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

\titlecontents

section [0.72em] \thecontentslabel. \contentspage\titlecontentssubsection [0em]    \thecontentslabel. \contentspage\titlecontentssubsubsection [0em]     \thecontentslabel. \contentspage

Family Floer superpotential’s critical values are eigenvalues of quantum product by c1c_{1}

Hang Yuan
Abstract

Abstract: In the setting of the non-archimedean SYZ mirror construction in [Yua20], we prove the folklore conjecture that the critical values of the mirror superpotential are the eigenvalues of the quantum multiplication by the first Chern class. Our result relies on a weak unobstructed assumption, but it is usually ensured in practice by Solomon’s results [Sol20] on anti-symmetric Lagrangians. Lastly, we note that some explicit examples are presented in the recent work [Yua22a].

1 Introduction

The mirror symmetry phenomenon does not merely focus on the Calabi-Yau setting. Assume that a compact symplectic manifold (X,ω)(X,\omega) is Fano (or more generally that KX-K_{X} is nef). The mirror of XX is expected to be a Landau-Ginzburg model (X,W)(X^{\vee},W^{\vee}) consisting of a space XX^{\vee} equipped with a global function WW^{\vee} called the superpotential. In the context of the homological mirror symmetry [Kon95], a celebrated folklore conjecture states that: (known to Kontsevich and Seidel, compare also [She16, §2.9])

Conjecture 1.1.

The critical values of the mirror Landau-Ginzburg (LG) superpotential WW^{\vee} are the eigenvalues of the quantum multiplication by the first Chern class of XX.

The conjecture is first proved by Auroux [Aur07] in the Fano toric case, concerning the Landau-Ginzburg mirror associated to a toric Lagrangian fibration in XX [CO06, FOOO10a].

Theorem 1.2.

Conjecture 1.1 holds for the family Floer mirror Landau-Ginzburg superpotential.

See §1.1 for the precise statement. In brief, we aim to prove the conjecture with nontrivial Maslov-0 quantum correction, and actually we don’t need to be limited to the family Floer scope. It looks like a “small” step to include Maslov-0 disks, but it is indeed difficult and requires some new ideas (§1.2). We also find new examples recently in [Yua22a] to justify and amplify the value of this paper.

The symplectic-geometric ideas we use are not original, but it is not our main point at all. Instead, a major point in this paper is to demonstrate that the classical ideas fit perfectly into the framework of the family Floer mirror construction in [Yua20]. For example, a key new ingredient for our proof is a preliminary version of “Fukaya category over affinoid coefficients” (i.e. the self Floer cohomology for 𝒰𝒟\mathscr{UD} in §3). Roughly, the coefficient we use is not \mathbb{C}, not the Novikov field Λ\Lambda, but an affinoid algebra over Λ\Lambda. The merit is that we can ignore all the affects of Maslov-0 disk obstruction up to affinoid algebra isomorphism on the coefficients using the ideas in the author’s thesis [Yua20]. We expect the affinoid coefficients should be important to develop Abouzaid’s family Floer functor [Abo17, Abo21] with Maslov-0 corrections. A work in progress is [Yua]. For the above speculative version of Fukaya category, one might study many topics like splitting generations [AFO+] and Bridgeland stability conditions (cf. [Smi17] [Kon]) in some similar ways in future studies. We also expect that it is related to the non-archimedean geometry, like the coherent analytic sheaves, on the family Floer mirror space, potentially being the analytification of an algebraic variety over the Novikov field. But unfortunately, the precondition of all the story is to allow the Maslov-0 disks in the course. Otherwise, we cannot even realize a simple algebraic variety like Y={x0x1=1+y}Y=\{x_{0}x_{1}=1+y\} as a family Floer mirror space and thus miss the new examples of Conjecture 1.1 in [Yua22a].

Refer to caption
Figure 1: A piece of contribution to the LG superpotential with / without the Maslov-0 quantum correction. When we perturb JJ, all the Maslov-0 disks (green) in the Feynman-diagrams of the minimal model AA_{\infty} algebras are continually created and annihilated for the wall-crossing. This is more like a sort of “quantum fluctuation”. This is why the LG superpotential is only well-defined up to affinoid algebra isomorphism.

(1) What is the difficulty of allowing Maslov-0 corrections?

First and foremost, the family Floer mirror superpotential is only well-defined up to affinoid algebra isomorphisms, or equivalently, up to family Floer analytic transition maps [Yua20]. This is because the Maslov-0 quantum correction give rise to certain mirror non-archimedean analytic topology. Note that the idea of uniting Maurer-Cartan sets, adopted by Fukaya and Tu [Fuk09, Fuk11, Tu14], cannot achieve the family Floer mirror non-archimedean analytic structure. We must go beyond their Maurer-Cartan picture; otherwise, although Fukaya and Tu did tell the correct local analytic charts, the local-to-global gluing was only set-theoretic. In fact, we never think of bounding cochains but work with formal power series in the corresponding affinoid algebra directly [Yua20]. One of many crucial flaws in the previous work concerns the lack of studies of the divisor axiom for AA_{\infty} structures111The earliest literature we find for open-string divisor axiom is due to Seidel [Sei06]. See also [Aur07, Fuk10].. The ud-homotopy theory developed in [Yua20] is indispensable for the mirror analytic gluing. Remarkably, the ud-homotopy theory plays the key role for the folklore conjecture once again (cf. §1.2).

Putting another way, we must work with the minimal model AA_{\infty} algebras, while paying attention to the wall-crossing phenomenon simultaneously (Figure 1). Without minimal model AA_{\infty} algebras, the superpotential would be only Ω0(L)\Omega^{0}(L)-valued and was not well-defined. This issue was not so fatal if the Maslov indices >0>0, but it is indeed fatal if we allow Maslov-0 correction. As another evidence, a minimal model AA_{\infty} algebra is intuitively like counting holomorphic pearly-trees (cf. [FOOO09]) which are also used in Sheridan’s proof of HMS of Calabi-Yau projective hypersurfaces [She15, §6.2]. The conventional ideas of the folklore conjecture cannot work or need to be improved for holomorphic pearly-trees or minimal model AA_{\infty} algebras, if nontrivial Maslov-0 disks exist (Figure 1). A daunting trouble is to handle the minimal model AA_{\infty} algebra’s Hochschild cohomology (see §1.2).

(2) Why does the family Floer method give a convincing approach to construct the mirror?

The answer in short is simply that we can prove the Conjecture 1.1 supported by examples below.

Gross and Siebert proposed a method of mirror construction in the algebro-geometric framework [GS11] based on the SYZ proposal [SYZ96]. Gross-Hacking-Keel [GHK15] also give a synthetic algebro-geometric construction of the mirror family for local Calabi-Yau surfaces. These approaches turn out to be successful in many aspects, such as [HK20, GHK13, GHKK18]. But, in our biased opinion, the folklore Conjecture 1.1 inevitably relies on Floer-theoretic or symplectic-geometric ideas in the end. The family Floer mirror construction [Yua20] can naturally include the Landau-Ginzburg superpotential. Such a Floer-theoretic mirror construction succeeded in rigorizing the T-duality of smooth Lagrangian torus fibration with full quantum correction considered, and leads to a mathematical precise statement of SYZ conjecture with singularities in [Yua22a, Yua22b, Yua23].

1.1 Main result

Let (X,ω)(X,\omega) be a symplectic manifold which is closed or convex at infinity. Let dimX=2n\dim_{\mathbb{R}}X=2n. Let LL be an oriented, spin, and compact Lagrangian submanifold satisfying the semipositive condition that prohibits any nontrivial negative Maslov index holomorphic disk. In practice, a graded or special Lagrangian submanifold is always semipositive by [Aur07, Lemma 3.1]. Note that LL is not necessarily a torus. Given an ω\omega-tame almost complex structure JJ, we can associate to LL a superpotential function

(1) WL:=WLJ=βπ2(X,L),μ(β)=2TE(β)Yβ𝔪0,βW_{L}:=W_{L}^{J}=\sum_{\beta\in\pi_{2}(X,L),\mu(\beta)=2}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{0,\beta}

where we use the minimal model AA_{\infty} algebra

𝔪=(𝔪k,β)k,βπ2(X,L)\mathfrak{m}=(\mathfrak{m}_{k,\beta})_{k\in\mathbb{N},\beta\in\pi_{2}(X,L)}

of the cochain-level AA_{\infty} algebra associated to LL. Here we work over the Novikov field Λ=((T))\Lambda=\mathbb{C}((T^{\mathbb{R}})), YY is a formal symbol, so WLΛ[[π1(L)]]W_{L}\in\Lambda[[\pi_{1}(L)]]. The reverse isoperimetric inequality [GS14] ensures that WLW_{L} converges on an analytic open neighborhood of H1(L;UΛ)H^{1}(L;U_{\Lambda}) in H1(L;Λ)H^{1}(L;\Lambda^{*}).

Although the summation in (1) runs only over Maslov-2 topological disks β\beta, it actually can involve all the possible Maslov-0 obstruction in the same time. This is simply the nature of minimal model AA_{\infty} algebras, and the idea is illustrated in Figure 1. In general, allowing the Maslov-0 disks, the superpotential function WLJW_{L}^{J} is not well-defined and rely on the various choices like JJ. But, as in [Yua20] (see Lemma 2.6), we can show that the WLW_{L} is well-defined up to affinoid algebra isomorphism, and we can also show that the critical points and critical values are preserved.

Assumption 1.3.

The formal power series222It is related to the weak Maurer-Cartan equations in the literature, but we forget the latter and only think QLJQ_{L}^{J}. QLJ:=μ(β)=0TE(β)Yβ𝔪0,βQ_{L}^{J}:=\sum_{\mu(\beta)=0}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{0,\beta} vanishes identically.

Notably, the recent results in [Yua24] provide strong support for considering the above assumption reasonable, and it is likely that the assumption holds in many situations. It does not say that the counts of Maslov-0 disks vanish. These disks still contribute to both the minimal model AA_{\infty} algebras and the wall-crossing phenomenon, but their contributions ultimately cancel out in the coefficient affinoid algebra. Moreover, it is not a restrictive assumption. Thanks to the work of Solomon [Sol20], a very useful sufficient condition of the above Assumption 1.3, which we often adopt in practice, is given as follows:

Assumption 1.4.

There is an anti-symplectic involution φ\varphi that preserves LL.

Indeed, such a φ\varphi gives a pairing on π2(X,L)\pi_{2}(X,L) via βφβ\beta\xleftrightarrow{}-\varphi_{*}\beta. Then, the Maslov-0 obstruction terms 𝔪0,β\mathfrak{m}_{0,\beta} in QLJQ_{L}^{J} are then canceled pairwise. In practice, such an involution φ\varphi is often not hard to find (see also [CBMS10]). Note that the vanishing of QLJQ_{L}^{J} does not rely on JJ due to [Yua20]. Beware that we have not discuss the family Floer setting or the SYZ setting at this moment. Note also that we do not require LL is a torus here. Now, we state the main result.

Theorem A (Theorem 5.5).

Under Assumption 1.3, if the cohomology ring H(L)H^{*}(L) is generated by H1(L;)H^{1}(L;\mathbb{Z}), then any critical value of WLW_{L} is an eigenvalue of the quantum product by c1c_{1}. Moreover, the same result still holds if we use different choices.

The main new progress and difficulty is that we allow Maslov-0 holomorphic disks. It is conceivable that if the above LL lives in a family of Lagrangian submanifolds {Lq}\{L_{q}\}, then one may extend the domain of WLW_{L} to a larger open analytic space. Up to the Fukaya’s trick, this amounts to study how WLW_{L} depends on the choice of JJ. Recall that WLW_{L} is only invariant up to affinoid algebra isomorphisms. This is because of a sort of ‘quantum fluctuation’, when we think of the minimal model AA_{\infty} algebra together with the wall-crossing phenomenon in the meantime (see Figure 1).

1.1.1 Family Floer mirror construction setting

As a byproduct of the above Theorem A, we can verify the family Floer mirror superpotential in [Yua20] respects Conjecture 1.1. We first briefly review it. Suppose (X,ω)(X,\omega) is a symplectic manifold which is closed or convex at infinity. Suppose there is a smooth Lagrangian torus fibration π0:X0B0\pi_{0}:X_{0}\to B_{0} in an open subset X0X_{0} in XX over a half-dimensional integral affine manifold B0B_{0}. Suppose that π0\pi_{0} is semipositive in the sense that any holomorphic disk bounding a fiber Lq:=π01(q)L_{q}:=\pi_{0}^{-1}(q) has a non-negative Maslov index. By [Aur07], a sufficient condition is to assume π0\pi_{0}-fibers are graded or special Lagrangians. For simplicity, we also require that the π0\pi_{0}-fibers satisfy Assumption 1.3 or 1.4. Then, we state the result of SYZ duality with quantum correction considered:

Theorem 1.5 ([Yua20]).

We can naturally associate to the pair (X,π0)(X,\pi_{0}) a triple 𝕏:=(X0,W0,π0)\mathbb{X}^{\vee}:=(X_{0}^{\vee},W_{0}^{\vee},\pi_{0}^{\vee}) consisting of a non-archimedean analytic space X0X_{0}^{\vee} over Λ\Lambda, a global analytic function W0W_{0}^{\vee}, and an affinoid torus fibration π0:X0B0\pi_{0}^{\vee}:X_{0}^{\vee}\to B_{0} such that the following properties hold:

  1. i)

    The analytic structure of X0X_{0}^{\vee} is unique up to isomorphism.

  2. ii)

    The integral affine structure on B0B_{0} from π0\pi_{0}^{\vee} coincides with the one from the fibration π0\pi_{0}.

  3. iii)

    The set of closed points in X0X_{0}^{\vee} coincides with qB0H1(Lq;UΛ)\bigcup_{q\in B_{0}}H^{1}(L_{q};U_{\Lambda}).

Notice that this mirror construction uses holomorphic disks in XX rather than just in X0X_{0}, and we should place π0\pi_{0} inside XX to understand the statement. For simplicity, we often omit the subscripts 0’s, so we often write (X,W,π)(X^{\vee},W^{\vee},\pi^{\vee}), if there is no confusion. The ud-homotopy theory and the category 𝒰𝒟\mathscr{UD} in [Yua20] is indispensable for the non-archimedean analytic topology in Theorem 1.5. The usual homotopy theory can only achieve a gluing in the set-theoretic level. Very intuitively, the family Floer theory is ‘algebrized’ by the category 𝒰𝒟\mathscr{UD}, and the choice issues are all controlled by the ud-homotopy relations in 𝒰𝒟\mathscr{UD}.

Theorem B.

Conjecture 1.1 holds for the mirror Landau-Ginzburg model in Theorem 1.5. Namely, any critical value of WW^{\vee} must be an eigenvalue of the quantum product by the first Chern class c1c_{1} of XX.

Roughly, by allowing the wall-crossing invariance, we can enlarge the domain of the superpotential and so potentially cover more eigenvalues. But, we cannot conversely conclude that all the eigenvalues of c1c_{1} are realized by a critical point of the superpotential. Note that the Λ\Lambda-coefficients in WW^{\vee} include the data of symplectic areas and can tell which chambers the critical points are situated in. In particular, the locations of critical points depend on the symplectic form ω\omega.

1.2 A new operator for Hochschild cohomology and ud-homotopy theory

The obstructions in Lagrangian Floer theory cause issues for both analysis and algebra. The analysis part should be of most importance and difficulty, concerning the complicated transversality issues of moduli spaces of pseudo-holomorphic curves. This is studied and developed by [FOOO10b, FOOO10c, FOOO20] and so on. In comparison, the algebra part should be minor. But, in our biased opinion, the intricacies and subtleties of the algebra part are somewhat underestimated.

The open string Floer theoretic invariants often depend on choices. Thus, we often must talk about certain equivalence relations, depending on which the information we miss can vary. For example, in the literature like [FOOO10b], we usually talk about AA_{\infty} algebras up to homotopy equivalence. It is fine in most cases, but it may be too coarse in case we need more information (e.g. non-archimedean analytic structure). The homotopy equivalence relation cannot distinguish between a minimal model AA_{\infty} algebra and its original. This is why the usual Maurer-Cartan picture [Fuk09, Fuk11, Tu14] cannot be enough for a local-to-global gluing for the family Floer mirror analytic structure, as explained before.

In our specific situation here for the folklore Conjecture 1.1, we must carefully distinguish between a minimal model AA_{\infty} algebra, denoted by (H(L),𝔪)(H^{*}(L),\mathfrak{m}), and its original, denoted by (Ω(L),𝔪ˇ)(\Omega^{*}(L),\check{\mathfrak{m}}). This is because the classic ideas for the folklore conjecture can only achieve connections between the quantum cohomology and the cochain-level Hochschild cohomology HH(Ω(L),𝔪ˇ)HH(\Omega^{*}(L),\check{\mathfrak{m}}). Meanwhile, the family Floer Landau-Ginzburg superpotential comes from the minimal model AA_{\infty} algebra (H(L),𝔪)(H^{*}(L),\mathfrak{m}).

There are some new challenges. For instance, the Hochschild cohomology is just not functorial, and in pure homological algebra, it is basically impossible to make any meaningful connections between the Hochschild cohomologies of a minimal model AA_{\infty} algebra and its original. Fortunately, in our specific geometric situation (with Assumption 1.3 or 1.4), we magically make an effective connection by the following operation (see (53))

Θ:φ(𝔦1{φ})𝔦=(1)𝔦1(𝔦,,𝔦,φ(𝔦,,𝔦),𝔦,,𝔦)\Theta:\varphi\mapsto(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}=\sum(-1)^{*}\ \mathfrak{i}^{-1}\big{(}\mathfrak{i},\dots,\mathfrak{i},\varphi(\mathfrak{i},\dots,\mathfrak{i}),\mathfrak{i},\dots,\mathfrak{i}\big{)}

where \diamond is the composition (see [She15]), the {}\{\} is the celebrated brace operation (see [Get93]), and the 𝔦1\mathfrak{i}^{-1} is the ud-homotopy inverse of the natural AA_{\infty} homotopy equivalence 𝔦:𝔪𝔪ˇ\mathfrak{i}:\mathfrak{m}\to\check{\mathfrak{m}} in the course of homological perturbation. A difficult part of the paper is to find the above operation Θ\Theta and establish the corresponding computations. As far as we know, such Θ\Theta does not appear in any literature and seems to be exclusive for our ud-homotopy theory in [Yua20].

Now, the ud-homotopy theory shows its necessity again. Moreover, it is interesting to note that the family Floer mirror analytic gluing in [Yua20] does not exploit the full information of ud-homotopy relations, while the techniques in the present paper (e.g. the self Floer cohomology for 𝒰𝒟\mathscr{UD} in §3) use more information of ud-homotopy relations.

1.3 Outline and major difficulties

If the Maslov indices 2\geq 2, the approach to Conjecture 1.1 is well-known to experts for a long time: see [Aur07, §6] or [RS17, §12.1] in toric cases and [FOOO19, §23] in general; see also [She16, Definition 2.3]. Roughly speaking, a key idea concerns a unital ring homomorphism

(2) QH(X)HH(CF(L,L))HF(L,L)QH^{*}(X)\to HH^{*}(CF(L,L))\to HF(L,L)

from the quantum cohomology to the self Floer cohomology through the Hochschild cohomology. But, when nontrivial Maslov-0 disks are allowed, we cannot obtain such a unital ring homomorphism as usual. It is even necessary to find a suitable way to generalize the definitions of HF(L,L)HF(L,L) and HH(CF(L,L))HH^{*}(CF(L,L)) in (2). The major difficulties are outlined as follows.

i

We want the self Floer cohomology HF(L,L)HF(L,L) to be defined in a ‘wall-crossing invariant’ way.

This is especially crucial for the global result, and we note that LL is not necessarily a torus here.

First, thanks to the Fukaya-Oh-Ohta-Ono’s foundation works, we can very generally define the self Floer cohomology HF((L,b),(L,b))HF((L,b),(L,b)) by choosing any odd-degree weak bounding cochain bb. However, it may be too general for our purpose. We want to work with a more restrictive situation, highlighting the semipositive condition and the wall-crossing phenomenon. By utilizing the related ideas about the category 𝒰𝒟\mathscr{UD}, we define a modified self Floer cohomology HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) for any cohomology-level AA_{\infty} algebra (H(L),𝔪)(H^{*}(L),\mathfrak{m}) in 𝒰𝒟\mathscr{UD}3). It is a unital Λ\Lambda-algebra, and it does not refer to a specific weak bounding cochain. Instead, whenever 𝐲H1(L;UΛ)\mathbf{y}\in H^{1}(L;U_{\Lambda}) admits a lift of weak bounding cochain bb in H1(L;Λ0)H^{1}(L;\Lambda_{0}) via exp:Λ0UΛ\exp:\Lambda_{0}\to U_{\Lambda}, we may define HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) in almost the same way as the above HF((L,b),(L,b))HF((L,b),(L,b)); moreover, it may be viewed as the restriction of HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) at the point 𝐲\mathbf{y}, accompanied by a natural unital ring homomorphism 𝐲:HF(L,𝔪)HF(L,𝔪,𝐲)\mathcal{E}_{\mathbf{y}}:\operatorname{HF}(L,\mathfrak{m})\to\operatorname{HF}(L,\mathfrak{m},\mathbf{y}).

Our modified Floer cohomology are wall-crossing invariant in the following sense. Recall that an analytic transition map ϕ\phi is determined by an AA_{\infty} homotopy equivalence :𝔪𝔪\mathfrak{C}:\mathfrak{m}\to\mathfrak{m}^{\prime} in 𝒰𝒟\mathscr{UD}. Then, the ϕ\phi and \mathfrak{C} can induce an isomorphism HF(L,𝔪)HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m})\cong\operatorname{HF}(L,\mathfrak{m}^{\prime}) as well. Further, we have the following: When the de Rham cohomology ring H(L)H^{*}(L) is generated by H1(L)H^{1}(L) (e.g. LL is topologically a torus), the HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) is non-vanishing if and only if 𝐲\mathbf{y} is a critical point of WW, where WW is the local superpotential defined by 𝔪\mathfrak{m} as before. This result also has the wall-crossing invariance. Set 𝐲=ϕ(𝐲)\mathbf{y}^{\prime}=\phi(\mathbf{y}) and W=ϕWW^{\prime}=\phi^{*}W; not only HF(L,𝔪,𝐲)HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y})\cong\operatorname{HF}(L,\mathfrak{m}^{\prime},\mathbf{y}^{\prime}) but also we know that if 𝐲\mathbf{y} is a critical point of WW, then 𝐲\mathbf{y}^{\prime} is a one of WW^{\prime} and vice versa. Finally, we remark that the non-archimedean analysis plays an important role for this result: Let >0\hbar>0 be a lower bound of the energies of nontrivial holomorphic disks, then we should first show the result modulo TT^{\hbar} and inductively refine it to TkT^{k\hbar} for k1k\geq 1. The limit for the adic topology in the Novikov field Λ\Lambda as kk\to\infty works in the end.

ii

The definition of HH(CF(L,L))HH^{*}(CF(L,L)) in the middle of (2) becomes problematic and ambiguous.

This issue is actually more difficult than (i). The moduli spaces of holomorphic disks first only give the chain-level AA_{\infty} algebra (Ω(L),𝔪ˇ)(\Omega^{*}(L),\check{\mathfrak{m}}). In contrast, the self Floer cohomology and the superpotential all resort to the cohomology-level (H(L),𝔪)(H^{*}(L),\mathfrak{m}), i.e. the minimal model AA_{\infty} algebra. Previously, when the Maslov indices were 2\geq 2, the homological perturbation largely degenerates so that the difference between (Ω(L),𝔪ˇ)(\Omega^{*}(L),\check{\mathfrak{m}}) and (H(L),𝔪)(H^{*}(L),\mathfrak{m}) is mild enough to achieve (2). Unfortunately, when the Maslov-0 disks are involved, such difference cannot be ignored, and the subsequent discrepancy between the two Hochschild cohomology rings HH(Ω(L),𝔪ˇ)HH^{*}(\Omega^{*}(L),\check{\mathfrak{m}}) and HH(H(L),𝔪)HH^{*}(H^{*}(L),\mathfrak{m}) becomes highly nontrivial. For instance, there is basically no chance to make a ring homomorphism between the two of them, as the Hochschild cohomology is neither covariant nor contravariant. This obstacle leads to a peculiar mismatch between the algebra and geometry. The moduli spaces with interior markings can produce a ring homomorphism from the quantum cohomology QH(X)QH^{*}(X) only to the chain-level Hochschild cohomology HH(Ω(L),𝔪ˇ)HH^{*}(\Omega^{*}(L),\check{\mathfrak{m}}) but not to the cohomology-level one HH(H(L),𝔪)HH^{*}(H^{*}(L),\mathfrak{m}).

In a word, the solution can be realized by the relevant technologies about the category 𝒰𝒟\mathscr{UD} introduced both in this paper and in [Yua20]. Not only the homological algebra but also the non-archimedean analysis and the geometric semipositive condition in 𝒰𝒟\mathscr{UD} are used in a significant way. Briefly, we will carefully design a map Θ\Theta between the underlying two Hochschild cochain complexes such that: although it does not precisely induce a ring homomorphism, the gap to become so is controlled by the ud-homotopy relations in 𝒰𝒟\mathscr{UD}. Next, applying a natural projection map \mathbb{P}, this gap is then controlled by the weak Maurer-Cartan formal power series and eliminated under Assumption 1.3. We can ultimately show that the composition map Θ\mathbb{P}\circ\Theta can induce an honest unital ring homomorphism Φ\Phi5).

(cochain-level)\textstyle{(\text{cochain-level})}QH(X)\textstyle{QH^{*}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HH(Ω(L),𝔪ˇ)\textstyle{HH^{*}(\Omega^{*}(L),\check{\mathfrak{m}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ:=[Θ]\scriptstyle{\Phi:=[\mathbb{P}\circ\Theta]}(cohomology-level)\textstyle{(\text{cohomology-level})}HH(H(L),𝔪)\textstyle{HH^{*}(H^{*}(L),\mathfrak{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HF(L,𝔪)\textstyle{\operatorname{HF}(L,\mathfrak{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐲\scriptstyle{\mathcal{E}_{\mathbf{y}}}HF(L,𝔪,𝐲)\textstyle{\operatorname{HF}(L,\mathfrak{m},\mathbf{y})}

Another technical point is that we need the so-called reduced Hochschild cohomologies (c.f. [FOOO19, §25]). Roughly speaking, it is designed to keep track of units (§4). Finally, the composition gives a unital ring homomorphism 𝕆𝐲:QH(X)HF(L,𝔪,𝐲)\mathbb{CO}_{\mathbf{y}}:QH^{*}(X)\to\operatorname{HF}(L,\mathfrak{m},\mathbf{y}), despite of all the troubles caused by Maslov-0 disks (§5).

2 Preliminaries and reviews

The Floer theory usually works over the Novikov field Λ=((T))\Lambda=\mathbb{C}((T^{\mathbb{R}})). It has a valuation map 𝗏:Λ{}\operatorname{\mathsf{v}}:\Lambda\to\mathbb{R}\cup\{\infty\}, sending a nonzero series aiTλi\sum a_{i}T^{\lambda_{i}} to the smallest λi\lambda_{i} with ai0a_{i}\neq 0 and sending the zero series to \infty. The Novikov ring is the valuation ring Λ0:=𝗏1[0,]\Lambda_{0}:=\operatorname{\mathsf{v}}^{-1}[0,\infty] whose maximal ideal is Λ+:=𝗏1(0,]\Lambda_{+}:=\operatorname{\mathsf{v}}^{-1}(0,\infty]. The multiplicative group of units is UΛ:=𝗏1(0)U_{\Lambda}:=\operatorname{\mathsf{v}}^{-1}(0). Note that UΛ=Λ+U_{\Lambda}=\mathbb{C}^{*}\oplus\Lambda_{+} and Λ0=Λ+\Lambda_{0}=\mathbb{C}\oplus\Lambda_{+}. The standard isomorphism /2πi\mathbb{C}^{*}\cong\mathbb{C}/2\pi i\mathbb{Z} naturally extends to UΛΛ0/2πiU_{\Lambda}\cong\Lambda_{0}/2\pi i\mathbb{Z}. In particular, for any yUΛy\in U_{\Lambda}, there exists some xΛ0x\in\Lambda_{0} with y=exp(x)y=\exp(x).

Let Λ=Λ{0}\Lambda^{*}=\Lambda\setminus\{0\}, and we consider the tropicalization map

(3) 𝔱𝔯𝔬𝔭𝗏×n:(Λ)nn\operatorname{\mathfrak{trop}}\equiv\operatorname{\mathsf{v}}^{\times n}:(\Lambda^{*})^{n}\to\mathbb{R}^{n}

The total space (Λ)n(\Lambda^{*})^{n} should be understood as not just an algebraic variety over Λ\Lambda but also a non-archimedean analytic space. Roughly, a non-archimedean analytic space is a locally ringed space whose structure sheaf is locally given by affinoid algebras. Affinoid algebras are quotients of the Tate algebras. The spectrum of an affinoid algebra is called an affinoid space or affinoid domain. Let Δ\Delta be a rational polyhedron in n\mathbb{R}^{n}, then the preimage 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta) is an affinoid (sub)domain in (Λ)n(\Lambda^{*})^{n}. We refer to the appendix in [Yua20] or the standard textbook [Bos14, Ber12] for further readings.

Finally, we note a useful observation in [Yua20] or [Yua24] as follows:

Lemma 2.1.

fΛ[[Y1±,,Yn±]]f\in\Lambda[[Y_{1}^{\pm},\dots,Y_{n}^{\pm}]] vanishes identically if and only if f|UΛn0f|_{U_{\Lambda}^{n}}\equiv 0.

Proof.

Let f=νmcνYνf=\sum_{\nu\in\mathbb{Z}^{m}}c_{\nu}Y^{\nu}. Note that val(cν)\mathrm{val}(c_{\nu})\to\infty i.e. |cν|0|c_{\nu}|\to 0. Arguing by contraction, suppose the sequence |cν||c_{\nu}| was nonzero, say, it had a maximal value |cν0|=1|c_{\nu_{0}}|=1 for some ν0n\nu_{0}\in\mathbb{Z}^{n}. May further assume cν0=1c_{\nu_{0}}=1. Then, |cν|1|c_{\nu}|\leqslant 1 for all ν\nu, so fΛ0[[Y±]]f\in\Lambda_{0}[[Y^{\pm}]]. Modulo the ideal of elements with norm <1<1, we get a power series f¯=νc¯νYν\bar{f}=\sum_{\nu}\bar{c}_{\nu}Y^{\nu} over the residue field \mathbb{C}. As cν0c_{\nu}\to 0, we have |cν|<1|c_{\nu}|<1 and c¯ν=0\bar{c}_{\nu}=0 for ν1\nu\gg 1. Hence, this f¯\bar{f} is just a Laurent polynomial over \mathbb{C} with c¯ν0=1\bar{c}_{\nu_{0}}=1. Meanwhile, the condition also tells that f¯(𝐲¯)\bar{f}(\mathbf{\bar{y}}) vanishes for all 𝐲¯()n\mathbf{\bar{y}}\in(\mathbb{C}^{*})^{n}; thus, f¯\bar{f} must be identically zero. This is a contradiction. ∎

2.1 Gapped AA_{\infty} algebras and the category 𝒰𝒟\mathscr{UD}

We revisit the notations and terminologies from [Yua20, §2], along with introducing some new concepts where necessary. While there are slight differences in the formulations, similar notions have been previously introduced in [FOOO10b] and [Fuk10].

2.1.1 Gappedness

A label group is a triple (𝔊,E,μ)(\mathfrak{G},E,\mu) consisting of an abelian group 𝔊\mathfrak{G} and two group homomorphisms E:𝔊E:\mathfrak{G}\to\mathbb{R} and μ:𝔊2\mu:\mathfrak{G}\to 2\mathbb{Z}. For instance, let XX and LL denote a symplectic manifold and a Lagrangian submanifold, then we may take 𝔊=π2(X,L)\mathfrak{G}=\pi_{2}(X,L) (or more precisely, 𝔊\mathfrak{G} is the image of Hurewicz map π2(X,L)H2(X,L)\pi_{2}(X,L)\to H_{2}(X,L)); the EE and μ\mu are the energy and Maslov index.

Assume C,CC,C^{\prime} are graded \mathbb{R}-vector spaces. Given kk\in\mathbb{N} and β𝔊\beta\in\mathfrak{G}, we use the notation 𝐂𝐂k,β(C,C)\operatorname{\mathbf{CC}}_{k,\beta}(C,C^{\prime}) to denote a copy of Hom(Ck,C)\operatorname{Hom}(C^{\otimes k},C^{\prime}) where β\beta is just an extra index. Then, consider the space

(4) 𝐂𝐂(C,C)kβ𝔊𝐂𝐂k,β(C,C)\operatorname{\mathbf{CC}}(C,C^{\prime})\subseteq\prod_{k\in\mathbb{N}}\prod_{\beta\in\mathfrak{G}}\operatorname{\mathbf{CC}}_{k,\beta}(C,C^{\prime})

consisting of the operator systems 𝔱=(𝔱k,β)\mathfrak{t}=(\mathfrak{t}_{k,\beta}) satisfying the following gappedness conditions:

  • (a)

    𝔱0,0=0\mathfrak{t}_{0,0}=0;

  • (b)

    if E(β)<0E(\beta)<0 or E(β)=0E(\beta)=0, β0\beta\neq 0, then 𝔱β:=(𝔱k,β)k\mathfrak{t}_{\beta}:=(\mathfrak{t}_{k,\beta})_{k\in\mathbb{N}} vanishes identically;

  • (c)

    for any E0>0E_{0}>0, there are only finitely many β\beta such that 𝔱β0\mathfrak{t}_{\beta}\neq 0 and E(β)E0E(\beta)\leq E_{0}.

Here a collection of multilinear maps is called an operator system 𝔱\mathfrak{t}. If it is further contained in 𝐂𝐂(C,C)\operatorname{\mathbf{CC}}(C,C^{\prime}), namely, it satisfies the above three conditions, the 𝔱\mathfrak{t} is called (𝔊\mathfrak{G}-)gapped. Unless we further specify, everything is gapped from now on. We often abbreviate it to 𝐂𝐂\operatorname{\mathbf{CC}}.

2.1.2 Signs and degrees

Define the shifted degree degx:=degx1\deg^{\prime}x:=\deg x-1 for xCx\in C. It is convenient to set

(5) x#=id#(x)=id#(x)=(1)degx1x=(1)degxxx^{\#}=\mathrm{id}^{\#}(x)=\mathrm{id}_{\#}(x)=(-1)^{\deg x-1}x=(-1)^{\deg^{\prime}x}x

Given ss\in\mathbb{N} and a multi-linear map ϕ\phi, we put

ϕ#s=ϕid#s\phi^{\#s}=\phi\circ\mathrm{id}^{\#s}

where id#s\mathrm{id}^{\#s} denotes the ss-iteration id#id#\mathrm{id}^{\#}\circ\cdots\circ\mathrm{id}^{\#} of id#\mathrm{id}^{\#}. If s=1s=1, we set ϕ#=ϕ#1\phi^{\#}=\phi^{\#1}. Let degϕ\deg\phi be the usual degree as a homogeneously-graded kk-multilinear operator among graded vector spaces. Then, the shifted degree is degϕ=degϕ+k1\deg^{\prime}\phi=\deg\phi+k-1. Note that ϕ#=(1)degϕid#ϕ\phi^{\#}=(-1)^{\deg^{\prime}\phi}\ \mathrm{id}^{\#}\circ\phi.

