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

Functors of wrapped Fukaya categories from Lagrangian correspondences

Yuan Gao1 1Department of Mathematics, Stony Brook University, Stony Brook NY, 11794, USA ygao@math.stonybrook.edu
Abstract.

We study wrapped Floer theory on product Liouville manifolds and prove that the wrapped Fukaya categories defined with respect to two different kinds of natural Hamiltonians and almost complex structures are equivalent. The implication is we can do quilted version of wrapped Floer theory, based on which we then construct functors between wrapped Fukaya categories of Liouville manifolds from certain classes of Lagrangian correspondences, by enlarging the wrapped Fukaya categories appropriately, allowing exact cylindrical Lagrangian immersions. For applications, we present a general Künneth formula, and also identify the Viterbo restriction functor with the functor associated to the completed graph of embedding of a Liouville sub-domain.

1. Introduction

This paper pushes further the discussion in [Gao17] studying Lagrangian correspondences between Liouville manifolds. The goal is to understand the natural functors between wrapped Fukaya categories arising from Lagrangian correspondences, with a motivation in investigating the functoriality properties of wrapped Fukaya categories, as well as the relation to the well-established functoriality properties of the devired categories of coherent sheaves via homological mirror symmetry.

1.1. Floer theory on product manifolds

As Lagrangian correspondences are simply Lagrangian submanifolds of the product manifold, it is natural to study first of all the wrapped Fukaya category of a product Liouville manifold and relate it to the wrapped Fukaya categories of both factors, 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N). To understand such a relation, it is natural to setup up wrapped Floer theory using the split Hamiltonian, which is the sum of admissible Hamiltonians on both factors, i.e. Hamiltonian of the form πMHM+πNHN\pi_{M}^{*}H_{M}+\pi_{N}^{*}H_{N}. This together with the product almost complex structure defines a version of wrapped Fukaya category of M×NM\times N, which we call the split model of wrapped Fukaya category and denote by 𝒲s(M×N)\mathcal{W}^{s}(M\times N).

However, as already addressed in [Gao17], there is a technical issue: the split Hamiltonian is a priori not admissible for wrapped Floer theory, meaning that it might have defined a different category compared to the quadratic Hamiltonian on the product M×NM\times N with respect to a natural cylindrical end. This issue was resolved on the cohomology level there. In this paper, we carry out a chain-level discussion, confirming that the wrapped Fukaya category defined with respect to the split Hamiltonians is quasi-isomorphic to the one defined with respect to a quadratic Hamiltonian. Thus up to canonical quasi-equivalence, there is no ambiguity in mentioning the wrapped Fukaya category of the product manifold. This would simplify many functoriality arguments in wrapped Floer theory, and expectantly in the study of homological mirror symmetry.

Theorem 1.1.

The split model 𝒲s(M×N)\mathcal{W}^{s}(M\times N) is quasi-equivalent to the ordinary wrapped Fukaya category 𝒲(M×N)\mathcal{W}(M\times N).

To be more precise, let 𝕃\mathbb{L} be a countable collection of admissible Lagrangian submanifolds of M×NM\times N and 𝒲s(𝕃)\mathcal{W}^{s}(\mathbb{L}) (resp. 𝒲(𝕃)\mathcal{W}(\mathbb{L})) be the full subcategory consisting of objects in 𝕃\mathbb{L}. Then there is a natural quasi-equivalence

(1.1) R:𝒲s(𝕃)𝒲(𝕃).R:\mathcal{W}^{s}(\mathbb{L})\to\mathcal{W}(\mathbb{L}).

As this AA_{\infty}-functor is constructed using an action-filtration argument, it will be called the action-restriction functor.

1.2. Functors associated to Lagrangian correspondences

Next, we investigate a specific class of Lagrangian correspondences, which are either products or cylindrical with respect to a natural choice of cylindrical end of the product M×NM^{-}\times N. Associated to every such Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N, we would like to construct an AA_{\infty}-functor from the wrapped Fukaya category of MM to that of NN. Technically, this is not always possible. To overcome this, we include wider class of objects in the wrapped Fukaya category, which are to be introduced in section 7. Details are to be presented later, but let us first illustrate the main spirit below.

Using a quilted version wrapped Floer cohomology, we first construct an AA_{\infty}-functor

𝒲(M×N)(𝒲(M),𝒲(N))bimod\mathcal{W}(M^{-}\times N)\to(\mathcal{W}(M),\mathcal{W}(N))^{bimod}

from the wrapped Fukaya category of the product to the AA_{\infty}-category of (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N))-bimodules. By purely algebraic considerations, this gives rise to an AA_{\infty}-functor

𝒲(M×N)func(𝒲(M),𝒲(N)lmod)\mathcal{W}(M^{-}\times N)\to func(\mathcal{W}(M),\mathcal{W}(N)^{l-mod})

by applying the Yoneda embedding on the second factor 𝒲(N)\mathcal{W}(N). Concretely, to each admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N, we associate an AA_{\infty}-bimodule PP_{\mathcal{L}} over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)) whose value on a pair (L,L)(L,L^{\prime}) is the quilted wrapped Floer cochain complex CW(L,,L)CW^{*}(L,\mathcal{L},L^{\prime}). Regarding LNL^{\prime}\subset N as a testing object, with small amount of homological algebra argument we then immediately get an AA_{\infty}-functor

(1.2) Φ:𝒲(M)𝒲(N)lmod.\Phi_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}(N)^{l-mod}.

Then we study the geometric composition of Lagrangian correspondences, which is in general a Lagrangian immersion. In order to be able to include these immersed Lagrangian submanifolds as objects of the wrapped Fukaya category, in section 4 we define the immersed wrapped Fukaya category of MM, denoted by 𝒲im(M)\mathcal{W}_{im}(M), whose objects are unobstructed proper exact cylindrical Lagrangian immersions of MM (with transverse self-intersections), together with bounding cochains for them. Floer theory for such Lagrangian immersions turns out to work over \mathbb{Z}, so that 𝒲im(M)\mathcal{W}_{im}(M) is an AA_{\infty}-category over \mathbb{Z} in the usual sense. We also extend this theory to cylindrical Lagrangian immersions with clean self-intersections in section 5.

The ordinary wrapped Fukaya category 𝒲(M)\mathcal{W}(M) can be embedded into 𝒲im(M)\mathcal{W}_{im}(M) as a full sub-category. Also, the AA_{\infty}-functors Φ\Phi_{\mathcal{L}} and Φ\Phi can be extended to the category of modules over the immersed wrapped Fukaya category

(1.3) Φ:𝒲(M)𝒲im(N)lmod,\Phi_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}_{im}(N)^{l-mod},

and

(1.4) Φ:𝒲(M×N)func(𝒲(M),𝒲im(N)lmod).\Phi:\mathcal{W}(M^{-}\times N)\to func(\mathcal{W}(M),\mathcal{W}_{im}(N)^{l-mod}).
Theorem 1.2.

Let M×N\mathcal{L}\subset M^{-}\times N be an admissible Lagrangian correspondence between Liouville manifolds MM and NN, such that the projection N\mathcal{L}\to N is proper. Then under some further generic geometric conditions, namely Assumption 7.10, we have:

  1. (i)

    For every object LOb𝒲(M)L\in Ob\mathcal{W}(M), there is a curved AA_{\infty}-algebra associated to the geometric composition LL\circ\mathcal{L}, defined in terms of wrapped Floer theory for Lagrangian immersions.

  2. (ii)

    The geometric composition LL\circ\mathcal{L} is always unobstructed, with a canonical choice of bounding cochain bb for it. Thus (L,b)(L\circ\mathcal{L},b) becomes an object of 𝒲im(N)\mathcal{W}_{im}(N). This bb is unique such that the next condition is satisfied.

  3. (iii)

    There is a natural AA_{\infty}-functor

    (1.5) Θ:𝒲(M)𝒲im(N),\Theta_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}_{im}(N),

    which represents Φ\Phi_{\mathcal{L}}. On the level of objects, it sends any Lagrangian submanifold LOb𝒲(M)L\in Ob\mathcal{W}(M) to the pair (L,b)Ob𝒲im(N)(L\circ\mathcal{L},b)\in Ob\mathcal{W}_{im}(N).

We expect that the above AA_{\infty}-functor extends over 𝒲im(M)\mathcal{W}_{im}(M) in a canonical and unique way. However, as that is not one of the main subjects of this paper and involves quite a lot of technicality, we will not try to give a proof.

In a more functorial form, the assignment of AA_{\infty}-functors to Lagrangian correspondences is functorial in the wrapped Fukaya category of the product manifold M×NM^{-}\times N:

Theorem 1.3.

Let 𝒜(M×N)\mathcal{A}(M^{-}\times N) be the full AA_{\infty}-subcategory whose objects are Lagrangian correspondences \mathcal{L} from MM to NN such that the projection N\mathcal{L}\to N is proper, which further satisfies Assumption 7.10. Then there is a canonical AA_{\infty}-functor

(1.6) Θ:𝒜(M×N)func(𝒲(M),𝒲im(N)),\Theta:\mathcal{A}(M^{-}\times N)\to func(\mathcal{W}(M),\mathcal{W}_{im}(N)),

such that

  1. (i)

    Θ\Theta represents Φ\Phi;

  2. (ii)

    Θ()=Θ\Theta(\mathcal{L})=\Theta_{\mathcal{L}} for every Ob𝒜(M×N)\mathcal{L}\in Ob\mathcal{A}(M^{-}\times N).

1.3. Some applications

As a particular application of the construction of functors, we present a well-known and expected (but not fully established) Künneth formula for wrapped Fukaya categories, which relate 𝒲(M×N)\mathcal{W}(M\times N) to the AA_{\infty}-tensor product 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N) in an appropriate sense. We also show that under the condition that both 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N) have finite collections of split-generators, there is a quasi-equivalence between 𝒲(M×N)\mathcal{W}(M\times N) and 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N). These results will be discussed in a more formal way in subsection 7.9.

An especially important instance of Lagrangian correspondence comes from Liouville sub-domains U0M0U_{0}\subset M_{0}: the graph of the natural inclusion can be completed to a Lagrangian correspondence between UU and MM. This is called the graph correspondence, denoted by Γ\Gamma. As an admissible Lagrangian correspondence from MM to UU, it satisfies the hypothesis of Theorem 1.2. Thus we obtain an AA_{\infty}-functor

(1.7) ΘΓ:𝒲(M)𝒲im(U).\Theta_{\Gamma}:\mathcal{W}(M)\to\mathcal{W}_{im}(U).

There is a full sub-category on which this functor takes a simpler form. First, there is the full sub-category (M)\mathcal{B}(M) consisting of exact cylindrical Lagrangian submanifolds LL whose primitive ff is locally constant near both M\partial M and U\partial U (Assumption 8.3). Then, we may further consider a full sub-category 0(M)\mathcal{B}_{0}(M) of (M)\mathcal{B}(M) whose objects satisfy an additional geometric condition (Assumption 8.4). Restricted to this sub-category, the functor ΘΓ\Theta_{\Gamma} induces a functor

ΘΓ:0(M)𝒲(N),\Theta_{\Gamma}:\mathcal{B}_{0}(M)\to\mathcal{W}(N),

whose image lies in the ordinary wrapped Fukaya category consisting of properly embedded exact cylindrical Lagrangian submanifolds, with zero bounding cochains. On the other hand, recall that the Viterbo restriction functor is defined on this sub-category (M)\mathcal{B}(M). More detailed definitions will be given in section 8. We shall see that their linear terms agree, when the Viterbo restriction functor is further restricted to 0(M)\mathcal{B}_{0}(M).

Theorem 1.4.

The AA_{\infty}-functor ΘΓ\Theta_{\Gamma} associated to the graph correspondence ΓM×U\Gamma\subset M^{-}\times U of the Liouville sub-domain restricts to an AA_{\infty}-functor on the full sub-category 0(M)\mathcal{B}_{0}(M)

(1.8) ΘΓ:0(M)𝒲(U).\Theta_{\Gamma}:\mathcal{B}_{0}(M)\to\mathcal{W}(U).

The linear term ΘΓ1\Theta_{\Gamma}^{1} is chain homotopic to the linear term r1r^{1} of the Viterbo restriction functor rr.

It is expected that ΘΓ\Theta_{\Gamma} and rr are in fact homotopic as AA_{\infty}-functors, when restricted to the full sub-category 0(M)\mathcal{B}_{0}(M). Our current method of proof needs improvement in order to prove this more general statement. Moreover, they should be indeed homotopic on the bigger sub-category (M)\mathcal{B}(M) in appropriate sense, if we can manage to show that the geometric composition with the graph correspondence is Floer-theoretically equivalent to the actual restriction of Lagrangian submanifolds. Such points are to be discussed in [Gao].

Going back to the definition of the Viterbo restriction functor, we recall that it is only defined for those Lagrangian submanifolds which satisfy Assumption 8.3. However, the restriction functor ΘΓ\Theta_{\Gamma} is defined on the whole wrapped Fukaya category, though the image of a Lagrangian submanifold is not necessarily simply its restriction to the sub-domain. However, it is conjectured that the Viterbo restriction functor can be extended to the whole wrapped Fukaya category via some kind of deformation theory, as stated in Conjecture 8.26. It is not completely known but a very interesting question to ask in what cases the extension agrees with the functor ΘΓ\Theta_{\Gamma}. This is a question for future research.

The structure of this paper is as follows. Sections 2 to 5 provide fundamental materials for the theory to work: section 2 reviews basic homological algebra regarding AA_{\infty}-modules and presents a version of AA_{\infty}-direct limit; section 3 recalls the definition of the wrapped Fukaya category and introduces a new class of objects necessary to establish representability of the functors; section 4 defines the immersed wrapped Fukaya category and section 5 generalizes the story to Lagrangian immersions with clean self-intersections. Section 6 studies Floer theory on product manifolds and proves the two models of wrapped Fukaya categories are equivalent. Section 7 formally sets up the framework of functoriality in Lagrangian correspondences and proves representability, with a discussion on the Künneth formula at the end. Section 8 concerns sub-domains and how the Viterbo restriction functor is related to our framework.

Acknowledgment.

This research is part of a project on studying functorial properties of wrapped Fukaya categories, which the author carries out in the graduate school at Stony Brook University. The author is grateful to his advisor Kenji Fukaya for his inspiration on this project, as well as for numerous helpful suggestions and ideas that author learned from him. The author would also like to express thanks to Mohammed Abouzaid for his insight that the theory should work over the integers, as well as to Mark McLean for his explanation and discussion on various constructions and structures related to wrapped Floer cohomology and the Viterbo functor.

2. Homological algebra preliminaries

2.1. AA_{\infty}-modules and bimodules

Most of the results about AA_{\infty}-categories in this section are well-known, which are included for the purpose of fixing notations and conventions for AA_{\infty}-categories. Let kk be a field. We shall work with AA_{\infty}-categories that are linear over kk. Let ChCh be the dg-category of chain complexes over kk, regarded as an AA_{\infty}-category whose higher order operations mk,k3m^{k},k\geq 3 are all zero. Here chain complexes are not necessarily bounded.

By definition, a non-unital left AA_{\infty}-module over 𝒜\mathcal{A} is a non-unital AA_{\infty}-functor 𝒜Chop\mathcal{A}\to Ch^{op}, where ChopCh^{op} is the opposite category of ChCh. All left AA_{\infty}-modules over 𝒜\mathcal{A} also form an AA_{\infty}-category (in fact a dg-category), 𝒜nulmod=nufunc(𝒜,Chop)\mathcal{A}^{nu-l-mod}=nu-func(\mathcal{A},Ch^{op}).

There are also non-unital right AA_{\infty}-modules over 𝒜\mathcal{A}, which are functors from 𝒜op\mathcal{A}^{op} to ChCh. These also form a dg-category, 𝒜nurmod=func(𝒜op,Ch)\mathcal{A}^{nu-r-mod}=func(\mathcal{A}^{op},Ch).

When 𝒜\mathcal{A} comes with strict units, homotopy units or cohomological units, there are also unital versions of left and right AA_{\infty}-modules:

𝒜lmod=func(𝒜,Chop),\mathcal{A}^{l-mod}=func(\mathcal{A},Ch^{op}),
𝒜rmod=func(𝒜op,Ch),\mathcal{A}^{r-mod}=func(\mathcal{A}^{op},Ch),

as unital AA_{\infty}-functors.

2.2. Representable modules

First, we recall the non-unital version of the Yoneda embedding. The left Yoneda functor is a non-unital AA_{\infty}-functor

(2.1) 𝔶l:𝒜𝒜numod\mathfrak{y}_{l}:\mathcal{A}\to\mathcal{A}^{nu-mod}

which sends an object YOb𝒜Y\in Ob\mathcal{A} to its left Yoneda module 𝒴l𝒜numod\mathcal{Y}^{l}\in\mathcal{A}^{nu-mod}, which is defined as follows. For objects XOb𝒜X\in Ob\mathcal{A},

(2.2) 𝒴l(X)=hom𝒜(Y,X),\mathcal{Y}^{l}(X)=hom_{\mathcal{A}}(Y,X),

and the module structure is given by the AA_{\infty}-structure maps of 𝒜\mathcal{A}:

(2.3) n𝒴ld:hom𝒜(Xd2,Xd1)hom𝒜(X0,X1)𝒴r(X0)𝒴r(Xd1),n_{\mathcal{Y}^{l}}^{d}:\hom_{\mathcal{A}}(X_{d-2},X_{d-1})\otimes\cdots\otimes\hom_{\mathcal{A}}(X_{0},X_{1})\otimes\mathcal{Y}^{r}(X_{0})\to\mathcal{Y}^{r}(X_{d-1}),
(2.4) n𝒴ld(ad1,,a1,b)=m𝒜d+1(ad1,,a1,b).n_{\mathcal{Y}^{l}}^{d}(a_{d-1},\cdots,a_{1},b)=m^{d+1}_{\mathcal{A}}(a_{d-1},\cdots,a_{1},b).

On the level of morphisms, 𝔶l1\mathfrak{y}_{l}^{1} assigns to a morphism chom𝒜(Y0,Y1)c\in hom_{\mathcal{A}}(Y_{0},Y_{1}) a pre-module homomorphism:

(𝔶l1(c))d:hom𝒜(Xd2,Xd1)hom𝒜(X0,X1)𝒴1l(X0)𝒴0l(Xd1),(\mathfrak{y}_{l}^{1}(c))^{d}:\hom_{\mathcal{A}}(X_{d-2},X_{d-1})\otimes\cdots\otimes\hom_{\mathcal{A}}(X_{0},X_{1})\otimes\mathcal{Y}_{1}^{l}(X_{0})\to\mathcal{Y}_{0}^{l}(X_{d-1}),
(2.5) (𝔶l1(c))d(ad1,,a1,b)=m𝒜d+1(ad1,,a1,b,c).(\mathfrak{y}_{l}^{1}(c))^{d}(a_{d-1},\cdots,a_{1},b)=m^{d+1}_{\mathcal{A}}(a_{d-1},\cdots,a_{1},b,c).

Higher order terms 𝔶lk\mathfrak{y}_{l}^{k} are defined in by analogous formulas:

(𝔶lk(ck,,c1))d:hom𝒜(Xd2,Xd1)hom𝒜(X0,X1)𝒴kl(X0)𝒴0l(Xd1),(\mathfrak{y}_{l}^{k}(c_{k},\cdots,c_{1}))^{d}:\hom_{\mathcal{A}}(X_{d-2},X_{d-1})\otimes\cdots\otimes\hom_{\mathcal{A}}(X_{0},X_{1})\otimes\mathcal{Y}_{k}^{l}(X_{0})\to\mathcal{Y}_{0}^{l}(X_{d-1}),
(2.6) (𝔶lk(ck,,c1))d(ad1,,a1,b)=m𝒜d+k(ad1,,a1,b,ck,,c1).(\mathfrak{y}_{l}^{k}(c_{k},\cdots,c_{1}))^{d}(a_{d-1},\cdots,a_{1},b)=m^{d+k}_{\mathcal{A}}(a_{d-1},\cdots,a_{1},b,c_{k},\cdots,c_{1}).

In case 𝒜\mathcal{A} is cohomologically unital, the image of the Yoneda functor lies in the AA_{\infty}-subcategory of 𝒜nulmod\mathcal{A}^{nu-l-mod} consisting of c-unital right 𝒜\mathcal{A}_{\infty}-modules over 𝒜\mathcal{A}. We denote this AA_{\infty}-subcategory by 𝒜lmod\mathcal{A}^{l-mod}.

An important notion is the representability of an AA_{\infty}-functor, in the sense of [Fuk02]. We briefly recall it here.

Definition 2.1.

A left AA_{\infty}-module over 𝒜\mathcal{A}, namely an AA_{\infty}-functor :𝒜Chop\mathcal{M}:\mathcal{A}\to Ch^{op} is said to be representable, if it is homotopic (or equivalently quasi-isomorphic) to a left Yoneda module 𝒴l=𝔶l(Y)\mathcal{Y}^{l}=\mathfrak{y}_{l}(Y) for some object YY of 𝒜\mathcal{A}, as AA_{\infty}-functors 𝒜Chop\mathcal{A}\to Ch^{op}.

There is also a right Yoneda functor

(2.7) 𝔶r:𝒜𝒜nurmod,\mathfrak{y}_{r}:\mathcal{A}\to\mathcal{A}^{nu-r-mod},

whose c-unital version becomes

(2.8) 𝔶r:𝒜𝒜rmod.\mathfrak{y}_{r}:\mathcal{A}\to\mathcal{A}^{r-mod}.

Similarly, a right AA_{\infty}-module over 𝒜\mathcal{A} is said to be representable, if it is homotopic to a right Yoneda module 𝒴r=𝔶r(Y)\mathcal{Y}^{r}=\mathfrak{y}_{r}(Y).

We will see in the following subsection that the notion of representability does not have ambiguity, up to quasi-isomorphism. That is, the representative is unique up to quasi-isomorphism.

2.3. Yoneda lemma

The key point related to Yoneda embedding we want to emphasize here is that whenever 𝒜\mathcal{A} is c-unital, the Yoneda embedding is cohomologically fully faithful, and therefore is an AA_{\infty}-homotopy equivalence to its image. This was first proved in [Fuk02] under the assumption that 𝒜\mathcal{A} is strictly unital, with a new proof given in [Sei08] in case 𝒜\mathcal{A} is only cohomologically unital. To see this, let \mathcal{M} be any c-unital left AA_{\infty}-module over 𝒜\mathcal{A} and consider the following cochain map

(2.9) λ:(Y)hom𝒜mod(𝒴l,)\lambda:\mathcal{M}(Y)\to hom_{\mathcal{A}^{mod}}(\mathcal{Y}^{l},\mathcal{M})
(2.10) λ(c)d(ad1,,a1,b)=nd+1(ad1,,a1,b,c).\lambda(c)^{d}(a_{d-1},\cdots,a_{1},b)=n_{\mathcal{M}}^{d+1}(a_{d-1},\cdots,a_{1},b,c).

Note in fact this definition also makes sense for general non-unital 𝒜\mathcal{A}_{\infty}-modules, but we will only emphasize its importance in the c-unital case.

Lemma 2.2.

Then the above cochain map λ\lambda is a quasi-isomorphism, for any object YY of 𝒜\mathcal{A}.

Proof.

The mapping cone of the cochain map λ\lambda is the following cochain complex:

(2.11) ((Y)hom𝒜lmod(𝒴l,)[1],(n10λm𝒜lmod1))[1].(\mathcal{M}(Y)\oplus\hom_{\mathcal{A}^{l-mod}}(\mathcal{Y}^{l},\mathcal{M})[-1],\begin{pmatrix}n^{1}_{\mathcal{M}}&0\\ \lambda&-m^{1}_{\mathcal{A}^{l-mod}}\end{pmatrix})[1].

Define a filtration on this cochain complex by first taking the subcomplex hom𝒜lmod(𝒴l,)[1]\hom_{\mathcal{A}^{l-mod}}(\mathcal{Y}^{l},\mathcal{M})[-1], then filtering it by its natural length filtration. Denote A=H(𝒜),M=H()A=H(\mathcal{A}),M=H(\mathcal{M}). Associated to this filtration there is a spectral sequence whose E1E_{1}-page is

(2.12) E1rs={Ms(Y), if r=0,X0,,Xr1HomRs(HomA(Xr2,Xr1)HomA(X0,X1)HomA(Y,X0),M(Xr1)), if r>0.E_{1}^{rs}=\begin{cases}M^{s}(Y),\text{ if }r=0,\\ \begin{split}\prod_{X_{0},\cdots,X_{r-1}}&Hom^{s}_{R}(Hom_{A}(X_{r-2},X_{r-1})\otimes\cdots\otimes Hom_{A}(X_{0},X_{1})\\ &\otimes Hom_{A}(Y,X_{0}),M(X_{r-1})),\text{ if }r>0.\end{split}\end{cases}

The differential d=d1rs:E1rsE1r+1,sd=d_{1}^{rs}:E_{1}^{rs}\to E_{1}^{r+1,s} is given by d(c)(b)=(1)|b|(c)d(c)(b)=(-1)^{|b|}(c) if r=0r=0, and

(2.13) d(t)(ar,,a1,b)=(1)|b|+rt(ar,,a1)b+(1)|b|t(ar,ar1,,a1b)+n(1)Δt(ar,,an+2an+1,,a1,b),\begin{split}d(t)(a_{r},\cdots,a_{1},b)&=(-1)^{|b|+*_{r}}t(a_{r},\cdots,a_{1})b+(-1)^{|b|}t(a_{r},a_{r-1},\cdots,a_{1}b)\\ &+\sum_{n}(-1)^{\Delta}t(a_{r},\cdots,a_{n+2}a_{n+1},\cdots,a_{1},b),\end{split}

where r=|a1|++|ar|r*_{r}=|a_{1}|+\cdots+|a_{r}|-r, and Δ=|an+2|++|ar|+|b|+n+1r\Delta=|a_{n+2}|+\cdots+|a_{r}|+|b|+n+1-r. In the above expression, aiaja_{i}a_{j} is the induced composition in the cohomology category AA, which is associative, and abab is the induced AA-module structure on MM from the structure of left AA_{\infty}-module on \mathcal{M}. This is the standard bar resolution of the cochain complex (2.11), which in the presence of cohomological unit of 𝒜\mathcal{A}, admits a contracting homotopy

(2.14) κ:E1r+1,sE1rs\kappa:E_{1}^{r+1,s}\to E_{1}^{rs}
(2.15) κ(t)(ar1,,a1,b)=t(ar1,,a1,b,eY),\kappa(t)(a_{r-1},\cdots,a_{1},b)=t(a_{r-1},\cdots,a_{1},b,e_{Y}),

where eYe_{Y} is a cochain in hom𝒜(Y,Y)\hom_{\mathcal{A}}(Y,Y) representing the identity morphism in H(hom𝒜(Y,Y))H(\hom_{\mathcal{A}}(Y,Y)). This shows that the spectral sequence degenerates at the E1E_{1}-page, which imples that the cochain complex (2.11) is acyclic, and therefore the cochain map λ\lambda is a quasi-isomorphism. ∎

Corollary 2.3.

If 𝒜\mathcal{A} is c-unital, then the (c-unital) left Yoneda functor

𝔶l:𝒜𝒜lmod\mathfrak{y}_{l}:\mathcal{A}\to\mathcal{A}^{l-mod}

is cohomologically fully faithful.

There is a parallel discussion for right Yoneda functors. That is, the right Yoneda functor

𝔶r:𝒜𝒜rmod\mathfrak{y}_{r}:\mathcal{A}\to\mathcal{A}^{r-mod}

is cohomologically fully faithful.

For this reason, we also call the Yoneda functor the Yoneda embedding. And there is no ambiguity for the notion of representability. This allows us to make the following definition.

Definition 2.4.

The AA_{\infty}-category 𝒜replmod\mathcal{A}^{rep-l-mod} (resp. 𝒜reprmod\mathcal{A}^{rep-r-mod}) of representable left (resp. right) AA_{\infty}-modules over 𝒜\mathcal{A}, is the full AA_{\infty}-subcategory of 𝒜lmod\mathcal{A}^{l-mod} whose objects are representable left (resp. right) AA_{\infty}-modules. Equivalently, it is the image of the left (resp. right) Yoneda embedding 𝔶l\mathfrak{y}_{l} (resp. 𝔶r\mathfrak{y}_{r}) inside 𝒜lmod\mathcal{A}^{l-mod} (resp. 𝒜rmod\mathcal{A}^{r-mod}).

2.4. Bimodules and functors

Let us relate the story of representable AA_{\infty}-modules to that of AA_{\infty}-functors. We shall be considering AA_{\infty}-functors of the following kind

(2.16) m:𝒜lmod.\mathcal{F}_{m}:\mathcal{A}\to\mathcal{B}^{l-mod}.

Such an AA_{\infty}-functor is called a module-valued functor, which says that it suffices to verify representability on objects.

Definition 2.5.

A module-valued functor

m:𝒜lmod\mathcal{F}_{m}:\mathcal{A}\to\mathcal{B}^{l-mod}

is said to be representable, if there exists an AA_{\infty}-functor

(2.17) :𝒜,\mathcal{F}:\mathcal{A}\to\mathcal{B},

such that 𝔶l\mathfrak{y}_{l}\circ\mathcal{F} is homotopic to m\mathcal{F}_{m} as AA_{\infty}-functors.

We have the following concrete criterion for representability of a module-valued functor.

Lemma 2.6.

A module-valued functor

m:𝒜lmod\mathcal{F}_{m}:\mathcal{A}\to\mathcal{B}^{l-mod}

is representable, if and only if m(X)\mathcal{F}_{m}(X) is a representable left \mathcal{B}-module of any object XOb𝒜X\in Ob\mathcal{A}.

Proof.

The ”only if” part is obvious by definition.

Now consider the ”if” part. Since for every XOb𝒜X\in Ob\mathcal{A}, the AA_{\infty}-module m(X)\mathcal{F}_{m}(X) over \mathcal{B} is representable, there exists an object Y=Y(X)Y=Y(X) and an AA_{\infty}-module homomorphism

(2.18) TX:m(X)𝒴l,T_{X}:\mathcal{F}_{m}(X)\to\mathcal{Y}^{l},

which is a quasi-isomorphism of AA_{\infty}-modules. We make a choice of a homotopy inverse KXK_{X} of TXT_{X} for each XX.

Let us recall what it means for m\mathcal{F}_{m} to be representable. There should exist an AA_{\infty}-functor

:𝒜,\mathcal{F}:\mathcal{A}\to\mathcal{B},

as well as an AA_{\infty}-natural transformation of degree 11

Thomfunc(𝒜,lmod)(m,𝔶l),T\in\hom_{func(\mathcal{A},\mathcal{B}^{l-mod})}(\mathcal{F}_{m},\mathfrak{y}_{l}\circ\mathcal{F}),

such that TT is a homotopy between m\mathcal{F}_{m} and 𝔶l\mathfrak{y}_{l}\circ\mathcal{F}.

To define these, we need to pick a homotopy inverse of the Yoneda embedding

𝔶l:lmod,\mathfrak{y}_{l}:\mathcal{B}\to\mathcal{B}^{l-mod},

when restricted to the image. That is, if we regard the Yoneda embedding as an AA_{\infty}-functor

𝔶l:replmod,\mathfrak{y}_{l}:\mathcal{B}\to\mathcal{B}^{rep-l-mod},

it is a quasi-isomorphism, so that we can choose a homotopy inverse,

(2.19) λ:replmod.\lambda_{\mathcal{B}}:\mathcal{B}^{rep-l-mod}\to\mathcal{B}.

We define \mathcal{F} as follows. On objects, (X)=Y=Y(X)\mathcal{F}(X)=Y=Y(X). For X0,X1Ob𝒜X_{0},X_{1}\in Ob\mathcal{A}, we define

(2.20) 1:hom𝒜(X0,X1)hom(Y0,Y1)\mathcal{F}^{1}:\hom_{\mathcal{A}}(X_{0},X_{1})\to\hom_{\mathcal{B}}(Y_{0},Y_{1})

as follows. For each ahom𝒜(X0,X1)a\in\hom_{\mathcal{A}}(X_{0},X_{1}), the morphism

m1(a)homlmod(m(X0),m(X1))\mathcal{F}_{m}^{1}(a)\in\hom_{\mathcal{B}^{l-mod}}(\mathcal{F}_{m}(X_{0}),\mathcal{F}_{m}(X_{1}))

is a AA_{\infty}-pre-module homomorphism, which can be composed with KX0K_{X_{0}} and TX1T_{X_{1}} in the dg-category lmod\mathcal{B}^{l-mod} by the structure map mlmod2m^{2}_{\mathcal{B}^{l-mod}} to get an AA_{\infty}-pre-module homomorphism

(2.21) 𝒢m(a)=mlmod2(TX1,mlmod2(m(a),KX0))homlmod(𝒴0l,𝒴1l).\mathcal{G}_{m}(a)=m^{2}_{\mathcal{B}^{l-mod}}(T_{X_{1}},m^{2}_{\mathcal{B}^{l-mod}}(\mathcal{F}_{m}(a),K_{X_{0}}))\in\hom_{\mathcal{B}^{l-mod}}(\mathcal{Y}_{0}^{l},\mathcal{Y}_{1}^{l}).

Since mlmod2m^{2}_{\mathcal{B}^{l-mod}} is associative, there is no ambiguity of this composition, and therefore this is well-defined in a unique way, once we fix a choice of a homotopy inverse KXK_{X} of TXT_{X} for every XOb𝒜X\in Ob\mathcal{A}, as this homotopy inverse is indepedent of aa. Note that

𝒢m(a)homlmod(𝒴0l,𝒴1l)\mathcal{G}_{m}(a)\in\hom_{\mathcal{B}^{l-mod}}(\mathcal{Y}_{0}^{l},\mathcal{Y}_{1}^{l})

is an AA_{\infty}-pre-module homomorphism between left Yoneda modules, which lie in the sub-category replmod\mathcal{B}^{rep-l-mod}. Thus we can apply λ\lambda_{\mathcal{B}} to 𝒢m(a)\mathcal{G}_{m}(a) to obtain

(2.22) 1(a)=λ1(𝒢m(a)).\mathcal{F}^{1}(a)=\lambda_{\mathcal{B}}^{1}(\mathcal{G}_{m}(a)).

This defines \mathcal{F} on morphisms.

For higher order terms k\mathcal{F}^{k}, we again follow the same strategy. That is, we define

(2.23) k:hom𝒜(Xk1,Xk)hom𝒜(X0,X1)hom(Y0,Yd)\mathcal{F}^{k}:\hom_{\mathcal{A}}(X_{k-1},X_{k})\otimes\cdots\otimes\hom_{\mathcal{A}}(X_{0},X_{1})\to\hom_{\mathcal{B}}(Y_{0},Y_{d})

to be the image of the composition of mk\mathcal{F}_{m}^{k} with KX0K_{X_{0}} and TXkT_{X_{k}} under λ\lambda_{\mathcal{B}}. That is, for aihom𝒜(Xi1,Xi)a_{i}\in\hom_{\mathcal{A}}(X_{i-1},X_{i}), we define

(2.24) k(ak,,a1)=λ1(mlmod2(TXk,mlmod2(mk(ak,,a1),KX0))).\mathcal{F}^{k}(a_{k},\cdots,a_{1})=\lambda_{\mathcal{B}}^{1}(m^{2}_{\mathcal{B}^{l-mod}}(T_{X_{k}},m^{2}_{\mathcal{B}^{l-mod}}(\mathcal{F}_{m}^{k}(a_{k},\cdots,a_{1}),K_{X_{0}}))).

It is a straightforward computation to check that ={k}k=1\mathcal{F}=\{\mathcal{F}^{k}\}_{k=1}^{\infty} satisfies the AA_{\infty}-functor equations.

Then we need to define the homotopy between 𝔶l\mathfrak{y}_{l}\circ\mathcal{F} and m\mathcal{F}_{m}. But this is clear: we simply take

A natural source of module-valued functors is given by AA_{\infty}-bimodules. Let 𝒫\mathcal{P} be a left-\mathcal{B} right-𝒜\mathcal{A} AA_{\infty}-bimodule, or simply called an AA_{\infty}-bimodule over (𝒜,)(\mathcal{A},\mathcal{B}). This notation is slightly misleading as we write 𝒜\mathcal{A} on the left and \mathcal{B} on the right, but we shall keep this convention as it is fitted into Floer theory which will be discussed later on. From 𝒫\mathcal{P} we can define an AA_{\infty}-functor

(2.25) 𝒫:𝒜lmod\mathcal{F}_{\mathcal{P}}:\mathcal{A}\to\mathcal{B}^{l-mod}

as follows. For each object XOb𝒜X\in Ob\mathcal{A}, we set

𝒫(X)=𝒫(X,).\mathcal{F}_{\mathcal{P}}(X)=\mathcal{P}(X,\cdot).

That is, 𝒫(X)\mathcal{F}_{\mathcal{P}}(X) is the left-\mathcal{B} module that takes value 𝒫(X,Y)\mathcal{P}(X,Y) for each object YObY\in Ob\mathcal{B}, and has module structure maps

(2.26) n𝒫(X)k:hom(Yk1,Yk)hom(Y0,Y1)𝒫(X,Y0)𝒫(X,Yk),n^{k}_{\mathcal{F}_{\mathcal{P}}(X)}:\hom_{\mathcal{B}}(Y_{k-1},Y_{k})\otimes\hom_{\mathcal{B}}(Y_{0},Y_{1})\otimes\mathcal{P}(X,Y_{0})\to\mathcal{P}(X,Y_{k}),
(2.27) n𝒫(X)k(bk,,b1,p)=n𝒫k,0(bk,,b1,p),n^{k}_{\mathcal{F}_{\mathcal{P}}(X)}(b_{k},\cdots,b_{1},p)=n^{k,0}_{\mathcal{P}}(b_{k},\cdots,b_{1},p),

where n𝒫k,ln^{k,l}_{\mathcal{P}} are the AA_{\infty}-bimodule structure maps of 𝒫\mathcal{P}. Next we define the action of 𝒫\mathcal{F}_{\mathcal{P}} on morphism spaces

(2.28) 𝒫l:hom𝒜(Xl1,Xl)hom𝒜(X0,X1)homlmod(𝒫(Xl),𝒫(X0))\mathcal{F}_{\mathcal{P}}^{l}:\hom_{\mathcal{A}}(X_{l-1},X_{l})\otimes\cdots\otimes\hom_{\mathcal{A}}(X_{0},X_{1})\to\hom_{\mathcal{B}^{l-mod}}(\mathcal{F}_{\mathcal{P}}(X_{l}),\mathcal{F}_{\mathcal{P}}(X_{0}))

by the formula

(2.29) (𝒫l(al,,a1))k(bk,,b1,p)=n𝒫k,l(bk,,b1,p,al,,a1).(\mathcal{F}_{\mathcal{P}}^{l}(a_{l},\cdots,a_{1}))^{k}(b_{k},\cdots,b_{1},p)=n^{k,l}_{\mathcal{P}}(b_{k},\cdots,b_{1},p,a_{l},\cdots,a_{1}).

By the AA_{\infty}-bimodule equations for n𝒫k,ln^{k,l}_{\mathcal{P}}, it is straightforward to verify:

Lemma 2.7.

The multilinear maps {𝒫l}l=1\{\mathcal{F}_{\mathcal{P}}^{l}\}_{l=1}^{\infty} form an AA_{\infty}-functor

𝒫:𝒜lmod.\mathcal{F}_{\mathcal{P}}:\mathcal{A}\to\mathcal{B}^{l-mod}.

All AA_{\infty}-bimodules over (𝒜,)(\mathcal{A},\mathcal{B}) form an AA_{\infty}-category, denoted by (𝒜,)bimod(\mathcal{A},\mathcal{B})^{bimod}. The above construction of a module-valued functor 𝒫\mathcal{F}_{\mathcal{P}} associated to an AA_{\infty}-bimodule 𝒫\mathcal{P} can be made on the level of the whole category (𝒜,)bimod(\mathcal{A},\mathcal{B})^{bimod}.

Proposition 2.8.

There is a canonical AA_{\infty}-functor

(2.30) :(𝒜,)bimodfunc(𝒜,lmod),\mathcal{F}:(\mathcal{A},\mathcal{B})^{bimod}\to func(\mathcal{A},\mathcal{B}^{l-mod}),

such that (𝒫)=𝒫\mathcal{F}(\mathcal{P})=\mathcal{F}_{\mathcal{P}} for every object 𝒫Ob(𝒜,)bimod\mathcal{P}\in Ob(\mathcal{A},\mathcal{B})^{bimod}.

In general, it is not possible to expect (X)\mathcal{F}(X) is representable for every XOb𝒜X\in Ob\mathcal{A}. But we might say that it is representable for some of the objects. This leads to the following definition.

Definition 2.9.

Let 𝒜0\mathcal{A}_{0} be a full AA_{\infty}-subcategory of 𝒜\mathcal{A}. We say that

:(𝒜,)bimodfunc(𝒜,lmod)\mathcal{F}:(\mathcal{A},\mathcal{B})^{bimod}\to func(\mathcal{A},\mathcal{B}^{l-mod})

is representable on 𝒜0\mathcal{A}_{0}, if (X)\mathcal{F}(X) is a representable left \mathcal{B}-module for every XOb𝒜0X\in Ob\mathcal{A}_{0}.

By the criterion stated in Lemma 2.6, this definition makes sense.

2.5. Cyclic element and bounding cochain

In this subsection, we introduce the notion of a cyclic element, which is key to proving the existence and uniqueness of a bounding cochain under certain assumptions. One origianl formulation due to Fukaya (Proposition 3.5 of [Fuk15]), works with filtered AA_{\infty}-algebras and filtered AA_{\infty}-modules over the Novikov ring. In our case, we consider curved AA_{\infty}-algebras and AA_{\infty}-modules over \mathbb{Z}. There is in general no intrinsic way of making the method work for arbitrary curved AA_{\infty}-algebras and AA_{\infty}-modules, so additional structures are required, to be explained below.

Let (C,mk)(C,m^{k}) be a curved AA_{\infty}-algebra over \mathbb{Z} (m00m^{0}\neq 0), where CC is a free \mathbb{Z}-module (of finite or infinite rank), and (D,nk)(D,n^{k}) a left AA_{\infty}-module over (C,mk)(C,m^{k}), where DD is also a free \mathbb{Z}-module. We need an additional condition, similar to the gappedness condition for filtered AA_{\infty}-algebras and filtered AA_{\infty}-modules. Although we work with usual AA_{\infty}-algebras and AA_{\infty}-modules, we still want to have some inductive structures, similar to those for filtered AA_{\infty}-algebras and filtered AA_{\infty}-modules. For us, the notion we need is a filtration which satisfies certain analogous conditions.

Definition 2.10.
  1. (i)

    A filtration FCF_{C} on (C,mk)(C,m^{k}) is an \mathbb{R}-filtration such that

    FCλCFCλC, if λ<λ,F_{C}^{\lambda^{\prime}}C\subset F_{C}^{\lambda}C,\text{ if }\lambda<\lambda^{\prime},

    and furthermore,

    (2.31) mk(FCλkCFCλ1C)FCj=1kλjC.m^{k}(F_{C}^{\lambda_{k}}C\otimes\cdots F_{C}^{\lambda_{1}}C)\subset F_{C}^{\sum_{j=1}^{k}\lambda_{j}}C.
  2. (ii)

    Given a filtration FCF_{C} on (C,mk)(C,m^{k}), a compatible filtration FDF_{D} on (D,nk)(D,n^{k}) is an \mathbb{R}-filtration such that

    FDλDFDλD, if λ<λ,F_{D}^{\lambda^{\prime}}D\subset F_{D}^{\lambda}D,\text{ if }\lambda<\lambda^{\prime},

    and furthermore,

    (2.32) nk(FCλkCFCλ1CFDλ)FDλ+j=1kλjD.n^{k}(F_{C}^{\lambda_{k}}C\otimes F_{C}^{\lambda_{1}}C\otimes F_{D}^{\lambda^{\prime}})\subset F_{D}^{\lambda^{\prime}+\sum_{j=1}^{k}\lambda_{j}}D.
Definition 2.11.

A filtration FCF_{C} is said to be discrete, if there is a discrete subset ΛC\Lambda_{C} of \mathbb{R} such that

(2.33) FCλC=FCλC, if [λ,λ]ΛC=,F_{C}^{\lambda}C=F_{C}^{\lambda^{\prime}}C,\text{ if }[\lambda,\lambda^{\prime}]\cap\Lambda_{C}=\varnothing,

for any λ<λ\lambda<\lambda^{\prime}. A similar definition applies to a compatible filtration FDF_{D}.

A discrete compatible filtration FDF_{D} is said to be strictly compatible, if ΛC=ΛD\Lambda_{C}=\Lambda_{D}.

In order to deform a curved AA_{\infty}-algebra to a non-curved AA_{\infty}-algebra, we need the notion of a bounding cochain.

Definition 2.12.

A bounding cochain for the curved AA_{\infty}-algebra (C,mk)(C,m^{k}) is an element bb, such that the inhomogeneous Maurer-Cartan equation is satisfied

(2.34) k=0mk(b,,b)=0,\sum_{k=0}^{\infty}m^{k}(b,\cdots,b)=0,

where the sum stops at a finite stage. That is, there exists KK such that for all k>0k>0, mk(b,,b)=0m^{k}(b,\cdots,b)=0 and

(2.35) k=0Kmk(b,,b)=0.\sum_{k=0}^{K}m^{k}(b,\cdots,b)=0.

In other words, bb is assumed to be nilpotent.

There is a very important class of examples for which nilpotent elements naturally exist. For example, if CC as a \mathbb{Z}-module has finitely many generators which lie in FC0F_{C}^{0}, then any such generator in the strictly positive part of the filtration is nilpotent, because the AA_{\infty}-structure maps increase the filtration. For geometric applications, we shall see that wrapped Floer cochain spaces have this property. Thus it does not harm to give a name for such a filtration.

Definition 2.13.

We say that the filtration FCF_{C} is bounded above, if ΛC\Lambda_{C} is bounded above. Equivalently, there exists λ+\lambda_{+} such that

FCλ+C=0.F_{C}^{\lambda_{+}}C=0.

Suppose that we have chosen a discrete filtration FCF_{C} on (C,mk)(C,m^{k}) as well as a discrete strictly compatible filtration FDF_{D} on (D,nk)(D,n^{k}). For simplicity, we assume that 0ΛD0\in\Lambda_{D}. Now we are going to introduce the key notion.

Definition 2.14.

We say that uDu\in D is a cyclic element, if the following properties are satisfied:

  1. (i)

    the map CDC\to D sending xx to n1(x;u)n^{1}(x;u) is a filtration-preserving isomorphism of \mathbb{Z}-modules, with its inverse also filtration-preserving;

  2. (ii)

    uFD0Du\in F_{D}^{0}D, but n0(u)FD0Dn^{0}(u)\notin F_{D}^{0}D.

Note that n0(FDλD)FDλDn^{0}(F_{D}^{\lambda}D)\subset F_{D}^{\lambda}D for every λ\lambda. Thus the second condition means when n0n^{0} acts on uu, it strictly increases the filtration.

The following result is proved in [Fuk15] in the case of filtered AA_{\infty}-algebras and filtered AA_{\infty}-modules (Proposition 3.5 of [Fuk15]) over the Novikov ring. We have a similar result in our case, which yields bounding cochains of ”finite type”.

Lemma 2.15.

Let (C,mk)(C,m^{k}) be a curved AA_{\infty}-algebra, and (D,nk)(D,n^{k}) a left AA_{\infty}-module over (C,mk)(C,m^{k}). Suppose that the filtrations FCF_{C} and FDF_{D} are bounded above. Suppose uDu\in D is a cyclic element such that uFD0u\in F_{D}^{0}. Then there exists a unique nilpotent bounding cochain bb of (C,mk)(C,m^{k}) such that

bFCλ1Cb\in F_{C}^{\lambda_{1}}C

for some λ1>0\lambda_{1}>0, and

(2.36) db(u)=0.d^{b}(u)=0.

Here db:DDd^{b}:D\to D is defined by

db(y)=k=0nk(b,,b;y).d^{b}(y)=\sum_{k=0}^{\infty}n^{k}(b,\cdots,b;y).
Proof.

Because the filtration FDF_{D} is strictly compatible with FCF_{C}, it is possible to find free generators yjy_{j} of DD and xjx_{j} of CC, such that yjFDλjDy_{j}\in F_{D}^{\lambda_{j}}D and xjFCλjCx_{j}\in F_{C}^{\lambda_{j}}C, where λjΛD=ΛC\lambda_{j}\in\Lambda_{D}=\Lambda_{C}.

We write u=u0y0++ulylu=u_{0}y_{0}+\cdots+u_{l}y_{l} in a unique way, where uju_{j}\in\mathbb{Z} and yjDy_{j}\in D are free generators of the free \mathbb{Z}-module DD, such that y0FD0Dy_{0}\in F_{D}^{0}D, and yjFDλjDy_{j}\in F_{D}^{\lambda_{j}}D but yjFDλDy_{j}\notin F_{D}^{\lambda}D if λ>λj\lambda>\lambda_{j}, where 0<λ1<<λl0<\lambda_{1}<\cdots<\lambda_{l}. Because of condition (ii) for a cyclic element, we have that n0(u0)=0n^{0}(u_{0})=0, so that n0(u)=j=1lujn0(yj)FDλ1n^{0}(u)=\sum_{j=1}^{l}u_{j}n^{0}(y_{j})\in F_{D}^{\lambda_{1}}.

Also, bb has a unique expression b=b1x1+blxlb=b_{1}x_{1}+\cdots b_{l}x_{l}, where bjb_{j}\in\mathbb{Z} and xjCx_{j}\in C are free generators of the free \mathbb{Z}-module CC, such that xjFCλjCx_{j}\in F_{C}^{\lambda_{j}}C but xjFCλCx_{j}\notin F_{C}^{\lambda}C if λ>λj\lambda>\lambda_{j}, where λ1<<λm\lambda_{1}<\cdots<\lambda_{m}.

Let us try to solve the equation (2.36) for such an element bb. This breaks down to a system of equations

(2.37) ki1,,ikjbi1bikujnk(xi1,,xik;yj)=0.\sum_{k}\sum_{i_{1},\cdots,i_{k}}\sum_{j}b_{i_{1}}\cdots b_{i_{k}}u_{j}n^{k}(x_{i_{1}},\cdots,x_{i_{k}};y_{j})=0.

Here the term nk(xi1,,xik;yj)FDλi1++λik+λjDn^{k}(x_{i_{1}},\cdots,x_{i_{k}};y_{j})\in F_{D}^{\lambda_{i_{1}}+\cdots+\lambda_{i_{k}}+\lambda_{j}}D, but not in a FDλDF_{D}^{\lambda^{\prime}}D for any smaller λ\lambda^{\prime}. Because FDF_{D} is bounded above, there can be at most finitely many such terms which are non-zero. Thus this system can be separated to finitely many equations, according to the filtration. We order these equations in an increasing order in terms of filtration. Rewrite every equation as the form

(2.38) biu0n1(xi;y0)+bi1bikujnk(xi1,,xik;yj)=0b_{i}u_{0}n^{1}(x_{i};y_{0})+\sum b_{i_{1}}\cdots b_{i_{k}}u_{j}n^{k}(x_{i_{1}},\cdots,x_{i_{k}};y_{j})=0

where the second sum is taken over all possible indices so that λi1++λik+λjλi\lambda_{i_{1}}+\cdots+\lambda_{i_{k}}+\lambda_{j}\leq\lambda_{i}. In particular, all ij<ii_{j}<i. When i=1i=1, there are no bb’s in the second sum, and the only possibly nonzero terms are of the form ujn0(uj)u_{j}n^{0}(u_{j}) such that λjλi+λ0\lambda_{j}\leq\lambda_{i}+\lambda_{0}. Because n1(;u)n^{1}(\cdot;u) is a filtration-preserving isomorphism of \mathbb{Z}-modules, the coefficient of the first term u0u_{0} is invertible: u0=±1u_{0}=\pm 1. Thus we can solve for a unique b1b_{1}. Then we consider i=2i=2 and argue in the same way to solve for b2b_{2}. We repeat this process until we solve for all bi,i=1,,lb_{i},i=1,\cdots,l.

It remains to prove that this solution is also a solution to the inhomogeneous Maurer-Cartan equation (2.34). Since bFCλ1Cb\in F_{C}^{\lambda_{1}}C for λ1>0\lambda_{1}>0, and FCF_{C} is bounded above, bb is automatically nilpotent, and the equation (2.34) can be reduced to (2.35). The strategy is to prove that

(2.39) kmk(b,,b)FCλC\sum_{k}m^{k}(b,\cdots,b)\in F_{C}^{\lambda^{\prime}}C

for every λ>0\lambda^{\prime}>0, which implies that it must vanish because FCF_{C} is bounded above: for λ\lambda^{\prime} large, FCλC=0F_{C}^{\lambda^{\prime}}C=0. From the AA_{\infty}-equations for nkn^{k} and mkm^{k} we get the following equation

(2.40) nk1(b,,b,nk2(b,,b;u)+nk1(b,,mk2(b,,b),,b);u)=0.\sum n^{k_{1}}(b,\cdots,b,n^{k_{2}}(b,\cdots,b;u)+\sum n^{k_{1}}(b,\cdots,m^{k_{2}}(b,\cdots,b),\cdots,b);u)=0.

The first sum vanishes because nk(b,,b;u)=0\sum n^{k}(b,\cdots,b;u)=0. Following the same kind of argument as before, we rewrite the above equation as the following:

(2.41) bi1bikujnk1(xi1,,xis,mk2(xis+1,,xis+k2),,xik;yj)=0.\sum b_{i_{1}}\cdots b_{i_{k}}u_{j}n^{k_{1}}(x_{i_{1}},\cdots,x_{i_{s}},m^{k_{2}}(x_{i_{s+1}},\cdots,x_{i_{s+k_{2}}}),\cdots,x_{i_{k}};y_{j})=0.

We may further write this equation as a system of equations, ordered by the filtration. Let us try to prove (2.39) by induction on λ\lambda, namely we take a discrete sequence λn+\lambda^{\prime}_{n}\to+\infty and prove (2.39) for λn\lambda^{\prime}_{n}. Recall that we have

(2.42) biu0n1(xi;y0)+bi1bikujnk(xi1,,xik;yj)=0b_{i}u_{0}n^{1}(x_{i};y_{0})+\sum b_{i_{1}}\cdots b_{i_{k}}u_{j}n^{k}(x_{i_{1}},\cdots,x_{i_{k}};y_{j})=0

If we assume (2.39) holds for any λ\lambda^{\prime} smaller than λi\lambda_{i}, it follows that the terms on the left hand side of (2.41) cancel with each other except that the following terms are left over:

(2.43) biu0n0(m1(xi);y0)+bi1biku0nk(mk(xi1,,xik);y0)b_{i}u_{0}n^{0}(m^{1}(x_{i});y_{0})+\sum b_{i_{1}}\cdots b_{i_{k}}u_{0}n^{k}(m^{k}(x_{i_{1}},\cdots,x_{i_{k}});y_{0})

where the second sum is taken over all possible indices such that λi1++λikλi\lambda_{i_{1}}+\cdots+\lambda_{i_{k}}\leq\lambda_{i}. As the map n0(;u)n^{0}(\cdot;u) is a filtration-preserving isomorphism, and y0y_{0} is the summand of uu in lowest filtration, this implies that

bim1(xi)+bi1bikmk(xi1,,xik)=0.b_{i}m^{1}(x_{i})+\sum b_{i_{1}}\cdots b_{i_{k}}m^{k}(x_{i_{1}},\cdots,x_{i_{k}})=0.

Since m1m^{1} does not decrease the filtration, this implies that (2.39) holds for λi\lambda_{i}.

2.6. Homotopy direct limit

In this subsection, we discuss an AA_{\infty}-analogue of homotopy direct limit, which is necessary for extending the action-restriction functor to the whole wrapped Fukaya categories, which is the main concern of section 6. The basic idea based on the existing AA_{\infty}-functors (6.21) is to take some kind of limit over all admissible Lagrangian submanifolds in M×NM\times N.

We temporarily depart from geometry and formalize the idea in a general situation of AA_{\infty}-categories and functors, not limited to Fukaya-type categories. Let 𝒜,\mathcal{A},\mathcal{B} be AA_{\infty}-categories consisting of countably many objects and let

id:𝒜d𝒜,i_{d}:\mathcal{A}_{d}\to\mathcal{A},
jd:dj_{d}:\mathcal{B}_{d}\to\mathcal{B}

be full AA_{\infty}-subcategories of 𝒜\mathcal{A} and \mathcal{B} respectively, both of which consist of dd objects: X1,,XdX_{1},\cdots,X_{d} for 𝒜d\mathcal{A}_{d} and Y1,,YdY_{1},\cdots,Y_{d} for d\mathcal{B}_{d}. We shall make the following assumption:

Assumption 2.16.

For d<dd<d^{\prime} the subcategory 𝒜d\mathcal{A}_{d} is also a full AA_{\infty}-subcategory of 𝒜d\mathcal{A}_{d^{\prime}} via embedding

id,d:𝒜d𝒜di_{d,d^{\prime}}:\mathcal{A}_{d}\to\mathcal{A}_{d^{\prime}}

and that the sequence of AA_{\infty}-subcategories

(2.44) 𝒜1𝒜d\mathcal{A}_{1}\subset\cdots\subset\mathcal{A}_{d}\subset\cdots

form a directed system such that

(2.45) 𝒜=limd𝒜d.\mathcal{A}=\lim\limits_{\begin{subarray}{c}\longrightarrow\\ d\to\infty\end{subarray}}\mathcal{A}_{d}.

The same assumption should hold for \mathcal{B}.

Now suppose that there exists an AA_{\infty}-functor

(2.46) d:𝒜dd\mathcal{F}_{d}:\mathcal{A}_{d}\to\mathcal{B}_{d}

sending XiX_{i} to YiY_{i}, and induces an isomorphism on cohomology categories. To be able to take the limit over dd to obtain an AA_{\infty}-functor

(2.47) :𝒜,\mathcal{F}:\mathcal{A}\to\mathcal{B},

we shall also need certain compatibility condition on the sequence of AA_{\infty}-functors d\mathcal{F}_{d}:

Assumption 2.17.

For d<dd<d^{\prime} the AA_{\infty}-functor d\mathcal{F}_{d^{\prime}} maps the subcategory 𝒜d\mathcal{A}_{d} of 𝒜d\mathcal{A}_{d^{\prime}} to the subcategory jd,d(d)j_{d,d^{\prime}}(\mathcal{B}_{d}) of d\mathcal{B}_{d^{\prime}}. Thus we may define

(2.48) jd,d1did,d:𝒜dd.j^{-1}_{d,d^{\prime}}\circ\mathcal{F}_{d^{\prime}}\circ i_{d,d^{\prime}}:\mathcal{A}_{d}\to\mathcal{B}_{d}.

This should be homotopic to d\mathcal{F}_{d} as AA_{\infty}-functors from 𝒜d\mathcal{A}_{d} to d\mathcal{B}_{d}.

Under Assumptions 2.16 and 2.17, it is almost straightforward to obtain the AA_{\infty}-functor (2.47) by taking the homotopy direct limit, in the sense we describe below. The following lemma is a straightforward consequence of these assumptions, which can be proved by using homological perturbation lemma.

Lemma 2.18.

For each dd and dd^{\prime}, there exists an auto-equivalence d,d\mathcal{H}_{d,d^{\prime}} of d\mathcal{B}_{d} such that the composition jd,dd,ddj_{d,d^{\prime}}\circ\mathcal{H}_{d,d^{\prime}}\circ\mathcal{F}_{d} strictly coincides with did,d\mathcal{F}_{d^{\prime}}\circ i_{d,d^{\prime}}.

Moreover, we make the following observation:

Lemma 2.19.

jd+1,d+2d+1,d+2jd,d+1d,d+1j_{d+1,d+2}\circ\mathcal{H}_{d+1,d+2}\circ j_{d,d+1}\circ\mathcal{H}_{d,d+1} is homotopic to jd,d+2j_{d,d+2}.

Proof.

We claim that the auto-equivalence d,d\mathcal{H}_{d,d^{\prime}} in Lemma 2.18 is actually homotopic to the identity functor, and that thus the lemma follows. The claim follows from the fact that jd,ddj_{d,d^{\prime}}\circ\mathcal{F}_{d} is homotopic to did,d\mathcal{F}_{d^{\prime}}\circ i_{d,d^{\prime}}, and the fact that the auto-equivalence d,d\mathcal{H}_{d,d^{\prime}} of d\mathcal{B}_{d} is introduced to make them strictly agree. ∎

We are thus led to the situation where we have homotopy directed systems of AA_{\infty}-categories and diagrams of AA_{\infty}-functors

(2.49) 𝒜1i1,2𝒜2i2,3𝒜3i3,41231j1,21,22j2,32,33j3,43,4\begin{CD}\mathcal{A}_{1}@>{i_{1,2}}>{}>&\mathcal{A}_{2}@>{i_{2,3}}>{}>&\mathcal{A}_{3}@>{i_{3,4}}>{}>&\cdots\\ @V{}V{\mathcal{F}_{1}}V&@V{}V{\mathcal{F}_{2}}V&@V{}V{\mathcal{F}_{3}}V\\ \mathcal{B}_{1}@>{j_{1,2}\circ\mathcal{H}_{1,2}}>{}>&\mathcal{B}_{2}@>{j_{2,3}\circ\mathcal{H}_{2,3}}>{}>&\mathcal{B}_{3}@>{j_{3,4}\circ\mathcal{H}_{3,4}}>{}>&\cdots\end{CD}

such that each square is strictly commutative. Then we can take the direct limits:

limd(𝒜d,id,d+1)=𝒜,\lim\limits_{\begin{subarray}{c}\longrightarrow\\ d\to\infty\end{subarray}}(\mathcal{A}_{d},i_{d,d+1})=\mathcal{A},
limd(d,jd,d+1d,d+1)=~,\lim\limits_{\begin{subarray}{c}\longrightarrow\\ d\to\infty\end{subarray}}(\mathcal{B}_{d},j_{d,d+1}\circ\mathcal{H}_{d,d+1})=\tilde{\mathcal{B}},

where ~\tilde{\mathcal{B}} is quasi-equivalent to \mathcal{B}, as well as

(2.50) limdd=:𝒜~.\lim\limits_{\begin{subarray}{c}\longrightarrow\\ d\to\infty\end{subarray}}\mathcal{F}_{d}=\mathcal{F}:\mathcal{A}\to\tilde{\mathcal{B}}.

This is the limit AA_{\infty}-functor obtained from the sequence d\mathcal{F}_{d}, with target different from but quasi-equivalent to \mathcal{B}.

3. The wrapped Fukaya category: revisited

3.1. Overview

There are several ways of defining the wrapped Fukaya category of a Liouville manifold. The basic approaches grow out of the definition of wrapped Floer cohomology, which either uses a cofinal family of Hamiltonians linear at infinity [AS10], or a single Hamiltonian quadratic at infinity [Abo10], or more generally a single Hamiltonian whose growth at infinity is faster than linear. A more functorial approach is via categorical colimits and localization, without having to specify a particular choice of Hamiltonian as introduced in [GPS17]. Of course, all these approaches give quasi-equivalent AA_{\infty}-categories.

At first, we set up the wrapped Fukaya category using a single Hamiltonian quadratic at infinity. Although this is not the best definition as it relies on working with a specific choice of Hamiltonian, it is convenient for us to compare two versions of wrapped Fukaya categories of the product manifold, as we just have to work with some chosen Hamiltonians, so that the difficulty in dealing with complicated triple colimits of functors is bypassed.

Also, this setup simplifies the construction of functors from Lagrangian correspondences, as they can be constructed in a single step, without having to be defined as a colimit of functors. However, regarding applications and calculation, it is better to give a definition in the framework of wrapped Fukaya category defined with respect to linear Hamiltonians. This definition, due to [AS10] will be reviewed in subsection LABEL:section:_linear_Hamiltonians.

3.2. Basic geometric setup

The geometric setup considered here is exactly the same as that in [Gao17]. We briefly mention it here mainly for the purpose of fixing notations. Consider a Liouville manifold MM which is the completion of a Liouville domain M0M_{0} with boundary M\partial M, which has a collar neighborhood M×(ϵ,1]\partial M\times(\epsilon,1] so that the Liouville vector field is equal to r\frac{\partial}{\partial r} in that neighborhood. We assume that MM is symplectically Calabi-Yau, namely 2c1(M)=0H2(M;)2c_{1}(M)=0\in H^{2}(M;\mathbb{Z}).

The admissible Lagrangian submanifolds are either closed exact Lagrangian submanifolds in the interior M0M_{0}, or cylindrical Lagrangian submanifolds of the form L=L0L×[1,+)L=L_{0}\cup\partial L\times[1,+\infty) where LM\partial L\subset\partial M is a Legendrian submanifold with respect to the contact structure induced from the Liouville one-form. To be more specific, for the latter kind of Lagrangian submanifold L0L_{0} of M0M_{0}, there should be a function ff on it so that df=λ|L0df=\lambda|_{L_{0}}, where λ\lambda is the Liouville form. Moreover, we require that ff has an extension to a neighborhood of LL in MM such that it is locally constant near L×(ϵ,+)\partial L\times(\epsilon,+\infty). In addition, we shall make the assumption that

(3.1) 2c1(M,L)=0H2(M,L;),2c_{1}(M,L)=0\in H^{2}(M,L;\mathbb{Z}),

which ensures the existence of gradings and spin structures on LL. We will fix a choice of grading and spin structure for every admissible Lagrangian submanifold. These conditions will ensure that the wrapped Fukaya category of MM is defined over \mathbb{Z}, and carries \mathbb{Z}-gradings.

3.3. Floer data and consistency

The moduli space of surfaces controlling the algebraic operations and relations in the Fukaya categories are the moduli spaces of stable marked nodal disks, which was studied in [FOOO09a] and proved to be cellular isomorphic to the moduli spaces of stable metric ribbon trees introduced by Stasheff [Stasheff], which is known to be the operad controlling AA_{\infty}-algebras.

Let ¯k+1\bar{\mathcal{M}}_{k+1} be the compactified moduli space of stable (k+1)(k+1)-marked disks. It is proved in [FOOO09a] that ¯k+1\bar{\mathcal{M}}_{k+1} is a compact smooth manifold with corners and a neighborhood of the stratum T\mathcal{M}_{T} with combinatorial type modeled on a stable ribbon tree TT with one root and kk-leaves is covered by the image of the gluing map:

(3.2) (1,0]e(T)×TUT¯k+1,(-1,0]^{e(T)}\times\mathcal{M}_{T}\supset U_{T}\to\bar{\mathcal{M}}_{k+1},

which is smooth and a diffeomorphism onto the image by shrinking UTU_{T} if necessary.

To define AA_{\infty}-operations on Floer cochain spaces, we need to study moduli spaces of (perturbed) pseudoholomorphic maps from nodal disks to MM with boundary in Lagrangian submanifolds. For this purpose, we also need to include the case where the domain is unstable, or contains unstable components. The unstable curve involved here is the infinite strip ZZ, whose automorphism is the additive group \mathbb{R}. To write down inhomogeneous Cauchy-Riemann equations and achieve transversality of the moduli spaces of solutions, we need several auxilliary data. We briefly recall the notion here.

Definition 3.1.

Given a semistable (k+1)(k+1)-marked nodal disk Sk¯S\in\bar{\mathcal{M}_{k}}, a Floer datum PSP_{S} for SS consists of

  1. (i)

    A collection of positive integers w0,,wkw_{0},\cdots,w_{k}.

  2. (ii)

    A time-shifting function ρS:S[1,+)\rho_{S}:\partial S\to[1,+\infty), which takes the value wjw_{j} over the jj-th strip-like end ϵj\epsilon_{j}.

  3. (iii)

    A basic one form αS\alpha_{S}, whose restriction to every smooth component of SS is closed, and whose pullback by ϵj\epsilon_{j} agrees with wjdtw_{j}dt.

  4. (iv)

    A Hamiltonian perturbation HS:S(M)H_{S}:S\to\mathcal{H}(M), whose pullback by ϵj\epsilon_{j} agrees with Hϕwjwj2\frac{H\circ\phi^{w_{j}}}{w_{j}^{2}}.

  5. (v)

    A domain-dependent perturbation of almost complex structures JS:S𝒥(M)J_{S}:S\to\mathcal{J}(M), whose pullback by ϵj\epsilon_{j} agrees with (ϕwj)Jt(\phi^{w_{j}})^{*}J_{t}.

such that over unstable components of SS, i.e. strips, all the three data restricts to translation-invariant data.

In order to ensure that the various operations constructed from moduli spaces of marked inhomogeneous pseudoholomorphic disks satisfy the AA_{\infty}-equations, we need to make sure that the Floer data chosen for the underlying semistable marked nodal disks are compatible with respect to gluing maps (3.2). Therefore the following notion is useful: for k3k\geq 3, a universal and consistent choice of Floer data is a choice of Floer data for all S¯k+1S\in\bar{\mathcal{M}}_{k+1} that varies smoothly with respect to SS in the compactified moduli space. The notion of universal and consistent choice of Floer data is extended also to the strip ZZ, as follows: when we glue in a strip at a boundary marked point of a stable marked nodal disk SS, we require that the (translation-invariant) Floer datum chosen on ZZ agree with that on the strip-like end for SS near that marked point. The space of choices of Floer data is convex, therefore by induction on the strata of the moduli space of stable marked nodal disks, we can construct universal and consistent choices of Floer data. More detailed explanation is given in [Sei08].

3.4. Inhomogeneous pseudoholomorphic disks

To define the AA_{\infty}-operations on the wrapped Fukaya category, we need to study the moduli spaces of inhomogeneous pseudoholomorphic disks with boundary mapped to several Lagrangian submanifolds. Make universal and consistent choices of Floer data PP for all semistable marked nodal disks. Denote by SS an element in the smooth part of the moduli space k+1\mathcal{M}_{k+1}. That is, SS is a smooth disk with boundary marked points (z0,,zk)(z_{0},\cdots,z_{k}) that are cyclically ordered on the boundary. Given admissible Lagrangian submanifolds L0,,LkL_{0},\cdots,L_{k}, consider the following inhomogeneous Cauchy-Riemann equation, for both SS and uu as variables:

(3.3) {(duαSXHS)0,1=0;u(z)ϕρS(z)Lj, if z lies in between zj and zj+1;limsuϵ0(s,)=ϕw0x0()𝒳(ϕw0L0,ϕw0Lk);lims+uϵj(s,)=ϕwjxj()𝒳(ϕwjLj1,ϕwjLj),j=1,,k.\begin{cases}(du-\alpha_{S}\otimes X_{H_{S}})^{0,1}=0;\\ u(z)\in\phi^{\rho_{S}(z)}L_{j},\text{ if $z$ lies in between $z_{j}$ and $z_{j+1}$};\\ \lim\limits_{s\to-\infty}u\circ\epsilon_{0}(s,\cdot)=\phi^{w_{0}}x_{0}(\cdot)\in\mathcal{X}(\phi^{w_{0}}L_{0},\phi^{w_{0}}L_{k});\\ \lim\limits_{s\to+\infty}u\circ\epsilon_{j}(s,\cdot)=\phi^{w_{j}}x_{j}(\cdot)\in\mathcal{X}(\phi^{w_{j}}L_{j-1},\phi^{w_{j}}L_{j}),&j=1,\cdots,k.\end{cases}

The solutions will sometimes also be called Floer’s disks.

Suppose for the moment k2k\geq 2. Let k+1(L0,,Lk;x0,,xk;P)\mathcal{M}_{k+1}(L_{0},\cdots,L_{k};x_{0},\cdots,x_{k};P) be the moduli space of solutions (S,u)(S,u) to the above equation with respect to the chosen Floer data PP, and let ¯k+1(L0,,Lk;x0,,xk;P)\bar{\mathcal{M}}_{k+1}(L_{0},\cdots,L_{k};x_{0},\cdots,x_{k};P) be its stable map compactification. It is proved in [Abo10] that for a generic choice of Floer data PP, the zero-dimensional and one-dimensional components of ¯k+1(L0,,Lk;x0,,xk;P)\bar{\mathcal{M}}_{k+1}(L_{0},\cdots,L_{k};x_{0},\cdots,x_{k};P) are compact smooth manifolds with corners of dimension

2k+deg(x0)deg(x1)deg(xk).2-k+\deg(x_{0})-\deg(x_{1})-\cdots-\deg(x_{k}).

In the unstable case k=1k=1, there is no moduli of SS, so we consider the set of solutions uu to the above equation. Since the Floer datum PZP_{Z} on the strip is chosen to be translation-invariant, we can quotient the parametrized moduli space by this automorphism group. We denote the quotient moduli space by 2(L0,L1;x0,x1;PZ)\mathcal{M}_{2}(L_{0},L_{1};x_{0},x_{1};P_{Z}) as well, and the corresponding Gromov bordification by ¯2(L0,L1;x0,x1;PZ)\bar{\mathcal{M}}_{2}(L_{0},L_{1};x_{0},x_{1};P_{Z}).

3.5. Identification of Floer cochain spaces with different weights

The ”count” of rigid elements in the moduli spaces ¯k+1(L0,,Lk;x0,,xk;P)\bar{\mathcal{M}}_{k+1}(L_{0},\cdots,L_{k};x_{0},\cdots,x_{k};P) defines operations of the following kind

(3.4) CW(ϕwkLk1,ϕwkLk;Hϕwkwk)CW(ϕw1L0,ϕw1L1;Hϕw1w1)CW(ϕw0L0,ϕw0Lk;Hϕw0w0).\begin{split}&CW^{*}(\phi^{w_{k}}L_{k-1},\phi^{w_{k}}L_{k};\frac{H\circ\phi^{w_{k}}}{w_{k}})\otimes\cdots\otimes CW^{*}(\phi^{w_{1}}L_{0},\phi^{w_{1}}L_{1};\frac{H\circ\phi^{w_{1}}}{w_{1}})\\ \to&CW^{*}(\phi^{w_{0}}L_{0},\phi^{w_{0}}L_{k};\frac{H\circ\phi^{w_{0}}}{w_{0}}).\end{split}

In order define an honest AA_{\infty}-category with morphism spaces being CW(L,L;H)CW^{*}(L,L^{\prime};H) between the two objects LL and LL^{\prime} so that the operations happen on these morphism spaces, we have to identify CW(ϕwL,ϕwL;Hϕww)CW^{*}(\phi^{w}L,\phi^{w}L^{\prime};\frac{H\circ\phi^{w}}{w}) with CW(L,L;H)CW^{*}(L,L^{\prime};H), in a canonical way. Because the Hamiltonian HH is quadratic in the radial coordinate of the cylindrical end, such an identification is easily achieved by noting that Hϕww2\frac{H\circ\phi^{w}}{w^{2}} behaves the same as HH in the cylindrical end where Reeb dynamics occur. Technically, the rescaled Hamiltonian differs from HH by a small amount that is supported in the compact part of MM, and this can be taken care of by using a compactly supported deformation of Hamiltonian functions, which gives rise to continuation maps that form an AA_{\infty}-quasi-isomorphism (of AA_{\infty}-bimodule structures on the wrapped Floer cochain spaces).

To summarize, these arguments in previous subsections together imply that the wrapped Fukaya category 𝒲(M)\mathcal{W}(M) is well-defined, up to quasi-isomorphism.

3.6. Winding Lagrangian submanifolds

Let us introduce a new class of Lagrangian submanifolds in the wrapped Fukaya category. These Lagrangian submanifolds come from geometric compositions of Lagrangian correspondences to be discussed in detail in section 7.4. This class includes in particular HH-perturbed cylindrical Lagrangian submanifolds in MM, i.e. ϕH1(L)\phi_{H}^{1}(L) for some cylindrical Lagrangian submanifold L=L0L×[1,+)L=L_{0}\cup\partial L\times[1,+\infty), where the Hamiltonian perturbation is the same as the one used to defined 𝒲(M)\mathcal{W}(M) - these HH-perturbed cylindrical Lagrangian submanifolds are geometric compositions with the diagonal. The picture of such a Lagrangian submanifold is one that winds around in the cylindrical end of MM. The point is, we want to include both cylindrical Lagrangian submanifolds and these perturbed cylindrical Lagrangian submanifolds as object in our wrapped Fukaya category, with respect to the same Liouville structure, although one can easily show that the perturbed ones are cylindrical with respect to a different Liouville structure.

Now let ϕH(L)\phi_{H}(L) be an HH-perturbed Lagrangian submanifold. Geometrically, it looks like winding around a cylindrical Lagrangian submanifold in the cylindrical end in accelerating speeds with respect to the radial coordinate. The first task is to define its self wrapped Floer cohomology. On the level of the underlying cochain groups, this is fairly straightforward: the self wrapped Floer complex of ϕH(L)\phi_{H}(L) can be thought of as the Lagrangian intersection Floer cochain complex of ϕH2(L)\phi_{H}^{2}(L) with ϕH(L)\phi_{H}(L). As the underlying cochain group, this is canonically isomorphic to CF(ϕH(L),L)CF^{*}(\phi_{H}(L),L) by applying the exact symplectomorphism ϕH1\phi_{H}^{-1}. However, the differential is a bit tricky to define. Recall that we have defined the differential on CW(L)CW^{*}(L) using inhomogeneous JJ-holomorphic strips with boundary on LL that converge to HH-chords at infinity. If one somehow wants to ”transport” the differential of CW(L)CW^{*}(L) to one of CW(ϕH(L))CW^{*}(\phi_{H}(L)), geometrically that will be to consider moduli space of inhomogeneous (ϕH)J(\phi_{H})^{*}J-holomorphic strips, instead of JJ-holomorphic strips. On possible way to resolve this is to refer to the Lagrangian intersection setup, where one defines CW(ϕH(L))CW^{*}(\phi_{H}(L)) as the Lagrangian intersection Floer complex of ϕH2(L)\phi_{H}^{2}(L) with ϕH(L)\phi_{H}(L). Thus the differential will be to use homogeneous JJ-holomorphic strips with boundary on ϕH2(L)\phi_{H}^{2}(L) and ϕH(L)\phi_{H}(L) which converge to the corresponding intersection points. An appropriate version of maximum principle in this setting is necessary. To establish that, we compare ϕH2(L)\phi_{H}^{2}(L) with ϕ2H(L)\phi_{2H}(L) and note they could only possibly differ by a compact-supported Hamiltonian isotopy. Since 2H2H is also admissible, the maximum principle applies to such JJ-holomorphic strips.

Alternatively, we could directly try to prove the compactness results of the moduli spaces of inhomogeneous pseudoholomorphic disks with boundary on (appropriate rescalings of) ϕH(L)\phi_{H}(L) by appealing to the action-energy equality. The most important fact is the following equality computing the new primitive function for ϕH(L)\phi_{H}(L) in terms of that of for LL.

Lemma 3.2.

Let ff be a primitive for LL. Then the following function

(3.5) f+ιXλf+\iota_{X}\lambda

is a primitive for ϕH(L)\phi_{H}(L). Here XX is the Hamiltonian vector field of HH.

Proof.

The proof is a straightforward calculation, based on the well-known fact that a Hamiltonian symplectomorphism is exact and adds to the primitive of the symplectic form the following:

(3.6) d01ιXtλ𝑑t.d\int_{0}^{1}\iota_{X_{t}}\lambda dt.

Now since our Hamiltonian is time-independent, this is simply equal to

(3.7) dιXλ.d\iota_{X}\lambda.

Thus a primitive for ϕH(L)\phi_{H}(L) can be taken to be

(3.8) f+ιXλf+\iota_{X}\lambda

The following estimate on the new primitive is the crucial step in proving compactness result for the relevant moduli spaces in wrapped Floer theory.

Lemma 3.3.

ιXλ\iota_{X}\lambda equals 2r22r^{2} in the cylindrical end M×[1,+)\partial M\times[1,+\infty). In particular, it is constant on any level hypersurface M×{r}\partial M\times\{r\}.

Proof.

This is a straightforward elementary calculation. ∎

Looking back at the action-energy equality (the action is computed with respect to the original primitive),

𝒜H,ϕH(L)(x)=01xλ+H(x(t))dt+fϕH(L)(x(1))fϕH(L)(x(0)),\mathcal{A}_{H,\phi_{H}(L)}(x)=\int_{0}^{1}-x^{*}\lambda+H(x(t))dt+f_{\phi_{H}(L)}(x(1))-f_{\phi_{H}(L)}(x(0)),

we find that for an HH-chord from ϕH(L)\phi_{H}(L) to itself which is contained in a level hypersurface, the last two terms contribute zero because the extra term ιXλ\iota_{X}\lambda in the primitive is constant on this hypersurface. Thus, the same estimates apply as if we were in the case of a cylindrical Lagrangian submanifold, and consequently the action-energy equality implies compactness results for the moduli spaces of inhomogeneous pseudoholomorphic strips with boundary on ϕH(L)\phi_{H}(L). Therefore, the self wrapped Floer cohomology of ϕH(L)\phi_{H}(L) is well-defined.

