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

\newaliascnt

lemmatheorem \aliascntresetthelemma \newaliascntcorollarytheorem \aliascntresetthecorollary \newaliascntconjecturetheorem \aliascntresettheconjecture \newaliascntpropositiontheorem \aliascntresettheproposition \newaliascntquestiontheorem \aliascntresetthequestion \newaliascntdefinitiontheorem \aliascntresetthedefinition \newaliascntnotationtheorem \aliascntresetthenotation \newaliascntremarktheorem \aliascntresettheremark \newaliascntextheorem \aliascntresettheex

The derived moduli stack of logarithmic flat connections

Francis Bischoff Exeter College and Mathematical Institute, University of Oxford; [email protected]
Abstract

We give an explicit finite-dimensional model for the derived moduli stack of flat connections on k\mathbb{C}^{k} with logarithmic singularities along a weighted homogeneous Saito free divisor. We investigate in detail the case of plane curves of the form xp=yqx^{p}=y^{q} and relate the moduli spaces to the Grothendieck-Springer resolution. We also discuss the shifted Poisson geometry of these moduli spaces. Namely, we conjecture that the map restricting a logarithmic connection to the complement of the divisor admits a shifted coisotropic structure and we construct a shifted Poisson structure on the formal neighbourhood of a canonical connection in the case of plane curves xp=yqx^{p}=y^{q}.

1 Introduction

Let DkD\subset\mathbb{C}^{k} be a hypersurface cut out by a reduced holomorphic function ff. In [30] Saito considers the subsheaf, usually denoted Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D), of holomorphic vector fields on k\mathbb{C}^{k} which preserve the ideal generated by ff. In general, it is coherent and closed under the Lie bracket, but may fail to be locally free. In fact, Saito provides a very explicit criterion for determining whether the sheaf is locally free. When it is, DD is said to be a free divisor and Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D), known as the logarithmic tangent bundle, defines a Lie algebroid. Examples of free divisors include smooth hypersurfaces, plane curves and simple normal crossings. In general, DD may be highly singular.

Let GG be a connected complex reductive group with Lie algebra 𝔤\mathfrak{g} and assume that DD is a free divisor which is homogeneous under a given \mathbb{C}^{*}-action on k\mathbb{C}^{k} with the property that all its weights are strictly positive. In this paper, we are interested in studying the moduli space of Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D)-representations on principal GG-bundles, also known as logarithmic flat connections. There is a standard way of defining this moduli space as the Maurer-Cartan locus of an infinite-dimensional differential graded Lie algebra (dgla) LD,𝔤L_{D,\mathfrak{g}} which is associated to DD and 𝔤\mathfrak{g}. Let Ωk1(logD)\Omega^{1}_{\mathbb{C}^{k}}(\text{log}D) denote the logarithmic cotangent bundle, which is the dual to Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D), and let Ωk(logD)=Ωk1(logD)\Omega^{\bullet}_{\mathbb{C}^{k}}(\text{log}D)=\wedge^{\bullet}\Omega^{1}_{\mathbb{C}^{k}}(\text{log}D) be the exterior algebra. This defines a commutative differential graded algebra when equipped with the Lie algebroid differential dd. Then LD,𝔤=Ωk(logD)𝔤L_{D,\mathfrak{g}}=\Omega^{\bullet}_{\mathbb{C}^{k}}(\text{log}D)\otimes\mathfrak{g} inherits the structure of a dgla. The Maurer-Cartan locus of this dgla is defined to be the following set

MC(LD,𝔤)={ωLD,𝔤1|dω+12[ω,ω]=0}.MC(L_{D,\mathfrak{g}})=\{\omega\in L^{1}_{D,\mathfrak{g}}\ |\ d\omega+\frac{1}{2}[\omega,\omega]=0\}.

Here, ωΩk1(logD)𝔤\omega\in\Omega^{1}_{\mathbb{C}^{k}}(\text{log}D)\otimes\mathfrak{g} is a Lie algebra valued 11-form, and it defines the following connection =d+ω\nabla=d+\omega, which has a logarithmic singularity along DD. It’s curvature is given by the following expression

F(ω)=dω+12[ω,ω].F(\omega)=d\omega+\frac{1}{2}[\omega,\omega].

The degree 0 component of the dgla is LD,𝔤0=Map(k,𝔤)L^{0}_{D,\mathfrak{g}}=Map(\mathbb{C}^{k},\mathfrak{g}), which is the Lie algebra of the infinite dimensional gauge group 𝔊=Map(k,G)\mathfrak{G}=Map(\mathbb{C}^{k},G). This group acts on the Maurer-Cartan locus, giving the correct equivalence between flat connections. As a result, the moduli space of flat logarithmic connections is defined to be the stack quotient

[MC(LD,𝔤)/𝔊].[MC(L_{D,\mathfrak{g}})/\mathfrak{G}].

Although this construction involves infinite dimensional spaces, in [4] we provide a purely finite dimensional model. More precisely, we show that the category of logarithmic flat connections with fixed residue data is equivalent to the stack quotient of an affine algebraic variety by the action of an algebraic group.

The purpose of the present paper is to provide a derived enhancement of the moduli stack. There are several different approaches to derived geometry in the literature, such as [16, 12, 19, 31, 28]. In this paper, we have opted to go with the notion of bundles of curved dgla’s, which requires relatively little technology and is sufficient for our purposes. Let us recall the definition from [2].

Definition \thedefinition.

A bundle of curved differential graded Lie algebras over a variety MM consists of a graded vector bundle \mathcal{L}^{\bullet} starting in degree 22, which is equipped with the following data

  1. 1.

    a section FΓ(M,2)F\in\Gamma(M,\mathcal{L}^{2}),

  2. 2.

    a degree 11 bundle map δ:[1]\delta:\mathcal{L}^{\bullet}\to\mathcal{L}^{\bullet}[1]

  3. 3.

    a smoothly varying graded Lie bracket [,][-,-] on the fibres of \mathcal{L}^{\bullet},

satisfying the following conditions

  1. 1.

    the Bianchi identity δF=0\delta{F}=0,

  2. 2.

    δ2=[F,]\delta^{2}=[F,-],

  3. 3.

    δ\delta is a graded derivation of the bracket [,][-,-].

If [M/𝒢][M/\mathcal{G}] is a stack, defined by the data of a Lie groupoid 𝒢\mathcal{G} over MM, then we define a bundle of curved dgla’s over the stack to be such a bundle over MM, which is equipped with an equivariant action of 𝒢\mathcal{G} preserving the data (F,δ,[,])(F,\delta,[-,-]). We will use the respective terminology of derived manifolds and derived stacks to refer to this data.

There is a standard way of constructing a derived manifold from the data of a dgla, and we can apply it to the case of LD,𝔤L_{D,\mathfrak{g}}. Namely, we take the base to be M=LD,𝔤1M=L_{D,\mathfrak{g}}^{1} and take the bundle \mathcal{L}^{\bullet} to be trivial with fibre given by the truncation LD,𝔤2L_{D,\mathfrak{g}}^{\bullet\geq 2}. The section is given by the curvature FF, and the bracket is simply the constant one inherited from LD,𝔤L_{D,\mathfrak{g}}. The bundle map δ\delta varies over MM, and above a point ωLD,𝔤1\omega\in L_{D,\mathfrak{g}}^{1}, it is given by the twisted differential δω=d+[ω,]\delta_{\omega}=d+[\omega,-]. Let us denote the resulting derived manifold D,𝔤\mathcal{M}_{D,\mathfrak{g}}. It can be further upgraded to a derived stack by noting that the gauge group 𝔊\mathfrak{G} lifts to an action on LD,𝔤2L_{D,\mathfrak{g}}^{\bullet\geq 2} via the adjoint representation.

A derived manifold (or stack) has an underlying classical truncation π0()\pi_{0}(\mathcal{M}), defined as the vanishing locus of the section FF. In the example under consideration, the classical truncation is given by the Maurer-Cartan locus, and hence

π0([D,𝔤/𝔊])=[MC(LD,𝔤)/𝔊].\pi_{0}([\mathcal{M}_{D,\mathfrak{g}}/\mathfrak{G}])=[MC(L_{D,\mathfrak{g}})/\mathfrak{G}].

For this reason, we say that [D,𝔤/𝔊][\mathcal{M}_{D,\mathfrak{g}}/\mathfrak{G}] is the derived moduli stack of flat logarithmic GG-connections. The main result of this paper is a finite-dimensional model of this derived stack.

Here is a brief description of this result. Given an element A𝔤A\in\mathfrak{g}, we consider the infinite dimensional derived moduli stack [D,𝔤(A)/𝔊][\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}] of GG-connections whose ‘residue’ is conjugate to AA. Details of this are given in Section 2. Let A=S+N0A=S+N_{0} be the Jordan decomposition, where SS is semisimple and N0N_{0} is nilpotent. In Section 3 we construct a finite dimensional dgla (U0,δS)(U_{0},\delta_{S}) associated to SS, with corresponding derived stack [𝒰S/Aut(S)][\mathcal{U}_{S}/Aut(S)]. This is interpreted as a certain sub-moduli space of flat connections on the fibre f1(1)f^{-1}(1). Then, associated to the element AA, we construct a derived substack [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] of the shifted tangent bundle T[1][𝒰S/Aut(S)]T[-1][\mathcal{U}_{S}/Aut(S)]. The main result is Theorem 3.1, which states that there is an equivalence of derived stacks

q:[𝒲(A)/Aut(S)][D,𝔤(A)/𝔊].q:[\mathcal{W}(A)/Aut(S)]\to[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}].

By this we mean that qq induces an equivalence between the groupoids of solutions to the MC equation, and given any solution ww, the derivative dqwdq_{w} is a quasi-isomorphism of tangent complexes.

In Section 5 we turn to the case of a plane curve defined by the function f=xpyqf=x^{p}-y^{q}. This is the simplest case above k=1k=1, and already it exhibits interesting behaviour. We construct an explicit derived stack [𝒬(A)/PS][\mathcal{Q}(A)/P_{S}] from the data of a parabolic subgroup PSP_{S} of the centralizer of exp(2πipqS)exp(\frac{2\pi i}{pq}S) and a representation H1(U0)H^{1}(U_{0}). In Theorem 5.1 we show that, given a condition on the eigenvalues of adSad_{S}, the derived stack [𝒬(A)/PS][\mathcal{Q}(A)/P_{S}] is equivalent to [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)]. For general SS, the moduli space can have extra components and we illustrate this in Example 5. The derived stack [𝒬(A)/PS][\mathcal{Q}(A)/P_{S}] can be interpreted in terms of spaces showing up in geometric representation theory, such as the Grothendieck-Springer resolution. Hence Theorem 5.1 can be viewed as a higher dimensional generalization of Boalch’s description in [5] of the moduli space of logarithmic connections on the disc.

Speculations about Poisson geometry

Going back to the work of Atiyah-Bott [1] and Goldman [15], we know that the moduli space of flat connections on a closed Riemann surface admits a symplectic structure. If the surface is punctured, then the moduli space admits a Poisson structure, whose symplectic leaves are obtained by fixing boundary conditions at the punctures [14]. This picture has since been generalized in several directions, including to the case of flat connections with singularities [6, 8, 7]. More recently, the moduli space of local systems on higher dimensional manifolds has been studied using tools from derived algebraic geometry. For a compact oriented manifold MM of dimension dd, the moduli space of local systems LocG(M)Loc_{G}(M) is a derived stack equipped with a shifted symplectic structure of degree 2d2-d [25]. If MM has a boundary M=N\partial M=N, then LocG(N)Loc_{G}(N) has a 3d3-d-shifted symplectic structure, and the restriction map r:LocG(M)LocG(N)r:Loc_{G}(M)\to Loc_{G}(N) has a Lagrangian structure [9], inducing on LocG(M)Loc_{G}(M) a 2d2-d-shifted Poisson structure [22]. We wish to generalize this picture to the case of logarithmic flat connections in higher dimensions.

In the above setting of a map f:kf:\mathbb{C}^{k}\to\mathbb{C}, the inverse image of the unit circle f1(S1)f^{-1}(S^{1}) is a manifold of dimension 2k12k-1, usually with boundary, and so LocG(f1(S1))Loc_{G}(f^{-1}(S^{1})) has a Poisson structure of degree 32k3-2k. Given a logarithmic flat connection, we can restrict it to f1(S1)f^{-1}(S^{1}) and take its holonomy. This should define a map

r:[𝒲(A)/Aut(S)]LocG(f1(S1)).r:[\mathcal{W}(A)/Aut(S)]\to Loc_{G}(f^{-1}(S^{1})). (1.1)

This map was studied by Boalch [5] in the special case of k=1k=1, where f1(S1)=S1f^{-1}(S^{1})=S^{1}. In this case LocG(S1)=G/GLoc_{G}(S^{1})=G/G has a 11-shifted symplectic structure, and the work of Boalch (suitably interpreted by [29]) shows that rr has a Lagrangian structure. In higher dimensions we conjecture that the map can be equipped with a shifted coisotropic structure in the sense of [21, 22].

Conjecture \theconjecture.

The map rr can be naturally equipped with a coisotropic structure.

In order to avoid the analytic issues that arise in taking the holonomy, it may be preferable to replace LocG(f1(S1))Loc_{G}(f^{-1}(S^{1})) with a moduli space of flat connections on the complement of DD.

In recent work [23, 24], Pantev and Toën studied the moduli spaces of local systems and flat connections on non-compact algebraic varieties. They constructed shifted Poisson structures and explained how to obtain the symplectic leaves by imposing suitable boundary conditions at infinity. Conjecture 1 may be viewed as providing another source of boundary conditions for the moduli spaces associated to f1(kD)f^{-1}(\mathbb{C}^{k}\setminus D). We hope that it may also be used in conjunction with their results, for example by considering the map rr in the presence of additional boundary conditions at the boundary of the fibres of ff.

One implication of the conjecture is that the moduli spaces [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] should admit 2(1k)2(1-k)-shifted Poisson structures. In Theorem 5.2 we provide evidence for the conjecture by constructing a 2-2-shifted Poisson structure on the formal neighbourhood of a special connection in the case of plane curves xp=yqx^{p}=y^{q}. Our construction is somewhat ad hoc, but it makes use of an invariant inner product on the Lie algebra 𝔤\mathfrak{g}, as well as the intersection pairing on the cohomology of the curve f1(1)f^{-1}(1). We have also not checked that our shifted Poisson structure fits into the formalism developed by [10]. We hope to address all these issues in future work.

Acknowledgements. I would like to thank Elliot Cheung for pointing me to the paper [2].

