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

Filtered Stokes G-local Systems in Nonabelian Hodge Theory on Curves

Pengfei Huang and Hao Sun
Abstract.

In the wild nonabelian Hodge correspondence on curves, filtered Stokes GG-local systems are regarded as the objects on the Betti side. In this paper, we demonstrate a construction of the moduli space of them, called the Betti moduli space, and it reduces to the wild character variety when the Betti weights are trivial. We study some particular examples including Eguchi–-Hanson space and the Airy equation together with the corresponding moduli spaces. Furthermore, we provide a proof of the correspondence among irregular singular GG-connections, Stokes GG-local systems, and Stokes GG-representations. This correspondence can be viewed as the GG-version of irregular Riemann–Hilbert correspondence on curves.

11footnotetext: Key words: Stokes local system, meromorphic connection, Betti moduli space, nonabelian Hodge correspondence22footnotetext: MSC2020: 14D20, 34M40

1. Introduction

1.1. Background

The study of the nonabelian Hodge correspondence on noncompact curves begins with Simpson [Sim90], where he introduced filtered regular Higgs bundles, filtered regular DXD_{X}-modules, and filtered local systems along with corresponding stability conditions to establish a one-to-one correspondence among them. This correspondence is famously known as the tame nonabelian Hodge correspondence, where tameness characterizes a polynomial growth condition of flat sections or the regularity of meromorphic connections. Under this framework, filtered local systems are local systems on a noncompact curve with additional structures called weighted filtrations (also called parabolic structures). They correspond to fundamental group representations satisfying certain compatibility conditions determined by weights (these conditions are trivial in some sense). The existence of weights makes the stability of filtered local systems not equivalent to the irreducibility of the corresponding fundamental group representations. In a similar way, Biquard–Boalch [BB04] generalized Simpson’s correspondence to a broader framework beyond tameness, called wildness, where meromorphic connection can be irregular. They established a one-to-one correspondence between (poly)stable filtered irregular Higgs bundles and (poly)stable filtered irregular DXD_{X}-modules under a “very good” condition. Although this work is known as the (unramified) wild nonabelian Hodge correspondence, it does not touch the objects from the Betti side.

To establish a comprehensive wild nonabelian Hodge correspondence, a crucial step involves figuring out the correct objects on the Betti side, which not only correspond to filtered irregular DXD_{X}-modules but also preserves stability conditions. Classically, the Riemann–Hilbert correspondence connects DXD_{X}-modules with local systems. The same idea holds in the wild case and it is well-known that there is a one-to-one correspondence between connections with irregular singularities (or called irregular DXD_{X}-modules) and Stokes local systems [Sib77, Mal78, Mal83, LR94], which is also known as the irregular Riemann–Hilbert correspondence. Moreover, this correspondence was studied systematically in a great generality by Boalch for reductive groups as the structure group [Boa14, Boa18, Boa21]. Inspired by previous works, the authors introduced (Betti) weights to Stokes local systems, which are called filtered Stokes local systems and regarded as the objects on the Betti side, and define its stability condition. Then, an (unramified) wild nonabelian Hodge correspondence at the level of categories was established [HS23a]. Furthermore, such a correspondence also holds for complex reductive groups as the structure groups. For the case of trivial Betti weights, filtered Stokes local systems from the Betti side reduce to Stokes local systems. In this case, Boalch constructed the moduli space of Stokes local systems by identifying Stokes local systems with representations of the fundamental groupoid of irregular curves, which is now known as the wild character variety. However, as the tame case Simpson ever observed, the stability of filtered Stokes GG-local systems no longer aligns with the irreducibility of the corresponding representations, which is different from the case of trivial Betti weights investigated by Boalch–Yamakawa [BY23]. Consequently, this leads to the fact that a construction of the moduli spaces of filtered Stokes GG-local systems becomes challenging.

In this paper, we provide a construction of this moduli space in detail, for both unramified and ramified cases and study some specific examples. Moreover, the (unramified) wild nonabelian Hodge correspondence considered in [HS23a] holds at the level of moduli spaces.

1.2. Main result

We introduce some notations first. Let GG be a connected complex reductive group with a given maximal torus TT. Let XX be a connected smooth projective algebraic curve over \mathbb{C} with a collection 𝑫\boldsymbol{D} of finite points. Denote by X𝑫:=X\𝑫X_{\boldsymbol{D}}:=X\backslash\boldsymbol{D} the punctured curve. We also fix a collection of irregular types 𝑸={Qx,x𝑫}\boldsymbol{Q}=\{Q_{x},x\in\boldsymbol{D}\} and a collection of weights 𝜽={θx,x𝑫}\boldsymbol{\theta}=\{\theta_{x},x\in\boldsymbol{D}\} labelled by points in 𝑫\boldsymbol{D}, where a weight is regarded as a rational cocharacter Hom(𝔾m,T){\rm Hom}(\mathbb{G}_{m},T)\otimes_{\mathbb{Z}}\mathbb{Q}.

To construct the moduli space of filtered Stokes GG-local systems, we first relate Stokes GG-local systems to Stokes GG-representations, and study the irregular Riemann–Hilbert correspondence for reductive groups GG on X𝑫X_{\boldsymbol{D}} in both unramified and ramified cases in §2 (Theorem 2.14 and 2.17). The proof is similar to [Boa14, Appendix], where the author studied the unramified case. Moreover, Hohl and Jakob recently proved the correspondence (Theorem 2.14) in a different way via Tannakian categories and we refer the reader to [HJ24, §3] for more details.

In §3, we review the stability conditions on filtered Stokes GG-local systems (Definition 3.4), which is the stability condition considered in the wild nonabelian Hodge correspondence [HS23a], and construct the moduli space of filtered Stokes GG-local systems. We would like to point out that King’s result [Kin94] inspires us to relate the stability condition of filtered Stokes GG-local systems to the stability condition in the sense of GIT (Proposition 3.9), which help us to construct the moduli sapce.

Theorem 1.1 (Theorem 3.10).

The moduli space B(X𝐃,G,𝐐,𝛉)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta}) of degree zero 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}} exists as a quasi-projective variety.

In §4, we study some examples of filtered Stokes GG-local systems together with their moduli spaces. Here is a summary of the results.

  1. (1)

    When the weights 𝜽\boldsymbol{\theta} are trivial, the moduli space B(X𝑫,G,𝑸,𝜽)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta}) is exactly the wild character variety, which has been studied in [Boa14, BY15, BY23, DDP18, HMW19].

  2. (2)

    When the irregular types 𝑸\boldsymbol{Q} are trivial, the moduli space B(X𝑫,G,𝑸,𝜽)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta}) is the moduli space of filtered GG-local systems [HS23b], which is the Betti moduli space in the tame nonabelian Hodge correspondence.

  3. (3)

    We equip Eguchi–Hanson space with distinct weights and find a filtered Stokes local system that is stable but not irreducible. This example shows that the Betti moduli spaces in the wild nonabelian Hodge correspondence may not be the wild character varieties.

  4. (4)

    We calculate the corresponding Stokes GG-representations of the classical Airy equation and show that the moduli space, where it lies, is a single point. Therefore, the corresponding irregular singular connection of the Airy equation is both rigid and physically rigid. This result is also obtained in [HJ24, Theorem 1.2.1] recently. Moreover, we prove that in the case of G=SL2()G={\rm SL}_{2}(\mathbb{C}) and a ramified irregular type QQ, the moduli space B(X𝑫,SL2(),Q,θ)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q,\theta) is always isomorphic to the wild character variety and does not depend on the weight (Proposition 4.4). This result can be generalized to G=SLn()G={\rm SL}_{n}(\mathbb{C}) in a certain extent (Remark 4.5).

In §5, we construct the Betti moduli space (Corollary 5.1) in the (unramified) wild nonabelian Hodge correspondence for principal bundles on curves given by the authors in [HS23a, Theorem in §1], and then we show the correspondence holds at the level of moduli spaces (Theorem 5.2). Moreover, the corresponding moduli space is shown admitting a hyperKälher structure (Theorem 5.3).

Acknowledgments. The authors would like to thank Konstantin Jakob, Yichen Qin and Xiaomeng Xu for helpful discussions. Pengfei Huang acknowledges funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No 101018839) and Deutsche Forschungsgemeinschaft (DFG, Projektnummer 547382045). Hao Sun is partially supported by National Key R&\&D Program of China (No. 2022YFA1006600).

2. Irregular Riemann–Hilbert Correspondence

Let XX be a connected smooth projective algebraic curve over \mathbb{C} (i.e. a connected compact Riemann surface) with a collection of finite points 𝑫\boldsymbol{D}. Denote by X𝑫:=X\𝑫X_{\boldsymbol{D}}:=X\backslash\boldsymbol{D} the punctured curve. In this section, we study the irregular Riemann–Hilbert correspondence for connected complex reductive groups GG on X𝑫X_{\boldsymbol{D}} in both unramified and ramified cases. Fixing a collection of irregular types 𝑸={Qx,x𝑫}\boldsymbol{Q}=\{Q_{x},x\in\boldsymbol{D}\}, we prove the equivalence of the following three categories on X𝑫X_{\boldsymbol{D}}

  • the category of GG-connections with irregular type 𝑸\boldsymbol{Q};

  • the category of Stokes GG-local systems with irregular type 𝑸\boldsymbol{Q};

  • the category of Stokes GG-representations with irregular type 𝑸\boldsymbol{Q}.

In §2.1, we first study the correspondence on a punctured disc 𝔻\mathbb{D}^{*}, and then in §2.2, we prove the equivalence of categories on X𝑫X_{\boldsymbol{D}}. In §2.3, we give an equivalent description of the space of Stokes GG-representations, which will be used to construct the moduli space of filtered Stokes GG-local systems in §3.

2.1. Local Correspondence

We fix some notations first

R={z},R^=[[z]],\displaystyle R=\mathbb{C}\{z\},\quad\quad\widehat{R}=\mathbb{C}[\![z]\!],
K=({z}),K^=((z)).\displaystyle K=\mathbb{C}(\!\{z\}\!),\quad\widehat{K}=\mathbb{C}(\!(z)\!).

Sometimes we add subscript ‘z’ or ‘w’ to emphasize the local coordinate, for instance, Rz={z}R_{z}=\mathbb{C}\{z\} and Rw={w}R_{w}=\mathbb{C}\{w\}. Let 𝔻=SpecR\mathbb{D}={\rm Spec}\,R and 𝔻^=SpecR^\widehat{\mathbb{D}}={\rm Spec}\,\widehat{R} be the disc and formal disc, respectively, and then

𝔻: punctured disc,\displaystyle\mathbb{D}^{\ast}:\text{ punctured disc},
𝔻^: formal punctured disc.\displaystyle\widehat{\mathbb{D}}^{\ast}:\text{ formal punctured disc}.

Let GG be a connected complex reductive group with a given maximal torus TT. Let 𝔤\mathfrak{g} (resp. 𝔱\mathfrak{t}) be the Lie algebra of GG (resp. TT). Denote by \mathcal{R} the set of roots.

Let VV be a GG-bundle on 𝔻\mathbb{D}^{\ast} with a GG-connection \nabla and a GG-connection in this paper is always assumed to be algebraic. To simplify the terminology, such a pair (V,)(V,\nabla) is also called a GG-connection. With respect to the local coordinate ‘z’, we write \nabla as

=d+A(z)dz,\displaystyle\nabla=d+A(z)dz,

where dd is the exterior differential and A(z)𝔤(K)A(z)\in\mathfrak{g}(K) is the connection form. Note that the connection form A(z)A(z) of \nabla is in 𝔤(K)\mathfrak{g}(K) because we are working on connections with irregular singularities. Two GG-connections 1=d+A1(z)dz\nabla_{1}=d+A_{1}(z)dz and 2=d+A2(z)dz\nabla_{2}=d+A_{2}(z)dz are (gauge) equivalent if there exists gG(K)g\in G(K) such that g1=2g\circ\nabla_{1}=\nabla_{2}, i.e.

A2(z)=gg1+gA1(z)g1.\displaystyle A_{2}(z)=-g^{\prime}\cdot g^{-1}+gA_{1}(z)g^{-1}.

Moreover, we say that 1\nabla_{1} and 2\nabla_{2} are formally (gauge) equivalent if there exists an element gG(K^)g\in G(\widehat{K}) such that g1=2g\circ\nabla_{1}=\nabla_{2}.

Now given a GG-connection =d+A(z)dz\nabla=d+A(z)dz, we consider the following set

G^():={gG(K^)|g is a G-connection}.\displaystyle\widehat{G}(\nabla):=\{g\in G(\widehat{K})\,|\,g\circ\nabla\emph{ is a $G$-connection}\}.

Equivalently, this set can be regarded as

G^()={gG(K^)|gg1+gA(z)g1𝔤(K)}.\displaystyle\widehat{G}(\nabla)=\{g\in G(\widehat{K})\,|\,-g^{\prime}\cdot g^{-1}+gA(z)g^{-1}\in\mathfrak{g}(K)\}.

Clearly,

G(K)G^()G(K^).\displaystyle G(K)\subseteq\widehat{G}(\nabla)\subseteq G(\widehat{K}).

The quotient set G^()/G(K)\widehat{G}(\nabla)/G(K) classifies all GG-connections which are formally equivalent to \nabla. Moreover, taking an arbitrary element G^()\nabla^{\prime}\in\widehat{G}(\nabla), we have G^()=G^()\widehat{G}(\nabla)=\widehat{G}(\nabla^{\prime}), and thus G^()/G(K)=G^()/G(K)\widehat{G}(\nabla)/G(K)=\widehat{G}(\nabla^{\prime})/G(K), which implies that the quotient does not depend on the choice of representatives.

When G=GLn()G={\rm GL}_{n}(\mathbb{C}), the Malgrange–Sibuya isomorphism theorem [Mal78, Mal83, Sib77] describes G^()/G(K)\widehat{G}(\nabla)/G(K) as a non-abelian cohomological set. The argument can be applied to complex reductive groups in the same way. We briefly review the setup and state the result. For convenience, let 𝔻=\mathbb{D}=\mathbb{C} and denote by 𝔻~𝔻\widetilde{\mathbb{D}}\rightarrow\mathbb{D} the real oriented blowup of 𝔻\mathbb{D} at 0𝔻0\in\mathbb{D}, i.e. 𝔻~=S1×[0,+)\widetilde{\mathbb{D}}=S^{1}\times[0,+\infty). Let UU be an open subset of S1S^{1}, and we define an open subset

U~={(ρ,θ)𝔻~|ρ>0,θU}\displaystyle\widetilde{U}=\{(\rho,\theta)\in\widetilde{\mathbb{D}}\,|\,\rho>0,\,\theta\in U\}

of 𝔻~\widetilde{\mathbb{D}}. Now, we define a nonabelian sheaf Λ\Lambda_{\nabla} on S1S^{1}, of which a germ gg at θS1\theta\in S^{1} is an (holomorphic) element in G(𝒪(U~))G(\mathcal{O}(\widetilde{U})) such that

  • gg is asymptotic to the identity on U~\widetilde{U} at 0;

  • g=g\circ\nabla=\nabla.

We refer the reader to [Mal83, §3] and [LR94, I.2] for more details about the construction of this sheaf. With a similar argument as in the case of G=GLn()G={\rm GL}_{n}(\mathbb{C}), we have the following Malgrange–Sibuya theorem for reductive groups:

Theorem 2.1 (Theorem A.1 in [Boa14]).

There exists a bijection between the sets G^()/G(K)\widehat{G}(\nabla)/G(K) and H1(S1,Λ)H^{1}(S^{1},\Lambda_{\nabla}).

To give a more precise description of the cohomology H1(S1,Λ)H^{1}(S^{1},\Lambda_{\nabla}), we first introduce irregular types. An element in the following form

Q(z)=qnzn/r++q1z1/r\displaystyle Q(z)=q_{-n}z^{-n/r}+\dots+q_{-1}z^{-1/r}

for some positive integers nn and rr, where qi𝔱q_{-i}\in\mathfrak{t}, is called an irregular type. Under the substitution z=wrz=w^{r}, we have

Q(w)=qnwn++q1w1.\displaystyle Q(w)=q_{-n}w^{-n}+\dots+q_{-1}w^{-1}.

The substitution z=wrz=w^{r} can be regarded as choosing a covering 𝔻𝔻\mathbb{D}\rightarrow\mathbb{D}. An irregular type is called unramified if r=1r=1, and ramified if r2r\geq 2.

Definition 2.2.

Given an irregular type QQ, a GG-connection =d+A(z)dz\nabla=d+A(z)dz is with irregular type QQ, if under the substitution z=wrz=w^{r}, the connection d+A(w)d(wd)d+A(w)d(w^{d}) is formally equivalent (under the action of G(K^w)G(\widehat{K}_{w})) to a connection in the form

d+dQ+b1dww\displaystyle d+dQ+b_{-1}\frac{dw}{w}

such that [Q,b1]=0[Q,b_{-1}]=0, where b1𝔤b_{-1}\in\mathfrak{g}.

It has been proven that any GG-connection is of a certain irregular type.

Theorem 2.3 ([BV83]).

Given any GG-connection d+A(z)dzd+A(z)dz, where A(z)=anzn+𝔤(Kz)A(z)=a_{-n}z^{-n}+\dots\in\mathfrak{g}(K_{z}), there exists a positive integer rr such that under the substitution z=wrz=w^{r}, the GG-connection d+A(w)dwd+A(w)dw is formally equivalent to d+B(w)dwd+B(w)dw (under the action of G(K^w)G(\widehat{K}_{w})), where B(w)=bnwn+𝔤(Kw)B(w)=b_{-n^{\prime}}w^{-n^{\prime}}+\dots\in\mathfrak{g}(K_{w}) such that

  • bi𝔱b_{i}\in\mathfrak{t} for i2i\leq-2;

  • bi=0b_{i}=0 for i0i\geq 0;

  • [bi,bj]=0[b_{i},b_{j}]=0.

Remark 2.4.

In the above theorem, although the integer rr is not unique, in this paper we always assume that the integer rr we choose is the smallest one.

Now we fix a GG-connection =d+A(z)dz\nabla=d+A(z)dz with irregular type QQ. In the case of G=GLn()G={\rm GL}_{n}(\mathbb{C}), Loday-Richaud gives a constructive description of H1(S1,Λ)H^{1}(S^{1},\Lambda_{\nabla}) from unipotent Lie groups [LR94]. The arguments can be applied to complex reductive groups as well. In the following, we only give the statement and refer the reader to [Boa14, BY15] for the description.

Given a root α\alpha\in\mathcal{R}, it determines a meromorphic function qα(z):=α(Q(z))q_{\alpha}(z):=\alpha(Q(z)). A direction dS1d\in S^{1} is an anti-Stokes direction (supported by α\alpha) if the meromorphic function exp(qα(z))\exp(q_{\alpha}(z)) has maximal decay as zz goes to zero in the direction. Denote by 𝔸\mathbb{A} the set of all anti-Stokes directions with respect to the given irregular type QQ. Given an anti-Stokes direction d𝔸d\in\mathbb{A}, let (d)\mathcal{R}(d)\subseteq\mathcal{R} be the subset of roots supporting dd, and for each α\alpha\in\mathcal{R}, let Uα:=exp(𝔤α)GU_{\alpha}:=\exp(\mathfrak{g}_{\alpha})\subseteq G be the corresponding unipotent subgroup. Denote by 𝕊tod\mathbb{S}{\rm to}_{d} the image of the product map α(d)UαG\prod_{\alpha\in\mathcal{R}(d)}U_{\alpha}\rightarrow G, which is a unipotent group [LR94, I.4.8]. We define

𝕊to(Q):=d𝔸𝕊tod.\displaystyle\mathbb{S}{\rm to}(Q):=\prod_{d\in\mathbb{A}}\mathbb{S}{\rm to}_{d}.
Theorem 2.5 (Theorem A.2 in [Boa14]).

There is a bijection

𝕊to(Q)H1(S1,Λ).\displaystyle\mathbb{S}{\rm to}(Q)\rightarrow H^{1}(S^{1},\Lambda_{\nabla}).
Remark 2.6.

In [Boa14, Appendix], although the author only deals with the unramified case, the results and arguments hold for the ramified case. Therefore, we only state Theorem 2.1 and Theorem 2.5 without a proof.

