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

Categorical enumerative invariants of the ground field

Junwu Tu111Partially supported by the NSFC grant 12071290, and STCSM grant 20JC1413300. Junwu Tu, Institute of Mathematical Sciences, ShanghaiTech University, Shanghai, 201210, China.
Abstract

Abstract: For an S1S^{1}-framed modular operad PP, we introduce its “Feynman compactification” denoted by FPFP which is a modular operad. Let {𝕄𝖿𝗋(g,n)}(g,n)\{\mathbb{M}^{\sf fr}(g,n)\}_{(g,n)} be the S1S^{1}-framed modular operad defined using moduli spaces of smooth curves with framings along punctures. We prove that the homology operad of F𝕄𝖿𝗋F\mathbb{M}^{\sf fr} is isomorphic to H(M¯)H_{*}(\overline{M}), the homology operad of the Deligne-Mumford operad. Using this isomorphism, we obtain an explicit formula of the fundamental class of [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}] in terms of Sen-Zwiebach’s string vertices. As an immediate application, we prove Costello’s categorical enumerative invariants of the ground field match with the Gromov-Witten invariants of a point.

1 Introduction

CEI of the algebra \mathbb{Q}.

Categorical enumerative invariants (CEI) were introduced by Costello [5] and recently made explicit in [2] and [3]. These are invariants associated with a pair (A,s)(A,s) where

  • AA is a cyclic AA_{\infty} algebra of dimension dd over a field 𝕂\mathbb{K} (of characteristic zero) that is proper, smooth and satisfies the Hodge-to-de-Rham degeneration.

  • ss is a choice of splitting of the non-commutative Hodge filtration of AA.

Explicitly, CEI take the following form

α1,,αng,nA,s𝕂,(g,n),  2g2+n>0\langle\alpha_{1},\cdots,\alpha_{n}\rangle_{g,n}^{A,s}\in\mathbb{K},\;\;\;\;\forall(g,n),\;\;2g-2+n>0

with α1,,αnHHd(A)[u]\alpha_{1},\ldots,\alpha_{n}\in HH_{*-d}(A)[u] in the dd-shifted Hochschild homology of AA adjoining a formal variable uu of homological degree 2-2. It was proved in [3] that CEI is only possibly non-zero if

(1) j=1n(|αj|)=2(g1)(3d)+2n\sum_{j=1}^{n}(-|\alpha_{j}|)=2(g-1)(3-d)+2n

In the cohomological degree convention, this is precisely the dimension axiom in Gromov-Witten theory. Indeed, the original motivation to define and study such invariants came from mirror symmetry following Kontsevich’s proposal [14]. It is expected that these invariants, when applied to Fukaya categories, should recover Gromov-Witten invariants at all genera.

At first glance, comparing these categorical invariants with Gromov-Witten invariants may look rather difficult since CEI are defined purely algebraically while the GW invariants involves moduli spaces of stable maps. The purpose of this paper is to demonstrate otherwise. We begin with the case when the ground field 𝕂=\mathbb{K}=\mathbb{Q} and the algebra AA is also \mathbb{Q}. In this case, the dimension dd is zero, and the algebra AA has a unique homogeneous splitting of the Hodge filtration which we still denote by ss. Furthermore, its Hochschild homology group is HH()=HH_{*}(\mathbb{Q})=\mathbb{Q} and so the insertions lie in the space HH()[u]=[u]HH_{*}(\mathbb{Q})[u]=\mathbb{Q}[u].

Theorem A. The categorical enumerative invariants of the algebra \mathbb{Q} (the field of rationals) matches the Gromov-Witten invariants of a point, i.e. we have

uk1,,ukng,n,s=[M¯g,n]ψ1k1ψnkn\langle u^{k_{1}},\ldots,u^{k_{n}}\rangle^{\mathbb{Q},s}_{g,n}=\int_{[\overline{M}_{g,n}]}\psi_{1}^{k_{1}}\cdots\psi_{n}^{k_{n}}

where ψj\psi_{j}’s are the ψ\psi-classes on the Deligne-Mumford moduli space M¯g,n\overline{M}_{g,n}. By the dimension formula (1), these invariants are only non-zero when k1++kn=3g3+nk_{1}+\cdots+k_{n}=3g-3+n.

Basic constructions and notations.

In order to prove the above theorem, it is necessary to have a better understanding of Sen-Zwiebach’s string vertices [18], since they play a central role in the construction of CEI.

We need to introduce some notations. For a topological space XX, denote by C(X)C_{*}(X) its normalized singular chain complex with coefficients in \mathbb{Q}. For each pair (g,n)×(g,n)\in\mathbb{N}\times\mathbb{N} such that 2g2+n>02g-2+n>0, denote by Mg,n𝖿𝗋M_{g,n}^{\sf fr} the moduli space of tuples (Σ,p1,,pn,φ1,,φn)(\Sigma,p_{1},\ldots,p_{n},\varphi_{1},\ldots,\varphi_{n}) where Σ\Sigma is a Riemann surface of genus gg; p1,,pnp_{1},\ldots,p_{n} are nn marked points in Σ\Sigma; and for each 1in1\leq i\leq n, φi:D2U(pi)Σ\varphi_{i}:D^{2}\rightarrow U(p_{i})\subset\Sigma is a holomorphic chart such that it extends to an open neighborhood of the unit disk in \mathbb{C}. We also require that the pair-wise intersections of the closures of the framed disks be empty, i.e. U(pi)¯U(pj)¯=,1i<jn\overline{U(p_{i})}\cap\overline{U(p_{j})}=\emptyset,\;\;\forall 1\leq i<j\leq n. As in Segal [17], sewing along the coordinate charts defines two types of composition maps:

Mg,n𝖿𝗋×Mg,n𝖿𝗋Mg+g,n+n2𝖿𝗋,   1in,1jn\displaystyle M_{g,n}^{\sf fr}\times M_{g^{\prime},n^{\prime}}^{\sf fr}\rightarrow M_{g+g^{\prime},n+n^{\prime}-2}^{\sf fr},\;\;\;1\leq i\leq n,1\leq j\leq n^{\prime}
Mg,n+2𝖿𝗋Mg+1,n𝖿𝗋,   1i<jn+2\displaystyle M_{g,n+2}^{\sf fr}\rightarrow M^{\sf fr}_{g+1,n},\;\;\;1\leq i<j\leq n+2

We shall denote both of them by cijc_{ij}, as no confusion can arise in this way. The collection of topological spaces {Mg,n𝖿𝗋}(g,n)\{M_{g,n}^{\sf fr}\}_{(g,n)} together with the two sewing operations form a modular operad defined by Getzler-Kapranov [10] in the category of topological spaces.

Applying the normalized singular chain functor CC_{*} yields a differential graded modular operad {C(Mg,n𝖿𝗋)}(g,n)\{C_{*}(M_{g,n}^{\sf fr})\}_{(g,n)}. For the first type composition, note that the functor CC_{*} is lax monoidal, i.e. for two topological spaces XX and YY, we have a naturally defined map C(X)C(Y)C(X×Y)C_{*}(X)\otimes C_{*}(Y)\rightarrow C_{*}(X\times Y), the Alexander-Whitney map. This enables us to define a composition map still denoted by

cij:C(Mg,n𝖿𝗋)C(Mg,n𝖿𝗋)𝖠𝖶C(Mg,n𝖿𝗋×Mg,n𝖿𝗋)C(Mg+g,n+n2𝖿𝗋).c_{ij}:C_{*}(M_{g,n}^{\sf fr})\otimes C_{*}(M_{g^{\prime},n^{\prime}}^{\sf fr})\stackrel{{\scriptstyle{\sf AW}}}{{\longrightarrow}}C_{*}(M_{g,n}^{\sf fr}\times M_{g^{\prime},n^{\prime}}^{\sf fr})\rightarrow C_{*}(M_{g+g^{\prime},n+n^{\prime}-2}^{\sf fr}).

Denote this differential graded modular operad by 𝕄𝖿𝗋\mathbb{M}^{\sf fr}, with 𝕄𝖿𝗋(g,n):=C(Mg,n𝖿𝗋)\mathbb{M}^{\sf fr}(g,n):=C_{*}(M_{g,n}^{\sf fr}).

The modular operad 𝕄𝖿𝗋\mathbb{M}^{\sf fr} is in fact an S1S^{1}-framed modular operad with the circle actions given by rotations of the local coordinate charts. Explicitly, the circle actions on 𝕄𝖿𝗋\mathbb{M}^{\sf fr} are given by degree one operators

Bi:𝕄𝖿𝗋(g,n)𝕄𝖿𝗋(g,n),  1in.B_{i}:\mathbb{M}^{\sf fr}(g,n)\rightarrow\mathbb{M}^{\sf fr}(g,n),\;\;1\leq i\leq n.

We refer to Section 2 for more details on the definition of S1S^{1}-framed modular operad and examples. Denote the homotopy quotient complex of the (S1)n(S^{1})^{n}-action by

𝕄S1𝖿𝗋(g,n):=(𝕄𝖿𝗋(g,n)[u11,,un1],+i=1nuiBi)\mathbb{M}^{\sf fr}_{S^{1}}(g,n):=\big{(}\mathbb{M}^{\sf fr}(g,n)[u_{1}^{-1},\ldots,u_{n}^{-1}],\partial+\sum_{i=1}^{n}u_{i}\cdot B_{i}\big{)}

with u1,,unu_{1},\ldots,u_{n} circle parameters of homological degree 2-2. Observe that the symmetric group SnS_{n} still acts on 𝕄S1𝖿𝗋(g,n)\mathbb{M}^{\sf fr}_{S^{1}}(g,n). Denote its further quotient space by 𝕄S1𝖿𝗋(g,n)Sn\mathbb{M}^{\sf fr}_{S^{1}}(g,n)_{S_{n}}.

String vertices.

With these preparations, we are ready to introduce one of the main constructions in defining CEI: Sen-Zwiebach’s differential graded Lie algebra (DGLA) [2, Section 3]. As a graded vector space, this DGLA is given by

𝔤:=((g,n)𝕄S1𝖿𝗋(g,n)Sn[1])[[,λ]]\mathfrak{g}:=\big{(}\bigoplus_{(g,n)}\mathbb{M}^{\sf fr}_{S^{1}}(g,n)_{S_{n}}[1]\big{)}[[\hbar,\lambda]]

where \hbar, λ\lambda are two formal variables both of homological degree 2-2. The differential of 𝔤\mathfrak{g} is of the form +i=1nuiBi+Δ\partial+\sum_{i=1}^{n}u_{i}\cdot B_{i}+\hbar\Delta with the operator Δ\Delta defined by all possible ways of “twisted sewing”:

Δ(αu1k1unkn):=1i<jnδki=0δkj=0cij(Biα)\Delta(\alpha\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}):=\sum_{1\leq i<j\leq n}\delta_{k_{i}=0}\cdot\delta_{k_{j}=0}\cdot c_{ij}(B_{i}\alpha)

The Lie bracket {,}\{-,-\} of 𝔤\mathfrak{g} is defined in a similar way:

{αu1k1unkn,βv1l1vnln}:=1in,1jnδki=0δlj=0cij(Biα,β)\{\alpha\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}},\beta\cdot v_{1}^{-l_{1}}\cdots v_{n^{\prime}}^{-l_{n^{\prime}}}\}:=\sum_{1\leq i\leq n,1\leq j\leq n^{\prime}}\delta_{k_{i}=0}\cdot\delta_{l_{j}=0}\cdot c_{ij}(B_{i}\alpha,\beta)

Observe that the circle operator BiB_{i} has homological degree one. This explains the shift by [1][1] in the definition of 𝔤\mathfrak{g}: to make the Lie bracket of homological degree zero. We refer to [2, Section 3] for more details of the construction of 𝔤\mathfrak{g}.

By definition, the string vertex is a Maurer-Cartan element of the form

𝒱:=(g,n)𝒱g,ngλ2g2+n\mathcal{V}:=\sum_{(g,n)}\mathcal{V}_{g,n}\hbar^{g}\lambda^{2g-2+n}

in the DGLA 𝔤\mathfrak{g} satisfying an initial condition of 𝒱0,3\mathcal{V}_{0,3}. The Maurer-Cartan equation satisfied by 𝒱\mathcal{V} is equivalent to the following system of equations usually known as the quantum master equation:

(2) (+i=1nuiBi)𝒱g,n+Δ𝒱g1,n+2+12g+g′′=g,n+n′′=n+2{𝒱g,n,𝒱g′′,n′′}=0(\partial+\sum_{i=1}^{n}u_{i}\cdot B_{i})\mathcal{V}_{g,n}+\Delta\mathcal{V}_{g-1,n+2}+\frac{1}{2}\sum_{g^{\prime}+g^{\prime\prime}=g,n^{\prime}+n^{\prime\prime}=n+2}\{\mathcal{V}_{g^{\prime},n^{\prime}},\mathcal{V}_{g^{\prime\prime},n^{\prime\prime}}\}=0

Costello proves that the above equations together with the initial condition of 𝒱0,3\mathcal{V}_{0,3} characterize 𝒱\mathcal{V} up to homotopy ( [5, Theorem 1]).

In the homological convention, a Maurer-Cartan element has degree 1-1, which implies that 𝒱g,n\mathcal{V}_{g,n} is of degree 6g6+2n6g-6+2n inside the equivariant chain complex 𝕄𝖿𝗋(g,n)[u11,,un1]\mathbb{M}^{\sf fr}(g,n)[u_{1}^{-1},\ldots,u_{n}^{-1}]. Thus explicitly, the string vertex at position (g,n)(g,n) is of the form

𝒱g,n=(k1,,kn)𝒱g,nk1,,knu1k1unkn,\mathcal{V}_{g,n}=\sum_{(k_{1},\ldots,k_{n})}\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}},

with the coefficient chains

𝒱g,nk1,,knC6g6+2n2k12kn(Mg,n𝖿𝗋).\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}\in C_{6g-6+2n-2k_{1}-\cdots-2k_{n}}(M_{g,n}^{\sf fr}).

By definition, it is also symmetric under permutations of the indices. Then, after unwinding the definition of CEI in [2, 3], in the case of Theorem A we have

uk1,,ukng,n,s={0,k1++kn3g3+n,n!|𝒱g,nk1,,kn|,k1++kn=3g3+n.\langle u^{k_{1}},\ldots,u^{k_{n}}\rangle^{\mathbb{Q},s}_{g,n}=\begin{cases}0,&\;\;\;k_{1}+\cdots+k_{n}\neq 3g-3+n,\\ n!\cdot|\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}|,&\;\;\;k_{1}+\cdots+k_{n}=3g-3+n.\end{cases}

Here in the second case, note that 𝒱g,nk1,,kn\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}} is a zero chain, hence it is a linear combination of points. The notation |||\cdot| stands for the sum of its coefficients.

Feynman compactification of 𝕄𝖿𝗋\mathbb{M}^{\sf fr}.

To prove Theorem A, it remains to relate these numbers with integrals of ψ\psi-classes on the Deligne-Mumford compactification M¯g,n\overline{M}_{g,n}. It is natural to do this on the level of operads.

In Section 2 we define the Feynman compactification of the S1S^{1}-framed modular operad 𝕄𝖿𝗋\mathbb{M}^{\sf fr}. Denote the resulting operad by F𝕄𝖿𝗋F\mathbb{M}^{\sf fr}. Intuitively speaking, this is an ordinary modular operad such that the circle actions on the composition maps of 𝕄𝖿𝗋\mathbb{M}^{\sf fr} have been universally trivialized. More precisely, an element of F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n) is given by a stable graph GΓ((g,n))G\in\Gamma((g,n)) together with decorations:

  • At a vertex vv, the decoration is by an element in the homotopy quotient 𝕄S1𝖿𝗋(g(v),𝖫𝖾𝗀(v)):=𝕄𝖿𝗋(g(v),𝖫𝖾𝗀(v))(S1)|𝖫𝖾𝗀(v)|\mathbb{M}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v)):=\mathbb{M}^{\sf fr}(g(v),{\sf Leg}(v))_{(S^{1})^{|{\sf Leg}(v)|}}.

  • At an edge ee, it is decorated by a homological degree 22 element DeD_{e}.

We think of the edge decoration DeD_{e} as giving a universal trivialization of the circle action on the composition map, by setting its boundary De\partial D_{e} to be exactly the twisted sewing operation along the edge ee, see Equation (6) for a precise definition.

Theorem B. There is an isomorphism of modular operads

H(F𝕄𝖿𝗋)H(M¯,)H_{*}(F\mathbb{M}^{\sf fr})\cong H_{*}(\overline{{M}},\mathbb{Q})

where M¯\overline{{M}} is the Deligne-Mumford modular operad [10, Section 6.2].

Remark 1.1.

This result may be viewed as a higher genus extension of the beautiful works by Dotsenko-Shadrin-Vallette [7] [8], Drummond-Cole [9] and Khoroshkin-Markarian-Shadrin [12] in genus zero. Indeed, we prove in Section 2 that a 𝕄𝖿𝗋\mathbb{M}^{\sf fr}-algebra VV together with a trivialization of the circle action induces a F𝕄𝖿𝗋F\mathbb{M}^{\sf fr}-algebra structure on the homotopy quotient VS1V_{S^{1}} (see Proposition 2 for details). Together with the theorem above, this implies that the homology H(VS1)H_{*}(V_{S^{1}}) carries a H(M¯)H_{*}(\overline{{M}})-algebra structure. However, the approach taken here which uses the Feynman compactification F𝕄F\mathbb{M} is different even in genus zero. It is worthwhile to explore the comparison with the previous works.

Remark 1.2.

We also point out that ”Feynman compactifications” are different from Getzler-Kapranov’s ”Feynman transforms” of modular operads [10]. Another rather interesting comparison question is to clarify the relationship between Theorem B with Getzler-Kapranov [10, Proposition 6.11].

To this point, we say a word about the proof of Theorem B. The idea is to exhibit a sequence of quasi-isomorphisms:

F𝕄𝖿𝗋(g,n)i𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝕀𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))pC(M¯g,n)\begin{CD}F{\mathbb{M}}^{\sf fr}(g,n)@>{i^{\sharp}}>{}>{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}@>{\mathbb{I}}>{}>{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}@>{p}>{}>C_{*}(\overline{M}_{g,n})\end{CD}

Here F𝕄𝖿𝗋(g,n)F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet} is a simplicial resolution of F𝕄𝖿𝗋(g,n)F{\mathbb{M}}^{\sf fr}(g,n) which we refer to as Mondello’s resolution as it’s first constructed by Mondello [15]. The hatted version 𝕄^𝖿𝗋(g,n)\widehat{\mathbb{M}}^{\sf fr}(g,n) is given by the normalized singular chain complex of the space M^g,n𝖿𝗋\widehat{M}_{g,n}^{\sf fr} constructed in [13]. Basically, this is the framed version of the oriented real blowup of M¯g,n\overline{M}_{g,n} along its boundary divisors, see Section 2 for more details. The first map ii^{\sharp} is constructed explicitly in Section 3, see Equation (13). Our main construction is the second map 𝕀\mathbb{I} defined in Section 4. The natural inclusion map Mg,n𝖿𝗋M^g,n𝖿𝗋M_{g,n}^{\sf fr}\hookrightarrow\widehat{M}_{g,n}^{\sf fr}, being a homotopy equivalence, does not respect the operad structures in the strict sense. The map 𝕀\mathbb{I} is precisely to interpolate between the two operadic compositions using higher coherent homotopies. The last map pp is a canonical projection map, see Paragraph 4. We prove in Section 4 that the composition p𝕀ip\mathbb{I}i^{\sharp} induces an isomorphism in homology that also respect the operadic composition on both sides.