We define the composition (see [She15, Definition 2.39]) and the Gerstenhaber product (see [Ger63])

(6) (𝔤𝔣)k,β=1k1++k=kβ0+β1++β=β𝔤,β0(𝔣k1,β1𝔣k,β)\displaystyle(\mathfrak{g}\diamond\mathfrak{f})_{k,\beta}=\sum_{\ell\geq 1}\sum_{k_{1}+\dots+k_{\ell}=k}\sum_{\beta_{0}+\beta_{1}+\cdots+\beta_{\ell}=\beta}\mathfrak{g}_{\ell,\beta_{0}}\circ(\mathfrak{f}_{k_{1},\beta_{1}}\otimes\cdots\otimes\mathfrak{f}_{k_{\ell},\beta_{\ell}})
(𝔤{𝔥})k,β=λ+μ+ν=kβ+β′′=β𝔤λ+μ+1,β(id#deg𝔥λ𝔥ν,β′′idμ)\displaystyle(\mathfrak{g}\{\mathfrak{h}\})_{k,\beta}=\sum_{\lambda+\mu+\nu=k}\sum_{\beta^{\prime}+\beta^{\prime\prime}=\beta}\mathfrak{g}_{\lambda+\mu+1,\beta^{\prime}}\circ(\mathrm{id}_{\#\deg^{\prime}\mathfrak{h}}^{\lambda}\otimes\mathfrak{h}_{\nu,\beta^{\prime\prime}}\otimes\mathrm{id}^{\mu})

The gappedness ensures that they are finite sums. For instance, if 𝔣0,0\mathfrak{f}_{0,0} could be non-zero, the sum in 𝔤𝔣\mathfrak{g}\diamond\mathfrak{f} would have infinite terms and was not well-defined.

2.1.3 AA_{\infty} structures

By a 𝔊\mathfrak{G}-gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}), we mean an operator system 𝔪=(𝔪k,β)\mathfrak{m}=(\mathfrak{m}_{k,\beta}) in 𝐂𝐂(C,C)\operatorname{\mathbf{CC}}(C,C) such that deg𝔪k,β=2μ(β)k\deg\mathfrak{m}_{k,\beta}=2-\mu(\beta)-k and 𝔪{𝔪}=0\mathfrak{m}\{\mathfrak{m}\}=0, namely

β+β′′=βλ+μ+ν=k𝔪λ+μ+1,β(id#λ𝔪ν,β′′idμ)=0\sum_{\beta^{\prime}+\beta^{\prime\prime}=\beta}\sum_{\lambda+\mu+\nu=k}\mathfrak{m}_{\lambda+\mu+1,\beta^{\prime}}(\mathrm{id}_{\#}^{\lambda}\otimes\mathfrak{m}_{\nu,\beta^{\prime\prime}}\otimes\mathrm{id}^{\mu})=0

A 𝔊\mathfrak{G}-gapped AA_{\infty} homomorphism from (C,𝔪)(C^{\prime},\mathfrak{m}^{\prime}) to (C,𝔪)(C,\mathfrak{m}) is an operator system 𝔣=(𝔣k,β)𝐂𝐂(C,C)\mathfrak{f}=(\mathfrak{f}_{k,\beta})\in\operatorname{\mathbf{CC}}(C^{\prime},C) such that deg𝔣k,β=1μ(β)k\deg\mathfrak{f}_{k,\beta}=1-\mu(\beta)-k and 𝔪𝔣=𝔣{𝔪}\mathfrak{m}\diamond\mathfrak{f}=\mathfrak{f}\{\mathfrak{m}^{\prime}\}, namely

10=j0j=kβ0+β1++β=β𝔪,β0(𝔣j1j0,β1𝔣jj1,β)=λ+μ+ν=kβ+β′′=β𝔣λ+μ+1,β(id#λ𝔪ν,β′′idμ)\displaystyle\sum_{\ell\geq 1}\sum_{\begin{subarray}{c}0=j_{0}\leq\cdots\leq j_{\ell}=k\\ \beta_{0}+\beta_{1}+\cdots+\beta_{\ell}=\beta\end{subarray}}\mathfrak{m}_{\ell,\beta_{0}}(\mathfrak{f}_{j_{1}-j_{0},\beta_{1}}\otimes\cdots\otimes\mathfrak{f}_{j_{\ell}-j_{\ell-1},\beta_{\ell}})=\sum_{\begin{subarray}{c}\lambda+\mu+\nu=k\\ \beta^{\prime}+\beta^{\prime\prime}=\beta\end{subarray}}\mathfrak{f}_{\lambda+\mu+1,\beta^{\prime}}(\mathrm{id}_{\#}^{\lambda}\otimes\mathfrak{m}^{\prime}_{\nu,\beta^{\prime\prime}}\otimes\mathrm{id}^{\mu})

As μ(β)\mu(\beta) is always an even integer, we know deg𝔪=1\deg^{\prime}\mathfrak{m}=1 and deg𝔣=0\deg^{\prime}\mathfrak{f}=0 in 2\mathbb{Z}_{2}. Restricting the 𝔪\mathfrak{m} or 𝔣\mathfrak{f} to the energy-zero part k𝐂𝐂k,0\prod_{k}\operatorname{\mathbf{CC}}_{k,0} (i.e. β=0\beta=0), what we obtain is called the reduction (C,𝔪¯)(C,\bar{\mathfrak{m}}) (resp. 𝔣¯:𝔪¯𝔪¯\bar{\mathfrak{f}}:\bar{\mathfrak{m}}^{\prime}\to\bar{\mathfrak{m}}) of 𝔪\mathfrak{m} (resp. 𝔣\mathfrak{f}). In the energy-zero part, we always have 𝔪1,0𝔪1,0=0\mathfrak{m}_{1,0}\circ\mathfrak{m}_{1,0}=0 and 𝔪1,0𝔣1,0=𝔣1,0𝔪1,0\mathfrak{m}_{1,0}\circ\mathfrak{f}_{1,0}=\mathfrak{f}_{1,0}\circ\mathfrak{m}^{\prime}_{1,0}. Hence, an AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}) naturally induces a cochain complex (C,𝔪1,0)(C,\mathfrak{m}_{1,0}) and an AA_{\infty} homomorphism 𝔣\mathfrak{f} also induces a cochain map 𝔣1,0\mathfrak{f}_{1,0}. Applying the homological perturbation to 𝔪\mathfrak{m} yields a new AA_{\infty} algebra 𝔪^\hat{\mathfrak{m}} with 𝔪^1,0=0\hat{\mathfrak{m}}_{1,0}=0, often called the minimal model.

Let PP be the convex hull of finite points in a Euclidean space such like a dd-simplex Δd\Delta^{d}. Denote by C(P,C)C^{\infty}(P,C) the set of all smooth maps from PP to CC, and we define

CP:=Ω(P)C(P)C(P,C)C_{P}:=\Omega^{*}(P)\otimes_{C^{\infty}(P)}C^{\infty}(P,C)

It naturally comes with the evaluation map

Evals:CPC\operatorname{Eval}^{s}:C_{P}\to C

for any sPs\in P and the inclusion map

Incl:CCP\operatorname{Incl}:C\to C_{P}

Alternatively, it can be regarded as the space of CC-valued differential forms on PP. For instance, when C=Ω(L)C=\Omega^{*}(L) for a manifold LL, we have a natural identification CPΩ(P×L)C_{P}\cong\Omega^{*}(P\times L) via ηxηx\eta\otimes x\xleftrightarrow{}\eta\wedge x. Also, the Evals\operatorname{Eval}^{s} and Incl\operatorname{Incl} are the pullbacks for the inclusion L{s}×LL\xhookrightarrow{}\{s\}\times L and the projection P×LLP\times L\to L.

A PP-pseudo-isotopy (of gapped AA_{\infty} algebras) on CC is simply a special sort of a gapped AA_{\infty} algebra structure 𝔐\mathfrak{M} on CPC_{P} such that the 𝔐\mathfrak{M} is Ω(P)\Omega^{*}(P)-linear (up to sign) and the 𝔐1,0\mathfrak{M}_{1,0} is the canonical differential on CPC_{P}. In the special case C=Ω(L)C=\Omega^{*}(L), the 𝔐1,0\mathfrak{M}_{1,0} coincides with the exterior derivative dP×Ld_{P\times L}. The notion of pseudo-isotopy is introduced in [Fuk10]. See also [Yua20] for the details.

2.1.4 Unitality and divisor axiom

An arbitrary operator system 𝔱𝐂𝐂\mathfrak{t}\in\operatorname{\mathbf{CC}} is called cyclically unital if, for any degree-zero element 𝐞\mathbf{e} and any (k,β)(0,0)(k,\beta)\neq(0,0), we have

CU[𝔱]k,β(𝐞;x1,,xk):=i=1k+1𝔱k+1,β(x1#,,xi1#,𝐞,xi,,xk)=0\textstyle\operatorname{CU}[\mathfrak{t}]_{k,\beta}(\mathbf{e};x_{1},\dots,x_{k}):=\sum_{i=1}^{k+1}\mathfrak{t}_{k+1,\beta}(x_{1}^{\#},\dots,x_{i-1}^{\#},\mathbf{e},x_{i},\dots,x_{k})=0

On the other hand, a gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}) is called (strictly) unital if there exists a degree-zero element 1, called a unit, such that

  1. 1.

    𝔪1,0(1)=0\mathfrak{m}_{1,0}(\text{1})=0

  2. 2.

    𝔪2,0(1,x)=(1)degx𝔪2,0(x,1)=x\mathfrak{m}_{2,0}(\text{1},x)=(-1)^{\deg x}\mathfrak{m}_{2,0}(x,\text{1})=x

  3. 3.

    𝔪k,β(,1,)=0\mathfrak{m}_{k,\beta}(\dots,\text{1},\dots)=0 for (k,β)(1,0),(2,0)(k,\beta)\neq(1,0),(2,0).

Assume (C,𝔪)(C^{\prime},\mathfrak{m}^{\prime}) and (C,𝔪)(C,\mathfrak{m}) have the units 1\text{1}^{\prime} and 1; then, a gapped AA_{\infty} homomorphism 𝔣:𝔪𝔪\mathfrak{f}:\mathfrak{m}^{\prime}\to\mathfrak{m} is called unital if 𝔣1,0(1)=1\mathfrak{f}_{1,0}(\text{1}^{\prime})=\text{1} and

𝔣k,β(,1,)=0\mathfrak{f}_{k,\beta}(\dots,\text{1}^{\prime},\dots)=0

for (k,β)(1,0)(k,\beta)\neq(1,0).

A gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}) is called a quantum correction to de Rham complex, or in abbreviation, a q.c.dR, if CC is Ω(M)\Omega^{*}(M) for some manifold MM so that 𝔪1,0=dM\mathfrak{m}_{1,0}=d_{M}, 𝔪2,0(x1,x2)=(1)degx1x1x2\mathfrak{m}_{2,0}(x_{1},x_{2})=(-1)^{\deg x_{1}}x_{1}\wedge x_{2}, and 𝔪k,0=0\mathfrak{m}_{k,0}=0 for k3k\geq 3. Geometrically, an AA_{\infty} algebra obtained by the moduli spaces or any of its minimal model can satisfy the q.c.dR conditions.

From now on, we always assume 𝔊=π2(X,L)\mathfrak{G}=\pi_{2}(X,L) and C=H(L)PC=H^{*}(L)_{P} or Ω(L)P\Omega^{*}(L)_{P} for some LL and PP. In either of two cases, we have a natural differential dd on CC for which there is a well-defined cap product β\partial\beta\cap\cdot on

Z1(C):={bCdb=0,degb=1}Z^{1}(C):=\{b\in C\mid db=0,\deg b=1\}

for any βπ2(X,L)\beta\in\pi_{2}(X,L). Now, we say an operator system 𝔱=(𝔱k,β)\mathfrak{t}=(\mathfrak{t}_{k,\beta}) in 𝐂𝐂(C,C)\operatorname{\mathbf{CC}}(C,C^{\prime}) satisfies the divisor axiom if for any bZ1(C)b\in Z^{1}(C) and (k,β)(0,0)(k,\beta)\neq(0,0),

DA[𝔱]k,β(b;x1,,xk):=i=1k+1𝔱k+1,β(x1,,xi1,b,xi,,xk)=βb𝔱k,β(x1,,xk)\textstyle\operatorname{DA}[\mathfrak{t}]_{k,\beta}(b;x_{1},\dots,x_{k}):=\sum_{i=1}^{k+1}\mathfrak{t}_{k+1,\beta}(x_{1},\dots,x_{i-1},b,x_{i},\dots,x_{k})=\partial\beta\cap b\cdot\mathfrak{t}_{k,\beta}(x_{1},\dots,x_{k})

2.1.5 The category 𝒰𝒟\mathscr{UD}

Given a symplectic manifold (X,ω)(X,\omega), let LL be an oriented spin Lagrangian submanifold. We remark that LL is not necessarily a Lagrangian torus. Set 𝔊=π2(X,L)\mathfrak{G}=\pi_{2}(X,L). For simplicity, we use a uniform notation 1 to denote all various constant-one functions in either H(L)PH^{*}(L)_{P} or Ω(L)P\Omega^{*}(L)_{P}. In [Yua20, §2], we introduced a category 𝒰𝒟\mathscr{UD} as follows:

  1. (I)

    An object 𝔪\mathfrak{m} in 𝒰𝒟\mathscr{UD} is a 𝔊\mathfrak{G}-gapped AA_{\infty} algebra with the following properties:

    • (I-0)

      it extends H(L)PH^{*}(L)_{P} or Ω(L)P\Omega^{*}(L)_{P};

    • (I-1)

      the constant-one function is a (strict) unit;

    • (I-2)

      the cyclical unitality;

    • (I-3)

      the divisor axiom;

    • (I-4)

      it is a PP-pseudo-isotopy;

    • (I-5)

      every βπ2(X,L)\beta\in\pi_{2}(X,L) in the set 𝖦𝔪:={β𝔪β0}\mathsf{G}_{\mathfrak{m}}:=\{\beta\mid\mathfrak{m}_{\beta}\neq 0\} satisfies μ(β)0\mu(\beta)\geq 0.

  2. (II)

    A morphism 𝔣\mathfrak{f} in 𝒰𝒟{\mathscr{UD}} is a 𝔊\mathfrak{G}-gapped AA_{\infty} homomorphism with the following properties:

    • (II-1)

      the strict unitalities with the constant-one functions as units;

    • (II-2)

      the cyclically unitality;

    • (II-3)

      the divisor axiom;

    • (II-4)

      β𝔣1,0(b)=βb\partial\beta\cap\mathfrak{f}_{1,0}(b)=\partial\beta\cap b for any divisor input bZ1(C)b\in Z^{1}(C);

    • (II-5)

      every βπ2(X,L)\beta\in\pi_{2}(X,L) in the set 𝖦𝔣:={β𝔊𝔣β0}\mathsf{G}_{\mathfrak{f}}:=\{\beta\in\mathfrak{G}\mid\mathfrak{f}_{\beta}\neq 0\} satisfies μ(β)0\mu(\beta)\geq 0.

We also call (I-5) and (II-5) the semipositive conditions. We write Obj𝒰𝒟\operatorname{Obj}\mathscr{UD} and Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} for the collections of objects and morphisms in 𝒰𝒟\mathscr{UD} respectively.

2.1.6 Homotopy theory and Whitehead theorem

For 𝒰𝒟\mathscr{UD}, there is the notion of ud-homotopy: We call

𝔣0,𝔣1Hom𝒰𝒟((C,𝔪),(C,𝔪))\mathfrak{f}_{0},\mathfrak{f}_{1}\in\operatorname{Hom}_{\mathscr{UD}}((C^{\prime},\mathfrak{m}^{\prime}),(C,\mathfrak{m}))

are ud-homotopic if there is

𝔉Hom𝒰𝒟((C,𝔪),(C[0,1],𝔐tri))\mathfrak{F}\in\operatorname{Hom}_{\mathscr{UD}}((C^{\prime},\mathfrak{m}^{\prime}),(C_{[0,1]},\mathfrak{M}^{\mathrm{tri}}))

such that

Eval0𝔉=𝔣0,Eval1𝔉=𝔣1\operatorname{Eval}^{0}\mathfrak{F}=\mathfrak{f}_{0}\quad,\quad\operatorname{Eval}^{1}\mathfrak{F}=\mathfrak{f}_{1}

where 𝔐tri\mathfrak{M}^{\mathrm{tri}} is the trivial pseudo-isotopy about 𝔪\mathfrak{m} and can be proved to be an object in 𝒰𝒟\mathscr{UD} as well. In this case, we write 𝔣0ud𝔣1\mathfrak{f}_{0}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathfrak{f}_{1}. Moreover, 𝔣0ud𝔣1\mathfrak{f}_{0}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathfrak{f}_{1} if and only if there exists (𝔣s)(\mathfrak{f}_{s}) and (𝔥s)(\mathfrak{h}_{s}) in 𝐂𝐂(C,C)\operatorname{\mathbf{CC}}(C^{\prime},C) such that: (see [Yua20, §2])

  1. (a)

    Every 𝔣sMor𝒰𝒟\mathfrak{f}_{s}\in\operatorname{Mor}\mathscr{UD};

  2. (b)

    dds𝔣s=𝔥s(id#𝔪id)+𝔪(𝔣s#𝔣s#𝔥s𝔣s𝔣s)\frac{d}{ds}\circ\mathfrak{f}_{s}=\sum\mathfrak{h}_{s}\circ(\mathrm{id}_{\#}^{\bullet}\otimes\mathfrak{m}^{\prime}\otimes\mathrm{id}^{\bullet})+\sum\mathfrak{m}\circ(\mathfrak{f}_{s}^{\#}\otimes\cdots\otimes\mathfrak{f}_{s}^{\#}\otimes\mathfrak{h}_{s}\otimes\mathfrak{f}_{s}\otimes\cdots\otimes\mathfrak{f}_{s});

  3. (c)

    The 𝔥s\mathfrak{h}_{s} satisfies the divisor axiom, the cyclical unitality, and (𝔥s)k,β(1)=0(\mathfrak{h}_{s})_{k,\beta}(\cdots\text{1}\cdots)=0 for all (k,β)(k,\beta);

  4. (d)

    deg(𝔥s)k,β=kμ(β)\deg(\mathfrak{h}_{s})_{k,\beta}=-k-\mu(\beta). For every β\beta with 𝔥β0\mathfrak{h}_{\beta}\neq 0, we have μ(β)0\mu(\beta)\geq 0.

In the context of 𝒰𝒟\mathscr{UD}, we also have the Whitehead theorem [Yua20, §3]:

Theorem 2.2 (Whitehead).

Fix 𝔣Hom𝒰𝒟((C,𝔪),(C,𝔪))\mathfrak{f}\in\operatorname{Hom}_{\mathscr{UD}}((C^{\prime},\mathfrak{m}^{\prime}),(C,\mathfrak{m})) such that 𝔣1,0\mathfrak{f}_{1,0} is a quasi-isomorphism of cochain complexes. Then, there exists 𝔤Hom𝒰𝒟((C,𝔪),(C,𝔪))\mathfrak{g}\in\operatorname{Hom}_{\mathscr{UD}}((C,\mathfrak{m}),(C^{\prime},\mathfrak{m}^{\prime})), unique up to ud-homotopy, such that 𝔤𝔣udidC\mathfrak{g}\diamond\mathfrak{f}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id}_{C^{\prime}} and 𝔣𝔤udidC\mathfrak{f}\diamond\mathfrak{g}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id}_{C}. We call 𝔤\mathfrak{g} a ud-homotopy inverse of 𝔣\mathfrak{f}.

2.2 An observation about weak Maurer-Cartan equations

The semipositive condition (I-5) says that for 𝔪Obj𝒰𝒟\mathfrak{m}\in\operatorname{Obj}\mathscr{UD}, we have 𝔪β0\mathfrak{m}_{\beta}\neq 0 only if μ(β)0\mu(\beta)\geq 0. Hence, deg𝔪0,β=2μ(β)2\deg\mathfrak{m}_{0,\beta}=2-\mu(\beta)\leq 2. If the 𝔪\mathfrak{m} is defined on H(L)H^{*}(L), then we may write as in [Yua20]:

(7) W1+Q:=μ(β)=2TE(β)Yβ𝔪0,β+μ(β)=0TE(β)Yβ𝔪0,βW\cdot\text{1}+Q:=\sum_{\mu(\beta)=2}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{0,\beta}+\sum_{\mu(\beta)=0}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{0,\beta}
Remark 2.3

The moduli space of holomorphic disks bounding a Lagrangian submanifold LL gives an AA_{\infty} algebra 𝔪ˇ\check{\mathfrak{m}} on Ω(L)\Omega^{*}(L) at first. But, we need to use its minimal model (H(L),𝔪)(H^{*}(L),\mathfrak{m}). Otherwise, the first summation in (7) would live in Λ[[π1(L)]]^Ω0(L)\Lambda[[\pi_{1}(L)]]\hat{\otimes}\Omega^{0}(L) and may not be in the form of W1W\cdot\text{1}.

Let 𝔞{\mathfrak{a}} denote the ideal in Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] generated by the components of QQ. The ideal 𝔞{\mathfrak{a}} vanishes under Assumption 1.3, but let us keep it for a moment to highlight the ideas. Basically, one may view (7) as the (weak) Maurer-Cartan equations with certain differences. We review the following lemma whose proof is also implicitly given in [Yua20]:

Lemma 2.4.

If 𝔣0ud𝔣1\mathfrak{f}_{0}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathfrak{f}_{1} from (H(L),𝔪)(H^{*}(L),\mathfrak{m}^{\prime}) to (H(L),𝔪)(H^{*}(L),\mathfrak{m}) and (𝔣0)1,0=(𝔣1)1,0(\mathfrak{f}_{0})_{1,0}=(\mathfrak{f}_{1})_{1,0}, then

βTE(β)Yβ((𝔣1)0,β(𝔣0)0,β)𝔞\textstyle\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}((\mathfrak{f}_{1})_{0,\beta}-(\mathfrak{f}_{0})_{0,\beta})\in{\mathfrak{a}}

In particular, under Assumption 1.3, we have

βTE(β)Yβ(𝔣0)0,β=βTE(β)Yβ(𝔣1)0,β\textstyle\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\mathfrak{f}_{0})_{0,\beta}=\textstyle\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\mathfrak{f}_{1})_{0,\beta}
Proof.

Let (𝔣s)(\mathfrak{f}_{s}) and (𝔥s)(\mathfrak{h}_{s}) be as in §2.1.6 with the conditions (a) (b) (c) (d) therein. For a basis of π1(L)\pi_{1}(L), we have Λ[[π1(L)]]Λ[[Y1±,,Ym±]]\Lambda[[\pi_{1}(L)]]\cong\Lambda[[Y_{1}^{\pm},\dots,Y_{m}^{\pm}]]. For a basis of H2(L)H^{2}(L), we write Q=ηQηηQ=\sum_{\eta}Q_{\eta}\cdot\eta, and then by definition, the ideal 𝔞{\mathfrak{a}} is generated by all the QηQ_{\eta}. By Lemma 2.1, it suffices to restrict

S:=TE(β)Yβ((𝔣1)0,β(𝔣0)0,β)S:=\sum T^{E(\beta)}Y^{\partial\beta}((\mathfrak{f}_{1})_{0,\beta}-(\mathfrak{f}_{0})_{0,\beta})

to an arbitrary point 𝐲=(y1,,ym)\mathbf{y}=(y_{1},\dots,y_{m}) in UΛmH1(L;UΛ)U_{\Lambda}^{m}\cong H^{1}(L;U_{\Lambda}). We can write yi=exiy_{i}=e^{x_{i}} for xiΛ0x_{i}\in\Lambda_{0}. We set b=(x1,,xm)b=(x_{1},\dots,x_{m}) for the basis. Applying the divisor axiom of 𝔣0,𝔣1\mathfrak{f}_{0},\mathfrak{f}_{1} and the condition (b) implies that

S(y1,,ym)\displaystyle S(y_{1},\dots,y_{m}) =k,βTE(β)((𝔣1)k,β(b,,b)(𝔣0)k,β(b,,b))\displaystyle=\sum_{k,\beta}T^{E(\beta)}\big{(}(\mathfrak{f}_{1})_{k,\beta}(b,\dots,b)-(\mathfrak{f}_{0})_{k,\beta}(b,\dots,b)\big{)}
=TE(β)01𝑑sdds(𝔣s)k,β(b,,b)\displaystyle=\sum T^{E(\beta)}\int_{0}^{1}ds\cdot\frac{d}{ds}\circ(\mathfrak{f}_{s})_{k,\beta}(b,\dots,b)
=TE(β1)01𝑑s(𝔥s)λ+μ+1,β1(b,,b,TE(β2)𝔪ν,β2(b,,b),b,b)\displaystyle=\sum T^{E(\beta_{1})}\int_{0}^{1}ds\cdot(\mathfrak{h}_{s})_{\lambda+\mu+1,\beta_{1}}(b,\dots,b,T^{E(\beta_{2})}\mathfrak{m}^{\prime}_{\nu,\beta_{2}}(b,\dots,b),b,\dots b)
+TE(β)01𝑑s𝔪((𝔣s)(b,,b),,(𝔥s),β0(b,,b),,(𝔣s)(b,,b))\displaystyle+\sum T^{E(\beta)}\int_{0}^{1}ds\cdot\mathfrak{m}\big{(}(\mathfrak{f}_{s})(b,\dots,b),\dots,(\mathfrak{h}_{s})_{\ell,\beta_{0}}(b,\dots,b),\dots,(\mathfrak{f}_{s})(b,\dots,b)\big{)}

Since deg(𝔥s),β0(b,,b)=μ(β0)\deg(\mathfrak{h}_{s})_{\ell,\beta_{0}}(b,\dots,b)=-\mu(\beta_{0}), the semipositive condition and the cyclical unitality of 𝔪\mathfrak{m} deduce that the second summation vanishes. Hence, for the operator :=01𝑑s𝔥s\mathfrak{H}:=\int_{0}^{1}ds\cdot\mathfrak{h}_{s}, it follows from the condition (c) that

S(y1,,ym)=βTE(β)𝐲β1,β(W(𝐲)1+ηQη(𝐲)η)=ηQη(𝐲)βTE(β)𝐲β1,β(η)S(y_{1},\dots,y_{m})=\sum_{\beta}T^{E(\beta)}\mathbf{y}^{\partial\beta}\mathfrak{H}_{1,\beta}\big{(}W(\mathbf{y})\cdot\text{1}+\sum_{\eta}Q_{\eta}(\mathbf{y})\cdot\eta\big{)}=\sum_{\eta}Q_{\eta}(\mathbf{y})\cdot\sum_{\beta}T^{E(\beta)}\mathbf{y}^{\partial\beta}\mathfrak{H}_{1,\beta}(\eta)

Now that the above equation holds for any 𝐲=(yi)UΛm\mathbf{y}=(y_{i})\in U_{\Lambda}^{m}, it also holds identically by Lemma 2.1. ∎

2.3 Wall-crossing from family Floer viewpoint: Review

Here we review the wall-crossing aspect of the family Floer framework in [Yua20]. In this section, we provide a summary of the basic points and refer to [Yua20] for the details.

2.3.1 AA_{\infty} structures

Let JJ be an ω\omega-tame almost complex structure in XX. Fix kk\in\mathbb{N} and βπ2(X,L)\beta\in\pi_{2}(X,L) with (k,β)(0,0),(1,0)(k,\beta)\neq(0,0),(1,0). We consider the moduli space k+1,β(J,L)\mathcal{M}_{k+1,\beta}(J,L) of the equivalence classes of (k+1)(k+1)-boundary-marked JJ-holomorphic genus-zero stable maps with one boundary component in LL in the class β\beta. Let 𝕄(J)\mathbb{M}(J) denote the collection of the moduli spaces k+1,β(J,L)\mathcal{M}_{k+1,\beta}(J,L) for all (k,β)(k,\beta). We call 𝕄(J)\mathbb{M}(J) a moduli system. Using the virtual techniques [FOOO20], it gives rise to a chain-level AA_{\infty} algebra denoted by (Ω(L),𝔪ˇJ,L)(\Omega^{*}(L),\check{\mathfrak{m}}^{J,L}). Suppose JJ is semipositive in the sense that there does not exist any JJ-holomorphic disk in π2(X,L)\pi_{2}(X,L) with negative Maslov index. Then, one can check that 𝔪ˇ=𝔪ˇJ,L\check{\mathfrak{m}}=\check{\mathfrak{m}}^{J,L} is an object in 𝒰𝒟=𝒰𝒟(L)\mathscr{UD}=\mathscr{UD}(L).

The family Floer theory only works with the cohomology-level AA_{\infty} algebras; see e.g. Remark 2.3. We need to perform the homological perturbation to the above 𝔪ˇJ,L\check{\mathfrak{m}}^{J,L}. Let gg be a metric. By the Hodge decomposition, one can find two cochain maps i(g):H(L)Ω(L)i(g):H^{*}(L)\to\Omega^{*}(L), π(g):Ω(L)H(L)\pi(g):\Omega^{*}(L)\to H^{*}(L) of degree zero and a map G(g):Ω(L)Ω(L)G(g):\Omega^{*}(L)\to\Omega^{*}(L) of degree 1-1 such that i(g)π(g)id=dLG+GdLi(g)\circ\pi(g)-\mathrm{id}=d_{L}\circ G+G\circ d_{L}, π(g)i(g)=id\pi(g)\circ i(g)=\mathrm{id}, G(g)G(g)=0G(g)\circ G(g)=0, G(g)i(g)=0G(g)\circ i(g)=0, and π(g)G(g)=0\pi(g)\circ G(g)=0. Now, we call 𝖼𝗈𝗇(g):=(i(g),π(g),G(g))\mathsf{con}(g):=(i(g),\pi(g),G(g)) the harmonic contraction. Applying the homological perturbation with this 𝖼𝗈𝗇(g)\mathsf{con}(g), the AA_{\infty} algebra 𝔪ˇJ,L\check{\mathfrak{m}}^{J,L} induces its minimal model denoted by (H(L),𝔪g,J,L)(H^{*}(L),\mathfrak{m}^{g,J,L}). It is an object in 𝒰𝒟\mathscr{UD} and also accompanied by a morphism 𝔦g,J,L:(H(L),𝔪g,J,L)(Ω(L),𝔪ˇJ,L)\mathfrak{i}^{g,J,L}:\ (H^{*}(L),\mathfrak{m}^{g,J,L})\to(\Omega^{*}(L),\check{\mathfrak{m}}^{J,L}) in 𝒰𝒟\mathscr{UD}. Next, for a path of metrics 𝐠=(gs)s[0,1]\mathbf{g}=(g_{s})_{s\in{[0,1]}}, there is a parameterized harmonic contraction 𝖼𝗈𝗇(𝐠)\mathsf{con}(\mathbf{g}); further using the parameterized moduli spaces for a path 𝑱=(Js)s[0,1]\boldsymbol{J}=(J_{s})_{s\in{[0,1]}}, we can naturally produce an AA_{\infty} homotopy equivalence :(H(L),𝔪g0,J0,L)(H(L),𝔪g1,J1,L)\mathfrak{C}:(H^{*}(L),\mathfrak{m}^{g_{0},J_{0},L})\to(H^{*}(L),\mathfrak{m}^{g_{1},J_{1},L}) in 𝒰𝒟\mathscr{UD} such that 1,0=id\mathfrak{C}_{1,0}=\mathrm{id}. It always comes from a pseudo-isotopy, and it does not depend on various choices up to the ud-homotopy relation.

2.3.2 Analytic coordinate change maps

We take an adjacent Lagrangian submanifold L~\tilde{L} in a Weinstein neighborhood 𝒰L\mathcal{U}_{L} of LL such that there exists a small isotopy FF supported near LL such that F(L)=L~F(L)=\tilde{L}. Choose a tautological 1-form λ\lambda on 𝒰L\mathcal{U}_{L} such that λ\lambda vanishes exactly on L~\tilde{L} and ω=dλ\omega=d\lambda. Then, one can use the Stokes’ formula to show E(β~)=E(β)βλ|LE(\tilde{\beta})=E(\beta)-\partial\beta\cap\lambda|_{L} for any βπ2(X,L)\beta\in\pi_{2}(X,L) and β~:=Fβπ2(X,L~)\tilde{\beta}:=F_{*}\beta\in\pi_{2}(X,\tilde{L}). Note that λ:=λ|L\lambda:=\lambda|_{L} is a closed one-form on LL. Conventionally, we distinguish between π2(X,L)\pi_{2}(X,L) and π2(X,L~)\pi_{2}(X,\tilde{L}) due to the energy change, but we may feel free to identify π1(L)\pi_{1}(L) with π1(L~)\pi_{1}(\tilde{L}) for adjacent fibers.

By [Fuk10], Fukaya’s trick essentially refers to the observation of a natural correspondence between a JJ-holomorphic curve uu with boundary on the Lagrangian LL and an FJF_{*}J-holomorphic curve FuF\circ u with boundary on L~\tilde{L}.

Convention 2.5.

We say :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) is an AA_{\infty} homotopy equivalence up to the Fukaya’s trick, if it is an AA_{\infty} homotopy equivalence in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} from (H(L),𝔪)(H^{*}(L),\mathfrak{m}) to (H(L),𝔪~F)(H^{*}(L),\tilde{\mathfrak{m}}^{F}) such that 1,0=F1\mathfrak{C}_{1,0}=F^{-1*} and the AA_{\infty} algebra 𝔪~F\tilde{\mathfrak{m}}^{F} is defined in view of Fukaya’s trick by

𝔪~k,βF=F𝔪~k,β~(F1,,F1)\tilde{\mathfrak{m}}^{F}_{k,\beta}=F^{*}\tilde{\mathfrak{m}}_{k,\tilde{\beta}}(F^{-1*},\dots,F^{-1*})

In practice, such a \mathfrak{C} is obtained in almost the same way as §2.3.1 but further using the Fukaya’s trick.

Suppose :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) is an AA_{\infty} homotopy equivalence in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} up to the Fukaya’s trick as above. We write W,Q,𝔞W,Q,{\mathfrak{a}} and W~,Q~,𝔞~\tilde{W},\tilde{Q},\tilde{\mathfrak{a}} as in (7) for 𝔪\mathfrak{m} and 𝔪~\tilde{\mathfrak{m}} respectively. We write

(8) 𝕱(Y)=TE(γ)Yγ0,γΛ[[π1(L)]]^H(L)\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)=\sum T^{E(\gamma)}Y^{\partial\gamma}\mathfrak{C}_{0,\gamma}\ \ \in\Lambda[[\pi_{1}(L)]]\hat{\otimes}H^{*}(L)

By the semipositive condition, we know any nonzero 0,γH1(L)\mathfrak{C}_{0,\gamma}\in H^{1}(L). By the gappedness, 0,0=0\mathfrak{C}_{0,0}=0. The following lemma describes the wall-crossing phenomenon precisely.

Lemma 2.6 ([Yua20]).

