1 Introduction

Kontsevich conjectured [26] that enumerative mirror symmetry, an equality between Gromov–Witten invariants of a space X and period integrals on Y (see [4]) is a consequence of a homological mirror symmetry:

$$\begin{aligned} Fuk(X) \cong D^bCoh(Y). \end{aligned}$$
(1)

This paper focusses on the symplectic side of mirror symmetry. Barannikov [2] shows that one can extract the Gromov–Witten invariants of X from a variation of semi-infinite Hodge structures (VSHS) associated to the quantum cohomology of X, together with a splitting of the Hodge filtration. This goes via the intermediary step of a Frobenius manifold. One approach to obtain enumerative invariants from the Fukaya category is thus to first associate a VSHS to it, and then to specify the correct splitting. It is by now well understood how to construct the structure of a VSHS on the cyclic homology of an \(A_\infty \)-category (see [16, 24], or [42]). Characterising the splitting has not been done in general, but results have been obtained in various settings.

Ganatra et al. [12] characterise the splitting when the VSHS is \(\mathbb {Z}\)-graded and of Hodge–Tate type over a one-dimensional base. The geometric setting one should think of is the Fukaya category of a Calabi–Yau. In this case the splitting is determined by the VSHS itself. Secondly Amorim–Tu [1] show how the grading operator on quantum cohomology classifies the correct splitting when the Hochschild cohomology ring of the Fukaya category is semi-simple. The grading operator constitutes extra data, so the splitting is not necessarily determined intrinsically by the VSHS. The main examples are all Fano: complex projective space, or quadric hypersurfaces.

1.1 Formal TEP-structures

Hertling [19, Section 2.5] defines TERP-structures. We will only need TEP-structures. Furthermore, rather than working with holomorphic functions, we work with formal power series in the equivariant parameter. Hence we call them formal TEP-structures.

Definition 1.1

(see Definition 2.1) Let \(\mathbb {K}\) be a field.

  1. 1.

    A formal pre-T-structure over a \(\mathbb {K}\)-algebra R, is a pair \((\mathcal {E}, \nabla )\). Here \(\mathcal {E}\) is an R[[u]]-module and \(\nabla : Der_{\mathbb {K}}R \otimes \mathcal {E} \rightarrow u^{-1} \mathcal {E}\) a flat connection.

  2. 2.

    If \(\mathcal {E}\) is free and finitely-generated, call this a formal T-structure.

  3. 3.

    A formal TE-structure is a formal T-structure together with an extension of the connection to a flat connection \(\nabla : Der_{\mathbb {K}}(R[[u]]) \otimes \mathcal {E} \rightarrow u^{-2} \mathcal {E}\).

  4. 4.

    A formal TEP-structure is a formal TE-structure equipped with polarisation, i.e. a symmetric, sesquilinear, covariantly constant pairing \(( \cdot , \cdot ): \mathcal {E} \otimes \mathcal {E} \rightarrow R[[u]]\), which restricts to a non-degenerate pairing \(( \cdot , \cdot ): \mathcal {E}/u\mathcal {E} \otimes \mathcal {E}/u\mathcal {E} \rightarrow R\).

Thus a VSHS in the sense of [2] is a formal TP-structure.

Remark 1.2

A TEP-structure can be formalised to yield a formal TEP-structure, this process forgets information (the Stokes’ data, see [36, §II.6]). The cyclic homology of an \(A_\infty \)-category only yields a formal TEP-structure, which is why we will always be talking about formal TEP-structures. For ease of reading, we omit the word ‘formal’ from now on. We hope this doesn’t cause any confusion.

Definition 1.3

The quantum TEP-structure is defined over \(R = \Lambda [[H^*(X)]]\), where \(\Lambda \) is a Novikov ring. It is given by the \(S^1\)-equivariant quantum cohomology \(QH^*(X;R)[[u]]\). The connection is as defined in [6], or see Sect. 4.2. The pairing is given by the sesquilinear extension of the Poincaré pairing.

Definition 1.4

The TEP-structure \(HC_*^-(\mathcal {C})\) associated to an R-linear \(A_\infty \)-category is as defined in [24] or see Sect. 3.

Remark 1.5

In general, the TEP-structure associated to an \(A_\infty \)-category is only a pre-TEP-structure. If the non-commutative Hodge-de Rham spectral sequence degenerates, it is actually a TEP-structure. This is conjectured to hold for smooth and compact \(A_\infty \)-categories, see [25]. In the \({\mathbb {Z}}\)-graded setting, Kaledin [23] proves this conjectures holds. Our \(A_\infty \)-categories will only be \({\mathbb {Z}}/2\)-graded, and we will always assume the Hodge-de Rham spectral sequence degenerates, and can thus drop the prefix ‘pre’.

1.2 Cyclic open–closed map

An essential ingredient in proving that the enumerative invariants obtained from the Fukaya category agree with the Gromov–Witten invariants is an isomorphism of TEP-structures called the (negative) cyclic open–closed map. Let X be a closed symplectic manifold. Let Fuk(X) denote the Fukaya category of X, which is an \(A_\infty \)-category over the Novikov ring \(\Lambda \). Assume there exists a bulk-deformed Fukaya category \(Fuk^t(X)\). By this we mean a Fukaya category which is linear over the ring \(R = \Lambda [[H^*(X)]]\).

Conjecture 1.6

There exists a cyclic open-map \(\mathcal{O}\mathcal{C}^{-}: HC^-_*(Fuk^t(X)) \rightarrow QH^*(X;R)[[u]]\). This is a morphism of TEP-structures over \(\Lambda [[H^*(X)]]\).

Such a morphism has not been constructed in general. Partial results exist: [7, 10] construct cyclic open–closed maps in a wide range of settings. Ganatra et al. [12] have announced work proving this is an isomorphism of TP-structures when X is a projective Calabi–Yau manifold. In their case, \(R = \Lambda \), so they consider no bulk-deformations. Furthermore, Ohta–Sanda [31] show that both TE-structures considered come from a new algebraic structure they define, a ‘CH-structure’. They show that an isomorphism of the CH-structures associated to the Fukaya category and quantum cohomology would imply an isomorphism of associated TE-structures.

We prove a local version of this conjecture, focussing on the cyclic homology of the \(A_\infty \)-algebra associated to a single Lagrangian. We use the same technical assumptions as used by Solomon and Tukachinsky [47]: the moduli spaces of holomorphic disks need to be smooth orbifolds with corners, and the evaluation maps at boundary points are assumed to be submersions, see Assumptions 4.18. For us the main example satisfying these conditions is \(X = \mathbb{C}\mathbb{P}^n\) and L a Lagrangian torus fibre (see Lemma 4.19 for a proof of the assumptions). Another class of examples is given by flag varieties and their products, with the Lagrangian given by the real locus (see [47, Example 1.5]). The \(A_\infty \)-algebra \(CF^*(L,L)\) we use is equal up to sign to the \(A_\infty \)-algebra defined by [47], see Remark 4.41.

Theorem 1.7

(See Theorem 4.71) Let \(L\subset X\) be an oriented, relatively-spin Lagrangian submanifold equipped with a \(U(\Lambda )\)-local system. Suppose there exists a complex structure J such that (LJ) satisfy Assumptions 4.18. Then there exists a bulk-deformed Fukaya \(A_\infty \)-algebra \(CF^*(L,L)\). This is an \(R=\Lambda [[H^*(X)]]\)-linear, curved and filtered \(A_\infty \)-algebra. Furthermore, there exist a cyclic open–closed map

$$\begin{aligned} \mathcal{O}\mathcal{C}^{-}: HC^-_*(CF^*(L,L)) \rightarrow QH^*(X;R)[[u]], \end{aligned}$$
(2)

which is a morphism of pre-TE-structures over R.

The argument we use to show that the cyclic open–closed map is a morphism of T-structures is due to Ganatra et al. [11], as announced in [12]. The argument simplifies in our setting as the \(A_\infty \)-category we use is cyclic and strictly unital. This ensures that our construction of the cyclic open–closed map does not require higher order terms in u, as opposed to [10]. This comes at the cost of working over a field containing \({\mathbb {R}}\). To shows that the cyclic open–closed map respects the connection in the u-directions an extra ingredient is needed, which is that the cyclic open–closed map respects (Euler-)gradings.

As each component of the Fukaya category \(Fuk(\mathbb{C}\mathbb{P}^n)\) is generated by the Clifford torus (but with different local systems), we can thus construct a bulk-deformed Fukaya category \(Fuk(\mathbb{C}\mathbb{P}^n;R)\) over \(R = \mathbb {C}[[H^*(\mathbb{C}\mathbb{P}^n)]]\). We can thus define a global cyclic open–closed map using our setup:

Corollary 1.8

There exists a cyclic open–closed map

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-: HC^-_*(Fuk(\mathbb{C}\mathbb{P}^n;R)) \rightarrow QH^*(\mathbb{C}\mathbb{P}^n;R)[[u]], \end{aligned}$$
(3)

which is an isomorphism of TE-structures over R.

Remark 1.9

One reason we adopt the rather restrictive technical assumptions of [47] is that we plan follow-up work in which we relate the results of this paper, which concern closed Gromov–Witten invariants, with the open Gromov–Witten invariants defined in [44]. Similar to [47, Remark 4.2] we expect that these restrictive technical assumptions can be removed, as their role is purely to simplify the analysis of moduli spaces of holomorphic disks.

1.3 Image of the cyclic open–closed map for monotone symplectic manifolds

For the remainder of the introduction, let X be a monotone symplectic manifold. It is then possible to define the Fukaya category and quantum cohomology over \(\mathbb {C}\) (rather than over a Novikov ring). For ease of exposition in this introduction, we set all bulk-parameters equal to zero and the Novikov parameter to 1, so that \(R = {\mathbb {K}}= \mathbb {C}\). Because we then only have a connection in the u-direction, we call \(QH^*(X)[[u]]\) an E-structure (see Sect. 2.1).

By definition (see [41] for example), \(Fuk(X;\mathbb {C}) = \bigoplus _w Fuk(X)_w\). Here \(Fuk(X)_w\) is a \(\mathbb {C}\)-linear \(A_\infty \)-category with objects monotone Lagrangians with disk potential \(w \in \mathbb {C}\). We consider \(Fuk(X)_w\) as a weakly curved \(A_\infty \)-algebra with curvature \(w \cdot 1\). For a monotone symplectic manifold, quantum cohomology can also be defined over \(\mathbb {C}\) (see [29]). As a vector-space we have \(QH^*(X;\mathbb {C}) = H^*(X;\mathbb {C})\). The first Chern class defines a map:

$$\begin{aligned} c_1 \star : QH^*(X;\mathbb {C}) \rightarrow QH^*(X;\mathbb {C}). \end{aligned}$$
(4)

Decompose quantum cohomology into generalised eigenspaces for this map:

$$\begin{aligned} QH^*(X;\mathbb {C}) = \bigoplus _w QH^*(X)_w. \end{aligned}$$
(5)

The following was first proved by Ritter–Smith:

Theorem 1.10

([33, Theorem 9.5]) The open–closed map satisfies \(\mathcal{O}\mathcal{C}(HH_*(Fuk(X)_w)) \subset QH^*(X)_w\).

A natural question to ask is how this result extends to cyclic homology:

Question

What is the image of the cyclic open closed map \({\mathcal{O}\mathcal{C}}^{-}(HC^-_*(Fuk(X)_w)) \subset QH^*(X)[[u]]\)?

One might naively think we would have \({\mathcal{O}\mathcal{C}}^{-}(HC^-_*(Fuk(X)_w)) \subset QH^*(X)_w[[u]]\), but the latter is not necessarily invariant under the connection in the u-direction, so this is incompatible with Conjecture 1.6. Instead, taking inspiration from [24, Section 2.28], we apply the Hukuhara–Levelt–Turrittin theorem (see [21, 27, 49]) to decompose the quantum E-structure as a direct sum of \(\nabla _{\frac{d}{du}}\)-invariant submodules indexed by the eigenvalues of \(c_1 \star \):

Lemma 1.11

(Hukuhara–Levelt–Turrittin) There exists a unique decomposition of \(QH^*(X)[[u]]\) into \(\mathbb {C}[[u]]\) submodules:

$$\begin{aligned} QH^*(X)[[u]]= \bigoplus _w QH^*(X)[[u]]_w, \end{aligned}$$
(6)

such that each \(QH^*(X)[[u]]_w\) is invariant under \(u^2\nabla _{\frac{d}{du}}\), and \(QH^*(X)[[u]]_{w}/uQH^*(X)[[u]]_w = QH^*(X)_w\).

Conjecture 1.6, along with a slight extension of the results in [27], then shows that the cyclic open–closed map respects this decomposition:

Corollary 1.12

(See 6.5) \({\mathcal{O}\mathcal{C}}^{-}(HC^-_*(Fuk(X)_w)) \subset QH^*(X)[[u]]_w\).

Since we don’t actually prove the full conjecture, the corollary that follows from our Theorem 1.7 is:

Corollary 1.13

Let \(L \subset X\) be a monotone Lagrangian with disk potential w. Suppose (XL) satisfies Assumptions 4.18, then:

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^{-}(HC^-(CF^*(L,L))) \subset QH^*(X)[[u]]_w. \end{aligned}$$
(7)

Remark 1.14

The Hukuhara–Levelt–Turrittin decomposition has appeared before in the study of mirror symmetry. It was used first in [20] and later in [24] to introduce the notion of a Hodge structure of exponential type.

1.4 Semi-simple quantum cohomology

If we additionally assume that \(QH^*(X;{\mathbb {C}})\) is a semi-simple \({\mathbb {C}}\)-algebra (isomorphic as a ring to a direct sum of copies of \({\mathbb {C}}\)), we can completely determine the E-structure \(QH^*(X)[[u]]\). To this end, for \(\phi \in \mathbb {C}[u^{-1}]\), let \(\mathcal {E}^{\phi }:= (\mathbb {C}[[u]], \nabla _{\frac{d}{du}})\) denote the 1-dimensional TE-structure (over \(R = {\mathbb {C}}\)), with connection given by \(\nabla _{\frac{d}{du}} = \frac{d}{du} + \frac{d\phi }{du}\). We show the following, which was already obtained by [6], see also [9, 48]:

Lemma 1.15

(See Corollary 6.9) Assume \(QH^*(X)\) is semi-simple, then there exists a basis \(v_i \in QH^*(X)[[u]]\) such that \(u^2\nabla _{\frac{d}{du}}v_i = w_iv_i\), where the \(w_i\) are the eigenvalues of \(c_1 \star \). We call the \(v_i\)\(w_i\)-flat sections’. Equivalently, there is an isomorphism of E-structures

$$\begin{aligned} QH^*(X)[[u]]_w \cong \mathcal {E}^{-\frac{w}{u}}. \end{aligned}$$
(8)

Remark 1.16

The semi-simplicity assumption is essential; diagonalisability of \(c_1\) is insufficient. This is because we need a special property of the grading operator \(\mu \) on quantum cohomology (see Lemma 6.8).

Definition 1.17

Given an E-structure \((\mathcal {E},\nabla )\) a splitting is a \({\mathbb {K}}\)-linear map \(s: \mathcal {E}/u\mathcal {E} \rightarrow \mathcal {E}\) splitting the natural projection \(\pi : \mathcal {E} \rightarrow \mathcal {E}/u\mathcal {E}\).

Example 1.18

The E-structure \(\mathcal {E}^{\phi }\) admits a splitting given by:

$$\begin{aligned} s: \mathbb {C} = \mathcal {E}^\phi /u\mathcal {E}^\phi&\rightarrow \mathcal {E}^\phi = {\mathbb {C}}[[u]]\nonumber \\ \alpha&\mapsto \alpha . \end{aligned}$$
(9)

Remark 1.19

A choice of splitting is equivalent to a choice of opposite subspace as used by Barannikov [2] to obtain a Frobenius manifold from a VSHS. See also [18, Section 2.1.7].

The quantum E-structure admits a canonical splitting. This splitting does not respect the decomposition of Lemma 1.11, but it is the one relevant for Gromov–Witten theory:

$$\begin{aligned} s^{GW}: QH^*(X)&\rightarrow QH^*(X)[[u]]\nonumber \\ \alpha&\mapsto \alpha . \end{aligned}$$
(10)

When the quantum cohomology is semi-simple, the \(w_i\)-flat sections define a second splitting \(s^{ss}: QH^*(X) \rightarrow QH^*(X)[[u]]\) given by:

$$\begin{aligned} v_i (\text {mod } u) \mapsto v_i. \end{aligned}$$
(11)

Note that whilst the \(v_i\) are not unique, the associated splitting is uniquely determined, as any two choices of the \(v_i\) are related by a constant matrix. This splitting preserves the decomposition of the quantum TE-structure:

$$\begin{aligned} s^{ss}(QH^*(X)_w) \subset QH^*(X)[[u]]_w. \end{aligned}$$
(12)

For a general E-structure, given two splittings \(s_1, s_2\), we obtain an element \(R \in Aut(\mathcal {E}/u\mathcal {E})[[u]]\) as \(R = \sum _{i \ge 0} u^iR_i\), with \(R_0 = Id\), and

$$\begin{aligned} s_1(\alpha ) = \sum _{i \ge 0} u^i s_2(R_i(\alpha )) \; \text {for all} \; \alpha \in \mathcal {E}/u\mathcal {E}. \end{aligned}$$
(13)

Such R is called an R-matrix.

Remark 1.20

R-matrices were used by Givental [17] and Teleman [48] to classify semi-simple TFT’s. See also [32, chapter 2] for the definition of R-matrices and their action on cohomological field theories. Their definition of an R-matrix involves an additional ‘symplectic’ property, namely that R preserves the polarisation. The group of such symplectic R-matrices is called the Givental loop group. We do not consider this polarisation, so our R-matrices need not be elements of the Givental loop group.

The two splittings on the quantum E-structure are thus related by an R-matrix \(R \in Aut(QH^*(X))[[u]]\). A short computation shows that this is indeed the same R-matrix as defined by Teleman [48] to recover all (including higher genus) Gromov–Witten invariants of X from its genus 0, 3-point invariants.

By Corollary 1.12, we find the following:

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-(HC^-_*(Fuk(X)_w)) \subset R(QH^*(X)_w[[u]]). \end{aligned}$$
(14)

The R-matrix thus tells us how to change the naive/constant decomposition of quantum cohomology to be compatible with the cyclic open–closed map.

Amorim and Tu show the categorical version of Lemma 1.15:

Lemma 1.21

([1, Corollary 3.8]) Let \(\mathcal {C} = \bigoplus _w \mathcal {C}_w\) be a direct sum of strictly unital, smooth, finite-dimensional, cyclic and weakly curved \(A_\infty \)-categories of curvature \(w \cdot 1\). Assume \(HH^*(\mathcal {C})\) is semi-simple. Then there exists a splitting:

$$\begin{aligned} s^{\mathcal {C}}:HC^-_*(\mathcal {C})/uHC^-_*(\mathcal {C}) = HH_*(\mathcal {C}) \rightarrow HC^-_*(\mathcal {C}) \end{aligned}$$
(15)

characterised by the equation \(u^2\nabla _{\frac{d}{du}}s(\alpha ) = ws(\alpha )\) for all \(\alpha \in HH_*(\mathcal {C}_w)\).

This lemma can be rephrased as the existence of an isomorphism of E-structures

$$\begin{aligned} HC_*^-(\mathcal {C}) \cong \bigoplus _w \mathcal {E}^{-\frac{w}{u}}. \end{aligned}$$
(16)

If \(QH^*(X)\) is semi-simple, and the closed–open map is an isomorphism, then \(HH^*(Fuk(X))\) is semi-simple. Thus the previous lemma is indeed what was expected from Conjecture 1.6 and Lemma 1.15.

In Sect. 6.1 we explain how our Conjecture 1.6, if proved in appropriate generality, can be used to give an alternative proof of the following theorem of Amorim-Tu.

Theorem 1.22

([1, Theorem 1.3]) Let X be a symplectic manifold with \(HH^*(Fuk(X))\) semi-simple. Then the category Fuk(X) together with the closed–open map determine the big quantum cohomology as a Frobenius manifold.

Amorim and Tu prove their theorem under the assumption that \({\mathcal{C}\mathcal{O}}\) is a ring isomorphism, and use the Dubrovin-Teleman reconstruction theorem ([6, 48]) of semi-simple Frobenius manifolds. Our proof instead uses \({\mathcal{O}\mathcal{C}}^-\) and assumes Conjecture 1.6, which allows us to avoid appealing to the reconstruction theorem.

1.5 Speculations on the general case

When the quantum cohomology is not semi-simple, a basis of w-flat sections does not necessarily exist. However, sometimes it is still possible to construct a non-trivial R-matrix. Consider the case when the Fukaya category of X splits as follows:

$$\begin{aligned} Fuk(X) \cong \bigoplus _i Fuk(Y_i), \end{aligned}$$
(17)

where the \(Y_i\) are (not necessarily monotone) symplectic manifolds. This is expected to hold when X is a blow up (see [50] for a proof in certain cases). Another example is the complete intersection of two quadric hypersurfaces in \(\mathbb{C}\mathbb{P}^5\) (see [43]). We conjecture:

Conjecture 1.23

When Fuk(X) splits up as above, then the Gromov–Witten invariants of X can be obtained from those of the \(Y_i\), together with the genus 0, 3 point invariants of X.

We will illustrate this conjecture when X is the complete intersection of two quadric hypersurfaces in \(\mathbb{C}\mathbb{P}^5\). The eigenvalue decomposition of the Fukaya category is as follows:

$$\begin{aligned} Fuk(X) = Fuk(X)_{-8} \oplus Fuk(X)_0 \oplus Fuk(X)_8. \end{aligned}$$
(18)

Smith proves an equivalence:

Theorem 1.24

([43, Theorem 1.1]) \(D^\pi Fuk(X)_0 \cong D^\pi Fuk(\Sigma _2)\), for \(\Sigma _2\) a genus 2 surface.

Assume that \(Fuk(X)_{\pm 8} \cong Fuk(pt)\), which [43, Section 1.6] expects. And note that the Fuk(pt) are considered here with curvature \(\pm 8\), so that \(HC^-_*(Fuk(X)_{\pm 8}) \cong \mathcal {E}^{\mp 8/u}\) (see Lemma 3.27). Also note that [42, Chapter 4] proves a natural isomorphism \(HC^{-}(D^\pi \mathcal {C}) \cong HC^{-}(\mathcal {C})\). We thus have an isomorphism of E-structures:

$$\begin{aligned} HC_*^{-}(Fuk(X)) \cong \mathcal {E}^{8/u} \oplus HC_*^{-}(Fuk(\Sigma _2)) \oplus \mathcal {E}^{-8/u}. \end{aligned}$$
(19)

The cyclic open–closed map then carries this isomorphism to an isomorphism of E-structures:

$$\begin{aligned} \Phi : QH^*(X)[[u]]&\cong \mathcal {E}^{8/u} \oplus QH^*(\Sigma _2)[[u]]\oplus \mathcal {E}^{-8/u},\end{aligned}$$
(20)
$$\begin{aligned} x&\mapsto (\Phi _1(x),\Phi _2(x), \Phi _3(x)). \end{aligned}$$
(21)

An explicit computation shows that \(\Phi \) is unique up to rescaling the \(\Phi _i\) by constants \(\lambda _i \in {\mathbb {C}}\). Thus, the following splitting is well-defined:

$$\begin{aligned} s_1:= \Phi ^{-1} \circ (s \oplus s^{GW} \oplus s) \circ \pi \circ \Phi : QH^*(X) \rightarrow QH^*(X)[[u]]. \end{aligned}$$
(22)

Here \(s: {\mathbb {C}}\rightarrow \mathcal {E}^{\pm 8/u}\) is as defined in Example 1.18, \(s^{GW}\) denotes the canonical splitting on \(QH^*(\Sigma _2)[[u]]\), and \(\pi \) is the map given by setting \(u = 0\). Let \(s_2 = s^{GW}: QH^*(X) \rightarrow QH^*(X)[[u]]\) be the canonical splitting. These splittings \(s_1\) and \(s_2\) are related by an R-matrix. In Appendix B we show how to compute this R-matrix. We conjecture:

Conjecture 1.25

This R-matrix recovers all (including higher genus) Gromov–Witten invariants of X from the genus 0, 3 point invariants of X and the all-genus Gromov–Witten invariants of \(\Sigma _2\).

1.6 Outline of the paper

Section 2 defines formal TEP-structures and related notations. In Sect. 2.2 we define semi-simple TEP-structures and interpret results of [1] using this language. Next in Sect. 3 we endow the cyclic homology of an \(A_\infty \)-algebra with a TE-structure. Section 4.3 outlines properties of the Fukaya category and the cyclic open–closed map which are sufficient to prove Conjecture 1.6 in a general setting. For a Fukaya category with a single Lagrangian we then construct a cyclic open–closed map satisfying these properties in Sect. 4.8. This relies on a structure equation for horocyclic operations, which we prove in Sect. 5. In Sect. 6 we study applications of Conjecture 1.6. In particular we show how Lemmas 1.11 and 1.15 follow from general considerations about TE-structures. We also explain an alternative proof of Theorem 1.22. In Appendix A we provide heuristics showing how a ‘standard’ definition of a Fukaya category (with multiple Lagrangians) can be modified to define a \({\mathbb {Z}}\)-graded category (but at the cost of enlarging the coefficient ring). We also outline why we expect the properties of Sect. 4.3 (which are sufficient to prove that the cyclic open–closed map is a morphism of TE-structures) to hold for this Fukaya category. In Appendix B we show there exists a unique R-matrix for the intersection of quadrics in \(\mathbb{C}\mathbb{P}^5\). Finally in Appendix 1 we prove a result which was missing in the literature about the orientation properties of gluing of holomorphic maps.

2 Formal TEP-structures

Let \(\mathbb {K}\) be a field of characteristic 0. Let R be a \({\mathbb {Z}}/2\)-graded commutative \({\mathbb {K}}\)-algebra.

Definition 2.1

(Formal T(EP)-structure)

  1. 1.

    A formal pre-T-structure over \(\mathcal {M} = Spec(R)\), is a pair \((\mathcal {E}, \nabla )\). Here \(\mathcal {E}\) is a \({\mathbb {Z}}/2\)-graded R[[u]]-module, with u of even degree and \(\nabla : Der_{{\mathbb {K}}}R \otimes \mathcal {E} \rightarrow u^{-1} \mathcal {E}\) a flat connection of even degree.

  2. 2.

    A formal pre-TE-structure is a formal pre-T-structure together with an extension of the connection to a flat connection \(\nabla : Der_{{\mathbb {K}}}(R[[u]]) \otimes \mathcal {E} \rightarrow u^{-2} \mathcal {E}\).

  3. 3.

    A formal pre-TP-structure is a formal pre-T-structure equipped with a polarisation, i.e. a covariantly constant pairing

    $$\begin{aligned} ( \cdot , \cdot )_{{\mathcal {E}}}: \mathcal {E} \otimes \mathcal {E} \rightarrow R[[u]], \end{aligned}$$

    which is R-linear, of even degree and u-sesquilinear. That is, for \(f(u) \in R[[u]]\), we have:

    $$\begin{aligned} f(u)(a,b)_{\mathcal {E}}= (f(u)a,b)_{\mathcal {E}}= (-1)^{|f||a|}(a,f(-u)b)_{\mathcal {E}}. \end{aligned}$$
  4. 4.

    For a formal pre-TEP-structure, we require that the pairing is also covariantly constant with respect to \(\nabla _{\partial _u}\). More precisely:

    $$\begin{aligned} (\nabla _{u\partial _u}a,b)_{\mathcal {E}}+ (a,\nabla _{u\partial _u}b)_{\mathcal {E}}= u\partial _u(a,b)_{\mathcal {E}}. \end{aligned}$$
  5. 5.

    If additionally \(\mathcal {E}\) is free and finitely-generated, we drop the prefix ‘pre’ from T- and TE-structures. Let \({\widetilde{{\mathcal {E}}}}= {\mathcal {E}}/u{\mathcal {E}}\). For a TP-structure, we additionally require that the restriction of the pairing \(( \cdot , \cdot )_{{\widetilde{{\mathcal {E}}}}}: {\widetilde{{\mathcal {E}}}}\otimes {\widetilde{{\mathcal {E}}}}\rightarrow R\) is non-degenerate.

As mentioned in the introduction, we will always be taking about formal T-structures, so we will forget about the ‘formal’. Additionally, we call a (pre-)TE(P)-structure with \(R = {\mathbb {K}}\) a (pre-)E(P)-structure.

Definition 2.2

A morphism of pre-T(EP)-structures is an R[[u]]-module map \(F: {\mathcal {E}}_1 \rightarrow {\mathcal {E}}_2\) which respects connections and the pairing (if one exists). A morphism of T(EP)-structures is the same as a morphism of pre-T(EP)-structures.

Definition 2.3

Let \({\mathcal {E}}\) be a pre-T-structure over spec(R). An Euler-grading on \({\mathcal {E}}\) consists of an even degree \({\mathbb {K}}\)-linear map:

$$\begin{aligned} Gr: {\mathcal {E}}\rightarrow {\mathcal {E}}, \end{aligned}$$
(23)

called the grading and a vector field \(E \in Der_{\mathbb {K}}R\) of even degree, called the Euler vector field, such that for \(f \in R\), \(a \in {\mathcal {E}}\) and \(X \in Der_{\mathbb {K}}R\):

$$\begin{aligned}&Gr(fa) = (2u\partial _u + 2E)(f)a + fGr(a), \end{aligned}$$
(24)
$$\begin{aligned}&[Gr,\nabla _X] = \nabla _{[2E,X]}. \end{aligned}$$
(25)

If \({\mathcal {E}}\) is a pre-TP-structure, we additionally require that \((2u \partial _u + 2E)(a,b)_{\mathcal {E}}= (Gr(a),b)_{\mathcal {E}}+ (a,Gr(b))_{\mathcal {E}}\).

Remark 2.4

An Euler-grading differs from a more standard definition of graded in that \({\mathcal {E}}\) is not required to admit a direct sum decomposition into graded pieces.

Definition 2.5

For Euler-graded pre-T-structures \({\mathcal {E}}_1\), \({\mathcal {E}}_2\) over R with grading operators \(Gr_1\) and \(Gr_2\), and Euler-vector field \(E_1 = E_2\), a morphism of Euler-graded pre-T-structures is a morphism of pre-T-structures F which additionally satisfies \(F \circ Gr_1 = Gr_2 \circ F\).

Definition 2.6

Given an Euler-graded pre-T-structure, we obtain an associated pre-TE-structure by setting:

$$\begin{aligned} \nabla _{\partial _u}:= \frac{1}{2u}Gr - \frac{1}{u}\nabla _E. \end{aligned}$$
(26)

A short computation shows the total connection is flat, showing this is a valid definition. As a morphism of Euler-graded pre-T-structures respects the grading and the connection, we find:

Lemma 2.7

A morphism of Euler-graded pre-T-structures is a morphism of associated pre-TE-structures.

Definition 2.8

An Euler-grading on a pre-TE-structure is an Euler-grading on the underlying T-structure, such that \(\nabla _{\partial _u} = \frac{1}{2u}Gr - \frac{1}{u}\nabla _E\).

2.1 E-structures

Throughout this subsection we take \({\mathbb {K}}= {\mathbb {C}}\).

Definition 2.9

(E(P)-structure) An E(P)-structure is a TE(P)-structure \({\mathcal {E}}\) over the \({\mathbb {K}}\)-algebra \(R = {\mathbb {K}}\), so that \({\mathcal {M}}= pt\). We thus only have a connection \(\nabla _{\frac{d}{du}}: {\mathcal {E}}\rightarrow u^{-2}{\mathcal {E}}\). For ease of notation we will often write \(\nabla \) for \(\nabla _{\frac{d}{du}}\) for an E(P)-structure.

Example 2.10

Let \(\mathcal {E}^{-w/u}\) be the 1-dimensional EP-structure \(\mathcal {E} = \mathbb {C}[[u]]\) with connection \(\nabla = \frac{d}{du} + \frac{w}{u^2}\) and pairing \((1,1) = 1\).

Definition 2.11

A splitting of an E-structure is a \({\mathbb {K}}\)-linear map \(s: {\widetilde{{\mathcal {E}}}}\rightarrow {\mathcal {E}}\) which splits the canonical map \(\pi : {\mathcal {E}}\rightarrow {\widetilde{{\mathcal {E}}}}\). If \({\mathcal {E}}\) is an EP-structure, we say s is P-compatible if \((s(a),s(b))_{\mathcal {E}}= (a,b)_{{\widetilde{{\mathcal {E}}}}}\) for all \(a,b \in {\widetilde{{\mathcal {E}}}}\).

As \({\mathcal {E}}\) is finitely generated and free, there always exists a splitting. A choice of splitting s defines an isomorphism:

$$\begin{aligned} \Phi _s: {\widetilde{{\mathcal {E}}}}[[u]]&\cong {\mathcal {E}}\nonumber \\ \sum _{i\ge 0} a_iu^i&\mapsto \sum _i s(a_i)u^i. \end{aligned}$$
(27)

Note that the sum on the right side makes sense as \({\mathcal {E}}\) is finitely generated. We can then write the connection on \({\mathcal {E}}\) as

$$\begin{aligned} \nabla =: \frac{d}{du} + A =: \frac{d}{du} + u^{-2}\sum _{i\ge 0} A_iu^i, \text { for some linear maps } A_i: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}. \end{aligned}$$
(28)

Call A the connection matrix and \(A_0\) the residue. Given two splittings \(s_1\) and \(s_2\) we obtain an isomorphism

$$\begin{aligned} R = \Phi _{s_1}^{-1} \circ \Phi _{s_2}: {\widetilde{{\mathcal {E}}}}[[u]] \rightarrow {\widetilde{{\mathcal {E}}}}[[u]]. \end{aligned}$$
(29)

Writing \(R = \sum _{i \ge 0} u^i R_i\) for linear maps \(R_i: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}[[u]]\), we find that \(R_0 = Id\). The splittings \(s_1\) and \(s_2\) are then related via:

$$\begin{aligned} s_2(\cdot ) = \sum _{i\ge 0} u^i s_1(R_i(\cdot )). \end{aligned}$$
(30)

The connection matrices are related via:

$$\begin{aligned} A^{s_2} = R^{-1} A^{s_1} R + R^{-1}\frac{dR}{du}, \end{aligned}$$
(31)

which shows that the residue \(A_0\) is independent of the choice of splitting. Such a matrix series R is called an R-matrix. Usually an extra condition, symplecticity, is imposed on R. This condition is satisfied when both splittings are P-compatible.