Geometric understanding of string vertices.

Using Theorem B, we can finally unlock the mystery of string vertices and the quantum master equation (2). In Section 5, we obtain a simple formula expressing the orbifold fundamental class of the symmetric Deligne-Mumford space M¯g,n/Sn\overline{M}_{g,n}/S_{n} in terms “disk bundles” over string vertices {𝒱g,n}(g,n)\{\mathcal{V}_{g,n}\}_{(g,n)}:

(3) [M¯g,n/Sn]=GΓ((g,n))1|𝖠𝗎𝗍(G)|eEGDevVG𝒱g(v),|𝖫𝖾𝗀(v)|Sym.~{}[\overline{M}_{g,n}/S_{n}]=\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\prod_{e\in E_{G}}D_{e}\otimes\prod_{v\in V_{G}}\mathcal{V}^{\operatorname{Sym}}_{g(v),|{\sf Leg}(v)|}.

where the summation is over isomorphism classes of stable graphs of type (g,n)(g,n), DeD_{e} is a degree 22 element associated to an edge of GG, and the superscript Sym\operatorname{Sym} denotes the symmetrization map. Note that the closedness of the right hand side follows from the quantum master equation (2). Note that the above equation makes sense under the isomorphism in Theorem B: the fundamental cycle [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}] lies in the SnS_{n}-invariants of H(M¯g,n,)H_{*}(\overline{M}_{g,n},\mathbb{Q}), while the expression on the right hand side of Equation (3) lies inside the SnS_{n}-invariants of H(F𝕄𝖿𝗋(g,n))H_{*}(F\mathbb{M}^{\sf fr}(g,n)).

Using the above identity of the fundamental classes, Theorem AA follows immediately after unwinding the definitions of CEI.

Geometrically, Equation (3) may be viewed as a decomposition of the fundamental class of the symmetric Deligne-Mumford space M¯g,n/Sn\overline{M}_{g,n}/S_{n} according to its boundary strata labeled by stable graphs. String vertices are simply a coherent choice of the complement of a tubular neighborhood of the boundary divisor M¯g,n/Sn\partial\overline{M}_{g,n}/S_{n} in the ambient space M¯g,n/Sn\overline{M}_{g,n}/S_{n} for all (g,n)(g,n) such that 2g2+n>02g-2+n>0. The coherence equation is precisely the quantum master equation in the Sen-Zwiebach DGLA. Once such a coherent choice is made, Equation (3) simply corresponds to the decomposition of the fundamental class [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}] in terms disk bundles over products of string vertices.

We illustrate this point of view in the following figures. The left figure corresponds to the cases when (g,n)=(0,4)(g,n)=(0,4) or (1,1)(1,1). For example, the symmetric quotient M¯0,4/S4\overline{M}_{0,4}/S_{4} is topologically a sphere. In it there is a point (denoted by \infty) representing a nodal sphere with 44 marked points. Thus the fundamental class is decomposed into a (closed) disk bundle DD_{\infty} over the point \infty and 𝒱0,4\mathcal{V}_{0,4} which is the closure of M¯0,4/S4D\overline{M}_{0,4}/S_{4}-D_{\infty}. The same decomposition holds in the case of M¯1,1\overline{M}_{1,1} with the point \infty given by the once punctured nodal elliptic curve. The figure on the right illustrates the situation when two boundary divisors cross, producing a codimension 22 strata in the moduli space. For example, the symmetric quotient M¯0,5/S5\overline{M}_{0,5}/S_{5} is decomposed into 𝒱0,5\mathcal{V}_{0,5}, disk bundles and double disk bundles.

[Uncaptioned image][Uncaptioned image]

This geometric point of view was already implicit in the original work of Sen-Zwiebach [18] and more recently the work of Costello-Zwiebach [6]. From this perspective, it is appropriate to think of the CEI at (g,n)(g,n) as a way of decomposing an integral over M¯g,n\overline{M}_{g,n} into integrals over poly-disk bundles on each degenerate strata whose combinatorial types are given by stable graphs in Γ((g,n))\Gamma((g,n)).

Organization of the paper.

In Section 2 we define Feynman compactifications of S1S^{1}-framed modular operads. In Section 3 we introduce a simplicial resolution of the Feynman compactification F𝕄𝖿𝗋F\mathbb{M}^{\sf fr}, which is a minor modification of the one constructed by Mondello [15]. In Section 4, we prove Theorem B. Section 5 is devoted to prove the main identity (3). In Section 6 we obtain Theorem A, as well as clarify some technicalities in generalizing it to Fukaya categories with semi-simple cohomology.

Notations and Conventions.

Throughout the paper, we work with chain complexes and homological degree convention. Unless otherwise stated, we take the ground field to be \mathbb{Q}, the field of rationals. For a topological space XX, the notation C(X)C_{*}(X) stands for its normalized singular chain complex, with coefficients in \mathbb{Q}.

When working with graphs, we follow notations in [10, Section2]. By a labeled graph, we shall mean a graph GG that is endowed with a genus labeling g:VGg:V_{G}\rightarrow\mathbb{N} on its set of vertices VGV_{G}. The genus of a labeled graph is defined by

vVGg(v)+dimH1(G).\sum_{v\in V_{G}}g(v)+\dim H^{1}(G).

If GG is a labeled graph, denote by

  • VGV_{G}: the set of vertices of GG.

  • EGE_{G}: the set of edges of GG.

  • 𝖫𝖾𝗀(v){\mathsf{Leg}}(v): the set of half-edges at a vertex vVGv\in V_{G}.

  • 𝖫𝖾𝗀(G){\mathsf{Leg}}(G): the set of all half-edges of GG.

  • 𝖫𝖾𝖺𝖿(G){\sf Leaf}(G): the set of leaves of GG.

Denote by Γ(g,n)\Gamma(g,n) the set of isomorphism classes of labeled graphs of genus gg and with nn leaves. For a vertex vVGv\in V_{G} of a graph, its valency is denoted by n(v)n(v). A graph is called stable if 2g(v)2+n(v)>02g(v)-2+n(v)>0 holds at every vertex vVGv\in V_{G}. The set of isomorphism classes of stable graphs is denoted by Γ((g,n))\Gamma((g,n)). If h:GGh:G\rightarrow G^{\prime} is a morphism of graphs, denote by EGGEGE_{G\rightarrow G^{\prime}}\subset E_{G} the set of edges that are contracted by hh, and denote by EGGc=EG\EGGE_{G\rightarrow G^{\prime}}^{c}=E_{G}\backslash E_{G\rightarrow G^{\prime}} the complement.

Acknowledgments.

The author is grateful for Lino Amorim, Andrei Caldararu, and Kevin Costello for discussions around the topic.

2 Feynman compactifications of S1S^{1}-framed modular operads

In this section, we introduce the notion of Feynman compactifications of S1S^{1}-framed modular operads. Intuitively speaking, the construction yields an ordinary modular operad such that the circle actions on the composition maps of the original operad have been universally trivialized.

S1S^{1}-framed modular operads.

Recall from [10] the notion of a stable 𝕊\mathbb{S}-module given by a collection {P(g,n)|n,g0,2g+n2<0}\{P(g,n)|n,g\geq 0,2g+n-2<0\} of chain complexes with an action of the symmetric group SnS_{n} on P(g,n)P(g,n). For a finite set II of cardinality nn, we set

P(g,I):=(f:{1,,n}IP(g,n))Sn,P(g,I):=\big{(}\bigoplus_{f:\{1,\ldots,n\}\rightarrow I}P(g,n)\big{)}_{S_{n}},

with the direct sum taken over all bijections from {1,,n}\{1,\ldots,n\} to II. For each stable graph GG, denote by

P(G):=vVGP(g(v),𝖫𝖾𝗀(v)).P(G):=\bigotimes_{v\in V_{G}}P(g(v),{\mathsf{Leg}}(v)).

A modular operad structure on the stable 𝕊\mathbb{S}-module PP is given by an extension of PP to a functor on the category of stable graphs, that is, for each h:GGh:G\rightarrow G^{\prime} we have a morphism P(h):P(G)P(G)P(h):P(G)\rightarrow P(G^{\prime}) such that P(h1h0)=P(h1)P(h0)P(h_{1}h_{0})=P(h_{1})P(h_{0}) whenever h1h_{1} and h0h_{0} are two composable morphisms of stable graphs.

We shall work with a S1S^{1}-framed version of modular operads. First, we introduce some terminologies. Let (C,)(C,\partial) be a chain complex. By an S1S^{1}-action on CC, we mean a homological degree one operator B:CC+1B:C_{\bullet}\rightarrow C_{\bullet+1} such that B2=0B^{2}=0 and B+B=0\partial B+B\partial=0. Let n1n\geq 1, by an (S1)n(S^{1})^{n}-action, we mean nn mutually commuting circle actions Bj(j=1,,n)B_{j}\;(j=1,\ldots,n), i.e. BiBj+BjBi=0B_{i}B_{j}+B_{j}B_{i}=0 for all 1i,jn1\leq i,j\leq n. Its homotopy quotient is defined as the chain complex

C(S1)n:=(C[u11,,un1],+j=1nBjuj)C_{(S^{1})^{n}}:=\Big{(}C[u_{1}^{-1},\ldots,u_{n}^{-1}],\partial+\sum_{j=1}^{n}B_{j}u_{j}\Big{)}

with the uju_{j}’s of homological degree 2-2. The [u1,,un]\mathbb{Q}[u_{1},\ldots,u_{n}]-module structure is given by the quotient of C[u1±,,un±]C[u_{1}^{\pm},\ldots,u_{n}^{\pm}] by the subspace jujC[u1,,un]\sum_{j}u_{j}\cdot C[u_{1},\ldots,u_{n}]. Observe that the inclusion map

CC[u11,,un1]=C(S1)nC\hookrightarrow C[u_{1}^{-1},\ldots,u_{n}^{-1}]=C_{(S^{1})^{n}}

is a map of chain complexes, which we refer to as the canonical quotient map.

With this preparation, an S1S^{1}-framed modular operad is given by a modular operad PP together with the following additional data:

  • There is an action of Sn(S1)nS_{n}\rtimes(S^{1})^{n} on each P(g,n)P(g,n), extending the SnS_{n}-action.

  • The morphism P(h):P(G)P(G)P(h):P(G)\rightarrow P(G^{\prime}) associated with a morphism h:GGh:G\rightarrow G^{\prime} with G,GΓ(g,n)G,G^{\prime}\in\Gamma(g,n) is Sn(S1)|𝖫𝖾𝗀(G)|S_{n}\rtimes(S^{1})^{|{\mathsf{Leg}}(G^{\prime})|}-equivariant. Here we implicitly identified 𝖫𝖾𝗀(G){\mathsf{Leg}}(G^{\prime}) as a subset of 𝖫𝖾𝗀(G){\mathsf{Leg}}(G) using the morphism hh (see [10, Section 2.13]).

  • The morphism P(h)P(h) is invariant under simultaneous rotation at the two ends of an edge. More precisley, we require that

    P(h)(Be,+α)=P(h)(Be,α).P(h)(B_{e,+}\alpha)=P(h)(B_{e,-}\alpha).

    Here αP(G)\alpha\in P(G) is of the form vVGαv\otimes_{v\in V_{G}}\alpha_{v} and Be,+B_{e,+} and Be,B_{e,-} are the two circle actions at the two half edges of a contracted edge eEGGe\in E_{G\rightarrow G^{\prime}}.

A first example.

For each (g,n)(g,n), let M^g,n\widehat{M}_{g,n} be the oriented real blowup along the boundary divisors in the Deligne-Mumford space M¯g,n\overline{M}_{g,n}. An element in M^g,n\widehat{M}_{g,n} is given by a stable curve (Σ,p1,,pn)M¯g,n(\Sigma,p_{1},\ldots,p_{n})\in\overline{M}_{g,n} together with a decoration at each nodal point by a unit tangent vector :

vαTxαΣnTyαΣn,v_{\alpha}\in T_{x_{\alpha}}\Sigma^{n}\otimes T_{y_{\alpha}}\Sigma^{n},

where ΣnΣ\Sigma^{n}\rightarrow\Sigma is the normalization of Σ\Sigma and {xα,yα}\{x_{\alpha},y_{\alpha}\} is the preimage of a nodal point labeled by α\alpha. We also consider a framed version of it denoted by M^g,n𝖿𝗋\widehat{M}_{g,n}^{\sf fr}. An element of this moduli space M^g,n𝖿𝗋\widehat{M}_{g,n}^{\sf fr} consists of a decorated stable Riemann surface (Σ,p1,,pn,αvα)M^g,n(\Sigma,p_{1},\ldots,p_{n},\prod_{\alpha}v_{\alpha})\in\widehat{M}_{g,n}, together with a framing around each marked point. That is, for each 1in1\leq i\leq n, a local coordinate system φi:𝔻2U(pi)Σ\varphi_{i}:\mathbb{D}^{2}\rightarrow U(p_{i})\subset\Sigma such that the biholomorphic maps φ\varphi’s extend to an open neighborhood of 𝔻2\mathbb{D}^{2}. We also require that the pair-wise intersection of the closures of the framed disks be empty, i.e. U(pi)¯U(pj)¯=\overline{U(p_{i})}\cap\overline{U(p_{j})}=\emptyset. The circle S1S^{1} acts on a local coordinate system by rotation: φφeiθ\varphi\mapsto\varphi\circ e^{i\theta}. This action is clearly a free action. It is also well-known that the set of local coordinate systems around any point pip_{i} is homotopy equivalent to S1S^{1}. Thus, the forgetful map

M^g,n𝖿𝗋M^g,n\widehat{M}_{g,n}^{\sf fr}\rightarrow\widehat{M}_{g,n}

is a homotopy (S1)n(S^{1})^{n}-bundle.

We define a S1S^{1}-equivariant modular operad 𝕄^𝖿𝗋\widehat{\mathbb{M}}^{\sf fr} by setting

𝕄^𝖿𝗋(g,n):=C(M^g,n𝖿𝗋).\widehat{\mathbb{M}}^{\sf fr}(g,n):=C_{*}(\widehat{M}_{g,n}^{\sf fr}).

Let h:GGh:G\rightarrow G^{\prime} be a morphism of stable graphs, the composition morphism is defined as follows. The map on the underlying stable curve is the product of inclusions

vVGM¯g(v),𝖫𝖾𝗀(v)=vVGvh1(v)M¯g(v),𝖫𝖾𝗀(v)vVGM¯g(v),𝖫𝖾𝗀(v)\prod_{v\in V_{G}}\overline{M}_{g(v),{\mathsf{Leg}}(v)}=\prod_{v^{\prime}\in V_{G^{\prime}}}\prod_{v\in h^{-1}(v^{\prime})}\overline{M}_{g(v),{\mathsf{Leg}}(v)}\rightarrow\prod_{v^{\prime}\in V_{G^{\prime}}}\overline{M}_{g(v^{\prime}),{\mathsf{Leg}}(v^{\prime})}

We keep the same framings at marked points, and the same decorations of tangent vectors at nodes corresponding to the edges of EGGcE^{c}_{G\rightarrow G^{\prime}}. For the decoration at a node corresponding to a contracted edge eαEGGe_{\alpha}\in E_{G\rightarrow G^{\prime}}, we decorate by the unit tangent direction of the vector

vα:=dφxα(ddz)dφyα(ddw)TxαΣnTyαΣn,v_{\alpha}:=d\varphi_{x_{\alpha}}(\frac{d}{dz})\otimes d\varphi_{y_{\alpha}}(\frac{d}{dw})\in T_{x_{\alpha}}\Sigma^{n}\otimes T_{y_{\alpha}}\Sigma^{n},

where φxα:𝔻zU(xα)\varphi_{x_{\alpha}}:\mathbb{D}_{z}\rightarrow U(x_{\alpha}) and φyα:𝔻wU(yα)\varphi_{y_{\alpha}}:\mathbb{D}_{w}\rightarrow U(y_{\alpha}) are the two framings at the two legs of the edge eαe_{\alpha}. This defines a map

ξGG:vVGM^g(v),𝖫𝖾𝗀(v)𝖿𝗋vVGM^g(v),𝖫𝖾𝗀(v)𝖿𝗋\xi_{G\rightarrow G^{\prime}}:\prod_{v\in V_{G}}\widehat{M}_{g(v),{\mathsf{Leg}}(v)}^{\sf fr}\rightarrow\prod_{v^{\prime}\in V_{G^{\prime}}}\widehat{M}_{g(v^{\prime}),{\mathsf{Leg}}(v^{\prime})}^{\sf fr}

which, by taking normalized singular chain functor, induces the desired map 𝕄^𝖿𝗋(h)\widehat{\mathbb{M}}^{\sf fr}(h).

A second example.

We define another example of S1S^{1}-equivariant modular operad 𝕄𝖿𝗋\mathbb{M}^{\sf fr} by setting

𝕄𝖿𝗋(g,n):=C(Mg,n𝖿𝗋)\mathbb{M}^{\sf fr}(g,n):=C_{*}(M_{g,n}^{\sf fr})

given by the moduli space of smooth curves of type (g,n)(g,n) together with a framing at each marked point. The composition morphism

𝕄𝖿𝗋(h):𝕄𝖿𝗋(G)𝕄𝖿𝗋(G)\mathbb{M}^{\sf fr}(h):\mathbb{M}^{\sf fr}(G)\rightarrow\mathbb{M}^{\sf fr}(G^{\prime})

associated to a contraction map h:GGh:G\rightarrow G^{\prime} is defined as follows. Again we keep the same framing at the leaves of GG and GG^{\prime}. For a contracted edge eEGGe\in E_{G\rightarrow G^{\prime}}, we sew the Riemann surfaces at the two ends of ee using the two framings φxα:𝔻zU(xα)\varphi_{x_{\alpha}}:\mathbb{D}_{z}\rightarrow U(x_{\alpha}) and φyα:𝔻wU(yα)\varphi_{y_{\alpha}}:\mathbb{D}_{w}\rightarrow U(y_{\alpha}). More precisely, we cut out U(xα)U(x_{\alpha}) and U(yα)U(y_{\alpha}) from the two surfaces, then identify a neighborhood of the boundary circle by the equation zw=1zw=1. This defines a map

ηGG:vVGMg(v),𝖫𝖾𝗀(v)𝖿𝗋vVGMg(v),𝖫𝖾𝗀(v)𝖿𝗋\eta_{G\rightarrow G^{\prime}}:\prod_{v\in V_{G}}M_{g(v),{\mathsf{Leg}}(v)}^{\sf fr}\rightarrow\prod_{v\in V_{G^{\prime}}}M_{g(v^{\prime}),{\mathsf{Leg}}(v^{\prime})}^{\sf fr}

which induces the desired map 𝕄𝖿𝗋(h)\mathbb{M}^{\sf fr}(h).