The assignment ϕ:Y~αTα,λYαexpα,𝕱\phi:\tilde{Y}^{\alpha}\mapsto T^{\langle\alpha,\lambda\rangle}Y^{\alpha}\exp\langle\alpha,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}\rangle defines an isomorphism

(9) ϕ=ϕ:Λ[[π1(L~)]]/𝔞~Λ[[π1(L)]]/𝔞\phi=\phi_{\mathfrak{C}}:\Lambda[[\pi_{1}(\tilde{L})]]/\tilde{\mathfrak{a}}\to\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}

such that ϕ(W~)=W\phi(\tilde{W})=W. Moreover, this map ϕ\phi only depends on the ud-homotopy class of \mathfrak{C}. Under Assumption 1.3, we actually have 𝔞=𝔞~=0{\mathfrak{a}}=\tilde{\mathfrak{a}}=0 and

ϕ=ϕ:Λ[[π1(L~)]]Λ[[π1(L)]]\phi=\phi_{\mathfrak{C}}:\Lambda[[\pi_{1}(\tilde{L})]]\to\Lambda[[\pi_{1}(L)]]

We use the formal symbols YαY^{\alpha} and Y~α\tilde{Y}^{\alpha} to distinguish the monomials in Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] and Λ[[π1(L~)]]\Lambda[[\pi_{1}(\tilde{L})]] respectively, where απ1(L)π1(L~)\alpha\in\pi_{1}(L)\cong\pi_{1}(\tilde{L}). From a different perspective, the coordinate change map ϕ\phi corresponds to a rigid analytic map

(10) ϕ:𝒰H1(L;Λ)H1(L~;Λ)\phi:\mathcal{U}\subset H^{1}(L;\Lambda^{*})\to H^{1}(\tilde{L};\Lambda^{*})

where the domain 𝒰\mathcal{U} is usually a proper subset but always contains H1(L;UΛ)H^{1}(L;U_{\Lambda}) by Gromov’s compactness. When H1(L;Λ)H^{1}(L;\Lambda^{*}) is identified with (Λ)n(\Lambda^{*})^{n}, and we often take the domain 𝒰\mathcal{U} to be a polytopal affinoid domain 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta) for a rational polyhedron Δ\Delta in n\mathbb{R}^{n}, c.f. (3). Concretely, the image point 𝐲~:=ϕ(𝐲)\tilde{\mathbf{y}}:=\phi(\mathbf{y}) in H1(L~;Λ)H^{1}(\tilde{L};\Lambda^{*}) is such that 𝐲~α=Tα,λ𝐲αexpα,𝕱(𝐲)\tilde{\mathbf{y}}^{\alpha}=T^{\langle\alpha,\lambda\rangle}\mathbf{y}^{\alpha}\exp\langle\alpha,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(\mathbf{y})\rangle for απ1(L)π1(L~)\alpha\in\pi_{1}(L)\cong\pi_{1}(\tilde{L}), where the 𝐲α\mathbf{y}^{\alpha} denotes the image of (𝐲,α)(\mathbf{y},\alpha) under the pairing H1(L;Λ)×π1(L)ΛH^{1}(L;\Lambda^{*})\times\pi_{1}(L)\to\Lambda^{*}. Further, chosen a basis, this means y~i=Tciyiexp(𝔉i(y1,,yn))\tilde{y}_{i}=T^{c_{i}}y_{i}\exp\big{(}\mathfrak{F}_{i}(y_{1},\dots,y_{n})\big{)}. One can check that whenever 𝐲H1(L;UΛ)\mathbf{y}\in H^{1}(L;U_{\Lambda}), we have 𝐲~H1(L~;UΛ)\tilde{\mathbf{y}}\in H^{1}(\tilde{L};U_{\Lambda}). The above lemma tells that W~(𝐲~)=W(𝐲)\tilde{W}(\tilde{\mathbf{y}})=W(\mathbf{y}). Conventionally, the ϕ\phi will be called a (family Floer) analytic coordinate change.

3 Self Floer cohomology for the category 𝒰𝒟\mathscr{UD}

3.1 Non-curved AA_{\infty} structures

3.1.1 Definition

Fix a cohomology-level AA_{\infty} algebra (H(L),𝔪)Obj𝒰𝒟(H^{*}(L),\mathfrak{m})\in\operatorname{Obj}\mathscr{UD}2.1.5). We define

(11) 𝐦k:=βπ2(X,L)TE(β)Yβ𝔪k,β:H(L)kH(L)^Λ[[π1(L)]]\mathbf{m}_{k}:=\sum_{\beta\in\pi_{2}(X,L)}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{k,\beta}\ :H^{*}(L)^{\otimes k}\to H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]

where we can further extend each 𝔪k,β\mathfrak{m}_{k,\beta} by linearity so that the 𝐦k\mathbf{m}_{k} is Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]]-linear. The 𝐦k\mathbf{m}_{k} is actually Λ0[[π1(L)]]\Lambda_{0}[[\pi_{1}(L)]]-linear, since E(β)0E(\beta)\geq 0 whenever 𝔪k,β0\mathfrak{m}_{k,\beta}\neq 0. Also, the semipositive condition (§2.1.5) tells that μ(β)0\mu(\beta)\geq 0 if 𝔪k,β0\mathfrak{m}_{k,\beta}\neq 0 for some kk. Remark that since we really use higher Maslov indices like μ(β)>2\mu(\beta)>2 here, there may be more information extracted from 𝒰𝒟\mathscr{UD} than in [Yua20].

Be cautious that 𝐦0=W1+Q\mathbf{m}_{0}=W\cdot\text{1}+Q by (7) is only available in the cohomology level (Remark 2.3). To develop a non-curved AA_{\infty} algebra, we need to exclude 𝐦0\mathbf{m}_{0} in our consideration. Notice that the coefficient ring Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] may be replaced by some quotient ring. Although the ideal 𝔞{\mathfrak{a}} vanishes under Assumption 1.3, we retain it to preserve a more general framework. The vanishing of 𝔞{\mathfrak{a}} becomes relevant only when considering the derivatives and critical points of WW.

Proposition 3.1.

The collection 𝐦={𝐦k}k1\mathbf{m}=\{\mathbf{m}_{k}\}_{k\geq 1} forms a (non-curved) AA_{\infty} algebra over Λ[[π1(L)]]/𝔞\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}.

Proof.

The AA_{\infty} associativity of 𝔪=(𝔪k,β)\mathfrak{m}=(\mathfrak{m}_{k,\beta}) deduces that (𝔪{𝔪})k,β=0(\mathfrak{m}\{\mathfrak{m}\})_{k,\beta}=0 for all pairs (k,β)(k,\beta). It implies that βTE(β)Yβ(𝔪{𝔪})k,β(x1,,xk)=0\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\mathfrak{m}\{\mathfrak{m}\})_{k,\beta}(x_{1},\dots,x_{k})=0 for any fixed kk. Therefore,

λ+μ+ν=kλ,μ0,ν1𝐦λ+μ+1(x1#,,xλ#,𝐦ν(xλ+1,,xλ+ν),,xk)\displaystyle\sum_{\begin{subarray}{c}\lambda+\mu+\nu=k\\ \lambda,\mu\geq 0,\nu\geq 1\end{subarray}}\mathbf{m}_{\lambda+\mu+1}\big{(}x_{1}^{\#},\dots,x_{\lambda}^{\#},\mathbf{m}_{\nu}(x_{\lambda+1},\dots,x_{\lambda+\nu}),\dots,x_{k}\big{)}
=λβTE(β)Yβ𝔪k+1,β(x1#,,xλ#,𝐦0,xλ+1,,xk)=0mod𝔞\displaystyle=-\sum_{\lambda}\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{k+1,\beta}(x_{1}^{\#},\dots,x_{\lambda}^{\#},\mathbf{m}_{0},x_{\lambda+1},\dots,x_{k})=0\mod{\mathfrak{a}}

The reasons are as follows. First, all of 𝔪0,β\mathfrak{m}_{0,\beta}-terms are collected on the right, which forms 𝐦0=W1+Q\mathbf{m}_{0}=W\cdot\text{1}+Q. Recall the ideal 𝔞{\mathfrak{a}} is generated by the components of QQ, so 𝐦0W1\mathbf{m}_{0}\equiv W\cdot\text{1} (mod 𝔞{\mathfrak{a}}). Finally, the unitality of 𝔪\mathfrak{m} deduces that 0λk𝔪k+1,β(x1#,,xλ#,1,xλ+1,,xk)=0\sum_{0\leq\lambda\leq k}\mathfrak{m}_{k+1,\beta}(x_{1}^{\#},\dots,x_{\lambda}^{\#},\text{1},x_{\lambda+1},\dots,x_{k})=0 for every β\beta. ∎

Corollary 3.2.

𝐦1𝐦1=0\mathbf{m}_{1}\circ\mathbf{m}_{1}=0, 𝐦1(𝐦2(x,y))+𝐦2(x#,𝐦1(y))+𝐦2(𝐦1(x),y)=0\mathbf{m}_{1}\big{(}\mathbf{m}_{2}(x,y)\big{)}+\mathbf{m}_{2}(x^{\#},\mathbf{m}_{1}(y))+\mathbf{m}_{2}(\mathbf{m}_{1}(x),y)=0, and

𝐦2(𝐦2(x,y),z)+𝐦2(x#,𝐦2(y,z))+𝐦1(𝐦3(x,y,z))+𝐦3(𝐦1(x),y,z)+𝐦3(x#,𝐦1(y),z)+𝐦3(x#,y#,𝐦1(z))=0\mathbf{m}_{2}(\mathbf{m}_{2}(x,y),z)+\mathbf{m}_{2}(x^{\#},\mathbf{m}_{2}(y,z))+\mathbf{m}_{1}(\mathbf{m}_{3}(x,y,z))+\mathbf{m}_{3}(\mathbf{m}_{1}(x),y,z)+\mathbf{m}_{3}(x^{\#},\mathbf{m}_{1}(y),z)+\mathbf{m}_{3}(x^{\#},y^{\#},\mathbf{m}_{1}(z))=0

Definition 3.3.

The self Floer cohomology of LL (associated to 𝔪\mathfrak{m}) is defined to be the 𝐦1\mathbf{m}_{1}-cohomology:

(12) HF(L,𝔪):=H(H(L)^Λ[[π1(L)]]/𝔞,𝐦1)\operatorname{HF}(L,\mathfrak{m}):=H^{*}\big{(}H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}},\mathbf{m}_{1}\big{)}

Moreover, [x][y]=[𝐦2(x#,y)][x]\cdot[y]=[\mathbf{m}_{2}(-x^{\#},y)] defines a ring structure on HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) with a unit [1][\text{1}].

Note that we may also replace the formal power series ring Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] by any polyhedral affinoid algebra contained in it. Since deg𝔪1,β=1μ(β)1(mod 2)\deg\mathfrak{m}_{1,\beta}=1-\mu(\beta)\equiv 1\ (\mathrm{mod}\ 2), the HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) is at least 2\mathbb{Z}_{2}-graded.

Suppose :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) is an AA_{\infty} homotopy equivalence up to the Fukaya’s trick in the sense of Convention 2.5 with respect to an isotopy FF such that F(L)=L~F(L)=\tilde{L}. Consider the two non-curved AA_{\infty} algebras 𝐦\mathbf{m} and 𝐦~\tilde{\mathbf{m}} obtained as in Proposition 3.1 using the two gapped AA_{\infty} algebras 𝔪\mathfrak{m} and 𝔪~\tilde{\mathfrak{m}} respectively. By virtue of Lemma 2.6, there is an \mathfrak{C}-induced isomorphism ϕ=ϕ:Λ[[π1(L~)]]/𝔞~Λ[[π1(L)]]/𝔞\phi=\phi_{\mathfrak{C}}:\Lambda[[\pi_{1}(\tilde{L})]]/\tilde{\mathfrak{a}}\cong\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}. Similar to (11), we define

(13) 𝐂k:=𝐂k(F):=F1βTE(β)Yβk,β:(H(L)^Λ[[π1(L)]]/𝔞)kH(L~)^Λ[[π1(L~)]]/𝔞~\mathbf{C}_{k}:=\mathbf{C}^{(F)}_{k}:=F^{-1*}\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\mathfrak{C}_{k,\beta}:\big{(}H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}\big{)}^{\otimes k}\to H^{*}(\tilde{L})\hat{\otimes}\Lambda[[\pi_{1}(\tilde{L})]]/\tilde{\mathfrak{a}}

where all the formal power series coefficients in YY will be afterward transformed to those in Y~\tilde{Y} via the isomorphism ϕ=ϕ\phi=\phi_{\mathfrak{C}}. Note that 𝐂0=F1𝕱\mathbf{C}_{0}=F^{-1*}\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}, c.f. (8). Recall that the collections 𝐦\mathbf{m} and 𝐦~\tilde{\mathbf{m}} exclude 𝐦0\mathbf{m}_{0} and 𝐦~0\tilde{\mathbf{m}}_{0}. Similarly, we set 𝐂={𝐂k}k1\mathbf{C}=\{\mathbf{C}_{k}\}_{k\geq 1} excluding 𝐂0\mathbf{C}_{0} as well.

Proposition 3.4.

The 𝐂=𝐂(F)\mathbf{C}=\mathbf{C}^{(F)} gives a (non-curved) AA_{\infty} homomorphism from 𝐦\mathbf{m} to 𝐦~\tilde{\mathbf{m}}.

Proof.

The AA_{\infty} associativity equations tells (𝔪~F)k,β=({𝔪})k,β(\tilde{\mathfrak{m}}^{F}\diamond\mathfrak{C})_{k,\beta}=(\mathfrak{C}\{\mathfrak{m}\})_{k,\beta} (c.f. Convention 2.5). Hence,

(14) βTE(β)Yβ(𝔪~F)k,β(x1,,xk)=βTE(β)Yβ({𝔪})k,β(x1,,xk)\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\tilde{\mathfrak{m}}^{F}\diamond\mathfrak{C})_{k,\beta}(x_{1},\dots,x_{k})=\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\mathfrak{C}\{\mathfrak{m}\})_{k,\beta}(x_{1},\dots,x_{k})

for any fixed k1k\geq 1. On the left side, the terms involving 0,β\mathfrak{C}_{0,\beta} form the series 𝕱\boldsymbol{\mathfrak{F}}_{\mathfrak{C}} as in (8). Besides, since the components of 𝕱\boldsymbol{\mathfrak{F}}_{\mathfrak{C}} are contained in H1(L)H^{1}(L), the divisor axiom of 𝔪~F\tilde{\mathfrak{m}}^{F} can be applied after we use Lemma 2.1. Therefore, the left side of (14) becomes:

1,β,β1,,β0=j0<j1<<j1<j=kTE(β)Yβexpβ,𝕱𝔪~,βF(TE(β1)Yβ1j1j0,β1TE(β)Yβjj1,β)(x1,,xk)\displaystyle\sum_{\begin{subarray}{c}\ell\geq 1,\beta^{\prime},\beta_{1},\dots,\beta_{\ell}\\ 0=j_{0}<j_{1}<\dots<j_{\ell-1}<j_{\ell}=k\end{subarray}}T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\exp\langle\partial\beta^{\prime},\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}\rangle\ \tilde{\mathfrak{m}}^{F}_{\ell,\beta^{\prime}}\big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}\mathfrak{C}_{j_{1}-j_{0},\beta_{1}}\otimes\cdots\otimes T^{E(\beta_{\ell})}Y^{\partial\beta_{\ell}}\mathfrak{C}_{j_{\ell}-j_{\ell-1},\beta_{\ell}}\big{)}(x_{1},\dots,x_{k})

On the other hand, using Lemma 2.6 yields that:

TE(β)Yβexpβ,𝕱𝔪~,β=TE(β~)Y~β~𝔪~,β=𝐦~\sum T^{E(\beta)}Y^{\partial\beta}\exp\langle\partial\beta,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}\rangle\tilde{\mathfrak{m}}_{\ell,\beta}=\sum T^{E(\tilde{\beta})}\tilde{Y}^{\partial\tilde{\beta}}\tilde{\mathfrak{m}}_{\ell,\beta}=\tilde{\mathbf{m}}_{\ell}

(Be careful to distinguish YY from Y~\tilde{Y}.) In summary, the left side of (14) further becomes:

0=j0<j1<<j1<j=kF𝐦~(𝐂j1j0(F)𝐂jj1(F))(x1,,xk)\displaystyle\sum_{0=j_{0}<j_{1}<\dots<j_{\ell-1}<j_{\ell}=k}F^{*}\tilde{\mathbf{m}}_{\ell}(\mathbf{C}^{(F)}_{j_{1}-j_{0}}\otimes\cdots\otimes\mathbf{C}^{(F)}_{j_{\ell}-j_{\ell-1}})(x_{1},\dots,x_{k})

Now, in the right side of (14), we collect all those terms involving 𝔪0,β\mathfrak{m}_{0,\beta} first, and their sum is equal to

β,0λkTE(β)Yβk+1,β(x1#,,xλ#,𝐦0,xλ+1,,xk)\displaystyle\sum_{\beta,0\leq\lambda\leq k}T^{E(\beta)}Y^{\partial\beta}\mathfrak{C}_{k+1,\beta}\big{(}x^{\#}_{1},\dots,x^{\#}_{\lambda},\mathbf{m}_{0},x_{\lambda+1},\dots,x_{k}\big{)}

But, 𝐦0=W1+QW1(mod𝔞)\mathbf{m}_{0}=W\cdot\text{1}+Q\equiv W\cdot\text{1}\ (\mathrm{mod}\ {\mathfrak{a}}) and k1k\geq 1, and the above summation vanishes by the unitality of \mathfrak{C}. To conclude, what remain on the right side of (14) are precisely as follows:

λ,μ0,ν1λ+μ+ν=kF𝐂λ+μ+1(F)(x1#,,xλ#,𝐦ν(xλ+1,,xλ+ν),,xk)\sum_{\begin{subarray}{c}\lambda,\mu\geq 0,\nu\geq 1\\ \lambda+\mu+\nu=k\end{subarray}}F^{*}\mathbf{C}^{(F)}_{\lambda+\mu+1}\big{(}x_{1}^{\#},\dots,x_{\lambda}^{\#},\mathbf{m}_{\nu}(x_{\lambda+1},\dots,x_{\lambda+\nu}),\dots,x_{k}\big{)}

Applying F1F^{-1*} to the two equations above, we complete the proof. ∎

Remark 3.5

It is worth mentioning that the equation (14) in the above proof assumes k1k\geq 1. In contrast, the gluing analytic coordinate change maps developed in [Yua20] (i.e. Lemma 2.6) essentially uses the same equation as (14) but only for k=0k=0. Therefore, we really use more information about the category 𝒰𝒟\mathscr{UD} here. Compare also Remark 3.7.

3.1.2 Invariance

In particular, Proposition 3.4 tells that 𝐦~1𝐂1=𝐂1𝐦1\tilde{\mathbf{m}}_{1}\circ\mathbf{C}_{1}=\mathbf{C}_{1}\circ\mathbf{m}_{1} and 𝐦~2(𝐂1(x1),𝐂1(x2))=𝐂1(𝐦2(x1,x2))\tilde{\mathbf{m}}_{2}(\mathbf{C}_{1}(x_{1}),\mathbf{C}_{1}(x_{2}))=\mathbf{C}_{1}(\mathbf{m}_{2}(x_{1},x_{2})). Therefore, we know that 𝐂1:H(L)^Λ[[π1(L)]]/𝔞H(L~)^Λ[[π1(L~)]]/𝔞~\mathbf{C}_{1}:H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}\to H^{*}(\tilde{L})\hat{\otimes}\Lambda[[\pi_{1}(\tilde{L})]]/\tilde{\mathfrak{a}} is a cochain map and further induces a Λ\Lambda-algebra homomorphism

(15) η:=[𝐂1]:HF(L,𝔪)HF(L~,𝔪~)\eta_{\mathfrak{C}}:=[\mathbf{C}_{1}]:\operatorname{HF}(L,\mathfrak{m})\to\operatorname{HF}(\tilde{L},\tilde{\mathfrak{m}})
Proposition 3.6.

The η\eta_{\mathfrak{C}} only relies on the ud-homotopy class of \mathfrak{C} in 𝒰𝒟\mathscr{UD}.

Proof.

We basically use the ideas in [Yua20]. Suppose ,\mathfrak{C},\mathfrak{C}^{\prime} are two AA_{\infty} homotopy equivalences from (H(L),𝔪)(H^{*}(L),\mathfrak{m}) to (H(L~),𝔪~)(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) up to the Fukaya’s trick in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} about some isotopy FF (Convention 2.5). Suppose also that they are ud-homotopic to each other. Our task is to prove η=η\eta_{\mathfrak{C}}=\eta_{\mathfrak{C}^{\prime}}.

By definition, there exist operator systems 𝔣s={(𝔣s)k,β}\mathfrak{f}_{s}=\{(\mathfrak{f}_{s})_{k,\beta}\} and 𝔥s={(𝔥s)k,β}\mathfrak{h}_{s}=\{(\mathfrak{h}_{s})_{k,\beta}\} for 0s10\leq s\leq 1 such that the conditions (a) (b) (c) (d) in §2.1.6 hold, where 𝔣0=\mathfrak{f}_{0}=\mathfrak{C} and 𝔣1=\mathfrak{f}_{1}=\mathfrak{C}^{\prime}. By the condition (b), we get

βTE(β)Yβdds(𝔣s)1,β(𝒙)\displaystyle\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\frac{d}{ds}(\mathfrak{f}_{s})_{1,\beta}(\boldsymbol{x}) =β1,β2TE(β1)Yβ1(𝔥s)1,β1(TE(β2)Yβ2𝔪1,β2(𝒙))\displaystyle=\sum_{\beta_{1},\beta_{2}}T^{E(\beta_{1})}Y^{\partial\beta_{1}}(\mathfrak{h}_{s})_{1,\beta_{1}}(T^{E(\beta_{2})}Y^{\partial\beta_{2}}\mathfrak{m}_{1,\beta_{2}}(\boldsymbol{x}))
+β1,β2TE(β1)Yβ1(𝔥s)2,β1(𝒙#,TE(β2)Yβ2𝔪0,β2)\displaystyle+\sum_{\beta_{1},\beta_{2}}T^{E(\beta_{1})}Y^{\partial\beta_{1}}(\mathfrak{h}_{s})_{2,\beta_{1}}\big{(}\boldsymbol{x}^{\#},T^{E(\beta_{2})}Y^{\partial\beta_{2}}\mathfrak{m}_{0,\beta_{2}}\big{)}
(16) +β1,β2TE(β1)Yβ1(𝔥s)2,β1(TE(β2)Yβ2𝔪0,β2,𝒙)\displaystyle+\sum_{\beta_{1},\beta_{2}}T^{E(\beta_{1})}Y^{\partial\beta_{1}}(\mathfrak{h}_{s})_{2,\beta_{1}}\big{(}T^{E(\beta_{2})}Y^{\partial\beta_{2}}\mathfrak{m}_{0,\beta_{2}},\boldsymbol{x}\big{)}
+βTE(β)Yβ1i,β0++β+β=βCU[𝔪~F],β0((𝔥s)0,β;(𝔣s)0,β1,,(𝔣s)1,βi(𝒙),,(𝔣s)0,β)\displaystyle+\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\sum_{1\leq i\leq\ell,\ \beta_{0}+\cdots+\beta_{\ell}+\beta^{\prime}=\beta}\operatorname{CU}[\tilde{\mathfrak{m}}^{F}]_{\ell,\beta_{0}}\big{(}(\mathfrak{h}_{s})_{0,\beta^{\prime}};(\mathfrak{f}_{s})_{0,\beta_{1}},\dots,(\mathfrak{f}_{s})_{1,\beta_{i}}(\boldsymbol{x}),\dots,(\mathfrak{f}_{s})_{0,\beta_{\ell}}\big{)}
+βTE(β)Yβ1i,β0++β=β𝔪~,β0F((𝔣s)0,β1,,(𝔥s)1,βi(𝒙),,(𝔣s)0,β)\displaystyle+\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\sum_{1\leq i\leq\ell,\ \beta_{0}+\cdots+\beta_{\ell}=\beta}\tilde{\mathfrak{m}}^{F}_{\ell,\beta_{0}}\Big{(}(\mathfrak{f}_{s})_{0,\beta_{1}},\dots,(\mathfrak{h}_{s})_{1,\beta_{i}}(\boldsymbol{x}),\dots,(\mathfrak{f}_{s})_{0,\beta_{\ell}}\Big{)}

The second and third sums are all zero modulo 𝔞{\mathfrak{a}} due to the condition (c) and the decomposition (7). The fourth sum vanishes because the semipositive condition ensures deg(𝔥s)0,β=μ(β)=0\deg(\mathfrak{h}_{s})_{0,\beta^{\prime}}=-\mu(\beta^{\prime})=0 and we can use the cyclical unitality. For the fifth sum, the semipositive condition also tells deg(𝔣s)0,βj=1μ(βj)=1\deg(\mathfrak{f}_{s})_{0,\beta_{j}}=1-\mu(\beta_{j})=1; by the divisor axiom and by Lemma 2.1, we conclude that the fifth sum is equal to

β0,β1TE(β0)Yβ0expβ0,𝕱𝔣s𝔪~1,β0F(TE(β1)Yβ1(𝔥s)1,β1(𝒙))\sum_{\beta_{0},\beta_{1}}T^{E(\beta_{0})}Y^{\partial\beta_{0}}\exp\langle\partial\beta_{0},\boldsymbol{\mathfrak{F}}_{\mathfrak{f}_{s}}\rangle\ \tilde{\mathfrak{m}}^{F}_{1,\beta_{0}}\big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}(\mathfrak{h}_{s})_{1,\beta_{1}}(\boldsymbol{x})\big{)}

Then, since 𝔣sudud\mathfrak{f}_{s}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathfrak{C}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathfrak{C}^{\prime}, it follows from Lemma 2.4 that 𝕱𝔣s𝕱=𝕱\boldsymbol{\mathfrak{F}}_{\mathfrak{f}_{s}}\equiv\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}=\boldsymbol{\mathfrak{F}}_{\mathfrak{C}^{\prime}} (mod 𝔞{\mathfrak{a}}), where we recall the notation in (8). Hence, by Lemma 2.6, after we transform into Y~\tilde{Y}-coefficients, this fifth sum is

Fβ0,β1TE(β0)Y~β0𝔪~1,β0(TE(β1)Yβ1F1(𝔥s)1,β1(𝒙))F^{*}\sum_{\beta_{0},\beta_{1}}T^{E(\beta_{0})}\tilde{Y}^{\partial\beta_{0}}\tilde{\mathfrak{m}}_{1,\beta_{0}}\big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}F^{-1*}(\mathfrak{h}_{s})_{1,\beta_{1}}(\boldsymbol{x})\big{)}

To conclude, only the first and fifth sum in (16) survive. Now, we set ξ=βTE(β)YβF101(𝔥s)1,β\xi=\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\textstyle F^{-1*}\int_{0}^{1}(\mathfrak{h}_{s})_{1,\beta}. Taking the integration from s=0s=0 to s=1s=1 and applying F1F^{-1*} to the both sides of the initial equation (16), we exactly obtain that 𝐂1(𝒙)𝐂1(𝒙)=ξ𝐦1+𝐦~1ξ{\mathbf{C}}^{\prime}_{1}(\boldsymbol{x})-\mathbf{C}_{1}(\boldsymbol{x})=\xi\circ\mathbf{m}_{1}+\tilde{\mathbf{m}}_{1}\circ\xi. In other words, the ξ\xi gives a cochain homotopy between the two cochain maps 𝐂1\mathbf{C}_{1} and 𝐂1\mathbf{C}^{\prime}_{1}. Particularly, the induced morphisms on the cohomology are the same η=η\eta_{\mathfrak{C}}=\eta_{\mathfrak{C}^{\prime}}. The proof is now complete. ∎

Remark 3.7

Similar to Remark 3.5, we indicate that the equation (16) without the input 𝒙\boldsymbol{x} is used in [Yua20] to prove that the analytic coordinate change maps are well-defined. In contrast, adding the input 𝒙\boldsymbol{x} here, we have extracted more information from the category 𝒰𝒟\mathscr{UD} than [Yua20].

3.1.3 Composition

Now, we consider three adjacent Lagrangian submanifolds L,L~L,\tilde{L}, and L~~\tilde{\tilde{L}}. Let FF and FF^{\prime} be small isotopies such that F(L)=L~F(L)=\tilde{L} and F(L~)=L~~F^{\prime}(\tilde{L})=\tilde{\tilde{L}}. Suppose :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) and :(H(L~),𝔪~)(H(L~~),𝔪~~)\mathfrak{C}^{\prime}:(H^{*}(\tilde{L}),\tilde{\mathfrak{m}})\to(H^{*}(\tilde{\tilde{L}}),\tilde{\tilde{\mathfrak{m}}}) are AA_{\infty} homotopy equivalences in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} up to Fukaya’s tricks. By Proposition 3.4, they give rise to two non-curved AA_{\infty} homomorphisms 𝐂={𝐂k}k1\mathbf{C}=\{\mathbf{C}_{k}\}_{k\geq 1} and 𝐂={𝐂k}k1\mathbf{C}^{\prime}=\{\mathbf{C}^{\prime}_{k}\}_{k\geq 1}. By Proposition 3.4 again, the composition gapped AA_{\infty} morphism \mathfrak{C}^{\prime}\diamond\mathfrak{C} in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} also induces a non-curved AA_{\infty} homomorphism, denoted by 𝐂𝐂=((𝐂𝐂)k)k1\mathbf{C}^{\prime}\diamond\mathbf{C}=((\mathbf{C}^{\prime}\diamond\mathbf{C})_{k})_{k\geq 1}. We call it the composition of 𝐂\mathbf{C}^{\prime} and 𝐂\mathbf{C}. The notation and term we use here can be justified as follows:

Proposition 3.8.

The non-curved AA_{\infty} homomorphism 𝐂𝐂=((𝐂𝐂)k)k1\mathbf{C}^{\prime}\diamond\mathbf{C}=((\mathbf{C}^{\prime}\diamond\mathbf{C})_{k})_{k\geq 1} satisfies that

(𝐂𝐂)k=1, 0=j0<j1<<j1<j=k𝐂(𝐂j1j0,,𝐂jj1)(\mathbf{C}^{\prime}\diamond\mathbf{C})_{k}=\sum_{\ell\geq 1,\ 0=j_{0}<j_{1}<\dots<j_{\ell-1}<j_{\ell}=k}\mathbf{C}^{\prime}_{\ell}\big{(}\mathbf{C}_{j_{1}-j_{0}},\dots,\mathbf{C}_{j_{\ell}-j_{\ell-1}}\big{)}
Proof.

Fix k1k\geq 1. By definition, we first know (𝐂𝐂)k=βTE(β)Yβ()k,β(\mathbf{C}^{\prime}\diamond\mathbf{C})_{k}=\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}(\mathfrak{C}^{\prime}\diamond\mathfrak{C})_{k,\beta}. Expand each ()k,β(\mathfrak{C}^{\prime}\diamond\mathfrak{C})_{k,\beta} by (6), and all the terms involving 0,β\mathfrak{C}_{0,\beta} form the series 𝕱\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}. Then, as in the proof of Proposition 3.4, the divisor axiom can be applied after we use Lemma 2.1, and we finally obtain

(𝐂𝐂)k=1,β,β1,,β0=j0<j1<<j1<j=kTE(β)Yβexpβ,𝕱,β(TE(β1)Yβ1j1j0,β1TE(β)Yβjj1,β)\displaystyle\scalebox{0.9}{$(\mathbf{C}^{\prime}\diamond\mathbf{C})_{k}$}=\sum_{\begin{subarray}{c}\ell\geq 1,\beta^{\prime},\beta_{1},\dots,\beta_{\ell}\\ 0=j_{0}<j_{1}<\dots<j_{\ell-1}<j_{\ell}=k\end{subarray}}\scalebox{0.9}{$T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\exp\langle\partial\beta^{\prime},\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}\rangle\mathfrak{C}^{\prime}_{\ell,\beta^{\prime}}\big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}\mathfrak{C}_{j_{1}-j_{0},\beta_{1}}\otimes\cdots\otimes T^{E(\beta_{\ell})}Y^{\partial\beta_{\ell}}\mathfrak{C}_{j_{\ell}-j_{\ell-1},\beta_{\ell}}\big{)}$}

Recall that Y~β=TE(β)Yβexpβ,𝕱\tilde{Y}^{\partial\beta^{\prime}}=T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\exp\langle\partial\beta^{\prime},\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}\rangle by Lemma 2.6. The proof is now complete. ∎

In particular, the above proposition infers that (𝐂𝐂)1=𝐂1𝐂1(\mathbf{C}^{\prime}\diamond\mathbf{C})_{1}=\mathbf{C}^{\prime}_{1}\circ\mathbf{C}_{1} is also a cochain map; hence,

(17) η=ηη:HF(L,𝔪)HF(L~,𝔪~)HF(L~~,𝔪~~)\eta_{\mathfrak{C}^{\prime}\diamond\mathfrak{C}}=\eta_{\mathfrak{C}^{\prime}}\circ\eta_{\mathfrak{C}}:\operatorname{HF}(L,\mathfrak{m})\to\operatorname{HF}(\tilde{L},\tilde{\mathfrak{m}})\to\operatorname{HF}(\tilde{\tilde{L}},\tilde{\tilde{\mathfrak{m}}})
Corollary 3.9.

In the context of Proposition 3.6, the η\eta_{\mathfrak{C}} is an isomorphism.

Proof.

The \mathfrak{C} admits a ud-homotopy inverse 1\mathfrak{C}^{-1} so that 1udid\mathfrak{C}^{-1}\diamond\mathfrak{C}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id} and 1udid\mathfrak{C}\diamond\mathfrak{C}^{-1}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id} (up to Fukaya’s tricks). By Proposition 3.6 and (17), we get η1η=η1=ηid=id\eta_{\mathfrak{C}^{-1}}\circ\eta_{\mathfrak{C}}=\eta_{\mathfrak{C}^{-1}\diamond\mathfrak{C}}=\eta_{\mathrm{id}}=\mathrm{id} and vice versa. ∎

3.2 Evaluation of self Floer cohomology

3.2.1 Background and review

In the literature, the self Floer cohomology is defined by a different way. Here let’s assume π1(L)\pi_{1}(L) has no torsion part. Instead of H(L)^Λ[[π1(L)]]/𝔞H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}} in Proposition 3.1, we previously work with CF(L;Λ):=H(L)^Λ\operatorname{CF}(L;\Lambda):=H^{*}(L)\hat{\otimes}\Lambda and consider a weak bounding cochain bHodd(L;Λ+)b\in H^{\mathrm{odd}}(L;\Lambda_{+}), i.e. k0βTE(β)𝔪k,β(b,,b)Λ+1\textstyle\sum_{k\geq 0}\sum_{\beta}T^{E(\beta)}\mathfrak{m}_{k,\beta}(b,\dots,b)\in\Lambda_{+}\cdot\text{1} for the constant-one function 1H0(L)\text{1}\in H^{0}(L). Then, we define

𝐦kb(x1,,xk)=TE(β)𝔪,β(b,,b,x1,b,,b,xk,b,,b)\mathbf{m}_{k}^{b}(x_{1},\dots,x_{k})=\sum T^{E(\beta)}\mathfrak{m}_{\bullet,\beta}(b,\dots,b,x_{1},b,\dots\ \dots,b,x_{k},b,\dots,b)

By condition, the 𝐦0bΛ+1\mathbf{m}_{0}^{b}\in\Lambda_{+}\cdot\text{1} will be eliminated. It is known in the literature that (see e.g. [FOOO10b, Proposition 3.6.10]): The collection {𝐦kb}k1\{\mathbf{m}_{k}^{b}\}_{k\geq 1} forms a (non-curved) AA_{\infty} algebra on H(L)^ΛH^{*}(L)\hat{\otimes}\Lambda. Then, the self Floer cohomology is usually defined by HF(L,b;Λ):=H(CF(L;Λ);𝐦1b)\operatorname{HF}(L,b;\Lambda):=H^{*}(\operatorname{CF}(L;\Lambda);\mathbf{m}_{1}^{b}).