2 Homogeneous free divisors and logarithmic flat connections

Assume that the given \mathbb{C}^{*} action on k\mathbb{C}^{k} has strictly positive weights. It is generated infinitesimally by an Euler vector field

E=i=1knizizi,E=\sum_{i=1}^{k}n_{i}z_{i}\partial_{z_{i}},

where ni>0n_{i}\in\mathbb{Z}_{>0} are positive integers. This vector field defines a weight grading on the holomorphic functions 𝒪k\mathcal{O}_{\mathbb{C}^{k}} (and more generally tensor fields) on k\mathbb{C}^{k}, such that the coordinate function ziz_{i} has weight nin_{i}. This grading will play an important role. Because of our assumption, each weight space is finite-dimensional over \mathbb{C}. We also assume that the function ff defining DD is homogeneous of weight rr: E(f)=rfE(f)=rf.

The \mathbb{C}^{*} action determines an action Lie algebroid k\mathbb{C}\ltimes\mathbb{C}^{k} which is generated by the Euler vector field. Because EE is a section of Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D), there is an induced Lie algebroid morphism

i:kTk(logD),(λ,z)λEz.i:\mathbb{C}\ltimes\mathbb{C}^{k}\to T_{\mathbb{C}^{k}}(-\text{log}D),\qquad(\lambda,z)\mapsto\lambda E_{z}.

The logarithmic 11-form dlogf=dffd\text{log}f=\frac{df}{f} is a closed section of Ωk1(logD)\Omega^{1}_{\mathbb{C}^{k}}(\text{log}D). Hence, it determines a Lie algebroid morphism

π:Tk(logD),V1rfV(f),\pi:T_{\mathbb{C}^{k}}(-\text{log}D)\to\mathbb{C},\qquad V\mapsto\frac{1}{rf}V(f),

where \mathbb{C} is considered as an abelian Lie algebra. The composition p=πi:kp=\pi\circ i:\mathbb{C}\ltimes\mathbb{C}^{k}\to\mathbb{C} is given by projection to the first factor. This has a section

j:k,λ(λ,0),j:\mathbb{C}\to\mathbb{C}\ltimes\mathbb{C}^{k},\qquad\lambda\mapsto(\lambda,0),

which is also a Lie algebroid morphism. Altogether, we have the following diagram of Lie algebroids:

k\mathbb{C}\ltimes\mathbb{C}^{k}\mathbb{C}Tk(logD)T_{\mathbb{C}^{k}}(-\text{log}D)iiπ\pippjj

Each Lie algebroid determines a differential graded Lie algebra, whose Maurer-Cartan locus consists of flat algebroid connections. Furthermore, each morphism of Lie algebroids determines a pullback morphism between dgla’s, and as a result, a pullback morphism between categories of representations, or more generally, derived moduli stacks of flat connections. This gives rise to the following diagram of (infinite-dimensional) derived stacks:

[(𝒪k𝔤)/𝔊][(\mathcal{O}_{\mathbb{C}^{k}}\otimes\mathfrak{g})/\mathfrak{G}][𝔤/G][\mathfrak{g}/G][D,𝔤/𝔊][\mathcal{M}_{D,\mathfrak{g}}/\mathfrak{G}]ii^{*}π\pi^{*}pp^{*}jj^{*}

In this diagram, [𝔤/G][\mathfrak{g}/G] is the moduli stack of 𝔤\mathfrak{g}-representations of \mathbb{C}. It is the stack quotient corresponding to the adjoint action of GG on its Lie algebra. [(𝒪k𝔤)/𝔊][(\mathcal{O}_{\mathbb{C}^{k}}\otimes\mathfrak{g})/\mathfrak{G}] is the moduli stack of 𝔤\mathfrak{g}-representations of r\mathbb{C}\ltimes\mathbb{C}^{r}. In both cases the derived structure is trivial because the Lie algebroids have rank 11.

Now fix an element A𝔤A\in\mathfrak{g}, let OA𝔤O_{A}\subset\mathfrak{g} be its adjoint orbit, and let GAGG_{A}\subseteq G be its centralizer subgroup. This determines a substack [OA/G][𝔤/G][O_{A}/G]\subset[\mathfrak{g}/G] which is Morita equivalent to BGABG_{A}. The preimage [D,𝔤(A)/𝔊]:=(ji)1(BGA)[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}]:=(j^{*}i^{*})^{-1}(BG_{A}) is the derived stack of logarithmic flat connections ω\omega whose ‘residue’ ji(ω)j^{*}i^{*}(\omega) lies in OAO_{A}. More precisely, the base of the derived manifold D,𝔤(A)\mathcal{M}_{D,\mathfrak{g}}(A) is given by

M(A)={ωΩk1(logD)𝔤|ji(ω)OA},M(A)=\{\omega\in\Omega^{1}_{\mathbb{C}^{k}}(\text{log}D)\otimes\mathfrak{g}\ |\ j^{*}i^{*}(\omega)\in O_{A}\},

with the bundle of curved dgla’s restricted from D,𝔤\mathcal{M}_{D,\mathfrak{g}}. The action of 𝔊\mathfrak{G} preserves M(A)M(A).

3 Finite dimensional model

Let A=S+N0A=S+N_{0} be the Jordan decomposition of AA, where SS is semisimple, N0N_{0} is nilpotent, and [S,N0]=0[S,N_{0}]=0. In this section we will construct a finite dimensional model for [D,𝔤(A)/𝔊][\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}].

The dgla LD,𝔤L_{D,\mathfrak{g}}

We start by analysing the structure of the dgla LD,𝔤L_{D,\mathfrak{g}}. Being constructed from the cdga Ωk(logD)\Omega^{\bullet}_{\mathbb{C}^{k}}(\text{log}D) and the Lie algebra 𝔤\mathfrak{g}, LD,𝔤L_{D,\mathfrak{g}} inherits their derivations. The basic ones are as follows:

  • the Lie algebroid differential dd, which has degree +1+1 and squares to 0,

  • the interior multiplication with the Euler vector field ιE\iota_{E}, which has degree 1-1 and squares to 0,

  • the adjoint action of SS, adSad_{S}, which has degree 0.

By taking commutator brackets we arrive at further derivations, such as LE=[ιE,d]L_{E}=[\iota_{E},d], the Lie derivative with respect to EE, which is a derivation of degree 0. We can also wedge any derivation by a differential form to obtain a new derivation. Let α0=1rdlogf\alpha_{0}=\frac{1}{r}d\text{log}f, which is a closed logarithmic 11-form. Then α0adS\alpha_{0}ad_{S} is a degree +1+1 derivation which squares to 0. Among the 55 derivations just described, almost all of them commute. The only two non-vanishing commutator brackets are the following:

[ιE,d]=LE,[ιE,α0adS]=adS.[\iota_{E},d]=L_{E},\qquad[\iota_{E},\alpha_{0}ad_{S}]=ad_{S}.

The second bracket follows as a consequence of the identity ιE(α0)=1\iota_{E}(\alpha_{0})=1. We are primarily interested in studying the dgla structure arising from

δS=d+α0adS,\delta_{S}=d+\alpha_{0}ad_{S},

which is a degree +1+1 derivation that squares to 0. We are also interested in the following degree 0 derivation

LS:=[ιE,δS]=LE+adS.L_{S}:=[\iota_{E},\delta_{S}]=L_{E}+ad_{S}.

This operator is diagonalisable in the sense that any element βLD,𝔤\beta\in L_{D,\mathfrak{g}} has a Taylor series expansion

β=uβu\beta=\sum_{u}\beta_{u}

where each term satisfies LS(βu)=uβuL_{S}(\beta_{u})=u\beta_{u}. Indeed, the operator adSad_{S} is diagonalizable on 𝔤\mathfrak{g} with finitely many eigenvalues since SS is semisimple. The eigenspaces of LEL_{E} are the weight spaces. We noted earlier that the weight degrees of holomorphic functions are strictly positive integers, and that each weight space is finite dimensional. As an operator on Ωk(logD)\Omega^{\bullet}_{\mathbb{C}^{k}}(\text{log}D), the eigenvalues of LEL_{E} may not be positive, but they are integers which are bounded below. Hence, the eigenvalues of LSL_{S} have the form ui+0u_{i}+\mathbb{Z}_{\geq 0}, for finitely many complex numbers uiu_{i}.

Let LD,𝔤,uL_{D,\mathfrak{g},u} denote the uu-eigenspace, and note that it is finite-dimensional. Because LSL_{S} is a derivation, the Lie bracket respects this decomposition:

[,]:LD,𝔤,u×LD,𝔤,vLD,𝔤,u+v.[-,-]:L_{D,\mathfrak{g},u}\times L_{D,\mathfrak{g},v}\to L_{D,\mathfrak{g},u+v}.

In particular, LD,𝔤,0L_{D,\mathfrak{g},0} is a finite-dimensional Lie subalgebra. The derivations δS\delta_{S} and ιE\iota_{E} commute with LSL_{S}, and hence preserve its eigenspaces. In particular, they restrict to LD,𝔤,0L_{D,\mathfrak{g},0}.

Now introduce the degree 0 derivation P=α0ιEP=\alpha_{0}\iota_{E}. This derivation satisfies P2=PP^{2}=P, and hence induces a decomposition LD,𝔤=ker(P)im(P)L_{D,\mathfrak{g}}=\text{ker}(P)\oplus\text{im}(P). Let U=ker(P)U=\text{ker}(P) and let I=im(P)I=\text{im}(P). With respect to the bracket, UU is a subalgebra and II is an abelian ideal.

Lemma \thelemma.

The derivation ιE\iota_{E} vanishes on UU. For every degree ii it defines an isomorphism

ιE:IiUi1,\iota_{E}:I^{i}\to U^{i-1},

with inverse given by multiplication by α0\alpha_{0}. Therefore, as a graded Lie algebra, LD,𝔤L_{D,\mathfrak{g}} is isomorphic to UU[1]U\ltimes U[-1], where UU acts on U[1]U[-1] via the adjoint action.

Proof.

It is clear that ker(ιE)U=ker(P)\text{ker}(\iota_{E})\subseteq U=\text{ker}(P). For the opposite inclusion, suppose that P(x)=0P(x)=0. Then ιE(x)\iota_{E}(x) is in the kernel of multiplication by α0\alpha_{0}. Since α0\alpha_{0} is a non-vanishing algebroid 11-form, ιE(x)\iota_{E}(x) must be of the form α0y\alpha_{0}\wedge y. But then

0=ιE2(x)=ιE(α0y)=yα0ιE(y),0=\iota^{2}_{E}(x)=\iota_{E}(\alpha_{0}\wedge y)=y-\alpha_{0}\wedge\iota_{E}(y),

which implies that ιE(x)=0\iota_{E}(x)=0, as required. The image of ιE\iota_{E} is contained in UU since ιE2=0\iota_{E}^{2}=0. To see surjectivity, we can explicitely construct the inverse as mulitplication by α0\alpha_{0}. Given xUx\in U, check that α0x=P(α0x)I\alpha_{0}\wedge x=P(\alpha_{0}\wedge x)\in I. Hence α0:Ui1Ii\alpha_{0}\wedge:U^{i-1}\to I^{i}. Then for xUx\in U, we have ιE(α0x)=x\iota_{E}(\alpha_{0}\wedge x)=x, and for P(y)IP(y)\in I we have α0ιEP(y)=P2(y)=P(y)\alpha_{0}\wedge\iota_{E}P(y)=P^{2}(y)=P(y).

Now define the isomorphism Ξ:LD,𝔤UU[1]\Xi:L_{D,\mathfrak{g}}\to U\ltimes U[-1] by the following formula in degree ii:

UiIiUiUi1,(x,y)(x,(1)iιE(y)).U^{i}\oplus I^{i}\to U^{i}\oplus U^{i-1},\qquad(x,y)\mapsto(x,(-1)^{i}\iota_{E}(y)).

This preserves Lie brackets. ∎

The commutator [P,LS]=0[P,L_{S}]=0. Therefore, the two operators can be simultaneously diagonalized. In particular, we have the decomposition LD,𝔤,0=U0I0L_{D,\mathfrak{g},0}=U_{0}\oplus I_{0}. The results of the previous lemma remain true for this subalgebra. Next, we have [P,δS]=α0LS[P,\delta_{S}]=\alpha_{0}L_{S}. If we re-write this as the following identity

PδS=α0LS+δSPP\delta_{S}=\alpha_{0}L_{S}+\delta_{S}P

then we can deduce that δS\delta_{S} preserves II. Indeed, applying this identity to an element of the form x=P(y)x=P(y), we obtain

PδS(x)=α0LSP(y)+δSP2(y)=α0PLS(y)+δSP(y)=δS(x).P\delta_{S}(x)=\alpha_{0}L_{S}P(y)+\delta_{S}P^{2}(y)=\alpha_{0}PL_{S}(y)+\delta_{S}P(y)=\delta_{S}(x).

On the other hand, the differential δS\delta_{S} does not preserve UU. But by applying the identity to an element xUx\in U, we compute that the ‘off-diagonal’ term is given by PδS(x)=α0LS(x)P\delta_{S}(x)=\alpha_{0}L_{S}(x). This term vanishes when we restrict to the subalgebra LD,𝔤,0L_{D,\mathfrak{g},0}. Hence, we obtain the following corollary.

Corollary \thecorollary.

The subalgebra U0U_{0} is preserved by δS\delta_{S}, and there is an isomorphism of dgla’s

(LD,𝔤,0,δS)(U0,δS)(U0,δS)[1].(L_{D,\mathfrak{g},0},\delta_{S})\cong(U_{0},\delta_{S})\ltimes(U_{0},\delta_{S})[-1].
Proof.

On the subspace LD,𝔤,0L_{D,\mathfrak{g},0} we have [ιE,δS]=0[\iota_{E},\delta_{S}]=0. This implies that the morphism Ξ\Xi from Lemma 3 is a chain map. ∎

The gauge group Aut(S)Aut(S)

Viewing S𝔤S\in\mathfrak{g} as a representation of \mathbb{C}, we can pull it back to obtain a representation pSp^{*}S of k\mathbb{C}\ltimes\mathbb{C}^{k}. Let Aut(S)Aut(S) be the subgroup of the gauge group 𝔊\mathfrak{G} consisting of gauge transformations which preserve pSp^{*}S:

Aut(S)={g𝔊|gpS=pS}.Aut(S)=\{g\in\mathfrak{G}\ |\ g\ast p^{*}S=p^{*}S\}.