The second task is to define the wrapped Floer cohomology HW(ϕH(L),L)HW^{*}(\phi_{H}(L),L^{\prime}), for any pair (ϕH(L),L)(\phi_{H}(L),L^{\prime}) where LL^{\prime} is either a closed exact Lagrangian submanifold in the interior M0M_{0}, or a cylindrical Lagrangian submanifold. Reasoning in the same way as above, we know that this can be defined as the Lagrangian intersection Floer cohomology of ϕH2(L)\phi_{H}^{2}(L) with LL^{\prime}. On the other hand, by a rescaling argument introduced in the previous subsection, that group is isomorphic to HW(L,L)HW^{*}(L,L^{\prime}). There is also an alternative construction by direct analysis on the compactness of the moduli spaces of inhomogeneous pseudoholomorphic strips with boundary on the pair (ϕH(L),L)(\phi_{H}(L),L^{\prime}) using estimates on the primitives, which can be done in a similar way to that for a single HH-perturbed Lagrangian submanifold.

The third task is to define the wrapped Floer cohomology HW(L,ϕH(L))HW^{*}(L^{\prime},\phi_{H}(L)). The story is parallel to the previous case, but there is an asymmetry. This time, HW(L,ϕH(L))HW^{*}(L^{\prime},\phi_{H}(L)) should be regarded as the Lagrangian intersection Floer cohomology of ϕH(L)\phi_{H}(L^{\prime}) with ϕH(L)\phi_{H}(L). In a generic situation where LL and LL^{\prime} intersect transversely and the Legendrian boundaries of LL^{\prime} and LL do not intersect, there is in fact a canonical chain-level isomorphism between CF(ϕH(L),ϕH(L))CF^{*}(\phi_{H}(L^{\prime}),\phi_{H}(L)) and CF(L,L)CF^{*}(L^{\prime},L), where both are defined with respect to JJ. This is because there are only finitely many intersections which are contained in the compact domain M0M_{0}, so distinguishing JJ and ϕHJ\phi_{H}^{*}J is irrelevant. In the special case where L=LL=L^{\prime}, further discussion is needed. We may define a Morse-Bott Floer complex by introducing a chain model for LL, say a Morse complex of a Morse function f:Lf:L\to\mathbb{R} without any critical point in the cylindrical end L×[1,+)\partial L\times[1,+\infty). For example, we can take ff to be the restriction of HH to LL, provided HH is generic. The differential counts (J,H)(J,H)-pseudoholomorphic strips, which become gradient flow lines provided HH is C2C^{2}-small. Thus the cohomology HW(L,ϕH(L))HW^{*}(L,\phi_{H}(L)) is isomorphic to the ordinary cohomology of LL.

The next task is to introduce AA_{\infty}-structures on the cochain groups underlying the above-mentioned wrapped Floer cohomology groups. There are several cases:

  1. (i)

    The self wrapped Floer cochain complex CW(ϕH(L))CW^{*}(\phi_{H}(L)) should be equipped with a structure of an AA_{\infty}-algebra. Because of our previous definition of this cochain complex as the Lagrangian intersection Floer cochain complex of ϕH2(L)\phi_{H}^{2}(L) with LL, it is inappropriate to use inhomogeneous pseudoholomorphic disks with boundary on LL.

  2. (ii)

    Regarding the wrapped Floer cochain complex CW(ϕH(L),L)CW^{*}(\phi_{H}(L),L^{\prime}) for a pair, it is equipped with an AA_{\infty}-bimodule structure over (CW(ϕH(L)),CW(L))(CW^{*}(\phi_{H}(L)),CW^{*}(L^{\prime})), defined by the moduli spaces of (J,H)(J,H)-holomorphic strips with punctures on both boundary components that are mapped to (ϕH(L),L)(\phi_{H}(L),L^{\prime}), together with (J,H)(J,H)-holomorphic disks with punctures with boundary on ϕH(L)\phi_{H}(L) or LL^{\prime}, which are joint to the strips at their boundary punctures.

  3. (iii)

    As for CW(L,ϕH(L))CW^{*}(L^{\prime},\phi_{H}(L)), there is a small difference. To equip this with a natural left-CW(L)CW^{*}(L^{\prime}) and right-CW(ϕH(L))CW^{*}(\phi_{H}(L)) AA_{\infty}-bimodule structure, we need a small modification of the setup of Floer complex for the pair (L,ϕH(L))(L^{\prime},\phi_{H}(L)), by defining the differential using the original setup for wrapped Floer cohomology, instead of JJ-holomorphic strips with boundary on (L,L)(L^{\prime},L). We consider CW(L,ϕH(L))CW^{*}(L^{\prime},\phi_{H}(L)) as being generated by time-one HH-chords from LL^{\prime} to ϕH(L)\phi_{H}(L), which are in one-to-one correspondence with intersection points LLL\cap L^{\prime}, hence finite. The differential counts rigid (J,H)(J,H)-holomorphic strips with boundary on (L,ϕH(L))(L^{\prime},\phi_{H}(L)), which is well-defined because there are only finitely many time-one HH-chords. Now the moduli space of (J,H)(J,H)-holomorphic strips with boundary on (L,ϕH(L))(L^{\prime},\phi_{H}(L)) is compatible with the moduli spaces of (J,H)(J,H)-holomorphic disks with boundary on LL^{\prime} or ϕH(L)\phi_{H}(L), meaning that the relevant pseudoholomorphic curves can be glued together in a coherent way. Therefore we obtain the desired AA_{\infty}-bimodule structure.

4. Wrapped Floer theory for Lagrangian immersions

4.1. Overview of immersed Lagrangian Floer theory

In this section, we extend wrapped Floer theory to certain classes of Lagrangian immersions. The main purpose of such an extension is to prove representability of functors associated to Lagrangian correspondences in general (to be discussed in section 7), though in many concrete and interesting cases, it is sufficient to study embedded Lagrangian submanifolds.

In order for the Lagrangian immersions in consideration to have well-behaved Floer theory, we must impose some conditions: they should satisfy a condition similar to being exact, be embedded in the cylindrical end of MM, and possibly have transverse self-intersections in the interior part of MM. Without loss of generality, we assume these self-intersections are at most double points. In general, there will be pseudoholomorphic disks bounded by the image of such a Lagrangian immersion, and these disks will interact with inhomogeneous pseudoholomorphic disks (solutions to Floer’s equation). Therefore, we should pick a good model for the compactifications of the relevant moduli spaces of disks. Fortunately, this can be done fairly directly, as the Lagrangian immersions we are going to consider still satisfy an ”exactness” condition, which will be introduced in the next subsection.

Wrapped Floer theory should contain information about Reeb dynamics on the boundary contact manifold, in addition to the cohomological generators of the Lagrangian submanifolds. The construction of the AA_{\infty}-structure maps involves both inhomogeneous pseudoholomorphic disks and homogeneous pseudoholomorphic disks. The entire picture would be an analogue of the setup of holomorphic curves in relative symplectic field theory (see [BEH+03]). Fortunately, there is a purely Floer-theoretic formulation, where we can construct moduli spaces of maps which satisfy certain Floer’s equation, and the virtual techniques used in proving transversality does not go beyond the theory of Kuranishi structures, because the pseudoholomorphic curves that we are dealing with are all of genus zero with connected boundary, considered by Fukaya-Oh-Ohta-Ono [FOOO09a], [FOOO09b].

Wrapped Floer theory assigns to such a Lagrangian immersion ι:LM\iota:L\to M a curved AA_{\infty}-algebra (CW(L,ι;H),mk)(CW^{*}(L,\iota;H),m^{k}) over \mathbb{Z} (compare to the case of general compact Lagrangian immersions studied by Akaho and Joyce [AJ10]). To define the immersed wrapped Fukaya category, we shall consider unobstructed Lagrangian immersions, i.e. those for which the curved AA_{\infty}-algebra CW(L,ι;H)CW^{*}(L,\iota;H) has a bounding cochain.

4.2. Gradings and spin structures

For Floer theory to carry an absolute \mathbb{Z}-grading as well as to have coefficients in \mathbb{Z}, we need to introduce the notions of gradings and spin structures.

Definition 4.1.

Say that the Lagrangian immersion ι:LM\iota:L\to M is graded, if the square phase function αL:LS1\alpha_{L}:L\to S^{1} has a lift α~L\tilde{\alpha}_{L} to \mathbb{R}. Here the square phase function is defined by sending xLx\in L to (dι)x(TxL)(d\iota)_{x}(T_{x}L), an element in the Lagrangian Grassmannian 𝒜𝒢(TM)\mathcal{LAG}(TM), then mapping that to S1S^{1} by pairing any orthonormal basis for the Lagrangian plane (dι)x(TxL)(d\iota)_{x}(T_{x}L) with the quadratic volume form, which is independent of the choice of an orthonormal basis. Such a lift is called a grading for this Lagrangian immersion ι:LM\iota:L\to M.

From now on we shall make the following assumption.

Assumption 4.2.

The manifold LL is spin with a chosen spin structure vv. Also, the Lagrangian immersion is graded in the sense of Definition 4.1 with a chosen grading α~L\tilde{\alpha}_{L}.

A grading for the Lagrangian immersion ι:LM\iota:L\to M defines an absolute Maslov index for each generator cc (note that if cc is a critical point of ff together with a capping half-disk ww, the disk Maslov index of ww agrees with the Morse index of ff at pp), which endows with the wrapped Floer cochain space a \mathbb{Z}-grading. A spin structure vv determines orientations on the moduli spaces we are going to introduce in subsection 4.5.

Remark 4.3.

The condition that the immersion ι:LM\iota:L\to M be graded implies that the Maslov class of ι:LM\iota:L\to M is zero. However, it does not prohibit the existence of holomorphic disks with boundary on ι(L)\iota(L) of non-zero Maslov indices.

4.3. The wrapped Floer cochain space for a cylindrical Lagrangian immersion

Let ι:LM\iota:L\to M be a Lagrangian immersion. To develop wrapped Floer theory for it, we shall make some geometric assumptions.

Definition 4.4.

A Lagrangian immersion ι:LM\iota:L\to M is said to have transverse self-intersections, if the self fiber product L×ιLL\times_{\iota}L is a smooth submanifold of L×LL\times L, and has a decomposition

(4.1) L×ιL=ΔLaLaL\times_{\iota}L=\Delta_{L}\coprod\coprod_{a}L_{a}

where ΔLL\Delta_{L}\cong L is the diagonal, and the disjoint union is a union of isolated points.

Let ι:LM\iota:L\to M be a Lagrangian immersion with transverse self-intersections which are at most double points. It is said to be cylindrical, if over the cylindrical end M×[1,+)\partial M\times[1,+\infty), ι\iota is an embedding, whose image is of the form l×[1,+)l\times[1,+\infty) for some Legendrian submanifold lMl\subset\partial M. More generally speaking, being cylindrical means the image is invariant under the Liouville flow outside of a compact set.

The Lagrangian immersion ι:LM\iota:L\to M is said to be exact, if there exists a function f:Lf:L\to\mathbb{R} such that df=ιλMdf=\iota^{*}\lambda_{M}.

Let ι:LM\iota:L\to M be an exact cylindrical Lagrangian immersion with transverse self-intersections. From now on we assume that all self-intersections are at most double points, in which case the decomposition of the self fiber product takes the form

(4.2) L×ιL=ΔLx{(p,p+):pp+,ι(p)=ι(p+)=x}.L\times_{\iota}L=\Delta_{L}\coprod\coprod_{x}\{(p_{-},p_{+}):p_{-}\neq p_{+},\iota(p_{-})=\iota(p_{+})=x\}.

The more classical version of Floer theory gives rise to a deformation of the singular chain complex of the Lagrangian submanifold. In addition to that, Reeb chords on the boundary contact manifold and the Legendrian boundary of the Lagrangian submanifolds should also be included as generators in wrapped Floer theory. In the case of a Lagrangian immersion, the self-intersection points bring in extra generators.

Given the above general idea, in order to choose a suitable chain model for the wrapped Floer cochain space for an exact cylindrical Lagrangian immersion, we consider the following geometric setup. We take a smooth Hamiltonian H:MH:M\to\mathbb{R} which is zero in the interior part M0M_{0} of the Liouville manifold MM. This in particular implies that all the Hamiltonian chords from ι(L)\iota(L) to itself which are contained in the interior part of MM are constant. Over the cylindrical end M×[1,+)\partial M\times[1,+\infty), more precisely on a smaller subset M×[1+ϵ,+)\partial M\times[1+\epsilon,+\infty) for some small ϵ>0\epsilon>0, the Hamiltonian HH is quadratic, i.e. of the form r2r^{2}.

Let 𝒫(M;ι(L))\mathcal{P}(M;\iota(L)) be the space of paths in MM from ι(L)\iota(L) to itself. Fix a choice of a basepoint xx_{*} in every connected component of 𝒫(M;ι(L))\mathcal{P}(M;\iota(L)). Let xx be a time-one HH-chord from ι(L)\iota(L) to itself (either constant or non-constant), which lies in a connected component of 𝒫(M;ι(L))\mathcal{P}(M;\iota(L)) where xx_{*} is located. A capping half-disk of ww with reference to xx_{*} is a map

(4.3) w:[0,1]×[0,1]M,w:[0,1]\times[0,1]\to M,

such that w(s,i)ι(L)w(s,i)\in\iota(L) for i=0,1i=0,1, and w(0,t)=x(t),w(1,t)=x(t)w(0,t)=x_{*}(t),w(1,t)=x(t). Now if xx is a non-constant time-one HH-chord contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty), there is a unique homotopy class of capping half-disk for xx, since ι\iota is an embedding there and is exact in the usual sense. For such a Hamiltonian chord, it is not necessary to specify the homotopy class of capping half-disks, just as in the setup for an embedded exact Lagrangian submanifold.

Because the Hamiltonian HH is constantly zero in the interior part, it is natural to introduce a Morse-Bott setup for wrapped Floer theory. There are many approaches, depending on the chain model for MM. If we use the singular chain model for the complex which computes the cohomology of MM, the transversality argument as in [FOOO09b] involves quite complicate process of choosing countably generated sub-complexes. To minimize the amount the work, we shall use the Morse complex for computing the cohomology of MM, so that there are finitely many generators of the complex which computes the cohomology of MM. Correspondingly, we define the wrapped Floer cochain space CW(L,ι;H)CW^{*}(L,\iota;H), which is a graded \mathbb{Z}-module, as follows.

Definition 4.5.

Denote by S(L,ι)S(L,\iota) the set of pairs (p,p+)(p_{-},p_{+}) so that ι(p)=ι(p+)\iota(p_{-})=\iota(p_{+}) but pp+p_{-}\neq p_{+}. Let ff be an auxiliary C2C^{2}-small Morse function on the fiber product L×ιL=ΔLS(L,ι)L\times_{\iota}L=\Delta_{L}\coprod S(L,\iota).

We define the wrapped Floer cochain space CW(L,ι;H)CW^{*}(L,\iota;H) to be the free \mathbb{Z}-module generated by the following two kinds of generators:

  1. (i)

    pairs (p,w)(p,w) where pCrit(f)p\in Crit(f) is a critical point of ff, and ww is a Γ\Gamma-equivalence class of capping half-disk for pp, i.e. a map from a one-punctured disk to MM with boundary on ι(L)\iota(L) which converges to pp at the puncture (see [FOOO09a] Chapter 2 for the definition of Γ\Gamma-equivalence);

  2. (ii)

    non-constant time-one HH-chords xx from ι(L)\iota(L) to itself contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty). The set of these non-constant Hamiltonian chords is denoted by 𝒳+(L,ι;H)\mathcal{X}_{+}(L,\iota;H).

For simplicity, we shall denote by cc any kind of generator.

The following lemma explains why this is a reasonable and meaningful definition.

Lemma 4.6.

Let KK be a C2C^{2}-small generic perturbation of HH supported in the interior part M0M_{0} of MM, which is also assumed to be non-degenerate. Then all the KK-chords starting from ι(L)\iota(L) to itself which are contained in the interior part of MM are constant and non-degenerate (hence isolated). Moreover, there is a natural one-to-one correspondence between these constant chords, and the critical points of the lift of KK to the fiber product L×ιL=ΔLS(L,ι)L\times_{\iota}L=\Delta_{L}\coprod S(L,\iota).

Proof.

Since KK is C2C^{2}-small in the interior part, all time-one KK-chords from ι(L)\iota(L) to itself which are contained in that region are constant. Non-degeneracy follows from the assumption, which then implies that all such KK-chords are isolated.

Note that as in (4.2), the fiber product is the union of the diagonal ΔLL\Delta_{L}\cong L and the set of pairs (p,p+)(p_{-},p_{+}), with pp+p_{-}\neq p_{+}, each of which corresponds to a self-intersection pp of the immersion ι\iota. The lift of KK to this fiber product is simply the pullback of KK by ι\iota on the main component ΔL\Delta_{L}, and constants on each isolated point (p,p+)(p_{-},p_{+}) with value equal to K(p)K(p) where p=ι(p)=ι(p+)Mp=\iota(p_{-})=\iota(p_{+})\in M. Since all the interior KK-chords from ι(L)\iota(L) to itself are constant, every such constant chord xx has image being a point pp on ι(L)\iota(L). If pp is not a self-intersection point of ι\iota, then it satisfies the condition that dKι(L)(p)=0dK_{\iota(L)}(p)=0, which implies that it is a critical point of the lift of KK to the fiber product L×ιLL\times_{\iota}L. ∎

On the other hand, non-constant time-one HH-chords are located in the cylindrical end M×[1,+)\partial M\times[1,+\infty), and correspond to Reeb chords on the contact boundary M\partial M, from the Legendrian submanifold ι(L)M\iota(L)\cap\partial M to itself. Thus Definition 4.5 is a good definition.

4.4. Pearly trees

The moduli spaces we use to set up wrapped Floer theory for a cylindrical Lagrangian immersion ι:LM\iota:L\to M combine two geometric configurations: both pseudoholomorphic disks and gradient flow trees. Moreover, there are two different kinds of disks - both inhomogeneous pseudoholomorphic disks and homogeneous pseudoholomorphic disks. To describe elements in the moduli spaces, we introduce the following objects as the underlying domains of ”pseudoholomorphic maps” to MM.

Definition 4.7.

A colored rooted tree with kk-leaves (k0k\geq 0) consists of the following data:

  1. (i)

    a planar oriented metric ribbon tree (T,V,E,r)(T,V,E,r) with k+1k+1 ends, where VV is the set of vertices, EE is the set of edges, and r:E0{}r:E\to\mathbb{R}_{\geq 0}\cup\{\infty\} is a length function;

  2. (ii)

    a decomposition of the set EE of edges into the set of exterior edges EextE_{ext} and the set of interior edges EintE_{int}, such that EextE_{ext} consists of k+1k+1 semi-infinite edges corresponding to the k+1k+1 ends: one is called the root e0e_{0} and the other kk are called the leaves e1,,eke_{1},\cdots,e_{k}, while the interior edges are finite;

  3. (iii)

    a coloring c:V{0,1}c:V\to\{0,1\} and a coloring d:Eext{0,1}d:E_{ext}\to\{0,1\}, which satisfy the property that if d(e)=1d(e)=1 for an enterior edge ee, then its endpoint must have color 11.

Note that we should also allow one exceptional case: TT has no vertices and only one edge ee which is infinite in both directions, with color d(e)=0d(e)=0. This does not quite fit into the definition of a colored rooted tree, but we shall still call it one. This infinite edge ee should be thought of as joining the root and one leaf together, so that it also comes with a preferred orientation.

The orientation on TT induces an orientation on every interior edge eEinte\in E_{int}, so the two endpoints of each edge can be naturally distinguished - one is called the source, denoted by s(e)s(e), the other is called the target, denoted by t(e)t(e). On the other hand, exterior edge are semi-infinite, and the root e0e_{0} has only the target t(e0)t(e_{0}) as its endpoint, while each eie_{i} has only the source s(ei)s(e_{i}).

Remark 4.8.

Note that in our definition, we do not require the valency val(v)val(v) of a vertex vv to be greater than or equal to 22. In fact, the presence of vertices vv with val(v)=1val(v)=1 will be important in the story.

Definition 4.9.

A colored rooted tree TT is called admissible, if it is obtained from a colored rooted tree T0T_{0} whose vertices all have color 11 by attachment of colored rooted trees TjT_{j} whose vertices all have color 0. These TjT_{j}’s are attached to T0T_{0} by edges eje_{j} (not the leaves) whose color are 0. Moreover, T0T_{0} and TjT_{j}’s are subtrees of TT.

For our purpose of setting up wrapped Floer theory, we shall consider only admissible colored rooted trees, and simply call them colored rooted trees by abuse of name.

Given a colored rooted tree with kk-leaves as above, we can construct from it a topological space STS_{T} canonically in the following way.

To every vertex vv, we assign a punctured disk Sv=D{zv,0,,zv,val(v)1}S_{v}=D\setminus\{z_{v,0},\cdots,z_{v,val(v)-1}\}, where each puncture zv,jz_{v,j} corresponds to an edge adjacent to vv.

To every interior edge ee, we assign a finite interval IeI_{e} of length r(e)r(e), joining the two disks (possibly with punctures) associated to s(e),t(e)s(e),t(e) at the punctures on Ss(e)S_{s(e)} and St(e)S_{t(e)} which correspond to ee. The length r(e)r(e) is allowed to be zero, in which case IeI_{e} topologically becomes a point, but we still think of ee as an edge combinatorially.

To every exterior edge ee with d(e)=0d(e)=0, we assign a semi-infinite interval Ie=(,0]I_{e}=(-\infty,0] if ee is the root e0e_{0}, or Ie=[0,+)I_{e}=[0,+\infty) if ee is any of the leaves ei,i=1,,ke_{i},i=1,\cdots,k. The semi-infinite interval IeI_{e} is attached at {0}\{0\} to the corresponding puncture on Ss(e)S_{s(e)} or St(e)S_{t(e)}.

Finally, to every exterior edge ee with d(e)=1d(e)=1, we assign a semi-infinite strip Ze=(,0]×[0,1]Z_{e}=(-\infty,0]\times[0,1] if ee is the root e0e_{0}, or Ze=[0,+)×[0,1]Z_{e}=[0,+\infty)\times[0,1] if ee is any of the leaves eie_{i}. This semi-infinite strip should be identified with a strip-like end near the corresponding puncture on Ss(e)S_{s(e)} or St(e)S_{t(e)}.

The topological space STS_{T} is the union of all the above configurations, which are glued together according to the combinatorial data of the tree TT.

One special case is when TT has only one vertex vv one root e0e_{0} and one leaf e1e_{1} with colors d(e0)=d(e1)=1d(e_{0})=d(e_{1})=1. In this case, STS_{T} should be a disk with two boundary punctures, which is identified with Z=×[0,1]Z=\mathbb{R}\times[0,1].

In the exceptional case, i.e. when TT has no vertices and only one edge ee which is infinite in both directions, we assign ST=Ie=S_{T}=I_{e}=\mathbb{R}.

In order for such geometric objects STS_{T} to have a reasonable moduli problem, we should then equip STS_{T} with an additional structure - a complex structure jvj_{v} on each disk component DvD_{v}. We briefly denote that by jj. We call (ST,j)(S_{T},j) or simply STS_{T} a (rooted) pearly tree with kk-leaves. However, these pearly trees do not have a moduli space because we do not impose stability conditions at the moment.

Pearly trees will be the underlying domains of the pseudoholomorphic maps in wrapped Floer theory of cylindrical Lagrangian immersions. However, they are not enough, as positive-dimensional families of pseudoholomorphic maps can degenerate to broken pseudoholomorphic maps. To describe those, we introduce broken colored rooted trees as well as broken pearly trees.

We need some terminology when talking about degeneration of colored rooted trees. When the length of an interior edge ee of a colored rooted tree TT tends to infinity, we obtain a pair of colored rooted trees T0,T1T_{0},T_{1}, so that ee breaks into a new leaf e0,newe_{0,new} of T0T_{0} and the root e1,0e_{1,0} of T1T_{1}. In such a picture, we say that the pair of root and leaf e1,0,e0,newe_{1,0},e_{0,new} is connected at infinity.

Now we formalize the definition of a broken colored rooted tree.

Definition 4.10.

A broken colored rooted tree is a tuple (T0,,Tm)(T_{0},\cdots,T_{m}), where each TiT_{i} is an admissible colored rooted tree, such that it satisfies the following conditions:

  1. (i)

    (rooting) The root e0,0e_{0,0} of T0T_{0} is not connected at infinity to any leaf of any Ti,i0T_{i},i\neq 0.

  2. (ii)

    (ordering) For each j0j\neq 0, there is a unique l(j)l(j) (which can be 0) such that the root ej,0e_{j,0} of TjT_{j} is connected at infinity to some (unique) leaf el(j),p(j)e_{l(j),p(j)} of Tl(j)T_{l(j)}.

  3. (iii)

    (compatible coloring) The root ej,0e_{j,0} of TjT_{j} and the leaf el(j),p(j)e_{l(j),p(j)} of Tl(j)T_{l(j)} which are connected at infinity should have the same coloring, i.e. d(ej,0)=d(el(j),p(j))d(e_{j,0})=d(e_{l(j),p(j)}).

Given a broken colored rooted tree (T0,,Tm)(T_{0},\cdots,T_{m}) as above, as well as mm positive real numbers ρ1,,ρm\rho_{1},\cdots,\rho_{m}, we may perform a gluing construction as follows. For each Tj,j0T_{j},j\neq 0 and the corresponding l(j)l(j), recall that we have identifications

(4.4) ej,0(,0],el(j),p(j)[0,+),\begin{split}&e_{j,0}\cong(-\infty,0],\\ &e_{l(j),p(j)}\cong[0,+\infty),\end{split}

cut off (,ρj/2](-\infty,-\rho_{j}/2] from ej,0e_{j,0} and [ρj/2,+)[\rho_{j}/2,+\infty) from el(j),p(j)e_{l(j),p(j)}, and glue the remaining intervals at the endpoints {ρj/2}{ρj/2}\{-\rho_{j}/2\}\sim\{\rho_{j}/2\}. We may suitably reparametrize the interval so it has a nicer form, but that is not important; the only important information is that the resulting edge has length ρj\rho_{j}. After doing this process for all j=1,,mj=1,\cdots,m, we obtain a colored rooted tree

(4.5) T=ρ1,,ρm(T0,,Tm).T=\sharp_{\rho_{1},\cdots,\rho_{m}}(T_{0},\cdots,T_{m}).

If the resulting colored rooted tree TT is admissible, we call this an admissible gluing, and call (T0,,Tm)(T_{0},\cdots,T_{m}) an admissible broken colored rooted tree. From now on we shall only consider admissible broken colored rooted trees, and call them broken colored rooted trees for simplicity. Partial gluings are also allowed, which again give us broken colored rooted trees. It can be defined in a similar way, but the gluing process is only done for a sub-collection of edges connected at infinity. Let J{1,,m}J\subset\{1,\cdots,m\} index such a sub-collection, and we denote the result of partial gluing by

(4.6) ρj:jJ(T0,,Tm).\sharp_{\rho_{j}:j\in J}(T_{0},\cdots,T_{m}).

Similar to the case of a colored rooted tree, we can assign to a broken colored rooted tree a topological space as follows.

Definition 4.11.

A broken pearly tree (ST0,,STm)(S_{T_{0}},\cdots,S_{T_{m}}) associated to a broken colored rooted tree (T0,,Tm)(T_{0},\cdots,T_{m}) is simply the union of pearly trees STiS_{T_{i}} associated to each component.

As the underlying broken colored rooted tree (T0,,Tm)(T_{0},\cdots,T_{m}) can be glued root-to-leaf in an admissible way, we can also glue the associated broken pearly tree to get a pearly tree. There are two cases. If d(ej,0)=d(el(j),p(j))=0d(e_{j,0})=d(e_{l(j),p(j)})=0, the 0-th end ϵj,0\epsilon_{j,0} of STjS_{T_{j}} is the negative half-ray (,0](-\infty,0] and the p(j)p(j)-th end ϵl(j),p(j)\epsilon_{l(j),p(j)}-th is the positive half-ray [0,+)[0,+\infty). In this case perform the gluing in the same way as we have done for the underlying trees. If d(ej,0)=d(el(j),p(j))=1d(e_{j,0})=d(e_{l(j),p(j)})=1, the 0-th end ϵj,0\epsilon_{j,0} of STjS_{T_{j}} is the negative infinite half-strip (,0]×[0,1](-\infty,0]\times[0,1] and the p(j)p(j)-th end ϵl(j),p(j)\epsilon_{l(j),p(j)}-th is the positive infinite half-strip [0,+)×[0,1][0,+\infty)\times[0,1]. We cut off (,ρj/2]×[0,1](-\infty,\rho_{j}/2]\times[0,1] from ϵj,0\epsilon_{j,0} and [ρj/2,+)×[0,1][\rho_{j}/2,+\infty)\times[0,1] from ϵl(j),p(j)\epsilon_{l(j),p(j)}, then glue the resulting finite strips along the boundary intervals {ρj/2}×[0,1]{ρj/2}×[0,1]\{-\rho_{j}/2\}\times[0,1]\sim\{\rho_{j}/2\}\times[0,1]. Doing this process for all jj, we obtain a pearly tree denoted by

(4.7) ST=ρ1,,ρm(ST0,,STm).S_{T}=\sharp_{\rho_{1},\cdots,\rho_{m}}(S_{T_{0}},\cdots,S_{T_{m}}).

Also, we can perform partial gluing in a similar way as we have done for broken colored rooted trees.

4.5. Moduli spaces of stable pearly trees

The moduli spaces involved in wrapped Floer theory for the Lagrangian immersion ι:LM\iota:L\to M are analogues and modifications of those used by [BO16] to set up linearized contact homology in Hamiltonian formulation, without circle action and symmetry in our case. In addition, the inhomogeneous pseudoholomorphic disks in the interior part which have asymptotic limits being the constant chords are also in consideration.

Two issues bring up complication in the construction of the moduli spaces. Unlike the case of an embedded exact Lagrangian submanifold, the image ι(L)\iota(L) in general bounds JJ-holomorphic disks, and limits of inhomogeneous pseudoholomorphic disks might bubble off homogeneous pseudoholomorphic disks with boundary on ι(L)\iota(L). These should be suitably packaged into the moduli spaces. On the other hand, the AA_{\infty}-structure maps are typically defined by appropriate counts of inhomogeneous pseudoholomorphic disks; in particular, we expect the zeroth order map m0m^{0} of the curved AA_{\infty}-algebra structure to be defined by inhomogeneous pseudoholomorphic disks with one puncture. This causes some potential problems, in particular in the verification of AA_{\infty}-relations, as the elements in some boundary strata do not satisfy Floer’s equations because they arise from disk bubbling.

In order to treat homogeneous and inhomogeneous pseudoholomorphic disks with one puncture in a uniform way, we have chosen our Hamiltonian HH to be constantly zero in the interior part of MM, so that inhomogeneous pseudoholomorphic disks which are contained in the interior part automatically satisfy the homogeneous Cauchy-Riemann equation. Because of the chain model we pick for the wrapped Floer cochain space, we shall construct a version of Morse-Bott moduli spaces combining pseudoholomorphic disks and gradient flow trees.

The issues can be resolved using the exactness condition for the Lagrangian immersion ι:LM\iota:L\to M. The key properties about the behavior of pseudoholomorphic disks are given by the following lemmas.

Lemma 4.12.

For any compatible almost complex structure JJ of contact type near the boundary M\partial M, all JJ-holomorphic disks with boundary on ι(L)\iota(L) are contained in the interior part of MM, and have to pass through a self-intersection point of ι:LM\iota:L\to M.

For any self-intersection point (p,p+)(p_{-},p_{+}) (pp+p_{-}\neq p_{+}) such that ι(p)=ι(p+)=p\iota(p_{-})=\iota(p_{+})=p, there are finitely many relative homology classes β\beta of JJ-holomorphic disks bounded by ι(L)\iota(L), with one boundary marked point mapped to the point pp.

Proof.

The first statement follows from the assumption that ι\iota is an embedding over the cylindrical end M×[1,+)\partial M\times[1,+\infty) of an exact Lagrangian submanifold of the form l×[1,+)l\times[1,+\infty), using a standard argument by the maximum principle.

The second statement follows from the exactness condition. For any such a JJ-holomorphic disk uu, the exactness condition df=ιλMdf=\iota^{*}\lambda_{M} implies that its energy is fixed:

E(u)=f(p)f(p+),E(u)=f(p_{-})-f(p_{+}),

by integration by parts. On the other hand, the energy is also equal to ω(β)\omega(\beta), which implies that there can only be finitely many such homology classes β\beta.

The next three lemmas all follow from maximum principle.

Lemma 4.13.

Let u:DMu:D\to M be a JJ-holomorphic disk with boundary on ι(L)\iota(L). Then u(D)u(D) cannot be entirely contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty).

Lemma 4.14.

Let u:SMu:S\to M be a JJ-holomorphic curve with boundary on ι(L)\iota(L), for any smooth Riemann surface SS with boundary S\partial S\neq\varnothing. If u(D)u(\partial D) is contained in a compact subset of the interior part of MM away from M\partial M, then the image of the entire disk u(D)u(D) must be contained in the interior part of MM away from M\partial M.

Lemma 4.15.

Let u:SMu:S\to M be a smooth map from a one-punctured disk to MM with boundary on ι(L)\iota(L), which satisfies the Floer’s equation

(4.8) (duγXH)0,1=0.(du-\gamma\otimes X_{H})^{0,1}=0.

If uu asymptotically converges to either a constant HH-chord, or a self-intersection point of ι\iota at the puncture, then u(S)u(S) must be entirely contained in the interior part of MM, where the Hamiltonian HH vanishes. In particular, uu has a natural smooth extension to the closed disk DD, and satisfies the JJ-holomorphic curve equation

(4.9) (du)0,1=0(du)^{0,1}=0

in the interior of DD.

Now let us proceed to describe elements in the moduli spaces. To write down the inhomogeneous Cauchy-Riemann equations, we shall need certain geometric data, e.g. Morse functions, Hamiltonians, and almost complex structures. Fix the original Hamiltonian HH, and let J=JL,ιJ=J_{L,\iota} be a compatible almost complex structure of contact type. The various geometric data needed for writing down the relevant equations are packaged in the following way.

Definition 4.16.

A Floer datum on a pearly tree STS_{T} consists of the following data:

  1. (i)

    A time-shifting function ρST:ST[1,+)\rho_{S_{T}}:\partial S_{T}\to[1,+\infty), where ST\partial S_{T} denotes the union of boundary components of the disk components of STS_{T}, as well as the intervals. ρST\rho_{S_{T}} should be equal to a constant over the strip-like end near each puncture.

  2. (ii)

    For each vertex vv with c(v)=0c(v)=0, a constant family of almost complex structures JSv=JJ_{S_{v}}=J.

  3. (iii)

    For each vertex vv with c(v)=1c(v)=1 and val(v)=1val(v)=1, a constant family of Hamiltonians HSv=HH_{S_{v}}=H and a constant family of almost complex structures JSv=JJ_{S_{v}}=J.

  4. (iv)

    For each vertex vv with c(v)=1c(v)=1 and val(v)=2val(v)=2, a time-dependent family of almost complex structures Jv,tJ_{v,t}, rescaled by weight wv,0=wv,1w_{v,0}=w_{v,1}. This is a time-dependent perturbation of JJ in the class of almost complex structures of contact type. Moreover, we require that the choices Jv,t=JtJ_{v,t}=J_{t} be the same for all such vertices vv, but possibly rescaled by differente weights according to the values of ρST\rho_{S_{T}}.

  5. (v)

    For each vertex vv with c(v)=1c(v)=1 and val(v)3val(v)\geq 3, a domain-dependent family of Hamiltonians HSvH_{S_{v}} and a domain-dependent family of almost complex structures JSvJ_{S_{v}}, such that HSvH_{S_{v}} and JSvJ_{S_{v}} asymptotically agrees with HH and respectively JtJ_{t} rescaled by weight wv,jw_{v,j}, over the strip-like end near each puncture zv,jz_{v,j}. Moreover, we require that the family of Hamiltonians HSvH_{S_{v}} be a compactly-supported domain-dependent perturbation of HH, i.e. HSv=HH_{S_{v}}=H in a neighborhood of the boundary Sv\partial S_{v}.

  6. (vi)

    For each interior edge ee, an ss-dependent family of Morse functions fe,sf_{e,s} on IeI_{e}.

  7. (vii)

    For each exterior edge ee with d(e)=0d(e)=0, a family of Morse functions fe,sf_{e,s} on LL parametrized by sIes\in I_{e}, which agrees with ff for |s|0|s|\gg 0.

  8. (viii)

    For each exterior edge ee with d(e)=1d(e)=1, a family of time-dependent Hamiltonians He,s,t=He,tH_{e,s,t}=H_{e,t}, and a family of time-dependent almost complex structures Je,s,t=Je,tJ_{e,s,t}=J_{e,t} which agree with HH and respectively JJ for |s|0|s|\gg 0. Moreover, as the semi-infinite strips are glued to the punctured disk Ss(e)S_{s(e)} or St(e)S_{t(e)}, we require that these data extend smoothly over the glued domain.

In the exceptional case where TT does not have vertices and has a single infinite edge ee of color d(e)=0d(e)=0, so that Ie=I_{e}=\mathbb{R}, we require fe,s=ff_{e,s}=f for all ss.

In order to define the AA_{\infty}-structures, a necessary condition is to make sure that the Floer data chosen for various on various pearly trees satisfy certain consistency conditions.

The simplest consistency condition to state is when two colored rooted trees T1,T2T_{1},T_{2} are glued together root-to-leaf in a way that the resulting colored rooted tree T=T10,i,ρT2T=T_{1}\sharp_{0,i,\rho}T_{2} is still admissible. The consistency condition means that the Floer datum on STS_{T} is obtained from gluing the Floer data on ST1S_{T_{1}} and ST2S_{T_{2}}. There are higher consistency conditions, which we refer the reader to [Sei08] for a detailed description.

Remark 4.17.

Such consistency conditions make sense because the Floer datum is designed so that perturbations are not put on the punctured disks SvS_{v} for vertices vv with c(v)=0c(v)=0, that is, the family of almost complex structure JSvJ_{S_{v}} on such a component is constant equal to JJ.

Furthermore, a consistent choice of Floer data is not for the purpose of achieving transversality, even the choices are ”generic” for all the components that allow non-constant families of perturbations of Hamiltonians and almost complex structures. Such a consistent choice is made mainly to ensure that the moduli spaces of pearly trees have good compactifications so that the boundary strata of the compactifications are products of moduli spaces of the same type, as we shall see below.

Having made a consistent choice of Floer data as above, we now describe elements in the moduli spaces.

Definition 4.18.

Let I{0,,k}I\subset\{0,\cdots,k\} be a subset. Let α:IS(L,ι)\alpha:I\to S(L,\iota) be a map, labeling those marked points which are mapped to some self-intersection point of ι\iota. A stable pearly tree map is a triple (ST,u,l)(S_{T},u,l) satisfying the following conditions:

  1. (i)

    STS_{T} is a pearly tree modeled on a colored rooted tree TT with kk-leaves.

  2. (ii)

    u:STMu:S_{T}\to M is a continuous map.

  3. (iii)

    For a vertex vv with c(v)=0c(v)=0, let uvu_{v} be the restriction of uu to the punctured disk SvS_{v} associated to vv. Then uvu_{v} satisfies the homogeneous Cauchy-Riemann equation

    (4.10) (duv)0,1=0,(du_{v})^{0,1}=0,

    with respect to the family JSvJ_{S_{v}} of almost complex structures. In case val(v)=2val(v)=2, JSv=JtJ_{S_{v}}=J_{t}.

  4. (iv)

    For a vertex vv with c(v)=1c(v)=1, let uvu_{v} be the restriction of uu to the punctured disk SvS_{v} associated to vv. Then uvu_{v} satisfies the inhomogeneous Cauchy-Riemann equation

    (4.11) (duvγvXHSv)0,1=0.(du_{v}-\gamma_{v}\otimes X_{H_{S_{v}}})^{0,1}=0.
  5. (v)

    For an interior edge ee, let ueu_{e} be the restriction of uu to the interval IeI_{e} associated to ee. Then ueu_{e} comes with a preferred lift u~e:IeL\tilde{u}_{e}:I_{e}\to L and satisfies the gradient flow equation:

    (4.12) du~eds+fe,s(u~e)=0.\frac{d\tilde{u}_{e}}{ds}+\nabla f_{e,s}(\tilde{u}_{e})=0.

    In the exceptional case where ST=S_{T}=\mathbb{R}, this gradient flow equation takes the form

    (4.13) du~eds+f(u~e)=0,\frac{d\tilde{u}_{e}}{ds}+\nabla f(\tilde{u}_{e})=0,

    by our choice of Floer datum.

  6. (vi)

    u(z)ϕMρST(z)ι(L)u(z)\in\phi_{M}^{\rho_{S_{T}}(z)}\iota(L), for zSTz\in\partial S_{T}.

  7. (vii)

    l:STLl:\partial S_{T}\to L is a continuous map, which specifies the boundary lifting condition, i.e. ιl=u|ST\iota\circ l=u|_{\partial S_{T}}.

  8. (viii)

    If ee is an exterior edge such that IeI_{e} is either half-infinite or \mathbb{R}, then lims±u~e(s,)=p\lim\limits_{s\to\pm\infty}\tilde{u}_{e}(s,\cdot)=p for some critical point pp of ff.

  9. (ix)

    If eie_{i} is the ii-th exterior edge with color d(ei)=1d(e_{i})=1, lims±uv(ei)ϵi(s,)=ϕMwixi()\lim\limits_{s\to\pm\infty}u_{v(e_{i})}\circ\epsilon_{i}(s,\cdot)=\phi_{M}^{w_{i}}x_{i}(\cdot), where v(ei)v(e_{i}) is the endpoint of ee, and ϵi\epsilon_{i} is the strip-like end associated to eie_{i}. Here xix_{i} is some non-constant time-one HH-chord from ι(L)\iota(L) to itself.

  10. (x)

    (limθ0l(e1θζi),limθ0l(e1θζi))=α(i)S(L,ι)(\lim\limits_{\theta\uparrow 0}l(e^{\sqrt{-1}\theta}\zeta_{i}),\lim\limits_{\theta\downarrow 0}l(e^{\sqrt{-1}\theta}\zeta_{i}))=\alpha(i)\in S(L,\iota) for iIi\in I. In addition, ι(α(i))=u~ei(zv(ei),j(v(ei))\iota(\alpha(i))=\tilde{u}_{e_{i}}(z_{v(e_{i}),j(v(e_{i})}). Here u~ei(zv(ei),j(v(ei)))\tilde{u}_{e_{i}}(z_{v(e_{i}),j(v(e_{i}))}) means the asymptotic value of u~ei\tilde{u}_{e_{i}} at the end of the half-infinite ray associated to the ii-th exterior edge eie_{i}, where zv(ei),j(v(ei))z_{v(e_{i}),j(v(e_{i}))} is the puncture on the punctured disk Sv(ei)S_{v(e_{i})} that corresponds to the exterior edge eie_{i}.

  11. (xi)

    If d(e)=0d(e)=0, then the preferred lift u~e\tilde{u}_{e} of ueu_{e} should be compatible with the restriction of ll to IeI_{e}, such that the above condition hold. That is, if ue~\tilde{u_{e}} is a constant map to a component of S(L,ι)S(L,\iota), i.e. a discrete point, then the restriction of ll to IeI_{e} should satisfy the same limiting condition in (x)(x).

  12. (xii)

    The homology class of uu together with its asymptotic non-constant Hamiltonian chords oppositely oriented, is βH2(M,ι(L);)\beta\in H_{2}(M,\iota(L);\mathbb{Z}).

The last condition in the definition of a stable pearly tree map needs some explanation. Since some of the asymptotic convergence conditions are non-constant Hamiltonian chords, which might be non-contractible relative ι(L)\iota(L), the map uu does not define a homology class in H2(M,ι(L);)H_{2}(M,\iota(L);\mathbb{Z}). However, if we compactify STS_{T} at infinity by adding {±}×[0,1]\{\pm\infty\}\times[0,1] to all exterior strip-like ends, and glue the oppositely oriented Hamiltonian chords to the map uu, then the resulting map defines a homology class in H2(M,ι(L);)H_{2}(M,\iota(L);\mathbb{Z}).

There is an obvious notion of two triples (ST,u,l)(S_{T},u,l) and (ST,u,l)(S^{\prime}_{T^{\prime}},u^{\prime},l^{\prime}) being isomorphic. First of all, there should be an isomorphism of rooted colored trees ϕ¯:TT\bar{\phi}:T\to T^{\prime}, an isomorphism ϕ:STST\phi:S_{T}\to S^{\prime}_{T^{\prime}} compatible with ϕ¯\bar{\phi}, such that uϕ=uu^{\prime}\circ\phi=u, and the pullback of the Floer datum chosen for STS^{\prime}_{T^{\prime}} by ϕ\phi agrees with that for STS_{T}.

Denote by cc either a pair (p,w)(p,w), where pp is a critical point of ff and ww is a homotopy class of capping half-disk, or a pair (x,w)(x,w), where xx is a non-constant HH-chord, and ww is a homotopy class of capping half-disk with reference to some chosen based point in the connected component of xx in the space 𝒫(M;ι(L))\mathcal{P}(M;\iota(L)) of paths in MM from ι(L)\iota(L) to itself. Let k+1(α,β;J,H;c0,,ck)\mathcal{M}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) be the moduli space of the above triples (ST,u,l)(S_{T},u,l) which satisfy the convergence conditions to C0,,ckC_{0},\cdots,c_{k} at the root e0e_{0} and leaves e1,,eke_{1},\cdots,e_{k}. Clearly, there is a decomposition according to the combinatorial type of TT:

(4.14) k+1(α,β;J,H;c0,,ck)=TT(α,β;J,H;c0,,ck).\mathcal{M}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k})=\coprod_{T}\mathcal{M}_{T}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}).

This description is only for the purpose of visualizing elements in the moduli space, but not for the purpose of proving transversality results by induction on combinatorial types.

4.6. Compactification: stable broken pearly trees

The moduli space of stable pearly trees k+1(α,β;J,H;c0,,ck)\mathcal{M}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is generally non-compact, from which one cannot expect to extract invariants. In order to define the desired curved AA_{\infty}-algebra whose structure constants are given by appropriate counts of elements in the moduli space, we need to compactify it.

Lemma 4.19.

The moduli space k+1(α,β;J,H;c0,,ck)\mathcal{M}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) has a natural compactification ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), which has a natural topology being compact and Hausdorff.

Lemma 4.20.

The compactification ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) consists of products of moduli spaces ¯ki+1(αi,βi;J,H;ci,0,,ci,ki)\bar{\mathcal{M}}_{k_{i}+1}(\alpha_{i},\beta_{i};J,H;c^{\prime}_{i,0},\cdots,c^{\prime}_{i,k_{i}}) of the same type. More precisely, the codimension-mm strata

Sm¯k+1(α,β;J,H;c0,,ck),S_{m}\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}),

i.e. those points of codimension exactly mm consist of the union of fiber products of the form

(4.15) ¯k0+1(α0,β0;J,H;c0,0,,c0,k0)×ׯkm+1(αm,βm;J,H;cm,0,,cm,km),\bar{\mathcal{M}}_{k_{0}+1}(\alpha_{0},\beta_{0};J,H;c_{0,0},\cdots,c_{0,k_{0}})\times\cdots\times\bar{\mathcal{M}}_{k_{m}+1}(\alpha_{m},\beta_{m};J,H;c_{m,0},\cdots,c_{m,k_{m}}),

where cj=cij,kjc_{j}=c_{i_{j},k_{j}} for some ij,kji_{j},k_{j}, for every j=0,,kj=0,\cdots,k. Here fiber products are taken over a discrete set of generators, so they are written as products. In fact, by the inductive construction, for any element in the codimension-mm strata, the underlying domain is a broken pearly tree (ST0,,STm)(S_{T_{0}},\cdots,S_{T_{m}}) consisting of exactly m+1m+1 pearly trees. Denote by uju_{j} the restriction of the stable map on STjS_{T_{j}}. For each j0j\neq 0, there is a unique l(j)l(j) for which uju_{j} converges over the 0-th end ϵj,0\epsilon_{j,0} (corresponding to the root) of STjS_{T_{j}} to the same generator as ul(j)u_{l(j)} converges over the end ϵl(j),p(j)\epsilon_{l(j),p(j)} (corresponding to some leaf) of STl(j)S_{T_{l(j)}}. That is to say, cj,0=cl(j),p(j)c_{j,0}=c_{l(j),p(j)} for some p(j)p(j).

Sketch of proof of Lemma 4.19 and Lemma 4.20.

The definition of the compactification is given in the statement of Lemma 4.20, which is valid because of the Gromov compactness theorem. This inductive fiber product structure makes sense, because the various factors satisfy an induction condition with respect to (E,k)(E,k) in lexicographic order, where EE the energy of the stable maps from broken pearly trees, and kk is the number of leaves. The fact that ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is compact follows from the standard maximum principle and Gromov compactness theorem, as explained below.

The proof of the rest is basically standard, so we shall simply explain the crucial reason. Fixing the homology class β\beta provides a priori energy bound for all elements in the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), so that there are only finitely many types of product moduli spaces that appear in the boundary strata of the compactfication. Thus by the maximum principle, there is a compact subset CC of MM depending on k,α,β,J,Hk,\alpha,\beta,J,H and the asymptotic convergence conditions c0,,ckc_{0},\cdots,c_{k} but not on individual maps in the moduli space, so that any element in ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is represented by a stable map whose image is contained in CC. We may take CC to be some sub-level set {rA}\{r\leq A\} by possibly enlarging the subset. Then Gromov compactness theorem applies.

Hausdorff-ness follows from the stability condition. For a detailed proof of Hausdorff-ness, we refer the reader to [FOOO09b].

Note that ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) naturally compactifies the Gromov bordification of the moduli space of inhomogeneous pseudoholomorphic disks with boundary on ι(L)\iota(L), encoding additional data that specify the boundary lifts to the preimage of the immersion, as introduced in [AJ10]. Here by the Gromov bordification, we mean the moduli space of broken stable inhomogeneous pseudoholomorphic disks as those in [Abo10] (see also [Sei08] in the unwrapped setting), which in our case is not compact because the limit of a sequence of a broken stable inhomogeneous pseudoholomorphic disk can bubble off homogeneous pseudoholomorphic disks with one marked points (such disks necessarily pass through some self-intersection point of ι:LM\iota:L\to M. In words, the above compactification is obtained by adding all such disks.

4.7. Kuranishi structures on the moduli spaces of stable pearly tree maps

Appropriate perturbation framework is required to prove transversality results for the moduli spaces

¯k+1(α,β;J,H;c0,,ck),\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}),

so that we can extract algebraic structures from these moduli spaces. Because of the presence of disk bubbles, in particular those disks with one marked point, unless we impose further geometric assumptions, it seems difficult to use traditional transversality methods, which use generic choices of Floer data which depend underlying Riemann surfaces in the underlying moduli spaces of domains. However, the domains of those pseudoholomorphic disks with one marked point are unstable, which do not form a good moduli space. The usual inductive argument for a universal and consistent choice of Floer data would fail, as gluing in a disk with one marked point will decrease the number of marked points. To overcome such difficulty, our strategy is to appropriately adapt the theory of Kuranishi structures [FO99] [FOOO09a], [FOOO09b], [FOOO12], [FOOO13] developed by Fukaya-Ono, Fukaya-Oh-Ohta-Ono. In our paper, nothing new about the theory of Kuranishi structures is invented; we shall only use their results to construct virtual fundamental chains of the relevant moduli spaces, so the argument is sketchy leaving technical details these monumental works.

Proposition 4.21.

The moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) has an oriented Kuranishi structure with boundary and corners. Moreover, the induced Kuranishi structure on the codimension-mm stratum Sm¯k+1(α,β;J,H;c0,,ck)S_{m}\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) agrees with the fiber product Kuranishi structures on the fiber products (4.15).

Sometimes we shall call a Kuranishi structure with boundary and corners briefly a Kuranishi structure for short. The proof of this proposition is based on minor modification of the momentous machinery developed in [FOOO09b], as the basic idea is the same despite the fact that the form of the moduli space presented is slightly different. As this is not the main focus of the current paper, we shall only outline the main steps. We expect to provide a more detailed treatment in the work on analytic setup of the currently used ”Morse-Bott” type setup of wrapped Floer theory.

The construction of a Kuranishi structure on ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is of an inductive manner, for which we outline the main steps as follows. For more detailed argument, in particular various estimates to make the gluing construction work, see Chapter 7 of [FOOO09b].

Given the geometric data α,β,J,H,c0,,ck\alpha,\beta,J,H,c_{0},\cdots,c_{k}, there is a uniform upper bound on the energy of all elements in the moduli space, say bounded by EE. The stable maps in

¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k^{\prime}+1}(\alpha^{\prime},\beta^{\prime};J,H;c^{\prime}_{0},\cdots,c^{\prime}_{k^{\prime}})

which appears as a factor of the fiber product (4.15) either have less energy, of fewer leaves. This fact allows one to build the Kuranishi structure by induction on lexicographic order of energy and number of leaves.

To build Kuranishi charts, we first look at a subset

(4.16) ¯k+1reg(α,β;J,H;c0,,ck)¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}^{reg}(\alpha,\beta;J,H;c_{0},\cdots,c_{k})\subset\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k})

consisting of elements whose underlying domain is a single pearly tree (these are called smooth points in the moduli space).

Lemma 4.22.

There exists a Kuranishi chart (U(ST,u,l),E(ST,u,l),s(ST,u,l),Γ(ST,u,l))(U(S_{T},u,l),E_{(S_{T},u,l)},s_{(S_{T},u,l)},\Gamma_{(S_{T},u,l)}) at each point [ST,u,l]¯k+1reg(α,β;J,H;c0,,ck)[S_{T},u,l]\in\bar{\mathcal{M}}_{k+1}^{reg}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), where the isotropy group Γ(ST,u,l)\Gamma_{(S_{T},u,l)} is trivial.

Proof.

For each point [ST,u,l]¯k+1reg(α,β;J,H;c0,,ck)[S_{T},u,l]\in\bar{\mathcal{M}}_{k+1}^{reg}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), the choice of Floer datum for STS_{T} defines a non-degenerate Fredholm problem, which is a combination (direct sum) of inhomogeneous Cauchy-Riemann operators and the linearized operators of the gradient flow equations. Let \mathcal{B} be the Banach manifold of consisting maps (ST,u,l)(S_{T},u,l) (without satisfying the equations (4.11) and (4.12)) of suitable weighted Sobolev class W1,pW^{1,p}, and \mathcal{E}\to\mathcal{B} be the Banach bundle whose fiber over (ST,u,l)(S_{T},u,l) is the space of one-forms on STS_{T} with values in uTMu^{*}TM, which are of class Wloc0,pW^{0,p}_{loc}. The equations (4.11) and (4.12) combine to define a section of this bundle, briefly denoted by ¯\bar{\partial}. The linearization at (ST,u,l)(S_{T},u,l) is the previously mentioned Fredholm operator

(4.17) D(ST,u,l)¯:T(ST,u,l)(ST,u,l).D_{(S_{T},u,l)}\bar{\partial}:T_{(S_{T},u,l)}\mathcal{B}\to\mathcal{E}_{(S_{T},u,l)}.

Choose a finite-dimensional subspace E(ST,u,l)E_{(S_{T},u,l)} of (ST,u,l)\mathcal{E}_{(S_{T},u,l)} containing the cokernel, so that

(4.18) D(ST,u,l)¯:T(ST,u,l)(ST,u,l)/E(ST,u,l)D_{(S_{T},u,l)}\bar{\partial}:T_{(S_{T},u,l)}\mathcal{B}\to\mathcal{E}_{(S_{T},u,l)}/E_{(S_{T},u,l)}

is surjective. A careful choice of E(ST,u,l)E_{(S_{T},u,l)}, based on the unique continuation theorem of solutions to elliptic equations, can be made that every vector in E(ST,u,l)E_{(S_{T},u,l)} is smooth and has compact support in the underlying domain STS_{T}, meaning that the support of every vector in E(ST,u,l)E_{(S_{T},u,l)} is contained in the union of the interiors of punctured disk components and the interiors of compact interval components. In fact, we may take E(ST,u,l)E_{(S_{T},u,l)} so that every vector vanishes on the interval components, by choosing the families of Morse functions generic enough. However, we will not use this fact in our construction.

Define U((ST,u,l))U((S_{T},u,l)) to be the space of nearby elements (ST,u,l)(S^{\prime}_{T^{\prime}},u^{\prime},l^{\prime})\in\mathcal{B} such that

(4.19) ¯((ST,u,l))E(ST,u,l),\bar{\partial}((S^{\prime}_{T^{\prime}},u^{\prime},l^{\prime}))\in E_{(S_{T},u,l)},

where E(ST,u,l)E_{(S_{T},u,l)} is identified with a subspace of (ST,u,l)\mathcal{E}_{(S^{\prime}_{T^{\prime}},u^{\prime},l^{\prime})} by parallel transport along minimal geodesics. This will be a Kuranishi neighborhood at [ST,u,l][S_{T},u,l], over which the obstruction space is E(ST,u,l)E_{(S_{T},u,l)}, so that the Kuranishi map is defined by ¯\bar{\partial}. Regarding the isotropy group, we may take Γ(ST,u,l)={1}\Gamma_{(S_{T},u,l)}=\{1\} to be the trivial group. This choice is fine because of the following lemma.

Lemma 4.23.

The automorphism group of a semistable pearly tree is torsion-free.

Next, we shall extend the construction to other strata in the compactification

¯k+1(α,β;J,H;c0,,ck).\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}).

We need to construct a Kuranishi chart at each point in

¯k+1(α,β;J,H;c0,,ck)¯k+1reg(α,β;J,H;c0,,ck).\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k})\setminus\bar{\mathcal{M}}^{reg}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}).

This is standard application of gluing theorems, as explained in [FOOO09b], which we briefly recall here. For simplicity, we consider the case of a pair of pearly tree maps [(ST0,u0,l0),(ST1,u1,l1)][(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})], which is an element in the codimension-one strata, such that the root e1,0e_{1,0} of ST1S_{T_{1}} can be glued to some leaf e0,je_{0,j} of ST0S_{T_{0}} in an admissible way, and the asymptotic convergence conditions of u0u_{0} at e0,je_{0,j} and of u1u_{1} at e1,0e_{1,0} agree.

Lemma 4.24.

At every such point [(ST0,u0,l0),(ST1,u1,l1)][(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})] in the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), there exists a Kuranishi chart

(U((ST0,u0,l0),(ST1,u1,l1)),E[(ST0,u0,l0),(ST1,u1,l1)],s[(ST0,u0,l0),(ST1,u1,l1)],Γ[(ST0,u0,l0),(ST1,u1,l1)]),\begin{split}&(U((S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})),E_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]},\\ &s_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]},\Gamma_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]}),\end{split}

such that Γ[(ST0,u0,l0),(ST1,u1,l1)]\Gamma_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]} is trivial.

Proof.

Consider the two Banach bundles ii\mathcal{E}_{i}\to\mathcal{B}_{i}, together with choices of obstruction spaces E(STi,ui,lj)E_{(S_{T_{i}},u_{i},l_{j})} making the linearized section surjective onto the quotient space (i)(STi,ui,lj)/(Ei)(STi,ui,lj)(\mathcal{E}_{i})_{(S_{T_{i}},u_{i},l_{j})}/(E_{i})_{(S_{T_{i}},u_{i},l_{j})} at the point (STi,ui,li)(S_{T_{i}},u_{i},l_{i}).

Consider the direct sum linearized operator

(4.20) D(ST0,u0,l0)¯0D(ST1,u1,l1)¯1:T(ST0,u0,l0)0T(ST1,u1,l1)1(0)(ST0,u0,l0)(1)(ST1,u1,l1).D_{(S_{T_{0}},u_{0},l_{0})}\bar{\partial}_{0}\oplus D_{(S_{T_{1}},u_{1},l_{1})}\bar{\partial}_{1}:T_{(S_{T_{0}},u_{0},l_{0})}\mathcal{B}_{0}\oplus T_{(S_{T_{1}},u_{1},l_{1})}\mathcal{B}_{1}\to(\mathcal{E}_{0})_{(S_{T_{0}},u_{0},l_{0})}\oplus(\mathcal{E}_{1})_{(S_{T_{1}},u_{1},l_{1})}.

After projecting to the quotient spaces, we obtain a surjective Fredholm operator

(4.21) D(ST0,u0,l0),(ST1,u1,l1)¯:T(ST0,u0,l0)0T(ST1,u1,l1)1(0)(ST0,u0,l0)/(E0)(ST0,u0,l0)(1)(ST1,u1,l1)/(E1)(ST1,u1,l1).\begin{split}&D_{(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})}\bar{\partial}:T_{(S_{T_{0}},u_{0},l_{0})}\mathcal{B}_{0}\oplus T_{(S_{T_{1}},u_{1},l_{1})}\mathcal{B}_{1}\\ &\to(\mathcal{E}_{0})_{(S_{T_{0}},u_{0},l_{0})}/(E_{0})_{(S_{T_{0}},u_{0},l_{0})}\oplus(\mathcal{E}_{1})_{(S_{T_{1}},u_{1},l_{1})}/(E_{1})_{(S_{T_{1}},u_{1},l_{1})}.\end{split}

To make sure that the Kuranishi chart at the point [(ST0,u0,l0),(ST1,u1,l1)][(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})] is compatible with the Kuranishi charts at nearby points that we have previously constructed, we need the following argument. Let (ST,u,l)(S_{T},u^{\prime},l) be the pre-gluing of (ST0,u0,l0)(S_{T_{0}},u_{0},l_{0}) and (ST1,u1,l1)(S_{T_{1}},u_{1},l_{1}) for some small gluing parameter ρ\rho. Here by pre-gluing, we mean gluing the two maps together on the glued pearly tree using cut-off functions to obtain a continuous map, which might not satisfy the desired equations on the glued region. For the underlying pearly trees, the process of gluing is easily described. Regarding the map uu^{\prime}, it is the result of pre-gluing gradient flow lines if the common asymptote of u0u_{0} and u1u_{1} is a critical point of ff, or the result of pre-gluing solutions to Floer’s equation if the common asymptote of u0u_{0} and u1u_{1} is a time-one Hamiltonian chord for HH.

Using the cut-off functions in the above gluing process, we may define maps

(4.22) Ii:(STi,ui,li)(ST,u,l),I_{i}:\mathcal{E}_{(S_{T_{i}},u_{i},l_{i})}\to\mathcal{E}_{(S_{T},u^{\prime},l)},

by cutting off the domain of the one-forms on STiS_{T_{i}} and including the resulting domain to STS_{T}. Recall that we have chosen the obstruction spaces (Ei)(STi,ui,li)(E_{i})_{(S_{T_{i}},u_{i},l_{i})} so that every vector has compact support. Therefore we may choose E(ST,u,l)E_{(S_{T},u^{\prime},l)} carefully so that the restriction of sum of I0I_{0} and I1I_{1}:

(4.23) I0+I1:(E0)(ST0,u0,l0)(E1)(ST1,u1,l1)E(ST,u,l)I_{0}+I_{1}:(E_{0})_{(S_{T_{0}},u_{0},l_{0})}\oplus(E_{1})_{(S_{T_{1}},u_{1},l_{1})}\to E_{(S_{T},u^{\prime},l)}

is injective. In this way we may regard

(4.24) (E0)(ST0,u0,l0)(E1)(ST1,u1,l1)E(ST,u,l).(E_{0})_{(S_{T_{0}},u_{0},l_{0})}\oplus(E_{1})_{(S_{T_{1}},u_{1},l_{1})}\subset E_{(S_{T},u^{\prime},l)}.

Consider maps u:STMu:S_{T}\to M in a C0C^{0}-neighborhood of uu^{\prime} of class W1,pW^{1,p}, which satisfy the following condition

(4.25) ¯u0mod(E0)(ST0,u0,l0)(E1)(ST1,u1,l1),\bar{\partial}u\equiv 0\mod(E_{0})_{(S_{T_{0}},u_{0},l_{0})}\oplus(E_{1})_{(S_{T_{1}},u_{1},l_{1})},

i.e. it satisfies gradient flow equation on interval components, JJ-holomorphic curve equation on disk components of color 0, and Floer’s equation on disk components of color 11, modulo the obstruction subspace (4.24).

Then the strategy for solving this equation (4.25) is standard in the theory of pseudoholomorphic curves (perhaps in more general setup of Fredholm problems). First, we may find an approximate solution u~\tilde{u} by using a cut-off function to interpolate any given two tangent vectors ViT(STi,ui,li)i,i=0,1V_{i}\in T_{(S_{T_{i}},u_{i},l_{i})}\mathcal{B}_{i},i=0,1, and use exponential map to get an element in (ST,u,l)\mathcal{B}_{(S_{T},u^{\prime},l)}. Then a standard implicit theorem argument (finding a right inverse to the linearized operator), together with exponential decay estimates as explained in detail in [FOOO09b], [FOOO12] provides an exact solution uu to the equation (4.25). This part is a combination of the standard proofs of gluing theorems in various setups: for gradient flow equations, for JJ-holomorphic curve equations, and for Floer’s equations. For complete details see section 7.1 of [FOOO09b], though the conditions are slightly different because we have fixed a set of generators as asymptotic convergence conditions for the stable pearly tree maps in consideration.

Now we define

(4.26) U((ST0,u0,l0),(ST1,u1,l1))U((S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1}))

to be the space of (ST,u,l)(S_{T},u,l) where uu is an exact solution of (4.25) obtained from perturbing the approximate solution u~\tilde{u}, which is close to the original pre-gluing uu^{\prime}. The Kuranishi map s[(ST0,u0,l0),(ST1,u1,l1)]s_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]} is simply the restriction of ¯\bar{\partial} to U((ST0,u0,l0),(ST1,u1,l1))U((S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})). Again, the isotropy group is taken to be the trivial group, Γ[(ST0,u0,l0),(ST1,u1,l1)]={1}\Gamma_{[(S_{T_{0}},u_{0},l_{0}),(S_{T_{1}},u_{1},l_{1})]}=\{1\}.

This completes the construction of a Kuranishi chart at a point in the codimension-one strata of ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}). For elements in strata of higher codimension, the construction can be done in the same way.

Now we shall gather all these Kuranishi charts and globalize the construction to a Kuranishi structure on the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}). Briefly denote by σ\sigma a point in the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), i.e. a stable broken pearly tree map. Since the moduli space is compact, we may choose finitely many elements, σ1,,σK\sigma_{1},\cdots,\sigma_{K}, so that their Kuranishi neighborhoods UσU_{\sigma} cover the moduli space

(4.27) iInt(Uσ)¯k+1(α,β;J,H;c0,,ck),\cup_{i}Int(U_{\sigma})\supset\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}),

where we take the Kuranishi neighborhoods UσU_{\sigma} to be closed neighborhoods of σi\sigma_{i} inside the Banach manifold \mathcal{B}. Now given any point σ¯k+1(α,β;J,H;c0,,ck)\sigma\in\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), we modify the previously chosen Kuranishi neighborhood U(σ)U(\sigma) and obstruction space EσE_{\sigma} so that the new ones fit well into a global structure.

First consider the case where the underlying broken pearly tree Σ=(ST0,,STm)\Sigma=(S_{T_{0}},\cdots,S_{T_{m}}) of σ\sigma is stable. We define a new obstruction space

(4.28) Eσ=i:σUσiEσi.E^{\prime}_{\sigma}=\oplus_{i:\sigma\in U_{\sigma_{i}}}E_{\sigma_{i}}.

Also, we modify the previously chosen U(σ)U(\sigma) to a new Kuranishi neighborhood UσU^{\prime}_{\sigma} to be the set of isomorphism classes of (ST,u,l)(S^{\prime}_{T^{\prime}},u^{\prime},l^{\prime}), where STS^{\prime}_{T^{\prime}} is close to STS_{T} in the moduli space of stable pearly trees (which is a smooth manifold with corners), and uu^{\prime} satisfies

(4.29) ¯u0modEσ,\bar{\partial}u^{\prime}\equiv 0\mod E^{\prime}_{\sigma},

while ll^{\prime} is still simply a lifting condition compatible with the restriction of uu^{\prime} to ST\partial S^{\prime}_{T^{\prime}}.

We shall choose these UσU^{\prime}_{\sigma} carefully such that the following condition is satisfied:

(4.30) σUσEσEσ.\sigma^{\prime}\in U^{\prime}_{\sigma}\implies E^{\prime}_{\sigma^{\prime}}\xhookrightarrow{}E^{\prime}_{\sigma^{\prime}}.

Here the inclusion map is defined using parallel transport with respect to the metric on the moduli space of underlying pearly trees, as well as the metric on MM determined by ω\omega and JJ. For such choices, we may assume that whenever σUσ\sigma^{\prime}\in U^{\prime}_{\sigma}, there is a neighborhood Uσ,σU^{\prime}_{\sigma^{\prime},\sigma} of σ\sigma^{\prime} in UσU^{\prime}_{\sigma^{\prime}}, together with a natural inclusion map Uσ,σUσU^{\prime}_{\sigma^{\prime},\sigma}\xhookrightarrow{}U^{\prime}_{\sigma}. Such

(4.31) (Uσ,Eσ,sσ,Γσ={1})(U^{\prime}_{\sigma},E^{\prime}_{\sigma},s_{\sigma},\Gamma_{\sigma}=\{1\})

will be the desired Kuranishi chart, so that there are natural coordinate changes whenever σUσ\sigma^{\prime}\in U^{\prime}_{\sigma} described as above. The compatibility of these coordinate changes with the Kuranishi maps sσ,sσs_{\sigma},s_{\sigma^{\prime}} holds automatically, because the Kuranishi map is simply the restriction of the section ¯\bar{\partial} on UσU^{\prime}_{\sigma}, as a subspace of a Banach manifold or (finite) direct products of Banach manifolds.

Now consider the case where the underlying broken pearly tree Σ=(ST0,,STm)\Sigma=(S_{T_{0}},\cdots,S_{T_{m}}) of σ\sigma is unstable. In this case, ”nearby” pearly trees are not unique - they occur in positive dimensional families parametrized by

(4.32) iLie(Aut(Σi))0,\prod_{i}Lie(Aut(\Sigma_{i}))_{0},

where Lie(Aut(Σi))0Lie(Aut(\Sigma_{i}))_{0} is a neighborhood of 0 in the Lie algebra of the group of automorphisms of the underlying broken pearly trees Σi\Sigma_{i} of σi\sigma_{i}, where the product is taken over all unstable underlying pearly trees Σi\Sigma_{i} of σi\sigma_{i}, among the chosen σ1,,σK\sigma_{1},\cdots,\sigma_{K}. However, the equation (4.29) is not invariant under the pullback by the automorphisms given by (exponentiating) elements in (4.32).

There are several ways of resolving this issue, among which we choose the method of ”canonical gauge fixing”, introduced in [FO99]. That method has the advantage that we only have to deal with the fixed σi\sigma_{i}’s, instead of performing some construction for each element σ\sigma, e.g. stabilizing the domain Σ\Sigma of σ\sigma.

The argument is as follows. For i=1,,Ki=1,\cdots,K, let Σi,un\Sigma_{i,un} be the union of unstable components of Σi\Sigma_{i}. Note that Σi,un\Sigma_{i,un} is not a broken pearly tree and might not be connected. Let Σi,un0\Sigma_{i,un}^{0} be Σi,un\Sigma_{i,un} with intervals [ρ/2,+),(,ρ/2][\rho/2,+\infty),(-\infty,-\rho/2] or half-strips [ρ/2,+)×[0,1],(,ρ/2]×[0,1][\rho/2,+\infty)\times[0,1],(-\infty,-\rho/2]\times[0,1] cut off from some exterior edges, so that we can identify Σi,un0\Sigma_{i,un}^{0} as a subset of Σ\Sigma whenever σ\sigma is close to σi\sigma_{i}, so that Σ\Sigma is obtained from partial gluing of the broken pearly tree Σi\Sigma_{i}, with suitable deformation of the almost complex structure.

Suppose σ\sigma is close to σi\sigma_{i} so that there is some gluing and deformation parameter ξ=(ρ,λ)\xi=(\vec{\rho},\lambda) and an isomorphism ϕ:ΣΣi,ξ\phi:\Sigma\to\Sigma_{i,\xi}, where Σi,ξ\Sigma_{i,\xi} is obtained from Σ\Sigma by partial gluing with respect to ρ\vec{\rho} and deformation of complex structure by λ\lambda. Let d2:M×M[0,)d^{2}:M\times M\to[0,\infty) be the Riemannian distance square function. Define a function on the space of such (ξ,ϕ)(\xi,\phi) by

(4.33) Fi(ξ,ϕ)=xΣi,un0d2(uξ(x),uϕ1(x))𝑑x,F_{i}(\xi,\phi)=\int_{x\in\Sigma_{i,un}^{0}}d^{2}(u_{\xi}(x),u\circ\phi^{-1}(x))dx,

where uξ:Σi,ξMu_{\xi}:\Sigma_{i,\xi}\to M is a map on Σi,ξ\Sigma_{i,\xi} obtained from deforming uu to an exact solution of (4.29). Using this FiF_{i}, we define a function on the neighborhood in the Lie algebra, Lie(Aut(Σi))0Lie(Aut(\Sigma_{i}))_{0}, by

(4.34) Fi(vξ,ϕ),vLie(Aut(Σi))0.F_{i}(v\xi,\phi),v\in Lie(Aut(\Sigma_{i}))_{0}.

Using the convexity of the Riemannian distance square, we can prove that FiF_{i} is convex with respect to vLie(Aut(Σi))0v\in Lie(Aut(\Sigma_{i}))_{0}, i.e. the Hessian matrix with respect to vv is positive definite. Moreover, there is a uniform constant c>0c>0 such that the Hessian matrix is bigger than cIcI uniformly, where II is the identity matrix. Now consider vLie(Aut(Σi))0v\in\partial Lie(Aut(\Sigma_{i}))_{0} on the boundary of this neighborhood. In that case it is not difficult to find a biholomorphic map ϕ\phi such that

(4.35) Fi(vξ,ϕ)Fi(ξ,ϕ)>c,vLie(Aut(Σi))0.F_{i}(v\xi,\phi)-F_{i}(\xi,\phi)>c,\forall v\in\partial Lie(Aut(\Sigma_{i}))_{0}.

Thus we may choose UσiU^{\prime}_{\sigma_{i}} small enough such that there is a unique vLie(Aut(Σi))0v\in Lie(Aut(\Sigma_{i}))_{0} such that Fi(vξ,ϕ)F_{i}(v\xi,\phi) has a unique local minimum with respect to ϕ\phi. Thus the gauge fixing condition can be chosen as follows:

(4.36) Fi(vξ,ϕ)Fi(ξ,ϕ),vLie(Aut(Σi))0.F_{i}(v\xi,\phi)\geq F_{i}(\xi,\phi),\forall v\in Lie(Aut(\Sigma_{i}))_{0}.

Requiring this condition uniquely determines (ξ,ϕ)(\xi,\phi).

After imposing this condition, for any such σ\sigma, the equation (4.29) has no ambiguity for elements close to σ\sigma, so that we can use it to define the Kuranishi neighborhood UσU^{\prime}_{\sigma}. This completes the construction of the desired Kuranishi chart (Uσ,Eσ,sσ,Γσ={1})(U^{\prime}_{\sigma},E^{\prime}_{\sigma},s_{\sigma},\Gamma_{\sigma}=\{1\}), which satisfies the desired properties as argued in the case where Σ\Sigma is stable.

The statement that the induced Kuranishi structure on Sm¯k+1(α,β;J,H;c0,,ck)S_{m}\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) agrees with the fiber product Kuranishi structures on the fiber products (4.15) follows immediately from the way in which the Kuranishi charts are constructed.

4.8. Perturbations by multisections

An important geometric assumption for us is that the ambient symplectic manifold MM is exact, which allows us to take a single-valued multisection to the obstruction bundle (system) of the Kuranishi space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) as in this case the isotropy group of every Kuranishi chart is trivial. Also see [FOOO13], in which notions of stacks and piecewise smooth global sections are introduced in order to deal with the more general case of spherically positive manifolds.

Going back to the construction of Kuranishi structures in the previous subsection, recall that we demand that the isotropy group Γ\Gamma for any Kuranishi chart on the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) be trivial.

The way of achieving transversality in the theory of Kuranishi structures is to perturb the Kuranishi map by multisections locally transverse to zero, then glue (the zero loci of) these multisections together according to the compatibility conditions with respect to the coordinate changes. In this way, we can obtain a virtual fundamental chain on the moduli space, as some kind of Cěch-type chain. Precisely it should be a dual of Cěch cochain, such discussions will not be made or used in this paper.

The following proposition can be essentially derived from [FOOO09b]. Moreover, we remark that both the statement and the proof here are much simpler and indeed simplified, because in [FOOO09b] they consider singular chains on the Lagrangian submanifolds, in whose intersection theory there are serious issues with transversality at the diagonal, forcing one to introduce a hierarchy of perturbations and choices of appropriate geometric chains to make the whole chain-level argument work. However, in our case we have chosen a different chain model involving a discrete set of generators satisfying certain non-degeneracy conditions (at the price of making the construction less canonical), and we have constructed Kuranishi structures depending on these generators as inputs and outputs, therefore there is not much trouble in arranging the perturbations by multisections to achieve transversality.

Proposition 4.25.

For all k,α,βk,\alpha,\beta and for each c0,,ckc_{0},\cdots,c_{k}, such that the moduli spaces ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) are non-empty as topological spaces (not after perturbation), there exist (systems of) single-valued multisections sα,β;J,H;c0,,cks_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}} for the given Kuranishi structure on ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}), such that they are transverse to zero.

In addition, these multisections are compatible with the fiber product multisections on the fiber product (4.15). This means that there are isomorphisms of multisections

(4.37) sα,β;J,H;c0,,cksα0,β0;J,H;c0,0,,c0,k0××sαm,βm;J,H;cm,0,,cm,km,s_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}}\cong s_{\alpha_{0},\beta_{0};J,H;c_{0,0},\cdots,c_{0,k_{0}}}\times\cdots\times s_{\alpha_{m},\beta_{m};J,H;c_{m,0},\cdots,c_{m,k_{m}}},

when restricting the multisection sα,β;J,H;c0,,cks_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}} to that codimension-mm stratum (4.15).

Note that the choices of these multisections depend on the inputs and outputs c0,,ckc_{0},\cdots,c_{k}. Therefore, it is not necessary to deal with transversality issues for various evaluation maps as in the setup using singular chains as in [FOOO09b], so the procedure described here is much simpler. In fact, the construction only uses abstract theory of Kuranishi structures, once we have constructed the Kuranishi structures on the moduli spaces ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) so that the Kuranishi structures are compatible at the boundary with the fiber product Kuranishi structures on (4.15).

4.9. The curved AA_{\infty}-algebra associated to an exact cylindrical Lagrangian immersion

Based on the discussion presented above, we may extract from the moduli spaces ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) a structure of a curve AA_{\infty}-algebra on the underlying wrapped Floer cochain space CW(L,ι;H)CW^{*}(L,\iota;H).

Proposition 4.26.

A coherent choice of single-valued multisections on ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) defines a structure of a curved AA_{\infty}-algebra on CW(L,ι;H)CW^{*}(L,\iota;H). Moreover, this curved AA_{\infty}-algebra is independent of the choice up to homotopy.

Sketch of proof.

We sketch the main steps of the proof while referring the reader to techniques developed in [FOOO09a], [FOOO09b]. The discussion in the previous subsections provides coherent system of Kuranishi structures on all the moduli spaces ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}). Also, by Proposition 4.25, there exist single-valued multisections, which are compatible at the boundary strata with fiber product multisections. Let us fix such a coherent choice of single-valued multisections.

To define the structure maps of the curved AA_{\infty}-algebra, we shall consider only the rigid cases, which means that we only consider generators cic_{i} of appropriate degrees such that the virtual dimension of the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is zero. The zero-sets of the chosen single-valued multisections define an integral virtual fundamental chain, which therefore gives us an integer number

aα,β;J,H;c0,,ck.a_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}}\in\mathbb{Z}.

Then we set

(4.38) mk(ck,,c1)=α,β,c0deg(c0)deg(c1)deg(ck)=2kaα,β;J,H;c0,,ckc0.m^{k}(c_{k},\cdots,c_{1})=\sum_{\begin{subarray}{c}\alpha,\beta,c_{0}\\ \deg(c_{0})-\deg(c_{1})-\cdots-\deg(c_{k})=2-k\end{subarray}}a_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}}c_{0}.