We now rephrase a theorem by Levelt [27, Chapter 2] in our setup. This theorem is the first step in the Hukuhara–Levelt–Turrittin decomposition. See for example [28] for a modern statement.

Theorem 2.12

Given an E-structure \({\mathcal {E}}\) there exists a unique decomposition \({\mathcal {E}}= \bigoplus _w {\mathcal {E}}_w\) where the w are the eigenvalues of the residue \(A_0: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\). This decomposition satisfies:

  • \(u^2\nabla ({\mathcal {E}}_w) \subset {\mathcal {E}}_w\) for all w,

  • \(\pi ({\mathcal {E}}_w) = {\widetilde{{\mathcal {E}}}}_w\), where \({\widetilde{{\mathcal {E}}}}_w\) denotes the w-generalised eigenspace of the residue.

The proof in [27, Chapter 2] is easily seen to apply in our situation. As we will need a specific form of the next term in the connection matrix, \(A_1\), we will provide a proof. The main result we need is:

Lemma 2.13

Let \(\{e_j\}\) be a basis for \({\mathcal {E}}\) such that the vectors \(\pi (e_j) \in {\widetilde{{\mathcal {E}}}}\) are generalised eigenvectors for the residue \(A_0\). Write the connection as \(\nabla = \frac{d}{du} + u^{-2}\sum _{i\ge 0} A_iu^i\) in this basis. Then there exists another basis \(\{v_j\}\) for \({\mathcal {E}}\) such that the following hold:

  • \(\pi (v_j) = \pi (e_j)\).

  • Write the connection as \(\nabla = \frac{d}{du} + u^{-2}\sum _{i\ge 0} \widetilde{A}_iu^i\) in the basis \(\{v_i\}\). Then each \(\widetilde{A}_i\) respects the eigenvalue decomposition of \({\widetilde{{\mathcal {E}}}}\), that is \(\widetilde{A}_i|_{{\widetilde{{\mathcal {E}}}}_w}: {\widetilde{{\mathcal {E}}}}_w \rightarrow {\widetilde{{\mathcal {E}}}}_w\).

  • \(\pi _w \circ \widetilde{A}_1|_{{\widetilde{{\mathcal {E}}}}_w} = \pi _ w \circ A_1|_{{\widetilde{{\mathcal {E}}}}_w}: {\widetilde{{\mathcal {E}}}}_w \rightarrow {\widetilde{{\mathcal {E}}}}_w\) for all w. Here \(\pi _w: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}_w\).

Proof of Lemma 2.13

Consider a new basis \(\lbrace P(e_j) \rbrace \) for some invertible \(\mathbb {C}[[u]]\)-linear map P. The new connection matrix is \(\widetilde{A} = P^{-1}AP - P^{-1}\frac{dP}{du}\). If we set \(P = id + u^mT_m\), we find (see [36, Theorem 5.7]): \(\widetilde{A}_s = A_s\) for \(s < m\),

$$\begin{aligned} \widetilde{A}_m = A_m + [A_0,T_m], \end{aligned}$$
(32)

and a more complicated expression for \(\widetilde{A}_{>m}\). Let \({\widetilde{{\mathcal {E}}}}_w\) for \(w \in \mathbb {C}\) be the generalised eigenspaces for \(A_0\), then let

$$\begin{aligned} \mathcal {W}_{A_0} = \left\{ \phi : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\mid 0 = \pi _w \circ \phi _{|_{{\widetilde{{\mathcal {E}}}}_w}}: {\widetilde{{\mathcal {E}}}}_w \rightarrow {\widetilde{{\mathcal {E}}}}_w \; \text {for all} \; w \right\} . \end{aligned}$$
(33)

These are the linear maps which vanish on the diagonal blocks of \(A_0\). A short computation shows that the restriction of the adjoint map \(ad_{A_0} = [A_0, \_]: \mathcal {W}_{A_0} \rightarrow \mathcal {W}_{A_0}\) is an isomorphism. Thus, there exists a \(T_m \in \mathcal {W}_{A_0}\) such that

$$\begin{aligned} \pi _w \circ [A_0,T_m]|_{{\widetilde{{\mathcal {E}}}}_{w'}} = -\pi _w \circ A_m|_{{\widetilde{{\mathcal {E}}}}_{w'}}: {\widetilde{{\mathcal {E}}}}_{w'} \rightarrow {\widetilde{{\mathcal {E}}}}_w. \end{aligned}$$
(34)

That is, all entries of \([A_0,T_m]\) and \(-A_m\) which are not in the diagonal blocks of \(A_0\) agree. We thus have \(\widetilde{A}_m({\widetilde{{\mathcal {E}}}}_w) \subset {\widetilde{{\mathcal {E}}}}_w\). We then find the \(T_m\) successively, starting with \(m=1\). Then set \(P = \prod _{m \ge 1} (id + u^mT_m)\), noting that this product is well-defined, as for each power of u, only finitely many terms in the product contribute. Then set \(v_j = P(e_j)\). This shows the first two properties.

For the final property, note that \(\widetilde{A}_1 = A_1 + [A_0,T_1]\). As we need that \(\widetilde{A}_1|_{{\widetilde{{\mathcal {E}}}}_w}: {\widetilde{{\mathcal {E}}}}_w \rightarrow {\widetilde{{\mathcal {E}}}}_w\), we can choose \(T_1\) to only have entries in the off-diagonal blocks. That is, the restriction \(T_1: {\widetilde{{\mathcal {E}}}}_w \rightarrow {\widetilde{{\mathcal {E}}}}_w\) vanishes. But then the same holds for \([A_0,T_1]\). \(\square \)

Proof of Theorem 2.12

Let \({\mathcal {E}}_w = \langle v_j | \pi (v_j) \in {\widetilde{{\mathcal {E}}}}_w \rangle \). By construction, the \({\mathcal {E}}_w\) are \(u^2 \nabla \)-invariant. Uniqueness of the decomposition follows from the following lemma.

Lemma 2.14

Let \(f: {\mathcal {E}}\rightarrow {\mathcal {E}}'\) be a morphism of E-structures. Then for any choice of decomposition by eigenvalues \({\mathcal {E}}= \oplus _w {\mathcal {E}}_w\) by eigenvalues of \(A_0\), and any choice of decomposition of \({\mathcal {E}}'\) by the eigenvalues of \(A_0'\), we have that \(f({\mathcal {E}}_w) \subset {\mathcal {E}}'_w\).

Remark 2.15

Levelt [27, Chapter 2] proves this lemma when f is an isomorphism. Our proof of the general case is very similar.

Lemma 2.14 follows directly from:

Lemma 2.16

Let \(f: {\mathcal {E}}\rightarrow {\mathcal {E}}'\) be a morphism of E-structures. Assume that the residues \(A_0\) and \(A'_0\) have no eigenvalues in common, then \(f = 0\).

Proof

Expand f in a basis for \({\mathcal {E}}\) and \({\mathcal {E}}'\) as a matrix \(F = \sum _i u^iF_i\). Expand the connections \(\nabla , \nabla '\) as usual with connection matrices A and \(A'\). As f respects connection, we obtain the equation:

$$\begin{aligned} u^2\frac{dF}{du} = FA - A'F. \end{aligned}$$
(35)

Expanding in powers of u, we find

$$\begin{aligned} F_0A_0 - A'_0F_0 = 0. \end{aligned}$$
(36)

As \(A'_0\) and \(A_0\) have no eigenvalues in common, this implies \(F_0 = 0\) (see [14, Chapter 8.1]). Next, compare coefficients of \(u^{m+1}\). This yields:

$$\begin{aligned} F_{m+1}A_0 - A'_0 F_{m+1} = L(F_0,F_1, \dots F_{m}), \end{aligned}$$
(37)

where \(L(F_0, \dots , F_{m-g})\) denotes a linear combination of the \(F_{\le m}\) with vanishing constant term. By induction we can assume \(F_0, \dots , F_m\) vanish, which implies \(F_{m+1} = 0\). \(\square \)

2.2 Semi-simple TEP-structures

In this section we will interpret results from [1] in the language of TEP-structures. For simplicity, let \({\mathbb {K}}= {\mathbb {C}}\).

Definition 2.17

(semi-simple E(P)-structure) An E(P)-structure is semi-simple if there exists an isomorphism of E(P)-structures \({\mathcal {E}}\cong \bigoplus _w \mathcal {E}^{-w/u}\). Here the values \(w \in {\mathbb {C}}\) are allowed to occur with multiplicity. Let \(\xi = u^2\nabla _{\frac{d}{du}}: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\) be the residue of the connection. Thus, \(\xi \) is given by multiplication by w on each summand \(\mathcal {E}^{-w/u}\).

The following two definitions are inspired by [1].

Definition 2.18

([1, Definition 3.9]) Let \({\mathcal {E}}\) be a semi-simple EP-structure with a specified element \(\omega \in {\widetilde{{\mathcal {E}}}}\). We call a \({\mathbb {C}}\)-linear map \(\mu : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\) a grading operator and say it is:

  1. 1.

    P-compatible: if \((\mu (x),y)_{{\widetilde{{\mathcal {E}}}}} +(x,\mu (y))_{{\widetilde{{\mathcal {E}}}}} = 0\) for all \(x,y \in {\widetilde{{\mathcal {E}}}}\),

  2. 2.

    \(\xi \)-compatible: if the restriction \(\pi _w \circ \mu : {\widetilde{{\mathcal {E}}}}_{w} \rightarrow {\widetilde{{\mathcal {E}}}}_{w}\) vanishes. Here \(\pi _w: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}_w\) denotes the projection onto the w-eigenspace of \(\xi \),

  3. 3.

    \(\omega \)-compatible: if \(\mu (\omega ) = r \omega \) for some \(r \in {\mathbb {C}}\) called the weight of \(\mu \).

Definition 2.19

([1, Definition 3.7]) Let \({\mathcal {E}}\) be an EP-structure with a specified element \(\omega \in {\widetilde{{\mathcal {E}}}}\), and \(s: {\widetilde{{\mathcal {E}}}}\rightarrow {\mathcal {E}}\) a splitting. We say the splitting is:

  1. 1.

    P-compatible: if \((s(a),s(b))_{\mathcal {E}}= (a,b)_{{\widetilde{{\mathcal {E}}}}}\) for all \(a,b \in {\widetilde{{\mathcal {E}}}}\),

  2. 2.

    Homogeneous: if \(\nabla _{\frac{d}{du}} s(a) \in u^{-1} Im(s) + Im(s)\) for all \(a \in {\widetilde{{\mathcal {E}}}}\),

  3. 3.

    \(\omega \)-compatible: if \(\nabla _{u\frac{d}{du}} s(\omega ) \in r s(\omega ) + u^{-1}Im(s)\) for some \(r \in {\mathbb {C}}\).

Example 2.20

The EP-structure \(\mathcal {E}^{-w/u}\) admits a canonical homogeneous, P- and \(\omega \)-compatible splitting given by \(s^{can}(1) = 1 \in {\mathcal {E}}\). Here we have not specified the element \(\omega \in {\widetilde{{\mathcal {E}}}}\), as the splitting is \(\omega \)-compatible for any choice of \(\omega \). This is because, by definition, \(\nabla _{u\frac{d}{du}} s^{can}(a) = u^{-1}ws^{can}(a)\) for all \(a \in {\widetilde{{\mathcal {E}}}}\).

Example 2.21

A semi-simple E-structure \({\mathcal {E}}\) comes with a canonical splitting induced by the isomorphism \(\Phi : {\mathcal {E}}\cong \bigoplus _w \mathcal {E}^{-w/u}\) and the splitting \(s^{can}\) on each \({\mathcal {E}}^{-w/u}\). This splitting is independent of the choice of isomorphism \(\Phi \) as any two such isomorphisms are related by an isomorphism \(\Psi : \bigoplus _w \mathcal {E}^{-w/u} \rightarrow \bigoplus _w \mathcal {E}^{-w/u}\) and any such \(\Psi \) is necessarily independent of u. We denote this splitting by \(s^{ss}\) (the semi-simple splitting) and note that it is homogeneous and \(\omega \)-compatible, for any \(\omega \in {\widetilde{{\mathcal {E}}}}\). If \({\mathcal {E}}\) is a semi-simple EP-structure, \(s^{ss}\) is also P-compatible.

Remark 2.22

A splitting is homogeneous if and only if the associated connection matrix \(A = u^{-2}\sum _{i \ge 0} u^iA_i\) satisfies \(A_i = 0\) for \(i \ge 2\).

Amorim and Tu show the following for EP-structures coming from the cyclic homology of an \(A_\infty \)-category, see [1, Theorem 3.10]. We state their result in our more general setup. The proof is identical.

Theorem 2.23

Let \({\mathcal {E}}\) be a semi-simple EP-structure with a specified element \(\omega \in {\widetilde{{\mathcal {E}}}}\). Then there exists a bijection between the set of homogeneous, P- and \(\omega \)-compatible splittings \(s: {\widetilde{{\mathcal {E}}}}\rightarrow {\mathcal {E}}\) and the set of P-, \(\xi \)- and \(\omega \)-compatible grading operators \(\mu : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\).

We refer the reader to [1, Theorem 3.10] for the details of the proof, but will say a few words about it. Given a splitting s as in the lemma, there exists a unique series \(R = \sum _{i\ge 0} u^iR_i\), where \(R_i: {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\) and \(R_0 = Id\) such that

$$\begin{aligned} s(a) = \sum _{i\ge 0} u^is^{can}(R_i(a)). \end{aligned}$$
(38)

The associated grading operator is then defined by \(\mu ^s = [\xi ,R_1]\), and one obtains the following relation on R:

$$\begin{aligned}{}[\xi ,R_{k+1}] = R_k(\mu ^s - k). \end{aligned}$$
(39)

One then checks all the required properties hold. Conversely, given a grading operator \(\mu : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\), [1] show that there exists a unique R-matrix solving Eq. (39) and then define the splitting \(s^{\mu }\) by Eq. (38).

Now let \(\mathcal {H}\) be an Euler-graded TEP-structure over a \({\mathbb {C}}\)-algebra R. Let \(\widetilde{{\mathcal {H}}} = {\mathcal {H}}/u{\mathcal {H}}\). The following definition is originally due to Saito [37, Definition 3.1].

Definition 2.24

([1, Definition 4.1]) An element \(\zeta \in {\mathcal {H}}\) is called a primitive form if it satisfies:

  • (Primitivity) The map defined by

    $$\begin{aligned} \rho ^{\zeta }: Der_{{\mathbb {C}}}(R) \rightarrow {\widetilde{{\mathcal {H}}}}, \; \rho ^{\zeta }(v) = [u \nabla _v \zeta ] \end{aligned}$$
    (40)

    is an isomorphism.

  • (Orthogonality) For any tangent vectors \(v_1, v_2 \in Der_{\mathbb {C}}(R)\), we have:

    $$\begin{aligned} (u\nabla _{v_1} \zeta , u\nabla _{v_2} \zeta )_{{\mathcal {H}}} \in R. \end{aligned}$$
    (41)
  • (Holonomicity) For any tangent vectors \(v_1, v_2, v_3 \in Der_{\mathbb {C}}(R)\), we have:

    $$\begin{aligned} (u\nabla _{v_1} u\nabla _{v_2} \zeta , u\nabla _{v_3} \zeta )_{{\mathcal {H}}} \in R \oplus u\cdot R. \end{aligned}$$
    (42)
  • (Homogeneity) There exists a constant \(r \in {\mathbb {C}}\) such that

    $$\begin{aligned} Gr(\zeta ) = 2r \zeta . \end{aligned}$$
    (43)

If \(\zeta \) only satisfies the Primitivity property, we will call \(\zeta \) a primitive element, and call the TEP-structure \({\mathcal {H}}\) primitive if such \(\zeta \) exists.

Definition 2.25

Let \(\mathcal {H}\) be a primitive TEP-structure over \(R = {\mathbb {C}}[[t_1, \dots , t_n]]\). Let \({\mathcal {E}}\) be the EP-structure \({\mathcal {E}}:= {\mathcal {H}}\otimes _R {\mathbb {C}}\), where \({\mathbb {C}}\) is an R-module under the map \(t_i \mapsto 0\). For \(\omega \in {\widetilde{{\mathcal {E}}}}\) say \(\omega \) is primitive if there exists a primitive element \(\zeta \in {\mathcal {H}}\) such that \(\zeta |_{t= u = 0} = \omega \).

Amorim and Tu [1, Theorem 4.2] also prove a bijection between primitive forms and splittings, which is a bijection originally established in [38]. We rephrase their theorem to apply to our setup. As already observed by [1, Remark 4.3], their proof applies to our more general setup (note that what they call a VSHS corresponds to what we call a TE-structure).

Theorem 2.26

Let \(\mathcal {H}\) be a Euler-graded, primitive TEP-structure over \(R = {\mathbb {C}}[[t_1, \dots , t_n]]\) and let \(\omega \in {\widetilde{{\mathcal {E}}}}\) be primitive. Then there exists a natural bijection between the following two sets:

$$\begin{aligned} \mathcal {P}&:= \{ \zeta \in {\mathcal {H}}| \zeta \text { is a primitive form with } \zeta |_{t= 0, u=0} = \omega \}, \end{aligned}$$
(44)
$$\begin{aligned} \mathcal {S}&:= \{ \text {homogeneous,}\, P-\, \text {and} \omega \text {-compatible splittings } s: {\widetilde{{\mathcal {E}}}}\rightarrow {\mathcal {E}}\}. \end{aligned}$$
(45)

Definition 2.27

Let \(\mathcal {H}\) be a TEP-structure over \(R = {\mathbb {C}}[[t_1, \dots , t_n]]\). We say \({\mathcal {H}}\) is semi-simple if the associated EP-structure \({\mathcal {E}}:= {\mathcal {H}}\otimes _R {\mathbb {C}}\) is semi-simple.

For a semi-simple, Euler graded and primitive TEP-structure \({\mathcal {H}}\) as above, with a choice of primitive \(\omega \in {\widetilde{{\mathcal {E}}}}\), Theorems 2.23 and 2.26 thus combine to give a bijection between the set of P-, \(\xi \)- and \(\omega \)-compatible grading operators \(\mu : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\) and the set of primitive forms \(\zeta \in {\mathcal {H}}\) with \(\zeta |_{t= 0, u=0} = \omega \).

Corollary 2.28

Let \(\mathcal {H}\) be a Euler-graded, primitive, semi-simple TEP-structure over \(R = {\mathbb {C}}[[t_1, \dots , t_n]]\) and let \(\omega \in {\widetilde{{\mathcal {E}}}}\) be primitive. Then there exists a natural bijection between the following two sets:

$$\begin{aligned} \mathcal {P}&:= \{ \zeta \in {\mathcal {H}}| \zeta \text { is a primitive form with } \zeta |_{t= 0, u=0} = \omega \},\end{aligned}$$
(46)
$$\begin{aligned} \mathcal {G}&:= \{\text {P-, }\xi \text {- and }\omega \text {-compatible grading operators }\mu : {\widetilde{{\mathcal {E}}}}\rightarrow {\widetilde{{\mathcal {E}}}}\}. \end{aligned}$$
(47)

The relevance of this bijection is that given a primitive form \(\zeta \) as above, Saito and Takahashi [39] endow Spec(R) with the structure of a Frobenius manifold. A grading operator \(\mu \) on a semi-simple TEP-structure over a ring R thus gives rise to a Frobenius manifold \({\mathcal {M}}_{\mu }\). In chapter 6 we will come back to this construction.

3 TE-structure on the cyclic homology of an \(A_\infty \)-algebra

In this section we will define a TE-structure on the cyclic homology of an \(A_\infty \)-algebra. All of the definitions can easily be extended to \(A_\infty \)-categories. Let \(S_n[k]\) be the set of all partitions of \(\lbrace 1, \dots k\rbrace \) into n ordered sets of the form \((1, 2, \dots , k_1)\), \((k_1 + 1, \dots , k_1 + k_2), \dots ,(k_1 + \dots + k_{n-1} + 1, \dots , k_1 + \dots + k_n)\). Let (i : n) denote the ith set of the partition. The size of (i : n) is \(k_i\). We allow for the case \(k_i = 0\).

Let \({\mathbb {K}}\) be a field, and assume \({\mathbb {K}}\) is complete when equipped with the trivial valuation. Let R be a \(\mathbb {Z}/2\)-graded \({\mathbb {K}}\)-algebra with a complete valuation \(\zeta _R: R \rightarrow {\mathbb {R}}_{\ge 0} \cup \{ \infty \}\). Let A be a \(\mathbb {Z}/2\)-graded R-module with a complete valuation \(\zeta _A: A \rightarrow {\mathbb {R}}_{\ge 0}\cup \{ \infty \}\). Let \(|\alpha |\) denote the degree of \(\alpha \in A\), and \(|\alpha |':= |\alpha | - 1\) the shifted degree. For \(\alpha = (\alpha _1,\dots ,\alpha _k)\) let \(\epsilon (\alpha ) = \sum _{j=1}^k |\alpha _j|'\). For a partition \(P \in S_n[k]\) let \(\epsilon _i = \epsilon (\alpha ^{(i:n)})\).

Let \(I \coprod J = [l]\) be a partition of [l] in the usual sense, not respecting the order of [l]. Equip the subsets I and J with the order induced from [l].

Definition 3.1

([47, Definition 1.1]) An n-dimensional, strictly unital, curved, cyclic, \(\mathbb {Z}/2\)-graded \(A_\infty \)-structure on A consists of:

  • R-multilinear maps \(\mathfrak {m}_k: A^{\otimes {k}} \rightarrow A[2-k]\). This means that for \(t \in R\) and \(\alpha _1, \dots , \alpha _k \in A\):

    $$\begin{aligned} \mathfrak {m}_k(\alpha _1, \dots , t\alpha _i, \dots , \alpha _k) = (-1)^{|t|(|\alpha _1|' + \dots + |\alpha _{i-1}|' + 1)}t\mathfrak {m}_k(\alpha _1, \dots , \alpha _i, \dots , \alpha _k).\nonumber \\ \end{aligned}$$
    (48)
  • A pairing \(\langle \;, \; \rangle : A \otimes A \rightarrow R[n]\).

  • An element \({\textbf{e}} \in A\) with \(|{\textbf{e}}| = 0\).

These satisfy the following relations:

  1. 1.

    The \(A_\infty \) relations hold:

    $$\begin{aligned} \sum _{P \in S_3[k]} (-1)^{\epsilon _1}\mathfrak {m}_{k_1 + 1 + k_3}(\alpha ^{(1:3)}, \mathfrak {m}_{k_2}(\alpha ^{(2:3)}),\alpha ^{(3:3)}) = 0.\end{aligned}$$
    (49)
  2. 2.

    The pairing \(\langle \;, \; \rangle \) is graded R-bilinear:

    $$\begin{aligned} \langle a \alpha _1, \alpha _2 \rangle = a \langle \alpha _1, \alpha _2 \rangle = (-1)^{|a||\alpha _1|'}\langle \alpha _1, a \alpha _2 \rangle . \end{aligned}$$
    (50)
  3. 3.

    The pairing is graded anti-symmetric:

    $$\begin{aligned} \langle \alpha _1, \alpha _2 \rangle = (-1)^{|\alpha _1|'|\alpha _2|'+1}\langle \alpha _2, \alpha _1 \rangle . \end{aligned}$$
    (51)
  4. 4.

    The pairing is cyclic: for \(\alpha = (\alpha _1, \dots , \alpha _k)\), we have

    $$\begin{aligned} \langle \mathfrak {m}_k(\alpha _1, \dots , \alpha _k), \beta \rangle = (-1)^{|\beta |'\epsilon (\alpha )} \langle \mathfrak {m}_k(\beta , \alpha _1, \dots , \alpha _{k-1}), \alpha _k \rangle . \end{aligned}$$
    (52)
  5. 5.

    The unit is strict:

    1. (a)

      \(\mathfrak {m}_k(\alpha _1, \dots , \alpha _{i-1}, {\textbf{e}}, \alpha _{i+1}, \dots , \alpha _k) = 0, \; \forall k \ne 0,2,\)

    2. (b)

      \(\mathfrak {m}_2({\textbf{e}},\alpha ) = \alpha = (-1)^{|\alpha |}\mathfrak {m}_2(\alpha , {\textbf{e}})\).

  6. 6.

    The \(A_\infty \)-operations respect the valuation:

    1. (a)

      \(\zeta _A({\mathfrak {m}}_k({\alpha })) \ge \zeta _A({\alpha })\),

    2. (b)

      \({\mathfrak {m}}_0 = w \cdot {\textbf{e}} + {\overline{{\mathfrak {m}}}}_0\), with \(\zeta _A({\overline{{\mathfrak {m}}}}_0) > 0\),

    3. (c)

      \(\zeta _R(\langle {\alpha }_1, {\alpha }_2 \rangle ) \ge \zeta _A({\alpha }_1) + \zeta _A({\alpha }_2)\).

Given an \(A_\infty \)-algebra A, recall the opposite \(A_\infty \)-algebra as defined by [42, definition 3.5]. As an R-module we have \(A^{op} = A\). But now:

$$\begin{aligned} \mathfrak {m}_k^{op}({\alpha }_1, \dots , {\alpha }_k):= (-1)^{\dagger ({\alpha })}\mathfrak {m}_k({\alpha }_k, \dots , {\alpha }_1). \end{aligned}$$
(53)

Here, for \({\alpha }= ({\alpha }_1, \dots , {\alpha }_k)\), we set \(\dagger ({\alpha }) = \sum _{1 \le i <j \le k} |{\alpha }_i|'|{\alpha }_j|'\). If A is strictly unital with unit \({\textbf{e}} \in A\), then \(A^{op}\) is strictly unital with unit \({\textbf{e}}^{op}:= -{\textbf{e}}\).

We can also define:

Definition 3.2

(\(negative A_\infty \)-algebra) Let the negative \(A_\infty \)-algebra be given by \(A^{-}:= A\) as R-modules, but \(\mathfrak {m}_k^{-}({\alpha }_1, \dots , {\alpha }_k) = (-1)^{k-1} \mathfrak {m}_k({\alpha }_1, \dots , {\alpha }_k)\).

Note that there is an isomorphism of \(A_\infty \)-algebras \(A \cong A^{-}\) given by \({\alpha }\mapsto -{\alpha }\). Combining both definitions we have

Definition 3.3

(negative-\(opposite A_\infty \)-algebra) The negative-opposite algebra associated to A is given by \((A^{op})^{-} =: A^{-op}\). This is then an n-dimensional, strictly unital, cyclic, \(\mathbb {Z}/2\)-graded \(A_\infty \)-algebra with unit \({\textbf{e}} \in A\).

3.1 Hochschild invariants

Let \((A,\mathfrak {m})\) be an n-dimensional, cyclic, strictly unital, curved, \(\mathbb {Z}/2\)-graded \(A_{\infty }\)-algebra over R. Let \(\mathfrak {r} \subset R\) be the maximal ideal of elements with positive valuation. Define the (reduced) Hochschild cochains of A:

$$\begin{aligned} CC^*(A):= \prod _{k=0}^\infty Hom_R \left( \left( \frac{A}{R\cdot {\textbf{e}}}[1]\right) ^{\otimes k}, A[1] \right) . \end{aligned}$$
(54)

Also define the uncompleted (reduced) Hochschild chains of A to be:

$$\begin{aligned} CC^{unc}_*(A):= \bigoplus _{k=0}^\infty A \otimes \left( \frac{A}{R \cdot {\textbf{e}}}\right) ^{\otimes k}. \end{aligned}$$
(55)

Following [5, Section 3.5], we define:

Definition 3.4

The completed reduced Hochschild chains and cochains are given by:

$$\begin{aligned} CC_*(A) = \varprojlim CC^{unc}_*(A/\mathfrak {r}^k). \end{aligned}$$
(56)

Remark 3.5

In the remainder of this chapter, we will recall and define various operations on Hochschild (co)chains. For simplicity, we will often define these operations only for the uncompleted chains. They descend to operations on the completed Hochschild chains as the \(A_\infty \)-operations and the pairing are assumed to respect the valuation.

Remark 3.6

We need to be careful about the R-linearity of Hochschild cochains. For \(\phi \in CC^*(A)\) and \(t\in R\) this means that:

$$\begin{aligned} \phi ({\alpha }_1, \dots , t{\alpha }_i, \dots , {\alpha }_k) = (-1)^{|t|(|{\alpha }_1|' + \dots + |{\alpha }_{i-1}|' + |\phi |')} t\phi ({\alpha }_1, \dots , {\alpha }_i, \dots , {\alpha }_k).\nonumber \\ \end{aligned}$$
(57)

Denote an element \(\alpha \in CC_*(A)\) by \(\alpha = \alpha _0[\alpha _1|\dots |\alpha _k]\), and for a subset \(I\subset \lbrace 1, \dots , k \rbrace \), write \(\alpha ^I\) for \(\bigotimes _{j \in I} \alpha _j\).

Definition 3.7

Hochschild homology is defined as: \(HH_*(A):= H^*(CC_*(A), b)\). Here the differential b is given by:

$$\begin{aligned} b(\alpha )&= \sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')} \mathfrak {m}_{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)})[\alpha ^{(2:3)}] \nonumber \\&\quad +\,\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _1 + |\alpha _0|'}\alpha _0[\alpha ^{(1:3)}|\mathfrak {m}_{k_2}(\alpha ^{(2:3)})|\alpha ^{(3:3)}]. \end{aligned}$$
(58)

Note here that in the second sum, terms with \(k_2 = 0\) are allowed.

Definition 3.8

The negative cyclic chain complex is given by

$$\begin{aligned} CC_*^-(A) = \varprojlim (CC^{unc}_*(A/\mathfrak {r}^k)[[u]], b+uB). \end{aligned}$$
(59)

Here the second differential B is defined by:

$$\begin{aligned} B(\alpha ) = \sum _{\begin{array}{c} P \in S_2[k] \end{array}} (-1)^{\epsilon _2(|\alpha _0|'+\epsilon _1)} 1[\alpha ^{(2:2)}|\alpha _0|\alpha ^{(1:2)}]. \end{aligned}$$
(60)

The homology of the negative cyclic chain complex is called the negative cyclic homology, denoted \(HC^-_*(A)\).

Hochschild cochains admit a differential too. First introduce the Gerstenhaber product, defined by:

$$\begin{aligned} \phi \circ \psi (\alpha _1, \dots , \alpha _k) = \sum _{P \in S_3[k]} (-1)^{|\psi |' \epsilon _1} \phi (\alpha ^{(1:3)} \otimes \psi (\alpha ^{(2:3)}) \otimes \alpha ^{(3:3)}). \end{aligned}$$
(61)

The Gerstenhaber bracket is then defined by: \([\phi , \psi ] = \phi \circ \psi - (-1)^{|\phi |'|\psi |'} \psi \circ \phi \). The \(A_\infty \)-structure maps \(\mathfrak {m}_k\) define a Hochschild chain \(\mathfrak {m} \in CC^2(A)\). The differential on \(CC^*(A)\) is then given by \([\mathfrak {m}, \_]\). Hochschild cohomology is defined as \(HH^*(A):= H^*(CC^*(A),[\mathfrak {m}, \_])\).

Finally we note that \(CC^*(A)\) admits an \(A_\infty \)-structure \(M^k\), defined by [16]. \(M^1\) is the differential \([\mathfrak {m}, \_]\). We will also need the \(M^2\) part of these operations.

Definition 3.9

The cup product on \(CC^*(A)\) is given by \(\psi \cup \phi := (-1)^{|\psi |}M^2(\psi ,\phi )\). Here

$$\begin{aligned} M^2(\psi ,\phi )(\alpha ):= & {} \sum _{P\in S_5[k]} (-1)^{|\psi |'\epsilon _1 + |\phi |'(\epsilon _1+\epsilon _2+\epsilon _3)}\mathfrak {m}(\alpha ^{(1:5)}\nonumber \\{} & {} \otimes \psi (\alpha ^{(2:5)})\otimes \alpha ^{(3:5)} \otimes \phi (\alpha ^{(4:5)})\otimes \alpha ^{(5:5)}). \end{aligned}$$
(62)

Since A is cyclic, we can consider the pairing:

$$\begin{aligned} (\;, \;): CC^*(A)&\otimes CC_*(A) \rightarrow R\nonumber \\ \phi&\otimes \alpha \mapsto (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1)}\langle \phi (\alpha _1, \dots , \alpha _k), \alpha _0 \rangle . \end{aligned}$$
(63)

Here, for \(\alpha \in CC_*(A)\) we use \(\widetilde{\alpha }\) to denote the tuple \((\alpha _1, \dots , \alpha _k)\).

Lemma 3.10

The pairing above descends to a pairing \(HH^*(A) \otimes HH_*(A) \rightarrow R\). Concretely, we have \(([\mathfrak {m},\phi ],\alpha ) + (-1)^{|\phi |}(\phi ,b(\alpha )) = 0\).