It is a finite-dimensional algebraic group whose Lie algebra is LD,𝔤,00L^{0}_{D,\mathfrak{g},0}. We recall the description of its Levi decomposition which was given in [4]. The automorphism group of jpS=Sj^{*}p^{*}S=S is GSG_{S}, the centralizer subgroup of SS in GG, which is reductive. The pullback functor jj^{*} defines a homomorphism

j:Aut(S)GS,gg(0),j^{*}:Aut(S)\to G_{S},\qquad g\mapsto g(0),

and the pullback functor pp^{*} defines a splitting. The kernel of jj^{*}, denoted Aut0(S)Aut_{0}(S), is the unipotent radical. Hence the isomorphism Aut(S)Aut0(S)GSAut(S)\cong Aut_{0}(S)\rtimes G_{S} provides the Levi decomposition.

Define the following gauge action of Aut(S)Aut(S) on LD,𝔤,01L^{1}_{D,\mathfrak{g},0}:

gx=gxg1δS(g)g1,g\ast x=gxg^{-1}-\delta_{S}(g)g^{-1},

where δS(g)g1=dgg1+α0(SgSg1).\delta_{S}(g)g^{-1}=dgg^{-1}+\alpha_{0}(S-gSg^{-1}).

Lemma \thelemma.

The gauge action of Aut(S)Aut(S) is well-defined. In terms of the decomposition LD,𝔤,01=U01I01L^{1}_{D,\mathfrak{g},0}=U^{1}_{0}\oplus I^{1}_{0} it is given by

g(x,y)=(gxg1δS(g)g1,gyg1),g\ast(x,y)=(gxg^{-1}-\delta_{S}(g)g^{-1},gyg^{-1}),

where xU01x\in U^{1}_{0} and yI01y\in I^{1}_{0}. Furthermore, Aut(S)Aut(S) acts on LD,𝔤,02L^{\bullet\geq 2}_{D,\mathfrak{g},0} by conjugation, preserving the decomposition U0I0U_{0}\oplus I_{0} and the Lie bracket.

Proof.

A computation shows that LS(gx)=g(LSx)g1L_{S}(g\ast x)=g(L_{S}x)g^{-1} for xLD,𝔤1x\in L^{1}_{D,\mathfrak{g}}, showing that LD,𝔤,01L^{1}_{D,\mathfrak{g},0} is preserved. Similarly, LS(gxg1)=g(LSx)g1L_{S}(gxg^{-1})=g(L_{S}x)g^{-1} for xLD,𝔤jx\in L^{j}_{D,\mathfrak{g}}, showing that the conjugation action preserves LD,𝔤,02L^{\bullet\geq 2}_{D,\mathfrak{g},0}. Next, for xLD,𝔤,0x\in L_{D,\mathfrak{g},0} we have P(gxg1)=gP(x)g1P(gxg^{-1})=gP(x)g^{-1}, implying that the conjugation also preserves U0U_{0} and I0I_{0}. Finally,

P(δS(g)g1)=α0(LE(g)g1+SgSg1),P(\delta_{S}(g)g^{-1})=\alpha_{0}(L_{E}(g)g^{-1}+S-gSg^{-1}),

which vanishes for gAut(S)g\in Aut(S). Hence δS(g)g1U01\delta_{S}(g)g^{-1}\in U^{1}_{0}. ∎

The finite-dimensional derived stack

Given the finite dimensional dgla LD,𝔤,0L_{D,\mathfrak{g},0} we obtain a derived manifold 𝒲S\mathcal{W}_{S}. The base manifold is the vector space WS=LD,𝔤,01W_{S}=L^{1}_{D,\mathfrak{g},0}, the bundle of curved dgla’s is the trivial bundle WS×LD,𝔤,02W_{S}\times L^{\bullet\geq 2}_{D,\mathfrak{g},0}, the curvature section is given by the standard formula FS(w)=δS(w)+12[w,w]F_{S}(w)=\delta_{S}(w)+\frac{1}{2}[w,w], and the twisted differential δ\delta is given by

δS,w=δS+[w,],\delta_{S,w}=\delta_{S}+[w,-],

for wWSw\in W_{S}. Furthermore, Lemma 3 gives an equivariant action of Aut(S)Aut(S) on WS×LD,𝔤,02W_{S}\times L^{\bullet\geq 2}_{D,\mathfrak{g},0}, preserving the bracket. It is also straightforward to check that this action preserves FSF_{S} and δ\delta. Hence, we obtain a derived stack [𝒲S/Aut(S)][\mathcal{W}_{S}/Aut(S)].

U0U_{0} is a sub-dgla of LD,𝔤,0L_{D,\mathfrak{g},0}, which is preserved by the action of Aut(S)Aut(S). Hence, it gives rise to a derived substack [𝒰S/Aut(S)][\mathcal{U}_{S}/Aut(S)] of [𝒲S/Aut(S)][\mathcal{W}_{S}/Aut(S)]. Furthermore, since I0I_{0} is an ideal of LD,𝔤,0L_{D,\mathfrak{g},0}, we also get a projection morphism [𝒲S/Aut(S)][𝒰S/Aut(S)].[\mathcal{W}_{S}/Aut(S)]\to[\mathcal{U}_{S}/Aut(S)].

Proposition \theproposition.

The derived stack [𝒲S/Aut(S)][\mathcal{W}_{S}/Aut(S)] is isomorphic to the shifted tangent bundle T[1][𝒰S/Aut(S)]T[-1][\mathcal{U}_{S}/Aut(S)].

Proof.

This follows from Corollary 3 and Lemma 3. ∎

We are actually interested in a substack of [𝒲S/Aut(S)][\mathcal{W}_{S}/Aut(S)] which is determined by the element A=S+N0A=S+N_{0}. Recall that the image of Aut(S)Aut(S) under jj^{*} is GSG_{S}, the centralizer of SS. This implies that for any element ωWS\omega\in W_{S}, the image ji(ω)𝔤S=Lie(GS)j^{*}i^{*}(\omega)\in\mathfrak{g}_{S}=Lie(G_{S}). We will require that this element be contained in GSN0G_{S}\ast N_{0}, the adjoint orbit of N0N_{0} in 𝔤S\mathfrak{g}_{S}. Namely, define

W(A)={ωWS|ji(ω)GSN0}.W(A)=\{\omega\in W_{S}\ |\ j^{*}i^{*}(\omega)\in G_{S}\ast N_{0}\}.

Let 𝒲(A)\mathcal{W}(A) be the derived manifold obtained by pulling back the bundle of curved dgla’s from WSW_{S} to W(A)W(A). The action of Aut(S)Aut(S) restricts to an action on this sub-manifold. Hence, we obtain a derived stack [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)].

Theorem 3.1.

[𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] is equivalent to [D,𝔤(A)/𝔊][\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}], the derived stack of logarithmic flat connections whose residue lies in the adjoint orbit OAO_{A} of AA.

4 Proof of Theorem 3.1

In this section we will give the proof of the equivalence between [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] and [D,𝔤(A)/𝔊][\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}]. There is a natural morphism

q:[𝒲(A)/Aut(S)][D,𝔤(A)/𝔊],q:[\mathcal{W}(A)/Aut(S)]\to[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}],

which we describe as follows:

  1. 1.

    The map on the base manifolds is given by the following formula

    q:W(A)M(A),ωα0S+ω.q:W(A)\to M(A),\qquad\omega\mapsto\alpha_{0}S+\omega.
  2. 2.

    The map on bundles of curved dgla is given by the inclusion LD,𝔤,02LD,𝔤2L^{\bullet\geq 2}_{D,\mathfrak{g},0}\to L^{\bullet\geq 2}_{D,\mathfrak{g}}.

  3. 3.

    The group Aut(S)Aut(S) includes into 𝔊\mathfrak{G} as a subgroup, and the map qq is equivariant.

In order to show that qq is an equivalence, we must show two things. First, there is an underlying functor between the classical groupoids:

π0(q):Aut(S)MC(𝒲(A))𝔊MC(D,𝔤(A)).\pi_{0}(q):Aut(S)\ltimes MC(\mathcal{W}(A))\to\mathfrak{G}\ltimes MC(\mathcal{M}_{D,\mathfrak{g}}(A)).

We need to show that this is an equivalence of categories. This is implied by [4, Theorem 5.5] and the following lemma.

Lemma \thelemma.

Let ωW(A)\omega\in W(A). Then ιE(ω)\iota_{E}(\omega) is nilpotent.

Proof.

For ωLD,𝔤,01\omega\in L_{D,\mathfrak{g},0}^{1}, we have ιE(ω)U00=Lie(Aut(S))\iota_{E}(\omega)\in U^{0}_{0}=Lie(Aut(S)). If ωW(A)\omega\in W(A) we have in addition that jιE(ω)GSN0j^{*}\iota_{E}(\omega)\in G_{S}\ast N_{0}, and so is nilpotent. Let ιE(ω)=Bs+Bn\iota_{E}(\omega)=B_{s}+B_{n} be the Jordan decomposition, where BsB_{s} is semisimple and BnB_{n} is nilpotent. Then j(Bs)=0j^{*}(B_{s})=0, so that BsLie(Aut0(S))B_{s}\in Lie(Aut_{0}(S)). But since Aut0(S)Aut_{0}(S) is unipotent, this implies that Bs=0B_{s}=0, and hence ιE(ω)\iota_{E}(\omega) is nilpotent. ∎

Second, a derived stack has a tangent complex at every point of its MC locus, and the map qq induces a chain map between the tangent complexes:

dqw:𝕋w[𝒲(A)/Aut(S)]𝕋q(w)[D,𝔤(A)/𝔊].dq_{w}:\mathbb{T}_{w}[\mathcal{W}(A)/Aut(S)]\to\mathbb{T}_{q(w)}[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}].

We need to show that this is a quasi-isomorphism at each point of the MC locus. We will do this by first constructing an explicit homotopy at the special point q(0)q(0) (which is generally not in our space), and then apply the homological perturbation lemma to obtain the quasi-isomorphism at all points.

The homotopy

Let a:LD,𝔤,0LD,𝔤a:L_{D,\mathfrak{g},0}\to L_{D,\mathfrak{g}} be the inclusion and let b:LD,𝔤LD,𝔤,0b:L_{D,\mathfrak{g}}\to L_{D,\mathfrak{g},0} be the projection to the degree 0 component. Both aa and bb are chain maps with respect to δS\delta_{S}, but in general only aa preserves the Lie bracket. Furthermore, ba=idLD,𝔤,0b\circ a=id_{L_{D,\mathfrak{g},0}}.

Recall that a given element βLD,𝔤\beta\in L_{D,\mathfrak{g}} has a Taylor expansion in the eigenvalues of LSL_{S}:

β=uβu,\beta=\sum_{u}\beta_{u},

where each term satisfies LS(βu)=uβuL_{S}(\beta_{u})=u\beta_{u}. As we saw, the eigenvalues have the form ui+0u_{i}+\mathbb{Z}_{\geq 0} for finitely many complex numbers uiu_{i}. For this reason the series

β=u01uβu\beta^{\prime}=\sum_{u\neq 0}\frac{1}{u}\beta_{u}

converges to a well-defined element of LD,𝔤L_{D,\mathfrak{g}}. We use this to define the following degree 1-1 operator

h:LD,𝔤iLD,𝔤i1,uβuιE(u01uβu).h:L^{i}_{D,\mathfrak{g}}\to L^{i-1}_{D,\mathfrak{g}},\qquad\sum_{u}\beta_{u}\mapsto\iota_{E}(\sum_{u\neq 0}\frac{1}{u}\beta_{u}).

The following lemma results from straightforward computation. It has the upshot that aa defines a quasi-isomorphism of dgla’s from (LD,𝔤,0,δS)(L_{D,\mathfrak{g},0},\delta_{S}) to (LD,𝔤,δS)(L_{D,\mathfrak{g}},\delta_{S}).

Lemma \thelemma.

The operator hh defines a homotopy between abab and idLD,𝔤id_{L_{D,\mathfrak{g}}}. In other words, it satisfies

[δS,h]=idLD,𝔤ab.[\delta_{S},h]=id_{L_{D,\mathfrak{g}}}-ab.

Furthermore, it satisfies the ‘side conditions’ ha=0h\circ a=0, bh=0b\circ h=0 and h2=0h^{2}=0. Finally, it vanishes on UU and sends IiI^{i} to Ui1U^{i-1}.

The perturbation

We will now perturb the differential δS\delta_{S} and show that aa continues to define a quasi-isomorphism. This is achieved by using the perturbation lemma [13].

Let wW(A)w\in W(A) satisfy the Maurer-Cartan equation δS(w)+12[w,w]=0\delta_{S}(w)+\frac{1}{2}[w,w]=0 and consider the perturbed differential δS,w=δS+[w,]\delta_{S,w}=\delta_{S}+[w,-]. This is a differential on LD,𝔤L_{D,\mathfrak{g}}, and we want an induced perturbation of the homotopy data (a,b,h,δS)(a,b,h,\delta_{S}) of the previous section.

Lemma \thelemma.

The endomorphism adwhad_{w}\circ h of LD,𝔤L_{D,\mathfrak{g}} is nilpotent.

Proof.

The element ww can be decomposed as w=γ+α0Nw=\gamma+\alpha_{0}N, where γU01\gamma\in U^{1}_{0} and NU00N\in U^{0}_{0}. By Lemma 4, NN is nilpotent. Recall from Lemma 4 that hh vanishes on UU and its image is contained in UU. Furthermore, since UU is a subalgebra of LD,𝔤L_{D,\mathfrak{g}}, it is preserved by adγad_{\gamma}. As a result hadγh=0h\circ ad_{\gamma}\circ h=0. Hence, it suffices to show that the operator α0adNh\alpha_{0}ad_{N}\circ h is nilpotent.

Now note that adNad_{N} and multiplication by α0\alpha_{0} commute. Since NU00N\in U^{0}_{0}, adNad_{N} also commutes with hh. This implies that

(hα0adN)k=(adN)kh~k,(h\circ\alpha_{0}ad_{N})^{k}=(ad_{N})^{k}\circ\tilde{h}^{k},

where h~\tilde{h} is the operator h~(β)=h(α0β)\tilde{h}(\beta)=h(\alpha_{0}\wedge\beta). But this will vanish for large enough kk since NN is nilpotent. ∎

The upshot of this lemma is that we can now define the following perturbed maps:

h\displaystyle h^{\prime} =p=0(hadw)ph,\displaystyle=\sum_{p=0}^{\infty}(-had_{w})^{p}h,
δ\displaystyle\delta^{\prime} =δS+p=0b(adwh)padwa,\displaystyle=\delta_{S}+\sum_{p=0}^{\infty}b(-ad_{w}h)^{p}ad_{w}a,
a\displaystyle a^{\prime} =p=0(hadw)pa,\displaystyle=\sum_{p=0}^{\infty}(-had_{w})^{p}a,
b\displaystyle b^{\prime} =p=0b(adwh)p.\displaystyle=\sum_{p=0}^{\infty}b(-ad_{w}h)^{p}.

