} small function H0H_{0}. In this region, 1(x,z){\cal M}_{1}(x,z) and 1(z,x){\cal M}_{1}(z,x) are moduli spaces of Morse trajectories, and so one of these spaces is empty. Thus, 1(x,z)×1(z,x){\cal M}_{1}(x,z)\times{\cal M}_{1}(z,x) is empty as well.

  • Remark 11)

    If MM is toric and ff is perfect, one can also just conclude that 1(x,z){\cal M}_{1}(x,z) and 1(z,x){\cal M}_{1}(z,x) are both empty for degree reasons. If MM is not toric, one can get around using almost-complex structures in 𝒥(Ω){\cal J}(\Omega). The only place we use this is in Lemma 15. An alternative approach is to “shrink” the size of the domain so that all periodic orbits occur at very small radius, and then take the continuation map to CF(Gn)CF^{*}(G_{n}). Because the domain is quite small, the value of the Novikov exponents appearing are approximately the change in action, and so, in a limit of “shrunk” Floer data, they are all non-negative. Then argue that this continuation map coincides with the continuation map 𝔠HG\mathfrak{c}^{HG}, and so the Novikov exponents contributing to 𝔠HG\mathfrak{c}^{HG} are also non-negative.

Corollary 5

The differential on symplectic cohomology is well-defined, and the resulting homology theory does not depend upon choices.

  • Remark 12)

    It is not clear to what extent the completed symplectic cohomology SH^(H)\widehat{SH^{*}}(H) depends on the choice of Morse perturbation ff. On the other hand, reduced symplectic cohomology is independent of this choice.

3.1.4 Some technical lemmas for computation

In order to compute SH¯(H)\overline{SH^{*}}(H), we need three standard results on the behavior of the Floer differential and continuation maps.

Define 𝔴(x)\mathfrak{w}(x) to be the winding number of a non-constant periodic orbit xx, viewed as a map from the circle into \mathbb{C}^{*}. Define 𝔴(x)\mathfrak{w}(x) of a constant orbit xx to be zero.

Lemma 7

Let uu be a solution of Floer’s equation (8) with lims±u(s,t)=x±(t)\lim\limits_{s\rightarrow\pm\infty}u(s,t)=x_{\pm}(t). Then 𝔴(x+)𝔴(x)\mathfrak{w}(x_{+})\geq\mathfrak{w}(x_{-}).

Proof.

We use the integrated maximum principal of [abouzaid]. Assume for contradiction that 𝔴(x)>𝔴(x+)\mathfrak{w}(x_{-})>\mathfrak{w}(x_{+}). Say that x±x_{\pm} lives in the sphere bundle of radius σ±\sigma_{\pm}. If x±x_{\pm} are non-constant then the RαR_{\alpha}-orbit underlying x±x_{\pm} has period 1k𝔴(x±)\frac{1}{k}\mathfrak{w}(x_{\pm}). From the earlier computation XHn=(hn(kπr2)+ρf)Rα+XfhX_{H_{n}}=(h_{n}^{\prime}(k\pi r^{2})+\rho^{*}f)R_{\alpha}+X_{f}^{h} it follows that hn(kπσ±2)+ρf(x±)=1k𝔴(x±)h_{n}^{\prime}(k\pi\sigma_{\pm}^{2})+\rho^{*}f(x_{\pm})=\frac{1}{k}\mathfrak{w}(x_{\pm}). The winding numbers are integers, and so

hn(kπσ2)+ρf(x)hn(kπσ+2)+ρf(x+)+1k.h_{n}^{\prime}(k\pi\sigma_{-}^{2})+\rho^{*}f(x_{-})\geq h_{n}^{\prime}(k\pi\sigma_{+}^{2})+\rho^{*}f(x_{+})+\frac{1}{k}.

By the smallness of ff we can assume that

hn(kπσ2)hn(kπσ+2)ρf(x+)ρf(x)+1k>0.h_{n}^{\prime}(k\pi\sigma_{-}^{2})-h_{n}^{\prime}(k\pi\sigma_{+}^{2})\geq\rho^{*}f(x_{+})-\rho^{*}f(x_{-})+\frac{1}{k}>0.

As hnh_{n} is convex, we deduce that σ>σ+\sigma_{-}>\sigma_{+}. If x+x_{+} is constant and xx_{-} is non-constant it follows immediately that σ>σ+=0\sigma_{-}>\sigma_{+}=0.

Choose a generic circle subbundle 𝒮{\cal S} of radius σ\sigma, with σ+<σ<σ\sigma_{+}<\sigma<\sigma_{-}, and on which B=0B=0 and iti_{t} is cylindrical. For example, if σ\sigma is close to σ+\sigma_{+} or σ\sigma_{-} these conditions will, by construction, be met. Let 𝒟{\cal D} be the region bounded by 𝒮{\cal S}, and denote Σ=u1(E𝒟)\Sigma=u^{-1}(E\setminus{\cal D}). Let v:ΣE𝒟v\colon\Sigma\longrightarrow E\setminus{\cal D} be the restriction of uu. We will equate Σ\Sigma with its image under the inclusion into ×S1\mathbb{R}\times S^{1} and use the coordinates (s,t)(s,t) induced on Int(Σ)\textnormal{Int}(\Sigma).

Let cx=hn(kπx2)hn(kπx2)kπx2c_{x}=h_{n}(k\pi x^{2})-h^{\prime}_{n}(k\pi x^{2})k\pi x^{2} be the yy-intercept of the tangent line to hn(kπr2)h_{n}(k\pi r^{2}) at kπx2k\pi x^{2}. Then on 𝒮{\cal S}, Hn=hn(kπσ2)kπσ2+(1+kπσ2)ρf+cσH_{n}=h_{n}^{\prime}(k\pi\sigma^{2})k\pi\sigma^{2}+(1+k\pi\sigma^{2})\rho^{*}f+c_{\sigma} (and XHn=(hn(kπσ2)+ρf)Rα+XfhX_{H_{n}}=(h_{n}^{\prime}(k\pi\sigma^{2})+\rho^{*}f)R_{\alpha}+X_{f}^{h}). In particular,

Hn|𝒮=(1+kπσ2)α(XHn)hn(kπσ2)+cσ.H_{n}\big{|}_{{\cal S}}=(1+k\pi\sigma^{2})\alpha(X_{H_{n}})-h_{n}^{\prime}(k\pi\sigma^{2})+c_{\sigma}.

Let Σ+\partial\Sigma_{+} be the union of the boundary components of Σ\Sigma mapping into 𝒮{\cal S}. Note that we have chosen JJ to be Ω\Omega-tame, so that

EJ(v):=12Σ(Ω(sv,Jsv)+Ω(tvXHn(v),J(tvXHn(v)))dsdt0E_{J}(v):=\frac{1}{2}\int_{\Sigma}\left(\Omega(\partial_{s}v,J\partial_{s}v\right)+\Omega\left(\partial_{t}v-X_{H_{n}}(v),J(\partial_{t}v-X_{H_{n}}(v))\right)ds\wedge dt\geq 0

Shuffling terms, we have

EJ(v)=ΣvΩvdHndt=Σ+(1+kπr2)vαHn(v(s,t))dtS1(1+kπr2)xαHn(x(t))dt=Σ+(1+kπσ2)vα(1+kπσ2)α(XHn)dt+(hn(kπσ2)cσ)dtS1(1+kπσ2)xαHn(x(t))dt\begin{split}E_{J}(v)&=\int_{\Sigma}v^{*}\Omega-v^{*}dH_{n}\otimes dt\\ &=\int_{\partial\Sigma_{+}}(1+k\pi r^{2})v^{*}\alpha-H_{n}(v(s,t))dt-\int_{S^{1}}(1+k\pi r^{2})x_{-}^{*}\alpha-H_{n}(x_{-}(t))dt\\ &=\int_{\partial\Sigma_{+}}(1+k\pi\sigma^{2})v^{*}\alpha-(1+k\pi\sigma^{2})\alpha(X_{H_{n}})\otimes dt+\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt\\ &\hskip 56.9055pt-\int_{S^{1}}(1+k\pi\sigma_{-}^{2})x_{-}^{*}\alpha-H_{n}(x_{-}(t))dt\end{split}

We consider this final equation in pieces. A solution v(s,t)v(s,t) of Floer’s equation satisfies

(dvXHndt)(0,1)=0.(dv-X_{H_{n}}\otimes dt)^{(0,1)}=0.

We have that, on 𝒮{\cal S}, drit=2kπrαdr\circ i_{t}=-2k\pi r\alpha and Bt=0B_{t}=0. The former implies that JRαJR_{\alpha} is proportional to r\partial_{r} and the latter implies that JXfhJX_{f}^{h} lives in the horiziontal distribution. So altogether, α(JXHn)=0\alpha(JX_{H_{n}})=0. Thus, the “Σ+\partial\Sigma_{+}” terms become

Σ+(1+kπσ2)α(dvXHndt)+\displaystyle\int_{\partial\Sigma_{+}}(1+k\pi\sigma^{2})\alpha(dv-X_{H_{n}}\otimes dt)+ (hn(kπσ2)cσ)dt\displaystyle\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt
=Σ+(1+kπσ2)αJ(dvXHndt)j+(hn(kπσ2)cσ)dt\displaystyle=\int_{\partial\Sigma_{+}}-(1+k\pi\sigma^{2})\alpha\circ J(dv-X_{H_{n}}\otimes dt)\circ j+\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt
=Σ+1+kπσ22kπσdrdvj+(hn(kπσ2)cσ)dt\displaystyle=\int_{\partial\Sigma_{+}}-\frac{1+k\pi\sigma^{2}}{2k\pi\sigma}dr\circ dv\circ j+\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt
Σ+(hn(kπσ2)cσ)dt,\displaystyle\leq\int_{\partial\Sigma_{+}}\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt,

where the last inequality follows because a vector ζ\zeta that is positively-oriented with respect to the boundary orientation satisfies drj(ζ)0dr\circ j(\zeta)\geq 0. We also have

Hn(x(t))=hn(kπσ2)kπσ2+cσ+(1+kπσ2)ρf(x(t))=(1+kπσ2)𝔴(x)khn(kπσ2)+cσH_{n}(x_{-}(t))=h_{n}^{\prime}(k\pi\sigma_{-}^{2})k\pi\sigma_{-}^{2}+c_{\sigma_{-}}+(1+k\pi\sigma_{-}^{2})\rho^{*}f(x_{-}(t))=(1+k\pi\sigma_{-}^{2})\frac{\mathfrak{w}(x_{-})}{k}-h_{n}^{\prime}(k\pi\sigma^{2})+c_{\sigma_{-}}

and so the “xx_{-}” terms become

S1(1+kπr2)xα+01Hn(x(t))dt\displaystyle-\int_{S^{1}}(1+k\pi r^{2})x_{-}^{*}\alpha+\int_{0}^{1}H_{n}(x_{-}(t))dt =𝔴(x)k(1+kπσ2)+𝔴(x)k(1+kπσ2)+cσhn(kπσ2)\displaystyle=-\frac{\mathfrak{w}(x_{-})}{k}(1+k\pi\sigma_{-}^{2})+\frac{\mathfrak{w}(x_{-})}{k}(1+k\pi\sigma_{-}^{2})+c_{\sigma_{-}}-h_{n}^{\prime}(k\pi\sigma^{2}_{-})
=S1(hn(kπσ2)+cσ)dt\displaystyle=\int_{S^{1}}\left(-h_{n}^{\prime}(k\pi\sigma_{-}^{2})+c_{\sigma_{-}}\right)dt

Altogether,

(21) EJ(v)\displaystyle E_{J}(v) Σ+(hn(kπσ2)cσ)dt+S1(hn(kπσ2)+cσ)dt.\displaystyle\leq\int_{\partial\Sigma_{+}}\left(h^{\prime}_{n}(k\pi\sigma^{2})-c_{\sigma}\right)dt+\int_{S^{1}}\left(-h_{n}^{\prime}(k\pi\sigma_{-}^{2})+c_{\sigma_{-}}\right)dt.

Σ\Sigma is a collection of bounded regions in \mathbb{C}^{*} and one unbounded region enclosing the origin. As dtdt is exact on \mathbb{C}^{*}, the bounded regions contribute nothing to the right-hand side of (21). Let Σ~\tilde{\Sigma} be the unbounded component. Near zero, Σ\Sigma is contained in a neighborhood of xx_{-}, and so all boundary components of Σ~\tilde{\Sigma} occur within the intersection of Σ\Sigma with some annulus [R,)×S1[R,\infty)\times S^{1}. Let 𝔥\mathfrak{h} be a function on Σ~\tilde{\Sigma} that is equal to hn(kπσ2)cσh_{n}^{\prime}(k\pi\sigma^{2})-c_{\sigma} on Σ~([R,)×S1)\tilde{\Sigma}\cap\left([R,\infty)\times S^{1}\right) and equal to hn(kπσ2)cσh_{n}^{\prime}(k\pi\sigma_{-}^{2})-c_{\sigma_{-}} for all radii s<Rs<R. Then

EJ(v)\displaystyle E_{J}(v) Σ𝔥dt=Σ~d(𝔥dt)\displaystyle\leq\int_{\partial\Sigma}\mathfrak{h}dt=\int_{\tilde{\Sigma}}d(\mathfrak{h}dt)
=Σ~((0,R]×S1)d(𝔥dt)+Σ~([R,)×S1)d(𝔥dt)\displaystyle=\int_{\tilde{\Sigma}\cap\left((0,R]\times S^{1}\right)}d(\mathfrak{h}dt)+\int_{\tilde{\Sigma}\cap\left([R,\infty)\times S^{1}\right)}d(\mathfrak{h}dt)
=Σ~((0,R]×S1)d(𝔥dt)\displaystyle=\int_{\tilde{\Sigma}\cap\left((0,R]\times S^{1}\right)}d(\mathfrak{h}dt)
=(hn(kπσ2)hn(kπσ2))+(cσcσ).\displaystyle=(h_{n}^{\prime}(k\pi\sigma^{2})-h_{n}^{\prime}(k\pi\sigma_{-}^{2}))+(c_{\sigma_{-}}-c_{\sigma}).

As hnh_{n} is convex by assumption and σ<σ\sigma<\sigma_{-}, both (hn(kπσ2)hn(kπσ2))<0\left(h_{n}^{\prime}(k\pi\sigma^{2})-h_{n}^{\prime}(k\pi\sigma_{-}^{2})\right)<0 and (cσcσ)<0\left(c_{\sigma_{-}}-c_{\sigma}\right)<0. It follows that

(22) EJ(v)\displaystyle E_{J}(v) (hn(kπσ2)hn(kπσ2))+(cσcσ)<0.\displaystyle\leq(h_{n}^{\prime}(k\pi\sigma^{2})-h_{n}^{\prime}(k\pi\sigma_{-}^{2}))+(c_{\sigma_{-}}-c_{\sigma})<0.

which yields the desired contradiction.

Lemma 7 says that the winding number is decreased by the Floer differential. Lemma 8, below, says that the winding number is decreased by a continuation map. Thus, the winding numbers provide an auxilliary filtration on SC(H)SC^{*}(H).

Lemma 8

Let uu be a solution of Floer’s equation with respect to a Hamiltonian (Hn)s(H_{n})_{s} such that s(Hn)s0\partial_{s}(H_{n})_{s}\leq 0 everywhere. Suppose lims±u(s,t)=x±(t)\lim\limits_{s\rightarrow\pm\infty}u(s,t)=x_{\pm}(t). Then 𝔴(x+)𝔴(x)\mathfrak{w}(x_{+})\geq\mathfrak{w}(x_{-}).

Proof.

The proof of Lemma 7 applies almost verbatim, except the energy EJ(v)E_{J}(v) will have an additional integral

Σs(Hn)sdsdt\int_{\Sigma}\partial_{s}(H_{n})_{s}ds\wedge dt

which is non-positive, by assumption. This does not affect the final inequality (22), from which a contradiction was derived.

Finally, the additional structure imposed on the continuation maps yields the following.

Lemma 9

Continuation maps act as the canonical inclusions, sending a periodic orbit of XHnX_{H_{n}} to the periodic orbit of XHn+1X_{H_{n+1}} represented by the same map S1ES^{1}\rightarrow E.

Proof.

Let uu be a Floer solution of the data ((Hn)s,J)((H_{n})_{s},J), where (Hn)s(H_{n})_{s} induces an action-increasing continuation map cn:CF(Hn)CF(Hn+1)c_{n}:CF^{*}(H_{n})\rightarrow CF^{*}(H_{n+1}). By assumption, hn=hn+1h_{n}=h_{n+1} on [0,kπRn2][0,k\pi R_{n}^{2}]. Thus, (Hn)s=Hn=Hn+1(H_{n})_{s}=H_{n}=H_{n+1} within the disk bundle of radius RnR_{n}. By index considerations, any Floer solution is either constant or leaves the disk bundle of radius RnR_{n}. The latter cannot happen, by Lemma 8 and its easier analogue: uu remains inside the smallest disk bundle containing both of its asymptotes. Counting the constant Floer solutions precisely describes the canonical inclusion.

4 Proof of Theorem 1

We prove Theorem 1 in two steps, using the relationships between SH¯(H)\overline{SH^{*}}(H) and the closely-related theories SH(H)SH^{*}(H) and SH^(H)\widehat{SH^{*}}(H). Namely, there is an inclusion

ϕ:SH¯(H)SH^(H).\phi:\overline{SH^{*}}(H)\hookrightarrow\widehat{SH^{*}}(H).

induced by the inclusion

ker()^SC^(H)\widehat{\ker(\partial)}\hookrightarrow\widehat{SC^{*}}(H)

such that the following diagram commutes.

(23) SH¯(H){\overline{SH^{*}}(H)}SH(H){SH^{*}(H)}SH^(H){\widehat{SH^{*}}(H)}ϕ\scriptstyle{\phi}η\scriptstyle{\eta}π\scriptstyle{\pi}

Here, η\eta and π\pi come from the inclusion of ker()\ker(\partial) into its completion.

We will prove

Proposition 1

The image of ϕ\phi is equal to the image of π\pi.

As ϕ\phi is injective, this implies that SH¯(H)\overline{SH^{*}}(H) is isomorphic to im(π)\textnormal{im}(\pi), in particular,

SH¯(H)SH(H)/ker(π).\overline{SH^{*}}(H)\simeq{\raisebox{1.99997pt}{$SH^{*}(H)$}\left/\raisebox{-1.99997pt}{$\ker(\pi)$}\right.}.

Recall from Corollary 3 that the P.S.S. map

ι:HF(H0)SH(E)\iota:HF^{*}(H_{0})\longrightarrow SH^{*}(E)

induced by the inclusion

CF(H0)SC(E)CF^{*}(H_{0})\hookrightarrow SC^{*}(E)

is surjective. The composition

(24) Φ:=πι:HF(H0)SH^(H)\Phi:=\pi\circ\iota:HF^{*}(H_{0})\longrightarrow\widehat{SH^{*}}(H)

is therefore surjective and its image coincides with im(ϕ)\textnormal{im}(\phi). Thus,

SH¯(H)HF(H0)/ker(Φ).\overline{SH^{*}}(H)\simeq{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\ker(\Phi)$}\right.}.

The remainder of this section is devoted to computing the kernel of Φ\Phi and proving Proposition 1. Recall from Theorem 4 the isomorphism

SH^(H)=limaSHa(H).\widehat{SH^{*}}(H)=\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}SH^{*}_{a}(H).

Under this isomorphism, the map ϕ\phi is the inverse limit of maps

ϕa:HF(H0)SHa(H)\phi_{a}:HF^{*}(H_{0})\longrightarrow SH^{*}_{a}(H)

induced by the chain-level quotient maps

SC(H)SCa(H)=SC(H)/SC(a,)(H).SC^{*}(H)\longrightarrow SC^{*}_{a}(H)={\raisebox{1.99997pt}{$SC^{*}(H)$}\left/\raisebox{-1.99997pt}{$SC^{*}_{(a,\infty)}(H)$}\right.}.

We will compute the kernel of each ϕa\phi_{a} and show that

ker(ϕ)=limaker(ϕa).\ker(\phi)=\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker(\phi_{a}).

4.1 Step I

Consider an auxiliary family of Hamiltonians, defined as follows. Recall the radii {Rn}\{R_{n}\} that are part of the data of the Hamiltonians {Hn}\{H_{n}\}. Define

Kn=GnnkπRn2.K_{n}=G_{n}-nk\pi R_{n}^{2}.

The subcomplex

n=0CF(Kn)SC(K)\bigoplus_{n=0}^{\infty}CF^{*}(K_{n})\subset SC^{*}(K)

has a valuation derived from the usual action on each Floer chain complex. We extend this non-trivially to all of SC(K)SC^{*}(K) by defining

𝒜(𝐪)=kπR2.{\cal A}({\bf q})=-k\pi R^{2}.

Geometrically, this corresponds to the shift

CF(Kn)𝐪CF(KnkπR2)CF^{*}(K_{n}){\bf q}\cong CF^{*}(K_{n}-k\pi R^{2})

and ensures that action is increased by choices of continuation maps

cnK:CF(Kn)𝐪CF(Kn+1).c_{n}^{K}:CF^{*}(K_{n}){\bf q}\longrightarrow CF^{*}(K_{n+1}).

See Figure 4.

  • Remark 13)

    The “space” between KiK_{i} and Ki1𝐪K_{i-1}{\bf q} at r=0r=0 is important for defining action-increasing continuation maps. Without the space, there is no generic monotone-increasing homotopy. See [seidel-biased].

\mathbb{R}kπr2k\pi r^{2}K0K_{0}K0𝐪K_{0}{\bf q}K1K_{1}K1𝐪K_{1}{\bf q}
Figure 4: The family of Hamiltonians KnK^{n}

Because action is increased by continuation, SC(K)SC^{*}(K) has subcomplexes SC(a,)(K)SC^{*}_{(a,\infty)}(K) and quotient complexes SCa(K)SC^{*}_{a}(K) analogous to the subcomplex (16) and quotient complex (17) of SC(H)SC^{*}(H).

Choices of action-increasing continuation maps

CF(Kn)CF(Hn)CF^{*}(K_{n})\longrightarrow CF^{*}(H_{n})

induce a continuation map

𝔠KH:SH(K)SH(H)\mathfrak{c}^{KH}:SH^{*}(K)\longrightarrow SH^{*}(H)

that descends to a map

𝔠aKH:SHa(K)SHa(H).\mathfrak{c}_{a}^{KH}:SH^{*}_{a}(K)\longrightarrow SH^{*}_{a}(H).

This induces maps on the long-exact sequences

{\dots}SH(a,)(H){SH^{*}_{(a,\infty)}(H)}SH(H){SH^{*}(H)}SHa(H){SH^{*}_{a}(H)}{\dots}{\dots}SH(a,)(K){SH^{*}_{(a,\infty)}(K)}SH(K){SH^{*}(K)}SHa(K){SH^{*}_{a}(K)}{\dots}πa\scriptstyle{\pi_{a}}πa\scriptstyle{\pi_{a}}𝔠KH\scriptstyle{\mathfrak{c}^{KH}}𝔠KHa\scriptstyle{\mathfrak{c}^{KH}_{a}}

such that each square commutes. The commutativity of the right-hand square gives

(25) πa𝔠KH=𝔠aKHπa.\pi_{a}\circ\mathfrak{c}^{KH}=\mathfrak{c}_{a}^{KH}\circ\pi_{a}.