Proof

We first write out:

$$\begin{aligned}&([\mathfrak {m},\phi ],\alpha ) = (\mathfrak {m} \circ \phi - (-1)^{|\phi |'|\mathfrak {m}|'}\phi \circ \mathfrak {m}, \alpha ) \end{aligned}$$
(64)
$$\begin{aligned}&= \sum _{P \in S_3[k]} (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha })+ 1) + |\phi |' \epsilon _1} \langle \mathfrak {m}(\alpha ^{(1:3)} \otimes \phi (\alpha ^{(2:3)}) \otimes \alpha ^{(3:3)}),\alpha _0\rangle \end{aligned}$$
(65)
$$\begin{aligned}&+ \sum _{P \in S_3[k]} (-1)^{|\phi |'+1+|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1) + \epsilon _1} \langle \phi (\alpha ^{(1:3)} \otimes \mathfrak {m}(\alpha ^{(2:3)}) \otimes \alpha ^{(3:3)}),\alpha _0\rangle . \end{aligned}$$
(66)

Next, we write out:

$$\begin{aligned} (\phi ,b(\alpha ))&= \sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')} (\phi ,\mathfrak {m}_{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)})[\alpha ^{(2:3)}])\end{aligned}$$
(67)
$$\begin{aligned}&\qquad +\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _1 + |\alpha _0|'} (\phi ,\alpha _0[\alpha ^{(1:3)}|\mathfrak {m}_{k_2}(\alpha ^{(2:3)})|\alpha ^{(3:3)}])\end{aligned}$$
(68)
$$\begin{aligned}&= \sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')+ A_{1}} \langle \phi (\alpha ^{(2:3)}),\mathfrak {m}_{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)}) \rangle \end{aligned}$$
(69)
$$\begin{aligned}&\qquad +\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _1 + |\alpha _0|' + |\alpha _0|\epsilon (\widetilde{\alpha })} \langle \phi (\alpha ^{(1:3)}\otimes \mathfrak {m}_{k_2}(\alpha ^{(2:3)})\otimes \alpha ^{(3:3)}),\alpha _0\rangle . \end{aligned}$$
(70)

Where \(A_1 = (\epsilon _2+1)(|\alpha _0|'+\epsilon _1+\epsilon _3)\). The terms (66) and (70) agree up to a factor \((-1)^{|\phi |'}\), as required. We then use cyclic symmetry to make (69) agree with (65), up to the same factor. We compute

$$\begin{aligned}&\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')+A_{1}} \langle \phi (\alpha ^{(2:3)}),\mathfrak {m}_{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)}) \rangle \end{aligned}$$
(71)
$$\begin{aligned}&\quad =\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')+A_{1} +A_2} \langle \mathfrak {m}_{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)}),\phi (\alpha ^{(2:3)}) \rangle , \end{aligned}$$
(72)

by 51, where \(A_2 = (\epsilon _2 + |\phi |')(|\alpha _0|'+\epsilon _1+\epsilon _3+1)\). And finally:

$$\begin{aligned}&=\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')+A_{1} +A_2 + A_3} \langle \mathfrak {m}(\alpha ^{(1:3)} \otimes \phi (\alpha ^{(2:3)}) \otimes \alpha ^{(3:3)}),\alpha _0\rangle , \end{aligned}$$
(73)

where \(A_3 = (|\alpha _0|'+\epsilon _3)(|\phi |'+\epsilon _2+\epsilon _1)\). Combining all the signs, we get the required equality. \(\square \)

We will also need the bilinear maps \(b^{1,1}, B^{1,1}: CC^*(A) \otimes CC_*(A) \rightarrow CC_*(A)\), defined for a Hochschild cochain \(\phi \in CC^*(A)\), and \(\alpha = \alpha _0[\alpha _1|\dots |\alpha _k] \in CC_*(A)\) by:

$$\begin{aligned} b^{1,1}(\phi ;\alpha ) = \sum _{\begin{array}{c} P \in S_5[k] \end{array}} (-1)^{\dagger } \mathfrak {m}_{k_1+k_3+k_5+2}(\alpha ^{(3:5)}\otimes \phi (\alpha ^{(4:5)})\otimes \alpha ^{(5:5)}\otimes \alpha _0 \otimes \alpha ^{(1:5)})[\alpha ^{(2:5)}], \end{aligned}$$
(74)

where \(\dagger = (\epsilon _3 + \epsilon _4 + \epsilon _5)(|\alpha _0|' + \epsilon _1 +\epsilon _2) +|\phi |'\epsilon _3\), and

$$\begin{aligned} B^{1,1}(\phi ;\alpha ) = \sum _{\begin{array}{c} P \in S_4[k] \end{array}} (-1)^{|\phi |'\epsilon _2 + (\epsilon _1 + |\alpha _0|')(\epsilon _2 + \epsilon _3 + \epsilon _4)} {\textbf{e}}[\alpha ^{(2:4)}|\phi (\alpha ^{(3:4)})|\alpha ^{(4:4)}|\alpha _0|\alpha ^{(1:4)}].\nonumber \\ \end{aligned}$$
(75)

There is a nice interplay between the cup product on \(CC^*(A)\), the \(A_\infty \)-module structure \(b^{1,1}\), and the pairing.

Lemma 3.11

Define the cap product \(\phi \cap \alpha = (-1)^{|\phi |}b^{1,1}(\phi ,\alpha )\). We then have:

$$\begin{aligned} (\phi \cup \psi ,\alpha ) = (\phi , \psi \cap \alpha ). \end{aligned}$$
(76)

Proof

This is an easy verification of signs using the cyclic symmetry of the pairing. \(\square \)

Definition 3.12

Set \(i\lbrace \phi \rbrace = b^{1,1}(\phi , \_ ) + u B^{1,1}(\phi , \_ ): CC^{-}_*(A) \rightarrow CC^-_*(A)\).

The curvature, or Lie derivative, \(\mathcal {L}: CC^*(A) \otimes CC_*(A) \rightarrow CC_*(A)\) defined in [16] can be written as:

$$\begin{aligned} \mathcal {L}_\phi (\alpha )&= \sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{\epsilon _3(\epsilon _2 + \epsilon _1 + |\alpha _0|')}\phi _{k_3 + 1 + k_1}(\alpha ^{(3:3)}\otimes \alpha _0 \otimes \alpha ^{(1:3)})[\alpha ^{(2:3)}]\nonumber \\&\quad +\sum _{\begin{array}{c} P \in S_3[k] \end{array}} (-1)^{|\phi |'(\epsilon _1 + |\alpha _0|')} \alpha _0[\alpha ^{(1:3)}|\phi _{k_2}(\alpha ^{(2:3)})|\alpha ^{(3:3)}]. \end{aligned}$$
(77)

Observe that \(\mathcal {L}_{\mathfrak {m}} = b\). An easy computation shows:

Lemma 3.13

For any \(\phi , \psi \in CC^*(A)\), we have: \([\mathcal {L}_\psi ,\mathcal {L}_\phi ] = \mathcal {L}_{[\psi ,\phi ]}\). In particular \([b,\mathcal {L}_\phi ] = \mathcal {L}_{[{\mathfrak {m}},\phi ]}\). Furthermore \([B,\mathcal {L}_\phi ] = 0\).

Remark 3.14

For any linear maps A and B of homogeneous degrees the commutator [AB] is defined as the supercommutator: \([A,B] = AB - (-1)^{|A||B|}BA\). By extending linearly, this defines the commutator for all linear maps.

Getzler shows:

Proposition 3.15

([16, Theorem 2.2])

$$\begin{aligned}{}[i\lbrace \phi \rbrace , b+uB ]= u \mathcal {L}_{\phi } + i\lbrace [\mathfrak {m}, \phi ]\rbrace . \end{aligned}$$
(78)

Getzler furthermore defines a connection in the base directions. We extend his definition to allow for the case when R is \({\mathbb {Z}}/2\)-graded.

Definition 3.16

The Getzler–Gauss–Manin connection is defined on the chain level by:

$$\begin{aligned} \nabla ^{GGM}: Der_{\mathbb {C}}R \otimes _\mathbb {K} CC^-_*(A)&\rightarrow u^{-1}CC^{-}_*(A),\nonumber \\ \nabla ^{GGM}_v(\alpha )&:= v(\alpha ) + (-1)^{|v|+1}u^{-1}i\{ v(\mathfrak {m}) \}(\alpha ). \end{aligned}$$
(79)

Here for a Hochschild cochain \(\phi \in CC^*(A)\) and a derivation \(v \in Der_{\mathbb {C}}R\), the Hochschild cochain \(v(\phi )\) is defined as

$$\begin{aligned} v(\phi )({\alpha }):= v(\phi ({\alpha })) - (-1)^{|\phi |'|v|} \phi (v({\alpha })). \end{aligned}$$
(80)

Getzler shows the connection descends to the level of cohomology and is flat. This endows \(HC^{-}_*(A)\) with a \({\mathbb {Z}}/2\)-graded pre-T-structure over Spec(R).

Sheridan proves the following holds over a field \({\mathbb {K}}\). Nothing in their proof breaks down if we work over a general ring. We thus have:

Theorem 3.17

([42, Theorem B.2]) Let \(F: \mathcal {C} \rightarrow \mathcal {D}\) be an R-linear \(A_\infty \)-morphism. Then

$$\begin{aligned} F_*: HC_*^-({\mathcal {C}}) \rightarrow HC^-_*(\mathcal {D}) \end{aligned}$$
(81)

is a morphism of pre-T-structures.

3.2 The u-connection

The pre-T-structure \(HC^-_*(A)\) has been extended to a pre-TE-structure by [24]. We give another interpretation of this definition. First recall the notion of an Euler-grading on an \(A_\infty \)-algebra:

Definition 3.18

An Euler-grading on an n-dimensional, strictly unital, cyclic, \(\mathbb {Z}/2\)-graded \(A_\infty \)-algebra A consists of an Euler vector field \(E \in Der_{\mathbb {K}}R\) of even degree and an even degree map \(Gr: A \rightarrow A\) such that

$$\begin{aligned} Gr \circ {\mathfrak {m}}_k = {\mathfrak {m}}_k \circ Gr + (2-k) {\mathfrak {m}}_k. \end{aligned}$$
(82)

and

$$\begin{aligned} Gr(f{\alpha }) = 2E(f){\alpha }+ fGr({\alpha }) \text { for } f\in R \text { and } {\alpha }\in hom(L_1,L_2). \end{aligned}$$
(83)

Furthermore, we require that \(Gr({\textbf{e}}) = 0\). An R-linear \(A_\infty \)-morphism \(F: A \rightarrow B\) is said to be Euler-graded if \(E_A = E_B\) and \(F^k \circ Gr_A = Gr_B \circ F^k + (k-1)F^k\).

Now suppose that A is Euler-graded. Consider \(Gr: A \rightarrow A\) as a length-1 Hochschild cochain. Then define the operator \(Gr^{-}: CC^-_*(A) \rightarrow CC_*^-(A)\) by

$$\begin{aligned} Gr^{-}:= \mathcal {L}_{Gr} + \Gamma + 2 u \frac{\partial }{\partial u}, \end{aligned}$$
(84)

where \(\Gamma (\alpha _0[\alpha _1|\dots |\alpha _k]) = -k \alpha _0[\alpha _1|\dots |\alpha _k]\) is the length operator on cyclic chains.

Lemma 3.19

The grading \(Gr^{-}\) descends to cyclic homology, and endows \(HC^-_*(A)\) with an Euler-graded T-structure.

Proof

Let \({\mathfrak {m}}' = \prod _k (2-k){\mathfrak {m}}_k \in CC^*(A)\). A short computation shows \([\Gamma ,b] = b - \mathcal {L}_{{\mathfrak {m}}'}\) and \([\Gamma ,B] = -B\). We thus have:

$$\begin{aligned}{}[Gr^-,b+uB]&= [\mathcal {L}_{Gr},b+uB] + [\Gamma ,b+uB] + 2 [u \frac{d}{du},b+uB] \end{aligned}$$
(85)
$$\begin{aligned}&= \mathcal {L}_{[Gr,{\mathfrak {m}}]} + b - \mathcal {L}_{m'} - uB + 2uB\end{aligned}$$
(86)
$$\begin{aligned}&= \mathcal {L}_{{\mathfrak {m}}'} + b -\mathcal {L}_{{\mathfrak {m}}'} +uB\end{aligned}$$
(87)
$$\begin{aligned}&= b+uB. \end{aligned}$$
(88)

The second equality follows by Lemma 3.13. This shows \(Gr^-\) descends to cyclic homology. Next, observe that for \(f \in R[[u]]\) and \({\alpha }\in CC_*^-(A)\), we have \(\mathcal {L}_{Gr}(f\alpha ) = 2E(f){\alpha }+ fGr({\alpha })\). This shows that

$$\begin{aligned} Gr^-(f{\alpha }) = (2u\partial _u + 2E)(f){\alpha }+ f Gr^- {\alpha }\end{aligned}$$
(89)

holds on the chain level. Next, for \(v \in Der_\mathbb {C} R\), we want to compute \([Gr^-,\nabla _v]\). To this end, first observe that, after picking a basis for A, we have: \([\mathcal {L}_{Gr},v]({\alpha }) = [2E,v]({\alpha })\). Furthermore, a direct computation shows:

$$\begin{aligned}{}[\mathcal {L}_{Gr} + \Gamma ,b^{1,1}(v({\mathfrak {m}}),\cdot )]&= b^{1,1}([2E,v]({\mathfrak {m}}),\cdot ) + 2b^{1,1}(v({\mathfrak {m}}),\cdot )\end{aligned}$$
(90)
$$\begin{aligned}&= B^{1,1}([2E,v]({\mathfrak {m}}),\cdot ). \end{aligned}$$
(91)

We thus have:

$$\begin{aligned}{}[Gr^-,\nabla _v]&= [\mathcal {L}_{Gr} + \Gamma + 2u\partial _u,v - B^{1,1}(v({\mathfrak {m}}),\cdot ) - u^{-1}b^{1,1}(v({\mathfrak {m}}),\cdot )]\end{aligned}$$
(92)
$$\begin{aligned}&= [\mathcal {L}_{Gr},v] - [\mathcal {L}_{Gr} + \Gamma ,b^{1,1}(v({\mathfrak {m}}),\cdot )] - [\mathcal {L}_{Gr} \nonumber \\&\quad + \Gamma ,B^{1,1}(v({\mathfrak {m}}),\cdot )] + 2u^{-1}b^{1,1}(v({\mathfrak {m}}),\cdot )\end{aligned}$$
(93)
$$\begin{aligned}&= [2E,v] - B^{1,1}([2E,v]({\mathfrak {m}}),\cdot ) -u^{-1} b^{1,1}([2E,v]({\mathfrak {m}}),\cdot )\end{aligned}$$
(94)
$$\begin{aligned}&= \nabla _{[2E,v]}. \end{aligned}$$
(95)

\(\square \)

Thus, as an Euler-graded pre-T-structure naturally admits an extension to a pre-TE-structure, any Euler-graded \(A_\infty \)-algebra naturally admits a pre-TE-structure on \(HC^-_*(A)\). For an arbitrary \(A_\infty \)-algebra \({\mathcal {C}}\), we will now define an Euler-graded deformation, and use this to define a u-connection on \(HC^-_*({\mathcal {C}})\).

Definition 3.20

Let \({\mathcal {C}}\) be any R-linear \(A_\infty \)-algebra. Define the \(R[s,s^{-1}]\)-linear \(A_\infty \)-algebra \({\mathcal {C}}^s:= {\mathcal {C}}\otimes _R R[s,s^{-1}]\), where s is of odd degree. The operations are defined by:

$$\begin{aligned} \mathfrak {m}_k^s(\alpha _1, \dots , \alpha _k):= s^{2-k}\mathfrak {m}_k(\alpha _1, \dots , \alpha _k), \end{aligned}$$
(96)

and extending s-linearly.

Lemma 3.21

Define \(Gr: hom({\mathcal {C}}^s, {\mathcal {C}}^s) \rightarrow hom({\mathcal {C}}^s, {\mathcal {C}}^s)\) by setting \(Gr(s^ka):= ks^ka\) for \(a \in hom({\mathcal {C}},{\mathcal {C}})\). This makes \({\mathcal {C}}^s\) a \({\mathbb {Z}}\)-graded algebra. In particular, by defining \(E = \frac{s}{2}\frac{d}{ds} \in Der_{{\mathbb {K}}} R[s,s^{-1}]\), \({\mathcal {C}}^s\) is an Euler-graded \(A_\infty \)-algebra.

Remark 3.22

The \(A_\infty \)-algebra \({\mathcal {C}}^s\) is also used in [5, section 3.1] to define the connection in the u direction.

Remark 3.23

The deformation \({\mathcal {C}}^s\) is canonical in the following sense: an \(A_\infty \)-morphism \(F: {\mathcal {C}}\rightarrow \mathcal {D}\) induces an \(A_\infty \)-morphism \(F^s: {\mathcal {C}}^s \rightarrow \mathcal {D}^s\) given by

$$\begin{aligned} F^s_k(\alpha _1, \dots , \alpha _k) = s^{1-k}F_k(\alpha _1, \dots , \alpha _k). \end{aligned}$$
(97)

This morphism is Euler-graded.

\({\mathcal {C}}^s\) is Euler-graded, so naturally comes with a u-connection \(\nabla ^s_{\frac{\partial }{\partial u}} = \frac{1}{2u}Gr - \frac{1}{u}\nabla ^{GGM}_E\) (see Definition 2.6). Define the u-connection on \(HC^-({\mathcal {C}})\) to be the restriction to \(s=1\) of \(\nabla ^s_{\frac{\partial }{\partial u}}\). One can check that indeed:

$$\begin{aligned} \nabla _{\frac{\partial }{\partial u}} = \frac{\partial }{\partial u} + \frac{\Gamma }{2u} + \frac{i\lbrace \mathfrak {m}' \rbrace }{2u^2}, \end{aligned}$$
(98)

where \(\mathfrak {m}' = \sum _k (2-k)\mathfrak {m}_k\). Call this the canonical u-connection associated to an \(A_\infty \)-algebra. This makes \(HC^-_*({\mathcal {C}})\) into a pre-TE-structure.

Remark 3.24

In the deformation \({\mathcal {C}}^s\), s has odd degree. We can also define the \(R[e,e^{-1}]\)-linear \(A_\infty \)-algebra \(\mathcal {C}^e:= {\mathcal {C}}\otimes _R R[e,e^{-1}]\), where e is of even degree. The operations are defined by:

$$\begin{aligned} {\mathfrak {m}}_k^e(a_1,\dots , a_k) = e^\frac{2 - k - |{\mathfrak {m}}_k(a_1, \dots , a_k)| + \sum _i |a_i|}{2} {\mathfrak {m}}_k(a_1,\dots ,a_k). \end{aligned}$$
(99)

Here |a| is 0 if a has even degree or 1 if a has odd degree. Note that we can divide by 2 because \({\mathfrak {m}}\) is \(\mathbb {Z}/2\)-graded. This is Euler-graded with \(E = e\partial _e\), grading operator \(Gr(e^ka) = (2k+|a|)e^ka\).

Lemma 3.25

Let \({\mathcal {C}}\) be an Euler-graded \(A_\infty \)-algebra over R with grading Gr and Euler-vector field E. Then the canonical u-connection agrees up to homotopy with the u-connection coming from the Euler-grading.

Proof

Choose an R-basis for all morphism spaces. This defines an operator

$$\begin{aligned} deg:= Gr - 2E:hom({\mathcal {C}},{\mathcal {C}}) \rightarrow hom({\mathcal {C}},{\mathcal {C}}), \end{aligned}$$
(100)

Let \(\nabla _{\frac{d}{du}}\) denote the canonical u-connection. The u-connection defined using the Euler grading is given by

$$\begin{aligned} \widetilde{\nabla }_{\frac{d}{du}} = \frac{Gr^{-}}{2u} - \nabla ^{GGM}_E. \end{aligned}$$
(101)

Using the definition of deg we can rewrite this as:

$$\begin{aligned} \widetilde{\nabla }_{\frac{d}{du}} = \frac{d}{du} + \frac{\Gamma +\mathcal {L}_{deg}}{2u} + \frac{i\{ E(\mathfrak {m}) \}}{u^2}. \end{aligned}$$
(102)

The properties of E and Gr show:

$$\begin{aligned}{}[{\mathfrak {m}},deg] = 2E({\mathfrak {m}})-{\mathfrak {m}}'. \end{aligned}$$
(103)

By the Cartan homotopy formula 3.15, we thus have:

$$\begin{aligned} \widetilde{\nabla }_{\frac{d}{du}} = \nabla _{\frac{d}{du}} + u^{-2}[i\{deg\},b+uB]. \end{aligned}$$
(104)

\(\square \)

In particular, if we define a u-connection \(\widetilde{\nabla }_{\frac{d}{du}}\) on \(HC^-({\mathcal {C}})\) by restricting the connection \(\nabla ^e\) coming from the Euler-grading on \({\mathcal {C}}^e\) to \(e = 1\), then \(\widetilde{\nabla }_{\frac{d}{du}}\) agrees with the canonical u-connection.

Lemma 3.26

Let \(F: \mathcal {C} \rightarrow \mathcal {D}\) be an R-linear \(A_\infty \)-morphism. Then \(F_*: HC_*^-({\mathcal {C}}) \rightarrow HC^-_*(\mathcal {D})\) is a morphism of pre-TE-structures.

Proof

Let \(F^s: \mathcal {C}^s \rightarrow \mathcal {D}^s\) be the induced Euler-graded morphism. Now apply Theorem 3.17 to \(F^s\) to find that it respects \(\nabla ^{GGM}_E\) up to homotopy. As we also have that \([F^s,Gr] = 0\), we find that \(F^s\) respects \(\nabla _{\partial _u}\) up to homotopy. Restriction to \(s=1\) shows the result. \(\square \)

We finish this section with a comparison between the E-structures associated to a weakly curved \(A_\infty \)-algebra and its uncurved associated \(A_\infty \)-algebra. We use this to conclude that the eigenvalue decomposition of the negative cyclic homology is trivial. For simplicity, here we assume A is a \(\mathbb {C}\)-linear \(A_\infty \)-algebra. Suppose that \((A,\mathfrak {m})\) is strictly unital and weakly curved, i.e. \(\mathfrak {m}_0 = w \cdot {\textbf{e}}\) for some \(w \in \mathbb {C}\). From \((A,\mathfrak {m})\) we can then obtain an uncurved \(A_\infty \)-algebra by setting \(\overline{\mathfrak {m}}_k = \mathfrak {m}_k\) for \(k\ge 1\), and \(\overline{\mathfrak {m}}_0 = 0\).

Lemma 3.27

\((A,\overline{\mathfrak {m}})\) is a unital, uncurved \(A_\infty \)-algebra with:

$$\begin{aligned} HH_*(A,\mathfrak {m})\cong & {} HH_*(A,\overline{\mathfrak {m}}),\; HH^*(A,\mathfrak {m}) \cong HH^*(A,\overline{\mathfrak {m}}) \text { and } HC^{-}_*(A,\mathfrak {m}) \nonumber \\\cong & {} HC^{-}_*(A,\overline{\mathfrak {m}}). \end{aligned}$$
(105)

Furthermore, there exist an isomorphism of pre-E-structures:

$$\begin{aligned} (HC^-_*(A,\mathfrak {m}),\nabla ^{\mathfrak {m}}) \cong (HC^-_*(A,\overline{\mathfrak {m}}),\nabla ^{\overline{\mathfrak {m}}}) \otimes \mathcal {E}^{-\frac{w}{u}}. \end{aligned}$$
(106)

Here on both sides the connection \(\nabla \) denotes the canonical connection defined above.

Proof

The Hochschild differentials satisfy \(b = {\overline{b}}\) as we are working with reduced chains. \(B = {\overline{B}}\) by definition. \(\mathfrak {m}' = \overline{\mathfrak {m}}' + 2w\cdot {\textbf{e}}\), and then from the fact that \({\textbf{e}}\) is a strict unit, we get that \(b^{1,1}(\mathfrak {m}',\_) = {\overline{b}}^{1,1}(\overline{\mathfrak {m}},\_) + 2w\cdot Id\). Furthermore \(B^{1,1}(\mathfrak {m}',\_) = {\overline{B}}^{1,1}(\overline{\mathfrak {m}}',\_)\) by definition of the reduced chains. The result then follows. \(\square \)

The residue of the connection \(\nabla _{\frac{d}{du}}\) is the map \(b^{1,1}(\mathfrak {m}', \_): HC^-(A)/uHC^-(A) = HH_*(A) \rightarrow HH_*(A)\). The following lemma shows that this decomposition is trivial, with the only eigenvalue being w. See also [34, Lemma 2.4] and [24, Section 2.2.7].

Lemma 3.28

Let \((A,\mathfrak {m})\) be a weakly curved \(A_\infty \)-algebra with finite dimensional Hochschild homology. Then the operator \(b^{1,1}(\mathfrak {m}',\_) - w\cdot Id: HH_*(A) \rightarrow HH_*(A)\) is nilpotent.

Proof

Lemma 3.27 allows us to reduce this to the uncurved case. Then, on the chain level, \(b^{1,1}(\mathfrak {m}',\_): CC_*(A) \rightarrow CC_*(A)\) reduces the length of the chain by at least 1. Take a basis for \(HH_*(A)\) and pick representatives \(\alpha _i \in CC_*(A)\). Let \(N = \max _i length(\alpha _i)\), then \((b^{1,1}(\mathfrak {m}',\_))^{N+1} = 0 \) on \(HH_*(A)\). \(\square \)

Corollary 3.29

Let A be a \({\mathbb {C}}\)-linear, strictly unital and weakly curved \(A_\infty \)-algebra with curvature \(w\cdot {\textbf{e}}\). Assume the Hodge-de Rham spectral sequence of A degenerates. Then, in the eigenvalue decomposition of Theorem 2.12, the E-structure \(HC^-_*(A)\) has just a single summand associated to the eigenvalue w.

For future reference, we rephrase a result by Amorim and Tu [1, Corollary 3.8] in the language of TEP-structures:

Theorem 3.30

If A is an n-dimensional, strictly unital, cyclic, \(\mathbb {Z}/2\)-graded, smooth and finite dimensional \(A_\infty \)-algebra with \(HH^*(A)\) semi-simple, then \(HC_*^-(A)\) is a semi-simple TEP-structure.

4 Cyclic open–closed map respects connections

4.1 Coefficient rings

Consider the Novikov ring

$$\begin{aligned} \Lambda&= \left\{ \sum _{i = 0}^{\infty } a_iQ^{\lambda _i} | a_i \in \mathbb {C}, \; \lambda _i \in \mathbb {R}_{\ge 0},\; \lim _{i\rightarrow \infty } \lambda _i = \infty \right\} . \end{aligned}$$
(107)

Let Q have degree 0. For a \({\mathbb {Z}}\)-graded \({\mathbb {C}}\)-vector space U, let \({\mathbb {C}}[[U]]\) be the ring of formal functions on the completion of U at the origin. Explicitly, let \(\{v_i\}_{i \in I}\) be a homogeneous basis for U, and \(\{v_i^*\}_{i \in I}\) the dual basis for \(U^*\). Let \(\{t_i\}_{i \in I}\) be formal variables of degree \(-|v_i|\), then we have an isomorphism:

$$\begin{aligned} {\mathbb {C}}[[t_i]]_{i \in I}&\cong {\mathbb {C}}[[U]],\nonumber \\ t_i&\mapsto v_i^*. \end{aligned}$$
(108)

Each formal vector field \(v \in {\mathbb {C}}[[U]] \otimes U\) on U can be viewed as a derivation \(\partial _v: {\mathbb {C}}[[U]] \rightarrow {\mathbb {C}}[[U]]\). In coordinates, if \(v = \sum _i f_i v_i\), for some \(f_i \in {\mathbb {C}}[[U]]\), then \(\partial _v = \sum _i f_i \partial _{t_i}\). Define the vector fields

$$\begin{aligned} \Gamma _U = \sum _i t_i \partial _{t_i} \text { and } E_U = \sum _i \frac{deg(t_i)}{2}t_i\partial _{t_i}. \end{aligned}$$
(109)

These are independent of the chosen basis.

For \(l \in \mathbb {Z}\), let U[l] denote the graded vector space with \(U[l]^i = U^{i+l}\). Then set:

$$\begin{aligned} Q_U:= \Lambda [[U[2]]]. \end{aligned}$$
(110)

Following [47], define the valuation \(\zeta _Q: Q_U \rightarrow {\mathbb {R}}_{\ge 0}\) by:

$$\begin{aligned} \zeta _Q\left( \sum _{j = 0}^{\infty } a_jQ^{\lambda _j} \prod _{i \in I} t_i^{l_{ij}}\right) = \inf _{\begin{array}{c} j \\ a_j\ne 0 \end{array}} (\lambda _j + \sum _{i \in I} l_{ij}). \end{aligned}$$
(111)

Let \(\mathcal {I}_U = \{ f \in Q_U | \zeta _Q(f) > 0 \} \subset Q_U\).

To account for gradings, we will also make use of the ‘universal Novikov ring’: \(\Lambda ^{e}:= \Lambda [e,e^{-1}]\), where e has degree 2. Let \(Q^{e}_U\) be defined using \(\Lambda ^e\) instead of \(\Lambda \).

Remark 4.1

A lot of our work is based on [47]. They use a different Novikov ring, more commonly used in Gromov–Witten theory. Instead of taking series in \(Q^{\mathbb {R}}\) they take series with terms \(T^{\beta }\) for \(\beta \in H_2(X,L)\). For them the monomial \(T^\beta \) has degree \(\mu (\beta )\), where \(\mu : H_2(X,L) \rightarrow \mathbb {Z}\) is the Maslov index. The graded map \(T^{\beta } \mapsto Q^{\omega (\beta )}e^{\mu (\beta )/2}\), allows us to compare their Novikov ring with the universal Novikov ring \(\Lambda ^e\). Note that \(\mu (\beta ) \in 2\mathbb {Z}\) as we assume our Lagrangian is orientable.

4.2 Quantum TE-structure

Let \((X,\omega )\) be a symplectic manifold and let \(U \subset H^*(X;\mathbb {C})\) be a graded \({\mathbb {C}}\)-vector subspace. For any ring R, let \(A^*(X;R)\) denote the space of differential forms on X with coefficients in R.

Definition 4.2

A bulk-deformation parameter over U is an element \(\gamma \in A^*(X;\mathcal {I}_U)\) with \(d\gamma = 0\), \(|\gamma | = 2\) and \([\gamma ] = \Gamma _U \in Q_U \otimes U\).

In Sect. 4.3 we will outline how a general proof that \({\mathcal{O}\mathcal{C}}^-\) respects T-structures can be upgraded to a result about TE-structures. In order for this to work, the T-structure needs to know about deformation in the direction of \(c_1(X)\), the first Chern class of X. To that end we make the following assumption:

Assumption 4.3

We assume there exists a \(Y\in Der_{\Lambda ^e} Q^e_U\) be such that \([Y(\gamma )] = c_1 \in H^*(X;Q_U^e)\). In this case, note that as \(|{\gamma }| = 2\), this implies \(|Y| = 0\).

Let \(\gamma \) be a bulk-deformation parameter over U. We now consider the quantum cohomology \(QH^*(X;Q^e_U)\). As a vector space this is just \(H^*(X;Q^e_U)\), but the product is given by the bulk-deformed quantum cup product \(\eta _1 \star _\gamma \eta _2\). A general reference for the construction of the quantum cup product is [29], however, our coefficient ring includes the universal Novikov parameter e, so we sketch how to modify the definition. See also Definition 4.35 for a construction in our specific setup. Recall from [30] that the quantum cup product is defined as a sum over curve classes \(\beta \in H_2(X)\):

$$\begin{aligned} \eta _1 \star \eta _2 = \sum _\beta Q^{\omega (\beta )}(\eta _1 \star \eta _2)_{\beta }. \end{aligned}$$
(112)

Here \((\eta _1 \star \eta _2)_{\beta }\) is defined by the equation

$$\begin{aligned} \int _X \eta \cup (\eta _1 \star \eta _2)_{\beta } = GW^\beta _{0,3}(\eta ,\eta _1,\eta _2) \text { for all } \eta \in H^*(X), \end{aligned}$$
(113)

where \(GW^\beta _{0,3}\) denotes the genus 0, 3 point Gromov–Witten invariant in curve class \(\beta \). One can then extend this definition to take into account bulk deformations \({\gamma }\), to obtain the product \(\star _{\gamma }\) on quantum cohomology \(QH^*(X;Q_U)\). We then define (see also Definition 4.35) the product on \(QH^*(X;Q_U^e)\) by:

$$\begin{aligned} \eta _1 \star _\gamma \eta _2 = \sum _\beta Q^{\omega (\beta )}e^{c_1(\beta )}(\eta _1 \star _\gamma \eta _2)_{\beta }, \end{aligned}$$
(114)

where \(c_1 = c_1(TX)\) is the first Chern class.

Definition 4.4

The quantum T-structure over \(Q^e_U \supset \Lambda ^e\) is given as a \(Q_U^e[[u]]\)-module by:

$$\begin{aligned} QH^*(X;Q^e_U)[[u]]. \end{aligned}$$
(115)

For \(v \in Der_{\Lambda }Q_U\) the quantum connection is defined by:

$$\begin{aligned} \nabla _v \eta = v(\eta ) - u^{-1}v(\gamma ) \star _\gamma \eta . \end{aligned}$$
(116)

We now wish to extend the quantum T-structure to be defined over \(Q^e_U \supset \Lambda \). To this end observe that \(Der_\Lambda (Q_U^e) = Q_U^e \otimes _{Q_U} Der_\Lambda Q_U \oplus Q^e_U\langle \partial _e \rangle \). Extend the connection by setting:

$$\begin{aligned} \nabla _{e\partial _e} \eta = e\partial _e (\eta ) - u^{-1} c_1 \star _\gamma \eta . \end{aligned}$$
(117)

Lemma 4.5

These above definitions make \(QH^*(X;Q_U^e)[[u]]\) into a T-structure over \(Q_U^e \supset \Lambda \).

Proof

The verification that \([\nabla _v,\nabla _w] = \nabla _{[v,w]}\) for \(v,w \in Der_{\Lambda } Q_U\) is standard, so we will not do it here. Instead we verify that \([\nabla _v,\nabla _{e\partial _e]} = 0\). The divisor equation for closed Gromov–Witten invariants shows that for \(v \in Der_\Lambda Q_U\) we have:

$$\begin{aligned} e\partial _e (v(\gamma ) \star _\gamma \eta ) = c_1 \star _\gamma v(\gamma ) \star _\gamma \eta + v({\gamma }) \star _\gamma e\partial _e (\eta ). \end{aligned}$$
(118)

We also find that:

$$\begin{aligned} v(c_1 \star _\gamma \eta ) = c_1 \star _\gamma v({\gamma }) \star _\gamma \eta + c_1 \star _\gamma v(\eta ). \end{aligned}$$
(119)

A direct verification then shows that \([\nabla _{e\partial _e},\nabla _v] = 0\) holds. \(\square \)

We will now define endow \(QH^*(X;Q_U^e)[[u]]\) with an Euler-graded T-structure.

Definition 4.6

Define the Euler vector field by \(E = e\partial _e + E_U\). Define the grading operator

$$\begin{aligned} Gr^-: QH^*(X;Q_U^e)[[u]] \rightarrow QH^*(X;Q_U^e)[[u]] \end{aligned}$$
(120)

by taking into account the cohomological degrees, the grading on the coefficient rings and the degree of u, but with the grading shifted down by n so that, for \(\eta \in H^*(X;{\mathbb {C}})\) and \(f \in Q_U^e[[u]]\), we have:

$$\begin{aligned} Gr^-(f\eta ) = (|f| + |\eta | - n)f\eta = (2u\partial _u + 2E)(f){\alpha }+ 2f\mu ({\alpha }). \end{aligned}$$
(121)

where \(\mu : H^p(X;Q_U^e) \rightarrow H^p(X;Q_U^e)\) is given by \(\mu (\eta ) = \frac{p-n}{2}\eta \).

A short computation then shows:

Lemma 4.7

The above definitions make \(QH^*(X;Q_U^e)[[u]]\) into an Euler-graded T-structure.