The perturbation lemma says that δ\delta^{\prime} defines a differential on LD,𝔤,0L_{D,\mathfrak{g},0}, that aa^{\prime} and bb^{\prime} define chain maps between (LD,𝔤,0,δ)(L_{D,\mathfrak{g},0},\delta^{\prime}) and (LD,𝔤,δS,w)(L_{D,\mathfrak{g}},\delta_{S,w}), and that the following equations are satisfied:

ba=idLD,𝔤,0,[δS,w,h]=idLD,𝔤ab,ha=0,bh=0,hh=0.b^{\prime}\circ a^{\prime}=id_{L_{D,\mathfrak{g},0}},\ \ [\delta_{S,w},h^{\prime}]=id_{L_{D,\mathfrak{g}}}-a^{\prime}\circ b^{\prime},\ \ h^{\prime}\circ a^{\prime}=0,\ \ b^{\prime}\circ h^{\prime}=0,\ \ h^{\prime}\circ h^{\prime}=0.

The following lemma identifies the perturbations.

Lemma \thelemma.

The perturbations are given by

a=a,b=b,δ=δS,w.a^{\prime}=a,\ \ b^{\prime}=b,\ \ \delta^{\prime}=\delta_{S,w}.

In particular, the inclusion a:(LD,𝔤,0,δS,w)(LD,𝔤,δS,w)a:(L_{D,\mathfrak{g},0},\delta_{S,w})\to(L_{D,\mathfrak{g}},\delta_{S,w}) is a quasi-isomorphism. Furthermore, hh^{\prime} vanishes on UU and sends II to UU.

Proof.

The element wLD,𝔤,01w\in L_{D,\mathfrak{g},0}^{1} and so adwad_{w} restricts to LD,𝔤,0L_{D,\mathfrak{g},0} and commutes with both aa and bb. As a result of this and the side conditions of Lemma 4, we have that hadwa=0had_{w}a=0 and badwh=0bad_{w}h=0. Plugging this into the definitions of the deformed maps gives

a\displaystyle a^{\prime} =ap0(hadw)p(hadwa)=a,\displaystyle=a-\sum_{p\geq 0}(-had_{w})^{p}(had_{w}a)=a,
δ\displaystyle\delta^{\prime} =δS+badwap0b(adwh)p1adw(hadwa)=δS,w,\displaystyle=\delta_{S}+bad_{w}a-\sum_{p\geq 0}b(-ad_{w}h)^{p-1}ad_{w}(had_{w}a)=\delta_{S,w},
b\displaystyle b^{\prime} =bp0(badwh)(adwh)p=b.\displaystyle=b-\sum_{p\geq 0}(bad_{w}h)(-ad_{w}h)^{p}=b.

The statement about hh^{\prime} follows from Lemma 4 and the fact that each term in the definition of hh^{\prime} starts and ends with hh. ∎

The quasi-isomorphism of tangent complexes

Consider a point wMC(𝒲(A))w\in MC(\mathcal{W}(A)). It has the form w=γ+α0Nw=\gamma+\alpha_{0}N, where γU01\gamma\in U^{1}_{0}, NU00N\in U^{0}_{0}, and ji(w)=N(0)GSN0j^{*}i^{*}(w)=N(0)\in G_{S}\ast N_{0}. It has corresponding point q(w)MC(D,𝔤(A))q(w)\in MC(\mathcal{M}_{D,\mathfrak{g}}(A)). In this section we will describe the morphism of tangent complexes dqw:𝕋w[𝒲(A)/Aut(S)]𝕋q(w)[D,𝔤(A)/𝔊]dq_{w}:\mathbb{T}_{w}[\mathcal{W}(A)/Aut(S)]\to\mathbb{T}_{q(w)}[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}] and show that it is a quasi-isomorphism.

We start by describing the tangent complexes. First, the tangent complex of [D,𝔤(A)/𝔊][\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}] is given as follows:

𝕋q(w)[D,𝔤(A)/𝔊]=LD,𝔤0Tq(w)M(A)LD,𝔤2LD,𝔤3\mathbb{T}_{q(w)}[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}]=L^{0}_{D,\mathfrak{g}}\to T_{q(w)}M(A)\to L^{2}_{D,\mathfrak{g}}\to L^{3}_{D,\mathfrak{g}}\to...

Note that the first term is LD,𝔤0=Lie(𝔊)L^{0}_{D,\mathfrak{g}}=Lie(\mathfrak{G}), and the second term is the subspace

Tq(w)M(A)={vLD,𝔤1|ji(v)T(S+N(0))OA},T_{q(w)}M(A)=\{v\in L^{1}_{D,\mathfrak{g}}\ |\ j^{*}i^{*}(v)\in T_{(S+N(0))}O_{A}\},

where we use the fact that ji(q(w))=S+N(0)j^{*}i^{*}(q(w))=S+N(0). The first map is the derivative of the gauge action, and a computation shows that it is equal to δS,w-\delta_{S,w}. The minus sign is due to the fact that we are making the gauge group act on the left. For simplicity we will replace this by δS,w\delta_{S,w}, since it does not affect the cohomology. The second map is the derivative of the curvature dFdF, and a calculation shows that it is given by δS,w\delta_{S,w}. Finally, all higher maps are given by δq(w)=δS,w\delta_{q(w)}=\delta_{S,w}. Therefore, the tangent complex is a subcomplex of (LD,𝔤,δS,w)(L_{D,\mathfrak{g}},\delta_{S,w}).

The tangent complex of [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] has a similar descriptions. It is given by

𝕋w[𝒲(A)/Aut(S)]=LD,𝔤,00TwW(A)LD,𝔤,02LD,𝔤,02\mathbb{T}_{w}[\mathcal{W}(A)/Aut(S)]=L^{0}_{D,\mathfrak{g},0}\to T_{w}W(A)\to L^{2}_{D,\mathfrak{g},0}\to L^{2}_{D,\mathfrak{g},0}\to...

As above, LD,𝔤,00=Lie(Aut(S))L^{0}_{D,\mathfrak{g},0}=Lie(Aut(S)) and the second term is the subspace

TwW(A)={vLD,𝔤,01|ji(v)TN(0)(GSN0)}.T_{w}W(A)=\{v\in L^{1}_{D,\mathfrak{g},0}\ |\ j^{*}i^{*}(v)\in T_{N(0)}(G_{S}\ast N_{0})\}.

Again all maps are given by δS,w\delta_{S,w} (the first map has a minus sign, which we remove for simplicity). Hence, the tangent complex is a subcomplex of (LD,𝔤,0,δS,w)(L_{D,\mathfrak{g},0},\delta_{S,w}).

The map dqwdq_{w} is easily seen to coincide with aa. Therefore, in order to prove that dqwdq_{w} is a quasi-isomorphism, it suffices to show that the homotopy data (a,b,h,δS,w)(a,b,h^{\prime},\delta_{S,w}) restricts to the tangent complexes.

Lemma \thelemma.

The maps (a,b,h,δS,w)(a,b,h^{\prime},\delta_{S,w}) restrict to 𝕋q(w)[D,𝔤(A)/𝔊]\mathbb{T}_{q(w)}[\mathcal{M}_{D,\mathfrak{g}}(A)/\mathfrak{G}] and 𝕋w[𝒲(A)/Aut(S)]\mathbb{T}_{w}[\mathcal{W}(A)/Aut(S)]. Therefore, dqwdq_{w} defines a quasi-isomorphism.

Proof.

Since the complexes are modified in degree 11, it suffices to restrict our attention to degrees 0,1,20,1,2. The above description of the tangent complexes and dqwdq_{w} immediately implies that aa and δS,w\delta_{S,w} restrict. To check that hh^{\prime} restricts, we only need to show that h(LD,𝔤2)h^{\prime}(L^{2}_{D,\mathfrak{g}}) is contained in Tq(w)M(A)T_{q(w)}M(A). But this follows because, by Lemma 4, the image of hh^{\prime} is contained in UU.

For the map bb, consider a point βTq(w)M(A)\beta\in T_{q(w)}M(A). This can be expanded as β=uβu\beta=\sum_{u}\beta_{u}, where each term satisfies LS(βu)=uβuL_{S}(\beta_{u})=u\beta_{u}. By definition b(β)=β0b(\beta)=\beta_{0}. Hence, we need to check that if ji(β)T(S+N(0))OAj^{*}i^{*}(\beta)\in T_{(S+N(0))}O_{A}, then ji(β0)TN(0)(GSN0)j^{*}i^{*}(\beta_{0})\in T_{N(0)}(G_{S}\ast N_{0}). These tangent spaces have the following descriptions

T(S+N(0))OA=Im(adS+N(0):𝔤𝔤),TN(0)(GSN0)=Im(adN(0):𝔤S𝔤S).T_{(S+N(0))}O_{A}=Im(ad_{S+N(0)}:\mathfrak{g}\to\mathfrak{g}),\qquad T_{N(0)}(G_{S}\ast N_{0})=Im(ad_{N(0)}:\mathfrak{g}_{S}\to\mathfrak{g}_{S}).

Now using the eigenvector expansion, we have

ji(β)=uji(βu)=adS+N(0)(Z),j^{*}i^{*}(\beta)=\sum_{u}j^{*}i^{*}(\beta_{u})=ad_{S+N(0)}(Z),

for some Z𝔤Z\in\mathfrak{g}. One can check that each term in the summand satisfies adS(ji(βu))=uji(βu)ad_{S}(j^{*}i^{*}(\beta_{u}))=uj^{*}i^{*}(\beta_{u}). Since adS:𝔤𝔤ad_{S}:\mathfrak{g}\to\mathfrak{g} is diagonalizable, we can decompose ZZ into eigenvectors as well: Z=uZuZ=\sum_{u}Z_{u}. And since adS+N(0)ad_{S+N(0)} commutes with adSad_{S}, it preserves the eigenspaces. Hence, we can match up the eigenvectors to get

jiβ0=adS+N(0)(Z0)=adN(0)(Z0),j^{*}i^{*}\beta_{0}=ad_{S+N(0)}(Z_{0})=ad_{N(0)}(Z_{0}),

where Z0𝔤SZ_{0}\in\mathfrak{g}_{S}. ∎

5 Plane curves xpyqx^{p}-y^{q}

In this section we give a detailed study of the case of plane curves. Consider

f=xpyq:2,f=x^{p}-y^{q}:\mathbb{C}^{2}\to\mathbb{C},

where pp and qq are relatively prime positive integers satisfying p<qp<q. This function is weighted homogeneous of degree qpqp for the Euler vector field E=qxx+pyyE=qx\partial_{x}+py\partial_{y}, which defines the weight grading on coordinates |x|=q|x|=q and |y|=p|y|=p. The logarithmic tangent bundle is generated by the vector fields EE and V=qyq1x+pxp1yV=qy^{q-1}\partial_{x}+px^{p-1}\partial_{y}, which satisfy [E,V]=(qppq)V[E,V]=(qp-p-q)V. Let w0=qppqw_{0}=qp-p-q. The logarithmic 11-form α0=1qpdlogf\alpha_{0}=\frac{1}{qp}d\text{log}f pairs with VV to give 0. Therefore, it can be completed to a dual basis α0,β\alpha_{0},\beta of the logarithmic cotangent bundle. The form α0\alpha_{0} is closed, and β\beta satisfies dβ=(p+qqp)α0βd\beta=(p+q-qp)\alpha_{0}\wedge\beta.

Cohomology of VV

Let 𝒪w\mathcal{O}_{w} denote the subspace of polynomial functions with weight ww with respect to EE. Note that any integer ww\in\mathbb{Z} has a unique decomposition w=aq+bp+cqpw=aq+bp+cqp, where a,b,ca,b,c\in\mathbb{Z}, 0a<p0\leq a<p and 0b<q0\leq b<q. This decomposition provides a useful way of indexing the weights because of the following lemma.

Lemma \thelemma.

Let w=aq+bp+cqpw=aq+bp+cqp, with the above restrictions on a,b,ca,b,c. The dimension of 𝒪w\mathcal{O}_{w} is max(c,1)+1\max(c,-1)+1, and a basis is given by

xa+cpyb,xa+(c1)pyb+q,,xayb+cq.x^{a+cp}y^{b},x^{a+(c-1)p}y^{b+q},...,x^{a}y^{b+cq}.

The vector field VV has weight w0w_{0}, and hence it defines a map

V:𝒪w𝒪w+w0.V:\mathcal{O}_{w}\to\mathcal{O}_{w+w_{0}}.
Lemma \thelemma.

The kernel of VV is [f]\mathbb{C}[f].

Proof.

A calculation shows that V(fc)=0V(f^{c})=0. Conversely, let gker(V)g\in\text{ker}(V). Because VV is homogeneous, it suffices to consider the case where gg is homogeneous of weight w>0w>0. The equation V(g)=0V(g)=0 implies that xg=pxp1h\partial_{x}g=px^{p-1}h and yg=qyq1h\partial_{y}g=-qy^{q-1}h, for a common polynomial hh. Therefore,

wg=E(g)=qxxg+pyyg=qp(xpyq)h,wg=E(g)=qx\partial_{x}g+py\partial_{y}g=qp(x^{p}-y^{q})h,

so that g=qpwfhg=\frac{qp}{w}fh. Hence hh is a function of weight wqpw-qp and it lies in the kernel of VV. The result now follows by induction on the weight. ∎

The Jacobian ideal of ff is generated by xp1x^{p-1} and yq1y^{q-1}. Let C=[x,y]/(xp1,yq1)C=\mathbb{C}[x,y]/(x^{p-1},y^{q-1}), considered as a \mathbb{C}-vector space. It has a natural basis of monomials xaybx^{a}y^{b}, where 0a<p10\leq a<p-1 and 0b<q10\leq b<q-1. Using this basis, CC is naturally graded by weight, and there is a weight preserving injective linear map C[x,y]C\to\mathbb{C}[x,y]. Consider the graded polynomial ring [f]\mathbb{C}[f], where ff has degree pqpq. Then [x,y]\mathbb{C}[x,y] is a graded [f]\mathbb{C}[f]-module and there is a morphism of graded [f]\mathbb{C}[f]-modules

[f]C[x,y].\mathbb{C}[f]\otimes_{\mathbb{C}}C\to\mathbb{C}[x,y].