It is a general fact of the oriented real blowup construction that the natural inclusion map Mg,n𝖿𝗋M^g,n𝖿𝗋M_{g,n}^{\sf fr}\hookrightarrow\widehat{M}_{g,n}^{\sf fr} is a homotopy equivalence, which induces a homotopy equivalence of chain complexes for each (g,n)(g,n):

Ig,n:𝕄𝖿𝗋(g,n)𝕄^𝖿𝗋(g,n).I_{g,n}:\mathbb{M}^{\sf fr}(g,n)\hookrightarrow\widehat{\mathbb{M}}^{\sf fr}(g,n).

However, observe that the operad structure of 𝕄^𝖿𝗋\widehat{\mathbb{M}}^{\sf fr} does not restrict to that of 𝕄𝖿𝗋\mathbb{M}^{\sf fr} since when composing, the former is by forming a nodal curve while the latter is by sewing using framings.

Twisted sewing operations.

Let PP be an S1S^{1}-equivariant modular operad. We may form the homotopy quotients: for each pair (g,n)(g,n) we set

P(g,n)(S1)n=(P(g,n)[u11,,un1],+j=1nBjuj)P(g,n)_{(S^{1})^{n}}=\Big{(}P(g,n)[u_{1}^{-1},\ldots,u_{n}^{-1}],\partial+\sum_{j=1}^{n}B_{j}u_{j}\Big{)}

This collection of stable 𝕊\mathbb{S}-module does not form a modular operad. This is evident in the previous two examples: for instance in the case of 𝕄𝖿𝗋\mathbb{M}^{\sf fr}, the sewing operation of two Riemann surfaces is ambiguously defined if we first quotient out the circle actions on the framings.

Nevertheless, there does exist a degree one composition by allowing a full S1S^{1}-twist of the composition map of PP. There are two types of compositions depending on whether we compose along a loop edge or a non-loop edge. We begin with the loop edge case as illustrated in Figure 11.

Refer to caption
Figure 1: Loop contraction

Indeed, we write down ρij:P(g,n)(S1)nP(g+1,n2)(S1)n2\rho_{ij}:P(g,n)_{(S^{1})^{n}}\rightarrow P(g+1,n-2)_{(S^{1})^{n-2}} by setting

(4) ρij(xu1k1unkn):={0ki0, or kj0P(cij)(Bix)u1k1unknki=kj=0~{}\rho_{ij}(x\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}):=\begin{cases}0&k_{i}\neq 0,\mbox{ or }\;k_{j}\neq 0\\ P(c_{ij})(B_{i}x)u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}&k_{i}=k_{j}=0\end{cases}

The case of a non-loop edge contraction is similar, see Figure 22. Explicitly we define a map of chain complexes ρij:P(g,n)(S1)nP(h,m)(S1)mP(g+h,n+m2)(S1)n+m2\rho_{ij}:P(g,n)_{(S^{1})^{n}}\otimes P(h,m)_{(S^{1})^{m}}\rightarrow P(g+h,n+m-2)_{(S^{1})^{n+m-2}} for each 1in,1jm1\leq i\leq n,1\leq j\leq m by

(5) ρij(xu1k1unkn,yv1l1vmlm)={0ki0, or lj0P(cij)(Bixy)u1k1unknv1l1vmlmki=lj=0\displaystyle~{}\begin{split}&\rho_{ij}(x\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}},y\cdot v_{1}^{-l_{1}}\cdots v_{m}^{-l_{m}})\\ =&\begin{cases}0&k_{i}\neq 0,\mbox{ or }l_{j}\neq 0\\ P(c_{ij})(B_{i}x\otimes y)u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}\cdot v_{1}^{-l_{1}}\cdots v_{m}^{-l_{m}}&k_{i}=l_{j}=0\end{cases}\end{split}

where hh is the composition map of cijc_{ij} is the graph map that contracts the edge connecting ii and jj.

Refer to caption
Figure 2: Non-loop contraction

Generalizing the above, for each stable graph GG, we put

PS1(G):=vVGP(g(v),𝖫𝖾𝗀(v))(S1)|𝖫𝖾𝗀(v)|P_{S^{1}}(G):=\bigotimes_{v\in V_{G}}P(g(v),{\sf Leg}(v))_{(S^{1})^{|{\sf Leg}(v)|}}

to be the homotopy quotient version of P(G)P(G), i.e. on each vertex of GG we put the homotopy quotient by the circle action. Furthermore, associated to an edge eGe\in G we also have the corresponding twisted composition map defined using ρij\rho_{ij}’s constructed above:

ρe:PS1(G)PS1(G/e)\rho_{e}:P_{S^{1}}(G)\rightarrow P_{S^{1}}(G/e)

Note that ρe\rho_{e} is of homological degree one.

Feynman compactifications of S1S^{1}-equivariant modular operads.

We now proceed to define the Feynman compactification FPFP of an S1S^{1}-framed modular operad PP. The underlying stable 𝕊\mathbb{S}-module of FPFP is given by

FP(g,n):=GΓ(g,n)(D(G)PS1(G))𝖠𝗎𝗍(G),FP(g,n):=\bigoplus_{G\in\Gamma(g,n)}\big{(}D(G)\otimes P_{S^{1}}(G)\big{)}_{{\sf Aut}(G)},

where D(G)D(G) is the ”cocycle” (see [10, Section 4]) defined by

D(G):=eEGDe.D(G):=\bigotimes_{e\in E_{G}}\mathbb{Q}\cdot D_{e}.

The notation De\mathbb{Q}\cdot D_{e} stands for the one-dimensional vector space generated by DeD_{e} at homological degree 22. Hence D(G)D(G) is also one-dimensional, at homological degree 2|EG|2|E_{G}|, generated by the tensor product

DG:=eEGDe.D_{G}:=\bigotimes_{e\in E_{G}}D_{e}.

The differential on FP(g,n)FP(g,n) is the given by +δ\partial+\delta with \partial the differential of PS1(G)P_{S^{1}}(G), while the extra map δ\delta is given by

(6) δ(DGα)=eEGDG/eρe(α),αPS1(G).\displaystyle~{}\delta(D_{G}\otimes\alpha)=\sum_{e\in E_{G}}D_{G/e}\otimes\rho_{e}(\alpha),\;\;\alpha\in P_{S^{1}}(G).

Observe that since ρe\rho_{e} has degree one while DeD_{e} each has degree two, the degree of δ\delta is 1-1 as required. The idea behind this definition of δ\delta is that each DeD_{e} should be thought of as a universal trivialization of the twisted sewing operation along ee.

Composition maps of FPFP.

Associated with a morphism h:GGh:G\rightarrow G^{\prime} of stable graphs, we shall define a composition map FP(h):FP(G)FP(G)FP(h):FP(G)\rightarrow FP(G^{\prime}) so that the stable 𝕊\mathbb{S}-module {FP(g,n)}(g,n)\{FP(g,n)\}_{(g,n)} forms a modular operad. By the structure of modular operads [10, Section 3], it suffices to construct FP(cij)FP(c_{ij}) for an edge contraction map cij:GG/ec_{ij}:G\rightarrow G/e as in Figure 11 and Figure 22. In the case of Figure 11, by definition of FPFP, we have

FP(G)\displaystyle FP(G) =HΓg,n+2(D(H)PS1(H))𝖠𝗎𝗍(H)\displaystyle=\bigoplus_{H\in\Gamma_{g,n+2}}\big{(}D(H)\otimes P_{S^{1}}(H)\big{)}_{{\sf Aut}(H)}
FP(G/e)\displaystyle FP(G/e) =KΓg+1,n(D(K)PS1(K))𝖠𝗎𝗍(K)\displaystyle=\bigoplus_{K\in\Gamma_{g+1,n}}\big{(}D(K)\otimes P_{S^{1}}(K)\big{)}_{{\sf Aut}(K)}

For a stable graph HΓg,n+2H\in\Gamma_{g,n+2}, we denote by HijΓg+1,nH_{ij}\in\Gamma_{g+1,n} the stable graph obtained from HH by forming a loop ee using the two legs labeled by ii and jj. Observe that D(Hij/e)D(H)D(H_{ij}/e)\cong D(H) since both graph have isomorphic set of edges. Then we set the composition map as

(7) FP(cij)(DHxu1k1un+2kn+2)={DHij(xu1k1un+2kn+2)(ui+uj)ki0, or kj0DHij/eP(cij)(x)u1k1un+2kn+2ki=kj=0\displaystyle~{}\begin{split}&FP(c_{ij})\big{(}D_{H}\otimes xu_{1}^{-k_{1}}\cdots u_{n+2}^{-k_{n+2}}\big{)}\\ =&\begin{cases}D_{H_{ij}}\otimes(x\cdot u_{1}^{-k_{1}}\cdots u_{n+2}^{-k_{n+2}})\cdot(u_{i}+u_{j})&k_{i}\neq 0,\mbox{ or }\;k_{j}\neq 0\\ D_{H_{ij}/e}\otimes P(c_{ij})(x)\cdot u_{1}^{-k_{1}}\cdots u_{n+2}^{-k_{n+2}}&k_{i}=k_{j}=0\end{cases}\end{split}
{Lemma}

The composition map FP(cij)FP(c_{ij}) is a chain map.

Proof 2.1.

For simplicity, we shall suppress the uu’s to only uiu_{i} and uju_{j} as other circle parameters are not relevant in the calculation. We need to verify that [+uB+δ,FP(cij)]=0[\partial+uB+\delta,FP(c_{ij})]=0. When the input is of the form DHxD_{H}\otimes x, one has

(+uB+δ)FP(cij)(DHx)\displaystyle(\partial+uB+\delta)FP(c_{ij})(D_{H}\otimes x) =(+uB+δ)(DHij/eP(cij)(x))\displaystyle=(\partial+uB+\delta)(D_{H_{ij}/e}\otimes P(c_{ij})(x))
=δDHij/eP(cij)(x)+DHij/eP(cij)(x)\displaystyle=\delta D_{H_{ij}/e}\otimes P(c_{ij})(x)+D_{H_{ij}/e}\otimes\partial P(c_{ij})(x)
FP(cij)(+uB+δ)(DHx)\displaystyle FP(c_{ij})(\partial+uB+\delta)(D_{H}\otimes x) =FP(cij)(δDHx+DHx)\displaystyle=FP(c_{ij})(\delta D_{H}\otimes x+D_{H}\otimes\partial x)
=δDHij/eP(cij)(x)+DHij/eP(cij)(x)\displaystyle=\delta D_{H_{ij}/e}\otimes P(c_{ij})(x)+D_{H_{ij}/e}\otimes P(c_{ij})(\partial x)

The modular operad structure P(cij)P(c_{ij}) is a chain map, which shows that the two computations indeed give the same answer. When the input is of the form DHxui1D_{H}\otimes xu_{i}^{-1} we compute the two compositions as

(+uB+δ)FP(cij)(DHxui1)\displaystyle(\partial+uB+\delta)FP(c_{ij})(D_{H}\otimes xu_{i}^{-1})
=\displaystyle= (+uB+δ)(DHijx)\displaystyle(\partial+uB+\delta)(D_{H_{ij}}\otimes x)
=\displaystyle= δDHijx+DHijx\displaystyle\delta D_{H_{ij}}\otimes x+D_{H_{ij}}\otimes\partial x
FP(cij)(+uB+δ)(DHxui1)\displaystyle FP(c_{ij})(\partial+uB+\delta)(D_{H}\otimes xu_{i}^{-1})
=\displaystyle= FP(cij)(δDHxui1+DHxui1+DHBix)\displaystyle FP(c_{ij})(\delta D_{H}\otimes xu_{i}^{-1}+D_{H}\otimes\partial xu_{i}^{-1}+D_{H}\otimes B_{i}x)
=\displaystyle= (δDHijxDHij/eP(cij)(Bix))+DHijx+DHij/eP(cij)(Bix)\displaystyle\big{(}\delta D_{H_{ij}}\otimes x-D_{H_{ij}/e}\otimes P(c_{ij})(B_{i}x)\big{)}+D_{H_{ij}}\otimes\partial x+D_{H_{ij}/e}\otimes P(c_{ij})(B_{i}x)
=\displaystyle= δDHijx+DHijx\displaystyle\delta D_{H_{ij}}\otimes x+D_{H_{ij}}\otimes\partial x

The case of DHxuj1D_{H}\otimes xu_{j}^{-1} is similar. In the general case when the input is DHxuikiujkjD_{H}\otimes xu_{i}^{-k_{i}}u_{j}^{-k_{j}} with ki,kj1k_{i},k_{j}\geq 1, the computation is easier and is left as an exercise.

The second type composition corresponding to Figure 22 is defined similarly:

FP(cij)(DH1xu1k1unkn,DH2yv1l1vmlm)\displaystyle FP(c_{ij})\big{(}D_{H_{1}}\otimes x\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}},D_{H_{2}}\otimes y\cdot v_{1}^{-l_{1}}\cdots v_{m}^{-l_{m}}\big{)}
=\displaystyle= {DHij(xu1k1unkn)yv1l1vmlm)(ui+vj)ki0, or lj0DHij/eP(cij)(x,y)u1k1unknv1l1vmlmki=kj=0\displaystyle\begin{cases}D_{H_{ij}}\otimes\big{(}x\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}})\otimes y\cdot v_{1}^{-l_{1}}\cdots v_{m}^{-l_{m}}\big{)}(u_{i}+v_{j})&k_{i}\neq 0,\mbox{ or }\;l_{j}\neq 0\\ D_{H_{ij}/e}\otimes P(c_{ij})(x,y)\cdot u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}\cdot v_{1}^{-l_{1}}\cdots v_{m}^{-l_{m}}&k_{i}=k_{j}=0\end{cases}

The verification that this defines a chain map is analogous to the previous lemma.

Trivializations of circle actions.

By a cyclic PP-algebra structure on a chain complex (V,b)(V,b) we mean the following data:

  • A circle action BB on the chain complex (V,b)(V,b).

  • An (S1)nSn(S^{1})^{n}\ltimes S_{n}-equivariant action map

    ρV:P(g,n)Vn.\rho_{V}:P(g,n)\rightarrow V^{\otimes n}.
  • A graded symmetric inner product ,:V2𝕂\langle-,-\rangle:V^{\otimes 2}\rightarrow\mathbb{K} such that bb is graded anti-self-adjoint and BB is graded self-adjoint.

The compatibility condition between ρV\rho_{V} and the inner product should be self-evident: for an elementary contraction cijc_{ij}, it corresponds to contracting two copies of VV at the half-edges labeled by ii and jj using the inner product.

A trivialization of the circle action BB (see [7, 12]) is given by a chain map

s:(V,b)(V[[u]],b+uB)s:(V,b)\rightarrow(V[[u]],b+uB)

of the form s(x)=x+R1(x)u+R2(x)u2+s(x)=x+R_{1}(x)u+R_{2}(x)u^{2}+\cdots with Rj:VV,(j1)R_{j}:V\rightarrow V,\;(j\geq 1) of homological degree 2j2j.

{Proposition}

Let VV be a cyclic PP-algebra and let s:VV[[u]]s:V\rightarrow V[[u]] be a trivialization of the circle action. Then this data induces a natural cyclic FPFP-algebra structure on VS1=V[u1]V_{S^{1}}=V[u^{-1}].

Proof 2.2.

Taking (S1)n(S^{1})^{n}-quotients of the action map

ρV:P(g,n)Vn\rho_{V}:P(g,n)\rightarrow V^{\otimes n}

yields an action of PS1(g,n)P_{S^{1}}(g,n) on VS1V_{S^{1}}, i.e. an action map still denoted by

ρ:PS1(g,n)VS1n.\rho:P_{S^{1}}(g,n)\rightarrow V_{S^{1}}^{\otimes n}.

To extend this action to FPFP it suffices to define the action of DeD_{e} associated with an edge eEGe\in E_{G} of a stable graph GG. Observe that the inverse operator of RR is another operator of the form

T=id+T1u+T2u2+.T=\operatorname{id}+T_{1}u+T_{2}u^{2}+\cdots.

Here the operator TjT_{j}’s are defined by the following identity

i+j=kTiSj={id if k=0,0 if k1.\sum_{i+j=k}T_{i}S_{j}=\begin{cases}\operatorname{id}\mbox{\;\; if \;\;}k=0,\\ 0\mbox{\;\; if \;\;}k\geq 1.\end{cases}

Solving the above recursively yields formulas of TjT_{j} in terms RiR_{i}’s. For example, we have

T1=R1,T2=R2+R1R1T_{1}=-R_{1},\;\;T_{2}=-R_{2}+R_{1}R_{1}

Using the RR’s and TT’s, we define a contraction map

H:=i0,j0Hi,j:V[u1]V[u1]𝕂H:=\sum_{i\geq 0,j\geq 0}H_{i,j}:V[u^{-1}]\otimes V[u^{-1}]\rightarrow\mathbb{K}

with component maps Hi,j:VuiVuj𝕂H_{i,j}:V\cdot u^{-i}\otimes V\cdot u^{-j}\rightarrow\mathbb{K} given by formula

Hi,j(xui,yuj):=(1)jl=0jRlTi+j+1lx,y.H_{i,j}(xu^{-i},yu^{-j}):=\langle(-1)^{j}\sum_{l=0}^{j}R_{l}T_{i+j+1-l}x,y\rangle.

Here we set R0=T0=idR_{0}=T_{0}=\operatorname{id}. Then we define the action of DeD_{e} on VS1V_{S^{1}} by the symmetrization of HH, i.e. DeD_{e} acts by

H𝗌𝗒𝗆(α,β):=12(H(α,β)+(1)|α||β|H(β,α)).H^{\sf sym}(\alpha,\beta):=\frac{1}{2}\big{(}H(\alpha,\beta)+(-1)^{|\alpha||\beta|}H(\beta,\alpha)\big{)}.

The verification that [b+uB,H][b+uB,H] is compatible with the boundary map δDe\delta D_{e} in Equation (6) is a calculation, and is done in [3, Proposition 4.5].