In [BY15], the authors define a local system \mathcal{I} on S1S^{1}, of which sections over sectors are functions in the form Q=i=1nqizi/rQ=\sum_{i=1}^{n}q_{-i}z^{-i/r}, where qi/rq_{-i/r}\in\mathbb{C} and n,rn,r\in\mathbb{N}. The sheaf \mathcal{I} can be regarded as a vast disjoint union of circle coverings of S1S^{1}, and each component of \mathcal{I} (i.e. an element in π0()\pi_{0}(\mathcal{I})) is a covering of S1S^{1}. Given a point pS1p\in S^{1}, the fiber p\mathcal{I}_{p} is a free \mathbb{Z}-module. We define a pro-tori 𝒯p:=Hom(p,)\mathcal{T}_{p}:={\rm Hom}(\mathcal{I}_{p},\mathbb{C}^{*}), where a pro-tori is an inverse limit of torus. We obtain a system 𝒯\mathcal{T} of pro-tori over S1S^{1}, of which 𝒯p\mathcal{T}_{p} is the fiber at pS1p\in S^{1}.

Definition 2.7.

An \mathcal{I}-graded GG-local system on S1S^{1} is a GG-local system LL on S1S^{1} together with a morphism 𝒯Aut(L)\mathcal{T}\rightarrow{\rm Aut}(L) of local systems over S1S^{1} factoring through an algebraic quotient of 𝒯\mathcal{T}.

Boalch–Yamakawa proved the following equivalence of categories.

Theorem 2.8 (Theorem 6 in [BY15]).

The category of GG-connections on 𝔻^\widehat{\mathbb{D}}^{*} is equivalent to the category of \mathcal{I}-graded GG-local systems on S1S^{1}.

Remark 2.9.

Let LL be an \mathcal{I}-graded GG-local system on S1S^{1}, and denote by \nabla the corresponding GG-connection on 𝔻^\widehat{\mathbb{D}}^{*}. We briefly state how irregular types of \nabla corresponds to morphisms 𝒯Aut(L)\mathcal{T}\rightarrow{\rm Aut}(L). We fix a point pS1p\in S^{1}, and then we obtain a morphism 𝒯pAut(Lp)G\mathcal{T}_{p}\rightarrow{\rm Aut}(L_{p})\cong G by restricting to pp. Moreover, we suppose that the image of the morphism lies in the maximal torus TT of GG. Since the image of 𝒯pT\mathcal{T}_{p}\rightarrow T is a quotient of 𝒯p\mathcal{T}_{p}, it is equivalent to a finite rank free \mathbb{Z}-submodule of p\mathcal{I}_{p}, of which the generators can be regarded as irregular types. In this sense, we say that a \mathcal{I}-graded GG-local system is of irregular type QQ, if the corresponding connection \nabla is with irregular type QQ. Note that in [BY15], it is called irregular classes of \mathcal{I}-graded GG-local systems, while in this paper, we use the terminology irregular types for convenience.

Recall that 𝔻~\widetilde{\mathbb{D}} is the real oriented blow-up of 𝔻\mathbb{D} at zero, and the zero point is usually denoted by xx. Denote by =S1\partial=S^{1} the boundary circle. We draw a concentric circle (a halo) \partial^{\prime} on 𝔻~\widetilde{\mathbb{D}}, and denote by \mathbb{H} the region between \partial and \partial^{\prime}. In other words, \mathbb{H} is regarded as a tubular neighbourhood of \partial with another boundary circle \partial^{\prime}. We puncture \partial^{\prime} at #𝔸\#\mathbb{A} many distinct points and denote them by {xd,d𝔸}\{x_{d},\,d\in\mathbb{A}\}. According to the anti-Stokes directions, we require that all the #𝔸x\#\mathbb{A}_{x} auxiliary small cilia between each anti-Stokes direction and its nearby puncture do not cross (see the following picture for example).

[Uncaptioned image]

Denote by 𝔻Q\mathbb{D}_{Q} the punctured surface obtained in the above way. Moreover, we have

𝔻Q𝔻~𝔻.\displaystyle\mathbb{D}_{Q}\hookrightarrow\widetilde{\mathbb{D}}\rightarrow\mathbb{D}.
Definition 2.10.

A Stokes GG-local system with irregular type QQ on 𝔻\mathbb{D}^{*} is a GG-local system LL on 𝔻Q\mathbb{D}_{Q} such that L|L|_{\mathbb{H}} is with irregular type QQ and the monodromy around each puncture xdx_{d} in 𝔻Q\mathbb{D}_{Q} lies in 𝕊tod\mathbb{S}{\rm to}_{d}.

Remark 2.11.

Since \mathbb{H} is a tubular neighbourhood of \partial, the fundamental group of \mathbb{H} is isomorphic to the fundamental group of \partial. Then, the category of GG-local systems on \mathbb{H} is equivalent to the category of GG-local systems on \partial. Thus, a GG-local system on \mathbb{H} with irregular type QQ actually means that the corresponding GG-local system on \partial is with irregular type QQ.

Theorem 2.12 (Local Correspondence).

The category of GG-connections with irregular type QQ on 𝔻\mathbb{D}^{*} is equivalent to the category of Stokes GG-local systems with irregular type QQ on 𝔻\mathbb{D}^{*}.

Proof.

We fix a GG-connection 0\nabla_{0} with irregular type QQ. Then, isomorphism classes of GG-connections with irregular type QQ are classified by G^(0)/G(K)\widehat{G}(\nabla_{0})/G(K). As we discussed above Theorem 2.1, the quotient set does not depend on the choice of 0\nabla_{0}. Given an arbitrary GG-connection \nabla^{\prime} with irregular type QQ, denote by LL^{\prime} (resp. L0L_{0}) the corresponding GG-local system of \nabla^{\prime} (resp. 0\nabla_{0}). Moreover, we regard L0L_{0} as a GG-local system on \mathbb{H}, while LL^{\prime} as a GG-local system on 𝔻\mathbb{D}^{*}. By Theorem 2.1 and 2.5, there is a bijection between 𝕊to(Q)=d𝔸𝕊tod\mathbb{S}{\rm to}(Q)=\prod_{d\in\mathbb{A}}\mathbb{S}{\rm to}_{d} and G^(0)/G(K)\widehat{G}(\nabla_{0})/G(K). Denote by (γd)d𝔸(\gamma_{d})_{d\in\mathbb{A}} the corresponding elements of \nabla^{\prime} in d𝔸𝕊tod\prod_{d\in\mathbb{A}}\mathbb{S}{\rm to}_{d}, where γd𝕊tod\gamma_{d}\in\mathbb{S}{\rm to}_{d}. Then, we glue L0L_{0} and LL^{\prime} via γd\gamma_{d} around each puncture xdx_{d}. Thus, we obtain a GG-local system on 𝔻Q\mathbb{D}_{Q}, which is clear a Stokes GG-local system with irregular type QQ on 𝔻\mathbb{D}^{*}.

On the other hand, given a Stokes GG-local system LL with irregular type QQ on 𝔻Q\mathbb{D}_{Q}, we obtain two GG-local systems LL^{\prime} and L0L_{0} by taking restrictions to 𝔻\mathbb{D}^{*} and \mathbb{H} respectively. Under isomorphisms, we suppose that L0L_{0} is the GG-local system given by the GG-connection 0\nabla_{0}. Since LL is a Stokes GG-local system with irregular type QQ, the monodromy γd\gamma_{d} around puncture xdx_{d} for each d𝔸d\in\mathbb{A} gives an element (γd)d𝔸𝕊to(Q)(\gamma_{d})_{d\in\mathbb{A}}\in\mathbb{S}{\rm to}(Q). Therefore, we obtain a GG-connection with irregular type QQ. ∎

2.2. Global Correspondence

Let XX be a connected smooth projective algebraic curve over \mathbb{C}. Let 𝑫\boldsymbol{D} be a given collection of finitely many distinct points on XX, which is also regarded as a reduced effective divisor on XX, and denote by X𝑫:=X\𝑫X_{\boldsymbol{D}}:=X\backslash\boldsymbol{D} the punctured curve, which is also called a noncompact curve. Let X~\widetilde{X} be the real oriented blow-up of XX at each puncture x𝑫x\in\boldsymbol{D}. It is equivalent to consider that X~\widetilde{X} is obtained from XX by replacing each puncture x𝑫x\in\boldsymbol{D} by an oriented boundary circle x\partial_{x}, of which points are considered to be oriented directions emanating from xx. Now we equip each puncture x𝑫x\in\boldsymbol{D} with an irregular type QxQ_{x}, and denote by 𝑸={Qx,x𝑫}\boldsymbol{Q}=\{Q_{x},\,x\in\boldsymbol{D}\} the collection with irregular types. For each x𝑫x\in\boldsymbol{D}, let 𝔸x\mathbb{A}_{x} be the set of anti-Stokes directions of QxQ_{x}. Then, we draw a concentric circle (a halo) x\partial^{\prime}_{x} on X~\widetilde{X} near x\partial_{x}. Denote by x\mathbb{H}_{x} the region between x\partial_{x} and x\partial^{\prime}_{x}, which is a tubular neighbourhood of x\partial_{x}. Then, we puncture x\partial^{\prime}_{x} at #𝔸x\#\mathbb{A}_{x} distinct points according to anti-Stokes directions such that all the #𝔸x\#\mathbb{A}_{x} auxiliary small cilia between each anti-Stokes direction and its nearby puncture do not cross. Denote by {xd,d𝔸x}\{x_{d},\,d\in\mathbb{A}_{x}\} the collection of punctures with respect to point x𝑫x\in\boldsymbol{D}. Finally, let X𝑸X~X_{\boldsymbol{Q}}\hookrightarrow\widetilde{X} be the punctured surface obtained in the above way, which is called the irregular curve given by 𝑸\boldsymbol{Q}.

Definition 2.13.

A Stokes GG-local system with irregular type 𝐐\boldsymbol{Q} on X𝑫X_{\boldsymbol{D}} is a GG-local system LL on X𝑸X_{\boldsymbol{Q}} such that for each puncture x𝑫x\in\boldsymbol{D}, the restriction L|xL|_{\mathbb{H}_{x}} is with irregular type QxQ_{x} (up to isomorphism) and the monodromy around each puncture xdx_{d} lies in the Stokes group 𝕊tod\mathbb{S}{\rm to}_{d} for every d𝔸xd\in\mathbb{A}_{x}.

Theorem 2.14 (Global Correspondence).

The category of GG-connections with irregular type 𝐐\boldsymbol{Q} on VV, is equivalent to the category of Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}}.

Proof.

This is an immediate result of the local correspondence (Theorem 2.12). ∎

If we do not fix a specific irregular type, then we have the following correspondence, which is regarded as the GG-version of the classical irregular Riemmann–Hilbert correspondence of Deligne, Malgrange, Sibuya, Loday-Richaud [LR94, Mal83, Sib77]. Moreover, Hohl and Jakob give a different proof of Theorem 2.14 via Tannakian categories [HJ24, §3].

Corollary 2.15.

The category of GG-connections on X𝐃X_{\boldsymbol{D}} is equivalent to the category of Stokes GG-local systems on X𝐃X_{\boldsymbol{D}}.

Now we will introduce the fundamental groupoid of X𝑸X_{\boldsymbol{Q}} and show that the category of Stokes GG-local systems with irregular type 𝑸\boldsymbol{Q} on X𝑫X_{\boldsymbol{D}} is equivalent to the category of specific GG-representations of the fundamental groupoid of X𝑸X_{\boldsymbol{Q}}. We first introduce two sets H()H(\partial) and HH determined by an irregular type. Given an irregular type QQ, denote by HH the centralizer of Q=qnzn/r++q1z1/rQ=q_{n}z^{-n/r}+\dots+q_{-1}z^{-1/r} in GG. More precisely,

H={kG|[k,qi]=0 for each i.}\displaystyle H=\{k\in G\,|\,[k,q_{-i}]=0\text{ for each $i$.}\}

Denote by H()GH(\partial)\subseteq G the subset of formal monodromies given by QQ. In the unramfied case, H=H()H=H(\partial). Boalch–Yamakawa proved the following result:

Lemma 2.16 (Lemma 15 in [BY15]).

The (H×H)(H\times H)-action on H()H(\partial) via (h1,h2)(m)=h1mh2(h_{1},h_{2})(m)=h_{1}mh_{2} gives a HH-bitorsor structure on H()H(\partial).

Now we fix a base point b0X𝑸b_{0}\in X_{\boldsymbol{Q}}, which is not in the boundary circle x\partial_{x} for each x𝑫x\in\boldsymbol{D}. For each boundary circle x\partial_{x} of X𝑸X_{\boldsymbol{Q}}, we choose a base point bxb_{x}, and denote by 𝒃:={b0,bx,x𝑫}\boldsymbol{b}:=\{b_{0},b_{x},x\in\boldsymbol{D}\} the set of base points. Let Π1(X𝑸,𝒃)\Pi_{1}(X_{\boldsymbol{Q}},\boldsymbol{b}) be the fundamental groupoid of X𝑸X_{\boldsymbol{Q}} with 𝒃\boldsymbol{b} as the set of base points. Here is an explicit description of generators of Π1(X𝑸,𝒃)\Pi_{1}(X_{\boldsymbol{Q}},\boldsymbol{b}):

  1. (1)

    α1,β1,,αg,βg\alpha_{1},\beta_{1},\cdots,\alpha_{g},\beta_{g} are loops based at b0b_{0} determined by the genus of XX;

  2. (2)

    for each x𝑫x\in\boldsymbol{D}, the simple closed loop γx\gamma_{x} based at bxb_{x} goes once around x\partial_{x};

  3. (3)

    for each d𝔸xd\in\mathbb{A}_{x}, the loop γx,d\gamma_{x,d} based at bxb_{x} goes once around the nearby puncture xdx_{d} so that xdx_{d} is the only puncture inside γx,d\gamma_{x,d};

  4. (4)

    for each base point bxb_{x}, the simple path γ0x\gamma_{0x} connects b0b_{0} and bxb_{x}.

For the relations of Π1(X𝑸,𝒃)\Pi_{1}(X_{\boldsymbol{Q}},\boldsymbol{b}), for each x𝑫x\in\boldsymbol{D}, we define

(\ast) μx=γ0x1γx(d𝔸xγx,d)γ0x,\mu_{x}=\gamma_{0x}^{-1}\cdot\gamma_{x}\cdot\left(\prod_{d\in\mathbb{A}_{x}}\gamma_{x,d}\right)\cdot\gamma_{0x},

which is a loop based at b0b_{0}. Then, the relation of Π\Pi is

(i=1g[αi,βi])(x𝑫μx)=id.\displaystyle\left(\prod_{i=1}^{g}[\alpha_{i},\beta_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu_{x}\right)={\rm id}.

In the above setup, γx\gamma_{x} is usually regarded as formal monodromy, while μx\mu_{x} is regarded as topological monodromy. In fact, the definition of Π1(X𝑸,𝒃)\Pi_{1}(X_{\boldsymbol{Q}},\boldsymbol{b}) does not depend on the choice of base points, for which use the notation Π\Pi as an abbreviation. Denote by Ω\Omega the free group generated by generators of Π\Pi and there is a natural surjection ΩΠ\Omega\rightarrow\Pi.

Let Hom(Π,G)\mathrm{Hom}(\Pi,G) be the space of GG-representations of Π\Pi. An element (point) ρHom(Π,G)\rho\in\mathrm{Hom}(\Pi,G) is called a Stokes GG-representation with irregular type 𝐐\boldsymbol{Q} on X𝑫X_{\boldsymbol{D}} if for each x𝑫x\in\boldsymbol{D} and d𝔸xd\in\mathbb{A}_{x}, we have ρ(γx)H(x)\rho(\gamma_{x})\in H(\partial_{x}), which is the set of formal monodromies given by QxQ_{x}, and ρ(γx,d)𝕊tod\rho(\gamma_{x,d})\in\mathbb{S}\mathrm{to}_{d}. Denote by Hom𝕊(Π,G)\mathrm{Hom}_{\mathbb{S}}(\Pi,G) the space of all Stokes GG-representations with irregular type 𝑸\boldsymbol{Q}, which is a smooth affine variety. Here is an equivalent description of Hom𝕊(Π,G)\mathrm{Hom}_{\mathbb{S}}(\Pi,G) with respect to generators and relations of Π\Pi. For each x𝑫x\in\boldsymbol{D}, we define

𝒜(Qx)=H(x)×𝕊to(Qx).\displaystyle\mathcal{A}(Q_{x})=H(\partial_{x})\times\mathbb{S}{\rm to}(Q_{x}).

We consider the closed subvariety

Hom𝕊(Ω,G):=((G×G)g×x𝑫(G×𝒜(Qx)))Hom(Ω,G).\displaystyle{\rm Hom}_{\mathbb{S}}(\Omega,G):=\left((G\times G)^{g}\times\prod_{x\in\boldsymbol{D}}(G\times\mathcal{A}(Q_{x}))\right)\subseteq{\rm Hom}(\Omega,G).

Given a GG-representation ρ:ΩG\rho:\Omega\rightarrow G, we use the following notations

ai=ρ(αi),bi=ρ(βi),ρ(γx)=hx,ρ(γx,d)=Sx,d,ρ(γ0x)=cx.\displaystyle a_{i}=\rho(\alpha_{i}),\ b_{i}=\rho(\beta_{i}),\ \rho(\gamma_{x})=h_{x},\ \rho(\gamma_{x,d})=S_{x,d},\ \rho(\gamma_{0x})=c_{x}.

Then Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi,G) includes all data

((ai,bi)1ig,(cx,hx,Sx,d)x𝑫,d𝔸x)Hom𝕊(Ω,G)\displaystyle((a_{i},b_{i})_{1\leq i\leq g},(c_{x},h_{x},S_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Omega,G)

such that

(i=1g[ai,bi])(x𝑫(cx1hx(d𝔸xSx,d)cx))=id.\displaystyle\left(\prod_{i=1}^{g}[a_{i},b_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}(c_{x}^{-1}h_{x}(\prod_{d\in\mathbb{A}_{x}}S_{x,d})c_{x})\right)=\mathrm{id}.

Recall that HxH_{x} is the stabilizer of QxQ_{x} for x𝑫x\in\boldsymbol{D}. We define

𝑯:=x𝑫Hx,𝑯():=x𝑫H(x).\displaystyle\boldsymbol{H}:=\prod_{x\in\boldsymbol{D}}H_{x},\ \ \boldsymbol{H}(\partial):=\prod_{x\in\boldsymbol{D}}H(\partial_{x}).

There is a (G×𝑯)(G\times\boldsymbol{H})-action on Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi,G) given as follows:

(g,(kx\displaystyle(g,(k_{x} )x𝑫)((ai,bi)1ig,(cx,hx,Sx,d)x𝑫,d𝔸x):=\displaystyle)_{x\in\boldsymbol{D}})\cdot((a_{i},b_{i})_{1\leq i\leq g},(c_{x},h_{x},S_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}):=
((gaig1,gbig1)1ig,(kxcxg1,kxhxkx1,kxSx,dkx1)x𝑫,d𝔸x).\displaystyle\hskip 50.00008pt((ga_{i}g^{-1},gb_{i}g^{-1})_{1\leq i\leq g},(k_{x}c_{x}g^{-1},k_{x}h_{x}k^{-1}_{x},k_{x}S_{x,d}k_{x}^{-1})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Under this action, Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi,G) becomes a (twsited) quasi-Hamiltonian (G×𝑯)(G\times\boldsymbol{H})-space with moment map

μ:Hom𝕊(Π,G)\displaystyle\mu:{\rm Hom}_{\mathbb{S}}(\Pi,G) G×𝑯(),\displaystyle\to G\times\boldsymbol{H}(\partial),
ρ\displaystyle\rho (x𝑫(cx1hx(d𝔸xSx,d)cx),(hx1)x𝑫).\displaystyle\mapsto(\prod_{x\in\boldsymbol{D}}\Big{(}c_{x}^{-1}h_{x}(\prod_{d\in\mathbb{A}_{x}}S_{x,d})c_{x}\Big{)},(h_{x}^{-1})_{x\in\boldsymbol{D}}).

As a result, the quotient B(X𝑫,G,𝑸):=Hom𝕊(Π,G)//(G×𝑯)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q}):={\rm Hom}_{\mathbb{S}}(\Pi,G)/\!\!/(G\times\boldsymbol{H}), which is called wild character variety, exhibits a structure of an algebraic Poisson variety with symplectic leaves [Boa14, BY15]. Two Stokes GG-representations with irregular type 𝑸\boldsymbol{Q} are isomorphic if they are in the same (G×𝑯)(G\times\boldsymbol{H})-orbit.