By construction,

Φ=𝔠KHΦK,\Phi=\mathfrak{c}^{KH}\circ\Phi^{K},

where ΦK:HF(K0)SH(K)\Phi^{K}:HF^{*}(K_{0})\longrightarrow{SH^{*}}(K) is the PSS map, defined similarly to (24). Φk\Phi^{k} associated action-filtered map

ΦaK:HF(K0)SHa(K).\Phi_{a}^{K}:HF^{*}(K_{0})\longrightarrow SH^{*}_{a}(K).

such that

(26) Φa=𝔠aKHΦaK.\Phi_{a}=\mathfrak{c}_{a}^{KH}\circ\Phi_{a}^{K}.

We now assume that kk is algebraically closed and has characteristic zero. Recall the following fact.

Lemma 10 (Lemma A.1 in [fooo])

Λ\Lambda is algebraically closed if kk is algebraically closed and of characteristic zero.

Denote by {λ1,..,λm}\{\lambda_{1},..,\lambda_{m}\} the eigenvalues of the map

𝒄𝟏𝑮𝓢:HF(H0)HF(H0)\bm{c_{-1}^{G}}\circ\bm{{\cal S}}:HF^{*}(H_{0})\longrightarrow HF^{*}(H_{0})

(including geometric multiplicities) and fix a Jordan basis

={𝐯𝟏λ𝟏,,𝐯𝐤λ𝟏λ𝟏,𝐯𝟏λ𝟐,,𝐯𝐤λ𝐦λ𝐦},{{\cal B}}=\left\{{\bf v_{1}^{\lambda_{1}}},...,{\bf v_{k_{\lambda_{1}}}^{\lambda_{1}}},{\bf v_{1}^{\lambda_{2}}},...,{\bf v_{k_{\lambda_{m}}}^{\lambda_{m}}}\right\},

so that {𝐯𝟏λ𝐢,,𝐯𝐤λ𝐢λ𝐢}\{{\bf v_{1}^{\lambda_{i}}},...,{\bf v_{k_{\lambda_{i}}}^{\lambda_{i}}}\} spans the invariant subspace associated with the eigenvalue λi\lambda_{i}.

Define a valuation

𝓐:HF(Kn)\displaystyle\bm{{\cal A}}:HF^{*}(K_{n}) \displaystyle\longrightarrow\mathbb{R}
𝐗\displaystyle{\bf X} minXCF(K)[X]=𝐗𝒜(X).\displaystyle\mapsto\min_{\begin{subarray}{c}X\in CF^{*}(K)\\ [X]={\bf X}\end{subarray}}{\cal A}(X).
Lemma 11

The valuation 𝓐\bm{{\cal A}} satisfies

  1. 1.

    𝓐(𝐗+𝐘)min(𝓐(𝐗),𝓐(𝐘))\bm{{\cal A}}({\bf X+Y})\geq\min(\bm{{\cal A}}({\bf X}),\bm{{\cal A}}({\bf Y})) and

  2. 2.

    There exists a constant CnC_{n}, such that, for any XCF(Kn)X\in CF^{*}(K_{n}) with [X]=𝐗[X]={\bf X},

    𝒜(X)𝓐(𝐗)Cn.{\cal A}(X)-\bm{{\cal A}}({\bf X})\leq C_{n}.
Proof.

Let ZCF(Kn)Z\in CF^{*}(K_{n}) be a representative of 𝐗+𝐘{\bf X}+{\bf Y} such that

𝒜(Z)=𝓐(𝐗+𝐘).{\cal A}(Z)=\bm{{\cal A}}({\bf X}+{\bf Y}).

Let XX be a representative of 𝐗{\bf X} and YY a representative of 𝐘{\bf Y} such that

Z=X+Y.Z=X+Y.

As 𝒜{\cal A} is a non-Archimedean valuation,

𝒜(Z)min(𝒜(X),𝒜(Y)).{\cal A}(Z)\geq\min({\cal A}(X),{\cal A}(Y)).

Thus,

𝓐(𝐗+𝐘)=𝒜(Z)min(𝒜(X),𝒜(Y))min(𝓐(𝐗),𝓐(𝐘)).\bm{{\cal A}}({\bf X}+{\bf Y})={\cal A}(Z)\geq\min({\cal A}(X),{\cal A}(Y))\geq\min(\bm{{\cal A}}({\bf X}),\bm{{\cal A}}({\bf Y})).

This shows item (1). Item (2) follows immediately from the fact that CF(Kn)CF^{*}(K_{n}) is finite-dimensional over Λ\Lambda.

Using Property (1) of Lemma 11, we can prove the following Lemma.

Lemma 12

Let λ\lambda be an eigenvalue of 𝐜𝟏𝐆𝓢\bm{c_{-1}^{G}}\circ\bm{{\cal S}} with ev(λ)>kπR2ev(\lambda)>k\pi R^{2}. For any constant κΛ\kappa\in\Lambda and any basis vector 𝐯𝐢λ{\bf v_{i}^{\lambda}} in {\cal B},

Φa(κ𝐯𝐢λ)=0.\Phi_{a}(\kappa{\bf v_{i}}^{\lambda})=0.
Proof.

From Equation (26), it suffices to show that

𝔠aKHΦaK(κ𝐯𝐢λ)=0.\mathfrak{c}_{a}^{KH}\circ\Phi_{a}^{K}(\kappa{\bf v_{i}^{\lambda}})=0.

Indeed, we will show that

(27) ΦaK(κ𝐯𝐢λ):=πaKϕK(κ𝐯𝐢λ)=0.\Phi_{a}^{K}(\kappa{\bf v_{i}^{\lambda}}):=\pi_{a}^{K}\circ\phi^{K}(\kappa{\bf v_{i}^{\lambda}})=0.

If ev(λ)=ev(\lambda)=\infty, that is, if λ=0\lambda=0, this follows immediately from Corollary 3. So suppose ev(λ)<ev(\lambda)<\infty. We will construct a cocycle XSC(K)X\in SC^{*}(K) such that

ϕK(κ𝐯𝐢λ)=𝐗\phi^{K}(\kappa{\bf v_{i}^{\lambda}})={\bf X}

and

πaK(𝐗)=0.\pi_{a}^{K}({\bf X})=0.

Let X0CF(H0)=CF(K0)X_{0}\in CF^{*}(H_{0})=CF^{*}(K_{0}) be a cocycle representing κ𝐯𝐢λ\kappa{\bf v_{i}^{\lambda}}. Define an infinite sequence {Xn𝐪CF(Kn)𝐪}n\left\{X_{n}{\bf q}\in CF^{*}(K_{n}){\bf q}\right\}_{n\in\mathbb{N}} by defining the base case

𝐪(X0𝐪)=X0\partial_{\bf q}(X_{0}{\bf q})=X_{0}

and inductively defining

Xn𝐪=cn1K(Xn1)𝐪.X_{n}{\bf q}=c_{n-1}^{K}(X_{n-1}){\bf q}.

Because the Floer differential commutes with continuation maps,

fl(Xn)=cn1Kc0fl(X0)=0.\partial^{fl}(X_{n})=c_{n-1}^{K}\circ...\circ c_{0}\circ\partial^{fl}(X_{0})=0.

Thus,

(n=0NXn𝐪)\displaystyle\partial\left(\sum_{n=0}^{N}X_{n}{\bf q}\right) =X0cNK(XN)+n=1NXncn1K(Xn1)\displaystyle=X_{0}-c_{N}^{K}(X_{N})+\sum_{n=1}^{N}X_{n}-c_{n-1}^{K}(X_{n-1})
=X0XN+1+n=1NXnXn\displaystyle=X_{0}-X_{N+1}+\sum_{n=1}^{N}X_{n}-X_{n}
=X0XN+1,\displaystyle=X_{0}-X_{N+1},

and so

[X0][Xn][X_{0}]\simeq[X_{n}]

for any nn. We will show that, for nn large, 𝒜(Xn)>a{\cal A}(X_{n})>a.

Denote by HFλ(H0)HF^{*}_{\lambda}(H_{0}) the invariant subspace corresponding to λ\lambda. In the fixed Jordan basis, the matrix of 𝒄𝟏𝑮𝓢|HFλ(H0)\bm{c_{-1}^{G}\circ{\cal S}}\big{|}_{HF^{*}_{\lambda}(H_{0})} is

A=[λ1λ100λ1λ].A=\left[\begin{array}[]{cccccc}\lambda&1&&&&\\ &\lambda&1&&0&\\ &&\ddots&\ddots&&\\ &0&&\ddots&\ddots&\\ &&&&\lambda&1\\ &&&&&\lambda\end{array}\right].

Ignoring non-zero 𝕂\mathbb{K}-scalar factors (these do not affect action),

(28) Akλ=[λkλλkλ1λλkλλkλ1λ20λkλλkλ1λkλ]A^{k_{\lambda}}=\left[\begin{array}[]{cccccc}\lambda^{k_{\lambda}}&\lambda^{k_{\lambda}-1}&&\dots&&\lambda\\ &\lambda^{k_{\lambda}}&\lambda^{k_{\lambda}-1}&&\dots&\lambda^{2}\\ &&\ddots&\ddots&&\vdots\\ &0&&\ddots&\ddots&\\ &&&&\lambda^{k_{\lambda}}&\lambda^{k_{\lambda}-1}\\ &&&&&\lambda^{k_{\lambda}}\end{array}\right]

and

AN=λNkλAkλA^{N}=\lambda^{N-k_{\lambda}}A^{k_{\lambda}}

for all NkλN\geq k_{\lambda}. View X0CF(G0)X_{0}\in CF^{*}(G_{0}) under the identification K0=G0K_{0}=G_{0}. Again ignoring non-zero 𝕂\mathbb{K}-scalar factors,

ANκ𝒗𝒊𝝀=κλNkλj=1iλkλi+j𝒗𝒋𝝀A^{N}\kappa\bm{v_{i}^{\lambda}}=\kappa\lambda^{N-k_{\lambda}}\sum_{j=1}^{i}\lambda^{k_{\lambda}-i+j}\bm{v_{j}^{\lambda}}

Using Property (1) of Lemma 11, we compute

(29) 𝓐((𝒄𝟏𝑮𝓢)N(κ𝒗𝒊𝝀))\displaystyle\bm{{\cal A}}((\bm{c_{-1}^{G}\circ{\cal S}})^{N}(\kappa\bm{v_{i}^{\lambda}})) =𝓐(κλNkλj=1iλkλi+j𝒗𝒋𝝀)\displaystyle=\bm{{\cal A}}\left(\kappa\lambda^{N-k_{\lambda}}\sum_{j=1}^{i}\lambda^{k_{\lambda}-i+j}\bm{v_{j}^{\lambda}}\right)
(30) ev(κλNkλ)+minj𝓐(λkλi+j𝒗𝒋𝝀)\displaystyle\geq ev(\kappa\lambda^{N-k_{\lambda}})+\min_{j}\bm{{\cal A}}(\lambda^{k_{\lambda}-i+j}\bm{v_{j}^{\lambda}})
(31) (Nkλ)ev(λ)+ev(κ)+(kλi)ev(λ)+minj𝒜(𝒗𝒋𝝀)\displaystyle\geq(N-k_{\lambda})\cdot ev(\lambda)+ev(\kappa)+(k_{\lambda}-i)\cdot ev(\lambda)+\min_{j}{\cal A}(\bm{v_{j}^{\lambda}})
(32) Nev(λ)+ev(κ)+𝓐(𝒗𝒊𝝀)𝓐(𝒗𝒊𝝀)+minj𝓐(𝒗𝒋𝝀)iev(λ)\displaystyle\geq N\cdot ev(\lambda)+ev(\kappa)+\bm{{\cal A}}(\bm{v_{i}^{\lambda}})-\bm{{\cal A}}(\bm{v_{i}^{\lambda}})+\min_{j}\bm{{\cal A}}(\bm{v_{j}^{\lambda}})-i\cdot ev(\lambda)
(33) =Nev(λ)+𝓐(κ𝒗𝒊𝝀)𝓐(𝒗𝒊𝝀)+minj𝓐(𝒗𝒋𝝀)iev(λ)\displaystyle=N\cdot ev(\lambda)+\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})-\bm{{\cal A}}(\bm{v_{i}^{\lambda}})+\min_{j}\bm{{\cal A}}(\bm{v_{j}^{\lambda}})-i\cdot ev(\lambda)
(34) =Nev(λ)+𝓐(κ𝒗𝒊𝝀)+C,\displaystyle=N\cdot ev(\lambda)+\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})+C,

where

C=𝓐(𝒗𝒊𝝀)+minj𝓐(𝒗𝒋𝝀)iev(λ).C=-\bm{{\cal A}}(\bm{v_{i}^{\lambda}})+\min_{j}\bm{{\cal A}}(\bm{v_{j}^{\lambda}})-i\cdot ev(\lambda).

Let XN=𝒮N(c1G𝒮)N(X0)CF(GN).X_{N}^{\prime}={\cal S}^{-N}\circ(c_{-1}^{G}\circ{\cal S})^{N}(X_{0})\in CF^{*}(G_{N}). By Theorem 6,

XN=cN1GcN2Gc0G(X0).{X_{N}^{\prime}}=c_{N-1}^{G}\circ c_{N-2}^{G}\circ...\circ c_{0}^{G}(X_{0}).

The map 𝒮{\cal S} preserves action, and so

𝒜(XN)=𝒜((c1G𝒮)N(X0)).{\cal A}(X_{N}^{\prime})={\cal A}((c_{-1}^{G}\circ{\cal S})^{N}(X_{0})).

Indeed, the representatives of [(c1G𝒮)N(X0)][(c_{-1}^{G}\circ{\cal S})^{N}(X_{0})] and [XN][X_{N}^{\prime}] are in bijective correspondence through the action-preserving map 𝒮N{\cal S}^{-N}. Thus,

(35) 𝓐([XN])=𝓐([(c1G𝒮)N(X0)]).\bm{{\cal A}}([X_{N}^{\prime}])=\bm{{\cal A}}([(c_{-1}^{G}\circ{\cal S})^{N}(X_{0})]).

Recall that [X0]=κ𝒗𝒊𝝀𝒋[X_{0}]=\kappa\bm{v_{i}^{\lambda_{j}}}. As (c1G𝒮)N(c_{-1}^{G}\circ{\cal S})^{N} is a chain map,

[(c1G𝒮)N(X0)]=(𝒄𝟏𝑮𝓢)N(κ𝒗𝒊𝝀𝒋).[(c_{-1}^{G}\circ{\cal S})^{N}(X_{0})]=\bm{(c_{-1}^{G}\circ{\cal S})}^{N}(\kappa\bm{v_{i}^{\lambda_{j}}}).

Combining (34) and (35),

𝓐([XN])𝓐(κ𝒗𝒊𝝀)+Nev(λ)+C.\bm{{\cal A}}([X_{N}^{\prime}])\geq\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})+N\cdot ev(\lambda)+C.

The canonical identification

CF(GN)CF(KN)CF^{*}(G_{N})\xrightarrow{\simeq}CF^{*}(K_{N})

descends to a map on cohomology that sends [XN][X_{N}^{\prime}] to [XN][X_{N}]. This isomorphism decreases action on-the-nose by NkπRN2-Nk\pi R_{N}^{2}, and so the action of [XN][X_{N}] is bounded by

𝓐([XN])𝓐([XN])NkπRN2,\bm{{\cal A}}([X_{N}])\geq\bm{{\cal A}}([X_{N}^{\prime}])-Nk\pi R_{N}^{2},

and so

𝓐([XN])Nev(λ)NkπRN2+𝓐(κ𝒗𝒊𝝀)+C.\bm{{\cal A}}([X_{N}])\geq N\cdot ev(\lambda)-Nk\pi R_{N}^{2}+\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})+C.

The difference ev(λ)kπRN2>0ev(\lambda)-k\pi R_{N}^{2}>0 is bounded below by the positive number ev(λ)kπR2>0ev(\lambda)-k\pi R^{2}>0, and 𝓐(κ𝒗𝒊𝝀)+C\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})+C is a constant. There therefore exists nn satisfying

n(ev(λ)kπRN2)+𝓐(κ𝒗𝒊𝝀)+C>a.n(ev(\lambda)-k\pi R_{N}^{2})+\bm{{\cal A}}(\kappa\bm{v_{i}^{\lambda}})+C>a.

Thus,

𝒜(Xn)𝓐(𝐗𝐧)>a.{\cal A}(X_{n})\geq\bm{{\cal A}}({\bf X_{n}})>a.

Lemma 13

Denote by HFλ(H0)HF^{*}_{\lambda}(H_{0}) the invariant subspace of HF(H0)HF^{*}(H_{0}) corresponding to an eigenvalue λ\lambda of 𝐜𝟏𝐆𝓢{\bf c_{-1}^{G}}\circ\bm{{\cal S}}.

ev(λ)>kπR2HFλ(H0)ker(Φa).\bigoplus_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\subset\ker(\Phi_{a}).
Proof.

The map Φa\Phi_{a} is a Λ0\Lambda_{0}-module homomorphism. Any element in ev(λ)>kπR2HFλ(H0)\ \bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0}) can be written as the sum of generalized eigenvectors κ𝐯𝐢λ\kappa{\bf v_{i}}^{\lambda} with ev(λ)>kπR2ev(\lambda)>k\pi R^{2} and κΛ\kappa\in\Lambda. Lemma 12 shows that

Φa(κ𝐯𝐢λ)=0.\Phi_{a}(\kappa{\bf v_{i}}^{\lambda})=0.

By linearity, Φa(X)=0\Phi_{a}(X)=0 as well.

4.2 Step II

We want to show that, as aa\rightarrow\infty,

ev(λ)kπR2HFλ(H0)ker(Φa)=.\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\cap\ker(\Phi_{a})=\mathop{\varnothing}.

Denote by HFλ,(a,)(H0)HF^{*}_{\lambda,(a,\infty)}(H_{0}) the intersection of HFλ(H0)HF^{*}_{\lambda}(H_{0}) with the image of the map

HF(a,)(H0)HF(H0).HF^{*}_{(a,\infty)}(H_{0})\longrightarrow HF^{*}(H_{0}).
Proposition 2

The kernel of Φa:HF(H0)SHa(H)\Phi_{a}:HF^{*}(H_{0})\longrightarrow SH^{*}_{a}(H) satisfies the following inclusions:

ev(λ)>kπR2HFλ(H0)ev(λ)kπR2HFλ,(a,)(H0)ker(Φa)\bigoplus_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\oplus\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda,(a,\infty)}(H_{0})\subset\ker(\Phi_{a})

and

ker(Φa)ev(λ)>kπR2HFλ(H0)ev(λ)kπR2HFλ,(aC,)(H0),\ker(\Phi_{a})\subset\bigoplus_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\oplus\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda,(a-C,\infty)}(H_{0}),

where C>0C>0 is a constant depending only upon the fixed Jordan basis {\cal B}.

The proof of Proposition 2 relies on four Lemmas that bound action.

4.2.1 Lemmas bounding action

Lemma 14

There exists a constant C>0C>0 such that, for any cocycles X,YCF(H0)X,Y\in CF^{*}(H_{0}) satisfying

[X]=[Y],[X]=[Y],

the difference in actions is no more than CC, that is,

|𝒜(X)𝒜(Y)|C.|{\cal A}(X)-{\cal A}(Y)|\leq C.
Proof.

This follows immediately from property (2) of Lemma 11 and the fact that HF(H0)HF^{*}(H_{0}) is finitely generated.

Lemma 15

The valuations of the Novikov-valued coefficients are not decreased by 𝔠HG\mathfrak{c}^{HG}.

Proof.

Let GnsG_{n}^{s} be a monotone homotopy between HnH_{n} and GnG_{n} of the form

Gns=𝔥ns(kπr2)+(1+kπr2)ρf.G_{n}^{s}=\mathfrak{h}_{n}^{s}(k\pi r^{2})+(1+k\pi r^{2})\rho^{*}f.

Thus, the homotopy only changes in the radial direction.

Let JsJ_{s} be a family of admissible almost-complex structures. A solution u(s,t)u(s,t) of the Floer equation

us+Js(utXHns)=0\frac{\partial u}{\partial s}+J_{s}\left(\frac{\partial u}{\partial t}-X_{H_{n}^{s}}\right)=0

descends to a solution v=ρu(s,t)v=\rho\circ u(s,t) of the Floer equation

(36) vs+ρJs(vtXf)=0.\frac{\partial v}{\partial s}+\rho_{*}J_{s}\left(\frac{\partial v}{\partial t}-X_{f}\right)=0.

Let v~\tilde{v} be the compactification of vv by its limit points, and let u~\tilde{u} be the compactification of uu by the limiting fiber disks, so that ρ(u~)=v~\rho_{*}(\tilde{u})=\tilde{v}. Since

Ω([u~])=ρω([u~])=ω([v~]),\Omega([\tilde{u}])=\rho^{*}\omega([\tilde{u}])=\omega([\tilde{v}]),

uu contributes a weight

TΩ[u~]=Tω[v~]T^{-\Omega[\tilde{u}]}=T^{-\omega[\tilde{v}]}

to cHGnc^{HG}_{n}. Action is increased by solutions of (36), and so

0ω([v~])+limsf(v(s,t))limsf(v(s,t)).0\leq-\omega([\tilde{v}])+\lim_{s\rightarrow-\infty}f(v(s,t))-\lim_{s\rightarrow\infty}f(v(s,t)).

In particular,

(37) maxx,yM|f(x)f(y)|ω([v~]).-\max_{x,y\in M}|f(x)-f(y)|\leq-\omega([\tilde{v}]).

By assumption, ff is 𝒞2{\cal C}^{2} small; assume that f>0f>0 and ff is small enough that

(38) maxxMf(x)>minAπ2(M)ω(A)0|ω(A)|.-\max_{x\in M}f(x)>-\min_{\begin{subarray}{c}A\in\pi_{2}(M)\\ \omega(A)\neq 0\end{subarray}}|\omega(A)|.

(37) and (38) imply

minAπ2(M)ω(A)0|ω(A)|<ω([v~]),-\min_{\begin{subarray}{c}A\in\pi_{2}(M)\\ \omega(A)\neq 0\end{subarray}}|\omega(A)|<-\omega([\tilde{v}]),

or

0ω([v~]).0\leq-\omega([\tilde{v}]).

Applying this argument to every nn, the Lemma follows.