3.2.2 Evaluation at a weak bounding cochain

For the family Floer theory and SYZ picture, we should mainly focus on those degree-one weak bounding cochains bb in H1(L;Λ+)H^{1}(L;\Lambda_{+}). Indeed, if LL is a Lagrangian torus fiber in the SYZ framework, its dual torus fiber is expected to be H1(L;U(1))H^{1}(L;U(1)), at least at the topological level. Now, we want to make connections with HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) in (12). In fact, since degb=1\deg b=1, the divisor axiom of 𝔪\mathfrak{m} implies that

𝐦kb(x1,,xk)=TE(β)eβb𝔪k,β(x1,,xk)\mathbf{m}_{k}^{b}(x_{1},\dots,x_{k})=\sum T^{E(\beta)}e^{\partial\beta\cap b}\mathfrak{m}_{k,\beta}(x_{1},\dots,x_{k})

and we can even actually require bH1(L;Λ0)b\in H^{1}(L;\Lambda_{0}) instead of H1(L;Λ+)H^{1}(L;\Lambda_{+}) without violating the convergence. There is a quotient map 𝗊:H1(L;Λ0)H1(L;UΛ)\mathsf{q}:H^{1}(L;\Lambda_{0})\to H^{1}(L;U_{\Lambda}) induced by exp:Λ0UΛ\exp:\Lambda_{0}\to U_{\Lambda}. For a basis of H1(L)H^{1}(L), we write b=(x1,,xm)b=(x_{1},\dots,x_{m}) for xiΛ0x_{i}\in\Lambda_{0}; the image point 𝐲=𝗊(b)\mathbf{y}=\mathsf{q}(b) is given by (y1,,ym)(y_{1},\dots,y_{m}) for yi=exiUΛy_{i}=e^{x_{i}}\in U_{\Lambda}. In general, let 𝐲α\mathbf{y}^{\alpha} be the image of (α,𝐲)(\alpha,\mathbf{y}) under the natural pairing π1(L)×H1(L;Λ)Λ\pi_{1}(L)\times H^{1}(L;\Lambda^{*})\to\Lambda^{*}. We also pick a basis so that π1(L)m\pi_{1}(L)\cong\mathbb{Z}^{m}. For α=(α1,,αm)π1(L)\alpha=(\alpha_{1},\dots,\alpha_{m})\in\pi_{1}(L) and 𝐲=(y1,,ym)H1(L;Λ)\mathbf{y}=(y_{1},\dots,y_{m})\in H^{1}(L;\Lambda^{*}), we have 𝐲α=y1α1ymαm\mathbf{y}^{\alpha}=y_{1}^{\alpha_{1}}\cdots y_{m}^{\alpha_{m}}. Now, we can introduce a slight variation of the above 𝐦kb\mathbf{m}_{k}^{b} as follows:

(18) 𝐦k𝐲(x1,,xk)=TE(β)𝐲β𝔪k,β(x1,,xk)\mathbf{m}_{k}^{\mathbf{y}}(x_{1},\dots,x_{k})=\sum T^{E(\beta)}\mathbf{y}^{\partial\beta}\mathfrak{m}_{k,\beta}(x_{1},\dots,x_{k})

Note that 𝐦0𝐲W(𝐲)1+Q(𝐲)\mathbf{m}^{\mathbf{y}}_{0}\equiv W(\mathbf{y})\cdot\text{1}+Q(\mathbf{y}) (7). By the divisor axiom and under the semipositive condition, Q(𝐲)=0Q(\mathbf{y})=0 if and only if the 𝐲\mathbf{y} admits a lift bb which is a weak bounding cochain. Hence, by the transformation yi=exiy_{i}=e^{x_{i}}, it simply goes back to the previous 𝐦kb\mathbf{m}_{k}^{b}. Nevertheless, the advantage of (18) is that we can further allow 𝐲\mathbf{y} to run over a small neighborhood of H1(L;UΛ)UΛnH^{1}(L;U_{\Lambda})\cong U_{\Lambda}^{n} in H1(L;Λ)(Λ)nH^{1}(L;\Lambda^{*})\cong(\Lambda^{*})^{n}.

Proposition 3.10.

If Q(𝐲)=0Q(\mathbf{y})=0, then 𝐦𝐲={𝐦k𝐲}k1\mathbf{m}^{\mathbf{y}}=\{\mathbf{m}_{k}^{\mathbf{y}}\}_{k\geq 1} forms a non-curved AA_{\infty} algebra on H(L)^ΛH^{*}(L)\hat{\otimes}\Lambda.

Proof.

We may argue just as Proposition 3.1 by the AA_{\infty} associativity and the unitality of 𝔪\mathfrak{m}. ∎

Particularly, 𝐦1𝐲𝐦1𝐲=0\mathbf{m}_{1}^{\mathbf{y}}\circ\mathbf{m}_{1}^{\mathbf{y}}=0, so we can define the self Floer cohomology of LL at 𝐲\mathbf{y}:

(19) HF(L,𝔪,𝐲):=H(H(L)^Λ,𝐦1𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}):=H^{*}(H^{*}(L)\hat{\otimes}\Lambda,\mathbf{m}^{\mathbf{y}}_{1})

As before, we can also show that [x1][x2]=[𝐦2𝐲(x1#,x2)][x_{1}]\cdot[x_{2}]=[\mathbf{m}^{\mathbf{y}}_{2}(-x_{1}^{\#},x_{2})] defines a ring structure on HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) with a unit [1][\text{1}]. Additionally, we can consider the evaluation map at 𝐲\mathbf{y}:

(20) 𝐲:H(L)^Λ[[π1(L)]]H(L)^ΛYαx𝐲αx\mathcal{E}_{\mathbf{y}}:H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]\to H^{*}(L)\hat{\otimes}\Lambda\qquad Y^{\alpha}\cdot x\mapsto\mathbf{y}^{\alpha}\cdot x

where απ1(L)\alpha\in\pi_{1}(L) and xH(L)x\in H^{*}(L). For the sake of convergence, we should replace Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] by some polyhedral affinoid algebra in it, and we also require each coefficient 𝐲\mathbf{y} is contained in the corresponding polyhedral affinoid domain. But, let’s make this point implicit for clarity. Here we also assume 𝐲\mathbf{y} admits a weak bounding cochain lift bb up to Fukaya’s trick, namely, Q(𝐲)=0Q(\mathbf{y})=0. Thus, the above 𝐲\mathcal{E}_{\mathbf{y}} also descends to a quotient map modulo 𝔞{\mathfrak{a}}, which is still denoted by 𝐲:H(L)^Λ[[π1(L)]]/𝔞H(L)^Λ\mathcal{E}_{\mathbf{y}}:H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}\to H^{*}(L)\hat{\otimes}\Lambda. Further, we can extend the definition of 𝐲\mathcal{E}_{\mathbf{y}} by setting (𝐲)1=𝐲(\mathcal{E}_{\mathbf{y}})_{1}=\mathcal{E}_{\mathbf{y}}, (𝐲)k=0(\mathcal{E}_{\mathbf{y}})_{k}=0, k1k\neq 1, and we can directly check the following:

Lemma 3.11.

The 𝐲\mathcal{E}_{\mathbf{y}} gives a non-curved AA_{\infty} homomorphism from 𝐦\mathbf{m} to 𝐦𝐲\mathbf{m}^{\mathbf{y}}. Moreover, the induced map 𝐲=[𝐲]:HF(L,𝔪)HF(L,𝔪,𝐲)\mathcal{E}_{\mathbf{y}}=[\mathcal{E}_{\mathbf{y}}]:\operatorname{HF}(L,\mathfrak{m})\to\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) is a unital Λ\Lambda-algebra homomorphism so that 𝐲([1])=[1]\mathcal{E}_{\mathbf{y}}([\text{1}])=[\text{1}].

Suppose :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) is an AA_{\infty} homotopy equivalence up to the Fukaya’s trick for a small isotopy FF with F(L)=L~F(L)=\tilde{L} (Convention 2.5). For the analytic map ϕ\phi in (10), the point 𝐲~=ϕ(𝐲)\tilde{\mathbf{y}}=\phi(\mathbf{y}) also admits a weak bounding cochain lift b~\tilde{b}, i.e. Q~(𝐲~)=0\tilde{Q}(\tilde{\mathbf{y}})=0. Then, we get two non-curved AA_{\infty} algebras {𝔪k𝐲}k1\{\mathfrak{m}_{k}^{\mathbf{y}}\}_{k\geq 1} and {𝔪~k𝐲~}k1\{\tilde{\mathfrak{m}}_{k}^{\tilde{\mathbf{y}}}\}_{k\geq 1} on H(L)^ΛH^{*}(L)\hat{\otimes}\Lambda and H(L~)^ΛH^{*}(\tilde{L})\hat{\otimes}\Lambda by Proposition 3.10. Just as (13), we define

(21) 𝐂k𝐲:=F1TE(β)𝐲βk,β:(H(L)^Λ)kH(L~)^Λ\mathbf{C}_{k}^{\mathbf{y}}:=F^{-1}\sum T^{E(\beta)}\mathbf{y}^{\partial\beta}\mathfrak{C}_{k,\beta}:\big{(}H^{*}(L)\hat{\otimes}\Lambda\big{)}^{k}\to H^{*}(\tilde{L})\hat{\otimes}\Lambda
Proposition 3.12.

The collection 𝐂𝐲={𝐂k𝐲}k1\mathbf{C}^{\mathbf{y}}=\{\mathbf{C}_{k}^{\mathbf{y}}\}_{k\geq 1} forms a (non-curved) AA_{\infty} homomorphism from 𝐦𝐲\mathbf{m}^{\mathbf{y}} to 𝐦~𝐲~\tilde{\mathbf{m}}^{\tilde{\mathbf{y}}} so that 𝐲~𝐂=𝐂𝐲𝐲\mathcal{E}_{\tilde{\mathbf{y}}}\diamond\mathbf{C}=\mathbf{C}^{\mathbf{y}}\diamond\mathcal{E}_{\mathbf{y}}. In particular, we have 𝐦~1𝐲~𝐂1𝐲=𝐂1𝐲𝐦1𝐲\tilde{\mathbf{m}}_{1}^{\tilde{\mathbf{y}}}\circ\mathbf{C}_{1}^{\mathbf{y}}=\mathbf{C}_{1}^{\mathbf{y}}\circ\mathbf{m}_{1}^{\mathbf{y}} and 𝐦~2𝐲~(𝐂1𝐲(x1),𝐂1𝐲(x2))=𝐂1𝐲(𝐦2𝐲(x1,x2))\tilde{\mathbf{m}}_{2}^{\tilde{\mathbf{y}}}(\mathbf{C}_{1}^{\mathbf{y}}(x_{1}),\mathbf{C}_{1}^{\mathbf{y}}(x_{2}))=\mathbf{C}_{1}^{\mathbf{y}}(\mathbf{m}_{2}^{\mathbf{y}}(x_{1},x_{2})).

Proof.

The proof is just the same as that of Proposition 3.4. ∎

Proposition 3.13.

The 𝐂1𝐲\mathbf{C}_{1}^{\mathbf{y}} similarly induces a Λ\Lambda-algebra homomorphism η𝐲=[𝐂1𝐲]:HF(L,𝔪,𝐲)HF(L,𝔪~,𝐲~)\eta_{\mathfrak{C}}^{\mathbf{y}}=[\mathbf{C}_{1}^{\mathbf{y}}]:\operatorname{HF}(L,\mathfrak{m},\mathbf{y})\to\operatorname{HF}(L,\tilde{\mathfrak{m}},\tilde{\mathbf{y}}) which only relies on the ud-homotopy class of \mathfrak{C} in 𝒰𝒟\mathscr{UD}. In particular, if \mathfrak{C} admits a ud-homotopy inverse, then η𝐲\eta^{\mathbf{y}} is an isomorphism.

Proof.

The proof is just the same as that of Proposition 3.6 and Corollary 3.9. ∎

The proposition tells that HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) is invariant under the various analytic coordinate change ϕ\phi (Lemma 2.6). From the viewpoint of Theorem 1.5, we should view 𝐲\mathbf{y} and 𝐲~=ϕ(𝐲)\tilde{\mathbf{y}}=\phi(\mathbf{y}) as the same point in the mirror analytic space XX^{\vee} but referring to different local charts.

3.3 Critical points of the superpotential

Take a formal power series F=j=1cjYαjF=\sum_{j=1}^{\infty}c_{j}Y^{\alpha_{j}} in Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] (where cjΛc_{j}\in\Lambda and αjπ1(L)\alpha_{j}\in\pi_{1}(L)). Given θH1(L)Hom(π1(L),)\theta\in H^{1}(L)\cong\operatorname{Hom}(\pi_{1}(L),\mathbb{R}), we define the logarithmic derivative along θ\theta of FF by

(22) DθF=j=1cjαj,θYαj\textstyle D_{\theta}F=\sum_{j=1}^{\infty}c_{j}\langle\alpha_{j},\theta\rangle Y^{\alpha_{j}}

One can easily check the following properties: (a) Dθ(FG)=FDθG+GDθFD_{\theta}(F\cdot G)=F\cdot D_{\theta}G+G\cdot D_{\theta}F; (b) Dθ(exp(F))=exp(F)DθFD_{\theta}(\exp(F))=\exp(F)\cdot D_{\theta}F. Given an ideal \mathcal{I} in Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]], we denote by

(23) D=the ideal generated by{DθFθH1(L;),F}D\mathcal{I}=\ \text{the ideal generated by}\ \{D_{\theta}F\mid\forall\ \theta\in H^{1}(L;\mathbb{Z}),\,\forall\ F\in\mathcal{I}\}

We note that D=0D\mathcal{I}=0, whenever =0\mathcal{I}=0.

Suppose we have an AA_{\infty} homotopy equivalence :(H(L),𝔪)(H(L~),𝔪~)\mathfrak{C}:(H^{*}(L),\mathfrak{m})\to(H^{*}(\tilde{L}),\tilde{\mathfrak{m}}) in Mor𝒰𝒟\operatorname{Mor}\mathscr{UD} up to the Fukaya’s trick about a small isotopy FF such that F(L)=L~F(L)=\tilde{L} (Convention 2.5). Recall that by Lemma 2.6, the analytic coordinate change map

ϕ:Λ[[π1(L~)]]/𝔞~Λ[[π1(L)]]/𝔞,Y~αTα,λYαexpα,𝕱(Y)\phi:\Lambda[[\pi_{1}(\tilde{L})]]/\tilde{\mathfrak{a}}\to\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}},\qquad\tilde{Y}^{\alpha}\mapsto T^{\langle\alpha,\lambda\rangle}Y^{\alpha}\exp\langle\alpha,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle

is an isomorphism such that ϕ(W~)=W\phi(\tilde{W})=W. Recall that W,Q,𝔞W,Q,{\mathfrak{a}} and W~,Q~,𝔞~\tilde{W},\tilde{Q},\tilde{\mathfrak{a}} are given as in (7) for 𝔪\mathfrak{m} and 𝔪~\tilde{\mathfrak{m}} respectively. Regard WW or W~\tilde{W} as a principal ideal, we can define DWDW or DW~D\tilde{W} just like (23).

Proposition 3.14.

The analytic coordinate change map ϕ\phi match the ideal DW~D\tilde{W} with DWDW.

Proof.

Suppose θH1(L)\theta\in H^{1}(L) and θ~H1(L~)\tilde{\theta}\in H^{1}(\tilde{L}) are FF-related. First, we compute the commutator of ϕ\phi and DD applied to a monomial Y~α\tilde{Y}^{\alpha}:

Dθ(ϕ(Y~α))ϕ(Dθ~Y~α)=Tα,λYαexpα,𝕱(Y)Dθα,𝕱(Y)D_{\theta}(\phi(\tilde{Y}^{\alpha}))-\phi(D_{\tilde{\theta}}\tilde{Y}^{\alpha})=T^{\langle\alpha,\lambda\rangle}Y^{\alpha}\exp\langle\alpha,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle D_{\theta}\langle\alpha,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle

Note that 𝕱(Y)=TE(γ)Yγ0,γ\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)=\sum T^{E(\gamma)}Y^{\partial\gamma}\mathfrak{C}_{0,\gamma} is contained in Λ[[π1(L)]]^H1(L)\Lambda[[\pi_{1}(L)]]\hat{\otimes}H^{1}(L) by (8). Next, we compute

DθWϕ(Dθ~W~)\displaystyle D_{\theta}W-\phi(D_{\tilde{\theta}}\tilde{W}) =Dθ(ϕ(W~))ϕ(Dθ~W~)\displaystyle=D_{\theta}(\phi(\tilde{W}))-\phi(D_{\tilde{\theta}}\tilde{W})
=μ(β)=2TE(β)Yβexpβ,𝕱(Y)Dθβ,𝕱(Y)𝔪~0,β\displaystyle=\sum_{\mu(\beta)=2}T^{E(\beta)}Y^{\partial\beta}\exp\langle\partial\beta,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle D_{\theta}\langle\partial\beta,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle\cdot\tilde{\mathfrak{m}}_{0,\beta}
=μ(β)=2TE(β)Yβexpβ,𝕱(Y)(γ0,μ(γ)=0TE(γ)γ,θYγβ,0,γ)𝔪~0,β\displaystyle=\sum_{\mu(\beta)=2}T^{E(\beta)}Y^{\partial\beta}\exp\langle\partial\beta,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle\Big{(}\sum_{\gamma\neq 0,\mu(\gamma)=0}T^{E(\gamma)}\langle\partial\gamma,\theta\rangle Y^{\partial\gamma}\langle\partial\beta,\mathfrak{C}_{0,\gamma}\rangle\Big{)}\tilde{\mathfrak{m}}_{0,\beta}
=γ0,μ(γ)=0TE(γ)γ,θYγμ(β)=2β,0,γTE(β)Yβexpβ,𝕱(Y)𝔪~0,β\displaystyle=\sum_{\gamma\neq 0,\mu(\gamma)=0}T^{E(\gamma)}\langle\partial\gamma,\theta\rangle Y^{\partial\gamma}\sum_{\mu(\beta)=2}\langle\partial\beta,\mathfrak{C}_{0,\gamma}\rangle\cdot T^{E(\beta)}Y^{\partial\beta}\exp\langle\partial\beta,\boldsymbol{\mathfrak{F}}_{\mathfrak{C}}(Y)\rangle\cdot\tilde{\mathfrak{m}}_{0,\beta}

For a fixed γ\gamma, we observe that the second sum over β\beta can be further simplified as follows:

Thus, DθWD_{\theta}W is contained in the (extension) image ideal ϕ(DW~)\phi(D\tilde{W}), so DWϕ(DW~)DW\subset\phi(D\tilde{W}). By Lemma 2.6, the inverse ϕ1\phi^{-1} can be defined by the same pattern as (9) using a ud-homotopy inverse 1\mathfrak{C}^{-1} of \mathfrak{C}. Hence, we can similarly obtain that DW~ϕ1(DW)D\tilde{W}\subset\phi^{-1}(DW). Putting them together, we conclude ϕ(DW~)=DW\phi(D\tilde{W})=DW. ∎

Definition 3.15.

We call 𝐲H1(L;Λ)\mathbf{y}\in H^{1}(L;\Lambda^{*}) a critical point of WW if 𝐲\mathbf{y} is contained in the convergence domain333When we substitute 𝐲\mathbf{y} into WW, the result W(𝐲)W(\mathbf{y}) is a series in the field Λ\Lambda. We require that it converges with respect to the non-Archimedean norm; see [Bos14, Page 10, Lemma 3]. and DθW(𝐲)=0D_{\theta}W(\mathbf{y})=0 for any θH1(L)\theta\in H^{1}(L).

Corollary 3.16.

𝐲\mathbf{y} is a critical point of WW if and only if 𝐲~:=ϕ(𝐲)\tilde{\mathbf{y}}:=\phi(\mathbf{y}) is a critical point of W~\tilde{W}. Thus, in Theorem 1.5, the critical points of WW^{\vee} are well-defined points in the mirror analytic space XX^{\vee}.

3.4 Non-vanishing of self Floer cohomology

For the lemma below, the basic ideas already exist in the literature, e.g. [FOOO10a, Lemma 13.1], [FOOO16, Lemma 2.4.20], [FOOO12, Theorem 2.3]. But, we want some slight modifications in our settings. Let (H(L),𝔪)Obj𝒰𝒟(H^{*}(L),\mathfrak{m})\in\operatorname{Obj}\mathscr{UD} be a minimal model AA_{\infty} algebra obtained by the moduli space of holomorphic disks and by the homological perturbation. Recall that 𝔪1,0=0\mathfrak{m}_{1,0}=0 and 𝔪2,0\mathfrak{m}_{2,0} agrees with the wedge product up to sign.

Define the 𝐦k\mathbf{m}_{k} from 𝔪=(𝔪k,β)\mathfrak{m}=(\mathfrak{m}_{k,\beta}) as in (11).

Lemma 3.17.

Suppose the de Rham cohomology ring H(L)H^{*}(L) is generated by H1(L;)H^{1}(L;\mathbb{Z}). Let {θ1,,θn}\{\theta_{1},\dots,\theta_{n}\} in H1(L;)H^{1}(L;\mathbb{Z}) be generators. For any xH(L)x\in H^{*}(L), there exist R1,,RnH(L)^Λ0[[π1(L)]]R_{1},\dots,R_{n}\in H^{*}(L)\hat{\otimes}\Lambda_{0}[[\pi_{1}(L)]] such that

𝐦1(x)=Dθ1WR1++DθnWRn(modD𝔞)\mathbf{m}_{1}(x)=D_{\theta_{1}}W\cdot R_{1}+\cdots+D_{\theta_{n}}W\cdot R_{n}\quad(\mathrm{mod}\ D{\mathfrak{a}})

By [MS12, Proposition 4.1.4], for a sufficient small number >0\hbar>0, any nontrivial holomorphic disk uu in π2(X,L)\pi_{2}(X,L) satisfies

(24) E(u)>>0E(u)>\hbar>0

In the first place, we prove a weaker statement as follows:

Sublemma 3.18.

Assume the same conditions in Lemma 3.17. For any xH(L)x\in H^{*}(L), there exist R1,,Rn,SH(L)^Λ0[[π1(L)]]R_{1},\dots,R_{n},S\in H^{*}(L)\hat{\otimes}\Lambda_{0}[[\pi_{1}(L)]] such that

𝐦1(x)=Dθ1WR1++DθnWRn+T𝐦1(S)(modD𝔞)\mathbf{m}_{1}(x)=D_{\theta_{1}}W\cdot R_{1}+\cdots+D_{\theta_{n}}W\cdot R_{n}+T^{\hbar}\mathbf{m}_{1}(S)\quad(\mathrm{mod}\ D{\mathfrak{a}})
Proof.

We perform induction on degx\deg x. When degx=0\deg x=0, x=c1H0(L)x=c\cdot\text{1}\in H^{0}(L), so 𝐦1(x)=0\mathbf{m}_{1}(x)=0. When degx=1\deg x=1, we write x=θH1(L)x=\theta\in H^{1}(L), and it follows from the divisor axiom of 𝔪\mathfrak{m} that

(25) 𝐦1(θ)=βTE(β)Yββ,θ𝔪0,β=DθW1+DθQDθW1(modD𝔞)\textstyle\mathbf{m}_{1}(\theta)=\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\langle\partial\beta,\theta\rangle\mathfrak{m}_{0,\beta}=D_{\theta}W\cdot\text{1}+D_{\theta}Q\equiv D_{\theta}W\cdot\text{1}\ \ (\mathrm{mod}\ D{\mathfrak{a}})

Suppose the lemma is true for degrees k1\leq k-1. When degx=k\deg x=k, we may write x=θxx=\theta\wedge x^{\prime} for some θH1(L)\theta\in H^{1}(L) by assumption, so degx=k1\deg x^{\prime}=k-1. Recall that 𝔪2,0(x1,x2)=(1)degx1x1x2\mathfrak{m}_{2,0}(x_{1},x_{2})=(-1)^{\deg x_{1}}x_{1}\wedge x_{2}. Therefore,

(26) x=θx=𝔪2,0(θ,x)=𝐦2(θ,x)+β0TE(β)Yβ𝔪2,β(θ,x)=:𝐦2(θ,x)+TS0x=\theta\wedge x^{\prime}=\mathfrak{m}_{2,0}(-\theta,x^{\prime})=\mathbf{m}_{2}(-\theta,x^{\prime})+\sum_{\beta\neq 0}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{2,\beta}(\theta,x^{\prime})\,\,=:\mathbf{m}_{2}(-\theta,x^{\prime})+T^{\hbar}S_{0}

By (24), if 𝔪2,β(θ,x)0\mathfrak{m}_{2,\beta}(\theta,x^{\prime})\neq 0, we must have E(β)>E(\beta)>\hbar. Hence, S0Λ0[[π1(L)]]S_{0}\in\Lambda_{0}[[\pi_{1}(L)]]. By Corollary 3.2,

(27) 𝐦1(𝐦2(θ,x))=𝐦2(𝐦1(θ),x)+𝐦2(θ,𝐦1(x))\mathbf{m}_{1}(\mathbf{m}_{2}(-\theta,x^{\prime}))=\mathbf{m}_{2}(\mathbf{m}_{1}(\theta),x^{\prime})+\mathbf{m}_{2}(\theta,\mathbf{m}_{1}(x^{\prime}))

Due to (25) and the unitality of 𝔪\mathfrak{m}, the first term on the right side is given by

(28) 𝐦2(𝐦1(θ),x)=𝐦2(DθW1,x)=DθWx(modD𝔞)\mathbf{m}_{2}(\mathbf{m}_{1}(\theta),x^{\prime})=\mathbf{m}_{2}(D_{\theta}W\cdot\text{1},x^{\prime})=D_{\theta}W\cdot x^{\prime}\ \ \ (\mathrm{mod}\ D{\mathfrak{a}})

On the other hand, the induction hypothesis implies that there exist R1,,Rn,SH(L)^Λ0[[π1(L)]]R_{1}^{\prime},\dots,R_{n}^{\prime},S^{\prime}\in H^{*}(L)\hat{\otimes}\Lambda_{0}[[\pi_{1}(L)]] such that 𝐦1(x)=Dθ1WR1++DθnWRn+T𝐦1(S)\mathbf{m}_{1}(x^{\prime})=D_{\theta_{1}}W\cdot R_{1}^{\prime}+\cdots+D_{\theta_{n}}W\cdot R_{n}^{\prime}+T^{\hbar}\mathbf{m}_{1}(S^{\prime}) modulo D𝔞D{\mathfrak{a}}. Then, using the equations (27, 28) with SS^{\prime} in place of xx^{\prime} there, we deduce that the second term on the right side of (27) is

𝐦2(θ,𝐦1(x))\displaystyle\mathbf{m}_{2}(\theta,\mathbf{m}_{1}(x^{\prime})) =Dθ1W𝐦2(θ,R1)++DθnW𝐦2(θ,Rn)+T𝐦2(θ,𝐦1(S))\displaystyle=D_{\theta_{1}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{1})+\cdots+D_{\theta_{n}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{n})+T^{\hbar}\mathbf{m}_{2}(\theta,\mathbf{m}_{1}(S^{\prime}))
=Dθ1W𝐦2(θ,R1)++DθnW𝐦2(θ,Rn)+T(𝐦1(𝐦2(θ,S))DθWS)(modD𝔞)\displaystyle=D_{\theta_{1}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{1})+\cdots+D_{\theta_{n}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{n})+T^{\hbar}\Big{(}\mathbf{m}_{1}(\mathbf{m}_{2}(-\theta,S^{\prime}))-D_{\theta}W\cdot S^{\prime}\Big{)}\ \ \ (\mathrm{mod}\ D{\mathfrak{a}})

Putting things together, we get

𝐦1(x)=𝐦1(𝐦2(θ,x)+TS0)=𝐦2(𝐦1(θ),x)+𝐦2(θ,𝐦1(x))+T𝐦1(S0)\displaystyle\mathbf{m}_{1}(x)=\mathbf{m}_{1}(\mathbf{m}_{2}(-\theta,x^{\prime})+T^{\hbar}S_{0})=\mathbf{m}_{2}(\mathbf{m}_{1}(\theta),x^{\prime})+\mathbf{m}_{2}(\theta,\mathbf{m}_{1}(x^{\prime}))+T^{\hbar}\mathbf{m}_{1}(S_{0})
=\displaystyle= DθW(xTS)+Dθ1W𝐦2(θ,R1)++DθnW𝐦2(θ,Rn)+T𝐦1(S0+𝐦2(θ,S))(modD𝔞)\displaystyle D_{\theta}W\cdot\big{(}x^{\prime}-T^{\hbar}S^{\prime}\big{)}+D_{\theta_{1}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{1})+\cdots+D_{\theta_{n}}W\cdot\mathbf{m}_{2}(\theta,R^{\prime}_{n})+T^{\hbar}\mathbf{m}_{1}\big{(}S_{0}+\mathbf{m}_{2}(-\theta,S^{\prime})\big{)}\ \ \ (\mathrm{mod}\ D{\mathfrak{a}})

In conclusion, if θ=c1θ1++cnθn\theta=c_{1}\theta_{1}+\cdots+c_{n}\theta_{n}, we set Rk:=ck(xTS)+𝐦2(θ,Rk)R_{k}:=c_{k}(x^{\prime}-T^{\hbar}S^{\prime})+\mathbf{m}_{2}(\theta,R_{k}^{\prime}) for 1kn1\leq k\leq n and S:=S0+𝐦2(θ,S)S:=S_{0}+\mathbf{m}_{2}(-\theta,S^{\prime}). The induction is now complete. ∎

Proof of Sublemma 3.18 \implies Lemma 3.17.

We note that 𝐦1\mathbf{m}_{1} is actually Λ0[[π1(L)]]\Lambda_{0}[[\pi_{1}(L)]]-linear by definition. Repeatedly using Sublemma 3.18 implies that for various series Ri(k),S(k)H(L)^Λ0[[π1(L)]]R_{i}^{(k)},S^{(k)}\in H^{*}(L)\hat{\otimes}\Lambda_{0}[[\pi_{1}(L)]],

𝐦1(x)\displaystyle\mathbf{m}_{1}(x) =Dθ1WR1(0)++DθnWRn(0)+T𝐦1(S(0))\displaystyle=D_{\theta_{1}}W\cdot R_{1}^{(0)}+\cdots+D_{\theta_{n}}W\cdot R_{n}^{(0)}+T^{\hbar}\mathbf{m}_{1}(S^{(0)})
𝐦1(S(0))\displaystyle\mathbf{m}_{1}(S^{(0)}) =Dθ1WR1(1)++DθnWRn(1)+T𝐦1(S(1))\displaystyle=D_{\theta_{1}}W\cdot R_{1}^{(1)}+\cdots+D_{\theta_{n}}W\cdot R_{n}^{(1)}+T^{\hbar}\mathbf{m}_{1}(S^{(1)})
\displaystyle\cdots
𝐦1(S(N1))\displaystyle\mathbf{m}_{1}(S^{(N-1)}) =Dθ1WR1(N)++DθnWRn(N)+T𝐦1(S(N))\displaystyle=D_{\theta_{1}}W\cdot R_{1}^{(N)}+\cdots+D_{\theta_{n}}W\cdot R_{n}^{(N)}+T^{\hbar}\mathbf{m}_{1}(S^{(N)})
\displaystyle\cdots

modulo the ideal D𝔞D{\mathfrak{a}}. Then, it follows that

𝐦1(x)=i=1nDθiWk=0TkRi(k)(modD𝔞)\mathbf{m}_{1}(x)=\sum_{i=1}^{n}D_{\theta_{i}}W\cdot\sum_{k=0}^{\infty}T^{k\hbar}R_{i}^{(k)}\ \ \ (\mathrm{mod}\ D{\mathfrak{a}})

Since Ri(k)Λ0[[π1(L)]]R_{i}^{(k)}\in\Lambda_{0}[[\pi_{1}(L)]], the summations are convergent for the adic topology. ∎

Theorem 3.19.

Suppose the de Rham cohomology ring H(L)H^{*}(L) is generated by H1(L;)H^{1}(L;\mathbb{Z}). Under Assumption 1.3, HF(L,𝔪,𝐲)0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})\neq 0 if and only if 𝐲\mathbf{y} is a critical point of WW.

Proof.

The assumption tells that 𝔞=D𝔞=0{\mathfrak{a}}=D{\mathfrak{a}}=0 (23). Suppose 𝐲\mathbf{y} is not a critical point: there exists θH1(L)\theta\in H^{1}(L) with DθW(𝐲)0D_{\theta}W(\mathbf{y})\neq 0. Just like (25), utilizing the divisor axiom of 𝔪\mathfrak{m} yields

𝐦1𝐲(θ)=TE(β)𝐲β𝔪1,β(θ)=DθW(𝐲)1\mathbf{m}_{1}^{\mathbf{y}}(\theta)=\sum T^{E(\beta)}\mathbf{y}^{\partial\beta}\mathfrak{m}_{1,\beta}(\theta)=D_{\theta}W(\mathbf{y})\cdot\text{1}

Thus, for θ:=1DθW(𝐲)θH1(L)^Λ\theta^{\prime}:=\frac{1}{D_{\theta}W(\mathbf{y})}\theta\in H^{1}(L)\hat{\otimes}\Lambda, we have 𝐦1𝐲(θ)=1\mathbf{m}_{1}^{\mathbf{y}}(\theta^{\prime})=\text{1}. The unit [1]=0[\text{1}]=0, so HF(L,𝔪,𝐲)=0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})=0.

Conversely, suppose HF(L,𝔪,𝐲)=0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})=0. So, there exists x0H(L)^Λx_{0}\in H^{*}(L)\hat{\otimes}\Lambda such that 𝐦1𝐲(x0)=1\mathbf{m}_{1}^{\mathbf{y}}(x_{0})=\text{1} in H(L)^ΛΛdimH(L)H^{*}(L)\hat{\otimes}\Lambda\cong\Lambda^{\oplus\dim H^{*}(L)}. By Lemma 3.17, we apply 𝐦1\mathbf{m}_{1} (instead of 𝐦1𝐲\mathbf{m}_{1}^{\mathbf{y}}) to x0x_{0}, so there exist R1,,RnH(L)^Λ[[π1(L)]]R_{1},\dots,R_{n}\in H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]] such that the equation

𝐦1(x0)=Dθ1WR1++DθnWRn\mathbf{m}_{1}(x_{0})=D_{\theta_{1}}W\cdot R_{1}+\cdots+D_{\theta_{n}}W\cdot R_{n}

holds in H(L)^Λ[[π1(L)]]H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]. Further, applying the evaluation map 𝐲\mathcal{E}_{\mathbf{y}} in (20) to the both sides, we get

01=𝐦1𝐲(x0)=Dθ1W(𝐲)R1(𝐲)++DθnW(𝐲)Rn(𝐲)0\neq\text{1}=\mathbf{m}_{1}^{\mathbf{y}}(x_{0})=D_{\theta_{1}}W(\mathbf{y})\cdot R_{1}(\mathbf{y})+\cdots+D_{\theta_{n}}W(\mathbf{y})\cdot R_{n}(\mathbf{y})

in H(L)^ΛH^{*}(L)\hat{\otimes}\Lambda. Hence, at least one of DθiW(𝐲)D_{\theta_{i}}W(\mathbf{y}) is nonzero, so the 𝐲\mathbf{y} cannot be a critical point of WW. ∎

Remark 3.20

The above theorem has a ‘wall-crossing invariance’ in the sense that it does not rely on the choice of 𝔪\mathfrak{m}. If 𝔪~\tilde{\mathfrak{m}} is another AA_{\infty} algebra in 𝒰𝒟\mathscr{UD} which admits an AA_{\infty} homotopy equivalence :𝔪𝔪~\mathfrak{C}:\mathfrak{m}\to\tilde{\mathfrak{m}} such that 1,0=id\mathfrak{C}_{1,0}=\mathrm{id}. By Lemma 2.6, we have an isomorphism ϕ\phi defined by \mathfrak{C} as in (9), and the W~=ϕW\tilde{W}=\phi^{*}W is exactly the superpotential associated to 𝔪~\tilde{\mathfrak{m}}. Let 𝐲~=ϕ(𝐲)\tilde{\mathbf{y}}=\phi(\mathbf{y}). It follows from Corollary 3.16 that 𝐲~\tilde{\mathbf{y}} is a critical point of W~\tilde{W} if and only if 𝐲\mathbf{y} is a critical point of WW. Finally, by Proposition 3.13, HF(L,𝔪,𝐲)0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})\neq 0 if and only if HF(L,𝔪~,𝐲~)0\operatorname{HF}(L,\tilde{\mathfrak{m}},\tilde{\mathbf{y}})\neq 0.