As the quantum T-structure is Euler-graded, Definition 2.6 endows it with a connection in the u-direction:

$$\begin{aligned} \nabla _{\partial _u} = \frac{Gr^-}{2u} - u^{-1}\nabla _E. \end{aligned}$$
(122)

This makes \(QH^*(X;Q_U^e)[[u]]\) a TE-structure. Writing out the definitions of \(Gr^-\) and E yields the formula:

$$\begin{aligned} \nabla _{\partial _u} \eta = \partial _u(\eta ) + u^{-1}\mu (\eta ) + u^{-2}(c_1 + E_U(\gamma ))\star _\gamma \eta . \end{aligned}$$
(123)

Remark 4.8

Defining \(\nabla _{Q\partial _Q}\eta := Q\partial _Q (\eta ) - u^{-1} [\omega ] \star _\gamma \eta \) extends the connection to include the Q-direction, to obtain a TE-structure over \(Q_U^e \supset {\mathbb {C}}\). We do not use this connection.

Remark 4.9

We can use the same formula 123 to define a TE-structure on \(QH^*(X;Q_U)[[u]]\). There then is a natural isomorphism of TE-structures over \(Q_U \supset \Lambda \):

$$\begin{aligned} QH^*(X;Q_U^e)[[u]] \otimes _{\Lambda ^e} \Lambda \cong QH^*(X;Q_U)[[u]]. \end{aligned}$$
(124)

Here \(\Lambda \) is considered as a \(\Lambda ^e\) module via the homomorphism \(\Lambda ^e \rightarrow \Lambda \) given by evaluation at \(e=1\).

There are alternative definitions of the quantum connections on \(QH^*(X;Q_U^e)[[u]]\) given by changing the signs:

Definition 4.10

$$\begin{aligned} \nabla ^*_v \eta&= v(\eta ) + u^{-1}v(\gamma ) \star _\gamma \eta \end{aligned}$$
(125)
$$\begin{aligned} \nabla ^*_{e\partial _e} \eta&= e\partial _e (\eta ) + u^{-1}c_1 \star _\gamma \eta \end{aligned}$$
(126)

The alternative connection in the u-direction is given by \(\nabla ^*_{\partial _u} = \frac{Gr^-}{2u} -u^{-1}\nabla ^*_E\). Writing the formulae out we find:

$$\begin{aligned} \nabla ^*_{\partial _u}\eta = \partial _u(\eta ) + u^{-1} \mu (\eta ) - u^{-2} (c_1 + E_U({\gamma })) \star _\gamma \eta . \end{aligned}$$
(127)

Define the Poincaré pairing

$$\begin{aligned} \langle \cdot , \cdot \rangle _X: QH^*(X;Q_U^e) \otimes QH^*(X;Q_U^e) \rightarrow Q_U^e, \end{aligned}$$
(128)

by \(\langle \eta _1, \eta _2 \rangle _X = \int _X \eta _1 \wedge \eta _2\). Now extend the Poincaré pairing u-linearly to a pairing

$$\begin{aligned} \langle \cdot , \cdot \rangle _X: QH^*(X;Q_U^e)[[u]]\otimes QH^*(X;Q_U^e)[[u]] \rightarrow Q_U^e[[u]]. \end{aligned}$$
(129)

We then have:

$$\begin{aligned} \langle \nabla ^*_{v} \eta _1, \eta _2 \rangle _X + (-1)^{|\eta _1||v|} \langle \eta _1, \nabla _{v}\eta _2 \rangle _X = v (\langle \eta _1, \eta _2 \rangle _X), \end{aligned}$$
(130)

for all \(v \in Der_\Lambda Q_U^e[[u]]\).

Remark 4.11

It is customary to extend the Poincaré pairing sesquilinearly to the quantum TE-structure, to obtain a TEP-structure where the polarisation can be matched up with the higher residue pairing on cyclic homology. However, since we don’t mention the polarisation in this paper, we use the u-linear extension as it simplifies the proof that the cyclic open–closed map is a morphism of TE-structures.

4.3 Outline of proof of Theorem 1.7

In this section we give an outline of the proof of Theorem 1.7. We will state sufficient conditions which imply that the cyclic open–closed map respects the connection \(\nabla _{\partial _u}\). We state these conditions in such a way that they should be easy to generalise to different geometric setups.

Let \(U \subset H^*(X;\mathbb {C})\) be a graded vector space, and \({\gamma }\) a bulk-deformation parameter over U satisfying Assumption 4.3. Let \(L \subset X\) be a Lagrangian submanifold. We define an Euler-graded \(A_\infty \)-algebra \(CF^*(L,L)[e]\) over \(Q^e:= Q_U[e,e^{-1}]\) in Sect. 4.7. The Euler vector field is given by \(E = e\partial _e + E_U\), where \(E_U\) is as in Sect. 4.1 and e is of degree 2. The Floer cochain complex \(CF^*(L,L)\) is then defined by restricting to \(e=1\): \(CF^*(L,L):= CF^*(L,L)[e] \otimes _{Q_U^e} Q_U\).

More generally, suppose there exists a bulk-deformed Fukaya category \(Fuk^t(X)\) defined over \(Q_U\). By using e to take into account the Maslov index of holomorphic disks, it should be possible to construct an Euler-graded Fukaya category \(Fuk^t(X)[e]\) over \(Q^e\). In Appendix A, we construct such an Euler-graded Fukaya category geometrically. In the appendix, U will be the 1-dimensional vector-space spanned by the first Chern class.

Assume there exists a cyclic open–closed map

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-: HC^-_*(Fuk^t(X)) \rightarrow QH^*(X;Q_U)[[u]], \end{aligned}$$
(131)

which is the restriction to \(e=1\) of a map

$$\begin{aligned} {\mathcal{O}\mathcal{C}}_e^-: HC^-_*(Fuk^t(X)[e]) \rightarrow QH^*(X;Q_U^e)[[u]]. \end{aligned}$$
(132)

In Sect. 4.8.3, we will construct a cyclic open–closed map by defining a chain level pairing (which we call the cyclic open–closed pairing):

$$\begin{aligned} \langle \cdot , {\mathcal{O}\mathcal{C}}_e^-(\cdot ) \rangle : (C^*(X;Q_U^e)[[u]]) \otimes CC_*^-(Fuk^t(X)[e]) \rightarrow Q_U^e[[u]]. \end{aligned}$$
(133)

We show that it satisfies:

$$\begin{aligned} \langle d\eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle + (-1)^{|\eta |} \langle \eta , {\mathcal{O}\mathcal{C}}_e^-((b+uB)(\alpha )) \rangle = 0. \end{aligned}$$
(134)

This ensures the chain level pairing descends to a pairing:

$$\begin{aligned} \langle \cdot , {\mathcal{O}\mathcal{C}}_e^-(\cdot ) \rangle : QH^*(X;Q_U^e)[[u]] \otimes HC_*^-(Fuk^t(X)[e]) \rightarrow Q_U^e[[u]]. \end{aligned}$$
(135)

We then apply Poincaré duality to the \(QH^*(X;Q^e)\) factor to obtain the map \({\mathcal{O}\mathcal{C}}_e^-\). It is uniquely determined by the property:

$$\begin{aligned} \langle \eta ,{\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle _X = \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle . \end{aligned}$$
(136)

On the left, the pairing is the Poincaré pairing on X, and on the right the pairing is the open–closed pairing.

We expect more generally that given a construction of a chain level cyclic open–closed map, it can be extended to \(CC_*^-(Fuk^t(X)[e])\) by taking the Maslov index of the holomorphic disks into account.

Suppose the cyclic open–closed pairing satisfies the following properties.

Assumptions 4.12

  1. 1.

    For \(v \in Der_{\Lambda ^e} Q_U^e\), there exists a pairing \(\langle \_, G_{v}(\_) \rangle : (C^*(X;Q_U^e)[[u]]) \otimes CC_*^-(Fuk^t(X)[e]) \rightarrow Q_U^e[[u]]\) such that for all \(\eta \) and \({\alpha }\) we have:

    $$\begin{aligned}{} & {} \langle \nabla ^*_{v} \eta , {\mathcal{O}\mathcal{C}}_e^- (\alpha ) \rangle + (-1)^{|\eta ||v|}\langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\nabla _{v} \alpha ) \rangle = v \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right) \nonumber \\{} & {} \qquad +\, u^{-1}\left( \langle d\eta , G_{v} (\alpha ) \rangle + (-1)^{|\eta | + |v|}\langle \eta , G_{v}\left( (b+uB)(\alpha )\right) \rangle \right) . \end{aligned}$$
    (137)

    Here \(\nabla ^*\) is the sign-changed connection from (125). By (130), this implies that \({\mathcal{O}\mathcal{C}}_e^-\) is a morphism of T-structures over \(Q_U^e \supset \Lambda ^e\).

  2. 2.

    There exists a \(\phi \in CC^1(Fuk^t(X)[e])\), such that \(Y(\mathfrak {m}^e) = e\partial _e({\mathfrak {m}}^e) + [{\mathfrak {m}}^e, \phi ]\).

  3. 3.

    For any \(\eta \) and \({\alpha }\) we have:

    $$\begin{aligned} \langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle = \langle \eta , e\partial _e({\mathcal{O}\mathcal{C}}_e^-)({\alpha }) \rangle + \langle {\alpha }, {\mathcal{O}\mathcal{C}}_e^-(\mathcal {L}_\phi (\eta )) \rangle . \end{aligned}$$
    (138)

    Here, for any \(v \in Der_{\Lambda } Q_U^e\), we define:

    $$\begin{aligned} \langle \eta , v({\mathcal{O}\mathcal{C}}_e^-)({\alpha }) \rangle := v(\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle ) - \langle v(\eta ), {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle - (-1)^{|\eta ||v|} \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(v(\alpha )) \rangle .\nonumber \\ \end{aligned}$$
    (139)
  4. 4.

    \({\mathcal{O}\mathcal{C}}_e^-\) respects the Euler-grading on cyclic invariants: \(Gr^{-} \circ {\mathcal{O}\mathcal{C}}_e^- = {\mathcal{O}\mathcal{C}}_e^- \circ Gr^{-}\).

Remark 4.13

Since we construct a cyclic open–closed map via a cyclic open–closed pairing, we have stated the required properties for the latter. An easy modification of the assumptions would allow them to be applied to an open–closed map defined directly as a chain map \(CC_*^-(Fuk^t(X)[e]) \rightarrow C^*(X;Q_U^e)[[u]]\).

Remark 4.14

To define the connections \(\nabla _v\) and \(\nabla _{e\partial _e}\) on cyclic homology, we need to choose a basis for all of the morphism spaces in \(Fuk^t(X)\). Assumption 2 is required to hold with respect to the same bases as used to define the connections. The same holds for assumption 3. On quantum cohomology, we take the derivatives with respect to the standard constant basis (i.e. one in \(H^*(X;{\mathbb {C}})\)).

Theorem 4.15

Suppose Assumptions 4.12 hold, then \(\mathcal{O}\mathcal{C}_e^-\), and hence \({\mathcal{O}\mathcal{C}}^-\), respects \(\nabla _{\partial _u}\) on homology.

First we will show that Assumptions 1, 2 and 3 show that \({\mathcal{O}\mathcal{C}}^e\) respects the connection \(\nabla _{e\partial _e}\). To this end define:

Definition 4.16

Define the pairing \(\langle \eta , G_{e\partial _e}({\alpha }) \rangle \) by

$$\begin{aligned} \langle \eta , G_{e\partial _e}({\alpha }) \rangle = \langle \eta , G_{Y}({\alpha })\rangle + (-1)^{{|\eta |}}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{\phi \}({\alpha })) \rangle \end{aligned}$$

Proposition 4.17

The pairing \(\langle \eta , G_{e\partial _e}({\alpha }) \rangle \) satisfies

$$\begin{aligned}{} & {} \langle \nabla ^*_{e\partial _e} \eta , {\mathcal{O}\mathcal{C}}_e^- (\alpha ) \rangle + \langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\nabla _{e\partial _e} \alpha ) \rangle = e\partial _e \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right) \\{} & {} \quad +\, u^{-1}\left( \langle d\eta , G_{e\partial _e} (\alpha ) \rangle + (-1)^{|\eta |}\langle \eta , G_{e\partial _e}\left( (b+uB)(\alpha )\right) \rangle \right) . \end{aligned}$$

Proof

Applying Assumption 1, with \(v = Y\) yields:

$$\begin{aligned} \begin{aligned}&[t]u^{-1} \langle c_1 \star \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle - u^{-1}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{Y({\mathfrak {m}}^e)\}\alpha ) \rangle = \langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle \\&\quad + u^{-1}\langle d\eta , G_{Y} (\alpha ) \rangle +\, u^{-1}(-1)^{|\eta |}\langle \eta , G_{Y}\left( (b+uB)(\alpha )\right) \rangle . \end{aligned} \end{aligned}$$
(140)

using Assumption 2, this gives:

$$\begin{aligned} \begin{aligned}{}[t] u^{-1} \langle c_1 \star \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle - u^{-1}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{e\partial _e({\mathfrak {m}}^e)\}\alpha ) \rangle -u^{-1}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{[{\mathfrak {m}}^e,\phi ]\}\alpha ) \rangle \\ =\,\langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle + u^{-1}\langle d\eta , G_{Y} (\alpha ) \rangle + u^{-1}(-1)^{|\eta |}\langle \eta , G_{Y}\left( (b+uB)(\alpha )\right) \rangle . \end{aligned} \end{aligned}$$
(141)

By the Cartan homotopy formula (Proposition 3.15) we can rewrite this as:

$$\begin{aligned} \begin{aligned}&[t] u^{-1} \langle c_1 \star \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle - u^{-1}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{e\partial _e({\mathfrak {m}}^e)\}\alpha ) \rangle \\&\quad = \langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle - \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\mathcal {L}_{\phi }\alpha ) \rangle \\&\qquad + u^{-1}\langle d\eta , G_{e\partial _e} (\alpha ) \rangle + u^{-1}(-1)^{|\eta |}\langle \eta , G_{e\partial _e}\left( (b+uB)(\alpha )\right) \rangle . \end{aligned} \end{aligned}$$
(142)

Then apply Assumption 3 to obtain:

$$\begin{aligned} \begin{aligned}&[t] u^{-1} \langle c_1 \star \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle - u^{-1}\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(i\{e\partial _e({\mathfrak {m}}^e)\}\alpha ) \rangle \\&\quad = \langle \eta , e\partial _e({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle + u^{-1}\langle d\eta , G_{e\partial _e} (\alpha ) \rangle \\&\qquad + u^{-1}(-1)^{|\eta |}\langle \eta , G_{e\partial _e}\left( (b+uB)(\alpha )\right) \rangle . \end{aligned} \end{aligned}$$
(143)

This implies the result, by rewriting \(\langle \eta , e\partial _e({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle \) using (139). \(\square \)

Proof of Theorem 4.15

The previous proposition implies \(\mathcal{O}\mathcal{C}^e_-\) respects \(\nabla _{e\partial _e}\) on homology. We will spell this out. By the properties of the Poincaré pairing (Eq. 130):

$$\begin{aligned} \langle \eta ,\nabla _{e\partial _e}({\mathcal{O}\mathcal{C}}_e^-({\alpha })) \rangle _X = e\partial _e(\langle \eta , {\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle ) - \langle \nabla ^*_{e\partial _e} \eta , {\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle . \end{aligned}$$
(144)

The above shows that on homology:

$$\begin{aligned} \langle \eta ,\nabla _{e\partial _e}({\mathcal{O}\mathcal{C}}_e^-({\alpha })) \rangle _X = \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\nabla _{e\partial _e}({\alpha })) \rangle , \end{aligned}$$
(145)

which shows that \({\mathcal{O}\mathcal{C}}_e^-(\nabla _{e\partial _e}({\alpha })) = \nabla _{e\partial _e}({\mathcal{O}\mathcal{C}}_e^-({\alpha }))\) on homology. As \(E = e\partial _e + E_U\), we combine this with Assumption 1, applied to \(v = E_U\) to find that the open–closed map respects \(\nabla _{E}\). Then, as \(\nabla _{\partial _u} = \frac{Gr^{-}}{2u} - u^{-1}\nabla _E\), and using the fact that \({\mathcal{O}\mathcal{C}}_e^-\) respects the Euler-grading (Assumption 4), we find that \({\mathcal{O}\mathcal{C}}_e^-\) respects the connection \(\nabla _{\partial _u}\). The statement about \({\mathcal{O}\mathcal{C}}^-\) follows by restriction to \(e = 1\).

4.4 Regularity assumptions

Let X be a 2n-dimensional symplectic manifold and J be an \(\omega \)-tame almost complex structure on X. Let \(L \subset X\) be an oriented Lagrangian equipped with a \(U(\Lambda )\)-local system and a relative spin structure \(\mathfrak {s}\). For us a relative spin structure comes with a choice of element \(w_{\mathfrak {s}} \in H^2(X;{\mathbb {Z}}/2)\) such that \(w_{\mathfrak {s}}|_L = w_2(TL) \in H^2(L;{\mathbb {Z}}/2)\).

For \(l \ge 0\), let \(\mathcal {M}_{l+1}(\beta )\) be the moduli space of stable J-holomorphic spheres with \(l+1\) marked points in homology class \(\beta \in H_2(X,\mathbb {Z})\). Let

$$\begin{aligned} ev_j^\beta : \mathcal {M}_{l+1}(\beta ) \rightarrow X \end{aligned}$$
(146)

be the evaluation map at the j’th marked point. For \(k\ge -1\), \(l\ge 0\), let \(\mathcal {M}_{k+1,l}(\beta )\) be the moduli space of J-holomorphic stable maps \((\mathbb {D},S^1) \rightarrow (X,L)\) in homology class \(\beta \in H_2(X,L)\) with one boundary component, \(k+1\) anti-clockwise ordered boundary marked points, and l interior marked points. Let

$$\begin{aligned} evb_i^\beta : \mathcal {M}_{k+1,l}(\beta ) \rightarrow L \text { and } evi_j^\beta : \mathcal {M}_{k+1,l}(\beta ) \rightarrow X \end{aligned}$$
(147)

be the evaluation maps at the i’th boundary and j’th interior marked points respectively. The relative spin structure determines an orientation on the moduli spaces \(\mathcal {M}_{k+1,l}(\beta )\), see [7, Chapter 8].

We will also need a moduli space of disks with a horocyclic constraint. Recall that a horocycle in a disk is given by a circle tangent to the boundary. These moduli spaces are similar to the ones used in [44, Chapter 3], where some of the marked points are constrained to lie on a geodesic in \(\mathbb {D}\). Our definition is entirely analogous, except that we replace ‘geodesic’ with ‘horocycle’. Let the smooth locus of \(\mathcal {M}_{k+1,l; \perp _i}(\beta ) \subset \mathcal {M}_{k+1,l}(\beta )\) be the subset defined by requiring the first and second interior marked points \(w_1\) and \(w_2\) to lie at \(-t\) and t respectively for \(t \in (-1,1)\) and fixing the i’th boundary point \(z_i\) at \(-i\). Equivalently, we require that \(z_i\), \(w_1\), \(w_2\) lie on a horocycle in anti-clockwise ordering. This moduli space also appeared in [13], where it was used to show that the closed–open map is an algebra homomorphism.

We now give a more formal definition of the moduli space \(\mathcal {M}_{k+1,l; \perp _i}(\beta )\) as a fibre product of known spaces. Consider the forgetful map \(\mathcal {M}_{k+1,l}(\beta ) \rightarrow {\mathcal {M}}_{1,2} = D^2\), only remembering the i’th boundary marked point, and the first interior marked point. Here the identification \({\mathcal {M}}_{1,2} \cong D^2\) is achieved by using an automorphism of the disk to map the boundary marked point to \(-i\), and the interior marked point to 0. Consider the inclusion \(I \hookrightarrow D^2\) given by the arc of the horocycle through \(-i\) and 0 with negative real part. This is a circle of radius \(\frac{1}{2}\) centred at \(-\frac{i}{2}\). The condition on the order of the marked points means that second interior lies on the semi-circle with negative real part. We then define:

$$\begin{aligned} \mathcal {M}_{k+1,l; \perp _i}(\beta ) = I \times _{D^2} \mathcal {M}_{k+1,l}(\beta ). \end{aligned}$$
(148)

Take the orientation on I to be the positive orientation, so that \(\partial I = \{1\} -\{ 0\}\). The orientation on \(\mathcal {M}_{k+1,l; \perp _i}(\beta )\) is then defined by the fibre-product orientation, as in [46, Section 2.2].

We assume the following:

Assumptions 4.18

  1. 1.

    \(\mathcal {M}_{l+1}(\beta )\) is a smooth orbifold with corners.

  2. 2.

    \(ev_0\) is a proper submersion.

  3. 3.

    \(\mathcal {M}_{k+1,l}(\beta )\) is a smooth orbifold with corners.

  4. 4.

    \(evb_0^\beta \) is a proper submersion.

  5. 5.

    \(\mathcal {M}_{k+1,l; \perp _i}(X,\beta )\) is a smooth orbifold with corners.

  6. 6.

    \(evb_0^\beta |_{\mathcal {M}_{k+1,l; \perp _i}(X,\beta )}\) is a proper submersion.

We will now show these assumptions hold in the following setup:

Lemma 4.19

The above assumptions hold for \(L \subset X\) a Lagrangian and a complex structure J with the following properties:

  • J is integrable.

  • There exists a Lie group \(G_X\) acting J-holomorphically and transitively on X.

  • There exist a Lie subgroup \(G_L \subset G_X\) whose action restricts to a transitive action on L.

Proof

This argument is the same as [44, Section 1.3.12], but for horocyclic rather than geodesic constraints. For Assumptions 4.18.1 and 4.18.2, [29, Proposition 7.4.3] show that if the above properties hold, all stable holomorphic maps in \(\mathcal {M}_{l+1}(\beta )\) are regular, it then follows from [35] that this space is a smooth orbifold with corners. As \(G_X\) acts on \(\mathcal {M}_{l+1}(\beta )\), \(G_X\) acts transitively on X, and \(ev_0\) is equivariant with respect to this action, \(ev_0\) is a proper submersion.

Solomon and Tukachinsky show Assumptions 4.18.3 and 4.18.4 hold in this situation by adapting the arguments for closed Riemann surfaces to Riemann surfaces with boundary (see [47, Remark 1.6]).

Furthermore, \(\mathcal {M}_{k+1,l; \perp _i}(\beta )\) is a smooth orbifold with corners as the maps \(I \hookrightarrow D^2\) and \(\mathcal {M}_{k+1,l}(\beta ) \rightarrow D^2\) are transverse (in the sense of [22, Definition 6.1]). This is clear for any curve mapping to the part of the horocycle not intersecting \(S^1\). The remaining case, for a point in the pre-image of \(-i \in D^2\) is more involved.

To this end, let \(u \in {\mathcal {M}}_{k+1,l}(\beta )\) be a pre-image of the point \(-i \in D^2\). Let \(z_0\) denote the boundary marked point, and \(w_1, w_2\) the interior marked points partaking in the horocycle. The stable curve underlying u has at least 3 components, with one ‘middle’ disk component which has the marked point \(z_0\). There is then additionally a ‘right’ configuration which has the point \(w_1\), and a ‘left’ one which has \(w_2\). Both the left and the right configurations either are disks connected to the middle disks, or have a disk component which connects them to the middle disk. Let \(p_1\) and \(p_2\) denote the boundary points where the right and left configurations are connected to the middle disk. Now use a gluing argument to open up the point \(p_1\). By stability, the new middle disk (which might have \(w_1\) on it) must now have at least 4 boundary marked points, or 2 boundary and 1 interior marked point. A one parameter family can then be constructed by moving the point \(z_0\) towards \(p_2\). The original map u arises from the case where \(z_0\) merges with \(p_2\) and a new disk bubble is formed. The map u lies in the interior of this one parameter family (as we can move \(z_0\) to either side of \(p_2\)), so it lies in the same stratum as the family. The forgetful map to \(D^2\) is given by the location of \(p_2\) in this case, so we see that the map at the level of tangent spaces is indeed surjective onto \(T_{-i}S^1\) as required.

Finally note that \(G_L\) acts on \(\mathcal {M}_{k+1,l; \perp _i}(X,\beta )\), and the evaluations maps are equivariant. As \(G_L\) acts transitively on L, \(evb_0^\beta |_{\mathcal {M}_{k+1,l; \perp _i}(X,\beta )}\) is a proper submersion. \(\square \)

Example 4.20

The simplest example is \((\mathbb{C}\mathbb{P}^n, T_{cl})\), where \(T_{cl}\) denotes the Clifford torus. Here the groups acting are \(G_X = SU(n+1)\) and \(G_L = T^{n}\). Other examples (see [47, Example 1.5]) are \((\mathbb{C}\mathbb{P}^n, \mathbb{R}\mathbb{P}^n)\), or more generally flag varieties and Grassmannians with L being the real locus. In these cases \(G_X\) is a complex matrix group and \(G_L\) the real subgroup. Another class of examples are the quadric hypersurfaces with real locus \(S^n\):

$$\begin{aligned} X_{2,n} = \left\{ \sum _{i=0}^n z_i^2 = z_{n+1}^2 \right\} \subset \mathbb{C}\mathbb{P}^{n+1}. \end{aligned}$$
(149)

In this case, \(G_X = SO(n+1, {\mathbb {C}})\) and \(G_L = SO(n+1, {\mathbb {R}})\), with the group acting in the obvious manner on the first \(n+1\) coordinates.

4.5 \(\mathfrak {q}\)-operations

This section follows [44, 47] closely. Let \(L \subset X\) be as in the previous section. Let

$$\begin{aligned} hol: H_1(L,\mathbb {Z}) \rightarrow U(\Lambda ) \end{aligned}$$
(150)

denote the monodromy representation of the local system on L.

Let \(A^*(L)\) denote differential forms on L with coefficients in \(\mathbb {C}\), and similarly for X. For \({\alpha }\in A^*(L)\), let \(|{\alpha }|\) denote its degree as a differential form, and similarly for differential forms on X. Also, let \(|{\alpha }|':= |{\alpha }| - 1\), and for an ordered set \({\alpha }= ({\alpha }_1, \dots , {\alpha }_k)\), write \(\epsilon ({\alpha }):= \sum _i |{\alpha }_i|' \in \mathbb {Z}\).

For \(k,l \ge 0\) and \(\beta \in H_2(X,L)\) with \((k,l,\beta ) \notin \lbrace (1,0, \beta _0), (0,0,\beta _0) \rbrace \), [47] define operations:

$$\begin{aligned} \mathfrak {q}_{k,l}^{ST,\beta }:A^*(L)^{\otimes k} \otimes A^*(X)^{\otimes l} \rightarrow A^*(L). \end{aligned}$$
(151)

We extend their definition to take into account the local system and set:

$$\begin{aligned} \mathfrak {q}_{k,l}^{ST,\beta }(\alpha _1 \otimes \dots \otimes \alpha _k; \gamma _1 \otimes \dots \otimes \gamma _l):= & {} (-1)^{\zeta (\alpha )} hol(\partial \beta ) (evb_0^\beta )_* \nonumber \\{} & {} \left( \bigwedge _{j=1}^l (evi^\beta _j)^* \gamma _j \wedge \bigwedge _{i=1}^k (evb_i^\beta )^* \alpha _j \right) .\nonumber \\ \end{aligned}$$
(152)

Here \(\zeta (\alpha ) = 1 + \sum _{j=1}^{k} j|\alpha _j|'\). The special cases are as follows:

$$\begin{aligned}&\mathfrak {q}_{0,0}^{ST,\beta } := - (evb_0^\beta )_* 1 \in A^*(L), \end{aligned}$$
(153)
$$\begin{aligned}&\mathfrak {q}_{1,0}^{ST,\beta _0} := d\alpha ,\end{aligned}$$
(154)
$$\begin{aligned}&\mathfrak {q}_{0,0}^{ST,\beta _0} := 0. \end{aligned}$$
(155)

For the cleanest staments, we will use a sign convention differing from [47].

Definition 4.21

Let the operations \({\mathfrak {q}}_{k,l}^\beta : A^*(L)^{\otimes k} \otimes A^*(X)^{\otimes l} \rightarrow A^*(L)\) be defined by:

$$\begin{aligned} {\mathfrak {q}}_{k,l}^\beta ({\alpha }_1, \dots , {\alpha }_k;{\gamma }_1, \dots , {\gamma }_l) = (-1)^{\dagger ({\alpha }) + k-1} {\mathfrak {q}}_{k,l}^{ST,\beta }({\alpha }_k, \dots , {\alpha }_1;{\gamma }_1, \dots , {\gamma }_l).\nonumber \\ \end{aligned}$$
(156)

Here, for \({\alpha }= ({\alpha }_1, \dots , {\alpha }_k)\), we set \(\dagger ({\alpha }) = \sum _{1 \le i <j \le k} |{\alpha }_i|'|{\alpha }_j|'\). This is the sign coming from reversing the order of \({\alpha }\).

[44] also define closed operations:

$$\begin{aligned} \mathfrak {q}_{\emptyset , l}^{ST,\beta }: A^*(X)^{\otimes l} \rightarrow A^*(X), \end{aligned}$$
(157)

by

$$\begin{aligned} \mathfrak {q}_{\emptyset , l}^{ST,\beta }(\gamma _1, \dots , \gamma _l):= (-1)^{w_{\mathfrak {s}}(\beta )} (ev_0^\beta )_*\left( \bigwedge _{j=1}^l (ev_j^\beta )^* \gamma _j\right) , \end{aligned}$$
(158)

with special cases:

$$\begin{aligned} \mathfrak {q}_{\emptyset , 1}^{ST,\beta _0}:= 0, \; \mathfrak {q}_{\emptyset , 0}^{ST, \beta _0}:= 0. \end{aligned}$$
(159)

We use these operations, without any sign change, so that \({\mathfrak {q}}_{\emptyset ,l} = {\mathfrak {q}}^{ST}_{\emptyset ,l}\). The quantum product \(\_ \star \_: A^*(X) \otimes A^*(X) \rightarrow A^*(X)\), is then given by

$$\begin{aligned} \gamma _1 \star \gamma _2 = \mathfrak {q}_{\emptyset ,2}(\gamma _1, \gamma _2). \end{aligned}$$
(160)

We also define new operations coming from the moduli spaces with horocyclic constraints \({\mathcal {M}}_{k+1,l,\perp _i}(\beta )\). We first define these using sign conventions similar to [47].

Definition 4.22

Let \(\mathfrak {q}^{ST,\beta }_{k,l;\perp _i}: A^*(L)^{\otimes k} \otimes A^*(X)^{\otimes l} \rightarrow A^*(L),\) be defined by

$$\begin{aligned} \mathfrak {q}^{ST,\beta }_{k,l;\perp _i}(\alpha _1, \dots , \alpha _k;\gamma _1, \dots , \gamma _l)= & {} (-1)^{\zeta (\alpha ) + {\zeta _{\perp }(\alpha ;\gamma )}}hol(\partial \beta ) (evb_0^\beta )_*\nonumber \\{} & {} \quad \left( \bigwedge _{j=1}^{l} (evi_j^{\beta })^* \gamma _j \wedge \bigwedge _{j=1}^{k} (evb_j^{\beta })^*\alpha _j\right) , \end{aligned}$$
(161)

where \(\zeta _{\perp }(\alpha ;\gamma ) = |\alpha |' + |\gamma | + n\). Then, we modify the sign convention as before. We set

$$\begin{aligned} {\mathfrak {q}}_{k,l,\perp _i}^\beta ({\alpha }_1, \dots , {\alpha }_k;{\gamma }_1, \dots , {\gamma }_l) = (-1)^{\dagger ({\alpha }) + k-1} {\mathfrak {q}}_{k,l,\perp _{k+1-i}}^{ST,\beta }({\alpha }_k, \dots , {\alpha }_1;{\gamma }_1, \dots , {\gamma }_l).\nonumber \\ \end{aligned}$$
(162)

The sign \(\dagger ({\alpha })\) is as before. When \(i = 0\), \(\perp _{k+1}\) should be interpreted as \(\perp _0\).

For all of the above \(\mathfrak {q}^\beta \) operations, set

$$\begin{aligned} \mathfrak {q}_{*} = \sum _\beta Q^{\omega (\beta )}e^{\frac{\mu (\beta )}{2}} \mathfrak {q}^\beta _{*}. \end{aligned}$$
(163)

Here \(\mu : H_2(X,L) \rightarrow \mathbb {Z}\) is the Maslov-class, and e is of degree 2. We thus have operations

$$\begin{aligned} \mathfrak {q}_{k,l}:A^*(L;\Lambda ^e)^{\otimes k} \otimes A^*(X;\Lambda ^e)^{\otimes l} \rightarrow A^*(L;\Lambda ^e). \end{aligned}$$
(164)

Let \(\langle \alpha _1, \alpha _2 \rangle _L = (-1)^{|\alpha _2|}\int _L \alpha _1 \wedge \alpha _2\) be the Poincaré pairing on L. [47] prove results about the operations \({\mathfrak {q}}^{ST}\), we state the analogous results for our operations \({\mathfrak {q}}\). These follows by a direct verification of signs from the results in [47].