To make sense of the formula (4.38), we must show that this sum is in fact finite. If k=0k=0, the zeroth order map

m0(1)=α,beta,c0deg(c)=2aα,β;J,H;c0c0m^{0}(1)=\sum_{\begin{subarray}{c}\alpha,beta,c_{0}\\ \deg(c)=2\end{subarray}}a_{\alpha,\beta;J,H;c_{0}}c_{0}

”counts” inhomogeneous pseudoholomorphic disks with one marked point, which necessarily pass through some self-intersection point. That is, c0c_{0} is of the form (p,w)(p,w). Then, Lemma 4.12 implies that there can only be finitely many such homology classes β\beta appearing so that the moduli space ¯1(α,β;J,H;c0)\bar{\mathcal{M}}_{1}(\alpha,\beta;J,H;c_{0}) is non-empty. For k1k\geq 1, we can again use the action-energy identity, which holds because the Lagrangian immersions ι:LM\iota:L\to M is exact, to show that if the inputs c1,,ckc_{1},\cdots,c_{k} are fixed, there are only finitely many possible c0c_{0} and finitely many possible homology classes β\beta, for which the moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is non-empty. This implies that the sum (4.38) is a finite sum. Thus, it gives rise to a well-defined multilinear map

mk:CW(L,ι;H)kCW(L,ι;H).m^{k}:CW^{*}(L,\iota;H)^{\otimes k}\to CW^{*}(L,\iota;H).

To verify that these maps satisfy the AA_{\infty}-equations, we need to study boundary strata of moduli spaces ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) of virtual dimension one. Recall that the multisection sα,β;J,H;c0,,cks_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}} is isomorphic to the fiber product multisections (4.37) on (4.15). In particular, if the virtual dimension of ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is one, the only boundary strata that we have to consider are of codimension one, and have the form

(4.39) ¯k0+1(α0,β0;J,H;c0,0,,c0,k0)ׯk1+1(α1,β1;J,H;c1,0,,c1,k1).\bar{\mathcal{M}}_{k_{0}+1}(\alpha_{0},\beta_{0};J,H;c_{0,0},\cdots,c_{0,k_{0}})\times\bar{\mathcal{M}}_{k_{1}+1}(\alpha_{1},\beta_{1};J,H;c_{1,0},\cdots,c_{1,k_{1}}).

The numbering of these generators is as follows. The root e1,0e_{1,0} of a stable pearly tree map in ¯k1+1(α1,β1;J,H;c1,0,,c1,k1)\bar{\mathcal{M}}_{k_{1}+1}(\alpha_{1},\beta_{1};J,H;c_{1,0},\cdots,c_{1,k_{1}}) is connected to some leaf e0,l(1)e_{0,l(1)} of a stable pearly tree map in ¯k0+1(α0,β0;J,H;c0,0,,c0,k0)\bar{\mathcal{M}}_{k_{0}+1}(\alpha_{0},\beta_{0};J,H;c_{0,0},\cdots,c_{0,k_{0}}). Thus c1,0=c0,l(1)=cnewc_{1,0}=c_{0,l(1)}=c_{new}, while the other generators agree with the original ones:

(4.40) c0,i=ci,i=0,,l(1),\displaystyle c_{0,i}=c_{i},i=0,\cdots,l(1),
(4.41) c1,j=cj+l(1)1,j=1,,k1,\displaystyle c_{1,j}=c_{j+l(1)-1},j=1,\cdots,k_{1},
(4.42) c0,i=ci+k11,i=l(1)+1,,k0.\displaystyle c_{0,i}=c_{i+k_{1}-1},i=l(1)+1,\cdots,k_{0}.

So we rewrite the fiber product as

(4.43) ¯k0+1(α0,β0;J,H;c0,,cl(1)1,cnew,cl(1)+k1,,ck0+k11)ׯk1+1(α1,β1;J,H;cnew,cl(1),,cl(1)+k11).\begin{split}&\bar{\mathcal{M}}_{k_{0}+1}(\alpha_{0},\beta_{0};J,H;c_{0},\cdots,c_{l(1)-1},c_{new},c_{l(1)+k_{1}},\cdots,c_{k_{0}+k_{1}-1})\\ &\times\bar{\mathcal{M}}_{k_{1}+1}(\alpha_{1},\beta_{1};J,H;c_{new},c_{l(1)},\cdots,c_{l(1)+k_{1}-1}).\end{split}

Thus we have isomorphisms of multisections:

(4.44) sα,β;J,H;c0,,cksα0,β0;J,H;c0,,cl(1)1,cnew,cl(1)+k1,,ck0+k11×sα1,β1;J,H;cnew,cl(1),,cl(1)+k11,\begin{split}&s_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}}\\ &\cong s_{\alpha_{0},\beta_{0};J,H;c_{0},\cdots,c_{l(1)-1},c_{new},c_{l(1)+k_{1}},\cdots,c_{k_{0}+k_{1}-1}}\times s_{\alpha_{1},\beta_{1};J,H;c_{new},c_{l(1)},\cdots,c_{l(1)+k_{1}-1}},\end{split}

when restricting the multisection sα,β;J,H;c0,,cks_{\alpha,\beta;J,H;c_{0},\cdots,c_{k}} to the boundary strata. This implies that the operations mkm^{k} defined in (4.38) satisfy the AA_{\infty}-equations.

The independence of choices of multisections up to homotopy is a consequence of general theory of Kuranishi structures. To prove that the curved AA_{\infty}-algebra is independent of choice of almost complex structure JJ up to homotopy, we may introduce parametrized moduli spaces of pearly tree maps with time-allocation, associated to such a homotopy {Jt}t[0,1]\{J_{t}\}_{t\in[0,1]}, and construct Kuranishi structures on these moduli spaces in a well-arranged way so that they define cobordisms between the Kuranishi structures on the moduli spaces of pearly tree maps ¯k+1(α,β;Ji,H;c0,,ck),i=0,1\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J_{i},H;c_{0},\cdots,c_{k}),i=0,1 with respect to different almost complex structures. Here recall that the generators c0,,ckc_{0},\cdots,c_{k} themselves do not depend on the almost complex structures used to define pearly tree maps, so both kinds of moduli spaces make sense.

4.10. Bounding cochains

Given a cylindrical Lagrangian immersion ι:LM\iota:L\to M, we have constructed a curved AA_{\infty}-algebra structure (CW(L,ι;H),mk)(CW^{*}(L,\iota;H),m^{k}) on the wrapped Floer cochain space. In general m1m^{1} does not square to zero because of the curvature term m00m^{0}\neq 0. In order to get a differential, we want to deform the operations mkm^{k} algebraically such that the deformed differential squares to zero, and moreover, the deformed structure maps of all orders satisfy the AA_{\infty}-equations without curvature. This idea was introduced into Lagrangian Floer theory by Fukaya-Oh-Ohta-Ono [FOOO09a], by finding solutions to the inhomogeneous Maurer-Cartan equation:

(4.45) k=0mk(b,,b)=0.\sum_{k=0}^{\infty}m^{k}(b,\cdots,b)=0.

In order for this equation to make sense, we shall further impose the condition that bb is nilpotent, so that the sum stops at a final stage: there exists KK such that mk(b,,b)=0m^{k}(b,\cdots,b)=0 for all k>Kk>K and

k=0Kmk(b,,b)=0.\sum_{k=0}^{K}m^{k}(b,\cdots,b)=0.

Because of the presence of the inhomogeneous term m0(1)m^{0}(1), this equation might not have any solution in general. But if it does have one solution bCW(L,ι;H)b\in CW^{*}(L,\iota;H) which is nilpotent, we call bb a bounding cochain for (L,ι)(L,\iota), and say (L,ι)(L,\iota) is unobstructed in the sense of wrapped Floer theory. In this case, we can deform the curved AA_{\infty}-algebra to an AA_{\infty}-algebra with vanishing curvature:

(4.46) mk;b(ck,,c1)=i0i0++ik=imk+i(b,,bik times,ck,b,,bik1 times,,c1,b,,bi0 times).m^{k;b}(c_{k},\cdots,c_{1})=\sum_{\begin{subarray}{c}i\geq 0\\ i_{0}+\cdots+i_{k}=i\end{subarray}}m^{k+i}(\underbrace{b,\cdots,b}_{\text{$i_{k}$ times}},c_{k},\underbrace{b,\cdots,b}_{\text{$i_{k-1}$ times}},\cdots,c_{1},\underbrace{b,\cdots,b}_{\text{$i_{0}$ times}}).

In particular, m1;bm^{1;b} squares to zero, and we can define a cohomology group of the graded complex CW(L,ι;H)CW^{*}(L,\iota;H) with respect to the differential m1;bm^{1;b}, which we denote by HW(L,ι,b;H)HW^{*}(L,\iota,b;H). We call this the wrapped Floer cohomology group of the Lagrangian immersion ι:LM\iota:L\to M with respect to the bounding cochain bb.

4.11. Wrapped Floer cochain space of a pair of Lagrangian immersions

Given a pair of exact cylindrical Lagrangian immersions with transverse self-intersections, we want to construct a curved AA_{\infty}-bimodule CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H) over the curved AA_{\infty}-algebras CW(Li,ιi;H)CW^{*}(L_{i},\iota_{i};H) associated to the two Lagrangian immersions ιi:LiM,i=0,1\iota_{i}:L_{i}\to M,i=0,1. Assume that these two Lagrangian immersions are not the same (the case where they are the same can be treated using the construction for a single exact cylindrical Lagrangian immersion). In a generic situation, we may assume that the constant HH-chords from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}), which are identical to intersection points of ι0(L0)\iota_{0}(L_{0}) with ι1(L1)\iota_{1}(L_{1}), do not coincide with the self-intersection points of ι0\iota_{0} or those of ι1\iota_{1}. Moreover, we may assume that this pair intersects transversely, in the following sense.

Definition 4.27.

Let (L0,ι0),(L1,ι1))(L_{0},\iota_{0}),(L_{1},\iota_{1})) be a pair of exact cylindrical Lagrangian immersions with transverse self-intersections. They are said to intersect transversely, if the following conditions are satisfied:

  1. (i)

    the intersection points of ι0(L0)ι1(L1)\iota_{0}(L_{0})\cap\iota_{1}(L_{1}) are different from the self-intersections of ι0:L0M\iota_{0}:L_{0}\to M and those of ι1:L1M\iota_{1}:L_{1}\to M;

  2. (ii)

    the images ι0(L0)\iota_{0}(L_{0}) and ι1(L1)\iota_{1}(L_{1}) intersect transversely;

  3. (iii)

    they do not intersection over the cylindrical end M×[1,+)\partial M\times[1,+\infty).

Now let us define the wrapped Floer cochain space for such a pair (L0,ι0),(L1,ι1))(L_{0},\iota_{0}),(L_{1},\iota_{1})).

Definition 4.28.

We define the wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H) to be the free \mathbb{Z}-module generated by the following two kinds of generators

  1. (i)

    an intersection point pp of ι0(L0)\iota_{0}(L_{0}) and ι1(L1)\iota_{1}(L_{1}) which is contained in the interior part M0M_{0}, or equivalently a constant HH-chord x=xpx=x_{p} at pp from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}) (recall HH is constantly zero in the interior part);

  2. (ii)

    non-constant time one HH-chords xx from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}) which is contained in the cylindrical end, where both ι0,ι1\iota_{0},\iota_{1} are assumed to be embeddings.

Since every interior intersection point of ι0(L0)\iota_{0}(L_{0}) and ι1(L1)\iota_{1}(L_{1}) is assumed to be different from any self-intersections of ι0\iota_{0} or ι1\iota_{1}, and since both Lagrangian immersions are exact, the above definition makes sense, without having to include capping half-disks for these intersection points. From now on we shall use the letter cc to denote any kind of generator.

4.12. Moduli spaces of Floer trajectories

To study wrapped Floer theory for a pair of exact cylindrical Lagrangian immersions, we shall introduce moduli spaces of stable broken Floer trajectories. In our situation, because the Lagrangian submanifolds are immersed, we also need to add some refinement of the data to the moduli spaces, which are similar to those for a single Lagrangian immersion.

Choose for each Lagrangian immersion ιi:LiM\iota_{i}:L_{i}\to M an admissible almost complex structure Ji,i=0,1J_{i},i=0,1 of contact type, using which the curved AA_{\infty}-algebras (CW(Li,ιi;H),mk)(CW^{*}(L_{i},\iota_{i};H),m^{k}) are defined. Also, choose a path of admissible almost complex structures JtJ_{t} of contact type connecting J0J_{0} and J1J_{1}. We shall consider the moduli spaces of the following kinds of maps.

Let I0{1,,k}I_{0}\subset\{1,\cdots,k\} and I1{1,,l}I_{1}\subset\{1,\cdots,l\}, and αi:IiS(Li,ιi)\alpha_{i}:I_{i}\to S(L_{i},\iota_{i}) be maps, labeling those marked points which are mapped to some self-intersection points of ι0\iota_{0} and ι1\iota_{1} respectively. Also, let βH2(M,ι0(L0)ι1(L1))\beta\in H_{2}(M,\iota_{0}(L_{0})\cup\iota_{1}(L_{1})) be a relative homology class.

Definition 4.29.

A (k,l)(k,l)-marked Floer trajectory is a quadruple (Σ,s,u,l)(\Sigma,\vec{s},u,\vec{l}) satisfying the following conditions:

  1. (i)

    Σ=×[0,1]\Sigma=\mathbb{R}\times[0,1] is the infinite strip.

  2. (ii)

    s=(s0,s1)\vec{s}=(\vec{s}^{0},\vec{s}^{1}) with s0=(s10,,sk0)\vec{s}^{0}=(s^{0}_{1},\cdots,s^{0}_{k}) and s1=(s11,,sl1)\vec{s}^{1}=(s^{1}_{1},\cdots,s^{1}_{l}) are collections of real numbers, such that

    sj0>sj+10,j=1,,k1,s^{0}_{j}>s^{0}_{j+1},j=1,\cdots,k-1,
    sj1<sj+11,j=1,,l1.s^{1}_{j}<s^{1}_{j+1},j=1,\cdots,l-1.
  3. (iii)

    u:ΣMu:\Sigma\to M is a continuous map.

  4. (iv)

    uu satisfies the inhomogeneous Cauchy-Riemann equation (Floer’s equation)

    us+Jt(utXH(u))=0.\frac{\partial u}{\partial s}+J_{t}(\frac{\partial u}{\partial t}-X_{H}(u))=0.
  5. (v)

    u(s,0)ι0(L0),u(s,1)ι1(L1)u(s,0)\in\iota_{0}(L_{0}),u(s,1)\in\iota_{1}(L_{1}), for all ss\in\mathbb{R}.

  6. (vi)

    uu asymptotically converges to time-one HH-chords x,x+x_{-},x_{+} from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}) as ss\to-\infty and respectively ++\infty, where these HH-chords might be constant or non-constant.

  7. (vii)

    The homology class of uu is βH2(M,ι0(L0)ι1(L1))\beta\in H_{2}(M,\iota_{0}(L_{0})\cup\iota_{1}(L_{1})).

  8. (viii)

    l=(l0,l1)\vec{l}=(l_{0},l_{1}) is a pair of smooth maps li:×{i}Li×ιiLil_{i}:\mathbb{R}\times\{i\}\to L_{i}\times_{\iota_{i}}L_{i}, such that u(s,i)=ιili(s)u(s,i)=\iota_{i}\circ l_{i}(s), for s×{i}{sji|jIi}s\in\mathbb{R}\times\{i\}\setminus\{s^{i}_{j}|j\in I_{i}\}. l\vec{l} is the boundary lifting condition of uu.

  9. (ix)

    (limssjili(s),limssjili(s))=αi(j)(\lim\limits_{s\uparrow s^{i}_{j}}l_{i}(s),\lim\limits_{s\downarrow s^{i}_{j}}l_{i}(s))=\alpha_{i}(j), for every jIij\in I_{i}.

  10. (x)

    (Σ,s,u,l)(\Sigma,\vec{s},u,l) is stable.

There is an obvious \mathbb{R}-action by translations on the set of all marked Floer trajectories (Σ,s,u,l)(\Sigma,\vec{s},u,l). We denote by

𝒩k,l((L0,ι0),(L1,ι1);α,β;{Jt}t,H;c,c+)\mathcal{N}_{k,l}((L_{0},\iota_{0}),(L_{1},\iota_{1});\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+})

the set of equivalence classes of such (Σ,s,u,l)(\Sigma,\vec{s},u,l). For convenience, we sometimes omit the notations (L0,ι0),(L1,ι1)(L_{0},\iota_{0}),(L_{1},\iota_{1}) in case no confusion can occur.

The above moduli space has a natural compactification, called the moduli space of stable broken Floer trajectories, to be described in two steps: first, we need to add trees of disk bubbles to each boundary of a Floer trajectory; second, we need to add broken Floer trajectories.

First consider stable unbroken Floer trajectories, which are marked Floer trajectories with stable pearly trees attached to the two boundary components of the strip ×[0,1]\mathbb{R}\times[0,1].

In order to describe these elements in detail, we add more data to marked Floer trajectories. Let (Σ,s,u,l)(\Sigma,\vec{s},u,l) be a marked Floer trajectory.

Definition 4.30.

A decoration for (Σ,s,u,l)(\Sigma,\vec{s},u,l) is an assignment of coloring c:s{0,1}c:\vec{s}\to\{0,1\}.

For each sjis^{i}_{j} with color c(sji)=0c(s^{i}_{j})=0, we attach a half-infinite ray Isji=[0,+)I_{s^{i}_{j}}=[0,+\infty) to Σ\Sigma at sjis^{i}_{j}. For each sjis^{i}_{j} with color c(sji)=1c(s^{i}_{j})=1, we remove sjis^{i}_{j} and add a half-infinite strip Zsji+=[0,+)×[0,1]Z_{s^{i}_{j}}^{+}=[0,+\infty)\times[0,1] as a strip-like end near the puncture sjis^{i}_{j}.

Now we extend the map uu to the decorated domain Σ~\tilde{\Sigma}. On each newly-added half-infinite ray IsjiI_{s^{i}_{j}}, we extend uu by a map usji:IsjiLiu_{s^{i}_{j}}:I_{s^{i}_{j}}\to L_{i}, which comes with a preferred lift usji~:IsjiLi×ιiLi\tilde{u_{s^{i}_{j}}}:I_{s^{i}_{j}}\to L_{i}\times_{\iota_{i}}L_{i}, which satisfies the gradient flow equation:

(4.47) dusjidt+fi,s(usji)=0,\frac{du_{s^{i}_{j}}}{dt}+\nabla f_{i,s}(u_{s^{i}_{j}})=0,

and asymptotically converges to a critical point of fif_{i}, where fi,sf_{i,s} a family of perturbations of fif_{i} parametrized by sIsjis\in I_{s^{i}_{j}}, so that fi,s=fif_{i,s}=f_{i} for s0s\gg 0. On each newly-added half-infinite strip Zsji+Z_{s^{i}_{j}}^{+}, we extend uu by a map usji:Zsji+Mu_{s^{i}_{j}}:Z_{s^{i}_{j}}^{+}\to M, such that usjiu_{s^{i}_{j}} maps the boundary to ιi(Li)\iota_{i}(L_{i}), and satisfies the inhomogeneous Cauchy-Riemann equation:

(4.48) susji+Ji(tusjiXH(usji))=0,\partial_{s}u_{s^{i}_{j}}+J_{i}(\partial_{t}u_{s^{i}_{j}}-X_{H}(u_{s^{i}_{j}}))=0,

and asymptotically converges to some time-one HH-chord xx.

These critical points of fif_{i} or non-constant HH-chords from ιi(Li)\iota_{i}(L_{i}) to itself should come equipped with choices of capping half-disks, which make them into generators cijc_{i}^{j} of the wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H).

We also extend the maps l=(l0,l1)\vec{l}=(l_{0},l_{1}) to these newly-added components, so that the extended map l~i\tilde{l}_{i} satisfies ιil~i=u~\iota_{i}\circ\tilde{l}_{i}=\tilde{u}. In particular, the extension to a half-infinite ray IsjiI_{s^{i}_{j}} is precisely the preferred lift u~sji\tilde{u}_{s^{i}_{j}}.

Definition 4.31.

We call the extended map (Σ~,s,u~,l~)(\tilde{\Sigma},\vec{s},\tilde{u},\vec{\tilde{l}}) a decorated Floer trajectory. Denote by

(4.49) 𝒩k,ldec(α,β;{Jt}t,H;c,c+;c10,,ck0,c11,,cl1)\mathcal{N}_{k,l}^{dec}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};c^{0}_{1},\cdots,c^{0}_{k},c^{1}_{1},\cdots,c^{1}_{l})

the moduli space of decorated Floer trajectories.

For convenience, denote by k=k0k=k_{0} and l=k1l=k_{1}. Let n0,n1,m0,m1n_{0},n_{1},m_{0},m_{1} and m0,1,,m0,n0m_{0,1},\cdots,m_{0,n_{0}}, and m1,1,,m1,n1m_{1,1},\cdots,m_{1,n_{1}} be non-negative integers such that

ki=mi+a=1nimi,a.k_{i}=m_{i}+\sum_{a=1}^{n_{i}}m_{i,a}.

Let Ai{1,,mi+ni}A_{i}\subset\{1,\cdots,m_{i}+n_{i}\} be a subset of nin_{i} elements and put

(4.50) Ai={σi(1),,σi(ni)},A_{i}=\{\sigma_{i}(1),\cdots,\sigma_{i}(n_{i})\},

where σi:{1,,ni}{1,,mi+ni}\sigma_{i}:\{1,\cdots,n_{i}\}\to\{1,\cdots,m_{i}+n_{i}\} is an injective map satisfying σi(a)<σi(a+1)\sigma_{i}(a)<\sigma_{i}(a+1). Let πi,1:(Crit(fi)𝒳+(ιi(Li),H))ki+ni(Crit(fi)𝒳+(ιi(Li),H))ni\pi_{i,1}:(Crit(f_{i})\coprod\mathcal{X}_{+}(\iota_{i}(L_{i}),H))^{k_{i}+n_{i}}\to(Crit(f_{i})\coprod\mathcal{X}_{+}(\iota_{i}(L_{i}),H))^{n_{i}} be the projection

πi,1(x1,,xki+ni)=(xσi(1),,xσi(ni)).\pi_{i,1}(x_{1},\cdots,x_{k_{i}+n_{i}})=(x_{\sigma_{i}(1)},\cdots,x_{\sigma_{i}(n_{i})}).

Let πi,2:(Crit(fi)𝒳+(ιi(Li),H))ki+ni(Crit(fi)𝒳+(ιi(Li),H))ki\pi_{i,2}:(Crit(f_{i})\coprod\mathcal{X}_{+}(\iota_{i}(L_{i}),H))^{k_{i}+n_{i}}\to(Crit(f_{i})\coprod\mathcal{X}_{+}(\iota_{i}(L_{i}),H))^{k_{i}} be the projection to the other factors.

For simplicity, denote by AA the collection of data n0,n1,m0,m1n_{0},n_{1},m_{0},m_{1}, m0,1,,m0,n0m_{0,1},\cdots,m_{0,n_{0}}, and m1,1,,m1,n1m_{1,1},\cdots,m_{1,n_{1}} as well as A0,A1A_{0},A_{1}. We put

(4.51) 𝒩k0,k1A(α,β;{Jt}t,H;c,c+;{ca0,ca,10,,ca,m0,a0}a=1n0,{cb1,cb,11,,cb,m1,b1}b=1n1)=ββ0,aβ1,b=βα0aα0,a=α0α1bα1,b=α1𝒩n0+m0,n1+m1dec(α,β;{Jt}t,H;x,x+;c10,,cn0+m00,c11,,cn1+m11)×(π0,1,π1,1)ev,ev0(a=1n0¯m0,a+1(α0,a,β0,a;J0,H;ca0,ca,10,,ca,m0,a0)×b=1n1¯m1,b+1(α1,b,β1,b;J1,H;cb1,cb,11,,cb,m1,b1)),\begin{split}&\mathcal{N}^{A}_{k_{0},k_{1}}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\{c^{0}_{a},c^{0}_{a,1},\cdots,c^{0}_{a,m_{0,a}}\}_{a=1}^{n_{0}},\{c^{1}_{b},c^{1}_{b,1},\cdots,c^{1}_{b,m_{1,b}}\}_{b=1}^{n_{1}})\\ &=\bigcup_{\beta^{\prime}\sharp\sum\beta_{0,a}\sharp\sum\beta_{1,b}=\beta}\coprod_{\begin{subarray}{c}\alpha^{\prime}_{0}\cup\bigcup_{a}\alpha_{0,a}=\alpha_{0}\\ \alpha^{\prime}_{1}\cup\bigcup_{b}\alpha_{1,b}=\alpha_{1}\end{subarray}}\mathcal{N}_{n_{0}+m_{0},n_{1}+m_{1}}^{dec}(\vec{\alpha}^{\prime},\beta^{\prime};\{J_{t}\}_{t},H;x_{-},x_{+};\\ &c^{0}_{1},\cdots,c^{0}_{n_{0}+m_{0}},c^{1}_{1},\cdots,c^{1}_{n_{1}+m_{1}})\\ &\times_{(\pi_{0,1},\pi_{1,1})\circ ev,\vec{ev}_{0}}(\prod_{a=1}^{n_{0}}\bar{\mathcal{M}}_{m_{0,a}+1}(\alpha_{0,a},\beta_{0,a};J_{0},H;c^{0}_{a},c^{0}_{a,1},\cdots,c^{0}_{a,m_{0,a}})\\ &\times\prod_{b=1}^{n_{1}}\bar{\mathcal{M}}_{m_{1,b}+1}(\alpha_{1,b},\beta_{1,b};J_{1},H;c^{1}_{b},c^{1}_{b,1},\cdots,c^{1}_{b,m_{1,b}})),\end{split}

where ev0=(ev0,,ev0)\vec{ev}^{0}=(ev^{0},\cdots,ev^{0}) are the evaluation maps at the 0-th marked point, and the moduli spaces

¯m0,a+1(α0,a,β0,a;J0,H;ca0,ca,10,,ca,m0,a0)\bar{\mathcal{M}}_{m_{0,a}+1}(\alpha_{0,a},\beta_{0,a};J_{0},H;c^{0}_{a},c^{0}_{a,1},\cdots,c^{0}_{a,m_{0,a}})

and

¯m1,b+1(α1,b,β1,b;J1,H;cb1,cb,11,,cb,m1,b1)\bar{\mathcal{M}}_{m_{1,b}+1}(\alpha_{1,b},\beta_{1,b};J_{1},H;c^{1}_{b},c^{1}_{b,1},\cdots,c^{1}_{b,m_{1,b}})

are moduli spaces of stable pearly tree maps.

To simplify the notations, we denote by cA0\vec{c}^{0}_{A} the collection {ca0,ca,10,,ca,m0,a0}a=1n0\{c^{0}_{a},c^{0}_{a,1},\cdots,c^{0}_{a,m_{0,a}}\}_{a=1}^{n_{0}} of generators and similarly cA1\vec{c}^{1}_{A} for the other collection. Also, denote by c0,c1\vec{c}^{0},\vec{c}^{1} for any possible collections among cA0,cA1c^{0}_{A},c^{1}_{A} for all type AA.

Definition 4.32.

We put

(4.52) 𝒩k0,k1unbr(α0,α1;β;c,c+;c0,c1)=A𝒩k0,k1A(α0,α1;β;c,c+;cA0,cA1),\mathcal{N}^{unbr}_{k_{0},k_{1}}(\alpha_{0},\alpha_{1};\beta;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})=\bigcup_{A}\mathcal{N}^{A}_{k_{0},k_{1}}(\alpha_{0},\alpha_{1};\beta;c_{-},c_{+};\vec{c}^{0}_{A},\vec{c}^{1}_{A}),

and call it the moduli space of stable unbroken Floer trajectories.

In words, the elements in this moduli space are Floer trajectories with trees of disks attached to each boundary of ×[0,1]\mathbb{R}\times[0,1]. We do not have to include sphere bubbles because the ambient symplectic manifold MM is exact.

4.13. Compactification: stable broken Floer trajectories

Next, as the limit of a sequence of stable unbroken Floer trajectories can break into several Floer trajectories, we introduce the moduli space of stable broken Floer trajectories.

Definition 4.33.

A stable broken Floer trajectory is a tuple

((Σ~(1),s(1),u~(1),l~(1)),,(Σ~(K),s(K),u~(K),l~(K)),((\tilde{\Sigma}(1),\vec{s}(1),\tilde{u}(1),\vec{\tilde{l}}(1)),\cdots,(\tilde{\Sigma}(K),\vec{s}(K),\tilde{u}(K),\vec{\tilde{l}}(K)),

where each (Σ~(a),s(a),u~(a),l~(a))(\tilde{\Sigma}(a),\vec{s}(a),\tilde{u}(a),\vec{\tilde{l}}(a)) is a stable unbroken Floer trajectory, such that the asymptotic convergence conditions match for successive stable unbroken Floer trajectories:

(4.53) c+(a)=c(a+1).c_{+}(a)=c_{-}(a+1).

The moduli space of stable broken Floer trajectories is denoted by

(4.54) 𝒩¯k0,k1(α,β;{Jt}t,H;c,c+;c0,c1)=Kα0(1),,α0(K)α0(1)α0(K)=α0α1(1),,α1(K)α1(1)α1(K)=α0β(1)β(K)=βk0,1,,k0,Kk0,1++k0,K=k0k1,1,,k1,Kk1,1++k1,K=k1a=1K1𝒩k0,a,k1,aunbr(α0(a),α1(a),β(a);{Jt}t,H;c(a),c+(a);c0(a),c1(a)),\begin{split}&\bar{\mathcal{N}}_{k_{0},k_{1}}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})\\ &=\coprod_{K}\coprod_{\begin{subarray}{c}\alpha_{0}(1),\cdots,\alpha_{0}(K)\\ \alpha_{0}(1)\cup\cdots\cup\alpha_{0}(K)=\alpha_{0}\end{subarray}}\coprod_{\begin{subarray}{c}\alpha_{1}(1),\cdots,\alpha_{1}(K)\\ \alpha_{1}(1)\cup\cdots\cup\alpha_{1}(K)=\alpha_{0}\end{subarray}}\coprod_{\beta(1)\sharp\cdots\sharp\beta(K)=\beta}\\ &\coprod_{\begin{subarray}{c}k_{0,1},\cdots,k_{0,K}\\ k_{0,1}+\cdots+k_{0,K}=k_{0}\end{subarray}}\coprod_{\begin{subarray}{c}k_{1,1},\cdots,k_{1,K}\\ k_{1,1}+\cdots+k_{1,K}=k_{1}\end{subarray}}\prod_{a=1}^{K-1}\\ &\mathcal{N}^{unbr}_{k_{0,a},k_{1,a}}(\alpha_{0}(a),\alpha_{1}(a),\beta(a);\{J_{t}\}_{t},H;c_{-}(a),c_{+}(a);\vec{c}^{0}(a),\vec{c}^{1}(a)),\end{split}

where c(1)=c,c(K)=c+c(1)=c_{-},c(K)=c_{+}. Here the disjoint union is taken over all K,α,c,k0,a,k1,aK,\vec{\alpha},\vec{c},k_{0,a},k_{1,a}.

Now it is standard to prove:

Lemma 4.34.

The moduli space of stable broken Floer trajectories

𝒩¯k0,k1(α,β;{Jt}t,H;c,c+;c0,c1)\bar{\mathcal{N}}_{k_{0},k_{1}}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})

is compact.

4.14. Kuranishi structures and single-valued multisections

The following propositions are essential for extracting algebraic structures from the moduli spaces of stable broken Floer trajectories. The proofs are similar to the case of moduli spaces of stable pearly tree maps in the previous section 4.7.

Proposition 4.35.

The moduli space of stable broken Floer trajectories

𝒩¯k0,k1(α0,α1,β;{Jt}t,H;c,c+;c0,c1)\bar{\mathcal{N}}_{k_{0},k_{1}}(\alpha_{0},\alpha_{1},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})

has an oriented Kuranishi structure which is compatible with the fiber product Kuranishi structures at its codimension one boundary strata:

(4.55) 𝒩¯k0,k1(α,β;{Jt}t,H;c,c+;c0,c1)k0+k0′′=k01ik0α0(c0,c′′′0,c′′0)=c0c0 is an i-tuple𝒩¯k0+1,k1(α0,α1,β;{Jt}t,H;c,c+;(c0,c′′0),c1)×ev0i,ev0¯k0′′+1(α0′′,β0;J0,H;c′′′0)k1+k1′′=k11ik1α1(c1,c′′′1,c′′1)=c1c1 is an i-tuple𝒩¯k0,k1+1(α0,α1,β;{Jt}t,H;c,c+;c0,c1,c′′1)×ev1i,ev0¯k1′′+1(α1′′,β1;J1,H;c′′′1)c+=c𝒩¯k0,k1(α0,α1,β;{Jt}t,H;c,c+;c0,c1)×𝒩¯k0′′,k1′′(α0′′,α1′′,β′′;{Jt}t,H;c,c+;c′′0,c′′1)\begin{split}&\partial\bar{\mathcal{N}}_{k_{0},k_{1}}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})\\ &\cong\coprod_{\begin{subarray}{c}k^{\prime}_{0}+k^{\prime\prime}_{0}=k_{0}\\ 1\leq i\leq k^{\prime}_{0}\\ \alpha_{0}\end{subarray}}\coprod_{\begin{subarray}{c}(\vec{c}^{\prime 0},\vec{c}^{\prime\prime\prime 0},\vec{c}^{\prime\prime 0})=\vec{c}^{0}\\ \text{$\vec{c}^{\prime 0}$ is an $i$-tuple}\end{subarray}}\bar{\mathcal{N}}_{k^{\prime}_{0}+1,k_{1}}(\alpha^{\prime}_{0},\alpha_{1},\beta^{\prime};\{J_{t}\}_{t},H;c_{-},c_{+};(\vec{c}^{\prime 0},\vec{c}^{\prime\prime 0}),\vec{c}^{1})\\ &\times_{ev_{0}^{i},ev^{0}}\bar{\mathcal{M}}_{k^{\prime\prime}_{0}+1}(\alpha^{\prime\prime}_{0},\beta_{0};J_{0},H;\vec{c}^{\prime\prime\prime 0})\\ &\cup\coprod_{\begin{subarray}{c}k^{\prime}_{1}+k^{\prime\prime}_{1}=k_{1}\\ 1\leq i\leq k^{\prime}_{1}\\ \alpha_{1}\end{subarray}}\coprod_{\begin{subarray}{c}(\vec{c}^{\prime 1},\vec{c}^{\prime\prime\prime 1},\vec{c}^{\prime\prime 1})=\vec{c}^{1}\\ \text{$\vec{c}^{\prime 1}$ is an $i$-tuple}\end{subarray}}\bar{\mathcal{N}}_{k_{0},k^{\prime}_{1}+1}(\alpha_{0},\alpha^{\prime}_{1},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{\prime 1},\vec{c}^{\prime\prime 1})\\ &\times_{ev_{1}^{i},ev^{0}}\bar{\mathcal{M}}_{k^{\prime\prime}_{1}+1}(\alpha^{\prime\prime}_{1},\beta_{1};J_{1},H;\vec{c}^{\prime\prime\prime 1})\\ &\cup\coprod_{c^{\prime}_{+}=c^{\prime}_{-}}\bar{\mathcal{N}}_{k^{\prime}_{0},k^{\prime}_{1}}(\alpha^{\prime}_{0},\alpha^{\prime}_{1},\beta^{\prime};\{J_{t}\}_{t},H;c_{-},c^{\prime}_{+};\vec{c}^{\prime 0},\vec{c}^{\prime 1})\\ &\times\bar{\mathcal{N}}_{k^{\prime\prime}_{0},k^{\prime\prime}_{1}}(\alpha^{\prime\prime}_{0},\alpha^{\prime\prime}_{1},\beta^{\prime\prime};\{J_{t}\}_{t},H;c^{\prime}_{-},c_{+};\vec{c}^{\prime\prime 0},\vec{c}^{\prime\prime 1})\end{split}
Proposition 4.36.

On the moduli space of stable broken Floer trajectories

𝒩¯k0,k1(α,β;{Jt}t,H;c,c+;c0,c1),\bar{\mathcal{N}}_{k_{0},k_{1}}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1}),

equipped with a Kuranishi structure as in Proposition 4.35, there exists a single-valued multisection

sk0,k1;α,β;c,c+;{Jt}t,H;c0,c1,s_{k_{0},k_{1};\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1}},

which is transverse to zero, and compatible with the fiber-product single-valued multisections on the Kuranishi spaces (4.55) at the boundary with the fiber-product multisections on (4.55).

4.15. The curved AA_{\infty}-bimodule associated to a pair of exact cylindrical Lagrangian immersions with transverse self-intersections

For a pair of exact cylindrical Lagrangian immersions as above, we shall construct a curved AA_{\infty}-bimodule structure on the wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H). The main result is stated as follows.

Proposition 4.37.

There is a natural curved AA_{\infty}-bimodule structure on the wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H) over the curved AA_{\infty}-algebras for ιi:LiM\iota_{i}:L_{i}\to M,

(CW(L0,ι0;H),CW(L1,ι1;H)).(CW^{*}(L_{0},\iota_{0};H),CW^{*}(L_{1},\iota_{1};H)).

The structure maps nk,ln^{k,l} are defined by appropriate counts of broken stable Floer trajectories.

Suppose both ιi:LiM\iota_{i}:L_{i}\to M are unobstructed with choices of bounding cochains biCW(Li,ιi;H)b_{i}\in CW^{*}(L_{i},\iota_{i};H). Then the (b0,b1)(b_{0},b_{1})-deformation nk,l;b0,b1n^{k,l;b_{0},b_{1}} defines a non-curved AA_{\infty}-bimodule over the deformed AA_{\infty}-algebras

((CW((L0,ι0;H),mk;b0),(CW((L1,ι1;H),mk;b1)).((CW^{*}((L_{0},\iota_{0};H),m^{k;b_{0}}),(CW^{*}((L_{1},\iota_{1};H),m^{k;b_{1}})).
Proof.

Consider moduli spaces of stable broken Floer trajectories which are of virtual dimension zero. Then the virtual fundamental chains associated to the chosen single-valued multisections sk0,k1;α,β;c,c+;{Jt}t,H;c0,c1s_{k_{0},k_{1};\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1}} gives rise to an integer number

(4.56) ak,l;α,β;c,c+;{Jt}t,H;c0,c1=(sk,l;α,β;c,c+;{Jt}t,H;c0,c1)1(0).a_{k,l;\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1}}=(s_{k,l;\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1}})^{-1}(0)\in\mathbb{Z}.

We define multilinear maps

(4.57) nk,l:CW(L0,ι0)kCW((L0,ι0),(L1,ι1))CW(L1,ι1)lCW((L0,ι0),(L1,ι1)),n^{k,l}:CW^{*}(L_{0},\iota_{0})^{\otimes k}\otimes CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1}))\otimes CW^{*}(L_{1},\iota_{1})^{\otimes l}\to CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1})),

for k=k0,l=k10k=k_{0},l=k_{1}\geq 0, by the formula:

(4.58) nk,l(ck0,,c10,c+,c11,,cl1)=α,β,cdim𝒩¯k0,k1(α,β;c,c+;{Jt}t,H;c0,c1)=0ak,l;α,β;c,c+;{Jt}t,H;c0,c1c.\begin{split}&n^{k,l}(c^{0}_{k},\cdots,c^{0}_{1},c_{+},c^{1}_{1},\cdots,c^{1}_{l})\\ &=\sum_{\begin{subarray}{c}\alpha,\beta,c_{-}\\ \dim\bar{\mathcal{N}}_{k_{0},k_{1}}(\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1})=0\end{subarray}}a_{k,l;\vec{\alpha},\beta;c_{-},c_{+};\{J_{t}\}_{t},H;\vec{c}^{0},\vec{c}^{1}}c_{-}.\end{split}

Again, this is a finite sum by the same argument as that for (4.38), so that the multilinear map nk,ln^{k,l} is well-defined.

These maps satisfy the AA_{\infty}-equations for a curved AA_{\infty}-bimodule, by Proposition (4.36).

In general, n0,0n^{0,0} does not square to zero because of the contribution of m0m^{0} from each (Li,ιi)(L_{i},\iota_{i}). To obtain a differential, we need to assume that both ι0:L0M\iota_{0}:L_{0}\to M and ι1:L1M\iota_{1}:L_{1}\to M are unobstructed, in which case we can deform the curved AA_{\infty}-bimodule structure by the chosen bounding cochains for ι0:L0M\iota_{0}:L_{0}\to M and ι1:L1M\iota_{1}:L_{1}\to M respectively:

(4.59) nk,l;b0,b1(ck0,,c10,c+,c11,,cl1)=i0,j0i0++ik=ij0++jl=jnk+i,l+j(b0,,b0ik times,ck0,b0,,b0ik1 times,,c10,b0,,b0i0 times,c+,b1,,b1j0 times,c11,b1,,b1j1 times,,cl1,b1,,b1jl times).\begin{split}&n^{k,l;b_{0},b_{1}}(c^{0}_{k},\cdots,c^{0}_{1},c_{+},c^{1}_{1},\cdots,c^{1}_{l})\\ &=\sum_{\begin{subarray}{c}i\geq 0,j\geq 0\\ i_{0}+\cdots+i_{k}=i\\ j_{0}+\cdots+j_{l}=j\end{subarray}}n^{k+i,l+j}(\underbrace{b_{0},\cdots,b_{0}}_{\text{$i_{k}$ times}},c^{0}_{k},\underbrace{b_{0},\cdots,b_{0}}_{\text{$i_{k-1}$ times}},\cdots,c^{0}_{1},\underbrace{b_{0},\cdots,b_{0}}_{\text{$i_{0}$ times}},\\ &c_{+},\underbrace{b_{1},\cdots,b_{1}}_{\text{$j_{0}$ times}},c^{1}_{1},\underbrace{b_{1},\cdots,b_{1}}_{\text{$j_{1}$ times}},\cdots,c^{1}_{l},\underbrace{b_{1},\cdots,b_{1}}_{\text{$j_{l}$ times}}).\end{split}

Because of the Maurer-Cartan equations that the bounding cochains satisfy, n0,0;b0,b1n^{0,0;b_{0},b_{1}} squares to zero, and thus defines a differential on CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H). We call the resulting cohomology group the wrapped Floer cohomology group of the pair of the exact cylindrical Lagrangian immersions ι0:L0M\iota_{0}:L_{0}\to M and ι1:L1M\iota_{1}:L_{1}\to M, with respect to the bounding cochains b0b_{0} and b1b_{1}, and denote it by HW((L0,ι0,b0),(L1,ι1,b1);H)HW^{*}((L_{0},\iota_{0},b_{0}),(L_{1},\iota_{1},b_{1});H).

5. Clean intersections

5.1. Clean self-intersections

This section extends wrapped Floer theory to a wider class of immersed Lagrangian submanifolds: exact cylindrical Lagrangian immersions with clean self-intersections. The construction of the moduli spaces follows exactly the same pattern, but there are additional data to be taken into account, which we introduce below.

Definition 5.1.

Let ι:LM\iota:L\to M be a Lagrangian immersion. We say that it has clean self-intersections, if the following conditions are satisfied:

  1. (i)

    The fiber product L×ιL={(p,q)L×L|ι(p)=ι(q)}L\times_{\iota}L=\{(p,q)\in L\times L|\iota(p)=\iota(q)\} is a smooth submanifold of L×LL\times L;

  2. (ii)

    At each point (p,q)L×ιL(p,q)\in L\times_{\iota}L, we have that

    (5.1) T(p,q)(L×ιL)={(V,W)TpL×TqL|dpι(V)=dqι(V)}T_{(p,q)}(L\times_{\iota}L)=\{(V,W)\in T_{p}L\times T_{q}L|d_{p}\iota(V)=d_{q}\iota(V)\}

For such ι:LM\iota:L\to M, we can decompose the fiber product L×ιLL\times_{\iota}L into connected components:

(5.2) L×ιL=aALa=ΔLaA{0}La,L\times_{\iota}L=\coprod_{a\in A}L_{a}=\Delta_{L}\cup\coprod_{a\in A\setminus\{0\}}L_{a},

indexed by an indexing set AA containing a special element 0. Here ΔL\Delta_{L} is the diagonal component, which is identified with LL and labeled by 0A0\in A, and LaL_{a}’s are some other components which are smooth manifolds (possibly of different dimensions). These LaL_{a}’s are called switching components.

In order to make wrapped Floer theory well behaved for such Lagrangian immersions, we must impose further conditions: being exact and cylindrical. Exactness is the same as before: there exists f:Lf:L\to\mathbb{R} such that df=ιλMdf=\iota^{*}\lambda_{M}, while being cylindrical is slightly different.

Definition 5.2.

Let ι:LM\iota:L\to M be an exact Lagrangian immersion with clean self-intersections. We say that it is exact, if there exists f:Lf:L\to\mathbb{R} such that df=ιλMdf=\iota^{*}\lambda_{M}.

We say that ι:LM\iota:L\to M is cylindrical, if there exists an embedded closed Legendrian submanifold ll of M\partial M such that the geometric image ι(L)\iota(L) satisfies

(5.3) ι(L)(M×[1,+))=l×[1,+),\iota(L)\cap(\partial M\times[1,+\infty))=l\times[1,+\infty),

and moreover, the preimage ι1(l×[1,+)\iota^{-1}(l\times[1,+\infty) is a union of copies of l×[1,+)l\times[1,+\infty), so that the restriction of ι\iota is a trivial discrete covering map.

For an exact cylindrical Lagrangian immersion ι:LM\iota:L\to M with clean self-intersections, we have a finer description of the decomposition of the fiber product L×ιLL\times_{\iota}L into its connected components. That is, we shall specify those connected components which are mapped to the cylindrical end l×[1,+)l\times[1,+\infty) of the image. We denote

(5.4) L×ιL=LaLb,L\times_{\iota}L=\coprod L_{a}\cup\coprod L_{b},

where the LbL_{b}’s are the connected components part of which are mapped to the cylindrical end l×[1,+)l\times[1,+\infty), including the diagonal component ΔLL\Delta_{L}\cong L, and the LaL_{a}’s are the ones which are not, and correspond to those self-intersections only contained in the interior part of MM.

Here is a more refined description of the components LbL_{b}. Suppose that ι1(l×[1,+)\iota^{-1}(l\times[1,+\infty) is a union of copies of l×[1,+)l\times[1,+\infty), labeled by iIi\in I for some index set II, which can be infinite. Then the labels bb correspond to pairs (i,j)(i,j) for i,jIi,j\in I, distinguishing the copies LbL_{b} in the fiber product.

For wrapped Floer theory to be well-defined and to give desired AA_{\infty}-structures, we shall from now on assume that the immersion ι:LM\iota:L\to M be proper. In particular, the covering of the cylindrical end l×[1,+)l\times[1,+\infty) has at most finite sheets, say dd-fold.

5.2. Local systems

Similar to the setup of Morse homology of a Morse-Bott function, we need to take into account certain local systems on the components LaL_{a} and LbL_{b} of the fiber product L×ιLL\times_{\iota}L, in order to obtain canonical orientations of the moduli spaces of pseudoholomorphic disks considered in wrapped Floer theory. Most of the definitions follow from Chapter 8 of [FOOO09b], so we just list the essential definitions that we need to fix notations, provide part of the proofs while leaving the full details to the reference.

Let LaL_{a} be any connected component different from the diagonal component ΔLL\Delta_{L}\cong L. Since ι\iota is a Lagrangian immersion, for each xLax\in L_{a} we get two Lagrangian subspaces

(5.5) λa,x,l=dιx(Tpr1(x)L),\displaystyle\lambda_{a,x,l}=d\iota_{x}(T_{pr_{1}(x)}L),
(5.6) λa,x,r=dιx(Tpr2(x)L)\displaystyle\lambda_{a,x,r}=d\iota_{x}(T_{pr_{2}(x)}L)

of Tι(x)MT_{\iota(x)}M, where pr1,pr2:LaL×ιLLpr_{1},pr_{2}:L_{a}\subset L\times_{\iota}L\to L are induced by the projections to the two factors.

Let 𝒫a,x\mathcal{P}_{a,x} be the set of all smooth maps λa,x:[0,1]𝒜𝒢(Tι(x)M)\lambda_{a,x}:[0,1]\to\mathcal{LAG}(T_{\iota(x)}M), such that λa,x(0)=λa,x,l,λa,x(1)=λa,x,r\lambda_{a,x}(0)=\lambda_{a,x,l},\lambda_{a,x}(1)=\lambda_{a,x,r}. Associated to each λa,x𝒫a,x\lambda_{a,x}\in\mathcal{P}_{a,x}, there are two Cauchy-Riemann operators

(5.7) ¯λa,x:L1,δp(Z;Tι(x)M,λa,x)Lδp(Z;Tι(x)M)\bar{\partial}_{\lambda_{a,x}}^{-}:L^{p}_{1,\delta}(Z_{-};T_{\iota(x)}M,\lambda_{a,x})\to L^{p}_{\delta}(Z_{-};T_{\iota(x)}M)

and

(5.8) ¯λa,x+:L1,δp(Z+;Tι(x)M,λa,x)Lδp(Z+;Tι(x)M)\bar{\partial}_{\lambda_{a,x}}^{+}:L^{p}_{1,\delta}(Z_{+};T_{\iota(x)}M,\lambda_{a,x})\to L^{p}_{\delta}(Z_{+};T_{\iota(x)}M)

on the weighted Sobolev spaces. Here

(5.9) Z={z|Re(z)0,|z|1}{z|Re(z)0,1Im(z)1},Z_{-}=\{z\in\mathbb{C}|Re(z)\leq 0,|z|\leq 1\}\cup\{z\in\mathbb{C}|Re(z)\geq 0,-1\leq Im(z)\leq 1\},

and

(5.10) Z+={z|Re(z)0,|z|1}{z|Re(z)0,1Im(z)1}.Z_{+}=\{z\in\mathbb{C}|Re(z)\geq 0,|z|\leq 1\}\cup\{z\in\mathbb{C}|Re(z)\leq 0,-1\leq Im(z)\leq 1\}.

The weighted Sobolev space L1,δp(Z;Tι(x)M,λa,x)L^{p}_{1,\delta}(Z_{-};T_{\iota(x)}M,\lambda_{a,x}) is the L1,δpL^{p}_{1,\delta}-completion of the space of smooth maps u:ZTι(x)Mu:Z_{-}\to T_{\iota(x)}M satisfying the following conditions:

  1. (i)

    u(s+1)λa,x,lu(s+\sqrt{-1})\in\lambda_{a,x,l}, for all s0s\geq 0;

  2. (ii)

    u(s1)λa,x,ru(s-\sqrt{-1})\in\lambda_{a,x,r}, for all s0s\geq 0;

  3. (iii)

    u(e1(π2+πt))λa,x(t)u(e^{\sqrt{-1}(\frac{\pi}{2}+\pi t)})\in\lambda_{a,x}(t), for t[0,1]t\in[0,1];

  4. (iv)

    Zeδ|Re(z)|up𝑑z𝑑z¯<\int_{Z_{-}}e^{\delta|Re(z)|}||\nabla u||^{p}dzd\bar{z}<\infty.

The other weighted Sobolev spaces are defined in similar fashion.

These operators ¯λa,x\bar{\partial}_{\lambda_{a,x}}^{-} and ¯λa,x+\bar{\partial}_{\lambda_{a,x}}^{+} are Fredholm. Consider their determinant lines:

(5.11) Θλa,x±=det(¯λa,x±).\Theta_{\lambda_{a,x}}^{\pm}=\det(\bar{\partial}_{\lambda_{a,x}}^{\pm}).

We wish to move xx as well as λa,x\lambda_{a,x}, so that ¯λa,x±\bar{\partial}_{\lambda_{a,x}}^{\pm} form a family index bundle Ind(D±)Ind(D^{\pm}), and the associated determinant line bundle det¯Ind(D±)\underline{\det}Ind(D^{\pm}) has fiber being (5.11). To make sense of this discussion, we shall first define the space over which the family index bundle is defined.

Definition 5.3.

Define a fiber bundle a,x𝒫a,x\mathcal{I}_{a,x}\to\mathcal{P}_{a,x} in the following five steps:

  1. (i)

    First define (Ia,x)λa,x(I_{a,x})_{\lambda_{a,x}} to be the space of all smooth maps σa,x:[0,1]×nTM\sigma_{a,x}:[0,1]\times\mathbb{R}^{n}\to TM such that for each t[0,1]t\in[0,1], the map σa,x(t,)\sigma_{a,x}(t,\cdot) is a linear isometry between n\mathbb{R}^{n} and λa,x(t)\lambda_{a,x}(t). That is, (Ia,x)λa,x(I_{a,x})_{\lambda_{a,x}} is the space of trivializations along the path λa,x\lambda_{a,x} of Lagrangian subspaces in Tι(x)MT_{\iota(x)}M.

  2. (ii)

    Let PSO(L)P_{SO}(L) be the principal SO(n)SO(n)-bundle associated to the tangent bundle of LL. For the given points p±p_{\pm}, let PSpin(L)p±P_{Spin}(L)_{p_{\pm}} be the double cover of the fiber PSO(L)p±P_{SO}(L)_{p_{\pm}} of PSO(L)P_{SO}(L) at p±p_{\pm}. If x=(p,p+)Lax=(p_{-},p_{+})\in L_{a}, we set

    (5.12) Px=(PSpin(L)p×PSpin(L)p+)/{±1}.P_{x}=(P_{Spin}(L)_{p_{-}}\times P_{Spin}(L)_{p_{+}})/\{\pm 1\}.
  3. (iii)

    Define a map (Ia,x)λa,xPSO(L)p×PSO(L)p+(I_{a,x})_{\lambda_{a,x}}\to P_{SO}(L)_{p_{-}}\times P_{SO}(L)_{p_{+}} as follows. For each σa,x\sigma_{a,x}, consider its restriction to the endpoints t=0,1t=0,1. By definition, σ(0,)\sigma(0,\cdot) is an isometry between n\mathbb{R}^{n} and dιx(TpL)d\iota_{x}(T_{p_{-}}L), hence is canonically identified as an element in the fiber PSO(L)pP_{SO}(L)_{p_{-}}. A parallel argument applies to t=1t=1.

  4. (iv)

    Then define (a,x)λa,x(\mathcal{I}_{a,x})_{\lambda_{a,x}} to be the fiber product

    (5.13) (a,x)λa,x=(Ia,x)λa,x×PSO(L)p×PSO(L)p+Px.(\mathcal{I}_{a,x})_{\lambda_{a,x}}=(I_{a,x})_{\lambda_{a,x}}\times_{P_{SO}(L)_{p_{-}}\times P_{SO}(L)_{p_{+}}}P_{x}.
  5. (v)

    Finally, we consider the union over all paths λa,x\lambda_{a,x}:

    (5.14) a,x=λa,x𝒫a,x(a,x)λa,x.\mathcal{I}_{a,x}=\cup_{\lambda_{a,x}\in\mathcal{P}_{a,x}}(\mathcal{I}_{a,x})_{\lambda_{a,x}}.

    This is the desired fiber bundle a,x𝒫a,x\mathcal{I}_{a,x}\to\mathcal{P}_{a,x}.

Lemma 5.4.

Suppose LL is spin with a chosen spin structure. Then the union xLaa,x\cup_{x\in L_{a}}\mathcal{I}_{a,x} restricts to a fiber bundle a\mathcal{I}_{a} over the 33-skeleton of LaL_{a}.

Proof.

Since LL is spin, there is a globally defined fiberwise double cover PSpin(L)P_{Spin}(L) of PSO(L)P_{SO}(L) over the 33-skeleton (La)[3](L_{a})_{[3]} of LaL_{a}, determined by the spin structure. Thus when defining a,x\mathcal{I}_{a,x}, the definition of PxP_{x} as in (5.12) can be made globally over (La)[3](L_{a})_{[3]}. ∎

From now on we shall always assume LL to be spin, with a chosen spin structure vv. Returning to the concern about family index bundles, by moving xx and λa,x\lambda_{a,x}, the operators ¯λa,x±\bar{\partial}_{\lambda_{a,x}}^{\pm} form a family index bundle Ind(D±)Ind(D^{\pm}) over a\mathcal{I}_{a}, whose determinant line bundle det¯Ind(D±)\underline{\det}Ind(D^{\pm}) is a real line bundle with fiber Θλa,x±\Theta_{\lambda_{a,x}}^{\pm}.

First observe that:

Lemma 5.5.

On each fiber a,x\mathcal{I}_{a,x} of the fiber bundle a(La)[3]\mathcal{I}_{a}\to(L_{a})_{[3]}, the pullback of the determinant line bundle det¯Ind(D±)\underline{\det}Ind(D^{\pm}) is trivial.

Proof.

Fix a reference point [λa,x,σa,x,s1,s2]a,x[\lambda_{a,x},\sigma_{a,x},s_{1},s_{2}]\in\mathcal{I}_{a,x}. Consider the family of operators D={¯λa,x,Z}D^{\prime-}=\{\bar{\partial}_{\lambda^{\prime}_{a,x},Z_{-}}\} parametrized by [λa,x,σa,x,s1,s2]a,x[\lambda^{\prime}_{a,x},\sigma^{\prime}_{a,x},s^{\prime}_{1},s^{\prime}_{2}]\in\mathcal{I}_{a,x}. By gluing DD^{\prime-} with ¯λa,x,Z+\bar{\partial}_{\lambda_{a,x},Z_{+}}, where the latter is a single operator regarded as a constant family, we obtain a family of Dolbeault operators with domain D2D^{2}, with boundary conditions given by the family of real sub-bundles parametrized by S1=D2S^{1}=\partial D^{2}, specified by the union of the paths λa,x\lambda_{a,x} and λa,x\lambda^{\prime}_{a,x}.

Since [λa,x,σa,x,s1,s2][\lambda_{a,x},\sigma_{a,x},s_{1},s_{2}] and [λa,x,σa,x,s1,s2][\lambda^{\prime}_{a,x},\sigma^{\prime}_{a,x},s^{\prime}_{1},s^{\prime}_{2}] determine the spin structures on the family of real sub-bundles consistently, the determinant line bundle of the family of Dolbeault operators is trivial. By definition, this family of Dolbeault operators is obtained from gluing DD^{\prime-} with a constant family of operators, so the determinant line bundle of the family DD^{\prime-} is also trivial, which completes the proof.

Lemma 5.6.

There exist local systems Θa±\Theta_{a}^{\pm} on LaL_{a}, such that their pullbacks to a\mathcal{I}_{a} are isomorphic to Ind(D±)Ind(D^{\pm}). Moreover, there is an isomorphism

(5.15) ΘaΘa+detTLa.\Theta_{a}^{-}\otimes\Theta_{a}^{+}\cong\det TL_{a}.
Proof.

Recall that 𝒫a,x\mathcal{P}_{a,x} is the space of smooth paths in 𝒜𝒢(Tι(x)M)\mathcal{LAG}(T_{\iota(x)}M) connecting the Lagrangian subspaces λa,x,l\lambda_{a,x,l} and λa,x,r\lambda_{a,x,r} in (5.5). Hence it is homotopy equivalent to the based loop-space of the Lagrangian Grassmannian 𝒜𝒢(n)\mathcal{LAG}(n) of linear Lagrangian subspaces of 2n\mathbb{R}^{2n}. Set Ia,x=λa,x(Ia,x)λa,xI_{a,x}=\cup_{\lambda_{a,x}}(I_{a,x})_{\lambda_{a,x}}. The fiber bundle Ia,x𝒫a,xI_{a,x}\to\mathcal{P}_{a,x} is homotopy equivalent to the free loop-space of the Lagrangian Grassmannian 𝒜𝒢(n)\mathcal{LAG}(n). Therefore, π0(Ia,x)=\pi_{0}(I_{a,x})=\mathbb{Z} so that Ia,xI_{a,x} has \mathbb{Z}-worth of connected components, labeled by Ia,x;kI_{a,x;k}.

Also recall that a,x\mathcal{I}_{a,x} is a double cover of Ia,xI_{a,x}, which is non-trivial. Let a,x;k\mathcal{I}_{a,x;k} be the pullback of Ia,x;kI_{a,x;k} to a,x\mathcal{I}_{a,x}, which is therefore connected as Ia,x;kI_{a,x;k} is connected. Let a;k\mathcal{I}_{a;k} be the union of a,x;k\mathcal{I}_{a,x;k} over x(La)[3]x\in(L_{a})_{[3]}. From the fibration

a,x;ka;k(La)[3],\mathcal{I}_{a,x;k}\xhookrightarrow{}\mathcal{I}_{a;k}\to(L_{a})_{[3]},

we obtain a long exact sequence of homotopy groups:

(5.16) π1(a,x;k)π1(a;k)π1((La)[3]){1}.\pi_{1}(\mathcal{I}_{a,x;k})\to\pi_{1}(\mathcal{I}_{a;k})\to\pi_{1}((L_{a})_{[3]})\to\{1\}.

This implies that π1(a;k)π1((La)[3])\pi_{1}(\mathcal{I}_{a;k})\to\pi_{1}((L_{a})_{[3]}) is surjective. Now we have a real line bundle Θa±\Theta_{a}^{\pm} on (La)[3](L_{a})_{[3]}, which is classified by a homomorphism π1((La)[3])/2\pi_{1}((L_{a})_{[3]})\to\mathbb{Z}/2, whose pullback to the homomorphism π1(a;k)/2\pi_{1}(\mathcal{I}_{a;k})\to\mathbb{Z}/2 defined by the real line bundle det¯Ind(D±)\underline{\det}Ind(D^{\pm}). The homomorphism π1(a;k)/2\pi_{1}(\mathcal{I}_{a;k})\to\mathbb{Z}/2 is well-defined because of Lemma 5.5.

Since any line bundle on the 33-skeleton has a unique extension to the whole space, we obtain the desired Θa±\Theta_{a}^{\pm} on LaL_{a}. ∎

Definition 5.7.

The local system Θa\Theta_{a}^{-} is called the orientation local system for the cleanly self-intersecting component LaL_{a}.

The next lemma explains how the local systems change under different choices of spin structures.

Lemma 5.8.

Let v1,v2v_{1},v_{2} be two spin structures on LL. Let Θa(v1),Θa(v2)\Theta_{a}^{-}(v_{1}),\Theta_{a}^{-}(v_{2}) be the orientation local systems defined by v1v_{1} and v2v_{2}, respectively. Then the local system

Θa(v1)Θa(v2)\Theta_{a}^{-}(v_{1})\otimes\Theta_{a}^{-}(v_{2})

is classified by the /2\mathbb{Z}/2-cohomology class

(5.17) pr1(v1v2)pr2(v1v2)H1(La,/2).pr_{1}^{*}(v_{1}-v_{2})-pr_{2}^{*}(v_{1}-v_{2})\in H^{1}(L_{a},\mathbb{Z}/2).

The proof of this lemma also follows from an argument by gluing families of elliptic operators. As we shall not quite use it, we refer the reader to Chapter 8 of [FOOO09b] for the detailed proof.

5.3. The wrapped Floer cochain space in the presence of clean self-intersections

Let ι:LM\iota:L\to M be a cylindrical proper Lagrangian immersion with clean self-intersections, for which we have fixed a grading and a spin structure for it. As in the case of a cylindrical Lagrangian immersion with transverse self-intersections, the wrapped Floer cochain space for (L,ι)(L,\iota) should capture both the topology of the fiber product L×ιLL\times_{\iota}L as well as the non-constant time-one Hamiltonian chords in the cylindrical end of MM starting and ending on ι(L)\iota(L). And because ι\iota is a covering to its cylindrical end, we should keep track of this information when studying the non-constant Hamiltonian chords.

To pick a chain model for the wrapped Floer cochain space, we choose an auxiliary Morse function faf_{a} on each connected component LaL_{a} of the fiber product L×ιLL\times_{\iota}L. Let pa,jp_{a,j} be the critical points of faf_{a}. The corresponding Morse complex computes the cohomology of the fiber product L×ιLL\times_{\iota}L. Let 𝒳+(ι(L);H)\mathcal{X}_{+}(\iota(L);H) be the set of all non-constant time-one HH-chords from ι(L)\iota(L) (the geometric image) to itself which are contained in the cylindrical end M×[1,+)M\times[1,+\infty). These non-constant time-one HH-chords naturally correspond to Reeb chords of all times from ll to itself on the contact boundary M\partial M.

Definition 5.9.

The wrapped Floer cochain space CW(L,ι;H)CW^{*}(L,\iota;H) is defined in a similar way to Definition 4.5 as the free \mathbb{Z}-module generated by the following two kinds of generators

  1. (i)

    (p,w)θp(p,w)\otimes\theta_{p}, where pCrit(fa)p\in Crit(f_{a}), ww is a Γ\Gamma-equivalence class of capping half-disks for pp, and θp(Θa)p\theta_{p}\in(\Theta_{a}^{-})_{p};

  2. (ii)

    (x,b)(x,b), where xx is a non-constant time-one HH-chord from ι(L)\iota(L) to itself which is contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty), and bb is a lifting index, corresponding to a pair (i,j)(i,j) where i,ji,j label the copies of the preimage of the covering ι\iota when restricted to the cylindrical end.

That is,

(5.18) CW(L,ι;H)=(aCM(La,fa;Θa)b=(i,j)=(1,1)(d,d)𝒳+(ι(L);H).CW^{*}(L,\iota;H)=(\bigoplus_{a}CM^{*}(L_{a},f_{a};\Theta_{a}^{-})\oplus\bigoplus_{b=(i,j)=(1,1)}^{(d,d)}\mathbb{Z}\mathcal{X}_{+}(\iota(L);H).

Here b𝒳+(ι(L);H)\bigoplus_{b}\mathbb{Z}\mathcal{X}_{+}(\iota(L);H) means the direct sum of several copies of 𝒳+(ι(L);H)\mathbb{Z}\mathcal{X}_{+}(\iota(L);H), one for each index b=(i,j)b=(i,j).

In the definition of the wrapped Floer cochain space, we include several copies of the free module generated by non-constant HH-chords, in order to keep track of which component of the preimage the boundary map is lifted to.

In particular, when ι:LM\iota:L\to M is a proper embedding, we see that this cochain space is isomorphic to the usual Morse-Bott wrapped Floer cochain space.

5.4. Moduli space of stable pearly trees and Kuranishi structures

To associate a curved AA_{\infty}-algebra to an exact cylindrical Lagrangian immersion with clean self-intersections, we use similar kind of moduli spaces of stable pearly tree maps.

Choosing a Morse cochain complex to compute the cohomology of the fiber product L×ιLL\times_{\iota}L means we have implicitly perturb the Morse-Bott submanifolds LaL_{a} to isolated critical points. This implies that we can define stable pearly tree maps in a similar way to those in the case of a cylindrical Lagrangian immersion with transverse self-intersections, while there are two differences, which are to be discussed below.

First, in Definition 4.18, we need to change the definition of the map α:IS(L,ι)\alpha:I\to S(L,\iota), as in the clean self-intersection case there is no longer a discrete set S(L,ι)S(L,\iota) of preimages of self-intersection points. We simply change it to a set-valued map

(5.19) α:I{La:aA{0}},\alpha:I\to\{L_{a}:a\in A\setminus\{0\}\},

among all the self-intersection components different from the diagonal component ΔLL\Delta_{L}\cong L.

Second, the condition (x)(x) in Definition 4.18 should also be modified to the following condition:

(5.20) (x):(limθ0l(e1θζi),limθ0l(e1θζi))α(i) for iI.In addition ,u~ei(zv(ei))=pa,j(i)ι(α(i)), for some critical point pa,j(i) on α(i).\begin{split}(x)^{\prime}:&(\lim\limits_{\theta\uparrow 0}l(e^{\sqrt{-1}\theta}\zeta_{i}),\lim\limits_{\theta\downarrow 0}l(e^{\sqrt{-1}\theta}\zeta_{i}))\in\alpha(i)\text{ for }i\in I.\\ &\text{In addition },\tilde{u}_{e_{i}}(z_{v(e_{i})})=p_{a,j(i)}\in\iota(\alpha(i)),\text{ for some critical point }p_{a,j(i)}\text{ on }\alpha(i).\end{split}

Third, we have the orientation local systems Θa\Theta_{a}^{-} on the Morse-Bott submanifolds LaL_{a} that needs to be taken into account. We shall explain the role of the orientation local systems in the rest of this subsection.

Let ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) be the moduli space of stable broken pearly tree maps defined in a similar way as that in section 4.5, with the map α\alpha replaced by (5.19), and the condition (x)(x) replaced by (x)(x)^{\prime} in (5.20). As in the case of a cylindrical Lagrangian immersion with transverse self-intersections, we have the following structure results on the moduli space.

Proposition 5.10.

The moduli space ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) has a Kuranishi structure with corners, which is compatible at the boundary with the fiber product Kuranishi structures on

¯k0+1(α0,β0;J,H;c0,0,,c0,k0)×ׯkm+1(αm,βm;J,H;cm,0,,cm,km)\bar{\mathcal{M}}_{k_{0}+1}(\alpha_{0},\beta_{0};J,H;c_{0,0},\cdots,c_{0,k_{0}})\times\cdots\times\bar{\mathcal{M}}_{k_{m}+1}(\alpha_{m},\beta_{m};J,H;c_{m,0},\cdots,c_{m,k_{m}})

as in (4.15).

Remark 5.11.

In the above proposition, we did not mention the word ”oriented”, so for the moment the statement holds for unoriented Kuranishi structures. In the rest of this subsection we shall describe the orientations on these Kuranishi structures.

Next, we shall orient ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) using the orientation local systems Θa\Theta_{a}^{-} on LaL_{a}. First we introduce some notations.

Definition 5.12.

The type function τ\tau is define by

(5.21) τ(i)={a,if ci corresponds to a critical point of fa on La;,if ci corresponds to a non-constant H-chord, appearing in the b-th copy.\tau(i)=\begin{cases}a,&\text{if }c_{i}\text{ corresponds to a critical point of }f_{a}\text{ on }L_{a};\\ \infty,&\text{if }c_{i}\text{ corresponds to a non-constant $H$-chord, appearing in the $b$-th copy}.\end{cases}

By convention, we set Θ=¯\Theta_{\infty}^{-}=\underline{\mathbb{R}} the trivial line bundle, and also detTL=¯\det T^{*}L_{\infty}=\underline{\mathbb{R}}.

Proposition 5.13.

The above Kuranishi structure is oriented. At each point

σ¯k+1(α,β;J,H;c0,,ck),\sigma\in\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}),

if (Uσ,Eσ,sσ,Γσ={1})(U_{\sigma},E_{\sigma},s_{\sigma},\Gamma_{\sigma}=\{1\}) is the Kuranishi chart, then we have a canonical isomorphism of the orientation line bundles

(5.22) detTUσdetEσoc0Θτ(0)(oc1Θτ(1)detTLτ(1))1(ockΘτ(k)detTLτ(k))1,\begin{split}&\det TU_{\sigma}\otimes\det E_{\sigma}^{*}\\ &\cong o_{c_{0}}\otimes\Theta_{\tau(0)}^{-}\otimes(o_{c_{1}}\otimes\Theta_{\tau(1)}^{-}\otimes\det T^{*}L_{\tau(1)})^{-1}\otimes\cdots\otimes(o_{c_{k}}\otimes\Theta_{\tau(k)}^{-}\otimes\det T^{*}L_{\tau(k)})^{-1},\end{split}

where ocio_{c_{i}} is the orientation line of the generator cic_{i}, which is the orientation line for a non-degenerate time-one HH-chord, if cic_{i} corresponds to a non-constant HH-chord, or the orientation line for a critical point of faf_{a}, if cic_{i} corresponds to a critical point of faf_{a}, defined by a choice of coherent orientations on the unstable manifolds of all the critical points, twisted by the orientation local system Θa\Theta_{a}^{-}.

Proof.

The proof is an application of the standard gluing theorem for Cauchy-Riemann operators. We glue the operation operators Dc0,Dc1+,,Dck+D_{c_{0}}^{-},D_{c_{1}}^{+},\cdots,D_{c_{k}}^{+} to the Fredholm operator

Dσ:W1,pLpD_{\sigma}:W^{1,p}\to L^{p}

and obtain a Cauchy-Riemann operator DD with Lagrangian boundary conditions. The operator DD can be deformed through Fredholm operators to the standard Dolbeault operator ¯\bar{\partial}, whose determinant line is canonically trivialized. On the one hand, we have by definition:

detDc0=oc0Θτ(0),\displaystyle\det D_{c_{0}}^{-}=o_{c_{0}}\otimes\Theta_{\tau(0)}^{-},
detDci+=oc01Θτ(i)+,i=1,,k.\displaystyle\det D_{c_{i}}^{+}=o_{c_{0}}^{-1}\otimes\Theta_{\tau(i)}^{+},i=1,\cdots,k.

On the other hand, at the point σ\sigma, there is a canonical isomorphism

detTUσdetEσdetDσ,\det TU_{\sigma}\otimes\det E_{\sigma}^{*}\cong\det D_{\sigma},

because TUσEσTU_{\sigma}\to E_{\sigma} is a finite-dimensional reduction of the Fredholm complex for DσD_{\sigma}. At last, recall that ΘaΘa+detTLa\Theta_{a}^{-}\otimes\Theta_{a}^{+}\cong\det TL_{a}. Thus we deduce that

detDσ(oc0Θτ(0))(oc1Θτ(1)detTLτ(1))1(ockΘτ(k)detTLτ(k))1\det D_{\sigma}\otimes(o_{c_{0}}\otimes\Theta_{\tau(0)}^{-})\otimes(o_{c_{1}}\otimes\Theta_{\tau(1)}^{-}\otimes\det T^{*}L_{\tau(1)})^{-1}\otimes\cdots\otimes(o_{c_{k}}\otimes\Theta_{\tau(k)}^{-}\otimes\det T^{*}L_{\tau(k)})^{-1}

is canonically trivialized. This implies the desired isomorphism (5.22).

Corollary 5.14.

The Kuranishi structure on ¯k+1(α,β;J,H;c0,,ck)\bar{\mathcal{M}}_{k+1}(\alpha,\beta;J,H;c_{0},\cdots,c_{k}) is oriented. Moreover, the induced orientation at the boundary agrees with the fiber product orientation.

From this point, the remaining part of the construction of a curved AA_{\infty}-algebra associated to (L,ι)(L,\iota) is the same as that in the case of a cylindrical Lagrangian immersion with transverse self-intersections.

5.5. A cleanly-intersecting pair

Now consider a pair (ι0:L0M,ι1:L1M)(\iota_{0}:L_{0}\to M,\iota_{1}:L_{1}\to M) of cylindrical proper Lagrangian immersions with clean self-intersections. The goal of this subsection is to define wrapped Floer cohomology for such a pair in the case where the two Lagrangian immersions intersect cleanly in the following sense.

Definition 5.15.

The pair (ι0:L0M,ι1:L1M)(\iota_{0}:L_{0}\to M,\iota_{1}:L_{1}\to M) is said to have clean intersections, if the following conditions are satisfied:

  1. (i)

    the fiber product

    L0×ι0,ι1L1L_{0}\times_{\iota_{0},\iota_{1}}L_{1}

    is a smooth manifold, possibly disconnected with different components having different dimensions,

    (5.23) L0×ι0,ι1L1=aCa.L_{0}\times_{\iota_{0},\iota_{1}}L_{1}=\coprod_{a}C_{a}.
  2. (ii)

    the tangent space of the fiber product at each point is given by

    (5.24) T(p0,p1)(L0×ι0,ι1L1)={(V0,V1)Tp0L0×Tp1L1:dp0ι0(V0)=dp1ι1(V1)}.T_{(p_{0},p_{1})}(L_{0}\times_{\iota_{0},\iota_{1}}L_{1})=\{(V_{0},V_{1})\in T_{p_{0}}L_{0}\times T_{p_{1}}L_{1}:d_{p_{0}}\iota_{0}(V_{0})=d_{p_{1}}\iota_{1}(V_{1})\}.

In wrapped Floer theory, it is important to keep track of the geometry of the cylindrical ends. Recall that for i=0,1i=0,1, ιi:LiM\iota_{i}:L_{i}\to M is assumed to be a discrete trivial covering of a cylindrical end li×[1,+)l_{i}\times[1,+\infty) over the cylindrical end M×[1,+)\partial M\times[1,+\infty) modeled on some Legendrian submanifold lil_{i} of M\partial M. And by the properness assumption, the covering is finitely-sheeted, say did_{i}-fold. We shall make the following assumption on how this pair intersects, distinguishing from the case of a single Lagrangian immersion with clean self-intersections.

Assumption 5.16.

ι0:L0M\iota_{0}:L_{0}\to M and ι1:L1M\iota_{1}:L_{1}\to M do not intersect in the cylindrical end M×[1,+)\partial M\times[1,+\infty). That is, the Legendrian boundaries l0,l1l_{0},l_{1} do not intersect in the contact boundary M\partial M.

This is essentially reduced to an assumption on Legendrian submanifolds of M\partial M, which can be achieved generically, in case l0l1l_{0}\neq l_{1}. The case l0=l1l_{0}=l_{1} but ι0ι1\iota_{0}\neq\iota_{1} is different and slightly more involved, which can be studied in a similar way but will not be discussed in this paper.

To set up wrapped Floer theory for such a pair of Lagrangian immersions, we need to choose a chain model for the wrapped Floer cochain space. We shall for each component CaC_{a} an auxiliary Morse function fa:Caf_{a}:C_{a}\to\mathbb{R}, which is C2C^{2}-small and satisfies the Morse-Smale condition. Let Crit(fa)Crit(f_{a}) be the set of critical points of faf_{a}. The specific choice will be made below.

Lemma 5.17.

There exists a C2C^{2}-small generic perturbation KK of HH so that all the time-one KK-chords that are contained in the interior part of MM are non-degenerate and constant. Moreover, these KK-chords correspond bijectively to the critical points of the lift of KK to CaC_{a}.

Thus it is natural to choose faf_{a} to be the lift of KK to CaC_{a}.

As in the case of a single cylindrical Lagrangian immersion with clean self-intersections, we have rank-one /2\mathbb{Z}/2-local systems Θa±\Theta_{a}^{\pm} on CaC_{a}. They satisfy ΘaΘa+TCa\Theta_{a}^{-}\otimes\Theta_{a}^{+}\cong TC_{a}. We call Θa\Theta_{a}^{-} the orientation local system on CaC_{a}.

Let 𝒳+(ι0(L0),ι1(L1);H)\mathcal{X}_{+}(\iota_{0}(L_{0}),\iota_{1}(L_{1});H) be the set of non-constant time-one HH-chords from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}). These HH-chords are contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty) and naturally correspond to Reeb chords on the contact manifold M\partial M from l0l_{0} to l1l_{1} of all lengths.

The wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H) for a pair of exact cylindrical Lagrangian immersions with clean intersections is defined in a way similar to Definition 4.28.

Definition 5.18.

The wrapped Floer cochain space CW((L0,ι0),(L1,ι1);H)CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H) for a pair of exact cylindrical Lagrangian immersions with clean intersections is the free \mathbb{Z}-module generated by the following two kinds of generators:

  1. (1)

    (p,w)θp(p,w)\otimes\theta_{p}, where pCrit(fa)p\in Crit(f_{a}), ww is a Γ\Gamma-equivalence class of capping half-disks for pp, and θp(Θa)p\theta_{p}\in(\Theta_{a}^{-})_{p};

  2. (2)

    (x,b)(x,b), where xx is a non-constant time-one HH-chord from ι0(L0)\iota_{0}(L_{0}) to ι1(L1)\iota_{1}(L_{1}) which is contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty), and bb is a lifting index, corresponding to a pair (i,j)(i,j) where ii labels the copy of the preimage of the covering ι0\iota_{0}, and jj labels the copy of the preimage of the cover ι1\iota_{1}, when restricted to the cylindrical end.

That is,

(5.25) CW((L0,ι0),(L1,ι1);H)=(aCM(Ca,fa;Θa)b=(i,j)=(1,1)(d0,d1)𝒳+(ι0(L0),ι1(L1);H).CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H)=(\bigoplus_{a}CM^{*}(C_{a},f_{a};\Theta_{a}^{-})\oplus\bigoplus_{b=(i,j)=(1,1)}^{(d_{0},d_{1})}\mathbb{Z}\mathcal{X}_{+}(\iota_{0}(L_{0}),\iota_{1}(L_{1});H).

5.6. Floer trajectories between a cleanly-intersecting pair

The moduli spaces of Floer trajectories are similar to those in the case of a transversely intersecting pair of cylindrical Lagrangian immersions with transverse self-intersections. So we just outline the definitions emphasizing the differences.

Choosing a Morse cochain complex to compute the cohomology of the fiber product L0×ι0,ι1L1L_{0}\times_{\iota_{0},\iota_{1}}L_{1} implicitly perturbs the Morse-Bott submanifolds CaC_{a} to isolated critical points. Thus the only difference is that in the clean intersecting case we have the orientation local systems Θa\Theta_{a}^{-} on the components CaC_{a} of the fiber product L0×ι0,ι1L1L_{0}\times_{\iota_{0},\iota_{1}}L_{1}. The discussion on the role of these orientation local systems is parallel to that in the case of a single cylindrical Lagrangian immersion with clean self-intersections. We simply state the results as the proofs are completely similar.

Proposition 5.19.

There exists a Kuranishi structure on the moduli space of stable broken Floer trajectories

𝒩¯k,l(α,β;{Jt}t,H;c,c+;c0,c1),\bar{\mathcal{N}}_{k,l}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1}),

which is compatible with the fiber product Kuranishi structure on the boundary (4.55).

Proposition 5.20.