3 Mondello’s resolution

Let 𝕄𝖿𝗋\mathbb{M}^{\sf fr} and 𝕄^𝖿𝗋\widehat{\mathbb{M}}^{\sf fr} be the two S1S^{1}-framed modular operads as defined in Paragraphs 2 and 2. In this section, we introduce a simplicial resolution denoted by F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet} of the Feynman compactification F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n). We call it Mondello’s resolution since its construction is a minor modification of Mondello’s construction in [15]. Similarly, we also have the resolution F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n)_{\bullet} of the Feynman compactification F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n). The main result of the section is the existence of the following quasi-isomorphisms:

F𝕄𝖿𝗋(g,n)\displaystyle F\mathbb{M}^{\sf fr}(g,n) 𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))\displaystyle\cong{\sf Tot}\big{(}F\mathbb{M}^{\sf fr}(g,n)_{\bullet}\big{)}
F𝕄^𝖿𝗋(g,n)\displaystyle F\widehat{\mathbb{M}}^{\sf fr}(g,n) 𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))\displaystyle\cong{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}

We shall mainly deal with the hatted version, the un-hatted version is completely parallel.

Mondello’s construction [15].

Let XX be a smooth algebraic variety with a simple 222Simplicity is not necessary for this construction, as was remarked by Mondello [15, Footnote 1]. normal crossing divisor

D=iID¯i,D=\bigcup_{i\in I}\overline{D}_{i},

i.e. we assume that each D¯i\overline{D}_{i} is a smooth irreducible divisor, and pair-wise intersect transversally. For a subset JIJ\subset I, denote by D¯J:=jJD¯j\overline{D}_{J}:=\bigcap_{j\in J}\overline{D}_{j}. Denote by D^J\widehat{D}_{J} the manifold with corners obtained by performing a real oriented blow-up of D¯J\overline{D}_{J} along a divisor iI\JD¯J{i}\bigcup_{i\in I\backslash J}\overline{D}_{J\cup\{i\}}. Note that the natural inclusion DJ:=D¯J\(iI\JD¯J{i})D^JD_{J}:=\overline{D}_{J}\backslash(\bigcup_{i\in I\backslash J}\overline{D}_{J\cup\{i\}})\hookrightarrow\widehat{D}_{J} is a homotopy equivalence. For a chain J0J1JkIJ_{0}\subsetneq J_{1}\subsetneq\cdots\subsetneq J_{k}\subset I of inclusions of subsets of II, define D^J0,,Jk:=D^J0×D¯J0D¯Jk\widehat{D}_{J_{0},\cdots,J_{k}}:=\widehat{D}_{J_{0}}\times_{\overline{D}_{J_{0}}}\overline{D}_{J_{k}}. Observe that the space

(8) D^J0,,Jk=iJk\J0bJk(SND¯i/X|D¯Jk)~{}\widehat{D}_{J_{0},\cdots,J_{k}}=\prod_{i\in J_{k}\backslash J_{0}}b_{J_{k}}^{*}\big{(}SN_{\overline{D}_{i}/X}|_{\overline{D}_{J_{k}}}\big{)}

is given by the (S1)d(S^{1})^{d}-bundle where SND¯i/XSN_{\overline{D}_{i}/X} is the circle bundle associated to the normal bundle of D¯i\overline{D}_{i} in XX, the map bJk:D^JkD¯Jkb_{J_{k}}:\widehat{D}_{J_{k}}\rightarrow\overline{D}_{J_{k}} is the natural projection, and d=|Jk\J0|d=|J_{k}\backslash J_{0}|. In loc. cit. the author constructs a simplicial topological space XX_{\bullet} with

Xk:=J0J1JkID^J0,,Jk,X_{k}:=\coprod_{J_{0}\subsetneq J_{1}\subsetneq\cdots\subsetneq J_{k}\subset I}\widehat{D}_{J_{0},\ldots,J_{k}},

such that its geometric realization |X||X_{\bullet}| is homotopy equivalent to XX.

Application to M¯g,n\overline{M}_{g,n}.

We apply the previous construction to the case with XX equals the Deligne-Mumford compactification M¯g,n\overline{M}_{g,n}, and DD the divisor corresponding to nodal curves. Instead of working with an index set II labeling the irreducible components of DD, it is most natural in this setup to work with stable graphs in Γ((g,n))\Gamma((g,n)). That is, for each GΓ((g,n))G\in\Gamma((g,n)), let us write M¯G\overline{M}_{G} for the image of the map

(9) ξG:vVG𝗂𝗇𝗍M¯g(v),|𝖫𝖾𝗀(v)|M¯GM¯g,n.~{}\xi_{G}:\prod_{v\in V_{G}^{\sf int}}\overline{M}_{g(v),|{\sf Leg}(v)|}\rightarrow\overline{M}_{G}\subset\overline{M}_{g,n}.

Denote by MG:=ξG(vVG𝗂𝗇𝗍Mg(v),𝗏𝖺𝗅(v))M_{G}:=\xi_{G}\big{(}\prod_{v\in V_{G}^{\sf int}}M_{g(v),{\sf val}(v)}\big{)} the locus of MG¯\overline{M_{G}} such that at each vertex of GG we have a smooth curve in Mg(v),|𝖫𝖾𝗀(v)|M_{g(v),|{\sf Leg}(v)|}. The oriented real blowup of M¯G\overline{M}_{G} along boundary divisors is denoted by M^G\widehat{M}_{G}. A general fact of oriented real blowup is that the natural inclusion map MGM^GM_{G}\hookrightarrow\widehat{M}_{G} is a homotopy equivalence. Let GGG\rightarrow G^{\prime} be a graph contraction map. It induces a corresponding inclusion map

ξGG:M¯GM¯G\xi_{G\rightarrow G^{\prime}}:\overline{M}_{G}\rightarrow\overline{M}_{G^{\prime}}

In dimension k0k\geq 0, we describe the set of non-degenerate kk-simplices of Mondello’s construction applied to M¯g,n\overline{M}_{g,n} in terms of stable graphs. For each kk-step graph contraction, we set

M^G0Gk:=M^GkM¯G0.\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}:=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}}.

Explicitly, an element of M^GkM¯G0\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}} consists of

  • A stable curve with marked points (Σ,p1,,pn)M¯G0(\Sigma,p_{1},\ldots,p_{n})\in\overline{M}_{G_{0}}. By definition, its dual graph GG admits a contraction map GG0G\rightarrow G_{0}.

  • For each edge that got contracted by the composition map GG0GkG\rightarrow G_{0}\rightarrow G_{k}, the corresponding node xΣx\in\Sigma is decorated by a unit tangent vector in the tensor product Tx+ΣnTxΣnT_{x_{+}}\Sigma^{n}\otimes T_{x_{-}}\Sigma^{n} where ΣnΣ\Sigma^{n}\rightarrow\Sigma is the normalization of the nodal curve Σ\Sigma, and {x+,x}\{x_{+},x_{-}\} is the pre-image of the nodal point xx.

Then the set of non-degenerate kk-simplices is given by

𝕏k:=limG0GkM^G0Gk\mathbb{X}_{k}:=\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}

where the colimit is taken over the category of isomorphisms of kk-step stable graph contractions.

Next, we describe the boundary maps of 𝕏\mathbb{X}_{\bullet}. There are three types of boundary maps in 𝕏k𝕏k1\mathbb{X}_{k}\rightarrow\mathbb{X}_{k-1}:

  • The first type boundary map is a map of the form

    (10) δ1:M^G0Gk=M^GkM¯G0M^G0Gk1=M^Gk1M¯G0\delta_{1}:\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}}\;\rightarrow\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}}=\widehat{M}_{G_{k-1}}\mid_{\overline{M}_{G_{0}}}

    given by the natural projection map which forgets the decorations at nodes that got contracted in the map Gk1GkG_{k-1}\rightarrow G_{k}.

  • The second type boundary map is given by

    (11) δ2:M^G0Gk=M^GkM¯G0M^G1Gk=M^GkM¯G1,\delta_{2}:\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}}\;\rightarrow\widehat{M}_{G_{1}\rightarrow\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{1}}},

    simply the inclusion map induced by the inclusion M¯G0M¯G1\overline{M}_{G_{0}}\hookrightarrow\overline{M}_{G_{1}}.

  • The third type boundary map is defined in the case k2k\geq 2 and for each 1jk11\leq j\leq k-1, and is given by the identity map

    (12) δ3:M^G0Gj^Gk=M^GkM¯G0M^G0Gj^Gk=M^GkM¯G0,\delta_{3}:\widehat{M}_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}}\;\rightarrow\widehat{M}_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}},

Circle/Disk bundles.

For each stable graph contraction G0GkG_{0}\rightarrow\cdots\rightarrow G_{k}, we shall work with a modified but homotopy equivalent version of M^G0Gk\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}. Let GGG\rightarrow G^{\prime} be a graph contraction. Denote by NGGN_{G\rightarrow G^{\prime}} the normal bundle associated with the embedding M¯GM¯G\overline{M}_{G}\hookrightarrow\overline{M}_{G^{\prime}}. Its rank is equal to d=|EGG|d=|E_{G\rightarrow G^{\prime}}|, the number of contracted edges in GGG\rightarrow G^{\prime}. Moreover, we denote by SNGGSN_{G\rightarrow G^{\prime}} the associated (S1)d(S^{1})^{d}-bundle, and DNGGDN_{G\rightarrow G^{\prime}} the associated DdD^{d}-bundle. For GG\rightarrow\star with \star the unique stable graph in Γ((g,n))\Gamma((g,n)) with one vertex, we write SNGSN_{G} and DNGDN_{G} for the corresponding circle/disk bundles. As observed by Mondello [15] (see Equation 8), we have

M^G0Gk=M^GkM¯G0=bG0(SNG0Gk),\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}=\widehat{M}_{G_{k}}\mid_{\overline{M}_{G_{0}}}=b_{G_{0}}^{*}(SN_{G_{0}\rightarrow G_{k}}),

where bG0:M^G0M¯G0b_{G_{0}}:\widehat{M}_{G_{0}}\rightarrow\overline{M}_{G_{0}} is the canonical projection map by definition of the oriented real blowup construction.

We shall make two modifications to the above space, both of which are homotopy equivalent to it by construction:

  • The first modification is to add a disk bundle in the normal direction of M¯Gk\overline{M}_{G_{k}} inside M¯g,n\overline{M}_{g,n}. More precisely, we use

    bG0(SNG0Gk×M¯G0DNGk|M¯G0).b_{G_{0}}^{*}\big{(}SN_{G_{0}\rightarrow G_{k}}\times_{\overline{M}_{G_{0}}}DN_{G_{k}}|_{\overline{M}_{G_{0}}}\big{)}.

    Since disks are contractible, it is clear that this is homotopy equivalent to M^G0Gk\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}.

  • The second modification is to put framings on the marked points, then realizing the above space as its (S1)n(S^{1})^{n}-quotient. Since the circle action on framings is a free action, the resulting space is also homotopy equivalent to M^G0Gk\widehat{M}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}.

More formally, for each k0k\geq 0, we define

𝕐k:=limG0Gk𝕐G0Gk𝖿𝗋/(S1)n\mathbb{Y}_{k}:=\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\mathbb{Y}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}/(S^{1})^{n}

where an element of 𝕐G0Gk𝖿𝗋\mathbb{Y}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}} consists of a stable curve with marked points together with framings (Σ,p1,,pn,φ1,,φn)M¯G0𝖿𝗋(\Sigma,p_{1},\ldots,p_{n},\varphi_{1},\ldots,\varphi_{n})\in\overline{M}^{\sf fr}_{G_{0}}. This stable curve’s dual graph GG admits a contraction map GG0G\rightarrow G_{0}. For each edge that got contracted by the composition map GG0GkG\rightarrow G_{0}\rightarrow G_{k}, we decorate the corresponding node by a unit tangent vector in the tensor product Tx+ΣnTxΣnT_{x_{+}}\Sigma^{n}\otimes T_{x_{-}}\Sigma^{n}. For each of the remaining node yy of Σ\Sigma, we decorate it by a tangent vector in the unit disk of Ty+ΣnTyΣnT_{y_{+}}\Sigma^{n}\otimes T_{y_{-}}\Sigma^{n}.

Analogous to 𝕏\mathbb{X}_{\bullet}, there are three types of simplicial boundary maps on 𝕐\mathbb{Y}_{\bullet} so that it is compatible with the boundary map of 𝕏\mathbb{X}_{\bullet} defined in Equations (10) (11) (12). We shall abuse the notations δi(i=1,2,3)\delta_{i}\,(i=1,2,3) for the three types of boundary maps.

  • For the first type boundary map, we set

    δ1:𝕐G0Gk𝖿𝗋/(S1)n𝕐G0Gk1𝖿𝗋/(S1)n\delta_{1}:\mathbb{Y}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}^{\sf fr}/(S^{1})^{n}\rightarrow\mathbb{Y}_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}}^{\sf fr}/(S^{1})^{n}

    to be the natural map induced by the inclusion map S1𝔻S^{1}\subset\mathbb{D} at the nodes that are contracted by the morphism Gk1GkG_{k-1}\rightarrow G_{k}.

  • The boundary map of the second type

    δ2:𝕐G0Gk𝖿𝗋/(S1)n𝕐G1Gk𝖿𝗋/(S1)n\delta_{2}:\mathbb{Y}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}^{\sf fr}/(S^{1})^{n}\rightarrow\mathbb{Y}_{G_{1}\rightarrow\cdots\rightarrow G_{k}}^{\sf fr}/(S^{1})^{n}

    is again induced by the inclusion M^G0M^G1\widehat{M}_{G_{0}}\rightarrow\widehat{M}_{G_{1}} while keeping decorations at nodes and framings at marked points.

  • The third type boundary map is again the identity map:

    δ3=id:𝕐G0Gk𝖿𝗋/(S1)n𝕐G0Gj^Gk𝖿𝗋/(S1)n.\delta_{3}=\operatorname{id}:\mathbb{Y}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}^{\sf fr}/(S^{1})^{n}\rightarrow\mathbb{Y}_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k}}^{\sf fr}/(S^{1})^{n}.

Note that by construction we have a homotopy equivalence of simplicial spaces between 𝕏\mathbb{X}_{\bullet} and 𝕐\mathbb{Y}_{\bullet}. By Mondello’s theorem, we obtain a quasi-isomophism of chain complexes

C(M¯g,n)𝖳𝗈𝗍(C(𝕐))C_{*}(\overline{M}_{g,n})\cong{\sf Tot}\big{(}C_{*}(\mathbb{Y}_{\bullet})\big{)}

Simplicial resolution of the Feynman compactification.

To link the simplicial complex C(𝕐)C_{*}(\mathbb{Y}_{\bullet}) with the Feynman compactification F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n) we observe that there is a discrepancy at nodes. Indeed, let us consider a node xx corresponding to a contracted edge in the dual graph of a surface Σ\Sigma. In the case of 𝕐\mathbb{Y}_{\bullet}, a node xx is decorated by elements in the tensor product Tx+ΣnTxΣnT^{*}_{x_{+}}\Sigma^{n}\otimes T^{*}_{x_{-}}\Sigma^{n}, while for F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n) the two circle actions on Tx+T^{*}_{x_{+}} and TxT^{*}_{x_{-}} has already been quotiented out. The two constructions are related by the following {Lemma}  Let MM and NN be two complexes with a circle action. Then there is a canonical homotopy equivalence

(MN)S1MS1NS1𝕂[ϵ],(M\otimes N)_{S^{1}}\cong M_{S^{1}}\otimes N_{S^{1}}\otimes\mathbb{K}[\epsilon],

where on the left hand side we take the homotopy quotient of the off-diagonal action, i.e. with circle operator BMBNB_{M}-B_{N}, and on the right hand side, the differential is given by bM+u1BM+bN+u2BN+(u1+u2)ϵb_{M}+u_{1}B_{M}+b_{N}+u_{2}B_{N}+(u_{1}+u_{2})\epsilon.

Proof 3.1.

This is standard Koszul duality between 𝕂[ϵ]\mathbb{K}[\epsilon] and its Koszul dual coalgebra u1𝕂[u1]u^{-1}\mathbb{K}[u^{-1}]. Thus for any S1S^{1}-module QQ, there is a chain equivalence

((Q)S1𝕂[ϵ],bQ+uBQ+uϵ)Q.\big{(}(Q)_{S^{1}}\otimes\mathbb{K}[\epsilon],b_{Q}+uB_{Q}+u\epsilon\big{)}\cong Q.

Apply this formula to our setting to obtain

MS1NS1𝕂[ϵ]\displaystyle M_{S^{1}}\otimes N_{S^{1}}\otimes\mathbb{K}[\epsilon] (MN)S1×S1𝕂[ϵ]\displaystyle\cong(M\otimes N)_{S^{1}\times S^{1}}\otimes\mathbb{K}[\epsilon]
((MN)S1)S1𝕂[ϵ]\displaystyle\cong\big{(}(M\otimes N)_{S^{1}}\big{)}_{S^{1}}\otimes\mathbb{K}[\epsilon]
(MN)S1.\displaystyle\cong(M\otimes N)_{S^{1}}.

where in the second equivalence we split the product S1S^{1}-action using BMBNB_{M}-B_{N} and BM+BNB_{M}+B_{N}.

We proceed to define a simplicial resolution of F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n). Let G0GkG_{0}\rightarrow\cdots\rightarrow G_{k} be a kk-step contraction. Define

F𝕄^G0Gk𝖿𝗋:=vVG0𝕄^S1𝖿𝗋(g(v),𝖫𝖾𝗀(v))eEG0GkC(S1)eEG0GkcC(D).F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}:=\bigotimes_{v\in V_{G_{0}}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v))\otimes\bigotimes_{e\in E_{G_{0}\rightarrow G_{k}}}C_{*}(S^{1})\otimes\bigotimes_{e^{\prime}\in E^{c}_{G_{0}\rightarrow G_{k}}}C_{*}(D).

That is,

  • For each vertex we decorate it by homotopy (S1)|𝖫𝖾𝗀(v)|(S^{1})^{|{\sf Leg}(v)|}-quotient of 𝕄^𝖿𝗋(g(v),𝖫𝖾𝗀(v))\widehat{\mathbb{M}}^{\sf fr}(g(v),{\sf Leg}(v)).

  • For each contracted edge eEG0Gke\in E_{G_{0}\rightarrow G_{k}}, we decorate by the cellular chain complex of circle:

    C(S1):=𝕂[ϵ]=𝕂𝟏𝕂ϵC_{*}(S^{1}):=\mathbb{K}[\epsilon]=\mathbb{K}\cdot{\mathbf{1}}\oplus\mathbb{K}\cdot\epsilon

    with ϵ\epsilon of homological degree 11.