Proposition 4.23

(Cyclic symmetry, see [47, Proposition 3.3]) For any \(\alpha = (\alpha _1, \dots , \alpha _{k+1})\) and \(\gamma = (\gamma _1, \dots , \gamma _l)\):

$$\begin{aligned}{} & {} \langle \mathfrak {q}_{k,l}(\alpha _1, \dots , \alpha _k; \gamma _1, \dots , \gamma _l),\alpha _{k+1} \rangle _L = (-1)^{|\alpha _{k+1}|'\epsilon _k(\alpha )} \langle \nonumber \\{} & {} \quad \mathfrak {q}_{k,l}(\alpha _{k+1}, \alpha _1, \dots , \alpha _{k-1}; \gamma _1, \dots , \gamma _l),\alpha _{k} \rangle _L \end{aligned}$$
(165)

Proposition 4.24

(Degree property, see [47, Proposition 3.4]) For any \(\alpha = (\alpha _1, \dots , \alpha _{k})\) and \(\gamma = (\gamma _1, \dots , \gamma _l)\):

$$\begin{aligned}&|\mathfrak {q}^\beta _{k,l}(\alpha _1, \dots , \alpha _k; \gamma _1, \dots , \gamma _l)| = 2 + \epsilon (\alpha ) -\mu (\beta ) + \sum _{j=1}^l (|\gamma _j| - 2) \end{aligned}$$
(166)
$$\begin{aligned}&\quad \equiv \epsilon (\alpha ) + \sum _{j=1}^l |\gamma _j| \; (\text {mod } 2) \end{aligned}$$
(167)

The last equality holds as L is orientable so the Maslov-index of any disk is even. By construction, we then have:

$$\begin{aligned} |{\mathfrak {q}}_{k,l}({\alpha },{\gamma })| = 2 + \epsilon ({\alpha }) + |{\gamma }| -2l. \end{aligned}$$
(168)

Proposition 4.25

(Unit property, see [47, Proposition 3.2]) For \(f \in A^0(L)\), \(\alpha _1, \dots , \alpha _k \in A^*(L;R)\) and \(\gamma \in A^*(X;Q)^{\otimes l}\), we have:

$$\begin{aligned} \mathfrak {q}^\beta _{k,l}(\alpha _1, \dots , \alpha _{i-1},f,\alpha _i, \dots , \alpha _k; \gamma ) = {\left\{ \begin{array}{ll} df &{}(k,l,\beta ) = (1,0,\beta _0)\\ (-1)^{|f|}f\alpha _1 &{}(k,l,\beta ) = (2,0,\beta _0), \; i=1\\ (-1)^{|\alpha _1||f|'}f\alpha _1 &{} (k,l,\beta ) = (2,0,\beta _0), \; i=2\\ 0 &{} \text {otherwise} \end{array}\right. }\nonumber \\ \end{aligned}$$
(169)

Proposition 4.26

(Top degree property, see [47, Proposition 3.12]) We have \((\mathfrak {q}^\beta _{k,l}(\alpha ;\gamma ))_n = 0\) for all lists \(\alpha \), \(\gamma \) for all \((k,l,\beta ) \notin \{ (1,0,\beta _0), (0,1,\beta _0), (2,0,\beta _0) \}\). Here \(\delta _n\) denotes the degree n part of a differential form \(\delta \in A^*(L;R)\).

Proposition 4.27

(Divisor property, see [47, Proposition 3.9]) Assume \(\gamma _1 \in A^2(X,L)\), \(d\gamma _1 = 0\), then

$$\begin{aligned} \mathfrak {q}^\beta _{k,l}(\alpha , \gamma ) = \left( \int _\beta \gamma _1\right) \cdot \mathfrak {q}^\beta _{k,l-1}\left( \alpha ; \bigotimes _{j \ge 2} \gamma _j\right) \end{aligned}$$
(170)

Proposition 4.28

(Energy-zero property, see [47, Proposition 3.8]) For \(k \ge 0\),

$$\begin{aligned} \mathfrak {q}_{k,l}^{\beta _0}(\alpha _1, \dots , \alpha _k; \gamma _1, \dots , \gamma _l) = {\left\{ \begin{array}{ll} d\alpha _1, &{}(k,l) = (1,0),\\ (-1)^{|\alpha _1|}\alpha _1 \wedge \alpha _2, &{}(k,l) = (2,0),\\ \gamma _1|_L, &{}(k,l) = (0,1),\\ 0, &{}\text {otherwise}. \end{array}\right. } \end{aligned}$$
(171)

Proposition 4.29

(Fundamental class property, see [47, Proposition 3.7]) For \(k \ge 0\),

$$\begin{aligned} \mathfrak {q}_{k,l}^{\beta }(\alpha _1, \dots , \alpha _k; 1, \gamma _1, \dots , \gamma _{l-1}) = {\left\{ \begin{array}{ll} 1, &{}(k,l,\beta ) = (0,1,\beta _0),\\ 0, &{} \;\text {otherwise} \end{array}\right. } \end{aligned}$$
(172)

Let \(\gamma = (\gamma _1, \dots ,\gamma _l)\) be a list of differential forms on X. For subsets \(I \sqcup J = \{ 1, \dots , l\}\), define \(sign^\gamma (I,J)\) by the equation

$$\begin{aligned} \bigwedge _{i\in I} \gamma _i \wedge \bigwedge _{j\in J} \gamma _j = (-1)^{sign^\gamma (I,J)}\bigwedge _{s\in [l]} \gamma _s, \end{aligned}$$
(173)

or explicitly

$$\begin{aligned} sign^\gamma (I,J) = \sum _{i \in I, j \in J, j<i} |\gamma _i||\gamma _j|. \end{aligned}$$
(174)

Proposition 4.30

(Structure equation for \({\mathfrak {q}}_{k,l}\), see[47, Proposition 2.4]) For any \(\alpha = (\alpha _1, \dots , \alpha _{k})\) and \(\gamma = (\gamma _1, \dots , \gamma _l)\):

$$\begin{aligned} 0&= \sum _{\begin{array}{c} P \in S_3[k] \\ (2:3) = \{ j \} \end{array}} (-1)^{|\gamma ^{(1:3)}| + 1}\mathfrak {q}_{k,l}(\alpha ;\gamma ^{(1:3)} \otimes d\gamma _j \otimes \gamma ^{(3:3)}) \end{aligned}$$
(175)
$$\begin{aligned}&\quad +\, \sum _{\begin{array}{c} P \in S_3[k] \\ I \sqcup J = [l] \end{array}} (-1)^{i(\alpha , \gamma , P, I)} \mathfrak {q}_{k_1+1+k_3,|I|}(\alpha ^{(1:3)}\otimes \mathfrak {q}_{k_2,|J|}(\alpha ^{(2:3)};\gamma ^{J})\otimes \alpha ^{(3:3)};\gamma ^I), \end{aligned}$$
(176)

where

$$\begin{aligned} i(\alpha ,\gamma , P, I) = (|\gamma _J| + 1)\epsilon _1 + |\gamma _I| + sign^\gamma (I,J). \end{aligned}$$
(177)

The new result we prove concerns the boundary of the moduli spaces \(\mathcal {M}_{k+1,l; \perp _0}(X,\beta )\). The proof is given in chapter 5.

Proposition 4.31

(structure equation for \({\mathfrak {q}}_{k,l,\perp _0}\))

$$\begin{aligned} 0&= \sum _{\begin{array}{c} S_3[l]\\ (2:3) = \{ i \} \end{array}} (-1)^{1+|\gamma ^{(1:3)}|} \mathfrak {q}_{k,l;\perp _0}(\otimes _{j=1}^{k} \alpha _j; \gamma ^{(1:3)} \otimes d\gamma _i \otimes \gamma ^{(3:3)}) \end{aligned}$$
(178)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2 =[l] \\ 1,2 \in J_2 \end{array}} (-1)^{sign^\gamma (J_1,J_2)}\mathfrak {q}_{k, j_1}(\alpha ; \mathfrak {q}_{\emptyset , j_2}(\gamma ^{J_2}) \otimes \gamma ^{J_1})\end{aligned}$$
(179)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2 =[l] \\ 1,2 \in J_2\\ P \in S_3[k] \end{array}} (-1)^{sign^\gamma (J_1,J_2) + \epsilon _1|\gamma ^{J_2}| + 1} \mathfrak {q}_{k_1+k_3+1, j_1} (\alpha ^{(1:3)} \nonumber \\&\qquad \otimes \mathfrak {q}_{k_2,j_2;\perp _0}(\alpha ^{(2:3)}; \gamma ^{J_2})\otimes \alpha ^{(3:3)}; \gamma ^{J_1})\end{aligned}$$
(180)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2=[l]\\ 1,2 \in J_1\\ P \in S_3[k] \end{array}} (-1)^{sign^\gamma (J_1,J_2) + \epsilon _1(|\gamma ^{J_2}|+1) + |\gamma ^{J_1}|} \mathfrak {q}_{k_1+k_3+1, j_1;\perp _0} (\alpha ^{(1:3)}\nonumber \\&\qquad \otimes \mathfrak {q}_{k_2,j_2}(\alpha ^{(2:3)};\gamma ^{J_2})\otimes \alpha ^{(3:3)}; \gamma ^{J_1})\end{aligned}$$
(181)
$$\begin{aligned}&\quad +\sum _{\begin{array}{c} J_1 \cup J_2 \cup J_3 = [l]\\ 1\in J_2,\;2\in J_3\\ P \in S_5[k] \end{array}} (-1)^{A_5}\mathfrak {q}_{k_1+k_3 + k_5+2, j_1} (\alpha ^{(1:5)} \otimes \mathfrak {q}_{k_2,j_2}(\alpha ^{(2:5)};\gamma ^{J_2})\nonumber \\&\qquad \otimes \alpha ^{(3:5)} \otimes \mathfrak {q}_{k_4,j_3}(\alpha ^{(4:5)};\gamma ^{J_3}) \otimes \alpha ^{(5:5)};\gamma ^{J_1}). \end{aligned}$$
(182)

Where

$$\begin{aligned} A_5 = sign^{\gamma }(J_1,J_2,J_3)+ |\gamma ^{J_2}|+ (|\gamma ^{J_2}| + 1)\epsilon _1 + (|\gamma ^{J_3}| + 1)(\epsilon _1 + \epsilon _2 + \epsilon _3) + 1\nonumber \\ \end{aligned}$$
(183)

Additionally we need the following proposition, which is similar to [44, Lemma 3.10] and is proven in Sect. 5.4.

Proposition 4.32

(Unit on the horocycle) For \(\alpha = (\alpha _0, \dots , \alpha _k) \in A^*(L)^{\otimes {k+1}}\) and \(\gamma \in A^*(X)^{\otimes l}\) we have:

$$\begin{aligned}&\langle \mathfrak {q}_{k,l}(\widetilde{\alpha };\gamma ),\alpha _0 \rangle _L = \sum _{P \in S_2[k]} (-1)^{\epsilon _1+|\gamma |} \langle \mathfrak {q}_{k+1,l,\perp _i}(\alpha ^{(1:2)},1,\alpha ^{(2:2)};\gamma ), \alpha _0 \rangle _L \end{aligned}$$
(184)
$$\begin{aligned}&\quad = \sum _{P \in S_2[k]} (-1)^{\epsilon _1+(\epsilon _1 + 1)(\epsilon _2 + |\alpha _0|')+|\gamma |} \langle \mathfrak {q}_{k+1,l,\perp _0}(\alpha ^{(2:2)},\alpha _0,\alpha ^{(1:2)};\gamma ), 1 \rangle _L. \end{aligned}$$
(185)

4.6 Bulk-deformed \(\mathfrak {q}\)-operations

Let \(U \subset H^*(X;{\mathbb {C}})\) be a graded vector subspace, and recall the definition of \(Q_U\) from Sect. 4.1. The same formulae as before then define \({\mathfrak {q}}\)-operations for differential forms with coefficients in \(Q_U^e\). Thus, for example we have:

$$\begin{aligned} \mathfrak {q}_{k,l}:A^*(L;Q_U^e)^{\otimes k} \otimes A^*(X;Q_U^e)^{\otimes l} \rightarrow A^*(L;Q_U^e). \end{aligned}$$
(186)

We then have:

Proposition 4.33

(Linearity, see [47, Proposition 3.1]) The \({\mathfrak {q}}\)-operations are multilinear, in the sense that for \(f \in Q_U^e\), \({\alpha }= ({\alpha }_1, \dots , {\alpha }_k)\) and \({\gamma }= ({\gamma }_1, \dots , {\gamma }_l)\), we have:

$$\begin{aligned} {\mathfrak {q}}_{k,l}({\alpha }_1, \dots , {\alpha }_{i-1}, f {\alpha }_i, \dots , {\alpha }_k; {\gamma }) = (-1)^{|f|(i + \sum _{j=1}^{i-1}|{\alpha }_j| + |{\gamma }|)}f{\mathfrak {q}}_{k,l}({\alpha }, {\gamma }),\nonumber \\ \end{aligned}$$
(187)

and

$$\begin{aligned} {\mathfrak {q}}_{k,l}({\alpha }; {\gamma }_1, \dots , f{\gamma }_i, \dots , {\gamma }_l) = (-1)^{|f|\sum _{j=1}^{i-1}|{\gamma }_j|}f{\mathfrak {q}}_{k,l}({\alpha }, {\gamma }). \end{aligned}$$
(188)

Definition 4.34

A bulk-deformation pair over U is a pair \((b,{\gamma })\). Here \(\gamma \in \mathcal {I}_UA^*(X;Q_U)\) is a bulk-deformation parameter over U and \(b \in \mathcal {I}_U A^*(L,Q_U^e)\) with \(|b| = 1\).

For a bulk-deformation pair \((b,{\gamma })\) define the bulk-deformed operations:

$$\begin{aligned}{} & {} \mathfrak {q}^{b,\gamma }_{k,l}({\alpha }_1, \dots , {\alpha }_k; {\gamma }_1, \dots , {\gamma }_l):= \sum _{\begin{array}{c} s,t \ge 0\\ s_0 + \dots + s_k = s \end{array}} \nonumber \\{} & {} \quad \frac{1}{t!}\mathfrak {q}_{k+s,l+t}(b^{\otimes s_0} \otimes {\alpha }_1 \otimes \dots \otimes {\alpha }_k \otimes b^{\otimes s_k}; {\gamma }_1 \otimes \dots \otimes {\gamma }_l \otimes \gamma ^{\otimes t}). \end{aligned}$$
(189)

Similarly define:

$$\begin{aligned} \mathfrak {q}^{\gamma }_{\emptyset ,l}({\gamma }_1, \dots , {\gamma }_l) = \sum _t \frac{1}{t!}\mathfrak {q}_{\emptyset ,l+t}({\gamma }_1 \otimes \dots \otimes {\gamma }_l \otimes {\gamma }^{\otimes t}). \end{aligned}$$
(190)

Finally we define the bulk-deformed horocyclic \(\mathfrak {q}\)-operations by:

$$\begin{aligned}{} & {} \mathfrak {q}^{b,\gamma }_{k,l,\perp _i}({\alpha }_1, \dots , {\alpha }_k; {\gamma }_1, \dots , {\gamma }_l):= \sum _{\begin{array}{c} s,t \ge 0\\ s_0 + \dots + s_k = s \end{array}} \nonumber \\{} & {} \quad \frac{1}{t!}\mathfrak {q}_{k+s,l+t,\perp _{i + \sum _j^{i-1} s_j}}(b^{\otimes s_0} \otimes {\alpha }_1 \otimes \dots \otimes {\alpha }_k \otimes b^{\otimes s_k}; {\gamma }_1 \otimes \dots \otimes {\gamma }_l \otimes \gamma ^{\otimes t}).\nonumber \\ \end{aligned}$$
(191)

We also define:

Definition 4.35

For a bulk-deformation parameter \({\gamma }_U\), the bulk-deformed quantum cup product is defined by:

$$\begin{aligned} \star _{{\gamma }_U}: H^*(X;Q_U^e) \otimes H^*(X;Q_U^e) \rightarrow H^*(X;Q_U^e), \end{aligned}$$
(192)

by setting \(\eta _1 \star _{{\gamma }_U} \eta _2 = {\mathfrak {q}}^{{\gamma }_U}_{\emptyset ,2}(\eta _1,\eta _2)\).

Remark 4.36

When \(({\gamma },b)\) is a bulk-deformation pair, the degree assumptions on \(\gamma \) and b imply that the properties (4.23)–(4.32), with the exception of the energy-zero property (4.28), all hold for the operations \(\mathfrak {q}^{b,\gamma }\) with the same signs as before.

We will need the following lemma later on; it follows from an easy verification of signs.

Lemma 4.37

For \(v \in Der_{\Lambda ^e} Q^e\), \(\alpha \in A^*(L;Q^e)^{\otimes k}\) and \(\eta \in A^*(X;Q^e)^{\otimes l}\) all of homogeneous degrees, we have:

$$\begin{aligned}&v({\mathfrak {q}}^{b,{\gamma }}_{k,l}(\alpha ;\eta )) = \sum _{\begin{array}{c} P \in S_3[k]\\ (2:3) = i \end{array}} (-1)^{|\eta ^{(1:3)}||v|} {\mathfrak {q}}^{b,{\gamma }}_{k,l}({\alpha };\eta ^{(1:3)}\otimes v(\eta _i) \otimes \eta ^{(3:3)}) \end{aligned}$$
(193)
$$\begin{aligned}&\quad + (-1)^{|\eta ||v|}{\mathfrak {q}}^{b,{\gamma }}_{k,l+1}({\alpha };\eta \otimes v(\gamma ))\end{aligned}$$
(194)
$$\begin{aligned}&\quad +\sum _{P \in S_2[k]} (-1)^{(|\eta |+\epsilon _1 + 1)|v| + 1} {\mathfrak {q}}^{b,{\gamma }}_{k+1,l}({\alpha }^{(1:2)}\otimes v(b) \otimes \alpha ^{(2:2)};\eta )\end{aligned}$$
(195)
$$\begin{aligned}&\quad + (-1)^{(|\eta |+ 1)|v|} {\mathfrak {q}}^{b,{\gamma }}_{k,l}(v({\alpha });\eta ). \end{aligned}$$
(196)

Here

$$\begin{aligned} v({\alpha }) = \sum _{\begin{array}{c} P \in S_3[k] \\ (2:3) = \{ j \} \end{array}} (-1)^{\epsilon _1|v|} {\alpha }^{(1:3)} \otimes v({\alpha }_j) \otimes {\alpha }^{(3:3)}. \end{aligned}$$
(197)

A similar lemma holds for \(v = e \partial _e\), here one gets an additional term, as the \({\mathfrak {q}}\) operations depend on e. Also note that \(\partial _e({\gamma }) = 0\) by definition of a bulk-parameter. First, define operations weighted by the Maslov index \(\mu \): \(\widetilde{\mathfrak {q}}_{k,l}^{b,{\gamma }} = \sum _{\beta } \mu (\beta )\mathfrak {q}_{k,l}^{b,{\gamma },\beta }\). We then have:

Lemma 4.38

For \(\alpha \in A^*(L;Q^e)^{\otimes k}\) and \(\eta \in A^*(X;Q^e)^{\otimes l}\) all of homogeneous degrees, we have:

$$\begin{aligned}&e\partial _e({\mathfrak {q}}^{b,{\gamma }}_{k,l}(\alpha ;\eta )) = \sum _{\begin{array}{c} P \in S_3[k]\\ (2:3) = i \end{array}} {\mathfrak {q}}^{b,{\gamma }}_{k,l}({\alpha };\eta ^{(1:3)}\otimes e\partial _e(\eta _i) \otimes \eta ^{(3:3)})\end{aligned}$$
(198)
$$\begin{aligned}&\quad -\sum _{P \in S_2[k]} {\mathfrak {q}}^{b,{\gamma }}_{k+1,l}({\alpha }^{(1:2)}\otimes e\partial _e(b) \otimes \alpha ^{(2:2)};\eta )\end{aligned}$$
(199)
$$\begin{aligned}&\quad + {\mathfrak {q}}^{b,{\gamma }}_{k,l}(e\partial _e({\alpha });\eta )\end{aligned}$$
(200)
$$\begin{aligned}&\quad + \frac{1}{2}\widetilde{\mathfrak {q}}_{k,l}^{b,{\gamma }}({\alpha };\eta ). \end{aligned}$$
(201)

4.7 Fukaya \(A_\infty \)-algebra

Let \(U \subset H^*(X;{\mathbb {C}})\) and \(({\gamma },b)\) be a bulk-deformation pair over U. Assume \({\gamma }\) satisfies Assumption 4.3, fixing the derivation \(Y\in Der_{\Lambda ^e} Q^e_U\) such that \([Y({\gamma })] = c_1\). For ease of notation, write \(Q = Q_U\).

Solomon and Tukachinsky [47, Theorem 1] construct an \(A_\infty \)-algebra \(A^{ST}\) using the operations \({\mathfrak {q}}^{ST}_{k,0}\). We have different sign conventions for our operations \({\mathfrak {q}}\), but the following still holds.

Definition 4.39

Let \((A:= CF^*(L,L)[e]:= A^*(L;Q^e),\mathfrak {m}_k:= {\mathfrak {q}}^{b,{\gamma }}_{k,0},\langle \;, \; \rangle _L, 1)\). It follows directly from the properties of the \({\mathfrak {q}}\) operations that this forms an n-dimensional, strictly unital and cyclic \(A_\infty \)-algebra. It follows from the degree property 4.24, and the definition that \(|e| = 2\) that this \(A_\infty \)-algebra is Euler-graded with Euler vector field \(E = e\partial _e + E_U\). The grading operator is defined by \(Gr(f{\alpha }) = (|f| + |{\alpha }|)f{\alpha }\) for \(f \in Q^e\) and \({\alpha }\in A^*(L)\). Furthermore, this \(A_\infty \)-algebra is (possibly) curved. The valuation \(\zeta _A\) is induced by the valuation \(\zeta _Q\), defined in (111).

Definition 4.40

Let \(CF^*(L,L):= CF^*(L,L)[e] \otimes _{Q^e} Q\) be the \(A_\infty \)-algebra obtained by setting \(e=1\).

Remark 4.41

Recall the definition of the negative-opposite of an \(A_\infty \)-algebra (Definition 3.3). Then \(CF^*(L,L)\) is related to \(A^{ST}\) as \(CF^*(L,L)^{-op} = A^{ST}\).

Whenever we have to pick a basis for A in order to compute derivatives, we will always pick a constant basis, i.e. one in \(A^*(L;\mathbb {C})\).

Recall the connection

$$\begin{aligned} \nabla ^{GGM}: Der_\Lambda ^e Q^e \otimes HC_*^-(A) \rightarrow u^{-1}HC_*^-(A). \end{aligned}$$
(202)

We will now define a connection \(\widetilde{\nabla }\) which agrees with \(\nabla ^{GGM}\) up to homotopy. First, for \(v \in Der_\mathbb {C} Q^e\) define the length zero Hochschild cochain \(\phi _v:= v(b) \in A\). Also let \(\mathfrak {m}_{v}(\alpha ):= \mathfrak {q}^{b,{\gamma }}_{k,1}(\alpha ,v(\gamma ))\). Lemma 4.37 then shows:

Corollary 4.42

For \(v \in Der_{\Lambda ^e} Q^e\) we have:

$$\begin{aligned} v(\mathfrak {m}) = \mathfrak {m}_v + [\phi _v,\mathfrak {m}]. \end{aligned}$$

Definition 4.43

For \(v \in Der_{\Lambda ^e Q^e}\), define \(J_v = i\{ \phi _v \}\).

The Cartan homotopy formula 3.15 then shows:

Corollary 4.44

The connection defined by:

$$\begin{aligned} \widetilde{\nabla }_v(\alpha ):= v(\alpha ) - \mathcal {L}_{\phi _v}(\alpha ) + (-1)^{|v|+1}u^{-1}i\{ {\mathfrak {m}}^{b,{\gamma }}_v \}({\alpha }) \end{aligned}$$
(203)

satisfies:

$$\begin{aligned} \nabla _v = \widetilde{\nabla }_v + u^{-1}[J_v,b+uB]. \end{aligned}$$
(204)

We finish this section by showing that the property 4.12.2 holds. To this end, set \(\widetilde{\mathfrak {m}}_k = \widetilde{{\mathfrak {q}}}^{b,{\gamma }}_{k,0}\). Lemma 4.38 then shows:

Lemma 4.45

\(2e\partial _e(\mathfrak {m}^{b,{\gamma }}) = \widetilde{\mathfrak {m}}^{b,{\gamma }} - [\mathfrak {m}^{b,{\gamma }},\phi _{2e\partial _e}]\).

Applying Lemma 4.42 to \(v= Y\), and using the divisor equation (4.27) to rewrite \(2{\mathfrak {m}}^{b,{\gamma }}_{Y} = \widetilde{{\mathfrak {m}}}^{b,{\gamma }}\), we find:

Lemma 4.46

The Fukaya \(A_\infty \)-algebra thus defined satisfies property 4.12.2.

$$\begin{aligned} Y({\mathfrak {m}}^{b,{\gamma }}) = e\partial _e ({\mathfrak {m}}^{b,{\gamma }}) + [{\mathfrak {m}}^{b,{\gamma }}, \phi _{e\partial _e} - \phi _{Y}]. \end{aligned}$$
(205)

4.8 Closed–open and open–closed maps

In [7] an open–closed map is defined directly. In our setup one would want to define it using operations \(\mathfrak {p}_k: CC_*(A) \rightarrow A^*(X)\) given by:

$$\begin{aligned} \mathfrak {p}_k(\alpha ) = (evi_1)_*\left( \bigwedge _{j=0}^k evb_{j}^*\alpha _j\right) , \end{aligned}$$
(206)

where now the push forward is along the interior evaluation \(evi_1: \mathcal {M}_{k+1,1} \rightarrow X\). However, in the present setup the push-forward along interior evaluation is not well defined, as \(evi_1:\mathcal {M}_{k+1,l}(\beta ) \rightarrow X\) need not be a submersion.

The approach we thus take is to first construct a closed–open map, and then dualise to obtain an open–closed map. The fact that the closed–open map is an algebra homomorphism (a result first shown by [8, 13]), will then translate into the result that the open–closed map is a morphism of \(QH^*(X)\) modules (see also [13]). We will then construct a cyclic open–closed map, which was also done in [10]. We then upgrade the result that the open–closed map is a morphism of \(QH^*(X)\)-module to show that the cyclic open–closed map is a morphism of T-structures, using an argument due to [11]. We then show Assumptions 4.12 hold for the cyclic open–closed map in order to conclude that it respects the connection in the u-direction.

4.8.1 The closed–open map

We define the closed–open map \({\mathcal{C}\mathcal{O}}_e: A^*(X;Q^e) \rightarrow CC^*(A)\) on the chain level. Set

$$\begin{aligned} {\mathcal{C}\mathcal{O}}_e(\eta )(\alpha _1, \dots , \alpha _k) = {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\alpha _1, \dots , \alpha _k; \eta ). \end{aligned}$$
(207)

It follows from the unit property, Proposition 4.25, that the Hochschild cochain \({\mathcal{C}\mathcal{O}}_e(\eta )\) is reduced. Furthermore, it follows from the degree property, Proposition 4.24, that \(|{\mathcal{C}\mathcal{O}}_e(\eta )| = |\eta |\; (mod\;2)\), so that the closed–open map is a \(\mathbb {Z}/2\)-graded map.

Lemma 4.47

The closed–open map is a chain map. That is: \([\mathfrak {m},{\mathcal{C}\mathcal{O}}_e(\eta )] = {\mathcal{C}\mathcal{O}}_e(d\eta )\).

Proof

By definition:

$$\begin{aligned}&[\mathfrak {m},{\mathcal{C}\mathcal{O}}_e(\eta )](\alpha _1, \dots , \alpha _k) = \sum _{P \in S_3[k]} (-1)^{\epsilon _1(1+|\eta |)}{\mathfrak {q}}^{b,{\gamma }}_{k_1+1+k_3,0}(\alpha ^{(1:3)}\otimes {\mathfrak {q}}^{b,{\gamma }}_{k_2,1}(\alpha ^{(2:3)};\eta )\otimes \alpha ^{(3:3)})\nonumber \\&\qquad + \sum _{P \in S_3[k]} (-1)^{\epsilon _1 +|\eta |}{\mathfrak {q}}^{b,{\gamma }}_{k_1+1+k_3,1}(\alpha ^{(1:3)}\otimes {\mathfrak {q}}^{b,{\gamma }}_{k_2,0}(\alpha ^{(2:3)})\otimes \alpha ^{(3:3)};\eta ), \end{aligned}$$
(208)

which, by Proposition 4.30 equals \({\mathcal{C}\mathcal{O}}_e(d\eta )\). \(\square \)

Let \({\mathcal{C}\mathcal{O}}_e: H^*(X;Q^e) \rightarrow HH^*(A)\) be the induced map on cohomology. Next up we will prove the following, which is originally due to [8, 13] in different setups:

Proposition 4.48

The closed–open map induces a unital algebra homomorphism on cohomology.

To this end, we first define:

Definition 4.49

Let the homotopy operator \(H:A^*(X;Q^e)^{\otimes 2} \rightarrow CC^*(A)\) be given by

$$\begin{aligned} H(\gamma _1, \gamma _2)(\alpha _1, \dots , \alpha _k) = {\mathfrak {q}}^{b,{\gamma }}_{k,2,\perp _0}(\alpha ;\gamma _1\otimes \gamma _2). \end{aligned}$$
(209)

Unitality follows from the fundamental class property 4.29. The following lemma immediately implies that the closed–open map respects the product.

Lemma 4.50

The homotopy operator H satisfies:

$$\begin{aligned} {\mathcal{C}\mathcal{O}}_e(\gamma _1 \star \gamma _2) = {\mathcal{C}\mathcal{O}}_e(\gamma _1) \cup {\mathcal{C}\mathcal{O}}_e(\gamma _2) + H(d(\gamma _1\otimes \gamma _2)) + [\mathfrak {m},H(\gamma _1,\gamma _2)]. \end{aligned}$$
(210)

Proof

We write down the terms one by one. Firstly:

$$\begin{aligned} {\mathcal{C}\mathcal{O}}_e(\gamma _1 \star \gamma _2)(\alpha ) = {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\alpha ;{\mathfrak {q}}^{b,{\gamma }}_{\emptyset ,2}(\gamma _1,\gamma _2)). \end{aligned}$$
(211)

Furthermore:

$$\begin{aligned}{} & {} {\mathcal{C}\mathcal{O}}_e(\gamma _1) \cup {\mathcal{C}\mathcal{O}}_e(\gamma _2)(\alpha ) = (-1)^{|\gamma _1|}M^2({\mathcal{C}\mathcal{O}}_e(\gamma _1),{\mathcal{C}\mathcal{O}}_e(\gamma _2)) \nonumber \\{} & {} \quad = \sum _{P\in S_5[k]} (-1)^{\star }{\mathfrak {q}}^{b,{\gamma }}_{k_1+k_3+k_5+2}(\alpha ^{(1:5)}\otimes {\mathfrak {q}}^{b,{\gamma }}_{k_2,1}(\alpha ^{(2:5)};\gamma _1)\otimes \alpha ^{(3:5)} \otimes {\mathfrak {q}}^{b,{\gamma }}_{k_4,1}(\alpha ^{(4:5)};\gamma _2) \otimes \alpha ^{(5:5)}),\nonumber \\ \end{aligned}$$
(212)

where \(\star = |\gamma _1| + (|\gamma _1|+1)\epsilon _1 + (|\gamma _2|+1)(\epsilon _1+\epsilon _2+\epsilon _3)\).

We then compute the homotopy terms:

$$\begin{aligned} H(d(\gamma _1\otimes \gamma _2)) = {\mathfrak {q}}^{b,{\gamma }}_{k,2,\perp _0}(\alpha ;d\gamma _1\otimes \gamma _2) + (-1)^{|\gamma _1|}{\mathfrak {q}}^{b,{\gamma }}_{k,2,\perp _0}(\alpha ;\gamma _1\otimes d\gamma _2). \end{aligned}$$
(213)

Finally we find:

$$\begin{aligned}&[\mathfrak {m},H(\gamma _1,\gamma _2)] = \mathfrak {m} \circ H(\gamma _1,\gamma _2) + (-1)^{|\gamma _1| + |\gamma _2| + 1}H(\gamma _1,\gamma _2) \circ \mathfrak {m}\end{aligned}$$
(214)
$$\begin{aligned}&\quad = \sum _{ P \in S_3[k]} (-1)^{\epsilon _1(|\gamma _1| + |\gamma _2|)} {\mathfrak {q}}^{b,{\gamma }}_{k_1+k_3+1,0} (\alpha ^{(1:3)} \otimes {\mathfrak {q}}^{b,{\gamma }}_{k_2,2;\perp _0}(\alpha ^{(2:3)}; \gamma _1 \otimes \gamma _2)\otimes \alpha ^{(3:3)}) \end{aligned}$$
(215)
$$\begin{aligned}&\qquad + \sum _{P \in S_3[k]} (-1)^{\epsilon _1 + |\gamma _1| + |\gamma _2| + 1} {\mathfrak {q}}^{b,{\gamma }}_{k_1+k_3+1,2;\perp _0} (\alpha ^{(1:3)} \otimes {\mathfrak {q}}^{b,{\gamma }}_{k_2,0}(\alpha ^{(2:3)})\otimes \alpha ^{(3:3)}; \gamma _1 \otimes \gamma _2). \end{aligned}$$
(216)

Lemma (4.50) then follows by applying the structure equation (4.31) with interior inputs \(\gamma _1 \otimes \gamma _2\). \(\square \)

We conclude this section with the following observations:

Lemma 4.51

\({\mathcal{C}\mathcal{O}}_e(c_1) = \widetilde{\mathfrak {m}}/2\).

Proof

Pick a representative \(\eta \in A^2(X,L)\) for the Maslov class \(\mu \). Then \(\eta \) also represents \(2c_1 \in H^*(X)\). The divisor property 4.27 shows that \({\mathcal{C}\mathcal{O}}_e({\gamma }_1) = \widetilde{\mathfrak {m}}\). \(\square \)

Furthermore, by definition of \({\mathfrak {m}}_v\) we have:

Lemma 4.52

For \(v \in Der_\mathbb {C} Q^e\) we have: \({\mathcal{C}\mathcal{O}}_e(v(\gamma )) = \mathfrak {m}_{v}\). It follows that \({\mathcal{C}\mathcal{O}}_e(E_U(\gamma )) = \mathfrak {m}_{E_U}\).

4.8.2 The open–closed map

The open–closed map will take the form \({\mathcal{O}\mathcal{C}}: HH_*(A) \rightarrow QH^*(X;Q^e)\). To this end, we first define the open–closed pairing \(A^*(X;Q^e) \otimes CC_*(A) \rightarrow Q^e\). We then show this descends to a pairing on (co)homology. Finally, by dualising the first factor, and using Poincaré duality, we obtain the open–closed map.

Definition 4.53

The open–closed pairing \(\langle \_,{\mathcal{O}\mathcal{C}}_e(\_) \rangle : A^*(X;Q^e) \otimes CC_*(A) \rightarrow Q^e\) is defined as

$$\begin{aligned} \langle \eta , {\mathcal{O}\mathcal{C}}_e(\alpha ) \rangle := ({\mathcal{C}\mathcal{O}}_e(\eta ),\alpha ) = (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1)}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };\eta ),\alpha _0 \rangle _L. \end{aligned}$$
(217)

Here the pairing \((\cdot ,\cdot )\) is as in Eq. (63).

We will show (Lemma 4.55) that the open–closed pairing descends to homology, so that the following makes sense.

Definition 4.54

The open–closed map \({\mathcal{O}\mathcal{C}}: HH_*(A) \rightarrow QH^*(X;Q^e)\) is defined by requiring that

$$\begin{aligned} \langle \eta ,{\mathcal{O}\mathcal{C}}({\alpha }) \rangle _X = \langle \eta , {\mathcal{O}\mathcal{C}}_e(\alpha ) \rangle . \end{aligned}$$
(218)

On the left, the pairing is the Poincaré pairing on X, and on the right the pairing is the open–closed pairing.