The action of VV on [x,y]\mathbb{C}[x,y] is [f]\mathbb{C}[f]-linear, so that the cokernel 𝖼𝗈𝗄𝖾𝗋(V)\mathsf{coker}(V) is also a [f]\mathbb{C}[f]-module. Post-composing with the quotient projection, we obtain the morphism

[f]C𝖼𝗈𝗄𝖾𝗋(V).\mathbb{C}[f]\otimes_{\mathbb{C}}C\to\mathsf{coker}(V).
Lemma \thelemma.

The morphism [f]C𝖼𝗈𝗄𝖾𝗋(V)\mathbb{C}[f]\otimes_{\mathbb{C}}C\to\mathsf{coker}(V) is an isomorphism of graded [f]\mathbb{C}[f]-modules.

Proof.

Since VV is homogeneous it suffices to consider a single weight at a time: we consider the cokernel of the map V:𝒪ww0𝒪wV:\mathcal{O}_{w-w_{0}}\to\mathcal{O}_{w}. Let w=aq+bp+cqpw=aq+bp+cqp, where 0a<p0\leq a<p, 0b<q0\leq b<q and c0c\geq 0, so that 𝒪w\mathcal{O}_{w} has dimension c+1c+1. Then ww0=(a+1)q+(b+1)p+(c1)qpw-w_{0}=(a+1)q+(b+1)p+(c-1)qp. If a<p1a<p-1 and b<q1b<q-1 then 𝒪ww0\mathcal{O}_{w-w_{0}} has dimension cc. Furthermore VV is injective because ww0w-w_{0} is not a multiple of qpqp. Hence 𝖼𝗈𝗄𝖾𝗋(V)w\mathsf{coker}(V)_{w} is 11-dimensional. If a=p1a=p-1 and b<q1b<q-1, then ww0=(b+1)p+cqpw-w_{0}=(b+1)p+cqp, so 𝒪ww0\mathcal{O}_{w-w_{0}} has dimension c+1c+1, VV is injective, and hence 𝖼𝗈𝗄𝖾𝗋(V)w=0\mathsf{coker}(V)_{w}=0. The same argument applies to the case a<p1a<p-1 and b=q1b=q-1. The only remaining case is a=p1a=p-1 and b=q1b=q-1. In this case ww0=(c+1)qpw-w_{0}=(c+1)qp and 𝒪ww0\mathcal{O}_{w-w_{0}} has dimension c+2c+2. But now VV has a 11-dimensional kernel and so 𝖼𝗈𝗄𝖾𝗋(V)w=0\mathsf{coker}(V)_{w}=0.

The upshot is that the cokernel is non-zero precisely when a<p1a<p-1 and b<q1b<q-1, in which case it is 11-dimensional. These dimensions match with the dimensions of [f]C\mathbb{C}[f]\otimes_{\mathbb{C}}C. Hence it suffices for us to prove that fcxaybf^{c}x^{a}y^{b} is not in the image of VV. We will do this by proving that the following map

M:𝒪ww0𝒪w,(λ,g)λfcxayb+V(g)M:\mathbb{C}\oplus\mathcal{O}_{w-w_{0}}\to\mathcal{O}_{w},\qquad(\lambda,g)\mapsto\lambda f^{c}x^{a}y^{b}+V(g)

is represented by a matrix with positive determinant, using the bases of Lemma 5. Applying VV to the element xa+1+ipyb+1+jqx^{a+1+ip}y^{b+1+jq} yields

p(1+b+jq)xa+(i+1)pyb+jq+q(1+a+ip)xa+ipyb+(j+1)q.p(1+b+jq)x^{a+(i+1)p}y^{b+jq}+q(1+a+ip)x^{a+ip}y^{b+(j+1)q}.

The salient thing to note is that the basis elements are consecutive and the coefficients are positive. Hence VV is represented by a (c+1)×c(c+1)\times c matrix such that column ii has positive entries in rows ii and i+1i+1 and 0 for the remaining rows. Using the binomial theorem, fcxayb=k=0c(1)kxa+(ck)pyb+kqf^{c}x^{a}y^{b}=\sum_{k=0}^{c}(-1)^{k}x^{a+(c-k)p}y^{b+kq}. The salient point here is that the terms are non-zero with alternating signs. These give the entries of the first column of the matrix MM. Computing the determinant of MM using the Laplace expansion along the first column shows that it is positive. ∎

The dgla (U0,δS)(U_{0},\delta_{S})

Now we choose a Lie algebra 𝔤\mathfrak{g} and a semisimple element S𝔤S\in\mathfrak{g}. This induces an eigenspace decomposition of the Lie algebra

𝔤=λ𝔤λ,\mathfrak{g}=\bigoplus_{\lambda}\mathfrak{g}_{\lambda},

where 𝔤λ\mathfrak{g}_{\lambda} is the eigenspace of adSad_{S} with eigenvalue λ\lambda. We will use the following convention: if λ\lambda is not an eigenvalue of adSad_{S}, then 𝔤λ=0\mathfrak{g}_{\lambda}=0. Note that the decomposition is preserved by the bracket: [𝔤λ,𝔤μ]𝔤λ+μ[\mathfrak{g}_{\lambda},\mathfrak{g}_{\mu}]\subseteq\mathfrak{g}_{\lambda+\mu}.

The dgla (U0,δS)(U_{0},\delta_{S}) has terms in degrees 0 and 11. They are given by

U00=w0𝒪w𝔤w,U01=w0𝒪w𝔤w0wβ,U^{0}_{0}=\bigoplus_{w\geq 0}\mathcal{O}_{w}\otimes\mathfrak{g}_{-w},\qquad U^{1}_{0}=\bigoplus_{w\geq 0}\mathcal{O}_{w}\otimes\mathfrak{g}_{w_{0}-w}\beta,

with δS\delta_{S} given by applying VV. We will sometimes drop β\beta from the notation.

Applying Lemmas 5 and 5 we obtain the following description of the cohomology of (U0,δS)(U_{0},\delta_{S}).

Corollary \thecorollary.

The cohomology of (U0,δS)(U_{0},\delta_{S}) is given as follows

H0(U0)=c0fc𝔤cpq,H1(U0)w0([f]C)w𝔤w0w.H^{0}(U_{0})=\bigoplus_{c\geq 0}f^{c}\mathfrak{g}_{-cpq},\qquad H^{1}(U_{0})\cong\bigoplus_{w\geq 0}(\mathbb{C}[f]\otimes C)_{w}\otimes\mathfrak{g}_{w_{0}-w}.

Furthermore, the graded Lie algebra H(U0)H^{\bullet}(U_{0}) with zero differential naturally embeds into (U0,δS)(U_{0},\delta_{S}) as a quasi-isomorphic sub-dgla.

Let a𝔤a\in\mathfrak{g} be a real semisimple element. Recall from [5] that this determines a parabolic subgroup of GG

P(a)={gG|limz0zagza exists in G along any ray},P(a)=\{g\in G\ |\ \lim_{z\to 0}z^{a}gz^{-a}\text{ exists in $G$ along any ray}\},

where zz\in\mathbb{C} and za=exp(log(z)a)z^{a}=\exp(\text{log}(z)a). Decomposing SS into real and imaginary parts, S=a+ibS=a+ib, we can define the following subgroup of GG

PS:=CG(e2πiqpS)P(aqp).P_{S}:=C_{G}(e^{\frac{-2\pi i}{qp}S})\cap P(\frac{-a}{qp}).

In this definition CG(e2πiqpS)C_{G}(e^{\frac{-2\pi i}{qp}S}) is the centralizer of e2πiqpSe^{\frac{-2\pi i}{qp}S} in GG. It is reductive but possibly disconnected. Let CSC_{S} denote the connected component of the identity. The group PSP_{S} is the parabolic subgroup of CG(e2πiqpS)C_{G}(e^{\frac{-2\pi i}{qp}S}) (or CSC_{S}) determined by the element aqp\frac{-a}{qp} and it is connected. The reductive quotient of PSP_{S} is GSG_{S}, the centralizer of SS in GG. Denote the quotient map χ:PSGS\chi:P_{S}\to G_{S}.

Lemma \thelemma.

The group PSP_{S} embeds into Aut(S)Aut(S) as the subgroup integrating H0(U0)H^{0}(U_{0}). The gauge action of PSP_{S} preserves H1(U0)U01H^{1}(U_{0})\subset U^{1}_{0} and is linear. Hence, we have a Lie subgroupoid

PSH1(U0)Aut(S)U01.P_{S}\ltimes H^{1}(U_{0})\subseteq Aut(S)\ltimes U^{1}_{0}.
Proof.

Let 2\mathbb{C}\ltimes\mathbb{C}^{2} be the Lie algebroid generated by the action of EE and let \mathbb{C}\ltimes\mathbb{C} be the Lie algebroid generated by the action of zzz\partial_{z}. The following defines a Lie algebroid morphism

f:2,(λ,x,y)(pqλ,f(x,y)),f:\mathbb{C}\ltimes\mathbb{C}^{2}\to\mathbb{C}\ltimes\mathbb{C},\qquad(\lambda,x,y)\mapsto(pq\lambda,f(x,y)),

and under this map, the logarithmic connection d+1qpSdlogzd+\frac{1}{qp}Sd\text{log}z pulls back to pSp^{*}S. As a result, the pullback defines an embedding of automorphism groups from Aut(d+1qpSdlogz)Aut(S)Aut(d+\frac{1}{qp}Sd\text{log}z)\to Aut(S). In [3, Proposition 3.4] it is shown that restricting an automorphism to 11\in\mathbb{C} defines an embedding of Aut(d+1qpSdlogz)Aut(d+\frac{1}{qp}Sd\text{log}z) into GG which identifies it with PSP_{S}. Furthermore, the Lie algebra of PSP_{S} is identified with c0zc𝔤cpq\bigoplus_{c\geq 0}z^{c}\mathfrak{g}_{-cpq}, and under the pullback, this is sent isomorphically to H0(U0)H^{0}(U_{0}). Finally, since the action of H0(U0)H^{0}(U_{0}) preserves H1(U0)H^{1}(U_{0}) and is linear, the same is true of PSP_{S}. ∎

Given the semisimple element S𝔤S\in\mathfrak{g}, we say that it is large enough if all the positive integer eigenvalues of adSad_{S} are strictly greater than w0w_{0}.

Proposition \theproposition.

The inclusion PSH1(U0)Aut(S)U01P_{S}\ltimes H^{1}(U_{0})\subseteq Aut(S)\ltimes U^{1}_{0} is a Morita equivalence if SS is large enough.

Proof.

First, because of the assumption on SS and the fact that 𝒪w0=0\mathcal{O}_{w_{0}}=0 (see Lemma 5), the vector space U01U^{1}_{0} has the following form

U01=w>w0𝒪w𝔤w0wβ.U^{1}_{0}=\bigoplus_{w>w_{0}}\mathcal{O}_{w}\otimes\mathfrak{g}_{w_{0}-w}\beta.

We now proceed in several steps.

  1. 1.

    Claim: The subspace H1(U0)H^{1}(U_{0}) intersects every orbit of Aut(S)Aut(S). Given γU01\gamma\in U^{1}_{0}, we need to find an element of Aut(S)Aut(S) which sends γ\gamma into H1(U0)H^{1}(U_{0}). We do this iteratively following the usual proof of the normal form for ODEs with Fuchsian singularities. First, we expand γ=w>w0γw\gamma=\sum_{w>w_{0}}\gamma_{w}, where γw𝒪w𝔤w0wβ\gamma_{w}\in\mathcal{O}_{w}\otimes\mathfrak{g}_{w_{0}-w}\beta. Given a weight w>w0w^{\prime}>w_{0}, let u𝒪ww0𝔤w0wu\in\mathcal{O}_{w^{\prime}-w_{0}}\otimes\mathfrak{g}_{w_{0}-w^{\prime}}, and consider the action of euAut(S)e^{u}\in Aut(S) on γ\gamma:

    euγ=euγeuV(eu)euβ.e^{u}\ast\gamma=e^{u}\gamma e^{-u}-V(e^{u})e^{-u}\beta.

    We claim that γ\gamma is modified in weights ww^{\prime} and higher. Indeed, expanding we get

    V(eu)eu=V(u)V(u)u+12V(u2)+V(e^{u})e^{-u}=V(u)-V(u)u+\frac{1}{2}V(u^{2})+...

    The first term has weight ww^{\prime}. All other terms have higher weights since ww0>0w^{\prime}-w_{0}>0. Expanding the term euγweue^{u}\gamma_{w}e^{-u} gives

    exp(adu)γw=γw+[u,γw]+12[u,[u,γw]]+exp(ad_{u})\gamma_{w}=\gamma_{w}+[u,\gamma_{w}]+\frac{1}{2}[u,[u,\gamma_{w}]]+...

    The second term has weight ww0+w>ww^{\prime}-w_{0}+w>w^{\prime}, since w>w0w>w_{0}. Note that the action on weight ww^{\prime} is given by γwγwV(u)β\gamma_{w^{\prime}}\mapsto\gamma_{w^{\prime}}-V(u)\beta.

    By Lemma 5, the element γw\gamma_{w^{\prime}} can be decomposed as

    γw=fcxaybXβ+V(u)β,\gamma_{w^{\prime}}=f^{c}x^{a}y^{b}\otimes X\beta+V(u)\beta,

    where fcxaybX([f]C)w𝔤w0wf^{c}x^{a}y^{b}\otimes X\in(\mathbb{C}[f]\otimes C)_{w^{\prime}}\otimes\mathfrak{g}_{w_{0}-w^{\prime}} and u𝒪ww0𝔤w0wu\in\mathcal{O}_{w^{\prime}-w_{0}}\otimes\mathfrak{g}_{w_{0}-w^{\prime}}. Then (euγ)w=γwV(u)βH1(U0)(e^{u}\ast\gamma)_{w^{\prime}}=\gamma_{w^{\prime}}-V(u)\beta\in H^{1}(U_{0}).

    Now starting with the lowest weight w>w0w^{\prime}>w_{0}, we iteratively act on γ\gamma by elements euAut(S)e^{u}\in Aut(S) so that the terms up to level ww^{\prime} lie in H1(U0)H^{1}(U_{0}). This will terminate after finitely many steps since U01U^{1}_{0} is finite-dimensional. The result is an element of H1(U0)H^{1}(U_{0}).

  2. 2.

