The quantitative nature of reduced Floer theory
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 sees all eigenspaces whose eigenvalues have size less than , up to rescaling by a fixed constant. Similarly, we show that the reduced symplectic cohomology of an annulus subbundle between radii and captures all eigenspaces whose eigenvalues have size between and , 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 ; 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 of a contact manifold .
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 . Ritter’s result surprisingly tied the symplectic cohomology of a line bundle , denoted by , 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 , of a particular action on quantum cohomology. Precisely, there is a ring isomorphism
(1) |
The action in question is given by quantum cup product with the pull-back of the first Chern class of the the line bundle .
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 -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 family, indexed by radius, that induces an family of metrics. We write for the reduced symplectic cohomology theory associated to the radius , where is the disk subbundle of radius .
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 lying in a particular circle subbundle. In [venkatesh-thesis], we showed that
(2) |
Thus, reduced symplectic cohomology contains local information about the disk bundle .
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 , denoted by and defined by
Quantum cohomology splits into generalized eigenspaces
where is the -generalized eigenspace of the pullback of acting by quantum cup product. Each non-zero eigenvalue has a “size”, denoted by , 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 , 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 be a weak+, negative, symplectic line bundle with negativity constant . The reduced symplectic cohomology of a disk subbundle is isomorphic as a -algebra to
Here, the negativity constant determines the line bundle 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, .
-
Remark 2)
If is monotone over a toric base, all non-zero eigenvalues have the same size. For , the split-generating Lagrangian lies in a circle bundle of radius [ritter-fano]. Thus, Theorem 1 is also a generalization of the equivalence (2).
The full theory 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 by . Completed symplectic cohomology for a trivial cobordism yields an invariant , defined in [venkatesh].
Theorem 2
If is furthermore a line bundle over a toric base, there is a vector-space isomorphism
(3) |
-
Remark 3)
The main result of [c-f-o] shows that, in the Liouville setting, the uncompleted symplectic cohomology of a trivial cobordism is isomorphic to the Rabinowitz Floer homology of . Expounding upon this result, the reduced theory 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 considers positively-traversed Reeb orbits at radius and negatively-traversed Reeb orbits at radius . One could just as easily switch the roles of and , instead considering the negatively-traversed Reeb orbits at radius and the positively-traversed Reeb orbits at radius . Abusing notation, we write for the completed cohomology theory that switches these roles. This theory is dual to and satisfies
where is the -generalized eigenspace of acting by quantum intersection product on the quantum homology .
Toric line bundles have a Landau-Ginzburg mirror , where is a rigid analytic space, and is an analytic function on . 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 is monotone, then there is an isomorphism
(4) |
A local statement of mirror symmetry for annulus subbundles follows as a corollary of Theorem 1 and the isomorphism (4).
Corollary 1
Let . An annulus subbundle in a monotone, negative line bundle over a toric base has a Landau-Ginzburg mirror such that
(5) |
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 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 of dimension that is weak+ monotone. Abusing notation, denote by the image of in under the Hurewicz homomorphism. Practically, is weak+ monotone if it satisfies at least one of four conditions on :
-
1.
there is a positive constant such that on ,
-
2.
,
-
3.
,
-
4.
or the minimal Chern number, defined to be
is at least .
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 be a Hamiltonian. induces a Hamiltonian vector field , which is the unique vector field satisfying
The smooth maps satisfying
are called the periodic orbits of . Assume that the periodic orbits of satisfy a generic non-degeneracy condition. Denote the set of orbits by . Define the universal Novikov ring over to be
(6) |
Define the Hamiltonian cochain complex to be a cochain complex with underlying vector space over generated by the periodic orbits of :
Let be the minimal Chern number. Grading is given by the cohomological Conley-Zehnder index and takes values in [seidel-biased].
To define the differential on , fix a capping of each periodic orbit . The action of is defined to be
(7) |
Let be an -tame almost-complex structure that is -compatible in neighborhoods of the periodic orbits of . Consider a solution of Floer’s equation
(8) |
The energy of is defined to be
(9) |
A finite-energy Floer solution converges asymptotically in to orbits in . Fix and denote by
the set of finite-energy Floer solutions with
The moduli space is equipped with a free -action whenever that translates the -coordinate of a Floer solution. Modding out by this action yields a moduli space
whose components can be indexed by their dimension. The zero-dimensional component, denoted by , is compact if is closed. In this case, the Floer differential is defined on by
where the sign is determined by fixing choices of orientations on each moduli space and the expression refers to the homology class of the sphere constructed by gluing , , and 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 is an open symplectic manifold, which, outside of a compact set, is modeled on the symplectization
(10) |
of a contact manifold with contact form . Fix a radius and a monotone-increasing sequence with . Let be a family of Hamiltonians satisfying the following constraints.
-
1.
is bounded uniformly by a fixed constant on ,
-
2.
is linear in with slope on , and
-
3.
the non-constant orbits of lie in some small neighborhood .
The first condition ensures that the action of a periodic orbit is not heavily influenced by , the second condition ensures that Floer solutions are well-behaved, and the final condition ensures that the set of periodic orbits cluster near the contact hypersurface . Assume that each is chosen so that all one-periodic orbits of are non-degenerate. Choose an almost-complex structure that is cylindrical on :
Under the assumptions on , , and , the Floer cochain complex defined in Section 2.1 is well-defined.
By assumption, outside a compact set of . Thus, there are chain maps, called continuation maps,
for each . Let be a formal variable of degree , satisfying . Define the symplectic cochain complex of to be
(11) |
with differential , where
The cohomology of , denoted by , is called the symplectic cohomology of .
A standard result in Floer theory is
Theorem 3
The symplectic cohomology is independent of choice of radius , family , and cylindrical almost-complex structure .
We will often write instead of .
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) | ||||
(13) | ||||
(14) |
Define a valuation on by
where are -valued coefficients. Extend this to a valuation on by
where each lies in a fixed and . The non-Archimedean metric on is given by
Following [groman], denote by the completion of with respect to . Denote by the closure of in . The quotient
is the reduced symplectic cohomology of the Floer data .
An alternative definition of reduced symplectic cohomology, and one which is prevalent in the literature, is to first complete with respect to , and then define on to be the natural extension of . This defines a symplectic cohomology theory
(15) |
To avoid confusion, we call the completed symplectic cohomology.