Since the closed–open map is a chain map (Lemma 4.47), and the pairing respects differentials (Lemma 3.10), the following is immediate.

Lemma 4.55

We have:

$$\begin{aligned} \langle d\eta , {\mathcal{O}\mathcal{C}}_e(\alpha ) \rangle + (-1)^{|\eta |} \langle \eta , {\mathcal{O}\mathcal{C}}_e(b(\alpha )) \rangle = 0. \end{aligned}$$
(219)

The open–closed pairing thus descends to (co)homology.

Ganatra [13] shows the closed–open map makes \(HH_*(A)\) into a \(QH^*(X)\)-module. We prove this in our setup.

Proposition 4.56

The open–closed map is a map of \(QH^*(X;Q^e)\)-modules.

To this end, first define:

Definition 4.57

Recall the map H from Definition 4.49. Then let

$$\begin{aligned} \langle \_,G(\_) \rangle :A^*(X)^{\otimes 2}&\otimes CC_*(A) \rightarrow R^e \end{aligned}$$
(220)
$$\begin{aligned} \gamma _1 \otimes \gamma _2&\otimes \alpha \mapsto \langle \gamma _1 \otimes \gamma _2, G(\alpha ) \rangle = (H(\gamma _1, \gamma _2),\alpha ). \end{aligned}$$
(221)

The following lemma follows directly from Lemma 4.50.

Lemma 4.58

The pairing G satisfies:

$$\begin{aligned} \langle \gamma _1 \star \gamma _2, {\mathcal{O}\mathcal{C}}_e(\alpha ) \rangle= & {} \langle \gamma _1, {\mathcal{O}\mathcal{C}}_e({\mathcal{C}\mathcal{O}}_e(\gamma _2) \cap \alpha )\rangle + \langle d(\gamma _1 \otimes \gamma _2), G(\alpha ) \rangle + (-1)^{|{\gamma }_1| |{\gamma }_2|}\nonumber \\{} & {} + \langle \gamma _1 \otimes \gamma _2, G(b(\alpha )) \rangle . \end{aligned}$$
(222)

Proof of Proposition 4.56

Lemma 4.58 shows that on homology we have::

$$\begin{aligned} \langle \gamma _1, {\mathcal{O}\mathcal{C}}_e({\mathcal{C}\mathcal{O}}_e(\gamma _2) \cap \alpha )\rangle _X&= \langle {\gamma }_1 \star {\gamma }_2,{\mathcal{O}\mathcal{C}}({\alpha }) \rangle _X \end{aligned}$$
(223)
$$\begin{aligned}&= \langle {\gamma }_1, {\gamma }_2 \star {\mathcal{O}\mathcal{C}}({\alpha }) \rangle _X. \end{aligned}$$
(224)

Thus \({\gamma }\star {\mathcal{O}\mathcal{C}}({\alpha }) = {\mathcal{O}\mathcal{C}}_e({\mathcal{C}\mathcal{O}}_e(\gamma ) \cap \alpha )\).

4.8.3 The cyclic open–closed map

In order to define the cyclic open–closed map, we first show the open–closed pairing descends to cyclic homology.

Definition 4.59

Extend the open–closed pairing (Definition 4.53) u-linearly to a map:

$$\begin{aligned} \langle \_,{\mathcal{O}\mathcal{C}}_e^-(\_) \rangle : (A^{*}(X;Q^e)[[u]]) \otimes CC^{-}(A) \rightarrow Q^e[[u]]. \end{aligned}$$
(225)

We then have:

Lemma 4.60

The open–closed pairing descends to a pairing \(QH_{-}^{*}(X;R^e) \otimes HC^{-}(A) \rightarrow Q^e[[u]]\).

Proof

We have already shown the open–closed pairing respects the first differential b. We are then done if \(\langle \eta ,\mathcal{O}\mathcal{C}(B(\alpha )) \rangle = 0\). Now

$$\begin{aligned} \langle \eta ,{\mathcal{O}\mathcal{C}}_e(B(\alpha )) \rangle&= \sum _{P \in S_2[k]} (-1)^{\epsilon _2(|\alpha _0|' + \epsilon _1)} ( {\mathcal{C}\mathcal{O}}_e(\eta ), 1[\alpha ^{(2:2)}|\alpha _0|\alpha ^{(1:2)}] ) \end{aligned}$$
(226)
$$\begin{aligned}&= \sum _{P \in S_2[k]} (-1)^{\epsilon _2(|\alpha _0|' + \epsilon _1)+ \epsilon (\alpha ) + |\eta |} \langle {\mathfrak {q}}^{b,{\gamma }}_{k+1,1}(\alpha ^{(2:2)}\otimes \alpha _0 \otimes \alpha ^{(1:2)};\eta ),1 \rangle _L\end{aligned}$$
(227)
$$\begin{aligned}&= 0. \end{aligned}$$
(228)

The first equality is by definition of B and \({\mathcal{O}\mathcal{C}}\). The second equality holds by definition of \({\mathcal{C}\mathcal{O}}_e\). The last equality follows by the top degree property 4.26. \(\square \)

Extend the Poincaré pairing on X u-linearly to a pairing \(QH^*(X;Q^e)[[u]] \otimes QH^*(X;Q^e)[[u]] \rightarrow Q^e[[u]]\).

Definition 4.61

The cyclic open–closed map \({\mathcal{O}\mathcal{C}}_e^-: HC^-_*(A) \rightarrow QH^*(X;Q^e)[[u]]\) is defined by requiring that

$$\begin{aligned} \langle \eta ,{\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle _X = \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha ) \rangle . \end{aligned}$$
(229)

On the left, the pairing is the Poincaré pairing on X, and on the right the pairing is the open–closed pairing.

We now prove:

Theorem 4.62

The cyclic open–closed map \({\mathcal{O}\mathcal{C}}_e^-: HC^{-}(A) \rightarrow (QH^*(X;Q_U)[[u]],\nabla )\) is a morphism of T-structures over \(Q^e \supset \Lambda ^e\).

Our proof follows the same ideas as outlined by Ganatra-Perutz-Sheridan in talks. First we observe that the same reasoning as for Lemma 4.60 shows that:

Lemma 4.63

\(\langle \eta , \mathcal{O}\mathcal{C}(B^{1,1}(\phi , \alpha )) \rangle = 0\) for any \(\eta \in A^*(X;Q^e)\), \(\alpha \in CC_*(A)\) and \(\phi \in CC^*(A)\).

Next we define:

Definition 4.64

For \(v \in Der_{\Lambda ^e} Q^e\), set

$$\begin{aligned} \langle \eta ,\widetilde{G}_{v}({\alpha })\rangle := (-1)^{|v||\eta |}\langle \eta \otimes v({\gamma }), G({\alpha }) \rangle , \end{aligned}$$
(230)

where we have extended G from Definition 4.57u-linearly.

We then have:

Lemma 4.65

The pairing \(\widetilde{G}_v\) satisfies:

$$\begin{aligned}{} & {} \langle \nabla ^*_{v} \eta , {\mathcal{O}\mathcal{C}}_e^- (\alpha ) \rangle + (-1)^{|\eta ||v|}\langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\widetilde{\nabla }_{v} \alpha ) \rangle = v \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right) \\{} & {} \quad + u^{-1}\left( \langle d\eta , \widetilde{G}_{v} (\alpha ) \rangle + (-1)^{|\eta | + |v|}\langle \eta , \widetilde{G}_{v}\left( (b+uB)(\alpha )\right) \rangle \right) . \end{aligned}$$

In order to prove this, we first show the following:

Lemma 4.66

We have:

$$\begin{aligned}{} & {} v \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right) + (-1)^{|\eta |+|v|}\langle \eta , \widetilde{G}_{v}(B(\alpha )) \rangle = \langle v(\eta ), {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \nonumber \\{} & {} \quad + (-1)^{|\eta ||v|}\langle \eta , {\mathcal{O}\mathcal{C}}_e^- \left( v(\alpha ) - \mathcal {L}_{\phi _v}({\alpha })\right) \rangle . \end{aligned}$$
(231)

Proof

Using Lemma 4.32, we find that:

$$\begin{aligned} \langle \eta , \widetilde{G}_{v}(B(\alpha )) \rangle = (-1)^{|\eta ||v| + |\alpha _0|(\epsilon (\widetilde{\alpha })+1) + |v| + |\eta | + 1}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,l+1}(\widetilde{{\alpha }};\eta \otimes v(\gamma )),\alpha _0 \rangle _L.\nonumber \\ \end{aligned}$$
(232)

We also write out the other terms.

$$\begin{aligned}&\langle v(\eta ), {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle = (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1)}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };v(\eta )),\alpha _0 \rangle _L, \end{aligned}$$
(233)
$$\begin{aligned}&\langle \eta , {\mathcal{O}\mathcal{C}}_e^- (v(\alpha )) \rangle \nonumber \\&= (-1)^{(|v|+|\alpha _0|)(\epsilon (\widetilde{\alpha }) + 1)}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };\eta ),v(\alpha _0) \rangle _L +(-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1)}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(v(\widetilde{\alpha });\eta ),\alpha _0 \rangle _L,\end{aligned}$$
(234)
$$\begin{aligned}&\langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\mathcal {L}_{\phi _v}({\alpha })) \rangle = \sum _{P \in S_2[k]} (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1) + |v|(\epsilon _1 + 1)}\nonumber \\&\quad \langle {\mathfrak {q}}^{b,{\gamma }}_{k+1,1}(\alpha ^{(1:2)} \otimes v(b) \otimes \alpha ^{(2:2)};\eta ),\alpha _0 \rangle _L. \end{aligned}$$
(235)

By definition of the open–closed pairing, we also have:

$$\begin{aligned} v \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right)= & {} (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1)}\langle v({\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };\eta )),\alpha _0 \rangle _L + (-1)^{|\alpha _0|(\epsilon (\widetilde{\alpha }) + 1) + |v|(\epsilon (\widetilde{{\alpha }}) +|\eta |+ 1)}\nonumber \\{} & {} \langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };\eta ),v(\alpha _0) \rangle _L. \end{aligned}$$
(236)

Then, apply Lemma 4.37 to compute \(v({\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{\alpha };\eta ))\). Keeping track of all the signs shows the result. \(\square \)

Proof of Lemma 4.65

As all the terms in the above equation are u-linear, we may assume \({\alpha }\) and \(\eta \) are independent of u, and then prove this order by order in u.

To verify the \(u^{-1}\) term, apply Lemma 4.58 with \(\gamma _1 = \eta \) and \(\gamma _2 = v(\gamma )\) and use Lemma 4.52 to compute \({\mathcal{C}\mathcal{O}}_e(v(\gamma )) = \mathfrak {m}_{v}\). Equality of the \(u^0\) terms is shown by Lemma 4.66.

Definition 4.67

Recall the definition of \(J_v\) from Definition 4.43. Then define the pairing:

$$\begin{aligned} \langle \eta ,G_{v}({\alpha })\rangle := \langle \eta ,\widetilde{G}_{v}({\alpha })\rangle + (-1)^{|\eta |'|v|' + 1} \langle \eta , {\mathcal{O}\mathcal{C}}^-_e(J_v({\alpha })) \rangle . \end{aligned}$$
(237)

Proposition 4.68

The pairing \(G_v\) satisfies:

$$\begin{aligned} \langle \nabla ^*_{v} \eta , {\mathcal{O}\mathcal{C}}_e^- (\alpha ) \rangle + (-1)^{|\eta ||v|}\langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\nabla _{v} \alpha ) \rangle = v \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right) \\ + u^{-1}\left( \langle d\eta , G_{v} (\alpha ) \rangle + (-1)^{|\eta | + |v|}\langle \eta , G_{v}\left( (b+uB)(\alpha )\right) \rangle \right) . \end{aligned}$$

Proof

This follows directly from Lemma 4.65, Corollary 4.44, and a verification of signs. \(\square \)

Proof of Theorem 4.62

By the properties of the Poincaré pairing 130:

$$\begin{aligned} \langle \eta ,\nabla _v({\mathcal{O}\mathcal{C}}_e^-({\alpha })) \rangle _X = (-1)^{|\eta ||v|} v(\langle \eta , {\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle ) - (-1)^{|\eta ||v|}\langle \nabla ^*_v \eta , {\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle .\nonumber \\ \end{aligned}$$
(238)

Lemma 4.68 then shows that on homology:

$$\begin{aligned} \langle \eta ,\nabla _v({\mathcal{O}\mathcal{C}}_e^-({\alpha })) \rangle _X = \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\nabla _v({\alpha })) \rangle , \end{aligned}$$
(239)

which shows that \({\mathcal{O}\mathcal{C}}_e^-(\nabla _v({\alpha })) = \nabla _v({\mathcal{O}\mathcal{C}}_e^-({\alpha }))\).

We next show that the Assumptions 4.12 hold in our setup, so that the open–closed map respects u-connections. Assumption 4.12(1) is Lemma 4.68. Assumption 4.12(2) is Lemma 4.46 with \(\phi = \phi _{e\partial _e} - \phi _{Y}\).

The following lemma shows Assumption 4.12(3) holds in our setup.

Lemma 4.69

For any \(\eta \) and \({\alpha }\) we have:

$$\begin{aligned} \langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha ) \rangle = \langle \eta , e\partial _e({\mathcal{O}\mathcal{C}}_e^-)({\alpha }) \rangle + \langle {\alpha }, {\mathcal{O}\mathcal{C}}_e^-(\mathcal {L}_\phi (\eta )) \rangle , \end{aligned}$$
(240)

Where \(\phi \) is as above.

Proof

First note that a computation similar to Lemma 4.66 shows that

$$\begin{aligned} e\partial _e \left( \langle \eta , {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle \right)= & {} \langle e\partial _e(\eta ), {\mathcal{O}\mathcal{C}}_e^-(\alpha )\rangle + \langle \eta , {\mathcal{O}\mathcal{C}}_e^- \left( v(\alpha ) -\mathcal {L}_{\phi _{e\partial _e}}({\alpha })\right) \rangle \nonumber \\{} & {} +\, \sum _\beta (-1)^{|\alpha _0|(\epsilon (\widetilde{{\alpha }}) +1)} \frac{\mu (\beta )}{2}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{{\alpha }};\eta ),{\alpha }_0 \rangle _L. \end{aligned}$$
(241)

We thus find that:

$$\begin{aligned} \langle \eta , e\partial _e({\mathcal{O}\mathcal{C}}_e^-)(\alpha )\rangle + \langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\mathcal {L}_{\phi _{e\partial _e}}({\alpha })) \rangle = \sum _\beta (-1)^{|\alpha _0|(\epsilon (\widetilde{{\alpha }}) +1)} \frac{\mu (\beta )}{2}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{{\alpha }};\eta ),{\alpha }_0 \rangle _L.\nonumber \\ \end{aligned}$$
(242)

Similarly, we have:

$$\begin{aligned} \langle \eta , Y({\mathcal{O}\mathcal{C}}_e^-)(\alpha )\rangle + \langle \eta , {\mathcal{O}\mathcal{C}}_e^- (\mathcal {L}_{\phi _{Y}}({\alpha })) \rangle&= (-1)^{|\alpha _0|(\epsilon (\widetilde{{\alpha }}) +1)} \langle {\mathfrak {q}}^{b,{\gamma }}_{k,2}(\widetilde{{\alpha }};\eta \otimes c_1),{\alpha }_0 \rangle _L\end{aligned}$$
(243)
$$\begin{aligned}&= \sum _\beta (-1)^{|\alpha _0|(\epsilon (\widetilde{{\alpha }}) +1)} \frac{\mu (\beta )}{2}\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{{\alpha }};\eta ),{\alpha }_0 \rangle _L. \end{aligned}$$
(244)

The last equality follows from the divisor property 4.27. The result follows. \(\square \)

Finally we show that Assumption 4.12(4) holds:

Lemma 4.70

We have: \(Gr^- \circ {\mathcal{O}\mathcal{C}}_e^- = {\mathcal{O}\mathcal{C}}_e^- \circ Gr^-\).

Proof

First observe that for \({\alpha }= {\alpha }_0[{\alpha }_1, \dots , {\alpha }_k]\), with \({\alpha }_i \in A^*(L;{\mathbb {C}})\), we have:

$$\begin{aligned} Gr^-({\alpha }) = (1 + \epsilon ({\alpha })){\alpha }\end{aligned}$$
(245)

We then compute

$$\begin{aligned} |\eta | + |{\mathcal{O}\mathcal{C}}_e^-({\alpha })| - 2n&= |\langle \eta ,{\mathcal{O}\mathcal{C}}_e^-({\alpha }) \rangle _X| \end{aligned}$$
(246)
$$\begin{aligned}&= |\langle {\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{{\alpha }};\eta ),{\alpha }_0 \rangle _L|\end{aligned}$$
(247)
$$\begin{aligned}&= |{\mathfrak {q}}^{b,{\gamma }}_{k,1}(\widetilde{{\alpha }};\eta )| + |{\alpha }_0| - n\end{aligned}$$
(248)
$$\begin{aligned}&= \epsilon (\widetilde{{\alpha }}) + |\eta | + |{\alpha }_0| - n. \end{aligned}$$
(249)

We thus have \(|{\mathcal{O}\mathcal{C}}_e^-({\alpha })| = 1 +\epsilon ({\alpha }) + n\). As the grading \(Gr^-\) on \(QH^*(X;Q^e_U)[[u]]\) is shifted down by n compared to the cohomological grading, we have

$$\begin{aligned} Gr^-({\mathcal{O}\mathcal{C}}_e^-({\alpha })) = (1+\epsilon ({\alpha })){\mathcal{O}\mathcal{C}}_e^-({\alpha }) = {\mathcal{O}\mathcal{C}}_e^-(Gr^-({\alpha })). \end{aligned}$$
(250)

\(\square \)

Theorem 4.15 thus implies that the cyclic open–closed map \({\mathcal{O}\mathcal{C}}_e^-\) is a morphism of TE-structures. Now recall that \(CF^*(L,L) = A \otimes _{Q^e} Q\). The cyclic open–closed map then restricts at \(e=1\) to a map

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-: HC^-_*(CF^*(L,L)) \rightarrow QH^*(X;Q). \end{aligned}$$
(251)

We have thus proved Theorem 1.7.

Theorem 4.71

(Theorem 1.7) Under Assumptions 4.18 the cyclic open–closed map

$$\begin{aligned} \mathcal{O}\mathcal{C}^-: HC^{-}(CF^*(L,L)) \rightarrow QH^*(X;Q_U)[[u]] \end{aligned}$$

is a morphism of TE-structures.

5 Analysis on the horocyclic moduli space

The main goal of this chapter is to prove Proposition 4.31. Finally, in the last section we proof Proposition 4.32.

The following sections use the method of proof explained to the author in an unpublished draft by Jake Solomon and Sara Tukachinsky. We prove the following result for the operations \({\mathfrak {q}}^{ST}_{k,l,\perp }\), which were defined using the sign convention similar to [47].

Proposition 5.1

$$\begin{aligned} 0&= \sum _{\begin{array}{c} S_3[l]\\ (2:3) = \{ i \} \end{array}} (-1)^{1+|\gamma ^{(1:3)}|} \mathfrak {q}^{ST}_{k,l;\perp _0}(\otimes _{j=1}^{k} \alpha _j; \gamma ^{(1:3)} \otimes d\gamma _i \otimes \gamma ^{(3:3)})\end{aligned}$$
(252)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2 =[l] \\ 1,2 \in J_2 \end{array}} (-1)^{sign^\gamma (J_1,J_2)}\mathfrak {q}^{ST}_{k, j_1}(\alpha ; \mathfrak {q}^{ST}_{\emptyset , j_2}(\gamma ^{J_2}) \otimes \gamma ^{J_1})\end{aligned}$$
(253)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2 =[l] \\ 1,2 \in J_2\\ P \in S_3[k] \end{array}} (-1)^{sign^\gamma (J_1,J_2) + \epsilon _1|\gamma ^{J_2}| + 1} \mathfrak {q}^{ST}_{k_1+k_3+1, j_1} (\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST}_{k_2,j_2;\perp _0}(\alpha ^{(2:3)}; \gamma ^{J_2})\nonumber \\&\quad \otimes \alpha ^{(3:3)}; \gamma ^{J_1})\end{aligned}$$
(254)
$$\begin{aligned}&\quad + \sum _{\begin{array}{c} J_1 \cup J_2=[l]\\ 1,2 \in J_1\\ P \in S_3[k] \end{array}} (-1)^{sign^\gamma (J_1,J_2) + \epsilon _1(|\gamma ^{J_2}|+1) + |\gamma ^{J_1}|} \mathfrak {q}^{ST}_{k_1+k_3+1, j_1;\perp _0} (\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST}_{k_2,j_2}(\alpha ^{(2:3)};\gamma ^{J_2})\nonumber \\&\quad \otimes \alpha ^{(3:3)}; \gamma ^{J_1})\end{aligned}$$
(255)
$$\begin{aligned}&\quad +\sum _{\begin{array}{c} J_1 \cup J_2 \cup J_3 = [l]\\ 1\in J_3,\;2\in J_2\\ P \in S_5[k] \end{array}} (-1)^{A_5}\mathfrak {q}^{ST}_{k_1+k_3 + k_5+2, j_1} (\alpha ^{(1:5)} \otimes \mathfrak {q}^{ST}_{k_2,j_2}(\alpha ^{(2:5)};\gamma ^{J_2})\otimes \alpha ^{(3:5)} \nonumber \\&\quad \otimes \mathfrak {q}^{ST}_{k_4,j_3}(\alpha ^{(4:5)};\gamma ^{J_3}) \otimes \alpha ^{(5:5)};\gamma ^{J_1}). \end{aligned}$$
(256)

Here

$$\begin{aligned} A_5 = sign^{\gamma }(J_1,J_2,J_3)+ |\gamma ^{J_2}|+ (|\gamma ^{J_2}| + 1)\epsilon _1 + (|\gamma ^{J_3}| + 1)(\epsilon _1 + \epsilon _2 + \epsilon _3)\nonumber \\ \end{aligned}$$
(257)

Proposition 4.31 then follows from the above by a direct verification of signs. In the following, we prove Proposition 5.1 assuming that the \(U(\Lambda )\)-local system is trivial. The general result then follows easily.

Recall that the orientation on \(\mathcal {M}_{k+1,l; \perp _0}(\beta )\) is defined by the fibre-product orientation, as defined in [46, Section 2.2]. We take the orientation on I to be the positive orientation, so that \(\partial I = \{1\} -\{ 0\}\). The boundary is then identified as:

$$\begin{aligned} \partial \mathcal {M}_{k+1,l; \perp _0}(\beta ) \cong \partial (I \times _{D^2} \mathcal {M}_{k+1,l}(\beta )) = \partial I \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) - I \times _{D^2} \partial \mathcal {M}_{k+1,l}(\beta ).\nonumber \\ \end{aligned}$$
(258)

We now further decompose each of the terms in the boundary. For each, we identity them with a fibre product of other moduli spaces (both with and without horocyclic constraints).

5.1 Signs of boundary components

In this section we identify boundary components of the moduli spaces with fibre products of different moduli spaces. We compute the difference in orientation between the induced orientation on the boundary components, and the fibre product orientation. First we consider the boundary components coming from \(I \times _{D^2} \partial \mathcal {M}_{k+1,l}(\beta )\). Let \(k = k_1 + k_2 + k_3\) and write \(\mathcal {M}_1:= \mathcal {M}_{k_1 + k_3+2,l}(\beta _1)\), \(\mathcal {M}_2:= \mathcal {M}_{k_2+1,l}(\beta _2)\). Finally write \(\mathcal {M}_{j,\perp } = I \times _{D^2} \mathcal {M}_{j}\). Let \(B_\perp \) be a boundary component where a disk bubbles off at the \((k_1 + 1)\)-th boundary point, with \(k_2\) of the boundary marked points and the interior marked points labelled by J. The boundary \(I \times _{D^2} \partial \mathcal {M}_{k+1,l}(\beta )\) can be decomposed into two components, \(B_{\perp ,1}\), where the bubbling is not at the zeroth marked point and \(B_{\perp ,2}\), where the bubbling is at the zeroth marked point.

Lemma 5.2

There exists diffeomorphisms:

$$\begin{aligned}&\phi _1: {{\mathcal {M}_{1,\perp }}_{evb_{k_1 + 1}^{\beta _1}}}{\times }_{evb_0^{\beta _2}} \mathcal {M}_2 \xrightarrow {\sim } B_{\perp ,1}, \end{aligned}$$
(259)
$$\begin{aligned}&\phi _2: {{\mathcal {M}_{1}}_{evb_{k_1 + 1}^{\beta _1}}}{\times }_{evb_0^{\beta _2}} \mathcal {M}_{2,\perp } \xrightarrow {\sim } B_{\perp ,2}. \end{aligned}$$
(260)

The maps \(\phi _j\) change the orientation by \(sign(\phi _j)\), where:

$$\begin{aligned} sign(\phi _1)&= (-1)^{1+k_2k_3 +k_1 + n},\end{aligned}$$
(261)
$$\begin{aligned} sign(\phi _2)&= (-1)^{k_3(k_2 + 1) + n + 1}. \end{aligned}$$
(262)

The proof of this lemma uses:

Proposition 5.3

([47, Proposition 2.8]) Let \(k,l \in \mathbb {Z}_{\ge 0}\). Let \(P \in S_3[k]\) and \(\beta _1 + \beta _2 = \beta \in H_2(X,L)\). Let \(I \sqcup J = [l]\) be a partition. Let \(B \subset \partial \mathcal {M}_{k+1,l}(\beta )\) be the boundary component where a disk bubbles off at the \(k_1 + 1\)-th boundary point, with \(k_2\) of the boundary marked points and the interior marked points labelled by J lying on the bubble disk. Then the canonical diffeomorphism

$$\begin{aligned} \theta : {{\mathcal {M}_{k_1 + k_3+1,l}(\beta _1)}_{evb_{k_1 + 1}^{\beta _1}}}{\times }_{evb_0^{\beta _2}} \mathcal {M}_{k_2+1,l}(\beta _2) \rightarrow B \end{aligned}$$
(263)

changes orientation by the sign \((-1)^{\delta _1}\), with

$$\begin{aligned} \delta _1 = k_2k_3 + k_1 + n. \end{aligned}$$
(264)

Proof of Lemma 5.2

We can decompose \(\phi _1\) as

$$\begin{aligned} \mathcal {M}_{1,\perp } \times _L \mathcal {M}_2 \rightarrow (I \times _{D^2} \mathcal {M}_1) \times _L \mathcal {M}_2 \xrightarrow {m_1} I \times _{D^2} (\mathcal {M}_1 \times _L \mathcal {M}_2) \xrightarrow {\hat{\theta }} I \times _{D^2} B \xrightarrow {t} B_{\perp ,1}.\nonumber \\ \end{aligned}$$
(265)

Here \(\hat{\theta }\) is the map induced by \(\theta \) from the Proposition 5.3. By [7, Lemma 8.2.3(4)], we have \(sign(\hat{\theta }) = sign(\theta )\). From Eq. (258) it is clear that \(sign(t) = -1\). Finally, from the associativity of the fibre product [7, Lemma 8.2.3(2)], \(sign(m_1) = 1\). Thus

$$\begin{aligned} sign(\phi _1) = (-1)^{1+k_2k_3 +k_1 + n}. \end{aligned}$$
(266)

Similarly, we decompose \(\phi _2\) as:

(267)

We can compute \(sign(m_2)\) as follows:

$$\begin{aligned} \mathcal {M}_1 \times _L (I \times _{D^2}\mathcal {M}_2)&= (-1)^{(k_2 + 1)(k_1+k_3)} (I \times _{D^2}\mathcal {M}_2) \times _L \mathcal {M}_1 \end{aligned}$$
(268)
$$\begin{aligned}&= (-1)^{(k_2 + 1)(k_1+k_3+1)} I \times _{D^2} (\mathcal {M}_2 \times _L \mathcal {M}_1)\end{aligned}$$
(269)
$$\begin{aligned}&= (-1)^{(k_2 + 1)(k_1+k_3+1) + k_2(k_1+k_3))} I \times _{D^2} (\mathcal {M}_1 \times _L \mathcal {M}_2). \end{aligned}$$
(270)

So that \(sign(m_2) = (-1)^{k_1+k_3+1}\), and thus:

$$\begin{aligned} sign(\phi _2) = (-1)^{k_3(k_2 + 1) + n + 1}. \end{aligned}$$
(271)

Now we consider the boundary components coming from \(\partial I \times _{D^2} \mathcal {M}_{k+1,l}(\beta )\). First we set up some notation regarding the moduli spaces of holomorphic disks. We have

$$\begin{aligned} \widetilde{\mathcal {M}}^{main}(\beta )= & {} \{ u:(D^2,S^1) \rightarrow (X,L)| u \text { is holomorphic and } u_*([D^2,S^1]) \nonumber \\= & {} \beta \in H_2(X,L) \}. \end{aligned}$$
(272)

The superscript main here denotes that this is the top-dimensional stratum of the moduli space. The moduli spaces \(\widetilde{\mathcal {M}}^{main}(\beta )\) are oriented using the relative spin structure as in [7, Chapter 8]. Adding marked points and quotienting by \(Aut(D^2)\) gives the moduli spaces \(\mathcal {M}_{k+1,l}^{main}(\beta )\) as open subsets

$$\begin{aligned} \mathcal {M}_{k+1,l}^{main}(\beta ) \subset \left( \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{ k+1} \times (D^2)^{ l}\right) /Aut(D^2). \end{aligned}$$
(273)

Here we need to be careful about the ordering in \((S^1)^{k+1}\). We stick to the convention in [7], so that \((S^{1})^{k+1} = S^1_0 \times S^1_1 \times \dots \times S^1_{k}\). Here \(S^1_i\) is the circle corresponding to the i’th boundary marked point. The orientation of a quotient by a Lie group is defined as in [7]. This means that the orientation on \(\mathcal {M}_{k+1,l}(\beta )\) is such that there exists an orientation preserving local diffeomorphism:

$$\begin{aligned} \mathcal {M}^{main}_{k+1,l}(\beta ) \times Aut(D^2) \cong \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{k+1} \times (D^2)^{ l}. \end{aligned}$$
(274)

Lemma 5.4

Let \(k\ge 1\) and \(l \ge 1\). Fixing the 0th boundary marked point at \(-i\), and the first interior marked point at 0 defines a local diffeomorphism:

$$\begin{aligned} \mathcal {M}^{main}_{k+1,l}(\beta ) \rightarrow \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{ k} \times (D^2)^{ l-1}, \end{aligned}$$
(275)

which changes orientation by \((-1)^k\).

Proof

Recall that in [7] the orientation on \(Aut(D^2)\) is given by considering the local diffeomorphism:

$$\begin{aligned} Aut(D^2)&\rightarrow (S^1)^{3}\nonumber \\ g&\mapsto (gz_0, g z_1, gz_2), \end{aligned}$$
(276)

for three points \(z_0, z_1, z_2 \subset S^1\) in counter-clockwise ordering. By definition, this map is orientation preserving. One can check that the map

$$\begin{aligned} Aut(D^2)&\rightarrow S^1 \times D^2\nonumber \\ g&\mapsto (g\cdot -i, g \cdot 0), \end{aligned}$$
(277)

is also orientation preserving. Now multiply both sides by \(\widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^k \times (D^2)^{l-1}\) on the left, and commute the various terms through to obtain a local diffeomorphism:

$$\begin{aligned} \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{ k} \times (D^2)^{ l-1} \times Aut(D^2) \cong (-1)^k \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{ k+1} \times (D^2)^{ l}.\nonumber \\ \end{aligned}$$
(278)

The sign \((-1)^k\) here comes from the change in ordering from \((S^1)^k \times S^1_0 \cong (-1)^k (S^1)^{k+1}\). Finally, apply Eq. (274) and cancel the factor \(Aut(D^2)\) to obtain the result. \(\square \)

Similar considerations show:

Lemma 5.5

Let \(k = k_1 + k_2 + k_3\). Fixing the three boundary marked points with indices \(0, k_1+1\) and \(k_1 + k_2 +2\), we obtain a local diffeomorphism:

$$\begin{aligned} \mathcal {M}_{k+3,l}^{main}(\beta ) \rightarrow \widetilde{\mathcal {M}}^{main}(\beta ) \times (S^1)^{k} \times (D^2)^{l-1}, \end{aligned}$$
(279)

which changes orientation by \((-1)^{k + k_2} = (-1)^{k_1 + k_3}\).

Lemma 5.6

Fixing the first three marked points at \(0,1,\infty \) gives an orientation preserving local diffeomorphism:

$$\begin{aligned} \mathcal {M}_{\emptyset ,l_1+1}^{main}(\beta ) \rightarrow \widetilde{\mathcal {M}}^{main}_{\emptyset }(\beta ) \times (S^2)^{l_1 - 2}. \end{aligned}$$
(280)

We now want to study the boundary components \(\partial I \times _{D^2} \mathcal {M}_{k+1,l}(\beta )\). Observe that:

$$\begin{aligned} \partial I \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) = \{1\} \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) - \{ 0\}\times _{D^2} \mathcal {M}_{k+1,l}(\beta ). \end{aligned}$$
(281)

First we look at the case where the two interior marked points collide. This corresponds to \(\{ 0 \} \times _{D^2} \mathcal {M}_{k+1,l}(\beta )\). Let \(B_{\perp ,3}\) be a boundary component where the interior marked points labelled by I bubble off on a sphere. Note that \(1,2 \in I\). Together with the output marked point on the sphere, this gives at least 3 marked points.

For gluing the moduli spaces of holomorphic maps \(\widetilde{\mathcal {M}}^{main}_{\emptyset }(\beta _1)\) and \(\widetilde{\mathcal {M}}^{main}(\beta _2)\), we use the following:

Proposition 5.7

(Lemma C.1) The gluing map

$$\begin{aligned} \widetilde{\mathcal {M}}^{main}_{\emptyset }(\beta _1) \times _{X} \widetilde{\mathcal {M}}^{main}(\beta _2) \rightarrow \widetilde{\mathcal {M}}^{main}(\beta _1 + \beta _2) \end{aligned}$$
(282)

is a local diffeomorphism which changes orientation by \((-1)^{w_{\mathfrak {s}}(\beta _1)}\).