Theorem 2.17 (Theorem A.3 in [Boa14]).

There is a one-to-one correspondence between (G×𝐇)(G\times\boldsymbol{H})-orbits in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi,G) and isomorphism classes of Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}}. Thus, the category of Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}} is equivalent to the category of Stokes GG-representations with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}}.

2.3. Stokes G-representations

In the previous subsection, we follow Boalch’s idea to construct the fundamental groupoid Π\Pi of X𝑸X_{\boldsymbol{Q}} with respect to a collection of base points 𝒃={b0,bxx𝑫}\boldsymbol{b}=\{b_{0},b_{x}\,x\in\boldsymbol{D}\}, and then in the definition of the fundamental groupoid Π\Pi, it has a path (a generator) γ0x\gamma_{0x} connecting b0b_{0} and bxb_{x} for each x𝑫x\in\boldsymbol{D}. In the following, we will define a fundamental group of X𝑸X_{\boldsymbol{Q}} with respect to a single base point and give an equivalent description of the space of Stokes GG-representations.

We define a free group Ω\Omega^{\prime} with generators

  1. (1)

    α1,β1,,αg,βg\alpha^{\prime}_{1},\beta^{\prime}_{1},\dots,\alpha^{\prime}_{g},\beta^{\prime}_{g};

  2. (2)

    γx\gamma^{\prime}_{x} for each x𝑫x\in\boldsymbol{D};

  3. (3)

    γx,d\gamma^{\prime}_{x,d} for each d𝔸xd\in\mathbb{A}_{x}.

Adding a relation

(\ast^{\prime}) (i=1g[αi,βi])(x𝑫μx)=id,\left(\prod_{i=1}^{g}[\alpha^{\prime}_{i},\beta^{\prime}_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu^{\prime}_{x}\right)={\rm id},

where

μx=γx(d𝔸xγx,d),\displaystyle\mu^{\prime}_{x}=\gamma^{\prime}_{x}\cdot\left(\prod_{d\in\mathbb{A}_{x}}\gamma^{\prime}_{x,d}\right),

we obtain a group Π\Pi^{\prime}. There is a morphism ΩΩ\Omega^{\prime}\rightarrow\Omega (resp. ΠΠ\Pi^{\prime}\rightarrow\Pi)

αiαi,βiβi,γxγ0x1γxγ0x,γx,dγ0x1γx,dγ0x,\displaystyle\alpha^{\prime}_{i}\mapsto\alpha_{i},\ \beta^{\prime}_{i}\mapsto\beta_{i},\ \gamma^{\prime}_{x}\mapsto\gamma_{0x}^{-1}\gamma_{x}\gamma_{0x},\ \gamma^{\prime}_{x,d}\mapsto\gamma_{0x}^{-1}\gamma_{x,d}\gamma_{0x},

which induces one Hom(Ω,G)Hom(Ω,G){\rm Hom}(\Omega,G)\rightarrow{\rm Hom}(\Omega^{\prime},G) (resp. Hom(Π,G)Hom(Π,G){\rm Hom}(\Pi,G)\rightarrow{\rm Hom}(\Pi^{\prime},G)). Therefore, the group Π\Pi^{\prime} can be regarded as the fundamental group of X𝑸X_{\boldsymbol{Q}} with respect to a given base point b0b_{0}. Given a GG-representation ρ:ΩG\rho^{\prime}:\Omega^{\prime}\rightarrow G, we introduce the following notations

ai=ρ(αi),bi=ρ(βi),hx=ρ(γx),Sx,d=ρ(γx,d).\displaystyle a^{\prime}_{i}=\rho^{\prime}(\alpha^{\prime}_{i}),\ b^{\prime}_{i}=\rho^{\prime}(\beta^{\prime}_{i}),\ h^{\prime}_{x}=\rho^{\prime}(\gamma^{\prime}_{x}),\ S^{\prime}_{x,d}=\rho^{\prime}(\gamma^{\prime}_{x,d}).

Consider the group x𝑫Gx\prod_{x\in\boldsymbol{D}}G_{x}, where Gx:=GG_{x}:=G. We define an action

(x𝑫Gx)×Hom(Ω,G)Hom(Ω,G)\displaystyle(\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}(\Omega^{\prime},G)\rightarrow{\rm Hom}(\Omega^{\prime},G)

via

(gx)x𝑫((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x):=((ai,bi)1ig,(gx1hxgx,gx1Sx,dgx)x𝑫,d𝔸x).\displaystyle(g_{x})_{x\in\boldsymbol{D}}\cdot((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}):=((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(g^{-1}_{x}h^{\prime}_{x}g_{x},g^{-1}_{x}S^{\prime}_{x,d}g_{x})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Consider the fiber product

((x𝑫Gx)×Hom(Ω,G))×Hom(Ω,G)((G×G)g×x𝑫𝒜(Qx)),\displaystyle\left((\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}(\Omega^{\prime},G)\right)\times_{{\rm Hom}(\Omega^{\prime},G)}\left((G\times G)^{g}\times\prod_{x\in\boldsymbol{D}}\mathcal{A}(Q_{x})\right),

where ((G×G)g×x𝑫𝒜(Qx))Hom(Ω,G)\left((G\times G)^{g}\times\prod_{x\in\boldsymbol{D}}\mathcal{A}(Q_{x})\right)\hookrightarrow{\rm Hom}(\Omega^{\prime},G) is the natural inclusion. The fiber product is a closed subvariety of (x𝑫Gx)×Hom(Ω,G)(\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}(\Omega^{\prime},G), and it includes all points

((gx)x𝑫,((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x))(x𝑫Gx)×Hom(Ω,G)\displaystyle((g_{x})_{x\in\boldsymbol{D}},((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}))\in(\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}(\Omega^{\prime},G)

such that

gx1hxgxH(x),gx1Sx,dgx𝕊tod\displaystyle g^{-1}_{x}h^{\prime}_{x}g_{x}\in H(\partial_{x}),\ g^{-1}_{x}S^{\prime}_{x,d}g_{x}\in\mathbb{S}{\rm to}_{d}

for each x𝑫x\in\boldsymbol{D} and d𝔸xd\in\mathbb{A}_{x}. Then, we define

Hom𝕊(Ω,G):=(((x𝑫Gx)×Hom(Ω,G))×Hom(Ω,G)((G×G)g×x𝑫𝒜(Qx)))|Hom(Ω,G).\displaystyle{\rm Hom}_{\mathbb{S}}(\Omega^{\prime},G):=\left(((\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}(\Omega^{\prime},G))\times_{{\rm Hom}(\Omega^{\prime},G)}((G\times G)^{g}\times\prod_{x\in\boldsymbol{D}}\mathcal{A}(Q_{x}))\right)\bigg{|}_{{\rm Hom}(\Omega^{\prime},G)}.

Clearly, Hom𝕊(Ω,G){\rm Hom}_{\mathbb{S}}(\Omega^{\prime},G) is a locally closed subset and includes all points

((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x)Hom(Ω,G)\displaystyle((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}(\Omega^{\prime},G)

satisfying the condition that for each x𝑫x\in\boldsymbol{D}, there exists gxGg_{x}\in G such that

gx1hxgxH(x),gx1Sx,dgx𝕊tod.\displaystyle g^{-1}_{x}h^{\prime}_{x}g_{x}\in H(\partial_{x}),\ g^{-1}_{x}S^{\prime}_{x,d}g_{x}\in\mathbb{S}{\rm to}_{d}.

Adding the relation (\ast^{\prime}2.3), we obtain a closed subvariety Hom𝕊(Π,G)Hom𝕊(Ω,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G)\hookrightarrow{\rm Hom}_{\mathbb{S}}(\Omega^{\prime},G). Furthermore, the natural GG-action on Hom(Ω,G){\rm Hom}(\Omega^{\prime},G) given by conjugation

g((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x):=((gaig1,gbig1)1ig,(ghxg1,gSx,dg1)x𝑫,d𝔸x)\displaystyle g\cdot((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}):=((ga^{\prime}_{i}g^{-1},gb^{\prime}_{i}g^{-1})_{1\leq i\leq g},(gh^{\prime}_{x}g^{-1},gS^{\prime}_{x,d}g^{-1})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})

induces a GG-action on Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G).

Proposition 2.18.

There is a one-to-one correspondence between (G×𝐇)(G\times\boldsymbol{H})-orbits in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi,G) and GG-orbits in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G).

Proof.

There is a natural morphism

Hom𝕊(Π,G)Hom𝕊(Π,G)\displaystyle{\rm Hom}_{\mathbb{S}}(\Pi,G)\rightarrow{\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G)

given by

((ai,bi)1ig,(cx,hx,Sx,d)x𝑫,d𝔸x)((ai,bi)1ig,(cx1hxcx,cx1Sx,dcx)x𝑫,d𝔸x).\displaystyle((a_{i},b_{i})_{1\leq i\leq g},(c_{x},h_{x},S_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\rightarrow((a_{i},b_{i})_{1\leq i\leq g},(c_{x}^{-1}h_{x}c_{x},c_{x}^{-1}S_{x,d}c_{x})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

In other words,

ai=ai,bi=bi,hx=cx1hxcx,Sx,d=cx1Sx,dcx.\displaystyle a^{\prime}_{i}=a_{i},\quad b^{\prime}_{i}=b_{i},\quad h^{\prime}_{x}=c^{-1}_{x}h_{x}c_{x},\quad S^{\prime}_{x,d}=c^{-1}_{x}S_{x,d}c_{x}.

Given an arbitrary element (g,(kx)x𝑫)G×𝑯(g,(k_{x})_{x\in\boldsymbol{D}})\in G\times\boldsymbol{H}, we have

(g,(kx\displaystyle(g,(k_{x} )x𝑫)((ai,bi)1ig,(cx,hx,Sx,d)x𝑫,d𝔸x)=\displaystyle)_{x\in\boldsymbol{D}})\cdot((a_{i},b_{i})_{1\leq i\leq g},(c_{x},h_{x},S_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})=
((gaig1,gbig1)1ig,(kxcxg1,kxhxkx1,kxSx,dkx1)x𝑫,d𝔸x).\displaystyle\hskip 50.00008pt((ga_{i}g^{-1},gb_{i}g^{-1})_{1\leq i\leq g},(k_{x}c_{x}g^{-1},k_{x}h_{x}k^{-1}_{x},k_{x}S_{x,d}k_{x}^{-1})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Moreover,

g((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x)=((gaig1,gbig1)1ig,(ghxg1,gSx,dg1)x𝑫,d𝔸x).\displaystyle g\cdot((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})=((ga^{\prime}_{i}g^{-1},gb^{\prime}_{i}g^{-1})_{1\leq i\leq g},(gh^{\prime}_{x}g^{-1},gS^{\prime}_{x,d}g^{-1})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Clearly, the image of

(g,(kx)x𝑫)((ai,bi)1ig,(cx,hx,Sx,d)x𝑫,d𝔸x)Hom𝕊(Π,G)\displaystyle(g,(k_{x})_{x\in\boldsymbol{D}})\cdot((a_{i},b_{i})_{1\leq i\leq g},(c_{x},h_{x},S_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Pi,G)

is

g((ai,bi)1ig,(hx,Sx,d)x𝑫,d𝔸x)Hom𝕊(Π,G).\displaystyle g\cdot((a^{\prime}_{i},b^{\prime}_{i})_{1\leq i\leq g},(h^{\prime}_{x},S^{\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G).

The proposition follows directly. ∎

We have the following corollary as a direct result of Theorem 2.17 and Proposition 2.18.

Corollary 2.19.

There is a one-to-one correspondence between GG-orbits in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G) and isomorphism classes of Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}}.

Terminology.

From now on, a representation in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G) will be called a Stokes GG-representation on X𝐃X_{\boldsymbol{D}}.

3. Moduli Space of Filtered Stokes G-Local Systems

In this section, we construct the moduli space of filtered Stokes GG-local systems with irregular type 𝑸\boldsymbol{Q} on X𝑫X_{\boldsymbol{D}}. By Proposition 2.18 and Corollary 2.19, it is equivalent to construct the moduli space for filtered Stokes GG-representations with irregular type 𝑸\boldsymbol{Q}. In §3.1, we first give the stability condition for filtered Stokes GG-representation (Definition 3.4) based on Ramanathan’s approach [Ram75, Ram96]. In §3.2, we give a third construction of the space of Stokes GG-representations, which will be used in the construction of the moduli space. In §3.3, we prove that the stability condition of filtered Stokes GG-representations is equivalent to a stability condition in the sense of GIT (Proposition 3.9), and then we follow King’s approach [Kin94, §2] to construct the moduli space (Theorem 3.10). In this section, since we always fix an irregular type 𝑸\boldsymbol{Q}, if there is no ambiguity, we use the terminology filtered Stokes GG-representations (or filtered Stokes GG-local systems) without mentioning the irregular type.

3.1. Stability Condition of Filtered Stokes G-local Systems

Recall that GG is a connected complex reductive group with a maximal torus TT. Denote by \mathcal{R} the set of roots. There is a natural pairing of cocharacters and characters

,:Hom(𝔾m,T)×Hom(T,𝔾m).\displaystyle\langle\cdot,\cdot\rangle:{\rm Hom}(\mathbb{G}_{m},T)\times{\rm Hom}(T,\mathbb{G}_{m})\rightarrow\mathbb{Z}.

This pairing can be extended to cocharacters and characters with rational coefficients, and a rational cocharacter in this paper is also called a weight. Now we fix a Borel subgroup BB, which includes TT. Let PP be a parabolic subgroup. Given a character χ\chi of PP and a cocharacter μ\mu of TT, we define

μ,χ:=g1μg,χ,\displaystyle\langle\mu,\chi\rangle:=\langle g^{-1}\mu g,\chi\rangle,

where gg satisfies BgPg1B\subseteq gPg^{-1} and g1μgg^{-1}\mu g is a cocharacter of PP. Furthermore, the definition μ,χ\langle\mu,\chi\rangle does not depend on the choice of such element gg.

Let {ei}\{e_{i}\} (resp. {ei}\{e^{*}_{i}\}) be a basis of Hom(𝔾m,T){\rm Hom}(\mathbb{G}_{m},T)\otimes_{\mathbb{Z}}\mathbb{Q} (resp. Hom(T,𝔾m){\rm Hom}(T,\mathbb{G}_{m})\otimes_{\mathbb{Z}}\mathbb{Q}) such that ei,ej=δij\langle e_{i},e^{*}_{j}\rangle=\delta_{ij}. Suppose that GG is semisimple for convenience, and then, {ei}\{e_{i}\} is regarded as a collection of simple coroots, while {ei}\{e^{*}_{i}\} is regarded as the set of the corresponding fundamental weights. Given a cocharacter θ\theta, a character χθ\chi_{\theta} is uniquely determined by the conditions

ei,χθ=θ,ei\displaystyle\langle e_{i},\chi_{\theta}\rangle=\langle\theta,e^{*}_{i}\rangle

for each ii, and similarly, a cocharacter θχ\theta_{\chi} is determined by a given character χ\chi by the conditions

θχ,ei=ei,χ\displaystyle\langle\theta_{\chi},e^{*}_{i}\rangle=\langle e_{i},\chi\rangle

for each ii. Clearly, we have

θ,χ=θχ,χθ.\displaystyle\langle\theta,\chi\rangle=\langle\theta_{\chi},\chi_{\theta}\rangle.

Now let θ\theta be a weight. It determines a parabolic subgroup

Pθ:={gG| the limit limt0θ(t)gθ(t)1 exists }\displaystyle P_{\theta}:=\{g\in G\,|\,\text{ the limit }\lim_{t\rightarrow 0}\theta(t)g\theta(t)^{-1}\text{ exists }\}

with Levi subgroup LθL_{\theta}. Here is another interpretation of PθP_{\theta}. Define

θ:={α|θ,α0}.\displaystyle\mathcal{R}_{\theta}:=\{\alpha\in\mathcal{R}\,|\,\langle\theta,\alpha\rangle\geq 0\}.

Then Pθ=T,Uα,αθP_{\theta}=\langle T,\,U_{\alpha},\,\alpha\in\mathcal{R}_{\theta}\rangle, i.e. PθP_{\theta} is generated by TT and UαU_{\alpha} for αθ\alpha\in\mathcal{R}_{\theta}. On the other hand, given a parabolic subgroup PGP\subseteq G, denote by P\mathcal{R}_{P} the set of roots of PP. Clearly, Pθ=θ\mathcal{R}_{P_{\theta}}=\mathcal{R}_{\theta}. Now we consider a special type of characters, which is called dominant characters and introduced by [Ram75, §2].

Definition 3.1.

Given a parabolic subgroup PP, a character χ\chi of PP is called dominant (resp. anti-dominant) if it is a positive (resp. negative) linear combination of fundamental weights given by roots in P\mathcal{R}_{P}.

In [HS23b], the authors proved the following lemma and a similar argument is also given in [MiR18, Lemma 2.2].

Lemma 3.2 (Lemma 4.6 in [HS23b]).

Given a weight θ\theta, the character χθ\chi_{\theta} is a dominant character of PθP_{\theta}. On the other hand, given a character χ\chi, if it is a dominant character of some parabolic subgroup PP, then PθχPP_{\theta_{\chi}}\supseteq P.

Definition 3.3.

Let 𝜽={θx,x𝑫}\boldsymbol{\theta}=\{\theta_{x},x\in\boldsymbol{D}\} be a collection of weights. A 𝛉\boldsymbol{\theta}-filtered Stokes GG-representation is a Stokes GG-representation ρ\rho^{\prime} such that the formal monodromy hx=ρ(γx)h^{\prime}_{x}=\rho^{\prime}(\gamma^{\prime}_{x}) is conjugate to an element in PθxP_{\theta_{x}} for every x𝑫x\in\boldsymbol{D}. The corresponding Stokes GG-local system is called a 𝛉\boldsymbol{\theta}-filtered Stokes GG-local system.

It is well-known that a connected complex reductive group GG is covered by its Borel subgroups. Clearly, the statement also holds for parabolic subgroups. Then, fixing an arbitrary parabolic subgroup PP, any gGg\in G is conjugate to an element in PP. Therefore, the space of 𝜽\boldsymbol{\theta}-filtered Stokes GG-representations can also be regarded as Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G).

Given a Stokes GG-representation ρ:ΠG\rho^{\prime}:\Pi^{\prime}\rightarrow G, a parabolic subgroup PP is compatible with ρ\rho^{\prime}, if there is a lifting

P{P}Π{\Pi^{\prime}}G{G}ρ\scriptstyle{\rho^{\prime}}

In other words, the representation ρ\rho^{\prime} is well-defined when restricted to PP. Let LL be the Levi subgroup of PP. If PP is compatible with ρ\rho^{\prime}, then ρ\rho^{\prime} is also well-defined by restricting to LL. Under the morphism PLGP\twoheadrightarrow L\rightarrow G, we obtain a GG-representation and denote it by ρL\rho^{\prime}_{L}.

Given a 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho^{\prime}, there exists gxGg_{x}\in G such that gxρ(γx)gx1Pθxg_{x}\rho^{\prime}(\gamma^{\prime}_{x})g^{-1}_{x}\in P_{-\theta_{x}} for each x𝑫x\in\boldsymbol{D}. Suppose that the parabolic subgroup PP is compatible with ρ\rho^{\prime}, and then BθxgxPgx1B_{-\theta_{x}}\subseteq g_{x}Pg^{-1}_{x}, where BθxPθxB_{-\theta_{x}}\subseteq P_{-\theta_{x}} is the Borel subgroup. Let χ\chi be a character of PP and the natural pairing is given as

θx,χ=gx1θxgx,χ=θx,gxχgx1.\displaystyle\langle\theta_{x},\chi\rangle=\langle g^{-1}_{x}\theta_{x}g_{x},\chi\rangle=\langle\theta_{x},g_{x}\chi g_{x}^{-1}\rangle.

We define the degree of a 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho^{\prime} as

deglocρ(P,χ):=𝜽,χ=x𝑫θx,χ.\displaystyle\deg^{\rm loc}\rho^{\prime}(P,\chi):=\langle\boldsymbol{\theta},\chi\rangle=\sum_{x\in\boldsymbol{D}}\langle\theta_{x},\chi\rangle.