    Claim: The inclusion functor PSH1(U0)Aut(S)U01P_{S}\ltimes H^{1}(U_{0})\to Aut(S)\ltimes U^{1}_{0} is fully-faithful. We need to show that given gAut(S)g\in Aut(S) and γH1(U0)\gamma\in H^{1}(U_{0}), if gγH1(U0)g\ast\gamma\in H^{1}(U_{0}), then gPSg\in P_{S}. Recall the Levi decomposition Aut(S)Aut0(S)GSAut(S)\cong Aut_{0}(S)\rtimes G_{S}, and note that GSPSG_{S}\subseteq P_{S}. It therefore suffices to work under the assumption that gAut0(S)g\in Aut_{0}(S). Since such a gg is unipotent, it has the form g=eug=e^{u}, for uw>0𝒪w𝔤wu\in\bigoplus_{w>0}\mathcal{O}_{w}\otimes\mathfrak{g}_{-w}. Expanding in weights, u=ww1uwu=\sum_{w\geq w_{1}}u_{w}, where w1>0w_{1}>0 is the lowest weight. From the above expressions for euγe^{u}\ast\gamma, we see that the lowest weight for which γ\gamma is modified is w0+w1w_{0}+w_{1}. The corresponding term is given by

    (euγ)w0+w1=γw0+w1+V(uw1).(e^{u}\ast\gamma)_{w_{0}+w_{1}}=\gamma_{w_{0}+w_{1}}+V(u_{w_{1}}).

    Since euγH1(U0)e^{u}\ast\gamma\in H^{1}(U_{0}), we must have V(uw1)=0V(u_{w_{1}})=0, implying that uw1H0(U0)u_{w_{1}}\in H^{0}(U_{0}) and euw1PSe^{u_{w_{1}}}\in P_{S}. Let γ~=euw1γH1(U0)\tilde{\gamma}=e^{u_{w_{1}}}\ast\gamma\in H^{1}(U_{0}), so that gγ=(eueuw1)γ~g\ast\gamma=(e^{u}e^{-u_{w_{1}}})\ast\tilde{\gamma}. Using the Baker-Campbell-Hausdorff formula and the fact that w1>0w_{1}>0, we see that eueuw1=eve^{u}e^{-u_{w_{1}}}=e^{v}, where the lowest weight of vv is strictly greater than w1w_{1}. Hence the result follows by induction on w1w_{1}.

  3. 3.

    Let γH1(U0)\gamma\in H^{1}(U_{0}). Claim: The inclusion (H(U0),adγ)(U0,δS+adγ)(H^{\bullet}(U_{0}),ad_{\gamma})\to(U_{0}^{\bullet},\delta_{S}+ad_{\gamma}) is a quasi-isomorphism. This follows by the homological perturbation lemma [13]. Let a:H(U0)U0a:H^{\bullet}(U_{0})\to U^{\bullet}_{0} be the inclusion. By Corollary 5, this is a quasi-isomorphism with respect to the 0 differential on the domain and δS\delta_{S} on the codomain. We have the decomposition U01=H1(U0)𝖨𝗆(δS)U^{1}_{0}=H^{1}(U_{0})\oplus\mathsf{Im}(\delta_{S}). Let CC be a complement to H0(U0)H^{0}(U_{0}), so that U00=H0(U0)CU^{0}_{0}=H^{0}(U_{0})\oplus C. It is possible to choose this complement compatible with the weight decomposition. Using the decomposition we define the projection b:U0H(U0)b:U^{\bullet}_{0}\to H^{\bullet}(U_{0}). The restriction δS|C:C𝖨𝗆(δS)\delta_{S}|_{C}:C\to\mathsf{Im}(\delta_{S}) is an isomorphism, and the inverse defines a map h:U01U00h:U^{1}_{0}\to U^{0}_{0} which has weight w0-w_{0}. These maps satisfy ba=idb\circ a=id, idab=[δS,h]id-a\circ b=[\delta_{S},h], as well as the side conditions ha=0h\circ a=0, bh=0b\circ h=0 and hh=0h\circ h=0.

    Now consider the map adγ:U00U01ad_{\gamma}:U^{0}_{0}\to U^{1}_{0} which will serve as a perturbation. Note that it restricts to a map H0(U0)H1(U0)H^{0}(U_{0})\to H^{1}(U_{0}). Expanding in the weights, γ=w>w0γw\gamma=\sum_{w>w_{0}}\gamma_{w}. Since the weight of hh is w0-w_{0}, it follows that adγhad_{\gamma}\circ h is an endomorphism of U01U^{1}_{0} which raises the weight of an element by at least 11. It follows that adγhad_{\gamma}\circ h is nilpotent, allowing us to apply the perturbation lemma. Using the formulas appearing above Lemma 4, we see that aa remains unperturbed, δS\delta_{S} is deformed to δS+adγ\delta_{S}+ad_{\gamma} and the zero differential is deformed to adγad_{\gamma}.

The moduli stack [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)]

Now we choose an element A=S+N0𝔤A=S+N_{0}\in\mathfrak{g}, which we write using the Jordan decomposition. This determines the derived stack [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)]. The base of the derived manifold is

W(A)={Cβ+Nα0U01U00α0|N(0)GSN0}.W(A)=\{C\beta+N\alpha_{0}\in U^{1}_{0}\oplus U^{0}_{0}\alpha_{0}\ |\ N(0)\in G_{S}\ast N_{0}\}.

The bundle of curved dgla’s is the trivial bundle W(A)×U01α0W(A)\times U^{1}_{0}\alpha_{0} with trivial dgla structure. The curvature section is given by F(Cβ+Nα0)=(V(N)+[C,N])βα0F(C\beta+N\alpha_{0})=(V(N)+[C,N])\beta\wedge\alpha_{0}. Applying Lemma 5 and Proposition 5 we can construct a smaller model for this derived stack.

By Lemma 5, the vector space H1(U0)H^{1}(U_{0}) is a linear representation of PSP_{S}. The Lie algebra Lie(PS)Lie(P_{S}) is likewise a representation, and the subspace dχ1(GSN0)d\chi^{-1}(G_{S}\ast N_{0}) is preserved by this action (recall that χ:PSGS\chi:P_{S}\to G_{S} is the projection to the reductive quotient). Let

Q(A)=H1(U0)×dχ1(GSN0),Q(A)=H^{1}(U_{0})\times d\chi^{-1}(G_{S}\ast N_{0}),

equipped with the action of PSP_{S}. The infinitesimal action of PSP_{S} on H1(U0)H^{1}(U_{0}) defines a PSP_{S}-equivariant map

FS:Q(A)H1(U0),(C,N)[C,N].F_{S}:Q(A)\to H^{1}(U_{0}),\qquad(C,N)\mapsto[C,N].

Viewing this as a section of the bundle Q(A)×H1(U0)Q(A)\times H^{1}(U_{0}) we get a derived manifold 𝒬(A)\mathcal{Q}(A), which represents the derived vanishing locus of FSF_{S}. This defines the derived stack

[𝒬(A)/PS].[\mathcal{Q}(A)/P_{S}].

By Corollary 5 and Lemma 5 this maps into [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)].

Theorem 5.1.

There is a map of derived stacks

i:[𝒬(A)/PS][𝒲(A)/Aut(S)].i:[\mathcal{Q}(A)/P_{S}]\to[\mathcal{W}(A)/Aut(S)].

For points (0,N)Q(A)(0,N)\in Q(A) the derivative di(0,N)di_{(0,N)} is a quasi-isomorphism of tangent complexes. Furthermore, if SS is large enough, then ii is an equivalence.

Proof.

To begin, assume that SS is large enough, so that PSH1(U0)Aut(S)U01P_{S}\ltimes H^{1}(U_{0})\subseteq Aut(S)\ltimes U^{1}_{0} is a Morita equivalence by Proposition 5. Now let Cβ+Nα0MC(𝒲(A))C\beta+N\alpha_{0}\in MC(\mathcal{W}(A)). Claim: If CH1(U0)C\in H^{1}(U_{0}), then (C,N)MC(𝒬(A))(C,N)\in MC(\mathcal{Q}(A)). Indeed, observe that F(Cβ+Nα0)=(δS(N)+adCβ(N))α0F(C\beta+N\alpha_{0})=(\delta_{S}(N)+ad_{C\beta}(N))\wedge\alpha_{0}, and that

(H(U0),adCβ)(U0,δS+adCβ)(H^{\bullet}(U_{0}),ad_{C\beta})\to(U_{0}^{\bullet},\delta_{S}+ad_{C\beta}) (5.1)

is a quasi-isomorphism. It follows that NH0(U0)N\in H^{0}(U_{0}), and therefore that the claim is verified. It is straightforward to deduce from this claim that the induced morphism

PSMC(𝒬(A))Aut(A)MC(𝒲(A))P_{S}\ltimes MC(\mathcal{Q}(A))\to Aut(A)\ltimes MC(\mathcal{W}(A))

is an equivalence of categories. Now given a point (C,N)MC(𝒬(A))(C,N)\in MC(\mathcal{Q}(A)), we need to show that the morphism of tangent complexes di(C,N):𝕋(C,N)[𝒬(A)/PS]𝕋(C,N)[𝒲(A)/Aut(S)]di_{(C,N)}:\mathbb{T}_{(C,N)}[\mathcal{Q}(A)/P_{S}]\to\mathbb{T}_{(C,N)}[\mathcal{W}(A)/Aut(S)] is a quasi-isomorphism. The differentials on these tangent complexes have the form d+adNα0d+ad_{N\alpha_{0}}. By an argument involving the perturbation lemma, it suffices to prove that di(C,N)di_{(C,N)} is a quasi-isomorphism with respect to the differentials dd. But for these differentials, di(C,N)di_{(C,N)} is a direct sum of Equation 5.1 and a shift of a subspace. Therefore, it is a quasi-isomorphism. Note that by Corollary 5, Equation 5.1 is quasi-isomorphism when C=0C=0 even if SS is not large enough. ∎

Remark \theremark.

Connections of the form (0,N)Q(A)(0,N)\in Q(A) are pullbacks by ff of connections on \mathbb{C} with a logarithmic pole at the origin.

The condition in Theorem 5.1 that SS is large enough is necessary. In the following Example we see that the moduli space can have extra components when the condition is not satisfied.

Example \theex.

Let f=x2y5f=x^{2}-y^{5}, let 𝔤=𝔤𝔩3\mathfrak{g}=\mathfrak{gl}_{3}, and let A=SA=S be a diagonal matrix with entries 0,10,1 and 1111. Note that SS is not large enough since 11 is a positive eigenvalue of adSad_{S} which is smaller than w0=3w_{0}=3. The subalgebra 𝔤S\mathfrak{g}_{S} consists of the diagonal matrices and

U00\displaystyle U_{0}^{0} =𝔤Sspan(x2E23,y5E23,xy3E13),\displaystyle=\mathfrak{g}_{S}\oplus\mathrm{span}_{\mathbb{C}}(x^{2}E_{23},y^{5}E_{23},xy^{3}E_{13}),
U01\displaystyle U_{0}^{1} =span(yE21,y2E12,xy4E23,x2y2E13,y7E13),\displaystyle=\mathrm{span}_{\mathbb{C}}(yE_{21},y^{2}E_{12},xy^{4}E_{23},x^{2}y^{2}E_{13},y^{7}E_{13}),

where EijE_{ij} are the elementary matrices. On the other hand, the δS\delta_{S} cohomology is given by

H0(U0)=𝔤Sspan(fE23),H1(U0)=span(yE21,y2E12,y2fE13).H^{0}(U_{0})=\mathfrak{g}_{S}\oplus\mathrm{span}_{\mathbb{C}}(fE_{23}),\ \ H^{1}(U_{0})=\mathrm{span}_{\mathbb{C}}(yE_{21},y^{2}E_{12},y^{2}fE_{13}).

An arbitrary element of W(A)W(A) has the form Cβ+Nα0C\beta+N\alpha_{0}, where

C\displaystyle C =C21yE21+C12y2E12+C23xy4E23+(C13(1)x2y2+C13(2)y7)E13,\displaystyle=C_{21}yE_{21}+C_{12}y^{2}E_{12}+C_{23}xy^{4}E_{23}+(C_{13}^{(1)}x^{2}y^{2}+C_{13}^{(2)}y^{7})E_{13},
N\displaystyle N =N13xy3E13+(N23(1)f+N23(2)(x2+y5))E23.\displaystyle=N_{13}xy^{3}E_{13}+(N_{23}^{(1)}f+N_{23}^{(2)}(x^{2}+y^{5}))E_{23}.

The Maurer-Cartan equation consists of the following coupled system of equations:

20N23(2)\displaystyle 20N_{23}^{(2)} =C21N13\displaystyle=-C_{21}N_{13}
5N13\displaystyle 5N_{13} =C12(N23(2)N23(1))\displaystyle=-C_{12}(N_{23}^{(2)}-N_{23}^{(1)})
6N13\displaystyle 6N_{13} =C12(N23(2)+N23(1)).\displaystyle=-C_{12}(N_{23}^{(2)}+N_{23}^{(1)}).

Adding the last two equations and substituting the result into the first yields (110C21C12)N23(2)=0(110-C_{21}C_{12})N_{23}^{(2)}=0. Assume first that C21C12110C_{21}C_{12}\neq 110. Then the equations simplify to N23(2)=N13=C12N23(1)=0N_{23}^{(2)}=N_{13}=C_{12}N_{23}^{(1)}=0, and NH0(U0)N\in H^{0}(U_{0}). The result is a 55-dimensional variety MaM_{a} with 22 irreducible components. Next, assume that C21C12=110C_{21}C_{12}=110. Then we can solve the equations to obtain N13=2C12N23(1)N_{13}=-2C_{12}N_{23}^{(1)} and N23(2)=11N23(1)N_{23}^{(2)}=11N_{23}^{(1)}. Hence, the result is a smooth irreducible 55-dimensional variety MbM_{b}.

An element of Aut(S)Aut(S) has the form

g=(u0λxy30vax2+by500w),g=\begin{pmatrix}u&0&\lambda xy^{3}\\ 0&v&ax^{2}+by^{5}\\ 0&0&w\end{pmatrix},

and it acts on C=(C21,C12,C23,C13(1),C13(2))C=(C_{21},C_{12},C_{23},C_{13}^{(1)},C_{13}^{(2)}) and N=(N13,N23(1),N23(2))N=(N_{13},N_{23}^{(1)},N_{23}^{(2)}) in the following way:

gC\displaystyle g\ast C =(vuC21,uvC12,vwC23λvuwC2110(a+b)w,uwC13(1)auvwC126λw,uwC13(2)buvwC125λw)\displaystyle=(\frac{v}{u}C_{21},\frac{u}{v}C_{12},\frac{v}{w}C_{23}-\frac{\lambda v}{uw}C_{21}-\frac{10(a+b)}{w},\frac{u}{w}C_{13}^{(1)}-\frac{au}{vw}C_{12}-\frac{6\lambda}{w},\frac{u}{w}C_{13}^{(2)}-\frac{bu}{vw}C_{12}-\frac{5\lambda}{w})
gN\displaystyle g\ast N =(uwN13,vwN23(1),vwN23(2)).\displaystyle=(\frac{u}{w}N_{13},\frac{v}{w}N_{23}^{(1)},\frac{v}{w}N_{23}^{(2)}).