Remark 5.8

This proposition is implicit in [44] and the statement was communicated to the author by Sara Tukachinsky. See also [15, Remark 2.7]. As far as the author is aware, the proof of this statement has not appeared in any literature before. We thus prove it in Appendix 1.

We then prove:

Proposition 5.9

The canonical local diffeomorphism

$$\begin{aligned} \phi _3: \mathcal {M}_{\emptyset ,l_1+1}(\beta _1) \times _X \mathcal {M}_{k+1,l_2+1}(\beta _2) \xrightarrow {\sim } B_{\perp ,3} \subset \partial \mathcal {M}_{k+1,l; \perp _0}(\beta ) \end{aligned}$$
(283)

changes orientation by \(sign(\phi _3) = (-1)^{1+w_{\mathfrak {s}}(\beta _1)}\). Here \(\beta = \beta _1 + \beta _2\).

Proof

Let (vu) be a stable map, where \(v: S^2 \rightarrow X\), \(u: (D^2,S^1) \rightarrow (X,L)\) and \(evi_0(v) = evi_0(u) = x \in X\). We will compute the change in orientation locally at (vu). By definition of the fibre product orientation, we have:

$$\begin{aligned} T_{(0,(v,u))} \{ 0 \} \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) \oplus T_0D^2 \cong T_u {\mathcal {M}}_{k+1,l}(\beta ). \end{aligned}$$
(284)

by Lemma 5.4 we have:

$$\begin{aligned}&T_u {\mathcal {M}}_{k+1,l}(\beta ) \oplus T_xX \cong (-1)^kT_{(v,u)}\widetilde{\mathcal {M}}(\beta ) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l - 1} \oplus T_xX. \end{aligned}$$
(285)

Now use Proposition 5.7 to rewrite this as:

$$\begin{aligned}&\cong (-1)^{w_{\mathfrak {s}}(\beta _1)+k} T_{(v,u)}\widetilde{\mathcal {M}}_{\emptyset }(\beta _1) \times _X \widetilde{\mathcal {M}}(\beta _2)\oplus T_xX \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l - 1}. \end{aligned}$$
(286)

By definition of the fibre product orientation, this is isomorphic to

$$\begin{aligned}&\cong (-1)^{w_{\mathfrak {s}}(\beta _1)+k} T_{v}\widetilde{\mathcal {M}}_{\emptyset }(\beta _1) \oplus T_u \widetilde{\mathcal {M}}(\beta _2) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l - 1}. \end{aligned}$$
(287)

Next, we rearrange the terms, to obtain:

$$\begin{aligned}&\cong (-1)^{w_{\mathfrak {s}}(\beta _1)+k} (T_{v}\widetilde{\mathcal {M}}_{\emptyset }(\beta _1) \oplus {\mathbb {C}}^{l_1 -2}) \oplus (T_u \widetilde{\mathcal {M}}(\beta _2) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l_2}) \oplus {\mathbb {C}}. \end{aligned}$$
(288)

By Lemmas 5.4 and 5.6, this is isomorphic to:

$$\begin{aligned}&\cong (-1)^{w_{\mathfrak {s}}(\beta _1)} T_{v}\mathcal {M}_{\emptyset ,l_1+1}(\beta _1) \oplus T_u \mathcal {M}_{k,l_2}(\beta _2) \oplus {\mathbb {C}}. \end{aligned}$$
(289)

Again, by definition of the fibre product orientation, this is isomorphic to:

$$\begin{aligned}&\cong (-1)^{w_{\mathfrak {s}}(\beta _1)} T_{(v,u)}\mathcal {M}_{\emptyset ,l_1+1}(\beta _1) \otimes _X T_u \mathcal {M}_{k,l_2}(\beta _2) \oplus T_xX \oplus {\mathbb {C}}. \end{aligned}$$
(290)

Thus, as \(T_0\mathbb {D} \cong {\mathbb {C}}\), and cancelling the terms \(T_xX\), we obtain:

$$\begin{aligned} T_{(0,(v,u))} \{ 0 \} \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) \cong (-1)^{w_{\mathfrak {s}}(\beta _1)} T_{(v,u)}\mathcal {M}_{\emptyset ,l_1+1}(\beta _1) \times _X T_u \mathcal {M}_{k,l_2}(\beta _2)\nonumber \\ \end{aligned}$$
(291)

The extra change in sign then comes from Eq. (281). \(\square \)

Next we consider the case when one of the horocyclicly constrained points collides with the boundary marked point. This corresponds to \(\{1 \} \times _{D^2} \mathcal {M}_{k+1,l}(\beta )\). Here two disks bubble off on either side of the disk. Let \(B_{\perp ,4}\) be this boundary component. We show:

Proposition 5.10

The map

$$\begin{aligned}{} & {} \phi _4: \mathcal {M}_{k_4+1, l_3}(\beta _3) \times _L \mathcal {M}_{k_1+k_3+k_5+3,l_1}(\beta _1) \times _L \mathcal {M}_{k_2+1,l_2}(\beta _2) \rightarrow B_{\perp ,4} \nonumber \\{} & {} \quad \subset \partial \mathcal {M}_{k+1,l,\perp _0}(\beta ) \end{aligned}$$
(292)

changes orientation with \(sign(\phi _4) = k_4(k_1 + k_2 + k_3) + k_2(k_3+k_5) +k_3\).

Applying [7, Lemma 8.3.5] twice shows:

Lemma 5.11

The gluing map:

$$\begin{aligned} \theta : \widetilde{\mathcal {M}}(\beta _3) \times _L \widetilde{\mathcal {M}}(\beta _1) \times _L \widetilde{\mathcal {M}}(\beta _2) \rightarrow \widetilde{\mathcal {M}}(\beta ), \end{aligned}$$
(293)

is an orientation preserving local diffeomorphism. Here \(\beta = \beta _1 + \beta _2 + \beta _3\).

Proof of 5.10

Let \(u = (u_1,u_2,u_3) \in B_{\perp ,4}\) be a stable map. For simplicity, write \({\mathcal {M}}_1 = \mathcal {M}_{k_1+k_3+k_5+3,l_1}\), \({\mathcal {M}}_2 = \mathcal {M}_{k_2+1, l_2}(\beta _2)\) and \(\mathcal {M}_{k_4+1, l_3}(\beta _3)\). We first note that by definition of the fibre product orientation:

$$\begin{aligned} T_{(1,u)} \{1\} \times _{D^2} \mathcal {M}_{k+1,l}(\beta ) \oplus T_1D^2 \cong T_u \mathcal {M}_{k+1,l}(\beta ). \end{aligned}$$
(294)

We then use Lemma 5.4 to write:

$$\begin{aligned}&T_u\mathcal {M}_{k+1,l}(\beta ) \oplus TL \oplus TL \cong (-1)^k \widetilde{\mathcal {M}}(\beta ) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l-1} \oplus TL \oplus TL, \end{aligned}$$
(295)

which, by Proposition 5.11 is isomorphic to:

$$\begin{aligned}&\cong (-1)^k T_u (\widetilde{\mathcal {M}}(\beta _3) \times _L \widetilde{\mathcal {M}}(\beta _1) \times _L \widetilde{\mathcal {M}}(\beta _2)) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l-1} \oplus TL \oplus TL. \end{aligned}$$
(296)

By applying the definition of the fibre product orientation twice, this is isomorphic to:

$$\begin{aligned}&\cong (-1)^{k+n} T_{u_3} \widetilde{\mathcal {M}}(\beta _3) \oplus T_{u_1}\widetilde{\mathcal {M}}(\beta _1) \oplus T_{u_2} \widetilde{\mathcal {M}}(\beta _2) \oplus \mathbb {R}^{k} \oplus \mathbb {C}^{l-1}. \end{aligned}$$
(297)

Commuting the various terms through, noting that \(\mathbb {C}\) is even dimensional, and \(\mathbb {R}\) is odd dimensional, this gives:

$$\begin{aligned} \begin{aligned} \cong (-1)^{k+n + A} (T_{u_3} \widetilde{\mathcal {M}}(\beta _3) \oplus {\mathbb {R}}^{k_4} \oplus {\mathbb {C}}^{l_3-1}) \oplus (T_{u_1}\widetilde{\mathcal {M}}(\beta _1) \oplus {\mathbb {R}}^{k_1+k_3 + k_5} \oplus {\mathbb {C}}^{l_1}) \\ \oplus (T_{u_2} \widetilde{\mathcal {M}}(\beta _2) \oplus \mathbb {R}^{k_2} \oplus \mathbb {C}^{l_2-1}) \oplus {\mathbb {C}}, \end{aligned} \end{aligned}$$
(298)

where \(A = k_4(k_1+k_2+k_3) + n(k_1+k_3+k_5) + k_2(k_3+k_5)\). Then apply Lemmas 5.4 and 5.5 to find:

$$\begin{aligned}&\cong (-1)^{k+n + A + B} T_{u_3} T_{u_3}{\mathcal {M}}_3 \oplus T_{u_1} {\mathcal {M}}_1 \oplus T_{u_2} {\mathcal {M}}_2 \oplus {\mathbb {C}}, \end{aligned}$$
(299)

where \(B = k_1 + k_2 + k_4 + k_5\). Finally, apply the definition of the fibre product orientation twice to obtain:

$$\begin{aligned}&\cong (-1)^{k+n + A+ B + n(k_1+k_3+k_5 + n)} T_u ({\mathcal {M}}_3 \times _L {\mathcal {M}}_1 \times _L {\mathcal {M}}_2) \oplus TL \oplus TL \oplus {\mathbb {C}}. \end{aligned}$$
(300)

The result then follows by cancelling the factors TL and noting that \(T_1D^2 \cong {\mathbb {C}}\). \(\square \)

5.2 Stokes’ theorem and push-forward

The next step is to apply Stokes’ theorem for the push-forward of differential forms.

Theorem 5.12

([46, Theorem 1]) Let M be a smooth orbifold with boundary. For a smooth submersion \(f: M\rightarrow N\) of relative dimension s, and \(\xi \in A^t(M)\). We have:

$$\begin{aligned} 0 = f_*(d\xi ) - d(f_*\xi ) + (-1)^{s+t}(f|_{\partial M})_*\xi . \end{aligned}$$
(301)

Recall the following facts about the push-forward of differential forms, see [46]:

Lemma 5.13

  1. 1.

    Let \(f:M\rightarrow N\) be a proper submersion, \(\alpha \in A^*(N)\), \(\beta \in A^*(M)\). Then:

    $$\begin{aligned} f_*(f^*\alpha \wedge \beta ) = \alpha \wedge f_*\beta . \end{aligned}$$
    (302)
  2. 2.

    Let

    (303)

    be a pull-back diagram of smooth maps, where g and f are proper submersions. Let \(\alpha \in A^*(P)\). Then:

    $$\begin{aligned} q_*p^*\alpha = f^*g_*\alpha . \end{aligned}$$
    (304)

    Similarly, if \(\beta \in A^*(M)\), then:

    $$\begin{aligned} p_*q^*\beta = (-1)^{(dim(M)-dim(N))(dim(P)-dim(N))}g^*f_*\beta . \end{aligned}$$
    (305)

To obtain the structure equations for the \(\mathfrak {q}^{ST}_{\perp }\) operations, we will apply Stokes’ theorem with \(M = \mathcal {M}_{k+1,l, \perp _0}(\beta )\), \(N = L\), \(f = evb_0\) and \(\xi = \bigwedge _{i=1} evi^*\gamma _i \wedge \bigwedge _{j=1} evb_j^*\alpha _j\). When it is clear which evaluation maps are used, we will simply write \(evi^*\gamma \) for \(\bigwedge _{i=1} evi^*\gamma _i\) and similarly for the boundary evaluations.

The first term in Stokes’ theorem is:

$$\begin{aligned} (evb_0)_*(d\xi )&= (evb_0)_*\left( \sum _{\begin{array}{c} S_3[l]\\ (2:3) = \{ i \} \end{array}} (-1)^{|\gamma ^{(1:3)}|} evi^*(\gamma ^{(1:3)} \wedge d\gamma _i \wedge \gamma ^{(3:3)})\wedge evb^*\alpha \right. \end{aligned}$$
(306)
$$\begin{aligned}&\left. \qquad \quad +\sum _{\begin{array}{c} S_3[k]\\ k_2 = 1 \end{array}} (-1)^{|\gamma | + \epsilon _1 + k_1}evi^*\gamma \wedge evb^* (\alpha ^{(1:3)} \wedge d\alpha _{k_1+1} \wedge \alpha ^{(3:3)})\right) , \end{aligned}$$
(307)

by definition of the \(\mathfrak {q}\) operations, this equals:

$$\begin{aligned}&= \sum _{\begin{array}{c} S_3[l]\\ (2:3) = \{ i \} \end{array}} (-1)^{|\gamma ^{(1:3)}|+\zeta (\alpha ) + \zeta _{\perp }(\alpha ,d\gamma )} \mathfrak {q}^{ST,\beta }_{k,l;\perp _0}(\otimes _{j=1}^{k} \alpha _j; \gamma ^{(1:3)} \otimes d\gamma _i \otimes \gamma ^{(3:3)}) \end{aligned}$$
(308)
$$\begin{aligned}&\qquad \quad + \sum _{\begin{array}{c} S_3[k]\\ k_2 = 1 \end{array}} (-1)^{|\gamma |+\zeta (\alpha )+1+\epsilon _1 + \zeta _{\perp }(d\alpha ,\gamma )} \mathfrak {q}^{ST,\beta }_{k,l;\perp _0}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST, \beta _0}_{1,0}(\alpha _{k_1+1}) \otimes \alpha ^{(3:3)}; \gamma ), \end{aligned}$$
(309)

by expanding the signs \(\zeta _{\perp }\), we find:

$$\begin{aligned}&= \sum _{\begin{array}{c} S_3[l]\\ (2:3) = \{ i \} \end{array}} (-1)^{n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + |\gamma ^{(1:3)}| + 1} \mathfrak {q}^{ST, \beta }_{k,l;\perp _0}(\otimes _{j=1}^{k} \alpha _j; \gamma ^{(1:3)} \otimes d\gamma _i \otimes \gamma ^{(3:3)}) \end{aligned}$$
(310)
$$\begin{aligned}&\qquad \quad + \sum _{\begin{array}{c} S_3[k]\\ k_2 = 1 \end{array}} (-1)^{n+|\gamma |+\zeta (\alpha ) + \epsilon (\alpha ) +\epsilon _1 + |\gamma |} \mathfrak {q}^{ST,\beta }_{k,l;\perp _0}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST,\beta _0}_{1,0}(\alpha _{k_1+1}) \otimes \alpha ^{(3:3)}; \gamma ). \end{aligned}$$
(311)

The second term in Stokes’ theorem reads:

$$\begin{aligned} -d((evb_0)_*\xi ) = -\mathfrak {q}^{ST,\beta _0}_{1,0}((evb_0)_*\xi )&= (-1)^{\zeta (\alpha ) + 1 + \zeta _{\perp }(\alpha ,\gamma )} \mathfrak {q}^{ST,\beta _0}_{1,0}(\mathfrak {q}^{ST,\beta }_{k,l,\perp _0}(\alpha ,\gamma )) \end{aligned}$$
(312)
$$\begin{aligned}&=(-1)^{n + |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + 1} \mathfrak {q}_{1,0}^{ST,\beta _0}(\mathfrak {q}^{ST,\beta }_{k,l,\perp _0}(\alpha ,\gamma )). \end{aligned}$$
(313)

The final term in Stokes’ theorem is given by restricting to the various boundary components \(B_{\perp ,i}\) for \(i = 1,2,3,4\). We first compute the overall sign \((-1)^{s+t}\). Note that \(|\xi | = |\gamma | + |\alpha |\), \(dim(M) \equiv k+1 \;(mod \; 2)\) and \(dim(L) = n\), so \(s \equiv k+1+n\). The overall sign is thus

$$\begin{aligned} s+t \equiv |\gamma | + \epsilon (\alpha ) + n + 1\; (mod \; 2). \end{aligned}$$
(314)

5.2.1 \(B_{\perp ,1}\)

Here we will show:

Lemma 5.14

The terms in Stokes’ theorem coming from \((evb_0|_{B_{1,\perp }})_*\xi \) are

$$\begin{aligned} \sum _{\begin{array}{c} P \in S_3[k]\\ \beta _1 + \beta _2 = \beta \\ \beta _2 \ne 0\\ J_1 \cup J_2 = [l]\\ \gamma _1,\gamma _2 \in J_1 \end{array}} (-1)^{C_1} \mathfrak {q}_{k_1+1+k_3,l_1,\perp }^{ST,\beta _1}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST,\beta _2}_{k_2,l_2}(\alpha ^{(2:3)};\gamma ^{J_2}) \otimes \alpha ^{(3:3)}; \gamma ^{J_1}), \end{aligned}$$
(315)

where \(C_1 = n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2) + (|\gamma ^{J_2}|+1)\epsilon _1 + |\gamma ^{J_1}|\).

Recall that

$$\begin{aligned} \phi _1: {{\mathcal {M}_{k_1 + k_3+2,l_1,\perp _0}(\beta _1)} _{evb_{k_1 + 1}^{\beta _1}}}{\times }_{evb_0^{\beta _2}} \mathcal {M}_{k_2+1,l_2}(\beta _2) \xrightarrow {\sim } B_{\perp ,1} \end{aligned}$$
(316)

changes orientation by \(\delta _1:= sign(\phi _1) = n + k_1 + k_2k_3\).

Denote everything associated with \(\mathcal {M}_{k_1 + k_3+2,l_1,\perp _0}(\beta _1)\) with a subscript 1 and everything associated with \(\mathcal {M}_{k_2+1,l_2}(\beta _2)\) with a subscript 2. Consider the commutative diagram:

(317)

Let

$$\begin{aligned} {\overline{\xi }}&= \phi _1^*\xi , \end{aligned}$$
(318)
$$\begin{aligned} \xi _1&= (evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*(\alpha ^{(1:3)}\wedge \alpha ^{(3:3)}),\end{aligned}$$
(319)
$$\begin{aligned} \xi _2&= (evi^2)^*\gamma ^{J_2} \wedge (evb^2)^*\alpha ^{(2:3)}. \end{aligned}$$
(320)

Define \(\delta _2\) by

$$\begin{aligned} p_1^*\xi _1 \wedge p_2^*\xi = (-1)^{\delta _2}{\overline{\xi }}. \end{aligned}$$
(321)

We thus find:

$$\begin{aligned} \delta _2 = sign^\gamma (J_1,J_2) + |\gamma ^{J_2}|(\epsilon _1 + k_1 + \epsilon _3 + k_3) + (\epsilon _2+k_2)(\epsilon _3 + k_3). \end{aligned}$$
(322)

Now compute

$$\begin{aligned} (evb_0|_{B_{\perp ,1}})_*\xi&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p_1)_*(p_1^*\xi _1 \wedge p_2^*\xi _2), \end{aligned}$$
(323)

by using the relation (302), this equals:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(\xi _1 \wedge (p_1)_*(p_2)^*\xi _2), \end{aligned}$$
(324)

using Eq. (304) we obtain:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(\xi _1 \wedge (evb^1_{k_1+1})^*(evb_0^2)_* \xi _2), \end{aligned}$$
(325)

which, by definition of \(\mathfrak {q}^{ST}\) equals:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 + \delta _3} (evb^1_0)_*(\xi _1 \wedge \mathfrak {q}^{ST,\beta _2}_{k_2,l_2}(\alpha ^{(2:3)};\gamma ^{J_2})), \end{aligned}$$
(326)

where \(\delta _3 = \zeta (\alpha ^{(2:3)})\). Expanding \(\xi _1\) and rearranging gives:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 +\delta _3 + \delta _4} (evb^1_0)_*((evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*(\alpha ^{(1:3)} \wedge \mathfrak {q}^{ST,\beta _2}_{k_2,l_2}(\alpha ^{(2:3)};\gamma ^{J_2}) \wedge \alpha ^{(3:3)})), \end{aligned}$$
(327)

where \(\delta _4 = (|\gamma ^{J_2}| + \epsilon _2)(\epsilon _3 + k_3)\). Finally by definition this equals:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 +\delta _3 + \delta _4 + \delta _5} \mathfrak {q}_{k_1+1+k_3,l_1,\perp }^{ST,\beta _1}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST,\beta _2}_{k_2,l_2}(\alpha ^{(2:3)};\gamma ^{J_2}) \otimes \alpha ^{(3:3)}; \gamma ^{J_1}), \end{aligned}$$
(328)

where \(\delta _5 = \zeta (\alpha ^{(1:3)},\mathfrak {q}^{ST}(\alpha ^{(2:3)};\gamma ^{J_2}),\alpha ^{(3:3)}) + \zeta _{\perp }(\alpha ^{(1:3)},\mathfrak {q}^{ST}(\alpha ^{(2:3)};\gamma ^{J_2}),\alpha ^{(3:3)})\). Adding all signs together with the sign in Stokes’ theorem, we get an overall sign:

$$\begin{aligned} C_1 = n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2) + (|\gamma ^{J_2}|+1)\epsilon _1 + |\gamma ^{J_1}|. \end{aligned}$$
(329)

Here we have used [47, Lemma 2.9] to compute

$$\begin{aligned}{} & {} \zeta (\alpha ^{(2:3)}) + \zeta (\alpha ^{(1:3)},\mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:3)};\gamma ^{J_2}),\alpha ^{(3:3)}) \nonumber \\{} & {} \quad = \zeta (\alpha ) + \epsilon (\alpha ) + \epsilon _1 + (k_1+1)|\gamma ^{J_2}| + k_3k_2. \end{aligned}$$
(330)

5.2.2 \(B_{\perp ,2}\)

We repeat the above argument for the boundary component \(B_{\perp ,2}\). We show:

Lemma 5.15

The terms in Stokes’ theorem coming from \((evb_0|_{B_{2,\perp }})_*\xi \) are

$$\begin{aligned} \sum _{\begin{array}{c} P \in S_3[k]\\ \beta _1 + \beta _2 = \beta \\ \beta _1 \ne 0\\ J_1 \cup J_2 = [l]\\ \gamma _1,\gamma _2 \in J_2 \end{array}} (-1)^{C_2} \mathfrak {q}_{k_1+1+k_3,l_1}^{ST,\beta _1}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST,\beta _2}_{k_2,l_2,\perp }(\alpha ^{(2:3)};\gamma ^{J_2}) \otimes \alpha ^{(3:3)}; \gamma ^{J_1}), \end{aligned}$$
(331)

where \( C_2 = n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2) + |\gamma ^{J_2}|\epsilon _1 + 1\).

Recall that

$$\begin{aligned} \phi _2: {{\mathcal {M}_{k_1 + k_3+2,l_1}(\beta _1)}_{evb_{k_1 + 1}^{\beta _1}}}{\times }_{evb_0^{\beta _2}} \mathcal {M}_{k_2+1,l_2,\perp _0}(\beta _2) \xrightarrow {\sim } B_{\perp ,2} \end{aligned}$$
(332)

changes orientation by \(\delta _1:= sign(\phi _1) = n + k_3 + k_2k_3 + 1\).

Denote everything associated with \(\mathcal {M}_{k_1 + k_3+2,l_1}(\beta _1)\) with a subscript 1, everything associated with \(\mathcal {M}_{k_2+1,l_2,\perp _0}(\beta _2)\) with a subscript 2. Consider the commutative diagram:

(333)

Let

$$\begin{aligned} {\overline{\xi }}&= \phi _2^*\xi , \end{aligned}$$
(334)
$$\begin{aligned} \xi _1&= (evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*(\alpha ^{(1:3)}\wedge \alpha ^{(3:3)}),\end{aligned}$$
(335)
$$\begin{aligned} \xi _2&= (evi^2)^*\gamma ^{J_2} \wedge (evb^2)^*\alpha ^{(2:3)}. \end{aligned}$$
(336)

\(\delta _2\) is as before:

$$\begin{aligned} p_1^*\xi _1 \wedge p_2^*\xi = (-1)^{\delta _2}{\overline{\xi }}. \end{aligned}$$
(337)

Now compute

$$\begin{aligned} (evb_0|_{B_{\perp ,2}})_*\xi&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p_1)_*(p_1^*\xi _1 \wedge p_2^*\xi _2) \end{aligned}$$
(338)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(\xi _1 \wedge (p_1)_*(p_2)^*\xi _2) \end{aligned}$$
(339)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(\xi _1 \wedge (evb^1_{k_1+1})^*(evb_0^2)_* \xi _2) \end{aligned}$$
(340)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 + \delta _3} (evb^1_0)_*(\xi _1 \wedge \mathfrak {q}^{ST,\beta _2}_{k_2,l_2,\perp }(\alpha ^{(2:3)};\gamma ^{J_2})), \end{aligned}$$
(341)

where \(\delta _3 = \zeta (\alpha ^{(2:3)}) + \zeta _{\perp }(\alpha ^{(2:3)}; \gamma ^{J_2})\). Expanding \(\xi _1\) and rearranging gives:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 +\delta _3 + \delta _4} (evb^1_0)_*((evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*(\alpha ^{(1:3)} \wedge \mathfrak {q}^{ST,\beta _2}_{k_2,l_2,\perp }(\alpha ^{(2:3)};\gamma ^{J_2}) \wedge \alpha ^{(3:3)})), \end{aligned}$$
(342)

where \(\delta _4 = (|\gamma ^{J_2}| + \epsilon _2 + 1)(\epsilon _3 + k_3)\). By definition of the \(\mathfrak {q}^{ST}\) operations, this equals:

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 +\delta _3 + \delta _4 + \delta _5} \mathfrak {q}_{k_1+1+k_3,l_1,\perp }^{ST,\beta _1}(\alpha ^{(1:3)} \otimes \mathfrak {q}^{ST,\beta _2}_{k_2,l_2,\perp }(\alpha ^{(2:3)};\gamma ^{J_2}) \otimes \alpha ^{(3:3)}; \gamma ^{J_1}), \end{aligned}$$
(343)

where \(\delta _5 = \zeta (\alpha ^{(1:3)},\mathfrak {q}^{ST}_{k_2,l_2,\perp _0}(\alpha ^{(2:3)};\gamma ^{J_2}),\alpha ^{(3:3)})\). Adding all signs together with the sign in Stokes’ theorem 5.12, we get an overall sign:

$$\begin{aligned} C_2 = n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2) + |\gamma ^{J_2}|\epsilon _1 + 1. \end{aligned}$$
(344)

5.2.3 \(B_{\perp ,3}\)

We show:

Lemma 5.16

The terms in Stokes’ theorem coming from \((evb_0|_{B_{3,\perp }})_*\xi \) are

$$\begin{aligned} \sum _{\begin{array}{c} \beta _1 + \beta _2 = \beta \\ J_1 \cup J_2 = [l]\\ \gamma _1,\gamma _2 \in J_1 \end{array}} (-1)^{C_3} \mathfrak {q}^{ST}_{k,l_2}(\alpha ;\mathfrak {q}^{ST}_{\emptyset ,l_1}(\gamma ^{J_1}) \otimes \gamma ^{J_2}), \end{aligned}$$
(345)

where \(C_3 = n + |\gamma | +\zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2)\).

Recall that

$$\begin{aligned} {\overline{\phi }}_3: \mathcal {M}_{\emptyset ,l_1+1}(\beta _1) \times _X \mathcal {M}_{k+1,l_2+1}(\beta _2) \xrightarrow {\sim } B_{\perp ,3} \subset \partial \mathcal {M}_{k+1,l; \perp _0}(\beta ) \end{aligned}$$
(346)

changes orientation by \(\delta _1:= sign(\phi _3) = w_{\mathfrak {s}}(\beta _1)+1\).

Denote everything associated with \(\mathcal {M}_{\emptyset ,l_1+1}(\beta _1)\) with a subscript 1, and those for \(\mathcal {M}_{k+1,l_2+1}(\beta _2)\) with a subscript 2. Consider the commutative diagram:

(347)

Let

$$\begin{aligned} {\overline{\xi }}&= \phi _3^*\xi , \end{aligned}$$
(348)
$$\begin{aligned} \xi _1&= (evi^1)^*\gamma ^{J_1},\end{aligned}$$
(349)
$$\begin{aligned} \xi _2&= (evi^2)^*\gamma ^{J_2} \wedge (evb^2)^*\alpha . \end{aligned}$$
(350)

\(\delta _2\) is defined by:

$$\begin{aligned} p_1^*\xi _1 \wedge p_2^*\xi = (-1)^{\delta _2}{\overline{\xi }}, \end{aligned}$$
(351)

so that:

$$\begin{aligned} \delta _2 = sign^\gamma (J_1,J_2). \end{aligned}$$
(352)

Now compute

$$\begin{aligned} (evb_0|_{B_{\perp ,3}})_*\xi&= (-1)^{\delta _1+\delta _2} (evb^2_0)_*(p_2)_*(p_1^*\xi _1 \wedge p_2^*\xi _2) \end{aligned}$$
(353)
$$\begin{aligned}&\quad = (-1)^{\delta _1+\delta _2+\delta _3} (evb^2_0)_*((p_2)_*(p^1)^*\xi _1 \wedge \xi _2), \end{aligned}$$
(354)

where \(\delta _3 = reldim(p_2)|\xi _1| \equiv 0\; (mod \; 2)\). Using the commutative diagram, we find:

$$\begin{aligned}&\quad = (-1)^{\delta _1+\delta _2} (evb^2_0)_*((evi^2_{1})^*(evi_0^1)_*\xi _1 \wedge \xi _2) \end{aligned}$$
(355)
$$\begin{aligned}&\quad = (-1)^{\delta _1+\delta _2+\delta _4} (evb^2_0)_*((evi^2_{1})^*(\mathfrak {q}^{ST,\beta _1}_{\emptyset ,l_1}(\gamma ^{J_1}) \wedge \xi _2)), \end{aligned}$$
(356)

where \(\delta _4 = w_{\mathfrak {s}}(\beta _1)\). By definition this equals:

$$\begin{aligned}&\quad = (-1)^{\delta _1+\delta _2 + \delta _4+\delta _5} \mathfrak {q}^{ST}_{k,l_2}(\alpha ;\mathfrak {q}^{ST}_{\emptyset ,l_1}(\gamma ^{J_1}) \otimes \gamma ^{J_2}), \end{aligned}$$
(357)

where \(\delta _5 = \zeta (\alpha )\). Adding all signs together with the sign in Stokes’ theorem 5.12, we get an overall sign:

$$\begin{aligned} C_3 = n + |\gamma | +\zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2). \end{aligned}$$
(358)

5.2.4 \(B_{\perp ,4}\)

Lemma 5.17

The terms in Stokes’ theorem coming from \((evb_0|_{B_{4,\perp }})_*\xi \) are

$$\begin{aligned}{} & {} \sum _{\begin{array}{c} P \in S_5[k]\\ \beta _1 + \beta _2+\beta _3 = \beta \\ J_1 \cup J_2 \cup J_3 = [l]\\ 1 \in J_3, \;2 \in J_2 \end{array}} \mathfrak {q}_{k_1+1+k_3+k_5,l_1}^{ST,\beta _1}(\alpha ^{(1:5)} \otimes \mathfrak {q}_{k_2,l_2}^{ST,\beta _2}(\alpha ^{(2:5)};\gamma ^{J_2}) \otimes \alpha ^{(3:5)} \nonumber \\{} & {} \quad \otimes \mathfrak {q}_{k_4,l_4}^{ST,\beta _4}(\alpha ^{(4:5)};\gamma ^{J_3}) \otimes \alpha ^{(5:5)}; \gamma ^{J_1}), \end{aligned}$$
(359)

where \(C_4 = n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2,J_3) + (|\gamma ^{J_2}|+1)\epsilon _1 + (|\gamma ^{J_3}|+1)(\epsilon _1 + \epsilon _2 + \epsilon _3) + |\gamma ^{J_2}|\).

Recall that

$$\begin{aligned} \phi _4: \mathcal {M}_{k_4+1, l_3}(\beta _3) \times _L \mathcal {M}_{k_1+k_3+k_5+3,l_1}(\beta _1) \times _L \mathcal {M}_{k_2+1,l_2}(\beta _2) \rightarrow B_{\perp ,4} \subset \partial \mathcal {M}_{k+1,l,\perp _0}(\beta )\nonumber \\ \end{aligned}$$
(360)

changes orientation by \(\delta _1:= sign(\phi _4) = k_4(k_1 + k_2 + k_3) + k_2(k_3+k_5) + k_3\).