4 Reduced Hochschild cohomology

4.1 Quantum cohomology: Review

4.1.1 Preparation

We will adopt the virtual language in [FOOO20], but the readers may also adopt their own conventions. By a smooth correspondence, we mean a tuple 𝔛=(𝒳,M,M0,f,f0)\mathfrak{X}=(\mathcal{X},M,M_{0},f,f_{0}) consisting of a compact metrizable space 𝒳\mathcal{X} equipped with a Kuranishi structure, two smooth manifolds M0M_{0} and MM, a strongly smooth map f:𝒳Mf:\mathcal{X}\to M and a weakly submersive strongly smooth map f0:𝒳M0f_{0}:\mathcal{X}\to M_{0}. We have a correspondence map of degree dimMvdim(𝒳)\dim M-\mathrm{vdim}(\mathcal{X}) between the space of differential forms:

(29) Corr𝔛Corr(𝒳;f,f0):Ω(M0)Ω(M)\operatorname{Corr}_{\mathfrak{X}}\equiv\operatorname{Corr}(\mathcal{X};f,f_{0}):\Omega^{*}(M_{0})\to\Omega^{*}(M)

One can define the fiber product 𝔛13\mathfrak{X}_{13} of two Kuranishi spaces 𝔛12\mathfrak{X}_{12} and 𝔛23\mathfrak{X}_{23}, and the composition formula [FOOO20, Theorem 10.21] implies

(30) Corr𝔛23(Corr𝔛12(h1)×h2)=Corr𝔛13(h1×h2)\operatorname{Corr}_{\mathfrak{X}_{23}}\big{(}\operatorname{Corr}_{\mathfrak{X}_{12}}(h_{1})\times h_{2}\big{)}=\operatorname{Corr}_{\mathfrak{X}_{13}}(h_{1}\times h_{2})

On the other hand, a smooth correspondence 𝔛\mathfrak{X} induces a boundary smooth correspondence 𝔛=(𝒳,M,M0,f|𝒳,f0|𝒳)\partial\mathfrak{X}=(\partial\mathcal{X},M,M_{0},f|_{\partial\mathcal{X}},f_{0}|_{\partial\mathcal{X}}). The Stokes’ formulas in the Kuranishi theory [FOOO20, Theorem 9.28] is

(31) dM0Corr𝔛Corr𝔛dM=Corr𝔛d_{M_{0}}\circ\operatorname{Corr}_{\mathfrak{X}}-\operatorname{Corr}_{\mathfrak{X}}\circ d_{M}=\operatorname{Corr}_{\partial\mathfrak{X}}

4.1.2 Definition

We briefly recall the basic aspects of the (small) quantum cohomology. A standard reference is [MS12]. Let tt be a formal symbol, and we define (c.f. [McL20])

(32) ΛX:={ibitAibi,AiH2(X), 0E(Ai)ω(Ai)}\Lambda^{X}:=\left\{\sum_{i\in\mathbb{N}}b_{i}t^{A_{i}}\mid b_{i}\in\mathbb{R},A_{i}\in H_{2}(X),\ 0\leq E(A_{i})\equiv\omega(A_{i})\to\infty\right\}

We may also replace H2(X)H2(X;)H_{2}(X)\equiv H_{2}(X;\mathbb{Z}) by the image of the Hurewicz map π2(X)H2(X)\pi_{2}(X)\to H_{2}(X). We call ΛX\Lambda^{X} the quantum Novikov ring of XX, since we reserve the term ‘Novikov ring’ for the Λ0\Lambda_{0}. Define

(33) QH(X;ΛX):=H(X)^ΛXandQH(X;Λ)=H(X)^ΛQH^{*}(X;\Lambda^{X}):=H^{*}(X)\hat{\otimes}\Lambda^{X}\quad\text{and}\quad QH^{*}(X;\Lambda)=H^{*}(X)\hat{\otimes}\Lambda

Take the moduli space (A)\mathcal{M}_{\ell}(A) of genus-0 stable maps of homology class AA with \ell marked points. Let

Ev=(Ev0,Ev1,,Ev1):(A)X\mathrm{Ev}=(\mathrm{Ev}_{0},\mathrm{Ev}_{1},\dots,\mathrm{Ev}_{\ell-1}):\mathcal{M}_{\ell}(A)\to X^{\ell}

be the evaluation maps at the marked points. The quantum product of QH(X;ΛX)QH^{*}(X;\Lambda^{X}) is defined as follows. First, in the chain level, we define for g1,g2Ω(X)g_{1},g_{2}\in\Omega^{*}(X):

(g1g2)A=Corr(3(A);(Ev1,Ev2),Ev0)(g1,g2)(g_{1}\boldsymbol{\ast}g_{2})_{A}=\operatorname{Corr}(\mathcal{M}_{3}(A);(\mathrm{Ev}_{1},\mathrm{Ev}_{2}),\mathrm{Ev}_{0})(g_{1},g_{2})

and define

g1g2=AH2(X)(g1g2)AtAg_{1}\boldsymbol{\ast}g_{2}=\sum_{A\in H_{2}(X)}(g_{1}\boldsymbol{\ast}g_{2})_{A}\cdot t^{A}

Since the moduli (A)\mathcal{M}_{\ell}(A) has no codimension-one boundary, it follows from the Stokes’ formula (31) that the above assignment :Ω(X)Ω(X)Ω(X)^ΛX\boldsymbol{\ast}:\Omega^{*}(X)\otimes\Omega^{*}(X)\to\Omega^{*}(X)\hat{\otimes}\Lambda^{X} is a cochain map and hence induces a product map :H(X)H(X)H(X)^ΛX\boldsymbol{\ast}:H^{*}(X)\otimes H^{*}(X)\to H^{*}(X)\hat{\otimes}\Lambda^{X}. Extending by linearity, we get the quantum product:

:QH(X;ΛX)×QH(X;ΛX)QH(X;ΛX)\boldsymbol{\ast}:QH^{*}(X;\Lambda^{X})\times QH^{*}(X;\Lambda^{X})\to QH^{*}(X;\Lambda^{X})

Abusing the notations, over the Novikov field Λ\Lambda, we also put

g1g2=AH2(X)(g1g2)ATE(A)g_{1}\boldsymbol{\ast}g_{2}=\sum_{A\in H_{2}(X)}(g_{1}\boldsymbol{\ast}g_{2})_{A}\cdot T^{E(A)}

By the same discussion, it defines a quantum product

:QH(X;Λ)×QH(X;Λ)QH(X;Λ)\boldsymbol{\ast}:QH^{*}(X;\Lambda)\times QH^{*}(X;\Lambda)\to QH^{*}(X;\Lambda)

The Novikov field Λ\Lambda is algebraically closed [FOOO10a, Appendix A], so we will work with the latter QH(X;Λ)QH^{*}(X;\Lambda) to study the eigenvalues. But, the first QH(X;ΛX)QH^{*}(X;\Lambda^{X}) is slightly more general and is still useful for our purpose. In either cases, it is standard that the product \boldsymbol{\ast} gives a ring structure. The leading term (g1g2)0(g_{1}\boldsymbol{\ast}g_{2})_{0} of g1g2g_{1}\boldsymbol{\ast}g_{2} is just the standard wedge (cup) product. Besides, the constant-one function 1XH0(X)\text{1}_{X}\in H^{0}(X) is the unit in the quantum cohomology ring [MS12, Proposition 11.1.11].

4.1.3 An example: n\mathbb{CP}^{n}

Let H2(n)\mathcal{H}\in H_{2}(\mathbb{CP}^{n}) be the standard generator represented by the line 1\mathbb{CP}^{1}. The cohomology H(n)H^{*}({\mathbb{CP}^{n}}) is a truncated polynomial ring generated by the class cH2(n)c\in H^{2}(\mathbb{CP}^{n}) such that c()=1c(\mathcal{H})=1 and cn+1=0c^{n+1}=0. It suffices to compute (cicj)(c^{i}\boldsymbol{\ast}c^{j})_{\ell\mathcal{H}} for 0i,jn0\leq i,j\leq n. When =0\ell=0, it corresponds to the usual cup product. When 1\ell\geq 1, it is standard that (see e.g. [MS12, Example 11.1.12])

X(cicj)ck=GW,3n(ci,cj,ck)={1,if=1,i+j+k=2n+10,otherwise\int_{X}(c^{i}\boldsymbol{\ast}c^{j})_{\ell\mathcal{H}}\cup c^{k}=\mathrm{GW}_{\ell\mathcal{H},3}^{\mathbb{CP}^{n}}(c^{i},c^{j},c^{k})=\begin{cases}1,\quad&\text{if}\ \ell=1,i+j+k=2n+1\\ 0,\quad&\text{otherwise}\end{cases}

To sum up,

(cicj)={ci+jif=0, 0i+jnci+jn1if=1,n+1i+j2n0otherwise(c^{i}\boldsymbol{\ast}c^{j})_{\ell\mathcal{H}}=\begin{cases}c^{i+j}&\text{if}\ \ell=0,\ 0\leq i+j\leq n\\ c^{i+j-n-1}&\text{if}\ \ell=1,\ n+1\leq i+j\leq 2n\\ 0&\text{otherwise}\end{cases}

Namely, the quantum product \boldsymbol{\ast} on QH(X;Λ)QH^{*}(X;\Lambda) (resp. QH(X;ΛX)QH^{*}(X;\Lambda^{X})) is given by

cicj={ci+j,if 0i+jnci+jn1TE(),(resp. ci+jn1t)ifn+1i+j2nc^{i}\boldsymbol{\ast}c^{j}=\begin{cases}c^{i+j},&\text{if}\ 0\leq i+j\leq n\\ c^{i+j-n-1}\cdot T^{E(\mathcal{H})},\quad(\text{resp. }\,\,c^{i+j-n-1}\cdot t^{\mathcal{H}})\ &\text{if}\ n+1\leq i+j\leq 2n\end{cases}

Recall that the first Chern class is c1=(n+1)cc_{1}=(n+1)\cdot c. To study the eigenvalues of cc\boldsymbol{\ast}, it would be better to work with QH(X;Λ)QH^{*}(X;\Lambda) instead of QH(X;ΛX)QH^{*}(X;\Lambda^{X}). The matrix of cc\boldsymbol{\ast} is then given by

[0TE()In0]\begin{bmatrix}0&T^{E(\mathcal{H})}\\ I_{n}&0\end{bmatrix}

where InI_{n} denotes the n×nn\times n identity matrix. So, the eigenvalues of cc\boldsymbol{\ast} solve the equation λnTE()=0\lambda^{n}-T^{E(\mathcal{H})}=0. Since c1=(n+1)cc_{1}=(n+1)c, we finally know the eigenvalues of c1c_{1}\boldsymbol{\ast} are

Ξs:=(n+1)TE(n+1)e2πisn+1fors{0,1,,n}\Xi_{s}:=(n+1)T^{E(\frac{\mathcal{H}}{n+1})}e^{\frac{2\pi is}{n+1}}\qquad\text{for}\ s\in\{0,1,\dots,n\}

4.2 Reduced Hochschild cohomology

4.2.1 Label grading

Fix a 𝔊\mathfrak{G}-gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}). We study the Hochschild cohomologies in our labeled setting §2.1.1. The 𝐂𝐂𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}\equiv\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C) is a subspace of the direct product k,β𝐂𝐂k,β\prod_{k,\beta}\operatorname{\mathbf{CC}}_{k,\beta} where each 𝐂𝐂k,β\operatorname{\mathbf{CC}}_{k,\beta} is just a copy of Hom(Ck,C)\operatorname{Hom}(C^{\otimes k},C) with an extra label β\beta. Note that every 𝐂𝐂k,β\operatorname{\mathbf{CC}}_{k,\beta} naturally embeds into 𝐂𝐂\operatorname{\mathbf{CC}}. Since deg𝔪k,β=2kμ(β)\deg\mathfrak{m}_{k,\beta}=2-k-\mu(\beta) involves β\beta, the operator system 𝔪\mathfrak{m} in 𝐂𝐂\operatorname{\mathbf{CC}} only has a well-defined degree modulo 2. To settle this, we introduce the following degree on 𝐂𝐂\operatorname{\mathbf{CC}}.

Definition 4.1.

The label degree |||\cdot| of a kk-multilinear map ϕ\phi in 𝐂𝐂k,β\operatorname{\mathbf{CC}}_{k,\beta} is defined to be

|ϕ|=degϕ+μ(β)|\phi|=\deg^{\prime}\phi+\mu(\beta)\ \in\mathbb{Z}

where degϕ\deg^{\prime}\phi is the shifted degree as a multilinear map (see §2.1.2). This is called the label grading.

Since μ(β)2\mu(\beta)\in 2\mathbb{Z}, |ϕ|degϕ|\phi|\equiv\deg^{\prime}\phi (mod 2\mathrm{mod}\ 2). But, if 𝔪=(𝔪k,β)\mathfrak{m}=(\mathfrak{m}_{k,\beta}) is a gapped AA_{\infty} algebra on CC, then deg𝔪k,β=2kμ(β)\deg\mathfrak{m}_{k,\beta}=2-k-\mu(\beta), deg𝔪k,β=1μ(β)\deg^{\prime}\mathfrak{m}_{k,\beta}=1-\mu(\beta), and so |𝔪k,β|=1|\mathfrak{m}_{k,\beta}|=1 is homogeneous. For a 𝔊\mathfrak{G}-gapped AA_{\infty} homomorphism 𝔣=(𝔣k,β)\mathfrak{f}=(\mathfrak{f}_{k,\beta}), one can similarly check |𝔣k,β|=0|\mathfrak{f}_{k,\beta}|=0.

4.2.2 Brace operations

The brace operation for n=2n=2 is defined by Gerstenhaber [Ger63], and the higher braces are defined by Kadeishvili [Kad] and Getzler [Get93]. We can adapt the definitions to our labeled setting. Given homogeneously-graded 𝔤,𝔣1,,𝔣n𝐂𝐂𝔊(C)\mathfrak{g},\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\in\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C), we define an element 𝔤{𝔣1,,𝔣n}\mathfrak{g}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\} in 𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C) by the formula:

(𝔤{𝔣1,,𝔣n})k,β=r0++rn=rr+t1++tn=kβ¯+β1++βn=β𝔤r+n,β¯((id#s0)r0(𝔣1#s1)t1,β1(id#s2)r2(id#sn1)rn1(𝔣n#sn)tn,βn(id#sn)rn)(\mathfrak{g}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\})_{k,\beta}=\sum_{\scalebox{0.75}{$\begin{subarray}{c}r_{0}+\cdots+r_{n}=r\\ r+t_{1}+\cdots+t_{n}=k\\ \bar{\beta}+\beta_{1}+\cdots+\beta_{n}=\beta\end{subarray}$}}\mathfrak{g}_{r+n,\bar{\beta}}\left((\mathrm{id}^{\#s_{0}})^{\otimes r_{0}}\otimes(\mathfrak{f}_{1}^{\#s_{1}})_{t_{1},\beta_{1}}\otimes(\mathrm{id}^{\#s_{2}})^{\otimes r_{2}}\otimes\cdots\otimes(\mathrm{id}^{\#s_{n-1}})^{\otimes r_{n-1}}\otimes(\mathfrak{f}_{n}^{\#s_{n}})_{t_{n},\beta_{n}}\otimes(\mathrm{id}^{\#s_{n}})^{\otimes r_{n}}\right)

where we denote sm=|𝔣m+1|++|𝔣n|s_{m}=|\mathfrak{f}_{m+1}|+\cdots+|\mathfrak{f}_{n}| for 0mn0\leq m\leq n (so sn=0s_{n}=0). Moreover, note that the gappedness conditions (b) (c) in §2.1.1 ensure the above sum is finite, but the condition (a) is not necessary and will be omitted in §4.2.4. The brace operation has degree zero for the label grading in the sense that for homogeneously-graded elements, we have

(34) |𝔤{𝔣1,,𝔣n}|=|𝔤|+|𝔣1|++|𝔣n||\mathfrak{g}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\}|=|\mathfrak{g}|+|\mathfrak{f}_{1}|+\cdots+|\mathfrak{f}_{n}|

Recall that we already adopt the brace operations to denote the Gerstenhaber product in §2.1.2. Also, we adopt the convention that

𝔤{}=𝔤\mathfrak{g}\{\}=\mathfrak{g}

By routine calculation, we can show that: (see [GJ94, Page 51] and [TT05, Proposition 2.3.2])

Proposition 4.2.

The brace operations satisfy the following property:

𝔥{𝔤1,,𝔤m}{𝔣1,,𝔣n}=(1)ϵ𝔥{𝔣1,,𝔣i1,𝔤1{𝔣i1+1,,𝔣j1},𝔣j1+1,,𝔣im,𝔤m{𝔣im+1,,𝔣jm},𝔣jm+1,,𝔣n}\mathfrak{h}\{\mathfrak{g}_{1},\dots,\mathfrak{g}_{m}\}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\}=\sum(-1)^{\epsilon}\ \mathfrak{h}\big{\{}\mathfrak{f}_{1},\dots,\mathfrak{f}_{i_{1}},\mathfrak{g}_{1}\{\mathfrak{f}_{i_{1}+1},\dots,\mathfrak{f}_{j_{1}}\},\mathfrak{f}_{j_{1}+1},\dots,\mathfrak{f}_{i_{m}},\mathfrak{g}_{m}\{\mathfrak{f}_{i_{m}+1},\dots,\mathfrak{f}_{j_{m}}\},\mathfrak{f}_{j_{m}+1},\dots,\mathfrak{f}_{n}\big{\}}

where the summation is over 0i1j1imjmn{0\leq i_{1}\leq j_{1}\leq\cdots\leq i_{m}\leq j_{m}\leq n} and ϵ=k=1m|𝔤k|=1ik|𝔣|\epsilon=\sum_{k=1}^{m}|\mathfrak{g}_{k}|\cdot\sum_{\ell=1}^{i_{k}}|\mathfrak{f}_{\ell}|.

The sign can be described briefly as follows: for each 𝔤k\mathfrak{g}_{k}, we gather all 𝔣j\mathfrak{f}_{j}’s that appear to the left of 𝔤k\mathfrak{g}_{k}, then compute the sum of the degrees of these 𝔣j\mathfrak{f}_{j}’s multiplied by the degree of 𝔤k\mathfrak{g}_{k}. A few special cases of Proposition 4.2 are as follows:

𝔥{𝔤}{𝔣}\displaystyle\mathfrak{h}\{\mathfrak{g}\}\{\mathfrak{f}\} =𝔥{𝔤{𝔣}}+𝔥{𝔤{},𝔣}+(1)|𝔤||𝔣|𝔥{𝔣,𝔤{}}\displaystyle=\mathfrak{h}\{\mathfrak{g}\{\mathfrak{f}\}\}+\mathfrak{h}\{\mathfrak{g}\{\},\mathfrak{f}\}+(-1)^{|\mathfrak{g}||\mathfrak{f}|}\mathfrak{h}\{\mathfrak{f},\mathfrak{g}\{\}\}
𝔥{𝔤}{𝔣1,𝔣2}\displaystyle\mathfrak{h}\{\mathfrak{g}\}\{\mathfrak{f}_{1},\mathfrak{f}_{2}\} =𝔥{𝔤{𝔣1,𝔣2}}+𝔥{𝔤{𝔣1},𝔣2}+(1)|𝔤||𝔣1|𝔥{𝔣1,𝔤{𝔣2}}\displaystyle=\mathfrak{h}\{\mathfrak{g}\{\mathfrak{f}_{1},\mathfrak{f}_{2}\}\}+\mathfrak{h}\{\mathfrak{g}\{\mathfrak{f}_{1}\},\mathfrak{f}_{2}\}+(-1)^{|\mathfrak{g}||\mathfrak{f}_{1}|}\mathfrak{h}\{\mathfrak{f}_{1},\mathfrak{g}\{\mathfrak{f}_{2}\}\}
+𝔥{𝔤{},𝔣1,𝔣2}+(1)|𝔤||𝔣1|𝔥{𝔣1,𝔤{},𝔣2}+(1)|𝔤|(|𝔣1|+|𝔣2|)𝔥{𝔣1,𝔣2,𝔤{}}\displaystyle+\mathfrak{h}\{\mathfrak{g}\{\},\mathfrak{f}_{1},\mathfrak{f}_{2}\}+(-1)^{|\mathfrak{g}||\mathfrak{f}_{1}|}\ \mathfrak{h}\{\mathfrak{f}_{1},\mathfrak{g}\{\},\mathfrak{f}_{2}\}+(-1)^{|\mathfrak{g}|(|\mathfrak{f}_{1}|+|\mathfrak{f}_{2}|)}\ \mathfrak{h}\{\mathfrak{f}_{1},\mathfrak{f}_{2},\mathfrak{g}\{\}\}

where we also recall the convention that 𝔤{}=𝔤\mathfrak{g}\{\}=\mathfrak{g}.

On the other hand, we define the Gerstenhaber bracket in our labeled setting as follows:

(35) [𝔣,𝔤]:=𝔣{𝔤}(1)|𝔣||𝔤|𝔤{𝔣}[\mathfrak{f},\mathfrak{g}]:=\mathfrak{f}\{\mathfrak{g}\}-(-1)^{|\mathfrak{f}||\mathfrak{g}|}\mathfrak{g}\{\mathfrak{f}\}

Due to (34), |[𝔣,𝔤]|=|𝔣|+|𝔤||[\mathfrak{f},\mathfrak{g}]|=|\mathfrak{f}|+|\mathfrak{g}|. One can easily check the graded skew-symmetry: [𝔣,𝔤]=(1)|𝔣||𝔤|[𝔤,𝔣][\mathfrak{f},\mathfrak{g}]=-(-1)^{|\mathfrak{f}||\mathfrak{g}|}[\mathfrak{g},\mathfrak{f}]. A tedious but straightforward computation yields the graded Jacobi identity:

(36) [𝔣,[𝔤,𝔥]]=[[𝔣,𝔤],𝔥]+(1)|𝔣||𝔤|[𝔤,[𝔣,𝔥]][\mathfrak{f},[\mathfrak{g},\mathfrak{h}]]=[[\mathfrak{f},\mathfrak{g}],\mathfrak{h}]+(-1)^{|\mathfrak{f}||\mathfrak{g}|}[\mathfrak{g},[\mathfrak{f},\mathfrak{h}]]

4.2.3 Hochschild cohomology as usual

Given a gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}), we define

(37) δ𝔪:=[𝔪,]:𝐂𝐂𝔊(C)𝐂𝐂𝔊(C)\delta_{\mathfrak{m}}:=[\mathfrak{m},\cdot]:\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C)\to\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C)

Clearly, δ𝔪δ𝔪=0\delta_{\mathfrak{m}}\circ\delta_{\mathfrak{m}}=0, and it is called a Hochschild differential; thereafter, we may call an operator system in 𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C) a Hochschild cochain. Since |𝔪|=1|\mathfrak{m}|=1, we have |δ𝔪|=1|\delta_{\mathfrak{m}}|=1, i.e. |δ𝔪𝔣|=|𝔣|+1|\delta_{\mathfrak{m}}\mathfrak{f}|=|\mathfrak{f}|+1. In contrast, degδ𝔪\deg^{\prime}\delta_{\mathfrak{m}} is only defined in 2\mathbb{Z}_{2}; this is also a reason why we introduce the label grading above. In addition, one can use (36) to show the graded Leibniz rule:

(38) δ𝔪[𝔣,𝔤]=[δ𝔪𝔣,𝔤]+(1)|𝔣|[𝔣,δ𝔪𝔤]\delta_{\mathfrak{m}}[\mathfrak{f},\mathfrak{g}]=[\delta_{\mathfrak{m}}\mathfrak{f},\mathfrak{g}]+(-1)^{|\mathfrak{f}|}[\mathfrak{f},\delta_{\mathfrak{m}}\mathfrak{g}]

In summary, (𝐂𝐂(C),δ𝔪,[,])(\operatorname{\mathbf{CC}}(C),\delta_{\mathfrak{m}},[,]) is a differential graded Lie algebra. Now, the Hochschild cohomology of a gapped AA_{\infty} algebra (C,𝔪)(C,\mathfrak{m}) is defined by

HH(C,𝔪):=H(𝐂𝐂(C),δ𝔪)HH^{*}(C,\mathfrak{m}):=H^{*}(\operatorname{\mathbf{CC}}(C),\delta_{\mathfrak{m}})

Next, we define the following Yoneda product on the cochain complex 𝐂𝐂(C)\operatorname{\mathbf{CC}}(C): (c.f. [Gan13])

(39) 𝔣𝔪𝔤=(1)|𝔣|+1𝔪{𝔣,𝔤}\mathfrak{f}\cup_{\mathfrak{m}}\mathfrak{g}=(-1)^{|\mathfrak{f}|+1}\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\}

Then, we can carry out a few computations using Proposition 4.2:

0=𝔪{𝔪}{𝔣,𝔤}\displaystyle 0=\mathfrak{m}\{\mathfrak{m}\}\{\mathfrak{f},\mathfrak{g}\} =𝔪{𝔪,𝔣,𝔤}+(1)|𝔣|𝔪{𝔣,𝔪,𝔤}+(1)|𝔣|+|𝔤|𝔪{𝔣,𝔤,𝔪}\displaystyle=\mathfrak{m}\{\mathfrak{m},\mathfrak{f},\mathfrak{g}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m},\mathfrak{g}\}+(-1)^{|\mathfrak{f}|+|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{m}\}
+𝔪{𝔪{𝔣},𝔤}+(1)|𝔣|𝔪{𝔣,𝔪{𝔤}}+𝔪{𝔪{𝔣,𝔤}}\displaystyle+\mathfrak{m}\{\mathfrak{m}\{\mathfrak{f}\},\mathfrak{g}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m}\{\mathfrak{g}\}\}+\mathfrak{m}\{\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\}\}
𝔪{𝔣,𝔤}{𝔪}=(1)|𝔣|+|𝔤|𝔪{𝔪,𝔣,𝔤}+(1)|𝔤|𝔪{𝔣,𝔪,𝔤}+𝔪{𝔣,𝔤,𝔪}+(1)|𝔤|𝔪{𝔣{𝔪},𝔤}+𝔪{𝔣,𝔤{𝔪}}\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\}\{\mathfrak{m}\}=(-1)^{|\mathfrak{f}|+|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{m},\mathfrak{f},\mathfrak{g}\}+(-1)^{|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m},\mathfrak{g}\}+\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{m}\}+(-1)^{|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f}\{\mathfrak{m}\},\mathfrak{g}\}+\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\{\mathfrak{m}\}\}

Hence, a direct calculation yields that δ𝔪(𝔣𝔤)=δ𝔪𝔣𝔤(1)|𝔣|𝔣δ𝔪𝔤\delta_{\mathfrak{m}}(\mathfrak{f}\cup\mathfrak{g})=\delta_{\mathfrak{m}}\mathfrak{f}\cup\mathfrak{g}-(-1)^{|\mathfrak{f}|}\mathfrak{f}\cup\delta_{\mathfrak{m}}\mathfrak{g}, and we get a well-defined Yoneda cup product on the Hochschild cohomology:

=𝔪:HH(C,𝔪)×HH(C,𝔪)HH(C,𝔪)\cup=\cup_{\mathfrak{m}}:HH^{*}(C,\mathfrak{m})\times HH^{*}(C,\mathfrak{m})\to HH^{*}(C,\mathfrak{m})

From (34) it follows that the label degree is ||=1|\cup|=1, namely, |𝔣𝔤|=|𝔣|+|𝔤|+1|\mathfrak{f}\cup\mathfrak{g}|=|\mathfrak{f}|+|\mathfrak{g}|+1.

Proposition 4.3.

The cup product 𝔪\cup_{\mathfrak{m}} defines a ring structure on the Hochschild cohomology HH(C,𝔪)HH^{*}(C,\mathfrak{m}).

Proof.

We aim to check the cup product is associative. In fact, by Proposition 4.2 again, we obtain

0=𝔪{𝔪}{𝔣,𝔤,𝔥}\displaystyle 0=\mathfrak{m}\{\mathfrak{m}\}\{\mathfrak{f},\mathfrak{g},\mathfrak{h}\} =𝔪{𝔪,𝔣,𝔤,𝔥}+(1)|𝔣|𝔪{𝔣,𝔪,𝔤,𝔥}+(1)|𝔣|+|𝔤|𝔪{𝔣,𝔤,𝔪,𝔥}+(1)|𝔣|+|𝔤|+|𝔥|𝔪{𝔣,𝔤,𝔥,𝔪}\displaystyle=\mathfrak{m}\{\mathfrak{m},\mathfrak{f},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|+|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{m},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|+|\mathfrak{g}|+|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h},\mathfrak{m}\}
+𝔪{𝔪{𝔣},𝔤,𝔥}+(1)|𝔣|𝔪{𝔣,𝔪{𝔤},𝔥}+(1)|𝔣|+|𝔤|𝔪{𝔣,𝔤,𝔪{𝔥}}\displaystyle+\mathfrak{m}\{\mathfrak{m}\{\mathfrak{f}\},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m}\{\mathfrak{g}\},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|+|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{m}\{\mathfrak{h}\}\}
+𝔪{𝔪{𝔣,𝔤},𝔥}+(1)|𝔣|𝔪{𝔣,𝔪{𝔤,𝔥}}+𝔪{𝔪{𝔣,𝔤,𝔥}}\displaystyle+\mathfrak{m}\{\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m}\{\mathfrak{g},\mathfrak{h}\}\}+\mathfrak{m}\{\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h}\}\}
𝔪{𝔣,𝔤,𝔥}{𝔪}\displaystyle\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h}\}\{\mathfrak{m}\} =(1)|𝔣|+|𝔤|+|𝔥|𝔪{𝔪,𝔣,𝔤,𝔥}+(1)|𝔤|+|𝔥|𝔪{𝔣,𝔪,𝔤,𝔥}+(1)|𝔥|𝔪{𝔣,𝔤,𝔪,𝔥}+𝔪{𝔣,𝔤,𝔥,𝔪}\displaystyle=(-1)^{|\mathfrak{f}|+|\mathfrak{g}|+|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{m},\mathfrak{f},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{g}|+|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{m},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{m},\mathfrak{h}\}+\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h},\mathfrak{m}\}
+(1)|𝔤|+|𝔥|𝔪{𝔣{𝔪},𝔤,𝔥}+(1)|𝔥|𝔪{𝔣,𝔤{𝔪},𝔥}+𝔪{𝔣,𝔤,𝔥{𝔪}}\displaystyle+(-1)^{|\mathfrak{g}|+|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{f}\{\mathfrak{m}\},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{h}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\{\mathfrak{m}\},\mathfrak{h}\}+\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h}\{\mathfrak{m}\}\}

It follows that

(1)|𝔤|((𝔣𝔤)𝔥𝔣(𝔤𝔥))=𝔪{δ𝔪𝔣,𝔤,𝔥}+(1)|𝔣|𝔪{𝔣,δ𝔪𝔤,𝔥}+(1)|𝔣|+|𝔤|𝔪{𝔣,𝔤,δ𝔪𝔥}+δ𝔪(𝔪{𝔣,𝔤,𝔥})\displaystyle(-1)^{|\mathfrak{g}|}\big{(}(\mathfrak{f}\cup\mathfrak{g})\cup\mathfrak{h}-\mathfrak{f}\cup(\mathfrak{g}\cup\mathfrak{h})\big{)}=\mathfrak{m}\{\delta_{\mathfrak{m}}\mathfrak{f},\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|}\mathfrak{m}\{\mathfrak{f},\delta_{\mathfrak{m}}\mathfrak{g},\mathfrak{h}\}+(-1)^{|\mathfrak{f}|+|\mathfrak{g}|}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\delta_{\mathfrak{m}}\mathfrak{h}\}+\delta_{\mathfrak{m}}\big{(}\mathfrak{m}\{\mathfrak{f},\mathfrak{g},\mathfrak{h}\}\big{)}

Descending to the cohomology completes the proof. ∎

4.2.4 Reduced Hochschild cohomology

In practice, we need a slight variant. The space 𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C) consists of those operator systems 𝔱=(𝔱k,β)\mathfrak{t}^{\prime}=(\mathfrak{t}^{\prime}_{k,\beta}). By the gappedness condition (a) in §2.1.1, we have to assume 𝔱0,0=0\mathfrak{t}^{\prime}_{0,0}=0. But, we must allow nontrivial component for (k,β)=(0,0)(k,\beta)=(0,0) later. Thus, we define

(40) 𝐂𝐂~(C)𝐂𝐂0,0(C)×𝐂𝐂𝔊(C)\operatorname{\mathbf{\widetilde{CC}}}(C)\subset\operatorname{\mathbf{CC}}_{0,0}(C)\times\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C)

to be the space of the operator systems 𝔱=(𝔱k,β)k,β𝔊\mathfrak{t}=(\mathfrak{t}_{k,\beta})_{k\in\mathbb{N},\beta\in\mathfrak{G}} such that 𝔱𝔱0,0(𝔱k,β)(k,β)(0,0)\mathfrak{t}-\mathfrak{t}_{0,0}\equiv(\mathfrak{t}_{k,\beta})_{(k,\beta)\neq(0,0)} is contained in 𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C) and

(41) 𝔱k,β(,1,)=0(k1,β)\mathfrak{t}_{k,\beta}(\dots,\text{1},\dots)=0\ \ \quad(\forall\ k\geq 1,\ \forall\ \beta)

We keep using the label grading on 𝐂𝐂𝔊(C)\operatorname{\mathbf{CC}}_{\mathfrak{G}}(C), and we further define the grading on C𝐂𝐂0,0(C)C\equiv\operatorname{\mathbf{CC}}_{0,0}(C) by

(42) |𝔱0,0|=deg𝔱0,0|\mathfrak{t}_{0,0}|=\deg^{\prime}\mathfrak{t}_{0,0}
Remark 4.4

Algebraically, the above version is needed for the unital ring structures. Geometrically, this is also necessary, because the operator 𝔮,0,0\mathfrak{q}_{\ell,0,0} (for the moduli space of constant maps with one boundary and \ell interior markings as in §4.3) should correspond to a term with label (k,β)=(0,0)(k,\beta)=(0,0).

Lemma 4.5.

The brace operations can be defined on 𝐂𝐂~(C)\operatorname{\mathbf{\widetilde{CC}}}(C).

Proof.