There are two things we can immediately note. First, by looking at the action on NN, we can see that there are orbits in MbM_{b} which do not intersect Q(A)Q(A). Hence, [Mb/Aut(S)][M_{b}/Aut(S)] is an extra component of the moduli space. Second, the product C21C12C_{21}C_{12} is invariant under the action and if we assume that C21C12110C_{21}C_{12}\neq 110, then it is always possible to send CC into H1(U0)H^{1}(U_{0}) (i.e. C23=0C_{23}=0 and C13(1)+C13(2)=0C_{13}^{(1)}+C_{13}^{(2)}=0). The subgroup of Aut(S)Aut(S) which stabilizes this locus is PSP_{S}, and hence [Ma/Aut(S)][M_{a}/Aut(S)] is contained in [MC(𝒬(A))/PS][MC(\mathcal{Q}(A))/P_{S}]. \blacksquare

Relation to the Grothendieck-Springer resolution

Recall that PSCSP_{S}\subseteq C_{S} is a parabolic subgroup with Levi factor GSG_{S}. Denote their Lie algebras 𝔭S\mathfrak{p}_{S}, 𝔠S\mathfrak{c}_{S} and 𝔤S\mathfrak{g}_{S}, respectively. Let \mathbb{P} denote the partial flag variety, defined as the set of parabolic subalgebras of 𝔠S\mathfrak{c}_{S} which are conjugate to 𝔭S\mathfrak{p}_{S}. It is isomorphic to CS/PSC_{S}/P_{S}. Consider the incidence variety

𝔠~={(x,𝔭)𝔠S×|x𝔭},\tilde{\mathfrak{c}}=\{(x,\mathfrak{p})\in\mathfrak{c}_{S}\times\mathbb{P}\ |\ x\in\mathfrak{p}\},

which is isomorphic to (CS×𝔭S)/PS(C_{S}\times\mathfrak{p}_{S})/P_{S} via the map which sends (g,x)CS×𝔭S(g,x)\in C_{S}\times\mathfrak{p}_{S} to the pair (Adg(x),Adg(𝔭S))𝔠~(Ad_{g}(x),Ad_{g}(\mathfrak{p}_{S}))\in\tilde{\mathfrak{c}}. When PSP_{S} is a Borel subgroup, 𝔠~\tilde{\mathfrak{c}} is the Grothendieck-Springer resolution (see e.g. [11] for details). The element N0𝔤SN_{0}\in\mathfrak{g}_{S} defines an adjoint orbit 𝒪\mathcal{O} in the Levi quotient of every parabolic 𝔭\mathfrak{p}\in\mathbb{P} [5, Lemma 1]. This lets us define the following subspace of 𝔠~\tilde{\mathfrak{c}}:

𝔠~N0={(x,𝔭)𝔠~|dχ(x)𝒪},\tilde{\mathfrak{c}}_{N_{0}}=\{(x,\mathfrak{p})\in\tilde{\mathfrak{c}}\ |\ d\chi(x)\in\mathcal{O}\},

where dχd\chi denotes the projection to the Levi quotient. The space Q(A)Q(A) is a PSP_{S}-equivariant vector bundle over dχ1(GSN0)d\chi^{-1}(G_{S}\ast N_{0}). Hence π:EA=(CS×Q(A))/PS𝔠~N0\pi:E_{A}=(C_{S}\times Q(A))/P_{S}\to\tilde{\mathfrak{c}}_{N_{0}} is a CSC_{S}-equivariant vector bundle and the map FSF_{S} gives rise to an equivariant section σS\sigma_{S} of π(EA)EA\pi^{*}(E_{A})\to E_{A}. In this way, we obtain a derived stack [A/CS][\mathcal{E}_{A}/C_{S}] which is equivalent to [𝒬(A)/PS][\mathcal{Q}(A)/P_{S}].

Example \theex.

Let G=GLnG=GL_{n} and let A=S=pqDA=S=pqD, where DD is a diagonal matrix with distinct integer eigenvalues. In this case, SS is large enough, CS=GC_{S}=G and PS=BP_{S}=B is a Borel subgroup. Hence \mathbb{P} is the flag variety FlnFl_{n}. Since N0=0N_{0}=0, 𝔤~0\tilde{\mathfrak{g}}_{0} is the Springer resolution, which is isomorphic to TFlnT^{*}Fl_{n}. We use Corollary 5 to compute the cohomology space H1(U0)H^{1}(U_{0}). The weights ww showing up in the decomposition have the form w=(p1)q+(q1)p+cqpw=(p-1)q+(q-1)p+cqp. But ([f]C)w=0(\mathbb{C}[f]\otimes C)_{w}=0 in this case, so that H1(U0)=0H^{1}(U_{0})=0. Hence, the moduli space [𝒲(A)/Aut(S)][\mathcal{W}(A)/Aut(S)] is equivalent to the quotient stack

[TFln/GLn].[T^{*}Fl_{n}/GL_{n}].

Note that the connections corresponding to the points of TFlnT^{*}Fl_{n} are all pulled-back from logarithmic connections on \mathbb{C}, where a similar classification is given by [5, Theorem A]. \blacksquare

Example \theex.

Let G=GLnG=GL_{n} and A=S=qprDA=S=\frac{qp}{r}D, where DD is a diagonal integer matrix with distinct eigenvalues and rr is a positive integer. In this case, the eigenvalues of e2πiqpSe^{\frac{-2\pi i}{qp}S} are rthr^{th} roots of unity, implying that CS=i=0r1GLmiC_{S}=\prod_{i=0}^{r-1}GL_{m_{i}}, with the factors indexed by the roots of unity. Since the eigenvalues of SS are assumed to be distinct, PS=i=0r1BmiP_{S}=\prod_{i=0}^{r-1}B_{m_{i}} is a product of Borels. Therefore, i=0r1Flmi\mathbb{P}\cong\prod_{i=0}^{r-1}Fl_{m_{i}} is a product of flag varieties and 𝔠~0i=0r1TFlmi\tilde{\mathfrak{c}}_{0}\cong\prod_{i=0}^{r-1}T^{*}Fl_{m_{i}} is the product of their cotangent bundles. Writing the eigenvalues of DD as rk+urk+u, with 0u<r0\leq u<r, the eigenvalues of adSad_{S} have the form qp(k1k2)+qpr(u1u2)qp(k_{1}-k_{2})+\frac{qp}{r}(u_{1}-u_{2}). We can ensure that SS is large enough by restricting the possible values of kik_{i}. For example, this is guaranteed if k1k20,±1k_{1}-k_{2}\neq 0,\pm 1 for all pairs of eigenvalues. By Corollary 5, the weights ww showing up in the decomposition of H1(U0)H^{1}(U_{0}) have the form

w=qp(1+k2k1)+qpr(u2u1)pq.w=qp(1+k_{2}-k_{1})+\frac{qp}{r}(u_{2}-u_{1})-p-q.

As in Example 5, we must have u2u1u_{2}\neq u_{1} in order to get a non-zero contribution.

Now we specialise to GL4GL_{4}, with r=pqr=pq and S=DS=D a diagonal matrix with entries

pqk1,pqk2,pqk3+p+q,pqk4+p+q,pqk_{1},\ \ pqk_{2},\ \ pqk_{3}+p+q,\ \ pqk_{4}+p+q,

such that k1k2k3k4k_{1}\ll k_{2}\ll k_{3}\ll k_{4}. Then CS=GL2×GL2C_{S}=GL_{2}\times GL_{2} and 1×1\mathbb{P}\cong\mathbb{P}^{1}\times\mathbb{P}^{1}. Let T1T_{1} and T2T_{2} be the tautological rank 22 vector bundles over 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} (corresponding respectively to the first and second factors). They are both trivial, but are equipped with tautological line subbundle bundles LiL_{i} with respective degrees (1,0)(-1,0) and (0,1)(0,-1). The cotangent bundle 𝔠~0\tilde{\mathfrak{c}}_{0} can be identified with Hom(T1/L1,L1)Hom(T2/L2,L2)Hom(T_{1}/L_{1},L_{1})\oplus Hom(T_{2}/L_{2},L_{2}), or alternatively, the nilpotent filtration-preserving endomorphisms of T1T_{1} and T2T_{2}.

The weights in the decomposition of H1(U0)H^{1}(U_{0}) have the form qp(1+kjki)qp(1+k_{j}-k_{i}), for j=3,4j=3,4 and i=1,2i=1,2. Hence, H1(U0)H^{1}(U_{0}) can be identified with the subspace of 𝔤𝔩4\mathfrak{gl}_{4} consisting of the upper right 2×22\times 2 block. Therefore,

EA=Hom(T1/L1,L1)Hom(T2/L2,L2)Hom(T2,T1),E_{A}=Hom(T_{1}/L_{1},L_{1})\oplus Hom(T_{2}/L_{2},L_{2})\oplus Hom(T_{2},T_{1}),

with section σS(a,b,c)=cbac\sigma_{S}(a,b,c)=c\circ b-a\circ c. \blacksquare

Tangent Lie bialgebra and shifted Poisson geometry

Let A=S𝔤A=S\in\mathfrak{g} be a semisimple element and consider the moduli space [𝒲(S)/Aut(S)][\mathcal{W}(S)/Aut(S)]. There is a distinguished point 0MC(𝒲(S))0\in MC(\mathcal{W}(S)) corresponding to the connection 11-form α0S\alpha_{0}S. In this section we focus our attention on a formal neighbourhood of this point in the moduli space and sketch the construction of a 2-2-shifted Poisson structure on this neighbourhood. According to the fundamental principal of derived deformation theory (see e.g. [17, 20, 27, 18]) this formal neighbourhood is encoded by the shifted tangent complex 𝕋0[𝒲(S)/Aut(S)][1]\mathbb{T}_{0}[\mathcal{W}(S)/Aut(S)][-1], equipped with its structure as a differential graded Lie algebra. By Theorem 5.1 this dgla is quasi-isomorphic to

𝕋0[𝒬(A)/PS][1]=H0(U0)H1(U0)H0(U0)+H1(U0),\mathbb{T}_{0}[\mathcal{Q}(A)/P_{S}][-1]=H^{0}(U_{0})\to H^{1}(U_{0})\oplus H^{0}(U_{0})_{+}\to H^{1}(U_{0}),

where the differential is 0 and H0(U0)+=c>0fc𝔤cpqH^{0}(U_{0})_{+}=\bigoplus_{c>0}f^{c}\mathfrak{g}_{-cpq}. By [26, Proposition 1.5], a 3-3-shifted Lie bialgebra structure on 𝕋0[𝒬(A)/PS][1]\mathbb{T}_{0}[\mathcal{Q}(A)/P_{S}][-1] gives rise to a 2-2-shifted Poisson structure on the formal neighbourhood. Hence, our strategy is to construct a Lie bialgebra structure on the tangent complex by embedding it into H(U0)H(U0)[1]H^{\bullet}(U_{0})\ltimes H^{\bullet}(U_{0})[-1], and then to realize this larger Lie algebra as a Lagrangian inside a 3-3-shifted Manin triple.

We construct the Manin triple in stages, starting with the following input data:

  • Choose an invariant inner product kk on the reductive Lie algebra 𝔤\mathfrak{g}. This induces a perfect pairing between the eigenspaces 𝔤λ\mathfrak{g}_{\lambda} and 𝔤λ\mathfrak{g}_{-\lambda}.

  • The \mathbb{C}-vector space C=[x,y]/(xp1,yq1)C=\mathbb{C}[x,y]/(x^{p-1},y^{q-1}) is canonically isomorphic to the degree 11 cohomology of f1(1)f^{-1}(1), which is a curve of genus g=12(p1)(q1)g=\frac{1}{2}(p-1)(q-1) with a single puncture. The isomorphism is given as follows:

    CH1(f1(1)),xayb(xaybβ)|f1(1).C\to H^{1}(f^{-1}(1)),\qquad x^{a}y^{b}\mapsto(x^{a}y^{b}\beta)|_{f^{-1}(1)}.

    By pulling back the intersection pairing on the curve, we obtain a symplectic form II on CC. Up to a scaling constant, it is given by the following formula

    I(xayb,xayb)=δ(a+a+2p)δ(b+b+2q)aq+bpw0,I(x^{a}y^{b},x^{a^{\prime}}y^{b^{\prime}})=\frac{\delta(a+a^{\prime}+2-p)\delta(b+b^{\prime}+2-q)}{aq+bp-w_{0}},

    where δ\delta is the delta function which evaluates to 11 at 0 and 0 otherwise. We can extend this to a [f]\mathbb{C}[f]-linear pairing on [f]C\mathbb{C}[f]\otimes_{\mathbb{C}}C.

Now consider the Lie algebra 𝔠=cfc𝔤cpq\mathfrak{c}=\bigoplus_{c\in\mathbb{Z}}f^{c}\mathfrak{g}_{-cpq}. This has two distinguished subalgebras: first 𝔟=H0(U0)\mathfrak{b}=H^{0}(U_{0}), where c0c\geq 0, and second 𝔟\mathfrak{b}^{-}, where c0c\leq 0. Note that 𝔠\mathfrak{c} can be viewed as a subalgebra of 𝔤\mathfrak{g} and so inherits the pairing kk. This defines a perfect pairing between 𝔟\mathfrak{b} and 𝔟\mathfrak{b}^{-}.

Next, define the following vector space

𝔎=c0ap20bq2fcxayb𝔤w0cpqaqbp.\mathfrak{K}=\bigoplus_{c\in\mathbb{Z}}\bigoplus_{\begin{subarray}{c}0\leq a\leq p-2\\ 0\leq b\leq q-2\end{subarray}}f^{c}x^{a}y^{b}\mathfrak{g}_{w_{0}-cpq-aq-bp}.