Denote everything associated with \(\mathcal {M}_{k_1+k_3+k_5+3,l_1}(\beta _1)\) with a subscript 1, everything associated with \(\mathcal {M}_{k_2+1,l_2}(\beta _2)\) with a subscript 2 and subscript 3 for \(\mathcal {M}_{k_4+1, l_3}(\beta _3)\). Consider the commutative diagram:

(361)

Let

$$\begin{aligned} {\overline{\xi }}&= \phi _4^*\xi , \end{aligned}$$
(362)
$$\begin{aligned} \xi _1&= (evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*(\alpha ^{(1:5)}\wedge \alpha ^{(3:5)}\wedge \alpha ^{(5:5)}),\end{aligned}$$
(363)
$$\begin{aligned} \xi _2&= (evi^2)^*\gamma ^{J_2} \wedge (evb^2)^*\alpha ^{(2:5)},\end{aligned}$$
(364)
$$\begin{aligned} \xi _3&= (evi^3)^*\gamma ^{J_3} \wedge (evb^3)^*\alpha ^{(4:5)}. \end{aligned}$$
(365)

\(\delta _2\) is defined by:

$$\begin{aligned} p_3^*\xi _3 \wedge p_1^*\xi _1 \wedge p_2^*\xi = (-1)^{\delta _2}{\overline{\xi }}, \end{aligned}$$
(366)

so that:

$$\begin{aligned} \delta _2= & {} sign^\gamma (J_1,J_2) + sign^\gamma (J_3,J_1\cup J_2) + |\gamma ^{J_1}|(k_4+\epsilon _4) \nonumber \\{} & {} \quad + |\gamma ^{J_2}|(k_5 +\epsilon _5 + k_4+\epsilon _4+k_3+\epsilon _3 + k_1 + \epsilon _1) \nonumber \\{} & {} \quad + (\epsilon _4 + k_4)(\epsilon _1+k_1+\epsilon _3+k_3) + (\epsilon _2+k_2)(k_5 +\epsilon _5 + k_4+\epsilon _4+k_3+\epsilon _3).\nonumber \\ \end{aligned}$$
(367)

Now compute

$$\begin{aligned}&(evb_0|_{B_{\perp ,4}})_*\xi = (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p_1)_*(p_3^*\xi _3 \wedge p_1^*\xi _1 \wedge p_2^*\xi _2) \end{aligned}$$
(368)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p^{13}_1)_*(p_{13})_*((p_{13})^*((p^{13}_3)^*\xi _3 \wedge (p^{13}_{1})^*\xi _1) \wedge p_2^*\xi _2) \end{aligned}$$
(369)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p^{13}_1)_*((p^{13}_3)^*\xi _3 \wedge (p^{13}_{1})^*\xi _1 \wedge (p_{13})_*p_2^*\xi _2) \end{aligned}$$
(370)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2} (evb^1_0)_*(p^{13}_1)_*((p^{13}_3)^*\xi _3 \wedge (p^{13}_{1})^*\xi _1 \wedge (p^{13}_{1})^*(evb^1_{k_1+1})^*(evb_0^2)_*\xi _2) \end{aligned}$$
(371)
$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 + \delta _3} (evb^1_0)_*((p^{13}_1)_*(p^{13}_3)^*\xi _3 \wedge \xi _1 \wedge (evb^1_{k_1+1})^*(evb_0^2)_*\xi _2), \end{aligned}$$
(372)

where \(\delta _3 = k_4(|\gamma ^{J_1}| + |\gamma ^{J_2}|+k_5 +\epsilon _5 +k_3+\epsilon _3 + k_1 + \epsilon _1+ \epsilon _2)\).

$$\begin{aligned}&= (-1)^{\delta _1+\delta _2 +\delta _3 + \delta _4} (evb^1_0)_*((evb^1_{k_1+k_3+k_5+2})^*(evb_0^3)_*\xi _3 \wedge \xi _1 \wedge (evb^1_{k_1+1})^*(evb_0^2)_*\xi _2)\nonumber \\ \end{aligned}$$
(373)

where \(\delta _4 = k_4(k_1+k_3+k_5)\)

$$\begin{aligned}&= (-1)^{\sum _{i=1}^5\delta _i} (evb^1_0)_*((evb^1_{k_1+k_3+k_5+2})^*(evb_0^3)_*\xi _3 \wedge \xi _1 \wedge (evb^1_{k_1+1})^*\mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2}))\nonumber \\ \end{aligned}$$
(374)

where \(\delta _5 = \zeta (\alpha ^{(2:5)})\)

$$\begin{aligned}&= (-1)^{\sum _{i=1}^6\delta _i} (evb^1_0)_*((evb^1_{k_1+k_3+k_5+2})^*\mathfrak {q}^{ST}_{k_4,l_4}(\alpha ^{(4:5)};\gamma ^{J_3}) \wedge \xi _1 \wedge (evb^1_{k_1+1})^*\nonumber \\&\quad \mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2})) \end{aligned}$$
(375)

where \(\delta _6 = \zeta (\alpha ^{(4:5)})\)

$$\begin{aligned}&= (-1)^{\sum _{i=1}^7\delta _i} (evb^1_0)_*((evi^1)^*\gamma ^{J_1} \wedge (evb^1)^*( \alpha ^{(1:5)} \wedge \mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2}) \wedge \alpha ^{(3:5)} \nonumber \\&\quad \wedge \mathfrak {q}^{ST}_{k_4,l_4}(\alpha ^{(4:5)};\gamma ^{J_3}) \wedge \alpha ^{(5:5)})) \end{aligned}$$
(376)

where \(\delta _7 = (|\gamma ^{J_1}|+\epsilon _1+k_1+\epsilon _3+k_3)(\epsilon _4 + |\gamma ^{J_3}|) + (|\gamma ^{J_2}|+\epsilon _2)(k_5 +\epsilon _5 +k_3+\epsilon _3+\epsilon _4 +|\gamma ^{J_3}|)\)

$$\begin{aligned}&=(-1)^{\sum _{i=1}^8\delta _i}\mathfrak {q}_{k_1+1+k_3+k_5,l_1}^{ST,\beta _1}(\alpha ^{(1:5)} \otimes \mathfrak {q}_{k_2,l_2}^{ST,\beta _2}(\alpha ^{(2:5)};\gamma ^{J_2}) \otimes \alpha ^{(3:5)} \nonumber \\&\quad \otimes \mathfrak {q}_{k_4,l_4}^{ST,\beta _4}(\alpha ^{(4:5)};\gamma ^{J_3}) \otimes \alpha ^{(5:5)}; \gamma ^{J_1}) \end{aligned}$$
(377)

where \(\delta _8 = \zeta (\alpha ^{(1:5)}, \mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2}), \alpha ^{(3:5)}, \mathfrak {q}^{ST}_{k_4,l_4}(\alpha ^{(4:5)};\gamma ^{J_3}), \alpha ^{(5:5)})\). Adding all signs together with the sign in Stokes’ theorem 5.12, we get an overall sign:

$$\begin{aligned} C_4= & {} n+ |\gamma | + \zeta (\alpha ) + \epsilon (\alpha ) + sign^\gamma (J_1,J_2,J_3) + (|\gamma ^{J_2}|+1)\epsilon _1\nonumber \\{} & {} \quad + (|\gamma ^{J_3}|+1)(\epsilon _1 + \epsilon _2 + \epsilon _3) + |\gamma ^{J_2}|. \end{aligned}$$
(378)

The above computation relies on the following lemma, which is similar to [47, Lemma 2.9]:

Lemma 5.18

Let \(P \in S_5[k]\), then

$$\begin{aligned}{} & {} \zeta (\alpha ^{(2:5)}) + \zeta (\alpha ^{(4:5)}) + \zeta (\alpha ^{(1:5)}, \mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2}), \alpha ^{(3:5)}, \mathfrak {q}^{ST}_{k_4,l_4}(\alpha ^{(4:5)};\gamma ^{J_3}), \alpha ^{(5:5)}) \nonumber \\{} & {} \quad \equiv \zeta (\alpha ) + \epsilon _2 + \epsilon _3 + k_3 +|\gamma ^{J_2}|(k_1 + 1) + |\gamma ^{J_3}|(k_1+k_3)\nonumber \\{} & {} \qquad + 1 + k_2(\epsilon _3 + \epsilon _4 + \epsilon _5) + k_4\epsilon _5 \;(mod \;2). \end{aligned}$$
(379)

Proof

All our computations here will be done modulo two. From the definition of \(\zeta \) we have:

$$\begin{aligned} \zeta (\alpha ^{(2:5)}) = 1 + \sum _{i = k_1 + 1}^{k_2}(i-k_1)|\alpha _i|' = 1 + k_1\epsilon _2 + \sum _{i = k_1 + 1}^{k_2} i|\alpha _i|'. \end{aligned}$$
(380)

Similarly we have:

$$\begin{aligned} \zeta (\alpha ^{(4:5)})= & {} 1 + \sum _{i = k_1+k_2+k_3 + 1}^{k_1+k_2+k_3+k_4}(i-(k_1+k_2+k_3))|\alpha _i|' \nonumber \\= & {} 1 +(k_1+k_2+k_3)\epsilon _4 + \sum _{i = k_1+k_2+k_3 + 1}^{k_1+k_2+k_3+k_4} i|\alpha _i|'. \end{aligned}$$
(381)

Finally:

$$\begin{aligned}{} & {} \zeta (\alpha ^{(1:5)}, \mathfrak {q}^{ST}_{k_2,l_2}(\alpha ^{(2:5)};\gamma ^{J_2}), \alpha ^{(3:5)}, \mathfrak {q}^{ST}_{k_4,l_4}(\alpha ^{(4:5)};\gamma ^{J_3}), \alpha ^{(5:5)}) \nonumber \\{} & {} \quad =1+ \sum _{i = 1}^{k_1}i|\alpha _i|' + (k_1 + 1)(\epsilon _2 + |\gamma ^{J_2}| + 1) \nonumber \\{} & {} \qquad + \sum _{i = k_1+k_2+1}^{k_1+k_2+k_3}(i+1 -k_2)|\alpha _i|' + (k_1+k_3+2)(\epsilon _4 + |\gamma ^{J_3}| + 1)\nonumber \\{} & {} \qquad + \sum _{i = k_1+k_2+k_3+k_4+1}^{k}(i+2 - k_2 - k_4)|\alpha _i|' \nonumber \\{} & {} \quad = 1+ \sum _{i = 1}^{k_1}i|\alpha _i|' + \sum _{i = k_1+k_2+1}^{k_1+k_2+k_3}i|\alpha _i|' + \sum _{i = k_1+k_2+k_3+k_4+1}^{k}i|\alpha _i|'+ (k_1 + 1)\nonumber \\{} & {} \qquad (\epsilon _2 + |\gamma ^{J_2}| + 1)+ (k_2+1) \epsilon _3 + (k_1+k_3)(\epsilon _4 + |\gamma ^{J_3}| + 1) + (k_2+k_4)\epsilon _5.\nonumber \\ \end{aligned}$$
(382)

Adding up these three terms gives the result. \(\square \)

5.3 Concluding the proof of the structure equation

The last step of the proof of Proposition 5.1 is to combine the terms coming from \(B_{\perp ,1}\) with term (311) coming from \((evb_0)_*(d\xi )\). Because the disk bubbles in \(B_{\perp ,1}\) must be stable, there are no disks contributing with \(\beta _2 =\beta _0\). These contributions are exactly provided by the term \((evb_0)_*(d\xi )\) in Stokes’ theorem. The same holds for \(B_{\perp ,2}\) and the terms coming from \(d(evb_0)_*\xi \). We then sum all the terms together with those from \(B_{\perp ,3}\) and \(B_{\perp ,4}\). Finally multiply by \((-1)^{n+|\gamma | + \zeta (\alpha ) + \epsilon (\alpha )}\) to get Proposition 5.1.

To prove Proposition 4.31, one easily adapts the signs. The main computation one needs is the following, which can be directly verified:

Lemma 5.19

We have:

$$\begin{aligned}{} & {} \dagger \left( \alpha ^{(1:5)},\mathfrak {q}^{ST}_{k_2,j_2}(\alpha ^{(2:5)};\gamma ^{J_2}), \alpha ^{(3:5)}, \mathfrak {q}^{ST}_{k_4,j_3}(\alpha ^{(4:5)};\gamma ^{J_3}), \alpha ^{(5:5)};\gamma ^{J_1}\right) + \dagger (\alpha ^{(2:5)}) \nonumber \\{} & {} \qquad + \dagger (\alpha ^{(4:5)}) \nonumber \\{} & {} \quad =\dagger ({\alpha }) + 1 + \epsilon _2 + \epsilon _4 + |\gamma ^{J_2}|(\epsilon _1 + \epsilon _3 + \epsilon _4 + \epsilon _5 +1) \nonumber \\{} & {} \qquad + |\gamma ^{J_2}|(\epsilon _1 + \epsilon _2 + \epsilon _3 + \epsilon _5 +1 + |\gamma ^{J_2}|). \end{aligned}$$
(383)

5.4 Unit on the horocycle

This proof follows that of [44, Lemma 3.10] very closely. Let \(p_i: \mathcal {M}_{k+2,l; \perp _i}(\beta ) \rightarrow \mathcal {M}_{k+1,l}(\beta )\) be the map given by forgetting the boundary marked point \(z_i\) and the horocyclic constraint. This map is injective, as there is a unique horocycle through any two interior points \(w_1\) and \(w_2\). Fixing the interior marked point \(w_1\) and the boundary marked points \(z_j\) for \(j \ne i\), we see that as \(z_i\) moves between the two adjacent boundary marked points \(z_j\) and \(z_{j+1}\), the point \(w_2\) sweeps out a lunar arc between the two horocycles through \(w_1\) and \(z_j\), and through \(w_1\) and \(z_{j+1}\) (see Fig. 1). As these lunar arcs cover the entire unit disk we see that the image of

$$\begin{aligned} p:= \sqcup _{i=1}^{k+1} p_i: \sqcup _{i=1}^{k+1} \mathcal {M}_{k+2,l; \perp _i}(\beta ) \rightarrow \mathcal {M}_{k+1,l}(\beta ) \end{aligned}$$
(384)

is an open dense subset.

Fig. 1
figure 1

the coloured regions show the areas swept out by \(w_2\)

We next compute the sign of \(p_i\):

Lemma 5.20

The map \(p_i: \mathcal {M}_{k+2,l; \perp _i}(\beta ) \rightarrow \mathcal {M}_{k+1,l}(\beta )\) is a local diffeomorphism with \(sign(p_i) = n + i\).

Proof

The orientation on \({\mathcal {M}}_{k+2,l,\perp _i}(\beta ) = I \times _{D^2} {\mathcal {M}}_{k+2,l}(\beta )\) is defined by the fibre-product orientation (see [46, Section 2.2]). At the level of tangent spaces, this means that we have a natural identification

$$\begin{aligned} T_t I \oplus T_u {\mathcal {M}}_{k+2,l}(\beta ) \cong T_t D^2 \oplus T_{(t,u)}(I \times _{D^2} {\mathcal {M}}_{k+2,l}(\beta )). \end{aligned}$$
(385)

To calculate the sign of \(p_i\), we multiply both sides by a factor of \(D^2\) and \(Aut(D^2)\). This does not change the sign. Let \({\mathbb {R}}_i\) denote the tangent space to the factor \(S^1_i\) in the definition of the orientation on \({\mathcal {M}}_{k+2,l}(\beta )\) (see Eq. 273). Thus, \(sign(p_i)\) is equal to the sign of the map:

$$\begin{aligned}{} & {} T_t I \oplus T_u \widetilde{{\mathcal {M}}}^{main}(\beta ) \oplus {\mathbb {R}}_0 \oplus \dots \oplus {\mathbb {R}}_{k+2} \oplus {\mathbb {C}}^l \oplus T_z Aut(D^2) \nonumber \\{} & {} \quad \rightarrow T_t D^2 \oplus T_u \widetilde{{\mathcal {M}}}^{main}(\beta ) \oplus {\mathbb {R}}_0 \oplus \dots \oplus {\mathbb {R}}_{i-1} \oplus {\mathbb {R}}_{i+1} \oplus \dots \nonumber \\{} & {} \quad \oplus {\mathbb {R}}_{k+2} \oplus {\mathbb {C}}^l \oplus T_z Aut(D^2). \end{aligned}$$
(386)

Bringing the summand \({\mathbb {R}}_i\) to the second position, which gives a sign of \((-1)^{i+n}\), and cancelling the obvious summands, we find that \(sign(p_i)\) is given by the sign of:

$$\begin{aligned} T_t I \oplus {\mathbb {R}}_i \rightarrow (-1)^{n + i} T_t D^2. \end{aligned}$$
(387)

By drawing a local picture, it is clear that the natural map \(T_t I \oplus {\mathbb {R}}_i \rightarrow T_t D^2\) is orientation preserving. We have thus shown that \(sign(p_i) = i + n\). \(\square \)

Proof of Proposition 4.32

Let \(evi_j\) and \(evb_j\) denote the evaluation maps for the space \({\mathcal {M}}_{k+1,l}(\beta )\). Use \(evi^i_j\) and \(evb^i_j\) for those of \({\mathcal {M}}_{k+2,l,\perp _i}\). The map \(p_i\) then satisfies \(evi_j \circ p_i = evi^i_i\) and:

$$\begin{aligned} evb_j \circ p_i = {\left\{ \begin{array}{ll} evb^i_j &{} \text {if }\, j<i\\ evb^i_{j+1} &{} \text {if}\, j > i \end{array}\right. } \end{aligned}$$
(388)

Let \(\xi = \bigwedge _{j=1}^l (evi_j)^*\gamma _j \wedge \bigwedge _{j=0}^{k} (evb_j)^* {\alpha }_j\) and \(\xi _i = p^*_i \xi \). Then as the images of \(p_i\) for \(1\le i \le k+1\) form an open dense set in \({\mathcal {M}}_{k+1,l}(\beta )\), we have:

$$\begin{aligned} pt_*\xi = \sum _{i=1}^{k+1} (-1)^{sign(p_i)} pt_* \xi _i. \end{aligned}$$
(389)

A short sign computation shows that

$$\begin{aligned} pt_* \left( (evb_0)_*\xi \wedge {\alpha }_0 \right) = (-1)^{\zeta (\widetilde{{\alpha }}) + |{\alpha }_0|(\epsilon (\widetilde{{\alpha }}) + 1)}\langle {\mathfrak {q}}^{ST,\beta }_{k,l}({\alpha }_1, \dots , {\alpha }_k; \gamma ), {\alpha }_0 \rangle _L.\nonumber \\ \end{aligned}$$
(390)

Take the partition with \((1:2) = (1, \dots , i-1)\). We then compute:

$$\begin{aligned} pt_* \xi _i&= (-1)^{\zeta (\widetilde{{\alpha }}) + |{\alpha }_0|(\epsilon (\widetilde{{\alpha }}) + 1) + 1 + |{\gamma }| + n + \epsilon _2 + i}\langle {\mathfrak {q}}^{ST,\beta }_{k+1,l,\perp _i}({\alpha }^{(1:2)},1, {\alpha }^{(2:2)}); \gamma ), {\alpha }_0 \rangle _L. \end{aligned}$$
(391)

Combining this, we have shown:

$$\begin{aligned} \langle \mathfrak {q}^{ST}_{k,l}(\widetilde{\alpha };\gamma ),\alpha _0 \rangle _L = \sum _{P \in S_2[k]} (-1)^{\epsilon _1+|\gamma |+1} \langle \mathfrak {q}^{ST}_{k+1,l,\perp _i}(\alpha ^{(1:2)},1,\alpha ^{(2:2)};\gamma ), \alpha _0 \rangle _L.\nonumber \\ \end{aligned}$$
(392)

Taking into account the sign changes between the operations \({\mathfrak {q}}^ST\) and \({\mathfrak {q}}\), the first result of Proposition 4.32 follows. The second equality follows from a cyclic symmetry property of the \(\mathfrak {q}_\perp \) operations, which is the direct analogue of Proposition 4.23.

6 Applications of Conjecture 1.6 in the monotone setting

Let \((X,\omega )\) be a 2n dimensional monotone symplectic manifold, i.e \(c_1 = c_1(TX) = \tau [\omega ] \in H^2(X;\mathbb {R})\) for some \(\tau \in \mathbb {R}_{>0}\). We can then consider quantum cohomology \(QH^*(X) = H^*(X;\mathbb {C})\) with coefficients in \(\mathbb {C}\), as for example in [41].

Definition 6.1

(Quantum E-structure) The quantum E-structure is the \(\mathbb {C}[[u]]\)-module \(QH^*(X)[[u]]\) with connection:

$$\begin{aligned} \nabla _{\frac{d}{du}} = \frac{d}{d u} + \frac{\mu }{u} + \frac{c_1 \star }{u^2}. \end{aligned}$$
(393)

Here \(\star \) is the quantum cup product, and \(\mu : QH^*(X) \rightarrow QH^*(X)\) is the grading operator with \(\mu (\alpha ) = \frac{p-n}{2}\alpha \) for \(\alpha \in QH^p(X)\). The residue of the connection is given by \(c_1 \star \). The canonical map \(\pi : QH^*(X)[[u]]\rightarrow QH^*(X)\) is given by evaluation at \(u=0\). We will often write \(\nabla \) for \(\nabla _{\frac{d}{du}}\).

Definition 6.2

Quantum cohomology admits a canonically defined splitting, given by:

$$\begin{aligned} s^{GW}: QH^*(X)&\rightarrow QH^*(X)[[u]]\nonumber \\ \alpha&\mapsto \alpha . \end{aligned}$$
(394)

As this is the splitting relevant for Gromov–Witten theory (see e.g. [18]), we call this the Gromov–Witten splitting.

Decompose quantum cohomology as a direct sum of generalised eigenspaces of \(c_1 \star \)

$$\begin{aligned} QH^*(X) = \bigoplus _w QH^*(X)_w. \end{aligned}$$
(395)

Proposition 2.12 shows that we can extend this decomposition by eigenvalues of \(c_1 \star \) to the E-structure \(QH^*(X)[[u]]\).

Lemma 6.3

There exists a unique decomposition of \(QH^*(X)[[u]]\):

$$\begin{aligned} QH^*(X)[[u]]= \bigoplus _w QH^*(X)[[u]]_w, \end{aligned}$$
(396)

which is compatible with the connection:

$$\begin{aligned} u^2\nabla : QH^*(X)[[u]]_w \rightarrow QH^*(X)[[u]]_w, \end{aligned}$$
(397)

and respects the eigenvalue decomposition of \(QH^*(X)\):

$$\begin{aligned} \pi (QH^*(X)[[u]]_w) = QH^*(X)_w. \end{aligned}$$
(398)

Remark 6.4

Note that in general \(QH^*(X)[[u]]_w \ne QH^*(X)_w[[u]]\). This is because the map \(u^2\nabla \) does not map \(QH^*(X)_w[[u]]\) to itself, as in general \(\mu \) will not, see Example 6.10. Similarly, the Gromov–Witten splitting does not respect the eigenvalue decomposition. So in general \(s^{GW}(QH^*(X)_w) \not \subset QH^*(X)[[u]]_w\). See Example 6.10.

Lemma 2.14 together with Corollary 3.29 then shows:

Corollary 6.5

Let X be a monotone symplectic manifold. Assume Conjecture 1.6 holds. Then the cyclic open–closed map respects the decompositions, that is:

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^{-}(HC^-_*(Fuk(X)_w)) \subset QH^*(X)[[u]]_w. \end{aligned}$$
(399)

6.1 Semi-simple quantum cohomology

Now assume additionally that the quantum cohomology ring \(QH^*(X)\) is semi-simple. This is the case for example for \(X = \mathbb{C}\mathbb{P}^n\). The main result is:

Theorem 6.6

Let X be a monotone symplectic manifold with semi-simple quantum cohomology. Then the EP-structure \(QH^*(X)[[u]]\) is semi-simple in the sense of Definition 2.17.

Remark 6.7

This theorem is a rephrasing of results in [6, 9].

In order to prove this theorem, we need the following lemma. This was first observed by Dubrovin [6], but see also [48, Remark 8.2 (iii)] or [9, Lemma 2.4.4].

Lemma 6.8

([6, Lemma 3.2]) When \(QH^*(X)\) is semi-simple, \(0 = \pi _w \circ \mu |_{QH_w^*(X)}: QH_w^*(X) \rightarrow QH_w^*(X)\). I.e. the diagonal blocks for a matrix representation of \(\mu \) vanish, when we choose a basis of eigenvectors for \(c_1 \star \).

Proof of theorem 6.6

By Proposition 2.12, after changing to a basis for the \({\mathbb {C}}[[u]]\)-modules \(QH^*(X)[[u]]_w\) the connection is given by

$$\begin{aligned} \nabla = \frac{d}{du} + u^{-2}c_1 \star + u^{-1}B_{-1} + \sum _{i\ge 0} B_i u^i. \end{aligned}$$
(400)

We also have that \(B_{-1}|_{QH^*(X)_w} = \pi _w \circ \mu |_{QH^*(X)_w} = 0\) by Lemma 6.8. As \(B_{-1}\) also respects the decomposition of \(QH^*(X)\) into eigenspaces for \(c_1 \star \), we find that \(B_{-1} = 0\). The connection on \(QH^*(X)[[u]]_w\) is thus given by:

$$\begin{aligned} \nabla = \frac{d}{du} + \frac{w}{u^2} + \sum _{i\ge 0} u^i B_i. \end{aligned}$$
(401)

The connection \(\widetilde{\nabla }^w\) on \(QH^*(X)[[u]]_w \otimes \mathcal {E}^{\frac{w}{u}}\) is then given by:

$$\begin{aligned} \widetilde{\nabla }^w = \frac{d}{du} + \sum _{i\ge 0} u^i B_i. \end{aligned}$$
(402)

As in [49, Chapter 2, case 1], there now exists a further change of basis so that the connection matrix on \(QH^*(X)[[u]]_w \otimes \mathcal {E}^{\frac{w}{u}}\) vanishes. Equivalently, there exists an isomorphism \(QH^*(X)[[u]]_w \cong \mathcal {E}^{-\frac{w}{u}}\). \(\square \)

We thus obtain the following result, which is a part of [9, Proposition 2.5.1].

Corollary 6.9

When \(QH^*(X)\) is semi-simple, there exists a basis \(v_i \in QH^*(X)[[u]]\) such that \(u^2\nabla _{\frac{d}{du}}v_i = w_iv_i\), where the \(w_i\) are the eigenvalues of \(c_1 \star \).

We call the basis \(\{v_{i}\}\) a w-flat basis. As in Example 2.21, we obtain a splitting:

$$\begin{aligned} s^{ss}: QH^*(X)&\rightarrow QH^*(X)[[u]]\nonumber \\ \pi (v_{i_j})&\rightarrow v_{i_j}. \end{aligned}$$
(403)

There then exists a unique transformation \(R = \sum _{i \ge 0} u^iR_i\), where \(R_i: QH^*(X) \rightarrow QH^*(X)\) such that

$$\begin{aligned} s^{ss}(\alpha ) = \sum _{i \ge 0} u^i s^{GW}(R_i(\alpha )). \end{aligned}$$
(404)

We will now show R agrees with the R-matrix as defined by Teleman  [48]. Start with a basis for \(QH^*(X)\) consisting of eigenvectors \(\{e_{i}\}\) for \(c_1 \star \). Thus in this basis for \(QH^*(X)[[u]]\) the connection reads:

$$\begin{aligned} \nabla = \frac{d}{du} + u^{-2} c_1 \star + u^{-1}\mu . \end{aligned}$$
(405)

The R-matrix then changes bases to \(v_{i} = R(e_{i})\), in which the connection can be expanded as

$$\begin{aligned} R^*\nabla = \frac{d}{du} + u^{-2} c_1 \star . \end{aligned}$$
(406)

We thus obtain the recursive relation for R:

$$\begin{aligned}{}[c_1 \star , R_{i+1}] + (\mu + i) R_i = 0, \end{aligned}$$
(407)

which agrees with the relation in [48, Proposition 8.5].

Example 6.10

Consider a basis 1, H for \(S^2\), where H is the point class. Then \(c_1 = 2 H\). The quantum multiplication table reads

$$\begin{aligned} 1 \star 1 = 1, \; 1 \star H = H, \; H\star H = 1. \end{aligned}$$
(408)

Thus \(QH^*(S^2) \cong \frac{\mathbb {C}[H]}{\langle H^2 - 1 \rangle }\), which is semi-simple. The decomposition into eigenspaces for \(c_1\) is:

$$\begin{aligned} QH^*(S^2) = QH^*(S^2)_{-2} \oplus QH^*(S^2)_2 = \langle H-1 \rangle \oplus \langle 1+H \rangle := \langle v \rangle \oplus \langle w \rangle .\nonumber \\ \end{aligned}$$
(409)

Now \(\mu (v) = \frac{w}{2}\) and \(\mu (w) = \frac{v}{2}\). Thus the \(\mathbb {C}[[u]]\)-modules \(\langle 1+H \rangle _{\mathbb {C}[[u]]}\) and \(\langle 1-H \rangle _{\mathbb {C}[[u]]}\) are not invariant under \(u^2\nabla \). In fact, we can solve the differential equation \(u^2 \nabla _{\frac{d}{du}} \widetilde{w} = 2 \widetilde{w}\) directly, and find

$$\begin{aligned} \widetilde{w} = \sum _{n\ge 0} u^n(\alpha _n w + \beta _n v), \end{aligned}$$
(410)

where

$$\begin{aligned} \alpha _{n+1} = \frac{4^{-2(n+1)}}{(n+1)!}\prod _{j=0}^{n} (4j^2 - 1) \; \text {and} \; \beta _n = -2n\alpha _n. \end{aligned}$$
(411)

Similarly we obtain a solution to \(u^2 \nabla _{\frac{d}{du}} \widetilde{v} = 2 \widetilde{v}\):

$$\begin{aligned} \widetilde{v} = \sum _{n\ge 0} u^n({\gamma }_n w + \delta _n v), \end{aligned}$$
(412)

where

$$\begin{aligned} {\gamma }_{n} = -2n\delta _{n} \; \text {and} \; \delta _{n+1} = \frac{(-1)^{n+1}4^{-2(n+1)}}{(n+1)!}\prod _{j=0}^{n} (4j^2 - 1). \end{aligned}$$
(413)

The R-matrix in the basis (vw) is thus:

$$\begin{aligned} R_{n+1} = \frac{4^{-2(n+1)}}{(n+1)!}\prod _{j=0}^{n} (4j^2 - 1) \begin{pmatrix} (-1)^{n+1}&{}(-1)^{n}2(n+1) \\ -2(n+1)&{}1 \\ \end{pmatrix}, \end{aligned}$$
(414)

which indeed agrees with the R-matrix computed in [1, Example 5.4] for the cyclic homology of \(Fuk(S^2)\).

We now rephrase [1, Theorem 5.9] and provide an alternative proof. The proof in [1] uses the closed–open map and the Dubrovin–Teleman reconstruction theorem [48]. Our proof instead uses the cyclic open–closed map and assumes Conjecture 1.6. In particular, it does not rely on the Dubrovin-Teleman reconstruction theorem.

Theorem 6.11

Let X be a symplectic manifold such that \({\mathcal{O}\mathcal{C}}: HH_*(Fuk(X)) \rightarrow QH^*(X)\) is an isomorphism and \(HH^*(Fuk(X))\) is semi-simple. Let \(\mu ^{{\mathcal{O}\mathcal{C}}} = {\mathcal{O}\mathcal{C}}^{-1} \circ \mu \circ {\mathcal{O}\mathcal{C}}: HH_*(Fuk(X)) \rightarrow HH_*(Fuk(X))\) be the pull-back of the grading operator \(\mu \) on \(QH^*(X)\). Then the Frobenius manifold \({\mathcal {M}}_{\mu ^{{\mathcal{O}\mathcal{C}}}}\) associated to \(\mu ^{{\mathcal{O}\mathcal{C}}}\) (see Sect. 2.2) is isomorphic to the big quantum cohomology of X. Here Fuk(X) denotes a non-bulk deformed Fukaya category defined over \(\Lambda \) (or \({\mathbb {C}}\) in the monotone case).

Proof

As the open–closed map is an isomorphism, so is the closed–open map \({\mathcal{C}\mathcal{O}}: QH^*(X) \rightarrow HH^*(Fuk(X))\). Thus, \(QH^*(X)\) is also semi-simple. Lemma 6.6 then shows \(QH^*(X)[[u]]\) is a semi-simple EP-structure. Let \(R = {\mathbb {C}}[[H^*(X)]]\) parametrise bulk-deformations. Let \(QH^*(X;R)[[u]]\) denote the quantum TEP-structure over R.

Let \(Fuk^t(X)\) denote the bulk-deformed Fukaya category. As \({\mathcal{C}\mathcal{O}}\) is an isomorphism, this is a versal deformation of Fuk(X), and can thus be extracted from the categorical data of Fuk(X).

Now apply the bijection between grading operators and primitive forms (Corollary 2.28) to the TEP-structure \(QH^*(X;R)[[u]]\). It follows directly from the definition of the product and metric in [39], that the Frobenius manifold associated to the grading operator \(\mu \) and the primitive \(\omega = 1 \in QH^*(X)\) is indeed the big quantum cohomology ring \(QH^*(X;R)\).

Amorim and Tu [1, Corollary 3.8] show that as \(HH^*(Fuk(X))\) is semi-simple, \(HC^-(Fuk^t(X))\) is a semi-simple TEP-structure. The grading operator \(\mu ^{{\mathcal{O}\mathcal{C}}}\) on \(HH_*(Fuk(X))\) is pulled back from the grading operator on \(QH^*(X)\). The primitive element \(\omega \in HH_*(Fuk(X))\) is defined as \({\mathcal{O}\mathcal{C}}^{-1}(1)\).

Now consider the bulk deformed cyclic open–closed map

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-: HC^-_*(Fuk^t(X)) \rightarrow QH^*(X;R)[[u]]. \end{aligned}$$
(415)

By conjecture 1.6, this is an isomorphism of TEP-structures. Furthermore, the cyclic open–closed map:

$$\begin{aligned} {\mathcal{O}\mathcal{C}}^-: HC^-(Fuk(X)) \rightarrow QH^*(X)[[u]]\end{aligned}$$
(416)

respects the grading operator and the primitive element \(\omega \). Thus, Corollary 2.28 equips the TEP-structures \(HC^-(Fuk^t(X))\) and \(QH^*(X;R)[[u]]\) with the same primitive form under the cyclic open–closed map. Hence, the associated Frobenius manifolds \({\mathcal {M}}_{\mu ^{{\mathcal{O}\mathcal{C}}}}\) and \(QH^*(X;R)\) are isomorphic. \(\square \)

6.2 Example: intersection of quadrics

We will now give an example where, even though the quantum cohomology is not semi-simple, it is still possible to construct an R-matrix. Let X be a complete intersection of two quadric hypersurfaces in \(\mathbb{C}\mathbb{P}^5\), which is a monotone symplectic manifold. The eigenvalue decomposition of the Fukuya category is as follows:

$$\begin{aligned} Fuk(X) = Fuk(X)_{-8} \oplus Fuk(X)_0 \oplus Fuk(X)_8. \end{aligned}$$
(417)

Smith proves an equivalence:

Theorem 6.12

([43, Theorem 1.1]) \(D^\pi Fuk(X)_0 \cong D^\pi Fuk(\Sigma _2)\), for \(\Sigma _2\) a genus 2 surface.

Assume that \(Fuk(X)_{\pm 8} \cong Fuk(pt)\), which [43, Section 1.6] expects. And note that the Fuk(pt) are considered here with curvature \(\pm 8\), so that \(HC^-_*(Fuk(X)_{\pm 4}) \cong \mathcal {E}^{\mp 8/u}\). Also note that [42, Chapter 4] proves a natural isomorphism \(HC_*^{-}(D^\pi \mathcal {C}) \cong HC_*^{-}(\mathcal {C})\). We thus have an isomorphism of TE-structures:

$$\begin{aligned} HC_*^{-}(Fuk(X)) \cong \mathcal {E}^{8/u} \oplus HC_*^{-}(Fuk(\Sigma _2)) \oplus \mathcal {E}^{-8/u}. \end{aligned}$$
(418)

As the TEP-structure associated to an \(A_\infty \)-category is canonical, Conjecture 1.6 implies:

Lemma 6.13

There exists an isomorphism of TEP-structures:

$$\begin{aligned} \Phi : QH^*(X)[[u]]\cong \mathcal {E}^{\frac{-8}{u}} \oplus QH^*(\Sigma _2)[[u]]\oplus \mathcal {E}^{\frac{8}{u}}. \end{aligned}$$
(419)

We prove Lemma 6.13 in Appendix B by a direct computation, providing evidence for the conjecture. Moreover, we show that this isomorphism is unique.