Furthermore, a parabolic subgroup PP is admissible with ρ\rho^{\prime} if PP is compatible with ρ\rho^{\prime} and for any character χ:P𝔾m\chi:P\rightarrow\mathbb{G}_{m} trivial on the center, we have deglocρ(P,χ)=0\deg^{\rm loc}\rho^{\prime}(P,\chi)=0.

We follow Ramanathan’s stability condition on principal bundles [Ram75, Ram96] to give the definition of stability condition on filtered Stokes GG-representations (also for filtered Stokes GG-local systems), which is called the RR-stability condition.

Definition 3.4.

A 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho^{\prime} is RR-semistable (resp. RR-stable), if for

  • any proper parabolic subgroup PGP\subseteq G compatible with ρ\rho^{\prime},

  • any nontrivial anti-dominant character χ:P𝔾m\chi:P\rightarrow\mathbb{G}_{m}, which is trivial on the center of PP,

we have

deglocρ(P,χ)0(resp.>0).\displaystyle\deg^{\rm loc}\rho^{\prime}(P,\chi)\geq 0\quad(\text{resp.}>0).

Moreover, two RR-semistable 𝜽\boldsymbol{\theta}-filtered Stokes GG-representations ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} are SS-equivalent if there exist parabolic subgroups P1P_{1} and P2P_{2} (with Levi subgroups L1L_{1} and L2L_{2}) admissible with ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} respectively such that the corresponding Stokes GG-representations (ρ1)L1(\rho^{\prime}_{1})_{L_{1}} and (ρ2)L2(\rho^{\prime}_{2})_{L_{2}} are conjugate under the action of GG.

Definition 3.5.

A 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho^{\prime} is of degree zero, if for any character χ\chi of GG, we have deglocρ(P,χ)=0\deg^{\rm loc}\rho^{\prime}(P,\chi)=0. Note that when GG is semisimple, this condition is always satisfied.

3.2. An Equivalent Construction

In this subsection, we give a third construction of the space of Stokes GG-representations. We define a free group Ω′′\Omega^{\prime\prime} generated by the following elements

  1. (1)

    α1′′,β1′′,,αg′′,βg′′\alpha^{\prime\prime}_{1},\beta^{\prime\prime}_{1},\dots,\alpha^{\prime\prime}_{g},\beta^{\prime\prime}_{g};

  2. (2)

    ιx′′,γx′′\iota^{\prime\prime}_{x},\gamma^{\prime\prime}_{x} for each x𝑫x\in\boldsymbol{D};

  3. (3)

    γx,d′′\gamma^{\prime\prime}_{x,d} for each d𝔸xd\in\mathbb{A}_{x}.

Given a relation

(′′\ast^{\prime\prime}) (i=1g[αi′′,βi′′])(x𝑫μx′′)=id,\left(\prod_{i=1}^{g}[\alpha^{\prime\prime}_{i},\beta^{\prime\prime}_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu^{\prime\prime}_{x}\right)={\rm id},

where

μx′′=ιx′′γx′′(d𝔸xγx,d′′),\displaystyle\mu^{\prime\prime}_{x}=\iota^{\prime\prime}_{x}\cdot\gamma^{\prime\prime}_{x}\cdot\left(\prod_{d\in\mathbb{A}_{x}}\gamma^{\prime\prime}_{x,d}\right),

we obtain a group Π′′\Pi^{\prime\prime}. The natural surjection Ω′′Π′′\Omega^{\prime\prime}\rightarrow\Pi^{\prime\prime} induces a closed embedding Hom(Π′′,G)Hom(Ω′′,G){\rm Hom}(\Pi^{\prime\prime},G)\hookrightarrow{\rm Hom}(\Omega^{\prime\prime},G). Moreover, given a GG-representation ρ′′:Ω′′G\rho^{\prime\prime}:\Omega^{\prime\prime}\rightarrow G (or ρ′′:Π′′G\rho^{\prime\prime}:\Pi^{\prime\prime}\rightarrow G), we introduce the following notations

ai′′=ρ′′(αi′′),bi′′=ρ′′(βi′′),ρ′′(ιx′′)=lx′′,ρ′′(γx′′)=hx′′,ρ′′(γx,d′′)=Sx,d′′.\displaystyle a^{\prime\prime}_{i}=\rho^{\prime\prime}(\alpha^{\prime\prime}_{i}),\ b^{\prime\prime}_{i}=\rho^{\prime\prime}(\beta^{\prime\prime}_{i}),\ \rho^{\prime\prime}(\iota^{\prime\prime}_{x})=l^{\prime\prime}_{x},\ \rho^{\prime\prime}(\gamma^{\prime\prime}_{x})=h^{\prime\prime}_{x},\ \rho^{\prime\prime}(\gamma^{\prime\prime}_{x,d})=S^{\prime\prime}_{x,d}.

We define a morphism ΩΩ′′\Omega^{\prime}\rightarrow\Omega^{\prime\prime} (resp. ΠΠ′′\Pi^{\prime}\rightarrow\Pi^{\prime\prime})

αα′′,ββ′′,γxιx′′γx′′,γx,dγx,d′′,\displaystyle\alpha^{\prime}\mapsto\alpha^{\prime\prime},\ \beta^{\prime}\mapsto\beta^{\prime\prime},\ \gamma^{\prime}_{x}\mapsto\iota^{\prime\prime}_{x}\gamma^{\prime\prime}_{x},\ \gamma^{\prime}_{x,d}\mapsto\gamma^{\prime\prime}_{x,d},

which induces a morphism Hom(Ω′′,G)Hom(Ω,G){\rm Hom}(\Omega^{\prime\prime},G)\rightarrow{\rm Hom}(\Omega^{\prime},G) (resp. Hom(Π′′,G)Hom(Π,G){\rm Hom}(\Pi^{\prime\prime},G)\rightarrow{\rm Hom}(\Pi^{\prime},G)). Taking the fiber product

Hom𝕊(Ω′′,G){{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)}Hom𝕊(Ω,G){{\rm Hom}_{\mathbb{S}}(\Omega^{\prime},G)}Hom(Ω′′,G){{\rm Hom}(\Omega^{\prime\prime},G)}Hom(Ω,G),{{\rm Hom}(\Omega^{\prime},G)\ ,}

we obtain a quasi-projective variety Hom𝕊(Ω′′,G){\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G), which includes all points

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)Hom(Ω′′,G)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}(\Omega^{\prime\prime},G)

satisfying the condition that for each x𝑫x\in\boldsymbol{D}, there exists gxGg_{x}\in G such that

gx1lx′′hx′′gxH(x),gx1Sx,d′′gx𝕊tod.\displaystyle g^{-1}_{x}l^{\prime\prime}_{x}h^{\prime\prime}_{x}g_{x}\in H(\partial_{x}),\ g^{-1}_{x}S^{\prime\prime}_{x,d}g_{x}\in\mathbb{S}{\rm to}_{d}.

Moreover, we define a (x𝑫Gx)(\prod_{x\in\boldsymbol{D}}G_{x})-action on Hom𝕊(Ω′′,G){\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G) via

(gx)x𝑫((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x):=((ai′′,bi′′)1ig,(gx1lx′′,hx′′gx,Sx,d′′)x𝑫,d𝔸x),\displaystyle(g_{x})_{x\in\boldsymbol{D}}\cdot((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}):=((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(g^{-1}_{x}l^{\prime\prime}_{x},h^{\prime\prime}_{x}g_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}),

where Gx=GG_{x}=G for each x𝑫x\in\boldsymbol{D}.

Now given a collection of weights 𝜽={θx,x𝑫}\boldsymbol{\theta}=\{\theta_{x},x\in\boldsymbol{D}\}, denote by 𝑷={Pθx,x𝑫}\boldsymbol{P}=\{P_{-\theta_{x}},x\in\boldsymbol{D}\} the collection of parabolic subgroups. We define a closed subvariety Hom𝕊(Ω′′,𝑷)Hom𝕊(Ω′′,G){\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})\subseteq{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G), of which points

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})

satisfy the condition that

lx′′Lθx,hx′′Pθx\displaystyle l^{\prime\prime}_{x}\in L_{\theta_{x}},\ h^{\prime\prime}_{x}\in P_{-\theta_{x}}

for each x𝑫x\in\boldsymbol{D}. We take the fiber product

Hom~𝕊(Ω′′,𝑷){\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})}Hom𝕊(Ω′′,𝑷){{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})}(x𝑫Gx)×Hom𝕊(Ω′′,G){(\prod_{x\in\boldsymbol{D}}G_{x})\times{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)}Hom𝕊(Ω′′,G).{{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)\ .}

Then we restrict it to Hom(Ω′′,G){\rm Hom}(\Omega^{\prime\prime},G) and define

Hom𝕊(Ω′′,[𝑷]):=Hom~𝕊(Ω′′,𝑷)|Hom𝕊(Ω′′,G).\displaystyle{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},[\boldsymbol{P}]):=\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})|_{{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)}.

The variety Hom𝕊(Ω′′,[𝑷]){\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},[\boldsymbol{P}]) includes all points

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)Hom𝕊(Ω′′,G)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)

such that for each x𝑫x\in\boldsymbol{D}, there exists gxGg_{x}\in G such that gx1lx′′Lθxg^{-1}_{x}l^{\prime\prime}_{x}\in L_{\theta_{x}} and hx′′gxPθxh^{\prime\prime}_{x}g_{x}\in P_{-\theta_{x}}. Then we obtain a closed subvariety Hom𝕊(Π′′,[𝑷])Hom𝕊(Ω′′,[𝑷]){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},[\boldsymbol{P}])\subseteq{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},[\boldsymbol{P}]) by adding the relation (′′\ast^{\prime\prime}3.2). Since the collection 𝑷\boldsymbol{P} of parabolic subgroups is determined by 𝜽\boldsymbol{\theta}, we would like to use the notation

Hom𝕊(Π′′,𝜽):=Hom𝕊(Π′′,[𝑷]).\displaystyle{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}):={\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},[\boldsymbol{P}]).

Define 𝑳=x𝑫Lθx\boldsymbol{L}=\prod_{x\in\boldsymbol{D}}L_{\theta_{x}}. There is a 𝑳\boldsymbol{L}-action on Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) defined as follows

(lx)x𝑫\displaystyle(l_{x})_{x\in\boldsymbol{D}}\,\cdot\, ((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})
:=\displaystyle:= ((ai′′,bi′′)1ig,(lx′′lx1,lxhx′′,Sx,d′′)x𝑫,d𝔸x)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x}l_{x}^{-1},l_{x}h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})

Then we define a (G×𝑳)(G\times\boldsymbol{L})-action on Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta})

(g,(lx)x𝑫)\displaystyle(g,(l_{x})_{x\in\boldsymbol{D}})\,\cdot\, ((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})
:=\displaystyle:= ((gai′′g1,gbi′′g1)1ig,(glx′′lx1,lxhx′′g1,gSx,d′′g1)x𝑫,d𝔸x).\displaystyle((ga^{\prime\prime}_{i}g^{-1},gb^{\prime\prime}_{i}g^{-1})_{1\leq i\leq g},(gl^{\prime\prime}_{x}l_{x}^{-1},l_{x}h^{\prime\prime}_{x}g^{-1},gS^{\prime\prime}_{x,d}g^{-1})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).
Lemma 3.6.

There is a one-to-one correspondence between (G×𝐋)(G\times\boldsymbol{L})-orbits in Hom𝕊(Π′′,𝛉){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) and GG-orbits in Hom𝕊(Π,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G). Therefore, (G×𝐋)(G\times\boldsymbol{L})-orbits in Hom𝕊(Π′′,𝛉){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) are in one-to-one correspondence with isomorphism classes of 𝛉\boldsymbol{\theta}-filtered Stokes GG-representations, and thus isomorphism classes of 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems.

Proof.

The surjective morphism

Hom(Π′′,G)Hom(Π,G)\displaystyle{\rm Hom}(\Pi^{\prime\prime},G)\rightarrow{\rm Hom}(\Pi^{\prime},G)

induces the surjection

Hom𝕊(Π′′,𝜽)Hom𝕊(Π,G)\displaystyle{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta})\rightarrow{\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G)

given by

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)((ai′′,bi′′)1ig,(lx′′hx′′,Sx,d′′)x𝑫,d𝔸x).\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\rightarrow((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x}h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Given an arbitrary element (g,(lx)x𝑫)G×𝑳(g,(l_{x})_{x\in\boldsymbol{D}})\in G\times\boldsymbol{L}, the image of

(g,(lx)x𝑫)((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)Hom𝕊(Π′′,𝜽)\displaystyle(g,(l_{x})_{x\in\boldsymbol{D}})\,\cdot\,((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta})

is exactly

g(ai′′,bi′′)1ig,(lx′′hx′′,Sx,d′′)x𝑫,d𝔸x)Hom𝕊(Π,G).\displaystyle g\,\cdot\,(a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x}h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G).

This finishes the proof of this lemma. ∎

3.3. Moduli Space

In this subsection, we follow King’s approach [Kin94] to construct the moduli space of filtered Stokes GG-local systems. We fix a collection of weights 𝜽\boldsymbol{\theta} and we suppose that dd is the common denominator of θx\theta_{x} for x𝑫x\in\boldsymbol{D}, i.e. dθxd\theta_{x} is a cocharacter for every x𝑫x\in\boldsymbol{D}. In the previous subsection, we construct a quasi-projective variety Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) with a natural (G×𝑳)(G\times\boldsymbol{L})-action such that the (G×𝑳)(G\times\boldsymbol{L})-orbits are in one-to-one correspondence with isomorphism classes of 𝜽\boldsymbol{\theta}-filtered Stokes GG-representations by Lemma 3.6. We will introduce a particular character χ𝜽:G×𝑳𝔾m\chi_{\boldsymbol{\theta}}:G\times\boldsymbol{L}\rightarrow\mathbb{G}_{m} such that a 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho^{\prime} is RR-semistable if and only if the corresponding representation ρ′′Hom𝕊(Π′′,𝜽)\rho^{\prime\prime}\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) is χ𝜽\chi_{\boldsymbol{\theta}}-semistable in the sense of GIT. Based on the equivalence of stability conditions, we use Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) to construct the moduli space.

We define a character

χ𝜽:G×𝑳𝔾m,\displaystyle\chi_{\boldsymbol{\theta}}:G\times\boldsymbol{L}\rightarrow\mathbb{G}_{m},

as

χ𝜽(g,(lx)x𝑫)=χ0(g)x𝑫χdθx(lx),\displaystyle\chi_{\boldsymbol{\theta}}(g,(l_{x})_{x\in\boldsymbol{D}})=\chi_{0}(g)\cdot\prod_{x\in\boldsymbol{D}}\chi_{-d\theta_{x}}(l_{x}),

where χ0\chi_{0} is the trivial character of GG and the character χdθx\chi_{-d\theta_{x}} is determined by the weight dθx-d\theta_{x}. Given a cocharacter λ:𝔾mG×𝑳\lambda:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L}, it is given by a cocharacter λ0\lambda_{0} of GG and a cocharacter λx\lambda_{x} of LθxL_{\theta_{x}} for each x𝑫x\in\boldsymbol{D}. Thus, the pairing λ,χ𝜽\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle is given by

λ,χ𝜽=λ0,χ0+x𝑫λx,χdθx=x𝑫λx,χdθx.\displaystyle\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=\langle\lambda_{0},\chi_{0}\rangle+\sum_{x\in\boldsymbol{D}}\langle\lambda_{x},\chi_{-d\theta_{x}}\rangle=\sum_{x\in\boldsymbol{D}}\langle\lambda_{x},\chi_{-d\theta_{x}}\rangle.

With respect to the (G×𝑳)(G\times\boldsymbol{L})-action and character χ𝜽\chi_{\boldsymbol{\theta}}, King defined the GIT quotient Hom𝕊(Π′′,𝜽)//(G×𝑳,χ𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta})/\!\!/(G\times\boldsymbol{L},\chi_{\boldsymbol{\theta}}), which parametrizes GIT equivalence classes of χ𝜽\chi_{\boldsymbol{\theta}}-semistable points in Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}). We refer the reader to [Kin94, §2] for more details about this construction. Applying [Kin94, Proposition 2.5, 2.6], we have the following equivalent description of χ𝜽\chi_{\boldsymbol{\theta}}-semistable points.

Lemma 3.7.

Denote by Δ\Delta the kernel of the (G×𝐋)(G\times\boldsymbol{L})-action on Hom𝕊(Π′′,𝛉){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}). A point ρ′′Hom𝕊(Π′′,𝛉)\rho^{\prime\prime}\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) is χ𝛉\chi_{\boldsymbol{\theta}}-semistable if and only if χ𝛉(Δ)={1}\chi_{\boldsymbol{\theta}}(\Delta)=\{1\} and any cocharacter λ\lambda of G×𝐋G\times\boldsymbol{L}, for which the limit limt0λ(t)ρ′′\lim_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists, satisfies λ,χ𝛉0\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle\geq 0. It is χ𝛉\chi_{\boldsymbol{\theta}}-stable if and only if any cocharacter λ\lambda, for which limt0λ(t)ρ′′\lim_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists and λ,χ𝛉=0\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=0, is in Δ\Delta. Moreover, two χ𝛉\chi_{\boldsymbol{\theta}}-semistable points ρ1′′\rho^{\prime\prime}_{1} and ρ2′′\rho^{\prime\prime}_{2} are GIT equivalent if and only if there are cocharacters λ1\lambda_{1} and λ2\lambda_{2} such that λ1,χ𝛉=λ2,χ𝛉=0\langle\lambda_{1},\chi_{\boldsymbol{\theta}}\rangle=\langle\lambda_{2},\chi_{\boldsymbol{\theta}}\rangle=0 and the limits limt0λ1(t)ρ1′′\lim_{t\rightarrow 0}\lambda_{1}(t)\cdot\rho^{\prime\prime}_{1} and limt0λ2(t)ρ2′′\lim_{t\rightarrow 0}\lambda_{2}(t)\cdot\rho^{\prime\prime}_{2} are in the same (G×𝐋)(G\times\boldsymbol{L})-orbit.

Remark 3.8.

We give a precise description of the cocharacter λ\lambda such that limt0λ(t)ρ′′\lim\limits_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists, and we regard ρ′′\rho^{\prime\prime} as a tuple

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x).\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}}).

Given a cocharacter λ:𝔾mG×𝑳\lambda:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L}, it is uniquely determined by a cocharacter λ0\lambda_{0} of GG and a cocharacter λx\lambda_{x} of LθxL_{\theta_{x}} for each x𝑫x\in\boldsymbol{D}. Suppose that the limit limt0λ(t)ρ′′\lim\limits_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists. Then the existence of the limits

limt0λ0(t)ai′′λ01(t),limt0λ0(t)bi′′λ01(t),limt0λ0(t)Sx,d′′λ01(t)\displaystyle\lim\limits_{t\rightarrow 0}\lambda_{0}(t)a^{\prime\prime}_{i}\lambda^{-1}_{0}(t),\ \lim\limits_{t\rightarrow 0}\lambda_{0}(t)b^{\prime\prime}_{i}\lambda^{-1}_{0}(t),\ \lim\limits_{t\rightarrow 0}\lambda_{0}(t)S^{\prime\prime}_{x,d}\lambda^{-1}_{0}(t)

implies that ai′′,bi′′,Sx,d′′Pλ0a^{\prime\prime}_{i},b^{\prime\prime}_{i},S^{\prime\prime}_{x,d}\in P_{\lambda_{0}}, and the existence of the limits

limt0λ0(t)lx′′λx1(t),limt0λx(t)hx′′λ01(t)\displaystyle\lim\limits_{t\rightarrow 0}\lambda_{0}(t)l^{\prime\prime}_{x}\lambda^{-1}_{x}(t),\ \lim\limits_{t\rightarrow 0}\lambda_{x}(t)h^{\prime\prime}_{x}\lambda^{-1}_{0}(t)

implies that lx′′hx′′Pλ0l^{\prime\prime}_{x}h^{\prime\prime}_{x}\in P_{\lambda_{0}}. Therefore, the corresponding representation ρHom𝕊(Π,G)\rho^{\prime}\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime},G) of ρ′′\rho^{\prime\prime} (under the morphism ΠΠ′′\Pi^{\prime}\rightarrow\Pi^{\prime\prime}) is compatible with Pλ0P_{\lambda_{0}}. Moreover, we claim that for each x𝑫x\in\boldsymbol{D}, there exists gxGg_{x}\in G such that λx(t)=gx1λ0(t)gx\lambda_{x}(t)=g^{-1}_{x}\lambda_{0}(t)g_{x}, gx1lx′′Lθxg^{-1}_{x}l^{\prime\prime}_{x}\in L_{\theta_{x}} and hx′′gxPθxh^{\prime\prime}_{x}g_{x}\in P_{-\theta_{x}}. Therefore, we have

(\bullet) λ,χ𝜽=x𝑫λx,χdθx=dx𝑫θx,χλx=dx𝑫θx,χλ0=d𝜽,χλ0.\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=\sum_{x\in\boldsymbol{D}}\langle\lambda_{x},\chi_{-d\theta_{x}}\rangle=-d\sum_{x\in\boldsymbol{D}}\langle\theta_{x},\chi_{\lambda_{x}}\rangle=-d\sum_{x\in\boldsymbol{D}}\langle\theta_{x},\chi_{\lambda_{0}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{0}}\rangle.