  • For each of the remaining edges in EG0GkcE_{G_{0}\rightarrow G_{k}}^{c}, we decorate by the cellular chain complex of the disk

    C(D):=𝕂𝟏𝕂ϵ𝕂D,C_{*}(D):=\mathbb{K}\cdot{\mathbf{1}}\oplus\mathbb{K}\cdot\epsilon\oplus\mathbb{K}\cdot D,

    with DD of homological degree 22. The differential acts on the 22-cell by (D)=ϵ\partial(D)=\epsilon.

The differential of F𝕄^G0Gk𝖿𝗋F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}} is not simply the tensor product differential. The extra differential comes from Lemma 3. Indeed, for every edge eEG0e\in E_{G_{0}} (no matter if it’s contracted by G0GkG_{0}\rightarrow G_{k} or not), the extra differential is given by

(α+𝟏eα)=(ue+α+)ϵeα+α+ϵe(ueα),\partial(\alpha_{+}\otimes{\mathbf{1}}_{e}\otimes\alpha_{-})=(u_{e_{+}}\alpha_{+})\otimes\epsilon_{e}\otimes\alpha_{-}+\alpha_{+}\otimes\epsilon_{e}\otimes(u_{e_{-}}\alpha_{-}),

with ue+u_{e_{+}} and ueu_{e_{-}} the two circle parameters acting on chains α+\alpha_{+} and α\alpha_{-} on the two ends of ee. For each k0k\geq 0, we define the set of kk-simplices by setting

F𝕄^𝖿𝗋(g,n)k:=limG0GkF𝕄^G0Gk𝖿𝗋F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{k}:=\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k}}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}

to obtain a simplicial chain complex F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}. Again, just like 𝕏\mathbb{X}_{\bullet} and 𝕐\mathbb{Y}_{\bullet}, it has three types of simplicial boundary maps:

  • The first type boundary map

    δ1:(F𝕄^𝖿𝗋)G0Gk(F𝕄^𝖿𝗋)G0Gk1\delta_{1}:(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}}

    is defined via the inclusion C(S1)C(D)C_{*}(S^{1})\hookrightarrow C_{*}(D) at the edges that are contracted by Gk1GkG_{k-1}\rightarrow G_{k}.

  • The second type boundary map

    δ2:(F𝕄^𝖿𝗋)G0Gk(F𝕄^𝖿𝗋)G1Gk\delta_{2}:(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow(F\widehat{\mathbb{M}}^{\sf fr})_{G_{1}\rightarrow\cdots\rightarrow G_{k}}

    is defined by applying Lemma 3 at the edges that are contracted by the map G0G1G_{0}\rightarrow G_{1}.

  • The third simplicial boundary map

    δ3:(F𝕄^𝖿𝗋)G0Gk(F𝕄^𝖿𝗋)G0Gj^Gk\delta_{3}:(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k}}

    is again the identity map.

In conclusion, for each pair (g,n)(g,n) such that 2g2+n>02g-2+n>0, we have defined a simplicial chain complex F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}. Denote its total complex by

𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n)):=(k0limG0Gk(F𝕄^𝖿𝗋)G0Gk[k],+δ).{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}:=\big{(}\bigoplus_{k\geq 0}\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k}}(F\widehat{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}[k],\partial+\delta\big{)}.

Mondello [15]’s simplicial resolution construction implies the following

{Theorem}

There is a homotopy equivalence of chain complexes

C(M¯g,n)𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n)).C_{*}(\overline{M}_{g,n})\cong{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}.

Homotopy colimit of F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}.

When k=0k=0, we have

F𝕄^𝖿𝗋(g,n)0=limGΓ((g,n))vVG𝕄^S1𝖿𝗋(g(v),𝖫𝖾𝗀(v))eEGC(D).F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{0}=\varinjlim_{G\in\Gamma((g,n))}\bigotimes_{v\in V_{G}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v))\otimes\bigotimes_{e\in E_{G}}C_{*}(D).

On the other hand the Feynman compactification is given by

F𝕄^𝖿𝗋(g,n)=limGΓ((g,n))vVG𝕄^S1𝖿𝗋(g(v),𝖫𝖾𝗀(v))eEGDe.F\widehat{\mathbb{M}}^{\sf fr}(g,n)=\varinjlim_{G\in\Gamma((g,n))}\bigotimes_{v\in V_{G}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v))\otimes\bigotimes_{e\in E_{G}}D_{e}\cdot\mathbb{Q}.

The latter space is clearly a subspace of the former. However, observe that neither the canonical inclusion map nor the canonical projection map is a map of complexes.

In the following proposition, we prove that the homotopy colimit of the simplicial chain complex F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet} is given by the Feynman compactification F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n).

{Proposition}

There exists a quasi-isomorphism of chain complexes

π:𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))F𝕄^𝖿𝗋(g,n)\pi^{\sharp}:{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\rightarrow F\widehat{\mathbb{M}}^{\sf fr}(g,n)

depicted as

F𝕄^𝖿𝗋(g,n)0δF𝕄^𝖿𝗋(g,n)1δπF𝕄^𝖿𝗋(g,n)\begin{CD}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{0}@<{\delta}<{}<F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{1}@<{\delta}<{}<\cdots\\ @V{\pi^{\sharp}}V{}V\\ F\widehat{\mathbb{M}}^{\sf fr}(g,n)\end{CD}
Proof 3.2.

Let us fix a stable graph G0Γ((g,n))kG_{0}\in\Gamma((g,n))_{k} and consider the subspace of the total complex given by

F𝕄^G0𝖿𝗋δ1+δ3limG0G1F𝕄^G0G1𝖿𝗋δ1+δ3limG0G1G2F𝕄^G0G1G2𝖿𝗋δ1+δ3\begin{CD}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}}@<{\delta_{1}+\delta_{3}}<{}<\displaystyle\varinjlim_{G_{0}\rightarrow G_{1}}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow G_{1}}@<{\delta_{1}+\delta_{3}}<{}<\displaystyle\varinjlim_{G_{0}\rightarrow G_{1}\rightarrow G_{2}}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow G_{1}\rightarrow G_{2}}@<{\delta_{1}+\delta_{3}}<{}<\cdots\end{CD}

which is endowed with the differential δ1+δ3\delta_{1}+\delta_{3}. The natural inclusion map

i:F𝕄^𝖿𝗋(G0)F𝕄^G0𝖿𝗋i:F\widehat{\mathbb{M}}^{\sf fr}(G_{0})\hookrightarrow F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}}

which decorates each edge ee by the disk cell DeD_{e} splits the projection map

π:F𝕄^G0𝖿𝗋F𝕄^𝖿𝗋(G0)\pi:F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}}\rightarrow F\widehat{\mathbb{M}}^{\sf fr}(G_{0})

induced by the projection map C(D)DeC_{*}(D)\rightarrow D_{e}\cdot\mathbb{Q} at every edges of G0G_{0}. We first construct a homotopy operator

h:limG0GkF𝕄^G0Gk𝖿𝗋limG0Gk+1F𝕄^G0Gk+1𝖿𝗋h:\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k}}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow\varinjlim_{G_{0}\rightarrow\cdots\rightarrow G_{k+1}}F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k+1}}

such that it satisfies the deformation retract identity idiπ=[δ1+δ3,h]\operatorname{id}-i\pi=[\delta_{1}+\delta_{3},h]. To construct such an hh, consider an element

αF𝕄^G0Gk𝖿𝗋=vVG0𝕄^S1𝖿𝗋(g(v),𝖫𝖾𝗀(v))eEG0GkC(S1)eEG0GkcC(D).\alpha\in F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}=\bigotimes_{v\in V_{G_{0}}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v))\otimes\bigotimes_{e\in E_{G_{0}\rightarrow G_{k}}}C_{*}(S^{1})\otimes\bigotimes_{e^{\prime}\in E^{c}_{G_{0}\rightarrow G_{k}}}C_{*}(D).

Let e1,,ele^{\prime}_{1},\ldots,e^{\prime}_{l} be edges in EG0GkcE^{c}_{G_{0}\rightarrow G_{k}} such that in α\alpha its decoration is in C(S1)C_{*}(S^{1}). Then we set

h(α)=αF𝕄^G0GkGk+1𝖿𝗋h(\alpha)=\alpha\in F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}\rightarrow G_{k+1}}

where Gk+1G_{k+1} is obtained from GkG_{k} by contracting the edges e1,,ele^{\prime}_{1},\ldots,e^{\prime}_{l}. Let us check the identity [h,δ1+δ3]=idiπ[h,\delta_{1}+\delta_{3}]=\operatorname{id}-i\pi. Indeed, if the set {e1,,el}\{e_{1}^{\prime},\ldots,e_{l}^{\prime}\} is not empty, we have α=δ1h(α)\alpha=\delta_{1}h(\alpha) and δ3h(α)=j=1k(α)G0Gj^Gk+1\delta_{3}h(\alpha)=\sum_{j=1}^{k}(\alpha)_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k+1}}. While in the other direction, we have

hδ1(α)\displaystyle h\delta_{1}(\alpha) =h((α)G0Gk1)=(α)G0Gk1Gk+1\displaystyle=h\big{(}(\alpha)_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}}\big{)}=(\alpha)_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}\rightarrow G_{k+1}}
hδ3(α)\displaystyle h\delta_{3}(\alpha) =j=1k1(α)G0Gj^GkGk+1\displaystyle=\sum_{j=1}^{k-1}(\alpha)_{G_{0}\rightarrow\cdots\widehat{G_{j}}\cdots\rightarrow G_{k}\rightarrow G_{k+1}}

This shows that [δ1+δ3,h]=id[\delta_{1}+\delta_{3},h]=\operatorname{id} if l>0l>0. Since if l>0l>0, we have iπ(α)=0i\pi(\alpha)=0, thus the previous identity is equivalent to the homotopy retract identity. In the case l=0l=0, we have h(α)=0h(\alpha)=0, and that

hδ1(α)\displaystyle h\delta_{1}(\alpha) ={h((α)G0Gk1)=(α)G0Gk1Gk=αk10k=0\displaystyle=\begin{cases}h\big{(}(\alpha)_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}}\big{)}=(\alpha)_{G_{0}\rightarrow\cdots\rightarrow G_{k-1}\rightarrow G_{k}}=\alpha&k\geq 1\\ 0&k=0\end{cases}
hδ3(α)\displaystyle h\delta_{3}(\alpha) =0\displaystyle=0

This also implies the desired identity [h,δ1+δ3]=idiπ[h,\delta_{1}+\delta_{3}]=\operatorname{id}-i\pi.

Finally, we evoke the differential +δ2\partial+\delta_{2} and use homological perturbation formula to compute the induced differential on the Feynman compactification F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n) which matches the definition given in Paragraph 2. Indeed, the perturbed differential turns out to have two terms:

πi+πδ2hi\pi\partial i+\pi\delta_{2}h\partial i

The first term corresponds to the boundary map on vertices of stable graphs in the Feynman compactification F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n). The extra differential δ\delta of F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n) (defined in Equation 6) applied to edges of stable graphs is precisely the second term, i.e.

δ=πδ2hi.\delta=\pi\delta_{2}h\partial i.

This is a calculation which we illustrate in the case when the underlying stable graph has only one edge ee joining two vertices decorated by α1\alpha_{1} and α2\alpha_{2}:

πδ2hi(Deα1α2)\displaystyle\pi\delta_{2}h\partial i(D_{e}\otimes\alpha_{1}\otimes\alpha_{2}) =πδ2h(Deα1α2)\displaystyle=\pi\delta_{2}h\partial(D_{e}\otimes\alpha_{1}\otimes\alpha_{2})
=πδ2(ϵeα1α2)\displaystyle=\pi\delta_{2}(\epsilon_{e}\otimes\alpha_{1}\otimes\alpha_{2})
=ρe(α1α2)\displaystyle=\rho_{e}(\alpha_{1}\otimes\alpha_{2})

This agrees with the formula of δ\delta in Equation 6, which finishes the proof.

W

e write down the perturbed inclusion and projection maps which will be used later in the paper.

(13) i=k0(h)kiπ=π+πδ2h\displaystyle~{}\begin{split}i^{\sharp}&=\sum_{k\geq 0}(h\partial)^{k}i\\ \pi^{\sharp}&=\pi+\pi\delta_{2}h\end{split}

Completely parallel to the construction of F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}, one may also define a simplicial resolution F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n)_{\bullet} for the Feynman compactification F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n). We also have the quasi-isomorphisms:

i\displaystyle i^{\sharp} :F𝕄𝖿𝗋(g,n)𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))\displaystyle:F\mathbb{M}^{\sf fr}(g,n)\rightarrow{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}
π\displaystyle\pi^{\sharp} :𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))F𝕄𝖿𝗋(g,n)\displaystyle:{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\rightarrow F\mathbb{M}^{\sf fr}(g,n)

4 Comparison with the Deligne-Mumford operad

In this section, we prove the following theorem.

{Theorem}

There is an isomorphism of modular operads

H(F𝕄𝖿𝗋)H(M¯,)H_{*}(F\mathbb{M}^{\sf fr})\cong H_{*}(\overline{{M}},\mathbb{Q})

where M¯\overline{{M}} is the Deligne-Mumford modular operad [10, Section 6.2].

The proof of this theorem occupies the rest of the section. We briefly outline the main idea here. By results of the previous section, we have isomorphisms:

H(F𝕄𝖿𝗋(g,n))\displaystyle H_{*}\big{(}F\mathbb{M}^{\sf fr}(g,n)\big{)} H(𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n)))\displaystyle\cong H_{*}\Big{(}{\sf Tot}\big{(}F\mathbb{M}^{\sf fr}(g,n)_{\bullet}\big{)}\Big{)}
H(F𝕄^𝖿𝗋(g,n))\displaystyle H_{*}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)\big{)} H(𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n)))H(M¯g,n)\displaystyle\cong H_{*}\Big{(}{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\Big{)}\cong H_{*}(\overline{M}_{g,n})

Thus in order to relate the first line with the second line, we shall construct a quasi-isomorphism

𝕀:𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))\mathbb{I}:{\sf Tot}\big{(}F\mathbb{M}^{\sf fr}(g,n)_{\bullet}\big{)}\cong{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}

Combining with the previous isomorphisms yields the isomorphism in Theorem 4. One then checks the compatibility of the modular operad structures.

Comparison between F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet} and F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n)_{\bullet}.

As was pointed in Paragraph 2, the inclusion map

Ig,n:𝕄𝖿𝗋(g,n)𝕄^𝖿𝗋(g,n)I_{g,n}:\mathbb{M}^{\sf fr}(g,n)\hookrightarrow\widehat{\mathbb{M}}^{\sf fr}(g,n)

does not respect the operadic composition map. This implies that the induced inclusion map between F𝕄^𝖿𝗋(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet} and F𝕄𝖿𝗋(g,n)F\mathbb{M}^{\sf fr}(g,n)_{\bullet} does not respect the simplicial structure. More precisely, the simplicial boundary map δ2\delta_{2} is not compatible with the inclusion map. However, the failure of this compatibility is actually homotopically trivial. Consider the simplest situation with h:Gg,nh:G\rightarrow\star_{g,n} with GG a stable graph with two vertices v1v_{1}, v2v_{2} and an unique edge ee. Consider the following (non-commutative) diagram:

Mg(v1),𝖫𝖾𝗀(v1)𝖿𝗋×Mg(v2),𝖫𝖾𝗀(v2)𝖿𝗋ηhMg,n𝖿𝗋IIM^g(v1),𝖫𝖾𝗀(v1)𝖿𝗋×M^g(v2),𝖫𝖾𝗀(v2)𝖿𝗋ξhM^g,n𝖿𝗋\begin{CD}M_{g(v_{1}),{\mathsf{Leg}}(v_{1})}^{\sf fr}\times M_{g(v_{2}),{\mathsf{Leg}}(v_{2})}^{\sf fr}@>{\eta_{h}}>{}>M_{g,n}^{\sf fr}\\ @V{I}V{}V@V{}V{I}V\\ \widehat{M}_{g(v_{1}),{\mathsf{Leg}}(v_{1})}^{\sf fr}\times\widehat{M}_{g(v_{2}),{\mathsf{Leg}}(v_{2})}^{\sf fr}@>{\xi_{h}}>{}>\widehat{M}_{g,n}^{\sf fr}\end{CD}

The above diagram is commutative up to homotopy. Indeed, let φ1:DzU(x)\varphi_{1}:D_{z}\rightarrow U(x) and φ2:DwU(y)\varphi_{2}:D_{w}\rightarrow U(y) be the two framings to be sewed. Fix 0<t10<t\leq 1, we may scale the two framings by setting

φ1t(z):=φ1(tz),φ2t(w):=φ2(tw).\varphi_{1}^{t}(z):=\varphi_{1}(t\cdot z),\;\;\;\varphi_{2}^{t}(w):=\varphi_{2}(t\cdot w).

The scaled framings are given by φ1t:Dzφ1(tD)U(x)\varphi_{1}^{t}:D_{z}\rightarrow\varphi_{1}(t\cdot D)\subset U(x) and φ2t:Dwφ1(tD)U(x)\varphi_{2}^{t}:D_{w}\rightarrow\varphi_{1}(t\cdot D)\subset U(x). Using the scaled framings we obtain a tt-dependent sewing map

ηht:Mg(v1),𝖫𝖾𝗀(v1)𝖿𝗋×Mg(v2),𝖫𝖾𝗀(v2)𝖿𝗋Mg,n𝖿𝗋\eta_{h}^{t}:M_{g(v_{1}),{\mathsf{Leg}}(v_{1})}^{\sf fr}\times M_{g(v_{2}),{\mathsf{Leg}}(v_{2})}^{\sf fr}\rightarrow M_{g,n}^{\sf fr}

It is clear that ηh1=ηh\eta_{h}^{1}=\eta_{h}.

{Lemma}

Let the notations be as above. Then we have

limt0Iηht=ξhI.\lim_{t\rightarrow 0}I\circ\eta_{h}^{t}=\xi_{h}\circ I.
Proof 4.1.

It is clear that on the underlying stable curve, as t0t\rightarrow 0 the limiting curve becomes a nodal curve with the node given by the image of xx and yy. The limiting decoration at this node is then given by the unit tangent direction vector of

dφ1t(ddz)dφ2t(ddw)=t2dφ1(ddz)dφ2(ddw).d\varphi_{1}^{t}(\frac{d}{dz})\otimes d\varphi_{2}^{t}(\frac{d}{dw})=t^{2}d\varphi_{1}(\frac{d}{dz})\otimes d\varphi_{2}(\frac{d}{dw}).

Thus the unit tangent direction is independent of tt, and thus have a limiting vector given by the unit tangent direction vector of dφ1(ddz)dφ2(ddw)d\varphi_{1}(\frac{d}{dz})\otimes d\varphi_{2}(\frac{d}{dw}), which is precisely the definition of ξhI\xi_{h}\circ I.