The Kuranishi structure on the moduli space

𝒩¯k,l(α,β;{Jt}t,H;c,c+;c0,c1)\bar{\mathcal{N}}_{k,l}(\vec{\alpha},\beta;\{J_{t}\}_{t},H;c_{-},c_{+};\vec{c}^{0},\vec{c}^{1})

is oriented. The orientation is defined as follows. If σ\sigma is any point in the moduli space with Kuranishi chart (Uσ,Eσ,sσ,Γσ={1})(U_{\sigma},E_{\sigma},s_{\sigma},\Gamma_{\sigma}=\{1\}), then there is an isomorphism

(5.26) detTUσdetEσocΘτ()(oc+Θτ(+)detTLτ(+))1i=1k(oci0Θτ(0,i)detTLτ(0,i))1j=1l(ocj1Θτ(1,j)detTLτ(1,j))1.\begin{split}\det TU_{\sigma}\otimes\det E_{\sigma}^{*}\cong&o_{c_{-}}\otimes\Theta_{\tau(-)}^{-}\otimes(o_{c_{+}}\otimes\Theta_{\tau(+)}^{-}\otimes\det T^{*}L_{\tau(+)})^{-1}\\ &\otimes\bigotimes_{i=1}^{k}(o_{c^{0}_{i}}\otimes\Theta_{\tau(0,i)}^{-}\otimes\det T^{*}L_{\tau(0,i)})^{-1}\\ &\otimes\bigotimes_{j=1}^{l}(o_{c^{1}_{j}}\otimes\Theta_{\tau(1,j)}^{-}\otimes\det T^{*}L_{\tau(1,j)})^{-1}.\end{split}

5.7. On the category level

Now let us summarize the results above and extend them to a categorical level. This yields to the definition of the (unobstructed) immersed wrapped Fukaya category 𝒲im(M)\mathcal{W}_{im}(M). The objects are pairs ((ι:LM),b)((\iota:L\to M),b) (or sometimes denoted by (L,ι,b)(L,\iota,b)), where ι:LM\iota:L\to M is an exact cylindrical Lagrangian immersion with transverse or clean self-intersections, and bb is a bounding cochain for it. The morphism space between a pair of objects is the wrapped Floer cochain space

hom𝒲im(M)((L0,ι0,b0),(L1,ι1,b1))=CW((L0,ι0),(L1,ι1);H),\hom_{\mathcal{W}_{im}(M)}((L_{0},\iota_{0},b_{0}),(L_{1},\iota_{1},b_{1}))=CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H),

which is independent of the bounding cochains as a \mathbb{Z}-module. However, we sometimes also write the morphism space as CW((L0,ι0,b0),(L1,ι1,b1))CW^{*}((L_{0},\iota_{0},b_{0}),(L_{1},\iota_{1},b_{1})) to remember the choices of bounding cochains. The first order structure map m𝒲im(M)1m^{1}_{\mathcal{W}_{im}(M)} for the pair ((L0,ι0,b0),(L1,ι1,b1))((L_{0},\iota_{0},b_{0}),(L_{1},\iota_{1},b_{1})) is given by the (b0,b1)(b_{0},b_{1})-deformed differential

n0,0;b0,b1:CW((L0,ι0),(L1,ι1);H)CW((L0,ι0),(L1,ι1);H)[1],n^{0,0;b_{0},b_{1}}:CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H)\to CW^{*}((L_{0},\iota_{0}),(L_{1},\iota_{1});H)[1],

where [1][1] means the map is of degree +1+1. Higher order structure maps

(5.27) m𝒲im(M)k:CW((Lk1,ιk1,bk1),(Lk,ιk,bk);H)CW((L0,ι0,b0),(L1,ι1,b1);H)CW((L0,ι0,b0),(Lk,ιk,bk);H)\begin{split}m^{k}_{\mathcal{W}_{im}(M)}&:CW^{*}((L_{k-1},\iota_{k-1},b_{k-1}),(L_{k},\iota_{k},b_{k});H)\otimes\cdots\otimes CW^{*}((L_{0},\iota_{0},b_{0}),(L_{1},\iota_{1},b_{1});H)\\ &\to CW^{*}((L_{0},\iota_{0},b_{0}),(L_{k},\iota_{k},b_{k});H)\end{split}

are defined by appropriate counts of inhomogeneous pseudoholomorphic disks, with insertions of bounding cochains bib_{i} on the boundary components of the disks. These disks are similar to those used in the definition of the ordinary wrapped Fukaya category, but now they are modeled as stable pearly tree maps as in subsection 4.5, with boundary conditions replaced by multiple Lagrangian immersions ιj:LjM,j=0,,k\iota_{j}:L_{j}\to M,j=0,\cdots,k.

Proposition 5.21.

With the bounding cochains bib_{i}’s taken into account, the structure maps m𝒲im(M)km^{k}_{\mathcal{W}_{im}(M)} satisfy the equations for a non-curved AA_{\infty}-category over \mathbb{Z}, i.e. m𝒲im(M)0=0m^{0}_{\mathcal{W}_{im}(M)}=0.

The more formal definitions are based on the construction of Kuranishi structures on the moduli spaces of these stable pearly tree maps, choices of single-valued multisections of the Kuranishi charts, for each tuple of Lagrangian immersions involved in the structure maps. Technically, we should work with a countable collection of Lagrangian immersions in order to be able to make consistent choices of (abstract) perturbations. In that way, we should obtain a curved AA_{\infty}-category 𝒲ob,im(M)\mathcal{W}_{ob,im}(M) whose unobstructed deformation gives 𝒲im(M)\mathcal{W}_{im}(M). Moreover, this curved AA_{\infty}-category should allow us to define the weakly unobstructed wrapped Fukaya category with bulk deformations. We will not discuss that construction in detail as it is not needed in this paper. However, the perspective of deformation theory will be of interest and addressed in an upcoming work on generalizing the Viterbo functor [Gao].

5.8. A quasi-embedding

A basic but important property of the immersed wrapped Fukaya category is that the ordinary wrapped Fukaya category, consisting of properly embedded exact cylindrical Lagrangian submanifolds, embeds into the immersed wrapped Fukaya category as a full sub-category.

Proposition 5.22.

The wrapped Fukaya category 𝒲(M)\mathcal{W}(M) quasi-embeds into the immersed wrapped Fukaya category 𝒲im(M)\mathcal{W}_{im}(M). That is, there is a canonical cohomologically fully faithful AA_{\infty}-functor

(5.28) j:𝒲(M)𝒲im(M).j:\mathcal{W}(M)\to\mathcal{W}_{im}(M).
Proof.

There are two definitions for the ordinary wrapped Fukaya category 𝒲(M)\mathcal{W}(M): one is using a Morse-Bott Hamiltonian as in the case of immersed wrapped Fukaya category, the other is using a non-degenerate Hamiltonian as usual.

If we use the first definition of 𝒲(M)\mathcal{W}(M), this is obvious, as 𝒲(M)\mathcal{W}(M) is naturally an AA_{\infty}-sub-category of 𝒲im(M)\mathcal{W}_{im}(M), thus we may simply take jj to be the natural inclusion.

Let us consider the second definition. Let KK be a non-degenerate time-independent Hamiltonian, which is used to define the wrapped Fukaya category 𝒲(M)\mathcal{W}(M). We shall define an AA_{\infty}-functor

j:𝒲(M)𝒲im(M),j:\mathcal{W}(M)\to\mathcal{W}_{im}(M),

which on the level of objects acts by L(L,0)L\mapsto(L,0). The zeroth order term vanishes, j0=0j^{0}=0, and the first order term

j1:CW(Li,Lj;K)CW((Li,0),(Lj,0);H)j^{1}:CW^{*}(L_{i},L_{j};K)\to CW^{*}((L_{i},0),(L_{j},0);H)

is defined as follows. On the one hand, CW(Li,Lj;K)CW^{*}(L_{i},L_{j};K) is generated by time-one KK-chords from LiL_{i} to LjL_{j}: these are either Hamiltonian chords that are contained in the interior part M0M_{0}, or those that are contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty). On the other hand, CW((Li,0),(Lj,0);H)CW^{*}((L_{i},0),(L_{j},0);H) is generated by two kinds of generators: critical points of auxiliary Morse functions on the components of the intersection LiLjL_{i}\cap L_{j}, as well as time-one KK-chord that are contained in the cylindrical end M×[1,+)\partial M\times[1,+\infty). Note that HH and KK agree in the cylindrical end, so that the corresponding Hamiltonian chords which are contained in the cylindrical end agree.

If LiL_{i} and LjL_{j} intersect transversely in the interior part M0M_{0}, these critical points are in one-to-one correspondence with the isolated intersection points of LiLjL_{i}\cap L_{j}. In this case, we define the map j1j^{1} as follows. j1j^{1} is identity for Hamiltonian chords that are contained in the cylindrical end. For time -one KK-chords that are contained in the interior part M0M_{0}, we define the map j1j^{1} by counting rigid elements in the following parametrized moduli space (p,x)\mathcal{M}(p,x) of inhomogeneous pseudoholomorphic maps

w:SM,w:S\to M,

where SS is a disk with one positive boundary puncture z+z_{+} and one boundary marked point z0z_{0}, satisfying the equation:

(dwdtXHS)0,1=0,(dw-dt\otimes X_{H_{S}})^{0,1}=0,

for a domain-dependent family of Hamiltonians HSH_{S} which is equal to KK near z+z_{+} and HH near zz_{-} (identifying SS with an infinite strip by removing zz_{-}, we have a natural coordinate t[0,1]t\in[0,1]), as well as the condition:

w(z0)=p,lims+wϵ+(s,)=x(),w(z_{0})=p,\lim\limits_{s\to+\infty}w\circ\epsilon_{+}(s,\cdot)=x(\cdot),

where pp is an intersection point of LiLjL_{i}\cap L_{j}, and xx is a time-one KK-chord from LiL_{i} to LjL_{j} which is contained in the interior part M0M_{0}.

If LiL_{i} and LjL_{j} intersect cleanly in the interior part M0M_{0}, there is a decomposition

LiLj=aCaL_{i}\cap L_{j}=\coprod_{a}C_{a}

into connected components, so that the wrapped Floer cochain space is generated by critical points of faf_{a} for some chosen Morse functions faf_{a} on CaC_{a}. There is an orientation local system Θa\Theta_{a}^{-} so that the Θa\Theta_{a}^{-}-twisted Morse complex of faf_{a} contributes to the wrapped Floer cochain space CW((Li,0),(Lj,0);H)CW^{*}((L_{i},0),(L_{j},0);H). In this case, we define the map j1j^{1} on generators of CW(Li,Lj;K)CW^{*}(L_{i},L_{j};K) that come from interior time-one KK-chords using slightly modified moduli spaces Ca((ξa)pa,x)\mathcal{M}_{C_{a}}((\xi_{a})_{p_{a}},x), defined as follows. Let S¯\bar{S} be union of a disk SS with one positive boundary puncture z+z_{+} and one boundary marked point z0z_{0} and a half-infinite ray T(,0]T\cong(-\infty,0] joint to SS at the point z0z_{0}. Let w:S¯Mw:\bar{S}\to M be a continuous map which is the union of two maps wS:SMw_{S}:S\to M and wT:TCaMw_{T}:T\to C_{a}\subset M, such that wSw_{S} is an inhomogeneous pseudoholomorphic map with two boundary components lying on LiL_{i} and LjL_{j} respectively, and wTw_{T} is a gradient flow line for faf_{a} on CaC_{a}. Moreover, the map ww satisfies the following conditions:

limswT(s)=pa,\displaystyle\lim\limits_{s\to-\infty}w_{T}(s)=p_{a},
wT(0)=wS(z0),\displaystyle w_{T}(0)=w_{S}(z_{0}),
lims+wSϵ+(s,)=x().\displaystyle\lim\limits_{s\to+\infty}w_{S}\circ\epsilon_{+}(s,\cdot)=x(\cdot).

In addition, we impose the following monodromy condition: the monodromy of ww around the boundary of S¯\bar{S} (starting from -\infty of the infinite-half ray ll to 0, going around S\partial S, then returning back to -\infty) is ξaΘa\xi_{a}\in\Theta_{a}^{-}. Counting rigid elements in such a moduli space defines the map j1j^{1} on generators of CW(Li,Lj;K)CW^{*}(L_{i},L_{j};K) coming from interior KK-chords.

Higher order terms can be defined in a similar way. ∎

Remark 5.23.

In the above proof we have constructed several moduli spaces (p,x)\mathcal{M}(p,x) and Ca(pa,x)\mathcal{M}_{C_{a}}(p_{a},x) using which we define the map j1j^{1}. Since the Lagrangian submanifolds Li,LjL_{i},L_{j} are embedded and exact, we can use classical transversality argument to prove that these moduli spaces are smooth manifolds of expected dimension for generic perturbation, so that we can actually count rigid elements.

6. Wrapped Floer theory in the product manifold

6.1. Overview

In view of Lagrangian correspondences as Lagrangian submanifolds in the product symplectic manifold, we need to study wrapped Floer theory of the product manifold in order to understand the construction of functors between the wrapped Fukaya categories. Then a somewhat technical but essential problem arises: there are two natural models for the wrapped Fukaya category of the product manifold, which are obviously equivalent. This causes some trouble in understanding wrapped Floer theory of the product manifold.

Denote by HM,N=πMHM+πNHNH_{M,N}=\pi_{M}^{*}H_{M}+\pi_{N}^{*}H_{N} the split Hamiltonian and JM,N=JM×JNJ_{M,N}=J_{M}\times J_{N} the product almost complex structure. Using HM,NH_{M,N} and JM,NJ_{M,N}, we can define a version of the wrapped Fukaya category of the product Liouville manifold M×NM\times N, which we call the split model of wrapped Fukaya category of M×NM\times N and denote it by 𝒲s(M×N)\mathcal{W}^{s}(M\times N). On the other hand, there is a natural choice of the cylindrical end Σ×[1,+)\Sigma\times[1,+\infty) for M×NM\times N (see [Gao17]), which allows us to define Hamiltonian functions KK that are quadratic at infinity. Therefore the ordinary wrapped Fukaya category of M×NM\times N can also be defined as usual.

However one important thing is not mentioned explicitly. That is, the classes of Lagrangian submanifolds for which we can define wrapped Floer cohomology and AA_{\infty}-algebras are a priori different with respect to the two kinds of Hamiltonians and almost complex structures. For the the split Hamiltonian HM,NH_{M,N} and product almost complex structure JM,NJ_{M,N}, the natural class of Lagrangian submanifolds are products of objects in 𝒲(M)\mathcal{W}(M) and those in 𝒲(N)\mathcal{W}(N). For the quadratic Hamiltonian KK and almost complex structure JJ, the natural class of Lagrangian submanifolds are cylindrical Lagrangian submanifolds with respect to the cylindrical end Σ×[1,+)\Sigma\times[1,+\infty). If we want to identify these two versions of wrapped Floer theory, we must ensure that both classes of objects can be included in each wrapped Fukaya category. One of the main results in [Gao17] confirms that this is possible, which we will review in subsection 6.2.

This section is devoted to proving Theorem 1.1, which says 𝒲s(M×N)\mathcal{W}^{s}(M\times N) is quasi-equivalent to 𝒲(M×N)\mathcal{W}(M\times N). Thus the ambiguity in differentiating these two models is removed, which allows us to better understand the functoriality properties of wrapped Fukaya categories with respect to Lagrangian correspondences.

Remark 6.1.

In this section all the wrapped Fukaya categories are to consist of embedded Lagrangian submanifolds, but the result should continue to hold in the immersed case, following similar lines of argument. Nonetheless, note that the action filtration no longer exists, but instead we need to use energy filtration instead.

The strategy of proof goes as follows. In [Gao17], the action-restriction map is constructed. We would like to extend it to an AA_{\infty}-functor, including it as the first-order term.

Fix an ordering on the collection 𝕃={i}i=1\mathbb{L}=\{\mathcal{L}_{i}\}_{i=1}^{\infty} and consider the finite collection 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d}. In the next subsection we will introduce certain geometric data which we call action-restriction data, which are used to define a AA_{\infty}-functor RdR_{d} from the AA_{\infty}-subcategory 𝒲ds\mathcal{W}^{s}_{d} of 𝒲s(𝕃)\mathcal{W}^{s}(\mathbb{L}) consisting of 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} as objects to the AA_{\infty}-subcategory 𝒲d\mathcal{W}_{d} of 𝒲(𝕃)\mathcal{W}(\mathbb{L}) with the same objects, which acts by identity on the level of objects, and induces quasi-isomorphisms on all morphism spaces. Therefore this AA_{\infty}-functor RdR_{d} is a quasi-isomorphism 𝒲ds𝒲d\mathcal{W}^{s}_{d}\to\mathcal{W}_{d}. Moreover, the AA_{\infty}-functor RdR_{d} will be cohomologically unital.

This would finish the proof of a weak version of Theorem 1.1 when M×NM\times N is non-degenerate, which means that the two models of wrapped Fukaya categories are both split-generated by finitely many Lagrangian submanifolds.

To deal with the general case, we need an additional argument with the help of homological algebra discussed in subsection 2.6. We will return to this point later.

Remark 6.2.

Of course, the quasi-isomorphism will depend on choices of action-restriction data. But for the purpose of comparing the two versions of wrapped Fukaya categories of the product, this does not matter, since the Fukaya category itself is only well-defined up to quasi-isomorphism, because of the flexibility in the choice of auxilliary data in Floer theory. The choices themselves form a contractible space, so we might work harder to show that different functors constructed in this way are homotopic, though we will not attempt to do so in this paper.

6.2. Well-definedness

One thing that needs explanation is why the same class of objects in 𝒲(M×N)\mathcal{W}(M\times N) can be included in 𝒲s(M×N)\mathcal{W}^{s}(M\times N), and vice versa. This is discussed in detail in [Gao17]. We recall some relevant notions.

Definition 6.3.

A Lagrangian submanifold M×N\mathcal{L}\subset M\times N is called admissible, if it is exact and cylindrical, and moreover satisfies c1(M×N,)=0c_{1}(M\times N,\mathcal{L})=0. More specifically, \mathcal{L} can be one of the following kinds:

  1. (i)

    a product L×LL\times L^{\prime}, where LL is an exact cylindrical Lagrangian submanifold of MM, and LL^{\prime} is one of NN;

  2. (ii)

    a cylindrical Lagrangian submanifold with respect to the natural cylindrical end Σ×[1,+)\Sigma\times[1,+\infty).

Remark 6.4.

Note that the notion of being cylindrical only depends on the Liouville structure, but not a particular choice of cylindrical end. In particular, class (i) belongs to class (ii). However, we distinguish them as it will be convenient when we discuss well-definedness of wrapped Floer theory.

One of the main results in [Gao17] is about the well-definedness of the two versions of wrapped Floer cohomology for an admissible Lagrangian submanifold M×N\mathcal{L}\subset M\times N. The argument can easily be generalized, using the same methods, to prove the following result:

Proposition 6.5.

Suppose M×N\mathcal{L}\subset M\times N is an admissible Lagrangian submanifold. Then wrapped Floer AA_{\infty}-algebra associated to \mathcal{L} is well-defined, with respect to either the split Hamiltonian HM,NH_{M,N} and product almost complex structure JM,NJ_{M,N}, or the quadratic Hamiltonian KK and almost complex structure JJ.

Sketch of proof.

This is proved in [Gao17]. We shall review some key arguments while referring to [Gao17] for details. There are two key issues in proving well-definedness: transversality and compactness of the moduli spaces of pseudoholomorphic disks.

Regarding transversality, perturbing the almost complex structure JM,NJ_{M,N} within the class of product almost complex structures 𝒥(M)×𝒥(N)\mathcal{J}(M)\times\mathcal{J}(N) might not ensure enough genericity to make the moduli spaces regular. However, when defining the wrapped Fukaya category, we shall take domain-dependent perturbations of almost complex structures in a suitable way. Namely, if we allow perturbations of JM,NJ_{M,N} in a neighborhood in 𝒥(M×N)\mathcal{J}(M\times N) instead of just product almost complex structures, then transversality can be achieved. On the other hand, if one of the Lagrangian submanifolds involved in mkm^{k} is a product Lagrangian L×LL\times L^{\prime}, it is true that transversaliy of the moduli spaces involved can be achieved by perturbations within the class of product almost complex structures. The argument is similar to Wehrheim and Woodward’s argument in quilted Floer cohomology (for compact monotone or exact Lagrangian submanifolds), combined with the transversality argument in wrapped Floer theory.

Compactness is the substantial issue that we need to think carefully about. Consider for simplicity the case of a pair (0,1)(\mathcal{L}_{0},\mathcal{L}_{1}) for which we want to define wrapped Floer cohomology using the split Hamiltonian HM,NH_{M,N} and product almost complex structure JM,NJ_{M,N}. For a pair of HM,NH_{M,N}-chords x¯0,x¯1\underline{x}_{0},\underline{x}_{1} from 0\mathcal{L}_{0} to 1\mathcal{L}_{1}, the moduli space (x¯0,x¯1)=~(x¯0,x¯1)/\mathcal{M}(\underline{x}_{0},\underline{x}_{1})=\tilde{\mathcal{M}}(\underline{x}_{0},\underline{x}_{1})/\mathbb{R} of inhomogeneous pseudoholomorphic strips from xx to yy has a natural Gromov bordification ¯(x¯0,x¯1)\bar{\mathcal{M}}(\underline{x}_{0},\underline{x}_{1}) by adding broken strips. To ensure that ¯(x¯0,x¯1)\bar{\mathcal{M}}(\underline{x}_{0},\underline{x}_{1}) is compact, the main ingredient in addition to Gromov compactness is the maximum principle, which prevents strips connecting x¯0\underline{x}_{0} and x¯1\underline{x}_{1} from escaping to infinity. By directly estimating the action of Hamiltonian chords, and using the action-energy equality to give an a priori estimate for the energy of inhomogeneous pseudoholomorphic disks. On the other hand, we need a C0C^{0}-bound, which can be achieved by comparing the split Hamiltonian HM,NH_{M,N} with an admissible one. The argument is done when the action-restriction map is constructed and proved to be a cochain map. Let us recall that here. Given x¯0\underline{x}_{0} and x¯1\underline{x}_{1}, we choose bb so that their action is greater than b-b, and correspondingly a large compact subset of M×NM\times N containing all Hamiltonian chords whose action is greater than b-b. Then we modify the Hamiltonian HM,NH_{M,N} to another one KbK_{b} which agrees with HM,NH_{M,N} inside of this compact set, and is quadratic in the radial coordinate on Σ×[B,+)\Sigma\times[B,+\infty) for some BB large enough (depending on bb). A corresponding almost complex structure is also constructed, which is of contact type in Σ×[B,+)\Sigma\times[B,+\infty). It is then shown that (HM,N,JM,N)(H_{M,N},J_{M,N})-pseudoholomorphic strips connecting x¯0\underline{x}_{0} and x¯1\underline{x}_{1} are in one-to-one correspondence with (Kb,Jb)(K_{b},J_{b})-pseudoholomorphic strips connecting x¯0\underline{x}_{0} and x¯1\underline{x}_{1}. In particular, maximum principle can be applied to give C0C^{0}-bound for (Kb,Jb)(K_{b},J_{b})-pseudoholomorphic strips, and therefore also (HM,N,JM,N)(H_{M,N},J_{M,N})-pseudoholomorphic strips. Such analysis is done in details in [Gao17].

The second part of the compactness result in wrapped Floer theory is that for fixed x¯1\underline{x}_{1}, the moduli space ¯(x¯0,x¯1)\bar{\mathcal{M}}(\underline{x}_{0},\underline{x}_{1}) is empty for all but finitely many x¯0\underline{x}_{0}’s. This can be proved again using the action-energy equality. In the case of a single MM, the statement follows from the fact that if there is an Hamiltonian chord xx lying on the level hypersurface M×{r}\partial M\times\{r\}, then its action is roughly r2-r^{2} (with a difference contributed from the primitives of the Lagrangian submanifolds, which are negligible for large rr).

Now in case of split Hamiltonian HM,NH_{M,N} on the product M×NM\times N, the Hamiltonian vector field XHM,NX_{H_{M,N}} no longer agrees with the Reeb vector field on the contact hypersurface we have chosen, because that hypersurface is a small deformation of the singular hypersurface (M×N)\partial(M\times N). As a consequence, if x¯=(x,x)\underline{x}=(x,x^{\prime}) is an HM,NH_{M,N}-chord, then xx is an HMH_{M}-chord and xx^{\prime} is an HNH_{N}-chord. In the cylindrican end, such a chord is no longer exactly located on some smooth contact hypersurface Σ×{r}\Sigma\times\{r\}. There are three places where such chords can occur. The first place is M×[1,+)×N×[1,+)\partial M\times[1,+\infty)\times\partial N\times[1,+\infty), where xx lies on some hypersurface M×{r1}\partial M\times\{r_{1}\} and xx^{\prime} on N×{r2}\partial N\times\{r_{2}\}. We can calculate its action to be r12r22+c-r_{1}^{2}-r_{2}^{2}+c, where cc is a uniformly bounded (i.e. independent of the radial coordinates r1,r2r_{1},r_{2}) constant coming from the primitives of 0\mathcal{L}_{0} and 1\mathcal{L}_{1}. The second place is int(M)×N×[1,+)int(M)\times\partial N\times[1,+\infty), where now xx is an HMH_{M}-chord in the interior of MM (there are finitely many such ”interior” chords, all of which have small action), and xx^{\prime} lies on some hypersurface N×{r2}\partial N\times\{r_{2}\}. The third place is M×[1,+)×N\partial M\times[1,+\infty)\times N. This case is similar to the second one.

These action computations yield the desired compactness result: since the action of such chords must also go to negative infinity as the positions of them tend to infinity, there cannot be infinitely many inhomogeneous pseudoholomorphic disks with different outputs which tend to infinity, as the energy of any inhomogeneous pseudoholomorphic disk is always positive.

We have thus shown that the wrapped Floer cohomology for an admissible pair (0,1)(\mathcal{L}_{0},\mathcal{L}_{1}) is well-defined using the split Hamiltonian HM,NH_{M,N} and product almost complex structure JM,NJ_{M,N} (but with generic domain-dependent perturbations). The same argument applies to the definition of higher structure maps mkm^{k}, by using action-energy equality for inhomogeneous pseudoholomorphic disks with more punctures.

6.3. Action-restriction data

As mentioned at the beginning of section 6 main technical issue in studying Floer theory on the product manifold M×NM\times N and relating that to quilted Floer theory is with regard to the choice of Hamiltonian functions. Having found the cylindrical end Σ×[1,+)\Sigma\times[1,+\infty), one can immediately set up wrapped Floer theory using a Hamiltonian that depends only on the radial coordinate rr in the cylindrical end and has quadratic growth for rr large enough. On the other hand, what is more directly related to quilted Floer theory is the split Hamiltonian, i.e. the sum of the two Hamiltonians on both factors MM and NN. However, the split Hamiltonian is not a priori admissible in the usual sense of wrapped Floer theory. Also, there is a similar issue with almost complex structures. Thus it is not immediately clear that the resulting two versions of wrapped Fukaya categories are equivalent.

By the invariance nature of Floer cohomology, it is expected that these two versions are equivalent. In [Gao17], we showed that these two types of geometric data define isomorphic wrapped Floer cohomology, by constructing a cochain quasi-isomorphism between the two versions of wrapped Floer cochain complexes. Let us briefly summarize the idea here. The action functional defines a filtration on the wrapped Floer complex CW(0,1;HM,N,JM,N)CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N}). Consider the truncated complex CW(b,a](0,1;HM,N,JM,N)CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N}). Starting from the split Hamiltonian, we may construct an admissible Hamiltonian KbK_{b} which agrees with the split Hamiltonian inside some large compact set CbC_{b} depending on bb, and outside a bigger compact set CbC^{\prime}_{b} is quadratic in the radial coordinate rr on Σ×[1,+)\Sigma\times[1,+\infty). Note that the wrapped Floer complexes as graded \mathbb{Z}-modules do not depend on the choice of almost complex structure. Thus the generators of CW(b,a](0,1;HM,N,JM,N)CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N}) are automatically generators of CW(b,a](0,1;Kb,Jb)CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K_{b},J_{b}) for any admissible JbJ_{b}. We therefore get a homomorphism of graded modules

(6.1) R¯b1:CW(b,a](0,1;HM,N,JM,N)CW(b,a](0,1;Kb,Jb).\bar{R}^{1}_{b}:CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N})\to CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K_{b},J_{b}).

We proved in [Gao17] that this is an isomorphism of graded modules by showing that some extra chords of the Hamiltonian KbK_{b} outside CbC_{b} but inside CbC^{\prime}_{b} have sufficiently positive action, and consequently those chords do not lie in the truncated complex. By modifying JM,NJ_{M,N} to an admissible almost complex structure JbJ_{b}, we proved that (6.1) is a cochain map.

To extend the above map to the whole wrapped Floer complex, we need to identify each truncated Floer complex CW(b,a](0,1;Kb,Jb)CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K_{b},J_{b}) with one defined with respect to a Hamiltonian and an almost complex structure independent of bb.

Lemma 6.6.

There exists a Hamiltonian KK which is quadratic in the radial coordinate on Σ×[1,+)\Sigma\times[1,+\infty) away from a compact set, and an almost complex structure of contact type with respect to Σ\Sigma, such that for any b>0b>0, there is a cochain homotopy equivalence

(6.2) hb:CW(b,a](0,1;Kb,Jb)CW(b,a](0,1;K,J).h_{b}:CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K_{b},J_{b})\to CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K,J).
Proof.

This is proved in [Gao17]. Let us briefly recall the argument here. The Hamiltonian KbK_{b} is quadratic in the radial coordinate rr on Σ×[1,+)\Sigma\times[1,+\infty) for rBr\geq B where BB is a suitable number depending quadratically on bb. We may use the time-log(B)-\log(B) Liouville flow to rescale the Hamiltonian so that the rescaled Hamiltonian is quadratic outside a neighborhood of r1r\leq 1. The resulting Hamiltonian KbK^{\prime}_{b} still depends on bb, but we can choose a particular Hamiltonian KK which agrees with KbK^{\prime}_{b} outside a neighborhood of r1r\leq 1, such that there is a compactly-supported homotopy between KbK^{\prime}_{b} and KK. This homotopy induces a cochain homotopy equivalence between the corresponding wrapped Floer complexes. ∎

Then we compose the above cochain homotopy equivalence with (6.1) (which we still denote by the same symbol) to obtain a cochain map:

(6.3) Rb1:CW(b,a](0,1;HM,N,JM,N)CW(b,a](0,1;K,J),R^{1}_{b}:CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N})\to CW^{*}_{(-b,a]}(\mathcal{L}_{0},\mathcal{L}_{1};K,J),
(6.4) Rb1=hbR¯b1,R^{1}_{b}=h_{b}\circ\bar{R}^{1}_{b},

for admissible KK and JJ independent of bb. These maps are shown to commute up to homotopy with natural inclusions of subcomplexes under action filtration, so that we are able to take the limit (homotopy direct limit) to obtain the desired cochain quasi-isomorphism on the whole wrapped Floer complex:

(6.5) R1:CW(0,1;HM,N,JM,N)CW(0,1;K,J).R^{1}:CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N},J_{M,N})\to CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1};K,J).

See the last section of [Gao17] for more details for the construction of this map.

Remark 6.7.

The way we define the action-restriction map is not via certain parametrized moduli space, but this map can alternatively be interpreted as the continuation map induced by a monotone homotopy from the split Hamiltonian to the admissible Hamiltonian, on the truncated wrapped Floer cochain complexes.

In the construction of the map Rb1R^{1}_{b} above, there are several parameters involved in the procedure of modifying the Hamiltonian (e.g. the size of the compact set, the precise behavior of the Hamiltonian KbK_{b}, etc.), such that we can obtain the desired estimate on the action of the extra chords. We call a choice of these parameters, or more essentially the collection of geometric data these parameters determine, an action-restriction datum. A precise definition is given below.

Definition 6.8.

An action-restriction datum for the strip Z=×[0,1]Z=\mathbb{R}\times[0,1] and the pair (0,1)(\mathcal{L}_{0},\mathcal{L}_{1}) of Lagrangian submanifolds of M×NM\times N consists of the following data:

  1. (i)

    a truncation (b,a](-b,a] of the wrapped Floer cochain space CW(0,1;HM,N)CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N}) using the action filtration;

  2. (ii)

    a large number AA such that all HM,NH_{M,N}-chords from 0\mathcal{L}_{0} to 1\mathcal{L}_{1} of action in (b,a](-b,a] are contained in the compact set {rA}\{r\leq A\}, and which satisfies an additional property to specified in the next condition;

  3. (iii)

    an admissible Hamiltonian function KbK_{b}, which is of split type in the compact set {rA}\{r\leq A\}, and is quadratic outside a small neighborhood of a larger compact set {rcA}\{r\leq cA\}; the number AA should be large enough such that all the extra chords of KbK_{b} compared to HM,NH_{M,N} have sufficiently positive action;

  4. (iv)

    an admissible almost complex structure JbJ_{b}, which agrees with the product almost complex structure JM,NJ_{M,N} in the compact set {rA}\{r\leq A\}, and is of contact type outside a neighborhood of the compact set {rcA}\{r\geq cA\};

  5. (v)

    a homotopy HsH_{s} between HM,NH_{M,N} and KbK_{b}, parametrized by ss\in\mathbb{R}, which has support within the compact set {rcA}\{r\leq cA\}, and constantly equals HM,NH_{M,N} for ss sufficiently negative and KbK_{b} for ss sufficiently positive;

  6. (vi)

    a deformation JsJ_{s} from JM,NJ_{M,N} to JbJ_{b}, parametrized by ss\in\mathbb{R}, which has support within the compact set {rcA}\{r\leq cA\}, and constantly equals HM,NH_{M,N} for ss sufficiently negative and JbJ_{b} for ss sufficiently positive.

The existence of such action-restriction data was proved in [Gao17], which was in turn used to prove:

Proposition 6.9.

The map (6.1) is a cochain isomorphism. Therefore, the maps (6.5) and (6.3) are cochain homotopy equivalence.

To define a higher-order analogue of the action restriction map (6.3) so that we can extend it to an AA_{\infty}-functor, we need an analogue of action-restriction datum for disks with more punctures.

Definition 6.10.

An action-restriction datum for a disk SS with (k+1)(k+1)-boundary punctures and a (k+1)(k+1)-tuple of Lagrangian submanifolds (0,,k)(\mathcal{L}_{0},\cdots,\mathcal{L}_{k}) consists of the following data:

  1. (i)

    a collection of positive integers wi,i=0,,kw^{i},i=0,\cdots,k, called weights;

  2. (ii)

    a smooth function ψS:S[1,cA]\psi_{S}:\partial S\to[1,cA] parametrizing moving Lagrangian boundary conditions ϕM×NψS(z)i\phi_{M\times N}^{\psi_{S}(z)}\mathcal{L}_{i}, when zz lies between ziz_{i} and zi+1z_{i+1}, which is equal to wiw_{i} near ziz_{i};

  3. (iii)

    truncations (bi,ai](-b_{i},a_{i}] of the action filtration on CW(i1,i;HM,N)CW^{*}(\mathcal{L}_{i-1},\mathcal{L}_{i};H_{M,N}), for i=1,,ki=1,\cdots,k, and a truncation (b0,a0](-b_{0},a_{0}] of CW(0,k;HM,N)CW^{*}(\mathcal{L}_{0},\mathcal{L}_{k};H_{M,N}), subject to the following condition:

    (6.6) i=1kbib0, and i=1kaia0;\sum_{i=1}^{k}b_{i}\leq b_{0},\text{ and }\sum_{i=1}^{k}a_{i}\leq a_{0};
  4. (iv)

    a sub-closed one-form αS\alpha_{S} on SS, which vanishes along the boundary S\partial S, and agrees with widtw^{i}dt over the ii-th strip-like end, such that the differential dαSd\alpha_{S} also vanishes in a small neighborhood of the boundary;

  5. (v)

    a large number AiA_{i} depending on bib_{i}, such that all HM,NH_{M,N}-chords from i1\mathcal{L}_{i-1} to i\mathcal{L}_{i} (i=1,,ki=1,\cdots,k) and those from 0\mathcal{L}_{0} to k\mathcal{L}_{k} within the action ranges (bi,ai](-b_{i},a_{i}] and (b0,a0](-b_{0},a_{0}] are contained in the compact set {rAi}\{r\leq A_{i}\}, where rr is the radial coordinate on the cylindrical end Σ×[1,+)\Sigma\times[1,+\infty);

  6. (vi)

    a modified Hamiltonian KbiK_{b_{i}} which agrees with the split Hamiltonian in the compact set {rAi}\{r\leq A_{i}\}, and is quadratic in rr outside {rcAi}\{r\leq cA_{i}\}. The choice of AiA_{i} above should satisfy the additional condition that the extra KbiK_{b_{i}}-chords in the region {rcAi}\{r\leq cA_{i}\} from i\mathcal{L}_{i} to i+1\mathcal{L}_{i+1} (respectively from 0\mathcal{L}_{0} to k\mathcal{L}_{k} all have sufficiently positive action, much bigger than aia_{i} (respectively a0a_{0});

  7. (vii)

    an SS-parametrized family of admissible Hamiltonians KS,bK_{S,\vec{b}}, such that over the ii-th strip-like end (i=1,,ki=1,\cdots,k), KS,bK_{S,\vec{b}} agrees with the split Hamiltonian HM,NH_{M,N} rescaled by weight wiw^{i}, and over the 0-th strip-like end, KS,bK_{S,\vec{b}} agrees with the modified Hamiltonian Kb0K_{b_{0}}, rescaled by weight w0w^{0};

  8. (viii)

    a modified almost complex structure JbiJ_{b_{i}}, which is product-type in {rAi}\{r\leq A_{i}\} and is generic of contact type outside {rcAi}\{r\leq cA_{i}\};

  9. (ix)

    an SS-parametrized family of admissible almost complex structures JS,bJ_{S,\vec{b}}, such that over the ii-th strip-like end (i=1,ki=1,\cdots k), it agrees with the product almost complex structure JM,NJ_{M,N}, and over the 0-th strip-like end, it agrees with the modified almost complex structure Jb0J_{b_{0}}.

In particular, when restricting to each strip-like end, the Hamiltonian and the almost complex structure should be independent of the ss-coordinate, at least for sufficiently large |s||s|. Here ss is the coordinate on ±\mathbb{R}_{\pm}.

Later, we shall slightly modify the definition of action-restriction data, as the abstract moduli spaces underlying inhomogeneous pseudoholomorphic curves used to define AA_{\infty}-functors consist of not only punctured disks but also an additional non-negative number associated to each punctured disk (see the next subsection). All the ingredients of an action-restriction datum will be the same, though.

Remark 6.11.

Although we need only to modify the Hamiltonian near the 0-th strip-like end, we include all the relevant information near other strip-like ends because it will be convenient to extend the action-restriction data to nodal disks and state the gluing process for action-restriction data.

With these definitions, one of the main results of [Gao17] can be interpreted as having established the existence, i.e. possibility of making a choice, of action-restriction data for the strip and any pair of admissible Lagrangian submanifolds, as well as those for the triangle and any triple of admissible Lagrangian submanifolds. As is expected, the existence in fact holds for any polygon and any tuple of admissible Lagrangian submanifolds, whose construction follows the same pattern as that in [Gao17].

6.4. Choosing action-restriction data for all curves

The underlying operad controlling AA_{\infty}-functors is commonly known as multiplihedra, introduced by Stasheff. Before constructing the action-restriction functor, we need a model for multiplihedra whose elements are the underlying domains for various inhomogeneous pseudoholomorphic maps used to define the action-restriction functor which relates the two versions of wrapped Fukaya categories of the product manifold. There are many equivalent models for the multiplihedra, among which the version used in [Syl15] to construct continuation functors of wrapped Fukaya categories seems mostly adaptable to our setup, because we have always used a single Hamiltonian to define the wrapped Fukaya category, instead of a cofinal family of linear Hamiltonians.

The multiplihedra are constructed as compactifications of moduli spaces 𝒩k+1\mathcal{N}_{k+1} of punctured disks equipped with a weight. We set 𝒩1+1\mathcal{N}_{1+1} to be a single-point set, which consists of the strip ZZ equiped with a positive strip-like end at ++\infty and a negative strip-like end at -\infty. For k2k\geq 2, define 𝒩k+1=k+1×+\mathcal{N}_{k+1}=\mathcal{M}_{k+1}\times\mathbb{R}_{+}.

We are going to describe a compactification 𝒩¯k+1\bar{\mathcal{N}}_{k+1} of 𝒩k+1\mathcal{N}_{k+1}, which serves as a model for the multiplihedra. These are constructed inductively in kk. Set 𝒩¯2=𝒩2\bar{\mathcal{N}}_{2}=\mathcal{N}_{2}. Suppose we have constructed 𝒩¯l+1\bar{\mathcal{N}}_{l+1} for l<kl<k, which has boundary and generalized corners being products of associahedra and multiplihedra of lower dimensions, and that for each element (S,w)𝒩¯l+1(S,w)\in\bar{\mathcal{N}}_{l+1} we have chosen strip-like ends near the punctures in a consistent way. Then we want to construct 𝒩¯k+1\bar{\mathcal{N}}_{k+1} as well as choose strip-like ends for every element therein, such that the following two conditions are satisfied:

  1. (i)

    The boundary and generalized corners of 𝒩¯k+1\bar{\mathcal{N}}_{k+1} are products of i+1\mathcal{M}_{i+1}’s and 𝒩j+1\mathcal{N}_{j+1}’s for i,j<ki,j<k.

  2. (ii)

    If an element of 𝒩¯k+1\bar{\mathcal{N}}_{k+1} lies in the boundary strata, then the strip-like ends we choose for it should agree with the ones chosen for the irreducible components of Σ\Sigma regarded as elements in i+1\mathcal{M}_{i+1}’s and 𝒩j+1\mathcal{N}_{j+1}’s for i,j<ki,j<k.

A more accurate description of the boundary strata is given as follows. Some boundary strata appear as ww goes to infinity, which have the form

(6.7) l+11×i=1l𝒩mi+1,\mathcal{M}_{l+1}^{1}\times\prod_{i=1}^{l}\mathcal{N}_{m_{i}+1},

where i=1lmi=d\sum_{i=1}^{l}m_{i}=d. Here l+11\mathcal{M}_{l+1}^{1} means a copy of l+1\mathcal{M}_{l+1} arising at w=w=\infty. Other boundary strata appear in the compactification of k+1\mathcal{M}_{k+1}, one copy for w=0w=0 and one for w=w=\infty. These products arise as boundary of the compactification via the gluing maps. For instance, a boundary chart in the case ww\to\infty is of the form

(6.8) U×i=1lUi×[0,a),U\times\prod_{i=1}^{l}U_{i}\times[0,a),

where Ul+11U\subset\mathcal{M}_{l+1}^{1}, and Ui𝒩mi+1U_{i}\subset\mathcal{N}_{m_{i}+1}. The last factor [0,a)[0,a) is the gluing parameter, where 0 corresponds to degenerate curves, i.e. tuples of curves in U×i=1lUiU\times\prod_{i=1}^{l}U_{i}. Given (Sl+1,(Sm1+1,w1),,(Sml+1,wl))U×i=1lUi(S_{l+1},(S_{m_{1}+1},w_{1}),\cdots,(S_{m_{l}+1},w_{l}))\in U\times\prod_{i=1}^{l}U_{i}, as well as a nonzero gluing parameter ρ(0,a)\rho\in(0,a), we would like to produce a pair (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1} by gluing the surfaces together. For every ii, glue the negative strip-like end of Smi+1S_{m_{i}+1} to the ii-th positive strip-like end ϵi\epsilon_{i} of Sl+1S_{l+1} simultaneously, with length li=e1ρwiwi(Sl+1)l_{i}=e^{\frac{1}{\rho}}-w_{i}-w_{i}(S_{l+1}), where wi(Sl+1)w_{i}(S_{l+1}) is the width of the ii-th positive strip-like end ϵi\epsilon_{i} of Sl+1S_{l+1}. This procedure is uniform for all elements in U×i=1lUiU\times\prod_{i=1}^{l}U_{i}, and provides gluing maps

(6.9) 1:l+11×i=1l𝒩mi+1×[0,1)𝒩¯k+1.\sharp_{1}:\mathcal{M}_{l+1}^{1}\times\prod_{i=1}^{l}\mathcal{N}_{m_{i}+1}\times[0,1)\to\bar{\mathcal{N}}_{k+1}.

The other kind of boundary stratum has the form

(6.10) 𝒩m+2×km+1,0mk2\mathcal{N}_{m+2}\times\mathcal{M}_{k-m+1},0\leq m\leq k-2

which also comes with gluing maps

(6.11) 2:𝒩m+2×km+1×[0,1)𝒩¯k+1.\sharp_{2}:\mathcal{N}_{m+2}\times\mathcal{M}_{k-m+1}\times[0,1)\to\bar{\mathcal{N}}_{k+1}.

The third type of boundary strata are obtained by compactifying all the k+1\mathcal{M}_{k+1} components. These k+1\mathcal{M}_{k+1} components come into the picture either in the first two type codimension-one boundary strata, as w0w\to 0 or as w+w\to+\infty, or in the place where ww remains finite as the compactification of k+1×{w}𝒮k+1=k+1×+\mathcal{M}_{k+1}\times\{w\}\subset\mathcal{S}_{k+1}=\mathcal{M}_{k+1}\times\mathbb{R}_{+}. In the latter case, consider a boundary stratum σ𝒩¯k+1\sigma\subset\partial\bar{\mathcal{N}}_{k+1} whose elements are modeled over some rooted tree TT with labeled leaves. A boundary chart for σ\sigma takes the form

(6.12) U×interior vertices vUv×interior edges e[0,ae),U\times\prod_{\text{interior vertices }v}U_{v}\times\prod_{\text{interior edges }e}[0,a_{e}),

where UU is a small open subset of 𝒩l+1\mathcal{N}_{l+1} where ll is the valency of the root, while others UvU_{v} are small open subsets of mv+1\mathcal{M}_{m_{v}+1}. When all the gluing parameters in interior edges e[0,ae)\prod_{\text{interior edges }e}[0,a_{e}) are nonzero, the above chart (6.12) is identified with a subset of 𝒩k+1\mathcal{N}_{k+1} via the gluing construction in the same way as that for ¯k+1\bar{\mathcal{M}}_{k+1}, with the extra +\mathbb{R}_{+} factor in 𝒩m+1U\mathcal{N}_{m+1}\supset U mapping to the +\mathbb{R}_{+} factor in 𝒩k+1\mathcal{N}_{k+1} by identity.

Let us slightly modify the definition of an action-restriction datum so that it is more adapted to the construction of AA_{\infty}-functors. An action-restriction datum is now chosen for elements in 𝒩k+1\mathcal{N}_{k+1}, with a potential extension to the compactification 𝒩¯k+1\bar{\mathcal{N}}_{k+1}.

Definition 6.12.

Let (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1}, with given Lagrangian labels. An action-restriction datum for (S,w)(S,w) and the Lagrangian labels is simply an action-restriction datum for SS as in Definition 6.10, which depends on w+w\in\mathbb{R}_{+}.

Although an action-restriction datum for (S,w)(S,w) is the same as one for SS, the dependence on ww matters when we make a choice of an action-restriction datum for every element in 𝒩k+1\mathcal{N}_{k+1}. In other words, it matters when we consider families of action-restriction data.

Definition 6.13.

A universal choice of action-restriction data over 𝒩k+1\mathcal{N}_{k+1} is a choice of an action-restriction datum for each (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1}, which depends smoothly on (S,w)(S,w) (note that 𝒩k+1=k+1×+\mathcal{N}_{k+1}=\mathcal{M}_{k+1}\times\mathbb{R}_{+} has a natural structure of smooth manifold).

In the next subsection, we shall discuss how the action-restriction data are extended to the compactification 𝒩¯k+1\bar{\mathcal{N}}_{k+1}.

6.5. Making choices of action-restriction data consistently

According to the previous summary of the results in [Gao17], the construction of a cochain homotopy equivalence of the two versions of wrapped Floer complexes works well for any pair among a fixed finite collection of Lagrangian submanifolds, as that only requires a choice of an action-restriction datum for the strip and that particular pair of Lagrangian submanifolds in consideration. Extending this to an AA_{\infty}-functor on the categorical level requires us to choose action-restriction data for many Lagrangian submanifolds.

The following observation is useful in arranging various action-restriction data: it is the parameters aa’s, bb’s, AA’s that are crucial in the definition of an action-restriction datum, and they essentially depend on the initially given split Hamiltonian, product almost complex structure, and Lagrangian submanifolds only. Moreover, we can choose the same parameters for all pairs such that the action-restriction maps are defined, which is possible because of the following three reasons:

  1. (i)

    we only need to fix aa’s not too small so that the truncated wrapped Floer complexes CW(b,a](i,j;HM,N,JM,N)CW^{*}_{(-b,a]}(\mathcal{L}_{i},\mathcal{L}_{j};H_{M,N},J_{M,N}) include all Hamiltonian chords from i\mathcal{L}_{i} to j\mathcal{L}_{j} that are contained in the interior part of M×NM\times N;

  2. (ii)

    we have to consider all possible bb’s satisfying the condition (6.6), thus these are not matters of choices;

  3. (iii)

    we only have to choose the numbers AiA_{i}’s, which determines the size of the compact set inside which the split Hamiltonian HM,NH_{M,N} agrees with the admissible one KbK_{b}, and that of the compact set outside which KbK_{b} is quadratic, to be sufficiently large in order for the action-restriction map to be defined.

However, it is practically difficult to directly construct the action-restriction functor on the whole wrapped Fukaya category. The reason is as follows. To define higher order maps and verify the AA_{\infty}-equations, we need to choose action-restriction data for all punctured disks and all Lagrangian submanifolds in a consistent way. But there are infinitely many Lagrangian submanifolds in the category, making it unlikely possible to arrange the choices of action-restriction data for all Lagrangian submanifolds simultaneously. Keeping track of all these infinitely many choices at one time might be a painful job.

To overcome this difficulty, we take the following maneuver. First, construct the desired AA_{\infty}-functors on a chain of full AA_{\infty}-subcategories of the wrapped Fukaya category, each of which consists of finitely many objects. We want the union of these subcategories to be the whole wrapped Fukaya category, which is possible because we have assumed that the wrapped Fukaya category is made up of a countable collection of objects. Second, prove that these AA_{\infty}-functors are compatible with each other under natural inclusions of these subcategories in the chain. Third, take the limit in an appropriate sense to obtain the desired AA_{\infty}-quasi-equivalence.

Let us start from the first step: considering the full AA_{\infty}-subcategory 𝒲ds\mathcal{W}_{d}^{s} (respectively 𝒲d\mathcal{W}_{d}) of 𝒲s(M×N)\mathcal{W}^{s}(M\times N) (respectively 𝒲(M×N)\mathcal{W}(M\times N)) consisting of dd Lagrangian submanifolds 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} from our collection. We may construct an AA_{\infty}-functor

(6.13) Rd:𝒲ds𝒲dR_{d}:\mathcal{W}_{d}^{s}\to\mathcal{W}_{d}

which is a quasi-isomorphism. As mentioned before, to verify the AA_{\infty}-equations, we need to arrange the action-restriction data in a consistent way.

To explain the precise meaning of this consistency, we first describe the way how the action-restriction data can be glued together when we glue the underlying curves. As in the case of Floer data for punctured disks, we introduce the concept of conformal equivalence for action-restriction data. To state this, we introduce some notations. If HH is a Hamiltonian, denote by HCH_{C} the time-logC\log C rescaling of HH,

HC=1CHψC,H_{C}=\frac{1}{C}H\circ\psi^{C},

where ψC\psi^{C} is the time-logC\log C Liouville flow. Similarly, if JJ is an almost complex structure, denote by JCJ_{C} the time-logC\log C rescaling of JJ,

JC=JψC.J_{C}=J\circ\psi^{C}.
Definition 6.14.

Two action-restriction data are said to be conformally equivalent, if there exist constants C,W>0C,W>0 such that:

  1. (i)

    the weights differ by scalar multiplication by WW, i.e. wi=Wwiw^{i}=Ww^{\prime i};

  2. (ii)

    the action filtrations agree, i.e. ai=ai,bi=bia_{i}=a^{\prime}_{i},b_{i}=b^{\prime}_{i};

  3. (iii)

    the Hamiltonians agree up to rescaling,

    (6.14) KS,b=1W(KS,b)C;K_{S,\vec{b}}=\frac{1}{W}(K^{\prime}_{S,\vec{b}})_{C};
  4. (iv)

    the almost complex structures agree up to rescaling,

    (6.15) JS,b=(JS,b)C.J_{S,\vec{b}}=(J^{\prime}_{S,\vec{b}})_{C}.

Fix a finite collection 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} in consideration. For example, suppose that we are given Sl+11S\in\mathcal{M}_{l+1}^{1} and (Si,wi)𝒩mi+1,i=1,,l(S_{i},w_{i})\in\mathcal{N}_{m_{i}+1},i=1,\cdots,l, all equipped with appropriate Lagrangian labels so that they match on the boundary components of SS and SiS_{i} that are to be glued together, in the way that the ii-th strip-like end of SS is glued with the 0-th strip-like end of SiS_{i}. Suppose that we have chosen an action-restriction datum for each (Si,wi)(S_{i},w_{i}), as well as a Floer datum for SS (which is involved in the definition of the AA_{\infty}-structure on 𝒲(M×N)\mathcal{W}(M\times N)).

Definition 6.15.

The action-restriction data on (Si,wi)(S_{i},w_{i}) and the Floer datum on SS are said to be conformally consistent, if there exist positive constants

C,W,C1,,Cl,W1,,Wl>0,C,W,C_{1},\cdots,C_{l},W_{1},\cdots,W_{l}>0,

such that the following conditions hold:

  1. (i)

    Let wiw^{i} be the weight on the ii-th strip-like end of SS, and wi0w_{i}^{0} the weight on the 0-th strip-like end of SiS_{i}. Then

    (6.16) Cwi=Ciwi0.Cw^{i}=C_{i}w_{i}^{0}.
  2. (ii)

    Let KS,iK_{S,i} be the asymptotic Hamiltonian of the SS-dependent family of Hamiltonians KSK_{S} over the ii-th strip-like end of SS, and KSi,bi0K_{S_{i},b_{i}^{0}} the asymptotic Hamiltonian of the SiS_{i}-dependent family of Hamiltonians KSi,biK_{S_{i},\vec{b}_{i}} over the 0-th strip-like end of SiS_{i}. Then

    (6.17) 1W(KS,i)C=1Wj(KSi,bi0)Ci.\frac{1}{W}(K_{S,i})_{C}=\frac{1}{W_{j}}(K_{S_{i},b_{i}^{0}})_{C_{i}}.
  3. (iii)

    Let JS,iJ_{S,i} be the asymptotic almost complex structure of the SS-dependent family of almost complex structures JSJ_{S} over the ii-th strip-like end of SS, and JSi,bi0J_{S_{i},b_{i}^{0}} the asymptotic almost complex structure of the SiS_{i}-dependent family of almost complex structures JSi,biJ_{S_{i},\vec{b}_{i}} over the 0-th strip-like end of SiS_{i}. Then

    (6.18) (JS,i)C=(JSi,bi0)Ci.(J_{S,i})_{C}=(J_{S_{i},b_{i}^{0}})_{C_{i}}.

Given every gluing parameter ρ(0,a)\rho\in(0,a), we may glue SS and (Si,wi)(S_{i},w_{i}) together to get a smooth punctured disk (Sρ,wρ)(S_{\rho},w_{\rho}). The glued action-restriction datum for (Sρ,wρ)(S_{\rho},w_{\rho}) is defined as follows. The weights are simply those CiwijC_{i}w_{i}^{j}’s, and Cw0Cw^{0}, removing those weights wi,wi0w^{i},w_{i}^{0} for the strip-like ends that are glued together. As for the truncations, over the strip-like ends of SρS_{\rho} that are obtained from SiS_{i}’s, we simply take the existing truncations (bij,aij](-b_{i}^{j},a_{i}^{j}]. But we have not specified a truncation for the 0-th strip-like end of SρS_{\rho} which comes from SS, as the Floer datum on SS does not contain such information. Nonetheless, the previous condition on Hamiltonians give natural truncation over that end. We may simply take b0=i=1lb0ib_{0}=\sum_{i=1}^{l}b^{i}_{0}, and take a0a_{0} not less than i=1a0i\sum_{i=1}a^{i}_{0} such that the truncated wrapped Floer complex CW(b0,a]CW^{*}_{(-b_{0},a]} is independent of aa when aa0a\geq a_{0}.

Because of the above conditions on families of Hamiltonians and almost complex structures over SS and SiS_{i}’s that they agree up to rescaling over the strip-like ends which are to be glued together, we may rescale these families KS,KSi,biK_{S},K_{S_{i},\vec{b}_{i}} and JS,JSi,biJ_{S},J_{S_{i},\vec{b}_{i}}, and take the union of the rescaled families on the glued surface Σρ\Sigma_{\rho} to obtain the families KΣρ,JΣρK_{\Sigma_{\rho}},J_{\Sigma_{\rho}}, which have all the desired properties.

There are more complicated gluings, which happen in higher-codimensional strata. But the corresponding gluing process for action-restriction data is the same.

Now let us formalize all the above ideas in the following definition.

Definition 6.16.

A universal and conformally consistent choice of action-restriction data is a choice of an action-restriction datum for every k1k\geq 1 and for every (representative of) element 𝒩¯k+1\bar{\mathcal{N}}_{k+1} and every (k+1)(k+1)-tuple of Lagrangian submanifolds (0,,k)(\mathcal{L}_{0},\cdots,\mathcal{L}_{k}) of M×NM\times N, which varies smoothly on 𝒩k+1\mathcal{N}_{k+1}, and satisfies the following conditions:

  1. (i)

    For an element (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1} that is sufficiently close to the boundary strata 𝒩¯k+1\partial\bar{\mathcal{N}}_{k+1}, then the choice of action-restriction datum for (S,w)(S,w) is conformally equivalent to the action-restriction datum induced by gluing of action-restriction data and Floer data;

  2. (ii)

    The following chart

    (6.19) U×interior vertices vUv×interior edges e(0,ae),U\times\prod_{\text{interior vertices }v}U_{v}\times\prod_{\text{interior edges }e}(0,a_{e}),

    for a deleted neighborhood of a boundary stratum σ𝒩¯k+1\sigma\subset\partial\bar{\mathcal{N}}_{k+1} which has a chart (6.12), the restriction of the action-restriction data to the main component (S,w)𝒮l+1(S,w)\in\mathcal{S}_{l+1} induces a family of action-restriction data for (S,w)(S,w) parametrized by

    U×e adjacent to the root(0,ae)×E.U\times\prod_{e\text{ adjacent to the root}}(0,a_{e})\times E.

    We require this family extends smoothly to

    U×e adjacent to the root[0,ae)×E,U\times\prod_{e\text{ adjacent to the root}}[0,a_{e})\times E,

    and that it agrees on U×e adjacent to the root{0}×EU\times\prod_{e\text{ adjacent to the root}}\{0\}\times E with the family of action-restriction data that was chosen for 𝒩l+1\mathcal{N}_{l+1}, up to a family of conformal rescalings.

Remark 6.17.

The ”universal and conformally consistent” condition is stated in a slightly more complicated way than that for Floer data as in [Abo10], because the compactification 𝒩¯k+1\bar{\mathcal{N}}_{k+1} does not have a structure of smooth manifold with corners. In fact it is a more generalized smooth space. This is the reason why we state the consistency condition in (ii) in such a complicated way, which in a concise language means the action-restriction data vary smoothly in each smooth chart of 𝒩¯k+1\bar{\mathcal{N}}_{k+1}.

Lemma 6.18.

Fix a finite collection of Lagrangian submanifolds 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} of M×NM\times N. Then universal and conformally consistent choices of action-restriction data exist, with Lagrangian labels chosen from this collection.

Proof.

The proof is an inductive argument based upon the inductive structure of the construction of the multiplihedra 𝒩¯k+1\bar{\mathcal{N}}_{k+1}, as well as the fact that the space of action-restriction data for any single element (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1} is contractible.

First, fix a choice of an action-restriction datum over the strip for every pair (i,j)(\mathcal{L}_{i},\mathcal{L}_{j}) of Lagrangian submanifolds in this collection, which is used to define the action-restriction map

(6.20) R1:CW(i,j;HM,N,JM,N)CW(i,j;K,J).R^{1}:CW^{*}(\mathcal{L}_{i},\mathcal{L}_{j};H_{M,N},J_{M,N})\to CW^{*}(\mathcal{L}_{i},\mathcal{L}_{j};K,J).

Such a choice is possible by the results of [Gao17]. Then we consider the 33-pointed disk S3S_{3}, which is the unique element in 3\mathcal{M}_{3}. For each w(0,+)w\in(0,+\infty), we consider the pair (S3,w)𝒩3(S_{3},w)\in\mathcal{N}_{3}. Let ji,i=0,1,2\mathcal{L}_{j_{i}},i=0,1,2 be the Lagrangian labels for the boundary condition. We choose an action-restriction datum for S3S_{3} and these Lagrangian labels, such that over the 0-th strip-like end, it agrees with the Floer datum for (j0,j2)(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{2}}) in 𝒲d\mathcal{W}_{d}, and over the ii-th strip-like ends (i=1,2i=1,2), it agrees with the Floer datum for (ji1,ji)(\mathcal{L}_{j_{i-1}},\mathcal{L}_{j_{i}}) in 𝒲ds\mathcal{W}^{s}_{d}. The existence of such an action-restriction datum is also proved in [Gao17]. Moreover, we can choose action-restriction data which vary smoothly with respect to ww, such that as w0w\to 0, the action-restriction datum converges to a Floer datum for 𝒲ds\mathcal{W}^{s}_{d}, and as ww\to\infty, the action-restriction datum converges to a Floer datum for 𝒲d\mathcal{W}_{d}.

To proceed, note that the gluing maps (6.9) and (6.11) are the key for us to run an inductive argument. The initial step is to obtain action-restriction data for all elements in 𝒩¯4\bar{\mathcal{N}}_{4} and all Lagrangian labels in the collection. Note that the choices of action-restriction data for S3S_{3} and all possible triples of Lagrangian labels determine choices of action-restriction data for elements in a neighborhood of the boundary stratum of 𝒩¯4\bar{\mathcal{N}}_{4} for the quadruple of Lagrangian labels ji,i=0,1,2,3\mathcal{L}_{j_{i}},i=0,1,2,3. We then extend the choices of action-restriction data over the whole 𝒩¯4\bar{\mathcal{N}}_{4}. This is possible because there are no obstructions: 𝒩¯4\bar{\mathcal{N}}_{4} is contractible, and the spaces of admissible Hamiltonians and of almost complex structures are contractible.

For the inductive step, suppose that we have made consistent choices of action-restriction data for elements in 𝒩¯mi+1,𝒩¯m+2\bar{\mathcal{N}}_{m_{i}+1},\bar{\mathcal{N}}_{m+2} as well as Floer data for elements in ¯l+1,km+1\bar{\mathcal{M}}_{l+1},\mathcal{M}_{k-m+1}, such that the Hamiltonians, almost complex structures, etc. agree over the strip-like ends that are to be glued together. We then use the gluing maps (6.9) and (6.11) to obtain action-restriction data for elements in a neighborhood of the boundary strata 𝒩¯k+1\partial\bar{\mathcal{N}}_{k+1} in ¯k+1\bar{\mathcal{M}}_{k+1}, in the way we described before Definitions 6.8 and 6.10. Since 𝒩¯k+1\bar{\mathcal{N}}_{k+1} is contractible and the spaces of admissible Hamiltonians and almost complex structures are contractible, we may extend the choices for all elements therein, without changing those for elements in the boundary 𝒩¯k+1\partial\bar{\mathcal{N}}_{k+1}. This completes the inductive step and therefore finishes the proof.

Remark 6.19.

Unlike the case of compact Lagrangian submanifolds, in general CW(0,1)CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1}) is different from CW(1,0)CW^{*}(\mathcal{L}_{1},\mathcal{L}_{0}) even on the level of cohomology, and in general there is no Poincare duality between these two complexes. Thus there is no need to worry about compatibility for the choices of action-restriction datum for (0,1)(\mathcal{L}_{0},\mathcal{L}_{1}) and that for (1,0)(\mathcal{L}_{1},\mathcal{L}_{0}) - these are simply two different, independent choices.

6.6. The action-restriction functor: partial definition

Suppose we have made a universal and conformally consistent choice of action-restriction data for all k1k\geq 1 and all 𝒩¯k+1\bar{\mathcal{N}}_{k+1}, with the Lagrangian labels for the boundary components of punctured disks chosen from a fixed finite collection of Lagrangian submanifolds 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} of M×NM\times N. Then we shall construct an AA_{\infty}-functor

(6.21) Rd:𝒲ds𝒲d,R_{d}:\mathcal{W}^{s}_{d}\to\mathcal{W}_{d},

with the following properties. On the level of objects, it acts as the identity. The first order map Rd1R_{d}^{1} is the action-restriction map (6.3), for each pair of Lagrangian submanifolds from the collection 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d}.

As regard for higher order structure maps, consider (k+1)(k+1)-tuple of Lagrangian submanifolds j0,,jk\mathcal{L}_{j_{0}},\cdots,\mathcal{L}_{j_{k}}, where ji{1,,d}j_{i}\in\{1,\cdots,d\}. Suppose we have chosen a universal and conformally consistent choice of action-restriction data, where the Lagrangian labels are limited to the cyclically-ordered tuple (j0,,jk)(\mathcal{L}_{j_{0}},\cdots,\mathcal{L}_{j_{k}}). For any (S,w)𝒩k+1(S,w)\in\mathcal{N}_{k+1}, the action-restriction datum from our choice sets up a moduli problem. Varying (S,w)(S,w) in the moduli space 𝒩k+1\mathcal{N}_{k+1} gives rise to a family over 𝒩k+1\mathcal{N}_{k+1}, which is a parametrized moduli space 𝒩k+1(x~0;x1,,xk)\mathcal{N}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}), provided appropriate asymptotic convergence conditions are given:

xiCW(bi,ai](ji1,ji;HM,N,JM,N),x_{i}\in CW^{*}_{(-b_{i},a_{i}]}(\mathcal{L}_{j_{i-1}},\mathcal{L}_{j_{i}};H_{M,N},J_{M,N}),

and

x~0CW(b0,a0](j0,jk;Kb,Jb).\tilde{x}_{0}\in CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K_{\vec{b}},J_{\vec{b}}).

The moduli space 𝒩k+1(x~0;x1,,xk)\mathcal{N}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) has a compactification 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}). The additional elements are either (equivalence classes of) broken inhomogeneous pseudoholomorphic maps from domains being elements in 𝒩¯k+1\bar{\mathcal{N}}_{k+1}, or broken inhomogeneous pseudoholomorphic maps which break out inhomogeneous pseudoholomorphic strips.

Combined with the results in the last section of [Gao17], we can use standard transversality argument to prove:

Proposition 6.20.

For generic choice of action-restriction data, the moduli space 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) satisfies the following properties:

  1. (i)

    If the virtual dimension is zero, 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) is a zero-dimensional compact smooth manifold, hence consists of finitely many points;

  2. (ii)

    If the virtual dimension is one, 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) is a one-dimensional compact topological manifold, hence is a disjoint union of finitely many circles and intervals.

By counting rigid solutions in the moduli space 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}), we define the following multilinear maps of degree 1k1-k

(6.22) R¯d,bk:CW(bk,ak](jk1,jk;HM,N,JM,N)CW(b1,a1](j0,j1;HM,N,JM,N)CW(b0,a0](j0,jk;Kb0,Jb0).\begin{split}\bar{R}_{d,\vec{b}}^{k}:&CW^{*}_{(-b_{k},a_{k}]}(\mathcal{L}_{j_{k-1}},\mathcal{L}_{j_{k}};H_{M,N},J_{M,N})\otimes\cdots\otimes CW^{*}_{(-b_{1},a_{1}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{1}};H_{M,N},J_{M,N})\\ &\to CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K_{b_{0}},J_{b_{0}}).\end{split}

Similar to the first-order map, we may compose this with a canonical cochain homotopy equivalence

hb0:CW(j0,jk;Kb0,Jb0)CW(j0,jk;K,J),h_{b_{0}}:CW^{*}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K_{b_{0}},J_{b_{0}})\to CW^{*}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J),

to obtain the following map

(6.23) Rd,bk:CW(bk,ak](jk1,jk;HM,N,JM,N)CW(b1,a1](j0,j1;HM,N,JM,N)CW(b0,a0](j0,jk;K,J).\begin{split}R_{d,\vec{b}}^{k}:&CW^{*}_{(-b_{k},a_{k}]}(\mathcal{L}_{j_{k-1}},\mathcal{L}_{j_{k}};H_{M,N},J_{M,N})\otimes\cdots\otimes CW^{*}_{(-b_{1},a_{1}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{1}};H_{M,N},J_{M,N})\\ &\to CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J).\end{split}
Remark 6.21.

An alternative interpretation of the action-restriction functor is helpful. Simply put, the action-restriction functor is a continuation functor associated to a particular kind of monotone homotopy from (HM,N,JM,N)(H_{M,N},J_{M,N}) to (K,J)(K,J). The reason we perform the construction in a slightly non-standard way as above is that it is not obviously clear that this kind of continuation functor is a homotopy equivalence, as (HM,N,JM,N)(H_{M,N},J_{M,N}) and (K,J)(K,J) behave somewhat differently at infinity.

To extend the above maps to the whole wrapped Floer complexes so that we can obtain the desired AA_{\infty}-functor (6.21), we are faced with two potential problems. First, we must ensure that these maps are compatible with natural inclusions of wrapped Floer complexes under action filtration. Second, the target wrapped Floer complex is defined with respect to a Hamiltonian KbK_{\vec{b}} and a one-parameter family of almost complex structures JbJ_{\vec{b}} which depend on the action filtration windows chosen in the action-restriction data. The solutions to these two problems will have to depend on the precise behavior of the choices of families of Hamiltonians and almost complex structures involved in the action-restriction data. We shall discuss these matters in greater detail in the next subsection.

6.7. Arranging geometric data in a compatible system

We shall now explain how to obtain the families of Hamiltonians and almost complex structures involved in the action-restriction data. In Definition (6.10), the condition (6.6) implies that we can define a new kk-th order ”multiplication” mbkm^{k}_{\vec{b}} on the truncated wrapped Floer complexes:

(6.24) mbk:CW(bk,ak](jk1,jk;Kbk,Jbk)CW(b1,a1](j0,j1;Kb1,Jb1)CW(b0,a0](J0,jk;Kb0,Jb0),\begin{split}m^{k}_{\vec{b}}&:CW^{*}_{(-b_{k},a_{k}]}(\mathcal{L}_{j_{k-1}},\mathcal{L}_{j_{k}};K_{b_{k}},J_{b_{k}})\otimes\cdots\otimes CW^{*}_{(-b_{1},a_{1}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{1}};K_{b_{1}},J_{b_{1}})\\ &\to CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{J_{0}},\mathcal{L}_{j_{k}};K_{b_{0}},J_{b_{0}}),\end{split}

whose definition is an obvious analogue of the operation mb,2b2m^{2}_{b,2b} introduced in [Gao17]. Roughly speaking, this counts rigid inhomogeneous pseudoholomorphic polygons with possibly different asymptotic data (Hamiltonians and almost complex structures) over the strip-like ends.

In [Gao17], we proved that mb,2b2m^{2}_{b,2b} is strictly compatible with the first order action-restriction maps R31R_{3}^{1} (here 33 is for a triple of Lagrangians). But the new operations mbkm^{k}_{\vec{b}} are not the AA_{\infty}-operations for the AA_{\infty}-subcategory 𝒲d\mathcal{W}_{d} of the wrapped Fukaya category 𝒲(M×N)\mathcal{W}(M\times N). In general, the maps R¯dk\bar{R}_{d}^{k} in (6.22) satisfy analogue of AA_{\infty}-equations with the AA_{\infty}-structure maps replaced by these new ones mbkm^{k}_{\vec{b}}, but not the original AA_{\infty}-structure maps. This is one of the main reasons that we need to introduce the higher order maps RdkR_{d}^{k} to adjust the failure of Rd1R_{d}^{1} from being an AA_{\infty}-functor with respect to the honest AA_{\infty}-structures.

Now let us start constructing the families of Hamiltonians and almost complex structures. The family of Hamiltonian functions is chosen as follows. We start from the families of Hamiltonians that are used to define AA_{\infty}-structures of the wrapped Fukaya categories 𝒲ds\mathcal{W}^{s}_{d} and 𝒲d\mathcal{W}_{d}. For the first one, recall from [Gao17] that for each (k+1)(k+1)-punctured disk, we may choosen the family of split Hamiltonians HSH_{S} to be ”essentially trivial”, meaning that the whole family is the rescaling by a (globally extended) time-shifting function of the single split Hamiltonian HM,NH_{M,N}, i.e.,

(6.25) HS,s=HM,NϕρS(s)ρ(s)2,H_{S,s}=\frac{H_{M,N}\circ\phi^{\rho_{S}(s)}}{\rho(s)^{2}},

where ρS:S[1,+)\rho_{S}:S\to[1,+\infty) is an extension of the time-shifting function to the whole surface SS. For the second, instead of considering a single family of Hamiltonians, we consider for each b0b_{0} the SS-family of admissible Hamiltonians HS,b0H_{S,b_{0}}, which can also be chosen to be of the form

(6.26) HS,b0,s=Kb0ϕρS(s)ρS(s)2.H_{S,b_{0},s}=\frac{K_{b_{0}}\circ\phi^{\rho_{S}(s)}}{\rho_{S}(s)^{2}}.

To find the family of Hamiltonians that is required in an action-restriction datum, our strategy is to first find a homotopy between the two Hamiltonian functions HM,NH_{M,N} and KbK_{\vec{b}}. Note that after suitable compactly-supported homotopy, Kb0K_{b_{0}} agrees with HM,NH_{M,N} inside a compact set, and differs from HM,NH_{M,N} by a product term cr1r2cr_{1}r_{2} in the region {r11,r21}\{r_{1}\geq 1,r_{2}\geq 1\}, where cc is a universal constant (roughly c=2c=2) and r1,r2r_{1},r_{2} are radial coordinates on the cylindrical ends of MM and NN respectively. We take a homotopy

Kb,t,t[0,1]K_{\vec{b},t},t\in[0,1]

from HM,NH_{M,N} to Kb0K_{b_{0}} such that in the region M×[B,+)×N×[B,+)\partial M\times[B,+\infty)\times\partial N\times[B,+\infty), it takes the form r12+r22+cχ(s)r1r2r_{1}^{2}+r_{2}^{2}+c\chi(s)r_{1}r_{2}, where χ:[0,1]\chi:\mathbb{R}\to[0,1] is a smooth increasing function which takes the value 0 for sufficiently negative s0s\ll 0, and takes the value 11 for sufficiently positive s0s\gg 0. In other regions as part of the cylindrical end Σ×[1,+)\Sigma\times[1,+\infty), we require that the homotopy is also an increasing homotopy. Then we apply the rescaling to this homotopy to obtain the desired family of Hamiltonians, using the following trick.

To get a family of Hamiltonians parametrized by the (k+1)(k+1)-punctured disk SS from the homotopy, we shall realize SS as a subset of 2\mathbb{R}^{2} to make use of χ\chi to construct the family. Embed SS into 2\mathbb{R}^{2} as a domain with boundary, such that the 0-th strip-like end is mapped to the following region

[A,+)×[12,12],[A,+\infty)\times[-\frac{1}{2},\frac{1}{2}],

for some large constant A>0A>0, and the ii-th strip-like end is mapped to the following region

(,A]×[C12,C+12](-\infty,-A]\times[C-\frac{1}{2},C+\frac{1}{2}]

where the CiC_{i}’s are some constants satisfying Ci<Ci+12C_{i}<C_{i+1}-2. Now the function χ:[0,1]\chi:\mathbb{R}\to[0,1] induces a function χS:S[0,1]\chi_{S}:S\to[0,1] by trivial extension to 2=×\mathbb{R}^{2}=\mathbb{R}\times\mathbb{R} and restriction to SS. Then we put

(6.27) KS,b,s=Kb,χS(s)ϕρS(s)ρS(s)2.K_{S,\vec{b},s}=\frac{K_{\vec{b},\chi_{S}(s)}\circ\phi^{\rho_{S}(s)}}{\rho_{S}(s)^{2}}.

This is the desired family of Hamiltonians KS,bK_{S,\vec{b}}.

The construction of the family of almost complex structures JS,bJ_{S,\vec{b}} follows the same pattern. We choose a small infinitesimal deformation YY of the almost complex structure in the space of admissible almost complex structures compatible, and add it (via the exponential map) to the product almost complex structure, such that the perturbed almost complex structure is generic, for the purpose of achieving transversality. Multiply YY by χS(s)\chi_{S}(s) and use the exponential map to get a deformation of almost complex structures from JM,NJ_{M,N} to Jb0J_{b_{0}}, so that we can apply similar trick to define the family JS,bJ_{S,\vec{b}}.

With these families of Hamiltonians and almost complex structures, we obtain consistent choices of action-restriction data for all k1k\geq 1 and for all 𝒩¯k+1\bar{\mathcal{N}}_{k+1}, with Lagrangian labels from the fixed collection 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d}. Now we can prove:

Lemma 6.22.

The multilinear maps (6.22) satisfy an analogue of AA_{\infty}-functor equations, with respect to the mkm^{k}’s for 𝒲(M)\mathcal{W}(M) and and the modified operations mbkm^{k}_{\vec{b}} for 𝒲(N)\mathcal{W}(N):

(6.28) mb01R¯bk(xk,,x1)=2lks11,,sl1s1++sl=kmbl,newl(R¯bslsl(xs1++sl,,xs1++sl1+1),,R¯bs1s1(xs1,,x1))+0sk1iR¯bs+1s+1(xk,,xi+ks+1,mks(xi+ks,,xi+1),xi,,x1).\begin{split}&m^{1}_{b_{0}}\circ\bar{R}^{k}_{\vec{b}}(x_{k},\cdots,x_{1})\\ =&\sum_{2\leq l\leq k}\sum_{\begin{subarray}{c}s_{1}\geq 1,\cdots,s_{l}\geq 1\\ s_{1}+\cdots+s_{l}=k\end{subarray}}m^{l}_{\vec{b}^{l,new}}(\bar{R}^{s_{l}}_{\vec{b}^{s_{l}}}(x_{s_{1}+\cdots+s_{l}},\cdots,x_{s_{1}+\cdots+s_{l-1}+1}),\cdots,\bar{R}^{s_{1}}_{\vec{b}^{s_{1}}}(x_{s_{1}},\cdots,x_{1}))\\ &+\sum_{0\leq s\leq k-1}\sum_{i}\bar{R}^{s+1}_{\vec{b}^{s+1}}(x_{k},\cdots,x_{i+k-s+1},m^{k-s}(x_{i+k-s},\cdots,x_{i+1}),x_{i},\cdots,x_{1}).\end{split}

Here the filtration numbers bb’s are determined by the following rule. In the first term on the right hand side of (6.28), we have bsi=(b0si,,bsisi)\vec{b}^{s_{i}}=(b^{s_{i}}_{0},\cdots,b^{s_{i}}_{s_{i}}), such that bl,new=(b0,b0s1,,b0sl)\vec{b}^{l,new}=(b_{0},b^{s_{1}}_{0},\cdots,b^{s_{l}}_{0}). On the left hand side, we have b=(b0,,bk)\vec{b}=(b_{0},\cdots,b_{k}). Moreover, we require that when deleting those b0sib^{s_{i}}_{0}’s for all ii and combining the rest together, (b0,b1s1,,bs1s1,,b1sl,,bslsl)(b_{0},b^{s_{1}}_{1},\cdots,b^{s_{1}}_{s_{1}},\cdots,b^{s_{l}}_{1},\cdots,b^{s_{l}}_{s_{l}}) should agree with (b0,b1,,bk)(b_{0},b_{1},\cdots,b_{k}) from the left hand side. In the second term, we have bs+1=(b0,,bi,bi+ks+1,,bk)\vec{b}^{s+1}=(b_{0},\cdots,b_{i},b_{i+k-s+1},\cdots,b_{k}).

Proof.