Completed symplectic cohomology has an equivalent definition as the algebraic completion induced by a filtration. By choosing everywhere, the action is increased by both the Floer differential and continuation maps. It therefore defines a filtration on . Define the subcomplex
(16) |
and quotient complex
(17) |
These complexes are modules over the positive Novikov ring
and define cohomology theories denoted, respectively, by and . Running over all , 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])
-
Remark 5)
In practice, we will take the inverse limit over a countable set . This simplifies computations.
-
Remark 6)
and now depend upon the families used to define them. As we will see, these cohomology theories are quantitative invariants encoding local information about domains contained in .
-
Remark 7)
In this paper we study , rather than , simply because this is the object that we can compute. We give examples in Section 5 of manifolds and families on for which
(18) We do not know in how much generality (18) holds. However, we believe to be at least as robust an invariant as . 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.
Figure 1: The functions and Let be the family of Morse functions pictured in Figure 1. Denote the Morse cochain complex of over a ring by . Choose continuation maps
that correspond to the inclusion of a subcomplex. Analogously to symplectic cohomology, define a new Morse-type complex
where is a formal variable of degree , satisfying . Denoting the Morse differential by , the differential on is given by
For , define
The assignment extends to a valuation on the Morse complex that, in turn, defines a non-Archimedean metric. We use these metrics to define reduced and completed Morse cohomology theories, denoted by and . It is straight-forward to see that
but
For any single ,
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
2.4 Rabinowitz Floer homology
Recall that a Liouville cobordism 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 with positive contact boundary such that
-
1.
symplectically embeds into and
-
2.
under this embedding, the boundary of is the positive boundary of .
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
As with symplectic cohomology, admits a filtration by action, given by the chain complexes . Define subcomplexes
with homology . Action-completed symplectic homology is defined to be
Let be a family of Hamiltonians corresponding to some radius , and let , as always, be a family of Hamiltonians corresponding to some radius . We define a map
as follows. There is a map
that is the direct limit of the inclusion maps
and a map
that is projection onto the first component. Similarly, there is a map
that is the inverse limit of the projection maps
and a map
that is inclusion into the first component. The map is the composition
that projects an infinite sum onto the “” component, maps this component onto via continuation, and finally includes into the symplectic cochain complex.
Without loss of generality assume that . Define the symplectic cohomology of the cobordism to be the cohomology of the cone of , the latter written as
and the completed symplectic cohomology to be the cohomology of the cone of ,
Denote these cohomology theories by , respectively .
-
Remark 8)
If , we can still follow the above recipe to define a symplectic cohomology theory. We continue to write , respectively . These theories are dual to , respectively .
-
Remark 9)
Note that, a priori,
The left-hand side is a quotient by , while the right-hand side is a quotient by . In the examples we consider, however, the two will coincide. See Section 5.
-
Remark 10)
Suppose that is exact and is a convex contact hypersurface as in (10). has associated to it a Floer-type invariant called Rabinowitz Floer homology, denoted by . Cieliebak-Frauenfelder-Oancea showed in [c-f-o] that there is an isomorphism
In the non-exact case, there is a completed version of Rabinowitz Floer homology associated to a contact hypersurface , which we denote by . This was first studied by Albers-Kang in [albers-k]. In Section 5 we will give examples of scenarios in which, for ,
(19) There are maps
whenever [cieliebak-o]. We expect the isomorphism (20) to generalize to an isomorphism
(20)
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 be a Hamiltonian whose orbits are each either constant and non-degenerate or transversely non-degenerate. For each non-constant orbit choose a generic perfect Morse function . A choice of -tame almost-complex structure defines cascades: tuples associated to a sequence of orbits with non-constant, such that
-
1.
-
2.
is a finite-energy Floer solution corresponding to the Floer data ,
-
3.
is in the stable manifold of (or if is constant), and
-
4.
is in the unstable manifold of (or if is constant).
Choose capping discs for and for . The union represents a homology class . Let and be constant orbits of or critical points of some functions and . The moduli space is the space of tuples representing class such that is in the stable manifold of (or is equal to a constant orbit ) and is in the stable manifold of (or equal to a constant orbit ). Each component of carries an action, induced by the -actions on each Floer trajectory. By Proposition 3.2 in [bourgeois-o],
is a manifold of the expected dimension. The Floer differential now counts
The continuation maps are similarly modified.
Instead of using cascades, one could just generically perturb the Hamiltonian. However, the -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 is a transversally-nondegenerate Hamiltonian, there exists a non-degenerate Hamiltonian – a perturbation of – such that
are chain-isomorphic.
3 Negative line bundles
Let be a symplectic manifold of dimension . Denote by the line bundle satisfying for some fixed . Such a line bundle is called negative. is a symplectic manifold; assume that is weak+ monotone. Following [oancea-leray] and [ritter-gromov], we construct a canonical symplectic form on through .
Let be an -compatible complex structure on . Let be a Hermitian metric on with induced Chern curvature . Define a radial coordinate by and a fiber-wise angular one-form on the complement of the zero-section by
so that
Note that defines a contact one-form on the unit circle bundle. Let
be a symplectic form on the complement of the zero section. Extend smoothly over the zero-section by
Then is a symplectic form on and .
We wish to compute the symplectic cohomology of . There is a simple family of Hamiltonians to take, which lead to an elegant computation of . This was developed by Ritter in [ritter-gromov] and [ritter-fano] using the -action on 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 be a family of functions defined as
as depicted in Figure 2. Choose a -small Morse function on . Define
Recall that, away from the zero section,
Let be the horizontal lift of the Hamiltonian vector field on , uniquely defined through the connection one-form . As
the Hamiltonian vector field of away from the zero section is
The Reeb orbits of have period , and so the periodic orbits of exist only where is an integer multiple of . By construction, is -small, and so has no periodic orbits away from the zero section. On the zero section , which can be identified with . Thus, the periodic orbits of correspond to the periodic orbits of . As is small, these correspond precisely to the critical points of the Morse function . It follows that there is a vector space isomorphism
for every .
Suppose that is a loop of Hamiltonians based at the identity. Then acts on the loopspace by . This lifts to an action on a cover of the loopspace . We fix to be the cover defined by the deck transformation group
In other words, is the group of cappings of each loop, under the equivalence relation if
-
•
,
-
•
, and
-
•
.
Let be the action on that rotates each fiber by . This is a Hamiltonian action generated by the Hamiltonian , preserving the radial coordinate , and so
Define the lift to preserve cappings of constant loops.