Let μ1,μ2,μ\mu_{1},\mu_{2},\dots\mu_{\ell} be eigenvalues of the operator

c1G𝒮:CF(G0)CF(G0),c_{-1}^{G}\circ{\cal S}:CF^{*}(G_{0})\longrightarrow CF^{*}(G_{0}),

indexed with multiplicities. Let B={u1μ1,u2μ1,,ujμμ}B=\{u_{1}^{\mu_{1}},u_{2}^{\mu_{1}},...,u_{j_{\mu_{\ell}}}^{\mu_{\ell}}\} be a Jordan eigenbasis for c1G𝒮c_{-1}^{G}\circ{\cal S}. Write the μ\muth-generalized eigenspace as CFμ(G0)CF_{\mu}^{*}(G_{0}).

Let XCF(G0)X\in CF^{*}(G_{0}), and write XX in the basis BB.

X=i,jcijuiμjX=\sum_{i,j}c_{ij}u_{i}^{\mu_{j}}

(where cijΛc_{ij}\in\Lambda and the range of ii depends on jj). Now write

X=xn𝒫(H0)dnxnX=\sum_{x_{n}\in{\cal P}(H_{0})}d_{n}x_{n}

(where dnΛd_{n}\in\Lambda.)

Lemma 16

There exists a constant CC\in\mathbb{R}, depending only on the choice of bases, such that

𝒜(X)mini,j𝒜(cijuiμj)+C.{\cal A}(X)\leq\min_{i,j}{\cal A}(c_{ij}u_{i}^{\mu_{j}})+C.
Proof.

Let eijne_{ijn} be constants determining the change-of-basis {xn}{\cal B}\rightarrow\{x_{n}\}, so that

xn=i,jeijnuiμj.x_{n}=\sum_{i,j}e_{ijn}u_{i}^{\mu_{j}}.

Take

C=minijn𝒜(eijn)+minij𝒜(uiμj).C=\min_{ijn}{\cal A}(e_{ijn})+\min_{ij}{\cal A}(u_{i}^{\mu_{j}}).

Then

i,jcijuiμj\displaystyle\sum_{i,j}c_{ij}u_{i}^{\mu_{j}} =X=ndnxn=ndni,jeijnuiμj=i,j(ndneijn)uiμj,\displaystyle=X=\sum_{n}d_{n}x_{n}=\sum_{n}d_{n}\sum_{i,j}e_{ijn}u_{i}^{\mu_{j}}=\sum_{i,j}\left(\sum_{n}d_{n}e_{ijn}\right)u_{i}^{\mu_{j}},

and so

cij=ndneijn.c_{ij}=\sum_{n}d_{n}e_{ijn}.

It follows that

𝒜(cijuiμj)\displaystyle{\cal A}(c_{ij}u_{i}^{\mu_{j}}) minn𝒜(dneijnuiμj)\displaystyle\geq\min_{n}{\cal A}(d_{n}e_{ijn}u_{i}^{\mu_{j}})
(minn𝒜(dn))+(minn𝒜(eijn))+(mini,j𝒜(uiμj))\displaystyle\geq\left(\min_{n}{\cal A}(d_{n})\right)+\left(\min_{n}{\cal A}(e_{ijn})\right)+\left(\min_{i,j}{\cal A}(u_{i}^{\mu_{j}})\right)
𝒜(X)+C.\displaystyle\geq{\cal A}(X)+C.

The inequality given by the last line follows from the definition of the action of XX and the observation that

minijn𝒜(eijn)minn𝒜(eijn)\min_{ijn}{\cal A}(e_{ijn})\leq\min_{n}{\cal A}(e_{ijn})

for any fixed i,ji,j. ∎

Lemma 17

If μ\mu is an eigenvalue of the map

c1G𝒮:CF(H0)CF(H0)c_{-1}^{G}\circ{\cal S}:CF^{*}(H_{0})\longrightarrow CF^{*}(H_{0})

then

ev(μ)0.ev(\mu)\geq 0.
Proof.

Let uμu^{\mu}\in{\cal B} be the eigenvector satisfying

c1G𝒮(uμ)=λuμ.c_{-1}^{G}\circ{\cal S}(u^{\mu})=\lambda u^{\mu}.

Let G0G_{0}^{\prime} be an upward shift of the Hamiltonian G0G_{0}:

G0=G0+δG_{0}^{\prime}=G_{0}+\delta

for some small δ>0\delta>0. Let c1Gc_{-1}^{G^{\prime}} be an action-increasing

c1G:CF(G1)CF(G0).c_{-1}^{G^{\prime}}:CF^{*}(G_{-1})\longrightarrow CF^{*}(G_{0}).

Continuation maps are homotopy invariant, so under the identification CF(G0)=F(G0)CF^{*}(G_{0})=F^{*}(G_{0}^{\prime}) induced by the equality 𝒫(G0)=𝒫(G0){\cal P}(G_{0})={\cal P}(G_{0}^{\prime}),

c1G𝒮(uμ)=μuμ.c_{-1}^{G^{\prime}}\circ{\cal S}(u^{\mu})=\mu u^{\mu}.

Modifying c1Gc_{-1}^{G} if necessary, we assume without loss of generality that

CF(G1){CF^{*}(G_{-1})}CF(G0){CF^{*}(G_{0}^{\prime})}CF(G0){CF^{*}(G_{0})}c1G\scriptstyle{c_{-1}^{G^{\prime}}}c1G\scriptstyle{c_{-1}^{G}}=\scriptstyle{=}

commutes. Clearly,

ev(μ)+𝒜G0(uμ)\displaystyle ev(\mu)+{\cal A}_{G_{0}}(u^{\mu}) =𝒜G0(μuμ)\displaystyle={\cal A}_{G_{0}}(\mu u^{\mu})
=𝒜G0(μuμ)δ\displaystyle={\cal A}_{G_{0}^{\prime}}(\mu u^{\mu})-\delta
𝒜G0(uμ)δ,\displaystyle\geq{\cal A}_{G_{0}}(u^{\mu})-\delta,

where the last inequality follows from the fact that action is increased by c1Gc_{-1}^{G^{\prime}}. Thus,

(39) ev(μ)δ.ev(\mu)\geq-\delta.

As (39) holds for any δ>0\delta>0, we conclude that

ev(μ)0.ev(\mu)\geq 0.

4.2.2 Proof of Proposition 2

Let

𝔠nHG:CF(Hn)CF(Gn)\mathfrak{c}_{n}^{HG}:CF^{*}(H_{n})\longrightarrow CF^{*}(G_{n})

be an action-increasing continuation map. The following Lemma simplifies the computation of Lemma 19.

Lemma 18

The continuation maps 𝔠nHG\mathfrak{c}_{n}^{HG} may be chosen to commute at the level of cochains with the continuation maps cnGc_{n}^{G} and cnc_{n}, that is,

𝔠nHGcn1=cnG𝔠n1HG.\mathfrak{c}_{n}^{HG}\circ c_{n-1}=c_{n}^{G}\circ\mathfrak{c}_{n-1}^{HG}.
Proof.

We sketch the proof. Let gns:g_{n}^{s}:\mathbb{R}\rightarrow\mathbb{R} be the map interpolating between the functions gng_{n} and gn+1g_{n+1} in the definition of GnG_{n} and Gn+1G_{n+1}. Define the continuation maps inductively as follows. Denote by 𝔥1s\mathfrak{h}_{1}^{s} a generic map interpolating between h1h_{1} and g1g_{1} so that

  1. 1.

    𝔥1s=g1s\mathfrak{h}_{1}^{s}=g_{1}^{s} on [0,R12][0,\frac{R_{1}}{2}] and

  2. 2.

    𝔥1s\mathfrak{h}_{1}^{s} is monotone-decreasing in ss elsewhere.

Define 𝔠1HG\mathfrak{c}_{1}^{HG} through 𝔥1s\mathfrak{h}_{1}^{s}.

Now let 𝔨n\mathfrak{k}_{n} be the function that is

  1. 1.

    equal to gng_{n} on [0,kπRn2][0,k\pi R_{n}^{2}] and

  2. 2.

    a translation of hnh_{n} on [kπRn2,)[k\pi R_{n}^{2},\infty).

See the teal curve in Figure 5. Let 𝔥ns\mathfrak{h}_{n}^{s} be a generic function interpolating between hnh_{n} and 𝔨n\mathfrak{k}_{n} that is

  1. 1.

    equal to 𝔥n1s\mathfrak{h}_{n-1}^{s} on [0,kπRn12][0,k\pi R_{n-1}^{2}] and

  2. 2.

    translation by a constant on (kπRn12,)(k\pi R_{n-1}^{2},\infty).

Define a continuation map 𝔠nHK\mathfrak{c}_{n}^{HK} through 𝔥n\mathfrak{h}_{n}.

Let 𝔨ns\mathfrak{k}_{n}^{s} be a generic function interpolating between 𝔨n\mathfrak{k}_{n} and gng_{n} that is

  1. 1.

    equal to gn1sg_{n-1}^{s} on [0,kπRn12][0,k\pi R_{n-1}^{2}] and

  2. 2.

    monotone-decreasing in ss on [kπRn12,)[k\pi R_{n-1}^{2},\infty).

Define a continuation map 𝔠nKG\mathfrak{c}_{n}^{KG} through 𝔨ns\mathfrak{k}_{n}^{s}. Set

𝔠nHG=𝔠nKG𝔠nHK.\mathfrak{c}_{n}^{HG}=\mathfrak{c}_{n}^{KG}\circ\mathfrak{c}_{n}^{HK}.

The Lemma follows from the integrated maximum principal of Lemma 7.

\mathbb{R}kπr2k\pi r^{2}h2h_{2}𝔨2\mathfrak{k}_{2}g2g_{2}
Figure 5: Defining 𝔠2HG\mathfrak{c}_{2}^{HG} through the intermediary function 𝔨2\mathfrak{k}_{2}

Denote by

πμ:CF(G0)CFμ(G0)\pi_{\mu}:CF^{*}(G_{0})\longrightarrow CF^{*}_{\mu}(G_{0})

the projection onto the μ\muth-generalized eigenspace. The following Lemma bounds the action of a cochain in an arbitrary Floer complex CF(Hn)CF^{*}(H_{n}) by the action of a cochain in CF(H0)CF^{*}(H_{0}).

Lemma 19

Let μ\mu be an eigenvalue with ev(μ)kπR2ev(\mu)\leq k\pi R^{2}. Let XCF(Hn)X\in CF^{*}(H_{n}) be a cocycle such that

Yn:=πμ𝒮n𝔠HG(X)0.Y_{n}:=\pi_{\mu}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X)\neq 0.

Let Y0Y_{0} be the unique element in CFμ(G0)CF^{*}_{\mu}(G_{0}) satisfying (c1G𝒮)n(Y0)=Yn(c_{-1}^{G}\circ{\cal S})^{n}(Y_{0})=Y_{n}. There exists a constant 𝒞{\cal C}, independent of μ\mu, satisfying

𝒜(X)𝒜(Y0)<𝒞.{\cal A}(X)-{\cal A}(Y_{0})<{\cal C}.

Before embarking on the proof of Lemma 19 in full generality, we illustrate the idea of the proof with the following simplified scenario. Suppose that Y0=y𝒫(H0)Y_{0}=y\in{\cal P}(H_{0}) is a bona-fide μ\mu-eigenvector, and X=x𝒫(Hn)X=x\in{\cal P}(H_{n}) is a cocycle of winding number 𝔴(x)=n\mathfrak{w}(x)=n satisfying

𝒮n𝔠nHK(x)=μnκy=(c1G𝒮)n(κy){\cal S}^{n}\circ\mathfrak{c}_{n}^{HK}(x)=\mu^{n}\kappa y=(c_{-1}^{G}\circ{\cal S})^{n}(\kappa y)

for some κΛ\kappa\in\Lambda. By Lemma 17, the valuation of the Novikov coefficient is increased by 𝔠nHG\mathfrak{c}_{n}^{HG} (in this case the Novikov coefficient is 11 and has valuation ev(1)=0ev(1)=0). Therefore,

0=ev(1)ev(μnκ)=nev(μ)+ev(κ).0=ev(1)\leq ev(\mu^{n}\kappa)=n\cdot ev(\mu)+ev(\kappa).

Assume that the circle bundle on which xx lives has radius close to RR. The action of xx, compared to the action of κy\kappa y, is

𝒜(x)𝒜(κy)\displaystyle{\cal A}(x)-{\cal A}(\kappa y) kπnR2ev(κ)\displaystyle\simeq-k\pi nR^{2}-ev(\kappa)
kπnR2+nev(μ)\displaystyle\leq-k\pi nR^{2}+n\cdot ev(\mu)
=n(ev(μ)kπR2)\displaystyle=n(ev(\mu)-k\pi R^{2})
<0\displaystyle<0

if ev(μ)<kπR2ev(\mu)<k\pi R^{2}. Taking 𝒞=0{\cal C}=0, this is precisely the statement of Lemma 19. Of course, we can never assume that we are working in such a simplified scenario. Namely, we must take XX to be a weighted sum of periodic orbits and Y0Y_{0} to be a weighted sum of generalized eigenvectors.

Proof.

Write

X=kκkxk,X=\sum_{k}\kappa_{k}x_{k},

where κkΛ\kappa_{k}\in\Lambda and xk𝒫(Hn)x_{k}\in{\cal P}(H_{n}). Because 𝔠nHG\mathfrak{c}_{n}^{HG} is a Λ\Lambda-linear morphism,

𝒮n𝔠HGn(X)=k=1m𝒮𝔠HGn(κkxk).{\cal S}^{n}\circ\mathfrak{c}^{HG}_{n}(X)=\sum_{k=1}^{m}{\cal S}\circ\mathfrak{c}^{HG}_{n}(\kappa_{k}x_{k}).

Write πμ𝒮n𝔠HGn(κkxk)\pi_{\mu}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}_{n}(\kappa_{k}x_{k}) in the basis BB:

(40) πμ𝒮n𝔠HGn(κkxk)=j=1jμκk,j,ujμ\pi_{\mu}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}_{n}(\kappa_{k}x_{k})=\sum_{j=1}^{j_{\mu}}\kappa_{k,j,}u_{j}^{\mu}

for some {κk,jΛ}\{\kappa_{k,j}\in\Lambda\}. By assumption,

(41) j=1jμ(kκk,j)ujμ=πμ𝒮n𝔠nHG(X)=(cG1𝒮)n(Y0).\sum_{j=1}^{j_{\mu}}\left(\sum_{k}\kappa_{k,j}\right)u_{j}^{\mu}=\pi_{\mu}\circ{\cal S}^{n}\circ\mathfrak{c}_{n}^{HG}(X)=(c_{G}^{-1}\circ{\cal S})^{n}(Y_{0}).

Write Y0Y_{0} in the fixed Jordan basis B:

Y0=j=1jμ𝔶juμj.Y_{0}=\sum_{j=1}^{j_{\mu}}\mathfrak{y}_{j}u^{\mu}_{j}.

We want to compare the action of some κkxk\kappa_{k}x_{k} with the action of some 𝔶jujμ\mathfrak{y}_{j}u_{j}^{\mu}. We do this through the auxilliary cochain YnY_{n}. By definition, and ignoring extra non-zero 𝕂\mathbb{K} factors,

Yn=j=1jμi=jmin(n+j,jμ)μni+j𝔶iujμY_{n}=\sum_{j=1}^{j_{\mu}}\sum_{i=j}^{\min(n+j,j_{\mu})}\mu^{n-i+j}\mathfrak{y}_{i}u_{j}^{\mu}

Thus, (41) produces equalities for each j{1,,jμ}j\in\{1,...,j_{\mu}\}:

(42) kκk,j=i=jmin(nj+j,jμ)μni+j𝔶i.\sum_{k}\kappa_{k,j}=\sum_{i=j}^{\min(n_{j}+j,j_{\mu})}\mu^{n-i+j}\mathfrak{y}_{i}.

Fix jj so that 𝔶j0\mathfrak{y}_{j}\neq 0. Lemma 17 states that ev(μ)>0ev(\mu)>0. As a consequence, the action of the right-hand-side of (42) is completely determined by one of the summands:

𝒜(i=jmin(n+j,jμ)μni+j𝔶i)=mini𝒜(μni+j𝔶i)nev(μ)+ev(𝔶j).{\cal A}(\sum_{i=j}^{\min(n+j,j_{\mu})}\mu^{n-i+j}\mathfrak{y}_{i})=\min_{i}{\cal A}(\mu^{n-i+j}\mathfrak{y}_{i})\leq n\cdot ev(\mu)+ev(\mathfrak{y}_{j}).

Thus, there exists some kk, which we denote by 𝔨\mathfrak{k}, satisfying

(43) ev(κ𝔨,j)nev(μ)+ev(𝔶j).ev(\kappa_{\mathfrak{k},j})\leq n\cdot ev(\mu)+ev(\mathfrak{y}_{j}).

Denote by ww the winding number 𝔴(x𝔨)\mathfrak{w}(x_{\mathfrak{k}}) of x𝔨x_{\mathfrak{k}}. View x𝔨x_{\mathfrak{k}} as an element of CF(Hw)CF^{*}(H_{w}) under the inclusion CF(Hw)CF(Hn)CF^{*}(H_{w})\hookrightarrow CF^{*}(H_{n}). By Lemma 18,

(44) 𝒮n𝔠HGn(κ𝔨x𝔨)\displaystyle{\cal S}^{n}\circ\mathfrak{c}^{HG}_{n}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}}) =𝒮n𝔠HGncn1HcwH(κ𝔨x𝔨)\displaystyle={\cal S}^{n}\circ\mathfrak{c}^{HG}_{n}\circ c_{n-1}^{H}\circ...\circ c_{w}^{H}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})
(45) =(c1G𝒮)nw𝒮w𝔠HGw(κ𝔨x𝔨).\displaystyle=(c_{-1}^{G}\circ{\cal S})^{n-w}\circ{\cal S}^{w}\circ\mathfrak{c}^{HG}_{w}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}}).

Again appealing to the Jordan normal form, write

(46) πμ𝒮w𝔠HGw(κ𝔨x𝔨)=jγj,𝔨ujμ.\pi_{\mu}\circ{\cal S}^{w}\circ\mathfrak{c}^{HG}_{w}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})=\sum_{j}\gamma_{j,\mathfrak{k}}u_{j}^{\mu}.

The orbits in 𝒫(G0){\cal P}(G_{0}) are constant and appear at energy level H=fρ0H=f\circ\rho\geq 0. Lemma 15 states that the valuation of the Novikov-coefficient is increased by 𝔠wHG\mathfrak{c}_{w}^{HG}. The action of 𝒜(𝔠wHG(κ𝔨x𝔨)){{\cal A}}(\mathfrak{c}_{w}^{HG}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})) is the valuation of the Novikov coefficient, plus the value of fρf\circ\rho. Together, this implies that

(47) ev(κ𝔨)𝒜(𝔠wHG(κ𝔨x𝔨)).ev(\kappa_{\mathfrak{k}})\leq{{\cal A}}(\mathfrak{c}_{w}^{HG}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})).

By Lemma 16, there is a constant CC such that

𝒜(𝔠HG(κ𝔨x𝔨))=𝒜(𝒮w𝔠HG(κ𝔨x𝔨))𝒜(γj,𝔨ujμ)+C{\cal A}(\mathfrak{c}^{HG}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}}))={\cal A}({\cal S}^{w}\circ\mathfrak{c}^{HG}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}}))\leq{\cal A}(\gamma_{j,\mathfrak{k}}u_{j}^{\mu})+C

for any jj. We conclude that

(48) ev(κ𝔨)𝒜(γj,𝔨ujμ)+Cev(\kappa_{\mathfrak{k}})\leq{\cal A}(\gamma_{j,\mathfrak{k}}u_{j}^{\mu})+C

for any jj.

Applying the map (c1G𝒮)nw(c_{-1}^{G}\circ{\cal S})^{n-w} to (46), ignoring non-zero 𝕂\mathbb{K} scalars as always, yields

πμ(c1G𝒮)nw𝒮w𝔠HGw(κ𝔨x𝔨)=j=1jμq=jmin(nw+j,jμ)μnwq+jγq,𝔨ujμ.\pi_{\mu}\circ(c_{-1}^{G}\circ{\cal S})^{n-w}\circ{\cal S}^{w}\circ\mathfrak{c}^{HG}_{w}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})=\sum_{j=1}^{j_{\mu}}\sum_{q=j}^{\min(n-w+j,j_{\mu})}\mu^{n-w-q+j}\gamma_{q,\mathfrak{k}}u_{j}^{\mu}.

Comparing this with (40) through the equivalence (44), and ignoring non-zero 𝕂\mathbb{K}-scalars,

κ𝔨,j=q=jmin(nw+j,jμ)μnw(qj)γq,𝔨.\kappa_{\mathfrak{k},j}=\sum_{q=j}^{\min(n-w+j,j_{\mu})}\mu^{n-w-(q-j)}\gamma_{q,\mathfrak{k}}.

There therefore exists some qq, denoted by QQ, with

(49) ev(μnw+jQγQ,𝔨)ev(κ𝔨,j).ev(\mu^{n-w+j-Q}\gamma_{Q,\mathfrak{k}})\leq ev(\kappa_{\mathfrak{k},j}).

Letting Q=jQ=j in (48),

(50) ev(κ𝔨)𝒜(γQ,𝔨uQμ)+C.ev(\kappa_{\mathfrak{k}})\leq{\cal A}(\gamma_{Q,\mathfrak{k}}u_{Q}^{\mu})+C.

Combining equations (43), (49), and (50),

ev(κ𝔨)\displaystyle ev(\kappa_{\mathfrak{k}}) 𝒜(μwn+Qjκ𝔨,juQμ)+C\displaystyle\leq{\cal A}(\mu^{w-n+Q-j}\kappa_{\mathfrak{k},j}u_{Q}^{\mu})+C
=(wn+Qj)ev(μ)+ev(κ𝔨,j)+𝒜(uQμ)+C\displaystyle=(w-n+Q-j)\cdot ev(\mu)+ev(\kappa_{\mathfrak{k},j})+{\cal A}(u_{Q}^{\mu})+C
(wn+Qj)ev(μ)+nev(μ)+ev(𝔶j)+𝒜(uQμ)+C\displaystyle\leq(w-n+Q-j)\cdot ev(\mu)+n\cdot ev(\mu)+ev(\mathfrak{y}_{j})+{\cal A}(u_{Q}^{\mu})+C
=wev(μ)+ev(𝔶j)+Qev(μ)+𝒜(uQμ)jev(μ)+C.\displaystyle=w\cdot ev(\mu)+ev(\mathfrak{y}_{j})+Q\cdot ev(\mu)+{\cal A}(u_{Q}^{\mu})-j\cdot ev(\mu)+C.