Here is a brief explanation for the claim. By construction, for each x𝑫x\in\boldsymbol{D}, there exists gxGg^{\prime}_{x}\in G such that gx1lx′′Lθxg^{\prime-1}_{x}l^{\prime\prime}_{x}\in L_{\theta_{x}} and hx′′gxPθxh^{\prime\prime}_{x}g^{\prime}_{x}\in P_{-\theta_{x}}. The existence of the limit

gx1(limt0λ0(t)lx′′hx′′λ01(t))gx=limt0(gx1λ0(t)gx)(gx1lx′′hx′′gx)(gx1λ01(t)gx)\displaystyle g^{\prime-1}_{x}(\lim_{t\rightarrow 0}\lambda_{0}(t)l^{\prime\prime}_{x}h^{\prime\prime}_{x}\lambda^{-1}_{0}(t))g^{\prime}_{x}=\lim_{t\rightarrow 0}(g^{\prime-1}_{x}\lambda_{0}(t)g^{\prime}_{x})(g^{\prime-1}_{x}l^{\prime\prime}_{x}h^{\prime\prime}_{x}g^{\prime}_{x})(g^{\prime-1}_{x}\lambda^{-1}_{0}(t)g^{\prime}_{x})

shows that gx1λ0(t)gxg^{\prime-1}_{x}\lambda_{0}(t)g^{\prime}_{x} is a cocharacter of some maximal torus in LθxL_{\theta_{x}}. We choose lxLθxl_{x}\in L_{\theta_{x}} and define gx:=gxlxg_{x}:=g^{\prime}_{x}l_{x} such that gx1λ0(t)gxg_{x}^{-1}\lambda_{0}(t)g_{x} and λx(t)\lambda_{x}(t) are cocharacters of the same maximal torus in LθxL_{\theta_{x}}. Then we consider the limits

gx1(limt0λ0(t)lx′′λx1(t))=limt0(gx1λ0(t)gx)(gx1lx′′)λx1(t)\displaystyle g^{-1}_{x}(\lim_{t\rightarrow 0}\lambda_{0}(t)l^{\prime\prime}_{x}\lambda^{-1}_{x}(t))=\lim_{t\rightarrow 0}(g^{-1}_{x}\lambda_{0}(t)g_{x})(g^{-1}_{x}l^{\prime\prime}_{x})\lambda^{-1}_{x}(t)
(limt0λx1(t)hx′′λ01(t))gx=limt0λx1(t)(hx′′gx)(gx1λ01(t)gx).\displaystyle(\lim_{t\rightarrow 0}\lambda^{-1}_{x}(t)h^{\prime\prime}_{x}\lambda_{0}^{-1}(t))g_{x}=\lim_{t\rightarrow 0}\lambda^{-1}_{x}(t)(h^{\prime\prime}_{x}g_{x})(g^{-1}_{x}\lambda^{-1}_{0}(t)g_{x}).

Note that

gx1lx′′=lx1gx1lx′′Lθx,hx′′gx=hx′′gxlxPθx,\displaystyle g^{-1}_{x}l^{\prime\prime}_{x}=l^{-1}_{x}g^{\prime-1}_{x}l^{\prime\prime}_{x}\in L_{\theta_{x}},\ h^{\prime\prime}_{x}g_{x}=h^{\prime\prime}_{x}g^{\prime}_{x}l_{x}\in P_{-\theta_{x}},

and gx1λ0(t)gxg^{-1}_{x}\lambda_{0}(t)g_{x} and λx(t)\lambda_{x}(t) are cocharacters of the same maximal torus in LθxL_{\theta_{x}}. Therefore, the existence of the above two limits imply gx1λ0(t)gx=λx(t)g^{-1}_{x}\lambda_{0}(t)g_{x}=\lambda_{x}(t).

Proposition 3.9.

Given a point ρ′′Hom𝕊(Π′′,𝛉)\rho^{\prime\prime}\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}), denote by ρ\rho^{\prime} the corresponding 𝛉\boldsymbol{\theta}-filtered Stokes GG-representation. Then, ρ\rho^{\prime} is RR-semistable (resp. RR-stable) of degree zero if and only if the point ρ′′\rho^{\prime\prime} is χ𝛉\chi_{\boldsymbol{\theta}}-semistable (resp. χ𝛉\chi_{\boldsymbol{\theta}}-stable). Moreover, two χ𝛉\chi_{\boldsymbol{\theta}}-semistable points ρ1′′\rho^{\prime\prime}_{1} and ρ2′′\rho^{\prime\prime}_{2} are GIT equivalent if and only if the corresponding RR-semistable 𝛉\boldsymbol{\theta}-filtered Stokes GG-representations ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} are SS-equivalent.

Proof.

We suppose that GG is semisimple first. We regard ρ′′\rho^{\prime\prime} as a tuple

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)Hom𝕊(Π′′,𝜽).\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}).

Suppose that the point ρ′′\rho^{\prime\prime} is χ𝜽\chi_{\boldsymbol{\theta}}-semistable. We choose a parabolic subgroup PP compatible with ρ\rho^{\prime} and pick an arbitrary anti-dominant character χ\chi of PP, which is trivial on the center of PP. We have

𝜽,χ=1dx𝑫dθx,χ=1dx𝑫dθx,χ=1dx𝑫λχ,χdθx=1dλχ,χ𝜽.\langle\boldsymbol{\theta},\chi\rangle=\frac{1}{d}\sum_{x\in\boldsymbol{D}}\langle d\theta_{x},\chi\rangle=\frac{1}{d}\sum_{x\in\boldsymbol{D}}\langle-d\theta_{x},-\chi\rangle=\frac{1}{d}\sum_{x\in\boldsymbol{D}}\langle\lambda_{-\chi},\chi_{-d\theta_{x}}\rangle=\frac{1}{d}\langle\lambda_{-\chi},\chi_{\boldsymbol{\theta}}\rangle.

The cocharacter λχ\lambda_{-\chi} and the element ρ′′\rho^{\prime\prime} determine a cocharacter λ:𝔾mG𝜽\lambda:\mathbb{G}_{m}\rightarrow G_{\boldsymbol{\theta}} such that

λ=(λ0,λx,x𝑫),λ0:=λχ,λx(t):=gx1λ0(t)gx,\displaystyle\lambda=(\lambda_{0},\lambda_{x},x\in\boldsymbol{D}),\ \lambda_{0}:=\lambda_{-\chi},\ \lambda_{x}(t):=g^{-1}_{x}\lambda_{0}(t)g_{x},

where gxg_{x} is given in Remark 3.8. By Lemma 3.2, we have PλχPP_{\lambda_{-\chi}}\supseteq P. By the compatibility of PP with ρ\rho^{\prime}, the limit limt0λ(t)ρ′′\lim_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists. Since ρ′′\rho^{\prime\prime} is χ𝜽\chi_{\boldsymbol{\theta}}-semistable by assumption, we have λχ,χ𝜽0\langle\lambda_{-\chi},\chi_{\boldsymbol{\theta}}\rangle\geq 0 by Lemma 3.7, and thus 𝜽,χ0\langle\boldsymbol{\theta},\chi\rangle\geq 0. Therefore, ρ\rho^{\prime} is RR-semistable.

To prove that ρ\rho^{\prime} is of degree zero, we suppose that ρ′′\rho^{\prime\prime} is χ𝜽\chi_{\boldsymbol{\theta}}-stable and ρ\rho^{\prime} is RR-stable for convenience. Given a character χ:G𝔾m\chi:G\rightarrow\mathbb{G}_{m}, the corresponding cocharacter λχ\lambda_{\chi} has the property that its image is in the center of GG, and thus in Δ\Delta. In this case, we always have λχ,χ𝜽=0\langle\lambda_{\chi},\chi_{\boldsymbol{\theta}}\rangle=0. Then the formula (\bullet3.8) implies

𝜽,χ=dλχ,χ𝜽=0.\displaystyle\langle\boldsymbol{\theta},\chi\rangle=-d\langle\lambda_{\chi},\chi_{\boldsymbol{\theta}}\rangle=0.

Therefore, ρ\rho^{\prime} is of degree zero.

For the other direction, we suppose that ρ\rho^{\prime} is RR-semistable of degree zero. Clearly, χ𝜽(Δ)={1}\chi_{\boldsymbol{\theta}}(\Delta)=\{1\} because ρ\rho^{\prime} is of degree zero. We take a cocharacter λ:𝔾mG×𝑳\lambda:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L} such that the limit limt0λ(t)ρ′′\lim\limits_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists. Remark 3.8 shows that

ai′′,bi′′,Sx,j′′,lx′′hx′′Pλ0.\displaystyle a^{\prime\prime}_{i},\,b^{\prime\prime}_{i},\,S^{\prime\prime}_{x,j},\,l^{\prime\prime}_{x}h^{\prime\prime}_{x}\in P_{\lambda_{0}}.

Therefore, Pλ0P_{\lambda_{0}} is compatible with ρ\rho^{\prime}. Also, χλ0\chi_{\lambda_{0}} is a dominant character of Pλ0P_{\lambda_{0}} by Lemma 3.2. Then the formula (\bullet3.8) gives

λ,χ𝜽=d𝜽,χλ00\displaystyle\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{0}}\rangle\geq 0

because ρ\rho^{\prime} is RR-semistable. This finishes the proof for the semistable case. The argument for the stable case is similar.

When GG is reductive, let R(G)R(G) be its radical. By Remark 3.8, a cocharacter λ:𝔾mG×𝑳\lambda:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L}, of which the limit limt0λ(t)ϕ\lim\limits_{t\rightarrow 0}\lambda(t)\cdot\phi exists, is uniquely determined by a cocharacter λ0:𝔾mG\lambda_{0}:\mathbb{G}_{m}\rightarrow G. Moreover, λ0\lambda_{0} is uniquely determined by a cocharacter λss\lambda_{ss} of the semisimple group [G,G][G,G] and a cocharacter λR(G)\lambda_{R(G)} of R(G)R(G). Thus, the formula (\bullet3.8) gives

λ,χ𝜽=d𝜽,χλ0=d(𝜽,χλss+𝜽,χλR(G)).\displaystyle\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{0}}\rangle=-d(\langle\boldsymbol{\theta},\chi_{\lambda_{ss}}\rangle+\langle\boldsymbol{\theta},\chi_{\lambda_{R(G)}}\rangle).

Note that there is a natural injection of characters

Hom(G,𝔾m)Hom(R(G),𝔾m),\displaystyle{\rm Hom}(G,\mathbb{G}_{m})\rightarrow{\rm Hom}(R(G),\mathbb{G}_{m}),

whose image is of finite index. Therefore, Hom(G,𝔾m)Hom(R(G),𝔾m){\rm Hom}(G,\mathbb{G}_{m})\otimes_{\mathbb{Z}}\mathbb{Q}\cong{\rm Hom}(R(G),\mathbb{G}_{m})\otimes_{\mathbb{Z}}\mathbb{Q}. By definition of degree zero, we know that for any character χ\chi of GG, we have 𝜽,χ=0\langle\boldsymbol{\theta},\chi\rangle=0. Then, 𝜽,χλR(G)=0\langle\boldsymbol{\theta},\chi_{\lambda_{R(G)}}\rangle=0. Therefore,

λ,χ𝜽=d𝜽,χλss=λss,χ𝜽,\displaystyle\langle\lambda,\chi_{\boldsymbol{\theta}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{ss}}\rangle=\langle\lambda_{ss},\chi_{\boldsymbol{\theta}}\rangle,

and we reduces it to the semisimple case.

For the second statement about SS-equivalence, we introduce some notations first. Given ρ′′Hom𝕊(Π′′,𝜽)\rho^{\prime\prime}\in{\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}), let λ:𝔾mG×𝑳\lambda:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L} be a cocharacter given by a tuple (λ0,λx,x𝑫)(\lambda_{0},\lambda_{x},x\in\boldsymbol{D}), where λ0:𝔾mG\lambda_{0}:\mathbb{G}_{m}\rightarrow G and λx:𝔾mLθx\lambda_{x}:\mathbb{G}_{m}\rightarrow L_{\theta_{x}} are cocharacters, such that the limit limt0λ(t)ρ′′\lim_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime} exists. Denote by PP the parabolic subgroup given by λ0\lambda_{0} with Levi subgroup LL. We introduce the following notations:

(ai′′)L:=limt0λ0(t)ai′′λ01(t),(bi′′)L:=limt0λ0(t)bi′′λ01(t),(Sx,d′′)L:=limt0λ0(t)Sx,d′′λ01(t),\displaystyle(a^{\prime\prime}_{i})_{L}:=\lim_{t\rightarrow 0}\lambda_{0}(t)a^{\prime\prime}_{i}\lambda^{-1}_{0}(t),\ (b^{\prime\prime}_{i})_{L}:=\lim_{t\rightarrow 0}\lambda_{0}(t)b^{\prime\prime}_{i}\lambda^{-1}_{0}(t),\ (S^{\prime\prime}_{x,d})_{L}:=\lim_{t\rightarrow 0}\lambda_{0}(t)S^{\prime\prime}_{x,d}\lambda^{-1}_{0}(t),
(lx′′)L:=limt0λ0(t)lx′′λx1(t),(hx′′)L:=limt0λx(t)hx′′λ01(t),(ρ′′)L:=limt0λ(t)ρ′′,\displaystyle(l^{\prime\prime}_{x})_{L}:=\lim_{t\rightarrow 0}\lambda_{0}(t)l^{\prime\prime}_{x}\lambda^{-1}_{x}(t),\ (h^{\prime\prime}_{x})_{L}:=\lim_{t\rightarrow 0}\lambda_{x}(t)h^{\prime\prime}_{x}\lambda^{-1}_{0}(t),\ (\rho^{\prime\prime})_{L}:=\lim_{t\rightarrow 0}\lambda(t)\cdot\rho^{\prime\prime},

and clearly,

(ai′′)L,(bi′′)L,(Sx,d′′)L,(lx′′)L(hx′′)LL.\displaystyle(a^{\prime\prime}_{i})_{L},\ (b^{\prime\prime}_{i})_{L},\ (S^{\prime\prime}_{x,d})_{L},\ (l^{\prime\prime}_{x})_{L}(h^{\prime\prime}_{x})_{L}\in L.

Denote by ρ:ΠG\rho^{\prime}:\Pi^{\prime}\rightarrow G the corresponding representation of ρ′′\rho^{\prime\prime} under the morphism ΠΠ′′\Pi^{\prime}\rightarrow\Pi^{\prime\prime}, and then (ρ)L(\rho^{\prime})_{L} is the corresponding representation of (ρ′′)L(\rho^{\prime\prime})_{L}. In the following, we suppose that GG is semisimple, and the reductive case can be reduced to the semisimple case as we discussed above.

Assume that ρ1′′\rho^{\prime\prime}_{1} and ρ2′′\rho^{\prime\prime}_{2} are GIT equivalent. By Lemma 3.7, there exist cocharacters λ1\lambda_{1} and λ2\lambda_{2} of G×𝑳G\times\boldsymbol{L} such that

  • λ1,χ𝜽=λ2,χ𝜽=0\langle\lambda_{1},\chi_{\boldsymbol{\theta}}\rangle=\langle\lambda_{2},\chi_{\boldsymbol{\theta}}\rangle=0,

  • the limits (ρ1′′)L1=limt0λ1(t)ρ1′′(\rho^{\prime\prime}_{1})_{L_{1}}=\lim\limits_{t\rightarrow 0}\lambda_{1}(t)\cdot\rho^{\prime\prime}_{1} and (ρ2′′)L2=limt0λ2(t)ρ2′′(\rho^{\prime\prime}_{2})_{L_{2}}=\lim\limits_{t\rightarrow 0}\lambda_{2}(t)\cdot\rho^{\prime\prime}_{2} exist,

  • (ρ1′′)L1(\rho^{\prime\prime}_{1})_{L_{1}} and (ρ2′′)L2(\rho^{\prime\prime}_{2})_{L_{2}} are in the same (G×𝑳)(G\times\boldsymbol{L})-orbit.

There exists (g,(lx)x𝑫)G×𝑳(g,(l_{x})_{x\in\boldsymbol{D}})\in G\times\boldsymbol{L} such that

(g,(lx)x𝑫)(limt0λ1(t)ρ1′′)=limt0λ2(t)ρ2′′.\displaystyle(g,(l_{x})_{x\in\boldsymbol{D}})\cdot(\lim\limits_{t\rightarrow 0}\lambda_{1}(t)\cdot\rho^{\prime\prime}_{1})=\lim\limits_{t\rightarrow 0}\lambda_{2}(t)\cdot\rho^{\prime\prime}_{2}.

By Lemma 3.6, we have

g(limt0λ10(t)ρ1)=limt0λ20(t)ρ2.\displaystyle g\cdot(\lim\limits_{t\rightarrow 0}\lambda_{10}(t)\cdot\rho^{\prime}_{1})=\lim\limits_{t\rightarrow 0}\lambda_{20}(t)\cdot\rho^{\prime}_{2}.

The cocharacters λ10\lambda_{10} and λ20\lambda_{20} determine parabolic subgroups P1P_{1} and P2P_{2} and Levi subgroups L1L_{1} and L2L_{2} respectively. Clearly,

limt0λi0(t)ρi=(ρi)Li,i=1,2,\displaystyle\lim\limits_{t\rightarrow 0}\lambda_{i0}(t)\cdot\rho^{\prime}_{i}=(\rho^{\prime}_{i})_{L_{i}},\ i=1,2\ ,

and thus

g(ρ1)L1=(ρ2)L2.\displaystyle g\cdot(\rho^{\prime}_{1})_{L_{1}}=(\rho^{\prime}_{2})_{L_{2}}.

Now we have to prove that ρi\rho^{\prime}_{i} is admissible with PiP_{i} based on the condition λi,χ𝜽=d𝜽,χλi0=0\langle\lambda_{i},\chi_{\boldsymbol{\theta}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{i0}}\rangle=0. It is equivalent to show that for any character χi:Pi𝔾m\chi_{i}:P_{i}\rightarrow\mathbb{G}_{m}, we have 𝜽,χi=0\langle\boldsymbol{\theta},\chi_{i}\rangle=0. With the same approach as in [Ram96] (for instance the proof of [Ram96, Lemma 3.5.8]), it is equivalent to choose a faithful embedding GGL(V)G\hookrightarrow{\rm GL}(V) and prove this property for general linear groups. Therefore, the argument can be proved in the same way as [HS23b, Lemma 3.22]. In conclusion, ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} are SS-equivalent.