Using the geometric homotopy ηht\eta_{h}^{t}, we obtain a homotopy operator of homological degree one:

𝕀1:𝕄𝖿𝗋(G)𝕄^𝖿𝗋(g,n).\mathbb{I}_{1}:\mathbb{M}^{\sf fr}(G)\rightarrow\widehat{\mathbb{M}}^{\sf fr}(\star_{g,n}).

The lemma above implies that [,𝕀1]=IηhξhI[\partial,\mathbb{I}_{1}]=I\circ\eta_{h}-\xi_{h}\circ I.

In general, for each kk-step contraction, consider the following diagram:

F𝕄G0Gk𝖿𝗋δ2F𝕄G1Gk𝖿𝗋IIF𝕄^G0Gk𝖿𝗋δ2F𝕄^G1Gk𝖿𝗋\begin{CD}F{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}@>{\delta_{2}}>{}>F{\mathbb{M}}^{\sf fr}_{G_{1}\rightarrow\cdots\rightarrow G_{k}}\\ @V{I}V{}V@V{I}V{}V\\ F\widehat{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}@>{\delta_{2}}>{}>F\widehat{\mathbb{M}}^{\sf fr}_{G_{1}\rightarrow\cdots\rightarrow G_{k}}\end{CD}

We may apply the previous construction to all the edges in EG0G1E_{G_{0}\rightarrow G_{1}} with the same parameter t[0,1]t\in[0,1] to yield a homotopy operator still denoted by

𝕀1:F𝕄G0Gk𝖿𝗋F𝕄^G1Gk𝖿𝗋\mathbb{I}_{1}:F{\mathbb{M}}^{\sf fr}_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow F\widehat{\mathbb{M}}^{\sf fr}_{G_{1}\rightarrow\cdots\rightarrow G_{k}}

of homological degree one such that [,𝕀1]=Iδ2δ2I[\partial,\mathbb{I}_{1}]=I\circ\delta_{2}-\delta_{2}\circ I. Using these homotopy operators and its higher extensions (introduced in the proof below), we prove the following

{Theorem}

There exists a quasi-isomorphism of chain complexes

𝕀:𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))\mathbb{I}:{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\rightarrow{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}
Proof 4.2.

We inductively construct a morphism of chain complexes of the form

𝕀=d0𝕀d:𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))\mathbb{I}=\sum_{d\geq 0}\mathbb{I}_{d}:{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\cong{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}

with 𝕀0:=I\mathbb{I}_{0}:=I. Recall that the total differential on both total complexes is of the form +δ\partial+\delta with \partial the internal differential, and δ=δ1+δ2+δ3\delta=\delta_{1}+\delta_{2}+\delta_{3} the simplicial boundary maps. From the definition, it is easy to verify that [δ1,𝕀0]=[δ3,𝕀0]=0[\delta_{1},\mathbb{I}_{0}]=[\delta_{3},\mathbb{I}_{0}]=0. But the boundary map δ2\delta_{2} is not compatible with 𝕀0\mathbb{I}_{0}. Its commutator is precisely bounded by the homotopy 𝕀1\mathbb{I}_{1} defined in the previous paragraph.

In general, for each d1d\geq 1, we shall define a map of homological degree dd of the form

𝕀d:(F𝕄𝖿𝗋)G0Gk(F𝕄^𝖿𝗋)GdGk\mathbb{I}_{d}:(F{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow(F\widehat{\mathbb{M}}^{\sf fr})_{G_{d}\rightarrow\cdots\rightarrow G_{k}}

such that

(14) [,𝕀d+1]=[δ2+δ3,𝕀d],[δ1,𝕀d]=0\displaystyle~{}\begin{split}[\partial,\mathbb{I}_{d+1}]&=-[\delta_{2}+\delta_{3},\mathbb{I}_{d}],\\ [\delta_{1},\mathbb{I}_{d}]&=0\end{split}

Indeed, associated with a dd-step stable graph contraction G0GdG_{0}\rightarrow\cdots\rightarrow G_{d} we have a family of sewing maps depending on dd-variables t1,,tdt_{1},\cdots,t_{d}

ηG0Gdt1,,td:F𝕄𝖿𝗋(G0)F𝕄^𝖿𝗋(Gd),  0t1td1,\eta_{G_{0}\rightarrow\cdots\rightarrow G_{d}}^{t_{1},\cdots,t_{d}}:F{\mathbb{M}}^{\sf fr}(G_{0})\rightarrow F\widehat{\mathbb{M}}^{\sf fr}(G_{d}),\;\;0\leq t_{1}\leq\ldots\leq t_{d}\leq 1,

where we use the tit_{i}-variable to scale the sewing operation at an edge that got contracted by Gi1GiG_{i-1}\rightarrow G_{i}. See Figure (3) for illustrations of this homotopy maps.

Refer to caption
Figure 3: Higher homotopies

We use this dd-parameter family to induce a higher homotopy operator:

𝕀d:(F𝕄𝖿𝗋)G0Gk(F𝕄^𝖿𝗋)GdGk.\mathbb{I}_{d}:(F{\mathbb{M}}^{\sf fr})_{G_{0}\rightarrow\cdots\rightarrow G_{k}}\rightarrow(F\widehat{\mathbb{M}}^{\sf fr})_{G_{d}\rightarrow\cdots\rightarrow G_{k}}.

To see the above identities (14) hold, observe that the facets defined by t1=0t_{1}=0 and td=1t_{d}=1 corresponds to [δ2,𝕀d1]-[\delta_{2},\mathbb{I}_{d-1}], the intermediate facets defined by ti=ti+1, 1id1t_{i}=t_{i+1},\;1\leq i\leq d-1 are 𝕀d1δ3=[δ3,𝕀d1]\mathbb{I}_{d-1}\delta_{3}=-[\delta_{3},\mathbb{I}_{d-1}]. Here we used that δ3𝕀d1=0\delta_{3}\mathbb{I}_{d-1}=0 since the image of 𝕀d1\mathbb{I}_{d-1} lies in the index Gd1GdG_{d-1}\rightarrow G_{d} while δ3\delta_{3} vanishes on such type of component. Putting the maps {𝕀d}d0\{\mathbb{I}_{d}\}_{d\geq 0} together, we obtain a map of chain complexes

𝕀:=d0𝕀d:𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n)).\mathbb{I}:=\sum_{d\geq 0}\mathbb{I}_{d}:{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\rightarrow{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}.

To argue that 𝕀\mathbb{I} is a quasi-isomorphism, we consider the increasing filtration on both sides defined by that the kk-th filtration consists of elements of simplicial degree less than or equal to kk. Observe that 𝕀\mathbb{I} preserves the filtration. Thus it induces a morphism of the associated spectral sequences on the two total complexes. But we already have an isomorphism in the E1E^{1}-page since the map 𝕀0\mathbb{I}_{0} is induced by the natural inclusion maps Mg,n𝖿𝗋M^g,n𝖿𝗋M_{g,n}^{\sf fr}\hookrightarrow\widehat{M}_{g,n}^{\sf fr} which are homotopy equivalences.

Operadic compositions.

To summarize the previous discussions, we have the following diagram

𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝕀𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))@ AiAAπF𝕄𝖿𝗋(g,n)F𝕄^𝖿𝗋(g,n)\begin{CD}{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}@>{\mathbb{I}}>{}>{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\\ @ Ai^{\sharp}AA@V{\pi^{\sharp}}V{}V\\ F{\mathbb{M}}^{\sf fr}(g,n)F\widehat{\mathbb{M}}^{\sf fr}(g,n)\end{CD}

where all the arrows are quasi-isomorphisms. To prove analyze the operadic compositions in Theorem 4, we shall define a canonical projection map

p:𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))C(M¯g,n).p:{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}\rightarrow C_{*}(\overline{M}_{g,n}).

This map is only non-zero at simplicial degree zero, in which case we need to define a map

p:(F𝕄^𝖿𝗋)GC(M¯g,n)p:(F\widehat{\mathbb{M}}^{\sf fr})_{G}\rightarrow C_{*}(\overline{M}_{g,n})

for each stable graph GΓ((g,n))G\in\Gamma((g,n)). Let α(F𝕄^𝖿𝗋)G\alpha\in(F\widehat{\mathbb{M}}^{\sf fr})_{G} with its underlying stable graph GG. Explicitly, such an element is of the form

α=eEGαevVGαveEGC(De)vVGC(M^g(v),n(v)𝖿𝗋)[u11,,un(v)1]\alpha=\prod_{e\in E_{G}}\alpha_{e}\otimes\prod_{v\in V_{G}}\alpha_{v}\in\prod_{e\in E_{G}}C_{*}(D_{e})\otimes\prod_{v\in V_{G}}C_{*}(\widehat{M}_{g(v),n(v)}^{\sf fr})[u_{1}^{-1},\ldots,u_{n(v)}^{-1}]

At a vertex vVGv\in V_{G}, its vertex decoration is of the form

αv=αv[0]+O(u1),\alpha_{v}=\alpha_{v}^{[0]}+O(u^{-1}),

with αv[0]C(Mg(v),n(v)𝖿𝗋)\alpha_{v}^{[0]}\in C_{*}(M_{g(v),n(v)}^{\sf fr}) the constant term in αvC(M^g(v),n(v)𝖿𝗋)[u11,,un(v)1]\alpha_{v}\in C_{*}(\widehat{M}_{g(v),n(v)}^{\sf fr})[u_{1}^{-1},\ldots,u_{n(v)}^{-1}]. Let us denote by

pv:C(M^g(v),n(v)𝖿𝗋)[u11,,un(v)1]C(M¯g(v),n(v))p_{v}:C_{*}(\widehat{M}_{g(v),n(v)}^{\sf fr})[u_{1}^{-1},\ldots,u_{n(v)}^{-1}]\rightarrow C_{*}(\overline{M}_{g(v),n(v)})

the map obtained by first projection onto the constant term, followed by the forgetful map 𝖥𝗈𝗋𝗀𝖾𝗍:M^g(v),n(v)𝖿𝗋M¯g(v),n(v){\sf Forget}:\widehat{M}_{g(v),n(v)}^{\sf fr}\rightarrow\overline{M}_{g(v),n(v)} which forgets both the framings and the nodal decorations. For each edge eEGe\in E_{G}, denote by

pe:C(De)p_{e}:C_{*}(D_{e})\rightarrow\mathbb{Q}

the projection map onto C0(D)=C_{0}(D)=\mathbb{Q}. Then the map pp is defined by

p(α)=ξG(eEGpe(αe)vVGpv(αv))p(\alpha)=\xi_{G}\big{(}\prod_{e\in E_{G}}p_{e}(\alpha_{e})\otimes\prod_{v\in V_{G}}p_{v}(\alpha_{v})\big{)}

where ξG\xi_{G} is the sewing map in Equation (9). Using the map pp, we consider the composition

p𝕀i:F𝕄𝖿𝗋(g,n)C(M¯g,n).p\mathbb{I}i^{\sharp}:F{\mathbb{M}}^{\sf fr}(g,n)\rightarrow C_{*}(\overline{M}_{g,n}).
{Proposition}

The induced map on homology

p𝕀i:H(F𝕄𝖿𝗋)H(M¯)p\mathbb{I}i^{\sharp}:H_{*}(F{\mathbb{M}}^{\sf fr})\rightarrow H_{*}(\overline{M})

is a morphism of modular operads.

Proof 4.3.

We shall only deal with the first type composition corresponding to a loop contraction cijc_{ij}. The case of non-loop contraction is analogous. In the loop case, we want to prove that the following diagram is commutative up to homotopy:

F𝕄𝖿𝗋(g,n+2)F𝕄𝖿𝗋(cij)F𝕄𝖿𝗋(g+1,n)p𝕀ip𝕀iC(M¯g,n+2)C(M¯)(cij)C(M¯g+1,n)\begin{CD}F{\mathbb{M}}^{\sf fr}(g,n+2)@>{F{\mathbb{M}}^{\sf fr}(c_{ij})}>{}>F{\mathbb{M}}^{\sf fr}(g+1,n)\\ @V{p\mathbb{I}i^{\sharp}}V{}V@V{p\mathbb{I}i^{\sharp}}V{}V\\ C_{*}(\overline{M}_{g,n+2})@>{C_{*}(\overline{M})(c_{ij})}>{}>C_{*}(\overline{M}_{g+1,n})\end{CD}

Let αF𝕄𝖿𝗋(g,n+2)\alpha\in F{\mathbb{M}}^{\sf fr}(g,n+2) be an element with its underlying stable graph GΓ((g,n+2))G\in\Gamma((g,n+2)). Let us explicitly write down the composition

p𝕀i:F𝕄𝖿𝗋(g,n)C(M¯g,n)p\mathbb{I}i^{\sharp}:F\mathbb{M}^{\sf fr}(g,n)\rightarrow C_{*}(\overline{M}_{g,n})

using Equations of pp, 𝕀\mathbb{I} and Equation (13). We have

𝕀i(α)\displaystyle\mathbb{I}i^{\sharp}(\alpha) =k0(e1,,ek)𝕀k(j=1kϵejeEG{e1,,ek}DevVGαv)GG/e1G/(e1,,ek)\displaystyle=\sum_{k\geq 0}\sum_{(e_{1},\ldots,e_{k})}\mathbb{I}_{k}\big{(}\prod_{j=1}^{k}\epsilon_{e_{j}}\otimes\prod_{e\in E_{G}-\{e_{1},\ldots,e_{k}\}}D_{e}\otimes\prod_{v\in V_{G}}\alpha_{v}\big{)}_{G\rightarrow G/e_{1}\rightarrow\cdots\rightarrow G/(e_{1},\ldots,e_{k})}

Since pep_{e} vanishes on DeD_{e}, we have

p𝕀i(α)=(e1,,eK)𝕀K(eEGϵevVGαv)GG/e1G/(e1,,eK)p\mathbb{I}i^{\sharp}(\alpha)=\sum_{(e_{1},\ldots,e_{K})}\mathbb{I}_{K}\big{(}\prod_{e\in E_{G}}\epsilon_{e}\otimes\prod_{v\in V_{G}}\alpha_{v}\big{)}_{G\rightarrow G/e_{1}\rightarrow\cdots\rightarrow G/(e_{1},\ldots,e_{K})}

where the summation is over orderings of the edge set EGE_{G}, and K=|EG|K=|E_{G}|. Observe that in the definition of 𝕀K\mathbb{I}_{K}, we required the sewing parameters t1,,tKt_{1},\ldots,t_{K} be ordered as 0t1,tK10\leq t_{1}\leq\ldots,\leq t_{K}\leq 1. Summing over ordering of the edges gets rid of this ordering condition. Thus, the above expression can be rewritten as

(15) p𝕀i(α)=𝖥𝗈𝗋𝗀𝖾𝗍(ηG([0,1]××[0,1]K copies×(eEGBe)(vVGαg(v),n(v)[0])))~{}p\mathbb{I}i^{\sharp}(\alpha)={\sf Forget}\Big{(}\eta_{G}\big{(}\underbrace{[0,1]\times\cdots\times[0,1]}_{\mbox{$K$ copies}}\times(\prod_{e\in E_{G}}B_{e})(\prod_{v\in V_{G}}\alpha_{g(v),n(v)}^{[0]})\big{)}\Big{)}

with

ηG:[0,1]××[0,1]K copies×vVGMg(v),n(v)𝖿𝗋M^g,n𝖿𝗋\eta_{G}:\underbrace{[0,1]\times\cdots\times[0,1]}_{\mbox{$K$ copies}}\times\prod_{v\in V_{G}}M^{\sf fr}_{g(v),n(v)}\rightarrow\widehat{M}_{g,n}^{\sf fr}

the sewing map depending on KK parameters as in the proof of Theorem 4. Intuitively speaking, for each eEGe\in E_{G}, we act on the chain vVGαg(v),n(v)[0]\prod_{v\in V_{G}}\alpha_{g(v),n(v)}^{[0]} by the circle action BeB_{e} and also sewing with the radius parameter te[0,1]t_{e}\in[0,1]. The forgetful map precisely collapses the circle at te=0t_{e}=0, which forms a disk [0,1]×S1/0×S1[0,1]\times S^{1}/{0}\times S^{1} over each edge eEGe\in E_{G}.

To this end, we shall finish the proof using the explicit formula of p𝕀ip\mathbb{I}i^{\sharp} above. Indeed, let αF𝕄𝖿𝗋(g,n+2)\alpha\in F{\mathbb{M}}^{\sf fr}(g,n+2) be such that at the half-edges labeled by ii and jj the uu-powers are ukiu^{-k_{i}} and ukju^{-k_{j}}. If either kik_{i} or kjk_{j} is strictly positive, the resulting composition is trivial in homology. In the interesting case when ki=kj=0k_{i}=k_{j}=0, denote by GijΓ(g+1,n)G_{ij}\in\Gamma(g+1,n) the graph obtained from GG by sewing the two leaves indexed by ii and jj. Then we see that the two compositions p𝕀iF𝕄𝖿𝗋(cij)(α)p\mathbb{I}i^{\sharp}F\mathbb{M}^{\sf fr}(c_{ij})(\alpha) and C(M¯)(cij)p𝕀i(α)C_{*}(\overline{M})(c_{ij})p\mathbb{I}i^{\sharp}(\alpha) are related by the homotopy

ηGij:[0,1]××[0,1]K+1 copies×vVGMg(v),n(v)𝖿𝗋M^g,n𝖿𝗋\eta_{G_{ij}}:\underbrace{[0,1]\times\cdots\times[0,1]}_{\mbox{$K+1$ copies}}\times\prod_{v\in V_{G}}M^{\sf fr}_{g(v),n(v)}\rightarrow\widehat{M}_{g,n}^{\sf fr}

This finishes the proof of proposition, which combined with Theorem 4 yields Theorem 4.

5 String vertices and fundamental classes

In this section, we use the Feynman compactification construction to provide a geometric extension of the algebraic trivialization of circle actions in [3]. Roughly speaking, we construct a compactified version denoted by 𝔤¯\overline{\mathfrak{g}} of Sen-Zwiebach’s DGLA 𝔤\mathfrak{g} (reviewed in the beginning of the section) so that we have an inclusion of DGLA’s

j:𝔤𝔤¯.j:\mathfrak{g}\rightarrow\overline{\mathfrak{g}}.

Then we construct an LL_{\infty} quasi-isomorphism 𝒦:𝔤¯𝔤¯𝗍𝗋𝗂𝗏\mathcal{K}:\overline{\mathfrak{g}}\rightarrow\overline{\mathfrak{g}}^{\sf triv} where the DGLA 𝔤¯𝗍𝗋𝗂𝗏\overline{\mathfrak{g}}^{\sf triv} has the same underlying space as 𝔤¯\overline{\mathfrak{g}}, but its BV differential and the Lie bracket are both zero. We prove the identity (3) stated in the introduction that the push-forward of the string vertex 𝒦j𝒱\mathcal{K}_{*}j_{*}\mathcal{V} consists of the fundamental class of [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}].