Suppose 𝔤,𝔣1,,𝔣n𝐂𝐂~(C)\mathfrak{g},\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\in\operatorname{\mathbf{\widetilde{CC}}}(C) for n1n\geq 1. We define 𝔥:=𝔤{𝔣1,,𝔣n}\mathfrak{h}:=\mathfrak{g}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\} exactly as in §4.2.2, and there will be at most finite extra terms from the 𝐂𝐂0,0\operatorname{\mathbf{CC}}_{0,0}. Depending on the place of 1, we decompose

𝔥k,β(,1,)=±𝔤s,β¯(,(𝔣i)ti,βi(,1,),)+±𝔤s,β¯(,(𝔣i)ti,βi,,1,,(𝔣i+1)ti+1,βi+1,)\mathfrak{h}_{k,\beta}(\dots,\text{1},\dots)=\sum\pm\mathfrak{g}_{s,\bar{\beta}}(\dots,(\mathfrak{f}_{i})_{t_{i},\beta_{i}}(\dots,\text{1},\dots),\dots)+\sum\pm\mathfrak{g}_{s,\bar{\beta}}(\dots,(\mathfrak{f}_{i})_{t_{i},\beta_{i}},\dots,\text{1},\dots,(\mathfrak{f}_{i+1})_{t_{i+1},\beta_{i+1}},\dots)

Hence, the condition (41) of 𝔥\mathfrak{h} just follows from that of 𝔤\mathfrak{g} and 𝔣i\mathfrak{f}_{i}’s. ∎

In particular, we can define 𝔤{𝔣}\mathfrak{g}\{\mathfrak{f}\} and [𝔤,𝔣]=𝔤{𝔣}±𝔣{𝔤}[\mathfrak{g},\mathfrak{f}]=\mathfrak{g}\{\mathfrak{f}\}\pm\mathfrak{f}\{\mathfrak{g}\} on 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}}. By comparison, 𝔤𝔣\mathfrak{g}\diamond\mathfrak{f} is not defined for 𝔣,𝔤𝐂𝐂~(C)\mathfrak{f},\mathfrak{g}\in\operatorname{\mathbf{\widetilde{CC}}}(C), since there may be infinite extra terms in (6) from the 𝐂𝐂0,0\operatorname{\mathbf{CC}}_{0,0}. We introduce 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}} mainly for the sake of closed-open operators, whereas we must still use 𝐂𝐂\operatorname{\mathbf{CC}} rather than 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}} to develop the homotopy theory of AA_{\infty} algebras for the category 𝒰𝒟\mathscr{UD}, e.g. the Whitehead theorem 2.2. Furthermore, the condition (41) does not exactly match the unitality of an AA_{\infty} algebra 𝔪\mathfrak{m} or that of an AA_{\infty} homomorphism 𝔣\mathfrak{f}. Indeed, we only have 𝔪𝔪2,0𝐂𝐂~(C)\mathfrak{m}-\mathfrak{m}_{2,0}\in\operatorname{\mathbf{\widetilde{CC}}}(C) and 𝔣𝔣1,0𝐂𝐂~(C)\mathfrak{f}-\mathfrak{f}_{1,0}\in\operatorname{\mathbf{\widetilde{CC}}}(C) in general. But, according to our trial-and-error, the condition (41) is indeed the correct one as suggested below:

Lemma 4.6.

Suppose (C,𝔪)Obj𝒰𝒟(C,\mathfrak{m})\in\operatorname{Obj}\mathscr{UD}. Then,

  1. 1.

    the Hochschild differential δ𝔪=[𝔪,]\delta_{\mathfrak{m}}=[\mathfrak{m},\cdot] still gives a differential map on 𝐂𝐂~(C)\operatorname{\mathbf{\widetilde{CC}}}(C);

  2. 2.

    the Yoneda cup product 𝔪\cup_{\mathfrak{m}} induces a map 𝐂𝐂~(C)×𝐂𝐂~(C)𝐂𝐂~(C)\operatorname{\mathbf{\widetilde{CC}}}(C)\times\operatorname{\mathbf{\widetilde{CC}}}(C)\to\operatorname{\mathbf{\widetilde{CC}}}(C).

Proof.

We only know 𝔪𝔪2,0𝐂𝐂~(C)\mathfrak{m}-\mathfrak{m}_{2,0}\in\operatorname{\mathbf{\widetilde{CC}}}(C), but we recall that 𝔪2,0(1,x)=(1)degx+1𝔪2,0(x,1)=x\mathfrak{m}_{2,0}(\text{1},x)=(-1)^{\deg^{\prime}x+1}\mathfrak{m}_{2,0}(x,\text{1})=x.

(1). Given 𝔣𝐂𝐂~(C)\mathfrak{f}\in\operatorname{\mathbf{\widetilde{CC}}}(C), it suffices to show δ𝔪𝔣𝐂𝐂~(C)\delta_{\mathfrak{m}}\mathfrak{f}\in\operatorname{\mathbf{\widetilde{CC}}}(C). By Lemma 4.5, we have [𝔪𝔪2,0,𝔣]𝐂𝐂~(C)[\mathfrak{m}-\mathfrak{m}_{2,0},\mathfrak{f}]\in\operatorname{\mathbf{\widetilde{CC}}}(C). Thus, it suffices to check the condition (41) for δ𝔪2,0(𝔣)=[𝔪2,0,𝔣]\delta_{\mathfrak{m}_{2,0}}(\mathfrak{f})=[\mathfrak{m}_{2,0},\mathfrak{f}]. Specifically, fix k1k\geq 1 and 1ik1\leq i\leq k; we aim to show

Δ:=([𝔪2,0,𝔣])k,β(x1,,xi1,1,xi,,xk1)\Delta:=([\mathfrak{m}_{2,0},\mathfrak{f}])_{k,\beta}(x_{1},\dots,x_{i-1},\text{1},x_{i},\dots,x_{k-1})

vanishes all the time. We check it by cases. Recall that we denote x#=(1)degxxx^{\#}=(-1)^{\deg^{\prime}x}x (5).

  • \ast

    When i1i\neq 1 or kk, using the condition (41) of 𝔣\mathfrak{f} deduces that

    Δ=\displaystyle\Delta= (1)deg𝔣𝔣k1,β(x1#,,xi2#,𝔪2,0(xi1,1),xi,xi+1,,xk1)\displaystyle-(-1)^{\deg^{\prime}\mathfrak{f}}\ \mathfrak{f}_{k-1,\beta}(x_{1}^{\#},\dots,x_{i-2}^{\#},\mathfrak{m}_{2,0}(x_{i-1},\text{1}),x_{i},x_{i+1},\dots,x_{k-1})
    (1)deg𝔣𝔣k1,β(x1#,,xi2#,xi1#,𝔪2,0(1,xi),xi+1,,xk1)\displaystyle-(-1)^{\deg^{\prime}\mathfrak{f}}\ \mathfrak{f}_{k-1,\beta}(x_{1}^{\#},\dots,x_{i-2}^{\#},x_{i-1}^{\#},\mathfrak{m}_{2,0}(\text{1},x_{i}),x_{i+1},\dots,x_{k-1})

    This vanishes as 𝔪2,0(x,1)=(1)degx+1x=x#\mathfrak{m}_{2,0}(x,\text{1})=(-1)^{\deg^{\prime}x+1}x=-x^{\#} and 𝔪2,0(1,x)=x\mathfrak{m}_{2,0}(\text{1},x)=x.

  • \ast

    When i=1i=1, using the condition (41) of 𝔣\mathfrak{f} deduces that

    Δ=\displaystyle\Delta= 𝔪2,0(1#deg𝔣,𝔣k1,β(x1,,xk1))(1)deg𝔣𝔣k1,β(𝔪2,0(1,x1),x2,,xk1)\displaystyle\mathfrak{m}_{2,0}(\text{1}^{\#\deg^{\prime}\mathfrak{f}},\mathfrak{f}_{k-1,\beta}(x_{1},\dots,x_{k-1}))-(-1)^{\deg^{\prime}\mathfrak{f}}\ \mathfrak{f}_{k-1,\beta}(\mathfrak{m}_{2,0}(\text{1},x_{1}),x_{2},\dots,x_{k-1})

    Note that 𝔪2,0(1,x)=x\mathfrak{m}_{2,0}(\text{1},x)=x and 1#deg𝔣=(1)deg𝔣1\text{1}^{\#\deg^{\prime}\mathfrak{f}}=(-1)^{\deg^{\prime}\mathfrak{f}}\text{1}. Thus, we also get Δ=0\Delta=0.

  • \ast

    When i=ki=k, we can similarly compute the sign. Using the condition (41) of 𝔣\mathfrak{f} deduces that

    Δ\displaystyle\Delta =𝔪2,0(𝔣k1,β(x1,,xk1),1)(1)deg𝔣𝔣k1,β(x1#,,xk2#,𝔪2,0(xk1,1))\displaystyle=\mathfrak{m}_{2,0}(\mathfrak{f}_{k-1,\beta}(x_{1},\dots,x_{k-1}),\text{1})-(-1)^{\deg^{\prime}\mathfrak{f}}\mathfrak{f}_{k-1,\beta}(x^{\#}_{1},\dots,x^{\#}_{k-2},\mathfrak{m}_{2,0}(x_{k-1},\text{1}))
    =((1)deg𝔣+degx1++degxk1+1(1)deg𝔣+degx1++degxk2+(degxk1+1))𝔣k1,β(x1,,xk1)=0\displaystyle=\scalebox{0.8}{$\big{(}(-1)^{\deg^{\prime}\mathfrak{f}+\deg^{\prime}x_{1}+\cdots+\deg^{\prime}x_{k-1}+1}-(-1)^{\deg^{\prime}\mathfrak{f}+\deg^{\prime}x_{1}+\cdots+\deg^{\prime}x_{k-2}+(\deg^{\prime}x_{k-1}+1)}\big{)}$}\cdot\mathfrak{f}_{k-1,\beta}(x_{1},\dots,x_{k-1})=0

(2). Given 𝔣,𝔤𝐂𝐂~(C)\mathfrak{f},\mathfrak{g}\in\operatorname{\mathbf{\widetilde{CC}}}(C), we aim to show 𝔪{𝔣,𝔤}𝐂𝐂~(C)\mathfrak{m}\{\mathfrak{f},\mathfrak{g}\}\in\operatorname{\mathbf{\widetilde{CC}}}(C). Since 𝔪𝔪2,0𝐂𝐂~(C)\mathfrak{m}-\mathfrak{m}_{2,0}\in\operatorname{\mathbf{\widetilde{CC}}}(C), applying Lemma 4.5 first implies that (𝔪𝔪2,0){𝔣,𝔤}𝐂𝐂~(C)(\mathfrak{m}-\mathfrak{m}_{2,0})\{\mathfrak{f},\mathfrak{g}\}\in\operatorname{\mathbf{\widetilde{CC}}}(C). However, the conditions (41) of 𝔣\mathfrak{f} and 𝔤\mathfrak{g} also imply that ((𝔪2,0){𝔣,𝔤})k,β(x1,,xi1,1,xi,,xk1)=0\big{(}(\mathfrak{m}_{2,0})\{\mathfrak{f},\mathfrak{g}\}\big{)}_{k,\beta}(x_{1},\dots,x_{i-1},\text{1},x_{i},\dots,x_{k-1})=0 for any k1k\geq 1 and 1ik1\leq i\leq k. Hence, (𝔪2,0){𝔣,𝔤}𝐂𝐂~(C)(\mathfrak{m}_{2,0})\{\mathfrak{f},\mathfrak{g}\}\in\operatorname{\mathbf{\widetilde{CC}}}(C). So, the proof of (2) is now complete. (Remark that actually one can show that 𝔪{𝔣1,,𝔣n}𝐂𝐂~(C)\mathfrak{m}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\}\in\operatorname{\mathbf{\widetilde{CC}}}(C) for any n2n\geq 2 and 𝔣1,,𝔣n𝐂𝐂~(C)\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\in\operatorname{\mathbf{\widetilde{CC}}}(C).) ∎

Combining Lemma 4.5 and 4.6 with the proof of Proposition 4.3 yields the following valid definition:

Definition 4.7.

The reduced Hochschild cohomology for (C,𝔪)Obj𝒰𝒟(C,\mathfrak{m})\in\operatorname{Obj}\mathscr{UD} is defined by

HH~(C,𝔪):=H(𝐂𝐂~(C),δ𝔪)\operatorname{\widetilde{HH}}(C,\mathfrak{m}):=H^{*}(\operatorname{\mathbf{\widetilde{CC}}}(C),\delta_{\mathfrak{m}})

Moreover, the cup product 𝔪\cup_{\mathfrak{m}} also gives rise to a ring structure on HH~(C,𝔪)\operatorname{\widetilde{HH}}(C,\mathfrak{m}).

4.2.5 Unitality

A major advantage of the reduced Hochschild cohomology is the unitality:

Lemma 4.8.

Fix (C,𝔪)Obj𝒰𝒟(C,\mathfrak{m})\in\operatorname{Obj}\mathscr{UD}. Then, HH~(C,𝔪)\operatorname{\widetilde{HH}}(C,\mathfrak{m}) is a unital ring such that the unit is given by

(43) 𝔢=(𝔢k,β)k,β𝔊𝐂𝐂~(C)\mathfrak{e}=(\mathfrak{e}_{k,\beta})_{k\in\mathbb{N},\beta\in\mathfrak{G}}\in\operatorname{\mathbf{\widetilde{CC}}}(C)

where 𝔢0,0=1\mathfrak{e}_{0,0}=\text{1} and 𝔢k,β=0\mathfrak{e}_{k,\beta}=0 for any (k,β)(0,0)(k,\beta)\neq(0,0). Moreover, its label degree is |𝔢|=1|\mathfrak{e}|=-1.

Proof.

Clearly, δ𝔪(𝔢)=0\delta_{\mathfrak{m}}(\mathfrak{e})=0. By (42), the degree of 𝔢\mathfrak{e} is |𝔢|=deg1=1|\mathfrak{e}|=\deg^{\prime}\text{1}=-1. Fix [𝔣]HH~(C,𝔪)[\mathfrak{f}]\in\operatorname{\widetilde{HH}}(C,\mathfrak{m}). Using the unitality of 𝔪\mathfrak{m} implies (𝔢𝔪𝔣)k,β=𝔪2,0(𝔢0,0,𝔣k,β)=𝔣k,β(\mathfrak{e}\cup_{\mathfrak{m}}\mathfrak{f})_{k,\beta}=\mathfrak{m}_{2,0}(\mathfrak{e}_{0,0},\mathfrak{f}_{k,\beta})=\mathfrak{f}_{k,\beta} Similarly, we can check 𝔣𝔪𝔢=𝔣\mathfrak{f}\cup_{\mathfrak{m}}\mathfrak{e}=\mathfrak{f}. ∎

4.2.6 ΛX\Lambda^{X}-module

From now on, we assume the label group is 𝔊=π2(X,L)\mathfrak{G}=\pi_{2}(X,L). The quantum Novikov ring ΛX\Lambda^{X} in (32) acts on 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}} by (tA𝔣)k,β=𝔣k,βA(t^{A}\cdot\mathfrak{f})_{k,\beta}=\mathfrak{f}_{k,\beta-A} for any Aπ2(X)A\in\pi_{2}(X). That is, the 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}} has a natural ΛX\Lambda^{X}-module structure. Specifically, if 𝔣\mathfrak{f} is supported in a single component 𝐂𝐂k,γ\operatorname{\mathbf{CC}}_{k,\gamma}, the tA𝔣t^{A}\cdot\mathfrak{f} is simply the same multilinear map for a different label, supported in 𝐂𝐂k,γ+A\operatorname{\mathbf{CC}}_{k,\gamma+A} instead. This will not violate (41) or any gappedness conditions since E(A)0E(A)\geq 0. Notice that the label degree (§4.2.1) will be changed by

(44) |tA𝔣|=|𝔣|+2c1(A)|t^{A}\cdot\mathfrak{f}|=|\mathfrak{f}|+2c_{1}(A)

since μ(γ+A)=μ(γ)+2c1(A)\mu(\gamma+A)=\mu(\gamma)+2c_{1}(A). Particularly, we always have (1)|tA𝔣|=(1)|𝔣|(-1)^{|t^{A}\cdot\mathfrak{f}|}=(-1)^{|\mathfrak{f}|}.

Proposition 4.9.

The ΛX\Lambda^{X}-action is compatible with the brace operation, namely, tA(𝔤{𝔣1,,𝔣n})=(tA𝔤){𝔣1,,𝔣n}=𝔤{𝔣1,,(tA𝔣i),,𝔣n}t^{A}\cdot\big{(}\mathfrak{g}\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\}\big{)}=(t^{A}\cdot\mathfrak{g})\{\mathfrak{f}_{1},\dots,\mathfrak{f}_{n}\}=\mathfrak{g}\{\mathfrak{f}_{1},\dots,(t^{A}\cdot\mathfrak{f}_{i}),\dots,\mathfrak{f}_{n}\} for 1in1\leq i\leq n.

Proof.

We only consider n=1n=1, and the others are similar. By linearity, we may assume 𝔤\mathfrak{g} and 𝔣\mathfrak{f} are supported in 𝐂𝐂k+1,γ\operatorname{\mathbf{CC}}_{k+1,\gamma} and 𝐂𝐂,η\operatorname{\mathbf{CC}}_{\ell,\eta} respectively, and we may also assume |𝔣|=p|\mathfrak{f}|=p and |𝔤|=q|\mathfrak{g}|=q. Without the labels, we simply have 𝔤{𝔣}(x1,,xk+)=i𝔤(x1#p,,xi#p,𝔣(xi+1,,xi+),,xk+)\mathfrak{g}\{\mathfrak{f}\}(x_{1},\dots,x_{k+\ell})=\sum_{i}\mathfrak{g}(x^{\#p}_{1},\dots,x_{i}^{\#p},\mathfrak{f}(x_{i+1},\dots,x_{i+\ell}),\dots,x_{k+\ell}). Then, one can easily check that tA(𝔤{𝔣})t^{A}\cdot(\mathfrak{g}\{\mathfrak{f}\}), 𝔤{tA𝔣}\mathfrak{g}\{t^{A}\cdot\mathfrak{f}\}, and (tA𝔤){𝔣}(t^{A}\cdot\mathfrak{g})\{\mathfrak{f}\} are the same multilinear map as above, supported in 𝐂𝐂k+,γ+η+A\operatorname{\mathbf{CC}}_{k+\ell,\gamma+\eta+A} with the same label. The signs are also the same thanks to (44). ∎

Corollary 4.10.

δ𝔪(tA𝔣)=tAδ𝔪𝔣\delta_{\mathfrak{m}}(t^{A}\cdot\mathfrak{f})=t^{A}\cdot\delta_{\mathfrak{m}}\mathfrak{f}.

Corollary 4.11.

The (reduced) Hochschild cohomology is a ΛX\Lambda^{X}-module.

4.3 From quantum cohomology to reduced Hochschild cohomology

We study the moduli spaces of pseudo-holomorphic curves. Let LL be an oriented relatively-spin Lagrangian submanifold in a symplectic manifold XX. Let JJ be an ω\omega-tame almost complex structure. Given ,k\ell,k\in\mathbb{N} and βπ2(X,L)\beta\in\pi_{2}(X,L), we denote by ,k+1,β(L,β)\mathcal{M}_{\ell,k+1,\beta}(L,\beta) the moduli space of all equivalence classes [𝐮,𝐳+,𝐳][\mathbf{u},\mathbf{z}^{+},\mathbf{z}] of JJ-holomorphic stable maps (𝐮,𝐳+,𝐳)(\mathbf{u},\mathbf{z}^{+},\mathbf{z}) of genus-0 with one boundary component in LL such that [𝐮]=β[\mathbf{u}]=\beta and 𝐳+=(z1+,,z+)\mathbf{z}^{+}=(z^{+}_{1},\dots,z^{+}_{\ell}), 𝐳=(z0,z1,,zk)\mathbf{z}=(z_{0},z_{1},\dots,z_{k}) are the interior and boundary marked points respectively. Moreover, the moduli space is simply not defined when (,k,β)(0,0,0),(0,1,0)(\ell,k,\beta)\neq(0,0,0),(0,1,0) for the sake of the stability. Note that with the notation in §2.3.1, we have

(45) 0,k+1,β(J,L)=k+1,β(J,L)\mathcal{M}_{0,k+1,\beta}(J,L)=\mathcal{M}_{k+1,\beta}(J,L)

The moduli space admits natural evaluation maps corresponding to the marked points:

(ev+,ev0,ev)=(ev1+,,ev+;ev0;ev1,,evk):,k+1,β(J,L)X×Lk+1(\mathrm{ev}^{+},\mathrm{ev}_{0},\mathrm{ev})=(\mathrm{ev}_{1}^{+},\dots,\mathrm{ev}_{\ell}^{+};\mathrm{ev}_{0};\mathrm{ev}_{1},\dots,\mathrm{ev}_{k}):\mathcal{M}_{\ell,k+1,\beta}(J,L)\to X^{\ell}\times L^{k+1}

where evi+([𝐮,𝐳+,𝐳])=𝐮(zi+)\mathrm{ev}^{+}_{i}([\mathbf{u},\mathbf{z}^{+},\mathbf{z}])=\mathbf{u}(z^{+}_{i}) and evi([𝐮,𝐳+,𝐳])=𝐮(zi)\mathrm{ev}_{i}([\mathbf{u},\mathbf{z}^{+},\mathbf{z}])=\mathbf{u}(z_{i}). The codimension-one boundary of ,k+1,β(J,L)\mathcal{M}_{\ell,k+1,\beta}(J,L) in the sense of Kuranishi structure and smooth correspondence (29) is given by the following union of the fiber products:

(46) ,k+1,β(J,L)=1,k1+1,β1(J,L)×eviev02,k2+1,β2(J,L)\partial\mathcal{M}_{\ell,k+1,\beta}(J,L)=\bigcup\mathcal{M}_{\ell_{1},k_{1}+1,\beta_{1}}(J,L)\ \ \ {}_{\mathrm{ev}_{0}}\times_{\mathrm{ev}_{i}}\mathcal{M}_{\ell_{2},k_{2}+1,\beta_{2}}(J,L)

where the union is taken over all the (1,2)(\ell_{1},\ell_{2})-shuffles of {1,,}\{1,\dots,\ell\} with 1+2=\ell_{1}+\ell_{2}=\ell, k1+k2=kk_{1}+k_{2}=k, β1+β2=β\beta_{1}+\beta_{2}=\beta, and 1ik21\leq i\leq k_{2}. Here a (1,2)(\ell_{1},\ell_{2})-shuffle is a permutation (μ1,,μ1,ν1,,ν2)(\mu_{1},\dots,\mu_{\ell_{1}},\nu_{1},\dots,\nu_{\ell_{2}}) of (1,2,,)(1,2,\dots,\ell) such that {μ1<<μ1}\{\mu_{1}<\cdots<\mu_{\ell_{1}}\} and {ν1<<ν2}\{\nu_{1}<\cdots<\nu_{\ell_{2}}\}.

Consider the smooth correspondence 𝔛=(𝒳,M,M0,f,f0)\mathfrak{X}=(\mathcal{X},M,M_{0},f,f_{0}) given by M0=LM_{0}=L, M=X×LkM=X^{\ell}\times L^{k}, f0=ev0f_{0}=\mathrm{ev}_{0}, f=(ev+,ev1,,evk)f=(\mathrm{ev}^{+},\mathrm{ev}_{1},\dots,\mathrm{ev}_{k}), and 𝒳=,k+1,β(J,L)\mathcal{X}=\mathcal{M}_{\ell,k+1,\beta}(J,L). For (,β)(0,0)(\ell,\beta)\neq(0,0) and for giΩ(X)g_{i}\in\Omega^{*}(X) and hiΩ(L)h_{i}\in\Omega^{*}(L), we define

𝔮,k,βJ,L(g1,,g,h1,,hk)=1!Corr(,k+1,β(J,L);(ev+,ev),ev0)(g1gh1hk)\mathfrak{q}^{J,L}_{\ell,k,\beta}(g_{1},\dots,g_{\ell},h_{1},\dots,h_{k})=\tfrac{1}{\ell!}\operatorname{Corr}(\mathcal{M}_{\ell,k+1,\beta}(J,L);(\mathrm{ev}^{+},\mathrm{ev}),\mathrm{ev}_{0})(g_{1}\wedge\cdots g_{\ell}\wedge h_{1}\wedge\cdots\wedge h_{k})

by (29). Alternatively, it may be also instructive to adopt the following notation:

𝔮,k,βJ,L(g1,,g;h1,,hk)=±1!(ev0)!(ev1+g1ev+gev1h1evkhk)\mathfrak{q}^{J,L}_{\ell,k,\beta}(g_{1},\dots,g_{\ell};h_{1},\dots,h_{k})=\pm\tfrac{1}{\ell!}(\mathrm{ev}_{0})_{!}(\mathrm{ev}^{+*}_{1}g_{1}\wedge\cdots\mathrm{ev}^{+*}_{\ell}g_{\ell}\wedge\mathrm{ev}^{*}_{1}h_{1}\wedge\cdots\wedge\mathrm{ev}^{*}_{k}h_{k})

If (,β)=(0,0)(\ell,\beta)=(0,0), then we exceptionally define 𝔮0,k,0J,L=0\mathfrak{q}^{J,L}_{0,k,0}=0 for k1,2k\neq 1,2, 𝔮0,1,0J,L(h)=dh\mathfrak{q}^{J,L}_{0,1,0}(h)=dh, and 𝔮0,2,0J,L(h1,h2)=(1)degh1h1h2\mathfrak{q}^{J,L}_{0,2,0}(h_{1},h_{2})=(-1)^{\deg h_{1}}h_{1}\wedge h_{2}.

Proposition 4.12.

The operators 𝔮,k,β:=𝔮,k,βJ,L\mathfrak{q}_{\ell,k,\beta}:=\mathfrak{q}^{J,L}_{\ell,k,\beta} satisfy the following properties:

  1. (i)

    deg𝔮,k,β=22kμ(β)\deg\mathfrak{q}_{\ell,k,\beta}=2-2\ell-k-\mu(\beta). The 𝔮,k,β0\mathfrak{q}_{\ell,k,\beta}\neq 0 only if E(β)0E(\beta)\geq 0.

  2. (ii)

    For any permutation σ𝒮\sigma\in\mathcal{S}_{\ell}, we have

    𝔮,k,β(g1,,g;h1,,hk)=(1)a𝔮,k,β(gσ(1),,gσ();h1,,hk)\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,g_{\ell};h_{1},\dots,h_{k})=(-1)^{a}\mathfrak{q}_{\ell,k,\beta}(g_{\sigma(1)},\dots,g_{\sigma(\ell)};h_{1},\dots,h_{k})

    where a=i<j;σ(i)>σ(j)deggideggja=\sum_{i<j;\sigma(i)>\sigma(j)}\deg g_{i}\deg g_{j}.

  3. (iii)

    When =0\ell=0, we have 𝔮0,k,β=𝔪ˇk,βJ,L\mathfrak{q}_{0,k,\beta}=\check{\mathfrak{m}}^{J,L}_{k,\beta}2.3.1).

  4. (iv)

    When 0\ell\neq 0 and β=0\beta=0, we have 𝔮,k,0=0\mathfrak{q}_{\ell,k,0}=0 except 𝔮1,0,0=ι:Ω(X)Ω(L)\mathfrak{q}_{1,0,0}=\iota^{*}:\Omega^{*}(X)\to\Omega^{*}(L) where ι:LX\iota:L\to X is the inclusion.

  5. (v)

    Let 1Ω0(L)\text{1}\in\Omega^{0}(L) be the constant-one function. Then, 𝔮,k,β(g1,,g;h1,,hi1,1,hi,,hk)=0\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,g_{\ell};h_{1},\dots,h_{i-1},\text{1},h_{i},\dots,h_{k})=0 except 𝔮0,2,0(1,h)=(1)degh𝔮0,2,0(h,1)=h\mathfrak{q}_{0,2,0}(\text{1},h)=(-1)^{\deg h}\mathfrak{q}_{0,2,0}(h,\text{1})=h.

  6. (vi)

    Given g1,,gΩ(X)g_{1},\dots,g_{\ell}\in\Omega^{*}(X) and h1,,hkΩ(L)h_{1},\dots,h_{k}\in\Omega^{*}(L),

    1i(1)𝔮,k,β(g1,,dgi,,g;h1,,hk)+β1+β2=βλ+μ+ν=k1+2=(1,2)-shuffles(μ1,,μ1,ν1,,ν2)\displaystyle\sum_{1\leq i\leq\ell}(-1)^{\dagger}\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,dg_{i},\dots,g_{\ell};h_{1},\dots,h_{k})+\sum_{\begin{subarray}{c}\scalebox{0.6}{$\beta_{1}+\beta_{2}=\beta$}\\ \scalebox{0.6}{$\lambda+\mu+\nu=k$}\\ \scalebox{0.6}{$\ell_{1}+\ell_{2}=\ell$}\end{subarray}}\sum_{\begin{subarray}{c}\scalebox{0.6}{$(\ell_{1},\ell_{2})\text{-shuffles}$}\\ \scalebox{0.6}{$(\mu_{1},\dots,\mu_{\ell_{1}},\nu_{1},\dots,\nu_{\ell_{2}})$}\end{subarray}}
    (1)𝔮1,λ+μ+1,β1(gμ1,,gμ1;h1,,hλ,𝔮2,ν,β2(gν1,,gν2;hλ+1,,hλ+ν),hλ+ν+1,,hk)=0\displaystyle(-1)^{\ast}\mathfrak{q}_{\ell_{1},\lambda+\mu+1,\beta_{1}}(g_{\mu_{1}},\dots,g_{\mu_{\ell_{1}}};h_{1},\dots,h_{\lambda},\mathfrak{q}_{\ell_{2},\nu,\beta_{2}}(g_{\nu_{1}},\dots,g_{\nu_{\ell_{2}}};h_{\lambda+1},\dots,h_{\lambda+\nu}),h_{\lambda+\nu+1},\dots,h_{k})=0

    where =j=1i1deggj\dagger=\sum_{j=1}^{i-1}\deg g_{j} and =j=11deggμj+j=1λ(deghj1)j=12(deggνj1)\ast=\sum_{j=1}^{\ell_{1}}\deg g_{\mu_{j}}+\sum_{j=1}^{\lambda}(\deg h_{j}-1)\cdot\sum_{j=1}^{\ell_{2}}(\deg g_{\nu_{j}}-1).

Sketch of proof.

Nothing is new here, and we only give rough explanations; see the various literature for more details: [ST16], [FOOO10b, Theorem 3.8.32], [FOOO16, §2.3], and [FOOO19, Chapter 17]. First, the items (i) (ii) (iii) should be clear. We explain (iv). If β=0\beta=0, the moduli spaces consist of constant maps into LL. Hence, ,k+1,β(J,L)L×,k+1\mathcal{M}_{\ell,k+1,\beta}(J,L)\cong L\times\mathcal{M}_{\ell,k+1}, and the evaluation maps just become the projection maps evipr:L×,k+1L\mathrm{ev}_{i}\equiv\mathrm{pr}:L\times\mathcal{M}_{\ell,k+1}\to L or the projection maps pre-composed with the inclusion evi+ιpr:L×,k+1LX\mathrm{ev}^{+}_{i}\equiv\iota\circ\mathrm{pr}:L\times\mathcal{M}_{\ell,k+1}\to L\to X. Then, 𝔮,k,0(g1,,g;h1,,hk)=±pr!pr(ιg1ιgh1hk)\mathfrak{q}_{\ell,k,0}(g_{1},\dots,g_{\ell};h_{1},\dots,h_{k})=\pm\mathrm{pr}_{!}\mathrm{pr}^{*}\big{(}\iota^{*}g_{1}\wedge\cdots\iota^{*}g_{\ell}\wedge h_{1}\wedge\cdots\wedge h_{k}\big{)}. Note that pr!pr0\mathrm{pr}_{!}\mathrm{pr}^{*}\neq 0 only if the fibers of pr\mathrm{pr} has dimension zero, namely, 2+k2=dim,k+1=02\ell+k-2=\dim\mathcal{M}_{\ell,k+1}=0. When 0\ell\neq 0, the only possibility is that =1,k=0\ell=1,k=0. In this case, pr=idL\mathrm{pr}=\mathrm{id}_{L}, and 𝔮1,0,0=ι\mathfrak{q}_{1,0,0}=\iota^{*}. Next, we can check the item (v) by studying the forgetful map of the marked point of 1 in a similar way, c.f. [Yua20, §6]. Eventually, we explain (vi). For the smooth correspondence 𝔛\mathfrak{X} as above, applying the Stokes’ formulas (31) to (46) implies that: d𝔮,k,β(g1,,g;h1,hk)±i𝔮,k,β(g1,,dgi,,g;h1,,hk)±i𝔮,k,β(g1,,g;h1,,dhi,,hk)=±(1,β1)0(2,β2)𝔮1,λ+μ+1,β1(gμ1,,gμ1;h1,hi,𝔮2,ν,β2(gν1,,gν2;hλ+1,,hλ+ν),hλ+ν+1,,hk)d\circ\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,g_{\ell};h_{1},\dots h_{k})\pm\sum_{i}\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,dg_{i},\dots,g_{\ell};h_{1},\dots,h_{k})\pm\sum_{i}\mathfrak{q}_{\ell,k,\beta}(g_{1},\dots,g_{\ell};h_{1},\dots,dh_{i},\dots,h_{k})=\pm\sum_{(\ell_{1},\beta_{1})\neq 0\neq(\ell_{2},\beta_{2})}\\ \mathfrak{q}_{\ell_{1},\lambda+\mu+1,\beta_{1}}(g_{\mu_{1}},\dots,g_{\mu_{\ell_{1}}};h_{1}\dots,h_{i},\mathfrak{q}_{\ell_{2},\nu,\beta_{2}}(g_{\nu_{1}},\dots,g_{\nu_{\ell_{2}}};h_{\lambda+1},\dots,h_{\lambda+\nu}),h_{\lambda+\nu+1},\dots,h_{k}). Now, as 𝔮0,1,0=𝔪ˇ1,0J,L=d\mathfrak{q}_{0,1,0}=\check{\mathfrak{m}}_{1,0}^{J,L}=d, the equation in (vi) then follows. ∎

For our purpose, we focus on the case =1\ell=1, then the equation in (vi) becomes the following:

(47)

Accordingly, we can define a linear map

(48) 𝔮^:Ω(X)𝐂𝐂~(Ω(L))\hat{\mathfrak{q}}:\Omega^{*}(X)\to\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L))

by setting 𝔮^(g)k,β=𝔮1,k,β(g;)\hat{\mathfrak{q}}(g)_{k,\beta}=\mathfrak{q}_{1,k,\beta}(g;\cdots). Here one can directly check that the condition (41) by Proposition 4.12 (v). Moreover, for the label grading (§4.2.1), the item (i) implies

(49) |𝔮^(g)|=|𝔮1,k,β(g)|=degg1|\hat{\mathfrak{q}}(g)|=|\mathfrak{q}_{1,k,\beta}(g)|=\deg g-1

By §4.2.6, we can extend it linearly and thus obtain a ΛX\Lambda^{X}-linear map Ω(X)^ΛX𝐂𝐂~(Ω(L))\Omega^{*}(X)\hat{\otimes}\Lambda^{X}\to\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)), still denoted by 𝔮^\hat{\mathfrak{q}}. Then, the equation (47) exactly implies

𝔮^(dg)+δ𝔪ˇJ,L(𝔮^(g))=0\hat{\mathfrak{q}}(dg)+\delta_{\check{\mathfrak{m}}^{J,L}}\big{(}\hat{\mathfrak{q}}(g)\big{)}=0

Consequently, the 𝔮^\hat{\mathfrak{q}} induces a ΛX\Lambda^{X}-linear map to the reduced Hochschild cohomology of 𝔪ˇJ,L\check{\mathfrak{m}}^{J,L}:

(50) [𝔮^]:QH(X;ΛX)H(X)^ΛXHH~(Ω(L),𝔪ˇJ,L)[\hat{\mathfrak{q}}]:QH^{*}(X;\Lambda^{X})\equiv H^{*}(X)\hat{\otimes}\Lambda^{X}\to\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}}^{J,L})

Recall that the ud-homotopy theory of AA_{\infty} algebras for 𝒰𝒟\mathscr{UD} must work on 𝐂𝐂\operatorname{\mathbf{CC}} in (4). In contrast, as 𝔮^(g)0,0=ιg\hat{\mathfrak{q}}(g)_{0,0}=\iota^{*}g can be nonzero by Proposition 4.12 (iv), we must work on 𝐂𝐂~\operatorname{\mathbf{\widetilde{CC}}} in (40) here.