For the other direction, suppose that ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} are SS-equivalent, and then there exist parabolic subgroups P1P_{1} and P2P_{2} (with Levi subgroups L1L_{1} and L2L_{2}) admissible with ρ1\rho^{\prime}_{1} and ρ2\rho^{\prime}_{2} respectively such that g(ρ1)L1=(ρ2)L2g\cdot(\rho^{\prime}_{1})_{L_{1}}=(\rho^{\prime}_{2})_{L_{2}} for some gGg\in G. Clearly, gP1g1=P2gP_{1}g^{-1}=P_{2}. We choose cocharacters λi0:𝔾mG\lambda_{i0}:\mathbb{G}_{m}\rightarrow G such that Pλi0=PiP_{\lambda_{i0}}=P_{i} for i=1,2i=1,2. We define cocharacters λi:𝔾mG×𝑳\lambda_{i}:\mathbb{G}_{m}\rightarrow G\times\boldsymbol{L} as

λi=(λi0,λix,x𝑫),\displaystyle\lambda_{i}=(\lambda_{i0},\lambda_{ix},x\in\boldsymbol{D}),

where λix:=gix1λi0gix:𝔾mLθx\lambda_{ix}:=g^{-1}_{ix}\lambda_{i0}g_{ix}:\mathbb{G}_{m}\rightarrow L_{\theta_{x}} and gixGg_{ix}\in G is given in Remark 3.8. Since PiP_{i} is compatible with ρi\rho^{\prime}_{i}, we have

λi,χ𝜽=d𝜽,χλi0=0\displaystyle\langle\lambda_{i},\chi_{\boldsymbol{\theta}}\rangle=-d\langle\boldsymbol{\theta},\chi_{\lambda_{i0}}\rangle=0

and the limit

limt0λi(t)ρi′′\displaystyle\lim_{t\rightarrow 0}\lambda_{i}(t)\cdot\rho^{\prime\prime}_{i}

exist for i=1,2i=1,2. By Lemma 3.7, we only have to prove that (ρ1′′)L1(\rho^{\prime\prime}_{1})_{L_{1}} and (ρ2′′)L2(\rho^{\prime\prime}_{2})_{L_{2}} are in the same G𝜽G_{\boldsymbol{\theta}}-orbit. Since g(ρ1)L1=(ρ2)L2g\cdot(\rho^{\prime}_{1})_{L_{1}}=(\rho^{\prime}_{2})_{L_{2}}, the key point is to find lxLθxl_{x}\in L_{\theta_{x}} for each x𝑫x\in\boldsymbol{D} such that

(g,(lx)x𝑫)(ρ1′′)L1=(ρ2′′)L2.\displaystyle(g,(l_{x})_{x\in\boldsymbol{D}})\cdot(\rho^{\prime\prime}_{1})_{L_{1}}=(\rho^{\prime\prime}_{2})_{L_{2}}.

Consider the element

gix1(lix′′)Li=limt0(gix1λi0(t)gix)(gix1lix′′)λix1(t).\displaystyle g^{-1}_{ix}(l^{\prime\prime}_{ix})_{L_{i}}=\lim_{t\rightarrow 0}(g^{-1}_{ix}\lambda_{i0}(t)g_{ix})(g^{-1}_{ix}l^{\prime\prime}_{ix})\lambda^{-1}_{ix}(t).

Since gix1λi0(t)gix=λix(t)g^{-1}_{ix}\lambda_{i0}(t)g_{ix}=\lambda_{ix}(t) and gix1lix′′Lθxg^{-1}_{ix}l^{\prime\prime}_{ix}\in L_{\theta_{x}}, we have gix1(lix′′)LiLθxg^{-1}_{ix}(l^{\prime\prime}_{ix})_{L_{i}}\in L_{\theta_{x}}. With a similar argument, we have (hix′′)LigixLθx(h^{\prime\prime}_{ix})_{L_{i}}g_{ix}\in L_{\theta_{x}}. Note that

g(l1x′′)L1(h1x′′)L1g1=(l2x′′)L2(h2x′′)L2.\displaystyle g(l^{\prime\prime}_{1x})_{L_{1}}(h^{\prime\prime}_{1x})_{L_{1}}g^{-1}=(l^{\prime\prime}_{2x})_{L_{2}}(h^{\prime\prime}_{2x})_{L_{2}}.

Reformulating the equation, we have

g2x1gg1x(g1x1(l1x′′)L1)((h1x′′)L1g1x)g1x1g1g2x=(g2x1(l2x′′)L2)((h2x′′)L2g2x).\displaystyle g^{-1}_{2x}gg_{1x}(g^{-1}_{1x}(l^{\prime\prime}_{1x})_{L_{1}})((h^{\prime\prime}_{1x})_{L_{1}}g_{1x})g^{-1}_{1x}g^{-1}g_{2x}=(g^{-1}_{2x}(l^{\prime\prime}_{2x})_{L_{2}})((h^{\prime\prime}_{2x})_{L_{2}}g_{2x}).

Therefore, g2x1gg1xLθxg^{-1}_{2x}gg_{1x}\in L_{\theta_{x}} because the normalizer of LθxL_{\theta_{x}} is itself. We define

lx:=((l2x′′)L2)1g(l1x′′)L1\displaystyle l_{x}:=((l^{\prime\prime}_{2x})_{L_{2}})^{-1}g(l^{\prime\prime}_{1x})_{L_{1}}

Clearly,

lx=(g2x1(l2x′′)L2)1(g2x1gg1x)(g1x1(l1x′′)L1)Lθx.\displaystyle l_{x}=(g^{-1}_{2x}(l^{\prime\prime}_{2x})_{L_{2}})^{-1}(g^{-1}_{2x}gg_{1x})(g^{-1}_{1x}(l^{\prime\prime}_{1x})_{L_{1}})\in L_{\theta_{x}}.

Since

lx=((l2x′′)L2)1g(l1x′′)L1=(h2x′′)L2g((h1x′′)L1)1,\displaystyle l_{x}=((l^{\prime\prime}_{2x})_{L_{2}})^{-1}g(l^{\prime\prime}_{1x})_{L_{1}}=(h^{\prime\prime}_{2x})_{L_{2}}g((h^{\prime\prime}_{1x})_{L_{1}})^{-1},

it is easy to check

(g,(lx)x𝑫)(ρ1′′)L1=(ρ2′′)L2.\displaystyle(g,(l_{x})_{x\in\boldsymbol{D}})\cdot(\rho^{\prime\prime}_{1})_{L_{1}}=(\rho^{\prime\prime}_{2})_{L_{2}}.

Therefore, (ρ1′′)L1(\rho^{\prime\prime}_{1})_{L_{1}} and (ρ2′′)L2(\rho^{\prime\prime}_{2})_{L_{2}} are in the same G𝜽G_{\boldsymbol{\theta}}-orbit. ∎

Under the equivalence of Stokes GG-representations and Stokes GG-local systems (Corollary 2.19), we obtain the moduli space of filtered Stokes GG-local systems.

Theorem 3.10.

The quasi-projective variety

B(X𝑫,G,𝑸,𝜽):=Hom𝕊(Π′′,𝜽)//(G×𝑳,χ𝜽)\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta}):={\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta})/\!\!/(G\times\boldsymbol{L},\chi_{\boldsymbol{\theta}})

is the moduli space of degree zero RR-semistable 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}}, of which points are in one-to-one correspondence with SS-equivalence classes of degree zero RR-semistable 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q}. There exists an open subset Bs(X𝐃,G,𝐐,𝛉)\mathcal{M}^{s}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta}), of which points correspond to isomorphism classes of degree zero RR-stable 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q}.

Proof.

By Proposition 3.9, the theorem follows directly from King’s result [Kin94, §2]. ∎

4. Examples

We study some examples of the moduli space of filtered Stokes GG-local systems in this section. In §4.1, we consider the case of trivial weights. We conclude that the RR-stability of filtered Stokes GG-representations (with trivial weights) is equivalent to the irreducibility of the corresponding representations (Lemma 4.1) and the moduli space is an affine variety, which is known as the wild character variety [Boa14, DDP18]. In §4.2, we consider the case of trivial irregular types. If all irregular types are trivial, then the discussion completely reduces to the tame case, of which the moduli space has been constructed in [HS23b]. In §4.3, we consider the Eguchi–Hanson space, of which the irregular type is unramified. In this case, we find a particular θ\theta-filtered Stokes GG-local system, which is RR-stable but not semisimple as a representation. This example shows that wild character varieties may not be the Betti moduli space in the nonabelian Hodge correspondence. In §4.4 and §4.5, we start from the Airy equation and study Stokes SL2(){\rm SL}_{2}(\mathbb{C})-local systems with ramified irregular type on X𝑫X_{\boldsymbol{D}}, where (X,𝑫)=(1,0)(X,\boldsymbol{D})=(\mathbb{P}^{1},0). In this case, we find that Stokes SL2(){\rm SL}_{2}(\mathbb{C})-representations are always irreducible. Therefore, the corresponding moduli space is exactly the wild character variety by Corollary 4.2.

4.1. Trivial weights

Fixing a collection of irregular types 𝑸\boldsymbol{Q}, let Π\Pi^{\prime} be the group defined in §2.3. Suppose that all weights are trivial, i.e. θx=0\theta_{x}=0 for x𝑫x\in\boldsymbol{D}. We use the notation 𝟎\boldsymbol{0} for the collection of trivial weights. In this case, Lθx=GL_{\theta_{x}}=G. Then,

G×𝑳=G×x𝑫G.\displaystyle G\times\boldsymbol{L}=G\times\prod_{x\in\boldsymbol{D}}G.

Moreover, the character χdθx\chi_{-d\theta_{x}} is also trivial, which implies that the character χ𝟎:G×𝑳𝔾m\chi_{\boldsymbol{0}}:G\times\boldsymbol{L}\rightarrow\mathbb{G}_{m} is trivial.

Lemma 4.1.

A degree zero 𝟎\boldsymbol{0}-filtered Stokes GG-representation ρ:ΠG\rho^{\prime}:\Pi^{\prime}\rightarrow G is RR-stable (resp. RR-semistable) if and only if it is an irreducible (resp. semisimple) representation.

Proof.

Given an irreducible representation ρ:ΠG\rho^{\prime}:\Pi^{\prime}\rightarrow G, it cannot be restricted to any proper nontrivial parabolic subgroup PP. Then it is RR-stable automatically. On the other hand, given a degree zero RR-stable 𝟎\boldsymbol{0}-filtered Stokes GG-representation ρ\rho^{\prime}, suppose that a nontrivial proper parabolic subgroup PP is compatible with ρ\rho^{\prime}. We choose an arbitrary nontrivial anti-dominant character χ:P𝔾m\chi:P\rightarrow\mathbb{G}_{m}, which is trivial on the center of PP. We have

deglocρ(P,χ)=𝟎,χ=0\displaystyle\deg^{\rm loc}\rho^{\prime}(P,\chi)=\langle\boldsymbol{0},\chi\rangle=0

because θx\theta_{x} is trivial for every x𝑫x\in\boldsymbol{D}. This contradicts the assumption that ρ\rho^{\prime} is RR-stable. Therefore, any nontrivial proper parabolic subgroup is not compatible with ρ\rho^{\prime}, which means that ρ\rho^{\prime} is irreducible. The proof for semistable case is similar. ∎

Corollary 4.2.

The moduli space

B(X𝑫,G,𝑸,𝟎)=Hom𝕊(Π′′,𝟎)//(G×𝑳,χ𝟎)\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{0})={\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{0})/\!\!/(G\times\boldsymbol{L},\chi_{\boldsymbol{0}})

is an affine variety, of which points correspond to isomorphism classes of semisimple Stokes GG-representations. There exists an open subset Bs(X𝐃,G,𝐐,𝟎)\mathcal{M}^{s}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{0}), of which points correspond to isomorphism classes of irreducible Stokes GG-representations.

Proof.

Since all weights θx\theta_{x} are trivial, Hom𝕊(Π′′,𝟎)=Hom𝕊(Π′′,G){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{0})={\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},G) is an affine variety. Since the character χ𝟎\chi_{\boldsymbol{0}} is trivial, the GIT quotient Hom𝕊(Π′′,𝟎)//(G×𝑳,χ𝟎){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{0})/\!\!/(G\times\boldsymbol{L},\chi_{\boldsymbol{0}}) is an affine variety. By Theorem 3.10, points in B(X,G,𝑸,𝟎)\mathcal{M}_{\rm B}(X,G,\boldsymbol{Q},\boldsymbol{0}) are in one-to-one correspondence with SS-equivalence classes of degree zero RR-semistable 𝟎\boldsymbol{0}-filtered Stokes GG-representations, which are exactly isomorphism classes of semisimple representations by Lemma 4.1. This finishes the proof of this corollary. ∎

The moduli space B(X,G,𝑸,𝟎)\mathcal{M}_{\rm B}(X,G,\boldsymbol{Q},\boldsymbol{0}) given above is exactly the wild character variety considered by Boalch [Boa14, §8 and §9]. The difference is that Boalch constructed the moduli space from the fundamental groupoid Π\Pi, while we construct the moduli space from Π\Pi^{\prime} and Π′′\Pi^{\prime\prime}. Moreover, we also refer the reader to [Boa14, Theorem 9.3] for another proof that ρHom𝕊(Π,G)\rho\in{\rm Hom}_{\mathbb{S}}(\Pi,G) is stable (in the sense of GIT) if and only if it is irreducible.

4.2. Trivial Irregular Types

Suppose that all irregular types are trivial, i.e. Qx=0Q_{x}=0 for any x𝑫x\in\boldsymbol{D}, and we use the notation 𝟎\boldsymbol{0} for the collection of trivial irregular types. In this case, the Betti moduli space in the wild case completely reduce to the tame case considered in [HS23b].

Recall that the generators of the fundamental groupoid Π\Pi of X𝑸X_{\boldsymbol{Q}} introduced in §2.2 are given by

  1. (1)

    αi,βi\alpha_{i},\beta_{i}, 1ig1\leq i\leq g;

  2. (2)

    γx\gamma_{x} for x𝑫x\in\boldsymbol{D};

  3. (3)

    γx,d\gamma_{x,d} for x𝑫x\in\boldsymbol{D} and d𝔸xd\in\mathbb{A}_{x};

  4. (4)

    γ0x\gamma_{0x} for x𝑫x\in\boldsymbol{D}.

Since all irregular types are trivial, there is no anti-Stokes directions and then the set 𝔸x\mathbb{A}_{x} is empty for every x𝑫x\in\boldsymbol{D}. Therefore, when irregular types are trivial, the group Π\Pi is generated by αi,βi,γx,γ0x\alpha_{i},\beta_{i},\gamma_{x},\gamma_{0x} with the relation

(i=1g[αi,βi])(x𝑫μx)=id,\displaystyle\left(\prod_{i=1}^{g}[\alpha_{i},\beta_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu_{x}\right)={\rm id},

where μx=γ0x1γxγ0x\mu_{x}=\gamma^{-1}_{0x}\cdot\gamma_{x}\cdot\gamma_{0x}. Following the same discussion, the group Π\Pi^{\prime} introduced in §2.3 is generated by

  1. (1)

    αi,βi\alpha^{\prime}_{i},\beta^{\prime}_{i} for 1ig1\leq i\leq g;

  2. (2)

    γx\gamma^{\prime}_{x} for x𝑫x\in\boldsymbol{D},

with the relation

(i=1g[αi,βi])(x𝑫μx)=id,\displaystyle\left(\prod_{i=1}^{g}[\alpha^{\prime}_{i},\beta^{\prime}_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu^{\prime}_{x}\right)={\rm id},

where μx=γx\mu^{\prime}_{x}=\gamma^{\prime}_{x}. Clearly, Π\Pi^{\prime} can be regarded as the fundamental group of X𝑫X_{\boldsymbol{D}} and the formal monodromy reduces to the topological monodromy. By Proposition 2.18, isomorphism classes of Stokes GG-local systems with irregular type 𝟎\boldsymbol{0} are in one-to-one correspondence with GG-orbits in Hom(Π,G){\rm Hom}(\Pi^{\prime},G), which implies that Stokes GG-local systems with irregular type 𝟎\boldsymbol{0} on X𝑫X_{\boldsymbol{D}} are exactly GG-local systems on X𝑫X_{\boldsymbol{D}}. For the construction of the moduli space, we introduce the third group Π′′\Pi^{\prime\prime} in §3.2, which is generated by

  1. (1)

    αi′′,βi′′\alpha^{\prime\prime}_{i},\beta^{\prime\prime}_{i} for 1ig1\leq i\leq g;

  2. (2)

    ιx′′,γx′′\iota^{\prime\prime}_{x},\gamma^{\prime\prime}_{x} for x𝑫x\in\boldsymbol{D}

with the relation

(i=1g[αi′′,βi′′])(x𝑫μx′′)=id,\displaystyle\left(\prod_{i=1}^{g}[\alpha^{\prime\prime}_{i},\beta^{\prime\prime}_{i}]\right)\cdot\left(\prod_{x\in\boldsymbol{D}}\mu^{\prime\prime}_{x}\right)={\rm id},

where μx′′=ιx′′γx′′\mu^{\prime\prime}_{x}=\iota^{\prime\prime}_{x}\gamma^{\prime\prime}_{x}. Fixing a collection of weights 𝜽={θx,x𝑫}\boldsymbol{\theta}=\{\theta_{x},x\in\boldsymbol{D}\}, the variety Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}) now parametrizes points

((ai′′,bi′′)1ig,(lx′′,hx′′)x𝑫)\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x})_{x\in\boldsymbol{D}})

such that for each x𝑫x\in\boldsymbol{D}, there exists gxGg_{x}\in G such that gx1lx′′Lθxg^{-1}_{x}l^{\prime\prime}_{x}\in L_{\theta_{x}} and hx′′gxPθxh^{\prime\prime}_{x}g_{x}\in P_{\theta_{x}}. Moreover, there is a natural (G×𝑳)(G\times\boldsymbol{L})-action on Hom𝕊(Π′′,𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta}). This construction is exactly the same as [HS23b, Construction 3.10], and it is easy to check that the stability condition in this special case is equivalent to that in [HS23b]. Therefore, the moduli space B(X,G,𝟎,𝜽)\mathcal{M}_{\rm B}(X,G,\boldsymbol{0},\boldsymbol{\theta}) of degree zero RR-semistable 𝜽\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝟎\boldsymbol{0} on X𝑫X_{\boldsymbol{D}} is exactly the moduli space B(X𝑫,G,𝜽)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{\theta}) of degree zero RR-semistable 𝜽\boldsymbol{\theta}-filtered GG-local systems on X𝑫X_{\boldsymbol{D}} [HS23b, Theorem 1.2].

4.3. “Weighted” Eguchi–Hanson Space

We consider an explicit example such that G=SL2()G=\mathrm{SL}_{2}(\mathbb{C}) and (X,𝑫)=(1,0)(X,\boldsymbol{D})=(\mathbb{P}^{1},0). Denote by α\alpha and α-\alpha the roots of SL2(){\rm SL}_{2}(\mathbb{C}), and let

U+:=Uα=(101),U:=Uα=(101).\displaystyle U_{+}:=U_{\alpha}=\begin{pmatrix}1&\ast\\ 0&1\end{pmatrix},\ U_{-}:=U_{-\alpha}=\begin{pmatrix}1&0\\ \ast&1\end{pmatrix}.

Given an irregular type with a pole of order 33 at z=0z=0

Q3=A3z3+A2z2+A1z,\displaystyle Q_{-3}=\frac{A_{3}}{z^{3}}+\frac{A_{2}}{z^{2}}+\frac{A_{1}}{z},

where the subscript of the irregular type is for its degree, the leading coefficient A3A_{3} is nontrivial. Since G=SL2()G={\rm SL}_{2}(\mathbb{C}), the leading coefficient A3A_{3} is automatically regular and semisimple. In this case, the irregular type Q3Q_{-3} has 66 anti-Stokes directions and

𝕊to(Q3)=(U+×U)3.\displaystyle\mathbb{S}{\rm to}(Q_{-3})=(U_{+}\times U_{-})^{3}.

Since this irregular type Q3Q_{-3} is in the unramified case, the centralizer of Q3Q_{-3} coincides with the set of formal monodromies given by Q3Q_{-3}, i.e.

H=H()={(aa1)|a}.\displaystyle H=H(\partial)=\left\{\begin{pmatrix}a&\\ &a^{-1}\end{pmatrix}\ \Big{|}\ a\in\mathbb{C}^{*}\right\}.

Therefore, the space Hom𝕊(Π1(XQ3),G)G×H×(U+×U)3{\rm Hom}_{\mathbb{S}}(\Pi_{1}(X_{Q_{-3}}),G)\subseteq G\times H\times(U_{+}\times U_{-})^{3} is a closed subvariety including

(c,h,(u+,i,u,i)1i3)G×H×(U+×U)3\displaystyle(c,h,(u_{+,i},u_{-,i})_{1\leq i\leq 3})\in G\times H\times(U_{+}\times U_{-})^{3}

such that c1(hi=13(u+,iu,i))c=idc^{-1}(h\prod_{i=1}^{3}(u_{+,i}u_{-,i}))c={\rm id}.