DGLA’s.

Recall the construction of Sen-Zwiebach’s DGLA:

𝔤:=((g,n)𝕄𝖿𝗋(g,n)Sn(S1)n[1])[[,λ]]\mathfrak{g}:=\big{(}\bigoplus_{(g,n)}\mathbb{M}^{\sf fr}(g,n)_{S_{n}\ltimes(S^{1})^{n}}[1]\big{)}[[\hbar,\lambda]]

with two formal variables \hbar and λ\lambda both of homological degree 2-2. By construction, its differential is +uB+Δ\partial+uB+\hbar\Delta with Δ\Delta defined using the twisted self-sewing map (4), and its Lie bracket is {,}\{-,-\} defined using the twisted sewing map (5). We refer to [2, Section 3] for more details of this construction.

Similar to the definition of 𝔤\mathfrak{g}, one may define a “compactified” version of 𝔤\mathfrak{g} by putting

𝔤¯:=((g,n)F𝕄𝖿𝗋(g,n)Sn[1])[[,λ]]\overline{\mathfrak{g}}:=\big{(}\bigoplus_{(g,n)}F\mathbb{M}^{\sf fr}(g,n)_{S_{n}}[1]\big{)}[[\hbar,\lambda]]

with the differential +uB+δ+Δ\partial+uB+\delta+\hbar\Delta (with δ\delta as in Equation (6)) and the Lie bracket {,}\{-,-\} defined in the same way as 𝔤\mathfrak{g}. Observe that there is an inclusion map

j:𝔤𝔤¯j:\mathfrak{g}\rightarrow\overline{\mathfrak{g}}

of DGLA’s onto the g,n\star_{g,n}-component of each F𝕄𝖿𝗋(g,n)SnF\mathbb{M}^{\sf fr}(g,n)_{S_{n}}.

We also define a trivialized version of 𝔤¯\overline{\mathfrak{g}}, denoted by 𝔤¯𝗍𝗋𝗂𝗏\overline{\mathfrak{g}}^{\sf triv} which has the same underlying graded vector space as 𝔤¯\overline{\mathfrak{g}}, but is endowed with differential +uB+δ\partial+uB+\delta and zero Lie bracket.

{Theorem}

There exists an LL_{\infty} quasi-isomorphism of DGLA’s:

𝒦:𝔤¯𝔤¯𝗍𝗋𝗂𝗏\mathcal{K}:\overline{\mathfrak{g}}\rightarrow\overline{\mathfrak{g}}^{\sf triv}

This theorem is the geometric “lift” of the algebraic version proved in [3, Theorem 4.2]. The proof is in complete parallel as well. In the following, we describe the construction of 𝒦\mathcal{K}. Roughly speaking, the map is simply inserting elements of 𝔤¯\overline{\mathfrak{g}} on vertices of stable graphs, as illustrated in Figure (4).

Construction of 𝒦\mathcal{K}.

We proceed to construct an LL_{\infty}-morphism 𝒦\mathcal{K}. First, we introduce more notations about graphs. Denote by Γ(g,n)m\Gamma(g,n)_{m} (respectively Γ((g,n))m\Gamma((g,n))_{m}) the set of labeled (respectively stable) graphs with mm vertices. For a labeled graph GΓ(g,n)mG\in\Gamma(g,n)_{m}, a marking of GG is a bijection

f:{1,,m}VG.f:\{1,\cdots,m\}\rightarrow V_{G}.

An isomorphism between two marked and labeled graphs is an isomorphism of the underlying labeled graphs that also preserves the marking map. Denote by Γ(g,n)~m\widetilde{\Gamma(g,n)}_{m} (respectively Γ((g,n))~m\widetilde{\Gamma((g,n))}_{m}) the set of isomorphism classes of marked (respectively stable) graphs.

For each integer m1m\geq 1, we shall define a degree zero linear map

𝒦m:Symm(𝔤¯[1])𝔤¯𝗍𝗋𝗂𝗏[1]\mathcal{K}_{m}:\operatorname{Sym}^{m}(\overline{\mathfrak{g}}[1])\rightarrow\overline{\mathfrak{g}}^{\sf triv}[1]

The shift by one of 𝔤¯\overline{\mathfrak{g}} is given by

𝔤¯[1]=((g,n)F𝕄𝖿𝗋(g,n)Sn[2])[[,λ]]\overline{\mathfrak{g}}[1]=\big{(}\bigoplus_{(g,n)}F\mathbb{M}^{\sf fr}(g,n)_{S_{n}}[2]\big{)}[[\hbar,\lambda]]

For each marked labeled graph (G,f)Γ(g,n)~m(G,f)\in\widetilde{\Gamma(g,n)}_{m} with f:{1,,m}VGf:\{1,\ldots,m\}\rightarrow V_{G} a marking on the vertices of GG, denote by vj:=f(j)v_{j}:=f(j) and nj:=|𝖫𝖾𝗀(vj)|n_{j}:=|{\sf Leg}(v_{j})|. Define a \mathbb{Q}-linear map

𝒦(G,f):j=1mF𝕄𝖿𝗋(g(vj),nj)Snj[2]g(vj)F𝕄𝖿𝗋(g,n)Sn[2]g.\mathcal{K}_{(G,f)}:\bigotimes_{j=1}^{m}F\mathbb{M}^{\sf fr}(g(v_{j}),n_{j})_{S_{n_{j}}}[2]\cdot\hbar^{g(v_{j})}\rightarrow F\mathbb{M}^{\sf fr}(g,n)_{S_{n}}[2]\cdot\hbar^{g}.
Refer to caption
Figure 4: Illustration of 𝒦(G,f)(γ1,γ2,γ3)\mathcal{K}_{(G,f)}(\gamma_{1},\gamma_{2},\gamma_{3})

For each 1jm1\leq j\leq m, let

γj(D(Hj)vVHj𝕄S1𝖿𝗋(g(v),𝖫𝖾𝗀(v)))𝖠𝗎𝗍(Hj)\gamma_{j}\in\big{(}D(H_{j})\bigotimes\bigotimes_{v\in V_{H_{j}}}\mathbb{M}_{S^{1}}^{\sf fr}(g(v),{\sf Leg}(v))\big{)}_{{\sf Aut}(H_{j})}

be an element of F𝕄𝖿𝗋(g(vj),nj)F\mathbb{M}^{\sf fr}(g(v_{j}),n_{j}), with its underlying stable graph given by HjΓ(g(vj),nj)H_{j}\in\Gamma(g(v_{j}),n_{j}). In order to insert the graphs HjH_{j}’s into GG, we need to make identification of leaves of HjH_{j} with 𝖫𝖾𝗀(vj){\sf Leg}(v_{j}). There is no canonical choice for such an identification, thus we take the symmetrization of the γj\gamma_{j}’s. We set

γjSym:=σj:{1,,nj}𝖫𝖾𝗀(vj)γjσj,\gamma_{j}^{\operatorname{Sym}}:=\sum_{\sigma_{j}:\{1,\ldots,n_{j}\}\rightarrow{\sf Leg}(v_{j})}\gamma_{j}^{\sigma_{j}},

where the notation γjσj\gamma_{j}^{\sigma_{j}} is the same element as γj\gamma_{j}, but labeled by an isomorphism σj:{1,,nj}𝖫𝖾𝗀(vj)\sigma_{j}:\{1,\ldots,n_{j}\}\rightarrow{\sf Leg}(v_{j}). Fixing the identifications σ1,,σm\sigma_{1},\ldots,\sigma_{m} on vertices of GG, denote by

G(H1σ,,Hmσm)G(H_{1}^{\sigma},\ldots,H_{m}^{\sigma_{m}})

the stable graph obtained by inserting HjH_{j} at vertex vjv_{j} through the isomorphism σj\sigma_{j}. With this preparation, we define

𝒦(G,f)(γ1,,γm):=σ1,,σmD(G)γ1σ1γmσm.\mathcal{K}_{(G,f)}(\gamma_{1},\ldots,\gamma_{m}):=\sum_{\sigma_{1},\ldots,\sigma_{m}}D(G)\otimes\gamma_{1}^{\sigma_{1}}\otimes\cdots\otimes\gamma_{m}^{\sigma_{m}}.

This makes sense because

  • There is a natural isomorphism

    D(G(H1σ1,,Hmσm))D(G)D(H1)D(Hm)D(G(H_{1}^{\sigma_{1}},\ldots,H_{m}^{\sigma_{m}}))\cong D(G)\otimes D(H_{1})\otimes\cdots\otimes D(H_{m})
  • We have VG(H1σ1,,Hmσm)=j=1mVHjV_{G(H_{1}^{\sigma_{1}},\ldots,H_{m}^{\sigma_{m}})}=\coprod_{j=1}^{m}V_{H_{j}}.

  • This map descends to coinvariants by graph automorphisms. Indeed, given an automorphism ρj:HjHj\rho_{j}:H_{j}\rightarrow H_{j}, it induces an isomorphism

    G(H1σ1,,Hmσm)G(H1σ1,,Hjρj.σj,,Hmσm),G(H_{1}^{\sigma_{1}},\ldots,H_{m}^{\sigma_{m}})\rightarrow G(H_{1}^{\sigma_{1}},\ldots,H_{j}^{\rho_{j}.\sigma_{j}},\ldots,H_{m}^{\sigma_{m}}),

    where ρj.σj\rho_{j}.\sigma_{j} is the identification given by first applying σj\sigma_{j} then followed by the automorphism ρj\rho_{j} acting on leaves of HjH_{j}.

  • This map is of homological degree zero after the shift by [2][2]. To see this, observe that at vertices the degrees are simply that in γj\gamma_{j}’s, but since D(G(H1σ1,,Hmσm))D(G)D(H1)D(Hm)D(G(H_{1}^{\sigma_{1}},\ldots,H_{m}^{\sigma_{m}}))\cong D(G)\otimes D(H_{1})\otimes\cdots\otimes D(H_{m}) the term D(G)D(G) has degree 2|EG|2|E_{G}|. Putting this together with the shift by [2][2] and the powers of \hbar, we have

    deg(𝒦(G,f))\displaystyle\deg(\mathcal{K}_{(G,f)}) =2|EG|2|VG|+22(gg(v1)g(vm))\displaystyle=2|E_{G}|-2|V_{G}|+2-2(g-g(v_{1})-\cdots-g(v_{m}))
    =2(1+h1(G))+22(gg(v1)g(vm))\displaystyle=2(-1+h_{1}(G))+2-2(g-g(v_{1})-\cdots-g(v_{m}))
    =2(h1(G)+g(v1)++g(vm)g)=0\displaystyle=2(h_{1}(G)+g(v_{1})+\cdots+g(v_{m})-g)=0

Finally, extending 𝒦(G,f)\mathcal{K}_{(G,f)} by λ\lambda-linearity, we define for each m1m\geq 1 a linear (in both λ\lambda and \hbar variables) by setting

(16) 𝒦m:=(g,n)(G,f)Γ(g,n)~m1|𝖠𝗎𝗍(G,f)|𝒦(G,f)~{}\mathcal{K}_{m}:=\sum_{\begin{subarray}{c}(g,n)\\ (G,f)\in\widetilde{\Gamma(g,n)}_{m}\end{subarray}}\frac{1}{|{\sf Aut}(G,f)|}\cdot\mathcal{K}_{(G,f)}

Proof of Equation (3).

In this subsection, we prove the main identity (3) that the push-forward of the String vertex 𝒱=(g,n)𝒱g,ngλ2g2+n\mathcal{V}=\sum_{(g,n)}\mathcal{V}_{g,n}\hbar^{g}\lambda^{2g-2+n} via the composition LL_{\infty} map

𝔤𝔤¯𝒦𝔤¯𝗍𝗋𝗂𝗏\mathfrak{g}\hookrightarrow\overline{\mathfrak{g}}\stackrel{{\scriptstyle\mathcal{K}}}{{\longrightarrow}}\overline{\mathfrak{g}}^{\sf triv}

gives exactly the fundamental classes [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}].

{Theorem}

Under the isomorphism in Theorem 4, we have the following formula expressing the fundamental class of M¯g,n/Sn\overline{M}_{g,n}/S_{n} in terms of String vertices:

[M¯g,n/Sn]=(𝒦𝒱)g,n=GΓ((g,n))1|𝖠𝗎𝗍(G)|eEGDevVG𝒱g(v),|𝖫𝖾𝗀(v)|Sym.[\overline{M}_{g,n}/S_{n}]=(\mathcal{K}_{*}\mathcal{V})_{g,n}=\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\prod_{e\in E_{G}}D_{e}\otimes\prod_{v\in V_{G}}\mathcal{V}^{\operatorname{Sym}}_{g(v),|{\sf Leg}(v)|}.
Proof 5.1.

Define a filtration on the chain complex 𝕄^𝖿𝗋(g,n)\widehat{\mathbb{M}}^{\sf fr}(g,n) by the setting k𝕄^𝖿𝗋(g,n)\partial_{k}\widehat{\mathbb{M}}^{\sf fr}(g,n) to be chains that are supported on the locus such that the underlying stable Riemann surfaces have at least kk nodes. Similarly, we also have the nodal filtration on the homotopy quotient k𝕄^S1𝖿𝗋(g,n)(k0)\partial_{k}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)\;(k\geq 0), and further SnS_{n}-quotient k𝕄^S1𝖿𝗋(g,n)Sn\partial_{k}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}. Following Costello [5, Theorem 9.0.7], one may form a DGLA on the homology of the associated graded

𝔫:=(g,n),k0H(𝖦𝗋k𝕄^S1𝖿𝗋(g,n)Sn)[1][[,λ]]\mathfrak{n}:=\bigoplus_{(g,n),k\geq 0}H_{*}\big{(}{\sf Gr}^{k}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\big{)}[1][[\hbar,\lambda]]

Its differential and Lie bracket are denoted by +Δ\partial+\hbar\Delta and {,}\{-,-\} respectively. The boundary map :Hd(𝖦𝗋k𝕄^S1𝖿𝗋(g,n)Sn)Hd1(𝖦𝗋k+1𝕄^S1𝖿𝗋(g,n)Sn)\partial:H_{d}\big{(}{\sf Gr}^{k}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\big{)}\rightarrow H_{d-1}\big{(}{\sf Gr}^{k+1}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\big{)} is the boundary map induced by the sequence

0𝖦𝗋k+1𝕄^S1𝖿𝗋(g,n)Snk𝕄^S1𝖿𝗋(g,n)Sn/k+2𝕄^S1𝖿𝗋(g,n)Sn𝖦𝗋k𝕄^S1𝖿𝗋(g,n)Sn00\rightarrow{\sf Gr}^{k+1}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\rightarrow\partial_{k}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}/\partial_{k+2}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\rightarrow{\sf Gr}^{k}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\rightarrow 0

The two operations Δ\Delta and {,}\{-,-\} are both defined using the twisted sewing operations. Observe that when k=0k=0 we have

H(𝖦𝗋0𝕄^S1𝖿𝗋(g,n)Sn)\displaystyle H_{*}\big{(}{\sf Gr}^{0}_{\partial_{\bullet}}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\big{)}
=\displaystyle= H(𝕄^S1𝖿𝗋(g,n)Sn,1𝕄^S1𝖿𝗋(g,n)Sn)\displaystyle H_{*}\big{(}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}},\partial_{1}\widehat{\mathbb{M}}_{S^{1}}^{\sf fr}(g,n)_{S_{n}}\big{)}
\displaystyle\cong H(M¯g,n/Sn,1M¯g,n/Sn)\displaystyle H_{*}\big{(}\overline{M}_{g,n}/S_{n},\partial_{1}\overline{M}_{g,n}/S_{n}\big{)}

The last isomorphism is induced by the canonical map M^g,nM¯g,n\widehat{M}_{g,n}\rightarrow\overline{M}_{g,n} that forgets the decoration at all nodes. The orbifold fundamental class [M¯g,n/Sn][\overline{M}_{g,n}/S_{n}], under this isomorphism, corresponds to the fundamental class [M^g,n/Sn]H6g6+2n(M^g,n/Sn,1M^g,n/Sn)[\widehat{M}_{g,n}/S_{n}]\in H_{6g-6+2n}(\widehat{M}_{g,n}/S_{n},\partial_{1}\widehat{M}_{g,n}/S_{n}). It was shown in Loc. Cit. that the element

𝒳:=(g,n)[M^g,n/Sn]gλ2g2+n\mathcal{X}:=\sum_{(g,n)}[\widehat{M}_{g,n}/S_{n}]\hbar^{g}\lambda^{2g-2+n}

of the DGLA 𝔫\mathfrak{n} satisfies the Maurer-Cartan equation and is uniquely determined by the equation after fixing the initial condition at (g,n)=(0,3)(g,n)=(0,3).