The reduced Hochschild cohomology ring HH~(Ω(L),𝔪ˇJ,L)\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}}^{J,L}) (Definition 4.7) admits a unital ring structure. The unit is the 𝔢\mathfrak{e} that is supported in 𝐂𝐂0,0\operatorname{\mathbf{CC}}_{0,0} (Lemma 4.8), and its label degree is 1-1. Meanwhile, QH(X;ΛX)H(X)^ΛXQH^{*}(X;\Lambda^{X})\equiv H^{*}(X)\hat{\otimes}\Lambda^{X} has a unital ring structure with the unit 1X\text{1}_{X} of degree 0.

Our goal is to show the 𝔮^\hat{\mathfrak{q}} gives a unital ring homomorphism. But, as we mentioned in the introduction, the main complication is not Theorem 4.13 below but that the moduli space geometry cannot make a ring homomorphism into the reduced Hochschild cohomology of the minimal model AA_{\infty} algebras. We will resolve this issue soon later by using several techniques about 𝒰𝒟\mathscr{UD}.

Theorem 4.13.

The map [𝔮^][\hat{\mathfrak{q}}] is a unital ring homomorphism of degree 1-1 such that 𝔮^(1X)=𝔢\hat{\mathfrak{q}}(\text{1}_{X})=\mathfrak{e}.

Sketch of proof.

The argument here is standard and well-known to many experts, and we simply rewrite it in our notations and languages. But, we will moreover discover that our notion of the reduced Hochschild cohomology fits perfectly as well. In short, Proposition 4.12 (v) corresponds to (41).

First off, the degree matches by (49). Next, we consider the forgetful map 𝔣𝔬𝔯𝔤𝔢𝔱:2,k+1,β(J,L)2,1{\mathfrak{forget}}:\mathcal{M}_{2,k+1,\beta}(J,L)\to\mathcal{M}_{2,1} which forgets the maps and all the incoming boundary marked points and then shrinks the resulting unstable domain components if any. Given 𝔵2,1\mathfrak{x}\in\mathcal{M}_{2,1}, we denote the fiber of 𝔣𝔬𝔯𝔤𝔢𝔱{\mathfrak{forget}} over 𝔵\mathfrak{x} by

2,k+1,β𝔵(J,L):=𝔣𝔬𝔯𝔤𝔢𝔱1(𝔵)\mathcal{M}^{\mathfrak{x}}_{2,k+1,\beta}(J,L):={\mathfrak{forget}}^{-1}(\mathfrak{x})

Notice that the 2,1\mathcal{M}_{2,1} is homeomorphic to the closed unit disk 𝔻\mathbb{D} and has a stratification (see [FOOO16, Figure 2.6.1]) which consists of two copies of (1,1)(-1,1), int(𝔻){0}\mathrm{int}(\mathbb{D})\setminus\{0\}, and three distinguished moduli points [Σ0][\Sigma_{0}], [Σ12][\Sigma_{12}], [Σ21][\Sigma_{21}]. Geometrically, we may assume one interior marking is at 0𝔻0\in\mathbb{D}, and the other is denoted by zz. Then, the three moduli points corresponding to the situations when z0,±1z\to 0,\pm 1 respectively. To be specific, we have the following fiber products:

(51) 2,k+1,β[Σ0](J,L)=γ+A=β3(A)×ev1+Ev01,k+1,γ(J,L)\mathcal{M}^{[\Sigma_{0}]}_{2,k+1,\beta}(J,L)=\bigcup_{\gamma+A=\beta}\mathcal{M}_{3}(A)\ \ \ {}_{{}_{\mathrm{Ev}_{0}}}\times_{\scalebox{0.66}{$\mathrm{ev}_{1}^{+}$}}\mathcal{M}_{1,k+1,\gamma}(J,L)
2,k+1,β[Σ12](J,L)=1i<jm+2m+m′′+m=kγ+γ′′+γ=β1,m+1,γ(J,L)×eviev00,m+3,γ(J,L)×ev0evj1,m′′+1,γ′′(J,L)\mathcal{M}^{[\Sigma_{12}]}_{2,k+1,\beta}(J,L)=\bigcup_{\begin{subarray}{c}1\leq i<j\leq m+2\\ m^{\prime}+m^{\prime\prime}+m=k\\ \gamma^{\prime}+\gamma^{\prime\prime}+\gamma=\beta\end{subarray}}\mathcal{M}_{1,m^{\prime}+1,\gamma^{\prime}}(J,L)\ \ \ {}_{\mathrm{ev}_{0}}\times_{\mathrm{ev}_{i}}\ \mathcal{M}_{0,m+3,\gamma}(J,L)\ \ \ {}_{\mathrm{ev}_{j}}\times_{\mathrm{ev}_{0}}\ \mathcal{M}_{1,m^{\prime\prime}+1,\gamma^{\prime\prime}}(J,L)

Now, we take a path 𝔶:[0,1]2,1\mathfrak{y}:{[0,1]}\to\mathcal{M}_{2,1} from [Σ0][\Sigma_{0}] to [Σ12][\Sigma_{12}] and define

2,k+1,β𝔶(J,L):=t[0,1]{t}×2,k+1,β𝔶(t)(J,L)\mathcal{M}^{\mathfrak{y}}_{2,k+1,\beta}(J,L):=\bigcup_{t\in{[0,1]}}\{t\}\times\mathcal{M}^{\mathfrak{y}(t)}_{2,k+1,\beta}(J,L)

Its codimensional-one boundary is given by the union of 2,k+1,β[Σ0](J,L)\mathcal{M}^{[\Sigma_{0}]}_{2,k+1,\beta}(J,L), 2,k+1,β[Σ12](J,L)\mathcal{M}^{[\Sigma_{12}]}_{2,k+1,\beta}(J,L), and

(52) m+m′′=kβ+β′′=β2,m+2,β𝔶(J,L)×ev0evi0,m′′+1,β′′(J,L)\bigcup_{\begin{subarray}{c}m^{\prime}+m^{\prime\prime}=k\\ \beta^{\prime}+\beta^{\prime\prime}=\beta\end{subarray}}\mathcal{M}^{\mathfrak{y}}_{2,m^{\prime}+2,\beta^{\prime}}(J,L)\ \ \ {}_{\mathrm{ev}_{i}}\times_{\mathrm{ev}_{0}}\mathcal{M}_{0,m^{\prime\prime}+1,\beta^{\prime\prime}}(J,L)
m+m′′=kβ+β′′=β0,m+2,β(J,L)×ev0evi2,m′′+1,β′′𝔶(J,L)\bigcup_{\begin{subarray}{c}m^{\prime}+m^{\prime\prime}=k\\ \beta^{\prime}+\beta^{\prime\prime}=\beta\end{subarray}}\mathcal{M}_{0,m^{\prime}+2,\beta^{\prime}}(J,L)\ \ \ {}_{\mathrm{ev}_{i}}\times_{\mathrm{ev}_{0}}\mathcal{M}^{\mathfrak{y}}_{2,m^{\prime\prime}+1,\beta^{\prime\prime}}(J,L)
Refer to caption
Figure 2: The figure for Equation (52) where the open circle indicates the outgoing marked point and the filled circles are the other boundary marked points.

Given g1,g2Ω(X)g_{1},g_{2}\in\Omega^{*}(X) and h1,,hkΩ(L)h_{1},\dots,h_{k}\in\Omega^{*}(L), we define

𝔔2,k,β(g1,g2;h1,,hk):=Corr(2,k+1,β𝔶(J,L);(ev1+,ev2+,ev),ev0)(g1,g2,h1,,hk)\mathfrak{Q}_{2,k,\beta}(g_{1},g_{2};h_{1},\dots,h_{k}):=\operatorname{Corr}(\mathcal{M}^{\mathfrak{y}}_{2,k+1,\beta}(J,L);(\mathrm{ev}^{+}_{1},\mathrm{ev}^{+}_{2},\mathrm{ev}),\mathrm{ev}_{0})(g_{1},g_{2},h_{1},\dots,h_{k})

by (29). Applying the Stokes’ formula (31) yields that

𝔔2,k,β(dg1,g2;h1,,hk)+𝔔2,k,β(g1,dg2;h1,,hk)\displaystyle\mathfrak{Q}_{2,k,\beta}(dg_{1},g_{2};h_{1},\dots,h_{k})+\mathfrak{Q}_{2,k,\beta}(g_{1},dg_{2};h_{1},\dots,h_{k})
=β+β′′=β±𝔔2,m+1,β(g1,g2;h1,,hλ,𝔪ˇm′′,β′′J,L(hλ+1,,hλ+ν),hλ+ν+1,,hk)\displaystyle=\sum_{\beta^{\prime}+\beta^{\prime\prime}=\beta}\pm\mathfrak{Q}_{2,m^{\prime}+1,\beta^{\prime}}(g_{1},g_{2};h_{1},\dots,h_{\lambda},\check{\mathfrak{m}}^{J,L}_{m^{\prime\prime},\beta^{\prime\prime}}(h_{\lambda+1},\dots,h_{\lambda+\nu}),h_{\lambda+\nu+1},\dots,h_{k})
+β+β′′=β±𝔪ˇm+1,βJ,L(h1,,hλ,𝔔2,m′′,β′′(g1,g2;hλ+1,,hλ+ν),hλ+ν+1,,hk)\displaystyle+\sum_{\beta^{\prime}+\beta^{\prime\prime}=\beta}\pm\check{\mathfrak{m}}^{J,L}_{m^{\prime}+1,\beta^{\prime}}(h_{1},\dots,h_{\lambda},\mathfrak{Q}_{2,m^{\prime\prime},\beta^{\prime\prime}}(g_{1},g_{2};h_{\lambda+1},\dots,h_{\lambda+\nu}),h_{\lambda+\nu+1},\dots,h_{k})
+γ+A=β±𝔮1,k,γ((g1g2)A;h1,,hk)\displaystyle+\sum_{\gamma+A=\beta}\pm\mathfrak{q}_{1,k,\gamma}((g_{1}\boldsymbol{\ast}g_{2})_{A};h_{1},\dots,h_{k})
+γ+γ′′+γ=β±𝔪ˇm+2,γJ,L(h1,,hi1,𝔮1,m,γ(g1;hi,,hi+m1),,𝔮1,m′′,γ′′(g2;hj+m,),,hk)\displaystyle+\sum_{\gamma^{\prime}+\gamma^{\prime\prime}+\gamma=\beta}\pm\ \check{\mathfrak{m}}^{J,L}_{m+2,\gamma}(h_{1},\dots,h_{i-1},\mathfrak{q}_{1,m^{\prime},\gamma^{\prime}}(g_{1};h_{i},\dots,h_{i+m^{\prime}-1}),\dots,\mathfrak{q}_{1,m^{\prime\prime},\gamma^{\prime\prime}}(g_{2};h_{j+m^{\prime}},\dots),\dots,h_{k})

In reality, the left side together with those terms with (m′′,β′′)=(1,0)(m^{\prime\prime},\beta^{\prime\prime})=(1,0) in the first sum of the right side corresponds to the left side of the Stokes’ formula (31); all others correspond to the right side of (31): the first and second summations correspond to (52), and the third and forth summations correspond to (51). As in (48), we define a map

𝔔^:Ω(X)Ω(X)𝐂𝐂~(Ω(L))\mathfrak{\hat{Q}}:\Omega^{*}(X)\otimes\Omega^{*}(X)\to\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L))

by setting (𝔔^(g1,g2))k,β=𝔔2,k,β(g1,g2;)\big{(}\mathfrak{\hat{Q}}(g_{1},g_{2})\big{)}_{k,\beta}=\mathfrak{Q}_{2,k,\beta}(g_{1},g_{2};\cdots). By Proposition 4.12 (v), the image of 𝔔^\mathfrak{\hat{Q}} satisfies the condition (41) for 𝐂𝐂~(Ω(L))\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)). Now, the above relation can be concisely written as follows:

𝔔^(dg1,g2)+𝔔^(g1,dg2)=δ𝔪ˇJ,L(𝔔^(g1,g2))±(𝔮^(g1g2)𝔮^(g1)𝔮^(g2))\mathfrak{\hat{Q}}(dg_{1},g_{2})+\mathfrak{\hat{Q}}(g_{1},dg_{2})=\delta_{\check{\mathfrak{m}}^{J,L}}(\mathfrak{\hat{Q}}(g_{1},g_{2}))\pm\big{(}\hat{\mathfrak{q}}(g_{1}\boldsymbol{\ast}g_{2})-\hat{\mathfrak{q}}(g_{1})\cup\hat{\mathfrak{q}}(g_{2})\big{)}

Accordingly, the map [𝔮^][\hat{\mathfrak{q}}] in (50) satisfies that for g1,g2H(X)g_{1},g_{2}\in H^{*}(X), we have

[𝔮^](g1g2)=[𝔮^](g1)[𝔮^](g2)[\hat{\mathfrak{q}}](g_{1}\boldsymbol{\ast}g_{2})=[\hat{\mathfrak{q}}](g_{1})\cup[\hat{\mathfrak{q}}](g_{2})

Ultimately, it remains to show the unitality. Let 1XΩ0(X)\text{1}_{X}\in\Omega^{0}(X) be the constant and 𝔢\mathfrak{e} be given in (43). Then, we aim to show 𝔮^(1X)=𝔢\hat{\mathfrak{q}}(\text{1}_{X})=\mathfrak{e}. In fact, suppose S:=𝔮^(1X)k,β(h1,,hk)=𝔮1,k,β(1X;h1,,hk)S:=\hat{\mathfrak{q}}(\text{1}_{X})_{k,\beta}(h_{1},\dots,h_{k})=\mathfrak{q}_{1,k,\beta}(\text{1}_{X};h_{1},\dots,h_{k}) is nonzero. If β=0\beta=0, then using Proposition 4.12 (iv) implies that k=0k=0 and S=𝔮1,0,0(1X)=1LS=\mathfrak{q}_{1,0,0}(\text{1}_{X})=\text{1}_{L}. If β0\beta\neq 0, it is defined by the moduli space 1,k+1,β(J,L)\mathcal{M}_{1,k+1,\beta}(J,L). We consider the forgetful map 𝔣𝔬𝔯𝔤𝔢𝔱:1,k+1,β(J,L)k+1,β(J,L){\mathfrak{forget}}:\mathcal{M}_{1,k+1,\beta}(J,L)\to\mathcal{M}_{k+1,\beta}(J,L). To produce a nonzero term, the dimension of the fiber of 𝔣𝔬𝔯𝔤𝔢𝔱{\mathfrak{forget}} needs to be zero just as what we discussed before. But this is impossible when β0\beta\neq 0. See also [Fuk10]. ∎

5 Closed-open maps with quantum corrections

5.1 The maps Θ\Theta and \mathbb{P}

Recall that the moduli space system of JJ-holomorphic disks bounding LL gives rise to a chain-level AA_{\infty} algebra (Ω(L),𝔪ˇJ,L)(\Omega^{*}(L),\check{\mathfrak{m}}^{J,L}). A metric gg induces the so-called harmonic contraction for which applying the homological perturbation to 𝔪ˇJ,L\check{\mathfrak{m}}^{J,L} yields the minimal model AA_{\infty} algebra (H(L),𝔪g,J,L)(H^{*}(L),\mathfrak{m}^{g,J,L}) along with an AA_{\infty} homotopy equivalence 𝔦g,J,L:𝔪g,J,L𝔪ˇJ,L\mathfrak{i}^{g,J,L}:\mathfrak{m}^{g,J,L}\to\check{\mathfrak{m}}^{J,L} (see §2.3.1)

For simplicity, we set 𝔪ˇ=𝔪ˇJ,L\check{\mathfrak{m}}=\check{\mathfrak{m}}^{J,L}, 𝔪=𝔪g,J,L\mathfrak{m}=\mathfrak{m}^{g,J,L}, and 𝔦=𝔦g,J,L\mathfrak{i}=\mathfrak{i}^{g,J,L}. Note that they are contained in 𝒰𝒟\mathscr{UD}. The Whitehead theorem can be generalized to 𝒰𝒟\mathscr{UD}; namely, there is a ud-homotopy inverse 𝔦1\mathfrak{i}^{-1} of 𝔦\mathfrak{i} in the sense of Theorem 2.2. Concerning (40), we define the following map:

(53) Θ:𝐂𝐂~(Ω(L))𝐂𝐂~(H(L))φ(𝔦1{φ})𝔦\Theta:\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L))\to\operatorname{\mathbf{\widetilde{CC}}}(H^{*}(L))\qquad\varphi\mapsto(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}

We apologize for the lack of motivation, but it is truly just found out by trial-and-error. We may assume φ=(φk,β)k0,βπ2(X,L)\varphi=(\varphi_{k,\beta})_{k\geq 0,\beta\in\pi_{2}(X,L)} is homogeneously-graded, say |φ|=p|\varphi|=p. Then, we explicitly have

Θ(φ)k,β=λ,μ,ν0ri+ki+si=kβ+β′′+αi+βi+γi=β𝔦λ+μ+1,β1(𝔦r1,α1#p𝔦rλ,αλ#pφν,β′′(𝔦k1,β1𝔦kν,βν)𝔦s1,γ1𝔦sμ,γμ)\displaystyle\Theta(\varphi)_{k,\beta}=\displaystyle\sum_{\scalebox{0.68}{$\begin{subarray}{c}\lambda,\mu,\nu\geq 0\\ \sum r_{i}+\sum k_{i}+\sum s_{i}=k\\ \beta^{\prime}+\beta^{\prime\prime}+\sum\alpha_{i}+\sum\beta_{i}+\sum\gamma_{i}=\beta\end{subarray}$}}\mathfrak{i}^{-1}_{\lambda+\mu+1,\beta^{\prime}}\circ\Big{(}\mathfrak{i}^{\#p}_{r_{1},\alpha_{1}}\otimes\cdots\otimes\mathfrak{i}^{\#p}_{r_{\lambda},\alpha_{\lambda}}\otimes\varphi_{\nu,\beta^{\prime\prime}}(\mathfrak{i}_{k_{1},\beta_{1}}\otimes\cdots\otimes\mathfrak{i}_{k_{\nu},\beta_{\nu}})\otimes\mathfrak{i}_{s_{1},\gamma_{1}}\otimes\cdots\otimes\mathfrak{i}_{s_{\mu},\gamma_{\mu}}\Big{)}

Briefly, this reads Θ(φ)=𝔦1(𝔦#p𝔦#pφ(𝔦𝔦)𝔦𝔦)\Theta(\varphi)=\sum\mathfrak{i}^{-1}(\mathfrak{i}^{\#p}\cdots\mathfrak{i}^{\#p}\ \varphi(\mathfrak{i}\cdots\mathfrak{i})\ \mathfrak{i}\cdots\mathfrak{i}). One can check that Θ\Theta is ΛX\Lambda^{X}-linear (§4.2.6). Further, given φ𝐂𝐂~(Ω(L))\varphi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)), the image Θ(φ)\Theta(\varphi) satisfies (41) and so lies in 𝐂𝐂~(H(L))\operatorname{\mathbf{\widetilde{CC}}}(H^{*}(L)), since 𝔦\mathfrak{i} and 𝔦1\mathfrak{i}^{-1} are unital. The label degree is |Θ|=0|\Theta|=0, i.e. |Θ(φ)|=|φ||\Theta(\varphi)|=|\varphi|. Moreover, for the 𝔢\mathfrak{e} in (43), we have

(54) Θ(𝔢)=𝔢\Theta(\mathfrak{e})=\mathfrak{e}

as Θ(𝔢)0,0=𝔦1,01(𝔢0,0)=𝔦1,01(1)=1=𝔢0,0\Theta(\mathfrak{e})_{0,0}=\mathfrak{i}^{-1}_{1,0}(\mathfrak{e}_{0,0})=\mathfrak{i}^{-1}_{1,0}(\text{1})=\text{1}=\mathfrak{e}_{0,0}. (Here we slightly abuse the notation 𝔢\mathfrak{e}.)

As said in the introduction, it is generally incorrect that the Θ\Theta would induce a ring homomorphism between the two reduced Hochschild cohomologies HH~(Ω(L),𝔪ˇ)\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}}) and HH~(H(L),𝔪)\operatorname{\widetilde{HH}}(H^{*}(L),\mathfrak{m}), but it is not far from so. Heuristically, if we could pretend that 𝔦1𝔦=id\mathfrak{i}^{-1}\diamond\mathfrak{i}=\mathrm{id}, then the Θ\Theta would induce an honest ring homomorphism. But, 𝔦1𝔦\mathfrak{i}^{-1}\diamond\mathfrak{i} is only ud-homotopic to id\mathrm{id}, especially when the nontrivial Maslov-0 disks are allowed. Shortly, the above says that the gap for producing an honest ring homomorphism can be controlled by the ud-homotopy relations in 𝒰𝒟\mathscr{UD}. Anyway, we start with some useful computations:

Lemma 5.1.

Given φ𝐂𝐂~(Ω(L))\varphi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)), we have

Θ(δ𝔪ˇφ)=((𝔪𝔦1){φ})𝔦(1)|φ|Θ(φ){𝔪}\Theta(\delta_{\check{\mathfrak{m}}}\varphi)=\big{(}(\mathfrak{m}\diamond\mathfrak{i}^{-1})\{\varphi\}\big{)}\diamond\mathfrak{i}-(-1)^{|\varphi|}\Theta(\varphi)\{\mathfrak{m}\}

Given φ,ψ𝐂𝐂~(Ω(L))\varphi,\psi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)), we have

Θ(φ𝔪ˇψ)+𝔦1{δ𝔪ˇφ,ψ}𝔦+(1)|φ|𝔦1{φ,δ𝔪ˇψ}𝔦=((𝔪𝔦1){φ,ψ})𝔦(1)|φ|+|ψ|(𝔦1{φ,ψ}𝔦){𝔪}\Theta(\varphi\cup_{\check{\mathfrak{m}}}\psi)+\mathfrak{i}^{-1}\{\delta_{\check{\mathfrak{m}}}\varphi,\psi\}\diamond\mathfrak{i}+(-1)^{|\varphi|}\mathfrak{i}^{-1}\{\varphi,\delta_{\check{\mathfrak{m}}}\psi\}\diamond\mathfrak{i}=\big{(}(\mathfrak{m}\diamond\mathfrak{i}^{-1})\{\varphi,\psi\}\big{)}\diamond\mathfrak{i}-(-1)^{|\varphi|+|\psi|}\big{(}\mathfrak{i}^{-1}\{\varphi,\psi\}\diamond\mathfrak{i}\big{)}\{\mathfrak{m}\}

Proof.

(i) May assume φ\varphi is homogeneously-graded and |φ|=p|\varphi|=p. First, by Proposition 4.2, we obtain 𝔦1{δ𝔪ˇφ}=𝔦1{𝔪ˇ{φ}}(1)p𝔦1{φ{𝔪ˇ}}=𝔦1{𝔪ˇ}{φ}(1)p𝔦1{φ}{𝔪ˇ}\mathfrak{i}^{-1}\{\delta_{\check{\mathfrak{m}}}\varphi\}=\mathfrak{i}^{-1}\{\check{\mathfrak{m}}\{\varphi\}\}-(-1)^{p}\mathfrak{i}^{-1}\{\varphi\{\check{\mathfrak{m}}\}\}=\mathfrak{i}^{-1}\{\check{\mathfrak{m}}\}\{\varphi\}-(-1)^{p}\mathfrak{i}^{-1}\{\varphi\}\{\check{\mathfrak{m}}\}. Because 𝔦1\mathfrak{i}^{-1} is an AA_{\infty} homomorphism from 𝔪ˇ\check{\mathfrak{m}} to 𝔪\mathfrak{m}, we have 𝔪𝔦1=𝔦1{𝔪ˇ}\mathfrak{m}\diamond\mathfrak{i}^{-1}=\mathfrak{i}^{-1}\{\check{\mathfrak{m}}\}. Moreover, using the fact that 𝔦:𝔪𝔪ˇ\mathfrak{i}:\mathfrak{m}\to\check{\mathfrak{m}} is an AA_{\infty} homomorphism, we have (𝔦1{φ}{𝔪ˇ})𝔦=(𝔦1{φ}𝔦){𝔪}=Θ(φ){𝔪}\big{(}\mathfrak{i}^{-1}\{\varphi\}\{\check{\mathfrak{m}}\}\big{)}\diamond\mathfrak{i}=\big{(}\mathfrak{i}^{-1}\{\varphi\}\diamond\mathfrak{i}\big{)}\{\mathfrak{m}\}=\Theta(\varphi)\{\mathfrak{m}\}.

(ii) May assume φ\varphi and ψ\psi are homogeneously-graded and |φ|=p,|ψ|=q|\varphi|=p,|\psi|=q. By Proposition 4.2,

𝔦1{𝔪ˇ}{φ,ψ}(1)p+q𝔦1{φ,ψ}{𝔪ˇ}=𝔦1{δ𝔪ˇφ,ψ}+(1)p𝔦1{φ,δ𝔪ˇψ}+𝔦1{φ𝔪ˇψ}\mathfrak{i}^{-1}\{\check{\mathfrak{m}}\}\{\varphi,\psi\}-(-1)^{p+q}\mathfrak{i}^{-1}\{\varphi,\psi\}\{\check{\mathfrak{m}}\}=\mathfrak{i}^{-1}\{\delta_{\check{\mathfrak{m}}}\varphi,\psi\}+(-1)^{p}\mathfrak{i}^{-1}\{\varphi,\delta_{\check{\mathfrak{m}}}\psi\}+\mathfrak{i}^{-1}\{\varphi\cup_{\check{\mathfrak{m}}}\psi\}

Applying 𝔦-\diamond\mathfrak{i} on both sides, and using the AA_{\infty} equations of both 𝔦1\mathfrak{i}^{-1} and 𝔦\mathfrak{i}, we complete the proof. ∎

Finally, we define the length-zero projection map:

(55) :𝐂𝐂~(H(L))H(L)^Λ[[π1(L)]]/𝔞𝔣βTE(β)Yβ𝔣0,β\mathbb{P}:\operatorname{\mathbf{\widetilde{CC}}}(H^{*}(L))\to H^{*}(L)\hat{\otimes}\Lambda[[\pi_{1}(L)]]/{\mathfrak{a}}\qquad\mathfrak{f}\mapsto\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\mathfrak{f}_{0,\beta}

One can check that \mathbb{P} is (Λ,ΛX)(\Lambda,\Lambda^{X})-linear in the sense that (tA𝔣)=TE(A)(𝔣)\mathbb{P}(t^{A}\cdot\mathfrak{f})=T^{E(A)}\mathbb{P}(\mathfrak{f}). For (43), we have

(56) (𝔢)=T0Y0𝔢0,0=𝔢0,0=1\mathbb{P}(\mathfrak{e})=T^{0}Y^{0}\mathfrak{e}_{0,0}=\mathfrak{e}_{0,0}=\text{1}

5.2 From reduced Hochschild cohomology to self Floer cohomology

There is really no hope to cook up a unital ring homomorphism from HH~(Ω(L),𝔪ˇ)\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}}) to HH~(H(L),𝔪)\operatorname{\widetilde{HH}}(H^{*}(L),\mathfrak{m}). Notwithstanding, the following result will be sufficient for our purpose:

Theorem 5.2.

The composite map Θ\mathbb{P}\Theta induces a unital ring homomorphism

Φ:=[Θ]:HH~(Ω(L),𝔪ˇ)HF(L,𝔪)\Phi:=[\mathbb{P}\Theta]:\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}})\to\operatorname{HF}(L,\mathfrak{m})
Proof.

We notice Θ(𝔢)=1\mathbb{P}\Theta(\mathfrak{e})=\text{1} by (54) and (56). The following two items are what we need:

(a). Θ(δ𝔪ˇφ)=𝐦1(Θ(φ))\mathbb{P}\Theta(\delta_{\check{\mathfrak{m}}}\varphi)=\mathbf{m}_{1}\big{(}\mathbb{P}\Theta(\varphi)\big{)}

(b). If φ,ψ𝐂𝐂~(Ω(L))\varphi,\psi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)) are δ𝔪ˇ\delta_{\check{\mathfrak{m}}}-closed, then

Θ(φ𝔪ˇψ)=𝐦2(Θ(φ),Θ(ψ))+𝐦1(((𝔦1{φ,ψ})𝔦))\mathbb{P}\Theta(\varphi\cup_{\check{\mathfrak{m}}}\psi)=\mathbf{m}_{2}\big{(}\mathbb{P}\Theta(\varphi),\mathbb{P}\Theta(\psi)\big{)}+\mathbf{m}_{1}\big{(}\mathbb{P}((\mathfrak{i}^{-1}\{\varphi,\psi\})\diamond\mathfrak{i})\big{)}

First, may assume φ𝐂𝐂~(Ω(L))\varphi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)) is homogeneously-graded and |φ|=p|\varphi|=p. Concerning the first half of Lemma 5.1, we compute that ((𝔪𝔦1){φ})𝔦)\mathbb{P}\left((\mathfrak{m}\diamond\mathfrak{i}^{-1})\{\varphi\})\diamond\mathfrak{i}\right) equals to

λ,μ0;β,β′′,α1,,αλ,γ1,,γμTE(β)Yβ𝔪λ+μ+1,β(TE(α1)Yα1(𝔦1𝔦)0,α1#p,,TE(αλ)Yαλ(𝔦1𝔦)0,αλ#p,\displaystyle\sum_{\scalebox{0.77}{$\begin{subarray}{c}\lambda,\mu\geq 0\ ;\ \beta^{\prime},\beta^{\prime\prime},\alpha_{1},\dots,\alpha_{\lambda},\gamma_{1},\dots,\gamma_{\mu}\end{subarray}$}}T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\mathfrak{m}_{\lambda+\mu+1,\beta^{\prime}}\Big{(}T^{E(\alpha_{1})}Y^{\partial\alpha_{1}}(\mathfrak{i}^{-1}\diamond\mathfrak{i})^{\#p}_{0,\alpha_{1}},\dots,T^{E(\alpha_{\lambda})}Y^{\partial\alpha_{\lambda}}(\mathfrak{i}^{-1}\diamond\mathfrak{i})^{\#p}_{0,\alpha_{\lambda}},
TE(β′′)Yβ′′((𝔦1{φ})𝔦)0,β′′,TE(γ1)Yγ1(𝔦1𝔦)0,γ1,,TE(γμ)Yγμ(𝔦1𝔦)0,γμ)\displaystyle T^{E(\beta^{\prime\prime})}Y^{\partial\beta^{\prime\prime}}\big{(}(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}\big{)}_{0,\beta^{\prime\prime}},T^{E(\gamma_{1})}Y^{\partial\gamma_{1}}(\mathfrak{i}^{-1}\diamond\mathfrak{i})_{0,\gamma_{1}},\dots,T^{E(\gamma_{\mu})}Y^{\partial\gamma_{\mu}}(\mathfrak{i}^{-1}\diamond\mathfrak{i})_{0,\gamma_{\mu}}\Big{)}

Moreover, since 𝔦1𝔦udid\mathfrak{i}^{-1}\diamond\mathfrak{i}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id}, we know αTE(α)Yα(𝔦1𝔦)0,ααTE(α)Yα(id)0,α0(mod𝔞)\sum_{\alpha}T^{E(\alpha)}Y^{\partial\alpha}(\mathfrak{i}^{-1}\diamond\mathfrak{i})_{0,\alpha}\equiv\sum_{\alpha}T^{E(\alpha)}Y^{\partial\alpha}(\mathrm{id})_{0,\alpha}\equiv 0\ (\mathrm{mod}\ {\mathfrak{a}}) due to Lemma 2.4. Accordingly, those terms with λ0\lambda\neq 0 or μ0\mu\neq 0 all vanish modulo 𝔞{\mathfrak{a}}. So, we have:

((𝔪𝔦1){φ})𝔦)=β+β′′=βTE(β)Yβ𝔪1,β(TE(β′′)Yβ′′((𝔦1{φ})𝔦)0,β′′)=𝐦1(Θ(φ))\mathbb{P}\left(\big{(}\mathfrak{m}\diamond\mathfrak{i}^{-1})\{\varphi\}\big{)}\diamond\mathfrak{i}\right)=\sum_{\beta^{\prime}+\beta^{\prime\prime}=\beta}T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\mathfrak{m}_{1,\beta^{\prime}}\big{(}T^{E(\beta^{\prime\prime})}Y^{\partial\beta^{\prime\prime}}\big{(}(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}\big{)}_{0,\beta^{\prime\prime}}\big{)}=\mathbf{m}_{1}(\mathbb{P}\Theta(\varphi))

By the first half of Lemma 5.1, it remains to prove

(Θ(φ){𝔪})=TE(β)Yβ(Θ(φ))1,β(TE(β′′)Yβ′′𝔪0,β′′)0(mod𝔞)\mathbb{P}(\Theta(\varphi)\{\mathfrak{m}\})=\sum T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\big{(}\Theta(\varphi)\big{)}_{1,\beta^{\prime}}\big{(}T^{E(\beta^{\prime\prime})}Y^{\partial\beta^{\prime\prime}}\mathfrak{m}_{0,\beta^{\prime\prime}}\big{)}\equiv 0\ (\mathrm{mod}\ {\mathfrak{a}})

Indeed, we first note that TE(β′′)Yβ′′𝔪0,β′′=W1\sum T^{E(\beta^{\prime\prime})}Y^{\partial\beta^{\prime\prime}}\mathfrak{m}_{0,\beta^{\prime\prime}}=W\cdot\text{1} modulo 𝔞{\mathfrak{a}}. Besides, all of 𝔦1,φ,𝔦\mathfrak{i}^{-1},\varphi,\mathfrak{i} satisfy the condition (41) except 𝔦1,0(1)=1\mathfrak{i}_{1,0}(\text{1})=\text{1} and 𝔦1,01(1)=1\mathfrak{i}^{-1}_{1,0}(\text{1})=\text{1}, so any term in the expansion of Θ(φ)1,β(1)((𝔦1{φ})𝔦)1,β(1)\Theta(\varphi)_{1,\beta^{\prime}}(\text{1})\equiv\big{(}(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}\big{)}_{1,\beta^{\prime}}(\text{1}) must vanish. This completes the proof of (a).