Given the weight

θ=(1212)𝔱,\displaystyle\theta=\begin{pmatrix}\frac{1}{2}&\\ &-\frac{1}{2}\end{pmatrix}\in\mathfrak{t},

it determines a parabolic subgroup of SL2()\mathrm{SL}_{2}(\mathbb{C}) as

Pθ={(ab0a1)|a,b}.\displaystyle P_{\theta}=\left\{\begin{pmatrix}a&b\\ 0&a^{-1}\end{pmatrix}\ |\ a\in\mathbb{C}^{*},b\in\mathbb{C}\right\}.

We consider a special θ\theta-filtered Stokes GG-representation

ρ=(c,h,(u+,i,u,i)1i3)Hom𝕊(Π1(XQ3),G)\displaystyle\rho=(c,h,(u_{+,i},u_{-,i})_{1\leq i\leq 3})\in{\rm Hom}_{\mathbb{S}}(\Pi_{1}(X_{Q_{-3}}),G)

such that c,h,u+,i,u,iPθc,h,u_{+,i},u_{-,i}\in P_{\theta} and at least one of them is not included in PθP_{-\theta}. For example,

u+,1=(1101),u+,2=(1101)\displaystyle u_{+,1}=\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\ u_{+,2}=\begin{pmatrix}1&-1\\ 0&1\end{pmatrix}

and the other elements are identity matrix. Clearly, if we take P=PθP=P_{\theta}, the parabolic subgroup PP is compatible with ρ\rho. Furthermore, taking any nontrivial anti-dominant character χ\chi of PP, it is easy to check

deglocρ(P,χ)=θ,χ>0,\displaystyle\deg^{\rm loc}\rho(P,\chi)=\langle\theta,\chi\rangle>0,

and the θ\theta-filtered Stokes GG-representation ρ\rho is RR-stable. By Theorem 3.10, it corresponds to a point in Bs(X𝑫,G,Q3,θ)\mathcal{M}^{s}_{\rm B}(X_{\boldsymbol{D}},G,Q_{-3},\theta).

On the other hand, this representation ρ\rho is indecomposable but not semisimple. As we discussed in §4.1, the classical wild character variety B(X𝑫,G,Q,0)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,Q,0) only parametrizes semisimple representations. Therefore, the 𝜽\boldsymbol{\theta}-filtered Stokes GG-representation ρ\rho does not correspond to a point in Bs(X𝑫,G,Q3,0)\mathcal{M}^{s}_{\rm B}(X_{\boldsymbol{D}},G,Q_{-3},0). This example illustrates the fact that

B(X𝑫,SL2(),Q3,0)B(X𝑫,SL2(),Q3,θ)\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-3},0)\ncong\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-3},\theta)

for general weight θ\theta. Moreover, the moduli space B(X𝑫,G,Q3,0)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,Q_{-3},0) is exactly the Eguchi–Hanson space considered in [Boa18, §2], where he concluded that this is a smooth affine variety of dimension 22.

4.4. Airy Equation

We start from the Airy equation

y′′(t)=ty(t).\displaystyle y^{\prime\prime}(t)=ty(t).

This equation corresponds to the following connection

=d+(0t10)dt.\displaystyle\nabla=d+\begin{pmatrix}0&-t\\ -1&0\end{pmatrix}dt.

We change the coordinate z=1tz=\frac{1}{t}

=d+(0z3z20)dz.\displaystyle\nabla=d+\begin{pmatrix}0&z^{-3}\\ z^{-2}&0\end{pmatrix}dz.

Clearly, \nabla has an irregular singularity at z=0z=0 (or t=t=\infty). Thus, the Airy equation corresponds to a connection on 1\{0}\mathbb{P}^{1}\backslash\{0\} with an irregular singularity at z=0z=0, at which the topological monodromy is trivial [vdPS03, Example 8.15]. We continue working locally on the local coordinate zz. Let

A(z)=(0z3z20)=(0100)z3+(0010)z2\displaystyle A(z)=\begin{pmatrix}0&z^{-3}\\ z^{-2}&0\end{pmatrix}=\begin{pmatrix}0&1\\ 0&0\end{pmatrix}z^{-3}+\begin{pmatrix}0&0\\ 1&0\end{pmatrix}z^{-2}

be the connection form of \nabla. We take (on a ramified cover)

g1=(z1/4z1/4)\displaystyle g_{1}=\begin{pmatrix}z^{1/4}&\\ &z^{-1/4}\end{pmatrix}

and get

g1=ddg1g11+(g1A(z)g11)dz=d+(0z5/2z5/20)dz14(1001)dzz\displaystyle\begin{aligned} g_{1}\circ\nabla&=d-dg_{1}\cdot g_{1}^{-1}+(g_{1}A(z)g_{1}^{-1})dz\\ &=d+\begin{pmatrix}0&z^{-5/2}\\ z^{-5/2}&0\end{pmatrix}dz-\frac{1}{4}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\frac{dz}{z}\end{aligned}

Then we take

g2=(12121212),\displaystyle g_{2}=\begin{pmatrix}\frac{1}{\sqrt{2}}&-\frac{1}{\sqrt{2}}\\ \\ \frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\end{pmatrix},

and obtain

g2(g1)=d+(z5/200z5/2)dz14(0110)dzz=d+d((23z3/20023z3/2))14(0110)dzz.\displaystyle\begin{aligned} g_{2}\circ(g_{1}\circ\nabla)&=d+\begin{pmatrix}-z^{-5/2}&0\\ 0&z^{-5/2}\end{pmatrix}dz-\frac{1}{4}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\frac{dz}{z}\\ &=d+d\Big{(}\begin{pmatrix}\frac{2}{3}z^{-3/2}&0\\ 0&-\frac{2}{3}z^{-3/2}\end{pmatrix}\Big{)}-\frac{1}{4}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\frac{dz}{z}.\end{aligned}

Therefore, the corresponding connection of the Airy equation is of irregular type

Q32=(230023)z32,\displaystyle Q_{-\frac{3}{2}}=\begin{pmatrix}\frac{2}{3}&0\\ 0&-\frac{2}{3}\end{pmatrix}z^{-\frac{3}{2}},

which is in the ramified case. In conclusion, the corresponding connection of the Airy equation is a SL2(){\rm SL}_{2}(\mathbb{C})-connection with irregular type Q32Q_{-\frac{3}{2}} on X𝑫X_{\boldsymbol{D}}, where (X,𝑫)=(1,0)(X,\boldsymbol{D})=(\mathbb{P}^{1},0), and thus corresponds to a Stokes SL2(){\rm SL}_{2}(\mathbb{C})-representation (local system) with irregular type Q32Q_{-\frac{3}{2}} on X𝑫X_{\boldsymbol{D}}.

4.5. Stokes SL2(){\rm SL}_{2}(\mathbb{C})-local Systems with Ramified Irregular Types

We start from the setup

G=SL2(),(X,𝑫)=(1,0),Q32=(230023)z32,\displaystyle G={\rm SL}_{2}(\mathbb{C}),\ (X,\boldsymbol{D})=(\mathbb{P}^{1},0),\ Q_{-\frac{3}{2}}=\begin{pmatrix}\frac{2}{3}&0\\ 0&-\frac{2}{3}\end{pmatrix}z^{-\frac{3}{2}},

which is the same as §4.4, and consider Stokes SL2(){\rm SL}_{2}(\mathbb{C})-representations with irregular type Q32Q_{-\frac{3}{2}} on X𝑫X_{\boldsymbol{D}}. Following the same notation as in §4.3, we have

𝕊to(Q32)=U×U+×U\displaystyle\mathbb{S}{\rm to}(Q_{-\frac{3}{2}})=U_{-}\times U_{+}\times U_{-}

and

H={(aa1)|a},H()={(aa1)|a}.\displaystyle H=\left\{\begin{pmatrix}a&\\ &a^{-1}\end{pmatrix}\ \Big{|}\ a\in\mathbb{C}^{*}\right\},\ H(\partial)=\left\{\begin{pmatrix}&a\\ -a^{-1}&\end{pmatrix}\ \Big{|}\ a\in\mathbb{C}^{*}\right\}.

A point (c,h,u,1,u+,u,2)G×H()×U×U+×U(c,h,u_{-,1},u_{+},u_{-,2})\in G\times H(\partial)\times U_{-}\times U_{+}\times U_{-} corresponds to a Stokes GG-representation in Hom𝕊(Π1(XQ32),G){\rm Hom}_{\mathbb{S}}(\Pi_{1}(X_{Q_{-\frac{3}{2}}}),G) if it satisfies the condition that c1hu,1u+u,2c=idc^{-1}hu_{-,1}u_{+}u_{-,2}c={\rm id}. Ignoring cc, we consider the equation

hu,1u+u,2=id.\displaystyle hu_{-,1}u_{+}u_{-,2}={\rm id}.

Let

h=(0aa10),u+,1=(10b1),u=(1c01),u+,2=(10d1).\displaystyle h=\begin{pmatrix}0&a\\ -a^{-1}&0\end{pmatrix},\ u_{+,1}=\begin{pmatrix}1&0\\ b&1\end{pmatrix},\ u_{-}=\begin{pmatrix}1&c\\ 0&1\end{pmatrix},\ u_{+,2}=\begin{pmatrix}1&0\\ d&1\end{pmatrix}.

We have

(0aa10)=(10b1)(1c01)(10d1).\displaystyle\begin{pmatrix}0&-a\\ a^{-1}&0\end{pmatrix}=\begin{pmatrix}1&0\\ b&1\end{pmatrix}\begin{pmatrix}1&c\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ d&1\end{pmatrix}.

Therefore,

b=a1,c=a,d=a1.\displaystyle b=a^{-1},\ c=-a,\ d=a^{-1}.

This means that the formal monodromy uniquely determines Stokes data in this case. Therefore, the moduli space

B(X𝑫,SL2(),Q32,0)pt\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-\frac{3}{2}},0)\cong{\rm pt}

is a point. Note that whatever the formal monodromy is, the corresponding representation is irreducible. By the discussion in §4.1, we have

B(X𝑫,SL2(),Q32,θ)B(X𝑫,SL2(),Q32,0)pt\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-\frac{3}{2}},\theta)\cong\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-\frac{3}{2}},0)\cong{\rm pt}

for any weight θ\theta. Moreover, the Stokes SL2(){\rm SL}_{2}(\mathbb{C})-representation corresponding to the Airy equation is the unique point in the moduli space B(X𝑫,SL2(),Q32,θ)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q_{-\frac{3}{2}},\theta) as we discussed in §4.4. As a result, the connection corresponds to the Airy equation is both rigid and physically rigid.

Remark 4.3.

Recently, we noticed that Hohl and Jakob studied physical rigidity of Kloosterman connections and applied the result to a GG-version Airy equation and get the same result for rigidity [HJ24, Theorem 1.2.1].

Now we suppose that QQ is a ramified irregular type. Since we work on SL2(){\rm SL}_{2}(\mathbb{C}), the leading coefficient of QQ is regular and semisimple. Denote by 𝔸\mathbb{A} the set of anti-Stokes directions of QQ, and for each d𝔸d\in\mathbb{A}, 𝕊tod\mathbb{S}{\rm to}_{d} is a product of U+U_{+} and UU_{-}. Let 𝕊to(Q):=d𝔸𝕊tod\mathbb{S}{\rm to}(Q):=\prod_{d\in\mathbb{A}}\mathbb{S}{\rm to}_{d}. At the same time, we have

H={(aa1)|a},H()={(aa1)|a}.\displaystyle H=\left\{\begin{pmatrix}a&\\ &a^{-1}\end{pmatrix}\ \Big{|}\ a\in\mathbb{C}^{*}\right\},\ H(\partial)=\left\{\begin{pmatrix}&a\\ -a^{-1}&\end{pmatrix}\ \Big{|}\ a\in\mathbb{C}^{*}\right\}.

A Stokes SL2(){\rm SL}_{2}(\mathbb{C})-representation with irregular type QQ can be regarded as a tuple (c,h,(Sd)d𝔸)(c,h,(S_{d})_{d\in\mathbb{A}}) such that

(\star) c1h(d𝔸Sd)c=id.c^{-1}h(\prod_{d\in\mathbb{A}}S_{d})c={\rm id}.

Since the product d𝔸Sd\prod_{d\in\mathbb{A}}S_{d} contains both (nontrivial) upper and lower triangular matrices, any representation corresponding a tuple (c,h,(Sd)d𝔸)(c,h,(S_{d})_{d\in\mathbb{A}}) satisfying the relation (\star4.5) is irreducible. Then we have the following proposition:

Proposition 4.4.

Let G=SL2()G={\rm SL}_{2}(\mathbb{C}). Given a ramified irregular type QQ, we have

B(X𝑫,SL2(),Q,θ)B(X𝑫,SL2(),Q,0)\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q,\theta)\cong\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q,0)

for any weight θ\theta.

Remark 4.5.

In the case of SLn(){\rm SL}_{n}(\mathbb{C}), if the leading coefficient of a ramified irregular type QQ is regular and semisimple, the same consequence holds as Proposition 4.4, i.e.

B(X𝑫,SLn(),Q,θ)B(X𝑫,SL2(),Q,0)\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{n}(\mathbb{C}),Q,\theta)\cong\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},{\rm SL}_{2}(\mathbb{C}),Q,0)

for any weight θ\theta.

5. Relation to Wild Nonabelian Hodge Correspondence on Curves

The authors established the wild nonabelian Hodge correspondence for principal bundles on curves at the level of categories. Based on the existence of the corresponding moduli spaces, the Dolbeault moduli space admits a hyperKälher structure. Since the wild nonabelian Hodge correspondence for principal bundles is only given when irregular types are unramfied, we always assume that the irregular types considered in this section are unramified.

5.1. Betti Moduli Spaces

We follow the construction given in §3.2 and define a map

Hom𝕊(Ω′′,𝑷)x𝑫Pθx\displaystyle{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})\rightarrow\prod_{x\in\boldsymbol{D}}P_{-\theta_{x}}

as

((ai′′,bi′′)1ig,(lx′′,hx′′,Sx,d′′)x𝑫,d𝔸x)(lx′′hx′′)x𝑫.\displaystyle((a^{\prime\prime}_{i},b^{\prime\prime}_{i})_{1\leq i\leq g},(l^{\prime\prime}_{x},h^{\prime\prime}_{x},S^{\prime\prime}_{x,d})_{x\in\boldsymbol{D},d\in\mathbb{A}_{x}})\rightarrow(l^{\prime\prime}_{x}h^{\prime\prime}_{x})_{x\in\boldsymbol{D}}.

We fix a collection of elements M𝜽={Mθx,x𝑫}M_{\boldsymbol{\theta}}=\{M_{\theta_{x}},x\in\boldsymbol{D}\} in the Levi subgroups, i.e. MθxLθxM_{\theta_{x}}\in L_{\theta_{x}}, which is regarded as an element in x𝑫Lθx\prod_{x\in\boldsymbol{D}}L_{\theta_{x}}. Consider the following composition of maps

Hom~𝕊(Ω′′,𝑷)Hom𝕊(Ω′′,𝑷)x𝑫Pθxx𝑫Lθx.\displaystyle\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})\rightarrow{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P})\rightarrow\prod_{x\in\boldsymbol{D}}P_{-\theta_{x}}\rightarrow\prod_{x\in\boldsymbol{D}}L_{\theta_{x}}.

Denote by Hom~𝕊(Ω′′,𝑷,M𝜽)Hom~𝕊(Ω′′,𝑷)\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P},M_{\boldsymbol{\theta}})\subseteq\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P}) the preimage of M𝜽x𝑫LθxM_{\boldsymbol{\theta}}\in\prod_{x\in\boldsymbol{D}}L_{\theta_{x}}. Taking the restriction

Hom𝕊(Ω′′,𝜽,M𝜽):=Hom~𝕊(Ω′′,𝑷,M𝜽)|Hom𝕊(Ω′′,G)\displaystyle{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{\theta},M_{\boldsymbol{\theta}}):=\widetilde{{\rm Hom}}_{\mathbb{S}}(\Omega^{\prime\prime},\boldsymbol{P},M_{\boldsymbol{\theta}})|_{{\rm Hom}_{\mathbb{S}}(\Omega^{\prime\prime},G)}

and adding the relation (′′\ast^{\prime\prime}3.2), we obtain a variety Hom𝕊(Π′′,𝜽,M𝜽){\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta},M_{\boldsymbol{\theta}}) together with an induced (G×𝑳)(G\times\boldsymbol{L})-action. We define

B(X𝑫,G,𝑸,𝜽,M𝜽):=Hom𝕊(Π′′,𝜽,M𝜽)//(G×𝑳,χ𝜽).\displaystyle\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta},M_{\boldsymbol{\theta}}):={\rm Hom}_{\mathbb{S}}(\Pi^{\prime\prime},\boldsymbol{\theta},M_{\boldsymbol{\theta}})/\!\!/(G\times\boldsymbol{L},\chi_{\boldsymbol{\theta}}).

As a direct result of Theorem 3.10, we obtain the Betti moduli space considered in the (unramified) wild nonabelian Hodge correspondence on noncompact curves [HS23a].

Corollary 5.1.

There exists a quasi-projective variety B(X𝐃,G,𝐐,𝛉,M𝛉)\mathcal{M}_{\rm B}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\theta},M_{\boldsymbol{\theta}}) as the moduli space of degree zero RR-semistable 𝛉\boldsymbol{\theta}-filtered Stokes GG-local systems with irregular type 𝐐\boldsymbol{Q} on X𝐃X_{\boldsymbol{D}} such that the Levi factors of formal monodromies around punctures are given by M𝛉M_{\boldsymbol{\theta}} (up to conjugation).

5.2. De Rham and Dolbeault Moduli Spaces

Fixing a collection of weights 𝜽\boldsymbol{\theta}, let 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}} be the corresponding parahoric group scheme on XX (see [BS15, §2] or [KSZ23, §2.1] for instance). We fix a collection of unramified irregular types 𝑸\boldsymbol{Q}. The moduli space Higgs(X,𝒢𝜽,𝑸)\mathcal{M}_{\rm Higgs}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q}) (resp. Conn(X,𝒢𝜽,𝑸)\mathcal{M}_{\rm Conn}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q})) of degree zero RR-semistable merohoric (= meromorphic and parahoric) 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors with irregular type 𝑸\boldsymbol{Q} (resp. degree zero RR-semistable meromorphic parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-connections with irregular type 𝑸\boldsymbol{Q}) exists. The approach of constructing these two moduli spaces is the same as that for logahoric (= logarithmic and parahoric) 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors [KSZ23, §6], where the authors constructed the moduli spaces of logahoric Higgs torsors via an equivalence between logarithmic Higgs torsors on XX and equivariant GG-Higgs bundles on an appropriate cover. We also want to remind the reader that it needs a more careful discussion for the corresponding Dolbeault and de Rham moduli spaces when irregular types are ramified because there are two covers needed to be taken into consideration: one is for weights and the other is for the irregular types.

Given the existence for the moduli spaces of merohoric Higgs torsors and merohoric connections, we define the Dolbeault and de Rham residue morphisms in order to obtain the desired Dolbeault and de Rham moduli spaces in the wild nonabelian Hodge correspondence [HS23a, Theorem in §1], and a similar construction in the parabolic case is given in [BGPMiR20, §7]. We take the Dolbeault side as an example. We define the following map

μDol:Higgs(X,𝒢𝜽,𝑸)\displaystyle\mu_{\mathrm{Dol}}:\mathcal{M}_{\mathrm{Higgs}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q}) x𝑫(𝔥x𝔩θx)/Lθx\displaystyle\longrightarrow\prod_{x\in\boldsymbol{D}}(\mathfrak{h}_{x}\cap\mathfrak{l}_{\theta_{x}})/L_{\theta_{x}}
(,φ)\displaystyle(\mathcal{E},\varphi) x𝑫Lθx[Resx(φ)],\displaystyle\longmapsto\prod_{x\in\boldsymbol{D}}L_{\theta_{x}}\cdot[\mathrm{Res}_{x}(\varphi)],

where 𝔥x\mathfrak{h}_{x} is the Lie algebra of the stabilizer HxH_{x} and 𝔩θx\mathfrak{l}_{\theta_{x}} is the Lie algebra of the Levi subgroup LθxL_{\theta_{x}}. Given a L𝜽L_{\boldsymbol{\theta}}-orbit 𝓞={𝒪x,x𝑫}x𝑫(𝔥x𝔩θx)/Lθx\boldsymbol{\mathcal{O}}=\{\mathcal{O}_{x},x\in\boldsymbol{D}\}\subseteq\prod_{x\in\boldsymbol{D}}(\mathfrak{h}_{x}\cap\mathfrak{l}_{\theta_{x}})/L_{\theta_{x}}, the fiber

Dol(X,𝒢𝜽,𝑸,𝓞):=μDol1(𝓞)\displaystyle\mathcal{M}_{\mathrm{Dol}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q},\boldsymbol{\mathcal{O}}):=\mu_{\mathrm{Dol}}^{-1}(\boldsymbol{\mathcal{O}})

is the moduli space of degree zero RR-semistable meromorphic parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors with residue data lying inside 𝓞\boldsymbol{\mathcal{O}}. If 𝓞\boldsymbol{\mathcal{O}} corresponds to the L𝜽L_{\boldsymbol{\theta}}-orbit of a collection of Levi factors φ𝜽\varphi_{\boldsymbol{\theta}}, by abuse of notation, the moduli space Dol(X,𝒢𝜽,𝑸,𝓞)\mathcal{M}_{\mathrm{Dol}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q},\boldsymbol{\mathcal{O}}) is also denoted as Dol(X,𝒢𝜽,𝑸,φ𝜽)\mathcal{M}_{\mathrm{Dol}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q},\varphi_{\boldsymbol{\theta}}).