This decomposes as 𝔎=𝔫+𝔥𝔫\mathfrak{K}=\mathfrak{n}_{+}\oplus\mathfrak{h}\oplus\mathfrak{n}_{-} according to whether cc is positive, zero, or negative, respectively. Note that H1(U0)=𝔫+𝔥H^{1}(U_{0})=\mathfrak{n}_{+}\oplus\mathfrak{h}. Viewing both 𝔠\mathfrak{c} and 𝔎\mathfrak{K} as subspaces of the Lie algebra [x,y]𝔤\mathbb{C}[x,y]\otimes\mathfrak{g}, we can show that 𝔎\mathfrak{K} is a representation of 𝔠\mathfrak{c}. By combining the symplectic form II on CC with the Lie bracket on 𝔤\mathfrak{g}, we define a symmetric 𝔠\mathfrak{c}-equivariant map ω:S2(𝔎)𝔠\omega:S^{2}(\mathfrak{K})\to\mathfrak{c} as follows

ω(fc1xa1yb1X1,fc2xa2yb2X2)=fc1+c2I(xa1yb1,xa2yb2)[X1,X2].\omega(f^{c_{1}}x^{a_{1}}y^{b_{1}}X_{1},f^{c_{2}}x^{a_{2}}y^{b_{2}}X_{2})=f^{c_{1}+c_{2}}I(x^{a_{1}}y^{b_{1}},x^{a_{2}}y^{b_{2}})[X_{1},X_{2}].

Post-composing this with the natural projection to 𝔟\mathfrak{b}^{-} defines a map μ:S2(𝔎)𝔟\mu:S^{2}(\mathfrak{K})\to\mathfrak{b}^{-}. This is 𝔟\mathfrak{b}-equivariant, where 𝔟\mathfrak{b}^{-} is a 𝔟\mathfrak{b}-representation by using the inner product to identify it with 𝔟\mathfrak{b}^{*}. With respect to the 𝔟\mathfrak{b} action on 𝔎\mathfrak{K}, both 𝔫+\mathfrak{n}_{+} and 𝔫+𝔥\mathfrak{n}_{+}\oplus\mathfrak{h} are sub-representations. Hence the quotient 𝔥=𝔫+𝔥𝔫+\mathfrak{h}=\frac{\mathfrak{n}_{+}\oplus\mathfrak{h}}{\mathfrak{n}_{+}} is naturally a 𝔟\mathfrak{b}-representation, and μ\mu descends to define a 𝔟\mathfrak{b}-equivariant map ν:S2(𝔥)𝔟\nu:S^{2}(\mathfrak{h})\to\mathfrak{b}^{-}. We now define a graded Lie algebra

L=𝔟(𝔎𝔥)[1]𝔟[2].L=\mathfrak{b}\oplus(\mathfrak{K}\oplus\mathfrak{h})[-1]\oplus\mathfrak{b}^{-}[-2].

The bracket is constructed from the bracket on 𝔟\mathfrak{b}, the 𝔟\mathfrak{b}-action on the other summands, the symmetric pairing μ\mu on 𝔎\mathfrak{K} and the symmetric pairing ν-\nu on 𝔥\mathfrak{h}. We set the bracket between 𝔎\mathfrak{K} and 𝔥\mathfrak{h} to be zero.

Lemma \thelemma.

The vector space LL with the bracket described above defines a graded Lie algebra. Let 𝔭+=𝔟(𝔫+𝔥)[1]\mathfrak{p}_{+}=\mathfrak{b}\oplus(\mathfrak{n}_{+}\oplus\mathfrak{h})[-1] and let 𝔭=(𝔫𝔥)[1]𝔟[2]\mathfrak{p}_{-}=(\mathfrak{n}_{-}\oplus\mathfrak{h})[-1]\oplus\mathfrak{b}^{-}[-2]. There are morphisms

𝔭+L,\displaystyle\mathfrak{p}_{+}\to L, (b,n,h)(b,n+h,h,0)\displaystyle\qquad(b,n,h)\mapsto(b,n+h,h,0)
𝔭L,\displaystyle\mathfrak{p}_{-}\to L, (n,h,u)(0,n+h,h,u).\displaystyle\qquad(n,h,u)\mapsto(0,n+h,-h,u).

These embed 𝔭±\mathfrak{p}_{\pm} as complementary subalgebras of LL. Furthermore, 𝔭+\mathfrak{p}_{+} is isomorphic to H(U0)H^{\bullet}(U_{0}).

Next we construct an inner product on LL. By combining the inner product kk on 𝔤\mathfrak{g} with the symplectic pairing II on CC, we define a skew symmetric pairing Ω:2𝔎\Omega:\wedge^{2}\mathfrak{K}\to\mathbb{C} as follows

Ω(fc1xa1yb1X1,fc2xa2yb2X2)=fc1+c2I(xa1yb1,xa2yb2)k(X1,X2).\Omega(f^{c_{1}}x^{a_{1}}y^{b_{1}}X_{1},f^{c_{2}}x^{a_{2}}y^{b_{2}}X_{2})=f^{c_{1}+c_{2}}I(x^{a_{1}}y^{b_{1}},x^{a_{2}}y^{b_{2}})k(X_{1},X_{2}).

This pairing is non-degenerate, restricts to a non-degenerate pairing on 𝔥\mathfrak{h} and defines a perfect pairing between 𝔫±\mathfrak{n}_{\pm}. We can now define a non-degenerate graded symmetric bilinear pairing

B:S2(L)[2].B:S^{2}(L)\to\mathbb{C}[-2].

More precisely, given y=(b1,k1,h1,u1),z=(b2,k2,h2,u2)Ly=(b_{1},k_{1},h_{1},u_{1}),z=(b_{2},k_{2},h_{2},u_{2})\in L, we set

B(y,z)=k(b1,u2)+k(u1,b2)+Ω(k1,k2)Ω(h1,h2).B(y,z)=k(b_{1},u_{2})+k(u_{1},b_{2})+\Omega(k_{1},k_{2})-\Omega(h_{1},h_{2}).
Lemma \thelemma.

The pairing BB is a non-degenerate invariant inner product on LL. Furthermore, 𝔭±\mathfrak{p}_{\pm} are complementary Lagrangian subalgebras. In other words, (L,𝔭+,𝔭)(L,\mathfrak{p}_{+},\mathfrak{p}_{-}) is a 2-2-shifted Manin triple.

Now let [ϵ]\mathbb{C}[\epsilon] be the cdga generated by a degree +1+1 variable ϵ\epsilon and let tr:[ϵ][1]tr:\mathbb{C}[\epsilon]\to\mathbb{C}[-1] be the trace map defined by sending the element a+bϵa+b\epsilon to bb. Tensoring with [ϵ]\mathbb{C}[\epsilon] defines a new triple of graded Lie algebras ([ϵ]L,[ϵ]𝔭+,[ϵ]𝔭)(\mathbb{C}[\epsilon]\otimes L,\mathbb{C}[\epsilon]\otimes\mathfrak{p}_{+},\mathbb{C}[\epsilon]\otimes\mathfrak{p}_{-}) and the bilinear form BB extends in the following way

B:S2([ϵ]L)[3],(fy,gz)(1)|y||g|tr(fg)B(y,z).B:S^{2}(\mathbb{C}[\epsilon]\otimes L)\to\mathbb{C}[-3],\qquad(fy,gz)\mapsto(-1)^{|y||g|}tr(fg)B(y,z).

In this way, we obtain a 3-3-shifted Manin triple. Note that [ϵ]𝔭+H(U0)H(U0)[1]\mathbb{C}[\epsilon]\otimes\mathfrak{p}_{+}\cong H^{\bullet}(U_{0})\ltimes H^{\bullet}(U_{0})[-1]. The tangent complex 𝕋0[𝒬(A)/PS][1]\mathbb{T}_{0}[\mathcal{Q}(A)/P_{S}][-1] is isomorphic to the subalgebra 𝔟(𝔫+𝔥)𝔟>0ϵ(𝔫+𝔥)ϵ\mathfrak{b}\oplus(\mathfrak{n}_{+}\oplus\mathfrak{h})\oplus\mathfrak{b}_{>0}\epsilon\oplus(\mathfrak{n}_{+}\oplus\mathfrak{h})\epsilon, where 𝔟>0\mathfrak{b}_{>0} consists of the elements fcXf^{c}X with c>0c>0. Let MM be the direct sum of this subalgebra with [ϵ]𝔭\mathbb{C}[\epsilon]\otimes\mathfrak{p}_{-}. Then MM is a coisotropic subalgebra of LL and M=𝔤0𝔟MM^{\perp}=\mathfrak{g}_{0}\subset\mathfrak{b}^{-}\subset M is an isotropic ideal. It follows that (M/M1,𝕋0[𝒬(A)/PS][1],[ϵ]𝔭/M)(M/M^{-1},\mathbb{T}_{0}[\mathcal{Q}(A)/P_{S}][-1],\mathbb{C}[\epsilon]\otimes\mathfrak{p}_{-}/M^{\perp}) is a 3-3-shifted Manin triple. Therefore, by [26, Lemma 1.3], the tangent complex obtains a 3-3-shifted Lie bialgebra structure.

Theorem 5.2.

The tangent complex 𝕋0[𝒬(A)/PS][1]\mathbb{T}_{0}[\mathcal{Q}(A)/P_{S}][-1] admits the structure of a 3-3-shifted Lie bialgebra. Therefore, the formal neighbourhood of 0[𝒲(S)/Aut(S)]0\in[\mathcal{W}(S)/Aut(S)] admits a 2-2-shifted Poisson structure.

References

  • [1] M. F. Atiyah and R. Bott. The Yang-Mills equations over Riemann surfaces. Philos. Trans. Roy. Soc. London Ser. A, 308(1505):523–615, 1983.
  • [2] Kai Behrend, Ionut Ciocan-Fontanine, Junho Hwang, and Michael Rose. The derived moduli space of stable sheaves. Algebra Number Theory, 8(4):781–812, 2014.
  • [3] Francis Bischoff. Lie groupoids and logarithmic connections. arXiv preprint arXiv:2010.03685, 2020.
  • [4] Francis Bischoff. Normal forms and moduli stacks for logarithmic flat connections. arXiv preprint arXiv:2209.00631, 2022.
  • [5] P. P. Boalch. Riemann–Hilbert for tame complex parahoric connections. Transformation groups, 16(1):27–50, 2011.
  • [6] Philip Boalch. Symplectic manifolds and isomonodromic deformations. Adv. Math., 163(2):137–205, 2001.
  • [7] Philip Boalch. Quasi-Hamiltonian geometry of meromorphic connections. Duke Math. J., 139(2):369–405, 2007.
  • [8] Philip P. Boalch. GG-bundles, isomonodromy, and quantum Weyl groups. Int. Math. Res. Not., (22):1129–1166, 2002.
  • [9] Damien Calaque. Lagrangian structures on mapping stacks and semi-classical TFTs. In Stacks and categories in geometry, topology, and algebra, volume 643 of Contemp. Math., pages 1–23. Amer. Math. Soc., Providence, RI, 2015.
  • [10] Damien Calaque, Tony Pantev, Bertrand Toën, Michel Vaquié, and Gabriele Vezzosi. Shifted Poisson structures and deformation quantization. J. Topol., 10(2):483–584, 2017.
  • [11] Neil Chriss and Victor Ginzburg. Representation theory and complex geometry. Modern Birkhäuser Classics. Birkhäuser Boston, Ltd., Boston, MA, 2010. Reprint of the 1997 edition.
  • [12] Ionuţ Ciocan-Fontanine and Mikhail Kapranov. Derived Quot schemes. Ann. Sci. École Norm. Sup. (4), 34(3):403–440, 2001.
  • [13] Marius Crainic. On the perturbation lemma, and deformations. arXiv preprint math/0403266, 2004.
  • [14] V. V. Fock and A. A. Rosly. Poisson structure on moduli of flat connections on Riemann surfaces and the rr-matrix. In Moscow Seminar in Mathematical Physics, volume 191 of Amer. Math. Soc. Transl. Ser. 2, pages 67–86. Amer. Math. Soc., Providence, RI, 1999.
  • [15] William M. Goldman. The symplectic nature of fundamental groups of surfaces. Adv. in Math., 54(2):200–225, 1984.
  • [16] M. Kapranov. Injective resolutions of BGBG and derived moduli spaces of local systems. J. Pure Appl. Algebra, 155(2-3):167–179, 2001.
  • [17] Maxim Kontsevich and Yan Soibelman. Deformation theory. Livre en préparation, 2002.
  • [18] Jacob Lurie. Moduli problems for ring spectra. In Proceedings of the International Congress of Mathematicians. Volume II, pages 1099–1125. Hindustan Book Agency, New Delhi, 2010.
  • [19] Jacob Lurie. Spectral algebraic geometry. preprint, 2018.
  • [20] Marco Manetti. Lie methods in deformation theory. Springer Monographs in Mathematics. Springer, Singapore, [2022] ©2022.
  • [21] Valerio Melani and Pavel Safronov. Derived coisotropic structures I: affine case. Selecta Math. (N.S.), 24(4):3061–3118, 2018.
  • [22] Valerio Melani and Pavel Safronov. Derived coisotropic structures II: stacks and quantization. Selecta Math. (N.S.), 24(4):3119–3173, 2018.
  • [23] Tony Pantev and Bertrand Toën. Poisson geometry of the moduli of local systems on smooth varieties. Publ. Res. Inst. Math. Sci., 57([3-4]):959–991, 2021.
  • [24] Tony Pantev and Bertrand Toën. Moduli of flat connections on smooth varieties. Algebr. Geom., 9(3):266–310, 2022.
  • [25] Tony Pantev, Bertrand Toën, Michel Vaquié, and Gabriele Vezzosi. Shifted symplectic structures. Publ. Math. Inst. Hautes Études Sci., 117:271–328, 2013.
  • [26] Slava Pimenov. Shifted poisson and batalin-vilkovisky structures on the derived variety of complexes. arXiv preprint arXiv:1511.00946, 2015.
  • [27] J. P. Pridham. Unifying derived deformation theories. Adv. Math., 224(3):772–826, 2010.
  • [28] Jonathan P. Pridham. Constructing derived moduli stacks. Geom. Topol., 17(3):1417–1495, 2013.
  • [29] Pavel Safronov. Quasi-Hamiltonian reduction via classical Chern-Simons theory. Adv. Math., 287:733–773, 2016.
  • [30] Kyoji Saito. Theory of logarithmic differential forms and logarithmic vector fields. J. Fac. Sci. Univ. Tokyo Sect. IA Math, 27(2):265–291, 1980.
  • [31] Bertrand Toën and Gabriele Vezzosi. Homotopical algebraic geometry. II. Geometric stacks and applications. Mem. Amer. Math. Soc., 193(902):x+224, 2008.