To this end, let us consider the following commutative diagram

F𝕄𝖿𝗋(g,n)i𝖳𝗈𝗍(F𝕄𝖿𝗋(g,n))𝕀𝖳𝗈𝗍(F𝕄^𝖿𝗋(g,n))pC(M¯g,n)𝗉𝗋𝖥𝗈𝗋𝗀𝖾𝗍𝕄^S1𝖿𝗋(g,n)qC(M^g,n)\begin{CD}F{\mathbb{M}}^{\sf fr}(g,n)@>{i^{\sharp}}>{}>{\sf Tot}\big{(}F{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}@>{\mathbb{I}}>{}>{\sf Tot}\big{(}F\widehat{\mathbb{M}}^{\sf fr}(g,n)_{\bullet}\big{)}@>{p}>{}>C_{*}(\overline{M}_{g,n})\\ @V{{\sf pr}}V{}V@A{{\sf Forget}}A{}A\\ \widehat{\mathbb{M}}^{\sf fr}_{S^{1}}(g,n)@>{q}>{}>C_{*}(\widehat{M}_{g,n})\end{CD}

where the downward arrow is the projection map onto the g,n\star_{g,n} component, the upward arrow is the map that forgets the nodal decoration, and the bottom horizontal arrow qq is the canonical quotient map followed by forgetting framings at punctures. Recall from Paragraph 4 the isomorphism in Theorem 4 is induced by the chain map p𝕀ip\mathbb{I}i^{\sharp}. By the previous discussion, it suffices to prove that

q𝗉𝗋𝕀i(𝒦𝒱)g,n=q𝗉𝗋𝕀i(GΓ((g,n))1|𝖠𝗎𝗍(G)|eEGDevVG𝒱g(v),|𝖫𝖾𝗀(v)|Sym)q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g,n}=q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\prod_{e\in E_{G}}D_{e}\otimes\prod_{v\in V_{G}}\mathcal{V}^{\operatorname{Sym}}_{g(v),|{\sf Leg}(v)|}\big{)}

satisfies the Maurer-Cartan equation in 𝔫\mathfrak{n}. Using the identity (15) we have

q𝗉𝗋𝕀i(𝒦𝒱)g,n\displaystyle q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g,n}
=\displaystyle= q(GΓ((g,n))1|𝖠𝗎𝗍(G)|ηG([0,1]××[0,1]|EG| copies×(eEGBe)(vVG𝒱g(v),n(v)Sym,[0])))\displaystyle q\Big{(}\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\cdot\eta_{G}\big{(}\underbrace{[0,1]\times\cdots\times[0,1]}_{\mbox{$|E_{G}|$ copies}}\times(\prod_{e\in E_{G}}B_{e})(\prod_{v\in V_{G}}\mathcal{V}^{\operatorname{Sym},[0]}_{g(v),n(v)})\big{)}\Big{)}

To finish the proof, we show that the Maurer-Cartan equation satisfied by 𝒱g,n\mathcal{V}_{g,n}’s implies the desired Maurer-Cartan equation of q𝗉𝗋𝕀i(𝒦𝒱)g,nq{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g,n}’s. Indeed, the leading coefficients 𝒱g,n[0]\mathcal{V}_{g,n}^{[0]} of string vertices satisfies the equation:

(17) 𝒱g,n[0]=12g+g′′=g,n+n′′=n{𝒱g,n[0],𝒱g′′,n′′[0]}Δ𝒱g1,n+2[0]~{}\partial\mathcal{V}_{g,n}^{[0]}=-\frac{1}{2}\sum_{g^{\prime}+g^{\prime\prime}=g,n^{\prime}+n^{\prime\prime}=n}\{\mathcal{V}_{g^{\prime},n^{\prime}}^{[0]},\mathcal{V}_{g^{\prime\prime},n^{\prime\prime}}^{[0]}\}-\Delta\mathcal{V}_{g-1,n+2}^{[0]}

The boundary of the expression q𝗉𝗋𝕀i(𝒦𝒱)g,nq{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g,n} consists of three types of terms:

  • (i)

    For an edge eEGe\in E_{G}, the corresponding parameter te=0t_{e}=0 (with a minus sign).

  • (ii)

    For an edge eEGe\in E_{G}, the corresponding parameter te=1t_{e}=1 (with a plus sign).

  • (iii)

    At a vertex vVGv\in V_{G}, there is 𝒱g(v),n(v)Sym,[0]\partial\mathcal{V}_{g(v),n(v)}^{\operatorname{Sym},[0]}.

We observe that using Equation (17), the (ii)(ii) and (iii)(iii) type terms cancel each other. Moreover, the (i)(i) term is precisely the twisted sewing operations used in defining the DGLA structure of 𝔫\mathfrak{n}. Hence depending on whether eEGe\in E_{G} is a separating or a non-separating edge, the (i)(i) term can be rewritten as

12g+g′′=g,n+n′′=n{q𝗉𝗋𝕀i(𝒦𝒱)g,n,q𝗉𝗋𝕀i(𝒦𝒱)g′′,n′′}Δq𝗉𝗋𝕀i(𝒦𝒱)g1,n+2.-\frac{1}{2}\sum_{g^{\prime}+g^{\prime\prime}=g,n^{\prime}+n^{\prime\prime}=n}\{q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g^{\prime},n^{\prime}},q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g^{\prime\prime},n^{\prime\prime}}\}-\Delta q{\sf pr}\mathbb{I}i^{\sharp}\big{(}\mathcal{K}_{*}\mathcal{V}\big{)}_{g-1,n+2}.

This finishes the proof.

6 CEI of the ground field

In this section, we use Equation (3) to prove Theorem A. Throughout the section, let AA be a cyclic AA_{\infty} algebra of dimension dd over a field 𝕂\mathbb{K} (of characteristic zero) that is proper, smooth and satisfies the Hodge-to-de-Rham degeneration.

A sketch of the definition of CEI.

CEI defined in [3]) takes a roundabout route due to the fact that the 22-dimensional Topological Conformal Field Theory (TCFT) structure on the reduced Hochschild chain complex C(A)C_{*}(A) has to have strictly positive number of inputs. This separation of inputs and outputs is the source of main difficulties in making Costello’s original definition [5] of categorical enumerative invariants explicit. Indeed, in [2] and [3], a Koszul type resolution is constructed for the Sen-Zwiebach Lie algebra (see Paragraph 5), giving a quasi-isomorphism of DGLA’s:

𝔤𝔤^,\mathfrak{g}\rightarrow\widehat{\mathfrak{g}},

where the right hand side DGLA 𝔤^\widehat{\mathfrak{g}} consists of Riemann surfaces with strictly positive number of inputs. The precise definition of 𝔤^\widehat{\mathfrak{g}} is not so much relevant to this paper: we shall see that in the particular case when the algebra is the ground field there is no need to use the Koszul resolution.

Similar to the construction of the Sen-Zwiebach Lie algebra 𝔤\mathfrak{g}, associated with the Hochschild chain complex C(A)[d]C_{*}(A)[d] and the Connes operator BB, we may form a DGLA

𝔥A:=Sym(C(A)[d])[1][[,λ]].\mathfrak{h}_{A}:=\operatorname{Sym}\big{(}C_{*}(A)[d]\big{)}[1][[\hbar,\lambda]].

In [2], we also introduced its Koszul resolution 𝔥^A\widehat{\mathfrak{h}}_{A}, and a canonical quasi-isomorphism of DGLA’s:

𝔥A𝔥^A.\mathfrak{h}_{A}\rightarrow\widehat{\mathfrak{h}}_{A}.

Again we omit the definition of 𝔥^A\widehat{\mathfrak{h}}_{A} and refer the details to [2].

In an ideal situation, the TCFT structure on C(A)C_{*}(A) would give us a commutative diagram

𝔤\mathfrak{g}𝔤^\widehat{\mathfrak{g}}𝔥A\mathfrak{h}_{A}𝔥^A\widehat{\mathfrak{h}}_{A}ι\iotaρ^A\widehat{\rho}_{A}ι\iotaρA\rho_{A}

with ρA\rho_{A} and ρ^A\widehat{\rho}_{A} the TCFT structure maps. In reality, the left vertical map ρA\rho_{A} is not there, which prevents us to obtain the push-forward (ρA)𝒱(\rho_{A})_{*}\mathcal{V} of the string vertex in order to define categorical enumerative invariants. In [3], we took the roundabout route:

  • (1.)

    We find string vertices 𝒱^\widehat{\mathcal{V}} in 𝔤^\widehat{\mathfrak{g}}.

  • (2.)

    Then push-forward 𝒱^\widehat{\mathcal{V}} to obtain a Maurer-Cartan element

    β^:=(ρ^A)𝒱^\widehat{\beta}:=(\widehat{\rho}_{A})_{*}\widehat{\mathcal{V}}

    of the DGLA 𝔥^A\widehat{\mathfrak{h}}_{A}.

  • (3.)

    The bottom map ι\iota is a quasi-isomorphism of DGLA’s when AA is smooth, proper, and satisfies the Hodge-to-de-Rham degeneration property. Hence, by homotopy invariance of the Maurer-Cartan moduli space, we obtain the desired Maurer-Cartan element β𝔥A\beta\in\mathfrak{h}_{A} with its defining property that ιβ\iota_{*}\beta is gauge-equivalent to β^\widehat{\beta}.

Categorical enumerative invariants of the ground field.

When AA is the ground field \mathbb{Q}, the above roundabout route is not necessary: we can actually define a map

ρA:𝔤𝔥A\rho_{A}:\mathfrak{g}\rightarrow\mathfrak{h}_{A}

to obtain the commutative diagram in the previous paragraph. Thus when computing the CEI, we may simply take the Maurer-Cartan element β=(ρA)𝒱\beta=(\rho_{A})_{*}\mathcal{V} in 𝔥A\mathfrak{h}_{A}. Explicitly, the map ρA\rho_{A} is just the augmentation map: at degree (g,n)(g,n), the map ρA:C(Mg,n𝖿𝗋)\rho_{A}:C_{*}(M_{g,n}^{\sf fr})\rightarrow\mathbb{Q} sends any point class in C0(Mg,n𝖿𝗋)C_{0}(M_{g,n}^{\sf fr}) to 11, and is zero otherwise.

For the splitting of the Hodge filtration, observe that the Connes operator BB acts on by zero on C(A)=C_{*}(A)=\mathbb{Q}. Thus, we may choose the splitting s:C(A)=C(A)[[u]]=[[u]]s:C_{*}(A)=\mathbb{Q}\rightarrow C_{*}(A)[[u]]=\mathbb{Q}[[u]] to be

s(1)=1.s(1)=1.

Since uu is of degree 2-2, such a splitting is characterized by the homogeneity condition. By Proposition 2, a splitting induces an extension of the TCFT to give a map

ρ¯A,s:𝔤¯𝔥A.\overline{\rho}_{A,s}:\overline{\mathfrak{g}}\rightarrow\mathfrak{h}_{A}.

In summary, we obtain a commutative diagram of LL_{\infty} morphisms between DGLA’s:

(18) 𝔤i𝔤¯𝒦𝔤¯𝗍𝗋𝗂𝗏ρ¯A,sρ¯A,s𝔥A𝒦A,s𝔥A𝗍𝗋𝗂𝗏~{}\begin{CD}\mathfrak{g}@>{i}>{}>\overline{\mathfrak{g}}@>{\mathcal{K}}>{}>\overline{\mathfrak{g}}^{\sf triv}\\ @V{\overline{\rho}_{A,s}}V{}V@V{\overline{\rho}_{A,s}}V{}V\\ \mathfrak{h}_{A}@>{\mathcal{K}_{A,s}}>{}>\mathfrak{h}_{A}^{\sf triv}\end{CD}

The construction of 𝒦\mathcal{K} is done in Section 5 while 𝒦A\mathcal{K}_{A} is done in [3].

{Theorem}

The CEI of (A=,s)(A=\mathbb{Q},s) agrees with the Gromov-Witten invariants of a point, i.e. we have

uk1,,ukng,n,s=M¯g,nψ1k1ψnkn\langle u^{k_{1}},\ldots,u^{k_{n}}\rangle_{g,n}^{\mathbb{Q},s}=\int_{\overline{M}_{g,n}}\psi_{1}^{k_{1}}\cdots\psi_{n}^{k_{n}}
Proof 6.1.

The string vertex 𝒱g,n\mathcal{V}_{g,n} is of the form

𝒱g,n=[k1,,kn0𝒱g,nk1,,knu1k1unkn],\mathcal{V}_{g,n}=[\sum_{k_{1},\ldots,k_{n}\geq 0}\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}],

with 𝒱g,nk1,,knC6g6+2n2k12kn(Mg,n𝖿𝗋)\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}\in C_{6g-6+2n-2k_{1}-\cdots-2k_{n}}(M_{g,n}^{\sf fr}). The bracket is to indicate the quotient map by the SnS_{n} action. In characteristic zero, we may assume that the chains 𝒱g,nk1,,kn\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}} are invariant under the action of SnS_{n}. Then we have

ρ¯A,si(𝒱g,n)=k1++kn=3g3+nρA(𝒱g,nk1,,kn)n!𝖠𝗎𝗍(k1,,kn)qk1qkn.\overline{\rho}_{A,s}i(\mathcal{V}_{g,n})=\sum_{k_{1}+\cdots+k_{n}=3g-3+n}\rho_{A}(\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}})\frac{n!}{{\sf Aut}(k_{1},\ldots,k_{n})}q_{k_{1}}\cdots q_{k_{n}}.

In the above formula, the right hand side takes value in 𝔥A=Sym([u1])\mathfrak{h}_{A}=\operatorname{Sym}(\mathbb{Q}[u^{-1}]), and the variable qkjq_{k_{j}} corresponds the basis ukju^{-k_{j}}. The combinatorial number 𝖠𝗎𝗍(k1,,kn){\sf Aut}(k_{1},\ldots,k_{n}) is the size of the set of nn-permutations that fixes the sequence (k1,,kn)(k_{1},\ldots,k_{n}). The extension formula in Proposition 2 of the splitting ss yields the zero map for the action of DeD_{e} on the edges of a stable graph. This implies that the bottom LL_{\infty} algebra map 𝒦A,s\mathcal{K}_{A,s} is the identity map, which shows that

𝒦A,sρ¯A,si(𝒱g,n)=k1++kn=3g3+nρA(𝒱g,nk1,,kn)n!𝖠𝗎𝗍(k1,,kn)qk1qkn.\mathcal{K}_{A,s}\overline{\rho}_{A,s}i(\mathcal{V}_{g,n})=\sum_{k_{1}+\cdots+k_{n}=3g-3+n}\rho_{A}(\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}})\frac{n!}{{\sf Aut}(k_{1},\ldots,k_{n})}q_{k_{1}}\cdots q_{k_{n}}.

By definition of CEI, we have

ul1,,ulng,n,s\displaystyle\langle u^{l_{1}},\ldots,u^{l_{n}}\rangle_{g,n}^{\mathbb{Q},s}
=\displaystyle= ql1qln(k1++kn=3g3+nρA(𝒱g,nk1,,kn)n!𝖠𝗎𝗍(k1,,kn)qk1qkn)\displaystyle\frac{\partial}{\partial q_{l_{1}}}\cdots\frac{\partial}{\partial q_{l_{n}}}\big{(}\sum_{k_{1}+\cdots+k_{n}=3g-3+n}\rho_{A}(\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}})\frac{n!}{{\sf Aut}(k_{1},\ldots,k_{n})}q_{k_{1}}\cdots q_{k_{n}}\big{)}
=\displaystyle= n!ρA(𝒱g,nl1,,ln)\displaystyle n!\rho_{A}(\mathcal{V}_{g,n}^{l_{1},\ldots,l_{n}})

On the other hand, by Theorem 5 we have

[M¯g,n/Sn]=(𝒦𝒱)g,n=GΓ((g,n))1|𝖠𝗎𝗍(G)|eEGDevVG𝒱g(v),|𝖫𝖾𝗀(v)|.[\overline{M}_{g,n}/S_{n}]=(\mathcal{K}_{*}\mathcal{V})_{g,n}=\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\prod_{e\in E_{G}}D_{e}\otimes\prod_{v\in V_{G}}\mathcal{V}_{g(v),|{\sf Leg}(v)|}.

This implies that

M¯g,nψ1l1ψnln\displaystyle\int_{\overline{M}_{g,n}}\psi_{1}^{l_{1}}\cdots\psi_{n}^{l_{n}}
=\displaystyle= n!u1l1ukln(GΓ((g,n))1|𝖠𝗎𝗍(G)|eEGDevVG𝒱g(v),|𝖫𝖾𝗀(v)|)\displaystyle n!\cdot u_{1}^{l_{1}}\cdots u_{k}^{l_{n}}\big{(}\sum_{G\in\Gamma((g,n))}\frac{1}{|{\sf Aut}(G)|}\prod_{e\in E_{G}}D_{e}\otimes\prod_{v\in V_{G}}\mathcal{V}_{g(v),|{\sf Leg}(v)|}\big{)}
=\displaystyle= n!u1l1ukln(k1,,kn0𝒱g,nk1,,knu1k1unkn)\displaystyle n!\cdot u_{1}^{l_{1}}\cdots u_{k}^{l_{n}}\big{(}\sum_{k_{1},\ldots,k_{n}\geq 0}\mathcal{V}_{g,n}^{k_{1},\ldots,k_{n}}u_{1}^{-k_{1}}\cdots u_{n}^{-k_{n}}\big{)}
=\displaystyle= n!ρA(𝒱g,nl1,,ln)\displaystyle n!\rho_{A}(\mathcal{V}_{g,n}^{l_{1},\ldots,l_{n}})

This proves the theorem.

References

  • [1] Amorim, L.; Tu, J., Categorical primitive forms of Calabi-Yau AA_{\infty}-categories with semi-simple cohomology. arXiv:1909.05319
  • [2] Caldararu, A.; Costello, K.; Tu, J., Categorical Enumerative Invariants, I: String vertices, arXiv:2009.06673.
  • [3] Caldararu, A.; Tu, J., Categorical Enumerative Invariants, II: Givental formula, arXiv:2009.06659.
  • [4] Costello, K., Topological conformal field theories and Calabi-Yau categories, Advances in Mathematics, Volume 210, Issue 1, 20 March 2007, 165-214.
  • [5] Costello, K., The partition function of a topological field theory, J. Topol. 2 (2009), no. 4, 779-822.
  • [6] Costello, K.; Zwiebach B. Hyperbolic String Vertices, preprint, arXiv:1909.0003.
  • [7] Dotsenko, V.; Shadrin, S.; Vallette, B., Givental group action on topological field theories and homotopy Batalin-Vilkovisky algebras. Adv. Math. 236 (2013), 224-256.
  • [8] Dotsenko, V.; Shadrin, S.; Vallette, B., Givental action and trivialisation of circle action. J. Ec. polytech. Math. 2 (2015), 213-246.
  • [9] Drummond-Cole, G. Homotopically trivializing the circle in the framed little disks. J. Topol. 7 (2014), no. 3, 641-676.
  • [10] Getzler, E.; Kapranov, M., Modular operads. Volume 110, Issue 1 January 1998 , pp. 65-125.
  • [11] Givental, A., Gromov-Witten invariants and quantization of quadratic Hamiltonians, Mosc. Math. J. 1 (2001), 551-568.
  • [12] Khoroshkin, A.; Markarian, N.; Shadrin, S., Hypercommutative operad as a homotopy quotient of BV. Comm. Math. Phys. 322 (2013), no. 3, 697-729.
  • [13] Kimura, T.; Stasheff, J.; Voronov, A., On operad structures of moduli spaces and string theory, Comm. Math. Phys. 171(1), 1-25 (1995).
  • [14] Kontsevich, M., Homological algebra of mirror symmetry, Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Zürich, 1994), 120-139, Birkhäuser, Basel, 1995.
  • [15] Mondello, G. A remark on the virtual homotopical dimension of some moduli spaces of stable Riemann surfaces, Journal of the European Mathematical Society, vol.10 (1) 2008, pp.231-241.
  • [16] Pandharipande, R.; Pixton, A.; Zvonkine, D., Relations on M¯g,n\overline{M}_{g,n} via 33-spin structures, J. Amer. Math. Soc. 28 (2015), 279-309.
  • [17] Segal, G. The Definition of Conformal Field Theory. Differential Geometrical Methods in Theoretical Physics. NATO ASI Series (Series C: Mathematical and Physical Sciences), vol 250 (1988). Springer, Dordrecht.
  • [18] Sen, A., Zwiebach, B., Quantum background independence of closed-string field theory, Nuclear Phys. B 423(2-3), (1994), 580-630
  • [19] Teleman, C., The structure of 2D semi-simple field theories, Invent. Math. 188 (2012), 525-588.