As a direct result of the wild nonabelian Hodge correspondence at the level of categories [HS23a, Theorem in §1], we formulate the wild nonabelian Hodge correspondence at the level of moduli spaces.

Theorem 5.2.

Let (𝛂,φ𝛂,𝐐~)(\boldsymbol{\alpha},\varphi_{\boldsymbol{\alpha}},\widetilde{\boldsymbol{Q}}), (𝛃,𝛃,𝐐)(\boldsymbol{\beta},\nabla_{\boldsymbol{\beta}},\boldsymbol{Q}), and (𝛄,M𝛄,𝐐)(\boldsymbol{\gamma},M_{\boldsymbol{\gamma}},\boldsymbol{Q}) be the local data for Dolbeault, de Rham and Betti, respectively. For each x𝐃x\in\boldsymbol{D}, these data are grouped into the following table,

   Dolbeault    de Rham    Betti
weights αx\alpha_{x} βx\beta_{x} γx\gamma_{x}
residues \\backslash monodromies φαx\varphi_{\alpha_{x}} βx\nabla_{\beta_{x}} MγxM_{\gamma_{x}}
irregular types Q~x\widetilde{Q}_{x} QxQ_{x} QxQ_{x}

Let βx=sβx+Yβx\nabla_{\beta_{x}}=s_{\beta_{x}}+Y_{\beta_{x}} be the Jordan decomposition, where sβxs_{\beta_{x}} is the semisimple part and YβxY_{\beta_{x}} is the nilpotent part. We complete YβxY_{\beta_{x}} into an 𝔰𝔩2\mathfrak{sl}_{2}-triple (Xβx,Hβx,Yβx)(X_{\beta_{x}},H_{\beta_{x}},Y_{\beta_{x}}). Suppose these data are subjected into the relations in the following table,

Dolbeault   de Rham Betti
weights 12(sβx+s¯βx)\frac{1}{2}(s_{\beta_{x}}+\bar{s}_{\beta_{x}}) βx\beta_{x} βx12(sβx+s¯βx)\beta_{x}-\frac{1}{2}(s_{\beta_{x}}+\bar{s}_{\beta_{x}})
residues \\backslash monodromies 12(sβxβx)+(YβxHβx+Xβx)\frac{1}{2}(s_{\beta_{x}}-\beta_{x})+(Y_{\beta_{x}}-H_{\beta_{x}}+X_{\beta_{x}}) sβx+Yβxs_{\beta_{x}}+Y_{\beta_{x}} exp(2π1(sβx+Yβx))\exp(-2\pi\sqrt{-1}(s_{\beta_{x}}+Y_{\beta_{x}}))
irregular types Q~x=12Qx\widetilde{Q}_{x}=\frac{1}{2}Q_{x} QxQ_{x} QxQ_{x}

Then the irregular Riemann–Hilbert correspondence (Theorem 2.14) gives rise to an isomorphism of complex analytic spaces

B(an)(X𝑫,G,𝑸,𝜸,M𝜸)dR(an)(X,𝒢𝜷,𝑸,𝜷),\displaystyle\mathcal{M}_{\mathrm{B}}^{\mathrm{(an)}}(X_{\boldsymbol{D}},G,\boldsymbol{Q},\boldsymbol{\gamma},M_{\boldsymbol{\gamma}})\cong\mathcal{M}_{\mathrm{dR}}^{\mathrm{(an)}}(X,\mathcal{G}_{\boldsymbol{\beta}},\boldsymbol{Q},\nabla_{\boldsymbol{\beta}}),

and we also have a homeomorphism of topological spaces

Dol(top)(X,𝒢𝜶,𝑸~,φ𝜶)dR(top)(X,𝒢𝜷,𝑸,𝜷).\displaystyle\mathcal{M}_{\mathrm{Dol}}^{\mathrm{(top)}}(X,\mathcal{G}_{\boldsymbol{\alpha}},\widetilde{\boldsymbol{Q}},\varphi_{\boldsymbol{\alpha}})\cong\mathcal{M}_{\mathrm{dR}}^{\mathrm{(top)}}(X,\mathcal{G}_{\boldsymbol{\beta}},\boldsymbol{Q},\nabla_{\boldsymbol{\beta}}).

As a direct application, the Dolbeault moduli space (and the de Rham and Betti moduli spaces as well) is hyperKähler, we will sketch a proof via wild harmonic principal bundles, weighted Sobolev spaces, and hyperKähler reduction. The hyperKähler geometry for the moduli spaces in the classical nonabelian Hodge theory has been well-studied decades ago, for example, [Hit87, Fuj91]. A similar property is also known for the tame and wild cases (for GLn()\mathrm{GL}_{n}(\mathbb{C}) or SLn()\mathrm{SL}_{n}(\mathbb{C}) as the structure group) from the work of Konno, Nakajima, Biquard, Biquard–Boalch, among others [Kon93, Nak96, Biq97, BB04]. We would like to mention that, besides the hyperKähler property, the completeness of hyperKähler metrics was also studied by Biquard–Boalch [BB04].

Theorem 5.3.

Dol(X,𝒢𝜽,𝑸,φ𝜽)\mathcal{M}_{\mathrm{Dol}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q},\varphi_{\boldsymbol{\theta}}) admits a hyperKähler structure.

Sketch of the proof.

To prove the theorem, here we first introduce some notions from analytic aspects, more details can be found in [HKSZ22, §4] and [HS23a, §2.2].

Let 𝒏={nx,x𝑫}\boldsymbol{n}=\{n_{x},\,x\in\boldsymbol{D}\} be a collection of integers labeled by points in 𝑫\boldsymbol{D}. Given a metrized GG-Higgs bundle (E,¯E,ϕ,h)(E,\bar{\partial}_{E},\phi,h) on X𝑫X_{\boldsymbol{D}}, it is called (𝛉,𝐧)(\boldsymbol{\theta},\boldsymbol{n})-adapted if hh is a 𝜽\boldsymbol{\theta}-adapted hermitian metric (see e.g. [HKSZ22, HS23a] for the definition of adaptedness), and for each x𝑫x\in\boldsymbol{D}, we have

zθx(znxϕ(z))zθxis bounded aszapproaches 0,\displaystyle z^{\theta_{x}}\cdot(z^{n_{x}}\cdot\phi(z))\cdot z^{-\theta_{x}}\ \text{is bounded as}\ z\ \text{approaches}\ 0,

where zz is the local coordinate vanishing at xx. In this case, hh induces an extension of (E,¯E,ϕ)(E,\bar{\partial}_{E},\phi) into a merohoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor (,φ)(\mathcal{E},\varphi) on XX, with φH0(X,(𝔤)KX(𝒏𝑫))\varphi\in H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K_{X}(\boldsymbol{nD})), where 𝒏𝑫:=x𝑫nxx\boldsymbol{nD}:=\sum_{x\in\boldsymbol{D}}n_{x}\cdot x.

Now we give an analytic proof via wild harmonic bundles. Let \mathcal{E} be a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX, and put E:=|X𝑫E:=\mathcal{E}|_{X_{\boldsymbol{D}}}, which is a GG-bundle on X𝑫X_{\boldsymbol{D}}. We fix a background (𝜽,𝒏)(\boldsymbol{\theta},\boldsymbol{n})-adapted GG-Higgs bundle (E,¯0,ϕ0,h)(E,\bar{\partial}_{0},\phi_{0},h) on X𝑫X_{\boldsymbol{D}} with (Levi factor of) residue data lying in 𝓞\boldsymbol{\mathcal{O}}. For each i0i\geq 0, let d0:=h+¯0d_{0}:=\partial_{h}^{\prime}+\bar{\partial}_{0} be the Chern connection. Denote by 𝒜i(E(𝔤))\mathcal{A}^{i}(E(\mathfrak{g})) the space of CC^{\infty}-sections of E(𝔤)ΩX𝑫iE(\mathfrak{g})\otimes\Omega^{i}_{X_{\boldsymbol{D}}}, and let Dkp𝒜i(E(𝔤))D^{p}_{k}\mathcal{A}^{i}(E(\mathfrak{g})) be the Sobolev completion of 𝒜i(E(𝔤))\mathcal{A}^{i}(E(\mathfrak{g})) with respect to the Sobolev norm Dkp\|\bullet\|_{D^{p}_{k}} (see [Kon93] for details).

Note that the metric hh is a section of KK-reduction of EE, i.e. h𝒜0(E/K)h\in\mathcal{A}^{0}(E/K), where KGK\subseteq G is the maximal compact subgroup. Let 𝔨\mathfrak{k} be the Lie algebra of KK, and then we have the Cartan decomposition 𝔤=𝔨i𝔨\mathfrak{g}=\mathfrak{k}\oplus i\mathfrak{k}, which induces the decomposition

E(𝔤)=E(𝔥)iE(𝔥).\displaystyle E(\mathfrak{g})=E(\mathfrak{h})\oplus iE(\mathfrak{h}).

Let F=hEF=h^{*}E be the KK-reduction, which is a KK-bundle with F×KGEF\times_{K}G\cong E. Define

𝔾\displaystyle\mathbb{G} ={gD2p𝒜0(E(G)):g𝒜0(F(K))},\displaystyle=\{g\in D_{2}^{p}\mathcal{A}^{0}(E(G)):\ g\in\mathcal{A}^{0}(F(K))\},
𝔸\displaystyle\mathbb{A} ={(dA,ϕ):dAd0D1p𝒜1(E(𝔥)),ϕϕ0D1p𝒜1,0(E(𝔤))},\displaystyle=\{(d_{A},\phi):d_{A}-d_{0}\in D^{p}_{1}\mathcal{A}^{1}(E(\mathfrak{h})),\phi-\phi_{0}\in D_{1}^{p}\mathcal{A}^{1,0}(E(\mathfrak{g}))\},

there is an adjoint action of the unitary gauge group 𝔾\mathbb{G} on 𝔸\mathbb{A}. The affine space 𝔸\mathbb{A} is of infinite dimensional, and it is hyperKähler. Indeed, its tangent space can be identified with D1p𝒜0,1(E(𝔤))D1p𝒜1,0(E(𝔤))D^{p}_{1}\mathcal{A}^{0,1}(E(\mathfrak{g}))\oplus D_{1}^{p}\mathcal{A}^{1,0}(E(\mathfrak{g})) with Riemannian metric given by

g((ξ1,η1),(ξ2,η2))=iX𝑫Tr(η2ξ1η1ξ2).\displaystyle g((\xi_{1},\eta_{1}),(\xi_{2},\eta_{2}))=i\int_{X_{\boldsymbol{D}}}\mathrm{Tr}(\eta_{2}\wedge\xi_{1}-\eta_{1}\wedge\xi_{2}).

The three complex structures are

I(ξ,η)=(iξ,iη),J(ξ,η)=(iη¯T,iξ¯T),K(ξ,η)=IJ(ξ,η)=(η¯T,ξ¯T).\displaystyle I(\xi,\eta)=(i\xi,i\eta),\ J(\xi,\eta)=(i\bar{\eta}^{\mathrm{T}},-i\bar{\xi}^{\mathrm{T}}),\ K(\xi,\eta)=IJ(\xi,\eta)=(-\bar{\eta}^{\mathrm{T}},\bar{\xi}^{\mathrm{T}}).

The action of 𝔾\mathbb{G} on 𝔸\mathbb{A} admits a hyperKähler moment map 𝝁=(μI,μJ,μK)\boldsymbol{\mu}=(\mu_{I},\mu_{J},\mu_{K}):

(5.1) {μI(dA,ϕ)=FA+[ϕ,ϕ],(μJ+iμK)(dA,ϕ)=2i¯Aϕ.\displaystyle\left\{\begin{aligned} \mu_{I}(d_{A},\phi)&=F_{A}+[\phi,\phi^{*}],\\ (\mu_{J}+i\mu_{K})(d_{A},\phi)&=2i\bar{\partial}_{A}\phi.\end{aligned}\right.

A solution to (5.1) is exactly a wild harmonic bundle on X𝑫X_{\boldsymbol{D}}. Then the hyperKähler quotient 𝝁1(0)/𝔾=(μI1(0)μJ1(0)μK1(0))/𝔾\boldsymbol{\mu}^{-1}(0)/\mathbb{G}=(\mu_{I}^{-1}(0)\cap\mu_{J}^{-1}(0)\cap\mu_{K}^{-1}(0))/\mathbb{G} admits a hyperKähler structure, as the moduli space of (𝜽,𝒏)(\boldsymbol{\theta},\boldsymbol{n})-adapted wild harmonic bundles on X𝑫X_{\boldsymbol{D}} with (Levi factor of) residue data lying in 𝓞\boldsymbol{\mathcal{O}}, which is naturally diffeomorphic to Dol(X,𝒢𝜽,𝑸,φ𝜽)\mathcal{M}_{\mathrm{Dol}}(X,\mathcal{G}_{\boldsymbol{\theta}},\boldsymbol{Q},\varphi_{\boldsymbol{\theta}}).

References

  • [BB04] O. Biquard and P. P. Boalch, Wild non-abelian Hodge theory on curves, Compos. Math. 140 (2004), no. 1, 179–204.
  • [BGPMiR20] O. Biquard, O. García-Prada, and I. Mundet i Riera, Parabolic Higgs bundles and representations of the fundamental group of a punctured surface into a real group, Adv. Math. 372 (2020), 107305, 70.
  • [Biq97] O. Biquard, Fibrés de Higgs et connexions intégrables: le cas logarithmique (diviseur lisse), Ann. Sci. École Norm. Sup. (4) 30 (1997), no. 1, 41–96.
  • [Boa14] P. P. Boalch, Geometry and braiding of Stokes data; fission and wild character varieties, Ann. of Math. (2) 179 (2014), no. 1, 301–365.
  • [Boa18] by same author, Wild character varieties, meromorphic Hitchin systems and Dynkin diagrams, Geometry and physics. Vol. II, Oxford Univ. Press, Oxford, 2018, pp. 433–454.
  • [Boa21] by same author, Topology of the Stokes phenomenon, Integrability, quantization, and geometry. I. Integrable systems, Proc. Sympos. Pure Math., vol. 103, Amer. Math. Soc., Providence, RI, [2021] ©2021, pp. 55–100.
  • [BS15] V. Balaji and C. S. Seshadri, Moduli of parahoric g-torsors on a compact riemann surface, J. Algebraic Geom. 24 (2015), no. 1, 1–49.
  • [BV83] D. G. Babbitt and V. S. Varadarajan, Formal reduction theory of meromorphic differential equations: a group theoretic view, Pacific J. Math. 109 (1983), no. 1, 1–80.
  • [BY15] P. P. Boalch and D. Yamakawa, Twisted wild character varieties, arXiv:1512.0809 (2015).
  • [BY23] by same author, Polystability of stokes representations and differential galois groups, arXiv:2301.09067 (2023).
  • [DDP18] D.-E. Diaconescu, R. Donagi, and T. Pantev, BPS states, torus links and wild character varieties, Comm. Math. Phys. 359 (2018), no. 3, 1027–1078.
  • [Fuj91] A. Fujiki, Hyper-Kähler structure on the moduli space of flat bundles, Prospects in complex geometry (Katata and Kyoto, 1989), Lecture Notes in Math., vol. 1468, Springer, Berlin, 1991, pp. 1–83.
  • [Hit87] N. J. Hitchin, The self-duality equations on a riemann surface, Proceedings of the London Mathematical Society 3 (1987), no. 1, 59–126.
  • [HJ24] A. Hohl and K. Jakob, Stokes phenomenon of kloosterman and airy connections, arXiv:2404.09582 (2024).
  • [HKSZ22] P. Huang, G. Kydonakis, H. Sun, and L. Zhao, Tame parahoric nonabelian hodge correspondence on curves, arXiv:2205.15475 (2022).
  • [HMW19] T. Hausel, M. Mereb, and M. L. Wong, Arithmetic and representation theory of wild character varieties, J. Eur. Math. Soc. (JEMS) 21 (2019), no. 10, 2995–3052.
  • [HS23a] P. Huang and H. Sun, Meromorphic parahoric Higgs torsors and filtered Stokes GG-local systems on curves, Adv. Math. 429 (2023), Paper No. 109183, 38.
  • [HS23b] by same author, Moduli spaces of filtered G-local systems on curves, arXiv:2304.09999 (2023).
  • [Kin94] A. D. King, Moduli of representations of finite-dimensional algebras, Quart. J. Math. Oxford Ser. (2) 45 (1994), no. 180, 515–530.
  • [Kon93] H. Konno, Construction of the moduli space of stable parabolic Higgs bundles on a Riemann surface, J. Math. Soc. Japan 45 (1993), no. 2, 253–276.
  • [KSZ23] G. Kydonakis, H. Sun, and L. Zhao, Logahoric Higgs torsors for a complex reductive group, Mathematische Annalen (2023).
  • [LR94] M. Loday-Richaud, Stokes phenomenon, multisummability and differential Galois groups, Ann. Inst. Fourier (Grenoble) 44 (1994), no. 3, 849–906.
  • [Mal78] B. Malgrange, Remarques sur les équations différentielles à points singuliers irréguliers (Séminaire Goulaouic-Schwartz 1976–1977), Séminaire Goulaouic-Schwartz (1977/1978), École Polytech., Palaiseau, 1978, pp. Exp. No. 25, 10.
  • [Mal83] by same author, La classification des connexions irrégulières à une variable, Mathematics and physics (Paris, 1979/1982), Progr. Math., vol. 37, Birkhäuser Boston, Boston, MA, 1983, pp. 381–399.
  • [MiR18] I. Mundet i Riera, Parabolic Higgs bundles for real reductive Lie groups: a very basic introduction, Geometry and physics. Vol. II, Oxford Univ. Press, Oxford, 2018, pp. 653–679.
  • [Nak96] H. Nakajima, Hyper-Kähler structures on moduli spaces of parabolic Higgs bundles on Riemann surfaces, Moduli of vector bundles (Sanda, 1994; Kyoto, 1994), Lecture Notes in Pure and Appl. Math., vol. 179, Dekker, New York, 1996, pp. 199–208.
  • [Ram75] A. Ramanathan, Stable principal bundles on a compact Riemann surface, Math. Ann. 213 (1975), 129–152.
  • [Ram96] by same author, Moduli for principal bundles over algebraic curves. I, Proc. Indian Acad. Sci. Math. Sci. 106 (1996), no. 3, 301–328.
  • [Sib77] Y. Sibuya, Stokes phenomena, Bull. Amer. Math. Soc. 83 (1977), no. 5, 1075–1077.
  • [Sim90] C. T. Simpson, Harmonic bundles on noncompact curves, J. Amer. Math. Soc. 3 (1990), no. 3, 713–770.
  • [vdPS03] M. van der Put and M. Singer, Galois theory of linear differential equations, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 328, Springer-Verlag, Berlin, 2003.

Max Planck Institute for Mathematics in the Sciences
Inselstraße 22, 04103 Leipzig, Germany
E-mail address: pfhwangmath@gmail.com


Department of Mathematics, South China University of Technology
381 Wushan Rd, Tianhe Qu, Guangzhou, Guangdong, China
E-mail address: hsun71275@scut.edu.cn