Second, we assume φ,ψ𝐂𝐂~(Ω(L))\varphi,\psi\in\operatorname{\mathbf{\widetilde{CC}}}(\Omega^{*}(L)) are δ𝔪ˇ\delta_{\check{\mathfrak{m}}}-closed and homogeneously-graded, say |φ|=p|\varphi|=p and |ψ|=q|\psi|=q. For the second half of Lemma 5.1, we compute that (((𝔪𝔦1){φ,ψ})𝔦)\mathbb{P}\left(((\mathfrak{m}\diamond\mathfrak{i}^{-1})\{\varphi,\psi\})\diamond\mathfrak{i}\right) equals to

(𝔪(,(𝔦1𝔦)#(p+q),,((𝔦1{φ})𝔦)#q,,(𝔦1𝔦)#q,,(𝔦1{ψ})𝔦,,𝔦1𝔦,))\displaystyle\mathbb{P}\Big{(}\mathfrak{m}\big{(}\dots,(\mathfrak{i}^{-1}\diamond\mathfrak{i})^{\#(p+q)},\dots,((\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i})^{\#q},\dots,(\mathfrak{i}^{-1}\diamond\mathfrak{i})^{\#q},\dots,(\mathfrak{i}^{-1}\{\psi\})\diamond\mathfrak{i},\dots,\mathfrak{i}^{-1}\diamond\mathfrak{i},\dots\big{)}\Big{)}
+\displaystyle+ (𝔪(,(𝔦1𝔦)#(p+q),,(𝔦1{φ,ψ})𝔦,,𝔦1𝔦,))\displaystyle\mathbb{P}\Big{(}\mathfrak{m}\big{(}\dots,(\mathfrak{i}^{-1}\diamond\mathfrak{i})^{\#(p+q)},\dots,(\mathfrak{i}^{-1}\{\varphi,\psi\})\diamond\mathfrak{i},\dots,\mathfrak{i}^{-1}\diamond\mathfrak{i},\dots\big{)}\Big{)}

Just as above, by applying Lemma 2.4 to 𝔦1𝔦udid\mathfrak{i}^{-1}\diamond\mathfrak{i}\stackrel{{\scriptstyle\mathrm{ud}}}{{\sim}}\mathrm{id}, we know it further equals to (modulo 𝔞{\mathfrak{a}})

β0,β1,β2TE(β0)Yβ0𝔪2,β0(TE(β1)Yβ1((𝔦1{φ})𝔦)0,β1,TE(β2)Yβ2((𝔦1{ψ})𝔦)0,β2)\displaystyle\textstyle\sum_{\beta_{0},\beta_{1},\beta_{2}}T^{E(\beta_{0})}Y^{\partial\beta_{0}}\mathfrak{m}_{2,\beta_{0}}\Big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}\big{(}(\mathfrak{i}^{-1}\{\varphi\})\diamond\mathfrak{i}\big{)}_{0,\beta_{1}},T^{E(\beta_{2})}Y^{\partial\beta_{2}}\big{(}(\mathfrak{i}^{-1}\{\psi\})\diamond\mathfrak{i}\big{)}_{0,\beta_{2}}\Big{)}
+\displaystyle+ β0,β1TE(β0)Yβ0𝔪1,β0(TE(β1)Yβ1(𝔦1{φ,ψ}𝔦)0,β1)=𝐦2(Θ(φ),Θ(ψ))+𝐦1(((𝔦1{φ,ψ})𝔦))\displaystyle\textstyle\sum_{\beta_{0},\beta_{1}}T^{E(\beta_{0})}Y^{\partial\beta_{0}}\mathfrak{m}_{1,\beta_{0}}\Big{(}T^{E(\beta_{1})}Y^{\partial\beta_{1}}\big{(}\mathfrak{i}^{-1}\{\varphi,\psi\}\diamond\mathfrak{i}\big{)}_{0,\beta_{1}}\Big{)}=\mathbf{m}_{2}\big{(}\mathbb{P}\Theta(\varphi),\mathbb{P}\Theta(\psi)\big{)}+\mathbf{m}_{1}\big{(}\mathbb{P}\big{(}(\mathfrak{i}^{-1}\{\varphi,\psi\})\diamond\mathfrak{i}\big{)}\big{)}

By Lemma 5.1, since δ𝔪ˇφ=δ𝔪ˇψ=0\delta_{\check{\mathfrak{m}}}\varphi=\delta_{\check{\mathfrak{m}}}\psi=0, it remains to prove that ((𝔦1{φ,ψ}𝔦){𝔪})0(mod𝔞)\mathbb{P}\big{(}(\mathfrak{i}^{-1}\{\varphi,\psi\}\diamond\mathfrak{i})\{\mathfrak{m}\}\big{)}\equiv 0\ (\mathrm{mod}\ {\mathfrak{a}}). Indeed, it can be proved by almost the same way as in part (a): first, TE(α)Yα𝔪0,α=W1(mod𝔞)\sum T^{E(\alpha)}Y^{\partial\alpha}\mathfrak{m}_{0,\alpha}=W\cdot\text{1}\ (\mathrm{mod}\ {\mathfrak{a}}); then, just use the conditions (41) of 𝔦1,φ,ψ,𝔦\mathfrak{i}^{-1},\varphi,\psi,\mathfrak{i}. Putting things together, the proof is now complete. ∎

5.3 Closed-open maps with Maslov-0 disks

By Theorem 4.13 and Theorem 5.2, the composition of [𝔮^][\hat{\mathfrak{q}}] and Φ[Θ]\Phi\equiv[\mathbb{P}\Theta]:

Ψ:\textstyle{\Psi:}QH(X;ΛX)\textstyle{QH^{*}(X;\Lambda^{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[𝔮^]\scriptstyle{[\hat{\mathfrak{q}}]\qquad}HH~(Ω(L),𝔪ˇ)\textstyle{\operatorname{\widetilde{HH}}(\Omega^{*}(L),\check{\mathfrak{m}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ=[Θ]\scriptstyle{\Phi=[\mathbb{P}\Theta]}HF(L,𝔪)\textstyle{\operatorname{HF}(L,\mathfrak{m})}

is a unital ring homomorphism. Recall that we have defined two versions of quantum cohomologies QH(X;ΛX)=H(X)^ΛXQH^{*}(X;\Lambda^{X})=H^{*}(X)\hat{\otimes}\Lambda^{X} and QH(X;Λ)=H(X)^ΛQH^{*}(X;\Lambda)=H^{*}(X)\hat{\otimes}\Lambda in (33). As HF(L,𝔪)\operatorname{HF}(L,\mathfrak{m}) is a Λ\Lambda-algebra, we can consider the Λ\Lambda-linear extension of the restriction of Ψ=Φ[𝔮^]\Psi=\Phi\circ[\hat{\mathfrak{q}}] on H(X;)H^{*}(X;\mathbb{R}), which gives a map:

(57) 𝕆:QH(X;Λ)HF(L,𝔪)\operatorname{{\mathbb{CO}}}:\quad QH^{*}(X;\Lambda)\to\operatorname{HF}(L,\mathfrak{m})
Proposition 5.3.

The 𝕆\operatorname{{\mathbb{CO}}} is a unital Λ\Lambda-algebra homomorphism and 𝕆(1X)=[1]\operatorname{{\mathbb{CO}}}(\text{1}_{X})=[\text{1}].

Proof.

Denote the quantum products in QH(X;ΛX)QH^{*}(X;\Lambda^{X}) and QH(X;Λ)QH^{*}(X;\Lambda) by X\boldsymbol{\ast}^{X} and \boldsymbol{\ast} respectively for a moment to distinguish. Recall that Θ\Theta and 𝔮^\hat{\mathfrak{q}} are ΛX\Lambda^{X}-linear, and \mathbb{P} is (Λ,ΛX)(\Lambda,\Lambda^{X})-linear. So, by construction, Ψ\Psi is also (Λ,ΛX)(\Lambda,\Lambda^{X})-linear, that is, Ψ(tAg)=TE(A)Ψ(g)\Psi(t^{A}\cdot g)=T^{E(A)}\Psi(g). Since 𝕆\operatorname{{\mathbb{CO}}} is the Λ\Lambda-linear extension, we see that Ψ(gXh)=𝕆(gh)\Psi(g{\boldsymbol{\ast}^{X}}h)=\operatorname{{\mathbb{CO}}}(g\boldsymbol{\ast}h) for any g,hH(X;)g,h\in H^{*}(X;\mathbb{R}). Let 𝒈,𝒉QH(X;Λ)H(X)^Λ\boldsymbol{g},\boldsymbol{h}\in QH^{*}(X;\Lambda)\equiv H^{*}(X)\hat{\otimes}\Lambda, and write 𝒈=i=1Tλigi\boldsymbol{g}=\sum_{i=1}^{\infty}T^{\lambda_{i}}g_{i} and 𝒉=j=1Tμjhj\boldsymbol{h}=\sum_{j=1}^{\infty}T^{\mu_{j}}h_{j}, where λi\lambda_{i}\nearrow\infty, μj\mu_{j}\nearrow\infty, and gi,hjH(X;)g_{i},h_{j}\in H^{*}(X;\mathbb{R}). Now, we have 𝕆(𝒈𝒉)=i,j1Tλi+μj𝕆(gihj)=i,j1Tλi+μjΨ(giXhj)=i,j1Tλi+μjΨ(gi)Ψ(hj)=𝕆(𝒈)𝕆(𝒉)\operatorname{{\mathbb{CO}}}(\boldsymbol{g}\boldsymbol{\ast}\boldsymbol{h})=\sum_{i,j\geq 1}T^{\lambda_{i}+\mu_{j}}\operatorname{{\mathbb{CO}}}(g_{i}\boldsymbol{\ast}h_{j})=\sum_{i,j\geq 1}T^{\lambda_{i}+\mu_{j}}\Psi(g_{i}{\boldsymbol{\ast}^{X}}h_{j})=\sum_{i,j\geq 1}T^{\lambda_{i}+\mu_{j}}\Psi(g_{i})\cdot\Psi(h_{j})=\operatorname{{\mathbb{CO}}}(\boldsymbol{g})\cdot\operatorname{{\mathbb{CO}}}(\boldsymbol{h}). Finally, 𝕆(1X)=Ψ(1X)=Φ[𝔮^](1X)=Φ(𝔢)=[1]\operatorname{{\mathbb{CO}}}(\text{1}_{X})=\Psi(\text{1}_{X})=\Phi\circ[\hat{\mathfrak{q}}](\text{1}_{X})=\Phi(\mathfrak{e})=[\text{1}]. ∎

On the other hand, just like the Gromov-Witten theory, the operator 𝔮\mathfrak{q} also satisfies the divisor axiom for the interior marked points. Specifically, if g1Z2(X)g_{1}\in Z^{2}(X) is a closed 2-form supported in XLX\setminus L, then

(58) 𝔮,k,β(g1,g2,,g;h1,,hk)=βg1𝔮1,k,β(g2,,g;h1,,hk)\mathfrak{q}_{\ell,k,\beta}(g_{1},g_{2},\dots,g_{\ell};h_{1},\dots,h_{k})=\textstyle\int_{\beta}g_{1}\cdot\mathfrak{q}_{\ell-1,k,\beta}(g_{2},\dots,g_{\ell};h_{1},\dots,h_{k})

Denote by c1=c1(X)H2(X)c_{1}=c_{1}(X)\in H^{2}(X) the first Chern class of XX. Note that there exists a cycle QQ in XLX\setminus L with βQ=μ(β)\beta\cap Q=\mu(\beta) and [Q][Q] is Poincare dual to 2c12c_{1} [FOOO19, §23.3]. Together with Proposition 4.12 (iii), we obtain

(59) (𝔮^(c1))k,β=μ(β)2𝔪ˇk,β\big{(}\hat{\mathfrak{q}}(c_{1})\big{)}_{k,\beta}=\tfrac{\mu(\beta)}{2}\cdot\check{\mathfrak{m}}_{k,\beta}
Proposition 5.4.

𝕆(c1)=W[1]\operatorname{{\mathbb{CO}}}(c_{1})=W\cdot[\text{1}].

Proof.

First, we compute in the cochain level:

Θ𝔮^(c1)\displaystyle\mathbb{P}\circ\Theta\circ\hat{\mathfrak{q}}(c_{1}) =((𝔦1{𝔮^(c1)})𝔦)=(𝔦1(𝔦#,,𝔦#,𝔮^(c1)𝔦,𝔦,,𝔦))\displaystyle=\mathbb{P}\Big{(}(\mathfrak{i}^{-1}\{\hat{\mathfrak{q}}(c_{1})\})\diamond\mathfrak{i}\Big{)}=\mathbb{P}\Big{(}\mathfrak{i}^{-1}\big{(}\mathfrak{i}^{\#},\dots,\mathfrak{i}^{\#},\ \hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i},\ \mathfrak{i},\dots,\mathfrak{i}\big{)}\Big{)}
=s,t0,𝖡π2(X,L)α+β+γi+ηi=𝖡TE(𝖡)Y𝖡𝔦s+t+1,α1(𝔦0,γ1#,,𝔦0,γs#,(𝔮^(c1)𝔦)0,β,𝔦0,η1,,𝔦0,ηt)\displaystyle=\sum_{s,t\geq 0,\mathsf{B}\in\pi_{2}(X,L)}\ \sum_{\alpha+\beta+\sum\gamma_{i}+\sum\eta_{i}=\mathsf{B}}T^{E(\mathsf{B})}Y^{\partial\mathsf{B}}\cdot\mathfrak{i}^{-1}_{s+t+1,\alpha}\big{(}\mathfrak{i}^{\#}_{0,\gamma_{1}},\dots,\mathfrak{i}^{\#}_{0,\gamma_{s}},\ \big{(}\hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i}\big{)}_{0,\beta},\ \mathfrak{i}_{0,\eta_{1}},\dots,\mathfrak{i}_{0,\eta_{t}}\big{)}

Since 𝔦0,γΩ1μ(γ)(L)\mathfrak{i}_{0,\gamma}\in\Omega^{1-\mu(\gamma)}(L), the semipositive condition implies that μ(γ)=0\mu(\gamma)=0 whenever 𝔦0,γ0\mathfrak{i}_{0,\gamma}\neq 0. By this observation and by (59), the term

(𝔮^(c1)𝔦)0,β=β0++β=β𝔮^(c1),β0(𝔦0,β1,,𝔦0,β)=β0++β=βμ(β0)2𝔪ˇ,β0(𝔦0,β1,,𝔦0,β)\big{(}\hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i}\big{)}_{0,\beta}=\sum_{\beta_{0}+\cdots+\beta_{\ell}=\beta}\hat{\mathfrak{q}}(c_{1})_{\ell,\beta_{0}}(\mathfrak{i}_{0,\beta_{1}},\dots,\mathfrak{i}_{0,\beta_{\ell}})=\sum_{\beta_{0}+\cdots+\beta_{\ell}=\beta}\tfrac{\mu(\beta_{0})}{2}\cdot\check{\mathfrak{m}}_{\ell,\beta_{0}}(\mathfrak{i}_{0,\beta_{1}},\dots,\mathfrak{i}_{0,\beta_{\ell}})

is contained in Ω2μ(β0)(L)\Omega^{2-\mu(\beta_{0})}(L). By the semipositive condition again, we may assume μ(β0)=0\mu(\beta_{0})=0 or 22. Hence, whenever it is nonzero, we may always assume μ(β0)=2\mu(\beta_{0})=2 and μ(βi)=0\mu(\beta_{i})=0 for other 1i1\leq i\leq\ell, in particular, μ(β)=2\mu(\beta)=2. Then, using the AA_{\infty} associativity relation 𝔪ˇ𝔦=𝔦{𝔪}\check{\mathfrak{m}}\diamond\mathfrak{i}=\mathfrak{i}\{\mathfrak{m}\} yields that

(𝔮^(c1)𝔦)0,β=μ(β0)=2𝔪ˇ,β0(𝔦0,β1,,𝔦0,β)=(𝔪ˇ,β0(𝔦0,β1,,𝔦0,β))|Ω0(L)=(𝔦1,β(𝔪0,β′′))|Ω0(L)\displaystyle\big{(}\hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i}\big{)}_{0,\beta}=\sum_{\mu(\beta_{0})=2}\check{\mathfrak{m}}_{\ell,\beta_{0}}(\mathfrak{i}_{0,\beta_{1}},\dots,\mathfrak{i}_{0,\beta_{\ell}})=\Big{(}\sum\check{\mathfrak{m}}_{\ell,\beta_{0}}(\mathfrak{i}_{0,\beta_{1}},\dots,\mathfrak{i}_{0,\beta_{\ell}})\Big{)}\Big{|}_{\Omega^{0}(L)}=\Big{(}\sum\mathfrak{i}_{1,\beta^{\prime}}(\mathfrak{m}_{0,\beta^{\prime\prime}})\Big{)}\Big{|}_{\Omega^{0}(L)}

Notice that μ(β)+μ(β′′)=μ(β)=2\mu(\beta^{\prime})+\mu(\beta^{\prime\prime})=\mu(\beta)=2. Thus,

βTE(β)Yβ(𝔮^(c1)𝔦)0,β\displaystyle\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\big{(}\hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i}\big{)}_{0,\beta} =μ(β)=2TE(β)Yβ(𝔮^(c1)𝔦)0,β\displaystyle=\sum_{\mu(\beta)=2}T^{E(\beta)}Y^{\partial\beta}\big{(}\hat{\mathfrak{q}}(c_{1})\diamond\mathfrak{i}\big{)}_{0,\beta}
=μ(β)=0TE(β)Yβ𝔦1,β(μ(β′′)=2TE(β′′)Yβ′′𝔪0,β′′)=𝔦1,0(W1)=W1(mod𝔞)\displaystyle=\sum_{\mu(\beta^{\prime})=0}T^{E(\beta^{\prime})}Y^{\partial\beta^{\prime}}\mathfrak{i}_{1,\beta^{\prime}}\big{(}\sum_{\mu(\beta^{\prime\prime})=2}T^{E(\beta^{\prime\prime})}Y^{\partial\beta^{\prime\prime}}\mathfrak{m}_{0,\beta^{\prime\prime}}\big{)}\ =\mathfrak{i}_{1,0}(W\cdot\text{1})=W\cdot\text{1}\quad(\mathrm{mod}\ {\mathfrak{a}})

where the terms with μ(β′′)=0\mu(\beta^{\prime\prime})=0 are killed modulo 𝔞{\mathfrak{a}}. Beware that we abuse the notation 1 to represent the constant-one functions in both Ω0(L)\Omega^{0}(L) and H0(L)H^{0}(L). Finally, back to the calculation at the start, using the unitality of 𝔦1\mathfrak{i}^{-1} further deduces that s=t=0s=t=0 and α=0\alpha=0 there. To conclude,

Θ𝔮^(c1)=𝔦1,01(W1)=W1(mod𝔞)\mathbb{P}\circ\Theta\circ\hat{\mathfrak{q}}(c_{1})=\mathfrak{i}^{-1}_{1,0}(W\cdot\text{1})=W\cdot\text{1}\qquad(\mathrm{mod}\ {\mathfrak{a}})

Passing to the cohomology, we get 𝕆(c1)=Ψ(c1)=W[1]\operatorname{{\mathbb{CO}}}(c_{1})=\Psi(c_{1})=W\cdot[\text{1}]. The proof is now complete. ∎

Suppose 𝐲H1(L;UΛ)\mathbf{y}\in H^{1}(L;U_{\Lambda}) can lift to a weak bounding cochain bH1(L;Λ0)b\in H^{1}(L;\Lambda_{0}); or equivalently, suppose the ideal 𝔞{\mathfrak{a}} vanishes at 𝐲\mathbf{y}, i.e. Q(𝐲)=0Q(\mathbf{y})=0. (Indeed, up to Fukaya’s trick, we may even allow 𝐲\mathbf{y} to lie in a neighborhood of H1(L;UΛ)H^{1}(L;U_{\Lambda}) in H1(L;Λ)H^{1}(L;\Lambda^{*}).) By Lemma 3.11, we get a unital Λ\Lambda-algebra homomorphism 𝐲:HF(L,𝔪)HF(L,𝔪,𝐲)\mathcal{E}_{\mathbf{y}}:\operatorname{HF}(L,\mathfrak{m})\to\operatorname{HF}(L,\mathfrak{m},\mathbf{y}). Accordingly, the composition

(60) 𝕆𝐲:=𝐲𝕆:\textstyle{\operatorname{{\mathbb{CO}}}_{\mathbf{y}}:=\mathcal{E}_{\mathbf{y}}\circ\operatorname{{\mathbb{CO}}}:}QH(X;Λ)\textstyle{QH^{*}(X;\Lambda)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝕆\scriptstyle{\operatorname{{\mathbb{CO}}}}HF(L,𝔪)\textstyle{\operatorname{HF}(L,\mathfrak{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐲\scriptstyle{\mathcal{E}_{\mathbf{y}}}HF(L,𝔪,𝐲)\textstyle{\operatorname{HF}(L,\mathfrak{m},\mathbf{y})}

is also a unital Λ\Lambda-algebra homomorphism. In particular,

(61) 𝕆𝐲(1X)=[1]\mathbb{CO}_{\mathbf{y}}(\text{1}_{X})=[\text{1}]

Moreover, by Proposition 5.4, we also have

(62) 𝕆𝐲(c1)=W(𝐲)[1]\mathbb{CO}_{\mathbf{y}}(c_{1})=W(\mathbf{y})\cdot[\text{1}]

After all the above preparations, there is no further obstacle for the proof of our local result:

Theorem 5.5 (Theorem A).

Suppose the cohomology ring H(L)H^{*}(L) is generated by H1(L;)H^{1}(L;\mathbb{Z}). Under Assumption 1.3, any critical value of WW is an eigenvalue of c1:QH(X;Λ)QH(X;Λ)c_{1}\boldsymbol{\ast}:QH^{*}(X;\Lambda)\to QH^{*}(X;\Lambda).

Proof.

Let 𝐲\mathbf{y} be a critical point of WW in the sense of Definition 3.15. We aim to show W(𝐲)ΛW(\mathbf{y})\in\Lambda is an eigenvalue of c1c_{1}\boldsymbol{\ast}. Arguing by contradiction, suppose a:=c1W(𝐲)1Xa:=c_{1}-W(\mathbf{y})\text{1}_{X} was invertible in QH(X;Λ)QH^{*}(X;\Lambda). Then, there would exist a1a^{-1} with aa1=1Xa{\boldsymbol{\ast}}a^{-1}=\text{1}_{X}. Notice that 𝕆𝐲(a)=𝕆𝐲(c1)W(𝐲)𝕆𝐲(1X)=0\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(a)=\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(c_{1})-W(\mathbf{y})\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(\text{1}_{X})=0 by (61) and (62). Since 𝕆𝐲\mathbb{CO}_{\mathbf{y}} is a unital Λ\Lambda-algebra homomorphism, we conclude that

[1]=𝕆𝐲(1X)=𝕆𝐲(aa1)=𝕆𝐲(a)𝕆𝐲(a1)=0\displaystyle[\text{1}]=\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(\text{1}_{X})=\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(a{\boldsymbol{\ast}}a^{-1})=\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(a)\cdot\operatorname{{\mathbb{CO}}}_{\mathbf{y}}(a^{-1})=0

Namely, we must have HF(L,𝔪,𝐲)=0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})=0. But, due to Theorem 3.19, the 𝐲\mathbf{y} being a critical point of WW implies that HF(L,𝔪,𝐲)0\operatorname{HF}(L,\mathfrak{m},\mathbf{y})\neq 0. This is a contradiction. ∎

5.4 Global result and the family Floer program

Lastly, we want to adapt Theorem 5.5 to the non-archimedean SYZ picture, showing Conjecture 1.1 for the analytic mirror Landau-Ginzburg model in Theorem 1.5. That is, we aim to show Theorem B.

5.4.1 General aspects

A brief review of Theorem 1.5 is as follows. First, using the moduli space of holomorphic disks and the homological perturbation, we obtain an AA_{\infty} algebra (H(L),𝔪)(H^{*}(L),\mathfrak{m}) in Obj𝒰𝒟(L)\operatorname{Obj}\mathscr{UD}(L) for each Lagrangian torus fiber L=Lq=π1(q)L=L_{q}=\pi^{-1}(q). Take a small rational polyhedron Δ\Delta in B0B_{0} near qq, and we can identify Δ\Delta with a subset in n\mathbb{R}^{n} via integral affine coordinates. Define the WW and 𝔞{\mathfrak{a}} for 𝔪\mathfrak{m} as in §2.2, but now the ideal 𝔞{\mathfrak{a}} vanishes under Assumption 1.3. Now, a local chart of XX^{\vee} is identified with a polyhedral affinoid domain 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta) for the fibration map 𝔱𝔯𝔬𝔭\operatorname{\mathfrak{trop}} in (3). From the rigid analytic geometry, we know 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta) can be recognized as the spectrum SpΛΔ\operatorname{Sp}\Lambda\langle\Delta\rangle of maximal ideals of the following affinoid algebra contained in Λ[[π1(L)]]Λ[[Y1±,,Yn±]]\Lambda[[\pi_{1}(L)]]\cong\Lambda[[Y_{1}^{\pm},\dots,Y_{n}^{\pm}]]:

(63) ΛΔ={απ1(L)cαYα𝗏(cα)+α,v,for anyvΔ}\Lambda\langle\Delta\rangle=\Big{\{}\textstyle\sum_{\alpha\in\pi_{1}(L)}c_{\alpha}Y^{\alpha}\mid\operatorname{\mathsf{v}}(c_{\alpha})+\langle\alpha,v\rangle\to\infty,\,\,\text{for any}\ v\in\Delta\Big{\}}

Recall that ΛΔ\Lambda\langle\Delta\rangle consists of all those formal power series in Λ[[π1(L)]]\Lambda[[\pi_{1}(L)]] which converge on 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta). Using the reverse isoperimetric inequalities [GS14] and shrinking Δ\Delta if necessary, we may require the 𝐦kβTE(β)Yβ𝔪k,β\mathbf{m}_{k}\equiv\sum_{\beta}T^{E(\beta)}Y^{\partial\beta}\mathfrak{m}_{k,\beta} maps H(L)^ΛΔH^{*}(L)\hat{\otimes}\Lambda\langle\Delta\rangle to H(L)^ΛΔH^{*}(L)\hat{\otimes}\Lambda\langle\Delta\rangle in (11). In particular, the W𝐦0W\equiv\mathbf{m}_{0} lies in ΛΔ\Lambda\langle\Delta\rangle or equivalently converges on 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta).

The affinoid space 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta) embeds into the mirror analytic space XX^{\vee} as a local chart. In reality, the transition maps ϕ\phi among all these local charts are given by restricting the homomorphisms in (9) to some affinoid algebra like (63). Eventually, by [Yua20], the various local charts glue to the mirror rigid analytic space XX^{\vee}.

5.4.2 Proof of the global result

Now, we are ready to prove Theorem B. Let λ:=W(𝐲)\lambda:=W^{\vee}(\mathbf{y}) be a critical value of the mirror Landau-Ginzburg superpotential WW^{\vee} at a point 𝐲\mathbf{y} in the mirror analytic space XX^{\vee}. In view of Corollary 3.16, we may choose a specific local chart of XX^{\vee} to present the critical point 𝐲\mathbf{y}. Actually, since X=qH1(Lq;UΛ)X^{\vee}=\bigcup_{q}H^{1}(L_{q};U_{\Lambda}) as a set, there is a Lagrangian fiber L=LqL=L_{q} so that its dual fiber H1(L;UΛ)H^{1}(L;U_{\Lambda}) contains the point 𝐲\mathbf{y}. Remark that one may also take any other adjacent Lagrangian fiber LL^{\prime} thanks to the Fukaya’s trick. Next, we find a nearby local chart which corresponds to an AA_{\infty} algebra (H(L),𝔪)Obj𝒰𝒟(H^{*}(L),\mathfrak{m})\in\operatorname{Obj}\mathscr{UD} associated to LL. As said before, choosing integral affine coordinates near qq, this local chart can be identified with a polyhedral affinoid domain 𝔱𝔯𝔬𝔭1(Δ)\operatorname{\mathfrak{trop}}^{-1}(\Delta), and the dual fiber H1(L;UΛ)H^{1}(L;U_{\Lambda}) is then identified with the central fiber 𝔱𝔯𝔬𝔭1(0)UΛn\operatorname{\mathfrak{trop}}^{-1}(0)\equiv U_{\Lambda}^{n}. The critical point can be also viewed as a point in 𝔱𝔯𝔬𝔭1(Δ)(Λ)n\operatorname{\mathfrak{trop}}^{-1}(\Delta)\subset(\Lambda^{*})^{n}, say 𝐲=(y1,,yn)\mathbf{y}=(y_{1},\dots,y_{n}). Let WW be the local expression of WW^{\vee} in this local chart. Recall that the nonvanishing of HF(L,𝔪,𝐲)\operatorname{HF}(L,\mathfrak{m},\mathbf{y}) does not rely on the local chart, and so does the critical value λ=W(𝐲)\lambda=W(\mathbf{y}). Now, it follows from Theorem 5.5 that λ\lambda is an eigenvalue of c1c_{1}. Ultimately, applying the same argument to all critical points of WW^{\vee}, we complete the proof of Theorem B.

Acknowledgment.

The author thanks Kenji Fukaya for his constant encouragement, and Ezra Getzler and Boris Tsygan for helpful comments on Hochschild cohomology. The author is also grateful to Mohammed Abouzaid, Denis Auroux, Andrew Hanlon, Mark McLean, Yuhan Sun, and Eric Zaslow for their interest and stimulating discussions. Special thanks to the anonymous referee for the meticulous review and invaluable suggestions.

References

  • [Abo17] Mohammed Abouzaid. The family Floer functor is faithful. Journal of the European Mathematical Society, 19(7):2139–2217, 2017.
  • [Abo21] Mohammed Abouzaid. Homological mirror symmetry without correction. Journal of the American Mathematical Society, 2021.
  • [AFO+] Mohammed Abouzaid, Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Quantum cohomology and split generation in Lagrangian Floer theory (in preparation).
  • [Aur07] Denis Auroux. Mirror symmetry and T-duality in the complement of an anticanonical divisor. Journal of Gökova Geometry Topology, 1:51–91, 2007.
  • [Ber12] Vladimir G Berkovich. Spectral theory and analytic geometry over non-Archimedean fields. Number 33. American Mathematical Soc., 2012.
  • [Bos14] Siegfried Bosch. Lectures on formal and rigid geometry, volume 2105. Springer, 2014.
  • [CBMS10] Ricardo Castaño-Bernard, Diego Matessi, and Jake P Solomon. Symmetries of Lagrangian fibrations. Advances in mathematics, 225(3):1341–1386, 2010.
  • [CO06] Cheol-Hyun Cho and Yong-Geun Oh. Floer cohomology and disc instantons of Lagrangian torus fibers in Fano toric manifolds. Asian Journal of Mathematics, 10(4):773–814, 2006.
  • [FOOO09] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Canonical models of filtered AA_{\infty}-algebras and Morse complexes. New perspectives and challenges in symplectic field theory, 49:201–227, 2009.
  • [FOOO10a] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian Floer theory on compact toric manifolds, I. Duke Mathematical Journal, 151(1):23–175, 2010.
  • [FOOO10b] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian intersection Floer theory: anomaly and obstruction, Part I, volume 1. American Mathematical Soc., 2010.
  • [FOOO10c] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian intersection Floer theory: anomaly and obstruction, Part II, volume 2. American Mathematical Soc., 2010.
  • [FOOO12] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Toric degeneration and nondisplaceable Lagrangian tori in S2×S2S^{2}\times S^{2}. International Mathematics Research Notices, 2012(13):2942–2993, 2012.
  • [FOOO16] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Lagrangian Floer Theory and Mirror Symmetry on Compact Toric Manifolds. Société Mathématique de France, 2016.
  • [FOOO19] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Spectral invariants with bulk, quasi-morphisms and Lagrangian Floer theory, volume 260. American Mathematical Society, 2019.
  • [FOOO20] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta, and Kaoru Ono. Kuranishi Structures and Virtual Fundamental Chains. Springer, 2020.
  • [Fuk09] Kenji Fukaya. Lagrangian surgery and rigid analytic family of Floer homologies. https://www.math.kyoto-u.ac.jp/ fukaya/Berkeley.pdf, 2009.
  • [Fuk10] Kenji Fukaya. Cyclic symmetry and adic convergence in Lagrangian Floer theory. Kyoto Journal of Mathematics, 50(3):521–590, 2010.
  • [Fuk11] Kenji Fukaya. Counting pseudo-holomorphic discs in Calabi-Yau 3-folds. Tohoku Mathematical Journal, Second Series, 63(4):697–727, 2011.
  • [Gan13] Sheel Ganatra. Symplectic cohomology and duality for the wrapped fukaya category. arXiv preprint arXiv:1304.7312, 2013.
  • [Ger63] Murray Gerstenhaber. The cohomology structure of an associative ring. Annals of Mathematics, pages 267–288, 1963.
  • [Get93] Ezra Getzler. Cartan homotopy formulas and the gauss-manin connection in cyclic homology. In Israel Math. Conf. Proc, volume 7, pages 65–78, 1993.
  • [GHK13] Mark Gross, Paul Hacking, and Sean Keel. Birational geometry of cluster algebras. arXiv preprint arXiv:1309.2573, 2013.
  • [GHK15] Mark Gross, Paul Hacking, and Sean Keel. Mirror symmetry for log Calabi-Yau surfaces I. Publications Mathématiques de l’IHES, 122(1):65–168, 2015.
  • [GHKK18] Mark Gross, Paul Hacking, Sean Keel, and Maxim Kontsevich. Canonical bases for cluster algebras. Journal of the American Mathematical Society, 31(2):497–608, 2018.
  • [GJ94] Ezra Getzler and John DS Jones. Operads, homotopy algebra and iterated integrals for double loop spaces. arXiv preprint hep-th/9403055, 1994.
  • [GS11] Mark Gross and Bernd Siebert. From real affine geometry to complex geometry. Annals of mathematics, pages 1301–1428, 2011.
  • [GS14] Yoel Groman and Jake P Solomon. A reverse isoperimetric inequality for J-holomorphic curves. Geometric and Functional Analysis, 24(5):1448–1515, 2014.
  • [HK20] Paul Hacking and Ailsa Keating. Homological mirror symmetry for log calabi-yau surfaces. arXiv preprint arXiv:2005.05010, 2020.
  • [Kad] T Kadeishvili. Structure of AA_{\infty}-algebra and Hochschild and Harrison cohomology. arXiv preprint math/0210331.
  • [Kon] Maxim Kontsevich. Bridgeland Stability over Non-Archimedean Fields. IHES website.
  • [Kon95] Maxim Kontsevich. Homological algebra of mirror symmetry. In Proceedings of the international congress of mathematicians, pages 120–139. Springer, 1995.
  • [McL20] Mark McLean. Birational Calabi-Yau manifolds have the same small quantum products. Annals of Mathematics, 191(2):439–579, 2020.
  • [MS12] Dusa McDuff and Dietmar Salamon. J-holomorphic curves and symplectic topology, volume 52. American Mathematical Soc., 2012.
  • [RS17] Alexander F Ritter and Ivan Smith. The monotone wrapped Fukaya category and the open-closed string map. Selecta Mathematica, 23(1):533–642, 2017.
  • [Sei06] Paul Seidel. A biased view of symplectic cohomology. Current developments in mathematics, 2006(1):211–254, 2006.
  • [She15] Nick Sheridan. Homological mirror symmetry for Calabi–Yau hypersurfaces in projective space. Inventiones mathematicae, 199(1):1–186, 2015.
  • [She16] Nick Sheridan. On the Fukaya category of a Fano hypersurface in projective space. Publications mathématiques de l’IHÉS, 124(1):165–317, 2016.
  • [Smi17] Ivan Smith. Stability conditions in symplectic topology. arXiv preprint arXiv:1711.04263, 2017.
  • [Sol20] Jake P Solomon. Involutions, obstructions and mirror symmetry. Advances in Mathematics, 367:107107, 2020.
  • [ST16] Jake P Solomon and Sara B Tukachinsky. Differential forms, Fukaya AA_{\infty} algebras, and Gromov-Witten axioms. arXiv preprint arXiv:1608.01304, 2016.
  • [SYZ96] A Strominger, S-T Yau, and E Zaslow. Mirror symmetry is T-duality. Nuclear Physics. B, 479(1-2):243–259, 1996.
  • [TT05] Dmitry Tamarkin and Boris Tsygan. The ring of differential operators on forms in noncommutative calculus. In Proceedings of Symposia in Pure Mathematics, volume 73. Citeseer, 2005.
  • [Tu14] Junwu Tu. On the reconstruction problem in mirror symmetry. Advances in Mathematics, 256:449–478, 2014.
  • [Yua] Hang Yuan. Fukaya category over affinoid coefficients. in preparation.
  • [Yua20] Hang Yuan. Family Floer program and non-archimedean SYZ mirror construction. arXiv preprint arXiv: 2003.06106, 2020.
  • [Yua22a] Hang Yuan. Family Floer mirror space for local SYZ singularities. arXiv preprint arXiv:2206.04652, 2022.
  • [Yua22b] Hang Yuan. Family Floer SYZ singularities for the conifold transition. arXiv preprint arXiv:2212.13948, 2022.
  • [Yua23] Hang Yuan. Family Floer SYZ conjecture for AnA_{n} singularity. arXiv preprint arXiv:2305.13554, 2023.
  • [Yua24] Hang Yuan. Non-archimedean analytic continuation of unobstructedness. arXiv preprint arXiv:2401.02577, 2024.