Let be any one-parameter family of -tame almost-complex structures. The action of on defined by
produces another one-parameter family of -tame almost complex structures. Recall that Floer data is generic if the Floer cochain complex 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 is generic, then so is . The action induces a chain isomorphism
Choose continuation maps
between Floer data and .
Theorem 6 (Ritter [ritter-gromov])
The following diagram induces a commutative diagram on the level of cohomology.
Corollary 2 (Ritter [ritter-gromov])
The symplectic cochain complex
is quasi-isomorphic to a complex
with differential
Note that the grading shift increases the Conley-Zehnder index by .
Corollary 3 (Ritter [ritter-gromov])
The uncompleted symplectic cohomology of is
Under this isomorphism, the map
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 with .
Theorem 7 (Ritter [ritter-gromov])
The P.S.S. isomorphism
identifies up to non-zero scalar the map with the action of quantum cup product with . If is the th-generalized eigenspace of the map , there is an isomorphism of -algebras
Indeed, Ritter shows that the scalar has valuation (see Theorem 67 of [ritter-gromov]).
Corollary 4
The P.S.S. isomorphism identifies a -generalized eigenspace of with a -generalized eigenspace of , and
3.1 Disk subbundles
We now use reduced symplectic cohomology to define an invariant of a disk subbundle contained in . Fix a radius , and let be the disk bundle of radius . As in Subsection 2.2, we will construct a chain complex generated by one-periodic orbits that cluster near the boundary of . We use a very precise family of Hamiltonians, as well as almost-complex structures, to compute .
3.1.1 The almost-complex structure
Let be the space of -families of almost-complex structures on 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 be the space of -families of almost-complex structures on compatible with . Choose and . The one-form determines a splitting of into a vertical component and horizontal component . Let be the space of -families of linear maps from to . Let be an open set comprised of small neighborhoods of the circle bundles on which live non-constant periodic orbits of each , as well as small neighborhoods of the constant orbits. Let be the elements with compact support in the complement of , and satisfying for all . 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 is almost-complex. Define
Transversality and regularity for 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 for which the finite-energy cascades of and , for any , are cut out transversally.
Lemma 2 (Albers-Kang: Lemma 2.15 in [albers-k])
All simple -holomorphic spheres are regular.
3.1.2 The Hamiltonians
Fix a constant and let be a family of functions where each is
-
1.
convex and monotone increasing on ,
-
2.
bounded in absolute value by on , and
-
3.
of slope on , for some .
Further assume that the sequence tends to as tends to . To simplify later proofs, we also choose such that
-
1.
and are monotone increasing on and
-
2.
on .
Choose a Morse function function that is -small. Define a family of Hamiltonians by
We assume that the one-periodic orbits of are transversally nondegenerate. For example, we can take each to be a smoothing of a piecewise linear function, as in Figure 3.
Let be the unique vector field on satisfying
Note that is the Reeb vector field of the contact form and the simply-covered orbits of have period . As
the Hamiltonian vector field of is
Thus, the one-periodic orbits of correspond bijectively to
-
1.
the -families of orbits of with period between and , lying in fibers above the critical points of , and
-
2.
the critical points of itself.
Impose a perfect Morse function on each -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 the union of all minimum and maximum distinguished orbits of , in addition to the constant orbits, and by the union .
Choose generic almost-complex structures in to define the Floer complexes . For each choose a generic -family of functions , monotonely decreasing in , with when and when . Set . Define continuation maps
through . In the notation of Section 2, this collection of data defines cochain complexes and .
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 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 for which the finite-energy cascades of and , for any , are cut out transversally.
Lemma 4 (Albers-Kang: Lemma 2.15 in [albers-k])
All simple -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 or between two fixed periodic orbits remain in a compact region of , for any .
Proof.
This follows from an integrated maximum principle, as in the proof of Lemma 7. Indeed, by assuming that is cylindrical and outside of , the integrated maximum principle tells us that all Floer solutions appearing in a cascade remain in .
∎
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 be a non-constant -holomorphic sphere of index . As the symplectic form on is exact away from the zero-section, must intersect the zero-section (else, by Stokes’ Theorem, would have zero symplectic area, contradicting that is non-constant). Assume for contradiction that leaves the zero-section. Recall that we chose cylindrical in a neighborhood of each circle bundle containing a non-constant periodic orbit. Choose a generic disc bundle containing no non-constant periodic orbits, but such that is cylindrical on . Suppose that leaves . Apply the integrated maximum principle from [abouzaid] to . This computation shows that the symplectic area of is negative, contradicting -holomorphicity. Thus, all -holomorphic spheres are constrained to lie on the zero section.
Denote by the moduli space of simple -holomorphic curves with , modded out by the group of biholomorphic automorphisms. By weak+ monotonicity, all -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 and . The index formula says that
As we consider families of almost-complex structures, the spheres in sweep out a codimension- set in . The relevant moduli spaces of Floer trajectories sweep out a set of dimension at m in . Thus, the space of relevant Floer trajectories is generically disjoint from . But any bubble must intersect a limiting (broken) Floer trajectory [salamon]. We conclude that any bubble belonging to must intersect a periodic orbit. By the first paragraph, the bubble must be contained in the zero-section and therefore intersect a constant orbit. has codimension in , and constant orbits have dimension zero. A generic perturbation of the Morse function therefore displaces the constant orbits from any such bubbles. We conclude that bubbling cannot occur through Chern- spheres.
The Chern- 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 ); and
-
•
choosing small enough, the Floer differential on is canonically identified with the Morse differential (index ).
It remains to consider the index- moduli spaces of cascades that appear in showing . Let be the two-dimensional component of the moduli space of cascades connecting period orbits and , and associated with generic Floer data . Suppose that bubbling occurs within the moduli space . By an argument in [salamon], any bubble must intersect a dimension- component of ; in particular, and any bubble passes through .
It follows from the first paragraph that is a constant orbit. Usually one could argue that an -family of -holomorphic spheres have codimension two in , and therefore do not generically intersect a zero-dimensional constant orbit. However, the requirement 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 without risking breaking the nice structure coming from the Morse function . We will instead show that no sequence of maps in converges to a broken trajectory. This will show that, even if “sees” bubbling, it does not affect the computation .
Suppose that a sequence of cascades converges to a broken cascade . By Lemma 7, is a constant orbit as well, and all curves remain in a region where the Floer dynamics are governed by the small function . In this region, and are moduli spaces of Morse trajectories, and so one of these spaces is empty. Thus, is empty as well.