The compactification 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) has codimension one boundary strata consisting of the following four kinds of products of moduli spaces:

  1. (i)
    l+11(x~0;x~1,,x~l)×i=1l𝒩si+1(x~i;xs1++si1+1,,xs1++si),l2;\begin{split}&\mathcal{M}_{l+1}^{1}(\tilde{x}_{0};\tilde{x}_{1},\cdots,\tilde{x}_{l})\\ \times\prod_{i=1}^{l}&\mathcal{N}_{s_{i}+1}(\tilde{x}_{i};x_{s_{1}+\cdots+s_{i-1}+1},\cdots,x_{s_{1}+\cdots+s_{i}}),l\geq 2;\end{split}
  2. (ii)
    𝒩k+1(x~0;x1,,xk)×21(x~0;x~0);\mathcal{N}_{k+1}(\tilde{x}^{\prime}_{0};x_{1},\cdots,x_{k})\times\mathcal{M}_{2}^{1}(\tilde{x}_{0};\tilde{x}^{\prime}_{0});
  3. (iii)
    𝒩s+2(x~0;x1,,xi,x,xi+ks+1,,xk)×ks+10(x;xi+1,,xi+ks),sk2;\begin{split}&\mathcal{N}_{s+2}(\tilde{x}_{0};x_{1},\cdots,x_{i},x^{\prime},x_{i+k-s+1},\cdots,x_{k})\\ \times&\mathcal{M}_{k-s+1}^{0}(x^{\prime};x_{i+1},\cdots,x_{i+k-s}),s\leq k-2;\end{split}
  4. (iv)
    20(xi;xi)×𝒩k+1(x~0;x1,,xi1,xi,xi+1,,xk).\mathcal{M}_{2}^{0}(x^{\prime}_{i};x_{i})\times\mathcal{N}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{i-1},x^{\prime}_{i},x_{i+1},\cdots,x_{k}).

Here the moduli spaces l+11()\mathcal{M}_{l+1}^{1}(\cdots) with superscript 11 consist of inhomogeneous pseudoholomorphic disks defined with respect to the Floer data (Kb,Jb)(K_{\vec{b}},J_{\vec{b}}).

The first two types appear when the domains degenerate, which occur in the compactification of 𝒩k+1\mathcal{N}_{k+1}. The last two types appear when a sequence of pseudoholomorphic maps breaks off pseudoholomorphic strips at one of the strip-like ends. Strata of higher codimensions correspond to further degenerations of the domains, and breaking off more pseudoholomorphic strips. Since for our purpose only codimension one strata need to be considered, we will not spell out the details for strata of higher codimensions.

Considering various operations defined by the moduli spaces appearing in the boundary strata of 𝒩¯k+1(x~0;x1,,xk)\bar{\mathcal{N}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}), we get the desired AA_{\infty}-equations (6.28).

We remark that in the above formula (6.28) there are have two kinds of terms, as we have combined the contributions from types (i), (ii) and the types (iii), (iv).

In order to extend the maps (6.23) over the whole wrapped Floer complexes, we need to check that these maps are compatible with each other (for different values of b=(b0,,bk)\vec{b}=(b_{0},\cdots,b_{k})) with respect to the natural inclusions

κi:CW(bi,ai](ji1,ji;HM,N,JM,N)CW(bi,ai](ji1,ji;HM,N,JM,N),\kappa_{i}:CW^{*}_{(-b_{i},a_{i}]}(\mathcal{L}_{j_{i-1}},\mathcal{L}_{j_{i}};H_{M,N},J_{M,N})\to CW^{*}_{(-b^{\prime}_{i},a^{\prime}_{i}]}(\mathcal{L}_{j_{i-1}},\mathcal{L}_{j_{i}};H_{M,N},J_{M,N}),

and

κ0:CW(b0,a0](j0,jk;K,J)CW(b0,a0](j0,jk;K,J)\kappa^{\prime}_{0}:CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J)\to CW^{*}_{(-b^{\prime}_{0},a^{\prime}_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J)

of the truncated wrapped Floer complexes, whenever bibi,aiaib^{\prime}_{i}\geq b_{i},a^{\prime}_{i}\leq a_{i}. In fact, we may fix once-for-all the aa’s at the beginning for all the truncated wrapped Floer complexes, as changing these numbers do not affect the wrapped Floer complexes as soon as they are chosen so that all the interior chords are included. Thus keeping track of aa’s is unnecessary.

Let us describe in more detail the compatibility conditions. Consider the two composed maps Rd,bk(κkκ1)R_{d,\vec{b}^{\prime}}^{k}\circ(\kappa_{k}\otimes\cdots\otimes\kappa_{1}) and κ0Rd,bk\kappa_{0}\circ R_{d,\vec{b}}^{k}. If they strictly agreed, then the maps Rd,bkR_{d,\vec{b}}^{k} would be the restriction of the single map (6.23) to the truncated wrapped Floer complexes. However, by the nature of our construction, we might have chosen different families of Hamiltonians and almost complex structures when constructing the maps Rd,bkR_{d,\vec{b}}^{k} on truncated wrapped Floer complexes for different values of b\vec{b}. Thus these two compositions in general differ from each other. The compatibility condition we require should therefore be phrased that the sequence of maps Rd,bk(κkκ1)R_{d,\vec{b}^{\prime}}^{k}\circ(\kappa_{k}\otimes\cdots\otimes\kappa_{1}) is homotopic to κ0Rd,bk\kappa_{0}\circ R_{d,\vec{b}}^{k}, for bibib^{\prime}_{i}\geq b_{i}.

Proposition 6.23.

Consider the maps Rd,bkR_{d,\vec{b}}^{k} defined in (6.23). Then there are multilinear maps Tb,bkT_{\vec{b},\vec{b}^{\prime}}^{k}, forming a homotopy between the two sequences of maps

Rd,bk(κkκ1)R_{d,\vec{b}^{\prime}}^{k}\circ(\kappa_{k}\otimes\cdots\otimes\kappa_{1})

and

κ0Rd,bk,\kappa_{0}\circ R_{d,\vec{b}}^{k},

in the sense of homotopy between AA_{\infty}-functors. That is, these homotopies satisfy the following analogue of AA_{\infty}-equations:

(6.29) Rd,bk(κkκ1)(xk,,x1)κ0Rd,bk(xk,,x1)=r,is1,,sr(1)mr(κ0Rd,bsr(xk,,xksr+1),,κ0Rd,bsi+1(xs1++si+1,,xs1++si+1),Tb,bsi(xs1++si,,xs1++si1+1),Rd,bsi(κs1++si1κs1++si2+1)(xs1++si1,,xs1++si2+1),,Rd,bs1(κs1κ1))+m,l(1)Tb,bkm+1(xk,,xm+l+1,μm(xm+l,,xl+1),xl,,x1).\begin{split}&R_{d,\vec{b}^{\prime}}^{k}\circ(\kappa_{k}\otimes\cdots\otimes\kappa_{1})(x_{k},\cdots,x_{1})-\kappa_{0}\circ R_{d,\vec{b}}^{k}(x_{k},\cdots,x_{1})\\ =&\sum_{r,i}\sum_{s_{1},\cdots,s_{r}}(-1)^{*}m^{r}(\kappa_{0}\circ R_{d,\vec{b}}^{s_{r}}(x_{k},\cdots,x_{k-s_{r}+1}),\cdots,\\ &\kappa_{0}\circ R_{d,\vec{b}}^{s_{i+1}}(x_{s_{1}+\cdots+s_{i+1}},\cdots,x_{s_{1}+\cdots+s_{i}+1}),T_{\vec{b},\vec{b}^{\prime}}^{s_{i}}(x_{s_{1}+\cdots+s_{i}},\cdots,x_{s_{1}+\cdots+s_{i-1}+1}),\\ &R_{d,\vec{b}^{\prime}}^{s_{i}}\circ(\kappa_{s_{1}+\cdots+s_{i-1}}\otimes\cdots\otimes\kappa_{{}_{s_{1}+\cdots+s_{i-2}+1}})(x_{s_{1}+\cdots+s_{i-1}},\cdots,x_{s_{1}+\cdots+s_{i-2}+1}),\\ &\cdots,R_{d,\vec{b}^{\prime}}^{s_{1}}\circ(\kappa_{s_{1}}\otimes\cdots\otimes\kappa_{1}))\\ +&\sum_{m,l}(-1)^{**}T_{\vec{b},\vec{b}^{\prime}}^{k-m+1}(x_{k},\cdots,x_{m+l+1},\mu^{m}(x_{m+l},\cdots,x_{l+1}),x_{l},\cdots,x_{1}).\end{split}

Here the symbols μk\mu^{k} (temporarily) denote the AA_{\infty}-structure maps in the wrapped Fukaya category defined with respect to (HM,N,JM,N)(H_{M,N},J_{M,N}), and mkm^{k} denote those in the wrapped Fukaya category defined with respect to (K,J)(K,J). The signs are

=s1++si1deg(x1)deg(xs1+si1),*=s_{1}+\cdots+s_{i-1}-\deg(x_{1})-\cdots-\deg(x_{s_{1}+\cdots s_{i-1}}),

and

=deg(x1)++deg(xl)l1.**=\deg(x_{1})+\cdots+\deg(x_{l})-l-1.
Sketch of proof.

Checking this kind of compatibility is essentially repetition of the argument in the last section of [Gao17], now inductively on kk. The key reasoning is that the spaces of admissible Hamiltonians and of almost complex structures are contractible. ∎

This compatibility condition stated in Proposition 6.23 then implies that the homotopy direct limit of Rd,bkR^{k}_{d,\vec{b}} exists as bi+b_{i}\to+\infty for all ii.

Corollary 6.24.

There exist multilinear maps

(6.30) Rdk:CW(jk1,jk;HM,N,JM,N)CW(j0,j1;HM,N,JM,N)CW(j0,jk;K,J),\begin{split}R^{k}_{d}:&CW^{*}(\mathcal{L}_{j_{k-1}},\mathcal{L}_{j_{k}};H_{M,N},J_{M,N})\otimes\cdots\otimes CW^{*}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{1}};H_{M,N},J_{M,N})\\ &\to CW^{*}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J),\end{split}

such that when restricted to any truncated wrapped Floer complex, it is homotopic to the maps (6.22).

Proof.

We can modify the maps Rd,bkR_{d,\vec{b}}^{k} in (6.22) by composing them with self-homotopy equivalences on the truncated wrapped Floer complexes

CW(bi,ai](ji1,ji;HM,N,JM,N)CW^{*}_{(-b_{i},a_{i}]}(\mathcal{L}_{j_{i-1}},\mathcal{L}_{j_{i}};H_{M,N},J_{M,N})

and also

CW(b0,a0](j0,jk;K,J),CW^{*}_{(-b_{0},a_{0}]}(\mathcal{L}_{j_{0}},\mathcal{L}_{j_{k}};K,J),

so that Rd,bk(κkκ1)R_{d,\vec{b}^{\prime}}^{k}\circ(\kappa_{k}\otimes\cdots\otimes\kappa_{1}) and κ0Rd,bk\kappa_{0}\circ R_{d,\vec{b}}^{k} strictly agree after such modification. Thus the direct limit of the modified maps exists. The homotopy direct limit

(6.31) Rdk=limbi+Rd,bkR^{k}_{d}=\varinjlim_{b_{i}\to+\infty}R^{k}_{d,\vec{b}}

is defined to be the direct limit of the modified maps. ∎

The AA_{\infty}-functor equations for the sequence of multilinear maps {Rdk}k=1\{R_{d}^{k}\}_{k=1}^{\infty} follow from the universal and conformally consistent choice of action-restriction data. Recall that we have analogues of AA_{\infty}-functor equations for the maps (6.22). By taking homotopy direct limit as above, we obtain the desired AA_{\infty}-equations for the maps RdkR_{d}^{k}.

We have thus completed the construction of the AA_{\infty}-quasi-isomorphism (6.21).

6.8. Extension to the whole wrapped Fukaya categories

To finish the proof of Theorem 1.1, we need to be able to extend the functors (6.21) to the whole wrapped Fukaya categories. The approach we take is to apply the homotopy direct limit procedure described in section 2.6 to obtain a homotopy direct limit functor

(6.32) R:𝒲s(𝕃)𝒲~(𝕃),R:\mathcal{W}^{s}(\mathbb{L})\to\tilde{\mathcal{W}}(\mathbb{L}),

from the sequence of AA_{\infty}-functors Rd:𝒲ds𝒲dR_{d}:\mathcal{W}_{d}^{s}\to\mathcal{W}_{d}. Here 𝒲~(𝕃)\tilde{\mathcal{W}}(\mathbb{L}) is an AA_{\infty}-category quasi-equivalent to the wrapped Fukaya category 𝒲(𝕃)\mathcal{W}(\mathbb{L}), consisting of the same objects, i.e. those Lagrangians in the collection 𝕃\mathbb{L}. Then it follows immediately that RR is a quasi-equivalence, since each RdR_{d} is a quasi-isomorphism.

It remains to check the two assumptions in section 2.6. Assumption (2.16) automatically holds by the definition of these AA_{\infty}-subcategories 𝒲ds\mathcal{W}^{s}_{d} of 𝒲s(𝕃)\mathcal{W}^{s}(\mathbb{L}) and 𝒲d\mathcal{W}_{d} of 𝒲(𝕃)\mathcal{W}(\mathbb{L}), because of the inductive nature of the choices of Floer data used in the definition of the wrapped Fukaya category. Therefore, to get the homotopy direct limit functor (6.32), it suffices to check Assumption 2.17.

We restrict to the case of wrapped Fukaya categories: 𝒜=𝒲s(𝕃),𝒜d=𝒲ds,=𝒲(𝕃),d=𝒲d\mathcal{A}=\mathcal{W}^{s}(\mathbb{L}),\mathcal{A}_{d}=\mathcal{W}^{s}_{d},\mathcal{B}=\mathcal{W}(\mathbb{L}),\mathcal{B}_{d}=\mathcal{W}_{d}, and the action-restriction functors d=Rd\mathcal{F}_{d}=R_{d} that we define on these subcategories. Note that Assumption (2.17) is rather weak, in the sense that it does not ask for a specific homotopy between the two AA_{\infty}-functors jd,d1Rdid,dj_{d,d^{\prime}}^{-1}\circ R_{d^{\prime}}\circ i_{d,d^{\prime}} and RdR_{d}. Thus the freedom in the choice of such a homotopy makes the process quite flexible.

Since there are at most dd^{\prime} Lagrangian submanifolds 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d^{\prime}} to consider at each time, the AA_{\infty}-functors RdR_{d} and RdR_{d^{\prime}} are quite concrete: they are determined by our choices of action-restriction data. There are two families of such choices - one for the collection of Lagrangian labels 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d}, denoted by D0D_{0}, the other for 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d^{\prime}}, denoted by D1D_{1} (meaning choices for all elements in 𝒩¯k+1\bar{\mathcal{N}}_{k+1} in a consistent way. We may also restrict D1D_{1} to the collection 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d} to obtain consistent choices of action-restriction data for 1,,d\mathcal{L}_{1},\cdots,\mathcal{L}_{d}, still denoted by D1D_{1}. The idea of proving that RdR_{d} and jd,d1Rdid,dj_{d,d^{\prime}}^{-1}\circ R_{d^{\prime}}\circ i_{d,d^{\prime}} are homotopic as AA_{\infty}-functors from 𝒲ds\mathcal{W}^{s}_{d} to 𝒲d\mathcal{W}_{d} is to choose a one-parameter family DtD_{t} of action-restriction data interpolating these two, and then use the resulting parametrized moduli spaces to construct the desired homotopy between the two action-restriction functors determined by D0D_{0} and D1D_{1} respectively. The existence of such one-parameter family is because the spaces of such Hamiltonians/almost complex structures are contractible.

To construct a homotopy between the functors associated to these two data D0,D1D_{0},D_{1}, we need a one-dimensional higher analogue of the multiplihedra, which we call the homotopehedra and denote by 𝒫¯k+1\bar{\mathcal{P}}_{k+1}. Define 𝒫k+1=𝒩k+1×[0,+)\mathcal{P}_{k+1}=\mathcal{N}_{k+1}\times[0,+\infty). The compactification 𝒫¯k+1\bar{\mathcal{P}}_{k+1} has boundary strata made of products of copies of i+1,𝒩j+1\mathcal{M}_{i+1},\mathcal{N}_{j+1} as well as 𝒫l+1\mathcal{P}_{l+1}.

Now we consider moduli space 𝒫k+1(x~0;x1,,xk)\mathcal{P}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}) of elements (S,w,t,u)(S,w,t,u) where (S,w,t)𝒫k+1(S,w,t)\in\mathcal{P}_{k+1}, and u:(S,w)Mu:(S,w)\to M is a continuation disk satisfying the Cauchy-Riemann equation with respect to the Hamiltonian and almost complex structure from the action-restriction datum DtD_{t}, which converges to some KK-chord x~0\tilde{x}_{0} over the 0-th strip-like end, and to some HM,NH_{M,N}-chord xix_{i} over the ii-th strip-like end.

There is a natural bordification 𝒫¯k+1(x~0;x1,,xk)\bar{\mathcal{P}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}), whose codimension one boundary strata consist of a union of product moduli spaces of the following form:

(6.33) 𝒫¯k+1(x~0;x1,,xk)=m,n𝒫km(x~0;x1,,xn,x,xn+m+1,,xk)×m+1(x;xn+1,,xn+m)r,si1+ir=kx~1,,x~rr+1(x~0;x~1,x~r)×(𝒩i1+1(x~1;x1,,xi1;D0)××𝒩is1+1(x~s1;xi1++is2+1,,xi1++is1;D0)×𝒫is+1(x~s;xi1++is1+1,,xi1++is;Dt)×𝒩is+1+1(x~s+1;xi1++is+1,,xi1++is+1;D1)××𝒩ir+1(x~r;xi1++ir1+1,,xk;D1))𝒩k+1(x~0;x1,,xk;D1)𝒩k+1(x~0;x1,,xk;D0).\begin{split}&\partial\bar{\mathcal{P}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k})\\ &=\coprod_{m,n}\mathcal{P}_{k-m}(\tilde{x}_{0};x_{1},\cdots,x_{n},x^{\prime},x_{n+m+1},\cdots,x_{k})\times\mathcal{M}_{m+1}(x^{\prime};x_{n+1},\cdots,x_{n+m})\\ &\cup\coprod_{\begin{subarray}{c}r,s\\ i_{1}+\cdots i_{r}=k\end{subarray}}\coprod_{\tilde{x}^{\prime}_{1},\cdots,\tilde{x}^{\prime}_{r}}\mathcal{M}_{r+1}(\tilde{x}_{0};\tilde{x}^{\prime}_{1},\cdots\tilde{x}^{\prime}_{r})\times(\mathcal{N}_{i_{1}+1}(\tilde{x}^{\prime}_{1};x_{1},\cdots,x_{i_{1}};D_{0})\\ &\times\cdots\times\mathcal{N}_{i_{s-1}+1}(\tilde{x}^{\prime}_{s-1};x_{i_{1}+\cdots+i_{s-2}+1},\cdots,x_{i_{1}+\cdots+i_{s-1}};D_{0})\\ &\times\mathcal{P}_{i_{s}+1}(\tilde{x}^{\prime}_{s};x_{i_{1}+\cdots+i_{s-1}+1},\cdots,x_{i_{1}+\cdots+i_{s}};D_{t})\\ &\times\mathcal{N}_{i_{s+1}+1}(\tilde{x}^{\prime}_{s+1};x_{i_{1}+\cdots+i_{s}+1},\cdots,x_{i_{1}+\cdots+i_{s+1}};D_{1})\\ &\times\cdots\times\mathcal{N}_{i_{r}+1}(\tilde{x}^{\prime}_{r};x_{i_{1}+\cdots+i_{r-1}+1},\cdots,x_{k};D_{1}))\\ &\cup\mathcal{N}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k};D_{1})\\ &\cup\mathcal{N}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k};D_{0}).\end{split}

Here the xx’s without tilde denote HM,NH_{M,N}-chords, while the x~\tilde{x}’s denote KK-chords.

By counting rigid elements in the moduli spaces 𝒫¯k+1(x~0;x1,,xk)\bar{\mathcal{P}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}), we construct from this one-parameter family of action-restriction data DtD_{t} a multilinear map

(6.34) Tk:CW(0,1;HM,N)CW(k1,k;HM,N)CW(0,k;K)T^{k}:CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1};H_{M,N})\otimes\cdots\otimes CW^{*}(\mathcal{L}_{k-1},\mathcal{L}_{k};H_{M,N})\to CW^{*}(\mathcal{L}_{0},\mathcal{L}_{k};K)

of degree k-k, whose (x~0;x1,,xk)(\tilde{x}_{0};x_{1},\cdots,x_{k}) component is the count of rigid elements in the moduli space 𝒫¯k+1(x~0;x1,,xk)\bar{\mathcal{P}}_{k+1}(\tilde{x}_{0};x_{1},\cdots,x_{k}).

Setting T0=0T^{0}=0, we claim that T={Tk}k=0T=\{T^{k}\}_{k=0}^{\infty} is a homotopy between the functors RdR_{d} and jd,d1Rdid,dj_{d,d^{\prime}}^{-1}\circ R_{d^{\prime}}\circ i_{d,d^{\prime}}, where the former is defined by the action-restriction datum D0D_{0}, and the latter by D1D_{1}. Verifying the relation m𝒬1(T)=jd,d1Rdid,dRdm^{1}_{\mathcal{Q}}(T)=j_{d,d^{\prime}}^{-1}\circ R_{d^{\prime}}\circ i_{d,d^{\prime}}-R_{d} amounts to looking at the boundary strata of the bordification 𝒫¯k+1(x0;x1,,xk)\bar{\mathcal{P}}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}) as above (6.33). This relation is precisely the condition for RdR_{d} and jd,d1Rdid,dj_{d,d^{\prime}}^{-1}\circ R_{d^{\prime}}\circ i_{d,d^{\prime}} to be homotopic as AA_{\infty}-functors. Thus we have verified Assumption 2.17, and thus completed the proof of Theorem 1.1.

Remark 6.25.

Here is a technical remark regarding the general definition of the wrapped Fukaya category that we use here. In practice, one starts with an at most countable collection of closed exact or cylindrical Lagrangian submanifolds, and declare them to be the objects of the wrapped Fukaya category. It is only in this sense our perturbation framework makes sense inductively, and the extension of the action-restriction functor to the whole wrapped Fukaya category is valid via the procedure we described in section 2.6.

In principle, one could allow all (not just a countable collection of) admissible Lagrangian submanifolds as objects of the wrapped Fukaya category, but should pay close attension to the transversality argument: one must carefully make choices of Floer data each time, and use Axiom of Choice uncountably many times to define the AA_{\infty}-structure on the wrapped Fukaya category. However, it seems technically difficult to extend the action-restriction functor in that setup. We will not discuss this technical point since it is not needed for most applications.

7. AA_{\infty}-functors associated to Lagrangian correspondences

7.1. Extension of quilted wrapped Floer cohomology to Lagrangian immersions

This section provides chain-level refinements of the construction of cohomological functors in [Gao17]. That is, we are going to prove that admissible Lagrangian correspondences give rise to functors between appropriate versions of wrapped Fukaya categories. For simplicity, we let the source of the functors be the wrapped Fukaya category of MM consisting of embedded exact cylindrical Lagrangian submanifolds. All AA_{\infty}-functors are to be understood as cohomologically unital AA_{\infty}-functors.

As an introductory part of the main construction, we first give a naive attempt in extending quilted wrapped Floer cohomology to exact cylindrical Lagrangian immersions with transverse or clean self-intersections. For our purpose of constructing functors from Lagrangian correspondences, we shall only consider the case where LML\subset M and M×N\mathcal{L}\subset M^{-}\times N are properly embedded, while LNL^{\prime}\subset N is replaced by an exact cylindrical Lagrangian immersion ι:LN\iota:L^{\prime}\to N.

One short-cut definition for the quilted wrapped Floer cochain space is given as follows. As the underlying \mathbb{Z}-module, the quilted wrapped Floer cochain space

CW(L,,(L,ι))CW^{*}(L,\mathcal{L},(L^{\prime},\iota))

is defined as the wrapped Floer cochain space

CW(,L×(L,ι)),CW^{*}(\mathcal{L},L\times(L^{\prime},\iota)),

for the pair of exact cylindrical Lagrangian immersions in M×NM^{-}\times N, where L×(L,ι)L\times(L^{\prime},\iota) is the obvious product Lagrangian immersion with clean self-intersections. This pair has clean intersections, whose wrapped Floer cochain space is defined in section 5.5. The quilted Floer ”differential” n0n^{0} is defined as the zeroth-order curved AA_{\infty}-structure map on the above wrapped Floer cochain space CW(,L×(L,ι))CW^{*}(\mathcal{L},L\times(L^{\prime},\iota)). Here we put the quotation mark because n0n^{0} might not square to zero in general. Alternatively, there is another straightforward definition, using moduli spaces of inhomogeneous pseudoholomorphic quilted strips, following the standard setup of quilted wrapped Floer theory. In fact, these inhomogeneous pseudoholomorphic quilted strips are in natural bijection to inhomogeneous pseudoholomorphic strips in the product manifold, and we can choose the same perturbations (by multisections) for both moduli spaces. Thus the second definition is equivalent to the first one.

The second definition is more suitable for discussing the AA_{\infty}-bimodule structure on the quilted wrapped Floer cochain space CW(L,,(L,ι))CW^{*}(L,\mathcal{L},(L^{\prime},\iota)). The details of the construction will be discussed in subsection 7.2.

7.2. The module-valued functors associated to Lagrangian correspondences

The guiding principle for constructing AA_{\infty}-functors from Lagrangian correspondences is to use moduli spaces of inhomogeneous pseudoholomorphic quilted maps. For these AA_{\infty}-functors to be defined over \mathbb{Z}, we must ensure that these moduli spaces carry coherent orientations. The reader is referred to the Appendix of [Gao17] for the discussion on orientations on the relevant moduli spaces of inhomogeneous pseudoholomorphic quilts, where the discussion focused on one particular kind of moduli space but can be easily generalized to all the other ones which we actually use in this paper. For a more general discussion on orientations of moduli spaces of pseudoholomorphic quilts, we refer the reader to [WW15a], but remark that our approach is independent because some of the quilted surfaces we use are not included there.

The starting point is to to relate the wrapped Fukaya category 𝒲(M×N)\mathcal{W}(M^{-}\times N) of the product manifold, to the dg-category of AA_{\infty}-bimodules over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)).

Proposition 7.1.

There is a canonical AA_{\infty}-functor

(7.1) Φ:𝒲(M×N)(𝒲(M),𝒲(N))bimod,\Phi:\mathcal{W}(M^{-}\times N)\to(\mathcal{W}(M),\mathcal{W}(N))^{bimod},

satisfying the following properties:

  1. (i)

    Φ\Phi is non-trivial for any non-trivial 𝒲(M×N)\mathcal{W}(M^{-}\times N);

  2. (ii)

    If either MM or NN is a point, Φ\Phi is the Yoneda functor for NN or MM;

  3. (iii)

    If =L×L\mathcal{L}=L\times L^{\prime} is a product Lagrangian correspondence, then the AA_{\infty}-bimodule Φ()\Phi(\mathcal{L}) splits. That is, there is an isomorphism of AA_{\infty}-bimodules

    (7.2) Φ()𝔶r(L)𝔶l(L),\Phi(\mathcal{L})\cong\mathfrak{y}_{r}(L)\otimes\mathfrak{y}_{l}(L^{\prime}),

    where 𝔶r\mathfrak{y}_{r} and 𝔶l\mathfrak{y}_{l} are the right and left Yoneda functors.

The idea of proof is to develop a quilted version of wrapped Floer theory, extending that in [Gao17]. The natural construction should yield an AA_{\infty}-functor from the split model of the wrapped Fukaya category 𝒲s(M×N)\mathcal{W}^{s}(M^{-}\times N). However, based on the results of section 6, that is equivalent to the ordinary wrapped Fukaya category 𝒲(M×N)\mathcal{W}(M^{-}\times N). Thus it does not matter which version of wrapped Fukaya category of the product manifold we use as far as algebraic structures are concerned, at least up to quasi-equivalence.

Definition 7.2.

A Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N from MM to NN is said to be admissible, if it is admissible for wrapped Floer theory in the product manifold M×NM^{-}\times N in the sense of Definition 6.3, i.e. is an object of the wrapped Fukaya category 𝒲(M×N)\mathcal{W}(M^{-}\times N).

As a result, by evaluating the above AA_{\infty}-functor at each given object \mathcal{L} of 𝒲(M×N)\mathcal{W}(M^{-}\times N), i.e. an admissible Lagrangian correspondence, we then get an AA_{\infty}-bimodule over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)). By purely algebraic consideration involving the Yoneda embedding, we have the following AA_{\infty}-functor associated to \mathcal{L}:

Corollary 7.3.

For any admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N, there is an associated AA_{\infty}-functor:

(7.3) Φ:𝒲(M)𝒲(N)lmod,\Phi_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}(N)^{l-mod},

to the dg-category of left AA_{\infty}-modules over 𝒲(N)\mathcal{W}(N).

In particular, we remark that such an AA_{\infty}-functor is defined for any admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N, without any properness assumption. However, as we shall see in the next subsection, a suitable properness assumption is needed in order to prove that this module-valued functor is representable, thus can be improved to a filtered AA_{\infty}-functor to the immersed wrapped Fukaya category 𝒲im(N)\mathcal{W}_{im}(N).

Now let us discuss in detail the construction of the bimodule-valued functor (7.1). On the level of objects, the functor Φ\Phi should assign an AA_{\infty}-bimodule Φ()\Phi(\mathcal{L}) over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)) to an admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N. The first order bimodule structure map of Φ()\Phi(\mathcal{L}) has already been constructed in [Gao17]. We now give an extension of that, defining the AA_{\infty}-bimodule structure maps of all orders in a uniform treatment.

Consider the quilted surface S¯k,l\underline{S}^{k,l} consisting of two patches S0k,S1lS^{k}_{0},S^{l}_{1}, where S0kS^{k}_{0} is a disk with (k+2)(k+2) boundary punctures z0,z01,,z0k,z0+z_{0}^{-},z_{0}^{1},\cdots,z_{0}^{k},z_{0}^{+}, and S1lS^{l}_{1} is a disk with (l+2)(l+2) boundary punctures z1,z11,,z1l,z1+z_{1}^{-},z_{1}^{1},\cdots,z_{1}^{l},z_{1}^{+}. Let I0±I_{0}^{\pm} be the boundary component of S0kS^{k}_{0} between z0+z_{0}^{+} and z0z_{0}^{-}, and I1±I_{1}^{\pm} the boundary component of S1lS^{l}_{1} between z1+z_{1}^{+} and z1z_{1}^{-}. The quilted surface is obtained by seaming the two patches along these two boundary components. After seaming the two patches, the quilted surface S¯k,l\underline{S}^{k,l} has (k+l)(k+l) positive strip-like ends ϵ01,,ϵ0k\epsilon_{0}^{1},\cdots,\epsilon_{0}^{k} and ϵ11,,ϵ1l\epsilon_{1}^{1},\cdots,\epsilon_{1}^{l} as well as two quilted ends (one positive and one negative), each of which consists of two strip-like ends. See the picture below.

L3L_{3}L2L_{2}L1L_{1}L0L_{0}\mathcal{L}L0L^{\prime}_{0}L1L^{\prime}_{1}L2L^{\prime}_{2}x3x^{3}x2x^{2}x1x^{1}(x,y)(x^{-},y^{-})(x+,y+)(x^{+},y^{+})y1y^{1}y2y^{2}
Figure 7.1. The quilted map defining the bimodule

Choosing a Floer datum for S¯k,l\underline{S}^{k,l} allows us to define inhomogeneous pseudoholomorphic quilted maps from S¯k,l\underline{S}^{k,l} to the pair (M,N)(M,N) with appropriate moving Lagrangian boundary conditions and asymptotic convergence conditions over the various ends. We shall choose Floer data for all (representatives of) such quilted surfaces in the moduli spaces, and extend the choices by automorphism-invariant Floer data on unstable components, i.e. quilted strips, of semistable quilted surfaces, such that they are compatible under gluing maps with the universal and consistent choices we made for disks.

Let k,l((x,y);(x+,y+),x,y)\mathcal{R}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) be the moduli space, namely the set of isomorphism classes of inhomogeneous pseudoholomorphic quilted maps (S¯k,l,u¯)(\underline{S}^{k,l},\underline{u}) as pictured in Figure 7.1. There is a natural Gromov bordification

(7.4) ¯k,l((x,y);(x+,y+),x,y),\bar{\mathcal{R}}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}),

which compactifies k,l((x,y);(x+,y+),x,y)\mathcal{R}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}). The codimension one boundary strata of ¯k,l((x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) are covered by a union of products of moduli spaces of the following form:

(7.5) 1ik(x,x′′)=xxnew¯ki+1,l((x,y);(x+,y+),x,xnew,y)ׯi+1(xnew,x′′)1jl(y,y′′)=yynew¯k,lj+1((x,y);(x+,y+),x,y,ynew)ׯj+1(ynew,y′′)k+k′′=k,l+l′′=l(x1+,y1+)¯k,l((x,y);(x1+,y1+),x,y)ׯk′′,l′′((x1+,y1+);(x+,y+),x′′,y′′)\begin{split}\coprod_{\begin{subarray}{c}1\leq i\leq k\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x}\end{subarray}}\coprod_{x_{new}}&\bar{\mathcal{R}}^{k-i+1,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x}^{\prime},x_{new},\vec{y})\times\bar{\mathcal{M}}_{i+1}(x_{new},\vec{x}^{\prime\prime})\\ \cup\coprod_{\begin{subarray}{c}1\leq j\leq l\\ (\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{y_{new}}&\bar{\mathcal{R}}^{k,l-j+1}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}^{\prime},y_{new})\times\bar{\mathcal{M}}_{j+1}(y_{new},\vec{y}^{\prime\prime})\\ \cup\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (x^{+}_{1},y^{+}_{1})\end{subarray}}&\bar{\mathcal{R}}^{k^{\prime},l^{\prime}}((x^{-},y^{-});(x^{+}_{1},y^{+}_{1}),\vec{x}^{\prime},\vec{y}^{\prime})\\ &\times\bar{\mathcal{R}}^{k^{\prime\prime},l^{\prime\prime}}((x^{+}_{1},y^{+}_{1});(x^{+},y^{+}),\vec{x}^{\prime\prime},\vec{y}^{\prime\prime})\end{split}

Here by the notation ¯ki+1,l((x,y);(x+,y+),x,xnew,y)\bar{\mathcal{R}}^{k-i+1,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x}^{\prime},x_{new},\vec{y}) we mean to insert the new Hamiltonian chord xnewx_{new} in every possible place that splits the tuple x\vec{x} of Hamiltonian chords to the two tuples x\vec{x}^{\prime} and x′′\vec{x}^{\prime\prime}, as long as the cyclic order is preserved. Similar remarks apply to the yy’s.

In [WW12], it is demonstrated that this kind of moduli space is locally modeled on a Fredholm complex. Since there are no disk bubbles, we can use the standard transversality argument to prove that for generic universal and consistent choices of Floer data, the Gromov bordification ¯k,l((x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) is a compact smooth manifold with corners of expected dimension

(7.6) k2+l2+deg((x,y))deg((x+,y+))deg(xi)deg(yj).k-2+l-2+\deg((x^{-},y^{-}))-\deg((x^{+},y^{+}))-\sum\deg(x^{i})-\sum\deg(y^{j}).

And moreover, we can arrange the perturbations so that every stratum is regular.

A finiteness result, which says that given inputs (x+,y+),x,y(x^{+},y^{+}),\vec{x},\vec{y}, the moduli spaces ¯k,l((x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) are empty for all but finitely many outputs (x,y)(x^{-},y^{-}), can be deduced from the action-energy equality, which plays a crucial role in the well-definedness of various maps in wrapped Floer theory ([AS10], [Abo10], also see [Gao17] in which we used the quilted version in special cases k1,l1k\leq 1,l\leq 1). This ensures that the count of rigid elements of all moduli spaces ¯k,l((x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) for fixed inputs (x+,y+),x,y(x^{+},y^{+}),\vec{x},\vec{y} is finite, which gives rise to a map

(7.7) nk|0|l:CW(Lk1,Lk)CW(L0,L1)CW(Lk,,L0)CW(Ll1,Ll)CW(L0,L1)CW(L0,,Ll).\begin{split}n^{k|0|l}&:CW^{*}(L_{k-1},L_{k})\otimes\cdots\otimes CW^{*}(L_{0},L_{1})\otimes CW^{*}(L_{k},\mathcal{L},L^{\prime}_{0})\\ &\otimes CW^{*}(L^{\prime}_{l-1},L^{\prime}_{l})\otimes\cdots\otimes CW^{*}(L^{\prime}_{0},L^{\prime}_{1})\to CW^{*}(L_{0},\mathcal{L},L^{\prime}_{l}).\end{split}

By analyzing the boundary strata of one dimensional moduli spaces as described in (7.5), we find that the operation nk|0|ln^{k|0|l} satisfies the following equation

(7.8) m1nk|0|l([x],[x+,y+],[y])=nki+1|0|l([x],mi([x′′]),[x+,y+],[y])+nk|0|lj+1([x],[x+,y+],[y],mj([y′′]))+k+k′′=k,l+l′′=l(x,x′′)=xnk|0|l([x],nk′′|0|l′′([x′′],[x+,y+],[y′′]),[y]).\begin{split}&m^{1}\circ n^{k|0|l}([\vec{x}],[x^{+},y^{+}],[\vec{y}])\\ &=\sum n^{k-i+1|0|l}([\vec{x}^{\prime}],m^{i}([\vec{x}^{\prime\prime}]),[x^{+},y^{+}],[\vec{y}])\\ &+\sum n^{k|0|l-j+1}([\vec{x}],[x^{+},y^{+}],[\vec{y}^{\prime}],m^{j}([\vec{y}^{\prime\prime}]))\\ &+\sum_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x}\end{subarray}}n^{k^{\prime}|0|l^{\prime}}([\vec{x}^{\prime}],n^{k^{\prime\prime}|0|l^{\prime\prime}}([\vec{x}^{\prime\prime}],[x^{+},y^{+}],[\vec{y}^{\prime\prime}]),[\vec{y}^{\prime}]).\end{split}

This precisely means that the operations nk|0|ln^{k|0|l} define an AA_{\infty}-bimodule structure on Φ()\Phi(\mathcal{L}) over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)).

Next we want to study what the AA_{\infty}-functor (7.1) does to morphisms, and how this is related with the AA_{\infty}-structure maps of 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N). A Floer cochain [γ]CW(0,1)[\gamma]\in CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1}) should give rise to an AA_{\infty}-bimodule homomorphism

(7.9) Φγ:Φ0Φ1.\Phi_{\gamma}:\Phi_{\mathcal{L}_{0}}\to\Phi_{\mathcal{L}_{1}}.

Moreover, this should be functorial in the wrapped Fukaya category of the product manifold M×NM^{-}\times N as stated in Proposition 7.1. More specifically, it means that there are multilinear maps

(7.10) nk|1|l:CW(0,1)hom(CW(Lk1,Lk)CW(L0,L1)CW(L0,0,Ll)CW(Ll1,Ll)CW(L0,L1),CW(Lk,1,L0)),\begin{split}n^{k|1|l}&:CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1})\to\hom(CW^{*}(L_{k-1},L_{k})\otimes\cdots\otimes CW^{*}(L_{0},L_{1})\\ &\otimes CW^{*}(L_{0},\mathcal{L}_{0},L^{\prime}_{l})\otimes CW^{*}(L^{\prime}_{l-1},L^{\prime}_{l})\\ &\otimes\cdots\otimes CW^{*}(L^{\prime}_{0},L^{\prime}_{1}),CW^{*}(L_{k},\mathcal{L}_{1},L^{\prime}_{0})),\end{split}

such that when evaluated on [γ]CW(0,1)[\gamma]\in CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1}), the resulting maps form the desired AA_{\infty}-bimodule homomorphism (7.9).

To define the maps (7.10), we shall study moduli spaces of inhomogeneous pseudoholomorphic maps from another kind of quilted surface S¯1,k,l\underline{S}^{1,k,l}, which we describe as follows. It has two patches S01,k,S11,lS^{1,k}_{0},S^{1,l}_{1}, where S01,kS^{1,k}_{0} is a disk with (k+3)(k+3) boundary punctures z0+,z0,z0p,z01,,z0kz_{0}^{+},z_{0}^{-},z_{0}^{p},z_{0}^{1},\cdots,z_{0}^{k}, and S11,lS^{1,l}_{1} is a disk with (l+3)(l+3) boundary punctures z1+,z1,z1p,z11,,z1kz_{1}^{+},z_{1}^{-},z_{1}^{p},z_{1}^{1},\cdots,z_{1}^{k}. We denote by I0,+I_{0,+} the boundary component of S01,kS^{1,k}_{0} between z0+z_{0}^{+} and z0pz_{0}^{p}, and by I0,I_{0,-} that between z0pz_{0}^{p} and z0z_{0}^{-}. We use similar notations for S1,lS^{1,l}. The quilted surface is obtained by seaming the two patches along the two pairs of boundary components (I0,+,I1,+)(I_{0,+},I_{1,+}) and (I0,,I1,)(I_{0,-},I_{1,-}).

Choose a Floer datum for S¯1,k,l\underline{S}^{1,k,l} so that we can write down the inhomogeneous Cauchy-Riemann equation for quilted maps u¯:S¯1,k,l(M,N)\underline{u}:\underline{S}^{1,k,l}\to(M,N):

(7.11) {(du0αS01,kXHS01,k)0,1=0(du1αS11,lXHS11,l)0,1=0u0(z)ϕMρS01,k(z)Li, if zS01,k lies between z0i and z0i+1u0(z)ϕMρS01,k(z)L0, if zS01,k lies between z0+ and z01 u0(z)ϕMρS01,k(z)Lk, if zS01,l lies between z0k and z0 u1(z)ϕNρS11,l(z)Lj, if zS11,l lies between z1j and z1j+1u1(z)ϕNρS11,l(z)L0, if zS11,l lies between z1+ and z11 u1(z)ϕNρS11,l(z)Ll, if zS11,l lies between z1l and z1 (u0(z),u1(z))(ϕMρS01,k(z)×ϕNρS11,l(z))1, if zS01,k lies between z0 and z0p(u0(z),u1(z))(ϕMρS01,k(z)×ϕNρS11,l(z))0, if zS01,k lies between z0p and z0+lims(u0ϵ0(s,),u1ϵ1(s,))=(x(),y())lims+(u0ϵ0+(s,),u1ϵ1+(s,))=(x+(),y+())lims+(u0ϵ0p(s,),u1ϵ1p(s,))=γ()lims+u0ϵ0i(s,)=xi(),i=1,,klims+u1ϵ1j(s,)=yj(),j=1,,l\begin{cases}(du_{0}-\alpha_{S^{1,k}_{0}}\otimes X_{H_{S^{1,k}_{0}}})^{0,1}=0\\ (du_{1}-\alpha_{S^{1,l}_{1}}\otimes X_{H_{S^{1,l}_{1}}})^{0,1}=0\\ u_{0}(z)\in\phi_{M}^{\rho_{S^{1,k}_{0}}(z)}L_{i},\text{ if $z\in\partial S^{1,k}_{0}$ lies between $z_{0}^{i}$ and $z_{0}^{i+1}$}\\ u_{0}(z)\in\phi_{M}^{\rho_{S^{1,k}_{0}}(z)}L_{0},\text{ if $z\in\partial S^{1,k}_{0}$ lies between $z_{0}^{+}$ and $z_{0}^{1}$ }\\ u_{0}(z)\in\phi_{M}^{\rho_{S^{1,k}_{0}}(z)}L_{k},\text{ if $z\in\partial S^{1,l}_{0}$ lies between $z_{0}^{k}$ and $z_{0}^{-}$ }\\ u_{1}(z)\in\phi_{N}^{\rho_{S^{1,l}_{1}}(z)}L^{\prime}_{j},\text{ if $z\in\partial S^{1,l}_{1}$ lies between $z_{1}^{j}$ and $z_{1}^{j+1}$}\\ u_{1}(z)\in\phi_{N}^{\rho_{S^{1,l}_{1}}(z)}L^{\prime}_{0},\text{ if $z\in\partial S^{1,l}_{1}$ lies between $z_{1}^{+}$ and $z_{1}^{1}$ }\\ u_{1}(z)\in\phi_{N}^{\rho_{S^{1,l}_{1}}(z)}L^{\prime}_{l},\text{ if $z\in\partial S^{1,l}_{1}$ lies between $z_{1}^{l}$ and $z_{1}^{-}$ }\\ (u_{0}(z),u_{1}(z))\in(\phi_{M}^{\rho_{S^{1,k}_{0}}(z)}\times\phi_{N}^{\rho_{S^{1,l}_{1}}(z)})\mathcal{L}_{1},\text{ if $z\in\partial S^{1,k}_{0}$ lies between $z_{0}^{-}$ and $z_{0}^{p}$}\\ (u_{0}(z),u_{1}(z))\in(\phi_{M}^{\rho_{S^{1,k}_{0}}(z)}\times\phi_{N}^{\rho_{S^{1,l}_{1}}(z)})\mathcal{L}_{0},\text{ if $z\in\partial S^{1,k}_{0}$ lies between $z_{0}^{p}$ and $z_{0}^{+}$}\\ \lim\limits_{s\to-\infty}(u_{0}\circ\epsilon_{0}^{-}(s,\cdot),u_{1}\circ\epsilon_{1}^{-}(s,\cdot))=(x^{-}(\cdot),y^{-}(\cdot))\\ \lim\limits_{s\to+\infty}(u_{0}\circ\epsilon_{0}^{+}(s,\cdot),u_{1}\circ\epsilon_{1}^{+}(s,\cdot))=(x^{+}(\cdot),y^{+}(\cdot))\\ \lim\limits_{s\to+\infty}(u_{0}\circ\epsilon_{0}^{p}(s,\cdot),u_{1}\circ\epsilon_{1}^{p}(s,\cdot))=\gamma(\cdot)\\ \lim\limits_{s\to+\infty}u_{0}\circ\epsilon_{0}^{i}(s,\cdot)=x^{i}(\cdot),i=1,\cdots,k\\ \lim\limits_{s\to+\infty}u_{1}\circ\epsilon_{1}^{j}(s,\cdot)=y^{j}(\cdot),j=1,\cdots,l\end{cases}

Here [xi]CW(Li1,Li),[yi]CW(Lj1,Lj)[x^{i}]\in CW^{*}(L_{i-1},L_{i}),[y^{i}]\in CW^{*}(L^{\prime}_{j-1},L^{\prime}_{j}) are Hamiltonian chords in MM and NN respectively, [γ]CW(0,1)[\gamma]\in CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1}) is a Hamiltonian chord in M×NM^{-}\times N with respect to the split Hamiltonian, and [(x,y)]CW(Lk,1,L0)[(x^{-},y^{-})]\in CW^{*}(L_{k},\mathcal{L}_{1},L^{\prime}_{0}), [(x+,y)]CW(L0,0,Ll)[(x^{+},y^{-})]\in CW^{*}(L_{0},\mathcal{L}_{0},L^{\prime}_{l}) are generalized chords for the corresponding Lagrangian boundary and seaming conditions. We omit suitable rescalings of the asymptotic Hamiltonian chords by the Liouville flow for the purpose of simplifying notation, but shall keep in mind that these can be chosen and have been chosen in a consistent way. The Lagrangian boundary conditions are ordered as follows: on the boundary of the first patch, Lk,,L0L_{k},\cdots,L_{0} are in order from the negative quilted puncture to the positive quilted puncture; on the boundary of the second patch, L0,,LlL^{\prime}_{0},\cdots,L^{\prime}_{l} are in order from the negative quilted puncture to the positive quilted puncture; on the seam, d,,0\mathcal{L}_{d},\cdots,\mathcal{L}_{0} are in order from the negative quilted puncture to the positive quilted puncture.

Let 1,k,l((x,y);x,γ,(x+,y+),y)\mathcal{R}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}) be the moduli space of solutions (S¯1,k,l,u¯)(\underline{S}^{1,k,l},\underline{u}) to the above equation. Here by S¯1,k,l\underline{S}^{1,k,l} in the triple we mean a complex structure on S¯1,k,l\underline{S}^{1,k,l} up to isomorphism. The Gromov bordification ¯1,k,l((x,y);x,γ,(x+,y+),y)\bar{\mathcal{R}}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}) is in fact a compactification, with its codimension one stratum covered by the following union of fiber products of moduli spaces

(7.12) 1ik(x,x′′)=xxnew¯1,ki+1,l((x,y);x,xnew,γ,(x+,y+),y)ׯi+1(xnew,x′′)1jl(y,y′′)=yynew¯1,l,lj+1((x,y);x,γ,(x+,y+),y)ׯj+1(ynew,y′′)γ1¯1,k,l((x,y);x,γ1,(x+,y+),y)ׯ(γ1,γ)k+k′′=k,l+l′′=l(x,x′′)=x,(y,y′′)=y(x1+,y1+)¯1,k,l((x,y);x,γ,(x1+,y1+),y)ׯ0,k′′,l′′((x1+,y1+);x′′,(x+,y+),y′′)k+k′′=k,l+l′′=l(x,x′′)=x,(y,y′′)=y(x1+,y1+)¯0,k,l((x,y);x,(x1+,y1+),y)ׯ1,k′′,l′′((x1+,y1+);x′′,γ,(x+,y+),y′′).\begin{split}&\coprod_{\begin{subarray}{c}1\leq i\leq k\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x}\end{subarray}}\coprod_{x_{new}}\bar{\mathcal{R}}^{1,k-i+1,l}((x^{-},y^{-});\vec{x}^{\prime},x_{new},\gamma,(x^{+},y^{+}),\vec{y})\times\bar{\mathcal{M}}_{i+1}(x_{new},\vec{x}^{\prime\prime})\\ &\cup\coprod_{\begin{subarray}{c}1\leq j\leq l\\ (\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{y_{new}}\bar{\mathcal{R}}^{1,l,l-j+1}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}^{\prime})\times\bar{\mathcal{M}}_{j+1}(y_{new},\vec{y}^{\prime\prime})\\ &\cup\coprod_{\gamma_{1}}\bar{\mathcal{R}}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma_{1},(x^{+},y^{+}),\vec{y})\times\bar{\mathcal{M}}(\gamma_{1},\gamma)\\ &\cup\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x},(\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{(x^{+}_{1},y^{+}_{1})}\bar{\mathcal{R}}^{1,k^{\prime},l^{\prime}}((x^{-},y^{-});\vec{x}^{\prime},\gamma,(x^{+}_{1},y^{+}_{1}),\vec{y}^{\prime})\\ &\times\bar{\mathcal{R}}^{0,k^{\prime\prime},l^{\prime\prime}}((x^{+}_{1},y^{+}_{1});\vec{x}^{\prime\prime},(x^{+},y^{+}),\vec{y}^{\prime\prime})\\ &\cup\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x},(\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{(x^{+}_{1},y^{+}_{1})}\bar{\mathcal{R}}^{0,k^{\prime},l^{\prime}}((x^{-},y^{-});\vec{x}^{\prime},(x^{+}_{1},y^{+}_{1}),\vec{y}^{\prime})\\ &\times\bar{\mathcal{R}}^{1,k^{\prime\prime},l^{\prime\prime}}((x^{+}_{1},y^{+}_{1});\vec{x}^{\prime\prime},\gamma,(x^{+},y^{+}),\vec{y}^{\prime\prime}).\end{split}

In this situation, the underlying quilted surface is not obtained by gluing patches tangentially, thus the limit of a sequence of inhomogeneous pseudoholomorphic quilted maps does not create a figure-eight bubble ([WW15b]). Therefore, the usual Sard-Smale theorem can be used to prove transversality. The upshot is that for generic choices of Floer data compatible with the choices made for puncture disks involved in the definition of wrapped Fukaya categories, these moduli spaces ¯1,k,l((x,y);x,γ,(x+,y+),y)\bar{\mathcal{R}}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}) are compact smooth manifolds with corners of expected dimension

(7.13) k1+l1+deg((x,y))deg((x+,y+))deg(α)deg(xi)deg(yj),k-1+l-1+\deg((x^{-},y^{-}))-\deg((x^{+},y^{+}))-\deg(\alpha)-\sum\deg(x^{i})-\sum\deg(y^{j}),

and moreover each stratum is regular. Counting rigid elements in the zero dimensional moduli space ¯1,k,l((x,y);x,γ,(x+,y+),y)\bar{\mathcal{R}}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}) gives rise to the desired map (7.10).

We then extend the construction to higher orders. For this purpose, we consider the quilted surface S¯d,k,l\underline{S}^{d,k,l} which consists of two patches S0d,k,S1d,lS^{d,k}_{0},S^{d,l}_{1}, where S0d,kS^{d,k}_{0} is a disk with (k+d+2)(k+d+2) boundary punctures z0+,z0,z0p1,,z0pd,z01,,z0kz_{0}^{+},z_{0}^{-},z_{0}^{p_{1}},\cdots,z_{0}^{p_{d}},z_{0}^{1},\cdots,z_{0}^{k}, and S1d,lS^{d,l}_{1} is a disk with (l+d+2)(l+d+2) boundary punctures z1+,z1,z1p1,,z1pd,z11,,z1kz_{1}^{+},z_{1}^{-},z_{1}^{p_{1}},\cdots,z_{1}^{p_{d}},z_{1}^{1},\cdots,z_{1}^{k}. After seaming these two patches together, the strip-like end near z0piz_{0}^{p_{i}} and the one near z1piz_{1}^{p_{i}} together form a quilted cylindrical end.

Consider the moduli spaces d,k,l((x,y);x,γd,,γ1,(x+,y+),y)\mathcal{R}^{d,k,l}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y}) of inhomogeneous pseudoholomorphic quilted maps with appropriate boundary conditions and asymptotic convergence conditions. These are similar to that in Figure 7.1, but now there are also punctures on the seam which have appropriate asymptotic convergence conditions to generalized chords γ\gamma’s.

By a standard argument using Gromov compactness theorem and the maximum principle, we may prove that the Gromov bordification

(7.14) ¯d,k,l((x,y);x,γd,,γ1,(x+,y+),y)\bar{\mathcal{R}}^{d,k,l}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y})

is compact. Thus it is possible to count rigid elements therein, which gives rise to multilinear maps

(7.15) nk|d|l:CW(d1,d)CW(0,1)hom(CW(Lk1,Lk)CW(L0,L1)CW(L0,0,Ll)CW(Ll1,Ll)CW(L0,L1),CW(Lk,d,L0))\begin{split}n^{k|d|l}&:CW^{*}(\mathcal{L}_{d-1},\mathcal{L}_{d})\otimes\cdots\otimes CW^{*}(\mathcal{L}_{0},\mathcal{L}_{1})\to\\ &\hom(CW^{*}(L_{k-1},L_{k})\otimes\cdots\otimes CW^{*}(L_{0},L_{1})\otimes CW^{*}(L_{0},\mathcal{L}_{0},L^{\prime}_{l})\\ &\otimes CW^{*}(L^{\prime}_{l-1},L^{\prime}_{l})\otimes\cdots CW^{*}(L^{\prime}_{0},L^{\prime}_{1}),CW^{*}(L_{k},\mathcal{L}_{d},L^{\prime}_{0}))\end{split}
Lemma 7.4.

The multilinear maps {nk|d|l}\{n^{k|d|l}\} satisfy the AA_{\infty}-functor equations for the AA_{\infty}-functor (7.1). More concretely, for varying k,lk,l and testing objects LiL_{i} and LjL^{\prime}_{j}, the multilinear maps nk|d|ln^{k|d|l} define for each dd-tuple of composable Floer cochains in 𝒲(M×N)\mathcal{W}(M^{-}\times N) a pre-bimodule homomorphism Φ0Φd\Phi_{\mathcal{L}_{0}}\to\Phi_{\mathcal{L}_{d}}; moreover, the assignments of pre-bimodule homomorphisms for dd-tuples of composable Floer cochains satisfy the AA_{\infty}-functor equations.

Proof.

To verify that the multilinear maps nk|d|ln^{k|d|l} satisfy the desired AA_{\infty}-equations, we look at the codimension one boundary strata of the moduli space (7.14). It is covered by a union of the following products of moduli spaces:

(7.16) ¯d,k,l((x,y);x,γd,,γ1,(x+,y+),y)1ik(x,x′′′,x′′)=xxnew¯i+1(xnew,x′′′)ׯd,ki+1,l((x,y);x,xnew,x′′γd,,γ1,(x+,y+),y)1jl(y,y′′′,y′′)=yynew¯j+1(ynew,y′′′)ׯd,k,lj+1((x,y);x,γd,,γ1,(x+,y+),y,ynew,y′′)0d1dk+k′′=k,l+l′′=l(x,x′′)=x,(y,y′′)=y¯d2,k′′,l′′((x,y);x′′,γd,,γd1+1,(x1+,y1+),y′′)ׯd1,k,l((x1+,y1+);x,γd1,,γ1,(x+,y+),y)d1+d2=d+10sd1γnew¯d2+1(γs+d2,,γs+1,γnew)ׯd1,k,l((x,y);x,γd,,γs+d2+1,γnew,γs,,γ1,(x+,y+)).\begin{split}&\partial\bar{\mathcal{R}}^{d,k,l}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y})\\ \cong&\coprod_{\begin{subarray}{c}1\leq i\leq k\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime\prime},\vec{x}^{\prime\prime})=\vec{x}\end{subarray}}\coprod_{x_{new}}\bar{\mathcal{M}}_{i+1}(x_{new},\vec{x}^{\prime\prime\prime})\\ &\times\bar{\mathcal{R}}^{d,k-i+1,l}((x^{-},y^{-});\vec{x}^{\prime},x_{new},\vec{x}^{\prime\prime}\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y})\\ &\cup\coprod_{\begin{subarray}{c}1\leq j\leq l\\ (\vec{y}^{\prime},\vec{y}^{\prime\prime\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{y_{new}}\bar{\mathcal{M}}_{j+1}(y_{new},\vec{y}^{\prime\prime\prime})\\ &\times\bar{\mathcal{R}}^{d,k,l-j+1}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y}^{\prime},y_{new},\vec{y}^{\prime\prime})\\ &\cup\coprod_{0\leq d_{1}\leq d}\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x},(\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\bar{\mathcal{R}}^{d_{2},k^{\prime\prime},l^{\prime\prime}}((x^{-},y^{-});\vec{x}^{\prime\prime},\gamma^{d},\cdots,\gamma^{d_{1}+1},(x^{+}_{1},y^{+}_{1}),\vec{y}^{\prime\prime})\\ &\times\bar{\mathcal{R}}^{d_{1},k^{\prime},l^{\prime}}((x^{+}_{1},y^{+}_{1});\vec{x}^{\prime},\gamma^{d_{1}},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y}^{\prime})\\ &\cup\coprod_{\begin{subarray}{c}d_{1}+d_{2}=d+1\\ 0\leq s\leq d_{1}\end{subarray}}\coprod_{\gamma_{new}}\bar{\mathcal{M}}_{d_{2}+1}(\gamma^{s+d_{2}},\cdots,\gamma^{s+1},\gamma_{new})\\ &\times\bar{\mathcal{R}}^{d_{1},k,l}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{s+d_{2}+1},\gamma_{new},\gamma^{s},\cdots,\gamma^{1},(x^{+},y^{+})).\end{split}

The above description of the codimension-one boundary strata of

¯d,k,l((x,y);x,γd,,γ1,(x+,y+),y),\bar{\mathcal{R}}^{d,k,l}((x^{-},y^{-});\vec{x},\gamma^{d},\cdots,\gamma^{1},(x^{+},y^{+}),\vec{y}),

similar to that in (7.16), implies the following series of identities which the operations nk|d|ln^{k|d|l} satisfy:

(7.17) m1nk|d|l([x],[γd],,[γ1],[x+,y+],[y])=nki+1|d|l([x],mi([x′′′]),[x′′],[γd],,[γ1],[x+,y+],[y])+nk|d|lj+1([x],[γd],,[γ1],[x+,y+],[y],mj([y′′′]),[y′′])+nk′′|d2|l′′([x′′],[γd],,[γd1+1],nk|d1|l([x],[γd1],,[γ1],[x+,y+],[y]),[y′′])+nk|d1|l([x,[γd],,[γs+d2+1],md2([γs+d2],,[γs+1]),[γs],,[γ1],[x+,y+])\begin{split}&m^{1}\circ n^{k|d|l}([\vec{x}],[\gamma^{d}],\cdots,[\gamma^{1}],[x^{+},y^{+}],[\vec{y}])\\ &=\sum n^{k-i+1|d|l}([\vec{x}^{\prime}],m^{i}([\vec{x}^{\prime\prime\prime}]),[\vec{x}^{\prime\prime}],[\gamma^{d}],\cdots,[\gamma^{1}],[x^{+},y^{+}],[\vec{y}])\\ &+\sum n^{k|d|l-j+1}([\vec{x}],[\gamma^{d}],\cdots,[\gamma^{1}],[x^{+},y^{+}],[\vec{y}^{\prime}],m^{j}([\vec{y}^{\prime\prime\prime}]),[\vec{y}^{\prime\prime}])\\ &+\sum n^{k^{\prime\prime}|d_{2}|l^{\prime\prime}}([\vec{x}^{\prime\prime}],[\gamma^{d}],\cdots,[\gamma_{d_{1}+1}],\\ &n^{k^{\prime}|d_{1}|l^{\prime}}([\vec{x}^{\prime}],[\gamma^{d_{1}}],\cdots,[\gamma^{1}],[x^{+},y^{+}],[\vec{y}^{\prime}]),[\vec{y}^{\prime\prime}])\\ &+\sum n^{k|d_{1}|l}([\vec{x},[\gamma^{d}],\cdots,[\gamma^{s+d_{2}+1}],\\ &m^{d_{2}}([\gamma^{s+d_{2}}],\cdots,[\gamma^{s+1}]),[\gamma^{s}],\cdots,[\gamma^{1}],[x^{+},y^{+}])\end{split}

Note in particular that the term nk′′|d2|l′′(,nk|d1|l(),)n^{k^{\prime\prime}|d_{2}|l^{\prime\prime}}(\cdots,n^{k^{\prime}|d_{1}|l^{\prime}}(\cdots),\cdots) accounts for the second order structure map m2m^{2} in the dg-category (𝒲(M),𝒲(N))bimod(\mathcal{W}(M),\mathcal{W}(N))^{bimod} of AA_{\infty}-bimodules over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)). Rewriting the above identity in a suitable way, combining the quilted Floer differentials and Floer differentials in both 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N) into the first order structure map in (𝒲(M),𝒲(N))bimod(\mathcal{W}(M),\mathcal{W}(N))^{bimod}, we obtain the desired AA_{\infty}-functor equations for the assignment Φ\mathcal{L}\mapsto\Phi_{\mathcal{L}}, from 𝒲(M×N)\mathcal{W}(M^{-}\times N) to the dg-category of AA_{\infty}-bimodules over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)).

We have thus completed the construction of the AA_{\infty}-functor (7.1). As mentioned before, by evaluation and the Yoneda embedding we obtain the module-valued functor (7.3) for each admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N.

Remark 7.5.

Note that in our construction, a Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N gives rise to a bimodule over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)), rather than (𝒲(M),𝒲(N))(\mathcal{W}(M^{-}),\mathcal{W}(N)). The sign is important and is due to the fact that the quilted inhomogeneous pseudoholomorphic maps are defined with respect to the almost complex structure with the correct sign, forcing the boundary conditions to be ordered in the desired way demanded by the structure of a bimodule over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)).

7.3. The quilted Floer bimodule for Lagrangian immersions

Now we would like to extend the AA_{\infty}-bimodule Φ()\Phi(\mathcal{L}) over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)) to an AA_{\infty}-bimodule over (𝒲(M),𝒲im(N))(\mathcal{W}(M),\mathcal{W}_{im}(N)). The construction can be viewed as a generalization of that in subsection 7.1.

Proposition 7.6.

The AA_{\infty}-bimodule Φ()\Phi(\mathcal{L}) over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)) extends to an AA_{\infty}-bimodule over (𝒲(M),𝒲im(N))(\mathcal{W}(M),\mathcal{W}_{im}(N)). That is, there is a AA_{\infty}-bimodule over (𝒲(M),𝒲im(N))(\mathcal{W}(M),\mathcal{W}_{im}(N)), which composed with the pullback

j:𝒲im(N)lmod𝒲(N)lmodj^{*}:\mathcal{W}_{im}(N)^{l-mod}\to\mathcal{W}(N)^{l-mod}

agrees with Φ()\Phi(\mathcal{L}), up to homotopy equivalence of bimodules. Here j:𝒲(N)𝒲im(N)j:\mathcal{W}(N)\to\mathcal{W}_{im}(N) is the quasi-embedding (5.28).

The construction of this AA_{\infty}-bimodule structure involves moduli spaces similar to k,l((x,y);(x+,y+),x,y)\mathcal{R}^{k,l}((x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) as (7.4), where now some of the conditions for the inhomogeneous pseudoholomorphic quilted maps are slightly modified. First, the Lagrangian submanifolds LjL^{\prime}_{j} as boundary conditions are replaced by the images of Lagrangian immersions ιj:LjN\iota_{j}:L^{\prime}_{j}\to N. Second, we need to include some additional information: the switching labels αj\alpha_{j} for the Lagrangian immersions ιj:LjN\iota_{j}:L^{\prime}_{j}\to N, and the relative homotopy class β\beta of the map. Third, the Hamiltonian chords yjy_{j} from Lj1L^{\prime}_{j-1} to LjL^{\prime}_{j} are now replaced by appropriate generators for the wrapped Floer cochain space CW((Lj1,ιj1),(Lj,ιj);H)CW^{*}((L^{\prime}_{j-1},\iota_{j-1}),(L^{\prime}_{j},\iota_{j});H), which are either critical points of an auxiliary Morse function on the intersection components, or non-constant time-one Hamiltonian chords contained in the cylindrical end of NN. Denoting these generators by the same letters y=(y1,,yl)\vec{y}=(y_{1},\cdots,y_{l}), we write the corresponding moduli space by

k,l(α,β;(x,y);(x+,y+),x,y),\mathcal{R}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}),

with the additional information included.

There is a natural compactification

¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y})

which is obtained from k,l(α,β;(x,y);(x+,y+),x,y)\mathcal{R}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) by adding stable broken inhomogeneous pseudoholomorphic disks in MM and stable broken pearly tree maps in NN with boundary on the image of one of the Lagrangian immersions ιj:LjN\iota_{j}:L^{\prime}_{j}\to N, as well as broken quilted maps. As an analogue to (7.5), the codimension-one boundary strata can thus be described as follows:

(7.18) 1ik(x,x′′)=xxnew¯1,ki+1,l((x,y);x,xnew,γ,(x+,y+),y)ׯi+1(xnew,x′′)1jl(y,y′′)=yynew¯1,l,lj+1((x,y);x,γ,(x+,y+),y)ׯj+1(ynew,y′′)γ1¯1,k,l((x,y);x,γ1,(x+,y+),y)ׯ(γ1,γ)k+k′′=k,l+l′′=l(x,x′′)=x,(y,y′′)=y(x1+,y1+)¯1,k,l((x,y);x,γ,(x1+,y1+),y)ׯ0,k′′,l′′((x1+,y1+);x′′,(x+,y+),y′′)k+k′′=k,l+l′′=l(x,x′′)=x,(y,y′′)=y(x1+,y1+)¯0,k,l((x,y);x,(x1+,y1+),y)ׯ1,k′′,l′′((x1+,y1+);x′′,γ,(x+,y+),y′′).\begin{split}&\coprod_{\begin{subarray}{c}1\leq i\leq k\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x}\end{subarray}}\coprod_{x_{new}}\bar{\mathcal{R}}^{1,k-i+1,l}((x^{-},y^{-});\vec{x}^{\prime},x_{new},\gamma,(x^{+},y^{+}),\vec{y})\times\bar{\mathcal{M}}_{i+1}(x_{new},\vec{x}^{\prime\prime})\\ &\cup\coprod_{\begin{subarray}{c}1\leq j\leq l\\ (\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{y_{new}}\bar{\mathcal{R}}^{1,l,l-j+1}((x^{-},y^{-});\vec{x},\gamma,(x^{+},y^{+}),\vec{y}^{\prime})\times\bar{\mathcal{M}}_{j+1}(y_{new},\vec{y}^{\prime\prime})\\ &\cup\coprod_{\gamma_{1}}\bar{\mathcal{R}}^{1,k,l}((x^{-},y^{-});\vec{x},\gamma_{1},(x^{+},y^{+}),\vec{y})\times\bar{\mathcal{M}}(\gamma_{1},\gamma)\\ &\cup\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x},(\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{(x^{+}_{1},y^{+}_{1})}\bar{\mathcal{R}}^{1,k^{\prime},l^{\prime}}((x^{-},y^{-});\vec{x}^{\prime},\gamma,(x^{+}_{1},y^{+}_{1}),\vec{y}^{\prime})\\ &\times\bar{\mathcal{R}}^{0,k^{\prime\prime},l^{\prime\prime}}((x^{+}_{1},y^{+}_{1});\vec{x}^{\prime\prime},(x^{+},y^{+}),\vec{y}^{\prime\prime})\\ &\cup\coprod_{\begin{subarray}{c}k^{\prime}+k^{\prime\prime}=k,l^{\prime}+l^{\prime\prime}=l\\ (\vec{x}^{\prime},\vec{x}^{\prime\prime})=\vec{x},(\vec{y}^{\prime},\vec{y}^{\prime\prime})=\vec{y}\end{subarray}}\coprod_{(x^{+}_{1},y^{+}_{1})}\bar{\mathcal{R}}^{0,k^{\prime},l^{\prime}}((x^{-},y^{-});\vec{x}^{\prime},(x^{+}_{1},y^{+}_{1}),\vec{y}^{\prime})\\ &\times\bar{\mathcal{R}}^{1,k^{\prime\prime},l^{\prime\prime}}((x^{+}_{1},y^{+}_{1});\vec{x}^{\prime\prime},\gamma,(x^{+},y^{+}),\vec{y}^{\prime\prime}).\end{split}

Following the same lines as in the construction of Kuranishi structures on the moduli space of inhomogeneous pseudoholomorphic disks that are used to constructe curved AA_{\infty}-structures for the immersed wrapped Fukaya category introduced in section 4, we can also construct Kuranishi structures on these moduli spaces. This compactification is obtained from k,l(α,β;(x,y);(x+,y+),x,y)\mathcal{R}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) by adding broken inhomogeneous pseudoholomorphic punctured disks in both MM and NN, as well as broken quilted maps. In particular, those broken inhomogeneous pseudoholomorphic disks in NN with boundary conditions given by the Lagrangian immersions ιj:LjN\iota_{j}:L^{\prime}_{j}\to N form moduli spaces which carry Kuranishi structures as discussed before. Thus, by the inductive nature of the construction of Kuranishi structures, it remains to build Kuranishi charts on codimension zero strata of various moduli spaces k,l(α,β;(x,y);(x+,y+),x,y)\mathcal{R}^{k^{\prime},l^{\prime}}(\vec{\alpha}^{\prime},\beta^{\prime};(x^{-},y^{-});(x^{+},y^{+}),\vec{x}^{\prime},\vec{y}^{\prime}) for kk,llk^{\prime}\leq k,l^{\prime}\leq l. That is, we need to build Kuranishi charts over the locus in the moduli spaces whose elements have smooth domains.

Suppose we are given an inhomogeneous pseudoholomorphic quilted map σ=(S¯,(u,v))\sigma=(\underline{S},(u,v)) in k,l(α,β;(x,y);(x+,y+),x,y)\mathcal{R}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}), where S¯\underline{S} is the moduli parameter of the underlying quilted surface, and (u,v)(u,v) is map to the manifold pair (M,N)(M,N). The linearized operator Dσ¯D_{\sigma}\bar{\partial} of the inhomogeneous Cauchy-Riemann equations for σ\sigma is Fredholm by non-degeneracy assumption. Thus the Fredholm complex, defined with respect to appropriate Sobolev norm W1,pW^{1,p}, has finite-dimensional reductions. We choose an obstruction space EσE_{\sigma}, which is a finite-dimensional subspace of Ω0,1(S¯;uTM×vTN)\Omega^{0,1}(\underline{S};u^{*}TM\times v^{*}TN) such that for each VEσV\in E_{\sigma}, the support of VV is contained in a closed subset of the domain S¯\underline{S} away from the boundary components and the seam. Then, following the lines in section 4.7, we can build a Kuranishi chart (Uσ,Eσ,sσ,ψσ,Γσ={1})(U_{\sigma},E_{\sigma},s_{\sigma},\psi_{\sigma},\Gamma_{\sigma}=\{1\}) at this point σ\sigma.

In order to modify these charts for all elements in the moduli space so that they together define a Kuranishi structure, we need to make sure that these charts glue well with Kuranishi charts for inhomogeneous pseudoholomorphic disks in MM and NN. To formulate this, consider the following moduli spaces

  1. (i)

    k,l(α,β;(x,y);(x+,y+),x,y)\mathcal{R}^{k^{\prime},l^{\prime}}(\vec{\alpha}^{\prime},\beta^{\prime};(x^{-},y^{-});(x^{+},y^{+}),\vec{x}^{\prime},\vec{y}^{\prime}),

  2. (ii)

    k′′+1(xnew,x′′)\mathcal{M}_{k^{\prime\prime}+1}(x_{new},\vec{x}^{\prime\prime}),

  3. (iii)

    l′′+1(α′′,β′′;ynew,y′′)\mathcal{M}_{l^{\prime\prime}+1}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{new},\vec{y}^{\prime\prime}),

where x=(x1,,xk)\vec{x}=(x_{1},\cdots,x_{k}), and x=(x1,,xi,xnew,xi+k′′+1,,xk)\vec{x}^{\prime}=(x_{1},\cdots,x_{i},x_{new},x_{i+k^{\prime\prime}+1},\cdots,x_{k}), and x′′=(xi+1,,xi+k′′)\vec{x}^{\prime\prime}=(x_{i+1},\cdots,x_{i+k^{\prime\prime}}); similarly for the yy’s. The union of the product moduli spaces

(7.19) ixnewjynewk,l(α,β;(x,y);(x+,y+),x,y)×k′′+1(xnew,x′′)×l′′+1(ynew,y′′)\coprod_{\begin{subarray}{c}i\\ x_{new}\end{subarray}}\coprod_{\begin{subarray}{c}j\\ y_{new}\end{subarray}}\mathcal{R}^{k^{\prime},l^{\prime}}(\vec{\alpha}^{\prime},\beta^{\prime};(x^{-},y^{-});(x^{+},y^{+}),\vec{x}^{\prime},\vec{y}^{\prime})\times\mathcal{M}_{k^{\prime\prime}+1}(x_{new},\vec{x}^{\prime\prime})\times\mathcal{M}_{l^{\prime\prime}+1}(y_{new},\vec{y}^{\prime\prime})

is a boundary stratum of the compactification ¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}), and under the gluing maps, it can be thickened to a neighborhood of the boundary in the compactification. The gluing happens near the ends for the quilted surface and respectively the punctured disks, over which the quilted map and respectively the inhomogeneous pseudoholomorphic map converge to xnewx_{new}; similarly for the ends with convergence condition ynewy_{new}. Since the gluing construction is local, the process is the same as gluing inhomogeneous pseudoholomorphic disks along strip-like ends. Since the various obstruction spaces are chosen such that the vectors have compact support away from the boundary of the disks and quilted surfaces, the obstruction spaces also glue well under the gluing map. Thus we may apply the process in section 4.7 to modify the Kuranishi charts so that they form a Kuranishi structure on the moduli space ¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}).

Moreover, such construction extends to boundary strata of higher codimension, by an inductive argument. This implies that we can construct fiber product Kuranishi structures on (7.18), which are are compatible with the Kuranishi structure on ¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}). This proves:

Proposition 7.7.

There exists an oriented Kuranishi structure on every such moduli space ¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}), which is compatible with the fiber product Kuranishi structures on (7.18). That is, the restriction of the Kuranishi structure on ¯k,l(α,β;(x,y);(x+,y+),x,y)\bar{\mathcal{R}}^{k,l}(\vec{\alpha},\beta;(x^{-},y^{-});(x^{+},y^{+}),\vec{x},\vec{y}) to (7.18) agrees with the fiber product Kuranishi structure.

Choosing single-valued multisections for these Kuranishi structures gives rise to virtual fundamental chains on these moduli spaces, using which the desired AA_{\infty}-bimodule structure maps are defined.

Corollary 7.8.

The AA_{\infty}-bimodule Φ()\Phi(\mathcal{L}) over (𝒲(M),𝒲(N))(\mathcal{W}(M),\mathcal{W}(N)) extends to an AA_{\infty}-bimodule over (𝒲(M),𝒲im(N))(\mathcal{W}(M),\mathcal{W}_{im}(N)). That is, there is an AA_{\infty}-bimodule over (𝒲(M),𝒲im(N))(\mathcal{W}(M),\mathcal{W}_{im}(N)), which composed with the pullback

j:𝒲im(N)lmod𝒲(N)lmodj^{*}:\mathcal{W}_{im}(N)^{l-mod}\to\mathcal{W}(N)^{l-mod}

agrees with Φ()\Phi(\mathcal{L}), up to homotopy of bimodules. Here j:𝒲(N)𝒲im(N)j:\mathcal{W}(N)\to\mathcal{W}_{im}(N) is the quasi-embedding (5.28).

Proof.

Such extension is presented above. Thus the only thing that we need to prove is that the pullback by jj^{*} agrees with Φ()\Phi(\mathcal{L}) up to homotopy. One way to prove this is that we apply virtual techniques to construction virtual fundamental chains on the moduli space of inhomogeneous pseudoholomorphic quilted maps (7.4), regarding all Lagrangian submanifolds as Lagrangian immersions. If we stick with classical transversality methods for those moduli spaces, the other way is a straightforward analogue of the proof of Proposition 5.22. As there is nothing essentially new, we leave the details to the interested reader.

By a parallel argument, we can also extend the the module-valued functor (7.3) to one with values in category of modules over the immersed wrapped Fukaya category:

Proposition 7.9.

There is a canonical extension of the AA_{\infty}-functor (7.3) to an AA_{\infty}-functor

(7.20) Φ:𝒲(M)𝒲im(N)lmod\Phi_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}_{im}(N)^{l-mod}

to the AA_{\infty}-category of left AA_{\infty}-modules over 𝒲im(N)\mathcal{W}_{im}(N). That is, the composition by the pullback

j:𝒲im(N)lmod𝒲(N)lmodj^{*}:\mathcal{W}_{im}(N)^{l-mod}\to\mathcal{W}(N)^{l-mod}

agrees with (7.3), up to homotopy.

Of course, this also follows from the process of converting a bimodule to a module-valued functor.

7.4. Geometric composition of Lagrangian correspondences

In order to obtain a functor that takes value in the actual immersed wrapped Fukaya category 𝒲im(N)\mathcal{W}_{im}(N) instead of the category of modules over it, we need to prove that the AA_{\infty}-functor (7.20) is representable, in the sense of [Fuk02]. It has been long noted that the geometric compositions of Lagrangian correspondences are natural candidates for the objects representing the modules defined above. A good reference for the definition and basic properties of geometric compositions is [WW12], in the case where the Hamiltonian perturbation is not present.

There are two issues in proving representability of (7.20). First, representability does not always hold if we only consider embedded Lagrangian submanifolds, as geometric compositions of Lagrangian correspondences are generally Lagrangian immersions. This is why we must go to the immersed wrapped Fukaya category and consider the AA_{\infty}-functor (7.20) instead of (7.3). Second, the geometric composition of Lagrangian correspondences might not have good geometric properties, thus not a priori admissible for wrapped Floer theory on the nose. For this, we must imposed further conditions so that they are admissible in the immersed wrapped Fukaya category.

Let us first recall the definition of geometric composition of Lagrangian correspondences. Given an admissible Lagrangian submanifold LML\subset M and an admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N, in a generic situation the fiber product over the graph Γ(ψHM)M×M\Gamma(\psi_{H_{M}})\subset M^{-}\times M over the Hamiltonian symplectomorphism ϕHM\phi_{H_{M}}

L=L×Γ(ϕHM)L^{\prime}=L\times_{\Gamma(\phi_{H_{M}})}\mathcal{L}

is a smooth submanifold of M×M×NM\times M\times N. The composition of the embedding LM×M×NL^{\prime}\subset M\times M\times N with the projection M×M×NNM\times M\times N\to N is a Lagrangian immersion:

ι:L×Γ(ϕHM)N.\iota:L\times_{\Gamma(\phi_{H_{M}})}\mathcal{L}\to N.

We call this Lagrangian immersion the geometric composition of LL with \mathcal{L}, under the perturbation by the Hamiltonian flow of HMH_{M}. By abuse of name, we sometimes also call the image of ι\iota the geometric composition for simplicity, and often denote it by LHML\circ_{H_{M}}\mathcal{L}.

If the geometric composition happens to be (properly) embedded, it comes with a natural choice of a primitive, which makes it an exact Lagrangian submanifold of NN. Because of the presence of the Hamiltonian perturbation, this primitive is slightly different from the naive sum, but instead takes the following form:

(7.21) g=f+F(ϕHM×idN)+iXHMλMg=f+F\circ(\phi_{H_{M}}\times id_{N})+i_{X_{H_{M}}}\lambda_{M}

where XHMX_{H_{M}} is the Hamiltonian vector field of HMH_{M}, thought of as a Hamiltonian pulled back to M×NM^{-}\times N. This formula is calculated in [Gao17], which follows directly from the formula for the change of the primitive for an exact Lagrangian submanifold under a Hamiltonian isotopy, Lemma 3.2.