Recall that the winding number of x𝔨x_{\mathfrak{k}} is 𝔴(x𝔨)=w\mathfrak{w}(x_{\mathfrak{k}})=w. Let x𝔨x_{\mathfrak{k}} lie in the circle bundle of radius r(x𝔨)r(x_{\mathfrak{k}}). We conclude that

𝒜(κ𝔨x𝔨)\displaystyle{\cal A}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}}) =ev(κ𝔨)+𝒜(x𝔨)\displaystyle=ev(\kappa_{\mathfrak{k}})+{\cal A}(x_{\mathfrak{k}})
=ev(κ𝔨)wkπr(x𝔨)2+01H(x𝔨)dt\displaystyle=ev(\kappa_{\mathfrak{k}})-wk\pi r(x_{\mathfrak{k}})^{2}+\int_{0}^{1}H(x_{\mathfrak{k}})dt
w(ev(μ)kπr(x𝔨)2)+01H(x)dt+Qev(μ)+𝒜(uQμ)+ev(𝔶j)jev(μ)+C.\displaystyle\leq w(ev(\mu)-k\pi r(x_{\mathfrak{k}})^{2})+\int_{0}^{1}H(x)dt+Q\cdot ev(\mu)+{\cal A}(u_{Q}^{\mu})+ev(\mathfrak{y}_{j})-j\cdot ev(\mu)+C.

Because ev(μ)<kπR2ev(\mu)<k\pi R^{2}, and because r(x𝔨)r(x_{\mathfrak{k}}) approaches RR as 𝔴\mathfrak{w} approaches infinity,

w(ev(μ)kπr(x𝔨)2)Cw(ev(\mu)-k\pi r(x_{\mathfrak{k}})^{2})\leq C^{\prime}

for some fixed C>0C^{\prime}>0 independent of xx.

Furthermore, there is a constant CC^{\prime\prime} bounding the expression

01H(x)dt+Qev(μ)+𝒜(uQμ)𝒜(ujμ)jev(μ)+CC.\int_{0}^{1}H(x)dt+Q\cdot ev(\mu)+{\cal A}(u_{Q}^{\mu})-{\cal A}(u_{j}^{\mu})-j\cdot ev(\mu)+C\leq C^{\prime\prime}.

Let

Cμ=C+C.C_{\mu}=C^{\prime}+C^{\prime\prime}.

Then

𝒜(κ𝔨x𝔨)ev(𝔶j)+𝒜(ujμ)+Cμ.{\cal A}(\kappa_{\mathfrak{k}}x_{\mathfrak{k}})\leq ev(\mathfrak{y}_{j})+{\cal A}(u_{j}^{\mu})+C_{\mu}.

Take

𝒞=maxμCμ.{\cal C}=\max_{\mu}C_{\mu}.

We have shown that, for any jj, we can find a kk satisfying

𝒜(κkxk)𝒜(𝔶jujμ)+𝒞.{\cal A}(\kappa_{k}x_{k})\leq{\cal A}(\mathfrak{y}_{j}u_{j}^{\mu})+{\cal C}.

By definition,

𝒜(X)𝒜(κkxk){\cal A}(X)\leq{\cal A}(\kappa_{k}x_{k})

for each kk. Combining this with the non-Archimedean nature of 𝒜{\cal A},

𝒜(X)minj𝒜(𝔶jujμ)+𝒞𝒜(Y0)+𝒞.{\cal A}(X)\leq\min_{j}{\cal A}(\mathfrak{y}_{j}u_{j}^{\mu})+{\cal C}\leq{\cal A}(Y_{0})+{\cal C}.

The result follows.

Lemma 19 shows us how cochains in a fixed eigensummand behave. However, we would like a result for general cochains. To this end, define π\pi_{\leq} to be the projection

π:CF(H0)ev(μ)kπR2CFμ(H0).\pi_{\leq}:CF^{*}(H_{0})\rightarrow\bigoplus_{ev(\mu)\leq k\pi R^{2}}CF^{*}_{\mu}(H_{0}).

Lemma 19 has the following easy corollary.

Corollary 6

Choose any XCF(Hn)X\in CF^{*}(H_{n}) such that

π𝒮n𝔠HG(X)0.\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X)\neq 0.

Let Yev(μ)kπR2CFμ(H0)Y\in\bigoplus\limits_{ev(\mu)\leq k\pi R^{2}}CF^{*}_{\mu}(H_{0}) satisfy

(c1G𝒮)n(Y)=π𝒮n𝔠HG(X).(c_{-1}^{G}\circ{\cal S})^{n}(Y)=\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X).

Then

𝒜(X)𝒜(Y)<C.{\cal A}(X)-{\cal A}(Y)<C.
Proof.

First note that, if

(c1G𝒮)n(Y)=π𝒮n𝔠HG(X).(c_{-1}^{G}\circ{\cal S})^{n}(Y)=\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X).

then

(c1G𝒮)n(πμ(Y))=πμ𝒮n𝔠HG(X).(c_{-1}^{G}\circ{\cal S})^{n}(\pi_{\mu}(Y))=\pi_{\mu}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X).

for each μ\mu with valuation ev(μ)kπR2ev(\mu)\leq k\pi R^{2}. By Lemma 19, there exists a constant C>0C>0 such that

𝒜(X)𝒜(πμ(Y))<C{\cal A}(X)-{\cal A}(\pi_{\mu}(Y))<C

for all μkπR2\mu\leq k\pi R^{2}. By the non-Archimedean property of 𝒜{\cal A},

𝒜(Y)minμ𝒜(πμ(Y)).{\cal A}(Y)\geq\min_{\mu}{\cal A}(\pi_{\mu}(Y)).

Combining these two inequalities,

𝒜(X0)𝒜(Y)𝒜(X)minμ𝒜(πμ(Y))<C.{\cal A}(X_{0})-{\cal A}(Y)\leq{\cal A}(X)-\min_{\mu}{\cal A}(\pi_{\mu}(Y))<C.

Finally, we would like a cohomological analogue of Corollary 6. Denote by

πλ:HF(G0)HFλ(G0)\pi_{\lambda}:HF^{*}(G_{0})\longrightarrow HF^{*}_{\lambda}(G_{0})

projection onto the λ\lambda-generalized eigenspace, and, abusing notation, denote by π\pi_{\leq} the projection

π:HF(H0)ev(λ)kπR2HF(H0).\pi_{\leq}:HF^{*}(H_{0})\rightarrow\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}(H_{0}).

Let us recall some basic linear algebra facts about maps acting on finite-dimensional vector spaces.

  1. 1.

    Any eigenvalue of 𝒄𝟏𝑮𝓢\bm{c_{-1}^{G}\circ{\cal S}} is an eigenvalue of c1G𝒮c_{-1}^{G}\circ{\cal S}.

  2. 2.

    Without loss of generality, we can assume that the collection of generalized eigenvectors {v1λ,,vkλλ}\{v_{1}^{\lambda},...,v_{k_{\lambda}}^{\lambda}\} are chosen to equal

    {ujλkλλ+im(),,ujλλ+im()}\left\{u_{j_{\lambda}-k_{\lambda}}^{\lambda}+\textnormal{im}(\partial),...,u_{j_{\lambda}}^{\lambda}+\textnormal{im}(\partial)\right\}
  3. 3.

    Restricting to the invariant subspace ker()\ker(\partial), we can abuse notation and assume that {\cal B} is an eigenbasis for ker()\ker(\partial).

  4. 4.

    If [X]HFλ(H)[X]\in HF^{*}_{\lambda}(H), then (πλ(X))=0\partial(\pi_{\lambda}(X))=0.

Corollary 7

Let XHF(Hn)X\in HF^{*}(H_{n}) such that

π𝓢n𝔠HG(X)0.\pi_{\leq}\circ\bm{{\cal S}}^{n}\circ\mathfrak{c}^{HG}(X)\neq 0.

Let Yev(λ)kπR2HFλ(G0)Y\in\bigoplus\limits_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(G_{0}) satisfy

(𝒄𝟏𝑮𝓢)n(Y)=π𝓢n𝔠HG(X).(\bm{c_{-1}^{G}\circ{\cal S}})^{n}(Y)=\pi_{\leq}\circ\bm{{\cal S}}^{n}\circ\mathfrak{c}^{HG}(X).

Let X0X_{0} be any cochain-level representative of XX. Then

𝒜(X0)𝓐(Y)<C.{\cal A}(X_{0})-\bm{{\cal A}}(Y)<C.
Proof.

Let Y0Y_{0} be a minimum-action cochain-level representative of YY. Note that

Y0(ev(λ)kπR2CFλ(H0))im().Y_{0}\in\left(\bigoplus_{ev(\lambda)\leq k\pi R^{2}}CF^{*}_{\lambda}(H_{0})\right)\cup\textnormal{im}(\partial).

Decompose Y0Y_{0} into components Z0Z_{0} and W0W_{0}, where

W0ev(λ)kπR2CFλ(H0)andZ0im(),W_{0}\in\bigoplus_{ev(\lambda)\leq k\pi R^{2}}CF^{*}_{\lambda}(H_{0})\hskip 28.45274pt\text{and}\hskip 28.45274ptZ_{0}\in\textnormal{im}(\partial),

and πλ(Z0)=0\pi_{\lambda}(Z_{0})=0 when ev(λ)kπR2ev(\lambda)\leq k\pi R^{2}. Clearly W0ker()W_{0}\in\ker(\partial) and

[(c1G𝒮)n(W0)]=π[𝒮n𝔠HG(X0)]=[π𝒮n𝔠HG(X0)].[(c_{-1}^{G}\circ{\cal S})^{n}(W_{0})]=\pi_{\leq}\circ[{\cal S}^{n}\circ\mathfrak{c}^{HG}(X_{0})]=[\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X_{0})].

Indeed, there exists

βim()ev(λ)kπR2CFλ(H0)\beta\in\textnormal{im}(\partial)\cap\bigoplus_{ev(\lambda)\leq k\pi R^{2}}CF^{*}_{\lambda}(H_{0})

such that

(c1G𝒮)n(W0)+β=π𝒮n𝔠HG(X0).(c_{-1}^{G}\circ{\cal S})^{n}(W_{0})+\beta=\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X_{0}).

As (c1G𝒮)n(c_{-1}^{G}\circ{\cal S})^{n} is an isomorphism on CFλ0(H0)CF^{*}_{\lambda\neq 0}(H_{0}) and im()\textnormal{im}(\partial) is an invariant subspace, β\beta has a well-defined inverse β1im()\beta^{-1}\in\textnormal{im}(\partial), and

(c1G𝒮)n(W0+β1)=π𝒮n𝔠HG(X0).(c_{-1}^{G}\circ{\cal S})^{n}(W_{0}+\beta^{-1})=\pi_{\leq}\circ{\cal S}^{n}\circ\mathfrak{c}^{HG}(X_{0}).

By Lemma 19, there exists C>0C^{\prime}>0 with

𝒜(X0)𝒜(W0+β1)<C.{\cal A}(X_{0})-{\cal A}(W_{0}+\beta^{-1})<C^{\prime}.

But [Y0]=[W0+β1][Y_{0}]=[W_{0}+\beta^{-1}], and so, by Lemma 14, there exists C>0C^{\prime\prime}>0 with

|𝒜(Y0)𝒜(W0+β1)|<C.|{\cal A}(Y_{0})-{\cal A}(W_{0}+\beta^{-1})|<C^{\prime\prime}.

Setting C=C+CC=C^{\prime}+C^{\prime\prime},

𝒜(X0)𝒜(Y0)<C.{\cal A}(X_{0})-{\cal A}(Y_{0})<C.

We are finally ready to prove Proposition 2.

Proof of Proposition 2: By Lemma 12,

ev(λ)>kπR2HFλ(H0)ker(Φa).\bigoplus_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\subset\ker(\Phi_{a}).

Clearly

HF(a,)(H0)ker(Φa).HF^{*}_{(a,\infty)}(H_{0})\subset\ker(\Phi_{a}).

By the linearity of Φa\Phi_{a}, it therefore suffices to show the inclusion

im(π|ker(Φa))ev(λ)kπR2HFλ,(aC,)(H0).\textnormal{im}\left(\pi_{\leq}\big{|}_{\ker(\Phi_{a})}\right)\subset\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda,(a-C,\infty)}(H_{0}).

Let

𝐗ev(λ)kπR2HFλ(H0){\bf X}\in\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(H_{0})

Suppose that 𝐗ker(Φa){\bf X}\in\ker(\Phi_{a}). Let YSC(H)Y\in SC^{*}(H) be a cocycle descending in cohomology to

[Y]=Φ(𝐗).[Y]=\Phi({\bf X}).

As

πa([Y])=Φa(𝐗)=0,\pi_{a}([Y])=\Phi_{a}({\bf X})=0,

we can assume without loss of generality that

(51) 𝒜(Y)>a.{\cal A}(Y)>a.

Let XCF(H0)X\in CF^{*}(H_{0}) be any representative of 𝐗{\bf X}, and let nn\in\mathbb{N} any integer with 𝔴(X)n\mathfrak{w}(X)\leq n. Under the PSS map,

[X]=[Y][X]=[Y]

in SH(H)SH^{*}(H). By homotopy-invariance of continuation maps,

(𝒄𝟏𝑮𝓢)n([X])=𝓢n𝔠HG([Y]).(\bm{c_{-1}^{G}\circ{\cal S}})^{n}([X])=\bm{{\cal S}}^{n}\circ\mathfrak{c}^{HG}([Y]).

By Corollary 7,

𝒜(Y)𝒜(X)<C,{\cal A}(Y)-{\cal A}(X)<C,

or

𝒜(X)>𝒜(Y)C>aC.{\cal A}(X)>{\cal A}(Y)-C>a-C.

It follows that

𝐗im(HF(aC,)(H0)HF(H0)).{\bf X}\in\textnormal{im}(HF^{*}_{(a-C,\infty)}(H_{0})\rightarrow HF^{*}(H_{0})).

4.2.3 The full PSS map

We have characterized ker(Φa)\ker(\Phi_{a}); we now want to characterize ker(Φ)\ker(\Phi).

Proposition 3

The isomorphism

im(Φa)HF(H0)/ker(Φa)\textnormal{im}(\Phi_{a})\cong{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\ker(\Phi_{a})$}\right.}

induces an isomorphism

(52) im(Φ)HF(H0)/limaker(Φa)\textnormal{im}(\Phi)\cong{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker({\Phi_{a}})$}\right.}

where the connecting maps for

limaker(Φa)\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker({\Phi_{a}})

are inclusions.

Proof.

Let

(53) πa,a:SHa(H)SHa(H)\pi_{a,a^{\prime}}:SH^{*}_{a}(H)\longrightarrow SH^{*}_{a^{\prime}}(H)

be the connecting map in the inverse limit. By definition,

Φa=πa,aΦa.\Phi_{a^{\prime}}=\pi_{a,a^{\prime}}\circ\Phi_{a}.

Thus,

ker(Φa)=ker(πa,aΦa)ker(Φa).\ker(\Phi_{a^{\prime}})=\ker(\pi_{a,a^{\prime}}\circ\Phi_{a})\supset\ker(\Phi_{a}).

Similarly,

πa,a(im(Φa))=im(Φa).\pi_{a,a^{\prime}}(\textnormal{im}(\Phi_{a}))=\textnormal{im}(\Phi_{a^{\prime}}).

Let

ιa,a:ker(Φa)ker(Φa){\iota_{a,a^{\prime}}}:\ker({\Phi_{a}})\longrightarrow\ker({\Phi_{a^{\prime}}})

be the inclusion. Abusing notation, let

πa,a:im(Φa)im(Φa)\pi_{a,a^{\prime}}:\textnormal{im}(\Phi_{a})\longrightarrow\textnormal{im}(\Phi_{a^{\prime}})

be the surjective map induced by restricting the connecting map (53). The diagram

0{0}ker(Φa){\ker({\Phi_{a}})}HF(H0){HF^{*}(H_{0})}im(Φa){\textnormal{im}(\Phi_{a})}0{0}0{0}ker(Φa){\ker({\Phi_{a^{\prime}}})}HF(H0){HF^{*}(H_{0})}im(Φa){\textnormal{im}(\Phi_{a^{\prime}})}0{0}ιa,a\scriptstyle{{\iota_{a,a^{\prime}}}}=\scriptstyle{=}πa,a\scriptstyle{{\pi_{a,a^{\prime}}}}

commutes. As the inverse limit is left-exact, there is an exact sequence

(54) 0limaker(Φa)HF(H0)limaim(Φa)lima1ker(Φa).0\longrightarrow\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker({\Phi_{a}})\longrightarrow HF^{*}(H_{0})\longrightarrow\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\textnormal{im}(\Phi_{a})\longrightarrow\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\ker({\Phi_{a}}).

We will show that lima1ker(Φa)=0\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\ker({\Phi_{a}})=0.

By Proposition 2 there exists some set of constants {Cλ,a}λkπR2\left\{C_{\lambda,a}\in\mathbb{R}\right\}_{\lambda\leq k\pi R^{2}}, such that

ker(Φa)=λ>kπR2HFλ(H0)λkπR2HFλ,(Cλ,a,)(H0)\ker({\Phi_{a}})=\bigoplus_{\lambda>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\oplus\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a},\infty)}(H_{0})

and limaCλ,a=\lim\limits_{a\rightarrow\infty}C_{\lambda,a}=\infty. Inverse limits commute with finite sums, so that

lima1ker(Φa)\displaystyle\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\ker({\Phi_{a}}) =lima1λ>kπR2HFλ(H0)lima1λkπR2HFλ,(Cλ,a,)(H0)\displaystyle=\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\bigoplus_{\lambda>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\oplus\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a},\infty)}(H_{0})
=lima1λkπR2HFλ,(Cλ,a,)(H0).\displaystyle=\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a},\infty)}(H_{0}).

The second equality holds because the connecting maps of the system

limaλ>kπR2HFλ(H0)\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\bigoplus_{\lambda>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})

are isomorphisms, and the system therefore satisfies the Mittag-Leffler condition.

Recall from Remark Remark 5) that the index aa belongs to a fixed countable sequence a1,a2,a3,a_{1},a_{2},a_{3},.... By definition,

(55) lima1λkπR2H\displaystyle\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}\bigoplus_{\lambda\leq k\pi R^{2}}H Fλ,(Cλ,a,)(H0)\displaystyle F^{*}_{\lambda,(C_{\lambda,a},\infty)}(H_{0})
=coker(i=1λkπR2HFλ,(Cλ,ai,)iιai,ai+1idi=1λkπR2HFλ,(Cλ,ai,)(H0)).\displaystyle=\textnormal{coker}\left(\prod_{i=1}^{\infty}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}\xrightarrow{\prod_{i}{\iota_{a_{i},a_{i+1}}}-{id}}\prod_{i=1}^{\infty}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0})\right).

Choose any element

i=1𝐗𝐢i=1λkπR2HFλ,(Cλ,ai,)(H0).\sum_{i=1}^{\infty}{\bf X_{i}}\in\prod_{i=1}^{\infty}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0}).

For each i+i\in\mathbb{N}_{+}, consider the formal sum

𝐘𝐢=j=i(1)ji+1𝐗𝐣.{\bf Y_{i}}=\sum_{j=i}^{\infty}(-1)^{j-i+1}{\bf X_{j}}.

As 𝐗𝐢λkπR2HFλ,(Cλ,ai,)(H0){\bf X_{i}}\in\bigoplus\limits_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0}), it has a cochain representative XiX_{i} such that 𝒜(Xi)>minλCλ,ai{\cal A}(X_{i})>\min\limits_{\lambda}C_{\lambda,a_{i}}. This implies that

limi𝒜(Xi)=.\lim_{i\rightarrow\infty}{\cal A}(X_{i})=\infty.

Thus,

j=i(1)jiXjλkπR2CFλ(H0).\sum_{j=i}^{\infty}(-1)^{j-i}X_{j}\in\bigoplus_{\lambda\leq k\pi R^{2}}CF^{*}_{\lambda}(H_{0}).

More particularly,

j=i(1)jiXjλkπR2CFλ,(Cλ,ai,)(H0),\sum_{j=i}^{\infty}(-1)^{j-i}X_{j}\in\bigoplus_{\lambda\leq k\pi R^{2}}CF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0}),

is a cochain representative whose cohomology class is given by 𝐘𝐢{\bf Y_{i}}. We conclude that

𝐘𝐢im(λkπR2HFλ,(Cλ,ai,)(H0)λkπR2HFλ(H0)).{\bf Y_{i}}\in\textnormal{im}\left(\bigoplus\limits_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0})\rightarrow\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\right).

Identify 𝐘𝐢{\bf Y_{i}} with some preimage in λkπR2HFλ,(Cλ,ai,)(H0)\bigoplus\limits_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{\lambda,a_{i}},\infty)}(H_{0}). There is a telescoping sum

i=1(ιai,ai+1id)(𝐘𝐢)\displaystyle\prod_{i=1}^{\infty}\left(\iota_{a_{i},a_{i+1}}-id\right)\left(\sum{\bf Y_{i}}\right) =i=1𝐘𝐢+𝟏𝐘𝐢\displaystyle=\sum_{i=1}^{\infty}{\bf Y_{i+1}}-{\bf Y_{i}}
=i=1𝐗𝐢+j=i+1(1)ji𝐗𝐣+(1)ji+1𝐗𝐣\displaystyle=\sum_{i=1}^{\infty}{\bf X_{i}}+\sum_{j=i+1}^{\infty}(-1)^{j-i}{\bf X_{j}}+(-1)^{j-i+1}{\bf X_{j}}
=i=1𝐗𝐢.\displaystyle=\sum_{i=1}^{\infty}{\bf X_{i}}.

We have shown that the map iιai,ai+1id\prod_{i}\iota_{a_{i},a_{i+1}}-id is surjective, and so, from the definition (55),

lima1ker(Φa)=0.\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}{}^{1}\ker({\Phi_{a}})=0.

Corollary 8

There is an isomorphism

im(Φ)HF(H0)/ev(λ)>kπR2HFλ(H0).\textnormal{im}(\Phi)\simeq{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}.
Proof.

It suffices to show the equality

limaker(Φa)=λ>kπR2HFλ(H0).\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker(\Phi_{a})=\bigoplus_{\lambda>k\pi R^{2}}HF^{*}_{\lambda}(H_{0}).

Proposition 3 shows that the connecting maps of limaker(𝚽𝒂)\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker(\bm{\Phi_{a}}) are inclusions; and so the inverse limit is the intersection:

limaker(𝚽𝒂)=aker(𝚽𝒂).\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\ker(\bm{\Phi_{a}})=\bigcap\limits_{a}\ker(\bm{\Phi_{a}}).

By Proposition 2,

ker(𝚽𝒂)=λ>kπR2HFλ(H0)λkπR2HFλ,(Ca,λ,)(H0)\ker(\bm{\Phi_{a}})=\bigoplus_{\lambda>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})\oplus\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{a,\lambda},\infty)}(H_{0})

for some constants Ca,λC_{a,\lambda} with limiting behavior limaCa,λ\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}C_{a,\lambda}\rightarrow\infty. The only element of CF(H0)CF^{*}(H_{0}) with arbitrarily large action is 0, and so

aλkπR2HFλ,(Ca,λ,)(H0)=0.\bigcap_{a}\bigoplus_{\lambda\leq k\pi R^{2}}HF^{*}_{\lambda,(C_{a,\lambda},\infty)}(H_{0})=0.

By definition,

HFλ(H0)HFλ(H0)=0HF^{*}_{\lambda}(H_{0})\cap HF^{*}_{\lambda^{\prime}}(H_{0})=0

whenever λλ\lambda\neq\lambda^{\prime}. The Corollary follows.