∎
-
Remark 11)
If is toric and is perfect, one can also just conclude that and are both empty for degree reasons. If is not toric, one can get around using almost-complex structures in . 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 . 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 , and so the Novikov exponents contributing to 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 depends on the choice of Morse perturbation . On the other hand, reduced symplectic cohomology is independent of this choice.
3.1.4 Some technical lemmas for computation
In order to compute , we need three standard results on the behavior of the Floer differential and continuation maps.
Define to be the winding number of a non-constant periodic orbit , viewed as a map from the circle into . Define of a constant orbit to be zero.
Lemma 7
Let be a solution of Floer’s equation (8) with . Then .
Proof.
We use the integrated maximum principal of [abouzaid]. Assume for contradiction that . Say that lives in the sphere bundle of radius . If are non-constant then the -orbit underlying has period . From the earlier computation it follows that . The winding numbers are integers, and so
By the smallness of we can assume that
As is convex, we deduce that . If is constant and is non-constant it follows immediately that .
Choose a generic circle subbundle of radius , with , and on which and is cylindrical. For example, if is close to or these conditions will, by construction, be met. Let be the region bounded by , and denote . Let be the restriction of . We will equate with its image under the inclusion into and use the coordinates induced on .
Let be the -intercept of the tangent line to at . Then on , (and ). In particular,
Let be the union of the boundary components of mapping into . Note that we have chosen to be -tame, so that
Shuffling terms, we have
We consider this final equation in pieces. A solution of Floer’s equation satisfies
We have that, on , and . The former implies that is proportional to and the latter implies that lives in the horiziontal distribution. So altogether, . Thus, the “” terms become
where the last inequality follows because a vector that is positively-oriented with respect to the boundary orientation satisfies . We also have
and so the “” terms become
Altogether,
(21) |
is a collection of bounded regions in and one unbounded region enclosing the origin. As is exact on , the bounded regions contribute nothing to the right-hand side of (21). Let be the unbounded component. Near zero, is contained in a neighborhood of , and so all boundary components of occur within the intersection of with some annulus . Let be a function on that is equal to on and equal to for all radii . Then
As is convex by assumption and , both and . It follows that
(22) |
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 .
Lemma 8
Let be a solution of Floer’s equation with respect to a Hamiltonian such that everywhere. Suppose . Then .
Proof.
The proof of Lemma 7 applies almost verbatim, except the energy will have an additional integral
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 to the periodic orbit of represented by the same map .
Proof.
Let be a Floer solution of the data , where induces an action-increasing continuation map . By assumption, on . Thus, within the disk bundle of radius . By index considerations, any Floer solution is either constant or leaves the disk bundle of radius . The latter cannot happen, by Lemma 8 and its easier analogue: 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 and the closely-related theories and . Namely, there is an inclusion
induced by the inclusion
such that the following diagram commutes.
(23) |
Here, and come from the inclusion of into its completion.
We will prove
Proposition 1
The image of is equal to the image of .
As is injective, this implies that is isomorphic to , in particular,
Recall from Corollary 3 that the P.S.S. map
induced by the inclusion
is surjective. The composition
(24) |
is therefore surjective and its image coincides with . Thus,
The remainder of this section is devoted to computing the kernel of and proving Proposition 1. Recall from Theorem 4 the isomorphism
Under this isomorphism, the map is the inverse limit of maps
induced by the chain-level quotient maps
We will compute the kernel of each and show that
4.1 Step I
Consider an auxiliary family of Hamiltonians, defined as follows. Recall the radii that are part of the data of the Hamiltonians . Define
The subcomplex
has a valuation derived from the usual action on each Floer chain complex. We extend this non-trivially to all of by defining
Geometrically, this corresponds to the shift
and ensures that action is increased by choices of continuation maps
See Figure 4.
-
Remark 13)
The “space” between and at is important for defining action-increasing continuation maps. Without the space, there is no generic monotone-increasing homotopy. See [seidel-biased].
Because action is increased by continuation, has subcomplexes and quotient complexes analogous to the subcomplex (16) and quotient complex (17) of .
Choices of action-increasing continuation maps
induce a continuation map
that descends to a map
This induces maps on the long-exact sequences
such that each square commutes. The commutativity of the right-hand square gives
(25) |
By construction,
where is the PSS map, defined similarly to (24). associated action-filtered map
such that
(26) |
We now assume that is algebraically closed and has characteristic zero. Recall the following fact.
Lemma 10 (Lemma A.1 in [fooo])
is algebraically closed if is algebraically closed and of characteristic zero.
Denote by the eigenvalues of the map
(including geometric multiplicities) and fix a Jordan basis
so that spans the invariant subspace associated with the eigenvalue .
Define a valuation
Lemma 11
The valuation satisfies
-
1.
and
-
2.
There exists a constant , such that, for any with ,
Proof.
Let be a representative of such that
Let be a representative of and a representative of such that
As is a non-Archimedean valuation,
Thus,
This shows item (1). Item (2) follows immediately from the fact that is finite-dimensional over .
∎
Using Property (1) of Lemma 11, we can prove the following Lemma.
Lemma 12
Let be an eigenvalue of with . For any constant and any basis vector in ,
Proof.
From Equation (26), it suffices to show that
Indeed, we will show that
(27) |
If , that is, if , this follows immediately from Corollary 3. So suppose . We will construct a cocycle such that
and
Let be a cocycle representing . Define an infinite sequence by defining the base case
and inductively defining
Because the Floer differential commutes with continuation maps,
Thus,
and so
for any . We will show that, for large, .
Denote by the invariant subspace corresponding to . In the fixed Jordan basis, the matrix of is
Ignoring non-zero -scalar factors (these do not affect action),
(28) |
and
for all . View under the identification . Again ignoring non-zero -scalar factors,
Using Property (1) of Lemma 11, we compute
(29) | ||||
(30) | ||||
(31) | ||||
(32) | ||||
(33) | ||||
(34) |
where
Let By Theorem 6,
The map preserves action, and so
Indeed, the representatives of and are in bijective correspondence through the action-preserving map . Thus,
(35) |
Recall that . As is a chain map,
The canonical identification
descends to a map on cohomology that sends to . This isomorphism decreases action on-the-nose by , and so the action of is bounded by
and so
The difference is bounded below by the positive number , and is a constant. There therefore exists satisfying
Thus,
∎
Lemma 13
Denote by the invariant subspace of corresponding to an eigenvalue of .