Depending on the geometry of LL and \mathcal{L}, the geometric composition might be or not be cylindrical, even if it is embedded. Therefore a proof of its admissibility in wrapped Floer theory is completely necessary. This is done in [Gao17] in the case where the geometric composition is properly embedded, with the above choice of the primitive. Technically, there we only proved that the wrapped Floer differential converges (and is in fact finite), but the same argument can be utilized to prove that higher order structure maps are also finite.

In general, the geometric composition ι:LHMN\iota:L\circ_{H_{M}}\mathcal{L}\to N is not an embedding, but we still expect it to have some favorable properties. This primitive gg still makes ι\iota an ”exact” Lagrangian immersion, in the sense that ιλN=dg\iota^{*}\lambda_{N}=dg.

There is also the notion of geometric composition in the usual sense. Instead of taking the fiber product over the graph Γ(ϕHM)\Gamma(\phi_{H_{M}}), we take the fiber product over the diagonal ΔM\Delta_{M}. The geometric composition LL\circ\mathcal{L} is the map

ι:L×ΔMN.\iota:L\times_{\Delta_{M}}\mathcal{L}\to N.

Generically this is a Lagrangian immersion, and is also exact in the generalized sense, where the primitive is

h=f+F.h=f+F.

In [Gao17], we proved well-definedness of wrapped Floer cohomology in the case where the geometric composition is properly embedded. The argument can be generalized to well-definedness of AA_{\infty}-structure maps of all orders. However, in this paper we shall take a slightly different point of view, in order to make our construction more functorial and canonical. In [Gao17], we considered the geometric composition under the large perturbation LHML\circ_{H_{M}}\mathcal{L}, and proved an isomorphism of Floer cohomology groups. While that definition is natural as one can find a natural one-to-one correspondence between the generators, we find it better to work with the geometric composition in the usual sense, when the whole categorical structure is in concern. Of course, the left-module structures associated to the geometric composition under large perturbation and the geometric composition in the usual sense are homotopy equivalent, so the essential difference is minor.

For the geometric composition of Lagrangian correspondences LL\circ\mathcal{L} to have well-defined wrapped Floer theory in general, we must have good control of the behavior of inhomogeneous pseudoholomorphic disks(modeled as stable pearly tree maps) bounded by the image of the geometric composition. For this purpose, we need to make sure that the geometry of LL\circ\mathcal{L} at infinity does not behave too badly. Thus it is natural to introduce the following assumption.

Assumption 7.10.

For the Lagrangian submanifold LL in consideration, the geometric composition LL\circ\mathcal{L} is a proper Lagrangian immersion with transverse or clean self-intersections, which is cylindrical in the generalized sense for a Lagrangian immersion.

When defining the wrapped Fukaya category 𝒲(M)\mathcal{W}(M), we have to specify a class of Lagrangian submanifolds as objects. Since Assumption 7.10 is generic, it is possible for us to choose a countable collection of Lagrangian submanifolds as objects of the wrapped Fukaya category 𝒲(M)\mathcal{W}(M), such that for every LL in this collection, Assumption 7.10 is satisfied. Then it follows almost from the definition that:

Proposition 7.11.

Under Assumption 7.10, there is a well-defined curved AA_{\infty}-algebra for the geometric composition ι:L×ΔMN\iota:L\times_{\Delta_{M}}\mathcal{L}\to N, in the sense of immersed wrapped Floer theory discussed in sections 4 and 5, for every LL from the collection of objects of 𝒲(M)\mathcal{W}(M).

Compared to the cohomological result in [Gao17], which works with the geometric composition under large perturbation, the assumption of Proposition 7.11 is in fact simpler as we assume the geometric composition LL\circ\mathcal{L} to be cylindrical. As discussed before, it is automatically exact, so the results of sections 4 and 5.

7.5. Unobstructedness of the geometric composition

We have shown that if N\mathcal{L}\to N is proper and if Assumption 7.10 is satisfied, there is a curved AA_{\infty}-algebra structure on the wrapped Floer cochain space CW(L;HN)CW^{*}(L\circ\mathcal{L};H_{N}) for the geometric composition LL\circ\mathcal{L}. To make it into an object of the immersed wrapped Fukaya category, we must also prove that this curved AA_{\infty}-algebra is unobstructed. The main result of this subsection says that the geometric composition LL\circ\mathcal{L} is unobstructed with a canonical and unique choice of bounding cochain bb satisfying a distinguished property.

Theorem 7.12.

Suppose LL is a properly embedded exact cylindrical Lagrangian submanifold of MM, and M×N\mathcal{L}\subset M^{-}\times N is a properly embedded exact cylindrical Lagrangian correspondence between MM and NN, such that the projection N\mathcal{L}\to N is proper. Let LL\circ\mathcal{L} be their geometric composition. Then LL\circ\mathcal{L} is unobstructed in the sense of wrapped Floer theory, with a canonical and unique choice of bounding cochain bb determined by LL and \mathcal{L}, with the property that bb gives rise to non-curved deformations for both the quilted Floer module CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) and the curved AA_{\infty}-algebra CW(L)CW^{*}(L\circ\mathcal{L}).

In particular, (L,b)(L\circ\mathcal{L},b) becomes an object in the immersed wrapped Fukaya category 𝒲im(N)\mathcal{W}_{im}(N).

The idea to prove Theorem 7.12 is to use Lemma 2.15 to provide an algebraic argument for the existence and uniqueness of such a bounding cochain. For that purpose, we shall first equip the quilted wrapped Floer cochain space CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) with a curved AA_{\infty}-module structure.

Lemma 7.13.

There is a natural curved left AA_{\infty}-module structure on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) over the curved AA_{\infty}-algebra (CW(L),mk)(CW^{*}(L\circ\mathcal{L}),m^{k}).

The construction involves moduli spaces of inhomogeneous pseudoholomorphic quilted maps of the following kind.

Consider the quilted surface S¯mod\underline{S}^{mod} consisting of two patches, S0mod,S1modS_{0}^{mod},S_{1}^{mod}, each of which is a disk with one negative puncture ziz_{i}^{-} and one positive puncture zi+z_{i}^{+}. Given generalized chords (x0,y0),(x1,y1)(x^{0},y^{0}),(x^{1},y^{1}) for (L,,L)(L,\mathcal{L},L\circ\mathcal{L}), consider quilted maps u¯:S¯mod(M,N)\underline{u}:\underline{S}^{mod}\to(M,N) satisfying the following conditions

(7.22) {u0s+JM(u0tXHM(u))=0u0s+JM(u0tXHM(u))=0u0(s,0)Lu1(s,1)L(u0(s,1),u1(s,0))limsu0(s,)=x0()lims+u0(s,)=x1()limsu1(s,)=y0()lims+u1(s,)=y1()\begin{cases}\frac{\partial u_{0}}{\partial s}+J_{M}\circ(\frac{\partial u_{0}}{\partial t}-X_{H_{M}}(u))=0\\ \frac{\partial u_{0}}{\partial s}+J_{M}\circ(\frac{\partial u_{0}}{\partial t}-X_{H_{M}}(u))=0\\ u_{0}(s,0)\in L\\ u_{1}(s,1)\in L\circ\mathcal{L}\\ (u_{0}(s,1),u_{1}(s,0))\in\mathcal{L}\\ \lim_{s\to-\infty}u_{0}(s,\cdot)=x^{0}(\cdot)\\ \lim_{s\to+\infty}u_{0}(s,\cdot)=x^{1}(\cdot)\\ \lim_{s\to-\infty}u_{1}(s,\cdot)=y^{0}(\cdot)\\ \lim_{s\to+\infty}u_{1}(s,\cdot)=y^{1}(\cdot)\\ \end{cases}

We also need to specify some additional data such as the switching conditions, the lifting conditions and the homology classes that are introduced in section 4.5. Quotienting by translations, we obtain moduli spaces of solutions to the above equation with these additional requirements.

To define a curved AA_{\infty}-module structure, we also need to study inhomogeneous pseudoholomorphic quilted maps of the same kind, but with more punctures. Break the boundary of S1modS_{1}^{mod} that gets mapped to LL\circ\mathcal{L} into several pieces, namely replace S1modS_{1}^{mod} by S1mod,kS_{1}^{mod,k}, a disk with (k+2)(k+2)-boundary punctures, z1,z11,,z1k,z1+z_{1}^{-},z_{1}^{1},\cdots,z_{1}^{k},z_{1}^{+}. We denote this new quilted surface by S¯mod,k\underline{S}^{mod,k}.

Since the map u1u_{1} on the second patch of the quilted surface has boundary condition being an immersed Lagrangian submanifold, it is necessary to include additional data α,β,l\alpha,\beta,l, etc. that indicate switching conditions of the boundary lifting, homology classes of the maps and so on. These data are analogous to those discussed in sections 4 and 5, so we omit the details here.

There is a natural compactification of the moduli space of these quilted maps, which we denote by

𝒬¯k(α,β;(x,y);y1,,yk,(x+,y+)).\bar{\mathcal{Q}}_{k}(\alpha,\beta;(x^{-},y^{-});y^{1},\cdots,y^{k},(x^{+},y^{+})).

A typical element in the compactified moduli space is a broken inhomogeneous pseudoholomorphic quilted map {(ui,vi)}\{(u_{i},v_{i})\} with trees of inhomogeneous pseudoholomorphic disks in NN attached to the lower boundary components of the second patches viv_{i}’s of the broken quilted map. In general, we have the following description of the codimension-one boundary strata of the above moduli space:

(7.23) 𝒬¯k(α,β;(x,y);y1,,yk,(x+,y+))0ikα1α2=αβ1β2=β(x1+,y1+)𝒬¯i(α1,β1;(x,y);y1,,yi,(x1+,y1+))×𝒬¯ki(α2,β2;(x1+,y1+);yi+1,,yk,(x+,y+))k1+k2=k+11ik1α1α2=αβ1β2=βynew𝒬¯k1(α1,β1;(x,y);y1,,yk1,ynew,yi+k2+1,,yk,(x+,y+))ׯi+1(α2,β2;ynew,yk1+1,,yi+k2).\begin{split}&\partial\bar{\mathcal{Q}}_{k}(\alpha,\beta;(x^{-},y^{-});y^{1},\cdots,y^{k},(x^{+},y^{+}))\\ \cong&\coprod_{0\leq i\leq k}\coprod_{\begin{subarray}{c}\alpha_{1}\cup\alpha_{2}=\alpha\\ \beta_{1}\sharp\beta_{2}=\beta\end{subarray}}\coprod_{(x^{+}_{1},y^{+}_{1})}\bar{\mathcal{Q}}_{i}(\alpha_{1},\beta_{1};(x^{-},y^{-});y^{1},\cdots,y^{i},(x^{+}_{1},y^{+}_{1}))\\ &\times\bar{\mathcal{Q}}_{k-i}(\alpha_{2},\beta_{2};(x^{+}_{1},y^{+}_{1});y^{i+1},\cdots,y^{k},(x^{+},y^{+}))\\ \cup&\coprod_{\begin{subarray}{c}k_{1}+k_{2}=k+1\\ 1\leq i\leq k_{1}\end{subarray}}\coprod_{\begin{subarray}{c}\alpha_{1}\cup\alpha_{2}=\alpha\\ \beta_{1}\sharp\beta_{2}=\beta\end{subarray}}\coprod_{y_{new}}\bar{\mathcal{Q}}_{k_{1}}(\alpha_{1},\beta_{1};(x^{-},y^{-});y^{1},\cdots,y^{k_{1}},\\ &y_{new},y^{i+k_{2}+1},\cdots,y^{k},(x^{+},y^{+}))\\ &\times\bar{\mathcal{M}}_{i+1}(\alpha_{2},\beta_{2};y_{new},y^{k_{1}+1},\cdots,y^{i+k_{2}}).\end{split}

Since LL and \mathcal{L} are exact, there are no pseudoholomorphic disks bubbling off LL or \mathcal{L}. Thus the above fiber products (7.23) cover all the boundary strata of the compactification.

As usual, we can construct Kuranishi structures on these moduli spaces compatibly and use single-valued multisections to define virtual fundamental chains.

Proposition 7.14.

There exists an oriented Kuranishi structure on the moduli space

𝒬¯k(α,β;(x,y);y1,,yk,(x+,y+)),\bar{\mathcal{Q}}_{k}(\alpha,\beta;(x^{-},y^{-});y^{1},\cdots,y^{k},(x^{+},y^{+})),

such that the induced Kuranishi structures on the boundary strata are isomorphic to the fiber product Kuranishi structures on (7.23).

In addition, we may choose single-valued multisections on these moduli spaces, such that the multisection on 𝒬¯k(α,β;(x,y);y1,,yk,(x+,y+))\bar{\mathcal{Q}}_{k}(\alpha,\beta;(x^{-},y^{-});y^{1},\cdots,y^{k},(x^{+},y^{+})) is compatible with the fiber product multisections at the boundary strata (7.23).

The virtual fundamental chains on these moduli spaces associated to a coherent choice of multisections then define a curved AA_{\infty}-module structure on the quilted wrapped Floer cochain space CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) over the curved AA_{\infty}-algebra CW(L)CW^{*}(L\circ\mathcal{L}).

To apply the result in section 2.5, we also need to find natural filtrations for the curved AA_{\infty}-algebra CW(L)CW^{*}(L\circ\mathcal{L}) and the curved AA_{\infty}-module CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) over it. These filtrations are given by the symplectic action functional.

Lemma 7.15.

The action filtration on CW(L)CW^{*}(L\circ\mathcal{L}) defines a discrete filtration for the curved AA_{\infty}-algebra (CW(L),mk)(CW^{*}(L\circ\mathcal{L}),m^{k}).

The action filtration on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) defines a discrete filtration for the curved AA_{\infty}-module (CW(L,,L),nk)(CW^{*}(L,\mathcal{L},L\circ\mathcal{L}),n^{k}) is compatible with the action filtration for (CW(L),mk)(CW^{*}(L\circ\mathcal{L}),m^{k}).

Moreover, these filtrations are bounded above.

Proof.

The proof of the fact that action filtrations are compatible with the curved AA_{\infty}-algebra structure and the curved AA_{\infty}-module structure follows immediately from the action-energy relation.

To prove that the action filtration on CW(L)CW^{*}(L\circ\mathcal{L}) is discrete, we recall that the generators of a wrapped Floer cochain space consist of two kinds: first, critical points of auxiliary Morse functions on components of the self fiber product of the preimage of the immersion; second, non-constant Hamiltonian chords in the cylindrical ends together with lifting indices. There are finitely many critical points, and we can arrange the primitive and choose the auxiliary Morse functions carefully so that their actions are different. On the other hand, ecause the Hamiltonian is non-degenerate in the cylindrical end, these non-constant Hamiltonian chords are non-degenerate and have a discrete action spectrum. A similar argument applies to show that the action filtration on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) is discrete.

Compatibility follows from the action-energy relation applied to inhomogeneous pseudoholomorphic quilted maps in the moduli spaces 𝒬¯k(α,β;(x,y);y1,,yk,(x+,y+))\bar{\mathcal{Q}}_{k}(\alpha,\beta;(x^{-},y^{-});y^{1},\cdots,y^{k},(x^{+},y^{+})), which are used to define this curved AA_{\infty}-module structure.

The fact that these filtrations are bounded above follows from the definition of the wrapped Floer cochain space: there are only finitely many free generators which have positive action, as those infinitely many generators, the non-constant Hamiltonian chords in the cylindrical end, all have negative action.

To finish the proof of Theorem 7.12, we need to find a cyclic element for the curved AA_{\infty}-module CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) over the curved AA_{\infty}-algebra CW(L)CW^{*}(L\circ\mathcal{L}). Recall from sections 4.3 and 5.3 that in the wrapped Floer cochain space CW(L)CW^{*}(L\circ\mathcal{L}), we have a distinguished generator - the minimum of chosen Morse function ff on the diagonal component ΔL×ΔM\Delta_{L\times_{\Delta_{M}}\mathcal{L}} of the self fiber product (L×ΔM)×ι(L×ΔM)(L\times_{\Delta_{M}}\mathcal{L})\times_{\iota}(L\times_{\Delta_{M}}\mathcal{L}). This corresponds to the fundamental chain of the manifold L×ΔML\times_{\Delta_{M}}\mathcal{L} in the singular chain model. This generator is the homotopy unit for the curved AA_{\infty}-algebra (CW(L),mk)(CW^{*}(L\circ\mathcal{L}),m^{k}).

As graded \mathbb{Z}-modules, we have that CW(L,,L)CW(L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L})\cong CW^{*}(L\circ\mathcal{L}), given by a natural bijective correspondence between the sets of generators. This follows directly from the definition of the geometric composition. Under this correspondence, we get a distinguished element eLCW(L,,L)e_{L\circ\mathcal{L}}\in CW^{*}(L,\mathcal{L},L\circ\mathcal{L}), corresponding to the homotopy unit of CW(L)CW^{*}(L\circ\mathcal{L}).

Lemma 7.16.

The element eLCW(L,,L)e_{L\circ\mathcal{L}}\in CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) defined above is a cyclic element.

Proof.

Recall that a cylic element has to satisfy two conditions. First, the map

CW(L)CW(L,,L)CW^{*}(L\circ\mathcal{L})\to CW^{*}(L,\mathcal{L},L\circ\mathcal{L})

defined by

xn1(x;eL)x\mapsto n^{1}(x;e_{L\circ\mathcal{L}})

is an isomorphism of \mathbb{Z}-modules. Second, eLe_{L\circ\mathcal{L}} lies in F0F^{0}, and applying n0n^{0} to eLe_{L\circ\mathcal{L}} should strictly increase the action filtration.

The first condition follows from the fact that eLe_{L\circ\mathcal{L}} corresponds to the homotopy unit of CW(L)CW^{*}(L\circ\mathcal{L}). Multiplication with the homotopy unit yields a self map on CW(L)CW^{*}(L\circ\mathcal{L}), which can be written as an upper-triangular matrix with respect to a basis for CW(L)CW^{*}(L\circ\mathcal{L}) ordered in increasing action. Moreover, the diagonal entries of this upper-triangular matrix are all equal to the identity. Now we consider the basis for CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) which corresponds to the chosen basis for CW(L)CW^{*}(L\circ\mathcal{L}) under the natural one-to-one correspondence between generators: each generalized chord for (L,,L)(L,\mathcal{L},L\circ\mathcal{L}) corresponds to a unique Hamiltonian chord from LL\circ\mathcal{L} to itself (the same applies to critical points). This basis is also ordered in increasing action, so that the map xn1(x;eL)x\mapsto n^{1}(x;e_{L\circ\mathcal{L}}) can be written as an upper-triangular matrix whose diagonal entries are all equal to the ”identity”, where this ”identity” means the natural one-to-one correspondence between generators.

Now let us check the second condition. First, the element eLe_{L\circ\mathcal{L}} itself is a free generator of the quilted wrapped Floer cochain space CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}), as it corresponds to the homotopy unit of CW(L)CW^{*}(L\circ\mathcal{L}) under the natural one-to-one correspondence between generators. Moreover, we can choose the primitive carefully such that eLe_{L\circ\mathcal{L}} has zero action. To prove that n0n^{0} applied eLe_{L\circ\mathcal{L}} strictly increases the action, it suffices to prove that there are no constant inhomogeneous pseudoholomorphic quilted strips with input being eLe_{L\circ\mathcal{L}}. Such a constant quilted strip, if existed, would correspond to a constant inhomogeneous pseudoholomorphic strip with boundary on the image of LL\circ\mathcal{L} with input being the homotopy unit of CW(L)CW^{*}(L\circ\mathcal{L}). But there are no such constant strips, because the homotopy unit is the minimum of the chosen Morse function on the diagonal component ΔL×ΔM\Delta_{L\times_{\Delta_{M}}\mathcal{L}} of the self fiber product. As a consequence, n0(eL)n^{0}(e_{L\circ\mathcal{L}}) can be written as a linear combination of some generators of CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}), none of which is any critical point on the diagonal component of the self fiber product, or any non-constant Hamiltonian chord in the cylindrical end. Therefore n0(eL)n^{0}(e_{L\circ\mathcal{L}}) can be written as a linear combination of generators, which correspond to critical points on the switching components of the self fiber product of the Lagrangian immersion LL\circ\mathcal{L} under the natural one-to-one correspondence between the generators of CW(L)CW^{*}(L\circ\mathcal{L}) and those of CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}). In particular, applying n0n^{0} to eLe_{L\circ\mathcal{L}} must strictly increase the action, because it is defined by counting non-constant inhomogeneous pseudoholomorphic quilted maps, which have strictly positive energy.

By a purely algebraic argument using Lemma 2.15, the existence of a cyclic element implies the unobstructedness of the geometric composition, with a unique bounding cochain bb satisfying the following property:

Corollary 7.17.

There exists a unique (nilpotent) bounding chain bCW(L)b\in CW^{*}(L\circ\mathcal{L}) such that bFϵb\in F^{\epsilon} for some ϵ>0\epsilon>0, and

db(eL)=0,d^{b}(e_{L\circ\mathcal{L}})=0,

where db()=k=0nk(b,,b;)d^{b}(\cdot)=\sum_{k=0}^{\infty}n^{k}(b,\cdots,b;\cdot).

Thus the proof of Theorem 7.12 is complete.

Remark 7.18.

Note that CW(L)CW^{*}(L\circ\mathcal{L}) with respect to the undeformed structure maps is generally not a curved AA_{\infty}-module over itself, because of the non-vanishing of the curvature term m0m^{0}. However, the bb-deformed structure maps mk;bm^{k;b} make CW(L,b)CW^{*}(L\circ\mathcal{L},b) an AA_{\infty}-module over itself, as the bb-deformed AA_{\infty}-algebra is non-curved.

Nonetheless, the curved AA_{\infty}-module structure on the quilted wrapped Floer cochain space CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}) is essentially different from that on CW(L,b)CW^{*}(L\circ\mathcal{L},b) as a AA_{\infty}-module over itself, although the underlying \mathbb{Z}-modules are isomorphic. The quilted Floer-theoretic setup is essential for this curved AA_{\infty}-module structure to exist.

Concerning the wrapped Fukaya category, we have the following vanishing result of the bounding cochain bb for the geometric composition, in the case where it is in fact a proper exact Lagrangian embedding.

Proposition 7.19.

If the geometric composition ι:LN\iota:L\circ\mathcal{L}\to N is a proper exact cylindrical Lagrangian embedding, whose primitive (coming from the primitive for LL and that for \mathcal{L}) extends to a function on NN which is locally constant in the cylindrical end of NN, then the bounding cochain bb from Theorem 7.12 vanishes.

Sketch of proof.

Recall that the bounding cochain bb is the unique solution to the equation

(7.24) n0;b(eL)=knk(b,,b;eL)=0,n^{0;b}(e_{L\circ\mathcal{L}})=\sum_{k}n^{k}(b,\cdots,b;e_{L\circ\mathcal{L}})=0,

with the property that bF>0b\in F^{>0}. This implies that for this choice of bounding cochain bb, the map

(7.25) gc1:CW(L,,(L,b))CW(L,b)gc^{1}:CW^{*}(L,\mathcal{L},(L\circ\mathcal{L},b))\to CW^{*}(L\circ\mathcal{L},b)

is a cochain map with respect to the deformed differentials on both sides. Recall that the deformed differential on the quilted wrapped Floer cochain space is

n0;b(x)=k0nk(b,,bk times;x),n^{0;b}(x)=\sum_{k\geq 0}n^{k}(\underbrace{b,\cdots,b}_{k\text{ times}};x),

while the deformed differential on the wrapped Floer cochain space CW(L,b)CW^{*}(L\circ\mathcal{L},b) is

m1;b(x)=k0,k10mk(b,,bk0 times,x,b,,bk1 times).m^{1;b}(x)=\sum_{k_{0},k_{1}\geq 0}m^{k}(\underbrace{b,\cdots,b}_{k_{0}\text{ times}},x,\underbrace{b,\cdots,b}_{k_{1}\text{ times}}).

Because of the assumption that ι\iota is a proper embedding, 0 is a bounding cochain for LL\circ\mathcal{L}. We want to prove that, if we choose 0 as the bounding cochain for LL\circ\mathcal{L}, the cyclic element is closed under the undeformed differential on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}). The proof will be separated in the following three lemmas.

The first lemma is a general statement about the bounding cochain bb, where we do not assume the geometric composition LL\circ\mathcal{L} is an embedding.

Lemma 7.20.

The bounding cochain bb from Theorem 7.12 is supported only at the critical points, or non-constant Hamiltonian chords in the cylindrical end which have positive action.

Proof.

The statement follows immediately from the condition that bF>0b\in F^{>0}.

The second lemma lists some equivalent conditions for 0 to be the desired bounding cochain under the assumption that the geometric composition is an embedding.

Lemma 7.21.

Suppose that the geometric composition ι:LN\iota:L\circ\mathcal{L}\to N is a proper exact cylindrical Lagrangian embedding. Then the following three conditions are equivalent:

  1. (i)

    The cyclic element eLe_{L\circ\mathcal{L}} is closed under the undeformed differential on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}).

  2. (ii)

    There does not exist an inhomogeneous pseudoholomorphic quilted strip with boundary condition (L,,L)(L,\mathcal{L},L\circ\mathcal{L}), which converges to eLe_{L\circ\mathcal{L}} over the positive quilted end (as input).

  3. (iii)

    There is no figure eight bubble which asymptotically converges to a generalized chord for (L,,,L)(L,\mathcal{L},\mathcal{L},L) over the negative quilted end.

Proof.

This lemma is proved in [Gao17]. Let us briefly recall it here.

(ii) implies (i) by the definition of the quilted Floer module structure map on CW(L,,L)CW^{*}(L,\mathcal{L},L\circ\mathcal{L}).

(iii) implies (ii) by a strip-shrinking argument. Suppose that there is no figure eight bubble as in (iii). If there is an inhomogeneous pseudoholomorphic quilted strip converging to eLe_{L\circ\mathcal{L}} over the positive quilted end, then by shrinking it we get an inhomogeneous pseudoholomorphic strip in NN with boundary on LL\circ\mathcal{L}, which asymptotically converges to the unit of CW(L)CW^{*}(L\circ\mathcal{L}) over the positive end. This is certainly not possible.

Thus it remains to prove that (i) implies (iii). Suppose on the contrary that eLe_{L\circ\mathcal{L}} is not closed under the undeformed quilted Floer differential. Then there exists a non-trivial inhomogeneous pseudoholomorphic strip which asymptotically converges to eLe_{L\circ\mathcal{L}}. By shrinking it we get two possible kinds of configurations: first, an inhomogeneous pseudoholomorphic strip in NN with boundary on LL\circ\mathcal{L}, which asymptotically converges to the unit of CW(L)CW^{*}(L\circ\mathcal{L}); second, a figure eight bubble asymptotic to a generalized chord for (L,,,L)(L,\mathcal{L},\mathcal{L},L). The first kind does not exist as previously argued, and the second kind does not exist by the assumption of (iii).

Remark 7.22.

Note that we have used a strip-shrinking argument, but not for the purpose of establishing an isomorphism between the moduli space of inhomogeneous pseudoholomorphic quilted strips and the moduli space of inhomogeneous pseudoholomophic strips. Rather, we use it to prove the non-existence of certain inhomogeneous pseudoholomorphic quilted strips by the given condition on the primitive for the geometric composition LL\circ\mathcal{L}. That is, the moduli space is empty because of a geometric constraint; as a consequence, it is isomorphic to the empty set as Kuranishi spaces.

Thus, to prove that the cyclic element eLe_{L\circ\mathcal{L}} is closed under the undeformed quilted wrapped Floer differential, it suffices to prove condition (iii). That uses the third lemma stated as below, which adds the assumption that the primitive is locally constant.

Lemma 7.23.

Suppose that LL\circ\mathcal{L} is a proper exact cylindrical Lagrangian embedding and that the primitive for LL\circ\mathcal{L} extends to a locally constant function in the cylindrical end of NN. Then any Hamiltonian chord from LL\circ\mathcal{L} which lies outside the compact domain N0N_{0} has negative action.

Proof.

Since the primitive is locally constant in the cylindrical end, the action of any Hamiltonian chord yy from LL\circ\mathcal{L} to itself contained in the cylindrical end is

𝒜(y)=yλN+HN(y(t))dt=r2<0,\mathcal{A}(y)=-\int y^{*}\lambda_{N}+H_{N}(y(t))dt=-r^{2}<0,

if yy lies on the level set N×{r}\partial N\times\{r\}. ∎

Suppose condition (iii) of Lemma (7.21) does not hold under the assumption of Proposition 7.19. Note that there is a uniform positive lower bound for the energy of any figure eight bubble. Thus if there were such a figure eight bubble, it would converge to some generalized chord for (L,,,L)(L,\mathcal{L},\mathcal{L},L) of positive action ϵ>0\geq\epsilon>0, for some uniform constant ϵ>0\epsilon>0 that is independent of individual pseudoholomorphic quilted maps but depends only on the background geometry - Liouville structures, Lagrangian submanifolds, Hamiltonians and almost complex structures. Then this generalized chord corresponds to some generator yy of CW(L)CW^{*}(L\circ\mathcal{L}) which has positive action. By the lemma above, this generator cannot be any non-constant Hamiltonian chord contained in the cylindrical end.

Thus, it suffices to prove that this generator yy cannot be a critical point either. The strategy is to consider Lagrangian Floer theory without wrapping. To carry out this idea, we consider the sub-complex CW0(L)CW^{*}_{0}(L\circ\mathcal{L}) of CW(L)CW^{*}(L\circ\mathcal{L}), generated by only critical points. Up to AA_{\infty}-homotopy equivalence, this is just the Morse complex of H|LH^{\prime}|_{L\circ\mathcal{L}} with its Morse AA_{\infty}-structure defined by counting gradient flow trees. There is a similar subspace CW0(L,,L)CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L}) of the quilted wrapped Floer cochain space, generated by generalized chords of low action (in absolute value), i.e. those generalized chords which correspond to Hamiltonian chords that are not contained in the cylindrical end of the product M×NM^{-}\times N. Since LL\circ\mathcal{L} is assumed to be a proper exact cylindrical Lagrangian embedding, CW0(L,,L)CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L}) is indeed a cochain complex equipped with the quilted Floer differential, which can be alternatively defined with respect to a pair of Hamiltonians on (M,N)(M,N) that are C2C^{2}-small in a compact set and linear at infinity of small slope less than the minimal length of a Reeb chord. Then the map (7.37) restricted to CW0(L,,L)CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L}) has image contained in the sub-complex CW0(L)CW^{*}_{0}(L\circ\mathcal{L}). Since these complexes can be identified with Floer complexes without Hamiltonian perturbations (or with small Hamiltonian perturbations if transversality is demanded) up to chain homotopy equivalences, this map (7.26) in fact becomes a cochain map

(7.26) gc:CW0(L,,L)CW0(L).gc:CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L})\to CW^{*}_{0}(L\circ\mathcal{L}).

after restriction without correction by any bounding cochain, by the argument of [LL13]. Moreover, it induces an isomorphism on cohomology groups. Note that the cyclic element eLe_{L\circ\mathcal{L}} in fact lies in the subspace CW0(L,,L)CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L}). In particular, it follows that eLe_{L\circ\mathcal{L}} is closed under the undeformed quilted Floer differential on CW0(L,,L)CW^{*}_{0}(L,\mathcal{L},L\circ\mathcal{L}). Thus we can argue by an analogue of Lemma (7.21) in the setup of quilted Floer theory without wrapping, and conclude that the generator yy in question cannot be any nonzero element in CW0(L)CW^{*}_{0}(L\circ\mathcal{L}). Therefore, y=0y=0 and there cannot be a figure eight bubble, which contradicts our assumption. The proof of Proposition 7.19 is now complete.

7.6. Representability

The previously constructed functor Φ\Phi_{\mathcal{L}} (7.20) is not good enough for understanding the functoriality properties of wrapped Fukaya categories, as modules over a non-proper AA_{\infty}-category can be very complicated. Thus we must find a more geometric replacement. In the case of compact monotone Lagrangian submanifolds in compact monotone symplectic manifolds, there are results from [WW12], [LL13] on the level of cohomology, which establish an isomorphism between the quilted Floer cohomology group and the Floer cohomology group of the geometric composition:

HF(L,,L)HF(L,L).HF^{*}(L,\mathcal{L},L^{\prime})\cong HF^{*}(L\circ\mathcal{L},L^{\prime}).

Now we would like to generalize this statement on the categorical level, aiming to prove that the Yoneda module associated to the geometric composition is homotopy equivalent to the module Φ(L)\Phi_{\mathcal{L}}(L) defined in terms of quilted wrapped Floer theory, and moreover that such homotopy equivalences are functorial in the wrapped Fukaya category of MM. Such a result can be improved to the statement that the module-valued functor Φ\Phi_{\mathcal{L}} (7.20) is representable.

In the previous subsection 7.5, we have shown that if the natural map N\mathcal{L}\to N is proper and if Assumption 7.10 holds, the geometric composition LL\circ\mathcal{L} is always unobstructed, and that there is a canonical choice of a bounding cochain bb for it, determined by LL and \mathcal{L}. Then, via the (left) Yoneda embedding

𝔶l:𝒲im(N)𝒲im(N)lmod,\mathfrak{y}_{l}:\mathcal{W}_{im}(N)\to\mathcal{W}_{im}(N)^{l-mod},

the distinguished object (L,b)(L\circ\mathcal{L},b) defines a left AA_{\infty}-module over 𝒲im(N)\mathcal{W}_{im}(N). The main result of this subsection claims that this AA_{\infty}-module is homotopy to the module Φ(L)\Phi_{\mathcal{L}}(L), which therefore yields the representability of the functor Φ\Phi_{\mathcal{L}} (7.20).

Theorem 7.24.

Suppose that M×N\mathcal{L}\subset M^{-}\times N is an admissible Lagrangian correspondence such that the map N\mathcal{L}\to N is proper, and Assumption 7.10 is satisfied. Then the AA_{\infty}-functor Φ\Phi_{\mathcal{L}} (7.20) is representable. That is, there exists a canonical AA_{\infty}-functor

(7.27) Ψ:𝒲(M)𝒲im(N)replmod\Psi_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}_{im}(N)^{rep-l-mod}

such that iΨi\circ\Psi_{\mathcal{L}} is homotopic to Φ\Phi_{\mathcal{L}} as AA_{\infty}-functors, where

i:𝒲im(N)replmod𝒲im(N)lmodi:\mathcal{W}_{im}(N)^{rep-l-mod}\to\mathcal{W}_{im}(N)^{l-mod}

is the obvious inclusion of the sub-category of representable modules to the category of all modules.

The proof of this theorem will occupy the rest of this subsection. An immediate consequence of this theorem is that we get a functor to the immersed wrapped Fukaya category 𝒲im(N)\mathcal{W}_{im}(N):

Corollary 7.25.

There is an AA_{\infty}-functor

(7.28) Θ:𝒲(M)𝒲im(N),\Theta_{\mathcal{L}}:\mathcal{W}(M)\to\mathcal{W}_{im}(N),

which represents the module-valued functor (7.20), in the sense that 𝔶lΘ\mathfrak{y}_{l}\circ\Theta_{\mathcal{L}} is homotopic to Φ\Phi_{\mathcal{L}}.

Proof.

The Yoneda lemma says that the left Yoneda functor is a homotopy equivalence onto its image, i.e.,

𝔶l:𝒲im(N)𝒲im(N)replmod\mathfrak{y}_{l}:\mathcal{W}_{im}(N)\to\mathcal{W}_{im}(N)^{rep-l-mod}

is a homotopy equivalence. Thus we may choose a homotopy inverse

λl:𝒲im(N)replmod𝒲im(N),\lambda_{l}:\mathcal{W}_{im}(N)^{rep-l-mod}\to\mathcal{W}_{im}(N),

and compose the functor Ψ\Psi_{\mathcal{L}} with λl\lambda_{l} to obtain the desired functor Θ\Theta_{\mathcal{L}} (7.28).

LL\circ\mathcal{L}LL\mathcal{L}L0L^{\prime}_{0}eexx^{\prime}(x,y)(x,y)
Figure 7.2. the quilted map defining the map gcgc

Our proof of Theorem 7.24 is a generalization of the proof of isomorphism of Floer cohomology groups under the geometric composition, as discussed in [Gao17]. For any properly embedded exact cylindrical Lagrangian submanifold L0NL^{\prime}_{0}\subset N, there is a canonical isomorphism

(7.29) gc:CW(L,,L0)CW(L,L0)gc:CW^{*}(L,\mathcal{L},L^{\prime}_{0})\to CW^{*}(L\circ\mathcal{L},L^{\prime}_{0})

of \mathbb{Z}-modules, which become a chain quasi-isomorphism if we equip the latter cochain space CW(L,L0)CW^{*}(L\circ\mathcal{L},L^{\prime}_{0}) with the differential deformed by the bounding cochain bb for the geometric composition LL\circ\mathcal{L} provided by Theorem 7.12. We call this map the geometric composition quasi-isomorphism. Recall that in [Gao17] we already defined this map in a slightly different form:

gc:CW(L,,L0)CW(LHM,L0)gc^{\prime}:CW^{*}(L,\mathcal{L},L^{\prime}_{0})\to CW^{*}(L\circ_{H_{M}}\mathcal{L},L^{\prime}_{0})

in case the geometric composisition is properly embedded, and proved that it is a quasi-isomorphism. This map is related to (7.29) via the homotopy equivalence of left modules associated to LL\circ\mathcal{L} and LHML\circ_{H_{M}}\mathcal{L} - these are modules over the curved AA_{\infty}-category 𝒲ob,im(M)\mathcal{W}_{ob,im}(M) whose objects are proper exact cylindrical Lagrangian immersions (possibly without bounding cochains). This curved AA_{\infty}-category has been introduced in a somewhat implicit way when we defined the immersed wrapped Fukaya category, before the contributions of the bounding cochains to the structure maps are included. In fact, the same proof applies to the current setup: the map (7.29) is defined using moduli spaces

(7.30) 𝒞((x,y);x;e)\mathcal{C}((x,y);x^{\prime};e)

of appropriate inhomogeneous pseudoholomorphic quilted maps, as pictured in Figure 7.2.

The asymptotic condition at the top quilted end (in Figure 7.2) is that the quilted map asymptotically converges to the generalized Hamiltonian chord ee for the triple (L,,L)(L,\mathcal{L},L\circ\mathcal{L}) representing the cyclic element e=eLCW(L,,L)e=e_{L\circ\mathcal{L}}\in CW^{*}(L,\mathcal{L},L\circ\mathcal{L}), which in turn corresponds to the homotopy unit of CW(L)CW^{*}(L\circ\mathcal{L}) under the \mathbb{Z}-module isomorphism (7.29). This cyclic element is discussed in the previous subsection, 7.5.

In order to see the contributions from the bounding cochain bb for LL\circ\mathcal{L}, we must modify these moduli spaces appropriately. Instead of looking at a single moduli space like (7.30), we consider a sequence of moduli spaces

(7.31) 𝒞k((x,y);b,,bk times ;x;e),\mathcal{C}_{k}((x,y);\underbrace{b,\cdots,b}_{k\text{ times }};x^{\prime};e),

where we add kk punctures to the boundary component of a quilted map as in Figure 7.2 which is mapped to LL\circ\mathcal{L}, and impose the asymptotic convergence conditions at these punctures to be given by the bounding cochain bb for LL\circ\mathcal{L}. By counting elements in these moduli spaces of virtual dimension zero, we get a map

(7.32) gc:CW(L,,L0)CW((L,b),L0).gc:CW^{*}(L,\mathcal{L},L^{\prime}_{0})\to CW^{*}((L\circ\mathcal{L},b),L^{\prime}_{0}).

The ”count” requires careful treatment. As the geometric composition is in general no longer an embedding, we cannot use domain-dependent perturbations of Hamiltonians and almost complex structures to achieve transversality of the moduli spaces. The count is instead given by virtual fundamental chains associated to a coherent choice of single-valued multisections for Kuranishi structures on the moduli spaces. Such constructions have been discussed several times and should be routine by now, so we leave the details to the interested reader.

Lemma 7.26.

The map (7.32) is a cochain map, where the differential on CW((L,b),L0)CW^{*}((L\circ\mathcal{L},b),L^{\prime}_{0}) is given by the bb-deformed structure map. That is,

gcn0=m1;bgc,gc\circ n^{0}=m^{1;b}\circ gc,

where n0n^{0} denotes the quilted Floer differential on CW(L,,L0)CW^{*}(L,\mathcal{L},L^{\prime}_{0}), and m1;bm^{1;b} is the bb-deformed Floer differential on CW(L,L0)CW^{*}(L\circ\mathcal{L},L^{\prime}_{0}).

Proof.

By looking at the codimension-one boundary strata of the moduli spaces 𝒞k((x,y);b,,bk times ;x;e)\mathcal{C}_{k}((x,y);\underbrace{b,\cdots,b}_{k\text{ times }};x^{\prime};e) (7.31), we find that the map gcgc as in (7.32) satisfies the following equation:

(7.33) gcn0((x,y))=m1;bgc((x,y))+db(e).gc\circ n^{0}((x,y))=m^{1;b}\circ gc((x,y))+d^{b}(e).

Because the cyclic element ee satisfies the condition that db(e)=0d^{b}(e)=0, the last term vanishes, so the map (7.32) is a cochain map.

With the capping half-disks taken into account, there is a well-defined single-valued action of the generators, so that we can use an action-filtration argument to prove that (7.32) is a cochain isomorphism, as follows. If we truncate the Floer complex using the action filtration, then this map can be written as an upper-triangular matrix with all diagonal entries equal to the ”identity”, as counting non-trivial inhomogeneous pseudoholomorphic quilted maps as above necessarily increases the action. Here by ”identity”, we mean the natural one-to-one correspondence between the set of generators for CW(L,,L0)CW^{*}(L,\mathcal{L},L^{\prime}_{0}) and that for CW(L,L0)CW^{*}(L\circ\mathcal{L},L^{\prime}_{0}).

In the general case where L0L^{\prime}_{0} is an exact cylindrical Lagrangian immersion, the proof of the map (7.32) being a cochain homotopy equivalence is in fact quite similar. The only difference is that the homotopy unit may no longer be closed under the previously-mentioned quilted Floer differential, because there are pseudoholomorphic disks with one marked point in NN with boundary on the image of L0L^{\prime}_{0}. However, L0L^{\prime}_{0} itself comes with a bounding cochain b0b^{\prime}_{0} for it to be an object of the immersed wrapped Fukaya category, and as long as we use the bounding cochain b0b^{\prime}_{0} on L0L^{\prime}_{0} to cancel the contribution of those disks, the homotopy unit of L0L^{\prime}_{0} becomes closed under the deformed quilted Floer differential. In this case, the map (7.32) takes the following form

(7.34) gc:CW(L,,(L0,b0))CW((L,b),(L0,b0)).gc:CW^{*}(L,\mathcal{L},(L^{\prime}_{0},b^{\prime}_{0}))\to CW^{*}((L\circ\mathcal{L},b),(L^{\prime}_{0},b^{\prime}_{0})).

Next, we shall construct a AA_{\infty}-pre-module homomorphism extending the map (7.34) as its first order term. We set gc0=0gc^{0}=0. For d2d\geq 2, we define multilinear maps gcdgc^{d}, for all possible (d1)(d-1)-tuple of testing objects (L0,b0),,(Ld1,bd1)(L^{\prime}_{0},b^{\prime}_{0}),\cdots,(L^{\prime}_{d-1},b^{\prime}_{d-1}) of 𝒲im(N)\mathcal{W}_{im}(N), as follows:

(7.35) gcd:CW(L,,(Ld1,bd1))CW((Ld2,bd2),(Ld1,bd1))CW((L0,b0),(L1,b1))CW((L,b),(L0,b0))\begin{split}gc^{d}:&CW^{*}(L,\mathcal{L},(L^{\prime}_{d-1},b^{\prime}_{d-1}))\otimes CW^{*}((L^{\prime}_{d-2},b^{\prime}_{d-2}),(L^{\prime}_{d-1},b^{\prime}_{d-1}))\\ &\otimes\cdots\otimes CW^{*}((L^{\prime}_{0},b^{\prime}_{0}),(L^{\prime}_{1},b^{\prime}_{1}))\to CW^{*}((L\circ\mathcal{L},b),(L^{\prime}_{0},b^{\prime}_{0}))\end{split}

defined by appropriate ”count” of elements in the moduli spaces

(7.36) 𝒞¯k,d,l0,,ld1((x,y);b0,,b0l0 times ,x1,b1,,b1l1 times ,xd1,bd1,,bd1ld1 times ;b,,bk times ;x;e)\bar{\mathcal{C}}_{k,d,l_{0},\cdots,l_{d-1}}((x,y);\underbrace{b^{\prime}_{0},\cdots,b^{\prime}_{0}}_{l_{0}\text{ times }},x^{\prime}_{1},\underbrace{b^{\prime}_{1},\cdots,b^{\prime}_{1}}_{l_{1}\text{ times }}\cdots,x^{\prime}_{d-1},\underbrace{b^{\prime}_{d-1},\cdots,b^{\prime}_{d-1}}_{l_{d-1}\text{ times }};\underbrace{b,\cdots,b}_{k\text{ times }};x^{\prime};e)

of quilted maps, defined in a way similar to (7.31), but with multiple punctures and Lagrangian labels on the boundary components of the second patch of the quilted surface. The corresponding asymptotic convergence conditions are given by generators xiCW((Li1,bi1),(Li,bi))x^{\prime}_{i}\in CW^{*}((L^{\prime}_{i-1},b^{\prime}_{i-1}),(L^{\prime}_{i},b^{\prime}_{i})), and also those given by bounding cochains biCW((Li,bi)),i=0,,d1b^{\prime}_{i}\in CW^{*}((L^{\prime}_{i},b^{\prime}_{i})),i=0,\cdots,d-1.

Lemma 7.27.

The maps {gcd}\{gc^{d}\} form an AA_{\infty}-pre-module homomorphism

(7.37) gc:Φ(L)𝔶l((L,b))gc:\Phi_{\mathcal{L}}(L)\to\mathfrak{y}_{l}((L\circ\mathcal{L},b))

from the left AA_{\infty}-module Φ(L)\Phi_{\mathcal{L}}(L) over 𝒲im(N)\mathcal{W}_{im}(N), to the left Yoneda module 𝔶l((L,b))\mathfrak{y}_{l}((L\circ\mathcal{L},b)) over 𝒲im(N)\mathcal{W}_{im}(N). Moreover, this AA_{\infty}-pre-module homomorphism is in fact an AA_{\infty}-module homomorphism.

Proof.

The verification of the AA_{\infty}-equations for pre-module homomorphisms can be done by looking at the boundary of the above-mentioned moduli spaces

𝒞¯k,d,l0,,ld1((x,y);b0,,b0l0 times ,x1,b1,,b1l1 times ,xd1,bd1,,bd1ld1 times ;b,,bk times ;x;e).\bar{\mathcal{C}}_{k,d,l_{0},\cdots,l_{d-1}}((x,y);\underbrace{b^{\prime}_{0},\cdots,b^{\prime}_{0}}_{l_{0}\text{ times }},x^{\prime}_{1},\underbrace{b^{\prime}_{1},\cdots,b^{\prime}_{1}}_{l_{1}\text{ times }}\cdots,x^{\prime}_{d-1},\underbrace{b^{\prime}_{d-1},\cdots,b^{\prime}_{d-1}}_{l_{d-1}\text{ times }};\underbrace{b,\cdots,b}_{k\text{ times }};x^{\prime};e).

For this, simply recall that the AA_{\infty}-module structure on CW(L,,)CW^{*}(L,\mathcal{L},\cdot) over 𝒲im(N)\mathcal{W}_{im}(N) is defined via suitable moduli spaces of inhomogeneous pseudoholomorphic quilted maps, while that on CW((L,b),)CW^{*}((L\circ\mathcal{L},b),\cdot) is defined via suitable moduli spaces of stable broken Floer trajectories in NN, with decorations by the bounding cochain bb. These are compatible with the compactification

𝒞¯k,d,l0,,ld1((x,y);b0,,b0l0 times ,x1,b1,,b1l1 times ,xd1,bd1,,bd1ld1 times ;b,,bk times ;x;e),\bar{\mathcal{C}}_{k,d,l_{0},\cdots,l_{d-1}}((x,y);\underbrace{b^{\prime}_{0},\cdots,b^{\prime}_{0}}_{l_{0}\text{ times }},x^{\prime}_{1},\underbrace{b^{\prime}_{1},\cdots,b^{\prime}_{1}}_{l_{1}\text{ times }}\cdots,x^{\prime}_{d-1},\underbrace{b^{\prime}_{d-1},\cdots,b^{\prime}_{d-1}}_{l_{d-1}\text{ times }};\underbrace{b,\cdots,b}_{k\text{ times }};x^{\prime};e),

meaning that the two kinds of moduli spaces arise in the boundary strata of this moduli space. This fact can be used to show that the maps defined above satisfy the AA_{\infty}-equations for pre-module homomorphisms.

Now if we deform the structure maps on CW(L,)CW^{*}(L\circ\mathcal{L},\cdot) by the bounding cochain bb, the cyclic element eLCW(L,,(L,b))e_{L\circ\mathcal{L}}\in CW^{*}(L,\mathcal{L},(L\circ\mathcal{L},b)) becomes closed under the bb-deformed Floer differential, which is the module differential for the left Yoneda module 𝔶l((L,b))\mathfrak{y}_{l}((L\circ\mathcal{L},b)). Thus the above AA_{\infty}-pre-module homomorphism is a cocycle in the functor category, i.e. an AA_{\infty}-module homomorphism.

Corollary 7.28.

The AA_{\infty}-module homomorphism gcgc (7.37) is a quasi-isomorphism of AA_{\infty}-modules.

Proof.

This follows from the fact that the first-order map gc1gc^{1} is an isomorphism of cochain complexes.

This module homomorphism gcgc is more than just a quasi-isomorphism, but indeed a homotopy equivalence. This statement is rather important because we are working over the integers. The proof is in fact very simple, based on our construction of gcgc.

Proposition 7.29.

The AA_{\infty}-module homomorphism

gc:Φ(L)𝔶l((L,b))gc:\Phi_{\mathcal{L}}(L)\to\mathfrak{y}_{l}((L\circ\mathcal{L},b))

is a homotopy equivalence of AA_{\infty}-modules.

Proof.

A homotopy inverse can be constructed using moduli spaces similar to (7.36), but we interchange the inputs and the outputs. That is, we regard xx^{\prime} as the input and (x,y)(x,y) as the output, and construct a sequence of multilinear maps of the form

(7.38) opd:CW((L,b),(Ld1,bd1))CW((Ld2,bd2),(Ld1,bd1))CW((L0,b0),(L1,b1))CW(L,,(L0,b0)).\begin{split}op^{d}:&CW^{*}((L\circ\mathcal{L},b),(L^{\prime}_{d-1},b^{\prime}_{d-1}))\otimes CW^{*}((L^{\prime}_{d-2},b^{\prime}_{d-2}),(L^{\prime}_{d-1},b^{\prime}_{d-1}))\otimes\cdots\\ &\otimes CW^{*}((L^{\prime}_{0},b^{\prime}_{0}),(L^{\prime}_{1},b^{\prime}_{1}))\to CW^{*}(L,\mathcal{L},(L^{\prime}_{0},b^{\prime}_{0})).\end{split}

These form an AA_{\infty}-module homomorphism

op:𝔶l((L,b))Φ(L).op:\mathfrak{y}_{l}((L\circ\mathcal{L},b))\to\Phi_{\mathcal{L}}(L).

Standard gluing argument in Floer theory implies that gc1op1gc^{1}\circ op^{1} and op1gc1op^{1}\circ gc^{1} are both chain homotopic to the identity, which implies that opop is the a homotopy inverse of gcgc.

Thus, we set

(7.39) Ψ(L)=𝔶l((L,b)),\Psi_{\mathcal{L}}(L)=\mathfrak{y}_{l}((L\circ\mathcal{L},b)),

the left Yoneda module of (L,b)Ob𝒲im(N)(L\circ\mathcal{L},b)\in Ob\mathcal{W}_{im}(N), for every object LOb𝒲(M)L\in Ob\mathcal{W}(M). The next step is to prove that such an AA_{\infty}-module homotopy equivalence of AA_{\infty}-modules is functorial in 𝒲(M)\mathcal{W}(M). For this purpose, we shall define multilinear maps

(7.40) Td:CW(Ld1,Ld)CW(L0,L1)hom𝒲im(N)lmod(Φ(L0),Ψ(Ld))[d]T^{d}:CW^{*}(L_{d-1},L_{d})\otimes\cdots\otimes CW^{*}(L_{0},L_{1})\to\hom_{\mathcal{W}_{im}(N)^{l-mod}}(\Phi_{\mathcal{L}}(L_{0}),\Psi_{\mathcal{L}}(L_{d}))[-d]

of degree d-d, which satisfy the equations for AA_{\infty}-pre-natural transformations. In more concrete terms, we shall define a multilinear map for all possible Floer cochains x1CW(L0,L1),,xdCW(Ld1,Ld)x_{1}\in CW^{*}(L_{0},L_{1}),\cdots,x_{d}\in CW^{*}(L_{d-1},L_{d}) as well as cylindrical Lagrangian immersions in NN equipped with bounding cochains (L0,b0),,(Lk1,bk1)(L^{\prime}_{0},b^{\prime}_{0}),\cdots,(L^{\prime}_{k-1},b^{\prime}_{k-1}):

(7.41) (Td(xd,,x1))k:CW((L0,b0),(Lk1,bk1))CW((Lk2,bk1),(Lk1,bk1))CW((L0,b0),(L1,b1))CW((Ld,bd),(L0,b0)),\begin{split}&(T^{d}(x_{d},\cdots,x_{1}))^{k}:CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L^{\prime}_{k-1},b^{\prime}_{k-1}))\\ &\otimes CW^{*}((L^{\prime}_{k-2},b^{\prime}_{k-1}),(L^{\prime}_{k-1},b^{\prime}_{k-1}))\otimes\cdots\otimes CW^{*}((L^{\prime}_{0},b^{\prime}_{0}),(L^{\prime}_{1},b^{\prime}_{1}))\\ &\to CW^{*}((L_{d}\circ\mathcal{L},b_{d}),(L^{\prime}_{0},b^{\prime}_{0})),\end{split}

which is linear with respect to each xix_{i}, and satisfies the following equation:

(7.42) m𝒲im(N)lmod1(Ψd(xd,,x1))+sm𝒲im(N)lmod2(Ψds(xd,,xs+1),Ψs(xs,,x1))=n,k(1)Ψdk+1(xd,,xn+k+1,m𝒲(M)k(xn+k,,xn+1),xn,,x1),\begin{split}&m^{1}_{\mathcal{W}_{im}(N)^{l-mod}}(\Psi_{\mathcal{L}}^{d}(x_{d},\cdots,x_{1}))\\ +&\sum_{s}m^{2}_{\mathcal{W}_{im}(N)^{l-mod}}(\Psi_{\mathcal{L}}^{d-s}(x_{d},\cdots,x_{s+1}),\Psi_{\mathcal{L}}^{s}(x_{s},\cdots,x_{1}))\\ =&\sum_{n,k}(-1)^{*}\Psi_{\mathcal{L}}^{d-k+1}(x_{d},\cdots,x_{n+k+1},m^{k}_{\mathcal{W}(M)}(x_{n+k},\cdots,x_{n+1}),x_{n},\cdots,x_{1}),\end{split}

where (only here) =|x1|++|xn|n*=|x_{1}|+\cdots+|x_{n}|-n.

We consider the following quilted surfaces S¯nf\underline{S}^{nf} consisting of two patches S0nf,S1nfS^{nf}_{0},S^{nf}_{1} where S0nfS^{nf}_{0} is a disk with (d+1)(d+1) positive boundary punctures z0+,1,z01,,z0dz_{0}^{+,1},z_{0}^{1},\cdots,z_{0}^{d}, and one negative puncture z0,2z_{0}^{-,2}, and S0nfS^{nf}_{0} is a disk with (k+1)(k+1) positive boundary punctures z1+,1,z11,,z1k1z_{1}^{+,1},z_{1}^{1},\cdots,z_{1}^{k-1}, and one negative boundary puncture z1,z1,2z_{1}^{-},z_{1}^{-,2}. We denote by I0+I_{0}^{+} the boundary component of S0nfS^{nf}_{0} between z0+,1z_{0}^{+,1} and z0+,2z_{0}^{+,2}, and by I1±I_{1}^{\pm} the boundary component of S1nfS^{nf}_{1} between z1z_{1}^{-} and z1+z_{1}^{+}. S¯nf\underline{S}^{nf} is obtained by seaming together the two patches along the pair (I0+,I1±)(I_{0}^{+},I_{1}^{\pm}) of boundary components. We need to consider semi-stable nodal quilted surfaces arising as domains of limits of stable maps from such quilted surfaces, and we denote them by the same symbol.

Suppose strip-like ends and quilted ends for all semistable nodal quilted surfaces S¯nf\underline{S}^{nf} have been chosen. Make consistent choices of Floer data for all such S¯nf\underline{S}^{nf}, requiring the choices to automorphism-invariant Floer data for semistable nodal quilted surfaces that are domains of stable maps to MM with Lagrangian boundary conditions that are to be specified below. Let ed=eLde_{d}=e_{L_{d}\circ\mathcal{L}} be the generator of CW(Ld,,(Ld,bd))CW^{*}(L_{d},\mathcal{L},(L_{d}\circ\mathcal{L},b_{d})) corresponding to the fundamental chain of LdL_{d}\circ\mathcal{L}. Consider the moduli space

(7.43) 𝒯d,k1nf(α,β;y;x1,,xd;y1,,yk1;(x+,y+);ed)\mathcal{T}^{nf}_{d,k-1}(\alpha,\beta;y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+});e_{d})

of triples (S¯nf,u¯,l1)(\underline{S}^{nf},\underline{u},l_{1}), satisfying the following conditions:

  1. (i)

    u¯:S¯nf(M,N)\underline{u}:\underline{S}^{nf}\to(M,N) is a quilted map with marked points or punctures z¯\underline{\vec{z}} satisfying the following equations:

    (7.44) {(du0αS0nfXHS0nf)0,1=0(du1αS1nfXHS1nf)0,1=0u0(z)ψMρS0nf(z)L0, if z lies between z0+,1 and z01 u0(z)ψMρS0nf(z)Li, if z lies between z0i and z0i+1 u0(z)ψMρS0nf(z)Ld, if z lies between z0d and z0+,2 u1(z)ψNρS1nf(z)ι0(L0), if z lies between z1+,1 and z11 u1(z)ψNρS1nf(z)ιj(Lj), if z lies between z1j and z1j+1 u1(z)ψNρS1nf(z)ιk1(Lk1), if z lies between z1k1 and z1 u1(z)ψNρS1nf(z)LdHM, if z lies between z1 and z1+,2 (u0(z),u1(z))(ψMρS0nf(z)×ψNρS1nf(z)), if z lies on the seam limsu1ϵ1(s,)=ψNw1y()lims+(u0ϵ0+,1(s,),u1ϵ1+,1(s,))=ψM×Nw+(x+(),y+())lims+u0ϵ0i(s,)=ψMw0ixi()lims+(u0ϵ0,2(s,),u1ϵ1,2(s,))=ψM×Nweed()lims+u1ϵ1j(s,)=ψNw1jyj()\begin{cases}(du_{0}-\alpha_{S^{nf}_{0}}\otimes X_{H_{S^{nf}_{0}}})^{0,1}=0\\ (du_{1}-\alpha_{S^{nf}_{1}}\otimes X_{H_{S^{nf}_{1}}})^{0,1}=0\\ u_{0}(z)\in\psi_{M}^{\rho_{S^{nf}_{0}}(z)}L_{0},\text{ if $z$ lies between $z_{0}^{+,1}$ and $z_{0}^{1}$ }\\ u_{0}(z)\in\psi_{M}^{\rho_{S^{nf}_{0}}(z)}L_{i},\text{ if $z$ lies between $z_{0}^{i}$ and $z_{0}^{i+1}$ }\\ u_{0}(z)\in\psi_{M}^{\rho_{S^{nf}_{0}}(z)}L_{d},\text{ if $z$ lies between $z_{0}^{d}$ and $z_{0}^{+,2}$ }\\ u_{1}(z)\in\psi_{N}^{\rho_{S^{nf}_{1}}(z)}\iota_{0}(L^{\prime}_{0}),\text{ if $z$ lies between $z_{1}^{+,1}$ and $z_{1}^{1}$ }\\ u_{1}(z)\in\psi_{N}^{\rho_{S^{nf}_{1}}(z)}\iota_{j}(L^{\prime}_{j}),\text{ if $z$ lies between $z_{1}^{j}$ and $z_{1}^{j+1}$ }\\ u_{1}(z)\in\psi_{N}^{\rho_{S^{nf}_{1}}(z)}\iota_{k-1}(L^{\prime}_{k-1}),\text{ if $z$ lies between $z_{1}^{k-1}$ and $z_{1}^{-}$ }\\ u_{1}(z)\in\psi_{N}^{\rho_{S^{nf}_{1}}(z)}L_{d}\circ_{H_{M}}\mathcal{L},\text{ if $z$ lies between $z_{1}^{-}$ and $z_{1}^{+,2}$ }\\ (u_{0}(z),u_{1}(z))\in(\psi_{M}^{\rho_{S^{nf}_{0}}(z)}\times\psi_{N}^{\rho_{S^{nf}_{1}}(z)})\mathcal{L},\text{ if $z$ lies on the seam }\\ \lim\limits_{s\to-\infty}u_{1}\circ\epsilon_{1}^{-}(s,\cdot)=\psi_{N}^{w_{1}^{-}}y^{-}(\cdot)\\ \lim\limits_{s\to+\infty}(u_{0}\circ\epsilon_{0}^{+,1}(s,\cdot),u_{1}\circ\epsilon_{1}^{+,1}(s,\cdot))=\psi_{M\times N}^{w^{+}}(x^{+}(\cdot),y^{+}(\cdot))\\ \lim\limits_{s\to+\infty}u_{0}\circ\epsilon_{0}^{i}(s,\cdot)=\psi_{M}^{w_{0}^{i}}x^{i}(\cdot)\\ \lim\limits_{s\to+\infty}(u_{0}\circ\epsilon_{0}^{-,2}(s,\cdot),u_{1}\circ\epsilon_{1}^{-,2}(s,\cdot))=\psi_{M\times N}^{w^{e}}e_{d}(\cdot)\\ \lim\limits_{s\to+\infty}u_{1}\circ\epsilon_{1}^{j}(s,\cdot)=\psi_{N}^{w_{1}^{j}}y^{j}(\cdot)\end{cases}
  2. (ii)

    l1:S1nf{z1i:iI}(L0×ι0L0)(Lk1×ιk1Lk1)((Ld)×ι(Ld))l_{1}:\partial S_{1}^{nf}\setminus\{z_{1}^{i}:i\in I\}\to(L^{\prime}_{0}\times_{\iota_{0}}L^{\prime}_{0})\cup\cdots(L^{\prime}_{k-1}\times_{\iota_{k-1}}L^{\prime}_{k-1})\cup((L_{d}\circ\mathcal{L})\times_{\iota}(L_{d}\circ\mathcal{L})) is a smooth map.

  3. (iii)

    ϕNρS1nf(l1(z))ιi=u1l1(z)\phi_{N}^{\rho_{S^{nf}_{1}(l_{1}(z))}}\iota_{i}=u_{1}\circ l_{1}(z), when zS1nfz\in\partial S^{nf}_{1} lies between z1iz_{1}^{i} and z1i+1z_{1}^{i+1}, for every 1ik21\leq i\leq k-2. If zz lies between z1+,1z_{1}^{+,1} and z1z_{1}, the corresponding Lagrangian immersion should be replaced by ι0:L0N\iota_{0}:L^{\prime}_{0}\to N. If zz lies between z1+,2z_{1}^{+,2}, the corresponding Lagrangian immersion should be replaced by ιk1:Lk1N\iota_{k-1}:L^{\prime}_{k-1}\to N. If zz lies between z1+,2z_{1}^{+,2} and z1+,1z_{1}^{+,1}, the corresponding Lagrangian immersion should be replaced by ι:LdN\iota:L_{d}\circ\mathcal{L}\to N.

  4. (iv)

    the relative homology class of u¯\underline{u} is β\beta.

  5. (v)

    the triple (S¯nf,u¯,l1)(\underline{S}^{nf},\underline{u},l_{1}) is stable, meaning that it has finite automorphism group.

The above conditions are analogous to those in the case of a single Lagrangian immersion for which we defined the moduli space of inhomogeneous pseudoholomorphic disks. Recall the relevant notations in section 4.5.

There is a natural compactification 𝒯¯d,k1nf(y;x1,,xd;y1,,yk1;(x+,y+))\bar{\mathcal{T}}^{nf}_{d,k-1}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+})) of this moduli space 𝒯d,k1nf(y;x1,,xd;y1,,yk1;(x+,y+))\mathcal{T}^{nf}_{d,k-1}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+})), which consists of broken quilted maps of the same type. In particular, the codimension-one boundary strata consist of union of fiber products

(7.45) 𝒯¯d,k1nf(y;x1,,xd;y1,,yk1;(x+,y+))𝒯¯d1,k11nf(y;x1,,xi,xnew,xi+k1+1,,xd;y1,,yj,ynew,yj+k2+1,,yk1;(x+,y+))ׯd2+1(xnew,xi+1,,xi+d2)ׯk2+1(ynew,yj+1,,yj+k2)𝒯¯d1,k11((x,y);x1,,xd1;y1,,yk11;(x1+,y1+))×𝒯¯d2,k21((x1+,y1+);xd1+1,,xd;yk1,,yk1;(x+,y+))\begin{split}&\partial\bar{\mathcal{T}}^{nf}_{d,k-1}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+}))\\ &\cong\coprod\bar{\mathcal{T}}^{nf}_{d_{1},k_{1}-1}(y^{-};x^{1},\cdots,x^{i},x^{new},x^{i+k_{1}+1},\cdots,x^{d};\\ &y^{1},\cdots,y^{j},y^{new},y^{j+k_{2}+1},\cdots,y^{k-1};(x^{+},y^{+}))\\ &\times\bar{\mathcal{M}}_{d_{2}+1}(x^{new},x^{i+1},\cdots,x^{i+d_{2}})\times\bar{\mathcal{M}}_{k_{2}+1}(y^{new},y^{j+1},\cdots,y^{j+k_{2}})\\ &\cup\coprod\bar{\mathcal{T}}_{d_{1},k_{1}-1}((x^{-},y^{-});x^{1},\cdots,x^{d_{1}};y^{1},\cdots,y^{k_{1}-1};(x^{+}_{1},y^{+}_{1}))\\ &\times\bar{\mathcal{T}}_{d_{2},k_{2}-1}((x^{+}_{1},y^{+}_{1});x^{d_{1}+1},\cdots,x^{d};y^{k_{1}},\cdots,y^{k-1};(x^{+},y^{+}))\end{split}

where the compactified moduli space 𝒯¯d,k1nf()\bar{\mathcal{T}}^{nf}_{d^{\prime},k^{\prime}-1}(\cdots) for d<d,k<kd^{\prime}<d,k^{\prime}<k are built inductively in this way. The stability condition ensures that the moduli space 𝒯¯d,k1nf(y;x1,,xd;y1,,yk1;(x+,y+))\bar{\mathcal{T}}^{nf}_{d,k-1}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+})) is compact and Hausdorff.

To include the contribution from the bounding cochains bjb^{\prime}_{j} for ιj:LjN\iota_{j}:L^{\prime}_{j}\to N, we modify elements in the moduli space 𝒯¯d,k1nf(y;x1,,xd;y1,,yk1;(x+,y+))\bar{\mathcal{T}}^{nf}_{d,k-1}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};(x^{+},y^{+})) by adding more punctures on each boundary component of S1nfS^{nf}_{1} that is mapped to the image of one of the Lagrangian immersions ιj:LjN\iota_{j}:L^{\prime}_{j}\to N and LdL_{d}\circ\mathcal{L}, and imposing the asymptotic convergence conditions at these additional puctures given by the bounding cochains bjb^{\prime}_{j}. The resulting moduli space is denoted by

(7.46) 𝒯¯d,k1,s0,,sk1,snf(y;x1,,xd;y1,,yk1;b0,,b0;;bk1,,bk1;(x+,y+)).\bar{\mathcal{T}}^{nf}_{d,k-1,s_{0},\cdots,s_{k-1},s}(y^{-};x^{1},\cdots,x^{d};y^{1},\cdots,y^{k-1};b^{\prime}_{0},\cdots,b^{\prime}_{0};\cdots;b^{\prime}_{k-1},\cdots,b^{\prime}_{k-1};(x^{+},y^{+})).

The picture of such a quilted map is shown in Figure 7.3, where we have omitted the bounding cochains, but shall remember that they are also included as suitable asymptotic convergence conditions on the boundary components of the second patch, with the prescribed number of punctures added.

L3L_{3}\circ\mathcal{L}L3L_{3}L2L_{2}L1L_{1}L0L_{0}\mathcal{L}L0L^{\prime}_{0}L1L^{\prime}_{1}L2L^{\prime}_{2}L3L^{\prime}_{3}yy^{-}(x3,e,y3,e)(x^{3,e},y^{3,e})x3x^{3}x2x^{2}x1x^{1}(x+,y+)(x^{+},y^{+})y1y^{1}y2y^{2}y3y^{3}
Figure 7.3. the quilted map defining the homotopy between two module-valued functors

Following the argument in section 4.7, we can construct Kuranishi structures on these moduli spaces, which are compatible with the fiber product Kuranishi structures at the boundary. By making a coherent choice of multisections on these Kuranishi spaces, we obtain the virtual fundamental chains, which give rise to the desired homotopy equivalence between the AA_{\infty}-modules iΨi\circ\Psi_{\mathcal{L}} and Φ\Phi_{\mathcal{L}}.

It is also possible to extend Φ,Ψ\Phi_{\mathcal{L}},\Psi_{\mathcal{L}} as well as Θ\Theta_{\mathcal{L}} to the immersed wrapped Fukaya category 𝒲im(M)\mathcal{W}_{im}(M), which we omit as it is not needed for our current purposes.

7.7. Categorification of the functors

So far we have discussed the representability for the AA_{\infty}-functor (7.3) associated to a single Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N. It is straightforward to generalize this functorially in the wrapped Fukaya category of the product manifold M×NM^{-}\times N. That is, we ask whether (7.1) is representable. The answer is affirmative, stated in Theorem 1.3.

Proof of Theorem 1.3.

The AA_{\infty}-functor (7.1) defines in a natural way an AA_{\infty}-functor

(7.47) Φ:𝒲(M×N)func(𝒲(M),𝒲(N)lmod).\Phi:\mathcal{W}(M^{-}\times N)\to func(\mathcal{W}(M),\mathcal{W}(N)^{l-mod}).

This statement is proved in a purely algebraic way in section 2.4. Summarizing the argument, we compose the AA_{\infty}-functor (7.1) with the algebraically-defined AA_{\infty}-functor

(7.48) (𝒲(M),𝒲(N))bimodfunc(𝒲(M),𝒲(N)lmod)(\mathcal{W}(M),\mathcal{W}(N))^{bimod}\to func(\mathcal{W}(M),\mathcal{W}(N)^{l-mod})

to obtain the desired functor.

In the previous two subsections, we have proved that if the projection N\mathcal{L}\to N is proper and if Assumption 7.10 holds for every LOb𝒲(M)L\in Ob\mathcal{W}(M), then the filtered module-valued functor Φ\Phi_{\mathcal{L}} is representable. Therefore, Φ\Phi is representable over the full subcategory 𝒜(M×N)𝒲(M×N)\mathcal{A}(M^{-}\times N)\subset\mathcal{W}(M^{-}\times N), in the sense of Definition 2.9. Thus we may rewrite the above AA_{\infty}-functor as

(7.49) Ψ:𝒜(M×N)func(𝒲(M),𝒲im(N)replmod).\Psi:\mathcal{A}(M^{-}\times N)\to func(\mathcal{W}(M),\mathcal{W}_{im}(N)^{rep-l-mod}).

Composing this with a homotopy inverse

λl:𝒲im(N)replmod𝒲im(N)\lambda_{l}:\mathcal{W}_{im}(N)^{rep-l-mod}\to\mathcal{W}_{im}(N)

of the left Yoneda functor

𝔶l:𝒲im(N)𝒲im(N)replmod,\mathfrak{y}_{l}:\mathcal{W}_{im}(N)\to\mathcal{W}_{im}(N)^{rep-l-mod},

we obtain the desired AA_{\infty}-functor Θ\Theta (1.6). Technically speaking, we shall require that Assumption 7.10 hold for every Lagrangian correspondence \mathcal{L} in the sub-category 𝒜(M×N)\mathcal{A}(M^{-}\times N), which again is a generic condition on the class of objects of the wrapped Fukaya category of the product manifold M×NM^{-}\times N.

7.8. A geometric realization of the cochain map for the correspondence functor

In practice, it is helpful to have a more direct and geometric construction of the functor (7.28), without referring to the implicit construction with the help of the Yoneda lemma. At this time there are still some technical issues in fully realizing this, but it is possible to construct a cochain map, which is homotopic to the first order map of (7.28). This construction is useful in some applications, for example when studying the relation to the Viterbo restriction functor.

Fix an admissible Lagrangian correspondence M×N\mathcal{L}\subset M^{-}\times N such that the projection N\mathcal{L}\to N is proper. Suppose LML\subset M is an admissible Lagrangian submanifold, which can be made as an object of 𝒲(M)\mathcal{W}(M). Recall that the geometric composition ι:LN\iota:L\circ\mathcal{L}\to N comes with a canonical and unique bounding cochain bb. Given a pair (L0,L1)(L_{0},L_{1}), we define a map

(7.50) Π:CW(L0,L1;HM)CW((L0,b0),(L1,b1);HU)\Pi_{\mathcal{L}}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L_{1}\circ\mathcal{L},b_{1});H_{U})

in the following way. Consider the moduli spaces

𝒰l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1)\mathcal{U}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1})

of quilted inhomogeneous pseudoholomorphic maps (S¯,(u,v))(\underline{S},(u,v)) in (M,N)(M,N), with the following properties:

  1. (i)

    The quilted surface S¯=(S0,S1\underline{S}=(S_{0},S_{1} has two patches. S0S_{0} is a disk with 33 punctures z01,z01,,z02,z_{0}^{1},z_{0}^{1,-},z_{0}^{2,-}, where z01,,z02,z_{0}^{1,-},z_{0}^{2,-} are special punctures. S1kS_{1}^{k} is a disk with 3+l0+l13+l_{0}+l_{1} punctures z11,,z10,1,,z10,l0,z10,z11,1,,z11,l1,z12,z_{1}^{1,-},z_{1}^{0,1},\cdots,z_{1}^{0,l_{0}},z_{1}^{0},z_{1}^{1,1},\cdots,z_{1}^{1,l_{1}},z_{1}^{2,-}, which are ordered in a counterclockwise order on the boundary, where z11,,z12,z_{1}^{1,-},z_{1}^{2,-} are special punctures. The quilted surface is obtained by seaming the two patches along the boundary component I0I_{0}^{-} of S0S_{0} between z01,,z02,z_{0}^{1,-},z_{0}^{2,-} and the boundary component I1I_{1}^{-} of S1S_{1} between z11,,z12,z_{1}^{1,-},z_{1}^{2,-}. Here, we regard the punctures z01z_{0}^{1} and z10z_{1}^{0} as being fixed, while z10,1,,z10,l0z_{1}^{0,1},\cdots,z_{1}^{0,l_{0}} and z11,1,,z11,l1z_{1}^{1,1},\cdots,z_{1}^{1,l_{1}} are allowed to move.

  2. (ii)

    u:S0Mu:S_{0}\to M is inhomogeneous pseudoholomorphic with respect to (HS0,JS0)(H_{S_{0}},J_{S_{0}}), for a family of Hamiltonians HS0H_{S_{0}} on MM parametrized by S0S_{0}, which agrees with HMH_{M} near z01z_{0}^{1}, and a family of almost complex structures JS0J_{S_{0}} parametrized by S0S_{0}, which agrees with JMJ_{M} near z01z_{0}^{1}.

  3. (iii)

    v:S1Nv:S_{1}\to N is inhomogeneous pseudoholomorphic with respect to (HS1,JS1)(H_{S_{1}},J_{S_{1}}), for a family of Hamiltonians HS1H_{S_{1}}, which agrees with HUH_{U} near each of the punctures z10,1,,z10,l0,z10,z11,1,,z11,l1z_{1}^{0,1},\cdots,z_{1}^{0,l_{0}},z_{1}^{0},z_{1}^{1,1},\cdots,z_{1}^{1,l_{1}}, and a family of almost complex structures JS1J_{S_{1}}, which agrees with JUJ_{U} each of the punctures

    z10,1,,z10,l0,z10,z11,1,,z11,l1.z_{1}^{0,1},\cdots,z_{1}^{0,l_{0}},z_{1}^{0},z_{1}^{1,1},\cdots,z_{1}^{1,l_{1}}.
  4. (iv)

    uu maps the boundary component of S0S_{0} between z01,z_{0}^{1,-} and z01z_{0}^{1} to L0L_{0}, the boundary component between z0jz_{0}^{j} and z0j+1z_{0}^{j+1} to LjL_{j} (for j=1,,k1j=1,\cdots,k-1), and the boundary component between z0kz_{0}^{k} and z02,z_{0}^{2,-} to LkL_{k}.

  5. (v)

    vv maps the boundary component of S1S_{1} between z11,z_{1}^{1,-} and z10,1z_{1}^{0,1}, the boundary component between z10,jz_{1}^{0,j} and z10,j+1z_{1}^{0,j+1} (for j=1,,l01j=1,\cdots,l_{0}-1) as well as the boundary component between z10,l0z_{1}^{0,l_{0}} and z10z_{1}^{0} to the image of the geometric composition L0HML_{0}\circ_{H_{M}}\mathcal{L}. vv maps the boundary component between z10z_{1}^{0} and z11,1z_{1}^{1,1}, the boundary component between z11,jz_{1}^{1,j} and z11,j+1z_{1}^{1,j+1} (for j=1,,l11j=1,\cdots,l_{1}-1) as well as the boundary component between z11,l1z_{1}^{1,l_{1}} and z12,z_{1}^{2,-} to the image of the geometric composition L1HML_{1}\circ_{H_{M}}\mathcal{L}.

  6. (vi)

    On the seam, the matching condition for (u,v)(u,v) is given by the Lagrangian correspondence \mathcal{L}.

  7. (vii)

    uu asymptotically converges to some time-one HMH_{M}-chord xx at z01z_{0}^{1}.

  8. (viii)

    vv asymptotically converges to some generator yy for CW(L0,L1)CW^{*}(L_{0}\circ\mathcal{L},L_{1}\circ\mathcal{L}) at z10z_{1}^{0}. In the case where yy is a time-one HNH_{N}-chord from the image of L0L_{0}\circ\mathcal{L} to that of L1L_{1}\circ\mathcal{L}, this condition is the same as those for uu. In the case where yy is a critical point (this happens only when L0=L1L_{0}=L_{1}), the domain S1S_{1} and the map vv have to be slightly modified, to be described later on.

  9. (ix)

    vv asymptotically converges to some generator y0,jy_{0,j} of CW(L0;HN)CW^{*}(L_{0}\circ\mathcal{L};H_{N}) at z10,jz_{1}^{0,j} for j=1,,l0j=1,\cdots,l_{0}, and to some generator y1,jy_{1,j} of CW(L1;HN)CW^{*}(L_{1}\circ\mathcal{L};H_{N}) at z11,jz_{1}^{1,j} for j=1,,l1j=1,\cdots,l_{1}.

  10. (x)

    Over the first quilted end, the quilted map (u,v)(u,v) asymptotically converges to the cyclic element e0e_{0} for (L0,,L0)(L_{0},\mathcal{L},L_{0}\circ\mathcal{L}). Over the second quilted end, the quilted map (u,v)(u,v) asymptotically converges to the cyclic element e1e_{1} for (L1,,L1)(L_{1},\mathcal{L},L_{1}\circ\mathcal{L}).

Now let us describe the necessary modification when L0=L1L_{0}=L_{1} and yy is a critical point of the chosen Morse function on the self fiber product of L0L_{0}\circ\mathcal{L}. In this case, we require that the family of Hamiltonians HS1H_{S_{1}} is chosen so that it vanishes near the strip-like end of z10z_{1}^{0}, so that the map vv converges to a point on the image of L1L_{1}\circ\mathcal{L}. The domain S1S_{1} should also be further modified, by attaching an infinite half ray (,0](-\infty,0] to it at the negative puncture (which now becomes a marked point as the Hamiltonian vanishes near there). Then we require that the lift of the map on the infinite half ray, which is a gradient flow, converges to yy at -\infty. To make the statement concise and unified, in both cases we shall briefly say that the map vv asymptotically converges to yy.