4.3 The product structure

Recall that SH(H)SH^{*}(H) has a product structure, called the pair of pants product. The product structure defined on an inverse limit of action-filtered Floer groups was first considered in [cieliebak-o].

Let Σ\Sigma be a Riemann surface with two positive punctures denoted by p1,p2p_{1},p_{2} and one negative puncture denoted by qq. Choose collar neighborhoods (0,ϵ]×S1(0,\epsilon]\times S^{1} of both p1p_{1} and p2p_{2} and [ϵ,0)×S1[-\epsilon,0)\times S^{1} of qq. Equip Σ\Sigma with a one-form β\beta such that β=w1dt\beta=w_{1}dt on the collar neighborhood of p1p_{1}, β=w2dt\beta=w_{2}dt on the collar neighborhood of p2p_{2}, and β=w0dt\beta=w_{0}dt on the collar neighborhood of qq, for fixed positive integers w0,w1,w2w_{0},w_{1},w_{2} satisfying w0w1+w2w_{0}\geq w_{1}+w_{2}. As shown in [ritter-fano], we may assume that β\beta also satisfies

dβ0.d\beta\leq 0.

For a fixed Hamiltonian HiH_{i}, let u:ΣEu:\Sigma\longrightarrow E satisfy

(duβXHi)(0,1)=0,(du-\beta\otimes X_{H_{i}})^{(0,1)}=0,

with respect to a generic Ω\Omega-tame, upper-triangular almost-complex structure. The energy of uu is defined as

E(u)=12Σ||duβXHi||2volΣE(u)=\frac{1}{2}\int_{\Sigma}||du-\beta\otimes X_{H_{i}}||^{2}vol_{\Sigma}

We call uu a pair-of-pants. The following Lemma is a standard application of the integrated maximum principal.

Lemma 20

Suppose uu has finite energy. The image of uu remains inside DRD_{R}.

We sketch the proof. It is essentially a mash-up of the proof of Lemma 7 and Lemma 9.7 in [ritter-fano].

Proof.

The positive punctures of uu converge to periodic orbits of HiH_{i}, and are therefore contained in the interior of DRD_{R}. The negative puncture of uu converges to a periodic orbit of nHinH_{i}, and is therefore also contained in the interior DRD_{R}. Let DσD_{\sigma} be a generic disk bundle in DRD_{R} containing the three asymptotic periodic orbits. Define Σ=u1(EDσ)\Sigma^{\prime}=u^{-1}(E\setminus D_{\sigma}), and suppose for contradiction that Σ\Sigma^{\prime} is non-empty. Then

E(u|Σ)\displaystyle E(u\big{|}_{\Sigma^{\prime}}) :=12Σ||duXHiβ||2volΣ\displaystyle:=\frac{1}{2}\int_{\Sigma^{\prime}}||du-X_{H_{i}}\otimes\beta||^{2}vol_{\Sigma^{\prime}}
=Σud((1+πr2)α)udHiβ\displaystyle=\int_{\Sigma^{\prime}}u^{*}d((1+\pi r^{2})\alpha)-u^{*}dH_{i}\wedge\beta
=Σud((1+πr2)α)ud(Hiβ)+uHidβ\displaystyle=\int_{\Sigma^{\prime}}u^{*}d((1+\pi r^{2})\alpha)-u^{*}d(H_{i}\wedge\beta)+u^{*}H_{i}d\beta
Σud((1+πr2)α)ud(Hiβ)\displaystyle\leq\int_{\Sigma^{\prime}}u^{*}d((1+\pi r^{2})\alpha)-u^{*}d(H_{i}\wedge\beta)
=Σu(1+πσ2)αuHiβ,\displaystyle=\int_{\partial\Sigma^{\prime}}u^{*}(1+\pi{\sigma}^{2})\alpha-u^{*}H_{i}\beta,

where the inequality follows because dβ0d\beta\leq 0 and Hi0H_{i}\geq 0 everywhere.

Let cσ=hi(kπσ2)hi(kπσ2)kπσ2c_{\sigma}=h_{i}(k\pi{\sigma}^{2})-h_{i}^{\prime}(k\pi{\sigma}^{2})k\pi{\sigma}^{2} be the yy-intercept of the tangent line to hih_{i} at kπσ2k\pi{\sigma}^{2}. Then on Dσ\partial D_{\sigma}, Hi=hi(kπσ2)kπσ2+(1+kπσ2)ρf+cσH_{i}=h_{i}^{\prime}(k\pi{\sigma}^{2})k\pi{\sigma}^{2}+(1+k\pi{\sigma}^{2})\rho^{*}f+c_{\sigma}. Thus,

E(u|Σ)\displaystyle E(u\big{|}_{\Sigma^{\prime}}) Σu(1+πσ2)αhi(kπσ2)kπσ2β+(1+kπσ2)ρfβ+cσβ\displaystyle\leq\int_{\partial\Sigma^{\prime}}u^{*}(1+\pi{\sigma}^{2})\alpha-h_{i}^{\prime}(k\pi{\sigma}^{2})k\pi{\sigma}^{2}\beta+(1+k\pi{\sigma}^{2})\rho^{*}f\beta+c_{\sigma}\beta
=Σ(1+kπσ2)α(duXHiβ)+(hi(kπσ2)cσ)β\displaystyle=\int_{\partial\Sigma^{\prime}}(1+k\pi{\sigma}^{2})\alpha(du-X_{H_{i}}\otimes\beta)+(h^{\prime}_{i}(k\pi{\sigma}^{2})-c_{\sigma})\beta
=Σ1+kπσ22kπσ2drduj+(hi(kπσ2)cσ)β\displaystyle=\int_{\partial\Sigma^{\prime}}-\frac{1+k\pi{\sigma}^{2}}{2k\pi{\sigma}^{2}}dr\circ du\circ j+(h_{i}^{\prime}(k\pi{\sigma}^{2})-c_{\sigma})\beta
Σ(hi(kπσ2)cσ)β\displaystyle\leq\int_{\partial\Sigma^{\prime}}(h_{i}^{\prime}(k\pi{\sigma}^{2})-c_{\sigma})\beta
=Σ(hi(kπσ2)cσ)dβ.\displaystyle=\int_{\Sigma^{\prime}}(h_{i}^{\prime}(k\pi{\sigma}^{2})-c_{\sigma})d\beta.

By assumption dβ0d\beta\leq 0 everywhere. Furthermore, hi0h_{i}^{\prime}\geq 0 on Σσ\Sigma_{\sigma}, and, as the yy-intercept of hih_{i}, cσ0c_{\sigma}\leq 0. Therefore,

Σ(hi(kπσ2)cσ)dβ0.\int_{\Sigma^{\prime}}(h_{i}^{\prime}(k\pi{\sigma}^{2})-c_{\sigma})d\beta\leq 0.

We conclude that

E(u|Σ)0.E(u\big{|}_{\Sigma^{\prime}})\leq 0.

But E(u|Σ)E(u\big{|}_{\Sigma^{\prime}}) is non-negative by definition. Therefore E(u|Σ)=0E(u\big{|}_{\Sigma^{\prime}})=0. This can only occur if

du=XHiβdu=X_{H_{i}}\otimes\beta

everywhere. As XHiX_{H_{i}} is contained in the horizontal distribution determined by α\alpha, the image of u|Σu\big{|}_{\Sigma^{\prime}} must be entirely contained in Σσ\Sigma_{\sigma}. But this contradicts the first observation made, that near the punctures of Σ\Sigma, uu approaches periodic orbits of HiH_{i} or nHinH_{i}.

Fix a Hamiltonian HH_{\ell}. For a fixed index ii\in\mathbb{N}, define a branch to be a cascade

𝐮𝐢=(cmii,umii,cmi1i,umi1,,c0i){\bf u^{i}}=(c_{m^{i}}^{i},u_{m^{i}}^{i},c_{m^{i}-1}^{i},u_{m^{i}-1},...,c_{0}^{i})

associated to a sequence of periodic orbits xmii,,x0ix_{m^{i}}^{i},...,x_{0}^{i} and a non-decreasing sequence of integer weights wmii,,w0iw_{m^{i}}^{i},...,w_{0}^{i}\in\mathbb{Z}, where

  1. 1.

    cjiim(xji)c_{j}^{i}\in im(x_{j}^{i}),

  2. 2.

    ujiu_{j}^{i} is a finite-energy Floer solution corresponding to an ss-family of Hamiltonians HjiH_{j}^{i}, where Hji=wjiHH_{j}^{i}=w_{j}^{i}H_{\ell} when s>>0s>>0 and Hji=wj+1iHH_{j}^{i}=w_{j+1}^{i}H_{\ell} when s<<0s<<0,

  3. 3.

    limsuji(s,0)\lim\limits_{s\rightarrow\infty}u_{j}^{i}(s,0) is in the stable manifold of cjic_{j}^{i} (or cji=limsuji(s,0)c_{j}^{i}=\lim\limits_{s\rightarrow\infty}u_{j}^{i}(s,0) if xjix_{j}^{i} is constant),

  4. 4.

    cjic_{j}^{i} is in the unstable manifold of limsuj+1i(s,0)\lim\limits_{s\rightarrow-\infty}u_{j+1}^{i}(s,0) (or cji=limsuj+1i(s,t)c_{j}^{i}=\lim\limits_{s\rightarrow-\infty}u_{j+1}^{i}(s,t) if xjix_{j}^{i} is constant).

Define a tree 𝐭{\bf t} to be a sequence (𝐮𝟎,𝐮𝟏,𝐮𝟐,v)({\bf u^{0}},{\bf u^{1}},{\bf u^{2}},v) where vv is a pair-of pants corresponding to weights w0=wm00w_{0}=w_{m^{0}}^{0}, w1=w01w_{1}=w_{0}^{1}, and w2=w02w_{2}=w_{0}^{2}; and where

  1. 1.

    taken in the collar neighborhood of p1p_{1}, limsv(s,0)\lim\limits_{s\rightarrow\infty}v(s,0) lies in the stable manifold of c01c_{0}^{1},

  2. 2.

    taken in the collar neighborhood of p2p_{2}, limsv(s,0)\lim\limits_{s\rightarrow\infty}v(s,0) lies in the stable manifold of c02c_{0}^{2},

  3. 3.

    and taken in the collar neighborhood of qq, limsv(s,0)\lim\limits_{s\rightarrow-\infty}v(s,0) lies in the unstable manifold of cm00c_{m^{0}}^{0}.

Let n0(x,y,z){\cal M}^{n}_{0}(x,y,z) be the moduli space of rigid trees 𝐭=(𝐮𝟎,𝐮𝟏,𝐮𝟐,v){\bf t}=({\bf u^{0}},{\bf u^{1}},{\bf u^{2}},v), where cm11c_{m^{1}}^{1} is in the stable manifold of xx, xm22x_{m^{2}}^{2} is in the stable manifold of yy, x00x_{0}^{0} is in the stable manifold of zz, and wm11=wm22=1w_{m^{1}}^{1}=w_{m^{2}}^{2}=1 and w00=nw_{0}^{0}=n. See Figure 6.

Refer to caption
Figure 6: A tree in n0(x,y,z){\cal M}^{n}_{0}(x,y,z).

A standard Corollary to Lemma 20 is that

Corollary 9

The moduli space n0(x,y,z){\cal M}^{n}_{0}(x,y,z) is compact.

Define a map

μ2,ni:CF(Hi)CF(Hi)\displaystyle\mu^{2,n}_{i}:CF^{*}(H_{i})\otimes CF^{*}(H_{i}) CF(nHi)\displaystyle\longrightarrow CF^{*}(nH_{i})
(x,y)\displaystyle(x,y) 𝐮0n(x,y,z)±TΩ((x~#𝐮#y~)#z~)z,\displaystyle\mapsto\sum_{{\bf u}\in{\cal M}_{0}^{n}(x,y,z)}\pm T^{-\Omega((-\tilde{x}\#{\bf u}\#-\tilde{y})\#\tilde{z})}z,

for each nn. Extend μ2,ni\mu^{2,n}_{i} to a map

μ2,ni,j:CF(Hi)CF(Hj)CF(nHi+j)\mu^{2,n}_{i,j}:CF^{*}(H_{i})\otimes CF^{*}(H_{j})\longrightarrow CF^{*}(nH_{i+j})

where μ2,ni,j(x,y)=μ2,ni+j(cj(x),ci(y))\mu^{2,n}_{i,j}(x,y)=\mu^{2,n}_{i+j}(c^{j}(x),c^{i}(y)). Recall that the continuation maps were chosen to be inclusions, so, μ2,ni,j\mu^{2,n}_{i,j} is just the composition of the inclusions

CF(Hi)CF(Hi+j)andCF(Hj)CF(Hi+j)CF^{*}(H_{i})\hookrightarrow CF^{*}(H_{i+j})\hskip 28.45274pt\text{and}\hskip 28.45274ptCF^{*}(H_{j})\hookrightarrow CF^{*}(H_{i+j})

with the pair-of-pants product.

As Hi+j0H_{i+j}\geq 0, there are action-increasing continuation maps

ηi+j:CF(nHi+j)CF((n+1)Hi+j)\eta_{i+j}:CF^{*}(nH_{i+j})\longrightarrow CF^{*}((n+1)H_{i+j})

Omitting signs, define a map

μ2:CF(Hi)[𝐪]CF(Hj)[𝐪]CF(4Hi+j)[𝐪]\mu^{2}:CF^{*}(H_{i})[{\bf q}]\otimes CF^{*}(H_{j})[{\bf q}]\longrightarrow CF^{*}(4H_{i+j})[{\bf q}]

by

μ(x,y)\displaystyle\mu(x,y) =ηi+j2μ2,2i,j(x,y)\displaystyle=\eta_{i+j}^{2}\circ\mu^{2,2}_{i,j}(x,y)
μ(x,y𝐪)=μ(x𝐪,y)\displaystyle\mu(x,y{\bf q})=\mu(x{\bf q},y) =ηi+j2μ2,2i,j(x,y)𝐪+ηi+jμ2,3i,j(x,y)\displaystyle=\eta_{i+j}^{2}\circ\mu^{2,2}_{i,j}(x,y){\bf q}+\eta_{i+j}\circ\mu^{2,3}_{i,j}(x,y)
μ(x𝐪,y𝐪)\displaystyle\mu(x{\bf q},y{\bf q}) =μ2,4i,j(x,y).\displaystyle=\mu^{2,4}_{i,j}(x,y).

See [abouzaid-s] for a careful treatment of signs. We omit the proof of the following Lemma, which is a standard cylinder-breaking analysis. See, for example, [ritter-s] in the case of the usual non-degenerate product structure in symplectic cohomology and Appendix A in [auroux] in the case of cascades.

Lemma 21

μ2\mu^{2} descends to a map

μ2:SH(H)SH(H)SH(4H)SH(H).{\mu^{2}}:SH^{*}(H)\otimes SH^{*}(H)\longrightarrow SH^{*}(4H)\cong SH^{*}(H).

The energy of a map u:ΣEu:\Sigma\longrightarrow E with positive punctures mapping to xx and yy and negative puncture mapping to zz is

0E(u)=𝒜(TΩ((x~#u#y~)#z~)z)𝒜(x)𝒜(y).0\leq E(u)={\cal A}(T^{-\Omega((-\tilde{x}\#u\#-\tilde{y})\#\tilde{z})}z)-{\cal A}(x)-{\cal A}(y).

Thus,

𝒜(TΩ((x~#u#y~)#z~)z)𝒜(x)+𝒜(y).{\cal A}(T^{-\Omega((-\tilde{x}\#u\#-\tilde{y})\#\tilde{z})}z)\geq{\cal A}(x)+{\cal A}(y).

It follows that μ2\mu^{2} induces a well-defined product μ^2\hat{\mu}_{2} on SC^(H)\widehat{SC^{*}}(H), and, in particular, on ker()^\widehat{\ker(\partial)}.

Lemma 22

The product μ^2\hat{\mu}_{2} induces a well-defined product on SH¯(H)\overline{SH^{*}}(H).

Proof.

The ordinary product satisfies

μ2(X,Y)=μ2(X,Y)+μ2(X,Y)\partial\circ\mu_{2}(X,Y)=\mu_{2}\circ(\partial X,Y)+\mu_{2}\circ(X,\partial Y)

for cochains X,YSC(H)X,Y\in SC^{*}(H). Suppose that X,Yker()^X,Y\in\widehat{\ker(\partial)}. Write

X=iXiandY=jYj,X=\sum_{i}X_{i}\hskip 28.45274pt\text{and}\hskip 28.45274ptY=\sum_{j}Y_{j},

where Xi,YjX_{i},Y_{j} are cocycles satisfying 𝒜(Xi){\cal A}(X_{i})\rightarrow\infty and 𝒜(Yj){\cal A}(Y_{j})\rightarrow\infty. Then

μ^2(X,Y)\displaystyle\partial\circ\hat{\mu}_{2}(X,Y) =i,jμ2(Xi,Yj)\displaystyle=\partial\circ\sum_{i,j}\mu_{2}(X_{i},Y_{j})
=i,jμ2(Xi,Yj)\displaystyle=\sum_{i,j}\partial\circ\mu_{2}(X_{i},Y_{j})
=i,jμ2(Xi,Yj)+μ2(Xi,Yj)\displaystyle=\sum_{i,j}\mu_{2}(\partial X_{i},Y_{j})+\mu_{2}(X_{i},\partial Y_{j})
=0.\displaystyle=0.

So μ^2\hat{\mu}_{2} is well-defined on ker()^\widehat{\ker(\partial)}. Now suppose Xim()¯X\in\overline{\textnormal{im}(\partial)}. Write

X=i(Zi)X=\sum_{i}\partial(Z_{i})

where 𝒜((Zi)){\cal A}(\partial(Z_{i}))\rightarrow\infty. Let Yker()^Y\in\widehat{\ker(\partial)}. Then

μ^2((Zi),Y)\displaystyle\hat{\mu}_{2}(\partial(Z_{i}),Y) =μ^2(Zi,Y)+μ^2(Zi,(Y))\displaystyle=\partial\circ\hat{\mu}_{2}(Z_{i},Y)+\hat{\mu}_{2}(Z_{i},\partial(Y))
=μ^2(Zi,Y).\displaystyle=\partial\circ\hat{\mu}_{2}(Z_{i},Y).

Thus,

𝒜(μ^2(Zi,Y))=𝒜(μ^2((Zi),Y))𝒜((Zi))+𝒜(Y).{\cal A}(\partial\circ\hat{\mu}_{2}(Z_{i},Y))={\cal A}(\hat{\mu}_{2}(\partial(Z_{i}),Y))\geq{\cal A}(\partial(Z_{i}))+{\cal A}(Y).

As 𝒜((Zi)){\cal A}(\partial(Z_{i}))\rightarrow\infty, we conclude that

𝒜(μ^2(Zi,Y)){\cal A}(\partial\circ\hat{\mu}_{2}(Z_{i},Y))\rightarrow\infty

as well. Summing over all ii,

μ^2(X,Y)=i(μ^2(Zi,Y))im()¯.\hat{\mu}_{2}(X,Y)=\sum_{i}\partial(\hat{\mu}_{2}(Z_{i},Y))\in\overline{\textnormal{im}(\partial)}.

The map

SC(H)ker(^)/im(dd)¯SC^{*}(H)\rightarrow{\raisebox{1.99997pt}{$\widehat{ker(\partial})$}\left/\raisebox{-1.99997pt}{$\overline{\textnormal{im}(dd)}$}\right.}

clearly intertwines μ2\mu_{2} and μ^2\hat{\mu}_{2}. Thus, the cohomology-level map

η:SH(H)SH¯(H)\eta:SH^{*}(H)\longrightarrow\overline{SH^{*}}(H)

is a ring isomorphism.

4.3.1 Finishing the proof

Proof of Theorem 1: Recall the injective map

SH¯(H)ϕSH^(H)limaSHa(H)\overline{SH^{*}}(H)\xrightarrow{\phi}\widehat{SH^{*}}(H)\simeq\lim_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}SH^{*}_{a}(H)

and the commutative diagram

(56) SH¯(H){\overline{SH^{*}}(H)}SH(H){SH^{*}(H)}SH^(H){\widehat{SH^{*}}(H)}ϕ\scriptstyle{\phi}η\scriptstyle{\eta}π\scriptstyle{\pi}

Given Corollary 8, it suffices to show that

im(π)=im(ϕ).\textnormal{im}(\pi)=\textnormal{im}(\phi).

Clearly

im(π)im(ϕ).\textnormal{im}(\pi)\subset\textnormal{im}(\phi).

We will show the reverse inclusion. Fix [X]SH¯(H)[X]\in\overline{SH^{*}}(H) and choose any cochain representative XX of [X][X]. By definition, there are cochains X1,X2,X3,X_{1},X_{2},X_{3},... in SC(H)SC^{*}(H) with 𝒜(Xi){\cal A}(X_{i})\rightarrow\infty, such that

X=i=1Xi.X=\sum_{i=1}^{\infty}X_{i}.

Let YiHF(H0)Y_{i}\in HF^{*}(H_{0}) be the descent of 𝒮n𝔠HG(Xi){\cal S}^{n}\circ\mathfrak{c}^{HG}(X_{i}) to cohomology. identifying YiY_{i} with its image in SH(G0)SH(H)SH^{*}(G_{0})\cong SH^{*}(H),

Xi=η(Yi).X_{i}=\eta(Y_{i}).

Write YiY_{i} in the fixed eigenspace decomposition:

Yi=Yi+Yi>,Y_{i}=Y_{i}^{\leq}+Y_{i}^{>},

where

Yiev(λ)kπR2HF(H0)andYi>ev(λ)>kπR2HF(H0)Y_{i}^{\leq}\in\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}(H_{0})\hskip 28.45274pt\text{and}\hskip 28.45274ptY_{i}^{>}\in\bigoplus_{ev(\lambda)>k\pi R^{2}}HF^{*}(H_{0})

Choose Ziev(λ)kπR2HF(H0)Z_{i}\in\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}(H_{0}) such that

(𝒄𝟏𝑮𝓢)n(Zi)=Yi.(\bm{c_{-1}^{G}\circ{\cal S}})^{n}(Z_{i})=Y_{i}^{\leq}.

By Corollary 7, there exists a constant CC such that

𝒜(Xi)𝓐(Zi)<C.{\cal A}(X_{i})-\bm{{\cal A}}(Z_{i})<C.

so that

(57) 𝓐(Zi)>𝒜(Xi)C.\bm{{\cal A}}(Z_{i})>{\cal A}(X_{i})-C.

Let

Z=i=1Zi.Z=\sum_{i=1}^{\infty}Z_{i}.

Equation (57), together with the assumption that 𝒜(Xi){\cal A}(X_{i})\rightarrow\infty, shows that ZZ is a well-defined element of HF(H0)HF^{*}(H_{0}).

Recall from Lemma 13 that

Φ(Y>)=0.\Phi(Y^{>})=0.