Proof.
The map is a -module homomorphism. Any element in can be written as the sum of generalized eigenvectors with and . Lemma 12 shows that
By linearity, as well.
∎
4.2 Step II
We want to show that, as ,
Denote by the intersection of with the image of the map
Proposition 2
The kernel of satisfies the following inclusions:
and
where is a constant depending only upon the fixed Jordan basis .
The proof of Proposition 2 relies on four Lemmas that bound action.
4.2.1 Lemmas bounding action
Lemma 14
There exists a constant such that, for any cocycles satisfying
the difference in actions is no more than , that is,
Proof.
This follows immediately from property (2) of Lemma 11 and the fact that is finitely generated.
∎
Lemma 15
The valuations of the Novikov-valued coefficients are not decreased by .
Proof.
Let be a monotone homotopy between and of the form
Thus, the homotopy only changes in the radial direction.
Let be a family of admissible almost-complex structures. A solution of the Floer equation
descends to a solution of the Floer equation
(36) |
Let be the compactification of by its limit points, and let be the compactification of by the limiting fiber disks, so that . Since
contributes a weight
to . Action is increased by solutions of (36), and so
In particular,
(37) |
By assumption, is small; assume that and is small enough that
(38) |
or
Applying this argument to every , the Lemma follows.
∎
Let be eigenvalues of the operator
indexed with multiplicities. Let be a Jordan eigenbasis for . Write the th-generalized eigenspace as .
Let , and write in the basis .
(where and the range of depends on ). Now write
(where .)
Lemma 16
There exists a constant , depending only on the choice of bases, such that
Proof.
Let be constants determining the change-of-basis , so that
Take
Then
and so
It follows that
The inequality given by the last line follows from the definition of the action of and the observation that
for any fixed . ∎
Lemma 17
If is an eigenvalue of the map
then
Proof.
Let be the eigenvector satisfying
Let be an upward shift of the Hamiltonian :
for some small . Let be an action-increasing
Continuation maps are homotopy invariant, so under the identification induced by the equality ,
Modifying if necessary, we assume without loss of generality that
commutes. Clearly,
where the last inequality follows from the fact that action is increased by . Thus,
(39) |
As (39) holds for any , we conclude that
∎
4.2.2 Proof of Proposition 2
Let
be an action-increasing continuation map. The following Lemma simplifies the computation of Lemma 19.
Lemma 18
The continuation maps may be chosen to commute at the level of cochains with the continuation maps and , that is,
Proof.
We sketch the proof. Let be the map interpolating between the functions and in the definition of and . Define the continuation maps inductively as follows. Denote by a generic map interpolating between and so that
-
1.
on and
-
2.
is monotone-decreasing in elsewhere.
Define through .
Now let be the function that is
-
1.
equal to on and
-
2.
a translation of on .
See the teal curve in Figure 5. Let be a generic function interpolating between and that is
-
1.
equal to on and
-
2.
translation by a constant on .
Define a continuation map through .
Let be a generic function interpolating between and that is
-
1.
equal to on and
-
2.
monotone-decreasing in on .
Define a continuation map through . Set
The Lemma follows from the integrated maximum principal of Lemma 7.
∎
Denote by
the projection onto the th-generalized eigenspace. The following Lemma bounds the action of a cochain in an arbitrary Floer complex by the action of a cochain in .
Lemma 19
Let be an eigenvalue with . Let be a cocycle such that
Let be the unique element in satisfying . There exists a constant , independent of , satisfying
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 is a bona-fide -eigenvector, and is a cocycle of winding number satisfying
for some . By Lemma 17, the valuation of the Novikov coefficient is increased by (in this case the Novikov coefficient is and has valuation ). Therefore,
Assume that the circle bundle on which lives has radius close to . The action of , compared to the action of , is
if . Taking , 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 to be a weighted sum of periodic orbits and to be a weighted sum of generalized eigenvectors.
Proof.
Write
where and . Because is a -linear morphism,
Write in the basis :
(40) |
for some . By assumption,
(41) |
Write in the fixed Jordan basis B:
We want to compare the action of some with the action of some . We do this through the auxilliary cochain . By definition, and ignoring extra non-zero factors,
Thus, (41) produces equalities for each :
(42) |
Fix so that . Lemma 17 states that . As a consequence, the action of the right-hand-side of (42) is completely determined by one of the summands:
Thus, there exists some , which we denote by , satisfying
(43) |
Denote by the winding number of . View as an element of under the inclusion . By Lemma 18,
(44) | ||||
(45) |
Again appealing to the Jordan normal form, write
(46) |
The orbits in are constant and appear at energy level . Lemma 15 states that the valuation of the Novikov-coefficient is increased by . The action of is the valuation of the Novikov coefficient, plus the value of . Together, this implies that
(47) |
Applying the map to (46), ignoring non-zero scalars as always, yields
Comparing this with (40) through the equivalence (44), and ignoring non-zero -scalars,
There therefore exists some , denoted by , with
(49) |
Letting in (48),
(50) |
Combining equations (43), (49), and (50),
Recall that the winding number of is . Let lie in the circle bundle of radius . We conclude that
Because , and because approaches as approaches infinity,
for some fixed independent of .
Furthermore, there is a constant bounding the expression
Let
Then
Take
We have shown that, for any , we can find a satisfying
By definition,
for each . Combining this with the non-Archimedean nature of ,
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 to be the projection
Lemma 19 has the following easy corollary.
Corollary 6
Choose any such that
Let satisfy
Then
Proof.
First note that, if
then
for each with valuation . By Lemma 19, there exists a constant such that
for all . By the non-Archimedean property of ,
Combining these two inequalities,
∎
Finally, we would like a cohomological analogue of Corollary 6. Denote by
projection onto the -generalized eigenspace, and, abusing notation, denote by the projection
Let us recall some basic linear algebra facts about maps acting on finite-dimensional vector spaces.
-
1.
Any eigenvalue of is an eigenvalue of .
-
2.
Without loss of generality, we can assume that the collection of generalized eigenvectors are chosen to equal
-
3.
Restricting to the invariant subspace , we can abuse notation and assume that is an eigenbasis for .
-
4.
If , then .
Corollary 7
Let such that
Let satisfy
Let be any cochain-level representative of . Then
Proof.