There is a natural stable map compactification of this moduli space, denoted by

(7.51) 𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1),\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1}),

which is constructed in an inductive nature. This compactification is obtained by adding all possible broken inhomogeneous pseudoholomorphic quilted maps. These broken quilted maps arise when energy escapes over the strip-like ends near the punctures (this phenomenon is often called strip breaking), or when the domains degenerate. There are several cases:

  1. (i)

    Inhomogeneous pseudoholomorphic disks bubbling off the boundary of the image of L0L_{0}\circ\mathcal{L}. The resulting broken quilted map has a main component which is similar to such a quilted map (S¯,(u,v))(\underline{S},(u,v)), with possibly less punctures l0l0l^{\prime}_{0}\leq l_{0}, and some other components consisting of trees of inhomogeneous pseudoholomorphic disks with boundary on the image of L0L_{0}\circ\mathcal{L}.

  2. (ii)

    Inhomogeneous pseudoholomorphic disks bubbling off the boundary of the image of L1L_{1}\circ\mathcal{L}. The resulting broken quilted map has a main component which is similar to such a quilted map (S¯,(u,v))(\underline{S},(u,v)), with possibly less punctures l1l1l^{\prime}_{1}\leq l_{1}, and some other components consisting of trees of inhomogeneous pseudoholomorphic disks with boundary on the image of L1L_{1}\circ\mathcal{L}.

  3. (iii)

    Inhomogeneous pseudoholomorphic strips breaking out at the strip-like end near z01z_{0}^{1}.

  4. (iv)

    Inhomogeneous pseudoholomorphic strips breaking out at the strip-like end near z10z_{1}^{0}.

  5. (v)

    Inhomogeneous pseudoholomorphic quilted strips breaking out at the quilted ends.

Thus, there is an isomorphism of the codimension-one boundary strata of the compactified moduli space:

(7.52) 𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1)x1deg(x1)=deg(x)+1¯(x1,x)×𝒰¯l0,l1(α,β;x1;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1)αα′′=αβ′′β′′=βy1deg(y1)=deg(y)1¯(α′′,β′′;y,y1)×𝒰¯l0,l1(α,β;x;y1;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1)1l0l0l0+l0′′=l0+10i0l0αα′′=αβ′′β′′=βy0,newdeg(y0,new)=deg(y0,i0+1)++deg(y0,i0+l0′′)+2l0′′𝒰¯l0,l1(α,β;x;y;y0,1,,y0,i0,y0,new,y0,i0+l0′′+1,,y0,l0;y1,1,,y1,l1;e0,e1)ׯl0′′+1(α′′,β′′;y0,new;y0,i0+1,,y0,i0+l0′′)1l1l1l1+l1′′=l1+11i1l1αα′′=αβ′′β′′=βy1,newdeg(y1,new)=deg(y1,i1+1)++deg(y1,i1+l1′′)+2l1′′𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,i1,y1,new,y1,i1+l1′′+1,,y1,l1)ׯl1′′+1(α′′,β′′;y1,new;y1,i1+1,,y1,i1+l1′′)0l0l0l0+l0′′=l0αα′′=αβ′′β′′=β(x0,new,y0,new)deg((x0,new,y0,new))=deg(e0)+1𝒩¯l0′′(α′′,β′′;y0,1,,y0,l0′′;(x0,new,y0,new),e0)×𝒰¯l0,l1(α,β;x;y;y0,l0′′+1,,y0,l0;y1,1,,y1,l1;(x0,new,y0,new),e1)0l1l1l1+l1′′=l1αα′′=αβ′′β′′=β(x1,new,y1,new)deg((x1,new,y1,new))=deg(e1)+1𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,(x1,new,y1,new))×𝒩¯l1′′(α′′,β′′;y1,l1+1,,y1,l1;(x1,new,y1,new),e1).\begin{split}&\partial\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1})\\ \cong&\coprod_{\begin{subarray}{c}x_{1}\\ \deg(x_{1})=\deg(x)+1\end{subarray}}\bar{\mathcal{M}}(x_{1},x)\\ &\times\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x_{1};y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1})\\ &\cup\coprod_{\begin{subarray}{c}\alpha^{\prime}\sharp\alpha^{\prime\prime}=\alpha\\ \beta^{\prime\prime}\sharp\beta^{\prime\prime}=\beta\end{subarray}}\coprod_{\begin{subarray}{c}y_{1}\\ \deg(y_{1})=\deg(y)-1\end{subarray}}\bar{\mathcal{M}}(\alpha^{\prime\prime},\beta^{\prime\prime};y,y_{1})\\ &\times\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha^{\prime},\beta^{\prime};x;y_{1};y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1})\\ &\cup\coprod_{\begin{subarray}{c}1\leq l^{\prime}_{0}\leq l_{0}\\ l^{\prime}_{0}+l^{\prime\prime}_{0}=l_{0}+1\end{subarray}}\coprod_{0\leq i_{0}\leq l^{\prime}_{0}}\coprod_{\begin{subarray}{c}\alpha^{\prime}\sharp\alpha^{\prime\prime}=\alpha\\ \beta^{\prime\prime}\sharp\beta^{\prime\prime}=\beta\end{subarray}}\coprod_{\begin{subarray}{c}y_{0,new}\\ \deg(y_{0,new})=\deg(y_{0,i_{0}+1})+\cdots+\deg(y_{0,i_{0}+l^{\prime\prime}_{0}})+2-l^{\prime\prime}_{0}\end{subarray}}\\ &\bar{\mathcal{U}}_{l^{\prime}_{0},l_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,1},\cdots,y_{0,i_{0}},y_{0,new},y_{0,i_{0}+l^{\prime\prime}_{0}+1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1})\\ &\times\bar{\mathcal{M}}_{l^{\prime\prime}_{0}+1}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{0,new};y_{0,i_{0}+1},\cdots,y_{0,i_{0}+l^{\prime\prime}_{0}})\\ &\cup\coprod_{\begin{subarray}{c}1\leq l^{\prime}_{1}\leq l_{1}\\ l^{\prime}_{1}+l^{\prime\prime}_{1}=l_{1}+1\end{subarray}}\coprod_{1\leq i_{1}\leq l^{\prime}_{1}}\coprod_{\begin{subarray}{c}\alpha^{\prime}\sharp\alpha^{\prime\prime}=\alpha\\ \beta^{\prime\prime}\sharp\beta^{\prime\prime}=\beta\end{subarray}}\coprod_{\begin{subarray}{c}y_{1,new}\\ \deg(y_{1,new})=\deg(y_{1,i_{1}+1})+\cdots+\deg(y_{1,i_{1}+l^{\prime\prime}_{1}})+2-l^{\prime\prime}_{1}\end{subarray}}\\ &\bar{\mathcal{U}}_{l_{0},l^{\prime}_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,i_{1}},y_{1,new},y_{1,i_{1}+l^{\prime\prime}_{1}+1},\cdots,y_{1,l_{1}})\\ &\times\bar{\mathcal{M}}_{l^{\prime\prime}_{1}+1}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{1,new};y_{1,i_{1}+1},\cdots,y_{1,i_{1}+l^{\prime\prime}_{1}})\\ &\cup\coprod_{\begin{subarray}{c}0\leq l^{\prime}_{0}\leq l_{0}\\ l^{\prime}_{0}+l^{\prime\prime}_{0}=l_{0}\end{subarray}}\coprod_{\begin{subarray}{c}\alpha^{\prime}\sharp\alpha^{\prime\prime}=\alpha\\ \beta^{\prime\prime}\sharp\beta^{\prime\prime}=\beta\end{subarray}}\coprod_{\begin{subarray}{c}(x_{0,new},y_{0,new})\\ \deg((x_{0,new},y_{0,new}))=\deg(e_{0})+1\end{subarray}}\\ &\bar{\mathcal{N}}_{l^{\prime\prime}_{0}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{0,1},\cdots,y_{0,l^{\prime\prime}_{0}};(x_{0,new},y_{0,new}),e_{0})\\ &\times\bar{\mathcal{U}}_{l^{\prime}_{0},l_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,l^{\prime\prime}_{0}+1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};(x_{0,new},y_{0,new}),e_{1})\\ &\cup\coprod_{\begin{subarray}{c}0\leq l^{\prime}_{1}\leq l_{1}\\ l^{\prime}_{1}+l^{\prime\prime}_{1}=l_{1}\end{subarray}}\coprod_{\begin{subarray}{c}\alpha^{\prime}\sharp\alpha^{\prime\prime}=\alpha\\ \beta^{\prime\prime}\sharp\beta^{\prime\prime}=\beta\end{subarray}}\coprod_{\begin{subarray}{c}(x_{1,new},y_{1,new})\\ \deg((x_{1,new},y_{1,new}))=\deg(e_{1})+1\end{subarray}}\\ &\bar{\mathcal{U}}_{l_{0},l^{\prime}_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l^{\prime}_{1}};e_{0},(x_{1,new},y_{1,new}))\\ &\times\bar{\mathcal{N}}_{l^{\prime\prime}_{1}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{1,l^{\prime}_{1}+1},\cdots,y_{1,l_{1}};(x_{1,new},y_{1,new}),e_{1}).\end{split}

Some notations need to be explained. Here

𝒰¯l0,l1(α,β;x;y;y0,l0′′+1,,y0,l0;y1,1,,y1,l1;(x0,new,y0,new),e1)\bar{\mathcal{U}}_{l^{\prime}_{0},l_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,l^{\prime\prime}_{0}+1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};(x_{0,new},y_{0,new}),e_{1})

is the moduli space of quilted maps of the same kind, except that the asymptotic convergence condition at the quilted end (z01,,z11,)(z_{0}^{1,-},z_{1}^{1,-}) is replaced by a new generalized chord (x0,new,y0,new)(x_{0,new},y_{0,new}) for (L0,,L0HM)(L_{0},\mathcal{L},L_{0}\circ_{H_{M}}\mathcal{L}); similarly for

𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,(x1,new,y1,new)).\bar{\mathcal{U}}_{l_{0},l^{\prime}_{1}}(\alpha^{\prime},\beta^{\prime};x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l^{\prime}_{1}};e_{0},(x_{1,new},y_{1,new})).

And

𝒩¯l0′′(α′′,β′′;y0,1,,y0,l0′′;(x0,new,y0,new),e0)\bar{\mathcal{N}}_{l^{\prime\prime}_{0}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{0,1},\cdots,y_{0,l^{\prime\prime}_{0}};(x_{0,new},y_{0,new}),e_{0})

is the moduli space of broken decorated inhomogeneous pseudoholomorphic quilted strips connecting (x0,new,y0,new)(x_{0,new},y_{0,new}) and e0e_{0}, with punctures on the boundary of the first patch; similarly for

𝒩¯l1′′(α′′,β′′;y1,l1+1,,y1,l1;(x1,new,y1,new),e1).\bar{\mathcal{N}}_{l^{\prime\prime}_{1}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{1,l^{\prime}_{1}+1},\cdots,y_{1,l_{1}};(x_{1,new},y_{1,new}),e_{1}).

When the virtual dimension is zero, the virtual fundamental chains of the moduli spaces 𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1)\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1}) give rise to multilinear maps

(7.53) al0,l1:CW(L0;HN)l0CW(L0,L1;HM)CW(L1;HN)l1CW(L0,L1;HN)).\begin{split}a_{l_{0},l_{1}}:&CW^{*}(L_{0}\circ\mathcal{L};H_{N})^{\otimes l_{0}}\otimes CW^{*}(L_{0},L_{1};H_{M})\otimes CW^{*}(L_{1}\circ\mathcal{L};H_{N})^{\otimes l_{1}}\\ &\to CW^{*}(L_{0}\circ\mathcal{L},L_{1}\circ\mathcal{L};H_{N})).\end{split}

By specializing y0,j=b0y_{0,j}=b_{0} and y1,j=b1y_{1,j}=b_{1}, we define

(7.54) Π(x)=l0,l1=0al0,l1(b0,,b0l0 times ;x;b1,,b1l1 times ).\Pi_{\mathcal{L}}(x)=\sum_{l_{0},l_{1}=0}^{\infty}a_{l_{0},l_{1}}(\underbrace{b_{0},\cdots,b_{0}}_{l_{0}\text{ times }};x;\underbrace{b_{1},\cdots,b_{1}}_{l_{1}\text{ times }}).

This is the definition of the map Π\Pi_{\mathcal{L}}.

The main observation is that the map Π\Pi_{\mathcal{L}} is homotopic to the cochain map Θ1\Theta_{\mathcal{L}}^{1}.

L0L_{0}\circ\mathcal{L}L1L_{1}\circ\mathcal{L}\mathcal{L}L0L_{0}L1L_{1}e0e_{0}e1e_{1}
Figure 7.4. the quilted map defining the cochain map
L(0)L^{\prime}(0)L(1)L^{\prime}(1)L(2)L^{\prime}(2)L(3)L^{\prime}(3)\mathcal{L}L0L_{0}L1L_{1}
Figure 7.5. the quilted map defining the linear term of the module-valued functor
L(0)L^{\prime}(0)L(1)L^{\prime}(1)L(2)L^{\prime}(2)L(3)L^{\prime}(3)L0L_{0}\circ\mathcal{L}L1L_{1}\circ\mathcal{L}
Figure 7.6. inhomogeneous pseudoholomorphic disk defining the module structure on the Yoneda module
L(0)L^{\prime}(0)L(1)L^{\prime}(1)L(2)L^{\prime}(2)L(3)L^{\prime}(3)L0L_{0}\circ\mathcal{L}L1L_{1}\circ\mathcal{L}\mathcal{L}L0L_{0}L1L_{1}e0e_{0}e1e_{1}
Figure 7.7. the quilted map for the composition of the cochain map with the Yoneda functor
Proposition 7.30.

The map Π\Pi_{\mathcal{L}} is a cochain map with respect to the usual Floer differential on CW(L0,L1;HM)CW^{*}(L_{0},L_{1};H_{M}) and the (b0,b1)(b_{0},b_{1})-deformed differential on CW(L0,L1;HN)CW^{*}(L_{0}\circ\mathcal{L},L_{1}\circ\mathcal{L};H_{N}).

Moreover, this map is chain homotopic to the first order map Θ1\Theta_{\mathcal{L}}^{1} of the AA_{\infty}-functor Θ\Theta_{\mathcal{L}}.

Proof.

The first statement follows from the facts that the b0b_{0}-deformed and b1b_{1}-deformed AA_{\infty}-structures have vanishing zeroth order terms, as well as the fact that the cyclic elements e0,e1e_{0},e_{1} are closed under the b0b_{0}-deformed and respectively the b1b_{1}-deformed quilted Floer differentials. The underlying geometric idea is as described below. In order to get the correct ”count” of elements, we shall impose the restricted asymptotic convergence conditions y0,j=b0y_{0,j}=b_{0} at z10,jz_{1}^{0,j} and y1,j=b1y_{1,j}=b_{1} at z11,jz_{1}^{1,j}. The meaning of this is as follows. If b0b_{0} (resp. b1b_{1}) is a formal linear combination of generators of CW(L0)CW^{*}(L_{0}\circ\mathcal{L}) (resp. CW(L1)CW^{*}(L_{1}\circ\mathcal{L})), then we consider these moduli spaces whose asymptotic convergence conditions are given by all such generators, and take the corresponding formal linear combination of the virtual fundamental chains with the same coefficients as those for b0b_{0} and b1b_{1}. The numerical effect is that when inserting the bounding cochains b0b_{0} and b1b_{1} for the corresponding asymptotic convergence conditions, the total contributions from any kind of the following moduli spaces ¯l0′′+1(α′′,β′′;y0,new;y0,i0+1,,y0,i0+l0′′)\bar{\mathcal{M}}_{l^{\prime\prime}_{0}+1}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{0,new};y_{0,i_{0}+1},\cdots,y_{0,i_{0}+l^{\prime\prime}_{0}}), or ¯l1′′+1(α′′,β′′;y1,new;y1,i1+1,,y1,i1+l1′′)\bar{\mathcal{M}}_{l^{\prime\prime}_{1}+1}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{1,new};y_{1,i_{1}+1},\cdots,y_{1,i_{1}+l^{\prime\prime}_{1}}), or 𝒩¯l0′′(α′′,β′′;y0,1,,y0,l0′′;(x0,new,y0,new),e0)\bar{\mathcal{N}}_{l^{\prime\prime}_{0}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{0,1},\cdots,y_{0,l^{\prime\prime}_{0}};(x_{0,new},y_{0,new}),e_{0}), or 𝒩¯l1′′(α′′,β′′;y1,l1+1,,y1,l1;(x1,new,y1,new),e1)\bar{\mathcal{N}}_{l^{\prime\prime}_{1}}(\alpha^{\prime\prime},\beta^{\prime\prime};y_{1,l^{\prime}_{1}+1},\cdots,y_{1,l_{1}};(x_{1,new},y_{1,new}),e_{1}) are all zero. In other words, no quilted strip breaking can occur numerically. Now, by looking at (7.52), we note that the only non-trivial contributions are from strip breaking at the strip-like ends near z01z_{0}^{1} or z10z_{1}^{0}, and therefore conclude that Π\Pi_{\mathcal{L}} is a cochain map.

The proof of the second statement is based upon the proof of representability of the module-valued functor Φ\Phi_{\mathcal{L}}. We shall compose this map Π\Pi_{\mathcal{L}} with the Yoneda functor and compare the resulting map to the linear term of module-valued functor Φ\Phi_{\mathcal{L}}. This will be done by analyzing how the relevant inhomogeneous pseudoholomorphic quilted maps can be related. Recall that the module-valued functor Φ\Phi_{\mathcal{L}} defines for each object LL of 𝒲(M)\mathcal{W}(M) an AA_{\infty}-module Φ(L)\Phi_{\mathcal{L}}(L) over 𝒲im(N)\mathcal{W}_{im}(N), with the module structure maps defined by ”counting” elements moduli spaces of quilted inhomogeneous pseudoholomorphic maps (u0,v0)(u^{0},v^{0}), whose first component u0u^{0} in MM is a strip with one boundary component mapped to LL, and second component v0v^{0} is a disk with several punctures, and the boundary components are mapped to objects of 𝒲im(N)\mathcal{W}_{im}(N), which are testing objects for the AA_{\infty}-module Φ\Phi_{\mathcal{L}}. Floer cochains in CW(L0,L1;HM)CW^{*}(L_{0},L_{1};H_{M}) give rise to pre-module homomorphisms from Φ(L0)\Phi_{\mathcal{L}}(L_{0}) to Φ(L1)\Phi_{\mathcal{L}}(L_{1}), which are defined by using moduli spaces of quilted inhomogeneous pseudoholomophic maps (u1,v1)(u^{1},v^{1}), as shown in Figure 7.5. These quilted maps are similar to the previous ones (u0,v0)(u^{0},v^{0}), but satisfying somewhat different conditions:

  1. (i)

    One boundary component of the domain of u1u^{1} is mapped to L0L_{0} and the other to L1L_{1}, with the prescribed Floer cochain in CW(L0,L1;HM)CW^{*}(L_{0},L_{1};H_{M}) as the asymptotic convergence condition at the puncture in between the two boundary components.

  2. (ii)

    The boundary components of the domain of v1v^{1} are mapped to testing objects of 𝒲im(N)\mathcal{W}_{im}(N). For instance, when defining the dd-th order map

    (7.55) (Φ1)d:CW(L0,L1;HM)hom(CW(L0,,(L(0),b(0)))CW((L(0),b(0)),(L(1),b(1)))CW((L(d2),b(d2)),(L(d1),b(d1))),CW(L1,,(L(d1),b(d1))),\begin{split}(\Phi^{1}_{\mathcal{L}})^{d}:&CW^{*}(L_{0},L_{1};H_{M})\\ &\to\hom(CW^{*}(L_{0},\mathcal{L},(L^{\prime}(0),b(0)))\otimes CW^{*}((L^{\prime}(0),b(0)),(L^{\prime}(1),b(1)))\otimes\cdots\\ &\otimes CW^{*}((L^{\prime}(d-2),b(d-2)),(L^{\prime}(d-1),b(d-1))),\\ &CW^{*}(L_{1},\mathcal{L},(L^{\prime}(d-1),b(d-1))),\end{split}

    the boundary conditions for v1v^{1} are given by (L(0),b(0)),,(L(d1),b(d1))(L^{\prime}(0),b(0)),\cdots,(L^{\prime}(d-1),b(d-1)). With bounding cochains included, these are understood as follows. There are additional several punctures on these boundary components, whose asymptotic convergence conditions are given by the bounding cochains b(j)b(j). Here (L(j),b(j))(L^{\prime}(j),b(j)) are general testing objects, which do not have to be the geometric compositions of L0L_{0} or L1L_{1} with \mathcal{L}.

  3. (iii)

    The asymptotic convergence conditions over the two quilted ends are the input and output for the modules Φ(L0)\Phi_{\mathcal{L}}(L_{0}) and Φ(L1)\Phi_{\mathcal{L}}(L_{1}).

On the other hand, recall that the proof of the representability of the module-valued functor Φ\Phi_{\mathcal{L}} uses the moduli spaces 𝒞¯(x;(x,y);e)\bar{\mathcal{C}}(x^{\prime};(x,y);e) (7.30) and 𝒞¯d(x;x1,,xd1;(x,y);e)\bar{\mathcal{C}}_{d}(x^{\prime};x^{\prime}_{1},\cdots,x^{\prime}_{d-1};(x,y);e) (7.36), whose count yields the homotopy equivalence (7.37). Such a quilted map in 𝒞¯d(x;x1,,xd1;(x,y);e)\bar{\mathcal{C}}_{d}(x^{\prime};x^{\prime}_{1},\cdots,x^{\prime}_{d-1};(x,y);e) converges to the cyclic element ee of the corresponding triple (L,,LHM)(L,\mathcal{L},L\circ_{H_{M}}\mathcal{L}) over one quilted end. Algebraically, the virtual fundamental chains of these moduli spaces give rise to a homotopy equivalence from the module Φ(L)\Phi_{\mathcal{L}}(L) to the Yoneda module 𝔶l((L,b))\mathfrak{y}_{l}((L\circ\mathcal{L},b)) for every LL. Now glue the other quilted end of two copies of such quilted maps (u0,v0)(u_{0},v_{0}) and (u1,v1)(u_{1},v_{1}) in the moduli spaces 𝒞¯(x;(x,y);e)\bar{\mathcal{C}}(x^{\prime};(x,y);e) (one for the triple (L0,,L0)(L_{0},\mathcal{L},L_{0}\circ\mathcal{L}) and the other for (L1,,L1HM)(L_{1},\mathcal{L},L_{1}\circ_{H_{M}}\mathcal{L})) to one quilted inhomogeneous pseudoholomorphic map (u1,v1)(u^{1},v^{1}), along the quilted ends over which (ui,vi)(u_{i},v_{i}) have a common asymptotic convergence condition with (u1,v1)(u^{1},v^{1}). The new quilted map (u~,v~)(\tilde{u},\tilde{v}) has two quilted ends, over which the asymptotic convergence conditions are given by the cyclic elements eie_{i} for the triple (Li,,Li)(L_{i},\mathcal{L},L_{i}\circ\mathcal{L}). Also, there are two new punctures on the domain of v~\tilde{v}, and correspondingly two new boundary components which are mapped to the image of the geometric composition L0L_{0}\circ\mathcal{L} and L1L_{1}\circ\mathcal{L} respectively. Over the strip-like ends of these two new punctures, the map v~\tilde{v} asymptotically converges to some generator of CW((L0,b0),(L(0),b(0)))CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L^{\prime}(0),b(0))) and one of CW((L1,b1),(L(d1),b(d1)))CW^{*}((L_{1}\circ\mathcal{L},b_{1}),(L^{\prime}(d-1),b(d-1)))., respectively.

Now let us look at the composition of Π\Pi_{\mathcal{L}} with the linear term of Yoneda functor

(7.56) 𝔶lΠ:CW(L0,L1;HM)CW((L0,b0),(L1,b1);HN)hom(𝔶l((L0,b0)),𝔶l((L1,b1))),\begin{split}\mathfrak{y}_{l}\circ\Pi_{\mathcal{L}}:&CW^{*}(L_{0},L_{1};H_{M})\\ &\to CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L_{1}\circ\mathcal{L},b_{1});H_{N})\\ &\to\hom(\mathfrak{y}_{l}((L_{0}\circ\mathcal{L},b_{0})),\mathfrak{y}_{l}((L_{1}\circ\mathcal{L},b_{1}))),\end{split}

which concretely consists of multilinear maps

(7.57) 𝔶lΠ:CW(L0,L1;HM)CW((L0,b0),(L1,b1);HN)hom(CW(L0,,(L(0),b(0)))CW((L(0),b(0)),(L(1),b(1)))CW((L(d2),b(d2)),(L(d1),b(d1))),CW(L1,,(L(d1),b(d1))),\begin{split}\mathfrak{y}_{l}\circ\Pi_{\mathcal{L}}:&CW^{*}(L_{0},L_{1};H_{M})\\ &\to CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L_{1}\circ\mathcal{L},b_{1});H_{N})\\ &\to\hom(CW^{*}(L_{0},\mathcal{L},(L^{\prime}(0),b(0)))\otimes CW^{*}((L^{\prime}(0),b(0)),(L^{\prime}(1),b(1)))\otimes\cdots\\ &\otimes CW^{*}((L^{\prime}(d-2),b(d-2)),(L^{\prime}(d-1),b(d-1))),\\ &CW^{*}(L_{1},\mathcal{L},(L^{\prime}(d-1),b(d-1))),\end{split}

Such maps are defined using moduli spaces of inhomogeneous pseudoholomorphic quilted maps which are obtained from gluing those quilted maps (u,v)(u,v) in the moduli spaces

𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1),\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1}),

with those inhomogeneous pseudoholomorphic maps which are used to define pre-module homomorphisms between the Yoneda modules. The inhomogeneous pseudoholomorphic maps of the latter kind are just ordinary AA_{\infty}-disks used to define the AA_{\infty}-structure on 𝒲im(N)\mathcal{W}_{im}(N), because the module structure maps and the pre-module homomoprhisms for Yoneda modules are precisely given by the original AA_{\infty}-structure maps for the corresponding objects in 𝒲im(N)\mathcal{W}_{im}(N), as shown in Figure 7.6. After gluing, we get a quilted map as shown in Figure 7.7.

It is then an immediate observation that such a quilted map is of the same type as the previously constructed one (u~,v~)(\tilde{u},\tilde{v}). Recall that the quilted maps (u0,v0)(u_{0},v_{0}) and (u1,v1)(u_{1},v_{1}) are shown in Figure 7.2. Gluing the two quilted maps (u0,v0)(u_{0},v_{0}) and (u1,v1)(u_{1},v_{1}) to (u1,v1)(u^{1},v^{1}) would replace the two quilted ends of (u1,v1)(u^{1},v^{1}) by two new quilted ends, over which the new map (u~,v~)(\tilde{u},\tilde{v}) asymptotically converges to the cyclic elements e0e_{0} and e1e_{1} respectively, and moreover create two new punctures on the component v~\tilde{v}, over which v~\tilde{v} asymptotically converges to generators of CW((L0,b0),(L(0),b(0)))CW^{*}((L_{0}\circ\mathcal{L},b_{0}),(L^{\prime}(0),b(0))) and CW((L1,b1),(L(d1),b(d1)))CW^{*}((L_{1}\circ\mathcal{L},b_{1}),(L^{\prime}(d-1),b(d-1))). Comparing this to Figure 7.7, we see that they are of the same type.

Such an argument can be generalized to the compactified moduli spaces in a straightforward way, by repeating the same process for broken quilted maps. It is clear that this process respects the structure of the boundary strata as described in (7.52). Translating the result algebraically, we conclude that the map 𝔶lΠ\mathfrak{y}_{l}\circ\Pi_{\mathcal{L}} is chain homotopic to Φ1\Phi_{\mathcal{L}}^{1}. Since Θ\Theta_{\mathcal{L}} represents the module-valued functor Φ\Phi_{\mathcal{L}}, we conclude that the map Π\Pi_{\mathcal{L}} is chain homotopic to Θ1\Theta_{\mathcal{L}}^{1}. Therefore the proof of the second statement is complete.

7.9. A Künneth formula for the wrapped Fukaya category

As in classical Floer cohomology theory, it is natural to expect that the wrapped Fukaya category of the product manifold 𝒲(M×N)\mathcal{W}(M\times N) can be expressed as a tensor product of 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N). This is some kind of Künneth formula in wrapped Floer theory. As an application of the construction of functors discussed before, in particular the bimodule-valued functors, we can phrase this in a precise way, where we take the statement in [GPS17].

There are some issues regarding the statement of the Künneth formula. First, 𝒲(M×N)\mathcal{W}(M\times N) generally has more objects than product Lagrangian submanifolds L×LL\times L^{\prime}, so in general we cannot expect this to be equivalent to the tensor product. However, passing to the split-closure gives some hope of establishing this kind of equivalence, provided that 𝒲(M×N)\mathcal{W}(M\times N) is split-generated by product Lagrangian submanifolds.

Second, the definition of AA_{\infty}-tensor product of AA_{\infty}-categories is delicate. Rather than giving a definition by certain ”universal property”, we shall take a particular model of the AA_{\infty}-tensor product 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N) so that the construction of bimodule-valued functors associated to Lagrangian correspondences can be untilized in the proof. For the definition of the AA_{\infty}-tensor product, we follow the construction of [SU04], in which a diagonal for the Stasheff associahedra is constructed. In particular, in this model of AA_{\infty}-tensor product 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N), the objects are pairs (L,L)(L,L^{\prime}) of Lagrangian submanifolds of MM and NN respectively (or formal products L×LL\times L^{\prime}, the underlying morphism spaces are the usual tensor products of wrapped Floer complexes

(7.58) hom((L0,L0),(L1,L1))=CW(L0,L1)CW(L0,L1),\hom((L_{0},L^{\prime}_{0}),(L_{1},L^{\prime}_{1}))=CW^{*}(L_{0},L_{1})\otimes CW^{*}(L^{\prime}_{0},L^{\prime}_{1}),

and the first order structure map is the tensor product wrapped Floer differential:

(7.59) m1=mL0,L11id+idmL0,L11.m^{1}=m^{1}_{L_{0},L_{1}}\otimes id+id\otimes m^{1}_{L^{\prime}_{0},L^{\prime}_{1}}.

However, we remark that all the different constructions are quasi-isomorphic (see [MS06]).

Proposition 7.31.

Assume both MM and NN are non-degenerate, i.e. their wrapped Fukaya categories satisfy Abouzaid’s generation criterion for finite collections of Lagrangian submanifolds. Then there is a canonical quasi-equivalence:

(7.60) 𝒲(M×N)perf(𝒲(M)𝒲(N))perf.\mathcal{W}(M\times N)^{perf}\to(\mathcal{W}(M)\otimes\mathcal{W}(N))^{perf}.

The outline of the proof goes as follows. First, non-degeneracy implies that there are finite collections of Lagrangian submanifolds L1,,LkL_{1},\cdots,L_{k} of MM and L1,,LlL^{\prime}_{1},\cdots,L^{\prime}_{l} of NN which split-generate their wrapped Fukaya categories, and moreover, the products Li×LjL_{i}\times L^{\prime}_{j} split-generate 𝒲(M×N)\mathcal{W}(M\times N). Thus it will be enough to consider the full AA_{\infty}-subcategory 𝒫\mathcal{P} of 𝒲(M×N)\mathcal{W}(M\times N) whose objects are products Li×LjL_{i}\times L^{\prime}_{j}. Second, the framework of Lagrangian correspondence gives us a bimodule-valued functor

(7.61) 𝒫(𝒲(M),𝒲(N))bimod,\mathcal{P}\to(\mathcal{W}(M^{-}),\mathcal{W}(N))^{bimod},

which is defined at the beginning of this section. Third, there is a canonical algebraically defined Yoneda-type AA_{\infty}-functor

(7.62) 𝒲(M)𝒲(N)(𝒲(M),𝒲(N))bimod,\mathcal{W}(M)\otimes\mathcal{W}(N)\to(\mathcal{W}(M^{-}),\mathcal{W}(N))^{bimod},

to the dg-category of AA_{\infty}-bimodules over (𝒲(M),𝒲(N))(\mathcal{W}(M^{-}),\mathcal{W}(N)). An appropriate version of Yoneda lemma says that this is cohomologically full and faithful, thus is a quasi-isomorphism onto the image. Fourth, note that the image of 𝒫\mathcal{P} under the AA_{\infty}-functor (7.61) lands in the image of 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N) under the AA_{\infty}-functor (7.62) in the dg-category of bimodules. Inverting the functor (7.62) on the image gives us an AA_{\infty}-functor

(7.63) 𝒫𝒲(M)𝒲(N).\mathcal{P}\to\mathcal{W}(M)\otimes\mathcal{W}(N).

Finally, using a direct argument by analyzing the moduli space of inhomogeneous pseudoholomorphic quilted strips, we can easily prove the Künneth formula on the level of cohomology. This isomorphism on cohomology can be rephrased as the statement that the above AA_{\infty}-functor is a quasi-isomorphism, which allows us to compare the images of 𝒫\mathcal{P} and 𝒲(M)𝒲(N)\mathcal{W}(M)\otimes\mathcal{W}(N) in the dg-category of bimodules (𝒲(M),𝒲(N))bimod(\mathcal{W}(M^{-}),\mathcal{W}(N))^{bimod}, and show that they are quasi-isomorphic. Passing to the split-closure, we get a quasi-equivalence:

(7.64) 𝒲(M×N)perf(𝒲(M)𝒲(N))perf.\mathcal{W}(M\times N)^{perf}\to(\mathcal{W}(M)\otimes\mathcal{W}(N))^{perf}.

While this is an outline, a complete detailed proof only requires careful writing out the formulas for the algebraic structures involving AA_{\infty}-tensor products that we use here. The reader is referred to [SU04] and [MS06] for detailed account of AA_{\infty}-tensor products and related algebraic results.

Remark 7.32.

In fact, a chain-level AA_{\infty}-quasi-equivalence between (the split-closure of) the split wrapped Fukaya category 𝒲s(M×N)\mathcal{W}_{s}(M\times N) and (the split-closure of) AA_{\infty}-tensor product of (suitable dg-replacements of) 𝒲(M)\mathcal{W}(M) and 𝒲(N)\mathcal{W}(N) was already established in [Gan13]. That is, Proposition 7.31 was proved there for the split wrapped Fukaya category 𝒲s(M×N)\mathcal{W}^{s}(M\times N) of the product manifold, with objects being products of exact cylindrical Lagrangian submanifolds of individual factors.

Without assuming non-degeneracy of MM or NN, there is a general version of Künneth formula for the wrapped Fukaya category.

Proposition 7.33.

There is a natural cohomologically fully faithful AA_{\infty}-bifunctor (or say AA_{\infty}-bilinear functor)

(7.65) 𝔎𝔲𝔫:𝒲(M)×𝒲(N)𝒲(M×N).\mathfrak{Kun}:\mathcal{W}(M)\times\mathcal{W}(N)\to\mathcal{W}(M\times N).

This determines a unique cohomologically fully faithful AA_{\infty}-functor

(7.66) 𝔎𝔲𝔫:𝒲(M)𝒲(N)𝒲(M×N),\mathfrak{Kun}^{\prime}:\mathcal{W}(M)\otimes\mathcal{W}(N)\to\mathcal{W}(M\times N),

up to homotopy.

Proof.

Define 𝒲(M×N)\mathcal{W}(M\times N) using the split model 𝒲s(M×N)\mathcal{W}^{s}(M\times N). Then the AA_{\infty}-bifunctor (7.65) can be defined in a straightforward way. On the level of objects, the bifunctor simply takes a pair (L,L)(L,L^{\prime}) to their direct product,

(L,L)L×L.(L,L^{\prime})\mapsto L\times L^{\prime}.

On the level of morphism spaces, the bifunctor

CW(L0,L1)×CW(L0,L1)CW(L0×L0,L1×L1)CW^{*}(L_{0},L_{1})\times CW^{*}(L^{\prime}_{0},L^{\prime}_{1})\to CW^{*}(L_{0}\times L^{\prime}_{0},L_{1}\times L^{\prime}_{1})

is defined as follows. For any basic Floer cochains xCW(L0,L1)x\in CW^{*}(L_{0},L_{1}) and xCW(L0,L1)x^{\prime}\in CW^{*}(L^{\prime}_{0},L^{\prime}_{1}) (here basic means that the Floer cochain is represented by a single Hamiltonian chord), the product x×xx\times x^{\prime} is naturally a Floer cochain in CW(L0×L0,L1×L1)CW^{*}(L_{0}\times L^{\prime}_{0},L_{1}\times L^{\prime}_{1}) since the latter is defined with respect to the split Hamiltonian. Higher-order terms are of the form

(7.67) 𝔎𝔲𝔫k,l:(CW(Lk1,Lk)CW(L0,L1))×(CW(Ll1,Ll)CW(L0,L1))CW(L0×L0,Lk×Ll),\begin{split}\mathfrak{Kun}^{k,l}:&(CW^{*}(L_{k-1},L_{k})\otimes\cdots\otimes CW^{*}(L_{0},L_{1}))\\ &\times(CW^{*}(L^{\prime}_{l-1},L^{\prime}_{l})\otimes\cdots\otimes CW^{*}(L^{\prime}_{0},L^{\prime}_{1}))\to CW^{*}(L_{0}\times L^{\prime}_{0},L_{k}\times L^{\prime}_{l}),\end{split}

which are defined as follows. Let xiCW(Li1,Li)x_{i}\in CW^{*}(L_{i-1},L_{i}) and xjCW(Lj1,Lj)x^{\prime}_{j}\in CW^{*}(L^{\prime}_{j-1},L^{\prime}_{j}) be basic Floer cochains. The image of (xkx1)×(xlx1)(x_{k}\otimes\cdots\otimes x_{1})\times(x^{\prime}_{l}\otimes\cdots\otimes x^{\prime}_{1}) under the map (7.67) is defined by counting inhomogeneous pseudoholomorphic (generalized) quilted maps of the following kind:

L4L_{4}L3L_{3}L2L_{2}L1L_{1}L0L_{0}L4L^{\prime}_{4}L3L^{\prime}_{3}L2L^{\prime}_{2}L1L^{\prime}_{1}L0L^{\prime}_{0}x4x_{4}x3x_{3}x2x_{2}x1x_{1}x4x^{\prime}_{4}x3x^{\prime}_{3}x2x^{\prime}_{2}x1x^{\prime}_{1}eL4×L4e_{L_{4}\times L^{\prime}_{4}}eL0×L0e_{L_{0}\times L^{\prime}_{0}}α\alpha
Figure 7.8. the quilted map defining the Künneth bifunctor

Here the output is αCW(L0×L0,Lk×Ll)\alpha\in CW^{*}(L_{0}\times L^{\prime}_{0},L_{k}\times L^{\prime}_{l}), and the asymptotic conditions near the quilted ends are given by the elements

eL0×L0CW(L0,L0×L0,L0),e_{L_{0}\times L^{\prime}_{0}}\in CW^{*}(L_{0},L_{0}\times L^{\prime}_{0},L^{\prime}_{0}),

and

eLk×LlCW(Lk,Lk×Ll,Ll)e_{L_{k}\times L^{\prime}_{l}}\in CW^{*}(L_{k},L_{k}\times L^{\prime}_{l},L^{\prime}_{l})

which correspond to the homotopy unit

1L0×L0CW(L0×L0),1_{L_{0}\times L^{\prime}_{0}}\in CW^{*}(L_{0}\times L^{\prime}_{0}),

and respectively

1Lk×LlCW(Lk×Ll).1_{L_{k}\times L^{\prime}_{l}}\in CW^{*}(L_{k}\times L^{\prime}_{l}).

This construction includes k=l=1k=l=1 as a special case, as such a quilted map is necessarily constant when k=l=1k=l=1.

It remains to prove that 𝔎𝔲𝔫\mathfrak{Kun} (7.65) as defined above is cohomologically fully faithful. This follows immediately from the definition of 𝔎𝔲𝔫\mathfrak{Kun}, as the first order map is the ”identity”: sending (x,x)(x,x^{\prime}) to the product x×xx\times x^{\prime}, which induces an isomorphism on wrapped Floer cohomology groups, proven in [Gao17] as a special case of the results in section 6.

The statement regarding 𝔎𝔲𝔫\mathfrak{Kun}^{\prime} (7.66) follows from the universal property of the AA_{\infty}-tensor product.

Now let us go back to the situation where both MM and NN are non-degenerate and compare the statement of Proposition 7.31 and that of Proposition 7.33. Non-degeneracy implies that the Kunneth AA_{\infty}-functor (7.66) is invertible after passing to the idempotent completion, and the inverse is given by the quasi-equivalence (7.60).

8. Liouville sub-domains and restriction functors

8.1. Sub-domains and restrictions of exact Lagrangian submanifolds

Let M0M_{0} be a Liouville domain, and MM its completion. A Liouville sub-domain U0U_{0} is a codimension zero exact symplectic submanifold of M0M_{0} with smooth boundary, such that the Liouville vector field on M0M_{0} points outward the boundary U0\partial U_{0} transversely. U0U_{0} is itself a Liouville domain with the induced Liouville structure. We may attach to U0U_{0} an infinite half-cylinder U0×[1,+)\partial U_{0}\times[1,+\infty) to the boundary to get a complete Liouville manifold UU. In this paper, we shall only consider Liouville sub-domains, and sometimes call them sub-domains for short.

Denote by MM^{-} the Liouville manifold with opposite symplectic form ω-\omega, with the Liouville form also being the opposite λM-\lambda_{M}, similarly for UU^{-}. There is a natural Lagrangian correspondence between MM and UU, defined as follows. Because U0U_{0} is a Liouville sub-domain, the graph of the embedding U0M0U_{0}\subset M_{0} is naturally an exact Lagrangian submanifold (with corners) of U0×M0U_{0}^{-}\times M_{0} with respect to the product exact one-form λU0×λM0-\lambda_{U_{0}}\times\lambda_{M_{0}}. The natural completion of that in U×MU^{-}\times M, is an exact cylindrical Lagrangian submanifold of U×MU^{-}\times M, with respect to the contact hypersurface Σ\Sigma. We denote this Lagrangian correspondence by Γ\Gamma. Alternatively, consider the natural embedding

(8.1) i:UMi:U\to M

induced by the inclusion U0M0U_{0}\subset M_{0}, whose is definition is given by the following formula:

(8.2) i(x)={x, if xU0,ψMr(i(y)), if x=(y,r)U×[1,+).i(x)=\begin{cases}x,&\text{ if }x\in U_{0},\\ \psi_{M}^{r}(i(y)),&\text{ if }x=(y,r)\in\partial U\times[1,+\infty).\end{cases}

Then Γ\Gamma is the graph of ii.

Equipped with the opposite primitive, it can also be regarded as an exact Lagrangian submanifold of M×UM^{-}\times U. To make Γ\Gamma into an admissible Lagrangian correspondence in Floer-theoretic sense, one needs a spin structure on it. A natural spin structure exists because we assume both MM and UU have vanishing first Chern classes. We shall give this Lagrangian correspondence a special name.

Definition 8.1.

The Lagrangian submanifold ΓM×U\Gamma\subset M^{-}\times U equipped with the natural spin structure is called the graph correspondence.

Given an exact Lagrangian submanifold L0L_{0} of M0M_{0}, possibly with non-empty boundary L0\partial L_{0}, the intersection L0=L0U0L^{\prime}_{0}=L_{0}\cap U_{0} is naturally an exact Lagrangian submanifold of U0U_{0}, possibly with non-empty boundary L0\partial L^{\prime}_{0}, even if L0L_{0} has empty boundary.

If L0L_{0} is either a closed exact Lagrangian submanifold, or an exact cylindrical Lagrangian submanifold, it can be naturally completed to an object LL of 𝒲(M)\mathcal{W}(M). However, in general this is not true for L0L^{\prime}_{0}, as the boundary L0\partial L^{\prime}_{0} might not behave nicely. In the next subsection we shall seek geometric assumptions such that L0L^{\prime}_{0} can be completed into an object of 𝒲(U)\mathcal{W}(U), so that we may attempt to define a restriction functor.

8.2. Restriction and the associated functor

First, note that the graph correspondence Γ\Gamma is admissible for wrapped Floer theory in the product M×UM^{-}\times U. This implies that the module-valued functor ΦΓ\Phi_{\Gamma} is well-defined. More importantly, note that the projection ΓU\Gamma\to U is proper, thus the module-valued functor ΦΓ\Phi_{\Gamma} is representable, and represented by the AA_{\infty}-functor

(8.3) ΘΓ:𝒲(M)𝒲im(U).\Theta_{\Gamma}:\mathcal{W}(M)\to\mathcal{W}_{im}(U).

In this and next subsections, we shall study this functor in a more specific way, compare it with the Viterbo restriction functor defined in [AS10], and prove Theorem 1.4. While the construction of the Viterbo restriction functor will not be repeated in this paper, we shall describe elements in the moduli spaces used to define it, in a sketchy way. More detailed definition and generalization of the Viterbo restriction functor will be discussed in the upcoming work [Gao].

Remark 8.2.

The graph correspondence Γ\Gamma can also be viewed as a Lagrangian correspondence from UU to MM. If we ask whether it defines a functor from 𝒲(U)\mathcal{W}(U) to 𝒲im(M)\mathcal{W}_{im}(M), the answer is not always yes. This is because the projection ΓM\Gamma\to M is not always proper, contrary to ΓU\Gamma\to U. In that case, the module-valued functor is not necessarily representable in our sense.

Consider an object LL in 𝒲(M)\mathcal{W}(M). Suppose it is the completion of an exact cylindrical Lagrangian submanifold L0L_{0} of M0M_{0} which intersects U0U_{0} nicely in the following sense.

Assumption 8.3.

L=L0U0\partial L^{\prime}=L_{0}\cap\partial U_{0} is a Legendrian submanifold of U0\partial U_{0}; and moreover, the primitive of L0L_{0} is locally constant near both M0\partial M_{0} and U0\partial U_{0}.

If L0L_{0} satisfies Assumption 8.3, then L0=L0U0L^{\prime}_{0}=L_{0}\cap U_{0} is an exact cylindrical Lagrangian submanifold of U0U_{0} and can therefore be completed to an exact cylindrical Lagrangian submanifold LUL^{\prime}\subset U. Because of the geometric nature of the AA_{\infty}-functor

ΘΓ:𝒲(M)𝒲im(U),\Theta_{\Gamma}:\mathcal{W}(M)\to\mathcal{W}_{im}(U),

we expect that it behaves like a restriction functor, and takes LL to LL^{\prime} on the level of objects.

However, the true story is a bit more complicated, and in fact the above expectation does not always hold true. In order for the geometric composition LHMΓL\circ_{H_{M}}\Gamma to be related to LL^{\prime}, we further need the following condition:

Assumption 8.4.

LL is invariant under the Liouville flow in the intermediate region M0int(U0)M_{0}\setminus int(U_{0}).

This assumption means that, the Lagrangian LL is the Liouville completion of the restriction L0=LU0L^{\prime}_{0}=L\cap U_{0}, in the sense that LL is the union of L0L^{\prime}_{0} and all trajectories of the Liouville flow on MM starting from points on L\partial L^{\prime}.

Example 8.5.

Let M0M_{0} and U0U_{0} be (critical) Weinstein domains. If M0int(U0)M_{0}\setminus int(U_{0}) is a Weinstein cobordism, and LL is the union of some cylinders over Legendrians and the stable or unstable manifolds of those critical points of a Lyapunov function which are contained in the cobordism M0int(U0)M_{0}\setminus int(U_{0}), then LL satisfies Assumption 8.4.

Under Assumption 8.4, it is easy to see that the geometric composition LΓL\circ\Gamma is precisely LL^{\prime}, the completion of the restriction L0L^{\prime}_{0} of LL to U0U_{0}. In fact, the geometric composition in the usual sense (without Hamiltonian perturbation) is the map

LΓ={(p1,p2,q)M×M×U:p1L,p1=p2,i(q)=p2}UL\circ\Gamma=\{(p_{1},p_{2},q)\in M\times M\times U:p_{1}\in L,p_{1}=p_{2},i(q)=p_{2}\}\to U

which sends (p1,p2,q)(p_{1},p_{2},q) to qUq\in U. In this case this is an exact Lagrangian embedding, whose image is

{qU:pM, such that pL,i(q)=p}.\{q\in U:\exists p\in M,\text{ such that }p\in L,i(q)=p\}.

By the definition of the map i:UMi:U\to M as in (8.2), this is the set of all points qUq\in U such that either qLU0q\in L\cap U_{0}, or q=(y,r)U×[1,+)q=(y,r)\in\partial U\times[1,+\infty) such that ψMr(y)=p\psi_{M}^{r}(y)=p. By Assumption 8.4, this is exactly LL^{\prime}.

Remark 8.6.

In general, every exact cylindrical Lagrangian LL which intersects U\partial U to a Legendrian submanifold is invariant under the Liouville flow in a neighborhood of U\partial U and one of M\partial M, after suitable Hamiltonian perturbation.

On the other hand, under assumption Assumption 8.3, we can show that the bounding cochain in fact vanishes by Proposition 7.19, so that the image of the functor ΘΓ\Theta_{\Gamma} lands in the completed wrapped Fukaya category 𝒲(U)\mathcal{W}(U) whose objects are embedded Lagrangian submanifolds. This is the first half of the statement of Theorem 1.4.

The full sub-category of 𝒲(M)\mathcal{W}(M) consisting of Lagrangian submanifolds which satisfy Assumption 8.3 and Assumption 8.4 will be denoted by 0(M)\mathcal{B}_{0}(M). This is a full sub-category of (M)\mathcal{B}(M). The restriction of the functor ΘΓ\Theta_{\Gamma} to this sub-category is

(8.4) ΘΓ:0(M)𝒲(U),\Theta_{\Gamma}:\mathcal{B}_{0}(M)\to\mathcal{W}(U),

such that the image of any LOb0(M)L\in Ob\mathcal{B}_{0}(M) is LOb𝒲(U)L^{\prime}\in Ob\mathcal{W}(U). This finishes the proof of the first half of Theorem 1.4.

8.3. The Viterbo restriction functor revisited

To prove the second half of Theorem 1.4, we shall briefly review the definition of the Viterbo restriction functor first, originally due to [AS10]. As the setup of the wrapped Fukaya category is slightly different (but equivalent), we shall consider inhomogeneous pseudoholomorphic maps which behave somewhat differently than cascade maps. However, they should yield an equivalent definition of the Viterbo restriction functor, although we will not attempt to include a proof.

The relevant inhomogeneous pseudoholomorphic maps that are used to define the Viterbo restriction functor will be called ”climbing disks”. Climbing disks are in fact very similar to cascades introduced in [AS10]. Roughly speaking, these are inhomogeneous pseudoholomorphic disks in MM, for which the Floer data consist of a family of Hamiltonians and a family of almost complex structures interpolating the Floer data on MM and those on UU, and the boundary conditions are given by a family of Lagrangian submanifolds interpolating LL and LL^{\prime}. In order to visualize Floer theory of UU in MM, we shall use a rescaling trick. More details are presented below.

To define climbing disks, we introduce the following geometric data. Conceptually, we want to consider geometric objects which reflect the way how the Liouville manifolds MM and UU can be interpolated in a suitable sense. The geometric data on MM and those on UU should be related in a one-parameter family. While in [AS10] they think of the size of the collar neighborhood of U\partial U inside U0U_{0} as the parameter (shrinking the sub-domain by a conformal factor ρ(0,1]\rho\in(0,1] which is taken to be the parameter), we fix the size of the collar neighborhood (as in [Vit99]), but rather change the height of the Hamiltonian and regard that as a parameter.

Note that the Liouville structure on U0U_{0} induced by that on M0M_{0} provides a symplectic embedding of a collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1] to U0U_{0} such that U×{1}\partial U\times\{1\} is mapped to the boundary U\partial U. Suppose we have chosen the quadratic Hamiltonian HMH_{M} on MM and HUH_{U} on UU, such that they agree on U0M0U_{0}\subset M_{0}. Define a family of piecewise smooth (and lower semi-continuous) Hamiltonians KAK_{A} depending on a parameter A[0,+)A\in[0,+\infty),

(8.5) KA={0, on U0(U×(1ϵ,1),A(r1+ϵ)2, if p=(y,r)U×(1ϵ,1),A, on M0U0,A+h(r), if x=(y,r)M×[1,+),K_{A}=\begin{cases}0,&\text{ on }U_{0}\setminus(\partial U\times(1-\epsilon,1),\\ A(r-1+\epsilon)^{2},&\text{ if }p=(y,r)\in\partial U\times(1-\epsilon,1),\\ A,&\text{ on }M_{0}\setminus U_{0},\\ A+h(r),&\text{ if }x=(y,r)\in\partial M\times[1,+\infty),\end{cases}

where h(r)h(r) is the same radial function as that for HMH_{M}, i.e. h(r)=r2h(r)=r^{2} if rr is slightly bigger than 11. Note that when A=0A=0, this is simply an admissible Hamiltonian on MM which vanishes in the interior part M0M_{0} and is quadratic in the cylindrical end M×[1,+)\partial M\times[1,+\infty). Let HAH_{A} be a C2C^{2}-small perturbation of KAK_{A}, which is smooth and non-degenerate, such that HAH_{A} agrees with KAK_{A} in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1] as well as in the cylindrical end M×[1,+)\partial M\times[1,+\infty).

We also need a suitable class of almost complex structures in order to write down the inhomogeneous Cauchy-Riemann equations for climbing disks. Suppose we have chosen almost complex structures JM=JMJ_{M}=J_{M} on MM and JU=JUJ_{U}=J_{U} on UU. What is needed is a family of almost complex structures interpolating these two families, defined below.

Definition 8.7.

An interpolating family of almost complex structures is a family JAJ_{A} of compatible almost complex structures on MM parametrized by A[0,+)A\in[0,+\infty) such that the following properties are satisfied:

  1. (i)

    J0=JMJ_{0}=J_{M};

  2. (ii)

    For each AA, JA=JMJ_{A}=J_{M} outside U0U_{0};

  3. (iii)

    For each AA, the restriction of JAJ_{A} to MU0M\setminus U_{0} is of contact type in the cylindrical end M×[1,+)\partial M\times[1,+\infty);

  4. (iv)

    There exists A0>0A_{0}>0 such that for A>A0A>A_{0}, the restriction of JAJ_{A} to U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]) agrees with (ψUA)JU(\psi_{U}^{A})_{*}J_{U}, and the restriction of JAJ_{A} to a collar neighborhood of U\partial U in MM is of contact type on some neighborhood of U×{1ϵ}U×[1ϵ,1]\partial U\times\{1-\epsilon\}\subset\partial U\times[1-\epsilon,1].

It is not difficult to show that such interpolating families exist, and in fact that they form a contractible space.

Given an exact cylindrical Lagrangian submanifold LML\subset M satisfying Assumption 8.3, the completion LL^{\prime} of its restriction L0=LU0L^{\prime}_{0}=L\cap U_{0} is exact cylindrical with a locally constant primitive near U\partial U. Call that constant cLc_{L}. We define a slightly shrunk Lagrangian submanifold

(8.6) L1ϵ={ψU1ϵL0, on U0(U×[1ϵ,1]),L×[1ϵ,1], on U×[1ϵ,1],LL^{1-\epsilon}=\begin{cases}\psi_{U}^{1-\epsilon}L^{\prime}_{0},&\text{ on }U_{0}\setminus(\partial U\times[1-\epsilon,1]),\\ \partial L^{\prime}\times[1-\epsilon,1],&\text{ on }\partial U\times[1-\epsilon,1],\\ L\end{cases}

This is well-defined, because LL is invariant under the Liouville flow in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1].

In Floer theory, when defining inhomogeneous pseudoholomorphic maps, we also need to perturb Hamiltonians and almost complex structures. Our strategy is to perturb almost complex structures, in a way such that they depend on domains of the inhomogeneous pseudoholomorphic maps. If the domain is a strip, then an interpolating family should also depend on an additional parameter tt. That is, an interpolating family for the strip is a family of almost complex structures JA,tJ_{A,t}, such that for every fixed tt, JA,tJ_{A,t} is an interpolating family in the sense of Definition 8.7.

Given HAH_{A} and JA,tJ_{A,t} defined as above, we now introduce a Floer datum on the strip. Such a Floer datum, roughly speaking, is a homotopy between (HM,JM,t)(H_{M},J_{M,t}) and (HA,JA,t)(H_{A},J_{A,t}), parametrized by ss\in\mathbb{R}.

Definition 8.8.

A Floer datum on the strip Z=×[0,1]Z=\mathbb{R}\times[0,1] consists of the following data:

  1. (i)

    A family of Hamiltonians HsH_{s} depending on ss\in\mathbb{R}, thought of as a family on the strip ZZ independent of tt, such that for s0s\gg 0, HsH_{s} agrees with HMH_{M}, and for s0s\ll 0, HsH_{s} agrees with HAH_{A}.

  2. (ii)

    A family of almost complex structures J(s,t)J_{(s,t)} parametrized by (s,t)Z(s,t)\in Z, such that for s0s\gg 0, J(s,t)J_{(s,t)} agrees with JM=JM,tJ_{M}=J_{M,t}, and for s0s\ll 0, J(s,t)J_{(s,t)} agrees with JA,tJ_{A,t}.

To define a climbing strip, we also need appropriate Lagrangian boundary conditions.

Definition 8.9.

A moving Lagrangian labeling for the strip Z=×[0,1]Z=\mathbb{R}\times[0,1] is a pair (L0,s,L1,s)(L_{0,s},L_{1,s}) where each Lj,sL_{j,s} is an exact Lagrangian isotopy (through exact cylindrical Lagrangian submanifolds) parametrized by ss\in\mathbb{R}, such that for s0s\gg 0, Lj,s=LjL_{j,s}=L_{j}, and for s0s\ll 0, Lj,s=Lj1ϵL_{j,s}=L_{j}^{1-\epsilon} as defined in (8.6).

Note that there is a natural choice of moving Lagrangian labeling as follows. Let λ:[1ϵ,1]\lambda:\mathbb{R}\to[1-\epsilon,1] be a smooth non-decreasing function which is 1ϵ1-\epsilon for s0s\ll 0 and is 11 for s0s\gg 0. Then we can define Lj,s=Ljλ(s)L_{j,s}=L_{j}^{\lambda(s)} in a similar way to that of (8.6) (replacing 1ϵ1-\epsilon by λ(s)\lambda(s)). Without loss of generality, we may assume that λ(s)=1\lambda(s)=1 for s0s\geq 0.

Definition 8.10.

A climbing strip is a smooth map

w:×[0,1]M,w:\mathbb{R}\times[0,1]\to M,

with the following properties:

  1. (i)

    ww satisfies the inhomogeneous Cauchy-Riemann equation:

    (8.7) sw+J(s,t)(twXHs(w))=0.\partial_{s}w+J_{(s,t)}(\partial_{t}w-X_{H_{s}}(w))=0.
  2. (ii)

    The boundary conditions for ww are:

    (8.8) w(s,0)L0,s,w(s,1)L1,s.w(s,0)\in L_{0,s},w(s,1)\in L_{1,s}.
  3. (iii)

    The asymptotic convergence conditions of ww are:

    limsw(s,)=xA(),\lim\limits_{s\to-\infty}w(s,\cdot)=x_{A}(\cdot),

    for a time-one HAH_{A}-chord xAx_{A} from L0,AL_{0,A} to L1,AL_{1,A}, and

    lims+w(s,)=x(),\lim\limits_{s\to+\infty}w(s,\cdot)=x(\cdot),

    for a time-one HMH_{M}-chord xx from L0L_{0} to L1L_{1}. We require that the HAH_{A}-chord xAx_{A} satisfy a further condition that

    (8.9) Aϵ2𝒜(xA)δ,-A\epsilon^{2}\leq\mathcal{A}(x_{A})\leq\delta,

    for a universal small constant δ\delta, whose meaning is to be explained later.

There are two conditions listed above which need further explanation. First, the asymptotic convergence conditions should be, as usual, Hamiltonian chords for the Hamiltonian HA,sH_{A,s}, as s±s\to\pm\infty. When s+s\to+\infty, HA,s=HMH_{A,s}=H_{M}, and the boundary conditions (L0,s,L1,s)(L_{0,s},L_{1,s}) agree with (L0,L1)(L_{0},L_{1}) when s0s\gg 0, so the asymptotic convergence condition as s+s\to+\infty is given by a time-one HMH_{M}-chord xx from L0L_{0} to L1L_{1}. When ss\to-\infty, HA,s=HAH_{A,s}=H_{A}, and the boundary conditions (L0,s,L1,s)(L_{0,s},L_{1,s}) agree with (L01ϵ,L11ϵ)(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon}) when s0s\ll 0, so the asymptotic convergence condition as s+s\to+\infty is given by a time-one HAH_{A}-chord xAx_{A} from L01ϵL_{0}^{1-\epsilon} to L11ϵL_{1}^{1-\epsilon}. Now we shall see how such a time-one HAH_{A}-chord is related to a time-one HUH_{U}-chord xx^{\prime} from L0L^{\prime}_{0} to L1L^{\prime}_{1}. Note that inside U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]), the Hamiltonian KAK_{A} vanishes, so that HAH_{A} is C2C^{2}-small, which can be taken to agree with 11ϵHUψU1ϵ\frac{1}{1-\epsilon}H_{U}\circ\psi_{U}^{1-\epsilon} (as HUH_{U} is also small, the rescaled Hamiltonian is close to HUH_{U}); in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1], the Hamiltonian HAH_{A} is the rescaling of a quadratic Hamiltonian on UU by a factor AA, i.e. 1AHUψUA\frac{1}{A}H_{U}\circ\psi_{U}^{A}; and outside U0U_{0}, HAH_{A} agrees with HM+Aϵ2H_{M}+A\epsilon^{2}. Regarding the Lagrangian boundary conditions, inside U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]) the Lagrangian submanifold Lj1ϵL_{j}^{1-\epsilon} agrees with ψU1ϵLj,0\psi_{U}^{1-\epsilon}L^{\prime}_{j,0}; and outside U0U_{0}, Lj1ϵL_{j}^{1-\epsilon} agrees with LjL_{j}. Thus, time-one HAH_{A}-chords from L01ϵL_{0}^{1-\epsilon} to L11ϵL_{1}^{1-\epsilon} inside U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]) naturally correspond to time-one HUH_{U}-chords from L0L^{\prime}_{0} to L1L^{\prime}_{1} inside U0U_{0}, and time-one Hamiltonian chords of HAH_{A} which are contained in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1] are in one-to-one correspondence with time-one Hamiltonian chords for HUH_{U} from L0L^{\prime}_{0} to L1L^{\prime}_{1} restricted to a finite part of the cylindrical end, U×[1,Aϵ+1]\partial U\times[1,A\epsilon+1], on which HUH_{U} is quadratic with leading coefficient 11.

The second point that needs explanation is the action constraint (8.9). The reason that we impose the constraint on the action of xx^{\prime} is that for the Hamiltonian KAK_{A}, the Hamiltonian chords are either constants in U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]) or M0U0)M_{0}\setminus U_{0}), and non-constant chords in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1] or in the cylindrical end M×[1,+)\partial M\times[1,+\infty). For the constant chords, the action is uniformly bounded by a small constant, denoted by δ\delta^{\prime}, which depends only on the primitives of the Lagrangian submanifolds and can be chosen very small. Thus if we make a small perturbation of KAK_{A} to HAH_{A}, the Hamiltonian chords which are contained inside U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]) also have action bounded by a small constant δ\delta. For the non-constant chords, we observe that the restriction of the Hamiltonian vector field XKAX_{K_{A}} to some level U×{r}U×[1ϵ,1]\partial U\times\{r\}\subset\partial U\times[1-\epsilon,1] is 2A(r1+ϵ)YU2A(r-1+\epsilon)Y_{\partial U}, where YUY_{\partial U} is the Reeb vector field on the contact boundary U\partial U of U0U_{0}. Thus a Hamiltonian chord on level r[1ϵ,1]r\in[1-\epsilon,1] has action A(r1+ϵ)2-A(r-1+\epsilon)^{2} (there is no extra contribution as the primitive is locally constant there), which is at least Aϵ2-A\epsilon^{2}. Thus the action constraint (8.9) for the Hamiltonian chord xAx_{A} for HAH_{A} is reasonable. By imposing this condition, we mean that for each AA, we only consider climbing strips ww whose asymptotic Hamiltonian chord at -\infty satisfies such constraint on its action.

For each given AA, the moduli space of such climbing strips is denoted by 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x), i.e.

(8.10) 𝒫A(xA,x)={w:w is a climbing strip with respect to the Floer datum (HA,s,JA,(s,t)), with Lagrangian boundary conditions given by (L0,s,L1,s), with asymptotic convergence conditions xA at  and x at +},\begin{split}\mathcal{P}^{A}(x_{A},x)=\{w:&w\text{ is a climbing strip with respect to the Floer datum }(H_{A,s},J_{A,(s,t)}),\\ &\text{ with Lagrangian boundary conditions given by }(L_{0,s},L_{1,s}),\\ &\text{ with asymptotic convergence conditions }x_{A}\text{ at }-\infty\text{ and }x\text{ at }+\infty\},\end{split}

Note that the equation (8.7) is not translation invariant, and we are not varying the parameter AA in this case, so the virtual dimension of the moduli space 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x) is

vdim𝒫A(xA,x)=deg(xA)deg(x).v-\dim\mathcal{P}^{A}(x_{A},x)=\deg(x_{A})-\deg(x).

If the virtual dimension is zero, the moduli space 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x) is a compact smooth manifold of dimension zero, for generic choices of Floer data. Thus, by counting rigid elements in this moduli space, we may define a map

(8.11) r~A1:CW(L0,L1;HM)CW(L01ϵ,L11ϵ;HA),\tilde{r}^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}),

by

(8.12) r~A1(x)=xA:deg(xA)=deg(x)w𝒫A(xA,x)ow,\tilde{r}^{1}_{A}(x)=\sum_{x_{A}:\deg(x_{A})=\deg(x)}\sum_{w\in\mathcal{P}^{A}(x_{A},x)}o_{w},

where ow:oxoxAo_{w}:o_{x}\to o_{x_{A}} is the canonical isomorphism of orientation lines induced by ww. Here CW(L01ϵ,L11ϵ;HA)CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}) is a sub-complex of the wrapped Floer complex CW(L01ϵ,L11ϵ;HA)CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}), generated by time-one HAH_{A}-chords from L01ϵL_{0}^{1-\epsilon} to L11ϵL_{1}^{1-\epsilon} which are contained in U0U_{0}. Because of the natural correspondence between time-one HAH_{A}-chords from L01ϵL_{0}^{1-\epsilon} to L11ϵL_{1}^{1-\epsilon} which are contained in U0U_{0} and time-one HUH_{U}-chords from L0L^{\prime}_{0} to L1L^{\prime}_{1} inside a finite part of UU, this space CW(L01ϵ,L11ϵ;HA)CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}) is related to the wrapped Floer complex CW(L0,L1;HU)CW^{*}(L^{\prime}_{0},L^{\prime}_{1};H_{U}) by the following map

(8.13) τA:CW(L01ϵ,L11ϵ;HA)CW(Aϵ2,δ)(L0,L1;HU),\tau_{A}:CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}),

for some fixed small δ>0\delta>0 such that all time-one HUH_{U}-chords have action less than δ\delta (this can be achieved by choosing the Hamiltonian and the primitives appropriately). This map is a chain-level isomorphism, and in fact sends any time-one HAH_{A}-chord xAx_{A} to its corresponding time-one HUH_{U}-chord xx^{\prime}. By composing the map r~A1\tilde{r}^{1}_{A} with τA\tau_{A}, we get a map

(8.14) rA1=τAr~A1:CW(L0,L1;HM)CW(Aϵ2,δ)(L0,L1;HU).r^{1}_{A}=\tau_{A}\circ\tilde{r}^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}).

To prove that rA1r^{1}_{A} is a cochain map, we need to study the compactification of the moduli space 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x). Because of the asymptotic behavior of the elements ww, it is natural to introduce a compactification

𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x)

of this moduli space, by adding broken climbing strips. In the codimension one boundary strata, broken climbing strips are of the following two types:

  1. (i)

    A pair (w,u)(w,u), where ww is a climbing strip, and uu is an inhomogeneous pseudoholomorphic strip in MM. ww and uu have a common asymptotic convergence condition x0x_{0} at the positive end of ww and the negative end of uu. This occurs because as s+s\to+\infty, the family HA,sH_{A,s} agrees with HMH_{M} (and similarly for the almost complex structures and Lagrangian boundary conditions), so when the energy of a sequence of climbing strips escapes from ++\infty, a (HM,JM)(H_{M},J_{M})-pseudoholomorphic strip breaks out.

  2. (ii)

    A pair (u,w)(u^{\prime},w), where uu^{\prime} is pseudoholomorphic with respect to the Floer datum (HA,JA)(H_{A},J_{A}), and ww is a climbing strip. uu^{\prime} and ww have a common asymptotic convergence condition xA,0x_{A,0} at the positive end of uu^{\prime} and the negative end of ww. This occurs because as ss\to-\infty, the family HA,sH_{A,s} agrees with HAH_{A} (and similarly for the almost complex structures and Lagrangian boundary conditions), so when the energy of a sequence of climbing strips escapes from ++\infty, a (HA,JA)(H_{A},J_{A})-pseudoholomorphic strip breaks out. Moreover, for such an inhomogeneous pseudoholomorphic strip, if the input xA,0x_{A,0} is in the sub-complex CW(L01ϵ,L11ϵ;HA)CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}), the output must also be in this sub-complex.

This implies that we have an isomorphism of the codimension one boundary strata:

(8.15) 𝒫¯A(xA,x)x0deg(x0)=deg(x)𝒫A(xA,x0)×(x0,x)xA,0deg(xA,0)=deg(xA)(xA,xA,0)×𝒫A(xA,0,x).\partial\bar{\mathcal{P}}^{A}(x_{A},x)\cong\coprod_{\begin{subarray}{c}x_{0}\\ \deg(x_{0})=\deg(x)\end{subarray}}\mathcal{P}^{A}(x_{A},x_{0})\times\mathcal{M}(x_{0},x)\cup\coprod_{\begin{subarray}{c}x_{A,0}\\ \deg(x_{A,0})=\deg(x_{A})\end{subarray}}\mathcal{M}(x_{A},x_{A,0})\times\mathcal{P}^{A}(x_{A,0},x).

Here (xA,xA,0)\mathcal{M}(x_{A},x_{A,0}) is the moduli space of (HA,JA)(H_{A},J_{A})-pseudoholomorphic strips with asymptotic convergence conditions xA,xA,0x_{A},x_{A,0}, which can be identified with the moduli space (x,x0)\mathcal{M}(x^{\prime},x^{\prime}_{0}) of (HU,JU)(H_{U},J_{U})-pseudoholomorphic strips in UU, by a similar rescaling argument.

In order for the above isomorphism to hold, we must assume from now on that we have make conformally consistent choice of Floer data for all kinds of strips. As the Lagrangian submanifolds involved are all exact cylindrical, standard transversality methods allow us to prove that, in the case of virtual dimension being one, the compactified moduli space 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x) of virtual dimension one is a compact topological manifold of dimension one, if the Floer data are chosen generically. Thus, by a standard gluing argument, based on on the structure of the codimension-one boundary strata of 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x) as shown in (8.15), we can show:

Lemma 8.11.

The map

rA1:CW(L0,L1;HM)CW(Aϵ2,δ)(L0,L1;HU)r^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})

is a cochain map.

Proof.

In addition to the standard gluing argument, one must ensure that no inhomogeneous pseudoholomorphic strip in U0U_{0} connecting two Hamiltonian chords for HAH_{A} inside U0U_{0} escapes from U0U_{0}. This is because that the hypersurface U×{1}\partial U\times\{1\} is pseudo-convex with respect to the chosen almost complex structure, so we can apply the maximum principle.

However, the problem is that this depends on an extra parameter AA. To remove this ambiguity, we introduce the following trick. Note that the action of any HAH_{A}-chord in the collar neighborhood ×[1ϵ,1]\partial\times[1-\epsilon,1] satisfies:

𝒜HA(xA)=A(r1+ϵ)2,\mathcal{A}_{H_{A}}(x_{A})=-A(r-1+\epsilon)^{2},

if xAx_{A} lies on U×{r}\partial U\times\{r\}. There is no contribution from the primitive ff, because ff is locally constant there, so that fA(xA(1))=fA(xA(0))f_{A}(x_{A}(1))=f_{A}(x_{A}(0)). For the corresponding HUH_{U}-chord xx^{\prime}, the same estimate is satisfied, thus we see that xx^{\prime} is a time-one HUH_{U}-chord contained in U×[1,Aϵ+1]\partial U\times[1,A\epsilon+1]. Thus, there exists δ>0\delta>0 sufficiently small, such that the map rA1r^{1}_{A} in fact descends to a sub-complex

(8.16) rA1:CW(L0,L1;HM)CW(Aϵ2,δ)(L0,L1;HU).r^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}).

To get rid of the dependence of AA, the idea is to take the direct limit as A+A\to+\infty. In order for such direct limit to exist, we must ensure that these maps are compatible with the natural inclusions

iA,A:CW(Aϵ2,δ)(L0,L1;HU)CW(Aϵ2,δ)(L0,L1;HU), for A<A.i_{A,A^{\prime}}:CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})\to CW^{*}_{(-A^{\prime}\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}),\text{ for }A<A^{\prime}.
Lemma 8.12.

There is a way of choosing families of Hamiltonians when defining the various maps involved, such that for every A<AA<A^{\prime}, the following diagram commutes up to chain homotopy

(8.17) CW(L0,L1;HM){CW^{*}(L_{0},L_{1};H_{M})}CW(L01ϵ,L11ϵ;HA){CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})}CW(Aϵ2,δ)(L0,L1;HU){CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})}CW(L01ϵ,L11ϵ;HA){CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A^{\prime}})}CW(Aϵ2,δ)(L0,L1;HU){CW^{*}_{(-A^{\prime}\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})}r~A1\scriptstyle{\tilde{r}^{1}_{A}}r~A1\scriptstyle{\tilde{r}^{1}_{A^{\prime}}}τA\scriptstyle{\tau_{A}}cA,A\scriptstyle{c_{A,A^{\prime}}}iA,A\scriptstyle{i_{A,A^{\prime}}}τA\scriptstyle{\tau_{A^{\prime}}}

Here

cA,A:CW(L01ϵ,L11ϵ;HA)CW(L01ϵ,L11ϵ;HA)c_{A,A^{\prime}}:CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A^{\prime}})

is the continuation map induced by a monotone homotopy between HAH_{A} and HAH_{A^{\prime}}, and

iA,A:CW(Aϵ2,δ)(L0,L1;HU)CW(Aϵ2,δ)(L0,L1;HU)i_{A,A^{\prime}}:CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})\to CW^{*}_{(-A^{\prime}\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U})

is the natural inclusion map.

Proof.

When defining the continuation map cA,Ac_{A,A^{\prime}}, we can choose the family of Hamiltonians HA,A,sH_{A,A^{\prime},s}, such that the composition of HA,sH_{A,s} and HA,A,sH_{A,A^{\prime},s} agrees with the family HA,sH_{A^{\prime},s}, after identifying the glued strip with ×[0,1]\mathbb{R}\times[0,1] by a suitable reparametrization. Thus, by a standard gluing argument, the left triangle commutes up to chain homotopy, which is unique up to higher homotopies.

To prove the homotopy commutativity of the right square, we need to study the continuation map cA,Ac_{A,A^{\prime}} in more details. For this, we first write down a specific choice of the homotopy HA,A,sH_{A,A^{\prime},s}, such that in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1], the homotopy takes the form

(8.18) HA,A,s(y,r)=((1λ(s))A+λ(s)A)(r1+ϵ)2,(y,r)U×[1ϵ,1],H_{A,A^{\prime},s}(y,r)=((1-\lambda(s))A+\lambda(s)A^{\prime})(r-1+\epsilon)^{2},(y,r)\in\partial U\times[1-\epsilon,1],

where λ:[0,1]\lambda:\mathbb{R}\to[0,1] is a smooth non-increasing function, such that λ(s)=1\lambda(s)=1 for s0s\ll 0, and λ(s)=0\lambda(s)=0 for s0s\gg 0. Let xAx_{A} be a time-one HAH_{A}-chord, which corresponds to a time-one HUH_{U}-chord xx^{\prime} under τA\tau_{A}. We shall first prove that, if AA^{\prime} is sufficiently close to AA, then the image of xAx_{A} under cA,Ac_{A,A^{\prime}} is the unique time-one HAH_{A^{\prime}}-chord which corresponds to the same time-one HUH_{U}-chord xx^{\prime} under τA\tau_{A^{\prime}}. There are two cases to consider:

  1. (i)

    xAx_{A} is a small perturbation of a constant Hamiltonian chord for KAK_{A}. Such a Hamiltonian chord is contained in U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]), where KAK_{A} and KAK_{A^{\prime}} are both zero. Thus we can take their small perturbations HAH_{A} and HAH_{A^{\prime}} such that they agree inside U0(U×[1ϵ,1])U_{0}\setminus(\partial U\times[1-\epsilon,1]). Since for a constant homotopy of Hamiltonians the continuation map is necessarily the identity, we conclude that cA,A(xA)c_{A,A^{\prime}}(x_{A}) must be the same Hamiltonian chord as xAx_{A}, now regarded as a Hamiltonian chord for HAH_{A^{\prime}}.

  2. (ii)

    xAx_{A} lies in the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1], and corresponds to a non-constant time-one HUH_{U}-chord xx^{\prime} in the cylindrical end. It is a general property shared by continuation maps that the count of rigid continuation strips does not jump except for a discrete set of AA’s. Thus for AA^{\prime} sufficiently close to AA, we have that cA,A(xA)=xAc_{A,A^{\prime}}(x_{A})=x_{A^{\prime}}, as cA,A=idc_{A,A}=id is the identity map (on the chain level), as it is the continuation map with respect to the constant homotopy of Hamiltonians. In particular, for our specific choice of homotopy of Hamiltonians (8.18), there is an explicit formula for such a continuation strip:

    (8.19) uA,A(s,t)=x(1λ(s))A+λ(s)A(t).u_{A,A^{\prime}}(s,t)=x_{(1-\lambda(s))A+\lambda(s)A^{\prime}}(t).

    In other words, if AA^{\prime} is sufficiently close to AA, then the unique continuation strip which asymptotically converges to xAx_{A} at ++\infty is the trace of all Hamiltonian chords of HA′′H_{A^{\prime\prime}} for varying A′′[A,A]A^{\prime\prime}\in[A,A^{\prime}], starting from the given time-one HAH_{A}-chord xAx_{A}.

This proves that if AA^{\prime} is sufficiently close to AA, the right square strictly commutes on the chain level.

For general A>AA^{\prime}>A, commutativity might not hold on strictly on the chain level, as there might be extra off-diagonal terms appearing in the continuation map. To prove homotopy commutativity we use the following trick. For each B[A,A]B\in[A,A^{\prime}], there is a small δB>0\delta_{B}>0 such that for B(BδB,B+δB)B^{\prime}\in(B-\delta_{B},B+\delta_{B}), we have that there is a choice of almost complex structure JBJ_{B^{\prime}} in a small neighborhood of JBJ_{B} homotopic to JBJ_{B}, for which the relevant moduli spaces of inhomogeneous pseudoholomorphic strips are regular, and the chain-level continuation map

cB,B:CW(L01ϵ,L11ϵ;HB,JB)CW(L01ϵ,L11ϵ;HB,JB),c_{B^{\prime},B}:CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{B^{\prime}},J_{B^{\prime}})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{B},J_{B}),

if BBB^{\prime}\leq B, or

cB,B:CW(L01ϵ,L11ϵ;HB,JB)CW(L01ϵ,L11ϵ;HB,JB)c_{B,B^{\prime}}:CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{B},J_{B})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{B^{\prime}},J_{B^{\prime}})

if BBB\leq B^{\prime}, are diagonal matrices with respect to the natural basis. The intervals (BδB,B+δB)(B-\delta_{B},B+\delta_{B}) form an open cover of [A,A][A,A^{\prime}], from which we choose a finite sub-cover, (B0δB0,B0+δB0),,(BNδBN,BN+δBN)(B_{0}-\delta_{B_{0}},B_{0}+\delta_{B_{0}}),\cdots,(B_{N}-\delta_{B_{N}},B_{N}+\delta_{B_{N}}), where A<B0<<BN<AA<B_{0}<\cdots<B_{N}<A^{\prime}. For each BjB_{j} and Bj+1B_{j+1}, (BjδBj,Bj+δBj)(Bj+1δBj+1,Bj+1+δBj+1)(B_{j}-\delta_{B_{j}},B_{j}+\delta_{B_{j}})\cap(B_{j+1}-\delta_{B_{j+1}},B_{j+1}+\delta_{B_{j+1}}) is non-empty, so we can pick a number BjB^{\prime}_{j} in this intersection. Then by a standard homotopy argument, the continuation map cBj,Bj+1c_{B_{j},B_{j+1}} is chain homotopic to the composition cBj,Bj+1cBj,Bj+1c_{B^{\prime}_{j},B_{j+1}}\circ c_{B^{\prime}_{j},B_{j+1}}. With respect to the natural basis (ordered according to action) for these Floer cochain complexes, both cBj,Bjc_{B_{j},B^{\prime}_{j}} and cBj,Bj+1c_{B^{\prime}_{j},B_{j+1}} can be written as matrices of the form (I,0)(I,0), where II is the identity matrix (the number of columns of cBj,Bjc_{B_{j},B^{\prime}_{j}} should be equal to the number of rows of cBj,Bj+1c_{B^{\prime}_{j},B_{j+1}}). Now the continuation map cA,Ac_{A,A^{\prime}} is chain homotopic to the composition cBN1,BNcB0,B1c_{B_{N-1},B_{N}}\circ\cdots\circ c_{B_{0},B_{1}}, and is therefore chain homotopic to a map κA,A\kappa_{A,A^{\prime}} which can be written as a matrix of the form (I,0)(I,0). Now it is easy to see that the map κA,A\kappa_{A,A^{\prime}}, when replacing cA,Ac_{A,A^{\prime}}, makes the right square strictly commutes, because we have κA,A(xA)=xA\kappa_{A,A^{\prime}}(x_{A})=x_{A^{\prime}} by definition, for the time-one HAH_{A}-chord xAx_{A} and the time-one HAH_{A^{\prime}}-chord xAx_{A^{\prime}} corresponding to the same time-one HUH_{U}-chord xx^{\prime}. Thus, for the continuation map cA,Ac_{A,A^{\prime}} itself, the right square commutes up to chain homotopy.