Thus,

ϕ(Xi)=ϕη(Yi)=π(Yi)=π(Yi)+π(Y>)=π(Yi)=π(Zi).\phi(X_{i})=\phi\circ\eta(Y_{i})=\pi(Y_{i})=\pi(Y_{i}^{\leq})+\pi(Y^{>})=\pi(Y_{i}^{\leq})=\pi(Z_{i}).

By linearity of Φ\Phi and η\eta,

π(Z)=ϕ(X).\pi(Z)=\phi(X).

Thus,

im(π)=im(ϕ).\textnormal{im}(\pi)=\textnormal{im}(\phi).

As ϕ\phi is injective and π=ϕη\pi=\phi\circ\eta, we conclude that η\eta is surjective. It follows that there is a ring isomorphism

SH¯(H)SH(H)/ker(η)=SH(H)/ker(π).\overline{SH^{*}}(H)\cong{\raisebox{1.99997pt}{$SH^{*}(H)$}\left/\raisebox{-1.99997pt}{$\ker(\eta)$}\right.}={\raisebox{1.99997pt}{$SH^{*}(H)$}\left/\raisebox{-1.99997pt}{$\ker(\pi)$}\right.}.

By Corollary 8,

SH(H)/ker(π)HF(H0)/ev(λ)>kπR2HFλ(H0).{\raisebox{1.99997pt}{$SH^{*}(H)$}\left/\raisebox{-1.99997pt}{$\ker(\pi)$}\right.}\simeq{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}.

Finally, by Corollary 4 and the PSS identification HF(H0)QH(E)HF^{*}(H_{0})\simeq QH^{*}(E),

HF(H0)/ev(λ)>kπR2HFλ(H0)QH(E)/ev(λ)>kπR2QHλ(E).{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}\simeq{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}QH^{*}_{\lambda}(E)$}\right.}.

5 Toric line bundles

We assume in this section that BB is a toric symplectic manifold. In particular, we will use the fact that BB has a perfect Morse function f:Bf:B\rightarrow\mathbb{R}, and the critical points of ff all have even index. We denote the Morse index of a critical point zz of ff by μf(z)\mu_{f}(z). Let {Hn:E}\{H_{n}:E\rightarrow\mathbb{R}\} be a family of functions defined as in subsection 3.1.2, using the perfect Morse function ff. Let JJ be an almost-complex structure of the form found in Subsection 3.1.1.

Proposition 4

There is an isomorphism

SH^(H)SH¯(H).\widehat{SH^{*}}(H)\simeq\overline{SH^{*}}(H).
  • Remark 14)

    In the author’s thesis [venkatesh], we computed SH^(H)\widehat{SH^{*}}(H) on monotone toric line bundles. Proposition 4 shows that the results of this paper encompass that of [venkatesh].

To prepare for the proof of Proposition 4 we recall results in [albers-k] about the structure of each Floer complex CF(Hn)CF^{*}(H_{n}). It is useful to first consider a \mathbb{Z}-graded Floer theory. Let SS be a formal variable of degree τ\tau, where τ\tau is the minimal Chern number of EE. The following results were proved by Albers-Kang for weakly monotone symplectic line bundles, but the proofs carry over verbatim to the more general case.

Lemma 23

The chain complex SC(H;Λ[S])SC^{*}(H;\Lambda[S]) is \mathbb{Z} graded. A generator x𝒫(Hn)x\in{\cal P}(H_{n}) with capping x~\tilde{x} has even degree, respectively odd degree, if xx is the minimum, respectively maximum, of a perfect Morse function on a transversally non-degenerate family of orbits. If x~\tilde{x} is chosen to be the fiber disk, then

|x|=2𝔴(x)+μf(ρx)+{1,0},|x|=-2\mathfrak{w}(x)+\mu_{f}(\rho\circ x)+\{-1,0\},

where the final term is 1-1 if xx is a minimum and 0 if xx is a maximum.

  • Remark 15)

    The grading in Lemma 23 is the cohomological Conley-Zehnder, shifted up by mm. We shift the grading for ease of notation; the grading on CF(H0)CF^{*}(H_{0}) now coincides with the grading on singular cochains.

  • Remark 16)

    Lemma 23 separates symplectic cohomology into an odd-graded component, denoted by SCodd(H)SC^{odd}(H), which consists of all “minimums”, and and even-graded component, denoted by SCeven(H)SC^{even}(H), which consists of all “maximums”, plus the constant orbits.

Lemma 24 (Albers-Kang [albers-k])

An element cSαTβxcS^{\alpha}T^{\beta}x, with c𝕂c\in\mathbb{K} and x𝒫(Hn)x\in{\cal P}(H_{n}) descends to an element

ρ(cSαTβx):=cSα+kβTβx\rho_{*}(cS^{\alpha}T^{\beta}x):=cS^{\alpha+k\beta}T^{\beta}x

of the Floer cochain complex associated with the Floer data {f,j}\{f,j\}. This cochain complex is \mathbb{R}-graded. The projected generator has degree

|ρ(cSαTβx)|=2τα+2τkβ+μf(ρx).|\rho_{*}(cS^{\alpha}T^{\beta}x)|=2\tau\alpha+2\tau k\beta+\mu_{f}(\rho\circ x).

Let pp\in\mathbb{R} and define

p=(j=0J(i=0ciTβi)Sαj)x|x𝐪𝒫(H),{0,1},|ρ(SαiTβjx)|p{\cal F}_{p}=\left\langle\left(\sum_{j=0}^{J}\left(\sum_{i=0}^{\infty}c_{i}T^{\beta_{i}}\right)S^{\alpha_{j}}\right)x\hskip 2.84544pt\bigg{|}\hskip 2.84544ptx{\bf q}^{\ell}\in{\cal P}(H),\ell\in\{0,1\},|\rho_{*}(S^{\alpha_{i}}T^{\beta_{j}}x)|\geq p\right\rangle
Lemma 25 (Albers-Kang [albers-k])

A solution u(s,t)u(s,t) of the Floer equation associated with the Floer data {Hn,J}\{H_{n},J\} on EE projects to a solution ρu\rho_{*}u of the Floer equation associated with the Floer data {f,j}\{f,j\} on MM.

Corollary 10

The complexes p{\cal F}_{p} define a filtration on SC(H)SC^{*}(H).

Note that this filtration induces a filtration on the subcomplex SC(a,)(H)SC^{*}_{(a,\infty)}(H) and the quotient complex SCa(H)SC^{*}_{a}(H). The following two Lemmas, stated for the full complex SC(H)SC^{*}(H), also apply to these smaller complexes.

Lemma 26 (Albers-Kang [albers-k])

The differential Λ[S]\partial_{\Lambda[S]} on SC(H;Λ[S])SC^{*}(H;\Lambda[S]) decomposes as

Λ[S]=0+1+2+\partial_{\Lambda[S]}=\partial_{0}+\partial_{1}+\partial_{2}+...

where i\partial_{i} increases the grading of the projected generator by ii.

Lemma 27 (Albers-Kang [albers-k])

Floer solutions contributing to the zeroth part of the differential, 0\partial_{0}, remain in a single fiber. If xMx_{M} is the maximum and xmx_{m} is the minimum of a perfect Morse function on an S1S^{1}-family of orbits, then 0(xm)=xM\partial_{0}(x_{m})=x_{M} and 0(xM)=0\partial_{0}(x_{M})=0.

Consider Lemma 23. If x𝒫(Hn)x\in{\cal P}(H_{n}) and

(j=0J(i=0ciTβi)Sαj)xSC(Hn)\left(\sum_{j=0}^{J}\left(\sum_{i=0}^{\infty}c_{i}T^{\beta_{i}}\right)S^{\alpha_{j}}\right)x\in SC^{\ell}(H_{n})

lies in a fixed degree {\ell}, then

αj=12(+2𝔴(x)μf(ρx)+{1,0})\alpha_{j}=\frac{1}{2}\left({\ell}+2\mathfrak{w}(x)-\mu_{f}(\rho\circ x)+\{-1,0\}\right)

where 𝔴(x)\mathfrak{w}(x) is the winding number of xx in the fiber and the contribution in {1,0}\{-1,0\} depends only on xx. So αj\alpha_{j} is a fixed constant independent of jj, which we denote by αxk\alpha_{x}^{k}.

Lemma 28

Let i=0ciSαxiTβixi𝐪ni\sum_{i=0}^{\infty}c_{i}S^{\alpha_{x_{i}}^{\ell}}T^{\beta_{i}}x_{i}{\bf q}^{n_{i}} be an element in the cochain complex limaSCa(H)\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}SC^{\ell}_{a}(H) or SC(H)SC^{*}(H), which consists of formal sums whose action limits to infinity (in the latter case, the sum includes only finitely many distinct “xix_{i}”). Assume ci0c_{i}\neq 0 for any ii. Then

limi|ρ(ciSαxiTβixi)|=.\lim_{i\rightarrow\infty}|\rho_{*}(c_{i}S^{\alpha_{x_{i}}^{\ell}}T^{\beta_{i}}x_{i})|=\infty.
Proof.

By the definition of limaSCa(H)\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}SC^{*}_{a}(H) and SC(H)SC^{*}(H),

limi𝒜(Tβixi)=.\lim_{i\rightarrow\infty}{\cal A}(T^{\beta_{i}}x_{i})=\infty.

Because each Hamiltonian HnH_{n} is bounded, this implies that

(58) limiβiD2(xi~)Ω=.\lim_{i\rightarrow\infty}\beta_{i}-\int_{D^{2}}(\widetilde{x_{i}})^{*}\Omega=\infty.

By construction, the winding number 𝔴(xi)\mathfrak{w}(x_{i}) satisfies

(59) 𝔴(xi)0.\mathfrak{w}(x_{i})\geq 0.

Thus,

(60) D2xi~Ω=πR2𝔴(xi)0.\int_{D^{2}}-\widetilde{x_{i}}^{*}\Omega=-\pi R^{2}\mathfrak{w}(x_{i})\leq 0.

Equations 58 and 60 imply that

limiβi=,\lim_{i\rightarrow\infty}\beta_{i}=\infty,

and so

(61) limi2τkβi=,\lim_{i\rightarrow\infty}2\tau k\beta_{i}=\infty,

where k>0k>0 is the constant satisfying c1E=k[ω]c_{1}^{E}=-k[\omega]. The choice of grading implies

=2𝔴(xi)+2ταxi+C,\ell=-2\mathfrak{w}(x_{i})+2\tau\alpha_{x_{i}}^{\ell}+C,

where CC is a bounded constant in the interval [1,dim(M)][-1,\dim_{\mathbb{R}}(M)]. Equation 59 now implies that

(62) 2ταxiC.2\tau\alpha_{x_{i}}^{\ell}\geq\ell-C.

Finally, by Lemma 24,

|ciSαxiTβiρxi|=2ταxi+2τkβi+μf(ρxi).|c_{i}S^{\alpha_{x_{i}}^{\ell}}T^{\beta_{i}}\rho\circ x_{i}|=2\tau\alpha_{x_{i}}^{\ell}+2\tau k\beta_{i}+\mu_{f}(\rho\circ x_{i}).

This grows like 2ταxi+2τkβi2\tau\alpha_{x_{i}}^{\ell}+2\tau k\beta_{i}, which, by observations (62) and (61), limits to infinity.

Let Λ[S]\partial_{\Lambda[S]} be the differential on SC(H;Λ[S])SC^{*}(H;\Lambda[S]), and let Λ\partial_{\Lambda} be the differential on SC(H;Λ)SC^{*}(H;\Lambda). Let

(63) qτ:SC(H;Λ[S])SC(H;Λ)q_{\tau}:SC^{*}(H;\Lambda[S])\longrightarrow SC^{*}(H;\Lambda)

be the covering map that sends SS to 11. Let 𝐪τ{\bf q_{\tau}} be the induced map on homology.

Fix a degree \ell\in\mathbb{Z} and consider the Λ\Lambda-linear map pτp_{\tau}^{\ell} given on generating orbits by

pτ:SCmod2τ(H;Λ)\displaystyle p_{\tau}^{\ell}:SC^{\ell\mod 2\tau}(H;\Lambda) SC(H;Λ[S])\displaystyle\longrightarrow SC^{\ell}(H;\Lambda[S])
x\displaystyle x Sαxx,\displaystyle\mapsto S^{\alpha_{x}}x,

where αx\alpha_{x} is uniquely determined by the degree formula

2αx+|x|=.2\alpha_{x}+|x|=\ell.

On the level of vector spaces, pτp_{\tau}^{\ell} is inverse to qτ|SC(H;Λ[S])q_{\tau}\big{|}_{SC^{\ell}(H;\Lambda[S])}. On the level of cochains,

pτ+1Λ=Λ[S]pτ,p_{\tau}^{\ell+1}\circ\partial_{\Lambda}=\partial_{\Lambda[S]}\circ p_{\tau}^{\ell},

and pτp_{\tau}^{\ell} descends to a map on SHmod2τ(H;Λ)SH^{\ell\mod 2\tau}(H;\Lambda), inverse to

𝒒𝝉|SH(H;Λ[S]):SH(H;Λ[S])SHmod2τ(H;Λ).\bm{q_{\tau}}\big{|}_{SH^{\ell}(H;\Lambda[S])}:SH^{\ell}(H;\Lambda[S])\longrightarrow SH^{\ell\mod 2\tau}(H;\Lambda).

Thus, 𝒒𝝉|SH(H;Λ[S])\bm{q_{\tau}}\big{|}_{SH^{\ell}(H;\Lambda[S])} is an isomorphism, and so the full map 𝒒𝝉\bm{q_{\tau}} is surjective on cohomology.

qτq_{\tau} respects the action filtration, inducing a surjective map

(64) SHa(H;Λ[S])SHa(H;Λ).SH^{*}_{a}(H;\Lambda[S])\longrightarrow SH^{*}_{a}(H;\Lambda).

This map restricts to an isomorphism

SHa(H;Λ[S])SH mod 2τa(H;Λ),SH^{\ell}_{a}(H;\Lambda[S])\longrightarrow SH^{\ell\text{ mod }2\tau}_{a}(H;\Lambda),

which induces an isomorphism

SH^(H;Λ[S])SH mod 2τ^(H;Λ).\widehat{SH^{\ell}}(H;\Lambda[S])\longrightarrow\widehat{SH^{\ell\text{ mod }2\tau}}(H;\Lambda).

But inverse limits commute with finite direct sums, and so

SH^(H;Λ)=12τSH mod 2τ^(H;Λ).\widehat{SH^{*}}(H;\Lambda)\cong\bigoplus\limits_{\ell=1}^{2\tau}\widehat{SH^{\ell\text{ mod }2\tau}}(H;\Lambda).

We conclude that the induced map

(65) SH^(H;Λ[S])SH^(H;Λ)\widehat{SH^{*}}(H;\Lambda[S])\longrightarrow\widehat{SH^{*}}(H;\Lambda)

is surjective as well.

Lemma 29

If x𝒫(Hn)x\in{\cal P}(H_{n}) is an even-graded periodic orbit then Λ(x)=0\partial_{\Lambda}(x)=0.

Proof.

First take coefficients in Λ[S]\Lambda[S]. By index conventions, xx is a maximum of a perfect Morse function on an S1S^{1}-family of orbits. Suppose for contradiction that Λ[S](x)=Y0\partial_{\Lambda[S]}(x)=Y\neq 0. By Lemma 28, we can write Y=i=p0YiY=\sum_{i=p_{0}}^{\infty}Y_{i}, where |ρYi|=i|\rho_{*}Y_{i}|=i and Yp00Y_{p_{0}}\neq 0. By definition, Λ[S]2(x)=0\partial_{\Lambda[S]}^{2}(x)=0, implying that Λ[S](Y)=0\partial_{\Lambda[S]}(Y)=0. Decompose Λ[S](Y)\partial_{\Lambda[S]}(Y) with respect to the filtration:

Λ[S](Y)=i=p0Λ[S](Y)i,\partial_{\Lambda[S]}(Y)=\sum_{i=p_{0}}^{\infty}\partial_{\Lambda[S]}(Y)_{i},

where |ρΛ[S](Y)i|=i|\rho_{*}\partial_{\Lambda[S]}(Y)_{i}|=i. Then Λ[S](Y)=0\partial_{\Lambda[S]}(Y)=0 if and only if Λ[S](Y)i=0\partial_{\Lambda[S]}(Y)_{i}=0 for every ii. In particular,

0(Yp0)=Λ[S](Y)p0=0.\partial_{0}(Y_{p_{0}})=\partial_{\Lambda[S]}(Y)_{p_{0}}=0.

However, by index considerations, each generator appearing in Yp0Y_{p_{0}} is a minimum of a perfect Morse function on an S1S^{1}-family of orbits. By Lemma 27,

0(Yp0)0.\partial_{0}(Y_{p_{0}})\neq 0.

A contradiction is reached.

Now take coefficients in Λ\Lambda. Recall the map qτq_{\tau} from Equation (63), which takes x(t)x(t) to itself. As qτq_{\tau} is a chain map,

Λ(x)=Λqτ(x)=qτΛ[S](x)=0.\partial_{\Lambda}(x)=\partial_{\Lambda}\circ q_{\tau}(x)=q_{\tau}\circ\partial_{\Lambda[S]}(x)=0.

Corollary 11

The connecting maps

πa,a:SHevena(H)SHevena(H)\pi_{a,a^{\prime}}:SH^{even}_{a}(H)\rightarrow SH^{even}_{a^{\prime}}(H)

are surjective.

Lemma 30

The kernel of \partial is 0 on the odd-component:

ker()|SCodd(H)=ker(^)|SCodd^(H;Λ)=0.\ker(\partial)\big{|}_{SC^{odd}(H)}=\ker(\widehat{\partial})\big{|}_{\widehat{SC^{odd}}(H;\Lambda)}=0.

As a consequence,

SHodd(H)=SHodd^(H)=0.SH^{odd}(H)=\widehat{SH^{odd}}(H)=0.
Proof.

First consider the uncompleted theory SHodd(H;Λ)SH^{odd}(H;\Lambda). The choices made ensure that

CFodd(H0;Λ)=0,CF^{odd}(H_{0};\Lambda)=0,

from which it is clear that HFodd(H0;Λ)=0HF^{odd}(H_{0};\Lambda)=0. Corollary 3 expresses an isomorphism

SH(H;Λ)HF(H0;Λ)/HF0(H0;Λ).SH^{*}(H;\Lambda)\cong{\raisebox{1.99997pt}{$HF^{*}(H_{0};\Lambda)$}\left/\raisebox{-1.99997pt}{$HF^{*}_{0}(H_{0};\Lambda)$}\right.}.

This isomorphism preserves the parity of the grading, and it follows that

SHodd(H;Λ)=HFodd(H0;Λ)/HFodd0(H0;Λ)=0.SH^{odd}(H;\Lambda)={\raisebox{1.99997pt}{$HF^{odd}(H_{0};\Lambda)$}\left/\raisebox{-1.99997pt}{$HF^{odd}_{0}(H_{0};\Lambda)$}\right.}=0.

It remains to show that

SHodd^(H;Λ)=0.\widehat{SH^{odd}}(H;\Lambda)=0.

First consider coefficients in Λ[S]\Lambda[S]. Choose a non-zero sequence {XaSCka(H;Λ[S])}\{X_{a}\in SC^{k}_{a}(H;\Lambda[S])\}, and assume for contradiction that each XaX_{a} is a cocycle and the connecting morphisms πa,a\pi_{a,a^{\prime}} send 𝐗𝐚{\bf X_{a}} to 𝐗𝐚{\bf X_{a^{\prime}}}, so that πa,a(𝐗𝐚)=𝐗𝐚{\pi_{a,a^{\prime}}}({\bf X_{a}})={\bf X_{a^{\prime}}}. By definition, there exists a coboundary (Z)\partial(Z) such that, at the chain level,

πa,a(Xa)=Xa+(Z).\pi_{a,a^{\prime}}(X_{a})=X_{a^{\prime}}+\partial(Z).

The proof of Lemma 29 implies that (Z)=0\partial(Z)=0, so

(66) πa,a(Xa)=Xa.\pi_{a,a^{\prime}}(X_{a})=X_{a^{\prime}}.

The action of the summands of the nested sequence {Xa}\{X_{a}\} approaches infinity. By Lemma 28, there exists some p0p_{0} such that Xap0X_{a}\in{\cal F}_{p_{0}} for all aa. Assume we have chosen the maximum such p0p_{0}, so that, for large enough aa, some non-zero summand Xp0,aX_{p_{0},a} of XaX_{a} achieves |ρXp0,a|=p0|\rho_{*}X_{p_{0},a}|=p_{0}. Without loss of generality, assume that we have fixed a large enough aa.

Let Ya=Λ[S](Xa)Y_{a}=\partial_{\Lambda[S]}(X_{a}). Write Xa=p=p0Xp,aX_{a}=\sum_{p=p_{0}}^{\infty}X_{p,a}, where |ρXp,a|=p|\rho_{*}X_{p,a}|=p, and similarly write Ya=i=p0Yp,aY_{a}=\sum_{i=p_{0}}^{\infty}Y_{p,a}. We want to show that Ya0Y_{a}\neq 0. It suffices to show that there exists a single non-zero summand Yp,aY_{p,a}.

The decomposition of the differential described in Lemma 26 yields

Yp,a=0(Xp,a)+1(Xp1,a)++pp0(Xp0,a).Y_{p,a}=\partial_{0}(X_{p,a})+\partial_{1}(X_{p-1,a})+...+\partial_{p-p_{0}}(X_{p_{0},a}).

In particular,

(0)Λ[S](Xp0,a)=Yp0,a.(\partial_{0})_{\Lambda[S]}(X_{p_{0},a})=Y_{p_{0},a}.

Assume a>𝒜((0)Λ[S](Xp0,a)a>{\cal A}((\partial_{0})_{\Lambda[S]}(X_{p_{0},a}).

Due to the grading conventions, Xp0,aX_{p_{0},a} is the minimum of a perfect Morse function on the underlying trajectory. Therefore, by Lemma 27, 0(Xp0,a)0\partial_{0}(X_{p_{0}},a)\neq 0. It follows that

Yp0,a0,Y_{p_{0},a}\neq 0,

and so

Ya=Λ[S](Xa)0.Y_{a}=\partial_{\Lambda[S]}(X_{a})\neq 0.

Thus, Λ[S](Xa)0\partial_{\Lambda[S]}(X_{a})\neq 0. As this holds for all sufficiently large aa, we conclude that ker(^)=0\ker(\hat{\partial})=0 and

SHodd^(H;Λ[S])=0.\widehat{SH^{odd}}(H;\Lambda[S])=0.

Combining the surjectivity of the map (65) with Lemma 29, it follows that

ker(^|SCodd^(H;Λ))=0\ker(\hat{\partial}\big{|}_{\widehat{SC^{odd}}(H;\Lambda)})=0

as well.

We end this subsection by proving Proposition 4.

Proof.

We have seen that there is an injective map

ϕ:SH¯(H)SH^(H)\phi:\overline{SH^{*}}(H)\rightarrow\widehat{SH^{*}}(H)

induced by the inclusion

ker()^SC^(H).\widehat{\ker(\partial)}\hookrightarrow\widehat{SC^{*}}(H).