Let be a minimum-action cochain-level representative of . Note that
Decompose into components and , where
and when . Clearly and
Indeed, there exists
such that
As is an isomorphism on and is an invariant subspace, has a well-defined inverse , and
By Lemma 19, there exists with
But , and so, by Lemma 14, there exists with
Setting ,
∎
We are finally ready to prove Proposition 2.
Proof of Proposition 2: By Lemma 12,
Clearly
By the linearity of , it therefore suffices to show the inclusion
Let
Suppose that . Let be a cocycle descending in cohomology to
As
we can assume without loss of generality that
(51) |
Let be any representative of , and let any integer with . Under the PSS map,
in . By homotopy-invariance of continuation maps,
By Corollary 7,
or
It follows that
∎
4.2.3 The full PSS map
We have characterized ; we now want to characterize .
Proposition 3
The isomorphism
induces an isomorphism
(52) |
where the connecting maps for
are inclusions.
Proof.
Let
(53) |
be the connecting map in the inverse limit. By definition,
Thus,
Similarly,
Let
be the inclusion. Abusing notation, let
be the surjective map induced by restricting the connecting map (53). The diagram
commutes. As the inverse limit is left-exact, there is an exact sequence
(54) |
We will show that .
By Proposition 2 there exists some set of constants , such that
and . Inverse limits commute with finite sums, so that
The second equality holds because the connecting maps of the system
are isomorphisms, and the system therefore satisfies the Mittag-Leffler condition.
Recall from Remark Remark 5) that the index belongs to a fixed countable sequence . By definition,
(55) | ||||
Choose any element
For each , consider the formal sum
As , it has a cochain representative such that . This implies that
Thus,
More particularly,
is a cochain representative whose cohomology class is given by . We conclude that
Identify with some preimage in . There is a telescoping sum
We have shown that the map is surjective, and so, from the definition (55),
∎
Corollary 8
There is an isomorphism
Proof.
It suffices to show the equality
Proposition 3 shows that the connecting maps of are inclusions; and so the inverse limit is the intersection:
By Proposition 2,
for some constants with limiting behavior . The only element of with arbitrarily large action is , and so
By definition,
whenever . The Corollary follows.
∎
4.3 The product structure
Recall that 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 be a Riemann surface with two positive punctures denoted by and one negative puncture denoted by . Choose collar neighborhoods of both and and of . Equip with a one-form such that on the collar neighborhood of , on the collar neighborhood of , and on the collar neighborhood of , for fixed positive integers satisfying . As shown in [ritter-fano], we may assume that also satisfies
For a fixed Hamiltonian , let satisfy
with respect to a generic -tame, upper-triangular almost-complex structure. The energy of is defined as
We call a pair-of-pants. The following Lemma is a standard application of the integrated maximum principal.
Lemma 20
Suppose has finite energy. The image of remains inside .
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 converge to periodic orbits of , and are therefore contained in the interior of . The negative puncture of converges to a periodic orbit of , and is therefore also contained in the interior . Let be a generic disk bundle in containing the three asymptotic periodic orbits. Define , and suppose for contradiction that is non-empty. Then
where the inequality follows because and everywhere.
Let be the -intercept of the tangent line to at . Then on , . Thus,
By assumption everywhere. Furthermore, on , and, as the -intercept of , . Therefore,
We conclude that
But is non-negative by definition. Therefore . This can only occur if
everywhere. As is contained in the horizontal distribution determined by , the image of must be entirely contained in . But this contradicts the first observation made, that near the punctures of , approaches periodic orbits of or .
∎
Fix a Hamiltonian . For a fixed index , define a branch to be a cascade
associated to a sequence of periodic orbits and a non-decreasing sequence of integer weights , where
-
1.
,
-
2.
is a finite-energy Floer solution corresponding to an -family of Hamiltonians , where when and when ,
-
3.
is in the stable manifold of (or if is constant),
-
4.
is in the unstable manifold of (or if is constant).
Define a tree to be a sequence where is a pair-of pants corresponding to weights , , and ; and where
-
1.
taken in the collar neighborhood of , lies in the stable manifold of ,
-
2.
taken in the collar neighborhood of , lies in the stable manifold of ,
-
3.
and taken in the collar neighborhood of , lies in the unstable manifold of .
Let be the moduli space of rigid trees , where is in the stable manifold of , is in the stable manifold of , is in the stable manifold of , and and . See Figure 6.

A standard Corollary to Lemma 20 is that
Corollary 9
The moduli space is compact.
Define a map
for each . Extend to a map
where . Recall that the continuation maps were chosen to be inclusions, so, is just the composition of the inclusions
with the pair-of-pants product.
As , there are action-increasing continuation maps
Omitting signs, define a map
by
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
descends to a map
The energy of a map with positive punctures mapping to and and negative puncture mapping to is
Thus,
It follows that induces a well-defined product on , and, in particular, on .
Lemma 22
The product induces a well-defined product on .
Proof.
The ordinary product satisfies
for cochains . Suppose that . Write
where are cocycles satisfying and . Then
So is well-defined on . Now suppose . Write
where . Let . Then
Thus,
As , we conclude that
as well. Summing over all ,
∎
The map
clearly intertwines and . Thus, the cohomology-level map
is a ring isomorphism.
4.3.1 Finishing the proof
Proof of Theorem 1: Recall the injective map
and the commutative diagram
(56) |
Given Corollary 8, it suffices to show that
Clearly
We will show the reverse inclusion. Fix and choose any cochain representative of . By definition, there are cochains in with , such that
Let be the descent of to cohomology. identifying with its image in ,
Write in the fixed eigenspace decomposition:
where
Choose such that
By Corollary 7, there exists a constant such that
so that
(57) |
Let
Equation (57), together with the assumption that , shows that is a well-defined element of .
Recall from Lemma 13 that
Thus,
By linearity of and ,
Thus,
As is injective and , we conclude that is surjective. It follows that there is a ring isomorphism
By Corollary 8,
Finally, by Corollary 4 and the PSS identification ,
∎
5 Toric line bundles
We assume in this section that is a toric symplectic manifold. In particular, we will use the fact that has a perfect Morse function , and the critical points of all have even index. We denote the Morse index of a critical point of by . Let be a family of functions defined as in subsection 3.1.2, using the perfect Morse function . Let be an almost-complex structure of the form found in Subsection 3.1.1.