As an immediate consequence of this lemma, we have the following:

Corollary 8.13.

The homotopy direct limit map

(8.20) r1=limA+rA1:CW(L0,L1;HM)CW(L0,L1;HU)r^{1}=\lim\limits_{A\to+\infty}r^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}(L^{\prime}_{0},L^{\prime}_{1};H_{U})

is well-defined. It is a cochain map.

Proof.

Simply note that the direct limit limA+CW(Aϵ2,δ)(L0,L1;HU)\lim\limits_{A\to+\infty}CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}) is homotopy equivalent to the whole wrapped Floer cochain complex CW(L0,L1;HU)CW^{*}(L_{0},L_{1};H_{U}).

The second item follows immediately from the fact that rA1r^{1}_{A} is a cochain map for every AA.

The map (8.20) is called the linear Viterbo restriction homomorphism.

8.4. Non-linear terms of the Viterbo restriction functor

The higher order terms of the Viterbo restriction functor are defined by counting inhomogeneous pseudoholomorphic maps of similar kind, whose domains are disks with several punctures. These are defined in a slightly different manner, as we shall also include AA as a parameter. Let k2k\geq 2 be a positive integer and let (Zk+1,A)𝒮k+1=k+1×(0,+)(Z^{k+1},A)\in\mathcal{S}^{k+1}=\mathcal{R}^{k+1}\times(0,+\infty) be an element of (the smooth part of) the multiplihedra, where Zk+1Z^{k+1} is a disk with k+1k+1 boundary punctures. The punctures are labeled by z0,,zkz_{0},\cdots,z_{k} in a counterclockwise order, where z0z_{0} is negative, while other punctures are positive. These punctures should come with chosen strip-like ends:

ϵ0:(,0]×[0,1]Zk+1,\epsilon_{0}:(-\infty,0]\times[0,1]\to Z^{k+1},

and

ϵj:[0,+)×[0,1]Zk+1,\epsilon_{j}:[0,+\infty)\times[0,1]\to Z^{k+1},

for j=1,,kj=1,\cdots,k. The boundary component between z0z_{0} and z1z_{1} is denoted by I0I_{0}^{-}, the boundary component between z0z_{0} and zkz_{k} is denoted by IkI_{k}^{-}, and for j=1,,k1j=1,\cdots,k-1, the boundary component between zjz_{j} and zj+1z_{j+1} is denoted by Ij+I_{j}^{+}.

The boundary conditions for (Zk+1,A)(Z^{k+1},A) is specified by the following definition.

Definition 8.14.

A moving Lagrangian label for (Zk+1,A)(Z^{k+1},A) is a collection of families of Lagrangian submanifolds, one for each boundary component of Zk+1Z^{k+1}. The assignment is as follows:

  1. (i)

    Assigned to Ij+I^{+}_{j}, the family of Lagrangian submanifolds is the constant family LjL_{j}, for j=1,,k1j=1,\cdots,k-1;

  2. (ii)

    Assigned to I0I^{-}_{0}, the family of Lagrangian submanifolds is the family L0,A,sL_{0,A,s};

  3. (iii)

    Assigned to IkI^{-}_{k}, the family of Lagrangian submanifolds is the family Lk,A,sL_{k,A,s}.

Here L0,A,s,Lk,A,sL_{0,A,s},L_{k,A,s} are the exact Lagrangian isotopies introduced in Definition LABEL:moving_Lagrangian_label.

To write down inhomogeneous Cauchy-Riemann equations adapted to our setup, we need to introduce an appropriate class of Floer data on these domains.

Definition 8.15.

A Floer datum on (Zk+1,A)(Z^{k+1},A) consists of the following data:

  1. (i)

    A collection of weights λ0,,λk\lambda_{0},\cdots,\lambda_{k}.

  2. (ii)

    A basic one-form βZk+1,A\beta_{Z^{k+1},A} on Zk+1Z^{k+1}, such that over the jj-th strip-like ends it agrees with λjdt\lambda_{j}dt. Here by a basic one-form we mean a sub-closed one-form which vanishes along the boundary of Zk+1Z^{k+1}, and whose differential vanishes in a neighborhood of the boundary of Zk+1Z^{k+1}.

  3. (iii)

    A family of Hamiltonians HZk+1,AH_{Z^{k+1},A} depending on (Zk+1,A)(Z^{k+1},A), such that near the jj-th strip-like end, it agrees with λjHA\lambda_{j}H_{A} up to some conformal rescaling factor ρj\rho_{j};

  4. (iv)

    A family of almost complex structures JZk+1,AJ_{Z^{k+1},A} depending on (Zk+1,A)(Z^{k+1},A), such that near the jj-th strip-like end, it agrees with JAJ_{A} up to the same conformal rescaling factor ρj\rho_{j}.

  5. (v)

    A shifting function ρ(Zk+1,A):Zk+1(0,+)\rho_{(Z^{k+1},A)}:\partial Z^{k+1}\to(0,+\infty) which takes the value ρj\rho_{j} on the boundary of the jj-th strip-like end.

Definition 8.16.

Suppose we have chosen a moving Lagrangian labeling for (Zk+1,A)(Z^{k+1},A), as well as a Floer datum on (Zk+1,A)(Z^{k+1},A). A climbing disk (with k+1k+1 punctures) is a triple (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}) where (Zk+1,A)𝒮k+1(Z^{k+1},A)\in\mathcal{S}^{k+1}, and

wk+1:Zk+1Mw^{k+1}:Z^{k+1}\to M

is a smooth map with the following properties:

  1. (i)

    wk+1w^{k+1} satisfies the inhomogeneous Cauchy-Riemann equation:

    (8.21) (dwβZk+1,AXHZk+1,A(w))+JZk+1,A(dwβZk+1,AXHZk+1,A(w))jZk+1,A=0.(dw-\beta_{Z^{k+1},A}\otimes X_{H_{Z^{k+1},A}}(w))+J_{Z^{k+1},A}\circ(dw-\beta_{Z^{k+1},A}\otimes X_{H_{Z^{k+1}},A}(w))\circ j_{Z^{k+1},A}=0.
  2. (ii)

    The boundary conditions for ww are given by the chosen moving Lagrangian labeling (L0,A,L1,,Lk1,Lk,A)(L_{0,A},L_{1},\cdots,L_{k-1},L_{k,A}).

  3. (iii)

    The asymptotic convergence conditions of wk+1w^{k+1} are:

    limswk+1ϵ0(s,)=x0,A()\lim\limits_{s\to-\infty}w^{k+1}\circ\epsilon_{0}(s,\cdot)=x_{0,A}(\cdot)

    for a time-one HAH_{A}-chord x0,Ax_{0,A} from L0,AL_{0,A} to Lk,AL_{k,A}, and

    lims+wk+1ϵj(s,)=xj()\lim\limits_{s\to+\infty}w^{k+1}\circ\epsilon_{j}(s,\cdot)=x_{j}(\cdot)

    for a time-one HMH_{M}-chord xjx_{j} from Lj1L_{j-1} to LjL_{j}, where j=1,,kj=1,\cdots,k. Similar to the case of climbing strips, we require that the HAH_{A}-chord xA,0x_{A,0} satisfy the following condition

    Aϵ2𝒜(x0,A)δ.-A\epsilon^{2}\leq\mathcal{A}(x_{0,A})\leq\delta.

The definition of a Floer datum easily generalizes to broken disks in the multiplihedra 𝒮¯k+1\bar{\mathcal{S}}^{k+1}. There is also a notion of universal and conformally consistent choices of Floer data, similar to that in the case for ordinary AA_{\infty}-disks and also that in the case of the action-restriction functor. Now let us assume from now on that universal and conformally consistent choices of Floer data have been made for all elements in 𝒮¯k+1\bar{\mathcal{S}}^{k+1} and for all kk.

Note that for each climbing disk (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}), the asymptotic convergence condition x0,Ax_{0,A} at the negative strip-like end depends on the moduli parameter AA. In order to define a moduli space of such climbing disks, such that the Hamiltonian chord at the negative strip-like end can be fixed as the output, we must find a way of identifying these x0,Ax_{0,A} for different AA. This issue can be resolved using the observation that time-one HAH_{A}-chords inside U0U_{0} can be identified with certain time-one HUH_{U}-chords. Let x0x^{\prime}_{0} be the time-one HUH_{U}-chord corresponding to x0,Ax_{0,A}, under the natural map

τA:CW(L0,A,Lk,A;HA)CW(Aϵ2,δ)(L0,Lk;HU).\tau_{A}:CW^{*}_{-}(L_{0,A},L_{k,A};H_{A})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{k};H_{U}).

As discussed in the proof of Lemma 8.12, the image of x0,Ax_{0,A} under the continuation map cA,Ac_{A,A^{\prime}} is a time-one HAH_{A^{\prime}}-chord x0,Ax_{0,A^{\prime}} corresponding to the same HUH_{U}-chord x0x^{\prime}_{0} under the isomorphism τA\tau_{A^{\prime}}. Thus, we may write that asymptotic convergence condition as

(8.22) limswk+1(s,)=τA(x0)(),\lim\limits_{s\to-\infty}w^{k+1}(s,\cdot)=\tau_{A}(x^{\prime}_{0})(\cdot),

for a fixed time-one HUH_{U}-chord x0x^{\prime}_{0} from L0L^{\prime}_{0} to LkL^{\prime}_{k}. Then, the moduli space of such climbing disks, with the given asymptotic convergence conditions x0x^{\prime}_{0} at the negative strip-like end, and x1,,xkx_{1},\cdots,x_{k} at the positive strip-like ends, is denoted by 𝒫k+1(x0;x1,,xk)\mathcal{P}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}). The virtual dimension of this moduli space is

vdim𝒫k+1(x0;x1,,xk)=deg(x0)deg(x1)deg(xk)+k1,v-\dim\mathcal{P}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k})=\deg(x^{\prime}_{0})-\deg(x_{1})-\cdots-\deg(x_{k})+k-1,

where k1k-1 is the dimension of the moduli space 𝒮k+1\mathcal{S}^{k+1} of the underlying domains (Zk+1,A)(Z^{k+1},A).

There is a natural compactification of the moduli space of climbing disks with k+1k+1 punctures, denoted by

𝒫¯k+1(x0;x1,,xk).\bar{\mathcal{P}}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}).

The elements in this compactified moduli space are broken climbing disks, generalizing the broken climbing strips. There are four types of such broken configurations:

  1. (i)

    A pair ((Zk+1,A,wk+1),u)((Z^{k+1},A,w^{k+1}),u), where (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}) is a climbing disk with k+1k+1 punctures, and uu is an inhomogeneous pseudoholomorphic strip in MM. wk+1w^{k+1} and uu have a common asymptotic convergence condition xnewx_{new} at some positive end of wk+1w^{k+1} and the negative end of uu, for some time-one HMH_{M}-chord xnewx_{new}. This occurs as the limit of a sequence of climbing disks where the energy escapes through some positive strip-like end.

  2. (ii)

    A pair (u,(Zk+1,A,wk+1))(u^{\prime},(Z^{k+1},A,w^{k+1})), where uu^{\prime} is an inhomogeneous pseudoholomorphic strip in UU with respect to the Floer datum (HA,JA)(H_{A},J_{A}), and (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}) is a climbing disk with k+1k+1 punctures. uu^{\prime} and wk+1w^{k+1} have a common asymptotic convergence condition xnewx^{\prime}_{new} at the positive end of uu^{\prime} and the negative end of wk+1w^{k+1} (up to a conformal rescaling), for some time-one HUH_{U}-chord xnewx^{\prime}_{new}. This occurs as the limit of a sequence of climbing disks as the energy escapes through the negative strip-like end.

  3. (iii)

    A pair ((Zkm+2,A,wkm+2),u)((Z^{k-m+2},A,w^{k-m+2}),u), where (Zkm+2,A,wkm+2)(Z^{k-m+2},A,w^{k-m+2}) is a climbing disk with km+2k-m+2 punctures (for some m>1m>1) and uu is an inhomogeneous pseudoholomorphic (m+1)(m+1)-punctured disk in MM. wkm+2w^{k-m+2} and uu have a common asymptotic convergence condition xnewx_{new} at some positive end of wkm+2w^{k-m+2}, and the negative end of uu, for some time-one HMH_{M}-chord xnewx_{new}. This occurs as the limit of a sequence of climbing disks for which the parameter AA tends to 0 or remains finite.

  4. (iv)

    A tuple (u,(Zs1+1,A1,ws1+1),,(Zsl+1,Al,wsl+1))(u^{\prime},(Z^{s_{1}+1},A_{1},w^{s_{1}+1}),\cdots,(Z^{s_{l}+1},A_{l},w^{s_{l}+1})), where (Zsj+1,Aj,wsj+1)(Z^{s_{j}+1},A_{j},w^{s_{j}+1}) is a climbing disk with sj+1s_{j}+1 punctures, and uu^{\prime} is an inhomogeneous pseudoholomorphic (l+1)(l+1)-punctured disk with respect to the Floer datum (HA,JA)(H_{A},J_{A}) (for some l>1l>1). uu^{\prime} and wsj+1w^{s_{j}+1} have a common asymptotic convergence condition xnew,jx^{\prime}_{new,j} at the jj-th positive end of uu^{\prime} (up to conformal rescaling), and the negative end of wsj+1w^{s_{j}+1}, for some time-one HUH_{U}-chord xnew,jx^{\prime}_{new,j}. This occurs as the limit of a sequence of climbing disks for which the parameter AA tends to ++\infty.

Note that the broken configurations of type (i) and type (iii) can be written in a uniform way, as (i) is the special case m=1m=1. And the broken configurations of type (ii) and type (iv) can be written in a uniform way, as (ii) is the special case l=1l=1. In case (ii), by an (HA,JA)(H_{A},J_{A})-pseudoholomorphic disk uu^{\prime} with l+1l+1 punctures, we mean it satisfies the inhomogeneous Cauchy-Riemann equation with respect to a domain-dependent family of Hamiltonians which agrees with HAH_{A} (up to rescaling) over the strip-like ends, and a domain-dependent family of almost complex structures which agrees with JAJ_{A} (up to rescaling) over the strip-like ends. Such an inhomogeneous pseudoholomorphic disk then corresponds to an (HU,JU)(H_{U},J_{U})-pseudoholomorphic disk under the conformal rescaling by Liouville flow. In case (iv), such an (HA,JA)(H_{A},J_{A})-pseudoholomorphic strip uu^{\prime} corresponds to an (HU,JU)(H_{U},J_{U})-pseudoholomorphic strip under conformal rescaling. Thus, we have an isomorphism of the codimension one boundary strata:

(8.23) 𝒫¯k+1(x0;x1,,xk)ixnewdeg(xnew)=deg(xi+1)++deg(xi+m)+2k𝒫km+2(x0;x1,,xi,xnew,xi+m+1,,xk)×m+1(xnew,xi+1,,xi+m)s1,,sls1++sl=kxnew,1,,xnew,ldeg(xnew,j)=deg(xs1++sj1+1)++deg(xs1++sj)+1sjl+1(x0,xnew,1,,xnew,l)×𝒫s1+1(xnew,1;x1,,xs1)××𝒫sl+1(xnew,l;xs1++sl1+1,,xk).\begin{split}&\partial\bar{\mathcal{P}}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k})\\ \cong&\coprod_{i}\coprod_{\begin{subarray}{c}x_{new}\\ \deg(x_{new})=\deg(x_{i+1})+\cdots+\deg(x_{i+m})+2-k\end{subarray}}\mathcal{P}_{k-m+2}(x^{\prime}_{0};x_{1},\cdots,x_{i},x_{new},x_{i+m+1},\cdots,x_{k})\\ &\times\mathcal{M}_{m+1}(x_{new},x_{i+1},\cdots,x_{i+m})\\ \cup&\coprod_{\begin{subarray}{c}s_{1},\cdots,s_{l}\\ s_{1}+\cdots+s_{l}=k\end{subarray}}\coprod_{\begin{subarray}{c}x^{\prime}_{new,1},\cdots,x^{\prime}_{new,l}\\ \deg(x^{\prime}_{new,j})=\deg(x_{s_{1}+\cdots+s_{j-1}+1})+\cdots+\deg(x_{s_{1}+\cdots+s_{j}})+1-s_{j}\end{subarray}}\mathcal{M}_{l+1}(x^{\prime}_{0},x^{\prime}_{new,1},\cdots,x^{\prime}_{new,l})\\ &\times\mathcal{P}_{s_{1}+1}(x^{\prime}_{new,1};x_{1},\cdots,x_{s_{1}})\times\cdots\times\mathcal{P}_{s_{l}+1}(x^{\prime}_{new,l};x_{s_{1}+\cdots+s_{l-1}+1},\cdots,x_{k}).\end{split}

Note that the components l+1(x0,xnew,1,,xnew,l)\mathcal{M}_{l+1}(x^{\prime}_{0},x^{\prime}_{new,1},\cdots,x^{\prime}_{new,l}) are moduli spaces of (HU,JU)(H_{U},J_{U})-pseudoholomorphic disks in UU, which enter this picture as the identification of the moduli spaces l+1(x0,A;xnew,1,A,,xnew,l,A)\mathcal{M}_{l+1}(x_{0,A};x_{new,1,A},\cdots,x_{new,l,A}) of (HA,JA)(H_{A},J_{A})-pseudoholomorphic disks, provided that a rigid climbing disk exists for that AA. As before, this isomorphism holds if the Floer data are chosen in a consistent way.

If we choose Floer data generically, the compactified moduli space 𝒫¯k+1(x0;x1,,xk)\bar{\mathcal{P}}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}), when the virtual dimension is zero or one, is a compact smooth/topological manifold of dimension zero/one. In particular, 𝒫k+1(x0;x1,,xk)\mathcal{P}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}) is compact if the dimension is zero, and therefore consists of finitely many points with natural orientations. Then by counting rigid elements in the moduli space 𝒫k+1(x0;x1,,xk)\mathcal{P}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}), we can define multilinear maps

(8.24) rk:CW(Lk1,Lk;HM)CW(L0,L1;HM)CW(L0,Lk;HU).r^{k}:CW^{*}(L_{k-1},L_{k};H_{M})\otimes\cdots\otimes CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}(L^{\prime}_{0},L^{\prime}_{k};H_{U}).

That is, for each rigid climbing disk (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}), we get a canonical isomorphism of orientation lines

o(Zk+1,A,wk+1):ox1oxkox0,A.o_{(Z^{k+1},A,w^{k+1})}:o_{x_{1}}\otimes\cdots\otimes o_{x_{k}}\to o_{x_{0,A}}.

We then compose it with the map τA\tau_{A} to get an isomorphism

ox1oxkox0,o_{x_{1}}\otimes\cdots\otimes o_{x_{k}}\to o_{x^{\prime}_{0}},

and sum over all such (Zk+1,A,wk+1)(Z^{k+1},A,w^{k+1}) and all x0x^{\prime}_{0} to get the desired map rkr^{k}.

Now consider the situation where the virtual dimension of 𝒫¯(x,x)\bar{\mathcal{P}}(x^{\prime},x) is one. Then the structure of the codimension one boundary strata of the moduli spaces 𝒫¯(x,x)\bar{\mathcal{P}}(x^{\prime},x) (8.15) and the moduli spaces 𝒫¯k+1(x0;x1,,xk)\bar{\mathcal{P}}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}) (8.23) immediately implies:

Lemma 8.17.

The multilinear maps

rk:CW(Lk1,Lk;HM)CW(L0,L1;HM)CW(L0,Lk;HU)r^{k}:CW^{*}(L_{k-1},L_{k};H_{M})\otimes\cdots\otimes CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}(L^{\prime}_{0},L^{\prime}_{k};H_{U})

satisfy the equations for AA_{\infty}-functors.

Proof.

The non-trivial part involves taking the direct limit limA+\lim\limits_{A\to+\infty} for various Floer cochain spaces and various maps, but that can be dealt with in a similar way as Lemma ∎

Varying the objects LjL_{j}’s in the full sub-category (M)\mathcal{B}(M), we get an AA_{\infty}-functor

(8.25) r:(M)𝒲(U).r:\mathcal{B}(M)\to\mathcal{W}(U).

To summarize, the spirit of our construction is as follows. Implicit, we have defined a version of homotopy direct limit

limA+𝒲(M;HA)\lim\limits_{A\to+\infty}\mathcal{W}_{-}(M;H_{A})

where 𝒲(M;HA)\mathcal{W}_{-}(M;H_{A}) is an AA_{\infty}-category whose morphism spaces are the truncated wrapped Floer complexes CW(L01ϵ,L11ϵ;HA)CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}), and whose AA_{\infty}-structure maps are defined by counting inhomogeneous pseudoholomorphic disks with respect to Floer data (HA,JA)(H_{A},J_{A}). Equivalently, this category is quasi-isomorphic to a truncated version of the wrapped Fukaya category of UU. There are natural functors

cA:𝒲(U)𝒲(M;HA),c_{A}:\mathcal{W}(U)\to\mathcal{W}_{-}(M;H_{A}),

which are induced by the continuation functors defined in terms of Floer theory in UU, given a homotopy of Hamiltonians from HUH_{U}, which is quadratic with leading coefficient 11, to a Hamiltonian which is quadratic with leading coefficient AA. These functors cAc_{A} are compatible with natural continuation functors

cA,A:𝒲(M;HA)𝒲(M;HA),c_{A,A^{\prime}}:\mathcal{W}_{-}(M;H_{A})\to\mathcal{W}_{-}(M;H_{A^{\prime}}),

so we can take the direct limit

limA+cA:𝒲(U)limA+𝒲(M;HA).\lim\limits_{A\to+\infty}c_{A}:\mathcal{W}(U)\to\lim\limits_{A\to+\infty}\mathcal{W}_{-}(M;H_{A}).

By an action filtration argument, this functor is a quasi-isomorphism. Thus we are able to find a homotopy inverse

k:limA+𝒲(M;HA)𝒲(U).k:\lim\limits_{A\to+\infty}\mathcal{W}_{-}(M;H_{A})\to\mathcal{W}(U).

From this point of view, the main part of our construction is to use the moduli spaces 𝒫k+1(x0;x1,,xk)\mathcal{P}_{k+1}(x^{\prime}_{0};x_{1},\cdots,x_{k}) to define an AA_{\infty}-functor

limA+rA:(M)limA+𝒲(M;HA).\lim\limits_{A\to+\infty}r_{A}:\mathcal{B}(M)\to\lim\limits_{A\to+\infty}\mathcal{W}_{-}(M;H_{A}).

By composing this with the functor kk, we get

r=klimA+rA:(M)𝒲(U).r=k\circ\lim\limits_{A\to+\infty}r_{A}:\mathcal{B}(M)\to\mathcal{W}(U).
Definition 8.18.

The AA_{\infty}-functor

r:(M)𝒲(U)r:\mathcal{B}(M)\to\mathcal{W}(U)

is called the Viterbo restriction functor.

8.5. Comparison between the linear terms

In this subsection, we shall prove the second half of Theorem 1.4. Formulated in a more precise way, what we need to prove is that the linear term of the functor ΘΓ\Theta_{\Gamma}

ΘΓ1:CW(L0,L1;HM)CW(L0,L1;HU)\Theta_{\Gamma}^{1}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}(L^{\prime}_{0},L^{\prime}_{1};H_{U})

is chain homotopic to the linear Viterbo restriction map

r1:CW(L0,L1;HM)CW(L0,L1;HU).r^{1}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}(L^{\prime}_{0},L^{\prime}_{1};H_{U}).

We shall from now on restrict the Viterbo restriction functor to the full sub-category 0(M)\mathcal{B}_{0}(M), and obtain an AA_{\infty}-functor

r:0(M)𝒲(U).r:\mathcal{B}_{0}(M)\to\mathcal{W}(U).

That is to say, we shall consider only Lagrangian submanifolds LL in the full sub-category 0(M)\mathcal{B}_{0}(M), i.e. they satisfy Assumption 8.3 and Assumption 8.4.

When comparing ΘΓ1\Theta_{\Gamma}^{1} to the linear term r1r^{1} of the Viterbo restriction functor, we shall instead work with the cochain map ΠΓ\Pi_{\Gamma} as defined in (7.50). The reason why we study this map is that its construction is straightforward and geometric, without explicitly referring to representability.

Recall that the map ΠΓ\Pi_{\Gamma} is defined using the moduli spaces 𝒰¯(x;x;e0,e1)\bar{\mathcal{U}}(x;x^{\prime};e_{0},e_{1}) of inhomogeneous pseudoholomorphic quilted maps, which are variants of the moduli spaces

𝒰¯l0,l1(α,β;x;y;y0,1,,y0,l0;y1,1,,y1,l1;e0,e1),\bar{\mathcal{U}}_{l_{0},l_{1}}(\alpha,\beta;x;y;y_{0,1},\cdots,y_{0,l_{0}};y_{1,1},\cdots,y_{1,l_{1}};e_{0},e_{1}),

where now there are no those l0+l1l_{0}+l_{1} punctures on the boundary components of the second patch, as the Lagrangian submanifolds are embedded and the bounding cochains b0b_{0} and b1b_{1} vanish. On the other hand, the linear Viterbo restriction map r1r^{1} is the homotopy direct limit of maps rA1r^{1}_{A}, constructed by appropriate count of elements in the moduli spaces of climbing strips 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x). The idea of proving that these two maps coincide up to chain homotopy is to compare these two kinds of relevant moduli spaces, so as to establish a natural bijective correspondence between them. For this purpose, we shall investigate the geometric conditions for inhomogeneous pseudoholomorphic quilted maps in the moduli spaces 𝒰¯(x;x;e0,e1)\bar{\mathcal{U}}(x;x^{\prime};e_{0},e_{1}), and prove that any inhomogeneous psuedoholomorphic quilted map can be converted to a climbing strip in a canonical way.

Let us first recall the picture of these inhomogeneous pseudoholomorphic quilted maps in 𝒰¯(x;x;e0,e1)\bar{\mathcal{U}}(x;x^{\prime};e_{0},e_{1}), or rather, the smooth locus 𝒰(x;x;e0,e1)\mathcal{U}(x;x^{\prime};e_{0},e_{1}). Let (u,v)(u,v) be an inhomogeneous pseudoholomorphic quilted map in 𝒰(x;x;e0,e1)\mathcal{U}(x;x^{\prime};e_{0},e_{1}). Then it satisfies the following conditions:

  1. (i)

    The quilted surface has two patches S¯=(S0,S1)\underline{S}=(S_{0},S_{1}). S0S_{0} is a punctured disk with a positive puncture z01z_{0}^{1} and two special punctures z0,1,z0,2z_{0}^{-,1},z_{0}^{-,2}, with chosen strip-like ends near the punctures. S1S_{1} is a punctured disk with a negative puncture z10z_{1}^{0} and two special punctures z1,1,z1,2z_{1}^{-,1},z_{1}^{-,2}. The quilted surface is obtained by seaming the two patches along the boundary component I0I_{0}^{-} of S0S_{0} between z0,1,z0,2z_{0}^{-,1},z_{0}^{-,2} and the boundary component I1I_{1}^{-} of S1S_{1} between z1,1,z1,2z_{1}^{-,1},z_{1}^{-,2}. The strip-like ends near special punctures form quilted strip-like ends for (z0,1,z1,1(z_{0}^{-,1},z_{1}^{-,1} and (z0,2,z1,2)(z_{0}^{-,2},z_{1}^{-,2}).

  2. (ii)

    u:S0Mu:S_{0}\to M is pseudoholomorphic with respect to the Floer datum (HS0,JS0)(H_{S_{0}},J_{S_{0}}), which maps the two (non-seam) boundary components to L0L_{0} and L1L_{1} repsectively, which asymptotically converges to a time-one HMH_{M}-chord xx from L0L_{0} to L1L_{1} at the puncture z01z_{0}^{1}.

  3. (iii)

    v:S1Uv:S_{1}\to U is pseudoholomorphic with respect to the Floer datum (HS1,JS1)(H_{S_{1}},J_{S_{1}}), which maps the two (non-seam) boundary components to LL^{\prime}, which asymptotically converges to an HUH_{U}-chord xx^{\prime} from L0L^{\prime}_{0} to L1L^{\prime}_{1} at z10z_{1}^{0}.

  4. (iv)

    Over the seam (I0,I1)(I_{0}^{-},I_{1}^{-}), the matching condition for the pair of maps (u,v)(u,v) is given by the Lagrangian correspondence Γ\Gamma, i.e., (u(z),v(z))Γ(u(z),v(z))\in\Gamma for zz on the seam.

  5. (v)

    At the two quilted strip-like ends, (u,v)(u,v) asymptotically converges to the unique generalized chord representing the cyclic ejCW(Lj,Γ,Lj)e_{j}\in CW^{*}(L_{j},\Gamma,L^{\prime}_{j}). In fact, eje_{j} comes from the fundamental chain of LjL^{\prime}_{j}, under the natural identification between the generalized intersections of (Lj,Γ,Lj)(L_{j},\Gamma,L^{\prime}_{j}) and the self-intersections of LjL^{\prime}_{j}.

Given these conditions, we find that the isomorphism class of the underlying quilted surface S¯\underline{S} is unique, i.e. the moduli space of such quilted surfaces is a singleton. Thus we often denote such a quilted map simply by (u,v)(u,v).

The main task of this subsection is to prove the following:

Proposition 8.19.

Suppose the virtual dimension of 𝒰(x;x;e0,e1)\mathcal{U}(x;x^{\prime};e_{0},e_{1}) is zero or one. Let A>0A>0 be any positive number such that

(8.26) 𝒜(x)Aϵ2.\mathcal{A}(x^{\prime})\geq-A\epsilon^{2}.

Among generic choices of Floer data which make 𝒰(x;x;e0,e1)\mathcal{U}(x;x^{\prime};e_{0},e_{1}) regular, we can make a specific choice of Floer data (without losing genericity), for which the moduli space 𝒰(x;x;e0,e1)\mathcal{U}(x;x^{\prime};e_{0},e_{1}) is isomorphic to a moduli space 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) of inhomogeneous pseudoholomorphic maps, which is orientedly cobordant to the moduli space of climbing strips 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x).

Remark 8.20.

For technical reasons, our construction does not immediately give rise to a climbing strip from a given inhomogeneous pseudoholomorphic quilted map, but rather yields a map satisfying the inhomogeneous Cauchy-Riemann equation with respect to the same Floer datum, but different boundary conditions. The moduli space of such inhomogeneous pseudoholomorphic maps are cobordant to the moduli space of climbing strips in a natural way by means of exact Lagrangian isotopies of the boundary conditions, which will be explained in the proof.

In order to turn such a quilted map (S¯,(u,v))(\underline{S},(u,v)) into an inhomogeneous pseudoholomorphic map in MM, we shall consider the Morse-Bott setup of both wrapped Floer cochain spaces CW(Lj;HM)CW^{*}(L_{j};H_{M}) and CW(Lj;HU)CW^{*}(L^{\prime}_{j};H_{U}), for j=0,1j=0,1. This process can be done without repeating the details by regarding LL^{\prime} as a Lagrangian immersion. Recall that the asymptotic conditions over the quilted ends are specified by the cyclic element ejCW(Lj,Γ,Lj)e_{j}\in CW^{*}(L_{j},\Gamma,L^{\prime}_{j}), which in turn corresponds to the homotopy unit of the AA_{\infty}-algebra CW(Lj;HU)CW^{*}(L^{\prime}_{j};H_{U}), under the map (7.29). This homotopy unit is represented by the minimum of a Morse function on LjL^{\prime}_{j}, whose image in UU is inside the interior part U0U_{0} where the Hamiltonian vanishes. Thus we may in fact demand that the families of Hamiltonians (HS0,HS1)(H_{S_{0}},H_{S_{1}}) be zero near the quilted ends: this condition is consistent with the Morse-Bott setup of the wrapped Floer theory for LjL_{j} and LjL^{\prime}_{j}.

A naive trial is to simply ”glue” the two components uu and vv together. Since Γ\Gamma is the completion of the graph of the natural inclusion U0M0U_{0}\to M_{0}, the condition that a point (p,q)M×U(p,q)\in M\times U lies on Γ\Gamma is equivalent to the condition that pp is the image of qq under the natural map i:UMi:U\to M. Thus the matching condition on Γ\Gamma in fact allows us to ”fold” the quilted map to obtain a map w¯\bar{w} in MM, after composing the second component vv with the embedding i:UMi:U\to M. That is, glue the two patches of the quilted surface together and define a map w~\tilde{w} which is uu on S0S_{0} and which is ivi\circ v on S1S_{1}. From the conditions for (S¯,(u,v))(\underline{S},(u,v)) one can immediately see that w~\tilde{w} satisfies the following properties:

  1. (i)

    The domain of the map w¯\bar{w} is a 44-punctured disk. One is a negative puncture ξ0\xi^{0}, which corresponds to the negative puncture of S1S_{1}. One is a positive puncture ξ1\xi^{1}, which corresponds to the positive puncture of S0S_{0}. The other two are special punctures ξ1,,ξ2,\xi^{1,-},\xi^{2,-}, which come from gluing the quilted punctures.

  2. (ii)

    w¯\bar{w} is smooth, as the matching condition on Γ\Gamma means that the maps uu and ivi\circ v agree on the seam along which the two patches are glued together.

  3. (iii)

    w¯\bar{w} has removable singularities at the special punctures ξ1,,ξ2,\xi^{1,-},\xi^{2,-}. This is because the families of Hamiltonians (HS0,HS1)(H_{S_{0}},H_{S_{1}}) are chosen such that they vanish over the two quilted ends, and the quilted map (u,v)(u,v) asymptotically converges to e0e^{\prime}_{0} and eke^{\prime}_{k}, which in the Morse-Bott chain model correspond to the fundamental chains of L0L^{\prime}_{0} and LkL^{\prime}_{k}.

We have thus obtained a map w¯\bar{w} to MM from the inhomogeneous pseudoholomorphic quilted map (u,v)(u,v). However, the domain of the map w¯\bar{w} is of an unfamiliar form, and it does not satisfy the desired inhomogeneous Cauchy-Riemann equation for a climbing strip. It is therefore necessary to perform suitable modification on the map w¯\bar{w}. By condition (iii), we may take conformal transformations mapping S0S_{0} to the positive half-strip Z+=[0,+)×[0,1]Z_{+}=[0,+\infty)\times[0,1], sending the positive puncture to ++\infty, and the two quilted punctures to the corner points (0,0),(0,1)(0,0),(0,1), and mapping S1S_{1} to the negative half-strip Z=(,0]×[0,1]Z_{-}=(-\infty,0]\times[0,1], sending the negative puncture to -\infty, and the two quilted punctures to the corner points (0,0),(0,1)(0,0),(0,1). Glue these two half-strips together along the common boundary {0}×[0,1]\{0\}\times[0,1], which correspond to the seam of the quilted surface. The result of gluing is a strip with two marked points z1,+=(0,0),z2,+=(0,1)z_{1,+}=(0,0),z_{2,+}=(0,1). This strip is the domain for our new map. In addition, condition (iii) also implies that the incidence conditions for these two marked points are given by the fundamental chains of L0L^{\prime}_{0} and L1L^{\prime}_{1}, thus these incidence conditions are in fact free conditions.

Now we shall construct Floer datum on this new domain from the original Floer datum for the quilted surface S¯\underline{S}. The guiding principle is to use the Liouville structure to rescale the original Floer datum to obtain the new one. Without loss of generality, we may assume that A>1A>1, as when taking the direct limit limA+\lim\limits_{A\to+\infty} we only have to consider sufficiently large AA. Fix a choice of a smooth increasing function

ρA:(,0][1,A]\rho_{A}:(-\infty,0]\to[1,A]

such that for s0s\ll 0, ρA(s)=A\rho_{A}(s)=A and ρA(s)=1\rho_{A}(s)=1 for ss close to 0. We can also extend this function to the whole of \mathbb{R} by setting ρA(s)=1\rho_{A}(s)=1 for all s0s\geq 0.

First let us discuss how to obtain a family of Hamiltonians from the given families of Hamiltonians (HS0,HS1)(H_{S_{0}},H_{S_{1}}) for the quilted surface. We define a family of Hamiltonians HA,sH_{A,s} on MM parametrized by ZZ, which depends only on the coordinate ss\in\mathbb{R}, by setting

(8.27) HA,s={1ρA(s)HZ,sψUρA(s)i1, on Z,HZ+,s, on Z+,H_{A,s}=\begin{cases}\frac{1}{\rho_{A}(s)}H_{Z_{-},s}\circ\psi_{U}^{\rho_{A}(s)}\circ i^{-1},&\text{ on }Z_{-},\\ H_{Z_{+},s},&\text{ on }Z_{+},\end{cases}

where HZH_{Z_{-}} is the family HS1H_{S_{1}} composed with the conformal transformation from ZZ_{-} to S1S_{1}, and HZ+H_{Z_{+}} is the family HS0H_{S_{0}} composed with the conformal transformation from Z+Z_{+} to S0S_{0}. Here ϵ\epsilon is the small constant that we have fixed for which the collar neighborhood U×[1ϵ,1]\partial U\times[1-\epsilon,1] embeds into U0U_{0}. And 1ρA(s)HZ,sψUρA(s)i1\frac{1}{\rho_{A}(s)}H_{Z_{-},s}\circ\psi_{U}^{\rho_{A}(s)}i^{-1} means a Hamiltonian on MM which takes value 1ρA(s)HZ,sψUρA(s)(v(s,t))\frac{1}{\rho_{A}(s)}H_{Z_{-},s}\circ\psi_{U}^{\rho_{A}(s)}(v(s,t)) at the point iv(s,t)Mi\circ v(s,t)\in M, and is constant (equal to Aϵ2A\epsilon^{2}) elsewhere. In fact, we do not have to specify what values the Hamiltonian take for points outside i(U)i(U), as the image of the point on the domain of the inhomogeneous pseudoholomorphic map at which the domain-dependent family of Hamiltonians is evaluated is inside i(U)i(U). A priori, the resulting family of Hamiltonians also depends on the tt-coordinate, but we may choose (HS0,HS1)(H_{S_{0}},H_{S_{1}}) appropriately such that the resulting families HZ+H_{Z_{+}} and HZH_{Z_{-}} is independent of tt (though this is indeed irrelevant for our purpose). This is possible, because of the following reasons: over the positive strip-like end S0S_{0}, HS0H_{S_{0}} agrees with HMH_{M}, and over the negative strip-like end of S1S_{1}, HS1H_{S_{1}} agrees with HUH_{U}, while over the quilted ends, (HS0,HS1)(H_{S_{0}},H_{S_{1}}) vanish - all of these are independent of tt.

For the almost complex structure, given the families of almost complex structures (JS0,JS1)(J_{S_{0}},J_{S_{1}}) for the quilted surface, we define a family of almost complex structures JA,(s,t)J_{A,(s,t)} on MM parametrized by (s,t)Z(s,t)\in Z, by setting

(8.28) JA,(s,t)={(ψUρA(s))JZ,(s,t)i1, on Z,JZ+,(s,t), on Z+.J_{A,(s,t)}=\begin{cases}(\psi_{U}^{\rho_{A}(s)})_{*}J_{Z_{-},(s,t)}\circ i^{-1},&\text{ on }Z_{-},\\ J_{Z_{+},(s,t)},&\text{ on }Z_{+}.\end{cases}

We must explain why (8.27) and (8.28) are well-defined. There are several cases to discuss, depending on the positions of the Hamiltonian chords xx^{\prime} and xx to which the quilted map (S¯,(u,v))(\underline{S},(u,v)) converges:

  1. (i)

    xx^{\prime} is a small perturbation of a constant Hamiltonian chord in U0U_{0}, and xx is a small perturbation of a constant Hamiltonian chord in M0M_{0}. In this case, by a maximum principle argument, it is necessary that, for the quilted map (u,v)(u,v), the image of uu is contained in M0M_{0} and the image of vv is contained in U0U_{0}. Thus, the matching condition on Γ\Gamma simply implies that u(0,t)=v(0,t)u(0,t)=v(0,t). If furthermore xx is contained in U0U_{0}, then in fact x=xx^{\prime}=x and the quilted map is constant. This is essentially the only case that needs some discussion. In this case, we may assume our choice of HMH_{M} agrees with the rescaling of HUH_{U} inside U0U_{0}, so that the above formula (8.27) is consistent. Now if xx is in M0U0M_{0}\setminus U_{0}, then such a quilted map (u,v)(u,v) is non-trivial. In this case we may choose (HS0,HS1)(H_{S_{0}},H_{S_{1}}) such that HZ,0=0H_{Z_{-},0}=0 and HZ+,0=0H_{Z_{+},0}=0. Thus it is automatic that the two formulas in (8.27) agree when s=0s=0.

  2. (ii)

    xx^{\prime} is a small perturbation of a constant Hamiltonian chord in U0U_{0}, while xx is a non-constant HMH_{M}-chord in the cylindrical end M×[1,+)\partial M\times[1,+\infty). This is similar to the previous case, since such a quilted map is necessarily non-constant, so that we have the freedom to choose the families (HS0,HS1)(H_{S_{0}},H_{S_{1}}) so that they are zero along the seam.

  3. (iii)

    xx^{\prime} is a non-constant HUH_{U}-chord in the cylindrical end U×[1,+)\partial U\times[1,+\infty), while xx is a small perturbation of a constant Hamiltonian chord in M0M_{0}. This xx^{\prime} corresponds to a time-one chord xAx_{A} for the rescaled Hamiltonian. Then there is in fact no such quilted map or climbing strip, as the action of xx^{\prime} (or xAx_{A}) is very negative, and the action of xx is positive, which is of course greater than the action of xx^{\prime} (or xAx_{A}). Thus it is not necessary to consider this case.

  4. (iv)

    xx^{\prime} is a non-constant HUH_{U}-chord in the cylindrical end U×[1,+)\partial U\times[1,+\infty), and xx is a non-constant HMH_{M}-chord in the cylindrical end M×[1,+)\partial M\times[1,+\infty). Now xx^{\prime} corresponds to a unique time-one HAH_{A}-chord xAx_{A}, and the quilted map is non-constant. The discussion is similar to those in cases (i) and (ii).

Thus we have shown that (8.27) is well-defined. For the family of almost complex structures, a parallel argument implies that (8.28) is well-defined.

To construct a climbing strip from the given quilted map (u,v)(u,v), and A>0A>0, we first define a map

(8.29) w0(s,t)={iψUρA(s)v(s,t), if s<0,u(s,t), if s0.w_{0}(s,t)=\begin{cases}i\circ\psi_{U}^{\rho_{A}(s)}\circ v(s,t),&\text{ if }s<0,\\ u(s,t),&\text{ if }s\geq 0.\end{cases}

This is well-defined and smooth, since when s=0s=0, we have ρA(0)=1\rho_{A}(0)=1 so that the first formula reads iv(0,t)i\circ v(0,t), which is equal to u(0,t)u(0,t) by the matching condition on Γ\Gamma. By definition, the boundary conditions for the map w0w_{0} are as follows:

(8.30) w0(s,j){i(ψUρA(s)Lj), if s<0,Lj, if s0,where j=0,1.w_{0}(s,j)\in\begin{cases}i(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j}),&\text{ if }s<0,\\ L_{j},&\text{ if }s\geq 0,\end{cases}\text{where }j=0,1.

Note that when s0s\leq 0 is close to zero, we have ρA(s)=1\rho_{A}(s)=1 so that i(ψUρA(s)Lj)=i(Lj)Lji(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j})=i(L^{\prime}_{j})\subset L_{j}, because of Assumption 8.4. Let us rewrite the boundary conditions as

w0(s,j)Lj,A,s,w_{0}(s,j)\in L_{j,A,s},

for

(8.31) Lj,A,s:={i(ψUρA(s)Lj), if s<0,Lj, if s0.L_{j,A,s}:=\begin{cases}i(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j}),&\text{ if }s<0,\\ L_{j},&\text{ if }s\geq 0.\end{cases}

When s0s\geq 0, the boundary condition can also be written as Lj,A,s=ψMρA(s)LjL_{j,A,s}=\psi_{M}^{\rho_{A}(s)}L_{j}, as we have extended ρA\rho_{A} such that ρA(s)=1\rho_{A}(s)=1 for all s0s\geq 0. Since ii is the map induced by the Liouville strucutre, it maps LjL^{\prime}_{j} into LjL_{j}, and i(ψUρA(s)Lj)i(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j}) is mapped into a unique exact cylindrical Lagrangian submanifold of MM whose cylindrical end is contained in that of LjL_{j}. In fact, the completion of i(ψUρA(s)Lj)i(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j}) is ψMALj\psi_{M}^{A}L_{j}, because LjL_{j} is assumed to be invariant under the Liouville flow in the Liouville cobordism M0int(U0)M_{0}\setminus int(U_{0}). Thus Lj,A,sL_{j,A,s} can be regarded as a family of Lagrangian submanifolds of MM parametrized by ss\in\mathbb{R}, if we understand i(ψUρA(s)Lj)i(\psi_{U}^{\rho_{A}(s)}L^{\prime}_{j}) as its completion. Moreover, this family is an exact Lagrangian isotopy. Such boundary conditions are slightly different from those for a climbing strip. The way to fix this is to use Lagrangian isotopy to move one kind of boundary conditions to the other - such an argument will be presented later.

Note that w0w_{0} satisfies the inhomogeneous Cauchy-Riemann equation on the positive half-strip Z+Z_{+}, with respect to the Floer datum (HA,s,JA,(s,t))(H_{A,s},J_{A,(s,t)}) defined as above, as on that region the Floer datum is simply given by the original one (HS0,JS0)(H_{S_{0}},J_{S_{0}}) for the quilted surface. On the negative half-strip ZZ_{-}, w0w_{0} might not be pseudoholomorphic, because an extra term appears when differentiating iψUρA(s)v(s,t)i\circ\psi_{U}^{\rho_{A}(s)}\circ v(s,t) with respect to ss, which is related to the differential of ψUρA(s)\psi_{U}^{\rho_{A}(s)}. By a straightforward calculation, we have

(8.32) s(iψUρA(s)v(s,t))=diψUρA(s)v(s,t)(dψUρA(s))v(s,t)ρA(s)(sv(s,t)),\partial_{s}(i\circ\psi_{U}^{\rho_{A}(s)}\circ v(s,t))=di_{\psi_{U}^{\rho_{A}(s)}\circ v(s,t)}\circ(d\psi_{U}^{\rho_{A}(s)})_{v(s,t)}\rho^{\prime}_{A}(s)(\partial_{s}v(s,t)),

where ρA(s)\rho^{\prime}_{A}(s) is the derivative of the function ρA\rho_{A} with respect to ss. On the other hand, in the inhomogeneous term contributed by the Hamiltonian vector field, there is not such a factor, as we have for s<0s<0,

(8.33) XHA,s=ρA(s)didψUρA(s)XHZ,X_{H_{A,s}}=\rho_{A}(s)di\circ d\psi_{U}^{\rho_{A}(s)}X_{H_{Z_{-}}},

as calculating the Hamiltonian vector field from the Hamiltonian does not involve differentiation with respect to the variable ss. In order to obtain a map which satisfies the inhomogeneous Cauchy-Riemann equation, we must therefore perturb the map w0w_{0}. In fact, we may choose the function ρA\rho_{A} suitably such that its derivative is small, say

|ρA(s)|<Cinj(M,g),|\rho^{\prime}_{A}(s)|<C\cdot inj(M,g),

where inj(M,g)inj(M,g) is the injectivity radius of a family of metrics g=g(s,t)g=g(s,t) on MM for which the Lagrangian submanifolds L0,A,s,L1,A,sL_{0,A,s},L_{1,A,s} are totally geodesic, and CC is an appropriate constant to be determined in the proof of the following perturbation lemma.

Lemma 8.21.

For any given A>0A>0 sufficiently large, and for each w0w_{0} defined as above, there is a unique map w~:ZM\tilde{w}:Z\to M closest to w0w_{0}, which satisfies the inhomogeneous Cauchy-Riemann equation:

(8.34) sw+JA,(s,t)(twXHA,s(w))=0,\partial_{s}w+J_{A,(s,t)}(\partial_{t}w-X_{H_{A,s}}(w))=0,

and the same boundary conditions L0,A,s,L1,A,sL_{0,A,s},L_{1,A,s} as those for w0w_{0}.

Proof.

For the point w0w_{0}\in\mathcal{B}, there is the exponential map

Expw0:OTw0,Exp_{w_{0}}:O\subset T_{w_{0}}\mathcal{B}\to\mathcal{B},

defined with respect to a family of metrics g=g(s,t)g=g(s,t) on MM parametrized by (s,t)×[0,1](s,t)\in\mathbb{R}\times[0,1], for which L0,A,sL_{0,A,s} and L1,A,sL_{1,A,s} are totally geodesic. Here \mathcal{B} is the Banach manifold (with respect to some Sobolev norm Wm,pW^{m,p}) of maps satisfying the same conditions for a climbing strip, except the inhomogeneous Cauchy-Riemann equation. We require that OO is an open neighborhood of zero such that the exponential map Expw0Exp_{w_{0}} is an isomorphism onto the image: the size of OO is at least the injectivity radius of the family of metrics gg:

inj(M,g)=min(s,t)×[0,1]inj(M,g(s,t)).inj(M,g)=\min_{(s,t)\in\mathbb{R}\times[0,1]}inj(M,g(s,t)).

In fact, g=g(s,t)g=g(s,t) is determined by the symplectic form and the chosen family of almost complex structures JA,(s,t)J_{A,(s,t)}. As a result, when s0s\ll 0 or s0s\gg 0, g(s,t)g(s,t) agrees with a family independent of ss. Therefore, the minimum over (s,t)×[0,1](s,t)\in\mathbb{R}\times[0,1] is indeed taken over a compact set, hence well-defined. The tangent space Tw0T_{w_{0}}\mathcal{B} is

Tw0={VWm,p(Z;w0TM;w0TL0,A,s,w0TL1,A,s}.T_{w_{0}}\mathcal{B}=\{V\in W^{m,p}(Z;w_{0}^{*}TM;w_{0}^{*}TL_{0,A,s},w_{0}^{*}TL_{1,A,s}\}.

Here the boundary conditions mean that V(s,0)Tw0(s,0)L0,A,s,V(s,1)Tw1(s,1)L1,A,sV(s,0)\in T_{w_{0}(s,0)}L_{0,A,s},V(s,1)\in T_{w_{1}(s,1)}L_{1,A,s}.

For any VOV\in O, consider the inhomogeneous Cauchy-Riemann equation

(8.35) sExpw0(V)+JA,(s,t)(tExpw0(V)XHA,s(Expw0(V)))=0.\partial_{s}Exp_{w_{0}}(V)+J_{A,(s,t)}(\partial_{t}Exp_{w_{0}}(V)-X_{H_{A,s}}(Exp_{w_{0}}(V)))=0.

We wish to find a solution VOV\in O to this equation. We denote the inhomogeneous Cauchy-Riemann operator with respect to the Floer datum (HA,s,JA,(s,t))(H_{A,s},J_{A,(s,t)}) by

¯A()=s()+JA,(s,t)(t()XHA,s()).\bar{\partial}_{A}(\cdot)=\partial_{s}(\cdot)+J_{A,(s,t)}(\partial_{t}(\cdot)-X_{H_{A,s}}(\cdot)).

Although w0w_{0} does not satisfy the equation ¯Aw0=0\bar{\partial}_{A}w_{0}=0, we have chosen a connection on the bundle w0TMw_{0}^{*}TM relative to (w0TL0,A,s,w0TL1,A,s)(w_{0}^{*}TL_{0,A,s},w_{0}^{*}TL_{1,A,s}) when defining the exponential map ExpExp, so that the linearized operator at w0w_{0}

(8.36) Dw0¯A:Tw0w0D_{w_{0}}\bar{\partial}_{A}:T_{w_{0}}\mathcal{B}\to\mathcal{E}_{w_{0}}

is well-defined, where w0\mathcal{E}_{w_{0}} is the space of (0,1)(0,1)-forms with values in w0TMw_{0}^{*}TM, namely w0=Wk,p(Z;w0TMΛZ0,1)\mathcal{E}_{w_{0}}=W^{k,p}(Z;w_{0}^{*}TM\otimes\Lambda^{0,1}_{Z}). The regularity assumption for the moduli space 𝒫A(x,x)\mathcal{P}^{A}(x^{\prime},x) implies that the linearized operator Dw¯AD_{w}\bar{\partial}_{A} is surjective for any ww which satisfies the equation ¯Aw=0\bar{\partial}_{A}w=0. However, this is an open condition - for those maps close to an actual solution, the linearized operator is also surjective. In particular, as w0w_{0} is close to an actual solution, Dw0¯AD_{w_{0}}\bar{\partial}_{A} is surjective. It follows that there is a bounded right inverse Qw0,AQ_{w_{0},A} of Dw0¯AD_{w_{0}}\bar{\partial}_{A}. Now if we set

C=1Qw0,A,C=\frac{1}{||Q_{w_{0},A}||},

then the vector

(8.37) V=Qw0(¯Aw0)V=Q_{w_{0}}(-\bar{\partial}_{A}w_{0})

is contained in the neighborhood OO. This is the desired solution. The meaning of this formula is as follows. As w0w_{0} does not satisfy the inhomogeneous Cauchy-Riemann equation, ¯Aw0\bar{\partial}_{A}w_{0} is non-zero vector in w0\mathcal{E}_{w_{0}}. Note that the exponential map ExpExp provides a linear structure on a neighborhood of w0w_{0} in \mathcal{B} by identifying that with OTw0O\subset T_{w_{0}}\mathcal{B}. Thus, we may ”add” the inverse image of the negative of this non-zero vector ¯Aw0\bar{\partial}_{A}w_{0} to the original map w0w_{0} to kill the deviation from being zero, so that the resulting map satisfies the inhomogeneous Cauchy-Riemann equation.

On the other hand, it is not hard to see such a solution VV is unique.

Let us now denote by 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) the moduli space of maps w~\tilde{w} as above satisfying the inhomogeneous Cauchy-Riemann equation with respect to the Floer datum (HA,s,JA,(s,t)(H_{A,s},J_{A,(s,t)} as well as the Lagrangian boundary conditions (L0,A,s,L1,A,s)(L_{0,A,s},L_{1,A,s}). This moduli space behaves very similarly to the moduli space of climbing strips 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x), and has a natural compactification 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x), by adding broken maps whose new components are inhomogeneous pseudoholomorphic strips in MM with respect to the Floer datum (HM,JM)(H_{M},J_{M}) and Lagrangian boundary conditions (L0,L1)(L_{0},L_{1}), or those with respect to the Floer datum (HA,JA,t)(H_{A},J_{A,t}) and Lagrangian boundary conditions given by the completions of (i(ψUAL0),i(ψUAL1)(i(\psi_{U}^{A}L^{\prime}_{0}),i(\psi_{U}^{A}L^{\prime}_{1}), whose asymptotic Hamiltonian chords are contained inside U0U_{0}. The latter kind of inhomogeneous pseudoholomorphic strips are also in one-to-one correspondence with inhomogeneous pseudoholomorphic strips in UU with respect to the Floer datum (HU,JU)(H_{U},J_{U}) and Lagrangian boundary conditions (L0,L1)(L^{\prime}_{0},L^{\prime}_{1}).

Thus, the above lemma provides a natural bijection between the moduli spaces

𝒰(x;x;e0,e1)𝒫~A(xA,x),\mathcal{U}(x;x^{\prime};e_{0},e_{1})\cong\tilde{\mathcal{P}}^{A}(x_{A},x),

when the virtual dimension is zero, for generic choices of Floer data. Then it remains to extend this bijection to the level of compactified moduli spaces.

Lemma 8.22.

Consider the situation where the virtual dimension of the moduli space 𝒰¯(x,x;e0,e1)\bar{\mathcal{U}}(x,x^{\prime};e_{0},e_{1}) is zero or one. Suppose we have chosen Floer data generically such that both the compactified moduli space 𝒰¯(x,x;e0,e1)\bar{\mathcal{U}}(x,x^{\prime};e_{0},e_{1}) and the compactified moduli space 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x) are regular. This means that these are compact smooth/topological manifolds of dimension zero/one. Then, among the generic choices of Floer data, there is a specific kind of choices for which there is a natural bijection from 𝒰¯(x,x;e0,e1)\bar{\mathcal{U}}(x,x^{\prime};e_{0},e_{1}) to 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x).

Moreover, when the virtual dimension is one, this bijection is an isomorphism of moduli spaces, in the sense that it comes with a natural virtual isomorphism of Fredholm complexes, and commutes with the gluing maps.

Proof.

The assignment (u,v)w(u,v)\mapsto w gives the natural bijection between elements of the uncompactified moduli spaces:

𝒰(x,x;e0,e1)𝒫~A(xA,x).\mathcal{U}(x,x^{\prime};e_{0},e_{1})\cong\tilde{\mathcal{P}}^{A}(x_{A},x).

whenever the virtual dimension is zero or one. If the virtual dimension is zero, these moduli spaces are compact, because we have chosen Floer data generically so that all such moduli spaces as well as moduli spaces of inhomogeneous pseudoholomorphic strips are regular. Thus there is nothing more to prove.

Now consider the case where the virtual dimension is one. Note that the two compactifications are both obtained by adding the same kinds of inhomogeneous pseudoholomorphic strips in MM or inhomogeneous pseudoholomorphic strips in UU. That is to say, there are isomorphisms:

𝒰¯(x,x;e0,e1)(x,x1)×𝒰(x,x1;e0,e1)𝒰(x1,x;e0,e1)×(x1,x).\begin{split}&\partial\bar{\mathcal{U}}(x,x^{\prime};e_{0},e_{1})\\ \cong&\coprod\mathcal{M}(x^{\prime},x^{\prime}_{1})\times\mathcal{U}(x,x^{\prime}_{1};e_{0},e_{1})\\ &\cup\coprod\mathcal{U}(x_{1},x^{\prime};e_{0},e_{1})\times\mathcal{M}(x_{1},x).\end{split}

and

𝒫~¯A(xA,x)~(xA,x1,A)×𝒫~A(x1,A,x)𝒫~A(xA,x1)×(x1,x).\begin{split}&\partial\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x)\\ \cong&\coprod\tilde{\mathcal{M}}(x_{A},x_{1,A})\times\tilde{\mathcal{P}}^{A}(x_{1,A},x)\\ &\cup\coprod\tilde{\mathcal{P}}^{A}(x_{A},x_{1})\times\mathcal{M}(x_{1},x).\end{split}

Here ~(xA,x1,A)\tilde{\mathcal{M}}(x_{A},x_{1,A}) is the moduli space of inhomogeneous pseudoholomorphic strips fAf_{A} with respect to the Floer datum (HA,JA,t)(H_{A},J_{A,t}) and Lagrangian boundary conditions given by the completions of (i(ψUAL0),i(ψUAL1)(i(\psi_{U}^{A}L^{\prime}_{0}),i(\psi_{U}^{A}L^{\prime}_{1}), whose asymptotic Hamiltonian chords are contained inside U0U_{0}. This is naturally isomorphic to (x,x1)\mathcal{M}(x^{\prime},x^{\prime}_{1}), the moduli space of inhomogeneous pseudoholomorphic strips ff^{\prime} in UU with respect to the Floer datum (HU,JU)(H_{U},J_{U}) and Lagrangian boundary conditions (L0,L1)(L^{\prime}_{0},L^{\prime}_{1}), when the time-one HUH_{U}-chords xx^{\prime} and x1x^{\prime}_{1} correspond to the time-one HAH_{A}-chords xAx_{A} and x1,Ax_{1,A} respectively. Thus, for any kind of broken quilted map (f,(u,v))(f^{\prime},(u,v)) or ((u,v),f)((u,v),f), there is a unique broken map (fA,w~)(f_{A},\tilde{w}) or (w~,f)(\tilde{w},f) associated to it, where fAf_{A} and ff^{\prime} correspond to each other under the above-mentioned isomorphism between ~(xA,x1,A)\tilde{\mathcal{M}}(x_{A},x_{1,A}) and (x,x1)\mathcal{M}(x^{\prime},x^{\prime}_{1}). In this way, the bijection extends over the compactifications.

This bijective correspondence naturally commutes with the gluing maps, because gluing happens near the usual strip-like ends, not the quilted ends.

As a corollary, this implies that when choosing Floer data generically in such special class, the counts of elements in these moduli spaces are equal. The algebraic consequence of this can be stated as follows. We have the wrapped Floer cochain space CW(ψMAL0,ψMAL1;HA)CW^{*}(\psi_{M}^{A}L_{0},\psi_{M}^{A}L_{1};H_{A}), on which the differential is defined by counting rigid elements in the moduli spaces ~(xA,x1,A)\tilde{\mathcal{M}}(x_{A},x_{1,A}) of inhomogeneous pseudoholomorphic strips. And we also have a sub-complex CW(ψMAL0,ψMAL1;HA)CW^{*}_{-}(\psi_{M}^{A}L_{0},\psi_{M}^{A}L_{1};H_{A}), generated by ”interior” Hamiltonian chords. In fact, for this sub-complex, one can equivalently write it as CW(i(ψUAL0),i(ψUAL1);HA)CW^{*}_{-}(i(\psi_{U}^{A}L^{\prime}_{0}),i(\psi_{U}^{A}L^{\prime}_{1});H_{A}), because any inhomogeneous pseudoholomorphic strip with asymptotic convergence conditions given by those Hamiltonian chords will be contained in the image of ii. In a similar way to the definition of the map

r~A1:CW(L0,L1;HM)CW(L01ϵ,L11ϵ;HA),\tilde{r}^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A}),

we may define a map

(8.38) tA1:CW(L0,L1;HM)CW(i(ψUAL0),i(ψUAL1);HA)t^{1}_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*}_{-}(i(\psi_{U}^{A}L^{\prime}_{0}),i(\psi_{U}^{A}L^{\prime}_{1});H_{A})

by counting rigid elements in the moduli space 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x). Then Lemma 8.22 implies that ΠΓ=tA1\Pi_{\Gamma}=t^{1}_{A}, when the former map is restricted to the sub-complex generated by those generators whose images under ΠΓ\Pi_{\Gamma} fall within the action filtration window (Aϵ2,δ)(-A\epsilon^{2},\delta).

The remaining task is to compare the map tA1t^{1}_{A} with rA1r^{1}_{A}. For that purpose, the underlying geometric idea is to relate the moduli space 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) to the moduli space 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x) of climbing strips. As mentioned before, the difference between a map w~\tilde{w} obtained from a quilted map (u,v)(u,v) and an actual climbing strip ww is that they have different boundary conditions. However, as we shall see, their moduli spaces are naturally cobordant to each other.

Lemma 8.23.

The moduli space 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) is orientedly cobordant to the moduli space 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x) of climbing strips. Moreover, the same holds for compactified moduli spaces 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x) and 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x), namely they are also cobordant.

Proof.

For each ss\in\mathbb{R}, the exact cylindrical Lagrangian submanifold Lj,A,sL_{j,A,s} as defined in (8.31) is exact Lagrangian isotopic to Lj,s=Ljλ(s)L_{j,s}=L_{j}^{\lambda(s)}. Let Lj,A,s,σL_{j,A,s,\sigma} be such an exact Lagrangian isotopy, parametrized by σ[0,1]\sigma\in[0,1]. It is possible to find such isotopies such that the two-dimensional family Lj,A,s,σL_{j,A,s,\sigma} parametrized by (s,σ)×[0,1](s,\sigma)\in\mathbb{R}\times[0,1] is smooth. Furthermore, Lj,A,s,σL_{j,A,s,\sigma} is constant (i.e. independent of both ss and σ\sigma) for s0s\gg 0, where Lj,s=LjL_{j,s}=L_{j}. The reason is as follows. Since LjL_{j} is assumed to be invariant under the Liouville flow in the Liouville cobordism M0int(U0)M_{0}\setminus int(U_{0}), the completion of i(ψUALj)i(\psi_{U}^{A}L^{\prime}_{j}) is exact Lagrangian isotopic to either LjL_{j} or Lj1ϵL_{j}^{1-\epsilon}. Recall that the completion of i(ψUALj)i(\psi_{U}^{A}L^{\prime}_{j}) is precisely ψMALj\psi_{M}^{A}L_{j}. This is isotopic to LjL_{j} via the exact Lagrangian isotopy ψMρA(s)Lj\psi_{M}^{\rho_{A}(s)}L_{j}, where ρA:[1,A]\rho_{A}:\mathbb{R}\to[1,A] is the previously used function which is AA for s0s\ll 0 and 11 for s0s\geq 0. On the other hand, Lj1ϵL_{j}^{1-\epsilon} is isotopic to LjL_{j} via the exact Lagrangian isotopy Ljλ(s)L_{j}^{\lambda(s)}. Thus we may compose these two isotopies to obtain an isotopy from ψMALj\psi_{M}^{A}L_{j} to Lj1ϵL_{j}^{1-\epsilon}. For each ss, we reparametrize these isotopies by σ[0,1]\sigma\in[0,1] to obtain an isotopy from Lj,A,sL_{j,A,s} to Lj,s=Ljλ(s)L_{j,s}=L_{j}^{\lambda(s)}, namely,

(8.39) Lj,A,s,σ={ψMρA((12σ)s)Lj, if σ[0,12],Ljλ12σ(s), if σ[12,1],L_{j,A,s,\sigma}=\begin{cases}\psi_{M}^{\rho_{A}((1-2\sigma)s)}L_{j},&\text{ if }\sigma\in[0,\frac{1}{2}],\\ L_{j}^{\lambda_{1-2\sigma}(s)},&\text{ if }\sigma\in[\frac{1}{2},1],\end{cases}

where λσ\lambda_{\sigma} is a non-decreasing homotopy between the function λ:[1ϵ,1]\lambda:\mathbb{R}\to[1-\epsilon,1] and the constant function 11. This two-dimensional family (8.39) then satisfies all desired properties.

We can then define a parametrized moduli space

𝒫+A(xA,x)\mathcal{P}_{+}^{A}(x_{A},x)

of pairs (σ,wσ)(\sigma,w_{\sigma}), where σ[0,1]\sigma\in[0,1] and wσ:ZMw_{\sigma}:Z\to M is an inhomogeneos pseudoholomorphic map with respect to the Floer datum (HA,s,JA,(s,t))(H_{A,s},J_{A,(s,t)}) and the Lagrangian boundary conditions (L0,A,s,σ,L1,A,s,σ)(L_{0,A,s,\sigma},L_{1,A,s,\sigma}). This moduli space provides the desired cobordism between the moduli spaces 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) and 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x).

To obtain a cobordism between the compactified moduli spaces, we just need to compactify the moduli space 𝒫+A(xA,x)\mathcal{P}_{+}^{A}(x_{A},x) in an appropriate way. Such a compactification can be obtained in a similar way to those for 𝒫~A(xA,x)\tilde{\mathcal{P}}^{A}(x_{A},x) and 𝒫A(xA,x)\mathcal{P}^{A}(x_{A},x), as described below. For each fixed σ\sigma, we have a moduli space 𝒫σA(xA,x)\mathcal{P}_{\sigma}^{A}(x_{A},x) of maps wσw_{\sigma}, such that when σ=0\sigma=0, 𝒫0A(xA,x)=𝒫~A(xA,x)\mathcal{P}_{0}^{A}(x_{A},x)=\tilde{\mathcal{P}}^{A}(x_{A},x), and when σ=1\sigma=1, 𝒫1A(xA,x)=𝒫A(xA,x)\mathcal{P}_{1}^{A}(x_{A},x)=\mathcal{P}^{A}(x_{A},x). Each such moduli space 𝒫σA(xA,x)\mathcal{P}_{\sigma}^{A}(x_{A},x) is compactified in the same way to 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x) and 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x). Thus we may define the compactification 𝒫¯+A(xA,x)\bar{\mathcal{P}}_{+}^{A}(x_{A},x) to the the union of these:

(8.40) 𝒫¯+A(xA,x)=σ[0,1]𝒫¯σA(xA,x).\bar{\mathcal{P}}_{+}^{A}(x_{A},x)=\cup_{\sigma\in[0,1]}\bar{\mathcal{P}}_{\sigma}^{A}(x_{A},x).

This compactified moduli space then provides the desired cobordism between 𝒫~¯A(xA,x)\bar{\tilde{\mathcal{P}}}^{A}(x_{A},x) and 𝒫¯A(xA,x)\bar{\mathcal{P}}^{A}(x_{A},x).

Corollary 8.24.

Under the assumption of Lemma 8.22, the cochain maps ΠΓ\Pi_{\Gamma} and r1r^{1} are chain homotopic.

Proof.

Note that there is a chain homotopy equivalence

kA:CW(i(ψUAL0),i(ψUAL1);HA)CW(L01ϵ,L11ϵ;HA)k_{A}:CW^{*}_{-}(i(\psi_{U}^{A}L^{\prime}_{0}),i(\psi_{U}^{A}L^{\prime}_{1});H_{A})\to CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})

defined by counting rigid elements in a moduli space of inhomogeneous pseudoholomorphic maps u:ZMu:Z\to M satisfying the equation

su+JA,t(tuXHA(u))=0,\partial_{s}u+J_{A,t}(\partial_{t}u-X_{H_{A}}(u))=0,

and the Lagrangian boundary conditions

u(s,j)Lj,A,,σ=ρ(s),j=0,1,u(s,j)\in L_{j,A,-\infty,\sigma=\rho(s)},j=0,1,

where Lj,A,,σ=ρ(s)L_{j,A,-\infty,\sigma=\rho(s)} is obtained from the family Lj,A,s,σL_{j,A,s,\sigma} by first specializing s=s=-\infty, and then substituting σ\sigma by ρ(s)\rho(s). Lj,A,,σ=ρ(s)L_{j,A,-\infty,\sigma=\rho(s)} can be regarded as an exact Lagrangian isotopy parametrized by ss\in\mathbb{R} such that for s0s\ll 0, Lj,A,,σ=ρ(s)=Lj1ϵL_{j,A,-\infty,\sigma=\rho(s)}=L_{j}^{1-\epsilon}, and for s0s\gg 0, Lj,A,,σ=ρ(s)=i(ψUALj)L_{j,A,-\infty,\sigma=\rho(s)}=i(\psi_{U}^{A}L^{\prime}_{j}).

We claim that r~A1\tilde{r}^{1}_{A} and kAtA1k_{A}\circ t^{1}_{A} are chain homotopic. By counting rigid elements in the moduli space 𝒫¯+A(xA,x)\bar{\mathcal{P}}_{+}^{A}(x_{A},x), we define a map

TA:CW(L0,L1;HM)CW1(L01ϵ,L11ϵ;HA)T_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*-1}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})

of degree 1-1. By a standard gluing argument, this is a chain homotopy between r~A1\tilde{r}^{1}_{A} and kAtA1k_{A}\circ t^{1}_{A}.

Composing TAT_{A} with the map

τA:CW(L01ϵ,L11ϵ;HA)CW(Aϵ2,δ)(L0,L1;HU),\tau_{A}:CW^{*}_{-}(L_{0}^{1-\epsilon},L_{1}^{1-\epsilon};H_{A})\to CW^{*}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}),

we get a map

SA:CW(L0,L1;HM)CW(Aϵ2,δ)1(L0,L1;HU).S_{A}:CW^{*}(L_{0},L_{1};H_{M})\to CW^{*-1}_{(-A\epsilon^{2},\delta)}(L^{\prime}_{0},L^{\prime}_{1};H_{U}).

Following a similar homotopy commutativity argument as in Lemma 8.12, we may take the direct limit of the directed system of maps SAS_{A} to get

S=limA+SA.S=\lim\limits_{A\to+\infty}S_{A}.

All the above maps have unique continuous extensions to the completed wrapped Floer cochain spaces, when the energy of inhomogeneous pseudoholomorphic maps is taken into account in the counting definition.

Combining Lemma 8.22 and Lemma 8.23, we conclude that ΠΓ\Pi_{\Gamma} is chain homotopic to rA1r^{1}_{A} when the former map is restricted to given sub-complex generated by those generators whose images under ΠΓ\Pi_{\Gamma} fall within the action filtration window (Aϵ2,δ)(-A\epsilon^{2},\delta). Such a chain homotopic is given by the map SAS_{A}. Taking the direct limit over AA, we conclude that ΠΓ\Pi_{\Gamma} is chain homotopic to r1r^{1}, where the chain homotopy is given by the map SS.

Since ΠΓ\Pi_{\Gamma} is chain homotopic to ΘΓ1\Theta_{\Gamma}^{1}, Theorem 1.4 is therefore complete.

Remark 8.25.

A point worth noting is that the Viterbo restriction functor is better defined as a colimit of continuation functors with respect to linear Hamiltonians (or the cascade definition as in [AS10]), if one wants to visualize the picture of the Hamiltonian dynamics more directly. We took the current approach simply because of the quadratic Hamiltonians are more convenient for the purpose of constructing functors from Lagrangian correspondences, so that these functors can be compared in the same setup.

8.6. Further questions

It is therefore natural to ask whether the functors ΘΓ\Theta_{\Gamma} and rr as a whole are homotopic to each other, not just limited to their linear terms. While the expectation is yes, an efficient way of proving this is yet to be discovered. At least, there is a very naive case where such coincidence can be easily verified. For example, consider the case where L0L_{0} is a closed exact Lagrangian submanifold that is contained in U0U_{0}. Then the action of Θ\Theta_{\mathcal{L}} is the identity. This can be easily proved using the maximum principle, which implies that any pseudoholomorphic disk in M0M_{0} with boundary on LL and its Hamiltonian perturbations must be contained in U0U_{0}. In the same way, one sees that the Viterbo restriction functor is also the identity functor on such Lagrangian submanifolds, which implies that the functor ΘΓ\Theta_{\Gamma} agrees with the Viterbo restriction functor on such objects as well. To solve this problem in general, the main difficulty is to find a workable geometric construction of the functor ΘΓ\Theta_{\Gamma}, as the definition of the cochain map ΠΓ\Pi_{\Gamma} does not seem to have a straightforward generalization to an AA_{\infty}-functor. Finding a suitable model of the moduli spaces of quilted surfaces based on which the functor ΘΓ\Theta_{\Gamma} can be constructed directly is the key step of solving this problem.

There are of courses many other Lagrangian submanifolds which do not satisfy the geometric conditions we have just discussed. First, there are non-compact exact cylindrical Lagrangian submanifolds of MM that does not have very nice restriction to U0U_{0}. A typical example is the cotangent fiber of an annulus restricted to the disjoint union of three cotangent fibers of a deformed sub-annulus, as illustrated in Example 4.2 of [AS10]. Second, there are closed exact Lagrangian submanifolds of MM which are not entirely contained in U0U_{0}. In such cases, the usual construction of the Viterbo restriction functor does not yield an AA_{\infty}-functor in general. However, by analyzing the failure of it being an AA_{\infty}-functor, we expect that there is an extension of the Viterbo restriction functor to such Lagrangia submanifolds. Spectacularly, such an extension is related to deformation theory of the wrapped Fukaya category of UU, and we phrase it as the following conjecture.

Conjecture 8.26.

Suppose U0M0U_{0}\subset M_{0} is a Liouville sub-domain. Let the wrapped Fukaya category of MM and that of UU consist of a suitable countable collection of Lagrangian submanifolds. Then there is a canonical deformation of the completed wrapped Fukaya category 𝒲(U)\mathcal{W}(U) of UU, denoted by 𝒲(U;B)\mathcal{W}(U;B), where BB is a collection of bounding cochains for objects in 𝒲(U)\mathcal{W}(U), such that there is a natural AA_{\infty}-functor

rB:𝒲(M)𝒲(U;B),r_{B}:\mathcal{W}(M)\to\mathcal{W}(U;B),

which agrees with the Viterbo restriction functor on the full sub-category (M)\mathcal{B}(M).

The study of such an extension also brings up the question when the extended Viterbo restriction functor can be identified with the functor ΘΓ\Theta_{\Gamma} and when not. That would require more thorough understanding of the bounding cochains in both pictures. Related topics will be discussed in the upcoming work [Gao].

References

  • [Abo10] Mohammed Abouzaid. A geometric criterion for generating the Fukaya category. Publ. Math. Inst. Hautes Études Sci., (112):191–240, 2010.
  • [AJ10] Manabu Akaho and Dominic Joyce. Immersed Lagrangian Floer theory. J. Differential Geom., 86(3):381–500, 2010.
  • [AS10] Mohammed Abouzaid and Paul Seidel. An open string analogue of Viterbo functoriality. Geom. Topol., 14(2):627–718, 2010.
  • [BEH+03] Frédéric Bourgeois, Tobias Ekholm, Helmut Hofer, Kris Wysocki, and Eduard Zehnder. Compactness results in symplectic field theory. Geom. Topol., 7:799–888, December 2003.
  • [BO16] Frédéric Bourgeois and Alexandru Oancea. S1-equivariant symplectic homology and linearized contact homology. International Mathematics Research Notices, page rnw029, 2016.
  • [FO99] Kenji Fukaya and Kaoru Ono. Arnold conjecture and Gromov-Witten invariant. J. Topology, 38(5):933–1048, 1999.
  • [FOOO09a] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian intersection Floer theory: anomaly and obstruction. Part I, volume 46 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, RI; International Press, Somerville, MA, 2009.
  • [FOOO09b] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian intersection Floer theory: anomaly and obstruction. Part II, volume 46 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, RI; International Press, Somerville, MA, 2009.
  • [FOOO12] K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono. Technical details on Kuranishi structure and virtual fundamental chain. arXiv:1209.4410, pages 1–257, 2012.
  • [FOOO13] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian Floer theory over integers: spherically positive symplectic manifolds. Pure and Applied Mathematics Quarterly, 9(2):189–289, 2013.
  • [Fuk02] Kenji Fukaya. Floer homology and mirror symmetry. II. In Minimal surfaces, geometric analysis and symplectic geometry (Baltimore, MD, 1999), volume 34 of Adv. Stud. Pure Math., pages 31–127. Math. Soc. Japan, Tokyo, 2002.
  • [Fuk15] Kenji Fukaya. SO(3)SO(3)-Floer homology of 3-manifolds with boundary 1. arXiv:1506.01435, pages 1–80, 2015.
  • [Gan13] Sheel Ganatra. Symplectic cohomology and duality for the wrapped Fukaya category. PhD thesis, MIT, 2013.
  • [Gao] Yuan Gao. Wrapped Fukaya categories, deformation, and the generalized Viterbo functor. in preparation.
  • [Gao17] Yuan Gao. Wrapped Floer cohomology and Lagrangian correspondences. arXiv:1703.04032, March 2017.
  • [GPS17] Sheel Ganatra, John Pardon, and Vivek Shende. Covariantly functorial Floer theory on Liouville sectors. arXiv:1706.03152, June 2017.
  • [LL13] Yankı Lekili and Max Lipyanskiy. Geometric composition in quilted Floer theory. Adv. Math., 236:1–23, 2013.
  • [MS06] Martin Markl and Steve Shnider. Associahedra, cellular W-construction and products of A{A}_{\infty}-algebras. Transactions of the American Mathematical Society, 358(6):2353–2372, 2006.
  • [Sei08] Paul Seidel. Fukaya categories and Picard-Lefschetz theory. Zurich Lectures in Advanced Mathematics. European Mathematical Society (EMS), Zürich, 2008.
  • [SU04] Samson Saneblidze and Ronald Umble. Diagonals on the permutahedra, multiplihedra and associahedra. Homology Homotopy Appl., 6(1):363–411, 2004.
  • [Syl15] Zachary Sylvan. On partially wrapped Fukaya categories. PhD thesis, UC Berkeley, 2015.
  • [Vit99] Claude Viterbo. Functors and computations in Floer homology with applications. I. Geom. Funct. Anal., 9(5):985–1033, 1999.
  • [WW12] Katrin Wehrheim and Chris T. Woodward. Floer cohomology and geometric composition of Lagrangian correspondences. Adv. Math., 230(1):177–228, 2012.
  • [WW15a] Katrin Wehrheim and Chris T. Woodward. Orientations for pseudoholomorphic quilts. arXiv:1503.07803, pages 1–56, 2015.
  • [WW15b] Katrin Wehrheim and Chris T. Woodward. Pseudoholomorphic Quilts. arXiv:0905.1369, pages 1–34, 2015.