Lemma 30 shows that SHodd^(H)=0\widehat{SH^{odd}}(H)=0, which implies that

SHodd¯(H)=0\overline{SH^{odd}}(H)=0

as well. By Lemma 29,

ker()=SCeven(H),\ker(\partial)=SC^{even}(H),

and so

ker(^)=SCeven^(H).\widehat{\ker(\partial})=\widehat{SC^{even}}(H).

It follows that the map ϕ\phi is also surjective.

5.1 Annulus subbundles

Recall that we defined the completed symplectic cohomology theory for trivial cobordisms in Subsection 2.4. In this section, we consider trivial cobordisms AR,RA_{R^{\prime},R} that lie between disk bundles of raddi RR^{\prime} and RR. We would like to know for which radii completed symplectic cohomology is non-vanishing.

Let QH(E)QH_{*}(E) be quantum homology, the dual of quantum cohomology. The dual of the map ρc1E\rho^{*}c_{1}^{E}\cup_{*}- is the quantum intersection product with the Poincaré dual of ρc1E\rho^{*}c_{1}^{E}, denoted by PD(ρc1E)PD(\rho^{*}c_{1}^{E})\cap_{*}-. These maps have the same eigenvalues, and we denote by QHλ(E)QH_{*}^{\lambda}(E) the λ\lambda-invariant subspace of QH(E)QH_{*}(E) under PD(ρc1E)PD(\rho^{*}c_{1}^{E})\cap_{*}-.

Theorem 8

If RRR^{\prime}\leq R, there is a vector-space isomorphism

SH^(AR,R)QH(E)/kπ(R)2ev(λ)orkπR2<ev(λ)QHλ(E).\widehat{SH^{*}}(A_{R^{\prime},R})\cong{\raisebox{1.99997pt}{$QH^{*}(E)$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{\begin{subarray}{c}k\pi(R^{\prime})^{2}\geq ev(\lambda)\\ \text{or}\\ k\pi R^{2}<ev(\lambda)\end{subarray}}QH^{*}_{\lambda}(E)$}\right.}.

If RRR^{\prime}\geq R, there is a vector-space isomorphism

SH^(AR,R)kπR2<ev(λ)kπ(R)2QH1λ(E).\widehat{SH^{*}}(A_{R^{\prime},R})\cong\bigoplus_{k\pi R^{2}<ev(\lambda)\leq k\pi(R^{\prime})^{2}}QH_{*-1}^{\lambda}(E).
  • Remark 17)

    In particular,

    SH^(AR,R)0\widehat{SH^{*}}(A_{R^{\prime},R})\neq 0

    precisely when the size of some eigenvalue lies between RR^{\prime} and RR.

By construction, there is a long exact sequence

(67) H(SC^(H)){H\left(\widehat{SC_{*}}(H^{\prime})\right)}H(SC^(H)){H\left(\widehat{SC^{*}}(H)\right)}SH^(AR,R){\widehat{SH^{*}}(A_{R^{\prime},R})}𝖈^\scriptstyle{\widehat{\bm{\mathfrak{c}}}}[1]\scriptstyle{[1]}

To compute SH^(AR,R)\widehat{SH^{*}}(A_{R^{\prime},R}) it therefore suffices to compute H(SC^(H))H\left(\widehat{SC_{*}}(H^{\prime})\right), H(SC^(H))H\left(\widehat{SC^{*}}(H)\right), and the connecting map 𝔠^\widehat{\mathfrak{c}}. As noted in Remark Remark 9), these new homology theories are not a priori the completed homology theories introduced thus far. However, in this simplified scenario, the definitions coincide.

Lemma 31

There are natural isomorphisms

H(SC^(H))SH^(H)H(\widehat{SC^{*}}(H))\cong\widehat{SH^{*}}(H)

and

H(SC^(H))SH^(H).H(\widehat{SC_{*}}(H))\cong\widehat{SH_{*}}(H).
Proof.

Applying the proof of Lemma 30 to a prospective cocycle in SCodd^(H)\widehat{SC^{odd}}(H) yields

Hodd(SC^(H))=0.H^{odd}\left(\widehat{SC^{*}}(H)\right)=0.

There is a Milnor exact sequence

0lima1SHevena(H)Heven(SC^(H))SHeven^(H)0.0\longrightarrow\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}SH^{even}_{a}(H)\longrightarrow H^{even}\left(\widehat{SC^{*}}(H)\right)\longrightarrow\widehat{SH^{even}}(H)\longrightarrow 0.

Corollary 11 says that the connecting maps

πa,a:SHevena(H)SHevena(H),a>a,{\pi_{a,a^{\prime}}}:SH^{even}_{a}(H)\longrightarrow SH^{even}_{a^{\prime}}(H),\hskip 28.45274pta>a^{\prime},

are surjective. Thus, the Mittag-Leffler condition is satisfied and

lima1SHevena(H)=0.\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}}\limits{}^{1}SH^{even}_{a}(H)=0.

The first isomorphism follows.

The second isomorphism follows from the exactness of direct limits.

We begin by computing SH(H)SH_{*}(H) and SH^(H)\widehat{SH_{*}}(H). The following Lemma is standard fare, but sets the stage for subsequent elaboration.

Lemma 32

Up to grading, SH(H)SH_{*}(H) is the dual homology theory to SH(H)SH^{*}(H).

Proof.

A standard Morse theory argument shows that the Floer chain complex (CF(Hi),fl)(CF_{*}(H_{i}),\partial_{fl}) is naturally isomorphic to the dual of the Floer cochain complex (CF(Hi),fl)(CF^{*}(H_{i}),\partial^{fl}), where fl\partial_{fl} counts rigid Floer cylinders with negative input and positive output. By definition, a generator xx of CF(H0)CF^{*}(H_{0}) corresponds to a generator xˇ\check{x} of CF(H0)CF_{*}(H_{0}). xˇ\check{x} is the morphism 𝟏x\mathbf{1}_{x} which on generators yy evaluates to

(68) 𝟏x(y)={1x=y0else\mathbf{1}_{x}(y)=\left\{\begin{array}[]{cc}1&x=y\\ 0&\text{else}\end{array}\right.

As direct products are dual to direct sums,

Hom(iCF(Hi)[𝐪],Λ)iHom(CF(Hi),Λ)[𝐪]iCF(Hi)[𝐪],Hom(\bigoplus_{i}CF^{*}(H_{i})[{\bf q}],\Lambda)\cong\prod_{i}Hom(CF^{*}(H_{i}),\Lambda)[{\bf q}]\cong\prod_{i}CF_{*}(H_{i})[{\bf q}],

with differential

(X+Y𝐪)=fl(X)+c(Y)+fl(Y).\partial(X+Y{\bf q})=\partial_{fl}(X)+c_{*}(Y)+\partial_{fl}(Y).

Here, cc_{*} is defined similarly to cc, but, as with fl\partial_{fl}, it counts the rigid Floer cylinders with negative input and positive output.

Poincaré duality in Floer theory yields natural isomorphisms

(CF(Hi),fl)(CF2m(Hi),fl).(CF_{*}(H_{i}),\partial_{fl})\cong(CF^{2m-*}(-H_{i}),\partial^{fl}).

See [c-f-o] for an elaboration on the grading, recalling that we have shifted their grading scheme by mm. These isomorphisms intertwine the continuation maps cc_{*} with the original continuation maps cc. Thus, there is a natural chain isomorphism

iCF(Hi)[𝐪]iCF2m(Hi)[𝐪].\prod_{i}CF_{*}(H_{i})[{\bf q}]\cong\prod_{i}CF^{2m-*}(-H_{i})[{\bf q}].

Just as 𝒄𝓢\bm{c\circ{\cal S}} acts on HF(H0)HF^{*}(H_{0}), the composition 𝓢𝒄\bm{{\cal S}\circ c} acts on HF(H0)HF^{*}(-H_{0}). The bijection

𝓢:HF(H0)HF(H0)\bm{{\cal S}}:HF^{*}(-H_{0})\longrightarrow HF^{*}(H_{0})

intertwines 𝒄𝓢\bm{c\circ{\cal S}} and 𝓢𝒄\bm{{\cal S}\circ c}, and it is easy to see that these two compositions have the same eigenvalues. We denote by HFλ(H0)HF^{*}_{\lambda}(-H_{0}) the λ\lambda-generalized eigenspace of the action of 𝓢𝒄\bm{{\cal S}\circ c} on HFλ(H0)HF^{*}_{\lambda}(-H_{0}).

Lemma 33

The isomorphism

SH(H)HF(H)/HF0(H)SH^{*}(H)\cong{\raisebox{1.99997pt}{$HF^{*}(H)$}\left/\raisebox{-1.99997pt}{$HF^{*}_{0}(H)$}\right.}

induces an isomorphism

SH(H)ev(λ)<HFλ(H0)SH_{*}(H)\cong\bigoplus_{ev(\lambda)<\infty}HF^{*}_{\lambda}(-H_{0})

such that the map Ψ:SH(H)HF(H0)\Psi:SH_{*}(H)\longrightarrow HF^{*}(-H_{0}) is the canonical inclusion.

Proof.

The dual of the inclusion

CF(H0)SC(H)CF^{*}(H_{0})\hookrightarrow SC^{*}(H)

is the projection

iCF(Hi)[𝐪]CF(H0).\prod_{i}CF_{*}(H_{i})[{\bf q}]\twoheadrightarrow CF_{*}(H_{0}).

Under Poincaré duality, this projection is a map

SC2m(H)CF2m(H0),SC_{2m-*}(H)\longrightarrow CF^{2m-*}(-H_{0}),

which descends to the map

(69) Ψ:SH2m(H)HF2m(H0).\Psi:SH_{2m-*}(H)\longrightarrow HF^{2m-*}(-H_{0}).

As Λ\Lambda is torsion free, SH(H;Λ)SH^{*}(H;\Lambda) is an injective module, and the Universal Coefficient Theorem yields a natural isomorphism

SH(H)Hom(SH(H),Λ).SH_{*}(H)\cong Hom(SH^{*}(H),\Lambda).

In particular, denoting by ι\iota^{\vee} the dual of the projection

ι:HF(H0)SH(H),\iota:HF^{*}(H_{0})\longrightarrow SH^{*}(H),

the following diagram commutes

SH2m(H){SH_{2m-*}(H)}Hom(SH(H),Λ){Hom(SH^{*}(H),\Lambda)}HF2m(H0){HF^{2m-*}(-H_{0})}Hom(HF(H0),Λ).{Hom(HF^{*}(H_{0}),\Lambda).}\scriptstyle{\simeq}Ψ\scriptstyle{\Psi}ι\scriptstyle{\iota^{\vee}}\scriptstyle{\simeq}

Identifying SH(H)SH^{*}(H) with the quotient complex

HF(H0)/HF0(H0)SH(H),{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$HF^{*}_{0}(H_{0})$}\right.}\cong SH^{*}(H),

the group SH2m(H)SH_{2m-*}(H) is precisely the subset of HF2m(H0)HF^{2m-*}(-H_{0}) on which HF0(H0)HF^{*}_{0}(H_{0}) vanishes, and Ψ\Psi is the canonical inclusion.

Let ={w1λ1,w2λ1,,wkλλk}{\cal B}^{\vee}=\left\{w_{1}^{\lambda_{1}},w_{2}^{\lambda_{1}},...,w_{k_{\lambda}}^{\lambda_{k}}\right\} be the basis of HF2m(H0)HF^{2m-*}(-H_{0}) dual to the fixed Jordan basis {\cal B}. By definition, {\cal B}^{\vee} is a generalized eigenbasis for the action on Hom(HF(H0),Λ)Hom(HF^{*}(H_{0}),\Lambda) dual to 𝒄𝓢\bm{c\circ{\cal S}}. A standard result in Floer theory shows that, under the identification HF(H0)Hom(HF(H0),Λ)HF_{*}(H_{0})\cong Hom(HF^{*}(H_{0}),\Lambda), the dual action is the map

𝓢1𝒄:HF(H0)HF(H0).\bm{{\cal S}}^{-1}\circ\bm{c_{*}}:HF_{*}(H_{0})\longrightarrow HF_{*}(H_{0}).

Under the Poincaré duality isomorphism HF(H0)HF2m(H0)HF_{*}(H_{0})\cong HF^{2m-*}(-H_{0}), the continuation maps 𝒄\bm{c} and 𝒄\bm{c_{*}} are identified, and 𝓢1𝒄\bm{{\cal S}}^{-1}\circ\bm{c_{*}} is identified with

𝓢𝒄:HF2m(H0)HF2m(H0).\bm{{\cal S}}\circ\bm{c}:HF^{2m-*}(-H_{0})\longrightarrow HF^{2m-*}(-H_{0}).

Thus, the dual basis {\cal B}^{\vee} is a generalized eigenbasis for the action of 𝓢𝒄\bm{{\cal S}}\circ\bm{c} on HF2m(H0)HF^{2m-*}(-H_{0}). By definition, HF0(H0)ker(wiλj)HF^{*}_{0}(H_{0})\subset\ker(w_{i}^{\lambda_{j}}) precisely when λj0\lambda_{j}\neq 0. Thus,

SH2m(H0)λ0HF2mλ(H0)=ev(λ)<HF2mλ(H0).SH_{2m-*}(H_{0})\cong\bigoplus_{\lambda\neq 0}HF^{2m-*}_{\lambda}(-H_{0})=\bigoplus_{ev(\lambda)<\infty}HF^{2m-*}_{\lambda}(-H_{0}).

Lemma 34

Up to grading, SH^(H)\widehat{SH_{*}}(H) is the dual homology theory to SH^(H)\widehat{SH^{*}}(H).

Proof.

Equip the uncompleted chain complex SC(H)SC^{*}(H) with the non- Archimedean metric given by

||X||=e𝒜(X).||X||=e^{-{\cal A}(X)}.

Completing SC(H)SC^{*}(H) with respect to ||||||\cdot|| yields a complex whose elements are formal sums

(70) {i=0κixi+diyi𝐪|xi,yi𝒫(H),𝒜(κixi),𝒜(diyi)}.\left\{\sum_{i=0}^{\infty}\kappa_{i}x_{i}+d_{i}y_{i}{\bf q}\hskip 2.84544pt\bigg{|}\hskip 2.84544ptx_{i},y_{i}\in{\cal P}(H),{\cal A}(\kappa_{i}x_{i})\rightarrow\infty,{\cal A}(d_{i}y_{i})\rightarrow\infty\right\}.

Because the connecting maps in the inverse limit lima\lim\limits_{\begin{subarray}{c}\leftarrow\\ a\end{subarray}} are surjections, SC(H)^\widehat{SC^{*}(H)} is isomorphic to the complex (70). The dual of SC^(H)\widehat{SC^{*}}(H) is bounded homomorphisms on SC(H)SC^{*}(H), that is,

HomΛ(SC^(H),Λ)=HombΛ(SC(H),Λ):={γHom(SC(H),Λ)|supXSC(H)||γ(X)||||X||<},Hom_{\Lambda}(\widehat{SC^{*}}(H),\Lambda)=Hom^{b}_{\Lambda}(SC^{*}(H),\Lambda):=\left\{\gamma\in Hom(SC^{*}(H),\Lambda)\hskip 5.69046pt\bigg{|}\hskip 5.69046pt\sup_{X\in SC^{*}(H)}\frac{||\gamma(X)||}{||X||}<\infty\right\},

equipped with the usual differential

(γ)=γ.\partial^{*}(\gamma)=\gamma\circ\partial.

We want to show that SC^(H)\widehat{SC_{*}}(H) is chain isomorphic to HombΛ(SC(H),Λ)Hom^{b}_{\Lambda}(SC^{*}(H),\Lambda).

Analogously to the statement (70) for cohomology, SC^(H)\widehat{SC_{*}}(H) is isomorphic to

(71) {i=0κixiˇ+diyiˇ𝐪|xiˇ,yiˇ𝒫(H),𝒜(κixiˇ),𝒜(diyiˇ)a for some fixed a independent of i}.\left\{\sum_{i=0}^{\infty}\kappa_{i}\check{x_{i}}+d_{i}\check{y_{i}}{\bf q}\hskip 2.84544pt\bigg{|}\hskip 2.84544pt\check{x_{i}},\check{y_{i}}\in{\cal P}(-H),{\cal A}(\kappa_{i}\check{x_{i}}),{\cal A}(d_{i}\check{y_{i}})\geq a\text{ for some fixed }a\in\mathbb{R}\text{ independent of }i\right\}.

Recall from the proof of Lemma 33 that Poincaré duality for Hamiltonian Floer theory is a chain isomorphism

(72) CF(Hi)CF2m(Hi)CF^{*}(-H_{i})\cong CF_{2m-*}(H_{i})

induced by the set bijection 𝒫(Hi)𝒫(Hi){\cal P}(-H_{i})\cong{\cal P}(H_{i}) via x(t)x(t)x(t)\mapsto x(-t). For each x𝒫(Hi)x\in{\cal P}(H_{i}), let xˇ\check{x} be the corresponding element of 𝒫(Hi){\cal P}(-H_{i}) under this bijection. Consider the map

(73) SC^(H)HombΛ(SC2m(H),Λ)\widehat{SC_{*}}(H)\longrightarrow Hom^{b}_{\Lambda}(SC^{2m-*}(H),\Lambda)

given on generators by

xiˇ𝟏xi,\check{x_{i}}\mapsto\mathbf{1}_{x_{i}},

where

𝟏xi(x)={1x=xi0xxi}\mathbf{1}_{x_{i}}(x)=\left\{\begin{array}[]{cc}1&x=x_{i}\\ 0&x\neq x_{i}\end{array}\right\}

To check that this map is well-defined, enumerate 𝒫(H)={x0,x1,x2,}{\cal P}(H)=\{x_{0},x_{1},x_{2},...\}, and write

Xˇ=i=0κixiˇ+dixiˇ𝐪SC^(H)\check{X}=\sum_{i=0}^{\infty}\kappa_{i}\check{x_{i}}+d_{i}\check{x_{i}}{\bf q}\in\widehat{SC_{*}}(H)

so that there exists aa\in\mathbb{R} with

𝒜(κixiˇ),𝒜(dixiˇ)a{\cal A}(\kappa_{i}\check{x_{i}}),{\cal A}(d_{i}\check{x_{i}})\geq a

for all ii. Then for all x𝒫(H)x\in{\cal P}(H),

ev(Xˇ(x))\displaystyle ev(\check{X}(x)) =ev(i=0κi𝟏xi(x)).\displaystyle=ev\left(\sum_{i=0}^{\infty}\kappa_{i}\mathbf{1}_{x_{i}}(x)\right).

If xxix\neq x_{i} for any ii, then Xˇ(x)=0\check{X}(x)=0 and

||Xˇ(x)||||x||=0.\displaystyle\frac{||\check{X}(x)||}{||x||}=0.

So suppose there exists ii such that x=xix=x_{i}. Then

ev(Xˇ(x))\displaystyle ev(\check{X}(x)) =ev(κi)\displaystyle=ev(\kappa_{i})
a𝒜(xiˇ)\displaystyle\geq a-{\cal A}(\check{x_{i}})
=a+𝒜(x).\displaystyle=a+{\cal A}(x).

where the last equality follows because the capping disk of xi(t)x_{i}(t) is the capping disk of xi(t)=xˇi(t)x_{i}(-t)=\check{x}_{i}(t) with reversed orientation. Thus,

||Xˇ(x)||||x||\displaystyle\frac{||\check{X}(x)||}{||x||} ea𝒜(x)e𝒜(x)\displaystyle\leq\frac{e^{-a-{\cal A}(x)}}{e^{-{\cal A}(x)}}
=ea.\displaystyle=e^{-a}.

Similarly,

𝒜(Xˇ(x𝐪))ea.{\cal A}(\check{X}(x{\bf q}))\leq e^{-a}.

Extending linearly, we find that Xˇ\check{X} is indeed a bounded operator.

To see that the map (73) is surjective, suppose γHombΛ(SC(H),Λ)\gamma\in Hom^{b}_{\Lambda}(SC^{*}(H),\Lambda), so that

eev(γ(X))+𝒜(X):=||γ(X)||||X||M.e^{-ev(\gamma(X))+{\cal A}(X)}:=\frac{||\gamma(X)||}{||X||}\leq M.

for some fixed M>0M>0. Then

ev(γ(X))𝒜(X)log(M).ev(\gamma(X))\geq{\cal A}(X)-\log(M).

Let

Xˇ=x𝒫(H)γ(x)xˇ+γ(x𝐪)xˇ𝐪\check{X}=\sum_{x\in{\cal P}(H)}\gamma(x)\check{x}+\gamma(x{\bf q})\check{x}{\bf q}

Then

𝒜(γ(x)xˇ)\displaystyle{\cal A}(\gamma(x)\check{x}) 𝒜(x)log(M)+𝒜(xˇ)=𝒜(x)log(M)𝒜(x)=log(M).\displaystyle\geq{\cal A}(x)-\log(M)+{\cal A}(\check{x})={\cal A}(x)-\log(M)-{\cal A}(x)=-\log(M).

Similarly,

𝒜(γ(x𝐪)xˇ)log(M),{\cal A}(\gamma(x{\bf q})\check{x})\geq-\log(M),

so XˇSC^(M).\check{X}\in\widehat{SC_{*}}(M).

The bijection (73) intertwines the differential by the Poincaré duality isomorphism (72) and a standard Morse theory argument.

Finally, Lemma 31 completes the isomorphism:

SH^(H)H(SC^(H))H(HombΛ(SC2m(H),Λ))H(HomΛ(SC2m^(H),Λ)).\widehat{SH_{*}}(H)\cong H\left(\widehat{SC_{*}}(H)\right)\cong H\left(Hom^{b}_{\Lambda}(SC^{2m-*}(H),\Lambda)\right)\cong H\left(Hom_{\Lambda}(\widehat{SC^{2m-*}}(H),\Lambda)\right).

Analogously to the computation of SH(H)SH_{*}(H), we can compute action-completed symplectic homology SH^(H)\widehat{SH_{*}}(H) as a subspace of HF(H0)HF^{*}(-H_{0}).

Proposition 5

There are isomorphisms

SH^(H)ev(λ)kπR2HFλ(H0).\widehat{SH_{*}}(H)\cong\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(-H_{0}).

and

SH(H)ev(λ)<HFλ(H0)SH_{*}(H)\cong\bigoplus_{ev(\lambda)<\infty}HF^{*}_{\lambda}(-H_{0})

under which the map

SH^(H)HF(H0)\widehat{SH_{*}}(H)\longrightarrow HF^{*}(-H_{0})

is the canonical inclusion.

Proof.

We showed the second isomorphism in Lemma 33. The first isomorphism is proved in an entirely analogous manner. For completion, we sketch this argument again.

By Lemma 34, there is an isomorphism

SH^(H)=H(HomΛ(SC2m^(H),Λ)).\widehat{SH_{*}}(H)=H\left(Hom_{\Lambda}(\widehat{SC^{2m-*}}(H),\Lambda)\right).