Proposition 4
There is an isomorphism
-
Remark 14)
In the author’s thesis [venkatesh], we computed 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 . It is useful to first consider a -graded Floer theory. Let be a formal variable of degree , where is the minimal Chern number of . 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 is graded. A generator with capping has even degree, respectively odd degree, if is the minimum, respectively maximum, of a perfect Morse function on a transversally non-degenerate family of orbits. If is chosen to be the fiber disk, then
where the final term is if is a minimum and if is a maximum.
-
Remark 15)
The grading in Lemma 23 is the cohomological Conley-Zehnder, shifted up by . We shift the grading for ease of notation; the grading on now coincides with the grading on singular cochains.
-
Remark 16)
Lemma 23 separates symplectic cohomology into an odd-graded component, denoted by , which consists of all “minimums”, and and even-graded component, denoted by , which consists of all “maximums”, plus the constant orbits.
Lemma 24 (Albers-Kang [albers-k])
An element , with and descends to an element
of the Floer cochain complex associated with the Floer data . This cochain complex is -graded. The projected generator has degree
Let and define
Lemma 25 (Albers-Kang [albers-k])
A solution of the Floer equation associated with the Floer data on projects to a solution of the Floer equation associated with the Floer data on .
Corollary 10
The complexes define a filtration on .
Note that this filtration induces a filtration on the subcomplex and the quotient complex . The following two Lemmas, stated for the full complex , also apply to these smaller complexes.
Lemma 26 (Albers-Kang [albers-k])
The differential on decomposes as
where increases the grading of the projected generator by .
Lemma 27 (Albers-Kang [albers-k])
Floer solutions contributing to the zeroth part of the differential, , remain in a single fiber. If is the maximum and is the minimum of a perfect Morse function on an -family of orbits, then and .
Consider Lemma 23. If and
lies in a fixed degree , then
where is the winding number of in the fiber and the contribution in depends only on . So is a fixed constant independent of , which we denote by .
Lemma 28
Let be an element in the cochain complex or , which consists of formal sums whose action limits to infinity (in the latter case, the sum includes only finitely many distinct “”). Assume for any . Then
Proof.
By the definition of and ,
Because each Hamiltonian is bounded, this implies that
(58) |
By construction, the winding number satisfies
(59) |
Thus,
(60) |
Equations 58 and 60 imply that
and so
(61) |
where is the constant satisfying . The choice of grading implies
where is a bounded constant in the interval . Equation 59 now implies that
(62) |
Finally, by Lemma 24,
This grows like , which, by observations (62) and (61), limits to infinity.
∎
Let be the differential on , and let be the differential on . Let
(63) |
be the covering map that sends to . Let be the induced map on homology.
Fix a degree and consider the -linear map given on generating orbits by
where is uniquely determined by the degree formula
On the level of vector spaces, is inverse to . On the level of cochains,
and descends to a map on , inverse to
Thus, is an isomorphism, and so the full map is surjective on cohomology.
respects the action filtration, inducing a surjective map
(64) |
This map restricts to an isomorphism
which induces an isomorphism
But inverse limits commute with finite direct sums, and so
We conclude that the induced map
(65) |
is surjective as well.
Lemma 29
If is an even-graded periodic orbit then .
Proof.
First take coefficients in . By index conventions, is a maximum of a perfect Morse function on an -family of orbits. Suppose for contradiction that . By Lemma 28, we can write , where and . By definition, , implying that . Decompose with respect to the filtration:
where . Then if and only if for every . In particular,
However, by index considerations, each generator appearing in is a minimum of a perfect Morse function on an -family of orbits. By Lemma 27,
A contradiction is reached.
Now take coefficients in . Recall the map from Equation (63), which takes to itself. As is a chain map,
∎
Corollary 11
The connecting maps
are surjective.
Lemma 30
The kernel of is on the odd-component:
As a consequence,
Proof.
First consider the uncompleted theory . The choices made ensure that
from which it is clear that . Corollary 3 expresses an isomorphism
This isomorphism preserves the parity of the grading, and it follows that
It remains to show that
First consider coefficients in . Choose a non-zero sequence , and assume for contradiction that each is a cocycle and the connecting morphisms send to , so that . By definition, there exists a coboundary such that, at the chain level,
The proof of Lemma 29 implies that , so
(66) |
The action of the summands of the nested sequence approaches infinity. By Lemma 28, there exists some such that for all . Assume we have chosen the maximum such , so that, for large enough , some non-zero summand of achieves . Without loss of generality, assume that we have fixed a large enough .
Let . Write , where , and similarly write . We want to show that . It suffices to show that there exists a single non-zero summand .
Due to the grading conventions, is the minimum of a perfect Morse function on the underlying trajectory. Therefore, by Lemma 27, . It follows that
and so
Thus, . As this holds for all sufficiently large , we conclude that and
Combining the surjectivity of the map (65) with Lemma 29, it follows that
as well.
∎
We end this subsection by proving Proposition 4.
Proof.
We have seen that there is an injective map
induced by the inclusion
Lemma 30 shows that , which implies that
as well. By Lemma 29,
and so
It follows that the map 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 that lie between disk bundles of raddi and . We would like to know for which radii completed symplectic cohomology is non-vanishing.
Let be quantum homology, the dual of quantum cohomology. The dual of the map is the quantum intersection product with the Poincaré dual of , denoted by . These maps have the same eigenvalues, and we denote by the -invariant subspace of under .
Theorem 8
If , there is a vector-space isomorphism
If , there is a vector-space isomorphism
-
Remark 17)
In particular,
precisely when the size of some eigenvalue lies between and .
By construction, there is a long exact sequence
(67) |
To compute it therefore suffices to compute , , and the connecting map . 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
and
Proof.
Applying the proof of Lemma 30 to a prospective cocycle in yields
There is a Milnor exact sequence
Corollary 11 says that the connecting maps
are surjective. Thus, the Mittag-Leffler condition is satisfied and
The first isomorphism follows.
The second isomorphism follows from the exactness of direct limits.
∎
We begin by computing and . The following Lemma is standard fare, but sets the stage for subsequent elaboration.
Lemma 32
Up to grading, is the dual homology theory to .
Proof.
A standard Morse theory argument shows that the Floer chain complex is naturally isomorphic to the dual of the Floer cochain complex , where counts rigid Floer cylinders with negative input and positive output. By definition, a generator of corresponds to a generator of . is the morphism which on generators evaluates to
(68) |
As direct products are dual to direct sums,
with differential
Here, is defined similarly to , but, as with , it counts the rigid Floer cylinders with negative input and positive output.
Poincaré duality in Floer theory yields natural isomorphisms
See [c-f-o] for an elaboration on the grading, recalling that we have shifted their grading scheme by . These isomorphisms intertwine the continuation maps with the original continuation maps . Thus, there is a natural chain isomorphism
∎
Just as acts on , the composition acts on . The bijection
intertwines and , and it is easy to see that these two compositions have the same eigenvalues. We denote by the -generalized eigenspace of the action of on .
Lemma 33
The isomorphism
induces an isomorphism
such that the map is the canonical inclusion.
Proof.
The dual of the inclusion
is the projection
Under Poincaré duality, this projection is a map
which descends to the map
(69) |
As is torsion free, is an injective module, and the Universal Coefficient Theorem yields a natural isomorphism
In particular, denoting by the dual of the projection
the following diagram commutes
Identifying with the quotient complex
the group is precisely the subset of on which vanishes, and is the canonical inclusion.
Let be the basis of dual to the fixed Jordan basis . By definition, is a generalized eigenbasis for the action on dual to . A standard result in Floer theory shows that, under the identification , the dual action is the map
Under the Poincaré duality isomorphism , the continuation maps and are identified, and is identified with
Thus, the dual basis is a generalized eigenbasis for the action of on . By definition, precisely when . Thus,
∎
Lemma 34
Up to grading, is the dual homology theory to .
Proof.
Equip the uncompleted chain complex with the non- Archimedean metric given by
Completing with respect to yields a complex whose elements are formal sums
(70) |
Because the connecting maps in the inverse limit are surjections, is isomorphic to the complex (70). The dual of is bounded homomorphisms on , that is,
equipped with the usual differential
We want to show that is chain isomorphic to .
Analogously to the statement (70) for cohomology, is isomorphic to
(71) |
Recall from the proof of Lemma 33 that Poincaré duality for Hamiltonian Floer theory is a chain isomorphism
(72) |
induced by the set bijection via . For each , let be the corresponding element of under this bijection. Consider the map
(73) |
given on generators by
where
To check that this map is well-defined, enumerate , and write
so that there exists with
for all . Then for all ,
If for any , then and
So suppose there exists such that . Then
where the last equality follows because the capping disk of is the capping disk of with reversed orientation. Thus,
Similarly,
Extending linearly, we find that is indeed a bounded operator.
To see that the map (73) is surjective, suppose , so that
for some fixed . Then
Let
Then
Similarly,
so
Analogously to the computation of , we can compute action-completed symplectic homology as a subspace of .
Proposition 5
There are isomorphisms
and
under which the map
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.
The map
is identified with the map dual to
denoted by
Under the identifications
and
is an inclusion
where is the subspace of on which vanishes. As in the proof of Lemma 33, is precisely
∎
It remains to compute the connecting map
We have seen that can be identified with the composition
Note, trivially, that the map is equivalent to the map . The map maps each -generalized eigenspace isomorphically onto a -generalized eigenspace. The map restricts to an isomorphism on , as this subspace contains no -generalized eigenvectors. Thus, the map restricts to an isomorphism
This shows
Lemma 35
The connecting map can be identified with the composition
It therefore has image
and kernel
We are now ready to prove Theorem 8.
Proof of Theorem 5: If , the map is injective by Lemma 35. The long exact sequence (67) and the isomorphism
of Lemma 31 show that
By Lemma 35 the right-hand side is
Finally, the isomorphism of Corollary 7 and the isomorphisms of Corollary 4 show the first statement in Theorem 8.
If , the map is surjective by Lemma 35. The long exact sequence (67) and the isomorphism
of Lemma 31 show that
By Lemma 35 the right-hand side is
Dualizing the isomorphism
and applying Poincaré duality yields an isomorphism
The first isomorphism intertwines with the map . The second isomorphism intertwines the dual maps: the quantum intersection product and . The result follows.
∎
Theorem 8 explicates a form of self-duality:
Corollary 12
The groups and 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 vanishes whenever . Their methods, in conjunction with the work of this section, show that, for and a monotone line bundle of negativity constant ,
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 inherits the structure of a toric variety from the base and the -action on the fibers. Its moment polytope can be described in terms of the moment polytope of (see Subsection 7.6 in [ritter-gromov] or Subsection 12.5 in [ritter-s]).
-
Example 1)
Let be the complex line bundle . is a toric variety whose image under the moment map is
The facet of lying in the plane is precisely .
has a conjectural Landau-Ginzberg mirror , where , and
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 and the Jacobian of :
(74) |
Suppose that is monotone, so that , and suppose . Then is monotone as well, with monotonicity constant . Suppose further that the superpotential 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 has a unique Lagrangian torus fiber that, when equipped with suitable choices of line bundles, split-generates the wrapped Fukaya category. sits inside the circle bundle of radius . The fiber that is mirror to , defined by , contains all critical points of . Open-string mirror symmetry matches each choice of line bundle with a critical point of .
-
Remark 19)
With monotone, the Chern classes and are related by a constant. The quantum cup products by and by therefore have the same eigenvalues, up to -scalar. Ritter showed in [ritter-fano] that all eigenvalues of have valuation in . The Floer-essential Lagrangian therefore sits inside the circle bundle whose radius satisfies
for some non-zero eigenvalue of . This is precisely the critical radius where non-vanishing symplectic cohomology theories occur. In particular,
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 be the annulus bundle between radii and in , with . The mirror of is
equipped with .
For and , denote by . We denote the ring of functions on in the variable by , where
Denote by the zeroes of the function . is an example of a Laurent domain: if is described by the functions
then is cut out by the inequalities
where
Laurent domains are examples of affinoid domains, and therefore satisfy a Nullstellensatz [tian]. In particular,
This occurs if and only if , that is, if and only if
Conversely, since is Morse by assumption, if then
Altogether,
But, as discussed in Remark Remark 19), is exactly the valuation of the non-zero eigenvalues of . Thus, Theorem 8 says that
Via the isomorphism of Equation (74), we conclude a closed-string mirror symmetry statement for subdomains:
Closed-string mirror symmetry If , then
-
Example 2)
Again let . The mirror of is
equipped with superpotential
(75) (76) (See Example 7.12 in [ritter-gromov] or Proposition 4.2 in [auroux].) Denote by the one-dimensional annulus . A straight-forward computation shows that
If , then for all . It follows that is a unit in , and so
Similarly, if , then for all , and so again
If then