The Universal Coefficient Theorem gives an isomorphism

H(HomΛ(SC2m^(H),Λ))HomΛ(SH2m^(H),Λ).H\left(Hom_{\Lambda}(\widehat{SC^{2m-*}}(H),\Lambda)\right)\cong Hom_{\Lambda}\left(\widehat{SH^{2m-*}}(H),\Lambda\right).

The map

SH^(H)SH(H)\widehat{SH_{*}}(H)\longrightarrow SH_{*}(H)

is identified with the map dual to

π:SH(H)SH^(H),\pi:SH^{*}(H)\longrightarrow\widehat{SH^{*}}(H),

denoted by

π:Hom(SH2m(H),Λ)Hom(SH2m^(H),Λ).\pi^{\vee}:Hom(SH^{2m-*}(H),\Lambda)\longrightarrow Hom(\widehat{SH^{2m-*}}(H),\Lambda).

Under the identifications

SH2m^(H)HF2m(H0)/ev(λ)>kπR2HF2mλ(H0)\widehat{SH^{2m-*}}(H)\cong{\raisebox{1.99997pt}{$HF^{2m-*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{2m-*}_{\lambda}(H_{0})$}\right.}

and

SH2m(H)HF2m(H0)/HF2m0(H0),SH^{2m-*}(H)\cong{\raisebox{1.99997pt}{$HF^{2m-*}(H_{0})$}\left/\raisebox{-1.99997pt}{$HF^{2m-*}_{0}(H_{0})$}\right.},

π\pi^{\vee} is an inclusion

Vev(λ)<HFλ(H0),V\hookrightarrow\bigoplus_{ev(\lambda)<\infty}HF^{*}_{\lambda}(-H_{0}),

where VV is the subspace of HF(H0)HF^{*}(-H_{0}) on which ev(λ)>kπR2HF2mλ(H0)\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{2m-*}_{\lambda}(H_{0}) vanishes. As in the proof of Lemma 33, VV is precisely

ev(λ)kπR2HFλ(H0).\bigoplus_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(-H_{0}).

It remains to compute the connecting map

𝖈^:SH^(H)SH^(H).\bm{\widehat{\mathfrak{c}}}:\widehat{SH_{*}}(H)\longrightarrow\widehat{SH^{*}}(H).

We have seen that 𝖈^\bm{\widehat{\mathfrak{c}}} can be identified with the composition

ev(λ)kπ(R)2HFλ(H0)HF(H0)𝐜HF(H0)HF(H0)/ev(λ)>kπR2HFλ(H0).\bigoplus_{ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(-H_{0})\hookrightarrow HF^{*}(-H_{0})\xrightarrow{{\bf c}}HF^{*}(H_{0})\twoheadrightarrow{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}.

Note, trivially, that the map 𝒄\bm{c} is equivalent to the map 𝒄𝓢𝓢1\bm{c}\circ\bm{{\cal S}}\circ\bm{{\cal S}}^{-1}. The map 𝓢1\bm{{\cal S}}^{-1} maps each λ\lambda-generalized eigenspace isomorphically onto a λ\lambda-generalized eigenspace. The map 𝒄𝓢\bm{c}\circ\bm{{\cal S}} restricts to an isomorphism on ev(λ)kπR2HFλ(H0)\bigoplus\limits_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(-H_{0}), as this subspace contains no 0-generalized eigenvectors. Thus, the map 𝒄\bm{c} restricts to an isomorphism

ev(λ)kπR2HFλ(H0)ev(λ)kπR2HFλ(H0).\bigoplus\limits_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(-H_{0})\rightarrow\bigoplus\limits_{ev(\lambda)\leq k\pi R^{2}}HF^{*}_{\lambda}(H_{0}).

This shows

Lemma 35

The connecting map 𝖈^\widehat{\bm{\mathfrak{c}}} can be identified with the composition

ev(λ)kπ(R)2HFλ(H0)ev(λ)kπ(R)2HFλ(H0)HF(H0)HF(H0)/ev(λ)>kπR2HFλ(H0).\bigoplus_{ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(-H_{0})\xrightarrow{\simeq}\bigoplus_{ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(H_{0})\hookrightarrow HF^{*}(H_{0})\twoheadrightarrow{\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{ev(\lambda)>k\pi R^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}.

It therefore has image

im(𝖈)=ev(λ)kπ(R)2HFλ(H0)/kπR2<ev(λ)kπ(R)2HFλ(H0).\textnormal{im}(\bm{\mathfrak{c}})={\raisebox{1.99997pt}{$\bigoplus\limits_{ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{k\pi R^{2}<ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(H_{0})$}\right.}.

and kernel

ker(𝖈)=kπR2<ev(λ)kπ(R)2HFλ(H0).\ker(\bm{\mathfrak{c}})=\bigoplus_{k\pi R^{2}<ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(-H_{0}).

We are now ready to prove Theorem 8.

Proof of Theorem 5: If RRR^{\prime}\leq R, the map 𝖈^\widehat{\bm{\mathfrak{c}}} is injective by Lemma 35. The long exact sequence (67) and the isomorphism

H(SC^(H))SH^(H)H(\widehat{SC^{*}}(H))\cong\widehat{SH^{*}}(H)

of Lemma 31 show that

SH^(AR,R)SH^(H)/im(𝖈^).\widehat{SH^{*}}(A_{R^{\prime},R})\cong{\raisebox{1.99997pt}{$\widehat{SH^{*}}(H)$}\left/\raisebox{-1.99997pt}{$\textnormal{im}(\widehat{\bm{\mathfrak{c}}})$}\right.}.

By Lemma 35 the right-hand side is

HF(H0)/ev(λ)>kπR2orev(λ)kπ(R)2HFλ(H0){\raisebox{1.99997pt}{$HF^{*}(H_{0})$}\left/\raisebox{-1.99997pt}{$\bigoplus\limits_{\begin{subarray}{c}ev(\lambda)>k\pi R^{2}\\ \text{or}\\ ev(\lambda)\leq k\pi(R^{\prime})^{2}\end{subarray}}HF^{*}_{\lambda}(H_{0})$}\right.}

Finally, the isomorphism HF(H0)QH(E)HF^{*}(H_{0})\cong QH^{*}(E) of Corollary 7 and the isomorphisms HFλ(H0)QHλ(E)HF^{*}_{\lambda}(H_{0})\cong QH^{*}_{\lambda}(E) of Corollary 4 show the first statement in Theorem 8.

If R>RR^{\prime}>R, the map 𝖈^\widehat{\bm{\mathfrak{c}}} is surjective by Lemma 35. The long exact sequence (67) and the isomorphism

H(SC^(H))SH^(H)H(\widehat{SC_{*}}(H))\cong\widehat{SH_{*}}(H)

of Lemma 31 show that

SH^(AR,R)ker(𝖈^).\widehat{SH^{*}}(A_{R^{\prime},R})\cong\ker(\bm{\widehat{\mathfrak{c}}}).

By Lemma 35 the right-hand side is

kπR2<ev(λ)kπ(R)2HFλ(H0).\bigoplus\limits_{k\pi R^{2}<ev(\lambda)\leq k\pi(R^{\prime})^{2}}HF^{*}_{\lambda}(-H_{0}).

Dualizing the isomorphism

QH(E)HF(H0)QH^{*}(E)\xrightarrow{\simeq}HF^{*}(H_{0})

and applying Poincaré duality yields an isomorphism

HF(H0)QH(E).HF^{*}(-H_{0})\xrightarrow{\simeq}QH_{*}(E).

The first isomorphism intertwines ρc1E\rho^{*}c_{1}^{E}\cup_{*}- with the map 𝐜𝓢\mathbf{c}\circ\bm{{\cal S}}. The second isomorphism intertwines the dual maps: the quantum intersection product PD(ρc1E)PD(\rho^{*}c_{1}^{E})\cap_{*}- and 𝓢c\bm{{\cal S}}\circ c. The result follows.

Theorem 8 explicates a form of self-duality:

Corollary 12

The groups SH^(AR,R)\widehat{SH^{*}}(A_{R^{\prime},R}) and SH^(AR,R)\widehat{SH^{*}}(A_{R,R^{\prime}}) are dual.

  • Remark 18)

    In [albers-k] Albers-Kang studied the line bundle associated to the prequantization bundle over a monotone base. They showed that the Rabinowitz Floer homology of a circle subbundle of radius RR vanishes whenever R<1πκR<\frac{1}{\sqrt{\pi\kappa}}. Their methods, in conjunction with the work of this section, show that, for R<RR^{\prime}<R and EE a monotone line bundle of negativity constant kk,

    SH^(AR,R){SH(H)R<1πkκR0else.\widehat{SH^{*}}(A_{R^{\prime},R})\simeq\left\{\begin{array}[]{cc}SH^{*}(H)&R^{\prime}<\frac{1}{\sqrt{\pi k\kappa}}\leq R\\ 0&\text{else}\end{array}\right..

    See also [venkatesh-thesis] for an application of the methods in [albers-k] to the case of a toric base.

5.2 Closed string mirror symmetry

The line bundle EE inherits the structure of a toric variety from the base BB and the \mathbb{C}^{*}-action on the fibers. Its moment polytope ΔE\Delta_{E} can be described in terms of the moment polytope ΔB\Delta_{B} of BB (see Subsection 7.6 in [ritter-gromov] or Subsection 12.5 in [ritter-s]).

  • Example 1)

    Let EE be the complex line bundle 𝒪(k)\xlongrightarrowρPm{\cal O}(-k)\xlongrightarrow{\rho}\mathbb{C}P^{m}. EE is a toric variety whose image under the moment map is

    ΔE:={(v1,,vm+1)m+1|vi0i{1,,m+1};v1vm+kvm+11}\Delta_{E}:=\left\{(v_{1},\dots,v_{m+1})\in\mathbb{R}^{m+1}\hskip 2.84544pt\bigg{|}\hskip 2.84544ptv_{i}\geq 0\hskip 1.42271pt\forall\hskip 1.42271pti\in\{1,\dots,m+1\};\hskip 2.84544pt-v_{1}-\dots-v_{m}+kv_{m+1}\geq-1\right\}

    The facet of ΔE\Delta_{E} lying in the m×{0}\mathbb{R}^{m}\times\{0\} plane is precisely ΔB\Delta_{B}.

EE has a conjectural Landau-Ginzberg mirror (E,W)(E^{\vee},W), where E:=ev1(Int(ΔE))E^{\vee}:=ev^{-1}(Int(\Delta_{E})), and

W:EΛW:E^{\vee}\longrightarrow\Lambda

is a superpotential that, to first order, is determined by the toric divisors. Closed-string mirror symmetry predicts an isomorphism between the symplectic cohomology of EE and the Jacobian of WW:

(74) SH(E;Λ)Λ[z1±,z2±,,zm+1±]/(z1W,z2W,,zm+1W)=:Jac(W).SH^{*}(E;\Lambda)\cong{\raisebox{1.99997pt}{$\Lambda[z_{1}^{\pm},z_{2}^{\pm},\dots,z_{m+1}^{\pm}]$}\left/\raisebox{-1.99997pt}{$(\partial_{z_{1}}W,\partial_{z_{2}}W,\dots,\partial_{z_{m+1}}W)$}\right.}=:Jac({W}).

Suppose that BB is monotone, so that c1TB=κ[ω]c_{1}^{TB}=\kappa[\omega], and suppose κ>k\kappa>k. Then EE is monotone as well, with monotonicity constant κk\kappa-k. Suppose further that the superpotential WW is Morse, with distinct critical values. In this case, computations in [ritter-fano] and [ritter-gromov] confirm Equation (74).

Open-string mirror symmetry was confirmed in [ritter-s]. They showed that the moment map μ:EΔE\mu:E\longrightarrow\Delta_{E} has a unique Lagrangian torus fiber LL that, when equipped with suitable choices of line bundles, split-generates the wrapped Fukaya category. LL sits inside the circle bundle of radius 1πk(κk)\frac{1}{\sqrt{\pi k(\kappa-k)}}. The fiber that is mirror to LL, defined by ev1μ(L)ev^{-1}\circ\mu(L), contains all critical points of WW. Open-string mirror symmetry matches each choice of line bundle with a critical point of WW.

  • Remark 19)

    With EE monotone, the Chern classes ρc1E=k[Ω]\rho^{*}c_{1}^{E}=-k[\Omega] and c1TE=(κk)[Ω]c_{1}^{TE}=(\kappa-k)[\Omega] are related by a constant. The quantum cup products by ρc1E\rho^{*}c_{1}^{E} and by c1TEc_{1}^{TE} therefore have the same eigenvalues, up to \mathbb{C}^{*}-scalar. Ritter showed in [ritter-fano] that all eigenvalues of c1TEc_{1}^{TE}\cup_{*}- have valuation in {0,1κk}\{0,\frac{1}{\kappa-k}\}. The Floer-essential Lagrangian LL therefore sits inside the circle bundle ΣR\Sigma_{R} whose radius satisfies

    kπR2=ev(λ)k\pi R^{2}=ev(\lambda)

    for some non-zero eigenvalue λ\lambda of ρc1E\rho^{*}c_{1}^{E}\cup_{*}-. This is precisely the critical radius where non-vanishing symplectic cohomology theories occur. In particular,

    SH^(AR1,R2)0LAR1,R2.\widehat{SH^{*}}(A_{R_{1},R_{2}})\neq 0\iff L\subset A_{R_{1},R_{2}}.

    The non-vanishing statement can be seen directly as a consequence of a closed-open map, as expounded upon in [venkatesh]. The vanishing statement, although it seems intimately related to a dearth of Floer-essential Lagrangians, does not seem to follow directly.

Closed-string mirror symmetry generalizes to domains of restricted size. Let AR1,R2A_{R_{1},R_{2}} be the annulus bundle between radii R1R_{1} and R2R_{2} in EE, with R1<R2R_{1}<R_{2}. The mirror of AR1,R2A_{R_{1},R_{2}} is

AR1,R2:={(z1,,zm+1)E|kπR12ev(zm+1)kπR22},A_{R_{1},R_{2}}^{\vee}:=\left\{(z_{1},\dots,z_{m+1})\in E^{\vee}\hskip 2.84544pt\bigg{|}\hskip 2.84544ptk\pi R_{1}^{2}\leq ev(z_{m+1})\leq k\pi R_{2}^{2}\right\},

equipped with W|AR1,R2{W}\big{|}_{A_{R_{1},R_{2}}^{\vee}}.

For I=(i1,,im+1)m+1I=(i_{1},\dots,i_{m+1})\in\mathbb{R}^{m+1} and 𝐳=(z1,,zm+1){\bf z}=(z_{1},\dots,z_{m+1}), denote (z1i1,,zm+1im+1)(z_{1}^{i_{1}},\dots,z_{m+1}^{i_{m+1}}) by 𝐳I{\bf z}^{I}. We denote the ring of functions on AR1,R2A_{R_{1},R_{2}}^{\vee} in the variable 𝐳{\bf z} by 𝒪(AR1,R2)𝐳{\cal O}(A_{R_{1},R_{2}}^{\vee})_{{\bf z}}, where

𝒪(AR1,R2)𝐳={i=0ci𝐳Ii|ciΛ;Iim+1;limiev(ci𝐳Ii)=𝐳AR1,R2}.{\cal O}(A_{R_{1},R_{2}}^{\vee})_{{\bf z}}=\left\{\sum_{i=0}^{\infty}c_{i}{\bf z}^{I_{i}}\hskip 2.84544pt\bigg{|}\hskip 2.84544ptc_{i}\in\Lambda;\hskip 2.84544ptI_{i}\in\mathbb{R}^{m+1};\hskip 2.84544pt\lim_{i\rightarrow\infty}ev(c_{i}{\bf z}^{I_{i}})=\infty\hskip 2.84544pt\forall\hskip 2.84544pt{\bf z}\in A_{R_{1},R_{2}}^{\vee}\right\}.

Denote by Z(ziW|A(R1,R2))Z\left(\partial_{z_{i}}W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right) the zeroes of the function ziW|A(R1,R2)\partial_{z_{i}}W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}. AR1,R2A_{R_{1},R_{2}}^{\vee} is an example of a Laurent domain: if ΔE\Delta_{E} is described by the functions

{v|v,n1(i)λ1,,v,ns(i)λs}\left\{v\hskip 5.69046pt\big{|}\hskip 5.69046pt\langle v,n_{1}^{(i)}\rangle\geq\lambda_{1},...,\langle v,n_{s}^{(i)}\rangle\geq\lambda_{s}\right\}

then AR1,R2A_{R_{1},R_{2}}^{\vee} is cut out by the inequalities

{||Tλ1i=1m+1zin1(i)||1,,||Tλsi=1m+1zins(i)||1,||TkπR12zm+1||1,TkπR22zm+1||1},\left\{||T^{-\lambda_{1}}\prod_{i=1}^{m+1}z_{i}^{n_{1}^{(i)}}||\geq 1,...,||T^{-\lambda_{s}}\prod_{i=1}^{m+1}z_{i}^{n_{s}^{(i)}}||\geq 1,||T^{-k\pi R_{1}^{2}}z_{m+1}||\geq 1,T^{-k\pi R_{2}^{2}}z_{m+1}||\leq 1\right\},

where

||z||=eev(z).||z||=e^{-ev(z)}.

Laurent domains are examples of affinoid domains, and therefore satisfy a Nullstellensatz [tian]. In particular,

Jac(W|A(R1,R2))=0if and only ifiZ(ziW|A(R1,R2))=.Jac\left(W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right)=0\hskip 14.22636pt\text{if and only if}\hskip 14.22636pt\bigcap_{i}Z\left(\partial_{z_{i}}W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right)=\mathop{\varnothing}.

This occurs if and only if Crit(W)AR1,R2=Crit(W)\cap A_{R_{1},R_{2}}^{\vee}=\mathop{\varnothing}, that is, if and only if

R2<1kπR2(kκ)orR1>1kπR2(kκ).R_{2}<\frac{1}{k\pi R^{2}(k-\kappa)}\hskip 14.22636pt\text{or}\hskip 14.22636ptR_{1}>\frac{1}{k\pi R^{2}(k-\kappa)}.

Conversely, since WW is Morse by assumption, if Crit(W)AR1,R2Crit(W)\subset A_{R_{1},R_{2}}^{\vee} then

Jac(W|A(R1,R2))𝒪(Crit(W))Jac(W).Jac\left(W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right)\simeq{\cal O}\left(Crit(W)\right)\simeq Jac(W).

Altogether,

Jac(W|A(R1,R2)){Jac(W)kπR12<1κk<kπR220elseJac\left(W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right)\simeq\left\{\begin{array}[]{cc}Jac(W)&k\pi R_{1}^{2}<\frac{1}{\kappa-k}<k\pi R_{2}^{2}\\ 0&\text{else}\end{array}\right.

But, as discussed in Remark Remark 19), 1κk\frac{1}{\kappa-k} is exactly the valuation of the non-zero eigenvalues of c1Ec_{1}^{E}\cup_{*}-. Thus, Theorem 8 says that

SH^(AR1,R2){SH(E)kπR12<1κk<kπR220else\widehat{SH^{*}}(A_{R_{1},R_{2}})\simeq\left\{\begin{array}[]{cc}SH^{*}(E)&k\pi R_{1}^{2}<\frac{1}{\kappa-k}<k\pi R_{2}^{2}\\ 0&\text{else}\end{array}\right.

Via the isomorphism SH(E)Jac(W)SH^{*}(E)\cong Jac(W) of Equation (74), we conclude a closed-string mirror symmetry statement for subdomains:

Closed-string mirror symmetry If R1<R2R_{1}<R_{2}, then

SH^(AR1,R2)Jac(W|A(R1,R2)).\widehat{SH^{*}}(A_{R_{1},R_{2}})\simeq Jac\left(W\big{|}_{A_{(R_{1},R_{2})^{\vee}}}\right).
  • Example 2)

    Again let E=𝒪(k)PmE={\cal O}(-k)\longrightarrow\mathbb{C}P^{m}. The mirror of EE is

    E:={(z1,,zm+1)(Λ)m+1|(ev(z1),,ev(zm+1))Δo},E^{\vee}:=\left\{(z_{1},\dots,z_{m+1})\in(\Lambda^{*})^{m+1}\hskip 2.84544pt\big{|}\hskip 2.84544pt\left(ev(z_{1}),\dots,ev(z_{m+1})\right)\in\Delta^{\mathrm{o}}\right\},

    equipped with superpotential

    (75) W:E\displaystyle W\colon E^{\vee} Λ\displaystyle\longrightarrow\Lambda
    (76) (z1,z2,,zm+1)\displaystyle(z_{1},z_{2},\dots,z_{m+1}) z1+z2++zm+zm+1+Tz11z21zm1zm+1k.\displaystyle\mapsto z_{1}+z_{2}+\dots+z_{m}+z_{m+1}+Tz_{1}^{-1}z_{2}^{-1}\dots z_{m}^{-1}z_{m+1}^{k}.

    (See Example 7.12 in [ritter-gromov] or Proposition 4.2 in [auroux].) Denote by AA the one-dimensional annulus kπR12ev(zm+1)kπR22k\pi R_{1}^{2}\leq ev(z_{m+1})\leq k\pi R_{2}^{2}. A straight-forward computation shows that

    Jac(W|AR1,R2)𝒪(A)/(1(k)kTzm+11m+k).Jac(W\big{|}_{A_{R_{1},R_{2}}^{\vee}})\cong{\raisebox{1.99997pt}{${\cal O}(A)$}\left/\raisebox{-1.99997pt}{$(1-(-k)^{k}Tz_{m+1}^{-1-m+k})$}\right.}.

    If πR22<11+mk\pi R_{2}^{2}<\frac{1}{1+m-k}, then ev(Tzm+11m+k)>0ev(Tz_{m+1}^{-1-m+k})>0 for all zm+1Az_{m+1}\in A. It follows that 1(k)kTzm+11m+k1-(-k)^{k}Tz_{m+1}^{-1-m+k} is a unit in 𝒪(A){\cal O}(A), and so

    Jac(W|AR1,R2)=0.Jac(W\big{|}_{A_{R_{1},R_{2}}^{\vee}})=0.

    Similarly, if πR12>11+mk\pi R_{1}^{2}>\frac{1}{1+m-k}, then ev(T1zm+11+mk)>0ev(T^{-1}z_{m+1}^{1+m-k})>0 for all zm+1Az_{m+1}\in A, and so again

    Jac(W|AR1,R2)=0.Jac(W\big{|}_{A_{R_{1},R_{2}}^{\vee}})=0.

    If πR1211+mkπR22\pi R_{1}^{2}\leq\frac{1}{1+m-k}\leq\pi R_{2}^{2} then

    Jac(W|AR1,R2)Λ[z±]/(1(k)kTz1m+k)SH(E)SH^(AR1,R2).Jac(W\big{|}_{A_{R_{1},R_{2}}^{\vee}})\cong{\raisebox{1.99997pt}{$\Lambda[z^{\pm}]$}\left/\raisebox{-1.99997pt}{$(1-(-k)^{k}Tz^{-1-m+k}$}\right.})\cong SH^{*}(E)\cong\widehat{SH^{*}}(A_{R_{1},R_{2}}).